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

Circle bundles with PSC over some four manifolds

Aditya Kumar Department of Mathematics, Johns Hopkins University. 3400 N. Charles Street, Baltimore, MD 21218, USA [email protected]  and  Balarka Sen School of Mathematics, Tata Institute of Fundamental Research. 1, Homi Bhabha Road, Mumbai-400005, India [email protected], [email protected]
Abstract.

We construct infinitely many examples of four manifolds with macroscopic dimension 44 equipped with circle bundles whose total spaces admit metrics of positive scalar curvature. Further, we verify that these bundles have macroscopic dimension at most 33. In particular, we answer a question of Gromov on the existence of circle bundles over enlargeable manifolds whose total spaces admit metrics of positive scalar curvature. Our constructions are based on techniques from symplectic geometry.

The second author is supported by the Department of Atomic Energy, Government of India, under project no.12-R&D-TFR-5.01-0500.

1. Introduction

1.1. Background and motivation

Definition 1 (Urysohn width).

Let XX be a metric space. The kk-Urysohn width of XX, denoted by UWk(X)\mathrm{UW}_{k}(X), is the infimum over t>0t>0 such that there is a kk-dimensional simplicial complex YY and a continuous map f:XYkf:X\to Y^{k} satisfying the property that the diameter of each fiber of ff is at most tt, i.e. for all yYy\in Y, we have diam(f1{y})t\mathrm{diam}(f^{-1}\{y\})\leq t.

If k<nk<n, then a small kk-Urysohn width intuitively means that the manifold is close to a kk-dimensional space. This notion of closeness is captured in the following definition.

Definition 2 (Macroscopic dimension).

Let XX be a metric space. The macroscopic dimension, denoted by dimmcX\dim_{mc}X, is the minimal kk such that UWk(X)<\mathrm{UW}_{k}(X)<\infty. In other words, dimmcXk\dim_{mc}X\leq k if there is a continuous map f:XYkf:X\to Y^{k} to a kk-dimensional simplicial complex YY, such that for all yYy\in Y, diamf1(y)<C\mathrm{diam}f^{-1}(y)<C for some uniform constant CC.

Remark.

Note that for a compact metric space XX, dimmc(X)=0\dim_{mc}(X)=0. Therefore, when one refers to the macroscopic dimension of a closed manifold MM, one typically means the macroscopic dimension of the universal cover M~\widetilde{M}.

The notions of Urysohn width and macroscopic dimension appear in several works of Gromov, with one of the aims being to capture the appropriate notion of smallness that must be exhibited by a manifold admitting a metric of positive scalar curvature (PSC). These also appear naturally in the study of non-collapsing and thick-thin decomposition. See also some of these recent works on the topic [ABG24, Gut11, Nab22, Sab22]. The primary motivation behind this article is the following conjecture formulated by Gromov for manifolds with positive scalar curvature [Gro96, Section 2½2\text{\textonehalf}].

Conjecture 1 (Gromov).

Let MM be a closed connected nn-dimensional manifold, admitting a PSC metric. Then the macroscopic dimension of the universal cover of MM is at most n2n-2, that is,

dimmcM~n2.\dim_{mc}\widetilde{M}\leq n-2.

Intuitively, the conjecture says that a manifold admitting a PSC metric, on large scales, looks like a manifold of at least two dimensions less. The primary evidence in this direction is the following simple observation: Given any closed manifold MM, consider the manifold M×Sr2M\times S^{2}_{r} with the product metric. Here, Sr2S^{2}_{r} denotes a sphere of radius rr. Then for rr small enough, the product metric on M×Sr2M\times S^{2}_{r} has positive scalar curvature. In light of the previous remark, we pass to the universal cover of M×Sr2M\times S^{2}_{r}, which is M~×Sr2\widetilde{M}\times S^{2}_{r}. Since the projection map to M~\widetilde{M} is continuous and each fiber is Sr2S^{2}_{r}, we have

UWn2(M×Sr2~)2r,\mathrm{UW}_{n-2}(\widetilde{M\times S^{2}_{r}})\leq 2r,

and therefore the macroscopic dimension dimmc(M×Sr2~)n2\dim_{mc}(\widetilde{M\times S^{2}_{r}})\leq n-2. Conjecture 1 asserts that all PSC manifolds must exhibit this behavior. Gromov’s conjecture has been verified in dimension 33 [GL83, LM23]. In higher dimensions there are several partial results under various assumptions (cf. [BD16, DD22]).

While Conjecture 1 arises from the simple observation that existence of a PSC metric is guaranteed given the presence of an S2S^{2} factor in the manifold, it is also natural to ask about the possible interaction between existence of PSC metrics and having an S1S^{1} factor. This has been widely investigated. For instance, PSC manifolds with a trivial S1S^{1} factor (i.e., M×S1M\times S^{1}) are amenable to the Schoen-Yau conformal descent argument [SY79], as they have a non-trivial codimension-11 homology class given by the Poincaré dual of the S1S^{1} factor. Such S1S^{1} factors are also useful in quantitative geometric problems related to scalar curvature. For example, building on an observation of Fisher-Colbrie and Schoen [FCS80], Gromov and Lawson introduced the toric symmetrization procedure [GL83] which was used to estimate the Lipschitz constant of a non-zero degree map from a PSC manifold. More recently, Gromov [Gro18] observed that toric symmetrization can also be used to prove width estimates for Riemannian bands. On the same theme, Rosenberg [Ros07, Conjecture 1.24] formulated the following conjecture regarding the relationship between PSC metrics and trivial S1S^{1} factors.

Conjecture 2 (Rosenberg).

Let MM be a closed connected nn-dimensional manifold. Then M×S1M\times S^{1} admits a PSC metric iff MM admits a PSC metric.

However, the situation is less clear when the S1S^{1} factor arises as a fiber of a non-trivial circle bundle. The key difficulty in studying existence of PSC metrics on non-trivial circle bundles is that, in this case, the Schoen-Yau conformal descent argument does not apply in a straightforward way.111Indeed, the integral homology class of the S1S^{1} fiber in nontrivial circle bundles is always torsion, with order given by the divisibility of the Euler class of the bundle. Therefore, its Poincaré dual is zero. In this direction, Gromov posed the following question [Gro18, pg. 661]:

Question 1.

[Gro18, pg. 661] “What happens to nontrivial bundles X¯Y\underline{X}\to Y?
[…] In general, the examples in [BH10] indicate a possibility of non-enlargeable circle bundles over enlargeable YY; yet, it seems hard(er) to find such examples, where the corresponding X¯\underline{X} would admit complete metrics with Scσ>0Sc\geq\sigma>0.”

In the negative direction we point out the recent comprehensive work [He23] on PSC metrics on non-trivial circle bundles due to He. In particular, he showed that circle bundles with homologically non-trivial fibers do not admit PSC metrics if the base space is a 11-enlargeable22211-enlargeable means that the maps to SnS^{n} in the definition of enlargeability are of degree 11. manifold of dimension n6n\leq 6 [He23, Proposition 4.12].

The main goal of this article is to study and exhibit examples of interactions between the existence of a positive scalar curvature metric and macroscopic dimension on non-trivial circle bundles over some 44-manifolds. In particular, we give a positive answer to Question 1. We construct these examples using techniques from symplectic geometry. The inspiration for our construction came from the following remark by Gromov [Gro23, pg. 8]:

The emergent picture of spaces with Sc.curv0Sc.curv\geq 0, where topology and geometry are intimately intertwined, is reminiscent of the symplectic geometry, but the former has not reached yet the maturity of the latter.

Granting this dictum of similarity between scalar curvature geometry and symplectic geometry, one is led to believe that the analogue of the Schoen-Yau descent for positive scalar curvature manifolds, in the symplectic case must be Donaldson’s symplectic divisor theorem [Don96], as it can be viewed as a codimension-22 descent-type result for symplectic manifolds. This analogy is strengthened when seen in conjunction with a recent result of Cecchini, Räde and Zeidler [CRZ23] which establishes a certain codimension-22 descent for positive scalar curvature manifolds of dimension 77. In fact, we show that circle bundles over Donaldson divisors in symplecic 66-manifolds fit neatly into the framework of applicability of [CRZ23].

1.2. Main results

We discuss the necessary symplectic preliminaries in Section 2.

In this article, we construct infinitely many examples of 44-manifolds ZZ with arbitrarily large Euler number |χ(Z)||\chi(Z)|, arising as Donaldson divisors (see Theorem 1) in symplectic 66-manifolds, such that ZZ does not admit a PSC metric and macroscopic dimension of Z~\widetilde{Z} is exactly 44. Nevertheless, we construct circle bundles over these ZZ,

S1EZ𝑝Z,S^{1}\hookrightarrow E_{Z}\xrightarrow{p}Z,

such that the 55-dimensional total space EZE_{Z} admits a PSC metric. An intriguing feature of these examples is that Z~\widetilde{Z} and EZ~\widetilde{E_{Z}} are quasi-isometric.333π1(EZ)\pi_{1}(E_{Z}) is a 𝐙m\mathbf{Z}_{m}-extension of π1(Z)\pi_{1}(Z), where mm is the divisibility of the Euler class of EZE_{Z} pulled back to the universal cover Z~\widetilde{Z} of ZZ. Therefore, π1(EZ)\pi_{1}(E_{Z}) and π1(Z)\pi_{1}(Z) are quasi-isometric. By the Švarc-Milnor lemma, EZ~\widetilde{E_{Z}} and Z~\widetilde{Z} must be quasi-isometric as well. Therefore, since dimmcZ~=4\dim_{mc}\widetilde{Z}=4, one may naively expect dimEZ~=4\dim\widetilde{E_{Z}}=4 as well (since EZE_{Z} is PSC, that would give a counterexample to Gromov’s Conjecture 1!). However, this is not the case. Indeed, while it is straightforward to see that macroscopic dimension is invariant under quasi-isometric homeomorphisms, it is not invariant under just a quasi-isometry. See also [KR23, Theorem F]. Thus, even though dimmcZ~=4\dim_{mc}\widetilde{Z}=4, and Z~\widetilde{Z} and EZ~\widetilde{E_{Z}} are quasi-isometric, strangely the macroscopic dimension drops for EZE_{Z}, even though EZE_{Z} is obtained from ZZ by taking a twisted product with S1S^{1}. Precisely, it turns out that dimmcEZ~3\dim_{mc}\widetilde{E_{Z}}\leq 3 in consonance with Gromov’s Conjecture 1. We show this using the natural deformation retraction from EZ~\widetilde{E_{Z}} to a 33-dimensional simplicial complex arising as the spine of the complement of the Donaldson divisor ZZ in the ambient symplectic 66-manifold.

Note that this phenomenon of the macroscopic dimension decreasing upon taking a (twisted) product with S1S^{1} is not possible in the case of a trivial product. Indeed,

dimmc(Z×S1~)=dimmc(Z~×𝐑)=dimmc(Z~)+1.\dim_{mc}(\widetilde{Z\times S^{1}})=\dim_{mc}(\widetilde{Z}\times\mathbf{R})=\dim_{mc}(\widetilde{Z})+1.

To the best of our knowledge examples exhibiting a drop in macroscopic dimension on twisting with S1S^{1} have not appeared before in literature. On a similar theme, we were informed by Guth [Gut24] that a drop in the Urysohn width upon taking a double cover, i.e. an S0S^{0} bundle, was observed by Balitskiy [Bal21, Theorem 2.4.4, pg. 36].

We now state the main result of this article:

Theorem A.

Let MM be an integral symplectic 44-manifold. Consider the symplectic manifold M×S2M\times S^{2}. Let Z4Z^{4} be any Donaldson divisor of M×S2M\times S^{2}. Then there is a circle bundle S1EZ𝑝ZS^{1}\hookrightarrow E_{Z}\xrightarrow{p}Z, such that the following statements hold:

  1. (1)

    The total space EZE_{Z} admits a metric of positive scalar curvature.

  2. (2)

    The macroscopic dimension of the total space EZE_{Z} is at most 3.

  3. (3)

    If b2+(M)>1b_{2}^{+}(M)>1, then ZZ does not admit a positive scalar curvature metric.

  4. (4)

    If MM admits a positive degree map to an aspherical manifold, then ZZ does not admit a positive scalar curvature metric.

  5. (5)

    If MM is a symplectically aspherical444See Definition 3. In particular, a symplectic manifold with π2=0\pi_{2}=0 (e.g. T4T^{4}) is symplectically aspherical. manifold and π1(M)\pi_{1}(M) is amenable, then the macroscopic dimension of ZZ is exactly 44.

We emphasize that Conclusions (1)(1) and (2)(2) make no assumptions either on the fundamental groups of MM and EZE_{Z}, or on the (non-)existence of spin or almost spin structures on the universal cover of EZE_{Z}. In particular, Conclusion (2)(2) does not follow from any existing results on Gromov’s conjecture (Conjecture 1). On the other hand, Conclusions (3)(3), (4)(4) and (5)(5) rest on Lemma 5, where we will show that there is a positive degree map from ZZ to MM. Then Conclusion (3)(3) follows from a theorem of Taubes [Tau94], Conclusion (4)(4) follows from the work of Chodosh, Li and Liokumovich [CLL23] and Conclusion (5)(5) follows from combining Proposition 6 with a result of Dranishnikov [Dra11].

By combining Theorem A with a Euler characteristic computation for Donaldson divisors carried out in Proposition 4, we obtain the following uniqueness corollary:

Corollary B.

There exists infinitely many 44-manifolds ZkZ_{k} and circle bundles EkE_{k} over ZkZ_{k} such that |χ(Zk)|O(k3)|\chi(Z_{k})|\sim O(k^{3}), and

  1. (1)

    dimmcZk=4\dim_{mc}Z_{k}=4 but dimmcEk=3\dim_{mc}E_{k}=3,

  2. (2)

    ZkZ_{k} is not PSC and EkE_{k} is PSC.

As a further corollary of Theorem A, we will construct examples sought after in Gromov’s Question 1.

Corollary C.

There exists enlargeable closed 44-manifolds ZZ and nontrivial circle bundles EZE\to Z such that EE admits a positive scalar curvature metric.

1.3. Organisation

In Section 2, we define the necessary symplectic preliminaries and make some observations regarding Donaldson divisors. In Section 3, we assemble other lemmas and propositions that we need to prove our main results. In Section 4, we prove Theorem A and Corollary C as well as provide explicit examples that are used towards proofs of Corollary B. The reader may first read Section 4, and refer to Sections 2 and 3, for the proofs of the results used in Section 4 as they arise.

Acknowledgements

The authors would like to thank Shihang He for several helpful comments on the first version of this paper. They would also like to thank Alexey Balitskiy, Alexander Dranishnikov, and Larry Guth for answering their questions and providing helpful remarks. The first author wishes to thank his advisor Yannick Sire for support and encouragement. The second author wishes to thank his advisor Mahan Mj, as well as his friends and colleagues Ritwik Chakraborty, and Sekh Kiran Ajij for several helpful conversations.

2. Donaldson divisors: definition and properties

Let (X,ω)(X,\omega) be a symplectic manifold. The symplectic form ω\omega is said to be integral if the cohomology class [ω]H2(X;𝐑)[\omega]\in H^{2}(X;\mathbf{R}) lies in the image of the change of coefficients homomorphism H2(X;𝐙)H2(X;𝐑)H^{2}(X;\mathbf{Z})\to H^{2}(X;\mathbf{R}).

Remark.

Any symplectic manifold (X,η)(X,\eta) admits a symplectic form ω\omega which is integral. To see this, choose a collection of closed 22-forms on XX representing a 𝐐\mathbf{Q}-basis of the vector space H2(X;𝐐)H2(X;𝐑)H^{2}(X;\mathbf{Q})\subset H^{2}(X;\mathbf{R}), and expresses the class [η]H2(X;𝐑)[\eta]\in H^{2}(X;\mathbf{R}) as a 𝐑\mathbf{R}-linear combination of the chosen basis elements. Approximate the real coefficients by rational numbers and then multiply by an integer to clear out the denominators. Let ω\omega be the resulting closed 22-form. Then ω\omega is an integral symplectic form on XX. Indeed, one only needs to verify that ω\omega is non-degenerate. But it is so, since non-degeneracy is both an open condition and is scale-invariant.

The following foundational result in symplectic geometry was established by Donaldson [Don96].

Theorem 1.

[Don96, Theorem 1] Suppose (X2n,ω)(X^{2n},\omega) is a symplectic manifold where ω\omega is integral. Then, for sufficiently large k1k\gg 1, there exists a symplectic submanifold ZXZ\subset X of codimension 22, such that ZZ represents the class Poincaré dual to k[ω]k[\omega].

The symplectic submanifolds Z(X,ω)Z\subset(X,\omega) appearing in the statement of the theorem are referred to as Donaldson divisors of (X,ω)(X,\omega). We briefly outline the construction of ZZ. Since ω\omega is an integral symplectic form, we can consider the pre-quantum line bundle

π:LX\pi:L\to X

for (X,ω)(X,\omega). This is a complex line bundle with a hermitian connection \nabla such that the curvature 22-form is given by F=2πiωF_{\nabla}=-2\pi i\omega. Thus, note that the first Chern class of the line bundle LL is c1(L)=[ω]c_{1}(L)=[\omega]. The symplectic submanifold Z2n2(X2n,ω)Z^{2n-2}\subset(X^{2n},\omega) is produced as the zero set of an asymptotically holomorphic section of LkL^{\otimes k}. For details, see [Don96, Theorem 5]. Note c1(Lk)=k[ω]c_{1}(L^{\otimes k})=k[\omega]. Since ZZ is the zero set of a section of LkL^{\otimes k}, we must have that ZZ represents the class Poincaré dual to k[ω]k[\omega].

2.1. Complement of Donaldson divisors

We note the following two simple but useful observations.

Lemma 2.1.

Let Lk|ZL^{\otimes k}|_{Z} denote the restriction of LkL^{\otimes k} to ZZ. Denote by 𝕊(Lk|Z)\mathbb{S}(L^{\otimes k}|_{Z}) the unit circle bundle of Lk|ZL^{\otimes k}|_{Z}. Then the following statements are true.

  1. (1)

    The tubular neighborhood ν(Z)\nu(Z) of ZXZ\subset X is diffeomorphic to Lk|ZL^{\otimes k}|_{Z}

  2. (2)

    The boundary ν(Z)\partial\nu(Z) is diffeomorphic to 𝕊(Lk|Z)\mathbb{S}(L^{\otimes k}|_{Z}).

Proof.

By the tubular neighborhood theorem, ν(Z)\nu(Z) is diffeomorphic to the normal bundle of ZXZ\subset X. By the construction of ZZ outlined above, Z=s1(𝟎),Z=s^{-1}(\mathbf{0}), where s:XLks:X\to L^{\otimes k} is a section transverse to the zero section 𝟎Lk\mathbf{0}\subset L^{\otimes k}. Since the normal bundle to 𝟎Lk\mathbf{0}\subset L^{\otimes k} is LkL^{\otimes k} itself, the normal bundle to ZZ in XX is Lk|ZL^{\otimes k}|_{Z} by transversality. Since ν(Z)LZk\nu(Z)\cong L^{\otimes k}_{Z}, the total spaces of the unit circle bundles are also diffeomorphic, i.e. ν(Z)𝕊(Lk|Z)\partial\nu(Z)\cong\mathbb{S}(L^{\otimes k}|_{Z}). ∎

The following is an analogue of the Lefschetz hyperplane theorem in the symplectic setting, which we shall use in the next section:

Proposition 2.

[Don96, Proposition 39] The inclusion j:Z2n2X2nj:Z^{2n-2}\hookrightarrow X^{2n} induces an isomorphism j:πk(Z)πk(X),j_{*}:\pi_{k}(Z)\to\pi_{k}(X), for all kn2k\leq n-2, and a surjection for k=n1k=n-1.

In fact, the following lemma can be deduced from the proof of Proposition 2 in [Don96, pg. 700-701]:

Lemma 2.2.

The complement X2nZX^{2n}\setminus Z deformation retracts to a simplicial complex of dimension nn.

Proof.

By the construction of ZXZ\subset X summarized above, there is an asymptotically holomorphic section s:XLks:X\to L^{\otimes k} such that Z=s1(𝟎)Z=s^{-1}(\mathbf{0}). Let ψ:XZ𝐑\psi:X\setminus Z\to\mathbf{R} be defined by ψ:=log|s|2\psi:=-\log|s|^{2}. During the course of the proof of [Don96, Proposition 39], the author shows that ψ\psi is a Morse function on XZX\setminus Z with critical points having index at most nn. Therefore, by Morse theory, XZX\setminus Z deformation retracts to an nn-dimensional simplicial complex. ∎

We deduce two corollaries from Lemma 2.2. These are central to the proof of Part (2) of Theorem A.

Corollary 3.

Let ν(Z)\nu(Z) be a tubular neighborhood of ZZ in X2nX^{2n}. Let W=Xν(Z)W=X\setminus\nu(Z). Then the following statements are true,

  1. (1)

    WW deformation retracts to a simplicial complex of dimension nn.

  2. (2)

    If n3n\geq 3, the inclusion WW\partial W\hookrightarrow W induces an isomorphism in π1\pi_{1}.

Proof.

For Part (1), observe W=ν(Z)\partial W=\partial\nu(Z). So, by Lemma 2.1, we have W𝕊(Lk|Z)\partial W\cong\mathbb{S}(L^{\otimes k}|_{Z}). Therefore, XZX\setminus Z is simply WW with a collar W×[0,)\partial W\times[0,\infty) attached. Consequently, XZX\setminus Z deformation retracts to WW. Since by Lemma 2.2, XZX\setminus Z deformation retracts to a nn-dimensional simplicial complex, so does WW.

For Part (2)(2), observe that the Morse function ψ\psi constructed in Lemma 2.2 gives a handlebody decomposition of WW by handles of index at most nn. By reversing the handlebody decomposition, we get that WW can be constructed from W\partial W by attaching handles of index at least nn. If n3n\geq 3, then we conclude that all these handles are of index at least 33. Since π1\pi_{1} is unchanged under attachements of handles of index bigger than 22, we conclude WW\partial W\hookrightarrow W induces an isomorphism in π1\pi_{1}. ∎

2.2. Topology of Donaldson divisors in dimension 66

In the next proposition, we show that Donaldson divisors of a 66-dimensional symplectic manifold (X6,ω)(X^{6},\omega) with integral symplectic form ω\omega, realizing the homology classes Poincaré dual to k[ω]k[\omega], are topologically distinct for distinct values of k1k\gg 1. We do this by calculating their Euler characteristic.

Proposition 4.

Let Zk(X6,ω)Z_{k}\subset(X^{6},\omega) denote a Donaldson divisor with [Zk]=PD(k[ω])[Z_{k}]=\mathrm{PD}(k[\omega]). Then, for sufficiently large values of k1k\gg 1, |χ(Zk)|O(k3)|\chi(Z_{k})|\sim O(k^{3}) is monotonically increasing in kk. In particular, for k,l1k,l\gg 1 sufficiently large, ZkZlZ_{k}\cong Z_{l} if and only if k=lk=l.

Proof.

From Lemma 2.1, we know that the normal bundle of ZkZ_{k} in XX is isomorphic to Lk|ZkL^{\otimes k}|_{Z_{k}}, where LL is a complex line bundle over XX with c1(L)=[ω]c_{1}(L)=[\omega]. Therefore,

TZkLk|ZkTX|ZkTZ_{k}\oplus L^{\otimes k}|_{Z_{k}}\cong TX|_{Z_{k}}

Let j:ZkXj:Z_{k}\hookrightarrow X denote the inclusion map. We compute the Chern classes:

c1(TZk)+c1(Lk|Zk)=jc1(TX)\displaystyle c_{1}(TZ_{k})+c_{1}(L^{\otimes k}|_{Z_{k}})=j^{*}c_{1}(TX) (2.1)
c2(TZk)+c1(TZk)c1(Lk|Zk)=jc2(TX)\displaystyle c_{2}(TZ_{k})+c_{1}(TZ_{k})\smile c_{1}(L^{\otimes k}|_{Z_{k}})=j^{*}c_{2}(TX) (2.2)

Note that c1(Lk|Zk)=jc1(Lk)=kj[ω]c_{1}(L^{\otimes k}|_{Z_{k}})=j^{*}c_{1}(L^{\otimes k})=k\cdot j^{*}[\omega]. Substituting this in (2.1), we obtain:

c1(TZk)=jc1(TX)kj[ω]c_{1}(TZ_{k})=j^{*}c_{1}(TX)-k\cdot j^{*}[\omega] (2.3)

Substituting (2.3) in (2.2), we obtain

c2(TZk)\displaystyle c_{2}(TZ_{k}) =jc2(TX)jc1(TX)c1(Lk|Zk)+kj[ω]c1(Lk|Zk)\displaystyle=j^{*}c_{2}(TX)-j^{*}c_{1}(TX)\smile c_{1}(L^{\otimes k}|_{Z_{k}})+k\cdot j^{*}[\omega]\smile c_{1}(L^{\otimes k}|_{Z_{k}})
=jc2(TX)kj(c1(TX)[ω])+k2j[ω]2\displaystyle=j^{*}c_{2}(TX)-k\cdot j^{*}(c_{1}(TX)\smile[\omega])+k^{2}\cdot j^{*}[\omega]^{\smile 2} (2.4)

Since the top Chern class is the Euler class, evaluating the cohomology classes appearing in either sides of (2.4) on [Zk]=kPD[ω][Z_{k}]=k\cdot\mathrm{PD}[\omega] we get:

χ(Zk)\displaystyle\chi(Z_{k}) =Zkc2(TZk)\displaystyle=\int_{Z_{k}}c_{2}(TZ_{k})
=k3M[ω]3k2Mc1(TX)[ω]2+kMc2(TX)[ω]\displaystyle=k^{3}\int_{M}[\omega]^{\smile 3}-k^{2}\int_{M}c_{1}(TX)\smile[\omega]^{\smile 2}+k\int_{M}c_{2}(TX)\smile[\omega]

Note that c1(TX),c2(TX),[ω]c_{1}(TX),c_{2}(TX),[\omega] are all independent of kk. Hence, χ(Zk)\chi(Z_{k}) is a cubic polynomial in kk with coefficients independent of kk, and a nonzero leading coefficient. Therefore, for sufficiently large values of k1k\gg 1, |χ(Zk)||\chi(Z_{k})| is monotonically increasing in kk, as required. ∎

Remark.

The fact that χ(Zk)O(k3)\chi(Z_{k})\sim O(k^{3}) in X6X^{6} should be put in context of the following standard observation: Let X=𝐂𝐏3X=\mathbf{CP}^{3} and Vk𝐂𝐏3V_{k}\subset\mathbf{CP}^{3} be a degree kk complex hypersurface, then χ(Vk)=k34k2+6k\chi(V_{k})=k^{3}-4k^{2}+6k.

2.3. Donaldson divisors of M×S2M\times S^{2}

Let (M4,η0)(M^{4},\eta_{0}) be an integral symplectic 44-manifold. We equip the manifold M×S2M\times S^{2} with the integral symplectic form ω=η0ω0\omega=\eta_{0}\oplus\omega_{0}, where ω0\omega_{0} denotes the standard area form on S2S^{2}. We now make the following observation, which will be crucial to the proof of Parts (4) and (5) of Theorem A.

Proposition 5.

The map f:ZMf:Z\to M, obtained by the composition

ZM×S2𝜋M,Z\hookrightarrow M\times S^{2}\xrightarrow{\pi}M,

has positive degree.

Proof.

Take any regular value pMp\in M. Then the oriented count #f1(p)\#f^{-1}(p) is equal to the oriented intersection number #([Z][{p}×S2])\#([Z]\cap[\{p\}\times S^{2}]). Therefore, we have

deg(f)\displaystyle\deg(f) =#([Z][{p}×S2])\displaystyle=\#([Z]\cap[\{p\}\times S^{2}])
=k{p}×S2ω(since Z is Poincaré dual to k[ω])\displaystyle=k\int_{\{p\}\times S^{2}}\omega\quad\text{(since $Z$ is Poincar\'{e} dual to $k[\omega]$)}
>0\displaystyle>0

The last inequality holds because {p}×S2\{p\}\times S^{2} is a symplectic submanifold of (M×S2,ω)(M\times S^{2},\omega). ∎

2.4. Symplectically aspherical manifolds

Definition 3.

A symplectic manifold (X,ω)(X,\omega) is said to be symplectically aspherical if for every smooth map ϕ:S2X\phi:S^{2}\to X,

S2ϕω=0\int_{S^{2}}\phi^{*}\omega=0
Remark.

Note that it is immediate from Definition 3 that symplectic submanifolds of symplectically aspherical manifolds are also symplectically aspherical. Since symplectic manifolds (X,ω)(X,\omega) with π2(X)=0\pi_{2}(X)=0 are automatically symplectically aspherical, together with the previous observation as well as Donaldson’s theorem (Theorem 1), this provides a large class of symplectically aspherical manifolds.

This class of manifolds was originally introduced by Floer in the context of the Arnol’d conjecture in symplectic geometry. We refer the reader to the survery article [KRT08] for detailed information. The key property of symplectically aspherical manifolds that we shall be using is the following:

Proposition 6.

Let (X2n,ω)(X^{2n},\omega) be an integral symplectic manifold which is symplectically aspherical. Then XX is a rationally essential manifold.555A dd-dimensional manifold MdM^{d} is rationally essential if for the map ϕ:MK(π1(M),1)\phi:M\to K(\pi_{1}(M),1) classifying the universal cover of MM, ϕ[M]0Hd(K(π1(M),1);𝐐)\phi_{*}[M]\neq 0\in H_{d}(K(\pi_{1}(M),1);\mathbf{Q})

Proof.

By attaching cells of dimension 3\geq 3 to XX, we kill the higher homotopy groups πi(X)\pi_{i}(X), i2i\geq 2 to obtain a CW-complex which is a model for a K(π1(X),1)K(\pi_{1}(X),1). By construction, there is a natural inclusion as a subcomplex

i:XK(π1(X),1)i:X\hookrightarrow K(\pi_{1}(X),1)

Let LL be the pre-quantum line bundle over XX, with c1(L)=[ω]c_{1}(L)=[\omega]. Since (X,ω)(X,\omega) is symplectically aspherical, LL must extend to the 33-skeleton of K(π1(X),1)K(\pi_{1}(X),1). Indeed, for any 33-cell of K(π1(X),1)K(\pi_{1}(X),1), let ϕ:S2X\phi:S^{2}\to X denote the attaching map. Then, we have:

c1(ϕL),[S2]=S2ϕω=0\langle c_{1}(\phi^{*}L),[S^{2}]\rangle=\int_{S^{2}}\phi^{*}\omega=0

Therefore, c1(ϕL)=0c_{1}(\phi^{*}L)=0. Consequently, ϕL\phi^{*}L must be the trivial bundle over the 22-sphere. Thus, we may extend LL as a trivial bundle over the 33-cell. Since complex line bundles over higher dimensional spheres are always trivial, we can extend LL over all the other higher dimension cells as well.

In this way, we obtain a line bundle LL^{\prime} over K(π1(X),1)K(\pi_{1}(X),1) such that iL=Li^{*}L^{\prime}=L. Let us denote α:=c1(L)H2(K(π1,1);𝐙)\alpha:=c_{1}(L^{\prime})\in H^{2}(K(\pi_{1},1);\mathbf{Z}). Then, iα=[ω]i^{*}\alpha=[\omega]. Consequently, iαn=[ωn]0i^{*}\alpha^{\wedge n}=[\omega^{\wedge n}]\neq 0. Thus, i:H2n(K(π1(X),1);𝐙)H2n(X;𝐙)i^{*}:H^{2n}(K(\pi_{1}(X),1);\mathbf{Z})\to H^{2n}(X;\mathbf{Z}) is nonzero. Since [ωn][\omega^{\wedge n}] is not torsion, αn\alpha^{\wedge n} must not be torsion either. Therefore, i:H2n(K(π1,1);𝐐)H2n(X;𝐐)i_{*}:H^{2n}(K(\pi_{1},1);\mathbf{Q})\to H^{2n}(X;\mathbf{Q}) is nonzero. Therefore, by the universal coefficients theorem, i:H2n(X;𝐐)H2n(K(π1(X),1);𝐐)i_{*}:H_{2n}(X;\mathbf{Q})\to H_{2n}(K(\pi_{1}(X),1);\mathbf{Q}) must be nonzero as well. ∎

3. Preliminary results

In this section we assemble all other results that we will require to prove Theorem A.

3.1. Nullcobordism and macroscopic dimension

We first prove a general lemma relating macroscopic dimension of a manifold with the homotopy type of a nullcobordism of the manifold. This will be used to prove of Part (2) of Theorem A.

Lemma 3.1.

Let WnW^{n} be a compact nn-dimensional manifold with boundary E=WE=\partial W. Suppose

  1. (1)

    WW deformation retracts to a kk-dimensional simplicial complex KWK\subset W,

  2. (2)

    The inclusion i:E=WWi:E=\partial W\hookrightarrow W induces an isomorphism in π1\pi_{1}.

Then, dimmcE~k\dim_{mc}\widetilde{E}\leq k.

Proof.

Let r:WKr:W\to K denote the deformation retract. Let f=ri:EKf=r\circ i:E\to K denote the composition with the inclusion i:EWi:E\hookrightarrow W. Since ii and rr both induce an isomorphism in π1\pi_{1}, so does ff. Therefore, ff lifts to a map between the respective universal covers:

f~:E~K~\displaystyle\widetilde{f}:\widetilde{E}\to\widetilde{K}

Choose an arbitrary Riemannian metric gg on WW, and let g~\widetilde{g} be its lift to the universal cover W~\widetilde{W}. We will show that the fibers of f~\widetilde{f} are uniformly bounded with respect to the metric g~\widetilde{g}. This will give us the desired conclusion, i.e. dimmcE~k\dim_{mc}\widetilde{E}\leq k.

To this end, let us assume without loss of generality that ff is a simplicial map. This can be arranged by the simplicial approximation theorem. Let p~K~\widetilde{p}\in\widetilde{K} be a point lying over pKp\in K. Then, we claim f~1(p~)\widetilde{f}^{-1}(\widetilde{p}) is isometric to f1(p)f^{-1}(p). Indeed, note that f1(p)Mf^{-1}(p)\hookrightarrow M induces the zero map on π1\pi_{1}, since otherwise there would be a loop γf1(p)\gamma\subset f^{-1}(p) which is not nullhomotopic in MM such that f[γ]=0π1(K)f_{*}[\gamma]=0\in\pi_{1}(K). This would be absurd, as ff induces an isomorphism on π1\pi_{1}. Thus, f1(p)f^{-1}(p) lifts homeomorphically in the cover to a unique subset in Y~\widetilde{Y} containing p~\widetilde{p}. By uniqueness, this lift is exactly f~1(p~)\widetilde{f}^{-1}(\widetilde{p}). Consequently, the covering projection must be a homeomorphism:

f~1(p~)f1(p)\widetilde{f}^{-1}(\widetilde{p})\to f^{-1}(p)

Since the metric on the cover is g~\widetilde{g}, obtained by lifting the metric gg on WW, the aforementioned homeomorphism is the desired isometry. Thus, the fibers of f~\widetilde{f} are isometric to the fibers of ff, which are uniformly bounded as EE is compact. Therefore, dimmcE~k\dim_{mc}\widetilde{E}\leq k. ∎

3.2. Positive scalar curvature and Donaldson divisors

We begin with the following result due to Cecchini, Räde and Zeidler in [CRZ23], which establishes a codimension-22 obstruction to existence of a PSC metric on a 77-manifold.

Theorem 2.

[CRZ23, Theorem 1.11] Let Y7Y^{7} be a closed connected manifold, and X5Y7X^{5}\subset Y^{7} be a submanifold such that

  1. (1)

    i:XYi:X\hookrightarrow Y has trivial normal bundle,

  2. (2)

    i:π1(X)π1(Y)i_{*}:\pi_{1}(X)\to\pi_{1}(Y) is injective,

  3. (3)

    i:π2(X)π2(Y)i_{*}:\pi_{2}(X)\to\pi_{2}(Y) is surjective

Then, if XX does not admit a PSC metric, YY does not admit a PSC metric either.

Let (M6,ω)(M^{6},\omega) be an integral symplectic 66-manifold. Let Z(M6,ω)Z\subset(M^{6},\omega) be a Donaldson divisor. Recall from Lemma 2.1 that the normal bundle of ZZ in MM is isomorphic to LkL^{\otimes k}, which is a nontrivial complex line bundle over ZZ. However, we will now consider the unit circle bundle of LkL^{\otimes k}, denoted by 𝕊(Lk)\mathbb{S}(L^{\otimes k}). In conjunction with it, we will consider the restriction of the unit circle bundle of LkL^{\otimes k} to ZZ, denoted by 𝕊(Lk|Z)\mathbb{S}(L^{\otimes k}|_{Z}). Note that 𝕊(Lk)\mathbb{S}(L^{\otimes k}) is a 77-manifold and 𝕊(Lk|Z)\mathbb{S}(L^{\otimes k}|_{Z}) is a 55-manifold.

In the following proposition, we shall verify that the pair (𝕊(Lk),𝕊(Lk|Z))(\mathbb{S}(L^{\otimes k}),\mathbb{S}(L^{\otimes k}|_{Z})) satisfies the hypotheses of Theorem 2.

Proposition 7.

Let us denote Y:=𝕊(Lk)Y:=\mathbb{S}(L^{\otimes k}) and X:=𝕊(Lk|Z)X:=\mathbb{S}(L^{\otimes k}|_{Z}). Then the pair (Y7,X5)(Y^{7},X^{5}) satisfies the hypotheses of Theorem 2. Therefore, if 𝕊(Lk)\mathbb{S}(L^{\otimes k}) is PSC, then so is 𝕊(Lk|Z)\mathbb{S}(L^{\otimes k}|_{Z}).

Before the proof of Proposition 7, we state and prove an elementary topology lemma.

Lemma 3.2.

Let π:EB\pi:E\to B be a complex line bundle, and 𝕊(E)E\mathbb{S}(E)\subset E be the unit circle bundle with respect to some choice of a fiberwise Riemannian metric on EE. Let πs:𝕊(E)B\pi_{s}:\mathbb{S}(E)\to B denote the bundle projection. Then, πsE\pi_{s}^{*}E is trivial over 𝕊(E)\mathbb{S}(E).

Proof.

By definition of the pullback construction,

πsE=𝕊(E)×BE\pi_{s}^{*}E=\mathbb{S}(E)\times_{B}E

The diagonal section σ:𝕊(E)𝕊(E)×BE\sigma:\mathbb{S}(E)\to\mathbb{S}(E)\times_{B}E defined by σ(e):=(e,e)\sigma(e):=(e,e) is a nowhere zero section of the circle bundle πsE𝕊(E)\pi_{s}^{*}E\to\mathbb{S}(E). Therefore, πsE\pi_{s}^{*}E is trivial over 𝕊(E)\mathbb{S}(E). ∎

Proof of Proposition 7.

First, we observe that the normal bundle of XYX\subset Y is trivial. Indeed, let E:=Lk|ZE:=L^{\otimes k}|_{Z} and πs:𝕊(E)=XZ\pi_{s}:\mathbb{S}(E)=X\to Z denote the bundle projection. Then,

NYX=πNMZπs(E)N_{Y}X=\pi^{*}N_{M}Z\cong\pi_{s}^{*}(E)

By Lemma 3.2, we conclude NYXN_{Y}X is trivial over 𝕊(E)=X\mathbb{S}(E)=X, as desired.

Next, we show ii induces an isomorphism on π1\pi_{1} and a surjection on π2\pi_{2}. To this end, consider the following commutative diagram:

S1{{S^{1}}}Y{{Y}}M{M}S1{{S^{1}}}X{{X}}Z{Z}id\scriptstyle{\mathrm{id}}i\scriptstyle{i}j\scriptstyle{j}

The rows are fiber bundles, therefore by applying the long exact sequence in homotopy and using naturality, we have the following diagram:

0{0}π2(Y){{\pi_{2}(Y)}}π2(M){{\pi_{2}(M)}}π1(S1)𝐙{{\pi_{1}(S^{1})\cong\mathbf{Z}}}π1(Y){{\pi_{1}(Y)}}π1(M){{\pi_{1}(M)}}0{0}0{0}π2(X){{\pi_{2}(X)}}π2(Z){{\pi_{2}(Z)}}π1(S1)𝐙{{\pi_{1}(S^{1})\cong\mathbf{Z}}}π1(X){{\pi_{1}(X)}}π1(Z){{\pi_{1}(Z)}}0{0}ϕM\scriptstyle{\phi_{M}}i\scriptstyle{i_{*}}j\scriptstyle{j_{*}}ϕZ\scriptstyle{\phi_{Z}}id\scriptstyle{\mathrm{id}}i\scriptstyle{i_{*}}j\scriptstyle{j_{*}}

Note, as indicated in the diagram above, j:π1(Z)π1(M)j_{*}:\pi_{1}(Z)\to\pi_{1}(M) is an isomorphism and j:π2(Z)π2(M)j_{*}:\pi_{2}(Z)\to\pi_{2}(M) is a surjection due to Proposition 2 specialized to the case of n=3n=3, as MM is a symplectic manifold of dimension 2n=62n=6.

  1. To show i:π2(X)π2(Y)i_{*}:\pi_{2}(X)\to\pi_{2}(Y) is surjective:

    Let απ2(Y)\alpha\in\pi_{2}(Y). Let απ2(M)\alpha^{\prime}\in\pi_{2}(M) be the image of α\alpha. Then ϕM(α)=0\phi_{M}(\alpha^{\prime})=0. Since j:π2(Z)π2(M)j_{*}:\pi_{2}(Z)\to\pi_{2}(M) is surjective, there exists βπ2(M)\beta^{\prime}\in\pi_{2}(M) such that j(β)=αj_{*}(\beta^{\prime})=\alpha^{\prime}. By commutativity,

    ϕZ(β)=ϕM(j(β))=ϕM(α)=0.\phi_{Z}(\beta^{\prime})=\phi_{M}(j_{*}(\beta^{\prime}))=\phi_{M}(\alpha^{\prime})=0.

    Therefore, β\beta^{\prime} lies in the kernel of ϕZ\phi_{Z}, which is equal to the image of π2(X)\pi_{2}(X) in π2(Z)\pi_{2}(Z). Thus, there exists βπ2(X)\beta\in\pi_{2}(X) such that image of β\beta is exactly β\beta^{\prime}. Now, the image of i(β)αi_{*}(\beta)-\alpha in π2(M)\pi_{2}(M) is j(β)α=0j_{*}(\beta^{\prime})-\alpha^{\prime}=0. Since π2(Y)\pi_{2}(Y) injects into π2(M)\pi_{2}(M), we conclude i(β)=αi_{*}(\beta)=\alpha. This proves the claim.

  2. To show i:π1(X)π1(Y)i_{*}:\pi_{1}(X)\to\pi_{1}(Y) is injective:

    Note that ϕM\phi_{M} and ϕZ\phi_{Z} have the same image and therefore the same cokernel. By the short five-lemma, i:π1(X)π1(Y)i_{*}:\pi_{1}(X)\to\pi_{1}(Y) is an isomorphism. In particular, it is injective. ∎

4. Proofs of the main results

4.1. Proof of Theorem A

In this section, we prove our main result. Let (M,η)(M,\eta) be an integral symplectic 44-manifold. Let ω0\omega_{0} be the standard area form on S2S^{2}. We equip M×S2M\times S^{2} with the integral symplectic form ω=ηω0\omega=\eta\oplus\omega_{0}. Let LL be the pre-quantum line bundle on the symplectic manifold (M×S2,ω)(M\times S^{2},\omega). Let Z(M×S2,ω)Z\subset(M\times S^{2},\omega) be a Donaldson divisor, with [Z]=PD(k[ω])[Z]=\mathrm{PD}(k\cdot[\omega]), for k1k\gg 1 sufficiently large. We set EZ:=𝕊(Lk|Z)E_{Z}:=\mathbb{S}(L^{\otimes k}|_{Z}) as the unit circle bundle of LkL^{\otimes k} restricted to ZZ (as introduced in Lemma 2.1 and Proposition 7).

  1. Proof of (1).

    As M×S2M\times S^{2} admits a PSC metric, and 𝕊(Lk)\mathbb{S}(L^{\otimes k}) is a circle bundle on M×S2M\times S^{2}, 𝕊(Lk)\mathbb{S}(L^{\otimes k}) also admits a PSC metric [BB83, Theorem C]. By Proposition 7, this implies that EZ=𝕊(Lk|Z)E_{Z}=\mathbb{S}(L^{\otimes k}|_{Z}) also admits a PSC metric.

  2. Proof of (2).

    Let ν(Z)\nu(Z) denote a tubular neighbourhood of ZM×S2Z\subset M\times S^{2}. Consider the manifold with boundary W:=(M×S2)ν(Z)W:=(M\times S^{2})\setminus\nu(Z). Note that W=ν(Z)\partial W=\partial\nu(Z). By Lemma 2.1, ν(Z)𝕊(Lk|Z)=EZ\partial\nu(Z)\cong\mathbb{S}(L^{\otimes k}|_{Z})=E_{Z}. Therefore WEZ\partial W\cong E_{Z}. In Corollary 3, we deduced that (W6,EZ)(W^{6},E_{Z}) satisfies the hypothesis of Lemma 3.1 with n=6n=6 and k=3k=3. Therefore, dimmcEZ~3\dim_{mc}\widetilde{E_{Z}}\leq 3.

  3. Proof of (3).

    By Proposition 5, the map f:ZMf:Z\to M is of positive degree. Moreover, by hypothesis b2+(M)>1b_{2}^{+}(M)>1. Since ff is of positive degree, the induced map

    f:H2(M;𝐐)H2(Z;𝐐)f^{*}:H^{2}(M;\mathbf{Q})\to H^{2}(Z;\mathbf{Q})

    is an injective map of quadratic spaces, where both domain and range are equipped with the quadratic form given by the cup product \smile. Therefore, b2+(Z)>1b_{2}^{+}(Z)>1. In [Tau94], Taubes proved that symplectic 44-manifolds with b2+>1b_{2}^{+}>1 have non-zero Seiberg-Witten invariant. Since having non-zero Seiberg-Witten invariant obstructs the existence of PSC metrics, this in particular implies ZZ does not admit a PSC metric.

  4. Proof of (4).

    By Proposition 5, the map f:ZMf:Z\to M is of positive degree. Moreover, by hypothesis MM admits a map of positive degree to an aspherical manifold. Composing these maps, we get a positive degree map from ZZ to an aspherical manifold. Then by [CLL23, Corollary 3], ZZ does not admit a PSC metric.

  5. Proof of (5).

    Since MM is symplectically aspherical, by Proposition 6, MM is a rationally essential manifold. Since the map f:ZMf:Z\to M if of positive degree by Proposition 5, ZZ must be rationally essential as well. Now, by Lemma 2,

    π1(Z)π1(M×S2)π1(M).\pi_{1}(Z)\cong\pi_{1}(M\times S^{2})\cong\pi_{1}(M).

    By hypothesis, π1(M)\pi_{1}(M) is amenable. Therefore, ZZ is a rationally essential manifold with amenable fundamental group. Hence, by [Dra11], dimmcZ=4\dim_{mc}Z=4.

4.2. Proof of Corollary B

Before we move to the proof of Corollary B we give several examples of 44-manifolds MM satisfying the hypotheses of Theorem A.

Example 8.

Let X4X^{4} be a T2T^{2}-bundle over T2T^{2}. Then M=X#m𝐂𝐏2¯M=X\#m\overline{\mathbf{CP}^{2}} satisfies the hypotheses required for Conclusions (4)(4) and (5)(5) of Theorem A.

Indeed, by a well-known observation of Thurston [Thu76], XX admits a symplectic structure. By the symplectic blowup construction due to Gompf [Gom95], MM also admits a symplectic structure. Note that MM is not aspherical for m>1m>1 but admits a degree 1 map onto the aspherical manifold T4T^{4}. The fundamental group π1(X)\pi_{1}(X) is a 𝐙2\mathbf{Z}^{2}-extension over 𝐙2\mathbf{Z}^{2}, therefore it is amenable.

As a trivial case of the above, one could consider X=T4X=T^{4}. Note that here we also have b2+=3>1b_{2}^{+}=3>1. Thus, MM satisfies the hypotheses required for Conclusions (3),(4)(3),(4) and (5)(5).

Example 9.

Let (X2n,ω)(X^{2n},\omega) be any symplectically aspherical manifold with amenable fundamental group. By repeated applications of Theorem 1 and Proposition 2, we can produce a symplectic submanifold M4X2nM^{4}\subset X^{2n} such that π1(M)π1(X)\pi_{1}(M)\cong\pi_{1}(X). Since XX is symplectically aspherical, so is MM. Therefore, MM satisfies the hypotheses required for Conclusion (5)(5) in Theorem A.

As an explicit example, one could take X=T6X=T^{6}. Considering the symplectic structure on T6T^{6} as the product T4×T2T^{4}\times T^{2}, we see that the projection MT4M\to T^{4} to the first factor must be a positive degree map by the same argument as in Proposition 5. Therefore, MM satisfies the hypotheses of Conclusions (3),(4)(3),(4) and (5)(5). One salient feature of this example is that π1(M)𝐙6\pi_{1}(M)\cong\mathbf{Z}^{6} has cohomological dimension 66. Therefore, the techniques in [BD16] are unable to provide apriori bounds on the macroscopic dimension of EZE_{Z}.

Example 10.

(Examples within examples) Take any manifold MM satisfying the hypotheses required for Conclusions (3),(4)(3),(4) and (5)(5) in Theorem A. Let ZM×S2Z\subset M\times S^{2} be a Donaldson divisor as in the statement of Theorem A. Then ZZ also satisfies the same hypotheses.

Proof of Corollary B.

Let MM be any manifold from Example 8, 9 or 10 above. Let k1k\gg 1 be sufficiently large. Let ZkM×S2Z_{k}\subset M\times S^{2} be the submanifold furnished by Donaldson’s theorem (Theorem 1), such that [Zk]=PD(k[ω])[Z_{k}]=\mathrm{PD}(k\cdot[\omega]). Let LL be the pre-quantum line bundle over MM, so that c1(L)=[ω]c_{1}(L)=[\omega]. Let EkE_{k} be the unit sphere bundle of LkL^{\otimes k} restricted to ZkZ_{k}. It follows from the proof of Theorem A that ZkZ_{k} and EkE_{k} satisfy Conclusions (1)(1) and (2)(2). Proposition 4 implies |χ(Zk)||\chi(Z_{k})|\to\infty as kk\to\infty, as desired. ∎

4.3. Proof of Corollary C

We begin by recalling the definition of enlargeable manifolds from Gromov-Lawson [GL83].

Definition 4.

A compact orientable Riemannian manifold MM of dimension nn is called ε1\varepsilon^{-1}-hyperspherical if there exists a ε\varepsilon-contracting map of positive degree from MM to the unit nn-sphere. Further, MM is called enlargeable if for every ε>0\varepsilon>0, there exists a finite cover of MM whch is ε1\varepsilon^{-1}-hyperspherical and spin.

The basic example of an enlargeable manifold is the unit circle S1𝐑/2π𝐙S^{1}\cong\mathbf{R}/2\pi\mathbf{Z}. Indeed, the nn-sheeted cover of 𝐑/2π𝐙\mathbf{R}/2\pi\mathbf{Z} is 𝐑/2πn𝐙\mathbf{R}/2\pi n\mathbf{Z}. The map 𝐑/2π𝐙𝐑/2πn𝐙\mathbf{R}/2\pi\mathbf{Z}\to\mathbf{R}/2\pi n\mathbf{Z} given as scaling by 1/n1/n is a 1/n1/n-contracting map, for any n1n\geq 1. Also, S1S^{1} is certainly a spin manifold. It is straightforward to see that

  1. (1)

    Product of enlargeable manifolds is enlargeable,

  2. (2)

    A manifold MM admitting a positive degree map to an enlargeable manifold is also enlargeable, provided MM admits a finite spin covering.

Therefore, TnS1××S1T^{n}\cong S^{1}\times\cdots\times S^{1} is enlargeable and any spin manifold admitting a positive degree map to TnT^{n} is also enlargeable. In [GL83], it is proved that all compact hyperbolic manifolds are enlargeable. The notion of enlargeability is important because of the main result of Gromov-Lawson [GL83, Theorem A], which shows that enlargeable manifolds do not admit PSC metrics.

Proof of Corollary C.

Let M=T4M=T^{4} as in Example 8 with m=0m=0. Let ω\omega denote the product symplectic form on T4×S2T^{4}\times S^{2}. Let k1k\gg 1 be a sufficiently large even integer. Then, by Theorem A, there exists a submanifold Z4(T4×S2,ω)Z^{4}\subset(T^{4}\times S^{2},\omega) such that [Z]=PD(k[ω])[Z]=\mathrm{PD}(k\cdot[\omega]), ZZ admits a positive degree map to T4T^{4}, and there exists a nontrivial circle bundle EZZE_{Z}\to Z such that the total space EZE_{Z} admits a metric of positive scalar curvature.

Since T4T^{4} is enlargeable and ZZ already admits a positive degree map to T4T^{4}, to show that ZZ is also enlargeable we just need to show that it is also spin, i.e. the second Stiefel-Whitney class is 0. Note that as ZZ is a symplectic manifold, the second Stiefel-Whitney class of ZZ is the first Chern class of ZZ modulo 22:

w2(Z)c1(Z)(mod2)w_{2}(Z)\equiv c_{1}(Z)\pmod{2}

By Equation (2.3) in Proposition 4, we have

c1(Z)=jc1(T4×S2)kj[ω]c_{1}(Z)=j^{*}c_{1}(T^{4}\times S^{2})-k\cdot j^{*}[\omega]

Next, we have c1(T4×S2)w1(T4×S2)0(mod2)c_{1}(T^{4}\times S^{2})\equiv w_{1}(T^{4}\times S^{2})\equiv 0\pmod{2}, as T4T^{4} and S2S^{2} are both spin, and product of spin manifolds is also spin. Since kk was moreover chosen to be even, we conclude c1(Z)0(mod2)c_{1}(Z)\equiv 0\pmod{2}. Thus, w2(Z)=0w_{2}(Z)=0. Hence, ZZ is spin. ∎

References

  • [ABG24] Hannah Alpert, Alexey Balitskiy, and Larry Guth. Macroscopic scalar curvature and codimension 2 width. J. Topol. Anal., 16(6):979–987, 2024.
  • [Bal21] Alexey Balitskiy. Bounds on Urysohn Width. PhD thesis, Massachusetts Institute of Technology, Cambridge, MA, 2021. Available at https://hdl.handle.net/1721.1/139312.
  • [BB83] L. Bérard-Bergery. Scalar curvature and isometry group. In Spectra of Riemannian Manifolds, pages 9–28. Kaigai Publications, 1983. Proc. Franco-Japanese seminar on Riemannian geometry (Kyoto, 1981).
  • [BD16] Dmitry Bolotov and Alexander Dranishnikov. On Gromov’s conjecture for totally non-spin manifolds. J. Topol. Anal., 8(4):571–587, 2016.
  • [BH10] Michael Brunnbauer and Bernhard Hanke. Large and small group homology. J. Topol., 3(2):463–486, 2010.
  • [CLL23] Otis Chodosh, Chao Li, and Yevgeny Liokumovich. Classifying sufficiently connected PSC manifolds in 4 and 5 dimensions. Geom. Topol., 27(4):1635–1655, 2023.
  • [CRZ23] Simone Cecchini, Daniel Räde, and Rudolf Zeidler. Nonnegative scalar curvature on manifolds with at least two ends. J. Topol., 16(3):855–876, 2023.
  • [DD22] Michelle Daher and Alexander Dranishnikov. On macroscopic dimension of non-spin 4-manifolds. J. Topol. Anal., 14(2):343–352, 2022.
  • [Don96] S. K. Donaldson. Symplectic submanifolds and almost-complex geometry. J. Differential Geom., 44(4):666–705, 1996.
  • [Dra11] A. N. Dranishnikov. Macroscopic dimension and essential manifolds. Tr. Mat. Inst. Steklova, 273(Sovremennye Problemy Matematiki):41–53, 2011.
  • [FCS80] Doris Fischer-Colbrie and Richard Schoen. The structure of complete stable minimal surfaces in 33-manifolds of nonnegative scalar curvature. Comm. Pure Appl. Math., 33(2):199–211, 1980.
  • [GL83] Mikhael Gromov and H. Blaine Lawson, Jr. Positive scalar curvature and the Dirac operator on complete Riemannian manifolds. Inst. Hautes Études Sci. Publ. Math., (58):83–196 (1984), 1983.
  • [Gom95] Robert E. Gompf. A new construction of symplectic manifolds. Ann. of Math. (2), 142(3):527–595, 1995.
  • [Gro96] M. Gromov. Positive curvature, macroscopic dimension, spectral gaps and higher signatures. In Functional analysis on the eve of the 21st century, Vol. II (New Brunswick, NJ, 1993), volume 132 of Progr. Math., pages 1–213. Birkhäuser Boston, Boston, MA, 1996.
  • [Gro18] Misha Gromov. Metric inequalities with scalar curvature. Geom. Funct. Anal., 28(3):645–726, 2018.
  • [Gro23] Misha Gromov. Four lectures on scalar curvature. In Perspectives in scalar curvature. Vol. 1, pages 1–514. World Sci. Publ., Hackensack, NJ, [2023] ©2023.
  • [Gut11] Larry Guth. Volumes of balls in large Riemannian manifolds. Ann. of Math. (2), 173(1):51–76, 2011.
  • [Gut24] Larry Guth, 2024. Personal communication.
  • [He23] Shihang He. Twisted S1S^{1} stability and positive scalar curvature obstruction on fiber bundles. arXiv preprint, 2023. Available at https://arxiv.org/abs/2303.12614.
  • [KR23] Thorben Kastenholz and Jens Reinhold. Simplicial volume and essentiality of manifolds fibered over spheres. J. Topol., 16(1):192–206, 2023.
  • [KRT08] Jarek Kȩdra, Yuli Rudyak, and Aleksy Tralle. Symplectically aspherical manifolds. J. Fixed Point Theory Appl., 3(1):1–21, 2008.
  • [LM23] Yevgeny Liokumovich and Davi 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.
  • [Nab22] Alexander Nabutovsky. Linear bounds for constants in Gromov’s systolic inequality and related results. Geom. Topol., 26(7):3123–3142, 2022.
  • [Ros07] Jonathan Rosenberg. Manifolds of positive scalar curvature: a progress report. In Surveys in differential geometry. Vol. XI, volume 11 of Surv. Differ. Geom., pages 259–294. Int. Press, Somerville, MA, 2007.
  • [Sab22] Stéphane Sabourau. Macroscopic scalar curvature and local collapsing. Ann. Sci. Éc. Norm. Supér. (4), 55(4):919–936, 2022.
  • [SY79] R. Schoen and S. T. Yau. On the structure of manifolds with positive scalar curvature. Manuscripta Math., 28(1-3):159–183, 1979.
  • [Tau94] Clifford Henry Taubes. The Seiberg-Witten invariants and symplectic forms. Math. Res. Lett., 1(6):809–822, 1994.
  • [Thu76] W. P. Thurston. Some simple examples of symplectic manifolds. Proc. Amer. Math. Soc., 55(2):467–468, 1976.