Skip to main content
Log in

Projectivity of the Witt vector affine Grassmannian

  • Published:
Inventiones mathematicae Aims and scope

Abstract

We prove that the Witt vector affine Grassmannian, which parametrizes W(k)-lattices in \(W(k)[\frac{1}{p}]^n\) for a perfect field k of characteristic p, is representable by an ind-(perfect scheme) over k. This improves on previous results of Zhu by constructing a natural ample line bundle. Along the way, we establish various foundational results on perfect schemes, notably h-descent results for vector bundles.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. For example, if \(\beta :G \rightarrow G_0\) is an isogeny of p-divisible groups over k, and a trivialization \(W(k)[\frac{1}{p}]^n \simeq D(G_0)[\frac{1}{p}]\) of the Dieudonné module of \(G_0\) has been fixed, then the induced map \(D(\beta )\) on Dieudonné modules defines a point of \({\mathrm {Gr}}^{W{\mathrm {aff}}}\). In the equal characteristic case, the relation has been obtained by Hartl and Viehmann [27, Theorem 6.3].

  2. The existence of this line bundle was already conjectured by Zhu.

  3. We are ultimately only interested in the statement at the level of \(\pi _0\). However, the language of spectra, or at least Picard groupoids, is critical to carry out the descent.

  4. The h-topology is not subcanonical (as e.g. nilimmersions are covers), so one needs to sheafify.

  5. In Lemma 6.2 below, we do a “big” construction, which however works in this truncated version of \({\mathrm {Perf}}\) as \(\kappa \) is a strong limit cardinal. One can also do a more careful construction in Lemma 6.2 to keep the rings smaller, and allow more general \(\kappa \).

  6. Recall that this includes the assumption that X and Y are qcqs.

  7. We recall that all objects of \({\mathrm {Perf}}\) are assumed to be qcqs.

  8. We apologize for the two very different uses of the word “perfect”.

  9. In [53], the language of Waldhausen categories was used. See [3, Notation 12.11] for a definition of the K-theory spectrum in terms of the stable \(\infty \)-category \({\mathrm {Perf}}(X)\).

  10. One could, however, truncate all spectra in degrees \(>1\), i.e. apply \(\tau _{\le 1}\), and work in the 2-category of groupoids.

  11. In fact, even if the line bundle is constructed using the K-theoretic approach, some of the geometric lemmas of this section are used.

  12. The rank of a valuation ring V is simply the Krull dimension of V, and can also be defined purely combinatorially from the value group of V. We refer to [10, Chapter 6] for more on valuation rings.

  13. If \(i>0\), then \(N=H^i(X,{\mathcal {O}}_X)\), and there is a short exact sequence \(0\rightarrow H^i(X,{\mathcal {O}}_X)\otimes _V k(s)\rightarrow H^i(X_s,{\mathcal {O}}_{X_s})\rightarrow {\mathrm {Tor}}_1^V(H^{i+1}(X,{\mathcal {O}}_X),k(s))\rightarrow 0\), where the middle term is zero; thus, \(N\otimes _V k(s)=0\). A similar argument works for \(i=0\), where \(N=H^0(X,{\mathcal {O}}_X)/V\).

  14. In fact, \(\lambda \mapsto n_\lambda \) is (up to shift \(i\mapsto i+1\)) the transpose of a partition, so the situation is symmetric.

  15. See Remark 7.4 for the definition of \(n_\lambda (i)\).

  16. Here, \({\mathcal {Q}}\) stands for the base extension of Q to varying fields.

  17. As we do not yet know that \({\mathrm {Gr}}_{\le \lambda }\) is representable, all of these assertions mean that they hold true after an arbitrary base-change to a representable S over \({\mathrm {Gr}}_{\le \lambda }\). Note that viewing \({\mathrm {Gr}}_{\le \lambda }\) as an object of a suitable ringed topos (the one associated to the site \({\mathrm {Perf}}\) ringed using \({\mathcal {O}}\) equipped with the v-topology) leads to a potentially different notion of a pushforward. Nevertheless, both these definitions coincide by Lemma 3.18.

  18. Cf. Lemma 3.6 for the meaning of this statement.

  19. For the construction, choose finite type models, apply the Stein factorization, and then go to the perfection.

  20. The notion of a hypercomplete sheaf is specific to working in the \(\infty \)-categorical setting. Roughly speaking, a sheaf in the \(\infty \)-categorical setup is only required to satisfy descent along Cech covers, while a hypersheaf is required to satisfy descent along hypercovers. If the sheaf takes on n-truncated values for some finite integer n, then the notions coincide.

  21. In this section, we depart from our standing conventions, and go “fully” derived. Thus, for an \(E_\infty \)-ring A (which could be a discrete ring), the notation \({\mathrm {Mod}}(A)\) refers to the stable \(\infty \)-category of A-module spectra (and coincides with the usual derived category D(A) when A is discrete). Likewise, all tensor products are always derived. The one exception is that “schemes” refers to ordinary schemes; when we need derived schemes, we say so.

  22. For detection of pseudo-coherence after passage to an h-cover, one argues as in Lemma 11.30 below.

  23. The only non-trivial bit is to check descent of compactness, i.e., show: given \(C \in {\mathrm {CAlg}}(D(A))\) satisfying \(C \otimes _A C \simeq C\), if \(C \otimes _A B\) is compact in \({\mathrm {CAlg}}(D(B))\), then C is compact in \({\mathrm {CAlg}}(D(A))\). For this, one first notes that the functor \({\mathrm {Map}}_A(C,-)\) on \({\mathrm {CAlg}}(D(A))\) takes on discrete values since \(C \otimes _A C \simeq C\): in fact, the values are always contractible or empty. The same also applies to \(C \otimes _A B^i \in {\mathrm {CAlg}}(D(B^i))\) for all i. The descent of compactness now follows as totalization of cosimplicial n-truncated spaces commute with filtered colimits for any finite n.

  24. An earlier version of this preprint asserted that one may choose \(N = d\). This is false: if \(R = k[x,y]/(xy)_{{\mathrm {perf}}}\), then \(d = 1\), but \({\mathrm {Tor}}_2^R(R/(x), R/(y)) \ne 0\). We thank Gabber for pointing out the mistake, the previous example, and the correct bound. Gabber has also pointed out that the global dimension N in Proposition 11.31 is \(\le 2d+1\), where equality is obtained in some examples; we do not prove that here.

  25. Indeed, after possibly adjusting our initial choice of \(P_0\) and \(R_0\), the map \(P_0 \rightarrow R_0\) is generically étale, so there exist some non-zero \(f \in P_0\) such that \(P_0 \rightarrow R_0\) is finite étale after inverting f. One then checks that such an f does the job.

  26. This means that the fibres are either contractible or empty.

  27. Lurie in [35, Remark 2.4.2.2] calls these commutative monoids. However, with [35, Notation 5.1.1.6], this becomes the same thing as an \({\mathbf {E}}_\infty \)-monoid.

  28. This is Segal’s category \(\Gamma \) (or its opposite, depending on references).

  29. i.e., of pointed spaces \(X\in {\mathcal {S}}_*\) such that \(\pi _j X=0\) for \(j<i\).

  30. For an \(\infty \)-categorical discussion of these matters, see recent work of Gepner–Groth–Nikolaus [20].

  31. The \(\infty \)-category \({\mathrm {Mon}}_{{\mathbf {E}}_\infty }\) admits all small limits, and \({\mathrm {Mon}}_{{\mathbf {E}}_\infty }\rightarrow {\mathcal {S}}_*\), \(X\mapsto X([1])\), preserves all small limits. In particular, this applies to the loop space \(X\mapsto \Omega X = *\times _X *\), giving a functor \(\Omega : {\mathrm {Mon}}_{{\mathbf {E}}_\infty }\rightarrow {\mathrm {Mon}}_{{\mathbf {E}}_\infty }\).

References

  1. Aberbach, I.M., Hochster, M.: Finite Tor dimension and failure of coherence in absolute integral closures. J. Pure Appl. Algebra 122(3), 171–184 (1997)

    Article  MathSciNet  Google Scholar 

  2. Asgharzadeh, M.: Homological properties of the perfect and absolute integral closures of Noetherian domains. Math. Ann. 348(1), 237–263 (2010)

    Article  MathSciNet  Google Scholar 

  3. Barwick, C.: On the algebraic K-theory of higher categories. J. Topol. (2012, to appear). arXiv:1204.3607

  4. Berthelot, P., Bloch, S., Esnault, H.: On Witt vector cohomology for singular varieties. Compos. Math. 143(2), 363–392 (2007)

    Article  MathSciNet  Google Scholar 

  5. Brylinski, J.-L., Deligne, P.: Central extensions of reductive groups by \({ K}_2\). Publ. Math. Inst. Hautes Études Sci. 94, 5–85 (2001)

    Article  Google Scholar 

  6. Bosch, S., Güntzer, U., Remmert, R.: Non-Archimedean analysis. In: Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 261. Springer, Berlin (1984) (A systematic approach to rigid analytic geometry)

  7. Bhatt, B.: Algebraization and Tannaka duality. Camb. J. Math. (2014, to appear). arXiv:1404.7483

  8. Bhatt, B., Halpern-Leistner, D.: Tannaka duality revisited (2015). arXiv:1507.01925

  9. Beauville, A., Laszlo, Y.: Conformal blocks and generalized theta functions. Commun. Math. Phys. 164(2), 385–419 (1994)

    Article  MathSciNet  Google Scholar 

  10. Bourbaki, N.: Éléments de mathématique. Masson, Paris (1985) (Algèbre commutative. Chapitres 5 à 7. [Commutative algebra. Chapters 5–7], Reprint)

  11. Bhatt, B., Scholze, P.: The pro-étale topology for schemes. Astérisque (2013, to appear). arXiv:1309.1198

  12. Bhatt, B., Schwede, K., Takagi, S.: The weak ordinarity conjecture and F-singularities. In: Proceedings of the conference in honor of Professor Yujiro Kawamata on his 60th birthday (2013, to appear). arXiv:1307.3763

  13. Carlsson, G.: Derived completions in stable homotopy theory. J. Pure Appl. Algebra 212(3), 550–577 (2008)

    Article  MathSciNet  Google Scholar 

  14. Chen, M., Kisin, M., Viehmann, E.: Connected components of affine Deligne–Lusztig varieties in mixed characteristic (2013). arXiv:1307.3845

  15. de Jong, A.J.: Smoothness, semi-stability and alterations. Inst. Hautes Études Sci. Publ. Math. 83, 51–93 (1996)

    Article  MathSciNet  Google Scholar 

  16. Faltings, G.: Algebraic loop groups and moduli spaces of bundles. J. Eur. Math. Soc. (JEMS) 5(1), 41–68 (2003)

    Article  MathSciNet  Google Scholar 

  17. Gabber, O.: Notes on some \(t\)-structures. In: Geometric Aspects of Dwork Theory. vol. I, II, pp. 711–734. Walter de Gruyter GmbH & Co. KG, Berlin (2004)

  18. Grothendieck, A., Dieudonné, J.: Éléments de géométrie algébrique. III. Étude cohomologique des faisceaux cohérents. I. Inst. Hautes Études Sci. Publ. Math. 11, 5–167 (1961)

    Google Scholar 

  19. Grothendieck, A., Dieudonné, J.: Éléments de géométrie algébrique. IV. Étude locale des schémas et des morphismes de schémas. III. Inst. Hautes Études Sci. Publ. Math. 28, 5–255 (1966)

    Article  Google Scholar 

  20. Gepner, D., Groth, M., Nikolaus, T.: Universality of multiplicative infinite loop space machines (2013). arXiv:1305.4550

  21. Gabber, O., Ramero, L.: Almost ring theory. In: Lecture Notes in Mathematics, vol. 1800. Springer, Berlin (2003)

  22. Greenberg, M.J.: Schemata over local rings. Ann. Math. 2(73), 624–648 (1961)

    Article  MathSciNet  Google Scholar 

  23. Haboush, W.J.: Infinite dimensional algebraic geometry: algebraic structures on \(p\)-adic groups and their homogeneous spaces. Tohoku Math. J. (2) 57(1), 65–117 (2005)

    Article  MathSciNet  Google Scholar 

  24. Hamacher, P.: The geometry of Newton strata in the reduction modulo p of Shimura varieties of PEL type (2013). arXiv:1312.0490

  25. Halpern-Leistner, D., Preygel, A.: Mapping stacks and categorical notions of properness. (2014). arXiv:1402.3204

  26. Huber, R.: A generalization of formal schemes and rigid analytic varieties. Math. Z. 217(4), 513–551 (1994)

    Article  MathSciNet  Google Scholar 

  27. Hartl, U., Viehmann, E.: The Newton stratification on deformations of local \(G\)-shtukas. J. Reine Angew. Math. 656, 87–129 (2011)

    Article  MathSciNet  Google Scholar 

  28. Keel, S.: Basepoint freeness for nef and big line bundles in positive characteristic. Ann. Math. (2) 149(1), 253–286 (1999)

    Article  MathSciNet  Google Scholar 

  29. Knudsen, F.F., Mumford, D.: The projectivity of the moduli space of stable curves. I. Preliminaries on “det” and “Div”. Math. Scand. 39(1), 19–55 (1976)

    Article  MathSciNet  Google Scholar 

  30. Kreidl, M.: On \(p\)-adic lattices and Grassmannians. Math. Z. 276(3–4), 859–888 (2014)

    Article  MathSciNet  Google Scholar 

  31. Kruckman, A.: Notes on ultra filters (2012). https://math.berkeley.edu/~kruckman/ultrafilters.pdf. Accessed 27 Nov 2016

  32. Lazarsfeld, R.: Positivity in algebraic geometry. I. In: Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics], vol. 48. Springer, Berlin (2004) (Classical setting: line bundles and linear series)

  33. Lurie, J.: Higher topos theory. In: Annals of Mathematics Studies, vol. 170. Princeton University Press, Princeton (2009)

  34. Lurie, J.: Derived algebraic geometry viii: Quasi-coherent sheaves and tannaka duality theorems (2014). http://www.math.harvard.edu/~lurie/papers/DAG-VIII.pdf. Accessed 27 Nov 2016

  35. Lurie, J.: Higher algebra (2014). http://www.math.harvard.edu/~lurie/papers/higheralgebra.pdf. Accessed 27 Nov 2016

  36. Liu, Y., Zheng, W.: Enhanced six operations and base change theorem for Artin stacks. (2012). arXiv:1211.5948

  37. Matsumoto, H.: Sur les sous-groupes arithmétiques des groupes semi-simples déployés. Ann. Sci. École Norm. Sup. 4(2), 1–62 (1969)

    Article  Google Scholar 

  38. Matthew, A.: The Galois group of a stable homotopy theory (2014). arXiv:1404.2156v1.pdf

  39. May, J.P.: \(E_{\infty }\) spaces, group completions, and permutative categories. In: New Developments in Topology (Proc. Sympos. Algebraic Topology, Oxford, 1972), pp. 61–93. London Math. Soc. Lecture Note Ser., No. 11. Cambridge University Press, London (1974)

  40. MacLane, S.: Natural associativity and commutativity. Rice Univ. Stud. 49(4), 28–46 (1963)

    MathSciNet  MATH  Google Scholar 

  41. Mehta, V.B., Ramanathan, A.: Frobenius splitting and cohomology vanishing for Schubert varieties. Ann. Math. (2) 122(1), 27–40 (1985)

    Article  MathSciNet  Google Scholar 

  42. May, J.P., Thomason, R.: The uniqueness of infinite loop space machines. Topology 17(3), 205–224 (1978)

    Article  MathSciNet  Google Scholar 

  43. Patel, D.: de Rham \({\cal{E}}\)-factors. Invent. Math. 190(2), 299–355 (2012)

  44. Pappas, G., Rapoport, M.: Twisted loop groups and their affine flag varieties. Adv. Math. 219(1), 118–198 (2008). (With an appendix by T. Haines and Rapoport)

    Article  MathSciNet  Google Scholar 

  45. Quillen, D.: Higher algebraic \(K\)-theory. I. In: Algebraic \(K\)-Theory, I: Higher \(K\)-Theories (Proc. Conf., Battelle Memorial Inst., Seattle, Wash., 1972), vol. 341, pp. 85–147. Lecture Notes in Mathematics. Springer, Berlin (1973)

  46. Raynaud, M., Gruson, L.: Critères de platitude et de projectivité. Techniques de “platification” d’un module. Invent. Math. 13, 1–89 (1971)

    Article  MathSciNet  Google Scholar 

  47. Rydh, D.: Submersions and effective descent of étale morphisms. Bull. Soc. Math. Fr. 138(2), 181–230 (2010)

    Article  MathSciNet  Google Scholar 

  48. Segal, G.: Categories and cohomology theories. Topology 13, 293–312 (1974)

    Article  MathSciNet  Google Scholar 

  49. Sato, M., Sato, Y.: Soliton equations as dynamical systems on infinite-dimensional Grassmann manifold. In: Nonlinear Partial Differential Equations in Applied Science (Tokyo, 1982), North-Holland Math. Stud., vol. 81, pp. 259–271. North-Holland, Amsterdam (1983)

  50. Steinberg, R.: Générateurs, relations et revêtements de groupes algébriques. In: Colloq. Théorie des Groupes Algébriques (Bruxelles, 1962), pp. 113–127. Librairie Universitaire, Louvain, Gauthier-Villars, Paris (1962)

  51. Segal, G., Wilson, G.: Loop groups and equations of KdV type. Inst. Hautes Études Sci. Publ. Math. 61, 5–65 (1985)

  52. The Stacks Project Authors. Stacks project (2015). http://stacks.math.columbia.edu. Accessed 27 Nov 2016

  53. Thomason, R.W., Trobaugh, T.: Higher algebraic \(K\)-theory of schemes and of derived categories. In: The Grothendieck Festschrift, Vol. III, vol. 88 of Progr. Math., pp. 247–435. Birkhäuser Boston, Boston, MA (1990)

  54. Voevodsky, V.: Homology of schemes. Selecta Math. (N.S.) 2(1), 111–153 (1996)

    Article  MathSciNet  Google Scholar 

  55. Voevodsky, V.: Homotopy theory of simplicial sheaves in completely decomposable topologies. (2000). http://www.math.uiuc.edu/K-theory/443/cdstructures.pdf. Accessed 27 Nov 2016

  56. Yanagihara, H.: Some results on weakly normal ring extensions. J. Math. Soc. Jpn. 35(4), 649–661 (1983)

    Article  MathSciNet  Google Scholar 

  57. Zhu, X.: Affine Grassmannians and the geometric satake in mixed characteristic. Ann. Math. (to appear, 2014). arXiv:1407.8519

Download references

Acknowledgements

This work started after the authors listened to a talk of Xinwen Zhu on his work at the MSRI, and the authors would like to thank him for asking the question on the existence of \({\mathcal {L}}\). They would also like to thank Akhil Mathew for enlightening conversations related to Sect. 11.2. Moreover, they wish to thank all the participants of the ARGOS seminar in Bonn in the summer term 2015 for their careful reading of the manuscript, and the many suggestions for improvements and additions. The first version of this preprint contained an error in the proof of Lemma 4.6, tracing back to an error in [18, Corollary 3.3.2]; we thank Christopher Hacon, and Linquan Ma (via Karl Schwede) and the anonymous referee for pointing this out. The authors are also indebted to the referee for providing numerous other comments that improved the readability of this paper. Finally, they would like to thank the Clay Mathematics Institute, the University of California (Berkeley), and the MSRI for their support and hospitality. This work was done while B. Bhatt was partially supported by NSF grant DMS 1340424 and P. Scholze was a Clay Research Fellow.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bhargav Bhatt.

Appendix: Determinants

Appendix: Determinants

Let R be a commutative ring. Our goal in this section is to recall the definition of the natural map

$$\begin{aligned} {\mathrm {det}}: K(R)\rightarrow {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R) \end{aligned}$$

from the K-theory spectrum of R to the Picard groupoid of graded line bundles \({\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R)\). Intuitively, this is just the map sending a finite projective R-module M to \(({\mathrm {det}}M, {\mathrm {rk}}M)\).

Recall that a symmetric monoidal category is a category C equipped with a functor \(\otimes : C\times C\rightarrow C\), a unit object \(1\in C\), as well as a unitality constraint \(\eta _X: 1\otimes X\cong X\), a commutativity constraint \(c_{X,Y}: X\otimes Y\cong Y\otimes X\) and an associativity constraint \(a_{X,Y,Z}: X\otimes (Y\otimes Z)\cong (X\otimes Y)\otimes Z\) functorially in \(X,Y,Z\in C\), satisfying certain compatibility conditions. Moreover, there is a notion of symmetric monoidal functor between symmetric monoidal categories, which is a functor which commutes with \(\otimes \) in the appropriate sense. As usual, there is also a notion of a natural transformation between symmetric monoidal functors. The definitions were originally given by MacLane [40].

Definition 12.1

A symmetric monoidal category C is called strict if \(c_{X,X}: X\otimes X\cong X\otimes X\) is the identity for all \(X\in C\).

In general, the axioms of a symmetric monoidal category say that \(c_{X,Y}\circ c_{Y,X} = {\mathrm {id}}\) for all \(X,Y\in C\), so in particular \(c_{X,X}\) is always an involution.

Note that whenever C is a symmetric monoidal category, the subcategory \(C^\simeq \subset C\) consisting of all objects, but only isomorphisms as morphisms, is another symmetric monoidal category (as all extra data concerns isomorphisms). Let us refer to symmetric monoidal categories all of whose morphisms are isomorphisms as symmetric monoidal groupoids.

Below, we will recall that any symmetric monoidal category C admits a K-theory spectrum K(C), and any symmetric monoidal functor \(F: C\rightarrow D\) induces a map of spectra \(K(F): K(C)\rightarrow K(D)\). Applying this to the inclusion \(C^\simeq \subset C\) will give an equivalence \(K(C^\simeq )\simeq K(C)\); actually, an equality \(K(C^\simeq ) = K(C)\). For this reason, we largely restrict to symmetric monoidal groupoids in the following.

The following examples are our main interest.

Example 12.2

Fix a commutative ring R.

  1. (i)

    The groupoid \({\mathcal {V}}{\mathrm {ect}}(R)\) of finite projective R-modules is a symmetric monoidal category with respect to \(\oplus \), unit 0, and the standard unitality, commutativity and associativity constraints. For example, \(c_{X,Y}: X\oplus Y\cong Y\oplus X\) sends (xy) to (yx). Note that \(c_{X,X}\) is not the identity map for \(X\ne 0\).

    We remark that the groupoid of finite projective R-modules is also symmetric monoidal with respect to the tensor product \(\otimes \); we will not use this symmetric monoidal structure.

  2. (ii)

    The groupoid \({\mathcal {P}}{\mathrm {ic}}(R)\) of line bundles over R (i.e., finite projective R-modules of rank 1) is a symmetric monoidal category with respect to \(\otimes \), unit 1, and the standard unitality, commutativity and associativity constraints. For example, \(c_{X,Y}: X\otimes Y\cong Y\otimes X\) sends \(x\otimes y\) to \(y\otimes x\). Note that \(c_{X,X}\) is the identity map for all \(X\in {\mathcal {P}}{\mathrm {ic}}(R)\).

  3. (iii)

    The groupoid \({\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R)\) of \({\mathbf {Z}}\)-graded line bundles. Here, an object is given by a pair (Lf) where \(L\in {\mathcal {P}}{\mathrm {ic}}(R)\), and \(f:{\mathrm {Spec}}(R) \rightarrow {\mathbf {Z}}\) is a locally constant function; the set \({\mathrm {Isom}}( (L,f), (M,g))\) is empty if \(f \ne g\), and given by \({\mathrm {Isom}}(L,M)\) otherwise. The unit is given by (1, 0). This groupoid is endowed with a symmetric monoidal structure \(\otimes \) where \((L,f) \otimes (M,g) {:=} (L \otimes M, f + g)\), and the commutativity constraint

    $$\begin{aligned} (L \otimes M, f+g) {=:} (L,f) \otimes (M,g) \simeq (M,g) \otimes (L,f) {:=} (M \otimes L, g+f) \end{aligned}$$

    determined by the rule

    $$\begin{aligned} \ell \otimes m \mapsto (-1)^{f \cdot g} m \otimes \ell , \end{aligned}$$

    and the obvious associativity constraint. Note that \(c_{(L,f),(L,f)}: (L,f)\otimes (L,f)\cong (L,f)\otimes (L,f)\) is given by multiplication with \((-1)^f\), and is thus not in general the identity.

The following proposition is central.

Proposition 12.3

There is a natural symmetric monoidal functor

$$\begin{aligned} {\mathrm {det}}: {\mathcal {V}}{\mathrm {ect}}(R)\rightarrow {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R) \end{aligned}$$

sending \(M\in {\mathcal {V}}{\mathrm {ect}}(R)\) to \(({\mathrm {det}}M,{\mathrm {rk}}M)\in {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R) = {\mathcal {P}}{\mathrm {ic}}(R) \times H^0({\mathrm {Spec}}(R),{\mathbf {Z}})\).

Note that this functor does not factor through \({\mathcal {P}}{\mathrm {ic}}(R)\).

Proof

All verifications are automatic, and the only critical observation is the following. If \(M, N\in {\mathcal {V}}{\mathrm {ect}}(R)\), then the commutativity constraint \(c_{M,N}: M\oplus N\cong N\oplus M\) swapping N and M induces multiplication by \((-1)^{{\mathrm {rk}}M {\mathrm {rk}}N}\) on

$$\begin{aligned} {\mathrm {det}}(M)\otimes {\mathrm {det}}(N)\cong {\mathrm {det}}(M\oplus N)\buildrel {{\mathrm {det}}c_{M,N}}\over \cong {\mathrm {det}}(N\oplus M)\cong {\mathrm {det}}(N)\otimes {\mathrm {det}}(M), \end{aligned}$$

if one identifies \({\mathrm {det}}(M)\otimes {\mathrm {det}}(N)\) and \({\mathrm {det}}(N)\otimes {\mathrm {det}}(M)\) using the usual commutativity constraint of \({\mathcal {P}}{\mathrm {ic}}(R)\). \(\square \)

Now we recall the construction of the K-theory spectrum. Let C be a symmetric monoidal category, with \(C^\simeq \subset C\) the underlying symmetric monoidal groupoid. By definition, \(K(C) = K(C^\simeq )\), so assume that C is a groupoid to start with. Recall that a groupoid is equivalent to a space whose only nonzero homotopy groups are \(\pi _0\) and \(\pi _1\). Concretely, this can be realized by the nerve construction, which associates to any category C the simplicial set N(C) whose n-simplices are chains of \(n-1\) morphisms,

$$\begin{aligned} X_0\buildrel f_0\over \rightarrow X_1\buildrel f_1\over \rightarrow \cdots \buildrel {f_{n-1}}\over \rightarrow X_n. \end{aligned}$$

In particular, the 0-simplices \(N(C)_0\) are the objects of C, and the 1-simplices \(N(C)_1\) are the morphisms of C. Higher simplices encode the composition law, and degenerate simplices encode identity morphisms. If C is a groupoid, then N(C) is a Kan complex, whose geometric realization |N(C)| is a (compactly generated) topological space whose only nonzero homotopy groups are \(\pi _0\) and \(\pi _1\). In fact, |N(C)| is homotopy equivalent to the disjoint union

$$\begin{aligned} \bigsqcup _{X\in C/\simeq } B{\mathrm {Aut}}(X) \end{aligned}$$

of the classifying spaces of the automorphism group \({\mathrm {Aut}}(X)\), over all isomorphism classes of objects \(X\in C\). In the following, we refer to Kan complexes as ‘spaces’. They are naturally organized into the \(\infty \)-category \({\mathcal {S}}\) of spaces. Here and in the following, we make use of the theory of \(\infty \)-categories, cf. [33], which were previously defined under the name of weak Kan complexes (by Boardman–Vogt) and quasicategories (by Joyal). Using this language, we can easily state precise definitions and theorems without having to go into the pain of detailed constructions. Giving detailed proofs is more elaborate, and we only try to convey the meaning of the statements below.

As C is symmetric monoidal, the space N(C) is equipped with an addition law \(N(\otimes ): N(C)\times N(C)=N(C\times C)\rightarrow N(C)\). Moreover, this addition law is commutative and associative up to coherent isomorphisms, as expressed by the commutativity and associativity constraints. However, the addition law is not strictly commutative and associative. Such spaces equipped with a coherently commutative and associative (but not necessarily invertible) addition law are known as (special) \(\Gamma \)-spaces, as defined by Segal [48], or as \({\mathbf {E}}_\infty \)-monoids, as defined by May [39]. Although our definition is essentially that of a special \(\Gamma \)-space, we prefer to call them \({\mathbf {E}}_\infty \)-monoids.Footnote 27

Definition 12.4

Let \({\mathrm {Fin}}_*\) be the category of finite pointed sets.Footnote 28 For \(n\ge 0\), let \([n] = \{0,1,\ldots ,n\}\in {\mathrm {Fin}}_*\) be the object pointed at 0. An \({\mathbf {E}}_\infty \)-monoid is a functor

$$\begin{aligned} X: N({\mathrm {Fin}}_*)\rightarrow {\mathcal {S}}\end{aligned}$$

of \(\infty \)-categories to the \(\infty \)-category \({\mathcal {S}}\) of Kan complexes, such that for all \([n]\in {\mathrm {Fin}}_*\), the natural map

$$\begin{aligned} X([n])\rightarrow \prod _{i=1}^n X([1]) \end{aligned}$$

is a weak equivalence, where the map is induced by the n maps \([n]\rightarrow [1]\) contracting everything to 0 except i.

The \({\mathbf {E}}_\infty \)-monoids are naturally organized into an \(\infty \)-category (a full subcategory of the \(\infty \)-category of functors from \(N({\mathrm {Fin}}_*)\) to \({\mathcal {S}}\)), which we call the \(\infty \)-category of \({\mathbf {E}}_\infty \)-monoids \({\mathrm {Mon}}_{{\mathbf {E}}_\infty }\).

For an \({\mathbf {E}}_\infty \)-monoid X, the 0-th space X([0]) is weakly contractible, as it is weakly equivalent to an empty product. We refer to X([1]) as the underlying space of the \({\mathbf {E}}_\infty \)-monoid X, and will sometimes confuse X with X([1]). It is pointed in the sense that it comes equipped with a map \(X([0])\rightarrow X([1])\) from a weakly contractible space; the space X([0]) is the ’unit’ of X([1]). Moreover, using the map \([2]\rightarrow [1]\) sending only 0 to 0, we get a natural map

$$\begin{aligned} X([1])\times X([1])\simeq X([2])\rightarrow X([1]) \end{aligned}$$

which gives an addition law on X([1]). The higher data precisely ensure that the unit is unital, and that the addition law is commutative and associative ’up to coherent homotopy’. In particular, for an \({\mathbf {E}}_\infty \)-monoid X, the set of connected components \(\pi _0 X {:=} \pi _0 X([1])\) forms a commutative monoid.

Construction 12.5

Let C be a symmetric monoidal groupoid. We define an \({\mathbf {E}}_\infty \)-monoid N(C) in the following way. For each finite pointed set \((S,s)\in {\mathrm {Fin}}_*\), let

$$\begin{aligned} N(C)(S) = N\left( \{(X_T)_{T\subset S{\setminus } \{s\}}, X_{T\sqcup T^\prime }\cong X_T\otimes X_{T^\prime }\}\right) \end{aligned}$$

be the nerve of the groupoid of objects \(X_T\in C\) for all \(T\subset S{\setminus } \{s\}\) which are equipped with compatible isomorphisms \(X_{T\sqcup T^\prime }\cong X_T\otimes X_{T^\prime }\) for any disjoint subsets \(T, T^\prime \subset S{\setminus } \{s\}\). For any morphism \(f: (S,s)\rightarrow (S^\prime ,s^\prime )\), let the map

$$\begin{aligned} N(C)(S)\rightarrow N(C)(S^\prime ) \end{aligned}$$

be given by

$$\begin{aligned} (X_T)_{T\subset S{\setminus } \{s\}}\mapsto (X_{f^{-1}(T^\prime )})_{T^\prime \subset S^\prime {\setminus } \{s^\prime \}}. \end{aligned}$$

Remark 12.6

Here, we have defined an actual functor from \({\mathrm {Fin}}_*\) to the category of Kan complexes. The naive definition of N(C)(S) would be \(N(C)(S) = N(C)^{S{\setminus } \{s\}}\), with the map

$$\begin{aligned} N(C)^{S{\setminus } \{s\}}\rightarrow N(C)^{S^\prime {\setminus } \{s^\prime \}} \end{aligned}$$

for a map \(f: (S,s)\rightarrow (S^\prime ,s^\prime )\) of pointed sets being given by

$$\begin{aligned} (X_t)_{t\in S{\setminus } \{s\}}\mapsto (\otimes _{t\in f^{-1}(t^\prime )} X_t)_{t^\prime \in S^\prime {\setminus } \{s^\prime \}}. \end{aligned}$$

This definition is not compatible with composition on the nose, but can be made into a functor of \(\infty \)-categories \(N(C): {\mathrm {Fin}}_*\rightarrow {\mathcal {S}}\). However, the verification of this fact is most easily done by constructing a weakly equivalent strict functor, as above.

We leave it to the reader to spell out the definition of compatibility of the isomorphisms \(X_{T\sqcup T^\prime }\cong X_T\otimes X_{T^\prime }\); verifying that \(N(C)([n])\rightarrow N(C)^n\) is an equivalence uses the precise axioms of a symmetric monoidal category.

Definition 12.7

An \({\mathbf {E}}_\infty \)-monoid X is called grouplike if the commutative monoid \(\pi _0(X)\) is a group. Let \({\mathrm {Mon}}_{{\mathbf {E}}_\infty }^{\mathrm {gp}}\subset {\mathrm {Mon}}_{{\mathbf {E}}_\infty }\) denote the full subcategory of grouplike \({\mathbf {E}}_\infty \)-monoids.

Grouplike \({\mathbf {E}}_\infty \)-monoids are equivalent to connective spectra.

Definition 12.8

The \(\infty \)-category of connective spectra \({\mathrm {Sp}}^{\ge 0}\) is given by the limit of

$$\begin{aligned} {\mathcal {S}}_*\buildrel \Omega \over \leftarrow {\mathcal {S}}_*^{\ge 1}\buildrel \Omega \over \leftarrow {\mathcal {S}}_*^{\ge 2}\buildrel \Omega \over \leftarrow \cdots \end{aligned}$$

in the \(\infty \)-category of \(\infty \)-categories, where \({\mathcal {S}}_*\) is the \(\infty \)-category of pointed spaces, \({\mathcal {S}}_*^{\ge i}\subset {\mathcal {S}}_*\) is the full subcategory of i-connected spacesFootnote 29, and \(\Omega : {\mathcal {S}}_*\rightarrow {\mathcal {S}}_*\), \(X\mapsto *\times _X *\) is the loop space functor.

The following theorem is due to Segal, [48, Proposition 3.4], cf. [35, Theorem 5.2.6.10, Remark 5.2.6.26]. The functor \({\mathrm {Mon}}_{{\mathbf {E}}_\infty }^{\mathrm {gp}}\rightarrow {\mathrm {Sp}}^{\ge 0}\) is usually called “the infinite loop space machine”.

Theorem 12.9

There is a natural equivalence of \(\infty \)-categories

$$\begin{aligned} {\mathrm {Mon}}_{{\mathbf {E}}_\infty }^{\mathrm {gp}}\simeq {\mathrm {Sp}}^{\ge 0}. \end{aligned}$$

May–Thomason [42], showed that there is essentially only one such equivalence (although there are many constructions).Footnote 30 Let us briefly sketch the definition of the functors in either direction. Let \(X\in {\mathrm {Sp}}^{\ge 0}\) be a connective spectrum. Then the natural map

$$\begin{aligned} X\sqcup X\rightarrow X\times X \end{aligned}$$

from the coproduct to the product, taken in the \(\infty \)-category of connective spectra (equivalently, of spectra), is an equivalence, cf. [35, Lemma 1.1.2.10]. Thus, the obvious map \(X\sqcup X\rightarrow X\) extends to a map \(X\times X\rightarrow X\), which is the ’addition law’ of the spectrum. Repeating the same arguments for arbitrary finite (co)products shows that the forgetful functor

$$\begin{aligned} {\mathrm {Mon}}_{{\mathbf {E}}_\infty }({\mathrm {Sp}}^{\ge 0})\rightarrow {\mathrm {Sp}}^{\ge 0} \end{aligned}$$

from the \(\infty \)-category of \({\mathbf {E}}_\infty \)-monoids in \({\mathrm {Sp}}^{\ge 0}\) to \({\mathrm {Sp}}^{\ge 0}\), is an equivalence, i.e. any (connective) spectrum comes equipped with a canonical \({\mathbf {E}}_\infty \)-monoid structure. Composing with the forgetful map

$$\begin{aligned} {\mathrm {Mon}}_{{\mathbf {E}}_\infty }({\mathrm {Sp}}^{\ge 0})\rightarrow {\mathrm {Mon}}_{{\mathbf {E}}_\infty }({\mathcal {S}}) = {\mathrm {Mon}}_{{\mathbf {E}}_\infty } \end{aligned}$$

gives a functor \({\mathrm {Sp}}^{\ge 0}\rightarrow {\mathrm {Mon}}_{{\mathbf {E}}_\infty }\), preserving all \(\pi _i\). As all \(\pi _i\), including \(\pi _0\), of a spectrum are groups, it follows that the functor takes values in \({\mathrm {Mon}}_{{\mathbf {E}}_\infty }^{\mathrm {gp}}\).

Conversely, start with an \({\mathbf {E}}_\infty \)-monoid \(X\in {\mathrm {Mon}}_{{\mathbf {E}}_\infty }\) (not necessarily grouplike). Then X admits a classifying space BX. Indeed, there is a natural functor \(\Delta ^{\mathrm {op}}\rightarrow {\mathrm {Fin}}_*\) (cf. [48, p. 295]) taking the m-simplex \(\Delta ^m\) to [m]. This allows one to view X as a simplicial space \(X^\prime : N(\Delta ^{\mathrm {op}})\rightarrow N({\mathrm {Fin}}_*)\rightarrow {\mathcal {S}}\), which can be turned into a space \(BX = \mathop {{\mathrm {colim}}}_{\Delta ^{\mathrm {op}}} X^\prime \in {\mathcal {S}}\). In fact, one checks directly that BX still carries a canonical \({\mathbf {E}}_\infty \)-monoid structure, cf. [48, Definition 1.3]. This gives a functor

$$\begin{aligned} B: {\mathrm {Mon}}_{{\mathbf {E}}_\infty }\rightarrow {\mathrm {Mon}}_{{\mathbf {E}}_\infty }. \end{aligned}$$

There is a natural map \(X\rightarrow \Omega BX\) of \({\mathbf {E}}_\infty \)-monoids.Footnote 31 The following is [48, Proposition 1.4].

Proposition 12.10

If \(X\in {\mathrm {Mon}}_{{\mathbf {E}}_\infty }\) is k-connected, then BX is \((k+1)\)-connected. Moreover, the natural map

$$\begin{aligned} X\rightarrow \Omega BX \end{aligned}$$

is an equivalence if and only if X is grouplike.

Remark 12.11

As \(\pi _0(\Omega BX) = \pi _1(X)\) is a group, \(\Omega BX\) is always a grouplike \({\mathbf {E}}_\infty \)-monoid. Also, BX is always 1-connected, and thus grouplike (as \(\pi _0 BX=0\) is a group).

Thus, if \(X\in {\mathrm {Mon}}_{{\mathbf {E}}_\infty }^{\mathrm {gp}}\) is a grouplike \({\mathbf {E}}_\infty \)-monoid, the sequence of spaces \(X,BX,B^2X,\ldots \) forms a connective spectrum, giving a functor

$$\begin{aligned} {\mathrm {Mon}}_{{\mathbf {E}}_\infty }^{\mathrm {gp}}\rightarrow {\mathrm {Sp}}^{\ge 0}. \end{aligned}$$

In particular, the construction shows that as with classical monoids, one can form the group completion.

Corollary 12.12

The full subcategory \({\mathrm {Mon}}_{{\mathbf {E}}_\infty }^{\mathrm {gp}}\subset {\mathrm {Mon}}_{{\mathbf {E}}_\infty }\) is a reflective subcategory, cf. [33, Remark 5.2.7.9], i.e. it admits a left adjoint, called the group completion,

$$\begin{aligned} {\mathrm {Mon}}_{{\mathbf {E}}_\infty }\rightarrow {\mathrm {Mon}}_{{\mathbf {E}}_\infty }^{\mathrm {gp}}: X\mapsto X^{\mathrm {gp}}=\Omega BX\ . \end{aligned}$$

Recall that \(K_0(C)\) is the group completion of the commutative monoid of objects of C up to isomorphism. Using the above machinery, we can erase the words “up to isomorphism”, and arrive at the definition of higher algebraic K-theory:

Definition 12.13

Let C be a symmetric monoidal category. The K-theory spectrum \(K(C)\in {\mathrm {Sp}}^{\ge 0}\) is defined to be the image under \({\mathrm {Mon}}_{{\mathbf {E}}_\infty }^{\mathrm {gp}}\simeq {\mathrm {Sp}}^{\ge 0}\) of the group completion \(N(C^\simeq )^{\mathrm {gp}}\) of the \({\mathbf {E}}_\infty \)-monoid \(N(C^\simeq )\) from Construction 12.5.

Clearly, this construction is functorial in C. There is a certain situation in which the group completion is unnecessary.

Definition 12.14

A symmetric monoidal category C is called a Picard groupoid if C is a groupoid, and any object \(X\in C\) admits an inverse \(X^{-1}\in C\) such that \(X\otimes X^{-1}\cong 1\).

The groupoids \({\mathcal {P}}{\mathrm {ic}}(R)\) and \({\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R)\) are examples of Picard groupoids (explaining the name), as one can form the inverse of a line bundle.

Proposition 12.15

A symmetric monoidal groupoid C is a Picard groupoid if and only if the \({\mathbf {E}}_\infty \)-monoid N(C) is grouplike.

Proof

One has an identification of \(\pi _0 N(C) = C/\simeq \) with the monoid (under \(\otimes \)) of objects of C up to isomorphism. But, by definition, C is a Picard groupoid if and only if \(C/\simeq \) is a group. \(\square \)

In particular, for a Picard groupoid, the space underlying K(C) is just the nerve N(C), which is 1-truncated. One can show that this induces an equivalence between Picard groupoids and 1-truncated connective spectra, cf. [43, §3]. The idea is that both can be identified with grouplike \({\mathbf {E}}_\infty \)-monoids in groupoids.

Because of this equivalence, we will often confuse a Picard groupoid with its K-theory spectrum, and in particular we continue to write \({\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R)\) for the corresponding (1-truncated) connective spectrum.

Definition 12.16

Let R be a commutative ring. The K-theory spectrum of R is \(K(R) = K({\mathcal {V}}{\mathrm {ect}}(R))\).

Corollary 12.17

There is a natural functorial map

$$\begin{aligned} {\mathrm {det}}: K(R)\rightarrow {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R) \end{aligned}$$

of connective spectra.

Proof

Apply the K-theory functor to

$$\begin{aligned} {\mathrm {det}}: {\mathcal {V}}{\mathrm {ect}}(R)\rightarrow {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R). \end{aligned}$$

\(\square \)

As \({\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R)\) is 1-truncated, the map \(K(R)\rightarrow {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R)\) factors canonically over the 1-truncation \(\tau _{\le 1}K(R)\). In fact, \({\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R)\) is, at least after Zariski sheafification, precisely the 1-truncation of K(R).

Proposition 12.18

The natural map

$$\begin{aligned} {\mathrm {det}}: \tau _{\le 1} K(R)\rightarrow {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R) \end{aligned}$$

of presheaves of groupoids on the category of affine schemes becomes an isomorphism after Zariski sheafification.

Proof

We need to prove that the maps

$$\begin{aligned} K_1(R) = \pi _1 K(R)\rightarrow \pi _1 {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R) \end{aligned}$$

and

$$\begin{aligned} K_0(R) = \pi _0 K(R)\rightarrow \pi _0 {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R) \end{aligned}$$

are isomorphisms after Zariski sheafification; equivalently, for local rings R. But if R is local, \(K_1(R) = R^\times = \pi _1 {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R)\), and \(K_0(R) = {\mathbf {Z}}= \pi _0 {\mathcal {P}}{\mathrm {ic}}^{\mathbf {Z}}(R)\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bhatt, B., Scholze, P. Projectivity of the Witt vector affine Grassmannian. Invent. math. 209, 329–423 (2017). https://doi.org/10.1007/s00222-016-0710-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-016-0710-4

Navigation