This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

scalar curvature rigidity of the four-dimensional sphere

Simone Cecchini Department of Mathematics, Texas A&M University [email protected] Jinmin Wang Department of Mathematics, Texas A&M University [email protected] Zhizhang Xie Department of Mathematics, Texas A&M University [email protected]  and  Bo Zhu Department of Mathematics, Texas A&M University [email protected]
Abstract.

Let (M,g)(M,g) be a four-dimensional closed connected oriented (possibly non-spin) Riemannian manifold with scalar curvature bounded below by n(n1)n(n-1). We prove that, if ff is a smooth distance non-increasing map of non-zero degree from (M,g)(M,g) to the unit four-sphere, then ff is an isometry. Following ideas of Gromov, we utilize μ\mu-bubbles and a version with coefficients of the rigidity of the three-sphere to rule out the case where all the inequalities are strict. Our proof of rigidity exploits monotonicity results for the harmonic map heat flow coupled with the Ricci flow due to Lee and Tam.

The second author is partially supported by NSF 1952693, 2247322 and NSFC 12171095.
The third author is partially supported by NSF 1952693 and 2247322.
The fourth author is partially supported by NSF 1952693 and 2247322.

1. Introduction

Extremality and rigidity properties of Riemannian manifolds with lower scalar curvature bounds have been the subject of intensive study in recent years. For a comprehensive overview of the subject, we refer to Gromov’s Four lectures on scalar curvature [gromov-four-lectures]. A cornerstone result in comparison geometry with scalar curvature is the rigidity of the round sphere in the spin setting, established by Llarull. Throughout this paper, we denote by g𝕊ng_{\mathbb{S}^{n}} the standard round metric on the nn-dimensional sphere 𝕊n\mathbb{S}^{n}.

Theorem 1.1 ([Llarull, Theorem B]).

Let (M,g)(M,g) be an nn-dimensional closed connected spin Riemannian manifold with Scgn(n1)\operatorname{Sc}_{g}\geq n(n-1). If f:(M,g)(𝕊n,g𝕊n)f\colon(M,g)\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}) is a smooth, distance non-increasing map of non-zero degree, then ff is an isometry.

This result illustrates the beautiful interplay between metric, curvature and topological information in scalar curvature geometry. Its proof relies on the Dirac operator method, requiring the hypothesis that MM is spin. A big open question in the field is whether the spin assumption can be dispensed with in Theorem 1.1. In this paper, we address this question affirmatively, at least in dimension four.

Theorem A.

Let (M,g)(M,g) be an four-dimensional closed connected oriented (possibly non-spin) Riemannian manifold with Scg12\operatorname{Sc}_{g}\geq 12. If f:(M,g)(𝕊n,g𝕊n)f\colon(M,g)\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}) is a smooth, distance non-increasing map of non-zero degree, then ff is an isometry.

Our approach intertwines various techniques from geometric analysis: minimal hypersurfaces, Ricci flow, and harmonic map heat flow. A key tool in our method is the utilization of μ\mu-bubbles, that are stable solutions to prescribed mean curvature problems. This technique, pioneered by Gromov [Gromov-macroscopic-dimension, Section 55/6\nicefrac{{5}}{{6}}], has been successfully used in addressing some challenging questions in scalar curvature geometry. Examples of its applications can be found in [li-polyhedron, chodosh2023generalized, gromov2020metrics, zhu_width_mu, chodosh-sufficiently-connected, cecchini-two-ends]. Drawing from Gromov’s ideas [gromov-four-lectures, Section 5], we utilize this technique to rule out the case where all the inequalities in A are strict. However, directly proving extremality and rigidity using μ\mu-bubbles poses a significant challenge. To overcome this difficulty, we exploit the monotonicity of the harmonic map heat flow coupled with the Ricci flow, recently established by Lee and Tam [LeeTam]. This approach enables us to reduce A to the situation where all the inequalities are strict, except when the metric is Einstein. Remarkably, Llarull’s rigidity theorem for Einstein manifolds follows from classical comparison geometry.

To further illustrate our strategy, let us compare it with Gromov’s perspective [gromov-four-lectures, Section 5.7]. We regard 𝕊n\mathbb{S}^{n} with two antipodal points removed as a warped product over 𝕊n1\mathbb{S}^{n-1}. From this viewpoint, Theorem 1.1 becomes a question about the scalar curvature rigidity of the degenerate spherical band

(𝕊n1×(π/2,π/2),cos2(t)g𝕊n1+dt2).\left(\mathbb{S}^{n-1}\times(-\pi/2,\pi/2),\cos^{2}(t)g_{\mathbb{S}^{n-1}}+dt^{2}\right). (1.1)

In the spin setting, this question has been recently addressed independently by Bär-Brendle-Hanke-Wang [Baer:2023aa] and by the second and third authors [wang2023scalar] utilizing the Dirac operator method. Gromov outlined an alternative approach to this problem employing μ\mu-bubbles. The crucial observation [gromov-four-lectures, Section 5.5] is that the variational formulas for μ\mu-bubbles, in this context, are related to the following stronger version of Theorem 1.1, due to Listing.

Theorem 1.2 ([Listing:2010te, Theorem 1]).

Let (M,g)(M,g) be an nn-dimensional closed connected spin Riemannian manifold. If f:(M,g)(𝕊n,g𝕊n)f\colon(M,g)\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}) is a smooth map of non-zero degree such that Scg(p)n(n1)dfp2\operatorname{Sc}_{g}(p)\geq n(n-1)\left\|df_{p}\right\|^{2} for any pMp\in M, then there exists a constant c>0c>0 such that f:(M,cg)(𝕊n,g𝕊n)f\colon(M,c\cdot g)\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}) is an isometry.

Here dfp\left\|df_{p}\right\| denotes the operator norm of the linear map dfpdf_{p}, see the discussion before Lemma 3.3. For the proof of this theorem, we also refer to [wang2023scalar, Theorem 3.3]. Since all three-dimensional oriented manifolds are spin, this observation enables us to use μ\mu-bubbles to study extremality and rigidity properties of spherical bands in dimension four without the spin condition.

Employing the outlined strategy to study the degenerate spherical band (1.1) poses two main challenges. Firstly, constructing suitable μ\mu-bubbles on open incomplete manifolds presents significant difficulties. While in dimension three this construction has been successfully carried out by Hu, Liu, and Shi [hu-3d-spherical] leveraging the extra control provided by the Gauss-Bonnet theorem, how to extend this approach to higher dimensions remains unclear. The authors intend to tackle this issue in future work. Secondly, it is implied in [gromov-four-lectures, Section 5.5] that the same construction is expected to carry over to degenerate bands over the torus, provided that the warping function is log-concave. However, [wang2023scalar, Example 4.1] shows that this condition is not sufficient for the rigidity of torical bands and, a fortiori, for the existence of a μ\mu-bubble with the desired properties. The authors also plan to investigate general conditions for the existence of μ\mu-bubbles in open incomplete bands in future work.

Instead of focusing on degenerate spherical bands, this paper adopts a different strategy to establish A, as outlined below.

  1. (1)

    We first rule out from A the case when all the inequalities are strict, following Gromov’s approach. This involves utilizing μ\mu-bubbles and Theorem 1.2 in dimension three. More precisely, we make use of these techniques to prove a comparison theorem with scalar and mean curvature bounds for compact spherical bands. Our comparison results are in the same spirit as [cecchini2022scalar, Raede23].

  2. (2)

    We then employ the harmonic map heat flow coupled with the Ricci flow to demonstrate that the general case of A reduces to the situation where all the inequalities are strict, unless the metric gg is Einstein with Ricg=3g\operatorname{Ric}_{g}=3g. Here, we make use of recent results of Lee and Tam [LeeTam], showing that the harmonic map heat flow coupled with the Ricci flow provides appropriate control of the Lipschitz constant with respect to the change of the scalar curvature under Ricci flow.

  3. (3)

    Finally, we prove A for Einstein manifolds, which follows as a consequence of Bishop’s volume comparison theorem.

This approach offers a novel perspective on proving A, circumventing the difficulties associated with degenerate spherical bands and exploiting powerful tools from geometric analysis.

The remainder of this paper is organized as follows. Section 2 reviews relevant results on the existence and properties of μ\mu-bubbles on compact Riemannian bands. In Section 3, we prove a comparison result for compact spherical bands with scalar and mean curvature bound. In Section 4, we use the harmonic map heat flow coupled with the Ricci flow to show that Theorem A can be reduced to the case where all inequalities become strict unless the metric gg is Einstein. Finally, in Section 5 we show that the comparison theorem from Section 3 suffices to rule out strict inequalities from A and that A holds for Einstein metrics. This completes the proof of A.

Acknowledgments.

We would like to thank Alessandro Carlotto, Man-Chun Lee, and Thomas Schick for helpful discussions.

2. μ\mu-bubbles with mean curvature bound

We review the properties of μ\mu-bubbles to be utilized in this paper. For more details, we refer the reader to the work of Gromov [gromov-four-lectures] and Zhu [zhu_width_mu].

Let us start by establishing notation and conventions. Let (X,g)(X,g) be an oriented Riemannian manifold. We stress that, throughout this paper, all manifolds are oriented and connected. We denote the Ricci curvature tensor by Ricg\operatorname{Ric}_{g}, and the scalar curvature by Scg\operatorname{Sc}_{g}. The Riemannian volume form is denoted by dVgdV_{g}. For an embedded hypersurface ZXZ\subset X, gZg_{Z} stands for the restriction of gg to ZZ. If the boundary X\partial X of XX is non-empty, Ag(X)\operatorname{A}_{g}(\partial X) denotes the second fundamental form with respect to the inward unit normal field ν\nu along X\partial X, and Hg(X)\operatorname{H}_{g}(\partial X) its trace, the mean curvature with respect to ν\nu. As per our convention, the boundary of the closed unit ball in n\operatorname{\mathbb{R}}^{n} has mean curvature equal to n1n-1.

A band is a compact connected manifold with boundary VV together with a decomposition V=V+V\partial V=\partial_{-}V\sqcup\partial_{+}V, where ±V\partial_{\pm}V are non-empty unions of components. A proper separating hypersurface for a band VV is a closed embedded hypersurface ΣVo\Sigma\subset{\kern 0.0ptV}^{\mathrm{o}} such that no connected component of VΣV\setminus\Sigma contains a path γ:[0,1]V\gamma\colon[0,1]\to V with γ(0)V\gamma(0)\in\partial_{-}V and γ(1)+V\gamma(1)\in\partial_{+}V. Note that a proper separating hypersurface Σ\Sigma for an orientable band VV is also orientable. Additionally, if (V,g)(V,g) is a Riemannian band and Σ\Sigma is a proper separating hypersurface for VV, we denote by (V^,g)(\hat{V},g) the Riemannian band isometrically embedded in (V,g)(V,g) such that V^=V\partial_{-}\hat{V}=\partial_{-}V and +V^=Σ\partial_{+}\hat{V}=\Sigma. We adopt the convention of denoting by Ag(Σ)\operatorname{A}_{g}(\Sigma) and Hg(Σ)\operatorname{H}_{g}(\Sigma) respectively the second fundamental form and the mean curvature on Σ\Sigma regarded as a union of components of V^\partial\hat{V}.

Let us now turn to μ\mu-bubbles. Let (V,g)(V,g) be an nn-dimensional Riemannian band. We denote by 𝒞(V)\mathcal{C}(V) the set of all Caccioppoli sets Ω^\hat{\Omega} in VV such that Ω^\hat{\Omega} contains an open neighborhood of V\partial_{-}V and Ω^+V=\hat{\Omega}\cap\partial_{+}V=\emptyset. For the notion of Caccioppoli sets and their properties, we refer the reader to [giusti-minimal-surfaces]. If Ω^\hat{\Omega} is a smooth Caccioppoli set in 𝒞(V)\mathcal{C}(V), then Ω^Vo\partial\hat{\Omega}\cap{\kern 0.0ptV}^{\mathrm{o}} is a separating hypersurface for VV. For any given smooth function μ\mu on VV, we consider the functional

𝒜μ(Ω^)=n1(Ω^Vo)Ω^μ𝑑n\mathcal{A}_{\mu}(\hat{\Omega})=\mathcal{H}^{n-1}(\partial^{\ast}\hat{\Omega}\cap{\kern 0.0ptV}^{\mathrm{o}})-\int_{\hat{\Omega}}\mu\,d\mathcal{H}^{n}

where k\mathcal{H}^{k} denotes the kk-dimensional Hausdorff measure on (V,g)(V,g), and Ω^\partial^{\ast}\hat{\Omega} denotes the reduced boundary of Ω^\hat{\Omega}. We say that a Caccioppoli set Ω𝒞(V)\Omega\in\mathcal{C}(V) is a μ\mu-bubble if it minimizes the functional 𝒜μ\mathcal{A}_{\mu}, that is, if Ω\Omega satisfies

𝒜μ(Ω)=inf{𝒜μ(Ω^):Ω^𝒞(V)}.\mathcal{A}_{\mu}(\Omega)=\inf\left\{\mathcal{A}_{\mu}(\hat{\Omega})\colon\hat{\Omega}\in\mathcal{C}(V)\right\}.

The next lemma states the existence of μ\mu-bubbles under strict mean curvature bound.

Lemma 2.1 ([zhu_width_mu, Section 2], [Raede23, Lemma 4.2]).

For n7n\leq 7, let (V,g)(V,g) be an nn-dimensional Riemannian band and μ:V\mu\colon V\to\operatorname{\mathbb{R}} a smooth function such that Hg(±V)>±μ|±V\operatorname{H}_{g}(\partial_{\pm}V)>\pm\mu|_{\partial_{\pm}V}. Then there exists a smooth μ\mu-bubble Ω\Omega.

In the next lemma we collect the properties of smooth μ\mu-bubbles that will be used in this paper.

Lemma 2.2 ([zhu_width_mu, Section 2],[Raede23, Lemmas 4.3 and 4.4]).

Let (V,g)(V,g) be an nn-dimensional Riemannian band, let μ:V\mu\colon V\to\operatorname{\mathbb{R}} be a smooth function, and let Ω\Omega be a smooth μ\mu-bubble. Then ΣΩVo\Sigma\coloneqq\partial\Omega\cap{\kern 0.0ptV}^{\mathrm{o}} is a proper separating hypersurface for VV satisfying

Σ(Hg(Σ)μ)u𝑑VgΣ=0\int_{\Sigma}\left(\operatorname{H}_{g}(\Sigma)-\mu\right)u\,dV_{g_{\Sigma}}=0 (2.1)

and

Σ|gΣu|2+12ScgΣu2dVgΣ12Σ(ScgHg2(Σ)+|Ag(Σ)|2+2Hg(Σ)μ+2g(gμ,ν))u2𝑑VgΣ\int_{\Sigma}\left|\nabla_{g_{\Sigma}}u\right|^{2}+\frac{1}{2}\operatorname{Sc}_{g_{\Sigma}}u^{2}\,dV_{g_{\Sigma}}\\ \geq\frac{1}{2}\int_{\Sigma}\left(\operatorname{Sc}_{g}-\operatorname{H}_{g}^{2}(\Sigma)+\left|\operatorname{A}_{g}(\Sigma)\right|^{2}+2\operatorname{H}_{g}(\Sigma)\mu+2g(\nabla_{g}\mu,\nu)\right)u^{2}\,dV_{g_{\Sigma}} (2.2)

for every uC(Σ)u\in C^{\infty}(\Sigma).

Remark 2.3.

Since the μ\mu-bubble Ω\Omega minimizes the functional 𝒜μ\mathcal{A}_{\mu}, Equation (2.1) follows from the first variation formula, whereas Inequality (2.2) follows from the second variation formula. See [zhu_width_mu, Section 2] and also  [Raede23, Lemmas 4.3 and 4.4].

3. scalar-mean bounds on spherical bands

We prove a comparison theorem for spherical bands with scalar and mean curvature bounds. Let I=[t,t+]I=[t_{-},t_{+}] be a compact interval, and let ϕ:I+\phi\colon I\to\operatorname{\mathbb{R}}_{+} be a smooth function. Consider the four-dimensional band 𝕊I3𝕊3×I\mathbb{S}^{3}_{I}\coloneqq\mathbb{S}^{3}\times I, equipped with the warped product metric

gϕ(p,t)ϕ2(t)g𝕊3(p)+dt2,(p,t)𝕊3×I=𝕊I3.g_{\phi}(p,t)\coloneqq\phi^{2}(t)g_{\mathbb{S}^{3}}(p)+dt^{2},\qquad\forall(p,t)\in\mathbb{S}^{3}\times I=\mathbb{S}^{3}_{I}.

Let hϕ:Ih_{\phi}\colon I\to\operatorname{\mathbb{R}} be the smooth function defined as

hϕ(t)3ddtlog(ϕ(t))=3ϕ(t)ϕ(t),tI.h_{\phi}(t)\coloneqq 3\frac{d}{dt}\log(\phi(t))=3\frac{\phi^{\prime}(t)}{\phi(t)},\qquad\forall t\in I. (3.1)

Utilizing the classical formula for the scalar curvature of warped product metrics [spingeometry, Ch.IV, Formula (6.15)],

Scgϕ(p,t)=6ϕ2(t)43hϕ2(t)2hϕ(t),(p,t)𝕊3×I.\operatorname{Sc}_{g_{\phi}}(p,t)=\frac{6}{\phi^{2}(t)}-\frac{4}{3}h_{\phi}^{2}(t)-2h_{\phi}^{\prime}(t),\qquad\forall(p,t)\in\mathbb{S}^{3}\times I. (3.2)

Additionally, ±𝕊I3=𝕊3×{t±}\partial_{\pm}\mathbb{S}^{3}_{I}=\mathbb{S}_{3}\times\{t_{\pm}\} and

Hgϕ(±𝕊I3)=±hϕ(t±).H_{g_{\phi}}(\partial_{\pm}\mathbb{S}^{3}_{I})=\pm h_{\phi}(t_{\pm}). (3.3)

We say that a positive smooth function ϕ(t)\phi(t) is log-concave if log(ϕ(t))\log(\phi(t)) is a concave function.

Proposition 3.1.

Let (V,g)(V,g) be a four-dimensional oriented Riemannian band. Let I=[t,t+]I=[t_{-},t_{+}] be a compact interval, and ϕ:I+\phi\colon I\to\operatorname{\mathbb{R}}_{+} be a smooth log-concave function. Let f:(V,g)(𝕊I3,gϕ)f\colon(V,g)\to(\mathbb{S}^{3}_{I},g_{\phi}) be a smooth map. Suppose

  1. (i)

    ff is distance non-increasing,

  2. (ii)

    f(±V)±𝕊I3f(\partial_{\pm}V)\subseteq\partial_{\pm}\mathbb{S}^{3}_{I},

  3. (iii)

    Scg>fScgϕ\operatorname{Sc}_{g}>f^{\ast}\operatorname{Sc}_{g_{\phi}},

  4. (iv)

    Hg(±V)>fHgϕ(±𝕊I3)\operatorname{H}_{g}(\partial_{\pm}V)>f^{\ast}\operatorname{H}_{g_{\phi}}(\partial_{\pm}\mathbb{S}^{3}_{I}).

Then ff has degree zero.

To prove this proposition, we will construct a separating hypersurface for VV with suitable properties forcing the degree of ff to be zero. Let us first introduce some additional notation. Let pr𝕊3:𝕊I3=𝕊3×I𝕊3\operatorname{pr}_{\mathbb{S}^{3}}\colon\mathbb{S}^{3}_{I}=\mathbb{S}^{3}\times I\to\mathbb{S}^{3} be the projection onto the first factor. Under the hypotheses of Proposition 3.1, let ZVZ\subset V be a closed hypersurface. Consider the smooth map

fZpr𝕊3fiZ:Z𝕊3,f_{Z}\coloneqq\operatorname{pr}_{\mathbb{S}^{3}}\circ f\circ i_{Z}\colon Z\to\mathbb{S}^{3}, (3.4)

where iZ:ZVi_{Z}\colon Z\hookrightarrow V denotes the inclusion of ZZ into VV. The next lemma contains well-known properties of the map fZf_{Z}, see [Raede23, Lemma 6.3]. We include a brief proof for clarity.

Lemma 3.2.

Let VV be a four-dimensional oriented band, let II be a compact interval, and let f:V𝕊I3f\colon V\to\mathbb{S}^{3}_{I} be a smooth map of non-zero degree such that f(±V)±𝕊I3f(\partial_{\pm}V)\subseteq\partial_{\pm}\mathbb{S}^{3}_{I}. Suppose Σ\Sigma is a proper separating hypersurface of VV. Then there exists a connected component YY of Σ\Sigma such that fY:Y𝕊3f_{Y}\colon Y\to\mathbb{S}^{3} has non-zero degree.

Proof.

Let α\alpha be a generator of H1(𝕊I3,𝕊I3;)=H^{1}(\mathbb{S}^{3}_{I},\partial\mathbb{S}^{3}_{I};\operatorname{\mathbb{Z}})=\operatorname{\mathbb{Z}}, and let [𝕊I3,𝕊I3]H4(𝕊I3,𝕊I3;)[\mathbb{S}^{3}_{I},\partial\mathbb{S}^{3}_{I}]\in H_{4}(\mathbb{S}^{3}_{I},\partial\mathbb{S}^{3}_{I};\operatorname{\mathbb{Z}}) be the fundamental class. Note that

(pr𝕊3)([𝕊I3,𝕊I3]α)=[𝕊3](\operatorname{pr}_{\mathbb{S}^{3}})_{\ast}([\mathbb{S}^{3}_{I},\partial\mathbb{S}^{3}_{I}]\frown\alpha)=[\mathbb{S}^{3}] (3.5)

as the Lefschetz dual of α\alpha is represented by any submanifold 𝕊3×{t}\mathbb{S}^{3}\times\{t\}, for tIt\in I. Since f(±V)±𝕊I3f(\partial_{\pm}V)\subseteq\partial_{\pm}\mathbb{S}^{3}_{I} and since Σ\Sigma is a proper separating hypersurface for VV, there exists a union of components of Σ\Sigma, that we denote by Σ\Sigma^{\prime}, such that [Σ][\Sigma^{\prime}] is Lefschetz dual to f(α)f^{\ast}(\alpha). Hence,

f([Σ])=f([V,V]f(α))=f([V,V])α=d[𝕊I3,𝕊I3]αf_{\ast}([\Sigma^{\prime}])=f_{\ast}([V,\partial V]\frown f^{\ast}(\alpha))=f_{\ast}([V,\partial V])\frown\alpha=d[\mathbb{S}^{3}_{I},\partial\mathbb{S}^{3}_{I}]\frown\alpha (3.6)

where d=deg(f)d=\deg(f) and [V,V]H4(V,V;)[V,\partial V]\in H_{4}(V,\partial V;\operatorname{\mathbb{Z}}) is the fundamental class. From (3.5) and (3.6), we deduce that (pr𝕊3f)([Σ])=d[𝕊3](\operatorname{pr}_{\mathbb{S}^{3}}\circ f)_{\ast}([\Sigma^{\prime}])=d[\mathbb{S}^{3}]. Therefore, if fYf_{Y} has degree zero for every component YY of Σ\Sigma, we must have d=0d=0. ∎

For a smooth map f:(X,g)(Y,h)f\colon(X,g)\to(Y,h) of Riemannian manifolds, we denote by dfp\left\|df_{p}\right\| the operator norm of the linear map dfp:(TpX,gp)(Tf(p)Y,hp)df_{p}\colon(T_{p}X,g_{p})\to(T_{f(p)}Y,h_{p}), that is, dfp\left\|df_{p}\right\| is the infimum of the numbers c>0c>0 such that hf(p)(dfpv,dfpv)1/2cgp(v,v)1/2h_{f(p)}(df_{p}v,df_{p}v)^{1/2}\leq cg_{p}(v,v)^{1/2}, for all vTpXv\in T_{p}X. We denote by df:X0\left\|df\right\|\colon X\to\operatorname{\mathbb{R}}_{\geq 0} the function whose value at a point pXp\in X is dfp\left\|df_{p}\right\|. Note that ff is distance non-increasing if and only if dfp1\left\|df_{p}\right\|\leq 1 for all pXp\in X.

Lemma 3.3.

Assume the same hypotheses as in Proposition 3.1. Then there exists a proper separating hypersurface Σ\Sigma for VV such that

Σ|gΣu|2+12(ScgΣ6dfΣ2)u2dVgΣ>0\int_{\Sigma}\left|\nabla_{g_{\Sigma}}u\right|^{2}+\frac{1}{2}\left(\operatorname{Sc}_{g_{\Sigma}}-6\left\|df_{\Sigma}\right\|^{2}\right)u^{2}\,dV_{g_{\Sigma}}>0 (3.7)

for all non-zero uC(Σ)u\in C^{\infty}(\Sigma).

Proof.

Let prI:𝕊I3=𝕊3×II\operatorname{pr}_{I}\colon\mathbb{S}^{3}_{I}=\mathbb{S}^{3}\times I\to I be the projection onto the second factor. Consider the smooth map

μϕhϕprIf:V,\mu_{\phi}\coloneqq h_{\phi}\circ\operatorname{pr}_{I}\circ f\colon V\to\operatorname{\mathbb{R}},

where hϕh_{\phi} is the function defined by (3.1). By Condition (iv) and Equation (3.3), Hg(±V)>fHgϕ(±𝕊I3)=±μϕ|±V\operatorname{H}_{g}(\partial_{\pm}V)>f^{\ast}\operatorname{H}_{g_{\phi}}(\partial_{\pm}\mathbb{S}^{3}_{I})=\pm\mu_{\phi}|_{\partial_{\pm}V}. By Lemma 2.1 and Lemma 2.2, there exists a smooth μϕ\mu_{\phi}-bubble Ω\Omega for VV, and ΣΩVo\Sigma\coloneqq\partial\Omega\cap{\kern 0.0ptV}^{\mathrm{o}} is a closed separating hypersurface for VV satisfying (2.1) and  (2.2). By (2.1), Hg(Σ)=μϕ\operatorname{H}_{g}(\Sigma)=\mu_{\phi}. Since |Ag(Σ)|2Hg2(Σ)n1\left|\operatorname{A}_{g}(\Sigma)\right|^{2}\geq\frac{\operatorname{H}_{g}^{2}(\Sigma)}{n-1},

Hg(Σ)2+|Ag(Σ)|2+2Hg(Σ)μϕμϕ2+μϕ2n1=nn1μϕ2.-\operatorname{H}_{g}(\Sigma)^{2}+\left|\operatorname{A}_{g}(\Sigma)\right|^{2}+2\operatorname{H}_{g}(\Sigma)\mu_{\phi}\geq\mu_{\phi}^{2}+\frac{\mu_{\phi}^{2}}{n-1}=\frac{n}{n-1}\mu_{\phi}^{2}.

From Inequality (2.2), we deduce that

Σ|gΣu|2+12ScgΣu2dVgΣ12Σ(Scg+43μϕ2+2g(gμϕ,ν))u2𝑑VgΣ\int_{\Sigma}\left|\nabla_{g_{\Sigma}}u\right|^{2}+\frac{1}{2}\operatorname{Sc}_{g_{\Sigma}}u^{2}\,dV_{g_{\Sigma}}\geq\frac{1}{2}\int_{\Sigma}\left(\operatorname{Sc}_{g}+\frac{4}{3}\mu_{\phi}^{2}+2g(\nabla_{g}\mu_{\phi},\nu)\right)u^{2}\,dV_{g_{\Sigma}} (3.8)

for all uC(Σ)u\in C^{\infty}(\Sigma). Let pΣp\in\Sigma. By Condition (i), |g(prIf)(p)|1|\nabla_{g}(\operatorname{pr}_{I}\circ f)(p)|\leq 1. Since ϕ\phi is log-concave, hϕ(prI(f(p)))0h_{\phi}^{\prime}(\operatorname{pr}_{I}(f(p)))\leq 0. Therefore,

hϕ(prI(f(p)))hϕ(prI(f(p)))|g(prIf)(p)|g(gμϕ(p),ν(p)).h_{\phi}^{\prime}(\operatorname{pr}_{I}(f(p)))\leq h_{\phi}^{\prime}(\operatorname{pr}_{I}(f(p)))\left|\nabla_{g}(\operatorname{pr}_{I}\circ f)(p)\right|\leq g(\nabla_{g}\mu_{\phi}(p),\nu(p)).

Thus,

43μϕ2(p)+2g(gμϕ(p),ν(p))\displaystyle\frac{4}{3}\mu_{\phi}^{2}(p)+2g(\nabla_{g}\mu_{\phi}(p),\nu(p))\geq 43hϕ2(prI(f(p)))+2hϕ(prI(f(p)))\displaystyle\frac{4}{3}h_{\phi}^{2}(\operatorname{pr}_{I}(f(p)))+2h_{\phi}^{\prime}(\operatorname{pr}_{I}(f(p))) (3.9)
=\displaystyle= 6ϕ2(prI(f(p)))Scgϕ(f(p))\displaystyle\frac{6}{\phi^{2}(\operatorname{pr}_{I}(f(p)))}-\operatorname{Sc}_{g_{\phi}}(f(p))

where in the last equality we used (3.2). Using the chain rule,

(dfΣ)p21ϕ2(prI(f(p))).\left\|(df_{\Sigma})_{p}\right\|^{2}\leq\frac{1}{\phi^{2}(\operatorname{pr}_{I}(f(p)))}. (3.10)

Using Condition (iii) and Inequalities (3.9) and (3.10), we deduce

Scg(p)+43μϕ2(p)+2g(gμϕ(p),ν(p))>6(dfΣ)p2,pΣ.\operatorname{Sc}_{g}(p)+\frac{4}{3}\mu_{\phi}^{2}(p)+2g\left(\nabla_{g}\mu_{\phi}(p),\nu(p)\right)>6\left\|(df_{\Sigma})_{p}\right\|^{2},\qquad\forall p\in\Sigma.

Finally, from the previous inequality and Inequality (3.8) we conclude that (3.7) holds for every non-zero uC(Σ)u\in C^{\infty}(\Sigma). ∎

Under the hypotheses of Proposition 3.1, let ZVZ\subset V be a closed hypersurface. Consider the second-order formally self-adjoint elliptic differential operator

gZΔgZ+18(ScgZ6dfZ2)\mathcal{L}_{g_{Z}}\coloneqq-\Delta_{g_{Z}}+\frac{1}{8}\left(\operatorname{Sc}_{g_{Z}}-6\left\|df_{Z}\right\|^{2}\right) (3.11)

where ΔgZ\Delta_{g_{Z}} denotes the Laplace-Beltrami operator of (Z,gZ)(Z,g_{Z}) and fZ:Z𝕊3f_{Z}\colon Z\to\mathbb{S}^{3} is defined by (3.4). Note that we adopt the sign convention for ΔgZ\Delta_{g_{Z}} such that

Z(Δgzu)u𝑑VgZ=Z|gZu|2𝑑VgZ,uC(Z).-\int_{Z}\left(\Delta_{g_{z}}u\right)u\,dV_{g_{Z}}=\int_{Z}\left|\nabla_{g_{Z}}u\right|^{2}\,dV_{g_{Z}},\qquad\forall u\in C^{\infty}(Z).

With this convention, ΔgZ-\Delta_{g_{Z}} has nonnegative spectrum. By classical elliptic theory, the spectrum of gZ\mathcal{L}_{g_{Z}} is a discrete set bounded from below consisting only of eigenvalues. Moreover, the eigenfunctions relative to each eigenvalue are smooth.

Lemma 3.4.

Suppose that

Z|gZu|2+12(ScgZ6dfZ2)u2dVgZ>0\int_{Z}\left|\nabla_{g_{Z}}u\right|^{2}+\frac{1}{2}\left(\operatorname{Sc}_{g_{Z}}-6\left\|df_{Z}\right\|^{2}\right)u^{2}\,dV_{g_{Z}}>0 (3.12)

for all non-zero uC(Z)u\in C^{\infty}(Z). Then the lowest eigenvalue of gZ\mathcal{L}_{g_{Z}} is positive.

Proof.

Let λ\lambda\in\operatorname{\mathbb{R}} be an eigenvalue of gZ\mathcal{L}_{g_{Z}}. We will show that λ>0\lambda>0. Let uC(Z)u\in C^{\infty}(Z) be an eigenfunction corresponding to λ\lambda, that is,

ΔgZu+λu=18(ScgZ6dfZ2)u\Delta_{g_{Z}}u+\lambda u=\frac{1}{8}\left(\operatorname{Sc}_{g_{Z}}-6\left\|df_{Z}\right\|^{2}\right)u

with u0u\not\equiv 0. We have

Z|gZu|2λu2dVgZ\displaystyle\int_{Z}\left|\nabla_{g_{Z}}u\right|^{2}-\lambda u^{2}\,dV_{g_{Z}} =Z(ΔgZu+λu)u𝑑VgZ\displaystyle=-\int_{Z}\left(\Delta_{g_{Z}}u+\lambda u\right)u\,dV_{g_{Z}}
=18Z(ScgZ6dfZ2)u2𝑑VgZ\displaystyle=-\frac{1}{8}\int_{Z}\left(\operatorname{Sc}_{g_{Z}}-6\left\|df_{Z}\right\|^{2}\right)u^{2}\,dV_{g_{Z}}
<14Z|gZu|2𝑑VgZ,\displaystyle<\frac{1}{4}\int_{Z}\left|\nabla_{g_{Z}}u\right|^{2}\,dV_{g_{Z}},

where in the last inequality we used (3.12). Therefore,

λZu2𝑑VgZ>34Z|gZu|2𝑑VgZ.\lambda\int_{Z}u^{2}\,dV_{g_{Z}}>\frac{3}{4}\int_{Z}\left|\nabla_{g_{Z}}u\right|^{2}\,dV_{g_{Z}}.

Since u0u\not\equiv 0, we conclude that λ>0\lambda>0. ∎

To prove Proposition 3.1, let us specialize Theorem 1.2 to the three-dimensional case. Since all three-dimensional oriented manifolds are spin, we obtain:

Proposition 3.5.

Let (M,g)(M,g) be a three-dimensional closed connected oriented Riemannian manifold. If f:(M,g)(𝕊3,g𝕊3)f\colon(M,g)\to(\mathbb{S}^{3},g_{\mathbb{S}_{3}}) is a smooth map of non-zero degree such that Scg(p)6dfp2\operatorname{Sc}_{g}(p)\geq 6\left\|df_{p}\right\|^{2} for all pMp\in M, then there exists a constant c>0c>0 such that f:(M,cg)(𝕊3,g𝕊3)f\colon(M,c\cdot g)\to(\mathbb{S}^{3},g_{\mathbb{S}^{3}}) is an isometry.

Remark 3.6.

Let (M,g)(M,g) be a three-dimensional closed connected oriented Riemannian manifold. Proposition 3.5 implies that there are no smooth maps f:(M,g)(𝕊3,g𝕊3)f\colon(M,g)\to(\mathbb{S}^{3},g_{\mathbb{S}_{3}}) of non-zero degree such that Scg(p)>6dfp2\operatorname{Sc}_{g}(p)>6\left\|df_{p}\right\|^{2} for all pMp\in M. In fact, we will only use this consequence in the proof of A.

We are now ready to prove Proposition 3.1. We proceed in a similar way as in [schoen-yau-psc-manifolds]. Under the hypotheses of Proposition 3.1, we use the spectral information from Lemma 3.4 to make a conformal change of the metric on (a suitable component of) the three-dimensional μ\mu-bubble in such a way that the new metric would contradict Proposition 3.5 if ff had non-zero degree.

Proof of Proposition 3.1.

Suppose, by contradiction, that ff has non-zero degree. Using Lemmas 3.2, 3.3 and 3.4, we choose a closed connected oriented hypersurface YY embedded in VV such that

  • \vartriangleright

    fY:Y𝕊3f_{Y}\colon Y\to\mathbb{S}^{3} has non-zero degree, where fYf_{Y} is the function defined by (3.4);

  • \vartriangleright

    the lowest eigenvalue λ\lambda of gY\mathcal{L}_{g_{Y}} is positive, where gYg_{Y} denotes the restriction of gg to YY, and where gY\mathcal{L}_{g_{Y}} is the operator defined by (3.11).

Let uu be an eigenfunction relative to λ\lambda, that is, 0uC(Y)0\not\equiv u\in C^{\infty}(Y) and satisfies

ΔgYu+18ScgYu=34dfY2u+λu.-\Delta_{g_{Y}}u+\frac{1}{8}\operatorname{Sc}_{g_{Y}}u=\frac{3}{4}\left\|df_{Y}\right\|^{2}u+\lambda u. (3.13)

By classical elliptic theory, uu doesn’t change sign. We can and we will assume that uu is strictly positive. Consider the conformal metric

g¯Yu4gY.\bar{g}_{Y}\coloneqq u^{4}\cdot g_{Y}.

The classical formula for the scalar curvature under conformal change [besse-einstein-manifolds, Corollary 1.161] gives

Scg¯Y=8u5(ΔgYu+18ScgYu).\operatorname{Sc}_{\bar{g}_{Y}}=8\cdot u^{-5}\left(-\Delta_{g_{Y}}u+\frac{1}{8}\operatorname{Sc}_{g_{Y}}u\right). (3.14)

We now compare the operator norms of dfYdf_{Y} with respect to the metrics gYg_{Y} and g¯Y\bar{g}_{Y}. Correspondingly, we use the notation dfYgY\left\|df_{Y}\right\|_{g_{Y}} and dfYg¯Y\left\|df_{Y}\right\|_{\bar{g}_{Y}}. A direct calculation shows that

dfYg¯Y2=u4dfYgY2.\left\|df_{Y}\right\|^{2}_{\bar{g}_{Y}}=u^{-4}\left\|df_{Y}\right\|^{2}_{g_{Y}}. (3.15)

Since uu is postive, from Eqs. 3.13, 3.14 and 3.15 we deduce

Scg¯Y=6u4dfYgY2+8λu4>6u4dfYgY2=6dfYg¯Y2.\displaystyle\operatorname{Sc}_{\bar{g}_{Y}}=6u^{-4}\left\|df_{Y}\right\|^{2}_{g_{Y}}+8\lambda u^{-4}>6u^{-4}\left\|df_{Y}\right\|^{2}_{g_{Y}}=6\left\|df_{Y}\right\|^{2}_{\bar{g}_{Y}}.

Hence, we constructed a three-dimensional closed oriented Riemannian manifold (Y,g¯Y)(Y,\bar{g}_{Y}) and a smooth map fY:(Y,g¯Y)(𝕊3,g𝕊3)f_{Y}\colon(Y,\bar{g}_{Y})\to(\mathbb{S}^{3},g_{\mathbb{S}^{3}}) of non-zero degree satisfying Scg¯Y>6dfYg¯Y2\operatorname{Sc}_{\bar{g}_{Y}}>6\left\|df_{Y}\right\|^{2}_{\bar{g}_{Y}}. By Remark 3.6, this contradicts Proposition 3.5. ∎

4. Ricci flow, harmonic map heat flow, and Einstein metrics

We employ estimates by Lee and Tam [LeeTam] on the harmonic map heat flow coupled with the Ricci flow to demonstrate that, under the hypotheses of A, if the metric gg is non-Einstein, then there exists a metric g~\tilde{g} and a function f~\tilde{f} satisfying the hypotheses of A with all inequalities being strict. For a comprehensive overview of the Ricci flow and the harmonic map heat flow, we recommend referring to [topping-lectures-ricci-flow] and [lin-wang-harmonic-maps], respectively. Recall that a metric gg is Einstein if Ricg=cg\operatorname{Ric}_{g}=cg for some constant cc, known as the proportionality constant of gg. For a smooth map f:(M,g)(N,h)f\colon(M,g)\to(N,h) of Riemannian manifolds, we denote by Lip(f)\operatorname{Lip}(f) the Lipschitz constant of ff.

Proposition 4.1.

Let (M,g)(M,g) be an nn-dimensional closed Riemannian manifold with Scgn(n1)\operatorname{Sc}_{g}\geq n(n-1), and let f:(M,g)(𝕊n,g𝕊n)f\colon(M,g)\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}) be a smooth, distance non-increasing map of non-zero degree. If the metric gg is non-Einstein, then there exists a Riemannian metric g~\tilde{g} on MM and a smooth map f~:(M,g~)(𝕊n,g𝕊n)\tilde{f}\colon(M,\tilde{g})\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}) of non-zero degree such that Scg~>n(n1)\operatorname{Sc}_{\tilde{g}}>n(n-1) and Lip(f~)<1\operatorname{Lip}(\tilde{f})<1.

Recall that a Ricci flow on a manifold MM is a smooth family of Riemannian metrics (gt)t[0,T](g_{t})_{t\in[0,T]} on MM satisfying the Ricci equation

tgt=2Ricgt.\partial_{t}g_{t}=-2\operatorname{Ric}_{g_{t}}.

The short-time existence and primary properties of the Ricci flow were established by Hamilton in his seminal work [hamilton-three-manifolds]. The following lemma summarizes well-known properties of the Ricci flow utilized in this paper. We include a brief proof for clarity.

Lemma 4.2.

Let (M,g)(M,g) be an nn-dimensional closed Riemannian manifold with Scgn(n1)\operatorname{Sc}_{g}\geq n(n-1). If (gt)t[0,T](g_{t})_{t\in[0,T]} is a Ricci flow on MM such that g0=gg_{0}=g, then

Scgtn(n1)12(n1)t,t[0,T].\operatorname{Sc}_{g_{t}}\geq\frac{n(n-1)}{1-2(n-1)t},\qquad\forall t\in[0,T]. (4.1)

Moreover, the previous inequality is strict for t(0,T]t\in(0,T], unless gg is an Einstein metric with Ricg=(n1)g\operatorname{Ric}_{g}=(n-1)g.

Proof.

Recall the orthogonal decomposition

Ricgt=Ric̊gt+Scgtngt,\operatorname{Ric}_{g_{t}}=\mathring{\operatorname{Ric}}_{g_{t}}+\frac{\operatorname{Sc}_{g_{t}}}{n}g_{t},

where Ric̊gt\mathring{\operatorname{Ric}}_{g_{t}} is the traceless component of the Ricci tensor of gtg_{t}. It is well-known that the scalar curvature Scgt\operatorname{Sc}_{g_{t}} evolves according to the equation

(tΔgt)Scgt=2|Ricgt|2=2|Ric̊gt|2+2Scgt2n2Scgt2n,\left(\partial_{t}-\Delta_{g_{t}}\right)\operatorname{Sc}_{g_{t}}=2|\operatorname{Ric}_{g_{t}}|^{2}=2|\mathring{\operatorname{Ric}}_{g_{t}}|^{2}+2\frac{\operatorname{Sc}_{g_{t}}^{2}}{n}\geq 2\frac{\operatorname{Sc}_{g_{t}}^{2}}{n}, (4.2)

see [topping-lectures-ricci-flow, Section 2.5]. Here, Δgt\Delta_{g_{t}} denotes the Laplace-Beltrami operator of (M,gt)(M,g_{t}). From (4.2), Scgt\operatorname{Sc}_{g_{t}} satisfies

(tΔgt)Scgt2Scgt2n.\left(\partial_{t}-\Delta_{g_{t}}\right)\operatorname{Sc}_{g_{t}}\geq 2\frac{\operatorname{Sc}_{g_{t}}^{2}}{n}.

The function ψ(t)=n(n1)/(12(n1)t)\psi(t)=n(n-1)/(1-2(n-1)t) solves

{ψ(t)=2ψ2(t)nψ(0)=n(n1).\begin{cases}\psi^{\prime}(t)=2\frac{\psi^{2}(t)}{n}\\ \psi(0)=n(n-1).\end{cases}

Since Scg0=Scgn(n1)=ψ(0)\operatorname{Sc}_{g_{0}}=\operatorname{Sc}_{g}\geq n(n-1)=\psi(0), from the weak minimum principle [topping-lectures-ricci-flow, Theorem 3.1.1] we deduce Inequality (4.1). To prove the last assertion, suppose Inequality (4.1) is an equality for some (p,T0)M×(0,T](p,T_{0})\in M\times(0,T]. By the strong minimum principle,

Scgt=ψ(t)=n(n1)12(n1)t,t[0,T0].\operatorname{Sc}_{g_{t}}=\psi(t)=\frac{n(n-1)}{1-2(n-1)t},\qquad\forall t\in[0,T_{0}].

See for example [mantegazza-mean-curvature-flow, Theorem 2.1.1]. Hence, tScgt=2Scgt2/n\partial_{t}\operatorname{Sc}_{g_{t}}=2\operatorname{Sc}_{g_{t}}^{2}/n, ΔgtScgt=0\Delta_{g_{t}}\operatorname{Sc}_{g_{t}}=0, and the inequality in (4.2) is an equality. Thus,

0=Ric̊gt=RicgtScgtngt=Ricgtψ(t)ngt,t[0,T0].0=\mathring{\operatorname{Ric}}_{g_{t}}=\operatorname{Ric}_{g_{t}}-\frac{\operatorname{Sc}_{g_{t}}}{n}g_{t}=\operatorname{Ric}_{g_{t}}-\frac{\psi(t)}{n}g_{t},\qquad\forall t\in[0,T_{0}].

This shows that gtg_{t} is an Einstein metric with proportionality constant Scgt/n=ψ(t)/n\operatorname{Sc}_{g_{t}}/n=\psi(t)/n for every t[0,T0]t\in[0,T_{0}]. In particular, g=g0g=g_{0} satisfies Ricg=(n1)g\operatorname{Ric}_{g}=(n-1)g. ∎

By utilizing estimates of Lee and Tam [LeeTam] for the harmonic map heat flow coupled with the Ricci flow, we derive the following lemma.

Lemma 4.3.

Let (M,g)(M,g) be an nn-dimensional closed Riemannian manifold, and let f:(M,g)(𝕊n,g𝕊n)f:(M,g)\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}) be a smooth, distance non-increasing map. Let (gt)t[0,T](g_{t})_{t\in[0,T]} be a Ricci flow on MM such that g0=gg_{0}=g. Then there exists T1(0,T]T_{1}\in(0,T] and a smooth family of smooth maps ft:(M,gt)(𝕊n,g𝕊n)f_{t}\colon(M,g_{t})\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}), for t[0,T1]t\in[0,T_{1}], such that f0=ff_{0}=f and

Lip(ft)112(n1)t,t[0,T1].\operatorname{Lip}(f_{t})\leq\frac{1}{1-2(n-1)t},\qquad\forall t\in[0,T_{1}]. (4.3)
Proof.

By [huang-tam-short-time-existence, Theorem 1.1], there exists T1(0,T]T_{1}\in(0,T] and a smooth family of smooth maps ft:(M,gt)(𝕊n,g𝕊n)f_{t}\colon(M,g_{t})\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}), for t[0,T1]t\in[0,T_{1}], such that

{tft=τ(ft)f0=f\begin{cases}\partial_{t}f_{t}=\tau(f_{t})\\ f_{0}=f\end{cases}

where τ(ft)\tau(f_{t}) is the tension field of the map ft:(M,gt)(𝕊n,g𝕊n)f_{t}\colon(M,g_{t})\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}). For the definition of the tension field, we refer to [huang-tam-short-time-existence, Section 4.1]. By applying [LeeTam, Theorem 2.1] with k=Ricgtk=-\operatorname{Ric}_{g_{t}} and κ=1\kappa=1, we conclude that each ftf_{t} satisfies (4.3). ∎

Proof of Proposition 4.1.

Suppose gg is not an Einstein metric with proportionality constant (n1)(n-1). Let (gt)t[0,T](g_{t})_{t\in[0,T]} be a Ricci flow on MM such that g0=gg_{0}=g. By Lemma 4.2,

Scgt>n(n1)12(n1)t,t(0,T].\operatorname{Sc}_{g_{t}}>\frac{n(n-1)}{1-2(n-1)t},\qquad\forall t\in(0,T].

By Lemma 4.3, there exist T1(0,T]T_{1}\in(0,T] and a smooth family of smooth maps ft:(M,gt)(𝕊n,g𝕊n)f_{t}\colon(M,g_{t})\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}), for t[0,T1]t\in[0,T_{1}], such that

Lip(ft)112(n1)t,t[0,T1].\operatorname{Lip}(f_{t})\leq\frac{1}{1-2(n-1)t},\qquad\forall t\in[0,T_{1}].

Since ff has non-zero degree, each ftf_{t} has non-zero degree as well. Let T2(0,T1]T_{2}\in(0,T_{1}] be fixed. Let ϵ>0\epsilon>0 be such that cϵ12(n1)T2ϵ>0c_{\epsilon}\coloneqq 1-2(n-1)T_{2}-\epsilon>0. By taking ϵ\epsilon sufficiently small, the metric cϵ1gT2c_{\epsilon}^{-1}g_{T_{2}} possesses the desired properties. ∎

5. Rigidity of the four-dimensional sphere

We combine the results from Sections 3 and 4 to establish the rigidity properties stated in A. First, we utilize Proposition 3.1 to address the case where all the inequalities in A are strict.

Lemma 5.1.

Let (M,g)(M,g) be a four-dimensional closed oriented Riemannian manifold with Scg>12\operatorname{Sc}_{g}>12. If f:(M,g)(𝕊4,g𝕊4)f\colon(M,g)\to(\mathbb{S}^{4},g_{\mathbb{S}^{4}}) is a strictly distance-decreasing smooth map, then ff has degree zero.

Before presenting the proof, we introduce some additional notation. For (0,π/2)\ell\in(0,\pi/2), consider the spherical band

(𝕊3,gcos)=(𝕊3×[,],cos2(t)g𝕊3+dt2).(\mathbb{S}^{3}_{\ell},g_{\cos})=\left(\mathbb{S}^{3}\times[-\ell,\ell],\cos^{2}(t)g_{\mathbb{S}^{3}}+dt^{2}\right).

Let p±p_{\pm} be antipodal points in 𝕊4\mathbb{S}^{4}. Utilizing the identification (𝕊4{p±},g𝕊4)(𝕊3×(π/2,π/2),cos2(t)g𝕊3+dt2)(\mathbb{S}^{4}\setminus\{p_{\pm}\},g_{\mathbb{S}^{4}})\cong(\mathbb{S}^{3}\times(-\pi/2,\pi/2),\cos^{2}(t)g_{\mathbb{S}^{3}}+dt^{2}), we view (𝕊3,gcos)(\mathbb{S}^{3}_{\ell},g_{\cos}) as a Riemannian band isometrically embedded in (𝕊4{p±},g𝕊4)(\mathbb{S}^{4}\setminus\{p_{\pm}\},g_{\mathbb{S}^{4}}).

Proof of Lemma 5.1.

Suppose, for the sake of contradiction, that ff has non-zero degree. Let δ(0,1)\delta\in(0,1) be such that Lip(f)1δ\operatorname{Lip}(f)\leq 1-\delta. Since the antipodal map on 𝕊4\mathbb{S}^{4} is an isometry, using the Brown-Sard theorem, we choose two antipodal points p±p_{\pm} in 𝕊4\mathbb{S}^{4} that are regular values of ff. For (0,π/2)\ell\in(0,\pi/2), we consider the spherical band (𝕊3,gcos)(\mathbb{S}^{3}_{\ell},g_{\cos}) isometrically embedded in (𝕊4{p±},g𝕊4)(\mathbb{S}^{4}\setminus\{p_{\pm}\},g_{\mathbb{S}^{4}}). Since MM is compact, f1({±p})f^{-1}(\{\pm p\}) is a finite set. Since p+p_{+} is a regular value of ff, we choose an open neighborhood U+U_{+} of p+p_{+} such that each component of f1(U+)f^{-1}(U_{+}) contains only one point in f1(p+)f^{-1}(p_{+}) and ff is a diffeomorphism when restricted to every component of f1(U+)f^{-1}(U_{+}). We choose an open neighborhood UU_{-} of pp_{-} with similar properties as U+U_{+}, such that UU+=U_{-}\cap U_{+}=\emptyset. Additionally, we choose L(π/(2+2δ),π/2)L\in(\pi/(2+2\delta),\pi/2) such that ±𝕊L3U±\partial_{\pm}\mathbb{S}^{3}_{L}\subset U_{\pm}. With this choice, V=f1(𝕊L3)V=f^{-1}(\mathbb{S}^{3}_{L}) is a four-dimensional oriented band with f1(±𝕊L3)=±Vf^{-1}(\partial_{\pm}\mathbb{S}^{3}_{L})=\partial_{\pm}V. Furthermore, the restriction fV=f|V:V𝕊L3f_{V}=f|_{V}\colon V\to\mathbb{S}^{3}_{L} is a smooth map of non-zero degree. We choose (L,π/2)\ell\in(L,\pi/2) such that

H(±V)>3tan()=Hgϕ(±𝕊3).\operatorname{H}(\partial_{\pm}V)>-3\tan(\ell)=H_{g_{\phi}}(\partial_{\pm}\mathbb{S}^{3}_{\ell}).

Let h:(𝕊L3,gcos)(𝕊3,gcos)h_{\ell}\colon(\mathbb{S}^{3}_{L},g_{\cos})\to(\mathbb{S}^{3}_{\ell},g_{\cos}) be the smooth map defined as

h(x,t)=(x,t/L)h_{\ell}(x,t)=(x,t\ell/L)

for (x,t)𝕊3×[,]=𝕊3(x,t)\in\mathbb{S}^{3}\times[-\ell,\ell]=\mathbb{S}^{3}_{\ell}. Note that Lip(h)<π/(2L)\operatorname{Lip}(h_{\ell})<\pi/(2L). Define fhfVf_{\ell}\coloneqq h_{\ell}\circ f_{V}. Since Lip(h)<π/(2L)\operatorname{Lip}(h_{\ell})<\pi/(2L) and L>π/(2+2δ)L>\pi/(2+2\delta), then Lip(f)<(1δ)π/(2L)<1\operatorname{Lip}(f_{\ell})<(1-\delta)\pi/(2L)<1. Finally, ff_{\ell} has non-zero degree, since it is the composition of maps of non-zero degree. Since Scg>12\operatorname{Sc}_{g}>12 and cos(t)\cos(t) is log-concave, this leads to a contradiction with Proposition 3.1. ∎

Next, we establish the scalar curvature rigidity of the nn-sphere for Einstein manifolds.

Lemma 5.2.

Let (M,g)(M,g) be an nn-dimensional closed connected oriented Einstein manifold with Ricg=(n1)g\operatorname{Ric}_{g}=(n-1)g. If f:(M,g)(𝕊n,g𝕊n)f\colon(M,g)\to(\mathbb{S}^{n},g_{\mathbb{S}^{n}}) is a smooth distance non-increasing map of non-zero degree, then ff is an isometry.

Proof.

By degree theory,

deg(f)𝕊n𝑑Vg𝕊n=Mf(dVg𝕊n).\deg(f)\int_{\mathbb{S}^{n}}\,dV_{g_{\mathbb{S}^{n}}}=\int_{M}f^{\ast}(dV_{g_{\mathbb{S}^{n}}}). (5.1)

Note that f(dVg𝕊n)=det(df)dVgf^{\ast}(dV_{g_{\mathbb{S}^{n}}})=\det(df)\,dV_{g}, where det(dfp)\det(df_{p}) is the determinant of dfp:TpMTf(p)𝕊ndf_{p}\colon T_{p}M\to T_{f(p)}\mathbb{S}^{n} as linear map of oriented inner product spaces. From (5.1),

|deg(f)|vol(𝕊n,g𝕊n)M|det(df)|𝑑Vg.\left|\deg(f)\right|\operatorname{vol}(\mathbb{S}^{n},g_{\mathbb{S}^{n}})\leq\int_{M}\left|\det(df)\right|\,dV_{g}. (5.2)

Let pMp\in M. Since ff is distance non-increasing, |det(dfp)|1\left|\det(df_{p})\right|\leq 1. Since Ricg=(n1)g\operatorname{Ric}_{g}=(n-1)g, by Bishop’s volume comparison [Petersen_Riemannian_geometry, Section 9.1.1, Lemma 35] vol(M,g)vol(𝕊n,g𝕊n)\operatorname{vol}(M,g)\leq\operatorname{vol}(\mathbb{S}^{n},g_{\mathbb{S}^{n}}). Hence, from (5.2) we deduce

|deg(f)|vol(𝕊n,g𝕊n)M|det(df)|𝑑Vgvol(M,g)vol(𝕊n,g𝕊n).\left|\deg(f)\right|\operatorname{vol}(\mathbb{S}^{n},g_{\mathbb{S}^{n}})\leq\int_{M}\left|\det(df)\right|\,dV_{g}\leq\operatorname{vol}(M,g)\leq\operatorname{vol}(\mathbb{S}^{n},g_{\mathbb{S}^{n}}).

Since deg(f)0\deg(f)\neq 0, it follows that |deg(f)|=1\left|\deg(f)\right|=1 and all inequalities must be equalities. Thus,

M|det(df)|𝑑Vg=vol(M,g).\int_{M}\left|\det(df)\right|\,dV_{g}=\operatorname{vol}(M,g).

Therefore, |det(dfp)|=1\left|\det(df_{p})\right|=1 for every pMp\in M. Since ff is distance non-increasing, we conclude that ff is a local isometry. Since 𝕊n\mathbb{S}^{n} is simply-connected, ff is an isometry. ∎

We are now in the position to prove our main theorem.

Proof of A.

Let us first show that gg must be Einstein with Ricg=3g\operatorname{Ric}_{g}=3g. Suppose, by contradiction, that this is not the case. By Proposition 4.1, there exists a Riemannian metric g~\tilde{g} on MM and a smooth map f~:(M,g~)(𝕊4,g𝕊4)\tilde{f}\colon(M,\tilde{g})\to(\mathbb{S}^{4},g_{\mathbb{S}^{4}}) of non-zero degree such that Scg~>12\operatorname{Sc}_{\tilde{g}}>12 and Lip(f)<1\operatorname{Lip}(f)<1, contradicting Lemma 5.1.

We conclude that gg is Einstein with Ricg=3g\operatorname{Ric}_{g}=3g. By Lemma 5.2, ff is an isometry. ∎

References