The K Ahler Geometry of Toric Manifolds: E-Mail Address: Apostolov - Vestislav@uqam - Ca
The K Ahler Geometry of Toric Manifolds: E-Mail Address: Apostolov - Vestislav@uqam - Ca
Vestislav Apostolov
E-mail address: apostolov.vestislav@uqam.ca
Contents
Introduction 5
Chapter 1. Delzant Theory 7
1. Hamiltonian group actions on symplectic manifolds 7
2. Hamiltonian actions of tori 10
3. The complex projective space 12
4. Toric symplectic manifolds from Delzant polytopes 17
5. Toric complex varieties from Delzant polytopes. Fans 19
6. Equivariant blow-up 20
7. Polarized projective toric varieties 22
8. Toric orbifolds 24
Chapter 2. Abreu–Guillemin Theory 27
1. The orbit space of a toric manifold 27
2. Toric Kähler metrics: local theory 27
3. The scalar curvature 30
4. Symplectic versus complex: the Legendre transform 32
5. The canonical toric Kähler metric 33
6. Toric Kähler metrics: compactification 34
Chapter 3. The Calabi Problem and Donaldson’s theory 37
1. The Calabi problem on a toric manifold 37
2. Donaldson–Futaki invariant 39
3. Uniqueness 40
4. K-stability 41
5. Toric test configurations 42
6. Uniform K-stability 45
7. Existence: an overview 48
Bibliography 59
3
Introduction
These lecture notes grew from the mini-courses I gave at the University
of Nantes in 2017 and at the CIRM in 2019. Their purpose is to present a
relatively self-contained and fast introduction to the theory of extremal Kähler
metrics, pioneered by E. Calabi in the 1950’s and developed over the years. The
framework of toric varieties, well-studied and used in both symplectic geometry
and algebraic geometry, offers a good testing ground and has been instrumental
to the general theory in many ways.
I am grateful to Alexandro Lepé, Lars Sektnan, Isaque Viza de Souza, and
Yicao Wang for their careful reading of previous versions of the manuscript and
for spotting and correcting errors, and suggesting many improvements. I have
benefited from numerous discussions with David Calderbank, Paul Gauduchon,
Éveline Legendre and Lars Sektnan on toric geometry.
5
CHAPTER 1
Delzant Theory
Example 1.2. Let M = S2 ⊂ R3 be the unit sphere endowed with the atlas of S2 given by the
stereographic projections from the north and the south poles, N = (0, 0, 1) and S = (0, 0, −1), respectively.
We denote by
2u 2v 1 − u2 − v 2 2
x= ,y = ,z = − , (u, v) ∈ R ,
1 + u2 + v 2 1 + u2 + v 2 1 + u2 + v 2
2ũ 2ṽ 1 − ũ2 − ṽ 2 2
x= 2 2
,y = 2 2
,z = , (ũ, ṽ) ∈ R
1 + ũ + ṽ 1 + ũ + ṽ 1 + ũ2 + ṽ 2
the corresponding chart patches on S2 . Recall that on S2 \ ({N } ∪ {S}), the transition between the coordiantes
u v 2
(u, v) and (ũ, ṽ) is given by the diffeomorphism ũ = u2 +v 2 , ṽ = u2 +v 2 of R \ {(0, 0)}. Then the 2-form
7
8 1. DELZANT THEORY
Example 1.4 (Hamiltonian flows). Let f be a smooth function on (M, ω). Using the non-degeneracy
of ω, we define a vector field by
−1
Xf := −ω (df ),
called the Hamiltonian vector field of f . Suppose that Xf is complete, i.e. its flow ϕt is defined for all t ∈ R
(this always holds if M is compact). Then, ρ(t) := ϕt defines a symplectic action of R on (M, ω). Indeed, we
have
d ∗ ∗
d
∗
(ϕ ω) = ϕs (ϕ ω)
dt |t=s t dt |t=0 t
∗
= ϕs LXf ω
∗
= ϕs dıXf + ıXf d)(ω)
∗
= ϕs d(−df ) = 0,
∗
ϕ0 (ω) = ω.
The previous two examples show that different groups (in particular Tm and
Rm ) can possibly give rise to “equivalent” symplectic actions, in the sense that
their images in Diff(M ) are the same. A way to normalize the situation is to
consider
Definition 1.2. A symplectic action of a Lie group G on (M, ω) is effective
if the homomorphism ρ : G → Diff(M ) has trivial kernel.
Thus, the symplectic action of Rm = {(t1 , . . . , tm )} on (R2m , ω0 ) defined in
Example 1.3 is not effective, whereas the action of Tm on the same manifold is.
In the sequel, we shall be interested in effective symplectic actions.
Let G be a Lie group acting smoothly on M . Denote by g = Lie(G) its Lie
algebra, i.e. g is the vector space of left-invariant vector fields on G. For any
ξ ∈ g, we denote by exp(tξ) the corresponding 1-parameter subgroup of G, and
by Xξ the vector field on M induced by the one parameter subgroup ρ(exp(tξ))
in Diff(M ), i.e.
d
Xξ (p) := ρ(exp(tξ))(p) .
dt |t=0
The vector field Xξ is called fundamental vector field of ξ ∈ g.
We recall that G also acts on itself by conjugation
ρ(g)(h) := ghg −1 , g ∈ G, h ∈ G,
fixing the unitary element e ∈ G and thus inducing a linear action on the vector
space Te G ∼= g, called the adjoint action
Ad : G → GL(g).
The induced linear action on the dual vector space Te∗ G ∼ = g∗ is given by
Ad∗ (g)(α) := α ◦ Ad(g −1 ) and is called the co-adjoint action. We are now
ready to give the definition of a hamiltonian action on (M, ω).
Definition 1.3. Let ρ : G → Diff(M ) be a smooth action of a Lie group G
on (M, ω). It is said to be hamiltonian if there exists a smooth map
µ : M → g∗ ,
called a momentum map, which satisfies the following two conditions.
(i) For any ξ ∈ g, the fundamental vector field Xξ satisfies
ω(Xξ , ·) = −dhµ, ξi,
where h·, ·i denotes the natural pairing between g and its dual g∗ .
1. HAMILTONIAN GROUP ACTIONS ON SYMPLECTIC MANIFOLDS 9
Remark 1.1. (a) Notice that the condition (i) of Definition 1.3 implies that
any fundamental vector field Xξ is hamiltonian (see Example 1.4) with respect
to the smooth function
(1) µξ (p) := hµ(p), ξi.
In particular, hamiltonian actions are symplectic.
(b) By a standard argument in Lie theory, the condition (ii) of Definition 1.3
can be equivalently expressed as
dµξ (Xη ) = −µ[η,ξ] .
For any two smooth functions f, h on (M, ω)
{f, h}ω := ω(Xf , Xh )
is the so-called Poisson bracket. Thus, the condition (ii) of Definition 1.3 is
equivalent to
(2) {µξ , µη }ω = −µ[ξ,η] .
An important example is the following
Example 1.5. Suppose G is a semi-simple compact Lie group with Lie algebra g, and T ⊂ G is a
maximal torus with Lie algebra t ⊂ g. Complexifying, we also have a semi-simple complex Lie group GC with
Lie algebra gc C := g ⊗ C, and a√Cartan sub-algebra
√
h := t ⊗ C. A specific example which one can bear in
mind is G = SU(n), T = {diag(e −1t1 , . . . , e −1tm ) : t1 + · · · + tm = 0}. Then GC = SL(n, C).
The general theory of semi-simple Lie algebras yields that gC admits a root decomposition
C
M C C
g =h⊕ (g−α ⊕ gα ),
α∈R+
∗
where R ⊂ h denotes the finite subspace of roots of gC and R = R+ ∪ (−R+ ) with R+ ∩ (−R+ ) = ∅ is the
choice of a subset R+ of positive roots. For α ∈ R, we have set
C C
gα := {u ∈ g : adx (u) = α(x)u, ∀x ∈ h}.
We denote by M an orbit in g for the adjoint action Ad : G → GL(g) and assume that M is not a point.
The general theory tells us that M is aways the orbit of a non-zero element x ∈ t, and that such x is unique
if we require moreover that −iα(x) ≥ 0 for all α ∈ R+ . If the unique x ∈ t determined as above also satisfies
−iα(x) > 0 for all α ∈ R+ , then the orbit M ∼ = G/T is principal for the adjoint action of G, and M is called
a flag manifold. In general, M ∼ = G/Gx where Gx is the subgroup of G fixing x (and whose Lie algebra is
gx = Ker(adx )). √
In the case G = SU(n), this corresponds to the basic fact that any hermitian matrix a ∈ −1su(n, C)
can be diagonalized by a conjugation with an element of SU(n). The diagonal matrix is uniquely determined
up to a permutation of its (real) eigenvalues, so we can normalize it by ordering the eigenvalues in decreasing
order. A hermitian matrix a determines a regular orbit for the action of G = SU(n) by conjugation precisely
when a has simple spectrum.
Each orbit M admits a natural symplectic form ω, called the Kirillov–Kostant–Souriau form, whose
definition we now recall. As the tangent space Ty M for any y ∈ M is generated by the fundamental vector
fields for the adjoint action of G, Ty M can be identified with the image my ⊂ g of the map
d
(3) u→ Adexp(tu) (y) = −ady (u).
dt |t=0
For u, v ∈ g, we denote by
hu, vi := tr(adu ◦ adv )
the Killing form of g. By assumption h·, ·i is negative definite (G being compact and semi-simple). We then
set
ωy (u, v) := −hy, [u, v]i,
where u, v ∈ Ty M ∼= Im(ady ). Then ω gives rise to a symplectic structure on M = G/Gx . The main fact is
Proposition 1.1 (Kirillov–Kostant–Souriau). The 2-form ω defines a symplectic structure on M =
G/Gx , the adjoint action of G on (M, ω) is hamiltonian with momentum map µ : M → g∗ identified with
the inclusion ı : M ⊂ g composed with the Killing form h·, ·i : g → g∗ .
10 1. DELZANT THEORY
Proof. By the argument in the proof of Lemma 1.1, the tangent space Tp O
of a principal orbit O = G(p) is ωp isotropic subspace of Tp M . Its dimension
(= dim G), therefore, is ≤ m.
Example 1.6. We return again to Example 1.3 and will now show that the symplectic action of Tm
on (R2m , ω0 ) is hamiltonian. We shall work on the dense open subset (C √
m 0
) = {(z1 , . . . , zm ) : zi 6= 0, ∀i =
1, . . . m} of Cm ∼
=R
2m
on which we introduce polar coordinates zi = ri e −1ϕi , i = 1, . . . , m. The symplectic
2-form ω0 then becomes
Xm
ω0 = ri dri ∧ dϕi ,
i=1
whereas the fundamental vector fields {X1 , . . . , Xm } associated to the standard basis of Rm = {(t1 , . . . , tm )} ∼
=
Lie(Tm ) are Xi = ∂ϕ∂
. It follows that ıXi ω0 = −ri dri = −d( 12 ri2 ). In other words, the smooth map
i ∗
µ : Cm → Rm ∼ m
= Lie(T ) defined by
1 2 2
(5) µ(z) := |z1 | , . . . , |zm |
2
is a momentum map for the action of Tm on (Cm )0 , and hence on Cm (by continuity). We also notice that
m
(6) Im(µ) = Cm := {(x1 , . . . , xm ) ∈ R : xi ≥ 0}.
1 2
Exercise 1.1. Consider the S -action on (S , ωS2 ) corresponding to the rotation around the z-axis of
R3 , see Example 1.2. Show that this is hamiltonian with momentum map given by the z-ccordinate.
2m+1
canonical round metric g S and the projection map (known as the Hopf fibra-
tion)
π : S2m+1 → CP m ,
2m+1
provides an example of a riemannian submersion between (S2m+1 , g S ) and
the (uniquely determined) riemannian metric g F S on CP m , introduced by the
property 2m+1
π ∗ gπ(z)
FS
= gzS |Hz ,
⊥
where at each point z ∈ S2m+1 , Hz = Ker(π∗ )z . We use a similar con-
struction in order to endow CP m with a symplectic form ω, using the pointwise
isomorphism (π∗ )z : Hz ∼ = Tπ(z) CP m and the restriction of ω0 to the subspace
Hz . As all tensor fields used in this construction are Tm+1 -invariant, and the
S1 -action√
(7) is obtained
√
by restricting the action of Tm+1 to its S1 subgroup
N = (e −1t ,...,e −1t ) ⊂ Tm+1 , there is a natural induced Tm = Tm+1 /N -
action on CP m = S2m+1 /N . With respect to this data, (CP m , ω F S , Tm ) is an
example of a symplectic toric manifold. In fact, this is just a very particular case
of a general fact in symplectic geometry, known as symplectic reduction, which
we shall present below.
To this end, let us summarize the construction of CP m : by Example 1.6,
(C m+1 , ω0 ) is a symplectic manifold endowed with the hamiltonian action on
Tm+1 with momentum map
1
µTm+1 (z) = (|z0 |2 , . . . , |zm |2 ).
2
N ⊂ Tm+1 is a circle subgroup and the momentum map for the action of N is
therefore
m
1X 2
µN (z) = |zi | .
2
i=0
Thus, 1
S2m+1 = µ−1
N
2
is the level set of the momentum map µN , and furthermore, N acts freely on
this level set. We are thus in the situation of the following fundamental result.
Proposition 1.2 (Marsden–Weinstein [41], Meyers [43]). Let N ⊂ G
be a closed normal subgroup of a compact group G acting in a hamiltonian way
on a symplectic manifold (M̃ , ω̃). Let µG : M̃ → g∗ be a momentum map for
the G-action, and µN : M̃ → n∗ the corresponding momentum map for the N -
action, obtained from µG by composing with the natural projection ı∗ : g∗ → n∗
(which is adjoint to the inclusion of the Lie algebras ı : n ⊂ g).
Suppose further that c̃ ∈ g∗ is a fixed point for the co-adjoint action of G on
g , such that ı∗ (c̃) = c is a regular point for µN , and that the action of N on
∗
−1
µN (c) is free.
Then, the N -invariant symplectic 2-form
ω̃, restricted to µ−1
N (c) defines a
symplectic form on the manifold M := µ−1N (c) /N , and the natural action of
G/N on M is hamiltonian with moment map, viewed as an N -invariant function
on µ−1
N (c), given by
µ = µG − c̃.
3. THE COMPLEX PROJECTIVE SPACE 15
(3) Show that the complex structure J induced on M = S2m+1 /S1 makes (M, J) biholomorphic to
CP m = (C \ {0})/C∗ endowed with its atlas of affine charts.
Hint. Show that the Dolbeault cohomology H 2,0 (M, J) = {0} and use Hodge decomposition theorem to
conclude that M admits a Kähler structure (ω 0 , J) with [ω 0 ] ∈ H 2 (M, Q). The conclusion then follows from
the Kodaira embedding theorem, see e.g. [26].
We notice that in order to construct M∆C we did not use the whole data from
6. Equivariant blow-up
We explain now how to blow-up equivariantly a fixed point of the action
of T on the complex manifold M∆ C constructed in Section 5. Recall that M C
∆
is endowed with a T-equivariant atlas of affine charts Cm v , parametrized by the
the vertices v of ∆, such that the action T on each chart Cm v is the standard
action of Tm on Cm , as described in Example 1.3. We notice that Cm v = C
m
is the complex manifold associated to (the fan of) the (unbounded) standard
cone Cm ⊂ Rm , and the standard lattice Zm ⊂ (Rm )∗ , via the construction in
Section 5: indeed, Cm has a unique vertex at the origin, and the inward primitive
6. EQUIVARIANT BLOW-UP 21
normals of the adjacent facets form the standard basis of (Rm )∗ , which define a
single chart Cm
0 .
We now blow-up the origin of Cm and obtain as a resulting manifold C m =
d
0
OCP m−1 (−1) → CP m−1 , the total space of the tautological bundle over CP m−1 .
The blow-down map b : OCP m−1 (−1) → Cm 0 is explicitly given by
b [w1 , . . . , wm ], ζ(w1 , . . . , wm ) = (ζw1 , . . . , ζwm ),
where [w1 , . . . , wm ] are homogeneous coordinates on CP m−1 , and ζ is a fibre-wise
coordinate of the tautological bundle (with respect to the generator (w1 , . . . , wm )).
The inverse map, defined on Cm 0 \ {0}, is
b−1 (z1 , . . . , zm ) = ([z0 , . . . , zm ], (z1 , . . . , zm )),
showing that the action of Tm on Cm 0 lifts to an action of T
m on C
d m , given in
0
the above coordinates by
√ √
(e −1t1 , . . . , e −1tm ) · [w1 , . . . , wm ]; ζ(w1 , . . . , wm ) =
(18) √ √ √ √
[e −1t1 w1 , . . . , e −1tm wm ], ζ(e −1t1 w1 , . . . , e −1tm wm ) ,
b 2 and C2 .
Figure 1. The polytopes of C 0
Figure 2
.
Exercise 1.7. Let MA(∆) ⊂ CP N be a toric polarized projective variety corresponding to a lattice
Delzant polytope ∆ ⊂ Rm with respect to the lattice Zm ⊂ Rm . Using that the sections of O(1) are identified
with linear functions in homogeneous coordinates [w0 , . . . , wN ] on CP N , show that the (C∗ )m -action on
MA(∆) defines a (C∗ )m -action on the vector space H 0 (MA(∆) , L) of holomorphic sections of L. Furthermore,
show that here is a basis {s0 , . . . , sN } of H 0 (MA(∆) , L), parametrized by the lattice points {λ(0) , . . . , λ(N ) }
in A(∆), such that (C∗ )m acts on sj by
(j) (j)
λ λ
(z1 , . . . , zm ) · sj = (z1 1 · · · zmm )sj .
8. Toric orbifolds
We briefly discuss here what happens with the Delzant construction reviewed
in Section 4 when one starts with a rational Delzant triple (∆, L, Λ), i.e satisfying
the weaker conditions (i)-(ii)-(iii)’ of Definition 1.5. There are two points in the
construction which deserve a special attention.
The first point is Claim 1. Consider for example the labelledPm Delzant simplex
1
(∆m , L) with L = {Lj (x) = xj , j = 1, . . . m, Lm+1 (x) = − j=1 xj + 2 }, corre-
sponding to (CP m , ω F S ), and change the standard lattice Λ = Zm ⊂ (Rm )∗ with
the lattice Λ0 = 12 Λ. The condition (iii)’ then holds for the triple (∆m , L, Λ0 ).
However, in this case,
N 0 = Ker(τ ) = S1 × Zm2 ,
m
where Z2 := (±1, ±1, . . . , ±1) ⊂ T m+1 . In particular, N 0 is no longer connected.
Following the construction further, the quotient space will be
M 0 = S2m+1 /N 0 = CP m /Zm
2 .
Such a space is an example of an orbifold.
Definition 1.11. An orbifold chart on a topological space M is a triple
(Û , Γ, ϕ) where:
• Û ⊂ Rn is an open subset;
• Γ ⊂ GL(n) is a finite subgroup acting on Û
• ϕ : Û /Γ → M is a homeomorphism between Û /Γ and an open subset
V ⊂ M.
We denote by ϕ̂ : Û → M the induced Γ-invariant map.
An embedding between two orbifold charts (Û1 , Γ1 , ϕ̂1 ) and (Û2 , Γ2 , ϕ̂2 ) is a
smooth embedding λ : Û1 → Û2 such that ϕ̂1 = ϕ̂2 ◦ λ.
Two orbifold charts (Û1 , Γ1 , ϕ̂1 ) and (Û2 , Γ2 , ϕ̂2 ) are compatible if there exists
an open subset V ⊂ V1 ∩ V2 (where we have set Vi = ϕ̂i (Ûi )) and an orbifold
chart (Û , Γ, ϕ̂) with ϕ̂(Û ) = V , and two embeddings λi : Û → Ûi , i = 1, 2.
An orbifold M of (real) dimension n is a Hausdorff paracompact topo-
logical space, endowed with an open covering of compatible orbifold charts
{(Ûi , Γi , ϕ̂i )}i∈I . A smooth function f : M → R on an orbifold M is defined
by the property that on each orbifold chart (Û , Γ, ϕ̂), it pulls-back by ϕ̂ to a
Γ-invariant smooth function on Û . One can define smooth tensors on M is a
similar way.
The above phenomenon can be remedy by taking Λmin = spanZ {u1 , . . . ud },
which is the minimal (under the inclusion) lattice for which the condition (iii)’
is satisfied for the given labelling L. The subgroup Nmin will be then a torus
and for any other choice of lattice Λ, the produced quotient space S/N will be
a quotient of S/Nmin by the finite abelian group Λ/Λmin . Geometrically, this
8. TORIC ORBIFOLDS 25
Abreu–Guillemin Theory
(Rm )∗ by using the dual basis {e∗1 , . . . , e∗m } and write x = (x1 , . . . , xm ) for the
elements of t∗ . As explained in the previous section, on M 0 , K1 , . . . , Km are
functionally independent, i.e. for each point p ∈ M 0 , (K1 ∧ · · · ∧ Km )(p) 6= 0.
We shall also identify the coordinate function xi = hx, ei i with the momentum
function hµ, ei i on M , i.e. we write
(19) ıKi ω = −dxi .
Let (g, J) be a T-invariant ω-compatible Kähler structure on M . Letting
Hij = g(Ki , Kj )
we get a T-invariant smooth function on M , which will tacitly identify with a
smooth function Hij (x) on ∆, see Fact 2 above. We denote by Hx = (Hij (x))
the corresponding symmetric matrix with smooth entries over ∆. In a more
intrinsic language, we regard H as a smooth function over ∆ with values in S 2 t∗
by letting
Hx (ξ1 , ξ2 ) := gp (Xξ1 , Xξ2 )
for any ξ1 , ξ2 ∈ t and any p ∈ µ−1 (x).
On ∆0 , H is positive definite and we denote by G := H−1 the inverse matrix;
equivalently, G is a smooth S 2 t-valued function on ∆0 .
Notice that, by (19),
(20) dxi (JKj ) = −ω(Ki , JKj ) = − − g(JKi , JKj ) = −g(Ki , Kj ) = −Hij (x).
We now consider the vector fields {K1 , . . . , Km , JK1 , . . . , JKm }. They form
a basis of Tp M at each p ∈ M 0 (because {K1 , . . . , Km } span an m-dimensional
space and {JK1 , . . . , JKm } span its g-orthogonal complement) and satisfy
(21) [Ki , Kj ] = [Ki , JKj ] = [JKi , JKj ] = 0,
where for the second identity we used that T preserves J, whereas for the third
identity we used the integrability of J, see (8). We now denote by
{θ1 , . . . , θm , Jθ1 , . . . , Jθm }
the dual basis of T ∗ M 0 , corresponding to
{K1 , . . . , Km , JK1 , . . . , JKm },
where for a 1-form θ we set Jθ(X) = −θ(JX), for any vector field X. The
commuting relation (21) is equivalent to
(22) dθi = 0 = d(Jθi ), i = 1, . . . m.
As the 1-forms Jθi satisfy
ıKj Jθi = 0, LKj Jθi = 0,
in terms of the T bundle structure µ : M 0 → ∆0, each Jθi is basic, i.e. Jθi =
µ∗ (αi ) of a closed 1-form αi on ∆0 . As ∆0 is contractible, we can write
(23) Jθi = −dyi
for some smooth function yi (x), defined on ∆0 up to an additive constant (as
usual, we omit the pull-back by µ in the notation). Furthermore, by (20), we
2. TORIC KÄHLER METRICS: LOCAL THEORY 29
find
m
X
−Jθi = dyi = Gij (x)dxj ,
j=1
(24) m
X
Jdxi = Hij (x)θj .
j=1
The 1-forms {θ1 , . . . , θm } on the other hand define a 1-form with values
t = Lie(T) by letting
m
X
(25) θ= θi ⊗ ei ,
i=1
we have by (24)
m
X m
X
dβ = dyi ∧ dxi = Gij dxi ∧ dxj = 0.
i=1 i,j=1
30 2. ABREU–GUILLEMIN THEORY
are closed. As θi are closed too (by the assumption √ that θ is flat), we get
a basis of Λ 1,0 (M 0 , J) of closed (1, 0)-forms −Jθ + −1θi , so writing
√ √ i √ locally
−Jθi + −1θi = dyi + −1dti we get holomorphic coordinates yi + −1ti for
J, i.e. J is integrable.
We have seen in Chapter 1 that the action of T ∼ = (S1 )m on (M, J) extends
to an (effective) holomorphic action of the complex torus TC ∼ = (C∗ )m . Fixing
a point p0 ∈ M , we can identify M with the orbit T (p0 ) ∼
0 0 C
= (C∗ )m . Using the
polar coordinates (ri , t0i ) on each C∗ , this identification gives rise the so-called
angular coordinates
t = (t01 , . . . , t0m ) : M 0 → t/2πΛ.
If another reference point is chosen, t varies by an additive constant in t. Writing
θ = dt, we have
Definition 2.1. For a fixed ω-compatible, T-invariant complex structure J
on (M, ω, T) (corresponding to a symplectic potential P u(x) on ∆0 ), and a base
point p0 ∈ M (giving rise to angular coordinates t = m
0
i=1 ti ei with respect to
a basis e = {e1 , . . . , em } of t), the functions {x1 , . . . , xm ; t1 , . . . , tm } on ∆0 × T
are called momentum-angle coordinates associated to (g, J).
We then compute
1
ρg = − ddc log det H
2
(30) 1X
=− Hij,ik dxk ∧ θj .
2
i,j,k
P
The formula for sg follows from (29), by using that ω = i dxi ∧ θi and (30).
32 2. ABREU–GUILLEMIN THEORY
ω = ddc φ,
where dc φ = Jdφ.
so that
m
X m
X m
X
dφ = d(xi u,i ) − u,i dxi = xi u,ij dxj = xi Gij dxj ,
i=1 i,j=1 i,j=1
5. THE CANONICAL TORIC KÄHLER METRIC 33
m
X
c
dd φ = dJdφ = d xi Gij Jdxj
i,j=1
Xm
=d xi Gij Hjk θk
i,j,k=1
m
X
= dxi ∧ θi = ω
i=1
To see this, recall that the flat metric g̃0 on R2d can be written in polar coordi-
nates (ri , ti ), i = 1, . . . , d as
d
X
g̃0 = dri2 + ri2 dt2i .
i=1
We have already observed in Example 1.6 that the momentum coordinates are
x̃i = ri2 /2.
Theorem 2.1 (Guillemin [28]). The symplectic potential of the induced
Kähler structure (g0 , J0 ) via the Delzant construction is
d
1X
u0 (x) = Lj log Lj .
2
j=1
Exercise 2.2 (Abreu’s boundary conditions [2]). Show that (33) and (34) are equivalent to
(35) G − G0 is smooth on ∆,
−1
(36) G0 G is smooth and nondegenerate on ∆.
Exercise 2.3. [2] Show that the conditions (35)-(36) are equivalent to
d
1X
(37) u− Lj log Lj is smooth on ∆.
2 j=1
d
Y
(38) det(Hess(u)) × Lj (x) is positive and smooth on ∆.
j=1
There is, however, a subtle point in the above theory (which I believe is often
neglected in the literature): in order to apply the sufficient conditions (33)-(34)
or (35)-(36), we need to use the angular coordinates defined by the initial metric
g0 . This is the main difficulty to show that the conditions are also necessary.
The following criterion is established in [3].
Proposition 2.1. A smooth, positive definite S 2 (t∗ )-valued function H on
∆0 corresponds via (27) to a T-invariant ω-compatible almost-Kähler structure
on M if and only if H satisfies the following conditions.
• [smoothness] H has a smooth extension as a S 2 (t∗ )-valued function on
∆;
• [boundary conditions] If x belongs to a co-dimension one face Fj ⊂ ∆
with normal uj ∈ t, then
Hx (uj , ·) = 0, dHx (uj , uj ) = 2uj .
• [positivity] Let F 0 ⊂ ∆ be the interior of a face of ∆ and denote by
tF ⊂ t the subspace spanned by the normals of all labels vanishing on F 0 .
Then, the restriction of H to F 0 , viewed as a S 2 (t0F )-valued function
for t0F = ann(tF ) ⊂ t∗ is positive definite.
Exercise 2.4. State and prove Proposition 2.1 in the case m = 1.
Theorem 2.2 (Abreu [1, 2]). The space of T-invariant, ω-compatible Kähler
structures (g, J) on (M, ω, T), modulo the action of the group of T-equivariant
symplectomorphisms, is parametrized by the space S(∆, L).
One can further refine Theorem 2.2.
Proposition 2.2 (Donaldson [22]). The functional space S(∆, L) of sym-
plectic potentials of globally defined T-invariant, ω-compatible Kähler metrics on
(M, ω, T) can be equivalently defined as the sub-space of the space C(∆) of convex
continious functions on ∆, such that
• [convexity] The restriction of u to the interior of any face of ∆ (includ-
ing ∆0 ) is a smooth strictly convex function.
• [asymptotic behaviour] u − 21 dj=1 Lj log Lj is smooth on ∆.
P
CHAPTER 3
Proof. The proof is elementary and uses integration by parts: recall that
for any smooth t∗ -valued function V = (V1 , . . . , Vm ) on t∗ , Stokes theorem gives
Z X m d Z
X
(42) Vj,j dv = − hV, dLk idσ,
∆ j=1 k=1 Fk
where we have used the convention (39) for dσ. We shall use the identity
Xm m
X m
X
(43) ϕ,ij Hij = ϕHij,ij − Vj,j ,
i,j=1 i,j=1 j=1
where
m
X m
X
(44) Vj := ϕ Hij,i − ϕ,i Hij .
i=1 i=1
It follows by (43) and (42) that
Z X m Z m
X d Z
X
(45) ϕ,ij Hij dv = ϕHij,ij + hV, dLk idσ.
∆ i,j=1 ∆ i,j=1 k=1 Fk
the condition (40) gives rise to a linear system with positive-definite symmetric
matrix
Z Xm Z Z
a0 xi dv + aj xj xi dv = 2 xi dσ
∆ j=1 ∆ ∂∆
(47) Z m Z Z
X
a0 dv + aj xj dv =2 dσ,
∆ j=1 ∆ ∂∆
We are going to show that s(x) = s(∆,L) . Notice that to this end we do not
assume that H satisfies the positivity conditions of Proposition 2.1 nor that
H = Hess(u)−1 for some u ∈ S(∆, L). Indeed, by Lemma 3.2 applied to an
affine-linear function ϕ = f , we get that s(x) satisfies the defining property
(40).
Definition 3.2. Let (∆, L) be a labelled compact convex simple polytope
in Rm , S(∆, L) the space of strictly convex smooth function on ∆0 satisfying the
conditions of Proposition 2.1 and s(∆,L) the extremal affine function of (∆, L).
Then the non-linear PDE
m
X
(49) s(u) := − (u,ij ),ij = s(∆,L)
i,j=1
2. Donaldson–Futaki invariant
We now introduce an obstruction to the existence of solutions of (49), due
to Donaldson [21].
40 3. THE CALABI PROBLEM AND DONALDSON’S THEORY
where we have used the convexity of ϕ for the inequality. Furthermore, as H > 0
on ∆0 , the inequality is strict unless ϕ,ij = 0, i.e. ϕ is affine-linear.
Exercise 3.1 (Donaldson [21]). Show that the statement of Proposition 3.2 holds true for continuous
convex functions ϕ which are smooth on the interior ∆0 and are not affine-linear.
3. Uniqueness
We show now that the solution of (49) is unique up to the addition of affine-
linear functions.
Theorem 3.1 (D. Guan [27]). Any two solutions u1 , u2 ∈ S(∆, L) of (49)
differ by an affine-linear function. In particular, on a compact toric Kähler
manifold or orbifold (M, ω, T), there exists at most one, up to a T-equivariant
isometry, ω-compatible T-invariant extremal Kähler metric (g, J).
Proof. Consider the functional
Z
(51) E(∆,L) (u) := F(∆,L) (u) − log det Hess(u) − log det Hess(u0 ) dv,
∆
referred to as the relative K-energy of (∆, L) with respect to T. It is well-defined
for elements u ∈ S(∆, L) by virtue of the equivalent boundary conditions (35)-
(36). Using the formula d log det A = tr A−1 dA for any non-degenerate matrix
A and Lemma 3.2, one computes the first variation of E(∆,L) at u in the direction
of u̇
Z X m
dE(∆,L) (u̇) =F(∆,L) (u̇) − Hiju u̇,ij dv
u ∆ i,j=1
Z h m
X i
= − Hiju − s(∆,L) u̇ dv,
∆ ,ij
i,j=1
showing that the critical points of E(∆,L) are precisely the solutions of (49).
Furthermore, using dA−1 = −A−1 dAA−1 , the second variation of E(∆,L) at u
in the directions of u̇ and v̇ is computed to be
Z
2
d E(∆,L) (u̇, v̇) = tr Hess(u) Hess(u̇) Hess(u) Hess(v̇) dv,
u ∆
4. K-STABILITY 41
showing that E(∆,L) is convex. In fact, as Hess(u) is positive definite and Hess(u̇)
is symmetric, the vanishing d2 E∆,L,f (u̇, u̇) = 0 is equivalent to u̇ being affine.
u
It follows from (35)-(36) that for any two elements u1 , u2 ∈ S(∆, L), u(t) =
tu1 + (1 − t)u2 , t ∈ [0, 1] is a curve in S(∆, L) with tangent vector u̇ = u1 − u2 .
Using the convexity of E(∆,L) , it follows that that if u1 and u2 are two solutions
of (49) (equivalently, u1 and u2 are critical points of E(∆,L) ), then u1 − u2 must
be affine.
Corollary 3.1. If (49) admits a solution u∗ ∈ S(∆, L), then the relative
K-energy E(∆,L) atteints its minimum at u∗ .
Proof. The arguments in the proof of Theorem 3.1 show that E(∆,L) is
convex on S(∆, L). The solution u∗ being a critical point of E(∆,L) , it is therefore
a global minima.
4. K-stability
Definition 3.4 (Toric K-stability). We say that a labelled compact con-
vex simple polytope (∆, L) in Rm is K-semistable if
∆00 )
Z Z
F(∆,L) (ϕ) = 2 ϕdσ − s(∆,L) ϕdv
∂∆ ∆
Z Z
=2 ϕdσ − s(∆,L) ϕdv.
∂∆0 ∆0
We now use (45) and (46) 0
P over ∆ with Hij satisfying the conditions of Propo-
sition 2.1 on ∆ and − ij Hij,ij = s(∆,L) . Noting that ϕ is affine over ∆0 , i.e.
ϕ,ij = 0 we have
Z Z X m
− s(∆,L) ϕdv = ϕ Hij,ij dv
∆0 ∆0 i,j=1
Z
=−2 ϕdσ
∂∆0
Z Xm
− ϕ hdH(dL, ei ), e∗i i dσ
F i=1
Z
+ H(dL, dL)dσ.
F
As ϕ vanishes on F , the term at the third line is zero, so that
Z
(53) F(∆,L) (ϕ) = H(dL, dL)dσ > 0,
F
as H is positive definite over ∆0 .
Thus motivated, the central conjecture is
Conjecture 3.1 (Donaldson [21]). (49) admits a solution in S(∆, L) if
and only if (∆, L) is K-stable.
gives rise to a smooth toric symplectic 2(m + 1)-dimensional manifold (M, Ω),
but the discussion below holds in the general orbifold case.
We denote by Tm+1 = Tm × S1(m+1) the corresponding torus, with Tm identi-
fied with the torus acting on M . Notice that ∆ is a facet of P, whose pre-image
is a smooth submanifold M̃ ⊂ M. The Delzant construction identifies the sta-
bilizer of points in M̃ 0 with the circle subgroup S1(m+1) ⊂ Tm+1 , so that with
respect to the induced action of Tm+1 /S1 ∼
= Tm , (M̃ , Ω| , is equivariantly
(m+1) M̃
isomorphic to (M, ω) by Delzant’s theorem. We shall thus assume without loss
of generality that M = M̃ and ω = Ω|M̃ .
Let us now choose an Ω-compatible Tm+1 -invariant complex structure J on
M, which induces a Tm -invariant ω-compatible complex structure J on (M, ω).
Donaldson [21, Proposition 4.2.1] shows that with respect to the C∗ -action ρ :
C∗ → Aut(M) induced by S1(m+1) , the complex (m + 1)-dimensional manifold
(M, Ω, ρ) is an example of a Kähler test configuration associated to the Kähler
manifold (M, ω), meaning that the following are satisfied:
Definition 3.5 (Kähler test configurations). A smooth Kȧhler test
configuration associated to a compact complex m-dimensional Kähler manifold
(M, ω) is a compact complex (m + 1)-dimensional Kȧhler manifold (M, Ω) en-
dowed with a C∗ -action ρ, such that
• there is a surjective holomorphic map π : M → CP 1 such that each
fibre Mτ := π −1 (τ ) ∼
= M for τ ∈ CP 1 \ {0};
• the C -action ρ on M is equivariant with respect to the standard C∗ -
∗
s(∆,L) the smooth function on M obtained by pulling back the extremal affine
linear function s(∆,L) associated to (M, ω). We then have
Lemma 3.3. In the above setting, the Donaldson–Futaki invariant (50) as-
sociated to the convex PL function f is given by
Z Ωm+1
m+1
(2π) F(∆,L) (f ) = − Scal(Ω) − s(∆,L)
M (m + 1)!
(55)
ωm
Z
+ (8π) .
M m!
Ωm+1
Z Z
s(∆,L) = (2π)m+1 s(∆,L) (x)dx ∧ dx0
M (m + 1)!
(57) ZP
= (2π)m+1 s(∆,L) (x)(R − f (x)dx.
∆
Using Lemma 2.3 for Scal(Ω) and (41) with ϕ = 1 on P, we compute
Ωm+1
Z Z
1
Scal(Ω) =2 dσP
(2π)m+1 M (m + 1)! ∂P
Z Z Z
(58) =2 dx + dσ(R−f )(∆) + (R − f )dσ∂∆
∆ (R−f )(∆) ∂∆
Z Z
=2 2 dx + (R − f )dσ∂∆ ,
∆ ∂∆
where for passing from the second line to the third we have used that dσ(R−f )(∆)
is defined by the equality
df ∧ dσ(R−f )(∆) = dx ∧ dx0 .
Lemma 3.3 now follows easily by combining (56), (57) and (58).
Remark 3.1. (a) In the special case when s(∆,L) = λ is a constant, which
means that (M, ω, J) has vanishing Futaki invariant, one can further specify
R the
dσ
expression in the rhs of (55) as follows. By (40) with f = 1 we obtain λ = 2 R∂∆
dv
R R ∆
whereas (41) yields ∆ s(u) = 2 ∂∆ dσ. It thus follows that
R
s(u)dv
λ = ∆R
dv
R ∆
Scal(ω)ω m
= MR m
Mω
c1 (M ) · [ω]m−1 [M ]
= 4πm ,
[ω]m [M ]
6. UNIFORM K-STABILITY 45
where Scal(ω) stands for the induced toric Kähler metric on M and u for its sym-
plectic potential. Substituting in (55), we obtain the following co-homological
expression for the Donaldson–Futaki invariant
F(M, Ω) := (2π)m F(∆,L) (f )
h c (M) · [Ω]m [M] c (M ) · [ω]m−1 (M ) Vol(M, Ω) i
1 1
= −2 −
m! (m − 1)! Vol(M, ω)
+ 4Vol(M, ω).
This formula makes sense for any Kähler test configuration (see Definition 3.5)
and only depends upon the deRham classes [Ω] on M and [ω] on M . By Corol-
lary 4.1, F(M, Ω) ≥ 0 on any toric Kähler test configuration (M, Ω) associ-
ated to (M, ω), with equality if and only if M corresponds to a single rational
affine linear function f . This is the original notion of K-stability going back
to Tian [53]. In this integral form, the Futaki invariant of a test configuration
was first used by Odaka [45, 46] and Wang [55] to study (possibly singular)
polarized projective test configurations.
(b) It turns out that the expression at the rhs of (55) makes sense for any
T-invariant Kähler test configuration associated to a Kähler manifold (M, ω)
endowed with a maximal compact torus T in its reduced group of complex au-
tomorphisms, and it merely depends upon the deRham classes [ω] and [Ω] and
the momentum image ∆ of M for that action of T. This leads to the notion of a
T-relative Donaldson–Futaki invariant F T (M, [Ω]) of a compatible test configu-
ration. This point of view has been taken and developed in [18, 17, 19, 20, 37]
for a general Kähler manifold, where an extension of Corollary 4.1 is also ob-
tained.
6. Uniform K-stability
Let C(∆) denote the set of continuous convex functions on ∆ (continuity
follows from convexity on the interior of ∆), C∞ (∆) ⊂ C(∆) the subset of those
functions which are smooth on the interior ∆0 , and S(∆, L) ⊂ C∞ (∆) the set
of symplectic potentials. Note that, by virtue of Proposition 2.2, if u ∈ S(∆, L)
and f ∈ C∞ (∆), with f smooth on all of ∆, then u + f ∈ S(∆, L). Conversely,
the difference of any two functions in S(∆, L) is a function in C∞ (∆) which is
smooth on ∆.
The affine linear functions act on C(∆) and C∞ (∆) by translation. Let
C ∗ (∆) be a slice for the action on C(∆) which is closed under positive linear
combinations, and C∞ ∗ (∆) the induced slice in C (∆). Then any f in C(∆) can
∞
be written uniquely as f = π(f )+g, where g is affine linear and π(f ) ∈ C ∗ (∆) for
a linear projection π. Functions in C ∗ (∆) are sometimes said to be normalized.
Example 3.1. [21] If x0 ∈ ∆0 is a fixed interior point, a slice as above is given by
n o
∗
C (∆) := f ∈ C(∆) : f (x) ≥ f (x0 ) = 0 .
Example 3.2. [49] Another natural choice for the slice is given by
n Z o
∗
C (∆) := f ∈ C(∆) : f (x)g(x) dv = 0 ∀ g affine linear .
∆
R Exercise 3.2. Donaldson considers in [21] the slice C ∗ (∆) of Example 3.1 with the norm ||f ||∗ :=
∂∆
f dσ. Show that || · ||∗ is tamed. Hint. Use that f is a positive convex function.
Remark 3.2. (a) For any tamed norm || · ||, both the spaces of PL convex
functions and smooth convex functions on the whole of ∆ are dense in C ∗ (∆).
(b) With respect to a tamed norm || · ||, F(∆,L) is well defined and continuous
on C ∗ (∆).
Proof. As F(∆,L) (f ) and and E(∆,L) (u) are unchanged by the addition of
an affine linear function, it suffices to prove the equivalence for normalized f
and u.
(i)⇒(ii) For any bounded function a on ∆, one can define a modified Futaki
invariant
R Fa by replacing the second integral (over ∆) in the formula (50) by
− ∆ a(x)f (x) dv. Similarly, one can define a modified K-energy Ea using Fa
instead of F(∆,L) in the formula (51). Donaldson [21] shows that Ea (which is
introduced on the space S(∆, L)) can be extended to C∞ (∆) (in fact on a slightly
larger space) taking values in (−∞, +∞].
For any bounded functions a, b, there is a constant C = Ca,b > 0 with
|Fa (f ) − Fb (f )| ≤ C||f || for all f ∈ C ∗ (∆), because || · || bounds the L1 norm on
C ∗ (∆) by assumption. Let us write C = (1+k)C −kC for an arbitrary k ≥ 0 and
take b to be the extremal affine-linear function b = s(∆,L) , so that Fb = F(∆,L) ,
whereas we can take a = s(u0 ) be the scalar curvature of the canonical potential
u0 ∈ S(∆, L). Thus, u0 trivially solves the equation s(u0 ) = a.
By assumption, |Fa (f ) − F(∆,L) (f )| ≤ Cλ−1 (1 + k)F(∆,L) (f ) − kC||f || for all
f ∈ C∞ ∗ (∆) and so F (f ) ≤ (1 + Cλ−1 (1 + k))F
a (∆,L) (f ) − kC||f ||. Turning this
around,
F(∆,L) (f ) ≥ εFa (f ) + δ||f ||,
where 0 < ε := (1 + Cλ−1 (1 + k))−1 < 1 and δ := kCλ(λ + C(1 + k))−1 . Notice
that δ is an injective function of k ∈ [0, ∞) with range [0, λ). For any normalized
6. UNIFORM K-STABILITY 47
7. Existence: an overview
7.1. The Chen–Cheng and He results. We review here a recent break-
through in the general existence theory of extremal Kähler metrics, due to Chen–
Cheng [12, 13, 14] in the constant scalar curvature case, with an enhancement
by He [32] to cover the extremal Kähler metric case. The key notion for these
results to hold is the properness of the relative K-energy with respect to a re-
ductive complex Lie group, first introduced by Tian [52] in the Fano case, and
adapted by Zhou–Zhu [59] to the general Kähler case. We start by explaining
the general setting.
but we shall use below another choice for H0 (M ). For any chosen normalization
H0 (M ), we will write
φ = φ0 + const,
where φ ∈ H(M ) and φ0 ∈ H0 (M ). Furthermore, for any σ ∈ Aut0 (M ), we
denote by σ[φ] ∈ H0 (M ) the unique ω0 -relative Kähler potential in H0 (M )
associated to the Kähler form σ ∗ (ωφ ).
We now fix a (real) connected compact subgroup K ⊂ Aut0 (M ) and denote
by G = K C ⊂ Aut0 (M ) its complexification, i.e. the smallest closed complex
subgroup in Aut(M ) containing K. By a standard averaging argument, we can
7. EXISTENCE: AN OVERVIEW 49
assume that the initial Kähler structure (g0 , J, ω0 ) is invariant under the action of
K, and consider the subspace H(M )K ⊂ H(M ) of K-invariant Kähler potentials
in H(M ). Thus, with a chosen normalisation H0 (M ), the space H0 (M )K :=
H0 (M ) ∩ H(M )K parametrizes the K-invariant Kähler metrics on (M, J) whose
Kähler forms belong to [ω0 ].
Let d be a distance on H(M ). For φ1 , φ2 ∈ H(M ), we let
(59) dG (φ1 , φ2 ) := inf d(φ1 , σ[φ2 ]),
σ∈G
where, we recall, φ0
∈ H0 (M )
denotes the normalized Kähler potential of ωφ for
0
the chosen normalization H (M ).
Definition 3.9 (Tian [52]). Let E K be a functional defined on the space
H(M )K and d a distance on H(M ). We say that E is G-proper with respect to
d if
• E K is bounded on H(M )K ;
• for any sequence φi ∈ H0 (M )K with dG (φ0 , φi ) → ∞, we have E K (φi ) →
∞.
It is useful to notice that if K0 ⊂ K is a compact subgroup with complex-
ification G0 ⊂ G, and E K is the restriction to H(M )K of a functional E K0 on
H(M )K0 which is G0 -proper, then E K is also G-proper. We also notice that the
above notion of properness depends upon the choice of normalization H0 (M ).
It is known (see e.g. [25]) that the group of complex automorphisms Aut0 (M )
admits a closed connected subgroup Autr (M ), called reduced group of automor-
phisms, whose Lie algebra is the space of real vector fields whose flow preserves
J, and which vanish somewhere on M . Furthermore, we let T ⊂ Autr (M ) be
a maximal (real) subtorus of Autr (M ), and denote by G = T C ⊂ Autr (M )
its complexification (which is a maximal complex subtorus of Autr (M )). By a
result of Calabi (see [25], Chapter 3), if there exists an extremal Kähler metric
50 3. THE CALABI PROBLEM AND DONALDSON’S THEORY
ωφ for some φ ∈ H(M ), then there is also an isometric extremal Kähler metric
ωφ̃ with φ̃ ∈ H(M )T . Thus, without loss, one can reduce the problem of finding
extremal Kähler metric in the de Rham class of ω0 to the related problem on
H0 (M )T . Furthermore, there is a natural functional E T : H0 (M )T → R, called
the T-relative K-energy and introduced by Mabuchi [40] and Guan [27], whose
critical points are precisely the Kähler potentials in H∗ (M )T , corresponding to
T-invariant extremal Kähler metrics in [ω0 ] (see [25], Chapter 4). Classically,
the relative K-energy is introduced with respect to a fixed maximal subgroup
K ⊂ Autr (M ), and acts on the space of Kähler potentials in H∗ (M )K , but it
is not difficult to see (using that the extremal vector field is central in the Lie
algebra of K) that its definition actually extends to H(M )T for any maximal
torus T ⊂ K.
The main result of [14, 32] can be then stated as follows.
Theorem 3.6 (Chen–Cheng [13], He [32]). Suppose T ⊂ Autr (M ) is
a maximal real torus and G = TC its complexification inside Aut(M ). If the
relative K-energy E T acting on H(M )T is G-proper with respect to the distance
d1 and the normalization (60), then there exists φ ∈ H(M )T such that ωφ is an
extremal Kähler metric.
We now turn to the toric case. Let (M, ω) be a smooth toric symplectic
manifold, classified by its Delzant polytope (∆, L, Λ). In this case T is a maximal
torus in Autr (M ) for any T-invariant compatible complex structure J on M .
We shall fix a point p0 on M 0 , corresponding to x0 ∈ ∆0 , a basis e of Λ, and
consider the slice
n o
∗ ∗
C∞ (∆) := f ∈ C∞ (∆) : f (x0 ) = 0 and dx0 f = 0
introduced in Example 3.1. We shall denote S ∗ (∆, L) := S(∆, L) ∩ C∞ ∗ (∆) the
Proof. We have already observed that the second part of the claim holds
for any path satisfying the weaker normalization condition dx0 ũ(t) = 0. Using
this and (62), it follows that I(φ(t)) = const. As S 0 (∆, L) is convex and u∗0 ∈
S 0 (∆, L), it follows that I(φ(t)) = 0.
Through the above correspondence, one can identify as in [21] the the relative
K-energy E T acting on H0 (M )T to the functional E(∆,L) defined in (51), acting
on S 0 (∆, L). We are now ready to state and prove our main observation in this
section.
Proposition 3.4. Suppose (∆, L) is a ∗-uniformly K-stable Delzant labelled
polytope, corresponding to toric Kähler manifold (M, ω0 , J). Then the relative
K-energy E T is TC -proper on H(M )T with respect to the distance d1 and the
normalization (60).
Proof. We use the bijection between the ω0 -relative Kähler potentials in
H0 (M ) and symplectic potentials ũ ∈ S 0 (∆, L) established in Lemma 3.4 in
order to deduce the G = TC properness of E T from the property (ii) in Proposi-
tion 3.3 of the functional E(∆,L) .
Let φj , j = 1, . . . , ∞ be a sequence in H0 (M )T with d1,G (0, φj ) → ∞. De-
note by ũj ∈ S 0 (∆, L) the corresponding sequence of normalized symplectic
potentials. We write
Z
∗ 1
ũj = uj + (u∗ − ũ∗j ) dv, j = 0, 1, . . . , ∞,
Vol(∆) ∆ 0
52 3. THE CALABI PROBLEM AND DONALDSON’S THEORY
where u∗j := π(ũj ) is the projection of ũj to the slice S ∗ (∆, L). Let us consider
the path ũj (t) =R (1 − t)u∗0 + tũj ∈ S 0 (∆, L). Using Lemma 3.4, the d1 -length of
this path is Cm ∆ |ũj − u∗0 | dv where Cm = (2π)m . Thus, we have
d1,G (0, φj ) ≤ d1 (0, φj )
Z
≤ Cm |ũj − u∗0 | dv
∆
(66)
Z Z
∗ ∗ 1
= Cm (uj − u0 ) + (u∗0 − ũ∗j ) dv dv
∆ Vol(∆) ∆
Z
≤ (Cm + 1) |u∗j − u∗0 | dv.
∆
∗
R
Thus, d1,G (0, φj ) → ∞ yields ∆ |uj | dv
→ ∞.
Now, by assumption, (∆, L) is ∗-uniform K-stable and by Proposition 3.3
(ii) we have
E T (φj ) = E(∆,L) (ũj ) ≥ δ||u∗j ||∗ + Cδ
Z
0
≥δ |u∗j | dv + Cδ → ∞,
∆
where for passing from the first line to the second we have used that ||·||∗ bounds
the L1 -norm, see Exercise 3.2.
Corollary 7.1. Suppose (∆, L) is a ∗-uniformly K-stable Delzant labelled
polytope, corresponding to a toric Kähler manifold (M, ω0 , J). Then (M, J) ad-
mits a T-invariant extremal Kähler metric whose Kähler form is in the deRham
class [ω0 ]. Equivalently, the Abreu equation (49) admits a solution in S(∆, L).
Proof. This follows from Theorem 3.6 and Proposition 3.4, using the bi-
jection between the ω0 relative Kähler potentials in H0 (M ) and symplectic po-
tentials ũ ∈ S 0 (∆, L) established in Lemma 3.4.
7.3. The continuity method. The main idea of the approach in [24] is to
use the continuity method over paths (∆, L(t)), t ∈ [0, 1] of labelled, compact,
convex, simple Delzant polytopes in Rm , where the labels Ltj (x) = hutj , xi + λtj
vary in a continuous way. We thus introduce
Definition 3.10. Let ∆ ⊂ (Rm )∗ be a given compact, convex, simple poly-
tope which has d codimension-one faces and
L = {L1 (x), · · · , Ld (x)}
7. EXISTENCE: AN OVERVIEW 53
where (u0 , f0 ) are the symplectic and Ricci potentials of the canonical metric
(g0 , J0 ). For t = 0, the existence of a solution u0 + ϕ ∈ S(∆, L) follows from
Yau’s theorem [57] of prescribing the Ricci form in a given Kähler class on
M∆ C = (M, J ). The openness follows too from a standard argument on M C .
0 ∆
The closedness reduces, via the method of [57], to a uniform C 0 (M ) bound for
the solutions ϕt of (69), see e.g. [54, 60]. The latter is derived in [51] by using
only the convexity of ∆ ⊂ (Rm )∗ .
|M | = M∆ /T ∼
C
= ∆. More precisely, for each vertex v we have maps Φu,v : Cm v →
m
∆ where ri = |zi | are the radial coordinates on Cv . We can thus define an
embedding of Φu,v : [0, ∞)m v → ∆ by letting ri = yi = u,i . The identification
(17) has now sense if we replace zi with ri , and thus we associate to ∆ a Hausdorff
topological space |M | which admits “uniformizing” charts Φv : Cm m
v /Tv → |M |,
precisely as in the definition (1.11) for orbifolds, but now the groups Γi ∼ = Tmv
are tori. Furthermore, we can define the sheaf of smooth functions of |M | by
pull back to each Cm v , etc. Using the relevant analysis in such charts, Legendre
has proved
56 3. THE CALABI PROBLEM AND DONALDSON’S THEORY
Theorem 3.9 (Legendre [38]). Theorem 3.8 holds true for any compact
convex simple labelled polytope (∆, L).
Proof. For any interior point x∗ ∈ ∆0 we let r(x∗ ) = (L1 (x∗ ), . . . , Ld (x∗ )) ∈
N (r) and
L (x) Ld (x)
1
L∗ := , . . . , ,
L1 (x∗ ) Ld (x∗ )
d Z d
R x dv X Z
∆ i
X
(70) Lj (x∗ ) xi dσLj = R ∗
Lj (x ) dσLj , i = 1, . . . , m.
j=1 Fj ∆ dv j=1Fj
In order to show that (70) has a unique solution we shall use the following
elementary
d Z
X Z
f dσLj uj = − dv u ∈ (Rm )∗ .
j=1 Fj ∆
The proof of Lemma 3.7 is left as an exercise, and can be obtained easily
starting from a triangulation of ∆ as the union of simplexes with a common
vertex at the interior of ∆.
We now suppose, without loss of generality, that 0 ∈ ∆0 , and the initial
labelling L satisfies Lj (0) = 1, i.e.
Lj (x) = huj , xi + 1, j = 1, . . . , d.
7. EXISTENCE: AN OVERVIEW 57
By Lemma 3.7,
d Z
1 X
x∗i = − huj , x∗ i xi dσLj
Vol(∆) Fj
j=1
d Z Z
1 X 1
=− Lj (x∗ ) xi dσLj + xi dσL
Vol(∆) Fj
Vol(∆) ∂∆
j=1
Z Z
(71) 1
= xi dσL − xi dσL∗ .
Vol(∆) ∂∆ ∂∆
Z X d Z Z
∗ ∗
0 = huj , x i dσLj = Lj (x ) dσLj − dσL
Fj j=1 Fj ∂∆
Z Z
= dσL − dσL∗
∂∆ ∂∆
and (70) becomes
Z Z Z Z
1
x∗i = dv xi dσL − xi dv dσL .
Vol(∆)2 ∆ ∂∆ ∆ ∂∆
Bibliography
[1] M. Abreu, Kähler geometry of toric varieties and extremal metrics, Internat. J. Math 9
(1998), 641–651.
[2] M. Abreu, Kähler metrics on toric orbifolds, J. Differential Geom. 58 (2001), 151–187.
[3] V. Apostolov, D. M. J. Calderbank, P. Gauduchon and C. Tønnesen-Friedman, Hamilton-
ian 2-forms in Kähler geometry II Global classification, J. Differential Geom. 68 (2004),
277–345.
[4] M. F. Atiyah, Convexity and commuting Hamiltonians, Bull. London Math. Soc. 14 (1982),
1–15.
[5] G. Bredon, Introduction to Compact Transformation Groups, Pure and Applied Mathe-
matics, 46, Academic Press, New York-London, 1972.
[6] E. Calabi, Extremal Kähler metrics, in Seminar on Differential Geometry, pp. 259–290,
Ann. of Math. Stud. 102, Princeton Univ. Press, Princeton, N.J., 1982.
[7] D.M.J. Calderbank, L. David and P. Gauduchon, The Guillemin formula and Kähler
metrics on toric symplectic manifolds, J. Symplectic Geom. 1 (2003), 767–784.
[8] D.M.J. Calderbank, K-energy and uniform stability for toric varieties, unpublished man-
uscript, 2006.
[9] A. Cannas da Silva, Symplectic toric manifolds, in “Symplectic geometry of integrable
Hamiltonian systems ”(Barcelona, 2001)’, 85–173, Adv. Courses Math. CRM Barcelona,
Birkhuser, Basel, 2003.
[10] B. Chen, A.-M. Li and L. Sheng, Uniform K-stability for extremal metrics on toric vari-
eties, Journal of Differential Equations 257 (2014), 1487–1500.
[11] B. Chen, A.-M. Li and L. Sheng, Extremal metrics on toric surfaces, preprint
arXiv:1008.2607.
[12] X. Chen, J. Cheng, On the constant scalar curvature Kähler metrics: apriori estimates,
arXiv:1712.06697.
[13] X. Chen, J. Cheng, On the constant scalar curvature Kähler metrics: existence results,
arXiv:1801.00656.
[14] X. Chen, J. Cheng, On the constant scalar curvature Kähler metrics: general automor-
phism group, arXiv:1801.05907.
[15] T. Darvas, The Mabuchi geometry of finite energy classes, Amer. J. Math. 139 (2017),
1275–1313.
[16] T. Delzant, Hamiltoniens périodiques et image convexe de l’application moment, Bull. Soc.
Math. France 116 (1988), 315–339.
[17] R. Dervan, Relative K-stability for Kähler manifolds, Math. Ann. 372 (2018), 859–889.
[18] R. Dervan and J. Ross, K-stability for Kähler manifolds, Math. res. Lett. 24 (2017), 689–
739.
[19] Z. Sjöström Dyrefelt, K-semistability of cscK manifolds with transcendental cohomology
class, to appear in J. Geom. Anal.
[20] Z. Sjöström Dyrefelt, On K-polystability of cscK manifolds with transcendental cohomology
class (with Appendix written by R. Dervan), to appear in Int. Math. Res. Not. (IMRN).
[21] S. K. Donaldson, Scalar curvature and stability of toric varieties, J. Differential Geom. 62
(2002), 289–349.
[22] S. K. Donaldson, Interior estimates for solutions of Abreu’s equation, Collect. Math. 56
(2005) 103–142.
[23] S.K. Donaldson, Extremal metrics on toric surfaces: a continuity method, J. Differential
Geom. 79 (2008), 389–432.
59
60 BIBLIOGRAPHY
[24] S.K. Donaldson, Constant scalar curvature metrics on toric surfaces, Geom. Funct. Anal.
19 (2009), 83–136.
[25] P. Gauduchon, Calabi’s extremal Kähler metrics: An elementary introduction, Lecture
Notes available upon request.
[26] P. Griffiths and J. Harris, Principles of Algebraic Geometry, Wiley Classics Library, John
Wiley and Sons, Inc., New York, 1994.
[27] D. Guan, On modified Mabuchi functional and Mabuchi moduli space of Kähler metrics
on toric bundles, Math. Res. Let. 6 (1999), 547–555.
[28] V. Guillemin, Kähler structures on toric varieties, J. Differential Geom. 40 (1994), 285–
309.
[29] V. Guillemin, Moment Maps and Combinatorial Invariants of Hamiltonian T n -spaces,
Progress in Mathematics 122, Birkhauser, Boston, 1994.
[30] V. Guillemin and S. Sternberg, Convexity properties of the moment mapping, Invent. Math.
67 (1982) 491–513.
[31] V. Guillemin and S. Sternberg, Riemann sums over polytopes, Ann. Inst. Fourier (Greno-
ble) 57 (2007), no. 7, 2183–2195.
[32] W. Hu, on Calabi’s extremal Kähler metrics and properness, arXiv:1801.07636.
[33] P. Heinzner and F. Loose, Reduction of complex Hamiltonian G-spaces, Geom. Funct.
Anal. 4 (1994) 288–297.
[34] Y. Karshon and E. Lerman, Non-compact symplectic toric manifolds, preprint (2009),
arXiv: 0907.2891.
[35] F. Kirwan, Cohomology of quotients in symplectic and algebraic geometry. Mathematical
Notes, 31. Princeton University Press, Princeton, NJ, 1984.
[36] S. Kobayashi and K. Nomizu, Foundations of Differential Geometry, I,II, Interscience
Publishers (1963).
[37] A. Lahdili, Kähler metrics with constant weighted scalar curvature and weighted K-stability,
arXiv:1808.07811.
[38] E. Legendre, Toric Kähler–Einstein metrics and convex compact polytopes, J. Geom. Anal.
26 (2016), 399–427.
[39] E. Lerman and S. Tolman, Hamiltonian torus actions on symplectic orbifolds and toric
varieties, Trans. Amer. Math. Soc. 349 (1997), 4201–423.
[40] T. Mabuchi, K-energy maps integrating Futaki invariants, Tohoku Math. J. (2) 38 (1986),
575–593.
[41] J. Marsden and A. Weinstein, Reduction of symplectic manifolds with symmetry, Rep.
Mathematical Phys. 5 (1974), 121–130.
[42] D. McDuff and D. Salamon, Introduction to Symplectic Topology, Oxford Mathematical
Monographs, Oxford University Press, New York, 1995.
[43] K. Meyer, Symmetries and integrals in mechanics, Dynamical Systems (Proc. Sympos.,
Univ. Bahia, Salvador, 1971), 259–272. Academic Press, New York, 1973.
[44] A. Newlander and L. Nirenberg, Complex analytic coordinates in almost complex manifolds,
Ann. of Math. 65 (1957), 391–404.
[45] Y. Odaka, The GIT stability stability of polarixed varieties via discrepency, Ann. Math.
(2) 177 (2013), 645–661.
[46] Y. Odaka, A generalization of the Ross-Thomas slope theory, Osaka J. Math. 50 (2013),
171–185.
[47] R. Sjamaar, Holomorphic slices, symplectic reduction and multiplicities of representations,
Ann. of Math. (2) 141 (1995), 87–129,
[48] G.W. Schwarz, Smooth functions invariant under the action of a compact Lie group, Topol-
ogy 14 (1975) 63–68.
[49] G. Székelyhidi, Extremal metrics and K-stability, PhD. Thesis, Imperial College London
(2006).
[50] W. Thurston, The Geometry and Topology of Three-Manifolds, New Jersey: Princeton
University Press, 1997.
[51] X.-J. Wang and X. Zhu, Kähler–Ricci solitons on toric manifolds with positive first Chern
class, Adv. Math. 188 (2004), 87–103.
[52] G. Tian, The K-energy on hypersurfaces and stability, Comm. Anal. Geom. 2 (1994),
239–265.
BIBLIOGRAPHY 61
[53] G. Tian, Kähler-Einstein metrics with positive scalar curvature, Inv. Math. 130 (1997),
1–37.
[54] G. Tian and X. Zhu, Uniqueness of Kȧhler–Ricci soliton, Acta Math. 184 (2000), 271–305.
[55] X. Wang Heigh and GIT weight, Math. res. Letter. 19 (2012), 909–926.
[56] R. O. Wells, Differential Analysis on Complex Manifolds, second edition, Graduate Texts
in Mathematics 65, Springer-Verlag, New York-Berlin, 1980.
[57] S.T. Yau, On the Ricci curvature of a compact Kähler manifold and the complex Monge-
Ampère equation, Comm. Pure Appl. Math. 31 (1978), 339–411.
[58] B. Zhou and X. Zhu, K-stability on toric manifolds, Proc. Amer. Math. Soc. 136 (2008),
3301–3307.
[59] B. Zhou and X. Zhu, Relative K-stability and modified K-energy on toric manifolds, Adv.
Math. 2019 (2008), 1327–1362.
[60] X. Zhu, Kähler–Ricci soliton type equations on compact complex manifolds with C1 (M ) >
0. J. Geom. Anal. 10 (2000), 759–774.