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

\stackMath

Smoothing, scattering, and a conjecture of Fukaya

Kwokwai Chan Department of Mathematics
The Chinese University of Hong Kong
Shatin
Hong Kong
[email protected]
Naichung Conan Leung The Institute of Mathematical Sciences and Department of Mathematics
The Chinese University of Hong Kong
Shatin
Hong Kong
[email protected]
 and  Ziming Nikolas Ma Department of Mathematics
Southern University of Science and Technology
Nanshan District
Shenzhen
China
[email protected]
Abstract.

In 2002, Fukaya [19] proposed a remarkable explanation of mirror symmetry detailing the SYZ conjecture [47] by introducing two correspondences: one between the theory of pseudo-holomorphic curves on a Calabi-Yau manifold Xˇ\check{X} and the multi-valued Morse theory on the base Bˇ\check{B} of an SYZ fibration pˇ:XˇBˇ\check{p}\colon\check{X}\to\check{B}, and the other between deformation theory of the mirror XX and the same multi-valued Morse theory on Bˇ\check{B}. In this paper, we prove a reformulation of the main conjecture in Fukaya’s second correspondence, where multi-valued Morse theory on the base Bˇ\check{B} is replaced by tropical geometry on the Legendre dual BB. In the proof, we apply techniques of asymptotic analysis developed in [7, 9] to tropicalize the pre-dgBV algebra which governs smoothing of a maximally degenerate Calabi-Yau log variety X0\prescript{0}{}{X}^{\dagger} introduced in [8]. Then a comparison between this tropicalized algebra with the dgBV algebra associated to the deformation theory of the semi-flat part XsfXX_{\mathrm{sf}}\subseteq X allows us to extract consistent scattering diagrams from appropriate Maurer-Cartan solutions.

1. Introduction

Two decades ago, in an attempt to understand mirror symmetry using the SYZ conjecture [47], Fukaya [19] proposed two correspondences:

  • Correspondence I: between the theory of pseudo-holomorphic curves (instanton corrections) on a Calabi–Yau manifold Xˇ\check{X} and the multi-valued Morse theory on the base Bˇ\check{B} of an SYZ fibration pˇ:XˇBˇ\check{p}\colon\check{X}\to\check{B}, and

  • Correspondence II: between deformation theory of the mirror XX and the same multi-valued Morse theory on the base Bˇ\check{B}.

In this paper, we prove a reformulation of the main conjecture [19, Conj 5.3] in Fukaya’s Correspondence II, where multi-valued Morse theory on the SYZ base Bˇ\check{B} is replaced by tropical geometry on the Legendre dual BB. Such a reformulation of Fukaya’s conjecture was proposed and proved in [7] in a local setting; the main result of the current paper is a global version of the main result in loc. cit. A crucial ingredient in the proof is a precise link between tropical geometry on an integral affine manifold with singularities and smoothing of maximally degenerate Calabi–Yau varieties.

The main conjecture [19, Conj. 5.3] in Fukaya’s Correspondence II asserts that there exists a Maurer–Cartan element of the Kodaira–Spencer dgLa associated to deformations of the semi-flat part XsfX_{\mathrm{sf}} of XX that is asymptotically close to a Fourier expansion ([19, Eq. (42)]), whose Fourier modes are given by smoothings of distribution-valued 1-forms defined by moduli spaces of gradient Morse flow trees which are expected to encode counting of non-trivial (Maslov index 0) holomorphic disks bounded by Lagrangian torus fibers (see [19, Rem. 5.4]). Also, the complex structure defined by this Maurer–Cartan element can be compactified to give a complex structure on XX. At the same time, Fukaya’s Correspondence I suggests that these gradient Morse flow trees arise as adiabatic limits of loci of those Lagrangian torus fibers which bound non-trivial (Maslov index 0) holomorphic disks. This can be reformulated as a holomorphic/tropical correspondence, and much evidence has been found [18, 20, 39, 40, 12, 11, 38, 10, 4].

The tropical counterpart of such gradient Morse flow trees are given by consistent scattering diagrams, which were invented by Kontsevich–Soibelman [36] and extensively used in the Gross–Siebert program [29] to solve the reconstruction problem in mirror symmetry, namely, the construction of the mirror XX from smoothing of a maximally degenerate Calabi–Yau variety X0\prescript{0}{}{X}. It is therefore natural to replace the distribution-valued 1-form in each Fourier mode in the Fourier expansion [19, Eq. (42)] by a distribution-valued 1-form associated to a wall-crossing factor of a consistent scattering diagram. This was exactly how Fukaya’s conjecture [19, Conj. 5.3] was reformulated and proved in the local case in [7].

In order to reformulate the global version of Fukaya’s conjecture, however, we must also relate deformations of the semi-flat part XsfX_{\mathrm{sf}} with smoothings of the maximally degenerate Calabi–Yau variety X0\prescript{0}{}{X}. This is because consistent scattering diagrams were used by Gross–Siebert [28] to study the deformation theory of the compact log variety X0\prescript{0}{}{X}^{\dagger} (whose log structure is specified by slab functions), instead of XsfX_{\mathrm{sf}}. For this purpose, we consider the open dense part

Xsf0:=μ1(W0)X0,\prescript{0}{}{X}_{\mathrm{sf}}:=\mu^{-1}(W_{0})\subset\prescript{0}{}{X},

where μ:X0B\mu\colon\prescript{0}{}{X}\rightarrow B is the generalized moment map in [43] and W0BW_{0}\subseteq B is an open dense subset such that BW0B\setminus W_{0} contains the tropical singular locus and all codimension 22 cells of BB.

Equipping Xsf0\prescript{0}{}{X}_{\mathrm{sf}} with the trivial log structure, there is a semi-flat dgBV algebra 𝖯𝖵,\prescript{}{}{\mathsf{PV}}^{*,*} governing its smoothings, and the general fiber of a smoothing is given by the semi-flat Calabi–Yau XsfX_{\mathrm{sf}} that appeared in Fukaya’s original conjecture [19, Conj. 5.3]. However, the Maurer–Cartan elements of 𝖯𝖵,\prescript{}{}{\mathsf{PV}}^{*,*} cannot be compactified to give complex structures on XX. On the other hand, in our previous work [8] we constructed a Kodaira–Spencer–type pre-dgBV algebra PV,\prescript{}{}{PV}^{*,*} which controls the smoothing of X0\prescript{0}{}{X}. A key observation is that a twisting of 𝖯𝖵,\prescript{}{}{\mathsf{PV}}^{*,*} by slab functions is isomorphic to the restriction of PV,\prescript{}{}{PV}^{*,*} to Xsf0\prescript{0}{}{X}_{\mathrm{sf}} (Lemma 5.10).

Our reformulation of the global Fukaya conjecture now claims the existence of a Maurer–Cartan element ϕ\phi of this twisted semi-flat dgBV algebra that is asymptotically close to a Fourier expansion whose Fourier modes give rise to the wall-crossing factors of a consistent scattering diagram. This conjecture follows from (the proof of) our main result, stated as Theorem 1.1 below, which is a combination of Theorem 4.18, the construction in §5.3.2 and Theorem 5.24:

Theorem 1.1.

There exists a solution ϕ\phi to the classical Maurer–Cartan equation (4.11) giving rise to a smoothing of the maximally degenerate Calabi–Yau log variety X0\prescript{0}{}{X}^{\dagger} over [[q]]\mathbb{C}[[q]], from which a consistent scattering diagram 𝒟(ϕ)\mathscr{D}(\phi) can be extracted by taking asymptotic expansions.

A brief outline of the proof of Theorem 1.1 is now in order. First, recall that the pre-dgBV algebra PV,\prescript{}{}{PV}^{*,*} which governs smoothing of the maximally degenerate Calabi–Yau variety X0\prescript{0}{}{X} was constructed in [8, Thm. 1.1 & §3.5], and we also proved a Bogomolov–Tian–Todorov–type theorem [8, Thm. 1.2 & §5] showing unobstructedness of the extended Maurer–Cartan equation (4.10), under the Hodge-to-de Rham degeneracy Condition 4.16 and a holomorphic Poincaré Lemma Condition 4.14 (both proven in [28, 17]). In Theorem 4.18, we will further show how one can extract from the extended Maurer–Cartan equation (4.10) a smoothing of X0\prescript{0}{}{X}, described as a solution ϕPV1,1(B)\phi\in\prescript{}{}{PV}^{-1,1}(B) to the classical Maurer–Cartan equation (4.11)

¯ϕ+12[ϕ,ϕ]+𝔩=0,\bar{\partial}\phi+\frac{1}{2}[\phi,\phi]+\mathfrak{l}=0,

together with a holomorphic volume form efωe^{f}\prescript{}{}{\omega} which satisfies the normalization condition

(1.1) Tefω=1,\int_{T}e^{f}\prescript{}{}{\omega}=1,

where TT is a nearby vanishing torus in the smoothing.

Next, we need to tropicalize the pre-dgBV algebra PV,\prescript{}{}{PV}^{*,*}. However, the original construction of PV,\prescript{}{}{PV}^{*,*} in [8] using the Thom–Whitney resolution [49, 14] is too algebraic in nature. Here, we construct a geometric resolution exploiting the affine manifold structure on BB. Using the generalized moment map μ:X0B\mu\colon\prescript{0}{}{X}\rightarrow B [43] and applying the techniques of asymptotic analysis (in particular the notion of asymptotic support) in [7], we define the sheaf 𝒯\mathscr{T}^{*} of monodromy invariant tropical differential forms on BB in §5.1. According to Definition 5.5, a tropical differential form can be regarded as a distribution-valued form supported on polyhedral subsets of BB. Using the sheaf 𝒯\mathscr{T}^{*}, we can take asymptotic expansions of elements in PV,\prescript{}{}{PV}^{*,*}, and hence connect differential geometric operations in dgBV/dgLa with tropical geometry. In this manner, we can extract local scattering diagrams from Maurer–Cartan solutions as we did in [7], but we need to glue them together to get a global object.

To achieve this, we need the aforementioned comparison between PV,\prescript{}{}{PV}^{*,*} and the semi-flat dgBV algebra 𝖯𝖵sf,\prescript{}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}} which governs smoothing of the semi-flat part Xsf0:=μ1(W0)X0\prescript{0}{}{X}_{\mathrm{sf}}:=\mu^{-1}(W_{0})\subset\prescript{0}{}{X} equipped with the trivial log structure. The key Lemma 5.10 says that the restriction of PV,\prescript{}{}{PV}^{*,*} to the semi-flat part is isomorphic to 𝖯𝖵sf,\prescript{}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}} precisely after we twist the semi-flat operator ¯\bar{\partial}_{\circ} by elements corresponding to the slab functions associated to the initial walls of the form:

ϕin=vρδv,ρlog(fv,ρ)dˇρ;\phi_{\mathrm{in}}=-\sum_{v\in\rho}\delta_{v,\rho}\otimes\log(f_{v,\rho})\partial_{\check{d}_{\rho}};

here the sum is over vertices in codimension one cells ρ\rho’s which intersect with the essential singular locus 𝒮e\mathscr{S}_{e} (defined in §3.3), δv,ρ\delta_{v,\rho} is a distribution-valued 11-form supported on a component of ρ𝒮e\rho\setminus\mathscr{S}_{e} containing vv, dˇρ\partial_{\check{d}_{\rho}} is a holomorphic vector field and fv,ρf_{v,\rho}’s are the slab functions associated to the initial walls. We remark that slab functions were used to specify the log structure on X0\prescript{0}{}{X} as well as the local models for smoothing X0\prescript{0}{}{X} in the Gross–Siebert program; see §2 for a review.

Now, the Maurer–Cartan solution ϕPV1,1(B)\phi\in\prescript{}{}{PV}^{-1,1}(B) obtained in Theorem 4.18 defines a new operator ¯ϕ\bar{\partial}_{\phi} on PV,\prescript{}{}{PV}^{*,*} which squares to zero. Applying the above comparison of dgBV algebras (Lemma 5.10) and the gauge transformation from Lemma 5.11, we show that, after restricting to W0W_{0}, there is an isomorphism

(PV1,1(W0),¯ϕ)(𝖯𝖵sf1,1(W0),¯+[ϕin+ϕs,])\left(\prescript{}{}{PV}^{-1,1}(W_{0}),\bar{\partial}_{\phi}\right)\cong\left(\prescript{}{}{\mathsf{PV}}^{-1,1}_{\mathrm{sf}}(W_{0}),\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot]\right)

for some element ϕs\phi_{\mathrm{s}}, where ‘s’ stands for scattering terms. From the description of 𝒯\mathscr{T}^{*}, the element ϕs\phi_{\mathrm{s}}, to any fixed order kk, is written locally as a finite sum of terms supported on codimension one walls/slabs (Definitions 5.13 and 5.14. For the purpose of a brief discussion in this introduction, we will restrict ourselves to a wall 𝐰\mathbf{w} below, though the same argument applies to a slab; see §5.3.2 for the details. In a neighborhood U𝐰U_{\mathbf{w}} of each wall 𝐰\mathbf{w}, the operator ¯+[ϕin+ϕs,]\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot] is gauge equivalent to ¯\bar{\partial}_{\circ} via some vector field θ𝐰𝖯𝖵sf1,0(W0)\theta_{\mathbf{w}}\in\prescript{}{}{\mathsf{PV}}^{-1,0}_{\mathrm{sf}}(W_{0}), i.e.

e[θ𝐰,]¯e[θ𝐰,]=¯+[ϕin+ϕs,].e^{[\theta_{\mathbf{w}},\cdot]}\circ\bar{\partial}_{\circ}\circ e^{-[\theta_{\mathbf{w}},\cdot]}=\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot].

Employing the techniques for analyzing the gauge which we developed in [7, 9, 37], we see that the gauge will jump across the wall, resulting in a wall-crossing factor Θ𝐰\varTheta_{\mathbf{w}} satisfying

e[θ𝐰,]|𝒞±={Θ𝐰|𝒞+on U𝐰𝒞+,idon U𝐰𝒞,e^{[\theta_{\mathbf{w}},\cdot]}|_{\mathcal{C}_{\pm}}=\begin{dcases}\varTheta_{\mathbf{w}}|_{\mathcal{C}_{+}}&\text{on $U_{\mathbf{w}}\cap\mathcal{C}_{+}$,}\\ \mathrm{id}&\text{on $U_{\mathbf{w}}\cap\mathcal{C}_{-}$,}\end{dcases}

where 𝒞±\mathcal{C}_{\pm} are the two chambers separated by 𝐰\mathbf{w}. Then from the fact that the volume form efωe^{f}\prescript{}{}{\omega} is normalized as in (1.1), it follows that ϕs\phi_{\mathrm{s}} is closed under the semi-flat BV operator Δ\prescript{}{}{\Delta}, and hence we deduce that the wall-crossing factor Θ𝐰\varTheta_{\mathbf{w}} lies in the tropical vertex group. This defines a scattering diagram 𝒟(ϕ)\mathscr{D}(\phi) on the semi-flat part W0W_{0} associated to ϕ\phi. Finally, we prove consistency of the scattering diagram 𝒟(ϕ)\mathscr{D}(\phi) in Theorem 5.24. We emphasize that the consistency is over the whole BB even though the diagram is only defined on W0W_{0}, because the Maurer–Cartan solution ϕ\phi is globally defined on BB.

Remark 1.2.

Our notion of scattering diagrams (Definition 5.17) is a little bit more relaxed than the usual notion defined in [36, 29] in two aspects: One is that we do not require the generator of the exponents of the wall-crossing factor to be orthogonal to the wall.111It seems reasonable to relax this orthogonality condition because one cannot require such a condition in more general settings [5, 37]. The other is that we allow possibly infinite number of walls/slabs approaching strata of the tropical singular locus. See the paragraph after Definition 5.17 for more details. In practice, this simply means that we are considering a larger gauge equivalence class (or equivalently, a weaker gauge equivalence), which is natural from the point of view of both the Bogomolov–Tian–Todorov Theorem and mirror symmetry (in the A-side, this amounts to flexibility in the choice of the almost complex structure). We also have a different, but more or less equivalent, formulation of the consistency of a scattering diagram; see Definition 5.21 and §5.3.1.

Along the way of proving Fukaya’s conjecture, besides figuring out the precise relation between the semi-flat part XsfX_{\mathrm{sf}} and the maximally degenerate Calabi–Yau log variety X0\prescript{0}{}{X}^{\dagger}, we also find the correct description of the Maurer–Cartan solutions near the singular locus, namely, they should be extendable to the local models prescribed by the log structure (or slab functions), as was hinted by the Gross–Siebert program. This is related to a remark by Fukaya [19, Pt. (2) after Conj. 5.3].

Another important point is that we have established in the global setting an interplay between the differential-geometric properties of the tropical dgBV algebra and the scattering (and other combinatorial) properties of tropical disks, which was speculated by Fukaya as well ([19, Pt. (1) after Conj. 5.3]) although he considered holomorphic disks instead of tropical ones.

Furthermore, by providing a direct linkage between Fukaya’s conjecture with the Gross–Siebert program [27, 28, 29] and Katzarkov–Kontsevich–Pantev’s Hodge theoretic viewpoint [33] through PV,\prescript{}{}{PV}^{*,*} (recall from [8] that a semi-infinite variation of Hodge structures can be constructed from PV,\prescript{}{}{PV}^{*,*}, using the techniques of Barannikov–Kontsevich [3, 2] and Katzarkov–Kontsevich–Pantev [33]), we obtain a more transparent understanding of mirror symmetry through the SYZ framework.

Remark 1.3.

A future direction is to apply the framework in this paper and the works [7, 8] to develop a local-to-global approach to understand genus 0 mirror symmetry. In view of the ideas of Seidel [46] and Kontsevich [35], and also recent breakthroughs by Ganatra–Pardon–Shende [25, 26, 24] and Gammage–Shende [22, 23], we expect that there is a sheaf of LL_{\infty} algebras on the A-side mirror to (the LL_{\infty} enhancement of) PV,\prescript{}{}{PV}^{*,*} that can be constructed by gluing local models. More precisely, a large volume limit of a Calabi–Yau manifold Xˇ\check{X} can be specified by removing from it a normal crossing divisor Dˇ\check{D} which represents the Kähler class of Xˇ\check{X}. This gives rise to a Weinstein manifold XˇDˇ\check{X}\setminus\check{D}, and produces a mirror pair XˇDˇX0\check{X}\setminus\check{D}\leftrightarrow\prescript{0}{}{X} at the large volume/complex structure limits. In [23], Gammage–Shende constructed a Lagrangian skeleton Λ(Φ)XˇDˇ\Lambda(\Phi)\subset\check{X}\setminus\check{D} from a combinatorial structure Φ\Phi called fanifold, which can be extracted from the integral tropical manifold BB equipped with a polyhedral decomposition 𝒫\mathscr{P} (here we assume that the gluing data ss is trivial). They also proved an HMS statement at the large limits. We expect that an A-side analogue of PV,\prescript{}{}{PV}^{*,*} can be constructed from the Lagrangian skeleton Λ(Φ)\Lambda(\Phi) in XˇDˇ\check{X}\setminus\check{D}, possibly together with a nice and compatible SYZ fibration on XˇDˇ\check{X}\setminus\check{D}, via gluing of local models. A local-to-global comparsion on the A-side and isomorphisms between the local models on the two sides should then yield an isomorphism of Frobenius manifolds.

Acknowledgement

We thank Kenji Fukaya, Mark Gross and Richard Thomas for their interest and encouragement, and also Helge Ruddat for useful comments on an earlier draft of this paper. We are very grateful to the anonymous referees for numerous constructive and extremely detailed comments/suggestions which have helped to greatly enhanced the exposition of the whole paper.

K. Chan was supported by grants of the Hong Kong Research Grants Council (Project No. CUHK14301420 & CUHK14301621) and direct grants from CUHK. N. C. Leung was supported by grants of the Hong Kong Research Grants Council (Project No. CUHK14301619 & CUHK14306720) and a direct grant (Project No. 4053400) from CUHK. Z. N. Ma was supported by National Science Fund for Excellent Young Scholars (Overseas). These authors contributed equally to this work.

List of notations

MM, MAM_{A} §2.1 lattice, MA:=MAM_{A}:=M\otimes_{\mathbb{Z}}A for any \mathbb{Z}-module AA
NN, NAN_{A} §2.1 dual lattice of MM, NA:=NAN_{A}:=N\otimes_{\mathbb{Z}}A for any \mathbb{Z}-module AA
(B,𝒫)(B,\mathscr{P}) Def. 2.2 integral tropical manifold equipped with a polyhedral
decomposition
Λσ\Lambda_{\sigma} §2.1 lattice generated by integral tangent vectors along σ\sigma
intre(τ)\mathrm{int}_{\mathrm{re}}(\tau) §2.1 relative interior of a polyhedron τ\tau
UτU_{\tau} §2.1 open neighborhood of intre(τ)\mathrm{int}_{\mathrm{re}}(\tau)
𝒬τ\mathscr{Q}_{\tau} §2.1 lattice generated by normal vectors to τ\tau
Sτ:Uτ𝒬τ,S_{\tau}\colon U_{\tau}\rightarrow\mathscr{Q}_{\tau,\real} §2.1 fan structure along τ\tau
Στ\Sigma_{\tau} §2.1 complete fan in 𝒬τ,\mathscr{Q}_{\tau,\real} constructed from SτS_{\tau}
KτσK_{\tau}\sigma §2.1 Kτσ=Sτ0(σUτ)K_{\tau}\sigma={}_{\geq 0}S_{\tau}(\sigma\cap U_{\tau}) is a cone in Στ\Sigma_{\tau} corresponding to σ\sigma
TxT_{x} §2.2 lattice of integral tangent vectors of BB at xx
Δi(τ)\Delta_{i}(\tau), Δˇi(τ)\check{\Delta}_{i}(\tau) Def. 2.9 monodromy polytope of τ\tau, dual monodromy polytope of τ\tau
𝒜𝑓𝑓\mathcal{A}\mathit{ff} Def. 2.5 sheaf of affine functions on BB
𝒫𝒫\mathcal{PL}_{\mathscr{P}} Def. 2.5 sheaf of piecewise affine functions on BB with respect to 𝒫\mathscr{P}
𝒫𝒫\mathcal{MPL}_{\mathscr{P}} Def. 2.6 sheaf of multi-valued piecewise affine functions on BB
with respect to 𝒫\mathscr{P}
φ\varphi Def. 2.7 strictly convex multi-valued piecewise linear function
τ1Σv\tau^{-1}\Sigma_{v} §2.3 localization of the fan Σv\Sigma_{v} at τ\tau
V(τ)V(\tau) §2.3 local affine scheme associated to τ\tau used for open gluing
PM(τ)\mathrm{PM}(\tau) §2.3 group of piecewise multiplicative maps on τ1Σv\tau^{-1}\Sigma_{v}
D(μ,ρ,v)D(\mu,\rho,v) Def. 2.15 number encoding the change of μPM(τ)\mu\in\mathrm{PM}(\tau) across ρ\rho through vv
Xτ0\prescript{0}{}{X}_{\tau} §2.3 closed stratum of X0\prescript{0}{}{X} associated to τ\tau
CτC_{\tau} §2.4 cone defined by the strictly convex function φ¯τ:Στ\bar{\varphi}_{\tau}\colon\Sigma_{\tau}\rightarrow\real
representing φ\varphi
P¯τ\bar{P}_{\tau} §2.4 monoid of integral points in CτC_{\tau}
q=zϱq=z^{\varrho} §2.4 parameter for a toric degeneration
𝒩ρ\mathcal{N}_{\rho} §2.4 line bundle on Xρ0\prescript{0}{}{X}_{\rho} having slab functions fρf_{\rho} as sections
fvρf_{v\rho} §2.4 local slab function associate to ρ\rho in the chart V(v)V(v)
ϰτ,i:Xτ0rτ,i\varkappa_{\tau,i}\colon\prescript{0}{}{X}_{\tau}\rightarrow\mathbb{P}^{r_{\tau,i}} §2.4 toric morphism induced from the monodromy polytope Δi(τ)\Delta_{i}(\tau)
Pτ,xP_{\tau,x} §2.4 toric monoid describing the local model of toric degeneration
near xXτ0x\in\prescript{0}{}{X}_{\tau}
Qτ,xQ_{\tau,x} §2.4 toric monoid isomorphic to Pτ,x/(ϱ+Pτ,x)P_{\tau,x}/(\varrho+P_{\tau,x})
𝒩τ\mathscr{N}_{\tau} §2.4 normal fan of a polytope τ\tau
μ:X0B\mu\colon\prescript{0}{}{X}\rightarrow B §3.1 generalized moment map
Υτ\Upsilon_{\tau} §3.2 coordinate chart on W(τ)BW(\tau)\subset B
𝒮\mathscr{S} (resp. 𝒮e\mathscr{S}_{e}) §3.3 (resp. essential) tropical singular locus in BB
ν:X0B\nu\colon\prescript{0}{}{X}\rightarrow B Def. 3.6 surjective map with ν(Z)𝒮e\nu(Z)\subset\mathscr{S}_{e}
𝒲={Wα}α\mathcal{W}=\{W_{\alpha}\}_{\alpha} §4 good cover (Condition 4.1) of BB with Vα:=ν1(Wα)V_{\alpha}:=\nu^{-1}(W_{\alpha}) being Stein
𝕍αk\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger} §4 kthk^{\text{th}}-order local smoothing model of VαV_{\alpha}
𝒢αk\prescript{k}{}{\mathcal{G}}_{\alpha}^{*} Def. 4.2 sheaf of kthk^{\text{th}}-order holomorphic relative log polyvector fields on 𝕍αk\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}
𝒦αk\prescript{k}{}{\mathcal{K}}_{\alpha}^{*} Def. 4.2 sheaf of kthk^{\text{th}}-order holomorphic log de Rham differentials on 𝕍αk\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}
𝒦αk\prescript{k}{\parallel}{\mathcal{K}}_{\alpha}^{*} §4.1 sheaf of kthk^{\text{th}}-order holomorphic relative log de Rham differentials on 𝕍αk\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}
ωαk\prescript{k}{}{\omega}_{\alpha} Def. 4.2 kthk^{\text{th}}-order relative log volume form on 𝕍αk\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}
Δαk\prescript{k}{}{\Delta}_{\alpha} §4.1 BV operator on 𝒢αk\prescript{k}{}{\mathcal{G}}_{\alpha}
PkVα,\prescript{k}{}{PV}^{*,*}_{\alpha} Def. 4.8 local sheaf of kthk^{\text{th}}-order polyvector fields
𝒜α,k\prescript{k}{}{\mathcal{A}}^{*,*}_{\alpha} Def. 4.9 local sheaf of kthk^{\text{th}}-order de Rham forms
PkV,\prescript{k}{}{PV}^{*,*} Def. 4.13 global sheaf of kthk^{\text{th}}-order polyvector fields from gluing of PkVα,\prescript{k}{}{PV}^{*,*}_{\alpha}’s
𝒜,k\prescript{k}{}{\mathcal{A}}^{*,*} Def. 4.13 global sheaf of kthk^{\text{th}}-order de Rham forms from gluing of 𝒜α,k\prescript{k}{}{\mathcal{A}}^{*,*}_{\alpha}’s
𝒯\mathscr{T}^{*} Def. 5.6 global sheaf of tropical differential forms on BB
W0W_{0} §5.2.1 semi-flat locus
𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} §5.2.1 sheaf of kthk^{\text{th}}-order semi-flat holomorphic relative vector fields
𝖪sfk\prescript{k}{}{\mathsf{K}}^{*}_{\mathrm{sf}} §5.2.1 sheaf of kthk^{\text{th}}-order semi-flat holomorphic log de Rham forms
𝔥k\prescript{k}{}{\mathfrak{h}} eqt. (5.2) sheaf of kthk^{\text{th}}-order semi-flat holomorphic tropical vertex Lie algebras
𝖯𝖵sf,k\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}} Def. 5.9 sheaf of kthk^{\text{th}}-order semi-flat polyvector fields
𝖠sf,k\prescript{k}{}{\mathsf{A}}^{*,*}_{\mathrm{sf}} Def. 5.9 sheaf of kthk^{\text{th}}-order semi-flat log de Rham forms
𝖳𝖫sfk\prescript{k}{}{\mathsf{TL}}^{*}_{\mathrm{sf}} Def. 5.12 sheaf of kthk^{\text{th}}-order semi-flat tropical vertex Lie algebras
(𝐰,Θ𝐰)(\mathbf{w},\Theta_{\mathbf{w}}) Def. 5.13 wall equipped with a wall-crossing factor
(𝐛,Θ𝐛)(\mathbf{b},\Theta_{\mathbf{b}}) Def. 5.14 slab equipped with a wall-crossing factor
𝒟\mathscr{D} Def. 5.17 scattering diagram
W0(𝒟)W_{0}(\mathscr{D}) §5.3.1 complement of joints in the semi-flat locus
𝔦\mathfrak{i} §5.3.1 the embedding 𝔦:W0(𝒟)B\mathfrak{i}\colon W_{0}(\mathscr{D})\rightarrow B
𝒪𝒟k\prescript{k}{}{\mathscr{O}}_{\mathscr{D}} §5.3.1 kthk^{\text{th}}-order wall-crossing sheaf associated to 𝒟\mathscr{D}
Notation 1.4.

We usually fix a rank ss lattice 𝐊\mathbf{K} together with a strictly convex ss-dimensional rational polyhedral cone Q𝐊=𝐊Q\subset\mathbf{K}=\mathbf{K}\otimes_{\mathbb{Z}}\real. We call Q:=Q𝐊Q:=Q\cap\mathbf{K} the universal monoid. We consider the ring R:=[Q]R:=\mathbb{C}[Q], a monomial element of which is written as qmRq^{m}\in R for mQm\in Q, and the maximal ideal 𝐦:=[Q{0}]\mathbf{m}:=\mathbb{C}[Q\setminus\{0\}]. Then Rk:=R/𝐦k+1\prescript{k}{}{R}:=R/\mathbf{m}^{k+1} is an Artinian ring, and we denote by R^:=limkRk\hat{R}:=\varprojlim_{k}\prescript{k}{}{R} the completion of RR. We further equip RR, Rk\prescript{k}{}{R} and R^\hat{R} with the natural monoid homomorphism QRQ\rightarrow R, mqmm\mapsto q^{m}, which gives them the structure of a log ring (see [29, Definition 2.11]); the corresponding log analytic spaces are denoted as SS^{\dagger}, Sk\prescript{k}{}{S}^{\dagger} and S^\hat{S}^{\dagger} respectively.

Furthermore, we let ΩS:=R𝐊\prescript{}{}{\Omega}^{*}_{S^{\dagger}}:=R\otimes_{\mathbb{C}}\bigwedge^{*}\mathbf{K}_{\mathbb{C}}, ΩSk:=Rk𝐊\prescript{k}{}{\Omega}^{*}_{S^{\dagger}}:=\prescript{k}{}{R}\otimes_{\mathbb{C}}\bigwedge^{*}\mathbf{K}_{\mathbb{C}} and Ω^S:=R^𝐊\hat{\Omega}^{*}_{S^{\dagger}}:=\hat{R}\otimes_{\mathbb{C}}\bigwedge^{*}\mathbf{K}_{\mathbb{C}} (here 𝐊=𝐊\mathbf{K}_{\mathbb{C}}=\mathbf{K}\otimes_{\mathbb{Z}}\mathbb{C}) be the spaces of log de Rham differentials on SS^{\dagger}, Sk\prescript{k}{}{S}^{\dagger} and S^\hat{S}^{\dagger} respectively, where we write 1m=dlogqm1\otimes m=d\log q^{m} for m𝐊m\in\mathbf{K}; these are equipped with the de Rham differential \partial satisfying (qm)=qmdlogqm\partial(q^{m})=q^{m}d\log q^{m}. We also denote by ΘS:=R𝐊\prescript{}{}{\Theta}_{S^{\dagger}}:=R\otimes_{\mathbb{C}}\mathbf{K}_{\mathbb{C}}^{\vee}, ΘS\prescript{}{}{\Theta}_{S^{\dagger}} and Θ^S\hat{\Theta}_{S^{\dagger}}, respectively, the spaces of log derivations, which are equipped with a natural Lie bracket [,][\cdot,\cdot]. We write n\partial_{n} for the element 1n1\otimes n with action n(qm)=(m,n)qm\partial_{n}(q^{m})=(m,n)q^{m}, where (m,n)(m,n) is the natural pairing between 𝐊\mathbf{K}_{\mathbb{C}} and 𝐊\mathbf{K}^{\vee}_{\mathbb{C}}.

2. Gross–Siebert’s cone construction of maximally degenerate Calabi–Yau varieties

This section is a brief review of Gross–Siebert’s construction of the maximally degenerate Calabi–Yau variety X0\prescript{0}{}{X} from the affine manifold BB and its log structures from slab functions [27, 28, 29].

2.1. Integral tropical manifolds

We first recall the notion of integral tropical manifolds from [29, §1.1]. Given a lattice MM of rank nn, a rational convex polyhedron σ\sigma is a convex subset in MM given by a finite intersection of rational (i.e. defined over MM_{\mathbb{Q}}) affine half-spaces. We usually drop the attributes “rational” and “convex” for polyhedra. A polyhedron σ\sigma is said to be integral if all its vertices lie in MM; a polytope is a compact polyhedron. The group 𝐀𝐟𝐟(M):=MGL(M)\mathbf{Aff}(M):=M\rtimes\mathrm{GL}(M) of integral affine transformations acts on the set of polyhedra in MM. Given a polyhedron σM\sigma\subset M, let Λσ,M\Lambda_{\sigma,\real}\subset M be the smallest affine subspace containing σ\sigma, and denote by Λσ:=Λσ,M\Lambda_{\sigma}:=\Lambda_{\sigma,\real}\cap M the corresponding lattice. The relative interior intre(σ)\mathrm{int}_{\mathrm{re}}(\sigma) refers to taking the interior of σ\sigma in Λσ,\Lambda_{\sigma,\real}. There is an identification Tσ,xΛσ,T_{\sigma,x}\cong\Lambda_{\sigma,\real} for the tangent space at xintre(σ)x\in\mathrm{int}_{\mathrm{re}}(\sigma). Write σ=σintre(σ)\partial\sigma=\sigma\setminus\mathrm{int}_{\mathrm{re}}(\sigma). Then a face of σ\sigma is the intersection of σ\partial\sigma with a supporting hyperplane. Codimension one faces are called facets.

Let LPoly¯\underline{\mathrm{LPoly}} be the category whose objects are integral polyhedra and morphisms consist of the identity and integral affine isomorphisms onto faces (i.e. an integral affine morphism τσ\tau\rightarrow\sigma which is an isomorphism onto its image and identifies τ\tau with a face of σ\sigma). An integral polyhedral complex is a functor 𝙵:𝒫LPoly¯\mathtt{F}\colon\mathscr{P}\rightarrow\underline{\mathrm{LPoly}} from a finite category 𝒫\mathscr{P} to LPoly¯\underline{\mathrm{LPoly}} such that every face of 𝙵(σ)\mathtt{F}(\sigma) still lies in the image of 𝙵\mathtt{F}, and there is at most one arrow τσ\tau\rightarrow\sigma for every pair τ,σ𝒫\tau,\sigma\in\mathscr{P}. By abuse of notation, we usually drop the notation 𝙵\mathtt{F} and write σ𝒫\sigma\in\mathscr{P} to represent an integral polyhedron in the image of the functor. From an integral polyhedral complex, we obtain a topological space B:=limσ𝒫σB:=\varinjlim_{\sigma\in\mathscr{P}}\sigma via gluing of the polyhedra along faces. We further assume that:

  1. (1)

    the natural map σB\sigma\rightarrow B is injective for each σ𝒫\sigma\in\mathscr{P}, so that σ\sigma can be identified with a closed subset of BB called a cell, and a morphism τσ\tau\rightarrow\sigma can be identified with an inclusion of subsets;

  2. (2)

    a finite intersection of cells is a cell; and

  3. (3)

    BB is an orientable connected topological manifold of dimension nn without boundary which in addition satisfies the condition that H1(B,)=0H^{1}(B,\mathbb{Q})=0.

Remark 2.1.

The condition H1(B,)=0H^{1}(B,\mathbb{Q})=0 will be used only in Theorem 4.18 to ensure that H1(X0,𝒪)=H1(B,)=0H^{1}(\prescript{0}{}{X},\mathcal{O})=H^{1}(B,\mathbb{C})=0, where X0\prescript{0}{}{X} is the degenerate Calabi–Yau variety that we are going to construct.222In his recent work [15], Felten was able to prove Theorem 4.18 without assuming that H1(B,)=0H^{1}(B,\mathbb{Q})=0. This corresponds to the condition that b1=0b_{1}=0 for smooth Calabi–Yau manifolds.

The set of kk-dimensional cells is denoted by 𝒫[k]\mathscr{P}^{[k]}, and the kk-skeleton by 𝒫[k]\mathscr{P}^{[\leq k]}. For every τ𝒫\tau\in\mathscr{P}, we define its open star by

Uτ:=στintre(σ),U_{\tau}:=\bigcup_{\sigma\supset\tau}\mathrm{int}_{\mathrm{re}}(\sigma),

which is an open subset of BB containing intre(τ)\mathrm{int}_{\mathrm{re}}(\tau). A fan structure along τ𝒫[nk]\tau\in\mathscr{P}^{[n-k]} is a continuous map Sτ:UτkS_{\tau}\colon U_{\tau}\rightarrow{}^{k} such that

  • Sτ1(0)=intre(τ)S^{-1}_{\tau}(0)=\mathrm{int}_{\mathrm{re}}(\tau),

  • for every στ\sigma\supset\tau, the restriction Sτ|intre(σ)S_{\tau}|_{\mathrm{int}_{\mathrm{re}}(\sigma)} is an integral affine submersion onto its image (meaning that it is induced by some epimorphism ΛσWk\Lambda_{\sigma}\rightarrow W\cap\mathbb{Z}^{k} for some vector subspace WkW\subset{}^{k}), and

  • the collection of cones {Kτσ:=Sτ0(σUτ)}στ\{K_{\tau}\sigma:={}_{\geq 0}S_{\tau}(\sigma\cap U_{\tau})\}_{\sigma\supset\tau} forms a complete finite fan Στ\Sigma_{\tau}.

Two fan structures along τ\tau are equivalent if they differ by composition with an integral affine transformation of k. If SτS_{\tau} is a fan structure along τ\tau and στ\sigma\supset\tau, then UσUτU_{\sigma}\subset U_{\tau} and there is a fan structure along σ\sigma induced from SτS_{\tau} via the composition:

UσUτkl,U_{\sigma}\hookrightarrow U_{\tau}\rightarrow\mathbb{R}^{k}\twoheadrightarrow\mathbb{R}^{l},

where kk/Sτ(σUτ)l\mathbb{R}^{k}\rightarrow\mathbb{R}^{k}/\real S_{\tau}(\sigma\cap U_{\tau})\cong\mathbb{R}^{l} is the quotient map.

Definition 2.2 ([29], Def. 1.2).

An integral tropical manifold is an integral polyhedral complex (B,𝒫)(B,\mathscr{P}) together with a fan structure SτS_{\tau} along each τ𝒫\tau\in\mathscr{P} such that whenever τσ\tau\subset\sigma, the fan structure induced from SτS_{\tau} is equivalent to SσS_{\sigma}.

Taking sufficiently small and mutually disjoint open subsets WvUvW_{v}\subset U_{v} for v𝒫[0]v\in\mathscr{P}^{[0]} and intre(σ)\mathrm{int}_{\mathrm{re}}(\sigma) for σ𝒫[n]\sigma\in\mathscr{P}^{[n]}, there is an integral affine structure on v𝒫[0]Wvσ𝒫[n]intre(σ)\bigcup_{v\in\mathscr{P}^{[0]}}W_{v}\cup\bigcup_{\sigma\in\mathscr{P}^{[n]}}\mathrm{int}_{\mathrm{re}}(\sigma). We will further choose the open subsets WvW_{v}’s and intre(σ)\mathrm{int}_{\mathrm{re}}(\sigma)’s so that the affine structure is defined outside a closed subset Γ\Gamma of codimension two in BB, as in [27, §1.3]. This affine structure allows us to use parallel transport to identify the tangent spaces TxBT_{x}B for different points xx outside the closed subset. For every τ\tau we choose a maximal cell στ\sigma\supset\tau and consider the lattice of normal vectors 𝒬τ=Λσ/Λτ\mathscr{Q}_{\tau}=\Lambda_{\sigma}/\Lambda_{\tau} (we suppress the dependence on σ\sigma because we will see that Λτ\Lambda_{\tau} is monodromy invariant under the monodromy transformation given by any two vertices of τ\tau and any two maximal cells containing τ\tau). We can identify 𝒬τ\mathscr{Q}_{\tau} with k\mathbb{Z}^{k} via SτS_{\tau}, and write the fan structure as Sτ:Uτ𝒬τ,S_{\tau}\colon U_{\tau}\rightarrow\mathscr{Q}_{\tau,\real}.

Example 2.3.

We take a 22-dimensional example from [1, Ex. 6.74] to illustrate the above definitions. Let Ξ\Xi be the convex hull of the points

p0=[111],p1=[311],p2=[131],p3=[113],p_{0}=\begin{bmatrix}-1\\ -1\\ -1\end{bmatrix},\ p_{1}=\begin{bmatrix}3\\ -1\\ -1\end{bmatrix},\ p_{2}=\begin{bmatrix}-1\\ 3\\ -1\end{bmatrix},\ p_{3}=\begin{bmatrix}-1\\ -1\\ 3\end{bmatrix},

so Ξ\Xi is a 33-simplex. Take BB (as a topological space) to be the boundary of Ξ\Xi. The polyhedral decomposition 𝒫\mathscr{P} is defined so that the integral points are vertices as shown in Figure 1.

Refer to caption
Figure 1. The polyhedral decomposition

Then we define affine coordinate charts on σ𝒫[n]intre(σ)v𝒫[0]Wv\bigcup_{\sigma\in\mathscr{P}^{[n]}}\mathrm{int}_{\mathrm{re}}(\sigma)\cup\bigcup_{v\in\mathscr{P}^{[0]}}W_{v} as follows. On intre(σ)\mathrm{int}_{\mathrm{re}}(\sigma), we take ψσ:intre(σ)Λσ,\psi_{\sigma}\colon\mathrm{int}_{\mathrm{re}}(\sigma)\rightarrow\Lambda_{\sigma,\real} which maps homeomorphically onto its image. At a vertex vv treated as a vector in 3, we let ψv:Wv3/3v\psi_{v}\colon W_{v}\subset{}^{3}\rightarrow{}^{3}/\real v, where 3/3v{}^{3}\rightarrow{}^{3}/\real v is the natural projection onto the quotient. By [1, Prop. 6.81], this gives an integral affine manifold with singularities. The affine structure can be extended to the complement of a subset Γ\Gamma consisting of 2424 points lying on the six edges of Ξ\Xi, with each edge containing 44 points (colored in red in Figure 1). The fan structure SτS_{\tau} can be defined similarly.

Locally near each singular point pΓp\in\Gamma contained in an edge ρ\rho, the affine structure is described as a gluing of two affine charts UI2{0}×0U_{\mathrm{I}}\subset{}^{2}\setminus\{0\}\times{}_{\geq 0} and UII20×0U_{\mathrm{II}}\subset{}^{2}\setminus 0\times{}_{\leq 0} as in [30, §3.2]. The change of coordinates from UIU_{\mathrm{I}} to UIIU_{\mathrm{II}} is given by the restriction of the map Υ\Upsilon from ({0})×(\real\setminus\{0\})\times\real to itself defined by

(x,y){(x,y),x<0(x,x+y),x>0.(x,y)\mapsto\begin{cases}(x,y),&x<0\\ (x,x+y),&x>0.\end{cases}

The fan structure Sρ:UρS_{\rho}\colon U_{\rho}\rightarrow\real is given as Sρ(x,y)=xS_{\rho}(x,y)=x and the fan Σρ\Sigma_{\rho} is the toric fan for 1\mathbb{P}^{1}. Figure 2 below illustrates the situation.

Refer to caption
Figure 2. Affine coordinate charts

With the structure of an integral tropical manifold, the corners and edges in Figure 1 are flattened via the affine coordinate charts, and we can view (B,𝒫)(B,\mathscr{P}) as the 2-sphere equipped with a polyhedral decomposition and with 2424 affine singularities. Such an affine structure with singularities also appears in the base BB of an SYZ fibration of a K3 surface.

Example 2.4.

A 33-dimensional example can be constructed as in [1, Ex. 6.74]. Take Ξ\Xi to be the convex hull of the points

p0=[1111],p1=[4111],p2=[1411],p3=[1141],p4=[1114],p_{0}=\begin{bmatrix}-1\\ -1\\ -1\\ -1\end{bmatrix},\ p_{1}=\begin{bmatrix}4\\ -1\\ -1\\ -1\end{bmatrix},\ p_{2}=\begin{bmatrix}-1\\ 4\\ -1\\ -1\end{bmatrix},\ p_{3}=\begin{bmatrix}-1\\ -1\\ 4\\ -1\end{bmatrix},\ p_{4}=\begin{bmatrix}-1\\ -1\\ -1\\ 4\end{bmatrix},

which gives a 44-simplex. Take BB (as a topological space) to be the boundary of Ξ\Xi. There are five 33-dimensional maximal cells intersecting along ten 22-dimensional facets. The polyhedral decomposition 𝒫\mathscr{P} on each facet is as in Figure 3.

Refer to caption
Figure 3. The polyhedral decomposition on a facet

The affine structure can be extended to the complement of codimension 2 closed subset Γ\Gamma whose intersection with a triangle in Figure 3 is a YY-shaped locus. Locally near each of these triangles, it looks like Figure 4a.

Refer to caption
(a) YY-vertex of type I
Refer to caption
(b) YY-vertex of type II

Ξ\Xi has ten 11-dimensional faces, each of which is an edge with affine length 55. The polyhedral decomposition 𝒫\mathscr{P} divides each edge into 55 intervals as we can see in Figure 3. Locally near each of these length 11 intervals, there are three 22-cells of 𝒫\mathscr{P} intersecting along it. The locus Γ\Gamma on each 22-cell intersects on the interval as shown in Figure 4b.

Definition 2.5 ([27], Def. 1.43).

An integral affine function on an open subset UBU\subset B is a continuous function φ\varphi on UU which is integral affine on Uintre(σ)U\cap\mathrm{int}_{\mathrm{re}}(\sigma) for σ𝒫[n]\sigma\in\mathscr{P}^{[n]} and on UWvU\cap W_{v} for v𝒫[0]v\in\mathscr{P}^{[0]}. We denote by 𝒜𝑓𝑓B\mathcal{A}\mathit{ff}_{B} (or simply 𝒜𝑓𝑓\mathcal{A}\mathit{ff}) the sheaf of integral affine functions on BB.

A piecewise integral affine function (abbrev. as PA-function) on UU is a continuous function φ\varphi on UU which can be written as φ=ψ+Sτ(φ¯)\varphi=\psi+S_{\tau}^{*}(\bar{\varphi}) on UUτU\cap U_{\tau} for every τ𝒫\tau\in\mathscr{P}, where ψ𝒜𝑓𝑓(UUτ)\psi\in\mathcal{A}\mathit{ff}(U\cap U_{\tau}) and φ¯\bar{\varphi} is a piecewise linear function on 𝒬τ,\mathscr{Q}_{\tau,\real} with respect to the fan Στ\Sigma_{\tau}. The sheaf of PA-functions on BB is denoted by 𝒫𝒫\mathcal{PL}_{\mathscr{P}}.

There is a natural inclusion 𝒜𝑓𝑓𝒫𝒫\mathcal{A}\mathit{ff}\hookrightarrow\mathcal{PL}_{\mathscr{P}}, and we let 𝒫𝒫\mathcal{MPL}_{\mathscr{P}} be the quotient:

0𝒜𝑓𝑓𝒫𝒫𝒫𝒫0.0\to\mathcal{A}\mathit{ff}\to\mathcal{PL}_{\mathscr{P}}\to\mathcal{MPL}_{\mathscr{P}}\to 0.

Locally, an element φΓ(B,𝒫𝒫)\varphi\in\Gamma(B,\mathcal{MPL}_{\mathscr{P}}) is a collection of piecewise affine functions {φU}\{\varphi_{U}\} such that on each overlap UVU\cap V, the difference φU|VφV|U\varphi_{U}|_{V}-\varphi_{V}|_{U} is an integral affine function on UVU\cap V.

Definition 2.6 ([27], Def. 1.45 and 1.47).

The sheaf 𝒫𝒫\mathcal{MPL}_{\mathscr{P}} is called the sheaf of multi-valued piecewise affine functions (abbrev. as MPA-funtions) of the pair (B,𝒫)(B,\mathscr{P}). A section φH0(B,𝒫𝒫)\varphi\in H^{0}(B,\mathcal{MPL}_{\mathscr{P}}) is said to be convex (resp. strictly convex) if for any vertex {v}𝒫\{v\}\in\mathscr{P}, there is a convex (resp. strictly convex) representative φv\varphi_{v} on UvU_{v}. (Here, convexity (resp. strict convexity) means if we take any maximal cone σUv\sigma\subset U_{v} with the affine function lσ:Uvl_{\sigma}\colon U_{v}\rightarrow\real defined by requiring φv|σ=lσ\varphi_{v}|_{\sigma}=l_{\sigma}, we always have φv(y)lσ(y)\varphi_{v}(y)\geq l_{\sigma}(y) (resp. φv(y)>lσ(y)\varphi_{v}(y)>l_{\sigma}(y)) for yUvσy\in U_{v}\setminus\sigma).

The set of all convex multi-valued piecewise affine functions gives a sub-monoid of H0(B,𝒫𝒫)H^{0}(B,\mathcal{MPL}_{\mathscr{P}}) under addition, denoted as H0(B,𝒫𝒫,)H^{0}(B,\mathcal{MPL}_{\mathscr{P}},\mathbb{N}); we let QQ be the dual monoid.

Definition 2.7 ([27], Def. 1.48).

The polyhedral decomposition 𝒫\mathscr{P} is said to be regular if there exists a strictly convex multi-valued piecewise linear function φH0(B,𝒫𝒫)\varphi\in H^{0}(B,\mathcal{MPL}_{\mathscr{P}}).

We always assume that 𝒫\mathscr{P} is regular with a fixed strictly convex φH0(B,𝒫𝒫)\varphi\in H^{0}(B,\mathcal{MPL}_{\mathscr{P}}).

2.2. Monodromy, positivity and simplicity

To describe monodromy, we consider two maximal cells σ±\sigma_{\pm} and two of their common vertices v±v_{\pm}. Taking a path γ\gamma going from v+v_{+} to vv_{-} through σ+\sigma_{+}, and then from vv_{-} back to v+v_{+} through σ\sigma_{-}, we obtain a monodromy transformation TγT_{\gamma}. As in [27, §1.5], we are interested in two cases. The first case is when v+v_{+} is connected to vv_{-} via a bounded edge ω𝒫[1]\omega\in\mathscr{P}^{[1]}. Let dωΛωd_{\omega}\in\Lambda_{\omega} be the unique primitive vector pointing to vv_{-} along ω\omega. For an integral tangent vector mTv+:=Tv+,Bm\in T_{v_{+}}:=T_{v_{+},\mathbb{Z}}B, the monodromy transformation TγT_{\gamma} is given by

(2.1) Tγ(m)=m+m,nωσ+σdωT_{\gamma}(m)=m+\langle m,n^{\sigma_{+}\sigma_{-}}_{\omega}\rangle d_{\omega}

for some nωσ+σ𝒬σ+σTv+n^{\sigma_{+}\sigma_{-}}_{\omega}\in\mathscr{Q}_{\sigma_{+}\cap\sigma_{-}}^{*}\subset T_{v_{+}}^{*}, where ,\langle\cdot,\cdot\rangle is the natural pairing between Tv+T_{v_{+}} and Tv+T_{v_{+}}^{*}. The second case is when σ+\sigma_{+} and σ\sigma_{-} are separated by a codimension one cell ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]}. Let dˇρ𝒬ρ\check{d}_{\rho}\in\mathscr{Q}_{\rho}^{*} be the unique primitive covector which is positive on σ+\sigma_{+}. The monodromy transformation is given by

(2.2) Tγ(m)=m+m,dˇρmv+vρT_{\gamma}(m)=m+\langle m,\check{d}_{\rho}\rangle m^{\rho}_{v_{+}v_{-}}

for some mv+vρΛτm^{\rho}_{v_{+}v_{-}}\in\Lambda_{\tau}, where τρ\tau\subset\rho is the smallest face of ρ\rho containing v±v_{\pm}. In particular, if we fix both v±ωρσ±v_{\pm}\in\omega\subset\rho\subset\sigma_{\pm}, one obtains the formula

(2.3) Tγ(m)=m+κωρm,dˇρdωT_{\gamma}(m)=m+\kappa_{\omega\rho}\langle m,\check{d}_{\rho}\rangle d_{\omega}

for some integer κωρ\kappa_{\omega\rho}.

Definition 2.8 ([27], Def. 1.54).

We say that (B,𝒫)(B,\mathscr{P}) is positive if κωρ0\kappa_{\omega\rho}\geq 0 for all ω𝒫[1]\omega\in\mathscr{P}^{[1]} and ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]} with ωρ\omega\subset\rho.

Following [27, Definition 1.58], we package the monodromy data into polytopes associated to τ𝒫[k]\tau\in\mathscr{P}^{[k]} for 1kn11\leq k\leq n-1. The simplest case is when ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]}, whose monodromy polytope is defined by fixing a vertex v0ρv_{0}\in\rho and setting

(2.4) Δ(ρ):=Conv{mv0vρ|vρ,v𝒫[0]}Λρ,,\Delta(\rho):=\mathrm{Conv}\{m^{\rho}_{v_{0}v}\ |\ v\in\rho,\ v\in\mathscr{P}^{[0]}\}\subset\Lambda_{\rho,\real},

where Conv\mathrm{Conv} refers to taking the convex hull. It is well-defined up to translation and independent of the choice of v0v_{0}. The normal fan of ρ\rho in Λρ,\Lambda_{\rho,\real}^{*} is a refinement of the normal fan of Δ(ρ)\Delta(\rho). Similarly, when ω𝒫[1]\omega\in\mathscr{P}^{[1]}, one defines the dual monodromy polytope by fixing σ0ω\sigma_{0}\supset\omega and setting

(2.5) Δˇ(ω):=Conv{nωσ0σ|σω,σ𝒫[n1]}𝒬ω,.\check{\Delta}(\omega):=\mathrm{Conv}\{n^{\sigma_{0}\sigma}_{\omega}\ |\ \sigma\supset\omega,\ \sigma\in\mathscr{P}^{[n-1]}\}\subset\mathscr{Q}_{\omega,\real}^{*}.

Again, this is well-defined up to translation and independent of the choice of σ0\sigma_{0}. The fan Σω\Sigma_{\omega} in 𝒬ω,\mathscr{Q}_{\omega,\real} is a refinement of the normal fan of Δˇ(ω)\check{\Delta}(\omega). For 1<dim(τ)<n11<\dim(\tau)<n-1, a combination of monodromy and dual monodromy polytopes is needed. We let 𝒫1(τ)={ω|ω𝒫[1],ωτ}\mathscr{P}_{1}(\tau)=\{\omega\ |\ \omega\in\mathscr{P}^{[1]},\ \omega\subset\tau\} and 𝒫n1(τ)={ρ|ρ𝒫[n1],ρτ}\mathscr{P}_{n-1}(\tau)=\{\rho\ |\ \rho\in\mathscr{P}^{[n-1]},\ \rho\supset\tau\}. For each ρ𝒫n1(τ)\rho\in\mathscr{P}_{n-1}(\tau), we choose a vertex v0ρv_{0}\in\rho and let

Δρ(τ):=Conv{mv0vρ|vτ,v𝒫[0]}Λτ,.\Delta_{\rho}(\tau):=\mathrm{Conv}\{m^{\rho}_{v_{0}v}\ |\ v\in\tau,\ v\in\mathscr{P}^{[0]}\}\subset\Lambda_{\tau,\real}.

Similarly, for each ω𝒫1(τ)\omega\in\mathscr{P}_{1}(\tau), we choose σ0τ\sigma_{0}\supset\tau and let

Δˇω(τ):=Conv{nωσ0σ|στ,σ𝒫[n1]}𝒬τ,.\check{\Delta}_{\omega}(\tau):=\mathrm{Conv}\{n^{\sigma_{0}\sigma}_{\omega}\ |\ \sigma\supset\tau,\ \sigma\in\mathscr{P}^{[n-1]}\}\subset\mathscr{Q}_{\tau,\real}^{*}.

These are well-defined up to translation and independent of the choices of v0v_{0} and σ0\sigma_{0} respectively.

Definition 2.9 ([27], Def. 1.60).

We say (B,𝒫)(B,\mathscr{P}) is simple if, for every τ𝒫\tau\in\mathscr{P}, there are disjoint non-empty subsets

Ω1,,Ωp𝒫1(τ),R1,,Rp𝒫n1(τ)\Omega_{1},\dots,\Omega_{p}\subset\mathscr{P}_{1}(\tau),\quad R_{1},\dots,R_{p}\subset\mathscr{P}_{n-1}(\tau)

(where pp depends on τ\tau) such that

  1. (1)

    for ω𝒫1(τ)\omega\in\mathscr{P}_{1}(\tau) and ρ𝒫n1(τ)\rho\in\mathscr{P}_{n-1}(\tau), κωρ0\kappa_{\omega\rho}\neq 0 if and only if ωΩi\omega\in\Omega_{i} and ρRi\rho\in R_{i} for some 1ip1\leq i\leq p;

  2. (2)

    Δρ(τ)\Delta_{\rho}(\tau) is independent (up to translation) of ρRi\rho\in R_{i} and will be denoted by Δi(τ)\Delta_{i}(\tau); similarly, Δˇω(τ)\check{\Delta}_{\omega}(\tau) is independent (up to translation) of ωΩi\omega\in\Omega_{i} and will be denoted by Δˇi(τ)\check{\Delta}_{i}(\tau);

  3. (3)

    if {e1,,ep}\{e_{1},\dots,e_{p}\} is the standard basis in p\mathbb{Z}^{p}, then

    Δ(τ):=Conv{i=1pΔi(τ)×{ei}},Δˇ(τ):=Conv{i=1pΔˇi(τ)×{ei}}\Delta(\tau):=\mathrm{Conv}\left\{\bigcup_{i=1}^{p}\Delta_{i}(\tau)\times\{e_{i}\}\right\},\quad\check{\Delta}(\tau):=\mathrm{Conv}\left\{\bigcup_{i=1}^{p}\check{\Delta}_{i}(\tau)\times\{e_{i}\}\right\}

    are elementary simplices (i.e. a simplex whose only integral points are its vertices) in (Λτp)\left(\Lambda_{\tau}\oplus\mathbb{Z}^{p}\right) and (𝒬τp)\left(\mathscr{Q}_{\tau}^{*}\oplus\mathbb{Z}^{p}\right) respectively.

We need the following stronger condition in order to apply [28, Thm. 3.21] in a later stage:

Definition 2.10.

We say (B,𝒫)(B,\mathscr{P}) is strongly simple if it is simple, and for every τ𝒫\tau\in\mathscr{P}, both Δ(τ)\Delta(\tau) and Δˇ(τ)\check{\Delta}(\tau) are standard simplices.

Example 2.11.

Consider the 22-dimensional example in Example 2.3. Following [1, Ex. 6.82(1)], we may choose the two adjacent vertices in Figure 1 to be v1=[111]Tv_{1}=\begin{bmatrix}-1&-1&-1\end{bmatrix}^{T} and v2=[011]Tv_{2}=\begin{bmatrix}0&-1&-1\end{bmatrix}^{T} which bound a 11-cell ρ\rho. The two adjacent maximal cells are given by σ+{b|w+,b=1}\sigma_{+}\subset\{b\ |\ \langle w_{+},b\rangle=1\} where w+=[001]Tw_{+}=\begin{bmatrix}0&0&-1\end{bmatrix}^{T} and σ{b|w,b=1}\sigma_{-}\subset\{b\ |\ \langle w_{-},b\rangle=1\} where w=[010]Tw_{-}=\begin{bmatrix}0&-1&0\end{bmatrix}^{T}. The tangent lattice Tv1T_{v_{1}} can be identified with 3/v1\mathbb{Z}^{3}/\mathbb{Z}\cdot v_{1} equipped with the basis e1=[100]Te_{1}=\begin{bmatrix}1&0&0\end{bmatrix}^{T}, e2=[010]Te_{2}=\begin{bmatrix}0&1&0\end{bmatrix}^{T}. If we let γ\gamma be a loop going from v1v_{1} to v2v_{2} through σ+\sigma_{+} and going back to v1v_{1} through σ\sigma_{-}, we have

Tγ(m)=m+[011]T,me1T_{\gamma}(m)=m+\langle\begin{bmatrix}0&1&-1\end{bmatrix}^{T},m\rangle e_{1}

for mTv1m\in T_{v_{1}}. Therefore, we have p=1p=1, Δ1(ρ)=Conv{0,e1}\Delta_{1}(\rho)=\mathrm{Conv}\{0,e_{1}\} and Δˇ1(ρ)=Conv{0,w+w}\check{\Delta}_{1}(\rho)=\mathrm{Conv}\{0,w_{+}-w_{-}\}. This is an example of a positive and strongly simple (B,𝒫)(B,\mathscr{P}) (Definitions 2.8 and 2.10).

Example 2.12.

Next we consider the two types of YY-vertex in Example 2.4.

We begin with YY-vertex of type II in Figure 4a. Following [1, Ex. 6.82(2)], the three vertices v1,v2,v3v_{1},v_{2},v_{3} can be chosen to be

v1=[1111]T,v2=[0111]T,v3=[1011]T,v_{1}=\begin{bmatrix}-1&-1&-1&-1\end{bmatrix}^{T},\ v_{2}=\begin{bmatrix}0&-1&-1&-1\end{bmatrix}^{T},\ v_{3}=\begin{bmatrix}-1&0&-1&-1\end{bmatrix}^{T},

and σ+{b|4w+,b=1}\sigma_{+}\subset\{b\in{}^{4}\ |\ \langle w_{+},b\rangle=1\}, σ{b|4w,b=1}\sigma_{-}\subset\{b\in{}^{4}\ |\ \langle w_{-},b\rangle=1\} are 33-cells of BB lying in the affine hyperplanes with dual vector w+=[0010]Tw_{+}=\begin{bmatrix}0&0&-1&0\end{bmatrix}^{T} and w=[0001]Tw_{-}=\begin{bmatrix}0&0&0&-1\end{bmatrix}^{T} respectively. If we identify TvT_{v} with Λσ+\Lambda_{\sigma_{+}} via parallel transport and choose the basis of Λσ+\Lambda_{\sigma_{+}} as

e1=[1000]T,e2=[0100]T,e3=[0001]T,e_{1}=\begin{bmatrix}1&0&0&0\end{bmatrix}^{T},\ e_{2}=\begin{bmatrix}0&-1&0&0\end{bmatrix}^{T},\ e_{3}=\begin{bmatrix}0&0&0&1\end{bmatrix}^{T},

then the monodromy transformations are given by

Tγ1=[101010001],Tγ2=[101011001],Tγ3=[100011001],T_{\gamma_{1}}=\begin{bmatrix}1&0&1\\ 0&1&0\\ 0&0&1\end{bmatrix},\ T_{\gamma_{2}}=\begin{bmatrix}1&0&-1\\ 0&1&-1\\ 0&0&1\end{bmatrix},\ T_{\gamma_{3}}=\begin{bmatrix}1&0&0\\ 0&1&1\\ 0&0&1\end{bmatrix},

where γi\gamma_{i} is the loop going from viv_{i} to vi+1v_{i+1} through σ+\sigma_{+} and going back to viv_{i} through σ\sigma_{-}, with indices of viv_{i}’s taken modulo 33. In this case, we have p=1p=1, Δ1(ρ)=Conv{0,e1,e2}\Delta_{1}(\rho)=\mathrm{Conv}\{0,e_{1},-e_{2}\} is a 22-simplex and Δˇ1(ρ)=Conv{0,w+w}\check{\Delta}_{1}(\rho)=\mathrm{Conv}\{0,w_{+}-w_{-}\} is a 11-simplex.

For the YY-vertex of type II in Figure 4b, we can choose

v1=[1111]T,v2=[0111]T,v_{1}=\begin{bmatrix}-1&-1&-1&-1\end{bmatrix}^{T},\ v_{2}=\begin{bmatrix}0&-1&-1&-1\end{bmatrix}^{T},

which are the end-points of a 11-cell τ\tau. We choose the three maximal cells σ1\sigma_{1}, σ2\sigma_{2} and σ3\sigma_{3} intersecting at τ\tau to be the 33-cells lying in affine hyperplanes defined by {b|wi,b=1}\{b\ |\ \langle w_{i},b\rangle=1\}, where

w1=[0010]T,w2=[0001]T,w3=[0100]T.w_{1}=\begin{bmatrix}0&0&-1&0\end{bmatrix}^{T},\ w_{2}=\begin{bmatrix}0&0&0&-1\end{bmatrix}^{T},\ w_{3}=\begin{bmatrix}0&-1&0&0\end{bmatrix}^{T}.

Let γ~i\tilde{\gamma}_{i} be the loop going from v1v_{1} to v2v_{2} through wiw_{i} and then going back to v1v_{1} through wi+1w_{i+1}, with indices taken to be modulo 33. Then the corresponding monodromy transformations are given by

Tγ1=[101010001],Tγ2=[110010001],Tγ3=[111010001],T_{\gamma_{1}}=\begin{bmatrix}1&0&1\\ 0&1&0\\ 0&0&1\end{bmatrix},\ T_{\gamma_{2}}=\begin{bmatrix}1&1&0\\ 0&1&0\\ 0&0&1\end{bmatrix},\ T_{\gamma_{3}}=\begin{bmatrix}1&-1&-1\\ 0&1&0\\ 0&0&1\end{bmatrix},

with respect to the basis

e1=[1000]T,e2=[0100]T,e3=[0010]T.e_{1}=\begin{bmatrix}1&0&0&0\end{bmatrix}^{T},\ e_{2}=\begin{bmatrix}0&1&0&0\end{bmatrix}^{T},\ e_{3}=\begin{bmatrix}0&0&-1&0\end{bmatrix}^{T}.

In this case, p=1p=1, Δ1(τ)=Conv{0,v2v1}\Delta_{1}(\tau)=\mathrm{Conv}\{0,v_{2}-v_{1}\} is a 11-simplex and Δˇ1(τ)=Conv{0,w1w2,w1w3}\check{\Delta}_{1}(\tau)=\mathrm{Conv}\{0,w_{1}-w_{2},w_{1}-w_{3}\} is a 22-simplex.

Both examples are positive and strongly simple.

Throughout this paper, we always assume that (B,𝒫)(B,\mathscr{P}) is positive and strongly simple. In particular, both Δi(τ)\Delta_{i}(\tau) and Δˇi(τ)\check{\Delta}_{i}(\tau) are standard simplices of positive dimensions, and ΛΔ1(τ)ΛΔp(τ)\Lambda_{\Delta_{1}(\tau)}\oplus\cdots\oplus\Lambda_{\Delta_{p}(\tau)} (resp. ΛΔˇ1(τ)ΛΔˇp(τ)\Lambda_{\check{\Delta}_{1}(\tau)}\oplus\cdots\oplus\Lambda_{\check{\Delta}_{p}(\tau)}) is an internal direct summand of Λτ\Lambda_{\tau} (resp. 𝒬τ\mathscr{Q}_{\tau}^{*}).

2.3. Cone construction by gluing open affine charts

In this subsection, we recall the cone construction of the maximally degenerate Calabi–Yau X0=X0(B,𝒫,s)\prescript{0}{}{X}=\prescript{0}{}{X}(B,\mathscr{P},s), following [27] and [29, §1.2]. For this purpose, we take 𝐊=\mathbf{K}=\mathbb{Z} and QQ to be the positive real axis in Notation 1.4. Throughout this paper, we will work in the category of analytic schemes.

We will construct X0\prescript{0}{}{X} as a gluing of affine analytic schemes V(v)V(v) parametrized by the vertices of 𝒫\mathscr{P}. For each vertex vv, we consider the fan Σv\Sigma_{v} and take the analytic affine toric variety

V(v):=Specan([Σv]),V(v):=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{v}]),

where Specan\mathrm{Spec}_{\mathrm{an}} means analytification of the algebraic affine scheme given by Spec\mathrm{Spec}. Here, the monoid structure for a general fan ΣM\Sigma\subset M is given by

p+q={p+q if p,qM are in a common cone of Σ, otherwise,p+q=\begin{cases}p+q&\text{ if }p,q\in M\text{ are in a common cone of }\Sigma,\\ \infty&\text{ otherwise},\end{cases}

and we set z=0z^{\infty}=0 in taking Spec([Σ])\mathrm{Spec}(\mathbb{C}[\Sigma]) (by abuse of notation, we use Σ\Sigma to stand for both the fan and the monoid associated to a fan if there is no confusion); in other words, the ring [Σ]\mathbb{C}[\Sigma] is defined explicitly as

[Σ]:=p|Σ|Mzp,zpzq={zp+q if p,qM are in a common cone of Σ,0 otherwise,\mathbb{C}[\Sigma]:=\bigoplus_{p\in|\Sigma|\cap M}\mathbb{C}\cdot z^{p},\quad z^{p}\cdot z^{q}=\begin{cases}z^{p+q}&\text{ if }p,q\in M\text{ are in a common cone of }\Sigma,\\ 0&\text{ otherwise},\end{cases}

where |Σ||\Sigma| denotes the support of the fan Σ\Sigma.

To glue these affine analytic schemes together, we need affine subschemes {V(τ)}\{V(\tau)\} associated to τ𝒫\tau\in\mathscr{P} with vτv\in\tau and natural open embeddings V(τ)V(ω)V(\tau)\hookrightarrow V(\omega) for vωτv\in\omega\subset\tau. First, for τ𝒫\tau\in\mathscr{P} such that vτv\in\tau, we consider the localization of Σv\Sigma_{v} at τ\tau defined by

τ1Σv:={Kvσ+Λτ,|Kvσ is a cone in Σv such that στ};\tau^{-1}\Sigma_{v}:=\{K_{v}\sigma+\Lambda_{\tau,\mathbb{R}}\,|\,K_{v}\sigma\text{ is a cone in }\Sigma_{v}\text{ such that }\sigma\supset\tau\};

here recall that Kvσ=Sv0(σUv)K_{v}\sigma={}_{\geq 0}S_{v}(\sigma\cap U_{v}) is the cone in Σv\Sigma_{v} (see the definition of a fan structure before Definition 2.2). This defines a new complete fan in Tv,T_{v,\real} consisting of convex, but not necessarily strictly convex, cones. If τ\tau contains another vertex vv^{\prime}, we can identify the fans τ1Σv\tau^{-1}\Sigma_{v} and τ1Σv\tau^{-1}\Sigma_{v^{\prime}} as follows: for each maximal στ\sigma\supset\tau, we identify the maximal cones Kvσ+Λτ,K_{v}\sigma+\Lambda_{\tau,\mathbb{R}} and Kvσ+Λτ,K_{v^{\prime}}\sigma+\Lambda_{\tau,\mathbb{R}} by identifying the tangent spaces TvTvT_{v}\cong T_{v^{\prime}} using parallel transport through στ\sigma\supset\tau. Patching these identifications for all στ\sigma\supset\tau together, we get a piecewise linear transformation from TvT_{v} to TvT_{v^{\prime}}, identifying the fans τ1Σv\tau^{-1}\Sigma_{v} and τ1Σv\tau^{-1}\Sigma_{v^{\prime}} and hence the corresponding monoids. This defines the affine analytic scheme

V(τ):=Specan([τ1Σv]),V(\tau):=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{v}]),

up to a unique isomorphism. Notice that τ1Σv\tau^{-1}\Sigma_{v} can be identified (non-canonically) with the fan Στ×Λτ,\Sigma_{\tau}\times\Lambda_{\tau,\real} in 𝒬τ,×Λτ,\mathscr{Q}_{\tau,\real}\times\Lambda_{\tau,\real}, so actually

V(τ)Specan([Λτ])×Specan([Στ]),V(\tau)\cong\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Lambda_{\tau}])\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}]),

where Specan([Λτ])Λτ()l\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Lambda_{\tau}])\cong\Lambda_{\tau}^{*}\otimes_{\mathbb{Z}}\mathbb{C}^{*}\cong(\mathbb{C}^{*})^{l} is a complex torus.

For any vωτv\in\omega\subset\tau, there is a map of monoids ω1Σvτ1Σv\omega^{-1}\Sigma_{v}\to\tau^{-1}\Sigma_{v} given by

p{p if pKvσ+Λω, for some στ, otherwisep\mapsto\begin{cases}p&\text{ if }p\in K_{v}\sigma+\Lambda_{\omega,\mathbb{R}}\text{ for some }\sigma\supset\tau,\\ \infty&\text{ otherwise}\end{cases}

(though there is no fan map from ω1Σv\omega^{-1}\Sigma_{v} to τ1Σv\tau^{-1}\Sigma_{v} in general), and hence a ring map

ιωτ:[ω1Σv][τ1Σv].\iota_{\omega\tau}^{*}\colon\mathbb{C}[\omega^{-1}\Sigma_{v}]\to\mathbb{C}[\tau^{-1}\Sigma_{v}].

This gives an open inclusion of affine schemes

ιωτ:V(τ)V(ω),\iota_{\omega\tau}\colon V(\tau)\hookrightarrow V(\omega),

and hence a functor F:𝒫𝐒𝐜𝐡anF\colon\mathscr{P}\to{\bf{Sch}}_{\mathrm{an}} defined by

F(τ):=V(τ),F(e):=ιωτ:V(τ)V(ω)F(\tau):=V(\tau),\quad F(e):=\iota_{\omega\tau}\colon V(\tau)\to V(\omega)

for ωτ\omega\subset\tau.

We can further introduce twistings of the gluing of the affine analytic schemes {V(τ)}τ𝒫\{V(\tau)\}_{\tau\in\mathscr{P}}. Toric automorphisms μ\mu of V(τ)V(\tau) are in bijection with the set of \mathbb{C}^{*}-valued piecewise multiplicative maps on Tv|τ1Σv|T_{v}\cap|\tau^{-1}\Sigma_{v}| with respect to the fan τ1Σv\tau^{-1}\Sigma_{v}. Explicitly, for each maximal cone σ𝒫[n]\sigma\in\mathscr{P}^{[n]} with τσ\tau\subset\sigma, there is a monoid homomorphism μσ:Λσ\mu_{\sigma}\colon\Lambda_{\sigma}\to\mathbb{C}^{*} such that if σ𝒫[n]\sigma^{\prime}\in\mathscr{P}^{[n]} also contains τ\tau, then μσ|Λσσ=μσ|Λσσ\mu_{\sigma}|_{\Lambda_{\sigma\cap\sigma^{\prime}}}=\mu_{\sigma^{\prime}}|_{\Lambda_{\sigma\cap\sigma^{\prime}}}. Denote by PM(τ)\mathrm{PM}(\tau) the multiplicative group of \mathbb{C}^{*}-valued piecewise multiplicative maps on Tv|τ1Σv|T_{v}\cap|\tau^{-1}\Sigma_{v}|. The group PM(τ)\mathrm{PM}(\tau) a priori depends on the choice of vτv\in\tau; however, for different choices, say vv and vv^{\prime}, the groups can be identified via the identification τ1Σvτ1Σv\tau^{-1}\Sigma_{v}\cong\tau^{-1}\Sigma_{v^{\prime}}. For ωτ\omega\subset\tau, there is a natural restriction map |τ:PM(ω)PM(τ)|_{\tau}\colon\mathrm{PM}(\omega)\rightarrow\mathrm{PM}(\tau) given by restricting to those maximal cells σω\sigma\supset\omega with στ\sigma\supset\tau.

Definition 2.13 ([29], Def. 1.18).

A choice of open gluing data (for the cone construction) for (B,𝒫)(B,\mathscr{P}) is a set s=(sωτ)ωτs=(s_{\omega\tau})_{\omega\subset\tau} of elements sωτPM(τ)s_{\omega\tau}\in\mathrm{PM}(\tau) such that

  1. (1)

    sττ=1s_{\tau\tau}=1 for all τ𝒫\tau\in\mathscr{P}, and

  2. (2)

    if ωτρ\omega\subset\tau\subset\rho, then

    sωρ=sτρsωτ|ρ.s_{\omega\rho}=s_{\tau\rho}\cdot s_{\omega\tau}|_{\rho}.

Two choices of open gluing data s,ss,s^{\prime} are said to be cohomologous if there exists a system {tτ}τ𝒫\{t_{\tau}\}_{\tau\in\mathscr{P}}, with tτPM(τ)t_{\tau}\in\mathrm{PM}(\tau) for each τ𝒫\tau\in\mathscr{P}, such that sωτ=tτ(tω|τ)1sωτs_{\omega\tau}=t_{\tau}(t_{\omega}|_{\tau})^{-1}s_{\omega\tau}^{\prime} whenever ωτ\omega\subset\tau.

The set of cohomology classes of choices of open gluing data is a group under multiplication, denoted as H1(𝒫,𝒬𝒫×)H^{1}(\mathscr{P},\mathscr{Q}_{\mathscr{P}}\otimes\mathbb{C}^{\times}). For sPM(τ)s\in\mathrm{PM}(\tau), we will denote also by ss the corresponding toric automorphism on V(τ)V(\tau) which is explicitly given by s(zm)=sσ(m)zms^{*}(z^{m})=s_{\sigma}(m)z^{m} for mστm\in\sigma\supset\tau. If ss is a choice of open gluing data, then we can define an ss-twisted functor Fs:𝒫𝐒𝐜𝐡anF_{s}\colon\mathscr{P}\to{\bf{Sch}}_{\mathrm{an}} by setting Fs(τ):=F(τ)=V(τ)F_{s}(\tau):=F(\tau)=V(\tau) on objects and Fs(ωτ):=F(ωτ)sωτ1:V(τ)V(ω)F_{s}(\omega\subset\tau):=F(\omega\subset\tau)\circ s_{\omega\tau}^{-1}\colon V(\tau)\to V(\omega) on morphisms. This defines the analytic scheme

X0=X0(B,𝒫,s):=limFs.\prescript{0}{}{X}=\prescript{0}{}{X}(B,\mathscr{P},s):=\lim_{\longrightarrow}F_{s}.

Gross–Siebert [27] showed that X0(B,𝒫,s)X0(B,𝒫,s)\prescript{0}{}{X}(B,\mathscr{P},s)\cong\prescript{0}{}{X}(B,\mathscr{P},s^{\prime}) as schemes when s,ss,s^{\prime} are cohomologous.

Remark 2.14.

Given τ𝒫[k]\tau\in\mathscr{P}^{[k]}, one can define a closed stratum ιτ:Xτ0X0\iota_{\tau}\colon\prescript{0}{}{X}_{\tau}\rightarrow\prescript{0}{}{X} of dimension kk by gluing together the kk-dimensional toric strata Vτ(ω)V(ω)=Specan([ω1Σv])V_{\tau}(\omega)\subset V(\omega)=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\omega^{-1}\Sigma_{v}]) corresponding to the cones Kvτ+Λω,K_{v}\tau+\Lambda_{\omega,\mathbb{R}} in ω1Σv\omega^{-1}\Sigma_{v}, for all ωτ\omega\subset\tau. Abstractly, it is isomorphic to the toric variety associated to the polyhedron τΛτ,\tau\subset\Lambda_{\tau,\real}. Also, for every pair ωτ\omega\subset\tau, there is a natural inclusion ιωτ:Xω0Xτ0\iota_{\omega\tau}\colon\prescript{0}{}{X}_{\omega}\rightarrow\prescript{0}{}{X}_{\tau}. One can alternatively construct X0\prescript{0}{}{X} by gluing along the closed strata Xτ0\prescript{0}{}{X}_{\tau}’s according to the polyhedral decomposition; see [27, §2.2].

We recall the following definition from [27], which serves as an alternative set of combinatorial data for encoding μPM(τ)\mu\in\mathrm{PM}(\tau).

Definition 2.15 ([27], Def. 3.25 and [29], Def. 1.20).

Let μPM(τ)\mu\in\mathrm{PM}(\tau) and ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]} with τρ\tau\subset\rho. For a vertex vτv\in\tau, we define

D(μ,ρ,v):=μσ(m)μσ(m)×,D(\mu,\rho,v):=\frac{\mu_{\sigma}(m)}{\mu_{\sigma^{\prime}}(m^{\prime})}\in\mathbb{C}^{\times},

where σ,σ\sigma,\sigma^{\prime} are the two unique maximal cells such that σσ=ρ\sigma\cap\sigma^{\prime}=\rho, mΛσm\in\Lambda_{\sigma} is an element projecting to the generator in 𝒬ρΛσ/Λρ\mathscr{Q}_{\rho}\cong\Lambda_{\sigma}/\Lambda_{\rho}\cong\mathbb{Z} pointing to σ\sigma^{\prime}, and mm^{\prime} is the parallel transport of mΛσm\in\Lambda_{\sigma} to Λσ\Lambda_{\sigma^{\prime}} through vv. D(μ,ρ,v)D(\mu,\rho,v) is independent of the choice of mm.

Let ρ𝒫[d1]\rho\in\mathscr{P}^{[d-1]} and σ+,σ\sigma_{+},\sigma_{-} be the two unique maximal cells such that σ+σ=ρ\sigma_{+}\cap\sigma_{-}=\rho. Let dˇρ𝒬ρ\check{d}_{\rho}\in\mathscr{Q}_{\rho}^{*} be the unique primitive generator pointing to σ+\sigma_{+}. For any two vertices v,vτv,v^{\prime}\in\tau, we have the formula

(2.6) D(μ,ρ,v)=μ(mvvρ)1D(μ,ρ,v)D(\mu,\rho,v)=\mu(m_{vv^{\prime}}^{\rho})^{-1}\cdot D(\mu,\rho,v^{\prime})

relating monodromy data to the open gluing data, where mvvρΛρm_{vv^{\prime}}^{\rho}\in\Lambda_{\rho} is as discussed in (2.2). The formula (2.6) describes the interaction between monodromy and a fixed μPM(τ)\mu\in\mathrm{PM}(\tau). We shall further impose the following lifting condition from [27, Prop. 4.25] relating svτ,svτPM(τ)s_{v\tau},s_{v^{\prime}\tau}\in\mathrm{PM}(\tau) and monodromy data:

Condition 2.16.

We say a choice of open gluing data ss satisfies the lifting condition if for any two vertices v,vτρv,v^{\prime}\in\tau\subset\rho with ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]}, we have D(svτ,ρ,v)=D(svτ,ρ,v)D(s_{v\tau},\rho,v)=D(s_{v^{\prime}\tau},\rho,v^{\prime}) whenever mvvρ=0m^{\rho}_{vv^{\prime}}=0.

2.4. Log structures

We need to equip the analytic scheme X0=X0(B,𝒫,s)\prescript{0}{}{X}=\prescript{0}{}{X}(B,\mathscr{P},s) with log structures. The main reference is [27, §3 - 5].

Definition 2.17.

Let XX be an analytic space, a log structure on XX is a sheaf of monoids X\mathcal{M}_{X} together with a homomorphism αX:X𝒪X\alpha_{X}\colon\mathcal{M}_{X}\rightarrow\mathcal{O}_{X} of sheaves of (multiplicative) monoids such that αX:α1(𝒪X)𝒪X\alpha_{X}\colon\alpha^{-1}(\mathcal{O}_{X}^{*})\rightarrow\mathcal{O}_{X}^{*} is an isomorphism. The ghost sheaf ¯X\overline{\mathcal{M}}_{X} of a log structure is defined as the quotient sheaf X/α1(𝒪X)\mathcal{M}_{X}/\alpha^{-1}(\mathcal{O}_{X}^{*}), whose monoid structure is written additively.

Example 2.18.

Let XX be an analytic space and DXD\subset X be a closed analytic subspace of pure codimension one. We denote by j:XDXj\colon X\setminus D\hookrightarrow X the inclusion. Then the sheaf of monoids

X:=j(𝒪XD)𝒪X,\mathcal{M}_{X}:=j_{*}(\mathcal{O}_{X\setminus D}^{*})\cap\mathcal{O}_{X},

together with the natural inclusion αX:X𝒪X\alpha_{X}\colon\mathcal{M}_{X}\rightarrow\mathcal{O}_{X} defines a log structure on XX.

We write XX^{\dagger} if we want to emphasize the log structure on XX. A general way to define a log structure is to take an arbitary homomorphism of sheaves of monoids

α~:𝒫𝒪X,\tilde{\alpha}\colon\mathcal{P}\rightarrow\mathcal{O}_{X},

and then define the associated log structure by

X:=(𝒫𝒪X)/{(p,α~(p)1)|pα~1(𝒪X)}.\mathcal{M}_{X}:=(\mathcal{P}\oplus\mathcal{O}_{X}^{*})/\{(p,\tilde{\alpha}(p)^{-1})\ |\ p\in\tilde{\alpha}^{-1}(\mathcal{O}_{X}^{*})\}.

In particular, this allows us to define log structures on an analytic space YY by pulling back those on another analytic space XX via a morphism f:YXf\colon Y\rightarrow X. More precisely, given a log structure on XX, the pullback log structure on YY is defined to be the log structure associated to the composition α~Y:f1(X)f1(𝒪X)𝒪Y\tilde{\alpha}_{Y}\colon f^{-1}(\mathcal{M}_{X})\rightarrow f^{-1}(\mathcal{O}_{X})\rightarrow\mathcal{O}_{Y}. For more details of the theory of log structures, readers are referred to, e.g., [27, §3].

Example 2.19.

Taking a toric monoid PP (i.e. P=CMP=C\cap M for a cone CMC\subset M), we can define α~:P¯𝒪Spec([P])\tilde{\alpha}\colon\underline{P}\rightarrow\mathcal{O}_{\operatorname{Spec}(\mathbb{C}[P])} by sending mzmm\mapsto z^{m}, where P¯\underline{P} is the constant sheaf with stalk PP. From this we obtain a log structure on the analytic toric variety Specan([P])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P]). Note that this is a special case of Example 2.18, where we take X=Specan([P])X=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P]) and DD to be the toric boundary divisor.

Before we describe the log structures on X0=X0(B,𝒫,s)\prescript{0}{}{X}=\prescript{0}{}{X}(B,\mathscr{P},s), let us first specify a ghost sheaf ¯\overline{\mathcal{M}} over X0\prescript{0}{}{X}. Recall that the polyhedral decomposition 𝒫\mathscr{P} is assumed to be regular, namely, there exists a strictly convex multi-valued piecewise linear function φH0(B,𝒫𝒫)\varphi\in H^{0}(B,\mathcal{MPL}_{\mathscr{P}}). For any τ𝒫\tau\in\mathscr{P}, we take a strictly convex representative φ¯τ\bar{\varphi}_{\tau} of φ\varphi on 𝒬τ,\mathscr{Q}_{\tau,\real}, and define

Γ(V(τ),¯):=P¯τ=Cτ(𝒬τ),\Gamma(V(\tau),\overline{\mathcal{M}}):=\bar{P}_{\tau}=C_{\tau}\cap(\mathscr{Q}_{\tau}\oplus\mathbb{Z}),

where Cτ:={(m,h)𝒬τ,|hφ¯τ(m)}C_{\tau}:=\{(m,h)\in\mathscr{Q}_{\tau,\real}\oplus\mathbb{R}\,|\,h\geq\bar{\varphi}_{\tau}(m)\}. For any ωτ\omega\subset\tau, we take an integral affine function ψωτ\psi_{\omega\tau} on UωU_{\omega} such that ψωτ+Sω(φ¯ω)\psi_{\omega\tau}+S_{\omega}^{*}(\bar{\varphi}_{\omega}) vanishes on KωτK_{\omega}\tau, and agrees with Sτ(φ¯τ)S_{\tau}^{*}(\bar{\varphi}_{\tau}) on all of σUτ\sigma\cap U_{\tau} for any στ\sigma\supset\tau. This induces a map CωCωτ:={(m,h)𝒬ω,|hψωτ(m)+φ¯ω(m)}C_{\omega}\rightarrow C_{\omega\tau}:=\{(m,h)\in\mathscr{Q}_{\omega,\real}\oplus\mathbb{R}\,|\,h\geq\psi_{\omega\tau}(m)+\bar{\varphi}_{\omega}(m)\} by sending (m,h)(m,h+ψωτ(m))(m,h)\mapsto(m,h+\psi_{\omega\tau}(m)), whose composition with the quotient map 𝒬ω,𝒬τ,\mathscr{Q}_{\omega,\real}\oplus\real\rightarrow\mathscr{Q}_{\tau,\real}\oplus\real gives a map CωCτC_{\omega}\rightarrow C_{\tau} of cones that corresponds to the monoid homomorphism P¯ωP¯τ\bar{P}_{\omega}\rightarrow\bar{P}_{\tau}. The P¯τ\bar{P}_{\tau}’s glue together to give the ghost sheaf ¯\overline{\mathcal{M}} over X0\prescript{0}{}{X}. There is a well-defined section ϱ¯Γ(X0,¯)\bar{\varrho}\in\Gamma(\prescript{0}{}{X},\overline{\mathcal{M}}) given by gluing (0,1)Cτ(0,1)\in C_{\tau} for each τ\tau.

One may then hope to find a log structure on X0\prescript{0}{}{X} which is log smooth and with ghost sheaf given by ¯\overline{\mathcal{M}}. However, due to the presence of non-trivial monodromies of the affine structure, this can only be done away from a complex codimension 22 subset ZX0Z\subset\prescript{0}{}{X} not containing any toric strata. Such log structures can be described by sections of a coherent sheaf 𝒮pre+\mathcal{LS}^{+}_{\mathrm{pre}} supported on the scheme-theoretic singular locus Xsing0X0\prescript{0}{}{X}_{\mathrm{sing}}\subset\prescript{0}{}{X}. We now describe the sheaf 𝒮pre+\mathcal{LS}^{+}_{\mathrm{pre}} and some of its sections called slab functions; readers are referred to [27, §3 and 4] for more details.

For every ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]}, we consider ιρ:Xρ0X0\iota_{\rho}\colon\prescript{0}{}{X}_{\rho}\rightarrow\prescript{0}{}{X}, where Xρ0\prescript{0}{}{X}_{\rho} is the toric variety associated to the polytope ρΛρ,\rho\subset\Lambda_{\rho,\real}. From the fact that the normal fan 𝒩ρΛρ,\mathscr{N}_{\rho}\subset\Lambda_{\rho,\real}^{*} of ρ\rho is a refinement of the normal fan 𝒩Δ(ρ)Λρ,\mathscr{N}_{\Delta(\rho)}\subset\Lambda_{\rho,\real}^{*} of the rρr_{\rho}-dimensional simplex Δ(ρ)\Delta(\rho) (as in §2.2), we have a toric morphism

(2.7) ϰρ:Xρ0rρ.\varkappa_{\rho}\colon\prescript{0}{}{X}_{\rho}\rightarrow\mathbb{P}^{r_{\rho}}.

Now, Δ(ρ)\Delta(\rho) corresponds to 𝒪(1)\mathcal{O}(1) on rρ\mathbb{P}^{r_{\rho}}. We let 𝒩ρ:=ϰρ(𝒪(1))\mathcal{N}_{\rho}:=\varkappa_{\rho}^{*}(\mathcal{O}(1)) on Xρ0\prescript{0}{}{X}_{\rho}, and define

(2.8) 𝒮pre+:=ρ𝒫[n1]ιρ,(𝒩ρ).\mathcal{LS}^{+}_{\mathrm{pre}}:=\bigoplus_{\rho\in\mathscr{P}^{[n-1]}}\iota_{\rho,*}(\mathcal{N}_{\rho}).

Sections of 𝒮pre+\mathcal{LS}^{+}_{\mathrm{pre}} can be described explicitly. For each v𝒫[0]v\in\mathscr{P}^{[0]}, we consider the open subscheme V(v)V(v) of X0\prescript{0}{}{X} and the local trivialization

𝒮pre+|V(v)=ρ:vρ𝒪Vρ(v),\mathcal{LS}^{+}_{\mathrm{pre}}|_{V(v)}=\bigoplus_{\rho:v\in\rho}\mathcal{O}_{V_{\rho}(v)},

whose sections over V(v)V(v) are given by (fvρ)vρ(f_{v\rho})_{v\in\rho}. Given v,vτv,v^{\prime}\in\tau where τ\tau corresponding to V(τ)V(\tau), these local sections obey the change of coordinates given by

(2.9) D(svτ,ρ,v)1svτ1(fvρ)=zmvvρD(svτ,ρ,v)1svτ1(fvρ),D(s_{v^{\prime}\tau},\rho,v^{\prime})^{-1}s_{v^{\prime}\tau}^{-1}(f_{v^{\prime}\rho})=z^{-m^{\rho}_{vv^{\prime}}}D(s_{v\tau},\rho,v)^{-1}s_{v\tau}^{-1}(f_{v\rho}),

where ρτ\rho\supset\tau and svτ,svτs_{v\tau},s_{v^{\prime}\tau} are part of the open gluing data ss. The section f:=(fvρ)vρf:=(f_{v\rho})_{v\in\rho} is said to be normalized if fvρf_{v\rho} takes the value 11 at the 0-dimensional toric stratum corresponding to a vertex vv, for all ρ\rho. We will restrict ourselves to normalized sections ff of 𝒮pre+\mathcal{LS}^{+}_{\mathrm{pre}}. The complex codimension 22 subset ZX0Z\subset\prescript{0}{}{X} is taken to be the zero locus of ff on Xsing0\prescript{0}{}{X}_{\mathrm{sing}}.

Only a subset of normalized sections of 𝒮pre+\mathcal{LS}^{+}_{\mathrm{pre}} corresponds to log structures. For every vertex v𝒫[0]v\in\mathscr{P}^{[0]} and τ𝒫[n2]\tau\in\mathscr{P}^{[n-2]} containing vv, we choose a cyclic ordering ρ1,,ρl\rho_{1},\dots,\rho_{l} of codimension one cells containing τ\tau according to an orientation of 𝒬τ,\mathscr{Q}_{\tau,\real}. Let dˇρi𝒬v\check{d}_{\rho_{i}}\in\mathscr{Q}_{v}^{*} be the positively oriented normal to ρi\rho_{i}. Then the condition for f=(fvρ)vρ𝒮pre+|V(v)f=(f_{v\rho})_{v\in\rho}\in\mathcal{LS}^{+}_{\mathrm{pre}}|_{V(v)} to define a log structure is given by

(2.10) i=1ldˇρifvρi|Vτ(v)=01,in 𝒬vΓ(Vτ(v)Z,𝒪Vτ(v)),\prod_{i=1}^{l}\check{d}_{\rho_{i}}\otimes f_{v\rho_{i}}|_{V_{\tau}(v)}=0\otimes 1,\quad\text{in }\mathscr{Q}_{v}^{*}\otimes\Gamma(V_{\tau}(v)\setminus Z,\mathcal{O}_{V_{\tau}(v)}^{*}),

where the group structure on 𝒬v\mathscr{Q}_{v}^{*} is additive and that on Γ(Vτ(v)Z,𝒪Vτ(v))\Gamma(V_{\tau}(v)\setminus Z,\mathcal{O}_{V_{\tau}(v)}^{*}) is multiplicative. If f=(fvρ)vρf=(f_{v\rho})_{v\in\rho} is a normalized section satisfying this condition, we call the fvρf_{v\rho}’s slab functions.

Theorem 2.20 ([27], Thm. 5.2).

Suppose that BB is compact and the pair (B,𝒫)(B,\mathscr{P}) is simple and positive. Let ss be a choice of open gluing data satisfying the lifting condition (Condition 2.16). Then there exists a unique normalized section fΓ(X0,𝒮pre+)f\in\Gamma(\prescript{0}{}{X},\mathcal{LS}^{+}_{\mathrm{pre}}) which defines a log structure on X0\prescript{0}{}{X} (i.e. satisfying the condition (2.10)).

From now on, we always assume that BB is compact. To describe the log structure in Theorem 2.20, we first construct some local smoothing models: For each vertex v𝒫[0]v\in\mathscr{P}^{[0]}, we represent the strictly convex piecewise linear function φ\varphi in a small neighborhood UU of vv by a strictly convex piecewise linear φv:𝒬v,\varphi_{v}\colon\mathscr{Q}_{v,\real}\to\mathbb{R} (so that φ=Sv(φv)\varphi=S_{v}^{*}(\varphi_{v})) and set

Cv:={(m,h)𝒬v,|hφv(m)},Pv:=Cv(𝒬v).\displaystyle C_{v}:=\{(m,h)\in\mathscr{Q}_{v,\real}\oplus\mathbb{R}\,|\,h\geq\varphi_{v}(m)\},\quad P_{v}:=C_{v}\cap(\mathscr{Q}_{v}\oplus\mathbb{Z}).

The element ϱ=(0,1)𝒬v\varrho=(0,1)\in\mathscr{Q}_{v}\oplus\mathbb{Z} gives rise to a regular function q:=zϱq:=z^{\varrho} on Specan([Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{v}]). We have a natural identification

V(v):=Specan([Σv])Specan([Pv]/q),V(v):=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{v}])\cong\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{v}]/q),

through which we view V(v)V(v) as the toric boundary divisor in Specan([Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{v}]) that corresponds to the holomorphic function qq, and πv:Specan([Pv])Specan([q])\pi_{v}\colon\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{v}])\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[q]) as a local model for smoothing V(v)V(v).

Using these local models, we can now describe the log structure around a point xX0Zx\in\prescript{0}{}{X}\setminus Z. On a neighborhood VV(v)ZV\subset V(v)\setminus Z of xx, the local smoothing model is given by composing the two inclusions :VV(v)\prescript{}{}{\flat}\colon V\hookrightarrow V(v) and V(v)Specan([Pv])V(v)\hookrightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{v}]). The natural monoid homomorphism Pv[Pv]P_{v}\rightarrow\mathbb{C}[P_{v}] defined by sending mzmm\mapsto z^{m} determines a log structure on Specan([Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{v}]) which restricts to one on the toric boundary divisor V(v)=Specan([Σv])V(v)=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{v}]). We further twist the inclusion :VV(v)\prescript{}{}{\flat}\colon V\hookrightarrow V(v) as

(2.11) zmhmzm for mΣv;z^{m}\mapsto h_{m}\cdot z^{m}\text{ for $m\in\Sigma_{v}$;}

here, for each mΣvm\in\Sigma_{v}, hmh_{m} is chosen as an invertible holomorphic function on VZero(zm;v)V\cap\mathrm{Zero}(z^{m};v), where we denote Zero(zm;v):={xV(v)|zm𝒪x}¯\mathrm{Zero}(z^{m};v):=\overline{\{x\in V(v)\ |\ z^{m}\in\mathcal{O}_{x}^{*}\}}, and such that they satisfy the relations

(2.12) hmhm=hm+m,on VZero(zm+m;v).h_{m}\cdot h_{m^{\prime}}=h_{m+m^{\prime}},\quad\text{on }V\cap\mathrm{Zero}(z^{m+m^{\prime}};v).

Then pulling back the log structure on V(v)V(v) via :VV(v)\prescript{}{}{\flat}\colon V\hookrightarrow V(v) produces a log structure on VV which is log smooth.

These local choices of hmh_{m}’s are also required to be determined by the slab functions fvρf_{v\rho}’s, up to equivalences. Here, we shall just give the formula relating them; see [27, Thm. 3.22] for details. For any ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]} containing vv and two maximal cells σ±\sigma_{\pm} such that σ+σ=ρ\sigma_{+}\cap\sigma_{-}=\rho, we take m+𝒬vKvσ+m_{+}\in\mathscr{Q}_{v}\cap K_{v}\sigma_{+} generating 𝒬ρ\mathscr{Q}_{\rho} with some m0𝒬vKvρm_{0}\in\mathscr{Q}_{v}\cap K_{v}\rho such that m0m+𝒬vKvσm_{0}-m_{+}\in\mathscr{Q}_{v}\cap K_{v}\sigma_{-}. Then the required relation is given by

(2.13) fvρ=hm02hm0m+hm0+m+|Vρ(v)V𝒪Vρ(v)(Vρ(v)V),f_{v\rho}=\frac{h_{m_{0}}^{2}}{h_{m_{0}-m_{+}}\cdot h_{m_{0}+m_{+}}}\Big{|}_{V_{\rho}(v)\cap V}\in\mathcal{O}^{*}_{V_{\rho}(v)}(V_{\rho}(v)\cap V),

which is independent of the choices of m0m_{0} and m+m_{+}.

By abuse of notation, we also let :V𝕍k\prescript{}{}{\flat}\colon V\rightarrow\prescript{k}{}{\mathbb{V}} be the kk-th order thickening of VV over [q]/qk+1\mathbb{C}[q]/q^{k+1} in the model Specan([Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{v}]) under the above embedding. Then there is a natural divisorial log structure on 𝕍k\prescript{k}{}{\mathbb{V}}^{\dagger} over Sk\prescript{k}{}{S}^{\dagger} coming from restriction of the log structure on Specan([Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{v}])^{\dagger} over SS^{\dagger} (i.e. Example 2.18, which is the same as the one given by Example 2.19 in this case). Restricting to VV reproduces the log structure we constructed above, which is the log structure of X0\prescript{0}{}{X}^{\dagger} over the log point S0\prescript{0}{}{S}^{\dagger} locally around xx. We have a Cartesian diagram of log spaces

(2.14) V\textstyle{V^{\dagger}\ \ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝕍k\textstyle{\prescript{k}{}{\mathbb{V}}^{\dagger}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S0\textstyle{\prescript{0}{}{S}^{\dagger}\ \ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sk\textstyle{\prescript{k}{}{S}^{\dagger}}

Next we describe the log structure around a singular point xZ(Xτ0ωτXω0)x\in Z\cap\left(\prescript{0}{}{X}_{\tau}\setminus\bigcup_{\omega\subset\tau}\prescript{0}{}{X}_{\omega}\right) for some τ\tau. Viewing f=ρ𝒫[n1]fρf=\sum_{\rho\in\mathscr{P}^{[n-1]}}f_{\rho} where fρf_{\rho} is a section of 𝒩ρ\mathcal{N}_{\rho}, we let Zρ=Z(fρ)Xρ0X0Z_{\rho}=Z(f_{\rho})\subset\prescript{0}{}{X}_{\rho}\subset\prescript{0}{}{X} and write Z=ρZρZ=\bigcup_{\rho}Z_{\rho}. For every τ𝒫\tau\in\mathscr{P}, we have the data Ωi\Omega_{i}’s, RiR_{i}’s, Δi(τ)\Delta_{i}(\tau) and Δˇi(τ)\check{\Delta}_{i}(\tau) described in Definition 2.9 because (B,𝒫)(B,\mathscr{P}) is simple. Since the normal fan 𝒩τΛτ,\mathscr{N}_{\tau}\subset\Lambda_{\tau,\real}^{*} of τ\tau is a refinement of 𝒩Δi(τ)Λτ,\mathscr{N}_{\Delta_{i}(\tau)}\subset\Lambda_{\tau,\real}^{*}, we have a natural toric morphism

(2.15) ϰτ,i:Xτ0rτ,i,\varkappa_{\tau,i}\colon\prescript{0}{}{X}_{\tau}\rightarrow\mathbb{P}^{r_{\tau,i}},

and the identification ιτρ(𝒩ρ)ϰτ,i(𝒪(1))\iota_{\tau\rho}^{*}(\mathcal{N}_{\rho})\cong\varkappa_{\tau,i}^{*}(\mathcal{O}(1)). By the proof of [27, Thm. 5.2], ιτρ(fρ)\iota_{\tau\rho}^{*}(f_{\rho}) is completely determined by the gluing data ss and the associated monodromy polytope Δi(τ)\Delta_{i}(\tau) where ρRi\rho\in R_{i}. In particular, we have ιτρ(fρ)=ιτρ(fρ)\iota_{\tau\rho}^{*}(f_{\rho})=\iota_{\tau\rho^{\prime}}^{*}(f_{\rho^{\prime}}) and ZρXτ0=ZρXτ0=:ZiτZ_{\rho}\cap\prescript{0}{}{X}_{\tau}=Z_{\rho^{\prime}}\cap\prescript{0}{}{X}_{\tau}=:Z_{i}^{\tau} for ρ,ρRi\rho,\rho^{\prime}\in R_{i}. Locally, if we write V(τ)=Specan([τ1Σv])V(\tau)=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{v}]) by choosing some vτv\in\tau, then, for each 1ip1\leq i\leq p, there exists an analytic function fv,if_{v,i} on V(τ)V(\tau) such that fv,i|Vρ(τ)=svτ1(fvρ)f_{v,i}|_{V_{\rho}(\tau)}=s_{v\tau}^{-1}(f_{v\rho}) for ρRi\rho\in R_{i}.

According to [28, §2.1], for each 1ip1\leq i\leq p, we have Δˇi(τ)𝒬τ,\check{\Delta}_{i}(\tau)\subset\mathscr{Q}_{\tau,\real}^{*}, which gives

(2.16) ψi(m)=inf{m,n|nΔˇi(τ)}.\psi_{i}(m)=-\inf\{\langle m,n\rangle\ |\ n\in\check{\Delta}_{i}(\tau)\}.

By convention, we write ψ0:=φ¯τ\psi_{0}:=\bar{\varphi}_{\tau}. By rearranging the indices ii’s, we can assume that xZ1τZrτx\in Z^{\tau}_{1}\cap\cdots\cap Z^{\tau}_{r} and xZr+1τZpτx\notin Z^{\tau}_{r+1}\cup\cdots\cup Z^{\tau}_{p}. We introduce the convention that ψx,i=ψi\psi_{x,i}=\psi_{i} for 0ir0\leq i\leq r and ψx,i0\psi_{x,i}\equiv 0 for r<idim(τ)r<i\leq\dim(\tau). Then the local smoothing model near xx is constructed as Specan([Pτ,x])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{\tau,x}]), where

(2.17) Pτ,x:={(m,a0,,al)𝒬τ×l+1|aiψx,i(m)},P_{\tau,x}:=\{(m,a_{0},\dots,a_{l})\in\mathscr{Q}_{\tau}\times\mathbb{Z}^{l+1}\ |\ a_{i}\geq\psi_{x,i}(m)\},

l=dim(τ)l=\dim(\tau), and the distinguished element ϱ=(0,1,0,,0)\varrho=(0,1,0,\dots,0) defines a family

Specan([Pτ,x])Specan([q])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{\tau,x}])\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[q])

by sending qzϱq\mapsto z^{\varrho}. The central fiber is given by Specan([Qτ,x])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[Q_{\tau,x}]), where

(2.18) Qτ,x={(m,a0,,al)|a0=ψx,0(m)}Pτ,x/(ϱ+Pτ,x)Q_{\tau,x}=\{(m,a_{0},\dots,a_{l})\ |\ a_{0}=\psi_{x,0}(m)\}\cong P_{\tau,x}/(\varrho+P_{\tau,x})

is equipped with the monoid structure

m+m={m+mif m+mQτ,x,otherwise.m+m^{\prime}=\begin{cases}m+m^{\prime}&\text{if }m+m^{\prime}\in Q_{\tau,x},\\ \infty&\text{otherwise.}\end{cases}

We have the ring isomorphism [Qτ,x][Στl]\mathbb{C}[Q_{\tau,x}]\cong\mathbb{C}[\Sigma_{\tau}\oplus\mathbb{N}^{l}] induced by the monoid isomorphism defined by sending (m,a0,a1,,al)(m,a1ψ1(m),,alψl(m))(m,a_{0},a_{1},\dots,a_{l})\mapsto(m,a_{1}-\psi_{1}(m),\dots,a_{l}-\psi_{l}(m)).

We also fix some isomorphism [τ1Σv][Στl]\mathbb{C}[\tau^{-1}\Sigma_{v}]\cong\mathbb{C}[\Sigma_{\tau}\oplus\mathbb{Z}^{l}] coming from the identification of τ1Σv\tau^{-1}\Sigma_{v} with the fan Στ=l{ω|lω is a cone of τ}\Sigma_{\tau}\oplus{}^{l}=\{\omega\oplus{}^{l}\ |\ \omega\text{ is a cone of }\tau\} in 𝒬τ,l\mathscr{Q}_{\tau,\real}\oplus{}^{l}. Taking a sufficiently small neighborhood VV of xx such that ZρV=Z_{\rho}\cap V=\emptyset if xZρx\notin Z_{\rho}, we define a map VSpecan([Qτ,x])V\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[Q_{\tau,x}]) by composing VSpecan([τ1Σv])Specan([Στl])V\hookrightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{v}])\cong\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}\oplus\mathbb{Z}^{l}]) with the map Specan([Στl])Specan([Στl])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}\oplus\mathbb{Z}^{l}])\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}\oplus\mathbb{N}^{l}]) described on generators by

(2.19) {zmhmzmif mΣτ;uifv,iif 1ir;uizizi(x)if r<il;\begin{cases}z^{m}\mapsto h_{m}\cdot z^{m}&\text{if }m\in\Sigma_{\tau};\\ u_{i}\mapsto f_{v,i}&\text{if }1\leq i\leq r;\\ u_{i}\mapsto z_{i}-z_{i}(x)&\text{if }r<i\leq l;\end{cases}

here uiu_{i} is the ii-th coordinate function of [l]\mathbb{C}[\mathbb{N}^{l}], ziz_{i} is the ii-th coordinate function of [l]\mathbb{C}[\mathbb{Z}^{l}] chosen so that (fv,izj)1ir,1jr\left(\frac{\partial f_{v,i}}{\partial{z_{j}}}\right)_{1\leq i\leq r,1\leq j\leq r} is non-degenerate on VV; also, each hmh_{m} is an invertible holomorphic functions on VZero(zm;v)V\cap\mathrm{Zero}(z^{m};v), and they satisfy the equations (2.12) and (2.13) where we replace fvρf_{v\rho} by

f~vρ={svτ1(fvρ)if xZρ,1if xZρ.\tilde{f}_{v\rho}=\begin{cases}s_{v\tau}^{-1}(f_{v\rho})&\text{if }x\notin Z_{\rho},\\ 1&\text{if }x\in Z_{\rho}.\end{cases}

Letting :V𝕍k\prescript{}{}{\flat}\colon V\rightarrow\prescript{k}{}{\mathbb{V}} be the kk-th order thickening of VV over [q]/qk+1\mathbb{C}[q]/q^{k+1} in the model Specan([Pτ,x])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{\tau,x}]) under the above embedding, we have a natural divisorial log structure on 𝕍k\prescript{k}{}{\mathbb{V}}^{\dagger} over Sk\prescript{k}{}{S}^{\dagger} induced from the inclusion Specan([Qτ,x])Specan([Pτ,x])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[Q_{\tau,x}])\hookrightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{\tau,x}]) (i.e. Example 2.18). Restricting it to VV gives the log structure of X0\prescript{0}{}{X}^{\dagger} over the log point S0\prescript{0}{}{S}^{\dagger} locally around xx.

3. A generalized moment map and the tropical singular locus on BB

In this section, we recall the construction of a generalized moment map μ:X0B\mu\colon\prescript{0}{}{X}\rightarrow B from [43, Prop. 2.1]. Then we construct some convenient charts on the base tropical manifold BB and study its singular locus.

3.1. A generalized moment map

From this point onward, we will assume the vanishing of an obstruction class associated to the open gluing data ss, namely, o(s)=1o(s)=1, where the obstruction class o(s)o(s) is written multiplicatively (see [27, Thm. 2.34]). Under this assumption, one can construct an ample line bundle \mathcal{L} on X0\prescript{0}{}{X} as follows: For each polytope τΛτ,\tau\subset\Lambda_{\tau,\real}, by identifying Xτ0\prescript{0}{}{X}_{\tau} (a closed stratum of X0\prescript{0}{}{X} described in Remark 2.14) with the projective toric variety associated to τ\tau, we obtain an ample line bundle τ\mathcal{L}_{\tau} on Xτ0\prescript{0}{}{X}_{\tau}. When the assumption holds, then there exists an isomorphism 𝐡ωτ:ιωτ(τ)ω\mathbf{h}_{\omega\tau}\colon\iota_{\omega\tau}^{*}(\mathcal{L}_{\tau})\cong\mathcal{L}_{\omega}, for every pair ωτ\omega\subset\tau, such that the isomorphisms 𝐡ωτ\mathbf{h}_{\omega\tau}’s satisfy the cocycle condition, i.e. 𝐡ωτιωτ(𝐡τσ)=𝐡ωσ\mathbf{h}_{\omega\tau}\circ\iota_{\omega\tau}^{*}(\mathbf{h}_{\tau\sigma})=\mathbf{h}_{\omega\sigma} for every triple ωτσ\omega\subset\tau\subset\sigma.333In fact, the vanishing of the obstruction class corresponds exactly to the validity of the cocycle condition. In particular, the degenerate Calabi–Yau X0=X0(B,𝒫,s)\prescript{0}{}{X}=\prescript{0}{}{X}(B,\mathscr{P},s) is projective.

Sections of \mathcal{L} correspond to the lattice points BBB_{\mathbb{Z}}\subset B. More precisely, given mBm\in B_{\mathbb{Z}}, there is a unique τ𝒫\tau\in\mathscr{P} such that mintre(τ)m\in\mathrm{int}_{\mathrm{re}}(\tau), and this determines a section ϑm,τ\vartheta_{m,\tau} of τ\mathcal{L}_{\tau} by toric geometry. This section extends uniquely as ϑm\vartheta_{m} to στ\sigma\supset\tau such that 𝐡τσ(ϑm)=ϑm,τ\mathbf{h}_{\tau\sigma}(\vartheta_{m})=\vartheta_{m,\tau}. Further extending ϑm\vartheta_{m} by 0 to other cells gives a section of \mathcal{L} corresponding to mm, called a (0th0^{\text{th}}-order) theta function. Now for a vertex v𝒫[0]v\in\mathscr{P}^{[0]}, we can trivialize \mathcal{L} over V(v)V(v) using ϑv\vartheta_{v} as the holomorphic frame. Then, for mm lying in a cell σ\sigma that contains vv, ϑm\vartheta_{m} is of the form gϑvg\vartheta_{v}, where gg is a constant multiple of zmz^{m}.

Under the above projectivity assumption, one can define a generalized moment map

(3.1) μ:X0B\mu\colon\prescript{0}{}{X}\rightarrow B

following [43, Prop. 2.1]: First of all, the theta functions {ϑm}mB\{\vartheta_{m}\}_{m\in B_{\mathbb{Z}}} defines an embedding of Φ:X0N\Phi\colon\prescript{0}{}{X}\hookrightarrow\mathbb{P}^{N}. Restricting to each closed toric stratum Xτ0X0\prescript{0}{}{X}_{\tau}\subset\prescript{0}{}{X}, the only non-zero theta functions are those corresponding to mBτm\in B_{\mathbb{Z}}\cap\tau. Also, there is an embedding 𝔧τ:𝚃τ:=Λτ,/Λτ,𝚄(1)N\mathfrak{j}_{\tau}\colon\mathtt{T}_{\tau}:=\Lambda_{\tau,\real}^{*}/\Lambda_{\tau,\mathbb{Z}}^{*}\hookrightarrow\mathtt{U}(1)^{N} of real tori such that the composition Φτ:Xτ0N\Phi_{\tau}\colon\prescript{0}{}{X}_{\tau}\rightarrow\mathbb{P}^{N} of Φ\Phi with the inclusion Xτ0X0\prescript{0}{}{X}_{\tau}\hookrightarrow\prescript{0}{}{X} is equivariant. The map μ\mu is then defined by setting

(3.2) μ|Xτ0(z):=1mBτ|ϑm(z)|2mBτ|ϑm(z)|2m,\mu|_{\prescript{0}{}{X}_{\tau}}(z):=\frac{1}{\sum_{m\in B_{\mathbb{Z}}\cap\tau}|\vartheta_{m}(z)|^{2}}\sum_{m\in B_{\mathbb{Z}}\cap\tau}|\vartheta_{m}(z)|^{2}\cdot m,

which can be understood as a composition of maps

Xτ0\textstyle{\prescript{0}{}{X}_{\tau}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φτ\scriptstyle{\Phi_{\tau}}N\textstyle{\mathbb{P}^{N}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ\scriptstyle{\mu_{\mathbb{P}}}()N\textstyle{({}^{N})^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d𝔧τ\scriptstyle{d\mathfrak{j}^{*}_{\tau}}Λτ,,\textstyle{\Lambda_{\tau,\real},}

where μ\mu_{\mathbb{P}} is the standard moment map for N\mathbb{P}^{N} and d𝔧τ:Λτ,Nd\mathfrak{j}_{\tau}\colon\Lambda_{\tau,\real}^{*}\rightarrow{}^{N} is the Lie algebra homomorphism induced by 𝔧τ:𝚃τ𝚄(1)N\mathfrak{j}_{\tau}\colon\mathtt{T}_{\tau}\rightarrow\mathtt{U}(1)^{N}.

Fixing a vertex v𝒫[0]v\in\mathscr{P}^{[0]}, we can naturally embed Λτ,Tv,\Lambda_{\tau,\real}\hookrightarrow T_{v,\real} for all τ\tau containing vv. Furthermore, we can patch the d𝔧τd\mathfrak{j}^{*}_{\tau}’s into a linear map d𝔧:()NTv,d\mathfrak{j}^{*}\colon({}^{N})^{*}\rightarrow T_{v,\real} so that μτ=d𝔧μΦτ\mu_{\tau}=d\mathfrak{j}^{*}\circ\mu_{\mathbb{P}}\circ\Phi_{\tau} for each τ\tau which contains vv. In particular, on the local chart V(τ)=Specan([τ1Σv])V(\tau)=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{v}]) associated with vτv\in\tau, we have the local description μ|V(τ)=d𝔧μΦ|V(τ)\mu|_{V(\tau)}=d\mathfrak{j}^{*}\circ\mu_{\mathbb{P}}\circ\Phi|_{V(\tau)} of the generalized moment map μ\mu.

We consider the amoeba 𝒜:=μ(Z)\mathcal{A}:=\mu(Z). As Xτ0Z=i=1pZiτ\prescript{0}{}{X}_{\tau}\cap Z=\bigcup_{i=1}^{p}Z^{\tau}_{i}, where ZiτZ^{\tau}_{i} is the zero set of a section of ϰτ,i(𝒪(1))\varkappa_{\tau,i}^{*}(\mathcal{O}(1)) (see the discussion right after equation (2.15)), we can see that 𝒜τ=i=1pμτ(Ziτ)\mathcal{A}\cap\tau=\bigcup_{i=1}^{p}\mu_{\tau}(Z^{\tau}_{i}) is a union of amoebas 𝒜iτ:=μτ(Ziτ)\mathcal{A}^{\tau}_{i}:=\mu_{\tau}(Z^{\tau}_{i}). It was shown in [43] that the affine structure defined right after Definition 2.2 extends to B𝒜B\setminus\mathcal{A}.

3.2. Construction of charts on BB

For any τ𝒫\tau\in\mathscr{P}, we have

μ(V(τ))=τωintre(ω)=:W(τ).\mu(V(\tau))=\bigcup_{\tau\subset\omega}\mathrm{int}_{\mathrm{re}}(\omega)=:W(\tau).

For later purposes, we would like to relate sufficiently small open convex subsets WW(τ)W\subset W(\tau) with Stein (or strongly 11-completed, as defined in [13]) open subsets UV(τ)U\subset V(\tau). To do so, we need to introduce a specific collection of (non-affine) charts on BB.

Recall that there are natural maps Λττ1Σv\Lambda_{\tau}\hookrightarrow\tau^{-1}\Sigma_{v} and τ1ΣvΣτ\tau^{-1}\Sigma_{v}\twoheadrightarrow\Sigma_{\tau}. By choosing a piecewise linear splitting 𝗌𝗉𝗅𝗂𝗍τ:Σττ1Σv\mathsf{split}_{\tau}\colon\Sigma_{\tau}\rightarrow\tau^{-1}\Sigma_{v}, we have an identification of monoids τ1ΣvΣτ×Λτ\tau^{-1}\Sigma_{v}\cong\Sigma_{\tau}\times\Lambda_{\tau}, which induces the biholomorphism

V(τ)=Specan([τ1Σv])Specan([Λτ])×Specan([Στ]),V(\tau)=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{v}])\cong\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Lambda_{\tau}])\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}]),

where Λτ,:=Specan([Λτ])Λτ()l\Lambda_{\tau,\mathbb{C}^{*}}^{*}:=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Lambda_{\tau}])\cong\Lambda_{\tau}^{*}\otimes_{\mathbb{Z}}\mathbb{C}^{*}\cong(\mathbb{C}^{*})^{l} is a complex torus. Fixing a set of generators {mi}i𝙱τ\{m_{i}\}_{i\in\mathtt{B}_{\tau}} of the monoid Στ\Sigma_{\tau}, which is not necessarily a minimal set, we can define an embedding Specan([Στ])|𝙱τ|\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])\hookrightarrow\mathbb{C}^{|\mathtt{B}_{\tau}|} as an analytic subset using the functions zmiz^{m_{i}}’s. We consider the real torus 𝚃τ,:=𝒬τ,/𝒬τ𝚄(1)nl\mathtt{T}_{\tau,\perp}:=\mathscr{Q}_{\tau,\real}^{*}/\mathscr{Q}_{\tau}^{*}\cong\mathtt{U}(1)^{n-l} and its action on Specan([Στ])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}]) defined by tzm=e2πi(t,m)zmt\cdot z^{m}=e^{2\pi i(t,m)}z^{m}, together with an embedding 𝚃τ,𝚄(1)|𝙱τ|\mathtt{T}_{\tau,\perp}\hookrightarrow\mathtt{U}(1)^{|\mathtt{B}_{\tau}|} of real tori via t(e2πi(t,mi))i𝙱τt\mapsto(e^{2\pi i(t,m_{i})})_{i\in\mathtt{B}_{\tau}}, so that Specan([Στ])|𝙱τ|\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])\hookrightarrow\mathbb{C}^{|\mathtt{B}_{\tau}|} is 𝚃τ,\mathtt{T}_{\tau,\perp}-equivariant.

We consider the moment map μ^τ:Specan([Στ])𝒬τ,\hat{\mu}_{\tau}\colon\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])\rightarrow\mathscr{Q}_{\tau,\real} defined by

(3.3) μ^τ:=i𝙱τ12|zmi|2mi,\hat{\mu}_{\tau}:=\sum_{i\in\mathtt{B}_{\tau}}\frac{1}{2}|z^{m_{i}}|^{2}\cdot m_{i},

which is obtained by composing the standard moment map |𝙱τ|0|𝙱τ|\mathbb{C}^{|\mathtt{B}_{\tau}|}\rightarrow{}^{|\mathtt{B}_{\tau}|}_{\geq 0}, (zi)i𝙱τ(12|zi|2)i𝙱τ(z_{i})_{i\in\mathtt{B}_{\tau}}\mapsto(\frac{1}{2}|z_{i}|^{2})_{i\in\mathtt{B}_{\tau}} with the projection |𝙱τ|𝒬τ,{}^{|\mathtt{B}_{\tau}|}\rightarrow\mathscr{Q}_{\tau,\real}, eimie_{i}\mapsto m_{i}. By [21, §4.2], μ^τ\hat{\mu}_{\tau} induces a homeomorphism between the quotient Specan([Στ])/𝚃τ,\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])/\mathtt{T}_{\tau,\perp} and 𝒬τ,\mathscr{Q}_{\tau,\real}. Taking product with the log map log:Λτ,Λτ,\log\colon\Lambda_{\tau,\mathbb{C}^{*}}^{*}\rightarrow\Lambda_{\tau,\real}^{*} (which is induced from the standard log map log:\log\colon\mathbb{C}^{*}\rightarrow\real defined by log(e2π(x+iθ))=x\log(e^{2\pi(x+i\theta)})=x), we obtain a map μτ:=(log,μ^τ):V(τ)Λτ,×𝒬τ,\mu_{\tau}:=(\log,\hat{\mu}_{\tau})\colon V(\tau)\rightarrow\Lambda_{\tau,\real}^{*}\times\mathscr{Q}_{\tau,\real},444It depends on the choices of the splitting 𝗌𝗉𝗅𝗂𝗍τ:Σττ1Σv\mathsf{split}_{\tau}\colon\Sigma_{\tau}\rightarrow\tau^{-1}\Sigma_{v} and the generators {mi}i\{m_{i}\}_{i}, but we omit these dependencies from our notations. and the following diagram

(3.4) V(τ)\textstyle{V(\tau)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ\scriptstyle{\mu}μτ\scriptstyle{\mu_{\tau}}Λτ,×𝒬τ,\textstyle{\Lambda_{\tau,\real}^{*}\times\mathscr{Q}_{\tau,\real}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Υτ\scriptstyle{\Upsilon_{\tau}}W(τ),\textstyle{W(\tau),}

where Υτ\Upsilon_{\tau} is a homeomorphism which serves as a chart.

The homeomorphism Υτ\Upsilon_{\tau} exists because if we fix a vertex vτv\in\tau, then we can equip V(τ)V(\tau) with an action by the real torus 𝚃n:=Tv,/Tv\mathtt{T}^{n}:=T^{*}_{v,\mathbb{R}}/T^{*}_{v} such that both μ\mu and μτ\mu_{\tau} induce homeomorphisms from the quotient V(τ)/𝚃nV(\tau)/\mathtt{T}^{n} onto the images. The restriction of Υτ\Upsilon_{\tau} to Λτ,×{o}\Lambda_{\tau,\real}^{*}\times\{o\}, where {o}\{o\} is the zero cone, is a homeomorphism onto intre(τ)W(τ)\mathrm{int}_{\mathrm{re}}(\tau)\subset W(\tau), which is nothing but (a generalized version of) the Legendre transform (see [21, §4.2] for the explicit formula); also, this homeomorphism is independent of the choices of the splitting 𝗌𝗉𝗅𝗂𝗍τ\mathsf{split}_{\tau} and the generators {mi}i𝙱τ\{m_{i}\}_{i\in\mathtt{B}_{\tau}}.

The dependences of the chart Υτ\Upsilon_{\tau} on the choices of the splitting 𝗌𝗉𝗅𝗂𝗍τ:Σττ1Σv\mathsf{split}_{\tau}\colon\Sigma_{\tau}\rightarrow\tau^{-1}\Sigma_{v} and the generators {mi}i\{m_{i}\}_{i} can be described as follows. First, if we choose another piecewise linear splitting 𝗌𝗉𝗅𝗂𝗍~τ:Σττ1Σv\widetilde{\mathsf{split}}_{\tau}\colon\Sigma_{\tau}\rightarrow\tau^{-1}\Sigma_{v}, then there is a piecewise linear map b:ΣτΛτ,b\colon\Sigma_{\tau}\rightarrow\Lambda_{\tau,\real} recording the difference between 𝗌𝗉𝗅𝗂𝗍τ\mathsf{split}_{\tau} and 𝗌𝗉𝗅𝗂𝗍~τ\widetilde{\mathsf{split}}_{\tau}. The two corresponding coordinate charts Υτ\Upsilon_{\tau} and Υ~τ\tilde{\Upsilon}_{\tau} are then related by a homeomorphism \gimel such that

(x,iyimi)=(x,iyie4πb(mi),xmi),\gimel\left(x,\sum_{i}y_{i}m_{i}\right)=\left(x,\sum_{i}y_{i}e^{4\pi\langle b(m_{i}),x\rangle}m_{i}\right),

where yi=12|zmi|2y_{i}=\frac{1}{2}|z^{m_{i}}|^{2} for some point zSpecan([Στ])z\in\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}]) and ii runs through miσm_{i}\in\sigma, via the formula Υ~τ=Υτ\tilde{\Upsilon}_{\tau}=\Upsilon_{\tau}\circ\gimel. Second, if we choose another set of generators m~j\tilde{m}_{j}’s, then the corresponding maps μ^τ,μ~τ:Specan([Στ])𝒬τ,\hat{\mu}_{\tau},\tilde{\mu}_{\tau}\colon\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])\rightarrow\mathscr{Q}_{\tau,\real} are related by a continuous map ^:𝒬τ,𝒬τ,\hat{\gimel}\colon\mathscr{Q}_{\tau,\real}\rightarrow\mathscr{Q}_{\tau,\real} which maps each cone σΣτ\sigma\in\Sigma_{\tau} back to itself. This is because both μ^τ,μ~τ\hat{\mu}_{\tau},\tilde{\mu}_{\tau} induce a homeomorphism between Specan([Στ])/𝚃τ,\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])/\mathtt{T}_{\tau,\perp} and 𝒬τ,\mathscr{Q}_{\tau,\real}.

Now suppose that ωτ\omega\subset\tau. We want to see how the charts Υω\Upsilon_{\omega}, Υτ\Upsilon_{\tau} can be glued together in a compatible manner. We first make a compatible choice of splittings. So we fix a vertex vωv\in\omega and a piecewise linear splitting 𝗌𝗉𝗅𝗂𝗍ω:Σωω1Σv\mathsf{split}_{\omega}\colon\Sigma_{\omega}\rightarrow\omega^{-1}\Sigma_{v}. We then choose a piecewise linear splitting 𝗌𝗉𝗅𝗂𝗍ωτ:ΣτΣω\mathsf{split}_{\omega\tau}\colon\Sigma_{\tau}\rightarrow\Sigma_{\omega} such that KτσK_{\tau}\sigma is mapped into KωσK_{\omega}\sigma for any στ\sigma\supset\tau. Together with the natural maps Λτ/Λωτ1Σω\Lambda_{\tau}/\Lambda_{\omega}\hookrightarrow\tau^{-1}\Sigma_{\omega} and τ1ΣωΣτ\tau^{-1}\Sigma_{\omega}\twoheadrightarrow\Sigma_{\tau}, we obtain an isomorphism τ1Σω(Λτ/Λω)×Στ\tau^{-1}\Sigma_{\omega}\cong(\Lambda_{\tau}/\Lambda_{\omega})\times\Sigma_{\tau}. By composing together 𝗌𝗉𝗅𝗂𝗍ωτ:ΣτΣω\mathsf{split}_{\omega\tau}\colon\Sigma_{\tau}\rightarrow\Sigma_{\omega}, 𝗌𝗉𝗅𝗂𝗍ω:Σωω1Σv\mathsf{split}_{\omega}\colon\Sigma_{\omega}\rightarrow\omega^{-1}\Sigma_{v} and the natural monoid homomorphism ω1Σvτ1Σv\omega^{-1}\Sigma_{v}\rightarrow\tau^{-1}\Sigma_{v}, we get a splitting 𝗌𝗉𝗅𝗂𝗍τ:Σττ1Σv\mathsf{split}_{\tau}\colon\Sigma_{\tau}\rightarrow\tau^{-1}\Sigma_{v}.

Using these choices of splittings, we have a biholomorphism

Specan([τ1Σω])(Λτ/Λω)×Specan([Στ])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{\omega}])\cong(\Lambda_{\tau}/\Lambda_{\omega})^{*}\otimes_{\mathbb{Z}}\mathbb{C}^{*}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])

which fits into the following diagram

(3.5) Λω,×Specan([Σω])\textstyle{\Lambda_{\omega,\mathbb{C}^{*}}^{*}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\omega}])\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}Specan([ω1Σv])\textstyle{\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\omega^{-1}\Sigma_{v}])}Λω,×Specan([τ1Σω])\textstyle{\Lambda_{\omega,\mathbb{C}^{*}}^{*}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{\omega}])\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}Specan([τ1Σv])\textstyle{\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{v}])\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}\scriptstyle{\cong}Specan([τ1Σv])\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{v}])\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sωτ1\scriptstyle{s_{\omega\tau}^{-1}}\scriptstyle{\cong}Fs(ωτ)\scriptstyle{F_{s}(\omega\subset\tau)}(ΛωΛτ/Λω)×Specan([Στ])\textstyle{(\Lambda_{\omega}\oplus\Lambda_{\tau}/\Lambda_{\omega})^{*}\otimes_{\mathbb{Z}}\mathbb{C}^{*}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])}Λτ,×Specan([Στ])\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\Lambda_{\tau,\mathbb{C}^{*}}^{*}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])}Λτ,×Specan([Στ]).\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\Lambda_{\tau,\mathbb{C}^{*}}^{*}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}]).}sωτ1\scriptstyle{s_{\omega\tau}^{-1}}

Here, the bottom left horizontal map is induced from a splitting (Λτ/Λω)Λτ(\Lambda_{\tau}/\Lambda_{\omega})\rightarrow\Lambda_{\tau} obtained by composing Λτ/Λωτ1Σω\Lambda_{\tau}/\Lambda_{\omega}\rightarrow\tau^{-1}\Sigma_{\omega} with the splitting τ1Σωτ1(ω1Σv)\tau^{-1}\Sigma_{\omega}\rightarrow\tau^{-1}(\omega^{-1}\Sigma_{v}), and then identifying with the image lattice Λτ\Lambda_{\tau}. The appearance of sωτs_{\omega\tau} in the diagram is due to the twisting of V(τ)V(\tau) by the open gluing data (sωτ)ωτ(s_{\omega\tau})_{\omega\subset\tau} when it is glued to V(ω)V(\omega).

We also have to make a compatible choice of the generators {mi}i𝙱ω\{m_{i}\}_{i\in\mathtt{B}_{\omega}} and {mi}i𝙱τ\{m_{i}\}_{i\in\mathtt{B}_{\tau}}. First note that the restriction of μ^ω\hat{\mu}_{\omega} to the open subset Specan([τ1Σω])Specan([Σω])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{\omega}])\subset\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\omega}]) depends only on the subcollection {mi}i𝙱ωτ\{m_{i}\}_{i\in\mathtt{B}_{\omega\subset\tau}} of {mi}i𝙱ω\{m_{i}\}_{i\in\mathtt{B}_{\omega}} which contains those mim_{i}’s that belong to some cone στ\sigma\supset\tau. We choose the set of generators {m~i}i𝙱τ\{\tilde{m}_{i}\}_{i\in\mathtt{B}_{\tau}} for Στ\Sigma_{\tau}, with 𝙱τ=𝙱ωτ\mathtt{B}_{\tau}=\mathtt{B}_{\omega\subset\tau}, to be the projection of {mi}i𝙱ωτ\{m_{i}\}_{i\in\mathtt{B}_{\omega\subset\tau}} through the natural map τ1ΣωΣτ\tau^{-1}\Sigma_{\omega}\rightarrow\Sigma_{\tau}. Each mim_{i} can be expressed as mi=𝗌𝗉𝗅𝗂𝗍ωτ(m~i)+bim_{i}=\mathsf{split}_{\omega\tau}(\tilde{m}_{i})+b_{i} for some biΛτ/Λωb_{i}\in\Lambda_{\tau}/\Lambda_{\omega}, through the splitting 𝗌𝗉𝗅𝗂𝗍ωτ:ΣτΣω\mathsf{split}_{\omega\tau}\colon\Sigma_{\tau}\rightarrow\Sigma_{\omega}. Notice that if miKωτm_{i}\in K_{\omega}\tau, then we have m~i=o\tilde{m}_{i}=o and hence biKωτb_{i}\in K_{\omega}\tau. By tracing through the biholomorphism in (3.5) and taking either the modulus or the log map, we have a map

:Λω,×(Λτ,/Λω,)×𝒬τ,Λω,×𝒬ω,\gimel\colon\Lambda_{\omega,\real}^{*}\times(\Lambda_{\tau,\real}/\Lambda_{\omega,\real})^{*}\times\mathscr{Q}_{\tau,\real}\rightarrow\Lambda_{\omega,\real}^{*}\times\mathscr{Q}_{\omega,\real}

satisfying

(3.6) (x1cωτ,1,x2cωτ,2,iyi|sωτ(𝗌𝗉𝗅𝗂𝗍ωτ(m~i))|2m~i)=(x1,iyie4πbi,x2mi),\gimel\left(x_{1}-c_{\omega\tau,1},x_{2}-c_{\omega\tau,2},\sum_{i}y_{i}|s_{\omega\tau}(\mathsf{split}_{\omega\tau}(\tilde{m}_{i}))|^{-2}\tilde{m}_{i}\right)=\left(x_{1},\sum_{i}y_{i}e^{4\pi\langle b_{i},x_{2}\rangle}m_{i}\right),

where yi=12|zm~i|2y_{i}=\frac{1}{2}|z^{\tilde{m}_{i}}|^{2}. Here, sωτPM(τ)s_{\omega\tau}\in\mathrm{PM}(\tau) is the part of the open gluing data associated to ωτ\omega\subset\tau, and cωτ=cωτ,1+cωτ,2Λτ,c_{\omega\tau}=c_{\omega\tau,1}+c_{\omega\tau,2}\in\Lambda_{\tau,\real}^{*} is the unique element representing the linear map log|sωτ|:Λτ,\log|s_{\omega\tau}|\colon\Lambda_{\tau,\real}\rightarrow\real defined by log|sωτ|(b)=log|sωτ(b)|\log|s_{\omega\tau}|(b)=\log|s_{\omega\tau}(b)|. For instance, the holomorphic function zmi[τ1Σω]z^{m_{i}}\in\mathbb{C}[\tau^{-1}\Sigma_{\omega}] is identified with zbizm~iz^{b_{i}}\cdot z^{\tilde{m}_{i}} in (Λτ/Λω)×Specan([Στ])(\Lambda_{\tau}/\Lambda_{\omega})^{*}\otimes_{\mathbb{Z}}\mathbb{C}^{*}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}]), resulting in the expression iyie4πbi,x2mi\sum_{i}y_{i}e^{4\pi\langle b_{i},x_{2}\rangle}m_{i} on the right hand side. We have Υτ=Υω\Upsilon_{\tau}=\Upsilon_{\omega}\circ\gimel, where we use the splitting (Λτ/Λω)Λτ(\Lambda_{\tau}/\Lambda_{\omega})\rightarrow\Lambda_{\tau} to obtain an isomorphism Λω,×(Λτ,/Λω,)Λτ,\Lambda_{\omega,\real}^{*}\times(\Lambda_{\tau,\real}/\Lambda_{\omega,\real})^{*}\cong\Lambda_{\tau,\real}^{*} and an identification of the domains of the two maps Υτ\Upsilon_{\tau} and Υω\Upsilon_{\omega}\circ\gimel.

Lemma 3.1.

There is a base \mathscr{B} of open subsets of BB such that the preimage μ1(W)\mu^{-1}(W) is Stein for any WW\in\mathscr{B}.

Proof.

First of all, it is well-known that analytic spaces associated to affine varieties are Stein. So V(τ)V(\tau) is Stein for any τ\tau. Now we fix a point xintre(τ)Bx\in\mathrm{int}_{\mathrm{re}}(\tau)\subset B. It suffices to show that there is a local base x\mathscr{B}_{x} of xx such that the preimage μ1(W)\mu^{-1}(W) is Stein for each WxW\in\mathscr{B}_{x}. We work locally on μ|V(τ):V(τ)W(τ)\mu|_{V(\tau)}\colon V(\tau)\rightarrow W(\tau). Consider the diagram (3.4) and write Υ1(x)=(x¯,o)\Upsilon^{-1}(x)=(\underline{x},o), where o𝒬τ,o\in\mathscr{Q}_{\tau,\real} is the origin. By [13, Ch. 1, Ex. 7.4], the preimage log1(W)\log^{-1}(W) under the log map log:()lΛτ,\log\colon(\mathbb{C}^{*})^{l}\rightarrow\Lambda_{\tau,\real}^{*} is Stein for any convex WΛτ,W\subset\Lambda_{\tau,\real}^{*} which contains x¯\underline{x}. Again by [13, Ch. 1, Ex. 7.4], any subset

j=1N{zSpecan([Στ])||fj(z)|<ϵ},\bigcap_{j=1}^{N}\{z\in\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])\ |\ |f_{j}(z)|<\epsilon\},

where fjf_{j}’s are holomorphic functions, is Stein. By taking fjf_{j}’s to be the functions zmjz^{m_{j}}’s associated to the set of all non-zero generators in {mj}j𝙱τ\{m_{j}\}_{j\in\mathtt{B}_{\tau}} and ϵ\epsilon sufficiently small, we have a subset

W={y|y=jyjmj with |yj|<ϵ22,where yj=12|zmj|2 at some point zSpecan([Στ])}W=\left\{y\ \Big{|}\ y=\sum_{j}y_{j}m_{j}\text{ with }|y_{j}|<\frac{\epsilon^{2}}{2},\ \text{where $y_{j}=\frac{1}{2}|z^{m_{j}}|^{2}$ at some point $z\in\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}])$}\right\}

of 𝒬τ,\mathscr{Q}_{\tau,\real} such that the preimage μ^τ1(W)\hat{\mu}_{\tau}^{-1}(W) is Stein. Therefore, we can construct a local base o\mathscr{B}_{o} of oo such that the preimage μ^τ1(W)\hat{\mu}_{\tau}^{-1}(W) is Stein for any WoW\in\mathscr{B}_{o}. Finally, since a product of Stein open subsets is Stein, we obtain our desired local base x\mathscr{B}_{x} by taking the products of these subsets. ∎

3.3. The tropical singular locus 𝒮\mathscr{S} of BB

We now specify a codimension 22 singular locus 𝒮B\mathscr{S}\subset B of the affine structure using the charts Υτ\Upsilon_{\tau} introduced in (3.4) for τ\tau such that dim(τ)<n\dim(\tau)<n. Given the chart Υτ\Upsilon_{\tau} that maps Λτ,\Lambda_{\tau,\real}^{*} to intre(τ)\mathrm{int}_{\mathrm{re}}(\tau), we define the tropical singular locus 𝒮\mathscr{S} by requiring that

(3.7) Υτ1(𝒮intre(τ))=ρ𝒩τ;dim(ρ)<dim(τ)((intre(ρ)+cτ)×{o}),\Upsilon_{\tau}^{-1}(\mathscr{S}\cap\mathrm{int}_{\mathrm{re}}(\tau))=\bigcup_{\begin{subarray}{c}\rho\in\mathscr{N}_{\tau};\\ \dim(\rho)<\dim(\tau)\end{subarray}}\big{(}(\mathrm{int}_{\mathrm{re}}(\rho)+c_{\tau})\times\{o\}\big{)},

where 𝒩τΛτ,\mathscr{N}_{\tau}\subset\Lambda_{\tau,\real}^{*} is the normal fan of the polytope τ\tau, and {o}\{o\} is the zero cone in Στ𝒬τ,\Sigma_{\tau}\subset\mathscr{Q}_{\tau,\real}; here, cτ=log|svτ|c_{\tau}=\log|s_{v\tau}| is the element in Λτ,\Lambda_{\tau,\real}^{*} representing the linear map log|svτ|:Λτ,\log|s_{v\tau}|\colon\Lambda_{\tau,\real}\rightarrow\real, which is independent of the vertex vτv\in\tau. A subset of the form 𝒮τ,ρ:=(intre(ρ)+cτ)×{o}\mathscr{S}_{\tau,\rho}:=(\mathrm{int}_{\mathrm{re}}(\rho)+c_{\tau})\times\{o\} in (3.7) is called a stratum of 𝒮\mathscr{S} in intre(τ)\mathrm{int}_{\mathrm{re}}(\tau). The locus 𝒮\mathscr{S} is independent of the choices of the splittings 𝗌𝗉𝗅𝗂𝗍τ\mathsf{split}_{\tau}’s and generators {mi}i𝙱τ\{m_{i}\}_{i\in\mathtt{B}_{\tau}} used to construct the charts Υτ\Upsilon_{\tau}’s.

Remark 3.2.

Our definition of the singular locus is similar to those in [27, 29]; the only difference is that our locus is a collection of polyhedra in Λτ,\Lambda_{\tau,\real}^{*}, instead of intre(τ)\mathrm{int}_{\mathrm{re}}(\tau). Note that Λτ,\Lambda_{\tau,\real}^{*} is homeomorphic to intre(τ)\mathrm{int}_{\mathrm{re}}(\tau) by the Legendre transform. This modification is needed for our construction of the contraction map 𝒞\mathscr{C} below, where we need to consider the convex open subsets in Λτ,\Lambda_{\tau,\real}^{*}, instead of those in intre(τ)\mathrm{int}_{\mathrm{re}}(\tau).

Lemma 3.3.

For ωτ\omega\subset\tau and a stratum 𝒮τ,ρ\mathscr{S}_{\tau,\rho} in intre(τ)\mathrm{int}_{\mathrm{re}}(\tau), the intersection of the closure 𝒮τ,ρ¯\overline{\mathscr{S}_{\tau,\rho}} in BB with intre(ω)\mathrm{int}_{\mathrm{re}}(\omega) is a union of strata of 𝒮\mathscr{S} in intre(ω)\mathrm{int}_{\mathrm{re}}(\omega).

Proof.

We consider the map \gimel described in equation (3.6) and take a neighborhood W=W1×𝒬ω,W=W_{1}\times\mathscr{Q}_{\omega,\real} of a point (x¯,o)(\underline{x},o) in Λω,×𝒬ω,\Lambda_{\omega,\real}^{*}\times\mathscr{Q}_{\omega,\real}, where W1W_{1} is some sufficiently small neighborhood of x¯\underline{x} in Λω,\Lambda_{\omega,\real}^{*}. By shrinking WW if necessary, we may assume that 1(W)=W1×(aintre(Kωτ))×𝒬τ,\gimel^{-1}(W)=W_{1}\times(a-\mathrm{int}_{\mathrm{re}}(K_{\omega}\tau^{\vee}))\times\mathscr{Q}_{\tau,\real}, where aa is some element in intre(Kωτ)(Λτ,/Λω,)-\mathrm{int}_{\mathrm{re}}(K_{\omega}\tau^{\vee})\subset(\Lambda_{\tau,\real}/\Lambda_{\omega,\real})^{*}. Writing cτ=cτ,1+cτ,2c_{\tau}=c_{\tau,1}+c_{\tau,2}, where cτ,1,cτ,2c_{\tau,1},c_{\tau,2} are the components of cτc_{\tau} according to the chosen decomposition Λτ,Λω,×(Λτ,/Λω,)\Lambda_{\tau,\real}^{*}\cong\Lambda_{\omega,\real}^{*}\times(\Lambda_{\tau,\real}/\Lambda_{\omega,\real})^{*}, the equality cτ,1+cωτ,1=cωc_{\tau,1}+c_{\omega\tau,1}=c_{\omega} follows from the compatibility of the open gluing data in Definition 2.13. If 𝒮τ,ρ\mathscr{S}_{\tau,\rho} intersects the open subset 1(W)\gimel^{-1}(W), then ρΛτ,\rho\subset\Lambda_{\tau,\real}^{*} must be the dual cone of some face ρωτ\rho^{\vee}\subset\omega\subset\tau in Λτ,\Lambda_{\tau,\real}^{*}. The intersection is of the form

(intre(ρ¯)+cτ,1)×(aintre(Kωτ))×{o}(\mathrm{int}_{\mathrm{re}}(\underline{\rho})+c_{\tau,1})\times(a-\mathrm{int}_{\mathrm{re}}(K_{\omega}\tau^{\vee}))\times\{o\}

for some ρ¯𝒩ω\underline{\rho}\in\mathscr{N}_{\omega} (cτ,2c_{\tau,2} is absorbed by aa), where ρ¯Λω,\underline{\rho}\subset\Lambda_{\omega,\real}^{*} is the dual cone of ρ\rho^{\vee} in Λω,\Lambda_{\omega,\real}^{*}, and hence we have W𝒮τ,ρ=((intre(ρ¯)+cτ,1)×(aintre(Kωτ))×{o})W\cap\mathscr{S}_{\tau,\rho}=\gimel((\mathrm{int}_{\mathrm{re}}(\underline{\rho})+c_{\tau,1})\times(a-\mathrm{int}_{\mathrm{re}}(K_{\omega}\tau^{\vee}))\times\{o\}). Therefore, the intersection of 𝒮τ,ρ¯\overline{\mathscr{S}_{\tau,\rho}} with Λω,\Lambda_{\omega,\real}^{*} in the open subset WΛω,×𝒬ω,W\subset\Lambda_{\omega,\real}^{*}\times\mathscr{Q}_{\omega,\real} is given by (ρ¯+cω)×{o}(\underline{\rho}+c_{\omega})\times\{o\}, which is a union of strata. ∎

The tropical singular locus 𝒮\mathscr{S} is naturally equipped with a stratification, where a stratum is given by 𝒮τ,ρ\mathscr{S}_{\tau,\rho} for some cone ρ𝒩τ\rho\subset\mathscr{N}_{\tau} of dim(ρ)<dim(τ)\dim(\rho)<\dim(\tau) for some τ𝒫[<n]\tau\in\mathscr{P}^{[<n]}. We use the notation 𝒮[k]\mathscr{S}^{[k]} to denote the set of kk-dimensional strata of 𝒮\mathscr{S}. The affine structure on v𝒫[0]Wvσ𝒫[n]intre(σ)\bigcup_{v\in\mathscr{P}^{[0]}}W_{v}\cup\bigcup_{\sigma\in\mathscr{P}^{[n]}}\mathrm{int}_{\mathrm{re}}(\sigma) introduced right after Definition 2.2 in §2.1 can be naturally extended to B𝒮B\setminus\mathscr{S} as in [29].

If we consider ωτρ\omega\subset\tau\subset\rho for some ω𝒫[1]\omega\in\mathscr{P}^{[1]} and ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]}, the corresponding monodromy transformation TγT_{\gamma} is non-trivial if and only if ωΩp\omega\in\Omega_{p} and ρRp\rho\in R_{p}, where pp is as in Definition 2.9. Therefore, the part of the singular locus 𝒮\mathscr{S} lying in Υτ1(intre(τ))=Λτ,×{o}\Upsilon_{\tau}^{-1}(\mathrm{int}_{\mathrm{re}}(\tau))=\Lambda_{\tau,\real}^{*}\times\{o\} is determined by the subsets Ωp\Omega_{p}’s. We may further define the essential singular locus 𝒮e\mathscr{S}_{e} to include only those strata contained in 𝒮[n2]\mathscr{S}^{[n-2]} with non-trivial monodromy around them. We observe that the affine structure can be further extended to B𝒮eB\setminus\mathscr{S}_{e}.

More explicitly, we have a projection

𝚒τ=𝚒τ,1𝚒τ,p:ΛτΛΔ1(τ)ΛΔp(τ),\mathtt{i}_{\tau}=\mathtt{i}_{\tau,1}\oplus\cdots\oplus\mathtt{i}_{\tau,p}\colon\Lambda_{\tau}^{*}\rightarrow\Lambda_{\Delta_{1}(\tau)}^{*}\oplus\cdots\oplus\Lambda_{\Delta_{p}(\tau)}^{*},

in which ΛΔ1(τ)ΛΔp(τ)\Lambda_{\Delta_{1}(\tau)}^{*}\oplus\cdots\oplus\Lambda_{\Delta_{p}(\tau)}^{*} can be treated as a direct summand as in §2.2. So we can consider the pullback of the fan 𝒩Δ1(τ)××𝒩Δp(τ)\mathscr{N}_{\Delta_{1}(\tau)}\times\cdots\times\mathscr{N}_{\Delta_{p}(\tau)} via the map 𝚒τ\mathtt{i}_{\tau}, and realize 𝒩τΛτ,\mathscr{N}_{\tau}\subset\Lambda_{\tau,\real}^{*} as a refinement of this fan. Similarly, we have 𝚒ˇτ=𝚒ˇτ,1𝚒ˇτ,p:𝒬τΛΔˇ1(τ)ΛΔˇp(τ)\check{\mathtt{i}}_{\tau}=\check{\mathtt{i}}_{\tau,1}\oplus\cdots\oplus\check{\mathtt{i}}_{\tau,p}\colon\mathscr{Q}_{\tau}^{*}\rightarrow\Lambda^{*}_{\check{\Delta}_{1}(\tau)}\oplus\cdots\oplus\Lambda^{*}_{\check{\Delta}_{p}(\tau)} and the fan 𝒩Δˇ1(τ)××𝒩Δˇp(τ)\mathscr{N}_{\check{\Delta}_{1}(\tau)}\times\cdots\times\mathscr{N}_{\check{\Delta}_{p}(\tau)} in 𝒬τ,\mathscr{Q}_{\tau,\real}^{*} under pullback via 𝚒ˇτ\check{\mathtt{i}}_{\tau}. The intersection 𝒮eintre(τ)\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\tau) can be described by replacing ρ𝒩τ\rho\in\mathscr{N}_{\tau} with the condition ρ𝚒τ1(𝒩Δ1(τ)××𝒩Δp(τ))\rho\in\mathtt{i}_{\tau}^{-1}(\mathscr{N}_{\Delta_{1}(\tau)}\times\cdots\times\mathscr{N}_{\Delta_{p}(\tau)}), with a stratum denoted by 𝒮e,τ,ρ\mathscr{S}_{e,\tau,\rho}. This gives a stratification on 𝒮e\mathscr{S}_{e}.

Lemma 3.4.

For ωτ\omega\subset\tau and a stratum 𝒮e,τ,ρ\mathscr{S}_{e,\tau,\rho} in intre(τ)\mathrm{int}_{\mathrm{re}}(\tau), the intersection of the closure 𝒮e,τ,ρ¯\overline{\mathscr{S}_{e,\tau,\rho}} in BB with intre(ω)\mathrm{int}_{\mathrm{re}}(\omega) is a union of strata of 𝒮e\mathscr{S}_{e} in intre(ω)\mathrm{int}_{\mathrm{re}}(\omega).

Proof.

Given ωτ\omega\subset\tau, we take a change of coordinate map \gimel together with a neighborhood WW as in the proof of Lemma 3.3. We need to show that W𝒮τ,ρ=((intre(ρ)+cτ,1)×(aintre(Kωτ))×{o})W\cap\mathscr{S}_{\tau,\rho}=\gimel((\mathrm{int}_{\mathrm{re}}(\rho)+c_{\tau,1})\times(a-\mathrm{int}_{\mathrm{re}}(K_{\omega}\tau^{\vee}))\times\{o\}) for some cone ρ𝚒τ1(i=1p𝒩Δi(τ))\rho\in\mathtt{i}_{\tau}^{-1}(\prod_{i=1}^{p}\mathscr{N}_{\Delta_{i}(\tau)}). Let Δ1(τ),,Δr(τ),,Δp(τ)\Delta_{1}(\tau),\dots,\Delta_{r}(\tau),\dots,\Delta_{p}(\tau) be the monodromy polytopes of τ\tau, and Δ1(ω),,Δr(ω),,Δp(ω)\Delta_{1}(\omega),\dots,\Delta_{r}(\omega),\dots,\Delta_{p^{\prime}}(\omega) be those of ω\omega such that Δj(ω)\Delta_{j}(\omega) is the face of Δj(τ)\Delta_{j}(\tau) parallel to Λω\Lambda_{\omega} for j=1,,rj=1,\dots,r. Then we have direct sum decompositions ΛΔ1(τ)ΛΔp(τ)Aτ=Λτ\Lambda_{\Delta_{1}(\tau)}\oplus\cdots\oplus\Lambda_{\Delta_{p}(\tau)}\oplus A_{\tau}=\Lambda_{\tau} and ΛΔ1(ω)ΛΔp(ω)Aω=Λω\Lambda_{\Delta_{1}(\omega)}\oplus\cdots\oplus\Lambda_{\Delta_{p^{\prime}}(\omega)}\oplus A_{\omega}=\Lambda_{\omega}. We can further choose an inclusion

𝖺ωτ:ΛΔr+1(ω)ΛΔp(ω)AωAτ;\mathsf{a}_{\omega\tau}\colon\Lambda_{\Delta_{r+1}(\omega)}\oplus\cdots\oplus\Lambda_{\Delta_{p^{\prime}}(\omega)}\oplus A_{\omega}\hookrightarrow A_{\tau};

in other words, for every j=r+1,,pj=r+1,\dots,p^{\prime}, any fRj𝒫n1(ω)f\in R_{j}\subset\mathscr{P}_{n-1}(\omega) in Definition 2.9 is not containing τ\tau. For every j=r+1,,pj=r+1,\dots,p and any fRj𝒫n1(τ)f\in R_{j}\subset\mathscr{P}_{n-1}(\tau), the element mv1v2fm^{f}_{v_{1}v_{2}} is zero for any two vertices v1,v2v_{1},v_{2} of ω\omega. We have the identification

Λτ/Λω=j=1r(ΛΔj(τ)/ΛΔj(ω))l=r+1pΛΔl(τ)coker(𝖺ωτ).\Lambda_{\tau}/\Lambda_{\omega}=\bigoplus_{j=1}^{r}(\Lambda_{\Delta_{j}(\tau)}/\Lambda_{\Delta_{j}(\omega)})\oplus\bigoplus_{l=r+1}^{p}\Lambda_{\Delta_{l}(\tau)}\oplus\mathrm{coker}(\mathsf{a}_{\omega\tau}).

As a result, any cone 𝚒τ1(j=1pρj)𝚒τ1(i=1p𝒩Δi(τ))\mathtt{i}^{-1}_{\tau}(\prod_{j=1}^{p}\rho_{j})\in\mathtt{i}^{-1}_{\tau}\big{(}\prod_{i=1}^{p}\mathscr{N}_{\Delta_{i}(\tau)}\big{)} of codimension greater than 0 intersecting 1(W)\gimel^{-1}(W) will be a pullback of a cone under the projection to ΛΔ1(τ),ΛΔr(τ),\Lambda_{\Delta_{1}(\tau),\real}^{*}\oplus\cdots\oplus\Lambda_{\Delta_{r}(\tau),\real}^{*}. Consider the commutative diagram of projection maps

(3.8) Λω,\textstyle{\Lambda_{\omega,\real}^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚙ω\scriptstyle{\mathtt{p}_{\omega}}Λτ,\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\Lambda_{\tau,\real}^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚙ωτ\scriptstyle{\mathtt{p}_{\omega\subset\tau}}𝚙τ\scriptstyle{\mathtt{p}_{\tau}}j=1rΛΔj(ω),\textstyle{\prod_{j=1}^{r}\Lambda_{\Delta_{j}(\omega),\real}^{*}}j=1rΛΔj(τ),,\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\prod_{j=1}^{r}\Lambda_{\Delta_{j}(\tau),\real}^{*},}Πωτ\scriptstyle{\Pi_{\omega\subset\tau}}

we see that, in the open subset 1(W)\gimel^{-1}(W), every cone of codimension greater than 0 coming from pullback via 𝚙τ\mathtt{p}_{\tau} is a further pullback via Πωτ𝚙τ\Pi_{\omega\subset\tau}\circ\mathtt{p}_{\tau}. As a consequence, it must be of the form ((intre(ρ)+cτ,1)×(aintre(Kωτ))×{o})\gimel((\mathrm{int}_{\mathrm{re}}(\rho)+c_{\tau,1})\times(a-\mathrm{int}_{\mathrm{re}}(K_{\omega}\tau^{\vee}))\times\{o\}) in WW. ∎

3.3.1. Contraction of 𝒜\mathcal{A} to 𝒮\mathscr{S}

We would like to relate the amoeba 𝒜=μ(Z)\mathcal{A}=\mu(Z) with the tropical singular locus 𝒮\mathscr{S} introduced above.

Assumption 3.5.

We assume the existence of a surjective contraction map 𝒞:BB\mathscr{C}\colon B\rightarrow B which is isotopic to the identity and satisfies the following conditions:

  1. (1)

    𝒞1(B𝒮)(B𝒮)\mathscr{C}^{-1}(B\setminus\mathscr{S})\subset(B\setminus\mathscr{S}) and the restriction 𝒞|𝒞1(B𝒮):𝒞1(B𝒮)B𝒮\mathscr{C}|_{\mathscr{C}^{-1}(B\setminus\mathscr{S})}\colon\mathscr{C}^{-1}(B\setminus\mathscr{S})\to B\setminus\mathscr{S} is a homeomorphism.

  2. (2)

    𝒞\mathscr{C} maps 𝒜\mathcal{A} into the essential singular locus 𝒮e\mathscr{S}_{e}.

  3. (3)

    For each τ𝒫\tau\in\mathscr{P}, we have 𝒞1(intre(τ))intre(τ)\mathscr{C}^{-1}(\mathrm{int}_{\mathrm{re}}(\tau))\subset\mathrm{int}_{\mathrm{re}}(\tau).

  4. (4)

    For each τ𝒫\tau\in\mathscr{P} with 0<dim(τ)<n0<\dim(\tau)<n, we have a decomposition

    τ𝒞1(B𝒮)=vτ[0]τv\tau\cap\mathscr{C}^{-1}(B\setminus\mathscr{S})=\bigcup_{v\in\tau^{[0]}}\tau_{v}

    of the intersection τ𝒞1(B𝒮)\tau\cap\mathscr{C}^{-1}(B\setminus\mathscr{S}) into connected components τv\tau_{v}’s, where each τv\tau_{v} is contractible and is the unique component containing the vertex vτv\in\tau.

  5. (5)

    For each τ𝒫\tau\in\mathscr{P} and each point xintre(τ)𝒮x\in\mathrm{int}_{\mathrm{re}}(\tau)\cap\mathscr{S}, 𝒞1(x)intre(τ)\mathscr{C}^{-1}(x)\subset\mathrm{int}_{\mathrm{re}}(\tau) is a connected compact subset.

  6. (6)

    For each τ𝒫\tau\in\mathscr{P} and each point xintre(τ)𝒮x\in\mathrm{int}_{\mathrm{re}}(\tau)\cap\mathscr{S}, there exists a local base x\mathscr{B}_{x} around xx such that (𝒞μ)1(W)V(τ)(\mathscr{C}\circ\mu)^{-1}(W)\subset V(\tau) is Stein for every WxW\in\mathscr{B}_{x}, and for any U𝒞1(x)U\supset\mathscr{C}^{-1}(x), we have 𝒞1(W)U\mathscr{C}^{-1}(W)\subset U for sufficiently small WxW\in\mathscr{B}_{x}.

Similar contraction maps appear in [43, Rem. 2.4] (see also [45, 44]).

When dim(B)=2\dim(B)=2, we can take 𝒞=id\mathscr{C}=\mathrm{id} because from [27, Ex. 1.62], we see that ZZ is a finite collection of points, with at most one point lying in each closed stratum Xτ0\prescript{0}{}{X}_{\tau}, and the amoeba 𝒜\mathcal{A} is exactly the image of ZZ under the generalized moment map μ\mu.

When dim(B)=3\dim(B)=3, the amoeba 𝒜\mathcal{A} can possibly be of codimension one and we need to construct a contraction map as shown in Figure 4.

Refer to caption
Figure 4. Contraction map 𝒞\mathscr{C} when dim(B)=3\dim(B)=3

For dim(τ)=1\dim(\tau)=1, again from [27, Ex. 1.62], we see that if 𝒜intre(τ)\mathcal{A}\cap\mathrm{int}_{\mathrm{re}}(\tau)\neq\emptyset, then there is exactly one Ω1\Omega_{1} and R1R_{1}, and Δ1(τ)\Delta_{1}(\tau) is a line segment of affine length 11. In this case, ZXτ0Z\cap\prescript{0}{}{X}_{\tau} consists of only one point, given by the intersection of the zero locus svτ1(fvρ)s_{v\tau}^{-1}(f_{v\rho}) with Vτ(τ)V(τ)\mathbb{C}^{*}\cong V_{\tau}(\tau)\subset V(\tau). Taking mm to be the primitive vector in Λτ\Lambda_{\tau} starting at vv that points into τ\tau, we can write svτ1(fvρ)=1+svτ1(m)zms_{v\tau}^{-1}(f_{v\rho})=1+s_{v\tau}^{-1}(m)z^{m}. Applying the log map log:\log\colon\mathbb{C}^{*}\rightarrow\real, we see that 𝒜intre(τ)=cτ\mathcal{A}\cap\mathrm{int}_{\mathrm{re}}(\tau)=c_{\tau}. Therefore, for an edge τ𝒫[1]\tau\in\mathscr{P}^{[1]}, we can define 𝒞\mathscr{C} to be the identity on τ\tau.

Refer to caption
Figure 5. Contraction at ρ\rho

On a codimension one cell ρ\rho such that intre(ρ)𝒜\mathrm{int}_{\mathrm{re}}(\rho)\cap\mathcal{A}\neq\emptyset (see Figure 5), we consider the log map log:Specan([Λρ])()2Λρ,2\log\colon\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Lambda_{\rho}])\cong(\mathbb{C}^{*})^{2}\rightarrow\Lambda_{\rho,\real}^{*}\cong{}^{2}, and take a sufficiently large polytope 𝙿\mathtt{P} (colored purple in Figure 5) so that 𝒜intre(𝙿)\mathcal{A}\setminus\mathrm{int}_{\mathrm{re}}(\mathtt{P}) is a disjoint union of legs. We first contract each leg to the tropical singular locus (colored blue in Figure 5) along the normal direction to the tropical singular locus. Next, we contract the polytope 𝙿\mathtt{P} to the 0-dimensional stratum of 𝒮e\mathscr{S}_{e}. Notice that the restriction of 𝒞\mathscr{C} to the tropical singular locus 𝒮\mathscr{S} is not the identity but rather a contraction onto itself. Once the contraction map is constructed for all codimension one cells ρ\rho, we can then extend it continuously to the whole of BB so that it is a diffeomorphism on intre(σ)\mathrm{int}_{\mathrm{re}}(\sigma) for every maximal cell σ\sigma. The map is chosen such that the preimage 𝒞1(x)\mathscr{C}^{-1}(x) for every point xintre(ρ)x\in\mathrm{int}_{\mathrm{re}}(\rho) is a convex polytope in 2. Therefore, given any open subset U2U\subset{}^{2} which contains 𝒞1(x)\mathscr{C}^{-1}(x), we can find some convex open neighborhood W1UW_{1}\subset U of 𝒞1(x)\mathscr{C}^{-1}(x) giving the Stein open subset log1(W1)()2\log^{-1}(W_{1})\subset(\mathbb{C}^{*})^{2}. By taking W=W1×W2W=W_{1}\times W_{2} in the chart Λρ,×𝒬ρ,\Lambda_{\rho,\real}^{*}\times\mathscr{Q}_{\rho,\real} as in the proof of Lemma 3.1, we have the open subset WW that satisfies condition (5) in Assumption 3.5.

In general, we need to construct 𝒞|intre(τ)\mathscr{C}|_{\mathrm{int}_{\mathrm{re}}(\tau)} inductively for each τ𝒫\tau\in\mathscr{P}, so that the preimage 𝒞1(x)intre(τ)\mathscr{C}^{-1}(x)\subset\mathrm{int}_{\mathrm{re}}(\tau) is convex in the chart Λτ,intre(τ)\Lambda_{\tau,\real}^{*}\cong\mathrm{int}_{\mathrm{re}}(\tau) and the codimension one amoeba 𝒜\mathcal{A} is contracted to the codimension 2 tropical singular locus 𝒮e\mathscr{S}_{e}. The reason for introducing such a contraction map is that we can modify the generalized moment map μ\mu to one which is more closely related with tropical geometry:

Definition 3.6.

We call the composition ν:=𝒞μ:X0B\nu:=\mathscr{C}\circ\mu\colon\prescript{0}{}{X}\rightarrow B the modified moment map.

One immediate consequence of property (6)(6) in Assumption 3.5 is that we have Rν()=ν()R\nu_{*}(\mathcal{F})=\nu_{*}(\mathcal{F}) for any coherent sheaf \mathcal{F} on X0\prescript{0}{}{X}, thanks to Lemma 3.1 and Cartan’s Theorem B:

Theorem 3.7 (Cartan’s Theorem B [6]; see e.g. Ch. IX, Cor. 4.11 in [13]).

For any coherent sheaf \mathcal{F} over a Stein space UU, we have H>0(U,)=0.H^{>0}(U,\mathcal{F})=0.

3.3.2. Monodromy invariant differential forms on BB

Outside of the essential singular locus 𝒮e\mathscr{S}_{e}, we have a nice integral affine manifold B𝒮eB\setminus\mathscr{S}_{e}, on which we can talk about the sheaf Ω\Omega^{*} of (-valued) de Rham differential forms. But in fact, we can extend its definition to 𝒮e\mathscr{S}_{e} as well using monodromy invariant differential forms.

We consider the inclusion ι:B0:=B𝒮eB\iota\colon B_{0}:=B\setminus\mathscr{S}_{e}\rightarrow B and the natural exact sequence

(3.9) 0¯𝒜𝑓𝑓ιΛB00,0\rightarrow\underline{\mathbb{Z}}\rightarrow\mathcal{A}\mathit{ff}\rightarrow\iota_{*}\Lambda_{B_{0}}^{*}\rightarrow 0,

where ΛB0\Lambda_{B_{0}}^{*} denotes the sheaf of integral cotangent vectors on B0B_{0}. For any τ𝒫\tau\in\mathscr{P}, the stalk (ιΛB0)x(\iota_{*}\Lambda_{B_{0}}^{*})_{x} at a point xintre(τ)𝒮ex\in\mathrm{int}_{\mathrm{re}}(\tau)\cap\mathscr{S}_{e} can be described using the chart Υτ\Upsilon_{\tau} in (3.4). Using the description in §3.3, we have x𝒮e,τ,ρ=intre(ρ)×{o}x\in\mathscr{S}_{e,\tau,\rho}=\mathrm{int}_{\mathrm{re}}(\rho)\times\{o\} for some ρ𝚒τ1(𝒩Δ1(τ)××𝒩Δp(τ))\rho\in\mathtt{i}_{\tau}^{-1}(\mathscr{N}_{\Delta_{1}(\tau)}\times\cdots\times\mathscr{N}_{\Delta_{p}(\tau)}). Taking a vertex vτv\in\tau, we can consider the monodromy transformations TγT_{\gamma}’s around the strata 𝒮e,η,ρ\mathscr{S}_{e,\eta,\rho}’s that contain xx in their closures. We can identify the stalk ι(ΛB0)x\iota_{*}(\Lambda_{B_{0}}^{*})_{x} as the subset of invariant elements of TvT_{v}^{*} under all such monodromy transformations. Since ρΛτ,\rho\subset\Lambda_{\tau,\real}^{*} is a cone, we have ΛρΛτ\Lambda_{\rho}\subset\Lambda_{\tau}^{*}. Using the natural projection map πvτ:TvΛτ\pi_{v\tau}\colon T_{v}^{*}\rightarrow\Lambda_{\tau}^{*}, we have the identification ι(ΛB0)xπvτ1(Λρ)\iota_{*}(\Lambda_{B_{0}}^{*})_{x}\cong\pi_{v\tau}^{-1}(\Lambda_{\rho}). There is a direct sum decomposition ι(ΛB0)x=Λρ𝒬τ\iota_{*}(\Lambda_{B_{0}}^{*})_{x}=\Lambda_{\rho}\oplus\mathscr{Q}_{\tau}^{*}, depending on a decomposition Tv=Λτ𝒬τT_{v}=\Lambda_{\tau}\oplus\mathscr{Q}_{\tau}. This gives the map

(3.10) 𝚡:Uxπvτ1(Λρ)\mathtt{x}\colon U_{x}\rightarrow\pi_{v\tau}^{-1}(\Lambda_{\rho})^{*}

in a sufficiently small neighborhood UxU_{x}, locally defined up to a translation in πvτ1(Λρ)\pi_{v\tau}^{-1}(\Lambda_{\rho})^{*}. We need to describe the compatibility between the map associated to a point x𝒮e,ω,ρx\in\mathscr{S}_{e,\omega,\rho} and that to a point x~𝒮e,τ,ρ~\tilde{x}\in\mathscr{S}_{e,\tau,\tilde{\rho}} such that 𝒮e,ω,ρ𝒮e,τ,ρ~¯\mathscr{S}_{e,\omega,\rho}\subset\overline{\mathscr{S}_{e,\tau,\tilde{\rho}}}.

The first case is when ω=τ\omega=\tau. We let x~intre(ρ~)×{o}Ux\tilde{x}\in\mathrm{int}_{\mathrm{re}}(\tilde{\rho})\times\{o\}\cap U_{x} for some ρρ~\rho\subset\tilde{\rho}. Then, after choosing suitable translations in πvτ1(Λρ)\pi_{v\tau}^{-1}(\Lambda_{\rho})^{*} for the maps 𝚡\mathtt{x} and 𝚡~\tilde{\mathtt{x}}, we have the following commutative diagram:

(3.11) Ux~Ux\textstyle{U_{\tilde{x}}\cap U_{x}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚡~\scriptstyle{\tilde{\mathtt{x}}}πvτ1(Λρ~)\textstyle{\pi_{v\tau}^{-1}(\Lambda_{\tilde{\rho}})^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚙\scriptstyle{\mathtt{p}}Ux\textstyle{U_{x}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚡\scriptstyle{\mathtt{x}}πvτ1(Λρ).\textstyle{\pi_{v\tau}^{-1}(\Lambda_{\rho})^{*}.}

The second case is when ωτ\omega\subsetneq\tau. Making use of the change of charts \gimel in equation (3.6), and the description in the proof of Lemma 3.4, we write

x~intre(ρ~)×{o}\tilde{x}\in\mathrm{int}_{\mathrm{re}}(\tilde{\rho})\times\{o\}

for some cone ρ~=𝚒τ1(j=1pρ~j)𝚒τ1(j=1pΛΔj(τ))\tilde{\rho}=\mathtt{i}_{\tau}^{-1}(\prod_{j=1}^{p}\tilde{\rho}_{j})\in\mathtt{i}_{\tau}^{-1}\big{(}\prod_{j=1}^{p}\Lambda_{\Delta_{j}(\tau)}^{*}\big{)} of positive codimension. In 1(W)\gimel^{-1}(W), we may assume ρ~\tilde{\rho} is the pullback of a cone ρ˘\breve{\rho} via Πωτ𝚙τ\Pi_{\omega\subset\tau}\circ\mathtt{p}_{\tau} as in equation (3.8). Since 𝒮e,ω,ρ𝒮e,τ,ρ~¯\mathscr{S}_{e,\omega,\rho}\subset\overline{\mathscr{S}_{e,\tau,\tilde{\rho}}}, we have ρ𝚙ω1(ρ˘)\rho\subset\mathtt{p}_{\omega}^{-1}(\breve{\rho}) and hence 𝚙ωτ1(Λρ)Λρ~\mathtt{p}_{\omega\subset\tau}^{-1}(\Lambda_{\rho})\subset\Lambda_{\tilde{\rho}}. Therefore, from 𝚙ωτπvτ=πvω\mathtt{p}_{\omega\subset\tau}\circ\pi_{v\tau}=\pi_{v\omega}, we obtain πvω1(Λρ)πvτ1(Λρ~)\pi_{v\omega}^{-1}(\Lambda_{\rho})\subset\pi_{v\tau}^{-1}(\Lambda_{\tilde{\rho}}), inducing the map 𝚙:πvτ1(Λρ~)πvω1(Λρ)\mathtt{p}\colon\pi_{v\tau}^{-1}(\Lambda_{\tilde{\rho}})^{*}\rightarrow\pi_{v\omega}^{-1}(\Lambda_{\rho})^{*}. As a result, we still have the commutative diagram (3.11) for a point x~\tilde{x} sufficiently close to xx.

Definition 3.8.

Given x𝒮ex\in\mathscr{S}_{e} as above, the stalk of Ω\Omega^{*} at xx is defined as the stalk Ωx:=(𝚡1Ω)x\Omega^{*}_{x}:=(\mathtt{x}^{-1}\Omega^{*})_{x} of the pullback of the sheaf of smooth de Rham forms on πvτ1(Λρ)\pi_{v\tau}^{-1}(\Lambda_{\rho})^{*}, which is equipped with the de Rham differential dd. This defines the complex (Ω,d)(\Omega^{*},d) of monodromy invariant smooth differential forms on BB. A section αΩ(W)\alpha\in\Omega^{*}(W) is a collection of elements αxΩx\alpha_{x}\in\Omega^{*}_{x}, xWx\in W such that each αx\alpha_{x} can be represented by 𝚡1βx\mathtt{x}^{-1}\beta_{x} in a small neighborhood Ux𝚙1(𝚄x)U_{x}\subset\mathtt{p}^{-1}(\mathtt{U}_{x}) for some smooth form βx\beta_{x} on 𝚄x\mathtt{U}_{x}, and satisfies the relation αx~=𝚡~1(𝚙βx)\alpha_{\tilde{x}}=\tilde{\mathtt{x}}^{-1}(\mathtt{p}^{*}\beta_{x}) in Ωx~\Omega^{*}_{\tilde{x}} for every x~Ux\tilde{x}\in U_{x}.

Example 3.9.

In the 22-dimensional case in Example 2.11, we consider a singular point

{x}=𝒮eintre(τ)\{x\}=\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\tau)

for some τ𝒫[1]\tau\in\mathscr{P}^{[1]}. In this case, we can take ρ\rho to be the 0-dimenisonal stratum in 𝒩τ=𝚒τ1(𝒩Δ1(τ))\mathscr{N}_{\tau}=\mathtt{i}_{\tau}^{-1}(\mathscr{N}_{\Delta_{1}(\tau)}) and we have ι(ΛB0)x=𝒬τ\iota_{*}(\Lambda_{B_{0}}^{*})_{x}=\mathscr{Q}_{\tau}^{*}. Taking a generator of 𝒬τ\mathscr{Q}_{\tau}^{*}, we get an invariant affine coordinate 𝚡:Ux\mathtt{x}\colon U_{x}\rightarrow\real which is the normal affine coordinate of τ\tau. The stalk Ωx\Omega^{*}_{x} is then identified with the pullback of the space of germs of smooth differential forms from (,0)(\real,0) via 𝚡\mathtt{x}. In particular, Ωx2=0\Omega^{2}_{x}=0.

For the YY-vertex of type II in Example 2.12, the situation is similar to the 22-dimensional case. For {x}=𝒮eintre(τ)\{x\}=\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\tau), we still have ι(ΛB0)x=𝒬τ\iota_{*}(\Lambda_{B_{0}}^{*})_{x}=\mathscr{Q}_{\tau}^{*}, and in this case, 𝚡:Ux2\mathtt{x}\colon U_{x}\rightarrow{}^{2} are the two invariant affine coordinates. We can identify Ωx\Omega^{*}_{x} as the pullback of the space of germs of smooth differential forms from (,20)({}^{2},0) via 𝚡\mathtt{x}.

For the YY-vertex of type I in Example 2.12, we use the identification Λτ,intre(τ)\Lambda_{\tau,\real}^{*}\cong\mathrm{int}_{\mathrm{re}}(\tau) via Υτ\Upsilon_{\tau} for the 22-dimensional cell τ\tau separating two maximal cells σ+\sigma_{+} and σ\sigma_{-}. In this case, 𝒮e\mathscr{S}_{e} is as shown (in blue color) in Figure 5 and 𝒩=𝚒τ1(𝒩Δ1(τ))\mathscr{N}=\mathtt{i}_{\tau}^{-1}(\mathscr{N}_{\Delta_{1}(\tau)}) is the fan of 2\mathbb{P}^{2}. If xx is the 0-dimensional stratum of 𝒮eintre(τ)\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\tau), we have ι(ΛB0)x=𝒬τ\iota_{*}(\Lambda_{B_{0}}^{*})_{x}=\mathscr{Q}_{\tau}^{*} and 𝚡:Ux\mathtt{x}\colon U_{x}\rightarrow\real as an invariant affine coordinate. If xx is a point on a leg of the YY-vertex, we have 𝚡=(𝚡1,𝚡2):Ux2\mathtt{x}=(\mathtt{x}_{1},\mathtt{x}_{2})\colon U_{x}\rightarrow{}^{2} with 𝚡1\mathtt{x}_{1} coming from a generator of Λρ\Lambda_{\rho} and 𝚡2\mathtt{x}_{2} coming from a generator of 𝒬τ\mathscr{Q}_{\tau}^{*}.

It follows from the definition that ¯Ω\underline{\real}\rightarrow\Omega^{*} is a resolution. We shall also prove the existence of a partition of unity.

Lemma 3.10.

Given any xBx\in B and a sufficiently small neighborhood UU, there exists ϱΩ0(U)\varrho\in\Omega^{0}(U) with compact support in UU such that 0ϱ10\leq\varrho\leq 1 and ϱ1\varrho\equiv 1 near xx. (Since Ω0\Omega^{0} is a subsheaf of the sheaf 𝒞0\mathcal{C}^{0} of continuous functions on BB, we can talk about the value f(x)f(x) for fΩ0(W)f\in\Omega^{0}(W) and xWx\in W.)

Proof.

If x𝒮ex\notin\mathscr{S}_{e}, the statement is a standard fact. So we assume that xintre(τ)𝒮ex\in\mathrm{int}_{\mathrm{re}}(\tau)\cap\mathscr{S}_{e} for some τ𝒫\tau\in\mathscr{P}. As above, we can write xintre(ρ)×{o}x\in\mathrm{int}_{\mathrm{re}}(\rho)\times\{o\}. Furthermore, since ρ\rho is a cone in the fan 𝚒τ1(𝒩Δ1(τ)××𝒩Δp(τ))\mathtt{i}_{\tau}^{-1}(\mathscr{N}_{\Delta_{1}(\tau)}\times\cdots\times\mathscr{N}_{\Delta_{p}(\tau)}), Λτ\Lambda_{\tau}^{*} has ΛΔ1(τ)ΛΔp(τ)\Lambda_{\Delta_{1}(\tau)}^{*}\oplus\cdots\oplus\Lambda_{\Delta_{p}(\tau)}^{*} as a direct summand, and the description of ι(ΛB0)x\iota_{*}(\Lambda_{B_{0}}^{*})_{x} is compatible with the direct sum decomposition of Λτ\Lambda_{\tau}^{*}. We may further assume that p=1p=1 and τ=Δ1(τ)\tau=\Delta_{1}(\tau) is a simplex.

If ρ\rho is not the smallest cone (i.e. the one consisting of just the origin in 𝒩τ\mathscr{N}_{\tau}), we have a decomposition Λτ=Λρ𝒬ρ\Lambda_{\tau}^{*}=\Lambda_{\rho}\oplus\mathscr{Q}_{\rho} and the natural projection 𝚙:Λτ𝒬ρ\mathtt{p}\colon\Lambda_{\tau}^{*}\rightarrow\mathscr{Q}_{\rho}. Then, locally near x0x_{0}, we can write the normal fan 𝒩τ\mathscr{N}_{\tau} as 𝚙1(Σρ)\mathtt{p}^{-1}(\Sigma_{\rho}) for some normal fan Σρ𝒬ρ\Sigma_{\rho}\subset\mathscr{Q}_{\rho} of a lower dimensional simplex. For any vector vv tangent to ρ\rho at x0x_{0} and the corresponding affine function lvl_{v} locally near x0x_{0}, we always have lvv>0\frac{\partial l_{v}}{\partial{v}}>0. This allows us to construct a bump function ϱ=vi(lvi(x)lvi(x0))2\varrho=\sum_{v_{i}}(l_{v_{i}}(x)-l_{v_{i}}(x_{0}))^{2} along the Λρ,\Lambda_{\rho,\real}-direction. So we are reduced to the case when ρ={o}\rho=\{o\} is the smallest cone in the fan 𝒩τ\mathscr{N}_{\tau}.

Now we construct the function ϱ\varrho near the origin o𝒩τo\in\mathscr{N}_{\tau} by induction on the dimension of the fan 𝒩τ\mathscr{N}_{\tau}. When dim(𝒩τ)=1\dim(\mathscr{N}_{\tau})=1, it is the fan of 1\mathbb{P}^{1} consisting of three cones -, {o}\{o\} and +. One can construct the bump function which is equal to 11 near oo and supported in a sufficiently small neighborhood of oo. For the induction step, we consider an nn-dimensional fan 𝒩τ\mathscr{N}_{\tau}. For any point xx near but not equal to oo, we have xintre(ρ)x\in\mathrm{int}_{\mathrm{re}}(\rho) for some ρ{o}\rho\neq\{o\}. Then we can decompose 𝒩τ\mathscr{N}_{\tau} locally as Λρ𝒬ρ\Lambda_{\rho}\oplus\mathscr{Q}_{\rho}. Applying the induction hypothesis to 𝒬ρ\mathscr{Q}_{\rho} gives a bump function ϱx\varrho_{x} compactly supported in any sufficiently small neighborhood of xx (for the Λρ\Lambda_{\rho} directions, we do not need the induction hypothesis to get the bump function). This produces a partition of unity {ϱi}\{\varrho_{i}\} outside oo. Finally, letting ϱ:=1iϱi\varrho:=1-\sum_{i}\varrho_{i} and extending it continuously to the origin oo gives the desired function. ∎

Lemma 3.10 produces a partition of unity for the complex (Ω,d)(\Omega^{*},d) of monodromy invariant differential forms on BB, which satisfies the requirement in Condition 4.7 below. In particular, the cohomology of (Ω(B),d)(\Omega^{*}(B),d) computes RΓ(B,¯)R\Gamma(B,\underline{\real}). Given a point xB𝒮ex\in B\setminus\mathscr{S}_{e}, we can take an element ϱxΩn(B)\varrho_{x}\in\Omega^{n}(B), compactly supported in an arbitrarily small neighborhood UxB𝒮eU_{x}\subset B\setminus\mathscr{S}_{e}, to represent a non-zero element in the cohomology Hn(Ω,d)=Hn(B,)H^{n}(\Omega^{*},d)=H^{n}(B,\mathbb{C})\cong\mathbb{C}.

4. Smoothing of maximally degenerate Calabi–Yau varieties via dgBV algebras

In this section, we review and refine the results in [8] concerning smoothing of the maximally degenerate Calabi–Yau log variety X0\prescript{0}{}{X}^{\dagger} over S^=Specan(R^)=Specan([[q]])\hat{S}^{\dagger}=\mathrm{Spec}_{\mathrm{an}}(\hat{R})^{\dagger}=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[[q]])^{\dagger} using the local smoothing models V𝕍kV^{\dagger}\rightarrow\prescript{k}{}{\mathbb{V}}^{\dagger}’s specified in §2.4. In order to relate with tropical geometry on BB, we will choose VV so that it is the pre-image ν1(W)\nu^{-1}(W) of an open subset WW in BB.

4.1. Good covers and local smoothing data

Given τ𝒫\tau\in\mathscr{P} and a point xintre(τ)Bx\in\mathrm{int}_{\mathrm{re}}(\tau)\subset B, we take a sufficiently small open subset WxW\in\mathscr{B}_{x}. We need to construct a local smoothing model on the Stein open subset V=ν1(W)V=\nu^{-1}(W).

  • If x𝒮ex\notin\mathscr{S}_{e}, then we can simply take the local smoothing 𝕍\prescript{}{}{\mathbb{V}}^{\dagger} introduced in (2.14) in §2.4.

  • If x𝒮eintre(τ)x\in\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\tau), we assume that 𝒞1(W)𝒜iτ\mathscr{C}^{-1}(W)\cap\mathcal{A}^{\tau}_{i}\neq\emptyset for i=1,,ri=1,\dots,r and 𝒞1(W)𝒜iτ=\mathscr{C}^{-1}(W)\cap\mathcal{A}^{\tau}_{i}=\emptyset for other ii’s. Note that 𝒞1(W)intre(τ)\mathscr{C}^{-1}(W)\cap\mathrm{int}_{\mathrm{re}}(\tau) may not be a small open subset in intre(τ)\mathrm{int}_{\mathrm{re}}(\tau) as we may contract a polytope 𝙿\mathtt{P} via 𝒞\mathscr{C} (Figure 5). If we write ΛΔ1(τ)ΛΔp(τ)Aτ=Λτ\Lambda_{\Delta_{1}(\tau)}\oplus\cdots\oplus\Lambda_{\Delta_{p}(\tau)}\oplus A_{\tau}=\Lambda_{\tau} as lattices, then for each direct summand ΛΔi(τ)\Lambda_{\Delta_{i}(\tau)}, we have a commutative diagram

    Λτ,\textstyle{\Lambda_{\tau,\mathbb{C}^{*}}^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚒τ,i,\scriptstyle{\mathtt{i}_{\tau,i,\mathbb{C}^{*}}}log\scriptstyle{\log}ΛΔi(τ),\textstyle{\Lambda_{\Delta_{i}(\tau),\mathbb{C}^{*}}^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}log\scriptstyle{\log}Λτ,\textstyle{\Lambda_{\tau,\real}^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚒τ,i,\scriptstyle{\mathtt{i}_{\tau,i,\real}}ΛΔi(τ),,\textstyle{\Lambda_{\Delta_{i}(\tau),\real}^{*},}

    so that both ZiτZ^{\tau}_{i} and 𝒜iτ\mathcal{A}^{\tau}_{i} are coming from pullbacks of some subsets under the projection maps 𝚒τ,i,\mathtt{i}_{\tau,i,\mathbb{C}^{*}} and 𝚒τ,i,\mathtt{i}_{\tau,i,\real} respectively. From this, we see that 𝒞1(W)𝒜1τ𝒜rτ\mathscr{C}^{-1}(W)\cap\mathcal{A}^{\tau}_{1}\cap\cdots\cap\mathcal{A}^{\tau}_{r}\neq\emptyset and ν1(W)Z1τZrτ\nu^{-1}(W)\cap Z^{\tau}_{1}\cap\cdots\cap Z^{\tau}_{r}\neq\emptyset while ν1(W)Ziτ=\nu^{-1}(W)\cap Z^{\tau}_{i}=\emptyset for other ii’s. Now we take ψx,i=ψi\psi_{x,i}=\psi_{i} for 1ir1\leq i\leq r and ψx,i=0\psi_{x,i}=0 otherwise accordingly. Then we can take Pτ,xP_{\tau,x} introduced in (2.17) and the map V=ν1(W)Specan([Στl])V=\nu^{-1}(W)\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}\oplus\mathbb{N}^{l}]) defined by

    (4.1) {zmhmzmif mΣτ;uifv,iif 1ir;uiziif r<il.\begin{cases}z^{m}\mapsto h_{m}\cdot z^{m}&\text{if }m\in\Sigma_{\tau};\\ u_{i}\mapsto f_{v,i}&\text{if }1\leq i\leq r;\\ u_{i}\mapsto z_{i}&\text{if }r<i\leq l.\end{cases}

    Note that the third line of this formula is different from that of equation (2.19) because we do not specify a point xZ1τZrτx\in Z^{\tau}_{1}\cap\cdots\cap Z^{\tau}_{r}. By shrinking WW if necessary, one can show that it is an embedding using an argument similar to [28, Thm. 2.6]. This is possible because we can check that the Jacobian appearing in the proof of [28, Thm. 2.6] is invertible for all point in ν1(x)=μ1(𝒞1(x))\nu^{-1}(x)=\mu^{-1}(\mathscr{C}^{-1}(x)), which is a connected compact subset by property (5)(5) in Assumption 3.5.

Condition 4.1.

An open cover {Wα}α\{W_{\alpha}\}_{\alpha} of BB is said to be good if

  1. (1)

    for each WαW_{\alpha}, there exists a unique τα𝒫\tau_{\alpha}\in\mathscr{P} such that WαxW_{\alpha}\in\mathscr{B}_{x} for some xintre(τ)x\in\mathrm{int}_{\mathrm{re}}(\tau);

  2. (2)

    Wαβ=WαWβW_{\alpha\beta}=W_{\alpha}\cap W_{\beta}\neq\emptyset only when τατβ\tau_{\alpha}\subset\tau_{\beta} or τβτα\tau_{\beta}\subset\tau_{\alpha}, and if this is the case, we have either intre(α)Wαβ\mathrm{int}_{\mathrm{re}}(\alpha)\cap W_{\alpha\beta}\neq\emptyset or intre(β)Wαβ\mathrm{int}_{\mathrm{re}}(\beta)\cap W_{\alpha\beta}\neq\emptyset.

Given a good cover {Wα}α\{W_{\alpha}\}_{\alpha} of BB, we have the corresponding Stein open cover 𝒱:={Vα}α\mathcal{V}:=\{V_{\alpha}\}_{\alpha} of X0\prescript{0}{}{X} given by Vα:=ν1(Wα)V_{\alpha}:=\nu^{-1}(W_{\alpha}) for each α\alpha. For each VαV_{\alpha}^{\dagger}, the infinitesimal local smoothing model is given as a log space 𝕍α\prescript{}{}{\mathbb{V}}_{\alpha}^{\dagger} over S^\hat{S}^{\dagger} (see (2.14)). Let 𝕍αk\prescript{k}{}{\mathbb{V}}_{\alpha} be the kthk^{\text{th}}-order thickening over Sk=Specan(R/𝐦k+1)\prescript{k}{}{S}^{\dagger}=\mathrm{Spec}_{\mathrm{an}}(R/\mathbf{m}^{k+1})^{\dagger} and j:VαZVαj\colon V_{\alpha}\setminus Z\hookrightarrow V_{\alpha} be the open inclusion. As in [8, §8], we obtain coherent sheaves of BV algebras (and modules) over VαV_{\alpha} from these local smoothing models. But for the purpose of this paper, we would like to push forward these coherent sheaves to BB and work with the open subsets WαW_{\alpha}’s. This leads to the following modification of [8, Def. 7.6] (see also [8, Def. 2.14 and 2.20]):

Definition 4.2.

For each k0k\in\mathbb{Z}_{\geq 0}, we define

  • the sheaf of kthk^{\text{th}}-order polyvector fields to be 𝒢αk:=νj(Θ𝕍αk/Sk)\prescript{k}{}{\mathcal{G}}_{\alpha}^{*}:=\nu_{*}j_{*}(\bigwedge^{-*}\Theta_{\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{k}{}{S}^{\dagger}}) (i.e. push-forward of relative log polyvector fields on 𝕍αk\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger});

  • the kthk^{\text{th}}-order log de Rham complex to be 𝒦αk:=νj(Ω𝕍αk/)\prescript{k}{}{\mathcal{K}}^{*}_{\alpha}:=\nu_{*}j_{*}(\Omega^{*}_{\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}/\mathbb{C}}) (i.e. push-forward of log de Rham differentials) equipped with the de Rham differential αk=\prescript{k}{}{\partial}_{\alpha}=\prescript{}{}{\partial} which is naturally a dg module over ΩSk\prescript{k}{}{\Omega}^{*}_{S^{\dagger}};

  • the local log volume form ωα\prescript{}{}{\omega}_{\alpha} as a nowhere vanishing element in νj(Ω𝕍α/S^n)\nu_{*}j_{*}(\Omega^{n}_{\prescript{}{}{\mathbb{V}}_{\alpha}^{\dagger}/\hat{S}^{\dagger}}) and the kthk^{\text{th}}-order volume form to be ωαk=ωα(mod 𝐦k+1)\prescript{k}{}{\omega}_{\alpha}=\prescript{}{}{\omega}_{\alpha}\ (\text{mod $\mathbf{m}^{k+1}$}).

Given k>lk>l, there are natural maps k,l:j(Θ𝕍αk/Sk)j(Θ𝕍αl/Sl)\prescript{k,l}{}{\flat}\colon j_{*}(\bigwedge^{-*}\Theta_{\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{k}{}{S}^{\dagger}})\rightarrow j_{*}(\bigwedge^{-*}\Theta_{\prescript{l}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{l}{}{S}^{\dagger}}) which induce the maps k,l:𝒢αk𝒢αl\prescript{k,l}{}{\flat}\colon\prescript{k}{}{\mathcal{G}}^{*}_{\alpha}\rightarrow\prescript{l}{}{\mathcal{G}}^{*}_{\alpha}. Before taking the push-forward μ\mu_{*}, each j(rΘ𝕍αk/Sk)j_{*}(\bigwedge^{r}\Theta_{\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{k}{}{S}^{\dagger}}) is a sheaf of flat Rk\prescript{k}{}{R}-modules with the property that j(rΘ𝕍αk/Sk)j(rΘ𝕍αk+1/Sk+1)Rk+1Rkj_{*}(\bigwedge^{r}\Theta_{\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{k}{}{S}^{\dagger}})\cong j_{*}(\bigwedge^{r}\Theta_{\prescript{k+1}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{k+1}{}{S}^{\dagger}})\otimes_{\prescript{k+1}{}{R}}\prescript{k}{}{R} by [17, Cor. 7.4 and 7.9]. In other words, we have a short exact sequence of coherent sheaves

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j(rΘ𝕍α0/S0)\textstyle{j_{*}(\bigwedge^{r}\Theta_{\prescript{0}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{0}{}{S}^{\dagger}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}qk+1\scriptstyle{\cdot q^{k+1}}j(rΘ𝕍αk+1/Sk+1)\textstyle{j_{*}(\bigwedge^{r}\Theta_{\prescript{k+1}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{k+1}{}{S}^{\dagger}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j(rΘ𝕍αk/Sk)\textstyle{j_{*}(\bigwedge^{r}\Theta_{\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{k}{}{S}^{\dagger}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

Applying μ\mu_{*}, which is exact, we get

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒢αr0\textstyle{\prescript{0}{}{\mathcal{G}}_{\alpha}^{-r}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}qk+1\scriptstyle{\cdot q^{k+1}}𝒢αrk+1\textstyle{\prescript{k+1}{}{\mathcal{G}}_{\alpha}^{-r}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒢αrk\textstyle{\prescript{k}{}{\mathcal{G}}_{\alpha}^{-r}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

As a result, we see that 𝒢αrk\prescript{k}{}{\mathcal{G}}_{\alpha}^{-r} is a sheaf of flat Rk\prescript{k}{}{R}-modules on WαW_{\alpha}, so we have 𝒢αrk+1Rk+1Rk𝒢αrk\prescript{k+1}{}{\mathcal{G}}_{\alpha}^{-r}\otimes_{\prescript{k+1}{}{R}}\prescript{k}{}{R}\cong\prescript{k}{}{\mathcal{G}}^{-r}_{\alpha} for each rr; a similar statement holds for 𝒦αrk\prescript{k}{}{\mathcal{K}}^{r}_{\alpha}.

A natural filtration 𝒦αk\prescript{k}{\bullet}{\mathcal{K}}^{*}_{\alpha} is given by 𝒦αsk:=ΩSsk𝒦αk[s]\prescript{k}{s}{\mathcal{K}}^{*}_{\alpha}:=\prescript{k}{}{\Omega}^{\geq s}_{S^{\dagger}}\wedge\prescript{k}{}{\mathcal{K}}^{*}_{\alpha}[s] and taking wedge product defines the natural sheaf isomorphism σ1rk:ΩSrkRk(𝒦α0k/𝒦α1k[r])𝒦αrk/𝒦αr+1k\prescript{k}{r}{\sigma}^{-1}\colon\prescript{k}{}{\Omega}^{r}_{S^{\dagger}}\otimes_{\prescript{k}{}{R}}(\prescript{k}{0}{\mathcal{K}}^{*}_{\alpha}/\prescript{k}{1}{\mathcal{K}}^{*}_{\alpha}[-r])\rightarrow\prescript{k}{r}{\mathcal{K}}^{*}_{\alpha}/\prescript{k}{r+1}{\mathcal{K}}^{*}_{\alpha}. We have the space 𝒦αk:=𝒦α0k/𝒦α1kνj(Ω𝐕αk/Sk)\prescript{k}{\parallel}{\mathcal{K}}^{*}_{\alpha}:=\prescript{k}{0}{\mathcal{K}}^{*}_{\alpha}/\prescript{k}{1}{\mathcal{K}}^{*}_{\alpha}\cong\nu_{*}j_{*}(\Omega^{*}_{\prescript{k}{}{\mathbf{V}}_{\alpha}^{\dagger}/\prescript{k}{}{S}^{\dagger}}) of relative log de Rham differentials.

There is a natural action vφv\mathbin{\lrcorner}\varphi for v𝒢αkv\in\prescript{k}{}{\mathcal{G}}_{\alpha}^{*} and φ𝒦k\varphi\in\prescript{k}{}{\mathcal{K}}^{*} given by contracting a logarithmic holomorphic vector field vv with a logarithmic holomorphic form φ\varphi. To simplify notations, for v𝒢α0kv\in\prescript{k}{}{\mathcal{G}}_{\alpha}^{0}, we often simply write vφv\varphi, suppressing the contraction \mathbin{\lrcorner}. We define the Lie derivative via the formula v:=(1)|v|(v)(v)\mathcal{L}_{v}:=(-1)^{|v|}\partial\circ(v\mathbin{\lrcorner})-(v\mathbin{\lrcorner})\circ\partial (or equivalently, (1)|v|v:=[,v](-1)^{|v|}\mathcal{L}_{v}:=[\prescript{}{}{\partial},v\mathbin{\lrcorner}]). By contracting with ωαk\prescript{k}{}{\omega}_{\alpha}, we get a sheaf isomorphism ωαk:𝒢αk𝒦αk\mathbin{\lrcorner}\prescript{k}{}{\omega}_{\alpha}\colon\prescript{k}{}{\mathcal{G}}_{\alpha}^{*}\rightarrow\prescript{k}{\parallel}{\mathcal{K}}^{*}_{\alpha}, which defines the BV operator Δαk\prescript{k}{}{\Delta}_{\alpha} by Δαk(φ)ωk:=αk(φωk)\prescript{k}{}{\Delta}_{\alpha}(\varphi)\mathbin{\lrcorner}\prescript{k}{}{\omega}:=\prescript{k}{}{\partial}_{\alpha}(\varphi\mathbin{\lrcorner}\prescript{k}{}{\omega}). We call it the BV operator because the BV identity:

(4.2) (1)|v|[v,w]:=Δ(vw)Δ(v)w(1)|v|vΔ(w)(-1)^{|v|}[v,w]:=\Delta(v\wedge w)-\Delta(v)\wedge w-(-1)^{|v|}v\wedge\Delta(w)

for v,w𝒢αkv,w\in\prescript{k}{}{\mathcal{G}}_{\alpha}^{*}, where we put Δ=Δαk\Delta=\prescript{k}{}{\Delta}_{\alpha}, defines a graded Lie bracket. This gives 𝒢αk\prescript{k}{}{\mathcal{G}}^{*}_{\alpha} the structure of a sheaf of BV algebras.

4.2. An explicit description of the sheaf of log de Rham forms

Here we apply the calculations in [28, 17] to give an explicit description of the stalk 𝒦α,xk\prescript{k}{}{\mathcal{K}}_{\alpha,x}^{*}.

Let us consider K=ν1(x)K=\nu^{-1}(x) and the local model near KK described in §4.1, with Pτ,xP_{\tau,x} and Qτ,xQ_{\tau,x} as in (2.17), (2.18) and an embedding VSpecan([Qτ,x])V\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[Q_{\tau,x}]). We may treat KVK\subset V as a compact subset of l=Specan([l])Specan([Qτ,x])\mathbb{C}^{l}=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\mathbb{N}^{l}])\hookrightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[Q_{\tau,x}]) via the identification Specan([Στl])Specan([Qτ,x])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}\oplus\mathbb{N}^{l}])\cong\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[Q_{\tau,x}]). For each mΣτm\in\Sigma_{\tau}, we denote the corresponding element (m,ψx,0(m),,ψx,l(m))Pτ,x(m,\psi_{x,0}(m),\dots,\psi_{x,l}(m))\in P_{\tau,x} by m^\hat{m} and the corresponding function by zm^[Pτ,x]z^{\hat{m}}\in\mathbb{C}[P_{\tau,x}]. Similar to [17, Lem. 7.14], the germs of holomorphic functions 𝒪𝕍k,K\mathcal{O}_{\prescript{k}{}{\mathbb{V}},K} near KK in the space 𝕍k=Specan([Pτ,x/qk+1])\prescript{k}{}{\mathbb{V}}=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{\tau,x}/q^{k+1}]) can be written as

(4.3) 𝒪𝕍k,K={mΣτ, 0ikαm,iqizm^|αm,i𝒪l(U)for some neigh. UK,supmΣτ{0}log|αm,i|𝚍(m)<},\mathcal{O}_{\prescript{k}{}{\mathbb{V}},K}=\Bigg{\{}\sum_{m\in\Sigma_{\tau},\ 0\leq i\leq k}\alpha_{m,i}q^{i}z^{\hat{m}}\,\Big{|}\,\alpha_{m,i}\in\mathcal{O}_{\mathbb{C}^{l}}(U)\ \text{for some neigh. $U\supset K$},\ \sup_{m\in\Sigma_{\tau}\setminus\{0\}}\frac{\log|\alpha_{m,i}|}{\mathtt{d}(m)}<\infty\Bigg{\}},

where 𝚍:Στ\mathtt{d}\colon\Sigma_{\tau}\rightarrow\mathbb{N} is a monoid morphism such that 𝚍1(0)=0\mathtt{d}^{-1}(0)=0, and it is equipped with the product zm^1zm^2:=zm^1+m^2z^{\hat{m}_{1}}\cdot z^{\hat{m}_{2}}:=z^{\hat{m}_{1}+\hat{m}_{2}} (but note that m1+m2^m^2+m^2\widehat{m_{1}+m_{2}}\neq\hat{m}_{2}+\hat{m}_{2} in general). Thus we have 𝒦α,x0k𝒢α,x0k𝒪𝕍k,K\prescript{k}{}{\mathcal{K}}^{0}_{\alpha,x}\cong\prescript{k}{}{\mathcal{G}}_{\alpha,x}^{0}\cong\mathcal{O}_{\prescript{k}{}{\mathbb{V}},K}.

To describe the differential forms, we consider the vector space =Pτ,x,\mathscr{E}=P_{\tau,x,\mathbb{C}}, regarded as the space of 11-forms on Specan([Pτ,xgp])()n+1\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[P_{\tau,x}^{\mathrm{gp}}])\cong(\mathbb{C}^{*})^{n+1}. Write dlogzpd\log z^{p} for pPτ,x,p\in P_{\tau,x,\mathbb{C}} and set 1:=dloguii=1l\mathscr{E}_{1}:=\mathbb{C}\langle d\log u_{i}\rangle_{i=1}^{l}, as a subset of \mathscr{E}. For an element m𝒬τ,m\in\mathscr{Q}_{\tau,\mathbb{C}}, we have the corresponding 11-form dlogzm^Pτ,x,d\log z^{\hat{m}}\in P_{\tau,x,\mathbb{C}} under the association between mm and zm^z^{\hat{m}}. Let 𝙿\mathtt{P} be the power set of {1,,l}\{1,\dots,l\} and write uI=iIuiu^{I}=\prod_{i\in I}u_{i} for I𝙿I\in\mathtt{P}. A computation for sections of the sheaf j(Ω𝕍k/r)j_{*}(\Omega^{r}_{\prescript{k}{}{\mathbb{V}}^{\dagger}/\mathbb{C}}) from [28, Prop. 1.12] and [17, Lem. 7.14] can then be rephrased as the following lemma.

Lemma 4.3 ([28, 17]).

The space of germs of sections of j(Ω𝕍k/)Kj_{*}(\Omega^{*}_{\prescript{k}{}{\mathbb{V}}^{\dagger}/\mathbb{C}})_{K} near KK is a subspace of 𝒪𝕍k,K\mathcal{O}_{\prescript{k}{}{\mathbb{V}},K}\otimes\bigwedge^{*}\mathscr{E} given by elements of the form

α=mΣτ0ikIαm,i,Iqizm^uIβm,I,βm,Im,I=(1,m,I2,m,Idlogq),\displaystyle\alpha=\sum_{\begin{subarray}{c}m\in\Sigma_{\tau}\\ 0\leq i\leq k\end{subarray}}\sum_{I}\alpha_{m,i,I}q^{i}z^{\hat{m}}u^{I}\otimes\beta_{m,I},\quad\beta_{m,I}\in\bigwedge\nolimits^{*}\mathscr{E}_{m,I}=\bigwedge\nolimits^{*}(\mathscr{E}_{1,m,I}\oplus\mathscr{E}_{2,m,I}\oplus\langle d\log q\rangle),

where 1,m,I=dloguiiI1\mathscr{E}_{1,m,I}=\langle d\log u_{i}\rangle_{i\in I}\subset\mathscr{E}_{1} and the subspace 2,m,I\mathscr{E}_{2,m,I}\subset\mathscr{E} is given as follows: we consider the pullback of the product of normal fans iI𝒩Δˇi(τ)\prod_{i\notin I}\mathscr{N}_{\check{\Delta}_{i}(\tau)} to 𝒬τ,\mathscr{Q}_{\tau,\real} and take 2,m,I=dlogzm^\mathscr{E}_{2,m,I}=\langle d\log z^{\hat{m}^{\prime}}\rangle for mσm,Im^{\prime}\in\sigma_{m,I}, where σm,I\sigma_{m,I} is the smallest cone in iI𝒩Δˇi(τ)𝒬τ,\prod_{i\notin I}\mathscr{N}_{\check{\Delta}_{i}(\tau)}\subset\mathscr{Q}_{\tau,\real} containing mm.

Here we can treat iI𝒩Δˇi(τ)𝒬τ,\prod_{i\notin I}\mathscr{N}_{\check{\Delta}_{i}(\tau)}\subset\mathscr{Q}_{\tau,\real} since iΛΔˇi(τ)\bigoplus_{i}\Lambda_{\check{\Delta}_{i}(\tau)} is a direct summand of 𝒬τ\mathscr{Q}_{\tau}^{*}. A similar description for j(Ω𝕍k/)Kj_{*}(\Omega^{*}_{\prescript{k}{}{\mathbb{V}}^{\dagger}/\mathbb{C}^{\dagger}})_{K} is simply given by quotienting out the direct summand dlogq\langle d\log q\rangle in the above formula for α\alpha. In particular, if we restrict ourselves to the case k=0k=0, a general element α\alpha can be written as

α=mΣτIαm,Izm^uIβm,I,βm,Im,I=(1,m,I2,m,I).\displaystyle\alpha=\sum_{m\in\Sigma_{\tau}}\sum_{I}\alpha_{m,I}z^{\hat{m}}u^{I}\otimes\beta_{m,I},\quad\beta_{m,I}\in\bigwedge\nolimits^{*}\mathscr{E}_{m,I}=\bigwedge\nolimits^{*}(\mathscr{E}_{1,m,I}\oplus\mathscr{E}_{2,m,I}).

One can choose a nowhere vanishing element

Ω=du1dulηu1ull1ndim(τ)2j(Ω𝕍0/n)K\Omega=du_{1}\cdots du_{l}\otimes\eta\in u_{1}\cdots u_{l}\otimes\wedge^{l}\mathscr{E}_{1}\otimes\wedge^{n-\dim(\tau)}\mathscr{E}_{2}\subset j_{*}(\Omega^{n}_{\prescript{0}{}{\mathbb{V}}^{\dagger}/\mathbb{C}^{\dagger}})_{K}

for some nonzero element ηndim(τ)2\eta\in\wedge^{n-\dim(\tau)}\mathscr{E}_{2}, which is well defined up to rescaling. Any element in j(Ω𝕍0/n)Kj_{*}(\Omega^{n}_{\prescript{0}{}{\mathbb{V}}^{\dagger}/\mathbb{C}^{\dagger}})_{K} can be written as fΩf\Omega for some f=mΣτfmzm^𝒪𝕍0,Kf=\sum_{m\in\Sigma_{\tau}}f_{m}z^{\hat{m}}\in\mathcal{O}_{\prescript{0}{}{\mathbb{V}},K}.

Recall that the subset KlK\subset\mathbb{C}^{l} is intersecting the singular locus Z1τ,,ZrτZ^{\tau}_{1},\dots,Z^{\tau}_{r} (as in §4.1), where uiu_{i} is the coordinate function of l\mathbb{C}^{l} with simple zeros along ZiτZ_{i}^{\tau} for i=1,,ri=1,\dots,r. There is a change of coordinates between a neighborhood of KK in l\mathbb{C}^{l} and that of KK in ()l(\mathbb{C}^{*})^{l} given by

{uifv,i|()lif 1ir;uiziif r<il.\begin{cases}u_{i}\mapsto f_{v,i}|_{(\mathbb{C}^{*})^{l}}&\text{if }1\leq i\leq r;\\ u_{i}\mapsto z_{i}&\text{if }r<i\leq l.\end{cases}

Under the map log:()ll\log\colon(\mathbb{C}^{*})^{l}\rightarrow{}^{l}, we have K=log1(𝒞)K=\log^{-1}(\mathscr{C}) for some connected compact subset 𝒞l\mathscr{C}\subset{}^{l}. In the coordinates z1,,zlz_{1},\dots,z_{l}, we find that dlogz1dlogzlηd\log z_{1}\cdots d\log z_{l}\otimes\eta can be written as fΩf\Omega near KK for some nowhere vanishing function f𝒪𝕍0,Kf\in\mathcal{O}_{\prescript{0}{}{\mathbb{V}},K}.

Lemma 4.4.

When KZ=K\cap Z=\emptyset (i.e. r=0r=0 in the above discussion), the top cohomology group n(j(Ω𝕍0/n)K,):=j(Ω𝕍0/n)K/Im()\mathcal{H}^{n}(j_{*}(\Omega^{n}_{\prescript{0}{}{\mathbb{V}}^{\dagger}/\mathbb{C}^{\dagger}})_{K},\prescript{}{}{\partial}):=j_{*}(\Omega^{n}_{\prescript{0}{}{\mathbb{V}}^{\dagger}/\mathbb{C}^{\dagger}})_{K}/\mathrm{Im}(\prescript{}{}{\partial}) is isomorphic to \mathbb{C}, which is generated by the element dlogz1dlogzlηd\log z_{1}\cdots d\log z_{l}\otimes\eta.

Proof.

Given a general element fΩf\Omega as above, first observe that we can write f=f0+f+f=f_{0}+f_{+}, where f+=mΣτ{0}fmzm^f_{+}=\sum_{m\in\Sigma_{\tau}\setminus\{0\}}f_{m}z^{\hat{m}} and f0𝒪l,Kf_{0}\in\mathcal{O}_{\mathbb{C}^{l},K}. We take a basis e1,,ese_{1},\dots,e_{s} of 𝒬τ,\mathscr{Q}_{\tau,\real}^{*}, and also a partition I1,,IsI_{1},\dots,I_{s} of the lattice points in Στ{0}\Sigma_{\tau}\setminus\{0\} such that ej,m0\langle e_{j},m\rangle\neq 0 for mIjm\in I_{j}. Letting

α=(1)ljmIjfmej,mzm^du1dulιejη,\alpha=(-1)^{l}\sum_{j}\sum_{m\in I_{j}}\frac{f_{m}}{\langle e_{j},m\rangle}z^{\hat{m}}du_{1}\cdots du_{l}\otimes\iota_{e_{j}}\eta,

we have (α)=f+Ω\prescript{}{}{\partial}(\alpha)=f_{+}\Omega. So we only need to consider elements of the form f0Ωf_{0}\Omega. If f0Ω=(α)f_{0}\Omega=\prescript{}{}{\partial}(\alpha) for some α\alpha, we may take α=jαjdu1duj^dulη\alpha=\sum_{j}\alpha_{j}du_{1}\cdots\widehat{du_{j}}\cdots du_{l}\otimes\eta for some αj𝒪l,K\alpha_{j}\in\mathcal{O}_{\mathbb{C}^{l},K}. Now this is equivalent to f0du1dul=(jαjdu1duj^dul)f_{0}du_{1}\cdots du_{l}=\prescript{}{}{\partial}\big{(}\sum_{j}\alpha_{j}du_{1}\cdots\widehat{du_{j}}\cdots du_{l}\big{)} as forms in Ωl,Kl\Omega^{l}_{\mathbb{C}^{l},K}. This reduces the problem to l\mathbb{C}^{l}.

Working in ()l(\mathbb{C}^{*})^{l} with coordinates ziz_{i}’s, we can write

𝒪()l,K={mlamzm|ml|am|ev,m<,for all vW, for some open W𝒞},\mathcal{O}_{(\mathbb{C}^{*})^{l},K}=\left\{\sum_{m\in\mathbb{Z}^{l}}a_{m}z^{m}\ \Big{|}\ \sum_{m\in\mathbb{Z}^{l}}|a_{m}|e^{\langle v,m\rangle}<\infty,\ \text{for all $v\in W$, for some open $W\supset\mathscr{C}$}\right\},

using the fact that KK is multi-circular. By writing Ω()l,K=𝒪()l,K1\Omega^{*}_{(\mathbb{C}^{*})^{l},K}=\mathcal{O}_{(\mathbb{C}^{*})^{l},K}\otimes\bigwedge^{*}\mathscr{F}_{1} with 1=dlogzii=1l\mathscr{F}_{1}=\langle d\log z_{i}\rangle_{i=1}^{l}, we can see that any element can be represented as cdlogz1dlogzlcd\log z_{1}\cdots d\log z_{l} in the quotient Ω()l,Kl/Im()\Omega^{l}_{(\mathbb{C}^{*})^{l},K}/\mathrm{Im}(\prescript{}{}{\partial}), for some constant cc. ∎

From this lemma, we conclude that the top cohomology sheaf n(𝒦0,)\mathcal{H}^{n}(\prescript{0}{\parallel}{\mathcal{K}}^{*},\prescript{}{}{\partial}) is isomorphic to the locally constant sheaf ¯\underline{\mathbb{C}} over B𝒮eB\setminus\mathscr{S}_{e}.

Lemma 4.5.

The volume element ω0\prescript{0}{}{\omega} is non-zero in n(𝒦0,)x\mathcal{H}^{n}(\prescript{0}{\parallel}{\mathcal{K}}^{*},\prescript{}{}{\partial})_{x} for every xBx\in B.

Proof.

We first consider the case when xintre(σ)x\in\mathrm{int}_{\mathrm{re}}(\sigma) for some maximal cell σ𝒫[n]\sigma\in\mathscr{P}^{[n]}. The toric stratum Xσ0\prescript{0}{}{X}_{\sigma} associated to σ\sigma is equipped with the natural divisorial log structure induced from its boundary divisor. Then the sheaf ΩXσ0/\Omega^{*}_{\prescript{0}{}{X}_{\sigma}^{\dagger}/\mathbb{C}^{\dagger}} of log derivations for X0\prescript{0}{}{X}^{\dagger} is isomorphic to nΛσ𝒪Xσ0\bigwedge^{n}\Lambda_{\sigma}\otimes_{\mathbb{Z}}\mathcal{O}_{\prescript{0}{}{X}_{\sigma}}. By [28, Lem. 3.12], we have ωx0=c(μσ)ν1(x)\prescript{0}{}{\omega}_{x}=c(\mu_{\sigma})_{\nu^{-1}(x)} in ν(ΩXσ0/n)x𝒦xn0\nu_{*}(\Omega^{n}_{\prescript{0}{}{X}_{\sigma}^{\dagger}/\mathbb{C}^{\dagger}})_{x}\cong\prescript{0}{\parallel}{\mathcal{K}}^{n}_{x}, where μσnΛσ,\mu_{\sigma}\in\bigwedge^{n}\Lambda_{\sigma,\mathbb{C}} is nowhere vanishing and cc is a non-zero constant cc. Thus X0|x\prescript{0}{}{X}|_{x} is non-zero in the cohomology as the same is true for μσν(ΩXσ0/n)x\mu_{\sigma}\in\nu_{*}(\Omega^{n}_{\prescript{0}{}{X}_{\sigma}^{\dagger}/\mathbb{C}^{\dagger}})_{x}. Next we consider a general point xintre(τ)x\in\mathrm{int}_{\mathrm{re}}(\tau). If the statement is not true, we will have ωx0=0(α)\prescript{0}{}{\omega}_{x}=\prescript{0}{}{\partial}(\alpha) for some α𝒦xn10\alpha\in\prescript{0}{\parallel}{\mathcal{K}}^{n-1}_{x}. Then there is an open neighborhood U𝒞1(x)U\supset\mathscr{C}^{-1}(x) such that this relation continues to hold. As Uintre(σ)U\cap\mathrm{int}_{\mathrm{re}}(\sigma)\neq\emptyset, for those maximal cells σ\sigma which contain the point xx, we can take a nearby point yUintre(σ)y\in U\cap\mathrm{int}_{\mathrm{re}}(\sigma) and conclude that cμσ=0(α)c\mu_{\sigma}=\prescript{0}{}{\partial}(\alpha) in ν(ΩXσ0/n)y\nu_{*}(\Omega^{n}_{\prescript{0}{}{X}_{\sigma}^{\dagger}/\mathbb{C}^{\dagger}})_{y}. This contradicts the previous case. ∎

Lemma 4.6.

Suppose that xWα𝒮ex\in W_{\alpha}\setminus\mathscr{S}_{e}. For an element of the form

ef(ωαk)𝒦α,xnke^{f}(\prescript{k}{}{\omega}_{\alpha})\in\prescript{k}{\parallel}{\mathcal{K}}^{n}_{\alpha,x}

with f𝒢α,x0k𝒪𝕍αk,xf\in\prescript{k}{}{\mathcal{G}}^{0}_{\alpha,x}\cong\mathcal{O}_{\prescript{k}{}{\mathbb{V}}_{\alpha},x} satisfying f0(mod 𝐦)f\equiv 0\text{(mod $\mathbf{m}$)}, there exist h(q)Rk=[q]/(qk+1)h(q)\in\prescript{k}{}{R}=\mathbb{C}[q]/(q^{k+1}) and v𝒢α,x1kv\in\prescript{k}{}{\mathcal{G}}^{-1}_{\alpha,x} with h,v0(mod 𝐦)h,v\equiv 0\text{(mod $\mathbf{m}$)} such that

(4.4) ef(ωαk)=ehev(ωαk)e^{f}(\prescript{k}{}{\omega}_{\alpha})=e^{h}e^{\mathcal{L}_{v}}(\prescript{k}{}{\omega}_{\alpha})

in 𝒦α,xnk\prescript{k}{\parallel}{\mathcal{K}}^{n}_{\alpha,x}, where we recall that v:=(1)|v|(v)(v)\mathcal{L}_{v}:=(-1)^{|v|}\partial\circ(v\mathbin{\lrcorner})-(v\mathbin{\lrcorner})\circ\partial.

Proof.

To simplify notations in this proof, we will drop the subscript α\alpha. We prove the first statement by induction on kk. The initial case is trivial. Assuming that this has been done for the (k1)st(k-1)^{\text{st}}-order, then, by taking an arbitrary lifting v~\tilde{v} of vv to the kthk^{\text{th}}-order, we have

eh+f+qkϵ(ωk)=ev~(ωk)e^{-h+f+q^{k}\epsilon}(\prescript{k}{}{\omega})=e^{\mathcal{L}_{\tilde{v}}}(\prescript{k}{}{\omega})

for some ϵ𝒪𝕍x0\epsilon\in\mathcal{O}_{\prescript{0}{}{\mathbb{V}}_{x}}. By Lemmas 4.4 and 4.5, we have ϵω0=cω0+(γ)\epsilon\prescript{0}{}{\omega}=c\prescript{0}{}{\omega}+\prescript{}{}{\partial}(\gamma) for some γ\gamma and some suitable constant cc. Letting θ(ω0)=γ\theta\mathbin{\lrcorner}(\prescript{0}{}{\omega})=\gamma and v˘=v~+qkθ\breve{v}=\tilde{v}+q^{k}\theta, we have

ev˘(ωk)=ev(ωk)qk(θ(ω0))=eh+f+cqk(ωk).e^{\mathcal{L}_{\breve{v}}}(\prescript{k}{}{\omega})=e^{\mathcal{L}_{v}}(\prescript{k}{}{\omega})-q^{k}\prescript{}{}{\partial}(\theta\mathbin{\lrcorner}(\prescript{0}{}{\omega}))=e^{-h+f+cq^{k}}(\prescript{k}{}{\omega}).

By defining h~(q):=h(q)cqk\tilde{h}(q):=h(q)-cq^{k} in [q]/(qk+1)\mathbb{C}[q]/(q^{k+1}), we obtain the desired expression. ∎

4.3. A global pre-dgBV algebra from gluing

One approach for smoothing X0\prescript{0}{}{X} is to look for gluing morphisms ψαβk:𝕍αk|Vαβ𝕍βk|Vαβ\prescript{k}{}{\psi}_{\alpha\beta}\colon\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}|_{V_{\alpha\beta}}\rightarrow\prescript{k}{}{\mathbb{V}}_{\beta}^{\dagger}|_{V_{\alpha\beta}} between the local smoothing models which satisfy the cocycle condition, from which one obtain a kthk^{\text{th}}-order thickening Xk\prescript{k}{}{X} over Sk\prescript{k}{}{S}^{\dagger}. This was done by Kontsevich–Soibelman [36] (in 2d) and Gross–Siebert [29] (in general dimensions) using consistent scattering diagrams. If such gluing morphisms ψαβk\prescript{k}{}{\psi}_{\alpha\beta}’s are available, one can certainly glue the global kthk^{\text{th}}-order sheaves 𝒢k\prescript{k}{}{\mathcal{G}}^{*}, 𝒦k\prescript{k}{}{\mathcal{K}}^{*} and the volume form ωk\prescript{k}{}{\omega}.

In [8], we instead took suitable dg-resolutions PkVα,:=Ω(𝒢αk)\prescript{k}{}{PV}^{*,*}_{\alpha}:=\Omega^{*}(\prescript{k}{}{\mathcal{G}}_{\alpha}^{*})’s of the sheaves 𝒢αk\prescript{k}{}{\mathcal{G}}_{\alpha}^{*}’s (more precisely, we used the Thom–Whitney resolution in [8, §3]) to construct gluings

gαβk:Ω(𝒢αk)|VαβΩ(𝒢βk)|Vαβ\prescript{k}{}{g}_{\alpha\beta}\colon\Omega^{*}(\prescript{k}{}{\mathcal{G}}_{\alpha}^{*})|_{V_{\alpha\beta}}\rightarrow\Omega^{*}(\prescript{k}{}{\mathcal{G}}_{\beta}^{*})|_{V_{\alpha\beta}}

of sheaves which only preserve the Gerstenhaber algebra structure but not the differential. The key discovery in [8] was that, as the sheaves Ω(𝒢αk)\Omega^{*}(\prescript{k}{}{\mathcal{G}}_{\alpha}^{*})’s are soft, such a gluing problem could be solved without any information from the complicated scattering diagrams. What we obtained is a pre-dgBV algebra555This was originally called an almost dgBV algebra in [8], but we later found the name pre-dgBV algebra from [16] more appropriate. PkV,(X)\prescript{k}{}{PV}^{*,*}(\prescript{}{}{X}), in which the differential squares to zero only modulo 𝐦=(q)\mathbf{m}=(q). Using well-known algebraic techniques [48, 33], we can solve the Maurer–Cartan equation and construct the thickening Xk\prescript{k}{}{X}. In this subsection, we will summarize the whole procedure, incorporating the nice reformulation by Felten [16] in terms of deformations of Gerstenhaber algebras.

To begin with, we assume the following condition holds:

Condition 4.7.

There is a sheaf (Ω,d)(\Omega^{*},d) of unital differential graded algebras (abbrev. as dga) (over or \mathbb{C}) over BB, with degrees 0L0\leq*\leq L for some LL, such that

  • the natural inclusion ¯Ω\underline{\real}\rightarrow\Omega^{*} (or ¯Ω\underline{\mathbb{C}}\rightarrow\Omega^{*}) of the locally constant sheaf (concentrated at degree 0) gives a resolution, and

  • for any open cover 𝒰={Ui}i\mathcal{U}=\{U_{i}\}_{i\in\mathcal{I}}, there is a partition of unity subordinate to 𝒰\mathcal{U}, i.e. we have {ρi}i\{\rho_{i}\}_{i\in\mathcal{I}} with ρiΓ(Ui,Ω0)\rho_{i}\in\Gamma(U_{i},\Omega^{0}) and supp(ρi)¯Ui\overline{\mathrm{supp}(\rho_{i})}\subset U_{i} such that {supp(ρi)¯}i\{\overline{\mathrm{supp}(\rho_{i})}\}_{i} is locally finite and iρi1\sum_{i}\rho_{i}\equiv 1.

It is easy to construct such an Ω\Omega^{*} and there are many natural choices. For instance, if BB is a smooth manifold, then we can simply take the usual de Rham complex on BB. In §3.3.2, the sheaf of monodromy invariant differential forms we constructed using the (singular) integral affine structure on BB is another possible choice for Ω\Omega^{*} (with degrees 0n0\leq*\leq n). Yet another variant, namely the sheaf of monodromy invariant tropical differential forms, will be constructed in §5.1; this links tropical geometry on BB with the smoothing of the maximally degenerate Calabi–Yau variety X0\prescript{0}{}{X}.

Let us recall how to obtain a gluing of the dg resolutions of the sheaves 𝒢αk\prescript{k}{}{\mathcal{G}}_{\alpha}^{*} and 𝒦αk\prescript{k}{}{\mathcal{K}}_{\alpha}^{*} using any possible choice of such an Ω\Omega^{*}. We fix a good cover 𝒲:={Wα}α\mathcal{W}:=\{W_{\alpha}\}_{\alpha} of BB and the corresponding Stein open cover 𝒱:={Vα}α\mathcal{V}:=\{V_{\alpha}\}_{\alpha} of X0\prescript{0}{}{X}, where Vα=ν1(Wα)V_{\alpha}=\nu^{-1}(W_{\alpha}) for each α\alpha.

Definition 4.8.

We define PkVαp,q=Ωq(𝒢αpk):=Ωq|Wα𝒢αpk\prescript{k}{}{PV}_{\alpha}^{p,q}=\Omega^{q}(\prescript{k}{}{\mathcal{G}}_{\alpha}^{p}):=\Omega^{q}|_{W_{\alpha}}\otimes\prescript{k}{}{\mathcal{G}}^{p}_{\alpha} and PkVα,=p,qPkVαp,q\prescript{k}{}{PV}_{\alpha}^{*,*}=\bigoplus_{p,q}\prescript{k}{}{PV}_{\alpha}^{p,q}, which gives a sheaf of dgBV algebras over WαW_{\alpha}. The dgBV structure (,¯α,Δα)(\wedge,\bar{\partial}_{\alpha},\prescript{}{}{\Delta}_{\alpha}) is defined componentwise by

(φv)(ψw)\displaystyle(\varphi\otimes v)\wedge(\psi\otimes w) :=(1)|v||ψ|(φψ)(vw),\displaystyle:=(-1)^{|v||\psi|}(\varphi\wedge\psi)\otimes(v\wedge w),
¯α(φv):=(dφ)v,\displaystyle\bar{\partial}_{\alpha}(\varphi\otimes v):=(d\varphi)\otimes v, Δα(φv):=(1)|φ|φ(Δv),\displaystyle\quad\prescript{}{}{\Delta}_{\alpha}(\varphi\otimes v):=(-1)^{|\varphi|}\varphi\otimes(\prescript{}{}{\Delta}v),

for φ,ψΩ(U)\varphi,\psi\in\Omega^{*}(U) and v,w𝒢αk(U)v,w\in\prescript{k}{}{\mathcal{G}}_{\alpha}^{*}(U) for each open subset UWαU\subset W_{\alpha}.

Definition 4.9.

We define 𝒜αp,qk=Ωq(𝒦αpk):=Ωq|Wα𝒦αpk\prescript{k}{}{\mathcal{A}}_{\alpha}^{p,q}=\Omega^{q}(\prescript{k}{}{\mathcal{K}}_{\alpha}^{p}):=\Omega^{q}|_{W_{\alpha}}\otimes\prescript{k}{}{\mathcal{K}}^{p}_{\alpha} and 𝒜α,k=p,q𝒜αp,qk\prescript{k}{}{\mathcal{A}}_{\alpha}^{*,*}=\bigoplus_{p,q}\prescript{k}{}{\mathcal{A}}_{\alpha}^{p,q}, which gives a sheaf of dgas over WαW_{\alpha} equipped with the natural filtration 𝒜α,k\prescript{k}{\bullet}{\mathcal{A}}_{\alpha}^{*,*} inherited from 𝒦αk\prescript{k}{\bullet}{\mathcal{K}}^{*}_{\alpha}. The structures (,¯α,α)(\wedge,\bar{\partial}_{\alpha},\prescript{}{}{\partial}_{\alpha}) are defined componentwise by

(φv)(ψw)\displaystyle(\varphi\otimes v)\wedge(\psi\otimes w) :=(1)|v||ψ|(φψ)(vw),\displaystyle:=(-1)^{|v||\psi|}(\varphi\wedge\psi)\otimes(v\wedge w),
¯α(φv):=(dφ)v,\displaystyle\bar{\partial}_{\alpha}(\varphi\otimes v):=(d\varphi)\otimes v, α(φv)=(1)|φ|φ(v),\displaystyle\quad\prescript{}{}{\partial}_{\alpha}(\varphi\otimes v)=(-1)^{|\varphi|}\varphi\otimes(\prescript{}{}{\partial}v),

for φ,ψΩ(U)\varphi,\psi\in\Omega^{*}(U) and v,w𝒦αk(U)v,w\in\prescript{k}{}{\mathcal{K}}_{\alpha}^{*}(U) for each open subset UWαU\subset W_{\alpha}.

There is an action of PkVα,\prescript{k}{}{PV}_{\alpha}^{*,*} on 𝒜α,k\prescript{k}{}{\mathcal{A}}_{\alpha}^{*,*} by contraction \mathbin{\lrcorner} defined by the formula

(φv)(ψw):=(1)|v||ψ|(φψ)(vw),(\varphi\otimes v)\mathbin{\lrcorner}(\psi\otimes w):=(-1)^{|v||\psi|}(\varphi\wedge\psi)\otimes(v\mathbin{\lrcorner}w),

for φ,ψΩ(U)\varphi,\psi\in\Omega^{*}(U), v𝒢αk(U)v\in\prescript{k}{}{\mathcal{G}}_{\alpha}^{*}(U) and w𝒦αk(U)w\in\prescript{k}{}{\mathcal{K}}_{\alpha}^{*}(U) for each open subset UWαU\subset W_{\alpha}. Note that the local holomorphic volume form ωαk𝒜αn,0k(Wα)\prescript{k}{}{\omega}_{\alpha}\in\prescript{k}{\parallel}{\mathcal{A}}_{\alpha}^{n,0}(W_{\alpha}) satisfies ¯α(ωαk)=0\bar{\partial}_{\alpha}(\prescript{k}{}{\omega}_{\alpha})=0, and we have the identity αk(ϕωαk)=Δαk(ϕ)ωαk\prescript{k}{}{\partial}_{\alpha}(\phi\mathbin{\lrcorner}\prescript{k}{}{\omega}_{\alpha})=\prescript{k}{}{\Delta}_{\alpha}(\phi)\mathbin{\lrcorner}\prescript{k}{}{\omega}_{\alpha} of operators.

The next step is to consider gluing of the local sheaves PkVα\prescript{k}{}{PV}_{\alpha}’s for higher orders kk. Similar constructions have been done in [8, 16] using the combinatorial Thom–Whitney resolution for the sheaves 𝒢αk\prescript{k}{}{\mathcal{G}}_{\alpha}’s. We make suitable modifications of those arguments to fit into our current setting.

First, since 𝕍αk|Vαβ\prescript{k}{}{\mathbb{V}}^{\dagger}_{\alpha}|_{V_{\alpha\beta}} and 𝕍βk|Vαβ\prescript{k}{}{\mathbb{V}}^{\dagger}_{\beta}|_{V_{\alpha\beta}} are divisorial deformations (in the sense of [28, Def. 2.7]) of the intersection Vαβ:=VαVβV^{\dagger}_{\alpha\beta}:=V^{\dagger}_{\alpha}\cap V^{\dagger}_{\beta}, we can use [28, Thm. 2.11] and the fact that VαβV_{\alpha\beta} is Stein to obtain an isomorphism ψαβk:𝕍αk|Vαβ𝕍βk|Vαβ\prescript{k}{}{\psi}_{\alpha\beta}\colon\prescript{k}{}{\mathbb{V}}^{\dagger}_{\alpha}|_{V_{\alpha\beta}}\rightarrow\prescript{k}{}{\mathbb{V}}^{\dagger}_{\beta}|_{V_{\alpha\beta}} of divisorial deformations which induces the gluing morphism ψαβk:𝒢αk|Wαβ𝒢βk|Wαβ\prescript{k}{}{\psi}_{\alpha\beta}\colon\prescript{k}{}{\mathcal{G}}_{\alpha}^{*}|_{W_{\alpha\beta}}\rightarrow\prescript{k}{}{\mathcal{G}}_{\beta}^{*}|_{W_{\alpha\beta}} that in turn gives ψαβk:PkVα|WαβPkVβ|Wαβ\prescript{k}{}{\psi}_{\alpha\beta}\colon\prescript{k}{}{PV}_{\alpha}|_{W_{\alpha\beta}}\rightarrow\prescript{k}{}{PV}_{\beta}|_{W_{\alpha\beta}}.

Definition 4.10.

A kthk^{\text{th}}-order Gerstenhaber deformation of P0V\prescript{0}{}{PV} is a collection of gluing morphisms gαβk:PkVα|WαβPkVβ|Wαβ\prescript{k}{}{g}_{\alpha\beta}\colon\prescript{k}{}{PV}_{\alpha}|_{W_{\alpha\beta}}\rightarrow\prescript{k}{}{PV}_{\beta}|_{W_{\alpha\beta}} of the form

gαβk=e[ϑαβ,]ψαβk\prescript{k}{}{g}_{\alpha\beta}=e^{[\vartheta_{\alpha\beta},\cdot]}\circ\prescript{k}{}{\psi}_{\alpha\beta}

for some θαβPkVβ1,0(Wαβ)\theta_{\alpha\beta}\in\prescript{k}{}{PV}_{\beta}^{-1,0}(W_{\alpha\beta}) with θαβ0(mod 𝐦)\theta_{\alpha\beta}\equiv 0\ (\text{mod $\mathbf{m}$}), such that the cocycle condition

gγαkgβγkgαβk=id\prescript{k}{}{g}_{\gamma\alpha}\circ\prescript{k}{}{g}_{\beta\gamma}\circ\prescript{k}{}{g}_{\alpha\beta}=\mathrm{id}

is satisfied.

An isomorphism between two kthk^{\text{th}}-order Gerstenhaber deformations {gαβk}αβ\{\prescript{k}{}{g}_{\alpha\beta}\}_{\alpha\beta} and {gαβk}αβ\{\prescript{k}{}{g}_{\alpha\beta}^{\prime}\}_{\alpha\beta} is a collection of automorphisms hαk:PkVαPkVα\prescript{k}{}{h}_{\alpha}\colon\prescript{k}{}{PV}_{\alpha}\rightarrow\prescript{k}{}{PV}_{\alpha} of the form

hαk=e[𝐛α,]\prescript{k}{}{h}_{\alpha}=e^{[\mathbf{b}_{\alpha},\cdot]}

for some 𝚋αPkVα1,0(Wα)\mathtt{b}_{\alpha}\in\prescript{k}{}{PV}_{\alpha}^{-1,0}(W_{\alpha}) with 𝚋α0(mod 𝐦)\mathtt{b}_{\alpha}\equiv 0(\text{mod $\mathbf{m}$}), such that

gαβkhαk=hβkgαβk.\prescript{k}{}{g}_{\alpha\beta}^{\prime}\circ\prescript{k}{}{h}_{\alpha}=\prescript{k}{}{h}_{\beta}\circ\prescript{k}{}{g}_{\alpha\beta}.

A slight modification of [16, Lem. 6.6], with essentially the same proof, gives the following:

Proposition 4.11.

Given a kthk^{\text{th}}-order Gerstenhaber deformation {gαβk}αβ\{\prescript{k}{}{g}_{\alpha\beta}\}_{\alpha\beta}, the obstruction to the existence of a lifting to a (k+1)st(k+1)^{\text{st}}-order deformation {gαβk+1}αβ\{\prescript{k+1}{}{g}_{\alpha\beta}\}_{\alpha\beta} lies in the Čech cohomology (with respect to the cover 𝒲={Wα}α\mathcal{W}=\{W_{\alpha}\}_{\alpha})

Hˇ2(𝒲,P0V1,0)(𝐦k+1/𝐦k+2).\check{H}^{2}(\mathcal{W},\prescript{0}{}{PV}^{-1,0})\otimes_{\mathbb{C}}(\mathbf{m}^{k+1}/\mathbf{m}^{k+2}).

The isomorphism classes of (k+1)st(k+1)^{\text{st}}-order liftings are in

Hˇ1(𝒲,P0V1,0)(𝐦k+1/𝐦k+2).\check{H}^{1}(\mathcal{W},\prescript{0}{}{PV}^{-1,0})\otimes_{\mathbb{C}}(\mathbf{m}^{k+1}/\mathbf{m}^{k+2}).

Fixing a (k+1)st(k+1)^{\text{st}}-order lifting {gαβk+1}αβ\{\prescript{k+1}{}{g}_{\alpha\beta}\}_{\alpha\beta}, the automorphisms fixing {gαβk}αβ\{\prescript{k}{}{g}_{\alpha\beta}\}_{\alpha\beta} are in

Hˇ0(𝒲,P0V1,0)(𝐦k+1/𝐦k+2).\check{H}^{0}(\mathcal{W},\prescript{0}{}{PV}^{-1,0})\otimes_{\mathbb{C}}(\mathbf{m}^{k+1}/\mathbf{m}^{k+2}).

Since Ωi\Omega^{i} satisfies Condition 4.7, we have Hˇ>0(𝒲,P0V1,0)=0\check{H}^{>0}(\mathcal{W},\prescript{0}{}{PV}^{-1,0})=0. In particular, we always have a set of compatible Gerstenhaber deformations g=(gk)k\prescript{}{}{g}=(\prescript{k}{}{g})_{k\in\mathbb{N}} where gk={gαβk}αβ\prescript{k}{}{g}=\{\prescript{k}{}{g}_{\alpha\beta}\}_{\alpha\beta} and any two of them are equivalent. Fixing such a set g\prescript{}{}{g}, we obtain a set {PkV}k\{\prescript{k}{}{PV}\}_{k\in\mathbb{N}} of Gerstenhaber algebras which is compatible, in the sense that there are natural identifications Pk+1VRk+1Rk=PkV\prescript{k+1}{}{PV}\otimes_{\prescript{k+1}{}{R}}\prescript{k}{}{R}=\prescript{k}{}{PV}.

We can also glue the local sheaves 𝒜α,k\prescript{k}{}{\mathcal{A}}^{*,*}_{\alpha}’s of dgas using g=(gk)k\prescript{}{}{g}=(\prescript{k}{}{g})_{k\in\mathbb{N}}. First, we can define ψαβk:𝒦αk|Wαβ𝒦βk|Wαβ\prescript{k}{}{\psi}_{\alpha\beta}\colon\prescript{k}{}{\mathcal{K}}_{\alpha}^{*}|_{W_{\alpha\beta}}\rightarrow\prescript{k}{}{\mathcal{K}}_{\beta}^{*}|_{W_{\alpha\beta}} using ψαβk:𝕍αk|Vαβ𝕍βk|Vαβ\prescript{k}{}{\psi}_{\alpha\beta}\colon\prescript{k}{}{\mathbb{V}}^{\dagger}_{\alpha}|_{V_{\alpha\beta}}\rightarrow\prescript{k}{}{\mathbb{V}}^{\dagger}_{\beta}|_{V_{\alpha\beta}}. For each fixed kk, we can write gαβk=e[ϑαβ,]ψαβk\prescript{k}{}{g}_{\alpha\beta}=e^{[\vartheta_{\alpha\beta},\cdot]}\circ\prescript{k}{}{\psi}_{\alpha\beta} as before. Then

(4.5) gk:=eϑαβψαβk:𝒜α,k|Wαβ𝒜β,k|Wαβ,\prescript{k}{}{g}:=e^{\mathcal{L}_{\vartheta_{\alpha\beta}}}\circ\prescript{k}{}{\psi}_{\alpha\beta}\colon\prescript{k}{}{\mathcal{A}}_{\alpha}^{*,*}|_{W_{\alpha\beta}}\rightarrow\prescript{k}{}{\mathcal{A}}_{\beta}^{*,*}|_{W_{\alpha\beta}},

where we recall that v:=(1)|v|(v)(v)\mathcal{L}_{v}:=(-1)^{|v|}\partial\circ(v\mathbin{\lrcorner})-(v\mathbin{\lrcorner})\circ\partial, preserves the dga structure (,α)(\wedge,\prescript{}{}{\partial}_{\alpha}) and the filtration on 𝒜α,k\prescript{k}{\bullet}{\mathcal{A}}_{\alpha}^{*,*}’s. As a result, we obtain a set of compatible sheaves {(𝒜,k,,)}k\{(\prescript{k}{}{\mathcal{A}}^{*,*},\wedge,\prescript{}{}{\partial})\}_{k\in\mathbb{N}} of dgas. The contraction action \mathbin{\lrcorner} is also compatible with the gluing construction, so we have a natural action \mathbin{\lrcorner} of PkV,\prescript{k}{}{PV}^{*,*} on 𝒜,k\prescript{k}{}{\mathcal{A}}^{*,*}.

Next, we glue the operators ¯α\bar{\partial}_{\alpha}’s and Δα\prescript{}{}{\Delta}_{\alpha}’s.

Definition 4.12.

A kthk^{\text{th}}-order pre-differential ¯\bar{\partial} on PkV,\prescript{k}{}{PV}^{*,*} is a degree (0,1)(0,1) operator obtained from gluing the operators ¯α+[ηα,]\bar{\partial}_{\alpha}+[\eta_{\alpha},\cdot] specified by a collection of elements ηαPkVα1,1(Wα)\eta_{\alpha}\in\prescript{k}{}{PV}^{-1,1}_{\alpha}(W_{\alpha}) such that ηα0(mod 𝐦)\eta_{\alpha}\equiv 0\ (\text{mod $\mathbf{m}$}) and

gβαk(¯β+[ηβ,])gαβk=(¯α+[ηα,]).\prescript{k}{}{g}_{\beta\alpha}\circ(\bar{\partial}_{\beta}+[\eta_{\beta},\cdot])\circ\prescript{k}{}{g}_{\alpha\beta}=(\bar{\partial}_{\alpha}+[\eta_{\alpha},\cdot]).

Two pre-differentials ¯\bar{\partial} and ¯\bar{\partial}^{\prime} are equivalent if there is a Gerstenhaber automorphism (for the deformation gk\prescript{k}{}{g}) h:PkV,PkV,h\colon\prescript{k}{}{PV}^{*,*}\rightarrow\prescript{k}{}{PV}^{*,*} such that h1¯h=¯h^{-1}\circ\bar{\partial}\circ h=\bar{\partial}^{\prime}.

Notice that we only have ¯20\bar{\partial}^{2}\equiv 0 (mod 𝐦)(\text{mod $\mathbf{m}$}), which is why we call it a pre-differential. Using the argument in [8, Thm. 3.34] or [16, Lem. 8.1], we can always lift any kthk^{\text{th}}-order pre-differential ¯k\prescript{k}{}{\bar{\partial}} to a (k+1)st(k+1)^{\text{st}}-order pre-differential. Furthermore, any two such liftings differ by a global element 𝔡P0V1,1𝐦k+1/𝐦k+2\mathfrak{d}\in\prescript{0}{}{PV}^{-1,1}\otimes\mathbf{m}^{k+1}/\mathbf{m}^{k+2}. We fix a set ¯:={¯k}k\bar{\partial}:=\{\prescript{k}{}{\bar{\partial}}\}_{k\in\mathbb{N}} of such compatible pre-differentials. For each kk, the action of ¯k\prescript{k}{}{\bar{\partial}} on 𝒜,k\prescript{k}{}{\mathcal{A}}^{*,*} is given by gluing of the action of ¯α+ηα\bar{\partial}_{\alpha}+\mathcal{L}_{\eta_{\alpha}} on 𝒜α,k\prescript{k}{}{\mathcal{A}}^{*,*}_{\alpha}. On the other hand, the elements

(4.6) 𝔩α:=¯α(ηα)+12[ηα,ηα]PkVα1,2(Wα)\mathfrak{l}_{\alpha}:=\bar{\partial}_{\alpha}(\eta_{\alpha})+\frac{1}{2}[\eta_{\alpha},\eta_{\alpha}]\in\prescript{k}{}{PV}^{-1,2}_{\alpha}(W_{\alpha})

glue to give a global element 𝔩PkV1,2(B)\mathfrak{l}\in\prescript{k}{}{PV}^{-1,2}(B), and for different kk’s, these elements are compatible. Computation shows that ¯2=[𝔩,]\bar{\partial}^{2}=[\mathfrak{l},\cdot] on PkV,\prescript{k}{}{PV}^{*,*} and ¯2=𝔩\bar{\partial}^{2}=\mathcal{L}_{\mathfrak{l}} on 𝒜,k\prescript{k}{}{\mathcal{A}}^{*,*}.

To glue the operators Δα\prescript{}{}{\Delta}_{\alpha}’s, we need to glue the local volume elements ωαk\prescript{k}{}{\omega}_{\alpha}’s to a global ωk\prescript{k}{}{\omega}. We consider an element of the form e𝔣αωαke^{\mathfrak{f}_{\alpha}\mathbin{\lrcorner}}\cdot\prescript{k}{}{\omega}_{\alpha}, where 𝔣αPkV0,0(Wα)\mathfrak{f}_{\alpha}\in\prescript{k}{}{PV}^{0,0}(W_{\alpha}) satisfies 𝔣α0(mod 𝐦)\mathfrak{f}_{\alpha}\equiv 0\ (\text{mod $\mathbf{m}$}). Given a kthk^{\text{th}}-order global volume element e𝔣αωαke^{\mathfrak{f}_{\alpha}\mathbin{\lrcorner}}\cdot\prescript{k}{}{\omega}_{\alpha}, we take a lifting e𝔣~αωαk+1e^{\tilde{\mathfrak{f}}_{\alpha}\mathbin{\lrcorner}}\cdot\prescript{k+1}{}{\omega}_{\alpha} such that

gαβk+1(e𝔣~αωαk+1)=e(𝔣~β𝔬αβ)ωβk+1,\prescript{k+1}{}{g}_{\alpha\beta}\big{(}e^{\tilde{\mathfrak{f}}_{\alpha}\mathbin{\lrcorner}}\cdot\prescript{k+1}{}{\omega}_{\alpha}\big{)}=e^{(\tilde{\mathfrak{f}}_{\beta}-\mathfrak{o}_{\alpha\beta})\mathbin{\lrcorner}}\cdot\prescript{k+1}{}{\omega}_{\beta},

for some element 𝔬αβP0V0,0(Wβ)𝐦k+1/𝐦k+2\mathfrak{o}_{\alpha\beta}\in\prescript{0}{}{PV}^{0,0}(W_{\beta})\otimes\mathbf{m}^{k+1}/\mathbf{m}^{k+2}. By construction, {𝔬αβ}αβ\{\mathfrak{o}_{\alpha\beta}\}_{\alpha\beta} gives a Čech 11-cycle in P0V0,0\prescript{0}{}{PV}^{0,0} which is exact. So there exist 𝔲α\mathfrak{u}_{\alpha}’s such that 𝔲β|Wαβ𝔲α|Wαβ=𝔬αβ\mathfrak{u}_{\beta}|_{W_{\alpha\beta}}-\mathfrak{u}_{\alpha}|_{W_{\alpha\beta}}=\mathfrak{o}_{\alpha\beta}, and we can modify 𝔣~α\tilde{\mathfrak{f}}_{\alpha} as 𝔣~α+𝔲α\tilde{\mathfrak{f}}_{\alpha}+\mathfrak{u}_{\alpha}, which gives the desired (k+1)st(k+1)^{\text{st}}-order volume element. Inductively, we can construct compatible volume elements ωk𝒜n,0k(B)\prescript{k}{}{\omega}\in\prescript{k}{\parallel}{\mathcal{A}}^{n,0}(B), kk\in\mathbb{N}. Any two such volume elements ωk\prescript{k}{}{\omega} and ωk\prescript{k}{}{\omega}^{\prime} differ by ωk=e𝔣ωk\prescript{k}{}{\omega}=e^{\mathfrak{f}\mathbin{\lrcorner}}\cdot\prescript{k}{}{\omega}^{\prime}, where 𝔣PkV0,0(B)\mathfrak{f}\in\prescript{k}{}{PV}^{0,0}(B) is some global element. Notice that ¯k(ωk)0\prescript{k}{}{\bar{\partial}}(\prescript{k}{}{\omega})\neq 0 unless mod 𝐦\mathbf{m}.

Using the volume element ω\prescript{}{}{\omega} (we omit the dependence on kk if there is no confusion), we may now define the global BV operator Δ\prescript{}{}{\Delta} by

(4.7) (Δφ)ω=(φω),(\prescript{}{}{\Delta}\varphi)\mathbin{\lrcorner}\prescript{}{}{\omega}=\prescript{}{}{\partial}(\varphi\mathbin{\lrcorner}\prescript{}{}{\omega}),

which can locally be written as Δαk+[𝔣α,]\prescript{k}{}{\Delta}_{\alpha}+[\mathfrak{f}_{\alpha},\cdot]. We have Δ2=0\prescript{}{}{\Delta}^{2}=0. The local elements

(4.8) 𝔫α:=Δαk(ηα)+¯α(𝔣α)+[ηα,𝔣α]\mathfrak{n}_{\alpha}:=\prescript{k}{}{\Delta}_{\alpha}(\eta_{\alpha})+\bar{\partial}_{\alpha}(\mathfrak{f}_{\alpha})+[\eta_{\alpha},\mathfrak{f}_{\alpha}]

glue to give a global element 𝔫PkV0,1(B)\mathfrak{n}\in\prescript{k}{}{PV}^{0,1}(B) which satisfies ¯Δ+Δ¯=[𝔫,]\bar{\partial}\prescript{}{}{\Delta}+\prescript{}{}{\Delta}\bar{\partial}=[\mathfrak{n},\cdot]. Also, the elements 𝔩\mathfrak{l} and 𝔫\mathfrak{n} satisfy the relation ¯(𝔫)+Δ(𝔩)=0\bar{\partial}(\mathfrak{n})+\prescript{}{}{\Delta}(\mathfrak{l})=0 by a local calculation.

In summary, we obtain pre-dgBV algebras (PkV,¯,Δ,)(\prescript{k}{}{PV},\bar{\partial},\prescript{}{}{\Delta},\wedge) and pre-dgas (𝒜k,¯,,)(\prescript{k}{}{\mathcal{A}},\bar{\partial},\prescript{}{}{\partial},\wedge) with a natural contraction action \mathbin{\lrcorner} of ¯k\prescript{k}{}{\bar{\partial}} on 𝒜,k\prescript{k}{}{\mathcal{A}}^{*,*}, and also volume elements ω\prescript{}{}{\omega}. We set

PV:=limkPkV,𝒜:=limk𝒜k,\prescript{}{}{PV}:=\varprojlim_{k}\prescript{k}{}{PV},\ \prescript{}{}{\mathcal{A}}:=\varprojlim_{k}\prescript{k}{}{\mathcal{A}},

and define a total de Rham operator 𝐝:𝒜,𝒜,\mathbf{d}\colon\prescript{}{}{\mathcal{A}}^{*,*}\rightarrow\prescript{}{}{\mathcal{A}}^{*,*} by

(4.9) 𝐝:=¯++𝔩,\mathbf{d}:=\bar{\partial}+\prescript{}{}{\partial}+\mathfrak{l}\mathbin{\lrcorner},

which preserves the filtration 𝒜,\prescript{}{\bullet}{\mathcal{A}}^{*,*}. Using the contraction ω:PV,𝒜+n,\prescript{}{}{\omega}\mathbin{\lrcorner}\colon\prescript{}{}{PV}^{*,*}\rightarrow\prescript{}{\parallel}{\mathcal{A}}^{*+n,*} to pull back the operator, we obtain the operator 𝐝=¯+Δ+(𝔩+𝔫)\mathbf{d}=\bar{\partial}+\prescript{}{}{\Delta}+(\mathfrak{l}+\mathfrak{n})\wedge acting on PV,\prescript{}{}{PV}^{*,*}. Direct computation shows that 𝐝2=0\mathbf{d}^{2}=0, and indeed it plays the role of the de Rham differential on a smooth manifold. Readers may consult [8, §4.2] for the computations and more details.

Definition 4.13.

We call PV,\prescript{}{}{PV}^{*,*} (resp. PkV,\prescript{k}{}{PV}^{*,*}) the sheaf of (resp. kthk^{\text{th}}-order) smooth relative polyvector fields over SS^{\dagger}, and 𝒜,\prescript{}{}{\mathcal{A}}^{*,*} (resp. 𝒜,k\prescript{k}{}{\mathcal{A}}^{*,*}) the sheaf of (resp. kthk^{\text{th}}-order) smooth forms over SS^{\dagger}. We denote the corresponding total complexes by PV=p+q=PVp,q\prescript{}{}{PV}^{*}=\bigoplus_{p+q=*}\prescript{}{}{PV}^{p,q} (resp. PkV\prescript{k}{}{PV}^{*}) and 𝒜=p+q=𝒜p,q\prescript{}{}{\mathcal{A}}^{*}=\bigoplus_{p+q=*}\prescript{}{}{\mathcal{A}}^{p,q} (resp. 𝒜k\prescript{k}{}{\mathcal{A}}^{*}).

4.4. Smoothing by solving the Maurer–Cartan equation

With the sheaf PV\prescript{}{}{PV}^{*} of pre-dgBV algebras defined, we can now consider the extended Maurer–Cartan equation

(4.10) (¯+tΔ)φ+12[φ,φ]+𝔩+t𝔫=0(\bar{\partial}+t\prescript{}{}{\Delta})\varphi+\frac{1}{2}[\varphi,\varphi]+\mathfrak{l}+t\mathfrak{n}=0

for φ=(φk)k\varphi=(\prescript{k}{}{\varphi})_{k}, where φkPkV0(B)[[t]]:=PkV0(B)[[t]]\prescript{k}{}{\varphi}\in\prescript{k}{}{PV}^{0}(B)[[t]]:=\prescript{k}{}{PV}^{0}(B)\otimes_{\mathbb{C}}\mathbb{C}[[t]]. Setting t=0t=0 gives the (classical) Maurer–Cartan equation

(4.11) ¯φ+12[φ,φ]+𝔩=0\bar{\partial}\varphi+\frac{1}{2}[\varphi,\varphi]+\mathfrak{l}=0

for φPV0(B)\varphi\in\prescript{}{}{PV}^{0}(B). To inductively solve these equations, we need two conditions, namely the holomorphic Poincaré Lemma and the Hodge-to-de Rham degeneracy.

We begin with the holomorphic Poincaré Lemma, which is a local condition on the sheaves j(Ω𝕍αk/)j_{*}(\Omega^{*}_{\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}/\mathbb{C}})’s. We consider the complex (j(Ω𝕍αk/)[u],α~)(j_{*}(\Omega^{*}_{\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}/\mathbb{C}})[u],\widetilde{\prescript{}{}{\partial}_{\alpha}}), where

α~(s=0lνsus):=s(ανs)us+sdlog(q)νsus1.\widetilde{\prescript{}{}{\partial}_{\alpha}}\left(\sum_{s=0}^{l}\nu_{s}u^{s}\right):=\sum_{s}(\prescript{}{}{\partial}_{\alpha}\nu_{s})u^{s}+sd\log(q)\wedge\nu_{s}u^{s-1}.

There is a natural exact sequence

(4.12) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔎¯αk\textstyle{\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j(Ω𝕍αk/)[u]\textstyle{j_{*}(\Omega^{*}_{\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}/\mathbb{C}})[u]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k,0~\scriptstyle{\widetilde{\prescript{k,0}{}{\flat}}}j(Ω𝕍α0/S0)\textstyle{j_{*}(\Omega^{*}_{\prescript{0}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{0}{}{S}^{\dagger}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0,\textstyle{0,}

where k,0~(s=0lνsus):=k,0(ν0)\widetilde{\prescript{k,0}{}{\flat}}(\sum_{s=0}^{l}\nu_{s}u^{s}):=\prescript{k,0}{}{\flat}(\nu_{0}) as elements in j(Ω𝕍α0/S0)j_{*}(\Omega^{*}_{\prescript{0}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{0}{}{S}^{\dagger}}).

Condition 4.14.

We say that the holomorphic Poincaré Lemma holds if at every point xX0x\in\prescript{0}{}{X}^{\dagger}, the complex (𝔎¯α,xk,α~)(\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha,x},\widetilde{\prescript{}{}{\partial}_{\alpha}}) is acyclic, where 𝔎¯α,xk\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha,x} denotes the stalk of 𝔎¯αk\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha} at xx.

The holomorphic Poincaré Lemma for our setting was proved in [28, proof of Thm. 4.1], but a gap was subsequently pointed out by Felten–Filip–Ruddat in [17], who used a different strategy to close the gap and give a correct proof in [17, Thm. 1.10]. From this condition, we can see that the cohomology sheaf (𝒦αk,αk)\mathcal{H}^{*}(\prescript{k}{\parallel}{\mathcal{K}}^{*}_{\alpha},\prescript{k}{}{\partial}_{\alpha}) is free over Rk=[q]/(qk+1)\prescript{k}{}{R}=\mathbb{C}[q]/(q^{k+1}) (cf. [34, Lem. 4.1]). We will need freeness of the cohomology H(𝒜k(B),𝐝)H^{*}(\prescript{k}{\parallel}{\mathcal{A}}^{*}(B),\mathbf{d}) over Rk\prescript{k}{}{R}, which can be seen by the following lemma (see [34] and [8, §4.3.2] for similar arguments).

Lemma 4.15.

Under Condition 4.14 (the holomorphic Poincaré Lemma), the natural map

k,0:H(𝒜k(B),𝐝)H(𝒜0(B),𝐝)\prescript{k,0}{}{\flat}\colon H^{*}(\prescript{k}{\parallel}{\mathcal{A}}^{*}(B),\mathbf{d})\rightarrow H^{*}(\prescript{0}{\parallel}{\mathcal{A}}^{*}(B),\mathbf{d})

is surjective for each k0k\geq 0.

Proof.

First of all, applying the functor ν\nu_{*} to the exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔎¯αk\textstyle{\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j(Ω𝕍αk/)[u]\textstyle{j_{*}(\Omega^{*}_{\prescript{k}{}{\mathbb{V}}_{\alpha}^{\dagger}/\mathbb{C}})[u]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k,0~\scriptstyle{\widetilde{\prescript{k,0}{}{\flat}}}j(Ω𝕍α0/S0)\textstyle{j_{*}(\Omega^{*}_{\prescript{0}{}{\mathbb{V}}_{\alpha}^{\dagger}/\prescript{0}{}{S}^{\dagger}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

gives the following exact sequence of sheaves on BB:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔎αk\textstyle{\prescript{k}{}{\mathfrak{K}}^{*}_{\alpha}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒦αk[u]\textstyle{\prescript{k}{}{\mathcal{K}}^{*}_{\alpha}[u]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k,0~\scriptstyle{\widetilde{\prescript{k,0}{}{\flat}}}𝒦α0\textstyle{\prescript{0}{}{\mathcal{K}}^{*}_{\alpha}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

This is true because every sheaf in the first exact sequence is a direct limit of coherent analytic sheaves, Rν!R\nu_{!} commutes with direct limits of sheaves, and Rν!=RνR\nu_{!}=R\nu_{*} as the fiber ν1(x)\nu^{-1}(x) is a compact Hausdorff topological space; see e.g. [32]. By taking a Cartan–Eilenberg resolution, we have the implication:

(𝔎¯α,xk,αk~) is acyclicRΓU((𝔎¯αk,αk~))=0(\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha,x},\widetilde{\prescript{k}{}{\partial}_{\alpha}})\text{ is acyclic}\Longrightarrow R\Gamma_{U}((\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha},\widetilde{\prescript{k}{}{\partial}_{\alpha}}))=0

for any open subset UU, where RΓUR\Gamma_{U} is the derived global section functor in the derived category of sheaves. In our case, U=ν1(W)U=\nu^{-1}(W) and we have RΓν1(W)=RΓWRνR\Gamma_{\nu^{-1}(W)}=R\Gamma_{W}\circ R\nu_{*}. Furthermore, we see that

Rν(𝔎¯αk,α~)=(𝔎αk,α~).R\nu_{*}(\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha},\widetilde{\partial_{\alpha}})=(\prescript{k}{}{\mathfrak{K}}^{*}_{\alpha},\widetilde{\partial_{\alpha}}).

This can be seen by taking a double complex C,C^{*,*} resolving (𝔎¯αk,α~)(\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha},\widetilde{\partial_{\alpha}}) such that ν(C,)\nu_{*}(C^{*,*}) computes Rν(𝔎¯αk,α~)R\nu_{*}(\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha},\widetilde{\partial_{\alpha}}). The spectral sequence associated to the double complex has the E1E_{1}-page given by Rqν(𝔎¯αpk)R^{q}\nu_{*}(\prescript{k}{}{\bar{\mathfrak{K}}}^{p}_{\alpha}), which is 0 if q>0q>0 because 𝔎¯αpk\prescript{k}{}{\bar{\mathfrak{K}}}^{p}_{\alpha} is a direct limit of coherent analytic sheaves. Therefore, ν(𝔎¯αk,α~)ν(C,)=Rν(𝔎¯αk,α~)\nu_{*}(\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha},\widetilde{\partial_{\alpha}})\rightarrow\nu_{*}(C^{*,*})=R\nu_{*}(\prescript{k}{}{\bar{\mathfrak{K}}}^{*}_{\alpha},\widetilde{\partial_{\alpha}}) is a quasi-isomorphism. Combining these, we obtain that RΓWi(𝔎αk,α~)=0R\Gamma^{i}_{W}(\prescript{k}{}{\mathfrak{K}}^{*}_{\alpha},\widetilde{\partial_{\alpha}})=0 for each ii.

Next, by Condition 4.7, (Ω|Wα𝔎αk)(\Omega^{*}|_{W_{\alpha}}\otimes\prescript{k}{}{\mathfrak{K}}^{*}_{\alpha}) is a resolution with a partition of unity, so the cohomology of the complex

(αk(W),¯α+α~):=((Ω|Wα𝔎αk)(W),¯α+α~)\left(\prescript{k}{}{\mathcal{B}}^{*}_{\alpha}(W),\bar{\partial}_{\alpha}+\widetilde{\prescript{}{}{\partial}_{\alpha}}\right):=\left((\Omega^{*}|_{W_{\alpha}}\otimes\prescript{k}{}{\mathfrak{K}}^{*}_{\alpha})(W),\bar{\partial}_{\alpha}+\widetilde{\prescript{}{}{\partial}_{\alpha}}\right)

computes RΓW(𝔎αk)R\Gamma_{W}(\prescript{k}{}{\mathfrak{K}}^{*}_{\alpha}). Through an isomorphism eηα:αkαke^{\eta_{\alpha}\mathbin{\lrcorner}}\colon\prescript{k}{}{\mathcal{B}}^{*}_{\alpha}\rightarrow\prescript{k}{}{\mathcal{B}}^{*}_{\alpha}, we can identify the operator:

𝐝α:=¯α+ηα+α~+ι¯α(ηα)+12[ηα,ηα]\mathbf{d}_{\alpha}:=\bar{\partial}_{\alpha}+\mathcal{L}_{\eta_{\alpha}}+\widetilde{\prescript{}{}{\partial}_{\alpha}}+\iota_{\bar{\partial}_{\alpha}(\eta_{\alpha})+\frac{1}{2}[\eta_{\alpha},\eta_{\alpha}]}

with ¯α+α~\bar{\partial}_{\alpha}+\widetilde{\prescript{}{}{\partial}_{\alpha}}, and hence deduce that (αk(W),𝐝α)(\prescript{k}{}{\mathcal{B}}_{\alpha}^{*}(W),\mathbf{d}_{\alpha}) is acyclic for any open subset WW.

Now, we consider the global sheaf (k,𝐝)(\prescript{k}{}{\mathcal{B}}^{*},\mathbf{d}) of complexes on BB obtained by gluing the local sheaves (αk,𝐝α)(\prescript{k}{}{\mathcal{B}}_{\alpha}^{*},\mathbf{d}_{\alpha}). We also have (𝒜k~,𝐝)(\widetilde{\prescript{k}{}{\mathcal{A}}^{*}},\mathbf{d}) obtained by gluing (Ω|Wα𝒦αk[u],𝐝α)(\Omega^{*}|_{W_{\alpha}}\otimes\prescript{k}{}{\mathcal{K}}^{*}_{\alpha}[u],\mathbf{d}_{\alpha}), and (𝒜0,𝐝)(\prescript{0}{\parallel}{\mathcal{A}}^{*},\mathbf{d}) obtained by gluing (Ω|Wα𝒦α0,𝐝α)(\Omega^{*}|_{W_{\alpha}}\otimes\prescript{0}{\parallel}{\mathcal{K}}^{*}_{\alpha},\mathbf{d}_{\alpha}). Then there is an exact sequence of complexes of sheaves

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k\textstyle{\prescript{k}{}{\mathcal{B}}^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒜k~\textstyle{\widetilde{\prescript{k}{}{\mathcal{A}}^{*}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒜0\textstyle{\prescript{0}{\parallel}{\mathcal{A}}^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

To see that the complex (k(B),𝐝)(\prescript{k}{}{\mathcal{B}}^{*}(B),\mathbf{d}) is acyclic, we consider the total Čech complex associated to the cover {Wα}α\{W_{\alpha}\}_{\alpha}. The associated spectral sequence has zero E1E_{1} page, thus (k(B),𝐝)(\prescript{k}{}{\mathcal{B}}^{*}(B),\mathbf{d}) is indeed acyclic. As a result, the map Hi(𝒜αk~(B),𝐝α)Hi(𝒜α0(B),𝐝α)H^{i}(\widetilde{\prescript{k}{}{\mathcal{A}}^{*}_{\alpha}}(B),\mathbf{d}_{\alpha})\rightarrow H^{i}(\prescript{0}{\parallel}{\mathcal{A}}^{*}_{\alpha}(B),\mathbf{d}_{\alpha}) is an isomorphism. Finally, surjectivity of the map k,0\prescript{k,0}{}{\flat} follows from the fact that the isomorphism Hi(𝒜αk~(B),𝐝α)Hi(𝒜α0(B),𝐝α)H^{i}(\widetilde{\prescript{k}{}{\mathcal{A}}^{*}_{\alpha}}(B),\mathbf{d}_{\alpha})\rightarrow H^{i}(\prescript{0}{\parallel}{\mathcal{A}}^{*}_{\alpha}(B),\mathbf{d}_{\alpha}) factors through k,0\prescript{k,0}{}{\flat}. ∎

The Hodge-to-de Rham degeneracy is a global Hodge-theoretic condition on X0\prescript{0}{}{X}^{\dagger}. We consider the Hodge filtration Frj(ΩX0/S0)=prj(ΩX0/S0p)F^{\geq r}j_{*}(\Omega^{*}_{\prescript{0}{}{X}^{\dagger}/\prescript{0}{}{S}^{\dagger}})=\bigoplus_{p\geq r}j_{*}(\Omega^{p}_{\prescript{0}{}{X}^{\dagger}/\prescript{0}{}{S}^{\dagger}}); the spectral sequence associated to it computes the hypercohomology of the complex of sheaves (j(ΩX0/S0),0)(j_{*}(\Omega^{*}_{\prescript{0}{}{X}^{\dagger}/\prescript{0}{}{S}^{\dagger}}),\prescript{0}{}{\partial})

Condition 4.16.

We say that the Hodge-to-de Rham degeneracy holds for X0\prescript{0}{}{X}^{\dagger} if the spectral sequence associated to the above Hodge filtration degenerates at E1E_{1}.

Under the assumption that (B,𝒫)(B,\mathscr{P}) is strongly simple (Definition 2.10), the Hodge-to-de Rham degeneracy for the maximally degenerate Calabi–Yau scheme X0\prescript{0}{}{X}^{\dagger} was proved in [28, Thm. 3.26]. This was later generalized to the case when (B,𝒫)(B,\mathscr{P}) is only simple (instead of strongly simple)666The subtle difference between the log Hodge group and the affine Hodge group when (B,𝒫)(B,\mathscr{P}) is just simple, instead of strongly simple, was studied in details by Ruddat in his thesis [42]. and further to log toroidal spaces in Felten–Filip–Ruddat [17] using different methods.

We consider the dgBV algebra P0V(B)[[t]]\prescript{0}{}{PV}^{*}(B)[[t]] equipped with the operator ¯+tΔ\bar{\partial}+t\prescript{}{}{\Delta}.

Lemma 4.17.

Under Condition 4.16 (the Hodge-to-de Rham degeneracy), H(P0V(B)[[t]],¯+tΔ)H^{*}(\prescript{0}{}{PV}^{*}(B)[[t]],\bar{\partial}+t\prescript{}{}{\Delta}) is a free [[t]]\mathbb{C}[[t]]-module.

Proof.

Recall that we are working with a good cover 𝒲={Wα}α\mathcal{W}=\{W_{\alpha}\}_{\alpha}, so that the inverse image Vα=ν1(Wα)V_{\alpha}=\nu^{-1}(W_{\alpha}) is Stein for each α\alpha. We have RΓν1(W)=RΓWRνR\Gamma_{\nu^{-1}(W)}=R\Gamma_{W}\circ R\nu_{*} and

Rν(j(ΩX0/S0),)=(𝒦0,).R\nu_{*}(j_{*}(\Omega^{*}_{\prescript{0}{}{X}^{\dagger}/\prescript{0}{}{S}^{\dagger}}),\prescript{}{}{\partial})=(\prescript{0}{\parallel}{\mathcal{K}}^{*},\prescript{}{}{\partial}).

If ν1(W)\nu^{-1}(W) is Stein, then RΓν1(W)(j(ΩX0/S0r))=Γν1(W)(j(ΩX0/S0r))R\Gamma_{\nu^{-1}(W)}(j_{*}(\Omega^{r}_{\prescript{0}{}{X}^{\dagger}/\prescript{0}{}{S}^{\dagger}}))=\Gamma_{\nu^{-1}(W)}(j_{*}(\Omega^{r}_{\prescript{0}{}{X}^{\dagger}/\prescript{0}{}{S}^{\dagger}})) and hence

RΓW(𝒦r0)=ΓW(𝒦r0).R\Gamma_{W}(\prescript{0}{\parallel}{\mathcal{K}}^{r})=\Gamma_{W}(\prescript{0}{\parallel}{\mathcal{K}}^{r}).

The hypercohomology of (j(ΩX0/S0),)(j_{*}(\Omega^{*}_{\prescript{0}{}{X}^{\dagger}/\prescript{0}{}{S}^{\dagger}}),\prescript{}{}{\partial}) is computed using the Čech double complex

𝒞ˇ(𝒱,j(ΩX0/S0))\check{\mathcal{C}}^{*}(\mathcal{V},j_{*}(\Omega^{*}_{\prescript{0}{}{X}^{\dagger}/\prescript{0}{}{S}^{\dagger}}))

with respect to the Stein open cover 𝒱={ν1(Wα)}α\mathcal{V}=\{\nu^{-1}(W_{\alpha})\}_{\alpha}. Similarly, the hypercohomology of the complex (𝒦0,)(\prescript{0}{\parallel}{\mathcal{K}}^{*},\prescript{}{}{\partial}) is computed using the Čech double complex 𝒞ˇ(𝒲,𝒦0)\check{\mathcal{C}}^{*}(\mathcal{W},\prescript{0}{\parallel}{\mathcal{K}}^{*}) with respect to the cover 𝒲={Wα}α\mathcal{W}=\{W_{\alpha}\}_{\alpha}; here, the Hodge filtration is induced from the filtration Fr𝒦0=pr𝒦p0F^{\geq r}\prescript{0}{\parallel}{\mathcal{K}}^{*}=\bigoplus_{p\geq r}\prescript{0}{\parallel}{\mathcal{K}}^{\geq p}.

These two Čech complexes, as well as their corresponding Hodge filtrations, are identified as 𝒦0(W)=j(ΩX0/S0r)(ν1(W))\prescript{0}{\parallel}{\mathcal{K}}^{*}(W)=j_{*}(\Omega^{r}_{\prescript{0}{}{X}^{\dagger}/\prescript{0}{}{S}^{\dagger}})(\nu^{-1}(W)) for each W=Wα1WαkW=W_{\alpha_{1}}\cap\cdots\cap W_{\alpha_{k}}. Hence, under Condition 4.16, we have E1E_{1} degeneracy also for 𝒞ˇ(𝒲,𝒦0)\check{\mathcal{C}}^{*}(\mathcal{W},\prescript{0}{\parallel}{\mathcal{K}}^{*}), or equivalently, that (𝒞ˇ(𝒲,𝒦0)[[t]],δ+t)(\check{\mathcal{C}}^{*}(\mathcal{W},\prescript{0}{\parallel}{\mathcal{K}}^{*})[[t]],\delta+t\prescript{}{}{\partial}) is a free [[t]]\mathbb{C}[[t]]-module. In view of the isomorphisms (𝒢0,Δ)(𝒦0,)(\prescript{0}{}{\mathcal{G}}^{*},\prescript{}{}{\Delta})\cong(\prescript{0}{\parallel}{\mathcal{K}},\prescript{}{}{\partial}) and

H(P0V(B)[[t]],¯+tΔ)H(𝒞ˇ(𝒲,𝒦0)[[t]],δ+t),H^{*}(\prescript{0}{}{PV}^{*}(B)[[t]],\bar{\partial}+t\prescript{}{}{\Delta})\cong H^{*}(\check{\mathcal{C}}^{*}(\mathcal{W},\prescript{0}{\parallel}{\mathcal{K}}^{*})[[t]],\delta+t\prescript{}{}{\partial}),

we conclude that H(P0V(B)[[t]],¯+tΔ)H^{*}(\prescript{0}{}{PV}^{*}(B)[[t]],\bar{\partial}+t\prescript{}{}{\Delta}) is a free [[t]]\mathbb{C}[[t]]-module as well. ∎

For the purpose of this paper, we restrict ourselves to the case that

φk=ϕk+t(fk),\prescript{k}{}{\varphi}=\prescript{k}{}{\phi}+t(\prescript{k}{}{f}),

where ϕkPkV1,1(B)\prescript{k}{}{\phi}\in\prescript{k}{}{PV}^{-1,1}(B) and fkPkV0,0(B)\prescript{k}{}{f}\in\prescript{k}{}{PV}^{0,0}(B). The extended Maurer-Cartan equation (4.10) can be decomposed, according to orders in tt, into the (classical) Maurer–Cartan equation (4.11) for ϕk\prescript{k}{}{\phi} and the equation

(4.13) ¯(fk)+[ϕk,fk]+Δ(ϕk)+𝔫=0.\bar{\partial}(\prescript{k}{}{f})+[\prescript{k}{}{\phi},\prescript{k}{}{f}]+\prescript{}{}{\Delta}(\prescript{k}{}{\phi})+\mathfrak{n}=0.
Theorem 4.18.

Suppose that both Conditions 4.14 and 4.16 hold. Then for any kthk^{\text{th}}-order solution φk=ϕk+t(fk)\prescript{k}{}{\varphi}=\prescript{k}{}{\phi}+t(\prescript{k}{}{f}) to the extended Maurer–Cartan equation (4.10), there exists a (k+1)st(k+1)^{\text{st}}-order solution φk+1=ϕk+1+t(fk+1)\prescript{k+1}{}{\varphi}=\prescript{k+1}{}{\phi}+t(\prescript{k+1}{}{f}) to (4.10) lifting φk\prescript{k}{}{\varphi}. The same statement holds for the Maurer–Cartan equation (4.11) if we restrict to ϕkPkV1,1(B)\prescript{k}{}{\phi}\in\prescript{k}{}{PV}^{-1,1}(B).

Proof.

The first statement follows from [8, Thm. 5.6] and [8, Lem. 5.12]: Starting with a kthk^{\text{th}}-order solution φk=ϕk+t(fk)\prescript{k}{}{\varphi}=\prescript{k}{}{\phi}+t(\prescript{k}{}{f}) for (4.10), one can always use [8, Thm. 5.6] to lift it to a general φk+1Pk+1V0(B)[[t]]\prescript{k+1}{}{\varphi}\in\prescript{k+1}{}{PV}^{0}(B)[[t]]. The argument in [8, Lem. 5.12] shows that we can choose φk+1\prescript{k+1}{}{\varphi} such that the component of φk+1|t=0\prescript{k+1}{}{\varphi}|_{t=0} in Pk+1V0,0(B)\prescript{k+1}{}{PV}^{0,0}(B) is zero. As a result, the component of ϕk+1+t(fk+1)\prescript{k+1}{}{\phi}+t(\prescript{k+1}{}{f}) in Pk+1V1,1(B)t(Pk+1V0,0(B))\prescript{k+1}{}{PV}^{-1,1}(B)\otimes t(\prescript{k+1}{}{PV}^{0,0}(B)) is again a solution to (4.10).

For the second statement, we argue that, given ϕk\prescript{k}{}{\phi}, there always exists fkPkV0,0(B)\prescript{k}{}{f}\in\prescript{k}{}{PV}^{0,0}(B) such that ϕk+t(fk)\prescript{k}{}{\phi}+t(\prescript{k}{}{f}) is a solution to (4.10). We need to solve the equation (4.13) by induction on the order kk. The initial case is trivial by taking f0=0\prescript{0}{}{f}=0. Suppose the equation can be solved for fj1\prescript{j-1}{}{f}. Then we take an arbitrary lifting f~j{\prescript{j}{}{\tilde{f}}} to the jthj^{\text{th}}-order. We can define an element 𝔬P0V0,0(B)\mathfrak{o}\in\prescript{0}{}{PV}^{0,0}(B) by

qj𝔬=¯(f~j)+[ϕj,f~j]+Δ(ϕj)+𝔫,q^{j}\mathfrak{o}=\bar{\partial}(\prescript{j}{}{\tilde{f}})+[\prescript{j}{}{\phi},\prescript{j}{}{\tilde{f}}]+\prescript{}{}{\Delta}(\prescript{j}{}{\phi})+\mathfrak{n},

which satisfies ¯(𝔬)=0\bar{\partial}(\mathfrak{o})=0. Therefore, the class [𝔬][\mathfrak{o}] lies in the cohomology

H1(P0V0,,¯)H1(X0,𝒪)H1(B,),H^{1}(\prescript{0}{}{PV}^{0,*},\bar{\partial})\cong H^{1}(\prescript{0}{}{X},\mathcal{O})\cong H^{1}(B,\mathbb{C}),

where the last equivalence is from [27, Prop. 2.37]. By our assumption in §2, we have H1(B,)=0H^{1}(B,\mathbb{C})=0, and hence we can find an element f˘\breve{f} such that ¯(f˘)=𝔬\bar{\partial}(\breve{f})=\mathfrak{o}. Letting fj=f~j+qjf˘(mod qj+1)\prescript{j}{}{f}=\prescript{j}{}{\tilde{f}}+q^{j}\cdot\breve{f}\ (\text{mod $q^{j+1}$}) proves the induction step from the (j1)st(j-1)^{\text{st}}-order to the jthj^{\text{th}}-order. Now, applying the first statement, we can lift the solution φk:=ϕk+t(fk)\prescript{k}{}{\varphi}:=\prescript{k}{}{\phi}+t(\prescript{k}{}{f}) to φk+1=ϕk+1+t(fk+1)\prescript{k+1}{}{\varphi}=\prescript{k+1}{}{\phi}+t(\prescript{k+1}{}{f}) which satisfies equation (4.10), and hence ϕk+1\prescript{k+1}{}{\phi} solves (4.11). ∎

From Theorem 4.18, we obtain a solution ϕPV1,1(B)\phi\in\prescript{}{}{PV}^{-1,1}(B) to the Maurer–Cartan equation (4.11), from which we obtain the sheaves ker(¯+[ϕ,])PkV,\ker(\bar{\partial}+[\phi,\cdot])\subset\prescript{k}{}{PV}^{*,*} and ker(¯+ϕ)𝒜,k\ker(\bar{\partial}+\mathcal{L}_{\phi})\subset\prescript{k}{\parallel}{\mathcal{A}}^{*,*} over BB. These sheaves are locally isomorphic to 𝒢αk\prescript{k}{}{\mathcal{G}}^{*}_{\alpha} and 𝒦αk\prescript{k}{\parallel}{\mathcal{K}}^{*}_{\alpha}, so we may treat them as obtained from gluing of the local sheaves 𝒢αk\prescript{k}{}{\mathcal{G}}^{*}_{\alpha}’s and 𝒦αk\prescript{k}{\parallel}{\mathcal{K}}^{*}_{\alpha}’s. From these, we can extract consistent and compatible gluings Φαβk:𝕍αk|Vαβ𝕍βk|Vαβ\prescript{k}{}{\varPhi}_{\alpha\beta}\colon\prescript{k}{}{\mathbb{V}}^{\dagger}_{\alpha}|_{V_{\alpha\beta}}\rightarrow\prescript{k}{}{\mathbb{V}}^{\dagger}_{\beta}|_{V_{\alpha\beta}} satisfying the cocycle condition, and hence obtain a kk-th order thichening Xk\prescript{k}{}{X} of X0\prescript{0}{}{X} over Sk\prescript{k}{}{S}^{\dagger}; see [8, §5.3]. Also, efωe^{f}\mathbin{\lrcorner}\prescript{}{}{\omega}, as a section of ker(¯+ϕ)\ker(\bar{\partial}+\mathcal{L}_{\phi}) over BB, defines a holomorphic volume form on the kk-th order thickening Xk\prescript{k}{}{X}.

4.4.1. Normalized volume form

For later purposes, we need to further normalize the holomorphic volume

Ω:=efωker(¯+ϕ)(B)𝒜n,0k(B)\varOmega:=e^{f}\mathbin{\lrcorner}\prescript{}{}{\omega}\in\ker(\bar{\partial}+\mathcal{L}_{\phi})(B)\subset\prescript{k}{\parallel}{\mathcal{A}}^{n,0}(B)

by adding a suitable power series h(q)(q)[[q]]h(q)\in(q)\subset\mathbb{C}[[q]] to ff, so that the condition that Tefω=1\int_{T}e^{f}\mathbin{\lrcorner}\prescript{}{}{\omega}=1, where TT is a nearby nn-torus in the smoothing, is satisfied.

The kthk^{\text{th}}-order Hodge bundle over Specan([q]/qk+1)\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[q]/q^{k+1}) is defined as the cohomology

k:=Hn(𝒜k,𝐝),\prescript{k}{}{\mathcal{H}}:=H^{n}(\prescript{k}{\parallel}{\mathcal{A}}^{*},\mathbf{d}),

equipped with a Gauss–Manin connection k\prescript{k}{}{\nabla}, where logqk\prescript{k}{}{\nabla}_{\frac{\partial}{\partial\log q}} is the connecting homomorphism of the long exact sequence associated to

(4.14) 0𝒜1kdlogq𝒜k𝒜k0;0\rightarrow\prescript{k}{\parallel}{\mathcal{A}}^{*-1}\otimes_{\mathbb{C}}\mathbb{C}\langle d\log q\rangle\rightarrow\prescript{k}{}{\mathcal{A}}^{*}\rightarrow\prescript{k}{\parallel}{\mathcal{A}}^{*}\rightarrow 0;

here dlogq\mathbb{C}\langle d\log q\rangle is the 11-dimensional graded vector space spanned by the degree 11 element dlogqd\log q. We denote ^:=limkk\widehat{\mathcal{H}}:=\varprojlim_{k}\prescript{k}{}{\mathcal{H}}. Restricting to the 0th0^{\text{th}}-order, we have N=logq0N=\prescript{0}{}{\nabla}_{\frac{\partial}{\partial\log q}}, which is a nilpotent operator acting on 0=Hn(𝒜0)n(X,jΩX/)\prescript{0}{}{\mathcal{H}}=H^{n}(\prescript{0}{\parallel}{\mathcal{A}}^{*})\cong\mathbb{H}^{n}(\prescript{}{}{X},j_{*}\Omega^{*}_{\prescript{}{}{X}^{\dagger}/\mathbb{C}^{\dagger}}), where X=X0\prescript{}{}{X}=\prescript{0}{}{X}. If we consider the top cohomoloy H2n(𝒜0)H^{2n}(\prescript{0}{\parallel}{\mathcal{A}}^{*}), which is 11-dimensional, we see that N=logq0=0N=\prescript{0}{}{\nabla}_{\frac{\partial}{\partial\log q}}=0. So logqk\prescript{k}{}{\nabla}_{\frac{\partial}{\partial\log q}} is a flat connection without log poles at q=0q=0. Hence, we can find a basis (order by order in qq) to identify H2n(𝒜k)H2n(𝒜0)[q]/qk+1H^{2n}(\prescript{k}{\parallel}{\mathcal{A}}^{*})\cong H^{2n}(\prescript{0}{\parallel}{\mathcal{A}}^{*})\otimes\mathbb{C}[q]/q^{k+1}, which also trivializes the flat connection \prescript{}{}{\nabla} as logq\frac{\partial}{\partial\log q}.

Since Hn(B,)H^{n}(B,\mathbb{C})\cong\mathbb{C}, we can fix a non-zero generator and choose a representative ϱΩn(B)\varrho\in\Omega^{n}(B). Then the element ϱ1𝒜nk(B)\varrho\otimes 1\in\prescript{k}{\parallel}{\mathcal{A}}^{n}(B) (which may simply be written as ϱ\varrho) represents a section [ϱ][\varrho] in ^\widehat{\mathcal{H}}. A direct computation shows that [ϱ]=0\prescript{}{}{\nabla}[\varrho]=0, i.e. it is a flat section to all orders. The pairing with the 0th0^{\text{th}}-order volume form ω0\prescript{0}{}{\omega} gives a non-zero element [ω0ϱ][\prescript{0}{}{\omega}\wedge\varrho] in H2n(𝒜0)H^{2n}(\prescript{0}{\parallel}{\mathcal{A}}^{*}).

Definition 4.19.

The volume form Ω=efω\varOmega=e^{f}\mathbin{\lrcorner}\prescript{}{}{\omega} is said to be normalized if [Ωϱ][\varOmega\wedge\varrho] is flat under \prescript{}{}{\nabla}.

In other words, we can write [Ωϱ]=[ω0ϱ][\varOmega\wedge\varrho]=[\prescript{0}{}{\omega}\wedge\varrho] under the identification

H2n(𝒜k)H2n(𝒜0)[q]/qk+1.H^{2n}(\prescript{k}{\parallel}{\mathcal{A}}^{*})\cong H^{2n}(\prescript{0}{\parallel}{\mathcal{A}}^{*})\otimes\mathbb{C}[q]/q^{k+1}.

By modifying ff to f+h(q)f+h(q), this can always be achieved. Further, after the modification, φ=ϕ+tf\varphi=\phi+tf still solves (4.10).

5. From smoothing of Calabi–Yau varieties to tropical geometry

5.1. Tropical differential forms

To tropicalize the pre-dgBV algebra PV,\prescript{}{}{PV}^{*,*}, we need to replace the Thom–Whitney resolution used in [8] by a geometric resolution. To do so, we first need to recall some background materials from our previous works [7, §4.2.3] and [9, §3.2]. Of crucial importance is the notion of differential forms with asymptotic support (which will be called tropical differential forms in this paper) that originated from multi-valued Morse theory and Witten deformations. Such differential forms can be regarded as distribution-valued forms supported on tropical polyhedral subsets. This key notion allows us to develop tropical intersection theory via differential forms, and in particular, define the intersection pairing between possibly non-transversal tropical polyhedral subsets simply using the wedge product.

Let UU be an open subset of MM. We consider the space Ωk(U):=Γ(U×>0,kTU)\Omega^{k}_{\hslash}(U):=\Gamma(U\times\mathbb{R}_{>0},\bigwedge^{\raisebox{-1.20552pt}{\scriptsize$k$}}T^{\vee}U), where we take 𝒞\mathcal{C}^{\infty} sections of kTU\bigwedge^{\raisebox{-1.20552pt}{\scriptsize$k$}}T^{\vee}U and \hslash is a coordinate on >0\mathbb{R}_{>0}. Let 𝒲k(U)Ωk(U)\mathcal{W}^{k}_{-\infty}(U)\subset\Omega^{k}_{\hslash}(U) be the subset of kk-forms α\alpha such that, for each qUq\in U, there exist a neighborhood qVUq\in V\subset U, constants Dj,VD_{j,V}, cVc_{V} and a sufficiently small real number 0>0\hslash_{0}>0 such that jαL(V)Dj,VecV/\|\nabla^{j}\alpha\|_{L^{\infty}(V)}\leq D_{j,V}e^{-c_{V}/\hslash} for all j0j\geq 0 and for 0<<00<\hslash<\hslash_{0}; here, the LL^{\infty}-norm is defined by αL(V)=supxVα(x)\|\alpha\|_{L^{\infty}(V)}=\sup_{x\in V}\|\alpha(x)\| for any section α\alpha of the tensor bundle TUkTUlTU^{\otimes k}\otimes T^{\vee}U^{\otimes l}, where we fix a constant metric on MM and use the induced metric on TUkTUlTU^{\otimes k}\otimes T^{\vee}U^{\otimes l}; j\nabla^{j} denotes an operator of the form xl1xlj\nabla_{\frac{\partial}{\partial x_{l_{1}}}}\cdots\nabla_{\frac{\partial}{\partial x_{l_{j}}}}, where \nabla is a torsion-free, flat connection defining an affine structure on UU and x=(x1,,xn)x=(x_{1},\dots,x_{n}) is an affine coordinate system (note that \nabla is not the Gauss–Manin connection in the previous section). Similarly, let 𝒲k(U)Ωk(U)\mathcal{W}^{k}_{\infty}(U)\subset\Omega^{k}_{\hslash}(U) be the set of kk-forms α\alpha such that, for each qUq\in U, there exist a neighborhood qVUq\in V\subset U, a constant Dj,VD_{j,V}, Nj,V>0N_{j,V}\in\mathbb{Z}_{>0} and a sufficiently small real number 0>0\hslash_{0}>0 such that jαL(V)Dj,VNj,V\|\nabla^{j}\alpha\|_{L^{\infty}(V)}\leq D_{j,V}\hslash^{-N_{j,V}} for all j0j\geq 0 and for 0<<00<\hslash<\hslash_{0}.

The assignment U𝒲k(U)U\mapsto\mathcal{W}^{k}_{-\infty}(U) (resp. U𝒲k(U)U\mapsto\mathcal{W}^{k}_{\infty}(U)) defines a sheaf 𝒲k\mathcal{W}^{k}_{-\infty} (resp. 𝒲k\mathcal{W}^{k}_{\infty}) on MM ([7, Defs. 4.15 & 4.16]). Note that 𝒲k\mathcal{W}^{k}_{-\infty} and 𝒲k\mathcal{W}^{k}_{\infty} are closed under the wedge product, x\nabla_{\frac{\partial}{\partial x}} and the de Rham differential dd. Since 𝒲k\mathcal{W}^{k}_{-\infty} is a dg ideal of 𝒲k\mathcal{W}^{k}_{\infty}, the quotient 𝒲/𝒲\mathcal{W}^{*}_{\infty}/\mathcal{W}^{*}_{-\infty} is a sheaf of dgas when equipped with the de Rham differential.

Now suppose UU is a convex open set. By a tropical polyhedral subset of UU, we mean a connected convex subset of UU which is defined by finitely many affine equations or inequalities over \mathbb{Q} of the form a1x1++anxnba_{1}x_{1}+\cdots+a_{n}x_{n}\leq b.

Definition 5.1 ([7], Def. 4.19).

A kk-form α𝒲k(U)\alpha\in\mathcal{W}_{\infty}^{k}(U) is said to have asymptotic support on a closed codimension kk tropical polyhedral subset PUP\subset U with weight ss\in\mathbb{Z}, denoted as α𝒲P,s(U)\alpha\in\mathcal{W}_{P,s}(U), if the following conditions are satisfied:

  1. (1)

    For any pUPp\in U\setminus P, there is a neighborhood pVUPp\in V\subset U\setminus P such that α|V𝒲k(V)\alpha|_{V}\in\mathcal{W}_{-\infty}^{k}(V).

  2. (2)

    There exists a neighborhood WPUW_{P}\subset U of PP such that α=h(x,)νP+η\alpha=h(x,\hslash)\nu_{P}+\eta on WPW_{P}, where νPkN\nu_{P}\in\bigwedge^{k}N is a non-zero affine kk-form (defined up to non-zero constant) which is normal to PP, h(x,)C(WP×)>0h(x,\hslash)\in C^{\infty}(W_{P}\times{}_{>0}) and η𝒲k(WP)\eta\in\mathcal{W}_{-\infty}^{k}(W_{P}).

  3. (3)

    For any pPp\in P, there exists a convex neighborhood pVUp\in V\subset U equipped with an affine coordinate system x=(x1,,xn)x=(x_{1},\dots,x_{n}) such that x:=(x1,,xk)x^{\prime}:=(x_{1},\dots,x_{k}) parametrizes codimension kk affine linear subspaces of VV parallel to PP, with x=0x^{\prime}=0 corresponding to the subspace containing PP. With the foliation {(PV,x)}xNV\{(P_{V,x^{\prime}})\}_{x^{\prime}\in N_{V}}, where PV,x={(x1,,xn)V|(x1,,xk)=x}P_{V,x^{\prime}}=\{(x_{1},\dots,x_{n})\in V\ |\ (x_{1},\dots,x_{k})=x^{\prime}\} and NVN_{V} is the normal bundle of VV, we require that, for all j0j\in\mathbb{Z}_{\geq 0} and multi-indices β=(β1,,βk)0k\beta=(\beta_{1},\dots,\beta_{k})\in\mathbb{Z}_{\geq 0}^{k}, the estimate

    x(x)β(supPV,x|j(ινPα)|)νPDj,V,βj+s|β|k2\int_{x^{\prime}}(x^{\prime})^{\beta}\left(\sup_{P_{V,x^{\prime}}}|\nabla^{j}(\iota_{\nu_{P}^{\vee}}\alpha)|\right)\nu_{P}\leq D_{j,V,\beta}\hslash^{-\frac{j+s-|\beta|-k}{2}}

    holds for some constant Dj,V,βD_{j,V,\beta} and ss\in\mathbb{Z}, where |β|=lβl|\beta|=\sum_{l}\beta_{l} and νP=x1xk\nu_{P}^{\vee}=\frac{\partial}{\partial x_{1}}\wedge\cdots\wedge\frac{\partial}{\partial x_{k}}.777For k=0k=0, we use the convention that νP=10N=\nu_{P}=1\in\bigwedge^{0}N=\real and also set νP=1\nu_{P}^{\vee}=1.

Observe that xl𝒲P,s(U)𝒲P,s+1(U)\nabla_{\frac{\partial}{\partial x_{l}}}\mathcal{W}_{P,s}(U)\subset\mathcal{W}_{P,s+1}(U) and (x)β𝒲P,s(U)𝒲P,s|β|(U)(x^{\prime})^{\beta}\mathcal{W}_{P,s}(U)\subset\mathcal{W}_{P,s-|\beta|}(U). It follows that

(x)βxl1xlj𝒲P,s(U)𝒲P,s+j|β|(U).(x^{\prime})^{\beta}\nabla_{\frac{\partial}{\partial x_{l_{1}}}}\cdots\nabla_{\frac{\partial}{\partial x_{l_{j}}}}\mathcal{W}_{P,s}(U)\subset\mathcal{W}_{P,s+j-|\beta|}(U).

The weight ss defines a filtration of 𝒲k\mathcal{W}^{k}_{\infty} (we drop the UU dependence from the notation whenever it is clear from the context):888Note that kk is equal to the codimension of PUP\subset U.

𝒲k𝒲P,1𝒲P,0𝒲P,1𝒲kΩk(U).\mathcal{W}_{-\infty}^{k}\subset\cdots\subset\mathcal{W}_{P,-1}\subset\mathcal{W}_{P,0}\subset\mathcal{W}_{P,1}\subset\cdots\subset\mathcal{W}_{\infty}^{k}\subset\Omega^{k}_{\hslash}(U).

This filtration, which keeps track of the polynomial order of \hslash for kk-forms with asymptotic support on PP, provides a convenient tool to express and prove results in asymptotic analysis.

Definition 5.2 ([9], Def. 3.10).

A differential kk-form α\alpha is in 𝒲~sk(U)\tilde{\mathcal{W}}_{s}^{k}(U) if there exist polyhedral subsets P1,,PlUP_{1},\dots,P_{l}\subset U of codimension kk such that αj=1l𝒲Pj,s(U)\alpha\in\sum_{j=1}^{l}\mathcal{W}_{P_{j},s}(U). If, moreover, dα𝒲~s+1k+1(U)d\alpha\in\tilde{\mathcal{W}}_{s+1}^{k+1}(U), then we write α𝒲sk(U)\alpha\in\mathcal{W}_{s}^{k}(U). For every ss\in\mathbb{Z}, let 𝒲s(U)=k𝒲s+kk(U)\mathcal{W}_{s}^{*}(U)=\bigoplus_{k}\mathcal{W}_{s+k}^{k}(U).

Example 5.3.

Let U=U=\real and xx be an affine coordinate on UU. Then we consider the \hslash-dependent 11-form

δ:=(1π)12ex2dx.\delta:=\left(\frac{1}{\hslash\pi}\right)^{\frac{1}{2}}e^{-\frac{x^{2}}{\hslash}}dx.

Direct calculations in [7, Lem 4.12] showed that δ𝒲11(U)\delta\in\mathcal{W}_{1}^{1}(U) has asymptotic support on the hyperplane PP defined by x=0x=0.

The hyperplane PP separates UU into two chambers H+H_{+} and HH_{-}. If we fix a base point in HH_{-} and apply the integral operator II in [7, Lem. 4.23], we obtain I(δ)W00(U)I(\delta)\in W^{0}_{0}(U) which has asymptotic support on H+PH_{+}\cup P, playing the role of a step function.

Taking finite products of elements of the above form, we obtain α𝒲kk(U)\alpha\in\mathcal{W}^{k}_{k}(U) with asymptotic support on arbitrary tropical polyhedral subsets of UU. Any forms obtained from a finite number of steps of applying the differential dd, applying the integral operator II and taking wedge product are in W0(U)W^{*}_{0}(U).

We say that two closed tropical polyhedral subsets P1,P2UP_{1},P_{2}\subset U of codimension k1,k2k_{1},k_{2} intersect transversally if the affine subspaces of codimension k1k_{1} and k2k_{2} which contain P1P_{1} and P2P_{2}, respectively, intersect transversally. This definition applies also when P1P2=P_{1}\cap P_{2}=\emptyset or Pi\partial P_{i}\neq\emptyset.

Lemma 5.4 ([7, Lem. 4.22]).
  1. (1)

    Let P1,P2,PUP_{1},P_{2},P\subset U be closed tropical polyhedral subsets of codimension k1k_{1}, k2k_{2} and k1+k2k_{1}+k_{2}, respectively, such that PP contains P1P2P_{1}\cap P_{2} and is normal to νP1νP2\nu_{P_{1}}\wedge\nu_{P_{2}}. Then 𝒲P1,s(U)𝒲P2,r(U)𝒲P,r+s(U)\mathcal{W}_{P_{1},s}(U)\wedge\mathcal{W}_{P_{2},r}(U)\subset\mathcal{W}_{P,r+s}(U) if P1P_{1} and P2P_{2} intersect transversally with P1P2P_{1}\cap P_{2}\neq\emptyset, and 𝒲P1,s(U)𝒲P2,r(U)𝒲k1+k2(U)\mathcal{W}_{P_{1},s}(U)\wedge\mathcal{W}_{P_{2},r}(U)\subset\mathcal{W}_{-\infty}^{k_{1}+k_{2}}(U) otherwise.

  2. (2)

    We have 𝒲s1k1(U)𝒲s2k2(U)𝒲s1+s2k1+k2(U)\mathcal{W}_{s_{1}}^{k_{1}}(U)\wedge\mathcal{W}_{s_{2}}^{k_{2}}(U)\subset\mathcal{W}_{s_{1}+s_{2}}^{k_{1}+k_{2}}(U). In particular, 𝒲0(U)𝒲(U)\mathcal{W}_{0}^{*}(U)\subset\mathcal{W}_{\infty}^{*}(U) is a dg subalgebra and 𝒲1(U)𝒲0(U)\mathcal{W}_{-1}^{*}(U)\subset\mathcal{W}_{0}^{*}(U) is a dg ideal.

Definition 5.5.

Let 𝒲s\mathcal{W}_{s}^{*} be the sheafification of the presheaf defined by U𝒲s(U)U\mapsto\mathcal{W}_{s}^{*}(U). We call the quotient sheaf 𝒯:=𝒲0/𝒲1\mathscr{T}^{*}:=\mathcal{W}_{0}^{*}/\mathcal{W}_{-1}^{*} the sheaf of tropical differential forms, which is a sheaf of dgas on MM with structures (,d)(\wedge,d).

From [9, Lem. 3.6], we learn that ¯𝒯\underline{\real}\rightarrow\mathscr{T}^{*} is a resolution. Furthermore, given any point xUx\in U and a sufficiently small neighborhood xWUx\in W\subset U, we can show that there exists f𝒲00(W)f\in\mathcal{W}_{0}^{0}(W) with compact support in WW and satisfying f1f\equiv 1 near xx (using an argument similar to the proof of Lemma 3.10). Therefore, 𝒯\mathscr{T}^{*} has a partition of unity subordinate to a given open cover. Replacing the sheaf of de Rham differential forms on Λρ1,𝒬τ,\Lambda_{\rho_{1},\real}^{*}\oplus\mathscr{Q}_{\tau,\real} by the sheaf 𝒯\mathscr{T}^{*} of tropical differential forms, we can construct a particular complex on the integral tropical manifold BB satisfying Condition 4.7, which dictates the tropical geometry of BB.

Definition 5.6.

Given a point xx as in §3.3.2 (with a chart as in equation (3.10)), the stalk of 𝒯\mathscr{T}^{*} at xx is defined as 𝒯x:=(𝚡1𝒯)x\mathscr{T}^{*}_{x}:=(\mathtt{x}^{-1}\mathscr{T}^{*})_{x}. This defines the complex (𝒯,d)(\mathscr{T}^{*},d) (or simply 𝒯\mathscr{T}^{*}) of monodromy invariant tropical differential forms on BB. A section α𝒯(W)\alpha\in\mathscr{T}^{*}(W) is a collection of elements αx𝒯x\alpha_{x}\in\mathscr{T}^{*}_{x}, xWx\in W such that each αx\alpha_{x} can be represented by 𝚡1βx\mathtt{x}^{-1}\beta_{x} in a small neighborhood Ux𝚙1(𝚄x)U_{x}\subset\mathtt{p}^{-1}(\mathtt{U}_{x}) for some tropical differential form βx\beta_{x} on 𝚄x\mathtt{U}_{x}, and satisfies the relation αx~=𝚡~1(𝚙βx)\alpha_{\tilde{x}}=\tilde{\mathtt{x}}^{-1}(\mathtt{p}^{*}\beta_{x}) in 𝒯x~\mathscr{T}^{*}_{\tilde{x}} for every x~Ux\tilde{x}\in U_{x}.

Notice that the definition of 𝒯\mathscr{T}^{*} requires the projection map 𝚙\mathtt{p} in equation (3.11) to be affine, while that of Ω\Omega^{*} in §3.3.2 does not. But like Ω\Omega^{*}, 𝒯\mathscr{T}^{*} satisfies Condition 4.7 and can be used for the purpose of gluing the sheaf PV\prescript{}{}{PV}^{*} of dgBV algebras in §4.3. In the rest of this section, we shall use the notations PV\prescript{}{}{PV}^{*} and 𝒜\prescript{}{}{\mathcal{A}}^{*} to denote the complexes of sheaves constructed using 𝒯\mathscr{T}^{*}.

5.2. The semi-flat dgBV algebra and its comparison with the pre-dgBV algebra PV,\prescript{}{}{PV}^{*,*}

In this section, we define a twisting of the semi-flat dgBV algebra by the slab functions (or initial wall-crossing factors) in §2.4, and compare it with the dgBV algebra we constructed in §4.3 using gluing of local smoothing models. The key result is Lemma 5.10, which is an important step in the proof of our main result.

We start by recalling some notations from §2.4. Recall that for each vertex vv, we fix a representative φv:Uv\varphi_{v}\colon U_{v}\rightarrow\real of the strictly convex multi-valued piecewise linear function φH0(B,𝒫𝒫)\varphi\in H^{0}(B,\mathcal{MPL}_{\mathscr{P}}) to define the cone CvC_{v} and the monoid PvP_{v}. The natural projection TvTvT_{v}\oplus\mathbb{Z}\rightarrow T_{v} induces a surjective ring homomorphism [ρ1Pv][ρ1Σv]\mathbb{C}[\rho^{-1}P_{v}]\rightarrow\mathbb{C}[\rho^{-1}\Sigma_{v}]; we denote by m¯ρ1Σv\bar{m}\in\rho^{-1}\Sigma_{v} the image of mρ1Pvm\in\rho^{-1}P_{v} under the natural projection. We consider 𝐕(τ)v:=Specan([τ1Pv])\mathbf{V}(\tau)_{v}:=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}P_{v}]) for some τ\tau containing vv, and write zmz^{m} for the function corresponding to mτ1Pvm\in\tau^{-1}P_{v}. The element ϱ\varrho together with the corresponding function zϱz^{\varrho} determine a family Specan([τ1Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}P_{v}])\rightarrow\mathbb{C}, whose central fiber is given by Specan([τ1Σv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{v}]). The variety 𝐕(τ)v=Specan([τ1Pv])\mathbf{V}(\tau)_{v}=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}P_{v}]) is equipped with the divisorial log structure induced by Specan([τ1Σv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\tau^{-1}\Sigma_{v}]), which is log smooth. We write 𝐕(τ)v\mathbf{V}(\tau)_{v}^{\dagger} if we need to emphasize the log structure.

Since BB is orientable, we can choose a nowhere vanishing integral element μΓ(B𝒮e,nTB,)\mu\in\Gamma(B\setminus\mathscr{S}_{e},\bigwedge^{n}T_{B,\mathbb{Z}}). We fix a local representative μvnTv\mu_{v}\in\bigwedge^{n}T_{v} for every vertex vv and μσnΛσ\mu_{\sigma}\in\bigwedge^{n}\Lambda_{\sigma} for every maximal cell σ\sigma. Writing μv=m1mn\mu_{v}=m_{1}\wedge\cdots\wedge m_{n}, we have the corresponding relative volume form

μv=dlogzm1dlogzmn\mu_{v}=d\log z^{m_{1}}\wedge\cdots\wedge d\log z^{m_{n}}

in Ω𝐕(τ)v/n\Omega^{n}_{\mathbf{V}(\tau)_{v}^{\dagger}/\mathbb{C}^{\dagger}}. Now the relative log polyvector fields can be written as

lΘ𝐕(τ)v/=mτ1Pvzmn1nl.\bigwedge\nolimits^{-l}\Theta_{\mathbf{V}(\tau)_{v}^{\dagger}/\mathbb{C}^{\dagger}}=\bigoplus_{m\in\tau^{-1}P_{v}}z^{m}\partial_{n_{1}}\wedge\cdots\wedge\partial_{n_{l}}.

The volume form μv\mu_{v} defines a BV operator via contraction (Δα)μv:=(αμv)(\prescript{}{}{\Delta}\alpha)\mathbin{\lrcorner}\mu_{v}:=\partial(\alpha\mathbin{\lrcorner}\mu_{v}), which is given explicitly by

Δ(zmn1nl)=j=1l(1)j1m,njzmn1^njnl.\prescript{}{}{\Delta}(z^{m}\partial_{n_{1}}\wedge\cdots\wedge\partial_{n_{l}})=\sum_{j=1}^{l}(-1)^{j-1}\langle m,n_{j}\rangle z^{m}\partial_{n_{1}}\wedge\cdots\widehat{\partial}_{n_{j}}\cdots\wedge\partial_{n_{l}}.

A Schouten–-Nijenhuis–type bracket is given by extending the following formulae skew-symmetrically:

[zm1n1,zm2n2]\displaystyle[z^{m_{1}}\partial_{n_{1}},z^{m_{2}}\partial_{n_{2}}] =zm1+m2m¯1,n2n1m¯2,n1n2,\displaystyle=z^{m_{1}+m_{2}}\partial_{\langle\bar{m}_{1},n_{2}\rangle n_{1}-\langle\bar{m}_{2},n_{1}\rangle n_{2}},
[zm,n]\displaystyle[z^{m},\partial_{n}] =m¯,nzm.\displaystyle=\langle\bar{m},n\rangle z^{m}.

This gives Θ𝐕(τ)v/\bigwedge^{-*}\Theta_{\mathbf{V}(\tau)_{v}^{\dagger}/\mathbb{C}^{\dagger}} the structure of a BV algebra.

5.2.1. Construction of the semi-flat sheaves

For each kk\in\mathbb{N}, we shall define a sheaf 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} (resp. 𝖪sfk\prescript{k}{}{\mathsf{K}}^{*}_{\mathrm{sf}}) of kthk^{\text{th}}-order semi-flat log vector fields (resp. semi-flat log de Rham forms) over the open dense subset W0BW_{0}\subset B defined by

W0:=σ𝒫[n]intre(σ)ρ𝒫0[n1]intre(ρ)ρ𝒫1[n1](intre(ρ)(𝒮intre(ρ))),W_{0}:=\bigcup_{\sigma\in\mathscr{P}^{[n]}}\mathrm{int}_{\mathrm{re}}(\sigma)\cup\bigcup_{\rho\in\mathscr{P}^{[n-1]}_{0}}\mathrm{int}_{\mathrm{re}}(\rho)\cup\bigcup_{\rho\in\mathscr{P}^{[n-1]}_{1}}\big{(}\mathrm{int}_{\mathrm{re}}(\rho)\setminus(\mathscr{S}\cap\mathrm{int}_{\mathrm{re}}(\rho))\big{)},

where 𝒫0[n1]\mathscr{P}^{[n-1]}_{0} consists of ρ\rho’s such that intre(ρ)𝒮e=\mathrm{int}_{\mathrm{re}}(\rho)\cap\mathscr{S}_{e}=\emptyset and 𝒫1[n1]\mathscr{P}^{[n-1]}_{1} of ρ\rho’s that intersect with 𝒮e\mathscr{S}_{e}. These sheaves use the natural divisorial log structure on 𝐕(ρ)v\mathbf{V}(\rho)_{v}^{\dagger} and will not depend on the slab functions fvρf_{v\rho}’s. This construction is possible because we are using the much more flexible Euclidean topology on W0W_{0}, instead of the Zariski topology on X0\prescript{0}{}{X}.

For σ𝒫[n]\sigma\in\mathscr{P}^{[n]}, recall that we have V(σ)=Specan([σ1Σv])V(\sigma)=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\sigma^{-1}\Sigma_{v}]) for some vσ[0]v\in\sigma^{[0]}. We also have Specan([σ1Σv])=Λσ,/Λσ\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\sigma^{-1}\Sigma_{v}])=\Lambda^{*}_{\sigma,\mathbb{C}}/\Lambda^{*}_{\sigma}, which is isomorphic to ()n(\mathbb{C}^{*})^{n}, because σ1Σv=Λσ,=Tv,\sigma^{-1}\Sigma_{v}=\Lambda_{\sigma,\real}=T_{v,\real}. The local kthk^{\text{th}}-order thickening

𝕍k(σ):=Specan([σ1Pv/qk+1])()n×Specan([q]/qk+1)\prescript{k}{}{\mathbb{V}}(\sigma)^{\dagger}:=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\sigma^{-1}P_{v}/q^{k+1}])\cong(\mathbb{C}^{*})^{n}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[q]/q^{k+1})

is obtained by identifying σ1Pv\sigma^{-1}P_{v} as Λσ×\Lambda_{\sigma}\times\mathbb{N}. Choosing a different vertex vv^{\prime}, we can use the parallel transport TvTvT_{v}\cong T_{v^{\prime}} from vv to vv^{\prime} within intre(σ)\mathrm{int}_{\mathrm{re}}(\sigma) and the difference φv|σφv|σ\varphi_{v}|_{\sigma}-\varphi_{v^{\prime}}|_{\sigma} between two affine functions to identify the monoids σ1Pvσ1Pv\sigma^{-1}P_{v}\cong\sigma^{-1}P_{v^{\prime}}. We take

𝖦sfk|intre(σ):=ν(Θ𝕍k(σ)/Sk)ν(𝒪𝕍k(σ))Λσ,.\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}}|_{\mathrm{int}_{\mathrm{re}}(\sigma)}:=\nu_{*}\Big{(}\bigwedge\nolimits^{-*}\Theta_{\prescript{k}{}{\mathbb{V}}(\sigma)^{\dagger}/\prescript{k}{}{S}^{\dagger}}\Big{)}\cong\nu_{*}(\mathcal{O}_{\prescript{k}{}{\mathbb{V}}(\sigma)^{\dagger}})\otimes\bigwedge\nolimits^{-*}\Lambda_{\sigma,\real}^{*}.

Next, we need to glue the sheaves 𝖦sfk|intre(σ)\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}}|_{\mathrm{int}_{\mathrm{re}}(\sigma)}’s along neighborhoods of codimension one cells ρ\rho’s. For each codimension one cell ρ\rho, we fix a primitive normal dˇρ\check{d}_{\rho} to ρ\rho and label the two adjacent maximal cells σ+\sigma_{+} and σ\sigma_{-} so that dˇρ\check{d}_{\rho} is pointing into σ+\sigma_{+}. There are two situations to consider.

The simpler case is when 𝒮eintre(ρ)=\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\rho)=\emptyset, where the monodromy is trivial. In this case, we have V(ρ)=Specan([ρ1Σv])V(\rho)=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}\Sigma_{v}]), with the gluing V(σ±)V(ρ)V(\sigma_{\pm})\hookrightarrow V(\rho) as described below Definition 2.13 using the open gluing data sρσ±s_{\rho\sigma_{\pm}}. We take the kthk^{\text{th}}-order thickening given by

𝕍k(ρ):=Specan([ρ1Pv/qk+1]),\prescript{k}{}{\mathbb{V}}(\rho)^{\dagger}:=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}/q^{k+1}])^{\dagger},

equipped with the divisorial log structure induced by V(ρ)V(\rho). We extend the open gluing data

sρσ±:Λσ±s_{\rho\sigma_{\pm}}\colon\Lambda_{\sigma_{\pm}}\rightarrow\mathbb{C}^{*}

to

sρσ±:Λσ±s_{\rho\sigma_{\pm}}\colon\Lambda_{\sigma_{\pm}}\oplus\mathbb{Z}\rightarrow\mathbb{C}^{*}

so that sρσ±(0,1)=1s_{\rho\sigma_{\pm}}(0,1)=1, which acts as an automorphism of Specan([σ1Σv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\sigma^{-1}\Sigma_{v}]). In this way we can extend the gluing V(σ±)V(ρ)V(\sigma_{\pm})\hookrightarrow V(\rho) to

Specan([σ±1Pv/qk+1])Specan([ρ1Pv/qk+1])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\sigma_{\pm}^{-1}P_{v}/q^{k+1}])\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}/q^{k+1}])

by twisting with the ring homomorphism induced by zmsρσ±(m)1zmz^{m}\rightarrow s_{\rho\sigma_{\pm}}(m)^{-1}z^{m}. On a sufficiently small neighborhood 𝒲ρ\mathscr{W}_{\rho} of intre(ρ)\mathrm{int}_{\mathrm{re}}(\rho), we take

𝖦sfk|𝒲ρ:=ν(Θ𝕍k(ρ)/Sk)|𝒲ρ.\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}}|_{\mathscr{W}_{\rho}}:=\nu_{*}\Big{(}\bigwedge\nolimits^{-*}\Theta_{\prescript{k}{}{\mathbb{V}}(\rho)^{\dagger}/\prescript{k}{}{S}^{\dagger}}\Big{)}\Big{|}_{\mathscr{W}_{\rho}}.

Choosing a different vertex vv^{\prime}, we may use parallel transport to identify the fans ρ1Σvρ1Σv\rho^{-1}\Sigma_{v}\cong\rho^{-1}\Sigma_{v^{\prime}}, and further use the difference φv|𝒲ρφv|𝒲ρ\varphi_{v}|_{\mathscr{W}_{\rho}}-\varphi_{v^{\prime}}|_{\mathscr{W}_{\rho}} to identify the monoids ρ1Pvρ1Pv\rho^{-1}P_{v}\cong\rho^{-1}P_{v^{\prime}}. One can check that the sheaf 𝖦sfk|𝒲ρ\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}}|_{\mathscr{W}_{\rho}} is well-defined.

The more complicated case is when 𝒮eintre(ρ)\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\rho)\neq\emptyset, where the monodromy is non-trivial. We write intre(ρ)𝒮=vintre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)\setminus\mathscr{S}=\bigcup_{v}\mathrm{int}_{\mathrm{re}}(\rho)_{v}, where intre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)_{v} is the unique component which contains the vertex vv in its closure. We fix one vv, the corresponding intre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)_{v}, and a sufficiently small open subset 𝒲ρ,v\mathscr{W}_{\rho,v} of intre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)_{v}. We assume that the neighborhood 𝒲ρ,v\mathscr{W}_{\rho,v} of intre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)_{v} intersects neither 𝒲v,ρ\mathscr{W}_{v^{\prime},\rho^{\prime}} nor 𝒲ρ\mathscr{W}_{\rho^{\prime}} for any possible vv^{\prime} and ρ\rho^{\prime}. Then we consider the scheme-theoretic embedding

V(ρ)=Specan([ρ1Σv])Specan([ρ1Pv])V(\rho)=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}\Sigma_{v}])\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}])

given by

zm{zm¯if m lies on the boundary of the cone ρ1Pv,0if m lies in the interior of the cone ρ1Pv.z^{m}\mapsto\begin{cases}z^{\bar{m}}&\text{if $m$ lies on the boundary of the cone $\rho^{-1}P_{v}$,}\\ 0&\text{if $m$ lies in the interior of the cone $\rho^{-1}P_{v}$.}\end{cases}

We denote by 𝖵k(ρ)v\prescript{k}{}{\mathsf{V}}(\rho)_{v}^{\dagger} the kthk^{\text{th}}-order thickening of V(ρ)|ν1(𝒲ρ,v)V(\rho)|_{\nu^{-1}(\mathscr{W}_{\rho,v})} in Specan([ρ1Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}]) and equip it with the divisorial log structure which is log smooth over Sk\prescript{k}{}{S}^{\dagger} (note that it is different from the local model 𝕍k(ρ)\prescript{k}{}{\mathbb{V}}(\rho)^{\dagger} introduced earlier in §4 because the latter depends on the slab functions fv,ρf_{v,\rho}, as we can see explicitly in §5.2.2, while the former doesn’t). We take

𝖦sfk|𝒲ρ,v:=Θ𝖵k(ρ)v/Sk.\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}}|_{\mathscr{W}_{\rho,v}}:=\bigwedge\nolimits^{-*}\Theta_{\prescript{k}{}{\mathsf{V}}(\rho)_{v}^{\dagger}/\prescript{k}{}{S}^{\dagger}}.

The gluing with nearby maximal cells σ±\sigma_{\pm} on the overlap intre(σ±)𝒲ρ,v\mathrm{int}_{\mathrm{re}}(\sigma_{\pm})\cap\mathscr{W}_{\rho,v} is given by parallel transporting through the vertex vv to relate the monoids σ±1Pv\sigma_{\pm}^{-1}P_{v} and ρ1Pv\rho^{-1}P_{v} constructed from PvP_{v}, and twisting the map Specan([σ±1Pv])Specan([ρ1Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\sigma^{-1}_{\pm}P_{v}])\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}]) with the open gluing data

zmsρσ±1(m)zm,z^{m}\mapsto s_{\rho\sigma_{\pm}}^{-1}(m)z^{m},

using previous liftings of sρσ±s_{\rho\sigma_{\pm}} to Λσ±\Lambda_{\sigma_{\pm}}\oplus\mathbb{Z}. We obtain a commutative diagram of holomorphic maps

V(σ±)|𝒟𝕍k(σ±)|𝒟V(ρ)|𝒟𝖵k(ρ)|𝒟,\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 18.5904pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\crcr}}}\ignorespaces{\hbox{\kern-18.5904pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{V(\sigma_{\pm})|_{\mathscr{D}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 42.5904pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 0.0pt\raise-29.44443pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 42.5904pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{\prescript{k}{}{\mathbb{V}}(\sigma_{\pm})^{\dagger}|_{\mathscr{D}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 69.09264pt\raise-27.49998pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-16.14063pt\raise-37.94443pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{V(\rho)|_{\mathscr{D}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 45.04016pt\raise-37.94443pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 45.04016pt\raise-37.94443pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{\prescript{k}{}{\mathsf{V}}(\rho)^{\dagger}|_{\mathscr{D}}}$}}}}}}}\ignorespaces}}}}\ignorespaces,

where 𝒟=ν1(𝒲ρ,vintre(σ±))\mathscr{D}=\nu^{-1}(\mathscr{W}_{\rho,v}\cap\mathrm{int}_{\mathrm{re}}(\sigma_{\pm})) and the vertical arrow on the right hand side respects the log structures. The induced isomorphism

ν(Θ𝖵k(ρ)v/Sk)ν(Θ𝕍k(σ±)v/Sk)\nu_{*}\Big{(}\bigwedge\nolimits^{-*}\Theta_{\prescript{k}{}{\mathsf{V}}(\rho)_{v}^{\dagger}/\prescript{k}{}{S}^{\dagger}}\Big{)}\cong\nu_{*}\Big{(}\bigwedge\nolimits^{-*}\Theta_{\prescript{k}{}{\mathbb{V}}(\sigma_{\pm})_{v}^{\dagger}/\prescript{k}{}{S}^{\dagger}}\Big{)}

of sheaves on the overlap 𝒲ρ,vintre(σ±)\mathscr{W}_{\rho,v}\cap\mathrm{int}_{\mathrm{re}}(\sigma_{\pm}) then gives the desired gluing for defining the sheaf 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} on W0W_{0}. Note that the cocycle condition is trivial here as there is no triple intersection of any three open subsets from intre(σ)\mathrm{int}_{\mathrm{re}}(\sigma), 𝒲ρ\mathscr{W}_{\rho} and 𝒲ρ,v\mathscr{W}_{\rho,v}.

Similarly, we can define the sheaf 𝖪sfk\prescript{k}{}{\mathsf{K}}^{*}_{\mathrm{sf}} of semi-flat log de Rham forms, together with a relative volume form ω0k𝖪sfnk(W0)\prescript{k}{}{\omega}_{0}\in\prescript{k}{\parallel}{\mathsf{K}}^{n}_{\mathrm{sf}}(W_{0}) obtained from gluing the local μv\mu_{v}’s specified by the element μ\mu as described in the beginning of §5.2.

It would be useful to write down elements of the sheaf 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} more explicitly. For instance, fixing a point xintre(ρ)vx\in\mathrm{int}_{\mathrm{re}}(\rho)_{v}, we may write

(5.1) 𝖦sf,xk=ν(𝒪𝖵k(ρ)v)xTv,,\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf},x}=\nu_{*}(\mathcal{O}_{\prescript{k}{}{\mathsf{V}}(\rho)_{v}})_{x}\otimes\bigwedge\nolimits^{-*}T^{*}_{v,\real},

and use n\partial_{n} to stand for the semi-flat holomorphic vector field associated to an element nTv,n\in T^{*}_{v,\real}.

Note that analytic continuation around the singular locus 𝒮eintre(ρ)\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\rho) acts non-trivially on the semi-flat sheaf 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} due to the presence of non-trivial monodromy of the affine structure. Below is a simple example.

Example 5.7.

We consider the local affine charts which appeared in Example 2.3, equipped with a strictly convex piecewise linear affine function φ\varphi on Σρ\Sigma_{\rho} whose change of slopes is 11. Let us study the analytic continuation of a local section along the loop γ\gamma which starts at a point b+b_{+}, as shown in Figure 6.

Refer to caption
Figure 6. Analytic continuation along γ\gamma

First, we can identify both ρ1Pv+\rho^{-1}P_{v_{+}} and ρ1Pv\rho^{-1}P_{v_{-}} with the monoid in the cone P={(x,y,z)|zφ(x)}P=\{(x,y,z)\ |\ z\geq\varphi(x)\} via parallel transport through σ+\sigma_{+}. Writing u=z(1,0,1)u=z^{(1,0,1)}, v=z(1,0,0)v=z^{(-1,0,0)}, w=z(0,1,0)w=z^{(0,-1,0)} and q=z(0,0,1)q=z^{(0,0,1)}, we have [P][u,v,w±,q]/(uvq)\mathbb{C}[P]\cong\mathbb{C}[u,v,w^{\pm},q]/(uv-q) . Now the analytic continuation of uν(𝒪𝖵k(ρ)v+)b+u\in\nu_{*}(\mathcal{O}_{\prescript{k}{}{\mathsf{V}}(\rho)_{v_{+}}})_{b_{+}} along γ\gamma (going from the chart UIIU_{\mathrm{II}} to the chart UIU_{\mathrm{I}} and then back to UIIU_{\mathrm{II}}) is given by as a sequence of elements:

u\textstyle{u\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sρσ+((1,0))1u\textstyle{s_{\rho\sigma_{+}}((1,0))^{-1}u\ignorespaces\ignorespaces\ignorespaces\ignorespaces}uw\textstyle{uw\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sρσ((1,0))1qv1w\textstyle{s_{\rho\sigma_{-}}((1,0))^{-1}qv^{-1}w\ignorespaces\ignorespaces\ignorespaces\ignorespaces}wu,\textstyle{wu,}

via the following sequence of maps between the stalks over b+,c+UIIb_{+},c_{+}\in U_{\mathrm{II}} and b,cUIb_{-},c_{-}\in U_{\mathrm{I}}:

ν(𝒪𝖵k(ρ)v+)b+ν(𝒪𝕍k(σ+))c+ν(𝒪𝖵k(ρ)v)bν(𝒪𝕍k(σ))cν(𝒪𝖵k(ρ)v+)b+.\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 27.17693pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&&\crcr}}}\ignorespaces{\hbox{\kern-27.17693pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{\nu_{*}(\mathcal{O}_{\prescript{k}{}{\mathsf{V}}(\rho)_{v_{+}}})_{b_{+}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 51.17693pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 51.17693pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{\nu_{*}(\mathcal{O}_{\prescript{k}{}{\mathbb{V}}(\sigma_{+})^{\dagger}})_{c_{+}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 130.10957pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 130.10957pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{\nu_{*}(\mathcal{O}_{\prescript{k}{}{\mathsf{V}}(\rho)_{v_{-}}})_{b_{-}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 205.9034pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 205.9034pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{\nu_{*}(\mathcal{O}_{\prescript{k}{}{\mathbb{V}}(\sigma_{-})^{\dagger}})_{c_{-}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 281.9916pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 281.9916pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{\nu_{*}(\mathcal{O}_{\prescript{k}{}{\mathsf{V}}(\rho)_{v_{+}}})_{b_{+}}}$}}}}}}}\ignorespaces}}}}\ignorespaces.

So we see that the analytic continuation along γ\gamma maps uu to wuwu.

𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} is equipped with the BV algebra structure inherited from Specan([ρ1Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}])^{\dagger} (as described in the beginning of §5.2), which agrees with the one induced from the volume form ω0k\prescript{k}{}{\omega}_{0}. This allows us to define the sheaf of semi-flat tropical vertex Lie algebras as

(5.2) 𝔥k:=Ker(Δ)|𝖦sf1k[1].\prescript{k}{}{\mathfrak{h}}:=\mathrm{Ker}(\prescript{}{}{\Delta})|_{\prescript{k}{}{\mathsf{G}}^{-1}_{\mathrm{sf}}}[-1].
Remark 5.8.

The sheaf can actually be extended over the non-essential singular locus 𝒮𝒮e\mathscr{S}\setminus\mathscr{S}_{e} because the monodromy around that locus acts trivially, but this is not necessary for our later discussion.

5.2.2. Explicit gluing away from codimension 22

When we define the sheaves 𝒢αk\prescript{k}{}{\mathcal{G}}^{*}_{\alpha}’s in §4.1, the open subset WαW_{\alpha} is taken to be a sufficiently small neighborhood of xintre(τ)x\in\mathrm{int}_{\mathrm{re}}(\tau) for some τ𝒫\tau\in\mathscr{P}. In fact, we can choose one of these open subsets to be the large open dense subset W0W_{0}. In this subsection, we construct the sheaves 𝒢0k\prescript{k}{}{\mathcal{G}}^{*}_{0} and 𝒦0k\prescript{k}{}{\mathcal{K}}^{*}_{0} on W0W_{0} using an explicit gluing of the underlying complex analytic space.

Over intre(σ)\mathrm{int}_{\mathrm{re}}(\sigma) for σ𝒫[n]\sigma\in\mathscr{P}^{[n]} or over 𝒲ρ\mathscr{W}_{\rho} for ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]} with 𝒮eintre(ρ)=\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\rho)=\emptyset, we have 𝒢0k=𝖦sfk\prescript{k}{}{\mathcal{G}}^{*}_{0}=\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}}, which was just constructed in §5.2.1. So it remains to consider ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]} such that 𝒮eintre(ρ)\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\rho)\neq\emptyset. The log structure of V(ρ)V(\rho)^{\dagger} is prescribed by the slab functions fv,ρΓ(𝒪Vρ(v))f_{v,\rho}\in\Gamma(\mathcal{O}_{V_{\rho}(v)})’s, which restrict to functions sv,ρ1(fv,ρ)s^{-1}_{v,\rho}(f_{v,\rho})’s on the torus Specan([Λρ])()n1\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Lambda_{\rho}])\cong(\mathbb{C}^{*})^{n-1}. Each of these can be pulled back via the natural projection Specan([ρ1Σv])Specan([Λρ])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}\Sigma_{v}])\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Lambda_{\rho}]) to give a function on Specan([ρ1Σv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}\Sigma_{v}]). In this case, we may fix the log chart V(ρ)|ν1(𝒲ρ,v)Specan([ρ1Pv])V(\rho)^{\dagger}|_{\nu^{-1}(\mathscr{W}_{\rho,v})}\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}])^{\dagger} given by the equation

zm{zm¯if dˇρ,m¯0 ,zm¯(svρ1(fv,ρ))dˇρ,m¯if dˇρ,m¯0 .z^{m}\mapsto\begin{dcases*}z^{\bar{m}}&\text{if $\langle\check{d}_{\rho},\bar{m}\rangle\geq 0$ },\\ z^{\bar{m}}\big{(}s^{-1}_{v\rho}(f_{v,\rho})\big{)}^{\langle\check{d}_{\rho},\bar{m}\rangle}&\text{if $\langle\check{d}_{\rho},\bar{m}\rangle\leq 0$ }.\end{dcases*}

Write 𝕍k(ρ)v\prescript{k}{}{\mathbb{V}}(\rho)_{v}^{\dagger} for the corresponding kthk^{\text{th}}-order thickening in Specan([ρ1Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}]), which gives a local model for smoothing V(ρ)|ν1(𝒲ρ,v)V(\rho)|_{\nu^{-1}(\mathscr{W}_{\rho,v})} (as in §4). We take

𝒢0k|𝒲ρ,v:=ν(Θ𝕍k(ρ)v/Sk).\prescript{k}{}{\mathcal{G}}_{0}^{*}|_{\mathscr{W}_{\rho,v}}:=\nu_{*}\Big{(}\bigwedge\nolimits^{-*}\Theta_{\prescript{k}{}{\mathbb{V}}(\rho)_{v}^{\dagger}/\prescript{k}{}{S}^{\dagger}}\Big{)}.

We have to specify the gluing on the overlap 𝒲ρ,vintre(σ±)\mathscr{W}_{\rho,v}\cap\mathrm{int}_{\mathrm{re}}(\sigma_{\pm}) with the adjacent maximal cells σ±\sigma_{\pm}. This is given by first using parallel transport through vv to relate the monoids σ±1Pv\sigma_{\pm}^{-1}P_{v} and ρ1Pv\rho^{-1}P_{v} as in the semi-flat case, and then an embedding Specan([σ±1Pv/qk+1])Specan([ρ1Pv/qk+1])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\sigma_{\pm}^{-1}P_{v}/q^{k+1}])\rightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}/q^{k+1}]) via the formula

(5.3) zm{sρσ+1(m)zmfor σ+ ,sρσ1(m)zm(svσ1(fv,ρ))dˇρ,m¯for σ ,z^{m}\mapsto\begin{dcases*}s_{\rho\sigma_{+}}^{-1}(m)z^{m}&\text{for $\sigma_{+}$ },\\ s_{\rho\sigma_{-}}^{-1}(m)z^{m}\big{(}s^{-1}_{v\sigma_{-}}(f_{v,\rho})\big{)}^{\langle\check{d}_{\rho},\bar{m}\rangle}&\text{for $\sigma_{-}$ },\end{dcases*}

where svσ±s_{v\sigma_{\pm}}, sρσ±s_{\rho\sigma_{\pm}} are treated as maps Λσ±\Lambda_{\sigma_{\pm}}\oplus\mathbb{Z}\rightarrow\mathbb{C}^{*} as before. We observe that there is a commutative diagram of log morphisms

V(σ±)|𝒟𝕍k(σ±)|𝒟V(ρ)|𝒟𝕍k(ρ)|𝒟,\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 18.72375pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\crcr}}}\ignorespaces{\hbox{\kern-18.72375pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{V(\sigma_{\pm})^{\dagger}|_{\mathscr{D}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 42.72375pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 0.0pt\raise-27.49998pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 42.72375pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{\prescript{k}{}{\mathbb{V}}(\sigma_{\pm})^{\dagger}|_{\mathscr{D}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 69.226pt\raise-27.49998pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-16.27399pt\raise-37.94443pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{V(\rho)^{\dagger}|_{\mathscr{D}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 45.17352pt\raise-37.94443pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 45.17352pt\raise-37.94443pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{\prescript{k}{}{\mathbb{V}}(\rho)^{\dagger}|_{\mathscr{D}}}$}}}}}}}\ignorespaces}}}}\ignorespaces,

where 𝒟=ν1(𝒲ρ,vintre(σ±))\mathscr{D}=\nu^{-1}(\mathscr{W}_{\rho,v}\cap\mathrm{int}_{\mathrm{re}}(\sigma_{\pm})). The induced isomorphism

ν(Θ𝕍k(ρ)v/Sk)ν(Θ𝕍k(σ±)v/Sk)\nu_{*}\Big{(}\bigwedge\nolimits^{-*}\Theta_{\prescript{k}{}{\mathbb{V}}(\rho)_{v}^{\dagger}/\prescript{k}{}{S}^{\dagger}}\Big{)}\cong\nu_{*}\Big{(}\bigwedge\nolimits^{-*}\Theta_{\prescript{k}{}{\mathbb{V}}(\sigma_{\pm})_{v}^{\dagger}/\prescript{k}{}{S}^{\dagger}}\Big{)}

of sheaves on the overlap 𝒟\mathscr{D} then provides the gluing for defining the sheaf 𝒢0k\prescript{k}{}{\mathcal{G}}^{*}_{0} on W0W_{0}. Hence, we obtain a sheaf 𝒢0k\prescript{k}{}{\mathcal{G}}^{*}_{0} of BV algebras, where the BV structure is inherited from the local models Specan([σ1Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\sigma^{-1}P_{v}]) and Specan([ρ1Pv])\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}]). Similarly, we can define the sheaf 𝒦0k\prescript{k}{}{\mathcal{K}}^{*}_{0} of log de Rham forms over W0W_{0}, together with a relative volume form ω0k𝒦0nk(W0)\prescript{k}{}{\omega}_{0}\in\prescript{k}{\parallel}{\mathcal{K}}^{n}_{0}(W_{0}) by gluing the local μv\mu_{v}’s.

5.2.3. Relation between the semi-flat dgBV algebra and the log structure

The difference between 𝒢0k\prescript{k}{}{\mathcal{G}}^{*}_{0} and 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} is that analytic continuation along a path γ\gamma in intre(σ±)intre(ρ)\mathrm{int}_{\mathrm{re}}(\sigma_{\pm})\cup\mathrm{int}_{\mathrm{re}}(\rho), where ρ=σ+σ\rho=\sigma_{+}\cap\sigma_{-}, induces a non-trivial action on 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} (the semi-flat sheaf) but not on 𝒢0k\prescript{k}{}{\mathcal{G}}^{*}_{0} (the corrected sheaf). This is because, near a singular point pΓp\in\Gamma of the affine structure on BB, there is another local model 𝒢αk\prescript{k}{}{\mathcal{G}}^{*}_{\alpha} for pWαp\in W_{\alpha} constructed in 4.1, where restrictions of sections are invariant under analytic continuation (cf. Example 5.7). This is in line with the philosophy that monodromy is being cancelled by the slab functions fv,ρf_{v,\rho}’s (which we also call initial wall-crossing factors). In view of this, we should be able to relate the sheaves 𝒢0k\prescript{k}{}{\mathcal{G}}^{*}_{0} and 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} by adding back the initial wall-crossing factors fv,ρf_{v,\rho}’s.

Recall that the slab function fv,ρf_{v,\rho} is a function on Vρ(v)Xρ0V_{\rho}(v)\subset\prescript{0}{}{X}_{\rho}, whose zero locus is Z1ρVρ(v)Z^{\rho}_{1}\cap V_{\rho}(v) for ρ\rho such that 𝒮eintre(ρ)\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\rho)\neq\emptyset. Also recall that, for ρ\rho containing vv, ρv\rho_{v} is the unique contractible component in ρ𝒞1(B𝒮)\rho\cap\mathscr{C}^{-1}(B\setminus\mathscr{S}) such that vρvv\in\rho_{v}, as defined in Assumption 3.5. Note that the inverse image μ1(ρv)Vρ(v)\mu^{-1}(\rho_{v})\subset V_{\rho}(v) under the generalized moment map μ\mu is also a contractible open subset. It contains the 0-dimensional stratum xvx_{v} in Vρ(v)V_{\rho}(v) that corresponds to vv. Since fv,ρ(xv)=1f_{v,\rho}(x_{v})=1, we can define log(fv,ρ)\log(f_{v,\rho}) in a small neighborhood of xvx_{v}, and it can further be extended to the whole of μ1(ρv)Vρ(v)\mu^{-1}(\rho_{v})\subset V_{\rho}(v) because this subset is contractible. Restricting to the open dense torus orbit Specan([Λρ])μ1(ρv)\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Lambda_{\rho}])\cap\mu^{-1}(\rho_{v}), we obtain log(svρ1(fv,ρ))\log(s_{v\rho}^{-1}(f_{v,\rho})), which can in addition be lifted to a section in 𝖦sf0k(𝒲ρ,v)=Γ(𝒲ρ,v,𝒪𝖵k(ρ)v)\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf}}(\mathscr{W}_{\rho,v})=\Gamma(\mathscr{W}_{\rho,v},\mathcal{O}_{\prescript{k}{}{\mathsf{V}}(\rho)_{v}}) for a sufficiently small 𝒲ρ,v\mathscr{W}_{\rho,v}.

Now we resolve the sheaves 𝒢0k\prescript{k}{}{\mathcal{G}}^{*}_{0} and 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} by the complex 𝒯\mathscr{T}^{*} introduced in §5.1. We let

𝖯𝖵sf,k:=𝒯|W0𝖦sfk\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}:=\mathscr{T}^{*}|_{W_{0}}\otimes\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}}

and equip it with ¯=d1\bar{\partial}_{\circ}=d\otimes 1, Δ\prescript{}{}{\Delta} and \wedge, making it a sheaf of dgBV algebras. Over the open subset 𝒲ρ,v\mathscr{W}_{\rho,v}, using the explicit description of 𝖦sfk|𝒲ρ,v\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}}|_{\mathscr{W}_{\rho,v}}, we consider the element

(5.4) ϕv,ρ:=δv,ρlog(svρ1(fv,ρ))dˇρ𝖯𝖵sf1,1k(𝒲ρ,v),\phi_{v,\rho}:=-\delta_{v,\rho}\otimes\log(s_{v\rho}^{-1}(f_{v,\rho}))\partial_{\check{d}_{\rho}}\in\prescript{k}{}{\mathsf{PV}}^{-1,1}_{\mathrm{sf}}(\mathscr{W}_{\rho,v}),

where δv,ρ\delta_{v,\rho} is any 11-form with asymptotic support in intre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)_{v} and whose integral over any curve transversal to intre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)_{v} going from σ\sigma_{-} to σ+\sigma_{+} is asymptotically 11; such a 1-form can be constructed using a family of bump functions in the normal direction of intre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)_{v} similar to Example 5.3 (see also [7, §4]). We can further extend the section ϕv,ρ\phi_{v,\rho} to the whole W0W_{0} by setting it to be 0 outside a small neighborhood of intre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)_{v} in 𝒲ρ,v\mathscr{W}_{\rho,v}.

Definition 5.9.

The sheaf of semi-flat polyvector fields is defined as

𝖯𝖵sf,k:=𝒯|W0𝖦sfk,\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}:=\mathscr{T}^{*}|_{W_{0}}\otimes\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}},

which is equipped with a BV operator Δ\prescript{}{}{\Delta}, a wedge product \wedge (and hence a Lie bracket [,][\cdot,\cdot]) and the operator

¯sf:=¯+[ϕin,]=¯+v,ρ[ϕv,ρ,],\bar{\partial}_{\mathrm{sf}}:=\bar{\partial}_{\circ}+[\phi_{\mathrm{in}},\cdot]=\bar{\partial}_{\circ}+\sum_{v,\rho}[\phi_{v,\rho},\cdot],

where ¯=d1\bar{\partial}_{\circ}=d\otimes 1 and ϕin:=v,ρϕv,ρ\phi_{\mathrm{in}}:=\sum_{v,\rho}\phi_{v,\rho}. We also define the sheaf of semi-flat log de Rham forms as

𝖠sf,k:=𝒯|W0𝖪sfk,\prescript{k}{}{\mathsf{A}}^{*,*}_{\mathrm{sf}}:=\mathscr{T}^{*}|_{W_{0}}\otimes\prescript{k}{}{\mathsf{K}}^{*}_{\mathrm{sf}},

equipped with \prescript{}{}{\partial}, \wedge,

¯sf:=¯+v,ρϕv,ρ,\bar{\partial}_{\mathrm{sf}}:=\bar{\partial}_{\circ}+\sum_{v,\rho}\mathcal{L}_{\phi_{v,\rho}},

and a contraction action \mathbin{\lrcorner} by elements in 𝖯𝖵sf,k\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}.

It can be easily checked that ¯sf2=[¯sf,Δ]=0\bar{\partial}_{\mathrm{sf}}^{2}=[\bar{\partial}_{\mathrm{sf}},\prescript{}{}{\Delta}]=0, so we have a sheaf of dgBV algebras.

On the other hand, we write

PkV0,:=𝒯|W0𝒢0k,\prescript{k}{}{PV}^{*,*}_{0}:=\mathscr{T}^{*}|_{W_{0}}\otimes\prescript{k}{}{\mathcal{G}}^{*}_{0},

which is equipped with the operators ¯0=d1\bar{\partial}_{0}=d\otimes 1, Δ\prescript{}{}{\Delta} and \wedge. The following important lemma is a comparison between the two sheaves of dgBV algebras.

Lemma 5.10.

There exists a set of compatible isomorphisms

Φ:PkV0,𝖯𝖵sf,k,k\varPhi\colon\prescript{k}{}{PV}^{*,*}_{0}\rightarrow\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}},\ k\in\mathbb{N}

of sheaves of dgBV algebras such that Φ¯0=¯sfΦ\varPhi\circ\bar{\partial}_{0}=\bar{\partial}_{\mathrm{sf}}\circ\varPhi for each kk\in\mathbb{N}.

There also exists a set of compatible isomorphisms

Φ:𝒜0,k𝖠sf,k,k\varPhi\colon\prescript{k}{}{\mathcal{A}}^{*,*}_{0}\rightarrow\prescript{k}{}{\mathsf{A}}^{*,*}_{\mathrm{sf}},\ k\in\mathbb{N}

of sheaves of dgas preserving the contraction action \mathbin{\lrcorner} and such that Φ¯0=¯sfΦ\varPhi\circ\bar{\partial}_{0}=\bar{\partial}_{\mathrm{sf}}\circ\varPhi for each kk\in\mathbb{N}. Furthermore, the relative volume form ω0k\prescript{k}{}{\omega}_{0} is identified via Φ\varPhi.

Proof.

Outside those intre(ρ)\mathrm{int}_{\mathrm{re}}(\rho)’s such that 𝒮eintre(ρ)\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\rho)\neq\emptyset, the two sheaves are identical. So we will take a component intre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)_{v} of intre(ρ)𝒮\mathrm{int}_{\mathrm{re}}(\rho)\setminus\mathscr{S} and compare the sheaves on a neighborhood 𝒲ρ,v\mathscr{W}_{\rho,v}.

We fix a point xintre(ρ)vx\in\mathrm{int}_{\mathrm{re}}(\rho)_{v} and describe the map Φ\Phi at the stalks of the two sheaves. First, the preimage K:=ν1(x)Λρ,/ΛρK:=\nu^{-1}(x)\cong\Lambda_{\rho,\real}^{*}/\Lambda_{\rho}^{*} can be identified as a real (n1)(n-1)-dimensional torus in Specan([Λρ])()n1\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Lambda_{\rho}])\cong(\mathbb{C}^{*})^{n-1}. We have an identification ρ1ΣvΣρ×Λρ\rho^{-1}\Sigma_{v}\cong\Sigma_{\rho}\times\Lambda_{\rho}, and we choose the unique primitive element mρΣρm_{\rho}\in\Sigma_{\rho} in the ray pointing into σ+\sigma_{+}. As analytic spaces, we write

Specan([Σρ])={uv=0}2,\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\rho}])=\{uv=0\}\subset\mathbb{C}^{2},

where u=zmρu=z^{m_{\rho}} and v=zmρv=z^{-m_{\rho}}, and

Specan([ρ1Σv])=()n1×{uv=0}.\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}\Sigma_{v}])=(\mathbb{C}^{*})^{n-1}\times\{uv=0\}.

The germ 𝒪V(ρ),K\mathcal{O}_{V(\rho),K} of analytic functions can be written as

𝒪V(ρ),K={a0+i=1aiui+i=1aivi|ai𝒪()n1(U)for neigh. UK,supi0log|ai||i|<}.\mathcal{O}_{V(\rho),K}=\left\{a_{0}+\sum_{i=1}^{\infty}a_{i}u^{i}+\sum_{i=-1}^{-\infty}a_{i}v^{-i}\ \Big{|}\ a_{i}\in\mathcal{O}_{(\mathbb{C}^{*})^{n-1}}(U)\ \text{for neigh. $U\supset K$},\ \sup_{i\neq 0}\frac{\log|a_{i}|}{|i|}<\infty\right\}.

Using the embedding V(ρ)|ν1(𝒲ρ,v)𝕍k(ρ)vV(\rho)|_{\nu^{-1}(\mathscr{W}_{\rho,v})}\rightarrow\prescript{k}{}{\mathbb{V}}(\rho)_{v}^{\dagger} in §5.2.2, we can write

𝒢0,x0k=𝒪𝕍k(ρ)v,K=\displaystyle\prescript{k}{}{\mathcal{G}}^{0}_{0,x}=\mathcal{O}_{\prescript{k}{}{\mathbb{V}}(\rho)_{v},K}=
{j=0k(a0,j+i=1ai,jui+i=1ai,jvi)qj|ai,j𝒪()n1(U)for neigh. UK,supi0log|ai,j||i|<},\displaystyle\left\{\sum_{j=0}^{k}(a_{0,j}+\sum_{i=1}^{\infty}a_{i,j}u^{i}+\sum_{i=-1}^{-\infty}a_{i,j}v^{-i})q^{j}\Big{|}\ a_{i,j}\in\mathcal{O}_{(\mathbb{C}^{*})^{n-1}}(U)\ \text{for neigh. $U\supset K$},\ \sup_{i\neq 0}\frac{\log|a_{i,j}|}{|i|}<\infty\right\},

with the relation uv=qlsvρ1(fv,ρ)uv=q^{l}s_{v\rho}^{-1}(f_{v,\rho}) (here ll is the change of slopes for φv\varphi_{v} across ρ\rho). For the elements (mρ,φv(mρ))(m_{\rho},\varphi_{v}(m_{\rho})) and (mρ,φv(mρ))(-m_{\rho},\varphi_{v}(-m_{\rho})) in ρ1Pv\rho^{-1}P_{v}, we have the identities (we omit the dependence on kk when we write elements in the stalks of sheaves):

z(mρ,φv(mρ))\displaystyle z^{(m_{\rho},\varphi_{v}(m_{\rho}))} =u,\displaystyle=u,
z(mρ,φv(mρ))\displaystyle z^{-(-m_{\rho},\varphi_{v}(-m_{\rho}))} =svρ1(fv,ρ)1v,\displaystyle=s_{v\rho}^{-1}(f_{v,\rho})^{-1}v,

describing the embedding 𝕍k(ρ)vSpecan([ρ1Pv])\prescript{k}{}{\mathbb{V}}(\rho)_{v}^{\dagger}\hookrightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}])^{\dagger}. For polyvector fields, we can write

𝒢0,xk=𝒢0,x0kTv,.\prescript{k}{}{\mathcal{G}}^{*}_{0,x}=\prescript{k}{}{\mathcal{G}}^{0}_{0,x}\otimes\bigwedge^{-*}T_{v,\real}^{*}.

The BV operator is described by the relations Δ(n)=0\prescript{}{}{\Delta}(\partial_{n})=0, [n1,n2]=0[\partial_{n_{1}},\partial_{n_{2}}]=0, and

(5.5) {[zm,n]=Δ(zmn)=m,nzmfor m with m¯Λρ,nTv,;[u,n]=Δ(un)=mρ,nufor nTv,;[v,n]=Δ(vn)=mρ,nv+n(logsvρ1(fv,ρ))vfor nTv,.\begin{cases}[z^{m},\partial_{n}]=\prescript{}{}{\Delta}(z^{m}\partial_{n})=\langle m,n\rangle z^{m}&\text{for $m$ with $\bar{m}\in\Lambda_{\rho},\ n\in T^{*}_{v,\real}$;}\\ [u,\partial_{n}]=\prescript{}{}{\Delta}(u\partial_{n})=\langle m_{\rho},n\rangle u&\text{for $n\in T^{*}_{v,\real}$};\\ [v,\partial_{n}]=\prescript{}{}{\Delta}(v\partial_{n})=\langle-m_{\rho},n\rangle v+\partial_{n}(\log s_{v\rho}^{-1}(f_{v,\rho}))v&\text{for $n\in T^{*}_{v,\real}$}.\end{cases}

Similarly, we can write down the stalk for 𝖦sf,xk=𝖦sf,xkTv,\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf},x}=\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf},x}\otimes\bigwedge^{-*}T_{v,\real}^{*}. As a module over 𝒪()n1,K[q]/(qk+1)\mathcal{O}_{(\mathbb{C}^{*})^{n-1},K}\otimes_{\mathbb{C}}\mathbb{C}[q]/(q^{k+1}), we have 𝖦sf,xk=𝒢0,xk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf},x}=\prescript{k}{}{\mathcal{G}}^{*}_{0,x}; the ring structure on 𝖦sf,x0k\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf},x} differs from that on 𝒢0,x0k\prescript{k}{}{\mathcal{G}}^{0}_{0,x} and is determined by the relation uv=qluv=q^{l}. The embedding 𝖵k(ρ)vSpecan([ρ1Pv])\prescript{k}{}{\mathsf{V}}(\rho)_{v}^{\dagger}\hookrightarrow\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\rho^{-1}P_{v}])^{\dagger} is given by

z(mρ,φv(mρ))\displaystyle z^{(m_{\rho},\varphi_{v}(m_{\rho}))} =u,\displaystyle=u,
z(mρ,φv(mρ))\displaystyle z^{-(-m_{\rho},\varphi_{v}(-m_{\rho}))} =v.\displaystyle=v.

The formulae for the BV operator are the same as that for 𝒢0,xk\prescript{k}{}{\mathcal{G}}^{*}_{0,x}, except that for the last equation in (5.5), we have [v,n]=Δ(vn)=mρ,nv[v,\partial_{n}]=\prescript{}{}{\Delta}(v\partial_{n})=\langle-m_{\rho},n\rangle v instead.

We apply the argument in [7, §4], where we considered a scattering diagram consisting of only one wall, to relate these two sheaves. We can find a set of compatible elements θ=(θk)k\theta=(\prescript{k}{}{\theta})_{k\in\mathbb{N}}, where θk𝖯𝖵sf1,0k(𝒲ρ,v)\prescript{k}{}{\theta}\in\prescript{k}{}{\mathsf{PV}}^{-1,0}_{\mathrm{sf}}(\mathscr{W}_{\rho,v}) for kk\in\mathbb{N}, such that eθ¯=¯sfe^{\theta}*\bar{\partial}_{\circ}=\bar{\partial}_{\mathrm{sf}} and Δ(θ)=0\prescript{}{}{\Delta}(\theta)=0. Explicitly, θ\theta is a step-function-like section of the form

θ={log(svρ1(fv,ρ))dˇρon intre(σ+)𝒲ρ,v,0on intre(σ)𝒲ρ,v.\theta=\begin{dcases}\log(s_{v\rho}^{-1}(f_{v,\rho}))\partial_{\check{d}_{\rho}}&\text{on $\mathrm{int}_{\mathrm{re}}(\sigma_{+})\cap\mathscr{W}_{\rho,v}$,}\\ 0&\text{on $\mathrm{int}_{\mathrm{re}}(\sigma_{-})\cap\mathscr{W}_{\rho,v}$.}\end{dcases}

For each kk\in\mathbb{N}, we also define θ0:=log(svρ1(fv,ρ))dˇρ\theta_{0}:=\log(s_{v\rho}^{-1}(f_{v,\rho}))\partial_{\check{d}_{\rho}}, as an element in 𝖦sf1k(𝒲ρ,v)\prescript{k}{}{\mathsf{G}}^{-1}_{\mathrm{sf}}(\mathscr{W}_{\rho,v}). Now we define the map Φx:PkV0,x,𝖯𝖵sf,x,k\varPhi_{x}\colon\prescript{k}{}{PV}^{*,*}_{0,x}\rightarrow\prescript{k}{}{\mathsf{PV}}_{\mathrm{sf},x}^{*,*} at the stalks by writing

PkV0,x,=𝒯x𝒢0,x0kTv,,\prescript{k}{}{PV}^{*,*}_{0,x}=\mathscr{T}_{x}^{*}\otimes\prescript{k}{}{\mathcal{G}}^{0}_{0,x}\otimes\bigwedge^{-*}T_{v,\real}^{*},

(and similarly for 𝖯𝖵sf,x,k\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf},x}), and extending the formulae

{Φx(α)=αfor α𝒯x,Φx(f)=e[θ,]f=ffor f𝒪()n1,K,Φx(u)=e[θθ0,]u,Φx(v)=e[θ,]v,Φx(n)=e[θθ0,]nfor nTv,\begin{cases}\varPhi_{x}(\alpha)=\alpha&\text{for $\alpha\in\mathscr{T}_{x}$},\\ \varPhi_{x}(f)=e^{[\theta,\cdot]}f=f&\text{for $f\in\mathcal{O}_{(\mathbb{C}^{*})^{n-1},K}$},\\ \varPhi_{x}(u)=e^{[\theta-\theta_{0},\cdot]}u,&\\ \varPhi_{x}(v)=e^{[\theta,\cdot]}v,&\\ \varPhi_{x}(\partial_{n})=e^{[\theta-\theta_{0},\cdot]}\partial_{n}&\text{for $n\in T_{v,\real}^{*}$}\end{cases}

through the tensor product \otimes and skew-symmetrically in n\partial_{n}’s.

To see that Φ\varPhi is the desired isomorphism, we check all the relations by computations:

  • Since e[θ,]¯e[θ,]=¯sfe^{[\theta,\cdot]}\circ\bar{\partial}_{\circ}\circ e^{-[\theta,\cdot]}=\bar{\partial}_{\mathrm{sf}}, we have

    ¯sfΦx(u)=e[θ,]¯(e[θ0,]u)=0;\bar{\partial}_{\mathrm{sf}}\varPhi_{x}(u)=e^{[\theta,\cdot]}\bar{\partial}_{\circ}(e^{-[\theta_{0},\cdot]}u)=0;

    similarly, we have ¯sf(Φx(v))=0=¯sf(Φx(n))\bar{\partial}_{\mathrm{sf}}(\varPhi_{x}(v))=0=\bar{\partial}_{\mathrm{sf}}(\varPhi_{x}(\partial_{n})). Hence, we have Φx¯0=¯sfΦx\varPhi_{x}\circ\bar{\partial}_{0}=\bar{\partial}_{\mathrm{sf}}\circ\varPhi_{x}.

  • We have e[θ0,]u=svρ1(fv,ρ)ue^{-[\theta_{0},\cdot]}u=s^{-1}_{v\rho}(f_{v,\rho})u and

    Φx(u)Φx(v)=e[θ,](svρ1(fv,ρ)u)e[θ,]v=svρ1(fv,ρ)e[θ,](uv)=qlsvρ1(fv,ρ)=Φx(uv),\varPhi_{x}(u)\varPhi_{x}(v)=e^{[\theta,\cdot]}(s^{-1}_{v\rho}(f_{v,\rho})u)e^{[\theta,\cdot]}v=s^{-1}_{v\rho}(f_{v,\rho})e^{[\theta,\cdot]}(uv)=q^{l}s^{-1}_{v\rho}(f_{v,\rho})=\varPhi_{x}(uv),

    i.e. the map Φx\varPhi_{x} preserves the product structure.

  • From the fact that Δ(θ)=0=Δ(θ0)\prescript{}{}{\Delta}(\theta)=0=\prescript{}{}{\Delta}(\theta_{0}), we see that e[θθ0,]e^{[\theta-\theta_{0},\cdot]} commutes with Δ\prescript{}{}{\Delta}, and hence Δ(Φx(n))=e[θθ0,]Δ(n)=0\prescript{}{}{\Delta}(\varPhi_{x}(\partial_{n}))=e^{[\theta-\theta_{0},\cdot]}\prescript{}{}{\Delta}(\partial_{n})=0. We also have [Φx(n1),Φx(n2)]=e[θθ0,][n1,n2]=0[\varPhi_{x}(\partial_{n_{1}}),\varPhi_{x}(\partial_{n_{2}})]=e^{[\theta-\theta_{0},\cdot]}[\partial_{n_{1}},\partial_{n_{2}}]=0.

  • Again from Δ(θ)=0=Δ(θ0)\prescript{}{}{\Delta}(\theta)=0=\prescript{}{}{\Delta}(\theta_{0}), we have

    Δ(Φx(u)Φx(n))\displaystyle\prescript{}{}{\Delta}(\varPhi_{x}(u)\varPhi_{x}(\partial_{n})) =Δ(e[θθ0,](un))=e[θθ0,](Δ(un))\displaystyle=\prescript{}{}{\Delta}(e^{[\theta-\theta_{0},\cdot]}(u\partial_{n}))=e^{[\theta-\theta_{0},\cdot]}\left(\prescript{}{}{\Delta}(u\partial_{n})\right)
    =mρ,ne[θθ0,](u)=mρ,nΦx(u)=Φx(Δ(un)).\displaystyle=\langle m_{\rho},n\rangle e^{[\theta-\theta_{0},\cdot]}(u)=\langle m_{\rho},n\rangle\varPhi_{x}(u)=\varPhi_{x}(\prescript{}{}{\Delta}(u\partial_{n})).
  • Finally, we have

    Δ(Φx(v)Φx(n))\displaystyle\prescript{}{}{\Delta}(\varPhi_{x}(v)\varPhi_{x}(\partial_{n})) =Δ(e[θθ0,]((e[θ0,]v)n))=e[θθ0,](Δ(svρ1(fv,ρ)vn))\displaystyle=\prescript{}{}{\Delta}\big{(}e^{[\theta-\theta_{0},\cdot]}((e^{[\theta_{0},\cdot]}v)\partial_{n})\big{)}=e^{[\theta-\theta_{0},\cdot]}\big{(}\prescript{}{}{\Delta}(s_{v\rho}^{-1}(f_{v,\rho})v\partial_{n})\big{)}
    =e[θθ0,](mρ,nsvρ1(fv,ρ)v+n(svρ1(fv,ρ))v)\displaystyle=e^{[\theta-\theta_{0},\cdot]}\big{(}\langle-m_{\rho},n\rangle s^{-1}_{v\rho}(f_{v,\rho})v+\partial_{n}(s^{-1}_{v\rho}(f_{v,\rho}))v\big{)}
    =mρ,n(e[θ,]v)+n(logsvρ1(fv,ρ))(e[θ,]v)\displaystyle=\langle-m_{\rho},n\rangle(e^{[\theta,\cdot]}v)+\partial_{n}\big{(}\log s^{-1}_{v\rho}(f_{v,\rho})\big{)}(e^{[\theta,\cdot]}v)
    =mρ,nΦx(v)+n(logsvρ1(fv,ρ))Φx(v)\displaystyle=\langle-m_{\rho},n\rangle\varPhi_{x}(v)+\partial_{n}\big{(}\log s^{-1}_{v\rho}(f_{v,\rho})\big{)}\varPhi_{x}(v)
    =Φx(Δ(vn)).\displaystyle=\varPhi_{x}(\prescript{}{}{\Delta}(v\partial_{n})).

We conclude that Φx:PkV0,x,𝖯𝖵sf,x,k\varPhi_{x}\colon\prescript{k}{}{PV}^{*,*}_{0,x}\rightarrow\prescript{k}{}{\mathsf{PV}}_{\mathrm{sf},x}^{*,*} is an isomorphism of dgBV algebras. We need to check that the map Φx\varPhi_{x} agrees with the isomorphism PkV0,|𝒞𝖯𝖵sf,k|𝒞\prescript{k}{}{PV}^{*,*}_{0}|_{\mathscr{C}}\rightarrow\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{C}} induced simply by the identity 𝒢0k|𝒞𝖦sfk|𝒞\prescript{k}{}{\mathcal{G}}^{*}_{0}|_{\mathscr{C}}\cong\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}}|_{\mathscr{C}}, where 𝒞=W0𝒮eintre(ρ)intre(ρ)\mathscr{C}=W_{0}\setminus\bigcup_{\mathscr{S}_{e}\cap\mathrm{int}_{\mathrm{re}}(\rho)\neq\emptyset}\mathrm{int}_{\mathrm{re}}(\rho). For this purpose, we consider two nearby maximal cells σ±\sigma_{\pm} such that σ+σ=ρ\sigma_{+}\cap\sigma_{-}=\rho. We have 𝕍k(σ±)=Specan([σ±1Pv]/qk+1)\prescript{k}{}{\mathbb{V}}(\sigma_{\pm})=\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\sigma^{-1}_{\pm}P_{v}]/q^{k+1}), and the gluing of 𝒢0k\prescript{k}{}{\mathcal{G}}^{*}_{0} over 𝒲ρ,vσ+\mathscr{W}_{\rho,v}\cap\sigma_{+} is given by parallel transporting through vv, and then by the formulae

(5.6) {zmsρσ+1(m)zmfor mΛρ,usρσ+1(mρ)zmρ,vqlsvσ+1(fv,ρ)sρσ+1(mρ)zmρ.\begin{cases}z^{m}\mapsto s^{-1}_{\rho\sigma_{+}}(m)z^{m}&\text{for $m\in\Lambda_{\rho}$},\\ u\mapsto s^{-1}_{\rho\sigma_{+}}(m_{\rho})z^{m_{\rho}},&\\ v\mapsto q^{l}s^{-1}_{v\sigma_{+}}(f_{v,\rho})s^{-1}_{\rho\sigma_{+}}(-m_{\rho})z^{-m_{\rho}}.\end{cases}

The only difference for gluing of 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} is the last equation in (5.6), which is now replaced by the formula vqlsρσ+1(mρ)zmρv\mapsto q^{l}s^{-1}_{\rho\sigma_{+}}(-m_{\rho})z^{-m_{\rho}}. Since we have

Φx(v)={svρ1(fv,ρ)von Uxintre(σ+),von Uxintre(σ)\varPhi_{x}(v)=\begin{dcases}s_{v\rho}^{-1}(f_{v,\rho})v&\text{on $U_{x}\cap\mathrm{int}_{\mathrm{re}}(\sigma_{+})$},\\ v&\text{on $U_{x}\cap\mathrm{int}_{\mathrm{re}}(\sigma_{-})$}\end{dcases}

on a sufficiently small neighborhood UxU_{x} of xx, we see that Φx(v)qlsvσ+1(fv,ρ)sρσ+1(mρ)zmρ\varPhi_{x}(v)\mapsto q^{l}s^{-1}_{v\sigma_{+}}(f_{v,\rho})s^{-1}_{\rho\sigma_{+}}(-m_{\rho})z^{-m_{\rho}} under the gluing map of 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} on Uxintre(σ+)U_{x}\cap\mathrm{int}_{\mathrm{re}}(\sigma_{+}). This shows the compatibility of Φx\varPhi_{x} with the gluing of 𝒢0k\prescript{k}{}{\mathcal{G}}^{*}_{0} and 𝖦sfk\prescript{k}{}{\mathsf{G}}^{*}_{\mathrm{sf}} over Uxintre(σ+)U_{x}\cap\mathrm{int}_{\mathrm{re}}(\sigma_{+}). A similar argument applies for Uxintre(σ)U_{x}\cap\mathrm{int}_{\mathrm{re}}(\sigma_{-}).

The proof for Φ:𝒜0,k𝖠sf,k\varPhi\colon\prescript{k}{}{\mathcal{A}}^{*,*}_{0}\rightarrow\prescript{k}{}{\mathsf{A}}^{*,*}_{\mathrm{sf}} is similar and will be omitted. The volume form is preserved under Φ\varPhi because we have Δ(θ)=0=Δ(θ0)\prescript{}{}{\Delta}(\theta)=0=\prescript{}{}{\Delta}(\theta_{0}). This completes the proof of the lemma. ∎

5.2.4. A global sheaf of dgLas from gluing of the semi-flat sheaves

We shall apply the procedure described in §4.3 to the semi-flat sheaves to glue a global sheaf of dgLas. First of all, we choose an open cover {Wα}α\{W_{\alpha}\}_{\alpha\in\mathscr{I}} satisfying the Condition 4.1, together with a decomposition =12\mathscr{I}=\mathscr{I}_{1}\sqcup\mathscr{I}_{2} such that 𝒲1={Wα}α1\mathcal{W}_{1}=\{W_{\alpha}\}_{\alpha\in\mathscr{I}_{1}} is a cover of the semi-flat part W0W_{0}, and 𝒲2={Wα}α2\mathcal{W}_{2}=\{W_{\alpha}\}_{\alpha\in\mathscr{I}_{2}} is a cover of a neighborhood of (τ𝒫[n2]τ)(ρ𝒮e𝒮intre(ρ))\big{(}\bigcup_{\tau\in\mathscr{P}^{[n-2]}}\tau\big{)}\cup\big{(}\bigcup_{\rho\cap\mathscr{S}_{e}\neq\emptyset}\mathscr{S}\cap\mathrm{int}_{\mathrm{re}}(\rho)\big{)}.

For each WαW_{\alpha}, we have a compatible set of local sheaves 𝒢αk\prescript{k}{}{\mathcal{G}}^{*}_{\alpha} of BV algebras, local sheaves 𝒦αk\prescript{k}{}{\mathcal{K}}^{*}_{\alpha} of dgas, and relative volume elements ωαk\prescript{k}{}{\omega}_{\alpha}, kk\in\mathbb{N} (as in §4.1). We can further demand that, over the semi-flat part W0W_{0}, we have 𝒢αk=𝒢0k|Wα\prescript{k}{}{\mathcal{G}}^{*}_{\alpha}=\prescript{k}{}{\mathcal{G}}^{*}_{0}|_{W_{\alpha}}, 𝒦αk=𝒦0k|Wα\prescript{k}{}{\mathcal{K}}^{*}_{\alpha}=\prescript{k}{}{\mathcal{K}}^{*}_{0}|_{W_{\alpha}} and ωαk=ω0k|Wα\prescript{k}{}{\omega}_{\alpha}=\prescript{k}{}{\omega}_{0}|_{W_{\alpha}}, and hence PkVα,=PkV0,|Wα\prescript{k}{}{PV}^{*,*}_{\alpha}=\prescript{k}{}{PV}^{*,*}_{0}|_{W_{\alpha}} and 𝒜α,k=𝒜0,k|Wα\prescript{k}{}{\mathcal{A}}^{*,*}_{\alpha}=\prescript{k}{}{\mathcal{A}}^{*,*}_{0}|_{W_{\alpha}} for α1\alpha\in\mathscr{I}_{1}.

Using the construction in §4.3, we obtain a Gerstenhaber deformation gαβk=e[θαβ,]ψαβk\prescript{k}{}{g}_{\alpha\beta}=e^{[\theta_{\alpha\beta},\cdot]}\circ\prescript{k}{}{\psi}_{\alpha\beta} specified by θαβPkVβ1,0(Wαβ)\theta_{\alpha\beta}\in\prescript{k}{}{PV}^{-1,0}_{\beta}(W_{\alpha\beta}), which give rise to sets of compatible global sheaves PkV,\prescript{k}{}{PV}^{*,*} and 𝒜,k\prescript{k}{}{\mathcal{A}}^{*,*}, kk\in\mathbb{N}. Restricting to the semi-flat part, we get two Gerstenhaber deformations PkV0,\prescript{k}{}{PV}^{*,*}_{0} and PkV,|W0\prescript{k}{}{PV}^{*,*}|_{W_{0}}, which must be equivalent as Hˇ>0(𝒲1,P0V1,0|W0)=0\check{H}^{>0}(\mathcal{W}_{1},\prescript{0}{}{PV}^{-1,0}|_{W_{0}})=0. So we have a set of compatible isomorphisms locally given by hα=e[𝐛α,]:PkV0,|WαPkV,|WαPkVα,h_{\alpha}=e^{[\mathbf{b}_{\alpha},\cdot]}\colon\prescript{k}{}{PV}^{*,*}_{0}|_{W_{\alpha}}\rightarrow\prescript{k}{}{PV}^{*,*}|_{W_{\alpha}}\cong\prescript{k}{}{PV}^{*,*}_{\alpha} for some 𝐛αPkV01,0(Wα)\mathbf{b}_{\alpha}\in\prescript{k}{}{PV}^{-1,0}_{0}(W_{\alpha}), for each kk\in\mathbb{N}, and they fit into the following commutative diagram

PkV0,|Wαβ\textstyle{\prescript{k}{}{PV}^{*,*}_{0}|_{W_{\alpha\beta}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathrm{id}}hα\scriptstyle{h_{\alpha}}PkV0,|Wαβ\textstyle{\prescript{k}{}{PV}^{*,*}_{0}|_{W_{\alpha\beta}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}hβ\scriptstyle{h_{\beta}}PkVα,|Wαβ\textstyle{\prescript{k}{}{PV}^{*,*}_{\alpha}|_{W_{\alpha\beta}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}gαβk\scriptstyle{\prescript{k}{}{g}_{\alpha\beta}}PkVβ,|Wαβ.\textstyle{\prescript{k}{}{PV}^{*,*}_{\beta}|_{W_{\alpha\beta}}.}

Since the pre-differential on PkV,|W0\prescript{k}{}{PV}^{*,*}|_{W_{0}} obtained from the construction in §4.3 is of the form ¯α+[ηα,]\bar{\partial}_{\alpha}+[\eta_{\alpha},\cdot] for some ηαPkV01,1(Wα)\eta_{\alpha}\in\prescript{k}{}{PV}^{-1,1}_{0}(W_{\alpha}), pulling back via hαh_{\alpha} gives a global element ηPkV01,1(W0)\eta\in\prescript{k}{}{PV}^{-1,1}_{0}(W_{0}) such that

hα1(¯α+[ηα,])hα=¯0+[η,].h_{\alpha}^{-1}\circ(\bar{\partial}_{\alpha}+[\eta_{\alpha},\cdot])\circ h_{\alpha}=\bar{\partial}_{0}+[\eta,\cdot].

Theorem 4.18 gives a Maurer–Cartan solution ϕPkV1,1(B)\phi\in\prescript{k}{}{PV}^{-1,1}(B) such that (¯+[ϕ,])2=0(\bar{\partial}+[\phi,\cdot])^{2}=0, together with a holomorphic volume form efωe^{f}\prescript{}{}{\omega}, compatible for each kk. We denote the pullback of ϕ\phi under hαh_{\alpha}’s to PkV01,1(W0)\prescript{k}{}{PV}^{-1,1}_{0}(W_{0}) as ϕ0\phi_{0}, and that of volume form to 𝒜0n,0k(W0)\prescript{k}{\parallel}{\mathcal{A}}^{n,0}_{0}(W_{0}) as egω0e^{g}\prescript{}{}{\omega}_{0}. We see that the equation

(¯0+η+ϕ0)egω0=0(\bar{\partial}_{0}+\mathcal{L}_{\eta+\phi_{0}})e^{g}\prescript{}{}{\omega}_{0}=0

is satisfied, or equivalently, that η+ϕ0+tg\eta+\phi_{0}+tg is a solution to the extended Maurer–Cartan equation 4.10.

Lemma 5.11.

If the holomorphic volume form efωe^{f}\prescript{}{}{\omega} is normalized in the sense of Definition 4.19, then we can find a set of compatible 𝒱PkV01,0(W0)\mathcal{V}\in\prescript{k}{}{PV}^{-1,0}_{0}(W_{0}), kk\in\mathbb{N} such that

e𝒱ω0=egω0.e^{-\mathcal{L}_{\mathcal{V}}}\prescript{}{}{\omega}_{0}=e^{g}\prescript{}{}{\omega}_{0}.

As a consequence, the Maurer–Cartan solution η+ϕ0+tg\eta+\phi_{0}+tg is gauge equivalent to a solution of the form ζ0+t0\zeta_{0}+t\cdot 0 for some ζ0PkV01,1(W0)\zeta_{0}\in\prescript{k}{}{PV}^{-1,1}_{0}(W_{0}), via the gauge transformation e[𝒱,]:PkV0,PkV0,e^{[\mathcal{V},\cdot]}\colon\prescript{k}{}{PV}^{*,*}_{0}\rightarrow\prescript{k}{}{PV}^{*,*}_{0}.

Proof.

We should construct 𝒱\mathcal{V} by induction on kk as in the proof of Lemma 4.6. Namely, suppose 𝒱\mathcal{V} is constructed for the (k1)st(k-1)^{\text{st}}-order, then we shall lift it to the kthk^{\text{th}}-order. We prove the existence of a lifting 𝒱xPkV0,x1,0\mathcal{V}_{x}\in\prescript{k}{}{PV}^{-1,0}_{0,x} at every stalk xW0x\in W_{0} and use partition of unity to glue a global lifting 𝒱\mathcal{V}.

First of all, we can always find a gauge transformation θPkV0,x1,0\theta\in\prescript{k}{}{PV}^{-1,0}_{0,x} such that

e[θ,]¯0e[θ,]=¯0+[η+ϕ0,].e^{-[\theta,\cdot]}\circ\bar{\partial}_{0}\circ e^{[\theta,\cdot]}=\bar{\partial}_{0}+[\eta+\phi_{0},\cdot].

So we have ¯0(eθegω0)=0\bar{\partial}_{0}(e^{\mathcal{L}_{\theta}}e^{g}\prescript{}{}{\omega}_{0})=0, which implies that eθegω0𝒦0,xnke^{\mathcal{L}_{\theta}}e^{g}\prescript{}{}{\omega}_{0}\in\prescript{k}{\parallel}{\mathcal{K}}^{n}_{0,x}. We can write eθegω0=ehω0e^{\mathcal{L}_{\theta}}e^{g}\prescript{}{}{\omega}_{0}=e^{h}\prescript{}{}{\omega}_{0} in the stalk at xx for some germ h𝒢0,x0kh\in\prescript{k}{}{\mathcal{G}}^{0}_{0,x} of holomorphic functions. Applying Lemma 4.6, we can further choose θ\theta so that h=h(q)(q)[q]/qk+1h=h(q)\in(q)\subset\mathbb{C}[q]/q^{k+1}. In a sufficiently small neighborhood UxU_{x}, we find an element ϱx𝒯n(Ux)\varrho_{x}\in\mathscr{T}^{n}(U_{x}) as in Definition 4.19. The fact that the volume form is normalized forces eh(q)[ω0ϱx]e^{h(q)}[\prescript{}{}{\omega}_{0}\wedge\varrho_{x}] to be constant with respect to the Gauss–Manin connection k\prescript{k}{}{\nabla}. Tracing through the exact sequence (4.14) on UxU_{x}, we can lift ω0\prescript{}{}{\omega}_{0} to 𝒦0nk(Ux)\prescript{k}{}{\mathcal{K}}^{n}_{0}(U_{x}) which is closed under \prescript{}{}{\partial}. As a consequence, we have logqk[ω0ϱx]=0\prescript{k}{}{\nabla}_{\frac{\partial}{\partial\log q}}[\prescript{}{}{\omega}_{0}\wedge\varrho_{x}]=0, and hence we conclude that h(q)=0h(q)=0.

Now we have to solve for a lifting 𝒱x\mathcal{V}_{x} such that eθe𝒱xω0=ω0e^{\mathcal{L}_{\theta}}e^{-\mathcal{L}_{\mathcal{V}_{x}}}\prescript{}{}{\omega}_{0}=\prescript{}{}{\omega}_{0} up to the kthk^{\text{th}}-order. This is equivalent to solving for a lifting uu satisfying euω0=ω0e^{\mathcal{L}_{u}}\prescript{}{}{\omega}_{0}=\prescript{}{}{\omega}_{0} for the kthk^{\text{th}}-order once the (k1)st(k-1)^{\text{st}}-order is given. Take an arbitrary lifting u~\tilde{u} to the kthk^{\text{th}}-order, and making use of the formula in [8, Lem. 2.8], we have

eu~ω0=exp(s=0δu~s(s+1)!Δ(u~))ω0,e^{\mathcal{L}_{\tilde{u}}}\prescript{}{}{\omega}_{0}=\exp\left(\sum_{s=0}^{\infty}\frac{\delta_{\tilde{u}}^{s}}{(s+1)!}\prescript{}{}{\Delta}(\tilde{u})\right)\prescript{}{}{\omega}_{0},

where δu~=[u~,]\delta_{\tilde{u}}=-[\tilde{u},\cdot]. From eu~ω0=ω0(mod 𝐦k)e^{\mathcal{L}_{\tilde{u}}}\prescript{}{}{\omega}_{0}=\prescript{}{}{\omega}_{0}\ (\text{mod $\mathbf{m}^{k}$}), we use induction on the order jj to prove that Δ(u~)=0\prescript{}{}{\Delta}(\tilde{u})=0 up to order (k1)(k-1). Therefore we can write

Δ(u~)=qkΔ(u˘)(mod 𝐦k)\prescript{}{}{\Delta}(\tilde{u})=q^{k}\prescript{}{}{\Delta}(\breve{u})\ (\text{mod $\mathbf{m}^{k}$})

for some u˘P0V0,x1,0\breve{u}\in\prescript{0}{}{PV}^{-1,0}_{0,x}, by the fact that the cohomology sheaf under Δ\prescript{}{}{\Delta} is free over Rk=[q]/(qk+1)\prescript{k}{}{R}=\mathbb{C}[q]/(q^{k+1}) (see the discussion right after Condition 4.14). Setting u=u~qku˘u=\tilde{u}-q^{k}\breve{u} will then solve the equation. ∎

The element 𝒱\mathcal{V} obtained in Lemma 5.11 can be used to conjugate the operator ¯0+[η+ϕ0,]\bar{\partial}_{0}+[\eta+\phi_{0},\cdot] to get ¯0+[ζ0,]\bar{\partial}_{0}+[\zeta_{0},\cdot], i.e.

e[𝒱,](¯0+[ζ0,])e[𝒱,]=¯0+[η+ϕ0,].e^{-[\mathcal{V},\cdot]}\circ(\bar{\partial}_{0}+[\zeta_{0},\cdot])\circ e^{[\mathcal{V},\cdot]}=\bar{\partial}_{0}+[\eta+\phi_{0},\cdot].

The volume form ω0\prescript{}{}{\omega}_{0} will be holomorphic under the operator ¯0+[ζ0,]\bar{\partial}_{0}+[\zeta_{0},\cdot]. From the equation (4.13), we observe that Δ(ζ0)=0\prescript{}{}{\Delta}(\zeta_{0})=0. Furthermore, the image of ζ0\zeta_{0} under the isomorphism Φ:PkV0,𝖯𝖵sf,k\varPhi\colon\prescript{k}{}{PV}^{*,*}_{0}\rightarrow\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}} in Lemma 5.10 gives ϕs𝖯𝖵sf1,1k(W0)\phi_{\mathrm{s}}\in\prescript{k}{}{\mathsf{PV}}^{-1,1}_{\mathrm{sf}}(W_{0}), and an operator of the form

(5.7) ¯+[ϕin+ϕs,]=¯+v,ρ[ϕv,ρ,]+[ϕs,],\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot]=\bar{\partial}_{\circ}+\sum_{v,\rho}[\phi_{v,\rho},\cdot]+[\phi_{\mathrm{s}},\cdot],

where ϕin=v,ρϕv,ρ\phi_{\mathrm{in}}=\sum_{v,\rho}\phi_{v,\rho}, that acts on 𝖯𝖵sf,k\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}.

Equipping with this operator, the semi-flat sheaf 𝖯𝖵sf,k\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}} can be glued to the sheaves PkVα,\prescript{k}{}{PV}^{*,*}_{\alpha}’s for α2\alpha\in\mathscr{I}_{2}, preserving all the operators. More explicitly, on each overlap W0α:=W0WαW_{0\alpha}:=W_{0}\cap W_{\alpha}, we have

(5.8) g0αk:𝖯𝖵sf,k|W0αPkV,|W0α\prescript{k}{}{g}_{0\alpha}\colon\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{W_{0\alpha}}\rightarrow\prescript{k}{}{PV}^{*,*}|_{W_{0\alpha}}

defined by

gαβkg0αk|Wαβ:=hβe[𝒱,]Φ1|Wαβ\prescript{k}{}{g}_{\alpha\beta}\circ\prescript{k}{}{g}_{0\alpha}|_{W_{\alpha\beta}}:=h_{\beta}\circ e^{-[\mathcal{V},\cdot]}\circ\varPhi^{-1}|_{W_{\alpha\beta}}

for β1\beta\in\mathscr{I}_{1}, which sends the operator ¯+[ϕin+ϕs,]\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot] to ¯α+[ηα+ϕ,]\bar{\partial}_{\alpha}+[\eta_{\alpha}+\phi,\cdot].

Definition 5.12.

We call 𝖳𝖫sfk:=Ker(Δ)[1]𝖯𝖵sf1,k[1]\prescript{k}{}{\mathsf{TL}}^{*}_{\mathrm{sf}}:=\mathrm{Ker}(\prescript{}{}{\Delta})[-1]\subset\prescript{k}{}{\mathsf{PV}}^{-1,*}_{\mathrm{sf}}[-1], equipped with the structure of a dgLa using ¯\bar{\partial}_{\circ} and [,][\cdot,\cdot] inherited from 𝖯𝖵sf1,k\prescript{k}{}{\mathsf{PV}}^{-1,*}_{\mathrm{sf}}, the sheaf of semi-flat tropical vertex differential graded Lie algebras (abbrev. as sf-TVdgLa).

Note that 𝖳𝖫sfk𝒯|W0𝔥k\prescript{k}{}{\mathsf{TL}}^{*}_{\mathrm{sf}}\cong\mathscr{T}^{*}|_{W_{0}}\otimes\prescript{k}{}{\mathfrak{h}}. Also, we have Δ(ϕs)=0\prescript{}{}{\Delta}(\phi_{\mathrm{s}})=0 since Δ(ζ0)=0\prescript{}{}{\Delta}(\zeta_{0})=0, and a direct computation shows that Δ(ϕin)=0\prescript{}{}{\Delta}(\phi_{\mathrm{in}})=0. Thus ϕin,ϕs𝖳𝖫sf1k(W0)\phi_{\mathrm{in}},\phi_{\mathrm{s}}\in\prescript{k}{}{\mathsf{TL}}^{1}_{\mathrm{sf}}(W_{0}), and the operator ¯+[ϕin+ϕs,]\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot] preserves the sub-dgLa 𝖳𝖫sfk\prescript{k}{}{\mathsf{TL}}^{*}_{\mathrm{sf}}.

From the description of the sheaf 𝒯\mathscr{T}^{*}, we can see that locally on UW0U\subset W_{0}, ϕs\phi_{\mathrm{s}} is supported on finitely many codimension one polyhedral subsets, called walls or slabs, which are constituents of a scattering diagram. This is why we use the subscript ‘s’ in ϕs\phi_{\mathrm{s}}, which stands for ‘scattering’.

5.3. Consistent scattering diagrams and Maurer–Cartan solutions

5.3.1. Scattering diagrams

In this subsection, we recall the notion of scattering diagrams introduced by Kontsevich–Soibelman [36] and Gross–Siebert [29], and make modifications to suit our needs. We begin with the notion of walls from [29, §2]. Let

𝒮^=(τ𝒫[n2]τ)(ρ𝒫[n1]ρ𝒮e𝒮intre(ρ))\hat{\mathscr{S}}=\left(\bigcup_{\tau\in\mathscr{P}^{[n-2]}}\tau\right)\cup\left(\bigcup_{\begin{subarray}{c}\rho\in\mathscr{P}^{[n-1]}\\ \rho\cap\mathscr{S}_{e}\neq\emptyset\end{subarray}}\mathscr{S}\cap\mathrm{int}_{\mathrm{re}}(\rho)\right)

be equipped with a polyhedral decomposition induced from 𝒫\mathscr{P} and 𝒮\mathscr{S}. For the exposition below, we will always fix k>0k>0 and consider all these structures modulo 𝐦k+1=(qk+1)\mathbf{m}^{k+1}=(q^{k+1}).

Definition 5.13.

A wall (𝐰,σ𝐰,dˇ𝐰,Θ𝐰)(\mathbf{w},\sigma_{\mathbf{w}},\check{d}_{\mathbf{w}},\Theta_{\mathbf{w}}) consists of

  • a maximal cell σ𝐰𝒫[n]\sigma_{\mathbf{w}}\in\mathscr{P}^{[n]},

  • a closed (n1)(n-1)-dimensional tropical polyhedral subset 𝐰\mathbf{w} of σ𝐰\sigma_{\mathbf{w}} such that

    intre(𝐰)(ρ𝒫[n1]ρ𝒮eintre(ρ))=,\mathrm{int}_{\mathrm{re}}(\mathbf{w})\cap\left(\bigcup_{\begin{subarray}{c}\rho\in\mathscr{P}^{[n-1]}\\ \rho\cap\mathscr{S}_{e}\neq\emptyset\end{subarray}}\mathrm{int}_{\mathrm{re}}(\rho)\right)=\emptyset,
  • a choice of a primitive normal dˇ𝐰\check{d}_{\mathbf{w}}, and

  • a section Θ𝐰\Theta_{\mathbf{w}} of the tropical vertex group exp(q𝔥k)\exp(q\cdot\prescript{k}{}{\mathfrak{h}}) over a sufficiently small neighborhood of 𝐰\mathbf{w}.

We call Θ𝐰\Theta_{\mathbf{w}} the wall-crossing factor associated to the wall 𝐰\mathbf{w}. We may write a wall as (𝐰,Θ𝐰)(\mathbf{w},\Theta_{\mathbf{w}}) for simplicity.

A wall cannot be contained in ρ\rho with ρ𝒮e\rho\cap\mathscr{S}_{e}\neq\emptyset. We define a notion of slabs for these subsets of codimension one strata ρ\rho intersecting 𝒮e\mathscr{S}_{e}. The difference is that we have an extra term Θv,ρ\varTheta_{v,\rho} coming from the slab function fv,ρf_{v,\rho}.

Definition 5.14.

A slab (𝐛,ρ𝐛,dˇρ,Ξ𝐛)(\mathbf{b},\rho_{\mathbf{b}},\check{d}_{\rho},\varXi_{\mathbf{b}}) consists of

  • an (n1)(n-1)-cell ρ𝐛𝒫[n1]\rho_{\mathbf{b}}\in\mathscr{P}^{[n-1]} such that ρ𝐛𝒮e\rho_{\mathbf{b}}\cap\mathscr{S}_{e}\neq\emptyset,

  • a closed (n1)(n-1)-dimensional tropical polyhedral subset 𝐛\mathbf{b} of ρ𝐛(ρ𝐛𝒮)\rho_{\mathbf{b}}\setminus(\rho_{\mathbf{b}}\cap\mathscr{S}),

  • a choice of a primitive normal dˇρ\check{d}_{\rho}, and

  • a section Ξ𝐛\varXi_{\mathbf{b}} of exp(q𝔥k)\exp(q\cdot\prescript{k}{}{\mathfrak{h}}) over a sufficiently small neighborhood of 𝐛\mathbf{b}.

The wall-crossing factor associated to the slab 𝐛\mathbf{b} is given by

Θ𝐛:=Θv,ρΞ𝐛,\Theta_{\mathbf{b}}:=\varTheta_{v,\rho}\circ\varXi_{\mathbf{b}},

where vv is the unique vertex such that intre(ρ)v\mathrm{int}_{\mathrm{re}}(\rho)_{v} contains 𝐛\mathbf{b} and

Θv,ρ=exp([log(svρ1(fv,ρ))dˇρ,])\varTheta_{v,\rho}=\exp([\log(s_{v\rho}^{-1}(f_{v,\rho}))\partial_{\check{d}_{\rho}},\cdot])

(cf. equation (5.4)). We may write a slab as (𝐛,Θ𝐛)(\mathbf{b},\Theta_{\mathbf{b}}) for simplicity.

Remark 5.15.

In the above definition, a slab is not allowed to intersect the singular locus 𝒮\mathscr{S}. This is different from the situation in [29, §2]. However, in our definition of consistent scattering diagrams, we will require consistency around each stratum of 𝒮\mathscr{S}.

Example 5.16.

We consider the 3-dimensional example shown in Figure 7, from which we can see possible supports of the walls and slabs. There are two adjacent maximal cells intersecting at ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]} with 𝒮eρ=𝒮ρ\mathscr{S}_{e}\cap\rho=\mathscr{S}\cap\rho colored in red. The 22-dimensional polyhedral subsets colored in blue can support walls and the polyhedral subset colored in green can support a slab because it is lying inside ρ\rho with 𝒮eρ\mathscr{S}_{e}\cap\rho\neq\emptyset.

Refer to caption
Figure 7. Supports of walls/slabs
Definition 5.17.

A (kthk^{\text{th}}-order) scattering diagram is a countable collection

𝒟={(𝐰i,Θi)}i{(𝐛j,Θj)}j\mathscr{D}=\{(\mathbf{w}_{i},\Theta_{i})\}_{i\in\mathbb{N}}\cup\{(\mathbf{b}_{j},\Theta_{j})\}_{j\in\mathbb{N}}

of walls or slabs such that the intersections of any two walls/slabs is at most an (n2)(n-2)-dimensional tropical polyhedral subset, and {𝐰iW0}i{𝐛jW0}j\{\mathbf{w}_{i}\cap W_{0}\}_{i\in\mathbb{N}}\cup\{\mathbf{b}_{j}\cap W_{0}\}_{j\in\mathbb{N}} is locally finite in W0W_{0}.

Our notion of scattering diagrams is more flexible than the one defined in [36, 29] in two ways: First, there is no relation between the affine direction orthogonal to a wall 𝐰\mathbf{w} or a slab 𝐛\mathbf{b} and its wall crossing factor. As a result, we cannot allow overlapping of walls/slabs in their relative interior because in that case their associated wall crossing factors are not necessarily commuting. Second, we only require that the intersection of 𝒟\mathscr{D} with W0W_{0} is a locally finite collection of W0W_{0}, which implies that we allow a possibly infinite number of walls/slabs approaching strata of 𝒮^\hat{\mathscr{S}}. In the construction of the scattering diagram 𝒟(φ)\mathscr{D}(\varphi) associated to a Maurer–Cartan solution φ\varphi below, all the walls/slabs will be compact subsets of W0W_{0}. These walls will not intersect 𝒮^\hat{\mathscr{S}}, as illustrated in Figure 7. However, there could be a union of infinitely many walls limiting to some strata of 𝒮^\hat{\mathscr{S}}. See also Remark 1.2.

Example 5.18.

For the 2-dimensional example shown in Figure 8, we see a vertex vv and its adjacent cells, and the singular locus 𝒮e\mathscr{S}_{e} consists of the red crosses. In our version of scattering diagrams, we allow infinitely many intervals limiting to {v}\{v\} or 𝒮e\mathscr{S}_{e}.

Refer to caption
Figure 8. Walls/slabs around 𝒮^\hat{\mathscr{S}}

Given a scattering diagram 𝒟\mathscr{D}, we can define its support as |𝒟|:=i𝐰ij𝐛j|\mathscr{D}|:=\bigcup_{i\in\mathbb{N}}\mathbf{w}_{i}\cup\bigcup_{j\in\mathbb{N}}\mathbf{b}_{j}. There is an induced polyhedral decomposition on |𝒟||\mathscr{D}| such that its (n1)(n-1)-cells are closed subsets of some walls or slabs, and all intersections of walls or slabs are lying in the union of the (n2)(n-2)-cells. We write |𝒟|[i]|\mathscr{D}|^{[i]} for the collection of all the ii-cells in this polyhedral decomposition. We may assume, after further subdividing the walls or slabs in 𝒟\mathscr{D} if necessary, that every wall or slab is an (n1)(n-1)-cell in |𝒟||\mathscr{D}|. We call an (n2)(n-2)-cell 𝔧|𝒟|[n2]\mathfrak{j}\in|\mathscr{D}|^{[n-2]} a joint, and a connected component of W0|𝒟|W_{0}\setminus|\mathscr{D}| a chamber.

Given a wall or slab, we shall make sense of wall crossing in terms of jumping of holomorphic functions across it. Instead of formulating the definition in terms of path-ordered products of elements in the tropical vertex group as in [29], we will express it in terms of the action by the tropical vertex group on the local sections of 𝖦sf0k\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf}}. There is no harm in doing so since we have the inclusion 𝖦sf1kDer(𝖦sf0k,𝖦sf0k)\prescript{k}{}{\mathsf{G}}^{-1}_{\mathrm{sf}}\hookrightarrow\mathrm{Der}(\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf}},\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf}}), i.e. a relative vector field is determined by its action on functions.

In this regard, we would like to define the (kthk^{\text{th}}-order) wall-crossing sheaf 𝒪𝒟k\prescript{k}{}{\mathscr{O}}_{\mathscr{D}} on the open set

W0(𝒟):=W0𝔧|𝒟|[n2]𝔧,W_{0}(\mathscr{D}):=W_{0}\setminus\bigcup_{\mathfrak{j}\in|\mathscr{D}|^{[n-2]}}\mathfrak{j},

which captures the jumping of holomorphic functions described by the wall-crossing factor when crossing a wall/slab. We first consider the sheaf 𝖦sf0k\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf}} of holomorphic functions over the subset W0|𝒟|W_{0}\setminus|\mathscr{D}|, and let

𝒪𝒟k|W0|𝒟|:=𝖦sf0k|W0|𝒟|.\prescript{k}{}{\mathscr{O}}_{\mathscr{D}}|_{W_{0}\setminus|\mathscr{D}|}:=\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf}}|_{W_{0}\setminus|\mathscr{D}|}.

To extend it through the walls/slabs, we will specify the analyic continuation through intre(𝐰)\mathrm{int}_{\mathrm{re}}(\mathbf{w}) for each 𝐰|𝒟|[n1]\mathbf{w}\in|\mathscr{D}|^{[n-1]}. Given a wall/slab 𝐰\mathbf{w} with two adjacent chambers 𝒞+\mathcal{C}_{+}, 𝒞\mathcal{C}_{-} and dˇ𝐰\check{d}_{\mathbf{w}} pointing into 𝒞+\mathcal{C}_{+}, and a point xintre(𝐰)x\in\mathrm{int}_{\mathrm{re}}(\mathbf{w}) with the germ Θ𝐰,x\Theta_{\mathbf{w},x} of wall-crossing factors near xx, we let

𝒪𝒟,xk:=𝖦sf,x0k,\prescript{k}{}{\mathscr{O}}_{\mathscr{D},x}:=\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf},x},

but with a different gluing to nearby chambers 𝒞±\mathcal{C}_{\pm}: in a sufficiently small neighborhood UxU_{x} of xx, the gluing of a local section f𝒪𝒟,xkf\in\prescript{k}{}{\mathscr{O}}_{\mathscr{D},x} is given by

(5.9) f|Ux𝒞±:={Θ𝐰,x(f)|Ux𝒞+on Ux𝒞+,f|Ux𝒞on Ux𝒞.f|_{U_{x}\cap\mathcal{C}_{\pm}}:=\begin{dcases}\Theta_{\mathbf{w},x}(f)|_{U_{x}\cap\mathcal{C}_{+}}&\text{on $U_{x}\cap\mathcal{C}_{+}$,}\\ f|_{U_{x}\cap\mathcal{C}_{-}}&\text{on $U_{x}\cap\mathcal{C}_{-}$.}\end{dcases}

In this way, the sheaf 𝒪𝒟k|W0|𝒟|\prescript{k}{}{\mathscr{O}}_{\mathscr{D}}|_{W_{0}\setminus|\mathscr{D}|} extends to W0(𝒟)W_{0}(\mathscr{D}).

Now we can formulate consistency of a scattering diagram 𝒟\mathscr{D} in terms of the behaviour of the sheaf 𝒪𝒟k\prescript{k}{}{\mathscr{O}}_{\mathscr{D}} over the joints 𝔧\mathfrak{j}’s and (n2)(n-2)-dimensional strata of 𝒮^\hat{\mathscr{S}}. More precisely, we consider the push-forward 𝔦(𝒪𝒟k)\mathfrak{i}_{*}(\prescript{k}{}{\mathscr{O}}_{\mathscr{D}}) along the embedding 𝔦:W0(𝒟)B\mathfrak{i}\colon W_{0}(\mathscr{D})\rightarrow B, and its stalk at xintre(𝔧)x\in\mathrm{int}_{\mathrm{re}}(\mathfrak{j}) and xintre(τ)x\in\mathrm{int}_{\mathrm{re}}(\tau) for strata τ𝒮^\tau\subset\hat{\mathscr{S}}. Similar to above, we can define the (lthl^{\text{th}}-order) sheaf 𝒪𝒟l\prescript{l}{}{\mathscr{O}}_{\mathscr{D}} by using 𝖦sf0l\prescript{l}{}{\mathsf{G}}^{0}_{\mathrm{sf}} and considering equation (5.9) modulo (q)l+1(q)^{l+1}. There is a natural restriction map k,l:𝔦(𝒪𝒟k)𝔦(𝒪𝒟l)\prescript{k,l}{}{\flat}\colon\mathfrak{i}_{*}(\prescript{k}{}{\mathscr{O}}_{\mathscr{D}})\rightarrow\mathfrak{i}_{*}(\prescript{l}{}{\mathscr{O}}_{\mathscr{D}}). Taking tensor product, we have k,l:𝔦(𝒪𝒟k)RkRl𝔦(𝒪𝒟l)\prescript{k,l}{}{\flat}\colon\mathfrak{i}_{*}(\prescript{k}{}{\mathscr{O}}_{\mathscr{D}})\otimes_{\prescript{k}{}{R}}\prescript{l}{}{R}\rightarrow\mathfrak{i}_{*}(\prescript{l}{}{\mathscr{O}}_{\mathscr{D}}), where Rk=[q]/(qk+1)\prescript{k}{}{R}=\mathbb{C}[q]/(q^{k+1}).

The proof of the following lemma will be given in Appendix §A.

Lemma 5.19 (Hartogs extension property).

We have

ι(𝒢00|W0)=𝒢00,\iota_{*}(\prescript{0}{}{\mathcal{G}}^{0}|_{W_{0}})=\prescript{0}{}{\mathcal{G}}^{0},

where ι:W0B\iota\colon W_{0}\rightarrow B is the inclusion. Moreover, for any scattering diagram 𝒟\mathscr{D}, we have

𝔦(𝒢00|W0(𝒟))=𝒢00,\mathfrak{i}_{*}(\prescript{0}{}{\mathcal{G}}^{0}|_{W_{0}(\mathscr{D})})=\prescript{0}{}{\mathcal{G}}^{0},

where 𝔦:W0(𝒟)B\mathfrak{i}\colon W_{0}(\mathscr{D})\rightarrow B is the inclusion.

Lemma 5.20.

The 0th0^{\text{th}}-order sheaf 𝔦(𝒪𝒟0)\mathfrak{i}_{*}(\prescript{0}{}{\mathscr{O}}_{\mathscr{D}}) is isomorphic to the sheaf 𝒢00\prescript{0}{}{\mathcal{G}}^{0}.

Proof.

In view of Lemma 5.19, we only have to show that the two sheaves are isomorphic on the open subset W0(𝒟)W_{0}(\mathscr{D}). Since we work modulo (q)(q), only the wall-crossing factor Θv,ρ\varTheta_{v,\rho} associated to a slab matters. So we take a point xintre(𝐛)intre(ρ)vx\in\mathrm{int}_{\mathrm{re}}(\mathbf{b})\subset\mathrm{int}_{\mathrm{re}}(\rho)_{v} for some vertex vv, and compare 𝒪𝒟,x0\prescript{0}{}{\mathscr{O}}_{\mathscr{D},x} with 𝒢x00=𝖦sf,x00\prescript{0}{}{\mathcal{G}}^{0}_{x}=\prescript{0}{}{\mathsf{G}}^{0}_{\mathrm{sf},x}. From the proof of Lemma 5.10, we have

𝒢x00\displaystyle\prescript{0}{}{\mathcal{G}}^{0}_{x} =𝖦sf,x00=𝒪𝕍k(ρ)v,K\displaystyle=\prescript{0}{}{\mathsf{G}}^{0}_{\mathrm{sf},x}=\mathcal{O}_{\prescript{k}{}{\mathbb{V}}(\rho)_{v},K}
={a0,j+i=1aiui+i=1aivi|ai𝒪()n1(U)for some neigh. UK,supi0log|ai||i|<},\displaystyle=\left\{a_{0,j}+\sum_{i=1}^{\infty}a_{i}u^{i}+\sum_{i=-1}^{-\infty}a_{i}v^{-i}\ \Big{|}\ a_{i}\in\mathcal{O}_{(\mathbb{C}^{*})^{n-1}}(U)\ \text{for some neigh. $U\supset K$},\ \sup_{i\neq 0}\frac{\log|a_{i}|}{|i|}<\infty\right\},

with the relation uv=0uv=0. The gluings with nearby maximal cells σ±\sigma_{\pm} of both 𝒢00\prescript{0}{}{\mathcal{G}}^{0} and 𝖦sf00\prescript{0}{}{\mathsf{G}}^{0}_{\mathrm{sf}} are simply given by the parallel transport through vv and the formulae

σ+:{zmsρσ+1(m)zmfor mΛρ,usρσ+1(mρ)zmρ,v0,σ:{zmsρσ1(m)zmfor mΛρ,u0,vsρσ1(mρ)zmρ\sigma_{+}\colon\begin{cases}z^{m}\mapsto s^{-1}_{\rho\sigma_{+}}(m)z^{m}&\text{for $m\in\Lambda_{\rho}$},\\ u\mapsto s^{-1}_{\rho\sigma_{+}}(m_{\rho})z^{m_{\rho}},&\\ v\mapsto 0,&\end{cases}\qquad\quad\sigma_{-}\colon\begin{cases}z^{m}\mapsto s^{-1}_{\rho\sigma_{-}}(m)z^{m}&\text{for $m\in\Lambda_{\rho}$},\\ u\mapsto 0,&\\ v\mapsto s^{-1}_{\rho\sigma_{-}}(-m_{\rho})z^{-m_{\rho}}&\end{cases}

in the proof of Lemma 5.10.

Now for the wall-crossing sheaf 𝒪𝒟,x0𝖦sf,x00\prescript{0}{}{\mathscr{O}}_{\mathscr{D},x}\cong\prescript{0}{}{\mathsf{G}}^{0}_{\mathrm{sf},x}, the wall-crossing factor Θv,ρ\varTheta_{v,\rho} acts trivially except on the two coordinate functions u,vu,v because m,dˇρ=0\langle m,\check{d}_{\rho}\rangle=0 for mΛρm\in\Lambda_{\rho}. The gluing of uu to the nearby maximal cells which obeys wall crossing is given by

u|Uxσ±:={u|Uxσ+on Uxσ+,Θv,ρ,x1(u)|Uxσ=0on Uxσ,u|_{U_{x}\cap\sigma_{\pm}}:=\begin{dcases}u|_{U_{x}\cap\sigma_{+}}&\text{on $U_{x}\cap\sigma_{+}$,}\\ \varTheta^{-1}_{v,\rho,x}(u)|_{U_{x}\cap\sigma_{-}}=0&\text{on $U_{x}\cap\sigma_{-}$,}\end{dcases}

in a sufficiently small neighborhood UxU_{x} of xx. Here, the reason that we have Θv,ρ,x1(u)|Uxσ=0\varTheta^{-1}_{v,\rho,x}(u)|_{U_{x}\cap\sigma_{-}}=0 on UxσU_{x}\cap\sigma_{-} is simply because we have u0u\mapsto 0 in the gluing of 𝖦sf00\prescript{0}{}{\mathsf{G}}^{0}_{\mathrm{sf}}. For the same reason, we see that the gluing of vv agrees with that of 𝒢00\prescript{0}{}{\mathcal{G}}^{0} and 𝖦sf00\prescript{0}{}{\mathsf{G}}^{0}_{\mathrm{sf}}. ∎

Definition 5.21.

A (kthk^{\text{th}}-order) scattering diagram 𝒟\mathscr{D} is said to be consistent if there is an isomorphism 𝔦(𝒪𝒟k)|Wα𝒢α0k\mathfrak{i}_{*}(\prescript{k}{}{\mathscr{O}}_{\mathscr{D}})|_{W_{\alpha}}\cong\prescript{k}{}{\mathcal{G}}^{0}_{\alpha} as sheaves of [q]/(qk+1)\mathbb{C}[q]/(q^{k+1})-algebras on each open subset WαW_{\alpha}.

The above consistency condition would imply that k,l:𝔦(𝒪𝒟k)𝔦(𝒪𝒟l)\prescript{k,l}{}{\flat}\colon\mathfrak{i}_{*}(\prescript{k}{}{\mathscr{O}}_{\mathscr{D}})\rightarrow\mathfrak{i}_{*}(\prescript{l}{}{\mathscr{O}}_{\mathscr{D}}) is surjective for any l<kl<k and hence 𝔦(𝒪𝒟k)\mathfrak{i}_{*}(\prescript{k}{}{\mathscr{O}}_{\mathscr{D}}) is a sheaf of free [q]/(qk+1)\mathbb{C}[q]/(q^{k+1})-modules on BB. We are going to see that 𝔦(𝒪𝒟k)\mathfrak{i}_{*}(\prescript{k}{}{\mathscr{O}}_{\mathscr{D}}) agrees with the push-forward of the sheaf of holomorphic functions on a (kthk^{\text{th}}-order) thickening Xk\prescript{k}{}{X} of the central fiber X0\prescript{0}{}{X} under the modified moment map ν\nu.

Let us elaborate a bit on the relation between this definition of consistency and that in [29]. Assuming we have a consistent scattering diagram in the sense of [29], then we obtain a kthk^{\text{th}}-order thickening Xk\prescript{k}{}{X} of X0\prescript{0}{}{X} which is locally modeled on the thickenings 𝕍αk\prescript{k}{}{\mathbb{V}}_{\alpha}’s by [28, Cor. 2.18]. Pushing forward via the modified moment map ν\nu, we obtain a sheaf of algebras over [q]/(qk+1)\mathbb{C}[q]/(q^{k+1}) lifting 𝒢00\prescript{0}{}{\mathcal{G}}^{0}, which is locally isomorphic to the 𝒢α0k\prescript{k}{}{\mathcal{G}}^{0}_{\alpha}’s. This consequence is exactly what we use to formulate our definition of consistency.

Lemma 5.22.

Suppose we have WWαWβW\subset W_{\alpha}\cap W_{\beta} such that V=ν1(W)V=\nu^{-1}(W) is Stein, and an isomorphism h:𝒢β0k|W𝒢α0k|Wh\colon\prescript{k}{}{\mathcal{G}}^{0}_{\beta}|_{W}\rightarrow\prescript{k}{}{\mathcal{G}}^{0}_{\alpha}|_{W} of sheaves of [q]/(qk+1)\mathbb{C}[q]/(q^{k+1})-algebras which is the identity modulo (q)(q). Then there is a unique isomorphism ψ:𝕍αk|V𝕍βk|V\psi\colon\prescript{k}{}{\mathbb{V}}_{\alpha}|_{V}\rightarrow\prescript{k}{}{\mathbb{V}}_{\beta}|_{V} of analytic spaces inducing hh.

Proof.

From the description in §2.4, we can embed both families 𝕍αk\prescript{k}{}{\mathbb{V}}_{\alpha}, 𝕍βk\prescript{k}{}{\mathbb{V}}_{\beta} over Specan([q]/(qk+1))\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[q]/(q^{k+1})) as closed analytic subschemes of N+1=N×q\mathbb{C}^{N+1}=\mathbb{C}^{N}\times\mathbb{C}_{q} and L+1=L×q\mathbb{C}^{L+1}=\mathbb{C}^{L}\times\mathbb{C}_{q} respectively, where projection to the second factor defines the family over [q]/(qk+1)\mathbb{C}[q]/(q^{k+1}). Let 𝒥α\mathcal{J}_{\alpha} and 𝒥β\mathcal{J}_{\beta} be the corresponding ideal sheaves, which can be generated by finitely many elements. We can take Stein open subsets UαN+1U_{\alpha}\subseteq\mathbb{C}^{N+1} and UβL+1U_{\beta}\subseteq\mathbb{C}^{L+1} such that their intersections with the subschemes give 𝕍αk|V\prescript{k}{}{\mathbb{V}}_{\alpha}|_{V} and 𝕍βk|V\prescript{k}{}{\mathbb{V}}_{\beta}|_{V} respectively. By taking global sections of the sheaves over WW, we obtain the isomorphism h:𝒪𝕍βk(V)𝒪𝕍αk(V)h\colon\mathcal{O}_{\prescript{k}{}{\mathbb{V}}_{\beta}}(V)\rightarrow\mathcal{O}_{\prescript{k}{}{\mathbb{V}}_{\alpha}}(V). Using the fact that UαU_{\alpha} is Stein, we can lift h(zi)h(z_{i})’s, where ziz_{i}’s are restrictions of coordinate functions to 𝕍βk|VUβ\prescript{k}{}{\mathbb{V}}_{\beta}|_{V}\subset U_{\beta}, to holomorphic functions on UαU_{\alpha}. In this way, hh can be lifted as a holomorphic map ψ:UαUβ\psi\colon U_{\alpha}\rightarrow U_{\beta}. Restricting to 𝕍αk|V\prescript{k}{}{\mathbb{V}}_{\alpha}|_{V}, we see that the image lies in 𝕍βk|V\prescript{k}{}{\mathbb{V}}_{\beta}|_{V}, and hence we obtain the isomorphism ψ\psi. The uniqueness follows from the fact the ψ\psi is determined by ψ(zi)=h(zi)\psi^{*}(z_{i})=h(z_{i}). ∎

Given a consistent scattering diagram 𝒟\mathscr{D} (in the sense of Definition 5.21), the sheaf 𝔦(𝒪𝒟k)\mathfrak{i}_{*}(\prescript{k}{}{\mathscr{O}}_{\mathscr{D}}) can be treated as a gluing of the local sheaves 𝒢α0k\prescript{k}{}{\mathcal{G}}^{0}_{\alpha}’s. Then from Lemma 5.22, we obtain a gluing of the local models 𝕍αk\prescript{k}{}{\mathbb{V}}_{\alpha}’s yielding a thickening Xk\prescript{k}{}{X} of X0\prescript{0}{}{X}. This justifies Definition 5.21.

5.3.2. Constructing consistent scattering diagrams from Maurer–Cartan solutions

We are finally ready to demonstrate how to construct a consistent scattering diagram 𝒟(φ)\mathscr{D}(\varphi) in the sense of Definition 5.21 from a Maurer–Cartan solution φ=ϕ+tf\varphi=\phi+tf obtained in Theorem 4.18. As in §5.2.4, we obtain a kthk^{\text{th}}-order Maurer–Cartan solution ζ0\zeta_{0} and define its scattered part as ϕs𝖳𝖫sf1k(W0)\phi_{\mathrm{s}}\in\prescript{k}{}{\mathsf{TL}}^{1}_{\mathrm{sf}}(W_{0}). From this, we want to construct a kthk^{\text{th}}-order scattering diagram 𝒟(φ)\mathscr{D}(\varphi).

We take an open cover {Ui}i\{U_{i}\}_{i} by pre-compact convex open subsets of W0W_{0} such that, locally on UiU_{i}, ϕin+ϕs\phi_{\mathrm{in}}+\phi_{\mathrm{s}} can be written as a finite sum

(ϕin+ϕs)|Ui=jαijvij,(\phi_{\mathrm{in}}+\phi_{\mathrm{s}})|_{U_{i}}=\sum_{j}\alpha_{ij}\otimes v_{ij},

where αij𝒯1(Ui)\alpha_{ij}\in\mathscr{T}^{1}(U_{i}) has asymptotic support on a codimension one polyhedral subset PijUiP_{ij}\subset U_{i}, and vij𝔥k(Ui)v_{ij}\in\prescript{k}{}{\mathfrak{h}}(U_{i}). We take a partition of unity {ϱi}i\{\varrho_{i}\}_{i} subordinate to the cover {Ui}i\{U_{i}\}_{i} such that supp(ϱi)\mathrm{supp}(\varrho_{i}) has asymptotic support on a compact subset CiC_{i} of UiU_{i}. As a result, we can write

(5.10) ϕin+ϕs=ij(ϱiαij)vij,\phi_{\mathrm{in}}+\phi_{\mathrm{s}}=\sum_{i}\sum_{j}(\varrho_{i}\alpha_{ij})\otimes v_{ij},

where each (ϱiαij)(\varrho_{i}\alpha_{ij}) has asymptotic support on the compact codimension one subset CiPijUiC_{i}\cap P_{ij}\subset U_{i}. The subset ijCiPij\bigcup_{ij}C_{i}\cap P_{ij} will be the support |𝒟||\mathscr{D}| of our scattering diagram 𝒟=𝒟(φ)\mathscr{D}=\mathscr{D}(\varphi).

We may equip |𝒟|:=ijCiPij|\mathscr{D}|:=\bigcup_{ij}C_{i}\cap P_{ij} with a polyhedral decomposition such that all the boundaries and mutual intersections of CiPijC_{i}\cap P_{ij}’s are contained in (n2)(n-2)-dimensional strata of |𝒟||\mathscr{D}|. So, for each (n1)(n-1)-dimensional cell τ\tau of |𝒟||\mathscr{D}|, if intre(τ)(CiPij)\mathrm{int}_{\mathrm{re}}(\tau)\cap(C_{i}\cap P_{ij})\neq\emptyset for some i,ji,j, then we must have τCiPij\tau\subset C_{i}\cap P_{ij}. Let 𝙸(τ):={(i,j)|τCiPij}\mathtt{I}(\tau):=\{(i,j)\ |\ \tau\subset C_{i}\cap P_{ij}\}, which is a finite set of indices. We will equip the (n1)(n-1)-cells τ\tau’s of |𝒟||\mathscr{D}| with the structure of walls or slabs.

We first consider the case of a wall. Take τ|𝒟|[n1]\tau\in|\mathscr{D}|^{[n-1]} such that intre(τ)intre(ρ)=\mathrm{int}_{\mathrm{re}}(\tau)\cap\mathrm{int}_{\mathrm{re}}(\rho)=\emptyset for all ρ\rho with ρ𝒮e\rho\cap\mathscr{S}_{e}\neq\emptyset. We let 𝐰=τ\mathbf{w}=\tau, choose a primitive normal dˇ𝐰\check{d}_{\mathbf{w}} of τ\tau, and give the labels 𝒞±\mathcal{C}_{\pm} to the two adjacent chambers 𝒞±\mathcal{C}_{\pm} so that dˇ𝐰\check{d}_{\mathbf{w}} is pointing into 𝒞+\mathcal{C}_{+}. In a sufficiently small neighborhood UτU_{\tau} of intre(τ)\mathrm{int}_{\mathrm{re}}(\tau), we have ϕin|Uτ=0\phi_{\mathrm{in}}|_{U_{\tau}}=0 and we may write

ϕs|Uτ=(i,j)𝙸(τ)(ϱiαij)vij,\phi_{\mathrm{s}}|_{U_{\tau}}=\sum_{(i,j)\in\mathtt{I}(\tau)}(\varrho_{i}\alpha_{ij})\otimes v_{ij},

where each (ϱiαij)(\varrho_{i}\alpha_{ij}) has asymptotic support on intre(τ)\mathrm{int}_{\mathrm{re}}(\tau). Since locally on UτU_{\tau} any Maurer–Cartan solution is gauge equivalent to 0, there exists an element θτ𝒯0(Uτ)q𝔥k(Uτ)\theta_{\tau}\in\mathscr{T}^{0}(U_{\tau})\otimes q\cdot\prescript{k}{}{\mathfrak{h}}(U_{\tau}) such that

e[θτ,]¯e[θτ,]=¯+[ϕs,].e^{[\theta_{\tau},\cdot]}\circ\bar{\partial}_{\circ}\circ e^{-[\theta_{\tau},\cdot]}=\bar{\partial}_{\circ}+[\phi_{\mathrm{s}},\cdot].

Such an element can be constructed inductively using the procedure in [37, §3.4.3], and can be chosen to be of the form

(5.11) θτ|Uτ𝒞±={θτ,0|Uτ𝒞+on Uτ𝒞+,0on Uτ𝒞,\theta_{\tau}|_{U_{\tau}\cap\mathcal{C}_{\pm}}=\begin{dcases}\theta_{\tau,0}|_{U_{\tau}\cap\mathcal{C}_{+}}&\text{on $U_{\tau}\cap\mathcal{C}_{+}$,}\\ 0&\text{on $U_{\tau}\cap\mathcal{C}_{-}$,}\end{dcases}

for some θτ,0q𝔥k(Uτ)\theta_{\tau,0}\in q\cdot\prescript{k}{}{\mathfrak{h}}(U_{\tau}). From this we obtain the wall-crossing factor associated to the wall 𝐰\mathbf{w}

(5.12) Θ𝐰:=e[θτ,0,].\Theta_{\mathbf{w}}:=e^{[\theta_{\tau,0},\cdot]}.
Remark 5.23.

Here we need to apply the procedure in [37, §3.4.3], which is a generalization of that in [7], because of the potential non-commutativity: [vij,vij]0[v_{ij},v_{ij^{\prime}}]\neq 0 for jjj\neq j^{\prime}.

For the case where τintre(ρ)v\tau\subset\mathrm{int}_{\mathrm{re}}(\rho)_{v} for some ρ\rho with ρ𝒮e\rho\cap\mathscr{S}_{e}\neq\emptyset, we will define a slab. We take UτU_{\tau} and 𝙸(τ)\mathtt{I}(\tau) as above, and let the slab 𝐛=τ\mathbf{b}=\tau. The primitive normal dˇρ\check{d}_{\rho} is the one we chose earlier for each ρ\rho. Again we work in a small neighborhood UτU_{\tau} of intre(τ)\mathrm{int}_{\mathrm{re}}(\tau) with two adjacent chambers 𝒞±\mathcal{C}_{\pm}. As in the proof of Lemma 5.10, we can find a step-function-like element θv,ρ\theta_{v,\rho} of the form

θv,ρ={log(svρ1(fv,ρ))dˇρon Uτ𝒞+,0on Uτ𝒞\theta_{v,\rho}=\begin{dcases}\log(s_{v\rho}^{-1}(f_{v,\rho}))\partial_{\check{d}_{\rho}}&\text{on $U_{\tau}\cap\mathcal{C}_{+}$},\\ 0&\text{on $U_{\tau}\cap\mathcal{C}_{-}$}\end{dcases}

to solve the equation e[θv,ρ,]¯e[θv,ρ,]=¯+[ϕin,]e^{[\theta_{v,\rho},\cdot]}\circ\bar{\partial}_{\circ}\circ e^{-[\theta_{v,\rho},\cdot]}=\bar{\partial}_{\circ}+[\phi_{\mathrm{in}},\cdot] on UτU_{\tau}. In other words,

Ψ:=e[θv,ρ,]:(𝖳𝖫sfk|Uτ,¯sf)(𝖳𝖫sfk|Uτ,¯)\varPsi:=e^{-[\theta_{v,\rho},\cdot]}\colon(\prescript{k}{}{\mathsf{TL}}^{*}_{\mathrm{sf}}|_{U_{\tau}},\bar{\partial}_{\mathrm{sf}})\rightarrow(\prescript{k}{}{\mathsf{TL}}^{*}_{\mathrm{sf}}|_{U_{\tau}},\bar{\partial}_{\circ})

is an isomorphism of sheaves of dgLas. Computations using the formula in [8, Lem. 2.5] then gives the identity

Ψ1(¯+[Ψ(ϕs),])Ψ=¯+[ϕin+ϕs,].\varPsi^{-1}(\bar{\partial}_{\circ}+[\varPsi(\phi_{\mathrm{s}}),\cdot])\circ\varPsi=\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot].

Once again, we can find an element θτ\theta_{\tau} such that

e[θτ,]¯e[θτ,]=¯+[Ψ(ϕs),],e^{[\theta_{\tau},\cdot]}\circ\bar{\partial}_{\circ}\circ e^{-[\theta_{\tau},\cdot]}=\bar{\partial}_{\circ}+[\varPsi(\phi_{\mathrm{s}}),\cdot],

and hence a corresponding element θτ,0q𝔥k(Uτ)\theta_{\tau,0}\in q\cdot\prescript{k}{}{\mathfrak{h}}(U_{\tau}) of the form (5.11). From this we get

(5.13) Ξ𝐛:=e[θτ,0,]\varXi_{\mathbf{b}}:=e^{[\theta_{\tau,0},\cdot]}

and hence the wall-crossing factor Θ𝐛:=Θv,ρΞ𝐛\Theta_{\mathbf{b}}:=\varTheta_{v,\rho}\circ\varXi_{\mathbf{b}} associated to the slab 𝐛\mathbf{b}.

Next we would like to argue that consistency of the scattering diagram 𝒟\mathscr{D} follows from the fact that ϕ\phi is a Maurer–Cartan solution. First of all, on the global sheaf PkV,\prescript{k}{}{PV}^{*,*} over BB, we have the operator ¯ϕ:=¯+[ϕ,]\bar{\partial}_{\phi}:=\bar{\partial}+[\phi,\cdot] which satisfies [Δ,¯ϕ]=0[\prescript{}{}{\Delta},\bar{\partial}_{\phi}]=0 and ¯ϕ2=0\bar{\partial}_{\phi}^{2}=0. This allows us to define the sheaf of kthk^{\text{th}}-order holomorphic functions as

𝒪ϕk:=Ker(¯ϕ)PkV0,0,\prescript{k}{}{\mathcal{O}}_{\phi}:=\mathrm{Ker}(\bar{\partial}_{\phi})\subset\prescript{k}{}{PV}^{0,0},

for each kk\in\mathbb{N}. It is a sequence of sheaves of commutative [q]/(qk+1)\mathbb{C}[q]/(q^{k+1})-algebras over BB, equipped with a natural map k,l:𝒪ϕk𝒪ϕl\prescript{k,l}{}{\flat}\colon\prescript{k}{}{\mathcal{O}}_{\phi}\rightarrow\prescript{l}{}{\mathcal{O}}_{\phi} for l<kl<k that is induced from the maps for PkV,\prescript{k}{}{PV}^{*,*}. By construction, we see that 𝒪ϕ0𝒢00ν(𝒪X0)\prescript{0}{}{\mathcal{O}}_{\phi}\cong\prescript{0}{}{\mathcal{G}}^{0}\cong\nu_{*}(\mathcal{O}_{\prescript{0}{}{X}}).

We claim that the maps k,l\prescript{k,l}{}{\flat}’s are surjective. To prove this, we fix a point xBx\in B and take an open chart WαW_{\alpha} containing xx in the cover of BB we chose at the beginning of §5.2.4. There is an isomorphism Φα:PkV,|WαPkVα,\varPhi_{\alpha}\colon\prescript{k}{}{PV}^{*,*}|_{W_{\alpha}}\cong\prescript{k}{}{PV}^{*,*}_{\alpha} identifying the differential ¯\bar{\partial} with ¯α+[ηα,]\bar{\partial}_{\alpha}+[\eta_{\alpha},\cdot] by our construction. Write ϕα=Φα(ϕ)\phi_{\alpha}=\varPhi_{\alpha}(\phi) and notice that ¯α+[ηα+ϕα,]\bar{\partial}_{\alpha}+[\eta_{\alpha}+\phi_{\alpha},\cdot] squares to zero, which means that ηα+ϕα\eta_{\alpha}+\phi_{\alpha} is a solution to the Maurer–Cartan equation for PkVα,(Wα)\prescript{k}{}{PV}^{*,*}_{\alpha}(W_{\alpha}). We apply the same trick as above to the local open subset WαW_{\alpha}, namely, any Maurer–Cartan solution lying in PkVα1,1(Wα)\prescript{k}{}{PV}^{-1,1}_{\alpha}(W_{\alpha}) is gauge equivalent to the trivial one, so there exists θαPkVα1,0(Wα)\theta_{\alpha}\in\prescript{k}{}{PV}^{-1,0}_{\alpha}(W_{\alpha}) such that

e[θα,]¯αe[θα,]=¯α+[ηα+ϕα,].e^{[\theta_{\alpha},\cdot]}\circ\bar{\partial}_{\alpha}\circ e^{-[\theta_{\alpha},\cdot]}=\bar{\partial}_{\alpha}+[\eta_{\alpha}+\phi_{\alpha},\cdot].

As a result, the map e[θα,]Φα:(PkV,|Wα,¯+[ϕ,])(PkVα,,¯α)e^{-[\theta_{\alpha},\cdot]}\circ\varPhi_{\alpha}\colon(\prescript{k}{}{PV}^{*,*}|_{W_{\alpha}},\bar{\partial}+[\phi,\cdot])\cong(\prescript{k}{}{PV}^{*,*}_{\alpha},\bar{\partial}_{\alpha}) is an isomorphism of dgLas, sending 𝒪ϕk\prescript{k}{}{\mathcal{O}}_{\phi} isomorphically onto 𝒢α0k\prescript{k}{}{\mathcal{G}}_{\alpha}^{0}.

We shall now prove the consistency of the scattering diagram 𝒟=𝒟(φ)\mathscr{D}=\mathscr{D}(\varphi) by identifying the associated wall-crossing sheaf 𝒪𝒟k\prescript{k}{}{\mathscr{O}}_{\mathscr{D}} with the sheaf 𝒪ϕk|W0(𝒟)\prescript{k}{}{\mathcal{O}}_{\phi}|_{W_{0}(\mathscr{D})} of kthk^{\text{th}}-order holomorphic functions.

Theorem 5.24.

There is an isomorphism Φ:𝒪ϕk|W0(𝒟)𝒪𝒟k\Phi\colon\prescript{k}{}{\mathcal{O}}_{\phi}|_{W_{0}(\mathscr{D})}\rightarrow\prescript{k}{}{\mathscr{O}}_{\mathscr{D}} of sheaves of [q]/(qk+1)\mathbb{C}[q]/(q^{k+1})-algebras on W0(𝒟)W_{0}(\mathscr{D}). Furthermore, the scattering diagram 𝒟=𝒟(φ)\mathscr{D}=\mathscr{D}(\varphi) associated to the Maurer–Cartan solution ϕ\phi is consistent in the sense of Definition 5.21.

Proof.

To prove the first statement, we first notice that there is a natural isomorphism

𝒪ϕk|W0|𝒟|𝒪𝒟k|W0|𝒟|,\prescript{k}{}{\mathcal{O}}_{\phi}|_{W_{0}\setminus|\mathscr{D}|}\cong\prescript{k}{}{\mathscr{O}}_{\mathscr{D}}|_{W_{0}\setminus|\mathscr{D}|},

so we only need to consider those points xintre(τ)x\in\mathrm{int}_{\mathrm{re}}(\tau) where τ\tau is either a wall or a slab. Since W0(𝒟)W0W_{0}(\mathscr{D})\subset W_{0}, we will work on the semi-flat locus W0W_{0} and use the model 𝖯𝖵sf,k\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}, which is equipped with the operator ¯+[ϕin+ϕs,]\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot]. Via the isomorphism

Φ:(PkV0,,¯ϕ)(𝖯𝖵sf,k,¯+[ϕin+ϕs,])\varPhi\colon(\prescript{k}{}{PV}^{*,*}_{0},\bar{\partial}_{\phi})\rightarrow(\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}},\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot])

from Lemma 5.10, we may write

𝒪ϕk|W0=Ker(¯ϕ)𝖯𝖵sf0,0k.\prescript{k}{}{\mathcal{O}}_{\phi}|_{W_{0}}=\mathrm{Ker}(\bar{\partial}_{\phi})\subset\prescript{k}{}{\mathsf{PV}}^{0,0}_{\mathrm{sf}}.

We fix a point xW0(𝒟)|𝒟|x\in W_{0}(\mathscr{D})\cap|\mathscr{D}| and consider the stalk at xx for both sheaves. In the above construction of walls and slabs from the Maurer–Cartan solution ϕ\phi, we first take a sufficiently small open subset UxU_{x} and then find a gauge transformation of the form Ψ=e[θτ,]\varPsi=e^{[\theta_{\tau},\cdot]} in the case of a wall, and of the form Ψ=e[θv,ρ,]e[θτ,]\varPsi=e^{[\theta_{v,\rho},\cdot]}\circ e^{[\theta_{\tau},\cdot]} in the case of a slab. We have

Ψ¯Ψ1=¯+[ϕin+ϕs,]\varPsi\circ\bar{\partial}_{\circ}\circ\varPsi^{-1}=\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot]

by construction, so this further induces an isomorphism

Ψ:𝖦sf0k|Ux𝒪ϕk|Ux\varPsi\colon\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf}}|_{U_{x}}\rightarrow\prescript{k}{}{\mathcal{O}}_{\phi}|_{U_{x}}

of [q]/(qk+1)\mathbb{C}[q]/(q^{k+1})-algebras.

It remains to see how the stalk Ψ:𝖦sf,x0k𝒪ϕ,xk\varPsi\colon\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf},x}\rightarrow\prescript{k}{}{\mathcal{O}}_{\phi,x} is glued to nearby chambers 𝒞±\mathcal{C}_{\pm}. For this purpose, we let

Ψ0:=e[θτ,0,]\Psi_{0}:=e^{[\theta_{\tau,0},\cdot]}

as in equation (5.12) in the case of a wall, and

Ψ0:=Θv,ρe[θτ,0,]\Psi_{0}:=\varTheta_{v,\rho}\circ e^{[\theta_{\tau,0},\cdot]}

as in (5.13) in the case of a slab. Then, the restriction of an element f𝖦sf,x0kf\in\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf},x} to a nearby chamber is given by

Ψ(f)={Ψ0(f)on Ux𝒞+,fon Ux𝒞\varPsi(f)=\begin{dcases}\Psi_{0}(f)&\text{on $U_{x}\cap\mathcal{C}_{+}$},\\ f&\text{on $U_{x}\cap\mathcal{C}_{-}$}\end{dcases}

in a sufficiently small neighborhood UxU_{x}. This agrees with the description of the wall-crossing sheaf 𝒪𝒟,xk\prescript{k}{}{\mathscr{O}}_{\mathscr{D},x} in equation (5.9). Hence we obtain an isomorphism 𝒪ϕk|W0(𝒟)𝒪𝒟k\prescript{k}{}{\mathcal{O}}_{\phi}|_{W_{0}(\mathscr{D})}\cong\prescript{k}{}{\mathscr{O}}_{\mathscr{D}}.

To prove the second statement, we first apply pushing forward via 𝔦:W0(𝒟)B\mathfrak{i}\colon W_{0}(\mathscr{D})\rightarrow B to the first statement to get the isomorphism

𝔦(𝒪ϕk|W0(𝒟))𝔦(𝒪𝒟k).\mathfrak{i}_{*}(\prescript{k}{}{\mathcal{O}}_{\phi}|_{W_{0}(\mathscr{D})})\cong\mathfrak{i}_{*}(\prescript{k}{}{\mathscr{O}}_{\mathscr{D}}).

Now, by the discussion right before this proof, we may identify 𝒪ϕk\prescript{k}{}{\mathcal{O}}_{\phi} with 𝒢α0k\prescript{k}{}{\mathcal{G}}^{0}_{\alpha} locally. But the sheaf 𝒢α0k\prescript{k}{}{\mathcal{G}}^{0}_{\alpha}, which is isomorphic to the restriction of 𝒢00[q]/(qk+1)\prescript{0}{}{\mathcal{G}}^{0}\otimes_{\mathbb{C}}\mathbb{C}[q]/(q^{k+1}) to WαW_{\alpha} as sheaves of [q]/(qk+1)\mathbb{C}[q]/(q^{k+1})-modules, satisfies the Hartogs extension property from W0(𝒟)WαW_{0}(\mathscr{D})\cap W_{\alpha} to WαW_{\alpha} by Lemma 5.19. So we have 𝔦(𝒪ϕk|W0(𝒟))𝒪ϕk\mathfrak{i}_{*}(\prescript{k}{}{\mathcal{O}}_{\phi}|_{W_{0}(\mathscr{D})})\cong\prescript{k}{}{\mathcal{O}}_{\phi}. Hence, we obtain

𝔦(𝒪𝒟k)|Wα(𝒪ϕk)|Wα𝒢α0k,\mathfrak{i}_{*}(\prescript{k}{}{\mathscr{O}}_{\mathscr{D}})|_{W_{\alpha}}\cong(\prescript{k}{}{\mathcal{O}}_{\phi})|_{W_{\alpha}}\cong\prescript{k}{}{\mathcal{G}}^{0}_{\alpha},

from which follows the consistency of the diagram 𝒟=𝒟(φ)\mathscr{D}=\mathscr{D}(\varphi). ∎

Remark 5.25.

From the proof of Theorem 5.24, we actually have a correspondence between step-function-like elements in the gauge group and elements in the tropical vertex group as follows. We fix a generic point xx in a joint 𝔧\mathfrak{j}, and consider a neighborhood of xx of the form Ux×DxU_{x}\times D_{x}, where UxU_{x} is a neighborhood of xx in intre(𝔧)\mathrm{int}_{\mathrm{re}}(\mathfrak{j}) and DxD_{x} is a disk in the normal direction of 𝔧\mathfrak{j}. We pick a compact annulus AxDxA_{x}\subset D_{x} surrounding xx, intersecting finitely many walls/slabs. We let τ1,,τs\tau_{1},\dots,\tau_{s} be the walls/slabs in anti-clockwise direction. For each τi\tau_{i}, we take an open subset 𝒲i\mathscr{W}_{i} just containing the wall τi\tau_{i} such that 𝒲iτi=𝒲i,+𝒲i,\mathscr{W}_{i}\setminus\tau_{i}=\mathscr{W}_{i,+}\cup\mathscr{W}_{i,-}. The following Figure 9 below illustrates the situation.

As in the proof of Theorem 5.24, there is a gauge transformation on each 𝒲i\mathscr{W}_{i} of the form

Ψi:(𝖯𝖵sf,k|𝒲i,¯)(𝖯𝖵sf,k|𝒲i,¯+[ϕin+ϕs,]),\varPsi_{i}\colon(\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i}},\bar{\partial}_{\circ})\rightarrow(\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i}},\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot]),

where Ψi=e[θv,ρ,]e[θτ,]\varPsi_{i}=e^{[\theta_{v,\rho},\cdot]}\circ e^{[\theta_{\tau},\cdot]} for a slab and Ψi=e[θτ,]\varPsi_{i}=e^{[\theta_{\tau},\cdot]} for a wall. These are step-function-like elements in the gauge group satisfying

Ψi={Θion 𝒲i,+,idon 𝒲i,,\varPsi_{i}=\begin{dcases}\Theta_{i}&\text{on $\mathscr{W}_{i,+}$,}\\ \mathrm{id}&\text{on $\mathscr{W}_{i,-}$,}\end{dcases}

where Θi\Theta_{i} is the wall crossing factor associated to τi\tau_{i}.

On the overlap 𝒲i,+=𝒲i𝒲i+1\mathscr{W}_{i,+}=\mathscr{W}_{i}\cap\mathscr{W}_{i+1} (where we set i+1=1i+1=1 if i=si=s), there is a commutative diagram

(𝖯𝖵sf,k|𝒲i,+,¯)\textstyle{(\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i,+}},\bar{\partial}_{\circ})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Θi\scriptstyle{\Theta_{i}}Ψi\scriptstyle{\varPsi_{i}}(𝖯𝖵sf,k|𝒲i,+,¯)\textstyle{(\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i,+}},\bar{\partial}_{\circ})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψi+1\scriptstyle{\varPsi_{i+1}}(𝖯𝖵sf,k|𝒲i,+,¯+[ϕin+ϕs,])\textstyle{(\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i,+}},\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot])\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathrm{id}}(𝖯𝖵sf,k|𝒲i,+,¯+[ϕin+ϕs,])\textstyle{(\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i,+}},\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot])}

allowing us to interpret the wall crossing factor Θi\Theta_{i} as the gluing between the two sheaves 𝖯𝖵sf,k|𝒲i\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i}} and 𝖯𝖵sf,k|𝒲i+1\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i+1}} over 𝒲i,+\mathscr{W}_{i,+}.

Notice that the Maurer–Cartan element ϕ\phi is global. On a small neighborhood WαW_{\alpha} containing Ux×DxU_{x}\times D_{x}, we have the sheaf (PkVα,,¯ϕ)(\prescript{k}{}{PV}_{\alpha}^{*,*},\bar{\partial}_{\phi}) on WαW_{\alpha}, and there is an isomorphism

e[θα,]:(PkVα,,¯α)(PkVα,,¯ϕ).e^{[\theta_{\alpha},\cdot]}\colon(\prescript{k}{}{PV}^{*,*}_{\alpha},\bar{\partial}_{\alpha})\cong(\prescript{k}{}{PV}^{*,*}_{\alpha},\bar{\partial}_{\phi}).

Composing with the isomorphism

(PkVα,|𝒲i,¯ϕ)(𝖯𝖵sf,k|𝒲i,¯+[ϕin+ϕs,]),(\prescript{k}{}{PV}^{*,*}_{\alpha}|_{\mathscr{W}_{i}},\bar{\partial}_{\phi})\cong(\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i}},\bar{\partial}_{\circ}+[\phi_{\mathrm{in}}+\phi_{\mathrm{s}},\cdot]),

we have a commutative diagram of isomorphisms

(𝖯𝖵sf,k|𝒲i,+,¯)Ψi,0(𝖯𝖵sf,k|𝒲i,+,¯)(PkVα,|𝒲i,+,¯α).\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 38.8742pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&\\&&\crcr}}}\ignorespaces{\hbox{\kern-38.8742pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{(\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i,+}},\bar{\partial}_{\circ})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 95.73177pt\raise 6.44008pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.34326pt\hbox{$\scriptstyle{\varPsi_{i,0}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 168.52295pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\kern 74.43823pt\raise-28.59079pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern 102.69858pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 168.52295pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{(\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i,+}},\bar{\partial}_{\circ})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\kern 132.96991pt\raise-28.59079pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}{\hbox{\kern-1.0pt\raise-39.83801pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 62.8742pt\raise-39.83801pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{(\prescript{k}{}{PV}^{*,*}_{\alpha}|_{\mathscr{W}_{i,+}},\bar{\partial}_{\alpha})}$}}}}}}}{\hbox{\kern 206.39716pt\raise-39.83801pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 1.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}\ignorespaces}}}}\ignorespaces.

This is a Čech-type cocycle condition between the sheaves 𝖯𝖵sf,k|𝒲i\prescript{k}{}{\mathsf{PV}}^{*,*}_{\mathrm{sf}}|_{\mathscr{W}_{i}}’s and PkVα,\prescript{k}{}{PV}^{*,*}_{\alpha}, which can be understood as the original consistency condition defined using path-ordered products in [36, 29]. In particular, taking a local holomorphic function in 𝒢α0k(Wα)\prescript{k}{}{\mathcal{G}}^{0}_{\alpha}(W_{\alpha}) and restricting it to Ux×AxU_{x}\times A_{x}, we obtain elements in 𝖦sf0k(𝒲i)\prescript{k}{}{\mathsf{G}}^{0}_{\mathrm{sf}}(\mathscr{W}_{i}) that jump across the walls according to the wall crossing factors Θi\Theta_{i}’s.

Refer to caption
Figure 9. Wall crossing around a joint 𝔧\mathfrak{j}

Appendix A The Hartogs extension property

The following lemma is an application of the Hartogs extension theorem [41].

Lemma A.1.

Consider the analytic space ()k×Specan([Στ])(\mathbb{C}^{*})^{k}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}]) for some τ\tau and an open subset of the form U×VU\times V, where U()kU\subset(\mathbb{C}^{*})^{k} and VV is a neighborhood of the origin oSpecan([Στ])o\in\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\tau}]). Let W:=V(ωVω)W:=V\setminus\big{(}\bigcup_{\omega}V_{\omega}\big{)}, where dim(ω)+2dim(Στ)\dim(\omega)+2\leq\dim(\Sigma_{\tau}) (i.e. WW is the complement of complex codimension 22 orbits in VV). Then the restriction 𝒪(U×V)𝒪(U×W)\mathcal{O}(U\times V)\rightarrow\mathcal{O}(U\times W) is a ring isomorphism.

Proof.

We first consider the case where dim(Στ)2\dim(\Sigma_{\tau})\geq 2 and W=V{0}W=V\setminus\{0\}. We can further assume that Στ\Sigma_{\tau} consists of just one cone σ\sigma, because the holomorphic functions on VV are those on VσV\cap\sigma that agree on the overlaps. So we can write

𝒪(U×W)={mΛσamzm|am𝒪()k(U)},\mathcal{O}(U\times W)=\left\{\sum_{m\in\Lambda_{\sigma}}a_{m}z^{m}\ \Big{|}\ a_{m}\in\mathcal{O}_{(\mathbb{C}^{*})^{k}}(U)\right\},

i.e. as Laurent series converging in WW. We may further assume that WW is a sufficiently small Stein open subset. Take f=mΛσamzm𝒪(U×W)f=\sum_{m\in\Lambda_{\sigma}}a_{m}z^{m}\in\mathcal{O}(U\times W). We have the corresponding holomorphic function mΛσam(u)zm\sum_{m\in\Lambda_{\sigma}}a_{m}(u)z^{m} on WW for each point uUu\in U, which can be extended to VV using the Hartogs extension theorem [41] because {0}\{0\} is a compact subset of VV such that W=V{0}W=V\setminus\{0\} is connected. Therefore, we have am(u)=0a_{m}(u)=0 for mσΛσm\notin\sigma\cap\Lambda_{\sigma} for each uu, and hence f=σΛσamzmf=\sum_{\sigma\cap\Lambda_{\sigma}}a_{m}z^{m} is an element in 𝒪(U×V)\mathcal{O}(U\times V).

For the general case, we use induction on the codimension of ω\omega to show that any holomorphic function can be extended through VωτVτV_{\omega}\setminus\bigcup_{\tau}V_{\tau} with dim(τ)<dim(ω)\dim(\tau)<\dim(\omega). Taking a point xVωτVτx\in V_{\omega}\setminus\bigcup_{\tau}V_{\tau}, a neighborhood of xx can be written as ()l×Specan([Σω])(\mathbb{C}^{*})^{l}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\omega}]). By the induction hypothesis, we know that holomorphic functions can already be extended through ()l×{0}(\mathbb{C}^{*})^{l}\times\{0\}. We conclude that any holomorphic function can be extended through VωτVτV_{\omega}\setminus\bigcup_{\tau}V_{\tau}. ∎

We will make use of the following version of the Hartogs extension theorem, which can be found in e.g. [31, p. 58], to handle extension within codimension one cells ρ\rho’s and maximal cells σ\sigma’s.

Theorem A.2 (Hartogs extension theorem, see e.g. [31]).

Let UnU\subset\mathbb{C}^{n} be a domain with n2n\geq 2, and AUA\subset U such that UAU\setminus A is still a domain. Suppose π(U)π(A)\pi(U)\setminus\pi(A) is a non-empty open subset, and π1(π(x))A\pi^{-1}(\pi(x))\cap A is compact for every xAx\in A, where π:nn1\pi\colon\mathbb{C}^{n}\rightarrow\mathbb{C}^{n-1} is projection along one of the coordinate direction. Then the natural restriction 𝒪(U)𝒪(UA)\mathcal{O}(U)\rightarrow\mathcal{O}(U\setminus A) is an isomorphism.

Proof of Lemma 5.19.

To prove the first statement, we apply Lemma A.1. So we only need to show that, for ρ𝒫[n1]\rho\in\mathscr{P}^{[n-1]}, a holomorphic function ff in Ux𝒮V(ρ)U_{x}\setminus\mathscr{S}\subset V(\rho) can be extended uniquely to UxU_{x}, where UxU_{x} is some neighborhood of xintre(ρ)𝒮x\in\mathrm{int}_{\mathrm{re}}(\rho)\cap\mathscr{S}. Writing V(ρ)=()n1×Specan([Σρ])V(\rho)=(\mathbb{C}^{*})^{n-1}\times\mathrm{Spec}_{\mathrm{an}}(\mathbb{C}[\Sigma_{\rho}]), we may simply prove that this is the case with Σρ\Sigma_{\rho} consisting of a single ray σ\sigma as in the proof of Lemma A.1. Thus we can assume that V(ρ)=()n1×V(\rho)=(\mathbb{C}^{*})^{n-1}\times\mathbb{C} and the open subset Ux=U×VU_{x}=U\times V for some connected UU. We observe that extensions of holomorphic functions from (U𝒮)×V(U\setminus\mathscr{S})\times V to U×VU\times V can be done by covering the former open subset with Hartogs’ figures.

To prove the second statement, we need to further consider extensions through intre(𝔧)\mathrm{int}_{\mathrm{re}}(\mathfrak{j}) for a joint 𝔧\mathfrak{j}. For those joints lying in some codimension one stratum ρ\rho, the argument is similar to the above. So we assume that σ𝔧=σ\sigma_{\mathfrak{j}}=\sigma is a maximal cell. We take a point xintre(𝔧)x\in\mathrm{int}_{\mathrm{re}}(\mathfrak{j}) and work in a sufficiently small neighborhood UU of xx. In this case, we may find a codimension one rational hyperplane ω\omega containing 𝔧\mathfrak{j}, together with a lattice embedding ΛωΛσ\Lambda_{\omega}\hookrightarrow\Lambda_{\sigma} which induces the projection π:()n()n1\pi\colon(\mathbb{C}^{*})^{n}\rightarrow(\mathbb{C}^{*})^{n-1} along one of the coordinate directions. Letting A=ν1(AU)A=\nu^{-1}(A\cap U) and applying Theorem A.2, we obtain extensions of holomorphic functions in UU. ∎

References

  • [1] P. Aspinwall, T. Bridgeland, A. Craw, M. R. Douglas, M. Gross, A. Kapustin, G. W. Moore, G. Segal, B. Szendrői, and P. M. H. Wilson, Dirichlet branes and mirror symmetry, Clay Mathematics Monographs, vol. 4, American Mathematical Society, Providence, RI; Clay Mathematics Institute, Cambridge, MA, 2009.
  • [2] S. Barannikov, Generalized periods and mirror symmetry in dimensions n>3n>3, preprint, arXiv:math/9903124.
  • [3] S. Barannikov and M. Kontsevich, Frobenius manifolds and formality of Lie algebras of polyvector fields, Internat. Math. Res. Notices (1998), no. 4, 201–215.
  • [4] S. Bardwell-Evans, M.-W. Cheung, H. Hong, and Y.-S. Lin, Scattering diagrams from holomorphic discs in log Calabi-Yau surfaces, preprint, arXiv:2110.15234.
  • [5] T. Bridgeland, Scattering diagrams, Hall algebras and stability conditions, Algebr. Geom. 4 (2017), no. 5, 523–561.
  • [6] H. Cartan, Variétés analytiques réelles et variétés analytiques complexes, Bull. Soc. Math. France 85 (1957), no. 19.7, 77.
  • [7] K. Chan, N. C. Leung, and Z. N. Ma, Scattering diagrams from asymptotic analysis on Maurer-Cartan equations, J. Eur. Math. Soc. (JEMS) 24 (2022), no. 3, 773–849.
  • [8] by same author, Geometry of the Maurer-Cartan equation near degenerate Calabi-Yau varieties, J. Differential Geom. 125 (2023), no. 1, 1–84.
  • [9] K. Chan and Z. N. Ma, Tropical counting from asymptotic analysis on Maurer-Cartan equations, Trans. Amer. Math. Soc. 373 (2020), no. 9, 6411–6450.
  • [10] M.-W. Cheung and Y.-S. Lin, Some examples of family Floer mirror, Adv. Theor. Math. Phys., to appear, arXiv:2101.07079.
  • [11] C.-H. Cho, H. Hong, and S.-C. Lau, Gluing localized mirror functors, J. Differential Geom., to appear, arXiv:1810.02045.
  • [12] by same author, Localized mirror functor for Lagrangian immersions, and homological mirror symmetry for a,b,c1\mathbb{P}^{1}_{a,b,c}, J. Differential Geom. 106 (2017), no. 1, 45–126.
  • [13] J.-P. Demailly, Complex analytic and differential geometry, 2012, https://www-fourier.ujf-grenoble.fr/ demailly/manuscripts/agbook.pdf.
  • [14] J. L. Dupont, Simplicial de Rham cohomology and characteristic classes of flat bundles, Topology 15 (1976), no. 3, 233–245.
  • [15] S. Felten, Global logarithmic deformation theory, preprint (2023), arXiv:2310.07949.
  • [16] by same author, Log smooth deformation theory via Gerstenhaber algebras, Manuscripta Math. 167 (2022), no. 1-2, 1–35.
  • [17] S. Felten, M. Filip, and H. Ruddat, Smoothing toroidal crossing spaces, Forum Math. Pi 9 (2021), Paper No. e7, 36.
  • [18] A. Floer, Morse theory for Lagrangian intersections, J. Differential Geom. 28 (1988), no. 3, 513–547.
  • [19] K. Fukaya, Multivalued Morse theory, asymptotic analysis and mirror symmetry, Graphs and patterns in mathematics and theoretical physics, Proc. Sympos. Pure Math., vol. 73, Amer. Math. Soc., Providence, RI, 2005, pp. 205–278.
  • [20] K. Fukaya and Y.-G. Oh, Zero-loop open strings in the cotangent bundle and Morse homotopy, Asian J. Math. 1 (1997), no. 1, 96–180.
  • [21] W. Fulton, Introduction to toric varieties, Annals of Mathematics Studies, vol. 131, Princeton University Press, Princeton, NJ, 1993, The William H. Roever Lectures in Geometry.
  • [22] B. Gammage and V. Shende, Mirror symmetry for very affine hypersurfaces, Acta Math. 229 (2022), no. 2, 287–346.
  • [23] by same author, Homological mirror symmetry at large volume, Tunis. J. Math. 5 (2023), no. 1, 31–71.
  • [24] S. Ganatra, J. Pardon, and V. Shende, Microlocal Morse theory of wrapped Fukaya categories, Ann. of Math., to appear, arXiv:1809.08807.
  • [25] by same author, Covariantly functorial wrapped Floer theory on Liouville sectors, Publ. Math. Inst. Hautes Études Sci. 131 (2020), 73–200.
  • [26] by same author, Sectorial descent for wrapped Fukaya categories, J. Amer. Math. Soc. 37 (2024), no. 2, 499–635.
  • [27] M. Gross and B. Siebert, Mirror symmetry via logarithmic degeneration data. I, J. Differential Geom. 72 (2006), no. 2, 169–338. MR 2213573 (2007b:14087)
  • [28] by same author, Mirror symmetry via logarithmic degeneration data, II, J. Algebraic Geom. 19 (2010), no. 4, 679–780. MR 2669728 (2011m:14066)
  • [29] by same author, From real affine geometry to complex geometry, Ann. of Math. (2) 174 (2011), no. 3, 1301–1428. MR 2846484
  • [30] by same author, An invitation to toric degenerations, Surveys in differential geometry. Volume XVI. Geometry of special holonomy and related topics, Surv. Differ. Geom., vol. 16, Int. Press, Somerville, MA, 2011, pp. 43–78. MR 2893676
  • [31] M. Jarnicki and P. Pflug, First steps in several complex variables: Reinhardt domains, EMS Textbooks in Mathematics, European Mathematical Society (EMS), Zürich, 2008.
  • [32] M. Kashiwara and P. Schapira, Sheaves on manifolds, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 292, Springer-Verlag, Berlin, 1994, With a chapter in French by Christian Houzel, Corrected reprint of the 1990 original.
  • [33] L. Katzarkov, M. Kontsevich, and T. Pantev, Hodge theoretic aspects of mirror symmetry, From Hodge theory to integrability and TQFT tt*-geometry, Proc. Sympos. Pure Math., vol. 78, Amer. Math. Soc., Providence, RI, 2008, pp. 87–174.
  • [34] Y. Kawamata and Y. Namikawa, Logarithmic deformations of normal crossing varieties and smoothing of degenerate Calabi-Yau varieties, Invent. Math. 118 (1994), no. 3, 395–409.
  • [35] M. Kontsevich, Symplectic geometry of homological algebra, preprint, available online.
  • [36] M. Kontsevich and Y. Soibelman, Affine structures and non-Archimedean analytic spaces, The unity of mathematics, Progr. Math., vol. 244, Birkhäuser Boston, Boston, MA, 2006, pp. 321–385.
  • [37] N. C. Leung, Z. N. Ma, and M. B. Young, Refined scattering diagrams and theta functions from asymptotic analysis of Maurer-Cartan equations, Int. Math. Res. Not. IMRN (2021), no. 5, 3389–3437.
  • [38] Y.-S. Lin, Correspondence theorem between holomorphic discs and tropical discs on K3 surfaces, J. Differential Geom. 117 (2021), no. 1, 41–92.
  • [39] G. Mikhalkin, Enumerative tropical algebraic geometry in 2\mathbb{R}^{2}, J. Amer. Math. Soc. 18 (2005), no. 2, 313–377. MR 2137980 (2006b:14097)
  • [40] T. Nishinou and B. Siebert, Toric degenerations of toric varieties and tropical curves, Duke Math. J. 135 (2006), no. 1, 1–51. MR 2259922 (2007h:14083)
  • [41] H. Rossi, Vector fields on analytic spaces, Ann. of Math. (2) 78 (1963), 455–467.
  • [42] H. Ruddat, Log Hodge groups on a toric Calabi-Yau degeneration, Mirror symmetry and tropical geometry, Contemp. Math., vol. 527, Amer. Math. Soc., Providence, RI, 2010, pp. 113–164.
  • [43] H. Ruddat and B. Siebert, Period integrals from wall structures via tropical cycles, canonical coordinates in mirror symmetry and analyticity of toric degenerations, Publ. Math. Inst. Hautes Études Sci. 132 (2020), 1–82.
  • [44] H. Ruddat and I. Zharkov, Topological Strominger-Yau-Zaslow fibrations, in preparation.
  • [45] by same author, Compactifying torus fibrations over integral affine manifolds with singularities, 2019–20 MATRIX annals, MATRIX Book Ser., vol. 4, Springer, Cham, [2021] ©2021, pp. 609–622.
  • [46] P. Seidel, Fukaya categories and deformations, Proceedings of the International Congress of Mathematicians, Vol. II (Beijing, 2002) (Beijing), Higher Ed. Press, 2002, pp. 351–360.
  • [47] A. Strominger, S.-T. Yau, and E. Zaslow, Mirror symmetry is TT-duality, Nuclear Phys. B 479 (1996), no. 1-2, 243–259.
  • [48] J. Terilla, Smoothness theorem for differential BV algebras, J. Topol. 1 (2008), no. 3, 693–702.
  • [49] H. Whitney, Geometric integration theory, Princeton University Press, Princeton, N. J., 1957.