Milnor K-theory

In mathematics, Milnor K-theory[1] is an algebraic invariant (denoted K ( F ) {\displaystyle K_{*}(F)} for a field F {\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 K 1 {\displaystyle K_{1}} and K 2 {\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

Motivation

After the definition of the Grothendieck group K ( R ) {\displaystyle K(R)} of a commutative ring, it was expected there should be an infinite set of invariants K i ( R ) {\displaystyle K_{i}(R)} called higher K-theory groups, from the fact there exists a short exact sequence

K ( R , I ) K ( R ) K ( R / I ) 0 {\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 1 {\displaystyle 1} and 2 {\displaystyle 2} . Then, if in a later generalization of algebraic K-theory was given, if the generators of K ( R ) {\displaystyle K_{*}(R)} lived in degree 1 {\displaystyle 1} and the relations in degree 2 {\displaystyle 2} , then the constructions in degrees 1 {\displaystyle 1} and 2 {\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 K ( R ) {\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 Q {\displaystyle \mathbb {Q} } , i.e. K n M ( F ) Q K n ( F ) Q {\displaystyle K_{n}^{M}(F)\otimes \mathbb {Q} \subseteq K_{n}(F)\otimes \mathbb {Q} } .[3] It turns out the natural map λ : K 4 M ( F ) K 4 ( F ) {\displaystyle \lambda :K_{4}^{M}(F)\to K_{4}(F)} fails to be injective for a global field F {\displaystyle F} [3]pg 96.

Definition

Note for fields the Grothendieck group can be readily computed as K 0 ( F ) = Z {\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

l : K 1 ( F ) F {\displaystyle l\colon K_{1}(F)\to F^{*}}

(the group of units of F {\displaystyle F} ) and observing the calculation of K2 of a field by Hideya Matsumoto, which gave the simple presentation

K 2 ( F ) = F F { l ( a ) l ( 1 a ) : a 0 , 1 } {\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 l ( a ) l ( a 1 ) {\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

K n M ( F ) = K 1 ( F ) K 1 ( F ) { l ( a 1 ) l ( a n ) : a i + a i + 1 = 1 } . {\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 K 1 ( F ) F {\displaystyle K_{1}(F)\cong F^{*}} modded out by the two-sided ideal generated by:

{ l ( a ) l ( 1 a ) : 0 , 1 a F } {\displaystyle \left\{l(a)\otimes l(1-a):0,1\neq a\in F\right\}}

so

n = 0 K n M ( F ) T ( K 1 M ( F ) ) { l ( a ) l ( 1 a ) : a 0 , 1 } {\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

Ring structure

The graded module K M ( F ) {\displaystyle K_{*}^{M}(F)} is a graded-commutative ring[1]pg 1-3.[4] If we write

( l ( a 1 ) l ( a n ) ) ( l ( b 1 ) l ( b m ) ) {\displaystyle (l(a_{1})\otimes \cdots \otimes l(a_{n}))\cdot (l(b_{1})\otimes \cdots \otimes l(b_{m}))}

as

l ( a 1 ) l ( a n ) l ( b 1 ) l ( b m ) {\displaystyle l(a_{1})\otimes \cdots \otimes l(a_{n})\otimes l(b_{1})\otimes \cdots \otimes l(b_{m})}

then for ξ K i M ( F ) {\displaystyle \xi \in K_{i}^{M}(F)} and η K j M ( F ) {\displaystyle \eta \in K_{j}^{M}(F)} we have

ξ η = ( 1 ) i j η ξ . {\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

l ( a ) 2 = l ( a ) l ( 1 ) {\displaystyle l(a)^{2}=l(a)l(-1)}
for l ( a ) K 1 ( F ) {\displaystyle l(a)\in K_{1}(F)} since l ( a ) l ( a ) = 0 {\displaystyle l(a)l(-a)=0} . Also, if a 1 + + a n {\displaystyle a_{1}+\cdots +a_{n}} of non-zero fields elements equals 0 , 1 {\displaystyle 0,1} , then
l ( a 1 ) l ( a n ) = 0 {\displaystyle l(a_{1})\cdots l(a_{n})=0}
There's a direct arithmetic application: 1 F {\displaystyle -1\in F} is a sum of squares if and only if every positive dimensional K n M ( F ) {\displaystyle K_{n}^{M}(F)} is nilpotent, which is a powerful statement about the structure of Milnor K-groups. In particular, for the fields Q ( i ) {\displaystyle \mathbb {Q} (i)} , Q p ( i ) {\displaystyle \mathbb {Q} _{p}(i)} with 1 Q p {\displaystyle {\sqrt {-1}}\not \in \mathbb {Q} _{p}} , all of its Milnor K-groups are nilpotent. In the converse case, the field F {\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

One of the core properties relating Milnor K-theory to higher algebraic K-theory is the fact there exists natural isomorphisms

K n M ( F ) CH n ( F , n ) {\displaystyle K_{n}^{M}(F)\to {\text{CH}}^{n}(F,n)}
to Bloch's Higher chow groups which induces a morphism of graded rings
K M ( F ) CH ( F , ) {\displaystyle K_{*}^{M}(F)\to {\text{CH}}^{*}(F,*)}
This can be verified using an explicit morphism[2]pg 181
ϕ : F CH 1 ( F , 1 ) {\displaystyle \phi :F^{*}\to {\text{CH}}^{1}(F,1)}
where
ϕ ( a ) ϕ ( 1 a ) = 0   in   CH 2 ( F , 2 )   for   a , 1 a F {\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
{ 1 } 0 CH 1 ( F , 1 ) { a } [ a ] CH 1 ( F , 1 ) {\displaystyle {\begin{aligned}\{1\}&\mapsto 0\in {\text{CH}}^{1}(F,1)\\\{a\}&\mapsto [a]\in {\text{CH}}^{1}(F,1)\end{aligned}}}
for [ a ] {\displaystyle [a]} the class of the point [ a : 1 ] P F 1 { 0 , 1 , } {\displaystyle [a:1]\in \mathbb {P} _{F}^{1}-\{0,1,\infty \}} with a F { 1 } {\displaystyle a\in F^{*}-\{1\}} . The main property to check is that [ a ] + [ 1 / a ] = 0 {\displaystyle [a]+[1/a]=0} for a F { 1 } {\displaystyle a\in F^{*}-\{1\}} and [ a ] + [ b ] = [ a b ] {\displaystyle [a]+[b]=[ab]} . Note this is distinct from [ a ] [ b ] {\displaystyle [a]\cdot [b]} since this is an element in CH 2 ( F , 2 ) {\displaystyle {\text{CH}}^{2}(F,2)} . Also, the second property implies the first for b = 1 / a {\displaystyle b=1/a} . This check can be done using a rational curve defining a cycle in C 1 ( F , 2 ) {\displaystyle C^{1}(F,2)} whose image under the boundary map {\displaystyle \partial } is the sum [ a ] + [ b ] [ a b ] {\displaystyle [a]+[b]-[ab]} for a b 1 {\displaystyle ab\neq 1} , showing they differ by a boundary. Similarly, if a b = 1 {\displaystyle ab=1} the boundary map sends this cycle to [ a ] [ 1 / a ] {\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
CH p ( F , q ) CH r ( F , s ) CH p + r ( F , q + s ) {\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
K M ( F ) CH ( F , ) {\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
K n M ( F ) CH n ( F , n ) {\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
K n ( X ) Q p CH p ( X , n ) Q {\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

K n M ( F ) K n ( F ) {\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 n 2 {\displaystyle n\leq 2} but not for larger n, in general. For nonzero elements a 1 , , a n {\displaystyle a_{1},\ldots ,a_{n}} in F, the symbol { a 1 , , a n } {\displaystyle \{a_{1},\ldots ,a_{n}\}} in K n M ( F ) {\displaystyle K_{n}^{M}(F)} means the image of a 1 a n {\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 { a , 1 a } = 0 {\displaystyle \{a,1-a\}=0} in K 2 M ( F ) {\displaystyle K_{2}^{M}(F)} for a F { 0 , 1 } {\displaystyle a\in F\setminus \{0,1\}} is sometimes called the Steinberg relation.

Representation in motivic cohomology

In motivic cohomology, specifically motivic homotopy theory, there is a sheaf K n , A {\displaystyle K_{n,A}} representing a generalization of Milnor K-theory with coefficients in an abelian group A {\displaystyle A} . If we denote A t r ( X ) = Z t r ( X ) A {\displaystyle A_{tr}(X)=\mathbb {Z} _{tr}(X)\otimes A} then we define the sheaf K n , A {\displaystyle K_{n,A}} as the sheafification of the following pre-sheaf[5]pg 4

K n , A p r e : U A t r ( A n ) ( U ) / A t r ( A n { 0 } ) ( U ) {\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 U × A n {\displaystyle U\times \mathbb {A} ^{n}} with coefficients in A {\displaystyle A} which are equidimensional and finite over U {\displaystyle U} (which follows straight from the definition of Z t r ( X ) {\displaystyle \mathbb {Z} _{tr}(X)} ). It can be shown there is an A 1 {\displaystyle \mathbb {A} ^{1}} -weak equivalence with the motivic Eilenberg-Maclane sheaves K ( A , 2 n , n ) {\displaystyle K(A,2n,n)} (depending on the grading convention).

Examples

Finite fields

For a finite field F = F q {\displaystyle F=\mathbb {F} _{q}} , K 1 M ( F ) {\displaystyle K_{1}^{M}(F)} is a cyclic group of order q 1 {\displaystyle q-1} (since is it isomorphic to F q {\displaystyle \mathbb {F} _{q}^{*}} ), so graded commutativity gives

l ( a ) l ( b ) = l ( b ) l ( a ) {\displaystyle l(a)\cdot l(b)=-l(b)\cdot l(a)}
hence
l ( a ) 2 = l ( a ) 2 {\displaystyle l(a)^{2}=-l(a)^{2}}
Because K 2 M ( F ) {\displaystyle K_{2}^{M}(F)} is a finite group, this implies it must have order 2 {\displaystyle \leq 2} . Looking further, 1 {\displaystyle 1} can always be expressed as a sum of quadratic non-residues, i.e. elements a , b F {\displaystyle a,b\in F} such that [ a ] , [ b ] F / F × 2 {\displaystyle [a],[b]\in F/F^{\times 2}} are not equal to 0 {\displaystyle 0} , hence a + b = 1 {\displaystyle a+b=1} showing K 2 M ( F ) = 0 {\displaystyle K_{2}^{M}(F)=0} . Because the Steinberg relations generate all relations in the Milnor K-theory ring, we have K n M ( F ) = 0 {\displaystyle K_{n}^{M}(F)=0} for n > 2 {\displaystyle n>2} .

Real numbers

For the field of real numbers R {\displaystyle \mathbb {R} } the Milnor K-theory groups can be readily computed. In degree n {\displaystyle n} the group is generated by

K n M ( R ) = { ( 1 ) n , l ( a 1 ) l ( a n ) : a 1 , , a n > 0 } {\displaystyle K_{n}^{M}(\mathbb {R} )=\{(-1)^{n},l(a_{1})\cdots l(a_{n}):a_{1},\ldots ,a_{n}>0\}}
where ( 1 ) n {\displaystyle (-1)^{n}} gives a group of order 2 {\displaystyle 2} and the subgroup generated by the l ( a 1 ) l ( a n ) {\displaystyle l(a_{1})\cdots l(a_{n})} is divisible. The subgroup generated by ( 1 ) n {\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

K 2 M ( C ) {\displaystyle K_{2}^{M}(\mathbb {C} )} is an uncountable uniquely divisible group.[7] Also, K 2 M ( R ) {\displaystyle K_{2}^{M}(\mathbb {R} )} is the direct sum of a cyclic group of order 2 and an uncountable uniquely divisible group; K 2 M ( Q p ) {\displaystyle K_{2}^{M}(\mathbb {Q} _{p})} is the direct sum of the multiplicative group of F p {\displaystyle \mathbb {F} _{p}} and an uncountable uniquely divisible group; K 2 M ( Q ) {\displaystyle K_{2}^{M}(\mathbb {Q} )} is the direct sum of the cyclic group of order 2 and cyclic groups of order p 1 {\displaystyle p-1} for all odd prime p {\displaystyle p} . For n 3 {\displaystyle n\geq 3} , K n M ( Q ) Z / 2 {\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 K 2 M ( F ) {\displaystyle K_{2}^{M}(F)} induced from the inclusion of a global field F {\displaystyle F} to its completions F v {\displaystyle F_{v}} , so there is a morphism

K 2 M ( F ) v K 2 M ( F v ) / ( max. divis. subgr. ) {\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 F {\displaystyle F} .

In addition, for a general local field F {\displaystyle F} (such as a finite extension K / Q p {\displaystyle K/\mathbb {Q} _{p}} ), the Milnor K-groups K n M ( F ) {\displaystyle K_{n}^{M}(F)} are divisible.

K*M(F(t))

There is a general structure theorem computing K n M ( F ( t ) ) {\displaystyle K_{n}^{M}(F(t))} for a field F {\displaystyle F} in relation to the Milnor K-theory of F {\displaystyle F} and extensions F [ t ] / ( π ) {\displaystyle F[t]/(\pi )} for non-zero primes ideals ( π ) Spec ( F [ t ] ) {\displaystyle (\pi )\in {\text{Spec}}(F[t])} . This is given by an exact sequence

0 K n M ( F ) K n M ( F ( t ) ) π ( π ) Spec ( F [ t ] ) K n 1 F [ t ] / ( π ) 0 {\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 π : K n M ( F ( t ) ) K n 1 F [ t ] / ( π ) {\displaystyle \partial _{\pi }:K_{n}^{M}(F(t))\to K_{n-1}F[t]/(\pi )} is a morphism constructed from a reduction of F {\displaystyle F} to F ¯ v {\displaystyle {\overline {F}}_{v}} for a discrete valuation v {\displaystyle v} . This follows from the theorem there exists only one homomorphism
: K n M ( F ) K n 1 M ( F ¯ ) {\displaystyle \partial :K_{n}^{M}(F)\to K_{n-1}^{M}({\overline {F}})}
which for the group of units U F {\displaystyle U\subset F} which are elements have valuation 0 {\displaystyle 0} , having a natural morphism
U F ¯ v {\displaystyle U\to {\overline {F}}_{v}^{*}}
where u u ¯ {\displaystyle u\mapsto {\overline {u}}} we have
( l ( π ) l ( u 2 ) l ( u n ) ) = l ( u ¯ 2 ) l ( u ¯ n ) {\displaystyle \partial (l(\pi )l(u_{2})\cdots l(u_{n}))=l({\overline {u}}_{2})\cdots l({\overline {u}}_{n})}
where π {\displaystyle \pi } a prime element, meaning Ord v ( π ) = 1 {\displaystyle {\text{Ord}}_{v}(\pi )=1} , and
( l ( u 1 ) l ( u n ) ) = 0 {\displaystyle \partial (l(u_{1})\cdots l(u_{n}))=0}
Since every non-zero prime ideal ( π ) Spec ( F [ t ] ) {\displaystyle (\pi )\in {\text{Spec}}(F[t])} gives a valuation v π : F ( t ) F [ t ] / ( π ) {\displaystyle v_{\pi }:F(t)\to F[t]/(\pi )} , we get the map π {\displaystyle \partial _{\pi }} on the Milnor K-groups.

Applications

Milnor K-theory plays a fundamental role in higher class field theory, replacing K 1 M ( F ) = F × {\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

K n M ( F ) H n ( F , Z ( n ) ) {\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:

K n M ( F ) / r H e t n ( F , Z / r ( n ) ) , {\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 n = 2 {\displaystyle n=2} and r = 2 {\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 W ( F ) Z / 2 {\displaystyle W(F)\to \mathbb {Z} /2} given by the dimension of a quadratic form, modulo 2. Milnor defined a homomorphism:

{ K n M ( F ) / 2 I n / I n + 1 { a 1 , , a n } a 1 , , a n = 1 , a 1 1 , a n {\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 a 1 , a 2 , , a n {\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 K n M ( F ) / 2 I n / I n + 1 {\displaystyle K_{n}^{M}(F)/2\to I^{n}/I^{n+1}} is an isomorphism.[11]

See also

References

  1. ^ a b c Milnor, John (1970-12-01). "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.
  2. ^ 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.
  3. ^ a b Shapiro, Jack M. (1981-01-01). "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.
  4. ^ Gille & Szamuely (2006), p. 184.
  5. ^ Voevodsky, Vladimir (2001-07-15). "Reduced power operations in motivic cohomology". arXiv:math/0107109.
  6. ^ 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.
  7. ^ An abelian group is uniquely divisible if it is a vector space over the rational numbers.
  8. ^ Mazza, Voevodsky, Weibel (2005), Theorem 5.1.
  9. ^ Voevodsky (2011).
  10. ^ Elman, Karpenko, Merkurjev (2008), sections 5 and 9.B.
  11. ^ Orlov, Vishik, Voevodsky (2007).

External links