Ramification of -power torsion points of formal groups
Abstract.
Let be a rational prime, let denote a finite, unramified extension of , let be the completion of the maximal unramified extension of , and let be some fixed algebraic closure of . Let be an abelian variety defined over , with good reduction, let denote the Néron model of over , and let be the formal completion of along the identity of its special fiber, i.e. the formal group of .
In this work, we prove two results concerning the ramification of -power torsion points on . One of our main results describes conditions on , base changed to , for which the field is a tamely ramified extension where denotes the group of -torsion points of over . This result generalizes previous work when is -dimensional and work of Arias-de-Reyna when is the Jacobian of certain genus 2 hyperelliptic curves.
Key words and phrases:
Abelian varieties, Formal Groups, Ramification1991 Mathematics Subject Classification:
11G10 (14K20, 11G25, 14L05)1. Introduction
In this work, we are interested in studying the ramification behaviour of the -torsion points of the formal group associated to an abelian variety over an unramified local field.
Let be a rational prime, let denote a finite, unramified extension of , let be the completion of the maximal unramified extension of , let be some fixed algebraic closure of , and let . Let be an abelian variety defined over , with good reduction, let denote the Néron model of over , and let be the formal completion of along the identity of its special fiber, i.e. the formal group of .
To state these results, we need the following definitions. In [Fon82], Fontaine studied the -module of Kähler differentials of over , or over . The -module is a torsion and -divisible -module, with a semi-linear action of . Let denote the canonical derivation, which is surjective. We denote by , the kernel of , which is an -sub-algebra of . Membership of an element of inside of reflects ramification properties of (see e.g., 2.2 and 2.3).
By using previous work of the authors [IMZ21], we are able to prove that the -points of the Tate module of are trivial, which implies that following theorem.
Theorem A.
Let be an abelian variety over with good reduction. Then there is such that for every and , we have .
For a more concretely description of what this means in terms of the coordinates of the torsion point , we refer the reader to 3.6.
Our second result gives conditions on for which one may take in A. More precisely, we describe a condition on a formal group of dimension over which implies that , the field of definition is tamely ramified. The condition is discussed in Section 4 and is related to a symmetric formal group law from [dR11].
Theorem B.
Let be a strict (4.1) formal group of dimension over . For , the field of definition is tamely ramified and . Moreover, is tamely ramified.
1.1. Related results
In [Ser72, Section 1], Serre showed that for an elliptic curve with good supersingular reduction, the field extension is tamely ramified, and his proof relies on a detailed study of the formal group attached to . In particular, Serre explicitly determined the -adic valuation of the points on (note that when has good supersingular reduction we have that where is the formal group of the Néron model of ), which allowed him to embed into a certain vector space on which the wild inertia group acts trivially.
In her thesis, Arias-de-Reyna generalized this approach, and in [dR11, Theorem 3.3], she showed that if there exists a positive rational number such that for all , the minimum of -adic valuation of the coordinates of equals , then the action of wild inertia on is trivial, and so is tamely ramified. She goes on to define the notion of a symmetric formal group law on a formal group of dimension 2, and then proves [dR11, Theorem 4.15] that if formal group of dimension 2 has a symmetric formal group law and height 4, then the above statement about the -adic valuation of the -torsion points holds. Later in [dR11, Theorem 5.9], she identifies a family of genus 2 curves whose Jacobians have associated formal groups with a symmetric formal group law and height 4, and hence their -torsion defines a tamely ramified extension. We also mention work of Rosen and Zimmerman [RZ89], in which the authors study the Galois group of where is a generic commutative formal group of dimension 1 and height .
In the global setting i.e., when working over a number field , Coleman [Col87] studied the ramification properties of torsion points on abelian varieties in relation to the Manin–Mumford conjecture. More precisely, he conjectured (loc. cit. Conjecture B) that for a smooth, projective, geometrically integral curve and any Galois stable torsion packet in , the field is unramified at a certain prime of . Coleman proved this conjecture when is large enough, and using work of Bogomolov, he provided a new proof of the Manin–Mumford conjecture.
1.2. Outline of paper
In Section 2, we recall the definition of the Fontaine integral, our previous work on the kernel of the Fontaine integral [IMZ21], and a different perspective on the Fontaine integral via work of Wintenberger. In Section 3, we prove A. We conclude in Section 4 with the definition of a strict formal group and our proof of B.
1.3. Conventions
We establish the following notations and conventions throughout the paper.
Fields
Fix a rational prime . Let denote the completion of maximal unramified extension of , let be a fixed algebraic closure of , and let denote the completion of with respect to the unique extension of the -adic valuation on (normalized such that ). For a tower of field extensions , we denote by and respectively the absolute Galois groups of and respectively. We denote .
Abelian varieties
We will consider an abelian variety defined over some subfield such that , with good reduction over . Let denote the Néron model of over and also denote by the formal completion of along the identity of its special fiber, i.e. the formal group of . We note that the formation of Néron models commutes with unramified base change. We will denote the Tate module of (resp. the Néron model of ) by (resp. ). We note that as -modules.
Formal groups
2. Fontaine integration for abelian varieties with good reduction
In this section, we recall the construction of the Fontaine integration as well as our previous work concerning the kernel of the Fontaine integral.
The differentials of the algebraic integers
First, we recall for the reader’s convenience the notation established above. Let denote the maximal unramified extension of , let be a fixed algebraic closure of , and let denote the completion of . Let denote the absolute Galois group of . We denote . Fix a finite extension of in . For a -representation , the -th Tate twist of is denoted by , which is just the tensor product of with the -fold product of the -adic cylcotomic character .
In [Fon82], Fontaine studied a fundamental object related to these choices, namely the -module of Kähler differentials of over , or over . The -module is a torsion and -divisible -module, with a semi-linear action of . Let denote the canonical derivation, which is surjective.
Important examples of algebraic differentials arise as follows: Let denote a compatible sequence of primitive th roots of unity in . Then
Next, we recall a theorem of Fontaine.
Theorem 2.1 ([Fon82, Théorème 1’]).
Let denote a compatible sequence of primitive th roots of unity in . For , write for some . The morphism defined by
is surjective and -equivariant with kernel
Moreover, and .
We denote by , the kernel of , which is an -sub-algebra of . Indeed, if , then , and so . In order to better understand , we recall a construction from the first and last author [IZ99].
Definition 2.2.
Let . Let be a finite extension which contains , let be a uniformizer of , and let be such that . Then, define
where denotes the different ideal of . Note that does not depend on , , or , and so it defines a function .
Lemma 2.3 (Properties of ).
The function from 2.2 satisfies the following properties.
-
(1)
If , then , and if , then we have equality.
-
(2)
If , then .
-
(3)
If and , then .
-
(4)
If , then if and only if .
-
(5)
For , if and only if .
-
(6)
The formula is well-defined and give a map , which makes the obvious diagram commutative.
We will use the follow properties of in our study of the Fontaine integral.
Lemma 2.4 ([IZ99, Lemma 2.2]).
Let be such that . Then there exists such that .
Proposition 2.5 ([IZ99, Theorem 2.2]).
Let be an algebraic extension. Then is deeply ramified (loc. cit. Definition 1.1) if and only if is unbounded.
The definition of Fontaine’s integration
We are now ready to define Fontaine’s integration. Let and respectively denote the -modules of invariant differentials on and respectively its Lie algebra. Note that being invariant implies that and where is the group law in .
Definition 2.6.
Let and . Each corresponds to a morphism , and hence we can pullback along this map giving us a Kähler differential . The sequence is a sequence of differentials in satisfying , and hence defines an element in .
The Fontaine integration map
is a non-zero, -equivariant map defined by
Remark 2.7.
Using Theorem 2.1 and the function from 2.2, we can give an alternative description of the Fontaine integration map. Let and . Each corresponds to a morphism , and hence we can pullback along this map giving us a Kähler differential .
For every , there is a maximal such that with where is some primitive -th root of unity. To see this, we first note that
for any primitive -th root of unity. This result follow from the definition of and a result of Tate [Tat67, Proposition 5] on the valuation of the different ideal of . By taking where denotes the greatest integer of the real number , we can use 2.3.(6) and 2.4 to deduce the above equality.
Now using Theorem 2.1, we have that
Moreover, using the definition of and this above interpretation, we can see that if (i.e., if is an unramified path), then . Indeed, it is clear from the definition of that .
The kernel of the Fontaine integral
In [IMZ21], we studied the kernel of . As noted in 2.7, we have that lies in , and in [IMZ21, Theorem 4.5, Theorem A.4], we showed that . In proving these results, we determined the kernel of the Fontaine integral when restricted to the Tate module of the formal group of . This result will play a role later on, and so we present it below.
Theorem 2.8 ([IMZ21, Theorem 5.5]).
Let be an abelian variety over with good reduction, let denote its Néron model, and let be the formal group of . The Fontaine integral restricted to the Tate module of is injective i.e., .
Another point of view on the Fontaine integration map
In this subsection, we give another perspective on the Fontaine integration map, which will naturally lead us towards an application of Theorem 2.8.
We keep all the notations from the previous sections and Subsection 1.3. Recall that we let and we have the -module with its canonical derivation . Note that is surjective and is -divisible, and let us denote .
Lemma 2.9.
Let denote the -adic completion of . Then, the exact sequence of -modules
induces another exact sequence:
where is an -algebra homomorphism and is seen as an ideal of of square .
Proof.
The statement follows from [Col12, Lemme 3.8] and also from [IZ99, Corollary 1.1], but we present another proof below.
We consider the diagram
The snake lemma gives the exact sequence of -modules:
By taking the projective limit with respect to of this exact sequence, we obtain the claim. ∎
Recall that we have the isomorphism . By 2.9, we have the short exact sequence
where is an ideal of such that .
By definition, we have
and hence we have the following short exact sequence of abelian groups with -action
Consider the following commutative diagram with exact rows
The snake lemma gives a -equivariant map
and by taking the projective limit over ’s, we obtain a map
Proposition 2.10.
The map obtained above
coincides with Fontaine’s integral, i.e. we have .
3. Consequences of Theorem 2.8: ramification of -power torsion points on
In this section, we use the interpretation of the Fontaine integral from 2.10 and Theorem 2.8 to deduce properties concerning the ramification of -power torsion points on the formal group of .
To begin, we recall the diagram
Above, we only wrote a piece of the snake lemma, and by writing more of it, we have an exact sequence of -modules
By taking projective limits, we have the exact sequence
(3.1) |
Therefore, Theorem 2.8 implies that .
To study consequences of this property, we will use another ring instead of .
Definition 3.1 ([Fon94]).
Let denote the projection map. Then, we define . In [Fon94, Remark 1.4.7], Fontaine gives the following construction of . Let us recall that
and that and are -modules. We make into a commutative ring by defining multiplication as follows: for , i.e. we require that is an ideal of of square . Then we have
By 3.1, we have an exact sequence of -modules
where , and the -adic completion of is . We note that we may construct the diagram above in the same way using instead of , which produces the exact sequence (LABEL:eqn:SESDf) with instead of . Instead of the exact sequence (3.1) above, we will have the following exact sequence
Again, the Theorem 2.8 implies that .
We will use this observation to deduce that . In order to do so, we need to show that , for all , which is accomplished through the following two lemmas.
Lemma 3.2.
Let , then there is with and such that .
Proof.
Recall that , i.e. we have an exact sequence
and this exact sequence splits over , i.e. the following diagram is cartesian and has exact rows
In particular, the section is defined by . Then defines a morphism , and if , then . ∎
In the next lemma, we show a converse of 3.2 at least for the formal group of .
Lemma 3.3.
Let denote the formal group of the abelian scheme and fix an integer. Let
and let be a point such that . Then for all .
Proof.
Let and denote by
the multiplication by on . Let , where is the ideal generated by , then we know is a finite flat -algebra, in particular is a free -module and with the co-multiplication of , is a finite flat group-scheme, so is an étale, therefore smooth, group-scheme over . This implies that the image in of the determinant of the matrix:
is a unit.
Let now , i.e. there is such that is not in . Let such that , i.e. , with and , for all . By the above assumption . As in we have for all , the Taylor formula implies that if we must have:
for every . But the determinant of the matrix
is a unit in , i.e. it is non-zero and . This is a contradiction. ∎
Remark 3.4.
We note that the group-scheme is not smooth over (for example could have nilpotents). We also remark that as is smooth, the map is surjective, but this is clear as .
3.2 and 3.3 imply that the map gives an isomorphism , for all . Combining this with the fact that , we have the following result.
Theorem 3.5 (A).
Let be an abelian variety over with good reduction. Then there is such that for every and , we have
Remark 3.6.
We observe that the above is a result regarding ramification properties of the -power torsion points of the formal group of our abelian variety with good reduction over (c.f. 2.5). More precisely, let and be as in the theorem above. Let with for . Theorem 3.5 says the following: let , let denote a uniformizer of and let denote the different ideal of . For every let be polynomials such that , for every . Then there is such that (i.e. ).
4. A theorem on the ramification type of the field obtained by adjoining a -torsion point of a formal group
In this subsection we study the ramification properties of the extension , where is a non-zero -power torsion point of the formal group of .
We continue to denote by the completion of the maximal unramified extension of in an algebraic closure of , which we denote . Let denote a formal group of dimension over . For example, can be the formal group of the Néron model of over .
To begin, we define the notion of a strict formal group.
Definition 4.1.
Consider the multiplication-by- map
on where each is a power series in with coefficients in . For each , define to be the form comprised of monomials of which have unit coefficient and minimal degree, where we consider each monomial to be of degree 1.
Let denote the degree of these forms , respectively, which we note are (possibly distinct) powers of . Let denote the reductions modulo of the forms .
Consider the system of equations
(4.1) |
We say that the formal group is strict if and the only solution to (4.1) is .
Remark 4.2.
If is a formal group of dimension , then it is clear that is strict since we will have that , where is the height of . Moreover, if is the product of -dimensional formal groups, then again is strict.
Remark 4.3.
We can given an equivalent characterization of strict as follows. Consider the matrix where the entry consists of the coefficient of in the linear form for each . The condition that be strict is equivalent to the determinant of being non-zero.
We refer the reader to [dR11, Remark 4.14] for an example of a -dimensional formal group of height where this condition holds, and here we note that the above degrees all equal . Moreover, we remark that the proof from loc. cit. holds for any -dimensional formal group of height where the degrees all equal .
With this definition, we can state our main result.
Theorem 4.4.
Let be a strict formal group of dimension . For , the field of definition is tamely ramified and . Moreover, is tamely ramified.
Proposition 4.5.
Let be a strict the formal group of dimension . For every , the coordinates of are not all in .
Proof.
Let be a non-zero -torsion point. By Theorem 4.4, we know that the extension is tamely ramified and that there exists some coordinate which is a uniformizer for . By [Ser79, Proposition III.6.13], we have that where is the different ideal of . Now since is a uniformizer for , we have that
where is the function defined in 2.2. By 2.3.(5), as desired. ∎
For the remainder of this section, we focus on proving Theorem 4.4. The proof can be broken down into three steps.
-
(1)
Given a non-zero -torsion point , we will carefully construct linear combinations of the which satisfy nice properties in terms of their valuations and distances between their -conjugates. See 4.6.
-
(2)
Next, we consider the change of variables (i.e., the isomorphism of formal groups) which sends the coordinate to the linear combination described above. We use the properties of the and the strictness of to precisely determine the valuation of and to estimate the valuation of the difference between them and their -conjugates.
-
(3)
Finally, we use Krasner’s lemma to deduce that one of the original coordinates of must be a uniformizer for the maximal order of , from which Theorem 4.4 follows.
Lemma 4.6.
Let be a formal group of dimension over . Let . There exist linear combinations of with coefficients in which satisfy:
-
(1)
,
-
(2)
,
-
(3)
for all where is the Galois closure of ,
and such the matrix representing the change of coordinates is invertible. Here the exponent indicates the transpose of a matrix.
Proof.
Let . Our proof will involve making a series of linear combinations. To begin, we will construct the element . First, consider all the linear combinations of the form
(4.1) |
By our assumptions on , the set of is infinite, with as in the above formula, and hence we may find one linear combination, call it , satisfying the following two conditions:
-
(1)
for all other linear combinations from ,
-
(2)
.
To show that condition (2) holds, consider the following. There are exactly embeddings of into the fixed algebraic closure , call them . Note that the vectors , , are distinct. Indeed, if for some the vectors and coincide, then and will coincide at and so they will coincide on , which is not the case.
Consider now, for each pair with and , the hyperplane given by
Since the vectors , are distinct, none of covers the full space. Denote by the union of these finitely many hyperplanes. Choose now any such that the point lies outside . Then we claim that the element
satisfies . Indeed, are distinct. For, if two of them are equal, say with , then is forced to lie in . Thus are distinct, so has at least distinct conjugates over . Hence and in conclusion i.e., has degree over .
We pause to note that the matrix representing the change of coordinates is invertible. Indeed, the matrix has units along the diagonal, the coefficients of the linear combination in the first row, and zeros elsewhere, hence the determinant is a unit.
We now look at the distances between these linear combinations and various of their conjugates over . Fix and consider the infimum of the values for the linear combinations as in (4.1), i.e. we look at . We first show that this infimum exists and is attained by some linear combination. Note that if , then since all linear combinations belong to , we have that for all linear combinations , i.e. . If we consider such that , then we at least have that for some linear combinations , for example for . We have that the set of possible valuations is discrete, and the set of values has a lower bound, namely . Thus, the infimum of the set of values is attained by some linear combination from (4.1), but note it need not be obtained by .
Let denote where is the Galois closure of . For a fixed , let denote one such linear combination attaining the minimum . We claim that we can find a linear combination of and of all of these where which simultaneously achieves these minima. To do this, consider all linear combinations
(4.2) |
We will choose the ’s such that each such will live in .
To achieve our desired simultaneous minima, we start with one , call it . First, we let . This linear combination might work in that it already attains the minimum at ; by this we mean that holds for all linear combinations from (4.1). If this is the case, then we set . Now suppose that does not attain the minimum at . In this case, we may use any unit and set . Indeed, for any unit and above, we have that
because . Let with unit such that . We have that for such , for all and that the with have the property that avoids a finite number of elements in .
For another automorphism , we proceed along the same lines, that is: if has the property that for all we set . If the above is not true, let , for some . Then as above we have: therefore realizes the minimum for , for all for which . What about for . The worst that can happen is that , i.e. if we denote by a uniformizer of , we have , with . Therefore Now the residue field of is , therefore by choosing such that we have and therefore realizes the minima for both and .
Continuing in this fashion, we arrive at the conclusion that there exist linear combinations of the form
(4.3) |
where which satisfy the following four conditions:
-
(1)
,
-
(2)
,
-
(3)
,
-
(4)
for all ,
as desired. Again, we pause to note that the matrix representing the change of coordinates is invertible. Indeed, the matrix has units along the diagonal, the coefficients of the linear combination in the first row, and zeros elsewhere, hence the determinant is clearly a unit.
We now wish to iterate this construction as follows. First, consider the set of all linear combinations
(4.4) |
Then, we can repeat the above construction to arrive at a linear combination satisfing:
-
(1)
for all other linear combinations from ,
-
(2)
.
Furthermore, we can follow the above construction to say that there exist linear combinations of the form
(4.5) |
where which satisfy the following four conditions:
-
(1)
,
-
(2)
,
-
(3)
,
-
(4)
for all ,
Note that the matrix representing the change of coordinates is invertible. Indeed, has units on the diagonal, the coefficients of the linear combination in the second row, and has zeros everywhere else. Although this matrix is not triangular, we can make it so by switching the second row with the first and interchanging the first and second columns; these operations will not change the determinant. After these operations, the matrix becomes triangular with units on the diagonal, and hence the will be invertible. Moreover, we see that the matrix representing the change of coordinates is invertible since .
We continue in this fashion for all of the remaining coordinates and arrive at our desired claim, namely that there exists linear combinations of with coefficients in which satisfy:
-
(1)
,
-
(2)
,
-
(3)
for all
and such the matrix representing the change of coordinates is invertible. ∎
We now complete the proof of Theorem 4.4.
Proof of Theorem 4.4.
Fix and let . We first remark that since is completion of the maximal unramified extension, we have that , and hence the extension is totally ramified. Indeed, this follows from that fact that the group-scheme is connected. In 4.6, we constructed linear combinations of satisfying certain properties. Let denote the same linear combinations of the coordinates (so if we were to evaluate at we would recover ). The last condition from 4.6 implies that the change of variables is an isomorphism of formal groups.
We claim that the isomorphism of formal groups will preserve strictness. As each of the are linear combinations of with coefficients in , this isomorphism of formal groups will act linearly on terms of minimal degree, and hence it changes by a linear transformation, which is invertible. We note that it also does the same to , therefore, it transforms the set of solutions of the system by an invertible transformation. Moreover, the system having or not having a single solution is the same before or after an isomorphism. We pause to note that it is crucial that the degrees from 4.1 are all equal
For the remainder of the proof, we work with this isomorphic formal group with coordinates . We note that the vector will reduce mod to the point because all have valuations strictly positive. But the have the same valuation so we can divide all of them by one of them, and consider the vector , which will not reduce the zero vector over the residue field. Since was assumed to be strict, the reduction of cannot be a common root of all of . Therefore, there exists an index for which is not zero in the residue field , and hence the valuation of equals the valuation of each of its individual monomials.
We now want to determine the valuation of , and hence the valuation of every other as they have the same valuation. By considering the equation , we have
We claim that where is the degree of . To see this choose a unit in that is a representative for the element in the residue field corresponding to , and similarly choose . Then is a unit in , because its image in the residue field is nonzero, by the above choice of . But this is a form (of degree ) so we can divide by inside , and re-denoting the ’s in consideration, we have that is a unit, and moreover,
Plugging this into the the above equation, we arrive at the equation
(4.6) |
The minimum valuation in the equality of (4.6) must be attained in at least two terms, and these terms are forced to be and . Our claim now follows since is a unit in .
We now want to study the relationship between the valuations where . Let which is a unit in . For each , we will consider (4.6) and
(4.7) |
where corresponds to terms strictly larger valuation. If we subtract equality (4.7) from (4.6), we arrive at the following:
(4.8) |
By condition (3) of 4.6, the valuation of the “interesting terms" from (4.8) will be larger than . Indeed, these interesting terms are in fact monomials of the form where some of the could be zero and the total degree is strictly greater than . In any case, we may deal with the difference of such monomials and their conjugates as follows. Suppose for example, that we have a term of the for . Then by adding and subtracting the term , the difference we need to deal with can then be written as times something of positive valuation, plus times something of positive valuation, and so we can use property (3) of 4.6 for this particular to get our desired claim.
We now arrive at the crucial claim of the proof. Recall that denotes the Galois group of the Galois closure of . We claim that for all . Assume that where . By the above discussion and condition (3) of 4.6, we can save the value from each term, and since there must be at least two terms of equal valuation in (4.8), we have that
Note that we cannot guarantee the equality of these valuations because there could be other terms of total minimal degree other than , but the above inequality will suffice. Using the above inequality and previous equality , and letting , we have that
Let . We have that , which by assumption is strictly greater than 0. We now arrive at a contradiction by considering the above inequality and the equation
and noting that all terms in the above have valuation strictly greater than . Therefore, we have that , and hence for all .
To conclude our proof, we use Krasner’s lemma [Ser79, Exercise II.2.1] to explicitly describe the extension and show that . Recall that our satisfies the equation (4.6). Consider the polynomial where as above. Note that is not a root of , but it satisfies the following inequality where the right side here is also equal to .
On the other hand, the roots of each have valuation exactly since is Eisenstein at . We now compute the valuation of the derivative of evaluated at a root in two different ways. First, and hence
Second, we directly compute that which yields
The first inequality and the second equality imply that for all . We also have that and hence
There are exactly terms here and since , it follows that at least one term must be strictly larger than . Without loss of generality, we may assume that Now we have the strict inequality
for all , and hence Krasner’s lemma implies that . Recall that we have just shown that for all . Moreover, we can just switch and to arrive at the inequality
and so we can apply Krasner’s lemma again to deduce that . Therefore, we have shown that and this extension is totally ramified of degree , hence tamely ramified. We also have that , which is a linear combination of with unit coefficients, is a uniformizer for and hence . Finally, we note that there exists some coordinate of which has valuation by condition (3) of 4.6 and hence is a uniformizer for as well.
The second statement of Theorem 4.4 follows because is the compositum of the fields as varies over points in and the compositum of tamely ramified extensions is again tamely ramified. ∎
References
- [Col87] Robert F. Coleman, Ramified torsion points on curves, Duke Mathematical Journal 54 (1987), no. 2, 615–640.
- [Col92] Pierre Colmez, Périodes -adiques des variétés abéliennes, Mathematische Annalen 292 (1992), no. 4, 629–644.
- [Col12] by same author, Une construction de , Rend. Semin. Mat. Univ. Padova 128 (2012), 109–130 (2013). MR 3076833
- [dR09] Sara Arias de Reyna, Galois representations and tame Galois realizations, Ph.D. thesis, Universidad de Barcelona, 2009.
- [dR11] by same author, Formal groups, supersingular abelian varieties and tame ramification, Journal of Algebra 334 (2011), no. 1, 84–100.
- [Fon94] Jean-Marc Fontaine, Exposé II : Le corps des périodes -adiques, Périodes -adiques - Séminaire de Bures, 1988 (Jean-Marc Fontaine, ed.), Astérisque, no. 223, Société mathématique de France, 1994, talk:2 (fr). MR 1293971
- [Fon82] by same author, Formes Différentielles et Modules de Tate des Variétés abéliennes sur les Corps Locaux., Inventiones mathematicae 65 (1981/82), 379–410 (fre).
- [Haz12] Michiel Hazewinkel, Formal groups and applications, AMS Chelsea Publishing, Providence, RI, 2012, Corrected reprint of the 1978 original. MR 2987372
- [IMZ21] Adrian Iovita, Jackson S. Morrow, and Alexandru Zaharescu, On -adic uniformization of abelian varieties with good reduction, Preprint, arXiv:2107.09165 (to appear in Compositio Mathematica) (September 20, 2021), with an appendix by Yeuk Hay Joshua Lam and Alexander Petrov.
- [IZ99] Adrian Iovita and Alexandru Zaharescu, Galois theory of , Compositio Mathematica 117 (1999), no. 1, 1–33.
- [RZ89] Michael Rosen and Karl Zimmermann, Torsion points of generic formal groups, Trans. Amer. Math. Soc. 311 (1989), no. 1, 241–253. MR 974776
- [Ser72] Jean-Pierre Serre, Propriétés galoisiennes des points d’ordre fini des courbes elliptiques, Invent. Math. 15 (1972), no. 4, 259–331. MR 387283
- [Ser79] by same author, Local fields, Graduate Texts in Mathematics, vol. 67, Springer-Verlag, New York, 1979, Translated from the French by Marvin Jay Greenberg.
- [Tat67] John Tate, -Divisible groups, Proceedings of a conference on Local Fields, Springer, 1967, pp. 158–183.
- [Win94] Jean-Pierre Wintenberger, Exposé IX : Théorème de comparaison -adique pour les schémas abéliens — I : Construction de l’accouplement de périodes, Périodes -adiques - Séminaire de Bures, 1988 (Jean-Marc Fontaine, ed.), Astérisque, no. 223, Société mathématique de France, 1994, talk:9 (fr). MR 1293978