• Nenhum resultado encontrado

Invariants of Hypersurface Singularities

No documento Introduction to (páginas 121-137)

2 Hypersurface Singularities

2.1 Invariants of Hypersurface Singularities

(2) The analytic algebras

Mf :=C{x}/j(f), Tf :=C{x}/f, j(f) are called theMilnor andTjurina algebra off, respectively.

(3) The numbers

μ(f) := dimCMf, τ(f) := dimCTf

are called theMilnor andTjurina number off, respectively.

The Milnor and the Tjurina algebra and, in particular, their dimensions play an important role in the study of isolated hypersurface singularities.

Let us consider some examples.

Example 2.1.1.(1) f =x1(x21+x32) +x23+. . .+x2n, n≥2, is called an E7- singularity (see the classification in Section 2.4). Since

j(f) = 3x21+x32, x1x22, x3, . . . , xn we see thatx31, x52∈j(f), in particular,f ∈j(f).

As C{x1, . . . , xn}/j(f)=C{x1, x2}/3x21+x32, x1x22 we can draw the monomial diagram ofj(f) in the 2-plane.

3 5

x1

x2

The monomials belonging to the shaded region are contained in j(f) and it is easy to see that none of the monomials below the shaded region belongs to j(f). The only relations between these monomials are 3x21≡ −x32modj(f) and, hence, 3x21x2≡ −x42 modj(f). It follows that 1, x1, x21, x2, x1x2, x21x2, x22 is aC-basis of bothMf andTf and, thus,μ(f) =τ(f) = 7.

(2) f =x5+y5+x2y2 has j(f) =5x4+ 2xy2,5y4+ 2x2y. We can compute aC-basis ofTf as 1, x, . . . , x4, xy, y, . . . , y4and aC-basis ofMf, which has an additional monomialy5. Hence, 10 =τ(f)< μ(f) = 11.

Such computations are quite tedious by hand, but can easily be done with a computer by using a computer algebra system which allows calculations in local rings. Here is theSingularcode:

ring r=0,(x,y),ds; // a ring with a local ordering poly f=x5+y5+x2y2;

ideal j=jacob(f);

vdim(std(j)); // the Milnor number //-> 11

ideal fj=f,j;

vdim(std(fj)); // the Tjurina number //-> 10

kbase(std(fj));

//-> _[1]=y4 _[2]=y3 _[3]=y2 _[4]=xy _[5]=y //-> _[6]=x4 _[7]=x3 _[8]=x2 _[9]=x _[10]=1

Moreover, iff satisfies a certain non-degeneracy (NND) property then there is a much more handy way to compute the Milnor number. Indeed, it can be read from the Newton diagram off (see Proposition 2.16 below).

Critical and Singular Points. LetU Cn be an open subset,f :U C a holomorphic function andx∈U. We set

j(f) :=

/∂f

∂x1

, . . . , ∂f

∂xn

0· O(U)⊂ O(U)

and define

Mf,x:=OCn,x/j(f)OCn,x, Tf,x:=OCn,x/f, j(f)OCn,x

to be the Milnor and Tjurina algebraof f at x. Furthermore, we introduce μ(f, x) := dimCMf,x, τ(f, x) := dimCTf,x,

and call these numbers theMilnor andTjurina number off atx.

It is clear thatμ(f, x)= 0 iff ∂x∂f

i(x) = 0 for alli, and that τ(f, x)= 0 iff additionallyf(x) = 0. Hence, we see thatμcounts the singular points of the function f, while τ counts the singular points of thezero set off, each with multiplicityμ(f, x), respectivelyτ(f, x). The following definition takes care of this difference:

Definition 2.2.LetU Cnbe open,f :U Ca holomorphic function, and X =V(f) =f1(0) the hypersurface defined by f inU. We call

Crit(f) := Sing(f) :=

x∈U

∂f

∂x1

(x) =. . .= ∂f

∂xn

(x) = 0

!

the set ofcritical, orsingular, points of f and Sing(X) :=

x∈U

f(x) = ∂f

∂x1

(x) =. . .= ∂f

∂xn

(x) = 0

!

the set ofsingular points of X.

A point x∈U is called an isolated critical point of f, if there exists a neighbourhoodV ofxsuch that Crit(f)∩V \ {x}=. It is called anisolated singular point of X if x∈X and Sing(X)∩V \ {x}=. Then we say also that the germ (X, x)(Cn, x) is anisolated hypersurface singularity.

Note that the definition of Sing(X), resp. Sing(f), is a special case of Definition 1.40, resp. 1.112.

Lemma 2.3.Letf :U Cbe holomorphic andx∈U, then the following are equivalent.

(a)xis an isolated critical point off, (b)μ(f, x)<∞,

(c)xis an isolated singularity of f1(f(x)) =V(f−f(x)), (d)τ(f−f(x), x)<∞.

Proof. (a), respectively (c) says thatxis an isolated point of the fibre over 0 (if it is contained in the fibre) of the morphisms

∂f

∂x1, . . . , ∂f

∂xn

:U −→Cn,

f−f(x), ∂f

∂x1, . . . , ∂f

∂xn

:U −→Cn+1, respectively. Hence, the equivalence of (a) and (b), respectively of (c) and (d), is a consequence of Proposition 1.70 or the Hilbert-R¨uckert Nullstellensatz 1.72.

Sinceμ(f, x)≤τ(f−f(x), x), the implication (b)(d) is evident. Finally, (c)(a) follows from the following lemma, which holds also for non-isolated

singularities.

Lemma 2.4.LetU Cn be open,f :U Ca holomorphic function,x∈U andf(x) = 0. Then there is a neighbourhoodV ofxin Usuch that

Crit(f)∩V ⊂f1(0).

In other words, the nearby fibresf1(t)∩V,t sufficiently small, are smooth.

Proof. Consider C= Crit(f) with its reduced structure. As a reduced com- plex space, the regular points of C, Reg(C), are open and dense in C by Corollary 1.111. Since ∂x∂f

i vanishes onCfori= 1, . . . , n,f is locally constant on the complex manifold Reg(C). A sufficiently small neighbourhood V ofx intersects only the connected components of Reg(C) having xin its closure.

Ifx /∈C the result is trivial. If x∈C then f|VC = 0, sincef is continuous

andf(x) = 0.

Hence, it cannot happen that the critical set of f (the dashed line) meets f1(0) as in the following picture.

Crit(f)

Semicontinuity of Milnor and Tjurina number.In the sequel we study the behaviour of μand τ under deformations. Loosely speaking, a deforma- tion of a power series f C{x}, usually called an unfolding, is given by a power series F C{x,t} such that, settingFt(x) =F(x,t),F0=f, while a deformation of the hypersurface germ f1(0) is given by any power series F C{x,t} satisfyingF01(0) =f1(0). So far, unfoldings and deformations are both given by a power seriesF, the difference appears later when we con- sider isomorphism classes of deformations. For the moment we only consider the power series F.

Definition 2.5.A power series F C{x,t}=C{x1, . . . , xn, t1, . . . , tk} is called an unfolding of f C{x1, . . . , xn} if F(x,0) =f(x). We use the no- tation

Ft(x) = F(x,t), t∈T ,

for the family of power series FtC{x} or, after choosing a representative F:U×T C, for the familyFt:U Cof holomorphic functions parame- trized by t∈T, where U Cn and T Ck are open neighbourhoods of the origin.

Theorem 2.6 (Semicontinuity of μand τ).

Let F C{x,t}be an unfolding off C{x},f(0) = 0, and assume that0is an isolated critical point off. Then there are neighbourhoodsU =U(0)Cn, V =V(0)C,T =T(0)Ck, such thatF converges on U×T and the fol- lowing holds for each t∈T:

(1)0∈Uis the only critical point off =F0:U →V, andFthas only isolated critical points inU.

(2) For eachy∈V,

μ(f,0)

xSing(Ft1(y))

μ(Ft,x) and

τ(f,0)

xSing(Ft1(y))

τ(Ft−y,x).

(3) Furthermore,

μ(f,0) =

xCrit(Ft)

μ(Ft,x).

Proof. (1) ChooseU such that0is the only critical point ofF0and consider the map

Φ:U×T Cn×T , (x,t) ∂Ft

∂x1(x), . . . ,∂Ft

∂xn(x),t

.

Then Φ1(0,0) = Crit(F0)× {0}={(0,0)} by the choice of U. Hence, by the local finiteness Theorem 1.66, Φ is a finite morphism if we choose U, T

V U f1(0)

f

T

t=0 t=t y1

y2

V

Ft1(y1) Ft1 (y2)

Fig. 2.5.Deformation of an isolated hypersurface singularity

to be sufficiently small. This implies that Φ has finite fibres, in particular, Crit(Ft)× {t}=Φ1(0,t) is finite.

(2) The first inequality follows from (3). For the second consider the map Ψ:U×T −→V ×Cn×T , (x,t)

Ft(x),∂Ft

∂x1(x), . . . ,∂Ft

∂xn(x),t

. Then Ψ1(0,0,0) = Sing(f01(0))× {0}={(0,0)} and, again by the local finiteness theorem, Sing

Ft1(y)

× {t}=Ψ1(y,0,t) is finite forU, V, T suf- ficiently small andy∈V, t∈T. Moreover, the direct image sheaf ΨOU×T

is coherent onV ×Cn×T. The semicontinuity of fibre functions (Theorem 1.81) implies that the function

ν(y,t) :=ν(ΨOU×T,(y,0,t))

=

(x,t)Ψ1(y,0,t)

dimCOU×T ,(x,t)/m(y,0,t)OU×T ,(x,t)

is upper semicontinuous. Since

OU×T ,(x,t)/m(y,0,t)OU×T ,(x,t)=OU,x

1 2

Ft−y,∂Ft

∂x1, . . . ,∂Ft

∂xn 3

we have ν(0,0) =τ(f,0) andν(y,t) =

xSing(Ft1(y))τ(Ft,x), and the re- sult follows.

(3) We consider again the morphismΦand have to show that the function

ν(t) :=ν(ΦOU×T,(0,t)) =

xCrit(Ft)

dimCOU,x

1 2∂Ft

∂x1

, . . . ,∂Ft

∂xn

3

is locally constant on T. Thus, by Theorems 1.81 and 1.82 we have to show that Φis flat at (0,0).

Since OU×T ,(0,0)=C{x1, . . . , xn, t1, . . . , tk} is a regular local ring, and since the n+k component functions ∂F∂xt

1, . . . ,∂F∂xt

n, t1, . . . , tk define a zero- dimensional, hence (n+k)-codimensional germ, the flatness follows from the

following proposition.

Proposition 2.7. (1) Let f = (f1, . . . , fk) : (X, x)(Ck,0) be a holomor- phic map germ andM a finitely generatedOX,x-module. ThenM isf-flat iff the sequencef1, . . . , fk isM-regular11.

In particular,f is flat ifff1, . . . , fk is a regular sequence.

(2) If (X, x) is the germ of an n-dimensional complex manifold, then f1, . . . , fk isOX,x-regular iffdim(f1(0), x) =n−k.

The proof is given in Appendix B.8.

Remark 2.7.1.Let (T,0)(Ck,0) be an arbitrary reduced analytic subgerm, and letF ∈ OCn×T ,0map tof C{x}(as in Theorem 2.6) under the canoni- cal surjectionOCn×T ,0→ OCn,0=C{x}. Then we can liftF toF∈ OCn×Ck,0

and apply Theorem 2.6 to obtain the semicontinuity of μ, resp.τ, for Fand all t in a neighbourhood of 0Ck. Since F : (Cn×T,0)(C,0) is the re- striction of F: (Cn×Ck,0)(C,0), statements (1), (2) and (3) hold for F and a sufficiently small representativeT of the germ (T,0).

Alternatively, we may apply the proof of Theorem 2.6 directly to an arbi- trary reduced germ (T,0). The flatness of the maps φandψfollows from the flatness of the mapsφandψ(associated toF) and the base change property for flatness (Propoisition 1.87 on page 89).

Example 2.7.2.(1) Consider the unfolding Ft(x, y) =x2−y2(t+y) of the cusp singularity f(x, y) =x2−y3. We compute Crit(Ft) ={(0,0),(0,−23t)} and Sing(Ft1(0)) ={(0,0)}. Moreover, μ(f) =τ(f) = 2, while for t= 0 we have μ(Ft,(0,0)) =τ(Ft,(0,0)) = 1 andμ(Ft,(0,−23t)) = 1.

(2) For the unfoldingFt(x, y) =x5+y5+tx2y2we compute the critical locus to be Crit(Ft) =V(5x4+ 2txy2,5y4+ 2tx2y). The only critical point ofF0is the origin 0= (0,0), and we haveμ(F0,0) =τ(F0,0) = 16. UsingSingular we compute that, for t= 0, Ft has a critical point at 0 with μ(Ft,0) = 11, τ(Ft,0) = 10, and five further critical points withμ=τ = 1 each. This shows that μ(F0,0) =

xCrit(Ft)μ(Ft,x) for eacht as stated in Theorem 2.6.

11 Recall that f1, . . . , fk is an M-regular sequence or M-regular iff f1 is a non- zerodivisor of M and fi is a non-zerodivisor of M/(f1M+. . .+fi1M) for i= 2, . . . , k.

V(f)

−→

2 3t

V(Ft)

Fig. 2.6.Deformation of a cusp singularity

But τ(F0,0) = 16>15 =

xCrit(Ft)τ(Ft−Ft(x),x), that is, even the

“total” Tjurina number is not constant.

(3) The local, respectively total, Milnor number can be computed inSingu- larby the same formulas but with a different choice of monomial ordering.

First, we work in the ringQ(t)[x, y]x,y, by choosing thelocal monomial or- deringds:

ring r=(0,t),(x,y),ds;

poly f=x5+y5;

poly F=f+tx2y2; // an unfolding of f LIB "sing.lib"; // load library

milnor(f); // (local) Milnor number of the germ (f,0) //-> 16

tjurina(f); // (local) Tjurina number of (f,0) //-> 16

milnor(F); // (local) Milnor number of F for generic t //-> 11

tjurina(F); // (local) Tjurina number of F for generic t //-> 10

To obtain the total (affine) Milnor, respectively Tjurina, number, we repeat the same commands in a ring with theglobal monomial orderingdp(imple- mentingQ(t)[x, y]):

ring R=(0,t),(x,y),dp;

poly F=x5+y5+tx2y2;

milnor(F); // global Milnor number of F_t for generic t //-> 16

tjurina(F); // global Milnor number of F_t for generic t //-> 10

Since the local and the global Tjurina number forFtcoincide, the hypersurface Ft1(0) has, for generict, the origin as its only singularity.

If the first inequality in Theorem 2.6 (2) happens to be an equality (for somey suffciently close to 0) then the fibreFt1(y) contains only one singular point:

Theorem 2.8.Let F C{x,t} be an unfolding of f x2C{x}. More- over, let T Ck and U Cn be open neighbourhoods of the origin, and let

Ft:U C, x →Ft(x) =F(x,t). If 0 is the only singularity of the special fibreF01(0) =f1(0) and, for allt∈T,

xSing(Ft1(0))

μ(Ft,x) = μ(f,0)

then all fibres Ft1(0),t∈T, have a unique singular point (with Milnor num- ber μ(f,0)).

This was proven independently by Lazzeri [Laz] and Gabri`elov [Gab1].

Right and Contact Equivalence.Now let us consider the behaviour ofμ andτ under coordinate transformation and multiplication with units.

Definition 2.9.Letf, g∈C{x1, . . . , xn}.

(1)f is calledright equivalent tog,f∼rg, if there exists an automorphismϕ ofC{x}such thatϕ(f) =g.

(2)f is calledcontact equivalent tog,f∼c g, if there exists an automorphism ϕofC{x}and a unitu∈C{x}such that f =u·ϕ(g)

If f, g∈ OCn,x then we sometimes also write (f, x)r(g, x), respectively (f, x)c(g, x).

Remark 2.9.1. (1) Of course, f∼r g implies f∼c g. The converse, however, is not true (see Exercise 2.1.3, below).

(2) Any ϕ∈AutC{x} determines a biholomorphic local coordinate change Φ= (Φ1, . . . , Φn) : (Cn,0)(Cn,0) by Φi=ϕ(xi), and, vice versa, any iso- morphism of germs Φ determines ϕ∈AutC{x} by the same formula. We have ϕ(g) =g◦Φand, hence,

f∼r g ⇐⇒ f =g◦Φ

for some biholomorphic map germΦ: (Cn,0)(Cn,0), that is, the diagram (Cn,0) Φ

= f

(Cn,0)

g

(C,0)

commutes. The notion of right equivalence results from the fact that, on the level of germs, the group of local coordinate changes acts from the right.

(3) Sincef andggenerate the same ideal inC{x}iff there is a unitu∈C{x} such thatf =u·g, we see thatf∼c gifff=ϕ(g)for someϕ∈AutC{x}. Moreover, since any isomorphism of analytic algebras lifts to the power series ring by Lemma 1.14, we get

f∼c g ⇐⇒ C{x}/f ∼=C{x}/gas analytic C-algebras.

Equivalently,f∼cgiff the complex space germs (f1(0),0) and (g1(0),0) are isomorphic.

Hence, f∼r g iff f and g define, up to a change of coordinates in (Cn,0), the same map germs (Cn,0)(C,0), while f∼c g iff f and g have, up to coordinate change, the same zero-fibre.

Lemma 2.10.Let f, g∈C{x1, . . . , xn}. Then

(1)f∼rg implies thatMf =Mg andTf =Tg as analytic algebras. In partic- ular, μ(f) =μ(g)andτ(f) =τ(g).

(2)f∼cg implies that Tf =Tg and henceτ(f) =τ(g).

Proof. (1) Ifg=ϕ(f) =f◦Φ, then (f◦Φ)

∂x1

(x), . . . ,∂(f◦Φ)

∂xn

(x)

= ∂f

∂x1

(Φ(x)), . . . , ∂f

∂xn

(Φ(x))

·DΦ(x), whereis the Jacobian matrix ofΦ, which is invertible in a neighbourhood of x. It follows that j(ϕ(f)) =ϕ(j(f)) and ϕ(f), j(ϕ(f))=ϕ(f, j(f)), which proves the claim.

(2) By the product rule we haveu·f, j(u·f)=f, j(f)for a unitu, which

together with (1) impliesTf =Tg.

In characteristic 0 it is even true thatf∼c g impliesμ(f) =μ(g), but this is more difficult. For an analytic proof we refer to [Gre] where the following formulas are shown (even for complete intersections):

μ(f) =

dimCOX,01, ifn= 1, dimCΩnX,01

nX,02, ifn≥2,

with (X,0) = (f1(0),0). Even more, μ(f) is a topological invariant of (f1(0),0) (cf. [Mil1] in general, respectively Section 3.4 for curves).

Example 2.10.1. (1) Consider the unfolding Ft(x, y) =x2+y2(t+y) with Crit(Ft) ={(0,0),(0,−23t)}. The coordinate changeϕt:x →x,y →y√

t+y, (t= 0), satisfiesϕt(x2+y2) =x2+y2(t+y) =Ft(x, y).

Hence, (Ft,0)r(x2+y2,0) for t= 0. Thus, we have τ(x2+y3,0) = 2, but for t= 0 we have τ(Ft,0) = 1, τ(Ft,(0,−23t)) = 1. Hence (Ft,0) and (Ft,(0,−23t)) are not contact equivalent to (f,0).

(2) Consider the unfoldingFt(x, y) =x2+y2+txy=x(x+ty) +y2. The co- ordinate change ϕ:x →x−12ty, y →y satisfies ϕ(Ft) =x2+y2(114t2), which is right equivalent tox2+y2 fort=±2. In particular, (Ft,0)r(F0,0) for all sufficiently smallt= 0.

(3) The Milnor number is not an invariant of the contact class in positive characteristic:f =xp+yp+1 hasμ(f) =, butμ((1 +x)f)<∞inK[[x, y]]

whereK is a field of characteristicp.

Quasihomogeneous Singularities. The class of those isolated hypersur- face singularities, for which the Milnor and Tjurina number coincide, at- tains a particular importance. Of course, an isolated hypersurface singularity (X, x)(Cn, x) belongs to this class ifff ∈j(f) for some (hence, by the chain rule, all) local equation(s)f C{x}=C{x1, . . . , xn}. In the following, we give a coordinate dependent description of this class:

Definition 2.11.A polynomial f =

αNnaαxαC[x] is called weighted homogeneous or quasihomogeneous) of type (w;d) = (w1, . . . , wn;d) if wi, d are positive integers satisfying

w-deg(xα) :=w,α=w1α1+. . .+wnαn=d

for each αNn withaα= 0. The numbers wi are called the weights and d theweighted degree or thew-degree off.

Note that this property is not invariant under coordinate changes (if the wi

are not all the same then it is not even invariant under linear coordinate changes).

In the above Example 2.1.1 (1),f is quasihomogeneous of type (6,4,9; 18), while in Example 2.1.1 (2),fis not quasihomogeneous, not even after a change of coordinates.

Remark 2.11.1.A quasihomogeneous polynomial f of type (w;d) obviously satisfies theEuler relation12

d·f = n

i=1

wixi

∂f

∂xi in C[x], and the relation

f(tw1x1, . . . , twnxn) =td·f(x1, . . . , xn) inC[x, t].

The Euler relation implies thatf is contained inj(f), hence,μ(f) =τ(f). The other relation implies that the hypersurfaceV(f)Cnis invariant under the C-action C×CnCn, (λ,x) →λ◦x:= (λw1x1, . . . , λwnxn). In particu- lar, the complex hypersurface V(f)Cn is contractible.

Moreover, Sing(f) and Crit(f) are also invariant underCand, hence, the union ofC-orbits. It follows that ifV(f) has an isolated singularity at0then 0is the only singular point of V(f). Furthermore, x →λ◦xmapsV(f−t) isomorphically onto V(f−λdt). Sincef ∈j(f), Sing(f) and Sing(V(f)) co- incide in this situation.

Definition 2.12.An isolated hypersurface singularity (X,0)(Cn,0) is called quasihomogeneous if there exists a quasihomogeneous polynomial f C[x] =C[x1, . . . , xn] such thatOX,0=C{x}/f.

12 The Euler relation generalizes theEuler formula for homogeneous polynomials f∈C[x0, . . . , xn]: x0 ∂f

∂x0 +. . .+xn ∂f

∂xn= deg(f)·f.

Lemma 2.13.Let f C[x] be quasihomogeneous and g∈C{x} arbitrary.

Thenf∼c g ifff∼r g.

Proof. Letf be weighted homogeneous of type (w1, . . . , wn;d). Iff∼c g then there exists a unitu∈C{x}and an automorphism ϕ∈AutC{x} such that u·f =ϕ(g). Choose ad-th rootu1/dC{x}. The automorphism

ψ:C{x} −→C{x}, xi →uwi/d·xi

yieldsψ f(x)

=f

uw1/dx1, . . . , uwn/dxn

=u·f(x) by Remark 2.11.1, im-

plying the result.

It is clear that for quasihomogeneous isolated hypersurface singularities the Milnor and Tjurina number coincide (since f ∈j(f)). It is a remarkable theorem of K. Saito [Sai] that for an isolated singularity the converse does also hold. Let (X, x)(Cn, x) be an isolated hypersurface singularity and let f C{x1, . . . , xn} be any local equation for (X, x), then

(X, x) quasihomogeneous ⇐⇒ μ(f) =τ(f).

Sinceμ(f) andτ(f) are computable, the latter equivalence gives an effective characterization of isolated quasihomogeneous hypersurface singularities.

Newton Non-Degenerate and Semiquasihomogeneous Singularities.

As mentioned before, for certain classes of singularities there is a much more handy way to compute the Milnor number. It can be read from the Newton diagram of an appropriate defining power series:

Definition 2.14.Let f =

αNnaαxαC{x}=C{x1, . . . , xn}, a0= 0.

Then the convex hull inRn of the support off, Δ(f) := conv αNnaα= 0

,

is called theNewton polytope off. We introduceK(f) := conv({0} ∪Δ(f)), and denote by K0(f) the closure of the set (K(f)(f))∪ {0}. Define the Newton diagram13 Γ(f,0)off at the originas the union of those faces of the polyhedral complexK0(f)∩Δ(f) through which one can draw a supporting hyperplane to Δ(f) with a normal vector having only positive coordinates.

Moreover, we introduce for a faceσ⊂Γ(f,0) thetruncation fσ :=

ασ

cαxα=

iσNn

cαxα,

that is, the sum of the monomials inf corresponding to the integral points in σ.

13 An equivalent definition is as follows: Define thelocal Newton polytope N(f,0) as

the convex hull of ,

αsupp(f)

α+ (R0)n. ThenΓ(f,0) is the union of the compact faces ofN(f,0).

Example 2.14.1.Letf =(y5+xy3+x2y2−x2y4+x3y−10x4y+x6).

Δ(f) K0(f) Γ(f,0)

In particular, the Newton diagram at0has three one-dimensional faces, with slopes2,−1,−13.

Definition 2.15.A power series f =

αNnaαxαmC{x} is called convenient if its Newton diagram Γ(f,0) meets all the coordinate axes. A convenient power seriesf is calledNewton non-degenerate (NND) at0if, for all faces σ⊂Γ(f,0), the hypersurface{fσ= 0} has no singular point in the torus (C)2.

In the above example, we have 3 truncations on one-dimensional faces σ of Γ(f,0), fσ=y5+xy3, xy3+x2y2+x3y and x3y+x6, respectively. None of the corresponding hypersurfaces {fσ= 0} is singular in (C)2, and the trun- cations at the 0-dimensional faces are monomials, hence define hypersurfaces having no singular point in the torus (C)2. However f is not Newton non- degenerate, since it is not convenient. In turn, x1f is NND. On the other hand,x1f +xy2is Newton degenerate, since its truncation at the face with slope1,y3+ 2xy2+x2y=y(x+y)2, is singular along the line{x+y= 0}. Proposition 2.16.Let f C{x1, . . . xn} be Newton non-degenerate. Then the Milnor number of f satisfies

μ(f) =n! Voln

K0(f) +

n i=1

(1)ni(n−i)!·Volni

K0(f)∩Hni

,

whereHi denotes the union of alli-dimensional coordinate planes, and where Voli denotes the i-dimensional Euclidean volume.

For a proof, we refer to [Kou, Thm. I(ii)]. The right-hand side of the formula is called the Minkowski mixed volume of the polytopeK0(f), or theNewton number off.

In the above Example 2.14.1, we compute μ(x1f) = 2· 19

2 11 + 1 = 9.

Note that, in general, the Newton number off gives a lower bound forμ(f) (cf. [Kou]).

Another important class of singularities is given by the class of semiquasi- homogeneous singularities, which is characterized by means of the Newton diagram, too:

Definition 2.17.A power seriesf C{x1, . . . , xn}is calledsemiquasihomo- geneous (SQH) at 0 (or, 0 is called a semiquasihomogeneous point of f) if there is a faceσ⊂Γ(f,0) of dimensionn−1 (called themain, or principal, face) such that the truncation fσ has no critical points in Cn\ {0}. fσ is called themain part, orprincipal part, off.

Note that fσ is a quasihomogeneous polynomial, hence, it is contained in the ideal generated by its partials. It follows thatfσ has no critical point in Cn\ {0} iff the hypersurface{fσ= 0} ⊂Cn has an isolated singularity at0.

In other words, due to Lemma 2.3,f is SQH iff we can write f =f0+g , μ(f0)<∞

withf0=fσa quasihomogeneous polynomial oftype(w;d) and all monomials ofgbeing ofw-degree at leastd+ 1.

We should like to point out that we do not require that the Newton di- agramΓ(f,0) meets all coordinate axes (as for NND singularities). For in- stance, xy+y3+x2y2C{x, y} is SQH with main part xy+y3 (which is w= (2,1)-weighted homogeneous of weighted degree 3); but it is not Newton non-degenerate, since the Newton diagram does not meet the x-axis. How- ever, the results of the next section show that each SQH power series is right equivalent to a convenient one.

Note that each convenient SQH power series f C{x, y} is NND, while for higher dimensions this is not true. For instancef = (x+y)2+xz+z2 is SQH (withf =f0) and convenient, but Newton degenerate (the truncation (x+y)2 at one of the one-dimensional faces has singular points in (C)3).

Corollary 2.18.Letf C{x}be SQH with principal partf0. Thenf has an isolated singularity at0andμ(f) =μ(f0).

Proof. Letf0C[x] be quasihomogeneous of type (w;d) and write f =f0+

i1

fi

withfi quasihomogeneous of type (w;d+i). Clearly,f is singular at0ifff0 is singular at0. Consider fort∈Cthe unfolding

Ft(x) :=f0(x) +

i1

tifi(x),

which satisfies F0=f0 andF1=f. Theorem 2.6 (1) implies that, for t0= 0 sufficiently small,Ft0has an isolated critical point at0. Since, for everyt∈C,

Ft(x1, . . . , xn) = 1

td ·f(tw1x1, . . . , twnxn), theC-actionx →(tw1x1, . . . , twnxn) maps

Crit(Ft)

x

∀i: |xi|< ε

|t|wi

! =

−→Crit(f)4

x∀i: |xi|< ε 5

. Hence, we can find some ε >0, independent of t, such that, for all |t| ≤1, Crit(Ft:Bε(0)C) ={0}. Finally, the statement follows from Theorem

2.6 (3).

Again, the SQH and NND property are both not preserved under analytic coordinate changes, for instance, x2−y3C{x, y} is SQH and NND, but (x+y)2−y3C{x, y}is neither SQH nor NND. Anyhow, we can make the following definition:

Definition 2.19.An isolated hypersurface singularity (X, x)(Cn, x), is called Newton non-degenerate (respectively semiquasihomogeneous), if there exists a NND (respectively SQH) power seriesf C{x}=C{x1, . . . , xn}such that OX,x=C{x}/f.

Exercises

Exercise 2.1.1.Let p1, . . . , pnZ1, and let f =xp11+. . .+xpnnC[x].

Show thatμ(f,0) = (p11)·. . .·(pn1).

More generally:

Exercise 2.1.2.Let p1, . . . , pnZ1, w= (w1, . . . , wn) with wi:=+

j=ipj

and d:=+n

i=1pi. Moreover, let f C[x1, . . . , xn] be a quasi-homogeneous polynomial of type (w;d) which has an isolated critical point at the origin.

Show thatμ(f,0) = (p11)·. . .·(pn1).

Exercise 2.1.3.Consider the unfolding

ft(x, y, z) =xp+yq+zr+txyz, 1 p+1

q +1 r <1.

Show that for allt, t = 0,ft

c ft butft

r ft.

Exercise 2.1.4.Show that μ−τ is lower semicontinuous in the following sense: with the notations and under the assumptions of Theorem 2.6, we have μ(f,0)−τ(f,0)≤μ(Ft,0)−τ(Ft,0).

Exercise 2.1.5.Let f =fd+fd+1, where fd, fd+1C[x] =C[x1, . . . , xn] are homogeneous polynomials of degree d,d+ 1, respectively. Assume that the system

∂fd

∂x1 =. . .= ∂fd

∂xn =fd+1= 0

has the origin as only solution. Show thatμ(f) =μ(f,0)<∞. Furthermore, if n= 2, show that μ(f) =d(d−1)−k, where k is the number of distinct linear divisors offd.

Exercise 2.1.6. (1) Letf C{x, y}be of orderd≥2 with a non-degenerate principal form14of degreed. Prove the following statements:

Iff is a polynomial of degree at mostd+ 1, then μ(f)−τ(f)

(k−1)2, ifd= 2k+ 1, (k−1)(k−2), ifd= 2k . Furthermore, show that this bound is sharp for d≤6.

In general, we have

μ(f)−τ(f)(d−4)(d−3)

2 .

(2) Improve the latter bound up to μ(f)−τ(f)

d5

k=1

min

(k+ 1)(k+ 2)

2 , d−5−k

! .

(3) Generalize the above bounds to semiquasihomogeneous plane curve sin- gularities.

(4)Generalize the above bounds to higher dimensions, for instance, prove that iff C{x}=C{x1, . . . , xn}is a polynomial of degree at mostd+ 1 with zero (d−1)-jet and a non-degenerated-form, then

μ(f)−τ(f)(k0+ 1)kn0, k0=

6(n−1)d−1 n

7

2, (n−1)d−2n−1 =nk0+

(the upper bound is the maximum number of integral points in a paral- lelepiped with sides parallel to the coordinate axes and inscribed into the simplex

(i1, . . . , in)Rni1+. . .+in≥d+ 1, max{i1, ..., in} ≤d−2 . Exercise 2.1.7.Under the hypotheses of Corollary 2.18, is it true thatg∼r f, respectivelyg∼c f?

14 Letf∈C{x, y}. Then we may writef=fd+fd+1+. . ., where d= ord(f) and fk is a homogeneous polynomial of degreek,k≥d. The polynomialfd is called theprincipal form (orprincipal d-form) of f. It is callednon-degenerate if the hypersurface{fd= 0}has no critical points inCn\ {0}.

No documento Introduction to (páginas 121-137)