This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Positive Ulrich Sheaves

Christoph Hanselka Universität Konstanz, Germany [email protected]  and  Mario Kummer Technische Universität Berlin, Germany [email protected]
Abstract.

We provide a criterion for a coherent sheaf to be an Ulrich sheaf in terms of a certain bilinear form on its global sections. When working over the real numbers we call it a positive Ulrich sheaf if this bilinear form is symmetric or hermitian and positive definite. In that case our result provides a common theoretical framework for several results in real algebraic geometry concerning the existence of algebraic certificates for certain geometric properties. For instance, it implies Hilbert’s theorem on nonnegative ternary quartics, via the geometry of del Pezzo surfaces, and the solution of the Lax conjecture on plane hyperbolic curves due to Helton and Vinnikov.

2010 Mathematics Subject Classification:
Primary: 14P05, 14J60; Secondary: 14M12, 12D15
The second author has been supported by the DFG under Grant No.421473641.

1. Introduction

A widespread principle in real algebraic geometry is to find and use algebraic certificates for geometric statements. For example, a sum of squares representation of a homogeneous polynomial p[x1,,xn]2dp\in{\mathbb{R}}[x_{1},\ldots,x_{n}]_{2d} of degree 2d2d is a finite sequence of polynomials g1,,gr[x1,,xn]dg_{1},\dots,g_{r}\in{\mathbb{R}}[x_{1},\ldots,x_{n}]_{d} such that

p=g12++gr2p=g_{1}^{2}+\ldots+g_{r}^{2}

and serves as an algebraic certificate of the geometric property that pp takes only nonnegative values at real points: p(x)0p(x)\geq 0 for all xnx\in{\mathbb{R}}^{n}. In an influential paper David Hilbert [Hil88] showed that the converse is true if (and only if) 2d=22d=2, or n=1n=1, or (2d,n)=(4,3)(2d,n)=(4,3). While in the first two cases this can be seen quite easily via linear algebra and the fundamental theorem of algebra respectively, the proof for the case (2d,n)=(4,3)(2d,n)=(4,3) of ternary quartics is nontrivial. There have been several different new proofs of this statement in the last twenty years: via the Jacobian of the plane curve defined by pp [PRSS04] relying on results of Arthur Coble [Cob82], using elementary techniques [PS12], and as a special case of more general results on varieties of minimal degree [BSV16].

Another instance, that has attracted a lot of attention recently, is the following. Let h[x0,,xn]dh\in{\mathbb{R}}[x_{0},\ldots,x_{n}]_{d} be a homogeneous polynomial for which there are real symmetric (or complex hermitian) matrices A0,,AnA_{0},\ldots,A_{n} and r>0r\in{\mathbb{Z}}_{>0} such that

hr=det(x0A0++xnAn)h^{r}=\det(x_{0}A_{0}+\ldots+x_{n}A_{n})

and e0A0++enAne_{0}A_{0}+\ldots+e_{n}A_{n} is positive definite for some en+1e\in{\mathbb{R}}^{n+1}. In this case, we say that hrh^{r} has a definite symmetric (or hermitian) determinantal representation and it is a certificate that hh is hyperbolic with respect to ee in the sense that the univariate polynomial h(tev)[t]h(te-v)\in{\mathbb{R}}[t] has only real zeros for all vn+1v\in{\mathbb{R}}^{n+1}: The minimal polynomial of a hermitian matrix has only real zeros. Peter Lax [Lax58] conjectured that for n=2n=2 and arbitrary d>0d\in{\mathbb{Z}}_{>0}, the following strong converse is true: Every hyperbolic polynomial h[x0,x1,x2]dh\in{\mathbb{R}}[x_{0},x_{1},x_{2}]_{d} has a definite symmetric determinantal representation (up to multiplication with a nonzero scalar). This conjecture was solved to the affirmative by Bill Helton and Victor Vinnikov [HV07]. Furthermore, every hyperbolic h[x0,x1,x2,x3]3h\in{\mathbb{R}}[x_{0},x_{1},x_{2},x_{3}]_{3} has a definite hermitian determinantal representation [BK07] and for every quadratic hyperbolic polynomial h[x0,,xn]2h\in{\mathbb{R}}[x_{0},\ldots,x_{n}]_{2} there is a r>0r\in{\mathbb{Z}}_{>0} such that hrh^{r} has definite symmetric (or hermitian) determinantal representation [NT12]. On the other hand, if d4d\geq 4 and n3n\geq 3 or if d3d\geq 3 and n42n\geq 42, there are hyperbolic polynomials h[x0,,xn]dh\in{\mathbb{R}}[x_{0},\ldots,x_{n}]_{d} such that no power hrh^{r} has a definite symmetric (or hermitian) determinantal representation [Brä11, Sau19]. The cases (d,n)(d,n) with d=3d=3 and 3<n<423<n<42 are open.

The condition of h[x0,,xn]dh\in{\mathbb{R}}[x_{0},\ldots,x_{n}]_{d} being hyperbolic with respect to ee can be phrased in geometric terms as follows: Let X=𝒱(h)nX={\mathcal{V}}(h)\subset\mathbb{P}^{n} be the hypersurface defined by hh. Then hh is hyperbolic with respect to ee if and only if the linear projection πe:Xn1\pi_{e}:X\to\mathbb{P}^{n-1} with center ee is real fibered in the sense that πe1(n1())X()\pi_{e}^{-1}(\mathbb{P}^{n-1}({\mathbb{R}}))\subset X({\mathbb{R}}). This leads to a natural generalization of hyperbolicity to arbitrary embedded varieties introduced in [SV18] and further studied in [KS20a]: A subvariety XnX\subset\mathbb{P}^{n} of dimension kk is hyperbolic with respect to a linear subspace EnE\subset\mathbb{P}^{n} of codimension k+1k+1 if XE=X\cap E=\emptyset and the linear projection πE:Xk\pi_{E}:X\to\mathbb{P}^{k} with center EE is real fibered. Furthermore, a polynomial p[x1,,xn]2dp\in{\mathbb{R}}[x_{1},\ldots,x_{n}]_{2d} is nonnegative if and only if the double cover XnX\to\mathbb{P}^{n} ramified along the zero set of pp is real fibered, where XX is defined as the zero set of y2py^{2}-p in a suitable weighted projective space. Thus both above mentioned geometric properties of polynomials, being nonnegative and being hyperbolic, can be seen as special instances of real fibered morphisms.

Let f:XYf:X\to Y be a morphism of schemes. A coherent sheaf {\mathcal{F}} on XX is called ff-Ulrich if there is a natural number r>0r>0 such that f𝒪Yrf_{*}{\mathcal{F}}\cong{\mathcal{O}}_{Y}^{r}. If XnX\subset\mathbb{P}^{n} is a closed subscheme, then one defines a coherent sheaf {\mathcal{F}} on XX to be an Ulrich sheaf per se if it is π\pi-Ulrich for any finite surjective linear projection π:Xk\pi:X\to\mathbb{P}^{k}. This is equivalent to Hi(X,(j))=0H^{i}(X,{\mathcal{F}}(-j))=0 for 1jdim(X)1\leq j\leq\dim(X) and all ii. See [Bea18] for an introduction to Ulrich sheaves. The question of which subvarieties of n\mathbb{P}^{n} carry an Ulrich sheaf is of particular interest in the context of Boij–Söderberg theory [ES11].

For real fibered morphisms f:XYf:X\to Y positive ff-Ulrich sheaves have been defined in [KS20a] and it was shown that a hypersurface in n\mathbb{P}^{n} carries a positive Ulrich sheaf if and only if it is cut out by a polynomial with a definite determinantal representation. For subvarieties XnX\subset\mathbb{P}^{n} of higher codimension supporting a positive Ulrich sheaf is equivalent to admitting some generalized type of determinantal representation that was introduced in [SV18] and motivated by operator theory. The existence of such a determinantal representation for XX implies that some power of the Chow form of XX has a definite determinantal representation. On the other hand, we will show that a real fibered double cover f:Xnf:X\to\mathbb{P}^{n} ramified along the zero set of a homogeneous polynomial pp admits a positive ff-Ulrich sheaf if and only if pp is a sum of squares. Therefore, the notion of positive Ulrich sheaves encapsulates both types of algebraic certificates for the above mentioned geometric properties of homogeneous polynomials, namely being a sum of squares and having a definite determinantal representation.

The main result of this article is a criterion for a coherent sheaf to be a positive Ulrich sheaf which implies all the above mentioned existence results on representations as a sum of squares and determinantal representations, and more. It only comprises a positivity criterion that can be checked locally as well as a condition on the dimension of the space of global sections (but suprisingly not of the higher cohomology groups) of the coherent sheaf at hand.

To this end, after some preparations in Section 3, we characterize Ulrich sheaves in terms of a certain bilinear mapping and its behavior on the level of global sections in Section 4, see in particular 4.8. In this part we work over an arbitrary ground field and we believe that these results can be of independent interest.

After we review some facts on the codifferent sheaf in Section 5 that will be important later on, we focus on varieties over {\mathbb{R}}. In Section 6 we recall some facts about real fibered morphisms. We then define positive Ulrich sheaves in Section 8 after some preparations in Section 7. 8.3 is the above mentioned convenient criterion for checking whether a sheaf is a positive Ulrich sheaf. In Section 9 we show that for a given polynomial having a determinantal or sum of squares representation is equivalent to the existence of a certain positive Ulrich sheaf. From this the result on determinantal representations of quadratic hyperbolic polynomials from [NT12] follows directly. In order to make our general theory also applicable to other cases of interest, we specialize to Ulrich sheaves of rank one on irreducible varieties in Section 10. 10.3 gives a convenient criterion for a Weil divisor giving rise to a positive Ulrich sheaf. Namely, under some mild assumptions, if f:XYf:X\to Y is a real fibered morphism and DD a Weil divisor on XX such that 2D2D differs from the ramification divisor of ff only by a principal divisor defined by a nonnegative rational function, then the sheaf associated to DD is a positive ff-Ulrich sheaf whenever its space of global sections has dimension deg(f)\deg(f). A particularly nice case is that of hyperbolic hypersurfaces: The existence of a definite determinantal representation is guaranteed by the existence of a certain interlacer, generalizing the construction for plane curves from [PV13], see 10.10.

In Section 11 we apply our theory to the case of curves and show how it easily implies the Helton–Vinnikov theorem on plane hyperbolic curves [HV07] as well as its generalization to curves of higher codimension from [SV18] using the 22-divisibility of the Jacobian. Finally, in Section 12 we consider the anticanonical map on real del Pezzo surfaces in order to reprove Hilbert’s theorem on ternary quartics [Hil88] and the existence of a hermitian determinantal representation on cubic hyperbolic surfaces [BK07]. We further prove a new result on quartic del Pezzo surfaces in 4\mathbb{P}^{4}. Apart from our general theory, the only ingredients for this part are basic properties of (real) del Pezzo surfaces as well as the Riemann–Roch theorem.

2. Preliminaries and notation

For any scheme XX and pXp\in X we denote by κ(p)\kappa(p) the residue class field of XX at pp. If XX is separated, reduced (but not necessarily irreducible) and of finite type over a field KK, we say that XX is a variety over KK. For any coherent sheaf {\mathcal{F}} on XX we denote by rankp()\operatorname{rank}_{p}({\mathcal{F}}) the dimension of the fiber of {\mathcal{F}} at pp considered as κ(p)\kappa(p)-vector space. If XX is irreducible with generic point ξ\xi, we simply denote rank()=rankξ()\operatorname{rank}({\mathcal{F}})=\operatorname{rank}_{\xi}({\mathcal{F}}). If XX is a scheme (over a field KK) and LL a field (extension of KK), then we denote by X(L)X(L) the set of all morphisms Spec(L)X\operatorname{Spec}(L)\to X of schemes (over KK). For a field KK we let Kn=Proj(K[x0,,xn])\mathbb{P}^{n}_{K}=\textrm{Proj}(K[x_{0},\ldots,x_{n}]) and if the field is clear from the context we omit the index and just write n\mathbb{P}^{n}. We say that a scheme is noetherian if it can be covered by a finite number of open affine subsets that are spectra of noetherian rings.

3. Bilinear mappings on coherent sheaves

Definition 3.1.

Let XX be a scheme and let 1,2{\mathcal{F}}_{1},{\mathcal{F}}_{2} and 𝒢{\mathcal{G}} be coherent sheaves on XX. A 𝒢{\mathcal{G}}-valued pairing of 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} is a morphism of coherent sheaves φ:12𝒢\varphi:{\mathcal{F}}_{1}\otimes{\mathcal{F}}_{2}\to{\mathcal{G}}. Let KK be a field and αX(K)\alpha\in X(K), i.e., a morphism α:Spec(K)X\alpha:\operatorname{Spec}(K)\to X. Then we get a bilinear map αφ\alpha^{*}\varphi on α1×α2\alpha^{*}{\mathcal{F}}_{1}\times\alpha^{*}{\mathcal{F}}_{2} with values in α𝒢\alpha^{*}{\mathcal{G}} which are just finite dimensional KK-vector spaces. We say that φ\varphi is nondegenerate at αX(K)\alpha\in X(K) if the map α1HomK(α2,α𝒢)\alpha^{*}{\mathcal{F}}_{1}\to\operatorname{Hom}_{K}(\alpha^{*}{\mathcal{F}}_{2},\alpha^{*}{\mathcal{G}}) induced by αφ\alpha^{*}\varphi is an isomorphism.

For the rest of this section, unless stated otherwise, let XX always be a geometrically integral scheme with generic point ξ\xi which is proper over a field KK.

Lemma 3.2.

Let {\mathcal{F}} be a coherent sheaf on XX which is generated by global sections. If there is a KK-basis of H0(X,)H^{0}(X,{\mathcal{F}}) which is also a κ(ξ)\kappa(\xi)-basis of ξ{\mathcal{F}}_{\xi}, then 𝒪Xr{\mathcal{F}}\cong{\mathcal{O}}_{X}^{r} where r=dimH0(X,)r=\dim H^{0}(X,{\mathcal{F}}).

Proof.

Let 𝒦{\mathcal{K}} be the kernel of the map 𝒪Xr{\mathcal{O}}_{X}^{r}\to{\mathcal{F}} that sends the unit vectors to the KK-basis of H0(X,)H^{0}(X,{\mathcal{F}}). Since {\mathcal{F}} is generated by global sections, this map is surjective. We thus obtain the short exact sequence

0𝒦𝒪Xr0.0\to{\mathcal{K}}\to{\mathcal{O}}_{X}^{r}\to{\mathcal{F}}\to 0.

Passing to the stalk at ξ\xi gives 𝒦ξ=0{\mathcal{K}}_{\xi}=0 by our assumption. Since 𝒦{\mathcal{K}} is torsion-free as a subsheaf of 𝒪Xr{\mathcal{O}}_{X}^{r}, this implies that 𝒦=0{\mathcal{K}}=0 and therefore 𝒪Xr{\mathcal{O}}_{X}^{r}\cong{\mathcal{F}}. ∎

Remark 3.3.

Let φ:1𝒪X2𝒪X\varphi:{\mathcal{F}}_{1}\otimes_{{\mathcal{O}}_{X}}{\mathcal{F}}_{2}\to{\mathcal{O}}_{X} be a pairing of the coherent sheaves 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2}. This induces a bilinear mapping

V1×V2KV_{1}\times V_{2}\to K

where Vi=H0(X,i)V_{i}=H^{0}(X,{\mathcal{F}}_{i}) since H0(X,𝒪X)=KH^{0}(X,{\mathcal{O}}_{X})=K.

Lemma 3.4.

Let 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} be coherent sheaves on XX, let Vi=H0(X,i)V_{i}=H^{0}(X,{\mathcal{F}}_{i}) and let s1,,srs_{1},\ldots,s_{r} be a basis of V1V_{1}. Let φ:1𝒪X2𝒪X\varphi:{\mathcal{F}}_{1}\otimes_{{\mathcal{O}}_{X}}{\mathcal{F}}_{2}\to{\mathcal{O}}_{X} be a pairing such that the induced bilinear mapping V1×V2KV_{1}\times V_{2}\to K is nondegenerate. Then the images of s1,,srs_{1},\ldots,s_{r} in the κ(ξ)\kappa(\xi)-vector space (1)ξ({\mathcal{F}}_{1})_{\xi} are linearly independent.

Proof.

Since the bilinear mapping V1×V2KV_{1}\times V_{2}\to K is nondegenerate, there is a basis t1,,trV2t_{1},\ldots,t_{r}\in V_{2} that is dual to s1,,srs_{1},\ldots,s_{r} with respect to this bilinear mapping. Suppose f1,,frκ(ξ)f_{1},\ldots,f_{r}\in\kappa(\xi) such that

f1s1++frsr=0.f_{1}s_{1}+\ldots+f_{r}s_{r}=0.

Tensoring with tjt_{j} and applying φ\varphi yields fjφ(sjtj)=0f_{j}\cdot\varphi(s_{j}\otimes t_{j})=0 and therefore fj=0f_{j}=0 since φ(sjsj)=1\varphi(s_{j}\otimes s_{j})=1 by assumption. ∎

Proposition 3.5.

Let 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} be coherent sheaves on XX, let Vi=H0(X,i)V_{i}=H^{0}(X,{\mathcal{F}}_{i}) and let 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} be generated by global sections. Let φ:1𝒪X2𝒪X\varphi:{\mathcal{F}}_{1}\otimes_{{\mathcal{O}}_{X}}{\mathcal{F}}_{2}\to{\mathcal{O}}_{X} be a pairing such that the induced bilinear mapping V1×V2KV_{1}\times V_{2}\to K is nondegenerate. Then:

  1. a)

    i𝒪Xr{\mathcal{F}}_{i}\cong{\mathcal{O}}_{X}^{r} where r=dimVir=\dim V_{i};

  2. b)

    the morphism 1om𝒪X(2,𝒪X){\mathcal{F}}_{1}\to\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{F}}_{2},{\mathcal{O}}_{X}) corresponding to φ\varphi is an isomorphism.

Proof.

Let s1,,srV1s_{1},\ldots,s_{r}\in V_{1} be a basis of V1V_{1}. By assumption, s1,,srs_{1},\ldots,s_{r} span the κ(ξ)\kappa(\xi)-vector space (1)ξ({\mathcal{F}}_{1})_{\xi} and by Lemma 3.4 they are linearly independent. Thus by Lemma 3.2 we have 1𝒪Xr{\mathcal{F}}_{1}\cong{\mathcal{O}}_{X}^{r}. The same argument applies to 2{\mathcal{F}}_{2} and Part b)b) then follows immediately from a). ∎

Lemma 3.6.

Let 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} be coherent torsion-free sheaves on XX. Assume that the image of Vi=H0(X,i)V_{i}=H^{0}(X,{\mathcal{F}}_{i}) spans the stalk (i)ξ({\mathcal{F}}_{i})_{\xi} as κ(ξ)\kappa(\xi)-vector space. Furthermore, let φ:1𝒪X2𝒪X\varphi:{\mathcal{F}}_{1}\otimes_{{\mathcal{O}}_{X}}{\mathcal{F}}_{2}\to{\mathcal{O}}_{X} be a pairing such that the induced bilinear mapping V1×V2KV_{1}\times V_{2}\to K is nondegenerate. Then the corresponding morphism 1om𝒪X(2,𝒪X){\mathcal{F}}_{1}\to\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{F}}_{2},{\mathcal{O}}_{X}) is injective.

Proof.

Let s1,,srV1s_{1},\ldots,s_{r}\in V_{1} be a basis of V1V_{1} and let t1,,trV2t_{1},\ldots,t_{r}\in V_{2} be the dual basis with respect to the bilinear mapping V1×V2KV_{1}\times V_{2}\to K. Let UXU\subset X be some open affine subset, A=𝒪X(U)A={\mathcal{O}}_{X}(U) and Mi=i(U)M_{i}={\mathcal{F}}_{i}(U). For any 0gM10\neq g\in M_{1} there is 0aA0\neq a\in A such that aga\cdot g is in the submodule of M1M_{1} that is spanned by the restrictions si|Us_{i}|_{U}:

ag=f1s1|U++frsr|Ua\cdot g=f_{1}\cdot s_{1}|_{U}+\ldots+f_{r}\cdot s_{r}|_{U}

for some fjAf_{j}\in A that are not all zero. Let for instance fi0f_{i}\neq 0, then

φ(agti|U)=afiφ(siti)0.\varphi(a\cdot g\otimes t_{i}|_{U})=a\cdot f_{i}\cdot\varphi(s_{i}\otimes t_{i})\neq 0.

This shows that the map MHomA(M,A)M\to\operatorname{Hom}_{A}(M,A) induced by φ\varphi in injective. ∎

Lemma 3.7.

Let {\mathcal{F}} be a coherent sheaf on XX and 𝒢{\mathcal{G}} a subsheaf with 𝒢ξ=ξ{\mathcal{G}}_{\xi}={\mathcal{F}}_{\xi}. Then the natural map om𝒪X(,𝒪X)om𝒪X(𝒢,𝒪X)\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{F}},{\mathcal{O}}_{X})\to\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{G}},{\mathcal{O}}_{X}) is injective.

Proof.

Let UXU\subset X be some open affine subset, A=𝒪X(U)A={\mathcal{O}}_{X}(U), M=(U)M={\mathcal{F}}(U) and N=𝒢(U)N={\mathcal{G}}(U). Consider a morphism φ:MA\varphi:M\to A such that φ|N=0\varphi|_{N}=0. For every gMg\in M there is a nonzero tAt\in A such that tgNt\cdot g\in N. Thus φ(tg)=tφ(g)=0\varphi(t\cdot g)=t\cdot\varphi(g)=0 and therefore φ(g)=0\varphi(g)=0. This shows that φ=0\varphi=0. ∎

Theorem 3.8.

Let 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} be coherent torsion-free sheaves on XX and let Vi=H0(X,i)V_{i}=H^{0}(X,{\mathcal{F}}_{i}). Let φ:1𝒪X2𝒪X\varphi:{\mathcal{F}}_{1}\otimes_{{\mathcal{O}}_{X}}{\mathcal{F}}_{2}\to{\mathcal{O}}_{X} be a pairing such that the induced bilinear mapping V1×V2KV_{1}\times V_{2}\to K is nondegenerate. If dimV1rank1\dim V_{1}\geq\operatorname{rank}{\mathcal{F}}_{1}, then 1𝒪Xr{\mathcal{F}}_{1}\cong{\mathcal{O}}_{X}^{r} and r=dimV1=rank1r=\dim V_{1}=\operatorname{rank}{\mathcal{F}}_{1}. Furthermore, 1om𝒪X(2,𝒪X){\mathcal{F}}_{1}\to\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{F}}_{2},{\mathcal{O}}_{X}) is an isomorphism.

Proof.

By Proposition 3.5a) it suffices to show that i{\mathcal{F}}_{i} is generated by global sections. Let 𝒢i{\mathcal{G}}_{i} be the subsheaf of i{\mathcal{F}}_{i} generated by its global sections ViV_{i}. We get the commutative diagram

𝒢1om𝒪X(𝒢2,𝒪X)1om𝒪X(2,𝒪X).\leavevmode\hbox to114.02pt{\vbox to53.75pt{\pgfpicture\makeatletter\hbox{\hskip 57.01001pt\lower-26.92593pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-57.01001pt}{-26.8261pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 9.62915pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-5.32361pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${{\mathcal{G}}_{1}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 9.62915pt\hfil&\hfil\hskip 59.38084pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-31.07533pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{G}}_{2},{\mathcal{O}}_{X})}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 35.38087pt\hfil\cr\vskip 18.00005pt\cr\hfil\hskip 8.96944pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-4.6639pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${{\mathcal{F}}_{1}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 8.96944pt\hfil&\hfil\hskip 58.72113pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-30.41562pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{F}}_{2},{\mathcal{O}}_{X})}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 34.72116pt\hfil\cr}}}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} { {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{{ {\pgfsys@beginscope \pgfsys@setdash{}{0.0pt}\pgfsys@roundcap\pgfsys@roundjoin{} {}{}{} {}{}{} \pgfsys@moveto{-2.07988pt}{2.39986pt}\pgfsys@curveto{-1.69989pt}{0.95992pt}{-0.85313pt}{0.27998pt}{0.0pt}{0.0pt}\pgfsys@curveto{-0.85313pt}{-0.27998pt}{-1.69989pt}{-0.95992pt}{-2.07988pt}{-2.39986pt}\pgfsys@stroke\pgfsys@endscope}} }{}{}{{}}{}{}{{}}\pgfsys@moveto{-47.38086pt}{3.33588pt}\pgfsys@lineto{-47.38086pt}{-15.73311pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{-47.38086pt}{-15.93309pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-37.55171pt}{11.50003pt}\pgfsys@lineto{-14.3517pt}{11.50003pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-14.15172pt}{11.50003pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-38.21143pt}{-24.3261pt}\pgfsys@lineto{-13.69199pt}{-24.3261pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-13.492pt}{-24.3261pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{21.62914pt}{-15.46638pt}\pgfsys@lineto{21.62914pt}{1.7337pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{1.0}{-1.0}{0.0}{21.62914pt}{1.93369pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}.

The homomorphism 𝒢1om𝒪X(𝒢2,𝒪X){\mathcal{G}}_{1}\to\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{G}}_{2},{\mathcal{O}}_{X}) is an isomorphism by Proposition 3.5b). By Lemma 3.4 and the condition on the dimension it follows that the image of Vi=H0(X,i)V_{i}=H^{0}(X,{\mathcal{F}}_{i}) spans the stalk (i)ξ({\mathcal{F}}_{i})_{\xi} as κ(ξ)\kappa(\xi)-vector space and that (𝒢i)ξ=(i)ξ({\mathcal{G}}_{i})_{\xi}=({\mathcal{F}}_{i})_{\xi}. Thus, the bottom and right maps in the diagram are also injective by Lemmas 3.6 and 3.7, respectively. This implies that 𝒢1=2{\mathcal{G}}_{1}={\mathcal{F}}_{2} and therefore 1{\mathcal{F}}_{1} is generated by global sections. ∎

Example 3.9.

This example is to illustrate that the assumption in 3.8 of being nondegenerate on global sections is crucial. Let X=1X={\mathbb{P}^{1}} and ξ\xi the generic point of 1{\mathbb{P}^{1}}. Consider the coherent torsion-free sheaf =𝒪1(1)𝒪1(1){\mathcal{F}}={\mathcal{O}}_{\mathbb{P}^{1}}(1)\oplus{\mathcal{O}}_{\mathbb{P}^{1}}(-1) on 1{\mathbb{P}^{1}}. We have dimH0(1,)=2=rankξ\dim H^{0}({\mathbb{P}^{1}},{\mathcal{F}})=2=\operatorname{rank}_{\xi}{\mathcal{F}}. On {\mathcal{F}} we define the pairing φ:𝒪X𝒪X\varphi:{\mathcal{F}}\otimes_{{\mathcal{O}}_{X}}{\mathcal{F}}\to{\mathcal{O}}_{X} that sends (a,b)(c,d)(a,b)\otimes(c,d) to ad+bcad+bc which is nondegenerate at ξ\xi. But the induced bilinear form on the global sections of {\mathcal{F}} is identically zero and ≇𝒪12{\mathcal{F}}\not\cong{\mathcal{O}}_{\mathbb{P}^{1}}^{2}.

4. Ulrich sheaves

Definition 4.1.

Let f:XYf:X\to Y be a morphism of schemes. A coherent sheaf {\mathcal{F}} on XX is ff-Ulrich if f𝒪Yrf_{*}{\mathcal{F}}\cong{\mathcal{O}}_{Y}^{r} for some natural number rr.

Remark 4.2.

Let fi:XiYf_{i}:X_{i}\to Y be finitely many morphisms of schemes. Let XX be the disjoint union of the XiX_{i} and f:XYf:X\to Y the morphism induced by the fif_{i}. A coherent sheaf {\mathcal{F}} on XX is ff-Ulrich if and only if |Xi{\mathcal{F}}|_{X_{i}} is fif_{i}-Ulrich for all ii. Thus we can usually restrict to the case when XX is connected.

A case of particular interest is when XX is an embedded projective variety.

Proposition 4.3 (Prop. 2.1 in [ESW03]).

Let XnX\subset\mathbb{P}^{n} be a closed subscheme and f:Xkf:X\to\mathbb{P}^{k} a finite surjective linear projection from a center that is disjoint from XX. Let {\mathcal{F}} be a coherent sheaf whose support is all of XX. Then the following are equivalent:

  1. (i)

    {\mathcal{F}} is ff-Ulrich;

  2. (ii)

    Hi(X,(i))=0H^{i}(X,{\mathcal{F}}(-i))=0 for i>0i>0 and Hi(X,(i1))=0H^{i}(X,{\mathcal{F}}(-i-1))=0 for i<ki<k;

  3. (iii)

    Hi(X,(j))=0H^{i}(X,{\mathcal{F}}(j))=0 for all 1ik11\leq i\leq k-1, jj\in{\mathbb{Z}}; H0(X,(j))=0H^{0}(X,{\mathcal{F}}(j))=0 for j<0j<0 and Hk(X,(j))=0H^{k}(X,{\mathcal{F}}(j))=0 for jkj\geq-k;

  4. (iv)

    the module M=iH0(X,(i))M=\oplus_{i\in{\mathbb{Z}}}H^{0}(X,{\mathcal{F}}(i)) of twisted global sections is a Cohen–Macaulay module over the polynomial ring S=K[x0,,xn]S=K[x_{0},\ldots,x_{n}] of dimension k+1k+1 whose minimal SS-free resolution is linear.

If {\mathcal{F}} as in 4.3 satisfies these equivalent conditions, then we say that {\mathcal{F}} is an Ulrich sheaf on XX without specifying the finite surjective linear projection ff as conditions (ii)(ii)(iv)(iv) do not depend on the choice of ff. A major open question in this context is the following:

Problem 4.4 (p. 543 in [ESW03]).

Is there an Ulrich sheaf on every closed subvariety XnX\subset\mathbb{P}^{n}?

We now want to apply the results from Section 3 to give a criterion for a sheaf to be Ulrich. For this we need a relative notion of nondegenerate bilinear mappings. Let f:XYf:X\to Y be a finite morphism of noetherian schemes. For any quasi-coherent sheaf 𝒢{\mathcal{G}} on YY we consider the sheaf om𝒪Y(f𝒪X,𝒢)\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{Y}}(f_{*}{\mathcal{O}}_{X},{\mathcal{G}}). Since this is a quasi-coherent f𝒪Xf_{*}{\mathcal{O}}_{X}-module, it corresponds to a quasi-coherent 𝒪X{\mathcal{O}}_{X}-module which we will denote by f!𝒢f^{!}{\mathcal{G}}. We recall the following basic lemma (cf. [Har77, III §6, Ex. 6.10]).

Lemma 4.5.

Let f:XYf:X\to Y be a finite morphism of noetherian schemes. Let {\mathcal{F}} be a coherent sheaf on XX and 𝒢{\mathcal{G}} be a quasi-coherent sheaf on YY. There is a natural isomorphism

fom𝒪X(,f!𝒢)om𝒪Y(f,𝒢)f_{*}\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{F}},f^{!}{\mathcal{G}})\to\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{Y}}(f_{*}{\mathcal{F}},{\mathcal{G}})

of quasi-coherent f𝒪Xf_{*}{\mathcal{O}}_{X}-modules.

Let f:XYf:X\to Y be a finite morphism of noetherian schemes. Let 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} be coherent sheaves on XX and consider an f!𝒪Yf^{!}{\mathcal{O}}_{Y}-valued pairing, i.e., a morphism 12f!𝒪Y{\mathcal{F}}_{1}\otimes{\mathcal{F}}_{2}\to f^{!}{\mathcal{O}}_{Y} of coherent 𝒪X{\mathcal{O}}_{X}-modules. This corresponds to a morphism 1om𝒪X(2,f!𝒪Y){\mathcal{F}}_{1}\to\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{F}}_{2},f^{!}{\mathcal{O}}_{Y}). Lemma 4.5 tells us that this gives us a morphism

f1om𝒪Y(f2,𝒪Y)f_{*}{\mathcal{F}}_{1}\to\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{Y}}(f_{*}{\mathcal{F}}_{2},{\mathcal{O}}_{Y})

which corresponds to an 𝒪Y{\mathcal{O}}_{Y}-valued pairing on the pushforwards f1f_{*}{\mathcal{F}}_{1} and f2f_{*}{\mathcal{F}}_{2}.

Remark 4.6.

Let us explain here the affine case in more detail. To that end let X=Spec(B)X=\operatorname{Spec}(B), Y=Spec(A)Y=\operatorname{Spec}(A) and f:XYf:X\to Y be induced by the finite ring homomorphism f#:ABf^{\#}:A\to B. Then f!𝒪Yf^{!}{\mathcal{O}}_{Y} is the sheaf on XX associated to the BB-module HomA(B,A)\operatorname{Hom}_{A}(B,A) whose BB-module structure is given by (bφ)(m)=φ(bm)(b\cdot\varphi)(m)=\varphi(b\cdot m) for all b,mBb,m\in B. If 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} are the coherent sheaves associated to the BB-modules M1M_{1} and M2M_{2}, then a morphism 12f!𝒪Y{\mathcal{F}}_{1}\otimes{\mathcal{F}}_{2}\to f^{!}{\mathcal{O}}_{Y} of coherent 𝒪X{\mathcal{O}}_{X}-modules corresponds to a homomorphism ψ:M1BM2HomA(B,A)\psi:M_{1}\otimes_{B}M_{2}\to\operatorname{Hom}_{A}(B,A) of BB-modules. This gives rise to the following AA-bilinear map on M1×M2M_{1}\times M_{2}:

(m1,m2)(ψ(m1,m2))(1).(m_{1},m_{2})\mapsto(\psi(m_{1},m_{2}))(1).

This gives the 𝒪Y{\mathcal{O}}_{Y}-valued pairing of the pushforwards f1f_{*}{\mathcal{F}}_{1} and f2f_{*}{\mathcal{F}}_{2}.

Definition 4.7.

Let f:XYf:X\to Y be a finite morphism of noetherian schemes. Let 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} be coherent sheaves on XX and consider an f!𝒪Yf^{!}{\mathcal{O}}_{Y}-valued pairing φ:12f!𝒪Y\varphi:{\mathcal{F}}_{1}\otimes{\mathcal{F}}_{2}\to f^{!}{\mathcal{O}}_{Y}. For a field KK we say that φ\varphi is nondegenerate at αY(K)\alpha\in Y(K) if the induced 𝒪Y{\mathcal{O}}_{Y}-valued pairing of the pushforwards f1f_{*}{\mathcal{F}}_{1} and f2f_{*}{\mathcal{F}}_{2} is nondegenerate at α\alpha.

Now let YY be a geometrically irreducible variety which is proper over a field KK. Let further f:XYf:X\to Y be a finite surjective morphism and 1{\mathcal{F}}_{1}, 2{\mathcal{F}}_{2} coherent sheaves on XX with a pairing 12f!𝒪Y{\mathcal{F}}_{1}\otimes{\mathcal{F}}_{2}\to f^{!}{\mathcal{O}}_{Y}. We have seen that this induces an 𝒪Y{\mathcal{O}}_{Y}-valued pairing on the pushforwards which in turn induces a KK-bilinear mapping

H0(X,1)×H0(X,2)K.H^{0}(X,{\mathcal{F}}_{1})\times H^{0}(X,{\mathcal{F}}_{2})\to K.
Theorem 4.8.

Let XX be an equidimensional variety over a field KK with irreducible components X1,,XsX_{1},\ldots,X_{s}. Let 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} be coherent torsion-free sheaves on XX and let Vi=H0(X,i)V_{i}=H^{0}(X,{\mathcal{F}}_{i}). Let YY be a geometrically irreducible variety which is proper over KK and f:XYf:X\to Y a finite surjective morphism. Assume that there is an f!𝒪Yf^{!}{\mathcal{O}}_{Y}-valued pairing of 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} such that the induced KK-bilinear mapping V1×V2KV_{1}\times V_{2}\to K is nondegenerate. Then the following are equivalent:

  1. (i)

    dimV1i=1sdeg(f|Xi)rank(1|Xi)\dim V_{1}\geq\sum_{i=1}^{s}\deg(f|_{X_{i}})\cdot\operatorname{rank}({\mathcal{F}}_{1}|_{X_{i}});

  2. (ii)

    1{\mathcal{F}}_{1} is ff-Ulrich.

Proof.

We will apply 3.8 to the coherent sheaves f1f_{*}{\mathcal{F}}_{1} and f2f_{*}{\mathcal{F}}_{2}. First we need that the fif_{*}{\mathcal{F}}_{i} are torsion-free. This follows from the assumptions that ff is finite and surjective, XX is equidimensional and i{\mathcal{F}}_{i} is torsion-free. Further by [Kol96, VI §2, Prop. 2.7] one has rank(f1)=i=1sdeg(f|Xi)rank(1|Xi).\operatorname{rank}(f_{*}{\mathcal{F}}_{1})=\sum_{i=1}^{s}\deg(f|_{X_{i}})\cdot\operatorname{rank}({\mathcal{F}}_{1}|_{X_{i}}).

5. The codifferent sheaf

In this section we recall some properties of f!𝒪Yf^{!}{\mathcal{O}}_{Y} and its relation to the codifferent sheaf. Most of the results should be well known, but for a lack of adequate references we will include (at least partial) proofs here.

Lemma 5.1.

Let f:XYf:X\to Y be a finite morphism of noetherian schemes.

  1. a)

    If ff is flat and both XX and YY are Gorenstein, then f!𝒪Yf^{!}{\mathcal{O}}_{Y} is a line bundle.

  2. b)

    If YY is a smooth variety over KK and XX is Gorenstein, then f!𝒪Yf^{!}{\mathcal{O}}_{Y} is a line bundle.

Proof.

We first note that b)b) is a special case of part a)a) because in this situation the morphism ff is automatically flat as Gorenstein implies Cohen-Macaulay. In order to prove part a)a) note that for every yYy\in Y the canonical module of 𝒪Y,y{\mathcal{O}}_{Y,y} is 𝒪Y,y{\mathcal{O}}_{Y,y} itself by [HK71, Satz 5.9], and for every xXx\in X the canonical module of 𝒪X,x{\mathcal{O}}_{X,x} is (f!𝒪Y)x(f^{!}{\mathcal{O}}_{Y})_{x} by [HK71, Satz 5.12]. Thus, again by [HK71, Satz 5.9], f!𝒪Yf^{!}{\mathcal{O}}_{Y} is a line bundle on XX if XX is Gorenstein. ∎

Remark 5.2.

The preceding lemma implies that the sheaf f!𝒪Yf^{!}{\mathcal{O}}_{Y} is a line bundle whenever f:XYf:X\to Y is a finite morphism of smooth varieties over a field KK.

Definition 5.3.

Let AA be a noetherian integral domain and ABA\subset B be a finite ring extension such that for each minimal prime ideal 𝔭{\mathfrak{p}} of BB we have 𝔭A=(0){\mathfrak{p}}\cap A=(0). Let KK be the quotient field of AA and let LL be the total quotient ring of BB. Then LL is a finite dimensional KK-vector space and we have the KK-linear map trL/K:LK\operatorname{tr}_{L/K}:L\to K that associates to every element xLx\in L the trace of the KK-linear map LL,aaxL\to L,\,a\mapsto ax. The codifferent of the ring extension ABA\subset B is the BB-module

Δ(B/A)={gL:trL/K(gB)A}.\Delta(B/A)=\{g\in L:\,\operatorname{tr}_{L/K}(g\cdot B)\subset A\}.

Clearly, the map

Δ(B/A)HomA(B,A),gtrL/K(g)\Delta(B/A)\to\operatorname{Hom}_{A}(B,A),\,g\mapsto\operatorname{tr}_{L/K}(g\cdot-)

is a homomorphism of BB-modules.

Now let YY be an integral noetherian scheme and f:XYf:X\to Y a finite morphism such the generic point of each irreducible component of XX is mapped to the generic point of YY. Let 𝒦X{\mathcal{K}}_{X} be the sheaf of total quotient rings of 𝒪X{\mathcal{O}}_{X}. By glueing the above we define the quasi-coherent subsheaf Δ(X/Y)\Delta(X/Y) of 𝒦X{\mathcal{K}}_{X} and we obtain a morphism of 𝒪X{\mathcal{O}}_{X}-modules Δ(X/Y)f!𝒪Y\Delta(X/Y)\to f^{!}{\mathcal{O}}_{Y}. We call Δ(X/Y)\Delta(X/Y) the codifferent sheaf.

Example 5.4.

Let AA be an integral domain and K=Quot(A)K=\operatorname{Quot}(A). Let fA[t]f\in A[t] be a monic polynomial over AA which has only simple zeros in the algebraic closure of KK and let B=A[t]/(f)B=A[t]/(f). Then the codifferent Δ(B/A)\Delta(B/A) is the fractional ideal generated by 1f\frac{1}{f^{\prime}} in the total quotient ring of BB. Here ff^{\prime} denotes the formal derivative of ff. This follows from a lemma often attributed to Euler, see [Ser79, III, §6].

Remark 5.5.

In order to construct a pairing 12f!𝒪Y{\mathcal{F}}_{1}\otimes{\mathcal{F}}_{2}\to f^{!}{\mathcal{O}}_{Y} it thus suffices by the discussion in 5.3 to construct a pairing 12𝒦X{\mathcal{F}}_{1}\otimes{\mathcal{F}}_{2}\to{\mathcal{K}}_{X} whose image is contained in Δ(X/Y)\Delta(X/Y).

Lemma 5.6.

Let XX and YY be varieties over a field of characteristic zero, XX equidimensional and YY irreducible. Then for any finite surjective morphism f:XYf:X\to Y the map Δ(X/Y)f!𝒪Y\Delta(X/Y)\to f^{!}{\mathcal{O}}_{Y} is an isomorphism of 𝒪X{\mathcal{O}}_{X}-modules.

Proof.

We can reduce to the affine case: Let ABA\subset B be a finite extension of finitely generated KK-algebras such that AA is an integral domain and 𝔭A=(0){\mathfrak{p}}\cap A=(0) for all minimal prime ideals 𝔭{\mathfrak{p}} of BB. Let EE and FF be the total quotient rings of AA and BB, respectively. Since BAEB\otimes_{A}E is finite dimensional as an EE-vector space and reduced, it is a ring of the form F1××FrF_{1}\times\cdots\times F_{r} for some finite field extensions of FiF_{i} of EE. Since no element of AA is a zero divisor in BB, we actually have that F=F1××FrF=F_{1}\times\cdots\times F_{r}. Thus we have an injective map HomA(B,A)HomE(F,E)\operatorname{Hom}_{A}(B,A)\to\operatorname{Hom}_{E}(F,E) that is given by tensoring with EE. By definition, the preimage of HomA(B,A)\operatorname{Hom}_{A}(B,A) under the map

ψ:FHomE(F,E),atrF/E(a)\psi:F\to\operatorname{Hom}_{E}(F,E),\,a\mapsto\operatorname{tr}_{F/E}(a\cdot-)

is exactly Δ(B/A)\Delta(B/A). It thus suffices to show that ψ\psi is an isomorphism because then the restriction of ψ\psi to Δ(B/A)\Delta(B/A) is the desired isomorphism Δ(B/A)HomA(B,A)\Delta(B/A)\to\operatorname{Hom}_{A}(B,A). The map ψ\psi is the direct sum of the maps

ψi:FiHomE(Fi,E),atrFi/E(a)\psi_{i}:F_{i}\to\operatorname{Hom}_{E}(F_{i},E),\,a\mapsto\operatorname{tr}_{F_{i}/E}(a\cdot-)

and thus it suffices to show that each ψi\psi_{i} is an isomorphism. We have that trFi/E(1)\operatorname{tr}_{F_{i}/E}(1) is the dimension of FiF_{i} as an EE-vector space. Since we work over a field of characteristic zero, this shows that each ψi\psi_{i} is a nonzero map. But since HomE(Fi,E)\operatorname{Hom}_{E}(F_{i},E) is one dimensional considered as a vector space over FiF_{i}, this implies that ψi\psi_{i} is an isomorphism. ∎

Remark 5.7.

5.6 is no longer true over fields of positive characteristic because the trace trL/K\operatorname{tr}_{L/K} is identically zero for inseparable field extensions KLK\subset L. This is one reason why we have not worked with the codifferent sheaf to begin with.

Proposition 5.8.

Let f:XYf:X\to Y be a finite surjective morphism of varieties over a field of characteristic zero. Let XX be equidimensional and Gorenstein and let YY be smooth and irreducible. Then Δ(X/Y)\Delta(X/Y) is an invertible subsheaf of 𝒦X{\mathcal{K}}_{X} and thus Δ(X/Y)=(R)\Delta(X/Y)={\mathcal{L}}(R) for some Cartier divisor RR on XX. This Cartier divisor is effective and its support consists exactly of those points where ff is ramified.

Proof.

By 5.6 Δ(X/Y)\Delta(X/Y) is isomorphic to f!𝒪Yf^{!}{\mathcal{O}}_{Y} which is an invertible sheaf by 5.1. Thus Δ(X/Y)\Delta(X/Y) is an invertible subsheaf of 𝒦X{\mathcal{K}}_{X} and we can write Δ(X/Y)=(R)\Delta(X/Y)={\mathcal{L}}(R) for some Cartier divisor RR on XX. We first show that RR is effective which is equivalent to the constant 11 being a global section of Δ(X/Y)\Delta(X/Y). This can be checked locally. We thus consider a finite ring extension ABA\subset B where AA is an integral domain. Furthermore, this ring extension is flat by the assumptions on XX and YY. Thus without loss of generality we can assume that BB is free as AA-module. Therefore, the AA-linear map BBB\to B given by multiplication with an element bBb\in B can be represented by a matrix having entries in AA. Using the notation of 5.3 this shows that trL/K(1B)A\operatorname{tr}_{L/K}(1\cdot B)\subset A. Thus the constant 11 is a global section of Δ(X/Y)\Delta(X/Y) and RR is effective. The image of 11 under the map H0(X,Δ(X/Y))H0(X,f!𝒪Y)H^{0}(X,\Delta(X/Y))\to H^{0}(X,f^{!}{\mathcal{O}}_{Y}) is just the trace map and the subscheme associated to RR is the zero set of this section. This consists exactly of the ramification points of ff, see for example [Sta20, Tag 0BW9] or [Kum16a, Rem. 2.2.19]. ∎

Definition 5.9.

In the situation of 5.8 we call the Cartier divisor RR on XX that corresponds to the invertible subsheaf Δ(X/Y)\Delta(X/Y) of 𝒦X{\mathcal{K}}_{X} the ramification divisor of ff.

Lemma 5.10.

Let f:XYf:X\to Y be a finite surjective morphism of varieties. Let XX be equidimensional and let YY be smooth and irreducible. Let ZXZ\subset X be of codimension at least two. Consider the open subset V=XZV=X\setminus Z and its inclusion ι:VX\iota\colon V\to X to XX. Then ι(Δ|V)=Δ\iota_{*}(\Delta|_{V})=\Delta where Δ=Δ(X/Y)\Delta=\Delta(X/Y).

Proof.

If we enlarge ZZ, then the statement becomes stronger, so we may assume that Z=π1(Z)Z=\pi^{-1}(Z^{\prime}) for some ZYZ^{\prime}\subset Y of codimension at least two. We write 𝒟=ι(Δ|V){\mathcal{D}}=\iota_{*}(\Delta|_{V}). Then Δ\Delta is a subsheaf of 𝒟{\mathcal{D}} and we need to show equality. To that end let UYU\subset Y be an affine open subset and W=π1(U)W=\pi^{-1}(U). Then we have the following:

  1. (1)

    𝒪X(W)𝒪X(WZ){\mathcal{O}}_{X}(W)\subset{\mathcal{O}}_{X}(W\setminus Z) and

  2. (2)

    𝒪Y(U)=𝒪Y(UZ){\mathcal{O}}_{Y}(U)={\mathcal{O}}_{Y}(U\setminus Z^{\prime}).

Letting LL be the total quotient ring of XX and KK the function field of YY, we get

𝒟(W)=Δ(π1(UZ))={aLtrL/K(a𝒪X(WZ))𝒪Y(UZ)}{\mathcal{D}}(W)=\Delta(\pi^{-1}(U\setminus Z^{\prime}))=\{a\in L\mid\operatorname{tr}_{L/K}(a{\mathcal{O}}_{X}(W\setminus Z))\subset{\mathcal{O}}_{Y}(U\setminus Z^{\prime})\,\}

and due to (1) and (2) the latter is contained in

{aLtrL/K(a𝒪X(W))𝒪Y(U)}=Δ(W).\{\,a\in L\mid\operatorname{tr}_{L/K}(a{\mathcal{O}}_{X}(W))\subset{\mathcal{O}}_{Y}(U)\,\}=\Delta(W).

Thus we have 𝒟(W)=Δ(W){\mathcal{D}}(W)=\Delta(W). Since f:XYf\colon X\to Y is affine, the sets W=π1(U)W=\pi^{-1}(U) for UYU\subset Y open and affine give an affine covering of XX and thus 𝒟=Δ{\mathcal{D}}=\Delta. ∎

6. Real fibered morphisms

In this section we recall the notion of real fibered morphisms, basic examples and some of their properties.

Definition 6.1.

Let f:XYf:X\to Y be a morphism of varieties over {\mathbb{R}}. If f1(Y())=X()f^{-1}(Y({\mathbb{R}}))=X({\mathbb{R}}), then we say that ff is real fibered.

Example 6.2.

Let p[x0,,xn]2dp\in{\mathbb{R}}[x_{0},\ldots,x_{n}]_{2d} be a homogeneous polynomial of degree 2d2d. Inside the weighted projective space (d,1,,1)\mathbb{P}(d,1,\ldots,1) we consider the hypersurface XX defined by y2=p(x0,,xn)y^{2}=p(x_{0},\ldots,x_{n}) and the natural projection π:Xn\pi:X\to\mathbb{P}^{n} onto the xx-coordinates. This is a double cover of n\mathbb{P}^{n} ramified at the hypersurface defined by p=0p=0. Clearly π\pi is real fibered if and only if pp is globally nonnegative, i.e., p(x)0p(x)\geq 0 for all xn+1x\in{\mathbb{R}}^{n+1}.

Hyperbolic polynomials yield another class of examples.

Definition 6.3.

Let h[x0,,xn]dh\in{\mathbb{R}}[x_{0},\ldots,x_{n}]_{d} be a homogeneous polynomial of degree dd and let en+1e\in{\mathbb{R}}^{n+1}. We say that hh is hyperbolic with respect to ee if the univariate polynomial h(tev)[t]h(te-v)\in{\mathbb{R}}[t] has only real zeros for all vn+1v\in{\mathbb{R}}^{n+1}. Note that this implies h(e)0h(e)\neq 0. A hypersurface XnX\subset\mathbb{P}^{n} is called hyperbolic with respect to [e][e] if its defining polynomial is hyperbolic with respect to ee.

Refer to caption
Figure 1. A plane quartic curve that is hyperbolic with respect to any point in the inner oval.
Example 6.4.

Let XnX\subset\mathbb{P}^{n} be a hypersurface and ene\in\mathbb{P}^{n} a point that does not lie on XX. Then the linear projection πe:Xn1\pi_{e}:X\to\mathbb{P}^{n-1} with center ee is real fibered if and only if XX is hyperbolic with respect to ee.

One can generalize this notion naturally to varieties of higher codimension.

Definition 6.5.

Let XnX\subset\mathbb{P}^{n} be a variety of pure dimension dd and EnE\subset\mathbb{P}^{n} a linear subspace of codimension d+1d+1 which does not intersect XX. We say that XX is hyperbolic with respect to EE if the linear projection πE:Xd\pi_{E}:X\to\mathbb{P}^{d} with center EE is real fibered.

An important feature of real fibered morphisms is the following.

Theorem 6.6 (Thm. 2.19 in [KS20a]).

Let f:XYf:X\to Y be a real fibered morphism of smooth varieties over {\mathbb{R}}. Then ff is unramified over X()X({\mathbb{R}}).

The following property of real fibered morphisms will come in handy later when we want to construct positive semidefinite bilinear forms.

Proposition 6.7.

Let YY be an irreducible smooth variety and let XX be an equidimensional variety over {\mathbb{R}}. Let f:XYf:X\to Y be a finite surjective real fibered morphism. Let KK and LL be the total quotient rings of YY and XX and trL/K:LK\operatorname{tr}_{L/K}:L\to K the trace map. If gLg\in L is nonnegative on X()X({\mathbb{R}}) (whereever it is defined), then trL/K(g)\operatorname{tr}_{L/K}(g) is nonnegative on Y()Y({\mathbb{R}}) (whereever it is defined).

Proof.

Assume that gLg\in L is nonnegative on X()X({\mathbb{R}}) (whereever it is defined). By generic freeness [Gro65, Lem. 6.9.2] there is a nonempty open affine subset UYU\subset Y such that B=𝒪X(f1(U))B={\mathcal{O}}_{X}(f^{-1}(U)) is a free AA-module where A=𝒪Y(U)A={\mathcal{O}}_{Y}(U) and such that gBg\in B. By a version of the Artin–Lang theorem [Bec82, Lem. 1.5] the real points of UU are dense in Y()Y({\mathbb{R}}) with respect to the euclidean topology because YY is smooth. Thus it suffices to show that trL/K(g)A\operatorname{tr}_{L/K}(g)\in A is nonnegative on every real point pp of UU. Let 𝔪A{\mathfrak{m}}\subset A be the corresponding maximal ideal. Then C=B/𝔪BC=B/{\mathfrak{m}}B is finite dimensional as vector space over =A/𝔪{\mathbb{R}}=A/{\mathfrak{m}} and Spec(C)\operatorname{Spec}(C) consists only of real points because ff is real fibered. Thus letting g¯C\overline{g}\in C be the residue class of gg and because gg is nonnegative on f1(p)f^{-1}(p), the quadratic form

C×C,(a,b)trC/(g¯ab)C\times C\to{\mathbb{R}},\,(a,b)\mapsto\operatorname{tr}_{C/{\mathbb{R}}}(\overline{g}\cdot a\cdot b)

is positive semidefinite by [PRS93, Thm. 2.1]. Thus in particular trC/(g¯)0\operatorname{tr}_{C/{\mathbb{R}}}(\overline{g})\geq 0. Finally, by flatness, we have that trL/K(g)¯=trC/(g¯)0\overline{\operatorname{tr}_{L/K}(g)}=\operatorname{tr}_{C/{\mathbb{R}}}(\overline{g})\geq 0. ∎

7. Symmetric and Hermitian bilinear forms

Definition 7.1.

Let XX be a scheme and let 1,2{\mathcal{F}}_{1},{\mathcal{F}}_{2} and 𝒢{\mathcal{G}} be coherent sheaves on XX. Let α:12\alpha:{\mathcal{F}}_{1}\to{\mathcal{F}}_{2} and β:𝒢𝒢\beta:{\mathcal{G}}\to{\mathcal{G}} be isomorphisms of sheaves of abelian groups. A 𝒢{\mathcal{G}}-valued pairing φ:12𝒢\varphi:{\mathcal{F}}_{1}\otimes{\mathcal{F}}_{2}\to{\mathcal{G}} is symmetric with respect to α\alpha and β\beta if we have for all local sections sis_{i} of i{\mathcal{F}}_{i} that

φ(s1s2)=β(φ(α1(s2)α(s1))).\varphi(s_{1}\otimes s_{2})=\beta(\varphi(\alpha^{-1}(s_{2})\otimes\alpha(s_{1}))).
Example 7.2.

Let =1=2{\mathcal{F}}={\mathcal{F}}_{1}={\mathcal{F}}_{2}. Then φ:12𝒢\varphi:{\mathcal{F}}_{1}\otimes{\mathcal{F}}_{2}\to{\mathcal{G}} is (skew-)symmetric if α:\alpha:{\mathcal{F}}\to{\mathcal{F}} is the identity id\textrm{id}_{\mathcal{F}} and β:𝒢𝒢\beta:{\mathcal{G}}\to{\mathcal{G}} is id𝒢\textrm{id}_{\mathcal{G}} (id𝒢-\textrm{id}_{\mathcal{G}}).

In order to define hermitian bilinear forms we have to set up some notation. Let AA be an {\mathbb{R}}-algebra and NN^{\prime} an AA-module. Further consider B=AB=A\otimes_{\mathbb{R}}{\mathbb{C}} and the BB-module N=NAB=NN=N^{\prime}\otimes_{A}B=N^{\prime}\otimes_{\mathbb{R}}{\mathbb{C}}. The complex conjugation on {\mathbb{C}} induces an automorphism σN:NN\sigma_{N}:N\to N of AA-modules whose fixed elements are exactly the elements of NN^{\prime}. If we have any BB-module MM, we can define another BB-module M¯\overline{M}. The elements of M¯\overline{M} are in bijection to those of MM and denoted by x¯\overline{x} for xMx\in M. The scalar multiplication is defined by letting bx¯:=σB(b)x¯b\cdot\overline{x}:=\overline{\sigma_{B}(b)\cdot x}. This implies that the map τM:MM¯,xx¯\tau_{M}:M\to\overline{M},\,x\mapsto\overline{x} is an isomorphism of AA-modules and we have M=M¯¯M=\overline{\overline{M}}.

These definitions carry over to the case of schemes by glueing. If XX is a scheme over {\mathbb{R}} we denote by X=X×Spec()Spec()X_{\mathbb{C}}=X\times_{\operatorname{Spec}({\mathbb{R}})}\operatorname{Spec}({\mathbb{C}}) the base change to {\mathbb{C}} and π:XX\pi:X_{\mathbb{C}}\to X the natural projection. If 𝒢=π𝒢{\mathcal{G}}=\pi^{*}{\mathcal{G}}^{\prime} for some quasi-coherent sheaf 𝒢{\mathcal{G}}^{\prime} on XX, we have an isomorphism σ𝒢:𝒢𝒢¯\sigma_{\mathcal{G}}:{\mathcal{G}}\to\overline{{\mathcal{G}}} of sheaves of abelian groups as above. For any quasi-coherent sheaf {\mathcal{F}} on XX_{\mathbb{C}} we obtain the quasi-coherent sheaf ¯\overline{{\mathcal{F}}} together with an isomorphism of sheaves of abelian groups τ:¯\tau_{\mathcal{F}}:{\mathcal{F}}\to\overline{{\mathcal{F}}}. We say that a 𝒢{\mathcal{G}}-valued pairing ¯𝒢{\mathcal{F}}\otimes\overline{{\mathcal{F}}}\to{\mathcal{G}} is hermitian if it is symmetric with respect to τ\tau_{\mathcal{F}} and σ𝒢\sigma_{\mathcal{G}}.

Remark 7.3.

Let XX be a scheme over {\mathbb{R}}. The complex conjugation induces an automorphism σ:𝒦X𝒦X\sigma:{\mathcal{K}}_{X_{\mathbb{C}}}\to{\mathcal{K}}_{X_{\mathbb{C}}} on the sheaf of total quotient rings of XX_{\mathbb{C}}. If {\mathcal{F}} is a subsheaf of 𝒦X{\mathcal{K}}_{X_{\mathbb{C}}}, then we can identify ¯\overline{{\mathcal{F}}} with the image of {\mathcal{F}} under σ\sigma.

8. Positive semidefinite bilinear forms

The criterion for a coherent sheaf being Ulrich that we have seen in Section 4 very much relies on the induced bilinear form on global sections being nondegenerate. Verifying this condition might be hard in general. However, in this section we show that when working over the real numbers we have a more convenient criterion at hand, namely positivity. In this section XX will always be a variety over {\mathbb{R}}.

Definition 8.1.

Let {\mathcal{F}} be a coherent sheaf on XX. Consider a symmetric bilinear form φ:𝒪X𝒪X\varphi:{\mathcal{F}}\otimes_{{\mathcal{O}}_{X}}{\mathcal{F}}\to{\mathcal{O}}_{X}. Let αX()\alpha\in X({\mathbb{R}}), i.e., a morphism α:Spec()X\alpha:\operatorname{Spec}({\mathbb{R}})\to X. Then we get a symmetric bilinear form αφ\alpha^{*}\varphi on the pullback α\alpha^{*}{\mathcal{F}} which is just a finite dimensional {\mathbb{R}}-vector space. We say that φ\varphi is positive semidefinite at αX()\alpha\in X({\mathbb{R}}) if αφ\alpha^{*}\varphi is positive semidefinite. We say that φ\varphi is positive semidefinite if it is positive semidefinite at every αX()\alpha\in X({\mathbb{R}}).

Analogously, let {\mathcal{F}} be a coherent sheaf on XX_{\mathbb{C}} and consider a hermitian bilinear form φ:𝒪X¯𝒪X\varphi:{\mathcal{F}}\otimes_{{\mathcal{O}}_{X}}\overline{{\mathcal{F}}}\to{\mathcal{O}}_{X}. We can consider any αX()\alpha\in X({\mathbb{R}}) also as a point of XX_{\mathbb{C}} that is fixed by the involution. Like this we obtain a hermitian bilinear form αφ\alpha^{*}\varphi on the pullback α\alpha^{*}{\mathcal{F}} which is just a finite dimensional {\mathbb{C}}-vector space. We say that φ\varphi is positive semidefinite at αX()\alpha\in X({\mathbb{R}}) if αφ\alpha^{*}\varphi is positive semidefinite. We say that φ\varphi is positive semidefinite if it is positive semidefinite at every αX()\alpha\in X({\mathbb{R}}).

A symmetric or hermitian bilinear form on a coherent sheaf induces a symmetric or hermitian bilinear form on the space of global sections. The next lemma shows how this behaves with respect to positivity.

Lemma 8.2.

Let XX be irreducible and proper over {\mathbb{R}} with generic point ξ\xi. Let φ:𝒪X𝒪X\varphi:{\mathcal{F}}\otimes_{{\mathcal{O}}_{X}}{\mathcal{F}}\to{\mathcal{O}}_{X} be a positive semidefinite symmetric bilinear form on the coherent sheaf {\mathcal{F}} and V=H0(X,)V=H^{0}(X,{\mathcal{F}}).

  1. a)

    If X()X({\mathbb{R}})\neq\emptyset, then the induced bilinear form V×VV\times V\to{\mathbb{R}} positive semidefinite.

  2. b)

    If X()X({\mathbb{R}}) is Zariski dense in XX, φ\varphi is nondegenerate at ξ\xi and {\mathcal{F}} is torsion-free, then the induced bilinear form V×VV\times V\to{\mathbb{R}} is positive definite and thus in particular nondegenerate.

The corresponding statements for hermitian bilinear forms hold true as well.

Proof.

For part a)a) we observe that if φ(s,s)=1\varphi(s,s)=-1 for some sVs\in V, then φ\varphi is not positive semidefinite at any point from X()X({\mathbb{R}}).

In order to show part b)b) we first observe that since X()X({\mathbb{R}}) is Zariski dense in XX, the field κ(ξ)\kappa(\xi) has an ordering PP. Consider a nonzero section sVs\in V. Since φ(s,s)0\varphi(s,s)\geq 0 at all points in X()X({\mathbb{R}}), Tarski’s principle implies that φ(s,s)\varphi(s,s) is also nonnegative with respect to the ordering PP when considered as an element of κ(ξ)\kappa(\xi). Thus the bilinear form induced by φ\varphi on the κ(ξ)\kappa(\xi)-vector subspace of ξ{\mathcal{F}}_{\xi} spanned by VV is also positive semidefinite (with respect to PP). But since φ\varphi is nondegenerate at ξ\xi, it is even positive definite (with respect to PP). Finally, because {\mathcal{F}} is torsion-free, the nonzero section ss is mapped to a nonzero element in the stalk ξ{\mathcal{F}}_{\xi} and therefore by positive definiteness φ(s,s)0\varphi(s,s)\neq 0. This shows the claim. ∎

Thus if we assume positive semidefiniteness, we only need to assure that our bilinear form is nondegenerate at the generic point rather than on global sections.

Theorem 8.3.

Let YY be a geometrically irreducible variety which is proper over {\mathbb{R}}. Let f:XYf:X\to Y be a finite surjective morphism where XX is an equidimensional variety over {\mathbb{R}} with X()X({\mathbb{R}}) Zariski dense in XX. Let X1,,XsX_{1},\ldots,X_{s} be the irreducible components of XX. Let {\mathcal{F}} be a coherent torsion-free sheaf on XX and let V=H0(X,)V=H^{0}(X,{\mathcal{F}}). Let φ:𝒪Xf!𝒪Y\varphi:{\mathcal{F}}\otimes_{{\mathcal{O}}_{X}}{\mathcal{F}}\to f^{!}{\mathcal{O}}_{Y} be a symmetric bilinear map which is nondegenerate at the generic point of each XiX_{i}. If the induced 𝒪Y{\mathcal{O}}_{Y}-valued bilinear form on ff_{*}{\mathcal{F}} is positive semidefinite, then the following are equivalent:

  1. (i)

    dimVi=1sdeg(f|Xi)rank(|Xi)\dim V\geq\sum_{i=1}^{s}\deg(f|_{X_{i}})\cdot\operatorname{rank}({\mathcal{F}}|_{X_{i}}).

  2. (ii)

    {\mathcal{F}} is an ff-Ulrich sheaf.

Proof.

Combining 8.2 and 3.8 it remains to show that the induced symmetric 𝒪Y{\mathcal{O}}_{Y}-valued bilinear form on ff_{*}{\mathcal{F}} is nondegenerate at the generic point of YY. But this follows from the assumption that φ\varphi is nondegenerate at the generic point of each XiX_{i} by 8.4 whose easy proof we leave as an exercise. ∎

Lemma 8.4.

Let A=K1××KrA=K_{1}\times\cdots\times K_{r} be the direct product of fields KiK_{i} each of which is a finite extension of the field KK. Let MM be a finitely generated AA-module. Then MV1××VrM\cong V_{1}\times\cdots\times V_{r} where each ViV_{i} is a finite dimensional KiK_{i}-vector space and the right-hand side is considered as an AA-module in the obvious way. A homomorphism φ:MHomK(M,K)\varphi:M\to\operatorname{Hom}_{K}(M,K) of AA-modules such that all induced maps ViHomK(Vi,K)V_{i}\to\operatorname{Hom}_{K}(V_{i},K) are isomorphisms is an isomorphism itself. In particular, the induced KK-bilinear form MKMKM\otimes_{K}M\to K is nondegenerate.

In the same manner we obtain the hermitian version.

Theorem 8.5.

Let YY be a geometrically irreducible variety which is proper over {\mathbb{R}}. Let f:XYf:X\to Y be a finite surjective morphism where XX is an equidimensional variety over {\mathbb{R}} with X()X({\mathbb{R}}) Zariski dense in XX. Let X1,,XsX_{1},\ldots,X_{s} be the irreducible components of XX_{\mathbb{C}}. Let {\mathcal{F}} be a coherent torsion-free sheaf on XX_{\mathbb{C}} and let V=H0(X,)V=H^{0}(X,{\mathcal{F}}). Let φ:𝒪X¯f!𝒪Y\varphi:{\mathcal{F}}\otimes_{{\mathcal{O}}_{X}}\overline{{\mathcal{F}}}\to f^{!}{\mathcal{O}}_{Y_{\mathbb{C}}} be a hermitian bilinear map which is nondegenerate at the generic point of each XiX_{i}. If the induced 𝒪Y{\mathcal{O}}_{Y_{\mathbb{C}}}-valued hermitian bilinear form on ff_{*}{\mathcal{F}} is positive semidefinite, then the following are equivalent:

  1. (i)

    dimVi=1sdeg(f|Xi)rank(|Xi)\dim V\geq\sum_{i=1}^{s}\deg(f|_{X_{i}})\cdot\operatorname{rank}({\mathcal{F}}|_{X_{i}}).

  2. (ii)

    {\mathcal{F}} is an ff-Ulrich sheaf.

Definition 8.6.

When the conditions, including those on X,YX,Y and ff, of 8.3 or 8.5 are satisfied we say that {\mathcal{F}} is a positive symmetric or hermitian ff-Ulrich sheaf, respectively.

In order to check the positivity condition the following lemma will be useful.

Lemma 8.7.

Let YY be an irreducible smooth variety with Y()Y({\mathbb{R}})\neq\emptyset and let XX be an equidimensional variety over {\mathbb{R}}. Let f:XYf:X\to Y be a finite surjective real fibered morphism. Let ss be a global section of 𝒦X{\mathcal{K}}_{X} which is nonnegative on X()X({\mathbb{R}}) and let 𝒟{\mathcal{D}} be the subsheaf of 𝒦X{\mathcal{K}}_{X} given by sΔ(X/Y)s\cdot\Delta(X/Y). Let {\mathcal{L}} be a subsheaf of 𝒦X{\mathcal{K}}_{X} such that 𝒟{\mathcal{L}}\cdot{\mathcal{L}}\subset{\mathcal{D}}. Then there is a f!𝒪Yf^{!}{\mathcal{O}}_{Y}-valued symmetric bilinear form on {\mathcal{L}} such that the induced 𝒪Y{\mathcal{O}}_{Y}-valued bilinear form on ff_{*}{\mathcal{L}} is positive semidefinite.

Proof.

By assumption we can define the symmetric Δ(X/Y)\Delta(X/Y)-valued bilinear form on {\mathcal{L}} that maps a pair of sections (g,h)(g,h) to ghs\frac{g\cdot h}{s}. Thus the induced 𝒪Y{\mathcal{O}}_{Y}-valued bilinear form maps (g,g)(g,g) to the trace of g2s\frac{g^{2}}{s} which is nonnegative by 6.7. ∎

Remark 8.8.

Let ι:Xn\iota:X\hookrightarrow\mathbb{P}^{n} be an embedding of a kk-dimensional projective variety and let {\mathcal{F}} be an Ulrich sheaf on XX, i.e. {\mathcal{F}} is π\pi-Ulrich for some finite surjective linear projection π:Xk\pi:X\to\mathbb{P}^{k}. One can show that

ι(om𝒪X(,π!𝒪k))xtnk(ι,ωn)(k+1).\iota_{*}(\operatorname{{\mathscr{H}om}}_{{\mathcal{O}}_{X}}({\mathcal{F}},\pi^{!}{\mathcal{O}}_{\mathbb{P}^{k}}))\cong{\mathcal{E}}xt^{n-k}(\iota_{*}{\mathcal{F}},\omega_{\mathbb{P}^{n}})(k+1).

Thus our notion of symmetry coincides with the one introduced in [ESW03, §3.1]. In particular, the property of being symmetric does not depend on the choice of the linear projection π\pi but the positivity condition does, as the next example shows.

Example 8.9.

Let Y=1Y=\mathbb{P}^{1} and X=𝒱(x02x12x22)2X={\mathcal{V}}(x_{0}^{2}-x_{1}^{2}-x_{2}^{2})\subset\mathbb{P}^{2}. We let fi:X1f_{i}:X\to\mathbb{P}^{1} be the linear projection with center eie_{i} where e1=[1:0:0]e_{1}=[1:0:0] and e2=[0:1:0]e_{2}=[0:1:0]. Note that f1f_{1} is real fibered but f2f_{2} is not. Let P=[1:1:0]XP=[1:1:0]\in X (Weil divisor on XX) and =(P){\mathcal{L}}={\mathcal{L}}(P) the corresponding invertible subsheaf of 𝒦X{\mathcal{K}}_{X}. A basis of the space of global sections VV of (P){\mathcal{L}}(P) is given by the two rational functions g1=1g_{1}=1 and g2=x0+x1x2g_{2}=\frac{x_{0}+x_{1}}{x_{2}}. The ramification divisor RiR_{i} of fif_{i} is given by R1=Q0+Q0¯R_{1}=Q_{0}+\overline{Q_{0}} and R2=Q1+Q2R_{2}=Q_{1}+Q_{2} with Q0=[0:1:i],Q1=[1:0:1] and Q2=[1:0:1].Q_{0}=[0:1:\textrm{i}],\,Q_{1}=[1:0:1]\textrm{ and }Q_{2}=[1:0:-1]. Denoting by Δi\Delta_{i} the codifferent sheaf associated to fif_{i}, we thus get that

(1) (a,b)abx0x1x0\displaystyle(a,b)\mapsto a\cdot b\cdot\frac{x_{0}-x_{1}}{x_{0}}

defines a symmetric bilinear form Δ1{\mathcal{L}}\otimes{\mathcal{L}}\to\Delta_{1}. Via the isomorphism Δ1f1!𝒪1\Delta_{1}\cong f_{1}^{!}{\mathcal{O}}_{\mathbb{P}^{1}} this induces an 𝒪1{\mathcal{O}}_{\mathbb{P}^{1}}-valued bilinear form on (f1)(f_{1})_{*}{\mathcal{L}}. With respect to the above basis of VV it is given by the matrix

(2002)\begin{pmatrix}2&0\\ 0&2\end{pmatrix}

which is positive definite. Thus {\mathcal{L}} is a positive symmetric f1f_{1}-Ulrich sheaf. On the other hand, via the isomorphism Δ1Δ2\Delta_{1}\to\Delta_{2} that is given by multiplication with x0x1\frac{x_{0}}{x_{1}} we get from (1) and Δ2f2!𝒪1\Delta_{2}\cong f_{2}^{!}{\mathcal{O}}_{\mathbb{P}^{1}} an 𝒪1{\mathcal{O}}_{\mathbb{P}^{1}}-valued bilinear form on (f2)(f_{2})_{*}{\mathcal{L}}. With respect to the above basis of VV it is given by the matrix

(2002)\begin{pmatrix}-2&0\\ 0&2\end{pmatrix}

which is not positive semidefinite. Thus although f1!𝒪1f2!𝒪1f_{1}^{!}{\mathcal{O}}_{\mathbb{P}^{1}}\cong f_{2}^{!}{\mathcal{O}}_{\mathbb{P}^{1}} as abstract line bundles, for checking the positivity condition we need to specify the morphism. Moreover, 10.2 will show that actually no nonzero symmetric f2!𝒪1f_{2}^{!}{\mathcal{O}}_{\mathbb{P}^{1}}-valued bilinear form on {\mathcal{L}} will induce a positive semidefinite bilinear form on (f2)(f_{2})_{*}{\mathcal{L}} since f2f_{2} is not real fibered.

Since later on we will focus on irreducible varieties, we close this section with an example for the reducible case. A systematic study of positive Ulrich sheaves on reducible hypersurfaces would be very interesting with regard to the so-called generalized Lax conjecture, see [Vin12, Con. 3.3]. We think it would be particularly beneficial to understand how the main result of [Kum17] fits into our context here.

Example 8.10.

Let l=x0x1l=x_{0}-x_{1} and h=x02(x12+x22+x32)h=x_{0}^{2}-(x_{1}^{2}+x_{2}^{2}+x_{3}^{2}). Let X=X1X2X=X_{1}\cup X_{2} where X1=𝒱(l)3X_{1}={\mathcal{V}}(l)\subset\mathbb{P}^{3} and X2=𝒱(h)3X_{2}={\mathcal{V}}(h)\subset\mathbb{P}^{3}. The linear projection

f:X2,[x0:x1:x2:x3][x1:x2:x3]f:X\to\mathbb{P}^{2},\,[x_{0}:x_{1}:x_{2}:x_{3}]\to[x_{1}:x_{2}:x_{3}]

with center e=[1:0:0:0]e=[1:0:0:0] is real fibered and we want to construct a positive symmetric ff-Ulrich sheaf on XX. The function field of 2\mathbb{P}^{2} is K=(x2x1,x3x1)K={\mathbb{R}}(\frac{x_{2}}{x_{1}},\frac{x_{3}}{x_{1}}) and the total quotient ring of XX is L=L1×L2L=L_{1}\times L_{2} where LiL_{i} is the function field of XiX_{i} for i=1,2i=1,2. We note that 𝒪X{\mathcal{O}}_{X} can be identified with the following subsheaf of 𝒦X{\mathcal{K}}_{X}:

𝒪X(U)={(a,b)𝒪X1(UX1)×𝒪X2(UX2):(ab)|X1X2U=0}{\mathcal{O}}_{X}(U)=\{(a,b)\in{\mathcal{O}}_{X_{1}}(U\cap X_{1})\times{\mathcal{O}}_{X_{2}}(U\cap X_{2}):\,(a-b)|_{X_{1}\cap X_{2}\cap U}=0\}

for UXU\subset X. We further define the subsheaf 𝒫{\mathcal{P}} of 𝒦X{\mathcal{K}}_{X} via

𝒫(U)={(a,b)𝒪X1(UX1)×𝒪X2(UX2):(a+b)|X1X2U=0}{\mathcal{P}}(U)=\{(a,b)\in{\mathcal{O}}_{X_{1}}(U\cap X_{1})\times{\mathcal{O}}_{X_{2}}(U\cap X_{2}):\,(a+b)|_{X_{1}\cap X_{2}\cap U}=0\}

for UXU\subset X. Note that 𝒫{\mathcal{P}} is a line bundle on XX and 𝒫𝒫=𝒪X{\mathcal{P}}\cdot{\mathcal{P}}={\mathcal{O}}_{X}. Finally let

(U)={(a,b)𝒪X1(UX1)×𝒪X2(UX2):{\mathcal{E}}(U)=\{(a,b)\in{\mathcal{O}}_{X_{1}}(U\cap X_{1})\times{\mathcal{O}}_{X_{2}}(U\cap X_{2}):
(ax3x2b)|X1X2UU2=0 and (a-\frac{x_{3}}{x_{2}}b)|_{X_{1}\cap X_{2}\cap U\cap U_{2}}=0\textrm{ and }
(x2x3ab)|X1X2UU3=0}.(\frac{x_{2}}{x_{3}}a-b)|_{X_{1}\cap X_{2}\cap U\cap U_{3}}=0\}.

The coherent sheaf {\mathcal{E}} is defined in such a way that 𝒫{\mathcal{E}}\cdot{\mathcal{E}}\subset{\mathcal{P}}.

Now let Vi2V_{i}\subset\mathbb{P}^{2} be the open affine set where xi0x_{i}\neq 0 and Ui=f1(Vi)U_{i}=f^{-1}(V_{i}) for i=1,2,3i=1,2,3. By 5.4 the codifferent sheaf Δ(X/2)\Delta(X/\mathbb{P}^{2}) is the subsheaf of 𝒦X{\mathcal{K}}_{X} generated by xi2De(lh)\frac{x_{i}^{2}}{\textrm{D}_{e}(l\cdot h)} on UiU_{i} for i=1,2,3i=1,2,3. As an element of L1×L2L_{1}\times L_{2} this is

(xi2x22+x32,xi22(x12+x22+x32x0x1))\left(-\frac{x_{i}^{2}}{x_{2}^{2}+x_{3}^{2}},\,\frac{x_{i}^{2}}{2\cdot(x_{1}^{2}+x_{2}^{2}+x_{3}^{2}-x_{0}x_{1})}\right)

on UiU_{i} for i=1,2,3i=1,2,3. Let {\mathcal{L}} be the subsheaf of 𝒦X{\mathcal{K}}_{X} generated by (xix1,xix1)(\frac{x_{i}}{x_{1}},\frac{x_{i}}{x_{1}}) on UiU_{i}. Consider

s=(x22+x32x12,2(x12+x22+x32x0x1)x12)s=\left(\frac{x_{2}^{2}+x_{3}^{2}}{x_{1}^{2}},\,\frac{2\cdot(x_{1}^{2}+x_{2}^{2}+x_{3}^{2}-x_{0}x_{1})}{x_{1}^{2}}\right)

which is nonnegative on X()X({\mathbb{R}}). We have s𝒫Δ(X/2){\mathcal{L}}\cdot{\mathcal{L}}\subset s\cdot{\mathcal{P}}\cdot\Delta(X/\mathbb{P}^{2}). Thus we have sΔ(X/2){\mathcal{F}}\cdot{\mathcal{F}}\subset s\cdot\Delta(X/\mathbb{P}^{2}) where ={\mathcal{F}}={\mathcal{E}}\cdot{\mathcal{L}}. This gives us a f!𝒪2f^{!}{\mathcal{O}}_{\mathbb{P}^{2}}-valued symmetric bilinear form on {\mathcal{L}} such that the induced 𝒪2{\mathcal{O}}_{\mathbb{P}^{2}}-valued bilinear form on ff_{*}{\mathcal{F}} is positive semidefinite by 8.7. Here are linearly independent global sections of {\mathcal{F}}:

(0,x0x1),(x2,x3),(x3,x2).(0,x_{0}-x_{1}),\,(x_{2},-x_{3}),\,(x_{3},x_{2}).

Thus {\mathcal{F}} is a positive symmetric ff-Ulrich sheaf.

9. Ulrich sheaves, determinantal representations and sums of squares

The main reason why we are interested in Ulrich sheaves is because they correspond to certain determinantal representations. For the applications we consider in this article, we are interested in the following situation. Let S=[y,x1,,xn]S={\mathbb{R}}[y,x_{1},\ldots,x_{n}] be the polynomial ring with the grading determined by letting deg(y)=e\deg(y)=e and deg(xi)=1\deg(x_{i})=1 for i=1,,ni=1,\ldots,n, let hSdeh\in S_{de} be an irreducible homogeneous element of degree ded\cdot e and X=𝒱(h)(e,1,,1)X={\mathcal{V}}(h)\subset\mathbb{P}(e,1,\ldots,1) the hypersurface in the weighted projective space corresponding to SS. Assume h(1,0,,0)=1h(1,0,\cdots,0)=1. The following proposition and the subsequent remark are well known among experts. But since we are not aware of a reference for the precise statement that we need, we include a proof for the sake of completeness. The case e=1e=1 is treated for example in [Bea00, Thm. A] and the proof for the general case works essentially in the same way.

Proposition 9.1.

Let f:Xn1f:X\to\mathbb{P}^{n-1} be the projection on the last coordinates and let {\mathcal{F}} be an ff-Ulrich sheaf on XX with rank()=r\operatorname{rank}({\mathcal{F}})=r. Then there is a square matrix AA of size drd\cdot r whose entries are homogeneous polynomials in the xix_{i} of degree ee such that hr=det(yIA)h^{r}=\det(y\cdot I-A). If {\mathcal{F}} is a positive symmetric (resp. hermitian) ff-Ulrich sheaf, then AA can be chosen to be symmetric (resp. hermitian).

Proof.

Let 𝒪X(1){\mathcal{O}}_{X}(1) be the pullback of 𝒪n1(1){\mathcal{O}}_{\mathbb{P}^{n-1}}(1) via ff and let M=iH0(X,(i))M=\oplus_{i\in{\mathbb{Z}}}H^{0}(X,{\mathcal{F}}(i)) be the module of twisted global sections. Since {\mathcal{F}} is an ff-Ulrich sheaf, we have that MM considered as a module over R=[x1,,xn]SR={\mathbb{R}}[x_{1},\ldots,x_{n}]\subset S is isomorphic to RdrR^{d\cdot r}. Multiplication with yy is an RR-linear map, homogeneous of degree ee, that can thus be represented by a square matrix AA of size drd\cdot r whose entries are real homogeneous polynomials in the xix_{i} of degree ee and whose minimal polynomial is hh. Thus hr=det(yIA)h^{r}=\det(y\cdot I-A). Further, a symmetric positive semidefinite bilinear form as in Theorem 8.3 yields a homomorphism of graded SS-modules (of degree zero)

ψ:MHomR(M,R)\psi:M\to\operatorname{Hom}_{R}(M,R)

which has the property that the induced RR-bilinear form

Rdr×Rdr\displaystyle R^{d\cdot r}\times R^{d\cdot r} R\displaystyle\to R
(a,b)\displaystyle(a,b) (ψ(a))(b)\displaystyle\mapsto(\psi(a))(b)

is symmetric, i.e., (ψ(a))(b)=(ψ(b))(a)(\psi(a))(b)=(\psi(b))(a). The degree zero part VV of MRRdrM\cong_{R}R^{d\cdot r} is the space of global sections of {\mathcal{F}} and the restriction of this symmetric bilinear form to VV is thus positive definite by 8.2. We can therefore choose an orthonormal basis of VV with respect to this bilinear form. Note that this will also be a basis of the RR-module RdrR^{d\cdot r} that is orthonormal with respect to ψ\psi. Since ψ\psi is a homomorphism of SS-modules, we have (ψ(a))(yb)=(ψ(ya))(b)(\psi(a))(y\cdot b)=(\psi(y\cdot a))(b), so multiplication with yy is self-adjoint with respect to our above defined symmetric bilinear form. Thus we can choose the representing matrix AA of the RR-linear map given by multiplication with yy to be symmetric. The hermitian case follows analogously. ∎

Remark 9.2.

The converse of 9.1 is also true: If hr=det(yIA)h^{r}=\det(y\cdot I-A), then the cokernel MM of the map SdrSdrS^{d\cdot r}\to S^{d\cdot r} given by yIAy\cdot I-A is supported on XX and MM considered as RR-module is just RdrR^{d\cdot r}. If AA is symmetric, then an isomorphism of MM with HomR(M,R)\operatorname{Hom}_{R}(M,R) as SS-modules is given by sending the standard basis of RdrR^{d\cdot r} to its dual basis. We argue analogously in the hermitian case.

A refined statement is true for possibly reducible subvarieties XnX\subset\mathbb{P}^{n} that are not necessarily hypersurfaces, see [KS20a, Thm. 5.7]. These correspond to so-called determinantal representations of Livsic-type introduced in [SV18] which are closely related to determinantal representations of the Chow form of XX, see also [ESW03]. The main applications of this article however concern irreducible varieties only.

Example 9.3.

The positive symmetric ff-Ulrich sheaf on the reducible cubic hypersurface X=𝒱((x0x1)(x02(x12+x22+x32))3X={\mathcal{V}}((x_{0}-x_{1})\cdot(x_{0}^{2}-(x_{1}^{2}+x_{2}^{2}+x_{3}^{2}))\subset\mathbb{P}^{3} from 8.10 gives the following symmetric definite determinantal representation

(x0+x1x3x2x3x0x10x20x0x1).\begin{pmatrix}x_{0}+x_{1}&x_{3}&-x_{2}\\ x_{3}&x_{0}-x_{1}&0\\ -x_{2}&0&x_{0}-x_{1}\end{pmatrix}.

Now we consider again the situation of 6.2 when hh is of the form y2p(x)y^{2}-p(x) where p[x1,,xn]2ep\in{\mathbb{R}}[x_{1},\ldots,x_{n}]_{2e} is a globally nonnegative polynomial.

Lemma 9.4.

Let p[x1,,xn]2ep\in{\mathbb{R}}[x_{1},\ldots,x_{n}]_{2e} be a globally nonnegative polynomial which is not a square. Let h=y2p(x)h=y^{2}-p(x) where yy has degree ee. If hr=det(yIA)h^{r}=\det(y\cdot I-A) for some r1r\geq 1 and a symmetric or hermitian matrix AA of size 2r2r whose entries are homogeneous of degree ee, then pp is a sum of 2r2r squares in the symmetric case and a sum of 4r14r-1 squares in the hermitian case.

Proof.

Since pp is not a square, we find that hh is irreducible. Thus hr=det(yIA)h^{r}=\det(y\cdot I-A) implies that hh is the minimal polynomial of AA, i.e., A2=pIA^{2}=p\cdot I. Letting aia_{i} be the iith column of AA we then have p=aitaip=a_{i}^{t}a_{i} in the symmetric case and p=aitai¯p=a_{i}^{t}\overline{a_{i}} in the hermitian case. Thus pp is a sum of 2r2r squares in the symmetric case and a sum of 4r14r-1 squares in the hermitian case. ∎

A converse of this last conclusion is given by the following lemma.

Lemma 9.5.

Let P=G12++Gn2P=G_{1}^{2}+\ldots+G_{n}^{2} where the GiAG_{i}\in A are elements of some commutative ring AA. There is a symmetric square matrix QQ of some size mm\in{\mathbb{N}} whose entries are {\mathbb{Z}}-linear combinations of the GiG_{i} such that PIm=Q2P\cdot I_{m}=Q^{2}.

Proof.

This follows from the basic properties of Clifford algebras. Recall that the Clifford algebra Cl0,n()Cl_{0,n}({\mathbb{R}}) is generated by e1,,ene_{1},\ldots,e_{n} satisfying eiej=ejeie_{i}\cdot e_{j}=-e_{j}\cdot e_{i} if iji\neq j and ei2=1e_{i}^{2}=-1. In particular, we have

x12++xn2=(x1e1++xnen)(x1e1++xnen)x_{1}^{2}+\ldots+x_{n}^{2}=-(x_{1}e_{1}+\ldots+x_{n}e_{n})\cdot(x_{1}e_{1}+\ldots+x_{n}e_{n})

for all xkx_{k}\in{\mathbb{R}}. For 1in1\leq i\leq n let AiA_{i} be the representing matrix of the map

Cl0,n()Cl0,n(),aeiaCl_{0,n}({\mathbb{R}})\to Cl_{0,n}({\mathbb{R}}),\,a\mapsto e_{i}\cdot a

with respect to the basis ei1eire_{i_{1}}\cdots e_{i_{r}} of Cl0,n()Cl_{0,n}({\mathbb{R}}) with 1i1<<irn1\leq i_{1}<\cdots<i_{r}\leq n and r0r\geq 0. Then one immediately verifies that AiA_{i} is a matrix having only entries in {0,±1}\{0,\pm 1\} satisfying AiT=AiA_{i}^{T}=-A_{i}. It follows that

(x12++xn2)IN=SSt(x_{1}^{2}+\ldots+x_{n}^{2})\cdot I_{N}=S\cdot S^{t}

for all xkx_{k}\in{\mathbb{R}} where S=x1A1++xnAnS=x_{1}A_{1}+\ldots+x_{n}A_{n} and NN is the dimension of Cl0,n()Cl_{0,n}({\mathbb{R}}). Now we can choose

Q=(0SSt0)Q=\begin{pmatrix}0&S\\ S^{t}&0\end{pmatrix}

and we get (x12++xn2)IN=Q2(x_{1}^{2}+\ldots+x_{n}^{2})\cdot I_{N}=Q^{2} for all xkx_{k}\in{\mathbb{R}}. Since the entries of the AiA_{i} are integers, the identity holds over every commutative ring. ∎

Putting all this together we get the following connection of sums of squares to Ulrich sheaves.

Theorem 9.6.

Let p[x1,,xn]2ep\in{\mathbb{R}}[x_{1},\ldots,x_{n}]_{2e} be a homogeneous polynomial of degree 2e2e which is not a square. Inside the weighted projective space (e,1,,1)\mathbb{P}(e,1,\ldots,1) we consider the hypersurface XX defined by y2=p(x0,,xn)y^{2}=p(x_{0},\ldots,x_{n}) and the natural projection π:Xn\pi:X\to\mathbb{P}^{n} onto the xx-coordinates. Then pp is a sum of squares of polynomials if and only if there is a positive symmetric (or hermitian) π\pi-Ulrich sheaf {\mathcal{F}} on XX. In that case, if rank()=r\operatorname{rank}({\mathcal{F}})=r then pp is a sum of 2r2r squares in the symmetric case and a sum of 4r14r-1 squares in the hermitian case.

Proof.

First assume that pp is a sum of squares. By 9.5 there is a symmetric square matrix AA of some size mm\in{\mathbb{N}} whose entries are homogeneous of degree ee in the variables x1,,xnx_{1},\ldots,x_{n} such that pIm=A2p\cdot I_{m}=A^{2}. Because pp is not a square, the polynomial h=y2ph=y^{2}-p is irreducible. Thus A2pIm=0A^{2}-p\cdot I_{m}=0 shows that hh is the minimal polynomial of AA. This implies that hr=det(yIA)h^{r}=\det(y\cdot I-A) for some r>0r>0. Thus there is a positive symmetric π\pi-Ulrich sheaf on XX by 9.2.

Now assume that there is a positive symmetric (or hermitian) π\pi-Ulrich sheaf {\mathcal{F}} on XX. Then by 9.1 there is a symmetric (resp. hermitian) matrix AA of size 2r2\cdot r whose entries are homogeneous polynomials in the xix_{i} of degree ee such that hr=det(yIA)h^{r}=\det(y\cdot I-A). Now the claim follows from 9.4. ∎

As a special case we get the following result by Netzer and Thom [NT12].

Corollary 9.7.

Let h[x0,,xn]2h\in{\mathbb{R}}[x_{0},\ldots,x_{n}]_{2} be a quadratic hyperbolic polynomial. Then hrh^{r} has a definite symmetric determinantal representation for some r>0r>0.

Proof.

If X=𝒱(h)X={\mathcal{V}}(h) is hyperbolic with respect to ene\in\mathbb{P}^{n}, then the linear projection πe:Xn1\pi_{e}:X\to\mathbb{P}^{n-1} is a real fibered double cover ramified along the zero set of a nonnegative quadratic polynomial pp. Since pp is a sum of squares, there is a positive symmetric πe\pi_{e}-Ulrich sheaf on XX by 9.6. 9.1 implies the claim. ∎

Remark 9.8.

Positive Ulrich sheaves on reciprocal linear spaces, i.e. the closure of the image of a linear space under coordinatewise inversion, were used in [KV19] to prove that a certain polynomial associated to a hyperplane arrangement, called the entropic discriminant, is a sum of squares. The relation of Ulrich sheaves and sums of squares used in [KV19] is a generalization of one direction of 9.6. Namely let f:Xnf:X\to\mathbb{P}^{n} be a finite surjective real fibered morphism such that f𝒪Xf_{*}{\mathcal{O}}_{X} is a sum of line bundles. Then, if there is a positive ff-Ulrich sheaf, then the polynomial defining the branch locus of ff is a sum of squares [KV19, Thm. 6.1].

10. Positive Ulrich sheaves of rank one on irreducible varieties

In this section let f:XYf:X\to Y always denote a finite surjective morphism of geometrically irreducible varieties which are proper over {\mathbb{R}}. We assume that YY is smooth and has a real point. We further assume that the singular locus of XX has codimension at least two. This allows us to speak about Weil divisors on XX. For a given Weil divisor DD on XX we denote by (D){\mathcal{L}}(D) the subsheaf of 𝒦X{\mathcal{K}}_{X} consisting of all rational functions with pole and zero orders prescribed by DD. We further denote (D)=dim(Γ(X,(D)))\ell(D)=\dim(\Gamma(X,{\mathcal{L}}(D))). Let ZYZ\subset Y be a closed subset of codimension at least two such that f1(Z)f^{-1}(Z) contains XsingX_{\textnormal{sing}}. Letting V=YZV=Y\setminus Z and U=f1(V)U=f^{-1}(V), the restriction f|U:UVf|_{U}:U\to V is a finite surjective morphism of smooth irreducible varieties. Thus Δ(U/V)\Delta(U/V) is invertible by 5.8 and thus corresponds to a Weil divisor RR on UU. Since XUX\setminus U has codimension at least two, we can also consider RR as a Weil divisor on XX which we call the ramification divisor of ff. 5.10 shows that the associated subsheaf (R){\mathcal{L}}(R) of 𝒦X{\mathcal{K}}_{X} is exactly Δ(X/Y)\Delta(X/Y). All this holds true for the complexification XX_{\mathbb{C}} as well. Complex conjugation gives an involution σ:XX\sigma:X_{\mathbb{C}}\to X_{\mathbb{C}} and for a Weil divisor DD on XX_{\mathbb{C}} we denote σ(D)\sigma(D) by D¯\overline{D}. Using the notation of Section 7 we have (D¯)=(D)¯{\mathcal{L}}(\overline{D})=\overline{{\mathcal{L}}(D)} as subsheaves of 𝒦X{\mathcal{K}}_{X_{\mathbb{C}}}.

Remark 10.1.

If D1D_{1} and D2D_{2} are Weil divisors on XX, then we have

(D1)(D2)(D1+D2){\mathcal{L}}(D_{1})\cdot{\mathcal{L}}(D_{2})\subset{\mathcal{L}}(D_{1}+D_{2})

considered as subsheaves of 𝒦X{\mathcal{K}}_{X}. But since (Di){\mathcal{L}}(D_{i}) is not necessarily an invertible sheaf, we do not have equality in general.

The relative notion of positive semidefiniteness introduced in Section 8 relates to the notion of being real fibered in the following way.

Proposition 10.2 (Thm. 5.11 in [KS20a]).

The following are equivalent:

  1. (i)

    There is a coherent sheaf {\mathcal{F}} on XX with Supp()=X\operatorname{Supp}({\mathcal{F}})=X and a symmetric nonzero f!𝒪Yf^{!}{\mathcal{O}}_{Y}-valued bilinear form on {\mathcal{F}} such that the induced bilinear form on ff_{*}{\mathcal{F}} is positive semidefinite,

  2. (ii)

    ff is real fibered, i.e., f1(Y())=X()f^{-1}(Y({\mathbb{R}}))=X({\mathbb{R}}).

In this situation we get the following convenient criterion for a Weil divisor to give rise to a positive Ulrich sheaf.

Theorem 10.3.

Let f:XYf:X\to Y be a real fibered finite surjective morphism of geometrically irreducible varieties which are proper over {\mathbb{R}}. Let YY be smooth and have a real point. Further let the singular locus of XX have codimension at least two. Let RR be the ramification divisor of ff and let ss be a rational function on XX which is nonnegative on X()X({\mathbb{R}}).

  1. (1)

    If DD is a Weil divisor on XX such that 2D+(s)=R2D+(s)=R and (D)deg(f)\ell(D)\geq\deg(f), then the sheaf (D){\mathcal{L}}(D) on XX is a positive symmetric ff-Ulrich sheaf.

  2. (2)

    If DD is a Weil divisor on XX_{\mathbb{C}} that satisfies D+D¯+(s)=RD+\overline{D}+(s)=R and (D)deg(f)\ell(D)\geq\deg(f), then the sheaf (D){\mathcal{L}}(D) on XX_{\mathbb{C}} is a positive hermitian ff-Ulrich sheaf.

Proof.

Let =(D){\mathcal{L}}={\mathcal{L}}(D) be the subsheaf of 𝒦X{\mathcal{K}}_{X} corresponding to DD. By assumption we can define the symmetric resp. hermitian Δ(X/Y)\Delta(X/Y)-valued bilinear form on {\mathcal{L}} that maps a pair of sections (g,h)(g,h) to the product sghs\cdot g\cdot h. This is clearly nondegenerate at the generic point of XX and it is positive semidefinite by 6.7. Then the claim follows from 8.3 and 8.5 respectively. ∎

Remark 10.4.

Let DD be a Weil divisor on XX or XX_{\mathbb{C}} such that 2D2D or D+D¯D+\overline{D}, respectively, is linearly equivalent to RR, i.e., they differ only by a principal divisor (g)(g). The signs that gg takes on real points of XX (up to global scaling) do only depend on the divisor class of DD. Indeed, if D=D+(f)D^{\prime}=D+(f) for some rational function ff, then 2D2D^{\prime} resp. D+D¯D+\overline{D} differs from RR by gf2g\cdot f^{2} or gff¯g\cdot f\overline{f} respectively.

Example 10.5.

Let LnL\subset\mathbb{P}^{n} be a linear subspace of dimension d<nd<n that is not contained in any coordinate hyperplane. We denote by L1L^{-1} its reciprocal, i.e. the (Zariski closure of the) image of LL under the rational map nn\mathbb{P}^{n}\dashrightarrow\mathbb{P}^{n} defined by coordinatewise inversion. It was shown by Varchenko [Var95] that L1L^{-1} is hyperbolic with respect to the orthogonal complement LL^{\perp} of LL. Further it was shown in [KV19] that there is a symmetric positive ff-Ulrich sheaf of rank one on L1L^{-1} where f:Xdf:X\to\mathbb{P}^{d} is the linear projection from LL^{\perp}. We want to outline how this also follows from 10.3, at least for generic LL. To this end let LL be the row span of a matrix A=(aij)A=(a_{ij}) of size (d+1)×(n+1)(d+1)\times(n+1) and assume that every maximal minor of AA is nonzero. Letting lj=i=1d+1aijxil_{j}=\sum_{i=1}^{d+1}a_{ij}x_{i} for j=1,,d+1j=1,\ldots,d+1 we can describe L1L^{-1} as the image of the rational map

ψ:dn,x(l1ld+1l1::l1ld+1ld+1).\psi:\mathbb{P}^{d}\dashrightarrow\mathbb{P}^{n},\,x\mapsto(\frac{l_{1}\cdots l_{d+1}}{l_{1}}:\cdots:\frac{l_{1}\cdots l_{d+1}}{l_{d+1}}).

Note that ψ\psi is defined in all points where at most one of the ljl_{j} vanishes. It follows from the proof of [SSV13, Cor. 5] and [SSV13, Rem. 31] that the ramification divisor RR of ff on L1L^{-1} is the proper transform under ψ\psi of the zero set ZdZ\subset\mathbb{P}^{d} of

P=Idet(AI)2jIlj2P=\sum_{I}\det(A_{I})^{2}\prod_{j\in I}l_{j}^{2}

where the sum is taken over all I{1,,n+1}I\subset\{1,\ldots,n+1\} of size ndn-d and AIA_{I} denotes the submatrix of AA obtained from erasing all columns indexed by II. Let HdH\subset\mathbb{P}^{d} be a generic hyperplane defined by a linear form GG and let DD be the divisor on L1L^{-1} defined as the proper transform of HH under ψ\psi. On d\mathbb{P}^{d} we have 2(nd)H=Z+(PG2(nd))2(n-d)H=Z+\left(\frac{P}{G^{2(n-d)}}\right) and our genericity assumption on AA implies that PP does not vanish entirely on any of the 𝒱(li,lj)d{\mathcal{V}}(l_{i},l_{j})\subset\mathbb{P}^{d} which comprise the locus where ψ\psi is not regular. Therefore, we have 2(nd)D=R+(PG2(nd))2(n-d)D=R+\left(\frac{P}{G^{2(n-d)}}\right) as divisors on L1L^{-1}. Clearly, the rational function PG2(nd)\frac{P}{G^{2(n-d)}} is nonnegative and ((nd)D)\ell((n-d)\cdot D) equals (nd)\binom{n}{d}, the number of monomials of degree ndn-d in d+1d+1 variables. Since this is also the degree of L1L^{-1} [PS06], 10.3(1) implies that ((nd)D){\mathcal{L}}((n-d)\cdot D) is a positive symmetric ff-Ulrich sheaf on L1L^{-1}.

The previous example leads to the following question. Let XnX\subset\mathbb{P}^{n} be an irreducible variety not contained in any coordinate hyperplane which is hyperbolic with respect to every linear subspace of codimension dim(X)+1\dim(X)+1 all of whose Plücker coordinates are positive. Then the image X1X^{-1} of XX under coordinatewise inversion is hyperbolic with respect to all these subspaces as well [KV19, Prop. 1.4].

Problem 10.6.

Given a symmetric positive Ulrich bundle on a variety XnX\subset\mathbb{P}^{n} as above. Does there exist one on X1X^{-1} of the same rank as well?

Remark 10.7.

Using [Kum13, Prop. 3.3.8] one can show that the answer to 10.6 is yes for hypersurfaces. It is also true for XX a linear subspace by [KV19].

In the case of hypersurfaces, 10.3 has a geometric interpretation in terms of so-called interlacers.

Definition 10.8.

Let g,h[x0,,xn]g,h\in{\mathbb{R}}[x_{0},\ldots,x_{n}] with d=deg(g)+1=deg(h)d=\deg(g)+1=\deg(h) be hyperbolic with respect to ee. If for all vn+1v\in{\mathbb{R}}^{n+1} we have that

a1b1a2bd1ada_{1}\leq b_{1}\leq a_{2}\leq\cdots\leq b_{d-1}\leq a_{d}

where the aia_{i} and bib_{i} are the zeros of h(tev)h(te-v) and g(tev)g(te-v) respectively, we say that gg interlaces hh, or that gg is an interlacer of hh. This definition carries over to hyperbolic hypersurfaces in the obvious way.

Refer to caption
Figure 2. A cubic hyperbolic plane curve (blue) interlaced by a plane hyperbolic conic (red).
Example 10.9.

If h[x0,,xn]h\in{\mathbb{R}}[x_{0},\ldots,x_{n}] is hyperbolic with respect to ee, then the directional derivative Deh\textrm{D}_{e}h of hh in direction ee is an interlacer of hh. This follows from Rolle’s theorem.

Corollary 10.10.

Let h[x0,,xn]dh\in{\mathbb{R}}[x_{0},\ldots,x_{n}]_{d} be hyperbolic with respect to ee and let X=𝒱(h)nX={\mathcal{V}}(h)\subset\mathbb{P}^{n} be the corresponding hypersurface. Assume that the singular locus of XX has dimension at most n3n-3. Let gg be an interlacer of hh and denote by GG the Weil divisor it defines on XX.

  1. (1)

    Assume that G=2DG=2D for some Weil divisor DD on XX. If the vector space of all p[x0,,xn]d1p\in{\mathbb{R}}[x_{0},\ldots,x_{n}]_{d-1} that vanish on DD has dimension at least dd, then hh has a definite symmetric determinantal representation.

  2. (2)

    Assume that G=D+D¯G=D+\overline{D} for some Weil divisor DD on XX_{\mathbb{C}}. If the vector space of all p[x0,,xn]d1p\in{\mathbb{C}}[x_{0},\ldots,x_{n}]_{d-1} that vanish on DD has dimension at least dd, then hh has a definite hermitian determinantal representation.

Proof.

The ramification divisor of the linear projection with center ee is the zero set of the directional derivative Deh\textrm{D}_{e}h on XX. Since gg is an interlacer, we have that the rational function Dehg\frac{\textrm{D}_{e}h}{g} is nonnegative on X()X({\mathbb{R}}) by [KPV15, Lem. 2.4]. Then the claim follows from 10.3 and 9.1 by our dimensional assumption. ∎

Remark 10.11.

When n=2n=2, the case of plane curves, the dimensional condition in 10.10 is automatically satisfied by Riemann–Roch. If g=Dehg=\textrm{D}_{e}h, then G=D+D¯G=D+\overline{D} by 6.6. In that case this recovers the result from [PV13, §4]. Note that the proof in [PV13] does not seem to generalize to higher dimensions as it uses Max Noether’s AF+BG theorem. It would be interesting to know when we actually need the dimensional condition in the case of higher dimensional hypersurfaces.

Problem 10.12.

Find hyperbolic hypersurfaces XnX\subset\mathbb{P}^{n} for which there is an interlacer whose zero divisor on XX is of the form D+D¯D+\overline{D} with (D)<deg(X)\ell(D)<\deg(X).

Example 10.13.

Let h[x0,,xn]nh\in{\mathbb{R}}[x_{0},\ldots,x_{n}]_{n} be the elementary symmetric polynomial of degree nn in x0,,xnx_{0},\ldots,x_{n}. It is hyperbolic with respect to every point in the positive orthant and g=x0hg=\frac{\partial}{\partial x_{0}}h is an interlacer. The zero divisor of gg on the hypersurface X=𝒱(h)nX={\mathcal{V}}(h)\subset\mathbb{P}^{n} is of the form 2D2D where D=1i<jnLijD=\sum_{1\leq i<j\leq n}L_{ij}. Here LijL_{ij} is the linear subspace 𝒱(xi,xj)X{\mathcal{V}}(x_{i},x_{j})\subset X. Each monomial x1xnxi\frac{x_{1}\cdots x_{n}}{x_{i}} for 1in1\leq i\leq n vanishes on DD. Thus by 10.10 the polynomial hh has a definite symmetric determinantal representation. This was known before and follows e.g. from the more general result that the bases generating polynomial of a regular matroid (in our case Un,n+1U_{n,n+1}) has a definite symmetric determinantal representation, see [COSW04, §8.2].

11. Smooth curves

In this section let f:XYf:X\to Y be a real fibered morphism between smooth irreducible curves that are projective over {\mathbb{R}} with Y()Y({\mathbb{R}}) Zariski dense in YY. Let RR be the ramification divisor of ff. We want to apply 10.3.

Lemma 11.1.

There is a divisor MM on XX and a nonnegative ss in the function field of XX such that R+(s)=2MR+(s)=2M.

Proof.

The proof is a projective version of the proof of [Han17, Cor. 4.2]. By 6.6 we have that ff is unramified at real points. Thus RR, considered as a Weil divisor, is a sum of nonreal points. Therefore, the Weil divisor RR^{\prime} that we obtain on the complexification X=X×X_{\mathbb{C}}=X\times_{\mathbb{R}}{\mathbb{C}} of XX is of the form R=Q+Q¯R^{\prime}=Q+\overline{Q} where QQ is some effective divisor and Q¯\overline{Q} its complex conjugate. Since the group Pic0(X)\textnormal{Pic}^{0}(X_{\mathbb{C}}) is divisible, there is a gg in the function field of XX_{\mathbb{C}} and a divisor NN on XX_{\mathbb{C}} such that QnP=2N+(g)Q-nP=2N+(g) where n=deg(Q)n=\deg(Q) and PP is any point on XX_{\mathbb{C}} with P=P¯P=\overline{P}. Thus (QnP)+(QnP¯)=2(N+N¯)+(gg¯)(Q-nP)+(\overline{Q-nP})=2(N+\overline{N})+(g\cdot\overline{g}) which implies

R=Q+Q¯=2(N+N¯+nP)+(gg¯).R^{\prime}=Q+\overline{Q}=2(N+\overline{N}+nP)+(g\cdot\overline{g}).

Since N+N¯+nPN+\overline{N}+nP is fixed by conjugation, it descends to a divisor MM on XX. The function s=gg¯s=g\cdot\overline{g} is a sum of two squares and thus nonnegative. ∎

From this we get the following theorem.

Theorem 11.2.

Let XX be a smooth irreducible curve that is projective over {\mathbb{R}}. For every real fibered f:X1f:X\to\mathbb{P}^{1} there is a positive symmetric ff-Ulrich line bundle.

Proof.

Let MM be the divisor from 11.1. By 10.3 we have to show that (M)deg(f)\ell(M)\geq\deg(f). By Hurwitz’s Theorem we have that 2g2=deg(R)2deg(f)2g-2=\deg(R)-2\deg(f) where gg is the genus of XX. Thus deg(M)=deg(f)+g1\deg(M)=\deg(f)+g-1 and by Riemann–Roch

(M)deg(M)g+1=deg(f).\ell(M)\geq\deg(M)-g+1=\deg(f).\qed
Corollary 11.3 (Thm. 7.2 in [SV18]).

The Chow form of every hyperbolic curve XnX\subset\mathbb{P}^{n} has a symmetric and definite determinantal representation.

Proof.

By [KS20a, Thm. 5.7] and [KS20a, Rem. 4.4] it suffices to show that there is a positive Ulrich bundle of rank 11 on XX. But this follows from the preceding theorem applied to the linear projection from an n2n-2-space of hyperbolicity if XX is smooth. Otherwise we can pass to the normalization of XX. ∎

Corollary 11.4 (Helton–Vinnikov Theorem [HV07]).

Every hyperbolic polynomial in three variables has a definite determinantal representation.

Example 11.5.

If the target is not 1\mathbb{P}^{1} as in 11.2, then there is in general no (positive symmetric) ff-Ulrich sheaf on XX. This fails already in the next easiest case, namely when XX and YY both have genus one. Indeed, let f:XYf:X\to Y be an unramified and real fibered double cover of an elliptic curve YY. Such maps exists, see for example [KS20b, Lem. 6.5], and by Riemann–Hurwitz XX is an elliptic curve as well. We claim that in this case there is actually no ff-Ulrich sheaf at all. Indeed, there is a line bundle {\mathcal{L}} on YY which is nontrivial and 22-torsion such that f𝒪X=𝒪Yf_{*}{\mathcal{O}}_{X}={\mathcal{O}}_{Y}\oplus{\mathcal{L}}. Then we have f=𝒪Xf^{*}{\mathcal{L}}={\mathcal{O}}_{X} and the projection formula implies that f=ff_{*}{\mathcal{F}}={\mathcal{L}}\otimes f_{*}{\mathcal{F}} for all coherent sheaves {\mathcal{F}} on XX. This excludes f=𝒪Yrf_{*}{\mathcal{F}}={\mathcal{O}}_{Y}^{r}.

12. Del Pezzo Surfaces

Recall that a del Pezzo surface is a smooth projective and geometrically irreducible surface whose anticanonical class is ample. We are interested in morphisms f:X2f:X\to\mathbb{P}^{2} where XX is a del Pezzo surface and the pullback f𝒪2(1)f^{*}{\mathcal{O}}_{\mathbb{P}^{2}}(1) is the anticanonical line bundle. It was shown in [Bea18, Prop. 4.1] that for such ff there exist ff-Ulrich line bundles. We will show that if ff is real fibered, then there are even positive hermitian ff-Ulrich line bundles. An introduction to the classical theory of del Pezzo surfaces can be found for example in [Dol12, Ch. 8] or [KSC04, §3.5]. This section further relies on the classification of real del Pezzo surfaces [Rus02].

Definition 12.1.

The degree of a del Pezzo surface XX is the self-intersection number KX.KXK_{X}.K_{X} of its canonical class KXK_{X}. A line on XX is an irreducible curve LXL\subset X such that L.L=L.KX=1L.L=L.K_{X}=-1.

Remark 12.2.

Note that if the anticanonical class KX-K_{X} of a del Pezzo surface XX is very ample, then a line LL on XX is mapped by the associated embedding to an actual line, i.e. a linear subspace of dimension one because L.KX=1L.K_{X}=-1.

Example 12.3.

These are examples of del Pezzo surfaces [KSC04, Thm. 3.36(7)]:

  1. (1)

    A smooth hypersurface of degree four in the weighted projective space (2,1,1,1)\mathbb{P}(2,1,1,1) is a del Pezzo surface of degree two.

  2. (2)

    A smooth cubic hypersurface in 3\mathbb{P}^{3} is a del Pezzo surface of degree three.

  3. (3)

    A smooth complete intersection of two quadrics in 4\mathbb{P}^{4} is a del Pezzo surface of degree four.

Furthermore, in case (1) the (complete) anticanonical linear system corresponds to the projection (2,1,1,1)2\mathbb{P}(2,1,1,1)\dashrightarrow\mathbb{P}^{2} restricted to our surface. Moreover, in the cases (2) and (3) the embeddings of the surfaces to 3\mathbb{P}^{3} and 4\mathbb{P}^{4} respectively correspond to the (complete) anticanonical linear system. These statements are for example shown in the course of the proof of [KSC04, Thm. 3.36].

Remark 12.4.

If XX is a del Pezzo surface over an algebraically closed field, then XX is isomorphic to either 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} or a blowup of 2\mathbb{P}^{2} in n8n\leq 8 points, see [KSC04, Ex. 3.56] or [Dol12, Cor. 8.1.17, Prop. 8.1.25]. A straight-forward calculation shows that the degree of XX is eight in the former and 9n9-n in the latter case.

Lemma 12.5.

Let X4X\subset\mathbb{P}^{4} be a smooth complete intersection of two quadrics in 4\mathbb{P}^{4} such that X()X({\mathbb{R}}) is homeomorphic to the disjoint union of two spheres. Then:

  1. (1)

    XX is contained in exactly five real singular quadrics;

  2. (2)

    one of these five quadrics has signature (2,2)(2,2) and the other four have signature (3,1)(3,1);

  3. (3)

    for exactly two of these singular quadrics the linear projection from its vertex realizes XX as a real fibered double cover of a hyperbolic quadratic hypersurface Q3Q\subset\mathbb{P}^{3}.

Proof.

The complex pencil λq0+μq1\lambda q_{0}+\mu q_{1} contains five singular quadrics as the qiq_{i} can be represented by symmetric 5×55\times 5 matrices. We will show that all of them are real. To this end recall that by 12.4 the complexification XX_{\mathbb{C}} of XX is isomorphic to the blowup of 2\mathbb{P}^{2} at five points p0,,p4p_{0},\ldots,p_{4}. The 16 lines on XX_{\mathbb{C}} correspond to the exceptional divisors E0,,E4E_{0},\ldots,E_{4}, the lines lijl_{ij} through pip_{i} and pjp_{j} for 0i<j40\leq i<j\leq 4 and the conic CC through all five points p0,,p4p_{0},\ldots,p_{4}, see for example [Man86, Thm. 26.2]. After relabeling if necessary, the complex conjugation on XX_{\mathbb{C}} interchanges E0E_{0} with CC, EiE_{i} with l0il_{0i} for 1i41\leq i\leq 4 and lijl_{ij} with lkll_{kl} for {i,j,k,l}={1,2,3,4}\{i,j,k,l\}=\{1,2,3,4\} and i<ji<j, k<lk<l, see [Rus02, Exp. 2, case n=3n=3]. We write

A1=E0+C,A2=l12+l34,A3=l13+l24,A4=l14+l23A_{1}=E_{0}+C,A_{2}=l_{12}+l_{34},A_{3}=l_{13}+l_{24},A_{4}=l_{14}+l_{23}

(divisors on XX) and note that all AiA_{i} belong to the same linear system. Similarly, the divisors Bi=Ei+l0iB_{i}=E_{i}+l_{0i} for 1i41\leq i\leq 4 on XX are also linearly equivalent to each other. Note that we did write Ai=Li+Li¯A_{i}=L_{i}+\overline{L_{i}} and Bi=Li+Li¯B_{i}=L_{i}^{\prime}+\overline{L_{i}^{\prime}} for suitable lines LiL_{i} and LiL_{i}^{\prime} on XX_{\mathbb{C}}. Each of the two linear systems realize XX as a conic bundle, i.e., define a morphism X1X\to\mathbb{P}^{1} all of whose fibers are isomorphic to a plane conic curve [Rus02, Exp. 2]. The four singular fibers of each bundle are exactly the AiA_{i} and BiB_{i} respectively. Therefore, for each connected component SS of X()X({\mathbb{R}}) there are exactly two values for ii and jj such that LiLi¯L_{i}\cap\overline{L_{i}} resp. LjLj¯L_{j}^{\prime}\cap\overline{L_{j}^{\prime}} is a point on SS. Our two conic bundle structures on XX induce a map X1×1X\to\mathbb{P}^{1}\times\mathbb{P}^{1} which is a double cover since AiBj=2A_{i}\cdot B_{j}=2. Since Ai+BjA_{i}+B_{j} is an anticanonical divisor, this double cover is a linear projection of XX to 3\mathbb{P}^{3} whose image is a hypersurface isomorphic to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}, i.e., defined by a quadric with signature (2,2)(2,2). The cone over this quadric in 4\mathbb{P}^{4} is one of our singular quadrics. The four divisors

Dj=L1+Lj¯=E0+l0jD_{j}=L_{1}+\overline{L_{j}^{\prime}}=E_{0}+l_{0j}

on XX_{\mathbb{C}} for 1j41\leq j\leq 4 also realize XX_{\mathbb{C}} as a conic bundle X1X_{\mathbb{C}}\to\mathbb{P}^{1}_{\mathbb{C}} and for each jj we consider the map X1×1X_{\mathbb{C}}\to\mathbb{P}^{1}_{\mathbb{C}}\times\mathbb{P}^{1}_{\mathbb{C}} associated to DjD_{j} in the first coordinate and Dj¯\overline{D_{j}} on the second. This corresponds to a morphism fj:XQf_{j}:X\to Q where Q3Q\subset\mathbb{P}^{3} is a hypersurface defined by a quadric of signature (3,1)(3,1). Since DjDj¯=2D_{j}\cdot\overline{D_{j}}=2, this is a double cover, and since Dj+Dj¯D_{j}+\overline{D_{j}} is anticanonical, the maps fjf_{j} correspond to linear projections of XX. This shows (2)(2) and (3)(3). We have

DjDj¯=L1L1¯+LjLj¯.D_{j}\cdot\overline{D_{j}}=L_{1}\cdot\overline{L_{1}}+L_{j}^{\prime}\cdot\overline{L_{j}^{\prime}}.

Thus in order to determine whether fi:XQf_{i}:X\to Q is real fibered or not, we have to check whether the two intersection points L1L1¯L_{1}\cap\overline{L_{1}} and LiLi¯L_{i}^{\prime}\cap\overline{L_{i}^{\prime}} lie on the same (not real fibered) or different connected components (real fibered) of X()X({\mathbb{R}}). As noted above, both cases occur for exactly two values of jj. ∎

With this preparation we are able to determine for which del Pezzo surfaces XX there is a real fibered morphism X2X\to\mathbb{P}^{2} whose corresponding linear system is anticanonical.

Proposition 12.6.

Let XX be a real del Pezzo surface and KK a canonical divisor on XX. There is a real fibered morphism f:X2f:X\to\mathbb{P}^{2} such that the pullback of a line is linearly equivalent to K-K on XX if and only if we have one of the following:

  1. (1)

    XX is a quartic surface in 4\mathbb{P}^{4} such that X()X({\mathbb{R}}) is homeomorphic to a disjoint union of two spheres;

  2. (2)

    XX is a cubic hypersurface in 3\mathbb{P}^{3} such that X()X({\mathbb{R}}) is homeomorphic to a disjoint union of a sphere and a real projective plane;

  3. (3)

    XX is a double cover of 2\mathbb{P}^{2} branched along a smooth plane quartic curve CC with C()=C({\mathbb{R}})=\emptyset so that X()X({\mathbb{R}}) is homeomorphic to a disjoint union of two real projective planes.

In particular, in each case X()X({\mathbb{R}}) has two connected components.

Proof.

Let d=K.Kd=K.K. For d=2d=2 the anticanonical map is a double cover of 2\mathbb{P}^{2} branched along a plane quartic curve CC, see 12.3. This is real fibered if and only if C()=C({\mathbb{R}})=\emptyset and X()X({\mathbb{R}}) is homeomorphic to a disjoint union of two real projective planes. In general if there exists such a morphism ff, then X()X({\mathbb{R}}) must be homeomorphic to the disjoint union of ss spheres and rr real projective planes such that d=2s+rd=2s+r by [KS20a, Cor. 2.20]. Going through the classification of real del Pezzo surfaces in [Rus02] we see that for d2d\neq 2 this is only possible for a complete intersection of two quadrics in 4\mathbb{P}^{4} such that X()X({\mathbb{R}}) is homeomorphic to a disjoint union of two spheres (d=4d=4) or a cubic hypersurface in 3\mathbb{P}^{3} such that X()X({\mathbb{R}}) is homeomorphic to a disjoint union of a sphere and a real projective plane (d=3d=3). This shows the “only if part”. It thus remains to show that the embedded surfaces in (1)(1) and (2)(2) are hyperbolic as these embeddings correspond to the anticanonical linear system, see 12.3. The case d=3d=3 is covered by [Vin12, Prop. 2.2]. For the case d=4d=4 we can compose a real fibered linear projection XQX\to Q from 12.5 to a hyperbolic hypersurface Q3Q\subset\mathbb{P}^{3} with the linear projection Q2Q\to\mathbb{P}^{2} from a point with respect to which QQ is hyperbolic. ∎

Lemma 12.7.

Let XX be a del Pezzo surface and KK a canonical divisor on XX. Let f:X2f:X\to\mathbb{P}^{2} be a real fibered morphism of degree dd such that the pullback of a line is linearly equivalent to K-K as constructed in the previous Proposition.

  1. a)

    The ramification divisor RR is linearly equivalent to 2K-2K.

  2. b)

    Let gg be a rational function on XX whose principal divisor is R+2KR+2K. Then gg has constant sign on each of the two connected components of X()X({\mathbb{R}}) and these signs are not the same.

Proof.

The ramification divisor is linearly equivalent to the canonical divisor on XX minus the pullback of the canonical divisor on 2\mathbb{P}^{2}. As the latter is 𝒪2(3){\mathcal{O}}_{\mathbb{P}^{2}}(-3) its pullback is 3K3K. This shows a)a).

For d=2d=2 part b)b) follows from the fact XX is a double cover of 2\mathbb{P}^{2} of the form y2=p(x)y^{2}=p(x) where p[x0,x1,x2]4p\in{\mathbb{R}}[x_{0},x_{1},x_{2}]_{4} is a globally positive quartic curve. The ramification divisor is given as the zero locus of yy whereas K-K is the zero set of a linear form l[x0,x1,x2]l\in{\mathbb{R}}[x_{0},x_{1},x_{2}]. Clearly yl2\frac{y}{l^{2}} has the desired properties.

In the case d=3d=3 our surface XX is the zero set of a hyperbolic polynomial hh and our morphism ff is the linear projection from a point e3e\in\mathbb{P}^{3} of hyperbolicity. Its ramification divisor is thus cut out by the interlacer Deh\textrm{D}_{e}h. Again K-K is the zero set of a linear form l[x0,x1,x2]l\in{\mathbb{R}}[x_{0},x_{1},x_{2}] and Dehl2\frac{\textrm{D}_{e}h}{l^{2}} has different sign on the two connected components of X()X({\mathbb{R}}).

In the case d=4d=4 we first note that by 12.5 we can assume (after a linear change of coordinates) that XX is cut out by q1=x02(x12+x22+x32)q_{1}=x_{0}^{2}-(x_{1}^{2}+x_{2}^{2}+x_{3}^{2}) and q2=x42pq_{2}=x_{4}^{2}-p for some quadratic form p[x0,x1,x2,x3]2p\in{\mathbb{R}}[x_{0},x_{1},x_{2},x_{3}]_{2} that is nonnegative on Q=𝒱(q1)3Q={\mathcal{V}}(q_{1})\subset\mathbb{P}^{3}. Our real fibered morphism f:X2f:X\to\mathbb{P}^{2} is then the composition of the two linear projections XQX\to Q with center [0:0:0:0:1][0:0:0:0:1] and Q2Q\to\mathbb{P}^{2} with center [1:0:0:0][1:0:0:0]. Thus the ramification locus of ff is cut out by x0x4x_{0}\cdot x_{4}. Again K-K is the zero set of a linear form l[x0,x1,x2]l\in{\mathbb{R}}[x_{0},x_{1},x_{2}] and x0x4l2\frac{x_{0}\cdot x_{4}}{l^{2}} has different sign on the two connected components of X()X({\mathbb{R}}). ∎

Let X4X_{4} be a complete intersection of two quadrics in 4\mathbb{P}^{4} such that X()X({\mathbb{R}}) is homeomorphic to a disjoint union of two spheres. We fix a sequence of morphisms

(2) X2X3X4\displaystyle X_{2}\to X_{3}\to X_{4}

where each map fi:XiXi+1f_{i}:X_{i}\to X_{i+1} is the blow up of XiX_{i} at real point on a connected component of Xi()X_{i}({\mathbb{R}}) that is homeomorphic to a sphere. Further let EiE_{i} be the exceptional divisor of fif_{i}. By the classification of real del Pezzo surfaces in [Rus02], we have that X3X_{3} is a cubic hypersurface in 3\mathbb{P}^{3} such that X()X({\mathbb{R}}) is homeomorphic to a disjoint union of a sphere and a real projective plane and X2X_{2} is a double cover of 2\mathbb{P}^{2} branched along a smooth plane quartic curve CC with C()=C({\mathbb{R}})=\emptyset so that X()X({\mathbb{R}}) is homeomorphic to a disjoint union of two real projective planes. Conversely, every such real del Pezzo surface fits in such a sequence of blow-ups.

Lemma 12.8.

Consider the cubic hypersurface X33X_{3}\subset\mathbb{P}^{3}.

  1. a)

    In addition to E3E_{3} there are two more real lines LL and LL^{\prime} on X3X_{3}. These three lines lie on a common plane.

  2. b)

    There are two different hyperplanes that contain LL and are tangential to a real point on the connected component of X3()X_{3}({\mathbb{R}}) that is homeomorphic to the sphere.

  3. c)

    The divisors H1H_{1} and H2H_{2} on X3X_{3} that are defined as the intersections with the hyperplanes from part b)b) are of the form Hi=L+Li+Li¯H_{i}=L+L_{i}+\overline{L_{i}} for some non-real lines LiL_{i} on (X3)(X_{3})_{\mathbb{C}}.

  4. d)

    The lines LiL_{i} and Li¯\overline{L_{i}} are disjoint from E3E_{3}. Furthermore, L1L_{1} and L2L_{2} are disjoint.

  5. e)

    Let ff be a rational function on X3X_{3} whose principal divisor is L2+L2¯L1L1¯L_{2}+\overline{L_{2}}-L_{1}-\overline{L_{1}}. Then ff has constant sign on each of the two connected components of X()X({\mathbb{R}}) and these signs are not the same.

Proof.

The number of real lines on X3X_{3} can be found for example in [Rus02, p. 302]. As they all must lie in the component of X()X({\mathbb{R}}) that is homeomorphic to 2{\mathbb{R}}\mathbb{P}^{2}, each two of them intersect in a point. Thus they all lie in a common plane H0H_{0} which proves a)a).

In the affine chart 3=(3H0)(){\mathbb{R}}^{3}=(\mathbb{P}^{3}\setminus H_{0})({\mathbb{R}}) the connected component of X3()X_{3}({\mathbb{R}}) that is homeomorphic to a sphere is the boundary of a compact convex set K3K\subset{\mathbb{R}}^{3}, namely KK is an affine slice of the hyperbolicity cone of the cubic that defines X3X_{3}. The hyperplanes containing LL correspond to a family of parallel affine hyperplanes in 3{\mathbb{R}}^{3}. Thus exactly two of them are tangent to KK. This shows b)b).

The zero divisors of these hyperplanes HiH_{i} contain besides LL a plane conic which has an isolated real point, namely the point of tangency. Thus the conic is a complex conjugate pair of lines LiL_{i} and Li¯\overline{L_{i}} which shows part c)c).

In order to show d)d) assume for the sake of a contradiction that LiL_{i} intersects E3E_{3}. Since LiHiL_{i}\subset H_{i} and E3H0E_{3}\subset H_{0}, this intersection point must lie on E3H0Hi=E3LE_{3}\cap H_{0}\cap H_{i}=E_{3}\cap L which implies that is real. But the only real point of LiL_{i} lies on the spherical component of X3()X_{3}({\mathbb{R}}). An analogous argument shows that L1L_{1} and L2L_{2} are disjoint.

Finally, let lil_{i} be the linear form that cuts out HiH_{i}. Then by construction p=l1l2p=l_{1}l_{2} is an interlacer of the polynomial defining X3X_{3}. Thus the rational function f=pl12f=\frac{p}{l_{1}^{2}} has constant sign on each of the two connected components of X()X({\mathbb{R}}) and these signs are not the same. Clearly the principal divisor corresponding to ff is L2+L2¯L1L1¯L_{2}+\overline{L_{2}}-L_{1}-\overline{L_{1}}. Therefore, we have shown part d)d). ∎

Refer to caption
Figure 3. A cubic hyperbolic hypersurface with two planes that contain a line on the pseudoplane (red) and are tangent to the spherical component (yellow).
Theorem 12.9.

Let XX be a del Pezzo surface and f:X2f:X\to\mathbb{P}^{2} a real fibered morphism of degree dd such that the pullback of a line is the anticanonical divisor class K-K. Then there is a positive hermitian ff-Ulrich line bundle.

Proof.

We put XX into a sequence of blow-ups as in (2). Since the lines LiL_{i} and Li¯\overline{L_{i}} on (X3)(X_{3})_{\mathbb{C}} from 12.8 are disjoint from E3E_{3}, they can be identified with some lines on (X4)(X_{4})_{\mathbb{C}} which we, by abuse of notation, also denote by LiL_{i} and Li¯\overline{L_{i}}. The same we do for the proper transforms of LiL_{i} and Li¯\overline{L_{i}} in (X2)(X_{2})_{\mathbb{C}}. We want to apply part b)b) of 10.3 to the divisor M=L2L1KM=L_{2}-L_{1}-K where KK is a canonical divisor on XX. Since XX is birational to X3X_{3} the rational function ff from part e)e) of 12.8 is also a rational function on XX and we have M+M¯=(f)2KM+\overline{M}=(f)-2K. By 12.7 there is a rational function gg on XX such that (g)=R+2K(g)=R+2K where RR is the ramification divisor. Furthermore, we can choose gg to have the same sign on each of X()X({\mathbb{R}}) as ff. Thus M+M¯=(fg)+RM+\overline{M}=(\frac{f}{g})+R and fg\frac{f}{g} is nonnegative. It thus remains to show that the dimension (M)\ell(M) of the space of global sections of (M){\mathcal{L}}(M) is at least dd. To this end we invoke the Theorem of Riemann–Roch for surfaces [Har77, Thm. 1.6]:

(M)+(KM)=12M.(MK)+1+pa+s(M)12M.(MK)+1=d.\ell(M)+\ell(K-M)=\frac{1}{2}M.(M-K)+1+p_{a}+s(M)\geq\frac{1}{2}M.(M-K)+1=d.

Now since the intersection product of (KM)(K-M) with the ample divisor K-K equals 2d<0-2d<0, it cannot be effective. Thus (KM)=0\ell(K-M)=0 and the claim follows. ∎

We now apply 12.9 to the three cases from 12.6. The following consequence is originally due to Buckley and Koşir [BK07].

Corollary 12.10.

Every hyperbolic polynomial h[x0,x1,x2,x3]h\in{\mathbb{R}}[x_{0},x_{1},x_{2},x_{3}] of degree three has a definite hermitian determinantal representation.

Proof.

First assume that the zero set of hh is smooth. Then we are in case (2)(2) of 12.6 and the claim follows from 12.9 and 9.1. For the singular case note that by [Nui68] the set of all hyperbolic polynomials is the closure of the smooth ones. Further by [PV13, Lem. 3.4] the set of hyperbolic polynomials with a definite hermitian determinantal representation is closed. ∎

The following consequence is originally due to Hilbert [Hil88].

Corollary 12.11.

Every nonnegative ternary quartic is a sum of three squares.

Proof.

First consider a nonnegative ternary quartic pp with smooth zero set. The hypersurface defined by y2py^{2}-p in (2,1,1,1)\mathbb{P}(2,1,1,1) is an instance of 12.6(3). Thus the claim follows from 12.9 and 9.6. The general case now follows from a limit argument as the set of sums of squares is closed in [x0,x1,x2]4{\mathbb{R}}[x_{0},x_{1},x_{2}]_{4}. ∎

Corollary 12.12.

The Chow form of a smooth hyperbolic surface in 4\mathbb{P}^{4} of degree four, which is a complete intersection of two quadrics, has a definite hermitian determinantal representation.

Proof.

Here we are in case (1) of 12.6. The claim follows from 12.9 together with a straight-forward adaption of the proof of [KS20a, Thm. 5.7] to the hermitian case and [KS20a, Rem. 4.4]. ∎

Remark 12.13.

We have seen that every nonnegative polynomial p[x0,x1,x2]4p\in{\mathbb{R}}[x_{0},x_{1},x_{2}]_{4} is a sum of squares and every hyperbolic polynomial h[x0,x1,x2,x3]3h\in{\mathbb{R}}[x_{0},x_{1},x_{2},x_{3}]_{3} has a definite hermitian determinantal representation, i.e., the associated real fibered morphisms admit a positive Ulrich sheaf. This is no longer true if we increase the degrees: Not every nonnegative polynomial p[x0,x1,x2]6p\in{\mathbb{R}}[x_{0},x_{1},x_{2}]_{6} is a sum of squares [Hil88] and there are hyperbolic polynomials h[x0,x1,x2,x3]4h\in{\mathbb{R}}[x_{0},x_{1},x_{2},x_{3}]_{4} such that no power hrh^{r} has a definite determinantal representation, take for example the polynomial considered in [Kum16b]. Double covers of 2\mathbb{P}^{2} ramified along plane sextic curves and quartic hypersurfaces in 3\mathbb{P}^{3} both belong to the class of K3 surfaces. So it would be very interesting to understand which real fibered morphisms X2X\to\mathbb{P}^{2} from a K3 surface XX admit a positive Ulrich bundle. Note that (not necessarily positive) Ulrich sheaves of rank two on K3 surfaces have been constructed in [AFO17]. Similarly, we can increase the dimensions: Not every nonnegative polynomial p[x0,x1,x2,x3]4p\in{\mathbb{R}}[x_{0},x_{1},x_{2},x_{3}]_{4} is a sum of squares [Hil88] and it is not known whether there are hyperbolic polynomials h[x0,x1,x2,x3,x4]3h\in{\mathbb{R}}[x_{0},x_{1},x_{2},x_{3},x_{4}]_{3} such that no power hrh^{r} has a definite determinantal representation, see [Sau19, §5] for cubic hyperbolic hypersurfaces. Double covers of 3\mathbb{P}^{3} ramified along quartic surfaces and cubic hypersurfaces in 4\mathbb{P}^{4} both belong to the class of Fano threefolds of index two. In [Bea18, §6] Ulrich sheaves of rank two on such threefolds have been constructed.

Problem 12.14.

Understand which finite surjective and real fibered morphisms XnX\to\mathbb{P}^{n} admit a positive Ulrich sheaf for XX a K3 surface or a Fano threefold of index two. If such exist, what are their ranks? Are there hyperbolic cubic hypersurfaces in 4\mathbb{P}^{4} that do not carry a positive Ulrich sheaf?

We conclude this section with some examples.

Example 12.15.

Let h=x03x0(2x12+2x22+x32)+x13+x1x22h=x_{0}^{3}-x_{0}(2x_{1}^{2}+2x_{2}^{2}+x_{3}^{2})+x_{1}^{3}+x_{1}x_{2}^{2}. The hypersurface X=𝒱(h)3X={\mathcal{V}}(h)\subset\mathbb{P}^{3} is hyperbolic with respect to e=[1:0:0:0]e=[1:0:0:0] and contains the real line L=𝒱(x0,x1)L={\mathcal{V}}(x_{0},x_{1}). The hyperplanes H1=𝒱(x0)H_{1}={\mathcal{V}}(x_{0}) and H2=𝒱(x0x1)H_{2}={\mathcal{V}}(x_{0}-x_{1}) contain LL and are tangent to the hyperbolicity cone of hh. The quadratic polynomial p=x0(x0x1)p=x_{0}(x_{0}-x_{1}) is an interlacer of hh and its zero divisor on XX is

D=2L+L1+L1¯+L2+L2¯D=2L+L_{1}+\overline{L_{1}}+L_{2}+\overline{L_{2}}

where L1=𝒱(x0,x1+ix2)L_{1}={\mathcal{V}}(x_{0},x_{1}+\textrm{i}x_{2}) and L2=𝒱(x0x1,x2+ix3)L_{2}={\mathcal{V}}(x_{0}-x_{1},x_{2}+\textnormal{i}x_{3}). Thus we have D=M+M¯D=M+\overline{M} with M=L+L1+L2M=L+L_{1}+L_{2}. The space of all quadrics vanishing on L,L1L,L_{1} and L2L_{2} is spanned by

x0(x0x1),x0(x2+ix3),(x0x1)(x1+ix2).x_{0}(x_{0}-x_{1}),x_{0}(x_{2}+\textnormal{i}x_{3}),(x_{0}-x_{1})(x_{1}+\textrm{i}x_{2}).

The minimal free resolution over S=[x0,x1,x2,x3]S={\mathbb{R}}[x_{0},x_{1},x_{2},x_{3}] of the ideal in S/(h)S/(h) generated by these quadrics has length one and is given by the matrix

(x0+x1x2ix3x1ix2x2+ix3x0x10x1+ix20x0).\begin{pmatrix}x_{0}+x_{1}&-x_{2}-\textnormal{i}x_{3}&-x_{1}-\textnormal{i}x_{2}\\ -x_{2}+\textnormal{i}x_{3}&x_{0}-x_{1}&0\\ -x_{1}+\textnormal{i}x_{2}&0&x_{0}\end{pmatrix}.

This matrix is hermitian and positive definite at ee. Its determinant is indeed hh.

Example 12.16.

Consider the following nonnegative ternary quartic

p=x04+2x02x12+2x0x13+x14+x02x222x0x1x22x12x22+x24p=x_{0}^{4}+2x_{0}^{2}x_{1}^{2}+2x_{0}x_{1}^{3}+x_{1}^{4}+x_{0}^{2}x_{2}^{2}-2x_{0}x_{1}x_{2}^{2}-x_{1}^{2}x_{2}^{2}+x_{2}^{4}

and let X(2,1,1,1)X\subset\mathbb{P}(2,1,1,1) be the corresponding double cover defined by y2=py^{2}=p. On XX we have the two lines

L1=𝒱(y+(1i)x12x22,x0+ix1) and L2=𝒱(y+ix1x2x12+x22,x0+ix2).L_{1}={\mathcal{V}}(y+(1-\textnormal{i})x_{1}^{2}-x_{2}^{2},x_{0}+\textnormal{i}x_{1})\textrm{ and }L_{2}={\mathcal{V}}(y+\textnormal{i}x_{1}x_{2}-x_{1}^{2}+x_{2}^{2},x_{0}+\textnormal{i}x_{2}).

The principal divisor associated to the rational function

f=x02+x22yx0x1x12+x22f=\frac{x_{0}^{2}+x_{2}^{2}}{y-x_{0}x_{1}-x_{1}^{2}+x_{2}^{2}}

is L2L1+L2¯L1¯L_{2}-L_{1}+\overline{L_{2}}-\overline{L_{1}}. As L1L_{1} and L2L_{2} are both non-real lines, this implies that ff has constant sign on each of the two connected components of X()X({\mathbb{R}}). Evaluating ff at points from the two different components, for example at

[y:x0:x1:x2]=[±1:1:0:0],[y:x_{0}:x_{1}:x_{2}]=[\pm 1:1:0:0],

shows that ff changes sign. Letting KK be a canonical divisor on XX, we have for M=L2L1KM=L_{2}-L_{1}-K that M+M¯=(f)2KM+\overline{M}=(f)-2K as in the proof of 12.9. We can realize the divisor class of MM by the ideal in [y,x0,x1,x2]/(y2p){\mathbb{R}}[y,x_{0},x_{1},x_{2}]/(y^{2}-p) that is generated by yx0x1x12+x22y-x_{0}x_{1}-x_{1}^{2}+x_{2}^{2} and (x0+ix1)(x0+ix2)(x_{0}+\textnormal{i}x_{1})(x_{0}+\textnormal{i}x_{2}). The minimal free resolution over the ring [y,x0,x1,x2]{\mathbb{R}}[y,x_{0},x_{1},x_{2}] has length one and is given by the matrix

(yx0x1x12+x22x02x1x2+i(x0x1+x0x2)x02x1x2i(x0x1+x0x2)y+x0x1+x12x22).\begin{pmatrix}y-x_{0}x_{1}-x_{1}^{2}+x_{2}^{2}&x_{0}^{2}-x_{1}x_{2}+\textrm{i}(x_{0}x_{1}+x_{0}x_{2})\\ x_{0}^{2}-x_{1}x_{2}-\textrm{i}(x_{0}x_{1}+x_{0}x_{2})&y+x_{0}x_{1}+x_{1}^{2}-x_{2}^{2}\end{pmatrix}.

Indeed, we have that A2=pIA^{2}=p\cdot I. Therefore

p=(x0x1+x12x22)2+(x02x1x2)2+(x0x1+x0x2)2.p=(x_{0}x_{1}+x_{1}^{2}-x_{2}^{2})^{2}+(x_{0}^{2}-x_{1}x_{2})^{2}+(x_{0}x_{1}+x_{0}x_{2})^{2}.

The key ingredient for this construction was the rational function ff. One way to find it is to blow down a real line of XX and then proceed as in 12.8.

Example 12.17.

The del Pezzo surface

X=𝒱(x02+x12+x22x32,x02+4x12+9x22x42)4X={\mathcal{V}}(x_{0}^{2}+x_{1}^{2}+x_{2}^{2}-x_{3}^{2},x_{0}^{2}+4x_{1}^{2}+9x_{2}^{2}-x_{4}^{2})\subset\mathbb{P}^{4}

of degree four is hyperbolic with respect to the line EE spanned by [0:0:0:1:0][0:0:0:1:0] and [0:0:0:0:1][0:0:0:0:1]. The rational function (x3x0)(x4x0)x02\frac{(x_{3}-x_{0})(x_{4}-x_{0})}{x_{0}^{2}} has different signs on the two connected components of X()X({\mathbb{R}}). Its corresponding divisor is of the form M+M¯+2KM+\overline{M}+2K for a suitable divisor MM on XX_{\mathbb{C}} with (M)=4\ell(M)=4. Thus (M){\mathcal{L}}(M) is a positive hermitian Ulrich line bundle by 10.3. As in [ESW03, Thm. 0.3] we obtain from that the following determinantal representation of the Chow form of XX, written in the Plücker coordinates:

(2x032x04+2x344x016ix02+4x136ix232x01+2ix022x14+2ix242ix124x01+6ix02+4x13+6ix232x032x04+2x3410ix122x012ix022x14+2ix242x012ix022x142ix2410ix122x03+2x04+2x344x01+6ix02+4x136ix232ix122x01+2ix022x142ix244x016ix02+4x13+6ix232x03+2x04+2x34)\begin{pmatrix}2x_{03}-2x_{04}+2x_{34}&4x_{01}-6\textrm{i}x_{02}+4x_{13}-6\textrm{i}x_{23}&-2x_{01}+2\textrm{i}x_{02}-2x_{14}+2\textrm{i}x_{24}&-2\textrm{i}x_{12}\\ 4x_{01}+6\textrm{i}x_{02}+4x_{13}+6\textrm{i}x_{23}&-2x_{03}-2x_{04}+2x_{34}&10\textrm{i}x_{12}&2x_{01}-2\textrm{i}x_{02}-2x_{14}+2\textrm{i}x_{24}\\ -2x_{01}-2\textrm{i}x_{02}-2x_{14}-2\textrm{i}x_{24}&-10\textrm{i}x_{12}&2x_{03}+2x_{04}+2x_{34}&-4x_{01}+6\textrm{i}x_{02}+4x_{13}-6\textrm{i}x_{23}\\ 2\textrm{i}x_{12}&2x_{01}+2\textrm{i}x_{02}-2x_{14}-2\textrm{i}x_{24}&-4x_{01}-6\textrm{i}x_{02}+4x_{13}+6\textrm{i}x_{23}&-2x_{03}+2x_{04}+2x_{34}\end{pmatrix}

We observe that it is hermitian and positive definite at EE.

References

  • [AFO17] Marian Aprodu, Gavril Farkas, and Angela Ortega. Minimal resolutions, Chow forms and Ulrich bundles on K3K3 surfaces. J. Reine Angew. Math., 730:225–249, 2017.
  • [Bea00] Arnaud Beauville. Determinantal hypersurfaces. volume 48, pages 39–64. 2000. Dedicated to William Fulton on the occasion of his 60th birthday.
  • [Bea18] Arnaud Beauville. An introduction to Ulrich bundles. Eur. J. Math., 4(1):26–36, 2018.
  • [Bec82] Eberhard Becker. Valuations and real places in the theory of formally real fields. In Real algebraic geometry and quadratic forms (Rennes, 1981), volume 959 of Lecture Notes in Math., pages 1–40. Springer, Berlin-New York, 1982.
  • [BK07] Anita Buckley and Tomaž Košir. Determinantal representations of smooth cubic surfaces. Geom. Dedicata, 125:115–140, 2007.
  • [Brä11] Petter Brändén. Obstructions to determinantal representability. Adv. in Math., 226(2):1202–1212, 2011.
  • [BSV16] Grigoriy Blekherman, Gregory G. Smith, and Mauricio Velasco. Sums of squares and varieties of minimal degree. J. Amer. Math. Soc., 29(3):893–913, 2016.
  • [Cob82] Arthur B. Coble. Algebraic geometry and theta functions, volume 10 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, R.I., 1982. Reprint of the 1929 edition.
  • [COSW04] Young-Bin Choe, James G. Oxley, Alan D. Sokal, and David G. Wagner. Homogeneous multivariate polynomials with the half-plane property. Adv. in Appl. Math., 32(1-2):88–187, 2004. Special issue on the Tutte polynomial.
  • [Dol12] Igor V. Dolgachev. Classical algebraic geometry. Cambridge University Press, Cambridge, 2012. A modern view.
  • [ES11] David Eisenbud and Frank-Olaf Schreyer. Boij-Söderberg theory. In Combinatorial aspects of commutative algebra and algebraic geometry, volume 6 of Abel Symp., pages 35–48. Springer, Berlin, 2011.
  • [ESW03] David Eisenbud, Frank-Olaf Schreyer, and Jerzy Weyman. Resultants and Chow forms via exterior syzygies. J. Amer. Math. Soc., 16(3):537–579, 2003.
  • [Gro65] A. Grothendieck. Éléments de géométrie algébrique. IV. Étude locale des schémas et des morphismes de schémas. II. Inst. Hautes Études Sci. Publ. Math., (24):231, 1965.
  • [Han17] Christoph Hanselka. Characteristic polynomials of symmetric matrices over the univariate polynomial ring. J. Algebra, 487:340–356, 2017.
  • [Har77] Robin Hartshorne. Algebraic geometry. Springer-Verlag, New York-Heidelberg, 1977. Graduate Texts in Mathematics, No. 52.
  • [Hil88] David Hilbert. Ueber   die   Darstellung   definiter   Formen   als   Summe   von   Formen- quadraten. Math. Ann., 32(3):342–350, 1888.
  • [HK71] Jürgen Herzog and Ernst Kunz, editors. Der kanonische Modul eines Cohen-Macaulay-Rings. Lecture Notes in Mathematics, Vol. 238. Springer-Verlag, Berlin-New York, 1971. Seminar über die lokale Kohomologietheorie von Grothendieck, Universität Regensburg, Wintersemester 1970/1971.
  • [HV07] J. William Helton and Victor Vinnikov. Linear matrix inequality representation of sets. Comm. Pure Appl. Math., 60(5):654–674, 2007.
  • [Kol96] János Kollár. Rational curves on algebraic varieties, volume 32 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer-Verlag, Berlin, 1996.
  • [KPV15] Mario Kummer, Daniel Plaumann, and Cynthia Vinzant. Hyperbolic polynomials, interlacers, and sums of squares. Math. Program., 153(1, Ser. B):223–245, 2015.
  • [KS20a] Mario Kummer and Eli Shamovich. Real fibered morphisms and Ulrich sheaves. J. Algebraic Geom., 29:167–198, 2020.
  • [KS20b] Mario Kummer and Rainer Sinn. Hyperbolic secant varieties of m-curves. arXiv preprint arXiv:2002.00486, 2020.
  • [KSC04] János Kollár, Karen E. Smith, and Alessio Corti. Rational and nearly rational varieties, volume 92 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2004.
  • [Kum13] Mario Kummer. Real Polynomials with Definite Determinantal Representation. Master’s Thesis, Universität Konstanz, Konstanz, 2013.
  • [Kum16a] Mario Kummer. From Hyperbolic Polynomials to Real Fibered Morphisms. PhD thesis, Universität Konstanz, Konstanz, 2016.
  • [Kum16b] Mario Kummer. A note on the hyperbolicity cone of the specialized Vámos polynomial. Acta Appl. Math., 144:11–15, 2016.
  • [Kum17] Mario Kummer. Determinantal representations and Bézoutians. Math. Z., 285(1-2):445–459, 2017.
  • [KV19] Mario Kummer and Cynthia Vinzant. The Chow form of a reciprocal linear space. Michigan Math. J., 68(4):831–858, 2019.
  • [Lax58] P. D. Lax. Differential equations, difference equations and matrix theory. Comm. Pure Appl. Math., 11:175–194, 1958.
  • [Man86] Yu. I. Manin. Cubic forms, volume 4 of North-Holland Mathematical Library. North-Holland Publishing Co., Amsterdam, second edition, 1986. Algebra, geometry, arithmetic, Translated from the Russian by M. Hazewinkel.
  • [NT12] Tim Netzer and Andreas Thom. Polynomials with and without determinantal representations. Linear Algebra Appl., 437(7):1579–1595, 2012.
  • [Nui68] Wim Nuij. A note on hyperbolic polynomials. Math. Scand., 23:69–72 (1969), 1968.
  • [PRS93] P. Pedersen, M.-F. Roy, and A. Szpirglas. Counting real zeros in the multivariate case. In Computational algebraic geometry (Nice, 1992), volume 109 of Progr. Math., pages 203–224. Birkhäuser Boston, Boston, MA, 1993.
  • [PRSS04] Victoria Powers, Bruce Reznick, Claus Scheiderer, and Frank Sottile. A new approach to Hilbert’s theorem on ternary quartics. C. R. Math. Acad. Sci. Paris, 339(9):617–620, 2004.
  • [PS06] Nicholas Proudfoot and David Speyer. A broken circuit ring. Beiträge Algebra Geom., 47(1):161–166, 2006.
  • [PS12] Albrecht Pfister and Claus Scheiderer. An elementary proof of Hilbert’s theorem on ternary quartics. J. Algebra, 371:1–25, 2012.
  • [PV13] Daniel Plaumann and Cynthia Vinzant. Determinantal representations of hyperbolic plane curves: an elementary approach. J. Symbolic Comput., 57:48–60, 2013.
  • [Rus02] Francesco Russo. The antibirational involutions of the plane and the classification of real del Pezzo surfaces. In Algebraic geometry, pages 289–312. de Gruyter, Berlin, 2002.
  • [Sau19] James Saunderson. Certifying polynomial nonnegativity via hyperbolic optimization. SIAM J. Appl. Algebra Geom., 3(4):661–690, 2019.
  • [Ser79] Jean-Pierre Serre. Local fields, volume 67 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1979. Translated from the French by Marvin Jay Greenberg.
  • [SSV13] Raman Sanyal, Bernd Sturmfels, and Cynthia Vinzant. The entropic discriminant. Adv. Math., 244:678–707, 2013.
  • [Sta20] The Stacks project authors. The stacks project. https://stacks.math.columbia.edu, 2020.
  • [SV18] E. Shamovich and V. Vinnikov. Livsic-type determinantal representations and hyperbolicity. Adv. Math., 329:487–522, 2018.
  • [Var95] A. Varchenko. Critical points of the product of powers of linear functions and families of bases of singular vectors. Compositio Math., 97(3):385–401, 1995.
  • [Vin12] Victor Vinnikov. LMI representations of convex semialgebraic sets and determinantal representations of algebraic hypersurfaces: past, present, and future. In Mathematical methods in systems, optimization, and control, volume 222 of Oper. Theory Adv. Appl., pages 325–349. Birkhäuser/Springer Basel AG, Basel, 2012.