Milnor K-theory
In mathematics, Milnor K-theory[1] is an algebraic invariant (denoted {\displaystyle K_{*}(F)} for a field {\displaystyle F}) defined by John Milnor (1970) as an attempt to study higher algebraic K-theory in the special case of fields. It was hoped this would help illuminate the structure for algebraic K-theory and give some insight about its relationships with other parts of mathematics, such as Galois cohomology and the Grothendieck–Witt ring of quadratic forms. Before Milnor K-theory was defined, there existed ad-hoc definitions for {\displaystyle K_{1}} and {\displaystyle K_{2}}. Fortunately, it can be shown Milnor K-theory is a part of algebraic K-theory, which in general is the easiest part to compute.[2]
Definition
[edit ]Motivation
[edit ]After the definition of the Grothendieck group {\displaystyle K(R)} of a commutative ring, it was expected there should be an infinite set of invariants {\displaystyle K_{i}(R)} called higher K-theory groups, from the fact there exists a short exact sequence
- {\displaystyle K(R,I)\to K(R)\to K(R/I)\to 0}
which should have a continuation by a long exact sequence. Note the group on the left is relative K-theory. This led to much study and as a first guess for what this theory would look like, Milnor gave a definition for fields. His definition is based upon two calculations of what higher K-theory "should" look like in degrees {\displaystyle 1} and {\displaystyle 2}. Then, if in a later generalization of algebraic K-theory was given, if the generators of {\displaystyle K_{*}(R)} lived in degree {\displaystyle 1} and the relations in degree {\displaystyle 2}, then the constructions in degrees {\displaystyle 1} and {\displaystyle 2} would give the structure for the rest of the K-theory ring. Under this assumption, Milnor gave his "ad-hoc" definition. It turns out algebraic K-theory {\displaystyle K_{*}(R)} in general has a more complex structure, but for fields the Milnor K-theory groups are contained in the general algebraic K-theory groups after tensoring with {\displaystyle \mathbb {Q} }, i.e. {\displaystyle K_{n}^{M}(F)\otimes \mathbb {Q} \subseteq K_{n}(F)\otimes \mathbb {Q} }.[3] It turns out the natural map {\displaystyle \lambda :K_{4}^{M}(F)\to K_{4}(F)} fails to be injective for a global field {\displaystyle F}[3] pg 96.
Definition
[edit ]Note for fields the Grothendieck group can be readily computed as {\displaystyle K_{0}(F)=\mathbb {Z} } since the only finitely generated modules are finite-dimensional vector spaces. Also, Milnor's definition of higher K-groups depends upon the canonical isomorphism
- {\displaystyle l\colon K_{1}(F)\to F^{*}}
(the group of units of {\displaystyle F}) and observing the calculation of K2 of a field by Hideya Matsumoto, which gave the simple presentation
- {\displaystyle K_{2}(F)={\frac {F^{*}\otimes F^{*}}{\{l(a)\otimes l(1-a):a\neq 0,1\}}}}
for a two-sided ideal generated by elements {\displaystyle l(a)\otimes l(a-1)}, called Steinberg relations. Milnor took the hypothesis that these were the only relations, hence he gave the following "ad-hoc" definition of Milnor K-theory as
- {\displaystyle K_{n}^{M}(F)={\frac {K_{1}(F)\otimes \cdots \otimes K_{1}(F)}{\{l(a_{1})\otimes \cdots \otimes l(a_{n}):a_{i}+a_{i+1}=1\}}}.}
The direct sum of these groups is isomorphic to a tensor algebra over the integers of the multiplicative group {\displaystyle K_{1}(F)\cong F^{*}} modded out by the two-sided ideal generated by:
- {\displaystyle \left\{l(a)\otimes l(1-a):0,1\neq a\in F\right\}}
so
- {\displaystyle \bigoplus _{n=0}^{\infty }K_{n}^{M}(F)\cong {\frac {T^{*}(K_{1}^{M}(F))}{\{l(a)\otimes l(1-a):a\neq 0,1\}}}}
showing his definition is a direct extension of the Steinberg relations.
Properties
[edit ]Ring structure
[edit ]The graded module {\displaystyle K_{*}^{M}(F)} is a graded-commutative ring[1] pg 1-3.[4] If we write
- {\displaystyle (l(a_{1})\otimes \cdots \otimes l(a_{n}))\cdot (l(b_{1})\otimes \cdots \otimes l(b_{m}))}
as
- {\displaystyle l(a_{1})\otimes \cdots \otimes l(a_{n})\otimes l(b_{1})\otimes \cdots \otimes l(b_{m})}
then for {\displaystyle \xi \in K_{i}^{M}(F)} and {\displaystyle \eta \in K_{j}^{M}(F)} we have
- {\displaystyle \xi \cdot \eta =(-1)^{i\cdot j}\eta \cdot \xi .}
From the proof of this property, there are some additional properties which fall out, like {\displaystyle l(a)^{2}=l(a)l(-1)} for {\displaystyle l(a)\in K_{1}(F)} since {\displaystyle l(a)l(-a)=0}. Also, if {\displaystyle a_{1}+\cdots +a_{n}} of non-zero fields elements equals {\displaystyle 0,1}, then {\displaystyle l(a_{1})\cdots l(a_{n})=0} There's a direct arithmetic application: {\displaystyle -1\in F} is a sum of squares if and only if every positive dimensional {\displaystyle K_{n}^{M}(F)} is nilpotent, which is a powerful statement about the structure of Milnor K-groups. In particular, for the fields {\displaystyle \mathbb {Q} (i)}, {\displaystyle \mathbb {Q} _{p}(i)} with {\displaystyle {\sqrt {-1}}\not \in \mathbb {Q} _{p}}, all of its Milnor K-groups are nilpotent. In the converse case, the field {\displaystyle F} can be embedded into a real closed field, which gives a total ordering on the field.
Relation to Higher Chow groups and Quillen's higher K-theory
[edit ]One of the core properties relating Milnor K-theory to higher algebraic K-theory is the fact there exists natural isomorphisms {\displaystyle K_{n}^{M}(F)\to {\text{CH}}^{n}(F,n)} to Bloch's Higher chow groups which induces a morphism of graded rings {\displaystyle K_{*}^{M}(F)\to {\text{CH}}^{*}(F,*)} This can be verified using an explicit morphism[2] pg 181 {\displaystyle \phi :F^{*}\to {\text{CH}}^{1}(F,1)} where {\displaystyle \phi (a)\phi (1-a)=0~{\text{in}}~{\text{CH}}^{2}(F,2)~{\text{for}}~a,1-a\in F^{*}} This map is given by {\displaystyle {\begin{aligned}\{1\}&\mapsto 0\in {\text{CH}}^{1}(F,1)\\\{a\}&\mapsto [a]\in {\text{CH}}^{1}(F,1)\end{aligned}}} for {\displaystyle [a]} the class of the point {\displaystyle [a:1]\in \mathbb {P} _{F}^{1}-\{0,1,\infty \}} with {\displaystyle a\in F^{*}-\{1\}}. The main property to check is that {\displaystyle [a]+[1/a]=0} for {\displaystyle a\in F^{*}-\{1\}} and {\displaystyle [a]+[b]=[ab]}. Note this is distinct from {\displaystyle [a]\cdot [b]} since this is an element in {\displaystyle {\text{CH}}^{2}(F,2)}. Also, the second property implies the first for {\displaystyle b=1/a}. This check can be done using a rational curve defining a cycle in {\displaystyle C^{1}(F,2)} whose image under the boundary map {\displaystyle \partial } is the sum {\displaystyle [a]+[b]-[ab]}for {\displaystyle ab\neq 1}, showing they differ by a boundary. Similarly, if {\displaystyle ab=1} the boundary map sends this cycle to {\displaystyle [a]-[1/a]}, showing they differ by a boundary. The second main property to show is the Steinberg relations. With these, and the fact the higher Chow groups have a ring structure {\displaystyle {\text{CH}}^{p}(F,q)\otimes {\text{CH}}^{r}(F,s)\to {\text{CH}}^{p+r}(F,q+s)} we get an explicit map {\displaystyle K_{*}^{M}(F)\to {\text{CH}}^{*}(F,*)} Showing the map in the reverse direction is an isomorphism is more work, but we get the isomorphisms {\displaystyle K_{n}^{M}(F)\to {\text{CH}}^{n}(F,n)} We can then relate the higher Chow groups to higher algebraic K-theory using the fact there are isomorphisms {\displaystyle K_{n}(X)\otimes \mathbb {Q} \cong \bigoplus _{p}{\text{CH}}^{p}(X,n)\otimes \mathbb {Q} } giving the relation to Quillen's higher algebraic K-theory. Note that the maps
- {\displaystyle K_{n}^{M}(F)\to K_{n}(F)}
from the Milnor K-groups of a field to the Quillen K-groups, which is an isomorphism for {\displaystyle n\leq 2} but not for larger n, in general. For nonzero elements {\displaystyle a_{1},\ldots ,a_{n}} in F, the symbol {\displaystyle \{a_{1},\ldots ,a_{n}\}} in {\displaystyle K_{n}^{M}(F)} means the image of {\displaystyle a_{1}\otimes \cdots \otimes a_{n}} in the tensor algebra. Every element of Milnor K-theory can be written as a finite sum of symbols. The fact that {\displaystyle \{a,1-a\}=0} in {\displaystyle K_{2}^{M}(F)} for {\displaystyle a\in F\setminus \{0,1\}} is sometimes called the Steinberg relation.
Representation in motivic cohomology
[edit ]In motivic cohomology, specifically motivic homotopy theory, there is a sheaf {\displaystyle K_{n,A}} representing a generalization of Milnor K-theory with coefficients in an abelian group {\displaystyle A}. If we denote {\displaystyle A_{tr}(X)=\mathbb {Z} _{tr}(X)\otimes A} then we define the sheaf {\displaystyle K_{n,A}} as the sheafification of the following pre-sheaf[5] pg 4 {\displaystyle K_{n,A}^{pre}:U\mapsto A_{tr}(\mathbb {A} ^{n})(U)/A_{tr}(\mathbb {A} ^{n}-\{0\})(U)} Note that sections of this pre-sheaf are equivalent classes of cycles on {\displaystyle U\times \mathbb {A} ^{n}} with coefficients in {\displaystyle A} which are equidimensional and finite over {\displaystyle U} (which follows straight from the definition of {\displaystyle \mathbb {Z} _{tr}(X)}). It can be shown there is an {\displaystyle \mathbb {A} ^{1}}-weak equivalence with the motivic Eilenberg-Maclane sheaves {\displaystyle K(A,2n,n)} (depending on the grading convention).
Examples
[edit ]Finite fields
[edit ]For a finite field {\displaystyle F=\mathbb {F} _{q}}, {\displaystyle K_{1}^{M}(F)} is a cyclic group of order {\displaystyle q-1} (since is it isomorphic to {\displaystyle \mathbb {F} _{q}^{*}}), so graded commutativity gives {\displaystyle l(a)\cdot l(b)=-l(b)\cdot l(a)} hence {\displaystyle l(a)^{2}=-l(a)^{2}} Because {\displaystyle K_{2}^{M}(F)} is a finite group, this implies it must have order {\displaystyle \leq 2}. Looking further, {\displaystyle 1} can always be expressed as a sum of quadratic non-residues, i.e. elements {\displaystyle a,b\in F} such that {\displaystyle [a],[b]\in F/F^{\times 2}} are not equal to {\displaystyle 0}, hence {\displaystyle a+b=1} showing {\displaystyle K_{2}^{M}(F)=0}. Because the Steinberg relations generate all relations in the Milnor K-theory ring, we have {\displaystyle K_{n}^{M}(F)=0} for {\displaystyle n>2}.
Real numbers
[edit ]For the field of real numbers {\displaystyle \mathbb {R} } the Milnor K-theory groups can be readily computed. In degree {\displaystyle n} the group is generated by {\displaystyle K_{n}^{M}(\mathbb {R} )=\{(-1)^{n},l(a_{1})\cdots l(a_{n}):a_{1},\ldots ,a_{n}>0\}} where {\displaystyle (-1)^{n}} gives a group of order {\displaystyle 2} and the subgroup generated by the {\displaystyle l(a_{1})\cdots l(a_{n})} is divisible. The subgroup generated by {\displaystyle (-1)^{n}} is not divisible because otherwise it could be expressed as a sum of squares. The Milnor K-theory ring is important in the study of motivic homotopy theory because it gives generators for part of the motivic Steenrod algebra.[6] The others are lifts from the classical Steenrod operations to motivic cohomology.
Other calculations
[edit ]{\displaystyle K_{2}^{M}(\mathbb {C} )} is an uncountable uniquely divisible group.[7] Also, {\displaystyle K_{2}^{M}(\mathbb {R} )} is the direct sum of a cyclic group of order 2 and an uncountable uniquely divisible group; {\displaystyle K_{2}^{M}(\mathbb {Q} _{p})} is the direct sum of the multiplicative group of {\displaystyle \mathbb {F} _{p}} and an uncountable uniquely divisible group; {\displaystyle K_{2}^{M}(\mathbb {Q} )} is the direct sum of the cyclic group of order 2 and cyclic groups of order {\displaystyle p-1} for all odd prime {\displaystyle p}. For {\displaystyle n\geq 3}, {\displaystyle K_{n}^{M}(\mathbb {Q} )\cong \mathbb {Z} /2}. The full proof is in the appendix of Milnor's original paper.[1] Some of the computation can be seen by looking at a map on {\displaystyle K_{2}^{M}(F)} induced from the inclusion of a global field {\displaystyle F} to its completions {\displaystyle F_{v}}, so there is a morphism{\displaystyle K_{2}^{M}(F)\to \bigoplus _{v}K_{2}^{M}(F_{v})/({\text{max. divis. subgr.}})} whose kernel finitely generated. In addition, the cokernel is isomorphic to the roots of unity in {\displaystyle F}.
In addition, for a general local field {\displaystyle F} (such as a finite extension {\displaystyle K/\mathbb {Q} _{p}}), the Milnor K-groups {\displaystyle K_{n}^{M}(F)} are divisible.
K*M(F(t))
[edit ]There is a general structure theorem computing {\displaystyle K_{n}^{M}(F(t))} for a field {\displaystyle F} in relation to the Milnor K-theory of {\displaystyle F} and extensions {\displaystyle F[t]/(\pi )} for non-zero primes ideals {\displaystyle (\pi )\in {\text{Spec}}(F[t])}. This is given by an exact sequence {\displaystyle 0\to K_{n}^{M}(F)\to K_{n}^{M}(F(t))\xrightarrow {\partial _{\pi }} \bigoplus _{(\pi )\in {\text{Spec}}(F[t])}K_{n-1}F[t]/(\pi )\to 0} where {\displaystyle \partial _{\pi }:K_{n}^{M}(F(t))\to K_{n-1}F[t]/(\pi )} is a morphism constructed from a reduction of {\displaystyle F} to {\displaystyle {\overline {F}}_{v}} for a discrete valuation {\displaystyle v}. This follows from the theorem there exists only one homomorphism {\displaystyle \partial :K_{n}^{M}(F)\to K_{n-1}^{M}({\overline {F}})} which for the group of units {\displaystyle U\subset F} which are elements have valuation {\displaystyle 0}, having a natural morphism {\displaystyle U\to {\overline {F}}_{v}^{*}} where {\displaystyle u\mapsto {\overline {u}}} we have {\displaystyle \partial (l(\pi )l(u_{2})\cdots l(u_{n}))=l({\overline {u}}_{2})\cdots l({\overline {u}}_{n})} where {\displaystyle \pi } a prime element, meaning {\displaystyle {\text{Ord}}_{v}(\pi )=1}, and {\displaystyle \partial (l(u_{1})\cdots l(u_{n}))=0} Since every non-zero prime ideal {\displaystyle (\pi )\in {\text{Spec}}(F[t])} gives a valuation {\displaystyle v_{\pi }:F(t)\to F[t]/(\pi )}, we get the map {\displaystyle \partial _{\pi }} on the Milnor K-groups.
Applications
[edit ]Milnor K-theory plays a fundamental role in higher class field theory, replacing {\displaystyle K_{1}^{M}(F)=F^{\times }\!} in the one-dimensional class field theory.
Milnor K-theory fits into the broader context of motivic cohomology, via the isomorphism
- {\displaystyle K_{n}^{M}(F)\cong H^{n}(F,\mathbb {Z} (n))}
of the Milnor K-theory of a field with a certain motivic cohomology group.[8] In this sense, the apparently ad hoc definition of Milnor K-theory becomes a theorem: certain motivic cohomology groups of a field can be explicitly computed by generators and relations.
A much deeper result, the Bloch-Kato conjecture (also called the norm residue isomorphism theorem), relates Milnor K-theory to Galois cohomology or étale cohomology:
- {\displaystyle K_{n}^{M}(F)/r\cong H_{\mathrm {et} }^{n}(F,\mathbb {Z} /r(n)),}
for any positive integer r invertible in the field F. This conjecture was proved by Vladimir Voevodsky, with contributions by Markus Rost and others.[9] This includes the theorem of Alexander Merkurjev and Andrei Suslin as well as the Milnor conjecture as special cases (the cases when {\displaystyle n=2} and {\displaystyle r=2}, respectively).
Finally, there is a relation between Milnor K-theory and quadratic forms. For a field F of characteristic not 2, define the fundamental ideal I in the Witt ring of quadratic forms over F to be the kernel of the homomorphism {\displaystyle W(F)\to \mathbb {Z} /2} given by the dimension of a quadratic form, modulo 2. Milnor defined a homomorphism:
- {\displaystyle {\begin{cases}K_{n}^{M}(F)/2\to I^{n}/I^{n+1}\\\{a_{1},\ldots ,a_{n}\}\mapsto \langle \langle a_{1},\ldots ,a_{n}\rangle \rangle =\langle 1,-a_{1}\rangle \otimes \cdots \otimes \langle 1,-a_{n}\rangle \end{cases}}}
where {\displaystyle \langle \langle a_{1},a_{2},\ldots ,a_{n}\rangle \rangle } denotes the class of the n-fold Pfister form.[10]
Dmitri Orlov, Alexander Vishik, and Voevodsky proved another statement called the Milnor conjecture, namely that this homomorphism {\displaystyle K_{n}^{M}(F)/2\to I^{n}/I^{n+1}} is an isomorphism.[11]
See also
[edit ]References
[edit ]- ^ a b c Milnor, John (1970年12月01日). "Algebraic K -theory and quadratic forms". Inventiones Mathematicae . 9 (4): 318–344. Bibcode:1970InMat...9..318M. doi:10.1007/BF01425486. ISSN 1432-1297. S2CID 13549621.
- ^ a b Totaro, Burt. "Milnor K-Theory is the Simplest Part of Algebraic K-Theory" (PDF). Archived (PDF) from the original on 2 Dec 2020.
- ^ a b Shapiro, Jack M. (1981年01月01日). "Relations between the milnor and quillen K-theory of fields". Journal of Pure and Applied Algebra. 20 (1): 93–102. doi:10.1016/0022-4049(81)90051-7 . ISSN 0022-4049.
- ^ Gille & Szamuely (2006), p. 184.
- ^ Voevodsky, Vladimir (2001年07月15日). "Reduced power operations in motivic cohomology". arXiv:math/0107109 .
- ^ Bachmann, Tom (May 2018). "Motivic and Real Etale Stable Homotopy Theory". Compositio Mathematica. 154 (5): 883–917. arXiv:1608.08855 . doi:10.1112/S0010437X17007710. ISSN 0010-437X. S2CID 119305101.
- ^ An abelian group is uniquely divisible if it is a vector space over the rational numbers.
- ^ Mazza, Voevodsky, Weibel (2005), Theorem 5.1.
- ^ Voevodsky (2011).
- ^ Elman, Karpenko, Merkurjev (2008), sections 5 and 9.B.
- ^ Orlov, Vishik, Voevodsky (2007).
- Elman, Richard; Karpenko, Nikita; Merkurjev, Alexander (2008), Algebraic and geometric theory of quadratic forms, American Mathematical Society, ISBN 978-0-8218-4329-1, MR 2427530
- Gille, Philippe; Szamuely, Tamás (2006). Central simple algebras and Galois cohomology. Cambridge Studies in Advanced Mathematics. Vol. 101. Cambridge: Cambridge University Press. ISBN 0-521-86103-9. MR 2266528. Zbl 1137.12001.
- Mazza, Carlo; Voevodsky, Vladimir; Weibel, Charles (2006), Lectures in Motivic Cohomology, Clay Mathematical Monographs, vol. 2, American Mathematical Society, ISBN 978-0-8218-3847-1, MR 2242284
- Milnor, John Willard (1970), "Algebraic K-theory and quadratic forms", Inventiones Mathematicae , 9 (4), With an appendix by John Tate: 318–344, Bibcode:1970InMat...9..318M, doi:10.1007/BF01425486, ISSN 0020-9910, MR 0260844, S2CID 13549621, Zbl 0199.55501
- Orlov, Dmitri; Vishik, Alexander; Voevodsky, Vladimir (2007), "An exact sequence for {\displaystyle K_{*}^{M}/2} with applications to quadratic forms", Annals of Mathematics , 165: 1–13, arXiv:math/0101023 , doi:10.4007/annals.2007.165.1, MR 2276765, S2CID 9504456
- Voevodsky, Vladimir (2011), "On motivic cohomology with {\displaystyle \mathbb {Z} /\ell }-coefficients", Annals of Mathematics , 174 (1): 401–438, arXiv:0805.4430 , doi:10.4007/annals.2011.174.1.11, MR 2811603, S2CID 15583705