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

\stackMath

Gluing 2\mathbb{Z}_{2}-Harmonic Spinors and Seiberg–Witten Monopoles on 3-Manifolds

Gregory J. Parker Department of Mathematics, Stanford University [email protected]
Abstract.

Given a 2\mathbb{Z}_{2}-harmonic spinor satisfying some genericity assumptions, this article constructs a 1-parameter family of two-spinor Seiberg–Witten monopoles converging to it after renormalization. The proof is a gluing construction beginning with the model solutions from [41]. The gluing is complicated by the presence of an infinite-dimensional obstruction bundle for the singular limiting linearized operator. This difficulty is overcome by introducing a generalization of Donaldson’s alternating method in which a deformation of the 2\mathbb{Z}_{2}-harmonic spinor’s singular set is chosen at each stage of the alternating iteration to cancel the obstruction components.

1. Introduction

The Uhlenbeck compactification of the moduli space of anti-self-dual (ASD) Yang-Mills instantons on a compact 4-manifold exemplifies a philosophy for constructing natural compactifications of moduli spaces that is ubiquitous in modern differential geometry. The construction has two main steps: first, K. Uhlenbeck’s compactness theorem shows that any sequence AnA_{n} of ASD instantons subconverges either to another instanton, or to a limiting object consisting of a background instanton AA_{\infty} of lower charge and bubbling data BB [57, 58]. Second, C. Taubes’s gluing results reverse this process, showing that each pair (A,B)(A_{\infty},B) is the limit of smooth ASD instantons [47]. Together, these results allow the construction of boundary charts which endow the moduli space with the structure of a smoothly stratified manifold. This structure of the moduli space is the basis for the celebrated applications of Yang-Mills theory to 3 and 4-dimensional topology [7].

C. Taubes’s more recent extension of Uhlenbeck’s compactness theorem to PSL(2,)\text{PSL}(2,\mathbb{C}) connections introduced a new type of non-compactness [50], which has since been shown to be quite general in 3 and 4 dimensional gauge theories. It is exhibited by most generalized Seiberg–Witten (SW) equations [15, 61], a class of equations that includes the Kapuastin–Witten equations [54, 55], the Vafa–Witten equations [53], the complex ASD equations [49], the Seiberg–Witten equations with multiple spinors [32, 52], and the ADHM Seiberg–Witten equations [65]. For these equations, a sequence of solutions need not converge but, after renormalization, subconverges either to another solution or to the limiting data of a Fueter section – a solution of a different elliptic PDE that is usually degenerate, and in many cases non-linear [8, 20, 48].

It is natural to ask whether this more subtle limiting process can be reversed by a gluing construction. An affirmative answer would provide an essential step in constructing compactifications of the moduli spaces of solutions to generalized Seiberg–Witten equations, which are expected to be necessary to study the conjectured relations of these equations to the geometry of manifolds cf. [59, 13, 63, 64, 22, 33, 23, 14]. Such a gluing result would produce, from a given Fueter section Φ\Phi, a family of solutions to the corresponding generalized Seiberg–Witten equation that converges to Φ\Phi after renormalization.

The purpose of this article is to prove a gluing result of this form in the case of the two-spinor Seiberg–Witten equations on a compact 3-manifold, where the corresponding Fueter sections are 2\mathbb{Z}_{2}-harmonic spinors. In most cases, 2\mathbb{Z}_{2}-harmonic spinors possess a singular set which is stable under perturbations, along which the relevant linearized operator degenerates. This degenerate operator carries an infinite-dimensional obstruction bundle, making the gluing problem considerably more challenging than most gluing problems in the literature.

Despite the presence of the infinite-dimensional obstruction bundle, the gluing can still be accomplished for a set of parameters with finite codimension. Geometrically, the infinite-dimensional obstruction arises because the location of the singular set may vary during the limiting process. This freedom plays an important role in previous work of the author [43], R. Takahashi [46], and S. Donaldson [12] on the deformation theory of 2\mathbb{Z}_{2}-harmonic spinors. To account for it here, deformations of the 2\mathbb{Z}_{2}-harmonic spinor’s singular set are included as an infinite-dimensional gluing parameter. This leads to an infinite-dimensional family of Seiberg–Witten equations coupled to embeddings of the singular set; the first-order effect of deforming the singular set may then be calculated by differentiating this family with respect to the embedding. The crucial result that allows the gluing to succeed is that the linearized deformations of the singular set perfectly pair with the infinite-dimensional obstruction, allowing it to be cancelled.

Once this analytic set-up is in place, the gluing is accomplished by adapting Donaldson’s alternating method [9] to the semi-Fredholm setting. The starting point is an approximate solution constructed by splicing the model solution constructed in [41] onto a neighborhood of the singular set. The gluing iteration is then a three-step cycle, which adds corrections to the approximate solution near the singular set, then away from it, and at the start of each new cycle, deforms the singular set to cancel the obstruction. The iteration is complicated by the fact that the linearized deformation operator displays a loss of regularity, which necessitates refining the approach to deforming the singular set in [43, 46, 12] by introducing specially adapted families of smoothing operators. The end result of the iteration is a family of Seiberg–Witten solutions converging to a given 2\mathbb{Z}_{2}-harmonic spinor after renormalization. The framework developed here may also be useful for addressing other geometric problems requiring the deformation of a singular set [12, 26, 62, 11].

1.1. The Seiberg–Witten Equations

Let (Y,g)(Y,g) be a closed, oriented, Riemannian 3-manifold, and fix a Spinc\text{Spin}^{c}-structure with spinor bundle SYS\to Y. Choose a rank 2 complex vector bundle EYE\to Y with trivial determinant, and fix a smooth background SU(2)SU(2)-connection BB on EE. The two-spinor Seiberg–Witten equations are the following equations for pairs (Ψ,A)Γ(SE)×𝒜U(1)(\Psi,A)\in\Gamma(S\otimes_{\mathbb{C}}E)\times\mathcal{A}_{U(1)} of an EE-valued spinor and a U(1)U(1) connection on det(S)\text{det}(S):

ABΨ\displaystyle\not{D}_{AB}\Psi =\displaystyle= 0\displaystyle 0 (1.1)
FA+12μ(Ψ,Ψ)\displaystyle\star F_{A}+\tfrac{1}{2}\mu(\Psi,\Psi) =\displaystyle= 0\displaystyle 0 (1.2)

where AB\not{D}_{AB} is the twisted Dirac operator formed using BB on EE and the Spinc connection induced by AA on SS, FAF_{A} is the curvature of AA, and μ:SEΩ1(i)\mu:S\otimes E\to\Omega^{1}(i\mathbb{R}) is a pointwise-quadratic map. The equations are invariant under U(1)U(1)-gauge transformations. Solutions of (1.11.2) are called monopoles.

Unlike for the standard (one-spinor) Seiberg–Witten equations, sequences of solutions to (1.11.2) may lack subsequences where ΨL2\|\Psi\|_{L^{2}} remains bounded, thus no subsequences can converge. The renormalization procedure alluded to above simply scales this L2L^{2}-norm to unity by setting Φ=εΨ\Phi=\varepsilon\Psi where ε=1ΨL2\varepsilon=\tfrac{1}{\|\Psi\|_{L^{2}}}. The re-normalized equations become

ABΦ\displaystyle\not{D}_{AB}\Phi =\displaystyle= 0\displaystyle 0 (1.3)
ε2FA+12μ(Φ,Φ)\displaystyle\star\varepsilon^{2}F_{A}+\tfrac{1}{2}\mu(\Phi,\Phi) =\displaystyle= 0\displaystyle 0 (1.4)
ΦL2\displaystyle\|\Phi\|_{L^{2}} =\displaystyle= 1\displaystyle 1 (1.5)

and diverging sequences are now described by the degenerating family of equations with parameter ε0\varepsilon\to 0. A theorem of Haydys–Walpuski [32] (Theorem 3.2 in Section 3.1) shows that sequences of solutions for which ε0\varepsilon\to 0 must converge, in an appropriate sense, to a solution of the ε=0\varepsilon=0 equations, i.e. to a pair (Φ,A)(\Phi,A) where Φ\Phi is a normalized harmonic spinor with pointwise values in μ1(0)\mu^{-1}(0), which is a 5-dimensional cone. Up to gauge, such solutions are equivalent (via the Haydys Correspondence, Section 3.2) to harmonic spinors valued in a vector bundle up to a sign ambiguity. The latter are 2\mathbb{Z}_{2}-harmonic spinors, which are the simplest non-trivial type of Fueter section.

A key feature of convergence for a sequence (Φi,Ai,εi)(\Phi_{i},A_{i},\varepsilon_{i}) is the concentration of curvature, which gives rise to a singular set. As εi0\varepsilon_{i}\to 0, curvature may concentrate along a closed subset 𝒵\mathcal{Z} of Hausdorff codimension 2, so that the LpL^{p}-norm of FAiF_{A_{i}} diverges on any neighborhood of 𝒵\mathcal{Z} for p>1p>1. In fact, [24] shows that 𝒵\mathcal{Z} represents the Poincaré dual of c1(S)-c_{1}(S) in H1(Y;)H_{1}(Y;\mathbb{Z}), hence it is necessarily non-empty if the Spinc\text{Spin}^{c} structure is non-trivial. Away from 𝒵\mathcal{Z}, the connections converge to a flat connection with holonomy in 2\mathbb{Z}_{2}, the data of which is equivalent to that of a real Euclidean line bundle Y𝒵\ell\to Y-\mathcal{Z}. This limiting connection and line bundle may have holonomy around 𝒵\mathcal{Z} that is the remnant of the curvature that has “bubbled” away. If this holonomy is non-trivial, the Dirac equation twisted by such a limiting connection is singular along 𝒵\mathcal{Z}; a 2\mathbb{Z}_{2}-harmonic spinor is, more accurately, a solution of such a singular equation on the complement of 𝒵\mathcal{Z}.

1.2. 2\mathbb{Z}_{2}-Harmonic spinors

Let (Y,g)(Y,g) be as in Section 1.1, and now fix a spin structure with spinor bundle S0YS_{0}\to Y. Let B0B_{0} be a zeroth order, \mathbb{R}-linear perturbation to the spin connection that commutes with Clifford multiplication. Given a closed submanifold 𝒵Y\mathcal{Z}\subset Y of codimension 2, choose a real Euclidean line bundle Y𝒵\ell\to Y-\mathcal{Z} and let AA be the unique flat connection with 2\mathbb{Z}_{2}-holonomy on \ell. The twisted spinor bundle S0S_{0}\otimes_{\mathbb{R}}\ell carries a Dirac operator A\not{D}_{A} twisted by AA on \ell, and perturbed using B0B_{0}. A 2\mathbb{Z}_{2}-harmonic spinor is a solution ΦΓ(S0)\Phi\in\Gamma(S_{0}\otimes_{\mathbb{R}}\ell) of the twisted Dirac equation on Y𝒵Y-\mathcal{Z} satisfying

AΦ=0andAΦL2.\not{D}_{A}\Phi=0\hskip 28.45274pt\text{and}\hskip 28.45274pt\nabla_{A}\Phi\in L^{2}. (1.6)

𝒵\mathcal{Z} is called the singular set of the 2\mathbb{Z}_{2}-harmonic spinor. A 2\mathbb{Z}_{2}-harmonic spinor is denoted by the triple (𝒵,A,Φ)(\mathcal{Z},A,\Phi) where 𝒵\mathcal{Z} is the singular set, AA is the unique flat connection determined by the line bundle \ell, and Φ\Phi is the spinor itself.

When 𝒵=\mathcal{Z}=\emptyset is empty, solutions of (1.6) are classical harmonic spinors whose study goes back to the work of Lichnerowicz [35], Atiyah-Singer [3], and Hitchin [27]. When the singular set is non-empty, the twisted Dirac operator is a type of degenerate elliptic operator known as an elliptic edge operator. This class of operators has little precedent in gauge theory, but extensive tools for their study have been developed in microlocal analysis [36, 39, 18]. Doan–Walpuski established the existence and abundance of solutions with 𝒵\mathcal{Z}\neq\emptyset on compact 3-manifolds in [16], and a stronger version of their result will appear in [31]. Additional examples have been constructed in [56, 25, 29].

The equations (1.6) do not carry an action of U(1)U(1)-gauge transformations; there is only a residual action of 2\mathbb{Z}_{2} by sign. In particular, (1.6) is not a gauge theory. On the other hand (1.6) is \mathbb{R}-linear, so admits a scaling action by +\mathbb{R}^{+} (note that the assumption that the perturbation B0B_{0} is \mathbb{R}-linear means that the Dirac operator is also only \mathbb{R}-linear). 2\mathbb{Z}_{2}-harmonic spinors are considered as equivalence classes modulo these two actions; the scaling action is eliminated by fixing the normalization condition (1.5, leaving a residual 2\mathbb{Z}_{2} action. Notice that (1.6) defines 2\mathbb{Z}_{2}-harmonic spinors without reference to the Seiberg–Witten equations; it is therefore not a priori clear whether an arbitrary 2\mathbb{Z}_{2}-harmonic spinor (1.6) should arise as a limit of (1.3)–(1.5).

1.3. The Gluing Problem

The gluing problem may now be stated more precisely. Fix a 2\mathbb{Z}_{2}-harmonic spinor (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}). The goal is to construct a family of solutions (Φε,Aε)(\Phi_{\varepsilon},A_{\varepsilon}) to (1.31.5) for sufficiently small ε>0\varepsilon>0 such that

(Φε,Aε)(𝒵0,A0,Φ0)(\Phi_{\varepsilon},A_{\varepsilon})\longrightarrow(\mathcal{Z}_{0},A_{0},\Phi_{0}) (1.7)

in the sense of Haydys–Walpuski’s compactness theorem (Theorem 3.2). In particular, this requires reconstructing a smooth EE-valued spinor Φε\Phi_{\varepsilon} (note Φ0\Phi_{0} is a section of a different bundle of real rank 4 over Y𝒵Y-\mathcal{Z}), and re-introducing the highly concentrated curvature by smoothing the singular connection A0A_{0}. The latter implicitly requires recovering the Spinc\text{Spin}^{c}-structure, which is lost in the limit as ε0\varepsilon\to 0.

In the simplest case, when 𝒵0=\mathcal{Z}_{0}=\emptyset and standard elliptic theory applies, Doan–Walpuski [15] showed that all classical harmonic spinors arise as limits of a family as in (1.7). Reversing the convergence by a gluing in the singular case 𝒵0\mathcal{Z}_{0}\neq\emptyset is far more challenging, and requires new analytic tools for elliptic edge operators and their desingularizations. The concentration of curvature along the singular set 𝒵0\mathcal{Z}_{0} as ε0\varepsilon\to 0 manifests by making the linearization of the ε=0\varepsilon=0 version of (1.3)–(1.5) a singular elliptic edge operator with an infinite-dimensional obstruction to solving. This prevents the application of the standard Fredholm approaches that have historically been used in gluing problems [7, 34, 38, 10].

The presence of the infinite-dimensional obstruction bundle does not mean that the gluing can only be accomplished for a subset of parameters of infinite codimension. Rather, this obstruction is an artifact arising from inadvertently fixing the singular set 𝒵0\mathcal{Z}_{0}. In fact, the location of the singular set is a degree of freedom that may also vary as ε0\varepsilon\to 0. Indeed, work of the author [43] and R. Takahashi [46] has shown that constructing families of 2\mathbb{Z}_{2}-harmonic spinors with respect to families of metrics gsg_{s} for s𝒮s\in\mathcal{S} requires allowing the singular set to depend on ss. Because the gluing problem is a de-singularization of the same situation, one anticipates the same phenomenon will occur. It is therefore necessary to include space of all possible singular sets nearby 𝒵0\mathcal{Z}_{0} as a parameter in the gluing construction. This approach has some precedent in the work of Pacard–Ritoré [44] on gluing problems in minimal surface theory arising from the Allen-Cahn and Yang-Mills-Higgs equations [5], though the obstruction in these situations is more tractable.

As explained in the introduction, the main idea of this paper is to show that the deformations of the singular set pair with the infinite-dimensional obstruction to create a Fredholm gluing theory. Key aspects of this approach rely on the theory developed in [41], and [43], and this article is in some sense the sequel to and culmination of these. The first, [43] develops the deformation theory for the singular set for 2\mathbb{Z}_{2}-harmonic spinors alone, without reference to Seiberg–Witten theory. The second, [41] constructs model solutions near the singular set 𝒵\mathcal{Z}, which are the starting point of the gluing construction. To keep this article self-contained, the relevant parts of [43] and [41] are reviewed in detail in Sections 46 and Section 7, respectively.

1.4. Main Results

To state the main result, we first describe several necessary assumptions on the starting data.

Let YY be a compact, oriented three-manifold. The two-spinor Seiberg–Witten equations (1.11.2) depend on a smooth background parameter pair p=(g,B)p=(g,B) of a Riemannian metric and an auxiliary SU(2)SU(2)-connection on EE in the space

𝒫={(g,B)|gMet(Y),B𝒜SU(2)(E)}\mathcal{P}=\{(g,B)\ |\ g\in\text{Met}(Y)\ ,\ B\in\mathcal{A}_{SU(2)}(E)\} (1.8)

of all such choices. The definition of a 2\mathbb{Z}_{2}-harmonic spinor makes reference to a similar parameter pair (g,B)(g,B) where BB is now a perturbation to the spin connection on S0S_{0} that is inherited from the SU(2)SU(2) connection denoted by the same symbol. The gluing construction can be carried out beginning from a 2\mathbb{Z}_{2}-harmonic spinor satisfying several conditions; these conditions constrain the parameters to lie in the complement of the locus 𝒫𝒫\mathcal{P}^{\prime}\subset\mathcal{P} admitting 2\mathbb{Z}_{2}-harmonic spinors with worse singular behavior.

Definition 1.1.

A 2\mathbb{Z}_{2}-harmonic spinor (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) with respect to a parameter pair p0=(g0,B0)p_{0}=(g_{0},B_{0}) is said to be regular if it satisfies the following three conditions:

  1. (i)

    (Smooth) the singular set 𝒵0Y\mathcal{Z}_{0}\subset Y is a smooth, embedded link, and the holonomy of A0A_{0} is equal to 1-1 around the meridian of each component.

  2. (ii)

    (Isolated) Φ0\Phi_{0} is the unique 2\mathbb{Z}_{2}-harmonic spinor for the pair (𝒵0,A0)(\mathcal{Z}_{0},A_{0}) with respect to p0=(g0,B0)p_{0}=(g_{0},B_{0}) up to normalization and sign.

  3. (iii)

    (Non-degenerate) Φ0\Phi_{0} has non-vanishing leading-order, i.e. there is a constant c>0c>0 such that

    |Φ0|cdist(,𝒵0)1/2.|\Phi_{0}|\geq c\cdot\text{dist}(-,\mathcal{Z}_{0})^{1/2}.

For a fixed singular set 𝒵0\mathcal{Z}_{0}, the twisted Dirac operator,

A0:H1(S0)L2(S0)\not{D}_{A_{0}}:H^{1}(S_{0}\otimes_{\mathbb{R}}\ell)\longrightarrow L^{2}(S_{0}\otimes_{\mathbb{R}}\ell) (1.9)

is semi-Fredholm, where H1H^{1} denotes the Sobolev space of sections whose covariant derivative is L2L^{2} with appropriate weights, and S0S_{0} is the spinor bundle of the spin structure hosting Φ0\Phi_{0} (see Section 4 for details). For weights requiring that a solution satisfy the integrability requirement in (1.6), this operator is left semi-Fredholm and has infinite-dimensional cokernel. This cokernel gives rise to the infinite-dimensional obstruction of the linearized Seiberg–Witten equations at ε=0\varepsilon=0.

As explained above, cancelling the obstruction requires deforming the singular set. Any singular set 𝒵\mathcal{Z} nearby 𝒵0\mathcal{Z}_{0} defines a flat connection A𝒵A_{\mathcal{Z}} whose holonomy is homotopic to A0A_{0}, thus it defines an accompanying real line bundle isomorphic to the original. Allowing the singular set to vary over the space of embedded links 𝒵\mathcal{Z} gives an infinite-dimensional family of Dirac operators parameterized by embeddings, which we combine into a universal Dirac operator

(𝒵,Φ)=A𝒵Φ,\not{\mathbb{D}}(\mathcal{Z},\Phi)=\not{D}_{A_{\mathcal{Z}}}\Phi, (1.10)

which is fully non-linear in the first argument and linear in the second. The theory of the universal Dirac operator was developed in detail in [43] and is reviewed in Section 6.

The idea that the derivative of (1.10) with respect to the embedding should cancel the cokernel was investigated for the twisted Dirac operator alone (as opposed to the Seiberg–Witten equations) in [43] (see also the work of Takahashi [46] and Donaldson [12], which take a different approach). The first main result of [43] is the following theorem, which makes precise the idea that the linearized deformations of the singular set at a 2\mathbb{Z}_{2}-harmonic spinor cancel the obstruction.

Theorem 1.2.

([43, Thm 1.3]) Let (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) be a regular 2\mathbb{Z}_{2}-harmonic spinor, and let Π0\Pi_{0} denote the projection onto the cokernel of (1.9). Then the cokernel component of the linearization of the universal Dirac operator with respect to deformations of 𝒵0\mathcal{Z}_{0}

Π0(d𝒵0)(𝒵0,Φ0):L2,2(𝒵0;N𝒵0)Coker(𝒵0)\Pi_{0}\circ(\text{d}_{\mathcal{Z}_{0}}\not{\mathbb{D}})_{(\mathcal{Z}_{0},\Phi_{0})}:L^{2,2}(\mathcal{Z}_{0};N\mathcal{Z}_{0})\longrightarrow\text{Coker}(\not{D}_{\mathcal{Z}_{0}}) (1.11)

is an elliptic pseudo-differential operator of order 12\tfrac{1}{2} and its Fredholm extension has index 1-1. ∎

Here, the domain is thought of as the tangent space at 𝒵0\mathcal{Z}_{0} to the space of embedded singular sets of Sobolev regularity (2,2)(2,2). Section 5 gives a precise characterization of the cokernel, which results in an equivalence Coker(𝒵0)Γ(𝒵0;𝒮)\text{Coker}(\not{D}_{\mathcal{Z}_{0}})\simeq\Gamma(\mathcal{Z}_{0};\mathcal{S}) with the space of sections of a vector bundle 𝒮𝒵0\mathcal{S}\to\mathcal{Z}_{0}. Viewed in this guise, (1.11) is a map between spaces of sections of vectors bundles on 𝒵0\mathcal{Z}_{0}, and the assertion that it is an elliptic pseudo-differential operator then has the standard meaning.

Definition 1.3.

A 2\mathbb{Z}_{2}-harmonic spinor is unobstructed if the Fredholm extension of (1.11) has trivial kernel.

Since the Seiberg-Witten equations in dimension 3 have index 0, one does not expect 1-parameter families of solutions (Φε,Aε)(\Phi_{\varepsilon},A_{\varepsilon}) converging to a 2\mathbb{Z}_{2}-harmonic spinor for the fixed parameter p0=(g0,B0)p_{0}=(g_{0},B_{0}), but rather for a 1-dimensional family of parameters pτ=(gτ,Bτ)p_{\tau}=(g_{\tau},B_{\tau}) which coincides with p0p_{0} at τ=0\tau=0. The third condition necessary for the gluing result is a requirement on such a 1-parameter family, and says that (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) arises as a transverse spectral crossing in this family.

The following theorem, proved in [43], ensures that this notion makes sense in an analogous way to classical harmonic spinors in three dimensions. The added complexity here is that the singular set 𝒵τ\mathcal{Z}_{\tau} of the spinor is defined implicitly as a function of the parameter τ\tau. The proof of the theorem uses the linearized result of Theorem 1.2 and the Nash-Moser Implicit Function Theorem.

Theorem 1.4.

([43] Corollary 1.5) Suppose that pτ=(gτ,Bτ)p_{\tau}=(g_{\tau},B_{\tau}) is a smooth path of parameters, and that (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) is a regular, unobstructed 2\mathbb{Z}_{2}-harmonic spinor with respect to p0p_{0}. Then, there is a τ0>0\tau_{0}>0 such that for τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}), there exist 2\mathbb{Z}_{2}-eigenvectors (𝒵τ,Aτ,Φτ)(\mathcal{Z}_{\tau},A_{\tau},\Phi_{\tau}) with eigenvalues Λτ\Lambda_{\tau}\in\mathbb{R} satisfying

AτΦτ=ΛτΦτ,\not{D}_{A_{\tau}}\Phi_{\tau}=\Lambda_{\tau}\Phi_{\tau}, (1.12)

where each member of the tuples is defined implicitly as a smooth function of τ\tau. Moreover, (𝒵τ,Aτ,Φτ)(\mathcal{Z}_{\tau},A_{\tau},\Phi_{\tau}) are regular and unobstructed for each τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}).

Definition 1.5.

The family of Dirac operators Aτ\not{D}_{A_{\tau}} is said to have transverse spectral crossing at τ=0\tau=0 if the family of 2\mathbb{Z}_{2}-eigenvectors (𝒵τ,Aτ,Φτ,Λτ)(\mathcal{Z}_{\tau},A_{\tau},\Phi_{\tau},\Lambda_{\tau}) has

Λ˙(0)0,\dot{\Lambda}(0)\neq 0,

where ˙\dot{\ } denotes the derivative with respect to τ\tau.

We now state the main result, which establishes the existence of two-spinor Seiberg-Witten solutions converging to a 2\mathbb{Z}_{2}-harmonic spinor satisfying the above conditions. Here, S0S_{0} denotes the spinor bundle of a Spin structure that hosts a 2\mathbb{Z}_{2}-harmonic spinor as in (1.6), while SS is reserved for the spinor bundle of the Spinc\text{Spin}^{c}-structure in the Seiberg–Witten equations.

Theorem 1.6.

Suppose that (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) is a regular, unobstructed 2\mathbb{Z}_{2}-harmonic spinor with respect to a parameter p0=(g0,B0)p_{0}=(g_{0},B_{0}), and that pτ=(gτ,Bτ)p_{\tau}=(g_{\tau},B_{\tau}) is a path of parameters such that the corresponding family (1.12) has transverse spectral crossing.

Then, for each orientation of 𝒵0\mathcal{Z}_{0}, there is a unique Spinc\text{Spin}^{c} structure with spinor bundle SYS\to Y, an ε0>0\varepsilon_{0}>0, and a family of configurations (Ψε,Aε)(\Psi_{\varepsilon},A_{\varepsilon}) for ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}) satisfying the following.

  • (1)

    The Spinc\text{Spin}^{c} structure is characterized by

    c1(S)=PD[𝒵0]and S|Y𝒵0S0.c_{1}(S)=-\text{PD}[\mathcal{Z}_{0}]\hskip 28.45274pt\text{and }\hskip 28.45274ptS|_{Y-\mathcal{Z}_{0}}\simeq S_{0}\otimes_{\mathbb{R}}\ell.
  • (2)

    The configurations (Ψε,Aε)Γ(SE)×𝒜U(1)(\Psi_{\varepsilon},A_{\varepsilon})\in\Gamma(S_{E})\times\mathcal{A}_{U(1)} solve two-spinor Seiberg-Witten equations

    AεΨε\displaystyle\not{D}_{A_{\varepsilon}}\Psi_{\varepsilon} =\displaystyle= 0\displaystyle 0
    FAε+12μ(Ψε,Ψε)\displaystyle\star F_{A_{\varepsilon}}+\tfrac{1}{2}{\mu(\Psi_{\varepsilon},\Psi_{\varepsilon})} =\displaystyle= 0\displaystyle 0

    on YY with respect to (gτ,Bτ)(g_{\tau},B_{\tau}) where τ=τ(ε)\tau=\tau(\varepsilon) is defined implicitly as a function of ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}) and satisfies either τ(ε)>0\tau(\varepsilon)>0 or τ(ε)<0\tau(\varepsilon)<0.

  • (3)

    The spinor has L2L^{2}-norm

    ΨεL2(Y)=1ε,\|\Psi_{\varepsilon}\|_{L^{2}(Y)}=\frac{1}{\varepsilon}, (1.13)

    and after renormalizing by setting Φε=εΨε\Phi_{\varepsilon}=\varepsilon\Psi_{\varepsilon}, the solutions converge to (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}), i.e.

    ΦεΦ0andAεA0\Phi_{\varepsilon}\to\Phi_{0}\hskip 28.45274pt\text{and}\hskip 28.45274ptA_{\varepsilon}\to A_{0}

    in Cloc(Y𝒵0)C^{\infty}_{loc}(Y-\mathcal{Z}_{0}) after applying gauge transformations defined on Y𝒵0Y-\mathcal{Z}_{0}, and |Φε||Φ0||\Phi_{\varepsilon}|\to|\Phi_{0}| in C0,α(Y)C^{0,\alpha}(Y) for some α>0\alpha>0.

Remark 1.7.

It is expected that all 2\mathbb{Z}_{2}-harmonic spinors are regular and unobstructed for generic parameters among those admitting 2\mathbb{Z}_{2}-harmonic spinors (see Section 1.5). We do not undertake the task of establishing the genericity results here. A partial result for the genericity of the non-degeneracy condition in Definition 1.1 is proved in [25] in the situation of 2\mathbb{Z}_{2}-harmonic 1-forms. It is straightforward to show (see Remark 1.11) that a generic path (gτ,Bτ)(g_{\tau},B_{\tau}) has transverse spectral crossing.

Remark 1.8.

By a simple diagonalization argument, Theorem 1.6 may be extended to the case of a 2\mathbb{Z}_{2}-harmonic spinor that is instead a limit of 2\mathbb{Z}_{2}-harmonic spinors satisfying the hypotheses of the theorem. In particular, the singular set of such a limiting 2\mathbb{Z}_{2}-harmonic spinor need not be embedded.

The discussion in the upcoming Section 1.5 suggests it is likely that every isolated 2\mathbb{Z}_{2}-harmonic spinor arises as such a limit. If this were the case, Theorem 1.6 would imply every isolated 2\mathbb{Z}_{2}-harmonic spinor on a compact 3-manifold arises as the limit of Seiberg–Witten monopoles (in fact in multiple ways—one for each Spinc\text{Spin}^{c} structure whose first chern class is Poincaré dual to some orientation of the singular set).

Remark 1.9.

The solutions (Ψε,Aε)(\Psi_{\varepsilon},A_{\varepsilon}) have curvature that is highly concentrated in a O(ε2/3)O(\varepsilon^{2/3}) neighborhood around a smooth embedded curve 𝒵ε\mathcal{Z}_{\varepsilon} that lies in a neighborhood UU of the singular set 𝒵τ(ε)\mathcal{Z}_{\tau(\varepsilon)}. Slightly surprisingly, the proof shows that the C0C^{0}-distance from 𝒵ε\mathcal{Z}_{\varepsilon} to 𝒵τ(ε)\mathcal{Z}_{\tau(\varepsilon)} is smaller than the concentration scale of the curvature, while C1C^{1}-distance is larger than the concentration scale. For the specific parameters used in the proof of Theorem 1.6, one finds that 𝒵ε\mathcal{Z}_{\varepsilon} lies in a O(ε23/24γ)O(\varepsilon^{23/24-\gamma}) neighborhood of 𝒵τ(ε)\mathcal{Z}_{\tau(\varepsilon)} in C0C^{0} and a O(ε11/24γ)O(\varepsilon^{11/24-\gamma}) neighborhood in C1C^{1} for some γ<<1\gamma<<1.

1.5. Wall-Crossing Formulas

The non-compactness of the moduli space SW\mathcal{M}_{SW} of solutions to (1.3)–(1.5) prevents the (signed) count of two-spinor Seiberg–Witten monopoles from being a topological invariant. In particular, the compactness theorem (Theorem 3.2) suggests that along a path of parameters pτp_{\tau} such that p0p_{0} admits a 2\mathbb{Z}_{2}-harmonic spinor, a family of monopoles may diverge so that the signed count of solutions changes; in fact, Theorem 1.6 shows that this necessarily happens if the 2\mathbb{Z}_{2}-harmonic spinor is regular and unobstructed.

Rather than being a topological invariant, it is conjectured that signed count #SW\#\mathcal{M}_{SW} is a chambered invariant with wall-crossing formulas. That is, it is conjectured that the subset 𝒲2𝒫\mathcal{W}_{\mathbb{Z}_{2}}\subseteq\mathcal{P} of parameters admitting 2\mathbb{Z}_{2}-harmonic spinors has codimension 1, and divides its complement in 𝒫\mathcal{P} into a collection of open chambers inside which the count is invariant, with a well-defined formula for how the count changes as it crosses the “wall” 𝒲2\mathcal{W}_{\mathbb{Z}_{2}}. This chambered invariant is conjectured to fit into a larger scheme of constructing invariants by summing chambered invariants with cancelling wall-crossing formulas (see [13, 33, 23, 14] for details and examples).

The main result of [43] provides a step towards confirming this picture by proving that 𝒲2\mathcal{W}_{\mathbb{Z}_{2}} indeed forms a “wall” near regular, unobstructed 2\mathbb{Z}_{2}-harmonic spinors.

Theorem 1.10.

( [43, Thm 1.4]) Suppose that (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) is a regular, unobstructed 2\mathbb{Z}_{2}-harmonic spinor with respect to p0𝒫p_{0}\in\mathcal{P}. Then there is an open neighborhood 𝒰\mathcal{U} of p0p_{0} such that ,

𝒲2𝒰𝒫\mathcal{W}_{\mathbb{Z}_{2}}\cap\mathcal{U}\subseteq\mathcal{P}

is locally a smooth Fréchet submanifold of codimension 1.

More generally, it is expected that 𝒲2\mathcal{W}_{\mathbb{Z}_{2}} is a stratified space with the following global structure. The top stratum 𝒲2reg\mathcal{W}{{}^{\text{reg}}}_{\mathbb{Z}_{2}} should consist of a disconnected Fréchet submanifold of codimension 1 whose components are labeled by isotopy classes of embedded links in YY, and where each parameter p𝒲2regp\in\mathcal{W}_{\mathbb{Z}_{2}}^{\text{reg}} admits a (unique) regular, unobstructed 2\mathbb{Z}_{2}-harmonic spinor. Confirming this expectation would, in particular, require confirming the prediction of Taubes that the singular set of a 2\mathbb{Z}_{2}-harmonic spinor is smooth for generic choices of smooth parameters [50, pg. 9]. Deeper strata of higher finite codimension are expected to consist of the locus where the singular set has the structure of an embedded graph with increasingly complicated self-intersections, and the loci where the regular or unobstructed condition fails. There should also be strata of infinite codimension where wilder singular behavior can occur. The work of [56, 16, 29, 28] support this picture.

Theorem 1.6 shows that the count #SW\#\mathcal{M}_{SW} changes along a path of parameters crossing 𝒲2reg\mathcal{W}^{\text{reg}}_{\mathbb{Z}_{2}} transversely, which provides a key step in confirming the conjectured wall-crossing formula. In particular, Theorem 1.6 constructs a subset of the parameterized moduli space over pτp_{\tau} for either τ0\tau\geq 0 or τ0\tau\leq 0 that is homeomorphic to a half-open interval [0,ε0)[0,\varepsilon_{0}). A complete proof of a wall-crossing formula would additionally require investigating orientations to determine the sign of the crossing as in [16], and showing that the homeomorphism to [0,ε0)[0,\varepsilon_{0}) determines a boundary chart for the moduli space. The latter involves showing the surjectivity of gluing, i.e. that the family of monopoles in Theorem 1.6 is the unique family converging to (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}), and will be the subject of future work.

Remark 1.11.

A path pτ=(gτ,Bτ)p_{\tau}=(g_{\tau},B_{\tau}) has transverse spectral crossing (Definition 1.5) at a regular, unobstructed 2\mathbb{Z}_{2}-harmonic spinor (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) if and only if pτ𝒲2regp_{\tau}\pitchfork\mathcal{W}_{\mathbb{Z}_{2}}^{\text{reg}} is a transverse intersection. The genericity of this condition follows easily from the latter characterization.

1.6. Outline

The article has 12 sections which build towards the proof of Theorem 1.6.

The proof is accomplished by generalizing Donaldson’s alternating method to the semi-Fredholm setting. Section 2 is a self-contained introduction to gluing using the alternating method and describes this generalization. Briefly, the alternating method decomposes the manifold into two regions

Y=Y+YY=Y^{+}\cup Y^{-}

each of which admits a model solution. These model solutions are spliced into a global approximate solution, which is then corrected to a true solution by alternating making corrections localized on Y+Y^{+} and YY^{-} using the linearized equations. We take Y+=N(𝒵0)Y^{+}=N(\mathcal{Z}_{0}) to be a tubular neighborhood of the singular set, and YY^{-} to be the complement of a slightly smaller tubular neighborhood.

In our semi-Fredholm setting, the linearized equations on YY^{-} have an infinite-dimensional obstruction coming from the cokernel of (1.9), and an addition step cancelling this obstruction by deforming the singular set is required. Thus the alternation becomes a three-step cyclic iteration:

{error with supp. on Y}\left\{\begin{matrix}\text{error with \ }\\ \text{supp. on }Y^{-}\end{matrix}\right\}{error on Yobstruction}\left\{\begin{matrix}\text{error on }Y^{-}\\ \perp\text{obstruction}\end{matrix}\right\}{error with supp. on Y+}.\left\{\begin{matrix}\text{error with \ }\\ \text{supp. on }Y^{+}\end{matrix}\right\}.deform 𝒵\mathcal{Z}solve on Y+\begin{matrix}\text{solve on $Y^{+}$}\end{matrix}solve on YY^{-} (1.14)

Sections 312 are devoted to setting up and carrying out this cyclic iteration scheme. In this, there are several key technical difficulties (described in detail in Section 2.5) that conspire to make naive attempts at the iteration (1.14) fail to converge. The most important of these is that the deformation operator (1.11) displays a loss of regularity [19]. Overcoming this loss of regularity requires developing a more sophisticated way of deforming singular sets than was used in [43, 46, 12]. These new deformations depend on the spectral decomposition of a linearized deformation. Somewhat surprisingly, this approach eliminates the need to use tame Fréchet spaces and Nash-Moser theory, which are the standard tools used to deal with a loss of regularity as in [43, 46, 12]. Section 2 ends with a glossary of notation.

Section 3 covers background necessary to set up the gluing argument, beginning by reviewing the compactness theorem for (1.3)–(1.5) and the Haydys Correspondence.

Sections 47 analyze the linearized Seiberg–Witten equations on Y±Y^{\pm}. Section 4 begins by studying the singular Dirac operator (1.9) for a fixed singular set, which appears as a direct summand of the linearized Seiberg–Witten equations on YY^{-}.

Section 5 considers the infinite-dimensional obstruction of (1.9). In particular, it is shown that there is a natural basis of the obstruction indexed by the eigenvalues of a 1-dimensional Dirac operator on 𝒵\mathcal{Z}, whose pointwise norms increasingly concentrate near 𝒵\mathcal{Z} as the eigenvalue increases.

Section 6 deals with the universal Dirac operator (1.10). The main goal is to calculate the derivative with respect to the embedding of the singular set to give a refined version of Theorem 1.2. This is done by defining a family of diffeomorphisms Fη:YYF_{\eta}:Y\to Y parameterized by linearized deformations ηTEmb(𝒵0;Y)\eta\in T\text{Emb}(\mathcal{Z}_{0};Y) such that the curves 𝒵η=Fη(𝒵0)\mathcal{Z}_{\eta}=F_{\eta}(\mathcal{Z}_{0}) form an open neighborhood of 𝒵0\mathcal{Z}_{0} in a sufficient space of embeddings. By the naturality of the Dirac operator, varying the singular set over curves 𝒵η\mathcal{Z}_{\eta} while fixing the metric g0g_{0} is equivalent to varying the metric over the family of pullback metrics gη=Fη(g0)g_{\eta}=F_{\eta}^{*}(g_{0}) while keeping 𝒵0\mathcal{Z}_{0} fixed:

(varying 𝒵ηfixed g0)𝒵𝒵Fηg𝒵0g(varying gηfixed 𝒵0).\begin{pmatrix}\text{varying }\mathcal{Z}_{\eta}\\ \text{fixed }g_{0}\end{pmatrix}\hskip 28.45274pt\frac{\partial}{\partial\mathcal{Z}}\not{D}_{\mathcal{Z}}\ \ \ \overset{F_{\eta}^{*}}{\Longleftrightarrow}\ \ \ \frac{\partial}{\partial g}\not{D}_{\mathcal{Z}_{0}}^{g}\hskip 28.45274pt\begin{pmatrix}\text{varying }g_{\eta}\\ \text{fixed }\mathcal{Z}_{0}\end{pmatrix}. (1.15)

The derivative with respect to the family of pullback metrics on the right is then calculated using a result of Bourguignon-Gauduchon [6] (see Theorem 6.4).

Section 7 describes the model solutions on Y+Y^{+} constructed in [41] and summarizes results about the linearized Seiberg–Witten equations at these.

Section 8 constructs the infinite-dimensional family of model solutions parameterized by deformations 𝒵0\mathcal{Z}_{0} alluded to in the introduction. This family is used to define a universal version of the Seiberg–Witten equations analogous to (1.10), which is differentiated using the same trick as in (1.15). Section 9 shows that this derivative of the universal Seiberg–Witten equations with respect to deformations of 𝒵0\mathcal{Z}_{0} is simply a small perturbation of the derivative of the universal Dirac operator (1.10). This allows the deformations to be used to cancel the obstruction on YY^{-}, just as in Theorem 1.2.

Sections 1011 use the results of Sections 49 to carry out the cyclic iteration scheme (1.14). The iteration is shown to converge to a smooth two-parameter family of Seiberg–Witten “eigenvectors” solving

SW(Φ,A)=(Λ(τ)+μ(ε,τ))χΦτε\text{SW}(\Phi,A)=\Big{(}\Lambda(\tau)+\mu({\varepsilon,\tau})\Big{)}\chi\frac{\Phi_{\tau}}{\varepsilon} (1.16)

for every pair (ε,τ)(\varepsilon,\tau), where μ(ε,τ)\mu(\varepsilon,\tau)\in\mathbb{R}, Λ(τ)\Lambda(\tau) is as in Theorem 1.4, and χ\chi is a smooth function on YY. Thus (1.16) are solutions of the Seiberg–Witten equations precisely when the one-dimensional obstruction Λ(τ)+μ(ε,τ)\Lambda(\tau)+\mu(\varepsilon,\tau) vanishes. Finally, Section 12 uses a trick due to T. Walpuski [60] to show the condition that this one-dimensional obstruction vanishes defines τ(ε)\tau(\varepsilon) implicitly as a function of ε\varepsilon, completing the proof of Theorem 1.6.

Acknowledgements

The author gratefully acknowledges the support and guidance of his Ph.D. advisor, Clifford Taubes throughout this project. This work also benefitted from the interest and expertise of many other people, including Aleksander Doan, Siqi He, Rafe Mazzeo, Tom Mrowka, and Thomas Walpuski. The author is supported by an NSF Mathematical Sciences Postdoctoral Research Fellowship (Award No. 2303102).

2. Gluing by the Alternating Method

This section reviews Donaldson’s Alternating method for gluing [9, Sec. 4], and introduces the semi-Fredholm generalization that is used to prove Theorem 1.6. The reader is referred to [10] for additional exposition of general approaches to gluing, and to [34, Ch. 19], and [7, 38, 60] for expositions of other particular instances of gluing results.

2.1. The Structure of Gluing Problems

Let YY be a compact manifold, and let 𝔉:𝒳𝒴\mathfrak{F}:\mathcal{X}\to\mathcal{Y} be a non-linear elliptic PDE viewed as a continuous map between Banach spaces 𝒳,𝒴\mathcal{X},\mathcal{Y} of sections of vector bundles. Suppose that YY is the union of two overlapping open regions

Y=Y+Y,Y=Y^{+}_{\circ}\cup Y^{-}_{\circ},

each of which hosts a solution Φ±\Phi^{\pm} (called model solutions) of 𝔉(Φ±)=0\mathfrak{F}(\Phi^{\pm})=0 on Y±Y^{\pm}_{\circ} (or more generally, a near solution, 𝔉(Φ±)0\mathfrak{F}(\Phi^{\pm})\approx 0).

The associated gluing problem is to produce a global solution Φ\Phi of 𝔉(Φ)=0\mathfrak{F}(\Phi)=0 on YY beginning with the model solutions. This is done by splicing the model solutions together to form an approximate solution

Φ1:=Φ+#εΦ\Phi_{1}:=\Phi^{+}\ \#_{\varepsilon}\ \Phi^{-}

on YY, which is then corrected to a true solution. Here, #ε\#_{\varepsilon} denotes a splicing procedure, usually performed using a cut-off function, which is allowed to depend on a parameter ε\varepsilon called the gluing parameter(s). The true solution is obtained by a particular gluing procedure, which is method for correcting approximate solutions,

ΦN\Phi_{N}\ \ ΦN+1:=ΦN+δΦN\ \ \Phi_{N+1}:=\Phi_{N}+\delta\Phi_{N}correct

that is applied iteratively starting from the initial approximation above to construct a sequence Φ1,Φ2,,\Phi_{1},\Phi_{2},\ldots, ΦN\Phi_{N} of successively improving approximations. Usually, the gluing procedure is a variation of Newton’s method which repeatedly solves some version of the linearized equations to produce corrections δΦN\delta\Phi_{N}.

If the resulting sequence ΦNΦε\Phi_{N}\to\Phi_{\varepsilon} converges in a sufficient function space for appropriate choices of the gluing parameter ε\varepsilon, the limit is a global solution of 𝔉(Φε)=0\mathfrak{F}(\Phi_{\varepsilon})=0 on YY (or more generally, a family of global solutions parameterized by appropriate choices of ε\varepsilon). Oftentimes, the gluing procedure and convergence are packaged into a suitable version of the Implicit Function Theorem or related contraction mapping.

The following example explains how the above framework applies in the context of gluing ASD instantons. This framework has been applied in dozens of other well-known gluing problems in geometric analysis, and many of these could be substituted as equally instructive examples.

Example 2.1.

(Gluing ASD Instantons, [47]) In the context of Uhlenbeck compactness on a compact 44-manifold X4X^{4}, one seeks to construct an ASD instanton of charge k+1k+1 by gluing two connections A±A^{\pm} with A+A^{+} being a “bubble”. In this case, A+A^{+} is the standard instanton of charge 11 and dilation parameter λ\lambda on a ball X+=Bλ(x0)X^{+}_{\circ}=B_{\sqrt{\lambda}}(x_{0}) of radius λ\sqrt{\lambda} around the bubbling point, and AA^{-} is an instanton of charge kk on the complement X=X4Bλ(x0)X^{-}_{\circ}=X^{4}-B_{\lambda}(x_{0}) of a smaller ball. Here, the dilation λ\lambda of the standard instanton is the gluing parameter. See [7, 17, 40] for details.

Notice the following two features of Example 2.1 that will be pertinent for the upcoming case of 2\mathbb{Z}_{2}-harmonic spinors. (1) the decomposition X4=X+λXX^{4}=X^{+}_{\circ}\cup_{\lambda}X^{-}_{\circ} depends on the gluing parameter λ\lambda, with X+X^{+}_{\circ} shrinking as λ0\lambda\to 0. (2) While the problem has a natural “invariant scale” of radius λ\lambda, the two regions X±X^{\pm}_{\circ} overlap in a “neck region” which extends from radius λ\lambda to λ\sqrt{\lambda}, and the much of the gluing analysis occurs there.

Returning now to the case of 2\mathbb{Z}_{2}-harmonic spinors, let (Y,g0)(Y,g_{0}) be a compact 3-manifold and (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) a 2\mathbb{Z}_{2}-harmonic spinor satisfying the hypotheses of Theorem 1.6. In this case the regions Y±Y^{\pm}_{\circ} are a tubular neighborhood of the singular set 𝒵0\mathcal{Z}_{0} whose radius depends on the parameter ε\varepsilon, and its complement. More specifically, let λ=λ(ε)\lambda=\lambda(\varepsilon) be a radius to be specified later, and set

Y+=Nλ(𝒵0)Y=YNλ4/3(𝒵0).Y^{+}_{\circ}=N_{\lambda}(\mathcal{Z}_{0})\hskip 56.9055ptY^{-}_{\circ}=Y-N_{\lambda^{4/3}}(\mathcal{Z}_{0}). (2.1)

The model solutions (Φε+,Aε+)(\Phi^{+}_{\varepsilon},A^{+}_{\varepsilon}) on Y+Y^{+}_{\circ} are the concentrating local family of model solutions constructed in [41] and reviewed in Section 7; the model solutions on YY^{-}_{\circ} are simply the limiting 2\mathbb{Z}_{2}-harmonic spinor (Φ0,A0)(\Phi_{0},A_{0}). In this case, we tacitly refer to the region Y+Y^{+} including the singular set as the “inside” region, and YY^{-} as the “outside” region; the overlap Y+YY^{+}_{\circ}\cap Y^{-}_{\circ} is referred to as the “neck” region as in Example 2.1. The gluing parameters in our situation are triples (ε,τ,𝒵)(\varepsilon,\tau,\mathcal{Z}) of the L2L^{2}-norm, the metric parameter, and a deformation of the singular set, and there is, more precisely, a decomposition (2.1) for each such triple.

2.2. The Alternating Method

The gluing procedure employed to prove Theorem 1.6 is a generalization of the alternating method.

The alternating method iteratively corrects approximate solutions by alternating making corrections localized to the two regions Y±Y^{\pm}_{\circ}. This method was first used by S. Donaldson [9] to give a different perspective on C. Taubes’s gluing theorem for ASD instantons (Example 2.1). More specifically, the successively approximations at the NthN^{\text{th}} stage have the form

ΦN=Φ1+χ+φN+χψN\Phi_{N}=\Phi_{1}+\chi^{+}\varphi_{N}+\chi^{-}\psi_{N} (2.2)

where φN,ψN\varphi_{N},\psi_{N} are corrections supported in the two regions respectively, and χ±\chi^{\pm} are cut-off function restricting them to their respective regions. The iteration may be summarized by the following schematic:

{𝔉(ΦN)=𝔢Nsupp(𝔢N)Y}\left\{\begin{matrix}\mathfrak{F}(\Phi_{N})=\mathfrak{e}_{N}\\ \text{supp}(\mathfrak{e}_{N})\subseteq Y^{-}\end{matrix}\right\}{𝔉(ΨN)=𝔢Nsupp(𝔢N)Y+}\left\{\begin{matrix}\mathfrak{F}(\Psi_{N})=\mathfrak{e}^{\prime}_{N}\\ \text{supp}(\mathfrak{e}^{\prime}_{N})\subseteq Y^{+}\end{matrix}\right\}solve on Y+NN+1,\begin{matrix}\text{solve on $Y^{+}$}\\ N\mapsto N+1,\end{matrix}solve on YY^{-}

where ΨN=Φ1+χ+φN+χψN+1\Psi_{N}=\Phi_{1}+\chi^{+}\varphi_{N}+\chi^{-}\psi_{N+1}. Thus a single stage of the overall iteration constitutes one alternation back and forth. In the notation of Section 2.1, one has δΦN=χ+δφN+χδψN\delta\Phi_{N}=\chi^{+}\delta\varphi_{N}+\chi^{-}\delta\psi_{N} where δφN=φNφN1\delta\varphi_{N}=\varphi_{N}-\varphi_{N-1} and likewise for ψ\psi. The iteration converges to a solution if the errors 𝔢N0\mathfrak{e}_{N}\to 0.

To explain further, let us give a more precise meaning to the steps of “solve on Y±Y^{\pm}”. The equations at a small perturbation Φ+φ\Phi+\varphi of an approximate solution Φ\Phi may be written

𝔉(Φ+φ)=𝔉(Φ)+Φ(φ)+Q(φ)\mathfrak{F}(\Phi+\varphi)=\mathfrak{F}(\Phi)+\mathcal{L}_{\Phi}(\varphi)+Q(\varphi) (2.3)

where Φ=dΦ𝔉\mathcal{L}_{\Phi}=\text{d}_{\Phi}\mathfrak{F} is the linearization of 𝔉\mathfrak{F} at Φ\Phi, and QQ the higher-order terms. The solving steps on Y±Y^{\pm}_{\circ} involve solving the linearized equation. Since Y±Y^{\pm}_{\circ} are open regions, it is often necessary to choose extensions Y±Y±Y^{\pm}_{\circ}\subseteq Y^{\pm} on which Φ\mathcal{L}_{\Phi} has sufficient invertibility properties. This is typically done by attaching a tubular ends to Y±\partial Y^{\pm}_{\circ} or imposing boundary conditions (see also Section 2.3). Write ±\mathcal{L}^{\pm} for the linearization acting on sections over sufficiently chosen extensions Y±Y^{\pm}. We assume that ±\mathcal{L}^{\pm} naturally extends the restriction of the linearization on Y±Y^{\pm}_{\circ}, and that the vector bundles and model solutions likewise admit natural extensions to Y±Y^{\pm}.

The alternating method, in its standard guise, requires the following two assumptions.

  • (I)

    The linearizations ±\mathcal{L}^{\pm} at the sequence of approximate solutions are uniformly invertible in the gluing parameter ε\varepsilon.

  • (II)

    The solution φ\varphi of ±φ=g\mathcal{L}^{\pm}\varphi=g decays away from the support of gg in both regions in the following sense. If supp(g)supp(dχ±)\text{supp}(g)\subseteq\text{supp}(d\chi^{\pm}), then

    σ(dχ)gCδg\|\sigma_{\mathcal{L}}(d\chi^{\mp})g\|\leq C\delta\|g\| (2.4)

    for some δ<1\delta<1, where σ\sigma_{\mathcal{L}} is the principal symbol.

Given that (I)–(II) hold, the starting point of the alternating method is an initial approximation Φ1\Phi_{1} so that the error 𝔉(Φ1)=𝔢1\mathfrak{F}(\Phi_{1})=\mathfrak{e}_{1} of this initial approximation is sufficiently small and supported where dχ+0d\chi^{\text{+}}\neq 0. The iteration then proceeds inductively, with a single step given as follows.

  1. (1)

    Let ψ\psi be the unique solution of

    (+Q)ψ=𝔢N(\mathcal{L}^{-}+Q)\psi=-\mathfrak{e}_{N} (2.5)

    on YY^{-} given by the Inverse Function Theorem.

  2. (2)

    Set ψN+1=ψN+ψ\psi_{N+1}=\psi_{N}+\psi, and ΨN=Φ1+χ+φN+χψN+1\Psi_{N}=\Phi_{1}+\chi^{+}\varphi_{N}+\chi^{-}\psi_{N+1}. This intermediate approximation satisfies

    𝔉(ΨN)=𝔢N𝔢N+dχψ+gN:=fN\mathfrak{F}(\Psi_{N})=\cancel{\mathfrak{e}_{N}-\mathfrak{e}_{N}}+d\chi^{-}\psi+g_{N}:=f_{N}

    where gN=χQ(ψ)Q(χψ)g_{N}=\chi^{-}Q(\psi)-Q(\chi^{-}\psi).

  3. (3)

    Since fNf_{N} is supported where dχ0d\chi^{-}\neq 0, the solution ψ\psi decays across the neck region by condition (II) so that fNCδ𝔢N\|f_{N}\|\leq C\delta\|\mathfrak{e}_{N}\| for some fixed factor δ<1\delta<1.

  4. (4)

    Repeat steps (1)–(3) on Y+Y^{+} to obtain a correction φ\varphi so φN+1=φN+φ\varphi_{N+1}=\varphi_{N}+\varphi, then set

    ΦN+1=Φ1+χ+φN+1+χψN+1.\Phi_{N+1}=\Phi_{1}+\chi^{+}\varphi_{N+1}+\chi^{-}\psi_{N+1}.

    The resulting error, 𝔢N+1\mathfrak{e}_{N+1}, then satisfies 𝔢N+1(Cδ)2𝔢N\|\mathfrak{e}_{N+1}\|\leq(C\delta)^{2}\|\mathfrak{e}_{N}\|.

If δ\delta is sufficiently small that (Cδ)<1(C\delta)<1, the sequence ΦN\Phi_{N} constructed by iterating Steps (1)–(4) converges to a limit Φ0\Phi_{0} which solves 𝔉(Φ0)=0\mathfrak{F}(\Phi_{0})=0.

Refer to captionχ+\chi^{+}χ\chi^{-}Y+Y^{+}YY^{-}𝔢N\mathfrak{e}_{N}𝔢N\mathfrak{e}^{\prime}_{N}𝔢N+1\mathfrak{e}_{N+1}
Figure 1. An illustration of the alternating iteration procedure in Steps (1)–(4) above. (Top) The cut-off functions χ±\chi^{\pm}, (red) the error terms with alternating support and decreasing norm, (blue/green) the decay of solutions across the neck region.

The content of proving a gluing result using the alternating method is showing a sufficient version of the hypotheses (I) and (II) are met. Verifying the statement (I) requires identifying suitable extensions Y±Y±Y^{\pm}_{\circ}\subseteq Y^{\pm} and proving that the restricted operator can be sufficiently inverted. The decay requirements in condition (II) can be concluded from knowledge of the linearization’s Green’s function, or from using weighted function spaces.

The main advantage of the alternating method over other gluing procedures, and indeed the reason it is suitable in our setting, is that it can effectively treat asymmetry between the two regions Y±Y^{\pm}. In particular, the method only requires analysis of ±\mathcal{L}^{\pm} in the two distinct regions separately, and never the analysis of a global linearization whose properties are an (in our case quite opaque) combination of those inherited from the two regions. This allows the asymmetric character of the equation in the two regions to be isolated and analyzed separately; indeed, [41] and [43] are most effectively viewed for the present purposes as manuals for the Seiberg–Witten theory on Y+Y^{+} and YY^{-} respectively.

Remark 2.2.

A slight variation on Steps (1)–(4) above is to solve the strictly linear equation at each step. Thus Step (1) is replaced by solving

ψ=eN+Q(ψN+φN)\mathcal{L}^{-}\psi=-e_{N}+Q(\psi_{N}+\varphi_{N})

where QQ denotes the higher-order error from the correction at the previous stage, and likewise for the inside. This formulation is equivalent, though it comes at the cost of disrupting the fact that the error terms are entirely supported where dχ±0d\chi^{\pm}\neq 0. In our case, the higher-order terms are sufficiently mild that the linearized version suffices.

2.3. Gluing as Non-linear Excision

The alternating method can be rephrased using the language of parametrix patching and contraction mappings. This rephrasing is helpful for addressing questions the regarding uniqueness of gluing and smooth dependence on parameters. It also clarifies the relationship of the alternating method to other gluing methods that use parametrices.

At first glance, the alternating method does not result in a contraction mapping or a Fredholm problem in an obvious way. Indeed, there is an infinite-dimensional ambiguity in the construction, coming from the freedom to alter (2.2) by φN,ψNφN+ξ,ψNξ\varphi_{N},\psi_{N}\mapsto\varphi_{N}+\xi,\psi_{N}-\xi on the overlap region. This ambiguity can be resolved by framing the alternating method as a non-linear analogue of the excision problem for the index of elliptic operators. With this perspective, it becomes clear that the redundancy in the description can be eliminated by considering the “virtual” gluing problem solved by φNψN\varphi_{N}-\psi_{N} on the neck region. The author learned this perspective on gluing from [40].

To explain further, let us briefly review the excision principle for elliptic operators. For convenience, assume here that 𝔉\mathfrak{F} is a first-order elliptic operator. Let S1,S2YS_{1},S_{2}\to Y be vector bundles whose sections form the domain and codomain of the linearization, so that Φ1:H1(S1)L2(S2)\mathcal{L}_{\Phi_{1}}:H^{1}(S_{1})\to L^{2}(S_{2}) is Fredholm. The closed manifold YY can be reconstructed from Y±Y^{\pm} by the following cut-and-paste procedure. Let U=Y+YY±U=Y^{+}_{\circ}\cap Y^{-}_{\circ}\subseteq Y^{\pm}_{\circ} be the overlap. Assume that the inclusion UY±U\hookrightarrow Y^{\pm} separates Y±UY^{\pm}-U into two connected components, so that we may write the extended domain Y±=Y±UN±Y^{\pm}=Y^{\pm}_{\circ}\cup_{U}N^{\pm}_{\circ} as the union of the original domain and the extension (this being a tubular end or collar neighborhood of Y±\partial Y^{\pm}_{\circ}). Cutting the domains Y±Y^{\pm} along UU and pasting them together again exchanging the second component,

Y+=Y+UN+Y=NUYY=Y+UYN=NUN+\displaystyle\begin{matrix}Y^{+}&=&Y^{+}_{\circ}\cup_{U}N^{+}_{\circ}\\ Y^{-}&=&N^{-}_{\circ}\cup_{U}Y^{-}_{\circ}\end{matrix}\hskip 42.67912pt\Leftrightarrow\hskip 42.67912pt\begin{matrix}Y=Y^{+}_{\circ}\cup_{U}Y^{-}_{\circ}\\ N=N^{-}_{\circ}\cup_{U}N^{+}_{\circ}\end{matrix} (2.6)

gives rise to the closed manifold YY and an extra copy of the neck region NN. As in Section 3.2, assume that the restrictions of S1,S2S_{1},S_{2} extend so that these bundles are well-defined on any of the four manifolds above; let +,,Φ1,N\mathcal{L}^{+},\mathcal{L}^{-},\mathcal{L}_{\Phi_{1}},\mathcal{L}_{N} denote the four manifestations of the linearized equations.

The cut-off functions χ±\chi^{\pm} extend by 0 or 11 to each piece. Using these, define maps

α:H1(Y+)H1(Y)H1(Y)H1(N)β:L2(Y)L2(N)L2(Y+)L2(Y)\alpha:H^{1}(Y^{+})\oplus H^{1}(Y^{-})\to H^{1}(Y)\oplus H^{1}(N)\hskip 42.67912pt\beta:L^{2}(Y)\oplus L^{2}(N)\to L^{2}(Y^{+})\oplus L^{2}(Y^{-})

by

α=(χ+χχχ+)β=(χ+χχχ+)\alpha=\begin{pmatrix}\chi^{+}&\chi^{-}\\ -\chi^{-}&\chi^{+}\end{pmatrix}\hskip 85.35826pt\beta=\begin{pmatrix}\chi^{+}&-\chi^{-}\\ \chi^{-}&\chi^{+}\end{pmatrix} (2.7)

which cut-and paste sections of S1,S2S_{1},S_{2} along with (2.6). These satisfy w:=αβ=(χ+)2+(χ)21w:=\alpha\circ\beta=(\chi^{+})^{2}+(\chi^{-})^{2}\geq 1.

Proposition 2.3 (Excision).

Assume that +,,Φ1\mathcal{L}^{+},\mathcal{L}^{-},\mathcal{L}_{\Phi_{1}} are Fredholm. Then N\mathcal{L}_{N} is also Fredholm, and the indices satisfy

ind(+)+ind()=ind(Φ1)+ind(N).\text{ind}(\mathcal{L}^{+})+\text{ind}(\mathcal{L}^{-})=\text{ind}(\mathcal{L}_{\Phi_{1}})+\text{ind}(\mathcal{L}_{N}).
Proof.

Let P±:L2(Y±)H1(Y±)P^{\pm}:L^{2}(Y^{\pm})\to H^{1}(Y^{\pm}) denote parametrices for ±\mathcal{L}^{\pm}, and define

P=α(P+,P)βw:L2(Y)L2(N)H1(Y)H1(N).P=\alpha\circ(P^{+},P^{-})\circ\frac{\beta}{w}:L^{2}(Y)\oplus L^{2}(N)\to H^{1}(Y)\oplus H^{1}(N). (2.8)

PP is Fredholm because P±P^{\pm} are and α,β\alpha,\beta are invertible. Setting L=(Φ1,N)L=(\mathcal{L}_{\Phi_{1}},\mathcal{L}_{N}), a quick calculation (see [34, Page 245]) shows that LPIdLP-\text{Id} and PLIdPL-\text{Id} are compact operators consisting of terms involving dχ±d\chi^{\pm}, hence LL is Fredholm. It follows that N\mathcal{L}_{N} is also Fredholm, and the indices satisfy ind(L)=ind(P)=ind(P+)ind(P)=ind(+)+ind()\text{ind}(L)=-\text{ind}(P)=-\text{ind}(P^{+})-\text{ind}(P^{-})=\text{ind}(\mathcal{L}^{+})+\text{ind}(\mathcal{L}^{-}). ∎

The correction to the approximate solution (2.2) constructed by the alternating method, is the H1(Y)H^{1}(Y)-component of α(φN,ψN)\alpha(\varphi_{N},\psi_{N}) with α\alpha as in (2.7). The relevance of excision to the alternating method is that it shows resolving the ambiguity on Y+YY^{+}_{\circ}\cap Y^{-}_{\circ} requires dictating the fate of the H1(N)H^{1}(N)-component of α(φN,ψN)\alpha(\varphi_{N},\psi_{N}). The alternation described in Section 2.2 implicitly requires that this component is zero. To see this, let ζ±:Y\zeta^{\pm}:Y\to\mathbb{R} be a partition of unity so that ζ=1\zeta^{\mp}=1 on supp(dχ±)\text{supp}(d\chi^{\pm}), and set

={(P+ζ+gPζg)|(g,0)L2(Y)L2(N)}H1(Y+)H1(Y)\mathcal{H}=\left\{\begin{pmatrix}P^{+}\zeta^{+}g\\ P^{-}\zeta^{-}g\end{pmatrix}\ \Big{|}\ (g,0)\in L^{2}(Y)\oplus L^{2}(N)\right\}\subseteq H^{1}(Y^{+})\oplus H^{1}(Y^{-}) (2.9)

i.e. \mathcal{H} is the image of (P+,P)β(P^{+},P^{-})\circ\beta where β\beta is as in (2.7) but now formed using ζ±\zeta^{\pm} in place of χ±\chi^{\pm}. A quick inspection of Steps (1)–(4) in Section 2.2 shows that the correction (φN,ψN)(\varphi_{N},\psi_{N}) lands in \mathcal{H} at each stage of the iteration.

At a functional analytic level, the alternating method constructs a sequence of approximate inverses to 𝔉\mathfrak{F} on the space \mathcal{H}. This is most easily understood as a slight extension of the linear setting (i.e. Q=0Q=0 in 2.3), which we now explain in detail. In the linear setting, Steps (1)–(4) produce a sequence of approximate inverses to Φ1α\mathcal{L}_{\Phi_{1}}\circ\alpha by a nested parametrix patching. (2.8) is an unsophisticated version of parametrix patching; indeed, it patches together the two parametrices P±P^{\pm} on their respective regions using the cut-off functions χ±\chi^{\pm}. When the L2(N)L^{2}(N)-component in the codomain vanishes, one may replace the L2(Y)L^{2}(Y)\to\mathcal{H} component of (2.8) by

P1=χ+P+ζ++χPζ.P_{1}=\chi^{+}P^{+}\zeta^{+}+\chi^{-}P^{-}\zeta^{-}. (2.10)

In this instances Φ1P1\mathcal{L}_{\Phi_{1}}P_{1} and P1Φ1P_{1}\mathcal{L}_{\Phi_{1}} fail to be the identity by compact terms involving dχ±,dζ±d\chi^{\pm},d\zeta^{\pm}. The alternating method corrects P1P_{1} to recursively cancel these errors, as the next proposition demonstrates.

Proposition 2.4.

Suppose that assumptions (I)–(II) from Section 2.2 hold. Then

Φ1α:L2(Y)\mathcal{L}_{\Phi_{1}}\circ\alpha:\mathcal{H}\to L^{2}(Y)

is invertible.

Proof.

Since ±\mathcal{L}^{\pm} are surjective, it may be assumed that P±P^{\pm} are injective; thus P1:L2(Y)P_{1}:L^{2}(Y)\to\mathcal{H} is an isomorphism, and in particular has index 0. Set

e1±:=dχ±.P±ζ±e^{\pm}_{1}:=d\chi^{\pm}.P^{\pm}\zeta^{\pm} (2.11)

so that Φ1P1=Id+e1++e1\mathcal{L}_{\Phi_{1}}P_{1}=\text{Id}+e^{+}_{1}+e^{-}_{1}. Then, with P1P_{1} as in (2.10) with P1±P_{1}^{\pm} being the two terms, recursively define

PN+1±:=PN±χP(dχ±.P±eN±fN)+χ±P±(dχ.PfNeN+1±)P_{N+1}^{\pm}:=P_{N}^{\pm}\ -\ \chi^{\mp}P^{\mp}(\underbrace{d\chi^{\pm}.P^{\pm}e_{N}^{\pm}}_{f^{\mp}_{N}})\ +\ \chi^{\pm}P^{\pm}(\underbrace{d\chi^{\mp}.P^{\mp}f_{N}^{\mp}}_{e^{\pm}_{N+1}}) (2.12)

and take PN=PN++PNP_{N}=P_{N}^{+}+P_{N}^{-}. Expanding shows

Φ1PN=Id+eN++eN\mathcal{L}_{\Phi_{1}}P_{N}=\text{Id}+e_{N}^{+}+e_{N}^{-}

is a telescoping series. Moreover, using (II) repeatedly shows that eN++eN=O(δN)e_{N}^{+}+e_{N}^{-}=O(\delta^{N}), where δ\delta is as in (2.4) because both eN±e_{N}^{\pm} are a nested composition of NN terms each of which gains a decay factor of δ\delta from condition (II). Provided Cδ<1C\delta<1 for CC depending on P±P^{\pm}, it follows that Φ1\mathcal{L}_{\Phi_{1}} is surjective and PNP_{N} injective. Since PNP_{N} differs from P1P_{1} by compact terms involving dχ±d\chi^{\pm}, PNP_{N} is index 0, hence it is an isomorphism thus so is Φ1\mathcal{L}_{\Phi_{1}}. ∎

The nested parametrix patching in the proof of Proposition 2.4 may be also be phrased as a contraction finding the true inverse of Φ1\mathcal{L}_{\Phi_{1}}. Indeed, for gL2(Y)g\in L^{2}(Y), the map dTg:\text{d}T_{g}:\mathcal{H}\to\mathcal{H} given by

dTg:=IdP1(Φ1(_)g)\text{d}T_{g}:=\text{Id}-P_{1}(\mathcal{L}_{\Phi_{1}}(\_)-g) (2.13)

applied iteratively starting from 00\in\mathcal{H} reconstructs the sequence of approximations PNgP_{N}g. For δ\delta sufficiently small, dTg2\text{d}T^{2}_{g} is a contraction. (In fact, on the subset of \mathcal{H} for which the error is supported where dχ±0d\chi\pm\neq 0, dTg\text{d}T_{g} itself behaves as a contraction, and the first application imposes this condition).

The non-linear version of the alternating method is essentially dentical, where T:T:\mathcal{H}\to\mathcal{H} is now given by

T:=IdP1𝔉(Φ1+_)T:=\text{Id}-P_{1}\mathfrak{F}(\Phi_{1}+\_) (2.14)

and where P1P_{1} is as in (2.10), with P±P^{\pm} now being inverses of the non-linear equations (2.5). In this case, 𝔉(ΦN)\mathfrak{F}(\Phi_{N}) on the sequence of approximate solutions constructed by Steps (1)–(4) in Section 2.2 expands to mirror the telescoping cancellation of Proposition 2.4, but (2.11) now also includes non-linear error terms. Again, T2T^{2} is a contraction, provided sufficient control on the non-linearity.

Remark 2.5.

In many gluing problems (including all previous gluing problems in gauge theory known to the author), a single iteration of the nested parametrix construction (2.10) resulting in P1P_{1} is sufficient. In these situations, the geometry (e.g. stretching a neck so that |dχ|0|d\chi|\to 0) or function spaces (appropriate weights) is such that in (2.11), e1±0e_{1}^{\pm}\to 0 in the appropriate norm as the gluing parameter ε0\varepsilon\to 0. When this is not the case (and indeed it is not for gluing 2\mathbb{Z}_{2}-harmonic spinors), the nested construction of Proposition 2.4 provides a more refined approximate inverse. The alternating method described in Sections 2.2 and 2.3 therefore subsumes the methods used in many standard gluing constructions with appropriate reorganization of the proof. More relevant here, it allows the upcoming generalization.

2.4. The Semi-Fredholm Alternating Method

As explained in the introduction, condition (I) in Section 2.2 fails for \mathcal{L}^{-} in the context of gluing 2\mathbb{Z}_{2}-harmonic spinors because the linearized Seiberg–Witten equation at a 2\mathbb{Z}_{2}-harmonic spinor are only semi-Fredholm and therefore not invertible. In fact, for effectively the same reason, condition (I) fails for +\mathcal{L}^{+} as well, but in this case what fails is the uniformity of invertibility, rather than invertibility itself. The gluing problem for 2\mathbb{Z}_{2}-harmonic spinors therefore departs the setting of standard gluing problems in several key ways, and requires the following more sophisticated generalization of the alternating method.

Replace assumptions (I)–(II) from Section 2.2 by the following weaker hypotheses:

  1. (I’)

    The linearizations +\mathcal{L}^{+} at the sequence of approximate solutions are invertible.

  2. (II’)

    Condition (II) from Section 2.2 continues to hold for constants δ±\delta^{\pm}.

  3. (III’)

    The linearizations \mathcal{L}^{-} at the sequence of approximate solutions are left semi-Fredholm with uniformly bounded left-inverses.

  4. (IV’)

    There is a family of operators FξF_{\xi} on configurations, parameterized by an infinite-dimensional parameter ξ𝔚\xi\in\mathfrak{W} which generate a corresponding family of approximate solutions Fξ(ΦN)F_{\xi}(\Phi_{N}). Moreover, the projection TΦN:=Π0dξ𝔉ΦNT_{\Phi_{N}}:=\Pi_{0}\circ\text{d}_{\xi}\mathfrak{F}_{\Phi_{N}} of the derivative

    TΦN:𝔚Coker()T_{\Phi_{N}}:\mathfrak{W}\longrightarrow\text{Coker}(\mathcal{L}^{-}) (2.15)

    is uniformly invertible at the family of approximate solutions, where 𝔉(ξ,Φ):=𝔉(Fξ(Φ))\mathfrak{F}(\xi,\Phi):=\mathfrak{F}(F_{\xi}(\Phi)), and Π0\Pi_{0} denotes the projection to the cokernel.

Notice that Item (I’) is identical to Item (I) from Section 2.2 for the inside except that it omits the requirement of uniformity. Note also the similarity of the map in Item (IV’) to that of (1.11) in Theorem 1.2 in the introduction.

The approximate solutions in the sequence (2.2) now also depends on the parameter ξ\xi that is adjusted in each step of the iteration, and have the form

ΦN=FξN(Φ1)+χ+φN+χψN.\Phi_{N}=F_{\xi_{N}}(\Phi_{1})+\chi^{+}\varphi_{N}+\chi^{-}\psi_{N}. (2.16)

The alternating iteration procedure is similarly adjusted to become the three-stage cycle (1.14), which may now be written as follows:

{𝔉(ξN,ΦN)=𝔢Nsupp(𝔢N)Y}\left\{\begin{matrix}\mathfrak{F}(\xi_{N},\Phi_{N})=\mathfrak{e}_{N}\\ \text{supp}(\mathfrak{e}_{N})\subseteq Y^{-}\end{matrix}\right\}{𝔉(ξN+1,ΦN)=𝔢NΠ0(𝔢N)=0}\left\{\begin{matrix}\mathfrak{F}(\xi_{N+1},\Phi_{N})=\mathfrak{e}^{\prime}_{N}\\ \Pi_{0}(\mathfrak{e}_{N}^{\prime})=0\end{matrix}\right\}{𝔉(ξN+1,ΨN)=𝔢N′′supp(𝔢N′′)Y+}.\left\{\begin{matrix}\mathfrak{F}(\xi_{N+1},\Psi_{N})=\mathfrak{e}^{\prime\prime}_{N}\\ \text{supp}(\mathfrak{e}_{N}^{\prime\prime})\subseteq Y^{+}\end{matrix}\right\}.solve on Y+\text{solve on }Y^{+}adjust ξ\xiNN+1\begin{matrix}N\mapsto N+1\end{matrix}\ \ \ solve on YY^{-} (2.17)

Commensurately, Steps (1)–(4) in the previous subsection are augmented by a Step (0) corresponding to the horizontal arrow in (2.17), and Step (1) is altered accordingly. Here, we retain the notation of Steps (1)–(4) from the previous subsection. Assuming that the codomain of \mathcal{L}^{-} is a Hilbert space, Π0\Pi_{0} may be viewed as the orthogonal projection to cokernel viewed as the orthogonal complement of the range, thus (1Π0)(1-\Pi_{0}) denotes the projection to the range of \mathcal{L}^{-}.

  1. (0)

    Let ξ\xi denote the unique solution of

    TΦN(ξ)=Π0(𝔢N),T_{\Phi_{N}}(\xi)=-\Pi_{0}(\mathfrak{e}_{N}), (2.18)

    and set ξN+1=ξN+ξ\xi_{N+1}=\xi_{N}+\xi.

  2. (1)

    Let ψ\psi be the unique solution of

    (ψ)=(1Π0)(𝔢N+d𝔉ΦN(ξ,0)).\mathcal{L}^{-}(\psi)=-(1-\Pi_{0})\left(\mathfrak{e}_{N}+\text{d}\mathfrak{F}_{\Phi_{N}}(\xi,0)\right). (2.19)

Steps (2)–(4) then proceed as before, with the following minor adjustment. Due to the lack of uniform invertibility of +\mathcal{L}^{+}, the constant CC in the decay factor (written CδC\delta in Step (3) in Section 2.2) may now depend on the gluing parameter ε\varepsilon. Thus the condition for the sequence {ΦN}\{\Phi_{N}\} for converge becomes that

|C(ε)δ+δ|<1.|C(\varepsilon)\delta^{+}\delta^{-}|<1. (2.20)

The decay constants δ±=δ±(ε)\delta^{\pm}=\delta^{\pm}(\varepsilon) may themselves depend on the gluing parameter.

The discussion of Section 2.3 carries over to the semi-Fredholm setting with analogous modifications. There are now three parametrices Pξ,P+,PP_{\xi},P^{+},P^{-} , with PξP_{\xi} being a parametrix for TΦNT_{\Phi_{N}}, and the cyclic iteration (2.17) may again be phrased as a non-linear analogue of the nested parametrix patching akin to how (2.14) is a non-linear version of (2.13). The linear version extending (2.13) in this case is constructed analogously to Proposition 2.4, replacing (2.9) by 𝔚\mathcal{H}\oplus\mathfrak{W} and (2.10) and (2.12) by a sequence

P1\displaystyle P_{1} :=\displaystyle:= Pξ\displaystyle P_{\xi} (2.21)
P2\displaystyle P_{2} :=\displaystyle:= Pξ+P(Idd𝔉Φ1P1)\displaystyle P_{\xi}+P^{-}(\text{Id}-\text{d}\mathfrak{F}_{\Phi_{1}}P_{1}) (2.22)
P3\displaystyle P_{3} :=\displaystyle:= Pξ+P(Idd𝔉Φ1P1)+P+(Idd𝔉Φ1P2)\displaystyle P_{\xi}+P^{-}(\text{Id}-\text{d}\mathfrak{F}_{\Phi_{1}}P_{1})+P^{+}(\text{Id}-\text{d}\mathfrak{F}_{\Phi_{1}}P_{2}) (2.23)
P4\displaystyle P_{4} :=\displaystyle:= Pξ+P(Idd𝔉Φ1P1)+P+(Idd𝔉Φ1P2)Pξ(Idd𝔉Φ1P3)\displaystyle P_{\xi}+P^{-}(\text{Id}-\text{d}\mathfrak{F}_{\Phi_{1}}P_{1})+P^{+}(\text{Id}-\text{d}\mathfrak{F}_{\Phi_{1}}P_{2})-P_{\xi}(\text{Id}-\text{d}\mathfrak{F}_{\Phi_{1}}P_{3})

where Pξ=TΦ11Π0ζP_{\xi}=T_{\Phi_{1}}^{-1}\Pi_{0}\zeta^{-}, P=χΦ11(1Π0)ζP^{-}=\chi^{-}\mathcal{L}_{\Phi_{1}}^{-1}(1-\Pi_{0})\zeta^{-} and P+=χ+Φ11ζ+P^{+}=\chi^{+}\mathcal{L}_{\Phi_{1}}^{-1}\zeta^{+}. Each successive approximation is defined by PN+1=PNP(Idd𝔉Φ1PN)P_{N+1}=P_{N}-P^{\circ}(\text{Id}-\text{d}\mathfrak{F}_{\Phi_{1}}P_{N}) with PP^{\circ} cycling between Pξ,PP_{\xi},P^{-}, and P+P^{+} mod 3. Equivalently, P3NP_{3N} is characterized by

(IdP3Nd𝔉Φ1)\displaystyle(\text{Id}-P_{3N}\text{d}\mathfrak{F}_{\Phi_{1}}) =\displaystyle= ((IdP+d𝔉Φ1)(IdPd𝔉Φ1)(IdPξd𝔉Φ1))N\displaystyle\left((\text{Id}-P^{+}\text{d}\mathfrak{F}_{\Phi_{1}})(\text{Id}-P^{-}\text{d}\mathfrak{F}_{\Phi_{1}})(\text{Id}-P_{\xi}\text{d}\mathfrak{F}_{\Phi_{1}})\right)^{N} (2.24)
=\displaystyle= (IdP3d𝔉Φ1)N,\displaystyle(\text{Id}-P_{3}\text{d}\mathfrak{F}_{\Phi_{1}})^{N}, (2.25)

which is dTN\text{d}T^{N} as in (2.13) for g=0g=0.

The main tasks in the proof of Theorem 1.6 are establishing the analogues of Lemma 2.3 and Proposition 2.4 in this setting. This is done in Sections 1011 in the Seiberg–Witten setting, where ξ\xi in Item (IV’) parameterizes small deformations of the the singular set 𝒵0\mathcal{Z}_{0}, and (2.15) is the operator in Theorem 1.2.

2.5. Key Technical Difficulties

Despite the fact that versions of the statements (I’)–(IV’) are established in [41, 43], there are still several key technical difficulties to overcome to carry out the alternating iteration. Addressing these in a systematic way to turn the scheme outlined in Section 2.4 into a proof of Theorem 1.6 occupies the bulk of Sections 411. In this subsection, we briefly sketch the most important technical problems that arise, and how they are solved.

2.5.1. Loss of Regularity

The main technicality that complicates the application of the cyclic iteration in Section 2.4 to prove Theorem 1.6 is the following:

(Difficulty I): The deformation operator (1.2) displays a loss of regularity of order 32\tfrac{3}{2}.

Loss of regularity is an intriguing (though sometimes technical) pheomenon in certain non-linear PDEs [19, 1, 30]. While the linearized deformation operator (1.11) is a pseudo-elliptic operator of order 1/21/2, the non-linear terms and the Range(A0)\text{Range}(\not{D}_{A_{0}})-components are of order >1/2>1/2. Thus one cannot choose function spaces for which it is simultaneously the case that the linear operator is Fredholm and the non-linear terms are bounded. This loss of regularity necessitates the use of a version of the Nash-Moser Implicit Function Theorem in [43] (see Sections 7 and 8 therein).

Here, the loss of regularity manifests differently, and (perhaps surprisingly) the use of a Nash-Moser type iteration involving smoothing operators is not necessary. In Section 5, it is shown that Coker(𝒵0)Γ(𝒵0;𝒮)\text{Coker}(\not{D}_{\mathcal{Z}_{0}})\simeq\Gamma(\mathcal{Z}_{0};\mathcal{S}) for a bundle 𝒮𝒵0\mathcal{S}\to\mathcal{Z}_{0} and that the cokernel increasingly concentrates around 𝒵0\mathcal{Z}_{0} as the Fourier index increases, so that the th\ell^{th} mode has characteristic length scale 1/||1/|\ell|. Consequently, a restriction on the support of a spinor away from 𝒵0\mathcal{Z}_{0} at distance R0R_{0} leads to unexpectedly rapid decay on the Fourier modes of its cokernel component once the Fourier index exceeds ||O(R01)|\ell|\geq O(R_{0}^{-1}). The error terms in the alternating iteration are supported where the cut-off dχ±0d\chi^{\pm}\neq 0 and therefore obey such a restriction. Consequently, it is possible to cancel the obstruction at each stage using deformations that lie in a large finite-dimensional space with Fourier modes only up to O(ε2/3)O(\varepsilon^{-2/3}). In particular, each deformation is smooth, and the use of smoothing operators is not required.

This does not, however, mean that the loss of regularity is inconsequential. If ξN\xi_{N} is the deformation correction at the NthN^{th} stage, given by (2.18), then the elliptic estimates for TΦ0T_{\Phi_{0}} gives a bound on ξN1/2\|\xi_{N}\|_{1/2} in the Sobolev space of regularity s=1/2s=1/2. Due to the loss of regularity, however, bounding other terms requires control of ξs,2\|\xi\|_{s,2} in the space of regularity s>2s>2. Because of the restriction on Fourier modes of ξN\xi_{N}, this norm is larger by a power of ε1\varepsilon^{-1}. The loss of regularity therefore manifests as adverse powers of ε1\varepsilon^{-1} in certain error terms.

More specifically, the problematic term is (1Π0)d𝔉ΦN(ξ,0)(1-\Pi_{0})\text{d}\mathfrak{F}_{\Phi_{N}}(\xi,0) in (2.19). In the proof of Theorem 1.6, this term takes the following form. Splitting the codomain of the linearization on YY^{-} as L2(S2)=Coker()Range()L^{2}(S_{2})=\text{Coker}(\mathcal{L}^{-})\oplus\text{Range}(\mathcal{L}^{-}) and supplementing the domain to be H1(S1)𝔚H^{1}(S_{1})\oplus\mathfrak{W}, the linearization can be written as the block matrix

d𝔉Φ1(ξ,ψ)=(TΦ10(1Π0)d𝔉Φ1(Φ1,A1))(ξψ).\text{d}\mathfrak{F}_{\Phi_{1}}(\xi,\psi)=\begin{pmatrix}T_{\Phi_{1}}&0\\ \boxed{(1-\Pi_{0})\text{d}\mathfrak{F}_{\Phi_{1}}}&\mathcal{L}^{-}_{(\Phi_{1},A_{1})}\end{pmatrix}\begin{pmatrix}\xi\\ \psi\end{pmatrix}. (2.26)

In this guise, there first two arrows of (2.17) (i.e. Step (0)–(1) in (2.18)) are precisely solving the first and second rows of (2.26) respectively. The problematic term is indicated by a box: in the Seiberg–Witten setting bounding this term requires a bound on ξ5/2,2\|\xi\|_{5/2,2}. Without careful set-up of the entire iteration scheme, the extraneous powers of ε1\varepsilon^{-1} from the loss of regularity make this term too large for the iteration to converge.

As explained briefly in Section 1.6, the key idea that overcomes this difficulty is a more sophisticated method of deforming the singular set:

(Solution I): Use mode-dependent deformations of the singular set.

These are introduced in Section 6.3. Just as the th\ell^{th} Fourier mode of the cokernel has characteristic length scale 1/||1/|\ell| as explained above, the mode-dependent deformations purposely manufacture a similar relationship for the diffeomorphisms used in (1.15). This link between the spatial domain on YY and the Fourier domain of the deformations results in unexpectedly strong control of the boxed term (the crucial result for this being Lemma 6.12), without disrupting the solvability of TΦ0T_{\Phi_{0}}.

2.5.2. Non-local Deformation Operator

The alternating iteration scheme only converges because of the decay property (II) and (II’) in Sections 2.2 and 2.4, which causes each iteration to reduce the error by a factor δ\delta of the decay across the neck region. In particular, in Section 2.2 the error on YY^{-} is supported where dχ+0d\chi^{+}\neq 0 and the solution in Step (1) decays across the region between the supports of dχ±d\chi^{\pm}. The first stage in (2.17) disrupts this property:

(Difficulty II): The projection and deformation operator (1Π0)(1-\Pi_{0}) and TΦ1T_{\Phi_{1}} are non-local.

In particular, both terms of (2.19) are supported globally on YY, regardless of the support of the error term in Step (0).

This at first makes it seem that there can be no decay factor δ\delta^{-} when solving on YY^{-}. Were the decay factor δ+\delta^{+} on Y+Y^{+} sufficiently strong, this would present no issue, but in our situation C(ε)δ+=ε1/24C(\varepsilon)\delta^{+}=\varepsilon^{-1/24} in (2.20) grows as ε0\varepsilon\to 0 due to the non-invertibility of the linearization on Y+Y^{+} and so δ\delta^{-} must at least compensate for this factor. This is achieved by

(Solution II): Sobolev spaces of spinors with “effective” support.

The notion of effective support (Definition 4.6) generalizes the standard support, and imposes constraints on how the norms of a spinor grow as a weight increases. The results of Sections 6, and 9.2 show that the mode-dependent deformations from Subsection 2.5.1 above mean that the deformation operator preserves effective support, and this is enough for a decay result.

2.5.3. Non-Uniform Invertibility

This final difficulty is the problem that was solved in [41], and the solution may be treated as a black box here. It still bears mentioning, however, because it makes Difficulties I and II above more unforgiving. Since the model solutions (Φε+,Aε+)(\Phi_{\varepsilon}^{+},A_{\varepsilon}^{+}) on Y+Y^{+} converge to (Φ0,A0)(\Phi_{0},A_{0}), the family of elliptic linearizations +\mathcal{L}^{+} converge to the degenerate elliptic linearization at ε=0\varepsilon=0 with infinite-dimensional obstruction.

(Difficulty III): The linearizations ε+\mathcal{L}^{+}_{\varepsilon} are not uniformly invertible in ε\varepsilon.

This situation is effectively mandated by the fact that there must be an infinite-dimensional subspace that becomes the infinite-dimensional obstruction in the limit, and the accompanying elliptic estimates necessarily fail to be uniform on this subspace (in any norms with locally defined norm density).

It is not overly difficult to achieve estimates showing that the norm of (+)1(\mathcal{L}^{+})^{-1} is O(εβ)O(\varepsilon^{-\beta}) for some β>0\beta>0. All naive approaches, however, frustratingly result in the constant β\beta that makes the powers of ε\varepsilon in (2.20) precisely cancel, preventing convergence (even once Solutions I and II are optimized). In order to achieve stronger control, one must identify quite precisely the subspace from which the infinite-dimensional obstruction arises, and treat this subspace independently. This was achieved in [41] by introducing boundary conditions which capture a certain high-dimensional subspace that approaches the obstruction.

(Solution III): Use twisted APS boundary conditions.

These boundary conditions are packaged in Lemma 7.9. Details may be found in [41, Section 7].

Appendix of Gluing Parameters

This appendix collects the notation and gluing parameters used throughout the remainder of the article. References to the section where the items are defined in the body of the text are included.

(1) Notation for the Seiberg–Witten Equations

  • SE=SES_{E}=S\otimes_{\mathbb{C}}E the two-spinor bundle (Section 3.1).

  • SRe,SImS^{\text{Re}},S^{\text{Im}} the real and imaginary subbundles of SES_{E}, defined over Y𝒵Y-\mathcal{Z} (Section 3.2).

  • Ω=Ω0(i)Ω1(i)\Omega=\Omega^{0}(i\mathbb{R})\oplus\Omega^{1}(i\mathbb{R}) the bundle of 𝔲(1)=i\mathfrak{u}(1)=i\mathbb{R}-valued 0 and 11-forms.

  • h0=(Φτ,Aτ)h_{0}=(\Phi_{\tau},A_{\tau}) the family of 2\mathbb{Z}_{2}-eigenvectors from Theorem (1.4), assumed to satisfy (1.5).

  • h1=(Φ1,A1)h_{1}=(\Phi_{1},A_{1}) the initial global approximate solutions on YY (Section 7.2).

  • (φ,a)(\varphi,a) used to denote corrections on Y+Y^{+} (Section 7.1).

  • (ψ,b)(\psi,b) used to denote corrections on YY^{-} (Sections 4.2 and 7.3).

  • (Φ,A)\mathcal{L}_{(\Phi,A)} the linearized SW equations at the re-normalized configuration (ε1Φ,A)(\varepsilon^{-1}\Phi,A) (Section 4.1).

  • ,d\not{\mathbb{D}},\text{d}\not{\mathbb{D}} the universal Dirac operator and its linearization (Section 6.1).

  • 𝕊𝕎,d𝕊𝕎\mathbb{SW},\text{d}\mathbb{SW} the universal Seiberg–Witten equations and their linearization (Section 8.2).

(2) Gluing Parameters

  • ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}) the L2L^{2}-norm parameter.

  • τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}) the coordinate along the parameter path pτ=(gτ,Bτ)p_{\tau}=(g_{\tau},B_{\tau}). Assumed to satisfy (1.5).

  • δ=ε1/48\delta=\varepsilon^{1/48} the convergence factor from a single cycle of (1.14).

  • γ,γ±,γ¯<<1\gamma,\gamma^{\pm},\underline{\gamma}<<1 small numbers, say O(106)O(10^{-6}).

  • η\eta a linearized deformation of 𝒵τ\mathcal{Z}_{\tau} (Section 6.1).

  • ξ=εη\xi=\varepsilon\eta a re-normalized linearized deformation (Section 9.1).

  • ν+=14106\nu^{+}=\tfrac{1}{4}-10^{-6} the inside weight.

  • ν=12106\nu^{-}=\tfrac{1}{2}-10^{-6} the outside weight.

(3) Regions and Cut-offs

  • Y+=Nλ(𝒵τ)Y^{+}=N_{\lambda}(\mathcal{Z}_{\tau}) the inside region: a tubular neighborhood of radius λ=ε1/2\lambda=\varepsilon^{1/2} (Section 7).

  • Y=YNλ(𝒵τ)Y^{-}\!=\!Y\!-\!N_{\lambda^{-}}(\mathcal{Z}_{\tau}) the outside region: the complement of a neighborhood of radius λ=ε2/3γ\lambda^{-}=\varepsilon^{2/3-\gamma^{-}}.

  • χ+\chi^{+} a logarithmic cut-off equal to 11 for rλ/4r\leq\lambda/4 and vanishing for rλ/2r\geq\lambda/2.

  • χ\chi^{-} a logarithmic cut-off equal to 11 for rλ/2r\geq\lambda^{-}/2 and vanishing for rλ/4r\leq\lambda^{-}/4.

  • 𝟙+\mathbb{1}^{+} the indicator function of the set rε2/3γr\leq\varepsilon^{2/3-\gamma^{-}}.

  • 𝟙=1𝟙+\mathbb{1}^{-}=1-\mathbb{1}^{+} the indicator function of the complementary region.

Important Remark.

The small numbers ε0,τ0\varepsilon_{0},\tau_{0} and γ\gamma are allowed to decrease a finite number of times between successive appearances over the course of the proof of Theorem 1.6. The proof of Proposition 11.3 in Section 11 collects the accumulated factors of γ\gamma. The numbers γ±,γ¯\gamma^{\pm},\underline{\gamma} remain fixed throughout.

3. 2\mathbb{Z}_{2}-Harmonic Spinors and Compactness

This section reviews the compactness properties of the two-spinor Seiberg–Witten equations from [32], and begins the set-up of the gluing analysis. More detailed expositions of the same material may be found in [41, 16, 61, 32, 52].

3.1. Compactness Theorem

Let YY be a compact, oriented, 3-manifold. With S,EYS,E\to Y as in Section 1.1 and p=(g,B)p=(g,B) as in (1.8), set SE:=SES_{E}:=S\otimes_{\mathbb{C}}E. Clifford multiplication on SS induces a Clifford multiplication γ:¯TYEnd(SE)\gamma:\underline{\mathbb{R}}\oplus T^{*}Y\to\text{End}(S_{E}) which acts trivially on EE. Define the moment map μ:SEΩ1(i)\mu:S_{E}\to\Omega^{1}(i\mathbb{R}) by

12μ(Ψ,Ψ)=j=13i2iγ(ej)Ψ,Ψej.\frac{1}{2}\mu(\Psi,\Psi)=\sum_{j=1}^{3}\frac{i}{2}\langle i\gamma(e^{j})\Psi,\Psi\rangle e^{j}. (3.1)

where {ej}\{e^{j}\} is a local orthonormal frame of TYT^{*}Y. Unlike for the single-spinor Seiberg–Witten equations, there are 0ΨΓ(SE)0\neq\Psi\in\Gamma(S_{E}) with μ(Ψ,Ψ)=0\mu(\Psi,\Psi)=0.

For the two-spinor Seiberg–Witten equations (1.11.2) to be an elliptic system on a 3-manifold, the an auxiliary 0-form is required. Revising the notation slightly, let A=(a0,A1)Ω0(i)𝒜U(1)A=(a_{0},A_{1})\in\Omega^{0}(i\mathbb{R})\otimes\mathcal{A}_{U(1)}, and denote

A=A1+γ(a0)FA=FA1da0.\not{D}_{A}=\not{D}_{A_{1}}+\gamma(a_{0})\hskip 56.9055pt\star F_{A}=\star F_{A_{1}}-da_{0}. (3.2)

where A1\not{D}_{A_{1}} is the Dirac operator on SES_{E} formed using the spin connection of gg, the connection A1A_{1}, and the background connection BB on EE, and FA1F_{A_{1}} is the curvature of A1A_{1}.

Definition 3.1.

The (extended) two-spinor Seiberg-Witten Equations for the parameter pp are the following equations for configurations (Ψ,A)Γ(SE)×(Ω0(i)×𝒜U(1))(\Psi,A)\in\Gamma(S_{E})\times(\Omega^{0}(i\mathbb{R})\times\mathcal{A}_{U(1)}):

AΨ\displaystyle\not{D}_{A}\Psi =\displaystyle= 0\displaystyle 0 (3.3)
FA+12μ(Ψ,Ψ)\displaystyle\star F_{A}+\tfrac{1}{2}\mu(\Psi,\Psi) =\displaystyle= 0\displaystyle 0 (3.4)

where FA\star F_{A} and A\not{D}_{A} are as in (3.2). These equations are invariant under the action of the gauge group 𝒢=Maps(Y;U(1))\mathcal{G}=\text{Maps}(Y;U(1)).

Notice that definition (3.1) of μ\mu differs by a sign from what is used by many other authors [37, 34, 15]. Consequently, the sign in the curvature equation (3.4) also reverses, leaving the equations unaltered. If (Ψ,A)(\Psi,A) solves (3.33.4) and Ψ0\Psi\neq 0, then integration by parts shows that a0=0a_{0}=0. For the purposes of Theorem 1.6, it therefore suffices to solve the extended equations. From here onward, we work exclusively with the extended equations.

Standard elliptic theory shows that a sequence of solutions to (3.33.4) with an a priori bound on the spinors’ L2L^{2}-norm admits a convergent subsequence [37, 34, 32]. In the case of the standard (one-spinor) Seiberg–Witten equations, such an a priori bound follows from the Weitzenböck formula, the maximum principle, and a pointwise identity for μ\mu. In fact, the bound in that case is ΨL2s02\|\Psi\|_{L^{2}}\leq\tfrac{s_{0}}{2} where s0s_{0} is the C0C^{0}-norm of the scalar curvature. In the case of two-spinors, the same pointwise identity for μ\mu fails, and there may be sequences of solutions (Ψi,Ai)(\Psi_{i},A_{i}) such that ΨiL2\|\Psi_{i}\|_{L^{2}}\to\infty which therefore admit no convergent subsequence.

The behavior of such sequences can be understood by renormalizing the spinor to have unit L2L^{2}-norm. Thus, with ε=1ΨL2\varepsilon=\frac{1}{\|\Psi\|_{L^{2}}} we define renormalized spinors

Φ:=εΨ\boxed{\Phi:=\varepsilon\Psi} (3.5)

so that ΦL2(Y)=1\|\Phi\|_{L^{2}(Y)}=1. As in the introduction, the re-normalized or blown-up Seiberg-Witten equations for a triple (Φ,A,ε)(\Phi,A,\varepsilon) become (1.31.5), where A=(a0,A1)A=(a_{0},A_{1}) is now as in (3.2) . The following theorem of Haydys–Walpuski describes the convergence behavior of sequences of solutions to the blown-up equations. The theorem states that sequences of solutions along which ε0\varepsilon\to 0 converge to solution of the ε=0\varepsilon=0-version of (1.31.5) away from a singular set 𝒵0\mathcal{Z}_{0}.

Theorem 3.2.

(Haydys–Walpuski [32, 42, 66, 51]) Suppose that (Φi,Ai,εi)(\Phi_{i},A_{i},\varepsilon_{i}) is a sequence of solutions to (1.31.5) with respect to a converging sequence of parameters pip0=(g0,B0)p_{i}\to p_{0}=(g_{0},B_{0}) such that εi0\varepsilon_{i}\to 0. Then there exists a triple (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) where

  • 𝒵0Y\mathcal{Z}_{0}\subset Y is a closed, rectifiable subset of Hausdorff codimension 2,

  • A0A_{0} is a flat U(1)U(1)-connection on Y𝒵0Y-\mathcal{Z}_{0} with holonomy in 2\mathbb{Z}_{2},

  • Φ0\Phi_{0} is a spinor on Y𝒵0Y-\mathcal{Z}_{0} such that |Φ0||\Phi_{0}| extends continuously to YY with 𝒵0=|Φ0|1(0)\mathcal{Z}_{0}=|\Phi_{0}|^{-1}(0),

satisfying

A0Φ0=0μ(Φ0,Φ0)=0Φ0L2=1\not{D}_{A_{0}}\Phi_{0}=0\hskip 42.67912pt\mu(\Phi_{0},\Phi_{0})=0\hskip 42.67912pt\|\Phi_{0}\|_{L^{2}}=1 (3.6)

on Y𝒵0Y-\mathcal{Z}_{0}, and after passing to a subsequence, ΦiΦ0\Phi_{i}\to\Phi_{0}, and AiA0A_{i}\to A_{0} in Cloc(Y𝒵0)C^{\infty}_{loc}(Y-\mathcal{Z}_{0}) modulo gauge transformations, and |Φi||Φ0||\Phi_{i}|\to|\Phi_{0}| in C0,α(Y)C^{0,\alpha}(Y) for some α>0\alpha>0.

Remark 3.3.

The above statement combines the original result of Haydys–Walpuski with refinements proved by Taubes [51], Zhang [66] and the author [42]. In [51], Taubes showed that the singular set 𝒵0\mathcal{Z}_{0} has finite 1-dimensional Hausdorff content; building on this Zhang showed in [66] that the singular set is rectifiable. In [42], the author improved the convergence to the limit from weak Lloc2,2L^{2,2}_{loc} for the spinor and weak Lloc1,2L^{1,2}_{loc} for the connection to ClocC^{\infty}_{loc} for both. Taubes also proved a four-dimensional version of Theorem 3.2 in [52], to which the same refinements apply.

Remark 3.4.

Theorem 3.2 is a particular instance of a family of similar compactness results for generalized Seiberg–Witten equations which originated with C. Taubes’s generalization of Uhlenbeck Compactness to PSL(2,)\text{PSL}(2,\mathbb{C}) connections. Generalizations and similar results for other generalized Seiberg–Witten equations can be found in [54, 53, 52, 49, 65, 61, 45].

3.2. The Haydys Corresondence

The limiting configurations (Φ0,A0)(\Phi_{0},A_{0}) in Theorem 3.2 are equivalent to 2\mathbb{Z}_{2}-harmonic spinors as defined in the Section 1.2. This equivalence is a particular instance of the Haydys Correspondence [21, 15].

A limiting configuration as in Theorem 3.2 induces a decomposition of the restriction of the two-spinor bundle SES_{E} to Y𝒵0Y-\mathcal{Z}_{0} as follows. Since A0A_{0} is flat with holonomy in 2\mathbb{Z}_{2}, det(S)|Y𝒵0¯\det(S)|_{Y-\mathcal{Z}_{0}}\simeq\underline{\mathbb{C}} is trivial, and S|Y𝒵0S|_{Y-\mathcal{Z}_{0}} admits a reduction of structure to SU(2)SU(2). Thus, there is a “charge conjugation” map JEnd(S|Y𝒵0)J\in\text{End}(S|_{Y-\mathcal{Z}_{0}}) such that J2=IdJ^{2}=-\text{Id}; since EE is an SU(2)SU(2)-bundle it admits a similar map, denoted jj. The product σ=Jj\sigma=J\otimes_{\mathbb{C}}j satisfies σ2=Id\sigma^{2}=\text{Id}, i.e. it is a real structure on SE|Y𝒵0S_{E}|_{Y-\mathcal{Z}_{0}}. Consequently, there is a decomposition

SE|Y𝒵0=SReSImS_{E}|_{Y-\mathcal{Z}_{0}}=S^{\text{Re}}\oplus S^{\text{Im}} (3.7)

where

SRe={12(Ψ+σΨ)|ΨΓ(SE|Y𝒵0)}SIm={12(ΨσΨ)|ΨΓ(SE|Y𝒵0)}S^{\text{Re}}=\{\tfrac{1}{2}(\Psi+\sigma\Psi)\ |\ \Psi\in\Gamma(S_{E}|_{Y-\mathcal{Z}_{0}})\}\hskip 56.9055ptS^{\text{Im}}=\{\tfrac{1}{2}(\Psi-\sigma\Psi)\ |\ \Psi\in\Gamma(S_{E}|_{Y-\mathcal{Z}_{0}})\}

are the “real” and “imaginary” subbundles respectively.

These subbundles have the following useful characterization, which is proved in [41, Sec. 2], and [16, Sec. 3].

Lemma 3.5.

Let (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) be a limiting configuration as in Theorem 3.2. The decomposition (3.7) satisfies the following:

  1. (1)

    The decomposition is parallel with respect to the connection A0\nabla_{A_{0}} induced by A0A_{0}.

  2. (2)

    Clifford multiplication by \mathbb{R}-valued forms preserves the decomposition, i.e.

    γ:(Ω0Ω1)()×SReSRe\gamma:(\Omega^{0}\oplus\Omega^{1})(\mathbb{R})\times S^{\text{Re}}\to S^{\text{Re}}

    and likewise for SImS^{\text{Im}}. Conversely, Clifford multiplication by ii\mathbb{R}-valued forms reverses it.

  3. (3)

    Φ0Γ(SRe)\Phi_{0}\in\Gamma(S^{\text{Re}}), and there exists a spin structure on YY with spinor bundle S0S_{0} and a real Euclidean line bundle Y𝒵0\ell\to Y-\mathcal{Z}_{0} such that

    SReS0S^{\text{Re}}\simeq S_{0}\otimes_{\mathbb{R}}\ell

    on Y𝒵0Y-\mathcal{Z}_{0}. Moreover, under this isomorphism, A0\nabla_{A_{0}} is taken to the connection formed from the spin connection on S0S_{0} and the unique flat connection on \ell, with an \mathbb{R}-linear perturbation commuting with γ\gamma arising from B0B_{0}. ∎

As a consequence of items (1) and (2) above, the Dirac operator on SES_{E} restricts to a Dirac operator

A0Re:Γ(SRe)Γ(SRe),\not{D}_{A_{0}}^{\text{Re}}:\Gamma(S^{\text{Re}})\to\Gamma(S^{\text{Re}}), (3.8)

and likewise for the imaginary part. When the subbundle in question is evident, we will omit the superscript from the notation. The isomorphism in Item (3) intertwines (3.8) and the Dirac operator on S0S_{0}\otimes_{\mathbb{R}}\ell formed using the connection in Item (3). This leads to the following equivalence, which is the manifestation of the Haydys Correspondence in this particular setting.

Corollary 3.6.

Suppose that 𝒵0Y\mathcal{Z}_{0}\subset Y is a smooth, embedded link. Then the data of a limiting configuration satisfying (3.6) as in Theorem 3.2 is equivalent to a 2\mathbb{Z}_{2}-Harmonic spinor (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) on SReS0S^{\text{Re}}\simeq S_{0}\otimes_{\mathbb{R}}\ell satisfying

A0Φ0=0A0Φ0L2\not{D}_{A_{0}}\Phi_{0}=0\hskip 56.9055pt\nabla_{A_{0}}\Phi_{0}\in L^{2}

with respect to (g0,B0)(g_{0},B_{0}).

Proof.

Except for the integrability condition, the corollary follows directly from isomorphism of item (3) in Lemma 3.5. The fact that Φ0L2\nabla\Phi_{0}\in L^{2} will follow from Lemma 4.5 in Section 4, which shows requirement that Φ0L2\nabla\Phi_{0}\in L^{2} is equivalent (for regular 2\mathbb{Z}_{2}-harmonic spinors) to requiring that |Φ0||\Phi_{0}| extend continuously over 𝒵0\mathcal{Z}_{0} with 𝒵0|Φ0|1(0)\mathcal{Z}_{0}\subset|\Phi_{0}|^{-1}(0). ∎

The purpose of the Haydys Correspondence is that it takes advantage of the gauge freedom to temper the singular nature of the limiting equations. To explain this further, limiting configurations are considered up to U(1)U(1) gauge transformations and solve the globally degenerate system of equations (3.6); indeed, the symbol (after gauge-fixing) in the curvature equation of (1.4) vanishes everywhere as ε0\varepsilon\to 0, leading to a loss of ellipticity. On the other side of the Haydys correspondence, 2\mathbb{Z}_{2}-harmonic spinors are considered only up to the action of 2\mathbb{Z}_{2} by sign on S0S_{0}\otimes_{\mathbb{R}}\ell, and solve the Dirac equation on S0S_{0}, which is a singular elliptic equation whose symbol degenerates only locally along 𝒵0\mathcal{Z}_{0} (see the local description in Section 4.1). While the first type of degeneracy appears to at first be rather intractable, the latter description places the problem in the well-studied class of elliptic edge problems [36, 39].

Remark 3.7.

Items (1) and (2) show that SReS^{\text{Re}} is a 4-dimensional real Clifford module on Y𝒵0Y-\mathcal{Z}_{0}. The isomorphism in Item (3) of Lemma 3.5 endows it with a complex structure, but not canonically so. In particular, the induced Dirac operator (3.8) is only \mathbb{R}-linear if the SU(2)SU(2)-connection BB is non-trivial (as it must be for condition (2) of Definition 1.1 to be met). In four-dimensions, the bundle SReS^{\text{Re}} need not be topologically isomorphic to the spinor bundle twisted by a real line bundle.

3.3. Recovering Spinc\text{Spin}^{c} Structures

By Corollary 3.6, a limiting configuration (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) gives rise to a 2\mathbb{Z}_{2}-harmonic spinor. The reverse is not immediately true because, in contrast to the Seiberg–Witten equations, the definition of a 2\mathbb{Z}_{2}-harmonic spinor makes no references to a Spinc\text{Spin}^{c} structure. The topological information of the Spinc\text{Spin}^{c} structure is lost in the limiting process of Theorem 3.2 and must be reconstructed before the gluing analysis begins.

Specifically, we seek a Spinc\text{Spin}^{c} structure with spinor bundle SS such that SReS^{\text{Re}} as defined by (3.7) satisfies the isomorphism of Item (3) from Lemma 3.5 for the twisted spinor bundle S0S_{0}\otimes_{\mathbb{R}}\ell that hosts Φ0\Phi_{0}. Given such an SS, Corollary 3.6 implies that (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) may be viewed as a (non-smooth) configuration on the subbundle SReSES^{\text{Re}}\subset S_{E} of two-spinor bundle formed from SS, and the gluing analysis begins from there.

The following lemma reconstructs the correct Spinc\text{Spin}^{c}-structure for the gluing, given an orientation of 𝒵0\mathcal{Z}_{0}. The proof may be found in Section 3 of [41]; see also [24] for more results in this direction.

Lemma 3.8.

Let (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) be a regular 2\mathbb{Z}_{2}-harmonic spinor on S0S_{0}\otimes_{\mathbb{R}}\ell. An orientation of 𝒵0\mathcal{Z}_{0} determines a unique Spinc\text{Spin}^{c}-structure with spinor bundle SYS\to Y satisfying the following criteria.

  1. (1)

    The first chern class satisfies

    c1(S)=PD[𝒵0]c_{1}(S)=-\text{PD}[\mathcal{Z}_{0}]

    with the specified orientation of 𝒵0\mathcal{Z}_{0}.

  2. (2)

    SS extends S0S_{0}\otimes_{\mathbb{R}}\ell in the sense that S|Y𝒵0S0S|_{Y-\mathcal{Z}_{0}}\simeq S_{0}\otimes_{\mathbb{R}}\ell. Moreover, there is an isomorphism

    S0SReSES_{0}\otimes_{\mathbb{R}}\ell\simeq S^{\text{Re}}\subset S_{E}

    where SE=SES_{E}=S\otimes_{\mathbb{C}}E, which makes Φ0\Phi_{0} a smooth section of SReY𝒵0S^{\text{Re}}\to Y-\mathcal{Z}_{0}.

Notice that we do not assume 𝒵0\mathcal{Z}_{0} is connected, thus there are 2k2^{k} possible choices of orientation when 𝒵0\mathcal{Z}_{0} has kk components.

Given Lemma 3.8, the data of a regular 2\mathbb{Z}_{2}-harmonic spinor (with an orientation of 𝒵0\mathcal{Z}_{0}) is equivalent to that of a limiting configuration (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) using the induced Spinc\text{Spin}^{c}-structure on YY. From here on, we therefore cease to distinguish between a regular (oriented) 2\mathbb{Z}_{2}-harmonic spinor (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) and the corresponding limiting configuration denoted (purposefully) by the same triple.

3.4. Adapted Coordinates

In order to describe Seiberg–Witten configurations in a neighborhood of 𝒵0\mathcal{Z}_{0}, we use adapted coordinate systems, constructed as follows.

Fix a component 𝒵j\mathcal{Z}_{j} of 𝒵0\mathcal{Z}_{0} with length |𝒵j||\mathcal{Z}_{j}| and an arclength parameterization p:S1𝒵jp:S^{1}\to\mathcal{Z}_{j}. Choose an orthonormal frame {n1,n2}\{n_{1},n_{2}\} for the pullback pN𝒵0p^{*}N\mathcal{Z}_{0} of the normal bundle, ordered so that {p˙,n1,n2}\{\dot{p},n_{1},n_{2}\} is an oriented frame of pTYp^{*}TY along 𝒵j\mathcal{Z}_{j}. Let Nr(𝒵0)N_{r}(\mathcal{Z}_{0}) denote the tubular neighborhood of radius rr around 𝒵0\mathcal{Z}_{0}, measured in the geodesic distance of g0g_{0}.

Definition 3.9.

A system of Fermi coordinates (t,x,y)(t,x,y) of radius r0<rinjr_{0}<r_{\text{inj}} where rinjr_{\text{inj}} is the injectivity radius of (Y,g0)(Y,g_{0}) is the diffeomorphism jS1×Dr0Nr0(𝒵0)\bigsqcup_{j}S^{1}\times D_{r_{0}}\simeq N_{r_{0}}(\mathcal{Z}_{0}) defined by

(t,x,y)Exp𝔭(t)(xn1+yn2).(t,x,y)\mapsto\text{Exp}_{\mathfrak{p}(t)}(xn_{1}+yn_{2}).

on each component of 𝒵0\mathcal{Z}_{0}, where t[0,|𝒵j|)t\in[0,|\mathcal{Z}_{j}|) is the normalized coordinate in the S1S^{1} direction. In these coordinates the metric g0g_{0} has the form

g0=dt2+dx2+dy2+O(r).g_{0}=dt^{2}+dx^{2}+dy^{2}\ +\ O(r). (3.9)

We denote the corresponding cylindrical coordinates by (t,r,θ)(t,r,\theta). Given a smooth family (gτ,𝒵τ)(g_{\tau},\mathcal{Z}_{\tau}) as in Theorem 1.4 it may be arranged that the Fermi coordinate systems depend smoothly on τ\tau.

A system of Fermi coordinates induces a trivialization of (Ω0Ω1)(i)(\Omega^{0}\oplus\Omega^{1})(i\mathbb{R}). The next lemma, which is proved in Section 3 of [41], provides a satisfactory trivialization of the two-spinor bundle SES_{E}.

Lemma 3.10.

In the neighborhood N=Nr0(𝒵j)N=N_{r_{0}}(\mathcal{Z}_{j}) of each component of 𝒵j\mathcal{Z}_{j}, there is a local trivialization

(SE)|NN×(2)(S\otimes_{\mathbb{C}}E)|_{N}\simeq N\times(\mathbb{C}^{2}\otimes_{\mathbb{C}}\mathbb{H})

with the following properties.

  • (1)

    The connection A0A_{0} has the form

    A0:=d+i2dθ+O(1).A_{0}:=\text{d}\ +\ \frac{i}{2}d\theta\ +\ O(1).
  • (2)

    There is an ϵj{0,1}\epsilon_{j}\in\{0,1\} such that the restriction of SReS^{\text{Re}} to NN is given by

    SRe|Nr0(𝒵j)={(αβ)1+eiθeiϵjt(β¯α¯)j|α,β:N}.S^{\text{Re}}\big{|}_{N_{r_{0}}(\mathcal{Z}_{j})}=\left\{\begin{pmatrix}\alpha\\ \beta\end{pmatrix}\otimes 1\ +\ e^{-i\theta}e^{-i\epsilon_{j}t}\begin{pmatrix}-\overline{\beta}\\ \overline{\alpha}\end{pmatrix}\otimes j\ \Big{|}\ \alpha,\beta:N\to\mathbb{C}\right\}.

Again, for the τ\tau-parameterized family of Theorem 1.4, and it may be assumed that this family of trivializations depends smoothly on τ\tau (using 𝒵τ\mathcal{Z}_{\tau} and AτA_{\tau} respectively).

4. The Singular Linearization

The linearized Seiberg–Witten equations play an essential role in carrying out the alternating iteration outlined in Section 2.4. This section introduces the linearized equations, which are a singular elliptic system at a 2\mathbb{Z}_{2}-harmonic spinor.

4.1. Singular Linearization

Differentiating, (3.33.4), the linearized (extended) Seiberg–Witten equations at a (renormalized) configuration (Φ,A)(\Phi,A) acting on a linearized deformation (φ,a)(\varphi,a) are

Aφ+γ(a)Φε\displaystyle\not{D}_{A}\varphi+\gamma(a)\tfrac{\Phi}{\varepsilon} =\displaystyle= 0\displaystyle 0 (4.1)
(d,d)a+μ(φ,Φ)ε\displaystyle(\star d,-d)a+\tfrac{\mu(\varphi,\Phi)}{\varepsilon} =\displaystyle= 0\displaystyle 0 (4.2)

where ε\varepsilon is as in (3.5).

To make (4.14.2) into an elliptic system, we impose the Ω0(i)\Omega^{0}(i\mathbb{R})-valued gauge-fixing condition

daiiφ,Φε=0,-d^{\star}a-\frac{i\langle i\varphi,\Phi\rangle}{\varepsilon}=0, (4.3)

where dd^{\star} denotes the adjoint of the exterior derivative. Extend (the polarization of) μ\mu to a bilinear map μ:SESE(Ω0Ω1)(i)\mu:S_{E}\otimes S_{E}\to(\Omega^{0}\oplus\Omega^{1})(i\mathbb{R}) by

μ(φ,ψ)\displaystyle\mu(\varphi,\psi) :=\displaystyle:= (iiφ,ψ,μ1(φ,ψ))\displaystyle(i\langle i\varphi,\psi\rangle,\mu_{1}(\varphi,\psi))

where μ1\mu_{1} is what were previously denoted by μ\mu.

Lemma 4.1.

Suppose that (Φ,A)(\Phi,A) is a smooth configuration on YY. Then the (extended, gauge-fixed) linearized Seiberg–Witten equations at (Φ,A)(\Phi,A) on a linearized deformation (φ,a)Γ(SE)(Ω0Ω1)(i)(\varphi,a)\in\Gamma(S_{E})\oplus(\Omega^{0}\oplus\Omega^{1})(i\mathbb{R}) take the form:

(Φ,A)(φ,a)=(Aγ(_)Φεμ(_,Φ)ε𝕕)(φa)\mathcal{L}_{(\Phi,A)}(\varphi,a)=\begin{pmatrix}\not{D}_{A}&\gamma(\_)\tfrac{\Phi}{\varepsilon}\\ \tfrac{\mu(\_,\Phi)}{\varepsilon}&\mathbb{d}\end{pmatrix}\begin{pmatrix}\varphi\\ a\end{pmatrix} (4.4)

where 𝕕a=(0ddd)(a0a1)\mathbb{d}a=\begin{pmatrix}0&-d^{\star}\\ -d&\star d\end{pmatrix}\begin{pmatrix}a_{0}\\ a_{1}\end{pmatrix}. Moreover, (Φ,A)\mathcal{L}_{(\Phi,A)} is a self-adjoint elliptic operator. ∎

Notice that the parameter ε\varepsilon is kept implicit in the notation (Φ,A)\mathcal{L}_{(\Phi,A)}

A 2\mathbb{Z}_{2}-harmonic spinor (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) (or eigenvector) is not a smooth configuration on YY, and the ellipticity in Lemma 4.1 fails when linearizing at these. At a 2\mathbb{Z}_{2}-harmonic spinor, however, the linearization admits a pleasing block form with respect to the decomposition of Lemma 3.5. Note in this case the operator acts on sections of bundles only defined over Y𝒵0Y-\mathcal{Z}_{0}.

Lemma 4.2.

The (extended, gauge-fixed) linearized Seiberg–Witten equations at (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) take the following form on a linearized deformation (φ1,φ2,a)Γ(SRe)Γ(SIm)(Ω0Ω1)(i)(\varphi_{1},\varphi_{2},a)\in\Gamma(S^{\text{Re}})\oplus\Gamma(S^{\text{Im}})\oplus(\Omega^{0}\oplus\Omega^{1})(i\mathbb{R}):

(Φ0,A0)(φ1,φ2,a)=(A0000A0γ(_)Φ0ε0μ(_,Φ0)ε.𝕕)(φ1φ2a)\mathcal{L}_{(\Phi_{0},A_{0})}(\varphi_{1},\varphi_{2},a)=\begin{pmatrix}\not{D}_{A_{0}}&0&0\\ 0&\not{D}_{A_{0}}&\gamma(\_)\frac{\Phi_{0}}{\varepsilon}\\ 0&\frac{\mu(\_,\Phi_{0})}{\varepsilon}.&\mathbb{d}\end{pmatrix}\begin{pmatrix}\varphi_{1}\\ \varphi_{2}\\ a\end{pmatrix} (4.5)

The same applies at an eigenvector (𝒵τ,Aτ,Φτ)(\mathcal{Z}_{\tau},A_{\tau},\Phi_{\tau}).

Proof.

The proof is an immediate consequence of Lemma 3.5. ∎

As explained in the introduction, the Dirac operator A0\not{D}_{A_{0}} at the singular connection A0A_{0} is a degenerate elliptic edge operator111An elliptic edge operator is an elliptic combination of the derivatives rr,θ,rtr\partial_{r},\partial_{\theta},r\partial_{t} in Fermi coordinates; technically speaking, rA0r\not{D}_{A_{0}} is the edge operator in question, but the factor of rr only shifts the weight. [36]. In the local coordinates and trivializations of Lemma 3.10, the degenerate nature becomes manifest. Near 𝒵0\mathcal{Z}_{0}, it has the form

A0=(it22¯it)+14r(0eiθeiθ0)+𝔡1+𝔡0\not{D}_{A_{0}}=\begin{pmatrix}i\partial_{t}&-2\partial\\ 2\overline{\partial}&-i\partial_{t}\end{pmatrix}\ +\ \frac{1}{4r}\begin{pmatrix}0&e^{-i\theta}\\ -e^{i\theta}&0\end{pmatrix}\ +\ \mathfrak{d}_{1}\ +\ \mathfrak{d}_{0} (4.6)

where 𝔡1=O(r)\mathfrak{d}_{1}=O(r)\nabla is a first order operator vanishing along 𝒵0\mathcal{Z}_{0}, and 𝔡0=O(1)\mathfrak{d}_{0}=O(1) is a bounded zeroth order operator. In particular, the second term, which arises from the non-trivial holonomy of A0A_{0}, is unbounded on L2L^{2}. Equivalently, rA0r\not{D}_{A_{0}} is a elliptic operator with L2L^{2}-bounded terms whose symbol degenerates along 𝒵0\mathcal{Z}_{0}. In fact, there are no function spaces with local norms for which the extension of this operator is Fredholm.

The remainder of Sections 45 develop the analysis of this operator. The lower 2×22\times 2 block in (4.5) is analyzed in Section 7.3, using a trick that reduces it to standard elliptic theory.

4.2. The Singular Dirac Operator

This section summarizes results from the elliptic edge theory of the singular Dirac operator (4.6). More generally, Lemmas 3.5 and 4.2 (and the expression 4.6) apply for any τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}), and we consider the family of operators

Aτ:Γ(Y𝒵τ;SRe)Γ(Y𝒵τ;SRe)\not{D}_{A_{\tau}}:\Gamma(Y-\mathcal{Z}_{\tau};S^{\text{Re}})\longrightarrow\Gamma(Y-\mathcal{Z}_{\tau};S^{\text{Re}}) (4.7)

where the dependence of SReS^{\text{Re}} on τ\tau is suppressed in the notation. More detailed discussion and proofs of the statements in this subsection may be found in Sections 2–4 of [43], and a discussion from the perspective of the microlocal analysis of elliptic edge operators is contained in [28].

To begin, consider the following function spaces. Let rr be a smooth weight function such that

r(y)={dist(y,𝒵τ)yNr0/2(𝒵τ)const.yYNr0(𝒵τ)r(y)=\begin{cases}\text{dist}(y,\mathcal{Z}_{\tau})\hskip 28.45274pty\in N_{r_{0}/2}(\mathcal{Z}_{\tau})\\ \text{const.}\hskip 48.36958pty\in Y-N_{r_{0}}(\mathcal{Z}_{\tau})\end{cases} (4.8)

where the distance is measured using the metric gτg_{\tau} (though we omit this from the notation), and r0r_{0} is as in Definition 3.9.

Definition 4.3.

For a constant ν\nu\in\mathbb{R}, the weighted edge Sobolev spaces (of regularity m=0,1m=0,1) are defined by

r1+νHe1(Y𝒵τ;SRe)\displaystyle r^{1+\nu}H^{1}_{e}(Y-\mathcal{Z}_{\tau};S^{\text{Re}}) :=\displaystyle:= {u|Y𝒵τ(|u|2+|u|2r2)r2νdV<}\displaystyle\Big{\{}\ u\ \ \Big{|}\ \ \int_{Y-\mathcal{Z}_{\tau}}\left(|\nabla u|^{2}+\frac{|u|^{2}}{r^{2}}\ \right)r^{-2\nu}dV\ <\ \infty\Big{\}} (4.9)
rνL2(Y𝒵τ;SRe)\displaystyle r^{\nu}L^{2}(Y-\mathcal{Z}_{\tau};S^{\text{Re}}) :=\displaystyle:= {v|Y𝒵τ|v|2r2νdV<}\displaystyle\Big{\{}\ v\ \ \Big{|}\ \ \int_{Y-\mathcal{Z}_{\tau}}{|v|^{2}}\ r^{-2\nu}dV\ <\ \infty\Big{\}} (4.10)

where \nabla denotes the covariant derivative on SReS^{\text{Re}} formed from AτA_{\tau} and the background pair (gτ,Bτ)(g_{\tau},B_{\tau}). These spaces are equipped with the norms given by the (positive) square root integrals required to be finite, and the Hilbert space structures arising from their polarizations.


Aτ\not{D}_{A_{\tau}} extends to a bounded linear operator

Aτ:r1+νHe1(Y𝒵τ;SRe)rνL2(Y𝒵τ;SRe).\not{D}_{A_{\tau}}:r^{1+\nu}H^{1}_{e}(Y-\mathcal{Z}_{\tau};S^{\text{Re}})\longrightarrow r^{\nu}L^{2}(Y-\mathcal{Z}_{\tau};S^{\text{Re}}). (4.11)

The results of [43, Prop 2.4], and [36, Thm 6.1] show the following.

Lemma 4.4.

For 12<ν<12-\tfrac{1}{2}<\nu<\tfrac{1}{2},

  1. (A)

    (4.11) is left semi-Fredholm (i.e. ker(Aτ)\ker(\not{D}_{A_{\tau}}) is finite-dimensional, and Range(Aτ)\text{Range}(\not{D}_{A_{\tau}}) is closed.)

  2. (B)

    There is a constant Cν>0C_{\nu}>0 such that the following “semi-elliptic” estimate holds:

    ur1+νHe1Cν(AτurνL2+πν(u)rνL2) for all ur1+νHe1\|u\|_{r^{1+\nu}H^{1}_{e}}\leq C_{\nu}\Big{(}\|\not{D}_{A_{\tau}}u\|_{r^{\nu}L^{2}}\ +\ \|\pi_{\nu}(u)\|_{r^{\nu}L^{2}}\Big{)}\hskip 42.67912pt\text{ for all }\ \ u\in r^{1+\nu}H^{1}_{e} (4.12)

    where πν\pi_{\nu} is the rνL2r^{\nu}L^{2}-orthogonal projection onto the finite-dimensional kernel.

  3. (C)

    When the assumptions of Theorem 1.6 hold, (4.12) holds uniformly for τ(τ0,τ0)\tau\in(\tau_{0},\tau_{0}) when πν\pi_{\nu} is replaced by the projection to the 1-dimensional eigenspace spanned by Φτ\Phi_{\tau}.

The finite-dimensional kernel for ν=0\nu=0 is, by definition, the set of 2\mathbb{Z}_{2}-harmonic spinors. The upcoming Lemma 4.5 implies that this space is independent of ν\nu in the range 12<ν<12-\tfrac{1}{2}<\nu<\tfrac{1}{2}.

A 2\mathbb{Z}_{2}-harmonic spinor is not necessarily smooth. Notice that the estimate (4.12) assumes a priori that ur1+νHe1u\in r^{1+\nu}H^{1}_{e} and does not imply that an rνL2r^{\nu}L^{2}-solution can be bootstrapped to ur1+νHe1u\in r^{1+\nu}H^{1}_{e}. Thus elliptic bootstrapping in the standard sense fails for Aτ\not{D}_{A_{\tau}}. Consequently, even for ν=0\nu=0, the kernel and cokernel of (4.11) need not coincide, despite the fact that Aτ\not{D}_{A_{\tau}} is formally self-adjoint. This is a general phenomenon for elliptic edge operators; the appropriate substitute for smoothness in this setting is the existence of polyhomogeneous expansions along the singular set 𝒵τ\mathcal{Z}_{\tau} as we now explain.

Fix a choice of Fermi coordinates near 𝒵τ\mathcal{Z}_{\tau} as in Definition 3.9. The results of [36, Sec. 7] imply the following regularity result about 2\mathbb{Z}_{2}-harmonic spinors (see also [26, App. A], [43, Section 3.3], and [18] for more general exposition).

Lemma 4.5.

Suppose that Φr1+νHe1(Y𝒵τ;SRe)\Phi\in r^{1+\nu}H^{1}_{e}(Y-\mathcal{Z}_{\tau};S^{\text{Re}}) is a 2\mathbb{Z}_{2}-harmonic spinor or eigenvector. Then Φ\Phi admits a local polyhomogenous expansion of the following form:

Φ\displaystyle\Phi \displaystyle\sim (c(t)d(t)eiθ)r1/2+n1k=2n2n+1p=0n1(cn,k,p(t)eikθdn,k,p(t)eikθeiθ)rn+1/2(logr)p\displaystyle\begin{pmatrix}c(t)\ \ \ \ \ \\ d(t)e^{-i\theta}\end{pmatrix}r^{1/2}\ +\ \sum_{n\geq 1}\sum_{k=-2n}^{2n+1}\sum_{p=0}^{n-1}\begin{pmatrix}\ \ c_{n,k,p}(t)e^{ik\theta}\ \ \ \ \ \\ \ \ d_{n,k,p}(t)e^{ik\theta}e^{-i\theta}\ \ \end{pmatrix}r^{n+1/2}(\log r)^{p} (4.13)

where c(t),d(t),ck,m,n(t),dk,m,n(t)C(S1;)c(t),d(t),c_{k,m,n}(t),d_{k,m,n}(t)\in C^{\infty}(S^{1};\mathbb{C}). Here, \sim denotes convergence in the following sense: for every NN\in\mathbb{N}, the partial sums

ΦN=nNk=2n2n+1p=0n1(cn,k,p(t)eikθdn,k,p(t)eikθeiθ)rn+1/2(logr)p\Phi_{N}=\sum_{n\leq N}\sum_{k=-2n}^{2n+1}\sum_{p=0}^{n-1}\begin{pmatrix}\ \ c_{n,k,p}(t)e^{ik\theta}\ \ \ \ \ \\ \ \ d_{n,k,p}(t)e^{ik\theta}e^{-i\theta}\ \ \end{pmatrix}r^{n+1/2}(\log r)^{p}

satisfy the pointwise bounds

|ΦΦN|\displaystyle|\Phi-\Phi_{N}| \displaystyle\leq CNrN+1+14|tαβ(ΦΦN)|CN,α,βrN+1+14|β|\displaystyle C_{N}r^{N+1+\tfrac{1}{4}}\hskip 56.9055pt|\nabla_{t}^{\alpha}\nabla^{\beta}(\Phi-\Phi_{N})|\leq C_{N,\alpha,\beta}r^{N+1+\tfrac{1}{4}-|\beta|} (4.14)

for constants CN,α,βC_{N,\alpha,\beta} determined by the background data and choice of local coordinates and trivialization. Here, β\beta is a multi-index of derivatives in the directions normal to 𝒵τ\mathcal{Z}_{\tau}. ∎

Notice that the non-degeneracy condition of Definition 1.1 is equivalent to the statement that the leading coefficients satisfy |c(t)|2+|d(t)|2>0|c(t)|^{2}+|d(t)|^{2}>0 for all t𝒵τt\in\mathcal{Z}_{\tau}. Note also that the existence of the expansion (4.13) implies that the second condition of (1.6) is equivalent to the requirement that |Φ0||\Phi_{0}| extends continuously to YY with 𝒵0=|Φ0|1(0)\mathcal{Z}_{0}=|\Phi_{0}|^{-1}(0) as in Theorem 3.2.

4.3. Polynomial Decay

This section establishes that solution of the Dirac equation in rHe1rH^{1}_{e} obey a decay property towards the singular set. This decay property is the precise manifestation of requirement (II) in the alternating method described in Section 2.2.

Because the projection to the obstruction bundle is a highly non-local operator on YY, cancelling the obstruction using the deformations of the singular set (the horizontal arrow in 2.17) disrupts the property that the error is supported where dχ+0d\chi^{+}\neq 0. The following generalization of support is needed to address this.

Definition 4.6.

A spinor gL2(Y𝒵τ;SRe)g\in L^{2}(Y-\mathcal{Z}_{\tau};S^{\text{Re}}) is said to have εβ\varepsilon^{\beta}-effective support if there is a constant CC such that for all ν(12,0],-\nu\in(-\tfrac{1}{2},0],

grνL2CεγενβgL2.\|g\|_{r^{\nu}L^{2}}\leq C\varepsilon^{-\gamma}\varepsilon^{-{\nu\beta}}\|g\|_{L^{2}}. (4.15)

holds.

The definition (4.10) implies this property holds holds if supp(g)\text{supp}(g) lies in the region where rCεβr\geq C\varepsilon^{\beta}.

The following lemma provides the relevant decay property for solutions of the Dirac equation in the outside region. In the statement, χ,ν\chi^{-},\nu^{-} are in Section 2.

Lemma 4.7.

Suppose that gL2Range(Aτ)g\in L^{2}\cap\text{Range}(\not{D}_{A_{\tau}}) has ε1/2\varepsilon^{1/2}-effective support, and let ψrHe1(Y𝒵τ;SRe)\psi\in rH^{1}_{e}(Y-\mathcal{Z}_{\tau};S^{\text{Re}}) be the unique Φτ\Phi_{\tau}-perpendicular solution to

Aτψ=g.\not{D}_{A_{\tau}}\psi=g.

Then

dχ.ψL2+(1χ)gL2Cε1/12γgL2.\|d\chi^{-}.\psi\|_{L^{2}}\ +\ \|(1-\chi^{-})g\|_{L^{2}}\leq C\varepsilon^{1/12-\gamma}\|g\|_{L^{2}}. (4.16)
Proof.

For the duration of the proof set, ν=ν=12γ\nu=\nu^{-}=\tfrac{1}{2}-\gamma, and χ=χ\chi=\chi^{-}. The elliptic estimate (4.12) from Lemma 4.4 applies to show that

ψr1+νHe1CgrνL2Cεγεν2gL2=Cε14γgL2.\|\psi\|_{r^{1+{\nu}}H^{1}_{e}}\ \leq\ C\|g\|_{r^{\nu}L^{2}}\ \leq\ C\varepsilon^{-\gamma}\varepsilon^{-\tfrac{\nu}{2}}\|g\|_{L^{2}}=C\varepsilon^{-\tfrac{1}{4}-\gamma^{\prime}}\|g\|_{L^{2}}.

Then, since χ\chi is a logarithmic cut-off function with |dχ.ψ|Cr|ψ||d\chi.\psi|\leq\frac{C}{r}|\psi|, and is supported where rε2/3γr\sim\varepsilon^{2/3-\gamma},

dχ.ψL2C(dχ0|ψ|2r2r2νr2νdV)12Cεν(23γ)ψr1+νHe1Cεν(23γ)ε14γgL2\|d\chi.\psi\|_{L^{2}}\leq C\left(\int_{d\chi\neq 0}\frac{|\psi|^{2}}{r^{2}}\frac{r^{2\nu}}{r^{2\nu}}\ dV\right)^{\tfrac{1}{2}}\leq C\varepsilon^{\nu(\tfrac{2}{3}-\gamma)}\|\psi\|_{r^{1+\nu}H^{1}_{e}}\leq C\varepsilon^{\nu(\tfrac{2}{3}-\gamma)}\varepsilon^{-\tfrac{1}{4}-\gamma^{\prime}}\|g\|_{L^{2}}

Relabeling γγ+νγ\gamma\mapsto\gamma^{\prime}+\nu\gamma and combining power of ε\varepsilon gives the bound on the first term of (4.16). Since (1χ)gL2=(1χ)AτψL2ψrHe1(χ1)\|(1-\chi)g\|_{L^{2}}=\|(1-\chi)\not{D}_{A_{\tau}}\psi\|_{L^{2}}\leq\|\psi\|_{rH^{1}_{e}(\chi\neq 1)}, a similar argument covers the second term. ∎

Remark 4.8.

The decay of solutions of ψ=g\not{D}\psi=g away from the support of gg can alternatively be obtained using the decay of the operator’s Schwartz kernel as in [36]. For our purposes, the above approach using weights is easier, and also applies to the inside region, where the properties of the Schwartz kernel of the linearization become quite opaque.

5. The Obstruction Bundle

This section characterizes the infinite-dimensional cokernel of (4.11) for the weight ν=0\nu=0. The cokernel may be canonically identified with the L2L^{2}-solutions of the formal adjoint operator, which for ν=0\nu=0 is Aτ\not{D}_{A_{\tau}} itself (now understood in a weak sense with domain L2L^{2}, see Section 2.2 of [43]).

5.1. The Obstruction Basis

To understand the form of the cokernel it is instructive for consider the model case of Y=S1×2Y^{\circ}=S^{1}\times\mathbb{R}^{2}. Endow YY^{\circ} with coordinates (t,x,y)(t,x,y) and the product metric where S1S^{1} has length 2π2\pi, and set 𝒵=S1×{0}\mathcal{Z}_{\circ}=S^{1}\times\{0\}. Let E=¯E=\underline{\mathbb{H}} be the trivial bundle equipped with the product connection. As in Section 3.2, SRe¯2H¯S^{\text{Re}}\subset\underline{\mathbb{C}}^{2}\otimes\underline{H} is given elements of the form 12(Ψ+σΨ)\tfrac{1}{2}(\Psi+\sigma\Psi). The Dirac operator takes the form (4.6) here with 𝔡1,𝔡0=0\mathfrak{d}_{1},\mathfrak{d}_{0}=0. By Section 3 of [43] or by direct computation, the L2L^{2}-kernel of A0\not{D}_{A_{0}} on SReS^{\text{Re}} is given by the L2L^{2} span of

Ψ:=12(Id+σ)(||eite||rr(eiθsgn()))\Psi_{\ell}^{\circ}:=\frac{1}{2}\left(\text{Id}+\sigma\right)\left({\sqrt{|\ell|}}\frac{e^{i\ell t}e^{-|\ell|r}}{\sqrt{r}}\begin{pmatrix}e^{-i\theta}\\ \text{sgn}(\ell)\end{pmatrix}\right) (5.1)

for \ell\in\mathbb{Z}. Note that any l2l^{2}-linear combination cΨ\sum c_{\ell}\Psi_{\ell}^{\circ} lies in L2L^{2} and is smooth away from 𝒵\mathcal{Z}_{\circ}, but fails to lie in rHe1rH^{1}_{e}, because r1/2L2\nabla r^{-1/2}\notin L^{2}. In this case, coker(Aτ)\text{coker}(\not{D}_{A_{\tau}}) may be identified with L2(𝒵;¯)L^{2}(\mathcal{Z}_{\circ};\underline{\mathbb{C}}) via Fourier series with Ψeit\Psi_{\ell}^{\circ}\mapsto e^{i\ell t}. 222Here, we are glossing over the fact that the =0\ell=0 modes are not in L2L^{2} on the non-compact YY^{\circ}; compactness of YY ameliorates this, see Section 4.3 of [43].

For a general compact manifold YY, the cokernel has a similar characterization given in the upcoming Propositions 5.2 and 5.3. In this case, there is the following caveat: because of the appearance of the 2\mathbb{Z}_{2}-harmonic spinor Φ0\Phi_{0} at τ=0\tau=0, the family of cokernels has a discontinuity at τ=0\tau=0 (the cokernel would jump in dimension here, were it finite-dimensional). Instead, we work with the following “thickening” of the family of cokernels.

Definition 5.1.

The Obstruction Space associated to the data (𝒵τ,gτ,Bτ)(\mathcal{Z}_{\tau},g_{\tau},B_{\tau}) is the (infinite-dimensional) subspace of L2(SRe)L^{2}(S^{\text{Re}}) defined by

𝐎𝐛(𝒵τ)\displaystyle{\bf Ob}(\mathcal{Z}_{\tau}) :=\displaystyle:= Span(ker(Aτ|L2),Φτ)\displaystyle\text{Span}\left(\ker(\not{D}_{A_{\tau}}|_{L^{2}})\ ,\ \Phi_{\tau}\right) (5.2)

This has an L2L^{2}-orthogonal decomposition Ob(𝒵τ)=Ob(𝒵τ)Φτ.\text{\bf Ob}(\mathcal{Z}_{\tau})=\text{\bf Ob}^{\perp}(\mathcal{Z}_{\tau})\oplus\mathbb{R}\Phi_{\tau}.

The Obstruction Bundle is the family of obstruction spaces parameterized by τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0})

Ob:=τOb(𝒵τ)(τ0,τ0).\text{\bf Ob}:=\bigsqcup_{\tau}\text{\bf Ob}(\mathcal{Z}_{\tau})\to(-\tau_{0},\tau_{0}).

The below proposition shows that Ob is a smooth Banach vector bundle, with fiber modeled on a space of spinors on 𝒵0\mathcal{Z}_{0}. For this purpose, let

𝒮τ𝒵τ\mathcal{S}_{\tau}\to\mathcal{Z}_{\tau} (5.3)

be rank 1 complex Clifford module whose fiber is the +i+i eigenspace of γ(dt)\gamma(dt) where dtdt is the normalized, oriented unit tangent vector to 𝒵τ\mathcal{Z}_{\tau} and γ\gamma is Clifford multiplication both defined using the metric gτg_{\tau}.

Proposition 5.2.

There is a family of continuous linear isomorphisms

Ξ:Ob(τ0,τ0)×(L2(𝒵0;𝒮0))\Xi:\text{\bf Ob}\to(-\tau_{0},\tau_{0})\times(L^{2}(\mathcal{Z}_{0};\mathcal{S}_{0})\oplus\mathbb{R})

that endow the obstruction bundle with the structure of a smooth Hilbert vector bundle over (τ0,τ0)(-\tau_{0},\tau_{0}). Moreover, Ξτ1\Xi_{\tau}^{-1} restricted to the \mathbb{R} factor is the inclusion of the span of Φτ\Phi_{\tau}.

Proof.

See Propositions 4.2 and 8.5 in [43]. ∎

The next proposition gives an explicit basis for Ob(𝒵τ)\text{\bf Ob}^{\perp}(\mathcal{Z}_{\tau}) indexed by ×π0(𝒵τ)\mathbb{Z}\times\pi_{0}(\mathcal{Z}_{\tau}), and gives bounds showing that the support of these basis elements increasingly concentrates near 𝒵τ\mathcal{Z}_{\tau} as the first index increases, just as in the model case of YY^{\circ}.

For the statement of the proposition, let 𝒮τ\mathcal{S}_{\tau} be as defined preceding Proposition 5.2). For simplicity, we state the proposition in the case that 𝒵τ\mathcal{Z}_{\tau} has a single component; the general case is the obvious extension with an additional index α\alpha ranging over π0(𝒵τ)\pi_{0}(\mathcal{Z}_{\tau}). A choice of arclength coordinate tt on 𝒵τ\mathcal{Z}_{\tau} induces a trivialization of 𝒵τ\mathcal{Z}_{\tau}, and there is a Dirac operator on 𝒮τ\mathcal{S}_{\tau} given by iti\partial_{t} in this trivialization. Let ϕ\phi_{\ell} denote the eigenfunctions of this Dirac operator, which are associated with eite^{i\ell t} for 2π/|𝒵τ|\ell\in 2\pi\mathbb{Z}/|\mathcal{Z}_{\tau}| in the trivialization.

Proposition 5.3.

([43, Prop 4.3] For τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}), there is a bounded linear isomorphism

obτι:L2(𝒵τ;𝒮τ)Ob(𝒵τ)\text{ob}_{\tau}\oplus\iota:L^{2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau})\oplus\mathbb{R}\longrightarrow\text{\bf Ob}(\mathcal{Z}_{\tau}) (5.4)

and a basis of Ob(𝒵τ)\text{\bf Ob}^{\perp}(\mathcal{Z}_{\tau}) which satisfy the following properties.

  1. (A)

    When Aτ\not{D}_{A_{\tau}} is complex linear, there is a complex basis Ψ\Psi_{\ell} of Ob(𝒵τ)\text{\bf Ob}^{\perp}(\mathcal{Z}_{\tau}) indexed by 2π/|𝒵τ|\ell\in 2\pi\mathbb{Z}/|\mathcal{Z}_{\tau}| such that the Ob(𝒵τ)\text{\bf Ob}(\mathcal{Z}_{\tau})-component of a spinor ψL2\psi\in L^{2} under (5.4) is given by

    obτ1(ΠObψ)=ψ,Ψϕ.ι1(ΠObψ)=ψ,ΦτΦτ.\text{ob}_{\tau}^{-1}(\Pi^{\text{Ob}}\psi)=\sum_{\ell}\langle\psi,\Psi_{\ell}\rangle_{{\mathbb{C}}}\phi_{\ell}.\hskip 85.35826pt\iota^{-1}(\Pi^{\text{Ob}}\psi)=\langle\psi,\Phi_{\tau}\rangle\Phi_{\tau}. (5.5)

    where ,\langle-,-\rangle_{\mathbb{C}} is the hermitian inner product. Moreover,

    Ψ=χΨ+ζ+ξ\Psi_{\ell}=\chi\Psi^{\circ}_{\ell}+\zeta_{\ell}+\xi_{\ell}

    where

    • Ψ\Psi_{\ell}^{\circ} are the L2L^{2}-orthonormalized Euclidean obstruction elements (5.1) and χ\chi is a cutoff function supported on a tubular neighborhood of 𝒵τ\mathcal{Z}_{\tau}.

    • ζ\zeta_{\ell} is a perturbation with L2L^{2}-norm O(1||)O(\tfrac{1}{|\ell|}) which decays exponentially away from 𝒵0\mathcal{Z}_{0} with exponent ec||e^{-c|\ell|}.

    • ξ\xi_{\ell} is a perturbation of L2L^{2}-norm O(1||N)O(\tfrac{1}{|\ell|^{N}}) for any N2N\geq 2.

  2. (B)

    In the case that A0\not{D}_{A_{0}} is only \mathbb{R}-linear, then there is a real basis

    Ψre=χΨ+ζre+ξreΨim=i(χΨ)+ζim+ξim\Psi_{\ell}^{\text{re}}=\chi\Psi^{\circ}_{\ell}+\zeta^{\text{re}}_{\ell}+\xi^{\text{re}}_{\ell}\hskip 56.9055pt\Psi_{\ell}^{\text{im}}=i(\chi\Psi^{\circ}_{\ell})+\zeta^{\text{im}}_{\ell}+\xi^{\text{im}}_{\ell}

    satisfying identical bounds where (5.5) uses the inner product ψ,Ψ=ψ,Ψre,+iψ,Ψim,.\langle\psi,\Psi_{\ell}\rangle_{\mathbb{C}}=\langle\psi,\Psi_{\ell}^{\text{re},{}}\rangle\ +\ i\langle\psi,\Psi_{\ell}^{\text{im},{}}\rangle.

A more precise meaning of the second and third bullet points is given in Proposition 4.3 of [43], but is not needed for our purposes here.

As a consequence of the concentration of the cokernel modes around 𝒵τ\mathcal{Z}_{\tau} as |||\ell|\to\infty, the obstruction space displays a rather different type of regularity than the Sobolev regularity of the spinors on Y𝒵τY-\mathcal{Z}_{\tau} (see Section 6.1 of [43] for more discussion). Regularity of a spinor Πτ(Ψ)Ob(𝒵τ)\Pi_{\tau}(\Psi)\in\text{\bf Ob}(\mathcal{Z}_{\tau}), means that obτ1Πτ(Ψ)Ls,2(𝒵τ;𝒮τ)L2(𝒵τ;𝒮τ)\text{ob}_{\tau}^{-1}\circ\Pi_{\tau}(\Psi)\in L^{s,2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau})\subseteq L^{2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau}) for some s>0s>0. By Proposition 5.3, this is a question of how fast the sequence of inner products (5.5) decays as |||\ell|\to\infty. Because of the concentration, latter depends both on the Sobolev regularity of Ψ\Psi and its rate of decay toward 𝒵τ\mathcal{Z}_{\tau}; in particular if a spinor Ψ\Psi is compactly supported in 𝒵τ\mathcal{Z}_{\tau}’s complement, its obstruction components enjoy high regularity, even if all its covariant derivatives fail to be integrable on Y𝒵τY-\mathcal{Z}_{\tau} in the normal sense. This improved regularity is crucial in Sections 911, where it is applied to the error terms in the alternating iteration which are compactly supported away from 𝒵τ\mathcal{Z}_{\tau} in the neck region.

The next lemma makes this improvement of regularity precise. Given L0L_{0}\in\mathbb{N}, let

πL0(ϕ)={ϕ||L00||>L0{\pi}_{L_{0}}(\phi_{\ell})=\begin{cases}\phi_{\ell}\hskip 28.45274pt|\ell|\leq L_{0}\\ 0\ \ \hskip 27.03003pt|\ell|>L_{0}\end{cases}

denote the Fourier projector to modes less than L0L_{0} in L2(𝒵τ;𝒮τ)L^{2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau}).

Lemma 5.4.

For any NN\in\mathbb{N}, there are CN,RN>0C_{N},R_{N}>0 such that the following is satisfied. If ΨL2(Y𝒵τ;SRe)\Psi\in L^{2}(Y-\mathcal{Z}_{\tau};S^{\text{Re}}) is a spinor and R<RNR<R_{N} is such that dist(supp(Ψ),𝒵τ)R\text{dist}(\text{supp}(\Psi),\mathcal{Z}_{\tau})\geq R. Then for any 0<γ<<10<\gamma<<1,

(1πL0)obτ1Πτ(Ψ)L2CN(L0)NΨL2\|(1-\pi_{L_{0}})\circ\text{ob}^{-1}_{\tau}\circ\Pi_{\tau}(\Psi)\|_{L^{2}}\leq\frac{C_{N}}{(L_{0})^{N}}\|\Psi\|_{L^{2}}

holds for any L0>R11γL_{0}>R^{-\tfrac{1}{1-\gamma}}. In particular, obτ1Πτ(Ψ)C(𝒵τ;𝒮τ)\text{ob}_{\tau}^{-1}\circ\Pi_{\tau}(\Psi)\in C^{\infty}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau}).

Proof.

By Proposition 5.3, the projection (1πlow)obτ1Πτ(Ψ)(1-\pi^{\text{low}})\circ\text{ob}_{\tau}^{-1}\circ\Pi_{\tau}(\Psi) is given by

||>L0Ψ,eϕ.\sum_{|\ell|>L_{0}}\langle\Psi_{\ell},e\rangle_{\mathbb{C}}\phi_{\ell}. (5.6)

The result then follows straightforwardly from using the decomposition Ψ=χΨ+ζ+ξ\Psi_{\ell}=\chi\Psi^{\circ}_{\ell}+\zeta_{\ell}+\xi_{\ell}. More specifically, using the first two bullet points of Item (A) in Proposition 5.3, the first two terms contribute O(L0Exp(L0R))O(L0Exp(L0γ))O(L_{0}\text{Exp}(-L_{0}R))\leq O(L_{0}\text{Exp}(-L_{0}^{\gamma})) since the assumption on L0L_{0} implies L0R>L0γL_{0}R>L_{0}^{\gamma}. For RNR_{N} sufficiently small, this dominates L0NL_{0}^{-N}. Using Cauchy-Schwartz on ξ\xi_{\ell} and the third bullet point of Item (A) with N=N+2N^{\prime}=N+2 then gives the desired bound. Applying the estimate repeatedly for L0=2kL_{0}=2^{k} shows that obτ1Πτ(Ψ)LN,2(𝒵τ;𝒮τ)\text{ob}^{-1}_{\tau}\circ\Pi_{\tau}(\Psi)\in L^{N,2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau}) for every NN. ∎

5.2. The Surjective Weights

By the general theory of elliptic edge operators ([36, Thm 6.1]), a first order elliptic edge operator LL

L:r1+νHe1rνL2L:r^{1+\nu}H^{1}_{e}\to r^{\nu}L^{2}

is semi-Fredholm provided that the weight ν\nu lies outside the discrete set I(L)I(L) of indicial roots. More specifically, there are critical weights ν¯<ν¯,\underline{\nu}<\overline{\nu}, such that (i) LL is left semi-Fredholm (finite-dimensional kernel and closed range) for ν¯<νI(L)\overline{\nu}<\nu\notin I(L), and (ii) is right semi-Fredholm (finite-dimensional cokernel) for ν¯>νI(L)\underline{\nu}>\nu\notin I(L). For the singular Dirac operator Aτ\not{D}_{A_{\tau}}, the critical weights (by [28, Prop 3.9]) are:

I(rAτ)=+12ν¯=ν¯=12.I(r\not{D}_{A_{\tau}})=\mathbb{Z}+\tfrac{1}{2}\hskip 71.13188pt\overline{\nu}=\underline{\nu}=-\tfrac{1}{2}. (5.7)

Thus when the weight ν\nu decreases past the critical weight 12-\tfrac{1}{2}, Aτ\not{D}_{A_{\tau}} flips from being left semi-Fredholm to being right semi-Fredholm.

Lemma 4.4 shows (4.11) is left semi-Fredholm for ν(12,12)\nu\in(-\tfrac{1}{2},\tfrac{1}{2}). The next lemma shows it is right semi-Frehdolm in the specific case ν=1\nu=-1.

Lemma 5.5.

For ν=1\nu=-1,

Aτ:He1(Y𝒵τ;SRe)r1L2(Y𝒵τ;SRe)\not{D}_{A_{\tau}}:H^{1}_{e}(Y-\mathcal{Z}_{\tau};S^{\text{Re}})\longrightarrow r^{-1}L^{2}(Y-\mathcal{Z}_{\tau};S^{\text{Re}}) (5.8)

has a finite-dimensional cokernel. Moreover, for uL2ker(Aτ|He1)u\perp_{L^{2}}\ker(\not{D}_{A_{\tau}}|_{H^{1}_{e}}), the elliptic estimate

uHe1CAτur1L2.\|u\|_{H^{1}_{e}}\leq C\|\not{D}_{A_{\tau}}u\|_{r^{-1}L^{2}}.

holds uniformly in τ\tau.

Proof.

See [36, Thm 6.1] and [28, Thm 3.18]. ∎

Corollary 5.6.

There is a closed subspace 𝒳τHe1\mathcal{X}_{\tau}\subseteq H^{1}_{e} such that

Aτ:𝒳τOb(𝒵τ)\not{D}_{A_{\tau}}:\mathcal{X}_{\tau}\to\text{\bf Ob}^{\perp}(\mathcal{Z}_{\tau})

is an isomorphism. In particular, there is a C>0C>0 such that if ΨOb(𝒵τ)\Psi\in\text{\bf Ob}^{\perp}(\mathcal{Z}_{\tau}), then there exists a unique uΨ𝒳τu_{\Psi}\in\mathcal{X}_{\tau} satisfying

AτuΨ=Ψ,uΨHe1CΨL2,\not{D}_{A_{\tau}}u_{\Psi}=\Psi,\hskip 56.9055pt\|u_{\Psi}\|_{H^{1}_{e}}\leq C\|\Psi\|_{L^{2}},

where CC is uniform in τ\tau.

Proof.

Self-adjointness implies that the cokernel of (5.8) is spanned by the 2\mathbb{Z}_{2}-harmonic spinors, which are orthogonal to Ob(𝒵τ)\text{\bf Ob}^{\perp}(\mathcal{Z}_{\tau}) by definition. The conclusion then follows directly from Lemma 5.5. ∎

Corollary 5.6 shows that there is a subspace of He1H^{1}_{e} consisting of singular spinors which hit the infinite-dimensional obstruction of (4.11). The next point is important, though. The spinors in this subspace of He1H^{1}_{e} cannot be used to cancel the obstruction components in the alternating iteration. Since these spinors grow across the overlap region, rather than decay, an alternating iteration scheme employing them will not satisfy requirement (II) of Section 3.2, and will therefore not converge.

Remark 5.7.

In fact, [36, Thm. 7.14] describes the form of uΨu_{\Psi} more precisely. That result implies that the maximal domain 𝒟={uL2|AτuL2}\mathcal{D}=\{u\in L^{2}\ |\ \not{D}_{A_{\tau}}u\in L^{2}\} consists of spinors of the form

𝒟={u=(c(t)zd(t)z¯)eiθ/2+v|vrHe1}.\mathcal{D}=\left\{u=\begin{pmatrix}\tfrac{c(t)}{\sqrt{z}}\\ \tfrac{d(t)}{\sqrt{\overline{z}}}\end{pmatrix}e^{-i\theta/2}\ +\ v\ \ \Big{|}\ \ v\in rH^{1}_{e}\right\}.

and Aτ:𝒟L2\not{D}_{A_{\tau}}:\mathcal{D}\to L^{2} is right semi-Fredholm. Proposition 5.3 implies that Ob(𝒵τ)\text{\bf Ob}^{\perp}(\mathcal{Z}_{\tau}) consists of L2L^{2}-spinors whose leading coefficients (modulo a compact operator) lie in the subspace where d(t)=iH(c(t))L1/2,2d(t)=iH(c(t))\in L^{-1/2,2}. The spinors uΨu_{\Psi}, modulo rHe1rH^{1}_{e} and compact operator, fill the subspace where d(t)=iH(c(t))d(t)=-iH(c(t)).

6. Deformations of Singular Sets

This section extends the theory of the singular Dirac operator to the case where the singular set varies. We begin by defining the universal Dirac operator \not{\mathbb{D}} as in (1.10), which is the infinite-dimensional family of singular Dirac operators parameterized by embedded singular sets. We then calculate the derivative d\text{d}\not{\mathbb{D}} with respect to deformations of the singular set and give a more precise version of Theorem 1.2.

6.1. The Universal Dirac Operator

Let Emb2,2(S1;Y)\text{Emb}^{2,2}(\bigsqcup S^{1};Y) denote the Banach manifold of embedded singular sets with Sobolev regularity (2,2)(2,2). The tangent space TEmb2,2(S1;Y)T\text{Emb}^{2,2}(\bigsqcup S^{1};Y) at a singular set 𝒵τ\mathcal{Z}_{\tau} is naturally identified with the space L2,2(𝒵τ;N𝒵τ)L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) of sections of the normal bundle N𝒵τN\mathcal{Z}_{\tau}. To begin, we construct families of charts around 𝒵τEmb2,2(S1;Y)\mathcal{Z}_{\tau}\in\text{Emb}^{2,2}(\bigsqcup S^{1};Y) using families of diffeomorphisms deforming 𝒵τ\mathcal{Z}_{\tau}.

Choose a family of diffeomorphisms

𝔽τ:L2,2(𝒵τ;N𝒵τ)\displaystyle\mathbb{F}_{\tau}:L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) \displaystyle\to Diff(Y)\displaystyle\text{Diff}(Y) (6.1)
η\displaystyle\eta \displaystyle\mapsto Fη:YY\displaystyle F_{\eta}:Y\to Y (6.2)

that associates to each ηL2,2(𝒵τ;N𝒵τ)\eta\in L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) a diffeomorphism FηF_{\eta} with F0=IdF_{0}=\text{Id} and such that dds|s=0Fsη\frac{d}{ds}\big{|}_{s=0}F_{s\eta} is a vector field extending η\eta smoothly to YY. We also assume that (6.1) depends smoothly on τ\tau. Then set

Expτ:L2,2(𝒵τ;N𝒵τ)\displaystyle\text{Exp}_{\tau}:L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) \displaystyle\to Emb2,2(𝒵τ;Y)\displaystyle\text{Emb}^{2,2}(\mathcal{Z}_{\tau};Y) (6.3)
η\displaystyle\eta \displaystyle\mapsto 𝒵η,τ:=Fη(𝒵τ).\displaystyle\mathcal{Z}_{\eta,\tau}:=F_{\eta}(\mathcal{Z}_{\tau}). (6.4)

Expτ\text{Exp}_{\tau} is a local diffeomorphism on a neighborhood of 0 by the Inverse Function Theorem (cf [43, Lem 5.3]). Fix ρ0\rho_{0} small enough that Expτ\text{Exp}_{\tau} is a diffeomorphism on the ball of radius ρ0\rho_{0} for for every τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}). In a slight abuse of notation, we use τ\mathcal{E}_{\tau} to denote both this ball and its image under Expτ\text{Exp}_{\tau}. We will use these charts with one choice of the family (6.1) in (6.12) and another more sophisticated choice in (6.21).

Each 𝒵τ\mathcal{Z}\in\mathcal{E}_{\tau} defines

  1. (1)

    A flat connection A𝒵A_{\mathcal{Z}} on Y𝒵Y-\mathcal{Z} with holonomy in 2\mathbb{Z}_{2} homotopic to that of AτA_{\tau}.

  2. (2)

    A bundle S𝒵ReS^{\text{Re}}_{\mathcal{Z}} and Dirac operator 𝒵:Γ(S𝒵Re)Γ(S𝒵Re)\not{D}_{\mathcal{Z}}:\Gamma(S^{\text{Re}}_{\mathcal{Z}})\to\Gamma(S^{\text{Re}}_{\mathcal{Z}}) as in Lemma 3.5 and (3.8).

  3. (3)

    Hilbert spaces rHe1(Y𝒵;S𝒵Re)rH^{1}_{e}(Y-\mathcal{Z};S^{\text{Re}}_{\mathcal{Z}}), and L2(Y𝒵;S𝒵Re)L^{2}(Y-\mathcal{Z};S^{\text{Re}}_{\mathcal{Z}}) as in Definition 4.3.

Note each of these depends implicitly on τ\tau; in particular the weights used in the norms are defined analogously to (4.8) using the geodesic distance of gτg_{\tau}.

Using these, define families of Hilbert spaces p1:e1(τ)τp_{1}:\mathbb{H}^{1}_{e}(\mathcal{E}_{\tau})\to\mathcal{E}_{\tau} and p0:𝕃2(τ)τp_{0}:\mathbb{L}^{2}(\mathcal{E}_{\tau})\to\mathcal{E}_{\tau} by

e1(τ)\displaystyle\mathbb{H}^{1}_{e}(\mathcal{E}_{\tau}) :=\displaystyle:= {(𝒵,u)|𝒵τ,urHe1(Y𝒵;S𝒵Re)}\displaystyle\{(\mathcal{Z},u)\ |\ \mathcal{Z}\in\mathcal{E}_{\tau}\ ,\ u\in rH^{1}_{e}(Y-\mathcal{Z};S^{\text{Re}}_{\mathcal{Z}})\}
𝕃2(τ)\displaystyle\mathbb{L}^{2}(\mathcal{E}_{\tau}) :=\displaystyle:= {(𝒵,v)|𝒵τ,vL2(Y𝒵;S𝒵Re)}\displaystyle\{(\mathcal{Z},v)\ |\ \mathcal{Z}\in\mathcal{E}_{\tau}\ ,\ v\in L^{2}(Y-\mathcal{Z};S^{\text{Re}}_{\mathcal{Z}})\}

with the p0,p1p_{0},p_{1} the obvious projections.

Lemma 6.1.

Each choice of a family 𝔽τ\mathbb{F}_{\tau} as in (6.1) determines trivializations

Υ:e1(τ)\displaystyle\Upsilon:\mathbb{H}^{1}_{e}(\mathcal{E}_{\tau}) \displaystyle\simeq τ×rHe1(Y𝒵τ;SRe)\displaystyle\mathcal{E}_{\tau}\times rH^{1}_{e}(Y-\mathcal{Z}_{\tau};S^{\text{Re}})
Υ:𝕃2(τ)\displaystyle\Upsilon:\mathbb{L}^{2}(\mathcal{E}_{\tau}) \displaystyle\simeq τ×L2(Y𝒵τ;SRe).\displaystyle\mathcal{E}_{\tau}\times\ L^{2}(Y-\mathcal{Z}_{\tau};S^{\text{Re}}).

Together, these trivializations endow the spaces on the left with the structure of locally trivial Hilbert vector bundles over τ\mathcal{E}_{\tau}.

Proof.

This is Lemma 5.1 in [43]. The proof is nearly identical to the proof of Lemma 8.2 in Section 8, which applies to the full Seiberg–Witten equations. ∎

The following definition gives a more precise meaning to the operator defined in (1.10).

Definition 6.2.

The Universal Dirac Operator is the section \not{\mathbb{D}} defined by

e1(τ).\mathbb{H}^{1}_{e}(\mathcal{E}_{\tau}).p1𝕃2(τ)p_{1}^{*}\mathbb{L}^{2}(\mathcal{E}_{\tau})(𝒵,u):=𝒵u\not{\mathbb{D}}(\mathcal{Z},u):=\not{D}_{\mathcal{Z}}u

The next proposition calculates the linearization of the universal Dirac operator

d(𝒵τ,Φτ):L2,2(𝒵τ;N𝒵τ)rHe1(Y𝒵τ;SRe)L2(Y𝒵τ;SRe)\text{d}\not{\mathbb{D}}_{(\mathcal{Z}_{\tau},\Phi_{\tau})}:L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau})\oplus rH^{1}_{e}(Y-\mathcal{Z}_{\tau};S^{\text{Re}})\ \longrightarrow\ L^{2}(Y-\mathcal{Z}_{\tau};S^{\text{Re}}) (6.5)

at the pair (𝒵τ,Φτ)(\mathcal{Z}_{\tau},\Phi_{\tau}) in the trivialization of Lemma 6.1. Since \not{\mathbb{D}} is linear on fibers, differentiation in the fiber directions is trivial. As explained in Section 1.6, the naturality of the Dirac operator with respect to diffeomorphisms allows us to recast differentiation with respect to the embedding as differentiation with respect to the metrics obtained via pullback by the diffeomorphisms FηF_{\eta}. For each fixed τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}) and s[1,1]s\in[-1,1], set

gsη:=Fsηgτg_{s\eta}:=F_{s\eta}^{*}g_{\tau}

and denote the derivative by g˙η=dds|s=0gsη\dot{g}_{\eta}=\frac{d}{ds}\Big{|}_{s=0}g_{s\eta}. There is again an implicit dependence on τ\tau.

Proposition 6.3.

The linearization (6.5) is given by

d(𝒵τ,Φτ)(η,ψ)\displaystyle\text{d}\not{\mathbb{D}}_{(\mathcal{Z}_{\tau},\Phi_{\tau})}(\eta,\psi) =\displaystyle= Φτ(η)+Aτψ\displaystyle\mathcal{B}_{\Phi_{\tau}}(\eta)\ +\ \not{D}_{A_{\tau}}\psi (6.6)

where

Φτ(η)=(12ijg˙η(ei,ej)ei.jτ+12dTrgτ(g˙η).+12divgτ(g˙η).+(Bτ,η).)Φτ\mathcal{B}_{\Phi_{\tau}}(\eta)=\left(-\frac{1}{2}\sum_{ij}\dot{g}_{\eta}(e_{i},e_{j})e^{i}.\nabla^{\tau}_{j}+\frac{1}{2}d\text{Tr}_{g_{\tau}}(\dot{g}_{\eta}).+\frac{1}{2}\text{div}_{g_{\tau}}(\dot{g}_{\eta}).+\mathcal{R}(B_{\tau},\eta).\right)\Phi_{\tau}\ \ (6.7)

in which (Bτ,η)\mathcal{R}(B_{\tau},\eta) is a smooth term that involves first derivatives of BτB_{\tau} and is zeroth order in η\eta, Clifford multiplication . is that of the metric gτg_{\tau}, and τ\nabla^{\tau} denotes the perturbation to the spin connection formed from (gτ,Bτ)(g_{\tau},B_{\tau}).

Proof Sketch..

(See [43, Section 5.2] for complete details). The derivative with respect to ψ\psi is immediate, becuase \not{\mathbb{D}} is linear in ψ\psi. Differentiating with respect to the family of metrics gηg_{\eta} requires associating the spinor bundles for distinct metrics. This is done as follows. For a fixed η\eta, let

X=(Y×[0,1],gsη+ds2)X=(Y\times[0,1],g_{s\eta}+ds^{2}) (6.8)

denote the metric cylinder on YY. For s=0s=0, the positive spinor bundle W+XW^{+}\to X is isomorphic to the spinor bundle of YY with the metric g0=gτg_{0}=g_{\tau}, while for s=1s=1 it is isomorphic to that with the metric gηg_{\eta}. Let 𝔗gτgη\mathfrak{T}_{g_{\tau}}^{g_{\eta}} denote the isomorphism between the two spinor bundles for gτ,gηg_{\tau},g_{\eta} defined by parallel transport along rays {y}×[0,1]\{y\}\times[0,1] using the spin connection on W+W^{+} (perturbed by Bs=Fsη(Bτ)B_{s}=F_{s\eta}^{*}(B_{\tau})).

The partial derivative with respect to η\eta is then given by

d(𝒵τ,Φτ)(η,0)=(dds|s=0𝔗gτgsηgsη(𝔗gτgsη)1)Φτ\text{d}\not{\mathbb{D}}_{(\mathcal{Z}_{\tau},\Phi_{\tau})}(\eta,0)=\left(\frac{d}{ds}\Big{|}_{s=0}\mathfrak{T}_{g_{\tau}}^{g_{s\eta}}\circ\not{D}_{g_{s\eta}}\circ(\mathfrak{T}_{g_{\tau}}^{g_{s\eta}})^{-1}\right)\Phi_{\tau} (6.9)

where gsη\not{D}_{g_{s\eta}} denotes the singular Dirac operator using the metric gsηg_{s\eta} and twisted around the fixed singular set 𝒵τ\mathcal{Z}_{\tau}. Minus a few additional details regarding pulling back the perturbation BτB_{\tau}, which gives rise to the term (Bτ,η)\mathcal{R}(B_{\tau},\eta), the proposition is now completed by the below theorem of Bourguignon-Gauduchon, which calculates (6.9). ∎

Generalizing the above situation slightly, let gsg_{s} be a family of metrics on YY, and denote by 𝔗g0gs\mathfrak{T}_{g_{0}}^{g_{s}} the association of spinor bundles in the proof of Proposition 6.3 by parallel transport on the cylinder (6.8).

Theorem 6.4.

(Bourguignon-Gauduchon [6]) The derivative of the Dirac operator with respect to the family of metrics gsg_{s} at s=0s=0 acting on a spinor Ψ\Psi is given by

(dds|s=0𝔗g0gsgs(𝔗g0gs)1)Ψ=12ijg˙s(ei,ej)ei.jg0Ψ+12dTrg0(g˙s).Ψ+12divg0(g˙s).Ψ\left(\frac{d}{ds}\Big{|}_{s=0}\mathfrak{T}_{g_{0}}^{g_{s}}\circ\not{D}_{g_{s}}\circ(\mathfrak{T}_{g_{0}}^{g_{s}})^{-1}\right)\Psi=-\frac{1}{2}\sum_{ij}\dot{g}_{s}(e_{i},e_{j})e^{i}.\nabla^{g_{0}}_{j}\Psi+\frac{1}{2}d\text{Tr}_{g_{0}}(\dot{g}_{s}).\Psi+\frac{1}{2}\text{div}_{g_{0}}(\dot{g}_{s}).\Psi (6.10)

where gs\not{D}_{g_{s}} is the Dirac operator of the metric gsg_{s}, and ei,ei,.,divg0,g0e_{i},e^{i},\ .\ ,\text{div}_{g_{0}},\nabla^{g_{0}} are respectively an orthonormal frame and co-frame, Clifford multiplication, the divergence of a symmetric tensor, and the spin connection of the metric g0g_{0}. ∎

In (6.10), the first term arises from differentiating the symbol/Clifford multiplication of the Dirac operator, while the last two terms arise from differentiating the spin connection.

6.2. The Deformation Operator

This subsection calculates the projection of the derivative (6.7) to the obstruction bundle.

Let πτ=Φτ,Φτ\pi_{\tau}=\langle\Phi_{\tau},-\rangle\Phi_{\tau} be the L2L^{2}-orthogonal projection onto the span of the eigenvector Φτ\Phi_{\tau}. Using the orthogonal splitting L2(Y𝒵τ)Ob(𝒵τ)RangeτL^{2}(Y-\mathcal{Z}_{\tau})\simeq\text{\bf Ob}(\mathcal{Z}_{\tau})\oplus\text{Range}^{\perp}_{\tau} where Rangeτ={ψRange(Aτ)|πτ(ψ)=0}\text{Range}^{\perp}_{\tau}=\{\psi\in\text{Range}(\not{D}_{A_{\tau}})\ |\ \pi_{\tau}(\psi)=0\}, the derivative (6.6) can be written as a block matrix:

d(𝒵τ,Φτ)=(ΠObΦτΛ(τ)πτ(1ΠOb)ΦτAτ):L2,2(𝒵τ;N𝒵τ)rHe1Ob(𝒵0)Rangeτ.\text{d}\not{\mathbb{D}}_{(\mathcal{Z}_{\tau},\Phi_{\tau})}=\begin{pmatrix}\Pi^{\text{Ob}}\mathcal{B}_{\Phi_{\tau}}&\Lambda(\tau)\pi_{\tau}\\ \\ (1-\Pi^{\text{Ob}})\mathcal{B}_{\Phi_{\tau}}&\not{D}_{A_{\tau}}\end{pmatrix}\ :\ \begin{matrix}L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau})\\ \oplus\\ rH^{1}_{e}\end{matrix}\ \ \longrightarrow\ \ \begin{matrix}\text{\bf Ob}(\mathcal{Z}_{0})\\ \oplus\\ \text{Range}_{\tau}^{\perp}.\end{matrix} (6.11)

Composing with the isomorphism obτ1ι:Ob(𝒵τ)L2(𝒵τ;𝒮τ)\text{ob}^{-1}_{\tau}\oplus\iota:\text{\bf Ob}(\mathcal{Z}_{\tau})\to L^{2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau})\oplus\mathbb{R} from Proposition 5.3 where 𝒮τ\mathcal{S}_{\tau} is as in (5.3), the top left block of (6.11) can be written as (TΦτ,πτ)(T_{\Phi_{\tau}},\pi_{\tau}) where TΦτT_{\Phi_{\tau}} is the composition:

L2,2(𝒵τ;N𝒵τ)L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau})Ob(𝒵τ)\text{\bf Ob}(\mathcal{Z}_{\tau})L2(𝒵τ;𝒮τ).L^{2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau}).ΠObΦτ\Pi^{\text{Ob}}\mathcal{B}_{\Phi_{\tau}}obτ1\text{ob}_{\tau}^{-1}TΦτT_{\Phi_{\tau}}

In particular, TΦτT_{\Phi_{\tau}} is an operator on sections of vector bundles over the fixed curve 𝒵τ\mathcal{Z}_{\tau}.

The form of TΦτT_{\Phi_{\tau}} depends on the selection of the family of diffeomorphisms 𝔽τ\mathbb{F}_{\tau}. This is effectively a choice of gauge on the bundle e1\mathbb{H}^{1}_{e}, as the trivializations of Lemma 6.1 also depend on 𝔽τ\mathbb{F}_{\tau} (cf Remark 6.11). One natural choice is as follows. Fix a system of Fermi coordinates as in Definition 3.9 with radius r0r_{0}, and choose a cut-off function χ:[0,)[0,1]\chi:[0,\infty)\to[0,1] equal to 1 for rr0/2r\leq r_{0}/2 and vanishing for rr0r\geq r_{0}. Then define a family of diffeomorphisms (6.1) by setting

Fη(t,z):=(t,z+χ(r)η(t)).F_{\eta}\left(t,z\right):=(t,z+\chi(r)\eta(t)). (6.12)

and extending FηF_{\eta} by the identity outside of rr0r\geq r_{0}. This is a diffeomorphism provided η2,2\|\eta\|_{2,2} is sufficiently small (see [43, Lem 5.3]).

The following theorem combines the statements of Theorem 6.1 and Lemma 6.7 from [43]. It is a refined version of Theorem 1.2 in the introduction.

Theorem 6.5.

([43]). For the family of diffeomorphisms (6.12), the operator TΦτT_{\Phi_{\tau}} is an elliptic pseudo-differential operator of order 12\tfrac{1}{2}, and its Fredholm extension

TΦτ:L1/2,2(𝒵τ;N𝒵τ)L2(𝒵τ;𝒮τ)T_{\Phi_{\tau}}:L^{1/2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau})\longrightarrow L^{2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau}) (6.13)

has index 0. Specifically, on sections ηC(𝒵τ;N𝒵τ)\eta\in C^{\infty}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}), it is given by the formula

TΦτ(η(t))=(3|𝒵τ|2(Δ+1)34𝒯Φτddt2)η(t)+Kη(t)T_{\Phi_{\tau}}(\eta(t))=\left(-\tfrac{3|\mathcal{Z}_{\tau}|}{2}(\Delta+1)^{-\tfrac{3}{4}}\ \circ\ \mathcal{T}_{\Phi_{\tau}}\circ\frac{d}{dt^{2}}\right)\eta(t)+K\eta(t) (6.14)

where |𝒵τ||\mathcal{Z}_{\tau}| is the length of 𝒵τ\mathcal{Z}_{\tau}, Δ\Delta is the Laplacian on 𝒮τ\mathcal{S}_{\tau}, ddt2\frac{d}{dt^{2}} is the second covariant derivative on Γ(N𝒵τ)\Gamma(N\mathcal{Z}_{\tau}) induced by the Levi-Civita connection of gτg_{\tau}, KK is a pseudo-differential operator of order 38\tfrac{3}{8} (hence compact), and 𝒯Φτ\mathcal{T}_{\Phi_{\tau}} is the zeroth order operator given by

𝒯Φτ(s(t))=H(cτ(t)s(t))s¯(t)dτ(t)\mathcal{T}_{\Phi_{\tau}}(s(t))=H(c_{\tau}(t)s(t))-\overline{s}(t)d_{\tau}(t)

where iH-iH is the Hilbert transform in the trivialization of 𝒮τ\mathcal{S}_{\tau} induced by the arclength parameterization, and cτ,dτc_{\tau},d_{\tau} are the leading coefficients of Φτ\Phi_{\tau} from the expansion (4.13). In particular, 𝒯Φτ\mathcal{T}_{\Phi_{\tau}} depends smoothly on τ\tau. ∎

The unobstructed condition in Definition 1.3 can be restated in terms of the operator TΦτT_{\Phi_{\tau}}.

Corollary 6.6.

If 2\mathbb{Z}_{2}-harmonic spinor (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) has unobstructed deformations, then TΦτT_{\Phi_{\tau}} is invertible for τ\tau sufficiently small.

Proof.

Definition 1.3 means that (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) is unobstructed if TΦ0T_{\Phi_{0}} is injective. Since it is index 0, (6.13) is invertible. It follows (by Lemma 8.17 of [43]) that TΦτT_{\Phi_{\tau}} is invertible for τ\tau sufficiently small. ∎

As a consequence of Theorem 6.5, the following version of standard elliptic estimates hold. They are proved by repeated differentiation (or integration by parts for m<2m<2).

Corollary 6.7.

For any m0m\geq 0, the extension

TΦτ:Lm+1/2,2(𝒵τ;𝒮τ)Lm,2(𝒵τ;𝒮τ)T_{\Phi_{\tau}}:L^{m+1/2,2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau})\to L^{m,2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau})

is Fredholm of index 0 and there are constants CmC_{m} such that it satisfies

ηm+1/2,2Cm(TΦτ(η)m,2+ηm+1/4,2).\|\eta\|_{m+1/2,2}\leq C_{m}\ (\|T_{\Phi_{\tau}}(\eta)\|_{{m,2}}+\|\eta\|_{m+1/4,2}). (6.15)

A more quantitative version of these elliptic estimates will also be needed, which is given in the next proposition. It says, roughly, that the constants CmC_{m} in the elliptic estimates (6.15) grow only as fast as the derivatives of the metric gτg_{\tau} and Φτ\Phi_{\tau} in the directions tangential to 𝒵τ\mathcal{Z}_{\tau}. In the statement of the proposition, g0g_{0} is used to denote the product metric in Fermi coordinates on Nr0(𝒵τ)N_{r_{0}}(\mathcal{Z}_{\tau}) defined using gτg_{\tau}. As in Definition 3.9, g0g_{0} differs from gτg_{\tau} by a symmetric tensor of size O(r)O(r).

Corollary 6.8.

Suppose that there is an M>1M>1 such that for each mm\in\mathbb{N} the bounds

|tm(gτg0)|Mmgτg0C3(Y)|tmΦτ|MmΦτC1(Y)|\partial^{m}_{t}(g_{\tau}-g_{0})|\leq M^{m}\|g_{\tau}-g_{0}\|_{C^{3}(Y)}\hskip 42.67912pt|\partial^{m}_{t}\Phi_{\tau}|\leq M^{m}\|\Phi_{\tau}\|_{C^{1}(Y)}

hold on Nr0(𝒵τ)N_{r_{0}}(\mathcal{Z}_{\tau}). Then there is a constant CmC_{m} independent of MM such that if TΦτ(η)=fT_{\Phi_{\tau}}(\eta)=f, then the following estimate holds for every m0m\geq 0:

ηm+1/2,2Cm(fm,2+ηm+1/4,2)+CmMm(f2+η1/4,2).\|\eta\|_{m+1/2,2}\leq C_{m}\ \left(\|f\|_{m,2}\ +\ \|\eta\|_{m+1/4,2}\right)\ +\ C_{m}M^{m}\left(\|f\|_{2}\ +\ \|\eta\|_{1/4,2}\right). (6.16)

Remark 6.9.

Note that there are several distinct notions of derivative being used. In the assumption of Corollary 6.8, t\partial_{t} refers to the derivative using the product connection d in Fermi coordinates. Moreover, the derivatives in e.g. 6.10 are 3-dimensional covariant derivatives on YY, while the derivatives in (6.14) and the norms Lm,2(𝒵τ;N𝒵τ)L^{m,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) are one-dimensional covariant derivatives denoted ddt\frac{d}{dt} on 𝒵τ\mathcal{Z}_{\tau} arising from the restriction of the three-dimensional covariant derivatives to 𝒮τ,N𝒵τ\mathcal{S}_{\tau},N\mathcal{Z}_{\tau}. It is not true that the isomorphism obτ\text{ob}_{\tau} intertwines the derivatives, i.e. tobτ(η)obτ(ddtη)\nabla_{t}\text{ob}_{\tau}(\eta)\neq\text{ob}_{\tau}(\frac{d}{dt}\eta), though Lemma 4.17 in [43] shows that the regularity of the two sides match, i.e.

obτ1ΨLm,2(𝒵τ)ΨOb(𝒵τ)Hm(Y𝒵τ).\text{ob}^{-1}_{\tau}\Psi\in L^{m,2}(\mathcal{Z}_{\tau})\ \ \Leftrightarrow\ \ \Psi\in\text{\bf Ob}(\mathcal{Z}_{\tau})\cap H^{m}(Y-\mathcal{Z}_{\tau}).

6.3. Mode-Dependent Deformations

This section introduces a more sophisticated choice of a family of diffeomorphisms (6.1) to deform singular sets. These new diffeomorphisms are constructed by taking a dyadic decomposition of the disks normal to 𝒵τ\mathcal{Z}_{\tau} by annuli with radii O((n+1)1)rO(n1)O((n+1)^{-1})\leq r\leq O(n^{-1}) for nn\in\mathbb{N}. At r=0r=0, the action of the diffeomorphisms is identical to (6.12), but for r0r\geq 0 the action is tempered so that on each dyadic annulus it depends on the Fourier modes η\eta_{\ell} of η\eta only up to ||=O(n)|\ell|=O(n). As explained in Section 1.4 these mode-dependent deformations are the crucial ingredient needed to make alternating iteration converge. More specifically, via the upcoming Proposition 6.12, they are tool that gives control over the loss of regularity described in Subsection 2.5.

To begin, we introduce the following notation. If f:[0,r0)f_{\ell}:[0,r_{0})\to\mathbb{R} is a family of smooth functions indexed by \ell\in\mathbb{Z}, let f¯\underline{f} denote the operator

f¯:L2(𝒵τ;)\displaystyle\underline{f}:L^{2}(\mathcal{Z}_{\tau};\mathbb{C}) \displaystyle\longrightarrow L2(Nr0(𝒵τ);)\displaystyle L^{2}(N_{r_{0}}(\mathcal{Z}_{\tau});\mathbb{C}) (6.17)
f¯[η]\displaystyle\underline{f}[\eta] :=\displaystyle:= pfp(r)ηpeipt.\displaystyle\sum_{p\in\mathbb{Z}}f_{p}(r)\eta_{p}e^{ipt}. (6.18)

where a fixed choice of Fermi coordinates is used to associate N𝒵τ¯N\mathcal{Z}_{\tau}\simeq\underline{\mathbb{C}}, and ηp\eta_{p} are the Fourier coefficients of η(t)\eta(t). f¯\underline{f} is a [0,r0)[0,r_{0})-parameterized family of pseudo-differential operators on L2(𝒵τ;)L^{2}(\mathcal{Z}_{\tau};\mathbb{C}) whose Fourier multiplier is given by {f(r)}\{f_{\ell}(r)\}_{\ell\in\mathbb{Z}} for each fixed rr.

Next, let R0>0R_{0}>0 be a large positive number to be specified shortly, and denote by χ:[0,)\chi:[0,\infty)\to\mathbb{R} a smooth cut-off function equal to 1 for rR0r\leq R_{0} and supported in the region where r2R0r\leq 2R_{0}. Assume it is chosen so that

|dχ|CR0.|d\chi|\leq\frac{C}{R_{0}}. (6.19)

Additionally, let χr0\chi_{r_{0}} denote a second smooth cut-ff function equal to 1 for rr0/2r\leq r_{0}/2 and supported in Nr0(𝒵τ)N_{r_{0}}(\mathcal{Z}_{\tau}). Then, for each \ell\in\mathbb{Z}, set

χ(r):=χ(||r)χr0(r).\chi_{\ell}(r):=\chi(|\ell|r)\chi_{r_{0}}(r). (6.20)

The family χ\chi_{\ell} gives rise an operator χ¯\underline{\chi} as in (6.17).

Using this, define a new family of diffeomorphisms by

F¯η(t,z):=(t,z+χ¯[η])\underline{F}_{\eta}(t,z):=\left(t,z+\underline{\chi}[\eta]\right) (6.21)

on Nr0(𝒵τ)N_{r_{0}}(\mathcal{Z}_{\tau}) and extended to YY by the identity. For η2,2\|\eta\|_{2,2} and r0r_{0} sufficiently small, the Inverse Function Theorem shows that F¯η\underline{F}_{\eta} is indeed a diffeomorphism on Nr0(𝒵τ)N_{r_{0}}(\mathcal{Z}_{\tau}). As in Sections 5.15.2, F¯η\underline{F}_{\eta} gives rise to (i) an exponential map Exp¯\underline{\text{Exp}},  (ii) trivializations Υ¯\underline{\Upsilon},  a (iii) a family of pullback metrics g¯η\underline{g}_{\eta},  (iv) a partial derivative ¯Φτ(η)\underline{\mathcal{B}}_{\Phi_{\tau}}(\eta),   and (v) a deformation operator T¯Φτ\underline{T}_{\Phi_{\tau}}.

The following corollary states that the results of Section 6.2 carry over to the underlined versions. It also dictates a lower bound for the value of the constant R0R_{0} used to define (6.17).

Corollary 6.10.

The operator T¯Φτ\underline{T}_{\Phi_{\tau}} is given by

T¯Φτ=TΦτ+TR0\underline{T}_{\Phi_{\tau}}=T_{\Phi_{\tau}}+T_{R_{0}}

where TR0T_{R_{0}} is a pseudo-differential operator of order 1/21/2 and for some KK\in\mathbb{N} and M>10M>10, it satisfies

TR0C(R0)KExp(R0/c)+CR0M\|T_{R_{0}}\|\leq C(R_{0})^{K}\text{Exp}(-R_{0}/c)\ +\ CR_{0}^{-M}

In particular, if (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) has unobstructed deformations then for R0R_{0} sufficiently large, T¯Φτ\underline{T}_{\Phi_{\tau}} is invertible for ττ0\tau\leq\tau_{0}, and the results of Corollaries 6.7 and 6.8 continue to hold.

Proof Sketch.

The idea of the proof is straightforward, given the following observation: the cokernel elements Ψ\Psi_{\ell} are nearly isolated in the th\ell^{th} Fourier mode and decay exponentially with with exponent 1/||1/|\ell|. Therefore, altering the deformation of the th\ell^{th} mode outside R0/||R_{0}/|\ell| for R0R_{0} large enough is a small perturbation. ∎

Remark 6.11.

(Cf. Remark 5.6 of [43] and Section 4.1 of [12]) There is an infinite-dimensional space of possible choices of a family of diffeomorphisms 𝔽\mathbb{F} of (6.1). Any two choices differ by pre-composing with a family of diffeomorphisms GηDiff(Y;𝒵τ)G_{\eta}\in\text{Diff}(Y;\mathcal{Z}_{\tau}) that fix 𝒵τ\mathcal{Z}_{\tau}. By pullbacks, this latter group acts by infinite-dimensional gauge transformations on the bundles of Lemma 6.1. Thus the choice of a family of diffeomorphisms is effectively a choice of gauge, and the expressions for the operator (6.13) depend on this choice. The choice of mode-dependent deformations may be understood as a gauge in which certain stronger estimates are available, akin to the elliptic estimates provided by the Coulomb gauge in standard situations.

The next lemma provides key bounds that are used to control the loss of regularity in Sections 911. The collection of terms it will be necessary to bound have two general types.

  • (A)

    Let Φ\Phi be a spinor such that there is an α>0\alpha>0, so that for all k0k\in\mathbb{Z}^{\geq 0}, |tkΦ|Ck1rα|\nabla^{k}_{t}\Phi|\leq C_{k_{1}}r^{\alpha} and |tkvΦ|Ckrα1|\nabla^{k}_{t}\nabla_{v}\Phi|\leq C_{k}r^{\alpha-1}, where v=xv=\partial_{x}, or y\partial_{y} in Fermi coordinates. Consider terms of the form

    MΦ(η):=mχ¯[dtnη]σjkΦ.M_{\Phi}(\eta):=\underline{\partial^{m}\chi}[d_{t}^{n}\eta]\sigma_{j}\nabla^{k}\Phi. (6.22)
  • (B)

    Let ψrHe1\psi\in rH^{1}_{e} and set k=0,1k=0,1. Consider terms of the form

    Mψ(η)=mχ¯[dtnη]σjkψ.M_{\psi}(\eta)=\underline{\partial^{m}\chi}[d_{t}^{n}\eta]\sigma_{j}\nabla^{k}\psi. (6.23)

In these expressions, mχ¯\underline{\partial^{m}\chi} denotes the operator (6.17) formed analogously to (6.20) but using mχ\partial^{m}\chi for a a multi-index of order mm, dtnηd_{t}^{n}\eta is the nthn^{th} derivative of ηC(𝒵τ;N𝒵τ)\eta\in C^{\infty}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}), and σj\sigma_{j} is chosen from a collection of fixed pointwise linear endomorphisms (e.g. γ(ej)\gamma(e^{j})). We say terms of the form (6.22) and (6.23) have weights

wA=m+n+k.wB=m+n{w_{A}=m+n+k.}\hskip 56.9055pt{w_{B}=m+n}

respectively.

Lemma 6.12.

Let ηC(𝒵τ;N𝒵τ)\eta\in C^{\infty}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) be a deformation, then the following bounds hold.

  1. (A)

    Suppose that MΦM_{\Phi} is a term of Type A as in (6.22) having weight wAw_{A}, and that β\beta\in\mathbb{R}. Then

    rβMΦ(η)L2(Nr0)CηwA(1+α+β)\|r^{\beta}M_{\Phi}(\eta)\|_{L^{2}(N_{r_{0}})}\leq C\|\eta\|_{w_{A}-(1+\alpha+\beta)}

    where ηs\|\eta\|_{s} denotes the Ls,2L^{s,2}-norm on 𝒵τ\mathcal{Z}_{\tau}. In particular, the terms of weight wA=2w_{A}=2 satisfy

    • χ¯[η′′]σjrβΦL2(Nr0)+aχ¯[η]σjrβΦL2(Nr0)+abχ¯[η]σjrβΦL2(Nr0)Cη1αβ\Big{\|}\underline{\chi}[\eta^{\prime\prime}]\sigma_{j}r^{\beta}\Phi\Big{\|}_{L^{2}(N_{r_{0}})}+\Big{\|}\underline{\partial_{a}\chi}[\eta^{\prime}]\sigma_{j}r^{\beta}\Phi\Big{\|}_{L^{2}(N_{r_{0}})}+\Big{\|}\underline{\partial_{ab}\chi}[\eta]\sigma_{j}r^{\beta}\Phi\Big{\|}_{L^{2}(N_{r_{0}})}\leq C\|\eta\|_{1-\alpha-\beta}

    • χ¯[η]σjrβbΦL2(Nr0)+aχ¯[η]σjrβbΦL2(Nr0)Cη1αβ\Big{\|}\underline{\chi}[\eta^{\prime}]\sigma_{j}r^{\beta}\nabla_{b}\Phi\Big{\|}_{L^{2}(N_{r_{0}})}+\Big{\|}\underline{\partial_{a}\chi}[\eta]\sigma_{j}r^{\beta}\nabla_{b}\Phi\Big{\|}_{L^{2}(N_{r_{0}})}\leq C\|\eta\|_{1-\alpha-\beta}

    where a,b=t,x,ya,b=t,x,y.

  2. (B)

    Suppose that Mψ(η)M_{\psi}(\eta) is a term of Type B as in (6.23) having weight wBw_{B}. Then, for some γ¯<<1\underline{\gamma}<<1

    rβMψ(η)L2(Nr0)CηwB+γ¯(1/2+β)ψrHe1.\|r^{\beta}M_{\psi}(\eta)\|_{L^{2}(N_{r_{0}})}\leq C\|\eta\|_{w_{B}+\underline{\gamma}-(1/2+\beta)}\|\psi\|_{rH^{1}_{e}}.

    In particular, for wB=2w_{B}=2,

    • χ¯[η′′]σjψrβL2(Nr0)+aχ¯[η]σjψrβL2(Nr0)+abχ¯[η]σjψrβL2(Nr0)Cη3/2+γ¯βψrL2\Big{\|}\underline{\chi}[\eta^{\prime\prime}]\sigma_{j}\psi r^{\beta}\Big{\|}_{L^{2}(N_{r_{0}})}+\Big{\|}\underline{\partial_{a}\chi}[\eta^{\prime}]\sigma_{j}\psi r^{\beta}\Big{\|}_{L^{2}(N_{r_{0}})}+\Big{\|}\underline{\partial_{ab}\chi}[\eta]\sigma_{j}\psi r^{\beta}\Big{\|}_{L^{2}(N_{r_{0}})}\leq C\|\eta\|_{3/2+\underline{\gamma}-\beta}\|\tfrac{\psi}{r}\|_{L^{2}}

    • χ¯[η]σjrβbψL2(Nr0)+aχ¯[η]σjrβbψL2(Nr0)Cη3/2+γ¯βψL2.\Big{\|}\underline{\chi}[\eta^{\prime}]\sigma_{j}r^{\beta}\nabla_{b}\psi\Big{\|}_{L^{2}(N_{r_{0}})}+\Big{\|}\underline{\partial_{a}\chi}[\eta]\sigma_{j}r^{\beta}\nabla_{b}\psi\Big{\|}_{L^{2}(N_{r_{0}})}\leq C\|\eta\|_{3/2+\underline{\gamma}-\beta}\|\nabla{\psi}\|_{L^{2}}.

Proof.

The lemma is proved working locally in Fermi coordinates and an accompanying trivialization as in Section 3.4. Thus η,Φ\eta,\Phi become \mathbb{C} and 2\mathbb{C}^{2}\otimes\mathbb{H}-valued functions respectively. The lemma is proved in the cases of wA=wB=2w_{A}=w_{B}=2; the general cases are similar.

(A) To begin, consider the term χ¯[η′′]\underline{\chi}[\eta^{\prime\prime}] and the case that β=0\beta=0. On each circle S1×{(x,y)}Nr0(𝒵τ)S^{1}\times\{(x,y)\}\subseteq N_{r_{0}}(\mathcal{Z}_{\tau}) the multiplication map L1,2(S1)×L2(S1)L2(S1)L^{1,2}(S^{1})\times L^{2}(S^{1})\to L^{2}(S^{1}) is uniformly boundedin (x,y)(x,y). Applying this to the product Φ(t,x,y)χ¯[η′′]\Phi(t,x,y)\underline{\chi}[\eta^{\prime\prime}] for each pair (x,y)(x,y),

χ¯[η′′]σjΦL2(Nr0)2\displaystyle\Big{\|}\underline{\chi}[\eta^{\prime\prime}]\sigma_{j}\Phi\Big{\|}^{2}_{L^{2}(N_{r_{0}})} =\displaystyle= D𝕣0S1|χ¯[η′′]σjΦ|2𝑑tr𝑑θ𝑑r\displaystyle\int_{D_{\mathbb{r}_{0}}}\int_{S^{1}}|\underline{\chi}[\eta^{\prime\prime}]\sigma_{j}\Phi|^{2}dt\ rd\theta dr
\displaystyle\leq Dr0|C0rα|2S1|χ¯[η′′]|2𝑑tr𝑑θ𝑑r\displaystyle\int_{D_{r_{0}}}|C_{0}r^{\alpha}|^{2}\cdot\int_{S^{1}}|\underline{\chi}[\eta^{\prime\prime}]|^{2}dt\ rd\theta dr
\displaystyle\leq C12Dr0|pηp(ip)2eiptχp(r)|2r2α+1𝑑θ𝑑r𝑑t\displaystyle C_{1}^{2}\int_{D_{r_{0}}}|\sum_{p\in\mathbb{Z}}\eta_{p}(ip)^{2}e^{ipt}\chi_{p}(r)|^{2}r^{2\alpha+1}d\theta drdt

where we have replaced the integral over S1S^{1} with a Dr0D_{r_{0}}-parameterized sum over Fourier modes via Plancherel’s Theorem.

Since the integrand is uniformly bounded, and ηp\eta_{p} is independent of (x,y)(x,y), the sum may be pulled outside, giving

\displaystyle\leq Cp|ηp|2|p|4Dr0|χp(r)|2r2α+1𝑑θ𝑑r𝑑t\displaystyle C\sum_{p\in\mathbb{Z}}|\eta_{p}|^{2}|p|^{4}\int_{D_{r_{0}}}|\chi_{p}(r)|^{2}r^{2\alpha+1}d\theta drdt (6.24)
\displaystyle\leq Cp|ηp|2|p|4[r2α+2]r=0R0/p\displaystyle C\sum_{p\in\mathbb{Z}}|\eta_{p}|^{2}|p|^{4}[r^{2\alpha+2}]_{r=0}^{R_{0}/p} (6.25)
\displaystyle\leq Cp|ηp|2|p|22α=Cη1α,2\displaystyle C\sum_{p\in\mathbb{Z}}|\eta_{p}|^{2}|p|^{2-2\alpha}=C\|\eta\|_{1-\alpha,2} (6.26)

as desired.

For the aχ¯[η]\underline{\partial_{a}\chi}[\eta^{\prime}] term, the proof is the same, except we use the bound |aχp|C|p||\partial_{a}\chi_{p}|\leq C|p|; this adds a factor of |p|2|p|^{2} to the integrand, but the same factor is lost by replacing η′′\eta^{\prime\prime} by η\eta^{\prime}. For the term abχ¯[η]\underline{\partial_{ab}\chi}[\eta] the same applies, now with |p|4|p|^{4}. Likewise, for the terms involving aΦ\nabla_{a}\Phi, one power of |p||p| is gained from using rα1r^{\alpha-1} instead of α\alpha, but the same factor is again lost since [¯η],aχ¯[η]\underline{[}\eta^{\prime}],\underline{\partial_{a}\chi}[\eta] both have one less derivative. The case with β0\beta\neq 0 is again similar.

(B) The proof in this case is essentially the same, except we use that multiplication C0(S1)×L2(S1)L2(S1)C^{0}(S^{1})\times L^{2}(S^{1})\to L^{2}(S^{1}) is bounded, and apply it to the term η′′ψ(t,x0,y0)\eta^{\prime\prime}\psi(t,x_{0},y_{0}) since ψ\psi does not necessarily have pointwise bounds. First, divide Dr0D_{r_{0}} into the sequence of annuli

An:={2R0n+1r2R0n}A_{n}:=\{\tfrac{2R_{0}}{n+1}\leq r\leq\tfrac{2R_{0}}{n}\}

for n1n\geq 1. Then

χ¯[η′′]σjψrβL2(Nr0)2\displaystyle\Big{\|}\underline{\chi}[\eta^{\prime\prime}]\sigma_{j}\psi r^{\beta}\Big{\|}^{2}_{L^{2}(N_{r_{0}})} =\displaystyle= n1AnS1|χ¯[η′′]σjψ|2r2β𝑑tr𝑑θ𝑑r\displaystyle\sum_{n\geq 1}\int_{A_{n}}\int_{S^{1}}|\underline{\chi}[\eta^{\prime\prime}]\sigma_{j}\psi|^{2}r^{2\beta}dt\ rd\theta dr
\displaystyle\leq n1supAnχ¯(η′′)C0(S1)AnS1|ψ|2r2β𝑑tr𝑑θ𝑑r\displaystyle\sum_{n\geq 1}\sup_{A_{n}}\|\underline{\chi}(\eta^{\prime\prime})\|_{C^{0}(S^{1})}\int_{A_{n}}\int_{S^{1}}|\psi|^{2}r^{2\beta}dt\ rd\theta dr
\displaystyle\leq n1supAnχ¯(η′′)C0(S1)2supAnr2+2βψrL2(An).\displaystyle\sum_{n\geq 1}\sup_{A_{n}}\|\underline{\chi}(\eta^{\prime\prime})\|^{2}_{C^{0}(S^{1})}\sup_{A_{n}}r^{2+2\beta}\|\tfrac{\psi}{r}\|_{L^{2}(A_{n})}.

Now, since χp=0\chi_{p}=0 on AnA_{n} for p2(n+1)p\geq 2(n+1), we see that

supAnχ¯(η′′)C0Cπ2(n+1)η′′C0Cπ2(n+1)η5/2+γ¯\sup_{A_{n}}\|\underline{\chi}(\eta^{\prime\prime})\|_{C^{0}}\leq C\|\pi^{2(n+1)}\eta^{\prime\prime}\|_{C^{0}}\leq C\|\pi^{2(n+1)}\eta\|_{{5/2+\underline{\gamma}}}

where π2(n+1)\pi^{2(n+1)} denotes the projection to Fourier modes |p|2(n+1)|p|\leq 2(n+1). Since supAnr2+2βCn2+2β\sup_{A_{n}}r^{2+2\beta}\leq\tfrac{C}{n^{2+2\beta}}, the restriction on Fourier modes implies Cη5/2+γ¯n22βCη3/2+γ¯βC\|\eta\|_{5/2+\underline{\gamma}}n^{-2-2\beta}\leq C\|\eta\|_{3/2+\underline{\gamma}-\beta}. Using this, the above is bounded by

\displaystyle\leq Cn1π2n+1η3/2+γ¯βψrL2(An)\displaystyle C\sum_{n\geq 1}\|\pi^{2n+1}\eta\|_{{3/2+\underline{\gamma}-\beta}}\|\tfrac{\psi}{r}\|_{L^{2}(A_{n})}
\displaystyle\leq Cη3/2+γ¯βψrL2(Nr0)\displaystyle C\|\eta\|_{{3/2+\underline{\gamma}-\beta}}\|\tfrac{\psi}{r}\|_{L^{2}(N_{r_{0}})}

as desired. Just as in (A), the aχ¯[η],abχ¯[η]\underline{\partial_{a}\chi}[\eta^{\prime}],\underline{\partial_{ab}\chi}[\eta] terms lose a power of nn from the derivative but gain one from the derivatives of the cut-off. The terms with ψ\nabla\psi are the same without the r2r^{2} factor. ∎

6.4. Non-Linear Terms

The universal Dirac operator \not{\mathbb{D}} is a fully non-linear function of the singular set 𝒵\mathcal{Z}. This section describes the non-linear terms. When solving the linearized equations in every step of the gluing iteration, these non-linear terms contribute a portion of the error for the subsequent step.

Given a configuration h0=(𝒵,Φ)e1h_{0}=(\mathcal{Z},\Phi)\in\mathbb{H}^{1}_{e},

(h0+(η,ψ))=h0+dh0(η,ψ)+Qh0(η,ψ)\not{\mathbb{D}}(h_{0}+(\eta,\psi))=\not{\mathbb{D}}h_{0}+\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,\psi)+Q_{h_{0}}(\eta,\psi)

where dh0\text{d}_{h_{0}}\not{\mathbb{D}} is the linearization (6.6) and Qh0Q_{h_{0}} is the non-linear term. Here, 𝒵+η\mathcal{Z}+\eta means 𝒵η\mathcal{Z}_{\eta} as in (6.4) using the chart centered at 𝒵\mathcal{Z}. The following lemma characterizes the non-linear term Qh0Q_{h_{0}}.

Lemma 6.13.

The non-linear term QQ has the following form:

Qh0(η,ψ)\displaystyle Q_{h_{0}}(\eta,\psi) =\displaystyle= ¯ψ(η)+MΦ(η,η)+Mψ(η,η)+FΦ+ψ(η)\displaystyle\underline{\mathcal{B}}_{\psi}(\eta)\ \ +\ \ M_{\Phi}(\eta,\eta)\ \ +\ \ M_{\psi}(\eta,\eta)\ \ +\ \ F_{\Phi+\psi}(\eta)

where

  • ¯ψ(η)\underline{\mathcal{B}}_{\psi}(\eta) is as in (6.7) with ψ\psi in place of Φτ\Phi_{\tau}.

  • MΦ(η,η)M_{\Phi}(\eta,\eta) is a term quadratic in η\eta and linear in Φ\Phi which is a finite sum of terms of the form

    m(y)a1(χ¯[η])MΦ(η)m(y)\cdot a_{1}(\underline{\chi}[\eta])\cdot M_{\Phi}(\eta)

    where m(y)C(Y;End(SRe))m(y)\in C^{\infty}(Y;\text{End}(S^{\text{Re}})), MΦ(η)M_{\Phi}(\eta) is a linear term of type A in the sense of Lemma 6.12 with weight wA=2w_{A}=2, and a1(χ¯[η])a_{1}(\underline{\chi}[\eta]) is a linear combination of χ¯[η],aχ¯[η]\underline{\chi}[\eta^{\prime}],\underline{\partial_{a}\chi}[\eta], and χ¯[η]\underline{\chi}[\eta].

  • Mψ(η,η)M_{\psi}(\eta,\eta) is the same with the final term replace by a term Mψ(η)M_{\psi}(\eta) of Type B in the sense of Lemma 6.12 with weight wB=2w_{B}=2.

  • FΦ+ψ(η)F_{\Phi+\psi}(\eta) has the same form but with higher-order dependence on η\eta, so that it satisfies a bound

    |FΦ+ψ(η)|CηC1(MΦ(η,η)+Mψ(η,η))|F_{\Phi+\psi}(\eta)|\leq C\|\eta\|_{C^{1}}\Big{(}M_{\Phi}^{\prime}(\eta,\eta)+M_{\psi}^{\prime}(\eta,\eta)\Big{)}

    where MΦ,MψM_{\Phi}^{\prime},M_{\psi}^{\prime} are of the same form as MΦ,MψM_{\Phi},M_{\psi} respectively.

Proof.

This formula is derived by substituting the formula for the pullback metric gη=Fχ¯(η)gg_{\eta}=F_{\underline{\chi}}(\eta)^{*}g into the non-linear version of Bourguignon-Gauduchon’s formula in Theorem 6.4. See Section 8.3 of [43]. ∎

To bound the non-linear terms later, we have the following analogue of Lemma 6.12. In the statement of the lemma, we tacitly use χ¯[η]\underline{\chi^{\prime}}[\eta] to denote a term having one derivative, i.e. a linear combination of χ¯[η]\underline{\chi}[\eta^{\prime}] and aχ¯[η]\underline{\partial_{a}\chi}[\eta]. χ¯′′[η]\underline{\chi}^{\prime\prime}[\eta] denotes the same but with up to second derivatives.

Lemma 6.14.

Retaining the assumptions and notation of Lemma 6.12, the following bounds hold for η,ξC(𝒵τ;N𝒵τ)\eta,\xi\in C^{\infty}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}).

  1. (A’)

    For a fixed small number number γ¯<<1\underline{\gamma}<<1,

    χ¯[η]χ¯[ξ]ΦL2(Nr0)\displaystyle\Big{\|}\underline{\chi^{\prime}}[\eta]\underline{\chi^{\prime}}[\xi]\nabla\Phi\Big{\|}_{L^{2}(N_{r_{0}})} \displaystyle\leq Cη3/2+γ¯ξ1α\displaystyle C\|\eta\|_{3/2+\underline{\gamma}}\cdot\|\xi\|_{1-\alpha}
    χ′′¯[η]χ¯[ξ]ΦL2(Nr0)+χ¯[η]χ′′¯[ξ]ΦL2(Nr0)\displaystyle\Big{\|}\underline{\chi^{\prime\prime}}[\eta]\underline{\chi^{\prime}}[\xi]\Phi\Big{\|}_{L^{2}(N_{r_{0}})}+\Big{\|}\underline{\chi^{\prime}}[\eta]\underline{\chi^{\prime\prime}}[\xi]\Phi\Big{\|}_{L^{2}(N_{r_{0}})} \displaystyle\leq C(η3/2+γ¯ξ1α+η1αξ3/2+γ¯)\displaystyle C(\|\eta\|_{3/2+\underline{\gamma}}\cdot\|\xi\|_{1-\alpha}+\|\eta\|_{1-\alpha}\cdot\|\xi\|_{3/2+\underline{\gamma}})
  2. (B’)

    Likewise,

    χ¯[η]χ¯[ξ]rβψL2(Nr0)\displaystyle\Big{\|}\underline{\chi^{\prime}}[\eta]\underline{\chi^{\prime}}[\xi]r^{\beta}\nabla\psi\Big{\|}_{L^{2}(N_{r_{0}})} \displaystyle\leq Cη3/2+γ¯βξ3/2+γ¯ψrL2\displaystyle C\|\eta\|_{3/2+\underline{\gamma}-\beta}\|\xi\|_{3/2+\underline{\gamma}}\|\tfrac{\psi}{r}\|_{L^{2}}
    χ′′¯[η]χ¯[ξ]rβψL2(Nr0)+χ¯[η]χ′′¯[ξ]rβψL2(Nr0)\displaystyle\Big{\|}\underline{\chi^{\prime\prime}}[\eta]\underline{\chi^{\prime}}[\xi]r^{\beta}\psi\Big{\|}_{L^{2}(N_{r_{0}})}+\Big{\|}\underline{\chi^{\prime}}[\eta]\underline{\chi^{\prime\prime}}[\xi]r^{\beta}\psi\Big{\|}_{L^{2}(N_{r_{0}})} \displaystyle\leq C(ξ3/2+γ¯η3/2+γ¯β+η3/2+γ¯ξ3/2+γ¯β)ψrL2\displaystyle C(\|\xi\|_{3/2+\underline{\gamma}}\|\eta\|_{3/2+\underline{\gamma}-\beta}+\|\eta\|_{3/2+\underline{\gamma}}\|\xi\|_{3/2+\underline{\gamma}-\beta})\|\tfrac{\psi}{r}\|_{L^{2}}
Proof.

Notice that the bound dχpC|p|d\chi_{p}\leq C|p| for each pp\in\mathbb{Z} implies that aχ¯[η]L2η1,2\|\underline{\partial_{a}\chi}[\eta]\|_{L^{2}}\lesssim\|\eta\|_{1,2}. In each case above, a factor with one derivative can therefore be pulled out using the Sobolev embedding χ[η]C0CηL3/2+γ¯,2\|\chi^{\prime}[\eta]\|_{C^{0}}\leq C\|\eta\|_{L^{3/2+\underline{\gamma},2}} or equivalently for ξ\xi, after which the proofs proceed as in Lemma 6.12. ∎

7. Concentrating Local Solutions

This section reviews the model solutions constructed in [41]. This ε\varepsilon-parameterized family of model solutions solve the Seiberg–Witten equations on a tubular neighborhood of the singular set, and locally converge to (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}), with curvature concentrating along 𝒵0\mathcal{Z}_{0} as ε0\varepsilon\to 0. Results are stated here without proof, with references to the relevant sections of [41]. These result here also include the trivial extension of [41] from a single 2\mathbb{Z}_{2}-harmonic spinor to the τ\tau-parameterized family of eigenvectors (1.12).

7.1. Hilbert Spaces and Boundary Conditions

This subsection defines Sobolev spaces with weights and boundary conditions on a tubular neighborhoods of each singular set 𝒵τ\mathcal{Z}_{\tau}. These (ε,τ)(\varepsilon,\tau)-parameterized tubular neighborhoods — which eventually host the model solutions — shrink in diameter as ε0\varepsilon\to 0 for each fixed τ\tau. More specifically, let

λ(ε)=ε1/2\boxed{\lambda(\varepsilon)=\varepsilon^{1/2}} (7.1)

and, immediately suppressing the dependence on ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), denote by Y+=Nλ(𝒵τ)Y^{+}=N_{\lambda}(\mathcal{Z}_{\tau}) the tubular neighborhood of the singular set for τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}). It is equipped with Fermi coordinates as in Definition 3.9 using the metric gτg_{\tau}.

The Sobolev norms depend on two weight functions, which we now describe. The latter of these is defined in terms of a “de-singularized spinor” Φτhε\Phi_{\tau}^{h_{\varepsilon}}, which is a precursor to the model solutions. Denote by r0r_{0} the radius of the Fermi coordinate chart.

  1. (1)

    With r=dist(,𝒵τ)r=\text{dist}(-,\mathcal{Z}_{\tau}), let RεR_{\varepsilon} be a smooth function defined as follows:

    Rε={κ2ε4/3+r2rr0/2constrr0,R_{\varepsilon}=\begin{cases}\sqrt{\kappa^{2}\varepsilon^{4/3}+r^{2}}\hskip 22.76228ptr\leq r_{0}/2\\ \text{const}\hskip 56.9055ptr\geq r_{0},\end{cases}

    where κ=κ(τ)\kappa=\kappa(\tau) is a uniformly bounded constant depending depending smoothly on τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}) defined in terms of the leading coefficients of Φτ\Phi_{\tau}.

  2. (2)

    For each τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}), let Φτhε\Phi_{\tau}^{h_{\varepsilon}} be the de-singularized spinor as defined in Definition 4.5 of [41]. The norm |Φτhε||\Phi^{h_{\varepsilon}}_{\tau}| is a smooth function on YY with the following properties for constants c1,c2c_{1},c_{2} (see [41, Sec. 4]):

    • (2a)

      For rc1ε2/3r\geq c_{1}\varepsilon^{2/3}, it satisfies ||Φτhε||Φτ||CExp(r3/2ε)\big{|}|\Phi^{h_{\varepsilon}}_{\tau}|-|\Phi_{\tau}|\big{|}\leq C\text{Exp}(-\tfrac{r^{3/2}}{\varepsilon}).

    • (2b)

      For rλr\leq\lambda, it is a monotonically increasing function of rr with a uniform lower bound |Φτhε|c2ε1/3|\Phi_{\tau}^{h_{\varepsilon}}|\geq c_{2}\varepsilon^{1/3}.

The weight function RεR_{\varepsilon} is approximately equal to rr on a tubular neighborhood of 𝒵τ\mathcal{Z}_{\tau}, being constant outside this neighborhood, and leveling off at rε2/3r\sim\varepsilon^{2/3} so that there is a uniform lower bound Rεc1ε2/3R_{\varepsilon}\geq c_{1}\varepsilon^{2/3}. The weight function |Φτhε||\Phi_{\tau}^{h_{\varepsilon}}| is exponentially close to |Φτ|r1/2|\Phi_{\tau}|\sim r^{1/2} for c1ε2/3rc_{1}\varepsilon^{2/3}\leq r. This latter weight function is effectively commensurate with |Φτhε|Rε|\Phi_{\tau}^{h_{\varepsilon}}|\sim\sqrt{R_{\varepsilon}}, but is used instead because it naturally appears in the Weitzenböck formula for the linearized operator.

Definition 7.1.

Let ν\nu\in\mathbb{R} be a weight. The “inside” Sobolev norms are defined by

(φ,a)Hε,ν1,+\displaystyle\|(\varphi,a)\|_{H^{1,+}_{\varepsilon,\nu}} :=\displaystyle:= (Y+(|φ|2+|a|2+|φ|2Rε2+|μ(φ,Φτhε)|2ε2+|a|2|Φτhε|2ε2)Rε2ν𝑑V)1/2\displaystyle\left(\int_{Y^{+}}\left(|\nabla\varphi|^{2}+|\nabla a|^{2}+\frac{|\varphi|^{2}}{R_{\varepsilon}^{2}}+\frac{|\mu(\varphi,\Phi^{h_{\varepsilon}}_{\tau})|^{2}}{\varepsilon^{2}}+\frac{|a|^{2}|\Phi^{h_{\varepsilon}}_{\tau}|^{2}}{\varepsilon^{2}}\right)R_{\varepsilon}^{2\nu}\ dV\right)^{1/2} (7.2)
(φ,a)Lε,ν2,+\displaystyle\|(\varphi,a)\|_{L^{2,+}_{\varepsilon,\nu}} :=\displaystyle:= (Y+(|φ|2+|a|2)Rε2ν𝑑V)1/2\displaystyle\left(\int_{Y^{+}}\left(|\varphi|^{2}+|a|^{2}\right)R_{\varepsilon}^{2\nu}\ dV\right)^{1/2} (7.3)

where the dependence of Φhε,dV\Phi^{h_{\varepsilon}},dV, and RεR_{\varepsilon} on τ\tau is suppressed in the notation. Both norms give rise to inner products via their polarizations. Because Nλ(𝒵τ)N_{\lambda}(\mathcal{Z}_{\tau}) is compact, these norms are equivalent to the standard L1,2L^{1,2} and L2L^{2} norms respectively (though not uniformly in ε,ν\varepsilon,\nu).

Next, Sobolev spaces are defined using these norms along with a set of boundary conditions on Y+\partial Y^{+} and orthogonality constraints in the interior. The boundary conditions are a twisted variation of APS boundary conditions; the details of these conditions are crucial for the proofs of the theorems in [41], but not needed here (see Remark 7.4 below). The reader is referred to Sections 7.1–7.3 of [41] for details, and in particular to Figure 2 of [41] which illustrates these conditions.

The following lemma is a restatement of Proposition 7.15 in [41] with unnecessary details omitted.

Lemma 7.2.

There exists a Hilbert subspace

HΦτ+L1/2,2(Y+)VH^{+}_{\Phi_{\tau}}\subseteq L^{1/2,2}(\partial Y^{+})\oplus V

where VL2(Y+)V\subseteq L^{2}(Y^{+}) is a subspace of complex dimension 1+2L0ε1/21+2L_{0}\varepsilon^{-1/2} with L0L_{0}\in\mathbb{R}, and a projection operator Π+:L1,2(Y+)HΦτ+\Pi^{+}:L^{1,2}(Y^{+})\to H^{+}_{\Phi_{\tau}} given by the direct sum of the boundary restriction and the L2L^{2}-orthogonal projection to VV, such that

((Φ,A),Π+):L1,2(Y+)L2(Y+)HΦτ+(\mathcal{L}_{(\Phi,A)}\ ,\ \Pi^{+}):L^{1,2}(Y^{+})\longrightarrow L^{2}(Y^{+})\oplus H^{+}_{\Phi_{\tau}} (7.4)

is Fredholm of Index 0 for any configuration (Φ,A)C(Y+)(\Phi,A)\in C^{\infty}(Y^{+}).


Definition 7.3.

Define the “inside” Hilbert spaces by

Hε,ν1,+\displaystyle H^{1,+}_{\varepsilon,\nu} :=\displaystyle:= {(φ,a)|(φ,a)Hε,ν1,+<,Π(φ,a)=0}\displaystyle\Big{\{}(\varphi,a)\ \big{|}\ \|(\varphi,a)\|_{H^{1,+}_{\varepsilon,\nu}}<\infty\ \ ,\ \ \Pi^{\mathcal{L}}(\varphi,a)=0\Big{\}}
Lε,ν2,+\displaystyle L^{2,+}_{\varepsilon,\nu} :=\displaystyle:= {(φ,a)|(φ,a)Lε,ν2,+<}.\displaystyle\Big{\{}(\varphi,a)\ \big{|}\ \|(\varphi,a)\|_{L^{2,+}_{\varepsilon,\nu}}<\infty\Big{\}}.

They are equipped with the norms from Definition 7.1, and the inner products arising from their polarizations. They are defined over the domain Y+=Nλ(𝒵τ)Y^{+}=N_{\lambda}(\mathcal{Z}_{\tau}), and depend implicitly on τ\tau.

Remark 7.4.

Although the details of the boundary conditions defined by Lemma 7.2 are peripheral to our purposes here, their importance for the overall gluing construction cannot be overstated, nor should the reader be lulled into dismissing the result of the lemma as an application of standard APS theory. Although the existence of some space HH such that the result of Lemma 7.2 about the index holds is a consequence of standard APS theory [2, 34, 4], the linearized equations at the model solutions (as in the upcoming Theorem 7.7) will necessarily fail to be invertible in a suitable sense for any but a very precisely crafted boundary condition. The reasons for this are explained in detail in Section 7.1.3 of [41], and the correct definition of Π+\Pi^{+} is the main challenge in [41].

7.2. Concentrating Local Solutions

This subsection states the main results of [41] about the model solutions and the linearization at these.

Theorem 1.2 of [41] first constructs model solutions (Φε,τ+,Aε,τ+)(\Phi^{+}_{\varepsilon,\tau},A^{+}_{\varepsilon,\tau}) on Nλ(𝒵τ)N_{\lambda}(\mathcal{Z}_{\tau}) that satisfy the Seiberg–Witten equations and the boundary conditions of Lemma 7.9. Using the cut-off function χ+\chi^{+} (recall this is equal to 11 for rλ/2r\leq\lambda/2 and vanishes for r3λ/4r\geq 3\lambda/4), these are are extended by the limiting eigenvector and connection to form global approximate solutions:

(Φ1,A1):=χ+(Φε,τ+,Aε,τ+)+(1χ+)(Φτ,Aτ).(\Phi_{1},A_{1}):=\chi^{+}\cdot(\Phi^{+}_{\varepsilon,\tau},A^{+}_{\varepsilon,\tau})+(1-\chi^{+})\cdot(\Phi_{\tau},A_{\tau}). (7.5)

On the left, the dependence on (ε,τ)(\varepsilon,\tau) is suppressed. The subscript “1” indicates that these are the first in the sequence of approximate solutions (2.2) arising in the alternating iteration.

Theorem 7.5.

([41], Theorem 1.2 ) Suppose that (𝒵τ,Aτ,Φτ)(\mathcal{Z}_{\tau},A_{\tau},\Phi_{\tau}) for τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}) are a family of 2\mathbb{Z}_{2}-harmonic eigenvectors satisfying the hypotheses of Theorem 1.6. Then, there exist approximate solutions (Φ1,A1)(\Phi_{1},A_{1}) smoothly parameterized by (ε,τ)(\varepsilon,\tau) and constructed as in (7.5) with the following properties.

  1. (I)

    They satisfy

    SW(Φ1ε,A1)=χΛ(τ)Φτε+e1+f1\text{SW}\left(\frac{\Phi_{1}}{\varepsilon},A_{1}\right)=\frac{\chi^{-}\Lambda(\tau)\Phi_{\tau}}{\varepsilon}+e_{1}+f_{1}

    where SW denotes the extended, gauge-fixed Seiberg–Witten equations with respect to (gτ,Bτ)(g_{\tau},B_{\tau}), and e1,f1e_{1},f_{1} are error terms obeying

    • (1)

      e1Γ(SRe)e_{1}\in\Gamma(S^{\text{Re}}), and supp(e1)supp(dχ+)\text{supp}(e_{1})\subseteq\text{supp}(d\chi^{+})

    • (2)

      e1L2(Y)Cε1/24+γ.\|e_{1}\|_{L^{2}(Y)}\leq C\varepsilon^{-1/24+\gamma}.

    • (3)

      f1L2(Y)CεM\|f_{1}\|_{L^{2}(Y)}\leq C\varepsilon^{M} for M>10M>10, and supp(f1)supp(dχ+)\text{supp}(f_{1})\subseteq\text{supp}(d\chi^{+}).

    for γ<<1\gamma<<1. Moreover, the derivatives τe1,τf1\partial_{\tau}e_{1},\partial_{\tau}f_{1} also satisfy (1)–(3).

  2. (II)

    There is a configuration (φ,a)Hε,01,+(\varphi,a)\in H^{1,+}_{\varepsilon,0} such that

    (Φ1,A1)=(Φτhε,Aτhε)+χ+(φ,a)(\Phi_{1},A_{1})=(\Phi^{h_{\varepsilon}}_{\tau},A^{h_{\varepsilon}}_{\tau})\ +\ \chi^{+}\cdot(\varphi,a)

    where

    (φ,a)Hε,01,++τ(φ,a)Hε,01,+Cε1/12γ.\|(\varphi,a)\|_{H^{1,+}_{\varepsilon,0}}\ +\ \|\partial_{\tau}(\varphi,a)\|_{H^{1,+}_{\varepsilon,0}}\leq C\varepsilon^{-1/12-\gamma}. (7.6)
  3. (III)

    The L2L^{2}-norm satisfies Φ1εL2(Y)=1ε+o(1)\|\tfrac{\Phi_{1}}{\varepsilon}\|_{L^{2}(Y)}=\tfrac{1}{\varepsilon}\ +\ o(1).

Remark 7.6.

The de-singularization process used to obtain (Φτhε,Aτhε)(\Phi^{h_{\varepsilon}}_{\tau},A^{h_{\varepsilon}}_{\tau}) smoothes the 2\mathbb{Z}_{2}-harmonic spinor and the accompanying singular connection AτA_{\tau}, thereby re-introducing a highly concentrated “bubble” of curvature. Briefly, these de-singularized configurations are exponentially close to (Φτ,Aτ)(\Phi_{\tau},A_{\tau}) outside a neighborhood of radius r=cε2/3r=c\varepsilon^{2/3}, and for r<cε2/3r<c\varepsilon^{2/3} may be characterized as follows. Φτhε\Phi^{h_{\varepsilon}}_{\tau} is smooth and non-vanishing with |Φτhε|>cε1/3|\Phi^{h_{\varepsilon}}_{\tau}|>c\varepsilon^{1/3}, and AτhεA^{h_{\varepsilon}}_{\tau} is smooth and its connection form in the trivialization (3.10) vanishes at r=0r=0. The curvature FAhεF_{A^{h_{\varepsilon}}} is smooth with C0C^{0}-norm of size O(ε4/3)O(\varepsilon^{-4/3}), and L2L^{2}-norm of size O(ε2/3)O(\varepsilon^{-2/3}). See [41, Fig.1, pg.30]. The model solutions (Φ1,A1)(\Phi_{1},A_{1}) are a small perturbations a as in Item (II) and do not disrupt this qualitative behavior.

The next theorem gives a precise statement of the invertibility of the linearization at the approximate solutions from Theorem 7.5. Let (Φ1,A1)\mathcal{L}_{(\Phi_{1},A_{1})} denote the (extended, gauge-fixed) linearization at the model solutions (recall from Section 4.1 that this means the linearization at the un-renormalized spinor with norm O(ε1)O(\varepsilon^{-1})), and let ±\mathcal{L}^{\pm} denote its restrictions to Y±Y^{\pm}.

Theorem 7.7.

([41], Theorems 1.4 & 7.1) For any ν[0,14)\nu\in\left[0,\tfrac{1}{4}\right),

(Φ1,A1)+:Hε,ν1,+(Y+)Lε,ν2,+(Y+)\mathcal{L}^{+}_{(\Phi_{1},A_{1})}:H^{1,+}_{\varepsilon,\nu}(Y^{+})\longrightarrow L^{2,+}_{\varepsilon,\nu}(Y^{+}) (7.7)

is Fredholm of Index 0. Moreover, there exists an ε0\varepsilon_{0} such that for ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), it is invertible, and there is a constant CνC_{\nu} such that the bound

(φ,a)Hε,ν1,+Cνε1/12+γ(Φ1,A1)+(φ,a)Lε,ν2,+\|(\varphi,a)\|_{H^{1,+}_{\varepsilon,\nu}}\leq\frac{C_{\nu}}{\varepsilon^{1/12+\gamma^{\prime}}}\|\mathcal{L}^{+}_{(\Phi_{1},A_{1})}(\varphi,a)\|_{L^{2,+}_{\varepsilon,\nu}} (7.8)

holds uniformly for τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}) where γ=23(14ν)+γν\gamma^{\prime}=\tfrac{2}{3}\left(\tfrac{1}{4}-\nu\right)+\gamma\nu. Moreover, τ(Φ1,A1)+\partial_{\tau}\mathcal{L}^{+}_{(\Phi_{1},A_{1})} is uniformly bounded on the spaces (7.7). ∎

Note in particular for ν+=14106\nu^{+}=\tfrac{1}{4}-10^{-6} is as in Section 2, then γ<<1\gamma^{\prime}<<1.

7.3. Outside Boundary Conditions

This subsection departs from [41] to discuss the linear theory for the Seiberg–Witten equations in YY^{-}. Recall from Lemma 4.2 that the linearized Seiberg–Witten equations at the limiting 2\mathbb{Z}_{2}-harmonic spinor take the form

(Φ0,A0)(φ1,φ2,a)=(A0000A0γ(_)Φ0ε0μ(_,Φ0)ε.𝕕)(φ1φ2a).\mathcal{L}_{(\Phi_{0},A_{0})}(\varphi_{1},\varphi_{2},a)=\begin{pmatrix}\not{D}_{A_{0}}&0&0\\ 0&\not{D}_{A_{0}}&\gamma(\_)\frac{\Phi_{0}}{\varepsilon}\\ 0&\frac{\mu(\_,\Phi_{0})}{\varepsilon}.&\mathbb{d}\end{pmatrix}\begin{pmatrix}\varphi_{1}\\ \varphi_{2}\\ a\end{pmatrix}. (7.9)

where φ=(φ1,φ2)SReSIm\varphi=(\varphi_{1},\varphi_{2})\in S^{\text{Re}}\oplus S^{\text{Im}}. In particular, the real component decouples from the imaginary and form components. The top left block is the operator that was studied in Section 4; the bottom block is a copy of the standard (i.e. single spinor) SW equations, can be reduced to standard elliptic theory by viewing it as a boundary-value problem using an adaptation of the boundary conditions of Lemma 7.9.

First, define the outside Sobolev norms with boundary conditions complementary to those of Definition 7.1. Let λ=ε2/3γ\lambda^{-}=\varepsilon^{2/3-\gamma}, and set Y=YNλ(𝒵τ)Y^{-}=Y-N_{\lambda^{-}}(\mathcal{Z}_{\tau}).

Definition 7.8.

Let ν\nu\in\mathbb{R} be a weight. The “outside” Sobolev norms are defined by

(φ1,φ2,a)Hε,ν1,\displaystyle\|(\varphi_{1},\varphi_{2},a)\|_{H^{1,-}_{\varepsilon,\nu}} :=\displaystyle:= (φ1r1νHe1(Y)2+Y(|φ2|2+|a|2+|φ2|2|Φ|2ε2+|a|2|Φ|2ε2)Rε2ν𝑑V)1/2\displaystyle\left(\|\varphi_{1}\|_{r^{1-\nu}H^{1}_{e}(Y)}^{2}\ +\ \int_{Y^{-}}\left(|\nabla\varphi_{2}|^{2}+|\nabla a|^{2}+\frac{|\varphi_{2}|^{2}|\Phi|^{2}}{\varepsilon^{2}}+\frac{|a|^{2}|\Phi|^{2}}{\varepsilon^{2}}\right)R_{\varepsilon}^{2\nu}\ dV\right)^{1/2}
(φ1,φ2,a)Lε,ν2,\displaystyle\|(\varphi_{1},\varphi_{2},a)\|_{L^{2,-}_{\varepsilon,\nu}} :=\displaystyle:= (φ1rνL2(Y)2+Y(|φ|2+|a|2)Rε2ν𝑑V)1/2\displaystyle\left(\|\varphi_{1}\|^{2}_{r^{-\nu}L^{2}(Y)}\ +\ \int_{Y^{-}}\left(|\varphi|^{2}+|a|^{2}\right)R_{\varepsilon}^{2\nu}\ dV\right)^{1/2}

where the dependence of Φτ,dV\Phi_{\tau},dV, and RεR_{\varepsilon} on τ\tau is suppressed in the notation. Here, =Aτ\nabla=\nabla_{A_{\tau}}.

Notice this norm is not, technically speaking, a norm on sections of a vector bundle over the manifold YY; in particular the section φ1\varphi_{1} in the real component is integrated over all Y𝒵τY-\mathcal{Z}_{\tau} whereas the other two components only over YY^{-}. Notice also the difference in the sign convention for the weight between these spaces and those of Section 4 (which adopts the conventions standard for edge operators); in particular rνL2=Lε,ν2r^{\nu}L^{2}=L^{2}_{\varepsilon,-\nu}. Finally, observe that |φ2||Φ|2=|μ(φ,Φ)|2|\varphi_{2}||\Phi|^{2}=|\mu(\varphi,\Phi)|^{2} since ΦSRe\Phi\in S^{\text{Re}} and that RεrR_{\varepsilon}\sim r on YY^{-}; thus this norm is equivalent to the Hε,ν1,+H^{1,+}_{\varepsilon,\nu}-norm on Y+YY^{+}\cap Y^{-} (since ΦΦhε\Phi-\Phi^{h_{\varepsilon}} is exponentially small there).

The following lemma describes the boundary conditions on (φ2,a)(\varphi_{2},a); no boundary condition is imposed on the φ1\varphi_{1}-component. In it, (Φ,A)Im\mathcal{L}^{\text{Im}}_{(\Phi,A)} denotes the linearized Seiberg-Witten equations at a smooth configuration (Φ,A)(\Phi,A) acting on tuples (0,φ2,a)(0,\varphi_{2},a).

Lemma 7.9.

There exists a Hilbert subspace

HΦτL1/2,2(Y;SImΩ)H^{-}_{\Phi_{\tau}}\subseteq L^{1/2,2}(\partial Y^{-};S^{\text{Im}}\oplus\Omega)

and a projection operator Π:L1,2(Y;SImΩ)HΦτ\Pi^{-}:L^{1,2}(Y^{-};S^{\text{Im}}\oplus\Omega)\to H^{-}_{\Phi_{\tau}} given by the composition of restriction to the boundary and the L2L^{2}-orthogonal projection such that

((Φ,A)Im,Π):L1,2(Y,SImΩ)L2(Y;SImΩ)HΦτ(\mathcal{L}^{\text{Im}}_{(\Phi,A)}\ ,\ \Pi^{-}):L^{1,2}(Y^{-},S^{\text{Im}}\oplus\Omega)\longrightarrow L^{2}(Y^{-};S^{\text{Im}}\oplus\Omega)\oplus H^{-}_{\Phi_{\tau}} (7.10)

is Fredholm of Index 0.∎

Definition 7.10.

Define the “outside” Hilbert spaces by

Hε,ν1,\displaystyle H^{1,-}_{\varepsilon,\nu} :=\displaystyle:= {(φ1,φ2,a)|(φ,a)Hε,ν1,<,Π(φ,a)=0}\displaystyle\Big{\{}(\varphi_{1},\varphi_{2},a)\ \big{|}\ \|(\varphi,a)\|_{H^{1,-}_{\varepsilon,\nu}}<\infty\ \ ,\ \ \Pi^{-}(\varphi,a)=0\Big{\}}
Lε,ν2,\displaystyle L^{2,-}_{\varepsilon,\nu} :=\displaystyle:= {(φ1,φ2,a)|(φ,a)Lε,ν2,<}.\displaystyle\Big{\{}(\varphi_{1},\varphi_{2},a)\ \big{|}\ \|(\varphi,a)\|_{L^{2,-}_{\varepsilon,\nu}}<\infty\Big{\}}.

They are equipped with the norms from Definition 7.1, and the inner products arising from their polarizations. They depend implicitly on τ\tau.

The following lemma establishes the invertibility of the bottom block matrix in (7.9), which we continue denote by (Φτ,Aτ)Im\mathcal{L}^{\text{Im}}_{(\Phi_{\tau},A_{\tau})}.

Lemma 7.11.

For 12<ν<12-\tfrac{1}{2}<\nu<\tfrac{1}{2}, the boundary-value problem

(Φτ,Aτ)Im:Hε,ν1,(Y;SImΩ)Lε,ν2,(Y;SImΩ)\mathcal{L}^{\text{Im}}_{(\Phi_{\tau},A_{\tau})}:H^{1,-}_{\varepsilon,\nu}(Y^{-};S^{\text{Im}}\oplus\Omega)\longrightarrow L^{2,-}_{\varepsilon,\nu}(Y^{-};S^{\text{Im}}\oplus\Omega) (7.11)

is invertible for ε0\varepsilon_{0} sufficiently small. Moreover, the estimate

(φ2,a)Hε,ν1,C(Φτ,Aτ)Im(φ2,a)Lε,ν2,\|(\varphi_{2},a)\|_{H^{1,-}_{\varepsilon,\nu}}\leq C\|\mathcal{L}_{(\Phi_{\tau},A_{\tau})}^{\text{Im}}(\varphi_{2},a)\|_{L^{2,-}_{\varepsilon,\nu}}

holds uniformly in ε,τ\varepsilon,\tau, and τ(Φτ,Aτ)Im\partial_{\tau}\mathcal{L}^{\text{Im}}_{(\Phi_{\tau},A_{\tau})} is uniformly bounded on the spaces (7.11). ∎

Proof Sketch.

The Weitzenböck formula ([41, Prop 2.13]) shows that

ImIm(φ2a)=(AτAτφ2𝕕𝕕a)+1ε2(γ(μ(φ2,Φτ))Φτμ(γ(a)Φτ,Φτ)))+1ε𝔅(φ2,a)\mathcal{L}^{\text{Im}}\mathcal{L}^{\text{Im}}\begin{pmatrix}\varphi_{2}\\ a\end{pmatrix}=\begin{pmatrix}\not{D}_{A_{\tau}}\not{D}_{A_{\tau}}\varphi_{2}\\ \mathbb{d}\mathbb{d}a\end{pmatrix}+\frac{1}{\varepsilon^{2}}\begin{pmatrix}\gamma(\mu(\varphi_{2},\Phi_{\tau}))\Phi_{\tau}\\ \mu(\gamma(a)\Phi_{\tau},\Phi_{\tau}))\end{pmatrix}+\frac{1}{\varepsilon}\mathfrak{B}(\varphi_{2},a) (7.12)

where 𝔅(φ2,a)\mathfrak{B}(\varphi_{2},a) schematically has terms of the form aAτΦτa\cdot\nabla_{A_{\tau}}\Phi_{\tau} and φAτΦτ\varphi\cdot\nabla_{A_{\tau}}\Phi_{\tau}. Taking the inner product of (7.12) with (φ2,a)(\varphi_{2},a) shows that

Im(φ2,a)L22=(φ2,a)Hε1,2+1ε(φ2,a),𝔅(φ2,a)+b.d. terms.\|\mathcal{L}^{\text{Im}}(\varphi_{2},a)\|^{2}_{L^{2}}=\|(\varphi_{2},a)\|^{2}_{H^{1,-}_{\varepsilon}}\ +\ \frac{1}{\varepsilon}\langle(\varphi_{2},a),\mathfrak{B}(\varphi_{2},a)\rangle\ +\ \text{b.d. terms}.

Using that Φτr1/2\Phi_{\tau}\sim r^{1/2} while Φτ,Aτr1/2\nabla\Phi_{\tau},A_{\tau}\sim r^{-1/2}, the terms involving 𝔅\mathfrak{B} and AτA_{\tau} are dominated by the ε2|Φτ|2\varepsilon^{-2}|\Phi_{\tau}|^{2} weight in (7.8) on YY^{-} where rε2/3γr\geq\varepsilon^{2/3-\gamma}, and may be absorbed. The remainder of the proof is showing that the subspace of Lemma 7.9 may be chosen so that the boundary terms vanish (this is non-trivial because of factor of eiθe^{-i\theta} in the local expression (4.13), but is similar to the arguments in [41, Section 7]). ∎

7.4. Exponential and Polynomial Decay

Solutions of

(Φ1,A1)+(φ,a)\displaystyle\mathcal{L}_{(\Phi_{1},A_{1})}^{+}(\varphi,a)\ =\displaystyle= g+\displaystyle g^{+} (7.13)

decay away from the support of g+g^{+}. This section gives precise decay results that provide the operative version of requirement (II) from Section 2.2. On Y+Y^{+}, real and imaginary/form components manifest two different types decay, respectively: (1) polynomial decay, akin to that of Lemma 4.7 and (2) exponential decay with exponent ε1\varepsilon^{-1}.

This next lemma address the polynomial decay. An analogous result holds for solutions of the equation in Lemma (7.11), but is not needed.

Lemma 7.12.

Suppose (φ,a)(\varphi,a) is the unique solutions of (7.13). If supp(g+){rcε2/3γ}\text{supp}(g^{+})\subseteq\{r\leq c\varepsilon^{2/3-\gamma}\}, then

dχ+.(φ,a)L2Cε1/24γgL2.\|d\chi^{+}.(\varphi,a)\|_{L^{2}}\leq C\varepsilon^{-1/24-\gamma}\|g\|_{L^{2}}.
Proof.

The proof follows an identical strategy to that of Lemma 4.7, with the following minor adjustments. First, Theorem 7.7 only applies for weights ν[0,14)\nu\in[0,\tfrac{1}{4}), the factor gained in the decay is reduced by half to ε1/24γ\varepsilon^{1/24-\gamma}. Second, the lack of uniform invertibility in (7.8) leads to an adverse factor of ε1/12γ\varepsilon^{-1/12-\gamma}. Multiplying these and redefining γ\gamma yields the assertion of the lemma. ∎

One could (perhaps justifiably) gripe that Lemma 7.12 does not, at first glance, constitute a “decay” result because the decay factor diverges as ε0\varepsilon\to 0. The point, however, is that this result still tempers that growth compared to the factor expected from (7.8), and does so sufficiently much that when combined with Lemma 4.7 the power of ε\varepsilon for a full cycle of (2.17) is positive.

The imaginary/form components of a solution to (7.13) obey a far stronger decay property. That this is so is a consequence of the intrinsic structure of generalized Seiberg–Witten equations, which allows them to be interpreted as a non-linear concentrating Dirac operator with degeneracy. This interpretation is developed extensively in [42], which establishes decay results in a quite general setting.

The following lemma specializes these results to the present case; it is proved in Appendix A of [42]. For the statement of the lemma, let KεY+𝒵τK_{\varepsilon}\Subset Y^{+}-\mathcal{Z}_{\tau} denote a family of compact subsets of the complement of the singular set. Set rK:=dist(Kε,𝒵τ)r_{K}:=\text{dist}(K_{\varepsilon},\mathcal{Z}_{\tau}). Let KεK^{\prime}_{\varepsilon} be a slightly larger family of compact sets so Y+NrK/2(𝒵τ)KεY^{+}-N_{r_{K}/2}(\mathcal{Z}_{\tau})\subseteq K_{\varepsilon}^{\prime}.

Lemma 7.13.

Let (φ,a)(\varphi,a) be the unique solution to (7.13), and write (φ,a)=(φ1,φ2,a)(\varphi,a)=(\varphi_{1},\varphi_{2},a) on Y𝒵τY-\mathcal{Z}_{\tau}. There exist constants C,cC,c such that if supp(g)(Kε)c\text{supp}(g)\subseteq(K_{\varepsilon}^{\prime})^{c}, then SImΩS^{\text{Im}}\oplus\Omega-components satisfy

(φ2,a)Cm(Kε)Cε2m+1rK3/2Exp(crK3/2ε)(φ,a)Hε,01,+\|(\varphi_{2},a)\|_{C^{m}(K_{\varepsilon})}\leq\frac{C}{\varepsilon^{2m+1}r_{K}^{3/2}}\text{Exp}\left(-\frac{cr_{K}^{3/2}}{\varepsilon}\right)\|(\varphi,a)\|_{H^{1,+}_{\varepsilon,0}} (7.14)

uniformly for τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}). Moreover, the configuration denoted by (φ,a)(\varphi,a) in Item (II) of Theorem 7.5 obeys the above.

Proof.

Both statements follow directly from Corollary A.2 of [42], which includes a similar conclusion for solutions of non-linear equations of the form in Theorem 7.5. ∎

8. Universal Seiberg–Witten Equations

This section uses the concentrating local solutions defined in Section 7 to construct an infinite-dimensional family of model solutions parameterized by deformations of the singular sets 𝒵τ\mathcal{Z}_{\tau}. This family is used to define a universal version of the Seiberg–Witten equations akin to the universal Dirac operator (1.10).

8.1. Hilbert Bundles

This subsection defines Hilbert bundles of Seiberg–Witten configurations over the space of embedded singular sets, analogously to Section 6.1.

Let τL2,2(𝒵τ;N𝒵τ)\mathcal{E}_{\tau}\subseteq L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) denote the ball on which Exp¯\underline{\text{Exp}} is a diffeomorphism as in Section 6.3. For each ξτ\xi\in\mathcal{E}_{\tau}, let F¯ξ\underline{F}_{\xi}, and 𝒵ξ,τ\mathcal{Z}_{\xi,\tau} be as in (6.3)–(6.4), and define

Yε,τ,ξ+:=F¯ξ[Nλ(𝒵τ)]Yε,τ,ξ:=YF¯ξ[Nλ(𝒵τ)]Y^{+}_{\varepsilon,\tau,\xi}:=\underline{F}_{\xi}[N_{\lambda}(\mathcal{Z}_{\tau})]\hskip 56.9055ptY^{-}_{\varepsilon,\tau,\xi}:=Y-\underline{F}_{\xi}[N_{\lambda^{-}}(\mathcal{Z}_{\tau})]

where λ,λ\lambda,\lambda^{-} are as in (7.1) and Definition 7.8 respectively. For each triple (ε,τ,ξ)(\varepsilon,\tau,\xi) there are weight functions defined analogously to those in Section 7.1:

Rε,τ,ξ\displaystyle R_{\varepsilon,\tau,\xi} :=\displaystyle:= RεF¯ξ1\displaystyle R_{\varepsilon}\circ\underline{F}^{-1}_{\xi} (8.1)
Φτ,ξhε\displaystyle\Phi_{\tau,\xi}^{h_{\varepsilon}} :=\displaystyle:= Υ¯1(ξ,Φτhε)\displaystyle\underline{\Upsilon}^{-1}(\xi,\Phi^{h_{\varepsilon}}_{\tau}) (8.2)

where RεR_{\varepsilon} is the weight from Item (1) in Section 7.1 (which depends implicitly on τ\tau), and Υ¯\underline{\Upsilon} is the analogue of 6.1 defined in the proof of Lemma 8.2 below.

Definition 8.1.

Let ν\nu\in\mathbb{R} be a weight. For each (ε,τ)(0,ε0)×(τ0,τ0)(\varepsilon,\tau)\in(0,\varepsilon_{0})\times(-\tau_{0},\tau_{0}) and each ξτ\xi\in\mathcal{E}_{\tau}, define Hilbert spaces

Hε,ν1,±(Yε,τ,ξ±)\displaystyle H^{1,\pm}_{\varepsilon,\nu}(Y^{\pm}_{\varepsilon,\tau,\xi}) :=\displaystyle:= {(φ,a)|(φ,a)Hε,ν1,±<}\displaystyle\Big{\{}\ \ \ \ \ (\varphi,a)\ \ \ \ \ \big{|}\ \ \ \ \ \ \ \|(\varphi,a)\|_{H^{1,\pm}_{\varepsilon,\nu}}<\infty\ \ \ \Big{\}}
Lε,ν2,±(Yε,τ,ξ±)\displaystyle L^{2,\pm}_{\varepsilon,\nu}(Y^{\pm}_{\varepsilon,\tau,\xi}) :=\displaystyle:= {(φ,a)|(φ,a)Lε,ν2,±<}\displaystyle\Big{\{}\ \ \ \ \ (\varphi,a)\ \ \ \ \ \big{|}\ \ \ \ \ \ \ \|(\varphi,a)\|_{L^{2,\pm}_{\varepsilon,\nu}}<\infty\ \ \ \ \Big{\}}

where the norms are defined analogously to Definitions 7.3 and 7.10 using the domains indicated and the weights (8.18.2). For each (ε,τ)(\varepsilon,\tau), denote the τ\mathcal{E}_{\tau}-parameterized families by

ε,ν1,±(τ)\displaystyle\mathbb{H}^{1,\pm}_{\varepsilon,\nu}(\mathcal{E}_{\tau}) :=\displaystyle:= {(𝒵ξ,τ,φ,a)|ξτ,(φ,a)Hε,ν1,±(Yε,τ,ξ±)}\displaystyle\Big{\{}(\mathcal{Z}_{\xi,\tau},\varphi,a)\ \ \ \hskip 11.38092pt|\ \ \xi\in\mathcal{E}_{\tau}\ ,\ (\varphi,a)\in H^{1,\pm}_{\varepsilon,\nu}(Y^{\pm}_{\varepsilon,\tau,\xi})\ \ \Big{\}} (8.3)
𝕃ε,ν2,±(τ)\displaystyle\mathbb{L}^{2,\pm}_{\varepsilon,\nu}(\mathcal{E}_{\tau}) :=\displaystyle:= {(𝒵ξ,τ,φ,a)|ξτ,(φ,a)Lε,ν2,±(Yε,τ,ξ±)}.\displaystyle\Big{\{}(\mathcal{Z}_{\xi,\tau},\varphi,a)\ \ \ \hskip 11.38092pt|\ \ \xi\in\mathcal{E}_{\tau}\ ,\ (\varphi,a)\in L^{2,\pm}_{\varepsilon,\nu}(Y^{\pm}_{\varepsilon,\tau,\xi})\ \ \ \Big{\}}. (8.4)

Finally, define global spaces by

ε,ν1:=ε,ν1,+(τ)ε,ν1,(τ)𝕃ε,ν2:={(𝒵ξ,τ,φ,a)|ξτ,(φ,a)Lε,ν2(Y)<}\mathbb{H}^{1}_{\varepsilon,\nu}:=\mathbb{H}^{1,+}_{\varepsilon,\nu}(\mathcal{E}_{\tau})\oplus\mathbb{H}^{1,-}_{\varepsilon,\nu}(\mathcal{E}_{\tau})\hskip 56.9055pt\mathbb{L}^{2}_{\varepsilon,\nu}:=\Big{\{}(\mathcal{Z}_{\xi,\tau},\varphi,a)\ |\ \xi\in\mathcal{E}_{\tau}\ ,\ \|(\varphi,a)\|_{L^{2}_{\varepsilon,\nu}(Y)}<\infty\Big{\}} (8.5)

where the Lε,ν2(Y)L^{2}_{\varepsilon,\nu}(Y)-norm is defined identically to the Lε,ν2,+(Y+)L^{2,+}_{\varepsilon,\nu}(Y^{+})-norm in Definition 7.1, but integrated over all YY.

The following lemma shows that pulling back by the family of diffeomorphisms F¯ξ\underline{F}_{\xi} trivializes these vector bundles.

Lemma 8.2.

The family of diffeomorphisms F¯ξ\underline{F}_{\xi} determines trivializations

Υ¯:ε,ν1,±(τ)\displaystyle\underline{\Upsilon}:\mathbb{H}^{1,\pm}_{\varepsilon,\nu}(\mathcal{E}_{\tau}) \displaystyle\longrightarrow τ×Hε,ν1,±(Yε,τ,0±)\displaystyle\mathcal{E}_{\tau}\ \times\ H^{1,\pm}_{\varepsilon,\nu}(Y^{\pm}_{\varepsilon,\tau,0}) (8.6)
Υ¯:𝕃ε,ν2,±(τ)\displaystyle\underline{\Upsilon}:\mathbb{L}^{2,\pm}_{\varepsilon,\nu}(\mathcal{E}_{\tau}) \displaystyle\longrightarrow τ×Lε,ν2,±(Yε,τ,0±)\displaystyle\mathcal{E}_{\tau}\ \times\ L^{2,\pm}_{\varepsilon,\nu}(Y^{\pm}_{\varepsilon,\tau,0}) (8.7)

which endow the spaces on the left with the structure of smooth Hilbert vector bundles. The same applies to 𝕃ε,ν2,ε,ν1\mathbb{L}^{2}_{\varepsilon,\nu},\mathbb{H}^{1}_{\varepsilon,\nu}.

Proof Sketch.

(See [43, Sec. 5.1] for further details). Fix ε,τ\varepsilon,\tau. To begin, we define Υ¯\underline{\Upsilon} on the spinor components. For each ξτ\xi\in\mathcal{E}_{\tau}, the Υ¯ξ\underline{\Upsilon}_{\xi} is the map induced on sections by a fiberwise isomorphism υ¯ξ:(SE)y1(SE)y2\underline{\upsilon}_{\xi}:(S_{E})_{y_{1}}\to(S_{E})_{y_{2}} where y1=F¯ξ(y2)y_{1}=\underline{F}_{\xi}(y_{2}).

For clarity, fix a spin structure with spinor bundle SgτS_{g_{\tau}} and a complex line bundle LL so that the spinor bundle of the Spinc\text{Spin}^{c}-structure is S=SgτLS=S_{g_{\tau}}\otimes_{\mathbb{C}}L. υ¯ξ\underline{\upsilon}_{\xi} is defined as the composition of the following three maps:

SgτLES_{g_{\tau}}\otimes L\otimes ESgξF¯ξ(LE)S_{g_{\xi}}\otimes\underline{F}_{\xi}^{*}(L\otimes E)F¯ξ(Sg0LE)\underline{F}_{\xi}^{*}(S_{g_{0}}\otimes L\otimes E)SgτLES_{g_{\tau}}\otimes L\otimes E𝔗gξgτ\mathfrak{T}_{g_{\xi}}^{g_{\tau}}𝔖ξ\mathfrak{S}_{\xi}F¯ξ\underline{F}_{\xi}^{*}

where

  • F¯ξ\underline{F}_{\xi}^{*} is the pullback by the diffeomorphism F¯ξ\underline{F}_{\xi}.

  • 𝔖ξ\mathfrak{S}_{\xi} is the canonical isomorphism F¯ξ(Sgτ)Sgξ\underline{F}_{\xi}^{*}(S_{g_{\tau}})\simeq S_{g_{\xi}} on the first factor, and the identity on LEL\otimes E.

  • 𝔗gξgτ\mathfrak{T}_{g_{\xi}}^{g_{\tau}} is the parallel transport map from the proof of Proposition (6.3).

In the third bullet, parallel transport over Z=Y×[0,1]Z=Y\times[0,1] on the bundles SgsξS_{g_{s\xi}}, F¯sξ(LE)\underline{F}_{s\xi}^{*}(L\otimes E) using the connections F¯sξBτ\underline{F}_{s\xi}^{*}B_{\tau} and F¯sξA\underline{F}_{s\xi}^{*}A_{\circ} on the second and third factors where AA_{\circ} is a fixed smooth background connection on LL.

The definition of Υ¯\underline{\Upsilon} on the forms component is identical with the simplification that the bundle of forms is canonically identified with its pullback so the second map 𝔖ξ\mathfrak{S}_{\xi} is not necessary. ∎

8.2. Concentrating Local Families

This subsection defines the universal Seiberg–Witten equations as sections of the vector bundles from Section 8.1. The equations are viewed as deformation equations around a universal family of concentrating approximate solutions.

Definition 8.3.

For τ(τ0,τ0)\tau\in(-\tau_{0},\tau_{0}) and ξτ\xi\in\mathcal{E}_{\tau}, define

  1. (A)

    The universal family of concentrating local solutions by

    (Φ1(ε,τ,ξ)ε,A1(ε,τ,ξ)):=Υ¯1(ξ,(Φ1(ε,τ)ε,A1(ε,τ)))\left(\frac{\Phi_{1}({\varepsilon,\tau,\xi})}{\varepsilon},A_{1}({\varepsilon,\tau,\xi})\right):=\underline{\Upsilon}^{-1}\left(\xi,\left(\frac{\Phi_{1}(\varepsilon,\tau)}{\varepsilon},A_{1}(\varepsilon,\tau)\right)\right) (8.8)

    i.e. as the pullback by Υ¯\underline{\Upsilon} of the constant section at the approximate solutions (Φ1(ε,τ),A1(ε,τ))(\Phi_{1}(\varepsilon,\tau),A_{1}(\varepsilon,\tau)) of Theorem 7.5.

  2. (B)

    The universal families of de-singularized configurations and 2\mathbb{Z}_{2}-eigenvectors by

    (Φτ,ξhε,Aτ,ξhε):=Υ¯1(ξ,(Φτhε,Aτhε))(Φτ,ξ,Aτ,ξ):=Υ¯1(ξ,(Φτ,Aτ))\left(\Phi^{h_{\varepsilon}}_{\tau,\xi},A^{h_{\varepsilon}}_{\tau,\xi}\right):=\underline{\Upsilon}^{-1}\left(\xi,(\Phi^{h_{\varepsilon}}_{\tau},A^{h_{\varepsilon}}_{\tau})\right)\hskip 42.67912pt\left(\Phi_{\tau,\xi},A_{\tau,\xi}\right):=\underline{\Upsilon}^{-1}\left(\xi,(\Phi_{\tau},A_{\tau})\right) (8.9)

    using de-singularized configurations (Φτhε,Aτhε)(\Phi^{h_{\varepsilon}}_{\tau},A^{h_{\varepsilon}}_{\tau}) and the 2\mathbb{Z}_{2}-eigenvectors (Φτ,Aτ)(\Phi_{\tau},A_{\tau}) respectively.

The family of de-singularized configurations was already used as the weight in definition (8.2). Strictly speaking, Υ¯\underline{\Upsilon} is defined to act on forms a(Ω0Ω1)(i)a\in(\Omega^{0}\oplus\Omega^{1})(i\mathbb{R}) rather than connections. It is easy to verify, via the definition of Υ¯\underline{\Upsilon} in the proof of Lemma 8.2 that a connection can be pulled back equally well. By construction, (8.8) have curvature highly concentrated around the deformed curves 𝒵ξ,τ\mathcal{Z}_{\xi,\tau}.

For each triple (ε,τ,ξ)(\varepsilon,\tau,\xi), there are accompanying cut-off functions χε,τ,ξ±:=χ±F¯ξ\chi^{\pm}_{\varepsilon,\tau,\xi}:=\chi^{\pm}\circ\underline{F}_{\xi}. To simplify notation, the latter two subscripts are omitted when they are clear in context. Consider the deformation equation of the Seiberg–Witten equations at the approximate solution (8.8), given by

SWε,τ,ξ(φ,a,ψ,b):=SWτ((Φ1(ε,τ,ξ)ε,A1(ε,τ,ξ))+χε+(φ,a)+χε(ψ,b))\text{SW}_{\varepsilon,\tau,\xi}(\varphi,a,\psi,b):=\text{SW}_{\tau}\left(\left(\frac{\Phi_{1}({\varepsilon,\tau,\xi})}{\varepsilon},A_{1}({\varepsilon,\tau,\xi})\right)+\chi^{+}_{\varepsilon}\cdot(\varphi,a)+\chi^{-}_{\varepsilon}\cdot(\psi,b)\right) (8.10)

for (φ,a)(Hε1,+)ξ(\varphi,a)\in(H^{1,+}_{\varepsilon})_{\xi} and (ψ,b)(Hε1,)ξ(\psi,b)\in(H^{1,-}_{\varepsilon})_{\xi} in the fibers over ξτ\xi\in\mathcal{E}_{\tau}, where SWτ\text{SW}_{\tau} denotes the Seiberg–Witten equations with parameter pτp_{\tau}. Additionally, denote the projection by p:ε,ν1τp:\mathbb{H}^{1}_{\varepsilon,\nu}\to\mathcal{E}_{\tau}.

Definition 8.4.

For each (ε,τ)(0,ε0)×(τ0,τ0)(\varepsilon,\tau)\in(0,\varepsilon_{0})\times(-\tau_{0},\tau_{0}), the universal Seiberg–Witten equations (resp. eigenvector equation) are the sections

ε,ν1(τ)\mathbb{H}^{1}_{\varepsilon,\nu}({\mathcal{E}_{\tau}})p𝕃ε,ν2(τ)p^{*}\mathbb{L}^{2}_{\varepsilon,\nu}({\mathcal{E}_{\tau}})𝕊𝕎,𝕊𝕎¯\mathbb{SW},\overline{\mathbb{SW}}

defined by

𝕊𝕎(ξ,φ,a,ψ,b)\displaystyle\mathbb{SW}(\xi,\varphi,a,\psi,b)\ \ :=\displaystyle:= SWε,τ,ξ(φ,a,ψ,b)\displaystyle\text{SW}_{\varepsilon,\tau,\xi}(\varphi,a,\psi,b) (8.11)
𝕊𝕎¯(ξ,φ,a,ψ,b,μ)\displaystyle\overline{\mathbb{SW}}(\xi,\varphi,a,\psi,b,\mu) :=\displaystyle:= SWε,τ,ξ(φ,a,ψ,b)μχεΦτ,ξε\displaystyle\text{SW}_{\varepsilon,\tau,\xi}(\varphi,a,\psi,b)-\mu\chi^{-}_{\varepsilon}\frac{\Phi_{\tau,\xi}}{\varepsilon} (8.12)

where the domain of the latter also includes the trivial summand μ¯\mu\in\underline{\mathbb{R}}.

To be more precise, (8.10) means the extended, gauge-fixed Seiberg–Witten equations subject to the gauge-fixing condition

dciiζ,Φξ,τhεε=0,-d^{\star}c-i\frac{\langle i\zeta,{\Phi_{\xi,\tau}^{h_{\varepsilon}}}\rangle}{\varepsilon}=0, (8.13)

where (ζ,c)=(ϕ1,a1)+χ+(φ,a)+χ(ψ,b)(\zeta,c)=(\phi_{1},a_{1})+\chi^{+}(\varphi,a)+\chi^{-}(\psi,b) where (εϕ1,a1)=(Φ1Φhε,A1Ahε)(\varepsilon\phi_{1},a_{1})=(\Phi_{1}-\Phi^{h_{\varepsilon}},A_{1}-A^{h_{\varepsilon}}) for each ξ,τ\xi,\tau. For ξ=0\xi=0, this is the gauge-fixing condition coincides with the one used in the proof of Theorem 7.5.

The purpose of introducing the family is that one only expects the deformation needed to correct (8.8) to a true solution to be small only in the weighted spaces for the correct deformation ξ\xi. Ultimately, each equation in the universal family (8.11) is simply the Seiberg–Witten equations on YY using the parameters (gτ,Bτ)(g_{\tau},B_{\tau}) acting on spaces with varying weights. In particular, to solve the SW equation on YY, it suffices to solve it for any parameter ξ\xi.

Corollary 8.5.

Suppose that ξτC(𝒵τ;N𝒵τ)\xi\in\mathcal{E}_{\tau}\cap C^{\infty}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) and that gC(Y)g\in C^{\infty}(Y), then

𝕊𝕎τ(ξ,φ,a,ψ,b)=gSWτ(Φ,A)=g{\mathbb{SW}_{\tau}}(\xi,\varphi,a,\psi,b)=g\ \ \Leftrightarrow\ \ \text{SW}_{\tau}(\Phi,A)=g

where (Φ,A)(\Phi,A) is the configuration on the right side of (8.10), and SWτ\text{SW}_{\tau} is the Seiberg–Witten equation using the parameters (gτ,Bτ)(g_{\tau},B_{\tau}). In particular, (Φ,A)(\Phi,A) is smooth.

The equivalent statement holds for the eigenvector equations 𝕊𝕎¯\overline{\mathbb{SW}}.

Proof.

The first statement is immediate from Definition 8.4. By Definition 8.1, (φ,a)Hε,ν1,+(\varphi,a)\in H^{1,+}_{\varepsilon,\nu} and (ψ,b)Hε,ν1,(\psi,b)\in H^{1,-}_{\varepsilon,\nu} imply that (Φ,A)L1,2(Y)(\Phi,A)\in L^{1,2}(Y). Since the background parameter (gτ,Bτ)(g_{\tau},B_{\tau}) is smooth, elliptic bootstrapping shows that (Φ,A)(\Phi,A) is also smooth.

In the case of the eigenvector equation, χΦτC\chi^{-}\Phi_{\tau}\in C^{\infty} because Φτ\Phi_{\tau} is smooth away from the singular set 𝒵τ\mathcal{Z}_{\tau} where χ=0\chi^{-}=0. Since ξ\xi is smooth, χΦτ,ξ\chi^{-}\Phi_{\tau,\xi} is as well and elliptic bootstrapping applies again. ∎

The universal Seiberg–Witten equations can also be extended to a map

𝕊𝕎¯:ε,ν1(τ×𝒳τ)p1𝕃ε,ν2\overline{\mathbb{SW}}:\mathbb{H}^{1}_{\varepsilon,\nu}(\mathcal{E}_{\tau}\times\mathcal{X}_{\tau})\to p_{1}^{*}\mathbb{L}^{2}_{\varepsilon,\nu} (8.14)

where 𝒳τ\mathcal{X}_{\tau} is the subspace from Lemma 5.6, and the bundle is the pullback of ε,ν1\mathbb{H}^{1}_{\varepsilon,\nu} by the projection τ×𝒳ττ\mathcal{E}_{\tau}\times\mathcal{X}_{\tau}\to\mathcal{E}_{\tau}. The extended map is defined by replacing ψ\psi in (8.10) by ψ+u\psi+u for u𝒳τu\in\mathcal{X}_{\tau}. This map factors through (8.12) since the map 𝒳τHε,ν1,\mathcal{X}_{\tau}\hookrightarrow H^{1,-}_{\varepsilon,\nu} by uχuu\mapsto\chi^{-}u is an inclusion.

8.3. Universal Linearization

This section calculates of the derivative of the universal Seiberg–Witten equations with respect to the deformation parameter.

Let h1=(𝒵τ,0,0)ε,ν1h_{1}=(\mathcal{Z}_{\tau},0,0)\in\mathbb{H}^{1}_{\varepsilon,\nu}. The (fiberwise component of) the linearization is a map

d𝕊𝕎h1:L2,2(𝒵τ)Hε,ν1,+(Y+)Hε,ν1,(Y)L2(Y),\text{d}{\mathbb{SW}}_{h_{1}}:L^{2,2}(\mathcal{Z}_{\tau})\ \oplus\ H^{1,+}_{\varepsilon,\nu}(Y^{+})\ \oplus\ H^{1,-}_{\varepsilon,\nu}(Y^{-})\ \longrightarrow\ L^{2}(Y), (8.15)

where we have used the canonical splitting of Th1ε,ν1T_{h_{1}}\mathbb{H}^{1}_{\varepsilon,\nu} along the zero-section. The next two propositions calculate this derivative in two steps, giving the analogues of Proposition 6.3 and Theorem 6.4 for the Seiberg–Witten equations.

In the following proposition, recall that (gξ,Bξ):=F¯ξ(gτ,Bτ)(g_{\xi},B_{\xi}):=\underline{F}_{\xi}^{*}(g_{\tau},B_{\tau}). Remember also that the model solution (Φ1,A1)(\Phi_{1},A_{1}) depends implicitly on (ε,τ)(\varepsilon,\tau).

Proposition 8.6.

In the local trivializations of Lemma 8.2, the derivative (8.15) is given by

d𝕊𝕎h1(ξ,φ,a,ψ,b)=𝔅h1(ξ)+h1(χε+(φ,a)+χε(ψ,b))\text{d}{\mathbb{SW}}_{h_{1}}(\xi,\varphi,a,\psi,b)=\mathfrak{B}_{h_{1}}(\xi)\ +\ \mathcal{L}_{h_{1}}\left(\chi_{\varepsilon}^{+}(\varphi,a)\ +\ \chi_{\varepsilon}^{-}(\psi,b)\right) (8.16)

where

  • 𝔅h1\mathfrak{B}_{h_{1}} is defined by

    𝔅h1(ξ)=dds|s=0((𝔗gτgξ(s))1SWpξ(s)𝔗gτgξ(s))(Φ1ε,A1)\mathfrak{B}_{h_{1}}(\xi)=\frac{d}{ds}\Big{|}_{s=0}\left((\mathfrak{T}_{g_{\tau}}^{g_{\xi}(s)})^{-1}\circ\text{SW}_{p_{\xi}(s)}\circ\mathfrak{T}_{g_{\tau}}^{g_{\xi}(s)}\right)\left(\frac{\Phi_{1}}{\varepsilon},A_{1}\right) (8.17)

    where 𝔗gτgξ(s)\mathfrak{T}^{g_{\xi}(s)}_{g_{\tau}} is the parallel transport map from Lemma 8.2, and SWpξ(s)\text{SW}_{p_{\xi(s)}} denotes the Seiberg-Witten equations with parameter pξ(s)=(gsξ,Bsξ)p_{\xi}(s)=(g_{s\xi},B_{s\xi}).

  • h1\mathcal{L}_{h_{1}} is the linearized Seiberg-Witten equations at (Φ1,A1)(\Phi_{1},A_{1}), given by (4.4)

Proof.

(see also [43, Prop. 5.5]) For the duration of the proof, we suppress the dependence on τ,ε,ν\tau,\varepsilon,\nu from the notation. Choose a path

γ:(s0,s0)\displaystyle\gamma:(-s_{0},s_{0}) \displaystyle\to 1()\displaystyle\mathbb{H}^{1}(\mathcal{E})
s\displaystyle s \displaystyle\mapsto (𝒵ζ(s),𝔭(s))\displaystyle(\mathcal{Z}_{\zeta(s)},\mathfrak{p}(s))

such that γ(0)=h1\gamma(0)=h_{1}, where ζ(s)=sξ+O(s2)\zeta(s)=s\xi+O(s^{2}). We may write 𝔭(s)=Υ¯ζ(s)1𝔮(s)\mathfrak{p}(s)=\underline{\Upsilon}_{\zeta(s)}^{-1}\mathfrak{q}(s), where 𝔮(s)=(φs,as,ψs,bs)H1,+H1,\mathfrak{q}(s)=(\varphi_{s},a_{s},\psi_{s},b_{s})\in H^{1,+}\oplus H^{1,-}. Using Definition 8.4 and (8.10), and then substituting Definition 8.8, the derivative (8.16) is then given by

dds|s=0Υ¯ζ(s)𝕊𝕎(𝒵ζ(s),𝔭(s))\displaystyle\frac{d}{ds}\Big{|}_{s=0}\underline{\Upsilon}_{\zeta(s)}\circ{\mathbb{SW}}(\mathcal{Z}_{\zeta(s)},\mathfrak{p}(s)) =\displaystyle= dds|s=0Υ¯ζ(s)SW((Φ1(ζ(s))ε,A1(ζ(s)))+χ±𝔭(s))\displaystyle\frac{d}{ds}\Big{|}_{s=0}\underline{\Upsilon}_{\zeta(s)}\circ\text{SW}\left(\left(\frac{\Phi_{1}({\zeta(s)})}{\varepsilon},A_{1}({\zeta(s)})\right)+\chi^{\pm}\mathfrak{p}(s)\right)
=\displaystyle= dds|s=0Υ¯ζ(s)SWΥ¯ζ(s)1((Φ1ε,A1)+χ+(φs,as)+χ(ψs,bs))\displaystyle\frac{d}{ds}\Big{|}_{s=0}\underline{\Upsilon}_{\zeta(s)}\circ\text{SW}\circ\underline{\Upsilon}_{\zeta(s)}^{-1}\left(\left(\frac{\Phi_{1}}{\varepsilon},A_{1}\right)+\chi^{+}(\varphi_{s},a_{s})+\chi^{-}(\psi_{s},b_{s})\right)

where Υ¯\underline{\Upsilon} is used to denote the trivialization of both 1\mathbb{H}^{1} and 𝕃2\mathbb{L}^{2}.

(LABEL:tocalculate6.6) appears as the rightmost vertical arrow in the commuting diagram below. The diagram decomposes Υ¯=𝔗gζg𝔖F¯ζ\underline{\Upsilon}=\mathfrak{T}_{g_{\zeta}}^{g}\circ\mathfrak{S}\circ\underline{F}_{\zeta}^{*} as in the proof of Lemma 8.2. It also abbreviates H1=H1,+H1,H^{1}=H^{1,+}\oplus H^{1,-} and Ω=(Ω0Ω1)(i)\Omega=(\Omega^{0}\oplus\Omega^{1})(i\mathbb{R}), and pζ=(gζ(s),Bζ(s))p_{\zeta}=(g_{\zeta}(s),B_{\zeta(s)}), and also uses SEhS_{E}^{h} to denote the spinor bundle formed using the metric hh.

H1(SEgζΩ)H^{1}(S_{E}^{g_{\zeta}}\oplus\Omega)L2(SEgζΩ)L^{2}(S_{E}^{g_{\zeta}}\oplus\Omega)H1(SEgΩ)H^{1}(S_{E}^{g}\oplus\Omega)L2(SEgΩ)L^{2}(S_{E}^{g}\oplus\Omega)H1(SEgΩ)H^{1}(S_{E}^{g}\oplus\Omega)L2(SEgΩ)L^{2}(S_{E}^{g}\oplus\Omega)(varying gζfixed 𝒵)\begin{pmatrix}\text{varying }g_{\zeta}\\ \text{fixed }\mathcal{Z}\end{pmatrix}(varying 𝒵ζfixed g)\begin{pmatrix}\text{varying }\mathcal{Z}_{\zeta}\\ \text{fixed }g\end{pmatrix}(fixed Sgfixed 𝒵)\begin{pmatrix}\text{fixed }S^{g}\\ \text{fixed }\mathcal{Z}\end{pmatrix}SWpζ\text{SW}_{p_{\zeta}}(𝔗gζg)1(\mathfrak{T}_{g_{\zeta}}^{g})^{-1}𝔖ζF¯ζ\mathfrak{S}_{\zeta}\circ\underline{F}_{\zeta}^{*}𝕊𝕎\mathbb{SW}𝔖ζF¯ζ\mathfrak{S}_{\zeta}\circ\underline{F}_{\zeta}^{*}(𝔗gζg)1(\mathfrak{T}_{g_{\zeta}}^{g})^{-1}ΥζSWΥζ1\Upsilon_{\zeta}\text{SW}\Upsilon_{\zeta}^{-1}

Expressing (LABEL:tocalculate6.6) as a composition using the vertical middle arrow in the diagram, and writing the Seiberg-Witten equations near a configuration (Φ,A)(\Phi,A) as

SW((Φ,A)+(φ,a))=SW(Φ,A)+(Φ,A)(φ,a)+Q(φ,a),\text{SW}\left((\Phi,A)+(\varphi,a)\right)=\text{SW}(\Phi,A)+\mathcal{L}_{(\Phi,A)}(\varphi,a)+Q(\varphi,a), (8.19)

the derivative is given by

=\displaystyle= dds|s=0Υζ(s)SWΥζ(s)1((Φ1ε,A1)+χ+(φs,as)+χ(ψs,bs))\displaystyle\frac{d}{ds}\Big{|}_{s=0}\Upsilon_{\zeta(s)}\circ\text{SW}\circ\Upsilon_{\zeta(s)}^{-1}\left(\left(\frac{\Phi_{1}}{\varepsilon},A_{1}\right)+\chi^{+}(\varphi_{s},a_{s})+\chi^{-}(\psi_{s},b_{s})\right) (8.21)
=\displaystyle= dds|s=0(𝔗gτgζ(s))1SWpζ(s)(𝔗gτgζ(s))((Φ1ε,A1)+χ+(φs,as)+χ(ψs,bs))\displaystyle\frac{d}{ds}\Big{|}_{s=0}(\mathfrak{T}_{g_{\tau}}^{g_{\zeta(s)}})^{-1}\circ\text{SW}_{p_{\zeta}(s)}\circ(\mathfrak{T}_{g_{\tau}}^{g_{\zeta(s)}})\left(\left(\frac{\Phi_{1}}{\varepsilon},A_{1}\right)+\chi^{+}(\varphi_{s},a_{s})+\chi^{-}(\psi_{s},b_{s})\right)
=\displaystyle= dds|s=0(𝔗ggζ(s))1SWpζ(s)(𝔗ggζ(s))(Φ1ε,A1)\displaystyle\frac{d}{ds}\Big{|}_{s=0}(\mathfrak{T}_{g}^{g_{\zeta(s)}})^{-1}\circ\text{SW}_{p_{\zeta}(s)}\circ(\mathfrak{T}_{g}^{g_{\zeta(s)}})\left(\frac{\Phi_{1}}{\varepsilon},A_{1}\right)
+dds|s=0(𝔗ggζ(s))1(h1(s)pζ(s)+Qpζ(s))(𝔗ggζ(s))(χ+(φs,as)+χ(ψs,bs)),\displaystyle+\ \frac{d}{ds}\Big{|}_{s=0}(\mathfrak{T}_{g}^{g_{\zeta(s)}})^{-1}\circ(\mathcal{L}^{p_{\zeta}(s)}_{h_{1}(s)}+Q^{p_{\zeta(s)}})\circ(\mathfrak{T}_{g}^{g_{\zeta(s)}})\left(\chi^{+}(\varphi_{s},a_{s})+\chi^{-}(\psi_{s},b_{s})\right),

where the last equality is an instance of (8.19). Here, h1(s)pζ(s)\mathcal{L}_{h_{1}(s)}^{p_{\zeta(s)}} is the linearization of the Seiberg–Witten equations at h1(s)=𝔗ggζ(s)(h1)h_{1}(s)=\mathfrak{T}_{g}^{g_{\zeta(s)}}(h_{1}) using the parameter pζ(s)p_{\zeta}(s).

Since ζ(s)=sξ+O(s2)\zeta(s)=s\xi+O(s^{2}), (8.21) is by definition 𝔅h1(ξ)\mathfrak{B}_{h_{1}}(\xi) after dropping O(s2)O(s^{2}) terms. Differentiating (8.21) using the product rule, all terms vanish except those differentiating (φs,as,ψs,bs)(\varphi_{s},a_{s},\psi_{s},b_{s}) since 𝔭(0)=0\mathfrak{p}(0)=0. Because 𝔗ggζ(0)=Id\mathfrak{T}_{g}^{g_{\zeta(0)}}=\text{Id} and QQ is quadratic, what remains is simply h1p0(φ˙,a˙,ψ˙,b˙)\mathcal{L}_{h_{1}}^{p_{0}}(\dot{\varphi},\dot{a},\dot{\psi},\dot{b}), giving the second bullet point. ∎

The next proposition gives a concrete formula for the term 𝔅h1(ξ)\mathfrak{B}_{h_{1}}(\xi) using Theorem 6.4. Set

g˙ξ=dds|s=0F¯sξgτB˙ξ=dds|s=0F¯sξBτ.\dot{g}_{\xi}=\frac{d}{ds}\Big{|}_{s=0}\underline{F}_{s\xi}^{*}g_{\tau}\hskip 56.9055pt\dot{B}_{\xi}=\frac{d}{ds}\Big{|}_{s=0}\underline{F}_{s\xi}^{*}B_{\tau}.
Proposition 8.7.

The term 𝔅h1(ξ)\mathfrak{B}_{h_{1}}(\xi) in (8.16) is given by

𝔅h1(ξ)=(1ε¯Φ1(ξ)𝒟¯A1(ξ)+μ¯Φ1(ξ))\mathfrak{B}_{h_{1}}(\xi)=\begin{pmatrix}\frac{1}{\varepsilon}\underline{\mathcal{B}}_{\Phi_{1}}(\xi)\\ \underline{\mathcal{D}}_{A_{1}}(\xi)+\underline{\mu}_{\Phi_{1}}(\xi)\end{pmatrix} (8.22)

where

  1. (1)

    ¯Φ1(ξ)\underline{\mathcal{B}}_{\Phi_{1}}(\xi) is the metric variation of the Dirac operator from Corollary 6.3,

    ¯Φ1(ξ)=(12ijg˙ξ(ei,ej)ei.j+12dTrgτ(g˙ξ).+12divgτ(g˙ξ).+(Bτ,ξ).)Φ1\underline{\mathcal{B}}_{\Phi_{1}}(\xi)=\left(-\frac{1}{2}\sum_{ij}\dot{g}_{\xi}(e_{i},e_{j})e^{i}.\nabla_{j}+\frac{1}{2}d\text{Tr}_{g_{\tau}}(\dot{g}_{\xi}).+\frac{1}{2}\text{div}_{g_{\tau}}(\dot{g}_{\xi}).+\mathcal{R}(B_{\tau},\xi).\right)\Phi_{1} (8.23)

    where Φ1=Φ1(ε,τ)\Phi_{1}=\Phi_{1}({\varepsilon,\tau}) is as in (8.8), . denotes the Clifford multiplication of gτg_{\tau}, and j\nabla_{j} is the covariant derivative on SES_{E} formed using the spin connection of gτg_{\tau}, A1A_{1}, and BτB_{\tau}.

  2. (2)

    𝔇¯A1(ξ)\underline{\mathfrak{D}}_{A_{1}}(\xi) is the metric variation of the de-Rham operator 𝕕\mathbb{d} given by

    𝔇¯A1(ξ)=(12ijg˙ξ(ei,ej)𝕔(ei)jLC+12𝕔(dTrgτ(g˙ξ))+12𝕔(divgτ(g˙ξ)))A1{\bf\underline{\mathfrak{D}}}_{A_{1}}(\xi)=\left(-\frac{1}{2}\sum_{ij}\dot{g}_{\xi}(e_{i},e_{j})\mathbb{c}(e^{i})\nabla^{\text{LC}}_{j}+\frac{1}{2}\mathbb{c}(d\text{Tr}_{g_{\tau}}(\dot{g}_{\xi}))+\frac{1}{2}\mathbb{c}(\text{div}_{g_{\tau}}(\dot{g}_{\xi}))\right)A_{1} (8.24)

    where A1=A1(ε,τ)A_{1}=A_{1}(\varepsilon,\tau) is the connection form of (8.8) in the trivialization of Lemma 3.10, 𝕔\mathbb{c} is the symbol of 𝕕\mathbb{d}, and LC\nabla^{\text{LC}} is the Levi-Civitas connection of gτg_{\tau}.

  3. (3)

    μ¯Φ1(ξ)\underline{\mu}_{\Phi_{1}}(\xi) is the metric variation of the moment map given by

    12μ¯Φ1(ξ)=(0,12jkig˙ξ(ej,ek)ej.Φ1,Φ1ε2iek)\frac{1}{2}\underline{\mu}_{\Phi_{1}}(\xi)=\left(0,-\frac{1}{2}\sum_{jk}\ \frac{\langle i\dot{g}_{\xi}(e_{j},e_{k})e^{j}.\Phi_{1},\Phi_{1}\rangle}{\varepsilon^{2}}ie^{k}\right) (8.25)

    where . and Φ1\Phi_{1} are as in Item (1). Here j,k=1,2,3j,k=1,2,3 are indices and i=1i=\sqrt{-1}.

Proof.

(1) The metric variation formula of Bourguignon-Gauduchon (Theorem 6.4) applies equally well to twisted Dirac operators, provided the connection on the twisting bundle remains fixed. The U(1)U(1)-connection in (8.17) (i.e. in the middle arrow of the diagram in the proof of Proposition 8.6) is the fixed connection A1A_{1} by Definition 8.8. The connection on EE differs from the fixed connection BτB_{\tau} by a zeroth order (in both ξ\xi and Φ1\Phi_{1}) term

BτB˙ξ:=(Bτ,ξ).B_{\tau}-\dot{B}_{\xi}:=\mathcal{R}(B_{\tau},\xi).

In fact, a quick calculation shows (Bτ,ξ)=ιξFBτ\mathcal{R}(B_{\tau},\xi)=\iota_{\xi}F_{B_{\tau}} is the contraction of the curvature with dds|s=0F¯sξ\frac{d}{ds}\Big{|}_{s=0}\underline{F}_{s\xi}.

(2) Let LL be as in the proof of Lemma 8.2. Let AA_{\circ} denote a smooth connection on LL that extends the product connection in the trivialization of Lemma 3.10. Then

ξFA1=ξFA+ξd(AA1)\star_{\xi}F_{A_{1}}=\star_{\xi}F_{A_{\circ}}+\star_{\xi}d(A_{\circ}-A_{1})

where ξ\star_{\xi} is the Hodge star of gξg_{\xi}. Since gξ=gτg_{\xi}=g_{\tau} outside the neighborhood Nr0(𝒵τ)N_{r_{0}}(\mathcal{Z}_{\tau}) where F¯ξ=Id\underline{F}_{\xi}=\text{Id}, while FAF_{A_{\circ}} vanishes inside Nr0(𝒵τ)N_{r_{0}}(\mathcal{Z}_{\tau}), the first term is zero. Consequently, the variation of the curvature (when supplemented with the Ω0(i)\Omega^{0}(i\mathbb{R}) component and gauge-fixing) reduces to the metric variation of 𝕕\mathbb{d}. The variation of this Dirac-type operator follows equally well from (Theorem 6.4), with the additional simplification that the form bundle does not depend on the metric.

(3) Let 𝔞ξ\mathfrak{a}_{\xi} be such that gξ=gτ(𝔞ξX,Y)g_{\xi}=g_{\tau}(\mathfrak{a}_{\xi}X,Y), then e~j=𝔞ξ1/2ej\widetilde{e}_{j}=\mathfrak{a}_{\xi}^{-1/2}e_{j} is an orthonormal frame for gξg_{\xi} where {ej}\{e_{j}\} is one of gτg_{\tau}. Expanding in Taylor series in the frame of gτg_{\tau} and differentiating yields a˙ξ(s)=12g˙ξ\dot{a}_{\xi}(s)=-\tfrac{1}{2}\dot{g}_{\xi}, just as in the symbol term of (6.4). The variation formula then follows from the definition of μ\mu (3.1). ∎

8.4. Non-Linear Terms

This section characterizes the non-linear terms in the universal Seiberg-Witten equations. The equations have quadratic non-linearities in fiber directions of Hε,ν1H^{1}_{\varepsilon,\nu}—these simply being the non-linearities of the original Seiberg-Witten equations, but are fully non-linear in the deformation parameter ξ\xi. There are also mixed terms fully non-linear in ξ\xi and linear or quadratic in the fiber directions.

The universal Seiberg-Witten equations at h=ε1¯h=\mathbb{H}^{1}_{\varepsilon}\oplus\underline{\mathbb{R}} may be written

𝕊𝕎(h)\displaystyle\mathbb{SW}\left(h\right) =\displaystyle= SW(Φ1ε,A1)+d𝕊𝕎h1(h)+h1(h)\displaystyle\text{SW}\left(\frac{\Phi_{1}}{\varepsilon},A_{1}\right)\ +\ \text{d}{\mathbb{SW}}_{h_{1}}(h)\ +\ \mathbb{Q}_{h_{1}}(h) (8.26)

where h1\mathbb{Q}_{h_{1}} consists of the non-linear terms. Write h=(ξ,φ,a)h=(\xi,\varphi,a) where (φ,a)=χ+(φ+,a+)+χ(ψ,b)(\varphi,a)=\chi^{+}(\varphi^{+},a^{+})+\chi^{-}(\psi,b) for (φ+,a+,ψ,b)Hε1,+Hε1,(\varphi^{+},a^{+},\psi,b)\in H^{1,+}_{\varepsilon}\oplus H^{1,-}_{\varepsilon} to simplify notation.

Proposition 8.8.

The non-linear term has the form

h1(h)=QSW(φ,a)+εQΦ1(ξ,φ)+εQA1(ξ,a)+Qa.φ(ξ,φ,a)+Qμ(ξ,φ,φ)\mathbb{Q}_{h_{1}}(h)=Q_{\text{SW}}(\varphi,a)\ +\ \varepsilon Q_{\Phi_{1}}(\xi,\varphi)\ +\ \varepsilon Q_{A_{1}}(\xi,a)\ +\ Q_{a.\varphi}(\xi,\varphi,a)\ +\ Q_{\mu}(\xi,\varphi,\varphi) (8.27)

where

  1. (1)

    QSWQ_{\text{SW}} is the standard non-linearity of the Seiberg–Witten equations given by (a.φ,μ(φ,φ))(a.\varphi,\mu(\varphi,\varphi)) using Clifford multiplication with respect to gτg_{\tau}.

  2. (2)

    QΦ1Q_{\Phi_{1}} is the non-linear portion of the metric variation of the Dirac operator A1\not{D}_{A_{1}} on SES_{E} as in Lemma 6.13, taking Φ=ε1Φ1\Phi=\varepsilon^{-1}\Phi_{1} and ψ=φ\psi=\varphi in the notation of that lemma.

  3. (3)

    QA1Q_{A_{1}} is the non-linear portion of the metric variation of the de-Rham operator 𝕕\mathbb{d} as in Lemma 6.13 taking Φ=A1\Phi=A_{1} and ψ=a\psi=a and appropriately substituting 𝐜𝐥{\bf{cl}} for Clifford multiplication.

  4. (4)

    Qa.φQ_{a.\varphi} is the non-linearity arising from the metric variation on the first component of QSWQ_{\text{SW}}, which has the form

    Qa.φ=F1(χ¯[ξ])Ma.φ(ξ)Q_{a.\varphi}=F_{1}(\underline{\chi}^{\prime}[\xi])\cdot M_{a.\varphi}(\xi)

    where Ma.φ(ξ)M_{a.\varphi}(\xi) is a term of Type B in the sense of (6.23) taking ψ=a.φ\psi=a.\varphi with weight wB=1w_{B}=1, and F1F_{1} is a C(Y)C^{\infty}(Y)-linear combination of 1,χ¯[ξ],aχ¯[ξ],1,\underline{\chi}[\xi^{\prime}],\underline{\partial_{a}\chi}[\xi], and higher order functions of these.

  5. (5)

    QμQ_{\mu} is the non-linearity arising from the metric variation of the moment map. Schematically, it has the form

    12g˙ξ(ej,ek)iej(Φ1+εφ),εφε2iek+qjk(ξ)iej(Φ1+εφ),Φ1+εφε2iek-\frac{1}{2}\frac{\langle\dot{g}_{\xi}(e^{j},e^{k})ie^{j}(\Phi_{1}+\varepsilon\varphi),\varepsilon\varphi\rangle}{\varepsilon^{2}}ie^{k}+\frac{\langle q_{jk}(\xi)ie^{j}(\Phi_{1}+\varepsilon\varphi),\Phi_{1}+\varepsilon\varphi\rangle}{\varepsilon^{2}}ie^{k}

    where qjkq_{jk} consists of quadratic and higher combinations of χ¯[ξ],aχ¯[ξ]\underline{\chi}[\xi^{\prime}],\underline{\partial_{a}\chi}[\xi].

Proof.

(1) is immediate, (2)–(3) follow directly from Lemma 6.13. (4)–(5) are derived by expanding the formula for g˙ξ\dot{g}_{\xi} as in Proposition 8.7 in Taylor series (see [43, Section 5.3]). ∎

9. Relating Deformation Operators

This section relates the deformation operator for the universal Seiberg–Witten equations from Section 8.3 (Proposition 8.6) to the deformation operator for the universal Dirac operator from Section 6.1 (Proposition 6.3). In the outside region YY^{-}, the former is a small perturbation of the latter.

9.1. Two Deformation Operators

Because of the renormalization of the spinor in the model solutions (8.8), it is convenient to introduce the following inverse normalization for the deformation. For ξL2,2(𝒵τ;N𝒵τ)\xi\in L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}), define η\eta by

ξ(t)=εη(t).\boxed{\xi(t)=\varepsilon\eta(t).} (9.1)

Next, with h0=(𝒵τ,Φτ,Aτ)h_{0}=(\mathcal{Z}_{\tau},\Phi_{\tau},A_{\tau}) and h1=(𝒵τ,Φ1,A1)h_{1}=(\mathcal{Z}_{\tau},\Phi_{1},A_{1}), define Ξ:L2,2(𝒵τ;N𝒵τ)Lε,ν2(Y)\Xi:L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau})\to L^{2}_{\varepsilon,\nu}(Y) by

d𝕊𝕎h1(ξ,0,0)\displaystyle\text{d}\mathbb{SW}_{h_{1}}(\xi,0,0) =\displaystyle= dh0(η,0)+Ξ(η)\displaystyle\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)\mathbb{\ }+\ \Xi(\eta) (9.2)
=\displaystyle= dh0(η,0)+Ξ+(η)+Ξ(η)\displaystyle\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)\mathbb{\ }+\ \Xi^{+}(\eta)\ +\ \Xi^{-}(\eta) (9.3)

where Ξ±(η)=Ξ(η)𝟙±\Xi^{\pm}(\eta)=\Xi(\eta)\mathbb{1}^{\pm} for 𝟙±\mathbb{1}^{\pm} as in Section 2.

Because of the exponential decay in Lemma 7.13, the (re-normalized) approximate solution (Φ1,A1)(\Phi_{1},A_{1}) is a small perturbation of the original 2\mathbb{Z}_{2}-eigenvector (Φτ,Aτ)(\Phi_{\tau},A_{\tau}) outside the invariant scale of r=ε2/3r=\varepsilon^{2/3}, thus one expects d𝕊𝕎h1(εη)\text{d}\mathbb{SW}_{h_{1}}(\varepsilon\eta) to also be a small perturbation of dh0(η)\text{d}\not{\mathbb{D}}_{h_{0}}(\eta) in this region. The following lemma gives a precise bound.

Lemma 9.1.

Let η(t)=ε1ξ(t)C(𝒵τ;N𝒵τ)\eta(t)=\varepsilon^{-1}\xi(t)\in C^{\infty}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) be a linearized deformation. There exists a constant CC such that

Ξ(η)Lε,ν2Cε11/12γη3/2+γ¯ν\|\Xi^{-}(\eta)\|_{L^{2}_{\varepsilon,\nu}}\leq C\varepsilon^{11/12-\gamma}\|\eta\|_{3/2+\underline{\gamma}-\nu} (9.4)

where γ,γ¯<<1\gamma,\underline{\gamma}<<1, with γ¯\underline{\gamma} as in Lemma 6.12.

Proof.

By Proposition 6.3, dh0(η,0)=¯Φτ(η)\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)=\underline{\mathcal{B}}_{\Phi_{\tau}}(\eta) and 1ε¯Φ1(ξ)=¯Φ1(η)\tfrac{1}{\varepsilon}\underline{\mathcal{B}}_{\Phi_{1}}(\xi)=\underline{\mathcal{B}}_{\Phi_{1}}(\eta). By Proposition 8.7, we can therefore write

Ξ(η)=(¯Φ1(η)¯Φτ(η)𝒟¯A1(η)+μ¯Φ1(η)).\Xi(\eta)=\begin{pmatrix}\underline{\mathcal{B}}_{\Phi_{1}}(\eta)-\underline{\mathcal{B}}_{\Phi_{\tau}}(\eta)\\ \underline{\mathcal{D}}_{A_{1}}(\eta)+\underline{\mu}_{\Phi_{1}}(\eta)\end{pmatrix}.

We first compare the spinor components. Let φ1\varphi_{1} denote the difference

Φ1ε=Φτε+φ1,\frac{\Phi_{1}}{\varepsilon}=\frac{\Phi_{\tau}}{\varepsilon}+\varphi_{1}, (9.5)

which satisfies φ1Hε1,+ε1/12γ\|\varphi_{1}\|_{H^{1,+}_{\varepsilon}}\leq\varepsilon^{-1/12-\gamma} on the support of 𝟙\mathbb{1}^{-} by (7.6) in Item (II) of Theorem 7.5 and the exponential bound on ΦhεΦτ\Phi^{h_{\varepsilon}}-\Phi_{\tau} from Item(2b) in Section 7.1.

Then

Φ1(ξ)Φτ(ξ)\displaystyle\mathcal{B}_{\Phi_{1}}(\xi)-\mathcal{B}_{\Phi_{\tau}}(\xi) =\displaystyle= εφ1(η)\displaystyle\varepsilon\mathcal{B}_{\varphi_{1}}(\eta) (9.6)
=\displaystyle= ε(12ijg˙η(ei,ej)ei.j+12dTrgτ(g˙η).+12divgτ(g˙η).+(Bτ,η).)φ1\displaystyle\varepsilon\left(-\frac{1}{2}\sum_{ij}\dot{g}_{\eta}(e_{i},e_{j})e^{i}.\nabla_{j}+\frac{1}{2}d\text{Tr}_{g_{\tau}}(\dot{g}_{\eta}).+\frac{1}{2}\text{div}_{g_{\tau}}(\dot{g}_{\eta}).+\mathcal{R}(B_{\tau},\eta).\right)\varphi_{1}

Each term in (9.6) is of Type B in the sense of Lemma 6.12, each with weight wB2w_{B}\leq 2. Applying Item (B) of that lemma with β=ν\beta=\nu shows that

εφ(η)𝟙Lε,ν2,Cεη3/2+γ¯ν(φ1rHe1(Y))Cε11/12γη3/2+γ¯ν.\|\varepsilon\mathcal{B}_{\varphi}(\eta)\mathbb{1}^{-}\|_{L^{2,-}_{\varepsilon,\nu}}\leq C\varepsilon\|\eta\|_{3/2+\underline{\gamma}-\nu}\Big{(}\|\varphi_{1}\|_{rH^{1}_{e}(Y^{-})}\Big{)}\leq C\varepsilon^{11/12-\gamma}\|\eta\|_{3/2+\underline{\gamma}-\nu}. (9.7)

as desired. In the final inequality, we have used that the rHe1rH^{1}_{e} and Hε1,+H^{1,+}_{\varepsilon} norms are comparable for rε2/3r\geq\varepsilon^{2/3}. This establishes the desired bound for the spinor components.

For the SIm(Ω0Ω1)S^{\text{Im}}\oplus(\Omega^{0}\oplus\Omega^{1}), components, note that that the difference from the (flat) limiting connection AτA_{\tau} and the imaginary components of Φ1\Phi_{1} are exponentially small on supp(𝟙)\text{supp}(\mathbb{1}^{-}) by Lemma 7.13; exponentially small here meaning O(ε3Exp(1εγ))O(\varepsilon^{-3}\text{Exp}(-\tfrac{1}{\varepsilon^{\gamma}})) for γ>0\gamma>0. Since 𝒟¯Aτ\underline{\mathcal{D}}_{A_{\tau}} depends only on the Ω0Ω1\Omega^{0}\oplus\Omega^{1}-component and μ¯Φ1\underline{\mu}_{\Phi_{1}} only depend on the SImS^{\text{Im}}-component by Item (2) of Lemma 3.5 (up to an exponentially small factor coming from the difference in (2a) in Section 7.1), the same argument as the spinor components shows these components satisfy (9.4) with an exponential factor in place of ε11/12\varepsilon^{11/12}, which may be absorbed once ε\varepsilon is sufficiently small. ∎

The situation for radii less than the invariant scale, i.e. for Ξ+\Xi^{+}, stands in contrast to that of Lemma 9.1: there, the two deformation operators bear no meaningful relation. Since the invariant scale shrinks rapidly as ε0\varepsilon\to 0, however, the norm of either deformation operator in this region decreases like a fixed positive power of ε\varepsilon so that Ξ+\Xi^{+} is effectively negligible, as the upcoming lemma shows. Since the weight function RεR_{\varepsilon} is almost constant (up to a factor of εγ\varepsilon^{-\gamma}) on supp(𝟙+)\text{supp}(\mathbb{1}^{+}), the lemma considers only the unweighted norms.

The proof utilizes the following re-scaling, which plays an essential role in the proof of Theorem 7.5 (see [41, Sec. 5.3]). There is a re-scaled coordinate ρ\rho satisfying c1ε2/3ρc2ε2/3c_{1}\varepsilon^{2/3}\leq\rho\leq c_{2}\varepsilon^{2/3} for constants c1,c2c_{1},c_{2} such that de-singularized solutions (Φhε,Ahε)(\Phi^{h_{\varepsilon}},A^{h_{\varepsilon}}) are given by

(Φhε,Ahε)=(ε1/3ΦH(ρ),AH(ρ))(\Phi^{h_{\varepsilon}},A^{h_{\varepsilon}})=(\varepsilon^{1/3}\Phi^{H}(\rho),A^{H}(\rho)) (9.8)

where ΦH,AH\Phi^{H},A^{H} are fixed, smooth, ε\varepsilon-independent functions on 𝒵τ×2\mathcal{Z}_{\tau}\times\mathbb{R}^{2} in Fermi coordinates (3.9). Moreover, ΦHρ1/2\Phi^{H}\sim\rho^{1/2} for ρ>>1\rho>>1, and AH=f(ρ)(dzzdz¯z¯)A^{H}=f(\rho)\left(\tfrac{dz}{z}-\tfrac{d\overline{z}}{\overline{z}}\right), where f(ρ)f(\rho) vanishes to second order at ρ=0\rho=0. See Section 4 of [41] for detailed proofs.

Lemma 9.2.

Let η(t)=ε1ξ(t)C(𝒵τ;N𝒵τ)\eta(t)=\varepsilon^{-1}\xi(t)\in C^{\infty}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) be a linearized deformation. There exists a constant CC such that

Ξ+(η)L2\displaystyle\|\Xi^{+}(\eta)\|_{L^{2}} \displaystyle\leq d𝕊𝕎h1(ξ,0)𝟙+L2+dh0(η,0)𝟙+L2\displaystyle\|\text{d}\mathbb{SW}_{h_{1}}(\xi,0)\mathbb{1}^{+}\|_{L^{2}}+\|\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)\mathbb{1}^{+}\|_{L^{2}} (9.9)
\displaystyle\leq Cεγ(ε1/3η1,2+ε11/12η3/2+γ¯,2+εη2,2+ε19/12η5/2+γ¯,2).\displaystyle C\varepsilon^{-\gamma}\Big{(}\varepsilon^{1/3}\|\eta\|_{1,2}\ +\ \varepsilon^{11/12}\|\eta\|_{3/2+\underline{\gamma},2}\ +\ \ \varepsilon\|\eta\|_{2,2}\ +\ \varepsilon^{19/12}\|\eta\|_{5/2+\underline{\gamma},2}\Big{)}. (9.10)

where γ,γ¯<<1\gamma,\underline{\gamma}<<1, with γ¯\underline{\gamma} as in Lemma 6.12.

Proof.

(9.9) is immediate from (9.2) and the triangle inequality, so it suffices to show both terms on the right side of (9.9) satisfy the desired bound.

To begin with the deformations of the singular Dirac operator, the formula (6.7) shows

dh0(η,0)𝟙+L2(g˙η)ijei.ΦτL2+dTrgτ(g˙η).ΦτL2+divgτ(g˙η).ΦτL2+(Bτ,η).ΦτL2.\|\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)\mathbb{1}^{+}\|_{L^{2}}\leq\|(\dot{g}_{\eta})_{ij}e^{i}.\nabla\Phi_{\tau}\|_{L^{2}}+\|d\text{Tr}_{g_{\tau}}(\dot{g}_{\eta}).\Phi_{\tau}\|_{L^{2}}+\|\text{div}_{g_{\tau}}(\dot{g}_{\eta}).\Phi_{\tau}\|_{L^{2}}+\|\mathcal{R}(B_{\tau},\eta).\Phi_{\tau}\|_{L^{2}}.

where summation is implicit in the first term. Since Φτ\Phi_{\tau} is polyhomogeneous with leading order r1/2r^{1/2} by Lemma 4.5, each of these terms is of Type A in the sense of Lemma 6.12, with weight wA2w_{A}\leq 2. Copying the proof of Item (A) of Lemma 6.12, but now only integrating over the support of 𝟙+\mathbb{1}^{+} shows that

(g˙η)ijei.ΦτL2Cε1/3γη1,2dTrgτ(g˙η).ΦτL2+divgτ(g˙η).ΦτL2Cε1γη2,2.\|(\dot{g}_{\eta})_{ij}e^{i}.\nabla\Phi_{\tau}\|_{L^{2}}\leq C\varepsilon^{1/3-\gamma}\|\eta\|_{1,2}\hskip 42.67912pt\|d\text{Tr}_{g_{\tau}}(\dot{g}_{\eta}).\Phi_{\tau}\|_{L^{2}}+\|\text{div}_{g_{\tau}}(\dot{g}_{\eta}).\Phi_{\tau}\|_{L^{2}}\leq C\varepsilon^{1-\gamma}\|\eta\|_{2,2}.

as desired. The term (τ,η)\mathcal{R}(\mathcal{B}_{\tau},\eta) is subsumed by these because it has weight wA=1w_{A}=1. This shows the second term of (9.9) satisfies the desired bound.

By (8.22) and the triangle inequality, the second term of (9.9) is bounded by

d𝕊𝕎h1(ξ)𝟙+L2¯Φ1(η)(I)L2+ε𝒟¯A1(η)(II)L2+εμ¯Φ1(η)(III)L2.\displaystyle\|\text{d}\mathbb{SW}_{h_{1}}(\xi)\mathbb{1}^{+}\|_{L^{2}}\leq\|\underbrace{\underline{\mathcal{B}}_{\Phi_{1}}(\eta)}_{\text{(I)}}\|_{L^{2}}\ +\ \varepsilon\|\underbrace{\underline{\mathcal{D}}_{A_{1}}(\eta)}_{\text{(II)}}\|_{L^{2}}\ +\ \varepsilon\|\underbrace{\underline{\mu}_{\Phi_{1}}(\eta)}_{\text{(III)}}\|_{L^{2}}.

We bound each term (I)–(III) separately. Beginning with (I), write (Φ1,A1)=(Φτhε,Aτhε)+(εφ1,a1)(\Phi_{1},A_{1})=(\Phi^{h_{\varepsilon}}_{\tau},A^{h_{\varepsilon}}_{\tau})+(\varepsilon\varphi_{1},a_{1}) where (φ1,a1)(\varphi_{1},a_{1}) are as in Item (2) of Theorem 7.5. The term (I) is comprised of four subterms (Ia)–(Id) as in (8.23); for each of these there is a leading order part coming from (Φhε,Ahε)(\Phi^{h_{\varepsilon}},A^{h_{\varepsilon}}) and a perturbation coming from (φ1,a1)(\varphi_{1},a_{1}). We begin with the leading order part of (Ia). Omitting indices and subscripts and writing N+=supp(𝟙+)N^{+}=\text{supp}(\mathbb{1}^{+}) for clarity,

g˙η.ΦhεL22\displaystyle\|\dot{g}_{\eta}.\nabla\Phi^{h_{\varepsilon}}\|^{2}_{L^{2}} \displaystyle\leq CN+(|χ¯(η)|2+|dχ¯(η)|2)|Φhε|2r𝑑r𝑑θ𝑑t\displaystyle C\int_{N^{+}}\left(|\underline{\chi}(\eta^{\prime})|^{2}+|\underline{d\chi}(\eta)|^{2}\right)|\nabla\Phi^{h_{\varepsilon}}|^{2}rdrd\theta dt
\displaystyle\leq CN+(|χ¯(η)|2+|dχ¯(η)|2)|ρΦhε|2ρ𝑑ρ𝑑θ𝑑t\displaystyle C\int_{N^{+}}\left(|\underline{\chi}(\eta^{\prime})|^{2}+|\underline{d\chi}(\eta)|^{2}\right)|\nabla_{\rho}\Phi^{h_{\varepsilon}}|^{2}\rho d\rho d\theta dt
\displaystyle\leq Cε2/3𝒵τ(|χ¯(η)|2+|dχ¯(η)|2)ρεγ|ρΦH|2ρ𝑑ρ𝑑θ𝑑t\displaystyle C\varepsilon^{2/3}\int_{\mathcal{Z}_{\tau}}\left(|\underline{\chi}(\eta^{\prime})|^{2}+|\underline{d\chi}(\eta)|^{2}\right)\int_{\rho\leq\varepsilon^{-\gamma}}|\nabla_{\rho}\Phi^{H}|^{2}\rho d\rho d\theta dt
\displaystyle\leq Cε2/3γη1,22\displaystyle C\varepsilon^{2/3-\gamma}\|\eta\|_{1,2}^{2}

where we have changed variables to the rescaled coordinate ρ\rho (in both the volume and \nabla) and then substituted (9.8). The last inequality follows from the fact that dχ¯(η)L2(S1)CηL2(S1)\|\underline{d\chi}(\eta)\|_{L^{2}(S^{1})}\leq C\|\eta^{\prime}\|_{L^{2}(S^{1})} (as in the proof of Lemma 6.14), and the fact that ΦHρ1/2\Phi^{H}\sim\rho^{1/2} for ρconst\rho\geq\text{const}. The same argument applies to the terms (Ib)–(Id), except there is an additional factor of ε2/3\varepsilon^{2/3} because there is no derivative to rescale, but the norm ηL2,2\|\eta\|_{L^{2,2}} is needed.

Turning now to the leading order of term (II), which is again comprised of three sub-terms (IIa)–(IIc) as in (8.24), a similar rescaling argument applies to show

ε2g˙ηAhεL22\displaystyle\varepsilon^{2}\|\dot{g}_{\eta}\nabla A^{h_{\varepsilon}}\|_{L^{2}}^{2} \displaystyle\leq ε2N+(|χ¯(η)|2+|dχ¯(η)|2)|ρf(ρ)(dzzdz¯z¯)|2ρ𝑑ρ𝑑θ𝑑t\displaystyle\varepsilon^{2}\int_{N^{+}}\left(|\underline{\chi}(\eta^{\prime})|^{2}+|\underline{d\chi}(\eta)|^{2}\right)|\nabla_{\rho}f(\rho)\left(\tfrac{dz}{z}-\tfrac{d\overline{z}}{\overline{z}}\right)|^{2}\rho d\rho d\theta dt
=\displaystyle= ε2ε4/3N+(|χ¯(η)|2+|dχ¯(η)|2)|G(ρ)|2ρ𝑑ρ𝑑θ𝑑tε2/3η1,22\displaystyle\varepsilon^{2}\varepsilon^{-4/3}\int_{N^{+}}\left(|\underline{\chi}(\eta^{\prime})|^{2}+|\underline{d\chi}(\eta)|^{2}\right)|G(\rho)|^{2}\rho d\rho d\theta dt\leq\varepsilon^{2/3}\|\eta\|^{2}_{1,2}

where we have substituted 1z=ε2/3ρeiθ\tfrac{1}{z}=\tfrac{\varepsilon^{2/3}}{\rho e^{i\theta}} (and likewise for z¯\overline{z}), then used the fact that f(ρ)f(\rho) vanishes to second order at ρ=0\rho=0 to combine these into a smooth bounded function G(ρ)G(\rho). The last inequality follows just as in term (I). Terms (IIb)–(IIc) proceed with the same alterations as (Ib)–(Id).

In both terms (I)–(II) the perturbation terms arising from (φ1,a1)(\varphi_{1},a_{1}), the same integration as for the leading order terms is used. First, for terms (Ia) and (IIa), a factor of η3/2+γ¯,2\|\eta\|_{3/2+\underline{\gamma},2} can be pulled out, reducing the integral to the Hε1,+H^{1,+}_{\varepsilon}-norm; for the other terms, the same applies to (Ib–Id) and (IIb)–(IIc) with η5/2+γ¯,2\|\eta\|_{5/2+\underline{\gamma},2} (and the weight in the norm gives an extra factor of ε2/3γ\varepsilon^{2/3-\gamma} for these).

The term (III) the terms of (8.25) of the form igξ.Φτ,φ1\langle ig_{\xi}.\Phi_{\tau},\varphi_{1}\rangle and igξ.φ1,Φτ\langle ig_{\xi}.\varphi_{1},\Phi_{\tau}\rangle is can be bounded simply by pulling an εη3/2+γ¯\varepsilon\|\eta\|_{3/2+\underline{\gamma}} out of the integral and using the weight on the Hε1,+H^{1,+}_{\varepsilon}-norm involving μ\mu. Finally, for the correction terms igξ.φ1,φ1\langle ig_{\xi}.\varphi_{1},\varphi_{1}\rangle, a bound by ε11/12η3/2+γ¯\varepsilon^{11/12}\|\eta\|_{3/2+\underline{\gamma}} follows from the interpolation inequality in the upcoming Lemma 9.4. ∎

9.2. The Range Component of Deformations

The previous subsection provided bounds on the perturbation term in (9.2); this subsection bounds the main term d(η,0)\text{d}\not{\mathbb{D}}(\eta,0). The range component (1Πτ)d(η,0)(1-\Pi_{\tau})\text{d}\not{\mathbb{D}}(\eta,0) is the crucial term affected by the loss of regularity whose significance was discussed in Subsection 2.5.1 (the boxed term in 2.26).

The key bound is provided by the following lemma. The lemma shows that, due to the use of the mode-dependent deformations, the derivative d(η,0)\text{d}\not{\mathbb{D}}(\eta,0) has effective support (Definition 4.6) equal to the support of the deformation η\eta in Fourier space.

Proposition 9.3.

Suppose that ηC(𝒵τ,N𝒵τ)\eta\in C^{\infty}(\mathcal{Z}_{\tau},N\mathcal{Z}_{\tau}) satisfies the following property: there is an MM\in\mathbb{R} such that ηm+1/2,2CMmη1/2,2\|\eta\|_{m+1/2,2}\leq CM^{m}\|\eta\|_{1/2,2} for all m>0m>0. Then deformation operator (6.3) of the universal Dirac operator satisfies

dh0(η,0)Lν2CMνη1/2,2\|\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)\|_{L^{2}_{-\nu}}\leq CM^{\nu}\|\eta\|_{1/2,2} (9.11)

for any weight ν>0\nu>0. In particular, it is effectively supported where r=O(M1)r=O(M^{-1}) in the sense of Definition 4.6.

Proof.

By Lemma 4.5, Φτ\Phi_{\tau} is polyhomogeneous. In particular, it satisfies the required bounds (6.22) uniformly in τ\tau, so that each term (6.7) of dh0(η,0)\text{d}_{h_{0}}\not{\mathbb{D}}(\eta,0) is a term of Type A with weight wA2w_{A}\leq 2 in the sense of Lemma 6.12. The conclusion follows directly from applying Item (A) of that lemma with β=ν\beta=-\nu and invoking the assumption on η\eta. ∎

9.3. Non-Linear Bounds

This subsection bounds the non-linear terms in Proposition 8.8.

The statement of the next lemma involves an auxiliary partition of unity defined as follows. With γ<<1\gamma<<1 being the constant such that 𝟙+\mathbb{1}^{+} is the indicator function of {rε2/3γ}\{r\leq\varepsilon^{2/3-\gamma}\}, let ζ±\zeta^{\pm} be a partition of unity consisting of two logarithmic cut-off functions such that |dζ±|Cr1|d\zeta^{\pm}|\leq Cr^{-1}, and ζ+=1\zeta^{+}=1 where {rε2/32γ}\{r\leq\varepsilon^{2/3-2\gamma}\} while supp(ζ+){r2ε2/3γ}\text{supp}(\zeta^{+})\subseteq\{r\leq 2\varepsilon^{2/3-\gamma}\}. The purpose of introducing ζ+\zeta^{+} is that it is equal to 11 on a neighborhood larger than the support of 𝟙+\mathbb{1}^{+} by a factor of εγ\varepsilon^{\gamma}; this extra buffer zone allows the exponential decay of Lemma 7.13 to take effect on supp(ζ)\text{supp}(\zeta^{-}) for configurations that decay away from the support of 𝟙+\mathbb{1}^{+} (see the proof of Corollary 10.8 in Section 10).

Lemma 9.4.

Let =QSW+QΦ1+QA1+Qa.φ+Qμ\mathbb{Q}=Q_{\text{SW}}+Q_{\Phi_{1}}+Q_{A_{1}}+Q_{a.\varphi}+Q_{\mu} be the non-linear terms from Proposition 8.8. Then these satisfy the following bounds.

  1. (I)

    QSWQ_{\text{SW}} satisfies

    QSW(φ,a)ζ+Cε1/3γ(φ,a)Hε12\displaystyle\ \ \ \ \ \|Q_{\text{SW}}(\varphi,a)\zeta^{+}\|\leq C\varepsilon^{1/3-\gamma}\|(\varphi,a)\|_{H^{1}_{\varepsilon}}^{2}\hskip 42.67912pt QSWRe(φ,a)ζL2Cε1/4(φIm,a)Hε12\displaystyle\|Q^{\text{Re}}_{\text{SW}}(\varphi,a)\zeta^{-}\|_{L^{2}}\leq C\varepsilon^{1/4}\|(\varphi^{\text{Im}},a)\|^{2}_{H^{1}_{\varepsilon}}
    QSWIm(φ,a)ζL2Cε1/4φReHε1(φIm,a)Hε1.\displaystyle\|Q^{\text{Im}}_{\text{SW}}(\varphi,a)\zeta^{-}\|_{L^{2}}\leq C\varepsilon^{1/4}\|\varphi^{\text{Re}}\|_{H^{1}_{\varepsilon}}\|(\varphi^{\text{Im}},a)\|_{H^{1}_{\varepsilon}}.
  2. (II)

    Provided ξ3/2+γ¯1\|\xi\|_{3/2+\underline{\gamma}}\leq 1, then QΦ1,QA1Q_{\Phi_{1}},Q_{A_{1}} satisfy

    QΦ1(η,φ)𝟙+L2\displaystyle\|Q_{\Phi_{1}}(\eta,\varphi)\mathbb{1}^{+}\|_{L^{2}} \displaystyle\leq Cε1γη3/2+γ¯(φHε1+ε1/3η1+εη2)+Cε5/3γη5/2+γ¯φHε1\displaystyle C\varepsilon^{1-\gamma}\|\eta\|_{3/2+\underline{\gamma}}\left(\|\varphi\|_{H^{1}_{\varepsilon}}\ +\ \varepsilon^{1/3}\|\eta\|_{1}+\varepsilon\|\eta\|_{2}\right)+C\varepsilon^{5/3-\gamma}\|\eta\|_{5/2+\underline{\gamma}}\|\varphi\|_{H^{1}_{\varepsilon}}
    QΦ1(η,φ)𝟙L2\displaystyle\|Q_{\Phi_{1}}(\eta,\varphi)\mathbb{1}^{-}\|_{L^{2}} \displaystyle\leq C(εη3/2+γ¯φHε1+εη1/2η3/2+γ¯),\displaystyle C\left(\varepsilon\|\eta\|_{3/2+\underline{\gamma}}\|\varphi\|_{H^{1}_{\varepsilon}}\ +\ \varepsilon\|\eta\|_{1/2}\|\eta\|_{3/2+\underline{\gamma}}\right),

    and identically for QA1Q_{A_{1}} with aHε1\|a\|_{H^{1}_{\varepsilon}} in place of φHε1\|\varphi\|_{H^{1}_{\varepsilon}}.

  3. (III)

    Provided ξ3/2+γ¯1\|\xi\|_{3/2+\underline{\gamma}}\leq 1, then

    Qa.φ(η,a,φ)L2\displaystyle\|Q_{a.\varphi}(\eta,a,\varphi)\|_{L^{2}} \displaystyle\leq Cεη3/2+γ¯QSW(φ,a)L2\displaystyle C\varepsilon\|\eta\|_{3/2+\underline{\gamma}}\|Q_{\text{SW}}(\varphi,a)\|_{L^{2}}
    Qμ(η,φ)L2\displaystyle\|Q_{\mu}(\eta,\varphi)\|_{L^{2}} \displaystyle\leq Cεη3/2+γ¯(φHe1+QSW(φ,a)L2).\displaystyle C\varepsilon\|\eta\|_{3/2+\underline{\gamma}}\left(\|\varphi\|_{H^{1}_{e}}+\|Q_{\text{SW}}(\varphi,a)\|_{L^{2}}\right).
Proof.

(I) Configurations on YY satisfy the interpolation inequality

uL4CuL21/4uL1,23/4.\|u\|_{L^{4}}\leq C\|u\|_{L^{2}}^{1/4}\|u\|_{L^{1,2}}^{3/4}. (9.12)

Since |ζ±|2|ζ±|1|\zeta^{\pm}|^{2}\leq|\zeta^{\pm}|\leq 1, and |dζ±|CRε1|d\zeta^{\pm}|\leq CR_{\varepsilon}^{-1} is bounded by the weight on the L2L^{2}-terms in the Hε1H^{1}_{\varepsilon}-norm, applying this shows, e.g.

QSW(ζ+φ,ζ+a)L2\displaystyle\|Q_{\text{SW}}(\zeta^{+}\varphi,\zeta^{+}a)\|_{L^{2}} \displaystyle\leq Cζ+φL21/4ζ+φL1,23/4ζ+aL21/4ζ+aL1,23/4\displaystyle C\|\zeta^{+}\varphi\|_{L^{2}}^{1/4}\ \|\zeta^{+}\varphi\|^{3/4}_{L^{1,2}}\ \|\zeta^{+}a\|^{1/4}_{L^{2}}\ \|\zeta^{+}a\|^{3/4}_{L^{1,2}}
\displaystyle\leq Cmax(ζ+Rε1/2)(φ,a)Hε12\displaystyle C\cdot\text{max}(\zeta^{+}R_{\varepsilon}^{1/2})\cdot\|(\varphi,a)\|^{2}_{H^{1}_{\varepsilon}}
\displaystyle\leq Cε1/3γ(φ,a)Hε12.\displaystyle C\varepsilon^{1/3-\gamma}\|(\varphi,a)\|^{2}_{H^{1}_{\varepsilon}}.

because the Hε1H^{1}_{\varepsilon}-norm dominates the L2L^{2}-norm with an extra weight of RεR_{\varepsilon}. On the support of ζ\zeta^{-}, the proof is the same but the fact that QSWQ_{\text{SW}} has at most one factor in SReS^{\text{Re}} means the weight gives an extra factor of ε1/4|Φτ|1/4Rε1/4ε1/4Rε1/8\varepsilon^{1/4}|\Phi_{\tau}|^{-1/4}R_{\varepsilon}^{1/4}\leq\varepsilon^{1/4}R_{\varepsilon}^{1/8} for QSWImQ_{\text{SW}}^{\text{Im}} on supp(ζ)\text{supp}(\zeta^{-}) and ε1/2|Φτ|1/2<ε1/4\varepsilon^{1/2}|\Phi_{\tau}|^{-1/2}<\varepsilon^{1/4} for QSWReQ_{\text{SW}}^{\text{Re}}.

(II) Follows from the same considerations as Lemmas 9.2 and 9.1 (with (φ,a)(\varphi,a) replacing (φ1,a1)(\varphi_{1},a_{1}) in 9.5), and employing Lemma 6.14 in place of Lemma 6.12, then invoking the assumption that ξ3/2+γ¯<1\|\xi\|_{3/2+\underline{\gamma}}<1.

(III) Using Taylor’s theorem on the terms involving ξ\xi, then pulling out a factor of ξ3/2+γ¯=εη3/2+γ¯\|\xi\|_{3/2+\underline{\gamma}}=\varepsilon\|\eta\|_{3/2+\underline{\gamma}} as in the proof of Lemma 6.14 reduces to terms of the form in (I) or of the form ε1|Φτ||φ|\varepsilon^{-1}|\Phi_{\tau}||\varphi| which are bounded by the Hε1H^{1}_{\varepsilon}-norm directly. ∎

10. Contraction Subspaces

This section and the next complete the bulk of the proof of Theorem 1.6. This is done by constructing three linear parametrices, to be denoted by Pξ,P+,PP_{\xi},P^{+},P^{-}, one each for the three steps of the cyclic iteration described in Section 2.4. These three parametrices are combined into a (non-linear) approximate inverse 𝔸:L2ε1(τ×𝒳τ)\mathbb{A}:L^{2}\to\mathbb{H}^{1}_{\varepsilon}(\mathcal{E}_{\tau}\times\mathcal{X}_{\tau}) of 𝕊𝕎¯\overline{\mathbb{SW}} as in (8.14), which is the non-linear analogue of (2.23). Then define

T=Id𝔸𝕊𝕎¯Λ,T=\text{Id}-\mathbb{A}\circ\overline{\mathbb{SW}}_{\Lambda}, (10.1)

where 𝕊𝕎¯Λ=𝕊𝕎¯χε1Λ(τ)Φτ\overline{\mathbb{SW}}_{\Lambda}=\overline{\mathbb{SW}}-\chi^{-}\varepsilon^{-1}{\Lambda(\tau)\Phi_{\tau}}, so that each successive application of TT carries out a cycle of the alternating iteration (cf. 2.14).

Proposition 10.1.

There exist closed subspaces T(ε1(τ×𝒳τ)\mathcal{H}\subseteq T(\mathbb{H}^{1}_{\varepsilon}(\mathcal{E}_{\tau}\times\mathcal{X}_{\tau}), and 𝔏𝕃ε2\mathfrak{L}\subseteq\mathbb{L}^{2}_{\varepsilon}, and a closed ball 𝒞\mathcal{C}\subseteq\mathcal{H} containing 0 such that the following hold for h𝒞h\in\mathcal{C} and NN\in\mathbb{N}, provided ε0,τ0\varepsilon_{0},\tau_{0} are sufficiently small.

  1. (A)

    The restriction T:𝒞𝒞T:\mathcal{C}\to\mathcal{C} is continuous and depends smoothly on (ε,τ)(\varepsilon,\tau).

  2. (B)

    𝕊𝕎¯Λ(TNh)𝔏δN𝕊𝕎¯Λ(h)𝔏,\|\overline{\mathbb{SW}}_{\Lambda}(T^{N}h)\|_{\mathfrak{L}}\leq\delta^{N}\|\overline{\mathbb{SW}}_{\Lambda}(h)\|_{\mathfrak{L}},

  3. (C)

    Th1Th2Cδh1h2\|Th_{1}-Th_{2}\|_{\mathcal{H}}\leq C\sqrt{\delta}\|h_{1}-h_{2}\|_{\mathcal{H}}

where δ=ε1/48\delta=\varepsilon^{1/48}. In particular, TT is a contraction for ε\varepsilon sufficiently small.

Note the proposition implicitly uses the exponential map from Section 6.1 to conflate an open subset of Tε1T\mathbb{H}^{1}_{\varepsilon} with ε1\mathbb{H}^{1}_{\varepsilon} so that (10.1) makes sense. The remainder of the current section constructs the spaces ,𝔏\mathcal{H},\mathfrak{L}, and Section 11 constructs the three parametrices to complete the proof of the proposition. Theorem 1.6 is deduced as a consequence of Proposition 10.1 in Section 12.

10.1. Three Fourier Regimes

The main challenge in the proof of Proposition 10.1 is controlling the loss of regularity of the deformation operator T¯Φτ\underline{T}_{\Phi_{\tau}} from Section (6.2). This requires careful analysis of three regimes of Fourier modes, which are dealt with differently; these are associated to three different length scales on YY.

There are two different spaces of sections on 𝒵τ\mathcal{Z}_{\tau} with Fourier decompositions, and these Fourier decompositions are linked to length scales on YY by two distinct mechanisms. First, the intrinsic concentration property of the obstruction (recall Proposition 5.3 and Lemma 5.4) links the length scales to the Fourier modes in Ob(𝒵τ)L2(𝒵τ;𝒮τ)\text{\bf Ob}(\mathcal{Z}_{\tau})\simeq L^{2}(\mathcal{Z}_{\tau};\mathcal{S}_{\tau}). Second, by fiat, the use of mode-dependent deformations (Section 6.3) links length scales on YY to the Fourier modes of a deformation ηL2,2(𝒵τ;N𝒵τ)\eta\in L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}).

The three length scales rcε2/3γ,cε2/3γrcε1/2r\leq c\varepsilon^{2/3-\gamma},c\varepsilon^{2/3-\gamma}\leq r\leq c\varepsilon^{1/2} and rcε1/2r\geq c\varepsilon^{1/2} of the inside, the outside, and the neck regions respectively give rise to three corresponding Fourier regimes.

Definition 10.2.

For a fixed γ<<1\gamma<<1, set

Llow=ε(1/2+γ)\displaystyle L^{\text{low}}=\varepsilon^{-(1/2+\gamma)} Lmed=ε2/3\displaystyle L^{\text{med}}=\varepsilon^{-2/3}

and for \ell\in\mathbb{Z} define

πlow(eit)\displaystyle{\pi}^{\text{low}}(e^{i\ell t}) =\displaystyle= {eit||Llow0||>Llowπmed(eit)={eitLlow<||Lmed0||>Lmed,||Llow\displaystyle\begin{cases}e^{i\ell t}\hskip 14.22636pt|\ell|\leq L^{\text{low}}\\ 0\hskip 17.07182pt\ \ |\ell|>L^{\text{low}}\end{cases}\hskip 28.45274pt\pi^{\text{med}}(e^{i\ell t})=\begin{cases}e^{i\ell t}\hskip 14.22636ptL^{\text{low}}<|\ell|\leq L^{\text{med}}\\ 0\hskip 22.76228pt|\ell|>L^{\text{med}},|\ell|\leq L^{\text{low}}\end{cases}
πhigh(eit)\displaystyle\pi^{\text{high}}(e^{i\ell t}) =\displaystyle= (Idπmedπlow)eit.\displaystyle(\text{Id}-\pi^{\text{med}}-\pi^{\text{low}})e^{i\ell t}.

We write e.g. Ψlow:=πlowobτ1(Ψ)\Psi^{\text{low}}:=\pi^{\text{low}}\circ\text{ob}_{\tau}^{-1}(\Psi) as shorthand for the projections of an obstruction element ΨOb(𝒵τ)\Psi\in\text{Ob}(\mathcal{Z}_{\tau}) to the corresponding Fourier regimes.


Lemma 5.4 yields the following statements for the projection operators of Definition 10.2: for ε\varepsilon sufficiently small, suppose that ψ1,ψ2\psi_{1},\psi_{2} satisfy dist(𝒵τ,supp(ψi))Ri\text{dist}(\mathcal{Z}_{\tau},\text{supp}(\psi_{i}))\geq R_{i} for i=1,2i=1,2 where R1=cε1/2R_{1}=c\varepsilon^{1/2} and R2=cε2/3γR_{2}=c\varepsilon^{2/3-\gamma}. Then for any MM\in\mathbb{N}

(1πlow)obτ1Πτ(ψ1)L2(𝒵τ)\displaystyle\|(1-\pi^{\text{low}})\circ\text{ob}_{\tau}^{-1}\circ\Pi_{\tau}(\psi_{1})\|_{L^{2}(\mathcal{Z}_{\tau})} \displaystyle\leq CMεMψ1L2(Y)\displaystyle C_{M}\varepsilon^{M}\|\psi_{1}\|_{L^{2}(Y)} (10.2)
πhighobτ1Πτ(ψ2)L2(𝒵τ)\displaystyle\|\pi^{\text{high}}\circ\text{ob}_{\tau}^{-1}\circ\Pi_{\tau}(\psi_{2})\|_{L^{2}(\mathcal{Z}_{\tau})} \displaystyle\leq CMεMψ2L2(Y).\displaystyle C_{M}\varepsilon^{M}\|\psi_{2}\|_{L^{2}(Y)}. (10.3)

hold for constants CMC_{M}. Specifically, these are obtained by applying Lemma 5.4 with taking γ=γ1\gamma=\gamma_{1} in the statement of the lemma, where γ1\gamma_{1} satisfies 23γ=23(11γ)-\tfrac{2}{3}-\gamma=-\tfrac{2}{3}(\tfrac{1}{1-\gamma^{\prime}}) for γ\gamma as in Definition 10.2.

Sections 5.2 and 6.2 described respectively two invertible elliptic operators, either of which can be used to cancel the obstruction:

T¯Φτ:L1/2,2(𝒵τ;N𝒵τ)\displaystyle\underline{T}_{\Phi_{\tau}}:L^{1/2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) \displaystyle\longrightarrow L2(𝒵τ;Sτ)\displaystyle L^{2}(\mathcal{Z}_{\tau};S_{\tau}) (10.4)
obτ1:𝒳τ\displaystyle\text{ob}_{\tau}^{-1}\not{D}:\ \ \ \ \ \ \ \ \ \ \ \mathcal{X}_{\tau}\ \ \ \ \ \ \ \ \displaystyle\longrightarrow L2(𝒵τ;Sτ).\displaystyle L^{2}(\mathcal{Z}_{\tau};S_{\tau}). (10.5)

The first is an isomorphism by Corollary 6.10, the second by Lemma 5.6. The pre-images of the three Fourier regimes in the codomain give rise to three corresponding regimes in each L1/2,2(𝒵τ;N𝒵τ)L^{1/2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) and 𝒳τ\mathcal{X}_{\tau}.

(10.4) is used to cancel the low and medium modes, while (10.5) is used to cancel the high modes. There is an important and delicate balance that must be struck here. The higher the Fourier mode, the more extreme the loss of regularity of (10.4) is, thus there is an upper limit on the modes for which (10.4) may be used to cancel the obstruction, else the alternating iteration will not converge. Conversely, there is a lower limit to the modes which may be solved for using (10.5). Because solutions of (10.5) grow across the neck region rather than decay, the modes solved for this way must be smaller than the dominant error by at least that growth factor. (10.2) and (10.3) provide such estimates, but only in the case of spinors with support restricted away from 𝒵τ\mathcal{Z}_{\tau}. The support of error terms coming from the cut-off functions dχ±d\chi^{\pm} in the alternating method therefore yield a lower bound on the range of modes where using (10.5) is permissible 333This lower bound is dictated by the somewhat arbitrary choice that dχ±d\chi^{\pm} are supported where r=(ε1/2)r=(\varepsilon^{1/2}) and r=O(ε2/3γ)r=O(\varepsilon^{2/3-\gamma}). Different choices, however, also change the powers of ε\varepsilon appearing elsewhere (e.g. Theorem 7.5). Different choices result in different ranges for the three regimes in Definition 10.2, but the central issue cannot be avoided. The key point is that together, the regions enclosed by these upper and lower bounds cover the entire spectrum in L2(𝒵τ;Sτ)L^{2}(\mathcal{Z}_{\tau};S_{\tau}).

Next, we will define a combined space and operator that cancels the obstruction using (10.4) for the low and medium modes and (10.5) in the high modes. First, however, one minor alteration is required. To achieve control on higher Sobolev norms, we replace (10.4) by an auxiliary operator T¯τ\underline{T}^{\circ}_{\tau} which is a small perturbation of it, constructed as follows. In Fermi coordinates (Definition 3.9) and the accompanying trivialization (Lemma 3.10) on Nr0(𝒵τ)N_{r_{0}}(\mathcal{Z}_{\tau}), smooth objects may be decomposed using Fourier modes in the tt-direction, leading to families of Fourier series smoothly parameterized by the normal coordinates (x,y)(x,y). Since d𝔻h0(η,0)\text{d}{\mathbb{D}}_{h_{0}}(\eta,0) is supported in Nr0(𝒵0)N_{r_{0}}(\mathcal{Z}_{0}), we may define T¯τ\underline{T}_{\tau}^{\circ} by restricting these Fourier modes. Write gτ=g0+h(t,x,y)g_{\tau}=g_{0}+h(t,x,y) where g0g_{0} is the product metric in Fermi coordinates, and set

g=g0+πlow(h)Φτ:=πlow(Φτ)g^{\circ}=g_{0}+\pi^{\text{low}}(h)\hskip 56.9055pt\Phi_{\tau}^{\circ}:=\pi^{\text{low}}(\Phi_{\tau}) (10.6)

where πlow\pi^{\text{low}} from Definition 10.2 applied for every fixed (x,y)Dr0(x,y)\in D_{r_{0}}. Let T¯τ\underline{T}_{\tau}^{\circ} be the deformation operator defined analogously to T¯Φτ\underline{T}_{\Phi_{\tau}} (Theorem 6.5) using (10.6) in place of gτ,Φτg_{\tau},\Phi_{\tau}. By construction, T¯τ\underline{T}^{\circ}_{\tau} obeys the hypotheses of Corollary (6.8) with M=ε1/2M=\varepsilon^{-1/2}, since gτ,Φτg_{\tau},\Phi_{\tau} are smooth in the tt-direction.

Corollary 10.3.

Let 𝔱τ\mathfrak{t}_{\tau}^{\circ} be such that T¯Φτ=T¯τ+𝔱τ\underline{T}_{\Phi_{\tau}}=\underline{T}_{\tau}^{\circ}+\mathfrak{t}^{\circ}_{\tau}. For any M>0M>0, there is a constant CMC_{M} depending on MM such that

𝔱τ(η)2CMεMη1/2,2.\|\mathfrak{t}^{\circ}_{\tau}(\eta)\|_{2}\leq C_{M}\varepsilon^{M}\|\eta\|_{1/2,2}.

In particular for ε0\varepsilon_{0} sufficiently small, T¯τ\underline{T}_{\tau}^{\circ} is invertible, and the estimates (6.16) hold uniformly in τ\tau. In this case, if T¯τ(ηlow)=Ψlow\underline{T}_{\tau}^{\circ}(\eta^{\text{low}})=\Psi^{\text{low}} and T¯τ(ηmed)=Ψmed\underline{T}_{\tau}^{\circ}(\eta^{\text{med}})=\Psi^{\text{med}}, then

ηlowm+1/2,2Cm(Llow)mΨlowL2ηmedm+1/2,2Cm(Lmed)mΨmedL2\|\eta^{\text{low}}\|_{m+1/2,2}\leq C_{m}(L^{\text{low}})^{-m}\|\Psi^{\text{low}}\|_{L^{2}}\hskip 56.9055pt\|\eta^{\text{med}}\|_{m+1/2,2}\leq C_{m}(L^{\text{med}})^{-m}\|\Psi^{\text{med}}\|_{L^{2}} (10.7)

uniformly in τ\tau for any m>0m>0. ∎

We use ηlow,ulow\eta^{\text{low}},u^{\text{low}} to denote deformations and spinors whose images under (10.4) and (10.5) fall in the corresponding regimes in L2(𝒵τ;Sτ)L^{2}(\mathcal{Z}_{\tau};S_{\tau}), and likewise for the medium and high regimes. Note that ηlow\eta^{\text{low}} need not precisely have modes restricted to the low regime in L2(𝒵τ;N𝒵τ)L^{2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau}) since T¯τ\underline{T}_{\tau}^{\circ} does not preserve support in Fourier space; for the purposes of estimates, however, (10.7) says this is effectively true.

Definition 10.4.

Let 𝔚ε,τL1/2,2(𝒵τ;N𝒵τ)𝒳τ\mathfrak{W}_{\varepsilon,\tau}\subseteq L^{1/2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau})\oplus\mathcal{X}_{\tau} be the subspace defined as the image of the following composition.

L2(𝒵τ;Sτ)L^{2}(\mathcal{Z}_{\tau};S_{\tau})L2(𝒵τ;Sτ)L2(𝒵τ;Sτ)\begin{matrix}L^{2}(\mathcal{Z}_{\tau};S_{\tau})\\ \oplus\\ L^{2}(\mathcal{Z}_{\tau};S_{\tau})\end{matrix}L1/2,2(𝒵τ;N𝒵τ)𝒳τ.\begin{matrix}L^{1/2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau})\\ \oplus\\ \mathcal{X}_{\tau}.\end{matrix}(πmed,πhigh)(\pi^{\text{med}},\pi^{\text{high}})(T¯τ,Aτ1obτ)(\underline{T}_{\tau}^{\circ},\not{D}^{-1}_{A_{\tau}}\text{ob}_{\tau})

Equip 𝔚ε,τ\mathfrak{W}_{\varepsilon,\tau} with the norm

(ηlow,ηmed,u)𝔚:=(ηlow1/2,22+ε1/3ηmed1/2,22+ε4/3uHe1)1/2.\|(\eta^{\text{low}},\eta^{\text{med}},u)\|_{\mathfrak{W}}:=\left(\|\eta^{\text{low}}\|_{1/2,2}^{2}\ +\ \varepsilon^{-1/3}\|\eta^{\text{med}}\|_{1/2,2}^{2}\ +\ \varepsilon^{-4/3}\|u\|_{H^{1}_{e}}\right)^{1/2}. (10.8)

By construction, (T¯τ,obτ1Aτ):𝔚ε,τL2(𝒵τ;Sτ)(\underline{T}_{\tau}^{\circ},\text{ob}_{\tau}^{-1}\not{D}_{A_{\tau}}):\mathfrak{W}_{\varepsilon,\tau}\to L^{2}(\mathcal{Z}_{\tau};S_{\tau}) is an isomorphism. The first two components give rise to a corresponding set of renormalized deformations ξ=εη\xi=\varepsilon\eta.

Here, we may alter πmed\pi^{\text{med}} so that the projection smoothly interpolates between 0 and 11 on the mode with Fourier index Lmed\left\lceil L^{\text{med}}\right\rceil. The same applies to the norm between the low and medium modes. It may therefore be arranged that the family 𝔚ε,τ\mathfrak{W}_{\varepsilon,\tau} form a smooth vector bundle over pairs (ε,τ)(\varepsilon,\tau).

10.2. Contraction Subspaces

This subsection defines the subspaces ,𝔏\mathcal{H},\mathfrak{L} whose existence is asserted in Proposition 10.1. The definitions have two tasks: (1) finding a closed subspace Tε1(τ×𝒳τ)\mathcal{H}\subseteq T\mathbb{H}^{1}_{\varepsilon}(\mathcal{E}_{\tau}\times\mathcal{X}_{\tau}) on which the linearization has index 0 (cf. Proposition 2.4), and (2) defining correctly weighted norms.

The upcoming weighted norms are chosen in hindsight and are necessary for the map TT to satisfy Items (B) and (C) of Proposition 10.1. The proof that a single application of TT reduces the error as in (B) relies the error having a certain effective support property (recall Definition 4.6): the majority of its effective support must be clustered where dχ+0d\chi^{+}\neq 0. The weights in the upcoming norms impose this property (without Difficulty II from Section 2.5, the weighted norms could simply weight the integral over support of dχ+d\chi^{+} differently — the use of effective support should be thought of as imposing a slight generalization of this).

Recall that Tε1(τ×𝒳τ)T\mathbb{H}^{1}_{\varepsilon}(\mathcal{E}_{\tau}\times\mathcal{X}_{\tau}) is isomorphic via Υ¯\underline{\Upsilon} (Lemma 8.2) to a trivial bundle with fiber Hε1,+Hε1,L2,2(𝒵τ;N𝒵τ)𝒳τH^{1,+}_{\varepsilon}\oplus H^{1,-}_{\varepsilon}\oplus L^{2,2}(\mathcal{Z}_{\tau};N\mathcal{Z}_{\tau})\oplus\mathcal{X}_{\tau}. Notice that 𝔚ε,τTε1(τ×𝒳τ)\mathfrak{W}_{\varepsilon,\tau}\subseteq T\mathbb{H}^{1}_{\varepsilon}(\mathcal{E}_{\tau}\times\mathcal{X}_{\tau}) despite the lower regularity in (10.8), since only a finite-dimensional subspace of deformations is included.

Definition 10.5.

Set

+\displaystyle\mathcal{H}^{+} :=\displaystyle:= {(Φ1,A1)1(g𝟙+)|gL2(Y+)}Hε1,+\displaystyle\left\{\mathcal{L}_{(\Phi_{1},A_{1})}^{-1}(g\mathbb{1}^{+})\ \ \Big{|}\ \ g\in L^{2}(Y^{+})\right\}\subseteq H^{1,+}_{\varepsilon}
\displaystyle\mathcal{H}^{-} :=\displaystyle:= {(Φτ,Aτ)1((1Πτ)g𝟙)|gL2(Y)}Hε1,\displaystyle\left\{\mathcal{L}_{(\Phi_{\tau},A_{\tau})}^{-1}((1-\Pi_{\tau})g\mathbb{1}^{-})\ \ \Big{|}\ \ g\in L^{2}(Y^{-})\right\}\subseteq H^{1,-}_{\varepsilon}

where \mathcal{H}^{-} uses the solution in Hε1,H^{1,-}_{\varepsilon} that is L2L^{2}-orthgonal to Φτ\Phi_{\tau} on YY^{-}. Equip these with the norms

(φ,a)+\displaystyle\|(\varphi,a)\|_{\mathcal{H}^{+}} =\displaystyle= ε1/12γ(Φ1,A1)(φ,a)𝟙+L2\displaystyle\varepsilon^{-1/12-\gamma}\|\mathcal{L}_{(\Phi_{1},A_{1})}(\varphi,a)\mathbb{1}^{+}\|_{L^{2}} (10.9)
(ψ,b)\displaystyle\|(\psi,b)\|_{\mathcal{H}^{-}} =\displaystyle= (ψRerHe12+ενψRe𝟙rHν12+ε1/2(0,ψIm,b)Hε1,)1/2.\displaystyle\left(\|\psi^{\text{Re}}\|^{2}_{rH^{1}_{e}}\ +\ \varepsilon^{\nu}\|\psi^{\text{Re}}\mathbb{1}^{-}\|^{2}_{rH^{1}_{-\nu}}\ +\ \varepsilon^{-1/2}\|(0,\psi^{\text{Im}},b)\|_{H^{1,-}_{\varepsilon}}\right)^{1/2}. (10.10)

where ν=ν=12γ\nu=\nu^{-}=\tfrac{1}{2}-\gamma^{-}.

Definition 10.6.

Define

ε,τ\displaystyle\mathcal{H}_{\varepsilon,\tau} :=\displaystyle:= +𝔚ε,τ𝔏ε,τ:=L2(Y)\displaystyle\mathcal{H}^{+}\oplus\mathcal{H}^{-}\oplus\mathfrak{W}_{\varepsilon,\tau}\oplus\mathbb{R}\hskip 51.21504pt\mathfrak{L}_{\varepsilon,\tau}:=L^{2}(Y)

and equip these with the norms

(h1,h2,η,μ)\displaystyle\|(h_{1},h_{2},\eta,\mu)\|_{\mathcal{H}} :=\displaystyle:= (h1+2+h22+η𝔚2+ε2|μ|2)1/2\displaystyle\left(\|h_{1}\|_{\mathcal{H}^{+}}^{2}\ +\ \|h_{2}\|_{\mathcal{H}^{-}}^{2}\ +\ \|\eta\|_{\mathfrak{W}}^{2}\ +\ \varepsilon^{-2}|\mu|^{2}\right)^{1/2}\vspace{.5cm} (10.11)
𝔢𝔏\displaystyle\vskip 6.0pt plus 2.0pt minus 2.0pt\|\mathfrak{e}\|_{\mathfrak{L}} :=\displaystyle:= (ε2/122γ𝔢𝟙+L22+𝔢Re𝟙L22+εν𝔢Re𝟙Lν22\displaystyle\Big{(}\varepsilon^{-2/12-2\gamma}\|\mathfrak{e}\mathbb{1}^{+}\|^{2}_{L^{2}}\ +\ \|\mathfrak{e}^{\text{Re}}\mathbb{1}^{-}\|^{2}_{L^{2}}\ +\ \varepsilon^{\nu}\|\mathfrak{e}^{\text{Re}}\mathbb{1}^{-}\|^{2}_{L^{2}_{-\nu}}\vskip 3.0pt plus 1.0pt minus 1.0pt (10.13)
+ε1/3𝔢Im𝟙L22+ε1/3πmedΠτ(𝔢)L22)1/2\displaystyle\ \ \ +\ \ \varepsilon^{-1/3}\|\mathfrak{e}^{\text{Im}}\mathbb{1}^{-}\|^{2}_{L^{2}}\ \ +\ \ \varepsilon^{-1/3}\|\pi^{\text{med}}\Pi_{\tau}(\mathfrak{e})\|^{2}_{L^{2}}\Big{)}^{1/2}

where 𝔢=(𝔢Re,𝔢Im)L2(SRe)L2(SImΩ)\mathfrak{e}=(\mathfrak{e}^{\text{Re}},\mathfrak{e}^{\text{Im}})\in L^{2}(S^{\text{Re}})\oplus L^{2}(S^{\text{Im}}\oplus\Omega). These families of spaces form smooth Banach vector bundles over pairs (ε,τ)(\varepsilon,\tau) for ε,τ\varepsilon,\tau sufficiently small.

Note that the weighted terms dictate that the ν\nu-weighted term in the norms is larger than the unweighted term by at most a factor of εν/2\varepsilon^{-\nu/2}, despite the fact that suprν=εν(2/3γ)\sup r^{-\nu}=\varepsilon^{-\nu(2/3-\gamma)} on supp(𝟙)\text{supp}(\mathbb{1}^{-}). This shows that configurations in 𝔏\mathfrak{L} have ε1/2\varepsilon^{1/2}-effective support. Note also that Theorem 7.5 implies that the pullback (10.9) is indeed a norm.

The following lemma shows the linearized Seiberg–Witten equations are bounded on the above spaces with operator norm independent of ε\varepsilon (up to a factor of γ)\gamma). This assertion is non-trivial: it means that the new norms control the loss of regularity of the deformation operator. Indeed, if one simply uses the L1/2,2,Hε1,±L^{1/2,2},H^{1,\pm}_{\varepsilon} and L2L^{2}-norms, the operator norm is only bounded by ε4/3\varepsilon^{-4/3}.

Lemma 10.7.

There is a constant CC independent of ε,τ\varepsilon,\tau such that the linearization d𝕊𝕎¯h1:𝔏\text{d}\overline{\mathbb{SW}}_{h_{1}}:\mathcal{H}\to\mathfrak{L} satisfies

d𝕊𝕎¯h1(h)𝔏Cεγh\|\text{d}\overline{\mathbb{SW}}_{h_{1}}(h)\|_{\mathfrak{L}}\leq C\varepsilon^{-\gamma}\|h\|_{\mathcal{H}} (10.14)

for some γ<<1\gamma<<1.

Proof.

By Proposition 8.7, the linearization on h=(ξ,φ,a,ψ,b,μ,u)h=(\xi,\varphi,a,\psi,b,\mu,u) where ξ=εη\xi=\varepsilon\eta may be written in the trivialization of Lemma 8.2 as

d𝕊𝕎¯h1(h)\displaystyle\text{d}\overline{\mathbb{SW}}_{h_{1}}(h) =\displaystyle= 1ε¯Φ1(ξ)+(Φ1,A1)(𝔭)+μχΦτε\displaystyle\tfrac{1}{\varepsilon}\underline{\mathcal{B}}_{\Phi_{1}}(\xi)\ +\ \mathcal{L}_{(\Phi_{1},A_{1})}(\mathfrak{p})\ +\ \mu\chi^{-}\frac{\Phi_{\tau}}{\varepsilon} (10.15)
=\displaystyle= dh0(η,0)+Ξ(η)+(Φ1,A1)(𝔭)+μχΦτε.\displaystyle\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)+\Xi(\eta)+\mathcal{L}_{(\Phi_{1},A_{1})}(\mathfrak{p})\ +\ \mu\chi^{-}\frac{\Phi_{\tau}}{\varepsilon}. (10.16)

where Ξ\Xi is as in (9.2), and 𝔭=χ+(φ,a)+χ(ψ+u,b)\mathfrak{p}=\chi^{+}(\varphi,a)+\chi^{-}(\psi+u,b) where (φ,a)Hε1,+(\varphi,a)\in H^{1,+}_{\varepsilon}, (ψ,b)Hε1,(\psi,b)\in H^{1,-}_{\varepsilon} and u𝒳τu\in\mathcal{X}_{\tau}. The proof now consists of four steps.

Step 1 ((φ,a)(\varphi,a) terms): Abbreviating (Φ1,A1)=\mathcal{L}_{(\Phi_{1},A_{1})}=\mathcal{L}, and using the definition of +\mathcal{H}^{+},

(χ+(φ,a))𝔏\displaystyle\|\mathcal{L}(\chi^{+}(\varphi,a))\|_{\mathfrak{L}} =\displaystyle= ε1/12γ(φ,a)𝟙+L2+dχ+φReL2+ενdχ+φReLν2\displaystyle\varepsilon^{-1/12-\gamma}\|\mathcal{L}(\varphi,a)\mathbb{1}^{+}\|_{L^{2}}+\|d\chi^{+}\varphi^{\text{Re}}\|_{L^{2}}+\ \varepsilon^{\nu}\|d\chi^{+}\varphi^{\text{Re}}\|_{L^{2}_{-\nu}}
+ε1/3dχ+(φIm,a)L2+ε1/3πmedΠτ(dχ+φRe)L2\displaystyle\ +\ \varepsilon^{-1/3}\|d\chi^{+}(\varphi^{\text{Im}},a)\|_{L^{2}}\ +\ \varepsilon^{-1/3}\|\pi^{\text{med}}\Pi_{\tau}(d\chi^{+}\varphi^{\text{Re}})\|_{L^{2}}\vskip 6.0pt plus 2.0pt minus 2.0pt
\displaystyle\leq (φ,a)++ε1/24γ(φ,a)𝟙+L2+dχ+φReL2\displaystyle\|(\varphi,a)\|_{\mathcal{H}^{+}}\ +\varepsilon^{1/24-\gamma}\|\mathcal{L}(\varphi,a)\mathbb{1}^{+}\|_{L^{2}}\ +\ \|d\chi^{+}\varphi^{\text{Re}}\|_{L^{2}}
O(ε3eεγ)(φ,a)H1,++O(εM)(φ,a)L2\displaystyle O(\varepsilon^{-3}e^{\varepsilon^{-\gamma}})\|(\varphi,a)\|_{H^{1,+}}\ +\ O(\varepsilon^{M})\|(\varphi,a)\|_{L^{2}}
\displaystyle\leq (φ,a)+.\displaystyle\|(\varphi,a)\|_{\mathcal{H}^{+}}.

where the five terms are bounded respectively as follows. The first by the definition (10.9) of the +\mathcal{H}^{+}-norm, the second is reduced to the case of the first by Lemma 7.12, and the third is identical to the second since dχ+d\chi^{+} is supported where r=O(ε1/2)r=O(\varepsilon^{1/2}). The fourth is exponentially small by Lemma (7.13) and Theorem 7.5 because the latter shows L2Hε1,++\|-\|_{L^{2}}\leq\|-\|_{H^{1,+}_{\varepsilon}}\leq\|-\|_{\mathcal{H}^{+}}, and the fifth is polynomially small by (10.2).

Step 2 ((ψ,b)(\psi,b) terms): Write (Φ1,A1)=(Φ0,A0)+K1\mathcal{L}_{(\Phi_{1},A_{1})}=\mathcal{L}_{(\Phi_{0},A_{0})}+K_{1}. The boundedness of the (Φ0,A0)(χ(ψ+u,b))\mathcal{L}_{(\Phi_{0},A_{0})}(\chi^{-}(\psi+u,b)) terms now follows similarly to the above, using the decay result of Lemma 4.7 in place of Lemma 7.12, and the boundedness of (7.11) and (4.11) for both ν=0,ν\nu=0,\nu^{-}. In this, the bound ψRe+urHε12h\|\psi^{\text{Re}}+u\|_{rH^{1}_{\varepsilon}}\leq 2\|h\|_{\mathcal{H}} is used, which comes from the ε4/3\varepsilon^{-4/3}-weight in (10.8),,

The only minor difference from Step 1 is for the πmed\pi^{\text{med}}-term in the 𝔏\mathfrak{L}-norm: this is zero on (ψ,b)(\psi,b) by definition, and on uu is dominated by the factor of ε4/3\varepsilon^{-4/3} in the u𝔚\|u\|_{\mathfrak{W}} norm. Finally, K1K_{1} is exponentially small on supp(𝟙+)\text{supp}(\mathbb{1}^{+}) by Lemma 7.13 (applied with Kε={rε2/3γ/2}K_{\varepsilon}=\{r\geq\varepsilon^{2/3-\gamma/2}\}).

Step 3 (μ\mu term): The term χε1μΦξ,τ\chi^{-}\varepsilon^{-1}\mu\Phi_{\xi,\tau} is obviously bounded by μ\|\mu\|_{\mathcal{H}} because the ε2\varepsilon^{-2} weight cancels the ε\varepsilon in the denominator, and ΦτL2L2ν\Phi_{\tau}\in L^{2}\cap L^{2}_{-\nu}. The πmed\pi^{\text{med}}-term is smaller, because πmed(χΦτ)L2=πmed((1χ)Φτ)L2Cε\|\pi^{\text{med}}(\chi^{-}\Phi_{\tau})\|_{L^{2}}=\|\pi^{\text{med}}((1-\chi^{-})\Phi_{\tau})\|_{L^{2}}\leq C\varepsilon using Lemma 4.5.

Step 4 (deformation terms): Proceeding now to the terms involving η=ε1ξ\eta=\varepsilon^{-1}\xi, Corollary 10.3 provides bounds on the higher Sobolev norms of η𝔚\eta\in\mathfrak{W}. Using these, Proposition 9.3 applied with ν=0,ν\nu=0,\nu^{-} and Proposition 9.1 show that

dh0(η,0)𝟙L2\displaystyle\|\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)\mathbb{1}^{-}\|_{L^{2}}\ \displaystyle\leq Cη1/2\displaystyle C\|\eta\|_{1/2} (10.17)
dh0(η,0)𝟙L2ν\displaystyle\|\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)\mathbb{1}^{-}\|_{L^{2}_{-\nu}} \displaystyle\leq C(εν(1/2+γ)ηlow1/2+Cε2ν/3ηmed1/2)Cεν/2γη𝔚\displaystyle C(\varepsilon^{-\nu(1/2+\gamma)}\|\eta^{\text{low}}\|_{1/2}+C\varepsilon^{-2\nu/3}\|\eta^{\text{med}}\|_{1/2})\leq C\varepsilon^{-\nu/2-\gamma}\|\eta\|_{\mathfrak{W}} (10.18)
Ξ(η)L2\displaystyle\|\Xi^{-}(\eta)\|_{L^{2}} \displaystyle\leq Cε11/12γ(ε1/2ηlow1/2+ε2/3ηmed1/2)Cε5/12γη𝔚.\displaystyle C\varepsilon^{11/12-\gamma}(\varepsilon^{-1/2}\|\eta^{\text{low}}\|_{1/2}+\varepsilon^{-2/3}\|\eta^{\text{med}}\|_{1/2})\leq C\varepsilon^{5/12-\gamma}\|\eta\|_{\mathfrak{W}}. (10.19)

Moreover, since rνε2ν/3εν/2εν/6r^{-\nu}\leq\varepsilon^{-2\nu/3}\leq\varepsilon^{-\nu/2}\varepsilon^{-\nu/6} on supp(𝟙)\text{supp}(\mathbb{1}^{-}), (10.19) shows that

εν/2Ξ(η)L2νCε4/12γη𝔚\varepsilon^{\nu/2}\|\Xi^{-}(\eta)\|_{L^{2}_{-\nu}}\leq C\varepsilon^{4/12-\gamma}\|\eta\|_{\mathfrak{W}} (10.20)

as well. Next, Proposition 9.2 (and the bounds of Corollary 10.3 again) shows that

dh0(η,0)𝟙++Ξ+(η)L2\displaystyle\|\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)\mathbb{1}^{+}+\Xi^{+}(\eta)\|_{L^{2}} \displaystyle\leq Cεγ(ε1/3η1+ε11/12η3/2+γ¯+εη2+ε19/12η5/2+γ¯)\displaystyle C\varepsilon^{-\gamma}\Big{(}\varepsilon^{1/3}\|\eta\|_{1}\ +\ \varepsilon^{11/12}\|\eta\|_{3/2+\underline{\gamma}}\ +\ \ \varepsilon\|\eta\|_{2}\ +\ \varepsilon^{19/12}\|\eta\|_{5/2+\underline{\gamma}}\Big{)} (10.21)
\displaystyle\leq Cε1/12γηlow1/2+Cηmed1/2\displaystyle C\varepsilon^{1/12-\gamma}\|\eta^{\text{low}}\|_{1/2}+C\|\eta^{\text{med}}\|_{1/2}
\displaystyle\leq Cε1/12γη𝔚.\displaystyle C\varepsilon^{1/12-\gamma}\|\eta\|_{\mathfrak{W}}. (10.22)

Together, (10.17)–(10.22) show that all but the πmed\pi^{\text{med}}-term in the 𝔏\mathfrak{L}-norm are bounded for dh0+Ξ\text{d}\not{\mathbb{D}}_{h_{0}}+\Xi.

For this final term, one has

πmedΠτdh0(η,0)𝟙L2πmedΠτdh0(η,0)L2+πmedΠτdh0(η,0)𝟙+L2\|\pi^{\text{med}}\Pi_{\tau}\text{d}_{h_{0}}\not{\mathbb{D}}(\eta,0)\mathbb{1}^{-}\|_{L^{2}}\ \leq\ \|\pi^{\text{med}}\Pi_{\tau}\text{d}_{h_{0}}\not{\mathbb{D}}(\eta,0)\|_{L^{2}}\ +\ \|\pi^{\text{med}}\Pi_{\tau}\text{d}_{h_{0}}\not{\mathbb{D}}(\eta,0)\mathbb{1}^{+}\|_{L^{2}} (10.23)

by the triangle inequality, and we bound each of these terms separately.

For the first term of (10.23),

ε1/6πmedΠτdh0(η,0)L2\displaystyle\varepsilon^{-1/6}\|\pi^{\text{med}}\Pi_{\tau}\text{d}_{h_{0}}\not{\mathbb{D}}(\eta,0)\|_{L^{2}} \displaystyle\leq ε1/6πmedT¯τ(η)L2+ε1/6𝔱τ(η)L2\displaystyle\varepsilon^{-1/6}\|\pi^{\text{med}}\underline{T}_{\tau}^{\circ}(\eta)\|_{L^{2}}\ +\ \varepsilon^{-1/6}\|\mathfrak{t}_{\tau}^{\circ}(\eta)\|_{L^{2}}
\displaystyle\leq Cηmed𝔚+O(εM)η𝔚\displaystyle C\|\eta^{\text{med}}\|_{\mathfrak{W}}+O(\varepsilon^{M})\|\eta\|_{\mathfrak{W}}

by Definition (10.4) and Lemma 10.3.

For the second term of (10.23), split η=ηlow+ηmed\eta=\eta^{\text{low}}+\eta^{\text{med}}; the portion involving ηmed\eta^{\text{med}} is obviously bounded because the same ε1/6\varepsilon^{-1/6} weight appears in (10.8) and in the 𝔏\mathfrak{L}-norm. Thus it suffices to consider the ηlow\eta^{\text{low}} terms. Write dh0=d+𝔡\text{d}\not{\mathbb{D}}_{h_{0}}=\text{d}^{\circ}\not{\mathbb{D}}+\mathfrak{d}^{\circ} where the latter are formed using the metrics of Lemma 10.3. Just as in that lemma, 𝔡CεM\|\mathfrak{d}\|\leq C\varepsilon^{M} and so the terms involving the latter are negligible. Recalling Proposition 5.3 now, πmed\pi^{\text{med}} is calculated by the sequence of inner-products Ψ,d(ηlow)𝟙+\langle\Psi_{\ell},\text{d}^{\circ}\not{\mathbb{D}}(\eta^{\text{low}})\mathbb{1}^{+}\rangle_{{\mathbb{C}}} for Llow<||LmedL^{\text{low}}<|\ell|\leq L^{\text{med}}. The second and third bullet points of Proposition 5.3 show that the ζ,ξ\zeta_{\ell},\xi_{\ell} terms of Ψ\Psi_{\ell} are smaller by a factor of ||1>ε1/2|\ell|^{-1}>\varepsilon^{-1/2}, which dominates the ε1/6\varepsilon^{-1/6} weight. For the leading order terms χΨ\chi\Psi_{\ell}^{\circ}, the restriction of Fourier modes on d\text{d}^{\circ}\not{\mathbb{D}} and Cauchy-Schwartz imply

χΨ,d(ηlow)𝟙+\displaystyle\langle\chi\Psi_{\ell}^{\circ},\text{d}^{\circ}\not{\mathbb{D}}(\eta^{\text{low}})\mathbb{1}^{+}\rangle_{{\mathbb{C}}} =\displaystyle= 0||2Llow\displaystyle 0\hskip 187.78836pt|\ell|\geq 2L^{\text{low}}
χΨ,d(ηlow)𝟙+\displaystyle\langle\chi\Psi_{\ell}^{\circ},\text{d}^{\circ}\not{\mathbb{D}}(\eta^{\text{low}})\mathbb{1}^{+}\rangle_{{\mathbb{C}}} \displaystyle\leq CΨL2(𝟙+)d(ηlow)L2(𝟙+)Llow||2Llow\displaystyle C\|\Psi_{\ell}^{\circ}\|_{L^{2}(\mathbb{1}^{+})}\|\text{d}^{\circ}\not{\mathbb{D}}(\eta^{\text{low}})\|_{L^{2}(\mathbb{1}^{+})}\hskip 28.45274ptL^{\text{low}}\leq|\ell|\leq 2L^{\text{low}}
\displaystyle\leq Cε1/62γε1/12γηlow1/2\displaystyle C\varepsilon^{1/6-2\gamma}\varepsilon^{1/12-\gamma}\|\eta^{\text{low}}\|_{1/2}

where the final line directly integrates (5.1) over supp(𝟙+)\text{supp}(\mathbb{1}^{+}), and invokes the second bound on the second term of (9.9) in Proposition 9.2, which applies equally well to d\text{d}^{\circ}\not{\mathbb{D}}. Returning to (10.23), the above combine to show

πmedΠτdh0(η,0)𝟙L2Cη𝔚\|\pi^{\text{med}}\Pi_{\tau}\text{d}\not{\mathbb{D}}_{h_{0}}(\eta,0)\mathbb{1}^{-}\|_{L^{2}}\leq C\|\eta\|_{\mathfrak{W}}

as desired. ∎

Lemma 9.4 may also be used to bound the nonlinear terms in terms of the ,𝔏\mathcal{H},\mathfrak{L}-norms.

Corollary 10.8.

Suppose that hNh_{N}\in\mathcal{H} satisfies hNCε1/20\|h_{N}\|_{\mathcal{H}}\leq C\varepsilon^{-1/20}. Then there is a C>0C>0 such that

  1. (A)

    (h)𝔏Cε2/12γh2\|\mathbb{Q}(h)\|_{\mathfrak{L}}\leq C\varepsilon^{2/12-\gamma}\|h\|_{\mathcal{H}}^{2},

  2. (B)

    dhN(h)𝔏Cε1/12γh\|\text{d}\mathbb{Q}_{h_{N}}(h)\|_{\mathfrak{L}}\leq C\varepsilon^{1/12-\gamma}\|h\|_{\mathcal{H}}

  3. (C)

    If h1,h2Cε1/20\|h_{1}\|_{\mathcal{H}},\|h_{2}\|_{\mathcal{H}}\leq C\varepsilon^{-1/20}, then (h1)(h2)𝔏Cε1/12γh1h2\|\mathbb{Q}(h_{1})-\mathbb{Q}(h_{2})\|_{\mathfrak{L}}\leq C\varepsilon^{1/12-\gamma}\|h_{1}-h_{2}\|_{\mathcal{H}}

hold uniformly in ε,τ\varepsilon,\tau.

Proof.

Let h=(ξ,φ,a,ψ,b,μ,u)h=(\xi,\varphi,a,\psi,b,\mu,u). Using that ψRe+urH12h\|\psi^{\text{Re}}+u\|_{rH^{1}}\leq 2\|h\|_{\mathcal{H}} as in the proof of the previous lemma, it suffices to prove the corollary with ψ\psi tacitly standing in for ψ+u\psi+u.

(A) Let 𝒬=QΦ+QA+Qa.φ+Qμ\mathcal{Q}=Q_{\Phi}+Q_{A}+Q_{a.\varphi}+Q_{\mu} be the latter four nonlinear terms in Proposition (8.8). Definition (10.8) means εη3/2+γ¯Cε1/2η𝔚\varepsilon\|\eta\|_{3/2+\underline{\gamma}}\leq C\varepsilon^{1/2}\|\eta\|_{\mathfrak{W}}, and Definition 10.5 and Theorem 7.5 imply (φ,a)H1ε(φ,a)\|(\varphi,a)\|_{H^{1}_{\varepsilon}}\leq\|(\varphi,a)\|_{\mathcal{H}}. That the same holds for (ψ,b)(\psi,b) is immediate from Definition 10.5. Item (II) and (III) of Lemma 9.4 therefore imply

𝒬(h)L2Cε1/2γh2.\|\mathcal{Q}(h)\|_{L^{2}}\leq C\varepsilon^{1/2-\gamma}\|h\|_{\mathcal{H}}^{2}.

and because the weights in the 𝔏\mathfrak{L}-norm are larger than the L2L^{2}-norm by at most ε1/6\varepsilon^{-1/6}, it follows that

𝒬(h)𝔏Cε3/12γh2.\|\mathcal{Q}(h)\|_{\mathfrak{L}}\leq C\varepsilon^{3/12-\gamma}\|h\|_{\mathcal{H}}^{2}.

Proceeding now to QSWQ_{\text{SW}}, there are four sub-terms coming from (χ+)2QSW(φ,a)(\chi^{+})^{2}Q_{\text{SW}}(\varphi,a), (χ)2QSW(ψ,b)(\chi^{-})^{2}Q_{\text{SW}}(\psi,b) and the cross-terms. We bound two pieces of each coming from the partition of unity ζ±\zeta^{\pm} in Proposition 9.4. By Item (I) of Proposition 9.4 and the above,

QSW(φ,a)ζ+𝔏ε1/6ζ+QSW(φ,a)ζ+L2Cε1/6ε4/12γ(φ,a)H1,+ε2Cε2/12γh2\|Q_{\text{SW}}(\varphi,a)\zeta^{+}\|_{\mathfrak{L}}\leq\varepsilon^{-1/6}\|\zeta^{+}Q_{\text{SW}}(\varphi,a)\zeta^{+}\|_{L^{2}}\leq C\varepsilon^{-1/6}\varepsilon^{4/12-\gamma}\|(\varphi,a)\|_{H^{1,+}_{\varepsilon}}^{2}\leq C\varepsilon^{2/12-\gamma}\|h\|_{\mathcal{H}}^{2}

and likewise for the other three sub-terms where ζ+>0\zeta^{+}>0.

Where ζ>0\zeta^{-}>0, the SReΩS^{\text{Re}}\oplus\Omega-components of (φ,a)(\varphi,a) are exponentially small by Lemma 7.13 applied with compact sets KεK_{\varepsilon} whose boundary lies halfway between supp(𝟙+)\text{supp}(\mathbb{1}^{+}) and supp(dζ+)\text{supp}(d\zeta^{+}). In this same region, the SImΩS^{\text{Im}}\oplus\Omega-components of (ψ,b)(\psi,b) are smaller by a factor of ε1/6\varepsilon^{1/6} by (10.10). Thus in this region, the right colum of Item (I) of Lemma 9.4 provdies the necessary bounds (in fact with ε3/12γ)\varepsilon^{3/12-\gamma}).

(B) The derivative dhN(h)\text{d}_{h_{N}}\mathbb{Q}(h) is given by the five terms of Proposition 8.8 viewed as multilinear functions of the arguments, with precisely one argument being chosen from the components of hh. The proof of the bound in (A) applies equally well in the bilinear setting to show that

dhN(h)𝔏Cε2/12γhNh\|\text{d}\mathbb{Q}_{h_{N}}(h)\|_{\mathfrak{L}}\leq C\varepsilon^{2/12-\gamma}\|h_{N}\|_{\mathcal{H}}\|h\|_{\mathcal{H}}

and the assertion follows. More specifically, the requirement that hNCε1/20\|h_{N}\|\leq C\varepsilon^{-1/20}, means εηN3/2+γ¯C\varepsilon\|\eta_{N}\|_{3/2+\underline{\gamma}}\leq C, hence the assumptions of Items (II) and (III) in Lemma 9.4 are satisfied, and the terms of dhN\text{d}\mathbb{Q}_{h_{N}} multi-linear in hNh_{N} can be bounded just as in that lemma.

(C) Follows from Item (B) applied to the family configurations hN=th1+(1t)h2h_{N}=th_{1}+(1-t)h_{2} and integration. ∎

11. The Alternating Iteration

This section proves Proposition 10.1 by carrying out the cyclic iteration outlined in Section 2.4. This is done by constructing three parametrices Pξ,P+,PP_{\xi},P^{+},P^{-}, one corresponding to each of the three stages of the iteration (2.17).

11.1. The Deformation Step

This subsection constructs the deformation parametrix PξP_{\xi}, and establishes the first of the three induction steps in the cyclic iteration (2.17). This step (which corresponds to the horizontal arrow in the diagram below (2.17)) shows the singular set 𝒵ξ,τ\mathcal{Z}_{\xi,\tau} can be adjusted to effectively cancel the obstruction components of an error term, without the error term growing much larger (in particular, this requires controlling the key term discussed in Section 2.5.1).

The following proposition is applied to the error terms inductively, beginning with the error 𝔢1\mathfrak{e}_{1} of the initial approximate solutions (Φ1,A1)(\Phi_{1},A_{1}) in Theorem 7.5. For the remainder of Section 11, we fix, once and for all, a choice of (ε,τ)(0,ε0)×(τ0,τ0)(\varepsilon,\tau)\in(0,\varepsilon_{0})\times(-\tau_{0},\tau_{0}) and omit this dependence from the subscripts in the notation where no confusion will arise.

Let Π:L2Ob(𝒵τ)\Pi^{\perp}:L^{2}\to\text{Ob}^{\perp}(\mathcal{Z}_{\tau}) be the L2L^{2}-orthogonal projection, so that Πτ=(Π,πτ)\Pi_{\tau}=(\Pi^{\perp},\pi_{\tau}) in the orthogonal decomposition in Definition 5.1 where πτ\pi_{\tau} is the L2L^{2}-orthogonal projection to the span of Φτ\Phi_{\tau}. Define the deformation parametrix

Pξ:𝔏𝔚Pξ:=(T¯τ,Aτ)1Π𝟙P_{\xi}:\mathfrak{L}\longrightarrow\mathfrak{W}\hskip 56.9055ptP_{\xi}:=\left(\underline{T}_{\tau}^{\circ},\not{D}_{A_{\tau}}\right)^{-1}\circ\Pi^{\perp}\circ\mathbb{1}^{-} (11.1)

where (T¯τ,Aτ)(\underline{T}_{\tau}^{\circ},\not{D}_{A_{\tau}}) is the map from Definition 10.4, with the map ob now kept implicit in the notation. In the following proposition, 𝕊𝕎¯Λ\overline{\mathbb{SW}}_{\Lambda} is as in (11.30).

Proposition 11.1.

PξP_{\xi} is a linear operator uniformly bounded in ε,τ\varepsilon,\tau and satisfies the following property. If for NN\in\mathbb{N}, hNh_{N}\in\mathcal{H} is a configuration with

  1. (I)

    hNCεγ𝔢1𝔏\|h_{N}\|_{\mathcal{H}}\leq C\varepsilon^{-\gamma}\|\mathfrak{e}_{1}\|_{\mathfrak{L}}

  2. (II)

    𝕊𝕎¯Λ(hN)=𝔢N\overline{\mathbb{SW}}_{\Lambda}(h_{N})=\mathfrak{e}_{N}  (resp. d𝕊𝕎¯h1(hN)=𝔢N\text{d}\overline{\mathbb{SW}}_{h_{1}}(h_{N})=\mathfrak{e}_{N}) where 𝔢N𝔏CδN1𝔢1𝔏\|\mathfrak{e}_{N}\|_{\mathfrak{L}}\leq C\delta^{N-1}\|\mathfrak{e}_{1}\|_{\mathfrak{L}},

then the updated configuration

hN=(IdPξ𝕊𝕎¯Λ)hNh_{N}^{\prime}=(\text{Id}-P_{\xi}\circ\overline{\mathbb{SW}}_{\Lambda})h_{N} (11.2)

satisfies

𝕊𝕎¯Λ(hN)=(1Πτ)𝔢N+𝔤N+λε1Φτ+𝔢N+1\overline{\mathbb{SW}}_{\Lambda}(h^{\prime}_{N})=(1-\Pi_{\tau})\mathfrak{e}_{N}^{\prime}\ +\ \mathfrak{g}^{\prime}_{N}\ +\ \lambda\varepsilon^{-1}\Phi_{\tau}\ +\ \mathfrak{e}_{N+1}

(resp. the same for d𝕊𝕎¯h1\text{d}\overline{\mathbb{SW}}_{h_{1}}), where 𝔢N,𝔢N+1𝔏\mathfrak{e}_{N}^{\prime},\mathfrak{e}_{N+1}\in\mathfrak{L}, and λ\lambda\in\mathbb{R} obey

  1. (1)

    𝔢N𝔏Cεγ𝔢N𝔏\|\mathfrak{e}_{N}^{\prime}\|_{\mathfrak{L}}\leq C\varepsilon^{-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}},  and   Π(𝔢N)=0\Pi(\mathfrak{e}^{\prime}_{N})=0.

  2. (2)

    𝔤N𝔏Cεγ𝔢N𝔏\|\mathfrak{g}_{N}^{\prime}\|_{\mathfrak{L}}\leq C\varepsilon^{-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}  and  𝔤N=𝔤N𝟙+\mathfrak{g}_{N}^{\prime}=\mathfrak{g}_{N}^{\prime}\mathbb{1}^{+}.

  3. (3)

    𝔢N+1𝔏Cε1/48δ𝔢N𝔏\|\mathfrak{e}_{N+1}\|_{\mathfrak{L}}\leq C\varepsilon^{1/48}\delta\|\mathfrak{e}_{N}\|_{\mathfrak{L}}

  4. (4)

    |λ|Cε1γ𝔢N𝔏|\lambda|\leq C\varepsilon^{1-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}.

Moreover, hNh_{N}^{\prime} continues to satisfy (I).

Proof.

The error term may be written 𝔢N=eN+fN+gN+ΨN\mathfrak{e}_{N}=e_{N}+f_{N}+g_{N}+\Psi_{N} where

gN\displaystyle g_{N} :=\displaystyle:= 𝔢N𝟙+fN:=𝔢NIm𝟙+πmedΠ(𝔢ReN𝟙)\displaystyle\mathfrak{e}_{N}\mathbb{1}^{+}\hskip 105.2751pt\ f_{N}:=\mathfrak{e}_{N}^{\text{Im}}\mathbb{1}^{-}+\pi^{\text{med}}\Pi(\mathfrak{e}^{\text{Re}}_{N}\mathbb{1}^{-})
ΨN\displaystyle\Psi_{N} :=\displaystyle:= πhighΠ(𝔢ReN𝟙)eN:=(1Π)𝔢NRe𝟙+πlowΠ(𝔢NRe𝟙)+πτ(𝔢NRe𝟙).\displaystyle\pi^{\text{high}}\Pi(\mathfrak{e}^{\text{Re}}_{N}\mathbb{1}^{-})\hskip 71.13188pte_{N}:=(1-\Pi)\mathfrak{e}_{N}^{\text{Re}}\mathbb{1}^{-}+\pi^{\text{low}}\Pi(\mathfrak{e}_{N}^{\text{Re}}\mathbb{1}^{-})+\pi_{\tau}(\mathfrak{e}_{N}^{\text{Re}}\mathbb{1}^{-}).

The definition of the 𝔏\mathfrak{L}-norm in (10.5) implies that gNL2Cε1/12γ𝔢N𝔏\|g_{N}\|_{L^{2}}\leq C\varepsilon^{1/12-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}  ,  fNL2Cε1/6𝔢N𝔏\|f_{N}\|_{L^{2}}\leq C\varepsilon^{1/6}\|\mathfrak{e}_{N}\|_{\mathfrak{L}} and Π(eN+fN+ΨN)L2νCεν/2𝔢N𝔏\|\Pi(e_{N}+f_{N}+\Psi_{N})\|_{L^{2}_{-\nu}}\leq C\varepsilon^{-\nu/2}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}. Moreover, (10.3) implies ΨNL2Cε10𝔢N𝔏\|\Psi_{N}\|_{L^{2}}\leq C\varepsilon^{10}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}.

Define

(η,u):=Pξ(Π(eN+fN+ΨN))(\eta,u):=P_{\xi}(\Pi^{\perp}(e_{N}+f_{N}+\Psi_{N}))

so that

T¯(ηlow)\displaystyle\underline{T}^{\circ}(\eta^{\text{low}})\ =\displaystyle= Π(eN)=πlowΠ(𝔢ReN𝟙)\displaystyle\Pi^{\perp}(e_{N})\hskip 2.84544pt\ =\ \pi^{\text{low}}\Pi^{\perp}(\mathfrak{e}^{\text{Re}}_{N}\mathbb{1}^{-}) (11.3)
T¯(ηmed)\displaystyle\underline{T}^{\circ}(\eta^{\text{med}}) =\displaystyle= Π(fN)=πmedΠ(𝔢ReN𝟙)\displaystyle\Pi^{\perp}(f_{N})\hskip 2.84544pt\ =\ \pi^{\text{med}}\Pi^{\perp}(\mathfrak{e}^{\text{Re}}_{N}\mathbb{1}^{-}) (11.4)
Aτu\displaystyle\not{D}_{A_{\tau}}u =\displaystyle= Π(ΨN)=πhighΠ(𝔢ReN𝟙)\displaystyle\Pi^{\perp}(\Psi_{N})\ =\ \pi^{\text{high}}\Pi^{\perp}(\mathfrak{e}^{\text{Re}}_{N}\mathbb{1}^{-}) (11.5)

where η=ηlow+ηmed\eta=\eta^{\text{low}}+\eta^{\text{med}}. The above estimates and the uniform bounds on (T¯,)1(\underline{T}^{\circ},\not{D})^{-1} coming from Corollary 6.8 (the version in Corollary 6.10) and Corollary 5.6 show that η,u\eta,u satisfy η𝔚C𝔢N𝔏\|\eta\|_{\mathfrak{W}}\leq C\|\mathfrak{e}_{N}\|_{\mathfrak{L}} and u𝔚Cε8𝔢N𝔏\|u\|_{\mathfrak{W}}\leq C\varepsilon^{8}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}.

We now proceed to calculate 𝕊𝕎¯Λ(hN)\overline{\mathbb{SW}}_{\Lambda}(h^{\prime}_{N}), where hN=hN(η,u)h_{N}^{\prime}=h_{N}-(\eta,u) as in (11.2). First, with ξ=εη\xi=\varepsilon\eta as previously, we compute:

d𝕊𝕎¯h1(ξ,0,0,0,0)\displaystyle\text{d}\overline{\mathbb{SW}}_{h_{1}}(\xi,0,0,0,0) =\displaystyle= d(η,0)+Ξ+(η)+Ξ(η)\displaystyle\text{d}\not{\mathbb{D}}(\eta,0)+\Xi^{+}(\eta)+\Xi^{-}(\eta) (11.6)
=\displaystyle= T¯(η)+𝔱(η)+(1Π)dh0(η)+πτdh0(η)+Ξ+(η)+Ξ(η)\displaystyle\underline{T}^{\circ}(\eta)+\mathfrak{t}^{\circ}(\eta)+(1-\Pi)\text{d}\not{\mathbb{D}}_{h_{0}}(\eta)+\pi_{\tau}\text{d}\not{\mathbb{D}}_{h_{0}}(\eta)+\Xi^{+}(\eta)+\Xi^{-}(\eta)
d𝕊𝕎¯h1(0,0,0,0,u)\displaystyle\text{d}\overline{\mathbb{SW}}_{h_{1}}(0,0,0,0,u) =\displaystyle= Aτ(χu)+K1\displaystyle\not{D}_{A_{\tau}}(\chi^{-}u)+K_{1} (11.7)
=\displaystyle= Aτu(1χ)Aτu+dχ.u+K1(χu)\displaystyle\not{D}_{A_{\tau}}u-(1-\chi^{-})\not{D}_{A_{\tau}}u+d\chi^{-}.u+K_{1}(\chi^{-}u)

where K1K_{1} is as in Step 2 of the proof of Lemma 10.14.

With these, we now compute:

𝕊𝕎¯Λ(hN)\displaystyle\overline{\mathbb{SW}}_{\Lambda}(h_{N}^{\prime}) =\displaystyle= 𝕊𝕎¯(h1)+dh1𝕊𝕎(hN)dh1𝕊𝕎(ξ,u)+(hN)(μ+Λ)χΦτε\displaystyle\overline{\mathbb{SW}}(h_{1})+\text{d}_{h_{1}}\mathbb{SW}(h_{N})\ -\ \text{d}_{h_{1}}\mathbb{SW}(\xi,u)\ +\ \mathbb{Q}(h_{N}^{\prime})\ -\ (\mu+\Lambda)\chi^{-}\frac{\Phi_{\tau}}{\varepsilon} (11.9)
=\displaystyle= 𝕊𝕎¯Λ(hN)dh1𝕊𝕎(ξ,u)+(hN)(hN)\displaystyle\ \overline{\mathbb{SW}}_{\Lambda}(h_{N})\ -\ \text{d}_{h_{1}}\mathbb{SW}(\xi,u)\ +\ \mathbb{Q}(h^{\prime}_{N})-\mathbb{Q}(h_{N})
=\displaystyle= 𝔢NT¯(ηlow)T¯(ηmed)Aτu(1Π)dh0(η)πτdh0(η)\displaystyle\mathfrak{e}_{N}\ -\ \underline{T}^{\circ}(\eta^{\text{low}})\ -\ \underline{T}^{\circ}(\eta^{\text{med}})\ -\ \not{D}_{A_{\tau}}u\ -\ (1-\Pi)\text{d}\not{\mathbb{D}}_{h_{0}}(\eta)-\pi_{\tau}\text{d}\not{\mathbb{D}}_{h_{0}}(\eta)
+Ξ+(η)+Zh(η,u)\displaystyle\ +\ \Xi^{+}(\eta)\ +\ \text{{Zh}}(\eta,u)
=\displaystyle= gN+Ξ+(η)+𝔢ImN𝟙+(1Π)(𝔢NRe𝟙+dh0(η))+πτ(𝔢NRe𝟙+dh0(η))\displaystyle g_{N}\ +\ \Xi^{+}(\eta)\ +\ \mathfrak{e}^{\text{Im}}_{N}\mathbb{1}^{-}\ +\ (1-\Pi)(\mathfrak{e}_{N}^{\text{Re}}\mathbb{1}^{-}+\text{d}\not{\mathbb{D}}_{h_{0}}(\eta))+\pi_{\tau}(\mathfrak{e}_{N}^{\text{Re}}\mathbb{1}^{-}+\text{d}\not{\mathbb{D}}_{h_{0}}(\eta))
+Zh(η,u)\displaystyle\ +\ \text{{Zh}}(\eta,u)

where (11.9) is obtained by substituting (11.6) and (11.7) with

Zh(η,u):=((hN)(hN))Ξ(η)𝔱(η)+(1χ)Aτudχ.uK1(χu)\text{{Zh}}(\eta,u):=\left(\mathbb{Q}(h^{\prime}_{N})-\mathbb{Q}(h_{N})\right)\ -\ \Xi^{-}(\eta)\ -\ \mathfrak{t}^{\circ}(\eta)\ +\ (1-\chi^{-})\not{D}_{A_{\tau}}u\ -\ d\chi^{-}.u\ -\ K_{1}(\chi^{-}u) (11.10)

and (11.9) from substituting (11.4)–(11.5).

Splitting up the terms of (11.9), define

𝔢N\displaystyle\mathfrak{e}_{N}^{\prime} :=\displaystyle:= 𝔢ImN𝟙+(1Π)(𝔢NRe𝟙+dh0(η))\displaystyle\mathfrak{e}^{\text{Im}}_{N}\mathbb{1}^{-}\ +\ (1-\Pi)(\mathfrak{e}_{N}^{\text{Re}}\mathbb{1}^{-}+\text{d}\not{\mathbb{D}}_{h_{0}}(\eta))
𝔤N\displaystyle\mathfrak{g}_{N}^{\prime} =\displaystyle= gN+Ξ+(η)\displaystyle g_{N}\ +\ \Xi^{+}(\eta)
λ\displaystyle\lambda :=\displaystyle:= επτ(𝔢NRe𝟙+dh0(η)),ΦτL2\displaystyle\varepsilon\langle\pi_{\tau}(\mathfrak{e}_{N}^{\text{Re}}\mathbb{1}^{-}+\text{d}\not{\mathbb{D}}_{h_{0}}(\eta))\ ,\ \Phi_{\tau}\rangle_{L^{2}}
𝔢N+1\displaystyle\mathfrak{e}_{N+1} :=\displaystyle:= Zh(η,u).\displaystyle\text{{Zh}}(\eta,u).

It now suffices to show that these satisfy the conclusions (1)–(4) of the proposition.

Beginning with (3), we have that

Zh(η,u)𝔏Cε1/48δ𝔢N𝔏.\|\text{{Zh}}(\eta,u)\|_{\mathfrak{L}}\leq C\varepsilon^{1/48}\delta\|\mathfrak{e}_{N}\|_{\mathfrak{L}}. (11.11)

By the bound u𝔚Cε8𝔢N𝔏\|u\|_{\mathfrak{W}}\leq C\varepsilon^{8}\|\mathfrak{e}_{N}\|_{\mathfrak{L}} from above and Corollary 10.3, all the terms of (11.10) except those involving ,Ξ+,Ξ\mathbb{Q},\Xi^{+},\Xi^{-} are bounded by, say, ε5𝔢N𝔏\varepsilon^{5}\|\mathfrak{e}_{N}\|_{\mathfrak{L}} and may be safely ignored. Corollary 9.4, and repeating the argument that obtained (10.19) for Ξ\Xi^{-} shows that Zh(η,u)𝔏Cε1/12γ𝔢N𝔏\|\text{{Zh}}(\eta,u)\|_{\mathfrak{L}}\leq C\varepsilon^{1/12-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}. Since γ<<1\gamma<<1 and ε1/48δ=ε1/24\varepsilon^{1/48}\delta=\varepsilon^{1/24}, (11.11) follows, which is conclusion (2).

(4) follows directly from the definition of λ\lambda and Cauchy-Schwartz, since 𝔢NRe𝟙L2𝔢N𝔏\|\mathfrak{e}_{N}^{\text{Re}}\mathbb{1}^{-}\|_{L^{2}}\leq\|\mathfrak{e}_{N}\|_{\mathfrak{L}} and

dh0(η)L2Cη1/2Cη𝔚\|\text{d}\not{\mathbb{D}}_{h_{0}}(\eta)\|_{L^{2}}\leq C\|\eta\|_{1/2}\leq C\|\eta\|_{\mathfrak{W}}

by Proposition 9.3.

For (1)–(2), the fact the Π(𝔢N)=0\Pi(\mathfrak{e}_{N}^{\prime})=0 and 𝔤N=𝔤N𝟙+\mathfrak{g}^{\prime}_{N}=\mathfrak{g}_{N}^{\prime}\mathbb{1}^{+} are immediate from the definitions. It remains to show that the asserted bounds hold. To re-iterate the cancellation that led to (11.9), adding and subtracting (11.4),(11.5), ΨN\Psi_{N} and λε1Φτ\lambda\varepsilon^{-1}\Phi_{\tau}, then using Π(d)=T+𝔱\Pi(\text{d}\not{\mathbb{D}})=T^{\circ}+\mathfrak{t}^{\circ} yields

𝔢N=𝔢NIm𝟙+eNRe𝟙+dh0(η)𝔱(η)ΨNλε1Φτ\mathfrak{e}_{N}^{\prime}=\mathfrak{e}_{N}^{\text{Im}}\mathbb{1}^{-}\ +\ e_{N}^{\text{Re}}\mathbb{1}^{-}\ +\ \text{d}\not{\mathbb{D}}_{h_{0}}(\eta)\ -\mathfrak{t}^{\circ}(\eta)\ -\ \Psi_{N}\ -\ \lambda\varepsilon^{-1}\Phi_{\tau}

Because 𝔱(η),ΨN\mathfrak{t}^{\circ}(\eta),\Psi_{N} are O(ε8)O(\varepsilon^{8}) these may be safely ignored as in the proof of (3). Since 𝔢ImN\mathfrak{e}^{\text{Im}}_{N} is unchanged, and πmedΠ(𝔢N)=0\pi^{\text{med}}\Pi(\mathfrak{e}_{N}^{\prime})=0, it suffices to bound the three terms on the top line in the norm (10.13).

dh0(η)𝟙+L2\displaystyle\|\text{d}\not{\mathbb{D}}_{h_{0}}(\eta)\mathbb{1}^{+}\|_{L^{2}}\ \displaystyle\leq Cε1/12γη𝔚Cε1/12γ𝔢N𝔏\displaystyle\ \ \ C\varepsilon^{1/12-\gamma}\|\eta\|_{\mathfrak{W}}\ \hskip 73.97733pt\leq\ C\varepsilon^{1/12-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}
(𝔢NRe+dh0(η))𝟙L2\displaystyle\|(\mathfrak{e}_{N}^{\text{Re}}+\ \text{d}\not{\mathbb{D}}_{h_{0}}(\eta))\mathbb{1}^{-}\|_{L^{2}}\ \displaystyle\leq C𝔢N𝔏+Cη𝔚C𝔢N𝔏\displaystyle\ \ \ C\|\mathfrak{e}_{N}\|_{\mathfrak{L}}\ \ \ \ +\ \ \ \ C\|\eta\|_{\mathfrak{W}}\ \ \ \ \ \ \ \ \ \ \ \leq\ \ C\|\mathfrak{e}_{N}\|_{\mathfrak{L}}
(𝔢NRe+dh0(η))𝟙L2ν\displaystyle\|(\mathfrak{e}_{N}^{\text{Re}}+\ \text{d}\not{\mathbb{D}}_{h_{0}}(\eta))\mathbb{1}^{-}\|_{L^{2}_{-\nu}} \displaystyle\leq Cεν/2𝔢N𝔏+Cεν/2γη𝔚Cεγεν/2𝔢N𝔏\displaystyle\ C\varepsilon^{-\nu/2}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}\ +C\varepsilon^{-\nu/2-\gamma}\|\eta\|_{\mathfrak{W}}\ \ \ \leq\ \ C\varepsilon^{-\gamma}\varepsilon^{-\nu/2}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}
gN+Ξ+(η)L2\displaystyle\|g_{N}+\Xi^{+}(\eta)\|_{L^{2}}\ \displaystyle\leq Cε1/12γ𝔢N𝔏+Cε1/12γη𝔚Cε1/12γ𝔢N𝔏\displaystyle C\varepsilon^{1/12-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}\ +\ C\varepsilon^{1/12-\gamma}\|\eta\|_{\mathfrak{W}}\ \leq\ C\varepsilon^{1/12-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}

where each line, the definition of the 𝔏\mathfrak{L}-norm is used in conjunction with, (10.22), (10.17), and (10.18). For the λε1Φτ\lambda\varepsilon^{-1}\Phi_{\tau}, the same bounds hold by (2) and the fact that ΦτL2L2ν\Phi_{\tau}\in L^{2}\cap L^{2}_{-\nu}. Conclusion (1) follows from the first three lines, and (2) from the fourth.

That (I) holds for hNh_{N}^{\prime} is immediate from the bounds η𝔚+u𝒳C𝔢N𝔏\|\eta\|_{\mathfrak{W}}\ +\ \|u\|_{\mathcal{X}}\leq C\|\mathfrak{e}_{N}\|_{\mathfrak{L}}, and the fact that it holds for hNh_{N}. Finally, the proof of the (resp. d𝕊𝕎¯\text{d}\overline{\mathbb{SW}}) statements is identical, omitting any mention of \mathbb{Q} and Λ\Lambda. ∎

11.2. The Outside Step

This subsection covers the second of the three stages of the cyclic induction (2.17). Now that the error terms are essentially orthogonal to the obstruction, solving in the outside can proceed using Lemma 4.4 and Proposition 7.11.

Define the outside parametrix PP^{-} by

P:𝔏P\displaystyle P^{-}:\mathfrak{L}\longrightarrow\mathcal{H}^{-}\oplus\mathbb{R}\hskip 56.9055ptP^{-} :=\displaystyle:= (1(Φτ,Aτ)(1Π),επτ)𝟙\displaystyle\left(\mathcal{L}^{-1}_{(\Phi_{\tau},A_{\tau})}(1-\Pi^{\perp})\ ,\ -\varepsilon\pi_{\tau}\right)\circ\mathbb{1}^{-} (11.12)

where πτ:𝔏\pi_{\tau}:\mathfrak{L}\to\mathbb{R} is understood to mean the coefficient of Φτ\Phi_{\tau} in Φτ\mathbb{R}\Phi_{\tau}. Notice that the first component indeed lands in \mathcal{H}^{-} by Definition (10.5).

Proposition 11.2.

PP\!^{-} is a linear operator uniformly bounded in ε,τ\varepsilon,\tau and satisfies the following property. If for NN\in\mathbb{N}, hNh^{\prime}_{N}\in\mathcal{H} is a configuration satisfying the conclusions of Proposition 11.1, then the updated configuration

hN=(IdP𝕊𝕎¯Λ)hNh_{N}^{\prime\prime}=(\text{Id}-P^{-}\circ\overline{\mathbb{SW}}_{\Lambda})h_{N}^{\prime} (11.13)

satisfies

𝕊𝕎¯Λ(hN)=𝔢N𝟙++𝔢N+1\overline{\mathbb{SW}}_{\Lambda}(h^{\prime\prime}_{N})=\mathfrak{e}_{N}^{\prime\prime}\mathbb{1}^{+}\ +\ \mathfrak{e}_{N+1}

(resp. the same for dh1𝕊𝕎¯\text{d}_{h_{1}}\overline{\mathbb{SW}}), where 𝔢N,𝔢N+1𝔏\mathfrak{e}_{N}^{\prime\prime},\mathfrak{e}_{N+1}\in\mathfrak{L} obey

  1. (1’)

    𝔢N𝔏Cεγ𝔢N𝔏\|\mathfrak{e}_{N}^{\prime\prime}\|_{\mathfrak{L}}\leq C\varepsilon^{-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}

  2. (2’)

    Item (2) from Proposition 11.1 continues to hold.

Moreover, hNh_{N}^{\prime\prime} continues to satisfy (I) from Proposition 11.1.

Proof.

As in the proof of Proposition 11.1, write 𝔢N+𝔤N=eN+fN+gN\mathfrak{e}_{N}^{\prime}+\mathfrak{g}_{N}^{\prime}=e^{\prime}_{N}+f^{\prime}_{N}+g^{\prime}_{N}, where

gN:=𝔤N𝟙+fN:=(𝔢N)Im𝟙eN:=(𝔢N)Re.g_{N}^{\prime}:=\mathfrak{g}_{N}^{\prime}\mathbb{1}^{+}\hskip 56.9055ptf_{N}^{\prime}:=(\mathfrak{e}^{\prime}_{N})^{\text{Im}}\mathbb{1}^{-}\hskip 56.9055pte_{N}^{\prime}:=(\mathfrak{e}_{N}^{\prime})^{\text{Re}}. (11.14)

Conclusions (1)–(4) of Proposition 11.1 mean that Π(eN)=0\Pi(e_{N}^{\prime})=0, that eNL2νCεγεν/2𝔢N𝔏\|e_{N}\|_{L^{2}_{-\nu}}\leq C\varepsilon^{-\gamma}\varepsilon^{-\nu/2}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}, for both ν=0,ν\nu=0,\nu^{-}, and that fNL2Cε1/6𝔢N𝔏\|f_{N}\|_{L^{2}}\leq C\varepsilon^{1/6}\|\mathfrak{e}_{N}\|_{\mathfrak{L}} and  gNL2Cε1/12γ𝔢N𝔏\|g_{N}\|_{L^{2}}\leq C\varepsilon^{1/12-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}, where 𝔢N\mathfrak{e}_{N} is the original error from Proposition 11.1.

Set

(ψ,b,μ):=P(eN+fN+λε1Φτ)(\psi,b,\mu):=P^{-}(e^{\prime}_{N}+f^{\prime}_{N}+\lambda\varepsilon^{-1}\Phi_{\tau})

so that

AτψRe\displaystyle\not{D}_{A_{\tau}}\psi^{\text{Re}} =\displaystyle= eN\displaystyle e_{N}^{\prime}
(Φτ,Aτ)Im(ψIm,b)\displaystyle\mathcal{L}_{(\Phi_{\tau},A_{\tau})}^{\text{Im}}(\psi^{\text{Im}},b) =\displaystyle= fN\displaystyle f_{N}^{\prime}
μ\displaystyle\mu =\displaystyle= επ(λε1Φτ)=λ\displaystyle-\varepsilon\pi(\lambda\varepsilon^{-1}\Phi_{\tau})=-\lambda

where ψ=(ψRe,ψIm)SReSIm\psi=(\psi^{\text{Re}},\psi^{\text{Im}})\in S^{\text{Re}}\oplus S^{\text{Im}}.

We now show that PP^{-} is uniformly bounded. Lemma 4.4 and Proposition 7.11 and the above bounds on eN,fNe^{\prime}_{N},f^{\prime}_{N} show that these unique solutions (where πτ(ψRe)=0\pi_{\tau}(\psi^{\text{Re}})=0), satisfy

ψRerH1e\displaystyle\|\psi^{\text{Re}}\|_{rH^{1}_{e}} \displaystyle\leq CeNL2Cεγ𝔢N𝔏\displaystyle C\|e_{N}^{\prime}\|_{L^{2}}\ \ \ \ \ \ \ \ \ \ \leq\ C\varepsilon^{-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}} (11.15)
εν/2ψRer1+νH1e\displaystyle\varepsilon^{\nu/2}\|\psi^{\text{Re}}\|_{r^{1+\nu}H^{1}_{e}} \displaystyle\leq Cεν/2eNL2νCεγ𝔢N𝔏\displaystyle C\varepsilon^{\nu/2}\|e_{N}^{\prime}\|_{L^{2}_{-\nu}}\ \ \hskip 5.12128pt\leq\ C\varepsilon^{-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}} (11.16)
ε1/6(ψIm,b)1,ε\displaystyle\varepsilon^{-1/6}\|(\psi^{\text{Im}},b)\|_{\mathcal{H}^{1,-}_{\varepsilon}} \displaystyle\leq Cε1/6fNL2Cεγ𝔢N𝔏\displaystyle C\varepsilon^{-1/6}\|f_{N}^{\prime}\|_{L^{2}}\ \ \ \leq\ C\varepsilon^{-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}} (11.17)
ε1|μ|\displaystyle\varepsilon^{-1}|\mu| \displaystyle\leq Cλε1ΦτL2Cεγ𝔢N𝔏.\displaystyle C\|\lambda\varepsilon^{-1}\Phi_{\tau}\|_{L^{2}}\ \hskip 7.96674pt\leq\ C\varepsilon^{-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}. (11.18)

To explain further how to obtain the bounds in the middle column, the condition that πτ(ψRe)=0\pi_{\tau}(\psi^{\text{Re}})=0 means that the ν=0\nu=0 version of Lemma 4.4 applies without the projection term, which yields (11.15). In turn, (11.15) and Cauchy-Schwartz show that πν(ψRe)CψL2CeNL2\pi_{\nu}(\psi^{\text{Re}})\leq C\|\psi\|_{L^{2}}\leq C\|e^{\prime}_{N}\|_{L^{2}} since ΦτL22ν\Phi_{\tau}\in L^{2}_{-2\nu}, after which (11.16) follows from Lemma 4.4 taking ν=ν\nu=\nu^{-}. (11.17) and (11.18) are immediate from Proposition 7.11 and the definition of μ\mu. The final column follows immediately from the bounds on (11.14) and conclusion (4) of Proposition 11.1.

To show that PP^{-} is uniformly bounded, it must be shown that 𝔢N+λε1Φτ𝔏\|\mathfrak{e}_{N}^{\prime}+\lambda\varepsilon^{-1}\Phi_{\tau}\|_{\mathfrak{L}} dominates each term in the middle column (since PP^{-} ignores 𝔤N\mathfrak{g}_{N}^{\prime}). Indeed, 𝔢N\mathfrak{e}_{N}^{\prime} is, by its construction in the proof of Proposition 11.1, L2L^{2}-orthogonal to Φτ\Phi_{\tau}. Therefore,

eNL2+λε1ΦτL2\displaystyle\|e_{N}^{\prime}\|_{L^{2}}+\|\lambda\varepsilon^{-1}\Phi_{\tau}\|_{L^{2}} \displaystyle\leq 𝔢N+λε1ΦτL2C𝔢N+λε1Φτ𝔏\displaystyle\ \ \ \ \ \ \|\mathfrak{e}_{N}^{\prime}+\lambda\varepsilon^{-1}\Phi_{\tau}\|_{L^{2}}\hskip 93.89418pt\ \leq C\|\mathfrak{e}_{N}^{\prime}+\lambda\varepsilon^{-1}\Phi_{\tau}\|_{\mathfrak{L}}\ \ \ \ \ \ \ (11.19)
eNL2ν\displaystyle\|e_{N}^{\prime}\|_{L^{2}_{-\nu}} \displaystyle\leq (eN+λε1Φτ)𝟙L2ν+(λε1Φτ)𝟙L2νC𝔢N+λε1Φτ𝔏.\displaystyle\|(e_{N}^{\prime}+\lambda\varepsilon^{-1}\Phi_{\tau})\mathbb{1}^{-}\|_{L^{2}_{-\nu}}+\|(\lambda\varepsilon^{-1}\Phi_{\tau})\mathbb{1}^{-}\|_{L^{2}_{-\nu}}\leq C\|\mathfrak{e}_{N}^{\prime}+\lambda\varepsilon^{-1}\Phi_{\tau}\|_{\mathfrak{L}}. (11.20)

The first of these is simply orthogonality along with fact that the 𝔏\mathfrak{L}-norm dominates the L2L^{2}-norm. The second is the triangle inequality, then invoking the first one along with the fact that the L2νL^{2}_{-\nu} and L2L^{2}-norms are equivalent on the 1-dimensional span of Φτ\Phi_{\tau}. That ε1/6fNL2\varepsilon^{-1/6}\|f_{N}^{\prime}\|_{L^{2}} in (11.17) is likewise bounded by the right side of (11.19) is immediate from the Definition of the 𝔏\mathfrak{L}-norm. This completes the claim that PP^{-} is uniformly bounded.

We now proceed to calculate 𝕊𝕎¯Λ(hN)\overline{\mathbb{SW}}_{\Lambda}(h^{\prime\prime}_{N}) where hN=hN(ψ,b,μ)h_{N}^{\prime\prime}=h_{N}-(\psi,b,\mu) as in (11.13). First,

d𝕊𝕎¯h1(ψ,b,μ)\displaystyle\text{d}\overline{\mathbb{SW}}_{h_{1}}(\psi,b,\mu) =\displaystyle= (Φτ,Aτ)χ(ψ,b)+𝒦1(ψ,b)με1χΦτ\displaystyle\mathcal{L}_{(\Phi_{\tau},A_{\tau})}\chi^{-}(\psi,b)\ +\ \mathcal{K}_{1}(\psi,b)\ -\ \mu\varepsilon^{-1}\chi^{-}\Phi_{\tau} (11.21)
=\displaystyle= AτψRe+(Φτ,Aτ)Im(ψIm,b)με1Φτ\displaystyle\not{D}_{A_{\tau}}\psi^{\text{Re}}\ +\ \mathcal{L}_{(\Phi_{\tau},A_{\tau})}^{\text{Im}}(\psi^{\text{Im}},b)\ -\ \mu\varepsilon^{-1}\Phi_{\tau}
+dχ(ψ,b)+K1(ψ,b)+(1χ)με1Φτ,\displaystyle\ +\ d\chi^{-}(\psi,b)\ +\ K_{1}(\psi,b)\ +\ (1-\chi^{-})\mu\varepsilon^{-1}\Phi_{\tau},

where we have used the fact that χ=1\chi^{-}=1 on the support of 𝟙\mathbb{1}^{-}. Using the conclusion of Proposition 11.1 and the definitions (11.14), then substituting (11.21) yields

𝕊𝕎¯Λ(hN)\displaystyle\overline{\mathbb{SW}}_{\Lambda}(h_{N}^{\prime\prime}) =\displaystyle= dh1𝕊𝕎(hN)dh1𝕊𝕎(ψ,b,μ)+(hN)(μN+μ+Λ)χΦτε\displaystyle\text{d}_{h_{1}}\mathbb{SW}(h_{N}^{\prime})\ -\ \text{d}_{h_{1}}\mathbb{SW}(\psi,b,\mu)\ +\ \mathbb{Q}(h_{N}^{\prime\prime})\ -\ (\mu_{N}+\mu+\Lambda)\chi^{-}\frac{\Phi_{\tau}}{\varepsilon}
=\displaystyle= 𝕊𝕎¯Λ(hN)dh1𝕊𝕎(ψ,b,μ)+(hN)(hN)\displaystyle\ \overline{\mathbb{SW}}_{\Lambda}(h_{N}^{\prime})\ -\ \text{d}_{h_{1}}\mathbb{SW}(\psi,b,\mu)\ +\ \mathbb{Q}(h_{N}^{\prime\prime})-\mathbb{Q}(h_{N}^{\prime})
=\displaystyle= eN+fN+gN+λε1ΦτAτψRe(Φτ,Aτ)Im(ψIm,b)+με1Φτ\displaystyle e_{N}^{\prime}\ +\ f_{N}^{\prime}\ +\ g_{N}^{\prime}\ +\ \lambda\varepsilon^{-1}\Phi_{\tau}\ -\ \not{D}_{A_{\tau}}\psi^{\text{Re}}\ -\ \mathcal{L}_{(\Phi_{\tau},A_{\tau})}^{\text{Im}}(\psi^{\text{Im}},b)\ +\ \mu\varepsilon^{-1}\Phi_{\tau}\ \ \
+((hN)(hN))+dχ(ψ,b)+K1(ψ,b)+(1χ)με1Φτ+𝔢N+1\displaystyle\ +\ \left(\mathbb{Q}(h_{N}^{\prime\prime})-\mathbb{Q}(h_{N}^{\prime})\right)\ +\ d\chi^{-}(\psi,b)\ +\ K_{1}(\psi,b)\ +\ (1-\chi^{-})\mu\varepsilon^{-1}\Phi_{\tau}\ +\ \mathfrak{e}_{N+1}
=\displaystyle= gN+((hN)(hN))+dχ(ψ,b)+K1(ψ,b)+(1χ)με1Φτ+𝔢N+1.\displaystyle g_{N}^{\prime}+\ \left(\mathbb{Q}(h_{N}^{\prime\prime})-\mathbb{Q}(h_{N}^{\prime})\right)\ +\ d\chi^{-}(\psi,b)\ +\ K_{1}(\psi,b)\ +\ (1-\chi^{-})\mu\varepsilon^{-1}\Phi_{\tau}\ +\ \mathfrak{e}_{N+1}.

Then (re)-define

𝔢N\displaystyle\mathfrak{e}_{N}^{\prime\prime} :=\displaystyle:= gN+dχ(ψ,b)+(1χ)με1Φτ\displaystyle g_{N}^{\prime}\ +\ d\chi^{-}(\psi,b)\ +\ (1-\chi^{-})\mu\varepsilon^{-1}\Phi_{\tau} (11.22)
𝔢N+1\displaystyle\mathfrak{e}_{N+1} \displaystyle\mapsto 𝔢N+1+((hN)(hN))+K1(ψ,b).\displaystyle\mathfrak{e}_{N+1}\ +\ \left(\mathbb{Q}(h_{N}^{\prime})-\mathbb{Q}(h_{N}^{\prime\prime})\right)\ +\ K_{1}(\psi,b). (11.23)

Notice that 𝔢N\mathfrak{e}_{N}^{\prime\prime} now includes the crucial alternating error term dχ(ψ,b)d\chi^{-}(\psi,b).

To conclude, we show 𝔢N\mathfrak{e}_{N}^{\prime\prime} and 𝔢N+1\mathfrak{e}_{N+1} satisfy conclusions (1)–(2). That (1) holds for the first term of 𝔢N\mathfrak{e}_{N}^{\prime\prime} in (11.22) is immediate from the bound on gNg^{\prime}_{N} from 11.14. The bounds on eNe_{N}^{\prime} in 11.14 also imply that eNe_{N}^{\prime} has ε1/2\varepsilon^{1/2}-effective support, hence Lemma 4.7 applies to show that

dχ.ψReL2Cε1/12γeNL2Cε1/122γ𝔢N𝔏.\|d\chi^{-}.\psi^{\text{Re}}\|_{L^{2}}\leq C\varepsilon^{1/12-\gamma}\|e_{N}^{\prime}\|_{L^{2}}\leq C\varepsilon^{1/12-2\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}. (11.24)

Note that the (1χ)𝔢N(1-\chi^{-})\mathfrak{e}_{N}^{\prime} term in Lemma 4.7 vanishes because χ=1\chi^{-}=1 on supp(𝟙)\text{supp}(\mathbb{1}^{-}). (11.24) also holds for (ψIm,b)(\psi^{\text{Im}},b) simply because of the ε1/6\varepsilon^{1/6} weight on these components in (11.17). Finally, for the third term of 𝔢N\mathfrak{e}_{N}^{\prime\prime} in (11.22) , notice that direct integration using the fact that |Φτ|Cr1/2|\Phi_{\tau}|\leq Cr^{1/2} shows that (1χ)ΦτL2Cε1γ\|(1-\chi^{-})\Phi_{\tau}\|_{L^{2}}\leq C\varepsilon^{1-\gamma}, and (11.18) therefore shows the third term of 𝔢N\mathfrak{e}_{N}^{\prime\prime} is bounded by the right hand side of (11.24) with an extra factor of ε11/12γ\varepsilon^{11/12-\gamma}. Combining these three shows that

𝔢N𝔏Cεγ𝔢N𝔏\|\mathfrak{e}_{N}^{\prime\prime}\|_{\mathfrak{L}}\leq C\varepsilon^{-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}

and 𝔢N=𝔢N𝟙\mathfrak{e}_{N}^{\prime\prime}=\mathfrak{e}_{N}^{\prime\prime}\mathbb{1}^{-} because χ=1\chi^{-}=1 outside supp(𝟙+)\text{supp}(\mathbb{1}^{+}).

Using the bounds (11.15)–(11.17), the terms involving \mathbb{Q} can be bounded identically to in the proof of Proposition 11.1. As in the proof of Lemma 10.14, K1K_{1} is exponentially small, hence negligible. Conclusion (2) follows after increasing γ\gamma slightly.

Finally, that (I) continues to hold for hNh_{N}^{\prime\prime} and of the (resp. d𝕊𝕎¯h1\text{d}\overline{\mathbb{SW}}_{h_{1}}) statements follows identically to in Proposition 11.1 using the uniform boundedness of PP^{-}. ∎

11.3. The Inside Step

The section completes the third stage of the cyclic induction by constructing P+P^{+}. Define

P+:𝔏+P+\displaystyle P^{+}:\mathfrak{L}\longrightarrow\mathcal{H}^{+}\hskip 56.9055ptP^{+} :=\displaystyle:= 1(Φ1,A1)𝟙+.\displaystyle\mathcal{L}^{-1}_{(\Phi_{1},A_{1})}\mathbb{1}^{+}. (11.25)

Definition 10.5 ensures that P+P^{+} lands in +\mathcal{H}^{+}.

Proposition 11.3.

P+P^{+} is a linear operator uniformly bounded in ε,τ\varepsilon,\tau and satisfies the following property. If for NN\in\mathbb{N}, hNh^{\prime\prime}_{N}\in\mathcal{H} is a configuration satisfying the conclusions of Proposition 11.2, then the updated configuration

hN+1=(IdP𝕊𝕎¯Λ)hNh_{N+1}=(\text{Id}-P^{-}\circ\overline{\mathbb{SW}}_{\Lambda})h_{N}^{\prime\prime} (11.26)

(resp. the same for d𝕊𝕎¯h1)\text{d}\overline{\mathbb{SW}}_{h_{1}}) satisfies the hypotheses of Proposition 11.1 with N=N+1N^{\prime}=N+1.

Proof.

Uniform boundedness is immediate from the definition of the +\mathcal{H}^{+}-norm in Definition 10.5. Let 𝔢N\mathfrak{e}_{N}^{\prime\prime} be as in the conclusion of Proposition 11.2, so that we may write

gN:=𝔢N𝟙+=𝔢Ng_{N}^{\prime\prime}:=\mathfrak{e}_{N}^{\prime\prime}\mathbb{1}^{+}=\mathfrak{e}_{N}^{\prime\prime}

where gNL2Cε1/12γ𝔢N𝔏\|g_{N}^{\prime\prime}\|_{L^{2}}\leq C\varepsilon^{1/12-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}.

Set

(φ,a):=P+(gN)(\varphi,a):=P^{+}(g_{N}^{\prime\prime})

so that

(Φ1,A1)(φ,a)=gN,\mathcal{L}_{(\Phi_{1},A_{1})}(\varphi,a)=g_{N}^{\prime\prime},

and

(φ,a)H1,+ε(φ,a)+=Cε1/12γgNCεγ𝔢N𝔏.\|(\varphi,a)\|_{H^{1,+}_{\varepsilon}}\ \leq\ \|(\varphi,a)\|_{\mathcal{H}^{+}}=C\varepsilon^{-1/12-\gamma}\|g_{N}^{\prime\prime}\|\ \leq\ C\varepsilon^{-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}. (11.27)

We now proceed to calculate 𝕊𝕎¯Λ(hN+1)\overline{\mathbb{SW}}_{\Lambda}(h_{N+1}) where hN+1=hN(φ,a)h_{N+1}=h_{N}-(\varphi,a) as in (11.13).

𝕊𝕎¯Λ(hN+1)\displaystyle\overline{\mathbb{SW}}_{\Lambda}(h_{N+1}) =\displaystyle= dh1𝕊𝕎(hN)dh1𝕊𝕎(φ,a)+(hN+1)(μN+μ+Λ)χΦτε\displaystyle\text{d}_{h_{1}}\mathbb{SW}(h_{N}^{\prime\prime})\ -\ \text{d}_{h_{1}}\mathbb{SW}(\varphi,a)\ +\ \mathbb{Q}(h_{N+1})\ -\ (\mu_{N}+\mu+\Lambda)\chi^{-}\frac{\Phi_{\tau}}{\varepsilon} (11.28)
=\displaystyle= 𝕊𝕎¯Λ(hN)dh1𝕊𝕎(ψ,b,μ)+(hN+1)(hN)\displaystyle\ \overline{\mathbb{SW}}_{\Lambda}(h_{N}^{\prime})\ -\ \text{d}_{h_{1}}\mathbb{SW}(\psi,b,\mu)\ +\ \mathbb{Q}(h_{N+1})-\mathbb{Q}(h_{N}^{\prime\prime})
=\displaystyle= gN(Φ1,A1)(φ,a)+((hN+1)(hN))+dχ+(φ,a)+𝔢N+1\displaystyle g_{N}^{\prime\prime}\ -\ \mathcal{L}_{(\Phi_{1},A_{1})}(\varphi,a)\ +\ \left(\mathbb{Q}(h_{N+1})-\mathbb{Q}(h_{N}^{\prime\prime})\right)\ +\ d\chi^{+}(\varphi,a)\ +\ \mathfrak{e}_{N+1}
=\displaystyle= ((hN+1)(hN))+dχ+(φ,a)+𝔢N+1\displaystyle\left(\mathbb{Q}(h_{N+1})-\mathbb{Q}(h_{N}^{\prime\prime})\right)\ +\ d\chi^{+}(\varphi,a)\ +\ \mathfrak{e}_{N+1}

The alternating error from dχ+d\chi^{+} has now been shifted back to the outside region.

Re-defining 𝔢N+1\mathfrak{e}_{N+1} to include all three terms of (11.28), we claim that it satisfies (II) of Proposition 11.1 with N=N+1N^{\prime}=N+1. In fact, one has the slightly stronger bound

𝔢N+1𝔏Cε1/48γδ𝔢N𝔏.\|\mathfrak{e}_{N+1}\|_{\mathfrak{L}}\leq C\varepsilon^{1/48-\gamma}\delta\|\mathfrak{e}_{N}\|_{\mathfrak{L}}. (11.29)

Indeed, the terms involving \mathbb{Q} in (11.28) may be bounded as in the proof of Proposition 11.1. Then, since dχ+d\chi^{+} is supported where r=O(ε1/2)r=O(\varepsilon^{1/2}), Lemma 7.12 implies

dχ+(φ,a)𝔏\displaystyle\|d\chi^{+}(\varphi,a)\|_{\mathfrak{L}} \displaystyle\leq dχ+(φ,a)L2+ενdχ+(φ,a)L2ν\displaystyle\|d\chi^{+}(\varphi,a)\|_{L^{2}}+\varepsilon^{\nu}\|d\chi^{+}(\varphi,a)\|_{L^{2}_{-\nu}}
\displaystyle\leq Cdχ+(φ,a)L2Cε1/24γgNL2Cε1/24γ𝔢N𝔏.\displaystyle C\|d\chi^{+}(\varphi,a)\|_{L^{2}}\leq C\varepsilon^{-1/24-\gamma}\|g_{N}^{\prime\prime}\|_{L^{2}}\leq C\varepsilon^{1/24-\gamma}\|\mathfrak{e}_{N}\|_{\mathfrak{L}}.

Finally, the original 𝔢N+1\mathfrak{e}_{N+1} in (11.28) already satisfies (11.29) by virtue of (2’) in Proposition 11.2. (11.29) follows.

Over the course of the proofs of (a single cycle of) Propositions 11.1, 11.2, and 11.3, the constants C,γC,\gamma have been increased a finite number of times. Let C0C_{0} being the original constant in Item (II) of Proposition 11.1 and C1,γ1C_{1},\gamma_{1} be the final versions of the constants appearing in (11.29). Since γ1<<1\gamma_{1}<<1 still, we may assume that

C1ε1/48γ1C0C_{1}\varepsilon^{1/48-\gamma_{1}}\leq C_{0}

once ε\varepsilon is sufficiently small, which reduces (11.29) to hypothesis (II) of Proposition 11.1 with N=N+1N^{\prime}=N+1 as desired. The proof of (I) and of the (resp. d𝕊𝕎¯h1\text{d}\overline{\mathbb{SW}}_{h_{1}}) statements again follows as in Propositions 11.1 and 11.2 using (11.27). ∎

11.4. Proof of Proposition 10.1

This section completes the proof of Proposition 10.1.

Analogously to (2.23), define 𝔸\mathbb{A} and 1=d𝔸\mathbb{P}_{1}=\text{d}\mathbb{A} by

𝔸\displaystyle\mathbb{A} =\displaystyle= Pξ+P(Id𝕊𝕎¯ΛPξ)+P+(Id𝕊𝕎¯ΛPξ𝕊𝕎¯ΛP(Id𝕊𝕎¯ΛPξ)),\displaystyle P_{\xi}\ +\ P^{-}(\text{Id}-\ \overline{\mathbb{SW}}_{\Lambda}P_{\xi})\ \ +\ \ P^{+}\left(\text{Id}-\ \overline{\mathbb{SW}}_{\Lambda}P_{\xi}\ -\ \ \overline{\mathbb{SW}}_{\Lambda}P^{-}(\text{Id}\ -\ \overline{\mathbb{SW}}_{\Lambda}P_{\xi})\right), (11.30)
1\displaystyle\mathbb{P}_{1} =\displaystyle= Pξ+P(Idd𝕊𝕎¯h1Pξ)+P+(Idd𝕊𝕎¯h1Pξd𝕊𝕎¯h1P(Idd𝕊𝕎¯h1Pξ)),\displaystyle P_{\xi}\ +\ P^{-}(\text{Id}-\text{d}\overline{\mathbb{SW}}_{h_{1}}P_{\xi})\ +\ P^{+}\left(\text{Id}-\text{d}\overline{\mathbb{SW}}_{h_{1}}P_{\xi}-\ \text{d}\overline{\mathbb{SW}}_{h_{1}}P^{-}(\text{Id}-\text{d}\overline{\mathbb{SW}}_{h_{1}}P_{\xi})\right), (11.31)

So that (analogously to 2.24),

T=Id𝔸𝕊𝕎¯Λ\displaystyle T=\text{Id}-\mathbb{A}\circ\ \overline{\mathbb{SW}}_{\Lambda}\ \ =\displaystyle= (IdP+𝕊𝕎¯Λ)(IdP𝕊𝕎¯Λ)(IdPξ𝕊𝕎¯Λ),\displaystyle(\text{Id}-P^{+}\ \overline{\mathbb{SW}}_{\Lambda}\ )(\text{Id}-P^{-}\overline{\mathbb{SW}}_{\Lambda}\ )(\text{Id}-P_{\xi}\ \overline{\mathbb{SW}}_{\Lambda}\ ), (11.32)
dT=Id1d𝕊𝕎¯h1\displaystyle\text{dT}=\text{Id}-\mathbb{P}_{1}\circ\text{d}\overline{\mathbb{SW}}_{h_{1}} =\displaystyle= (IdP+d𝕊𝕎¯h1)(IdPd𝕊𝕎¯h1)(IdPξd𝕊𝕎¯h1).\displaystyle(\text{Id}-P^{+}\text{d}\overline{\mathbb{SW}}_{h_{1}})(\text{Id}-P^{-}\text{d}\overline{\mathbb{SW}}_{h_{1}})(\text{Id}-P_{\xi}\text{d}\overline{\mathbb{SW}}_{h_{1}}). (11.33)

Thus applying TT carries one complete cycle of the iteration (2.17), and dT\text{d}T one complete cycle of the linearized iteration (the resp. statements in Propositions 11.111.3).

The next lemma is the analogue of Proposition 2.4.

Lemma 11.4.

The linearization d𝕊𝕎¯h1:𝔏\text{d}\overline{\mathbb{SW}}_{h_{1}}:\mathcal{H}\to\mathfrak{L} is invertible, and there is a constant CC independent of ε,γ\varepsilon,\gamma such that

hCεγd𝕊𝕎¯h1(h)𝔏\|h\|_{\mathcal{H}}\leq C\varepsilon^{-\gamma}\|\text{d}\overline{\mathbb{SW}}_{h_{1}}(h)\|_{\mathfrak{L}} (11.34)

holds.

Proof.

Let 0:𝔏\mathbb{P}_{0}:\mathfrak{L}\to\mathcal{H} be defined by

0:=P+(Pξ0P)\mathbb{P}_{0}:=P^{+}\oplus\begin{pmatrix}P_{\xi}&0\\ *&P^{-}\end{pmatrix}

where =P(1Π)dh0Pξ*=-P^{-}(1-\Pi^{\perp})\text{d}\not{\mathbb{D}}_{h_{0}}P_{\xi}. Since the map β:𝔏(Y)𝔏(Y+)𝔏(Y)\beta:\mathfrak{L}(Y)\to\mathfrak{L}(Y^{+})\oplus\mathfrak{L}(Y^{-}) defined by β(𝔢):=(𝔢𝟙+,𝔢𝟙)\beta(\mathfrak{e}):=(\mathfrak{e}\mathbb{1}^{+},\mathfrak{e}\mathbb{1}^{-}), which is built into the definitions of Pξ,P±P_{\xi},P^{\pm} is an isomorphism, and so is each of Pξ,P,P+P_{\xi},P^{-},P^{+}, it follows that 0\mathbb{P}_{0} is an isomorphism hence Fredholm with index 0.

Then, the calculations (11.6), (11.7), and (11.21) and the subsequent bounds (e.g. 11.11) in the proofs of Propositions 11.1 and 11.2 show that

10\displaystyle\mathbb{P}_{1}-\mathbb{P}_{0} =\displaystyle= (P++P)(Ξ+𝔱)PξP+(dχP(Idd𝕊𝕎¯h1Pξ))+\displaystyle-(P^{+}+P^{-})(\Xi+\mathfrak{t}^{\circ})P_{\xi}\ -\ P^{+}(d\chi^{-}P^{-}(\text{Id}-\text{d}\overline{\mathbb{SW}}_{h_{1}}P_{\xi}))\ +\ \star (11.35)
(P++P)O(ε8).\displaystyle-(P^{+}+P^{-})O(\varepsilon^{8}).

where \star consists of terms involving the 1-dimensional span of Φτ\Phi_{\tau}, and the O(ε8)O(\varepsilon^{8}) accounts for terms involving uu and K1K_{1}. The terms in the top line of (11.35) involving are compact, because they factor through either the compact inclusion 1,L2\mathcal{H}^{1,-}\hookrightarrow L^{2} or through the finite-dimensional spaces spanned by η𝔚\eta\in\mathfrak{W} and Φτ\Phi_{\tau} respectively. Since 01Cεγ\|\mathbb{P}_{0}^{-1}\|\leq C\varepsilon^{-\gamma}, the O(ε8)O(\varepsilon^{8}) terms may be safely ignored. It follows that 1\mathbb{P}_{1} is Fredholm of index 0 once ε\varepsilon is sufficiently small.

The same argument shows that N\mathbb{P}_{N} defined by

IdNd𝕊𝕎¯h1=dTN\text{Id}-\mathbb{P}_{N}\circ\text{d}\overline{\mathbb{SW}}_{h_{1}}=\text{d}T^{N}

is likewise Fredholm of Index 0. By construction (cf 2.212.24), N\mathbb{P}_{N} applies NN stages of the (resp. d𝕊𝕎¯h1\text{d}\overline{\mathbb{SW}}_{h_{1}}) alternating iteration from Propositions 11.111.3 starting from 𝔢1=Idd𝕊𝕎¯h10\mathfrak{e}_{1}=\text{Id}-\text{d}\overline{\mathbb{SW}}_{h_{1}}\mathbb{P}_{0}. We conclude that

d𝕊𝕎¯h1N=Id+O(δN1)\text{d}\overline{\mathbb{SW}}_{h_{1}}\mathbb{P}_{N}=\text{Id}+O(\delta^{N-1})

It follows as in the proof of Proposition 2.4 that d𝕊𝕎¯h1\text{d}\overline{\mathbb{SW}}_{h_{1}} is an isomorphism. (11.34) follows from the bound 01Cεγ\|\mathbb{P}_{0}^{-1}\|\leq C\varepsilon^{-\gamma} above, and statement (I) in Propositions 11.111.3 with hNh_{N} being the correction from N0\mathbb{P}_{N}-\mathbb{P}_{0}, since 𝔢1𝔏Cεγ\|\mathfrak{e}_{1}\|_{\mathfrak{L}}\leq C\varepsilon^{-\gamma}. ∎

Proof of Proposition 10.1.

Let 𝒞=Br(0)\mathcal{C}=B_{r}(0)\subseteq\mathcal{H} of radius r=ε1/20r=\varepsilon^{-1/20}, and let TT be given by (10.1) with 𝔸\mathbb{A} defined by (11.30).

(A) is deduced assuming (C) as follows. Let h𝒞h\in\mathcal{C}, then since 𝕊𝕎¯Λ(0)𝔏Cε1/24γ\|\overline{\mathbb{SW}}_{\Lambda}(0)\|_{\mathfrak{L}}\leq C\varepsilon^{-1/24-\gamma} by Theorem 7.5,

T(h)T(h)T(0)+T(0)Cδh+Cε1/242γr,\|T(h)\|_{\mathcal{H}}\leq\|T(h)-T(0)\|_{\mathcal{H}}+\|T(0)\|_{\mathcal{H}}\leq C\sqrt{\delta}\|h\|_{\mathcal{H}}+C\varepsilon^{-1/24-2\gamma}\leq r,

where the bound on T(0)T(0) is a consequence of (I) in Propositions 11.111.3 with 𝔢1=𝕊𝕎¯Λ(0)\mathfrak{e}_{1}=\overline{\mathbb{SW}}_{\Lambda}(0). Thus T:𝒞𝒞T:\mathcal{C}\to\mathcal{C} preserves 𝒞\mathcal{C}. Continuity of TT is immediate from (C). Smooth dependence on (ε,τ)(\varepsilon,\tau) is immediate from the smooth dependence of the Seiberg–Witten equations on pτp_{\tau}, and the smooth dependence of (Φ1,A1)(\Phi_{1},A_{1}) and (Φτ,Aτ)(\Phi_{\tau},A_{\tau}) and the linearized equations at these and their inverses used to construct TT.

(B) By (11.32), applying TT constitutes a full cycle of the three-stage iteration carried out by Propositions 11.111.3. The conclusion follows from applying these inductively.

(C) Let 𝔮\mathfrak{q} be such that

T(h)=dT(h)+𝔮(h).T(h)=\text{d}T(h)+\mathfrak{q}(h).

where dT\text{d}T is as in (11.33). The same argument as (B), now using the (resp. d𝕊𝕎¯h1\text{d}\overline{\mathbb{SW}}_{h_{1}}) statements in Propositions 11.111.3 shows that

d𝕊𝕎¯h1(dT(h))𝔏Cδd𝕊𝕎¯h1(h)𝔏.\|\text{d}\overline{\mathbb{SW}}_{h_{1}}(\text{d}T(h))\|_{\mathfrak{L}}\leq C\delta\|\text{d}\overline{\mathbb{SW}}_{h_{1}}(h)\|_{\mathfrak{L}}.

Therefore since dT\text{d}T is linear, Lemmas 10.14 and 11.4 shows

T(h1)T(h2)\displaystyle\|T(h_{1})-T(h_{2})\|_{\mathcal{H}} \displaystyle\leq Cεγd𝕊𝕎¯h1(T(h1)T(h2))𝔏\displaystyle C\varepsilon^{-\gamma}\|\text{d}\overline{\mathbb{SW}}_{h_{1}}(T(h_{1})-T(h_{2}))\|_{\mathfrak{L}}
\displaystyle\leq Cεγd𝕊𝕎¯h1(dT(h1)dT(h2))𝔏+Cεγd𝕊𝕎¯h1(𝔮(h1)𝔮(h2))𝔏\displaystyle C\varepsilon^{-\gamma}\|\text{d}\overline{\mathbb{SW}}_{h_{1}}(\text{d}T(h_{1})-\text{d}T(h_{2}))\|_{\mathfrak{L}}\ +\ C\varepsilon^{-\gamma}\|\text{d}\overline{\mathbb{SW}}_{h_{1}}(\mathfrak{q}(h_{1})-\mathfrak{q}(h_{2}))\|_{\mathfrak{L}}
\displaystyle\leq Cεγδd𝕊𝕎¯h1(h1h2)𝔏+Cεγd𝕊𝕎¯h1(𝔮(h1)𝔮(h2))𝔏\displaystyle C\varepsilon^{-\gamma}\delta\|\text{d}\overline{\mathbb{SW}}_{h_{1}}(h_{1}-h_{2})\|_{\mathfrak{L}}\ +\ C\varepsilon^{-\gamma}\|\text{d}\overline{\mathbb{SW}}_{h_{1}}(\mathfrak{q}(h_{1})-\mathfrak{q}(h_{2}))\|_{\mathfrak{L}}
\displaystyle\leq Cε2γδh1h2+Cε2γ𝔮(h1)𝔮(h2).\displaystyle C\varepsilon^{-2\gamma}\delta\|h_{1}-h_{2}\|_{\mathcal{H}}\ +\ C\varepsilon^{-2\gamma}\|\mathfrak{q}(h_{1})-\mathfrak{q}(h_{2})\|_{\mathcal{H}}.

Since Cε2γδ<δC\varepsilon^{-2\gamma}\delta<\sqrt{\delta} once ε\varepsilon is sufficiently, small, the following bound completes (C):

𝔮(h1)𝔮(h2)Cε1/126γh1h2.\|\mathfrak{q}(h_{1})-\mathfrak{q}(h_{2})\|_{\mathcal{H}}\leq C\varepsilon^{1/12-6\gamma}\|h_{1}-h_{2}\|_{\mathcal{H}}. (11.36)

To prove (11.36), we calculate 𝔮=TdT\mathfrak{q}=T-\text{d}T by expanding (11.32) and (11.33), with 𝕊𝕎¯Λ=d𝕊𝕎¯h1++Λ\overline{\mathbb{SW}}_{\Lambda}=\text{d}\overline{\mathbb{SW}}_{h_{1}}+\mathbb{Q}+\Lambda where the latter is shorthand for Λ=Λ(τ)ε1χΦτ\Lambda=\Lambda(\tau)\varepsilon^{-1}\chi^{-}\Phi_{\tau}. Writing 𝕃=d𝕊𝕎¯h1,\mathbb{L}=\text{d}\overline{\mathbb{SW}}_{h_{1}}, this shows

𝔮\displaystyle\mathfrak{q} =\displaystyle= P+Pξ+P++PPξ𝕃+PPξ+P𝕃Pξ\displaystyle P^{-}\mathbb{Q}\ +\ P_{\xi}\mathbb{Q}\ +\ P^{+}\mathbb{Q}\ +\ P^{-}\mathbb{Q}P_{\xi}\mathbb{L}\ +\ P^{-}\mathbb{Q}P_{\xi}\mathbb{Q}\ +\ P^{-}\mathbb{L}P_{\xi}\mathbb{Q}
+\displaystyle+ P+P𝕃+P+Pξ𝕃+P+P+P+Pξ+P+𝕃P+P+𝕃Pξ\displaystyle P^{+}\mathbb{Q}P^{-}\mathbb{L}\ +\ P^{+}\mathbb{Q}P_{\xi}\mathbb{L}\ +\ P^{+}\mathbb{Q}P^{-}\mathbb{Q}\ +\ P^{+}\mathbb{Q}P_{\xi}\mathbb{Q}\ +\ P^{+}\mathbb{L}P^{-}\mathbb{Q}\ +\ P^{+}\mathbb{L}P_{\xi}\mathbb{Q}
+\displaystyle+ P+𝕃PPξ𝕃+P+𝕃P𝕃Pξ+P+P𝕃Pξ𝕃+P+𝕃PPξ\displaystyle P^{+}\mathbb{L}P^{-}\mathbb{Q}P_{\xi}\mathbb{L}\ +\ P^{+}\mathbb{L}P^{-}\mathbb{L}P_{\xi}\mathbb{Q}+P^{+}\mathbb{Q}P^{-}\mathbb{L}P_{\xi}\mathbb{L}\ +\ P^{+}\mathbb{L}P^{-}\mathbb{Q}P_{\xi}\mathbb{Q}
+\displaystyle+ P+PPξ𝕃+P+P𝕃Pξ+P+PPξ\displaystyle P^{+}\mathbb{Q}P^{-}\mathbb{Q}P_{\xi}\mathbb{L}\ +\ P^{+}\mathbb{Q}P^{-}\mathbb{L}P_{\xi}\mathbb{Q}\ +\ P^{+}\mathbb{Q}P^{-}\mathbb{Q}P_{\xi}\mathbb{Q}
+\displaystyle+ 𝔮Λ\displaystyle\mathfrak{q}_{\Lambda}

where 𝔮Λ\mathfrak{q}_{\Lambda} is the same collection of terms replacing each instance of \mathbb{Q} with Λ\Lambda. Since Λ\Lambda is constant, qΛ(h1)qΛ(h2)=0q_{\Lambda}(h_{1})-q_{\Lambda}(h_{2})=0. Because each remaining term has at least one factor of \mathbb{Q}, (11.36) now follows from the boundedness of Pξ,P±P_{\xi},P^{\pm} in Propositions 11.111.3, the boundedness of 𝕃\mathbb{L} from Lemma 10.14, along with Items (A) and (C) of Corollary 10.8. ∎

12. Gluing

This final section concludes the proof Theorem 1.6 using the cyclic iteration from Section 9. The iteration leads to glued solutions of a “Seiberg–Witten eigenvector” equation for every pair of parameters (ε,τ)(\varepsilon,\tau). These fail to solve the true Seiberg–Witten equations by a 1-dimensional obstruction coming from a multiple of the eigenvector. The condition that this obstruction vanish defines τ(ε)\tau(\varepsilon) implicitly as a function of ε\varepsilon, completing the proof of the theorem.

12.1. Glued Configurations

Proposition 10.1 and the Banach fixed-point theorem (with smooth dependence on parameters) immediately imply the following.

Corollary 12.1.

For every pair (ε,τ)(ε0,ε0)×(τ0,τ0)(\varepsilon,\tau)\in(-\varepsilon_{0},\varepsilon_{0})\times(-\tau_{0},\tau_{0}) with ε0,τ0\varepsilon_{0},\tau_{0} sufficiently small, there exist tuples (𝒵,Φ,A,μ)(\mathcal{Z},\Phi,A,\mu) depending smoothly on ε,τ\varepsilon,\tau where

  1. (1)

    𝒵(ε,τ)=𝒵τ,ξ(ε,τ)\mathcal{Z}(\varepsilon,\tau)=\mathcal{Z}_{\tau,\xi(\varepsilon,\tau)} is the singular set arising from a linearized deformation ξ(ε,τ)C(𝒵τ,N𝒵τ)\xi(\varepsilon,\tau)\in C^{\infty}(\mathcal{Z}_{\tau},N\mathcal{Z}_{\tau}),

  2. (2)

    μ(ε,τ)\mu(\varepsilon,\tau)\in\mathbb{R}, and

  3. (3)

    (Φ(ε,τ),A(ε,τ))C(Y;SEΩ)(\Phi(\varepsilon,\tau),A(\varepsilon,\tau))\in C^{\infty}(Y;S_{E}\oplus\Omega)

that satisfy

SW(Φ(ε,τ)ε,A(ε,τ))=(Λ(τ)+μ(ε,τ)χεΦτ,ξ(ε,τ)ε\text{SW}\left(\frac{\Phi(\varepsilon,\tau)}{\varepsilon},A(\varepsilon,\tau)\right)=(\Lambda(\tau)+\mu(\varepsilon,\tau)\chi^{-}_{\varepsilon}\cdot\frac{\Phi_{\tau,\xi(\varepsilon,\tau)}}{\varepsilon} (12.1)

where Φτ,ξ(ε,τ)\Phi_{\tau,\xi(\varepsilon,\tau)} is as in Definition 8.3, and χε=χε,τ,ξ(ε,τ)\chi^{-}_{\varepsilon}=\chi^{-}_{\varepsilon,\tau,\xi(\varepsilon,\tau)} cut-off function (8.10).

Proof.

Part (C) of Proposition 10.1 and the Banach fixed-point theorem give a smoothly varying family of fixed points h=(ξ,φ,a,ψ,b,μ,u)1e×𝒳h=(\xi,\varphi,a,\psi,b,\mu,u)\in\mathbb{H}^{1}_{e}\times\mathcal{X}. By (B) of Proposition 10.1, these satisfy, 𝕊𝕎¯Λ(h)=0\overline{\mathbb{SW}}_{\Lambda}(h)=0. Setting

(Φ(ε,τ),A(ε,τ))=(Φ1(ε,τ,ξ),A1(ε,τ,ξ))+χε+(φ,a)+χε(ψ+u,b),(\Phi(\varepsilon,\tau),A(\varepsilon,\tau))=(\Phi_{1}({\varepsilon,\tau,\xi}),A_{1}({\varepsilon,\tau,\xi}))\ +\ \chi_{\varepsilon}^{+}(\varphi,a)\ +\ \chi_{\varepsilon}^{-}(\psi+u,b),

where ξ=ξ(ε,τ)\xi=\xi(\varepsilon,\tau), the conclusion follows the Definition (8.12) of 𝕊𝕎¯\overline{\mathbb{SW}} and Corollary 8.5 with right-hand side g=ε1χεΛ(τ)Φτ,ξ(ε,τ)g=\varepsilon^{-1}\chi_{\varepsilon}^{-}\cdot\Lambda(\tau)\Phi_{\tau,\xi(\varepsilon,\tau)}, which satisfies gC(Y)g\in C^{\infty}(Y) as in the proof of that corollary. ∎

12.2. The One-Dimensional Obstruction

The configurations (12.1) solve the Seiberg–Witten equations if and only if

Λ(τ)+μ(ε,τ)=0\Lambda(\tau)+\mu(\varepsilon,\tau)=0 (12.2)

is satisfied. The next lemma shows that the assumption of transverse spectral crossing (Definition 1.5) means the condition (12.2) defines τ\tau implicitly as a function of ε\varepsilon.

Lemma 12.2.

The solutions (Φ(ε,τ),A(ε,τ),μ(ε,τ))(\Phi(\varepsilon,\tau),A(\varepsilon,\tau),\mu(\varepsilon,\tau)) from Proposition 12.1 depend smoothly on (ε,τ)(0,ε0)×(τ0,τ0)(\varepsilon,\tau)\in(0,\varepsilon_{0})\times(-\tau_{0},\tau_{0}). Moreover,

|μ(ε,τ)|+|τμ(ε,τ)|Cε2/3.|\mu(\varepsilon,\tau)|\ +\ |\partial_{\tau}\mu(\varepsilon,\tau)|\leq C\varepsilon^{2/3}. (12.3)

holds uniformly.

Proof.

The standard proof of the Banach fixed point theorem with smooth dependence on parameters shows that a smooth family of fixed points hτh_{\tau} satisfy

τh(τ)C(τTτ)h(τ).\|\partial_{\tau}h(\tau)\|_{\mathcal{H}}\leq C\|(\partial_{\tau}T_{\tau})h(\tau)\|_{\mathcal{H}}.

Because of the weight on μ\mu in (10.6), it therefore suffices to show that

(τTτ)h(τ)Cε1/4.\|(\partial_{\tau}T_{\tau})h(\tau)\|_{\mathcal{H}}\leq C\varepsilon^{-1/4}. (12.4)

By Theorem 7.5 Item (I), Lemma 7.11, the expression Proposition 8.7 (using Item (II) of Theorem 7.5), and the proofs of Propositions 9.1, 9.2, and 9.3 (which may be repeated equally well with (τΦ1,τA1)(\partial_{\tau}\Phi_{1},\partial_{\tau}A_{1})) show that

τ(Φ1,A1)++τ(Φ1,A1)+τd𝕊𝕎¯h1(ξ)Cεγ\|\partial_{\tau}\mathcal{L}_{(\Phi_{1},A_{1})}^{+}\|\ +\ \|\partial_{\tau}\mathcal{L}_{(\Phi_{1},A_{1})}^{-}\|+\|\partial_{\tau}\text{d}\overline{\mathbb{SW}}_{h_{1}}(\xi)\|\leq C\varepsilon^{-\gamma}

are uniformly bounded as maps the H1,±εL2H^{1,\pm}_{\varepsilon}\to L^{2} and 𝔚L2\mathfrak{W}\to L^{2}. A similar argument applies to τ\partial_{\tau}\mathbb{Q}. Together with, Item (I) of Theorem 7.5, these show that τ𝕊𝕎¯Λ\partial_{\tau}\overline{\mathbb{SW}}_{\Lambda} is bounded by CεγC\varepsilon^{-\gamma} at h𝒞h\in\mathcal{C}. Differentiating Id=Pτ1Pτ\text{Id}=P_{\tau}^{-1}P_{\tau} for all three parametrices and using the bounds from Propositions 11.111.3 yields bounds on τPξ,τP±\partial_{\tau}P_{\xi},\partial_{\tau}P^{\pm} by at most Cε1/12γC\varepsilon^{-1/12-\gamma}. Using the product rule on (11.32) and combining these yields (12.4). ∎

The proof of the following lemma is essentially identical the treatment of [60, Eq. (10.6)].

Lemma 12.3.

If the family of parameters pτ=(gτ,Bτ)p_{\tau}=(g_{\tau},B_{\tau}) has a transverse spectral crossing, then (12.2) implicitly defines a function τ(ε)\tau(\varepsilon) so that

Λ(τ(ε))+μ(ε,τ(ε))=0\Lambda(\tau(\varepsilon))+\mu(\varepsilon,\tau(\varepsilon))=0

for ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}), and either τ(ε)>0\tau(\varepsilon)>0 or τ(ε)<0\tau(\varepsilon)<0.

Proof.

The assumption that τ=0\tau=0 is a transverse spectral crossing means that Λ(0)=0\Lambda(0)=0 and Λ˙(0)0\dot{\Lambda}(0)\neq 0. By the inverse function theorem, there is an inverse Λ1\Lambda^{-1} defined on an open neighborhood of τ=0\tau=0. Set Γ(ε)=Λτ(ε)\Gamma(\varepsilon)=\Lambda\circ\tau(\varepsilon), so that (12.2) becomes the condition that

Γ(ε)+μ(ε,Λ1Γ(ε))=0.\Gamma(\varepsilon)+\mu(\varepsilon,\Lambda^{-1}\circ\Gamma(\varepsilon))=0.

This is an equation for a real number Γ\Gamma depending on a parameter ε\varepsilon. (12.3) implies that this equation can be solved for Γ\Gamma\in\mathbb{R} using the Inverse Function Theorem with smooth dependence on the parameter ε\varepsilon; thus τ(ε)=Λ1Γ(ε)\tau(\varepsilon)=\Lambda^{-1}\circ\Gamma(\varepsilon) solves (12.2).

To see the sign of τ(ε)\tau(\varepsilon), we expand (12.2) in series. From the proof of Proposition 10.1 implies that μ\mu is a sum μ(ε,τ)=μ1(ε,τ)+δμ2(ε,τ)+δ2μ3(ε,τ)+\mu(\varepsilon,\tau)=\mu_{1}(\varepsilon,\tau)+\delta\mu_{2}(\varepsilon,\tau)+\delta^{2}\mu_{3}(\varepsilon,\tau)+\ldots. Meanwhile Λ(τ)\Lambda(\tau) is smooth and may be expanded in Taylor series.

0\displaystyle 0 =\displaystyle= Λ(τ)+μ(ε,τ)\displaystyle\Lambda(\tau)+\mu(\varepsilon,\tau)
=\displaystyle= Λ˙(0)τ+ε2/3μ¯1(ε,0)+O(τ2)+O(ε2/3δ)+O(ε2/3τ),\displaystyle\dot{\Lambda}(0)\tau+\varepsilon^{2/3}\overline{\mu}_{1}(\varepsilon,0)+O(\tau^{2})+O(\varepsilon^{2/3}\delta)+O(\varepsilon^{2/3}\tau),

where μi=ε2/3μ¯i\mu_{i}=\varepsilon^{2/3}\overline{\mu}_{i}. It follows that for ε\varepsilon sufficiently small, τ\tau has the opposite sign of μ¯1(ε,0)Λ˙(0)\frac{\overline{\mu}_{1}(\varepsilon,0)}{\dot{\Lambda}(0)}. ∎

Proof of Theorem 1.6.

Part (1) follows directly from Lemma 3.8. Along the family of parameters τ(ε)\tau(\varepsilon) satisfying (12.2) constructed in Lemma 12.3, the glued configurations of Proposition 12.1 satisfy the (extended) Seiberg–Witten equations. Integrating by parts shows the 0-form component a0a_{0} vanishes, and setting (Ψε,Aε)=(ε1Φ(ε,τ(ε)),A(ε,τ(ε)))(\Psi_{\varepsilon},A_{\varepsilon})=(\varepsilon^{-1}{\Phi(\varepsilon,\tau(\varepsilon))}\ ,\ A(\varepsilon,\tau(\varepsilon))) yields the solutions in Part (2).

The glued configurations have ΨεL2=ε1+O(ε1/24γ)\|\Psi_{\varepsilon}\|_{L^{2}}=\varepsilon^{-1}+O(\varepsilon^{-1/24-\gamma}), and this norm depends smoothly on ε\varepsilon. By re-parameterizing, it may be assumed that εΨεL2=1\|\varepsilon\Psi_{\varepsilon}\|_{L^{2}}=1 as in (1.13). As ε0\varepsilon\to 0, Theorem 3.2 shows that after renormalization, (Ψε,Aε)(\Psi_{\varepsilon},A_{\varepsilon}) converges in ClocC^{\infty}_{loc} to a 2\mathbb{Z}_{2}-harmonic spinor for the parameter p0=(g0,B0)p_{0}=(g_{0},B_{0}) and that εΨε|Φ0|\|\varepsilon\Psi_{\varepsilon}\|\to|\Phi_{0}| in C0,αC^{0,\alpha}. Since (𝒵0,A0,Φ0)(\mathcal{Z}_{0},A_{0},\Phi_{0}) is regular, it is the unique 2\mathbb{Z}_{2}-harmonic spinor for this parameter and must therefore be the limit. This establishes Part (3). ∎

Remark 12.4.

Recalling the construction of μ\mu from Step 5 of the proof of Proposition 11.2 with N=1N=1, the leading order obstruction in the proof of Lemma 12.3 appears to be

ε2/3μ¯1(ε,0)=εdh0(η1,0)+dχ+.φ1,Φ0,\varepsilon^{2/3}\overline{\mu}_{1}(\varepsilon,0)=\varepsilon\langle\text{d}_{h_{0}}\not{\mathbb{D}}(\eta_{1},0)\ +\ d\chi^{+}.\varphi_{1},\Phi_{0}\rangle, (12.5)

where φ1\varphi_{1} is as in Item (II) of Theorem 7.5 for τ=0\tau=0. In fact, since Φ0=O(r1/2)\Phi_{0}=O(r^{1/2}), simple integration shows that the second term is smaller by a factor of ε1/2\varepsilon^{1/2} than the bound on dh0(η1,0)\text{d}_{h_{0}}\not{\mathbb{D}}(\eta_{1},0) obtained in the proof of Proposition 11.1. The latter bound, however, might not be optimal.

References

  • AG [07] Serge Alinhac and Patrick Gérard, Pseudo-differential Operators and the Nash-Moser Theorem, vol. 82, American Mathematical Soc., 2007.
  • APS [75] Michael F Atiyah, Vijay K Patodi, and Isadore M Singer, Spectral asymmetry and Riemannian geometry. I, Mathematical Proceedings of the Cambridge Philosophical Society, vol. 77, Cambridge University Press, 1975, pp. 43–69.
  • AS [63] Michael F Atiyah and Isadore M Singer, The index of elliptic operators on compact manifolds, Bulletin of the American Mathematical Society 69 (1963), no. 3, 422–433.
  • BB [11] Christian Bär and Werner Ballmann, Boundary value problems for elliptic differential operators of first order, arXiv preprint (2011), arXiv:1101.1196.
  • BdP [23] Marco Badran and Manuel del Pino, Entire solutions to 4 dimensional Ginzburg–Landau equations and codimension 2 minimal submanifolds, Advances in Mathematics 435 (2023), 109365.
  • BG [92] Jean-Pierre Bourguignon and Paul Gauduchon, Spineurs, Opérateurs de Dirac et Variations de Métriques, Communications in Mathematical Physics 144 (1992), no. 3, 581–599.
  • DK [97] Simon Kirwan Donaldson and Peter B Kronheimer, The Geometry of Four-Manifolds, Oxford university press, 1997.
  • Doa [19] Aleksander Doan, Monopoles and Fueter sections on Three-manifolds, 2019.
  • Don [86] Simon K Donaldson, Connections, cohomology and the intersection forms of 4-manifolds, Journal of Differential Geometry 24 (1986), no. 3, 275–341.
  • Don [16] Simon Donaldson, Calabi-Yau metrics on Kummer surfaces as a model glueing problem, arXiv preprint (2016), arXiv:1007.4218.
  • Don [17] Simon Donaldson, Adiabatic limits of co-associative Kovalev–Lefschetz fibrations, Algebra, Geometry, and Physics in the 21st Century: Kontsevich Festschrift (2017), 1–29.
  • Don [21] by same author, Deformations of Multivalued Harmonic Functions, The Quarterly Journal of Mathematics 72 (2021), no. 1-2, 199–235.
  • DS [11] Simon Donaldson and Ed Segal, Gauge Theory in Higher Dimensions, II, Surveys in differential geometry Volume XVI. Geometry of special holonomy and related topics. 16 (2011), 1–41.
  • DW [19] Aleksander Doan and Thomas Walpuski, On Counting Associative Submanifolds and Seiberg-Witten Monopoles, Pure and Applied Mathematics Quarterly 15 (2019), no. 4, 1047–1133.
  • DW [20] by same author, Deformation Theory of the Blown-up Seiberg–Witten Equation in Dimension Three, Selecta Mathematica 26 (2020), no. 3, 1–48.
  • DW [21] by same author, On the Existence of Harmonic 2\mathbb{Z}_{2}-Spinors, Journal of Differential Geometry 117 (2021), no. 3, 395–449.
  • FU [12] Daniel S Freed and Karen K Uhlenbeck, Instantons and Four-manifolds, vol. 1, Springer Science & Business Media, 2012.
  • Gri [01] Daniel Grieser, Basics of the b-calculus, Approaches to singular analysis, Springer, 2001, pp. 30–84.
  • Ham [82] Richard S Hamilton, The Inverse Function Theorem of Nash and Moser, Bulletin (New Series) of the American Mathematical Society 7 (1982), no. 1, 65–222.
  • Hay [08] Andriy Haydys, Nonlinear Dirac operator and Quaternionic Analysis, Communications in Mathematical Physics 281 (2008), 251–261.
  • Hay [12] by same author, Gauge theory, Calibrated Geometry and Harmonic Spinors, Journal of the London Mathematical Society 86 (2012), no. 2, 482–498.
  • Hay [15] by same author, Fukaya-Seidel Category and Gauge Theory, J. Symplectic Geom. 13 (2015), no. 1, 151–207.
  • Hay [17] by same author, G2 Instantons and the Seiberg-Witten Monopoles, arXiv: 1703.06329.
  • Hay [19] by same author, The Infinitesimal Multiplicities and Orientations of the Blow-up set of the Seiberg-Witten Equation with Multiple Spinors, Advances in Mathematics 5 (2019), 193–218.
  • [25] Siqi He, Existence of Nondegenerate 2\mathbb{Z}_{2}-Harmonic 1-forms via 3\mathbb{Z}_{3} Symmetry, arXiv Preprint (2022), arXiv:2202.12283.
  • [26] by same author, The Branched Deformations of the Special Lagrangian Submanifolds, arXiv Preprint (2022), arXiv 2202.12282.
  • Hit [74] Nigel Hitchin, Harmonic Spinors, Advances in Mathematics 14 (1974), no. 1, 1–55.
  • [28] Andriy Haydys, Rafe Mazzeo, and Ryosuke Takahashi, An index theorem for /2\mathbb{Z}/2-harmonic spinors branching along a graph, arXiv preprint (2023), arXiv:2310.15295.
  • [29] by same author, New examples of /2\mathbb{Z}/2 harmonic 1-forms and their deformations, arXiv preprint (2023), arXiv:2307.06227.
  • Hor [76] Lars Hormander, The Boundary Problems of Physical Geodesy, Archive for Rational Mechanics and Analysis 62 (1976), no. 1, 1–52.
  • HP [24] Siqi He and Gregory Parker, 2\mathbb{Z}_{2}-Harmonic Spinors on Connect Sums, Forthcoming work (2024).
  • HW [15] Andriy Haydys and Thomas Walpuski, A Compactness Theorem for the Seiberg–Witten Equation with Multiple Spinors in Dimension Three, Geometric and Functional Analysis 25 (2015), no. 6, 1799–1821.
  • Joy [16] Dominic Joyce, Conjectures on counting associative 3-folds in G2G_{2}-manifolds, arXiv preprint (2016), arXiv:1610.09836.
  • KM [07] Peter B Kronheimer and Tomasz Mrowka, Monopoles and Three-Manifolds, vol. 10, Cambridge University Press Cambridge, 2007.
  • Lic [63] A. Lichnerowicz, Spineurs Harmoniques, C.R. Acad. Sci. Paris, Series A–B (1963), no. 257, 7–9.
  • Maz [91] Rafe Mazzeo, Elliptic Theory of Differential Edge Operators I, Communications in Partial Differential Equations 16 (1991), no. 10, 1615–1664.
  • Mor [96] John W Morgan, The Seiberg-Witten Equations and Applications to the Topology of Smooth Four-manifolds, vol. 44, Princeton University Press, 1996.
  • MS [12] Dusa McDuff and Dietmar Salamon, J-holomorphic curves and symplectic topology, vol. 52, American Mathematical Soc., 2012.
  • MV [14] Rafe Mazzeo and Boris Vertman, Elliptic Theory of Differential Edge Operators, II: Boundary Value Problems, Indiana University Mathematics Journal (2014), 1911–1955.
  • MW [19] Tomasz Mrowka and Donghao Wang, Lectures Notes on Math 18.937, 2019.
  • Par [22] Gregory J. Parker, Concentrating Local Solutions of the Two-Spinor Seiberg-Witten Equations on 3-Manifolds, arXiv Preprint (2022), arXiv:2210.08148.
  • [42] Gregory J Parker, Concentrating Dirac Operators and Generalized Seiberg-Witten Equations, arXiv preprint (2023), arXiv:2307.00694.
  • [43] by same author, Deformations of 2\mathbb{Z}_{2}-Harmonic Spinors on 3-Manifolds, arXiv preprint (2023), arXiv:2301.06245.
  • PR [03] Frank Pacard and Manuel Ritore, From constant mean curvature hypersurfaces to the gradient theory of phase transitions, Journal of Differential Geometry 64 (2003), no. 3, 359–423.
  • SN [23] Ahmad Reza Haj Saeedi Sadegh and Minh Lam Nguyen, The three-dimensional Seiberg-Witten equations for 3/23/2-spinors: a compactness theorem, arXiv preprint (2023), arXiv:2311.11902.
  • Tak [15] Ryosuke Takahashi, The Moduli Space of S1S^{1}-type Zero Loci for /2\mathbb{Z}/2-Harmonic Spinors in Dimension 3, arXiv Preprint (2015), arXiv 1503.00767.
  • Tau [82] Clifford Henry Taubes, Self-dual Yang-Mills connections on non-self-dual 4-manifolds, Journal of Differential Geometry 17 (1982), no. 1, 139–170.
  • Tau [99] by same author, Nonlinear Generalizations of a 3-Manifold’s Dirac Operator, AMS IP Studies in Advanced Mathematics 13 (1999), 475–486.
  • [49] by same author, Compactness Theorems for SL(2,)SL(2,{\mathbb{C}}) Generalizations of the 4-dimensional Anti-Self Dual Equations, arXiv Preprint (2013), arXiv:1307.6447.
  • [50] by same author, PSL(2;)\text{PSL}(2;\mathbb{C}) Connections on 3-manifolds with L2L^{2} Bounds on Curvature, Cambridge Journal of Mathematics 1 (2013), 239–397.
  • Tau [14] by same author, The Zero Loci of /2\mathbb{Z}/2-Harmonic Spinors in Dimension 2, 3 and 4, arXiv Preprint (2014), arXiv:1407.6206.
  • Tau [16] by same author, On the Behavior of Sequences of Solutions to U(1)U(1)-Seiberg-Witten Systems in Dimension 4, arXiv Preprint (2016), arXiv:1610.07163.
  • Tau [17] by same author, The Behavior of Sequences of Solutions to the Vafa-Witten Equations, arXiv Preprint (2017), arXiv:1702.04610.
  • Tau [18] by same author, Sequences of Nahm pole Solutions to the SU(2)SU(2) Kapustin-Witten Equations, arXiv Preprint (2018), arXiv:1805.02773.
  • Tau [19] by same author, The \mathbb{R}-invariant solutions to the Kapustin–Witten equations on (0,)×2×(0,\infty)\times\mathbb{R}^{2}\times\mathbb{R} with generalized Nahm pole asymptotics, arXiv preprint arXiv:1903.03539 (2019).
  • TW [20] Clifford Henry Taubes and Yingying Wu, Examples of Singularity Models of /2\mathbb{Z}/2-harmonic 1-forms and Spinors in Dimension 3, arXiv Preprint (2020), arXiv 2001.00227.
  • [57] Karen K Uhlenbeck, Connections with LpL^{p} bounds on curvature, 1982, pp. 31–42.
  • [58] by same author, Removable singularities in Yang-Mills fields, Communications in Mathematical Physics 83 (1982), 11–29.
  • VW [94] Cumrun Vafa and Edward Witten, A Strong Coupling Test of S-Duality, Nuclear Physics B 431 (1994), no. 1-2, 3–77.
  • Wal [17] Thomas Walpuski, G2G_{2}-Instantons, Associative Submanifolds and Fueter Sections, Communications in Analysis and Geometry 25 (2017), no. 4, 847–893.
  • Wal [23] by same author, Lectures on Generalised Seiberg–Witten Equations, 2023.
  • Wan [22] Yuanqi Wang, Atiyah classes and the essential obstructions in deforming a singular G2G_{2}-instanton, Mathematische Zeitschrift 300 (2022), no. 3, 2997–3021.
  • Wit [11] Edward Witten, Fivebranes and Knots, Quantum Topology 3 (2011), no. 1, 1–137.
  • Wit [12] by same author, Khovanov Homology and Gauge Theory, Proceedings of the Freedman Fest 18 (2012), 291–308.
  • WZ [21] Thomas Walpuski and Boyu Zhang, On the Compactness Problem for a Family of Generalized Seiberg-Witten Equations in Dimension 3, Duke Mathematical Journal 170(17) (2021), 239–397.
  • Zha [17] Boyu Zhang, Rectifiability and Minkowski Bounds for the Zero Loci of 2\mathbb{Z}_{2}-Harmonic Spinors in Dimension 4, arXiv Preprint (to appear in Comm. Analysis and Geom.) (2017), arXiv 2202.12282.