You are on page 1of 34

FIBRATIONS BY LAGRANGIAN TORI

FOR MAXIMAL CALABI–YAU DEGENERATIONS AND BEYOND

JAVIER FERNÁNDEZ DE BOBADILLA AND TOMASZ PEŁKA

Abstract. We introduce a general technique to construct Lagrangian torus fibrations in degenerations


of Kähler manifolds. We show that such torus fibrations naturally occur at the boundary of the
A’Campo space. This space extends a degeneration over a punctured disc to its real oriented blowup,
and is equipped with an exact modification of a given Kähler form which becomes fiberwise symplectic
at the boundary. Therefore, it allows to move Lagrangian tori from the boundary using the symplectic
connection. In the setting of maximal Calabi–Yau degenerations, we prove that our tori have some
arXiv:2312.13248v1 [math.AG] 20 Dec 2023

asymptotic properties expected from the (conjectural) Strominger–Yau–Zaslow fibration.

Contents
1. Introduction 1
2. Preliminaries on the essential skeleton 5
3. The A’Campo space and its C ∞ structures 8
4. Fiberwise Kähler forms 12
5. Lagrangian fibration at radius zero 17
6. Lagrangian fibrations at positive radius 21
7. The Calabi–Yau case 27
References 33

1. Introduction
The Kontsevich–Soibelman conjecture [KS01] predicts existence of certain special Lagrangian torus
fibrations in maximal Calabi–Yau degenerations. In this article, we prove an asymptotic version of
this conjecture. In order to state our main results, we briefly recall the relevant definitions; for more
details see Section 2 or e.g. [AB+ 09, Gro13, Li22b].
A maximal Calabi–Yau degeneration is a complex projective morphism f ◦ : X ◦ −→ D∗ , where D∗
is a punctured disc, such that the fibers Xz := (f ◦ )−1 (z) are Calabi–Yau n-folds, and the monodromy
has Jordan blocks of maximal size n + 1. Applying the semistable reduction theorem [KK+ 73] after

Pchange we can extend f to a model f : X −→
a finite base D such that X is smooth, the central
fiber D = i Di is a reduced snc divisor, and f |X\D = f ◦ . Since each fiber of f ◦ is Calabi–Yau,
there exists a holomorphic section Ω ∈ Γ(X, KX/D ) with zeros only at the central fiber.PMultiplying
Ω by a holomorphic function on the base D∗ , we can assume that the divisor of Ω is i ai Di , with
mini ai = 0. We say that Di is essential if ai = 0. The essential skeleton ∆S is a subcomplex of the
dual complex of D spanned by vertices corresponding to essential divisors. It is a closed n-dimensional
pseudomanifold [NX16], so the complement ∆◦S of the union of its codimension 2 faces is a topological
manifold. This implies that any (n − 1)-dimensional face F n−1 of ∆◦S is the intersection of exactly
two n-dimensional faces F0n , F1n of ∆◦S . In Section 5 we construct the expanded essential skeleton ES
from ∆◦S by replacing any (n − 1)-dimensional face F n−1 by F n−1 × [0, 1] and identifying F n−1 × {i}
with the subset F n−1 ⊆ Fin for each i ∈ {0, 1}; see Figure 3(c) for an example. Such ES is a smooth
manifold with boundary and corners, whose interior is homeomorphic to ∆◦S .

2020 Mathematics Subject Classification. Primary: 14D06; Secondary: 14J32, 32Q25, 53C23, 53D12.
J.F.B. was supported by the Basque Government through the BERC 2022-2025 program, by the Ministry of Science
and Innovation: BCAM Severo Ochoa accreditation CEX2021-001142-S/MICIN/AEI/10.13039/501100011033, and by
the Spanish Ministry of Science, Innovation and Universities, project reference PID2020-114750GB-C33. J.F.B. and T.P.
were supported by the semester “Singularity Theory and Low Dimensional Topology” at Rényi Institute, Budapest.
1
FIBRATIONS BY LAGRANGIAN TORI 2

n(n−1)
Let t := −1/ log |f |. It is known, cf. [KS01, §3] or Section 7.2, that volΩ n ı n
new := t ( 2 ) (−1) 2 Ω∧Ω
restricts to a volume form on each Xz such that the limit of the volume of Xz is finite as z → 0. Let
ωX be any Kähler form in X. Yau’s proof of the Calabi conjecture [Yau77] implies that there exists a
family of potentials φCY CY c CY is Ricci
z : Xz −→ R such that the Calabi–Yau metric ωXz := ωX |Xz + dd φz

flat for all z. Ricci flatness is equivalent to the existence of a function c : D −→ R such that
CY n
(1) 1
n! (ωXz ) = c(z) · volΩ
new |Xz .

The rescaling in the above definition of volΩ


new ensures that limz→0 c(z) exists and is nonzero.
Motivated by Mirror Symmetry, Strominger, Yau and Zaslow [SYZ96] predicted that Calabi–Yau
3-folds should admit ωXCY -Lagrangian fibrations by tori which are special with respect to Ω, cf. [Gro13].
z
A Lagrangian submanifold L is special of phase ̟ ∈ S1 if the imaginary part Im(̟Ω)|L vanishes. It
was realized by several authors, see e.g. [GW00, KS01], that such tori fibrations are an emergent feature
of a maximal Calabi–Yau degeneration, in the sense that they should appear on Xz for |z| sufficiently
small. More precise versions of the problem (e.g. taking into account inevitable singularities of some
tori) were formulated e.g. in [GS03, Joy03, Li20], with strong evidence provided by various constructions
in particular contexts, such as [Rua07, CBM09, RZ21, EM21, MPS23]. We do not attempt to give a
complete list here, but rather refer the reader to surveys [Aur09, Cha14, Li22b] or [AB+ 09, §7].
This article is mostly inspired by the formulation due to Kontsevich and Soibelman [KS01, §3]. In
short, the main aspects of their prediction are as follows; for details see Conjectures 1 and 2 loc. cit.
1. Gromov–Hausdorff limit. On each fiber Xz , consider the metric induced by the Calabi–Yau form
CY , rescaled so that the diameter of X is independent of z. Then the Gromov–Hausdorff limit of X
ωX z z z
as z → 0 is a metric space ∆ containing a smooth affine n-manifold ∆◦ whose complement in ∆ has
Hausdorff codimension at most 2. Moreover, the Riemannian metric on ∆◦ has a potential satisfying
the real Monge-Ampère equation, i.e. it is locally given by a matrix Hess(K) for some function K of
the affine coordinates; such that det Hess(K) is constant.
2. Special Lagrangian fibration.There is δ > 0 such that for any z ∈ D∗δ , there is a subset Xzsm ⊆ Xz
and a fibration Xzsm−→ ∆◦by special Lagrangian tori. Moreover, the pairs (Xz , Xz \ Xzsm ) equipped
with the rescaled Calabi–Yau metric as in (1) converge in the Gromov–Hausdorff sense to (∆, ∆ \ ∆◦ ).
It is expected that as z → 0, the volume of the region Xzsm asymptotically fills the whole volume of Xz
[Li22b, §2.5]. Here, the volume is computed with respect to volΩ new |Xz or equivalently, by Ricci-flatness
1 CY )n .
condition (1), with respect to n! (ωX z

In the non-archimedean setting, Nicaise, Xu and Yu [NXY19] constructed fibrations by tori over
∆◦S ,
i.e. over the complement of the codimension 2 faces of the essential skeleton ∆S . In turn, for the
Fermat family, Li [Li22a] constructed special Lagrangian fibrations over codimension 0 faces of ∆S , cf.
[HJ+ 22]. Given these insights it is possible to expect that the pair (∆, ∆◦ ) predicted by Kontsevich
and Soibelman may coincide with (∆S , ∆◦S ). This is the second source of inspiration for us.
In this article we introduce a general technique to produce fibrations by Lagrangian tori. We
summarize it in Theorem 6.13. In the setting of maximal Calabi–Yau degenerations, this theorem
specifies to Theorem 1.1 below, and with little additional arguments given in Section 7, it implies the
subsequent Theorem 1.2. The conjunction of Theorems 1.1 and 1.2 provides an asymptotic version of
Kontsevich and Soibelman prediction in the general case of maximal Calabi-Yau degenerations.
We denote by Clog := [0, ∞) × S1 the real oriented blowup of C at the origin, and by Dδ,log ⊆ Clog
the preimage of Dδ in Clog .
Theorem 1.1. Let f ◦ : X ◦ −→ C∗ be a maximal Calabi–Yau degeneration admitting a semi-stable
model. There is a model f : X −→ Dδ of f ◦ , with a family of holomorphic volume forms Ω as above,
such that the following holds. Let ωX be any Kähler form on X. There exists a family of potentials
φq : X ◦ −→ R, smoothly depending on a parameter q ∈ [0, 1], such that the family of 2-forms ωq :=
ωX + ddc φq satisfies the following properties.
(a) Kähler potential. For every z ∈ D∗δ and every q ∈ [0, 1], the restriction ωq |Xz is Kähler.
(b) Lagrangian fibration. There is a smooth family of maps lz,q : Xz,q sm −→ E , parametrized by
S
sm is a submanifold of X of codimension zero, with boundary and corners,
q ∈ (0, 1], where each Xz,q z
and each lz,q is a Lagrangian torus fibration with respect to the form ωq .
FIBRATIONS BY LAGRANGIAN TORI 3

(c) Gromov–Hausdorff limit. Choose a smooth path (z, q) : [0, δ) −→ Dδ,log × [0, 1] such that |z(0)| =
q(0) = 0 and |z(h)|, q(h) > 0 for h > 0, see Figure 1(a). For each h > 0 let Xh be the fiber Xz(h)
equipped with a Kähler metric given by ωq(h) , rescaled so that its diameter is independent of h. Let
sm
Xhsm := Xz(h),q(h) be the region introduced in (b). Then the Gromov-Hausdorff limit limh→0 (Xh , Xh \
Xh ) is the essential skeleton (∆S , ∆S \ ∆◦S ), equipped with a positive multiple of the euclidean metric.
sm

The euclidean metric on ∆S is restricted from the dual complex of D, viewed as a subset of some
euclidean space RN . Clearly, it has a potential satisfying the real Monge-Ampère equation for the
natural affine structure of the maximal faces, as required by [KS01, Conjecture 1(c)], see (1) above.
The base ES in Theorem 1.1(b) is the expanded essential skeleton, described in the beginning of the
introduction. It contains, as an open subset, the union of relative interiors of n-dimensional faces ∆S .
We denote this open subset by ∆gen gen sm gen
S , let Xz,q ⊆ Xz,q be the preimage of ∆S through the Lagrangian
gen
torus fibration lz,q from Theorem 1.1(b), and refer to Xz,q as the generic region of the fiber Xz . The
next result asserts that the generic region is large, and the family ωq approximates there the conjectural
behavior of the Ricci-flat forms. For a path as in Theorem 1.1(c) we put Xhgen := Xz(h),q(h)
gen
⊆ Xh .
Theorem 1.2. We keep the notation and assumptions from Theorem 1.1. Fix a smooth path as in
Theorem 1.1(c), and for h > 0 let Xhgen ⊆ Xh be the generic region introduced above. Then as h → 0,
we have the following.
(a) Volume filling. The volume of Xhgen with respect to volΩ
new converges to the total volume of Xh .
(b) Asymptotic Ricci-flatness. Let ch : Xhgen −→ R>0 be a smooth family of functions satisfying
1 n Ω
n! (ωq(h) ) = ch · volnew . Then ch converges to a positive constant.
(c) Lagrangian tori are asymptotically special. Assume that the path γ is special, see Definition 6.8:
this condition holds e.g. if |z(h)| = h for all h > 0, see Figure 1(b). Then there is a phase ̟ ∈ S1
such that for every b ∈ ∆gen −1
S , the fibers Lh := lh (b) ⊆ Xh
gen
of the Lagrangian torus fibration lh from
Theorem 1.1(b) satisfy
Im(̟Ω)|Lh
lim = 0.
h→0 Re(̟Ω)|Lh

In particular, the Lagrangian tori Lh ⊆ Xhgen asymptotically minimize volume in their homology classes,
see Corollary 7.8 for a precise statement.
(d) Maslov class. All the above Lagrangian tori, i.e. all fibers of lh , have trivial Maslov class.
The Maslov class of the Lagrangian torus L is the element of H 1 (L, Z) induced by the phase map
̟ : L −→ S1 , defined so that ̟ · Ω|L is a volume form on L, see [AB+ 09, §3.6.1.2]. Triviality of Maslov
class is important to define grading in Lagrangian Floer homology, see [AB+ 09, §8.3.3].
The specialty assumption in Theorem 1.2(c) is a mild technical condition implying that the path is
not very tangent to the q-axis, see Figure 1(b).

q(h) q(h)

|z(h)| |z(h)|

(a) Path in Theorems 1.1(c) and 1.2(a),(b). (b) Special path in Theorem 1.2(c).

Figure 1. Paths in Theorems 1.1(c) and 1.2.

The main novelty of this article is a general technique that allows to construct fibrations by La-
grangian tori in Kähler degenerations, where fibers do not need to be Calabi–Yau nor compact. Theo-
rems 1.1 and 1.2 above are obtained by combining this technique with special geometric properties of
maximal Calabi–Yau degenerations. Now, we briefly describe the construction.
Let f : X −→ C be a holomorphic map whose fibers have dimension n, and the unique singular
fiber D := f −1 (0) is simple normal crossings, with irreducible components D1 , . . . , DN . Let ωX be a
Kähler form on X. Choose a set of indices S ∈ {1, ..., N } corresponding to components of D which
we call essential. In the maximal Calabi–Yau degeneration case the set S corresponds, as above, to
components of minimal discrepancy, but the construction works for any choice of S. Let ∆D be the
FIBRATIONS BY LAGRANGIAN TORI 4

dual complex of D and let ∆S be the essential skeleton, defined as the subcomplex of ∆D spanned by
vertices corresponding to essential components.
In [FdBP22], we introduced the A’Campo space A of f , which fits into the diagram of smooth
manifolds with boundary and smooth maps
A X
(2) fA π f

Clog C
where fA is a locally trivial fibration, extending the family of smooth fibers over C∗ over the “radius-
zero” circle ∂Clog . The space A is constructed by a “hybrid” construction: it compactifies f : X ◦ −→ C∗
with the “radius-zero” fibration combining the Kato–Nakayama space of the log structure (X, D) and
the dual complex ∆D of F the snc divisor D. By construction, any “radius-zero” fiber Aθ for θ ∈ ∂Clog
splits naturally as Aθ = I A◦I,θ , where the disjoint union is indexed by all faces ∆I of ∆D . Each of
the pieces A◦I,θ admits a concrete geometric description (see Lemma 3.5), and in particular:
(i) If dim ∆I = n (so ∆I is a maximal face) then we have A◦I,θ = ∆I × (S1 )n ,
(ii) If dim ∆I = n − 1 then we have A◦I,θ = ∆I × (XI◦ )log , where (XI◦ )log −→ XI◦ is an (S1 )n−1 -bundle
over a compact Riemann surface with as many punctures as maximal faces contain ∆I .
Furthermore, in [FdBP22] we introduced a closed fiberwise symplectic form ωA as an exact mod-
ification of π ∗ ωX , which allowed to compare the usual symplectic monodromy with the radius-zero
model via the symplectic connection. However, compatibility of ωA with the complex structure was
not proved. In Section 4 we introduce a family of closed 2-forms
(3) ωqε = π ∗ ωX + ε · (q · ω ♯ + (1 − q) · ω ♭ ), q ∈ [0, 1],
which at positive radius are fiberwise equal to ωqε = π ∗ ωX + ddc φεq for some potential φεq , so they are
compatible with the complex structure. Moreover, we prove that for q > 0 they are fiberwise symplectic;
and fiberwise Kähler at positive radius. In the setting of maximal Calabi–Yau degenerations, this
will imply Theorem 1.1(a). The proof of the remaining statements follows the general philosophy
of [FdBP22]: we perform the main constructions at radius zero, where the structure of A and the
symplectic form is very well known and special, and then pull them to positive radius by the symplectic
connection. In the remaining part of the introduction, we highlight the main ideas and point to the
sections of the paper where they are implemented.
For any maximal face ∆I of ∆S , close to the interior of A◦I,θ ⊆ ∂A the form ω0ε is non-degenerate, see
Proposition 4.3(c), and as can be seen from the definition (21) of ω ♭ , the induced metric ω0ε resembles
asymptotically the semi-flat one described in [SYZ96], cf. [Li22b, p. 4]. This is the reason for asymptotic
Ricci-flatness claimed in Theorem 1.2(b). We prove it in Proposition 7.3.
Moreover, tori arising in the product structure (i) become Lagrangian with respect to ωqε for any
q ∈ [0, 1], see Proposition 5.6. Thus they can be pulled to positive radius by the symplectic connection
of ωqε , which results in Lagrangian torus fibration claimed in Theorem 1.1(b), see Proposition 6.3. In
the maximal Calabi–Yau degeneration case, as q approaches 0 faster than the radius (see Figure 1(b))
we prove in Proposition 7.7 that the Lagrangian tori become asymptotically special, as claimed in
Theorem 1.2(c). As a consequence, they have trivial Maslov class, as claimed in Theorem 1.2(d).
For any face ∆I of dimension n − 1, see case (ii) above, we take a proper Morse function hI : XI◦ −→
(0, ∞) and prove that for each c ∈ (0, ∞), the preimage of h−1 (c) in (XI◦ )log is isotropic for ωqε for any
q ∈ [0, 1]. Thus again
F ◦we get a Lagrangian torus fibration, away from the preimage of Crit(hI ).
Define Asm θ = I AI,θ , where the union is indexed by all faces ∆I of ∆S of dimension n and n − 1.
An appropriate choice of Morse functions hI guarantees that the Lagrangian fibrations defined above in
each of the pieces A◦I,θ glue to a Lagrangian fibration l: Asmθ −→ ES admitting finitely many singular
fibers. Here ES is the expanded skeleton, endowed with a structure of an ivy-like manifold, see Section
5.1: it is a smooth manifold with some singularities corresponding to critical points of hI .
In the maximal Calabi–Yau degeneration case we can take a model with XI◦ ∼ = C∗ , see Proposition
2.1(c.i), and we have a canonical choice of Morse functions hI with no critical points, see Remark
5.4. With this choice, the map l becomes an honest Lagrangian fibration over a smooth manifold ES ,
which agrees with the one described at the beginning of the introduction. Pulling back the Lagrangian
fibration l by the symplectic connection associated with ωqε for every q ∈ (0, 1], one gets Theorem 1.1(b),
FIBRATIONS BY LAGRANGIAN TORI 5

see Proposition 6.3. Moreover, just like in the semi-flat model, after a suitable rescaling we see that
the above Lagrangian tori collapse as z → 0, both with respect to the metric, and to the Calabi–Yau
volume. This implies the Gromov–Hausdorff convergence and volume filling, claimed in Theorem 1.1(c)
and 1.2(a). We prove these results in Propositions 6.12 and 7.2, respectively.
Acknowledgments. This work grew during visits of the first author to IMPAN, Warsaw, the second
author to BCAM, Bilbao, and both authors to Rényi Institute, Budapest and CIRM, Marseille. We
are grateful to each of these institutions for hospitality and excellent working conditions. We would
also like to thank Duco van Straten for inspiring conversations at CIRM which motivated this work.

2. Preliminaries on the essential skeleton


In this section, we set up the notation for the remaining part of the article, and, in case of maximal
Calabi–Yau degenerations, choose an appropriate model for which our Lagrangian tori will have no
singularities. The latter is achieved in Proposition 2.1 by applying the Minimal Model Program.
We work in the following setting. Let (X, ωX ) be a Kähler manifold of dimension n + 1, and let
f : X −→ C
be a holomorphic function with unique critical value 0 ∈ C, such that the fiber D := f −1 (0) is snc.
Let D1 , . . . , DN be the irreducible components of D, and write
N
X
(4) D= mi Di .
i=1
Moreover, we choose a subset S ⊆ {1, . . . , N }. In practice, we will either take S to be {1, . . . , N }, or,
in the setting of Theorem 1.1, the set of indices of essential components of D, see formula (7).
For z ∈ C∗ we write Xz = f −1 (z). Since we will be interested only in fibers Xz for |z| small enough,
we will often replace X by the preimage f −1 (Dδ ) for some δ > 0 such that f |X\D : X \ D −→ D∗δ is a
submersion; and keep shrinking δ > 0 whenever needed.
Throughout the article, the adjective fiberwise will always refer to the fibers of f |X\D , or its extension

fA to the A’Campo space, introduced in Section 3. We write ı = −1.
2.1. Dual complex of an snc divisor
PN T
Let D = i=1 mi Di be an snc divisor as above. For I ⊆ {1, . . . , N } we put XI = i∈I Di ,
S ◦ F
XI◦ = XI \ j6∈I Dj . This way, XI = X I and X = I XI◦ is a stratification of X.
P
Let ∆ = {(w1 , . . . , wN ) ∈ RN : i wi = 1} be the standard (N − 1)-dimensional simplex. For a
subset I ⊆ S{1, . . . , N } we put ∆I = {(w1 , . . . , wN ) ∈ ∆ : wi = 0 for i 6= I}. The dual complex of D
is ∆D := {I:XI 6=∅} ∆I ⊆ ∆. We equip it with a metric induced by the standard Riemannian metric
P 2 N
S
i (dwi ) on R . For a given subset S ⊆ {1, . . . , N }, we put ∆S = ∆D ∩ ( I⊆S ∆I ).
Since the divisor D is snc, for every face ∆I of ∆D we have #I = codimX XI = n+1−dim XI 6 n+1.
We say that the face ∆I is maximal if #I = n + 1 (so XI is a point), and submaximal if #I = n (so
XI is a curve). We will be mostly interested in the case when ∆D has at least one maximal face.
P
Each face ∆I , viewed as a subset of the plane { i wi = 1, wi = 0 for i 6∈ I} ⊆ RN is a manifold
with boundary and corners, see [BS73, Appendice] for precise definitions. By Proposition 3.1 loc. cit,
we can and do think of a manifold with boundary and corners as a subset of a usual smooth manifold,
mapped by each local chart to an open subset of Rp × Rm−p >0 for some 0 6 p 6 m. We use the notions
of smooth functions, tangent space etc. inherited from the ambient smooth manifold.
2.2. Maximal Calabi–Yau degenerations and their models
A smooth projective complex manifold X is Calabi–Yau if its canonical divisor KX is trivial, i.e.
there is a nowhere vanishing holomorphic n-form Ω ∈ Ωn,0 (X). Yau’s solution to the Calabi conjecture
[Yau77] asserts that any Kähler class in H 1,1 (X) ∩ H 2 (X, R) admits a unique representative whose
associated metric is Ricci flat, see [GHJ03, §5]. We refer to this representative as the Calabi–Yau form,
and denote it by ωCY . Ricci-flatness of the metric gCY := ωCY ( · , J · ) is equivalent to the fact that
there exists a holomorphic volume form Ω satisfying the complex Monge-Ampère equation
n Ω
n 1
(5) 1
n! · ωCY = vol , where volΩ := 2ı · (−1) 2 n(n−1) · Ω ∧ Ω,
FIBRATIONS BY LAGRANGIAN TORI 6

cf. [GHJ03, formula 11] or [AB+ 09, Proposition 6.25].


Let f ◦ : X ◦ −→ C∗ be a projective morphism whose fibers Xz := (f ◦ )−1 (z) are Calabi–Yau manifolds
of dimension n. We say that f is a Calabi–Yau degeneration if f ◦ is meromorphic at 0, i.e. it admits an
extension to a model f : X −→ C. It is maximal if the monodromy transformation T : H n (Xz ; Q) −→
H n (Xz , Q) satisfies (T − id)n 6= 0, (T − id)n+1 = 0, cf. [KS01, §3.1].
Resolving singularities, we see that every f ◦ admits a model f : X −→ C which is snc, i.e. such
that X is a smooth Kähler manifold, and the special fiber f −1 (0) is an snc divisor. Moreover, every
f ◦ admits, after a finite base change, an snc model which is semistable, i.e. its special fiber is reduced
[KK+ 73, Chapter II]. We note that after a finite base change, a maximal Calabi–Yau degeneration
is still maximal. Indeed, the general fiber remains the same, while the monodromy transformation is
replaced by its iterate, so its Jordan blocks remain of the same size.
We now introduce the essential skeleton ∆S of f ◦ , which agrees with the one used in [MN15, NX16],
see also [NXY19, §2.1]. Let f : X −→ C be an snc model of f ◦ . We write its special fiber as D =
PN
i=1 mi Di , see formula (4), and use notation from Section 2.1. By adjunction, for every z 6= 0 the
smooth fiber Xz satisfies KX |Xz = KXz = 0, so the canonical bundle of X admits a meromorphic
section Θ with no zeros or poles off D. We choose one such Θ, and let νi ∈ Z be an integer such that
Θ has a zero of order νi − 1 along Di . Thus we have a linear equivalence
N
X
(6) KX + Dred = νi Di .
i=1
Note that the coefficients νi depend on the choice of the section Θ. If we choose another section, say
Θ′ , then Θ′ = uf k · Θ for some holomorphic unit u ∈ OX ∗ and integer k ∈ Z, so the corresponding

coefficients are νi′ = νi + kmi . It follows that the set


 
νi νj
(7) S= i: = min
mi j mj

does not depend on Θ. We say that a component Di of D is essential if i ∈ S; a face ∆I of the dual
complex ∆D is essential if I ⊆ S. The essential skeleton ∆S ⊆ ∆D is the union of all essential faces.
In [MN15, §4.7], the essential skeleton ∆S (or rather its homeomorphic image in the Berkovich
space) is called the Kontsevich–Soibelman skeleton Sk(X ◦ , Ω), where
Θ f ·Θ
(8) Ω := =
d log f df
νi
is a chosen section of the vertical canonical bundle. As explained in §4.7.3 loc. cit, the quotient m i
equals wtΩ (x) + 1, where x is the divisorial valuation of C(X ◦ ) corresponding to Di , and wtΩ is
the weight S function introduced in §4.3.5 loc. cit. The essential skeleton is defined in 4.10 loc. cit. as
◦ := ◦ ◦
Sk(X ) Ω Sk(X , Ω), so since KX is Calabi–Yau, this definition agrees with Sk(X , Ω), see §4.10.2
loc. cit. Hence up to a homeomorphism, our essential skeleton ∆S agrees with the one in loc. cit.
Theorem 4.7.5 and Corollary 3.2.4 loc. cit. show that the homeomorphism type of the essential
skeleton ∆S , as well as its piecewise affine structure inherited from ∆D (see §3.2 loc. cit), do not depend
on the snc model X. By [NX16, Theorem 4.1.10], the Calabi–Yau degeneration f ◦ is maximal if and
only if dim ∆S = n. In this case, Theorem 4.2.4(3) loc. cit. asserts that ∆S is a closed pseudomanifold.
We remark that by [MN15, Proposition 4.3.4] the property of being essential depends only on the
corresponding valuation of C(X ◦ ), and not on a particular snc model. More precisely, if a valuation of
C(X ◦ ) corresponds to a component E (resp. E ′ ) of a special fiber of some snc model f (resp. f ′ ), then
E is essential if and only if E ′ is,.
The following proposition is known to experts, cf. e.g. [NXY19, p. 969]. For the reader’s convenience,
we include its proof.
Proposition 2.1. Let f ◦ : X ◦ −→ C∗ be a Calabi–Yau degeneration admitting a semistable snc model.
Then f ◦ admits a (not necessarily semistable) snc model f : X −→ C with the following properties.
(a) Every essential component Di of the special fiber D = f −1 (0) has multiplicity one in D.
(b) The canonical bundle KX admits a section Θ such that, denoting as above by νi − 1 the order of
Θ along a component Di of D, we have νi > 0, with equality if and only if Di is essential.
(c) Let ∆I be a submaximal, essential face of the dual complex ∆D . Then every maximal face of ∆D
containing ∆I is essential, too, and the following hold.
FIBRATIONS BY LAGRANGIAN TORI 7

(i) If f ◦ is a maximal degeneration then (XI , XI◦ ) ∼


= (P1 , C∗ ).
(ii) If f is not a maximal degeneration then XI = XI◦ is an elliptic curve.

Proof. Let fe: X e −→ C be a semi-stable snc model of f ◦ , with special fiber D.


e By [NX16, Theorem
e e e
2.3.6(1)], applying f -MMP as in [HX13] to the pair (X, D) we get a good dlt model f : X −→ C of
f ◦ , with special fiber D being the direct image of D;e in particular connected and reduced. The fact
that f is a good dlt model implies that X is normal and Q-factorial, the pair (X, D) is dlt, and the
Weil divisor KX + D is nef, see [NX16, §2.3]. Since D is numerically trivial, it follows that KX is nef,
P
too. Let D 1 , . . . be the irreducible components of D, so D = i D i .
−1
For z 6= 0 the fiber f (z) is Calabi–Yau, so KX |X\D is trivial. Hence we have a linear equivalence
P
of Weil divisors KX = i ai Di for some integers ai .
We claim that all integers ai are equal. Suppose the contrary, and say that a1 = maxi ai > a2 .
Because D is connected, we can order its components in such a way that D 1 meets D 2 . Choose
a curve C ⊆ D 1 such that C ∩ D 2 6= ∅ and C 6⊆ D i for all i 6= 1. Since KX is nef, we have
P P P
0 6 KX · C = i ai Di · C. Subtracting i a1 D i · C = a1 D · C = 0, we get 0 6 i>1 (ai − a1 )D i · C.
Since ai 6 a1 and C 6⊆ D i , each summand is non-positive, hence they are all zero. By assumption
C · D2 > 0, so a2 = a1 ; a contradiction.
Therefore, we have a linear equivalence KX = a1 D = 0. The fiber D is a trivial Cartier divisor,
so the same holds for KX . In particular, the Weil divisor KX + D is Cartier. Recall also that X is
Q-factorial, so all components of D are Q-Cartier. Now [NXY19, Theorem 4.5] implies that

(9) each one-dimensional stratum of D is contained in the snc locus of X,


T
where we consider the standard stratification of D by intersections i∈I D i , see §1.10 loc. cit.
Let π : (X, D) −→ (X, D) be a log resolution, i.e. a proper morphism which is an isomorphism over
the locus where X is smooth and D is snc; such that X is smooth and D := P π ∗ D is snc. Such a
resolution exists, see [Kol13, Theorem 10.45(2)] and references there. Write D = N i=1 mi Di , and split
P P
{1, . . . , N } = E ⊔ P, where Exc π = i∈E Di and π∗−1 D = i∈P Di . Note that mi = 1 for all i ∈ P.
Recall that KX + D is a Cartier divisor. Hence KX + Dred − π ∗ (KX + D) is a Cartier divisor, too. Its
push-forward on X is trivial since π∗ Dred = D. Thus for some integers νi we have a linear equivalence

X N
X

KX + Dred = π (KX + D) + νi Di = νi Di ,
i∈E i=1

where for the second equality we put νi = 0 for i ∈ P, and use the linear equivalences KX = D = 0.
We now choose a section Θ of KX with zeros of order νi − 1 along each Di , see formula (6).
Since the pair (X, D) is dlt and the resolution π is an isomorphism over its snc locus, we have νi > 0
νi ν
for all i ∈ E, see [NX16, §2.3.1]. Thus P = {i : νi = 0} = {i : m i
= minj mjj }. We conclude that P is
the set (7), i.e. i ∈ P if and only if a component Di of D is essential, cf. [NX16, Theorem 3.3.3]. Now
conditions (a), (b) follow from the fact that (mi , νi ) = (1, 0) for all i ∈ P.
Let now ∆I be a submaximal face of the dual complex of D, so dim XI = 1. Assume that ∆I is
essential, so I ⊆ P. By definition of P, the morphism π is an isomorphism at a general point of XI ,
so dim π(XI ) = dim XI = 1. Now condition (9) implies that π(XI ) is contained in the snc locus of D,
so π is an isomorphism near XI . It follows that for every component Dj of D meeting XI , we have
j ∈ P, hence DjPis essential. This proves the first part of (c) and shows that (KX + D)|XI = 0.
Write BI = ( i6∈I mi Di )|XI , so BI is a divisor on XI supported on XI \ XI◦ . Since each component
Di meeting XI has i ∈ P, so mi = 1, we see that BI is reduced. Applying Lemma 2.2 below to J = ∅,
we get KXI + BI = (KX + D)|XI = 0. It follows that either BI = 0 and XI = XI◦ is an elliptic curve;
or BI 6= 0 and (XI , XI◦ ) ∼= (P1 , C∗ ). Since by [NX16, Theorem 4.2.4(3)], the essential skeleton ∆S is
a pseudomanifold (with boundary) we have dim ∆S = n − 1 in the first case and dim ∆S = n in the
second case, see condition (1) of the definition of pseudomanifold given in §4.1.2 loc. cit. By Theorem
4.1.10 loc. cit., the second case happens if and only if f ◦ is a maximal degeneration. 

In the above proof, we used the following elementary lemma.


FIBRATIONS BY LAGRANGIAN TORI 8

P
Lemma 2.2. Let D be anT snc divisor on a smooth projective
P variety X. Write D = N i=1 mi Di , and for
I ⊆ {1, . . . , N } put XI = i∈I Di , X∅ = X, BI = i6∈I mi Di |XI . Fix two subsets J ⊆ I ⊆ {1, . . . , N },
and assume that mi = 1 for all i ∈ I. Then KXI + BI = (KXJ + BJ )|XI .
Proof. By induction on #I \J, we can assume that
P I = J ⊔{i}. By Padjunction, we have KXI = (KXJ +
Di |XJ )|XI = KXJ |XI + Di |XI . Moreover, BI = j6∈I mj Dj |XI = j6∈J mj Dj |XI − mi Di |XI = BJ |XI −
Di |XI since mi = 1 by assumption. Adding the two equalities we get KXI + BI = (KXJ + BJ )|XI . 
Example 2.3 (Hesse pencil). Consider a family of elliptic curves X = {(x31 + x32 + x33 ) · z = x1 x2 x3 } ⊆
P2 × C, and let f : X −→ C be the projection to the z-coordinate. The restriction f ◦ of f over C∗
is a maximal Calabi–Yau degeneration, and f is its model satisfying all conditions of Proposition 2.1.
Indeed, the special fiber D = f −1 (0) is just a union of three non-concurrent lines D1 , D2 , D3 ⊆ P2 ;
and KX = 0, so each Di is essential. We conclude that the essential skeleton is a triangle.
Fix p ∈ D and let π : X ′ −→ X be a blowup at p, so f ′ := f ◦ π is another model of f ◦ . Put
D = (f ′ )−1 (0), Di′ = π∗−1 Di and D0′ = Exc π, so KX ′ = D0′ . Let m′i , νi′ be as in formulas (4), (6), so

νi′ /m′i = 1 for i ∈ {1, 2, 3} and ν0′ = 2.


Assume first that p is a smooth point of D. Then m′0 = 1, so ν0′ /m′0 = 2 > 1, i.e. D0′ is not essential.
Hence the model f ′ satisfies conditions 2.1(a) and (b), but does not satisfy 2.1(c). Note that the
component Di′ meeting D0′ satisfies (Di′ , (Di′ \ (D ′ − Di′ )) ∼
= (P1 , C∗∗ ), so 2.1(c.i) fails, too. The essential

skeleton ∆S computed using this model is still a triangle, in agreement with [MN15, Theorem 4.7.5].
Assume now that p is a singular point of D. Then m′0 = 2, so ν0′ /m′0 = 1, i.e. all components of D ′
are essential. This model satisfies conditions 2.1(b) and (c), but does not satisfy 2.1(a). Now ∆′S is a
square obtained by subdividing the triangle ∆S , again in agreement with [MN15, Theorem 4.7.5].

3. The A’Campo space and its C ∞ structures


Let f : X −→ C be as in the beginning of Section 2, i.e. f is a holomorphic function whose unique
singular fiber D := f −1 (0) is snc; and f |X\D : X \ D −→ D∗δ is a submersion. In [FdBP22, §3], we
introduced the A’Campo space A of f , which fits into the diagram (2). In particular, it allows to extend
the monodromy of f to “radius zero”, i.e. to the monodromy of the restriction f |∂A : ∂A −→ ∂Clog .
The above radius-zero monodromy has simpler dynamical properties than the natural one at positive
radius, see Proposition 7.2 loc. cit. Most importantly, it is induced by a symplectic connection with
respect to a certain fiberwise symplectic form ωA ε . Our aim is to improve this form so that it becomes

fiberwise Kähler at positive radius.

3.1. Definition of the A’Campo space


P
In this section, we review the construction of the A’Campo space A. Recall that D = N i=1 Di is
the irreducible decomposition of the special fiber D.
We begin with the notion of an adapted chart [FdBP22, Definition 3.1]. For a holomorphic chart
(zi1 , . . . , zin ) : UX −→ Cn+1 of X, we define its associated index set as S = {i ∈ {1, . . . , N } : Di ∩ UX 6=
∅}. We say Q thatmithis chart is adapted to f if for every i ∈ {1, . . . , N } we have UX ∩ Di = {zi = 0} and
f |UX = i∈S zi in case S 6= ∅. We also require UX to be small enough, so that log |zi | < 1 for all
i ∈ S, and each coordinate zi extends to a continuous function on the compact closure U X .
For any i ∈ S we have the following smooth functions on UX \ D, see [FdBP22, §3.1.2]
zi −1
(10) ri = |zi |, θi = , si = log ri , ti = .
ri mi s i
We call ri and θi the radial and angular coordinates of UX . We also define global functions on X \ D
−1 f
(11) t= , g = η(t), θ= ,
log |f | |f |
where η : [0, 1] −→ [0, 1] is given by η(τ ) = (1 − log τ )−1 . The key property of this auxiliary function
is that its inverse is a smooth function whose all derivatives vanish at the origin.
Eventually, for each i ∈ S we put
t
(12) wi = , ui = η(wi ), vi = ti − ui , and σi = t2i + ti u2i .
ti
FIBRATIONS BY LAGRANGIAN TORI 9

The functions ri , wi and vi should be thought of as “natural”, “tropical” and “hybrid” coordinates,
respectively. For more intuition, we refer the reader to [FdBP22, Figures 2 and 3]. The last function
σi plays an auxiliary role.
p
In order to define the A’Campo space as a C 1 -manifold, it is convenient to choose an atlas {UX }p
p
consisting of adapted charts, a subordinate partition of unity {τ }p , and define
X X
ui = τ p upi , v i = τ p vip ,
p p
p
where for every p such that UX ∩ Di 6= ∅, the functions upi , vip
are the ones defined for UX p
by formulas
p p p
(12); and for p such that UX ∩ Di = ∅, we put ui = vi = 0.
We define a topological space Γ as the closure of the graph of (u1 , . . . , uN ) : X \ D −→ RN . The
A’Campo space A is a fibered product over X of Γ and the Kato–Nakayama space Xlog of (X, D), see
diagram (13). This way, A is a topological space equipped with a continuous map π : A −→ X which is
a homeomorphism over X \ D; such that for any adapted chart UX ⊆ X, all basic functions (10)–(12)
extend to continuous functions on the preimage π −1 (UX ) [FdBP22, Lemma 3.6(a)]. Crucially for us,
the value of each wi , ui on π −1 (UX ∩ D) does not depend on the choice of the adapted chart UX , see
Lemma 3.6(d),(e) loc. cit.

A Γ := graph(µ) X × RN
p
π
(g,θ)∼fA µ
(13) Xlog X RN
(u1 ,...,uN )
flog f

Clog C
F F
The stratification X = I XI◦ lifts to a decomposition A = A◦I , where A◦I := π −1 (XI◦ ).
The C 1 -atlas on A is introduced as follows. For an adapted chart UX with associated index set
1
S, we cover its preimage by open sets Ui = {wi > n+2 }, i ∈ S, which will be used as domains of
smooth charts on A. To make the notation more compact, we put ϑ := ((θi )i∈S , (zj )j∈{i1 ,...,in }\S ). For
k ∈ {1, . . . , n + 1} we write Qk,n+1 = [0, ∞) × Rk−1 × (S1 )k × Cn+1−k . Now, we define the C 1 -charts as
(14) ψi = (g, (vj )j∈S\{i} ; ϑ) : Ui −→ Qk,n+1 .
Results of [FdBP22, §3.3] show that these charts induce a C 1 -structure on the A’Campo space, which
p
is independent of the choice of {(UX , τ p )}p . The C ∞ -charts are defined in §3.4 loc. cit. as
(15) ψ i = (g, (v j )j∈S\{i} ; ϑ) : Ui −→ Qk,n+1 .
These charts upgrade the above C 1 -structure to a C ∞ one, which does depend on the choice of the
p
adapted atlas {UX }p and partition of unity {τ p }p , see Remark 3.32 loc. cit.
3.2. Distance functions
Our next goal is to introduce another C ∞ -atlas on A, compatible with the C 1 -structure given by
charts (14), which will allow us to introduce our fiberwise Kähler form.
We work in a fixed open set WX ⊆ X such that the restriction f |W X is proper, has connected fibers;
and f |W X \D : W X \ D −→ D∗δ is a submersion. If f is an snc model of a Calabi–Yau degeneration,
which is the case of most interest for us, we simply take WX = X. For another example, let X be an
embedded resolution of an isolated hypersurface singularity, and take for WX the preimage of some
neighborhood of the origin. We can and do assume that δ < exp(− maxi mi ).
p
Now, we introduce global distance functions rbi . To this end, we fix an atlas of adapted charts {UX }p ,
p p p
and denote by S = {i : UX ∩ Di 6= ∅} the associated index set of UX . We define open sets
[ p
[
(16) Ri = UX , and Ri◦ = Ri \ Rj ,
{p : i∈S p } j6=i

so Ri is a neighborhood of Di , and Ri◦


consists of points “far away” from the remaining part of D.
p
Refining the atlas {UX }p if needed, we can and do assume the following properties.
(i) For every p and every i ∈ S p the radial coordinate rip of UX
p
satisfies rip < 1e .
FIBRATIONS BY LAGRANGIAN TORI 10

p
(ii) For every maximal face ∆I of ∆D , the point XI lies in exactly one chart UX .

(iii) For every i ∈ {1, . . . , N } the open set Ri is non-empty.
Since the set W X ∩ D is compact, shrinking δ > 0 if needed we can further assume that
S
(iv) Every fiber f −1 (z) ∩ W X , z ∈ Dδ is contained in N ◦
i=1 Ri and meets Ri for each i ∈ {1, . . . , N }.
p
Eventually, we choose a subordinate partition of unity {τ p }p , put ri = 1e for i 6∈ S p , and define
X
rbi = τ p rip : X −→ [0, 1e ].
p

Lemma 3.1 (Distance functions). The functions rbi defined above have the following properties.
(a) For every point x ∈ Di and every adapted chart UX containing x, there is a neighborhood VX of x
in UX and a smooth function λ : V X −→ (0, ∞) such that the radial coordinate ri of UX satisfies
rbi |VX = λri |VX .
(b) Let ∆I be a maximal face of ∆D . There is an adapted chart UX around the point XI with
coordinates ri such that on UX we have ri = rbi if i ∈ I and rbi = 1e otherwise.
(c) Let UX ⊆ Ri◦ be an adapted chart meeting Di . Then its radial coordinate ri is equal to rbi .
(d) We have rbi < 1e on Ri and rbi = 1e on X \ Ri .
(e) We have rbimi > |f | and rbimi = |f | on Ri◦ ∩ f −1 (Dδ ).
p P
Proof. (a) Put Px = {p : UX ∩U X 6= ∅}, so rbi |U X = p∈Px τ p rip . Since the point x lies on Di , shrinking
p
the chart UX around x if necessary, we can assume that every chart UX meeting U X meets Di , too,
p p p
i.e. Px ⊆ {p : i ∈ S }. Fix p ∈ Px and put V = UX ∩ UX . Recall that, by definition of the adapted
p rp p
charts, both ri and rip extend continuously to the compact closure V . Put λp := rii : V −→ (0, ∞).
P
Now the function λ := p∈Px τ p λp = rrbii is defined on the whole compact closure U X and satisfies the
required equality rbi = λri . Moreover, each summand τ p λp is non-negative, and at each point of U X ,
at least one of them does not vanish; so λ is a positive function on U X , as needed. S
p p q
(b) By property (ii) the point XI lies in UX for a unique p. Now, take UX = UX \ q6=p U X .
(c) Because UX is an adapted chart with associated index {i}, we have ri = |f |1/mi . Since UX ⊆ Ri◦ ,
p
the same holds for all charts UX meeting UX , so rbi = |f |1/mi , too. Thus rbi = ri , as claimed.
(d) By property (i), on Ri we have τ p rip < τ p e−1 6 e−1 if UX p
meets Di and τ p rip = τ p e−1 6 e−1
p
otherwise. Since there is at least one chart UX meeting Di , we get rbi |Ri < e−1 . On the other hand, on
p
X \ Ri we have τ ri = τ e for all p, so rbi |X\Ri = e−1 , as claimed.
p p −1
p Q
(e) On each adapted chart UX we have |f | = j∈S p (rjp )mi 6 (rip )mi for any i ∈ S p , since rip < e−1 < 1
p
by property (i). In turn, if i 6∈ S p then on UX we have (rip )mi = e−mi > δ > |f | by our assumption
about δ. Thus for every i, we have τ p rip > τ p |f |1/mi on UXp
. Away from UX p
both sides of this inequality
P p p P p 1/mi
are zero, so it holds on the entire X. Now rbi = p τ ri > p τ |f | = |f |1/mi , as claimed. The
last assertion follows from part (c). 
Remark 3.2. The distance functions rbi can also be constructed as follows. Fix a hermitian metric on
the normal bundle NDi to Di in X; and identify a tubular neighborhood Ri of Di with the total space
of the disc bundle in NDi . Next, define rbi |Ri as the distance function in each fiber, composed with a
smooth bijection [0, 1] −→ [0, 1e ] whose all derivatives at 1 vanish. Then extend rbi to X by rbi |X\Ri = 1e .
Nonetheless, this alternative construction should not be regarded as more canonical than the one in
Lemma 3.1. Indeed, choosing a hermitian metric on NDi boils down to choosing a partition of unity.
3.3. A new smooth structure
Now, we repeat the definitions of basic functions (10) and (12) using global distance functions rbi
instead of the local ri . That is, we put
−1 t
(17) sbi = log rbi , b
ti = bi = , u
, w bi = η(wbi ), bvi = b
ti − u bi = −dc sbi .
bi , and α
mi sbi b
ti
The 1-form α bi , which will be used in the next section, is the analogue of the angular form αi from
[FdBP22, §4.1]. We use the convention dc = ı(∂ − ∂). Note that Lemma 3.1(e) implies that w bi ∈ [0, 1].
To write the new smooth charts, we replace v i in formula (15) by vbi . That is, we define
(18) ψbi = (g, (b
vj )j∈S\{i} ; ϑ) : Ui −→ Qk,n+1 .
FIBRATIONS BY LAGRANGIAN TORI 11

With these definitions, we have an analogue of [FdBP22, Proposition 3.33].


Proposition 3.3. The collection (ψbi ) for all adapted charts UX and all i in the associated index set
of UX , is a C ∞ atlas on the A’Campo space A. This atlas is compatible with the C 1 -atlas given by
charts (14). The space A endowed with this C ∞ structure has the following properties.
(a) The map π : A −→ X is smooth. Its restriction π|A\∂A : A \ ∂A −→ X \ D is a diffeomorphism.
(b) The map (g, θ) : A −→ [0, 1)×S1 is a smooth submersion. In particular, fA : A −→ Clog is smooth.
(c) For every i ∈ {1, . . . , N } the functions vbi , w
bi and the 1-form α
bi are smooth on A.
The proof of Proposition 3.3, which we outline below, follows the same steps as the proof of [FdBP22,
Proposition 3.33]. The key point is that we can compare the functions b bi , vbi with local functions
ti , w
ti , wi , vi in exactly the same way as we do in loc. cit. More precisely, we have the following formulas.
Lemma 3.4. Let UX be an adapted chart meeting the divisor Di . For a subset I of the associated index
◦ = U ∩ X ◦ . Moreover, put a := −m log λ ∈ C ∞ (U ), where λ is the function from
set of UX put UX,I X I i X
Lemma 3.1(a). Then the functions introduced in (10), (12) and (17) satisfy the following identities.
ati 2
(a) On UX , we have ti − b
ti = 1−at i
, see [FdBP22, Lemma 3.9(c) and formula (35)]
(b) On UX \ D, we have w bi − wi = at, see [FdBP22, Lemma 3.9(c)].
u2i log(1+ati )
(c) On UX , we have u
bi = ui + 1−u i log(1+ati )
, see [FdBP22, formula (34)]
at2 u2 log(1+at ) i
(d) On UX \ D, we have vi − vbi = 1+at i
i
+ 1−u
i
i log(1+ati )
, see [FdBP22, Lemma 3.26(c)].
∞ ◦
(e) There is a bounded function b ∈ C (UX,I ) such that on UX,I ◦ we have vi − bvi = σi b, where
2 2
σi = ti + ti ui , see [FdBP22, Lemmas 3.26(d) and 3.35(a)].
(f) Fix ε ∈ (0, 1). There are bounded functions c, q ∈ C ∞ (UX,I ◦ ), and a bounded 1-form γ ∈ Ω1 (U ◦ )
X,I
ε ◦
such that d(vi − vbi ) = ti · c dvi + ti · q dg + σi · γ on UX,I , see [FdBP22, Lemmas 3.26(e), 3.35(b)].
◦ we have a fiberwise equality db
(g) On UX,I vi = σi mi (1+cti ) dsi +σi γ for some bounded c ∈ C ∞ (UX,I◦ ),
◦ ).
γ ∈ Ω1 (UX,I
(h) There is a smooth 1-form β ∈ Ω1 (UX ) such that on UX \ D we have α bi = dθi + β. In particular,
the 2-form db αi extends to a smooth form on X, see [FdBP22, Lemma 4.5(a),(d)].
Proof. To get (a)–(f) and (h), we repeat the computation in the quoted part of [FdBP22], with the
following modification. In loc. cit, we took another adapted chart UX ′ , and on the intersection U ∩U ′ ,
X X
′ ′
we compared ti , wi etc. with the corresponding functions ti , wi etc. defined for UX′ . Each computation

was based on the existence of smooth function λ : U X ∩ U X −→ (0, ∞) satisfying an equality ri′ = λri .
Now, we work on UX instead of UX ∩ UX ′ , and use the analogous equality r bi = λri given by Lemma
3.1(a). Note that (e) and (f) are also analogues of [FdBP22, Lemma 3.35], with v i replaced by vbi .
For (g), we use a fiberwise inequality dg = 0. As in [FdBP22, p. 40], from definitions (12) we obtain
a fiberwise equality dvi = σi mi dsi . Substituting those two equalities to part (f) proves (g). 
Proof of Proposition 3.3. We describe the minor adjustment needed to adapt the proof of [FdBP22,
Proposition 3.33] to the current setting.
Fix an adapted chart UX meeting a divisor Di . The pair (UX , i) comes equipped with a “transition”
function a ∈ C ∞ (UX ), introduced in Lemma 3.4. We define an algebra Ai in the same way as in
formula (40) loc. cit, with the set Ti replace by singleton {a}. This algebra keeps all the properties
listed in loc. cit. Indeed, in the proof one only needs a ∈ Ti to be a smooth function on UX , small
enough so that the denominators in Lemma 3.4(d) are bounded away from zero and infinity. Lemma
3.4(d) implies that this algebra Ai contains functions vbi − vi .
The modified definition of Ai leads to the new definition of algebras Ii , J , P and the matrix ring
M. We introduce a matrix Ψ in the same way as in formula (41) loc. cit, but with v i replaced by vbi .
The fact that the functions vbi − vi belong to Ai implies that this new matrix Ψ belongs to the new
ring M: to see this, we repeat the proof of Lemma 3.47 loc. cit. Now, parts (a), (b) and smoothness
of vbi follow by repeating the proofs on p. 35 loc. cit. with this modified matrix Ψ.
Smoothness of the 1-form α bi follows from Lemma 3.4(h): indeed, the function θi is is smooth since
it is one of the coordinates of the chart (18); and the 1-form β is smooth as a pullback of a smooth
1-form on UX by π, which is smooth by (a). Similarly, to see that the function w bi is smooth, by Lemma
3.4(b) it is enough to prove smoothness of each local function wi : indeed, the function t = exp(1 − g−1 )
is smooth by (b), and the function a is smooth by (a), as it is a pullback of a smooth function from
FIBRATIONS BY LAGRANGIAN TORI 12

X. Now, we note that wi belongs to the algebra Wi′ introduced in p. 32 loc. cit., hence to the algebra
Ai · Wi′ · P, see §3.4.5 loc. cit. The proof of Lemma 3.49 loc. cit, which is valid in the new setting, too,
shows that all elements of this algebra are smooth, so wi is smooth, as needed. 

3.4. The rounded dual complex


The rounded dual complex ∆ e D is the (homeomorphic) image of the dual complex ∆D by the map
(−η, . . . , −η), see [FdBP22, §3.2.2]. For a face ∆I of the dual complex ∆D , we define its rounded
version ∆ e I as the image of ∆I by the same map. The set ∆ e I is a manifold with boundary and corners,
diffeomorphic to a standard simplex ∆I . Since the spaces ∆ e D, ∆e I are diffeomorphic to ∆D , ∆I one
can interchangeably think about each of them. Due to the nature of our smooth coordinate systems it
will be more convenient for us to formulateF statements using the rounded F versions.
Recall that the stratification X = I XI◦ lifts to a decomposition A = I A◦I given by A◦I = π −1 (XI◦ ).
Lemma 3.5 below gives a handy description of each piece A◦I , cf. Lemma 3.10(c) or Proposition 3.33(d)
of [FdBP22]. To state this result, we introduce some notation.
Let τ : Xlog −→ X be the natural map from the Kato–Nakayama space, see diagram (13). For
I ⊆ {1, . . . , N } put (XI◦ )log = τ −1 (XI◦ ). The restriction of τ to (Xlog )◦I is a smooth (S1 )#I -bundle over
−1
XI◦ . Denoting by πlog the projection A −→ Xlog , we have πlog (XI◦ )log = A◦I .
Lemma 3.5. For every nonempty subset I ⊆ {1, . . . , N }, the map
(19) (πlog , (b eI
vi )i∈I ) : A◦I −→ (XI◦ )log × ∆
is a diffeomorphism of manifolds with boundary and corners.
Proof. In A◦I we have b ti = 0 for i ∈ I, so vbi = −bui = ui = −η(wi ) for i ∈ I, where ui , wi are functions
introduced in (12), for any adapted chart with i in its associated index set. Hence the map (19) indeed
takes values in (XI◦ )log × ∆e I . It is surjective by [FdBP22, Lemma 3.10(c)]. It is injective because a
point in the A’Campo space is uniquely determined by its image in Xlog and the values of all functions
ui ; and since on A◦I we have ui = 0 for i 6∈ I, these data are determined by the image of the map (19),
as needed. To see that (19) is a local diffeomorphism, fix a point x ∈ A◦I and a chart (18) around x,
with coordinates (bvj )j∈I\{i} . At the image of x in ∆ e I , we can choose a smooth chart with coordinates
−η(wj ) for j ∈ I \ {i}. In these coordinates, the map (19) is the identity, as needed. 
Remark 3.6. The space Γ in diagram (13) can also be defined as the closure of the graph of the map
(b bN ) : X \ D −→ RN . Let π
u1 , . . . , u b: Ab −→ X be the A’Campo space obtained this way. Since the
limit of wi as we approach D does not depend on the chosen adapted chart [FdBP22, Lemma 3.6(d)],
the new boundary ∂ A b is equal to the original ∂A, as a subset of Xlog × RN . Thus we can define a
map Φ b: A b −→ A as the identity on ∂ A,
b and π −1 ◦ πb on Ab \ ∂A
b = X \ D. Exactly as in the proof of
[FdBP22, Proposition 3.7(e)], we see that Φ b is a homeomorphism. Moreover, if we define C ∞ -charts ψei
on A b by the same formula as (18), we have ψei = ψbi ◦ Φ, b so they endow A b with a C ∞ -structure.
This way, we see that the A’Campo space A, as a smooth manifold, depends only on the initial
function f : X −→ C and the choice of the distance functions rbi as in Lemma 3.1. Formula (14) for
the C 1 -charts show that the C 1 -structure on A is independent of this choice.

4. Fiberwise Kähler forms


Let f : X −→ C be as in Section 2, and let A be the A’Campo space of f , equipped with the smooth
structure given by charts (18). In this section, we introduce exact forms ω ♯ and ω ♭ appearing in formula
(3), and prove that the resulting forms ωqε are indeed fiberwise symplectic, see Proposition 4.3.
The form ω ♯ is an analogue of the form ωE introduced in [FdBP22, formula (44)]. It is designed so
that for any sufficiently small ε > 0, the form π ∗ ωX + εω ♯ extends to a fiberwise symplectic form on A,
and in particular yields a well-behaved monodromy at radius zero. The additional feature of ω ♯ with
respect to ωE from loc. cit. is its fiberwise J-compatibility, which we prove in Lemma 4.1. In turn, the
form ω ♭ is designed so that the induced metric degenerates at radius zero to the euclidean one on ∆S ,
as required by Theorem 1.1(c). It is inspired by the semi-flat models surveyed in [Li22b, §2.1].
We identify A \ ∂A with X \ D via diffeomorphism π|A\∂A : A \ ∂A −→ X \ D, and denote by J the
induced complex structure. We fix an open set WX as in Section 3.2, i.e. such that f |W X : W X −→ Dδ
FIBRATIONS BY LAGRANGIAN TORI 13

is proper, has connected fibers, and restricts to a submersion over D∗δ . We put W = π −1 (WX ) and
write Wz for the fiber fA−1 (z) ∩ W . We also identify D∗δ with its preimage in Dδ,log .

4.1. The “A’Campo form” ω ♯


Recall that by Proposition 3.3, for each i ∈ {1, . . . , N } we have a smooth function vbi ∈ C ∞ (A) and
a smooth angular 1-form α bi ∈ Ω1 (A), introduced in formula (17). We now define smooth forms on A:
N
X

(20) λ = vbi · α
bi , ω ♯ = dλ♯ .
i=1

Lemma 4.1. There is a smooth function φ : W \ ∂A −→ R satisfying a fiberwise equality λ♯ = dc φ. In


particular, ω ♯ is fiberwise J-compatible, i.e. we have an equality ω ♯ (J · , J · ) = ω ♯ ( · , · ) on the tangent
space to the fiber Wz for all z ∈ D∗δ .
Proof. Clearly, it is enough to prove the first assertion. If D has only one component D1 then
sb1 = −m1 t−1 is fiberwise constant, hence λ♯ = −b v1 · dc sb1 is fiberwise zero. Therefore, we can assume
that D has at least two components. Fix i ∈ {1, . . . , N }.
By definition (17), the function vbi : WX \ D −→ R is equal to the composition v̌i ◦ (b si , t), where
v̌i : T −→ R is defined as v̌i (š, ť) = −(mi š)−1 − η(−mi · šť),
and T ⊆ R2 is the image of the map (b si , t) : WX \ D −→ R2 .
We claim that T is a union of some horizontal segments (or half-lines), ending on the vertical line
š = −1. By Lemma 3.1(d) we have sbi 6 −1, and the equality sbi = −1 holds on X \ Ri . Since D has
at least two components, the set X \ Di contains a nonempty set Rj◦ for some j 6= i, which meets each
fiber by property (iv) of the atlas used to define the functions rbi . Thus the equality sbi = −1 is attained
somewhere on each fiber, and the claim follows since each fiber is connected. R −1
Therefore, we can define a smooth function φ̌i : T −→ R by the formula φ̌i (š, ť) = š v̌i (s, ť) ds.
Fundamental theorem of calculus gives ∂∂š φ̌i = −v̌i . We define a smooth function φi : WX \ D −→ R
si , t). Since ∂∂š φ̌i = −v̌i , we have a fiberwise equality dφi = −v̌i ◦ (b
as φi := φ̌i ◦ (b si = −b
si , t) db si , so
vi db
c c
PN
d φi = −b b by definition (17) of α
vi d sbi = vbi α bi . Eventually, we put φ = i=1 φi , so we have a fiberwise
c
PN i ♯
equality d φ = i=1 vbi α bi = λ , as needed. 

4.2. The “semi-flat form” ω ♭


As in Section 2.1, we fix any subset S ⊆ {1, . . . , N }: in the setting of Calabi–Yau degenerations, it
will correspond to essential divisors. Now, we define smooth forms on A:
X
(21) λ♭ = − mi w bi , ω ♭ = dλ♭ .
bi · α
i∈S
1 P
Lemma 4.2. Put φ♭= − 2t bi2 . Then
i∈S w the form dc φ♭ extends to a smooth form on A, and we
P
have an equality λ♭ = dc φ♭ + 21 i∈S w bi2 · dθ. In particular, we have a fiberwise equality λ♭ = dc φ♭ , so
the form ω ♭ is fiberwise J-compatible.
bi = −mi sbi t and −dc sbi = α
Proof. Using identities w bi from (17), we get
dc w
bi = −mi dc (b bi − mi sbi dc t = mi t α
si t) = mi t α bi dc t.
bi + 1t w
1 2 P
Put φ♭i = − 2t bi , so φ♭ = i∈S φ♭i . The above equality gives
w
dc φ♭i = − 1t w
bi dc w
bi + 1
b2 dc t
w
2t2 i
= −mi w bi −
bi α 1
b2 dc t.
w
2t2 i
Moreover, we have an equality
dc t = −t2 dθ.
Indeed, putting s = log |f | we see that dc s = −dθ and t = −s−1 , so dc t = −dc s−1 = s−2 dc s = −t2 dθ,
as claimed. Substituting this to the previous equality, we conclude that
dc φ♭i = −mi w bi2 dθ.
bi αi + 21 w
By Proposition 3.3(b),(c), the functions θ, w bi are smooth in A, hence so is dc φ♭i . The
bi and the form α
result follows after taking a sum over all i ∈ S. 
FIBRATIONS BY LAGRANGIAN TORI 14

4.3. The form ωqε is fiberwise Kähler


Recall that we have fixed a Kähler form ωX on X, and introduced 2-forms ω ♯ and ω ♭ by formulas
(20) and (21). We now introduce a family of smooth 2-forms on A by formula (3), which reads as
ωqε = π ∗ ωX + ε · (q · ω ♯ + (1 − q) · ω ♭ ), q ∈ [0, 1].
The following proposition summarizes key properties of this family.
Proposition 4.3. There is an ε0 > 0 and δ′ > 0 such that the following holds.
(a) For every q ∈ [0, 1] and ε ∈ [0, ε0 ], the form ωqε is Kähler on the fiber Wz for every z ∈ D∗δ′
(b) For every q ∈ (0, 1] and ε ∈ (0, ε0 ], the form ωqε is symplectic on the fiber Wz for every z ∈ Dδ′ ,log .
(c) Let Aint ◦
S be the union of Int∂A AI for all faces ∆I of ∆S . Then for every ε ∈ (0, ε0 ] there is a
neighborhood Vε of W ∩ Aint ε
S in A such that the restriction ω0 |Vε is fiberwise symplectic.
The proof of Proposition 4.3(a) follows the steps of [FdBP22, §4.2]. The starting point is the following
analogue of formula (53) loc. cit.
Lemma 4.4. Let UX be an adapted chart with associated index set I, and let U = π −1 (UX ). Then
on every fiber fA−1 (r, θ) ∩ U for r > 0 we have equalities
P
(a) ω ♯ = i∈I σi · (mi (1 + ci ti ) dsi ∧ dθi + γi ∧ dθi + mi (1 + ci ti ) dsi ∧ βi ) + ω̌,
P
(b) ω ♭ = i∈I∩S t · (m2i dsi ∧ dθi + dai ∧ dθi + dsi ∧ βi ) + ω e,
for some bounded functions ai , ci and bounded forms βi , γi , ω̌, ω e on UX \ D.
Proof. (a) By definition (20) of ω ♯ we have
X
ω♯ = bi + ω̌ ′ ,
vi ∧ α
db
i∈I
P PN
where ω̌ ′
= i6∈I db vi ∧ α
bi + i=1 vbi dαi . Lemma 3.4(g),(h) implies that the forms db bi for i 6∈ I and
vi , α
αi for all i are bounded on UX \ D, cf. [FdBP22, Lemma 4.5]. Thus the form ω̌ is bounded on UX \ D.
db
Now part (a) follows by substituting the fiberwise formulas for db vi and αbi from Lemma 3.4(g),(h).
(b) By definition (21) of ω ♭ we have
X
ω♭ = − mi dwbi ∧ α e ′,
bi + ω
i∈I∩S
P P
e ′ = − i∈S\I mi dw
where ω bi ∧ α
bi − i∈S mi w e ∈ Ω2 (UX \ D) is bounded.
αi , so as before, the form ω
bi db
By Lemma 3.4(b) we have w bi = wi + tai = −mi si t + tai for some smooth function ai ∈ C ∞ (UX ).
bi in Lemma 3.4(h), we get part (b).
Combining it with the formula for α 
Proof of Proposition 4.3(a). Lemmas 4.1 and 4.2 imply that ωqε |A\∂A is fiberwise J-compatible.
Thus it remains to prove that ωqε tames J on each fiber Wz , z ∈ D∗δ′ . This result is analogous to
[FdBP22, Proposition 4.2]. We now explain how to adapt its proof.
Covering W X with finitely many adapted charts, we see that it is enough to work locally, in an
adapted chart UX . Say that its associated index set is {1, . . . , k}. Reordering the components of D
if needed, we can assume that S ∩ {1, . . . , k} = {1, . . . , l} for some 1 6 l 6 k. We group terms from
Lemma 4.4 as follows:
Xk
ω1 := σi · (mi (1 + cti ) dsi ∧ dθi + γi ∧ dθi + mi (1 + cti ) dsi ∧ βi ),
i=1
(22) l
X
ω2 := t · (m2i dsi ∧ dθi + dai ∧ dθi + dsi ∧ βi ),
i=1
ω3,q := q · ω̌ + (1 − q) · ω
e for q ∈ [0, 1].
For a 2-form ω, we denote by ω J the symmetric part of the form ω( · , J · ). With this notation, we can
split the formula (3) defining ωqε as follows.
(ωqε )J = q · ( 13 π ∗ ωX
J
+ εω1J ) + (1 − q) · ( 13 π ∗ ωX
J
+ εω2J ) + ( 31 π ∗ ωX
J J
+ εω3,q ).
Therefore, to prove Proposition 4.3(a) it is enough to prove the following three assertions.
(i) There exists ε1 > 0 such that for every ε ∈ [0, ε1 ] the form 31 π ∗ ωX
J + εω J is positive definite.
1
FIBRATIONS BY LAGRANGIAN TORI 15

(ii) There exists ε2 > 0 such that for every ε ∈ [0, ε2 ] the form 31 π ∗ ωX
J + εω J is positive definite.
2
(iii) There exists ε3 > 0 such that for every ε ∈ [0, ε3 ] and every q ∈ [0, 1] the form 13 π ∗ ωJ + εω3,q
J is

positive definite.
To see assertion (iii), recall that ω3,q ∈ Ω2 (UX \D) is a family of bounded forms, smoothly parametrized
by q ∈ [0, 1]. Since the interval [0, 1] is compact, this family is uniformly bounded, so for sufficiently
small ε > 0, positive definiteness of π ∗ ωXJ implies that the sum 1 π ∗ ω + εω J is positive definite, too.
3 J 3,q
To see assertion (i), note that the definition of ω1 is entirely analogous to formula (53) in [FdBP22].
Repeating the proof of Proposition 4.3 loc. cit. from this formula onward, we get assertion (i).
It remains to prove (ii). The definition of ω2 is still similar to the aforementioned formula (53). We
now explain how to adapt the proof of Proposition 4.3 loc. cit. to this setting.
Recall that dc si = −dθi and dc θi = dc si . Applying these rules to the definition (22) of ω2 , we get
l
X
(23) ω2J = t · m2i ds2i + m2i dθi2 + dsi · βi′ + dθi · βl+i

i=1

for some 1-forms βi′ bounded with respect to the natural coordinates on UX \ D.
Consider a basis {ds1 , . . . , dsl ; dθ1 , . . . , dθl , dx1 , . . . , dx2(n+1−l) } of the cotangent space T ∗ (U \ ∂A),
where xj stand for real and imaginary parts of the remaining coordinates of UX . Let {ν1 , . . . , ν2l ,
ξ1 , . . . , ξ2(n+1−l) } be the corresponding dual basis of T (U \ ∂A). Let || · || be the maximum norm in
T (U \ ∂A) with respect to this basis, and let || · ||X be a maximum norm in T UX with respect to the
natural coordinates. As in [FdBP22, Lemma 4.10(e),(f)], we note the following two properties.
(a) For i ∈ {1, . . . , 2l}, we have ||π∗ νi ||X → 0 as r1 , . . . , rk → 0.
(b) Let Ξ be the subspace spanned by {ξ1 , . . . , ξ2(n+1−l) }. Then there is a constant K0 > 0 such that
for every ξ ∈ Ξ we have ||π∗ ξ||X > K0 ||ξ||.
The matrix of ω2J in this basis has a block form
 
A Q
t· .
Q P
We claim that, after possibly shrinking the chart UX , the entries of each block have the following
properties, cf. [FdBP22, Lemma 4.10(a)-(d)].
(c) The diagonal terms of A are bounded from below by 41 .
(d) The off-diagonal terms of A converge to 0 as r1 , . . . , rk → 0.
(e) All entries of the blocks Q and P are bounded functions on U \ ∂A.
These properties follow easily from formula (23). Indeed, the (i, j)-term of A for i, j ∈ {1, . . . , l} is the
function δij m2i + βi′ (νj ) + βj′ (νi ), where δij is the Kronecker delta, and mi = mi if i 6 l and mi = mi−l
if i > l. Property (a) and boundedness of the 1-forms βi′ imply that βi′ (νj ) → 0 as we approach the
origin of the chart, so after possibly shrinking UX , we get properties (c) and (d). To see property (e),
we note that P = 0, and the (i, j)-term of Q for i ∈ {1, . . . , 2l}, j ∈ {1, . . . , 2(n + 1 − k)} is βi′ (ξj ).
Since the 1-form βi′ and the vector field ξj are bounded with respect to natural coordinates on UX , so
is the function βi′ (ξj ), as needed.
It is immediate to see that all properties (a)–(e) are preserved by the first 2l steps of the Gaussian
diagonalization. Indeed, at each step we add to νi or ξj a linear combination of vectors ν1 , . . . , ν2l ,
with bounded coefficients, so properties (a), (b) are kept. In the matrix, to each entry of A and Q we
add a function converging to 0 as r1 , . . . , rk → 0; and to each entry of P we add a bounded function.
Hence after possibly shrinking UX at each step, we see that properties (c)–(e) are preserved, too.
This way, we get a basis of T (U \ ∂A) satisfying (a), (b), such that the matrix of ω2J in this basis is
 ′ 
A 0
t· ,
0 P′
where A′ is a diagonal matrix with entries bounded from below by 14 ; and all entries of P ′ are bounded.
Fix a nonzero vector v ∈ T (U \ ∂A), and decompose it as v = ν + ξ, according to the above splitting.
The above block-diagonal form of ω2J shows that
ω2J (ν, ν) > t · 41 ||ν||2 , |ω2J (ξ, ξ)| 6 t · K||ξ||2
for some constant K > 0, uniform on U \ ∂A.
FIBRATIONS BY LAGRANGIAN TORI 16

Now, we repeat the proof of [FdBP22, Proposition 4.3], starting on p. 44 loc. cit, with inequalities
1 ∗ J
(58) and (59) replaced by the above ones; and the forms π ∗ ωX J, ω
E replaced by 3 π ωX , ω2 . The
proof goes through without any further changes, except for the second-to-last line. There, in the
inequality bounding C, the number ε is now multiplied by t: i.e. the modified inequality reads as
C > ( 13 K1 − εtK)||ξ||2 . Shrinking δ > 0 we can make t as small as needed, which proves (ii). 

We now move on to the proof of Proposition 4.3(b),(c), that is, fiberwise non-degeneracy of the
forms ωqε for q > 0 on W ∩ ∂A; and of ω0ε on W ∩ Aint S . We follow the proof of [FdBP22, Proposition
eqε playing the role of ωA
4.12], with ω ε . First, we introduce some notation as in Lemma 4.13 loc. cit.

Fix a smooth chart (18), call it U . Reordering the components of D if needed, we can assume that
the associated index set of this chart is S = {1, . . . , k}, and the index i excluded in the formula (18) is
1. Fix a subset I ⊆ S containing 1, say I = {1, . . . , l}. Then as in [FdBP22, Lemma 4.13], we see that
at each point of UI◦ := U ∩ A◦I we have smooth coordinates

(b
v2 , . . . , vbl , sl+1 , . . . , sk , θ1 , . . . , θk , zi1 , . . . , zin+1−k ).
∂ ∂
These coordinates provide a splitting T UI◦ = VI ⊕ ΘI ⊕ ZI , where VI = span{ ∂b vl }, ΘI =
v2 , . . . , ∂b
∂ ∂
span{ ∂θ1 , . . . , ∂θl }, and π∗ maps ZI isomorphically to the tangent bundle to the stratum π(U ) ∩ XI◦ .
We have the following analogue of [FdBP22, Lemma 4.14].

Lemma 4.5. On UI◦ , the following holds.


(a) For every β ∈ Ω∗ (UX ∩ XI◦ ) we have π ∗ β|VI ⊕ΘI = 0.
(b) For every j > l we have dw bj = 0 and db vj |VI ⊕ΘI = 0.
(c) For every i ∈ {1, . . . , k}, j ∈ {1, . . . , l} we have α bi |VI = 0 and α bi ( ∂θ∂ j ) = δij .
P ζ(ui ) −1
(d) We have db v1 = − li=2 ζ(u 1)
vi , where ζ = (η −1 )′ : [0, 1] ∋ s 7→ s−2 e1−s ∈ [0, ∞).
db
(e) For every i ∈ {1, . . . , l} we have dw bi |ΘI ⊕ZI = 0 and db vi |ΘI ⊕ZI = 0.
(f) For every i ∈ {1, . . . , l} we have ωqε ( · , dθ∂ i ) = εci db
vi , where ci = q + (1 − q) · mi ζ(ui ) if i ∈ S and
ci = q if i 6∈ S.

Proof. The computations are exactly the same as in [FdBP22, Lemma 4.14]. The additional assertions
about the function w bi follow from the fact that on each piece A◦J of ∂A, we have w
bi = 0 if i 6∈ J and
vbi = −η(w
bi ) otherwise; so wbi is either zero or a re-parametrization of vbi . 

Proof of Proposition 4.3(b),(c). First, as in [FdBP22, Lemma 4.15] we prove that, after possibly
shrinking ε > 0 and the chart UX , the restriction ωqε |ZI is nondegenerate. To see this, we note that
P
by Lemma 4.5(b), (e) we have dw bi |ZI = 0 for all i ∈ {1, . . . , N }, so ω ♭ |ZI = − N i=1 mi w αi , which is
bi db
bounded in the natural coordinates of π(UX,I ◦ ). This way, we can repeat the proof of Lemma 4.15 loc.

cit., treating the contribution from ω ♭ as a part of the form ω̌1 .


With non-degeneracy of ωqε |ZI at hand, we now repeat the proof of Proposition 4.12 loc. cit, replacing
each v i for i ∈ {1, . . . , l} by ci db vi . We now outline each step of the proof.
Fix a point x ∈ UI , and assume that q > 0 or x ∈ Int∂A A◦I and I ⊆ S (so we are in the setting
of Proposition 4.3(b) or (c), respectively). Fix a nonzero vector ν ∈ Tx (∂A) which is ωqε -orthogonal
to the radius-zero fiber F := fA−1 (0, θ). We need to prove that ν 6∈ Tx F . In our coordinates, we have
P P Pl
Tx F = ker[ li=1 mi dθi ] and ν = li=2 ai ∂b ∂
vi +

i=1 bi ∂θi + ξ for some ai , bi ∈ R and a vector ξ ∈ ZI ,
see formulas (65) and (66) loc. cit.
We claim that at the point x, the coefficients c1 , . . . , cl from Lemma 4.5(f) are positive. This is clear
if q > 0. If q = 0 then by assumption I ⊆ S, so ci = mi ζ(ui ) for all i, and x ∈ Int∂A A◦I , so ui > 0 for
all i, too, which proves the claim.
Next, we claim that ai = 0 for all i ∈ {2, . . . , l}. Let νiθ := m1i ∂θ∂ i − m11 ∂θ∂ 1 ∈ T F . Lemma 4.5(f)
ci c1 ci c1
gives 0 = ωqε (ν, νiθ ) = ε · ( m i
ai − m 1
v1 (ν)), so m
db i
ai = m 1
v1 (ν). If ai 6= 0 then db
db vi (ν) 6= 0 and by
P l ζ(u ) P l ζ(u )m c
Lemma 4.5(d) we have −1 = dbv11(ν) · i=2 ζ(u1i ) ai = i=2 ζ(u1i )mi1 c1i > 0, a contradiction.
Now, using non-degeneracy of ωqε |ZI , we prove that ξ = 0. Fix ξ ′ ∈ ZI . By Lemma 4.5(b),(f) we
have ωqε (ξ ′ , ∂θ∂ i ) = εci db
vi (ξ ′ ) = 0, so ωqε (ξ ′ , ν) = ωqε (ξ ′ , ξ) since all the coefficients ai vanish. Thus
non-degeneracy of ωqε |ZI indeed implies ξ = 0.
FIBRATIONS BY LAGRANGIAN TORI 17

∂ P ∂
Eventually, we compute the coefficients bi . Lemma 4.5(f) gives 0 = ωqε ( ∂b
vi , ν) = j bj ·εcj db
vj ( ∂b
vi ) =
ζ(ui )
ε(bi ci − b1 c1 ζ(u 1)
) by Lemma 4.5(d). Hence there is a number b such that for all i we have bi = b · ζ(u i)
ci ,
P P
and b 6= 0 since ν 6= 0. Now i mi dθi (ν) = b i mi ζ(u i)
ci 6= 0, so ν 6∈ Tx F , as claimed. 
Remark 4.6 (Monodromy). For q = 1, computing the number b in the above proof shows P that the
ε ∂ 1 ∂
ω1 -symplectic lift to ∂A of the angular vector field ∂θ on ∂Dδ,log is given by P
i ζ(ui ) ∂θi ,
i mi ζ(ui )
which is precisely formula (48) from [FdBP22]. Thus the radius-zero monodromy for ω1ε is exactly the
same as in loc. cit, which means that one prove the results of loc. cit. using the form ω1ε instead of ωA ε.

Remark 4.7 (One can make ωqε Kähler). Replacing ωqε by ωqε + C · dg ∧ dθ for C ≫ 0, we can guarantee
that the form ωqε is symplectic (and Kähler away from ∂A), not just fiberwise symplectic (or Kähler).
Remark 4.8 (One can keep ωqε = ωX over |z| > δ0 ). Fix δ0 ∈ (0, δ′ ) and a smooth function
ρ : [0, ∞) −→ [0, 1] such that ρ(0) = 1 and ρ(r) = 0 for r > δ0 . Following an analogy with
formula (46) in [FdBP22], we can define λ♯δ0 = ρ(|f |)λ♯ , λ♭δ0 = ρ(|f |)λ♭ , and eventually ωqε,δ0 =
π ∗ ωX + ε · (q · dλ♯δ0 + (1 − q) · dλ♭δ0 ). This way, we have fiberwise equality ωqε,δ0 = π ∗ ωX over C \ Dδ0 ;
and in general ωqε,δ0 = ωqε for some ε′ ∈ [0, ε], with ε′ = ε > 0 on ∂A. Thus by Proposition 4.3, the

form ωqε,δ0 is fiberwise symplectic for all ε ∈ (0, ε0 ].

5. Lagrangian fibration at radius zero

P sections, let f : X −→ Dδ be a holomorphic function such that the unique singular


As in the previous
fiber f −1 (0) = N i=1 mi Di is snc. Choose a subset S ⊆ {1, . . . , N }, and recall that ∆S denotes the
subcomplex of the dual complex ∆D spanned by vertices with indices in S. Let A be the A’Campo
space of f , equipped with the family of forms ωqε introduced in formula (3).
In this section, we introduce the expanded skeleton ES , together with a map
G
(24) lS : Asm
S −→ ES , where Asm S = A◦I ,
#I>n, I⊆S
i.e.Asm
S is the union of boundary pieces of the A’Campo space which correspond to maximal and
submaximal faces of the skeleton ∆S . Moreover, we introduce the generic region Agen sm
S ⊆ AS as
G
Agen
S = Int∂A A◦I .
#I=n+1, I⊆S

and write A◦I,θ , Asm


S,θ andAgen
S,θfor the intersections of the above regions with a radius-zero fiber fA−1 (0, θ).
Proposition 5.6, which is the main result of this section, asserts that the restriction of lS to Asm S,θ is
a Lagrangian torus fibration with respect to each form ωqε , with some mild singularities.
As a set, ES is a disjoint union of maximal faces of ∆S and products ∆I × CI , where ∆I is a
submaximal face, and CI is a certain graph, namely a Reeb space of a fixed Morse function on the
Riemann surface XI◦ , defined in Example 5.1. These pieces, glued together in a natural way, give rise
to a manifold with boundary, corners and some mild branching coming from branching vertices of CI .
To describe this structure in precise terms, we introduce a notion of an ivy-like manifold.
In the setting of maximal Calabi–Yau degenerations, Proposition 2.1(c.i) will give a canonical choice
of lS , which is a Lagrangian torus fibration in the usual sense.

5.1. Ivy-like manifolds


An open graph is a topological space C obtained as a quotient of a finite disjoint union of closed
and semi-closed intervals, by some equivalence relation identifying only some endpoints of different
intervals. The images of those endpoints are called vertices of C, a degree of such vertex is the number
of intervals it lies in. The set of vertices of degree at least 3 is called a ramification locus and denoted
Ram(C); similarly, the set of vertices of degree 1 is called the boundary of C, and denoted by ∂C.
Clearly, C \ Ram(C) is a 1-dimensional manifold with boundary ∂C. Since we are only interested in
C as a topological space, we can and do assume that no vertex has degree 2.
An ivy is an open graph C together with a proper continuous map p : C −→ (0, ∞) whose restriction
to each interval constituting C is a homeomorphism onto its image. Note that the restriction of p to
C \ (Ram(C) ⊔ ∂C) is a local homeomorphism. We choose a smooth structure on C \ Ram(C) such
that the restriction p|C\Ram(C) becomes a diffeomorphism onto its image.
FIBRATIONS BY LAGRANGIAN TORI 18

Example 5.1 (Stein factorization of a Morse function, see Figure 2). The above definition is motivated
by the following, classical construction, explained in more detail in [Sae22, §2] and references therein.
Let M be a Riemann surface, and let h : M −→ R be a proper Morse function. The Reeb space of h is
the quotient space C := M/∼ , where for two points x, y ∈ M we put x ∼ y if x and y lie in the same
connected component of a fiber of h. The quotient map h : M −→ C is called a Stein factorization of
h. Writing h = p ◦ h, it is easy to see that (C, p) is an ivy, and that the restriction
−1
h: M \ h (Ram(C)) −→ C \ Ram(C)
−1
is smooth. For c 6∈ ∂C ∪ Ram(C), the fiber Mc := h (c) is a connected submanifold of dimension 1,
i.e. an embedded circle. For c ∈ ∂C the fiber Mc is a single point, either a maximum or a minimum of
h. Eventually, for c ∈ Ram(C) the fiber Mc is a connected union of circles, meeting at saddle points of
h where, since h is Morse, Mc is locally of the form {xy = 0}. Choosing “top” and “bottom” branch of
each such singularity, we conclude that Mc is an immersed circle (but the immersion is not canonical).

∂C

h p

M Ram(C) C R Crit(h)

Figure 2. Stein factorization of a Morse function through its Reeb space

We now return to abstract definitions. An ivy-like chart (with boundary and corners) is a home-
omorphism ϕ : V −→ U × C from a topological space V to a product U × C of a smooth manifold
U (with boundary and corners) and an ivy p : C −→ (0, ∞). The ramification locus of such chart is
Ram(V ) := ϕ−1 (U × Ram(C)), and outer boundary is ∂out V := ϕ−1 (U × ∂C). Note that V \ Ram(V )
inherits a structure of a smooth manifold with boundary and corners from U × (C \ Ram(C)).
We say that a continuous, surjective map ψ : N −→ V from a smooth manifold (with boundary
and corners) to an ivy-like chart V as above is an ivy-like submersion if itSlocally comes from a Stein
factorization, i.e. the following holds. There is an open covering N = i Ni and for each i, there
is a smooth manifold Ni′ (with boundary and corners), a Riemann surface Mi , and a diffeomorphism
τi : Ni′ ×Mi −→ Ni so that the composition ϕ|ψ(Ni ) ◦ψ|Ni ◦τi is of the form (ϕ′ , hi ), where ϕ′ : Ni′ −→ U
is a submersion onto its image, and hi is the Stein factorization of some Morse function hi : Mi −→ R,
see Example 5.1. Clearly, an ivy-like submersion restricts to a usual submersion over V \ Ram(V ).
It will be convenient to call a usual chart a homeomorphism ϕ : V −→ U for some smooth manifold
U (with boundary and corners). In this case we put Ram(V ) = ∂out V = ∅.
An ivy-like manifold S(with boundary and corners) is a Hausdorff topological space E together with
an open covering E = i Vi , such that each Vi is a domain of a usual or ivy-like chart ϕi ; for every
i 6= j we have Ram(Vi ) ∩ Ram(Vj ) = ∅; and the transition maps ϕj |Vi ∩Vj ◦ ϕ−1 i |ϕi (Vi ∩Vj ) are smooth, as
maps between usual manifolds (with boundary and corners). We define its ramification locus Ram(E)
and outer boundary ∂out E as the union of, respectively, all ramification loci and outer boundaries of
the charts. A collection {(Vi , ϕi )} is called an ivy-like atlas. Note that the subspace E \ Ram(E) is a
manifold (with boundary and corners) in the usual sense; and ∂out E ⊆ ∂E.
A continuous, surjective map ψ : N −→ E from a smooth manifold N with boundary and corners to
an ivy-like manifold E as above is an ivy-like submersion if for every i, the composition ϕi ◦ ψ|ψ−1 (Vi )
is a (usual, or ivy-like) submersion. In particular, ψ restricts to a usual submersion over E \ Ram(E).
Definition 5.2 (Ivy-like Lagrangian torus fibrations). Let (Y, ω) be a symplectic manifold of dimension
2n, and let ι : L −→ Y be an embedding (or an immersion). We say that its image ι(L) is an (immersed)
isotropic submanifold if ι∗ ω = 0. It is (immersed) Lagrangian if additionally dim L = n.
FIBRATIONS BY LAGRANGIAN TORI 19

Let E be an ivy-like manifold (with boundary and corners). We say that a map l: Y −→ E is
an ivy-like Lagrangian torus fibration if it is an ivy-like submersion and for every p ∈ E, and every
connected component T of the fiber l−1 (p) the following hold.
(a) If p ∈ E \ (Ram E ∪ ∂out E) then T is Lagrangian torus (S1 )n .
(b) If p ∈ Ram E then T is an immersed Lagrangian torus (S1 )n .
(c) if p ∈ ∂out E then T is an isotropic torus (S1 )n−1 .
We say that l is a Lagrangian torus fibration if Ram E ∪ ∂out E = ∅. In this case, E is a manifold (with
boundary and corners), the map l is a usual submersion and all fibers of l are Lagrangian tori.
5.2. Construction of the expanded skeleton
We return to the setting of Section 4, that is, we let f : X −→ C Pbe a holomorphic function whose
fibers over z 6= 0 are Kähler manifolds of dimension n; f −1 (0) = N i=1 mi Di is snc; and we let A be
the associated A’Campo space with fiberwise symplectic forms ωqε introduced in formula (3).
Our goal is to define the expanded skeleton ES together with a map (24), and prove that the latter
is an ivy-like Lagrangian torus fibration. We first do it piecewise, that is, we define a space EI◦ and a
map lI◦ : A◦I −→ EI◦ , for every maximal or submaximal face ∆I of ∆S .
Let ∆J be a maximal face of ∆S , so the stratum XJ is a point. Fix an adapted chart around XJ such
that its preimage UJ is covered by smooth charts (18). On UJ we have a free, smooth (S1 )n+1 -action,
given in each of those charts by a translation in θi -directions. Denote the quotient map by
(25) qJ : UJ −→ QJ , and define lJ◦ := qJ |A◦J : A◦J −→ EJ◦ ,
where EJ◦ = qJ (A◦J ). Clearly, QJ is a smooth manifold, with global coordinates (b vi )i∈J . By Lemma 3.5,

EJ is a codimension zero submanifold with boundary and corners of QJ , which in the above coordinates
gets identified with rounded simplex ∆ e J . In particular, E ◦ is diffeomorphic to the maximal face ∆J .
J
Let now ∆I be a submaximal face of ∆S , so the corresponding stratum XI is a Riemann surface.
We choose a proper Morse function hI : XI◦ −→ (0, ∞) satisfying the following condition: for every
point p ∈ XI \ XI◦ there is a neighborhood Up of p in XI , and a holomorphic coordinate z ∈ OXI (Up )
at p such that on Up \ {p} ⊆ XI◦ we have hI = |z| or |z|−1 . Clearly, such a Morse function hI exists.
Let hI = pI ◦ hI be the Stein factorization of hI , so pI : CI −→ (0, ∞) is its Reeb space, see Example
5.1. We define an ivy-like chart EI◦ := ∆ e I × CI , and an ivy-like submersion
(26) vi )i∈I , hI ◦ π) : A◦I −→ EI◦ .
lI◦ := ((b
Lemma 5.3. The above maps lI◦ glue to an ivy-like submersion lS : Asm
S −→ ES .
Proof. Fix a maximal face ∆J of ∆S and a submaximal face ∆I contained in ∆J . We explain how to
glue lJ◦ with lI◦ . The result will follow by repeating this construction for all such inclusions ∆I ( ∆J .
Say that J = {1, . . . , n + 1} and I = J \ {1}. By Lemma 3.1(b), there is an adapted chart UX
around the point XJ whose coordinates (z1 , . . . , zn+1 ) satisfy |zi | = rbi , so for each i ∈ J we have
vbi = vi . Moreover, shrinking the chart UX if needed we can assume that the Morse function hI equals
|ez |±1 for some holomorphic coordinate ze of XI at the point XJ . Write ze = λ0 ·λ(z1 )·z1 , where λ0 ∈ C∗ ,
and λ is a holomorphic function on a neighborhood of 0 ∈ C such that λ(0) = 1. Put z = λ−1 0 ·z e, so z
is another holomorphic coordinate of XI at XJ , and we have z = λ(z1 ) · z1 .
Put U = π −1 (UX ) and UI = AI ∩ U ⊆ ∂A, where AI = π −1 (XI ). The holomorphic coordinate z
pulls back to a smooth function on UI , which we denote by the same letter. Since |z| = |λ−1 ±1
0 | · hI ,
the fibers of |z| and hI give the same foliations of U ∩ A◦I .
1
Recall from formula (18) that U ∩ ∂A is covered by charts Vi := {wi > n+2 } ∩ ∂A, with coordinates

S ◦
((vj )j6=i , ϑ). Since on AI we have w1 = 0, the set i6=1 Vi is an open neighborhood of AI ∩ U in UI .
Fix i ∈ I, say i = n + 1, and put V := Vn+1 . The smooth coordinates on V are (v1 , . . . , vn ; ϑ). We
define a function p : V −→ R piecewise, as p = v1 on A◦J and p = m1 −1 ◦
log |z| on AI . We claim that, after
shrinking UX around XJ if necessary,
(27) the map (p, v2 , . . . , vn , ϑ) is a smooth coordinate system on V.
Once (27) is shown, we conclude as follows. The foliation of V given by the projection to the first n
coordinates agrees with the foliations of V ∩ A◦J and V ∩ A◦I given by lJ◦ and lI◦ , respectively. Hence
F a submersion V −→ EV , onto some smooth manifold EV . Now, we define a smooth
these maps glue to
manifold EI,J as i∈I EVi /∼ , where ∼ identifies two points whenever they are images of the same point
FIBRATIONS BY LAGRANGIAN TORI 20
S S
in i6=1 Vi . This way, we get a submersion i∈I Vi −→ EI,J , which restricts to lI◦ and lJ◦ . Eventually,
we glue to EI,J a usual chart lJ◦ (Int∂A A◦J ) ⊆ EJ◦ and an ivy-like chart EI◦ .
It remains to prove claim (27). Clearly, the function p is smooth on the interior of A◦J , where it
equals the coordinate v1 ; and on A◦I , where it is a pullback of a smooth function from XI◦ . It remains
to prove that all derivatives of v1 − p approach 0 as we approach A◦J from A◦I , i.e. as v1 ց 0.
Recall that z = λ(z1 ) · z1 for some non-vanishing, holomorphic function λ such that λ(0) = 1.
−1 −1 −1 v1
Moreover, on A◦I we have v1 = t1 = m1 log |z1 | , so p = m1 log |z| = m1 log |λ(z1 )|+m1 log |z1 | = 1−v1 m1 log |λ(z1 )| .
1 1v2 m log |λ(z )|
1
Thus p − v1 = 1−v 1 m1 log |λ(z1 )|
. By Leibniz rule, it is enough to show that log |λ(z1 )| and all its
derivatives approach 0 as v1 ց 0. By assumption we have log |λ(0)| = log 1 = 0, so by chain rule
−1 ◦
it remains to show that all derivatives of z1 vanish as v1 ց 0. Since v1 = m1 log |z1 | on AI , we have
z1 = exp(−m−1 −1
1 v1 + 2πı · θ1 ), where θ1 is a smooth coordinate on U . Thus dz1 = exp(−m1 v1 ) ·
−1 −1

exp(2πı·θ1 )·(m−1 −2
1 v1 dv1 +2πı dθ1 ). The result follows since all derivatives of the one-variable function
−1 −1 −2
v1 7→ exp(−m1 v1 )v1 vanish at 0. 
Remark 5.4. If for some submaximal face ∆I of ∆S we have (XI , XI◦ ) ∼ = (P1 , C∗ ), then we have
a canonical choice of the Morse function hI : namely, we let hI := |zI |, where zI is a meromorphic
function on XI with a simple zero at one point of XI \ XI◦ and a simple pole at the other one. Since
any two such functions zI differ by a multiplication by a nonzero complex number or taking the inverse,
for any such zI we get the same foliation of XI◦ , hence the same map lI◦ . Note that this function hI
has no critical points, so the ivy-like submersion (26) is a submersion in the usual sense.
As a consequence, if for all submaximal faces ∆I of ∆S we have (XI , XI◦ ) ∼ = (P1 , C∗ ), then the above
sm
canonical choice of Morse functions yields a canonical submersion lS : AS −→ ES (in the usual, not
just ivy-like sense). This happens e.g. if f is an snc model of a maximal Calabi–Yau degeneration
chosen in Proposition 2.1.
5.3. Lagrangian fibration at radius zero
We now prove that the map lS : Asm S −→ ES constructed in Section 5.2 restricts to an ivy-like
Lagrangian torus fibration on Asm
S,θ for each θ ∈ S1 . We begin by describing the fibers of lS .
Lemma 5.5. Fix a point b ∈ ES and let Lb = lS−1 (b). Then the following hold.
(a) The fiber Lb is the image of an immersion ιb : (S1 )k −→ A, where k = n + 1 if b 6∈ ∂out E and
k = n otherwise. If b 6∈ Ram(E) then the immersion ιb is an embedding.
(b) For every q ∈ [0, 1] and every ε > 0, we have ι∗b ωqε = 0.
(c) Fix θ ∈ S1 and let Fθ = fA−1 (0, θ) be the radius zero fiber. Then the restriction of ιb to the preimage
of Lb ∩ Fθ is an immersion ιθb : T −→ Fθ , where T ⊆ (S1 )k is a disjoint union of gcd{mi : i ∈ I}
tori (S1 )k−1 . If b 6∈ Ram(E) then ιθb is an embedding.
Proof. We have b ∈ EI◦ for some I such that ∆I is a maximal or submaximal face of ∆S .
Assume first that the face ∆I is maximal. Then by definition (25) of lI◦ , its fiber Lb is contained in
some chart (18) with associated index set I, and it is parametrized by the coordinates (θi )i∈I . Thus
Lb is an embedded torus (S1 )n+1 , which proves (a). The tangent space Tp Lb at every point p ∈ Lb is
spanned by coordinate vectors ( ∂θ∂ i )i∈I ; so by Lemma 4.5(e) we have db vi |Lb = 0 for every i ∈ I. Thus
Lemma 4.5(f) gives ωqε |Lb = 0, which P proves (b). Part (c) follows from the fact that in coordinates
(18), the fiber Fθ is given by equation i∈I mi θi = θ.
Assume now that ∆I is submaximal. By definition (26) of lI◦ , its fiber Lb is contained in A◦I , and
its image through the diffeomorphism (19) is the preimage in (XI◦ )log of a fiber of hI : XI◦ −→ CI . The
latter is either an embedded circle if b 6∈ Ram EI◦ ∪ ∂out EI◦ , an immersed circle if b ∈ Ram EI◦ , or a
point if b ∈ ∂out EI◦ . Since (XI◦ )log −→ XI◦ is an (S1 )n -bundle, it follows that Lb is, respectively: an
embedded torus (S1 )n+1 ; an immersed torus (S1 )n+1 ; or an embedded torus (S1 )n . This proves (a).
Fix a point p ∈ Lb , and let V ⊆ Tp A◦I be the tangent space to a branch of Lb at p. The point p
is contained in some chart (18) with associated index set I. Its coordinate vectors ∂θ∂ i ∈ Tp A◦I satisfy
dbvj ( ∂θ∂ i ) = 0 for j ∈ I and π∗ ( ∂θ∂ i ) = 0, see Lemma 4.5(e) and (a), so by definition (26) of lI◦ we have

∂θi ∈ V for i ∈ I. Since db vi |V = 0, Lemma 4.5(f) gives ωqε ( · , dθ∂ i )|V = 0. However, the vectors ( ∂θ∂ i )i∈I
span a subspace of V of codimension dim V − #I 6 (n + 1) − n = 1, so ε
P the 2-form ωq vanishes on V ,
as claimed in (b). Part (c) follows as before from the fact that Fθ = { i∈I mi θi = θ}. 
FIBRATIONS BY LAGRANGIAN TORI 21

The above construction is now summarized in the following result. Recall for a radius-zero fiber
Fθ := fA−1 (0, θ) ⊆ ∂A, the region Asm ◦
S,θ ⊆ Fθ is the union pieces AI ∩ Fθ such that #I > n and I ⊆ S,
i.e. ∆I is a maximal or submaximal face of ∆S .
Proposition 5.6 (Lagrangian fibration at radius zero). Let lSθ : Asm S,θ −→ ES be the restriction to Fθ
of the map (24) constructed in Lemma 5.3, for some collection of Morse functions (hI ). Fix ε ∈ (0, ε0 ]
as in Proposition 4.3(b). Then the following hold.
(a) For every q ∈ (0, 1], the map lSθ is an ivy-like Lagrangian torus fibration with respect to the
symplectic form ωqε introduced in formula (3), see Definition 5.2.
(b) If ∆I is a maximal face of ∆S then the image lSθ (A◦I ∩ Fθ ) is diffeomorphic to ∆I . Moreover, the
restriction of lSθ to Int∂A A◦I ∩ Fθ is a Lagrangian torus fibration with respect to ω0ε .
(c) If ∆I is a submaximal face of ∆S then the image lSθ (A◦I ∩ Fθ ) is diffeomorphic to ∆I × CI , where
CI is the Reeb space of hI , see Example 5.1.
(d) Assume that for every submaximal face ∆I of ∆S we have (XI , XI◦ ) ∼ = (P1 , C∗ ). Then the canonical
θ
map lS , chosen in Remark 5.4, is a Lagrangian torus fibration.
Proof. By Lemma 5.5, the map lSθ is an ivy-like Lagrangian torus fibration with respect to ωqε whenever
the latter is a symplectic form. Together with Proposition 4.3, this proves (a) and the second statement
of (b). Lemma 5.5 implies also that each fiber of lS meets Fθ , so lSθ (A◦I ∩ Fθ ) = lS (A◦I ) = EI◦ , which
by construction is diffeomorphic to ∆I if the latter is a maximal face; and to ∆I × CI otherwise. This
proves (b) and (c). Part (d) follows from Remark 5.4. 
Example 5.7 (Hesse pencil, see Figure 3). Consider the Hesse pencil {(x31 + x32 + x33 ) · z = x1 x2 x3 }
from Example 2.3. Recall that the special fiber is a triangle D1 + D2 + D3 , where each Di ∼
= P1 is line
in P . In the A’Campo space, each radius-zero fiber F is a torus, divided into six cylinders F ∩ A◦I .
2

Choose S = {1, 2, 3}, so ∆S = ∆D is a triangle in R3 . As in Remark 5.4, choose the Morse functions
hi : Di \ (D − Di ) −→ R without critical points. Now, the expanded skeleton ES is a circle, divided in
six parts: three disjoint closed segments corresponding to the intersection points Di ∩ Dj , and three
open ones, corresponding to each line Di .

A◦1,2 A◦2 ◦
E1,2 E2◦ ∆2
D2 ∆1,2

D1 A◦1 A◦2,3 E1◦ ◦


E2,3 ∆1 ∆2,3

D3 ◦
∆1,3
A◦1,3 A◦3 E1,3 E3◦ ∆3

(a) special fiber D ⊆ X (b) radius-zero fiber F ⊆ A (c) expanded skeleton ES (d) essential skeleton ∆S

Figure 3. Example 5.7: Hesse pencil {(x31 + x32 + x33 ) · z = x1 x2 x3 }.

Remark 5.8 (Submaximal Calabi–Yau degenerations). Our results are most powerful in case dim ∆S =
n + 1, e.g. for maximal Calabi–Yau degenerations. In fact, if dim ∆S < n then our results are vacuous.
Consider the case dim ∆S = n, so ∆S has submaximal faces, but no maximal one. Then the expanded
skeleton ES is a disjoint union of ivy-like charts ∆I × CI , where ∆I is a submaximal face of ∆S , and
CI is a Reeb space of the chosen Morse function hI : XI◦ −→ [0, 1]. Since the Riemann surface XI◦ is
closed, hI has a maximum, so ∂out ES 6= ∅ and therefore lSθ has fibers of lower dimension.
Nonetheless, in the setting of Calabi–Yau degenerations this can be easily fixed. Indeed, assume
that f is an snc model chosen in Proposition 2.1. Then XI◦ is an elliptic curve, diffeomorphic to S1 ×S1 .
Replacing hI by one of the projections, we get ES to be a disjoint union of products S1 × ∆I ; and the
same argument as in Lemma 5.5(b) shows that lSθ is a Lagrangian fibration.

6. Lagrangian fibrations at positive radius


We keep the notation and assumptions from Section 5. In Proposition 5.6, we have constructed an
ivy-like Lagrangian torus fibration lS of the part of the radius-zero fiber corresponding to maximal
FIBRATIONS BY LAGRANGIAN TORI 22

and submaximal faces of ∆S . In this section, we use the symplectic connection to transport those
Lagrangian tori to positive radius, and study their properties.
In order to apply Proposition 4.3, we need to choose an open subset WX ⊆ X such that f |W X is
proper; and the flows of the vector fields lifted by the symplectic connection do not escape WX . If f
was proper to begin with, we simply take WX = X. In general, we take any open subset VX ⊆ X with
compact closure, and obtain WX by slightly modifying VX near its boundary. This modification was
introduced in [FdBP22, Setting 5.12], we summarize it in Setting 6.1 below. The reader only interested
in the proper case may skip this part and substitute WX = X in what follows.
Setting 6.1 (see Figure 4 or [FdBP22, Figure 5]). Let VX be a domain with smooth boundary, such
that f |V X is proper, f −1 (0) is transverse to ∂VX , and the intersection f −1 (0) ∩ ∂V X is contained in
the union of the open sets Ri◦ introduced in formula (16). Shrinking δ > 0 if needed we get that
S ◦ −1
i Ri contains some neighborhood CX of ∂V X . Put C := π (CX ). By Lemma 3.1(d),(e), on C each
distance function rbi is either equal to 1e or |f |1/mi , hence is fiberwise constant. Thus each form αbi is
fiberwise zero there, so formulas (20) and (21) imply that ω ♯ and ω ♭ are fiberwise zero. Eventually,
definition (3) of ωqε gives a fiberwise equality ωqε |C = π ∗ ωX |C for every q and ε.

CX ′
CX ∂V X

WX

f
δ

Figure 4. Setting 6.1: modifying the domain in case f is not proper.

Now, for r < δ let Φr : CX ∩ f −1 (0) −→ X be the time r flow of the symplectic lift of the radial

vector field ∂r on a fixed ray of C, with respect to ωX : note that it is well defined since the restriction
−1
of f (0) to the collar CX is smooth. Shrinking δ > 0 if needed, we get that for some neighborhood
C0′ of f −1 (0) ∩ ∂V X in f −1 (0) ∩ CX , the image Φr (C0′ ) is contained in CX for all r < δ. Now, let CX ′

be the union of the images Φr (C0′ ), for all r < δ and all rays in C. Put W0 = f −1 (0) ∩ (VX ∪ C0′ ) and
define WX as the union of CX ′ and all connected components of X \ C ′ which do not meet the union
X
of images Φr (∂W0 ), see [FdBP22, Setting 5.12] for details.
Having fixed WX and δ > 0 as above, we choose ε0 > 0 and shrink δ > 0 further so that they satisfy
Proposition 4.3. That is, for every q ∈ (0, 1] and ε ∈ (0, ε0 ], the forms ωqε are fiberwise symplectic on
fA−1 (Dδ,log ) ∩ W , where W = π −1 (WX ). The above choice of WX guarantees that the the symplectic

lift of ∂r with respect to each ωqε defines a flow W ∩ ∂A −→ W for all times r < δ. Indeed, if W 6= A
then near ∂W we have fiberwise equality ωqε = π ∗ ωX , so the above flow is equal to the one of π ∗ ωX ,
so it does not escape W = π −1 (WX ) by construction of WX .
Notation 6.2. In the following, we replace X and its A’Campo space A by WX and its preimage in
A. As usual, we identify X \ D with its preimage A \ ∂A. We fix one ε ∈ (0, ε0 ], and suppress it in the
notation. This way, for every q ∈ (0, 1] we have a flow
(28) Φq• : ∂A −→ A
of the ωqε -symplectic lift of the radial vector field on Clog , with respect to the submersion (g, θ).
6.1. Moving Lagrangian tori by the symplectic connection
z
Fix z ∈ D∗δ , and put θ = |z| , τ = η(− log1|z| ), so the flow Φqτ maps the radius-zero fiber Fθ = fA−1 (0, θ)
F F
to the smooth fiber Xz = f −1 (z). We push the decomposition Fθ = I AI,θ to Xz = I (Xzq )◦I , where
FIBRATIONS BY LAGRANGIAN TORI 23

(Xzq )◦I = Φqτ (A◦I,θ ); and we define


(Xzq )sm q sm
S = Φτ (AS,θ ), (Xzq )gen q gen
S = Φτ (AS,θ ).

This way, (Xzq )sm q ◦


S is the union of pieces (Xz )I for all maximal and submaximal faces ∆I of the skeleton
q gen
∆S ; and (Xz )S is the union of IntXz (Xzq )◦I for all maximal faces of ∆S . Note that (Xzq )sm S is a
submanifold of Xzq of codimension 0 with boundary and corners; and (Xzq )gen S is its open subset.
We remark that the sets IntXz (Xzq )◦I play a similar role as the sets EI introduced in [Li22b, §3.1].
We have the following immediate consequence of Proposition 5.6(a).
Proposition 6.3. For every q ∈ (0, 1], the composition
(29) lSθ ◦ (Φqτ )−1 : (Xzq )sm
S −→ ES

is an ivy-like Lagrangian torus fibration with respect to the Kähler form ωqε on Xz . Its restriction to
(Xzq )gen
S is a Lagrangian torus fibration in the usual sense.
Moreover, if for every submaximal face ∆I of ∆S we have (XI , XI◦ ) ∼
= (P1 , C∗ ) then for the canonical
θ
map lS defined in Remark 5.4, the composition (29) is a Lagrangian torus fibration in the usual sense.
Remark 6.4 (Case q = 0). By Proposition 4.3(a),(c), the form ω0ε is fiberwise symplectic both at
positive radius, and on Agen
S . Therefore, we have a flow

Φ0τ : Agen 0 gen


S,θ −→ (Xz )S ⊆ Xz .

Moreover, by Proposition 5.6(b) we have a Lagrangian torus fibration


lSθ ◦ (Φ0τ )−1 : (Xz0 )gen gen
S −→ ∆S ,

where ∆gen
S is the union of relative interiors of maximal faces of ∆S .

6.2. The associated metrics


Motivated by [KS01, Conjectures 1 and 2], we study the behavior of Kähler metrics
gq = ωqε ( · , J · )
on Xz as z → 0. For maximal Calabi–Yau degenerations, [KS01, Conjecture 1] expects that the
fiberwise Ricci-flat metric should converge, after appropriate re-scaling, to one satisfying a real Monge–
Ampére equation. In our setting, an example of such metric is given by ω ♭ , so looking at formula (3), it
is reasonable to expect such limiting behavior as q → 0. This motivates considering a family of Kähler
metrics parametrized both by z ∈ D∗δ and q ∈ (0, 1]; and studying its limit as q, |z| → 0.
Before we state our result in Proposition 6.12, we need some preparations. We denote by Tvert A
the vertical tangent bundle to the A’Campo space, i.e. the kernel of the map dfA : T A −→ T Clog . As
∗ A, its symmetric powers by S k T ∗ A, etc.
usual, we denote its dual by Tvert vert
Recall that we identify A \ ∂A with its image X \ D; in particular we have there the standard
complex structure J induced from X. We study the following sections of ⊗2 T ∗ (A \ ∂A):
g ♯ = ω ♯ ( · , J · ), g♭ = ω ♭ ( · , J · ), gq = ωqε ( · , J · )

and gnew = t · g♯ , gnew

= t · g♭ , gq,new = t · gq .

Since ω ♯ and ω ♭ are fiberwise J-compatible, the above forms are fiberwise symmetric, i.e. their images
∗ (A \ ∂A) lie in S 2 T ∗ (A \ ∂A). We denote these images by the same letters.
in Tvert vert

Lemma 6.5. The following hold.


(a) On A \ ∂A, we have fiberwise equalities
N 
X  X 1 
♯ 1 2 2 ♭ 2
(30) g = (db bi · mi α
vi ) + σ bi + ǧ and g = · (dw
bi ) + t · m2i α
b2i + ge,
mi σ
bi t
i=1 i∈S

b =b
where σ t2i + b b2i , see formula (12); and ge, ǧ are smooth, symmetric 2-forms on A which are
ti u
bounded with respect to the natural coordinates of X.
♭ , g♯ 2 ∗ 2 ∗
(b) The sections gnew new and gq,new of S Tvert (A \ ∂A) extend to smooth sections of S Tvert A.
FIBRATIONS BY LAGRANGIAN TORI 24

(c) On ∂A, we have fiberwise equalities


N
X X
♯ 1 ′
(31) gnew = η (w bi )2
bi )(dw ♭
and gnew = bi )2 ,
(dw
mi
i=1 i∈S

where η ′ (w bi )2
bi )(dw extends to a smooth form on ∂A, vanishing at the zero locus of w
bi .
bci = α
Proof. Put α bi ◦ J. Definitions (17) give fiberwise equalities
1
(32) bi = −mi tb
w si , so dc w
bi = mi t α
bi bci = −
and α dw
bi .
mi t
♭ ♭
P
P we prove all claims about g . By formula (21) we have ω = − i∈S mi dw
First, bi ∧ α bi + ω e , where
ω
e = i∈S mi w αi . Substituting formulas (32) we get the second equality of (30), with e
bi db g := ω e ( · , J · ).
By Lemma 3.4(h), each form db αi is a pullback of a smooth 2-form on X; hence db αi ( · , J · ) is a smooth
2 ∗
section of ⊗ T X. It follows that e 2 ∗
g extends to a smooth section of S T A, bounded in the natural
coordinates of X, as claimed in (a). Multiplying the result by t, we get a fiberwise equality
X 

gnew = bi )2 + t2 · m2i α
(dw b2i + t · e
g.
i∈S

By Proposition 3.3(c), the functions w bi and forms α bi are smooth, so the right-hand side extends to a
2 ∗
smooth section of S T A, as claimed in (b). Substituting t = 0 we get the second equality of (31).
Now, we prove the claims about g ♯ . Recall from formula (17) that vbi = bti − u
bi , where u
bi = η(w
bi ) =
bi ))−1 . We have the following equalities, cf. [FdBP22, Lemma 3.21(g)]:
(1 − log(w
ui = η ′ (w
db bi = −(1 − log(w
bi ) dw bi−1 ) dw
bi ))−2 · (−w bi = u bi−1 dw
b2i w ti · t−1 dw
b2i b
bi = u bi ,
where in the last equality we used an identity t = b bi . This identity gives also a fiberwise equality
ti w
db bi−1 ) = −tw
ti = d(tw bi−2 dw
bi = −b
t2i · t−1 dw
bi . Combining those two, we get a fiberwise equality
(33) db t−2
vi = −bi ·t
−1
bi − b
dw b2i · t−1 dw
ti u bi = −t−1 σ
bi dw
bi ,
cf. [FdBP22, Lemma 3.22(e)]. Using formula (32), we get fiberwise equalities
1
(34) dc vbi = −mi σ
bi α
bi bci =
and α db
vi .
mi σbi
P P
By definition (20) of ω ♯ , we have ω ♯ = N i=1 db bi + ω̌, where ω̌ = N
vi ∧ α i=1 v αi . Substituting equalities
bi db
(34) we get the remaining part of (a), where like before ǧ := ω̌( · , J · ) is a smooth section of S 2 Tvert ∗ A,

bounded in natural coordinates of X. Multiplying by t, we get a fiberwise equality


XN   N  
♯ t 2 2 (33) X 1 2
gnew = (db σi · m i α
vi ) + tb bi + t · ǧ = − dw bi · db σi · m i α
vi + tb bi + t · ǧ.
mi σb mi
i=1 i=1

Recall that by Proposition 3.3(c) the functions vbi , w


bi and the forms α
bi are smooth. By Lemma
6.6 below, the functions tb
σi are smooth, too, so the right-hand side extends to a smooth, fiberwise
symmetric 2-form on A, as claimed in (b). Substituting t = 0, we get
N
X
♯ 1
gnew |∂A = − bi · db
dw vi .
mi
i=1

It remains to show that −dw vi is a smooth extension of η ′ (w


bi · db bi )2 to ∂A, vanishing on {w
bi )(dw bi = 0}.
The forms dw bi and db vi are smooth on ∂A. Clearly, dw bi = 0 on the zero locus of w bi , so −dw bi · db
vi is
a smooth form on ∂A, vanishing on {w bi = 0}. On the remaining part of ∂A we have b ti = tw bi−1 = 0, so
vi = −db
db ui = −η ′ (w bi , which proves the claim.
bi ) dw 
In the above proof, we used Lemma 6.6 below, which asserts that the functions tb σi are smooth.
However, the forthcoming applications require only the fact that those functions are continuous on
A and vanish on ∂A, which follows directly from boundedness of each σ bi = bt2i + b b2i . Therefore, we
ti u
give only an outline of the proof of Lemma 6.6, which is similar to the case of functions w bi treated in
Proposition 3.3(c). The key idea is that the function t allows to “flatten” all derivatives of σ bi .
σi for i ∈ {1, . . . , N } extend to smooth functions on A.
Lemma 6.6. The functions tb
FIBRATIONS BY LAGRANGIAN TORI 25

Proof. We work in the preimage of an adapted chart UX , say with associated index set {1, . . . , k};
1
and consider its subset U1 = {w1 > n+2 }; with coordinates (18) given by (g; vb2 , . . . , vbk ; ϑ).
Assume first that i > k, i.e. Di ∩ UX = ∅. Then rbi |U1 > 0, so b ti = −(mi rbi )−1 is smooth on U1 . Thus
bi |U1 = b
u ti |U1 − vbi |U1 is smooth by Proposition 3.3(c), and therefore σ bi |U1 is smooth, too.
Assume now that i = 1. We have w b1 |U1 > 0 by definition of U1 , so since w b1 and η|(0,1] are smooth,
so is their composition u b
b1 . By Proposition 3.3(c), so are the functions t1 = vb1 − u b1 and σ b1 .
It remains to consider the case i ∈ {2, . . . , k}, i.e. when vbi is a coordinate of U1 . Now σ bi might not
be smooth, but we will show that tb σi is. Using the identity t = ti wi , we write
(35) tb t2i + tib
σi = wi · (ti b b2i ).
ti u
Recall that in the proof of Proposition 3.3(c) we have introduced a slight modification of the algebra Ai
from [FdBP22, formula (40)]. The functions ti and ui belong to this algebra. Indeed, in the notation
of loc. cit. we have ti , ui ∈ Wi , ti ∈ Ri by formula (38) and ui ∈ Wi by Lemma 3.41(a). Now our
modification of Ai was done in such a way that the functions a used in loc. cit. are exactly those in
Lemma 3.4. Thus Lemma 3.4(a),(c) implies that b ti − ti ∈ Ai and u bi − ui ∈ Ai . We conclude that
b bi ∈ Ai , so tb
ti , u σi ∈ wi · Ai by formula (35).
It follows that the function tbσi lies in the algebra Wi′ · Ai · P introduced in p. 32 loc. cit: here Wi′
and P are some algebras containing wi and 1, respectively. As observed in the proof of Proposition
3.3(c), the proof of Lemma 3.49 loc. cit. shows that all elements of the algebra Wi′ · Ai · P are smooth
on U1 . Thus tb σi is smooth on U1 ; as needed. 
6.3. Admissible paths
We have set up a bi-parametric approach to study the degeneration of Kähler manifolds Xzq , where
the first parameter z determines the fiber Xz of the degeneration, and the second one, q, determines
the Kähler form ωqε within the family (3). To prove our main results, we choose some paths in this
bi-parametric model. First, we distinguish admissible paths as those used in Theorem 1.1(c) and
Theorem 1.2(a),(b), that is, we pose the following definition.
Definition 6.7 (see Figure 1(a)). A smooth path γ = (γ1 , γ2 ) : [0, ρ) −→ Clog × [0, 1] is admissible if
|γ1 (0)| = γ2 (0) = 0 and |γ1 (τ )| > 0, γ2 (τ ) > 0 for τ > 0.
1

Let (r, θ) be coordinates on Clog = [0, ∞) × S1 . In our setting, it is natural to replace r by t = e |f | ,
see formula (11): indeed, both the volume volΩ new and the metrics gq,new , defined to give a finite nonzero
limits as |f | → 0, required a rescaling by a power of t, see Sections 1 and 6.2. So, we endow Clog with
1
another smooth coordinate system, not compatible with (r, θ), given by (t, θ) = (e− r , θ).
For any admissible path γ : [0, ρ) −→ Clog × [0, 1], the restriction γ|[0,ρ′ ) for some ρ′ > 0 can be
parametrized, in the above coordinates (t, θ) of Clog , as γ(h) = (h, θ(h), q(h)), where θ : [0, ρ′ ) −→ S1
and q : [0, ρ′ ) −→ [0, 1] are continuous functions whose restrictions to (0, ρ′ ) are smooth. Such a
parametrization will be called normal. We can now distinguish paths used in Theorem 1.2(c).
Definition 6.8. Let γ(h) = (h, θ(h), q(h)) be a normal parametrization of an admissible path γ. We
say that γ is special if the function q is C 2 and q ′ (0) = 0.
Example 6.9 (see Figure 1(b)). Let γ = (γ1 , γ2 ) : [0, ρ) −→ Clog × [0, 1] be an admissible path such
that for all τ ∈ [0, ρ) we have |γ1 (τ )| = τ , where the absolute value refers to the natural coordinates
(r, θ) on Clog , see Theorem 1.2(c). We claim that γ is special. Indeed, the normal parametrization of γ
1 1
is (h, θ(h), q(h)), where θ(h) is the argument of γ1 (e− h ), and q(h) = γ2 (e− h ). Now q is smooth since
1 1
γ2 is; and q ′ (h) = −h−2 e− h γ2 (e− h ) → 0 as h → 0, as needed.
Notation 6.10. Let h 7→ (h, θ(h), q(h)) be a normal parametrization of an admissible path. We
denote by Xhγ the fiber (t, θ)−1 (h, θ(h)) ⊆ A, equipped with a 2-form ωhγ := ωq(h) ε , defined in formula

(3). Recall that ε > 0 is a number fixed at the beginning of this section, sufficiently small for so that
Proposition 4.3 holds. For a fixed q > 0, we write Ψqh = Φqη(h) , so Ψq• is a reparametrization of the flow
(28) in such a way that Ψqh maps X0γ to Xhγ . We define regions (Xhγ )gen γ sm γ
S ⊆ (Xh )S ⊆ Xh as in Section
q(h)
6.1, i.e. as images of (X0γ )gen
S and (X0γ )sm γ sm
S by the flow Ψh . Thus on (Xh )S we have an ivy-like
Lagrangian torus fibration (29), with respect to the form ωhγ . We write gh,newγ
:= h · ωhγ ( · , J · ) for the
rescaled metric on Xh . We denote by Aγ ⊆ A the union of fibers Xh ; so Aγ , and put Aγ+ := Aγ \ X0γ .
γ γ
FIBRATIONS BY LAGRANGIAN TORI 26

6.4. The Gromov–Hausdorff limit


With the above notation, we have the following consequence of Lemma 6.5.
Lemma 6.11. Let γ be an admissible path. Then for every h > 0 we have
γ
X
gh,new = (1 − q(h)) · ε · bi )2 + e
(dw gh ,
i∈S
P
where e gh is a Riemannian metric on Xhγ . In particular, g0,new
g0 = 0, and for h > 0, e γ
= ε· bi )2 .
i∈S (dw

Proof. Let gX = ωX ( · , J · ) be the Kähler metric on X. Recall that ωhγ = ωq(h)


ε , so formula (3) gives

γ
gh,new = h · π ∗ gX + q(h) · ε · gnew
♯ ♭
+ (1 − q(h)) · ε · gnew .

Substituting the second equality of (31) we get the required formula, with geh := h · π ∗ gX + q · ε · gnew .
q(h)·ε
Since the path γ is admissible, we have q(h) > 0 for h > 0, so geh = h · ω1 ( · , J · ) is a Riemannian
metric by Proposition 4.3(a); and q(0) = 0, so e
g0 = 0, as claimed. 
We can now prove the main result of this section. For a definition and basic properties of the
Gromov–Hausdorff convergence we refer to [BBI01, §7.3, 7.4].
Proposition 6.12. Let γ be an admissible path. For h > 0, we equip the fiber Xhγ with a Riemannian
γ
metric 1ε · gh,new . Then we have Gromov–Hausdorff convergence:
lim (Xhγ , Xhγ \ (Xhγ )sm >2
S ) = (∆S , ∆S ),
h→0

where ∆S ⊆ ∆D ⊆ RN is the essential skeleton, see Section 2.1, equipped with the standard metric of
RN ; and ∆S>2 ⊆ ∆S is the union of faces of ∆S of codimension at least 2.
Proof. For i ∈ {1, . . . , N } put w bi′ = w
bi if i ∈ S and wbi′ = 0 otherwise, and consider a smooth map
µ = (w ′ bN ) : A −→ R . It is easy to see that µ(A◦I ) = ∆I for every face ∆I of ∆S , cf. Lemma
b1 , . . . , w ′ N

3.5 or [FdBP22, Lemma 3.10], so µ(X0γ ) = ∆S and µ(X0γ \ (X0γ )sm >2
S ) = ∆S .
For any h ∈ [0, δ), the restriction of µ to a fiber Xhγ gives maps of compact metric spaces
µh : Xhγ −→ ∆S ∪ µ(Xhγ ), and µh>2 : Xhγ \ (Xhγ )sm >2 γ γ sm
S −→ ∆S ∪ µ(Xh \ (Xh )S ).
γ
Let dh , dstd be the distance functions associated to the metric 1
ε · gh,new on Xhγ and to the euclidean
1
metric on ∆S , respectively. Lemma 6.11 implies that for any two points x, y ∈ Xhγ we have (1 − q(h)) 2 ·
dstd (µh (x), µh (y)) 6 dh (x, y) 6 dstd (µh (x), µh (y)) + Dh , where Dh is the diameter of Xhγ with respect
to the metric e gh , hence Dh → 0 as h → 0. Since q(h) → 0 as h → 0, too, we conclude that the
distortion of µh and µ>2 h , see [BBI01, Definition 7.1.4], converges to 0 as h → 0.
Continuity of µ implies that for every η > 0 there is δ0 > 0 such that for every h ∈ [0, δ0 ], each
point of ∆S is at distance at most η to µ(Xhγ ). This means that µ(Xhγ ) is an η-net for ∆S ∪ µ(Xhγ ), see
[BBI01, Definition 1.6.1]. Similarly, µ(Xhγ \ (Xhγ )sm >2 γ γ sm
S ) is an η-net for ∆S ∪ µ(Xh \ (Xh )S ). Shrinking
>2
δ0 > 0 if needed, we infer that µh and µh are η-isometries, see Definition 7.3.27 loc. cit. The required
Gromov–Hausdorff convergence follows from Corollary 7.3.28(2) loc. cit. 
6.5. Summary
Before we specialize to the Calabi–Yau case, we summarize the general results obtained so far.
Theorem 6.13. Let (X, ωX ) be a Kähler manifold, let f : X −→ C be a holomorphic function whose
unique singular fiber f −1 (0) is snc, and let D1 , . . . , DN be its irreducible components. Choose a subset
S ⊆ {1, . . . , N } so that
T ∆S has a maximal or a submaximal face, i.e. there is I ⊆ S such that
#I > dim X − 1 and i∈I Di 6= ∅. Moreover, choose an open subset V ⊆ X such that f |V is proper,
let W be its slight modification as in Setting 6.1 (one can take W = X if f is proper), and let
Wz = f −1 (z) ∩ W . Then there is a δ > 0 such that for every ε > 0 small enough, the following hold.
(a) Kähler potential. For every z ∈ D∗δ and every q ∈ [0, 1], the form ωqε |Wz defined in (3) is Kähler.
(b) Lagrangian fibration. For every z ∈ D∗δ and every q ∈ (0, 1], formula (29) defines an ivy-like
Lagrangian torus fibration of a codimension zero submanifold (Wzq )sm
S ⊆ Wz with boundary and corners,
with respect to the Kähler form ωqε .
FIBRATIONS BY LAGRANGIAN TORI 27

(c) Gromov–Hausdorff limit. Fix a path (z, q) : [0, δ) −→ Clog × [0, 1] such that |z(0)| = q(0) = 0 and
z(h), q(h) > 0 for h > 0. For each h > 0, let Wh be the fiber Wz(h) with the Kähler metric given by
ε , rescaled so that its diameter is independent of h. Then we have Gromov–Hausdorff convergence
ωq(h)

lim (Wh , Wh \ Whsm ) = (∆S , ∆S>2 ),


h→0

where ∆S is the subcomplex of the dual complex of f −1 (0) spanned by vertices indexed by S, equipped
with some multiple of the euclidean metric; and ∆>2
S is the union of its faces of codimension at least 2.

Proof. Parts (a), (b) and (c) are proved in Propositions 4.3(a), 6.3 and 6.12, respectively. 
Proof of Theorem 1.1. Let f be a model constructed in Proposition 2.1; and let ∆S be the essential
skeleton. Now the result follows from Theorem 6.13 and the fact that for the canonical choice of lS
described in Remark 5.4, formula (29) defines a usual Lagrangian torus fibration. 

7. The Calabi–Yau case


From now on, we assume that f ◦ is a maximal Calabi–Yau degeneration admitting a semi-stable
snc model. Recall from Section 2.2 that the latter property can always be achieved after a finite
base change, which preserves maximality. We let f : X −→ C be the snc model of f ◦ constructed in
Proposition 2.1. We use notation introduced in Section 2.2. P
By Proposition 2.1(b) we have a holomorphic (n + 1)-form Θ on X whose divisor is i (νi − 1)Di
with mini νi = 0. Thus the subcomplex ∆S ⊆ ∆D for S = {i : νi = 0} is the essential skeleton, see
formula (7). As f ◦ is maximal, [NX16, Theorem 4.1.10] gives dim ∆S = n, i.e. ∆S has a maximal face.
Θ
We choose a section Ω of the vertical canonical bundle given by Ω := d log f , see formula (8). Its
restriction to any smooth fiber is a holomorphic n-form, which we denote by Ω, too. We denote by
volΩ the induced fiberwise volume form, defined in formula (5). We put
(36) Ωnew = tn · Ω and volΩ n Ω
new = t · vol .

7.1. Almost all volume is fibered by Lagrangian tori


In Proposition 6.3 we have constructed a Lagrangian torus fibration on the subset (Xz )sm S of the
fiber Xz . Proposition 7.2 below shows that as z → 0, the volume of (Xz )sm S – and even of its subset
(Xz )gen
S – approaches the whole volume of the fiber. The proof follows a well-known local computation,
see e.g. [BJ17, §3] or [Li22b, §3.1], which we perform in the A’Campo space. The key point is that the
suitably rescaled fiberwise volume forms (36) extend to finite, nonzero ones at radius zero.
Lemma 7.1. Let ∆I be a maximal essential face. The following hold.
Vn ∗ V2n ∗
(a) The forms Ωnew and volΩ new extend to smooth sections of C TC,vert A and Tvert A, respectively.
(b) The restrictions Ωnew |∂A and volnew |∂A are zero outside of Agen
S .
(c) Let UX be an adapted chart with associated index set I, let U = π −1 (UX ), and let wj , θj be the
corresponding functions introduced in (10). Then for some nonvanishing holomorphic function
c ∈ OX∗ (U ), and every i ∈ I, we have a fiberwise equality on U
X
^
(37) Ωnew = ±c (dwj − t2 ıdθj ).
j∈I\{i}

(d) The number c0 := c(XI ) does not depend on the maximal essential face ∆I , up to a sign.
(e) On U ∩ A◦I we have fiberwise equalities
^ ^
(38) Ωnew = ± c0 · dwj and volΩ n 2
new = (−1) · |c0 | · (dwj ∧ dθj ).
j∈I\{i} j∈I\{i}

Proof. First, we prove the assertions about Ωnew . Choose an adapted chart UX with associated index
Θ
set J, say J = {1, . . . , k}, fix an index i ∈ J, say i = 1; and let U = π −1 (UX ). Recall that Ω = d log f,
where Θ is a holomorphic (n + 1)-form with zero of order νj − 1 along each Dj . There is a nonvanishing
holomorphic function c ∈ OX ∗ (U ) such that
X
k
Y k
Y
ν −1 ν
(39) Θ|UX = c · zj j dz1 ∧ · · · ∧ dzk ∧ ζ = c · zj j d log z1 ∧ d log z2 ∧ · · · ∧ d log zk ∧ ζ,
j=1 j=1
FIBRATIONS BY LAGRANGIAN TORI 28

∂ log f
where ζ = dzik+1 ∧ · · · ∧ dzin+1 comprises the remaining coordinates of UX . Since ∂ log z1 = m1 , we have
k
Θ c Y νj
Ω= = zj d log z2 ∧ · · · ∧ d log zk ∧ ζ.
d log f m1
j=1

Formulas (12) give log zj = sj + ıθj = − tm1 j wj + ıθj , so we have a fiberwise equality
1 1
d log zj = − dwj + ı dθj = − (dwj − tmj ıdθj ) .
tmj tmj
Substituting it to the above expression for Ω, we get a fiberwise equality
k
c · (−1)k−1 Y νj
(40) Ω = t1−k · zj · (dw2 − tm2 ıdθ2 ) ∧ · · · ∧ (dwk − tmk ıdθk ) ∧ ζ.
m1 · . . . · mk
j=1

By Proposition 2.1(b), we have νj > 0 for all j, with equality if and only if j ∈ S. Hence the function
Qk νj
j=1 zj is smooth on UX , and equals 0 on UX ∩ XJ unless J ⊆ S. The function c and the form ζ are
n
smooth on UX by definition. Since Vn wj∗, θj are smooth in U ⊆ A, formula (40) shows that t Ω = Ωnew
extends to a smooth section of C TC,vert A, and Ωnew |U ∩A◦J = 0 unless k = n + 1 and J ⊆ S, i.e.
unless ∆J is a maximal, essential face. This proves the first part of (a) and (b). Moreover, if ∆J is a
maximal, essential face, i.e. J = {1, . . . , n + 1} ⊆ S then (mi , νi ) = (1, 0) for all i ∈ J by Proposition
2.1(a),(b). Substituting this to formula (40) we get part (c). Furthermore, substituting t = 0 gives the
following fiberwise equality on U ∩ ∂A
Ωnew = c · (−1)n · dw2 ∧ · · · ∧ dwn+1 .
Note that on U ∩ ∂A \ π −1 (Dj ) we have dwj = 0, so on U ∩ ∂A \ A◦J both sides of the above equality
are zero. On A◦J , the function c above is constant, equal to the value of c ∈ OX
∗ (U ) at the point X .
X J
This proves the first equality in (38).
Now, we prove the assertions about volΩ
new . Using formula (40) we get a fiberwise equality

Yk ^k ^k
|c|2
Ω ∧ Ω = t2−2k · |zj |2νj
· (dw j − tm j ıdθ j ) ∧ ζ ∧ (dwj + tmj ıdθj ) ∧ ζ =
m21 · . . . · m2k j=1 j=2 j=2
1 k k
1−k |c|2 · (−1)(k−1)(n+1−k) · (2ı)k−1 (−1) 2 (k−1)(k−2) Y 2νj
^
=t · · |zj | · (dwj ∧ mj dθj ) ∧ ζ ∧ ζ,
m21 · . . . · m2k j=1 j=2

where the sign (−1)(k−1)(n+1−k) comes from moving the (n+1−k)-form ζ to the right; and the constant
1
(2ı)k−1 (−1) 2 (k−1)(k−2) comes from multiplying out the 1-forms. Like before, we conclude that tn Ω ∧ Ω
V
extends to a smooth section of 2n Tvert∗ A, and vanishes on U ∩ ∂A unless ∆ is a maximal face of ∆ .
J S
In the latter case we have (mj , νj ) = (1, 0) for all j ∈ J, so the above expression reads as
1
tn Ω ∧ Ω = |c|2 · (2ı)n (−1) 2 n(n−1) · (dw2 ∧ dθ2 ) ∧ · · · ∧ (dwn+1 ∧ dθn+1 ).
which ends the proof of (a), (b) and (e). The remaining part (d) is shown in the proof of [BJ17,
Theorem 7.1], let us recall the argument here for completeness.
Fix two maximal faces ∆I and ∆J of ∆S . We claim that c(XI ) = ±c(XJ ). Since the essential
skeleton is a pseudomanifold [NX16, Theorem 4.2.4(3)], we can assume that ∆I and ∆J meet along a
common submaximal face ∆I ′ , see part (3) of the definition of pseudomanifold given in §4.1.2 loc. cit.
Say that I ′ = {2, . . . , n + 1}, I = I ′ ⊔ {1}, J = I ′ ⊔ {n + 2}. By Proposition 2.1(c.i), the stratum
XI ′ is isomorphic to P1 , and XI , XJ are two points on XI ′ . By Proposition 2.1(b) we have νi = 0 for
i ∈ {1, . . . , n + 2}, so in an adapted chart at a point p ∈ XI ′ , the local formula (39) defining c reads as
Θ = c · d log z2 ∧ · · · ∧ d log zn+1 ∧ ζ,
where up to a sign we have ζ = d log z1 if {p} = XI , ζ = d log zn+2 if {p} = XJ , and otherwise ζ = dz
for a holomorphic coordinate z on XI ′ . It follows that the logarithmic form c · ζ is the residue of Θ
along XI ′ ; with poles at XI and XJ whose residues are equal, respectively, ±c(XI ) and ±c(XJ ). By
the Poincaré residue theorem, those residues add up to 0, so c(XI ) = ±c(XJ ), as claimed. 
FIBRATIONS BY LAGRANGIAN TORI 29

We can now prove the title result of this section. Recall that in Proposition 6.3, we have constructed
a Lagrangian torus fibration of a region (Xzq )sm
S ⊆ Xz , which maps the generic region (Xz )S
q gen
to the
union of maximal faces of ∆S . The following result shows that this region is large.
Proposition 7.2. Fix q ∈ (0, 1], and let (Xzq )gen
S be as in Section 6.1. Then we have
R
(Xzq )gen volΩ
S
lim R Ω
= 1.
Xz vol
|z|→0

Proof. Put θ = z
|z| , Fθ = fA−1 (0, θ). The sets (Xzq )gen
S and Xz are images of Agen
S,θ and Fθ by the flow
(28). By Lemma 7.1 the form tn volΩ extends to a fiberwise form volΩ
new on A, whose restriction to Fθ
is zero outside of Agen
S,θ . Therefore, we have
Z Z Z Z
n Ω Ω Ω
lim t vol = volnew = volnew = lim tn volΩ ,
|z|→0 (Xzq )gen Agen Fθ |z|→0 Xz
S S,θ

which ends the proof. 

7.2. Asymptotic Ricci flatness


Recall that on each smooth fiber Xz of f ◦ , the cohomology class of ωX is represented by a unique
Ricci-flat form ωCY , and this condition can be written equivalently as
1 n Ω
n! ωCY = c · volnew ,
for some positive function c on the base D∗δ , see formulaR (5). The volume form on the right-hand side is
1 n
rescaled so that the total volume on both sides equals Xz n! ωX . To see that the appropriate rescaling
is indeed of the same order as volΩnew , note that it should integrate to a nonzero number on all fibers
in the A’Campo space, including the radius-zero ones. This is true for volΩ new by Lemma 7.1(e).
Alternatively, one can use [KS01, Definition 1]: indeed, in the notation of loc. cit. the rescaled
volume form should be of order ((log |q|)m |q|2k )−1 volΩ , where q = f (so − log |q| = t−1 ), m = n since
f ◦ is maximal, and k = mini νi , which equals 0 by Proposition 2.1(b).
Proposition 7.3 below, roughly speaking, asserts that our forms ω γ give some approximation of ωCY
in the generic region (Xzq )gen γ
S . Of course, due to some choices made in the definition of ω , we cannot
γ
expect an exact equality ω = ωCY at positive radius (nor off the generic region, see Remark 7.4).
Proposition 7.3. Let γ be an admissible path, see Definition 6.7. There is a continuous function
c+ : Aγ+ ∪ (X0γ )gen γ γ gen
S −→ R>0 such that on each fiber Xh and on (X0 )S , we have we have an equality
γ n
(41) 1
n! (ωh ) = c+ · volΩ
new

and the restriction of c+ to (X0γ )gen


S is a positive constant. This constant equals εn · (n + 1) · |c0 |−2 ,
where ε > 0 is a number fixed in Notation 6.2, and c0 ∈ C∗ is the number in Lemma 7.1(d).
γ γ gen
Proof. By Lemma 7.1(a),(e) the form volΩ new extends to X0 and restricts to a volume form on (X0 )S ,
so we have a continuous function c+ : Aγ+ ∪ (X0γ )gen S −→ R satisfying fiberwise equality (41). By
Proposition 4.3(a), (ωhγ )n is a volume form on each fiber Xhγ for h > 0, so the function c+ is positive
there. It remains to prove (41) at radius zero, i.e. prove that on (X0γ )gen
S we have
1 ε n
n! (ω0 ) = (εn · (n + 1) · |c0 |−2 ) · volΩ
new .

Recall that (X0γ )gen


S is the union of pieces Int∂A A◦I ∩ X0γ for all maximal faces ∆I of the essential
skeleton ∆S . Thus it is enough to show that the above equality holds fiberwise on each A◦I .
Fix a maximal, essential face ∆I . By Lemma 3.1(b), on A◦I we have w bi = wi , α bi = dθi for i ∈ I and
P
w
bi = 0, αbi = 0 for i 6= I. Thus ω = − i∈I dwi ∧ dθi by formula (21). Since π(A◦I ) = XI◦ is a point,

definition (3) of ωqε shows that on A◦I we have ω0ε = ε · ω ♭ , and therefore
X ^ (38)
1
n! (ω ε n
0 ) = (−ε)n
· (dwj ∧ dθj ) = (εn · (n + 1) · |c0 |−2 ) · volΩ
new ,
i∈I j∈I\{i}

as needed. 
FIBRATIONS BY LAGRANGIAN TORI 30

Remark 7.4. The function c+ does not extend to the remaining part X0γ \ (X0γ )gen S . Indeed, on
gen Ω γ
∂A \ AS the form volnew is zero by Lemma 7.1, while ω0 has some contributions from pullbacks of
forms from X, namely: the initial Kähler form ωX , the forms β from Lemma 3.4(h), and the forms
αi , cf. Lemma
db S 4.4(b). In particular, on the subset Ri◦ ∩ X0γ , which can be identified with an open
1 γ 1 n
subset of Di \ j6=i Dj , the form n! (ω0 )n restricts to the standard volume form n! ωX .
The following easy consequence of Proposition 7.3 will be useful in the next section.
Lemma 7.5. Fix h > 0. Let Lh ⊆ Xhγ be a fiber of a Lagrangian torus fibration (29), and let volgnew be
γ √
a Riemannian volume form induced by the metric gh,new . Then we have volgnew |Lh = c+ · |Ωnew |Lh |.
Proof. Fix a point x ∈ Lh and let e1 , . . . , en be an orthonormal basis of Tx Lh . Since Lh is Lagrangian
with respect to a J-compatible form h · ωhγ , the collection e1 , . . . , en , Je1 , . . . , Jen is an orthonormal
basis for Tx Xhγ with respect to the metric gh,new
γ
= h · ωhγ ( · , J · ) see formulas (1.2) and (1.2)’ in [HL82].
Substituting it to the formula (41) multiplied by tn , we get 1 = c+ · |Ωnew (e1 , . . . , en )|2 , as needed. 
7.3. Generic Lagrangian tori are asymptotically special
We recall the notion of special Lagrangian submanifolds, introduced in [HL82]. Let Y be a Calabi–
Yau manifold, with a holomorphic volume form ΩY and a Ricci-flat Kähler form ωCY , see equation
(5). A Lagrangian submanifold L of Y is special of phase ̟ ∈ S1 if Im(̟ · ΩY )|L = 0. Condition
(5) guarantees that special Lagrangian submanifolds are calibrated with respect to Re(̟ · ΩY ), see
[GHJ03, §8.1] in particular they minimize volume in their homology class, see Proposition 7.1 loc. cit.
The specialty condition makes the geometry of Lagrangians much more rigid, see [AB+ 09, §6.1]; and
indeed the Lagrangian tori in SYZ conjecture are expected to be special, see paragraph (2) in the
introduction. In this section, we prove that for our tori specialty holds asymptotically, in the generic
region, as claimed in Theorem 1.2(c).
To give a precise statement in Proposition 7.7 below, we introduce some notation. Fix an admissible
path γ, see Definition 6.7, and a point b ∈ ES . Consider a family of Lagrangian torus fibers Lh ⊆ Xhγ
over b. It follows from Lemma 7.5 that for h > 0, the form Ω does not vanish on Lh , so at each point it
is a nonzero complex multiple of a volume form on Lh . This allows us to pose the following definition.
Definition 7.6. Let volh be a volume form on Lh . Let ch : Lh −→ C∗ be the smooth function satisfying
volh = ch · Ω|Lh .
ch
We define the phase map as ̟h := |ch | : Lh −→ S1 .
Clearly, this definition does not depend on the choice of the volume form volh . We let Lγ ⊆ Aγ be
the union of fibers Lh for all h > 0; and let L+ = L \ L0 , so the fiberwise phase map yields a smooth
map ̟ : L+ −→ S1 . We can now state the main result of this section, which implies Theorem 1.2(c).
Proposition 7.7. Let γ be a special admissible path, see Definition 6.8. Let b be an interior point of
a maximal, essential face ∆I , and let L ⊆ Aγ be the family of fibers of the Lagrangian fibration (29)
over b. Then the phase map ̟ : L+ −→ S1 introduced in Definition 7.6 extends to a continuous map
L −→ S1 , whose value on L0 is a constant
c0
(42) ̟|L0 = ± ın · ,
|c0 |
where c0 ∈ C∗ is the number in formula (38), and the sign depends only on the face ∆I .
In the generic region (Xhγ )gen
S our form satisfies the “asymptotic Ricci flatness” condition (41), so
the “asymptotically special” Lagrangian tori Lh should “asymptotically” minimize volume there. And
indeed, we will see that the arguments of [HL82, §III.1] yield the following corollary.
Corollary 7.8. Let γ be a special admissible path, let b be an interior point of a maximal face of ∆S ,
and for h > 0 let Lh ⊆ (Xhγ )gen g
S be the fiber over b of the Lagrangian torus fibration (29). Let volnew
γ
be a Riemannian volume form on Lh associated to the metric gnew . Then we have a convergence
R g
Lh volnew
(43) lim R g = 1,
Γ volnew
h→0 minΓ∈[Lh ]

where [Lh ] is the class of Lh in Hn ((Xhγ )gen


S ; R).
FIBRATIONS BY LAGRANGIAN TORI 31

The remaining part of this section is devoted to the proof of Proposition 7.7 and Corollary 7.8.
We work in (Xhγ )gen
I for a fixed maximal essential face ∆I , say I = {1, . . . , n + 1}. By Lemma 3.1(b),
there is an adapted chart UX around the point XI◦ such that for each i ∈ I, the distance function rbi
equals the radial coordinate ri of UX . Thus on U := π −1 (UX ) the functions w bi , vbi and the form α
bi
introduced in formula (17) for each i ∈ I are equal to the functions wi , vi and the form dθi introduced
for UX in formula (12). Moreover, for i 6∈ I the forms db vi , dwbi and αbi are identically zero. As a
consequence, on U formulas (20) and (21) read as
X X
(44) ω♯ = dvi ∧ dθi and ω ♭ = dwi ∧ dθi .
i∈I i∈I
Since the radius-zero torus L0 is a compact subset of U ∩Int∂A A◦I ,
we can and do work in a neighborhood
V of L0 in U where each function wi is bounded away from 0. As a consequence of this boundedness,
we have the following convenient coordinate system.
Lemma 7.9. The smooth map (t, θ; w1 , . . . , wn , θ1 , . . . , θn ) endows V with a new smooth structure,
fiberwise compatible with the one inherited from A. This smooth structure has the following properties.
(a) For every i ∈ I, the functions vi , wi and θi are smooth.
(b) The map π : V −→ X is smooth.
(c) For every q ∈ [0, 1], the form ωqε is smooth, closed and fiberwise symplectic.
∂ ∂ ∂
(d) The coordinate vector field ∂t satisfies dwi ( ∂t ) = 0 and dθi ( ∂t ) = 0 for all i ∈ I.
S P
Proof. Let V = i∈I Vi be the covering of V by charts (18). Recall that on V we have θ = i θi
and vbi = vi for all i ∈ I, so the coordinate system (18) on Vi is equivalent to (g, θ, (vj )j∈Ii , (θj )j∈Ii ),
where Ii = I \ {i}. Since t = exp(1 − g−1 ) is a reparametrization of g, it is clear that the map
(t, θ, (vj )j∈Ii , (θj )j∈Ii ) endows V with a smooth structure, fiberwise compatible with the one from A.
We claim that it is equivalent to the one in the statement of the lemma. By definition (12) of vj ,
we have vj = twj−1 − η(wj ), so dvj = wj−1 dt − (twj−2 + η ′ (wj )) dwj . The coefficient twj−2 + η ′ (wj ) is
smooth P and positive since P wj > 0, so we can smoothly change each coordinate vj for wj . Eventually,
since j∈I wj = 1 and j∈I θj = θ, we can change wn+1 , θn+1 for wi , θi , as needed.
Thus the map (t, θ; w1 , . . . , wn , θ1 , . . . , θn ) gives a smooth structure on the whole V , equivalent to
(t, θ, (vj )j∈Ii , (θj )j∈Ii ) for each i ∈ I. This implies property (a). We now prove the remaining ones.
(b) It is enough to show that the coordinate functions ri cos(2πθi ), ri sin(2πθi ) are smooth for each
i ∈ I. To see this, recall that θi is smooth by (a); and ri = exp(−t−1 −1
i ) = exp(wi t ) is smooth because
wi is positive and smooth by (a).
(c) Part (a) and formula (44) show that the forms ω ♭ and ω ♯ are smooth and closed. By (b), the
form π ∗ ωX is smooth and closed, too, hence so is ωqε . Since the new smooth structure on V is fiberwise
compatible with the one in A, fiberwise non-degeneracy of ωqε follows from Proposition 4.3.
(d) For i 6= n + 1, this property
P follows from theP definition of the coordinate chart; for i = n + 1 we
use relations wn+1 = 1 − ni=1 wi and θn+1 = θ − ni=1 θi . 
Lemma 7.10. Let Q ⊆ C ∞ (V ) be the ideal of functions p such that for every k ∈ Z, tk p extends to a
smooth function on V , with respect to the smooth structure from Lemma 7.9. The following hold.
(a) The ideal Q is closed under taking partial derivatives.

(b) For every smooth form β on UX we have π ∗ β( ∂t ) ∈ Q.
(c) Let νt be the symplectic lift, with respect to the form ω0ε and the submersion (t, θ), of the unit
0

radial vector field on Clog . Then the vector field νt0 − ∂t has coefficients in Q.
(d) Let Ψ0 : ∂V × [0, ρ) −→ V be the flow of νt0 , where ∂V := V ∩ ∂A. Then for every integer k > 0
there is a smooth map ψk : ∂V × [0, ρ) −→ V such that in coordinates of Lemma 7.9 we have
Ψ0h (0, w; h) = (h, w) + hk · ψk (0, w; h),
where we write w := (θ, w1 , . . . , wn , θ1 , . . . , θn ) for the remaining coordinates of a point in V .
Proof. (a) Let ∂ be one of the partial derivatives. For every p ∈ Q and k ∈ Z we have tk · ∂p =
∂(tk p) − ktk−1 p · ∂t, which is a smooth function by definition of Q, so ∂p ∈ Q, as needed.
(b) Since Q is an ideal, it is sufficient to check (b) for the coordinate forms d(ri cos(2πθi )) and
d(ri sin(2πθi )). By (a) it is enough to check that ri ∈ Q. As in the proof of Lemma 7.9(b), we write
ri = exp(−wi t−1 ) and use the fact that wi is bounded from below by a positive number.
FIBRATIONS BY LAGRANGIAN TORI 32


(c) Since the vector field ∂t − νt0 is vertical, and the form ω0ε is fiberwise non-degenerate, it is enough

to prove that for every vertical vector field ν we have ω0ε ( ∂t − νt0 , ν) ∈ Q.
∂ ∂
Since νt0 is orthogonal to the fibers, we have ω0ε ( ∂t − νt0 , ν) = ω0ε ( ∂t , ν). Recall from formula (3) that
P
ω0 = π ωX + εω . Writing ωX = i βi ∧ βi for some smooth 1-forms βi , βi′ on UX , we see from (b)
ε ∗ ♭ ′
∂ ∂
that π ∗ ωX ( ∂t , ν) ∈ Q. In turn, formula (44) and Lemma 7.9(d) imply that ω ♭ ( ∂t , ν) = 0, as needed.

0
(d) By (c) we have νt − ∂t = t ν for some smooth vertical vector field ν. We treat tk ν as a time-
k

dependent
R h vector field on ∂V , and look at its integral curves. A point at time h has coordinates of the
form 0 sk a(s) ds for some smooth, time-dependent function a, namely a coordinate of ν. The result
follows since the first k − 1 derivatives with respect to h of such an integral vanish at h = 0. 
Consider a product V × [0, 1], where V is equipped with the smooth structure from Lemma 7.9,
and let q be the projection onto [0, 1]. The family of forms (ωqε ) yields a smooth form on V × [0, 1],
which is fiberwise symplectic with respect to the submersion (t, θ, q). Let νt be the symplectic lift of
e −→ V × [0, 1] be the flow of νt . Here
the radial vector field on Clog , and let Ψ : ∂V × [0, 1] × [0, δ)
δe = exp(−δ−1 ) > 0 is a small positive number such that the flow Ψh is defined for all times h ∈ [0, δ).
e
q e Note that for q > 0, this flow agrees with
Let Ψ be the restriction of Ψ to the slice ∂A × {q} × [0, δ).
q
Ψ introduced in Notation 6.10, once we identify the slice ∂A × {q} × [0, δ) e with ∂A × [0, δ).
e In turn,
0
the flow Ψ is described in Lemma 7.10(d).
Now, we fix a special admissible path γ, see Definition 6.8, and let [0, ρ) ∋ h 7→ (h, θ(h), q(h)) ∈
Clog × [0, 1] be its normal parametrization. We put Γ := ∂V × {(q(h), h) : h ∈ [0, ρ)}. Since γ is special,
Γ is a C 2 submanifold of ∂V × [0, 1] × [0, ρ).
Lemma 7.11. Define a smooth 1-form aj on ∂V × [0, 1] × [0, ρ) piecewise, as h1 ((Ψqh )∗ dwj − dwj ) for

h > 0 and as the Lie derivative Lνt ( ∂w j
) for h = 0; where h is the projection to the last coordinate.
γ −2 γ
Let aj = aj |Γ . Then the form h · aj extends to a continuous form on Γ.

Proof. We need to prove that aγj and its first derivative with respect to h vanishes at h = 0. Since
the path γ is admissible, the value h = 0 is attained precisely on the slice ∂V × {(0, 0)}. Thus
aγj |h=0 = aj |(h,q)=(0,0) = 0 by Lemma 7.10(d). By chain rule
∂aγj ∂aj ∂aj
= + · q ′ (0) = 0
∂h h=0 ∂h (h,q)=(0,0) ∂q (h,q)=(0,0)
∂aj
since ∂h |(h,q)=(0,0) = 0 by Lemma 7.10(d) and q ′ (0) = 0 by specialty of γ, see Definition 6.8. 

Recall that Lh ⊆ Xhγ is the fiber of the ωhγ -Lagrangian fibration (29) over the fixed point b ∈ ES .
q(h)
Thus in our notation, Lh × {q(h)} = Ψh (L0 × {q(h)}).
Lemma 7.12. Let Ωj := dwj − t2 ıdθj be a factor in formula (37). There is a continuous, complex
valued 1-form βj on Γ such that for every small h > 0 we have the equality of restrictions
q(h) ∗
(45) (Ψh ) (Ωj |Lh ) = h2 · (h · βj − ıdθj )|L0 .
Proof. By definition of the Lagrangian torus fibration (24), see Section 5.2, its fiber L0 is a level set
of (w1 , . . . , wn+1 ), so (dwj )|L0 = 0. Thus on L0 we have the equality of restrictions
q(h) ∗ q(h)
!
−2 q(h) ∗ −2 (Ψh ) dwj − dwj (Ψh )∗ dθj − dθj
h · (Ψh ) Ωj = h · h · −ı· − ıdθj =
h h
 
= h h−2 aγj + ıbγj − ıdθj .

where aγj is the smooth form defined in Lemma 7.11; and bγj is defined analogously, with θj instead of
wj . Lemma 7.11 shows that βj := h−2 aγj + ıbγj extends to a continuous form on Γ, as needed. 
V
Proof of Proposition 7.7. Let vol0 be a volume form on L0 given by ni=1 dθi . Taking a wedge
product in formula (45) and applying Lemma 7.1(c), we get the following equality on L0 :
q(h) ∗ q(h)
(Ψh ) (Ωnew |Lh ) = ±(c ◦ Ψh ) · h2n · ((−ı)n vol0 + h · β),
FIBRATIONS BY LAGRANGIAN TORI 33

for some continuous, complex-valued n-form β on L0 . Write β = b · vol0 , for some continuous, complex-
q(h)
valued function b. Let ch = c ◦ Ψh , so ch is a complex-valued function approaching a constant c0 as
h → 0, see Lemma 7.1(d). Now for small h > 0, the phase map ̟ from Definition 7.6 equals
±ch h2n ((−ı)n + hb) ch 1 − h · bın h→0
̟= = ±ın · · ̟,
e where ̟
e = −−−−→ 1,
|ch h2n ((−ı)n + hb)| |ch | |1 − h · bın |
c0
so ̟ extends to a continuous map L −→ S1 , whose value at L0 equals ±ın · |c0 | , as claimed. 

Proof of Corollary 7.8. We follow closely the argument of [HL82, §III.1], proving that the special
Lagrangians of phase ̟0 are Re(̟0 Ω)-calibrated, hence volume minimizing. The only difference in
our setting is that some bounds hold only up to asymptotically small errors.
We work in (Xhγ )gen
I for a fixed maximal essential face ∆I . Dividing the chosen holomorphic form Θ
by the constant (42), we can assume that the value of the phase map ̟ on L0 is 1. Lemma 7.5 gives
(46) √1
c+ · volgnew |Lh = Re Ωnew |Lh + β, where β → 0 as h → 0.
Now, we argue as in [HL82, Lemma 1.9]. Fix h > 0, x ∈ Lh , and let e1 , . . . , en be an orthonormal basis
of Tx Lh . Let H be another n-dimensional subspace of Tx Xhγ , let ε1 , . . . , εn be an orthonormal basis
of H, and let M be a linear automorphism of Tx Xhγ defined by M ei = εi , M Jei = Jεi . Note that M
is C-linear, so we can compute both the complex and real determinant of M , and they are related by
| detC M |2 = detR M . Using Lemma 7.5, we compute
(Re Ωnew (ε1 , . . . , εn ))2 6 |Ωnew (ε1 , . . . , εn )|2 = | detC M |2 · |Ωnew (e1 , . . . , en )|2 = detR M · c−1 −1
+ 6 c+
by Hadamard inequality and the fact that the vectors ε1 , . . . , εn have length one. Thus on n-dimensional
subspaces of Tx Xhγ we have an inequality
(47) Re Ωnew 6 √1
c+ · volgnew .
On Tx Lh , the above inequality becomes the “asymptotic” equality (46), cf. [HL82, Theorem 1.10].
Now fix Γ ∈ [Lh ]. The form Ω, hence R Re Ω is closed.
R Its rescaled version Re Ωnew is therefore
fiberwise closed, so Stokes theorem gives Lh Re Ωnew = Γ Re Ωnew . We conclude that
Z Z Z Z Z (47)
Z Z
1 g (46) 1 g
c+ · volnew = Re Ωnew + β= Re Ωnew + β 6 c+ · volnew + β.
√ √
Lh Lh Lh Γ Lh Γ Lh

The required convergence (43) follows since limh→0 β = 0 and limh→0 c+ is a positive constant. 
Proof of Theorem 1.2. Parts (a) and (b) are proved in Propositions 7.2 and 7.3. To see the first
statement of (c), take for ̟ ∈ S1 the inverse of the constant (42) and apply Proposition 7.7: note
that the path with |z(h)| = h is special by Example 6.9. The second statement of (c) is proved in
Corollary 7.8. Part (d) follows from (c) since the Maslov class is locally constant and the base ES of
the Lagrangian fibration lh is connected. 

References
+
[AB 09] P. Aspinwall, T. Bridgeland, et al., Dirichlet branes and mirror symmetry, Clay Mathematics Monographs,
vol. 4, American Mathematical Society, Providence, RI; Clay Mathematics Institute, Cambridge, MA, 2009.
[Aur09] D. Auroux, Special Lagrangian fibrations, wall-crossing, and mirror symmetry, Geometry, analysis, and al-
gebraic geometry: 40 years of the Journal of Differential Geometry, Surv. Differ. Geom., vol. 13, Int. Press,
Somerville, MA, 2009, pp. 1–47.
[BBI01] D. Burago, Y. Burago, and S. Ivanov, A course in metric geometry, Graduate Studies in Mathematics, vol. 33,
American Mathematical Society, Providence, RI, 2001.
[BJ17] S. Boucksom and M. Jonsson, Tropical and non-Archimedean limits of degenerating families of volume forms,
J. Éc. polytech. Math. 4 (2017), 87–139.
[BS73] A. Borel and J.-P. Serre, Corners and arithmetic groups, Comment. Math. Helv. 48 (1973), 436–491.
[CBM09] R. Castaño Bernard and D. Matessi, Lagrangian 3-torus fibrations, J. Differential Geom. 81 (2009), no. 3,
483–573. MR 2487600
[Cha14] K. Chan, The Strominger-Yau-Zaslow conjecture and its impact, Selected Expository Works of Shing-Tung
Yau with Commentary, Adv. Lect. Math., vol. 2, Int. Press, Somerville, MA, 2014, p. 1183–1208.
[EM21] J. Evans and M. Mauri, Constructing local models for Lagrangian torus fibrations, Ann. H. Lebesgue 4 (2021),
537–570.
[FdBP22] J. Fernández de Bobadilla and T. Pełka, Symplectic monodromy at radius zero and equimultiplicity of µ-
constant families, arXiv:2204.07007, 2022.
FIBRATIONS BY LAGRANGIAN TORI 34

[GHJ03] M. Gross, D. Huybrechts, and D. Joyce, Calabi-Yau manifolds and related geometries, Universitext, Springer-
Verlag, Berlin, 2003, Lectures from the Summer School held in Nordfjordeid, June 2001.
[Gro13] M. Gross, Mirror symmetry and the Strominger-Yau-Zaslow conjecture, Current developments in mathematics
2012, Int. Press, Somerville, MA, 2013, pp. 133–191.
[GS03] M. Gross and B. Siebert, Affine manifolds, log structures, and mirror symmetry, Turkish J. Math. 27 (2003),
no. 1, 33–60.
[GW00] M. Gross and P. Wilson, Large complex structure limits of K3 surfaces, J. Differential Geom. 55 (2000), no. 3,
475–546.
[HJ+ 22] J. Hultgren, M. Jonsson, et al., Tropical and non-Archimedean Monge-Ampère equations for a class of Calabi-
Yau hypersurfaces, arXiv:2208.13697, 2022.
[HL82] R. Harvey and B. Lawson, Calibrated geometries, Acta Math. 148 (1982), 47–157.
[HX13] C. Hacon and C. Xu, Existence of log canonical closures, Invent. Math. 192 (2013), no. 1, 161–195.
[Joy03] D. Joyce, Singularities of special Lagrangian fibrations and the SYZ conjecture, Comm. Anal. Geom. 11 (2003),
no. 5, 859–907.
[KK+ 73] G. Kempf, Knudsen, et al., Toroidal embeddings. I, Lecture Notes in Mathematics, Vol. 339, Springer-Verlag,
Berlin-New York, 1973.
[Kol13] J. Kollár, Singularities of the minimal model program, Cambridge Tracts in Mathematics, vol. 200, Cambridge
University Press, Cambridge, 2013, With a collaboration of S. Kovács.
[KS01] M. Kontsevich and Y. Soibelman, Homological mirror symmetry and torus fibrations, Symplectic geometry
and mirror symmetry (Seoul, 2000), World Sci. Publ., River Edge, NJ, 2001, pp. 203–263.
[Li20] Y. Li, Metric SYZ conjecture and non-archimedean geometry, arXiv:2007.01384, 2020.
[Li22a] , Strominger–Yau–Zaslow conjecture for Calabi–Yau hypersurfaces in the Fermat family, Acta Math.
229 (2022), no. 1, 1–53.
[Li22b] , Survey on the metric SYZ conjecture and non-Archimedean geometry, Internat. J. Modern Phys. A
37 (2022), no. 17, Paper No. 2230009, 44.
[MN15] M. Mustaţă and J. Nicaise, Weight functions on non-Archimedean analytic spaces and the Kontsevich-
Soibelman skeleton, Algebr. Geom. 2 (2015), no. 3, 365–404.
[MPS23] E. Mazzon and L. Pille-Schneider, Toric geometry and integral affine structures in non-archimedean mirror
symmetry, arXiv:2110.04223v2, 2023.
[NX16] J. Nicaise and C. Xu, The essential skeleton of a degeneration of algebraic varieties, Amer. J. Math. 138
(2016), no. 6, 1645–1667.
[NXY19] J. Nicaise, C. Xu, and T. Y. Yu, The non-archimedean SYZ fibration, Compos. Math. 155 (2019), no. 5,
953–972.
[Rua07] W.-D. Ruan, Generalized special Lagrangian torus fibration for Calabi-Yau hypersurfaces in toric varieties. I,
Commun. Contemp. Math. 9 (2007), no. 2, 201–216.
[RZ21] H. Ruddat and I. Zharkov, Compactifying torus fibrations over integral affine manifolds with singularities,
2019–20 MATRIX annals, MATRIX Book Ser., vol. 4, Springer, Cham, 2021, pp. 609–622.
[Sae22] O. Saeki, Reeb spaces of smooth functions on manifolds, Int. Math. Res. Not. IMRN (2022), no. 11, 8740–8768.
[SYZ96] A. Strominger, S. T. Yau, and E. Zaslow, Mirror symmetry is T-duality, Nuclear Physics B 479 (1996), no. 1,
243–259.
[Yau77] S. T. Yau, Calabi’s conjecture and some new results in algebraic geometry, Proc. Nat. Acad. Sci. U.S.A. 74
(1977), no. 5, 1798–1799.

Javier Fernández de Bobadilla: (1) IKERBASQUE, Basque Foundation for Science, Euskadi Plaza, 5,
48009 Bilbao, Basque Country, Spain; (2) BCAM, Basque Center for Applied Mathematics, Mazarredo
14, 48009 Bilbao, Basque Country, Spain; (3) Academic Colaborator at UPV/EHU.
Email address: jbobadilla@bcamath.org

Tomasz Pełka: Institute of Mathematics, University of Warsaw, Banacha 2, 02-097 Warsaw, Poland
Email address: tpelka@mimuw.edu.pl

You might also like