• Aucun résultat trouvé

Distribution of logarithmic spectra of the equilibrium energy

N/A
N/A
Protected

Academic year: 2022

Partager "Distribution of logarithmic spectra of the equilibrium energy"

Copied!
28
0
0

Texte intégral

(1)

HAL Id: hal-00816341

https://hal.archives-ouvertes.fr/hal-00816341v3

Submitted on 14 Oct 2013

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

energy

Huayi Chen, Catriona Maclean

To cite this version:

Huayi Chen, Catriona Maclean. Distribution of logarithmic spectra of the equilibrium energy.

manuscripta mathematica, Springer Verlag, 2015, 146 (3-4), pp.365-394. �10.1007/s00229-014-0712-8�.

�hal-00816341v3�

(2)

EQUILIBRIUM ENERGY

Huayi Chen & Catriona MacLean

Abstract. — Let L be a big invertible sheaf on a complex projective variety, equipped with two continuous metrics. We prove that the distribution of the eigen- values of the transition matrix between theL2 norms onH0(X, nL) with respect to the two metriques converges (in law) asngoes to infinity to a Borel probability measure onR. This result can be thought of as a generalization of the existence of the energy at the equilibrium as a limit, or an extension of Berndtsson’s results to the more general context of graded linear series and a more general class of line bundles.

Our approach also enables us to obtain ap-adic analogue of our main result.

Contents

1. Introduction. . . 1

2. Slopes of a vector space equipped with two norms.. . . 4

3. The non-Archimedean analogue.. . . 11

4. Equilibrium energy.. . . 17

5. Asymptotic distributions of logarithmic sections. . . 24

References. . . 27

1. Introduction

LetXbe a complex projective variety (or alternatively, a projective variety defined over a complete non-archimedean field), and let L be an invertible sheaf on X. We assume thatLisbig, or in other words, that the volume ofL, defined by

vol(L) := lim

n→+∞

rkCH0(X, nL) nd/d! ,

is strictly positive. (In the above formula,dis the dimension ofX.) In [2], Berman and Boucksom studied the Monge-Ampère energy functional of two continuous metrics on

(3)

L,ϕandψ. The equilibrium Monge-Ampère energy of the pair(ϕ, ψ)is an invariant defined by

Eeq(ϕ, ψ) := 1 d+ 1

Xd j=0

Z

X(C)

(P ϕ−P ψ)c1(L, P ψ)jc1(L, P ϕ)d−j.

In the above formula P φdenotes, for any continuous metricφ onL, the supremum of all (possibly singular) plurisubharmonic continuous metrics bounded above by φ.

This invariant describes the asymptotic behaviour of the ratio of the volumes of the unit balls in the linear systems H0(X, nL) (n > 1) with respect to the L2 norms induced by the metrics ϕand ψ, respectively. More precisely, given any probability measure µ on X(C), equivalent to Lesbegue measure in every local chart, we have that (cf. [2, Théorème A])

(1) lim

n→+∞

d!

nd+1lnvol(B2(nL, ϕ, µ))

vol(B2(nL, ψ, µ))=Eeq(ϕ, ψ),

where for any continuous metricφonL, the expressionB2(nL, φ, µ)denotes the unit ball in H0(X, nL)with respect to the followingL2 norm :

∀s∈H0(X, nL), ksk2L2φ,µ:=

Z

X(C)ksk2(x)µ(dx).

All volumes are calculated with respect to some arbitrary Haar measure onH0(X, nL).

The logarithm which appears in the left hand side of equation (1) is equal (up to multiplication by a constant) to the mean of the logarithms of the eigenvalues of the transition matrix betweenϕn andψn. Hereϕn andψn are theL2 norms induced by ϕandψrespectively on the spaceH0(X, nL).

In [3], Berndtson studies the distribution of eigenvalues of this transition matrix in the case whereϕandψare Kähler metrics on an ample line bundleL. He establishes the folowing fact by a detailled analysis of geodesics in the space of Kähler metrics onL: ifdn= dimVn and the numbersλj are the logarithms of the eigenvalues of the transition matrics betweenϕn andψn then the probability measure

νn=d−1n X

j

δλj

converges asntends to infinity to a measure onRwhich is defined using the Monge- Ampère geodesic linking ϕandψ inHL, the space of Hermitian metrics inL.

The log-ratio of the volumes appearing in (1) is an analogue of the degree function in Arakelov geometry. Let E be a vector space of finite rank over a number field K equipped with a family of norms k.kv, where v runs over the set MK of places of K, andk.kv is a norm on E⊗K Kv which is ultrametric when v is a finite place and Euclidean or Hermitian when v is real or complex. Moreover, we suppose that E contains a latticeE (ie. a maximal rank sub-OK module, where OK is the ring of algebraic integers in K) such that for all but a finite number of places v the norm k.kv comes from theOK-module structure onE.

(4)

The dataE= (E,(k.kv))is called aHermitian adelic bundle onKand itsArakelov degree is defined to be the weighted sum

(2) deg(E) :=d − X

v∈MK

nvlnks1∧ · · · ∧srkv,

where (s1, . . . , sr)is a basis of E over K and the weight nv is the rank of Kv as a vector space overQv. The Arakelov degree is well-defined due to the product formula

∀a∈K×, X

v∈MK

nvln|a|v= 0.

We refer the reader to [4, Appendice A] and [12] for a detailled exposition of this theory. In the context of complex geometry, we consider a finite dimensional vector spaceV equipped with two Hermitian normsϕandψ. We can write the log-ratio of the volumes of the unit balls (with respect toϕand ψrespectively) as

(3) lnvol(B(V, ϕ))

vol(B(V, ψ))=deg(V, ϕ, ψ) :=d −lnks1∧ · · · ∧srkϕ+ lnks1∧ · · · ∧srkψ, where (s1, . . . , sr) is a basis for V. This expression is independent of the choice of basis by the elementary product formula

∀a∈C×, ln|a| −ln|a|= 0.

With this notation, equation (1) can be rewritten iin the form

n→+∞lim

ddeg(H0(X, nL), L2ϕ,µ, L2ψ,µ)

nd+1/d! =Eeq(ϕ, ψ).

which is an analogue of the arithmetic Hilbert-Samuel theorem for Hermitian invert- ible bundles on a projective arithmetic variety.

Given the results in [3] and the convergences proved in [9], it is natural to wonder whether the eigenvalue distribution of the metric L2ϕ,µ with respect toL2ψ,µ is well- behaved asymptotically in more general situations. Unlike the arithmetic case, it does not seem possible to reformulate these spectra functorially, which is a key step in the proof of the convergence results in [9]. This phenomenon arises essentially because of the presence of a negative weight in formula (3), absent in the arithmetic case.

The interested reader will find counter-examples illustrating this phenomenon in [10], remark 5.10 and §4, example 2.

The main contribution of this article is the introduction of a truncation method for studying the asymptotic behaviour of eigenvalues of transition matrices between two metrics. We consider truncations of ψ by dilatations of ϕ, which enables us to prove the following result (cf. theorem 5.2infra) :

Main theorem. — Let k be the field of complex numbers or a complete non- archimedean field, and let X be a projective k-variety of dimension d>1. Let L be a big invertible OX-module with two continuous metrics ϕ and ψ. For any integer n>1, letϕn andψn be the sup norms onH0(X, nL)with respect to metricsϕandψ

(5)

respectively. Moreover, letϕn andψn be hermitian norms onH0(X, nL)such that(1) max(d(ϕn, ϕn), d(ψn, ψn)) = o(n). Let Zn be the map from {1, . . . , h0(X, nL)} to R sending i to the logarithm of the ith eigenvalue (with multiplicity) of ψn with respect toϕn, considered as a random variable on the set{1, . . . , h0(X, nL)}with the uniform distribution. Then the sequence of random variables(Zn/n)n>1 converges in law to a probability distribution onRdepending only on the pair (ϕ, ψ).

We recall that by definition the sequence of random variablesZn/nconverges in law if for any continuous bounded functionhdefined onRthe sequence(E[h(Zn/n)])n>1

converges in R.

The theorem 5.2 proved in §5 is a little bit stronger than this statement. It is valid for any sub-graded linear system of subspacesVn ⊂H0(X, nL)satisfying conditions (a)-(c) of section §4.3. Moreover, it also applies to the functionsµbidefined in §2.3, and which are associated to pairs of (possibly non-hermitian) norms. When these norms are in fact hermitian, these functions are equal to the logarithms of the eigenvalues of the transition matrix (cf. §2.2). This asymptotic distribution is a fine geometric invariant which measures the degree of non-proportionality of the metricsϕ andψ.

It should be useful in the variational study of metrics on an invertible sheaf.

The proof of the main theorem uses various techniques drawn from algebraic and arithmetic geometry. In§2-3 we introduce a Harder-Narisimhan type theory for finite dimensional vector spaces with two norms. This can be thought of as a geometric reformulation of the eigenvalues of the transition matrix between two Hermitian norms. This construction has the advantage of being valid for non-Hermitian norms, which allows us to work directly with the sup norm. Moreover, it is analogous to the Harder-Narasimhan filtration and polygon of a vector bundle on a smooth projective curve. In particular, the truncation results (propositions 2.8 and 3.8) are crucial for the proof of the main theorem. Another important ingredient is the existence of the equilibrium energy as a limit, presented in §4. For a complete complex graded system, this is a result of Berman and Boucksom [2] (where the limit is described.) Here, we use the Okounkov filtered semi-groups point of view developped in [6], analogous to that of Witt-Nystrom [17]. The combination of this method with the Harder- Naransimhan formalism is extremely flexible and enables us to prove the existence of the equilibrium energy as a limit in the very general setting of a graded linear system on both complex and non-archimedean varieties (cf. theorem 4.5 and its corollary 4.6).

Finally, in Section §5 we prove a general version of the main theorem (cf. theorem 5.2 and remark 5.3).

2. Slopes of a vector space equipped with two norms.

In this section we develop the formalism of slopes and Harder-Narasimham filtra- tions for finite dimensional complex vector spaces equipped with a pair of norms.

1. Ifη and η are two norms on a finite-dimensional complex vector space V, then d(η, η)is defined to be sup

06=s∈V

|lnkskηlnkskη|. This is a distance on the set of norms onV.

(6)

2.1. Slopes and the Harder-Narasimhan filtration.— LetCH be the class of triplets V = (V, ϕ, ψ), where V is a finite dimensional complex vector space and ϕ andψare two Hermitian norms onV. For anyV = (V, ϕ, ψ)∈ CH, we letdeg(Vd )be the real number defined by

(4) −lnks1∧ · · · ∧srkϕ+ lnks1∧ · · · ∧srkψ,

where (s1, . . . , sr) is a basis of V. If the V is non trivial then we let bµ(V) be the quotientdeg(Vd )/rk(V), which we call theslope ofV. Unless otherwise specified, for any subspaceW ofV we will letW be the vector spaceW equipped with the induced norms andV /W be the quotient spaceV /W equipped with the quotient norms. The following relationship holds for any subspaceW ofV :

(5) deg(Vd ) =deg(Wd ) +ddeg(V /W).

Proposition 2.1. — Let V = (V, ϕ, ψ) be a non-trivial element of CH. There is a unique subspaceVdes inV satisfying the following properties:

(1) for any subspace W ⊂V we have thatµ(Wb )6µ(Vb des),

(2) if W is a subspace of V such that µ(Wb ) =bµ(Vdes)thenW ⊂Vdes.

Proof. — Letλbe the norm of the identity map from(V,k.kϕ)to(V,k.kψ). We will prove that the set

Vdes={x∈V : kxkψ =λkxkϕ}

satisfies the conditions of the proposition. We start by checking that Vdes is a non- trivial subspace ofV. By definition,Vdes contains at least one non-zero vector and is stable under multiplication by a complex scalar. We need to check thatVdesis stable under addition. Let xand y be two elements of Vdes. As the norms ϕ and ψ are Hermitian, the parallelogram law states that

kx+ykϕ+kx−ykϕ= 2(kxkϕ+kykϕ), kx+ykψ+kx−ykψ= 2(kxkψ+kykψ).

Since λis the norm of the identity map from (V,k.kϕ) to (V,k.kψ)we have that kx+ykψ6λkx+ykϕandkx−ykψ6λkx−ykϕ. Asxandy are vectors inVdes we have thatkxkψ =λkxkϕ andkykψ=λkykϕ. It follows that

2λ(kxkϕ+kykϕ) = 2(kxkψ+kykψ) =kx+ykψ+kx−ykψ

6λ(kx+ykϕ+kx−ykϕ) = 2λ(kxkϕ+kykϕ), sokx+ykψ=λkx+ykϕ and it follows thatx+y∈Vdes.

In particular, as the norms k.kψ and k.kϕ are proportional with ratio λ on Vdes

we have that bµ(Vdes) = ln(λ). If W is a subspace of V then the identity map from (ΛrW,k.kψ)to (ΛrW,k.kϕ)has norm 6λr (by Hadamard’s inequality). This inequality is an equality if and only if the normsk.kψ andk.kϕare proportional with ratioλ- or in other words, the spaceW is contained inVdes.

Let V be a non-trivial element of CH. Let µbmax(V) be the slope of Vdes, which we call the maximal slope ofV. We say thatV issemi-stable ifµbmax(V) =bµ(V), or

(7)

equivalentlyVdes=V. The elementV is semi-stable if and only if the two norms on V are proportional.

For any V, non-trivial element of CH, we construct recursively a sequence of subspaces ofV of the form

(6) 0 =V0⊂V1⊂. . .⊂Vn =V

such thatVi/Vi−1 = (V /Vi−1)des for anyi∈ {1, . . . , n} using the quotient norms on V /Vi−1. Each of the subquotientsVi/Vi−1 is a semi-stable element ofCH. Moreover, ifµi is the slope ofVi/Vi−1, then we have that:

(7) µ1> µ2> . . . > µn.

These numbers are called the intermediate slopes of V. The flag (6) is called the Harder-Narasimhan filtration ofV.

Definition 2.2. — LetV be a non-zero element ofCHwith its Harder-Narasimham filtration and intermediate slopes as defined in (6) and (7). We letZV be the random variable with values in{µ1, . . . , µn}such that

P(ZVi) =rk(Vi/Vi−1) rk(V) for anyi∈ {1, . . . , n}.

Remark 2.3. — The above constructions are analogues of Harder-Narasimhan the- ory for vector bundles on a regular projective curve or hermitian adelic bundles on a number field. As in [9, §2.2.2], we can include the Harder-Narasimhan filtration and the intermediate slopes in a decreasing Rfiltration. However, contrary to the geometric and arithmetic cases, thisR-filtration is not functorial, as explained in [10, remarque 5.10]. Moreover, the Harder-Narasimhan filtration is not necessarily the only filtration whose sub-quotients are semi-stable with strictly decreasing slopes, as can be seen using the counter-example in [10, §4 exemple 2].

LetV be an element ofCHof rankr >0. We letPeV be the function on[0, r]whose graph is the upper boundary of the convex closure of the set of points of the form (rk(W),deg(Wd )). This is a concave function which is affine on each interval[i−1, i]

(i ∈ {1, . . . , r}). We call it the Harder-Narasimhan polygon of V. By definition, PeV(0) = 0andPeV(r) =ddeg(V). For anyi∈ {1, . . . , r}, we let bµi(V)be the slope of the function PeV on the interval[i−1, i], which we call the ith slope of V. We also introduce a normalised version of the Harder-Narasimhan polygon, defined by

PV(t) := 1

rk(V)PeV(trk(V)), t∈[0,1].

It follows from the relationPeV(r) =ddeg(V)that

(8) ddeg(V) =µb1(V) +· · ·+bµr(V).

(8)

Moreover, the distribution of the random variableZV (cf. definition 2.2) is given by 1

r Xr i=1

δbµi(V), whereδa is a Dirac measure supported ata.

The normalised polygonPV, the random variableZV and the slopes ofV are linked by the following simple formula.

(9) PV(1) =E[ZV] =µ(Vb ).

2.2. The link with eigenvalues.— The Harder-Narasimhan polygon and its inter- mediate slopes can be thought of as an intrinsic interpretation of the eigenvalues and eigenspaces of the transition matrix between two Hermitian norms. Indeed, given an elementV = (V, ϕ, ψ)∈ CH and an orthonormal basise= (ei)ri=1 forV with respect tok.kϕwe can construct a Hermitian matrix

AV,e=





he1, e1iψ he1, e2iψ · · · he1, eriψ

he2, e1iψ he2, e2iψ · · · he1, eriψ

... ... . .. ... her, e1iψ her, e2iψ · · · her, eriψ





If the intermediate slopes ofV are µ1> µ2> . . . > µn then the eigenvalues ofAV,e

are e1, . . . ,en. Moreover, if for any i∈ {1, . . . , n} we let V(i) be the eigenspace associated to the eigenvalueei of the endomorphism ofV given by the matrixAV,e

in the basisethen the flag

0(V(1)(V(1)+V(2)(. . .(V(1)+· · ·+V(n)=V is the Harder-Narasimhan filtration ofV.

The fact thatϕandψ are proportional on each of the subspacesV(i)implies that there is a complete flag (which will be in general finer than the Harder-Narasimhan filtration ofV)

0 =V0(V1(. . .(Vr=V

such that ddeg(Vi/Vi−1) = µbi(V). In particular, PV(i) = deg(Vd i). This can be thought of as a version of the Courant-Fischer theorem for symmetric or Hermitian matrices.

2.3. Generalisation to arbitrary norms. — The above constructions can be naturally generalised to spaces equipped with two arbitrary norms. Let C be the class of finite-dimensional complex vector spaces equipped with two (not necessarily Hermitian) norms. For any non-trivial elementV = (V, ϕ, ψ)in C we let ddeg(V)be the number

lnvol(B(V, ϕ)) vol(B(V, ψ)),

where B(V, ϕ) and B(V, ψ) are the unit balls with respect to the norms ϕ and ψ respectively, and vol is a Haar measure on V. This definition is independent of the choice of Haar measure vol. Moreover, when the norms ϕ and ψ are Hermitian,

(9)

it is equal to the number defined in (4). As in the Hermitian case, we let PeV be the concave function on[0,rk(V)] whose graph is the upper boundary of the convex closure of the set of points of the form (rk(W),ddeg(W)), where W is a member of the set of subspaces of V. For any i ∈ {1, . . . ,rk(V)}, we letµbi(V) be the slope of the functionPeV on the interval[i−1, i]. The equality (8) holds in this more general context. We letZV be a random variable whose law is an average of Dirac masses at the intermediate slopesµbi(V):

the law ofZV is 1 rk(V)

rk(V)

X

i=1

δbµi(V).

We can also introduce a normalised Harder-Narasimhan polygon :

(10) PV(t) = 1

rk(V)PeV(trk(V)), t∈[0,1].

Equation (9) holds in this more general setting : we have that (11) PV(1) =E[ZV] =µ(Vb ).

The following result compares Harder-Narasimhan polygons.

Proposition 2.4. — Let (V, ϕ, ψ) be an element ofC. If ψ is another norm on V such that(2) ψthen we have thatPe(V,ϕ,ψ)6Pe(V,ϕ,ψ)as functions on[0,rk(V)].

Proof. — For any subspaceW onV we have thatB(W, ψ)⊂ B(W, ψ)and it follows that deg(W, ϕ, ψ)d ≥ deg(W, ϕ, ψd ). The point (rk(W),deg(W, ϕ, ψd )) is therefore always below the graph ofPe(V,ϕ,ψ). It follows thatPe(V,ϕ,ψ)6Pe(V,ϕ,ψ).

Remark 2.5. — Similarly, if(V, ϕ, ψ)is an element ofC andϕ is another norm on V such thatϕ>ϕthen we have thatPe(V,ϕ,ψ)6Pe(V,ϕ,ψ)as functions on[0,rk(V)].

IfV is a non-trivial finite dimensional complex vector space andψandψ are two norms onV then we denote byd(ψ, ψ)the quantity

sup

06=x∈V

lnkxkψ−lnkxkψ

.

Corollary 2.6. — Let (V, ϕ, ψ) be a non-trivial element of C and let ϕ and ψ be two norms onV. For anyt∈[0,rk(V)]we have that

|Pe(V,ϕ,ψ)(t)−Pe(V,ϕ)(t)|6(d(ϕ, ϕ) +d(ψ, ψ))t.

Proof. — We denote byψ1 andψ2the norms onV such that k.kψ1 = e−d(ψ,ψ)k.kψ andk.kψ2 = ed(ψ,ψ)k.kψ. We have thatψ12. Moreover, for anyt∈[0,1]we have that

Pe(V,ϕ,ψ1)(t) =Pe(V,ϕ,ψ)(t)−d(ψ, ψ)t, Pe(V,ϕ,ψ1)(t) =Pe(V,ϕ,ψ)(t) +d(ψ, ψ)t.

2. In other words, for anyxV we have thatkxkψ6kxkψ.

(10)

By the above proposition we have that

|Pe(V,ϕ,ψ)(t)−Pe(V,ϕ,ψ)(t)|6d(ψ, ψ)t.

By the same argument we have that

|Pe(V,ϕ,ψ)(t)−Pe(V,ϕ)(t)|6d(ϕ, ϕ)t.

The sum of these two inequalities gives the required result.

2.4. John and Löwner norms. — Whilst the elements of C do not generally satisfy (5), John and Löwner’s ellipsoid technique enables us to show that (5) holds up to an error term. Indeed, if V is a complex vector space of rank r > 0 and φ is a norm onV then we can find two Hermitian norms φL andφJ which satisfy the following properties (cf. [16, page 84]) :

√1

rk.kφJ 6k.kφ6k.kφJ, k.kφL 6k.kφ6√ rk.kφL

Proposition 2.7. — Let V = (V, ϕ, ψ)be an element of C. If 0 =V0(V1(. . .(Vn =V is a flag of subspaces of V then we have that

(12)

ddeg(V)− Xn

i=1

ddeg(Vi/Vi−1)

6rk(V) ln(rk(V)).

Proof. — Consider the object(V, ϕL, ψJ), which is an element ofCH. The following equation holds by formula (5)):

deg(V, ϕd L, ψJ) = Xn i=1

deg(Vd i/Vi−1, ϕL, ψJ).

Moreover, sinceϕL6ϕandψJ>ψ, we have that

deg(Vd i/Vi−1, ϕ, ψ)6ddeg(Vi/Vi−1, ϕL, ψJ).

It follows that

(13) ddeg(V, ϕL, ψJ)>

Xn i=1

deg(Vd i/Vi−1, ϕ, ψ).

Moreover, it follows from the relations k.kϕ6p

rk(V)k.kϕL and k.kψ>(p

rk(V))−1k.kψJ

that

(14) deg(V, ϕ, ψ)d >deg(V, ϕd L, ψJ)−rk(V) ln(rk(V)).

It follows from the inequalities (13) and (14) that ddeg(V)>

Xn i=1

deg(Vd i/Vi−1)−rk(V) ln(rk(V)).

(11)

Applying the same argument to(V, ϕJ, ψL)we can show that ddeg(V)6

Xn i=1

deg(Vd i/Vi−1) + rk(V) ln(rk(V)).

This completes the proof of the proposition.

2.5. Truncation. — LetV be a finite-dimensional complex vector space. Ifϕis a norm onV andais a real number then we letϕ(a)be the norm onV such that

∀x∈V, kxkϕ(a)= eakxkϕ.

Ifϕandψare two norms onV then we denote byϕ∨ψthe norm onV such that

∀x∈V, kxkϕ∨ψ= max(kxkϕ,kxkψ).

Proposition 2.8. — Let V = (V, ϕ, ψ) be a non-zero element of C and let a be a real number. We have that

deg(V, ϕ, ψd ∨ϕ(a))−

rk(V)

X

i=1

max(µbi(V), a)

62 rk(V) ln(rk(V)) +rk(V) 2 ln(2).

Proof. — Letrbe the dimension ofV. We will first prove the following equation.

(15)

Xr i=1

max(µbi(V), a) = sup

t∈[0,r]

PeV(t)−at +ar.

As the functionPeV(t)−at is affine on each segment[i−1, i] (i∈ {1, . . . , r}) we get that

sup

t∈[0,r]

PeV(t)−at

= max

i∈{0,...,r}

X

16j6i

b

µi(V)−a

= Xr i=1

max µbi(V)−a,0 .

It follows that sup

t∈[0,r]

PeV(t)−at +ar=

Xr i=1

max µbi(V)−a,0 +a

= Xr i=1

max(bµi(V), a).

We start by proving the proposition in the special case whereϕandψare Hermitian.

There is then a basis e = (ei)ri=1 which is orthonormal for ϕ and orthogonal forψ.

For anyi∈ {1, . . . , r}setλi =keikψ. Without loss of generality, we may assume that λ1>. . . >λr. We therefore have that bµi(V) = ln(λi)for anyi∈ {1, . . . , r}. Letψ be the Hermitian norm onV such that

kx1e1+· · ·+xrerk2ψ = Xr i=1

x2imax(λ2i,e2a).

We then have that

kx1e1+· · ·+xrerk2ψ∨ϕ(a)= max Xr

i=1

x2iλ2i, Xr

i=1

x2ie2a

. and it follows that

k.kψ∨ϕ(a)6k.kψ 6√

2k.kψ∨ϕ(a).

(12)

It follows that

ddeg(V, ϕ, ψ∨ϕ(a))6ddeg(V, ϕ, ψ)6 r

2log(2) +deg(V, ϕ, ψd ∨ϕ(a)), and hence ddeg(V, ϕ, ψ∨ϕ(a))−

Xr i=1

max(µbi(V), a) 6 r

2log(2).

We now deal with the general case. We choose Hermitian normsϕ1 andψ1such that d(ϕ, ϕ1)6 12log(r)andd(ψ, ψ1)6 12log(r). Applying the above result to(V, ϕ1, ψ1) we get that

deg(V, ϕ1, ψ1∨ϕ1(a))− Xr i=1

max(µbi(V, ϕ1, ψ1), a) 6 r

2log(2).

Moreover, we have that

d(ψ∨ϕ(a), ψ1∨ϕ1(a))6max d(ϕ, ϕ1), d(ψ, ψ1) and hence

|deg(V, ϕ, ψ∨ϕ(a))−deg(V, ϕ1, ψ1∨ϕ1(a))|6d(ϕ, ϕ1)r+ max(d(ϕ, ϕ1), d(ψ, ψ1))r, which is bounded above byrln(r). Moreover, by 2.7 we get that

ePV(t)−Pe(V,ϕ11)(t)6 d(ϕ, ϕ1) +d(ψ, ψ1)

t6tln(r)6rln(r), and hence

sup

t∈[0,r]

PeV(t)−at

− sup

t∈[0,r]

Pe(V,ϕ11)(t)−at

6rln(r).

By (15) we get that

ddeg(V, ϕ, ψ∨ϕ(a))− Xr i=1

max(µbi(V), a)

62rln(r) +r 2log(2).

3. The non-Archimedean analogue.

In this section, we develop an analogue for slopes for finite dimensional vector spaces over a non-archimedean field kequipped with two ultrametric norms. Let k be a field equipped with a complete non-archimedean absolute value function |.|. 3.1. Ultrametric norms.— Let V be a k-vector space of dimension r equipped with a normk.k. Askis assumed to be complete the topology onV is induced by any isomorphismkr→V. (We refer the reader to [7, I.§2, n3] theorem 2 and the remark on page I.15 for a proof of this fact). In particular, any subspace ofV is closed (cf.

loc. cit. corollary 1 of theorem 2).

Let (V,k.k) be a finite-dimensional ultra-normed vector space on k. A basis e= (ei)ri=1 ofV is said to beorthogonal if the following holds :

∀(λ1, . . . , λr)∈kr, kλ1e1+· · ·+λrerk= max

i∈{1,...,r}i| · keik.

(13)

An ultra-normed vector space does not necessarily have an orthogonal basis, but the following proposition shows that an asymptotic version of Gram-Schmidt”s algorithm is still valid in this context.

Proposition 3.1. — Let (V,k.k)be a ultra-normedk-vector space of dimension r>

1. Let

0 =V0(V1(V2(. . .(Vr=V

be a full flag of subspaces of V. For any ε ∈]0,1[ there is a basis e = (ei)ri=1 compatible with the flag(3) such that

(16) ∀(λ1, . . . , λr)∈kr, kλ1e1+· · ·+λrerk>(1−ε) max

i∈{1,...,r}i| · keik. Proof. — We proceed by induction on r, the dimension of V. The case r = 1 is trivial. Assume the proposition holds for all spaces of dimension< rfor somer>2.

Applying the induction hypothesis toVr−1 and the flag0 =V0(V1(. . .(Vr−1 we get a basis(e1, . . . , er−1)compatible with the flag such that

(17) ∀(λ1, . . . , λr−1)∈kr−1, kλ1e1+· · ·+λr−1er−1k>(1−ε) max

i∈{1,...,r−1}i|·keik. Letxbe an element of V \Vr−1and lety be a point inVr−1 such that

(18) kx−yk6(1−ε)−1dist(x, Vr−1).

(The distance between xand Vr−1 is strictly positive because Vr−1 is closed in V.) We chooseer=x−y. The basise1, . . . , eris compatible with the flag0 =V0(V1( . . . (Vr =V. Let (λ1, . . . , λr) be an element of kr: we wish to find a lower bound for the norm ofz=λ1e1+· · ·+λrer. By (18) we have that

kzk>|λr| ·dist(x, Vr−1)>(1−ε)|λr| · kerk.

This provides our lower bound whenkλrerk>kλ1e1+· · ·+λr−1er−1k. Ifkλrerk<

1e1+· · ·+λr−1er−1kthen we havekzk=kλ1e1+· · ·+λr−1er−1kbecause the norm is ultrametric. By the induction hypothesis (17) we have thatkzk>(1−ε)|λi| · keik for anyi∈ {1, . . . , r−1}. This completes the proof of the proposition.

Remark 3.2. — If k is locally compact then the above equation holds for ε = 0 since the distance appearing in (18) is then attained.

Definition 3.3. — LetV be an ultra-normed k-vector space andα a real number in ]0,1]. We say that a basis e = (e1, . . . , er) of V is α-orthogonal if for any (λ1, . . . , λr)∈kr we have that

1e1+· · ·+λrerk>αmax(|λ1| · ke1k, . . . ,|λr| · kerk).

By definition,1-orthogonality is the same thing as orthogonality.

For any ultra-normed finite dimensionalk-vector spaceV we letV= Homk(V, k) be its dual space with the operator norm. This is also a finite-dimensional ultra- normedk-vector space.

3. We say that a basiseis compatible with a full flag0 =V0 (V1(V2 (. . .(Vr=V if for anyi∈ {1, . . . , r}, we have thatcard(Vie) =i.

(14)

Proposition 3.4. — Let V be a finite-dimensional ultranormed k-vector space and let (e1, . . . , er) be an α-orthogonal basis for V for some α ∈]0,1]. The dual basis (e1, . . . , er)is thenα-orthogonal.

Proof. — Consider anr-tuple(λ1, . . . , λr)∈kr. We then have that ei1e1+· · ·+λrer) =λi.

It follows that

(19) keik= sup

1,...,λr)6=(0,...,0)

i|

1e1+· · ·+λrerk 6α−1keik−1 Considerξ=λ1e1 +· · ·+λrer inV. Asξ(ei) =λi we get that

kξk>|λi|/keik>α|λi| · keik. This completes the proof of the proposition.

Let(V1, ϕ1)and (V2, ϕ2) be two ultra-normed finite-dimensionalk-vector spaces.

We can then identifyV1⊗V2withHomk(V1, V2)and equip it with the operator norm.

This ultrametric norm onV1⊗V2, is called thetensor norm and is denoted byϕ1⊗ϕ2. We can construct a tensor norm for a tensor product of multiple ultra-normed spaces recursively (cf. [12, remarque 3.8]).

Proposition 3.5. — Let V andW be two finite dimensionel ultra-normed k-vector spaces and let α be a real number in ]0,1]. If (si)ni=1 and (tj)mj=1 are α-orthogonal bases ofV andW respectively then(si⊗tj)16i6n

16j6m

is anα2-orthogonal basis ofV⊗W. Proof. — Consider(aij)16i6n

16j6m

ann×mmatrix with coefficients in k. LetT be the tensor Pn

i=1

Pm

j=1aijsi⊗tj. We seek a lower bound for the operator norm of T, considered as ak-linear map fromVtoW. Let(si)ni=1 be the dual basis of(si)ni=1. By Proposition 3.4 this is aα-orthogonal basis ofV. For anyi∈ {1, . . . , n}we have that

kT(si)k=

Xm j=1

aijtj

>α max

j∈{1,...,m}|aij| · ktjk. This proves that

kTk>ksik−1α max

j∈{1,...,m}|aij| · ktjk>α2 max

j∈{1,...,m}|aij| · ksik · ktjk,

where the second inequality follows from (19). This completes the proof of the proposition.

For anyk-vector spaceV of finite dimensionrequipped with an ultrametric norm k.kwe equip det(V) = Λr(V)with the quotient norm induced by the canonical map V⊗r → Λr(V). The ultrametric Hadamard inequality implies that for any basis (e1, . . . , er)ofV we have that

(20) ke1∧ · · · ∧erk6ke1⊗ · · · ⊗erk= Yr i=1

keik.

(15)

Equality holds when the basis(e1, . . . , er)is orthogonal. If the basis isα-orthogonal then we have that

(21) ke1∧ · · · ∧erk>αr Yr i=1

keik.

This follows from(4)proposition 3.5.

3.2. Arakelov degree and the Harder-Narasimhan polygon.— LetCkbe the class of finite dimensional k-vector spaces equipped with two ultrametric norms. If V = (V,k.kϕ,k.kψ)is an element ofCk we letdeg(Vd )be the following number

−lnks1∧ · · · ∧srkϕ+ lnks1∧ · · · ∧srkψ,

where(s1, . . . , sr)is an arbitraryk-basis ofV. This construction does not depend on the choice of(s1, . . . , sr)by the trivial product formula

∀a∈k×, −ln|a|+ ln|a|= 0.

If V is non-trivial we let bµ(V) be the quotient deg(Vd )/rk(V). It follows from Proposition 3.1, Hadamard’s inequality (20) and the inverse Hadamard inequality (21) that for any flag of subspaces ofV

0 =V0(V1(. . .(Vn =V we have that

(22) ddeg(V) =

Xn i=1

deg(Vd i/Vi−1), using the subquotient norms.

We let PeV be the function on [0, r] whose graph is the concave upper bound of the set of points in R2 of the form (rk(W),deg(Wd )), whereW runs over the set of subspaces of V. This is a concave function which is affine on each piece [i−1, i]

(i∈ {1, . . . , r}). For any i∈ {1, . . . , r} we letµbi(V)be the slope of this function on [i−1, i]. We introduce a normalised version ofPeV by setting

∀t∈[0,1], PV(t) = 1

rk(V)PeV(trk(V)).

LetZV be a random variable whose probability law is given by 1

rk(V)

rk(V)

X

i=1

δbµi(V).

The following equality also holds in the non-archimedean case (23) µ(Vb ) =PV(1) =E[ZV].

4. By induction, Proposition 3.5 implies that(eσ(1)⊗ · · · ⊗eσ(r))16σ(1),...,σ(r)6ris anαr-basis ofV⊗r. Ifξ=P

σaσeσ(1)⊗ · · · ⊗eσ(r)is an element ofV⊗nwhose image inΛrV ise1∧ · · · ∧er

thenP

σ∈Sraσ(−1)sgn(σ)= 1whereSris ther-th symmetric group. There is at least oneσSr such that|aσ|>1. It follows thatkξk>αrQr

i=1keik.

(16)

The results of §2.3 – in particular Proposition 2.4 and Corollary 2.6 – still hold for members ofCk (and their proofs are similar). We now summarise these properties:

Proposition 3.6. — Let (V, ϕ, ψ) be an element ofCk. Letϕ and ψ be two ultra- metric norms on V.

(1) IfψthenPe(V,ϕ,ψ)6Pe(V,ϕ,ψ). (2) IfϕthenPe(V,ϕ,ψ)6Pe(V,ϕ,ψ). (3) In general we have that

∀t∈[0,rk(V)], eP(V,ϕ,ψ)(t)−Pe(V,ϕ)(t)6 d(ϕ, ϕ) +d(ψ, ψ) t.

The following proposition can be seen as an ultrametric analogue of the Courant- Fischer theorem.

Proposition 3.7. — Let V = (V, ϕ, ψ)be an element ofCk of dimensionr>1. For any i∈ {1, . . . , r} we have that

(24) PeV(i) = sup

W⊂V rk(W)=i

ddeg(W).

Proof. — We prove by induction on r that for any α ∈]0,1[ there is a basis in V which is α-orthogonal for both ϕ and ψ. The case r = 1 is trivial. Suppose that this statement has been proved for all elements of Ck of dimension< r. We choose e1∈V \ {0} such that

(25) ke1kψ

ke1kϕ 6(√4

α)−1 inf

06=x∈V

kxkψ

kxkϕ.

By Proposition 3.1 there is a subspaceW ⊂V of dimensionr−1 such that (26) ∀y∈W, ke1+ykϕ>√4

αmax(ke1kϕ,kykϕ).

By (25) and (26) we have that

∀y∈W, ke1+ykψ>√4

α· ke1+ykϕ·ke1kψ

ke1kϕ

>√

α· ke1kψ.

Moreover, asψis ultrametric, ifkykψ>ke1kψ thenke1+ykψ=kykψ. It follows that (27) ke1+ykψ>√

αmax(ke1kψ,kykψ).

By the induction hypothesis, there is a basis(e2, . . . , er)ofW which is√

α-orthogonal for both ϕand ψ. Let us prove that (e1, . . . , er) isα-orthogonal for both ϕ andψ.

Let(λ1, . . . , λr)be an element of kr withλ16= 0. By (26) we have that kλ1e1+· · ·+λrerkϕ>√4

α·max(|λ1| · ke1kϕ,kλ2e2+· · ·+λrerkϕ)

>√4

α·max(|λ1| · ke1kϕ,√

α·max(|λ2| · ke2kϕ,· · · ,|λr| · kerkϕ))

>αmax(|λ1| · ke1kϕ, . . . ,|λr| · kerkϕ),

where the second inequality comes from the fact that (e2, . . . , er) is √

α-orthogonal forW for the normk.kϕ. This inequality also holds whenλ1= 0(here we use directly

(17)

the fact that (e2, . . . , er) is an√

α-orthogonal basis). Similarly, it follows from (27) that (e1, . . . , er)is anα-orthogonal basis forV with respect tok.kψ.

We now prove the proposition. We deal first with the case where there is a basis (s1, . . . , sr)which is orthogonal for both ϕ and ψ. For any i ∈ {1, . . . , r} let ai be the logarithm of the ratio ksikψ/ksikϕ. Without loss of generality we may assume that a1 >a2 >. . . >ar. For any integer m∈ {1, . . . , r} the vectors si1∧ · · · ∧sim

(16i1< . . . < im6r) form a basis forΛm(V)which is orthogonal for bothϕabd ψ. For anym-dimensional subspaceW ⊂V, writing a non-zero element ofΛmW as a sum of elements of the formsi1∧ · · · ∧sim enables us to prove that

deg(Wd )6a1+· · ·+am,

and equality is achieved whenW is generated by thes1, . . . , sm.

In the general case, the above proposition enables us to construct a sequence of objects(V, ϕn, ψn)(n∈N) inCk such that(5)

n→+∞lim d(ϕn, ϕ) +d(ψn, ψ) = 0

and for anynthere is a basis ofV which is orthogonal for both ϕn andψn. On the one hand,

Pe(V,ϕnn)(i) = sup

W⊂V rk(W)=i

deg(W, ϕd n, ψn), and on the other, we have that

eP(V,ϕnn)(t)−Pe(V,ϕ,ψ)(t)6(d(ϕn, ϕ) +d(ψn, ψ))t, and for any subspaceW ⊂V we have that

ddeg(W, ϕn, ψn)−ddeg(W, ϕ, ψ)6(d(ϕn, ϕ) +d(ψn, ψ)) rk(W).

Asn→+∞we obtain the desired result.

3.3. Truncation. — The results in §2.5 are still valid for elements ofCk. The proof is simpler because we only consider ultrametric norms. LetV be a finite dimensional k-vector space and let ϕ be an ultrametric norm on V. For any real number a let ϕ(a)be the norm onV such that

∀x∈V, kxkϕ(a)= eakxkϕ.

Ifϕandψare two ultrametric norms onV we letϕ∨ψbe the norm onV such that

∀x∈V, kxkϕ∨ψ= max(kxkϕ,kxkψ).

This is also an ultrametric norm onV.

5. As in the complex case, for any pair(η, η)of ultrametric norms onV we set

d(η, η) = sup

06=x∈V

lnkxkηlnkxkη

.

(18)

Proposition 3.8. — Let V = (V, ϕ, ψ)be a non-trivial element ofCk and letabe a real number. We have that

deg(V, ϕ, ψd ∨ϕ(a)) =

rk(V)

X

i=1

max(µbi(V), a)

Proof. — We start with the case whereV has a basis(e1, . . . , er)which is orthogonal for bothϕandψ. Without loss of generality, we have that

lnkeikψ keikϕ

=µbi(V).

The basis(e1, . . . , er)is also orthogonal for the normψ∨ϕ(a)and we have that lnkeikψ

keikϕ = max(µbi(V), a).

The result follows. In general, we can approximate (ϕ, ψ) by pairs of norms ((ϕn, ψn))n>1 such that for every n the vector space V has a basis orthogonal for bothϕn andψn (cf. the proof of Proposition 3.7). Passing to the limitn→+∞, we get our result.

4. Equilibrium energy.

In this section we fix a fieldk, which is eitherCwith the usual absolute value, or a complete field equipped with a non-archimedean absolute value. Ifk=Cthen we will denote byCk the classC defined in §2.3.

4.1. Monomial bases and the Okounkov semi-group.— In this subsection we recall the construction of the Okounkov semi-group of a graded linear series. We refer the reader to [15, 14, 5, 13] for more details.

Consider an integral projective schemeX of dimensiond>1defined over the field k. We assume that the scheme X has a regular rational pointx: the local ringOX,x is then a regular local ring of dimensiond. We fix a regular sequence(z1, . . . , zd)in its maximal ideal mx. The formal completion of OX,x with respect to the maximal idealmxis isomorphic to the algebra of formal series in the parametersz1, . . . , zd (cf.

[11, proposition 10.16]).

If we choose a monomial ordering(6)6onNd we obtain a decreasingNd-filtration (called theOkounkov filtration)FonObX,xsuch thatFα(ObX,x)is the ideal generated by monomials of the formzβ such thatβ >α. This filtration is multiplicative: we have that

Fα(ObX,x)Fβ(ObX,x)⊂ Fα+β(ObX,x).

6. This is a total order6onNdsuch that06αfor anyαNdandα6αimpliesα+β6α for anyα, α andβinNd.

(19)

The filtrationF induces by grading an Nd-graded algebragr(OX,x)which is iso- morphic(7) to k[z1, . . . , zd]. In particular, for any α∈ Nd, grα(OX,x)is a rank-one vector space onk.

IfLis an inversibleOX-module then on taking a local trivialisation in a neighbour- hood ofxwe can identifyLx withOX,x. The filtrationF then induces a decreasing Nd-filtration onH0(X, L)which is independent of the choice of trivialisation. For any s ∈H0(X, L)we denote by ord(s)the upper bound of the set of α∈ Nd such that s∈ FαH0(X, L). We have that

∀s, s ∈H0(X, L), ord(s+s)>min ord(s),ord(s) .

Moreover, for anys∈H0(X, L)and anya∈k× we have thatord(s) = ord(as).

Let L be an invertible OX-module. We let V(L) be the graded ring L

n>0H0(X, nL) (with the additive notation for the tensor product of invert- ible sheaves). By agraded linear system ofLwe mean a graded subalgebra ofV(L).

Any graded linear systemV ofLcan be identified, on choosing a local trivialisation ofLaroundx, with a graded subalgebra of the algebra of polynomialsOX,x[T]. The filtration F induces a decreasing Nd-filtration on each homogeneous piece Vn. We denote by gr(V) the Nd+1-graded k-algebra induced by this filtration. This is an Nd+1-graded subalgebra ofgr(OX,x)[T]∼=k[z1, . . . , zd, T]. In particular, the elements (n, α) ∈Nd+1 such that gr(n,α)(V) 6={0} form a sub-semigroup ofNd+1 which we denote by Γ(V). For any n∈Nwe denote byΓ(Vn)the subset ofNd of elementsα such that(n, α)∈Γ(V).

4.2. Monomial norms.— As above, we consider an integral projective schemeXof dimensiond>1overSpeck. We fix a regular rational point x∈X(k)(it is assumed that such a point exists), a system of parametersz= (z1, . . . , zd)atxand a monomial order onNd. LetLbe an invertibleOX-module and letV be a graded linear system of L. We assume that every k vector spaceVn is equipped with two norms ϕn and ψn, which are ultrametric if k is non-archimedean. We assume moreover that these norms are submultiplicative - ie. that for any(n, m)∈N2and any(sn, sm)∈Vn×Vm

we have that

(28) ksn⊗smkϕn+m6ksnkϕn· ksmkϕm, ksn⊗smkψn+m 6ksnkψn· ksmkψm. In this subsection, we study the asymptotic behaviour of deg(Vd n, ϕn, ψn). As the Harder-Narasimhan filtration is not functorial in Ck we cannot study this problem directly using the method developped in [6]. We will avoid this problem by using Okounkov filtrations. The norms ϕn and ψn induce quotient norms on each of the sub-quotients grα(Vn) (α ∈ Γ(Vn)) which by abuse of notation we will continue to denote byϕn andψn. By the results of previous sections, notably (12) and (22), we have that

ddeg(Vn, ϕn, ψn)− X

α∈Γ(Vn)

deg(grd α(Vn), ϕn, ψn)

6Ak(rk(Vn)),

7. This follows from the fact thatOX,xis dense inObX,x.

(20)

where Ak(r) = rln(r) ifk =Cand Ak(r) = 0 ifk is non-archimedean. We deduce that

(29) lim

n→+∞

bµ(Vn, ϕn, ψn)

n − 1

n#Γ(Vn) X

α∈Γ(Vn)

ddeg(grα(Vn), ϕn, ψn) = 0.

Moreover, if we equip the space

gr(Vn) = M

α∈Γ(Vn)

grα(Vn)

with normsϕˆn andψˆn such that the rank1 subspaces grα(Vn)are orthogonal with the induced (subquotient) normsϕn andψn respectively then we have that

X

α∈Γ(Vn)

ddeg(grα(Vn), ϕn, ψn) =deg(gr(Vd n),ϕˆn,ψˆn).

The problem then reduces to the study of the k-algebra of the semi-group Γ(V) (isomorphic to gr(V)) with the appropriate norms. As the semi-group Γ(V) is a multiplicative basis for the algebrak[Γ(V)]we can construct a new normηn on each space gr(Vn) which is multiplicative: for any γ ∈ Γ(V) we let sγ be the canonical image ofγ∈Γ(V)in the algebrak[Γ(V)]and we equip

gr(Vn) = M

α∈Γ(Vn)

ks(n,α)

with the norm ηn such that the vectorssn,α are orthogonal of norm1. Using these auxillary norms we can write the degreedeg(gr(Vd n),ϕˆn,ψˆn)as a difference

deg(gr(Vd n),ϕˆn, ηn)−ddeg(gr(Vn),ψˆn, ηn), or alternatively

X

α∈Γ(Vn)

deg(grd α(Vn), ϕn, ηn)− X

α∈Γ(Vn)

ddeg(grα(Vn), ψn, ηn).

It is easy to see that the real valued functions (n, α) 7→ ddeg(grα(Vn), ϕn, ηn) and (n, α)7→deg(grd α(Vn), ϕn, ηn)defined onΓ(V)are superadditive, so their asymptotic behaviour can be studied using the methods developped in [6].

4.3. Limit theorem. — In this subsection we fix an integer d > 1 and a sub- semigroupΓinNd+1. For any integern∈Nwe denote byΓnthe set{α∈Nd|(n, α)∈ Γ}. We suppose that the semi-groupΓverifies the following conditions (cf. [14, §2.1]) : (a) Γ0={0},

(b) there is a finite subsetB in {1} ×Nd such thatΓis contained in the sub-monoid ofNd+1generated byB,

(c) the groupZd+1 is generated byΓ.

Références

Documents relatifs

Abstract : We consider the two-dimensional problem of recovering globally in time the heat distribution on the surface of a layer inside of a heat conducting body from two

(1) Corollary 2.2 is true only for non-Archimedean dynamics as the ex- ample φ(z) = z 3 + z shows: its complex Julia set is non-empty and it admits a lot of repelling periodic

The second permanence property of Section 1, in particular Proposition 1.9, can be applied to obtain a solution of the Levi problem for various other locally convex spaces :

tinuous functions. The method used here generalises the results from reference [9]. For additional information we refer the reader to [1]. We remark that the notations used

The ~-product EeF, of two locally convex spaces E, F is the space of all continuous linear maps from to F endowed with the topology of uniform convergence.. on the

We prove that there exists an equivalence of categories of coherent D-modules when you consider the arithmetic D-modules introduced by Berthelot on a projective smooth formal scheme X

First of all, recall that a k-quasiaffinoid space is a pair (%', v) consisting of a locally ringed space % and an open immersion v of ^U in a A-affinoid space X.. We claim that the

It means that subspace clustering algorithms can determine a set of relevant dimensions (relevant subspace) locally for each part of the search space or subset of data vectors..