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

Some refinements of the Deligne–Illusie theorem

Piotr Achinger Institute of Mathematics of the Polish Academy of Sciences
ul. Śniadeckich 8, 00-656 Warsaw, Poland
pachinger@impan.pl
 and  Junecue Suh 1156 High Street
University of California in Santa Cruz
Santa Cruz CA 95064 USA
jusuh@ucsc.edu
Abstract.

We extend the results of Deligne and Illusie on liftings modulo p2p^{2} and decompositions of the de Rham complex in several ways. We show that for a smooth scheme XX over a perfect field kk of characteristic p>0p>0, the truncations of the de Rham complex in max(p1,2)\max(p-1,2) consecutive degrees can be reconstructed as objects of the derived category in terms of its truncation in degrees at most one (or, equivalently, in terms the obstruction class to lifting modulo p2p^{2}). Consequently, these truncations are decomposable if XX admits a lifting to W2(k)W_{2}(k), in which case the first nonzero differential in the conjugate spectral sequence appears no earlier than on page max(p,3)\max(p,3) (these corollaries have been recently strengthened by Drinfeld, Bhatt–Lurie, and Li–Mondal). Without assuming the existence of a lifting, we describe the gerbes of splittings of two-term truncations and the differentials on the second page of the conjugate spectral sequence, answering a question of Katz.

The main technical result used in the case p>2p>2 belongs purely to homological algebra. It concerns certain commutative differential graded algebras whose cohomology algebra is the exterior algebra, dubbed by us abstract Koszul complexes, of which the de Rham complex in characteristic pp is an example.

In the appendix, we use the aforementioned stronger decomposition result to prove that Kodaira–Akizuki–Nakano vanishing and Hodge–de Rham degeneration both hold for FF-split (p+1)(p+1)-folds.

1. Introduction

1.1. Decompositions of the de Rham complex

In [DI87], Deligne and Illusie showed that for a smooth scheme XX over a perfect field kk of characteristic p>0p>0, a flat lifting of the Frobenius twist X=FkXX^{\prime}=F_{k}^{*}X to W2(k)W_{2}(k) induces a splitting of the truncation of the de Rham complex in degrees [0,1][0,1], i.e. an isomorphism in the derived category

𝒪XΩX/k1[1]\xlongrightarrowτ1(FX/k,ΩX/k).\mathcal{O}_{X^{\prime}}\oplus\Omega^{1}_{X^{\prime}/k}[-1]\xlongrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\tau_{\leq 1}(F_{X/k,*}\Omega^{\bullet}_{X/k}).

Using the algebra structure of the de Rham complex, they further show that it induces an isomorphism

i<pΩX/ki[i]\xlongrightarrowτ<p(FX/k,ΩX/k).\bigoplus_{i<p}\Omega^{i}_{X^{\prime}/k}[-i]\xlongrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\tau_{<p}(F_{X/k,*}\Omega^{\bullet}_{X/k}).

With their method, it is unclear if one could extend this further to an isomorphism between i0ΩX/ki[i]\bigoplus_{i\geq 0}\Omega^{i}_{X^{\prime}/k}[-i] and FX/k,ΩX/kF_{X/k,*}\Omega^{\bullet}_{X/k} if dimXp\dim X\geq p, i.e. whether the de Rham complex ΩX/k\Omega^{\bullet}_{X/k} is decomposable. As a step further, Deligne and Illusie prove using duality that this is the case if dimX=p\dim X=p.

It is as of today an open problem whether there exists a smooth XX over kk liftable to W2(k)W_{2}(k), necessarily of dimension dimX>p\dim X>p, for which the de Rham complex is not decomposable. In this paper, as a small contribution to this question, we investigate the ways in which the truncation τ1(FX/k,ΩX/k)\tau_{\leq 1}(F_{X/k,*}\Omega^{\bullet}_{X/k}) determines the truncations τ[a,b](FX/k,ΩX/k)\tau_{[a,b]}(F_{X/k,*}\Omega^{\bullet}_{X/k}). Our first result is the following:

Theorem 1.1.

Let XX be a smooth scheme over a perfect field kk of characteristic p>0p>0 which is liftable to W2(k)W_{2}(k). Then the truncations

τ[a,b](FX/k,ΩX/k)\tau_{[a,b]}(F_{X/k,*}\Omega^{\bullet}_{X/k})

are decomposable for ab<a+p1a\leq b<a+p-1 when p>2p>2, and for aba+1a\leq b\leq a+1 when p=2p=2.

The above result immediately implies that in the conjugate spectral sequence

(1.1) E2ij=Hi(X,ΩX/kj)Hi+j(X,ΩX/k)E_{2}^{ij}=H^{i}(X^{\prime},\Omega^{j}_{X^{\prime}/k})\quad\Rightarrow\quad H^{i+j}(X,\Omega^{\bullet}_{X/k})

the differentials drijd_{r}^{ij} are zero for 2r<p2\leq r<p when p>2p>2, and for r=2r=2 when p=2p=2. As a sample corollary, we obtain the following criterion for degeneration of spectral sequences.

Corollary 1.2.

For XX as in Theorem 1.1, suppose that

Hi(X,ΩX/kj)=0for |ij|p.H^{i}(X,\Omega^{j}_{X/k})=0\quad\text{for }\quad|i-j|\geq p.

Then the conjugate spectral sequence (1.1) degenerates. If moreover XX is proper over kk, then the Hodge to de Rham spectral sequence

E1ij=Hj(X,ΩX/ki)Hi+j(X,ΩX/k)E^{ij}_{1}=H^{j}(X,\Omega^{i}_{X/k})\quad\Rightarrow\quad H^{i+j}(X,\Omega^{\bullet}_{X/k})

degenerates as well.

1.2. Truncations of the de Rham complex

Our methods give information about truncations of the de Rham complex without assuming liftability modulo p2p^{2}. Our results in this direction are the strongest and most explicit for truncations in two consecutive degrees. Namely, for a general smooth XX over kk (not necessarily liftable to W2(k)W_{2}(k)) and for q1q\geq 1, the truncated complex τ[q1,q](FX/k,ΩX/k)\tau_{[q-1,q]}(F_{X/k,*}\Omega^{\bullet}_{X/k}) can be described as the mapping fiber of δq[q]\delta^{q}[-q] for a map

δq:ΩX/kqΩX/kq1[2]i.e.δqExt2(ΩX/kq,ΩX/kq1)\delta^{q}\colon\Omega^{q}_{X^{\prime}/k}\longrightarrow\Omega^{q-1}_{X^{\prime}/k}[2]\quad\text{i.e.}\quad\delta^{q}\in\operatorname{Ext}^{2}(\Omega^{q}_{X^{\prime}/k},\Omega^{q-1}_{X^{\prime}/k})

which is the “cup product” with the negative of the deformation obstruction class

(1.2) δ1=obs(X/k/W2(k))Ext2(ΩX/k1,𝒪X)H2(X,TX/k)\delta^{1}=-\operatorname{obs}(X^{\prime}/k/W_{2}(k))\quad\in\quad\operatorname{Ext}^{2}(\Omega^{1}_{X^{\prime}/k},\mathcal{O}_{X^{\prime}})\simeq H^{2}(X^{\prime},T_{X^{\prime}/k})

to the existence of a lifting of XX^{\prime} to W2(k)W_{2}(k) (see Corollary 4.3). The result in particular implies that the two-term truncation τ[q1,q](FX/k,ΩX/k)\tau_{[q-1,q]}(F_{X/k,*}\Omega^{\bullet}_{X/k}) is decomposable if XX^{\prime} lifts to W2(k)W_{2}(k), and yields a description the differentials on the second page of the conjugate spectral sequence — answering a natural question of Katz.

Theorem 1.3 (see Corollary 4.4).

In the above situation, the differential

d2ij:Hi(X,ΩX/kj)Hi+2(X,ΩX/kj1)d_{2}^{ij}\colon H^{i}(X^{\prime},\Omega^{j}_{X^{\prime}/k})\longrightarrow H^{i+2}(X^{\prime},\Omega^{j-1}_{X^{\prime}/k})

in the conjugate spectral sequence (1.1) is induced by the cup product with the negative of the obstruction class obs(X/k/W2(k))\operatorname{obs}(X^{\prime}/k/W_{2}(k)).

Deligne and Illusie [DI87, §3] define the gerbe of splittings scK\operatorname{sc}K of a two-term complex KK, and relate the gerbe of splittings of τ1(FX/k,ΩX/k)\tau_{\leq 1}(F_{X/k,*}\Omega^{\bullet}_{X/k}) to the gerbe of liftings of XX^{\prime} to W2(k)W_{2}(k). This provides a “categorification” of the equality (1.2). In the same vein, for p>2p>2, our description of the class of τ[q1,q](FX/k,ΩX/k)\tau_{[q-1,q]}(F_{X/k,*}\Omega^{\bullet}_{X/k}) can be upgraded to a morphism of gerbes (see Theorem 3.9)

q:sc(τ1(FX/k,ΩX/k))sc(τ[q1,q](FΩX/k)).\wedge^{q}\colon\operatorname{sc}(\tau_{\leq 1}(F_{X/k,*}\Omega^{\bullet}_{X/k}))\longrightarrow\operatorname{sc}(\tau_{[q-1,q]}(F_{*}\Omega^{\bullet}_{X/k})).

Let us now discuss longer truncations of the de Rham complex. The assertion of Theorem 1.1 is subsumed by a recent beautiful observation of Drinfeld [Dri20, §5.12.1] (a proof will appear in Bhatt and Lurie’s forthcoming [BL21]): a lifting of XX^{\prime} to W2(k)W_{2}(k) induces a μp\mu_{p}-action on the de Rham complex FX/k,ΩX/kF_{X/k,*}\Omega^{\bullet}_{X/k}, which one can use to show that the truncations τ[qp+1,q](FX/k,ΩX/k)\tau_{[q-p+1,q]}(F_{X/k,*}\Omega^{\bullet}_{X/k}) are decomposable for all qq (even more recently, Li and Mondal [LM21] found an independent proof). However, the method of proof of Theorem 1.1 is completely different and provides interesting information even if XX is not liftable to W2(k)W_{2}(k). It is deduced from the following result (when p>2p>2) and Corollary 4.3 (when p=2p=2) alluded to above.

Theorem 1.4.

Let XX be a smooth scheme over a perfect field kk of characteristic p>0p>0, let qq be an integer, and let m<pm<p. One then has an isomorphism in the derived category of XX^{\prime}:

τ[qm,q](FX/k,ΩX/k)τqm(LΓq(τ1FX/k,ΩX/k))\tau_{[q-m,q]}(F_{X/k,*}\Omega^{\bullet}_{X/k})\quad\simeq\quad\tau_{\geq q-m}(L\Gamma^{q}(\tau_{\leq 1}F_{X/k,*}\Omega^{\bullet}_{X/k}))

where LΓqL\Gamma^{q} is the derived qq-th divided power.

1.3. Abstract Koszul complexes

The proof of Theorem 1.4 has very little to do with algebraic geometry. To state the main technical result behind it, we need the notion of an abstract Koszul complex (Definition 2.1), which is a certain commutative differential graded algebra (cdga) KK in a ringed topos for which the multiplication induces isomorphisms

q1(K)\xlongrightarrowq(K)for allq0.\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}(K)\xlongrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\operatorname{\cal{H}}^{q}(K)\quad\text{for all}\quad q\geq 0.

Thanks to the Cartier isomorphism, the de Rham complexes FX/k,ΩX/kF_{X/k,*}\Omega^{\bullet}_{X/k} in characteristic p>0p>0 are examples of such, and hence the result below immediately implies Theorem 1.4.

Theorem 1.5 (see Theorem 2.8).

Let KK be an abstract Koszul complex in a ringed topos (X,𝒪)(X,\mathcal{O}) satisfying the flatness condition (2.1), and let qm1q\geq m\geq 1 be integers such that m!m! is invertible in 𝒪\mathcal{O}. Suppose that either q=mq=m or that m+1m+1 is a nonzerodivizor in 𝒪\mathcal{O}. Then there exists an isomorphism in the derived category

(1.3) τqm(LΓq(τ1(K)))\xlongrightarrowτ[qm,q](K).\tau_{\geq q-m}\left(L\Gamma^{q}(\tau_{\leq 1}(K))\right)\xlongrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\tau_{[q-m,q]}(K).

In (1.3), LΓqL\Gamma^{q} is again the derived qq-th divided power, and the source of the map can be more concretely realized as τqm\tau_{\geq q-m} of the Koszul complex

i(𝒵1𝒦)Γ𝓆𝒾(𝒦0)degree iq1(𝒵1𝒦)𝒦0𝓆(𝒵1𝒦)degree 𝓆0\cdots\to\underbrace{\operatorname{\mbox{\large$\wedge$}}^{i}(\cal{Z}^{1}K)\otimes\Gamma^{q-i}(K^{0})}_{\text{degree $i$}}\to\cdots\to\operatorname{\mbox{\large$\wedge$}}^{q-1}(\cal{Z}^{1}K)\otimes K^{0}\to\underbrace{\operatorname{\mbox{\large$\wedge$}}^{q}(\cal{Z}^{1}K)}_{\text{degree }q}\to 0\to\cdots

For m=1m=1 (and assuming that 22 is a nonzerodivisor), we again obtain more refined information regarding τ[q1,q]K\tau_{[q-1,q]}K, including the differentials on the second page of the spectral sequence

E2ij=Hi(X,j(K))Hi+j(X,K)(see Corollary 4.3).E_{2}^{ij}=H^{i}(X,\operatorname{\cal{H}}^{j}(K))\quad\Rightarrow\quad H^{i+j}(X,K)\qquad\text{(see Corollary~{}\ref{cor:delta-q-formula})}.

As observed by Kato [Kat89], logarithmic de Rham complexes are abstract Koszul complexes, and hence Theorem 1.1 works also in the log case. The inspiration for Theorem 2.8 came from the result of Steenbrink [Ste95, §2.8] describing the nearby cycle complex RΨ𝐐R\Psi\mathbf{Q} for a complex semistable degeneration in terms of the logarithmic structure; see also [AO20, §4]. It is an interesting question whether Steenbrink’s result can be extended to work with integral coefficients; the nearby cycles RΨ𝐙R\Psi\mathbf{Z} are co-connective EE_{\infty}-algebra versions of abstract Koszul complexes, but we do not know whether they admit cdga models (see Remark 2.10 and Example 2.3). An affirmative answer would give an application unrelated to the Deligne–Illusie theorem, refining [AO20, Theorem 4.2.2(1)], providing a description of the two-step truncations τ[q1,q]\tau_{[q-1,q]} of certain logarithmic nearby cycle complexes.

1.4. The case p=2p=2 (Theorem 4.1)

The description of the truncations τ[q1,q](FX/k,ΩX/k)\tau_{[q-1,q]}(F_{X/k,*}\Omega^{\bullet}_{X/k}) and its corollary, Theorem 1.3, can be deduced from the “abstract Koszul complex” machinery and Theorem 1.4, but only for p>2p>2. In contrast, the assertion of Theorem 1.5 is vacuous if 2𝒪=02\cdot\mathcal{O}=0. Accordingly, the computation of the class of τ[q1,q]FX/k,ΩX/k\tau_{[q-1,q]}F_{X/k,*}\Omega^{\bullet}_{X/k}, occupying the entire Section 4 is much harder in the case p=2p=2, and uses more information about the de Rham complex than merely its abstract Koszul complex structure. For this technical point, we highlight the passage from (4.5) to (4.6).

It could be worthwile to extend the methods used in the case of p=2p=2 in order to “compute” the truncations τ[qp+1,p]\tau_{[q-p+1,p]} in pp consecutive degrees, and it would be interesting to extract the exact abstract properties of the de Rham complex in positive characteristic needed for the proof. Its relationship with the aforementioned result of Drinfeld, Bhatt–Lurie, and Li–Mondal remains elusive.

The results concerning the truncations τ[q1,q](FX/k,ΩX/k)\tau_{[q-1,q]}(F_{X/k,*}\Omega^{\bullet}_{X/k}) and Theorem 1.3, including the case p=2p=2, presented here are due to the second author and appeared in his 2007 Ph.D. thesis. After the first author proved Theorem 1.4, the authors decided to publish their results together.

1.5. Application to FF-split (p+1)(p+1)-folds

As an illustration of this circle of ideas, using the refinement of the Deligne–Illusie theorem due to Drinfeld, Bhatt–Lurie, and Li–Mondal, we prove in Appendix A that the Kodaira–Akizuki–Nakano vanishing theorem and the degeneration of the Hodge to de Rham spectral sequence both hold for FF-split (p+1)(p+1)-folds in characteristic pp.

Acknowledgements

The first author is grateful to Luc Illusie, Shizhang Li, Arthur Ogus, and Vadim Vologodsky for useful comments and stimulating discussions. The second author thanks Nicholas Katz for raising the question that we answer in Theorem 1.3, and Pierre Deligne and Luc Illusie for helpful comments.

The first author was supported by NCN OPUS grant number UMO-2015/17/B/ST1/02634.

Notation

If KK is a cochain complex in an abelian category, we write 𝒵𝒾𝒦=ker(𝒹:𝒦𝒾𝒦𝒾+1)\cal{Z}^{i}K=\ker(d\colon K^{i}\to K^{i+1}), 𝒾𝒦=im(𝒹:𝒦𝒾1𝒦𝒾)\cal{B}^{i}K=\operatorname{im}(d\colon K^{i-1}\to K^{i}), and i(K)=𝒵𝒾𝒦/𝒾𝒦\operatorname{\cal{H}}^{i}(K)=\cal{Z}^{i}K/\cal{B}^{i}K, and denote by τb(K)\tau_{\leq b}(K) the subcomplex

τb(K)=[Kb1𝒵𝒷𝒦0],\tau_{\leq b}(K)=\left[\cdots\to K^{b-1}\to\cal{Z}^{b}K\to 0\to\cdots\right],

by τa(K)\tau_{\geq a}(K) the quotient K/τa(K)K/\tau_{\leq a}(K), and define τ[a,b](K)=τaτb(K)\tau_{[a,b]}(K)=\tau_{\geq a}\tau_{\leq b}(K). We call KK decomposable if it is isomorphic in the derived category to the complex with zero differential i(K)[i]\bigoplus\operatorname{\cal{H}}^{i}(K)[-i].

A commutative differential graded algebra (cdga) is an associative graded ring K=n𝐙KnK=\bigoplus_{n\in\mathbf{Z}}K^{n} which is graded-commutative (i.e. xy=(1)mnyxxy=(-1)^{mn}yx for xKmx\in K^{m}, yKn{y\in K^{n}}), endowed with a differential d:KKd\colon K\to K mapping KnK^{n} to Kn+1K^{n+1} and satisfying d(xy)=dxy+(1)nxdyd(xy)=dx\cdot y+(-1)^{n}x\cdot dy for xKnx\in K^{n}. We say that KK is coconnective if Kn=0K^{n}=0 for n<0n<0.

2. Abstract Koszul complexes

2.1. Definition and examples

We work in a ringed topos (X,𝒪)(X,\mathcal{O}).

Definition 2.1 (Abstract Koszul complex).

A coconnective commutative differential graded 𝒪\mathcal{O}-algebra KK is called an abstract Koszul complex if the following conditions are satisfied:

  1. (i)

    𝒪0(K)\mathcal{O}\to\operatorname{\cal{H}}^{0}(K) is an isomorphism,

  2. (ii)

    For every q1q\geq 1, the multiplication map 1(K)qq(K)\operatorname{\cal{H}}^{1}(K)^{\otimes q}\to\operatorname{\cal{H}}^{q}(K) factors through an isomorphism

    μq:q1(K)\xlongrightarrowq(K).\mu^{q}:\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}(K)\xlongrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\operatorname{\cal{H}}^{q}(K).
Example 2.2 (De Rham complex in characteristic p>0p>0).

Let XX be a smooth scheme over a perfect field kk of characteristic p>0p>0, and let FX/k:XXF_{X/k}\colon X\to X^{\prime} be its relative Frobenius. Let K=FX/k,ΩX/kK=F_{X/k,*}\Omega^{\bullet}_{X/k} be the de Rham complex, treated as a cdga over 𝒪X\mathcal{O}_{X^{\prime}}. Then the Cartier isomorphisms

C:i(FX/k,ΩX/k)\xlongrightarrowΩX/kiC\colon\operatorname{\cal{H}}^{i}(F_{X/k,*}\Omega^{\bullet}_{X/k})\xlongrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\Omega^{i}_{X^{\prime}/k}

are multiplicative, and hence KK is an abstract Koszul complex over (X,𝒪X)(X^{\prime},\mathcal{O}_{X^{\prime}}).

More generally, if f:(X,𝒳)(𝒮,𝒮)f\colon(X,\cal{M}_{X})\to(S,\cal{M}_{S}) is a morphism of fine log schemes over 𝐅p\mathbf{F}_{p} which is smooth and of Cartier type, then the log de Rham complex FX/S,Ω(X,𝒳)/(𝒮,𝒮)F_{X/S,*}\Omega^{\bullet}_{(X,\cal{M}_{X})/(S,\cal{M}_{S})} is an abstract Koszul complex [Kat89, Theorem 4.12].

Example 2.3 (Nearby cycle complexes, see e.g. [Ste95, §2]).

Let XX be a complex manifold and let D=DαD=\bigcup D_{\alpha} be a divisor with simple normal crossings on XX. Let j:U=XDXj\colon U=X\setminus D\hookrightarrow X be the complementary open immersion, and let K=Rj𝐐UK=Rj_{*}\mathbf{Q}_{U}. Since we are working with rational coefficients, we can find a cdga model for KK (e.g. [KM95, Part II, Corollary 1.5]). The purity theorem implies that

Rij𝐐U=i(𝐐Dα),R^{i}j_{*}\mathbf{Q}_{U}=\operatorname{\mbox{\large$\wedge$}}^{i}\left(\bigoplus\mathbf{Q}_{D_{\alpha}}\right),

and hence any cdga model of KK is an abstract Koszul complex over (X,𝐐X)(X,\mathbf{Q}_{X}). Moreover, one has an isomorphism in the derived category [Ste95, Lemma 2.7] (see also [AO20, §4])

τ1(R1j𝐙U)[𝒪Xexpgp],\tau_{\leq 1}(R^{1}j_{*}\mathbf{Z}_{U})\quad\simeq\quad\left[\mathcal{O}_{X}\xrightarrow{\exp}\cal{M}^{\rm gp}\right],

where gp\cal{M}^{\rm gp} is the sheaf of meromorphic functions without zeros or poles on UU. Variants of this contruction exist for the nearby cycle complexes RΨ𝐐R\Psi\mathbf{Q} for a semistable degeneration over a disc, and there exist analogs in \ell-adic étale cohomology.

In the following, we will work with abstract Koszul complexes satisfying the additional flatness condition:

(2.1) The 𝒪\mathcal{O}-modules K0K^{0}, 1𝒦\cal{B}^{1}K, 𝒵1𝒦\cal{Z}^{1}K, and 1(K)\operatorname{\cal{H}}^{1}(K) are flat.

In particular, this implies that the modules q(K)qq(K)\operatorname{\cal{H}}^{q}(K)\simeq\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{q}(K) are flat for all q0q\geq 0. The above condition is satisfied in the situation of Examples 2.2 and 2.3.

2.2. Koszul complexes

Our goal is to show that to a certain extent, the underlying complex of an abstract Koszul complex satisfying the flatness condition (2.1) is determined by its truncation in degrees 1\leq 1 (Theorem 2.8). We achieve this using the notion of the Koszul complex of a map u:PQu:P\to Q, see [Ill71, Chapitre I, §4.3] and [KS04, §1.1–1.2].

Recall first that if MM and NN are 𝒪\mathcal{O}-modules over, then for every q0q\geq 0 there is a natural decomposition of the divided (resp. exterior) power

Γq(MN)=a+b=qΓaMΓbN(resp. q(MN)=a+b=qaMbN)\Gamma^{q}(M\oplus N)=\bigoplus_{a+b=q}\Gamma^{a}M\otimes\Gamma^{b}N\qquad\text{(resp.\ $\operatorname{\mbox{\large$\wedge$}}^{q}(M\oplus N)=\bigoplus_{a+b=q}\operatorname{\mbox{\large$\wedge$}}^{a}M\otimes\operatorname{\mbox{\large$\wedge$}}^{b}N$)}

In what follows, we will use the comultiplication maps

ηq:ΓqMMΓq1M(resp. ηq:qM(q1M)M)\eta^{q}:\Gamma^{q}M\longrightarrow M\otimes\Gamma^{q-1}M\qquad\text{(resp.\ $\eta^{q}:\operatorname{\mbox{\large$\wedge$}}^{q}M\longrightarrow(\operatorname{\mbox{\large$\wedge$}}^{q-1}M)\otimes M)$}

obtained as the composition of Γq\Gamma^{q} (resp. q\operatorname{\mbox{\large$\wedge$}}^{q}) of the diagonal map MMMM\to M\oplus M and the projection to the (a,b)=(1,q1)(a,b)=(1,q-1)-part (resp. (a,b)=(q1,1)(a,b)=(q-1,1)-part) in the above decomposition of Γq(MM)\Gamma^{q}(M\oplus M) (resp. of q(MM)\operatorname{\mbox{\large$\wedge$}}^{q}(M\oplus M)). Explicitly, we have

ηq(x1[e1]xr[er])\displaystyle\eta^{q}(x_{1}^{[e_{1}]}\cdot\ldots\cdot x_{r}^{[e_{r}]}) =i=1rxi(x1[e1]xi[ei1]xr[er]),\displaystyle=\sum_{i=1}^{r}x_{i}\otimes(x_{1}^{[e_{1}]}\cdot\ldots\cdot x_{i}^{[e_{i}-1]}\cdot\ldots\cdot x_{r}^{[e_{r}]}), (e1++er=q)\displaystyle(e_{1}+\cdots+e_{r}=q)
ηq(x1xq)\displaystyle\eta^{q}(x_{1}\wedge\ldots\wedge x_{q}) =i=1q(1)i1x1xi^xqxi.\displaystyle=\sum_{i=1}^{q}(-1)^{i-1}x_{1}\wedge\ldots\wedge\widehat{x_{i}}\wedge\ldots\wedge x_{q}\otimes x_{i}.

Sometimes we omit the superscript qq when it is clear from the context.

Definition 2.4 (Koszul complex Kosq(u)\operatorname{Kos}^{q}(u)).

Let u:PQu:P\to Q be a map of 𝒪\mathcal{O}-modules, and let q0q\geq 0 be an integer. Then the qq-th Koszul complex Kosq(u)\operatorname{Kos}^{q}(u) is the cochain complex whose ii-th term is

Kosq(u)i=i(Q)Γqi(P),\operatorname{Kos}^{q}(u)^{i}=\operatorname{\mbox{\large$\wedge$}}^{i}(Q)\otimes\Gamma^{q-i}(P),

with differential d:Kosq(u)iKosq(u)i+1d\colon\operatorname{Kos}^{q}(u)^{i}\to\operatorname{Kos}^{q}(u)^{i+1} defined as the composition

i(Q)Γqi(P)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{i}(Q)\otimes\Gamma^{q-i}(P)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1η\scriptstyle{1\otimes\eta}i(Q)PΓqi1(P)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{i}(Q)\otimes P\otimes\Gamma^{q-i-1}(P)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1u1\scriptstyle{1\otimes u\otimes 1}i(Q)QΓqi1(P)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{i}(Q)\otimes Q\otimes\Gamma^{q-i-1}(P)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{\wedge\otimes 1}i+1(Q)Γqi1(P).\textstyle{\operatorname{\mbox{\large$\wedge$}}^{i+1}(Q)\otimes\Gamma^{q-i-1}(P).}

Concretely, with e1++er=qie_{1}+\cdots+e_{r}=q-i, x1,,xrPx_{1},\ldots,x_{r}\in P, and yiQy\in\operatorname{\mbox{\large$\wedge$}}^{i}Q:

d(yx1[e1]xr[er])=j=1ryu(xj)(x1[e1]xj[ej1]xr[er]).d\left(y\otimes x_{1}^{[e_{1}]}\cdot\ldots\cdot x_{r}^{[e_{r}]}\right)=\sum_{j=1}^{r}y\wedge u(x_{j})\otimes\left(x_{1}^{[e_{1}]}\cdot\ldots\cdot x_{j}^{[e_{j}-1]}\cdot\ldots\cdot x_{r}^{[e_{r}]}\right).
Proposition 2.5.

Let u:PQu\colon P\to Q be a map of flat 𝒪\mathcal{O}-modules, and let F(u)=[P𝑢Q]F(u)=[P\xrightarrow{u}Q] be the two-term cochain complex with PP in degree zero (the mapping fiber). There exist natural isomorphisms in the derived category

Kosq(u)LΛq(F[1])[q]LΓq(F)\operatorname{Kos}^{q}(u)\quad\simeq\quad L\Lambda^{q}(F[1])[-q]\quad\simeq\quad L\Gamma^{q}(F)

where LΛqL\Lambda^{q} (resp. LΓqL\Gamma^{q}) is the derived exterior (resp. divided) power.

Proof.

Combine [KS04, Corollary 1.2.7] with [Ill71, I 4.3.2.1]. ∎

Corollary 2.6.

For a map u:PQu\colon P\to Q between flat 𝒪\mathcal{O}-modules, the complex Kosq(u)\operatorname{Kos}^{q}(u), treated as an object of the derived category, depends only on F(u)=[PQ]F(u)=[P\to Q] up to quasi-isomorphism. In particular, if F(u)F(u) is decomposable, then so is Kosq(u)\operatorname{Kos}^{q}(u).

Proposition 2.7 (cf. [Ste95, Lemma 1.4]).

Let u:PQu\colon P\to Q be a map of 𝒪\mathcal{O}-modules. There exist unique arrows

i(Q)Γqi(keru)𝒵𝒾(Kos𝓆(𝓊))andα𝒾:𝒾(cok𝓊)Γ𝓆𝒾(ker𝓊)𝒾(Kos𝓆(𝓊))\operatorname{\mbox{\large$\wedge$}}^{i}(Q)\otimes\Gamma^{q-i}(\ker u)\longrightarrow\cal{Z}^{i}(\operatorname{Kos}^{q}(u))\quad\text{and}\quad\alpha^{i}:\operatorname{\mbox{\large$\wedge$}}^{i}(\operatorname{cok}u)\otimes\Gamma^{q-i}(\ker u)\longrightarrow\operatorname{\cal{H}}^{i}(\operatorname{Kos}^{q}(u))

making the following diagram commute

i(Q)Γqi(P)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{i}(Q)\otimes\Gamma^{q-i}(P)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kosq(u)i\textstyle{\operatorname{Kos}^{q}(u)^{i}}i(Q)Γqi(keru)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{i}(Q)\otimes\Gamma^{q-i}(\ker u)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒵𝒾(Kos𝓆(𝓊))\textstyle{\cal{Z}^{i}(\operatorname{Kos}^{q}(u))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i(coku)Γqi(keru)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{i}(\operatorname{cok}u)\otimes\Gamma^{q-i}(\ker u)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αi\scriptstyle{\alpha^{i}}i(Kosq(u)).\textstyle{\operatorname{\cal{H}}^{i}(\operatorname{Kos}^{q}(u)).}

Moreover, the map αi\alpha^{i} is an isomorphism if PP, QQ, keru\ker u, imu\operatorname{im}u, and coku\operatorname{cok}u are all flat.

Proof.

The first assertion is straightforward. The second is reduced as in [Ill71, I 4.3.1.6] to the case where PP, QQ, keru\ker u, imu\operatorname{im}u, and coku\operatorname{cok}u are free 𝒪\mathcal{O}-modules of finite rank. In this case, splitting the surjection QcokuQ\to\operatorname{cok}u one can write u=uu′′u=u^{\prime}\oplus u^{\prime\prime} where u:Pimuu^{\prime}\colon P\to\operatorname{im}u and u′′:0cokuu^{\prime\prime}\colon 0\to\operatorname{cok}u. The assertion then holds for uu^{\prime} (by [Ill71, 4.3.1.6]) and for u′′u^{\prime\prime} (trivially), for all qq, and then the assertion for u=uu′′u=u^{\prime}\oplus u^{\prime\prime} follows from the isomorphism [Ill71, 4.3.1.5]

Kos(u)=Kos(u)Kos(u′′)\operatorname{Kos}^{\bullet}(u)=\operatorname{Kos}^{\bullet}(u^{\prime})\otimes\operatorname{Kos}^{\bullet}(u^{\prime\prime})

where Kos(u)=q0Kosq(u)[q]\operatorname{Kos}^{\bullet}(u)=\bigoplus_{q\geq 0}\operatorname{Kos}^{q}(u)[q]. ∎

2.3. Truncations of abstract Koszul complexes

The following theorem is the main result of this section.

Theorem 2.8.

Let mm be an integer such that m!m! is invertible in 𝒪\mathcal{O}, and let qmq\geq m. Suppose that either q=mq=m, or that m+1m+1 is not a zero divisor in 𝒪\mathcal{O}. Let KK be an abstract Koszul complex on (X,𝒪)(X,\mathcal{O}) satisfying the flatness condition (2.1), and write

τ1K=[K0Z1K]\tau_{\leq 1}K=\left[K^{0}\xrightarrow{\partial}Z^{1}K\right]

for its truncation in degrees 1\leq 1. Then the multiplication maps

Kosq()i=i(𝒵1𝒦)Γ𝓆𝒾(𝒦0)=𝒾(𝒵1𝒦)Sym𝓆𝒾(𝒦0)𝒦𝒾\operatorname{Kos}^{q}(\partial)^{i}=\operatorname{\mbox{\large$\wedge$}}^{i}(\cal{Z}^{1}K)\otimes\Gamma^{q-i}(K^{0})=\operatorname{\mbox{\large$\wedge$}}^{i}(\cal{Z}^{1}K)\otimes\operatorname{Sym}^{q-i}(K^{0})\longrightarrow K^{i}

for qmiqq-m\leq i\leq q (where we can identify Γqi\Gamma^{q-i} with Symqi\operatorname{Sym}^{q-i} as qimq-i\leq m, so that (qi)!(q-i)! is invertible in 𝒪\mathcal{O}) induce a quasi-isomorphism

(2.2) τqm(LΓq(τ1(K))=τqmKosq()\xlongrightarrowτ[qm,q](K).\tau_{\geq q-m}(L\Gamma^{q}(\tau_{\leq 1}(K))=\tau_{\geq q-m}\operatorname{Kos}^{q}(\partial)\xlongrightarrow{\,\smash{\raisebox{-2.79857pt}{$\scriptstyle\sim$}}\,}\tau_{[q-m,q]}(K).
Proof.

The multiplication maps define a morphism of “naive truncations”

Kosq()qm=[\textstyle{\operatorname{Kos}^{q}(\partial)^{\geq q-m}=[\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ\scriptstyle{\mu}qm(𝒵1𝒦)Sym𝓂(𝒦0)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{q-m}(\cal{Z}^{1}K)\otimes\operatorname{Sym}^{m}(K^{0})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q(𝒵1𝒦)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{q}(\cal{Z}^{1}K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}]\textstyle{]}τq(K)qm=[\textstyle{\tau_{\leq q}(K)^{\geq q-m}=[}Kqm\textstyle{K^{q-m}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ZqK\textstyle{Z^{q}K}].\textstyle{].}

To obtain the desired morphism μ:τqmKosq()τ[qm,q](K)\mu\colon\tau_{\geq q-m}\operatorname{Kos}^{q}(\partial)\to\tau_{[q-m,q]}(K), we need to check that the map

Kosq()qmKqm\operatorname{Kos}^{q}(\partial)^{q-m}\longrightarrow K^{q-m}

takes the image of Kosq()qm1Kosq()qm\operatorname{Kos}^{q}(\partial)^{q-m-1}\to\operatorname{Kos}^{q}(\partial)^{q-m} into 𝓆𝓂𝒦=𝒹𝒦𝓆𝓂1\cal{B}^{q-m}K=dK^{q-m-1}. This is clear if q=mq=m, so suppose that (m+1)(m+1) is not a zero divisor.

Let zKosq()qmz\in\operatorname{Kos}^{q}(\partial)^{q-m} be the image of wKosq()qm1w\in\operatorname{Kos}^{q}(\partial)^{q-m-1}, and consider (m+1)w(m+1)w as an element of the submodule

qm1(𝒵1𝒦)Sym𝓂+1(𝒦0)𝓆𝓂1(𝒵1𝒦)Γ𝓂+1(𝒦0).\operatorname{\mbox{\large$\wedge$}}^{q-m-1}(\cal{Z}^{1}K)\otimes\operatorname{Sym}^{m+1}(K^{0})\quad\subseteq\quad\operatorname{\mbox{\large$\wedge$}}^{q-m-1}(\cal{Z}^{1}K)\otimes\Gamma^{m+1}(K^{0}).

Let uKqm1u\in K^{q-m-1} be the image of (m+1)w(m+1)w under the multiplication map

qm1(𝒵1𝒦)Sym𝓂+1(𝒦0)𝒦𝓆𝓂1.\operatorname{\mbox{\large$\wedge$}}^{q-m-1}(\cal{Z}^{1}K)\otimes\operatorname{Sym}^{m+1}(K^{0})\longrightarrow K^{q-m-1}.

Then du=(m+1)μ(z)du=(m+1)\mu(z) in KqmK^{q-m}, where μ(z)\mu(z) is the image of zz under the multiplication map, and hence μ(z)\mu(z) gives an (m+1)(m+1)-torsion class in qm(K)\operatorname{\cal{H}}^{q-m}(K). Since by assumption qm(K)qm1(K)\operatorname{\cal{H}}^{q-m}(K)\simeq\operatorname{\mbox{\large$\wedge$}}^{q-m}\operatorname{\cal{H}}^{1}(K) is flat and m+1m+1 is not a zero divisor, μ(z)dKmq1\mu(z)\in dK^{m-q-1} as desired.

Finally, the maps induced by μ:τqmKosq()τ[qm,q](K)\mu\colon\tau_{\geq q-m}\operatorname{Kos}^{q}(\partial)\to\tau_{[q-m,q]}(K) on cohomology can, thanks to Proposition 2.7, be identified with the maps

μi:i1(K)i(K)forqmiq,\mu^{i}\colon\operatorname{\mbox{\large$\wedge$}}^{i}\operatorname{\cal{H}}^{1}(K)\longrightarrow\operatorname{\cal{H}}^{i}(K)\quad\text{for}\quad q-m\leq i\leq q,

which are isomorphisms by assumption. ∎

Remark 2.9.

Implicit in the above proof is the subcomplex Kosq~(u)\widetilde{\operatorname{Kos}^{q}}(u) of Kosq(u)\operatorname{Kos}^{q}(u) whose ii-th term equals i(Z1K)SymqiK\operatorname{\mbox{\large$\wedge$}}^{i}(Z^{1}K)\otimes\operatorname{Sym}^{q-i}K. The two complexes agree in degrees qm\geq q-m, and more generally the quotient Kosq(u)qi/Kosq~(u)qi\operatorname{Kos}^{q}(u)^{q-i}/\widetilde{\operatorname{Kos}^{q}}(u)^{q-i} is annihilated by i!i!. This subcomplex probably does not have any “derived meaning,” (for example, it is not clear that it is decomposable if τ1(K)\tau_{\leq 1}(K) is), but its advantage is that there is a multiplication map μ:Kosq~(u)K\mu\colon\widetilde{\operatorname{Kos}^{q}}(u)\to K.

Remark 2.10.

Our proof of Theorem 2.8 makes use of an explicit model of the cdga KK. Thus, for example, if KK and KK^{\prime} are equivalent cdgas to which the theorem applies, it is not obvious whether the isomorphisms (2.2) we obtain for KK and KK^{\prime} are compatible. More importantly, it does not apply to the more general case of coconnective EE_{\infty}-algebras or cosimplicial commutative rings whose cohomology algebras satisfy axioms (i)–(ii) of Definition 2.1.

Corollary 2.11.

Let KK be an abstract Koszul complex, and let mm be such that m!m! is invertible in 𝒪\mathcal{O}. Suppose that τ1(K)\tau_{\leq 1}(K) is decomposable. Then for ab<a+ma\leq b<a+m, the complex τ[a,b](K)\tau_{[a,b]}(K) is decomposable. Moreover, the complex τm(K)\tau_{\leq m}(K) is decomposable as well.

2.4. Application to de Rham cohomology

We now establish some of the straightforward consequences for de Rham cohomology mentioned in the introduction. The remaining ones shall be established at the end of Section 4.

Proof of Theorem 1.1.

Let K=FX/k,ΩX/kK=F_{X/k,*}\Omega^{\bullet}_{X/k}. By Example 2.2, this is an abstract Koszul complex over the ringed space (X,𝒪X)(X^{\prime},\mathcal{O}_{X^{\prime}}). By [DI87, Théorème 2.1], the liftability assumption implies that the complex τ1(K)\tau_{\leq 1}(K) is decomposable. Corollary 2.11 with m=p1m=p-1 implies that τ[a,b](K)\tau_{[a,b]}(K) is decomposable for ab<a+p1a\leq b<a+p-1, as desired. ∎

Proof of Corollary 1.2.

The differentials on the ErE_{r}-page of (1.1) depend only on the truncations τ[a,b](ΩX/k)\tau_{[a,b]}(\Omega^{\bullet}_{X/k}) with ab<a+ra\leq b<a+r, and hence all differentials on the pages ErE_{r} with r<pr<p vanish. Suppose that drij0d^{ij}_{r}\neq 0, then in particular Hi(X,ΩX/kj)H^{i}(X,\Omega^{j}_{X/k}) and Hi+r(X,ΩX/kjr+1)H^{i+r}(X,\Omega^{j-r+1}_{X/k}) are both nonzero, and hence

|ij|<pand|(i+r)(jr+1)|=|(ij)+2r1|<p,|i-j|<p\quad\text{and}\quad|(i+r)-(j-r+1)|=|(i-j)+2r-1|<p,

which implies r<pr<p. Therefore (1.1) degenerates.

For XX proper over kk, one can deduce the degeneration of the Hodge to de Rham spectral sequence as in [DI87, Corollaire 2.4]. ∎

Remark 2.12.

As in [DI87, §4] and [Kat89, Theorem 4.12(2)], analogous assertions hold for a smooth and separated morphism of 𝐅p\mathbf{F}_{p}-schemes XSX\to S, or more generally for a smooth morphism of Cartier type f:(X,𝒳)(𝒮,𝒮)f\colon(X,\cal{M}_{X})\to(S,\cal{M}_{S}) between fine log schemes over 𝐅p\mathbf{F}_{p}, assuming that there exists a fine log scheme (S~,𝒮~)(\widetilde{S},\cal{M}_{\widetilde{S}}) over 𝐙/p2𝐙\mathbf{Z}/p^{2}\mathbf{Z} such that S~\widetilde{S} is flat over 𝐙/p2𝐙\mathbf{Z}/p^{2}\mathbf{Z} and a smooth lifting

f~:(X~,𝒳~)(𝒮~,𝒮~)\tilde{f}^{\prime}\colon(\widetilde{X}{}^{\prime},\cal{M}_{\widetilde{X}{}^{\prime}})\to(\widetilde{S},\cal{M}_{\widetilde{S}})

of ff^{\prime} (the base change ff under the absolute Frobenius (S,𝒮)(𝒮,𝒮)(S,\cal{M}_{S})\to(S,\cal{M}_{S})). Here 𝐅p\mathbf{F}_{p} and 𝐙/p2𝐙\mathbf{Z}/p^{2}\mathbf{Z} are given the trivial log structure.

3. Truncations in two consecutive degrees and gerbes of splittings

In the following, we make a more detailed analysis of the truncations τ[q1,q]K\tau_{[q-1,q]}K for an abstract Koszul complex KK, as well as their associated gerbes of splittings. We keep working in the category of modules in a ringed topos (X,𝒪)(X,\mathcal{O}).

3.1. First order attachment maps

For a complex KK and an integer qq, the truncation

τ[q1,q]K=[0Kq1/𝓆1𝒵𝓆𝒦0]\tau_{[q-1,q]}K=[\cdots\to 0\to K^{q-1}/\cal{B}^{q-1}\to\cal{Z}^{q}K\to 0\to\cdots]

fits inside the functorial exact triangle

q1(K)[1q]τ[q1,q]Kq(K)[q]δKq[q]q1(K)[2q]\operatorname{\cal{H}}^{q-1}(K)[1-q]\longrightarrow\tau_{[q-1,q]}K\longrightarrow\operatorname{\cal{H}}^{q}(K)[-q]\xrightarrow{\delta^{q}_{K}[-q]}\operatorname{\cal{H}}^{q-1}(K)[2-q]

yielding a morphism

δKq:q(K)q1(K)[2].\delta^{q}_{K}\colon\operatorname{\cal{H}}^{q}(K)\longrightarrow\operatorname{\cal{H}}^{q-1}(K)[2].

such that δKq[q]\delta^{q}_{K}[-q] is the unique morphism making the above triangle distinguished (see [AO20, Proposition 2.1.1]). Thus the truncation τ[q1,q]K\tau_{[q-1,q]}K is determined by the map δKq\delta^{q}_{K}, as the mapping fiber of δKq[q]\delta^{q}_{K}[-q]; it is decomposable if and only if δKq=0\delta^{q}_{K}=0. We note for future reference the effect of the shift functor on the maps δKq\delta^{q}_{K}:

(3.1) δK[p]q=(1)pδKp+q.\delta^{q}_{K[p]}=(-1)^{p}\delta^{p+q}_{K}.

The maps δKq\delta^{q}_{K} describe the differentials on the second page of the spectral sequence

E2pq=Hp(X,q(K))Hp+q(X,K).E_{2}^{pq}=H^{p}(X,\operatorname{\cal{H}}^{q}(K))\quad\Rightarrow\quad H^{p+q}(X,K).

Namely, the differential

d2pq:Hp(X,q(K))Hp+2(X,q1(K))=Hp(X,q1(K)[2])d_{2}^{pq}\colon H^{p}(X,\operatorname{\cal{H}}^{q}(K))\longrightarrow H^{p+2}(X,\operatorname{\cal{H}}^{q-1}(K))=H^{p}(X,\operatorname{\cal{H}}^{q-1}(K)[2])

is the map induced by δKq\delta^{q}_{K} on Hp(X,)H^{p}(X,-).

3.2. Gerbe of splittings

We recall the gerbe of splittings described in [DI87]. Let

K=[K0𝑑K1]K=[K^{0}\xrightarrow{d}K^{1}]

be a two-term complex (i.e. Ki=0K^{i}=0 for i0,1i\neq 0,1), and suppose that the two conditions below hold

  1. (1)

    1(K){\cal{H}}^{1}(K) is locally free of finite rank, and

  2. (2)

    the projection of K0K^{0} onto 1=im(d){\cal{B}}^{1}=\operatorname{im}(d) locally admits a section.

One then constructs the gerbe sc(K)\operatorname{sc}(K) under Hom¯(1(K),0(K))\underline{\operatorname{Hom}}(\operatorname{\cal{H}}^{1}(K),\operatorname{\cal{H}}^{0}(K)) over XX [DI87, §3.2] as the stackification of the prestack sc(K)\operatorname{sc}^{\prime}(K) whose objects are local splittings

s:1(K)K1s\colon\operatorname{\cal{H}}^{1}(K)\longrightarrow K^{1}

of the projection K1=𝒵1𝒦1(𝒦)K^{1}=\cal{Z}^{1}K\to\operatorname{\cal{H}}^{1}(K), and where morphisms sss\to s^{\prime} are maps

h:1(K)K0h\colon\operatorname{\cal{H}}^{1}(K)\longrightarrow K^{0}

such that dh=ssdh=s^{\prime}-s. The automorphisms of an object ss are then identified with Hom¯(1(K),0(K))\underline{\operatorname{Hom}}(\operatorname{\cal{H}}^{1}(K),\operatorname{\cal{H}}^{0}(K)), and this makes sc(K)\operatorname{sc}(K) into a gerbe under Hom¯(1(K),0(K))\underline{\operatorname{Hom}}(\operatorname{\cal{H}}^{1}(K),\operatorname{\cal{H}}^{0}(K)). We denote by

clscKH2(X,Hom¯(1(K),0(K)))=Ext2(1(K),0(K))=Hom(1(K),0(K)[2])\operatorname{cl}\operatorname{sc}K\quad\in\quad H^{2}(X,\underline{\operatorname{Hom}}(\operatorname{\cal{H}}^{1}(K),\operatorname{\cal{H}}^{0}(K)))=\operatorname{Ext}^{2}(\operatorname{\cal{H}}^{1}(K),\operatorname{\cal{H}}^{0}(K))=\operatorname{Hom}(\operatorname{\cal{H}}^{1}(K),\operatorname{\cal{H}}^{0}(K)[2])

the class of the gerbe scK\operatorname{sc}K. The following result relates this class to the map δK1\delta^{1}_{K} defined previously.

Lemma 3.1 ([DI87, Proposition 3.3]).

Let K=[K0K1]K=[K^{0}\to K^{1}] be a two-term complex satisfying (1) and (2) above. Then, one has the following equality in Hom(1(K),0(K)[2])\operatorname{Hom}(\operatorname{\cal{H}}^{1}(K),\operatorname{\cal{H}}^{0}(K)[2]):

clscK=δK1.\operatorname{cl}\operatorname{sc}K=-\delta^{1}_{K}.

A bit more generally, given an integer qq and a complex KK satisfying the following conditions:

  1. (1)

    Ki=0K^{i}=0 for iq1,qi\neq q-1,q,

  2. (2)

    q(K){\cal{H}}^{q}(K) is locally free of finite rank, and

  3. (3)

    the projection of Kq1K^{q-1} onto q=im(d){\cal{B}}^{q}=\operatorname{im}(d) locally admits a section.

Then we denote by sc[q1,q](K)\operatorname{sc}_{[q-1,q]}(K) the gerbe of splittings of the complex

0Kq1Kq0\cdots\to 0\to K^{q-1}\to K^{q}\to 0\to\cdots

concentrated in degrees 0 and 11 and with dd being equal to the original differential of KK, rather than (1)q1(-1)^{q-1} times that; this convention has the consequence that

clsc[q1,q](K)=(1)q1clsc(K[q1])\operatorname{cl}\operatorname{sc}_{[q-1,q]}(K)=(-1)^{q-1}\operatorname{cl}\operatorname{sc}(K[q-1])

in H2(X,Hom¯(q,q1))H^{2}(X,\underline{\operatorname{Hom}}({\cal{H}}^{q},{\cal{H}}^{q-1})). Combined with Lemma 3.1 and (3.1), this implies the following generalization of Lemma 3.1:

clsc[q1,q](K)=δKq.\operatorname{cl}\operatorname{sc}_{[q-1,q]}(K)=-\delta^{q}_{K}.

When there is no confusion as to what qq is, we simply write sc(K)\operatorname{sc}(K) for sc[q1,q](K)\operatorname{sc}_{[q-1,q]}(K).

3.3. Truncated Koszul complexes

Let K=[K0𝑑K1]K=[K^{0}\xrightarrow{d}K^{1}] be a two-term complex of modules over (𝒳,𝒪)(\mathcal{X},\mathcal{O}), and let q1q\geq 1. Using the Koszul complex, one can build another two-term complex, concentrated in degrees [q1,q][q-1,q]:

τq1(Kosq(d))=[0q1K1K0d(q2K1Γ2K0)qK10].\tau_{\geq q-1}(\operatorname{Kos}^{q}(d))=\left[\cdots\to 0\to\frac{\operatorname{\mbox{\large$\wedge$}}^{q-1}K^{1}\otimes K^{0}}{d(\operatorname{\mbox{\large$\wedge$}}^{q-2}K^{1}\otimes\Gamma^{2}K^{0})}\to\operatorname{\mbox{\large$\wedge$}}^{q}K^{1}\to 0\to\cdots\right].

By Proposition 2.7 we have morphisms

(3.2) αq:q1(K)q(Kosq(d))andαq1:(q11(K))0(K)q1(Kosq(d))\alpha^{q}\colon\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}(K)\longrightarrow\operatorname{\cal{H}}^{q}(\operatorname{Kos}^{q}(d))\quad\text{and}\quad\alpha^{q-1}\colon(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1}(K))\otimes\operatorname{\cal{H}}^{0}(K)\longrightarrow\operatorname{\cal{H}}^{q-1}(\operatorname{Kos}^{q}(d))

which are isomorphisms if K0K^{0}, K1K^{1}, 0(K)\operatorname{\cal{H}}^{0}(K), 1(K)\operatorname{\cal{H}}^{1}(K) are flat. The following result describes the maps δKosq(d)q\delta^{q}_{\operatorname{Kos}^{q}(d)} and hence the truncation τq1(Kosq(d))\tau_{\geq q-1}(\operatorname{Kos}^{q}(d)).

Proposition 3.2.

Let K=[K0𝑑K1]K=[K^{0}\xrightarrow{d}K^{1}] be a two-term complex and let q1q\geq 1. Suppose that K0K^{0}, K1K^{1}, 0(K)\operatorname{\cal{H}}^{0}(K), and 1(K)\operatorname{\cal{H}}^{1}(K) are flat. Then the following diagram is commutative

q1(K)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ηq\scriptstyle{\eta^{q}}αq\scriptstyle{\alpha^{q}}(q11(K))1(K)\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1}(K))\otimes\operatorname{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idδK1\scriptstyle{\mathrm{id}\otimes\delta^{1}_{K}}(q11(K))0(K)[2]\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1}(K))\otimes\operatorname{\cal{H}}^{0}(K)[2]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αq1\scriptstyle{\alpha^{q-1}}q(Kosq(d))\textstyle{\operatorname{\cal{H}}^{q}(\operatorname{Kos}^{q}(d))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δKosq(d)q\scriptstyle{\delta^{q}_{\operatorname{Kos}^{q}(d)}}q1(Kosq(d))[2].\textstyle{\operatorname{\cal{H}}^{q-1}(\operatorname{Kos}^{q}(d))[2].}

In other words, using the identifications (3.2), we have the equality

δKosq(d)q=(idδK1)ηq\delta^{q}_{\operatorname{Kos}^{q}(d)}=(\mathrm{id}\otimes\delta^{1}_{K})\circ\eta^{q}

of maps q1(K)(q11(K))0(K)[2]\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}(K)\to(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1}(K))\otimes\operatorname{\cal{H}}^{0}(K)[2].

Proof.

Let us abbreviate i(K)\operatorname{\cal{H}}^{i}(K) to i\operatorname{\cal{H}}^{i}. We first check that the two-term complexes τq1Kosq(d)\tau_{\geq q-1}\operatorname{Kos}^{q}(d) and ((q11)K)[1q]((\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1})\otimes K)[1-q] form the middle square inside a commutative diagram with exact rows

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(q11)0\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1})\otimes\operatorname{\cal{H}}^{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathrm{id}}(q1K1)K0d((q2K1)Γ2(K0))\textstyle{{\displaystyle\frac{(\operatorname{\mbox{\large$\wedge$}}^{q-1}K^{1})\otimes K^{0}}{d((\operatorname{\mbox{\large$\wedge$}}^{q-2}K^{1})\otimes\Gamma^{2}(K^{0}))}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}β\scriptstyle{\beta}qK1\textstyle{\operatorname{\mbox{\large$\wedge$}}^{q}K^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha}q1\textstyle{\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ηq\scriptstyle{\eta^{q}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(q11)0\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1})\otimes\operatorname{\cal{H}}^{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(q11)K0\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1})\otimes K^{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id(1)q1d\scriptstyle{\mathrm{id}\otimes(-1)^{q-1}d}(q11)K1\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1})\otimes K^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(q11)1\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1})\otimes\operatorname{\cal{H}}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

We define the maps α\alpha and β\beta as follows. The map β\beta is uniquely determined by

β(z1zq1w mod d(q2K1Γ2(K0)))=[z1][zq]w.\beta(z_{1}\wedge\ldots\wedge z_{q-1}\otimes w\,\,\text{ mod }\,\,d(\operatorname{\mbox{\large$\wedge$}}^{q-2}K^{1}\otimes\Gamma^{2}(K^{0})))=[z_{1}]\wedge\ldots\wedge[z_{q}]\otimes w.

It is well-defined because an elements of the form

d(z1zq1w[2])\displaystyle d(z_{1}\wedge\ldots\wedge z_{q-1}\otimes w^{[2]}) =z1zq1dww,or\displaystyle=z_{1}\wedge\ldots\wedge z_{q-1}\wedge dw\otimes w,\qquad\text{or}
d(z1zq1wv)\displaystyle d(z_{1}\wedge\ldots\wedge z_{q-1}\otimes wv) =z1zq1dwv+z1zq1dvw.\displaystyle=z_{1}\wedge\ldots\wedge z_{q-1}\wedge dw\otimes v+z_{1}\wedge\ldots\wedge z_{q-1}\wedge dv\otimes w.

are sent to zero, since [dw]=0=[dv][dw]=0=[dv]. The map α\alpha is the composition

qK1ηq(q1K1)K1proj.id(q11)K1.\operatorname{\mbox{\large$\wedge$}}^{q}K^{1}\xrightarrow{\eta^{q}}(\operatorname{\mbox{\large$\wedge$}}^{q-1}K^{1})\otimes K^{1}\xrightarrow{\text{proj.}\otimes\mathrm{id}}(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1})\otimes K^{1}.

The commutativity of the left and rightmost squares is trivial to check. To see that the middle square commutes, we take (the class of) z1zq1w(q1K1)K0z_{1}\wedge\ldots\wedge z_{q-1}\otimes w\in(\operatorname{\mbox{\large$\wedge$}}^{q-1}K^{1})\otimes K^{0}, and compute

α(d(z1zq1w))\displaystyle\alpha(d(z_{1}\wedge\ldots\wedge z_{q-1}\otimes w)) =α(z1zq1dw)\displaystyle=\alpha(z_{1}\wedge\ldots\wedge z_{q-1}\wedge dw)
=i=1q(1)i1[z1][zi]^[zq1][dw]zi+(1)q1[z1][zq1]dw\displaystyle=\sum_{i=1}^{q}(-1)^{i-1}[z_{1}]\wedge\ldots\widehat{[z_{i}]}\ldots\wedge[z_{q-1}]\wedge[dw]\otimes z_{i}+(-1)^{q-1}[z_{1}]\wedge\ldots\wedge[z_{q-1}]\otimes dw
=(1)q1[z1][zq1]dw=(1)q1(idd)(β(z1zq1w)).\displaystyle=(-1)^{q-1}[z_{1}]\wedge\ldots\wedge[z_{q-1}]\otimes dw=(-1)^{q-1}(\mathrm{id}\otimes d)(\beta(z_{1}\wedge\ldots\wedge z_{q-1}\otimes w)).

Now, thanks to (3.1), our commutative diagram of complexes translates into a commutative square in the derived category

q1\textstyle{\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δKosq(d)q\scriptstyle{\delta^{q}_{\operatorname{Kos}^{q}(d)}}ηq\scriptstyle{\eta^{q}}(q11)0[2]\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1})\otimes\operatorname{\cal{H}}^{0}[2]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\mathrm{id}}(q1)1\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1})\otimes\operatorname{\cal{H}}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idδK1\scriptstyle{\mathrm{id}\otimes\delta^{1}_{K}}(q11)0[2]\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1})\otimes\operatorname{\cal{H}}^{0}[2]}

This implies the required assertion. ∎

3.4. Two-term truncations of abstract Koszul complexes

The following result relates the maps δKq\delta^{q}_{K} and δK1\delta^{1}_{K} for a cdga KK.

Proposition 3.3.

Suppose 22 is a nonzerodivisor in 𝒪\mathcal{O}. Let KK be a coconnective commutative differential graded algebra such that K0K^{0}, 𝒵1𝒦\cal{Z}^{1}K, 0(K)\operatorname{\cal{H}}^{0}(K), 1(K)\operatorname{\cal{H}}^{1}(K). Let q1q\geq 1 be an integer such that q1(K)\operatorname{\cal{H}}^{q-1}(K) is flat. Then, the following diagram commutes

q1(K)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}mult.ηq\scriptstyle{\eta^{q}}(q11(K))1(K)\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1}(K))\otimes\operatorname{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idδK1\scriptstyle{\mathrm{id}\otimes\delta^{1}_{K}}(q11(K))0(K)[2]\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1}(K))\otimes\operatorname{\cal{H}}^{0}(K)[2]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}mult.q(K)\textstyle{\operatorname{\cal{H}}^{q}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δKq\scriptstyle{\delta^{q}_{K}}q1(K)[2].\textstyle{\operatorname{\cal{H}}^{q-1}(K)[2].}
Proof.

Write τ1K=[K0𝒵1𝒦]\tau_{\leq 1}K=[K^{0}\xrightarrow{\partial}\cal{Z}^{1}K]. The proof of Theorem 2.8 (with m=1m=1) provides a morphism of complexes

μ:τq1Kosq()τ[q1,q]K.\mu\colon\tau_{\geq q-1}\operatorname{Kos}^{q}(\partial)\longrightarrow\tau_{[q-1,q]}K.

By functoriality of the maps δq\delta^{q}, we have a commutative square

q(τq1Kosq())\textstyle{\operatorname{\cal{H}}^{q}(\tau_{\geq q-1}\operatorname{Kos}^{q}(\partial))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ\scriptstyle{\mu}q1(τq1Kosq())[2]\textstyle{\operatorname{\cal{H}}^{q-1}(\tau_{\geq q-1}\operatorname{Kos}^{q}(\partial))[2]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i.e.q1(K)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(q11(K))0(K)[2]\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1}(K))\otimes\operatorname{\cal{H}}^{0}(K)[2]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q(τ[q1,q]K)\textstyle{\operatorname{\cal{H}}^{q}(\tau_{[q-1,q]}K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q1(τ[q1,q]K)[2]\textstyle{\operatorname{\cal{H}}^{q-1}(\tau_{[q-1,q]}K)[2]}q(K)\textstyle{\operatorname{\cal{H}}^{q}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q1(K)[2].\textstyle{\operatorname{\cal{H}}^{q-1}(K)[2].}

The assertion then follows from Proposition 3.2. ∎

Remark 3.4.

The proof of [AO20, Theorem 4.2.2(1)] implies the assertion of Proposition 3.3 under the stronger assumption that q!q! is invertible in 𝒪\mathcal{O}. However, the argument does not use the cdga structure of KK, only a weaker structure of a commutative monoid in the derived category K𝐋KKK\otimes^{\mathbf{L}}K\to K. In particular, the assertion holds for some EE_{\infty}-algebras which are not a priori equivalent to cdgas.

Remark 3.5.

In [AO20, Lemma 2.1.1], it is shown that the maps δKq\delta^{q}_{K} are compatible with the derived tensor product in the following way. If KK and LL are complexes and i,ji,j are integers such that i(K)\operatorname{\cal{H}}^{i}(K) and j(L)\operatorname{\cal{H}}^{j}(L) are flat 𝒪\mathcal{O}-modules, then the following square commutes.

i(K)j(L)\textstyle{\operatorname{\cal{H}}^{i}(K)\otimes\operatorname{\cal{H}}^{j}(L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δKi1+(1)iδLj\scriptstyle{\delta^{i}_{K}\otimes 1+(-1)^{i}\otimes\delta^{j}_{L}}i1(K)[2]j(L)i(K)j1(L)[2]\textstyle{\operatorname{\cal{H}}^{i-1}(K)[2]\otimes\operatorname{\cal{H}}^{j}(L)\,\,\oplus\,\,\operatorname{\cal{H}}^{i}(K)\otimes\operatorname{\cal{H}}^{j-1}(L)[2]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i+j(K𝐋L)\textstyle{\operatorname{\cal{H}}^{i+j}(K\otimes^{\mathbf{L}}L)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δK𝐋Li+j\scriptstyle{\delta_{K\otimes^{\mathbf{L}}L}^{i+j}}i+j1(K𝐋L)[2]\textstyle{\operatorname{\cal{H}}^{i+j-1}(K\otimes^{\mathbf{L}}L)[2]}

If q!q! is invertible in 𝒪\mathcal{O}, so that q1(K)\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}(K) is a direct summand of 1(K)q\operatorname{\cal{H}}^{1}(K)^{\otimes q}, the assertion of Proposition 3.3 can be deduced from this result.

For illustration, let us see how to do this for q=2q=2. We set L=KL=K and i=j=1i=j=1 in the above diagram, obtaining the middle square of the diagram below.

21(K)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{2}\operatorname{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}xy12(xyyx)\scriptstyle{x\wedge y\,\mapsto\,\frac{1}{2}(x\otimes y-y\otimes x)}η2\scriptstyle{\eta^{2}}1(K)1(K)\textstyle{\operatorname{\cal{H}}^{1}(K)\otimes\operatorname{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idδK1\scriptstyle{\mathrm{id}\otimes\delta^{1}_{K}}1(K)0(K)[2]\textstyle{\operatorname{\cal{H}}^{1}(K)\otimes\operatorname{\cal{H}}^{0}(K)[2]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}12(shuffle,id)\scriptstyle{\frac{1}{2}(\mathrm{shuffle},\mathrm{id})}1(K)1(K)\textstyle{\operatorname{\cal{H}}^{1}(K)\otimes\operatorname{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δK111δK1\scriptstyle{\delta^{1}_{K}\otimes 1-1\otimes\delta^{1}_{K}}0(K)[2]1(K)1(K)0(K)[2]\textstyle{\operatorname{\cal{H}}^{0}(K)[2]\otimes\operatorname{\cal{H}}^{1}(K)\,\,\oplus\,\,\operatorname{\cal{H}}^{1}(K)\otimes\operatorname{\cal{H}}^{0}(K)[2]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2(K𝐋K)\textstyle{\operatorname{\cal{H}}^{2}(K\otimes^{\mathbf{L}}K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δK𝐋K2\scriptstyle{\delta^{2}_{K\otimes^{\mathbf{L}}K}}1(K𝐋K)[2]\textstyle{\operatorname{\cal{H}}^{1}(K\otimes^{\mathbf{L}}K)[2]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2(K)\textstyle{\operatorname{\cal{H}}^{2}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δK2\scriptstyle{\delta^{2}_{K}}1(K)[2]\textstyle{\operatorname{\cal{H}}^{1}(K)[2]}

Here, the bottom square certifies the functoriality of δ2\delta^{2} with respect to the multiplication map K𝐋KKK\otimes^{\mathbf{L}}K\to K. Commutativity of the top square is easy to check. Then, commutativity of the exterior square gives the required assertion.

Corollary 3.6.

Suppose 22 is a nonzerodivisor in 𝒪\mathcal{O}. Let KK be an abstract Koszul complex satisfying the flatness condition (2.1) and let q1q\geq 1. We have the following commutative diagram

q1(K)\textstyle{\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

\displaystyle\sim

ηq\scriptstyle{\eta^{q}}(q11(K))1(K)\textstyle{(\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1}(K))\otimes\operatorname{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idδK1\scriptstyle{\mathrm{id}\otimes\delta^{1}_{K}}q11(K)[2]\textstyle{\operatorname{\mbox{\large$\wedge$}}^{q-1}\operatorname{\cal{H}}^{1}(K)[2]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

\displaystyle\sim

q(K)\textstyle{\operatorname{\cal{H}}^{q}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δKq\scriptstyle{\delta^{q}_{K}}q1(K)[2].\textstyle{\operatorname{\cal{H}}^{q-1}(K)[2].}

In other words, using the vertical identifications, we have the equality

δKq=(idδK1)ηq\delta^{q}_{K}=(\mathrm{id}\otimes\delta^{1}_{K})\circ\eta^{q}

in Hom(q(K),q1(K)[2])\operatorname{Hom}(\operatorname{\cal{H}}^{q}(K),\operatorname{\cal{H}}^{q-1}(K)[2]).

Corollary 3.7.

Suppose 22 is a nonzerodivisor in 𝒪\mathcal{O}. Let KK be an abstract Koszul complex satisfying the flatness condition (2.1) and let q1q\geq 1. Then

clsc[q1,q](K)=ηq(clsc(τ1K)).\operatorname{cl}\operatorname{sc}_{[q-1,q]}(K)=\eta^{q}(\operatorname{cl}\operatorname{sc}(\tau_{\leq 1}K)).
Corollary 3.8.

Suppose that 22 is a nonzerodivisor in 𝒪\mathcal{O}. Let KK be an abstract Koszul complex satisfying the flatness condition (2.1). Then the differential

d2pq:Hp(X,q(K))Hp+2(X,q1(K))d_{2}^{pq}\colon H^{p}(X,\operatorname{\cal{H}}^{q}(K))\to H^{p+2}(X,\operatorname{\cal{H}}^{q-1}(K))

equals the cup product with the class

δK1=clsc(τ1K)H2(X,Hom¯(1(K),0(K)))\delta^{1}_{K}=-\operatorname{cl}\operatorname{sc}(\tau_{\leq 1}K)\in H^{2}(X,\underline{\operatorname{Hom}}(\operatorname{\cal{H}}^{1}(K),\operatorname{\cal{H}}^{0}(K)))

followed by evaluation.

3.5. Morphisms of gerbes of splittings

Let KK be an abstract Koszul complex satisfying the flatness condition (2.1). In Corollary 3.7, under the assumption that 22 is a nonzerodivisor in 𝒪\mathcal{O}, we calculated the gerbe classes clsc[q1,q]K\operatorname{cl}\operatorname{sc}_{[q-1,q]}K in terms of the class clsc(τ1K)\operatorname{cl}\operatorname{sc}(\tau_{\leq 1}K). Below, we promote this equality into a morphism of gerbes.

Theorem 3.9.

For each integer q1q\geq 1, there is a morphism

q:sc(τ1K)sc(τ[q1,q]K)\wedge^{q}:\operatorname{sc}(\tau_{\leq 1}K)\longrightarrow\operatorname{sc}(\tau_{[q-1,q]}K)

of gerbes over XX, under which the obstruction classes correspond by the relation

clsc(τ[q1,q]K)=ctrq(clscτ1(K)),\operatorname{cl}\operatorname{sc}(\tau_{[q-1,q]}K)=\operatorname{ctr}^{q}(\operatorname{cl}\operatorname{sc}\tau_{\leq 1}(K)),

where ctrq:Hom¯(1,0)Hom¯(q,q1)\operatorname{ctr}^{q}:\underline{\operatorname{Hom}}({\cal{H}}^{1},{\cal{H}}^{0})\to\underline{\operatorname{Hom}}({\cal{H}}^{q},{\cal{H}}^{q-1}) denotes the morphism which maps a local section ff of the source to the one of the target by the formula

ctrq(f):ω1ωqj=1q(1)j1f(ωj)ω1ωj1ωj+1ωq.\operatorname{ctr}^{q}(f):\omega_{1}\wedge\cdots\wedge\omega_{q}\mapsto\sum_{j=1}^{q}(-1)^{j-1}f(\omega_{j})\omega_{1}\wedge\cdots\wedge\omega_{j-1}\wedge\omega_{j+1}\wedge\cdots\wedge\omega_{q}.

(Compare the formula for ctrq\operatorname{ctr}^{q} with the explicit formula for ηq\eta^{q} in Section 2.2.)

Notations. Before proceeding to the proof, we gather some notations concerning Čech cohomology. We denote by 𝒞ˇ(U,K)\check{\cal{C}}(U_{\bullet},K^{\bullet}) the Čech resolution of a complex KK^{\bullet} with respect to a hypercovering UU_{\bullet}. The differential induced by that of KK^{\bullet} will still be denoted by dd, while the Čech differential on the component 𝒞ˇ(Up,Kq)\check{\cal{C}}(U_{p},K^{q}):

(1)qi=0p+1(1)idi(-1)^{q}\sum_{i=0}^{p+1}(-1)^{i}d_{i}^{\ast}

will be denoted by dˇ\check{d}. Then the total differential

D=d+dˇD=d+\check{d}

is the differential of the total complex 𝒞ˇ(U,K)\check{\cal{C}}(U_{\bullet},K^{\bullet}).

When we compute the obstruction classes, we will use some notations which may not be standard. As usual, for each integer m1m\geq-1, we denote by [m][m] the set of integers ii such that 0im0\leq i\leq m (empty set for [1][-1]). And we denote by dij:[m2][m]d_{ij}:[m-2]\to[m] the unique increasing injection omitting ii and jj, where 0i<jm0\leq i<j\leq m. For example, for m=2m=2, we have

d02=d2d0=d0d1:[0][2]d_{02}=d_{2}\circ d_{0}=d_{0}\circ d_{1}:[0]\to[2]

(which maps 0 onto 11), where di:[m1][m]d_{i}:[m-1]\to[m] denotes the unique increasing injection omitting ii.

On the other hand, we denote by pri:[0][m]\operatorname{pr}_{i}:[0]\to[m] (resp. prij:[1][m]\operatorname{pr}_{ij}:[1]\to[m]) the unique map sending 0 to ii (resp. 0 to ii and 11 to jj), for 0im0\leq i\leq m (resp. for 0i<jm0\leq i<j\leq m).

Proof of Theorem 3.9.

In order to prove Theorem 3.9, we first describe the morphism, show that it is well-defined, and then calculate the obstruction class.

Construction of the functor q\wedge^{q}. We construct q:scτ1Kscτ[q1,q]K\wedge^{q}:\operatorname{sc}\tau_{\leq 1}K\to\operatorname{sc}\tau_{[q-1,q]}K by stackifying a morphism between the corresponding prestacks: scτ1Kscτ[q1,q]K\operatorname{sc}^{\prime}\tau_{\leq 1}K\to\operatorname{sc}^{\prime}\tau_{[q-1,q]}K.

Given an object of sc(τ1K)\operatorname{sc}^{\prime}(\tau_{\leq 1}K) over UU, that is, a section s:1𝒵1s:{\cal{H}}^{1}\to{\cal{Z}}^{1} of the projection 𝒵11{\cal{Z}}^{1}\to{\cal{H}}^{1} over UU, we define q(s)\wedge^{q}(s) as the composite morphism

qq(1)\textstyle{{\cal{H}}^{q}\simeq\wedge^{q}({\cal{H}}^{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q(s)\scriptstyle{\wedge^{q}(s)}q(𝒵1)\textstyle{\wedge^{q}({\cal{Z}}^{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}prod\scriptstyle{\mathrm{prod}}𝒵q,\textstyle{{\cal{Z}}^{q},}

where prod\mathrm{prod} means product; it is clearly a section of 𝒵qq{\cal{Z}}^{q}\to{\cal{H}}^{q} over UU.

Let s0s_{0} and s1s_{1} be two objects of sc(τ1K)\operatorname{sc}^{\prime}(\tau_{\leq 1}K) over UU and let hh be a homotopy from s0s_{0} to s1s_{1}. Then we need to define a corresponding homotopy q(h)\wedge^{q}(h) from q(s0)\wedge^{q}(s_{0}) to q(s1)\wedge^{q}(s_{1}). We first define a map

q(h):(1)qKq1/q1\wedge^{\prime q}(h):({\cal{H}}^{1})^{\otimes q}\longrightarrow K^{q-1}/{\cal{B}}^{q-1}

by letting it send ω1ωq\omega_{1}\otimes\cdots\otimes\omega_{q} (where ωj\omega_{j} are local sections of 1{\cal{H}}^{1}) to the class of

j=1q(1)j1h(ωj)s0(ω1)s0(ωj1)s1(ωj+1)s1(ωq)\sum_{j=1}^{q}(-1)^{j-1}h(\omega_{j})s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{j-1})\wedge s_{1}(\omega_{j+1})\wedge\cdots\wedge s_{1}(\omega_{q})

modulo q1{\cal{B}}^{q-1}.

It is easy to show that it factors through q1\wedge^{q}{\cal{H}}^{1} : If, say, ω1=ω2=ω\omega_{1}=\omega_{2}=\omega, then the alternating sum on the right reduces to the difference of the first two terms

h(ω)s1(ω)s1(ω3)s1(ωq)h(ω)s0(ω)s1(ω3)s1(ωq),h(\omega)s_{1}(\omega)\wedge s_{1}(\omega_{3})\wedge\cdots\wedge s_{1}(\omega_{q})-h(\omega)s_{0}(\omega)\wedge s_{1}(\omega_{3})\wedge\cdots\wedge s_{1}(\omega_{q}),

which is equal to

h(ω)dh(ω)s1(ω3)s1(ωq)h(\omega)dh(\omega)\wedge s_{1}(\omega_{3})\wedge\cdots\wedge s_{1}(\omega_{q})

which is a coboundary since 22 is invertible. Thus we defined q(h)\wedge^{q}(h).

(1)q\textstyle{({\cal{H}}^{1})^{\otimes q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q(h)\scriptstyle{\wedge^{\prime q}(h)}q(1)\textstyle{\wedge^{q}({\cal{H}}^{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q(h)\scriptstyle{\wedge^{q}(h)}Kq1/q1\textstyle{K^{q-1}/{\cal{B}}^{q-1}}

Then the following calculation shows that q(h)\wedge^{q}(h) is really a homotopy:

(q(s1)q(s0))(ω1ωq)\displaystyle\left(\wedge^{q}(s_{1})-\wedge^{q}(s_{0})\right)(\omega_{1}\wedge\cdots\wedge\omega_{q}) =j=1qs0(ω1)s0(ωj1){s1(ωj)s0(ωj)}s1(ωj+1)s1(ωq)\displaystyle=\sum_{j=1}^{q}s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{j-1})\wedge\{s_{1}(\omega_{j})-s_{0}(\omega_{j})\}\wedge s_{1}(\omega_{j+1})\wedge\cdots\wedge s_{1}(\omega_{q})
=j=1qs0(ω1)s0(ωj1){dh(ωj)}s1(ωj+1)s1(ωq)\displaystyle=\sum_{j=1}^{q}s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{j-1})\wedge\{dh(\omega_{j})\}\wedge s_{1}(\omega_{j+1})\wedge\cdots\wedge s_{1}(\omega_{q})
=j=1q(1)j1dh(ωj)s0(ω1)s0(ωj1)s1(ωj+1)s1(ωq)\displaystyle=\sum_{j=1}^{q}(-1)^{j-1}dh(\omega_{j})\wedge s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{j-1})\wedge s_{1}(\omega_{j+1})\wedge\cdots\wedge s_{1}(\omega_{q})
=d[q(h)(ω1ωq)].\displaystyle=d[\wedge^{q}(h)(\omega_{1}\wedge\cdots\wedge\omega_{q})].

Functoriality of q\wedge^{q}. Now in order to show that the morphism q\wedge^{q} is a functor, we must show that it is compatible with the composition of homotopies; so let h:s0s1h:s_{0}\Rightarrow s_{1} and h:s1s2h^{\prime}:s_{1}\Rightarrow s_{2} be two such in the source. We first define a second homotopy operator:

H2q(h,h):(1)q\displaystyle H_{2}^{q}(h,h^{\prime}):({\cal{H}}^{1})^{\otimes q} \displaystyle\to Kq2\displaystyle K^{q-2}
ω1ωq\displaystyle\omega_{1}\otimes\cdots\otimes\omega_{q} \displaystyle\mapsto 1j<kq(1)j+k+1h(ωj)h(ωk)s(j,k),\displaystyle\sum_{1\leq j<k\leq q}(-1)^{j+k+1}h(\omega_{j})h^{\prime}(\omega_{k})s(j,k),

where s(j,k)s(j,k) is equal to

s0(ω1)s0(ωj1)s1(ωj+1)s1(ωk1)s2(ωk+1)s2(ωq).s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{j-1})\wedge s_{1}(\omega_{j+1})\wedge\cdots\wedge s_{1}(\omega_{k-1})\wedge s_{2}(\omega_{k+1})\wedge\cdots\wedge s_{2}(\omega_{q}).

To show that q(h+h)\wedge^{q}(h+h^{\prime}) and q(h)+q(h)\wedge^{q}(h)+\wedge^{q}(h^{\prime}) are the same homotopies, it suffices to demonstrate the formula

[q(h+h){q(h)+q(h)}](ω1ωq)=dH2q(ω1ωq).[\wedge^{q}(h+h^{\prime})-\{\wedge^{q}(h)+\wedge^{q}(h^{\prime})\}](\omega_{1}\wedge\cdots\wedge\omega_{q})=dH_{2}^{q}(\omega_{1}\otimes\cdots\otimes\omega_{q}).

One expands the left hand side and groups the terms involving hh and hh^{\prime} separately:

j=1q(1)j1{h(ωj)+h(ωj)}s0(ω1)s0(ωj1)s2(ωj+1)s2(ωq)\displaystyle\sum_{j=1}^{q}(-1)^{j-1}\{h(\omega_{j})+h^{\prime}(\omega_{j})\}s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{j-1})\wedge s_{2}(\omega_{j+1})\wedge\cdots\wedge s_{2}(\omega_{q})
j=1q(1)j1h(ωj)s0(ω1)s0(ωj1)s1(ωj+1)s1(ωq)\displaystyle\,\,-\sum_{j=1}^{q}(-1)^{j-1}h(\omega_{j})s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{j-1})\wedge s_{1}(\omega_{j+1})\wedge\cdots\wedge s_{1}(\omega_{q})
j=1q(1)j1h(ωj)s1(ω1)s1(ωj1)s2(ωj+1)s2(ωq)\displaystyle\,\,-\sum_{j=1}^{q}(-1)^{j-1}h^{\prime}(\omega_{j})s_{1}(\omega_{1})\wedge\cdots\wedge s_{1}(\omega_{j-1})\wedge s_{2}(\omega_{j+1})\wedge\cdots\wedge s_{2}(\omega_{q})
=\displaystyle= j=1q(1)j1h(ωj)s0(ω1)s0(ωj1){s2(ωj+1)s2(ωq)s1(ωj+1)s1(ωq)}\displaystyle\sum_{j=1}^{q}(-1)^{j-1}h(\omega_{j})s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{j-1})\wedge\{s_{2}(\omega_{j+1})\wedge\cdots\wedge s_{2}(\omega_{q})-s_{1}(\omega_{j+1})\wedge\cdots\wedge s_{1}(\omega_{q})\}
+\displaystyle+ k=1q(1)k1h(ωk){s0(ω1)s0(ωk1)s1(ω1)s1(ωk1)}s2(ωk+1)s2(ωq).\displaystyle\sum_{k=1}^{q}(-1)^{k-1}h^{\prime}(\omega_{k})\{s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{k-1})-s_{1}(\omega_{1})\wedge\cdots\wedge s_{1}(\omega_{k-1})\}\wedge s_{2}(\omega_{k+1})\wedge\cdots\wedge s_{2}(\omega_{q}).

Now the differences in the curly brackets are themselves alternating sums, so

=\displaystyle= j=1q(1)j1h(ωj)s0(ω1)s0(ωj1)\displaystyle\sum_{j=1}^{q}(-1)^{j-1}h(\omega_{j})s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{j-1})\wedge
{k>j(1)k(j+1)dh(ωk)s1(ωj+1)s1(ωk1)s2(ωk+1)s2(ωq)}\displaystyle\,\,\,\,\wedge\left\{\sum_{k>j}(-1)^{k-(j+1)}dh^{\prime}(\omega_{k})\wedge s_{1}(\omega_{j+1})\wedge\cdots\wedge s_{1}(\omega_{k-1})\wedge s_{2}(\omega_{k+1})\wedge\cdots\wedge s_{2}(\omega_{q})\right\}
+\displaystyle+ k=1q(1)k1h(ωk){j<k(1)j1(dh(ωj))s0(ω1)s0(ωj1)s1(ωj+1)s1(ωk1)}\displaystyle\sum_{k=1}^{q}(-1)^{k-1}h^{\prime}(\omega_{k})\left\{\sum_{j<k}(-1)^{j-1}(-dh(\omega_{j}))\wedge s_{0}(\omega_{1})\wedge\cdots\wedge s_{0}(\omega_{j-1})\wedge s_{1}(\omega_{j+1})\wedge\cdots s_{1}(\omega_{k-1})\right\}\wedge
s2(ωk+1)s2(ωq)\displaystyle\,\,\,\,\wedge s_{2}(\omega_{k+1})\wedge\cdots\wedge s_{2}(\omega_{q})
=\displaystyle= 1j<kq(1)j+k+1{h(ωj)dh(ωk)+h(ωk)dh(ωj)}s(j,k),\displaystyle\sum_{1\leq j<k\leq q}(-1)^{j+k+1}\{h(\omega_{j})dh^{\prime}(\omega_{k})+h^{\prime}(\omega_{k})dh(\omega_{j})\}\wedge s(j,k),

and this is now equal to:

(3.3) =dH2q(h,h)(ω1ωq).=dH_{2}^{q}(h,h^{\prime})(\omega_{1}\otimes\cdots\otimes\omega_{q}).

This completes the proof of the fact that q\wedge^{q} is a functor.

Calculation of obstruction classes. Finally, we relate the obstruction elements. Let UXU_{\bullet}\to X be an open hypercovering such that one has

  1. (1)

    A section s:1𝒵1s:{\cal{H}}^{1}\to{\cal{Z}}^{1} of the canonical projection 𝒵11{\cal{Z}}^{1}\to{\cal{H}}^{1} over U0U_{0} and

  2. (2)

    A homotopy h:1K0h:{\cal{H}}^{1}\to K^{0} over U1U_{1} satisfying

    (3.4) d1sd0s=dhd_{1}^{\ast}s-d_{0}^{\ast}s=dh

Then by definition, the class of

obs=obs1=d0hd1h+d2hΓ(U2,Hom¯(1,0))\operatorname{obs}=\operatorname{obs}_{1}=d_{0}^{\ast}h-d_{1}^{\ast}h+d_{2}^{\ast}h\in\Gamma(U_{2},\underline{\operatorname{Hom}}({\cal{H}}^{1},{\cal{H}}^{0}))

in H2(X,Hom¯(1,0))H^{2}(X,\underline{\operatorname{Hom}}({\cal{H}}^{1},{\cal{H}}^{0})) is clscτ1K\operatorname{cl}\operatorname{sc}\tau_{\leq 1}K. On the other hand, by applying q\wedge^{q} to ss and hh, one sees that the class of

obsq=d0(qh)d1(qh)+d2(qh)Γ(U2,Hom¯(q,q1))\operatorname{obs}_{q}=d_{0}^{\ast}(\wedge^{q}h)-d_{1}^{\ast}(\wedge^{q}h)+d_{2}^{\ast}(\wedge^{q}h)\in\Gamma(U_{2},\underline{\operatorname{Hom}}({\cal{H}}^{q},{\cal{H}}^{q-1}))

in H2(X,Hom¯(q,q1))H^{2}(X,\underline{\operatorname{Hom}}({\cal{H}}^{q},{\cal{H}}^{q-1})) is clscτ[q1,q]K\operatorname{cl}\operatorname{sc}\tau_{[q-1,q]}K.

Now let ω1ωq\omega_{1}\wedge\cdots\wedge\omega_{q} be a local section of qq1{\cal{H}}^{q}\simeq\wedge^{q}{\cal{H}}^{1}. Then the evaluation of obsq\operatorname{obs}_{q} at ω1ωq\omega_{1}\wedge\cdots\wedge\omega_{q} is equal to

j=1q(1)j+1d0h(ωj)d01s(ω1)d01s(ωj1)d02s(ωj+1)d02s(ωq)\displaystyle\sum_{j=1}^{q}(-1)^{j+1}d_{0}^{\ast}h(\omega_{j})d_{01}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{01}^{\ast}s(\omega_{j-1})\wedge d_{02}^{\ast}s(\omega_{j+1})\wedge\cdots d_{02}^{\ast}s(\omega_{q})
\displaystyle- j=1q(1)j+1d1h(ωj)d01s(ω1)d01s(ωj1)d12s(ωj+1)d12s(ωq)\displaystyle\sum_{j=1}^{q}(-1)^{j+1}d_{1}^{\ast}h(\omega_{j})d_{01}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{01}^{\ast}s(\omega_{j-1})\wedge d_{12}^{\ast}s(\omega_{j+1})\wedge\cdots d_{12}^{\ast}s(\omega_{q})
+\displaystyle+ j=1q(1)j+1d2h(ωj)d02s(ω1)d02s(ωj1)d12s(ωj+1)d12s(ωq).\displaystyle\sum_{j=1}^{q}(-1)^{j+1}d_{2}^{\ast}h(\omega_{j})d_{02}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{02}^{\ast}s(\omega_{j-1})\wedge d_{12}^{\ast}s(\omega_{j+1})\wedge\cdots d_{12}^{\ast}s(\omega_{q}).

One groups the terms around the second sum and gets

j=1q(1)j+1(d0hd1h+d2h)(ωj)\displaystyle\sum_{j=1}^{q}(-1)^{j+1}(d_{0}^{\ast}h-d_{1}^{\ast}h+d_{2}^{\ast}h)(\omega_{j})\cdot
d01s(ω1)d01s(ωj1)d12s(ωj+1)d12s(ωq)\displaystyle\,\,\cdot d_{01}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{01}^{\ast}s(\omega_{j-1})\wedge d_{12}^{\ast}s(\omega_{j+1})\wedge\cdots\wedge d_{12}^{\ast}s(\omega_{q})
+\displaystyle+ j=1q(1)j+1d0h(ωj)d01s(ω1)d01s(ωj1)\displaystyle\sum_{j=1}^{q}(-1)^{j+1}d_{0}^{\ast}h(\omega_{j})d_{01}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{01}^{\ast}s(\omega_{j-1})\wedge
{d02s(ωj+1)d02s(ωq)d12s(ωj+1)d12s(ωq)}\displaystyle\,\,\wedge\{d_{02}^{\ast}s(\omega_{j+1})\wedge\cdots\wedge d_{02}^{\ast}s(\omega_{q})-d_{12}^{\ast}s(\omega_{j+1})\wedge\cdots\wedge d_{12}^{\ast}s(\omega_{q})\}
+\displaystyle+ j=1q(1)j+1d2h(ωj){d02s(ω1)d02s(ωj1)d01s(ω1)d01s(ωj1)}\displaystyle\sum_{j=1}^{q}(-1)^{j+1}d_{2}^{\ast}h(\omega_{j})\{d_{02}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{02}^{\ast}s(\omega_{j-1})-d_{01}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{01}^{\ast}s(\omega_{j-1})\}\wedge
d12s(ωj+1)d12s(ωq).\displaystyle\,\,\wedge d_{12}^{\ast}s(\omega_{j+1})\wedge\cdots\wedge d_{12}^{\ast}s(\omega_{q}).

The first alternating sum reduces to the “main” term we want, when taken modulo the coboundaries (q1{\cal{B}}^{q-1}). In the last two sums, one first notes that, as hh is a homotopy from d0sd_{0}^{\ast}s to d1sd_{1}^{\ast}s, it follows that d0hd_{0}^{\ast}h is a homotopy from d0d0sd_{0}^{\ast}d_{0}^{\ast}s to d0d1sd_{0}^{\ast}d_{1}^{\ast}s, that is,

d0h:d01sd02s.d_{0}^{\ast}h:d_{01}^{\ast}s\Rightarrow d_{02}^{\ast}s.

Similarly, d2hd_{2}^{\ast}h is a homotopy from d02sd_{02}^{\ast}s to d12sd_{12}^{\ast}s. Essentially by repeating the last three equalities leading up to (3.3), this time with a minus sign, one sees that the last two sums add up to

dH2q(d0h,d2h)(ω1ωq),-dH^{q}_{2}(d_{0}^{\ast}h,d_{2}^{\ast}h)(\omega_{1}\otimes\cdots\omega_{q}),

which is a coboundary. Therefore, reducing modulo q1{\cal{B}}^{q-1}, one gets

ev(obsq,ω1ωq)=ev(ctrq(obs1),ω1ωq).\operatorname{ev}(\operatorname{obs}_{q},\omega_{1}\wedge\cdots\wedge\omega_{q})=\operatorname{ev}(\operatorname{ctr}^{q}(\operatorname{obs}_{1}),\omega_{1}\wedge\cdots\wedge\omega_{q}).

This means

clscτ[q1,q]K=ctrqclscτ1K,\operatorname{cl}\operatorname{sc}\tau_{[q-1,q]}K=\operatorname{ctr}^{q}\operatorname{cl}\operatorname{sc}\tau_{\leq 1}K,

which completes the proof. ∎

Remark 3.10.

The construction of the map between gerbes can also be carried out using the language of higher topos theory [Lur09]. Let us give a brief outline.

Let p:YZp\colon Y\to Z be a map in the homotopy category of spaces, or more generally in any \infty-category 𝒞\cal{C}. On can then build the space sc(p)\operatorname{sc}(p) of splittings of pp as the homotopy fiber of

p:Hom𝒞(Z,Y)Hom𝒞(Z,Z).p\colon\operatorname{Hom}_{\cal{C}}(Z,Y)\longrightarrow\operatorname{Hom}_{\cal{C}}(Z,Z).

over the identity idZ{\rm id}_{Z}. Similarly, if p:YZp\colon Y\to Z is a map in the derived \infty-category of a ringed topos (X,𝒪)(X,\mathcal{O}), one obtains a sheaf of spaces sc¯(p)\underline{\operatorname{sc}}(p) of splittings of pp.

In the special case when YY is a two-term complex K=[K0𝑑K1]K=[K^{0}\xrightarrow{d}K^{1}] satisfying the conditions in §3.2 and YZY\to Z is the projection K1(K)[1]K\to\operatorname{\cal{H}}^{1}(K)[-1], then sc¯(p)\underline{\operatorname{sc}}(p) is a sheaf of groupoids (a stack) and can be identified with the gerbe of splittings scK\operatorname{sc}K.

Applying the functor τq1LΓq\tau_{\geq q-1}L\Gamma^{q} to the map p:K1(K)[1]p\colon K\to\operatorname{\cal{H}}^{1}(K)[-1] one obtains (simply by functoriality) a morphism of sheaves of spaces

sc¯(p)sc¯(τq1LΓq(p)).\underline{\operatorname{sc}}(p)\longrightarrow\underline{\operatorname{sc}}(\tau_{\geq q-1}L\Gamma^{q}(p)).

By inspection, the map τq1LΓq(p)\tau_{\geq q-1}L\Gamma^{q}(p) is the projection

τ[q1,q]Kosq(d)q(Kosq(d))[q]=q1(K)[q].\tau_{[q-1,q]}\operatorname{Kos}^{q}(d)\longrightarrow\operatorname{\cal{H}}^{q}(\operatorname{Kos}^{q}(d))[-q]=\operatorname{\mbox{\large$\wedge$}}^{q}\operatorname{\cal{H}}^{1}(K)[-q].

This way one obtains by abstract nonsense a morphism of gerbes scKsc(τ[q1,q]Kosq(d))\operatorname{sc}K\to\operatorname{sc}(\tau_{[q-1,q]}\operatorname{Kos}^{q}(d)).

In the case when KK is an abstract Koszul complex satisfying the flatness condition (2.1), the morphism of gerbes scτ1Kscτ[q1,q]K\operatorname{sc}\tau_{\leq 1}K\to\operatorname{sc}\tau_{[q-1,q]}K obtained this way should agree with the one constructed in Theorem 3.9, though we did not check it.

4. Gerbes of splittings of the de Rham complex

Our method of explicating the truncations τ[q1,q]K\tau_{[q-1,q]}K for an abstract Koszul complex KK in terms of the truncation τ1K\tau_{\leq 1}K requires that 22 is a nonzerodivisor. In this section, we describe these two-term truncations in the case of the de Rham complex in characteristic p>0p>0 by calculating the class

clsc(τ[q1,q]FΩX/S)\operatorname{cl}\operatorname{sc}\left(\tau_{[q-1,q]}F_{\ast}\Omega_{X/S}^{\bullet}\right)

The calculation uses more information about the de Rham complex than it being an abstract Koszul complex, namely the nature of the Cartier isomorphism (which we use only for p=2p=2). As a corollary, we deduce that τ[q1,q](FΩX/S)\tau_{[q-1,q]}(F_{*}\Omega^{\bullet}_{X/S}) is decomposable if τ1(FΩX/S)\tau_{\leq 1}(F_{*}\Omega^{\bullet}_{X/S}) is, and obtain a description of the d2d_{2} differentials in the conjugate spectral sequence.

Theorem 4.1.

Let SS be a scheme of characteristic p>0p>0 and X/SX/S a smooth separated scheme of finite type. Then for each integer qq, the class

clsc(τ[q1,q]FΩX/S)\operatorname{cl}\operatorname{sc}(\tau_{[q-1,q]}F_{\ast}\Omega^{\bullet}_{X/S})

is the image of the class

clsc(τ1FΩX/S)\operatorname{cl}\operatorname{sc}(\tau_{\leq 1}F_{\ast}\Omega^{\bullet}_{X/S})

under the contraction map (described in Theorem 3.9).

Proof.

We put K=FΩX/SK=F_{*}\Omega_{X/S}^{\bullet}, with F:XXF:X\to X^{\prime} the relative Frobenius of X/SX/S.

To calculate the class, we take an open hypercovering UXU_{\bullet}\to X^{\prime} such that

  1. (1)

    Over U0U_{0}, one has a section s:1𝒵1s:{\cal{H}}^{1}\to{\cal{Z}}^{1} of the projection 𝒵11{\cal{Z}}^{1}\to{\cal{H}}^{1} and a section

    σ(q):q(1)q\sigma^{(q)}:{\cal{H}}^{q}\to({\cal{H}}^{1})^{\otimes q}

    of the canonical projection (1)qq({\cal{H}}^{1})^{\otimes q}\to{\cal{H}}^{q}; and

  2. (2)

    Over U1U_{1}, one has a homotopy h:1K0h:{\cal{H}}^{1}\to K^{0} such that

    dh=d1sd0s:1𝒵1.dh=d_{1}^{\ast}s-d_{0}^{\ast}s:{\cal{H}}^{1}\to{\cal{Z}}^{1}.

(Let us recall that 1{\cal{H}}^{1}, and hence q=q1{\cal{H}}^{q}=\wedge^{q}{\cal{H}}^{1} for all integers qq, are locally free over 𝒪X{\cal{O}}_{X^{\prime}}.) The locally free kernel of the projection (1)qq({\cal{H}}^{1})^{\otimes q}\to{\cal{H}}^{q} being denoted by q{\cal{I}}^{q}, the 11-cocycle

d0σ(q)d1σ(q)Γ(U1,Hom¯𝒪X(q,q))d_{0}^{\ast}\sigma^{(q)}-d_{1}^{\ast}\sigma^{(q)}\in\Gamma\left(U_{1},\underline{\operatorname{Hom}}_{{\cal{O}}_{X^{\prime}}}({\cal{H}}^{q},{\cal{I}}^{q})\right)

represents the obstruction, in H1(X,Hom¯(q,q))=Ext𝒪X1(q,q)H^{1}(X^{\prime},\underline{\operatorname{Hom}}({\cal{H}}^{q},{\cal{I}}^{q}))=\operatorname{Ext}^{1}_{{\cal{O}}_{X^{\prime}}}({\cal{H}}^{q},{\cal{I}}^{q}), to the global existence of a section.

Let us calculate the class

clscτ[q1,q]KH2(X,Hom¯(q,q1))\operatorname{cl}\operatorname{sc}\tau_{[q-1,q]}K\in H^{2}(X^{\prime},\underline{\operatorname{Hom}}({\cal{H}}^{q},{\cal{H}}^{q-1}))

in characteristic p2p\geq 2. For ease of notation, we denote σ(q)\sigma^{(q)} simply by σ\sigma when no confusion is likely.

To do so, we may choose the composite morphism

q\textstyle{{\cal{H}}^{q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ(q)\scriptstyle{\sigma^{(q)}}(1)q\textstyle{({\cal{H}}^{1})^{\otimes q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sq\scriptstyle{s^{\otimes q}}(𝒵1)q\textstyle{({\cal{Z}}^{1})^{\otimes q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\wedge}𝒵q\textstyle{{\cal{Z}}^{q}}

which we denote by (sq)σ(s^{\wedge q})\circ\sigma, as the section of the projection 𝒵qq{\cal{Z}}^{q}\to{\cal{H}}^{q} over U0U_{0}.

Then one forms (the negative of) the Čech difference

d1(sq)d1σd0(sq)d0σ\displaystyle d_{1}^{\ast}(s^{\wedge q})\circ d_{1}^{\ast}\sigma-d_{0}^{\ast}(s^{\wedge q})\circ d_{0}^{\ast}\sigma
=\displaystyle= [(d1s)q(d0s)q]d0σ(d1s)q(d0σd1σ).\displaystyle[(d_{1}^{\ast}s)^{\wedge q}-(d_{0}^{\ast}s)^{\wedge q}]\circ d_{0}^{\ast}\sigma-(d_{1}^{\ast}s)^{\wedge q}\circ(d_{0}^{\ast}\sigma-d_{1}^{\ast}\sigma).

One notes that the second term is zero, since the image of d0σd1σd_{0}^{\ast}\sigma-d_{1}^{\ast}\sigma is contained in q{\cal{I}}^{q}, which in turn is annihilated by (d1s)q(d_{1}^{\ast}s)^{\wedge q}, for the wedge product is strictly graded commutative.

Then one expresses the remaining first term as the differential of something:

((d1s)q(d0s)q)(ω1ωq)\displaystyle((d_{1}^{\ast}s)^{\wedge q}-(d_{0}^{\ast}s)^{\wedge q})(\omega_{1}\otimes\cdots\otimes\omega_{q})
=\displaystyle= j=1q(d0s)ω1(d0s)ωj1(d1sd0s)ωj(d1s)ωj+1(d1s)ωq\displaystyle\sum_{j=1}^{q}(d_{0}^{\ast}s)\omega_{1}\wedge\cdots\wedge(d_{0}^{\ast}s)\omega_{j-1}\wedge(d_{1}^{\ast}s-d_{0}^{\ast}s)\omega_{j}\wedge(d_{1}^{\ast}s)\omega_{j+1}\wedge\cdots\wedge(d_{1}^{\ast}s)\omega_{q}
=\displaystyle= dj=1q(1)j+1h(ωj)(d0s)ω1(d0s)ωj1(d1s)ωj+1(d1s)ωq.\displaystyle d\sum_{j=1}^{q}(-1)^{j+1}h(\omega_{j})(d_{0}^{\ast}s)\omega_{1}\wedge\cdots\wedge(d_{0}^{\ast}s)\omega_{j-1}\wedge(d_{1}^{\ast}s)\omega_{j+1}\wedge\cdots\wedge(d_{1}^{\ast}s)\omega_{q}.

One defines η=η(q)=η(ω1ωq)\eta=\eta^{(q)}=\eta(\omega_{1}\otimes\cdots\otimes\omega_{q}) to be

j=1q(1)j+1h(ωj)(d0s)ω1(d0s)ωj1(d1s)ωj+1(d1s)ωq\sum_{j=1}^{q}(-1)^{j+1}h(\omega_{j})(d_{0}^{\ast}s)\omega_{1}\wedge\cdots\wedge(d_{0}^{\ast}s)\omega_{j-1}\wedge(d_{1}^{\ast}s)\omega_{j+1}\wedge\cdots\wedge(d_{1}^{\ast}s)\omega_{q}

in order to have a commutative diagram

q\textstyle{{\cal{H}}^{q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d1(sq)d1σd0(sq)d0σ\scriptstyle{d_{1}^{\ast}(s^{\wedge q})\circ d_{1}^{\ast}\sigma-d_{0}^{\ast}(s^{\wedge q})\circ d_{0}^{\ast}\sigma}d0σ\scriptstyle{d_{0}^{\ast}\sigma}𝒵q\textstyle{{\cal{Z}}^{q}}(1)q\textstyle{({\cal{H}}^{1})^{\otimes q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η¯\scriptstyle{\overline{\eta}}Kq1/q1\textstyle{K^{q-1}/{\cal{B}}^{q-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d},

in which η¯\overline{\eta} means the composite of η\eta followed by Kq1Kq1/q1K^{q-1}\to K^{q-1}/{\cal{B}}^{q-1}.

With this, we calculate the class of the gerbe by forming the Čech difference

(4.1) (d0d1+d2)(η¯d0σ)\displaystyle(d_{0}^{\ast}-d_{1}^{\ast}+d_{2}^{\ast})(\overline{\eta}\circ d_{0}^{\ast}\sigma) =\displaystyle= d0η¯d01σd1η¯d01σ+d2η¯d02σ\displaystyle d_{0}^{\ast}\overline{\eta}\circ d_{01}^{\ast}\sigma-d_{1}^{\ast}\overline{\eta}\circ d_{01}^{\ast}\sigma+d_{2}^{\ast}\overline{\eta}\circ d_{02}^{\ast}\sigma
(4.2) =\displaystyle= (d0d1+d2)η¯d01σd2η¯(d01σd02σ).\displaystyle(d_{0}^{\ast}-d_{1}^{\ast}+d_{2}^{\ast})\overline{\eta}\circ d_{01}^{\ast}\sigma-d_{2}^{\ast}\overline{\eta}\circ(d_{01}^{\ast}\sigma-d_{02}^{\ast}\sigma).

Let us put obs1=(d0d1+d2)h\operatorname{obs}_{1}=(d_{0}^{\ast}-d_{1}^{\ast}+d_{2}^{\ast})h, which represents the class of scτ1K\operatorname{sc}\tau_{\leq 1}K. Then the first summand of (4.2) can be expressed in terms of obs1\operatorname{obs}_{1}:

(d0d1+d2)η¯(ω1ωq)\displaystyle(d_{0}^{\ast}-d_{1}^{\ast}+d_{2}^{\ast})\overline{\eta}(\omega_{1}\otimes\cdots\otimes\omega_{q})
=\displaystyle= j=1q(1)j+1d0h(ωj)d01s(ω1)d01s(ωj1)d02s(ωj+1)d02s(ωq)\displaystyle\sum_{j=1}^{q}(-1)^{j+1}d_{0}^{\ast}h(\omega_{j})d_{01}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{01}^{\ast}s(\omega_{j-1})\wedge d_{02}^{\ast}s(\omega_{j+1})\wedge\cdots\wedge d_{02}^{\ast}s(\omega_{q})
j=1q(1)j+1d1h(ωj)d01s(ω1)d01s(ωj1)d12s(ωj+1)d12s(ωq)\displaystyle-\sum_{j=1}^{q}(-1)^{j+1}d_{1}^{\ast}h(\omega_{j})d_{01}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{01}^{\ast}s(\omega_{j-1})\wedge d_{12}^{\ast}s(\omega_{j+1})\wedge\cdots\wedge d_{12}^{\ast}s(\omega_{q})
+j=1q(1)j+1d2h(ωj)d02s(ω1)d02s(ωj1)d12s(ωj+1)d12s(ωq)\displaystyle+\sum_{j=1}^{q}(-1)^{j+1}d_{2}^{\ast}h(\omega_{j})d_{02}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{02}^{\ast}s(\omega_{j-1})\wedge d_{12}^{\ast}s(\omega_{j+1})\wedge\cdots\wedge d_{12}^{\ast}s(\omega_{q})
=\displaystyle= j=1q(1)j+1obs1(ω1)d01s(ω1)d01s(ωj1)d12s(ωj+1)d12s(ωq)\displaystyle\sum_{j=1}^{q}(-1)^{j+1}\operatorname{obs}_{1}(\omega_{1})d_{01}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{01}^{\ast}s(\omega_{j-1})\wedge d_{12}^{\ast}s(\omega_{j+1})\wedge\cdots\wedge d_{12}^{\ast}s(\omega_{q})
+j=1q(1)j+1d0h(ωj)d01s(ω1)d01s(ωj1){(d02s(qj)d12s(qj))(ωj+1ωq)}\displaystyle+\sum_{j=1}^{q}(-1)^{j+1}d_{0}^{\ast}h(\omega_{j})d_{01}^{\ast}s(\omega_{1})\wedge\cdots\wedge d_{01}^{\ast}s(\omega_{j-1})\wedge\left\{(d_{02}^{\ast}s^{\wedge(q-j)}-d_{12}^{\ast}s^{\wedge(q-j)})(\omega_{j+1}\otimes\cdots\otimes\omega_{q})\right\}
+j=1q(1)j+1d2h(ωj){(d02s(j1)d01s(j1))(ω1ωj1)}d12s(ωj+1)s(ωq).\displaystyle+\sum_{j=1}^{q}(-1)^{j+1}d_{2}^{\ast}h(\omega_{j})\left\{(d_{02}^{\ast}s^{\wedge(j-1)}-d_{01}^{\ast}s^{\wedge(j-1)})(\omega_{1}\otimes\cdots\otimes\omega_{j-1})\right\}\wedge d_{12}^{\ast}s(\omega_{j+1})\wedge\cdots\wedge s(\omega_{q}).

Again as in the three equalities leading up to (3.3), the differences in the curly brackets are themselves alternating sums, and one sees that the sum of the last two alternating sums is equal to

dH2q(d0h,d2h)(ω1ωq),-dH_{2}^{q}(d_{0}^{\ast}h,d_{2}^{\ast}h)(\omega_{1}\otimes\cdots\otimes\omega_{q}),

hence is zero modulo q1{\cal{B}}^{q-1}. On the other hand, the first alternating sum is equal to

ev(ctr(obs1),ω1ωq).\operatorname{ev}(\operatorname{ctr}(\operatorname{obs}_{1}),\omega_{1}\wedge\cdots\wedge\omega_{q}).

Now we analyze the second summand in (4.2). It is the cup product of two cohomology classes:

η¯|qΓ(U1,Hom¯(q,q1))\displaystyle\overline{\eta}|_{{\cal{I}}^{q}}\in\Gamma(U_{1},\underline{\operatorname{Hom}}({\cal{I}}^{q},{\cal{H}}^{q-1})) representing [η¯|q]Ext1(q,q1) and\displaystyle[\overline{\eta}|_{{\cal{I}}^{q}}]\in\operatorname{Ext}^{1}({\cal{I}}^{q},{\cal{H}}^{q-1})\,\,\mbox{ and }
d0σd1σΓ(U1,Hom¯(q,q))\displaystyle d_{0}^{\ast}\sigma-d_{1}^{\ast}\sigma\in\Gamma(U_{1},\underline{\operatorname{Hom}}({\cal{H}}^{q},{\cal{I}}^{q})) representing [σ]Ext1(q,q).\displaystyle[\sigma]\in\operatorname{Ext}^{1}({\cal{H}}^{q},{\cal{I}}^{q}).

When qq is less than p=char.(S)p=\mathrm{char.}\,(S), [σ][\sigma] is zero, for in this case one disposes of a canonical section of

(1)qq,({\cal{H}}^{1})^{\otimes q}\to{\cal{H}}^{q},

namely the anti-symmetrization.

On the other hand, if pp is odd, then [η¯|q][\overline{\eta}|_{{\cal{I}}^{q}}] is zero, because (even more strongly) η¯\overline{\eta} itself kills q{\cal{I}}^{q}: for example, it maps a local section

ωωω3ωq\omega\otimes\omega\otimes\omega_{3}\otimes\cdots\otimes\omega_{q}

of q{\cal{I}}^{q} to the element

[h(ω)d1s(ω)h(ω)d0s(ω)]d1s(ω3)d1s(ωq)=h(ω)dh(ω)ω3ωq(modq1)\left[h(\omega)d_{1}^{\ast}s(\omega)-h(\omega)d_{0}^{\ast}s(\omega)\right]\wedge d_{1}^{\ast}s(\omega_{3})\wedge\cdots\wedge d_{1}^{\ast}s(\omega_{q})\quad=\quad h(\omega)dh(\omega)\wedge\omega_{3}\wedge\cdots\wedge\omega_{q}\pmod{{\cal{B}}^{q-1}}

which is a coboundary when 22 is invertible (d(h(ω)2)=2h(ω)dh(ω)d(h(\omega)^{2})=2h(\omega)dh(\omega)).

So, let us restrict our attention to the case p=2p=2 and show that the class [η¯|q][\overline{\eta}|_{{\cal{I}}^{q}}] is still zero. First, one can easily check that η¯:(1)qKq1/q1\overline{\eta}:({\cal{H}}^{1})^{\otimes q}\to K^{q-1}/{\cal{B}}^{q-1}, and a fortiori η¯|q:q𝒵q1/q1=q1\overline{\eta}|_{{\cal{I}}^{q}}:{\cal{I}}^{q}\to{\cal{Z}}^{q-1}/{\cal{B}}^{q-1}={\cal{H}}^{q-1}, is symmetric in the sense that any element of the form

ω1ωj1(ωjωj+1+ωj+1ωj)ωj+2ωq\omega_{1}\otimes\cdots\otimes\omega_{j-1}\otimes(\omega_{j}\otimes\omega_{j+1}+\omega_{j+1}\otimes\omega_{j})\otimes\omega_{j+2}\otimes\cdots\otimes\omega_{q}

(when 1=11=-1, adding is subtracting) maps to zero under η¯\overline{\eta}. Therefore, one has a commutative diagram

q\textstyle{{\cal{I}}^{q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q/𝒥q\textstyle{{\cal{I}}^{q}/{\cal{J}}^{q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}η~\scriptstyle{\widetilde{\eta}}q1\textstyle{{\cal{H}}^{q-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(1)q\textstyle{({\cal{H}}^{1})^{\otimes q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Symq(1)\textstyle{\operatorname{Sym}^{q}({\cal{H}}^{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Kq1/q1\textstyle{K^{q-1}/{\cal{B}}^{q-1}}

where the composite of the two horizontal arrows in the first row (resp. in the second row) is equal to η¯|q\overline{\eta}|_{{\cal{I}}^{q}} (resp. η¯\overline{\eta}), and 𝒥q{\cal{J}}^{q} denotes the (locally free) kernel of the projection (1)qSymq(1)({\cal{H}}^{1})^{\otimes q}\to\operatorname{Sym}^{q}({\cal{H}}^{1}).

We get a notational advantage by taking the quotient by 𝒥q{\cal{J}}^{q} : now q/𝒥q{\cal{I}}^{q}/{\cal{J}}^{q} is generated by the images of local sections of the form

(4.3) ωωω3ωq\omega\otimes\omega\otimes\omega_{3}\otimes\cdots\omega_{q}

Such a local section is mapped under η~\widetilde{\eta} onto

(4.4) h(ω)dh(ω)d0s(ω3)d0s(ωq)(modq1)\displaystyle h(\omega)dh(\omega)\wedge d_{0}^{\ast}s(\omega_{3})\wedge\cdots\wedge d_{0}^{\ast}s(\omega_{q})\pmod{{\cal{B}}^{q-1}}
(4.5) =\displaystyle= [h(ω)dh(ω)(mod1)]ω3ωq\displaystyle[h(\omega)dh(\omega)\pmod{{\cal{B}}^{1}}]\wedge\omega_{3}\wedge\cdots\omega_{q}

We prove that [η¯|q][\overline{\eta}|_{{\cal{I}}^{q}}] is zero by finding a 0-cochain zz with coefficients in Hom¯(q,q1)\underline{\operatorname{Hom}}({\cal{I}}^{q},{\cal{H}}^{q-1}), that is, a section of this sheaf over U0U_{0}, such that η¯|q=d0zd1z\overline{\eta}|_{{\cal{I}}^{q}}=d_{0}^{\ast}z-d_{1}^{\ast}z. As we know that η¯|q\overline{\eta}|_{{\cal{I}}^{q}} factors through η~:q/𝒥qq1\widetilde{\eta}:{\cal{I}}^{q}/{\cal{J}}^{q}\to{\cal{H}}^{q-1}, it suffices to find z~:q/𝒥qq1\widetilde{z}:{\cal{I}}^{q}/{\cal{J}}^{q}\to{\cal{H}}^{q-1} such that

η~=d1z~d0z~.\widetilde{\eta}=d_{1}^{\ast}\widetilde{z}-d_{0}^{\ast}\widetilde{z}.

But from (4.5), one sees that

(4.6) η~(ωωω3ωq)=(C1Wdh(ω))ω3ωq.\widetilde{\eta}(\omega\otimes\omega\otimes\omega_{3}\otimes\cdots\otimes\omega_{q})=\left(C^{-1}W^{\ast}dh(\omega)\right)\wedge\omega_{3}\wedge\cdots\wedge\omega_{q}.

We denote here by WW the base-change of the absolute Frobenius endomorphism of SS, so that the diagram

X\textstyle{X^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}W\scriptstyle{W}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}frobS\scriptstyle{\mathrm{frob}_{S}}S\textstyle{S}

is cartesian, by WW^{\ast} the pull-back morphism of differential forms

ΩX/S1WΩX/S1\Omega^{1}_{X/S}\longrightarrow W_{\ast}\Omega^{1}_{X^{\prime}/S}

and by C1C^{-1} the (inverse) Cartier operation

ΩX/S11(FΩX/S)\Omega^{1}_{X^{\prime}/S}\longrightarrow{\cal{H}}^{1}(F_{\ast}\Omega_{X/S}^{\bullet})

(cf. [Kat70, §7] and recall p=2p=2). Thus the last expression is the same as

C1(d1sω)ω3ωqC1(d0sω)ω3ωq,C^{-1}(d_{1}^{\ast}s\omega)\wedge\omega_{3}\wedge\cdots\wedge\omega_{q}-C^{-1}(d_{0}^{\ast}s\omega)\wedge\omega_{3}\wedge\cdots\wedge\omega_{q},

and one is led to define over U0U_{0}

z~:q/𝒥qq1\widetilde{z}:{\cal{I}}^{q}/{\cal{J}}^{q}\longrightarrow{\cal{H}}^{q-1}

so that it maps the local section (4.3) modulo 𝒥q{\cal{J}}^{q} to

C1(sω)ω3ωq.C^{-1}(s\omega)\wedge\omega_{3}\wedge\cdots\wedge\omega_{q}.

As pointed out earlier, local sections of the form (4.3) generate q/𝒥q{\cal{I}}^{q}/{\cal{J}}^{q}, so such z~\widetilde{z} is unique if exists at all. Now its existence can be shown locally: if one has a basis e1,,ede_{1},\cdots,e_{d} of 1{\cal{H}}^{1} over 𝒪X{\cal{O}}_{X^{\prime}}, then the images of the sections

{ej1ejq:1j1jqd with at least one repetition }\{e_{j_{1}}\otimes\cdots\otimes e_{j_{q}}:1\leq j_{1}\leq\cdots\leq j_{q}\leq d\,\,\mbox{ with at least one repetition }\,\,\}

under qq/𝒥q{\cal{I}}^{q}\to{\cal{I}}^{q}/{\cal{J}}^{q} form a local basis of q/𝒥q{\cal{I}}^{q}/{\cal{J}}^{q}, and then one can let z~\widetilde{z} map the class of ej1ejqe_{j_{1}}\otimes\cdots\otimes e_{j_{q}} to

C1(s(ωjk))( the rest ),C^{-1}(s(\omega_{j_{k}}))\wedge(\mbox{ the rest }),

where jkj_{k} is an index that repeats: If two or more indices repeat, whichever one is chosen, the result is zero, and if an index repeats itself three or more times, it doesn’t matter which consecutive terms are chosen, for 1=11=-1 and the sign doesn’t matter.

Then one needs to show that any local section of the form (4.3) is mapped as desired under z~\widetilde{z} thus defined. One expresses the sections ω,ω3,,ωq\omega,\omega_{3},\cdots,\omega_{q} as linear combinations of the {ei}\{e_{i}\} and one sees that it boils down to showing the linearity in each variable ω3,,ωq\omega_{3},\cdots,\omega_{q}, which is evident, as well as the linearity “in the variable ωω\omega\otimes\omega,” which is less so.

Let ω=αξ+βθ\omega=\alpha\xi+\beta\theta, where α,β\alpha,\beta are sections of 𝒪X{\cal{O}}_{X^{\prime}} and ξ,θ\xi,\theta sections of 1{\cal{H}}^{1}. Then one calculates

C1W(s(αξ+βθ))\displaystyle C^{-1}W^{\ast}(s(\alpha\xi+\beta\theta)) =\displaystyle= C1W(Fαs(ξ)+Fβs(θ))\displaystyle C^{-1}W^{\ast}(F^{\ast}\alpha\cdot s(\xi)+F^{\ast}\beta\cdot s(\theta))
=\displaystyle= C1(α2Ws(ξ)+β2Ws(θ))\displaystyle C^{-1}(\alpha^{2}W^{\ast}s(\xi)+\beta^{2}W^{\ast}s(\theta))
=\displaystyle= (Fα)2C1Ws(ξ)+(Fβ)2C1Ws(θ),\displaystyle(F^{\ast}\alpha)^{2}C^{-1}W^{\ast}s(\xi)+(F^{\ast}\beta)^{2}C^{-1}W^{\ast}s(\theta),

where F:𝒪XF𝒪XF^{\ast}:{\cal{O}}_{X^{\prime}}\to F_{\ast}{\cal{O}}_{X} is the canonical pull-back morphism; here one uses the fact that WFW\circ F is equal to the absolute second power Frobenius of XX.

On the other hand, if one expands ωω\omega\otimes\omega as α2ξξ+β2θθ+αβ(ξθ+θξ)\alpha^{2}\xi\otimes\xi+\beta^{2}\theta\otimes\theta+\alpha\beta(\xi\otimes\theta+\theta\otimes\xi), then the last term is symmetric (i.e., lies in 𝒥2{\cal{J}}^{2}), and hence we get the same result this way.

This can also be explained with the following diagram

𝒵1FΩX/S1\textstyle{{\cal{Z}}^{1}\subseteq F_{\ast}\Omega^{1}_{X/S}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F(W)\scriptstyle{F_{\ast}(W^{\ast})}FWΩX/S1\textstyle{F_{\ast}W_{\ast}\Omega^{1}_{X^{\prime}/S}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FWC1\scriptstyle{F_{\ast}W_{\ast}C^{-1}}FW1(K)=(FX)1(K)\textstyle{F_{\ast}W_{\ast}{\cal{H}}^{1}(K)=(F_{X^{\prime}})_{\ast}{\cal{H}}^{1}(K)}1(K)\textstyle{{\cal{H}}^{1}(K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}s\scriptstyle{s}

over U0U_{0}, where FXF_{X^{\prime}} denotes the absolute second power Frobenius of XX^{\prime}; it shows that the map

ωC1Ws(ω)\omega\mapsto C^{-1}W^{\ast}s(\omega)

is 22-linear, while “extracting” ω\omega out of ωω\omega\otimes\omega would be 212^{-1}-linear; hence these nonlinearities cancel each other and the map ωωC1Ws(ω)\omega\otimes\omega\mapsto C^{-1}W^{\ast}s(\omega) is linear.

This shows that z~\widetilde{z}, hence the 0-cochain zz which is obtained by composing z~\widetilde{z} with the projection qq/𝒥q{\cal{I}}^{q}\to{\cal{I}}^{q}/{\cal{J}}^{q}, is well-defined and has the desired property. Therefore the class [η¯|q][\overline{\eta}|_{{\cal{I}}^{q}}] is zero, and the only thing that contributes to the class (4.2) is the qqth contraction of obs1\operatorname{obs}_{1}. This ends the proof. ∎

By the construction (Theorem 3.9) and the calculation (Theorem 4.1), we immediately get:

Corollary 4.2.

With the notations as in Theorem 4.1, suppose that X/SX^{\prime}/S is liftable to S~\widetilde{S}. Then for each integer qq, the truncation τ[q1,q]FΩX/S\tau_{[q-1,q]}F_{\ast}\Omega^{\bullet}_{X/S} of length 22 is decomposable in the derived category D(X,𝒪X)D(X^{\prime},{\cal{O}}_{X^{\prime}}).

Proof.

This follows from [DI87, 3.5] (which identifies the obstruction to liftability to the decomposability of τ1FX/S,ΩX/S\tau_{\leq 1}F_{X/S,\ast}\Omega^{\bullet}_{X/S}) and Theorems 3.9 and 4.1 (which relate the decomposability of τ[0,1]FX/S,ΩX/S\tau_{[0,1]}F_{X/S,\ast}\Omega^{\bullet}_{X/S} with that of τ[q1,q]FX/S,ΩX/S\tau_{[q-1,q]}F_{X/S,\ast}\Omega^{\bullet}_{X/S}). ∎

In particular, we extend the (special) case of Corollary 3.6 applied to the de Rham complex in characteristic p>2p>2, even to the case of p=2p=2.

Corollary 4.3.

Let XX be a smooth variety over a perfect field kk. Then we have the equality

δFX/k,ΩX/kq=(idδK1)ηq\delta^{q}_{F_{X/k,*}\Omega^{\bullet}_{X/k}}=(\mathrm{id}\otimes\delta^{1}_{K})\circ\eta^{q}

in Hom(ΩX/kq,ΩX/kq1[2])\operatorname{Hom}(\Omega^{q}_{X^{\prime}/k},\Omega^{q-1}_{X^{\prime}/k}[2]).

Finally, we answer the question of Katz:

Corollary 4.4.

Let SS be a scheme of characteristic p>0p>0, f:XSf:X\to S a smooth separated morphism of finite type, and f:XSf^{\prime}:X^{\prime}\to S (resp. F:XXF:X\to X^{\prime}) the base-change of ff by the Frobenius endomorphism of SS (resp. the relative Frobenius). Suppose S~\widetilde{S} is a flat 𝐙/p2{\mathbf{Z}}/p^{2}-scheme whose reduction modulo pp yields SS. Then the morphism in the conjugate spectral sequence

d2ij:RifΩX/SjRi+2fΩX/Sj1,d^{ij}_{2}:R^{i}f^{\prime}_{\ast}\Omega^{j}_{X^{\prime}/S}\longrightarrow R^{i+2}f^{\prime}_{\ast}\Omega^{j-1}_{X^{\prime}/S},

where one identifies j(FΩX/S){\cal{H}}^{j}(F_{\ast}\Omega_{X/S}^{\bullet}) with ΩX/Sj\Omega^{j}_{X^{\prime}/S} via the Cartier isomorphism, can be canonically regarded as the cup product with the additive inverse of the obstruction class (in H2(X,TX/S)H^{2}(X^{\prime},T_{X^{\prime}/S})) to lifting X/SX^{\prime}/S over S~\widetilde{S}.

Proof.

We first remark that by [DI87, 3.9] it can be directly seen that the obstruction class to lifting does not depend on the choice of a flat 𝐙/p2\mathbf{Z}/p^{2}-lifting S~\widetilde{S} of SS. Then the corollary follows from [DI87, 3.5] and Theorems 3.9 and 4.1. ∎

Appendix A FF-split schemes of dimension p+1p+1

Let kk be a perfect field of characteristic p>0p>0. As mentioned in the introduction, Drinfeld, Bhatt–Lurie, and Li–Mondal have obtained the following result.

Theorem A.1 ([Dri20, §5.12.1], [BL21], [LM21, Corollary 5.5]).

Let XX be a smooth scheme over a perfect field kk of characteristic p>0p>0. Suppose that XX is liftable to W2(k)W_{2}(k). Then the truncations

τ[qp+1,q]FX/k,ΩX/k\tau_{[q-p+1,q]}F_{X/k,*}\Omega^{\bullet}_{X/k}

are decomposable for all qq.

Below, we employ this in order to show Kodaira–Akizuki–Nakano vanishing and Hodge–de Rham degeneration for FF-split smooth projective schemes of dimension at most p+1p+1.

Recall [MR85] that a kk-scheme XX is FF-split if the morphism FX:𝒪XFX,𝒪XF_{X}^{*}\colon\mathcal{O}_{X}\to F_{X,*}\mathcal{O}_{X} is a split injection. Since kk is perfect, this is equivalent to the splitting of FX/k:𝒪XFX/k,𝒪XF_{X/k}^{*}\colon\mathcal{O}_{X^{\prime}}\to F_{X/k,*}\mathcal{O}_{X}. It is well-known that every FF-split scheme over kk admits a flat lifting to W2(k)W_{2}(k) [Ill96, §8.5].

If XX is FF-split and if LL is a line bundle on XX, then tensoring the split injection 𝒪XFX,𝒪X\mathcal{O}_{X}\to F_{X,*}\mathcal{O}_{X} with LL and taking cohomology shows that for all ii, Hi(X,L)H^{i}(X,L) is a direct summand of Hi(X,LF𝒪X)H^{i}(X,L\otimes F_{*}\mathcal{O}_{X}). By the projection formula and the fact that Frobenius is affine this latter summand equals Hi(X,FL)=Hi(X,Lp)H^{i}(X,F^{*}L)=H^{i}(X,L^{p}), and hence the Frobenius pull-back maps

F:Hi(X,L)Hi(X,Lp)F^{*}\colon H^{i}(X,L)\longrightarrow H^{i}(X,L^{p})

are injective. Thus if Hi(X,Lm)=0H^{i}(X,L^{m})=0 for m0m\gg 0, then already Hi(X,L)=0H^{i}(X,L)=0. Consequently, if XX is moreover smooth (or just Gorenstein) and projective, then Hi(X,L1)=0H^{i}(X,L^{-1})=0 for i<dimXi<\dim X and LL ample, i.e. Kodaira vanishing holds on XX. Similar reasoning with L=𝒪XL=\mathcal{O}_{X} shows that

F:Hi(X,𝒪X)Hi(X,𝒪X)F^{*}\colon H^{i}(X,\mathcal{O}_{X})\longrightarrow H^{i}(X,\mathcal{O}_{X})

is bijective for all i0i\geq 0.

Theorem A.2 (Kodaira–Akizuki–Nakano vanishing).

Let XX be a smooth projective scheme over kk of dimension d=p+1d=p+1. If XX is FF-split, then Kodaira–Akizuki–Nakano vanishing holds for XX, i.e. for every ample line bundle LL, we have

Hi(X,L1ΩX/kj)=0fori+j<d=p+1.H^{i}(X,L^{-1}\otimes\Omega^{j}_{X/k})=0\quad\text{for}\quad i+j<d=p+1.
Proof.

By Serre vanishing, the assertion holds for LpmL^{p^{m}} for m0m\gg 0. Therefore we may assume that it holds for LpL^{p}. Following [DI87, Proof of Lemme 2.9], we form the complex K=(L)1FX/k,ΩX/kK^{\bullet}=(L^{\prime})^{-1}\otimes F_{X/k,*}\Omega^{\bullet}_{X/k} where LL^{\prime} is the pull-back of LL to XX^{\prime}, and write the two spectral sequences

(A.1) E1ijI=Hj(X,(L)1FX/k,ΩX/ki)Hi+j(X,K){}_{I}E_{1}^{ij}=H^{j}(X^{\prime},(L^{\prime})^{-1}\otimes F_{X/k,*}\Omega^{i}_{X/k})\quad\Rightarrow\quad H^{i+j}(X^{\prime},K^{\bullet})

and

(A.2) E2ijII=Hi(X,(L)1ΩX/kj)Hi+j(X,K).{}_{II}E_{2}^{ij}=H^{i}(X^{\prime},(L^{\prime})^{-1}\otimes\Omega^{j}_{X^{\prime}/k})\quad\Rightarrow\quad H^{i+j}(X^{\prime},K^{\bullet}).

Now the projection formula gives E1ijI=Hj(X,LpΩX/ki){}_{I}E_{1}^{ij}=H^{j}(X,L^{-p}\otimes\Omega^{i}_{X/k}), which vanishes for i+jpi+j\leq p by assumption. Consequently the abutment Hr(X,K)=0H^{r}(X^{\prime},K^{\bullet})=0 for rpr\leq p.

We now investigate the second spectral sequence. Since XX is FF-split, it lifts to W2(k)W_{2}(k). Theorem A.1 implies that the differentials on ErijII{}_{II}E_{r}^{ij} are zero for rpr\leq p. For dimensional reasons, there are no nonzero differentials for r>p+1r>p+1, and the only two nonzero differentials on Ep+1ijII{}_{II}E_{p+1}^{ij} are

dp+10,p:H0(X,(L)1ΩX/kp)Hp+1(X,(L)1)d_{p+1}^{0,p}\colon H^{0}(X^{\prime},(L^{\prime})^{-1}\otimes\Omega^{p}_{X^{\prime}/k})\longrightarrow H^{p+1}(X^{\prime},(L^{\prime})^{-1})

and

dp+10,p+1:H0(X,(L)1ωX/k)Hp+1(X,(L)1ΩX/k1).d_{p+1}^{0,p+1}\colon H^{0}(X^{\prime},(L^{\prime})^{-1}\otimes\omega_{X^{\prime}/k})\longrightarrow H^{p+1}(X^{\prime},(L^{\prime})^{-1}\otimes\Omega^{1}_{X^{\prime}/k}).

We will show that dp+10,p=0d_{p+1}^{0,p}=0, which will then imply that in (A.2) we have

E2ijII=Ep+1ijII=EijII=0for i+jp.{}_{II}E_{2}^{ij}={}_{II}E_{p+1}^{ij}={}_{II}E_{\infty}^{ij}=0\qquad\text{for $i+j\leq p$.}

Note that Hp(X,K)=0H^{p}(X^{\prime},K^{\bullet})=0 implies 0=Ep+20,pII=ker(dp+10,p)0={}_{II}E_{p+2}^{0,p}=\ker(d_{p+1}^{0,p}), i.e. dp+10,pd_{p+1}^{0,p} is injective.

By Lemma A.3 below applied to E=L1E=L^{-1} and the map FX/k:XXF_{X^{-}/k}\colon X^{-}\to X where X=(Fk)1(X)X^{-}=(F_{k})^{-1}(X) is the Frobenius untwist of XX, we have a commutative square induced by FX/kF_{X/k}^{*}:

H0(X,(L)1ΩX/kp)\textstyle{H^{0}(X^{\prime},(L^{\prime})^{-1}\otimes\Omega^{p}_{X^{\prime}/k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}dp+10,p(L)\scriptstyle{d_{p+1}^{0,p}(L)}Hp+1(X,(L)1)\textstyle{H^{p+1}(X^{\prime},(L^{\prime})^{-1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FX/k\scriptstyle{F_{X^{\prime}/k}^{*}}0=H0(X,LpΩX/kp)\textstyle{0=H^{0}(X^{\prime},L^{-p}\otimes\Omega^{p}_{X/k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}dp+10,p(Lp)\scriptstyle{d_{p+1}^{0,p}(L^{p})}Hp+1(X,Lp)\textstyle{H^{p+1}(X^{\prime},L^{-p})}

(in fact the left vertical map is zero). Here the bottom map is deduced similarly for (the pull-back to XX^{-} of) LpL^{p} in place of LL. The vertical right map is injective because XX is FF-split, and the horizontal maps are injective by the previous paragraph. We conclude that H0(X,(L)1ΩX/kp)=0H^{0}(X^{\prime},(L^{\prime})^{-1}\otimes\Omega^{p}_{X^{\prime}/k})=0. ∎

In the proof above, as well as in the proof of Hodge–de Rham degeneration below, we need the following functoriality result.

Lemma A.3.

For a vector bundle EE on a smooth kk-scheme XX, write K(E)K^{\bullet}(E) for the complex EFX/k,ΩX/kE^{\prime}\otimes F_{X/k,*}\Omega^{\bullet}_{X/k}. Let f:YXf\colon Y\to X be a map of smooth kk-schemes, then ff induces a map of complexes fK(E)K(fE)f^{*}K^{\bullet}(E)\to K^{\bullet}(f^{*}E) and hence a map of spectral sequences

E2ijII(E)\textstyle{{}_{II}E_{2}^{ij}(E)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hi(X,EΩX/kj)\textstyle{H^{i}(X^{\prime},E^{\prime}\otimes\Omega^{j}_{X^{\prime}/k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\displaystyle\Rightarrow}Hi+j(X,K(E))\textstyle{H^{i+j}(X^{\prime},K^{\bullet}(E))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E2ijII(fE)\textstyle{{}_{II}E_{2}^{ij}(f^{*}E)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hi(Y,(fE)ΩY/kj)\textstyle{H^{i}(Y^{\prime},(f^{*}E)^{\prime}\otimes\Omega^{j}_{Y^{\prime}/k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\displaystyle\Rightarrow}Hi+j(Y,K(fE)),\textstyle{H^{i+j}(Y^{\prime},K^{\bullet}(f^{*}E)),}

where the maps Hi(X,EΩX/kj)Hi(Y,(fE)ΩY/kj)H^{i}(X^{\prime},E^{\prime}\otimes\Omega^{j}_{X^{\prime}/k})\to H^{i}(Y^{\prime},(f^{*}E)^{\prime}\otimes\Omega^{j}_{Y^{\prime}/k}) are induced by the composition

Hi(X,EΩX/kj)Hi(X,(f)(E)(f)ΩX/kj)Hi(X,(f)(E)ΩY/kj).H^{i}(X^{\prime},E^{\prime}\otimes\Omega^{j}_{X^{\prime}/k})\longrightarrow H^{i}(X^{\prime},(f^{\prime})^{*}(E^{\prime})\otimes(f^{\prime})^{*}\Omega^{j}_{X^{\prime}/k})\longrightarrow H^{i}(X^{\prime},(f^{\prime})^{*}(E^{\prime})\otimes\Omega^{j}_{Y^{\prime}/k}).
Proof.

Obvious. ∎

Theorem A.4 (Hodge–de Rham degeneration).

Let XX be a smooth projective scheme over kk of dimension d=p+1d=p+1. If XX is FF-split, then the Hodge to de Rham spectral sequence

E1ijI=Hj(X,ΩX/ki)Hi+j(X,ΩX/k){}_{I}E_{1}^{ij}=H^{j}(X,\Omega^{i}_{X/k})\quad\Rightarrow\quad H^{i+j}(X,\Omega^{\bullet}_{X/k})

degenerates.

Proof.

Since XX is proper, it is enough to show that the conjugate spectral sequence

E2ijII=Hi(X,ΩX/kj)Hi+j(X,ΩX/k){}_{II}E_{2}^{ij}=H^{i}(X^{\prime},\Omega^{j}_{X^{\prime}/k})\quad\Rightarrow\quad H^{i+j}(X,\Omega^{\bullet}_{X/k})

degenerates. Since XX is FF-split, it lifts to W2(k)W_{2}(k), and then Theorem A.1 implies that the differentials on the page ErijII{}_{II}E_{r}^{ij} of the conjugate spectral sequence are zero for rpr\leq p.

Since dimX=p+1\dim X=p+1, the only possibly nonzero differentials in this spectral sequence are therefore

dp+10,p:H0(X,ΩX/kp)Hp+1(X,𝒪X)d_{p+1}^{0,p}\colon H^{0}(X^{\prime},\Omega^{p}_{X^{\prime}/k})\longrightarrow H^{p+1}(X^{\prime},\mathcal{O}_{X^{\prime}})

and

dp+10,p+1:H0(X,ωX/k)Hp+1(X,ΩX/k1).d_{p+1}^{0,p+1}\colon H^{0}(X^{\prime},\omega_{X^{\prime}/k})\longrightarrow H^{p+1}(X^{\prime},\Omega^{1}_{X^{\prime}/k}).

We will show that dp+10,p=0d_{p+1}^{0,p}=0. Indeed, by functoriality of the above maps with respect to Frobenius (Lemma A.3 with E=𝒪XE=\mathcal{O}_{X} and the relative Frobenius FX/kF_{X/k}) gives a commutative square

H0(X,ΩX/kp)\textstyle{H^{0}(X^{\prime},\Omega^{p}_{X^{\prime}/k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FX/k\scriptstyle{F_{X/k}^{*}}dp+10,p\scriptstyle{d_{p+1}^{0,p}}Hp+1(X,𝒪X)\textstyle{H^{p+1}(X^{\prime},\mathcal{O}_{X^{\prime}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FX/k\scriptstyle{F_{X/k}^{*}}H0(X,ΩX/kp)\textstyle{H^{0}(X,\Omega^{p}_{X/k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}dp+10,p for X\scriptstyle{d_{p+1}^{0,p}\text{ for }X^{-}}Hp+1(X,𝒪X)\textstyle{H^{p+1}(X,\mathcal{O}_{X})}

where X=(Fk1)XX^{-}=(F^{-1}_{k})^{*}X is again the Frobenius untwist of XX. Since the Frobenius is zero on Ωi\Omega^{i} for i>0i>0, the left vertical map is zero. On the other hand, since XX is FF-split, the right vertical map is an isomorphism. Therefore the top map dp+10,pd_{p+1}^{0,p} is zero.

Finally, we obtain the vanishing of dp+10,p+1d_{p+1}^{0,p+1} by comparing dimensions and duality. Indeed, we have

dimHp+2(X,ΩX/k)\displaystyle\dim H^{p+2}(X,\Omega^{\bullet}_{X/k}) =dimHp(X,ΩX/k)\displaystyle=\dim H^{p}(X,\Omega^{\bullet}_{X/k}) (Poincaré duality)
=i+j=pdimHi(X,ΩX/kj)\displaystyle=\sum_{i+j=p}\dim H^{i}(X^{\prime},\Omega^{j}_{X^{\prime}/k}) (since dp+10,p=0d_{p+1}^{0,p}=0)
=i+j=p+2dimHi(X,ΩX/kj).\displaystyle=\sum_{i+j=p+2}\dim H^{i}(X^{\prime},\Omega^{j}_{X^{\prime}/k}). (Serre duality),\displaystyle\text{(Serre duality)},

so dp+10,p+1=0d_{p+1}^{0,p+1}=0. ∎

Remark A.5 (See [LM21, Theorem 5.4] and [BL21]).

In fact, the results of Drinfeld, Bhatt–Lurie, and Li–Mondal yield more than we have stated in Theorem A.1. Namely, for XX smooth over kk and liftable to W2(k)W_{2}(k), there exists a decomposition in the derived category

FX/k,ΩX/ki=0p1KiF_{X/k,*}\Omega^{\bullet}_{X/k}\simeq\bigoplus_{i=0}^{p-1}K_{i}

where j(Ki)=0\operatorname{\cal{H}}^{j}(K_{i})=0 unless ii and jj are congruent modulo pp. This implies that in the conjugate spectral sequence, as well as in the second spectral sequence used in the proof of Theorem A.2), the only nonzero differentials may appear on pages ErE_{r} where rr is congruent to one modulo pp. We only used this with rpr\leq p, and it would be interesting to obtain new vanishing and degeneration theorems using this stronger fact.

References

  • [AO20] Piotr Achinger and Arthur Ogus, Monodromy and log geometry, Tunis. J. Math. 2 (2020), no. 3, 455–534. MR 4041282
  • [BL21] Bhargav Bhatt and Jacob Lurie, Absolute prismatic cohomology, in preparation, 2021.
  • [DI87] Pierre Deligne and Luc Illusie, Relèvements modulo p2p^{2} et décomposition du complexe de de Rham, Invent. Math. 89 (1987), no. 2, 247–270. MR 894379
  • [Dri20] Vladimir Drinfeld, Prismatisation, arXiv preprint arXiv:2005.04746, 2020.
  • [Ill71] Luc Illusie, Complexe cotangent et déformations. I, Lecture Notes in Mathematics, Vol. 239, Springer-Verlag, Berlin-New York, 1971. MR 0491680
  • [Ill96] by same author, Frobenius et dégénérescence de Hodge, Introduction à la théorie de Hodge, Panor. Synthèses, vol. 3, Soc. Math. France, Paris, 1996, pp. 113–168. MR 1409820
  • [Kat70] Nicholas M. Katz, Nilpotent connections and the monodromy theorem: Applications of a result of Turrittin, Inst. Hautes Études Sci. Publ. Math. (1970), no. 39, 175–232. MR 291177
  • [Kat89] Kazuya Kato, Logarithmic structures of Fontaine-Illusie, Algebraic analysis, geometry, and number theory (Baltimore, MD, 1988), Johns Hopkins Univ. Press, Baltimore, MD, 1989, pp. 191–224. MR 1463703
  • [KM95] Igor Kříž and J. P. May, Operads, algebras, modules and motives, Astérisque (1995), no. 233, iv+145pp. MR 1361938
  • [KS04] Kazuya Kato and Takeshi Saito, On the conductor formula of Bloch, Publ. Math. Inst. Hautes Études Sci. (2004), no. 100, 5–151. MR 2102698
  • [LM21] Shizhang Li and Shubhodip Mondal, On endomorphisms of the de Rham cohomology functor, in preparation, 2021.
  • [Lur09] Jacob Lurie, Higher topos theory, Annals of Mathematics Studies, vol. 170, Princeton University Press, Princeton, NJ, 2009. MR 2522659
  • [MR85] Vikram B. Mehta and Annamalai Ramanathan, Frobenius splitting and cohomology vanishing for Schubert varieties, Ann. of Math. (2) 122 (1985), no. 1, 27–40. MR 799251
  • [Ste95] Joseph H. M. Steenbrink, Logarithmic embeddings of varieties with normal crossings and mixed Hodge structures, Math. Ann. 301 (1995), no. 1, 105–118. MR 1312571