ZZ - arXiv.org

advertisement
arXiv:1604.06655v1 [math.CV] 22 Apr 2016
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS ON
S 1 -SYMMETRIC KAEHLER MANIFOLDS
STEVE ZELDITCH AND PENG ZHOU
Abstract. This article is concerned with asymptotics of equivariant Bergman kernels and partial Bergman
kernels for polarized projective Kähler manifolds invariant under a Hamiltonian holomorphic S 1 action.
Asymptotics of partial Bergman kernel are obtained in the allowed region A resp. forbidden region F , generalizing results of Shiffman-Zelditch, Shiffman-Tate-Zelditch and Pokorny-Singer for toric Kähler manifolds.
The main result gives scaling asymptotics of equivariant Bergman kernels and partial Bergman kernels in
the transition region around the interface ∂A, generalizing recent work of Ross-Singer on partial Bergman
kernels, and refining the Ross-Singer transition asymptotics to apply to equivariant Bergman kernels.
This article is concerned with the asymptotics of partial Bergman kernels for positive Hermitian holomorphic line bundles (L, h) → (M, ω) over a Kähler manifold of complex dimension m carrying a Hamiltonian
holomorphic S 1 action
exp t ξH : T × M → M, ιξH ω = dH, exp tξH (z) := e2πit z,
where H : M → P0 := H(M ) ⊂ R is the Hamiltonian and ξH is its Hamilton vector field. The T-action1
preserves the data (L, h) and can be ‘quantized’ to give a unitary representation of T
Uk (θ) = eikθĤk : T × H 0 (M, Lk ) → H 0 (M, Lk )
(1)
on the spaces H 0 (X, Lk ) of holomorphic sections of the tensor powers Lk , equipped with the L2 norm Hilbhk
induced by the Hermitian metric h. The self-adjoint generator of Uk (θ) is denoted by
i
(2)
∇ξ : H 0 (M, Lk ) → H 0 (M, Lk ),
2πk H
where ∇ξH s is the covariant derivative of a section s and Hs is the product of s with H [Ko, GS]. When
ξH generates a holomorphic T action, Ĥk preserves holomorphic sections and coincides with the Toeplitz
operator
Ĥk s = Πk Ĥk Πk s, s ∈ H 0 (M, Lk )
Ĥk := H +
with principal symbol H (see §1.3). Here,
Πk : L2 (M, Lk ) → H 0 (M, Lk )
is the orthogonal projection (or Szegö -Bergman kernel)
We define the weight spaces by
Vk (j) = {s ∈ H 0 (M, Lk ) : Uk (θ)s = eijθ s} = {s ∈ H 0 (M, Lk ) : Ĥk s =
j
s};
k
(3)
it is known that Vk (j) 6= {0} if and only if kj ∈ P0 = H(M ), and their dimensions have been computed in
articles on “quantization commutes with reduction” [GS]. In Lemma 2.1 we show that H(M ) = [0, a] for
a positive integer a which is equal to the symplectic area of a generic C∗ orbit. We define the associated
weight space projections (termed equivariant Bergman kernels)
Πk,j (z, w) : L2 (M, Lk ) → Vk (j).
(4)
Research partially supported by NSF grant and DMS-1541126 and by the Stefan Bergman trust .
1 We denote it by T rather than by S 1 because we use that notation for a different circle action on L∗ . We also use the
terms Bergman kernel and Szegö kernel interchangeably.
1
2
STEVE ZELDITCH AND PENG ZHOU
These equivariant Bergman kernels are the smallest components of the full Bergman kernel (or Szegö projector)
X
Πk,j : L2 (M, Lk ) → H 0 (M, Lk )
(5)
Πk (z, w) =
j
∈P0
j: k
to possess strong asymptotic expansions when kj → E for some value E of H. The norm contraction of
Πk,j (z, z) on the diagonal is denoted by Πk,j (z) and is called the equivariant density of states.2 As proved
1
in Theorems 0.1 and 0.2, the normalized equivariant density of states k −m+ 2 Πk,j (z) resembles a Gaussian
bump concentrated on the energy level H −1 (E) in the sense of being essentially equal to 1 on H −1 ( kj ) and
having “Gaussian decay” e−kbE (z) away from H −1 ( kj ) along gradient lines σ → e−σ/2 · z of H, where bE is
defined by (9) and is a non-negative smooth strictly convex function vanishing only when z ∈ H −1 (E). This
is the analogue for S 1 actions of the result of [STZ] showing that joint eigensections z α of the torus action of
a toric Kähler manifold are Gaussian-like bumps centered on the tori µ−1 (α) (the inverse image of α ∈ Zm
under the moment map µ), a fact also used in [PS, RS].
The partial Bergman kernels of the title are projectors
X
Πk,j (z, w).
(6)
Π|kP (z, w) :=
j: kj ∈P
onto subspaces
Hk,P :=
M
j
∈P
j: k
Vk (j) ⊂ H 0 (M, Lk )
(7)
corresponding to proper sub-intervals P ⊂ P0 = H(M ). They behave like sums of Gaussian bumps centered
at the the inverse images H −1 ( kj ) of the “lattice points” kj ∈ P . An illustration of this bump in complex
dimension one is given in Figure 7.10 of [W].
The main problem is to relate the asymptotic properties of Π|kP (z, w) to the geometry of the Hamiltonian
flow of H and its complexification as a C∗ action. The analogous problem for toric Kähler manifolds was
studied in [ShZ], with P a sub-polytope of the Delzant moment polytope of (M, ω). As in the toric case, we
prove in Theorem 0.3the norm contraction Π|kP (z) of Π|kP (z, z) has standard asymptotics in the allowed
region AP and exponentially decaying asymptotics in the forbidden region FP where
AP := {z ∈ M : H(z) ∈ P }, FP := M \AP .
On the boundary, or “interface” ∂AP , Ross-Singer in [RS] showed that k −m Π|kP (z) decreases from ∼ 1 to
1
∼ 0 in a tube of radius k − 2 . In the
√ special case where the minimum set of H is a complex hypersurface,
Theorem 1.2 of [RS] asserts that if k(H(z) − ǫ) is bounded, then
Z √k(H(z)−ǫ)
t2
1
1
−
−n
e 2|ξH (z)|2 dt + O(k − 2 ).
k Π|kP (z) = p
(8)
2
2π|ξH (z)| −∞
Here, |ξH | is the norm of ξH with respect to the Kähler metric ω. This is the picture physicists have in mind
when speaking of the Landau levels being filled in H −1 (P ). For instance, in figure 7.11 of [W] is illustrated
the density profile of a droplet (with filling fraction 1) when the first
√ m angular momentum levels are filled.
It pictures the characteristic function of a disc of radius rm = 2mℓB (where ℓB is a certain “magnetic
length”) smoothly dropping off to zero in a thin shell around the boundary. The shape of the transition
curve is the incomplete Gaussian integral. A similar discussion may be found in [J].
One of the prinicipal motivations for this article is to establish this transition law for all Hamiltonian
holomorphic S 1 actions, with no conditions on the fixed point set. We obtain the interface asymptotics
from the Gaussian asymptotics of the equivariant kernels (4). The Gaussian asymptotics of the equivariant
Bergman kernels in Theorems 0.1 and 0.2 are used in Theorem 0.4 and Corollary 0.5 to give incomplete
Gaussian integral asymptotics for partial Bergman kernels (8), which are essentially integrals of equivariant
Bergman kernels. In Theorem 0.3 we give exponentially decaying asymptotics of the partial Bergman kernels
Π|kP in the forbidden region when the metric h is real analytic. The exponentially decaying asymptotics
2The norm contraction of any kernel K(z, w) on the diagonal is denoted K(z).
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
3
in the forbidden region is again governed by the ‘action’ bE (z) (9) from z to H −1 (E). Asymptotics of
equivariant Szegö kernels have been studied by R. Paoletti in several settings, of which the closest to this
article are contained in [P, P2]. They were not explicitly defined or studied in [RS]; as discussed in §0.6,
they constructed kernels Gj,k which play the role of Πk,j .
0.1. Asymptotics of equivariant Bergman kernels. To state our results we introduce some notation.
We assume that E is a regular energy level, i.e. H has no critical points on E. In §2.1 and §2.2 the complex
and real Morse theory of Hamiltonians generating holomorphic S 1 actions is reviewed. The Hamiltonian
and gradient flows of H commute and generate a C∗ action. There is an open dense subset Mmax in which
the C∗ orbits flow into the ‘source’, which is the minimum set of the Hamiltonian. More relevant for us is
E
the open dense set Mmax
of points whose gradient orbits intersect H −1 (E). Note that H −1 (E) disconnects
E
M into two sides, {H > E} and {H < E}, and z ∈ Mmax
if either the forward or backward gradient line
−1
through z intersects H (E).
For simplicity of notation we denote the T action by eiθ · z and its C∗ action by ew z = eρ+iθ · z. We
E
denote the infintesimal generators of the R+ , resp. T, action by ∂ρ , ∂θ . For z ∈ Mmax
, let τE (z) ∈ R be such
that
z = eτE (z) · zE
where zE ∈ H −1 (E). Thus, τE (z) is like the travel time along the gradient flow of H from z to the intersection
zE ∈ H −1 (E). We refer to §2.4 for further details. As in [ShZ] (reviewed in §2.5) define
Z 2τE (z) h
i
bE (z) = −2EτE (z) +
H(e−σ/2 · z) · dσ .
(9)
0
For each point z ∈ M that is not a fixed point of T, we fix a local C∗ -invariant holomorphic section
eL ∈ Γ(U, L) in an open neighborhood U of z and define the Kahler potential ϕ by e−ϕ = keL k2h .
Our first result is the precise statement that k −m+1/2 Πk,j (z, z) is a kind of Gaussian beam along H −1 (E),
i.e. in the tangential directions along H −1 (E) it is essentially constant while it has Gaussian decay at speed
k in the normal directions (i.e. along ∇H lines).
Theorem 0.1. Let ω be a T-invariant C ∞ Kähler metric, and let H generate the holomorphic Hamiltonian
E
T action. Let E be a regular value of H. Assume that | kj −E| ≤ Ck , and that z ∈ Mmax
. Then the equivariant
density of states has the following asymptotics.

q
2
−1
m− 12

)),
z ∈ H −1 (E),
k

π∂ 2 ϕ(zE ) (1 + O(k

ρ
Πk,j (z) ∼
q


2
−1
 k m− 12 e−kbE (z)
)),
π∂ 2 ϕ(zE ) (1 + O(k
ρ
z∈
/ H −1 (E).
Here, zE = e−τE (z) · z ∈ H −1 (E) and bE (z) is defined by (9).
1
At first sight the order k m− 2 may seem odd, since there are ≃ k lattice points kj in H(M ), since the
dimensions of the weight spaces have orders k m−1 , and since the sum over j gives the full Bergman kernel
density of states, Πk (z) ≃ k m A0 (z). The ‘paradox’ is resolved
pointwise,
√ by the fact that the asymptotics are
−1 j
while dimensions are their integrals, and Πk,j (z, z) is O( k) times its average at the center H ( k ) of its
bump due√
to its transverse Gaussian decay. Given z ∈ H −1 (E), Πk,j (z) is negligible asymptotically except
for the O( k) lattice points kj within O( √1k ) of E and these restore the full Bergman kernel.
The next result concerns scaling asymptotics of the equivariant Bergman kernels in a
of H −1 (E).
C
√
-neighborhood
k
Theorem 0.2. Let ω be a C ∞ T-invariant Kähler metric, and let H generate the holomorphic T action.
If kj = E + O( k1 ), and if z0 ∈ H −1 (E), then for u ∈ R,
1
Πk,j (e
u
√
k
2k m− 2
2
e−u
· z0 ) = q
2π∂ρ2 ϕ(z0 )
∂ρ2 ϕ(z0 )
(1 + O(k −1/2 )).
4
STEVE ZELDITCH AND PENG ZHOU
0.2. Asymptotics of partial Bergman kernels. In this section, we state analogues of Theorem 0.1 for
partial Bergman kernels (6) and of Theorem 1.2 of [ShZ] for partial Bergman kernels of toric Kähler manifolds.
Our aim is to obtain exponentially accurate asymptotics in the forbidden region.
The definition of partial Bergman kernel (6) makes sense for any proper subinterval P = [E1 , E2 ] ⊂ P0 =
H(M ) but the asymptotics are cleaner for special subintervals and sequences of powers k. As discussed in
§1.4, the pairs (k, j) indexing Πk,j run over a semi-lattice in Z+ × Z+ with kj ∈ P0 . The asymptotics are
cleanest for pairs {n·(j0 , k0 ) : n ∈ Z+ } along a rational ray in the semi-lattice. For a general interval [E1 , E2 ],
the pairs {(j, k) : kj ∈ [E1 , E2 ]} is the semi-lattice in the conic wedge in R2+ formed by rays xy = E1 (resp. E2 ).
The cleanest asymptotics hold for rational intervals [E1 , E2 ] = [ ap , pb ] and for powers Lkp , (k = 1, 2, 3, . . . ).
These are the analogues of the lattice sub-polytopes in [ShZ]. We first state the result for such intervals.
We also assume that the metric is real analytic. The proof is given in §6 and is essentially the same proof
as that in [ShZ] (Theorem 1.2). In Theorem 0.6 we prove a more general result with slightly less accurate
asymptotics.
Theorem 0.3. Let P = [ ap , pb ] ⊂ P0 = H(M ). Assume ω is real analytic with analytic radius ρa . Assume
that z ∈ Mmax is within the analyticity threshold of §(3.4). Then the density of states of the partial Bergman
kernel is given by the asymptotic formulas: for k ∈ pZ

 k m (c0 + c1 k −1 + c2 k −2 + · · · ),
for z ∈ H −1 (P )
Π|kP (z) ∼
 k m e−kbE (z) c (z) + O(k −1 ) ,
for z ∈ X1+ ∩ dist(z, H−1 (P)) < ρa
0
E
where cj ∈ C ∞ (Mmax
), and bE (z) is defined in (9) and E is the endpoint of P closer to H(z). Furthermore,
E
the remainder estimates are uniform on compact subsets of Mmax
.
The proof is based on the Euler-MacLaurin formula and that explains why we choose special intervals
and sequences. The asymptotics improve on Theorem 1.1 of [RS] by giving the exponentially accurate
asymptotics in the forbidden region and is the analogue of the mass density results of [ShZ].
0.3. Interface asymptotics. Interface asymptotics concerns the sums,
Z
√
√
√
X √ j
f ( k( − E))Πk,j (eβ/ k · z0 ) =
fˆ(t)e−iE kt Πk (eit/ k · z, z)dt
k
R
j
where z = eβ/
√
k
(10)
z0 and where fˆ ∈ L1 (R), so that the integral on the right side converges.
Theorem 0.4. Let ω be a C ∞√T-invariant Kähler metric, and let H generate the holomorphic T action.
Fix E ∈ H(M ), and let z = eβ/ k · z0 for some z0 ∈ H −1 (E), β ∈ R, and let f ∈ Cb (R). Then
lim k −m
k→∞
X
j: kj ∈P0
Z ∞
−β
− 21 √ 2x
√
√ j
2 ϕ(z )
∂ρ
0
f (x)e
f ( k( − E))Πk,j (eβ/ k · z0 ) =
k
−∞
√
∂ρ2 ϕ(z0 )
!2
2dx
q
.
2π∂ρ2 ϕ(z0 )
The asymptotics also hold if f = 1[E1 ,E2 ] is the characteristic function of an interval (possibly unbounded).
The above result assumes only f ∈ Cb (R) without any differentiability and get weak limit convergence
without bound of error term. In §8.2, we assume f ∈ C ∞ (R) and uses f (x) · 1P (x) as the test function,
where P is any closed interval on R, there we get O(k −1/2 ) error term.
Corollary 0.5. Using the same notation as Theorem 0.4, if we set f (x) = 1[0,∞] (x), then
Z β √∂ρ2 ϕ(z0 )
√
X
2
dt
e−t /2 √ + O(k −1/2 ).
k −m
Πk,j (eβ/ k · z0 ) =
2π
−∞
j>kE
This is very closed related to the interface asymptotics (8) of Ross-Singer. In §2.4 we discuss the Kähler
E
potential ϕ for ω in Mmax
, and show that
1
(11)
H = ∂ρ φ, π∂ρ2 φ = |∇H|2 .
2
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
5
Hence, the leading factor in the asymptotics agrees with Ross-Singer’s formula up to normalization difference
(see the equation for I on page 25 in [RS]).
Remark: The asymptotics are parallel to the scaling asymptotics of [HZZ] for Wigner distributions of
eigenspace projections Π~,E around the energy surface H −1 (E) ⊂ T ∗ Rd of a model Schroedinger operator
Ĥh (the isotropic Harmonic Oscillator) on Rd . The Hamiltonian flow is the Hamiltonian T action on Cd
generated by H(z) = kzk2 . However, the partly Airy scaling results in the real domain [HZZ] pertain to the
Schroedinger representation, while Theorem 0.2 refers to the Bargmann-Fock space representation.
0.4. Summary of results. The next result sums up the preceding results and is slightly more general. By
contrast with the complete asymptotic expansion in Theorem 0.3, it gives only a leading order term in the
forbidden region. Howevever it does not require real analyticity of the metric and works for general intervals.
Theorem 0.6. Let ω be a C ∞ T-invariant metric. Let z ∈ M , not
H(M ). Then we have

−∞
)

Πk (z) + O(k


2τ (z,j0 /k) −1
−1

(z)(1
−
e
Π
k,j

0
) (1 + O(k ))

√
k(H(z)−E1 )
Π|kP (z) = k m Φ
√
(1 + O(k −1/2 ))

∂ρ2 ϕ(z)/2



√


k(E2 −H(z))
k m Φ
√
(1 + O(k −1/2 ))
2
∂ρ ϕ(z)/2
a fixed point of T, and P = [E1 , E2 ] ⊂
z ∈ H −1 (int(P ))
z∈
/ H −1 (P )
|H(z) − E1 | < Ck −1/2
|H(z) − E2 | < Ck −1/2
where j0 = j(z, k, P ) is the j ∈ kP ∩Z which minimizes |H(z)− j/k|, and Φ(a) = P(X < a) is the cumulative
density of the standard Gaussian X ∼ N (0, 1).
0.5. Bernstein polynomials. In essence, Theorem 0.3 concerns kernels of the form,
Z
X j
fˆ(t)e−iEt Πk (eit/k z, z)dt
f ( − E)Πk,j (z, z) =
k
R
j
(12)
which are the Bernstein-Szasz analytic functions of [Ze2, F]. In contrast, Theorem 0.4 is essentially about
kernels of the form (10). More precisely, Bernstein polynomials in the sense of [Ze2, F] are functions
Bk (f )(z) := f (Ĥk )Πk (w, z)|w=z = f (Dφ )Πk (rφ z, z)|φ=0 =
Nk
X
j
f ( )Πk,j (z),
k
j=1
where f is smooth. Here, f (Ĥk ) on H 0 (M, Lk ) is defined by the spectral theorem,
Z
f (Ĥk ) =
fˆ(τ )eiτ Ĥk dτ,
(13)
(14)
R
so that f (Ĥk )ŝk,j = f kj ŝk,j if sk,j is an eigensection of Ĥk . Partial Bergman kernels are Bernsteinpolynomials in the case where f is a step function 1P .
Proposition 0.7. Let T be a holomorphic Hamiltonian circle action eiφ z on a polarized projective Kähler
manifold (M, L, ω), let ξ be its infinitesimal generator and let Ĥk be its linearization on H 0 (M, Lk ). Let 1P
be the characteristic function of a proper subinterval P ⊂ H(M ). Then,
Π|kP (z) =
1
Πk (z) 1P (Ĥk )Πk (w, z)|w=z .
When f ∈ C0∞ (R), it is proved in [Ze2, F] that
f (k −1 Dφ )Πk (reiφ z, z)|φ=0 → f (H(z)), (k → ∞).
(15)
Since 1P above is a characteristic function with a jump, the limit formula (15) cannot be true. In fact, there
is a kind of mean value formula at the jump involving incomplete Gaussian integrals. In the classical setting
of Bernstein polynomials on [−1, 1] (corresponding to M = CP1 with the Fubini-Study metric), the jump
formula is proved in [Ch, L, Lev]. The interface asymptotics of [RS] and of this article are generalizations of
Theorem 1.5.2 of [L] in the one-variable setting.
6
STEVE ZELDITCH AND PENG ZHOU
We now consider Bernstein polynomials which sum over sub-intervals but with respect to a smooth test
function f (x) on R. Let P = [E1 , E2 ] ⊂ H(M ) be a sub-interval and define
X
j
(16)
Πk,P,f (z) :=
f ( )Πk,j (z).
k
j∈kP ∩Z
The results on partial Bergman kernels generalize to Bernstein polynomials as follows:
Proposition 0.8. With the notation above, let z ∈ Mmax and f a smooth test function on R. Then

f (H(z))Πk (z) + O(k −1 )
z ∈ H −1 (int(P ))




2τ (z,E) −1

) (1 + O(k −1 )) z ∈
/ H −1 (P )

√ ) (z)(1 − e f (Ei )Πk,j(z,k,P
k(H(z)−E1 )
Πk,P,f (z) = f (E1 )k m Φ
√
(1 + O(k −1/2 ))
|H(z) − E1 | < Ck −1/2

∂ρ2 ϕ(z)/2



√


k(E2 −H(z))
f (E2 )k m Φ
√
(1 + O(k −1/2 ))
|H(z) − E2 | < Ck −1/2
2
(17)
∂ρ ϕ(z)/2
where Ei ∈ P is the point closest to H(z), j0 = j(z, k, P ) is the j ∈ kP ∩ Z which minimizes |H(z) − j/k|,and
Φ(a) = P(X < a) is the cumulative density of the standard Gaussian X ∼ N (0, 1).
Proposition 0.8 can be proved by a small modification of the proof of Theorem 0.6 in §7. In each case,
after one does the localization, one may Taylor expand f (z) around the localization point with appropriate
error terms O(k −1 ) or O(k −1/2 ). We omit further details.
0.6. Remarks on the proof, on related work and open problems. All of the properties of the
equivariant Bergman kernels and partial Bergman kernels are derived from the Boutet de Monvel -Sjöstrand
parametrix for the Bergman-Szegö kernels of Lk [BSj]. The equivariant Bergman kernels are expressed as
Fourier coefficients of the full Bergman kernel and stationary phase in the complex domain is used to get
the exponentially decaying asymptotics.
n
In [RS] the role of the equivariant kernels is played by the terms Gn,k (z, w)σ n (z) ⊗ σ(w) in Definition
5.21, where σ is defined in Section 5.2 as the section σ ∈ H 0 (Y, O(Y )) defining a hypersurface component of
Fix(T). We do not assume Fix(T) contains a hypersurface component and do not make use of σ. We also
do not make any constructions of Gn,k or construct special parametrices adapted to the hypersurface Y , as
is done in [RS].
The analysis in Ross-Singer [RS] was largely motivated by a more general unsolved problem of determining
Bergman kernel asymptotics for subspaces of sections defined in terms of vanishing order along a divisor
Y ⊂ M . The partial Bergman kernels are Schwartz kernels of the orthogonal projections
2
k
0
k
tk
ΠY,t
k (z, w) : L (X, L ) → H (X, O(L ) ⊗ IY )
(18)
ρY,t
(z) := ΠY,t
k,
k (z, z)hk
hk
z ⊗hz
(19)
onto the subspace of s ∈ H 0 (M, Lk ) which vanish to order tk on a complex hypersurface Y . The main
question is to find the asymptotics of the density of states,
defined by contracting the Szegö kernel along the diagonal with the metric. The asymptotics depend on
whether z lies in the allowed region At far from the divisor Y or whether it lies in the forbidden region Ft
near the divisor Y but it is more difficult to define these regions in the absence of a T action.
The general definition of allowed/forbidden regions (due to R. Berman [Ber0] and developed several
articles of Ross-Witt-Nystrom) is that the allowed region is the set At := DY,t = {φe,Y,t = φ} where a
certain equilibrium potential φe,Z,t equals the original Kähler potential. As pointed out in [PS], in this
generality, there is no information about the smoothness of ∂DY,t nor about the transition behaviour of
(19) near ∂DY,t . In [Ber0] it is suggested to employ singular Hermitian metrics with singularities and with
negative curvature concentrated along the divisor on Y . At the present time, this program has only partially
been carried out in [Ber0, RS, Co] and remains largely open.
The approach we take is to define partial Bergman kernels and allowed and forbidden regions in terms
of spectral theory of Toeplitz operators. When the Hamiltonian flow is holomorphic and periodic and when
one component of the fixed point set M T of the T action is a divisor Y , then the allowed and forbidden
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
7
regions are those defined above in terms of the Hamiltonian and the interface asymptotics are given by (8)
in [RS]. The hypersurface Y is necessarily the minimum set of the classical Hamiltonian. The link between
the spectral theory and the definition of partial Bergman kernels in terms of vanishing order is given in the
following Proposition (closely related to Lemma 5.4 of [RS]. )
Proposition 0.9. Suppose that the minimum set of H is a complex hypersurface Y . Then H 0 (X, O(Lk ) ⊗
L
IYtk ) = j≥tk Vk (j) is the direct sum of eigenspaces of Ĥk for eigenvalue ≥ t.
Of course, it is only in special cases that the minimum set of H is a hypersurface. To take a simple
model example, the hypersurface {z1 = 0} is a component of the fixed point set of the S 1 action on Cm
defined by eiθ (z1 , . . . , zm ) = (eiθ z1 , z2 , . . . , zm ), but for the ‘isotropic Harmonic oscillator’ eiθ (z1 , . . . , zm ) =
(eiθ z1 , eiθ z2 , . . . , eiθ zm ) generated by H = ||Z||2 /2 only {0} is in the fixed point set or minimum set of H.
The spectral viewpoint of this article generalizes in many respects to any Hamiltonian H : M → R on
any compact Kähler manifold. The Hamiltonian need not generate an S 1 action. In this case, the gradient
flow and Hamiltonian flow do not commute or define a C∗ action, so much of the analysis of this article does
not generalize. However it can be replaced by a more difficult analysis using Toeplitz operators [ZZ].
Partial Bergman kernels corresponding to intervals of eigenvalues are closely related to Bergman kernels
for the Kähler symplectic cut of M in H −1 (P ) in the sense of [BGL]. In a related article, we explain the
precise relationship.
0.7. Acknowledgements. We would like to thank R. Paoletti, J. Ross, and M. Singer for their comments
on earlier versions.
1. Hermitian line bundles, Kaehler potentials and geometric quantization
In this section, we review some elementary facts about the Kahler geometry and geometric quantization,
and also establish our notations and conventions.
1.1. Hermitian line bundles. Let (L, h) → (M, ω = c1 (L)) be an ample line bundle with a positive
hermitian metric over a projective Kahler manifold. For any z ∈ M , and any open neighborhood U of z on
which L is trivial, we may choose a local trivialization eL ∈ Γ(U, L), that is eL (z) 6= 0 for all z ∈ U . Then
we may define the corresponding local Kahler potential ϕ : U → R as
h(eL (z), eL (z)) =: e−ϕ(z) := h(z).
The Chern connection associated to h is
∇ : L → L ⊗ A1 (M )
where
∇ = ∇(1,0) + ∇(0,1) ,
¯
∇(0,1) = ∂,
∇(1,0) = ∂ + A(1,0) ,
A(1,0) = ∂ log h = −∂ϕ
and the curvature is
¯ log h = ∂ ∂ϕ.
¯
F∇ (z) = dA(1,0) = d∂ log h = ∂∂
We choose the Kahler form ω = Ric(L), more precisely
ω = c1 (L) =
i
i ¯
−1 c
F∇ =
∂ ∂ϕ =
dd ϕ.
2π
2π
4π
¯ such that dc f = df ◦ J.
where dc = i(∂ − ∂),
It often simplifies the analysis to lift sections of L → M and operators on sections to the unit circle
bundle Xh of the Hermitian metric h, so that geometric pre-quantization of the S 1 action is pullback of
scalar functions under a flow. In the next section we discuss the geometric aspects of the lift and in §3.1 we
discuss the analytic aspects.
8
STEVE ZELDITCH AND PENG ZHOU
1.2. The disc bundle and the circle bundle of L∗ . Let (L∗ , h∗ ) be the dual bundle to L with the
induced hermitian metric h∗ , which we will also denote as h from now on. Let e∗L ∈ Γ(U, L∗ ) be the dual
frame to eL , then we can define the disc and circle bundle in L∗ :
Dh = D(L∗ ) := {(z, λ) | z ∈ M, λ ∈ L∗z , kλkh ≤ 1},
X = Xh = ∂Dh .
∗
The disc bundle Dh is strictly pseudoconvex in L , and hence Xh inherits the structure of a strictly pseudoconvex CR manifold. Let ρ be a smooth function defined in a neighborhood of ∂D inside D, such that
ρ > 0 in Do , ρ = 0 on ∂D and dρ 6= 0 near ∂D. For example, one may take ρ(x) = 1 − |λ̂|2 eϕ(z) or
ρ(x) = − log(|λ̂|2 eϕ(z) ), where x = (z, λ̂e∗L (z)). Associated to Xh is the contact form3
1
¯ X = dθ + i∂ϕ(z),
¯
α = i∂ρ|X = −i∂ρ|
π∗ ω =
dα
(20)
2π
where we used (z, θ) to denote z, eiθ e∗L (z)/ke∗L (z)kh ∈ Xh , and we abused notation by omitting π ∗ in
¯
The Reeb vector field R is uniquely defined by α(R) = 1, ιR dα = 0; here it is R = ∂θ , the fiberwise
π ∗ ∂ϕ(z).
rotation. Since later we will use ∂θ for the generator of the holomorphic circle action on M , we will always
refer to the Reeb flow by R, and the group action by rθ := exp(θR).
A section sk of Lk determines an equivariant function ŝk on L∗ by the rule
ŝk (λ) = λ⊗k , sk (z) , λ ∈ L∗z , z ∈ M ,
where λ⊗k = λ ⊗ · · · ⊗ λ.
1.3. Lifting the Hamiltonian flow to a contact flow on Xh . Let H be a Hamiltonian function on
(M, ω). Let ξH be the Hamiltonian vector field associated to H, that is,
dH(Y ) = ω(ξH , Y )
for all vector field Y on M . The sign convention for the Hamiltonian vector field and the corresponding
Poisson bracket is
df (Y ) = ω(ξf , Y ), {f, g} = −ω(ξf , ξg )
this choice ensures that [ξf , ξg ] = ξ{f,g} .
The purpose of this section is to lift ξH to a contact vector field ξ̂H on Xh and to lift the Hamiltonian T
h
h
action to a contact T action. Recall that the horizontal lift is defined by ξH
∈ ker α and π∗ ξH
= ξH . We
1
also denote the Reeb vector field generating the canonical S action on Xh → M by R.
Lemma 1.1. Define
Then ξˆH is a contact vector field.
h
ξˆH = ξH
− 2πHR
h
= ξH , it suffices to check that ξˆH preserve the contact form α. By (20),
Proof. Since π∗ ξˆH = π∗ ξH
∗
h −2πHR α = (ιξ h −2πHR ◦ d + d ◦ ιξ h −2πHR )α = ιξ h π (2πω) + d(−2πHα(R)) = 0.
Lξ̂H α = LξH
H
H
H
In Lemma 2.7 we prove that the lifted flow is periodic of period 2π.
d̂
i
ξ̂H (ŝ). If H defines a holomorphic S 1 action, then the
Lemma 1.2. For any C ∞ section s of Lk , Hk s = 2πk
spectrum of Ĥk is given by
j
j
Sp(Ĥk ) = { : j ∈ N, ∈ H(M )}.
k
k
d
Proof. We write ξH = ξ. It is well-known that ∇ξ s = dŝ(ξ h ); we refer to [KN] for the proof. The equation
follows from the fact s lifts to an equivariant function satisfying Rŝ = ikŝ.
Recall the definition (1) of Uk (θ) = eikθĤk , acting on C ∞ (M, Lk ), we have
3If we used two different defining functions ρ and ρ , the induced αs would differ as well. However, if ρ = f (ρ ), then
1
2
1
2
α1 = f ′ (0)α2 only differ by a constant factor. The two choices of ρ given here differ by a function f (x) = − log(1 − x), with
f ′ (0) = 1, hence the resulting α is well-defined.
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
9
Corollary 1.3. For any smooth section s ∈ C ∞ (M, Lk ), and for any t ∈ R,
ξˆH
ξˆH
)ŝ = ŝ ◦ exp(θ ).
U\
k (θ)s = exp(−θ
2π
2π
As mentioned above, if we have a holomorphic S 1 action, then the lifted flow is periodic of period 2π
(Lemma 2.7). It follows that Uk (θ) is periodic of period 2π.
1.4. The S 1 × S 1 action on Xh and its weights. The Reeb flow and the lifted T action together define
an S 1 × S 1 action on Xh . Its weights form the semi-lattice {(j, k) ∈ Z+ × Z+ , j ∈ kP0 }. This lifting and the
approximation of energy levels by rays in Z+ × Z+ is discussed in detail in [STZ] for toric varieties, and the
same discussion applies almost verbatim to T actions.
The asymptotics of the equivariant Bergman kernels Πk,j involves pairs (jn , kn ) of lattice points along
a “ray” in the joint lattice. The simplest rays are the “rational rays” where j/k ∈ Q ∩ P0 . Somewhat
more complicated are “irrational rays” where kjnn → E ∈
/ Q. In this case we consider lattice points with
j
C
| k − E| ≤ k .
2. Kaehler manifolds carrying C∗ actions
We begin by reviewing the geometry of C∗ actions on Kähler manifolds and give examples where at least
one component of the fixed point set is a hypersurface. We also consider the possible Hamiltonians H which
generate such actions.
2.1. Bialnycki-Birula decomposition. Let (M, ω) be a compact Kähler manifold equipped with a holomorphic C∗ action. The generator of the C∗ action ξ ∈ H 0 (M, T 1,0 ) is a holomorphic vector field. A
holomorphic T action which preserves ω is necessarily an isometric T action for the Kähler metric. The
closure of a non-trivial C∗ orbit contains two fixed points and is a topological S 2 called a gradient sphere.
A free gradient sphere is one whose generic point has trivial stabilizer.
By Frankel’s theorem [F], if the action has a fixed point, then the real S 1 action is Hamiltonian. We
denote the Hamiltonian by H : M → R and its Hamilton vector field by ξ = ξH . Let F1 , · · · , Fr be the
connected components of the fixed point set M T . Each Fj is a compact totally geodesic Kähler submanifold
of (M, ω). Set
Mi+ := {x ∈ M : lim tx ∈ Fi }, Mi− := {x ∈ M : lim tx ∈ Fi }.
t→∞
t→0
The so-called Bialynicki-Birula decomposition [BB, CS] states that the strata of the disjoint decpomposition
[
[
M=
Mi+ =
Mi−
(21)
are locally closed analytic submanifolds. In Theorem II of [CS] it is proved that
T (Mj±)|Fj = N (Fj )± ⊕ T Fj ,
where N (Fj )+ (resp. N (Fj )− ) is the weight space decomposition with positive (resp. negative) weights.
Moreover, there is precisely one component M1+ of the plus-decomposition (called the source), resp. one
component M1− of the minus decomposition (called the sink) such that the associated stratra is Zariski open.
We will denote the maximal stratum M1− that flows to the sink by Mmax . To paraphrase [BBS], the C∗
action gives a ’flow’ from the source to the sink, and the ’flowlines’ are closures of ’generic’ orbits and limits
of such closures.
2.2. Morse theory and gradient flow. The same decomposition can be obtained from the real Morse
theory of the Hamiltonian H. Kirwan proved that H 2 is a minimally degenerate Morse function. Since we
are dealing with a real-valued moment map, we may simply use H and it is also a minimally degenerate
(perfect) Bott-Morse function. The gradient flow of H with respect to the Kähler metric induces a Morse
stratification of X, and in [Ki2, Y] it is proved that this stratification is the same as the Biralynicki-Birula
decomposition. That is, the Morse stratum
Sj± = {x ∈ M : lim exp(t∇H) · x ∈ Fj }
t→±∞
10
STEVE ZELDITCH AND PENG ZHOU
is the same as Mj± . We note that
M T = {x : dH(x) = 0}
so that Fj are the components of the critical point set. The sink corresponds to the minimum set of H. In
[RS] it is assumed that one of the components of M T is a hypersurface, and this hypersurface is necessarily
the minimum set of H.
E
Above we defined the open dense set Mmax
of points whose forward or backward gradient trajectories
−1
intersect H (E). Its complement consists of the stable/unstable manifolds of critical points other than the
minimum/maximum. These gradient trajectories can get hung up at the other critical points and not make
it to H −1 (E).
We will also identify the Lie algebra g of T with R, such that −2π∂θ ∈ g 7→ 1 ∈ Z. Let H be the
corresponding Hamiltonian for ξH = −2π∂θ . H is determined only up to an additive constant. We fix the
indeterminacy in H by defining H = 0 on its mimimum set, the source M1+ .
Remark: (Remark on periods) By definition, the vector field ∂θ of the action eiθ z has period 2π, so the
above convention makes the period of ξH equal to 1.
2.3. The image H(M ). We normalized H so that the minimum of H is zero. The question then arises
what is the maximum value of H or equivalently what is the interval H(M ). It must be a “lattice polytope”,
i.e. an interval which integer endpoints. Thus, the maximum of H must be an integer.
R
Lemma 2.1. Let z ∈ Mmax and let Oz ≃ CP1 be the compactification of C∗ z. Let ω(Oz ) = Oz ω. Then
• ω(Oz ) is a positive integer and is constant in z for z ∈ Mmax .
• max H = ω(Oz );
Proof. For each z ∈ Mmax we obtain a polarized Kähler CP1 by (Oz , L|Oz , ω|Oz ) and it must be the case
that ω|Oz ∈ H 2 (Oz , Z). This proves the first statement. We then restrict H : Oz →R R. It generates the
S 1 action restricted to Oz . Hence ω|Oz = (2π)−1 dH ∧ dθ. If ω(Oz ) = M , then M = Oz (2π)−1 dH ∧ dθ =
R Hmax
dH = Hmax , or Hmax = M .
0
2.4. The Hamiltonian and the T-invariant Kähler potentials. Following §0.1, for any w = eρ+iθ ∈ C∗ ,
denote the C∗ action on M by z 7→ eρ+iθ · z. If we choose a local slice S of the C∗ action (necessarily a
symplectic manifold), then we may define slice-orbit coordinates (ρ, θ, y) by letting y be coordinates on the
slice and identifying
eiθ+ρ y = z.
(22)
For instance, if we choose a slice SE of the T action on H −1 (E) then we may use SE × C∗ to give local
E
coordinates on a neighborhood of SE . Also, H −1 (E) is a slice of the gradient flow or R+ action on Mmax
−1
and we use the coordinates (ρ, zE ) ∈ R × H (E) as well.
E
As in the introduction, for z ∈ Mmax
, we define
τE (z) ∈ R := the unique time s.t. z = eτE (z) · zE , zE ∈ H −1 (E).
(23)
As in §0.1, we denote the two global vector fields ∂ρ , ∂θ (not be confused with the Reeb flow R on the
circle bundle), such that
∂θ f (z) :=
d
|θ=0 f (eiθ · z),
dθ
∂ρ f (z) :=
d
|ρ=0 f (eρ · z),
dρ
∀f ∈ C ∞ (M )
(24)
Since the C∗ action is holomorphic, we have ∂θ = J∂ρ .
On any simply connected open set we may define a local T-invariant Kähler potential ϕ for ω, so that
i
¯ We mainly use the properties of ϕ restricted to a C∗ orbit F . Our choice of coordinates is such
∂ ∂ϕ.
ω = 2π
that on F ,
1 2
i ¯∗
∂ ∂ιF ϕ =
∂ ϕ dρ ∧ dθ.
(25)
ι∗F ω =
2π
4π ρ
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
11
Recall that the gradient vector field ∇H is related to the Hamiltonian vector field ξH , for any Y ∈
V ect(M ),
dH(Y ) = ω(ξH , Y ), dH(Y ) = g(Y, ∇H), g(X, Y ) = ω(X, JY )
hence ∇H = JξH = −2πJ∂θ = 2π∂ρ . Thus the limit point of downward gradient flow ∇H is the same as
limρ→−∞ eρ · z = z∞ ∈ M1+ .
The following lemma relates H with the local Kahler potential.
Lemma 2.2. Fix any z ∈ Mmax . Then
H(z) =
1
∂ρ ϕ(z)
2
Proof. As in [GS] (5.5) we define a T-invariant potential using a T -invariant holomorphic section s ∈
0
HT
(M, L) in the sense of Lemma 1.3, i.e. so that Ĥs = 0. Then,
i
i
i
∇ξ s + Hs =
hA, ξH is + Hs =
h−∂ϕ, −2π∂θ is + Hs
2π H
2π
2π
where we used the Chern connection 1-form with respect to the basis frame s is given by A = −∂ϕ, and our
convention of ξH = −2π∂θ (see above). Hence
0 = Ĥs =
i
−J
−1
1
¯
H = −ih∂θ , ∂ϕi = h∂θ , (∂¯ − ∂)ϕi = h∂θ ,
(∂ + ∂)ϕi
= hJ∂θ ,
(dϕ)i = ∂ρ ϕ,
(26)
2
2
2
2
This definition is unambiguous because any two T-invariant holomorphic sections give the same Hamiltonian
1
0
∗
2 ∂ρ ϕ(z). Indeed, let s1 , s2 ∈ H (M, L) be two T-invariant (hence C invariant) holomorphic sections. Then
s1 = f s2 for some C∗ -invariant meromorphic function f . Then ∂ρ f = ∂θ f = 0, so ϕ1 (z) = − log ks1 k2 =
− log |f |2 + ϕ2 (z), and ∂ρ ϕ1 (z) = ∂ρ (− log |f |2 + ϕ2 (z)) = ∂ρ ϕ2 (z).
2.5. The second derivative of ϕ and the action integral bE . We now consider the relation of ∂ρ2 ϕ
and bE (z) (9). Let O(z) denote the C∗ orbit of z, and let OR (z) denote the gradient trajectory of z. If
OR (z) ∩ H −1 (E) 6= ∅, then they intersect at the unique point zE (23).
Lemma 2.3. If the C∗ orbit of z intersects H −1 (E), let z = eτE (z) · zE where τE (z) ∈ R and zE ∈ H −1 (E),
then
bE (z) = ϕ(z) − ϕ(zE ) − τE (z)∂ρ ϕ(zE ).
(27)
∂ρ bE (z) = ∂ρ ϕ(z) − ∂ρ ϕ(zE ),
∂ρ2 bE (z) = ∂ρ2 ϕ(z)
(28)
Hence, bE (z) is a strictly convex function in z, with minimum at z = zE and bE (zE ) = 0.
Proof. By Lemma 2.2, we have H(z) = 12 ∂ρ ϕ(z), hence 2E = ∂ρ ϕ(zE ). Hence from (9), we have
Z τE (z)
Z 2τE (z) h
i
H(e−σ/2 · z) · dσ = −2EτE (z) + 2
[H(eσ · zE )] · dσ
bE (z) = −2EτE (z) +
0
=
−τE (z)∂ρ ϕ(zE ) +
Z
0
τE (z)
0
[∂ρ ϕ(eσ · z)] · dσ = ϕ(z) − ϕ(zE ) − τE (z)∂ρ ϕ(zE ).
The other two identities follow from ∂ρ τE (z) = 1 and ∂ρ ϕ(zE ) = 0.
From Lemma 2.2 and the fact that ∇H = 2π∂ρ (see §2.4), we get
Corollary 2.4. π −1 ∂ρ2 ϕ = |∇H|2 = |ξH |2 .
Remark: A closely related formula, used below, is that
Z σ
Z σ
Z σ
d
d
∇H(es (z0 )) · es z0 ds = π
|∇H(es (z0 ))|2 ds.
H(es z0 )ds =
H(eσ z0 ) − H(z0 ) =
ds
0
0
0 ds
(29)
12
STEVE ZELDITCH AND PENG ZHOU
2.6. The Leafwise Symplectic Potential and bE . In this section we relate bE (z) to leaf-wise symplectic
potentials. To define the symplectic potentials we use slice-orbit coordiantes (θ, ρ, y) as in (22) and pull
back the Kahler potential (25) to C∗ , by ϕ(ρ, θ; y) = ϕ(eρ+iθ · y). Since the Kahler potential is T invariant,
ϕ(ρ, θ; y) is θ independent and is convex in ρ, and will be denoted by ϕ(ρ; y) relative to a slice SE ⊂ H −1 (E).
Note that ϕ(ρ; eiθ y) = ϕ(ρ; y) and so ϕ(ρ, zE ) is defined for all zE ∈ H −1 (E).
The leafwise symplectic potential is defined to be the Legendre transformation of ϕ(ρ; y),
u(I; y) = sup(Iρ − ϕ(ρ; y))
(30)
ρ∈R
Since ϕ(ρ; y) is a smooth convex function in ρ, we have
u(I; y) = Iρ(I; y) − ϕ(ρ(I; y); y),
where ρ(I; y) is s.t. I = ∂ρ ϕ(ρ(I; y); y).
(31)
The Legendre transformation is an involution,
ϕ(ρ; y) = sup(ρI − u(I; y)) = ρI(ρ; y) − u(I(ρ; y); y),
I∈R
where I(ρ; y) is s.t. ρ = ∂I u(I(ρ; y); y)
(32)
Also their second derivatives are related by
∂I2 u(I; y) = ∂I ρ(E; y) =
1
1
= 2
∂ρ I(ρ; y)
∂ρ ϕ(ρ; y)
(33)
where I = I(ρ; y) and ρ = ρ(I; y). We use the notation I as in the “action-variable’ dual to the angle variable
θ; (31) implies that I/2 lies in the range of H (see Lemma (2.6)).
The following Lemma relates bE (z) with the symplectic potential, and can be easily verified using Lemma
2.3.
Lemma 2.5. Let zE ∈ H −1 (E) and use gradient flow-coordinates z = eρ · zE . Then
bE (z) = −u(2H(z), zE ) − ϕ(zE ) + 2(H(z) − E)τE (z),
(34)
Using the symplectic potential, one can easily derive the dependence of bE (z) in terms of E for fixed z.
Lemma 2.6. Fix z ∈ M and E ∈ H(M ), such that the R∗ orbit of z intersects H −1 (E) at zE , and let
τ (z, E) = τE (z) ∈ R be such that eτE (z) · zE = z. Then b(z, E) = bE (z) can be written as
b(z, E) = ϕ(z) + u(2E; z)
(35)
and we have the following properties
∂E b(z, E) = −2τ (z, E),
2
∂E
b(z, E) =
4
∂ρ2 ϕ(zE )
.
(36)
Hence b(z, E) is a strictly convex function in E with minimum being 0 at E = H(z).
Proof. First we claim that ρ(2E; z) = −τ (z, E). Indeed by the definition of ρ(2E; z) in (31), we have
2E = ∂ρ ϕ(ρ(2E; z); z) = ∂ρ ϕ(eρ(2E;z) · z) = 2H(eρ(2E;z) · z)
and τ (z, E) by definition satisfies E = H(zE ) = H(e−τ (z,E) · z), hence ρ(2E; z) = −τ (z, E). Using (31) we
have
u(2E; z) = 2Eρ(2E; z) − ϕ(ρ(2E; z); z) = 2E(−τ (z, E)) − ϕ(e−τ (z,E) · z) = −2τ (z, E)E − ϕ(zE )
Combined with (27), this proves (35). Next, using ρ = ∂I u(I(ρ; y); y) from (32), we have
∂E (b(z, E)) = 2∂I u(I; z)|I=2E = 2ρ(2E; z) = −2τ (z, E),
and using (33) we have
2
∂E
(b(z, E)) = 4∂I2 u(2E; z) =
4
∂ρ2 ϕ(ρ(2E; z); z)
=
4
∂ρ2 ϕ(zE )
.
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
13
2.7. Periodicity of the lifted flow. We can now prove the periodicity statement in §1.3. Recall that the
h
contact vector field is ξ̂H = ξH
− 2πR.
Lemma 2.7. The lifted flow exp tξˆH is 1-periodic, or equivalently, Uk (θ) is 2π-periodic.
Proof. The equivalence follows from Lemma 1.3. By our choice of generator, the common period of all ξH h
orbits equals 1, hence flow by ξH
return to the same fiber. Since on the circle bundle, the horizonal vector
h
field ξH and vertical Reeb vector field R commute, and H(z) is constant along the ξH orbit, we may first
h
h
flow by ξH
for time 1, then by −2πHR for time 1. Let θγ be defined such that exp(ξH
)(z, λ) = (z, eiθγ λ).
Then,
Z 2π
Z
h−∂ϕ, ∂θ idθ = 2πH
θγ = i A = i
γ
0
where we used identities from (26). Hence flowing by −2πHR sends (z, λ) 7→ (z, e−i2πH λ) = (z, e−iθγ λ), the
two eiθγ factor cancels, hence ξ̂H has period 1.
2.8. Examples. To illustrate the variety of S 1 -Kähler manifolds, we first start with model linear cases
and then proceed to other types of examples.
(0): Linear actions on Cm On the non-compact Kähler manifold Cm with Euclidean metric the linear S 1
actions have the form,
eiθ · (Z1 , . . . , Zm ) = (eib1 Z1 , . . . , eibm Zm ), bj ∈ Z,
P
2
with Hamiltonians H =
j bj |Zj | . Extreme cases are where all bm = 0 except b1 = 1, in which case
the fixed point set is a hypersurface Z1 = 0, and the isotropic Harmonic oscillator with all bj = 1 and
Hamiltonian |Z|2 with fixed point set {0}.
(i) Standard S 1 actions on CPm They arise from subgroups S 1 ⊂ SU (m + 1) of the form
eiθ · [Z0 , . . . , Zm ] = [eib0 Z0 , . . . , eibm Zm ],
bj ∈ Z.
With no loss of generality it is assumed that b0 = 0. When bj 6= bk when j 6= k, the action has m + 1 isolated
fixed points, Pj = [0, . . . , 0, zj , 0, . . . , 0]. The weights at Pj are {bj − bi }j6=i . The Hamiltonian is
µ~b ([Z0 : · · · : Zm ]) =
b1 |Z1 |2 + · · · + bm |Zm |2
.
|Z|2
(ii) Cubic hypersurface in CP4
This example is taken from [Ki2]. Consider the cubic hypersurface X ⊂ CP4 ,
x3 + y 3 + z 3 = u2 v,
in CP4 = {[x, y, z, u, v]} and let C∗ act on X via the action on CP4 ,
t · [x, y, z, u, v] = [t−1 x, t−1 y, t−1 z, t−3 u, t3 v].
Then X T has three fixed point components,
F1 = {[0, 0, 0, 1, 0]}, F2 = {[x, y, z, 0, 0] : x3 + y 3 + z 3 = 0}, F3 = {[0, 0, 0, 0, 1]},
of which two (F1 , F3 ) are isolated fixed points and F2 is a hypersurface in X, i.e. a curve. The point
P = [0, 0, 0, 0, 1] is singular.
The corresponding stable sets Sj = {x ∈ X : limt→0 t · x ∈ Fj } are

S1 = {[x, y, z, u, v] ∈ X : u 6= 0},





S2 = {[x, y, z, u, v] ∈ X : u = 0, (x, y, z) 6= 0},





S3 = {[0, 0, 0, 0, 1]},
Here, S1 is Zariski open in X, S2 is of codimension one, and S3 = F3 is a point.
The Hamiltonian H : X → R is the restriction of the Hamiltonian for the T action on CPm above.
14
STEVE ZELDITCH AND PENG ZHOU
(iii) Ruled surfaces [HS]
Let Mg be a Riemann surface of genus g, equipped with a constant curvature metric. Let L → M be
a holomorphic line bundle. L carries a natural C∗ action. Projectivize each line Lz → PLz ≃ CP1 to get
PL. It still carries a C∗ action. Examples of S 1 -invariant Kähler metrics are the constant scalar curvature
metrics.
3. The Szegö kernel and the Boutet de Monvel-Sjöstrand parametrix
This section is preparation for Theorem 0.1 and the subsequence asymptotic results. The equivariant
Bergman kernels Πk,j have two positive integer indices, indicating a lattice point in Z+ × Z+ 4. The asymptotics in k for a fixed energy level E implicitly involve pairs (jn , kn ) of lattice points along a “ray” in the joint
lattice. The simplest rays are the “rational rays” where j/k ∈ Q ∩ [0, 1). Somewhat more complicated are
/ Q. For these one needs to have control over the approximation | kj − E|.
“irrational rays” where kjnn → E ∈
To deal with these issues, it seems cleanest to lift the sections of H 0 (M, Lk ), resp. the equivariant kernels
Πk,j , as equivariant functions (resp. kernels) on the unit circle bundle Xh → M associated to the Hermitian
line bundle (L∗ , h∗ ), see §1.2. This circle bundle carries a canonical S 1 action. The Hamiltonian T action
also lifts to Xh and thus the two commuting circle actions define an S 1 × S 1 action, whose weights form the
semi-lattice of {(j, k) ∈ Z+ × Z+ }. This lifting and the approximation of energy levels by rays in Z+ × Z+ is
discussed in detail in [STZ] for toric varieties, and the same discussion applies almost verbatim to T actions.
We therefore summarize the key points and refer to [STZ] for further details.
3.1. The Szegö kernel and the Bergman kernel. We now discuss the analytic aspects of the lift to the
circle bundle Xh and the disc bundle Dh in §1.2 and §1.3. We define the Hardy space H2 (Xh ) ⊂ L2 (Xh ) of
square-integrable CR functions on Xh , i.e., functions that are annihilated by the Cauchy-Riemann operator
∂¯b and are L2 with respect to the inner product
Z
(37)
F1 F2 dVX , F1 , F2 ∈ L2 (X) .
hF1 , F2 i =
X
Equivalently, H2 (X) is the space of boundary values of holomorphic functions on D that are in L2 (X). Here,
Xh is given the contact volume form
m
1 α
α
ωm
dα
dVX =
=
∧
∧ dVM , where dVM =
.
(38)
m! 2π
2π
2π
m!
L∞
The S 1 action on X commutes with ∂¯b ; hence H2 (X) =
H2 (X) where H2 (X) = {F ∈ H2 (X) :
k=0
k
k
F (rθ x) = eikθ F (x)}. As mentioned in §1.3, a section sk of Lk determines an equivariant function ŝl on L∗
by the rule
ŝk (λ) = λ⊗k , sk (z) , λ ∈ L∗z , z ∈ M ,
where λ⊗k = λ ⊗ · · · ⊗ λ. We henceforth restrict ŝ to X and then the equivariance property takes the form
ŝk (rθ x) = eikθ ŝk (x). The map s 7→ ŝ is a unitary equivalence between H 0 (M, Lk ) and Hk2 (X). (This follows
from (38)–(37) and the fact that α = dθ along the fibers of π : X → M .)
We define the (lifted) Szegö kernel Π̂(x, y) to be the (Schwarz) kernel of the orthogonal projection Π̂k :
L2 (X) → H2 (X). It is given by
Z
Π̂F (x) =
Π̂(x, y)F (y)dVX (y) , F ∈ L2 (X) .
(39)
X
2
The Fourier components Π̂k : L (X) → Hk2 (X) of the Szegö projector can be extracted from Π̂(x, y) by
Z 2π
dθ
(40)
Π̂k (x, y) =
e−ikθ Π̂(rθ x, y)
2π
0
The Szegö (or Bergman)5 kernel Πk (z, w) for the orthogonal projection Πk : L2 (M, Lk ) → H 0 (M, Lk ) can
be obtained via the isometry of H 0 (M, Lk ) ∼
= Hk2 (X).
4The T action in general has Z weights, since we have chosen H such that H(M ) ≥ 0, the corresponding weight are in Z .
+
5In the setting of line bundles, we use the terms interchangeably.
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
15
In a local coordinate patch U with a holomorphic frame eL ∈ Γ(U, L), we introduce two scalar kernels
Kk (z, w) and Bk (z, w), with respect to the holomorphic frame and unitary frame:
Πk (z, w) =: Kk (z, w) · ekL (z) ⊗ ekL (w)
=: Bk (z, w) ·
ekL (w)
ekL (z)
⊗
kekL (z)kh
kekL (w)kh
The Bergman density function Πk (z) is the contraction of Πk (z, w) with the hermitian metric on the diagonal,
Πk (z) := Bk (z, z)(:= Πk (z, z)),
(41)
where in the second equality we record a standard abuse of notation in which the diagonal of the Szegö
kernel is identified with its contraction.
From the parametrix for Π̂ one can derive semi-classical parameterices for the Fourier components and
thus for the semi-classical Szegö kernels on H 0 (M, Lk ) (see [Ze1], or [BBSj] for a direct construction).
3.2. Equivariant Szegö kernels. In this section, we provide some elementary facts about the equivariant
Szegö kernels (4).
Let eL is a local T-invariant holomorphic frame and define

k
ek (w)
(z)

 Πk,j (z, w) = Kk,j (z, w) · ekL (z) ⊗ ekL (w) = Bk,j (z, w) · keekL(z)k
⊗ kekL(w)k ,
h
h
L
L
(42)


Πk,j (z, z) = Πk,j (z) = Bk,j (z, z).
In Theorem 0.1, resp. Theorem 0.3, we give exponentially decaying asymptotics of equivariant Szegö
kernels, resp. partial Bergman kernels, on the far off-diagonal, i.e. on neighborhoods of the (anti-)diagonal
in M × M which do not shrink as k → ∞. The question arises, to which do such results require real
analyticity of the metric and in which neighborhoods do they hold? A crucial point is that although the
asymptotics pertain to off-diagonal pairs (z, w), the points z, w lie on the same C∗ orbit. The following result
shows that we have exponential off-diagonal asymptotics as long as z, w lie on the same C∗ orbit.
Lemma 3.1. For all α, β ∈ C, we have
and
Kk,j (eα · z, eβ · w) = ej(α+β̄) Kk,j (z, w)
α
Bk,j (eα · z, eβ · w) = ej(α+β̄)−k(ϕ(e
·z)−ϕ(z))/2−k(ϕ(eβ ·w)−ϕ(w))/2
Bk,j (z, w)
Proof. This is immediate from the definition of K and B.
In Propositions 4.1 and 4.3, this will be used to obtain sharp exponential asymptotics for equivariant
Bergman kernels for general C ∞ metrics.
3.3. The Boutet de Monvel-Sjöstrand parametrix. Near the diagonal in Xh × Xh , there exists a
parametrix due to Boutet de Monvel-Sjöstrand [BSj] for the Szegö kernel of the form,
Z
e−σψ̃(x,y) χ(x, y)s(x, y, σ)dσ + R̂(x, y).
(43)
Π̂(x, y) =
R+
Here, χ(x, y) is a smooth cutoff supported in a neighborhood of the diagonal that is discussed below. When
the Kähler metric ω is real analytic, the phase ψ̃ is constructed from the Kähler potential φ(z) of ω0 by
ψ̃(x, y) = ψ̃((z, λ), (w, µ)) = 1 − λµ̄eψ(z,w̄)
where ψ(z, w̄) is the analytic extension of φ(z) = ψ(z, z̄) into the complexification M × M̄ of M .
s∼
∞
X
(44)
6
Also,
σ m−n sn (x, y)
n=0
6ψ(z, w̄) is holomorphic in z, anti-holomorphic in w; the bar over w indicates that the function is holomorphic in w̄.
(45)
16
STEVE ZELDITCH AND PENG ZHOU
is an analytic symbol in the sense of [Sj]. Finally, the remainder term R̂(x, y) is real analytic in a neighborhood
of the diagonal. If ω is only C ∞ then ψ(z, w) is defined by an almost-analytic extension and the remainder
R is C ∞ .
If we substitute the first term of (43) into (40), one obtains the oscillatory integral,
Z Z 2π
Π̂k (x, y) ∼
eσψ(x,rθ y) eikθ s(x, rθ y, σ)dθdσ,
(46)
R+
0
Changing variables σ → kσ and eliminating the dθdσ integral by the stationary phase method gives, at least
formally, the off-diagonal expansion for the full Szegö kernel on M ,
Kk (z, w) = ekψ(z,w̄) k m (1 + k −1 a1 (z, w̄) + · · · ).
(47)
We view the expansion as defining a semi-classical parametrix for the Szegö kernel Πk (z, w) near the diagonal.
A direct construction of the parametrix is given in [BBSj] (where (47) is stated as (2.2)).
3.4. Analyticity thresholds. Theorem 0.3 gives exponentially decaying asymptotics in the forbidden
region and one may suspect that such accurate asymptotics require real analyticity assumptions on ω. In
fact, at least to leading order the exponentially decaying asymptotics are proved for general C ∞ metrics in
§7. The reason why this is possible stems ultimately from Lemma 3.1. Namely, although the proof uses the
off-diagonal of the Bergman kernel it only uses it along the C∗ orbits, and essentially only along R+ orbits.
But equivariance gives the off-diagonal asymptotics along these orbits. Real analyticity is assumed in §6
and appears naturally in the Boutet-Sjoestrand parametrix construction, so we briefly compare Bergman
kernels in the C ∞ and real analytic cases. This is not the place for a detailed discussion of Bergman kernels
in the real analytic setting, so we do not provide proofs of the statements (referring to [Sj, Kan] for further
discussion).
When the Kähler metric is real analytic, the phase ψ(z, w) of (47) (see also (43)) is the holomorphic
extension of the Kähler potential φ to a neighborhood of the anti-diagonal in M × M̄ , and coefficients
ai (z, w) are holomorphic functions in a neighborhood of the (anti-)diagonal. In the general C ∞ case, ψ is
the almost-holomorphic extension of the Kähler potential φ. The series (47) is the stationary phase expansion
of the integral in (43) and the remainder after truncating at the N th term is O(|ekψ(z,w̄ |k m−N ). However, to
obtain an expansion for the actual Szegö kernel Πk (z, w) one also needs a remainder estimate on the Fourier
coefficients of R̂(x, y) in (43). In the real analytic case, R̂(x, y) is analytic in a neighborhood of the diagonal
and its Fourier coefficients decay at the rate e−kδ where δ is the width of the neighborhood.
The integral in (43) is formally equal to the sum
ψe−m
m
X
n=0
sn (x, y)ψen + log ψe
∞
X
n=0
sn ψen .
(48)
According to [Kash] (above Theorem 0), the infinite series converges.7 It is possible to show that the series
converges sufficiently close to the “characteristic conoid” ψe = 0 using the Cauchy majorant method and the
fact that s is an analytic symbol (see [Kan] and [Sj] for further discussion). The difference between (48) and
the Bergman kernel, which we call the remainder term, is real analytic in a neighborhood {d(x, y) < δ} of
the diagonal (with respect to the distance function on Xh ).
Definition: We define the analytic radius ρa to be the largest ρ > 0 such that all of the following conditions
hold:
• The Kähler potential φ(z) admits an analytic continuation ψ(z, w) to the neighborhood r(z, w) < ρa
in M × M .
• The remainder R(z, w) in (46) is real analytic for r(z, w) < ρa .
Further we define:
Definition: A point z is within the analytic threshold with respect to an energy level E or an interval P if
7It also appears to be said that the series converges to the Bergman kernel, but this is not correct; Bergman kernels for
hyperbolic quotients show that there may exist a non-zero real analytic remainder.
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
17
• r(z, H −1 (E)) < ρa , resp. r(z, H −1 (P )) < ρa ;
• bP (z) < δ, where δ is defined in Proposition 3.2.
Here, r is the distance function of the Kähler metric. We choose the cutoff χ(x, y) to be equal to 1
in r(z, w) ≤ ρa and zero on a slightly larger neighborhood. The following Proposition is proved by using
analytic stationary phase on the first term and standard Paley-Wiener estimates on Rk (z, w).
Proposition 3.2. Let h be a real analytic T-invariant metric on L → M and let φ be the ‘open-orbit’
i
¯
Kähler potential of ω = 2π
∂ ∂ϕ.
If ϕ(z) admits an analytic extension ψ(z, w) to a δ-neighborhood of the
(anti-)diagonal in M × M , then in this neighborhood,
Πk (z, w) = ekψ(z,w) Ak (z, w) + Rk (z, w),
(49)
where
Ak (z, w) ∼ k m
∞
X
aj (z, w)k −j
j=0
is a semi-classical analytic symbol of order m with a0 = 1 and where Rk (z, w) is holomorphic in the neighborhood and |Rk (z, w)| ≤ Ce−δk .
3.5. Spectral interpretation of vanishing order: Proof of Propositon 0.9. The following is very
similar to Lemma 5.4 of [RS]:
Proposition 3.3. Let Y be the minimum set of H and assume it is a complex hypersurface.
H 0 (X, O(Lk ) ⊗ IYtk ) is the direct sum of eigenspaces for eigenvalues of Ĥk of eigenvalue ≥ t.
Then
Proof. Since Y is T-invariant, both Lk and IYtk are T invariant and so the action fo T on H 0 (X, O(Lk )⊗IYtk )
decomposes into weight spaces,
M
Vk (j) ⊗ IYtk .
H 0 (X, O(Lk ) ⊗ IYtk ) =
j: kj ∈P0
Thus it suffices to determine the summands which are non-zero. An element s ∈ Vk (j) transforms by wj
under the action of w ∈ C∗ . We restrict it to C∗ orbit which tends to a point y ∈ Y . In holomorphic
coordinates (w, y) where w = 0 on Y , it is given by cy wj . Thus it vanishes to order ≥ tk if and only if
j ≥ tk.
3.6. Equivariant kernels and symplectic reduction. Equivariant Bergman kernels are closely related
to Bergman kernels for the reductions of the level sets H −1 ( kj ). For instance, the space of invariant sections
0
Vk (0) := HT
(M, Lk ) = {s ∈ H 0 (M, Lk ) : eiθ s = s}.
(50)
is isomorphic in a canonical way to the space of holomorphic sections of the reduced line bundle LT on the re0
duced space ME := H −1 (E)/S 1 , i.e. Vk (0) ≃ H 0 (ME , LkT }, so that dim HT
(M, Lk ) = Vol(H−1 (E)/S1 ) km−1 .
One might expect partial Bergman kernels to be closely related to Bergman kernels of Kähler cuts [BGL].
4. Equivariant Bergman kernels: Proofs of Theorem 0.1 and Theorem 0.2
In this section, we prove that the equivariant Bergman
kernel Πk,j (z, z) forms a Gaussian bump around
√
the hypersurface H −1 (j/k), with decay width ∼ 1/ k.
18
STEVE ZELDITCH AND PENG ZHOU
4.1. Proof of Theorem 0.1. Theorem 0.1 follows from the following two propositions. The first one
establishes decay property of Πk,j (z) away from the real hypersurface H −1 (j/k). Recall the definition of
bE (z) (9).
Proposition 4.1. Let ω be a C ∞ Kähler metric, and let H generate a holomorphic Hamiltonian S 1 action.
Fix (k, j) and z ∈ H −1 (E) where E = j/k. Then, for any α ∈ R, we have
α
Πk,j (eα · z) = e−kbE (e
Proof. Using Lemma 3.1 with z = w, α = β ∈ R, we have
z)
α
Bk,j (eα · z, eα · z) = e2jα−k(ϕ(e
Πk,j (z)
·z)−ϕ(z))
Bk,j (z, z)
Now, write j = kE = kH(z) = k∂ρ ϕ/2, we have
α
Πk,j ((eα ·z) = Bk,j (eα ·z, eα ·z) = e−k(ϕ(e
·z)−ϕ(z)−α∂ρ ϕ(z))
α
Bk,j (z, z) = e−kbE (z) Bk,j (z, z) = e−kbE (e
z)
Πk,j (z)
Remark: This proof does not use the parametrix.
The next proposition studies the kernel when z ∈ H −1 (j/k).
Proposition 4.2. Let ω be a C ∞ Kähler metric, and let H generate a holomorphic Hamiltonian S 1 action.
Fix E ∈ H(M ) and a sequence {jk } such that | jkk − E| < C/k. Let z ∈ H −1 (E), then we have
s
2
m−1/2
(1 + O(1/k))
Πk,j (z) = k
π∂ρ2 ϕ(z)
Proof. Let Ek = j/k, and zk ∈ H −1 (Ek ), such that z = eρk zk . We have |ρk | = O(1/k). Indeed,
Z ρk
1
1
C/k > |Ek − E| = |∂ρ ϕ(zk ) − ∂ρ ϕ(z)| = |
∂ρ2 ϕ(es z)ds| > C ′ |ρk |
2
2 0
where we used the fact ϕ is psh and T-invariant, to get ∂ρ2 ϕ strictly positive, hence |ρk | = O(1/k). Then
using Proposition 4.1, we get
ρk
zk )
Πk,j (z) = Πk,j (zk )e−kbEk (e
2
= Πk,j (zk )e−kO(ρk ) = Πk,j (zk , zk )(1 + O(1/k))
Next, we evaluate Πk,j (zk , zk ) using the stationary phase method.
Z 2π
dθ
Kk (eiθ · z, z)e−ijθ
Kk,j (z, z) =
2π
0
Z 2π
iθ
dθ
m
= k
ekψ(e z,z) e−ijθ (1 + O(1/k))
2π
Z0 π
iθ
dθ
= km
ek(ψ(e z,z)−iEk θ) (1 + O(1/k))
2π
−π
We claim that θ = 0 is the unique critical point. Indeed, using the T invariance of ϕ and ψ, we have
ψ(eiθ z, z) = ψ(eiθ/2 z, e−iθ/2z). Define the function Ψ : C → C by
Ψ(τ ) = ψ(eτ /2 · z, eτ̄/2 · z) − Eτ
(51)
Then the phase function is
ψ(eiθ z, z) − iEk θ = Ψ(iθ).
To verify the claim that θ = 0 is a critical point for Ψ(iθ), suffice to check Ψ′ (0) = 0. From the restriction
of Ψ to R, we have, for all t ∈ R,
Ψ(t) = ψ(et/2 z, et/2 z) − tE = ϕ(et/2 z) − tE = ϕ(z) +
t2 2
∂ ϕ(z) + O(t3 )
8 ρ
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
19
Hence θ = 0 is a critical point of Ψ(iθ), with Hessian ∂θ2 |θ=0 Ψ(iθ) = −Ψ′′ (0) = − 14 ∂ρ2 ϕ(z). Applying the
standard stationary phase method, we get
s
2
kϕ(z) m−1/2
Kk,j (z, z) = e
k
(1 + O(1/k))
π∂ρ2 ϕ(z)
The above two propositions finish the proof of Theorem 0.1.
Remark: If ϕ(z) is real analytic, then Ψ(τ ) is holomorphic when Im(τ ) is small enough. If ϕ is only smooth,
then Ψ(τ ) is an almost analytic extension of Ψ|R . Although the proof uses the parametrix, it only uses Ψ in
the real domain and only uses the C ∞ remainder. Hence, it does not require real analyticity.
4.2. Proof of Theorem 0.2. Theorem 0.2 follows from the following
Proposition 4.3. Let ω be a C ∞ Kähler metric, and let H generate a holomorphic Hamiltonian S 1 action.
Then for any fixed k, j, z ∈ H −1 (j/k) and α ∈ R, we have
Πk,j (eα/
√
k
· z) = e−
α2
2
∂ρ2 ϕ(z)
Πk,j (z)(1 + O(k −1/2 ))
Proof. This follows from Proposition 4.1, and Lemma 2.3. We Taylor expand bE (eα z) around α = 0, to get
bE (eα z) =
α
then we plug in to the exponent e−kbE (e
Πk,j (e
√
α/ k
·z)
α2 2
∂ ϕ(z) + g3 (z, α),
2 ρ
(52)
.
α/
· z) = e−kbE (e
√
k
·z)
Πk,j (z)
= e
2
−k( α
2k
= e
2
− α2
∂ρ2 ϕ(z)
α
Πk,j (z)(1 + kg3 (z, √ ))
k
α2
2
∂ρ2 ϕ(z)
Πk,j (z)(1 + Oz (k −1/2 ))
= e−
∂ρ2 ϕ(z)+g3 (z, √αk )
Πk,j (z)
5. Partial Bergman Kernel: Localization of sums
As above, let ω be a C ∞ Kähler metric, and let H generate a holomorphic Hamiltonian S 1 action. In
this section we consider the sums the partial Bergman kernels (6) and the related sums (10) and (13). We
prove several localization formulae for these sums. Roughly speaking a localization formula that, for a given
M
contribute to the leading order asymptotics. We break up
z, only terms in the sums with | kj − H(z)| < √
k
E
into the cases where z is a general element of Mmax
and where z is √Ck -close to H −1 (E). In the first two
√ j
statements we assume the weights have the form f ( k( k − E)) for applications to Theorem 0.4, but the
proofs are valid for more general weights as stated in (3) below.
We start by proving a localization result for rather wide windows of indices and then proceed to sums
over narrower windows of indices.
1
5.1. Wide sums: | kj − H(z)| ≤ k − 2 +ǫ . We fix a standard smooth cut-off function χ : R → [0, 1], such that
χ(x) = 1 for |x| ≤ 1 and χ(x) = 0 for |x| ≥ 2.
Lemma 5.1 (Localization in spectral sum). Fix a generic z ∈ M that is not a fixed point of T, and an
arbitrary small 1/2 ≫ ǫ > 0, we have
X |j/k − H(z)|
Πk,j (z) = O(k −∞ )
(53)
1−χ
k −1/2+ǫ
j/k∈H(M)
20
STEVE ZELDITCH AND PENG ZHOU
Proof. From Proposition 4.1, we have
X |j/k − H(z)|
1−χ
Πk,j (z) =
k −1/2+ǫ
j/k∈H(M)
−τ (z,j/k)
|j/k − H(z)|
e−kb(z,j/k) Πk,j (zj )
1−χ
k −1/2+ǫ
X
j/k∈H(M)
−1
where zj = e
· z ∈ H (j/k). If j/k > H(z) and in the support of 1 − χ then monotonicity of b(z, E)
in E for E > H(z) (Lemma 2.6), we have
1 2
kb(z, j/k) > kb(z, H(z) + k −1/2+ǫ ) = ∂E
b(z, H(z))k 2ǫ + O(k −1/2+3ǫ ) > Ck 2ǫ
2
Similar result holds for j/k < H(z) in the support of 1 − χ. Hence
X 2ǫ
|j/k − H(z)|
e−kb(z,j/k) Πk,j (zj ) = O(k m+1 e−Ck ) = O(k −∞ )
1−χ
k −1/2+ǫ
j/k∈H(M)
5.2. Narrow windows. In this section we prove a localization formula for more general weights and narrower windows of indices.
E
Proposition 5.2. Let z ∈ Mmax
. Then:
(1) For any R there exists M such that
X √ j
f ( k( − E))Πk,j (z) =
k
j
X
j
−H(z)|<
j:| k
√ M log k
√ j
f ( k( − E))Πk,j (z) + O(k −R ).
k
k
(2) Let z0 ∈ H −1 (E). For any ǫ > 0 there exists M so that
√
√ j
P
−m
Πk,j (eβ/( k) z0 )
j f ( k( k − E))k
=
P
j
−E|≤
j:| k
√
√
√ M f ( k( j − E))k −m Πk,j (eβ/( k) z0 ) + Rk (z0 , β, M ), where |Rk (z0 , β, M)| ≤ ǫ.
k
(54)
k
(3) The same localization formulae are valid for weights of the form f ( kj ).
Proof. Proof of (1) For each j in the sum, write z0 = z0 (j, k) := e−τj/k (z)/2 z ∈ H −1 (j/k). By Proposition
4.1,
τj/k (z)/2
z0 )
Πk,j (eτj/k (z)/2 · z0 ) = e−kbj/k (e
Πk,j (z0 ).
By Theorem 0.1 we then get,
q
τj/k (z)/2
1
2
−1
z0 )
)),
z∈
/ H −1 (E),
(55)
Πk,j (z) = k m− 2 e−kbj/k (e
π∂ρ2 ϕ(z0 ) (1 + O(k
where ∂ρ2 ϕ(z0 ) = |∇H(z0 (j, k)|.
By (27),
bj/k (z) =
Z
0
τj/k (z)
σ −τj/k (z)
(H(e e
z) − j/k)dσ =
Z
0
τj/k (z) Z σ
0
|∇H(es · e−τj/k (z) (z)|2 dsdσ,
(56)
If H(z) = E then bE (z) = 0 and bj/k increases as |E − j/k| increases.
We now claim that for any ǫ > 0 there exists δ > 0 so that if |H(z) − j/k| ≥ ǫ then bj/k (z) ≥ δ. Indeed,
it is visible from (56) that bj/k (z) increases as τj/k increases or equivalently as |H(z) − j/k| increases.
It follows first that for any ǫ > 0 there exists δ > 0 so that
X
X √
√
f ( k(j/k − E))Πk,j (z, z) + O(e−δk ).
f ( k(j/k − E))Πk,j (z, z) =
j
j:|j/k−H(z)|<ǫ
By assumption, {H = H(z)} is a regular level set of H, and so there exists a neighborhood of z of the
form {H(z) − ǫ < H < H(z) + ǫ} in which |∇H|2 ≥ γ for some γ > 0. For j/k ∈ [H(z) − ǫ, H(z) + ǫ],t
bj/k (z) ≥ γτj/k (z)2 /2.
(57)
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
21
Moreover, by (29), if |∇H|2 ≤ C then (taking into account the sign of σ),
C|σ| ≥ |H(eσ z0 ) − H(z0 )| ≥ γ|σ|.
(58)
Hence
C|τj/k (z)| ≥ |H((eτj/k (z)/2 · z0 (j, k)) − j/k| ≥ γ|τj/k (z)|,
Combining with (57) gives
bj/k (z) ≥ (
It follows that if
γ 2
) |H((eτj/k (z)/2 · z0 (j, k)) − j/k|2 .
C
(59)
r
(60)
γ M log k
M log k
=⇒ bj/k (z) ≥ ( )2
.
k
C
k
γ 2
If we choose M so that M ( C
) ≥ m + R then by (55),
X
√
f ( k(j/k − E))Πk,j (z, z) ≤ CR k −R ,
√ M log k
|j/k − H(z)| ≥
j:|j/k−H(z)|≥
k
completing the proof of (1).
Proof of (2)
√
To prove (54) we modify the proof of case (1). Since z = eβ/ k z0 , it suffices to use the bounds for τ j (z0 )
k
and b j (z0 ). But we cannot use the logarithm log k in the numerator. Rather, we use we choose
k
r
γ M
j
M
=⇒ b j (z) ≥ ( )2 .
(61)
| − E| ≥
k
k
k
C k
γ 2
Since we divided by k m to normalize the sum, it suffices to choose M so that M ( C
) ≥ log Rǫ and then by
(55),
X
√ j
Rk (z, β, M ) :=
f ( k( − E))k −m Πk,j (z, z) ≤ Ce−Rǫ .
k
√M
j
j:| k −H(z)|≥
k
Here C is a bound on |∇H|−1 and other coefficients in the formula (which are independent of k). By choosing
Rǫ sufficiently large we can make Ce−Rǫ ≤ ǫ. This concludes the proof of (2).
Proof of (3) In the proofs of (1)-(2) we did not use the specific form of f . Only the localization properties
of the kernels Πk,j (z) were used. Hence the same proofs apply to weights f ( kj ).
6. Proof of Theorem 0.3: Euler-MacLaurin approach
In this section we prove Theorem 0.3 by the technique of [ShZ]. It uses what we called “polytope
characters” to sift out the weights in the given interval. In dimension one we refer to these polytope
characters as interval characters. For this result, we need to assume (for the first time) that the metric is
real analytic.
Given a proper rational subinterval P = [ ap , pb ] ⊂ H(M ) the interval characters χkpP defined on (C∗ ) by
X
X
ejw w ∈ C.
(62)
χkpP (ew ) =
ejw =
j∈kpP
j∈k[a,b]
Proposition 0.7 is essentially equivalent to the following
Lemma 6.1. For a proper rational subinterval P ⊂ H(M ),
Z
1
Π|kpP (z, w) =
Πkp (z, eiφ · w)χkpP (eiφ ) dφ ,
(2π) T
(63)
22
STEVE ZELDITCH AND PENG ZHOU
The analysis of the partial Bergman kernels in the forbidden region depends on the asymptotic properties
of interal characters near the real domain, i.e. for w lying in
S(ε) := {w ∈ C : |Imw| < ε}.
The main result is that polytope characters are given by oscillatory integrals over P :
Proposition 6.2. Let P = [ ap , pb ] ⊂ H(M ) be a proper rational subinterval of H(M ). Then there exists
ε > 0 such that the interval characters (62) are oscillatory integrals
Z
1
w
ekhw,xi [A0 (x, w)k + A1 (x, w) + · · · ] dx + (eakw + ebkw ) , for all w ∈ S(ε) ,
χkpP (e ) =
2
[a,b]
where the Al are analytic functions on P × S(ε) that are holomorphic in w and algebraic in x, and
i) Al (x, w) ∈ R whenever w ∈ R,
ii) A0 (x, 0) = 1 and A0 (x, w) = L(w) where
∞
X 1
S/2
=1+
b2k S 2k .
L(S) =
tanh S/2
(2k)!
k=1
Proof. This is proved in Proposition 3.1 of [ShZ] using the Euler-MacLaurin formula for sums over lattice
points. In dimension one the proof is elementary because one has the exact formula (see [KSW]),
Z bk
Z b
′
X
ekwb − ekwa
1
.
(64)
ewα = L(w)
ewx dx = kL(w)
ekwx dx = (ew/2 + e−w/2 ) w/2
2
e
− e−w/2
ak
a
α∈k[a,b]
Here,
′
X
[a,b]
The details are given in §10.2.
f (x) =
X
1
f (j),
(f (a) + f (b)) +
2
j∈(a,b)
Combining Lemma 6.1 and Proposition 6.2, we obtain the following representation,
Proposition 6.3. There exists a semi-classical symbol Ak (z, w, x) of order 1 so that
Z
Z
iθ
1
Π|kpP (z, z) =
eikhθ,xi ekpψ(z,e z) Ak (z, eiθ z, x)dxdθ + (Πkp,ka (z) + Πkp,kb (z)) + R|kpP (z, z), (65)
2
[a,b] T
where the phase ψ(z, w) is the off-diagonal almost analytic extension of the Kähler potential φ of h = e−φ
and where R|kP (z, z) = O(k −∞ ) in the general C ∞ case, resp. O(e−δk ) in the real analytic case.
Proof. By Proposition 6.2, there exists a Fourier integral representation of the partial Bergman kernel in the
form,
1
Πkp (z, eiφ · w) eikxφ [A0 (x, w)k + A1 (x, w) + · · · ] + (eiakφ + eibkφ ) dx dφ .
2
T [a,b]
(66)
The stated formula follows by substituting the Boutet de Monvel-Sjostrand parametrix and remainder for
the Bergman kernels of H 0 (M, Lk ).
Π|kpP (z, w) =
1
(2π)
Z Z
The phase ΨA is given by
ΨA (φ, x; z) = ihφ, xi + ψ(eiφ z, z).
(67)
We evaluate the integral by the method of stationary phase on half-spaces (see §10.1 and [Hö, Ch. 7]). We
first find the critical points of ΨA . By Lemma 2.2, the (interior) critical point equations
dx ΨA = iφ = 0,
yield
dφ ΨA |φ=0 = ix − iH(z) = 0
φ = 0,
x = H(z) .
(68)
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
23
We note that ReΨA ≤ 0 and ΨA = 0 at the critical point. The Hessian HΨA of ΨA (with respect to the
variables φ, x) is of the form


C i
2
 , where C = ∂ ΨA
.
(69)
HΨA |(0,x) = 
∂φ2 (0,x)
i 0
It follows that
det HΨA |(0,x) = 1,
that the inverse Hessian is given by


HΨA |−1
(0,x) =
0
−i
−i
C
(70)

,
and that Hessian operator in the stationary phase expansion is given by
∂2
∂2
+C
.
∂φ∂x
∂x∂x
By [Hö, Theorem 7.7.5], for z ∈ AP , we have a stationary phase expansion of the form
X
1
,
k m+j−ℓ Hℓ g3j
Π|kpP (z, z) ∼
j!ℓ!2ℓ
φ=0,x=H(z)
H = −i
2ℓ≥3j≥0
where g3 is the third order Taylor remainder of the phase ΨA . We note that g3 is a function of φ alone, but
that every φ derivative is accompanied by an x-derivative in H. It follows that only the very first term of
the summation is non-zero.
We now let z ∈ FP be a point in the classically forbidden region. Then the phase has no critical points.
We deform the integral over T by changing eiφ z to eiφ+τ /2 z. By the triangle inequality,
ReΨz (φ, E) < ReΨz (0, E) ,
for t 6= 0 .
Thus for each fixed τ and z, the maximal value of ReΨ occurs where t = 0. Then
dt Ψz = −iE + dφ ψ e−τ −iφ z, z .
(71)
(72)
Hence the critical-point equation is equivalent to:
E = H(e−τ /2 · z), dE Ψz = τ − iφ = 0.
(73)
To obatain a critical point at which ReΨ is maximal, we need to choose τ so that
q(z) := e−τ /2 z ∈ ∂P.
As in [ShZ], we then decompose the integral (65) into two parts:
Z Z
Π|kpP (z, z) = Ik′ + Ik′′ ,
Ik′ = k
ekpΨ(φ,x;τz ,z) ρ(x)A(φ, x, k; τz ) dx dφ ,
R
(74)
P
where ρ is supported in a small neighborhood of q(z) and is ≡ 1 near q(z). It follows from (72) that
dφ Ψ(0, x; τz , z) 6= 0 for x 6= q(z). Since dφ Ψ does not vanish on the support of 1 − ρ(x) and supR×P Ψ =
Ψ(0, q(z)) = −bP (z), we conclude by performing the φ integration first, that
Ik′′ = e−kpbP (z) · O(k −∞ )
Ik′ .
(75)
(see [Hö]). Hence, we need consider only
We evaluate Ik′ by applying stationary phase on a manifold with boundary as in §10.1. Since dφ Ψ(0, q(z); τz , z) =
e 0) = 0. The real part of the phase Ψ takes its maximum at e−τz /2 z, and hence
0, it follows that dy Ψ(0,
hypothesis (4) of the lemma holds. Also, hypothesis (2) is satisfied since dφ Ψ(0, x; τz , z) 6= 0 for x 6= q(z),
e y) 6= 0 for y 6= 0. The Hessian is non-degenerate due to the convexity of φ along C∗ orbits.
and hence dy Ψ(0,
The decay function bP (z) is given by
bP (z) = −Ψ(0, q(z)) .
(76)
The expansion
(77)
Ik1 (z) = ekpΨ(0,q(z);τz ,z) c0 + c1 k −1 + · · · + cl k −l + O(k −l−1 ) ,
24
STEVE ZELDITCH AND PENG ZHOU
stated in Theorem 0.3 then follows by Lemma 10.1 and (75). The pre-factor of k cancels due to the integration
1
in θ and integration of x, each of which produces a factor of k − 2 .
1
To complete the proof, we need to add the endpoint terms, 2 (Πak,k (z) + Πbk,k (z)) + R|kpP (z, z). But the
asymptotics of these terms are given by Theorems 0.1 and 0.2.
7. Proof of Theorem 0.6
In this section we prove Theorem 0.6 using Proposition 5.2 and asymptotics of equivariant Bergman
kernels in Theorem
0.1 and Theorem 0.2. Let P = [E1 , E2 ] ⊂ H(M ), and recall the partial Bergman density
P
as Π|kP (z) = j/k∈P Πk,j (z). The advantage of this proof is that it does not use or require real analyticity
of the metric. We refer to Theorem 0.6 for further notation.
We fix a standard smooth cut-off function χ : R → [0, 1], such that χ(x) = 1 for |x| ≤ 1 and χ(x) = 0 for
|x| ≥ 2.
Proof of Theorem 0.6. (Allowed Region). If z is in the allowed region, z ∈ H −1 (int(P)), we may use the
localization formula for the sum to write
X
|H(z) − j/k|
Πk,j (z) + O(k −∞ )
(78)
Πk,P (z) =
χ
k −1/2+δ
j∈kP ∩Z
However, this is the same as the full Bergman kernel, up to another O(k −∞ ) error term.
(Forbidden Region). If z is in the forbidden region, z ∈
/ H −1 (P ), then only terms with |H(z) − j/k|
small will contribute. For each k, let j0 := j0 (z, k, P ) ∈ kP ∩ Z be the integer j such that |H(z) − j/k|
smallest. Let zj ∈ H −1 (j/k) and τj be such that, z = eτj zj . Since we assumed H(z) < H(zj ), τj < 0. Then
using Proposition 4.1, we have
X
e−kb(z,j/k) Πk,j (zj )
Πk,P (z)
=
(79)
e−kb(z,j0 /k) Πk,j0 (zj0 ) j∈kP ∩Z e−kb(z,j0 /k) Πk,j0 (zj0 )
If z < E1 , then H(z) < j0 /k ≤ j/k; if z > E2 , then H(z) > j0 /k ≥ j/k. In both cases, b(z, j/k)−b(z, j0 /k) ≥
0. In fact, for any 1 ≫ ǫ > 0, we claim that
X
Πk,P (z)
|j/k − E0 |
e−kb(z,j/k) Πk,j (zj )
+ O(k −∞ )
(80)
=
χ
k −1+ǫ
e−kb(z,j0 /k) Πk,j0 (zj0 )
e−kb(z,j0 /k) Πk,j0 (zj0 )
j∈kP ∩Z
′
Indeed, consider first the case z < E1 . Let E0 := E1,k
:= j0 /k, then by Taylor expansion of b(z, E) in E,
there exists δ, C > 0, such that ∀|E − E0 | < δ
(2)
(2)
b(z, E) = b(z, E0 ) + (E − E0 )∂E b(z, E0 ) + Rb (z, E, E0 ), |Rb (z, E, E0 )| ≤ C|E − E0 |2
Then if (j − j0 ) > k ǫ , and k large enough such that k −1+ǫ < δ, then
(2)
k[b(z, j/k) − b(z, j0 /k)] > k[b(z, j + k ǫ /k) − b(z, j0 /k)] = ∂E b(z, j0 /k)k ǫ + kRb (z, E0 + k ǫ−1 , E0 )
(81)
(82)
(2)
Since ∂E b(z, j0 /k) = −2τ (z, E0 ) > 0, and kRb (z, E0 + k ǫ−1 , E0 ) < Ck −1+2ǫ ≪ k ǫ , we have
X
|j − j0 |
e−kb(z,j/k) Πk,j (zj )
= O(k −∞ ).
1
−
χ
kǫ
e−kb(z,j0 /k) Πk,j0 (zj0 )
(83)
j∈kP ∩Z
This finishes the proof of the localization claim (80).
Next, we claim that the sum in (80) can be approximated by an infinite geometric series with O(1/k)
error. Indeed, using Proposition 4.2, we have
s
∂ρ2 ϕ(zj0 )
Πk,j (zj )
=
(84)
+ O(1/k) = 1 + R1 (zj , zj0 ) + O(k −1 )
Πk,j0 (zj0 )
∂ρ2 ϕ(zj )
where |R1 (zj , zj0 )| < C|H(zj ) − E0 | = k −1 · C|j − j0 |. And
e−kb(z,j/k)
= e−A(j−j0 ) (1 + R2 (j, j0 )),
e−kb(z,j0 /k)
A = ∂E b(z, E0 ) = 2|τ (z, E0 )|
(85)
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
(2)
where R2 (j, j0 ) = ekRb
=
=
25
(z,j/k,j0 /k)
− 1 < Ck · |j/k − E0 |2 = k −1 · C|j − j0 |2 . Hence we get
X
|j/k − E0 |
e−kb(z,j/k) Πk,j (zj )
χ
k −1+ǫ
e−kb(z,j0 /k) Πk,j0 (zj0 )
j∈kP ∩Z


X
X
|j/k
−
E
|
|j/k
−
E
|
0
0
 ((1 + O(k −1 ))
e−A(j−j0 ) (1 + k −1 R(j − j0 , z0 ))χ
=
e−A(j−j0 ) χ
k −1+ǫ
k −1+ǫ
j∈kP ∩Z
j∈kP ∩Z
X
−A(j−j0 )
−1
−A −1
−1
e
(1 + O(k )) = (1 − e ) (1 + O(k ))
j∈kP ∩Z
where R(m, z0 ) has at most polynomial growth in m, hence is integrable against the exponential decaying
factor. This finishes the proof for the forbidden region case.
(Interface). Finally, we consider the√boundary case, |H(z) − Ei | < Ck −1/2 , i = 1 or 2. Again, assume
z is closed to the left boundary, z = eβ/ k · z0 , where H(z0 ) = zE . The right boundary case is identical. By
the Localization Lemma 5.1
X
|H(z) − j/k|
+ O(k −∞ )
Πk,P (z) =
Πk,j (z)χ
k −1/2+ǫ
j∈kP ∩Z
X
|H(z) − j/k|
=
e−kb(z,j/k) Πk,j (zj )χ
+ O(k −∞ )
k −1/2+ǫ
j∈kP ∩Z
Next we Taylor expand b(z, E) around E = H(z), ∃δ, C > 0, such that for all |H(z) − E| < δ, we have
b(z, E) =
4
|E − H(z)|2
|E − H(z)|2 2
(3)
(3)
∂E b(z, E) + Rb (z, E) =
+ Rb (z, E)
2
2
∂ρ2 ϕ(z)
(86)
(3)
where we have used Lemma 2.6, and |Rb (z, E)| < C|E − H(z)|3 . Define
√
uj = k(j/k − H(z))
(87)
we have
(3)
kb(z, j/k) = A2 u2j + kRb (z, j/k),
A2 =
2
∂ρ2 ϕ(z)
(88)
and
(3)
kRb (z, j/k) < Ck −1/2 u3j < Ck −1/2+3ǫ .
(89)
where we used χ(uj /k ǫ ) > 0 only uj < 2k ǫ . Abbrevating the cut-off function by χj , we have
Πk,P (z)
Πk,j0 (zj0 )
=
X
2
j∈kP ∩Z
=
X
(3)
e−A2 uj ekRb
= 

= 
X
Πk,j (zj )
· χ(uj /k ǫ )
Πk,j0 (zj0 )
2
e
−A2 u2j
e
−A2 u2j
j∈kP ∩Z
X
·
e−A2 uj (1 + k −1/2 R(uj ))χ(uj /k ǫ )
j∈kP ∩Z

(z,j/k)
j∈kP ∩Z

χ(uj /k ǫ ) (1 + O(k −1/2 ))

 (1 + O(k −1/2 ))
where R(uj ) has at most polynomial growth in uj , hence is integrable against the Gaussian decaying factor,
and removing the cut-off only will introduce an exponentially small error. Finally, we replace sum with
26
STEVE ZELDITCH AND PENG ZHOU
√
integral over u. Since uj+1 − uj = k, and the integrand is smooth and has bounded derivative, the
difference between the integral and the summation is again O(k −1/2 )
Z √k(E2 −H(z))
√
2u2
Πk,P (z)
kdu(1 + O(k −1/2 ))
(90)
exp − 2
= √
Πk,j0 (zj0 )
∂
ϕ(z)
k(E1 −H(z))
ρ
Using our assumption that z is closed to the left boundary, we may extend the upper limit of the integral to
∞, with an O(k −∞ ) error. Then
s
!
Z ∞
2
2u2
m
Πk,P (z) = k
exp − 2
du (1 + O(k −1/2 ))
√
π∂ρ2 ϕ(z)
∂ρ ϕ(z)
k(E1 −H(z))
!
s
4k
m
(H(z) − E1 ) (1 + O(k −1/2 ))
= k Φ
∂ρ2 ϕ(z)
where Φ(a) = P(X < a), X ∼ N (0, 1) is the cumulative density of standard Gaussian.
8. Interface asymptotics: Proof of Theorem 0.4
We now give two rather different proofs of Theorem 0.4. The first is Fourier analytic and illustrates
the scaling. The second uses the Gaussian bump asymptotics of the equivariant kernels and localization
formulae.
8.1. Fourier analytic proof.
Proof. First, we assume fˆ ∈ L1 (R). Then by the Fourier inversion formula and Proposition 3.2, f ∈ Cb (R)
and
√ j
P
If (z, k) :=
(z, z)
j f ( k( k√− E))Πk,j√
R
−iE kt
it/ k
ˆ
=
Πk (e
z, z)dt√
R fR(t)e
√
√
√
(91)
dt
m ∞ ˆ
−it( kE) kψ(eit/(2 k) ·z,e−it/(2 k) ·z)−kϕ(z)
= k −∞ f (t)e
e
Ak (eit/2 k z, z) 2π
√
√
R
∞
dt
+ k m −∞ fˆ(t)e−it( kE) Rk (eit/2 k z, z) 2π
√
√
Since Rk ∈ k −∞ C ∞ (M × M ), the second term is O(k −∞ ). We note that t → Πk (eit/ k z, z) is 2π k-periodic
(similarly for the parametrix and remainder terms), so the integrals converge when fˆ ∈ L1 (Rd ).
As in (51), define the function
√
√
√
(92)
Ψ(τ, z) = −τ ( kE) + kψ(eτ /2 k · z, eτ̄/2 k · z) − kϕ(z)
where the phase function in If is Ψ(it). If ϕ(z) is real analytic, then Ψ(τ ) is holomorphic when Im(τ ) is
small enough. If ϕ is only
smooth, then Ψ(τ ) is an almost analytic extension of Ψ|R .
√
Recall that z = eβ/( k) z0 with H(z0 ) = E. Then as k → ∞,
√
√
√
√
√
Ψ(τ, eβ/( k) z0 ) = −τ ( kE) + k ψ(e(τ /2+β)/ k · z0 , e(τ̄ /2+β)/ k · z0 ) − ϕ(eβ/ k · z0 )
"
2
τ /2 + β
1√
1 τ /2 + β
√
√
k∂ρ ϕ(z0 )) + k
= −τ (
∂ρ ϕ(z0 ) +
∂ρ2 ϕ(z0 )
2
2
k
k
#
2
β
1
β
2
√
∂ρ ϕ(z0 ) −
∂ρ ϕ(z0 ) + G3 (τ, z, β)
− √
2
k
k
1
=
((τ /2 + β)2 − β 2 )∂ρ2 ϕ(z0 ) + g3 (z, τ, β),
2
where
G3 = O(k −1/2 (|β|3 + |τ /2 + β|3 ), g3 = O(k −1/2 (|β|3 + |τ |3 )).
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
27
We now make the further assumption that fˆ ∈ Cc (R) (compactly supported continuous functions). Substituting τ = it, using the Plancherel theorem and the Taylor expansion above,
Z ∞
i
h 1
√
2
2
2
If (eβ/( k) z0 , k) = k m
fˆ(t) e 2 ((it/2+β) −β )∂ρ ϕ(z0 ) eg3 dt (1 + O(k −1 ))
−∞
Z ∞
i
h 1
2
2
2
1
fˆ(t) e 2 ((it/2+β) −β )∂ρ ϕ(z0 ) dt + O(k m− 2 ))
= km
−∞
Z ∞
Z ∞
2
2
2
1
dx
1
= km
f (x)
e−itx+ 2 ((it/2+β) −β )∂ρ ϕ(z0 ) dt
+ O(k m− 2 ))
2π
−∞
−∞
s
2 ϕ(z ))2
Z ∞
(2x−β∂ρ
0
−
2
2 ϕ(z )
2∂ρ
0
= km
f (x)
e
dx + O(k m−1/2 ))
2
π∂ρ ϕ(z0 )
−∞
This completes the proof when fˆ ∈ Cc (R).
We then observe that if we multiply If (z, k) by k −m (or more precisely divide by I1 (z, k) to get a
probability measure), then k −m If (z, k) is a bounded linear functional on f ∈ Cb (R) (bounded continuous
functions, equipped with the sup norm). The proof above shows that for f ∈ C0 (R) (vanishing at infinity)
with the property that fˆ ∈ Cc (R),
k
−m
If (e
√
β/( k)
z0 , k) →
Z
∞
−∞
s
f (x)
−
2
e
π∂ρ2 ϕ(z0 )
2 ϕ(z ))2
(2x−β∂ρ
0
2 ϕ(z )
2∂ρ
0
dx.
(93)
We now extend the convergence result in several steps to C0 (R), then to Cb (R) and then to characteristic
functions of infinite intervals. The extension to C0 (R) is immediate from the fact that the subspace of
f ∈ C0 (R) with the property that fˆ ∈ Cc (R) is dense in C0 (R) and the measures have uniform bounds. To
extend the convergence to Cb (R) we use:
√
Lemma 8.1. The family of measures µk,z0 ,β (f ) := k −m If (eβ/( k) z0 , k) is tight, i.e. for any ǫ > 0 and
A ∈ R+ , there exists Mǫ , K so that µk,z0 ,β (|t| ≥ Mǫ ) ≤ ǫ for all k ≥ K and all z0 ∈ H −1 (E), |β| ≤ A.
√
Proof Tightness follows from (2) of Proposition 5.2 with z = eβ/ k z0 where H(z0 ) = E. If f ∈ Cb (R)
M
while
and Supp(f) ⊂ [Mǫ , ∞] then the non-zero terms in the first sum come from j : | kj − E)| ≤ √
k
√ j
j
M
f ( k( k − E)) = 0 unless | k − E)| > √k . Since there are no such terms, it follows from Proposition 5.2(2)
that µk,z0 ,β (|t| ≥ M ) ≤ ǫ.
k −m If (eβ/(
=k
−m
P
√
k)
z0 , k) := k −m
P
j
√
√
f ( k( kj − E))Πk,j (eβ/( k) z0 )
(94)
√
√
√ M f ( k( j − E))Πk,j (eβ/( k) z0 ) + Rk (z, β, M ), where |Rk (z, β, M)| ≤ ǫ.
j
k
j:| −E|≤
k
k
It further follows from the Lemma that for any f ∈ Cb (R) and ǫ > 0, there exists fM ∈ C0 (R) so that
µk,z0 ,β (f − fM ) < ǫ for k sufficiently large. Hence, the limit formula (93) is valid for f ∈ Cb (R). By the
Portmanteau theorem, it therefore extends to characteristic functions of both finite and infinite intervals.
This concludes the proof of the Lemma and hence of Theorem 0.4.
8.2. Sharp localized interface asymptotics. We now give another approach to interface asymptotics.
√
Proposition 8.2. Let f (x) be a smooth function on R, E0 ∈ H(M ) a regular energy level, z = eβ/ k · z0
where z0 ∈ H −1 (E0 ) and β ∈ R. P = [a, b] ⊂ R any closed interval (possibly infinitely). Let fP (x) =
28
STEVE ZELDITCH AND PENG ZHOU
f (x)1P (x). Then
X
Πk,E0 ,P,f (z) :=
j∈kH(M)∩Z
=
k
Z
m
√
fP ( k(j/k − E0 ))Πk,j (z)
b
f (u)e
− 12 (u/A−2Aβ)2
a
where A = A(z0 ) =
q
∂ρ2 ϕ(z0 )/2.
du
√
2πA
!
(1 + O(k −1/2 ))
Proof. The proof follows verbatim
to the interface case in Proposition 0.6. First, we reduce the general case
√
to β = 0 case. Define uj = k(j/k − E0 ). Then using directly Lemma 3.1, we have
√
√
β2 2
Πk,j (eβ/ k · z0 )
= e2jβ/ k−k(ϕ(z)−ϕ(z0 )) = e2βuj − 2 ∂ρ ϕ(z0 ) (1 + O(k −1/2 ))
Πk,j (z0 )
Thus, if we define
g(u) = f (u)e2βu−
β2
2
∂ρ2 ϕ(z0 )
Then
Πk,E0 ,P,f (z) = Πk,E0 ,P,g (z0 )(1 + O(k −1/2 ))
Hence we assume z = z0 , β = 0 from now on.
Let j0 = j0 (z, k) be the closest integer to kE0 . By localization lemma 5.1
X
√
|E0 − j/k|
Πk,j (z)fP ( k(j/k − E0 ))χ
+ O(k −∞ )
Πk,P,E,f (z) =
k −1/2+ǫ
j∈kH(M)∩Z
X
√
|E0 − j/k|
+ O(k −∞ )
e−kb(z,j/k) Πk,j (zj )fP ( k(j/k − E0 ))χ
=
k −1/2+ǫ
j∈kH(M)∩Z
Next we Taylor expand b(z, j/k) around E = E0 , ∃δ, C > 0, such that for all |E0 − E| < δ, we have
b(z, E) =
|E − E0 |2 2
4
|E − E0 |2
(3)
(3)
∂E b(z, E0 ) + Rb (z, E, E0 ) =
+ Rb (z, E, E0 )
2
2
∂ρ2 ϕ(z0 )
(95)
(3)
where we have used Lemma 2.6, and |Rb (z, E, E0 )| < C|E − E0 |3 . we have
(3)
kb(z, j/k) = A2 u2j + kRb (z, j/k),
A2 =
2
∂ρ2 ϕ(z0 )
(96)
and
(3)
kRb (z, j/k) < Ck −1/2 u3j < Ck −1/2+3ǫ .
(97)
where we used χ(uj /k ǫ ) > 0 only uj < 2k ǫ . Abbrevating the cut-off function by χj , we have
X
(3)
2
Πk,j (zj )
Πk,P,E0 ,f (z)
=
· χ(uj /k ǫ )
fP (uj )e−A2 uj ekRb (z,j/k) ·
Πk,j0 (zj0 )
Πk,j0 (zj0 )
j∈Z


X
2
= 
fP (uj )e−A2 uj  (1 + O(k −1/2 ))
j∈Z
=
Z
b
a
√
2u2
kdu(1 + O(k −1/2 ))
f (u) exp − 2
∂ρ ϕ(z)
The arguments between the steps are the same as the interface case in Proposition 0.6, hence is omitted.
Finally we use Proposition 4.2 for Πk,j0 (zj0 ) to finish the proof for the β = 0 case. The β 6= 0 case is obtained
by replacing f (u) by f (u)e2βu−
β2
2
∂ρ2 ϕ(z0 )
, as explained in the beginning.
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
29
9. Example: The Bargmann-Fock model
In this section we illustrate the results in the Bargmann-Fock model of the line bundle Cm × C → Cm
with Kähler potential ϕ = kzk2 .
Z
2
Hk2 = Hh2 k = {f ∈ O(Cm ) :
|f (z)|2 e−k||z|| dm(z) < ∞}
BF
m
where dm = (ω) /m! is Lebesgue measure, and ω =
in §2.8, the linear S 1 actions on Cm have the form,
Cm
i
¯
2π ∂ ∂ϕ
=
i
2π dz ∧ dz̄
= π −1
P
j
dxj ∧ dyj . As mentioned
eiθ · (z1 , . . . , zm ) = (eib1 θ z1 , . . . , eibm θ zm ), bj ∈ Z,
Pm
P
with Hamiltonians H = 21 ∂ρ |ρ=0 j=1 (|ebj ρ zj |2 ) = j bj |zj |2 . We only consider the diagonal T action and
Hamiltonian H(z) = ||z||2 , i.e. the isotropic harmonic oscillator in the Bargmann-Fock
representation.
The usual quantum Hamiltonian for the harmonic oscillator is ~ N̂ +
is the number or Euler operator with eigenvalues/eigenfunctions
m
2
∂
where ~ = 1/k where N̂ = Z · ∂Z
N̂ z α = |α|z α .
P
where |α| :=
i αi . Since we chose our normailzation of H to have minimum 0, we will drop the m/2
constant, and define Hk = k1 N̂ . It is an elliptic S 1 action in the sense that its moment map H is proper and
all weight spaces
αm
Hk,j = Span{z α = z1α1 · · · zm
, |α| = α1 + · · · + αm = j}
are finite dimensional.
We will fix the constant section 1 ∈ Γ(Cm , C) as the holomorphic reference frame, then an orthonormal
basis is given by
s
m
Y
k αi +1 αi
α
z
cα z =
(αi )! i
i=1
thus the full Bergman kernel
Πk (z, w) = Kk (z, w),
Kk (z, w) :=
X k |α|+m z α w α
= k m ekz·w̄ .
α!
m
α∈Z≥0
and the equivariant Bergman kernels are
Πk,j (z, w) = Kk,j (z, w),
Kk,j (z, w) :=
X k |α|+m z α w α
.
α!
α:|α|=j
The equivariant kernel is obtained from the full kernel by
Z 2π
1
Kk,j (z, w) =
e−ijθ K(eiθ z, w)dθ
2π 0
(98)
And Bergman density Bk (z) = k m , and the equivariant Bergman density is
X k m+|α| z α w α
2
Πk,j (z) = Kk,j (z, z)k1k (z)k2hk (z) = e−kkzk
α!
α:kαk=j
Lemma 9.1. As k → ∞, and E = j/k, the equivariant Bergman kernel is
kE
Z
j
e · kzk2
−ijθ keiθ kzk2
mk
m
2j
m−1/2
e
dθ = k
Kk,j (z, z) = k − e
kzk ≃ k
(2πE)−1/2
j!
E
T
and the equivariant Bergman kernel is
Bk,j (z) = Kk,j (z, z)e
−kkzk2
=k
m−1/2
The maximum of Bk,j (z) is obtained, when kzk2 = E.
−1/2
(2πE)
kzk2
E
kE
e−k(kzk
2
−E)
30
STEVE ZELDITCH AND PENG ZHOU
Proof. In fact, Kk,j (z, w) is U (m)-invariant and so Kk,j (z, w) is a function of z · w. It is also homogeneous
of degree 2j so it is a constant multiple C = Ck,j,m of (z · w)k . The constant may be determined from the
fact that
Z
Z ∞
where dim Vk (j) =
S
d−1
j+m−1
m−1
d
2
Bk,j (z)dm(z) = π −m
dim Vk (j) =
Cm
0
e−kr · Cr2j · r2m−1 ω2m−1 dr
is the number of partitions of j in m parts, ωd−1 =
⊂ R . A straightforward computation gives Ck,j,m =
km+j
j!
2π d/2
Γ(d/2)
is the volume of
. Hence Kk,j (z, z) is the j-th term of the
m kkzk2
Taylor expansion k e
. But it is useful to compute the integral using the general method, which we will
explain next.
Let E := Ek,j = kj . The first equality follows from (98). We then change θ to θ + iτ so that the complex
phase is
Ψz,τ (θ) = −iE(θ + iτ ) + ei(θ+iτ ) ||z||2 .
The critical point equation is
∂
Ψz,τ (θ) = −E + ei(θ+iτ ) ||z||2 = 0 ⇐⇒ Ee−iθ = e−τ kzk2.
i∂θ
Since the right side is positive real, the only possible solution is θ = 0 and for this we need to choose τ so
that eτ = kzk2 /E = H(z)/E. With this choice of τ , and by deforming the contour to this |w| = eτ ∈ C, the
phase becomes −iE(θ + i log(kzk2 /E)) + Eeiθ and we have a non-degenerate critical point at θ = 0 and an
asymptotic expansion,
kE
Z
e · kzk2
m
−ikE(θ+i log(kzk2 /E)) kEeiθ
m−1/2
(2πE)−1/2 .
Kk,j (z, z) = k − e
e
dθ ≃ k
E
T
where we used the stationary phase formula for dθ integral. The result agrees with the exact one after
applying Stirling formula.
The statement about the maximum of Bk,j (z) can be obtained by solving
1
d
log Bk,j (z) = −1 + E/kzk2 = 0
d|z|2 k
Indeed, the maximum of Bk,j (z) occurs when kzk2 = E.
Now we scale the equivariant Bargmann-Fock kernels around H −1 (E) and prove Theorem 0.2 in this case.
Let z0 ∈ H −1 (E), i.e. ||z0 ||2 = E and fix u ∈ R.
Z
u
u
k eiθ kz0 k2 (1+ √u )2 −kz0 k2 (1+ √u )2
k
k
Πk,j (z0 (1 + √ ), z0 (1 + √ )) = k m − e−ikEθ e
dθ
(99)
k
k
T
As k → ∞,
√
u
u
eiθ kz0 k2 (1 + √ )2 − kz0 k2 (1 + √ )2 = E(iθ − θ2 /2 + e3 (θ))(1 + 2u/ k + u2 /k))
k
k
so the phase has the form kEΨ with
√
√
√
2
Ψ = iθ 2u/ k + u2 /k − θ2 1 + 2u/ k + u2 /k + e3 (iθ) 1 + 2u/ k + u2 /k ,
where ex = 1 + x + x2 /2! + e3 (x). We localize around θ = 0 using a cutoff χ ∈ C0∞ (−1, 1) and change
variables θ → k −1/2 θ to get
Z
√
θ2
−1/2
−1
k
(2π)
χ(θ/ k) eiθ(2uE)−E 2 Ak (θ, u)dθ,
R
where A is a semi-classical symbol of order zero. Here, we absorbed the other terms,
2
e
2
√
E( iθ|u|
− θ√ u −
k
k
3
iθ
√ )+O(1/k)
3! k
into A. Since A0 (θ, u) = 1 as k → ∞ the integral tends to
Z
2
θ2
−1/2
−1
k
(2π)
eiθ(2Eu)−E 2 dθ = (2πkE)−1/2 e−2u E (1 + O(k −1/2 )),
R
INTERFACE ASYMPTOTICS OF PARTIAL BERGMAN KERNELS
Thus
31
!
2
e−2Eu
√
+ O(k −1/2 )
2πE
u
u
Πk,j (z0 (1 + √ ), z0 (1 + √ )) = k m−1/2
k
k
proving Theorem 0.2 in this case.
10. Appendix
10.1. Stationary phase on a half-space.
Lemma 10.1. Let Ψ(ξ, y), A(ξ, y) ∈ C ∞ (Rk≥0 × Rs ) such that dy Ψ(0, 0) = 0, A has compact support and
1)
2)
3)
4)
Then
Z
∂Ψ
∂ξj
6= 0 on Supp (A), for 1 ≤ j ≤ k,
dy Ψ(0, y) 6= 0 for (0, y) ∈ Supp (A) \ {0},
det Hy Ψ(0, 0) 6= 0 (where Hy denotes the Hessian with respect to y),
ReΨ ≤ ReΨ(0, 0) on Supp (A).
Rs
Z
∞
0
eN Ψ(ξ,y) A(ξ, y) dξ dy = N −k−s/2 eN Ψ(0,0) [c0 + c1 N −1 + c2 N −2 + · · · + cl N −l + O(N −l−1 )]
for l = 1, 2, 3, . . . , where
(2π)s/2 A
c0 = p
Qk
det(−Hy Ψ) j=1 ∂ψ/∂ξj
Proof. Integrating by parts,
Z Z +∞
Z
1
A
eN Ψ A dξ1 dy =
eN Ψ
N Rs
∂Ψ/∂ξ1
Rs 0
ξ1 =0
1
dy −
N
Z
Rs
.
ξ=y=0
Z
0
+∞
e
NΨ
A
∂
dξ1 dy .
∂ξ1 ∂Ψ/∂ξ1
(100)
Applying the stationary phase expansion [Hö, Th. 7.7.5] to the first term of (100) and iterating, we obtain
the desired expansion.
10.2. Euler-MacLaurin in dimension one. The Euler-MacLaurin formula states (see [KSW]),
Z b+h2
′
X
∂
∂
f (x) = L(
f (t)dt|h1 =h2 =0 .
)L(
∂h1
∂h2 ) a−h1
[a,b]
In the case of exponential functions,
′
X
eλx = L(λ)
Z
b
eλx dx =
a
[a,b]
1 λ/2
eλb − eλa
(e
+ e−λ/2 ) λ/2
.
2
e
− e−λ/2
We then sum over j ∈ k[a, b] and let y = kx, to obtain the formula in Proposition 6.2,
Lemma 10.2.
′
X
k[a,b]
eλx = L(λ)
Z
bk
ak
eλx dx = kL(λ)
Z
b
ekλx dx =
a
ekλb − ekλa
1 λ/2
(e
+ e−λ/2 ) λ/2
2
e
− e−λ/2
References
[Ber0]
[BBSj]
[BB]
[BBS]
[BSj]
[BGL]
R. J. Berman, Bergman kernels and equilibrium measures for ample line bundles, arXiv:0704.1640.
R. J. Berman, B. Berndtsson and J. Sjoestrand, A direct approach to Bergman kernel asymptotics for positive line
bundles, Ark. Mat. 46 (2008), no. 2, 197–217 ( arXiv:math/0506367.)
A. Bialynicki-Birula, Some theorems on actions of algebraic groups. Ann. of Math. (2) 98 (1973), 480-497.
A. Biaynicki-Birula and A. Sommese, Quotients by C∗ and SL(2, C) actions. Trans. Amer. Math. Soc. 279 (1983), no.
2, 773-800.
L. Boutet de Monvel and J. Sjöstrand, Sur la singularité des noyaux de Bergman et de Szegö, Asterisque 34–35
(1976), 123–164.
D. Burns, V. Guillemin and E. Lerman, Kaehler cuts, arXiv:math/0212062.
32
STEVE ZELDITCH AND PENG ZHOU
[CS]
[Ch]
[Co]
[Fe]
[F]
[GS]
[HZZ]
[Hö]
[Hö2]
[HS]
[J]
[Kan]
[KSW]
[Kash]
[Ki1]
[Ki2]
[KN]
[Ko]
[Lev]
[L]
[P]
[P2]
[PS]
[RS]
[ShZ]
[STZ]
[Sj]
[Tay]
[W]
[Y]
[Ze1]
[Ze2]
[ZZ]
J.B. Carrell and A. J. Sommese, C∗ -actions. Math. Scand. 43 (1978/79), no. 1, 49-59.
M.I. Chlodovsky, Sur la reprsentation des fonctions discontinues par les polynmes de M.S. Bernstein, Fund. Math. 13
(1929) 6272.
D. Coman and G. Marinescu, On the first order asymptotics of partial Bergman kernels arXiv:1601.00241.
R. Feng, Szasz analytic functions and noncompact Kähler toric manifolds. J. Geom. Anal. 22 (2012), no. 1, 107-131.
T. Frankel, Fixed points and torsion on Kähler manifolds. Ann. of Math. (2) 70 1959 1-8.
V. Guillemin and S. Sternberg, Geometric quantization and multiplicities of group representations. Invent. Math. 67
(1982), no. 3, 515538.
B. Hanin, S. Zelditch and P. Zhou, Wigner distributions of the spectral functions of Harmonic Oscillators (in preparation).
L. Hörmander, The Analysis of Linear Partial Differential Operators, I, Springer Verlag, N.Y., 1983.
L. Hörmander, The multinomial distribution and some Bergman kernels, in ‘Geometric Analysis of PDE and Several
Complex Variables’, Contemp. Math., 368, 249–265 Amer. Math. Soc., Providence, RI, 2005.
A. D. Hwang and M. A. Singer, A momentum construction for circle-invariant Kähler metrics. Trans. Amer. Math.
Soc. 354 (2002), no. 6, 2285-2325.
J.K. Jain, Composite Fermions, Cambridge Univ. Press (2007), and “Composite Fermions And The Fractional Quantum Hall Effect: A Tutorial”, online at www.phys.psu.edu/ jain/cftutorial.pdf.
A. Kaneko, Introduction to Kashiwaras microlocal analysis for the Bergman kernel, Lecture Notes in Math., Korea
Advanced Institute of Science and Technology, 1989.
Y. Karshon, S. Sternberg, and J. Weitsman, Euler-Maclaurin with remainder for a simple integral polytope. Duke
Math. J. 130 (2005), no. 3, 401-434.
Kashiwara, M. Analyse micro-locale du noyau de Bergman. Seminaire Goulaouic-Schwartz (1976/1977), Equations
aux derivees partielles et analyse fonctionnelle, Exp. No. 8, 10 pp. Centre Math., Ecole Polytech., Palaiseau, 1977.
F. C. Kirwan, Cohomology of quotients in symplectic and algebraic geometry. Mathematical Notes, 31. Princeton
University Press, Princeton, NJ, 1984.
F.C. Kirwan, Intersection homology and torus actions. J. Amer. Math. Soc. 1 (1988), no. 2, 385-400.
S. Kobayashi and K. Nomizu, Foundations of differential geometry. Vol I. Interscience Publishers, John Wiley &
Sons, New York-London 1963.
B. Kostant, Quantization and unitary representations. I. Prequantization.Lectures in modern analysis and applications,
III, pp. 87208. Lecture Notes in Math., Vol. 170, Springer, Berlin, 1970.
Levikson, B. On the behavior of a certain class of approximation operators for discontinuous functions. Acta Math.
Acad. Sci. Hungar. 33 (1979), no. 3-4, 299306.
Lorentz, G. G. Bernstein polynomials. Second edition. Chelsea Publishing Co., New York, 1986.
R. Paoletti, Lower-order asymptotics for Szeg and Toeplitz kernels under Hamiltonian circle actions. Recent advances
in algebraic geometry, 321-369, London Math. Soc. Lecture Note Ser., 417, Cambridge Univ. Press, Cambridge, 2015.
R. Paoletti, Local scaling asymptotics for the Gutzwiller trace formula in Berezin-Toeplitz quantization
(arXiv:1601.02128).
F. Pokorny and M. Singer, Toric partial density functions and stability of toric varieties. Math. Ann. 358 (2014), no.
3-4, 879-923.
J. Ross and M. Singer, Asymptotics of Partial Density Functions for Divisors, arXiv:1312.1145.
B Shiffman and S. Zelditch, Random polynomials with prescribed Newton polytope. J. Am. Math. Soc. 17(1), 49-108.
B. Shiffman, T. Tate, and S; Zelditch, Distribution laws for integrable eigenfunctions. Ann. Inst. Fourier (Grenoble)
54 (2004), no. 5, 1497-1546.
J. Sjöstrand, Singularités analytiques microlocales. Asterisque, 95, 1–166, Astérisque, 95, Soc. Math. France, Paris,
1982.
M.E. Taylor, Pseudodifferential operators. Princeton Mathematical Series, 34. Princeton University Press, Princeton,
N.J., 1981.
X-G. Wen, Quantum Field Theory of Many-Body Systems, Oxford Grad Texts (2004).
Q. Yang, Morse and semistable stratifications of Khler spaces by C∗ -actions. Monatsh. Math. 155 (2008), no. 1, 79-95.
S. Zelditch, Szegö kernels and a theorem of Tian, Int. Math. Res. Notices 6 (1998), 317–331.
S. Zelditch, Bernstein polynomials, Bergman kernels and toric Khler varieties. J. Symplectic Geom. 7, 1-26 (2009).
S. Zelditch and P. Zhou, Partial Bergman kernels and Hamiltonians (in preparation).
Department of Mathematics, Northwestern University, Evanston, IL 60208, USA
E-mail address: zelditch@math.northwestern.edu
Download