TANT 9 - Non-Archimedean complete fields
Hello there! These are notes for the ninth class of the course "Topics in algebra and number theory" held in Block 4 of the academic year 2017/18 at the University of Copenhagen.
In the previous lectures we have analyzed the set ΣK of places of a field K, and we have completely characterized it when K is a number field. In order to do so we defined in the fourth lecture the notion of completion of a valued field (K,ϕ) and we have seen that every complete, Archimedean field is isomorphic to (R,|⋅|) or to (C,‖⋅‖).
So, what about the non-Archimedean case? Do we have a similar classification result? The answer is a resounding no! More precisely, we have an infinite number of non-Archimedean complete fields which are not isomorphic, as we will see by the end of the lecture.
Recall that for every non-Archimedean absolute value ϕ:K→R≥0 defined on a field K we have that the unit ball Aϕ is a local ring, i.e. a ring with a unique maximal ideal. We have also seen that if ϕ(K×)⊆R>0 is a discrete subgroup then Aϕ is a discrete valuation ring, which implies in particular that the maximal ideal mϕ⊆Aϕ is principal.
We would like thus if for every non-Archimedean absolute value ϕ:K→R≥0 the value group ϕ(K×) was discrete. Unfortunately this is not the case, as the following example shows.
Example 1 Let H≤R>0 be any subgroup, and let F be any field. We define the group ring F[H] to be the set of all formal sums x=∑h∈Hxh[h] where xh∈F and xh=0 for all but a finite number of h∈H. For x,y∈F[H] we define x+y:=∑h∈H(xh+yh)[h]andx⋅y:=∑h∈H(∑h1h2=hxh1yh2)[h] and we see immediately that these operations turn F[H] into a ring. Moreover, this ring is an integral domain. Indeed let x,y∈F[H]∖{0} and let h1:=max{h∈H∣xh≠0} and analogously h2:=max{h∈H∣yh≠0}. Then clearly (x⋅y)h1h2=xh1yh2≠0, because if ab=h1h2 then either a=h1 and b=h2 or either a>h1 or b>h2. This implies that x⋅y≠0, and thus we can define F(H):=Frac(F[H]).
Inspired by the previous proof we can define an absolute value ϕ:F(H)→R≥0 by setting ϕ(0)=0 and ϕ(x)=max{h∈H∣xh≠0} for all x∈F[H]. It is clear that ϕ(xy)=ϕ(x)ϕ(y) and that ϕ(x+y)≤max(ϕ(x),ϕ(y)). Thus ϕ is a non-Archimedean absolute value and ϕ(F(H)×)=H.
The previous exercise shows that every subgroup H≤R>0 can appear as a value group of some non-Archimedean absolute value ϕ:K→R≥0. Moreover, we can assume that K is complete with respect to ϕ using the following exercise.
Exercise 2 Let (K,ϕ) be a non-Archimedean valued field. Prove that its value group and its residue field don't change after completion, i.e. if (Kϕ,Φ) is a completion of (K,ϕ) prove that ϕ(K×)=Φ(K×ϕ) and that κϕ=κΦ. (Hint: prove that for every x∈Kϕ we can find a,b∈K satisfying ϕ(a−x)<ϕ(x) and ϕ(b−x)<1).
We see thus that there is no hope to classify non-Archimedean complete fields!
Nevertheless, there is some hope in the case when the value group is discrete. Thus we give a special name to this case.
Definition 3 Let (K,ϕ) be a valued field. We say that it is discretely valued if ϕ(K×)⊆R>0 is a discrete subgroup.
Definition 4 Let (K,ϕ) be a complete, non-Archimedean, discretely valued field. A uniformizer of (K,ϕ) is any generator of the principal ideal mϕ.
We have already seen that if π∈mϕ is a uniformizer of a complete, non-Archimedean, discretely valued field (K,ϕ) we can write x=uπn for all x∈K×, where u∈A×ϕ and n∈Z. This allows us to show that every element of K admits an expansion as a "power series" in π. For this we will need the content of the following exercise.
Exercise 5 Let (K,ϕ) be a non-Archimedean field, and let {xj}j∈N⊆K be any sequence. Prove that the sequence of partial sums {∑nj=0xj}n⊆K is a Cauchy sequence if and only if {xj}→0.
Exercise 6 Let (K,ϕ) be any valued field, and let {xn}⊆K be a sequence converging to x∞. Prove that ϕ(x∞=limn→+∞ϕ(xn). Hint: prove first that ϕ(a−b)≥|ϕ(a)−ϕ(b)| for all a,b∈K.
Exercise 7 Let (K,ϕ) be a non-Archimedean valued field, and let x,y∈K with ϕ(x)≠ϕ(y). Prove that ϕ(x+y)=max(ϕ(x),ϕ(y)).
Proposition 8 Let (K,ϕ) be a non-Archimedean, complete, discretely valued field. For every k∈N≥1 let πk∈K be a generator of mkϕ and let S⊆K be a set of representatives for the quotient Aϕ/mϕ which contains 0. Then we have that Aϕ={+∞∑k=0akπk∣ak∈S} and if ∑+∞k=0akπk=∑+∞k=0bkπk then ak=bk for all k∈N.
Proof Since mkϕ=πkAϕ=πkAϕ we have that ϕ(πk)=ϕ(π)k→0 as k→+∞. Thus using Exercise 4 we see that for every sequence {ak}⊆S the series ∑+∞k=0akπk converges in K. Using Exercise 5 and Exercise 6 we see now that ϕ(+∞∑k=0akπk)=limn→∞ϕ(n∑k=0akπk)=ϕ(πk0)=ϕ(π)k0wherek0:=min{k∈N∣ak≠0}.
This shows that ∑+∞k=0akπk∈Aϕ.
Let now x∈Aϕ. We can construct inductively a sequence {ak}k⊆S as follows. First of all let a0∈S be such that x≡a0modmϕ. Then since mϕ=π1Aϕ we can write x=a0+π1x1 for some x1∈Aϕ. Now we define a1∈S such that x1≡a1modmϕ and we write x=a0+π1a1+π2x2. We can go on inductively by defining ak∈S to be such that xk≡akmodmϕ and xn+1:=(a0−π1a1−⋯−anπn)/πn+1. It is thus true that x≡∑nj=0ajπjmodmn+1ϕ, which implies that the sequence of partial sums {∑nj=0ajπj} converges to x as n→+∞.
Let now {ak},{bk}⊆K be two distinct sequences. Then we can use what we have proved in the first paragraph of this proof to show that ϕ(+∞∑k=0akπk−+∞∑k=0bkπk)=ϕ(πk0)≠0wherek0:=min{k∈N∣ak≠bk} which implies that ∑+∞k=0akπk≠∑+∞k=0bkπk. Q.E.D.
Corollary 9 Let (K,ϕ) be a non-Archimedean, complete, discretely valued field. Then for every x∈K there exists a unique sequence {ak}+∞k=k0⊆S (where k0∈Z and ak0≠0 ) such that x=∑+∞k=k0akπk. Moreover x∈Aϕ if and only if k0≥0.
Proof We know already that x=uπk0 for unique u∈A×ϕ and k0∈Z. Moreover we can use the previous proposition with πk=πk to show that there exists a unique sequence {bj}+∞j=0⊆S such that u=∑+∞j=0bjπj. The fact that ϕ(u)=1 gives us that b0≠0. Thus we have that x=∑+∞k=k0ajπk where aj:=bj−k0. Q.E.D.
Definition 10 A topological field is a field K which is also a topological space such that the maps K×K→K(x,y)↦x+yandK×K→K(x,y)↦xyandK×→K×x↦x−1 are continuous.
Among all topological fields we need to pay special attention to the ones which are locally compact.
Definition 11 A topological space X is said to be locally compact if for every point x∈X there exist an open subset U⊆X and a compact subset C⊆X such that x∈U⊆C.
We will see later on that topological fields which are locally compact are automatically valued fields, i.e. the topology comes from an absolute value ϕ:K→R≥0. This absolute value comes from the Haar measure that we can define on K thanks to the fact that it is locally compact. For this reason, locally compact fields deserve a special name.
Definition 12 A local field K is a topological field which is locally compact as a topological space.
Even if we cannot prove at the moment that every local field is a valued field, we can prove a nice classification result about complete, non-Archimedean valued fields which are locally compact. In order to do so we need to introduce the field of formal Laurent series over a given field.
Example 13 Let F be a field. We can define the ring of formal power series with coefficients in F as F[[T]]:={+∞∑j=0ajTj∣aj∈F} with the usual operations of sum and product of series, namely +∞∑j=0ajTj++∞∑j=0bjTj:=+∞∑j=0(aj+bj)Tjand+∞∑j=0ajTj⋅+∞∑j=0bjTj:=+∞∑j=0(∑k+l=jakbl)Tj.
One can prove quite easily that F[[T]] is an integral domain and that its field of fraction is the field of formal Laurent series F((T)):={+∞∑j=j0ajTj∣aj∈F,j0∈Z} with the same operations defined above.
We will see in the next lecture how this field arises as a completion of the field F(T) of rational functions with coefficients in F with respect to the absolute value |⋅|t that we defined in the second lecture. We will use this result in the proof of the following theorem.
Theorem 14 Let (K,ϕ) be a non-trivial valued field and suppose that Tϕ is locally compact. Then (K,ϕ) is complete. Moreover, if ϕ is Archimedean then K≅R or K≅C and if ϕ is non-Archimedean then (K,ϕ) is either a finite extension of Qp or a finite extension of Fp((T))
Proof Observe first of all that every closed ball Bε(x):={y∈K∣ϕ(y−x)≤ε} is compact. Indeed since K is locally compact there exists δ∈R>0 such that Bδ(0) is compact. Let now z∈K be any element with ϕ(z)>1, which exists because ϕ is not trivial. Then the balls Bϕ(z)nδ(0) are compact for all n∈N since Bϕ(z)nδ(0)=(μz)n(Bδ(0)), where μz:K→K is the (continuous) multiplication map defined by μz(y):=z⋅y. Thus for every ε>0 the ball Bε(0) is compact because it is a closed subspace of Bϕ(z)nδ(0) for n sufficiently big. Finally we only need to observe that Bε(x)=tx(Bε(0)) for all x∈K, where tx:K→K is the (continuous) translation map defined by tx(y)=x+y.
The compactness of the closed balls implies immediately that K is complete. Indeed let {xn}⊆K be any Cauchy sequence. Since this sequence is Cauchy we can find n0∈N such that xn∈B1(xn0) for all n≥n0. Thus the Cauchy sequence {yn}⊆K defined by yn:=xn+n0 is convergent in B1(xn0) (because every compact metric space is complete, see here), which clearly implies that {xn} converges in K.
Now, if ϕ is Archimedean then we can use Theorem 16 of the fourth lecture to conclude that K is isomorphic either to R or to C. Suppose now that ϕ is non-Archimedean. Then if char(K)=0 we have an embedding Q↪K and the absolute value ϕ restricts to a non-Archimedean absolute value on Q. Now we can use Theorem 7 of the third lecture to see that this restriction must be equivalent to |⋅|p for some prime p∈N. Using the fact that Qp is the completion of Q with respect to |⋅|p (which we will prove in the next lecture) we obtain an embedding Qp↪K. To prove that the degree of this extension is finite we will need to use the fact that if F is a locally compact topological field, and V is a topological vector space over F which is locally compact then V is finite dimensional (Terence Tao proves this when F=R, but the proof generalizes to the general case).
Suppose now that char(K)=p>0. This implies that we get an embedding Fp↪K. Let now π∈K be a uniformizer. Then π is transcendental over Fp, because if this was algebraic then it would be finite, and thus Fp(π) would be a finite field, but Exercise 8 of the second lecture tells us that every valuation on a finite field is trivial, whereas we have that ϕ(π)<1. Thus we have that Fp(π) is isomorphic to the field of rational functions Fp(T) and we get an embedding Fp(T)↪K. This implies (as before) that ϕ restricts to an absolute value ψ on Fp(T) with ψ(T)<1. But we have only one such absolute value, namely |⋅|T, and thus we get an embedding Fp((T))↪K because (as we will prove in the next lecture) Fp((T)) is (isomorphic to) the completion of Fp(T) with respect to |⋅|T. We can apply again the theorem on the finite dimensionality of locally compact vector spaces to prove that this is a finite extension. Q.E.D.
A simple consequence of the previous theorem is that every non-Archimedean valued field (K,ϕ) such that Tϕ is locally compact is also complete and its residue field κϕ is finite. It is quite easy to prove that the following also holds.
Proposition 15 Let (K,ϕ) be a non-Archimedean, complete field whose residue field κϕ is finite. Then Tϕ is locally compact.
Proof We prove first of all that Aϕ is compact. Indeed let {Uα}α∈A be an open cover of Aϕ and suppose by contradiction that it does not admit a finite subcover. Observe now that Aϕ=(x1+πAϕ)∪⋯∪(xn+πAϕ) where π∈mϕ is a uniformizer and {x1,…,xn}⊆Aϕ are representatives for the finite quotient Aϕ/mϕ. Since {x1,…,xn} are finite there exists j0∈{1,…,n} such that xj0+πAϕ is not covered by a finite number of Uα's. Since we have that xj0+πAϕ=xj0+πx1+π2Aϕ∪⋯∪xj0+πxn+π2Aϕ there exists j1∈{1,…,n} such that xj0+πxj1+π2Aϕ is not covered by a finite number of Uα's. Going on like this we can always find jk∈{1,…,n} such that xj0+⋯+πkxjk+πk+1Aϕ is not covered by a finite number of Uα's. Moreover the series ∑+∞k=0xjkπk converges (why?) and we have that x:=∑+∞k=0xjkπk∈Aϕ. Thus x∈Uα for some α∈A and since Uα is open this implies that x+πnAϕ⊆Uα for some n∈N, because {x+πnAϕ}n is a system of open neighborhoods of x. But this is a contradiction, because it would imply that xj0+⋯+πn−1xjn−1+πnAϕ is contained in a finite number of Uα's.
To conclude we only need to observe that for every point x∈K we have that x∈x+πAϕ⊆x+Aϕ where x+πAϕ is clearly open and x+Aϕ is clearly compact. Q.E.D.
In the previous lectures we have analyzed the set ΣK of places of a field K, and we have completely characterized it when K is a number field. In order to do so we defined in the fourth lecture the notion of completion of a valued field (K,ϕ) and we have seen that every complete, Archimedean field is isomorphic to (R,|⋅|) or to (C,‖⋅‖).
So, what about the non-Archimedean case? Do we have a similar classification result? The answer is a resounding no! More precisely, we have an infinite number of non-Archimedean complete fields which are not isomorphic, as we will see by the end of the lecture.
Recall that for every non-Archimedean absolute value ϕ:K→R≥0 defined on a field K we have that the unit ball Aϕ is a local ring, i.e. a ring with a unique maximal ideal. We have also seen that if ϕ(K×)⊆R>0 is a discrete subgroup then Aϕ is a discrete valuation ring, which implies in particular that the maximal ideal mϕ⊆Aϕ is principal.
We would like thus if for every non-Archimedean absolute value ϕ:K→R≥0 the value group ϕ(K×) was discrete. Unfortunately this is not the case, as the following example shows.
Example 1 Let H≤R>0 be any subgroup, and let F be any field. We define the group ring F[H] to be the set of all formal sums x=∑h∈Hxh[h] where xh∈F and xh=0 for all but a finite number of h∈H. For x,y∈F[H] we define x+y:=∑h∈H(xh+yh)[h]andx⋅y:=∑h∈H(∑h1h2=hxh1yh2)[h] and we see immediately that these operations turn F[H] into a ring. Moreover, this ring is an integral domain. Indeed let x,y∈F[H]∖{0} and let h1:=max{h∈H∣xh≠0} and analogously h2:=max{h∈H∣yh≠0}. Then clearly (x⋅y)h1h2=xh1yh2≠0, because if ab=h1h2 then either a=h1 and b=h2 or either a>h1 or b>h2. This implies that x⋅y≠0, and thus we can define F(H):=Frac(F[H]).
Inspired by the previous proof we can define an absolute value ϕ:F(H)→R≥0 by setting ϕ(0)=0 and ϕ(x)=max{h∈H∣xh≠0} for all x∈F[H]. It is clear that ϕ(xy)=ϕ(x)ϕ(y) and that ϕ(x+y)≤max(ϕ(x),ϕ(y)). Thus ϕ is a non-Archimedean absolute value and ϕ(F(H)×)=H.
The previous exercise shows that every subgroup H≤R>0 can appear as a value group of some non-Archimedean absolute value ϕ:K→R≥0. Moreover, we can assume that K is complete with respect to ϕ using the following exercise.
Exercise 2 Let (K,ϕ) be a non-Archimedean valued field. Prove that its value group and its residue field don't change after completion, i.e. if (Kϕ,Φ) is a completion of (K,ϕ) prove that ϕ(K×)=Φ(K×ϕ) and that κϕ=κΦ. (Hint: prove that for every x∈Kϕ we can find a,b∈K satisfying ϕ(a−x)<ϕ(x) and ϕ(b−x)<1).
We see thus that there is no hope to classify non-Archimedean complete fields!
Nevertheless, there is some hope in the case when the value group is discrete. Thus we give a special name to this case.
Definition 3 Let (K,ϕ) be a valued field. We say that it is discretely valued if ϕ(K×)⊆R>0 is a discrete subgroup.
Complete, discretely valued, non-Archimedean fields
Let (K,ϕ) be a complete, non-Archimedean, discretely valued field. Then we know that mϕ⊆Aϕ is a principal ideal. This allows us to give the following definition.Definition 4 Let (K,ϕ) be a complete, non-Archimedean, discretely valued field. A uniformizer of (K,ϕ) is any generator of the principal ideal mϕ.
We have already seen that if π∈mϕ is a uniformizer of a complete, non-Archimedean, discretely valued field (K,ϕ) we can write x=uπn for all x∈K×, where u∈A×ϕ and n∈Z. This allows us to show that every element of K admits an expansion as a "power series" in π. For this we will need the content of the following exercise.
Exercise 5 Let (K,ϕ) be a non-Archimedean field, and let {xj}j∈N⊆K be any sequence. Prove that the sequence of partial sums {∑nj=0xj}n⊆K is a Cauchy sequence if and only if {xj}→0.
Exercise 6 Let (K,ϕ) be any valued field, and let {xn}⊆K be a sequence converging to x∞. Prove that ϕ(x∞=limn→+∞ϕ(xn). Hint: prove first that ϕ(a−b)≥|ϕ(a)−ϕ(b)| for all a,b∈K.
Exercise 7 Let (K,ϕ) be a non-Archimedean valued field, and let x,y∈K with ϕ(x)≠ϕ(y). Prove that ϕ(x+y)=max(ϕ(x),ϕ(y)).
Proposition 8 Let (K,ϕ) be a non-Archimedean, complete, discretely valued field. For every k∈N≥1 let πk∈K be a generator of mkϕ and let S⊆K be a set of representatives for the quotient Aϕ/mϕ which contains 0. Then we have that Aϕ={+∞∑k=0akπk∣ak∈S} and if ∑+∞k=0akπk=∑+∞k=0bkπk then ak=bk for all k∈N.
Proof Since mkϕ=πkAϕ=πkAϕ we have that ϕ(πk)=ϕ(π)k→0 as k→+∞. Thus using Exercise 4 we see that for every sequence {ak}⊆S the series ∑+∞k=0akπk converges in K. Using Exercise 5 and Exercise 6 we see now that ϕ(+∞∑k=0akπk)=limn→∞ϕ(n∑k=0akπk)=ϕ(πk0)=ϕ(π)k0wherek0:=min{k∈N∣ak≠0}.
This shows that ∑+∞k=0akπk∈Aϕ.
Let now x∈Aϕ. We can construct inductively a sequence {ak}k⊆S as follows. First of all let a0∈S be such that x≡a0modmϕ. Then since mϕ=π1Aϕ we can write x=a0+π1x1 for some x1∈Aϕ. Now we define a1∈S such that x1≡a1modmϕ and we write x=a0+π1a1+π2x2. We can go on inductively by defining ak∈S to be such that xk≡akmodmϕ and xn+1:=(a0−π1a1−⋯−anπn)/πn+1. It is thus true that x≡∑nj=0ajπjmodmn+1ϕ, which implies that the sequence of partial sums {∑nj=0ajπj} converges to x as n→+∞.
Let now {ak},{bk}⊆K be two distinct sequences. Then we can use what we have proved in the first paragraph of this proof to show that ϕ(+∞∑k=0akπk−+∞∑k=0bkπk)=ϕ(πk0)≠0wherek0:=min{k∈N∣ak≠bk} which implies that ∑+∞k=0akπk≠∑+∞k=0bkπk. Q.E.D.
Corollary 9 Let (K,ϕ) be a non-Archimedean, complete, discretely valued field. Then for every x∈K there exists a unique sequence {ak}+∞k=k0⊆S (where k0∈Z and ak0≠0 ) such that x=∑+∞k=k0akπk. Moreover x∈Aϕ if and only if k0≥0.
Proof We know already that x=uπk0 for unique u∈A×ϕ and k0∈Z. Moreover we can use the previous proposition with πk=πk to show that there exists a unique sequence {bj}+∞j=0⊆S such that u=∑+∞j=0bjπj. The fact that ϕ(u)=1 gives us that b0≠0. Thus we have that x=∑+∞k=k0ajπk where aj:=bj−k0. Q.E.D.
Local fields
Until now, we have only dealt with valued fields (K,ϕ) but, as we have seen, these are special cases of topological fields.Definition 10 A topological field is a field K which is also a topological space such that the maps K×K→K(x,y)↦x+yandK×K→K(x,y)↦xyandK×→K×x↦x−1 are continuous.
Among all topological fields we need to pay special attention to the ones which are locally compact.
Definition 11 A topological space X is said to be locally compact if for every point x∈X there exist an open subset U⊆X and a compact subset C⊆X such that x∈U⊆C.
We will see later on that topological fields which are locally compact are automatically valued fields, i.e. the topology comes from an absolute value ϕ:K→R≥0. This absolute value comes from the Haar measure that we can define on K thanks to the fact that it is locally compact. For this reason, locally compact fields deserve a special name.
Definition 12 A local field K is a topological field which is locally compact as a topological space.
Even if we cannot prove at the moment that every local field is a valued field, we can prove a nice classification result about complete, non-Archimedean valued fields which are locally compact. In order to do so we need to introduce the field of formal Laurent series over a given field.
Example 13 Let F be a field. We can define the ring of formal power series with coefficients in F as F[[T]]:={+∞∑j=0ajTj∣aj∈F} with the usual operations of sum and product of series, namely +∞∑j=0ajTj++∞∑j=0bjTj:=+∞∑j=0(aj+bj)Tjand+∞∑j=0ajTj⋅+∞∑j=0bjTj:=+∞∑j=0(∑k+l=jakbl)Tj.
One can prove quite easily that F[[T]] is an integral domain and that its field of fraction is the field of formal Laurent series F((T)):={+∞∑j=j0ajTj∣aj∈F,j0∈Z} with the same operations defined above.
We will see in the next lecture how this field arises as a completion of the field F(T) of rational functions with coefficients in F with respect to the absolute value |⋅|t that we defined in the second lecture. We will use this result in the proof of the following theorem.
Theorem 14 Let (K,ϕ) be a non-trivial valued field and suppose that Tϕ is locally compact. Then (K,ϕ) is complete. Moreover, if ϕ is Archimedean then K≅R or K≅C and if ϕ is non-Archimedean then (K,ϕ) is either a finite extension of Qp or a finite extension of Fp((T))
Proof Observe first of all that every closed ball Bε(x):={y∈K∣ϕ(y−x)≤ε} is compact. Indeed since K is locally compact there exists δ∈R>0 such that Bδ(0) is compact. Let now z∈K be any element with ϕ(z)>1, which exists because ϕ is not trivial. Then the balls Bϕ(z)nδ(0) are compact for all n∈N since Bϕ(z)nδ(0)=(μz)n(Bδ(0)), where μz:K→K is the (continuous) multiplication map defined by μz(y):=z⋅y. Thus for every ε>0 the ball Bε(0) is compact because it is a closed subspace of Bϕ(z)nδ(0) for n sufficiently big. Finally we only need to observe that Bε(x)=tx(Bε(0)) for all x∈K, where tx:K→K is the (continuous) translation map defined by tx(y)=x+y.
The compactness of the closed balls implies immediately that K is complete. Indeed let {xn}⊆K be any Cauchy sequence. Since this sequence is Cauchy we can find n0∈N such that xn∈B1(xn0) for all n≥n0. Thus the Cauchy sequence {yn}⊆K defined by yn:=xn+n0 is convergent in B1(xn0) (because every compact metric space is complete, see here), which clearly implies that {xn} converges in K.
Now, if ϕ is Archimedean then we can use Theorem 16 of the fourth lecture to conclude that K is isomorphic either to R or to C. Suppose now that ϕ is non-Archimedean. Then if char(K)=0 we have an embedding Q↪K and the absolute value ϕ restricts to a non-Archimedean absolute value on Q. Now we can use Theorem 7 of the third lecture to see that this restriction must be equivalent to |⋅|p for some prime p∈N. Using the fact that Qp is the completion of Q with respect to |⋅|p (which we will prove in the next lecture) we obtain an embedding Qp↪K. To prove that the degree of this extension is finite we will need to use the fact that if F is a locally compact topological field, and V is a topological vector space over F which is locally compact then V is finite dimensional (Terence Tao proves this when F=R, but the proof generalizes to the general case).
Suppose now that char(K)=p>0. This implies that we get an embedding Fp↪K. Let now π∈K be a uniformizer. Then π is transcendental over Fp, because if this was algebraic then it would be finite, and thus Fp(π) would be a finite field, but Exercise 8 of the second lecture tells us that every valuation on a finite field is trivial, whereas we have that ϕ(π)<1. Thus we have that Fp(π) is isomorphic to the field of rational functions Fp(T) and we get an embedding Fp(T)↪K. This implies (as before) that ϕ restricts to an absolute value ψ on Fp(T) with ψ(T)<1. But we have only one such absolute value, namely |⋅|T, and thus we get an embedding Fp((T))↪K because (as we will prove in the next lecture) Fp((T)) is (isomorphic to) the completion of Fp(T) with respect to |⋅|T. We can apply again the theorem on the finite dimensionality of locally compact vector spaces to prove that this is a finite extension. Q.E.D.
A simple consequence of the previous theorem is that every non-Archimedean valued field (K,ϕ) such that Tϕ is locally compact is also complete and its residue field κϕ is finite. It is quite easy to prove that the following also holds.
Proposition 15 Let (K,ϕ) be a non-Archimedean, complete field whose residue field κϕ is finite. Then Tϕ is locally compact.
Proof We prove first of all that Aϕ is compact. Indeed let {Uα}α∈A be an open cover of Aϕ and suppose by contradiction that it does not admit a finite subcover. Observe now that Aϕ=(x1+πAϕ)∪⋯∪(xn+πAϕ) where π∈mϕ is a uniformizer and {x1,…,xn}⊆Aϕ are representatives for the finite quotient Aϕ/mϕ. Since {x1,…,xn} are finite there exists j0∈{1,…,n} such that xj0+πAϕ is not covered by a finite number of Uα's. Since we have that xj0+πAϕ=xj0+πx1+π2Aϕ∪⋯∪xj0+πxn+π2Aϕ there exists j1∈{1,…,n} such that xj0+πxj1+π2Aϕ is not covered by a finite number of Uα's. Going on like this we can always find jk∈{1,…,n} such that xj0+⋯+πkxjk+πk+1Aϕ is not covered by a finite number of Uα's. Moreover the series ∑+∞k=0xjkπk converges (why?) and we have that x:=∑+∞k=0xjkπk∈Aϕ. Thus x∈Uα for some α∈A and since Uα is open this implies that x+πnAϕ⊆Uα for some n∈N, because {x+πnAϕ}n is a system of open neighborhoods of x. But this is a contradiction, because it would imply that xj0+⋯+πn−1xjn−1+πnAϕ is contained in a finite number of Uα's.
To conclude we only need to observe that for every point x∈K we have that x∈x+πAϕ⊆x+Aϕ where x+πAϕ is clearly open and x+Aϕ is clearly compact. Q.E.D.
Conclusions and references
In this lecture we managed to:- prove that every element of a non-Archimedean, complete, discretely valued field (K,ϕ) can be written in a unique way as a power series in the uniformizer π∈K;
- give a characterization of local fields as finite extensions of R, Qp and Fp((T)).
- Section 2 of these notes by Stevenhagen;
- these notes by Clark;
- these notes by Sutherland;
- these notes by Bouyer.
Comments
Post a Comment