1 Introduction

In this paper, we study three-dimensional manifolds with positive scalar curvature. The scalar curvature of a Riemannian manifold is the average of the Ricci curvatures which in turn is the average of the sectional curvatures. It can be determined more simply by taking the following limit:

$$\begin{aligned} \mathrm{Scal}(p)= \lim _{r\rightarrow 0} 30 \frac{{\text {Vol}}_{{\mathbb {E}}^3}(B(0,r))-{\text {Vol}}_{M^3}(B(p, r))}{r^2{\text {Vol}}_{{\mathbb {E}}^3}(B(0,r))}, \end{aligned}$$
(1)

where \({\text {Vol}}_{{\mathbb {E}}^3}(B(0,r))=(4/3) \pi r^3\) and \({\text {Vol}}_{M^3}(B(p, r))\) is the Hausdorff measure of the ball about p of radius r in our manifold, \(M^3\).

In [16], Gromov asks the following pair of deliberately vague questions which we paraphrase here: Given a class of Riemannian manifolds, \({\mathcal {B}}\), what is the weakest notion of convergence such that a sequence of manifolds, \(M_j \in {\mathcal {B}}\), subconverges to a limit\(M_\infty \in {\mathcal {B}}\)where now we will expand\({\mathcal {B}}\)to include singular metric spaces? What is this generalized class of singular metrics spaces that should be included in\({\mathcal {B}}\)? Gromov points out that when \({\mathcal {B}}\) is the class of Riemannian manifolds with nonnegative sectional curvature then the “best known” answer to this question is Gromov–Hausdorff convergence and the singular limit spaces are then Alexandrov spaces with nonnegative Alexandrov curvature. When \({\mathcal {B}}\) is the class of Riemannian manifolds with nonnegative Ricci curvature, one uses Gromov–Hausdorff and metric measure convergence to obtain limits which are metric measure spaces with generalized nonnegative Ricci curvature as in work of Cheeger–Colding [8]. Work towards defining classes of singular metric measure spaces with generalized notions of nonnegative Ricci has been completed by Ambrosio-Gigli-Savare, Lott-Villani, Sturm and others [1, 21, 28].

Gromov then writes that “the most tantalizing relation\({\mathcal {B}}\)is expressed with the scalar curvature by\(\mathrm{Scal}\ge k\)” [16]. Bamler [4] and Gromov [15] have proven that under \(C^0\) convergence to smooth Riemannian limits \(\mathrm{Scal}\ge 0\) is preserved. In order to find the weakest notion of convergence which preserves \(\mathrm{Scal}\ge 0\) in some sense, Gromov has suggested that one might investigate intrinsic flat convergence [16]. The intrinsic flat distance was first defined in work of the third author with Wenger [31], who also proved that for noncollapsing sequences of manifolds with nonnegative Ricci curvature, intrinsic flat limits agree with Gromov–Hausdorff and metric measure limits [30]. Intrinsic flat convergence is a weaker notion of convergence in the sense that there are sequences of manifolds with no Gromov–Hausdorff limit that have intrinsic flat limits, including Ilmanen’s example of a sequence of three spheres with positive scalar curvature [31]. The third author has investigated intrinsic flat limits of manifolds with nonnegative scalar curvature under additional conditions with Lee, Huang, LeFloch and Stavrov [17, 19, 20, 27]. These papers support Gromov’s suggestion in the sense that the limits obtained in these papers have generalized nonnegative scalar curvature.

Here we construct a sequence of Riemannian manifolds, \(M_j^{3}\), with positive scalar curvature that converges in the intrinsic flat, metric measure and Gromov–Hausdorff sense to a singular limit space, Y, which fails to satisfy (1) [Example 6.1]. In fact, the limit space is a sphere with a pulled thread:

$$\begin{aligned} Y= {{\mathbb {S}}}^3 / \sim \text { where } a\sim b \text { iff } a,b \in C, \end{aligned}$$
(2)

where C is one geodesic in \({\mathbb {S}}^3\) (see Sect. 4). The scalar curvature about the point \(p_0 =[C(t)]\) formed from the pulled thread is computed in Lemma 6.3 to be

$$\begin{aligned} \lim _{r\rightarrow 0} \,\,\frac{\,\,{\text {Vol}}_{{\mathbb {E}}^3}(B(0,r))-{\text {Vol}}_{M^3}(B(p, r))\,\,}{r^2{\text {Vol}}_{{\mathbb {E}}^3}(B(0,r))}\,\,=\,\,-\infty . \end{aligned}$$
(3)

In this sense, the limit space does not have generalized nonnegative scalar curvature.

We construct our sequence using a new method we call sewing developed in Propositions 3.13.3. Before we can sew the manifolds, the first two authors construct short tunnels between points in the manifolds building on prior work of Gromov–Lawson [12] and Schoen–Yau [32]. The details of this construction are in the Appendix. In a subsequent paper [7], we will extend this sewing technique to also provide examples whose limit spaces fail to satisfy the Scalar Torus Rigidity Theorem [12, 32] and the Positive Mass Rigidity Theorem [33]. These examples, all constructed using the sewing techniques developed in this paper, demonstrate that Gromov–Hausdorff and Intrinsic Flat limit spaces of noncollapsing sequences of manifolds with positive scalar curvature may fail to satisfy key properties of nonnegative scalar curvature.

In light of these counter examples and the aforementioned positive results towards Gromov’s conjecture, the third author has suggested in [26] to adapt the class \({\mathcal {B}}\). There it is proposed that the initial class of smooth Riemannian manifolds in \({\mathcal {B}}\) should have nonnegative scalar curvature, a uniform lower bound on volume (as assumed implicitly by Gromov), and also a uniform lower bound on the minimal area of a closed minimal surface in the manifold, \(\text {MinA}(M)\). The sequences of \(M_j^{3}\) we construct using our new sewing methods have positive scalar curvature and a uniform lower bound on volume, but \(\text {MinA}(M_j)\rightarrow 0\). Intuitive reasons as to why a uniform lower bound on \(\text {MinA}(M_j)\) is a natural condition are described in [26] along with a collection of related conjectures and open problems. Here we will simply propose the following possible revision of Gromov’s vague conjecture:

Conjecture 1.1

Suppose a sequence of Riemannian manifolds, \(M^3_j\), have

$$\begin{aligned} \mathrm{Scal}_j\ge 0, {\text {Vol}}(M_j)\ge V_0>0, \text { and } \text {MinA}(M_j)\ge A_0>0, \end{aligned}$$
(4)

and the sequence converges in the intrinsic flat sense, \(M_j {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} M_\infty \).

Then at every point \(p\in M_\infty \) we have

$$\begin{aligned} \lim _{r\rightarrow 0} \,\,\frac{\,\,{\text {Vol}}_{{\mathbb {E}}^3}(B(0,r))-{\text {Vol}}_{Y}(B(p, r))\,\,}{r^2{\text {Vol}}_{{\mathbb {E}}^3}(B(0,r))}\,\,\ge \,\, 0. \end{aligned}$$
(5)

This paper is part of the work towards Jorge Basilio’s doctoral dissertation at the CUNY Graduate Center conducted under the advisement of Professors Józef Dodziuk and Christina Sormani. We would like to thank Jeff Jauregui, Marcus Khuri, Sajjad Lakzian, Dan Lee, Raquel Perales, Conrad Plaut, and Catherine Searle for their interest in this work.

2 Background

In this section, we first briefly review Gromov–Lawson and Schoen–Yau’s work. We then review Gromov–Hausdorff, Metric Measure, and Intrinsic Flat Convergence covering the key definitions as well as theorems applied in this paper to prove our example converges with respect to all three notions of convergence.

2.1 Gluing Gromov–Lawson and Schoen–Yau Tunnels

Using different techniques, Gromov–Lawson and Schoen–Yau described how to construct tunnels diffeomorphic to \({{\mathbb {S}}}^2 \times [0,1]\) with metric tensors of positive scalar curvature that can be glued smoothly into three-dimensional spheres of constant sectional curvature [12, 32]. See Fig. 1. These tunnels are the first crucial piece for our construction.

Fig. 1
figure 1

The tunnel

Here we need to explicitly estimate the volume and diameter of these tunnels. So the first and second authors prove the following lemma in the appendix.

Lemma 2.1

Let \(0<\delta /2 < 1\). Given a complete Riemannian manifold, \(M^3\), that contains two balls \(B(p_i,\delta /2)\subset M^3\), \(i=1,2\), with constant positive sectional curvature \(K \in (0,1]\) on the balls, and given any \(\epsilon >0\), there exists a \(\delta _0>0\) sufficiently small so that we may create a new complete Riemannian manifold, \(N^3\), in which we remove two balls and glue in a cylindrical region, U, between them:

$$\begin{aligned} N^3=M^3 \setminus \left( B(p_1,\delta /2)\cup B(p_2,\delta /2)\right) \sqcup U, \end{aligned}$$
(6)

where \(U=U(\delta _0)\) has a metric of positive scalar curvature (See Fig. 1) with

$$\begin{aligned} {\text {Diam}}(U) \le h=h(\delta ), \end{aligned}$$
(7)

where

$$\begin{aligned} h(\delta )=O(\delta ), \end{aligned}$$
(8)

hence,

$$\begin{aligned} \lim _{\delta \rightarrow 0} h(\delta )=0 \text { uniformly for } K\in (0,1]. \end{aligned}$$
(9)

The collars \(C_i= B(p_i,\delta /2) \setminus B(p_i,\delta _0)\) identified with subsets of \(N^3\) have the original metric of constant curvature and the tunnel \(U'=U\setminus (C_1\cup C_2)\) has arbitrarily small diameter \(O(\delta _0)\) and volume \(O(\delta _0^3)\). Therefore with appropriate choice of \(\delta _0\), we have

$$\begin{aligned} (1-\epsilon ) 2{\text {Vol}}(B(p,\delta /2)) \le {\text {Vol}}(U) \le (1+\epsilon ) 2{\text {Vol}}(B(p,\delta /2)) \end{aligned}$$
(10)

and

$$\begin{aligned} (1-\epsilon ) {\text {Vol}}(M) \le {\text {Vol}}(N) \le (1+\epsilon ) {\text {Vol}}(M). \end{aligned}$$
(11)

We note that if \(M^3\) has positive scalar curvature then so does \(N^3\) and that, after inserting the tunnel, \(\partial B(p_1,\delta /2)\) and \(\partial B(p_2,\delta /2)\) are arbitrarily close together because of (9). Note that we have restricted to three dimensions here and required constant sectional curvature on the balls for simplicity. The first two authors will generalize these conditions in future work. This lemma suffices for proving all the examples in this paper.

2.2 Review GH Convergence

Gromov introduced the Gromov–Hausdorff distance in [14].

First recall that \(\varphi : X \rightarrow Y\) is distance preserving iff

$$\begin{aligned} d_Y(\varphi (x_1), \varphi (x_2)) = d_X(x_1, x_2) \qquad \forall x_1, x_2 \in X. \end{aligned}$$
(12)

This is referred to as a metric isometric embedding in [19] and is distinct from a Riemannian isometric embedding.

Definition 2.2

(Gromov) The Gromov–Hausdorff distance between two compact metric spaces \(\left( X, d_X\right) \) and \(\left( Y, d_Y\right) \) is defined as

$$\begin{aligned} d_{GH}\left( X,Y\right) := \inf \, d^Z_H\left( \varphi \left( X\right) , \psi \left( Y\right) \right) , \end{aligned}$$
(13)

where Z is a complete metric space, and \(\varphi : X \rightarrow Z\) and \(\psi :Y\rightarrow Z\) are distance preserving maps and where the Hausdorff distance in Z is defined as

$$\begin{aligned} d_{H}^Z\left( A,B\right) = \inf \{ \epsilon >0: A \subset T_\epsilon \left( B\right) \text { and } B \subset T_\epsilon \left( A\right) \}. \end{aligned}$$
(14)

Gromov proved that this is indeed a distance on compact metric spaces: \(d_{GH}\left( X,Y\right) =0\) iff there is an isometry between X and Y. When studying metric spaces which are only precompact, one may take their metric completions before studying the Gromov–Hausdorff distance between them.

We write

$$\begin{aligned} X_j {\mathop {\longrightarrow }\limits ^{ \text {GH}}} X_\infty \,\,\,\text { iff }\,\,\, d_{GH}(X_j, X_\infty ) \rightarrow 0. \end{aligned}$$
(15)

Gromov proved that if \(X_j {\mathop {\longrightarrow }\limits ^{ \text {GH}}} X_\infty \) then there is a common compact metric space Z and distance preserving maps \(\varphi _j: X_j \rightarrow Z\) such that

$$\begin{aligned} d^Z_H(\varphi _j(X_j), \varphi _\infty (X_\infty ))\rightarrow 0. \end{aligned}$$
(16)

We say \(p_j \in X_j\) converges to \(p_\infty \in X_\infty \) if there is such a set of maps such that \(\varphi _j(p_j)\) converges to \(\varphi _\infty (p_\infty )\) as points in Z. These limits are not uniquely defined but they are useful and every point in the limit space is a limit of such a sequence in this sense.

Theorem 2.3

(Gromov) Suppose \(\epsilon _j \rightarrow 0\). If a sequence of metric spaces \((X_j, d_j)\) have \(\epsilon _j\) almost isometries

$$\begin{aligned} F_j: X_j \rightarrow X_\infty \end{aligned}$$
(17)

such that

$$\begin{aligned} |d_\infty (F_j(p), F_j(q)) - d_j(p,q)| \le \epsilon _j \qquad \forall p,q\in X_j \end{aligned}$$
(18)

and

$$\begin{aligned} X_\infty \subset T_{\epsilon _j}(F_j(X_j)) \end{aligned}$$
(19)

then

$$\begin{aligned} X_j {\mathop {\longrightarrow }\limits ^{ \text {GH}}} X_\infty . \end{aligned}$$
(20)

Note that \(p_j \in X_j\) converges to \(p_\infty \in X_\infty \) if \(F_j(p_j) \rightarrow p_\infty \in X_\infty \).

Gromov’s Compactness Theorem states that a sequence of manifolds with nonnegative Ricci (or Sectional) Curvature, and a uniform upper bound on diameter, has a subsequence which converges in the Gromov–Hausdorff sense to a geodesic metric space [14]. If a sequence of manifolds has nonnegative sectional curvature, then they satisfy the Toponogov Triangle Comparison Theorem. Taking the limits of the points in the triangles, one sees that the Gromov–Hausdorff limit of the sequence also satisfies the triangle comparison. Thus the limit spaces are Alexandrov spaces with nonnegative Alexandrov curvature (cf. [5]).

2.3 Review of Metric Measure Convergence

Fukaya introduced the notion of metric measure convergence of metric measure spaces \((X_j, d_j, \mu _j)\) in [10]. He assumed the sequence converged in the Gromov–Hausdorff sense as in (16) and then required that the push forwards of the measures converge as well,

$$\begin{aligned} \varphi _{j*}\mu _j \rightarrow \varphi _{\infty *}\mu _\infty \text { weakly as measures in } Z. \end{aligned}$$
(21)

Cheeger–Colding proved metric measure convergence of noncollapsing sequences of manifolds with Ricci uniformly bounded below in [8] where the measure on the limit is the Hausdorff measure. They proved metric measure convergence by constructing almost isometries and showing the Hausdorff measures of balls about converging points converge:

$$\begin{aligned} \text { If } p_j \rightarrow p_\infty \text { then } {\mathcal {H}}^m(B(p_j,r)) \rightarrow {\mathcal {H}}^m(B(p_\infty ,r)). \end{aligned}$$
(22)

They also studied collapsing sequences obtaining metric measure convergence to other measures on the limit space. Cheeger and Colding applied this metric measure convergence to prove that limits of manifolds with nonnegative Ricci curvature have generalized nonnegative Ricci curvature. In particular they prove the limits satisfy the Bishop–Gromov Volume Comparison Theorem and the Cheeger–Gromoll Splitting Theorem.

Sturm, Lott, and Villani then developed the CD(k,n) notion of generalized Ricci curvature on metric measure spaces in [21, 28]. In [29], Sturm extended the study of metric measure convergence beyond the consideration of sequences of manifolds which already converge in the Gromov–Hausdorff sense, using the Wasserstein distance. This is also explored in Villani’s text [34]. CD(k,n) spaces converge in this sense to CD(k,n) spaces. RCD(k,n) spaces developed by Ambrosio-Gigli-Savare are also preserved under this convergence [1]. RCD(k,n) spaces are CD(k,n) spaces which also require that the tangent cones almost everywhere are Hilbertian. There has been significant work studying both of these classes of spaces proving they satisfy many of the properties of Riemannian manifolds with lower bounds on their Ricci curvature.

2.4 Review of Integral Current Spaces

The Intrinsic Flat Distance is defined and studied in [31] by applying sophisticated ideas of Ambrosio–Kirchheim [2] extending earlier work of Federer–Fleming [9]. Limits of Riemannian manifolds under intrinsic flat convergence are integral current spaces, a notion introduced by the third author and Stefan Wenger in [31].

Recall that Federer–Fleming first defined the notion of an integral current as an extension of the notion of a submanifold of Euclidean space [9]. That is a submanifold \(\psi : M^m \rightarrow {\mathbb {E}}^N\) can be viewed as a current \(T=\psi _{\#}[M ]\) acting on m-forms as follows:

$$\begin{aligned} T(\omega )= \psi _{\#}[M ](\omega ) = [M ](\psi ^*\omega ) =\int _M \psi ^*\omega . \end{aligned}$$
(23)

If \(\omega = f\,d\pi _1 \wedge \cdots \wedge d\pi _m\) then

$$\begin{aligned} T(\omega )=\psi _{\#}[M ](\omega ) = \int _M f\circ \psi \,d(\pi _1\circ \psi ) \wedge \cdots \wedge d(\pi _m\circ \psi ). \end{aligned}$$
(24)

They define boundaries of currents as \(\partial T(\omega ) = T (d\omega )\) so that then the boundary of a submanifold with boundary is exactly what it should be. They define integer rectifiable currents more generally as countable sums of images under Lipschitz maps of Borel sets. The integral currents are integer rectifiable currents whose boundaries are integer rectifiable.

Ambrosio–Kirchheim extended the notion of integral currents to arbitrary complete metric space [2]. As there are no forms on metric spaces, they use deGeorgi’s tuples of Lipschitz functions,

$$\begin{aligned} T(f, \pi _1,\ldots , \pi _m)=\psi _{\#}[M ](f, \pi _1,\ldots , \pi _m)= \int _M f\circ \psi \,d(\pi _1\circ \psi ) \wedge \cdots \wedge d(\pi _m\circ \psi ). \end{aligned}$$
(25)

This integral is well defined because Lipschitz functions are differentiable almost everywhere. They define boundary as follows:

$$\begin{aligned} \partial T(f, \pi _1,\ldots , \pi _m)= T(1,f, \pi _1,\ldots , \pi _m) \end{aligned}$$
(26)

which matches with

$$\begin{aligned} d(f\,d\pi _1 \wedge \cdots \wedge d\pi _m) =1\,df \wedge d\pi _1 \wedge \cdots \wedge d\pi _m. \end{aligned}$$
(27)

They also define integer rectifiable currents more generally as countable sums of images under Lipschitz maps of Borel sets. The integral currents are integer rectifiable currents whose boundaries are integer rectifiable.

The notion of an integral current space was introduced in [31].

Definition 2.4

An m-dimensional integral current space, \(\left( X,d, T\right) \), is a metric space, (Xd) with an integral current structure \(T \in {{\mathbf {I}}}_m\left( {\bar{X}}\right) \) where \({\bar{X}}\) is the metric completion of X and \(\mathrm{set}(\hbox {T})=\hbox {X}\). Given an integral current space \(M=\left( X,d,T\right) \) we will use \(\mathrm{set}\left( \hbox {M}\right) \) or \(X_M\) to denote X, \(d_M=d\) and \([\![M ]\!]=T \). Note that \(\mathrm{set}\left( \partial \hbox {T}\right) \subset {\bar{\hbox {X}}}\). The boundary of \(\left( X,d,T\right) \) is then the integral current space:

$$\begin{aligned} \partial \left( X,d_X,T\right) := \left( \mathrm{set}\left( \partial \hbox {T}\right) , d_{{\bar{\mathrm{X}}}}, \partial \hbox {T}\right) . \end{aligned}$$
(28)

If \(\partial T=0\) then we say \(\left( X,d,T\right) \) is an integral current without boundary.

A compact-oriented Riemannian manifold with boundary, \(M^m\), is an integral current space, where \(X=M^m\), d is the standard metric on M and T is integration over M. In this case \({{\mathbf {M}}}(M)={\text {Vol}}(M)\) and \(\partial M\) is the boundary manifold. When M has no boundary, \(\partial M=0\).

Ambrosio–Kirchheim defined the mass \({{\mathbf {M}}}(T)\) and the mass measure ||T|| of a current in [2]. We apply the same notions to define a mass for an integral current space. Applying their theorems we have

$$\begin{aligned} {{\mathbf {M}}}(M)={{\mathbf {M}}}(T)=\int _X \theta _T(x)\lambda (x) d{\mathcal {H}}^m(x), \end{aligned}$$
(29)

where \(\lambda (x)\) is the area factor and \(\theta _T\) is the weight. In particular, \(\lambda (x)=1\) when the tangent cone at x is Euclidean which is true on a Riemannian manifold where the weight is also 1. This is true almost everywhere in the examples in this paper as well. The mass measure, ||T||, is a measure on X and satisfies

$$\begin{aligned} ||T||(A)=\int _A \theta _T(x)\lambda (x) d{\mathcal {H}}^m(x). \end{aligned}$$
(30)

2.5 Review of the Intrinsic Flat Distance

The Intrinsic Flat distance was defined in work of the third author and Stefan Wenger [31] as a new distance between Riemannian manifolds based upon the Federer–Fleming flat distance [9] and the Gromov–Hausdorff distance [14].

Recall that the Federer–Fleming flat distance between m-dimensional integral currents \(S,T\in {{\mathbf {I}}}_m\left( Z\right) \) is given by

$$\begin{aligned} d^Z_{F}\left( S,T\right) := \inf \{{{\mathbf {M}}}\left( U\right) +{{\mathbf {M}}}\left( V\right) : S-T=U+\partial V \}, \end{aligned}$$
(31)

where \(U\in {{\mathbf {I}}}_m\left( Z\right) \) and \(V\in {{\mathbf {I}}}_{m+1}\left( Z\right) \).

In [31], the third author and Wenger imitate Gromov’s definition of the Gromov–Hausdorff distance (which he called the intrinsic Hausdorff distance) by replaced the Hausdorff distance by the Flat distance:

Definition 2.5

(Sormani and Wenger [31]) For \(M_1=\left( X_1,d_1,T_1\right) \) and \(M_2=\left( X_2,d_2,T_2\right) \in {\mathcal {M}}^m\) let the intrinsic flat distance be defined:

$$\begin{aligned} d_{{{\mathcal {F}}}}\left( M_1,M_2\right) := \inf d_F^Z \left( \varphi _{1\#} T_1, \varphi _{2\#} T_2 \right) , \end{aligned}$$
(32)

where the infimum is taken over all complete metric spaces \(\left( Z,d\right) \) and distance preserving maps \(\varphi _1 : \left( {\bar{X}}_1,d_1\right) \rightarrow \left( Z,d\right) \) and \(\varphi _2: \left( {\bar{X}}_2,d_2\right) \rightarrow \left( Z,d\right) \) and the flat norm \(d_F^Z\) is taken in Z. Here \({\bar{X}}_i\) denotes the metric completion of \(X_i\) and \(d_i\) is the extension of \(d_i\) on \({\bar{X}}_i\), while \(\varphi _\# T\) denotes the push forward of T.

They then prove that this distance is 0 iff the spaces are isometric with a current preserving isometry. They say

$$\begin{aligned} M_j {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} M_\infty \text { iff } d_{{\mathcal {F}}}(M_j, M_\infty ) \rightarrow 0. \end{aligned}$$
(33)

And prove that this happens iff there is a complete metric space Z and distance preserving maps \(\varphi _j: M_j \rightarrow Z\) such that

$$\begin{aligned} d_F^Z(\varphi _{j\#}T_j,\varphi _{\infty \#}T_\infty ) \rightarrow 0. \end{aligned}$$
(34)

Note that in contrast to Gromov’s embedding theorem as stated in (16), the Z here is only complete and not compact.

There is a special integral current space called the zero space,

$$\begin{aligned} \mathbf{{0}}=(\emptyset , 0,0). \end{aligned}$$
(35)

Following the definition above, \(M_j {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} \mathbf{{0}}\) iff \(d_{{\mathcal {F}}}(M_j, \mathbf{{0}}) \rightarrow 0\) which implies there is a complete metric space Z and distance preserving maps \(\varphi _j: M_j \rightarrow Z\) such that

$$\begin{aligned} d_F^Z(\varphi _{j\#}T_j, 0) \rightarrow 0 \end{aligned}$$
(36)

Note that in this case the manifolds disappear and points have no limits.

Combining Gromov’s Embedding Theorem with Ambrosio–Kirchheim’s Compactness Theorem one has:

Theorem 2.6

(Sormani and Wenger [31]) Given a sequence of m-dimensional integral current spaces \(M_j=\left( X_j, d_j, T_j\right) \) such that \(X_j\) are equibounded and equicompact and with uniform upper bounds on mass and boundary mass. A subsequence converges in the Gromov–Hausdorff sense \(\left( X_{j_i}, d_{j_i}\right) {\mathop {\longrightarrow }\limits ^{ \text {GH}}} \left( Y,d_Y\right) \) and in the intrinsic flat sense \(\left( X_{j_i}, d_{j_i}, T_{j_i}\right) {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} \left( X,d,T\right) \), where either \(\left( X,d,T\right) \) is an m-dimensional integral current space with \(X \subset Y\) or it is the \(\mathbf{0}\) current space.

Note that in [30], the third author and Wenger prove if the \(M_j\) have nonnegative Ricci curvature then in fact the intrinsic flat and Gromov–Hausdorff limits agree. Matveev and Portegies have extended this to more general lower bounds on Ricci curvature in [22]. With only lower bounds on scalar curvature the limits need not agree as seen in the Appendix of [31]. There are also sequences of manifolds with nonnegative scalar curvature that have no Gromov–Hausdorff limit but do converge in the intrinsic flat sense (cf. Ilmanen’s example presented in [31] and also [18]).

In [35], Wenger proved that any sequence of Riemannian manifolds with a uniform upper bound on diameter, volume, and boundary volume has a subsequence which converges in the intrinsic flat sense to an integral current space (cf. [31]). It is possible that the limit space is just the \(\mathbf {0}\) space which happens for example when the volumes of the manifolds converge to 0.

Note that when \(M_j {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} M_\infty \) the masses are lower semicontinuous:

$$\begin{aligned} \liminf _{j\rightarrow \infty } {{\mathbf {M}}}(M_j) \ge {{\mathbf {M}}}(M_\infty ), \end{aligned}$$
(37)

where the mass of an integral current space is just the mass of the integral current structure. The mass is just the volume when M is a Riemannian manifold and can be computed using (29) otherwise. As there is not equality here, intrinsic flat convergence does not imply metric measure convergence.

In [23], Portegies has proven that when a sequence converges in the intrinsic flat sense and in addition \({{\mathbf {M}}}(M_j)\) is assumed to converge to \({{\mathbf {M}}}(M_\infty )\), then the spaces do converge in the metric measure sense, where the measures are taken to be the mass measures.

2.6 Useful Lemmas and Theorems Concerning Intrinsic Flat Convergence

The following lemmas, definitions, and theorems appear in work of the third author [25], although a few (labeled only as c.f. [25]) were used within proofs in older work of the third author with Wenger [30]. All are proven rigorously in [25].

Lemma 2.7

(c.f. Sormani [25]) A ball in an integral current space, \(M=\left( X,d,T\right) \), with the current restricted from the current structure of the Riemannian manifold is an integral current space itself,

(38)

for almost every \(r > 0\). Furthermore,

$$\begin{aligned} B(p,r) \subset \mathrm{set}(\hbox {S}(\hbox {p,r}))\subset {\bar{\hbox {B}}}(p,r)\subset \hbox {X}. \end{aligned}$$
(39)

Lemma 2.8

(c.f. Sormani [25]) When M is a Riemannian manifold with boundary

(40)

is an integral current space for all \(r > 0\).

Definition 2.9

(c.f. Sormani [25]) If \(M_i=(X_i, d_i,T_i) {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} M_\infty =(X_\infty , d_\infty ,T_\infty )\), then we say \(x_i\in X_i\) are a converging sequence that converge to \(x_\infty \in {\bar{X}}_\infty \) if there exists a complete metric space Z and distance preserving maps \(\varphi _i:X_i\rightarrow Z\) such that

$$\begin{aligned} \varphi _{i\#} T_i {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} \varphi _{\infty \#}T_\infty \text { and } \varphi _i(x_i) \rightarrow \varphi _\infty (x_\infty ). \end{aligned}$$
(41)

If we say collection of points, \(\{p_{1,i}, p_{2,i},\ldots ,p_{k,i}\}\), converges to a corresponding collection of points, \(\{p_{1,\infty }, p_{2,\infty },\ldots ,p_{k,\infty }\}\), if \(\varphi _{i}(p_{j,i}) \rightarrow \varphi _\infty (p_{j, \infty })\) for \(j=1,\ldots ,k\).

Definition 2.10

(c.f. Sormani [25] ) If \(M_i=(X_i, d_i,T_i) {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} M_\infty =(X_\infty , d_\infty ,T_\infty )\), then we say \(x_i\in X_i\) are Cauchy if there exists a complete metric space Z and distance preserving maps \(\varphi _i:M_i\rightarrow Z\) such that

$$\begin{aligned} \varphi _{i\#} T_i {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} \varphi _{\infty \#}T_\infty \text { and } \varphi _i(x_i) \rightarrow z_\infty \in Z. \end{aligned}$$
(42)

We say the sequence is disappearing if \(z_\infty \notin \varphi _\infty (X_\infty )\). We say the sequence has no limit in \({\bar{X}}_\infty \) if \(z_\infty \notin \varphi _\infty ({\bar{X}}_\infty )\).

Lemma 2.11

(c.f. Sormani [25]) If a sequence of integral current spaces, \(M_i=\left( X_i,d_i,T_i\right) \in {\mathcal {M}}_0^m\), converges to an integral current space, \(M=\left( X,d,T\right) \in {\mathcal {M}}_0^m\), in the intrinsic flat sense, then every point x in the limit space X is the limit of points \(x_i\in M_i\). In fact, there exists a sequence of maps \(F_i: X \rightarrow X_i\) such that \(x_i=F_i(x)\) converges to x and

$$\begin{aligned} \lim _{i\rightarrow \infty } d_i(F_i(x), F_i(y))= d(x,y). \end{aligned}$$
(43)

Lemma 2.12

(c.f. Sormani [25]) If \(M_j {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} M_\infty \) and \(p_j \rightarrow p_\infty \in {\bar{X}}_\infty \), then for almost every \(r_\infty >0\) there exists a subsequence of \(M_j\) also denoted \(M_j\) such that

(44)

are integral current spaces for \(j\in \{1,2,\ldots ,\infty \}\) and we have

$$\begin{aligned} S(p_j,r_\infty ) {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} S(p_\infty ,r_\infty ). \end{aligned}$$
(45)

If \(p_j\) are Cauchy with no limit in \({\bar{X}}_\infty \) then there exists \(\delta >0\) such that for almost every \(r \in (0,\delta )\) such that \(S(p_j,r)\) are integral current spaces for \(j\in \{1,2,\ldots \}\) and we have

$$\begin{aligned} S(p_j,r) {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} 0. \end{aligned}$$
(46)

If \(M_j {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} \mathbf {0}\) then for almost every r and for all sequences \(p_j\) we have (46).

Theorem 2.13

(c.f. Sormani [25]) Suppose \(M_i=(X_i, d_i, T_i)\) are integral current spaces and

$$\begin{aligned} M_i {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} M_\infty , \end{aligned}$$
(47)

and suppose we have Lipschitz maps into a compact metric space Z,

$$\begin{aligned} F_i: X_i \rightarrow Z \text { with } {\text {Lip}}(F_i)\le K, \end{aligned}$$
(48)

then a subsequence converges to a Lipschitz map

$$\begin{aligned} F_\infty : X_\infty \rightarrow Z \text { with }{\text {Lip}}(F_\infty )\le K. \end{aligned}$$
(49)

More specifically, there exists distance preserving maps of the subsequence, \(\varphi _i: X_i \rightarrow Z\), such that

$$\begin{aligned} d_F^Z(\varphi _{i\#} T_i , \varphi _\infty T_\infty )\rightarrow 0 \end{aligned}$$
(50)

and for any sequence \(p_i\in X_i\) converging to \(p\in X_\infty \) (i.e., \(d_Z(\varphi _i(p_i), \varphi _\infty (p))\rightarrow 0\)), we have

$$\begin{aligned} \lim _{i\rightarrow \infty }F_i(p_i)=F_\infty (p_\infty ). \end{aligned}$$
(51)

Theorem 2.14

(c.f. Sormani [25]) Suppose \(M^m_i=(X_i, d_i, T_i)\) are integral current spaces which converge in the intrinsic flat sense to a nonzero integral current space \(M^m_\infty =(X_\infty , d_\infty , T_\infty )\). Suppose there exists \(r_0>0\) and a sequence \(p_i \in M_i\) such that for almost every \(r\in (0, r_0)\) we have integral current spaces, \(S(p_i,r)\), for all \(i\in {\mathbb {N}}\) and

$$\begin{aligned} \liminf _{i\rightarrow \infty } d_{{\mathcal {F}}}(S(p_i,r),\mathbf{{0}}) =h_0>0. \end{aligned}$$
(52)

Then there exists a subsequence, also denoted \(M_i\), such that \(p_{i}\) converges to \(p_\infty \in {\bar{X}}_\infty \).

Theorem 2.15

(c.f. Sormani [25]) Let \(M_i=(X_i, d_i, T_i)\) and \(M'_i=(X'_i,d'_i,T_i)\) be integral current spaces with

$$\begin{aligned} {{\mathbf {M}}}(M_i)\le V_0 \text { and }{{\mathbf {M}}}(\partial M_i) \le A_0 \end{aligned}$$
(53)

such that

$$\begin{aligned} M_i {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} M_\infty \text { and } M'_i {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} M'_\infty . \end{aligned}$$
(54)

Fix \(\delta >0\). Let \(F_i: M_i \rightarrow M'_i\) be continuous maps which are isometries on balls of radius \(\delta \):

$$\begin{aligned} \forall x\in X_i, \,\, F_i: {\bar{B}}(x,\delta ) \rightarrow {\bar{B}}(F_i(x),r)\text { is an isometry} \end{aligned}$$
(55)

Then, when \(M_\infty \ne \mathbf{{0}}\), we have \(M'_\infty \ne \mathbf{{0}}\) and there is a subsequence, also denoted \(F_i\), which converges to a (surjective) local current preserving isometry.

$$\begin{aligned} F_\infty : {\bar{X}}_\infty \rightarrow {\bar{X}}'_\infty \text { satisfying} \ \mathrm{(55)}. \end{aligned}$$
(56)

More specifically, there exists distance preserving maps of the subsequence \(\varphi _i: X_i \rightarrow Z, \varphi '_i: X'_i \rightarrow Z'\), such that

$$\begin{aligned} d_F^Z(\varphi _{i\#} T_i , \varphi _\infty T_\infty )\rightarrow 0 \text { and } d_F^{Z'}(\varphi '_{i\#} T'_i , \varphi '_\infty T'_\infty )\rightarrow 0 \end{aligned}$$
(57)

and for any sequence \(p_i\in X_i\) converging to \(p\in X_\infty \):

$$\begin{aligned} \lim _{i\rightarrow \infty } \varphi _i(p_i)=\varphi _\infty (p) \in Z \end{aligned}$$
(58)

we have

$$\begin{aligned} \lim _{i\rightarrow \infty }\varphi _i'(F_i(p_i))=\varphi _\infty '(F_\infty (p_\infty )) \in Z'. \end{aligned}$$
(59)

When \(M_\infty =\mathbf{{0}}\) and \(F_i\) are surjective, we have \(M'_\infty =\mathbf{{0}}\).

3 Sewing Riemannian Manifolds with Positive Scalar Curvature

The main technique we will introduce in this paper is the construction of three-dimensional manifolds with positive scalar curvature through a process we call “sewing” which involved gluing a sequence of tunnels along a curve. We apply Lemma 2.1 which constructs Gromov–Lawson Schoen–Yau tunnels. The lemma is proven in the Appendix.

3.1 Gluing Tunnels Between Spheres

We begin by gluing tunnels between arbitrary collections of pairs of spheres as in Fig. 2.

Proposition 3.1

Given a complete Riemannian manifold, \(M^3\), and \(A_{0} \subset M^3\) a compact subset with an even number of points \(p_{i} \in A_{0}\), \(i = 1, \ldots , n\), with pairwise disjoint contractible balls \(B(p_i,\delta )\) which have constant positive sectional curvature K, for some \(\delta >0\), define \(A_{\delta } = T_{\delta }(A_{0})\) and

$$\begin{aligned} A_{\delta }' = A_{\delta } \Bigg \backslash \left( \bigcup _{i=1}^n B(p_i,\delta /2) \right) \sqcup \bigcup _{i=1}^{n/2} U_i, \end{aligned}$$
(60)

where \(U_i\) are the tunnels as in Lemma 2.1 connecting \(\partial B(p_{2j+1},\delta /2)\) to \(\partial B(p_{2j+2},\delta /2)\) for \(j=0,1,\ldots ,n/2-1\). Then given any \(\epsilon >0\), shrinking \(\delta \) further, if necessary, we may create a new complete Riemannian manifold, \(N^3\),

$$\begin{aligned} N^3 = (M^3 \setminus A_{\delta }) \sqcup A_{\delta }' \end{aligned}$$
(61)

satisfying

$$\begin{aligned} (1-\epsilon ){\text {Vol}}(A_{\delta }) \le {\text {Vol}}(A_{\delta }')\le {\text {Vol}}(A_{\delta })(1+\epsilon ) \end{aligned}$$
(62)

and

$$\begin{aligned} (1-\epsilon ){\text {Vol}}(M^3)\le {\text {Vol}}(N^3) \le {\text {Vol}}(M^3) (1+\epsilon ). \end{aligned}$$
(63)

If, in addition, \(M^3\) has nonnegative or positive scalar curvature, then so does \(N^3\). In fact,

$$\begin{aligned} \inf _{x\in M^3} \mathrm{Scal}_x \ge \min \left\{ 0, \inf _{x\in N^3} \mathrm{Scal}_x\right\} \end{aligned}$$
(64)

If \(\partial M^3 \ne \emptyset \), the balls avoid the boundary and \(\partial M^3\) is isometric to \(\partial N^3\).

Definition 3.2

We say that we have glued the manifold to itself with a tunnel between the collection of pairs of sphere \(\partial B(p_i,\delta )\) to \(\partial B(p_{i+1},\delta )\) for \(i=1\) to \(n-1\). See Fig. 2.

Fig. 2
figure 2

Gluing two spheres with a tunnel

Proof

For simplicity of notation, set \(A=A_{\delta }\) and \(A'=A_{\delta }'\).

By induction on n and Lemma 2.1, we see that \(N^{3}\) can be given a metric of positive scalar curvature whenever \(M^{3}\) has positive scalar curvature.

Using the fact that the balls are pairwise disjoint and of the same volume, and (10) from Lemma 2.1, we have the volume of \(A'\) can be estimated:

$$\begin{aligned} {\text {Vol}}(A')&= {\text {Vol}}(A) - \sum _{i=1}^{n} {\text {Vol}}(B(p_i,\delta /2)) + \sum _{i=1}^{n/2} {\text {Vol}}(U_i) \\&= {\text {Vol}}(A) + \frac{n}{2} \cdot ({\text {Vol}}(U_i) - 2 {\text {Vol}}(B(p_i,\delta /2)))\\&\le {\text {Vol}}(A) + \frac{n}{2} \cdot (2{\text {Vol}}(B(p_i,\delta /2)) \cdot \epsilon )\\&= {\text {Vol}}(A) + \epsilon \cdot (n{\text {Vol}}(B(p_i,\delta /2))) \qquad (\text {by } (20))\\&\le {\text {Vol}}(A) + \epsilon {\text {Vol}}(A) \end{aligned}$$

which yields the right-hand side of (62).

Similarly,

$$\begin{aligned} {\text {Vol}}(A')&= {\text {Vol}}(A) - \sum _{i=1}^{n} {\text {Vol}}(B(p_i,\delta /2)) + \sum _{i=1}^{n/2} {\text {Vol}}(U_i) \\&= {\text {Vol}}(A) + \frac{n}{2} \cdot ({\text {Vol}}(U_i) - 2 {\text {Vol}}(B(p_i,\delta /2))) \\&\ge {\text {Vol}}(A) + \frac{n}{2} \cdot (-2{\text {Vol}}(B(p_i,\delta /2)) \cdot \epsilon )\\&= {\text {Vol}}(A) - \epsilon \cdot (n{\text {Vol}}(B(p_i,\delta /2))) \qquad (\text {by } (20))\\&\ge {\text {Vol}}(A) - \epsilon {\text {Vol}}(A) \end{aligned}$$

which yields the left-hand side of (62).

To estimate the volume of N we will use the volume estimates for \(A'\). Using (10) from Lemma 2.1 again, we have

$$\begin{aligned} {\text {Vol}}(N)&= {\text {Vol}}(M) - {\text {Vol}}(A) + {\text {Vol}}(A') \\&\le {\text {Vol}}(M) - {\text {Vol}}(A) + (1+\epsilon ){\text {Vol}}(A)\\&= {\text {Vol}}(M) + \epsilon {\text {Vol}}(A) \qquad (\text {by } (11))\\&\le {\text {Vol}}(M) + \epsilon {\text {Vol}}(M), \end{aligned}$$

which yields the right-hand side of (63).

Similarly,

$$\begin{aligned} {\text {Vol}}(N)&= {\text {Vol}}(M) - {\text {Vol}}(A) + {\text {Vol}}(A') \\&\ge {\text {Vol}}(M) - {\text {Vol}}(A) + (1-\epsilon ){\text {Vol}}(A)\\&= {\text {Vol}}(M) -\epsilon {\text {Vol}}(A) \qquad (\text {by } (11))\\&\ge {\text {Vol}}(M) -\epsilon {\text {Vol}}(A), \end{aligned}$$

which yields the left-hand side of (63).

Finally, observe that (64) follows since Lemma 2.1 shows that the tunnels \(U_{i}\) have positive scalar curvature. \(\square \)

3.2 Sewing Along a Curve

We now describe our process we call sewing along a curve, where a sequence of balls is taken to be located along curve much like holes created when stitching a thread. We glue a sequence of tunnels to the boundaries of these balls as in Fig. 3. We say that we have sewn the manifold along the curve C through the given balls. By gluing tunnels in this precise way, we are able to shrink the diameter of the edited tubular neighborhood around the curve because travel along the curve can be conducted efficiently through the tunnels.

Fig. 3
figure 3

Sewing a manifold through eight balls along a curve

Proposition 3.3

Given a complete Riemannian manifold, \(M^3\), and \(A_{0}\subset M^3\) Riemannian isometric to an embedded curve, \(C:[0,1]\rightarrow {\mathbb {S}}^3_K\) possibly with \(C(0)=C(1)\) and parametrized proportional to arclength, in a standard sphere of constant sectional curvature K, define \(A_{a} = T_{a}(A_{0})\) as in Proposition 3.1 and assume that \(A_{a}\) is Riemannian isometric to \(T_{a}(C) \subset {\mathbb {S}}^3_K\). Then, given any \(\epsilon >0\) there exists n sufficiently large and \(\delta =\delta (\epsilon ,n,C,K)>0\) sufficiently small as in (66) so that we can “sew along the curve” to create a new complete Riemannian manifold \(N^3\),

$$\begin{aligned} N^3 = (M^3 \setminus A_{\delta } )\sqcup A_{\delta }', \end{aligned}$$
(65)

exactly as in Proposition 3.1, for

$$\begin{aligned} \delta =\delta (\epsilon ,n,C,K) \text { such that } \delta <a,\, \lim _{n\rightarrow \infty } n \cdot h(\delta )= 0, \text { and } \lim _{n\rightarrow \infty } n \cdot \delta = 0, \end{aligned}$$
(66)

where h is defined in Lemma 2.1 and the disjoint balls \(B(p_i,\delta )\) are to be centered at

$$\begin{aligned} p_{2j+1}=C\left( \frac{j}{n}+\frac{\delta }{L(C)}\right) \qquad p_{2j+2}=C\left( \frac{j+1}{n} -\frac{\delta }{L(C)}\right) \qquad j=0,1,\ldots ,n-1 \end{aligned}$$
(67)

and

$$\begin{aligned} A_{\delta }'= A_{\delta } \Bigg \backslash \left( \bigcup _{i=1}^{2n} B(p_i,\delta /2)\right) \sqcup \bigcup _{j=0}^{n-1} U_{2j+1}. \end{aligned}$$
(68)

Thus, the tunnels \(U_{2j+1}\) connect \(\partial B(p_{2j+1},\delta )\) to \(\partial B(p_{2j+2},\delta )\) for \(j=0,1,\ldots , n-1\).

Furthermore,

$$\begin{aligned} (1-\epsilon ){\text {Vol}}(A_{\delta }) \le {\text {Vol}}(A_{\delta }')\le {\text {Vol}}(A_{\delta })(1+\epsilon ) \end{aligned}$$
(69)

and

$$\begin{aligned} (1-\epsilon ) {\text {Vol}}(M^3)\le {\text {Vol}}(N^3) \le {\text {Vol}}(M^3) (1+\epsilon ) \end{aligned}$$
(70)

and

$$\begin{aligned} {\text {Diam}}(A_{\delta }')\le H(\delta )= L(C)/n + (n+1)\, h(\delta )+(5n+2)\, \delta . \end{aligned}$$
(71)

Since

$$\begin{aligned} \lim _{\delta \rightarrow 0} H(\delta )=0 \text { uniformly for } K\in (0,1], \end{aligned}$$
(72)

we say we have sewn the curve, \(A_{0}\), arbitrarily short.

If, in addition, \(M^3\) has nonnegative or positive scalar curvature, then so does \(N^3\). In fact,

$$\begin{aligned} \inf _{x\in M^3} \mathrm{Scal}_x \ge \min \left\{ 0, \inf _{x\in N^3} \mathrm{Scal}_x\right\} \end{aligned}$$
(73)

If \(\partial M^3 \ne \emptyset \), the balls avoid the boundary and \(\partial M^3\) is isometric to \(\partial N^3\).

Proof

By the fact that C is embedded, for n sufficiently large, the balls in the statement are disjoint even when \(C(0)=C(1)\) so we may apply Proposition 3.1 to get (69) and (70).

For simplicity of notation, let \(A=A_{\delta }\) and \(A'=A_{\delta }'\).

We now verify the diameter estimate of \(A'\), (71). To do this, we define sets \(C_i \subset A'\) which correspond to the sets \(\partial B(p_i,\delta /2) \subset A\) which are unchanged because they are the boundaries of the edited regions:

$$\begin{aligned} C_i \cup C_{i+1} = \partial U_{i}, \end{aligned}$$
(74)

whenever i is an odd value. Let

$$\begin{aligned} U=\bigcup _{j=0}^{n-1} U_{2j+1}. \end{aligned}$$
(75)

Let x and y be arbitrary points in \(A'\). We claim that there exists \(j,k \in \{1,\ldots ,2n\}\) such that

$$\begin{aligned} d_{A'}(x,C_j)< \delta + L(C)/(2n)+h(\delta ) \text { and } d_{A'}(y,C_k) < \delta + L(C)/(2n)+h(\delta ). \end{aligned}$$
(76)

By symmetry we need only prove this for x. Note that in case I where

$$\begin{aligned} x \in A' \setminus U = A \Bigg \backslash \bigcup _{i=1}^{2n} B(p_i,\delta /2) \end{aligned}$$
(77)

we can view x as a point in A. Let \(\gamma _1 \subset A\) be the shortest path from x to the closest point \(c_x \in C[0,1]\) so that \(L(\gamma _1) < \delta \).

If

$$\begin{aligned} \gamma _1 \cap B(p_j,\delta /2) \ne \emptyset \end{aligned}$$
(78)

then

$$\begin{aligned} d_{A'\setminus U}(x,C_j)<\delta \end{aligned}$$
(79)

and we have that (76) holds. Otherwise, still in Case I, if (78) fails then we have

$$\begin{aligned} d_{A'\setminus U}(x,C_j)\le & {} d_{A'\setminus U}(x,c_x) + d(c_x,C_j) \qquad \text {(by the triangle inequality)} \end{aligned}$$
(80)
$$\begin{aligned}< & {} \delta +\frac{L(C)}{2n}, \end{aligned}$$
(81)

where the last inequality follows from \(d_{A'\setminus U}(x,c_x) \le L(\gamma _1) <\delta \) and the fact that \(c_x \in C([0,1])\) is at most L(C) / (2n) away from the boundary of the nearest tunnel.

Alternatively, we have case II where \(x \in U\). In this case, there exists j such that \(x \in U_{2j+1}\) and so

$$\begin{aligned} d_{A'}(x,C_{2j+1}) \le {\text {Diam}}(U_{2j+1}) \le h(\delta ). \end{aligned}$$
(82)

Thus, we have the claim in (76).

We now proceed to prove (71) by estimating \(d_{A'}(x,y)\) for \(x,y \in A'\). If \(j=k\) in (76), then \(d_{A'}(x,y) \le 2(\delta + L(C)/(2n)+h(\delta ))\) and we are done. Otherwise, by (76) and the triangle inequality, we have

$$\begin{aligned} d_{A'}(x,y)\le & {} d_{A'}(x,C_j) + d_{A'}(y,C_k) + \sup \{ d_{A'}(z,w) \mid z \in C_j, w\in C_k \} \end{aligned}$$
(83)
$$\begin{aligned}\le & {} 2(\delta + L(C)/(2n)+h(\delta )) + \sup \{ d_{A'}(z,w) \mid z \in C_j, w\in C_k \}. \end{aligned}$$
(84)

Without loss of generality, we may assume that \(j<k\) and that j is odd. Thus, \(C_j \subset \partial U_j\). If k is also odd then by the triangle inequality

$$\begin{aligned} \sup \{ d_{A'}(z,w) \mid z \in C_j, w\in C_k \}\le & {} {\text {Diam}}(U_j) + {\text {dist}}(U_j,U_{j+2}) \nonumber \\&+ {\text {Diam}}(U_{j+2}) +\cdots +{\text {Diam}}(U_{k-2}) \nonumber \\&+ {\text {dist}}(U_{k-2},U_k) \end{aligned}$$
(85)

and, when k is even,

$$\begin{aligned} \sup \{ d_{A'}(z,w) \mid z \in C_j, w\in C_k \}\le & {} {\text {Diam}}(U_j) + {\text {dist}}(U_j,U_{j+2}) \nonumber \\&+ {\text {Diam}}(U_{j+2}) +\cdots +{\text {Diam}}(U_{k-2}) \nonumber \\&+ {\text {dist}}(U_{k-2},U_{k-1}) + {\text {Diam}}(U_{k-1}). \end{aligned}$$
(86)

We know that \({\text {Diam}}(U_j) = \cdots ={\text {Diam}}(U_k) \le h(\delta )\) from (7) of Lemma 2.1, and that the distance between any two adjacent tunnels is the same, and that there are at most n tunnels. Thus, in either case (85) or (86) we have

$$\begin{aligned} \sup \{ d_{A'}(z,w) \mid z \in C_j, w\in C_k \} \le n\, h(\delta ) + n\cdot {\text {dist}}(U_j,U_{j+2}). \end{aligned}$$
(87)

and by construction the distance between adjacent tunnels is

$$\begin{aligned} {\text {dist}}(U_j,U_{j+2})\le & {} {\text {Diam}}(C_{j+1}) + {\text {dist}}(C_{j+1},C_{j+2}) + {\text {Diam}}(C_{j+2}) \end{aligned}$$
(88)
$$\begin{aligned}\le & {} \pi (\delta /2) + \delta + \pi (\delta /2) < 5\delta \end{aligned}$$
(89)

since the balls \(B(p_i,\delta /2)\) have constant sectional curvature K.

Therefore, combining (84), (87), and (89) we conclude that

$$\begin{aligned} d_{A'}(x,y) \le 2(\delta + L(C)/(2n)+h(\delta ))+ n\, h(\delta ) + 5n \delta \end{aligned}$$
(90)

which is the desired diameter estimate (71).

We observe that by our choice of \(\delta \) satisfying (66) and the fact that \(h(\delta ) = O(\delta )\) from Lemma 2.1 we have that (72) holds.

Finally, observe that (73) follows since Lemma 2.1 shows that the tunnels \(U_{i}\) have positive scalar curvature. \(\square \)

Fig. 4
figure 4

A two sphere with the equator pulled to a point

4 Pulled String Spaces

The following notion of a pulled string metric space captures the idea that if a metric space is a patch of cloth and a curve in the patch is sewn with a string, then one can pull the string tight, identifying the entire curve as a single point, thus creating a new metric space. This notion was first described to the third author by Burago when they were working ideas related to [6]. See Fig. 4.

Proposition 4.1

The notion of a metric space with a pulled string is a metric space \((Y, d_Y)\) constructed from a metric space \((X,d_X)\) with a curve \(C:[0,1]\rightarrow X\), so that

$$\begin{aligned} Y = X \setminus C[0,1] \sqcup \{p_0\}, \qquad p_0=C(0), \end{aligned}$$
(91)

where for \(x_i \in Y\) we have

$$\begin{aligned} d_Y(x, p_0) = \min \{ d_X(x, C(t)) : \, t\in [0,1]\} \end{aligned}$$
(92)

and for \(x_i \in X \setminus C[0,1]\) we have

$$\begin{aligned} d_Y(x_1, x_2) =\min \left\{ \, d_X(x_1, x_2), \min \{d_X(x_1, C(t_1)) + d_X(x_2, C(t_2)): \, t_i \in [0,1] \}\, \right\} . \end{aligned}$$
(93)

If (XdT) is a Riemannian manifold then \((Y,d,\psi _\#T)\) is an integral current space whose mass measure is the Hausdorff measure on Y and

$$\begin{aligned} {\mathcal {H}}_Y^m(Y)={\mathcal {H}}_X^m(X)-{\mathcal {H}}_X^m(K). \end{aligned}$$
(94)

If \((X, d_X, T)\) is an integral current space then \((Y, d_Y, \psi _{\#}T)\) is also an integral current space where \(\psi : X\rightarrow Y\) such that \(\psi (x)=x\) for all \(x\in X\setminus C[0,1]\) and \(\psi (C(t))=p_0\) for all \(t\in [0,1]\). So that

$$\begin{aligned} {{\mathbf {M}}}(\psi _{\#}T)={{\mathbf {M}}}(T) \end{aligned}$$
(95)

We will in fact prove this proposition as a consequence of two lemmas about spaces with arbitrary compact subsets pulled to a point. Lemma 4.2 proves such a space is a metric space and Lemma 4.3 proves (94) and (95).

4.1 Pulled String Spaces Are Metric Spaces

Lemma 4.2

Given a metric space \((X, d_X)\) and a compact set \(K \subset X\), we may define a new metric space \((Y, d_Y)\) by pulling the set K to a point \(p_0 \in K\) by setting

$$\begin{aligned} Y := X \setminus K \sqcup \{p_0\}, \qquad p_0 \in K \, \text {fixed}, \end{aligned}$$
(96)

and, for \(x \in Y\), we have

$$\begin{aligned} d_Y(x, p_0) = \min \{ d_X(x, y) : \, y\in K\} \end{aligned}$$
(97)

and, for \(x_i \in Y \setminus \{p_0\}\), we have

$$\begin{aligned} d_Y(x_1, x_2) =\min \left\{ d_X(x_1, x_2), \min \{d_X(x_1, y_1) + d_X(x_2, y_2): \, y_i\in K\} \right\} . \end{aligned}$$
(98)

Proof

We first prove that \((Y,d_Y)\) is a metric space. By definition, it is easy to see that \(d_Y\) is nonnegative and symmetric. To prove that \(d_Y\) satisfies the axiom of positivity, assume \(x_1=x_2\). Then either \(x_i=p_0\), and \(d_Y(x_1,x_2)=0\) by Definitions (96)–(97), or \(x_i \ne p_0\) and \(d_X(x_1,x_2)=0\) so by (98) we have \(d_Y(x_1,x_2) \le d_X(x_1,x_2)=0\). Conversely, if \(d_Y(x_1,x_2)=0\) then either \(d_X(x_1,x_2)=0\) or

$$\begin{aligned} 0= \min \{ d_X(x_1,y_1)+d_X(x_2,y_2) \mid y_i \in K \}. \end{aligned}$$
(99)

In the first case, \(x_1=x_2\) since \(d_X\) is a metric, so assume otherwise. Then \(d_X(x_1,x_2) \ne 0\) and (99) holds. Being that (99) is a sum of nonnegative numbers, it follows that \(d_X(x_1,y_1)=0\) and \(d_X(x_2,y_2)=0\) for some \(y_i \in K\). Hence, \(x_i=y_i\) which is impossible by the definition of Y unless \(x_1=x_2=p_0\) which yields a contradiction. This proves that \(d_Y\) satisfies positivity.

Next, let us note that by virtue of (97) and (98), we always have

$$\begin{aligned} d_Y(x_1,x_2) \le d_X(x_1,x_2), \qquad \forall \, x_1,x_2 \in Y \end{aligned}$$
(100)

and

$$\begin{aligned} \text {if }d_Y(x_1,x_2) \ne d_X(x_1,x_2) \implies d_Y(x_1,x_2) = d_X(x_1,y_1)+d_X(x_2,y_2). \end{aligned}$$
(101)

for some \(y_i \in K\).

We now verify the triangle inequality: for any \(x_1, x_2, x_3 \in Y\), we need to prove

$$\begin{aligned} d_Y(x_1,x_2) \le d_Y(x_1,x_3) + d_Y(x_3,x_2). \end{aligned}$$
(102)

It will be convenient to define \(y_i \in K\) such that

$$\begin{aligned} d_X(x_i,y_i) = \min \{ d_X(x_i,y) \mid y \in K\} \text { for }i=1,2,3. \end{aligned}$$
(103)

Assume in Case I that \(d_Y(x_1,x_2) \ne d_X(x_1,x_2)\). Then by (101) and (103),

$$\begin{aligned} d_Y(x_1,x_2) = d_X(x_1,y_1)+d_X(x_2,y_2). \end{aligned}$$
(104)

We have three possibilities: (i) \(d_Y(x_1,x_3) \ne d_X(x_1,x_3)\) and \(d_Y(x_2,x_3) \ne d_X(x_2,x_3)\); (ii) \(d_Y(x_1,x_3) = d_X(x_1,x_3)\) and \(d_Y(x_2,x_3) = d_X(x_2,x_3)\); and (iii) (without loss of generality) \(d_Y(x_1,x_3) \ne d_X(x_1,x_3)\) and \(d_Y(x_2,x_3) = d_Y(x_2,x_3)\).

In Case I (i), we have

$$\begin{aligned} d_Y(x_1,x_2)= & {} d_X(x_1,y_1) + d_X(x_2,y_2) \qquad (\text {by} (104))\\\le & {} d_X(x_1,y_1) +d_X(x_3,y_3)+ d_X(x_2,y_2) + d_X(x_3,y_3)\\= & {} d_Y(x_1,x_3) + d_Y(x_2,x_3). \qquad (\text {by assumption (i)}, (101), \hbox {and} (103)) \end{aligned}$$

In Case I (ii), we have

$$\begin{aligned} d_Y(x_1,x_2)\le & {} d_X(x_1,x_2) \qquad (\text {by} (100))\\\le & {} d_X(x_1,x_3) + d_X(x_2,x_3) \\= & {} d_Y(x_1,x_3) + d_Y(x_2,x_3). \qquad \text {(by assumption (ii))} \end{aligned}$$

In Case I (iii), we have

$$\begin{aligned} d_X(x_2,y_2)= & {} \min \{ d_X(x_2,K) \mid y \in K\} \qquad (\text {by} (103)) \nonumber \\\le & {} d_X(x_2,y_3) \nonumber \\\le & {} d_X(x_2,x_3) + d_X(x_3,y_3) \end{aligned}$$
(105)
$$\begin{aligned}\le & {} d_Y(x_2,x_3) + d_X(x_3,y_3) \qquad \text {(by assumption (iii))} \end{aligned}$$
(106)

so that

$$\begin{aligned} d_Y(x_1,x_2)= & {} d_X(x_1,y_1) + d_X(x_2,y_2) \qquad (\text {by} (104)) \\\le & {} d_X(x_1,y_1)+ d_Y(x_2,x_3) + d_X(x_3,y_3) \qquad (\text {by} (106))\\= & {} d_Y(x_1,x_3)+d_Y(x_2,x_3). \qquad \text {(by assumption (iii))} \end{aligned}$$

This proves the triangle inequality, (102), in Case I. Next, we assume, in Case II, that \(d_Y(x_1,x_2) = d_X(x_1,x_2)\).

Again, we have three possibilities: (i) \(d_Y(x_1,x_3) \ne d_X(x_1,x_3)\) and \(d_Y(x_2,x_3) \ne d_X(x_2,x_3)\); (ii) \(d_Y(x_1,x_3) = d_X(x_1,x_3)\) and \(d_Y(x_2,x_3) = d_X(x_2,x_3)\); and (iii) (without loss of generality) \(d_Y(x_1,x_3) \ne d_X(x_1,x_3)\) and \(d_Y(x_2,x_3) = d_Y(x_2,x_3)\).

In Case II (i), we have

$$\begin{aligned} d_Y(x_1,x_2)&= d_X(x_1,x_2) \\&\le d_X(x_1,y_1)+d_X(x_2,y_2) \qquad (\text {by }(104)) \\&\le d_X(x_1,y_1)+d_X(x_3,y_3) +d_X(x_2,y_2)+d_X(x_3,y_3) \\&= d_Y(x_1,x_3) + d_Y(x_2,x_3). \qquad (\text {by assumption (i)}, (101),\hbox { and }(103)) \end{aligned}$$

In Case II (ii), (102) follows immediately from the triangle inequality for \(d_X\).

Finally, in Case II (iii),

$$\begin{aligned} d_Y(x_1,x_2)&= d_X(x_1,x_2) \\&\le d_X(x_1,y_1)+d_X(x_2,y_3) \qquad (\text {by } (104)) \\&\le d_X(x_1,y_1)+d_X(x_2,x_3) + d_X(x_3,y_3) \\&= d_Y(x_1,x_3) + d_Y(x_2,x_3), \qquad \text {(by assumption (iii)}, (101),\hbox { and }(103)) \end{aligned}$$

which completes the proof. \(\square \)

4.2 Hausdorff Measures and Masses of Pulled String Spaces

Lemma 4.3

If \((X, d_X, T)\) is an integral current space with a compact subset \(K \subset X\) then \((Y, d_Y, \psi _{\#}T)\) is also an integral current space where \((Y, d_Y)\) is defined as in Lemma 4.2 and where \(\psi : X\rightarrow Y\) such that \(\psi (x)=x\) for all \(x\in X\setminus K\) and \(\psi (q)=p_0\) for all \(q\in K\). In addition

$$\begin{aligned} {{\mathbf {M}}}(\psi _{\#}T)={{\mathbf {M}}}(T) - ||T||(K) \end{aligned}$$
(107)

If \((X,d_X,T)\) is a Riemannian manifold then \((Y,d_Y,\psi _\#T)\) is an integral current space whose mass measure is the Hausdorff measure on Y and

$$\begin{aligned} {\mathcal {H}}_Y^m(Y)={\mathcal {H}}_X^m(X)-{\mathcal {H}}_X^m(K). \end{aligned}$$
(108)

Proof

We must show that \((Y,d_Y,\psi _\# T)\) is an integral current space. We first observe that \(\psi \) as defined in the statement of the proposition is a 1-Lipschitz function: for \(x,y \in X\setminus K\), there is no ambiguity so we may view them as elements of \(Y\setminus \{p_0\}\) and \(d_Y(\psi (x),\psi (y))=d_Y(x,y) \le d_X(x,y)\) by definition of \(d_Y\). Otherwise, we may assume, without loss of generality, that \(x \in K\) and \(y \notin K\). In this case, \(d_Y(\psi (x),\psi (y))=d_Y(p_0,\psi (y))=d_Y(p_0,y)=\min \{d_X(z,y) : z \in K\} \le d_X(x,y)\), as \(x\in K\). Thus, \(\psi _\# T\) is an integral current on Y since \(\psi \) is a 1-Lipschitz function and the well-known inequality

$$\begin{aligned} \Vert \psi _\# T\Vert \le {\text {Lip}}(\psi )^m \Vert T\Vert \end{aligned}$$
(109)

implies that \(\psi _\# T\) has finite mass because T does. To show that \((Y,d_Y,\psi _\# T)\) is an integral current space there remains to show that it is completely settled, or \(\psi _\# T\) has positive density at \(p_0\).

Let \(f:Y \rightarrow {\mathbb {R}}\) be a bounded Lipschitz map and \(\pi _j:Y \rightarrow {\mathbb {R}}\) be Lipschitz maps. Then

$$\begin{aligned} (\psi _\# T)(f,\pi _1,\ldots ,\pi _m)&= T(f \circ \psi , \pi _1 \circ \psi , \ldots , \pi _m \circ \psi ) \\&= T(f\cdot {1}_{X\setminus K} + f(p_0)\cdot {1}_K, \pi _1 \circ \psi , \ldots , \pi _m \circ \psi )\\&= T(f\cdot {1}_{X\setminus K}, \pi _1 \circ \psi , \ldots , \pi _m \circ \psi )\\&\quad + f(p_0)T({1}_K, \pi _1 \circ \psi , \ldots , \pi _m \circ \psi )\\&= T(f\cdot {1}_{X\setminus K}, \pi _1 \circ \psi , \ldots , \pi _m \circ \psi ) + 0 \end{aligned}$$

by locality since \(\pi _i \circ \psi \) are constant on \(\{{1}_K \ne 0\}\) (see [2]) so

So, using the characterization of mass from [2], (2.6) of Proposition 2.7,

because \(\psi (x)=x\) on \(X\setminus K\), so since \({{\mathbf {M}}}(\cdot )=\Vert \cdot \Vert \),

where the supremum is taken over all Borel partitions \(\{A_{j}\}\) of X such that \(X = \cup _j A_j\) and all Lipschitz functions \(\pi _i^j \in {\text {Lip}}(X)\) with \({\text {Lip}}(\pi _i^j) \le 1\), then continuing

$$\begin{aligned} (\psi _\# T)(f,\pi _1,\ldots ,\pi _m)&=\sup \left\{ \sum _{j=1}^\infty |T({1}_{X\setminus K} \cdot {1}_{A_j},\pi _1^j,\ldots ,\pi _m^j)| \right\} \\&=\sup \left\{ \sum _{j=1}^\infty |T({1}_{\tilde{A}_j},{\tilde{\pi }}_1^j,\ldots ,{\tilde{\pi }}_m^j)| \right\} , \end{aligned}$$

where the second supremum is taken over all Borel partitions \(\{\tilde{A}_{j}\}\) of \(X\setminus K\) such that \(X\setminus K = \cup _j \tilde{A}_j\) and all Lipschitz functions \({\tilde{\pi }}_i^j \in {\text {Lip}}(X\setminus K)\) with \({\text {Lip}}({\tilde{\pi }}_i^j) \le 1\). So, by the characterization of mass we have

$$\begin{aligned} (\psi _\# T)(f,\pi _1,\ldots ,\pi _m)&=\sup \left\{ \sum _{j=1}^\infty |T({1}_{\tilde{A}_j},{\tilde{\pi }}_1^j,\ldots ,{\tilde{\pi }}_m^j)| \right\} \\&=||T||(X\setminus K)\\&= ||T||(X) - ||T|(K) \\&={{\mathbf {M}}}(T)-||T||(K), \end{aligned}$$

which proves (107).

Finally, assume that the m-dimensional integral current space \((X,d_X,T)\) is a Riemannian manifold. We show that the mass measure of \((Y,d_Y,\psi _\#T)\) is the Hausdorff measure on \((Y,d_Y)\).

We claim that

(110)

First, observe that since \(\psi \) is 1-Lipschitz,

$$\begin{aligned} {\mathcal {H}}_Y^m (\psi (X\setminus K)) \le ({\text {Lip}}(\psi ))^{m} {\mathcal {H}}_X^m(X\setminus K), \end{aligned}$$

by Proposition 3.1.4 on page 37 from [3], hence

$$\begin{aligned} {\mathcal {H}}_Y^m (Y\setminus \{p_0\}) \le {\mathcal {H}}_X^m(X\setminus K). \end{aligned}$$

Thus, there remains to show the opposite inequality in (110).

Define sets

$$\begin{aligned} C_{j} = \{ y \in Y \mid d_{Y}(y,p_{0}) \ge 1/j \} \end{aligned}$$

for each \(j \in {\mathbb {N}}\). Then the \(C_{j}\) are closed sets, \(C_{j} \subset C_{j+1}\) and \(Y\setminus \{p_{0}\} = \cup _{j \in {\mathbb {N}}} C_{j}\). So we may use Theorem 1.1.18 from [3]:

$$\begin{aligned} {\mathcal {H}}_Y^m (Y\setminus \{p_0\}) = {\mathcal {H}}_Y^m ( \cup _{j \in {\mathbb {N}}} C_{j}) = \lim _{j \rightarrow \infty } {\mathcal {H}}_Y^m (C_{j}). \end{aligned}$$
(111)

Consider, for each \(j \in N\),

$$\begin{aligned} D_{j} = \psi ^{-1}(C_{j}) = \{ x \in X \mid d_{X}(x,K) \ge 1/j \} \end{aligned}$$

which are closed in X, \(D_{j} \subset D_{j+1}\), and \(X \setminus K = \cup _{j \in {\mathbb {N}}} D_{j}\). Using Theorem 1.1.8 from [3] again:

$$\begin{aligned} {\mathcal {H}}_X^m (X\setminus K) = {\mathcal {H}}_X^m ( \cup _{j \in {\mathbb {N}}} D_{j}) = \lim _{j \rightarrow \infty } {\mathcal {H}}_X^m (D_{j}). \end{aligned}$$
(112)

Next, we claim that

$$\begin{aligned} {\mathcal {H}}_X^m (D_{j}) \le {\mathcal {H}}_Y^m(C_{j}), \qquad j \in {\mathbb {N}}. \end{aligned}$$
(113)

Fix j. Fix \(\delta < \frac{1}{2j}\). Let \(\{ E_{l} \}_{l \in {\mathbb {N}}}\) be a countable cover of \(C_{j}\) with \({\text {Diam}}(E_{l})<\delta \), for all l. Then

$$\begin{aligned} {\text {dist}}(E_{l},p_{0})>\frac{1}{2j}, \qquad l \in {\mathbb {N}}. \end{aligned}$$
(114)

To see this, assume otherwise. Then since \({\text {dist}}_{Y}(p_{0},E_{l})<\frac{1}{2j}\) and the definition of distance (as an infimum), there is \(e \in E_{l}\) such that \(d_{Y}(p_{0},e)<\frac{1}{2j}\). Now, we also know that \(E_{l} \cap C_{j} \ne \emptyset \). So, there is \(c \in C_{j} \cap E_{l}\). So, \(d_{Y}(e,c) \le {\text {Diam}}_{Y}(E_{l})< \delta < \frac{1}{2j}\). Also, by the triangle inequality, \(d_{Y}(p_{0},c) \le d_{Y}(p_{0},e)+d_{Y}(e,c) < 1/j\). But this contradicts that \(c \in C_{j}\) as by definition of \(C_{j}\), \(d_{Y}(p_{0},c)>1/j\).

Next, we show that

$$\begin{aligned} {\text {Diam}}_{Y}(E_{l}) = {\text {Diam}}_{X}(\psi ^{-1}(E_{l})), \end{aligned}$$
(115)

i.e., \(\psi ^{-1}\) is an isometry when restricted to \(\{E_{l}\}\). In fact, we prove

$$\begin{aligned} d_{X}(\psi ^{-1}(a),\psi ^{-1}(b)) = d_{Y}(a,b), \qquad \forall \, a,b \in E_{l}, j \in {\mathbb {N}}. \end{aligned}$$

Let \(a,b \in E_{l}\). Then since \({\text {Diam}}(E_{l})<\delta <\frac{1}{2j}\) we have \(d_{Y}(a,b) \le {\text {Diam}}_{Y}(E_{l})< \delta < \frac{1}{2j}\), so

$$\begin{aligned} d_{Y}(a,b) < \frac{1}{2j}. \end{aligned}$$
(116)

By definition of the distance \(d_{Y}\), since \(\psi ^{-1}(a)=a\) and \(\psi ^{-1}(b)=b\),

$$\begin{aligned} d_{Y}(a,b) = \min \big \{ d_{X}(a,b),\, \min \{\, d_{X}(a,k_{1})+d_{X}(b,k_{2}) \mid k_{i} \in K\}\, \big \}. \end{aligned}$$

If \(d_{Y}(a,b)=d_{X}(a,b)\), we’re done. If not, then there exists \(k_{1},k_{2} \in K\) so that

$$\begin{aligned} d_{Y}(a,b)=d_{X}(a,k_{1})+d_{X}(b,k_{2}). \end{aligned}$$
(117)

By (114),

$$\begin{aligned} d_{Y}(a,p_{0}) \ge \frac{1}{2j} \qquad \text {and} \qquad d_{Y}(b,p_{0}) \ge \frac{1}{2j} \end{aligned}$$

which implies

$$\begin{aligned} {\text {dist}}_{X}(a,K) \ge \frac{1}{2j} \qquad \text {and} \qquad {\text {dist}}_{X}(b,K) \ge \frac{1}{2j}. \end{aligned}$$

But then

$$\begin{aligned} \frac{1}{j}&\le {\text {dist}}_{X}(a,K) + {\text {dist}}_{X}(b,K) \\&\le d_{X}(a,k_{1})+d_{X}(b,k_{2}) \\&= d_{X}(a,b) \qquad (\text {by } (117)) \\&< \frac{1}{j}, \qquad (\text {by }(116)) \end{aligned}$$

which is a contradiction.

Next, observe that \(\{\psi ^{-1}(E_{l})\}_{l \in {\mathbb {N}}}\) is necessarily a cover of \(D_{j}\) so

$$\begin{aligned} {\mathcal {H}}_{X}^{m}(D_{j})&\le \sum _{l=1}^{\infty } \omega _{m} \left( \frac{{\text {Diam}}_{X}(\psi ^{-1}(E_{l}))}{2}\right) ^{m} \\&= \sum _{l=1}^{\infty } \omega _{m} \left( \frac{{\text {Diam}}_{Y}(E_{l})}{2}\right) ^{m}. \qquad (\text {by }(115))\\ \end{aligned}$$

Taking the infimum over all covers of \(C_{j}\) with diameters less than \(\delta \) gives

$$\begin{aligned} {\mathcal {H}}_{X}^{m}(D_{j}) \le {\mathcal {H}}_{Y,\delta }^{m}(C_{j}) \end{aligned}$$

then taking the limit as \(\delta \rightarrow 0\) shows

$$\begin{aligned} {\mathcal {H}}_{X}^{m}(D_{j}) \le {\mathcal {H}}_{Y}^{m}(C_{j}) \end{aligned}$$

which proves the claim (113).

To finish, we take the limit in (113) as \(j \rightarrow \infty \) and use (111) and (112) to complete the proof. \(\square \)

5 Sewn Manifolds Converging to Pulled Strings

In this section, we consider sequences of sewn manifolds being sewn increasingly tightly and prove they converge in the Gromov–Hausdorff and Intrinsic Flat sense to metric spaces with pulled strings.

To be more precise, we consider the following sequences of increasingly tightly sewn manifolds:

Definition 5.1

Given a single Riemannian manifold, \(M^3\), with a curve, \(A_0=C([0,1])\subset M\), with a tubular neighborhood \(A=T_a(A_0)\) which is Riemannian isometric to a tubular neighborhood of a compact set \(V \subset {\mathbb {S}}^3_K\), in a standard sphere of constant sectional curvature K, satisfying the hypothesis of Proposition 3.3. We can construct its sequence of increasingly tightly sewn manifolds, \(N_j^3\), by applying Proposition 3.3 taking \(\epsilon =\epsilon _j \rightarrow 0\), \(n=n_j \rightarrow \infty \), and \(\delta =\delta _j\rightarrow 0\) to create each sewn manifold, \(N^3=N_j^3\) and the edited regions \(A_{\delta }'=A_{\delta _{j}}'\) which we simply denote by \(A_{j}'\). This is depicted in Fig. 5. Since these sequences \(N_j^3\) are created using Proposition 3.3, they have positive scalar curvature whenever \(M^3\) has positive scalar curvature, and \(\partial N_j^3=\partial M^3\) whenever \(M^3\) has a nonempty boundary.

Fig. 5
figure 5

A sequence of increasingly tightly sewn manifolds

In this section, we prove Lemmas 5.5, 5.6, and 5.7 which immediately imply the following theorem:

Theorem 5.2

The sequence \(N_j^3\) as in Definition 5.1 converges in the Gromov–Hausdorff sense

$$\begin{aligned} N_j^{3} {\mathop {\longrightarrow }\limits ^{ \text {GH}}} N_\infty , \end{aligned}$$
(118)

the metric measure sense

$$\begin{aligned} N_j^{3} {\mathop {\longrightarrow }\limits ^{ \text {mGH}}} N_\infty , \end{aligned}$$
(119)

and the intrinsic flat sense

$$\begin{aligned} N_j^{3} {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} N_\infty , \end{aligned}$$
(120)

where \(N_\infty \) is the metric space created by pulling the string, \(A_0=C([0,1])\subset M\), to a point as in Proposition 4.1.

In fact, our lemmas concern more general sequences of manifolds which are constructed from a given manifold M and scrunch a given compact set \(K\subset M\) down to a point as follows:

Definition 5.3

Given a single Riemannian manifold, \(M^3\), with a compact set, \(A_0\subset M\). A sequence of manifolds,

$$\begin{aligned} N_j^3= (M^3 \setminus A_{\delta _j})\sqcup A'_{\delta _j} \end{aligned}$$
(121)

is said to scrunch \(A_0\) down to a point if \(A_{\delta }=T_{\delta }(A_0)\) and \(A'_\delta \) satisfies:

$$\begin{aligned} (1-\epsilon ){\text {Vol}}(A_{\delta }) \le {\text {Vol}}(A_{\delta }')\le {\text {Vol}}(A_{\delta })(1+\epsilon ) \end{aligned}$$
(122)

and

$$\begin{aligned} (1-\epsilon ) {\text {Vol}}(M^3)\le {\text {Vol}}(N^3) \le {\text {Vol}}(M^3) (1+\epsilon ) \end{aligned}$$
(123)

and

$$\begin{aligned} {\text {Diam}}(A_{\delta }')\le H \end{aligned}$$
(124)

where \(\epsilon =\epsilon _j \rightarrow 0\) and where \(H=H_j \rightarrow 0\) and \(2\delta _j<H_j\).

Note that by Proposition 3.3, a sequence of increasingly tightly sewn manifolds sewn along a curve C([0, 1]) as in Definition 5.1 is a sequence of manifolds which scrunches \(A_0=C([0,1])\) down to a point as in Definition 5.3. So we will prove lemmas about sequences of manifolds which scrunch a compact set and then apply them to prove Theorem 5.2 in the final subsection of this section.

5.1 Constructing Surjective Maps to the Limit Spaces

Before we prove convergence of the scrunched sequence of manifolds to the pulled thread space, we construct surjective maps from the sequence to the proposed limit space.

Lemma 5.4

Given \(M^3\) a compact Riemannian manifold (possibly with boundary) and a smooth embedded compact zero to three-dimensional submanifold \(A_0\subset M^3\) (possibly with boundary), and \(N_j\) as in Definition 5.3. Then for j sufficiently large there exist surjective Lipschitz maps

$$\begin{aligned} F_j: N_j^3 \rightarrow N_\infty \text { with } {\text {Lip}}(F_j) \le 4, \end{aligned}$$
(125)

where \(N_\infty \) is the metric space created by taking \(M^3\) and pulling \(A_0\) to a point \(p_0\) as in Lemmas 4.24.3.

Note that when \(A_0\) is the image of a curve, \(N_\infty \), is a pulled thread space as in Proposition 4.1.

Proof

First observe that by the construction in Definition 5.3 there are maps

$$\begin{aligned} P_j: M^3 \rightarrow N_\infty \end{aligned}$$
(126)

which are Riemannian isometries on regions which avoid \(A_0\) and map \(A_0\) to \(p_0\). These define Riemannian isometries

$$\begin{aligned} P_j: N_j^3 \setminus A_j' \,\tilde{=}\,M^3 \setminus T_{\delta _j}(A_{0}) \rightarrow N_\infty ^3 \setminus T_{\delta _j}(p_0). \end{aligned}$$
(127)

In addition, sufficiently small balls lying in these regions are isometric to convex balls in \(M^3\).

Observe also that for \(\delta >0\) sufficiently small, the exponential map:

$$\begin{aligned} exp: \{(p,v):\,\,p\in A_0,\,\, v\in V_p \,\,|v|<2\delta \} \rightarrow T_{2\delta }(A_0) \end{aligned}$$
(128)

is invertible where

$$\begin{aligned} V_p=\{v\in T_pM:\,\, d_M(exp_p(tv),p)=d_M(exp_p(tv),A_0)\}. \end{aligned}$$
(129)

Taking \(\delta =\delta _{A_0}>0\) even smaller (depending on the submanifold \(A_0\)), we can guarantee that \(\forall v_i\in V_p, |v_i|<2\delta _{A_0}, t_i\in (0,1)\) we have

$$\begin{aligned} d_M(exp_{p_1}(t_1v_1), exp_{p_2}(t_2v_2)) \le 2 d_M(exp_{p_1}(v_1), exp_{p_2}(v_2)) + 2|t_1-t_2|. \end{aligned}$$
(130)

This is not true unless \(A_0\) is a smooth embedded compact submanifold with either no boundary or a smooth boundary.

Define \(F_j: N_j^3 \rightarrow N_\infty \) as follows:

$$\begin{aligned} F_j(x)=P_j(x) \qquad \forall x \in N_j^3\setminus T_{\delta _j}(A_j') \end{aligned}$$
(131)

and

$$\begin{aligned} F_j(x)=p_0 \qquad \forall x\in A_j'. \end{aligned}$$
(132)

Between these two regions, we take

$$\begin{aligned} F_j(x)=f_j(P_j(x)) \qquad \forall x \in T_{\delta _j}(A_j')\setminus A_j', \end{aligned}$$
(133)

where \(f_j: N_\infty \rightarrow N_\infty \) is a surjective map:

$$\begin{aligned} f_j: Ann_{p_0}(\delta _j, 2\delta _j) \rightarrow B_{2\delta _j}(p_{0})\setminus \{p_0\} \end{aligned}$$
(134)

which takes a point q to

$$\begin{aligned} f_j(q)=\gamma _q\left( (d_{N_\infty }(p_0, q)- \delta _j) /\delta _j \right) , \end{aligned}$$
(135)

where \(\gamma _q\) is the unique minimal geodesic from \(\gamma _q(0)=p_0\) to \(\gamma _q(1)=q\). Here we are assuming \(\delta _j<\delta _{A_0}\). So

$$\begin{aligned} d_{N_\infty }(p_0, P_j(x))=d_{M^3}(A_0, x) \end{aligned}$$
(136)

and

$$\begin{aligned} \gamma _q(t)=P_j(exp_{q'}(tv')) \text { where } P_j(exp_{q'}(v'))=q. \end{aligned}$$
(137)

In particular for \(x\in \partial T_{\delta _j}(A_j')\),

$$\begin{aligned} f_j(P_j(x))=\gamma _{P_j(x)}((2\delta _j-\delta _j)/\delta _j)=\gamma _{P_j(x)}(1)=P_j(x) \end{aligned}$$
(138)

and for \(x\in \partial A_j'\),

$$\begin{aligned} f_j(P_j(x))=\gamma _{P_j(x)}((\delta _j-\delta _j)/\delta _j)=\gamma _{P_j(x)}(0)=p_0 \end{aligned}$$
(139)

so that \(F_j\) is continuous.

We claim

$$\begin{aligned} {\text {Lip}}(F_j)= & {} 0 \text { on } A_j' \end{aligned}$$
(140)
$$\begin{aligned} {\text {Lip}}(F_j)\le & {} 4 \text { on } T_{\delta _j}(A_j')\setminus A_j' \end{aligned}$$
(141)
$$\begin{aligned} {\text {Lip}}(F_j)= & {} 1\text { on } N_j \setminus T_{\delta _j}(A_j'). \end{aligned}$$
(142)

Only the middle part is difficult. By the definition of \(d_{N_\infty }\), we have the following two possibilities

$$\begin{aligned} \text {Case I:} \quad d_{N_\infty }(q_1,q_2)= & {} d_{M}(P_j^{-1}(q_1), P_j^{-1}(q_2)) \end{aligned}$$
(143)
$$\begin{aligned} \text {Case II:}\quad d_{N_\infty }(q_1, q_2)= & {} d_M(P_j^{-1}(q_1), A_0)+ d_M(P_j^{-1}(q_2), A_0). \end{aligned}$$
(144)

In Case II, we see that the minimal geodesic from \(q_1\) to \(q_2\) passes through \(p_0\). Since \(f_j(q_1)\) and \(f_j(q_2)\) lie on this geodesic, we have

$$\begin{aligned} d_{N_\infty }(f_j(q_1), f_j(q_2))\le d_{N_\infty }(q_1,q_2). \end{aligned}$$
(145)

In Case I, we apply (130) with

$$\begin{aligned} t_i=(d_{M}(P_j^{-1}(q_i),A_0)- \delta _j) /\delta _j \end{aligned}$$
(146)

because \(t_{i} \in (0,1)\) due to (141) so that by the reverse triangle inequality

$$\begin{aligned} |t_1-t_2|= & {} |d_{M}(P_j^{-1}(q_1),A_0)-d_{M}(P_j^{-1}(q_2),A_0)|/\delta _j \end{aligned}$$
(147)
$$\begin{aligned}\le & {} d_{M}(P_j^{-1}(q_1),q_2)/\delta _j \end{aligned}$$
(148)
$$\begin{aligned}\le & {} d_{N_\infty }(q_1,q_2) \end{aligned}$$
(149)

to see that

$$\begin{aligned} d_{N_\infty }(f_j(q_1), f_j(q_2))\le & {} d_{M}(P_j^{-1}(f_j(q_1)), P_j^{-1}(f_j(q_2))) \end{aligned}$$
(150)
$$\begin{aligned}\le & {} 2 d_M(P_j^{-1}(q_1),P_j^{-1}(q_2))+ 2|t_1-t_2| \text { by }(130), \end{aligned}$$
(151)
$$\begin{aligned}\le & {} 2 d_{N_\infty }(q_1,q_2) + 2|t_1-t_2| \text { by Case I hypothesis,} \end{aligned}$$
(152)
$$\begin{aligned}\le & {} 4d_{N_\infty }(q_1,q_2). \end{aligned}$$
(153)

This gives our claim.

We claim \({\text {Lip}}(F_j)\le 4\) everywhere. Given \(x_1,x_2\in N_j^3\), we have a minimizing geodesic \(\eta :[0,1]\rightarrow N_j\) such that \(\eta (0)=x_1\) and \(\eta (1)=x_2\). Then

$$\begin{aligned} d_{N_\infty }(F_j(x_1), F_j(x_2)) \le L(F_j \circ \eta ). \end{aligned}$$
(154)

Since \(|(F_j\circ \eta )'(t)| \le 2|\eta '(t)|\) by our localized Lipschitz estimates and because the function \(F_j\) is continuous, we are done. \(\square \)

5.2 Constructing Almost Isometries

See Sect. 2.2 for a review of the Gromov–Hausdorff distance.

Lemma 5.5

Given \(N_j^{3}\) as in Definition 5.3, the maps \(F_j: N_j^3 \rightarrow N_\infty \) defined in (131)–(133) in the proof of Lemma 5.4 are \(H_j\)-almost isometries with \(\lim _{j\rightarrow \infty }H_j=0\). Thus

$$\begin{aligned} N_j {\mathop {\longrightarrow }\limits ^{ \text {GH}}} N_\infty . \end{aligned}$$
(155)

Proof

Before we begin the proof recall that

$$\begin{aligned} {\text {Diam}}(A_j')\le H_j \rightarrow 0 \end{aligned}$$
(156)

in (124) of Definition 5.3.

By Theorem 2.3 of Gromov, to prove (155) it suffices to show that \(F_j\) are \(H_j\)-almost isometries. To see this, examine \(x,y\in N_j\) and join them by a minimizing curve \(\sigma :[0,1]\rightarrow N_j\).

If \(\sigma [0,1]\subset N_j \setminus A'_j\), then by (131) we have

$$\begin{aligned} L(\sigma )=L(F_j\circ \sigma ) \end{aligned}$$
(157)

and so

$$\begin{aligned} d_{N_j}(x,y)\ge d_{N_\infty }(F_j(x),F_j(y)). \end{aligned}$$
(158)

Otherwise we have

$$\begin{aligned} d_{N_j}(x,y)\ge & {} d_{N_j}(x, A'_j) + d_{N_j}(y, A'_j) \qquad T_{\delta _j}(A'_j) \text { to } A'_j \end{aligned}$$
(159)
$$\begin{aligned}= & {} d_{N_\infty }(F_j(x), B_{\delta _j}(p_0)) + d_{N_\infty }(F_j(y), B_{\delta _j}(p_0)) \end{aligned}$$
(160)
$$\begin{aligned}= & {} d_{N_\infty }(F_j(x), p_0)-\delta _j + d_{N_\infty }(F_j(y), p_0)-\delta _j \end{aligned}$$
(161)
$$\begin{aligned}\ge & {} d_{N_\infty }(F_j(x), F_j(y)) -2\delta _j. \end{aligned}$$
(162)

Next we join \(F_j(x)\) to \(F_j(y)\) by a minimizing curve \(\gamma \). If \(\gamma [0,1]\subset N_\infty \setminus B_{\delta _j}(p_0)\) then there is a curve \(\eta \) such that \(\gamma =F_j\circ \eta \) with \(\eta [0,1]\subset N_j \setminus A'_j\) and so by (131)

$$\begin{aligned} d_{N_j}(x,y)\le L(\eta )=L(\gamma ) =d_{N_\infty }(F_j(x), F_j(y)). \end{aligned}$$
(163)

Otherwise we have

$$\begin{aligned} d_{N_j}(x,y)\le & {} d_{N_j}(x, A_j')+{\text {Diam}}(A_j') + d_{N_j}(y, A_j') \end{aligned}$$
(164)
$$\begin{aligned}\le & {} d_{N_j}(x, A_j')+H_{j}+ d_{N_j}(y, A_j') \end{aligned}$$
(165)
$$\begin{aligned}= & {} d_{N_\infty }(F_j(x), B_{\delta _j}(p_0)) + d_{N_\infty }(F_j(y), B_{\delta _j}(p_0)) + H_j\end{aligned}$$
(166)
$$\begin{aligned}\le & {} L(\gamma ) +H_j =d_{N_\infty }(F_j(x),F_j(y)) + H_j. \end{aligned}$$
(167)

Hence, \(F_j\) is an \(H_j\) isometry since \(2\delta _{j}<H_{j}\). \(\square \)

5.3 Metric Measure Convergence

Recall metric measure convergence as reviewed in Sect. 2.3.

Lemma 5.6

Given \(N_j^3 \rightarrow N_\infty \) as in Lemma 5.4 endowed with the Hausdorff measures, then we have metric measure convergence if \(A_0\) has \({\mathcal {H}}^{3}\)-measure 0.

Proof

Recall the maps \(F_j: N_j^3 \rightarrow N_\infty \) defined in (131)–(133) in the proof of Lemma 5.4. We need only show that for almost every \(p\in N_\infty \) and for almost every \(r<r_p\) sufficiently small we have

$$\begin{aligned} {\mathcal {H}}^3(B(p,r))=\lim _{j\rightarrow \infty }{\mathcal {H}}^3(B(p_j,r)), \end{aligned}$$
(168)

where \(F_j(p_j)=p\) and that for any sequence \(p_{0j}\rightarrow p_0\) we have \(r_0\) sufficiently small that for all \(r<r_0\)

$$\begin{aligned} {\mathcal {H}}^3(B(p_0,r))=\lim _{j\rightarrow \infty }{\mathcal {H}}^3(B(p_{0j},r)). \end{aligned}$$
(169)

In fact, take any \(p\ne p_0\) in \(N_{\infty }\) and choose

$$\begin{aligned} r<r_p<d_{N_\infty ^3}(p,p_0)/2. \end{aligned}$$
(170)

Then for j large enough that \(\delta _j< r_p\) we have

$$\begin{aligned} B(p,r)\cap B(p_0,\delta _j)=\emptyset . \end{aligned}$$
(171)

Thus

$$\begin{aligned} B(p_j,r)\cap A_j'=\emptyset . \end{aligned}$$
(172)

Thus by (131), \(F_j\) is an isometry from \(B(p_j,r) \subset N_j^3\) onto \(B(p,r)\subset N_\infty \) and so we have

$$\begin{aligned} {\mathcal {H}}^3(B(p,r))={\mathcal {H}}^3(B(p_j,r))\qquad \forall r<r_p. \end{aligned}$$
(173)

Next we examine \(p_0\). Observe that by (108)

$$\begin{aligned} {\mathcal {H}}_{N_\infty }^3(B(p_0,r)) = {\mathcal {H}}_{M}^3(T_r(A_0))-{\mathcal {H}}_{M}^3(A_0) = {\text {Vol}}_M(T_r(A_0)\setminus A_0). \end{aligned}$$
(174)

For any \(p_{0,j}\rightarrow p_0\), we have by (125)

$$\begin{aligned} r_j=d_{N_j}(p_{0,j}, A_j')\le 4d_{N_\infty }(F_j(p_{0,j}), p_0) \rightarrow 0 \end{aligned}$$
(175)

Thus

$$\begin{aligned} B(p_{0,j},r) \subset T_{r+r_j}(A_j'). \end{aligned}$$
(176)

So

$$\begin{aligned} {\text {Vol}}_{N_j}(B(p_{0,j},r))\le & {} {\text {Vol}}_{N_j}(T_{r+r_j}(A_j')) \end{aligned}$$
(177)
$$\begin{aligned}\le & {} {\text {Vol}}_{N_j}(T_{r+r_j}(A_j')\setminus A_j') +{\text {Vol}}_{N_j}(A_j')\end{aligned}$$
(178)
$$\begin{aligned}= & {} {\text {Vol}}_M\left( T_{r+r_j+\delta _j}(A_0) \setminus T_{\delta _j}(A_0)\right) +{\text {Vol}}_{N_j}(A_j'). \end{aligned}$$
(179)

Thus

$$\begin{aligned} \limsup _{j\rightarrow \infty }{\text {Vol}}_{N_j}(B(p_{0,j},r))\le & {} {\text {Vol}}_M\left( T_{r}(A_0) \setminus A_0\right) +\limsup _{j\rightarrow \infty } {\text {Vol}}_{N_j}(A_j') \end{aligned}$$
(180)
$$\begin{aligned}= & {} {\mathcal {H}}^3(B(p_0,r)) \end{aligned}$$
(181)

since we claim that

$$\begin{aligned} \lim _{j\rightarrow \infty }{\text {Vol}}_{N_j}(A_j') = 0. \end{aligned}$$
(182)

This follows because \(\epsilon _{j} \rightarrow 0\) and (122) implies

$$\begin{aligned} (1-\epsilon _j) {\text {Vol}}_M(A_{\delta _{j}}) \le {\text {Vol}}_{N_j}(A_j') \le (1+\epsilon _j){\text {Vol}}_{M}(A_{\delta _{j}}). \end{aligned}$$
(183)

The assumption that \({\mathcal {H}}^{3}(A_{0})=0\) then implies (182) after taking the limit.

Similarly, we have for j sufficiently large

$$\begin{aligned} T_{r-H_j-r_j}(A_j') \subset B(p_{0,j},r). \end{aligned}$$
(184)

So

$$\begin{aligned} {\text {Vol}}_{N_j}(B(p_{0,j},r))\ge & {} {\text {Vol}}_{N_j}(T_{r-H_j-r_j}(A_j')) \end{aligned}$$
(185)
$$\begin{aligned}= & {} {\text {Vol}}_{N_j}(T_{r-H_j-r_j}(A_j')\setminus A_j') +{\text {Vol}}_{N_j}(A_j')\end{aligned}$$
(186)
$$\begin{aligned}= & {} {\text {Vol}}_M\left( T_{r-H_j-r_j+\delta _j}(A_0) \setminus T_{\delta _j}(A_0)\right) +{\text {Vol}}_{N_j}(A_j').\qquad \end{aligned}$$
(187)

Thus

$$\begin{aligned} \liminf _{j\rightarrow \infty }{\text {Vol}}_{N_j}(B(p_{0,j},r))\ge & {} {\text {Vol}}_M\left( T_{r}(A_0) \setminus A_0\right) +\liminf _{j\rightarrow \infty }{\text {Vol}}_{N_j}(A_j') \end{aligned}$$
(188)
$$\begin{aligned}= & {} {\mathcal {H}}^3(B(p_0,r)), \,\,\, \text { by }(182) \end{aligned}$$
(189)

which completes the proof. \(\square \)

5.4 Intrinsic Flat Convergence

For a review of intrinsic flat convergence see Sect. 2.5.

Lemma 5.7

Let \(N_j^3 {\mathop {\longrightarrow }\limits ^{ \text {GH}}} N_\infty \) be exactly as in Lemmas 5.4 and 5.5 where we assume M is compact and we have a compact set, \(A_0\subset M\setminus \partial M\). Then there exists an integral current space N such that \({\bar{N}}\) is isometric to \(N_\infty \) and

$$\begin{aligned} N_j {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} N. \end{aligned}$$
(190)

and when \(A_0\) has Hausdorff measure 0

$$\begin{aligned} {{\mathbf {M}}}(N_j) \rightarrow {{\mathbf {M}}}(N)={\mathcal {H}}^3(N). \end{aligned}$$
(191)

When \(A_0=C([0,1])\) then \(N=N_\infty \).

Proof

By (123), we have uniformly bounded volume

$$\begin{aligned} {\text {Vol}}(N_j^3) \le 2 {\text {Vol}}(M^3). \end{aligned}$$
(192)

Since \(\partial N_j^3=\partial M^3\), we have uniformly bounded boundary volume

$$\begin{aligned} {\text {Vol}}(\partial N_j^3) ={\text {Vol}}(\partial M^3). \end{aligned}$$
(193)

Combining this with Lemma 5.5 and Theorem 2.6, there exists an integral current space N possibly \(N=\mathbf{{0}}\) such that a subsequence

$$\begin{aligned} N_j {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} N. \end{aligned}$$
(194)

We claim that \(N\ne \mathbf{{0}}\). If not, then by the final line in Lemma 2.12, for any sequence \(p_j\in N_j\) and almost every r, \(S(p_j,r) {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} \mathbf{{0}}\). However, taking \(p_j\) and r such that

$$\begin{aligned} B(p_j,r)\subset N_j^3 \setminus A_j' \end{aligned}$$
(195)

we know there is some \(p \in M^3\) with \(B(p,r) \subset N_\infty \setminus \{p_0\}\) that \(d_{{\mathcal {F}}}(S(p_j,r), S(p,r))=0\) for \(p\in M^3\), so \(S(p_j,r) {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} S(p,r) \ne \mathbf{{0}}\) which is a contradiction.

By Theorem 2.13, we know that after possibly taking a subsequence we obtain a limit map

$$\begin{aligned} F_\infty : N \rightarrow N_\infty . \end{aligned}$$
(196)

We claim that \(F_\infty \) is distance preserving. Let \(p,q\in N\). By Theorem 2.11, we have \(p_j,q_j \in N_j\) converging to pq in the sense of Definition 2.9, i.e.,

$$\begin{aligned} d_{N_j}(p_j, q_j ) \rightarrow d_{N}(p,q). \end{aligned}$$
(197)

Since the \(F_j\) are \(\epsilon _j\)-almost isometries and \(\epsilon _j \rightarrow 0\), we have

$$\begin{aligned} d_{N_\infty }(F_j(p_j), F_j(q_j)) \rightarrow d_N(p,q). \end{aligned}$$
(198)

By the definition of \(F_\infty \) we have \(F_j(p_j)\rightarrow F_\infty (p)\) and \(F_j(q_j)\rightarrow F_\infty (q)\). Thus

$$\begin{aligned} d_{N_\infty }(F_\infty (p), F_\infty (q)) = d_N(p,q). \end{aligned}$$
(199)

We claim that \(F_\infty \) maps onto at least \(N_\infty \setminus \{p_0\}\). Let \(x \in N_\infty \setminus \{p_0\}\). Since \(F_j\) are surjective, there exists \(x_j\in N_j\) such that \(F_j(x_j)=x\). Since \(x\ne p_0\), we may define

$$\begin{aligned} r=\min \{d_{N_\infty }(x,p_0)/3, \text {ConvexRad}_M(x)\}, \end{aligned}$$
(200)

where \(\text {ConvexRad}_M(x)\) is the convexity radius about x viewed as a point in M. Then there exists j sufficiently large such that \(\delta _j<r\) so that

$$\begin{aligned} B(x_j, r) \subset N_j \setminus T_{\delta _j}(A_j'). \end{aligned}$$
(201)

Furthermore, these balls are isometric to the convex ball \(B(x,r)\subset M^3\).

So

$$\begin{aligned} d_{{\mathcal {F}}}(S(x_j,r), \mathbf{{0}}) =d_{{\mathcal {F}}}(S(x,r), \mathbf{{0}}) >0. \end{aligned}$$
(202)

Thus by Theorem 2.14 with \(h_0=d_{{\mathcal {F}}}(S(x,r), \mathbf{{0}})\), and \(N_j {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} N\), a subsequence of the \(x_j\) converges to \(x_\infty \in N\). By the definition of \(F_\infty \), we have \(F_j(x_j) \rightarrow F_\infty (x_\infty ) \in N_\infty \). But since \(F_j(x_j)=x\) it follows that \(F_\infty (x_\infty )=x\), hence \(F_\infty \) maps onto \(N_\infty \setminus p_0\).

Taking the metric completions of N and \(N_\infty \setminus \{p_0\}\), we have an isometry

$$\begin{aligned} F_\infty : {\bar{N}} \rightarrow N_\infty . \end{aligned}$$
(203)

Since \(N_j\) are Riemannian manifolds,

$$\begin{aligned} {{\mathbf {M}}}([\![N_j ]\!])={\text {Vol}}(N_j)={\mathcal {H}}^3(N_j). \end{aligned}$$
(204)

By the lower semicontinuity of mass and the metric measure convergence of \(N_j\) to N we know that

$$\begin{aligned} {{\mathbf {M}}}([\![N_{\infty } ]\!])\le \liminf _{j\rightarrow \infty }{{\mathbf {M}}}([\![N_j ]\!])={\mathcal {H}}^3(N). \end{aligned}$$
(205)

On the other hand by (29)

$$\begin{aligned} {{\mathbf {M}}}([\![N_{\infty } ]\!])\ge {\mathcal {H}}^3(N) \end{aligned}$$
(206)

because almost every tangent cone is Euclidean and it has integer weight everywhere. Thus we have (191). In fact, equality in these inequalities implies that N has weight one everywhere.

Recall that the set of an integral current space only includes points of positive density. Since

$$\begin{aligned} \liminf _{r\rightarrow 0} \frac{{\text {Vol}}_{N_\infty }(B(p_0,r))}{r^3} =\liminf _{r\rightarrow 0} \frac{{\text {Vol}}_{M}(T_r(A_0)\setminus A_0)}{r^3} \end{aligned}$$
(207)

Thus N is isometric to \(N_\infty \) when this liminf is positive and N is isometric to \(N_\infty \setminus \{p_0\}\) when this liminf is 0. When \(A_0=C([0,1])\) is a curve in a 3-dimensional Riemannian manifold we have

$$\begin{aligned} \liminf _{r\rightarrow 0} \frac{{\text {Vol}}_{M}(T_r(A_0)\setminus A_0)}{r^3} = \liminf _{r\rightarrow 0} \frac{\pi r^2 L(C)}{r^3} = + \infty >0. \end{aligned}$$
(208)

Thus N is isometric to \(N_\infty \).

Thus N does not depend on the subsequence in (194) and in fact the original sequence (given a consistent orientation) converges in the intrinsic flat sense to N. \(\square \)

5.5 The Proof of Theorem 5.2

Proof

In Proposition 3.3, we show that given any \(\epsilon _j \rightarrow 0\) we can find \(n_j \rightarrow \infty \) and \(\delta _j \rightarrow 0\) so fast that \(\delta _jn_j \rightarrow 0\) and we have \(h(\delta _j)n_j \rightarrow 0\) as well such that the sewn manifolds:

$$\begin{aligned} N_j^3 = (M^3 \setminus A_{\delta _j} )\sqcup A_{\delta _j}', \end{aligned}$$
(209)

satisfy:

$$\begin{aligned} (1-\epsilon ){\text {Vol}}(A_{\delta }) \le {\text {Vol}}(A_{\delta }')\le {\text {Vol}}(A_{\delta })(1+\epsilon ) \end{aligned}$$
(210)

and

$$\begin{aligned} (1-\epsilon ) {\text {Vol}}(M^3)\le {\text {Vol}}(N^3) \le {\text {Vol}}(M^3) (1+\epsilon ) \end{aligned}$$
(211)

and

$$\begin{aligned} {\text {Diam}}(A_{\delta }')\le H(\delta )= L(C)/n + (n+1)\, h(\delta )+(5n+2)\, \delta , \end{aligned}$$
(212)

where

$$\begin{aligned} \lim _{\delta \rightarrow 0} H(\delta )=0 \text { uniformly for } K\in (0,1]. \end{aligned}$$
(213)

Thus we have a sequence \(N_j\) which is scrunching a set \(A_0=C([0,1])\) to a point as in Definition 5.3.

Lemma 5.5 implies that

$$\begin{aligned} N_j {\mathop {\longrightarrow }\limits ^{ \text {GH}}} N_\infty , \end{aligned}$$
(214)

where \(N_\infty \) is the pulled string space. Lemma 5.6 implies we have metric measure to \(N_\infty \) convergence because \(A_0=C([0,1])\) has \({\mathcal {H}}^{3}\)-measure 0.

Lemma 5.7 implies that

$$\begin{aligned} N_j {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} N_\infty \end{aligned}$$
(215)

and

$$\begin{aligned} {{\mathbf {M}}}(N_j) \rightarrow {{\mathbf {M}}}(N_\infty )={\mathcal {H}}^3(N), \end{aligned}$$
(216)

completing the proof of Theorem 5.2. \(\square \)

6 Sewing a Sphere to Obtain our Limit Space

Here we construct the specific example of a sequence of manifolds with positive scalar curvature that converges to a limit space which fails to have generalized nonnegative scalar curvature as discussed in the introduction. More specifically:

Example 6.1

We define a sequence \(N_j^3\) of manifolds with positive scalar curvature constructed from the standard \({\mathbb {S}}^3\) sewn along a closed geodesic \(C:[0,1]\rightarrow {\mathbb {S}}^3\) with \(\delta =\delta _j \rightarrow 0\) as in Proposition 3.3. Then by Theorem 5.2 we have

$$\begin{aligned} N_j^3 {\mathop {\longrightarrow }\limits ^{ \text {mGH}}} N_\infty \text { and } N_j^3 {\mathop {\longrightarrow }\limits ^{{\mathcal {F}}}} N_\infty , \end{aligned}$$
(217)

where \(N_\infty \) is the metric space created by taking the standard sphere and pulling the geodesic to a point as in Proposition 4.1. By Lemma 6.3 below we see that at the pulled point \(p_0\in N_\infty \), we have (3). Thus we have produces a sequence of three-dimensional manifolds with positive scalar curvature converging to a limit space which fails to satisfy generalized scalar curvature defined using limits of volumes of balls as in (1).

Remark 6.2

Note that with \(\delta _j \rightarrow 0\), the neck in the center of the tunnels has a rotationally symmetric minimal surface whose area is \(\le 4\pi \delta _j^2\) which converges to 0. So this sequence, and in fact any sewn sequence created as in Definition 5.1, has \(\text {MinA}(N_j) \rightarrow 0\).

Lemma 6.3

At the pulled point \(p_0\in N_\infty \) of Example 6.1 we have

$$\begin{aligned} \lim _{r\rightarrow 0} \left( \frac{{\text {Vol}}_{{\mathbb {E}}^3}(B(0,r)) - {\text {Vol}}_{N_\infty }(B(p_0,r))}{r^2 {\text {Vol}}_{{\mathbb {E}}^3}(B(0,r))}\right) =-\infty . \end{aligned}$$
(218)

Proof

First, observe that

$$\begin{aligned} {\text {Vol}}_{N_\infty }(B(p_0,r))= & {} {\mathcal {H}}_{N_\infty }^3\left( B(p_0,r)\right) \end{aligned}$$
(219)
$$\begin{aligned}= & {} {\mathcal {H}}_{N_\infty }^3\left( B(p_0,r) \setminus \{p_0\}\right) \end{aligned}$$
(220)
$$\begin{aligned}= & {} {\mathcal {H}}_{{\mathbb {S}}^3}^3\left( \,T_r(C([0,1]))\,\right) . \end{aligned}$$
(221)

Since C([0, 1]) is a closed geodesic of length \(2\pi \) in a three-dimensional sphere, we have

$$\begin{aligned} \lim _{r\rightarrow 0}\,\, \frac{\,{\mathcal {H}}_{{\mathbb {S}}^3}^3\left( \,T_r(C([0,1]))\,\right) \,}{ 2\pi (\pi r^2)}\,\,\,=\,\,\,1. \end{aligned}$$
(222)

Thus

$$\begin{aligned} \lim _{r\rightarrow 0} \frac{{\text {Vol}}_{{\mathbb {E}}^3}(B(0,r)) - {\text {Vol}}_{N_\infty }(B(p_0,r))}{r^2 {\text {Vol}}_{{\mathbb {E}}^3}(B(0,r))} =\lim _{r\rightarrow 0} \frac{(4/3)\pi r^3 - 2\pi (\pi r^2)}{(4/3)\pi r^5 } =-\infty \end{aligned}$$
(223)

as claimed. \(\square \)