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

Riemannian 3-spheres that are hard to sweep out by short curves

Omar Alshawa and Herng Yi Cheng
Abstract

We construct a family of Riemannian 3-spheres that cannot be “swept out” by short closed curves. More precisely, for each L>0L>0 we construct a Riemannian 3-sphere MM with diameter and volume less than 1, so that every 2-parameter family of closed curves in MM that satisfies certain topological conditions must contain a curve that is longer than LL. This obstructs certain min-max approaches to bound the length of the shortest closed geodesic in Riemannian 3-spheres.

We also find obstructions to min-max estimates of the lengths of orthogonal geodesic chords, which are geodesics in a manifold that meet a given submanifold orthogonally at their endpoints. Specifically, for each L>0L>0, we construct Riemannian 3-spheres with diameter and volume less than 1 such that certain orthogonal geodesic chords that arise from min-max methods must have length greater than LL.

1 Introduction

How long is the shortest closed geodesic in a closed Riemannian manifold MM that is not a constant loop? M. Gromov asked whether its length, denoted by scg(M)\operatorname{scg}(M), can be bounded in terms of volume vol(M)\operatorname{vol}(M) of MM [22]. If that is impossible, can scg(M)\operatorname{scg}(M) be bounded by some function of vol(M)\operatorname{vol}(M) and the diameter diam(M)\operatorname{diam}(M) of MM?

Consider the example where MM is a Riemannian 3-sphere. Suppose that MM can be “swept out” by short curves, in the following sense. The unit 3-sphere S34S^{3}\subset\mathbb{R}^{4} is partitioned into subspaces Sst1=S3(2×{(s,t)})S^{1}_{st}=S^{3}\cap(\mathbb{R}^{2}\times\{(s,t)\}) for s,ts,t\in\mathbb{R} which are circles when s2+t2<1s^{2}+t^{2}<1 and points when s2+t2=1s^{2}+t^{2}=1. Suppose that there exists a continuous map F:S3MF:S^{3}\to M with nonzero degree such that every closed curve F(Sst1)F(S^{1}_{st}) is shorter than ϕ(vol(M),diam(M))\phi(\operatorname{vol}(M),\operatorname{diam}(M)) for some function ϕ\phi. Then a standard technique called the min-max method would imply that scg(M)ϕ(vol(M),diam(M))\operatorname{scg}(M)\leq\phi(\operatorname{vol}(M),\operatorname{diam}(M)). This method begins from the intuitive picture of MM being “swept out” by a 2-parameter family of “short” closed curves F(Sst1)F(S^{1}_{st}). Roughly speaking, a length-shortening process is applied to the entire family of curves F(Sst1)F(S^{1}_{st}) continuously, and the property that degF0\deg F\neq 0 would force one of the curves to converge to a short but non-constant closed geodesic of length at most ϕ(vol(M),diam(M))\phi(\operatorname{vol}(M),\operatorname{diam}(M)).

However, we have ruled out this approach to bounding scg(M)\operatorname{scg}(M). We accomplished this by constructing Riemannian 3-spheres with small diameter and volume, but which cannot be swept out by short closed curves in the aforementioned manner.

Theorem 1.1 (Main result).

For any L>0L>0, there exists a Riemannian 3-sphere M=(S3,g)M=(S^{3},g) of diameter and volume at most 11 with the following property: for any continuous map F:S3MF:S^{3}\to M with nonzero degree, one of the closed curves F(Sst1)F(S^{1}_{st}) must be longer than LL.

Remark 1.2.

More general versions of Theorem 1.1 can be proven due to the following observation. A crucial ingredient in proving Theorem 1.1 is the fact that sSst1\bigcup_{s}S^{1}_{st} is a 2-sphere for all 0<t<10<t<1, because we will apply the Jordan curve theorem to those 2-spheres. In fact, it can be verified that our proof of this theorem will work even when {Sst1}s,t\{S^{1}_{st}\}_{s,t} is replaced by certain more general foliations {Zst}s,t\{Z_{st}\}_{s,t} of S3S^{3} by 1-cycles ZstZ_{st}, as long as for “almost all” tt, sZst\bigcup_{s}Z_{st} is a 2-sphere.

Theorem 1.1 should be appraised within the broader context of min-max methods and their applications to the study of the geometry of closed geodesics, minimal submanifolds, and other minimal objects. We will briefly survey these results in this Introduction and suggest potential implications of our theorem. We will also explain how our constructions were motivated by counterexamples to related conjectures.

The ideas behind our proof of Theorem 1.1 also helped us to prove a result related to the geometry of orthogonal geodesic chords, which may be seen as a “relative” analogue of closed geodesics. Given a Riemannian manifold MM and a closed submanifold NMN\subset M, an orthogonal geodesic chord is a geodesic in MM that starts and ends on NN, and which is orthogonal to NN at its endpoints. What is the length of the shortest orthogonal geodesic chord in this situation? In Section 1.5 we will present another result, Theorem 1.3, which obstructs certain min-max approaches to estimating the length of the shortest orthogonal geodesic chord.

When MM is a Riemannian manifold that is not simply connected, scg(M)\operatorname{scg}(M) is at most the systole of MM, which is the length of the shortest non-contractible loop. Gromov proved that when MM satisfies a topological condition called essentialness, its systole is at most c(n)vol(M)nc(n)\sqrt[n]{\operatorname{vol}(M)} where n=dimMn=\dim M and c(n)c(n) is a dimensional constant [22]. Essential manifolds include all closed surfaces that are not simply connected, as well as all real projective spaces and all tori. A. Nabutovsky showed that the inequality holds for c(n)=nc(n)=n [42].

On the other hand, when MM is simply connected, scg(M)\operatorname{scg}(M) cannot be bounded by a systole. In this sense, it is more difficult to bound scg(M)\operatorname{scg}(M) when MM is a Riemannian sphere. Nevertheless, such bounds are possible for Riemanian 2-spheres (S2,g)(S^{2},g). For instance, scg(S2,g)4diam(S2,g)\operatorname{scg}(S^{2},g)\leq 4\operatorname{diam}(S^{2},g), as proven by Nabutovsky and R. Rotman [43], and independently by S. Sabourau [55]. Rotman also proved that scg(S2,g)42area(S2,g)\operatorname{scg}(S^{2},g)\leq 4\sqrt{2}\sqrt{\operatorname{area}(S^{2},g)} [53]. These results are the sharpest known versions of such bounds for 2-spheres, which were first proven by C. B. Croke and then strengthened by several authors; the early history around these bounds is surveyed in [15, Section 4].

For manifolds MM of dimension at least 3, bounds on scg(M)\operatorname{scg}(M) have been proven under various conditions on the curvature of MM in [3, 58, 14, 44, 60, 54, 52]. On the other hand, no curvature-free bounds for scg(M)\operatorname{scg}(M) are known for simply-connected manifolds MM of dimension at least 3. For such manifolds, the min-max methods sketched before Theorem 1.1 are a main avenue for studying scg(M)\operatorname{scg}(M).

1.1 Min-max theory for sweepouts by free loops

The approach to bounding scg(M)\operatorname{scg}(M) by min-max methods is essentially a quantitative version of the proof that MM contains a closed geodesic. The existence of closed geodesics in MM was proven by applying a “geometric calculus of variations” to the free loop space ΛM\Lambda M, which is the space of continuous maps S1MS^{1}\to M given an appropriate topology [41]. Let Λ0M\Lambda^{0}M denote the space of constant loops on MM. Since closed geodesics are critical points in ΛM\Lambda M with respect to the length functional, we can, in some sense, find those critical points by applying the calculus of variations to continuous maps f:(Dd,Dd)(ΛM,Λ0M)f:(D^{d},\partial D^{d})\to(\Lambda M,\Lambda^{0}M), where ff represents a nontrivial element in πd(ΛM,Λ0M)\pi_{d}(\Lambda M,\Lambda^{0}M), and d1d\geq 1 is the minimal degree such that πd(ΛM,Λ0M)0\pi_{d}(\Lambda M,\Lambda^{0}M)\neq 0. We call such a map ff a sweepout of MM by free loops. The notion of a sweepout allows us to define the min-max value

λ(M)=inff:(Dd,Dd)(ΛM,Λ0M)0[f]πd(ΛM,Λ0M)sup(lengthf).\lambda(M)=\inf_{\begin{subarray}{c}f:(D^{d},\partial D^{d})\to(\Lambda M,\Lambda^{0}M)\\ 0\neq[f]\in\pi_{d}(\Lambda M,\Lambda^{0}M)\end{subarray}}\sup(\operatorname{length}\circ f). (1.1)

It can be proven that λ(M)\lambda(M) is the length of some nonconstant closed geodesic in MM. Indeed, a similar argument was used by A. I. Fet and L. A. Lyusternik to prove that MM must have at least one nonconstant closed geodesic [40].111Surveys of their proof are available in [8, 49]. Thus scg(M)λ(M)\operatorname{scg}(M)\leq\lambda(M), and a natural strategy to bound scg(M)\operatorname{scg}(M) is to bound λ(M)\lambda(M) by constructing a single sweepout of MM that is “efficient” in the sense that every free loop in the sweepout has bounded length.

Such a bound on λ(M)\lambda(M) may be possible for some combinations of geometric parameters of MM but not others. For example, as a consequence of work by Y. Liokumovich, Nabutovsky, and Rotman, every Riemannian 2-sphere M=(S2,g)M=(S^{2},g) satisfies λ(M)664area(S2,g)+2diam(S2,g)\lambda(M)\leq 664\sqrt{\operatorname{area}(S^{2},g)}+2\operatorname{diam}(S^{2},g) [37, Theorem 1.3]. Yet λ(S2,g)\lambda(S^{2},g) cannot be bounded in terms of diameter alone or area alone: Sabourau presented a family of Riemannian 2-spheres (S2,g)(S^{2},g) for which, after scaling, area(S2,g)1\operatorname{area}(S^{2},g)\leq 1 but λ(S2,g)\lambda(S^{2},g) grows without bound [55, Remark 4.10]. In a similar vein, Liokumovich constructed a family of Riemannian 2-spheres (S2,g)(S^{2},g) for which diam(S2,g)1\operatorname{diam}(S^{2},g)\leq 1 but λ(S2,g)\lambda(S^{2},g) grows without bound [32]. Evidently, the geometry of closed geodesics in MM can be probed by either constructing efficient sweepouts, or by constructing “pathological” Riemannian manifolds on which any sweepout must be inefficient.

Theorem 1.1 gives a family of Riemannian 3-spheres MM for which diam(M),vol(M)1\operatorname{diam}(M),\operatorname{vol}(M)\leq 1 but λ(M)\lambda(M) grows without bound. Such results proving that λ(M)\lambda(M) is “large” should be understood with the caveat that even if λ(M)\lambda(M) is large, MM may still contain short closed geodesics. This is illustrated by the aforementioned results on scg(S2,g)\operatorname{scg}(S^{2},g) and λ(S2,g)\lambda(S^{2},g). Furthermore, min-max values are in some sense the lengths of closed geodesics of “positive Morse index,” so they do not lead to estimates of the lengths of closed geodesics that are local minima of the length functional.

Our result also does not imply that scg(M)\operatorname{scg}(M) cannot be bounded by a fixed function of diam(M)\operatorname{diam}(M) and vol(M)\operatorname{vol}(M). What we can conclude is that if such a bound can be proven using min-max methods, then it is not enough to use sweepouts that represent nontrivial elements of π2(ΛM,Λ0M)\pi_{2}(\Lambda M,\Lambda^{0}M). More complicated types of sweepouts would be necessary, such as maps f:(X,X0)(ΛM,Λ0M)f:(X,X_{0})\to(\Lambda M,\Lambda^{0}M) where XX is not a disk, or sweepouts by objects like 1-cycles that generalize free loops.

1.2 Comparing our constructions to disks whose boundaries are hard to contract

Our constructions are inspired by some Riemannian nn-disks whose boundaries are “hard to contract” through surfaces of controlled (n1)(n-1)-volume. These Riemannian disks were constructed to answer questions of Gromov [23] and P. Papasoglu [50] which were special cases of the following general question: Given a Riemannian nn-disk DD, can D\partial D be homotoped to a point through DD while passing through only surfaces whose (n1)(n-1)-volumes are bounded by a given combination of the geometric parameters of DD and D\partial D?

Riemannian 2-disks whose boundaries are “hard to contract” through curves of length bounded in terms of the disk’s diameter were constructed by S. Frankel and M. Katz. More precisely, for each each C>0C>0 they constructed a Riemannian 2-disk DC2D^{2}_{C} such that diam(DC2),length(DC2)1\operatorname{diam}(D^{2}_{C}),\operatorname{length}(\partial D^{2}_{C})\leq 1, but every nullhomotopy of DC2\partial D^{2}_{C} must pass through a curve longer than CC [18].222Nevertheless, the boundary of a Riemannian 2-disk DD can always be contracted through curves of length at most length(D)+200diam(D)max{1,lnarea(D)diam(D)}\operatorname{length}(\partial D)+200\operatorname{diam}(D)\max\Big{\{}1,\ln\frac{\sqrt{\operatorname{area}(D)}}{\operatorname{diam}(D)}\Big{\}}, as proven by Liokumovich, Nabutovsky, and Rotman [37]. Similarly in dimension 3, P. Glynn-Adey and Z. Zhu constructed a family of Riemannian 3-disks DC3D^{3}_{C}, ranging over each C>0C>0, such that diam(DC3),vol(DC3)10\operatorname{diam}(D^{3}_{C}),\operatorname{vol}(D^{3}_{C})\leq 10 and area(DC3)=4π\operatorname{area}(\partial D^{3}_{C})=4\pi, but every null-homotopy of DC3\partial D^{3}_{C} must pass through some surface of area greater than CC [21]. Their construction of DC3D^{3}_{C} was inspired by a previous construction by D. Burago and S. Ivanov of Riemannian 3-tori with small asymptotic isoperimetric constants [9].

Our constructions in Theorem 1.1 are essentially a combination of DC3D^{3}_{C} and a modification of DC2D^{2}_{C} by Liokumovich [34]. Let us compare our constructions with DC3D^{3}_{C} and their predecessor constructions by Burago and Ivanov. Burago and Ivanov constructed Riemannian 3-tori that contained 3 intertwining solid tori, each homeomorphic to D2×S1D^{2}\times S^{1}. The metric on each solid torus is a product metric of a short metric on the S1S^{1} factor with a Euclidean metric on the D2D^{2} factor scaled by a large positive number. In contrast, DC3D^{3}_{C} is a Riemannian 3-disk containing two disjoint and linked solid tori T1T2T_{1}\cup T_{2}, each TiT_{i} being homeomorphic to D2×S1D^{2}\times S^{1}. The metric on each TiT_{i} is the product of a short metric on the S1S^{1} factor and a hyperbolic metric gg on the D2D^{2} factor with negative curvature of a large magnitude, so that (D2,g)(D^{2},g) has small diameter but large area. The metric MM we constructed is similar to DC3D^{3}_{C}, except that in each TiT_{i} we replace the hyperbolic metric on the D2D^{2} factor with another metric with small diameter and large area.

Our choice of the metric on the D2D^{2} factor in each TiT_{i} can be motivated from the fact that we are considering maps F:S3MF:S^{3}\to M of nonzero degree and studying the lengths of the loops F(Sst1)F(S^{1}_{st}). Each solid torus TiT_{i} only “sees” part of each loop, namely F(Sst1)TiF(S^{1}_{st})\cap T_{i}, which, assuming that the intersection is transverse, is either a closed curve or a union of arcs with endpoints on Ti\partial T_{i}. In other words, F(Sst1)TiF(S^{1}_{st})\cap T_{i} is a relative 1-cycle in (Ti,Ti)(T_{i},\partial T_{i}). These relative 1-cycles form a 2-parameter family ranging over ss and tt; we will demonstrate that some 1-parameter subfamily of those relative 1-cycles will project onto D2D^{2} to produce a “sweepout of D2D^{2} by relative 1-cycles,” a notion that will be defined in Section 1.3. Proving this involves certain technical complications that we will resolve.

In light of the above, we will construct MM so that D2D^{2} factor in each TiT_{i} has a Riemannian metric that is difficult to sweep out by short relative 1-cycles. This would imply that one of the relative 1-cycles in some TiT_{i}, and therefore one of the F(Sst1)F(S^{1}_{st})’s, has to be long. This metric on D2D^{2} will be adapted from Riemannian 2-spheres that are hard to sweep out by short 1-cycles, which were constructed by Liokumovich [34] following inspiration from [18, 32].

1.3 Min-max theory for sweepouts by cycles

Roughly speaking, if we take the min-max theory for sweepouts by free loops and replace free loops by relative cycles, we obtain a new min-max theory called Almgren-Pitts min-max theory that has been central to the study of minimal submanifolds.

More precisely, for any Riemannian nn-manifold MM with boundary, consider the space of relative flat kk-cycles in a MM with coefficients in GG, denoted by 𝒵k(M,M;G)\mathcal{Z}_{k}(M,\partial M;G). Intuitively one can think of it as the group of relative singular kk-cycles in MM endowed with a topology where two relative cycles are “close” when their difference can be filled by a (k+1)(k+1)-chain of small volume in MM; formal definitions are available in [16, 17]. One can define a sweepout of MM by relative kk-cycles to be a continuous map f:N𝒵k(M,M;G)f:N\to\mathcal{Z}_{k}(M,\partial M;G) from a simplicial complex NN so that f(ι)0f^{*}(\iota)\neq 0, where ιHnk(𝒵k(M,M;G);G)\iota\in H^{n-k}(\mathcal{Z}_{k}(M,\partial M;G);G) is the fundamental cohomology class of 𝒵k(M,M;G)\mathcal{Z}_{k}(M,\partial M;G). More generally, for any integer p1p\geq 1, ff is called a pp-sweepout by cycles if f(ιp)0f^{*}(\iota^{p})\neq 0, where ιp\iota^{p} denotes the pthp^{\text{th}} cup power. A min-max value called the pp-width can be defined as:

widthpk(M;G)=inff:N𝒵k(M,M;G)f(ιp)0sup(volkf).\operatorname{width}_{p}^{k}(M;G)=\inf_{\begin{subarray}{c}f:N\to\mathcal{Z}_{k}(M,\partial M;G)\\ f^{*}(\iota^{p})\neq 0\end{subarray}}\sup(\operatorname{vol}_{k}\circ f). (1.2)

Henceforth we will write widthpk(M)\operatorname{width}_{p}^{k}(M) to denote widthpk(M;)\operatorname{width}_{p}^{k}(M;\mathbb{Z}) when MM is orientable, and widthpk(M;2)\operatorname{width}_{p}^{k}(M;\mathbb{Z}_{2}) otherwise. The pp-widths form a non-decreasing sequence: width1k(M)width2k(M)width3k(M)\operatorname{width}_{1}^{k}(M)\leq\operatorname{width}_{2}^{k}(M)\leq\operatorname{width}_{3}^{k}(M)\leq\dotsb

The pp-widths of a closed Riemannian manifold MM are realized as the volumes of minimal submanifolds in MM, which may contain a “small” singular set. The study of pp-widths have led to existence proofs for minimal submanifolds and the solutions of several conjectures about minimal submanifolds. Almgren-Pitts Min-max theory and some its applications to these conjectures are surveyed in [12, 48].

Since every free loop in MM is an integral 1-cycle, we have width11(M)λ(M)\operatorname{width}_{1}^{1}(M)\leq\lambda(M). Like λ(M)\lambda(M), curvature-free bounds on the pp-widths of MM in terms of geometric parameters of MM are much better understood when dimM=2\dim M=2. For every closed Riemannian surface SS and p1p\geq 1 we have scg(S)widthp1(S)\operatorname{scg}(S)\leq\operatorname{width}_{p}^{1}(S) due to the recent work of O. Chodosh and C. Mantoulidis [11]. F. Balacheff and Sabourau proved that width11(S)108genus(S)+1area(S)\operatorname{width}_{1}^{1}(S)\leq 10^{8}\sqrt{\operatorname{genus}(S)+1}\sqrt{\operatorname{area}(S)} [2]. However, Liokumovich proved that width11(S)\operatorname{width}_{1}^{1}(S) (and, by extension, widthp1(S)\operatorname{width}_{p}^{1}(S) for all pp) cannot be bounded solely in terms of diam(S)\operatorname{diam}(S), answering a question of Sabourau [55]: the counterexamples from [32] can be adapted into a family of Riemannian surfaces SS for which diam(S)=1\operatorname{diam}(S)=1 but whose values of width11(S)\operatorname{width}_{1}^{1}(S) grow without bound [34]. This exemplifies the potential for bounds on λ\lambda, and counterexamples to conjectured bounds, to shed light on the pp-widths, and by extension the geometry of minimal submanifolds.

We adapted Liokumovich’s Riemannian surfaces of small diameter but large 1-width into Riemannian 2-disks with small diameter but large 1-width. These Riemannian 2-disks are hard to sweep out by short relative 1-cycles, and they serve as ingredients in our construction, as explained at the end of Section 1.2.

For manifolds of dimension at least 3, bounds on some of their pp-widths have been established under certain curvature assumptions [22, 24, 13, 20, 56, 36, 38]. However, there are no known curvature-free bounds on the pp-widths of manifolds of dimension 3 and above in terms of geometric parameters such as diameter and volume. In fact, for closed 3-manifolds MM and any pp, widthp2(M;2)\operatorname{width}_{p}^{2}(M;\mathbb{Z}_{2}) cannot be bounded in terms of diam(M)\operatorname{diam}(M) and vol(M)\operatorname{vol}(M). This can be shown in two steps: roughly speaking, MM can be “cut” into two regions of equal volume by a hypersurface of area width12(M;2)\operatorname{width}_{1}^{2}(M;\mathbb{Z}_{2}).333This follows due to an argument adapted from [34, p. 396]. On the other hand, Papasoglu and E. Swenson constructed, for each C>0C>0, a Riemannian 3-sphere of diameter and volume at most 1 for which any such “cutting hypersurface” must have area greater than CC [51].444Riemannian 3-disks that are “hard to cut” in this manner were independently constructed by Glynn-Adey and Zhu [21], except that they only studied cutting hypersurfaces that were embedded disks.

The preceding argument would not apply to widths widthpk(M)\operatorname{width}_{p}^{k}(M) where kdimM2k\leq\dim M-2. Nevertheless, our Theorem 1.1 may serve as a first step towards proving that the pp-widths widthp1\operatorname{width}_{p}^{1} of Riemannian 3-spheres cannot be bounded in terms of diameter and volume.

1.4 Min-max theory for sweepouts by slicing

Our result may also offer insights in the study of waists, which are another class of min-max values that were defined by Gromov in [22]. They arise from sweepouts of Riemannian nn-manifolds with boundary MM by relative kk-cycles that are obtained as “slices” of MM. More precisely, given a continuous map MnkM\to\mathbb{R}^{n-k}, its fibers can be considered as “slices” of MM. We can define the kk-waist of orientable nn-manifolds MM to be

waistk(M)=infσ:Mnksuptnkvolk(σ1(t)),\operatorname{waist}_{k}(M)=\inf_{\sigma:M\to\mathbb{R}^{n-k}}\sup_{t\in\mathbb{R}^{n-k}}\operatorname{vol}_{k}(\sigma^{-1}(t)),

where the infimum is taken over maps σ\sigma whose fibers are Lipschitz kk-cycles, such that the map nk𝒵k(M,M;)\mathbb{R}^{n-k}\to\mathcal{Z}_{k}(M,\partial M;\mathbb{Z}) given by tσ1(t)t\mapsto\sigma^{-1}(t) is continuous.

The known bounds on waists in terms of the geometric parameters of MM follow a pattern similar to the other min-max values that we introduced earlier. By the definitions we have width1k(M)waistk(M)\operatorname{width}_{1}^{k}(M)\leq\operatorname{waist}_{k}(M). For a Riemannian 2-sphere (S2,g)(S^{2},g), Liokumovich proved that waist1(S2,g)52area(S2,g)\operatorname{waist}_{1}(S^{2},g)\leq 52\sqrt{\operatorname{area}(S^{2},g)} [33].555A similar bound on waist1\operatorname{waist}_{1} for other closed Riemannian surfaces is implicit via the monotone sweepouts involved in the proof of [35, Theorem 1.1]. For manifolds of dimension at least 3, some bounds on waists have been obtained under curvature assumptions [38, 36, 56].

There have been more results on waistk(M)\operatorname{waist}_{k}(M) when k=dimM1k=\dim M-1; the aforementioned Riemannian 3-spheres that are “hard to cut” by hypersurfaces demonstrate that waist2(M)\operatorname{waist}_{2}(M), which is bounded from below by width12(M)\operatorname{width}_{1}^{2}(M), cannot be bounded in terms of diam(M)\operatorname{diam}(M) and vol(M)\operatorname{vol}(M) when dimM=3\dim M=3. On the flipside, it is currently an open question whether waistk(M)\operatorname{waist}_{k}(M) can be bounded in terms of the geometric parameters of MM when dimM3\dim M\geq 3 and kdimM2k\leq\dim M-2. Within this context, L. Guth conjectured that when MM is a Riemannian 3-torus, then waist1(M)Cvol(M)3\operatorname{waist}_{1}(M)\leq C\sqrt[3]{\operatorname{vol}(M)} for some constant CC [25, Section 7]. Our Theorem 1.1 could serve as a first step towards disproving this conjecture.

Our result can be contrasted with a recent result of Nabutovsky, Rotman, and Sabourau which proves bounds on another min-max value related to waists. To define this min-max value, for each closed Riemannian nn-manifold MM and an integer k0k\geq 0, consider a continuous map F:NMF:N\to M of nonzero degree from a closed nn-dimensional pseudomanifold NN. “Slice” NN similar to before using a continuous map σ:NK\sigma:N\to K to a finite simplicial complex KK whose fibers σ1(t)\sigma^{-1}(t) are kk-dimensional simplicial complexes. Then define

Wk(M)=infσ:NKF:NMdegf0suptKvolk(F(σ1(t))),W_{k}(M)=\inf_{\begin{subarray}{c}\sigma:N\to K\\ F:N\to M\\ \deg f\neq 0\end{subarray}}\>\sup_{t\in K}\>\operatorname{vol}_{k}(F(\sigma^{-1}(t))),

where the infimum range over all σ\sigma and FF that satisfy the stipulated criteria.

These min-max values are related to waists via Wk(M)waistk(M)W_{k}(M)\leq\operatorname{waist}_{k}(M). The bounds Wk(M)cnvol(M)nW_{k}(M)\leq c_{n}\sqrt[n]{\operatorname{vol}(M)} and Wk(M)cndiam(M)W_{k}(M)\leq c_{n}^{\prime}\operatorname{diam}(M) were proven by Nabutovsky, Rotman, and Sabourau for some dimensional constants cnc_{n} and cnc_{n}^{\prime} [47]. However, it remains to be seen whether Wk(M)W_{k}(M) is attained as the volume of some minimal object. Since each F(σ1(t))F(\sigma^{-1}(t)) may not be a loop or a cycle (see [47, Example 1.2]), the min-max theory for minimal submanifolds does not fit here. For the same reason, there is no direct comparison between W1(M)W_{1}(M) and λ(M)\lambda(M) or widthp1(M)\operatorname{width}_{p}^{1}(M).

1.5 Orthogonal geodesic chords and min-max theory for sweepouts by paths

The techniques we used to prove Theorem 1.1 can be adapted for yet another min-max theory that arises from using sweepouts by paths instead of free loops or 1-cycles. This has led to a result about the geometry of orthogonal geodesic chords, which were defined earlier as geodesics in a Riemannian manifold MM that meet a fixed submanifold NN orthogonally at its endpoints. Orthogonal geodesic chords are 1-dimensional analogues of free boundary minimal submanifolds, which are submanifolds PMP\subset M with vanishing mean curvature such that PN\partial P\subset N and P\partial P meets NN orthogonally.666Some results about the existence and regularity of free boundary minimal submanifolds are surveyed in [31]. Orthogonal geodesic chords are also related to brake orbits, special types of periodic orbits in certain Hamiltonian systems [19].

When dimN=0\dim N=0, an orthogonal geodesic chord is simply a geodesic with specified endpoints. The existence of such geodesics and bounds on their length were studied in [57, 59, 46, 45, 10, 4]. Lyusternik and L. Schnirelmann proved that every convex domain MnM\subset\mathbb{R}^{n} with boundary NN contains nn orthogonal geodesic chords [39]. W. Bos extended this existence result to Riemannian nn-disks MM with convex boundary NN [7]. For Riemannian 2-disks MM with strictly convex boundary NN, J. Hass and P. Scott [27] and D. Ko [29] showed that one can even arrange the orthogonal geodesic chords to be simple, that is, avoid self-intersection.

The existence and geometry of orthogonal geodesic chords may be probed using min-max techniques as follows. Let ΩNM\Omega_{N}M denote the space of piecewise smooth paths in MM that start and end on NN, topologized as in [41, p. 88]. Let d1d\geq 1 be the smallest degree for which πd(ΩNM,Λ0N)0\pi_{d}(\Omega_{N}M,\Lambda^{0}N)\neq 0. Then define

λrel(M,N)=inff:(Dd,Dd)(ΩNM,Λ0N)0[f]πd(ΩNM,Λ0N)sup(Ef),\lambda_{\mathrm{rel}}(M,N)=\inf_{\begin{subarray}{c}f:(D^{d},\partial D^{d})\to(\Omega_{N}M,\Lambda^{0}N)\\ 0\neq[f]\in\pi_{d}(\Omega_{N}M,\Lambda^{0}N)\end{subarray}}\sup(E\circ f), (1.3)

where EE is the energy of a path, E(α)=01α(t)2𝑑tE(\alpha)=\int_{0}^{1}\lVert\alpha^{\prime}(t)\rVert^{2}\,dt. (Using length instead of energy would not significantly affect the resulting min-max theory, because the Hölder inequality relates length to energy.) X. Zhou proved that when MM is a complete and homogeneously regular Riemannian manifold and NN is a closed submanifold such that π1(M,N)=0\pi_{1}(M,N)=0 and π2(M,N)0\pi_{2}(M,N)\neq 0, then λrel(M,N)>0\lambda_{\mathrm{rel}}(M,N)>0 is the energy of an orthogonal geodesic chord [61].

Recent results have estimated the lengths of orthogonal geodesics chords in various spaces MM and NN, where NN may or may not be M\partial M. When MM is a Riemannian 2-disk with convex boundary NN, I. Beach proved that MM contains at least two distinct simple orthogonal geodesic chords whose lengths are bounded by f(diam(M),area(M),length(M))f(\operatorname{diam}(M),\operatorname{area}(M),\operatorname{length}(\partial M)) for some function ff [5]. In addition, recent work by Beach, H. C. Peruyero, E. Griffin, M. Kerr, Rotman, and C. Searle implies that when MM is a closed Riemannian manifold and NN is an analytic 2-sphere embedded in MM, then MM contains an orthogonal geodesic chord whose length is bounded by f(dimM,diam(M),diam(N),area(N))f^{\prime}(\dim M,\operatorname{diam}(M),\operatorname{diam}(N),\operatorname{area}(N)) for some function ff^{\prime} [6]. Both of these results were obtained by proving that either λrel(M,N)\lambda_{\mathrm{rel}}(M,N) is bounded by the relevant function (ff or ff^{\prime}) due to the existence of a sweepout of MM by curves of energy (equivalently, length) bounded by that function, or else the obstruction to the existence of such a sweepout is an orthogonal geodesic chord of length bounded by that function.

When MM is a Riemannian 3-sphere and NN is an embedded sphere, we proved that λrel(M,N)\lambda_{\mathrm{rel}}(M,N) cannot be bounded by any function of vol(M)\operatorname{vol}(M), diam(M)\operatorname{diam}(M), and diam(N)\operatorname{diam}(N).

Theorem 1.3.

For any E>0E>0, there exists a Riemannian 3-sphere MM of diameter and volume at most 1 that contains an embedded 2-sphere NN of diameter at most 1 such that λrel(M,N)\lambda_{\mathrm{rel}}(M,N). MM also contains an embedded circle γ\gamma of length at most 1 such that λrel(M,γ)>E\lambda_{\mathrm{rel}}(M,\gamma)>E.

2 The Construction of our 3-Spheres

For any L>0L>0, Liokumovich constructed a Riemannian 2-sphere ShS_{h} of diameter at most 1 for which width11(Sh)>L\operatorname{width}_{1}^{1}(S_{h})>L [34, pg. 2]. ShS_{h} is roughly constructed as follows. Take a unit disk BB in a hyperbolic plane, and embed a regular ternary tree with unit edge length and height hh in 2\mathbb{R}^{2}. Glue BB to a small tubular neighbourhood of the tree in 2\mathbb{R}^{2} by identifying their boundaries. (The curvature of BB must be chosen such that the boundaries have the same length.) The result is ShS_{h} (see fig. 1(a)).

The metric of ShS_{h} has the symmetries of an equilateral triangle, generated by reflections and rotations. A plane of reflection cuts ShS_{h} into two isometric Riemannian disks, one of which we denote by DhD_{h} (see fig. 1(b)). It can be verified that diamDh12\operatorname{diam}D_{h}\leq\frac{1}{2}.

(a) (b)
Refer to caption Refer to caption
(c) (d)
Refer to caption Refer to caption
Figure 1: (a) A 2-sphere ShS_{h} of small diameter but large width from [34]. (b) A Riemannian disk DhD_{h} whose double is ShS_{h}. (c) A 1-cycle (blue) in ShS_{h}, which is the double of the relative 1-cycle AA (blue) in DhD_{h}, shown in (d).

The key property of DhD_{h} is that it has a small diameter but large width:

Lemma 2.1.

For any C>0C>0, there exists some hh such that width11(Dh)>C\operatorname{width}_{1}^{1}(D_{h})>C.

Proof.

Consider a sweepout by relative 1-cycles f:N𝒵1(Dh,Dh;)f:N\to\mathcal{Z}_{1}(D_{h},\partial D_{h};\mathbb{Z}); unpacking the definition, this implies that for some loop g:S1Ng:S^{1}\to N, the gluing homomorphism sends gfg\circ f to mm times of the fundamental class of DhD_{h}, for some m0m\neq 0. Consider the map f:N𝒵1(Sh;)f^{\prime}:N\to\mathcal{Z}_{1}(S_{h};\mathbb{Z}) where the 1-cycle f(x)f^{\prime}(x) (see fig. 1(c)) is the double of the relative 1-cycle f(x)f(x) (see fig. 1(d)), which is obtained by subtracting a reflected copy of f(x)f(x) from f(x)f(x). (The subtraction ensures that the orientations match up where the two copies of f(x)f(x) meet.) Then it can be verified that the gluing homomorphism sends fgf^{\prime}\circ g to ±m\pm m times of the fundamental class of ShS_{h}. The lemma then follows from the fact that ShS_{h} can have arbitrarily large width11\operatorname{width}_{1}^{1}. ∎

For i=1,2i=1,2, define Ti=Dh×S1T_{i}=D_{h}\times S^{1} to be solid tori, and let ghg^{\prime}_{h} be the product metric of DhD_{h} with a sufficiently short metric on S1S^{1}. Embed the disjoint union T1T2T_{1}\cup T_{2} into S3S^{3} in the manner of a Hopf link (see fig. 2), and denote their union by LTLT. Cover S3S^{3} by two open sets U1U_{1} and U2U_{2} so that LTU1U2LT\subset U_{1}\setminus U_{2}, and let ϕ1,ϕ2\phi_{1},\phi_{2} be a partition of unity subordinate to this cover. Extend ghg^{\prime}_{h} over U1U_{1} via a smooth bump function. Our Riemannian 33-sphere is then Mh=(S3,gh)M_{h}=(S^{3},g_{h}), where gh=ghϕ1+g0ϕ2g_{h}=g^{\prime}_{h}\phi_{1}+g_{0}\phi_{2} and g0g_{0} is the metric of a round sphere of sufficiently small radius. It can be verified that vol(Mh),diam(Mh)1\operatorname{vol}(M_{h}),\operatorname{diam}(M_{h})\leq 1.

3 Plan of the Proof of Theorem 1.1

In order to discuss the plan, we first make some notation clear. Let πi:TiDh\pi_{i}:T_{i}\rightarrow D_{h} be a projection onto the first factor. Recall the definition of Sst1S^{1}_{st} from the beginning of the Introduction. Define St2=S3(3×{t})S^{2}_{t}=S^{3}\cap(\mathbb{R}^{3}\times\{t\}).

Consider a map F:S3MhF:S^{3}\to M_{h} of nonzero degree. After perturbing FF, each Xi=F1(Ti)X_{i}=F^{-1}(\partial T_{i}) will be a smooth manifold containing a family of 1-cycles XiSt2X_{i}\cap S^{2}_{t}. Their images F(XiSt2)F(X_{i}\cap S^{2}_{t}) will form a sweepout of Ti\partial T_{i}. An analysis of degF|Xi\deg F|_{X_{i}} and the following lemma of algebraic topology guarantees the existence of a circle CC in some XiSt2X_{i}\cap S^{2}_{t} that maps to a non-contractible loop F(C)F(C) on Ti\partial T_{i}:

Lemma 3.1.

Let XX be a closed, orientable, and connected surface of genus at least 1 and consider a degree nonzero map α:XS1×S1\alpha:X\to S^{1}\times S^{1}. Then for any Morse function ϕ:X\phi:X\to\mathbb{R}, there exists some tt\in\mathbb{R} and a circle CC embedded in ϕ1(t)\phi^{-1}(t) such that α|C\alpha|_{C} is not nullhomotopic.

This result follows from [21, Lemma 2.2]. Nevertheless, at the end of this section we will sketch a proof for the case where α\alpha is a diffeomorphism, in order to articulate the fundamental reason why it is true.

As shown in fig. 2, we find that F|CF|_{C} has a nonzero linking number with the core curve of either T1T_{1} or T2T_{2}. As a consequence of Lemma 4.3, there will exist a surface ΣSt2\Sigma\subset S^{2}_{t} such that F(Σ)F(\Sigma) is contained within some TiT_{i}. In addition, Σ\Sigma will be partitioned into the curves ΣSst1\Sigma\cap S^{1}_{st}, and the projections of the curves F(ΣSst1)F(\Sigma\cap S^{1}_{st}) onto DhD_{h} via πi\pi_{i} will give a sweepout of DhD_{h} by relative 1-cycles. fig. 4 illustrates Σ\Sigma and F(Σ)F(\Sigma), where Σ\Sigma corresponds to Σ1\Sigma_{1} in the figure.

Refer to caption
Figure 2: When F|C:CT1F|_{C}:C\to\partial T_{1} (red curve) is not nullhomotopic, it has a nonzero linking number with either the core curve γ1\gamma_{1} of T1T_{1} (left picture) or the core curve γ2\gamma_{2} of T2T_{2} (right picture).

One of the curves in this sweepout by relative 1-cycles must be long, by Lemma 2.1. As these relative 1-cycles are orthogonal projections of the curves F(ΣSst1)F(\Sigma\cap S^{1}_{st}), it must be the case that some F(Sst1)F(S^{1}_{st}) is long as well.

Proof sketch of Lemma 3.1 when α\alpha is a diffeomorphism.

Since α\alpha is a diffeomorphism, let us identify XX with S1×S1S^{1}\times S^{1}. Let the critical values of ϕ\phi be t0<t1<<tmt_{0}<t_{1}<\dotsb<t_{m}. For each ii, ϕ1(ti)\phi^{-1}(t_{i}) is a disjoint union of wedge sums of circles. Denote those circle wedge summands by Ci,1,,Ci,kiC_{i,1},\dotsc,C_{i,k_{i}} (see fig. 3(a)). Let us assume that each circle Ci,jC_{i,j} is contractible in XX and derive a contradiction. In other words, we assume that each Ci,jC_{i,j} is the boundary of some 2-chain Di,jD_{i,j} in XX that is the continuous image of a 2-disk (see fig. 3(c)).

We can decompose the fundamental cycle of XX into the sum of 2-chains Ai=ϕ1([ti1,ti])A_{i}=\phi^{-1}([t_{i-1},t_{i}]) for i=1,,mi=1,\dotsc,m (see fig. 3(b)). Each Ai\partial A_{i} is composed of circles Ci,jC_{i^{\prime},j^{\prime}} (for i=i1,ii^{\prime}=i-1,i), and for each of these circles we can glue in Di,jD_{i^{\prime},j^{\prime}}. After “capping off” each circle in Ai\partial A_{i}, AiA_{i} becomes a 2-cycle that is the union of images of 2-spheres. To illustrate, in fig. 3(b), we can see that A1A_{1} and A3A_{3} have been capped off into spheres, while A2A_{2} has been capped off into the union of two spheres.

Refer to caption
Refer to caption
Refer to caption
Figure 3: (a) Level sets of a Morse function on a 2-torus XX. (b) “Capping off” each circle in each level set produces a decomposition of the fundamental cycle of XX into a sum of images of spheres. (c) The images of disks used for “capping off.”

However, as π2(X)=0\pi_{2}(X)=0, all of these images of 2-spheres are null-homologous. If we add up all of these images of spheres, then the caps Di,jD_{i,j} cancel each other out and the result is the fundamental cycle of XX. Thus we have expressed the fundamental cycle as a sum of null-homologous 2-cycles, which gives a contradiction. ∎

4 Main Result

We begin by proving that continuous maps S3MS^{3}\to M of nonzero degree can be perturbed to “geometrically nice” maps.

Lemma 4.1.

Consider any map F:S3MF:S^{3}\rightarrow M of nonzero degree such that every curve F(Sst1)F(S^{1}_{st}) is shorter than LL for some L>0L>0. Then for any δ>0\delta>0 and any closed submanifold YMY\subset M, FF is homotopic to a smooth map F^\hat{F} transverse to YY such that:

  1. 1.

    For any ss and tt, the lengths of F(Sst1)F(S^{1}_{st}) and F^(Sst1)\hat{F}(S^{1}_{st}) differ by at most δ\delta.

  2. 2.

    The sets F^1(Y)St2\hat{F}^{-1}(Y)\cap S^{2}_{t} are the level sets of some Morse function on F^1(Y)\hat{F}^{-1}(Y).

Proof.

We can approximate FF by a sequence of smooth maps Fi:S3MF_{i}:S^{3}\to M for i=1,2,i=1,2,\dotsc that are transverse to YY, so that each Xi=Fi1(Y)X_{i}=F_{i}^{-1}(Y) is a smooth manifold. It can be verified from the standard arguments for such approximations (e.g. in [30]), together with the fact that every F(Sst1)F(S^{1}_{st}) is shorter than LL, that each FiF_{i} can be chosen to satisfy (1). Thinking of XiX_{i} as a submanifold of 4\mathbb{R}^{4}, the function dp:Xi4d_{p}:X_{i}\to\mathbb{R}^{4} defined by dp(x)=xp2d_{p}(x)=\lVert x-p\rVert^{2} is Morse for generic p4p\in\mathbb{R}^{4} [41, p. 36]. Let us pick p(0,0,0,2)p\approx(0,0,0,2); then the rotational symmetry of S3S^{3} implies that the level sets of dpd_{p} are the intersections of XiX_{i} with hyperplanes orthogonal to pp. Therefore we may modify FiF_{i} and dpd_{p} by precomposing them with an isometry S3S3S^{3}\to S^{3} that is close to the identity, until those level sets become XiSt2X_{i}\cap S^{2}_{t}. Finally, we may choose F^\hat{F} to be FiF_{i} for sufficiently large ii. ∎

Thus we may replace FF by its perturbation F^\hat{F}. Henceforth we will assume that FF satisfies the properties in Lemma 4.1. Let X=F1(T1T2)X=F^{-1}(\partial T_{1}\cup\partial T_{2}). Then like any level set of a Morse function on a surface, each XSt2X\cap S^{2}_{t} is a disjoint union of circles and at most one figure eight (a wedge sum of two circles).

The proofs of the following two lemmas were inspired by the proof of [21, Lemma 2.3]. However, that proof contained an inaccuracy, so we give our own proofs.

Lemma 4.2.

Suppose that one of the circles CC embedded in XSt2X\cap S^{2}_{t} is such that F|CF|_{C} has nonzero linking number mm with the core curve of TiT_{i} for some i=1,2i=1,2. Then F1(Ti)St2F^{-1}(T_{i})\cap S^{2}_{t} contains a surface Σ\Sigma such that (πiF):H2(Σ,Σ)H2(Dh,Dh)(\pi_{i}\circ F)_{*}:H_{2}(\Sigma,\partial\Sigma)\rightarrow H_{2}(D_{h},\partial D_{h}) is a nonzero map.

Proof.

Without loss of generality, assume F|CF|_{C} has linking number m0m\neq 0 with the core curve of T1T_{1}. CC then bounds a disk Δ\Delta in St2S^{2}_{t} such that F(Δ)F(\Delta) has intersection number mm with the core curve of T1T_{1}.777With reference to Remark 1.2 about generalizations of Theorem 1.1, this is the part of our proof of Theorem 1.1 that requires the Jordan curve theorem. ΔF1(T1)\Delta\cap F^{-1}(T_{1}) is a union of disjoint surfaces Σ1Σk\Sigma_{1}\cup\dotsb\cup\Sigma_{k}, each corresponding to an intersection number mim_{i} between F(Σi)F(\Sigma_{i}) and the core curve of T1T_{1}. This is reflected in fig. 4. Then m=m1++mkmi0m=m_{1}+\dotsb+m_{k}\implies m_{i}\neq 0 for some mim_{i}. Thus Σi\Sigma_{i} is the desired surface. ∎

Refer to caption
Figure 4: An illustration of some elements from the proof of Lemma 4.2.

We then use Lemma 4.2 to prove the following statement:

Lemma 4.3.

For some tt and for some i=1,2i=1,2, F1(Ti)St2F^{-1}(T_{i})\cap S^{2}_{t} contains a surface Σ\Sigma such that (πiF):H2(Σ,Σ)H2(Dh,Dh)(\pi_{i}\circ F)_{*}:H_{2}(\Sigma,\partial\Sigma)\rightarrow H_{2}(D_{h},\partial D_{h}) is a nonzero map.

Proof.

Let X1=F1(T1)X_{1}=F^{-1}(\partial T_{1}). By Lemma 4.1 we can assume that X1St2=ϕ1(t)X_{1}\cap S^{2}_{t}=\phi^{-1}(t) for some Morse function ϕ:X1\phi:X_{1}\to\mathbb{R}. After choosing an orientation on T1\partial T_{1} and equipping X1X_{1} with the preimage orientation, it can be verified that the map F|X1:X1T1F|_{X_{1}}:X_{1}\to\partial T_{1} has degree equal to degF0\deg F\neq 0. (To see this, note that the preimages of a regular value of FF have neighbourhoods that are foliated by the St2S^{2}_{t}’s.) Consequently, for some connected component X1X_{1}^{\prime} of X1X_{1}, degF|X10\deg F|_{X_{1}^{\prime}}\neq 0. Since π2(T2)=0\pi_{2}(\partial T_{2})=0, the genus of X1X_{1}^{\prime} must be at least 1. Thus Lemma 3.1 gives some tt\in\mathbb{R} and a circle CC embedded in ϕ1(t)\phi^{-1}(t) such that F|CF|_{C} is not contractible in T1\partial T_{1}.

In other words, F:π1(C)π1(T1)F_{*}:\pi_{1}(C)\cong\mathbb{Z}\rightarrow\pi_{1}(\partial T_{1}) is a nonzero map with F(1)=(n1,n2)F_{*}(1)=(n_{1},n_{2}), where the first factor is a multiple of the generator homotopic to Dh\partial D_{h} and the second factor is a multiple of the generator homotopic through T1T_{1} to its core curve.

Then if n10n_{1}\neq 0, we have that F|CF|_{C} has nonzero linking number with the core curve of T1T_{1}, and so we may apply Lemma 4.2 to obtain Σ\Sigma as needed.

If n1=0n_{1}=0, then n20n_{2}\neq 0 where F|CF|_{C} has a nonzero linking number with the core curve of T2T_{2}, and so we again apply Lemma 4.2 to obtain Σ\Sigma. ∎

We have now proven the existence of a surface Σ\Sigma on St2S^{2}_{t} that maps into one of the tori in a “nice” way. We use this property to define a continuous family of 1-cycles on DhD_{h} and show that one of those 1-cycles must be long. This will eventually imply that one of the F(Sst1)F(S^{1}_{st})’s must also be long, leading to a proof of Theorem 1.1.

Proof of Theorem 1.1.

Consider any L>0L>0. Consider the Riemannian 3-sphere Mh=(S3,gh)M_{h}=(S^{3},g_{h}) that was constructed in Section 2, with hh chosen such that width11(Dh)>L+1\operatorname{width}_{1}^{1}(D_{h})>L+1, as in Lemma 2.1. Suppose for the sake of contradiction that for some continuous map F:S3MhF:S^{3}\to M_{h}, every curve F(Sst1)F(S^{1}_{st}) is shorter than LL. Recall that by applying a perturbation, we may assume that FF satisfies the properties of Lemma 4.1, for some δ<12\delta<\frac{1}{2}. We will arrive at a contradiction by proving that some F(Sst1)F(S^{1}_{st}) (for the perturbed FF) is longer than L+1L+1.

Applying Lemma 4.3, we obtain a surface Σ\Sigma in St2S^{2}_{t} for some tt such that F(Σ)TiF(\Sigma)\subset T_{i} for some ii, and (πiF):H2(Σ,Σ)H2(Dh,Dh)(\pi_{i}\circ F)_{*}:H_{2}(\Sigma,\partial\Sigma)\to H_{2}(D_{h},\partial D_{h}) is a nonzero map. With this, we may define a continuous family of relative 1-cycles K:[1,1]𝒵1(Dh,Dh;)K:[-1,1]\rightarrow\mathcal{Z}_{1}(D_{h},\partial D_{h};\mathbb{Z}) by K(s)=πi(F(Sst1Σ))K(s)=\pi_{i}(F(S^{1}_{st}\cap\Sigma)). Since K(±1)=0K(\pm 1)=0, we can think of KK as a map S1Z1(Dh,Dh)S^{1}\to Z_{1}(D_{h},\partial D_{h}).

We will show that KK gives a sweepout of DhD_{h} by relative 1-cycles by appealing to the Almgren isomorphism theorem, which implies that there is a natural ismorphism Γ:π1(𝒵1(Dh,Dh;))H2(Dh,Dh)\Gamma:\pi_{1}(\mathcal{Z}_{1}(D_{h},\partial D_{h};\mathbb{Z}))\xrightarrow{\cong}H_{2}(D_{h},\partial D_{h}) [1].888A modern proof of the Almgren isomorphism theorem is available in [26]. Γ\Gamma is induced by “gluing” a family of relative 1-cycles in DhD_{h} into a relative 2-cycle in DhD_{h} (see fig. 5). Gluing together the relative 1-cycles πi(F(Sst1Σ))\pi_{i}(F(S^{1}_{st}\cap\Sigma)) gives πi(F(Σ))\pi_{i}(F(\Sigma)), which represents a nontrivial class in H2(Dh,Dh)H_{2}(D_{h},\partial D_{h}) because (πiF):H2(Σ,Σ)H2(Dh,Dh)(\pi_{i}\circ F)_{*}:H_{2}(\Sigma,\partial\Sigma)\to H_{2}(D_{h},\partial D_{h}) is a nonzero map. By the Almgren isomorphism theorem, KK represents a nonzero element of π1(𝒵1(Dh,Dh;))\pi_{1}(\mathcal{Z}_{1}(D_{h},\partial D_{h};\mathbb{Z})), and thus K:H1(S1)H1(𝒵1(Dh,Dh;))K_{*}:H_{1}(S^{1})\to H_{1}(\mathcal{Z}_{1}(D_{h},\partial D_{h};\mathbb{Z})) is also a nonzero map. By definition, KK gives a sweepout of DhD_{h} by relative 1-cycles.

Refer to caption
Refer to caption
Figure 5: A 1-parameter family of 1-cycles in a Riemannian 2-disk (left) that “glues” into a relative 2-cycle (right), which in this case represents the relative fundamental class of the disk.

By Lemma 2.1, some K(s)K(s) must be longer than L+1L+1. As K(s)K(s) is an orthogonal projection of part of F(Sst1)F(S^{1}_{st}) onto DhD_{h}, we obtain that F(Sst1)F(S^{1}_{st}) must be longer than L+1L+1 as well. This gives a contradiction. ∎

5 Orthogonal Geodesic Chords

To prove Theorem 1.3, we will construct a sequence of Riemannian 3-spheres M¯h=(S3,g¯h)\bar{M}_{h}=(S^{3},\bar{g}_{h}) with small diameter and volume, so that λrel(M¯h,N)\lambda_{\mathrm{rel}}(\bar{M}_{h},N) and λrel(M¯h,γ)\lambda_{\mathrm{rel}}(\bar{M}_{h},\gamma) are large for some embedded 2-sphere NN and embedded circle γ\gamma. Consider three solid tori T1T_{1}, T2T_{2}, and T3T_{3} embedded in S3S^{3} and linked as shown in fig. 6. The embedded 2-sphere NN is chosen so that it intersects T2T_{2} in two 2-disks. NN also separates S3S^{3} into two closed 2-balls, B1B_{1} and B2B_{2}, so that BiB_{i} contains TiT_{i} in its interior. γ\gamma is chosen so that it links with T1T_{1} as shown in the figure.

Similarly to our previous construction of the metric ghg_{h} from Section 2, give each solid torus the product metric on Dh×S1D_{h}\times S^{1}, where the length of the S1S^{1} is sufficiently small so that the volume of the product metric is at most 110\frac{1}{10}. Extend this metric on T1T2T3T_{1}\cup T_{2}\cup T_{3} to the entire S3S^{3} as in Section 2, so that it is sufficiently small away from some open neighbourhood of T1T2T3T_{1}\cup T_{2}\cup T_{3}. This defines a metric g¯h\bar{g}_{h} on S3S^{3} with diameter and volume at most 1. As a Riemannian submanifold, NN has diameter at most 1 due to our choice of metric on M¯h\bar{M}_{h}.

Refer to caption
Figure 6: NN separates S3S^{3} into two connected components, each containing one solid torus. The remaining solid torus intersects NN in two disks.

Recall from the Introduction that ΩNS3\Omega_{N}S^{3} denotes the space of piecewise smooth paths in S3S^{3} whose endpoints lie on NN. One can prove that

πi(ΩNS3,Λ0N)πi+1(S3,N),\pi_{i}(\Omega_{N}S^{3},\Lambda^{0}N)\cong\pi_{i+1}(S^{3},N), (5.1)

by generalizing the proof that πi(ΩS3)πi+1(S3)\pi_{i}(\Omega S^{3})\cong\pi_{i+1}(S^{3}). Thus we have π1(ΩNS3,Λ0N)=0\pi_{1}(\Omega_{N}S^{3},\Lambda^{0}N)=0 but π2(ΩNS3,Λ0N)0\pi_{2}(\Omega_{N}S^{3},\Lambda^{0}N)\neq 0.

Consider some continuous map f:(D2,D2)(ΩNS3,Λ0N)f:(D^{2},\partial D^{2})\to(\Omega_{N}S^{3},\Lambda^{0}N) that represents a nonzero class in π2(ΩNS3,Λ0N)\pi_{2}(\Omega_{N}S^{3},\Lambda^{0}N). Similar to the discussion in the Plan of the Proof, ff induces a map F:(D2×I/)D3S3F:(D^{2}\times I/{\sim})\simeq D^{3}\to S^{3}, where (p,r)(p,r)(p,r)\sim(p,r^{\prime}) for all pD2p\in\partial D^{2}. The quotient map q:D2×ID3q:D^{2}\times I\to D^{3} can be chosen so that it sends p×Ip\times I to D3(×{p})D^{3}\cap(\mathbb{R}\times\{p\}), where D3D^{3} is regarded as the unit 3-disk in 4\mathbb{R}^{4}. Define Dst1=D3(×{(s,t)})D^{1}_{st}=D^{3}\cap(\mathbb{R}\times\{(s,t)\}) and Dt2=D3(×{t})D^{2}_{t}=D^{3}\cap(\mathbb{R}\times\{t\}).

Suppose that every F(Dst1)F(D^{1}_{st}) is shorter than LL for some L>0L>0. By taking the double of FF to get a map S3S3S^{3}\to S^{3}, we may apply Lemma 4.1 to conclude that for any δ>0\delta>0 we can approximate FF by a smooth map F^:D3S3\hat{F}:D^{3}\to S^{3} that is transverse to Y=T1T2Y=\partial T_{1}\cup\partial T_{2} so that:

  1. I.

    For each ss and tt, the curves F(Dst1)F(D^{1}_{st}) and F^(Dst1)\hat{F}(D^{1}_{st}) differ in length by less than δ\delta.

  2. II.

    Each set F^1(Y)Dt2\hat{F}^{-1}(Y)\cap D^{2}_{t} is a level set of a Morse function on F^1(Y)\hat{F}^{-1}(Y).

In particular, each Xi=F^1(Ti)X_{i}=\hat{F}^{-1}(\partial T_{i}) is a closed submanifold of S3S^{3}, which we endow with the preimage orientation. Henceforth we will write FF to mean F^\hat{F}.

Lemma 5.1.

For some i=1,2i=1,2, the map F|Xi:XiTiF|_{X_{i}}:X_{i}\to\partial T_{i} has nonzero degree.

Proof.

The isomorphism π2(ΩNS3,Λ0N)π3(S3,N)\pi_{2}(\Omega_{N}S^{3},\Lambda^{0}N)\cong\pi_{3}(S^{3},N) from eq. 5.1 sends [f]0[f]\neq 0 to [F]0[F]\neq 0. The Hurewicz homomorphism φ:π3(S3,N)H3(S3,N)\varphi:\pi_{3}(S^{3},N)\to H_{3}(S^{3},N) sends [F][F] to a class αH3(S3,N)H~3(S3/N)\alpha\in H_{3}(S^{3},N)\cong\tilde{H}_{3}(S^{3}/N)\cong\mathbb{Z}\oplus\mathbb{Z}. Statement (II) above implies that α=(k1,k2)\alpha=(k_{1},k_{2})\in\mathbb{Z}\oplus\mathbb{Z}, where ki=degF|Xik_{i}=\deg F|_{X_{i}}. (To see this, note that the preimages of a regular value of FF have neighbourhoods that are foliated by the Dt2D^{2}_{t}’s.) Therefore it suffices to prove that φ\varphi is an isomorphism. Consider the following commutative diagram whose rows are long exact sequences and whose columns are Hurewicz maps:

π3(N){\pi_{3}(N)}π3(S3){\pi_{3}(S^{3})}π3(S3,N){\pi_{3}(S^{3},N)}π2(N){\pi_{2}(N)}π2(S3)0{\overbrace{\pi_{2}(S^{3})}^{0}}H3(N)0{\underbrace{H_{3}(N)}_{0}}H3(S3){H_{3}(S^{3})}H3(S3,N){H_{3}(S^{3},N)}H2(N){H_{2}(N)}H2(S3)0{\underbrace{H_{2}(S^{3})}_{0}}φ3\scriptstyle{\varphi_{3}}φ\scriptstyle{\varphi}φ2\scriptstyle{\varphi_{2}} (5.2)

φ2\varphi_{2} and φ3\varphi_{3} are isomorphisms by the Hurewicz theorem. The Five Lemma implies that φ\varphi is also an isomorphism. (We use a stronger version of the Five Lemma [28, p. 129], in which the leftmost column is only required to be surjective and the rightmost column is only required to be injective.) ∎

Proof of Theorem 1.3.

Consider any E>0E>0. The Hölder inequality implies that length(γ)E(γ)\operatorname{length}(\gamma)\leq\sqrt{E(\gamma)} for piecewise smooth curves γ\gamma parametrized over II. Lemma 2.1 allows us to choose some hh such that width11(Dh)>E+1\operatorname{width}_{1}^{1}(D_{h})>\sqrt{E}+1. Consider the Riemannian 3-sphere M=(S3,g¯h)M=(S^{3},\bar{g}_{h}) defined at the beginning of this section.

Let us first prove that λrel(M,N)>E\lambda_{\mathrm{rel}}(M,N)>E. Suppose for the sake of contradiction that λrel(M,N)E\lambda_{\mathrm{rel}}(M,N)\leq E. Then some nonzero class in π2(ΩNM,N)\pi_{2}(\Omega_{N}M,N) is represented by a map f:(D2,D2)(ΩNM,N)f:(D^{2},\partial D^{2})\to(\Omega_{N}M,N) such that each f(p)f(p) is shorter than E+12\sqrt{E}+\frac{1}{2}. As explained previously, Lemma 4.1 implies that ff corresponds to a map FF satisfying statements (I) and (II), and so that every curve F(Dst1)F(D^{1}_{st}) is shorter than E+12\sqrt{E}+\frac{1}{2}.

Without loss of generality, Lemma 5.1 implies that F|X1:X1T1F|_{X_{1}}:X_{1}\to\partial T_{1} has nonzero degree. An argument adapted from the proof of Lemma 4.3 implies that for some tt, X1Dt2X_{1}\cap D^{2}_{t} contains some embedded circle CC such that F|C:CT1F|_{C}:C\to\partial T_{1} is not nullhomotopic. Thus F(C)F(C) winds around the core curve of TjT_{j} a nonzero number of times, where j=1j=1 or 3. (Here we used the fact that T1T_{1} and T3T_{3} are linked.) Let πj:TjDh\pi_{j}:T_{j}\to D_{h} denote the canonical projection. The proof of Lemma 4.2 works nearly verbatim to prove that F1(Tj)Dt2F^{-1}(T_{j})\cap D^{2}_{t} contains a surface Σ\Sigma such that (πjF):H2(Σ,Σ)H2(Dh,Dh)(\pi_{j}\circ F)_{*}:H_{2}(\Sigma,\partial\Sigma)\to H_{2}(D_{h},\partial D_{h}) is a nonzero map. (The crucial fact is that CC bounds a disk in Dt2D^{2}_{t}.)

As in the proof of Theorem 1.1, DhD_{h} is swept out by a family of relative 1-cycles πj(F(Dst1Σ))\pi_{j}(F(D^{1}_{st}\cap\Sigma)) so as a consequence of Lemma 2.1, one of the curves F(Dst1)F(D^{1}_{st}) must be longer than E+1\sqrt{E}+1, giving a contradiction.

Next we prove that λrel(M,γ)>E\lambda_{\mathrm{rel}}(M,\gamma)>E. Note that π1(ΩγM,Λ0γ)π2(M,γ)0\pi_{1}(\Omega_{\gamma}M,\Lambda^{0}\gamma)\cong\pi_{2}(M,\gamma)\neq 0, so we consider a map of pairs h:(I,I)(ΩγM,Λ0γ)h:(I,\partial I)\to(\Omega_{\gamma}M,\Lambda^{0}\gamma) that represents a nonzero class in π1(ΩγM,Λ0γ)\pi_{1}(\Omega_{\gamma}M,\Lambda^{0}\gamma). Suppose for the sake of contradiction that every curve h(p)h(p) has length at most E+12\sqrt{E}+\frac{1}{2}. Similar to previous arguments, hh induces a map H:D2(I×I/)MH:D^{2}\simeq(I\times I/{\sim})\to M, where (p,r)(p,r)(p,r)\sim(p,r^{\prime}) for all pIp\in\partial I. The quotient map I×ID2I\times I\to D^{2} can be chosen to send {t}×I\{t\}\times I to It=D2(×{t})I_{t}=D^{2}\cap(\mathbb{R}\times\{t\}). HH represents a nonzero class in π2(M,γ)\pi_{2}(M,\gamma), and each H(It)H(I_{t}) has length at most E+12\sqrt{E}+\frac{1}{2}.

The long exact sequence of homotopy groups of the pair (M,γ)(M,\gamma) reveals that the boundary map π2(M,γ)π1(γ)\pi_{2}(M,\gamma)\to\pi_{1}(\gamma) is an isomorphism, so H|D2H|_{\partial D^{2}} winds around γ\gamma a nonzero number of times. Since γ\gamma is linked with the core curve of T1T_{1}, the surface H(D2)H(D^{2}) has a nonzero intersection number with that core curve. Similarly to the previous arguments, HH can be perturbed to a homotopic map that is transverse to T1\partial T_{1} while changing the lengths of the curves H(It)H(I_{t}) only slightly. The argument used to prove Lemma 4.2 implies the existence of some surface ΣH1(T1)\Sigma\subset H^{-1}(T_{1}) such that (π1H):H2(Σ,Σ)H2(Dh,Dh)(\pi_{1}\circ H)_{*}:H_{2}(\Sigma,\partial\Sigma)\to H_{2}(D_{h},\partial D_{h}) is a nonzero map. As before, DhD_{h} is now swept out by relative 1-cycles πj(H(ItΣ))\pi_{j}(H(I_{t}\cap\Sigma)), and similar arguments as before show that one of the curves H(It)H(I_{t}) must be longer than E+1\sqrt{E}+1, giving a contradiction. ∎

Acknowledgements

The authors would like to thank Regina Rotman for suggesting the problem to us, and for helpful conversations. The first author was supported by the University of Toronto Excellence Award. The second author was supported by the Vanier Canada Graduate Scholarship.

References

  • [1] F. J. Almgren, Jr. The homotopy groups of the integral cycle groups. Topology, 1:257–299, 1962.
  • [2] F. Balacheff and S. Sabourau. Diastolic and isoperimetric inequalities on surfaces. Ann. Sci. Éc. Norm. Supér. (4), 43(4):579–605, 2010.
  • [3] W. Ballmann, G. Thorbergsson, and W. Ziller. Existence of closed geodesics on positively curved manifolds. J. Differential Geometry, 18(2):221–252, 1983.
  • [4] I. Beach. Short simple geodesic loops on a 2-sphere, 2024. Preprint.
  • [5] I. Beach. Short simple orthogonal geodesic chords on a 2-disk with convex boundary, 2024.
  • [6] I. Beach, H. C. Peruyero, E. Griffin, M. Kerr, R. Rotman, and C. Searle. Lengths of the orthogonal geodesic chords on riemannian manifolds, 2024.
  • [7] W. Bos. Kritische Sehnen auf Riemannschen Elementarraumstücken. Math. Ann., 151:431–451, 1963.
  • [8] R. Bott. Lectures on Morse theory, old and new. In Proceedings of the 1980 Beijing Symposium on Differential Geometry and Differential Equations, Vol. 1, 2, 3 (Beijing, 1980), pages 169–218. Science Press, Beijing, 1982.
  • [9] D. Burago and S. Ivanov. On asymptotic isoperimetric constant of tori. Geom. Funct. Anal., 8(5):783–787, 1998.
  • [10] H. Y. Cheng. Curvature-free linear length bounds on geodesics in closed Riemannian surfaces. Trans. Amer. Math. Soc., 375(7):5217–5237, 2022.
  • [11] O. Chodosh and C. Mantoulidis. The pp-widths of a surface. Publ. Math. Inst. Hautes Études Sci., 137:245–342, 2023.
  • [12] F. Codá Marques. Minimal surfaces: variational theory and applications. In Proceedings of the International Congress of Mathematicians—Seoul 2014. Vol. 1, pages 283–310. Kyung Moon Sa, Seoul, 2014.
  • [13] F. Codá Marques and A. Neves. Existence of infinitely many minimal hypersurfaces in positive Ricci curvature. Invent. Math., 209(2):577–616, 2017.
  • [14] C. B. Croke. Area and the length of the shortest closed geodesic. J. Differential Geom., 27(1):1–21, 1988.
  • [15] C. B. Croke and M. Katz. Universal volume bounds in Riemannian manifolds. In Surveys in differential geometry, Vol. VIII (Boston, MA, 2002), volume 8 of Surv. Differ. Geom., pages 109–137. Int. Press, Somerville, MA, 2003.
  • [16] H. Federer and W. H. Fleming. Normal and integral currents. Ann. of Math. (2), 72:458–520, 1960.
  • [17] W. H. Fleming. Flat chains over a finite coefficient group. Trans. Amer. Math. Soc., 121:160–186, 1966.
  • [18] S. Frankel and M. Katz. The morse landscape of a riemannian disk. In Annales de l’institut Fourier, volume 43, pages 503–507, 1993.
  • [19] R. Giambò, F. Giannoni, and P. Piccione. Orthogonal geodesic chords, brake orbits and homoclinic orbits in Riemannian manifolds. Adv. Differential Equations, 10(8):931–960, 2005.
  • [20] P. Glynn-Adey and Y. Liokumovich. Width, Ricci curvature, and minimal hypersurfaces. J. Differential Geom., 105(1):33–54, 2017.
  • [21] P. Glynn-Adey and Z. Zhu. Subdividing three-dimensional Riemannian disks. J. Topol. Anal., 9(3):533–550, 2017.
  • [22] M. Gromov. Filling Riemannian manifolds. J. Differential Geom., 18(1):1–147, 1983.
  • [23] M. Gromov. Asymptotic invariants of infinite groups. Technical report, P00001028, 1992.
  • [24] L. Guth. The width-volume inequality. Geom. Funct. Anal., 17(4):1139–1179, 2007.
  • [25] L. Guth. Metaphors in systolic geometry. Preprint, 2010. https://arxiv.org/abs/1003.4247.
  • [26] L. Guth and Y. Liokumovich. Parametric inequalities and weyl law for the volume spectrum. Geometry & Topology. To appear.
  • [27] J. Hass and P. Scott. Shortening curves on surfaces. Topology, 33(1):25–43, 1994.
  • [28] A. Hatcher. Algebraic topology. Cambridge University Press, Cambridge, 2002.
  • [29] D. Ko. Existence and morse index of two free boundary embedded geodesics on riemannian 2-disks with convex boundary, 2023. Preprint.
  • [30] J. M. Lee. Introduction to smooth manifolds, volume 218 of Graduate Texts in Mathematics. Springer, New York, second edition, 2013.
  • [31] M. M.-C. Li. Free boundary minimal surfaces in the unit ball: recent advances and open questions. In Proceedings of the International Consortium of Chinese Mathematicians 2017, pages 401–435. Int. Press, Boston, MA, [2020] ©2020.
  • [32] Y. Liokumovich. Spheres of small diameter with long sweep-outs. Proceedings of the American Mathematical Society, 141(1):309–312, 2013.
  • [33] Y. Liokumovich. Slicing a 2-sphere. J. Topol. Anal., 6(4):573–590, 2014.
  • [34] Y. Liokumovich. Surfaces of small diameter with large width. Journal of Topology and Analysis, 6(03):383–396, 2014.
  • [35] Y. Liokumovich. Families of short cycles on Riemannian surfaces. Duke Math. J., 165(7):1363–1379, 2016.
  • [36] Y. Liokumovich and D. Maximo. Waist inequality for 3-manifolds with positive scalar curvature. In Perspectives in scalar curvature. Vol. 2, pages 799–831. World Sci. Publ., Hackensack, NJ, [2023] ©2023.
  • [37] Y. Liokumovich, A. Nabutovsky, and R. Rotman. Contracting the boundary of a riemannian 2-disc. Geometric and Functional Analysis, 25:1543–1574, 2015.
  • [38] Y. Liokumovich and X. Zhou. Sweeping out 3-manifold of positive Ricci curvature by short 1-cycles via estimates of min-max surfaces. Int. Math. Res. Not. IMRN, (4):1129–1152, 2018.
  • [39] L. Lyusternik and L. Schnirelmann. Topological methods in variational problems and their application to the differential geometry of surfaces. Uspehi Matem. Nauk (N.S.), 2(1(17)):166–217, 1947.
  • [40] L. A. Lyusternik and A. I. Fet. Variational problems on closed manifolds. Doklady Akad. Nauk SSSR (N.S.), 81:17–18, 1951.
  • [41] J. Milnor. Morse theory, volume No. 51 of Annals of Mathematics Studies. Princeton University Press, Princeton, NJ, 1963. Based on lecture notes by M. Spivak and R. Wells.
  • [42] A. Nabutovsky. Linear bounds for constants in Gromov’s systolic inequality and related results. Geom. Topol., 26(7):3123–3142, 2022.
  • [43] A. Nabutovsky and R. Rotman. The length of the shortest closed geodesic on a 2-dimensional sphere. Int. Math. Res. Not., (23):1211–1222, 2002.
  • [44] A. Nabutovsky and R. Rotman. Upper bounds on the length of a shortest closed geodesic and quantitative Hurewicz theorem. J. Eur. Math. Soc. (JEMS), 5(3):203–244, 2003.
  • [45] A. Nabutovsky and R. Rotman. Linear bounds for lengths of geodesic loops on Riemannian 2-spheres. J. Differential Geom., 89(2):217–232, 2011.
  • [46] A. Nabutovsky and R. Rotman. Length of geodesics and quantitative Morse theory on loop spaces. Geom. Funct. Anal., 23(1):367–414, 2013.
  • [47] A. Nabutovsky, R. Rotman, and S. Sabourau. Sweepouts of closed Riemannian manifolds. Geom. Funct. Anal., 31(3):721–766, 2021.
  • [48] A. Neves. New applications of min-max theory. In Proceedings of the International Congress of Mathematicians—Seoul 2014. Vol. II, pages 939–957. Kyung Moon Sa, Seoul, 2014.
  • [49] A. Oancea. Morse theory, closed geodesics, and the homology of free loop spaces. In Free loop spaces in geometry and topology, volume 24 of IRMA Lect. Math. Theor. Phys., pages 67–109. Eur. Math. Soc., Zürich, 2015. With an appendix by Umberto Hryniewicz.
  • [50] P. Papasoglu. Contracting thin disks. J. Topol. Anal., 11(4):965–970, 2019.
  • [51] P. Papasoglu and E. Swenson. A surface with discontinuous isoperimetric profile and expander manifolds. Geom. Dedicata, 206:43–54, 2020.
  • [52] H.-B. Rademacher. Upper bounds for the critical values of homology classes of loops. Manuscripta Math., 174(3-4):891–896, 2024.
  • [53] R. Rotman. The length of a shortest closed geodesic and the area of a 2-dimensional sphere. Proc. Amer. Math. Soc., 134(10):3041–3047, 2006.
  • [54] R. Rotman. Positive Ricci curvature and the length of a shortest periodic geodesic. J. Geom. Anal., 34(6):Paper No. 167, 27, 2024.
  • [55] S. Sabourau. Filling radius and short closed geodesics of the 2-sphere. Bull. Soc. Math. France, 132(1):105–136, 2004.
  • [56] S. Sabourau. Volume of minimal hypersurfaces in manifolds with nonnegative Ricci curvature. J. Reine Angew. Math., 731:1–19, 2017.
  • [57] J.-P. Serre. Homologie singulière des espaces fibrés. Applications. Ann. of Math. (2), 54:425–505, 1951.
  • [58] A. Treibergs. Estimates of volume by the length of shortest closed geodesics on a convex hypersurface. Invent. Math., 80(3):481–488, 1985.
  • [59] A. S. Švarc. Geodesic arcs on Riemann manifolds. Uspehi Mat. Nauk, 13(6(84)):181–184, 1958.
  • [60] N. Wu and Z. Zhu. Length of a shortest closed geodesic in manifolds of dimension four. J. Differential Geom., 122(3):519–564, 2022.
  • [61] X. Zhou. On the free boundary min-max geodesics. Int. Math. Res. Not. IMRN, (5):1447–1466, 2016.