Hostname: page-component-848d4c4894-ttngx Total loading time: 0 Render date: 2024-06-02T00:58:52.731Z Has data issue: false hasContentIssue false

Gabber rigidity in hermitian K-theory

Published online by Cambridge University Press:  27 April 2023

Markus Land*
Affiliation:
Mathematisches Institut, Ludwig-Maximilians-Universität München, Theresienstraße 39, 80333 München, Germany markus.land@math.lmu.de
Rights & Permissions [Opens in a new window]

Abstract

We note that Gabber's rigidity theorem for the algebraic K-theory of henselian pairs also holds true for hermitian K-theory with respect to arbitrary form parameters.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press on behalf of The Royal Society of Edinburgh

Let $R$ be a commutative ring and $\mathfrak {m} \subseteq R$ an ideal such that $(R,\mathfrak {m})$ is a henselian pair. Standard examples include henselian local rings like valuation rings of complete nonarchimedean fields as well as pairs where $R$ is $\mathfrak {m}$-adically complete or where $\mathfrak {m}$ is locally nilpotent. We write $F = R/\mathfrak {m}$ and let $n$ be a natural number which is invertible in $R$. Then Gabber's rigidity theorem [Reference Gabber5] says that the canonical map

\[ K(R)/n \longrightarrow K(F)/n \]

is an equivalence; this result was preceded by work of Suslin [Reference Suslin13] who showed this conclusion for henselian valuation rings. See also [Reference Clausen, Mathew and Morrow4] for an extension of this result, involving topological cyclic homology, to the case where $n$ need not be invertible in $R$ and a general discussion of henselian pairs. The purpose of this short note is to use the results of [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle2, Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle3] as well as [Reference Harpaz, Nikolaus and Shah6] to show that Gabber's rigidity property also holds true for hermitian K-theory, a.k.a. Grothendieck–Witt theory.

To state the main result, let $\lambda$ be a form parameter over $R$ in the sense of [Reference Schlichting12, §3], see also [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle1, definition 4.2.26]. In loc. cit. it is explained that such a form parameter $\lambda$ is equivalently described by a Poincaré structure ${\unicode{x03D9}} ^{\mathrm {g}\lambda }_R$ in the sense of [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle1] on $\mathscr {D}^{\mathrm {p}}(R)$ which sends projective $R$-modules to discrete spectra. Here, $\mathscr {D}^{\mathrm {p}}(R)$ denotes the stable $\infty$-category of perfect complexes over $R$. We will assume that the $\mathbb {Z}$-module with involution over $R$ underlying the form parameter $\lambda$ is given by $\pm R$, that is, given by the $R$-module $R$ with $C_2$-action either the identity or multiplication by $-1$, viewed as an $R\otimes R$-module via the multiplication map. There is then an induced form parameter on $F$ whose associated Poincaré structure on $\mathscr {D}^{\mathrm {p}}(F)$ we will denote by ${\unicode{x03D9}} ^{\mathrm {g}\lambda }_F$, see remark 4 below for details. The construction is made so that the extension of scalars functor canonically refines to a Poincaré functor $(\mathscr {D}^{\mathrm {p}}(R),{\unicode{x03D9}} ^{\mathrm {g}\lambda }_R) \to (\mathscr {D}^{\mathrm {p}}(F),{\unicode{x03D9}} ^{\mathrm {g}\lambda }_F)$ and therefore a map on Grothendieck–Witt theory. Standard examples of form parameters capture the notion of quadratic, even and symmetric forms (as well as their skew-quadratic, skew-even and skew-symmetric cousins) with associated Poincaré structures ${\unicode{x03D9}} ^{\pm \mathrm {gq}}, {\unicode{x03D9}} ^{\pm \mathrm {ge}}$ and ${\unicode{x03D9}} ^{\pm \mathrm {gs}}$. A further example is provided by the Burnside Poincaré structure ${\unicode{x03D9}} ^\mathrm {b}$ whose L-theory was calculated explicitly for $\mathbb {Z}$ in [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle3, example 1.3.18] and whose $0$’th Grothendieck–Witt group was studied for commutative rings with 2 invertible in the PhD thesis of Dylan Madden [Reference Madden10]. With this notation fixed, we have the following result.

Theorem 1 Let $(R,\mathfrak {m})$ be a henselian pair, $F=R/\mathfrak {m}$ and let $n$ be a natural number invertible in $R$. Then the canonical map

\[ \mathrm{GW}(R;{\unicode{x03D9}}^{\mathrm{g}\lambda}_R)/n \longrightarrow \mathrm{GW}(F;{\unicode{x03D9}}^{\mathrm{g}\lambda}_F)/n \]

is an equivalence.

Proof. The main result of [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle2] gives a diagram of horizontal fibre sequences

and by Gabber rigidity, the left vertical map becomes an equivalence after tensoring with $\mathbb {S}/n$. Therefore, the statement of the theorem is equivalent to the statement that the map

\[ \mathrm{L}(R;{\unicode{x03D9}}^{\mathrm{g}\lambda}_R)/n \longrightarrow \mathrm{L}(F;{\unicode{x03D9}}^{\mathrm{g}\lambda}_R)/n \]

is an equivalence. We then consider the diagram

where ${\unicode{x03D9}} ^{\mathrm {q}}_{\pm R}$ denotes the homotopy quadratic Poincaré structure associated with the invertible module with involution $\pm R$ which is part of the form parameter $\lambda$, and likewise for ${\unicode{x03D9}} ^{\mathrm {q}}_{\pm F}$. We now observe that the formula for relative L-theory obtained in [Reference Harpaz, Nikolaus and Shah6] shows that the top and bottom horizontal cofibres are $\mathbb {S}[\tfrac {1}{n}]$-modules.

Indeed, [Reference Harpaz, Nikolaus and Shah6] shows that the cofibre of the top horizontal arrow is a filtered colimit of objects of the form

\[ \mathrm{Eq} \left( \mathrm{map}_{R}(T\otimes_R T,R) \rightrightarrows (\Sigma^{1-\sigma}\mathrm{map}_{R}(T\otimes_R T,R))_{hC_2} \right) \]

for $T \in \mathscr {D}^{\mathrm {p}}(R)$, the bottom horizontal cofibre is described similarlyFootnote 1. Since $\mathrm {map}_{R}(T\otimes _R T,R)$ is canonically an $R$-module and $n$ is invertible in $R$, it is also an $\mathbb {S}\left [\tfrac {1}{n}\right ]$-module. Moreover, since $\mathrm {Mod}\left (\mathbb {S}\left [\tfrac {1}{n}\right ]\right ) \subseteq \mathrm {Sp}$ is a full subcategory closed under colimits and limits, both terms in the equalizer, and therefore also the equalizer itself belong to $\mathrm {Mod}\left (\mathbb {S}\left [\tfrac {1}{n}\right ]\right )$. Consequently, the horizontal maps in the above diagram become equivalences upon tensoring with $\mathbb {S}/n$. The statement of the main theorem is therefore equivalent to the statement that the left vertical map in the above commutative square is an equivalence. This is a consequence of the work of Wall's [Reference Wall14] as explained in [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle3, prop. 2.3.7 and remark 2.3.8].

Remark 2 Restricting the situation above to form parameters rather than general Poincaré structures on $\mathscr {D}^{\mathrm {p}}(R)$ was merely a cosmetic choice to obtain a result about classical Grothendieck–Witt theory: Indeed, it is again a consequence of the main theorem of [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle2] that the diagram

is a pullback diagram for any Poincaré structure ${\unicode{x03D9}}$ on $\mathscr {D}^{\mathrm {p}}(R)$ whose $\mathbb {Z}$-module with involution over $R$ is given by $\pm R$. The proof presented above therefore shows that for any ring $R$ in which $n$ is invertible, the canonical map

\[ \mathrm{GW}(R;{\unicode{x03D9}}_{{\pm} R}^\mathrm{q})/n \longrightarrow \mathrm{GW}(R;{\unicode{x03D9}})/n \]

is an equivalence so that Gabber rigidity also holds for the Poincaré structure ${\unicode{x03D9}}$.

In particular, Gabber rigidity also applies to the homotopy symmetric Poincaré structure ${\unicode{x03D9}} ^{\pm \mathrm {s}}$ as well as the Tate Poincaré structure ${\unicode{x03D9}} ^{\mathrm {t}}_R$, see [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle1, example 3.2.12].

Remark 3 Rigidity in hermitian K-theory has of course been studied in several works before, see for instance [Reference Hornbostel and Yagunov7Reference Karoubi9, Reference Yagunov15] for the case of rings with involution. The main purpose here is to show how to use the formalism of Poincaré categories and the main result of [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle2] to reduce rigidity in hermitian K-theory to rigidity in algebraic K-theory and L-theory in a way that allows to treat general form parameters.

Remark 4 In this remark, we describe how extension of scalars can be used to prolong a form parameter over $R$ along a map $R \to R'$ of rings. It is here that the assumption on the underlying module with involution is used. Indeed, we will describe a general construction on Hermitian structures, and the assumption is used to ensure that the given Poincaré structure is sent to a Poincaré structure rather than merely a Hermitian structure.

Namely, in [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle1, §3.3], we have shown that the category of Hermitian structures on $\mathscr {D}^{\mathrm {p}}(R)$ is equivalent to the category $\mathrm {Mod}_{\mathrm {N}(R)}(\mathrm {Sp}^{C_2}) = \mathrm {Mod}(\mathrm {N}(R))$, that is, the category of modules over the multiplicative normFootnote 2 $\mathrm {N}(R)$ in the category $\mathrm {Sp}^{C_2}$ of genuine $C_2$-spectra. Moreover, the category $\mathrm {Mod}(\mathrm {N}(R))$ is equipped with a canonical $t$-structure whose heart is equivalent to the category of (possibly degenerate) form parameters over $R$, see [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle1, remark 4.2.27]. Objects in $\mathrm {Mod}(\mathrm {N}(R))$ are described by triples $(M,N, \alpha )$ where

  1. $M$ is an object of $\mathrm {Mod}_{R\otimes R}(\mathrm {Sp}^{BC_2})$, where $R\otimes R$ is an algebra in spectra with $C_2$-action where the action flips the two tensor factors,

  2. $N$ is an object of $\mathrm {Mod}(R)$ and

  3. $\alpha$ is a map $N \to M^{tC_2}$ of $R$-modules,

see [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle1]; the Poincaré structures then consist of the above triples where $M$ is invertible in the sense of [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle1, def. 3.1.4]. We warn the reader that caution has to be taken in regards to how $M^{tC_2}$ is to be viewed as an $R$-module, see e.g. [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle3, p. 7] for the details. An object $(M,N, \alpha )$ is connective in the canonical $t$-structure on $\mathrm {Mod}(\mathrm {N}(R))$ if and only if $M$ and $N$ are connective.

The Poincaré structure associated with the triple $(M,N,\alpha )$ is denoted by ${\unicode{x03D9}} ^\alpha _M$. Assuming that $M$ is in the image of the canonical functor $\mathrm {Fun}(BC_2,\mathrm {Mod}(R)) \to \mathrm {Mod}_{R\otimes R}(\mathrm {Fun}(BC_2,\mathrm {Sp}))$, the triple

\[ (M',N', \alpha') = (R'\otimes_R M, R'\otimes_R N, R'\otimes_R N \to R'\otimes_R M^{tC_2} \to (R'\otimes_R M)^{tC_2}) \]

gives rise to a Poincaré structure on $\mathscr {D}^{\mathrm {p}}(R')$ for which the extension of scalar functor canonically refines to a Poincaré functor $(\mathscr {D}^{\mathrm {p}}(R),{\unicode{x03D9}} ^\alpha _M) \to (\mathscr {D}^{\mathrm {p}}(R'),{\unicode{x03D9}} ^{\alpha '}_{M'})$, see [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle1, lemma 3.4.3]. Now, if $(M,N,\alpha )$ was associated with a form parameter, then the same need not be true for the triple $(M',N',\alpha ')$: Indeed, this is the case if and only if ${\unicode{x03D9}} ^{\alpha '}_{M'}(R')$ is a discrete spectrum which in general need not be the case (but by construction it is always a connective spectrum). However, we may consider the composite

\[ M'_{hC_2} \longrightarrow {\unicode{x03D9}}^{\alpha'}_{M'}(R') \longrightarrow \tau_{{\leq} 0}{\unicode{x03D9}}^{\alpha'}_{M'}(R') \]

and denote its cofibre by $N''$. The pushout diagram of spectra

and the fact that $(M')^{hC_2}$ is coconnective shows that there is a canonical map $\alpha ''\colon N'' \to (M')^{tC_2}$. By construction, the triple $(M', N'', \alpha '')$ is an object of $\mathrm {Mod}(\mathrm {N}(R'))^{\heartsuit }$ and in fact identifies with $\tau _{\leq 0}(M',N',\alpha ')$. This object determines a Poincaré structure ${\unicode{x03D9}} ^{\mathrm {g}\lambda '}$ associated with a form parameter $\lambda '$ over $R'$ for which the extension of scalars functor refines to a Poincaré functor

\[ (\mathscr{D}^{\mathrm{p}}(R),{\unicode{x03D9}}^{\mathrm{g}\lambda}) \longrightarrow (\mathscr{D}^{\mathrm{p}}(R'),{\unicode{x03D9}}^{\mathrm{g}\lambda'}). \]

To give an example of this construction, we recall the genuine Poincaré structures ${\unicode{x03D9}} ^{\geq m}_{\pm R}$ which, for $m=0,1,2$ are the Poincaré structures ${\unicode{x03D9}} ^{\mathrm {gq}}_{\pm R}$, ${\unicode{x03D9}} ^{\mathrm {ge}}_{\pm R}$ and ${\unicode{x03D9}} ^{\mathrm {gs}}_{\pm R}$ associated with the classical (skew-) quadratic, even and symmetric form parameter over $R$, respectively, see [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle3, remark R.3 and R.5]. In this case, the extension of scalars functor associated with a ring map $R \to R'$ indeed sends ${\unicode{x03D9}} ^{\geq m}_{\pm R}$ to ${\unicode{x03D9}} ^{\geq m}_{\pm R'}$.

Remark 5 For the Poincaré structures ${\unicode{x03D9}} ^{\geq m}_{\pm R}$ one can give the following argument that the map

\[ \mathrm{GW}(R;{\unicode{x03D9}}^{{\geq} m}_{{\pm} R})/n \longrightarrow \mathrm{GW}(F;{\unicode{x03D9}}^{{\geq} m}_{{\pm} R})/n \]

is an equivalence without appealing to the general formula for relative L-theory of [Reference Harpaz, Nikolaus and Shah6]. Namely, in [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle3, prop. 3.1.14] we have shown that the map

\[ \mathrm{L}(R;{\unicode{x03D9}}^{\mathrm{q}}_{{\pm} R})\left[\tfrac{1}{2}\right] \longrightarrow \mathrm{L}(R;{\unicode{x03D9}}^{{\geq} m}_{{\pm} R})\left[\tfrac{1}{2}\right] \]

is an equivalence for all $m \in \mathbb {Z}$. Therefore, the proof of the theorem applies in the case where 2 does not divide $n$. In the case where $2$ divides $n$, we deduce that $2$ is invertible in $R$ in which case already the map ${\unicode{x03D9}} ^{\mathrm {q}}_{\pm R} \to {\unicode{x03D9}} ^{\geq m}_{\pm R}$ is an equivalence of Poincaré structures, see [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle3, remark R.4].

Remark 6 Suppose that $R$ is an associative ring which is $\mathfrak {m}$-adically complete for an ideal $\mathfrak {m} \subset R$. Then the result of Wall, see again [Reference Calmés, Dotto, Harpaz, Hebestreit, Land, Moi, Nardin, Nikolaus and Steimle3, prop. 2.3.7], says that the map $\mathrm {L}^{\pm \mathrm {q}}(R) \to \mathrm {L}^{\pm \mathrm {q}}(R/\mathfrak {m})$ is an equivalence. To the best of our knowledge, it is not known whether also the map $K(R)/n \to K(R/\mathfrak {m})/n$ is an equivalence. However, if it is, this argument shows that the same is true for Grothendieck–Witt theory and vice versa.

Acknowledgements

The author was supported by the Danish National Research Foundation through the Copenhagen Centre for Geometry and Topology (DNRF151).

Footnotes

1 In [Reference Harpaz, Nikolaus and Shah6], the authors in fact show that the relative L-theory in question, also known as normal L-theory, is given by the $C_2$-geometric fixed points of the real topological cyclic homology of $(\mathscr {D}^{\mathrm {p}}(R),{\unicode{x03D9}} ^{\mathrm {g}\lambda }_R)$. The equalizer formula is then reminiscent of the Nikolaus–Scholze formula for TC [Reference Nikolaus and Scholze11].

2 Also known as the Hill–Hopkins–Ravenel norm.

References

Calmés, B., Dotto, E., Harpaz, Y., Hebestreit, F., Land, M., Moi, K., Nardin, D., Nikolaus, T. and Steimle, W.. Hermitian K-theory for stable $\infty$-categories I: foundations. arXiv:2009.07223, 2020.Google Scholar
Calmés, B., Dotto, E., Harpaz, Y., Hebestreit, F., Land, M., Moi, K., Nardin, D., Nikolaus, T. and Steimle, W.. Hermitian K-theory for stable $\infty$-categories II: cobordism categories and additivity. arXiv:2009.07224, 2020.Google Scholar
Calmés, B., Dotto, E., Harpaz, Y., Hebestreit, F., Land, M., Moi, K., Nardin, D., Nikolaus, T. and Steimle, W.. Hermitian K-theory for stable $\infty$-categories III: Grothendieck–Witt groups of rings. arXiv:2009.07225, 2020.Google Scholar
Clausen, D., Mathew, A. and Morrow, M.. $K$-theory and topological cyclic homology of henselian pairs. J. Am. Math. Soc. 34 (2021), 411473.CrossRefGoogle Scholar
Gabber, O.. $K$-theory of henselian local rings and henselian pairs. Algebraic $K$-theory, commutative algebra, and algebraic geometry (Santa Margherita Ligure, 1989). Contemp. Math., vol. 126, pp. 59–70 (Providence, RI: Amer. Math. Soc., 1992).CrossRefGoogle Scholar
Harpaz, Y., Nikolaus, T. and Shah, J.. Real topological cyclic homology and normal L-theory. In preparation, 2022.Google Scholar
Hornbostel, J. and Yagunov, S.. Rigidity for henselian local rings and $\Bbb A^1$-representable theories. Math. Z. 255 (2007), 437449.CrossRefGoogle Scholar
Jardin, J. F.. A rigidity theorem for $L$-theory. Preprint (University of Chicago, 1983). https://www.uwo.ca/math/faculty/jardine/preprints/preprint-l-rigid.pdf.Google Scholar
Karoubi, M.. Relations between algebraic $K$-theory and Hermitian $K$-theory. In Proceedings of the Luminy Conference on Algebraic $K$-theory (Luminy, 1983), vol. 34, pp. 259–263 (1984).CrossRefGoogle Scholar
Madden, D.. Burnside form rings and the K-theory of forms. Ph.D. thesis (University of Warwick, 2021). http://homepages.warwick.ac.uk/∼masiap/Dylan_Madden_thesis.pdf.Google Scholar
Nikolaus, T. and Scholze, P.. On topological cyclic homology. Acta Math. 221 (2018), 203409.CrossRefGoogle Scholar
Schlichting, M.. Higher $K$-theory of forms I. From rings to exact categories. J. Inst. Math. Jussieu 20 (2021), 12051273.CrossRefGoogle Scholar
Suslin, A. A.. On the $K$-theory of local fields. Proceedings of the Luminy Conference on Algebraic $K$-theory (Luminy, 1983), vol. 34, pp. 301–318 (1984).CrossRefGoogle Scholar
Wall, C. T. C.. On the classification of hermitian forms. III. Complete semilocal rings. Invent. Math. 19 (1973), 5971.CrossRefGoogle Scholar
Yagunov, S.. Suslin's work on the K-theory of local fields. Revisited, arXiv:2201.00728, 2022.Google Scholar