<<

arXiv:2103.07178v1 [math.AP] 12 Mar 2021 iiy lxnrffFnhlinequalities. Alexandroff-Fenchel bility; ueflw” rn ubrSH 1879/3-1. SCHE number ”Quermassinte Grant Project flows”; ture foundation); research German (DFG, .Itouto n results and Introduction 1. hswr a atal uddfne yte”etceForsc ”Deutsche the by funded funded partially was work This ..Ams osatcrauefunctions curvature constant inequality Almost Heintze-Karcher the results in 4.2. stability Stability sets geometric level the for of 4.1. estimates Proofs umbilicity almost of 4. Proof Notation and 3. Definitions eigenvalues Steklov for 2. Estimates problem overdetermined 1.7. Serrin’s inequalities Alexandroff-Fenchel 1.6. functions Non-convex curvature constant of 1.5. Hypersurfaces inequality Heintze-Karcher 1.4. sets The level and functions 1.3. Defining Stability vs. 1.2. Rigidity 1.1. e od n phrases. and words Key Date problem eigenvalue Steklov References order fourth source a nonlinear in Acknowledgments with Stability problem Serrin’s in 4.5. Stability inequalit Alexandroff-Fenchel non-convex in 4.4. Stability 4.3. ac 5 2021. 15, March : TBLT RMRGDT I UMBILICITY VIA RIGIDITY FROM STABILITY rtradnncne lxnrffFnhlinequalities. Alexandroff-Fenchel non-convex form bi and the space involving erator in problem Steklov theorem Examp a bubble problem, constraint. overdetermined soap geometric Alexandroff’s a pr in given stability sph stability hypersurface, for new ask a and problems of Those old ness many analysis. and of geometry treatment in application arising unified As a hypersurface. for the by allows bounded set open the for h asoffdsac oagoei peeo nebde hyp embedded an estim of an the sphere is geodesic to results a key to distance the Hausdorff of the One spaceforms. of persurfaces Abstract. L 2 nr ftetaeesHsinoeao fadfiigfunctio defining a of operator Hessian traceless the of -norm ecnie ag fgoercsaiiypolm o hy- for problems stability geometric of range a consider We opbbl hoe;Sri’ rbe;Selvpolm St problem; Steklov problem; Serrin’s theorem; bubble Soap UINSCHEUER JULIAN Contents 1 rlpeevn oa curva- local preserving gral Lpaeop- -Laplace t relating ate rclclose- erical hsresult this , hungsgemeinschaft” e include les ,Serrin’s s, ersurface ies oblems n 30 29 29 27 25 22 20 20 15 13 12 11 10 a- 7 6 4 2 2 2 J. SCHEUER

1. Introduction and results 1.1. Rigidity vs. Stability. This paper is about rigidity theorems in clas- sical hypersurface theory and their stability properties. An example is the soap bubble theorem, first proved by Alexandroff: Every constant mean cur- vature hypersurface embedded in Euclidean space, hyperbolic space or in a half-sphere of dimension n + 1, n 2, must be a round sphere, see [Ale62]. Way earlier Liebmann proved the convex≥ case [Lie00] for hypersurfaces of the Euclidean space. Even earlier Jellet [Jel53] proved the two-dimensional case for starshaped surfaces, following up the treatment of surfaces of revolution due to Delaunay [Del41]. A related famous result, which follows from Liebmann’s theorem and the Codazzi equation, is the so-called Nabelpunktsatz seemingly proven by Darboux1: For n 2, every totally umbilic closed, immersed hypersurface of the (n + 1)-dimensional≥ Euclidean space is a round sphere. Both of these theorems are instances of so-called rigidity results: A partic- ular functional defined on a subclass N of the class of closed, immersed and orientable hypersurfacesF M of a givenI Riemannian N dictates their global shape. Abstractly this can be described as follows. Let = f : S R S N G { → | ⊂ } be the set of all functions defined on a subset of N. Then both of the above examples describe particular level sets of a functional : . F IN → G In Liebmann’s theorem we have = M Rn+1 : M closed,h> 0 , (M)= H. IN { ⊂ } F1 Here h is the second fundamental form of M and H its trace with respect to the induced metric. In Alexandroff’s theorem we have = M N : M closed, embedded , IN { ⊂ } where N is Euclidean, hyperbolic or spherical. In the Nabelpunktsatz we instead use (M)= A˚ 2 = A 2 1 H2, F2 | | | | − n where A is the Weingarten operator. Then the above theorems can be phrased as 1 1 (1.1) − ( f = c: c R )= M N : M = sphere = − ( f 0 ). F1 { ∈ } { ⊂ } F2 { ≡ } In the course of this paper, we will encounter many more examples of such functionals. In this abstract context, the next natural question is whether we can describe enlarged preimage sets of . (1.1) characterises the preimages of particular constant functions. We areF interested in the description of sets like 1( f c < ǫ: c R ), for some small ǫ and a norm to be F − {k − k ∈ } k · k 1I did not find a clean reference. STABILITY FROM RIGIDITY VIA UMBILICITY 3 specified. Formally, if 2B is the power set of a set B, then the associated question of stability aims to relax the constancy in the following sense: 1 Is the map − : 2G 2I continuous at all arguments, the values of which F → we can characterise? Of course, the question of continuity has to be addressed with respect to topologies to be specified within particular problems. For example, the stability question for the Nabelpunktsatz would trans- late to whether a hypersurface is close to sphere, whenever it is almost umbilical. This question has received plenty of attention, initiated by early works of Reshetnyak [Res68] and Pogorelov [Pog73] for convex surfaces with a follow up by Leichtweiß, see [Lei99]. The study of almost umbilical hyper- surfaces in the non-convex class seems pioneered by works of De Lellis and M¨uller [DLM05, DLM06], who give a W 2,2-closeness, resp. a C0-closeness in terms of an L2-pinching of A˚ 2 in Rn+1. The W 2,2-closeness result was reproved with the help of the| Willmore| flow in [KS20]. The version for hyperbolic and spherical spaces was treated in [CZ14] and some optimality results in [CJ15]. Further results of almost umbilical type in terms of other Lp-norms were deduced in [Per11, Rot13, Rot15, RS18, Sch15]. In partic- ular we will make use of the following estimate from [RS18, Thm. 1.1] for a closed, connected, oriented and immersed hypersurface M Rn+1 and p>n 2: ⊂ ≥ A˚ α (1.2) dist(M,S ) CR k kp , R ≤ H α k kp with constants α = α(n,p) and C = C(n,p, A p, M ). Here p is the Lp-norm with respect to the surface area measure.k k | | k · k For the soap bubble theorem most of the stability results available con- trol the Hausdorff-distance of the hypersurface to a sphere (or an array of spheres) in terms of various pinching assumptions on the mean curvature. In the class of convex hypersurfaces such results are pioneered by Koutroufiotis [Kou71] and Moore [Moo73], also see [Arn93] for a rather convex geometric approach. In the non-convex class the geometry of the stability problem is more complicated, since bubbling (closeness to a connected sum of spheres with small necks) can occur in case the curvature of the hypersurfaces is un- controlled. In terms of a C0-pinching condition, stability results in the Eu- clidean and hyperbolic space were proven by Ciraolo/Vezzoni [CV18, CV20], where in both cases the authors used a quantitative analysis of Alexandroff’s proof by moving planes. In [CM17] similar stability results were obtained, but with a proof that relies on the stability analysis of Ros’ proof of the soap bubble theorem via the Heintze-Karcher inequality [Ros87]. In parallel to these developments, various stability results by Magnani- ni/Poggesi appeared relaxing the C0-pinching condition to an L1-pinching condition. The key is, that the soap bubble theorem can also be proved by integral methods, see [Ros87] and also compare [Rei82]. With the help 4 J. SCHEUER of integral formulae, which are valid for solutions f of the torsion Dirichlet problem ∆¯ f =1 inΩ f =0 on ∂Ω, Magnanini/Poggesi [MP19, MP20a, MP20b] were able to prove that what they call Cauchy-Schwarz deficit ˚¯ 2f 2 := ¯ 2f 2 1 (∆¯ f)2 |∇ | |∇ | − n+1 can be controlled by the L1 deficit of the mean curvature with respect to a constant. Comparison of f with the squared distance function gave closeness of ∂Ω to a sphere after employing various estimates for harmonic functions. The key idea of the present paper is that it is possible to control the distance to a sphere by an L2-norm of the Cauchy-Schwarz deficit, without relying on f being a solution to a PDE. This flexibility will allow us to treat a broader range of stability results, such as an extension of the results by Magnanini/Poggesi to spaceforms of constant curvature, stability results for the deficit in the Heintze-Karcher inequality in spaceforms, a double- stability result for Serrin’s overdetermined problem with nonlinear source term, as well as a stability result in bi-Laplace Steklov eigenvalue problems. Everything comes with a price: The cost of this flexibility is that in par- ticular geometric problems, some constants may be improved with methods tailor-made for the problem at hand. Let’s continue this introduction with a precise description of the results and more background information on the specific geometric problems. 1.2. Defining functions and level sets. A key theorem in our approach, which can be viewed as a generalisation and quantitative version of a result due to Reilly [Rei80, Lemma 3], that a function f on a domain Ω Rn+1, which satisfies for some constant L ⊂ ¯ 2 f = Lg,¯ f ∂Ω = 0, ∇ | is the squared distance function to a suitable point up to normalisation constants, is as follows: 1.1. Theorem. Let n 2 and let (N n+1, g¯) be a smooth conformally flat Riemannian domain, i.e.≥ N Rn+1 is open and ⊂ g¯ = e2ψ , h· ·i 2 with ψ C∞(N). Let Ω ⋐ N be an open set with C -boundary and let f C2(∈Ω)¯ satisfy ∈ ¯ (1.3) f ∂Ω = 0, f ∂Ω = 0. | ∇ | 6 Then for every connected component M of ∂Ω there exist constants α = α(n) and 1 2 1 M n ¯ f ,Ω C = C n, ψ ,Ω, M n ¯ ψ ,Ω, | | k∇ k∞ , k k∞ | | k∇ k∞ min ¯ f M |∇ | ! STABILITY FROM RIGIDITY VIA UMBILICITY 5 such that α 1 1 2 n+1 dist(M,S) C M n ˚¯ f ≤ | | min ¯ f n+1 ˆ |∇ |  M |∇ | Ω  for a totally umbilical hypersurface S N. ⊂ Here ¯ is the Levi-Civita connection ofg ¯, M is the surface area of M, ∇ | | ,Ω the sup-norm and dist the Hausdorff distance of sets. Ω ⋐ N means k · k∞ that Ω¯ is compact and Ck(Ω)¯ is the usual space of k-times differentiable functions in Ω where all derivatives are continuous up to the boundary. A function with the properties (1.3) is called defining function for Ω in the sequel. In many applications one has an estimate on the L2-norm of the traceless Hessian, hence we also state the following corollary. 1.2. Corollary. Under the assumptions of Theorem 1.1, for every connected component M of ∂Ω there exist constants α = α(n) and 1 2 1 M n ¯ f ,Ω C = C n, ψ ,Ω, M n ¯ ψ ,Ω, | | k∇ k∞ , k k∞ | | k∇ k∞ min ¯ f M |∇ | ! such that − − α 1 α(n 1) 1 2 2 dist(M,S) C M n ˚¯ f ≤ | | min ¯ f 2 ˆ |∇ |  M |∇ | Ω  for a totally umbilical hypersurface S. Theorem 1.1 will be proven with the help of an almost umbilical type theorem [RS18] in Section 3. The key feature of Corollary 1.2 is the following procedure: Whenever a rigidity result is proven by discarding an L2-norm of the Cauchy-Schwarz deficit of a defining function for Ω, you get an associated stability result. We will see plenty of examples throughout the following discussion and there are many more slumbering in the literature. 1.3. Remark. (i) Note that results like the one above can not hold without some dependence of the constant C on the Hessian of f, due to the possibility of bubbling, see for example [CV18, Rem. 5.2] for further details. Hence all of the stability results in this paper will come with some sort of curvature dependence. This also means that the choice of the Ln+1-norm in Theorem 1.1 is not essential and can be replaced by any Lp norm of the traceless Hessian. For scaling reasons the formu- lation in terms of the Ln+1-norm gives the most aesthetic form. (ii) The proof of this theorem crucially depends on the almost-umbilicity type estimate (1.2), which holds for a large class of hypersurfaces. In case that more properties are being given, such as convexity or mean- convexity, one can say more about the value of α, see for example [Lei99] for estimates in the convex class. 6 J. SCHEUER

(iii) Note that in the Euclidean space one may scale the problem to M = min ¯ f = 1 | | M |∇ | whenever convenient. 1.3. The Heintze-Karcher inequality. The following inequality, known today as the Heintze-Karcher inequality, was proved by Ros [Ros87] upon inspiration by related inequalities of Heintze/Karcher [HK78]. It goes as follows. For a closed and mean convex hypersurface M = ∂Ω Rn+1 we have ⊂ 1 (n + 1) Ω = u, ˆM H1 ≥ | | ˆM with equality precisely when M is a round sphere. Here u is the support function u = x,ν h i and nH1 = H. In particular this trivially yields a new proof of Alexandroff’s theorem, after invoking the formula

M = uH1. | | ˆM As to the soap bubble theorem, there was of course nothing new to be discovered from that method, however, Ros [Ros87] used this method to prove that constant Hk hypersurfaces must be spheres. We will recall that proof in Section 4 in the course of proving the stability result. In fact, Ros proved this inequality in ambient spaces of non-negative , however the equality case forces the ambient space to be Euclidean. For constant curvature spaceforms, refined versions of the Heintze-Karcher in- equality have emerged during the recent years. Using flow methods, Brendle [Bre13] proved that for closed and mean-convex hypersurfaces M = ∂Ω of the half-sphere or the hyperbolic space there holds ϑ ′ u, ˆM H1 ≥ ˆM with equality precisely on geodesic balls. Here

u =g ¯(ϑ∂r,ν) and the metric of the ambient space is written in polar coordinates as g¯ = dr2 + ϑ2(r)σ, with the round metric σ. This result was, among others, recovered by Qiu/Xia [QX15] via elliptic methods, i.e. by employing a new Reilly type formula, see (4.2) for further details. Similarly as above, those inequalities can be used to reprove Alexandroff’s theorem in spaceforms. But also on other occasions the Heintze-Karcher inequality has proven useful, for ex- ample for the deduction of isoperimetric type inequalities in Riemannian , see for example [SX19]. STABILITY FROM RIGIDITY VIA UMBILICITY 7

Following the previous discussion, it becomes an interesting question whether we have an associated stability result. Namely, does almost-equality of the Heintze-Karcher inequality imply any type of closeness to the sphere? For the Euclidean case there are the recent works of Magnanini/Poggesi [MP19, MP20a]. They prove that for a C2-hypersurface M = ∂Ω there holds τn 1 2 dist(M,S) C (n + 1) Ω , ≤ ˆ H − | |  M 1  for some τn and with a constant C depending on n, a lower inball radius bound for Ω at M and on the diameter of Ω. In Section 4 we will prove the following theorem for hypersurfaces of non-Euclidean spaceforms MK of n+1 sectional curvature K, where M1 = S+ is the open northern hemisphere. The proof applies to K = 0 too, but that result would not be new. In the following, if a constant depends on maxM r, in case K = 1 this means the constant may blow up at the equator r = π/2 . { } n 1.4. Theorem. Let n 2, β > 0 and M MK , K = 1, be a closed, embedded and strictly mean≥ convex C2,β-hypersurface⊂ with uniform± interior sphere condition with radius ρ. Then there exist constants α = α(n) and 1 C = C(n, β, ρ− , M , max r M , A β,M ), such that | | | | | α 1−α(n−1) ϑ′ dist(M,S) C M n u , ≤ | | ˆ H − ˆ  M 1 M  for a suitable geodesic sphere S. Here A is the H¨older norm of the Weingarten operator on M. | |β,M 1.4. Hypersurfaces of constant curvature functions. In [Ale62], Alex- androff did not only prove the soap bubble theorem. The robustness of the moving plane method allowed to treat much more general elliptic equations, also see [Vos56] and [Hsi56] as well as [Bre13, DM19, KLP18, Mon99, WX14] fur further developments. His theorem says that every embedded hypersur- face with

(1.4) F (κ1,...,κn) = const is a sphere. Here F is a strictly monotone function of the principal curvatures κi of the hypersurface. Particular important examples are the higher order mean curvatures 1 Hk = n κi1 ...κik ,H0 := 1, k 1 i1< 0 and M MK, K = 1, be a closed, em- bedded C2,β-hypersurface≥ with uniform interior⊂ sphere condition± with radius ρ. Then there exist constants α = α(n) and C as in Theorem 1.4, such that α − − 1 α(n 1) n M ϑ′ dist(M,S) C M n ϑ′ ´ H1 ≤ | | (n + 1) ϑ′ −  Ω + 1,M ´ for a suitable geodesic sphere S, where f+ denotes the positive part of a function f. Despite the wide variety of stability results concerning the soap bubble theorem, i.e. F = H1, much less seems to be known when it comes to stability results for other F . While in the convex class there are many re- sults available, e.g. [Arn93, GS91, Kou71, KU01, Moo73], in the non-convex class for more general F there is [CRV18], who quantified the moving plane method and obtained spherical closeness in terms of a C0-pinching condi- tion, generalising the corresponding results for F = H from [CV18, CV20]. In Section 4.2 we extend the class of stability results for non-convex hyper- surfaces and prove stability for a variety of nonlinear curvature operators, employing the method of almost umbilicity. Namely we treat the case of curvature quotients. We show that the distance to a sphere can be con- 1 trolled by an L -norm of the deviation of F = Hk+1/Hℓ, 0 ℓ k, from a constant. This weakens the pinching assumption in [CRV18]≤ from≤ C0 to L1. A particular highlight of this theorem is that in case ℓ = k the curvature STABILITY FROM RIGIDITY VIA UMBILICITY 9 assumptions do not imply the ellipticity of the operator F . This is some- what surprising and means that there is no chance to prove this result with previous methods, such as moving planes. In the following we denote 1 H 1 = − H1 and 2 ∂ Hk+1 Hk+1,n1 = . ∂κn∂κ1 n 1.6. Theorem. Let n 2, 1 k n 1 and 0 ℓ k. Let M MK , K = 0, 1, 1, be a closed,≥ immersed≤ ≤ and− orientable≤ hypersurface,≤ such⊂ that κ Γ and− in case ℓ k 1 suppose κ Γ . Then there exist constants ∈ k ≤ − ∈ k+1 α = α(n) and C = C(n, M , maxM r, A ,M , minM Hk+1,n1), such that | | k k∞ α M ϑ′Hk Hk+1 dist(M,S) C ϑ′Hℓ 1 ´ ≤ − ϑ′Hℓ 1 − Hℓ  M − + 1,M ´ for a suitable geodesic sphere S.

1.7. Remark. Note that for ℓ 1, this theorem quantifies the result in ≥ [KL01], i.e. immersions with constant quotient Hk+1/Hℓ are spheres. This is because if H k+1 = , Hℓ H then due to the Hsiung identities we have

ϑ′Hk = uHk+1 = uHℓ = ϑ′Hℓ 1 ˆM ˆM H ˆM H ˆM − and hence we are comparing to the “correct” constant. Note however, that something interesting happens when ℓ = 0. In this case the correct constant must be

M ϑ′Hk Hk+1 = ´ . M u Hence the theorem does provide a stability´ result in spaceforms for the result of Ros [Ros87]. However, we do not obtain that every immersed, constant Hk+1 hypersurface, 1 k n 2, must be a sphere. To my knowledge this question is still open.≤ Since≤ this− statement is wrong for k = 0, it seems that very different methods would be required to attack this open problem. However we do obtain the following corollary in this direction, which says that for k 1, a constant Hk+1 immersed closed hypersurface, on which the Heintze-Karcher≥ inequality holds, must be a sphere. 1.8. Corollary. Let n 2 and 1 k n 2. Let M n M , K = 0, 1, 1, ≥ ≤ ≤ − ⊂ K − be a closed, immersed and orientable hypersurface. If Hk+1 = const and ϑ u ′ , ˆM ≤ ˆM H1 10 J. SCHEUER then M is a round sphere. Proof. If H = , then k+1 H M ϑ′Hk M ϑ′Hk = ´ ´ H M u ≥ M ϑ′H 1 ´ ´ − and hence M ϑ′Hk ´ Hk+1 = 0. ϑ′H 1 −  M − + ´ 

1.5. Non-convex Alexandroff-Fenchel inequalities. Let Ω ⋐ Rn+1 be a convex domain. The classical Steiner formula [Ste13] says that the volume expansion of ǫ-parallel bodies is given by n+1 n + 1 Ω+¯ ǫB = W (Ω)ǫk ǫ 0. | | k k ∀ ≥ Xk=0   Here B is the standard unit ball. The coefficients Wk(Ω) are called the quermassintegrals of Ω.¯ It can be shown easily that in case M = ∂Ω is of class C2, then 1 W0(Ω) = Ω , Wk(Ω) = Hk 1, 1 k n + 1, | | n + 1 ˆM − ≤ ≤ e.g. [Sch14]. The classical Alexandroff-Fenchel inequalities [Ale37, Ale38] in this setting say that

n−k W (Ω) W (Ω) n+1−k k+1 k , 0 k n, W (B) ≥ W (B) ≤ ≤ k+1  k  with equality precisely if Ω is a ball. For k = 0 this is the isoperimet- ric inequality. With the help of inverse curvature flows, those inequalities were generalized by Guan/Li [GL09] to domains which are starshaped and k-convex, i.e. the curvatures of M lie in the Garding cone Γk. It is still open whether the starshapedness assumption can be dropped without re- placement. Due to the characterisation of the equality case, we may also ask for stability properties of these inequalities. In the convex case, the Haus- dorff distance to a sphere in terms of the Alexandroff-Fenchel discrepancy was estimated by Schneider [Sch89] and Groemer/Schneider [GS91], also see [Iva15, Iva16] for similar stability results in the class of convex bodies. However, except for the case of the isoperimetric inequality, for which the stability question has been studied extensively, e.g. [FMP08, FMP10, Iva14], I am not aware of a stability result for the non-convex Alexandroff-Fenchel inequalities of Guan/Li. In Section 4.3 we will combine a curvature flow with the almost umbilicity theorem from [RS18] to prove a stability result for the non-convex Alexandroff-Fenchel inequalities. STABILITY FROM RIGIDITY VIA UMBILICITY 11

n 1.9. Theorem. Let n 2 and 1 k n. Let M M0 be a closed and 2 ≥ ≤ ≤ ⊂ starshaped C -hypersurface, such that κ Γk. Then there exist constants 1 ∈ 1 α = α(n) and C = C(n, (minM r)− , r M 2, ,M , minM dist(κ, ∂Γk)− ) with k | k ∞ n−k α W (Ω) W (Ω) n−k+1 dist(M,S ) CR k+1 k R ≤ W (B) − W (B) k+1  k  ! for a sphere SR with radius R. 1.6. Serrin’s overdetermined problem. Very related to the soap bub- ble theorem and the Heintze-Karcher inequality are overdetermined elliptic problems. From the unique solvability of the Dirichlet problem ∆¯ f =1 inΩ (1.5) f =0 on ∂Ω it is a natural question to ask how much information a further condition, such as a Neumann condition, gives. Interestingly, Serrin [Ser71] proved that the additional condition

∂νf = const implies that f is, up to some normalisation, the squared radial distance and that Ω must be a ball. Immediately after this result appeared, Weinberger [Wei71] gave a simplified proof by showing that a good choice of test function yield the Cauchy-Schwarz deficit of f to vanish, from which the result fol- lows. The stability results of the present paper yield into a similar direction, as we use a quantitative version of this argument employing Corollary 1.2. Serrin’s result was generalized to the hyperbolic space and the hemisphere by Kumaresan/Prajapat [KP98] via moving planes. Using arguments simi- lar to that of Weinberger, similar results in a class of warped products and in spaceforms were obtained in [Ron18] and in [CV19]. The stability question associated to Serrin’s problem asks, whether (M)= ∂ f R F | ν − | can serve as a pinching quantity for spherical closeness. In the Euclidean space, under a C0-condition on ¯ f this was answered affirmatively in [ABR99, BNST08a, CMV16]. When|∇ it| comes to pinching in terms of an inte- gral quantity Magnanini/Poggesi [MP20b] have shown the following identity for solutions of (1.5):

˚2 2 1 2 2 ( f) ¯ f = ((∂ν f) R )(∂ν f ∂νq), ˆΩ − |∇ | 2 ˆM − − where q is any squared distance function and R is a suitable constant. A similar approach was taken by Feldman [Fel18] and an interesting alternative approach via quermassintegrals in [BNST08b]. This beautiful formula shows that pinches the Cauchy-Schwarz deficit and hence we could directly apply CorollaryF 1.2 to obtain spherical closeness, although the latter is not what is done in [MP20b], but the authors proceed via harmonic functions; note that 12 J. SCHEUER h = f q is harmonic. Although their approach is well suitable for Serrin’s problem,− the approach via Corollary 1.2 seems to be more flexible when it comes to other overdetermined problems. In particular, the flexibility of Corollary 1.2 allows us to prove the following double stability result for Serrin’s problem, which allows for nonlinear right hand sides. 2 1.10. Theorem. Let n 2, Ω ⋐ M0 be an open set with connected C - boundary, uniform interior≥ sphere condition with radius ρ and ∂Ω = 1. Let φ C2(( , 0]) be positive and suppose that f C2(Ω)¯ satisfies| | ∈ −∞ ∈ ∆¯ f = φ(f) in Ω f = 0 on ∂Ω.

2 1 There exist constants α(n) and C = C(n, ¯ f ,Ω, φ ,Ω, 1/φ ,Ω, ρ− ) and a round sphere S, such that k∇ k∞ k k∞ k k∞ dist(∂Ω,S) C ¯ f R + φ φ(0) + Φ fφ α , ≤ k|∇ |− k1,∂Ω k − k1,Ω k − k1,Ω where  1 0 R = φ(f), Φ(f)= φ(s) ds. M ˆ − ˆ | | Ω f 1.11. Remark. Of course, the second derivative of f can be estimated in terms of φ and Ω using elliptic estimates. But as we do not aim to keep track on how Ω enters precisely in this estimate, we make the statement as general as possible. 1.7. Estimates for Steklov eigenvalues. The Steklov eigenvalue prob- lem, introduced by Vladimir Steklov [Ste02] models the steady state tem- perature in a domain Ω Rn+1, such that the flux at the boundary is proportional to the temperature,⊂ i.e. ∆¯ f =0 inΩ

∂νf = µf on M = ∂Ω. The connection of the Steklov problem to free boundary minimal surfaces was treated by Fraser/Schoen [FS11, FS16]. Various upper bounds for the first Steklov eigenvalue µ1 have been obtained in the literature, for example [IM11, Rot20, WX09]. For example we have the result of Wang/Xia [WX09], which says that a uniformly convex domain with a lower bound c on the principal curvatures of M yields the estimate λ (M) (1.6) µ 1 λ (M)+ λ (M) nc2 , 1 ≤ nc 1 1 − p   with equality precisely if M is ap round sphere.p Here λ1 is the first nonzero eigenvalue of the Laplacian on M. We refer to the survey [GP17] for a much broader description of this vastly researched topic. Although it is possible to obtain a stability result for (1.6) with the help of an almost-CMC result, here we will focus on another problem, that can be solved by direct application STABILITY FROM RIGIDITY VIA UMBILICITY 13 of Corollary 1.2. Namely we prove a stability result for a Steklov problem involving the bi-Laplace: ∆¯ 2w =0 inΩ (1.7) w = ∆¯ w µ ∂ w =0 on ∂Ω. − 1 ν This problem arises from applications in elasticity. The following theorem is a stability version of the rigidity result [WX09, Thm. 1.2], that whenever H1 c and µ1 = (n+1)c, then ∂Ω is a round sphere. Note that the following theorem≥ even improves this rigidity result, in the sense that we do not need a priori to assume the lower bound on H1. 1.12. Theorem. Let n 2 and Ω ⋐ M be an open set with smooth and ≥ 0 connected boundary and ∂Ω = 1. Let µ1 be the first nonzero eigenvalue of the problem (1.7) and suppose| | w is an eigenfunction normalised to min ¯ w = 1. ∂Ω |∇ |

Then there exist constants α = α(n) and C = C(n, w 2, ,Ω), such that for some round sphere S there holds k k ∞ dist(∂Ω,S) C (µ (n + 1)H ) α . ≤ k 1 − 1 +k1,∂Ω

2. Definitions and Notation Let us collect some notation we use throughout the paper. Let N Rn+1 be open andg ¯ be a Riemannian metric on N. Then we call the pair⊂ (N, g¯) a Riemannian domain. Let ¯ denote the Levi-Civita connection ofg ¯. For a hypersurface x: M∇ ֒ N, induced geometric quantities such as the induced metric and its connection→ are distinguished from their ambient counterparts by missing an overbar, e.g. g =g ¯∗, and ∆. We write M for the volume of the Riemannian manifold (M, g).∇ The second fundamental| | form h of M in N are defined by the Gaussian formula ¯ Y = Y h(X,Y )ν, ∇X ∇X − where ν is a given local smooth normal, which will be chosen as “outward pointing”, whenever such a choice is possible. For the associated Weingarten operator we write A, i.e. there holds h(X,Y )= g(A(X),Y )= g(X, A(Y )). This operator is g-selfadjoint and its eigenvalues are called the principal curvatures and usually written as

κ = (κ1,...,κn). A key quantity in this paper is the traceless Weingarten operator 1 A˚ = A H id, − n 14 J. SCHEUER

ij where H = tr A = g hij is the mean curvature, indices denote components with respect to a local frame (ei)1 i n, ≤ ≤ gij = g(ei, ej), hij = h(ei, ej), and where (gij ) is the “inverse” of g, i.e. i ik δj = g gkj i ˚ with the Kronecker-delta δj. M is called totally umbilic, if A = 0. It is well known that in spaceforms all closed, totally umbilic hypersurfaces are geodesic spheres. Let us also note that latin indices denote components for hypersurfaces M N, and we use greek indices to distinguish indices of ambient quantities⊂ defined on N, e.g

g¯αβ =g ¯(eα, eβ). For a set U N and T a tensor field defined on U we use the standard essential sup-norm⊂ T ,U = sup T , k k∞ U | | where norms of tensors are formed with respect to the metric at hand. For a tensor T on a submanifold M N and 1 p< we use the normalized Lp-norm ⊂ ≤ ∞ 1 1 p T = T p . k kp,M M ˆ | | | | M  We avoid to write the volume element in such integrals, as the domain is always understood to be equipped with the standard volume element coming from the induced metric in question. For β > 0, we denote by T β,M the H¨older norm with exponent β, which can simply be defined using| coordinate| charts and the standard definition of H¨older norms in Euclidean space. Often the Riemannian domains we are working with are models for the (n + 1)-dimensional Euclidean space, the hyperbolic space and the half- sphere, i.e. the upper hemisphere of the standard round unit sphere of the Euclidean space. We denote all of them by MK , where K is the constant sectional curvature of the respective space. Introducing polar coordinates around a respective origin in these spaces, the ambient metricsg ¯ can all be written as g¯ = dr2 + ϑ2(r)σ, where σ is the round metric on the sphere Sn and r, K = 0 ϑ(r)= sin r, K = 1 sinh r, K = 1. − For an immersed, orientable hypersurface M M the higher order mean  K curvatures ⊂ 1 Hk = n κi1 ...κik k 1 i1<

ϑ′Hk = uHk+1, k = 0,...,n 1. ˆM ˆM − Here u is the support function

u =g ¯(ϑ∂r,ν) with a smooth unit normal ν and we use the convention H0 = 1. Finally, a comment on constants. We consistently use the letter C for a generic constant that might blow up with its arguments. For example, if we write C = C(n, M , A ,M ), | | k k∞ then we understand C as a function of the indicated arguments that is bounded, as long the arguments are bounded. One exception will be made: In case a result involves an estimate on the unit half-sphere M1 and maxM r belongs to the arguments of C, then C is bounded as long as maxM r is bounded away from π/2. This translates into saying that M has controlled distance from the equator. Occasionally we write Cn = C(n). Similarly, small generic constants are denoted by ǫ0, and they are bounded away from zero, provided their arguments are bounded away from zero.

3. Proof of almost umbilicity estimates for level sets In this section we prove Theorem 1.1. Proof. Let M be a connected component of ∂Ω and define 1 ǫ = ˚¯ 2f n+1. min ¯ f n+1 ˆ |∇ | M |∇ | Ω Note that ǫ is by definition scale invariant under both scalings Ω λΩ, f λf. 7→ 7→ In nature, this theorem is really a stability result, meaning that for the case of large ǫ there is nothing to prove. Indeed, using a simple case distinction invoking [Top08], one can see that it suffices to prove the following: 16 J. SCHEUER

There exist constants α = α(n), and C,ǫ0 > 0 bounded from above 1 resp. below in dependence of upper bounds for ψ ,Ω, M n ¯ ψ ,Ω and ∞ ∞ 1 1 2 k k | | k∇ k (minM ¯ f )− M n ¯ f ,Ω, such that ǫ ǫ0 implies |∇ | | | k∇ k∞ ≤ 1 α dist(M,S) C M n ǫ ≤ | | for a suitable totally umbilic hypersurface S. Hence in the following we prove the latter statement. Since M is a C2- level set of f we can relate the Hessian of f to the second fundamental form of M. A similar calculation appeared in [CM14]: Let (xi)1 i n be local coordinate vectors for M and without loss of generality suppose≤ ≤ that the outward pointing unit normal ν is ¯ f ν = ∇ , − ¯ f |∇ | otherwise treat f instead of f. Differentiating f along these coordinates we get − 0= ¯ f(x ), 0= ¯ 2f(x ,x )+ ¯ f(x )= ¯ 2f(x ,x ) ¯ f(ν)h , ∇ i ∇ i j ∇ ij ∇ i j − ∇ ij and hence ¯ 2f(x ,x )= ¯ f h . ∇ i j −|∇ | ij We obtain ¯ f ˚h = ¯ f h 1 Hg −|∇ | ij −|∇ | ij − n ij ¯ 2 1 kl ¯ 2 = f(xi,xj) g  f(xk,xl)gij ∇ − n ∇ = ˚¯ 2f(x ,x )+ 1 ∆¯ fg¯(x ,x ) ∇ i j n+1 i j 1 gkl ˚¯ 2f(x ,x )+ 1 ∆¯ fg¯(x ,x ) g − n ∇ k l n+1 k l ij   = ˚¯ 2f(x ,x )+ 1 ∆¯ fg¯(x ,x ) 1 gkl ˚¯ 2f(x ,x )g 1 ∆¯ fg ∇ i j n+1 i j − n ∇ k l ij − n+1 ij = ˚¯ 2f(x ,x )+ 1 ˚¯ 2f(ν,ν)g , ∇ i j n ∇ ij where we used gklxαxβ =g ¯αβ νανβ. k l − In turn we get the pointwise relation

(3.1) ¯ f 2 A˚ 2 = ˚¯ 2f(x ,x )˚¯ 2f(x ,x )gikgjl 1 (˚¯ 2f(ν,ν))2. |∇ | | | ∇ i j ∇ k l − n ∇ Now we construct a one-sided open neighbourhood Ω of M which is diffeomorphic to U ⊂ 1 Φ− ( )= I M, U × where I = f( ) and on which U ¯ 1 ¯ 1 f min f , Cn− Mt M Cn Mt , |∇ |≥ 2 M |∇ | | | ≤ | |≤ | | STABILITY FROM RIGIDITY VIA UMBILICITY 17 where Mt = Φ(t, M). The diffeomorphism Φ and are constructed as follows: For ξ M solve the initial value problem U ∈ ¯ f(Φ(t,ξ)) ∂ Φ(t,ξ)= ∇ t ¯ f(Φ(t,ξ)) 2 |∇ | Φ(0,ξ)= ξ on the largest open time interval Iξ = (0,tξ), such that ¯ f(Φ(t,ξ)) 1 min ¯ f |∇ |≥ 2 M |∇ | see [Ger06a, Thm. 12.6.2]. For all ξ M and 0 t

minM ¯ f T (3.2) √ǫ |∇ | = n+2 1 2 n 1 ≤ 2 M n ¯ f ,Ω 2 M n minM ¯ f | | k∇ k∞ | | |∇ | and define n 1 t := 2 M n min ¯ f √ǫ T. | | M |∇ | ≤ Hence we may integrate (3.1) and use the co-area formula [EG15, Thm. 3.13]: t t ˚¯ 2 n+1 ¯ n ˚ n+1 f ˚¯ 2 n+1 f A ds |∇ ¯ | ds = f . ˆ0 ˆMs |∇ | | | ≤ ˆ0 ˆMs f ˆ |∇ | |∇ | U Hence t 2 n(min ¯ f ) 1 M A˚ n+1 ds ǫ − − s n+1,Ms M |∇ | ˆ0 | |k k ≤ and thus there exists s [0,t], such that ∈ n+1 n ǫ 1 Ms A˚ 2 min ¯ f = M − n √ǫ. | |k kn+1,Ms ≤ M |∇ | t | | We want to apply [RS18, Thm. 3.1], which requires comparison of A˚ with the mean curvature of the hypersurface viewed in flat Euclidean space. Let Hˆ denote the mean curvature of a hypersurface viewed as a hypersurface in the Euclidean space. We use the Michael-Simon inequality [MS73] to estimate the L1-norm of Hˆ from below: n+1 n+1 1 n+1 (3.3) Ms − n cn,ψ Hˆ dµˆ cn,ψ Hˆ | | ≤ M ˆ | | ≤ k kn+1,Ms | s| Ms  to conclude 1 1 n+1 M − n Ms − n+1 n+1 (3.4) A˚ | | | | Hˆ √ǫ cn,ψ Hˆ √ǫ. k kn+1,Ms ≤ Hˆ n+1 k kn+1,Ms ≤ k kn+1,Ms k kn+1,Ms Now [RS18, Thm. 3.1] says that there exist constants C andǫ ˜0 depending 1 n ˆ on n, Ms A n+1,Ms and ψ ,Ω, such that | | k k k k∞ n+1 n+1 (3.5) A˚ Hˆ ˜ǫ0 k kn+1,Ms ≤ k kn+1,Ms STABILITY FROM RIGIDITY VIA UMBILICITY 19 implies CR γ dist(Ms,SR) A˚ , ≤ Hˆ γ k kn+1,Ms k kn+1,Ms where SR is a Euclidean sphere and γ = γ(n) a constant. Hence by choosing the size of ǫ in (3.4) we have to ensure that (3.5) holds. However, we also have to ensure that this does not introduce dependencies which are not allowed for ǫ. Therefore we have to estimate the term involving Aˆ in the arguments ofǫ ˜0. Using eψA = Aˆ + eψdψ(ν) id, [Ger06b, Prop. 1.1.11], we see that

n+1 1 (n+1)ψ n+1 Aˆ = Ms − e A dψ(ν) id dµˆ n+1,Ms ˆ k k | | Ms | − | n+1 n+1 (3.6) cn,ψ A + ¯ ψ ≤ k kn+1,Ms k∇ kn+1,Ms   (n+1) 2 n+1 n+1 cn,ψ (min ¯ f )− ¯ f + ¯ ψ . ≤ M |∇ | k∇ k ,Ω k∇ kn+1,Ms  ∞  Also invoking (3.2), we see that the required assumption is ǫ˜ min ¯ f √ǫ min 0 , M |∇ | . n+2 1 2 ≤ cn,ψ 2 M n ¯ f ,Ω ! | | k∇ k∞ Given this choice of ǫ0, we conclude that γ 1 γ dist(M ,S ) CRǫ 2(n+1) C M n ǫ 2(n+1) , s R ≤ ≤ | | where in the last estimate we used the isoperimetric inequality. It remains to transfer this estimate to M. Recall that n 1 s 2 M n min ¯ f √ǫ ≤ | | M |∇ | and that every point ξ M is connected to a unique point in Ms by the curve Φ( ,ξ), where ∈ · ξ = Φ(0,ξ) M, Φ(s,ξ) M . ∈ ∈ s Along this curve, we estimate Φ by 1 (3.7) dist (Φ(s,ξ), Φ(0,ξ)) c M n √ǫ g¯ ≤ n| | and hence 1 α dist(M,S ) C M n ǫ , R ≤ | | for some α = α(n).  3.1. Remark. We used the bound on the second derivatives of f on two occasions. Once to estimate T from below. In this step it would be possible to get away with a W 2,p-bound for p>n+1 and the Kondrachov embedding theorem. The second occasion is (3.6). However, note that here we are actually using the bound only on the subset . This observation can become U 20 J. SCHEUER valuable in further applications. In particular compare the appendix in [CM17].

4. Proofs of the geometric stability results 4.1. Stability in the Heintze-Karcher inequality. The proof of Theo- rem 1.4 is built upon solving

∆¯ f + (n + 1)Kf =1 inΩ (4.1) f =0 on ∂Ω and applying Corollary 1.2. Therefore we need a lower bound on ¯ f along ∂Ω. For this purpose we prove the following quantified Hopf|∇ bou|ndary lemma in some greater generality for further applications later on.

4.1. Lemma. Let Ω ⋐ MK , K = 0, 1, 1, be an open set satisfying a uniform interior sphere condition with radius −ρ. Let f satisfy ∆¯ f + af b in Ω ≥ f 0 on ∂Ω, ≤ where b> 0 and a C0(Ω)¯ satisfies ∈ a (n + 1) max(0, K). ≤ Then f < 0 in Ω and for all x ∂Ω with f(x ) = 0 we have 0 ∈ 0 1 1 ¯ f(x0) ǫ0(n− , b, ρ, (max a)− ) > 0. |∇ |≥ Ω¯ Proof. The proof follows well known comparison strategies, e.g. as in [GT01, Lemma 3.4] in conjunction with the choice of a good test function. (i) First we prove that f < 0 in Ω. For K 0 this follows from the strong maximum principle applied to f. If K = 1, we≤ have

∆¯ ϑ′ = (n + 1)ϑ′. − By assumption Ω¯ lies in the northern open hemisphere and hence ϑ′ > 0. The function z = f/ϑ′ solves ¯ ¯ ∆f f ¯ 2 ¯ ¯ ∆z = 2 ∆ϑ′ z, ϑ′ ϑ′ − ϑ′ − ϑ′ ∇ ∇ b (n + 1)f f 2 − + (n + 1) ¯ z, ¯ ϑ′ ≥ ϑ′ ϑ′ − ϑ′ ∇ ∇ b 2 = ¯ z, ¯ ϑ′ . ϑ′ − ϑ′ ∇ ∇ Since z 0 on M, z and hence f are both negative in Ω due to the strong maximum≤ principle. (ii) Let B = Bρ(y0) Ω, 0 <ρ<π/4, be an interior ball touching at x ∂Ω. By an ambient⊂ isometry we may assume that our polar coordinates 0 ∈ are centred at y0 and in these coordinates the metric is given by g¯ = dr2 + ϑ2(r)σ. STABILITY FROM RIGIDITY VIA UMBILICITY 21

Let Θ be a primitive of ϑ vanishing on ∂B, i.e. r ρ Θ(r)= ϑ(s) ds ϑ(s) ds. ˆ0 − ˆ0 Θ satisfies ∆Θ=(¯ n + 1)ϑ′. For δ > 0 the function w = f δΘ − satisfies ∆¯ w = ∆¯ f δ(n + 1)ϑ′ b af δ(n + 1)ϑ′ − ≥ − − = b aw δaΘ δ(n + 1)ϑ′ − − − b aw, ≥ 2 − 1 1 if δ = δ(n− , b, (maxΩ¯ a)− ) is chosen suitably small. As w = f 0 on ∂B, we may apply part (i) of this proof to w and obtain w 0 in B≤. Hence ≤ f B δΘ B. | ≤ | Now define CΘ(r) v(x)= e− 1, − where r = r(x). Then in the region ρ/2 < r < ρ we can estimate CΘ 2 CΘ 2 ∆¯ v = Ce− ∆Θ¯ + C e− ¯ Θ − |∇ | CΘ 2 2 = e− ( C(n + 1)ϑ′ + C ¯ Θ ) − |∇ | CΘ 2 2 ρ e− ( C(n + 1)ϑ′ + C ϑ ( )) ≥ − 2 0, ≥ 1 provided C = C(n, ρ− ) is chosen large enough. Since ρ

f ∂Bρ/2 δ ϑ(s) ds, | ≤− ˆρ/2 1 there exists a constant ǫ = ǫ(δ, ρ, b, (max a)− ), such that b f ∂Bρ/2 + ǫv ∂Bρ/2 0 and ǫav B 2 . | | ≤ | ≥− Hence we have ∆(¯ f + ǫv)+ a(f + ǫv) b + ǫ∆¯ v + ǫav b . ≥ ≥ 2 Step (i) applied to the domain B¯ B gives ρ\ ρ/2 f + ǫv 0 in B¯ B . ≤ ρ\ ρ/2 Taking the outward pointing derivative at x0 we get CΘ(ρ) 0 ∂ f + ǫ∂ v = ∂ f ǫCe− ϑ(ρ) ≤ ν ν ν − and ∂ f ǫCϑ(ρ). ν ≥  22 J. SCHEUER

Proof of Theorem 1.4. Write M = ∂Ω and define

ϑ′ M H1 δ = ´ 1. M u − ´ We need the following Reilly-type formula due to Qiu/Xia [QX15, equ. (16)]. There holds for a arbitrary C2(Ω)-function¯ f, recall Section 2 for notation,

2 2 2 ϑ′(∆¯ f + (n + 1)Kf) ϑ′ ¯ f + Kfg¯ ˆΩ − ˆΩ |∇ | 2 (4.2) = ϑ′(2∂ν f∆f + nH1(∂ν f) + h( f, f) + 2nKf∂νf) ˆM ∇ ∇ 2 2 + ¯ νϑ′( f nKf ). ˆM ∇ |∇ | − Let f be a solution to (4.1). With the help of

∆¯ ϑ′ = (n + 1)Kϑ′, − partial integration and the divergence theorem we deduce

ϑ′ = ϑ′∂ν f. ˆΩ ˆM Hence from the H¨older inequality we get

2 ϑ′ 2 ϑ′ ϑ′H (∂ f) ˆ ≤ ˆ H ˆ 1 ν  Ω  M 1 M ϑ′ 1 1 ˚2 2 = ϑ′ ϑ′ ¯ f ˆ H n + 1 ˆ − n ˆ |∇ | M 1  Ω Ω  and thus 2 Ω ϑ′ ϑ′ ´ = (1+ δ) u = (1+ δ)(n + 1) ϑ′. 1 1 ˚¯ 2 2 ≤ ˆM H1 ˆM ˆΩ n+1 Ω ϑ′ n Ω ϑ′ f ´ − ´ |∇ | Rearranging gives

˚2 2 δn n ϑ′ ϑ′ ¯ f ϑ′ = u . ˆ |∇ | ≤ (1 + δ)(n + 1) ˆ (1 + δ)(n + 1)2 ˆ H − Ω Ω M  1  Schauder theory [GT01] implies that the C2-norm of f is controlled by the C2,α-regularity of M and we note that M can be covered by a controlled number of charts, see [Per11, Ch. 1]. Plugging this into Corollary 1.2 gives the result. 

4.2. Almost constant curvature functions. We start with the proof of Theorem 1.5, which is related to the proof of Theorem 1.4. STABILITY FROM RIGIDITY VIA UMBILICITY 23

Proof of Theorem 1.5. Let f be a solution to (4.1) and then use f in the generalised Reilly formula (4.2). We obtain

˚2 2 n 2 ϑ′ ¯ f = ϑ′ n ϑ′H1(∂νf) ˆΩ |∇ | n + 1 ˆΩ − ˆM 2 n 2 = ϑ′ n ϑ′H (∂ f) (n + 1) ϑ ˆ − ˆ 1 ν Ω ′  Ω  M ´ 2 n 2 = ϑ′∂ f n ϑ′H (∂ f) (n + 1) ϑ ˆ ν − ˆ 1 ν Ω ′  M  M ´ n M ϑ′ 2 2 ´ ϑ′(∂ν f) n ϑ′H1(∂ν f) ≤ (n + 1) Ω ϑ′ ˆM − ˆM ´ 2 = ϑ′( H1)(∂ν f) , ˆM H− where n M ϑ′ = ´ . H (n + 1) Ω ϑ′ The C2-norm of f as well as the gradient´ are controlled from above and below in the same way as in the proof of Theorem 1.4. Hence the result follows from Theorem 1.1.  Rigidity results involving curvature functionals are often achieved by ap- plying Newton’s inequality for the normalized elementary symmetric poly- nomials of the principal curvatures 2 Hk+1Hk 1 Hk , − ≤ n [New07], also see [HLP34] for the case of arbitrary κ R . If κ Γk, the case of equality at a point x M occurs precisely if the∈ point x is∈ umbilic, i.e. all principal curvatures∈ are the same. In order to extract a stability result via the method of umbilicity, we need to control the error term in Newton’s inequality and relate it to the traceless second fundamental form. The following lemma provides such control:

4.2. Lemma. Let n 2, (κi)1 i n Γk with ≥ ≤ ≤ ∈ κ κ . 1 ≤···≤ n Let 1 H = κ ...κ , 1 k n, k n i1 ik ≤ ≤ k 1 i1<

Proof. From a simple calculation already done in [CGLS21] we obtain 2 2 2 ˚ 2 2 Hk Hk+1Hk 1 cn (κi κj) Hk+1,ij cn A Hk+1,n1. − − ≥ − ≥ | | Xi,j Denote by σ the elementary symmetric polynomial. Since κ Γ , we have k ∈ k (4.3) 0 <σk = κnσk,n + σk+1,n = κnκ1σk,n1 + (κn + κ1)σk+1,n1 + σk+2,n1 and

(4.4) 0 <σk,1 = κnσk,n1 + σk+1,n1. Now we suppose by contradiction, that at some κ Γ we had ∈ k σk+1,n1(κ) = 0. First of all, this is only possible in case n 3 and the smallest eigenvalue ≥ is κ1 0. Write λ = (κ2,...,κn 1) and let Sm be the m-th elementary polynomial≤ in λ. Then (4.3) implies−

0 < κnκ1Sk 2 + Sk − and from (4.4) we obtain Sk 2 > 0, which gives k 2 and Sk > 0. In case k = n 1 this is a contradiction.− For k n 2, we≥ obtain the contradiction − ≤ − 2 0

STABILITY FROM RIGIDITY VIA UMBILICITY 25

1 where C = C(n, A ,M , M , (minM Hk+1,n1)− , max r) and where (3.3) k k∞ | | was used with Ms replaced by M. Here Hˆ is the mean curvature of M when considered in the Euclidean space. The proof is complete after invoking [RS18, Thm. 3.1] and (3.3) again.  4.3. Stability in non-convex Alexandroff-Fenchel inequalities. In or- der to prove Theorem 1.9, we use a particular case of the inverse curvature flows studied by Gerhardt [Ger90] and Urbas [Urb90]. Their result goes as follows. Given a smooth closed and starshaped hypersurface M Rn+1, ⊂ such that the principal curvatures satisfy κ Γk, there exists a time- dependent family of starshaped hypersurfaces, parametriz∈ ed over a sphere, x˜ : [0, ) Sn Rn+1, ∞ × → such that the flow equation

Hk 1 ∂tx˜ = − ν Hk is satisfied and the rescaled hypersurfaces t x = e− x˜ smoothly converge to a sphere. The rescaled flow evolves, up to tangential reparametrisation, according to the flow equation

Hk 1 (4.5) ∂ x = − u ν. t H −  k  Guan/Li [GL09] have proved the Alexandroff-Fenchel inequalities for M with the help of Gerhardt’s and Urbas’ result, but they used a slightly different rescaling ofx ˜. However, the same result can be proved by using (4.5). A detailed account of the flow (4.5), which also provides an alternative argument for Gerhardt’s and Urbas’ result, is given in [Sch21]. There it is also shown that (4.5) keeps the quermassintegral Wk(Ωt) fixed and decreases Wk+1(Ωt). Here n ∂Ωt = x(t, S ) =: Mt. Since (4.5) converges to a round sphere, the proof of the Alexandroff-Fenchel inequalities is complete. In this section we make this proof quantitative and prove Theorem 1.9. Proof of Theorem 1.9. We may assume M is smooth. The result then fol- lows from approximation in C2. We are given a starshaped hypersurface M = ∂Ω Rn+1 with κ Γ , and define ǫ> 0 by ⊂ ∈ k n−k W (Ω) W (Ω) n−k+1 k+1 = k + ǫ. W (B) W (B) k+1  k  For 1 k n +1 define ≤ ≤ Wk(Ω) W˜ k(Ω) = . Wk(B) 26 J. SCHEUER

Let (Mt)0 t< be the solution to (4.5) with M0 = M and suppose without loss of generality≤ ∞ that

M ∂B in C∞, t . t → →∞ Then we use the variation formulae for the W˜ k,

d Hk 1 W˜ = c H − u , dt k+1 ˆ k+1 H − Mt  k  see for example [Sch21, Lemma 5.2]. Here c = cn,k is a constant. We obtain

∞ Hk+1Hk 1 ∞ Hk+1Hk 1 c − H = c − uH ˆ ˆ H − k ˆ ˆ H − k+1 0 Mt  k  0 Mt  k  d = ∞ W˜ (Ω ) dt ˆ k+1 t (4.6) 0 dt = 1 W˜ (Ω) − k+1 n−k = W˜ (Ω) n−k+1 W˜ (Ω) k − k+1 = ǫ. − Denote by r the function on the unit sphere, by which M is graphically parametrised. Further, write H F = k . Hk 1 − From the estimates in [Sch21, Lemma 4.6] we see that there exists a constant 1 C, which depends on osc r(0, ), (min u(0, ))− , 1/F (0, ) and A ,M such that · · · k k∞ 1 u C. F − ≤ Hence dist(M , M) Ct t 0. t ≤ ∀ ≥ From (4.6) we get

Hk+1Hk 1 min Hk − c√ǫ. s [0,√ǫ] ˆM − Hk ≤ ∈ s   Let M ǫ be the hypersurface where the minimum on the left is attained, in particular we obtain

ǫ Hk+1Hk 1 (4.7) dist(M , M) C√ǫ and Hk − C√ǫ. ≤ ˆ ǫ − H ≤ M  k  Then, using Lemma 4.2, 2 ˚ 2 1 C C max ǫ H H A ˚ n+1 ˚ 2 M k k+1,n1| | ǫ A ǫ A ǫ M ˆ ǫ | | ≤ M ˆ ǫ | | ≤ M min ǫ H ˆ ǫ H | | M | | M | | M k+1,n1 M k n+1 C H ǫ √ǫ. ≤ k kn+1,M The proof is complete after combining (4.7) with [RS18, Thm. 1.1].  STABILITY FROM RIGIDITY VIA UMBILICITY 27

4.4. Stability in Serrin’s problem with nonlinear source. The proof of Theorem 1.10 proceeds in the spirit of the integral approaches involv- ing so-called P -functions. The crucial ingredient is the following integral identity, which is a refined version of the ones in [BNST08a, MP20b]. 2 4.3. Lemma. Let Ω ⋐ M0 be an open set with C -boundary and suppose φ is a C2 function on an interval of the real line. Let f C2(Ω) C1(Ω)¯ satisfy ∈ ∩ ∆¯ f = φ(f) in Ω (4.8) f = 0 on M = ∂Ω. Then we have the equation 2 ˚2 2 1 2 2 R ( f) ¯ f = (∂ν f ∂νq) ¯ f R + (φ φ(0)) ˆΩ − |∇ | 2 ˆM − |∇ | − 2 ˆΩ − 3 n  φ(0) + (φ φ(0)) Φ fφ + (Φ fφ), ˆ − 2 − n + 1 2 ˆ − Ω   Ω where f 1 φ(0) Φ(f)= φ(s) ds, R = ∂ f, q(x)= x 2. ˆ M ˆ ν 2(n + 1)| | 0 | | M Proof. We write φ = φ f. Accounting for the nonlinearity, we define the P -function as ◦ 1 φ P = ¯ f 2 f. 2|∇ | − n + 1 Then 1 2 1 ∆¯ P = dφ( ¯ f)+ ¯ 2f 2 f∆¯ φ dφ( ¯ f) (∆¯ f)2 ∇ |∇ | − n + 1 − n + 1 ∇ − n + 1 n 1 1 = − dφ( ¯ f)+ ˚¯ 2f 2 f∆¯ φ. n + 1 ∇ |∇ | − n + 1 Thus, using integration by parts, n 1 1 ( f) ˚¯ 2f 2 = ( f)∆¯ P + − fdφ( ¯ f) f 2∆¯ φ ˆΩ − |∇ | ˆΩ − n + 1 ˆΩ ∇ − n + 1 ˆΩ 1 2 = ∂νf ¯ f φP + fdφ( ¯ f) 2 ˆM |∇ | − ˆΩ ˆΩ ∇ 1 2 2 = (∂ν f ∂νq) ¯ f R φP + fdφ( ¯ f) 2 ˆM − |∇ | − − ˆΩ ˆΩ ∇ 2  1 2 R + ¯ f ∂ν q + (φ φ(0)), 2 ˆΩ|∇ | 2 ˆΩ − where φ(0) q = x 2. 2(n + 1)| | From the Pohozaev identity, see [Str90, p. 156], we obtain 28 J. SCHEUER

˚2 2 1 2 2 ( f) ¯ f = (∂ν f ∂ν q) ¯ f R φP + fdφ( ¯ f) ˆΩ − |∇ | 2 ˆM − |∇ | − − ˆΩ ˆΩ ∇ n 1 φ(0)  R2 φ(0) Φ+ − fφ + (φ φ(0)) − ˆΩ n + 1 2 ˆΩ 2 ˆΩ − 1 2 2 3 n 2 = (∂ν f ∂ν q) ¯ f R + Φφ fφ 2 ˆM − |∇ | − 2 ˆΩ − n + 1 ˆΩ n 1 φ(0)  R2 φ(0) Φ+ − fφ + (φ φ(0)). − ˆΩ n + 1 2 ˆΩ 2 ˆΩ − Further rearranging gives 1 3 2 ( f) ˚¯ 2f 2 = (∂ f ∂ q) ¯ f 2 R2 + Φ φ φ(0) ˆ − |∇ | 2 ˆ ν − ν |∇ | − 2 ˆ − 3 Ω M Ω   n n 1  R2 fφ φ − φ(0) + (φ φ(0)) − n + 1 ˆ − 2n 2 ˆ − Ω   Ω 2 1 2 2 R = (∂ν f ∂ν q) ¯ f R + (φ φ(0)) 2 ˆM − |∇ | − 2 ˆΩ − 3 n  φ(0) + (φ φ(0)) Φ fφ + (Φ fφ). ˆ − 2 − n + 1 2 ˆ − Ω   Ω 

Proof of Theorem 1.10. Let f C2(Ω)¯ be a solution of (4.8). As φ > 0, ∈ Lemma 4.1 gives a lower bound on min∂Ω ¯ f in terms of min φ and ρ. As in the proof of Theorem 1.1, it suffices to| prove∇ | Theorem 1.10, when

1 ǫ = ( ¯ f R + φ φ(0) + Φ fφ ) 2 k|∇ |− k1,M k − k1,Ω k − k1,Ω is smaller than some ǫ0 > 0 that depends on the same quantities as C. Note that f < 0 in Ω due to the maximum principle. Hence, we fix some ǫ0, such that the level sets

M ǫ = f = ǫ , 0 <ǫ<ǫ0, − { − } are connected, non-empty, bound

Ω ǫ = f < ǫ − { − } and such that we have a uniform lower bound on ¯ f in Ω Ω ǫ. We recall that we are assuming a second derivative bound for|∇ that| purpo\ −se. We apply Corollary 1.2 to the function f and obtain a constant C, such that for some sphere S, α ˚2 2 dist(M ǫ,S) C ¯ f . − ≤ ˆ |∇ |  Ω−ǫ  STABILITY FROM RIGIDITY VIA UMBILICITY 29

Invoking that f < ǫ in Ω ǫ and Lemma 4.3, we estimate further: − − α C ˚2 2 dist(M ǫ,S) ( f) ¯ f − ≤ ǫα ˆ − |∇ |  Ω−ǫ  C α ( f) ˚¯ 2f 2 ≤ ǫα ˆ − |∇ |  Ω  C ¯ f R + φ φ(0) + Φ fφ α ≤ ǫα k|∇ |− k1,M k − k1,Ω k − k1,Ω = Cǫα.  As in the proof of Theorem 1.1, in particular recall (3.7), we obtain

dist(∂Ω, M ǫ) Cǫ. − ≤ Hence we complete the proof by α dist(∂Ω,S) dist(∂Ω, M ǫ) + dist(M ǫ,S) Cǫ + Cǫ . ≤ − − ≤ 

4.5. Stability in a fourth order Steklov eigenvalue problem. The proof of Theorem 1.12 makes the rigidity case from [WX09] quantitative. Proof of Theorem 1.12. Let w be an eigenfunction to the proposed equation, i.e. ∆¯ 2w =0 inΩ w = ∆¯ w µ ∂ w =0 on M = ∂Ω. − 1 ν It is known that w > 0 in Ω as well as ∂ν w < 0 on M, [BGM06, Thm. 1]. Then ¯ 2 Ω(∆w) µ1 = ´ 2 M (∂νw) and plugging this into (4.2) with K´= 0, we get

n 2 n 2 2 ˚2 2 µ1 (∂νw) = ∆¯ w = nH1(∂ν w) + ¯ w . n + 1 ˆM n + 1 ˆΩ ˆM ˆΩ|∇ | This implies 

˚2 2 2 (n + 1) ¯ w = n (µ1 (n + 1)H1)(∂ν w) . ˆΩ|∇ | ˆM − Now the result follows from Corollary 1.2. 

Acknowledgments This work was made possible through a research scholarship JS received from the DFG and which was carried out at Columbia University in New York. JS would like to thank the DFG, Columbia University and especially Prof. Simon Brendle for their support. 30 J. SCHEUER

References [ABR99] Amandine Aftalion, J´erˆome Busca, and Wolfgang Reichel, Approximate radial symmetry for overdetermined boundary value problems, Adv. Differ. Equ. 4 (1999), no. 6, 907–932. [Ale37] Alexandr Alexandroff, Zur Theorie der gemischten Volumina von konvexen K¨orpern. II. Neue Ungleichungen zwischen den gemischten Volumina und ihre Anwendungen, Rec. Math. [Mat. Sbornik] N.S. 2 (1937), no. 6, 1205–1238. [Ale38] Alexandr Alexandroff, Zur Theorie der gemischten Volumina von konvexen K¨orpern. III. Die Erweiterung zweier Lehrs¨atze Minkowkis ¨uber die konvexen Polyeder auf die beliebigen konvexen K¨orper, Rec. Math. [Mat. Sbornik] N.S. 3 (1938), no. 1, 27–46. [Ale62] Alexandr Alexandroff, A characteristic property of spheres, Ann. Mat. Pura Appl. 58 (1962), no. 4, 303–315. [Arn93] Randolf Arnold, On the Aleksandrov-Fenchel inequality and the stability of the sphere, Mon. Math. 115 (1993), no. 1–2, 1–11. [BC97] Joao Barbosa and Antonio Colares, Stability of hypersurfaces with constant r-mean curvature, Ann. Glob. Anal. Geom. 15 (1997), no. 3, 277–297. [BDC84] Joao Barbosa and Manfredo Do Carmo, Stability of hypersurfaces with con- stant mean curvature, Math. Z. 185 (1984), no. 3, 339–353. [BDCE88] Joao Barbosa, Manfredo Do Carmo, and Jost-Hinrich Eschenburg, Stability of hypersurfaces of constant mean curvature in Riemannian manifolds, Math. Z. 197 (1988), no. 1, 123–138. [BGM06] Elvise Berchio, Filippo Gazzola, and Enzo Mitidieri, Positivity preserving property for a class of biharmonic elliptic problems, J. Differ. Equ. 229 (2006), no. 1, 1–23. [BNST08a] Barbara Brandolini, Carlo Nitsch, Paolo Salani, and Cristina Trombetti, On the stability of the Serrin problem, J. Differ. Equ. 245 (2008), no. 6, 1566– 1583. [BNST08b] Barbara Brandolini, Carlo Nitsch, Paolo Salani, and Cristina Trombetti, Serrin-type overdetermined problems: An alternative proof, Arch. Rat. Mech. Anal. 190 (2008), no. 2, 267–280. [Bre13] Simon Brendle, Constant mean curvature surfaces in warped product mani- folds, Publ. Math. de l’IHES 117 (2013), no. 1, 247–269. [CGLS21] Chuanqiang Chen, Pengfei Guan, Junfang Li, and Julian Scheuer, A fully-nonlinear flow and quermassintegral inequalities in the sphere, arxiv:2102.07393, 2021. [Cho02] Jaigyoung Choe, Sufficient conditions for constant mean curvature surfaces to be round, Math. Ann. 323 (2002), no. 1, 143–156. [CJ15] Xu Cheng and Vazquez Juarez, Optimal constants of L2 inequalities for closed nearly umbilical hypersurfaces in space forms, Geom. Dedic. 177 (2015), no. 1, 189–211. [CM14] Tobias Colding and , Ricci curvature and monotonicity for harmonic functions, Calc. Var. Partial Differ. Equ. 49 (2014), no. 3–4, 1045– 1059. [CM17] Giulio Ciraolo and Francesco Maggi, On the shape of compact hypersurfaces with almost-constant mean curvature, Commun. Pure Appl. Math. 70 (2017), no. 4, 665–716. [CMV16] Giulio Ciraolo, Rolando Magnanini, and Vincenzo Vespri, H¨older stability for Serrin’s overdetermined problem, Ann. Mat. Pura Appl. 195 (2016), no. 4, 1333–1345. STABILITY FROM RIGIDITY VIA UMBILICITY 31

[CRV18] Giulio Ciraolo, Alberto Roncoroni, and Luigi Vezzoni, Quantitative stability for hypersurfaces with almost constant curvature in space forms, to appear in Ann. Mat. Pura Appl., arxiv:1812.00775, 2018. [CV18] Giulio Ciraolo and Luigi Vezzoni, A sharp quantitative version of Alexandrov’s theorem via the method of moving planes, J. Eur. Math. Soc. 20 (2018), no. 2, 261–299. [CV19] Giulio Ciraolo and Luigi Vezzoni, On Serrin’s overdetermined problem in space forms, Manuscr. Math. 159 (2019), no. 3–4, 445–452. [CV20] Giulio Ciraolo and Luigi Vezzoni, Quantitative stability for hypersurfaces with almost constant mean curvature in the hyperbolic space, Indiana Univ. Math. J. 69 (2020), no. 4, 1105–1153. [CZ14] Xu Cheng and Detang Zhou, Rigidity for closed totally umbilical hypersurfaces in space forms, J. Geom. Anal. 24 (2014), no. 3, 1–9. [Del41] Charles Delaunay, Sur la surface de revolution dont la courbaure moyenne est constante, J. Math. Pures Appl. 6 (1841), 309–314. [DLM05] Camillo De Lellis and Stefan M¨uller, Optimal rigidity estimates for nearly umbilical surfaces, J. Differ. Geom. 69 (2005), 75–110. [DLM06] Camillo De Lellis and Stefan M¨uller, A C0-estimate for nearly umbilical sur- faces, Calc. Var. Partial Differ. Equ. 26 (2006), no. 3, 283–296. [DM19] Matias Delgadino and Francesco Maggi, Alexandrov’s theorem revisited, Anal. & PDE 12 (2019), no. 6, 1613–1642. [EG15] Lawrence C. Evans and Ronald F. Gariepy, Measure theory and fine properties of functions, revised ed., Studies in advanced mathematics, CRC Press, Boca Raton, FL, 2015. [Fel18] William Feldman, Stability of Serrin’s problem and dynamic stability of a model for contact angle motion, SIAM J. Math. Anal. 50 (2018), no. 3, 3303– 3326. [FMP08] Nicola Fusco, Francesco Maggi, and Aldo Pratelli, The sharp quantitative isoperimetric inequality, Ann. Math. 168 (2008), no. 3, 941–980. [FMP10] Alessio Figalli, Francesco Maggi, and Aldo Pratelli, A mass transportation approach to quantitative isoperimetric inequalities, Invent. Math. 182 (2010), no. 1, 167–211. [FN18] Francisco Fontenele and Roberto N´u˜nes, A characterization of round spheres in space forms, Pac. J. Math. 297 (2018), no. 1, 67–78. [FS11] Ailana Fraser and Richard Schoen, The first Steklov eigenvalue, conformal geometry, and minimal surfaces, Adv. Math. 226 (2011), no. 5, 4011–4030. [FS16] Ailana Fraser and Richard Schoen, Sharp eigenvalue bounds and minimal surfaces in the ball, Invent. Math. 203 (2016), no. 3, 823–890. [Ger90] Claus Gerhardt, Flow of nonconvex hypersurfaces into spheres, J. Differ. Geom. 32 (1990), no. 1, 299–314. [Ger06a] Claus Gerhardt, Analysis II, International Series in Analysis, International Press of Boston Inc., 2006. [Ger06b] Claus Gerhardt, Curvature problems, Series in Geometry and Topology, vol. 39, International Press of Boston Inc., Sommerville, 2006. [Ger07] Claus Gerhardt, Curvature flows in semi-Riemannian manifolds, Geometric flows (Huai-Dong Cao and Shing-Tung Yau, eds.), Surveys in Differential Geometry, vol. 12, International Press of Boston Inc., 2007, pp. 113–166. [GL09] Pengfei Guan and Junfang Li, The quermassintegral inequalities for k-convex starshaped domains, Adv. Math. 221 (2009), no. 5, 1725–1732. [GP17] Alexandre Girouard and Iosif Polterovich, Spectral geometry of the Steklov problem, Shape optimization and spectral theory (Warsaw-) (Antoine Henrot, ed.), De Gruyter Open, 2017, pp. 120–148. 32 J. SCHEUER

[GS91] Helmut Groemer and Rolf Schneider, Stability estimates fo some geometric inequalities, Bull. London Math. Soc. 23 (1991), 67–74. [GT01] David Gilbarg and Neil Trudinger, Elliptic partial differential equations of sec- ond order, reprint of the 1998 edition ed., Classics in Mathematics, Springer, 2001. [HK78] Ernst Heintze and Hermann Karcher, A general comparison theorem with applications to volume estimates for submanifolds, Ann. Sci. Ec.´ Norm. Sup´er. (4) 11 (1978), no. 4, 451–470. [HLP34] Godfrey Hardy, John Littlewood, and George Polya, Inequalities, Cambridge University Press, Cambridge, 1934. [Hop89] Heinz Hopf, Differential geometry in the large, Lecture notes in mathematics, vol. 1000, Springer-Verlag, Berlin-Heidelberg, 1989. [Hsi54] Chuan-Chih Hsiung, Some integral formulas for closed hypersurfaces, Math. Scand. 2 (1954), 286–294. [Hsi56] Chuan-Chih Hsiung, Some integral formulas for closed hypersurfaces in Rie- mannian space, Pac. J. Math. 6 (1956), no. 2, 291–299. [HTY83] Wu-Yi Hsiang, Zhen-Huan Teng, and Wen-Ci Yu, New examples of constant mean curvature immersions of (2k − 1)-spheres into Euclidean 2k-space, Ann. Math. 117 (1983), no. 3, 609–625. [IM11] Said Ilias and Ola Makhoul, A Reilly inequality for the first Steklov eigenvalue, Differ. Geom. Appl. 29 (2011), no. 5, 699–708. [Iva14] Mohammad N. Ivaki, On the stability of the p-affine isoperimetric inequality, J. Geom. Anal. 24 (2014), no. 4, 1898–1911. [Iva15] Mohammad N. Ivaki, Stability of the Blaschke-Santal´oinequality in the plane, Mon. Math. 177 (2015), no. 3, 451–459. [Iva16] Mohammad N. Ivaki, The planar Busemann-Petty centroid inequality and its stability, Trans. Am. Math. Soc. 368 (2016), no. 5, 3539–3563. [Jel53] John Jellet, Sur la surface dont la courbure moyenne est constante, J. Math. Pures Appl. 18 (1853), 163–167. [KL01] Sung-Eun Koh and Seung-Won Lee, Addendum to the paper: Sphere theorem by mean of the ratio of mean curvature functions, Glasgow Math. J. 43 (2001), no. 2, 275–276. [KLP18] Kwok-Kun Kwong, Hoojoo Lee, and Juncheol Pyo, Weigthed Hsiung- Minkowski formulas and rigidity of umbilical hypersurfaces, Math. Res. Lett. 25 (2018), no. 2, 297–616. [Kor88] Nicholas Korevaar, Sphere theorems via Alexandrov for constant Weingarten curvature hypersurfacs - Appendix to a note of A. Ros, J. Differ. Geom. 27 (1988), no. 2, 221–223. [Kou71] Dimitri Koutroufiotis, Ovaloids which are almost spheres, Commun. Pure Appl. Math. 24 (1971), no. 3, 289–300. [KP98] Somas Kumaresan and Jyotshana Prajapat, Serrin’s result for hyperbolic space and sphere, Duke Math. J. 91 (1998), no. 1, 17–28. [KS20] Ernst Kuwert and Julian Scheuer, Asymptotic estimates for the Willmore flow with small energy, Int. Math. Res. Not. (2020), doi:10.1093/imrn/rnaa015/5766440. [KU01] Sung-Eun Koh and Teakwan Um, Almost spherical convex hypersurfaces in R4, Geom. Dedic. 88 (2001), no. 1–3, 67–80. [Lei99] Kurt Leichtweiß, Nearly umbilical ovaloids in the n-space are close to spheres, Result. Math. 36 (1999), no. 1-2, 102–109. [Lie00] Heinrich Liebmann, Ueber die Verbiegung der geschlossenen Fl¨achen positiver Kr¨ummung, Math. Ann. 53 (1900), no. 1–2, 81–112. [Mon99] Sebasti´an Montiel, Unicity of constant mean curvature hypersurfaces in some Riemannian manifolds, Indiana Univ. Math. J. 48 (1999), no. 2, 711–748. STABILITY FROM RIGIDITY VIA UMBILICITY 33

[Moo73] John Moore, Almost spherical convex hypersurfaces, Trans. Am. Math. Soc. 180 (1973), 347–358. [MP19] Rolando Magnanini and Giorgio Poggesi, On the stability for Alexandrov’s soap bubble theorem, J. Anal. Math. 139 (2019), no. 1, 179–205. [MP20a] Rolando Magnanini and Giorgio Poggesi, Nearly optimal stability for Serrin’s problem and the soap bubble theorem, Calc. Var. Partial Differ. Equ. 59 (2020), no. 1, 35. [MP20b] Rolando Magnanini and Giorgio Poggesi, Serrin’s problem and Alexandrov’s soap bubble theorem: Enhanced stability via integral identities, Indiana Univ. Math. J. 69 (2020), no. 4, 1181–1205. [MS73] James Michael and Leon Simon, Sobolev and mean-value inequalities on gen- eralized submanifolds of Rn, Commun. Pure Appl. Math. 26 (1973), no. 3, 361–379. [New07] Isaac Newton, Arithmetica universalis: sive de compositione et resolutione arithmetica liber; Cui accessit Halleiana aequationum radices arthmetice in- veniendi methodus, Cantabriga, Tooke, 1707. [Per11] Daniel Perez, On nearly umbilical hypersurfaces, Ph.D. thesis, Zuerich, 2011. [Pog73] Aleksei Pogorelov, Extrinsic geometry of convex surfaces, Translations of mathematical monographs, vol. 35, American Mathematical Society, 1973. [QX15] Guohuan Qiu and Chao Xia, A generalization of Reilly’s formula and its applications to a new Heintze-Karcher type inequality, Int. Math. Res. Not. 2015 (2015), no. 17, 7608–7619. [Rei70] Robert Reilly, Extrinsic rigidity theorems for compact submanifolds of the sphere, J. Differ. Geom. 4 (1970), no. 4, 487–497. [Rei80] Robert Reilly, Geometric applications of the solvability of Neumann problems on a Riemannian manifold, Arch. Rat. Mech. Anal. 75 (1980), no. 1, 23–29. [Rei82] Robert Reilly, Mean curvature, the Laplacian, and soap bubbles, Am. Math. Mon. 89 (1982), no. 3, 80–188 + 197–198. [Res68] Yurii Reshetnyak, Some estimates for almost umbilical surfaces, Sib. Math. J. 9 (1968), no. 4, 671–682. [Ron18] Alberto Roncoroni, A Serrin-type symmetry result on model manifolds: An extension of the Weinberger argument, C.R. Math. 356 (2018), no. 6, 648–656. [Ros87] Antonio Ros, Compact hypersurfaces with constant higher order mean curva- tures, Rev. Mat. Iberoam. 3 (1987), no. 3-4, 447–453. [Rot13] Julien Roth, A remark on almost umbilical hypersurfaces, Arch. Math. 49 (2013), no. 1, 1–7. [Rot15] Julien Roth, A new result about almost umbilical hypersurfaces of real space forms, Bull. Aust. Math. Soc. 91 (2015), no. 1, 145–154. [Rot20] Julien Roth, Reilly-type inequalities for Paneitz and Steklov eigenvalues, Po- tential Anal. 53 (2020), no. 3, 773–798. [RS18] Julien Roth and Julian Scheuer, Explicit rigidity of almost-umbilical hyper- surfaces, Asian J. Math. 22 (2018), no. 6, 1075–1088. [Sch89] Rolf Schneider, Stability in the Aleksandrov-Fenchel-Jessen theorem, Mathe- matika 36 (1989), no. 1, 50–59. [Sch14] Rolf Schneider, Convex bodies: The Brunn-Minkowski theory, 2. ed., Ency- clopedia of Mathematics and its Applications, no. 151, Cambridge University Press, 2014. [Sch15] Julian Scheuer, Quantitative oscillation estimates for almost-umbilical closed hypersurfaces in Euclidean space, Bull. Aust. Math. Soc. 92 (2015), no. 1, 133–144. [Sch21] Julian Scheuer, Extrinsic curvature flows and applications, 2019-20 Matrix Annals (David Wood, Jan De Gier, Cheryl Praeger, and Terence Tao, eds.), Matrix Book Series, vol. 4, Springer Nature, 2021, pp. 747–772. 34 J. SCHEUER

[Ser71] James Serrin, A symmetry problem in potential theory, Arch. Rat. Mech. Anal. 43 (1971), no. 4, 304–318. [Ste02] Vladimir Steklov, Sur les problemes fondamentaux de la physique mathema- tique (suite et fin), Ann. Sci. Ec.´ Norm. Sup´er. (3) 19 (1902), 191–259. [Ste13] Jakob Steiner, Jacob Steiner’s gesammelte Werke: Herausgegeben auf Veran- lassung der k¨oniglich preussischen Akademie der Wissenschaften, Cambridge Library Collection - Mathematics, vol. 2, Cambridge University Press, 2013. [Str90] Michael Struwe, Variational methods, Springer Berlin Heidelberg, 1990. [SX19] Julian Scheuer and Chao Xia, Locally constrained inverse curvature flows, Trans. Am. Math. Soc. 372 (2019), no. 10, 6771–6803. [Top08] Peter Topping, Relating diameter and mean curvature for submanifolds of Euclidean space, Comment. Math. Helv. 83 (2008), 539–546. [Urb90] John Urbas, On the expansion of starshaped hypersurfaces by symmetric func- tions of their principal curvatures, Math. Z. 205 (1990), no. 1, 355–372. [Vos56] Knud Voss, Einige differentialgeometrische Kongruenzs¨atze f¨ur geschlossene Fl¨achen und Hyperfl¨achen, Math. Ann. 131 (1956), no. 2, 180–218. [Wei71] Hans Weinberger, Remark on the preceding paper of Serrin, Arch. Rat. Mech. Anal. 43 (1971), no. 4, 319–320. [Wen86] Henry Wente, Counterexample to a conjecture of H. Hopf, Pac. J. Math. 121 (1986), no. 1, 193–243. [WX09] Qiaoling Wang and Changyu Xia, Sharp bounds for the first non-zero Stekloff eigenvalues, J. Funct. Anal. 257 (2009), no. 8, 2635–2644. [WX14] Jie Wu and Chao Xia, On rigidity of hypersurfaces with constant curvature functions in warped product manifolds, Ann. Glob. Anal. Geom. 46 (2014), no. 1, 1–22.

Cardiff University, School of Mathematics, Senghennydd Road, Cardiff CF24 4AG, Wales Email address: [email protected]