|
Abstract
Transseries provide a universal framework for the formal
asymptotics of regular solutions to ordinary differential equations at
infinity. More general functional equations such as
may have solutions that grow faster than any iterated exponential and
thereby faster than any transseries.
In order to develop a truly universal framework for the
asymptotics of regular univariate functions at infinity, we therefore
need a generalization of transseries: hyperseries. Hyperexponentials
and hyperlogarithms play a central role in such a program. The first
non-trivial hyperexponential and hyperlogarithm are
and its functional inverse
,
where
satisfies the above equation. Formally,
such functions
and
can
be introduced for any ordinal
.
For instance,
,
,
,
and
satisfies
.
In the present work, we construct a field of hyperseries
that is closed under and
for all ordinals
. This
generalizes previous work by Schmeling [29] in the case
when
, as well as the
previous construction of the field of logarithmic hyperseries by van
den Dries, van der Hoeven, and Kaplan [12].
In order to get our hands on a complex mathematical expression, we first
need to simplify it as much as possible. This can often be achieved by
eliminating parts that are asymptotically negligible. For instance, when
studying the expression for large values of
, we may compute the
approximations
,
, and then
.
Such approximations rely on our ability to determine and compare
asymptotic rates of growth.
Is it possible to develop a universal framework for this kind of
asymptotic simplification? This sounds like a difficult problem in
general, especially for multivariate functions or functions with an
irregular growth like . On
the other hand, the problem might become tractable for univariate
functions
in a neighborhood of infinity
, provided that
is constructed using a limited number of well-behaved primitives.
An important first step towards a systematic asymptotic calculus of this
kind was made by Hardy in [20, 21], based on
earlier ideas by du Bois-Reymond [14, 15, 16]. We say that is an
-function if it is constructed
from
and the real numbers
using the field operations, exponentiation, and logarithms. Given two
non-zero germs of
-functions
,
at
infinity, Hardy proved that exactly one of the relations
,
,
or
holds, where
Hardy also observed [20, p. 22] that “The only scales
of infinity that are of any practical importance in analysis are those
which may be constructed by means of the logarithmic and exponential
functions.” In other words, Hardy suggested that the framework of
-functions not only allows
for the development of a systematic asymptotic calculus, but that this
framework is also sufficient for all “practical” purposes.
Hardy went on [20, chapter V] with the examination of
possible counterexamples, through the exploration of pathological
functions whose asymptotic behavior does not conform to any
logarithmico-exponential scale. Here he made a distinction between
irregular asymptotic behavior (such as oscillating functions) and
regular asymptotic behavior that yet cannot be described in terms of
-functions. Basic examples of
the second type were already constructed by du Bois-Reymond and Hardy
[20, chapter II]. For instance, let
for
and let
for each
. Then
grows faster than any
-function.
In order to formalize the concept of “regular” growth at
infinity, we focus on classes of (germs of) functions that are stable
under common calculus operations such as addition, multiplication,
differentiation, and composition. The class of -functions indeed satisfies these conditions, but it
is interesting to investigate whether there exist larger classes of
functions with similar properties.
Two particular settings that have received a lot of attention are Hardy
fields (i.e. fields of germs of real continuously
differentiable functions at infinity that are closed under
differentiation [7]) and germs of definable functions in
-minimal structures [10].
Each of these settings excludes oscillatory behavior in a strong sense.
For instance, although the function
does not
oscillate for large values of
,
its second derivative does, so the germ of this function at infinity
does not belong to a Hardy field.
With a more precise definition of regularity at hand, one may re-examine
the existence of regular functions whose asymptotic growth falls outside
any scale of -functions. For
instance, the function
from above is not even
continuous and thereby not sufficiently regular. Nevertheless, it was
shown by Kneser [28] that the functional equation
has a real analytic solution on .
This provides us with a more natural candidate for a regular function
that grows faster than any
-function.
Indeed, it was shown by Boshernitzan [6] that Kneser's
solution belongs to a Hardy field. The functional inverse
of
frequently occurs in the
complexity analysis of recursive algorithms that use exponential size
reductions. For instance, the fastest known algorithm [22]
for multiplying two polynomials of degree
in
runs in time
.
This shows that Hardy's framework of
-functions
is insufficient, even for practical purposes.
Another example of a regular function that is not asymptotic to any
-function is the functional
inverse of
. This fact was
actually raised as a question by Hardy and only proved in [23]
and [13]. More explicit examples of such functions, like
, were given in [23].
It turns out that the class of
-functions
lacks several important closure properties (e.g. functional
inversion and integration), which makes it unsuitable as a universal
framework for asymptotic calculus.
The class of transseries forms a better candidate for such a universal
framework. A transseries is a formal object that is constructed from
(with
)
and the real numbers, using exponentiation, logarithms, and
infinite sums. One example of a transseries is
Depending on conditions satisfied by their supports, there are different types of transseries. The first constructions of fields of transseries are due to Dahn and Göring [9] and Écalle [17]. More general constructions were proposed subsequently by van der Hoeven and his student Schmeling [23, 29].
Transseries form a natural “infinitary” extension of the
concept of an -function. The
transseries are closed under integration and functional inversion [23, 13]. They also satisfy a differential
intermediate value property [26, Chapter 9]. However,
transseries are only defined formally, so their analytic meaning is not
necessarily obvious. One technique for associating an analytic meaning
to certain divergent transseries is accelero-summation [17],
a generalization of Borel summation [5]. An alternative
technique is based on differential algebra and model theory [27,
1]. In this paper, we focus on formal asymptotic
computations, without worrying about analytic counterparts.
Despite the excellent closure properties of transseries for the
resolution of differential equations, the functional equation (1.1)
still does not have a transseries solution. In order to establish a
universal formal framework for asymptotic calculus, we therefore need to
incorporate extremely fast growing functions such as , as well as formal solutions
,
,
etc. to the following equations:
The fast growing functions ,
,
are
called hyperexponentials. Their functional inverses
,
,
are called hyperlogarithms and they
grow extremely slowly. The first construction of a field of generalized
transseries that is closed under
and
for all
was accomplished in [29]. Here we understand that
and
.
The hyperlogarithms ,
, etc. obviously satisfy the
functional equations
In addition, we have a simple formula for their derivatives
where and
for all and
.
The formula (1.4) is eligible for generalization to
arbitrary ordinals
. Taking
, we note that the function
does not satisfy any functional equation. Yet
any truly universal formal framework for asymptotic calculus should
accommodate functions such as
for the simple
reason that it is possible to construct models with good properties in
which they exist. For instance, by [6], there exist Hardy
fields with infinitely large functions that grow more slowly than
for all
.
The construction of the field of logarithmic
hyperseries in [12] was the first step towards the
incorporation of hyperlogarithms
with arbitrary
. The field
is the smallest non-trivial field of generalized power series over
that is closed under all hyperlogarithms
and infinite real power products. It turns out that
is a proper class and that
is
closed under differentiation, integration, and composition.
The main purpose of the present paper is the construction of a field of
general hyperseries that is also closed under the functional inverse
of
for every ordinal
. Our construction strongly
relies on properties of the field
of logarithmic
hyperseries. Intuitively speaking, the reason is that the derivative of
can be expressed as the composition of a
logarithmic hyperseries with
:
and similarly for all higher derivatives. One key aspect of our approach
is therefore to construct increasingly large fields
of hyperseries simultaneously with compositions
where denotes the class of positive infinitely
large elements of
.
The main result of this paper is the construction of a field of hyperseries that is closed under all hyperlogarithms
and all hyperexponentials
. Does this end our quest for a universal formal
framework for asymptotic calculus? Not quite yet. First of all, it
remains to be shown that
is closed under all
common calculus operations, such as differentiation and composition.
Secondly, the field
does not contain any
solution to the functional equation
Fortunately, it is possible to construct fields of transseries with “nested” solutions
to such equations [29, Section 2.5]. Something similar is possible for hyperseries; although this is beyond the scope of the present paper, we introduce the fundamental concepts that we expect to use for this generalization.
One may also wonder whether there exist natural models for
hyperexponential functions and hyperseries. We already noted that Kneser
constructed a real analytic solution of the
equation (1.1). Schmeling also constructed real analytic
solutions
,
,
of (1.2),
(1.3), etc. Écalle introduced a systematic technique
for the construction of quasi-analytic solutions to these and more
general iteration equations [17]. In general, it seems
unlikely that there exist any “privileged” regular solutions
at infinity.
Another interesting model for hyperexponentiation is Conway's field of surreal numbers [8]. The
field
is a non-standard extension of the field
which contains the class
of ordinal numbers. The arithmetic operations are defined in a
surprisingly “simple” way, using transfinite induction.
Nevertheless, the field
has a remarkably rich
structure; e.g. it is real-closed. The exponential function
on the reals has been extended to
by Gonshor [18] and this extension preserves the first order properties of
[11].
A “simplest” surreal solution to (1.1) with
good properties has been constructed in [3]. This solution
is only defined on , but in
view of our previous remarks on real solutions of (1.1), is
interesting to note that we may indeed consider it as “the”
privileged solution on
.
Constructing each hyperlogarithm and hyperexponential
on surreal numbers involves overcoming many technical difficulties. Our
results from this paper reduce this to the simpler task of defining
partial hyperlogarithms on
,
which satisfy a short list of axioms.
We finally note that Berarducci and Mantova also defined a derivation
with respect to
on
[4]. This derivation again has good
model theoretic properties [2]. However, although the
derivation
satisfies
for
all
, it was pointed out in
[1] that it does not satisfy
for all
, for
“reasonable” definitions of
.
Indeed, for the definition of
from [3],
we have
.
Stated differently, the hyperexponential structure on
reveals that
is not the ultimate
derivation on
with respect to
that we might hope for. One major motivation behind the work in this
paper is precisely the construction of a better derivation on
, as well as a composition. The
plan, which has been detailed in [1], is to construct an
isomorphism between
and a suitable field of
hyperseries with a natural derivation and composition with respect to
. The present paper can be
regarded as one important step in this direction.
In order to construct a field of hyperseries that is closed under all
hyperexponentials and all hyperlogarithms
, we follow the common approach of
starting with an arbitrary field of hyperseries and then closing it off
via a transfinite sequence of extensions.
Now closing off under hyperlogarithms turns out to be much easier than
closing off under hyperexponentiation. For this reason, and following
[23, 29, 12], it is actually
convenient to do this once and for all and only work with fields of
hyperseries that are already closed under all hyperlogarithms. In
particular, the smallest field of hyperseries of this type with an
element is the field of logarithmic hyperseries
from [12], where
for all
.
The next step is to work out the technical definition of a “field
of hyperseries” that will be suitable for the hyperexponential
extension process. Quite naturally, such a field should be a Hahn field
of generalized power series, where
is a totally ordered monomial group: see Subsection 2.2 for basic definitions and reminders. For reasons mentioned
in the previous subsection, we also require the existence of a
composition law
, where
. For each
, this allows us to define a function
by setting
for
.
We say that is a hyperserial field if
the composition
satisfies a list of natural
axioms such as associativity and restricted Taylor expansions; see
Section 6 for the full list of axioms. The only non-obvious
axiom for traditional fields of transseries states that a transseries
is a monomial if and only if
(i.e. the support of
only contains
infinitely large elements). The only non-obvious axiom
, then we
require that the support of
satisfies
for any
-atomic
element
. Here
is said to be
-atomic
if
is a monomial for all
.
Our definition of hyperserial fields is similar to the definition of
fields of transseries from [24, 29], with a
few differences. The old definition includes an additional axiom of
well-nestedness T4 which is important for the
definition of derivations and compositions, but which is not required
for the purposes of the present paper. Of course, our current
presentation is based on the composition law . Finally, we use a slightly different technical
notion of confluence. We say that
is
confluent if, for all ordinals
and all
, there exist a
-atomic element
and
with
.
We refer to Remark 3.6 for a discussion of the differences
with the definition from [29].
Given an ordinal , it turns
out that the hyperlogarithm
is entirely
determined by its restriction to the set of
-atomic elements. The field
together with these restricted hyperlogarithms is called the
hyperserial skeleton of
.
The fact that the logarithm can be recovered from its restriction to
is a well-known fact. Indeed, we first recover
on
,
since
and
for all
. For all
,
, and
infinitesimal
, we then have
Hyperserial skeletons can also be defined in an abstract manner,
i.e. without knowledge of a hyperserial field of which it
is the skeleton. Precise definitions will be given in Section 3;
for now it suffices to know that an abstract hyperserial skeleton is a
field of generalized power series together with
partially defined functions
that satisfy
suitable axioms. The first key result of this paper is the construction
of an exact correspondence between abstract hyperserial skeletons and
hyperserial fields. This correspondence preserves confluence for a
suitable analogue of the confluence axiom for hyperserial skeletons.
Sections 4 and 5 contain the core of this
construction. In Section 4, we first show how to extend the
partial functions of a confluent hyperserial
skeleton
to all of
.
In Section 5, we prove that any confluent hyperserial
skeleton
can be endowed with a well-behaved
composition law
. In Section
6, we complete our construction of a correspondence between
hyperserial skeletons and hyperserial fields. More precisely, we prove:
Theorem is a confluent hyperserial skeleton, then there
is a unique function
such that
is a confluent hyperserial field with
Theorem be a hyperserial field of force
. Then the skeleton
of
is a hyperserial skeleton. Moreover, if
is confluent, then so is its skeleton and
coincides with the unique composition from Theorem 1.1.
Sections 7 and 8 are devoted to the closure of
a confluent hyperserial field under hyperexponentiation. In view of
Theorem 1.1, it suffices to operate on the level of
hyperserial skeletons instead of hyperserial fields. In Section 7,
we investigate when the hyperexponential of an element in already exists in
.
This gives us a criterion under which the extended hyperlogarithms
are bijective. In Section 8, we prove
our main theorem that every confluent hyperserial skeleton has a minimal
extension whose extended hyperlogarithm functions are bijective:
Theorem be a confluent hyperserial skeleton. Then
has a confluent extension
such that the function
is bijective for each ordinal .
Moreover, if
is another confluent extension and
if the extended function
is bijective for each
, then there is a unique
embedding of
into
over
.
Corollary. There exists a minimal hyperserial
extension of that is closed under
for all ordinals
.
We work in von Neumann-Gdel-Bernays
set theory with Global Choice (NBG), which is a conservative extension
of ZFC. In this set theory, all proper classes are in bijective
correspondence with the class
of all
ordinal numbers. We will sometimes write
for
elements that are either ordinals or equal to the class
of ordinals. In that case, we write
instead of
. We make the
convention that
. If
is a successor ordinal, then we define
to be the unique ordinal with
. If
is a limit
ordinal, then we define
.
Recall that every ordinal has a unique Cantor
normal form
where ,
and
with
.
The ordinals
are called the exponents of the Cantor normal form and the integers
its coefficients. We write
(resp.
) if each
exponent
of the Cantor normal form of
satisfies
(resp.
). We also define
to be the unique ordinal with
and with
for some
.
Note that
if and only if
.
is a well-based set, i.e. a set which is
well-ordered for the reverse order .
We regard elements of
as
well-based series
where
for each
.
By [19], the class
is a field for
the operations
Note that each sum has finite support. We say
that
is a field of well-based series and that
is the monomial group of
. An element
is called a monomial.
An increasing union of fields of well-based series is not, in general, a
field of well-based series. However, this is always true if the union is
indexed over :
Lemma be a family of linearly ordered abelian groups such
that
whenever
.
Set
for each
,
so
for
.
Set
. Then
Proof. Set .
Clearly,
, so it remains to
show the other inclusion. Let
.
For each
, let
be the least
with
. Set
Then .
to be the dominant monomial of . We give
the structure
of a totally ordered field by setting
We define the asymptotic relations ,
,
, and
on
by
The monomial group is naturally embedded in
as an ordered group and
for all non-zero .
For and
,
we set
. We say that a series
is a truncation of
, denoted
, if there is
with
. We have
if and only
if
(which holds vacuously when
). We finally define
Series in are called infinitesimal and series in
are called positive
infinite. Each
can be decomposed
uniquely as
, where
,
,
and
is infinitesimal.
is a well-based set, and
is finite for all
.
Then we may define the sum of
as the series
If and
are families,
then we define their product as the family
.
By [25, Proposition 3.3], if
and
are well-based, then so is their product, and we
have
We will frequently use the following facts regarding families for
and
.
Lemma .
Let be another field of well-based series. If
is
-linear,
then we say that
is strongly linear if for every well-based family
in
𝕋, the family
in
is well-based, and
If is a function, then we say that it is
well-based if for any well-based family
in
, the family
in
is well-based. Then
extends uniquely into a strongly linear map
[25, Proposition 3.5]. Moreover,
is strictly increasing whenever
is strictly increasing and it is a ring morphism whenever
for all
[25,
Corollary 3.8]. In particular, if
for all
and
is strictly increasing,
then
is well-based. Hence:
Proposition
If extends
(so
extends
), then
the operator support of a function
is the set
. If
is a well-based set, then
is well-based; see [12, Lemma 2.9].
Definition to be relatively
well-based if
is well-based.
Proposition be relatively well-based. Assume that
and that
is strictly increasing.
Then
is well-based and its strongly linear
extension
is injective.
Proof. Given a well-based subset , we have to show that
is a
well-based family. We have
so is a well-based subset of
. For any
,
the set of pairs
with
forms a finite antichain. Since any
with
induces such a pair
,
it follows that the set of all such
is also
finite. This completes the proof that
is
well-based. To see that
is injective, let
and take
with
. The assumption that
is strictly increasing gives
,
whence
.
![]() |
(2.1) |
and for where
,
, and
, we set
.
Proposition and
we have
Proof. For ,
the first two relations follow from basic power series manipulations;
see [25, Corollary 16]. The extension to the general case
when
is straightforward and left to the reader.
Assume now that and
. Since
,
it suffices to show that
.
Write
where
,
, and
. Since
,
we have
, so either
, or
and
, or
and
. If
, then
,
so
. If
and
, then
and
. If
and
, then
, so
.
Thus, the extended real power operation gives
the structure of a multiplicative ordered
-vector space. Accordingly, we say
that
has real powers.
and elements , we say that
is defined at
if the
family
is well-based.
Lemma is uncountable and let
be a power series which is defined on
.
If
, then
for some
.
Proof. We prove this by induction on . If
,
then set
and write
.
Suppose that
and let
be
the set of
with
.
Fix
and let
be such that
for all
.
Since
, there exists an index
with
.
Then
, whence
In particular, is countable, whereas
is uncountable, so
.
Now suppose that and write
. Assume that
.
By the induction hypothesis, we can find
and
such that
.
Fix such elements
and let
be the set of
such that
. By the special case when
, we see that
.
Thus,
for some
.
A central object in our work is the field
of logarithmic hyperseries of [12],
equipped with its natural derivation
and
composition
. We briefly
recall its definition and some of its properties.
We call the field of logarithmic
hyperseries of strength
.
If
are ordinals with
, then we let
denote the
interval
and we let
denote the subgroup
As in [12], we write ,
and
We will sometimes write and
. We have natural inclusions
, which give natural inclusions
.
where . Note that
. For
and
, we sometimes write
. Equipped with its
derivation, the field
is an
-field with small derivation, so for
, we have
Moreover, is well-based, which implies the
following variant of [12, Lemma 2.13]:
Lemma , let
be a field of well-based series, and let
be a
strongly linear field embedding. For
and
with
, the
family
is well-based. Moreover, the map
is a strongly linear field embedding.
Proof. Since is well-based and
is a strongly linear field embedding, the set
is well-based. Thus
is
well-based and
, since
. Let
.
For each
, we have
Since is well-based and
, it follows that
is
well-based and that the map
is well-defined and
strongly linear. For all
, we
also have
which shows that preserves multiplication.
For , the map
is a strongly linear field embedding. As a consequence
this map preserves the relations
and
[12, Lemma 6.6].
For and
,
we have
[12, Proposition 7.14].
For and
,
we have
[12, Corollary 7.5].
For and
,
we have
and
[12, Section 1.4].
For ordinals , we have
[12, Corollary 5.11].
For any successor ordinal ,
we have
[12, Lemma 5.8].
The constant term of vanishes if
is a limit ordinal [12, Lemma 5.8].
For and
with
, the family
is well-based, and
![]() |
[12, Proposition 8.1] |
The uniqueness follows from [12, Theorem 1.3]. By [12,
Proposition 7.8], the derivation also satisfies the chain rule: for all
and
,
we have
As we will see, the field equipped with the
composition
is hyperserial.
For , the unique composition
from above restricts to a composition
. For
,
the map
defined by
is a
strongly linear field embedding with image
by
[12, Lemma 5.13]. Accordingly, for
, we let
denote
the unique series in
with
. Note that
for all
and that, more generally,
for
and
.
For
and
we have
![]() |
(2.2) |
where is the derivation
on
(see [12, Section 5.1]). Let
. Then
is strongly linear and
, so,
by [12, Lemma 2.2],
![]() |
(2.3) |
In particular,
![]() |
(2.4) |
Lemma , each
, and each
,
we have
.
Proof. Since is closed under
taking derivatives and the derivation preserves infinitesimals, it
suffices to prove the lemma for
.
We have
, so
Since and
,
this yields
. Since
as well, we have
.
Since the map
maps
onto
, we conclude that
.
We let be an ordered field of well-based series
with real powers. Let
be an ordinal with
. Given a structure
where
are partial
functions on
, we consider
the following axioms for
:
Suppose satisfies all axioms
. We set
for all
and we extend this notation to the case
when
, by setting
![]() |
(3.1) |
For , we call
the class of
-atomic elements. Note that
for all
. We let
be
the identity function with
and, for
with Cantor normal form
,
we define
. Here we
understand that
whenever
,
, and
so on until
.
Proposition with
,
we have
Proof. Given and
, let us first show by induction on
that
is defined and in
. This holds for
by definition. Let
and assume that the assertion
holds strictly below
. If
, then
. Assume
and let
,
and
be such that
.
We have
so
by
definition. In particular
,
so our inductive hypothesis on
applied to
gives that
is a monomial.
Given such that
and
for all
,
let us next show by induction that
.
This is clear if
. Let
be such that the statement holds strictly below
. If
is a
successor, then for
and
, we have
so for all
, whence
. Assume now that
is a limit
and let
. Then
for all
, so
the induction hypothesis yields
.
We again conclude that
.
Let be an ordered field of well-based series
with real powers, let
, and
let
be partial functions
on
which satisfy the axioms
. We consider the following
axioms, where
is an ordinal with
.
We define a logarithm as follows:
![]() |
(3.2) |
Then is an ordered
-vector
subspace of
. For
with
, we
consider the following axiom:
Remark for
to be the unique monomial
with
, hence the name. Note
that the axiom
is a consequence of
and
, we have
.
Definition . A hyperserial
skeleton of force
is a structure
where
is an ordered
field of well-based series with real powers and
are partial functions on
which satisfy
, as well as
with
.
Note that a hyperserial skeleton of force 0 is just a field of
well-based series with real powers and that is a
hyperserial skeleton of force
if and only if
is a hyperserial skeleton of force
for each ordinal
.
We will often write
to denote a hyperserial
skeleton (of force
), where
it is implied that for
, the
term
refers to the
-th
hyperlogarithm on
.
Definition and
be hyperserial skeletons of
force
. We say that a
function
is an
embedding of force
if it is a strongly linear strictly increasing ring morphism with
for each
such that
and such that
If is an embedding of force
, then we say that
is
an extension of
of
force
.
In this subsection, let be a hyperserial
skeleton of force
and let
with
. We inductively define
the notion of
-confluence in conjunction with functions
and the classes
, as follows:
Definition is said to be
-confluent if
is non-trivial. The function
maps each
to its dominant monomial
.
For each
, we set
Let with
,
let
, and suppose
is
-confluent
for all
.
If is a successor, then we define
to be the class of series
with
for some .
If is a limit, then we define
to be the class of series
with
for some .
If each class contains an
-atomic element, then we say that
is
-confluent.
We will see that each class
contains at most one
-atomic element, which we
denote by
.
Remark -confluence is somewhat
stronger than the similar notion of
-confluence
from [29], due to the extra requirement that we have maps
.
Lemma with
and suppose
is
-confluent.
Then the function
is well-defined. Moreover, we
have
for all
and
.
Proof. We prove this by induction on , noticing that the case
is trivial. Assume that this is the case for all ordinals
and let
. To
see that
is well-defined, let
with
. We need to show
.
Assume that is a successor. Take
with
Since
is
-atomic for each
and since
is well-defined by our
induction hypothesis, we have
for each
. It follows by induction on
that
for each
and, likewise,
for each
, so
.
As both
and
are
monomials, we have
. The
axiom
implies that
is
injective and thus
.
Assume now that is a limit and take
with
. Since
is
-atomic
and since
is well-defined by our induction
hypothesis, we have
.
Likewise,
, so
. As both
and
are monomials, we have
.
Since
is injective by
, we conclude that
.
As to our second assertion, consider and
with
. If
is a successor, then the inductive hypothesis
implies
,
so
and
.
If
is a limit, then
, so
and
.
Corollary with
. If
is
-confluent,
then
for all
.
Proposition with
.
If
is
-confluent
for all
, then the class
is convex for each
.
Moreover, if
is
-confluent,
then
is non-decreasing.
Proof. We prove this by induction on . Let
.
It is clear that
is convex and that
is increasing. Let
and assume that
the result holds for all
. By
the monotonicity axioms, each function
is
strictly increasing on
(when
, one also needs to use
for
).
As the composition of non-decreasing functions is non-decreasing, the
function
is non-decreasing for each
and each
. We
deduce that
is non-decreasing and that the
classes
are convex.
If is
-confluent
for all
, then the
proposition implies that the classes
with
form a partition of
into
convex subclasses. If
is also
-confluent, then we have the following explicit
decomposition for all
:
Definition is said to be confluent if it is
-confluent
for each
with
.
An extension/embedding
of force
is confluent if
is confluent.
Note that if , then 𝕋
is confluent if and only if it is
-confluent.
Let be an ordinal and set
. The goal of this section is to check that
is a confluent hyperserial skeleton of force
. This is immediate for
, so we assume that
.
Definition and for
, let
. Given
, set
We will show that is a hyperserial skeleton by
checking that the axioms are satisfied. We begin with the domain of
definition axioms.
Lemma satisfies
,
for all
.
Proof. We prove this by induction on . The case when
is
immediate. For
, consider an
infinite monomial
. We have
, which is a monomial if and
only
for some
.
Conversely, for each
we have
. Now let
and suppose
that the lemma holds for all non-zero ordinals less than
. Assume that
is a
limit. We have
, whence
Assume now that is a successor. If
, then
where
, and we clearly have
for all
,
whence
. Conversely, let
. Then
where
. If
is a limit, then
, whence
and
.
If
is a successor, then
for some
and some
,
so
Since , we see that
Since , we must have
, so
.
For and
,
note that
. Note also that
the notions of
-atomicity and
-atomicity coincide in
whenever
is a limit with
. This will not be the case
in general.
Proposition satisfies
.
Proof. Let and let
. By Remark 3.2, we may assume
. We have
for some
. Let
be a sequence of real numbers. We have
This sum coincides with where
.
Proposition satisfies
.
Proof. Let and let
. We have
for some
. Write
where
,
, and
if
is a limit. We claim
.
If
is a successor, then since
, we have
If is a limit, then
, so
Now we move on to verification of are
and possibly 1 (if
), so
for
all
and
,
which proves
, we have
so
As to with
.
We have
for some
with
. Write
where
,
, and
if
is a limit. The argument above gives
. If
,
then
and if
,
then
and
.
In either case,
Recall that for and
, we write
for the real
exponent of
in
.
Given
, we define
to be the least ordinal with
;
see also [12, p. 23].
Proposition is
-confluent.
More precisely, for
and
, we have
![]() |
(3.3) |
Proof. We first note that is
-confluent as
is not trivial. We proceed by induction on
. Take
.
If
, then we have
where
is
-atomic, so
and
is
-confluent.
It remains to note that
.
Now suppose that and assume that
is
-confluent
and satisfies (3.3) for all
.
Suppose
is a successor, so
. Write
with
and with
if
is a limit. We have
, so
and .
Now suppose is a limit, so there is
with
. By
hypothesis, we have that
and so
Again, this yields .
Theorem is a confluent hyperserial skeleton of force
.
Proof. Using the identity
for , the field
is easily seen to satisfy
satisfies
by [12, Lemma 5.6]. Using Propositions 3.13,
3.14 and 3.15, we conclude that
is a confluent hyperserial skeleton of force
.
Corollary is a confluent hyperserial skeleton of force
.
Let . The purpose of the next
two sections is to prove the following theorem:
Theorem be a confluent hyperserial skeleton of force
. There is a unique function
satisfying:
We claim that it suffices to prove the theorem in the case when . The case when
can then be proved as follows: let
be a
confluent hyperserial skeleton of force
, there exists
a unique composition
that satisfies
, the composition
extends
,
by uniqueness. For any
and
, we have
for some
, so we may define
and this definition does not depend on
. It is straightforward to check that this
defines the unique composition
which satisfies
C1
,
C2
,
C3
, and
C4
.
Throughout this section, we fix an ordinal and a
hyperserial skeleton
of force
. We fix also
such that
is
-confluent
and we set
. We assume that
Theorem 4.1 holds for
,
so we have a unique composition
satisfying
,
,
, and
. For
and
, we write
.
In light of Lemma 2.9, the expression
makes sense for each
. Our
main goal is to prove the following result:
Proposition to
such that for all
,
, and
with
, we have
We will also prove that satisfies the extension
of
(Proposition 4.13), that
has Taylor expansions around every point (Proposition
4.15) and that it is strictly increasing on
(Lemma 4.17).
Our extension will heavily depend on Taylor series expansions, so it is
convenient to introduce some notation for that. Let
be such that
for all
. Let
and
with
. By Lemma 2.8
,
in place of
,
and
, we have that the family
is well-based. We define
and
Note that and
.
By [25, Corollary 16], we have
![]() |
(4.1) |
![]() |
(4.2) |
![]() |
(4.3) |
Proposition into an ordered group
embedding
with
for all
,
and
extends the natural logarithm on
.
For , writing
for
,
, and
,
we have
Proof. The uniqueness and the fact that is a morphism are proven in [29, Example
2.1.3 and Lemma 2.1.4]. To see that
is
order-preserving, we need only check that
for
all
. If
, then
and we have
. If
and
, then we have
, since
. If
, we
have
, so
.
For , we often write
in place of
.
For
, we define
to be the unique element of
with
. If
, then we sometimes use
instead of
.
Proposition , we have
.
Proof. Let ,
where
,
, and
.
If
, then
, so
.
If
, then
. If
,
then
, since
,
,
and
. Thus
If , then
and
. Hence
and
. If
, then
is negative and
infinite, so
.
Remark satisfies the properties of
transseries fields in [29, Definition 2.2.1] except
possibly for the axiom
.
Proposition , and
, we have
.
Proof. First, note that for all
: if
, then this is just axiom
, then
; if
,
then
. Now, writing
with
,
, and
,
we have
![]() |
![]() |
![]() |
|
![]() |
![]() |
(by (4.3)) | |
![]() |
![]() |
(by (![]() |
|
![]() |
![]() |
![]() |
Proposition and
with
, the family
is
well-based, with
Proof. For and
, we have
and
. For
,
we have
So the family is well-based with
. The proposition follows, as
Assume now that . Let us
revisit the notion of confluence.
Lemma and suppose that
for some
. Then
for all
with
.
Proof. We first show that .
Take
and
such that
. We have
where . Set
, so
.
We have
where . Thus,
.
Now, fix with
and set
. By
and
, we have
Lemma 2.9 in conjunction with the fact that gives us that
,
so
.
Proposition , we have
Proof. We fix .
Since
, we know by Lemma 4.8 that it is enough to show that
We
proceed by induction on
. If
, then
and
An easy induction on yields
for each
, whence the result.
Now suppose that . If
is a successor, then for each
there is some
with
.
By our inductive assumption applied to
,
we have that
for some
. By Lemma 4.8, we have
and an easy induction on
gives us
that
. Thus, we have that
for some
.
Likewise,
for some
.
By replacing
and
with
and invoking Lemma 4.8, we may
assume that
. Since
, we have
. On the other hand, given
, if
for some
, then take some
with
. By Lemma 4.8,
we have
, so
.
If is a limit, then for each
there is
with
.
By our inductive assumption applied to
,
we have that
for some
, so
by Lemma 4.8.
Thus
and likewise,
for
some
. By replacing
and
with
and invoking Lemma 4.8, we may assume that
. Since
,
we have
. On the other hand,
given
, if
for some
, then take some
with
.
By Lemma 4.8, we have
,
so
.
Proposition 4.9 in conjunction with Lemma 4.8 gives us the following corollary:
Corollary there is
such
that
for all . Moreover, if
for some
and some
, then
.
Definition and let
with
. We define
As discussed at the beginning of the section, the series exists in
by Lemmas 2.8
and 2.9. To prove Proposition 4.2 all that
remains is to show:
Lemma .
Proof. Let ,
,
be
as in Definition 4.11 and suppose that
for some
. Set
. We need to show that
Without loss of generality, we may assume that . Now
Since , this yields
. Let
considered as formal power series and
in
Then
so it suffices to show that .
For each
, we have
so for all
,
and we conclude that
by Lemma 2.7.
We end this section with various extensions of the validity of Taylor
series expansions and a proof that is strictly
increasing on
.
Proposition is a successor. For
, we have
.
Proof. By Corollary 4.10, there is some
such that
.
We may write
Note that is
-atomic,
so
. For
we have
so for all
.
It follows that
This concludes the proof.
Lemma ,
, and
,
we have
Proof. By applying to
for
, we have
Arguing as in the proof of Lemma 4.12, it is enough to show that
as power series in . Let
. We have
Therefore,
Since , we are done by Lemma
2.7.
Proposition and
with
, we have
Proof. Let with
. Set
.
By Lemmas 2.8 and 2.9, the series
exists in
. We
claim that
. As
by
, it
suffices to show that
as power series in . But this
follows from Lemma 2.7, by noting that the equality holds
when evaluated at any element of
.
Now set and let
,
so
. By definition of
and the above claim, we have
Using , it follows that
By Lemma 4.14, we conclude that
where the last equality follows from the definition of .
Lemma , let
, and let
with
Then where
.
Proof. Set ,
so
. We have
where the first and third equalities follow from the definition of and where the second equality holds by Lemma 4.14.
Lemma is strictly increasing on
.
Proof. By induction on ,
we may assume that
is strictly increasing on
for all
(the
case follows from Proposition 4.3). As a
composition of strictly increasing functions is strictly increasing, the
function
is strictly increasing on
for all
. Given
, let us show that
. We start with the case when
and take
with
and
. Then
is infinitesimal and positive by our induction
hypothesis. By Lemma 4.16, we have
Since , we have
, so
.
Now we turn to the case when .
Set
and
and take an
ordinal
with
Set , so
Repeated applications of (2.4) with
in place of
gives
,
so
and
Since , we have
so
. Thus,
. All together, this shows
that
. Likewise, we have
. By the monotonicity axiom
, so
.
Throughout this section, stands for a fixed
ordinal and
for a fixed confluent hyperserial
skeleton of force
. We let
. Our aim is to construct a
well-behaved external composition
that satisfies
has relatively well-based support for all
. Throughout the section, we make
the inductive assumption that Theorem 4.1 holds for all
and that the mapping
has
relatively well-based support for all
and
.
Here is a
-confluent
hyperserial skeleton of force
.
The field
is the field of well-based series of
real powers of the variable
,
with real coefficients.
Lemma is well-based, then
is
well-based for all
.
Proof. Let with
,
and
. Note that
.
Since
and
are both
well-based, we are done.
Given and
,
the family
is well-based by the above lemma, so
we may define
One easily verifies that this composition satisfies
and
. We next prove
Proposition ,
, and
.
We have
.
Proof. Write where
,
,
and
. We have
, so
.
We also have
Since by
,
we only need to show that
.
Now
for some
,
so
by Proposition 2.6.
Corollary ,
, and
.
We have
.
Proof. We have
where the second equality follows from Proposition 5.2 and
the third follows from strong linearity of composition with .
Proposition ,
and
with
,
we have
.
Proof. We first handle the case when , where
.
We have
, so
For , we also have
, so
.
Therefore,
Now, Lemma 2.8 gives that the the map
is well-based and strongly linear, so for a general
, we have
The above results show that satisfies
, and
.
To complete the proof of Theorem 4.1 for
, it remains to show uniqueness.
Proposition is unique to satisfy
,
,
, and
Proof. Let be a composition
satisfying conditions
,
,
,
and
Write
,
where
,
, and
.
By strong linearity, it suffices to verify that
for any monomial in
. Given
, the condition
implies
We have , so
We have by
and
by
, so
. Likewise,
, so
For the remainder of this section, we assume that . By the results in Section 4, we have
a well-defined extension of
to all of
for each
.
Indeed, for
and
,
take
with
with
(so
if
is
a limit). Then we may set
.
Given and
,
we have by
that
,
so we set
. Clearly, the map
is an embedding of monomial groups which
preserves real powers, and by
,
this embedding is order-preserving as well. For
, we set
.
By Proposition 2.3, we have:
Lemma is a strongly linear ordered field
embedding.
By Lemma 2.8, the series
exists in
for any
.
Lemma ,
, and
.
We have
Proof. If is a limit ordinal,
then the lemma follows from
for any ordinal
with
, so
we may assume that
is a successor. We prove the
lemma by induction on
. The
lemma is immediate when
, so
suppose
and take
and
with
.
Our induction hypothesis yields
Note that and
.
We claim that
When , this just follows from
. When
, this is Proposition 4.7 with
in place of
and
in place of
.
When
, this is Definition 4.11 with
in place of
and
in place of
.
Since
and
,
it remains to show that
. Now
as series in
.
Indeed, for
, we have
We conclude by Lemma 2.7.
Lemma 5.6 shows that . In the general situation when
, our next goal is to show that the family
is well-based. For the remainder of this subsection,
we fix
. By
-confluence and Corollary 4.10,
take
and
such that
for all
.
If
is a successor, we can arrange that
. Set
,
set
, and set
.
Lemma and let
.
If
is a successor or
, then the expression
is
defined and equal to
.
Proof. Suppose is a successor
ordinal, so
. Then
, so
is
defined. As the maps
and
are strongly linear, we may assume that
is a
monomial
. Since
for
, we have
Now suppose that is a limit and that
for some
. By
increasing
, we may assume
that
, so
. Then
and
give
Lemma from
such that
for each with
.
Proof. Fix with
. We first claim that
If is a limit, then take
with
. Then
, so
gives
and gives
,
thereby proving the claim. If
is a successor,
then take
with
.
By Lemma 5.7 and the fact that
, we have
Having proved our claim, let be given and set
. Lemma 2.9
yields
, whence
. If
is a limit, then
, where
is as above. So in both the successor and limit cases, we may apply
Lemma 5.8 with
in place of
to get
This gives
It remains to show that the family is
well-based. Since
is a well-based family in
and
is strongly linear, the
family
is well-based. Since
is well-based and infinitesimal, the family
is
well-based. We conclude that the family
is
well-based.
Proposition be a sequence of real numbers. Then the family
is well-based and the series
lies in
.
Proof. We will show the following:
For each , the family
is well-based and
The family is well-based and
The proposition follows from (a) and (b), since the union of finitely
many well-based families is well-based and is
closed under finite sums.
To see why (a) holds, let and note that
Since is well-based,
gives that
is well-based. We have
Set . We claim that
. If
,
then
and
If , then
gives
As for (b), let . By Lemma 5.9, there exists a well-based family
from
such that
The families and
are
well-based by Lemma 5.6 and the fact that
. Since the family
is
also well-based, it follows that
is again
well-based. In particular,
is well-based. Now
Since and
are
infinitesimal for all
, we
may write
where By (4.1), we have
. Furthermore, P
implies
We conclude that .
Let . In light of Proposition
5.10, we define
We note that the map is an embedding of ordered
multiplicative groups for each
.
Our next objective is to show that the map
extends by strong linearity to a map
which
satisfies
is a relatively
well-based mapping, by using a similar “gluing” technique as
for Proposition 5.10. Recall that our second induction
hypothesis from the beginning of this section stipulated that the
mapping
is relatively well-based for all
and
.
Proposition be the map
.
Then
is relatively well-based.
Proof. Let be the restriction
of
to
and for
, let
be
the restriction of
to
. Since
it suffices to show that each and
are relatively well-based. For the
, fix
.
Our induction hypothesis implies that the map
is
relatively well-based. By Lemma 5.8 with
in place of
, we have
It follows that is also relatively well-based
with
.
Now for . Let
. By Lemma 5.9, we have a
well-based family
from
such that
Exponentiating both sides, we obtain
so . The set
is well-based, infinitesimal and does not depend on . Since
for all , we conclude that
is well-based.
We already noted that the map from Proposition
5.11 is an order-preserving multiplicative embedding. By
Proposition 2.5, it follows that
is
well-based, so it extends uniquely into an order-preserving and strongly
linear embedding
. Taking
for all
,
this proves
extends the unique composition
of Theorem 4.1 for
.
Proposition and
,
we have
.
Proof. As in Proposition 5.2, it suffices
to prove that holds for each
. For such
,
we have
By Proposition 4.6, we also have . By injectivity of the logarithm, we conclude that
.
Lemma and all
, we have
.
Proof. First, we note that for , we have
where the last equality uses the definition of . Now, let
and write
with
,
, and
. Then
and
Here we used the facts that and that composition
with
commutes with powers and infinite
sums.
Proposition satisfies
,
, and
we have
.
Proof. We will show by induction on
that
for all
,
all
, and all
. If
,
then this follows from Proposition 5.12 and strong
linearity.
Let , let
and
be fixed, and assume that the proposition
holds whenever
for some
. By strong linearity, it suffices to prove that
for all
.
Lemma 5.13 gives
Using the injectivity of and strong linearity,
we may thus reduce to the case when
for
. Our induction hypothesis takes
care of the case when
is a limit ordinal or when
, so we may assume that
, where
. By the inductive definitions of
and
, we may further reduce
to the case when
. Lemma 5.13 takes care of the case
,
so we may assume that
. In
summary, we thus need to show that
,
where
.
Set . We claim that
. We have
, where
,
, and
if
is a limit ordinal. As
, we have
This gives
where the first equality in the second line follows from Proposition 4.13.
Having proved our claim, let us now show that . Take
with
, and
.
We have
As and
for all
, by Lemma 2.9, our
induction hypothesis applied to
gives
for . Along with
Using also our claim that ,
we obtain
It remains to show that . Now
so and
.
We may thus apply Lemma 4.16 to
and
, to conclude that
.
Proposition ,
and
with
,
we have
Proof. Since holds for
, we need only consider the case
when
is a successor and
. We prove the result by induction on
. By Proposition 4.7 (when
) or Proposition 4.15
(when
), we have
. Assume that
and write
with
.
We have
Since , the induction
hypothesis yields
Since , it remains to show
that
. It is enough to show
that
as power series in
. This follows from Lemma 2.7, since
for all .
Proposition satisfies
, all
and all
with
, we
have
Proof. Fix and
with
. Let
be the map given by
We need to show that for all
. By Lemma 2.8, the map
is strongly linear, so it suffices to show that
for all
. Since
is injective, it is enough to show that
. Now
by
Lemma 5.13 and
by [12,
Lemma 8.3]. By Proposition 5.15 and strong linearity, we
have
We conclude that .
To conclude our proof of Theorem 4.1, we prove the
uniqueness of .
Proposition is unique to satisfy
Proof. Let be a composition
satisfying conditions
. We first show that
. Write
, with
,
, and
. By
For , we have
, so
Thus, it remains to show that .
Using
, we see
that
Likewise .
Now we turn to the task of showing that for
. We make the inductive
assumption that for
and
, we have
(if
, this is Proposition 5.5). By
strong linearity, it suffices to verify that
for
any monomial
. As
and likewise for
,
it suffices to show this only for
.
Given
, we have by
Thus, it suffices to show that for all
. By our induction hypothesis, we
only need to handle the case that
is a successor
and
. If
, then by Proposition 4.9, there
is an ordinal
with
.
Our inductive hypothesis and Lemma 2.9 yield
![]() |
![]() |
![]() |
|
![]() |
![]() |
![]() |
(for ![]() |
Thus,
![]() |
![]() |
![]() |
|
![]() |
![]() |
(by ![]() |
|
![]() |
![]() |
||
![]() |
![]() |
||
![]() |
![]() |
Now suppose and assume by induction that
for all
.
Take
with
.
Then
This concludes the proof.
We are now in a position to prove Theorems 1.1 and 1.2.
Let be a field of well-based series, let
, and let
be a function. For
and
, we define
as follows: set
, set
if
, and set
if
. For
with
, we define
to be the class of series
with
for all
.
We say that
is a hyperserial field of
force
if the following axioms
are satisfied:
for all
,
, and
.
Axioms , so they are assumed to hold trivially when
. We say that
is confluent if
and if for all
with
and
all
, there exist
and
with
For the remainder of this section, we fix a hyperserial field of force
. For
each
, we define the function
. The skeleton of
is defined to be the structure
equipped with the real power operation on
given
by
Theorem of
is a
hyperserial skeleton. Moreover, if
is confluent,
then so is its skeleton and
coincides with the
unique composition from Theorem 4.1.
When , then the skeleton of
is just the field
itself
with the real power operation on
.
Clearly, this is a hyperserial skeleton, as there are no axioms to
verify. Moreover, it is 0-confluent so long as
is, so Theorem 6.1 follows from Proposition 5.5,
since
clearly satisfies
,
,
, and
.
Therefore, we may assume that
.
We will verify the various hyperserial skeleton axioms over the next few
lemmas, beginning with the Domain of Definition axioms:
Lemma satisfies the axioms
.
Proof. By definition, is the
class of
with
.
Since
by
and let us assume that
holds for all
.
If
is a limit, then
Suppose is a successor. The inclusion
holds by definition, so we show the other inclusion. Let
and let
.
Take
and
with
. Then
, so
, by
our inductive assumption. Repeated applications of
.
Since
is arbitrary, this gives
.
Now for the functional equations, asymptotics, regularity, and monotonicity axioms:
Lemma satisfies the axioms
.
Proof. Given and
, we have
so be a successor ordinal and let
, so
is defined and
lies in
. The axiom
so in
and
for all
. By
and
are mutually cofinal for each
.
The regularity axioms
therefore follow from
Lemma satisfies the axioms
.
Proof. The axiom .
For
, let
and take
with
.
We need to show
We first consider the case that for some
with
.
Then
. By (2.4), we have
, where
and
. Since
, we have
Since , we have
, so
.
The axiom
, so
.
Thus,
Now we handle the case that for all
with
. We claim
that the sets
are mutually cofinal. Let be given and take
with
and
. Then
by assumption,
so
and
. This proves the
cofinality claim. Now,
and likewise,
. Thus,
In particular, , as
desired.
Before proving the infinite powers axioms, we need a lemma:
Lemma with
,
, and
. Then
where is as defined in Subsection 4.1.
Proof. Set ,
so
and
.
The axiom
We have by
. Hence
Given , we have
, so for
,
we have
Thus, .
Lemma satisfies the axioms
with
.
Proof. let with
, let
and let
be a sequence of real numbers. We need to show that
, where
for
and where
.
Set
. We may assume that
and, by negating each
if need
be, we further assume that
.
Hence
is defined. The axioms
so it remains to show that .
For each
, we have
. This gives
. Take
and
with
. Lemma 6.5
gives
The axiom . If
,
then
, so
. If
,
we have
. As we have
established that
, it follows
that
and
.
Thus
, as desired.
This shows that is a hyperserial skeleton of
force
. Now we turn to
confluence. First, we need a lemma:
Lemma and let
.
If
, then
and
for all
with
. In particular,
for all
with
.
Proof. The proof is essentially the same as the proof
of Lemma 4.8. Take and
with
. By Lemma
6.5, we have
so . Set
, so
.
Again, Lemma 6.5 gives
so . Now set
and fix
with
.
We have
Since , we have
.
Lemma is confluent. Then
is
confluent as well.
Proof. The skeleton is
-confluent since
is non-trivial. Let
with
and assume that
is
-confluent
for all
. We also make the
inductive assumption that for
and
, we have
for some
. Let
and take
and
with
. We will show that
. We first handle the case that
is a successor. Take
with
. Lemma 6.7 gives
. By assumption, we have
for some
,
so
, again by Lemma 6.7.
Induction on
gives
for
all
, so
and . The case that
is a limit is similar, though this time we take
with
and use that
to see that . Since
was arbitrary, this gives that
is
-confluent.
Proof of Theorem 6.1. Lemmas 6.2,
6.3, 6.4, and 6.6 show that of
is a hyperserial skeleton.
The composition
clearly satisfies
is confluent, then
is
confluent by Lemma 6.8 and Proposition 5.17
implies that
coincides with the unique
composition from Theorem 4.1.
Given a confluent hyperserial skeleton of force
, it is clear that the unique
composition
in Theorem 4.1
satisfies all of the hyperserial field axioms except for possibly
Lemma be a confluent hyperserial skeleton of force
and let
be the composition
law established in Theorem 4.1. Then the function
is strictly increasing for all ordinals
.
Thus, the unique composition in Theorem 4.1 satisfies
Theorem is a confluent hyperserial skeleton of force
, then there is a unique
function
such that
is a
confluent hyperserial field of force
with
skeleton
.
Our goal for Sections 7 and 8 is to prove
Theorem 1.3. We will actually prove a
“relative” version of the theorem, for which we first need a
few more definitions. Given a confluent hyperserial skeleton of force
, we
let
be the composition from Theorem 4.1.
For each
, we let
be the map given by
.
Definition be a confluent hyperserial skeleton of force
and let
.
We say that
has force
if for each
, the function
is
bijective.
Note that if has force
, then
is bijective for all
.
Remark is a confluent
hyperserial skeleton of force
.
Given a set-sized field of transseries
,
we recall that the exponential function cannot be total [23,
Proposition 2.2]. Thus, any confluent hyperserial skeleton of force
with
is necessarily a proper
class.
Remark
be a hyperserial skeleton of force
.
Then
is hyperserial of force
if and only if
is hyperserial of force
for all
.
Similarly,
is hyperserial of force
if and only if
is hyperserial of
force
for all
.
We can now state the relative version of Theorem 1.3 that we are after:
Theorem be a confluent hyperserial skeleton of force
and let
.
Then
has a confluent extension
of force
with the following property:
if
is another confluent hyperserial skeleton of
force
and if
is an
embedding of force
, then
there is a unique embedding
of force
that extends
.
Theorem 1.3 follows from Theorem 7.4 by taking
. Throughout Section 7
and Subsections 8.1, 8.2, 8.3,
and 8.4, we fix a confluent hyperserial skeleton
of force
and an ordinal
, and we make the induction
hypothesis that Theorem 7.4 holds for
. Note that this holds trivially if
. Our main objective is to show
that Theorem 7.4 still holds for
instead of
.
For this, we have to show how to define missing hyperexponentials of the
form with
.
In this section, we start by giving a formula for hyperexponentials
that are already defined in
. In the next section, we show how to adjoin the
missing hyperexponentials to
.
Before we continue, let us fix some notation. Let
Given , we set
Note that and that
.
Given
, we set
and we view ,
, and
as functions from
to
.
We define
and
analogously.
Given , we say that
is defined if
.
If
is of force
,
then
is defined for all
and
. Lemma 4.17
tells us that
is strictly increasing; in
particular, it is injective. We let
be its
functional inverse, which is again strictly increasing. We may also
consider
as a partially defined function on
.
Our induction hypothesis, that exists, has the
following consequence:
Lemma , the function
is strictly increasing on
.
Proof. Let with
. By our inductive assumption,
and
both exist in
.
As
and
are strictly
increasing on
and
,
we have
and
In this subsection, we study the range of the functions
for
and give a formula for their partial
functional inverses. We fix
and set
. We also fix
.
For
, we define series
inductively by
Intuitively speaking, behaves like
, whereas the sum
behaves like
for
,
and thereby provides a functional inverse of
on
a neighborhood of
.
Proposition with
.
Then the family
is well-based and
for
.
Proof. Consider the derivative
on
. We claim that
for all
. This
is clear for
. Assuming that
the claim holds for a given
,
we have
In light of this claim, we have .
Recall that
has well-based operator support
as an operator on
,
so
Consider the strongly linear map
and set
so is well-based and
. For
,
we have
, so for
, there exist
with
This gives us
and it follows that
As , we have
, so we deduce that
is
well-based and that
for
.
For our next result, we need a combinatorial lemma for power series over
a differential field. Let be a differential
field. Then the ring
is naturally equipped with
two derivations:
We also have a composition given by
for . This composition
cooperates with our derivations as follows:
Lemma and
.
Write
and assume that we have
for each and
,
where
. Then
.
Proof. The last two assumptions give us the following identities
We claim that . Indeed, we
have
Write . The identity
yields
for each
. Since
,
we conclude that
and
for
.
![]() |
(7.3) |
Proof. We have
Consider the formal power series
Writing , we have
Thus, it suffices to show that .
Let and
.
Then by factoring out
from the inner sum and
reindexing, we have
Note that the sequence satisfies the identities:
Since , the sequence
satisfies the identities
Setting , we have
Using Lemma 7.7, we conclude that .
Proposition is a bijection from
to
.
Proof. Let and let
. We have
,
so
Since , we have
so . This gives
.
Conversely, given , Lemma 7.8 yields
. Let us
show by induction on
that
. We have
.
Assuming that
, we have
We have
so . This gives
It follows that for each
, so
.
Since
, we conclude that
.
Definition is
-truncated
if it is purely infinite, i.e. if
. For
,
we say that
is
-truncated if
for all
and
all
. Let
denote the class of
-truncated
series in
.
In Subsection 3.3, we showed for
that the class
can be partitioned into convex
subclasses
, each of which
contains a unique
-atomic
element
. In this section, we
describe a different partition of
into convex
subclasses, each of which will contain a unique
-truncated series
.
We will then show that
is bijective provided
that
. This was done in [29] for
, but we
provide a short proof below. First, for
,
set
, so
is 1-truncated and
. We also
set
.
Proposition ,
we have
if and only if
. Thus, the function
is
bijective if and only if
.
Proof. Let and let
, and
with
. We have
,
since
. Thus, we have
if and only if
,
since
is an additive subgroup of
.
As a related fact for later use, we also note the following:
Lemma , we have
. Thus,
if and only if
. Moreover,
is a bijection between
and
for each
.
Proof. Given ,
write
, where
,
,
and
. We have
and, as
is a subgroup of
, it follows that
.
Thus,
is 1-truncated. If
, then
and if
, then
,
so
. Thus,
, as desired. The fact that
if and only if
follows from this and the fact
that
is injective. Now assume that
and let
. Then
so . By Proposition 7.11,
.
For the remainder of this subsection, we assume that .
Lemma . If
is a successor, then
.
Proof. For and
, we have
and
so
. Assume now
that
is a successor and let
and
. Again,
. Take
with
. Then for all
and
, we have
so .
Lemma and let
.
Then
is
-truncated
if and only if
for all
.
Proof. We have for all
since the series
is infinite.
Let
and let
.
By Lemma 7.5, the function
is
strictly increasing, so we have
if and only if
, hence the result.
By Lemma 7.14 and
is
-truncated for all
. The axiom
is
-truncated
for
.
Lemma with
and let
. Then
.
Proof. Take with
. Then Lemma 7.5 gives
, so it is enough to prove that
. For this, we may show that
in
.
As the map
is order-preserving, it is enough to
show that
This follows from Lemma 7.5 and the fact that .
Definition , we define
Proposition form a partition of
into convex subclasses.
Proof. Let .
The convexity of
follows immediately from the
definition of
and Lemma 7.5. Let
. We claim that
, from which it follows by symmetry that
. This clearly holds if
, so assume that
.
We first show that . Let
and let
with
for some
. Given
with
, we have
, whence
. Therefore,
by Lemma 7.5, so for all such
.
If is a successor, take
with
. Then
and since
, we have
If is a limit, take
with
, so that
. Let us show that
. Suppose for contradiction that
. By (2.4), we have
Since , we have
, so
Therefore,
This means that : a
contradiction since
.
Now let and let us show
. This is clear if
or if
, so we assume that
,
,
and
are pairwise distinct. By our claim, we have
and
,
so take
with
and
. Note that
thus, . Lemmas 7.5
and 7.15 yield
If , then
by definition. If
, then
, so
by our claim.
Proposition . Then the class
contains exactly one
-truncated
element.
Proof. Let us first show that
contains a
-truncated
element. Suppose that
itself is not
-truncated, let
be
greatest such that
for some
. Setting
,
we have
, so
by Lemma 7.15. Our assumption on
therefore yields
, whence
.
We claim that is
-truncated.
Fix
. By definition of
, we have
for all
. Since
, Lemma 7.15 gives
for all
, so
. By definition, this means
that
for all
.
Since
, we have
, by Lemma 7.15. Thus,
, as claimed.
Now let be
-truncated
series with
. We need to show
that
. Take
with
. For
, we have
since
is
-truncated.
Therefore,
so by Lemma 7.15. Thus
. Since
,
we deduce
, so
. We also have
,
so the same argument gives
and we conclude that
.
For , we define
to be the unique
-truncated
series in
. Note that this
definition extends the previous definition of
. It follows from the proof of Proposition 7.18
that
for all
and that
Proof. We have if and only if
for some
.
Since
for each
,
this is in turn equivalent to
by Lemma 7.5
. Thus,
if and only if
for some
, and it remains to
show that
for some
if
and only if
for some
. This follows from the fact that if
, then
,
so
.
Proposition we have
.
Proof. Let .
Then there is
with
.
Thus,
and so
by
Proposition 7.9. Therefore,
by
Proposition 7.19.
Corollary on
.
Thus, for
, we have
if and only if
.
Proof. Let .
Then
by Proposition 7.20 and
is
-truncated
by
.
The fact that
if and only if
follows from this and the fact that
is
injective.
Proposition is a confluent hyperserial skeleton of
force
. Then
for all
. In particular, if
is defined for
,
then
is defined on
.
Proof. We prove this by induction on . Let
.
By Proposition 7.20, we need only prove that
. Let
.
By Proposition 7.19, there is a
with
. By Proposition 7.9, there is a
with
. Since
is hyperserial
of force
, the
hyperexponential
is defined and
Finally, since , Lemma 4.8 and Proposition 4.9 imply
.
Corollary is a confluent hyperserial skeleton of
force
. Then we have
whenever one of the sides is defined.
Corollary
has force
.
For all , the function
is defined on
.
For all and
,
the hyperexponential
is defined for some
.
For all , we have
.
Proof. The equivalence between ) and
)
follows from Proposition 7.22 and the fact that we have
for all . The equivalence
between
) and
) follows directly from Proposition 7.22.
The equivalence between
) and
) follows from Corollary 7.21.
Throughout this subsection, we let and we set
and
.
Given
, it will be convenient
to introduce the following notations:
Lemma ,
, and
.
We have
Proof. We claim that if ,
then
and
Assuming that , we have
whence . Now
so . Since
, we have
,
so Lemma 7.5 gives
as desired. Composing with gives that if
, then
and
From which it follows that .
Corollary with
.
Then
.
Proof. We have .
Let
with
.
We have
, so
by Lemma 7.5. Since by Lemma 7.25 and
is convex, we are done.
Lemma and each
,
we have
Proof. Take with
. Since
,
we have
by Lemma 7.5. This gives
Thus, . Likewise, since
, we have
so . Lemma 7.5
gives
, so we have
By Lemma 7.25, both and
are elements of
.
Since
is convex, this means that it also
contains
and
.
We have the following useful consequence:
Corollary be such that
for some
. Then
Proof. Take with
. Then
We have by Lemma 7.27 and we have
by Lemma 7.25, so
. Likewise,
.
Since
is convex, this yields
. Since
by Lemma 7.27, we conclude that
.
Corollary with
.
Then
and
.
Proof. As is strictly
increasing, we have
, which
gives
. We first claim that
. If
, then Corollary 7.28 gives that
, so we may focus on the case when
. Suppose towards
contradiction that
and that
for some
. Then
Since , we have
, so
.
Since
, we have
so by Proposition 7.19, a
contradiction.
From our claim, we get . This
yields
, as
. Take
with
. Lemma 7.25 gives
so since
is convex and
is strictly increasing.
In this section, is a confluent hyperserial
skeleton of force
and
. Given
we consider a class
of
-truncated
series without
-hyperexponentials
in
and show how to extend
into a minimal confluent hyperserial skeleton
of force
that contains
for all series
.
Most of the work in the case
has already been
done in [29], but Subsection 8.1 contains a
self-contained treatment for our setting. For the construction of
in Subsections 8.1, 8.2, 8.3, and 8.4, we recall that we made the
induction hypothesis that Theorem 7.4 holds for
. After the construction of
, we conclude this section with the
proofs of Theorems 7.4 and 1.3.
Let be the class of all 1-truncated series
for which
is not defined. Let
be a non-empty subclass of
and let
be the
-subspace
of
generated by
and
. By Lemma 7.12,
consists only of 1-truncated series.
We let be the restriction of
to
. Note the following:
We claim that satisfies
, where
. Then
,
so
by definition. This gives
, which contradicts the fact that
satisfies
By definition, we have if and only if
, so
satisfies
Since for
,
the axiom
Extending .
For
with
,
we have
if and only if
, so
if and only if
if and only if
.
Accordingly, we set
This ensures that holds. Note that if
, then
, so
and
. Thus,
and
. We also have the following:
For , we have
so satisfies
.
For , we have
, since
. Thus
which proves .
satisfies
.
To see this, let
with
and let
. We want to show
that
. Since
and
by
. Now
Since and since
satisfies
, we have
Let . Since
and
satisfies
, the hyperlogarithm
is
-truncated by Lemma 7.14. It follows from Lemma 7.13 that
is also
-truncated,
so
satisfies
.
Let and let
be a
sequence of real numbers. To show that
satisfies
, we need to
show that the sum
is in
. We have
Since and since
satisfies
, we have
. It remains to note that
and that
is closed under
finite sums.
Since and
are both
monomials, they must be equal. The axiom
gives
.
Now ,
,
,
, and
hold
for each
, since they hold in
. Furthermore,
holds if
; this
is clear if
and the above proof of
still goes through when
.
Thus,
is a hyperserial skeleton of force
which extends
.
Proposition . Then
is
-confluent.
Proof. Clearly, is
-confluent. Let
and
take
with
.
We have
. Let
and take
with
. We have
and, by
assumption,
, so
By definition, implies
, so
. We
have
so and, more generally,
for
. Thus, the skeleton
is
-confluent.
Let us summarize:
Proposition is a confluent hyperserial skeleton of
force
. It is an extension of
of force
with
.
Using the composition from Theorem 4.1, we can check
whether an embedding of confluent hyperserial
skeletons is of force
without having to verify
that
for all
.
Lemma be a confluent hyperserial skeleton of force
with the external composition
from Theorem 4.1 and let
be a
strongly linear field embedding. Suppose that
, that
for all
and all
, and
that
for all
and all
. Then
is an embedding of force
.
Proof. We will show by induction on
that
. For
, this holds as
is
order-preserving. Let
and assume that
for all
. If
is a limit, then by
, we have
Suppose is a successor and let
. We have
for all
by
. Our
induction hypothesis gives
for all
. Applying
again gives
, so
.
Proposition be a confluent hyperserial skeleton of force
and let
be an embedding
of force
. If
, then there is a unique embedding
of force that extends
.
Proof. As is hyperserial of
force
, we have an external
composition
. Since
,
is
-linear, and
is an
-subspace of
containing
, we
have
.
Since , we have
so
. Thus,
is a monomial for
by Lemma 7.12. We define a map
by setting
It is routine to check that is an embedding of
ordered monomial groups with
-powers.
By Proposition 2.3, this embedding
uniquely extends into a strongly linear field embedding of
into
, which we
will still denote by
. Note
that if
, then
, so
extends
.
We now prove that is a force
embedding. By Lemma 8.3, we need only show that
commutes with logarithms and hyperlogarithms. Given
, we have
Now let with
and let
. If
, then
,
so we automatically have
,
since
extends
.
If
, then
, so
Let us finally assume that is another embedding
of force
that extends
. To see that
,
it suffices to show that
for
. Now
so .
From this point until Subsection 8.5, we let and set
Note that if
is a
successor and
if
is a
limit. Let
be the class of all
-truncated series
for which
is not defined. Let
be a non-empty subclass of
.
Consider
and
.
We have
by Corollary 7.21. Since
contains a unique
-truncated
element,
is
-truncated
and
, it follows that
. Thus, we have
If is a successor, then let
be the smallest class containing
such that
whenever
and
. By Lemma 7.13, the class
also consists only of
-truncated
series. If
is a limit, then set
. Note that
.
For the remainder of this subsection (with the exception of Proposition
8.30) we assume that
.
Remark and suppose
is a successor.
If
for some
,
so
is defined in
,
then for each
, we have
so . In particular,
.
We define the group as follows. If
is a limit, then
is the
group generated by the elements
with
and satisfying the relations
.
Hence
is the group of products
of is finite. If
is a
successor, then let
be the equivalence relation
on
defined by
We let be the group of formal products
for which the hypersupport is well-based and
is finite. Given
,
we note that
, whence
. Hence
is
indeed a group.
For , we define
and
. We set
if
,
which happens if and only if
.
The following facts will be used frequently, where
range over
:
for
,
, and if
then
If or
,
then
,
If and
or if
and
then
.
Let denote the direct product
. We denote by
a
general element
of this group, where we
implicitly understand that
and
; we also identify
and
with
and
, respectively. In the special case when
, we write
.
Remark is a successor and consider
as above. The advantage of the representation of
as an infinite product of terms of the form
with
is that such a representation
is unique. Alternatively, it is possible to represent
as a finite product of terms of the form
with
, but uniqueness is lost,
since
.
Nevertheless, we may construct a privileged representation as a finite
product as follows. Since is finite, there exist
with
for
and
. Since
is well-based, we may also take
for all
. Then
Fix and set
.
For each
, we have
, so
Set
This gives us the finite representation
Note that .
![]() |
(I) | ||
![]() |
(II) | ||
![]() |
(III) | ||
![]() |
(IV) |
We define the relation on
by
if and only if
.
Proposition is an order on
that
extends the orderings on both
and
.
Proof. By definition, the relation
extends the orderings on
and
. In order to show that
is an order, it suffices to check that
is a
total positive cone on
.
Let . By the definition of
and the fact that
,
it is clear that
and
cannot both be in
. Let us
show that either
or
. Assume that
.
If
and
or
and
, then
satisfies (III) or (IV). Suppose that
,
,
and
. Then
since
, so
. Since
and
, we conclude that
satisfies (II). If
,
, and
then
satisfies (I), for similar reasons.
Now let . We will show that
. If both
and
satisfy one of the last two rules, then this
is clear. Thus, we may assume without loss of generality that
satisfies either rule (I) or rule (II). We consider the
following cases:
Case 1: and
both satisfy (I) or they both satisfy (II). Suppose that they both
satisfy (I). Then
and
, so we need to verify that
. By Corollary 7.26, we have
. Since
, we also have
,
whence
. The case when
and
both satisfy (II) is
similar.
Case 2: satisfies (I) and
satisfies (III) or (IV). We have
, so if
,
then
satisfies (IV). Suppose that
. If
,
then
and if
,
then
, so
as
is strictly increasing. Since
satisfies rule (I) and
,
we have
so satisfies (I).
Case 3: satisfies (II) and
satisfies (III) or (IV). We have
, so if
,
then
satisfies (IV). Suppose that
. If
,
then
, so
and
. Since
satisfies rule (II), we have
,
so
Hence satisfies (II).
Case 4: One of the monomials
and
satisfies (I) and the other one satisfies
(II). Without loss of generality, we may assume that
satisfies (I) and
satisfies (II). Let us first
consider the case when
. Then
, so
and
. Since
and
, we deduce that
, so
by
Corollary 7.29. Since
satisfies
(II), we have
, so
and satisfies (II).
Let us now consider the case when .
If
, then
satisfies (III). If
and
, then
satisfies (IV). If
, then
, so
,
contradicting that
. It
remains to consider the case that
.
Then
, so
as
is strictly increasing. Since
and
, we deduce
that
, so
. Since
,
we have
, so
. This gives
so satisfies (I).
Remark and
, we
have
Since , we also have
. More generally, for
, we have
This is because by Corollary 7.28
with
.
It is easy to check that this defines a real power operation on . Note that
for each non-zero
.
Now that we have defined an ordering and a real power operation on , we let
. Then
is a field of
well-based series extending
.
When
, we write
.
In this subsection, we extend the hyperlogarithms
from
to
,
while verifying that they satisfy the axioms for hyperserial skeletons.
We separate various cases as a function of
,
including the case of the ordinary logarithm when
.
In each case, we start with the definition of the domain of the extended hyperlogarithm
on
and then define
on the
elements of
which do not already lie in
. We next check that
satisfies the domain definition axioms
, as well as the other axioms for hyperserial
skeletons.
We extend to
by setting
for . Note that
if and only if
.
We claim that
is well-defined. Let
and
be as in Remark 8.6,
so
and
Each sum exists in
,
since the support
is a strictly decreasing
sequence in
. Thus,
is well-defined. If
,
then we note that
Finally, we extend log to all of by setting
. We let
be the restriction of
to the class
, so
satisfies
Using the definition of real powers, it is straightforward to check that
satisfies
. If
, then
.
If
, then
by Corollary 7.21. In either case,
for all
. Hence,
for and
follows from
satisfies
Lemma satisfies
Proof. Given ,
we must show that
. We
proceed by case distinction:
If , then
since
satisfies
If , then
and
If , then
and
. Thus
since
.
If
, then
, so Remark 8.8 again yields
. In either case,
.
Suppose ,
, and
.
We have
, so it is enough
to show that
and
. We have
by Lemma 7.25, so ,
whence
and
.
Since
, we also have
, so
Suppose ,
, and
.
This time, we need to show that
and
. Using that
and that
, we have
, so
If and
,
then
. So the result
follows from the fact that
and
.
Lemma satisfies
Proof. Given ,
we need to show that
. If
, then
so
since
satisfies
, then
,
so
. Since
we have . If
, then
.
Consider now the case that ,
, and
. Since
,
we need to show that
. For
each
, we have
by Lemma 7.25. As ,
this gives
. If
, then we have
by Remark 8.8. If ,
so
is defined in
,
then for each
, we have
since is strictly increasing. As
is arbitrary, this gives
.
Finally, suppose ,
, and
.
The same argument as above gives
for
, so
and
.
This sum is well-defined, as for
. For
,
we set
This sum is also well-defined, as is well-based
and
for all
,
and
with
.
If
, then note that
, so
and whenever
.
Finally, we extend
to all of
by setting
for . As before, we let
be the restriction of
to
, so
satisfies
The axiom (and thus
) follow easily from the definition of
and the axiom
for each
. Let us prove that
. Given
, we need to show that
. Since
,
it suffices to show that
.
Since
, this further reduces
to showing that
. But this
follows from the fact that
. The proof that
is identical to cases 3–5 of Lemma
8.9. Let us now show that
also
satisfies
Lemma
satisfies
Proof. We have for
and
for
. It follows that
for
so long as
.
Suppose that
,
, and
.
Then
, so it is enough to
show that
. As shown in the
argument that
. By Lemma 7.27, we also have
.
Thus,
, so
; see Remark 8.8. The case that
,
, and
is similar.
Given with
,
we decompose
, and define
Note that and
whenever
is a limit ordinal. More generally, we have
whenever (including the case when
).
Lemma satisfies
for each
.
Proof. We prove this by induction on , beginning with
.
Let
, so
If , then either
or
. If
, then
if
and only if
. If
, then
if and only if
. It remains to note that
for all
.
Now assume that and that
holds for all
. Since
satisfies
for each
, we may focus on elements of the
form
where
and
. For the remainder of the proof,
we fix such an element. If
is a successor, then
we need to show that
if and only if
. One direction is clear: if
then
for each
.
For the other direction, if
,
then
, so write
and note that
is a monomial if and
only if
. If
is a limit, then
for all
if and only if
, so we have
if and only if
for all
.
Lemma
satisfies
for each
.
Proof. Let and
with
and
.
Since
satisfies
for each
, it suffices to show that
. Decomposing
, we have
,
so
Let , let
, and let
.
We note that
has no infinitesimal terms in its
support, so
is satisfied since it holds in
. To see that
satisfies
, suppose that
is a successor and write
. Then
Lemma satisfies
for each
with
.
Proof. Let with
, let
with
, and let
.
We want to show that
If , then this holds because
satisfies
.
Consider the following cases:
If and
,
then write
and
.
We have
Since , we have
. If
, then
.
If
, then
, so either
or
and
.
In both cases, we have
.
If and
,
then we must have
by Remark 8.8.
Writing
, we have
, so
. By Corollary 7.28, we have
, so
again by Remark 8.8. In particular, .
If and
,
then
. Arguing as in the
previous case, we have
.
Lemma satisfies
for each
.
Proof. Let and let
be a sequence of real numbers. Consider the sum
. We need to show that
. If
,
then
. Suppose
with
. Then
for all
,
so
where .
Note that implies
for
some
. Moreover, if
is a limit, then
.
In other words,
We define
The proofs of Lemmas 8.12 and 8.15 can be
amended to show that satisfies
and
; just replace
with
. Since
satisfies
,
, and
, it suffices to check these axioms for elements of
the form
, where
and
. If
, then
and
if
, then
is
-truncated by Lemma 7.13 (recall that
is a successor),
whence
by Corollary 7.21. In either
case, we have
, so
satisfies
. As
for
, suppose that
is a successor. We have
if and we have
otherwise.
Lemma
satisfies
.
Proof. Let ,
, and
. If
,
then we have
If , then
. Since
,
we have
by Remark 8.8.
Lemma
satisfies
.
Proof. Let with
and let
. We
need to show that
We proceed by case distinction.
If , then this holds
because
satisfies
.
Suppose and
for some
and some
.
Then
If , then either
or
and
. In either case, we have
. If
and
, then
so
by Remark 8.8. A similar argument handles the case that
and
.
Finally, if
and
, then again, either
or
and
.
In the first case, we have
,
so
since both
and
are monomials by Lemma 7.13 and
Corollary 7.21. In the second case, we have
since
and
.
Suppose for some
and
some
and
.
Then
since
.
For each
, we have
by Lemma 7.25, so
. If
,
then
by Remark 8.8. Hence,
If , then
since
is strictly increasing.
Since
is arbitrary, this gives
, so
Suppose and
for some
and some
.
Then
, so similar
arguments as above give
for each
. Again, we conclude that
Lemma
satisfies
Proof. If ,
let
, with
. We have
If , then either
or
. If
, then
if
and only if
. If
, then
if and only if
. By Remark 8.5,
we have
for all , where
by Lemma 7.13 and Corollary 7.21.
If is a successor, then let
and
. We need to show that
if and only if
.
This holds since
and since whenever
,
by Corollary 7.21.
Finally, if is a non-zero limit, then we use
that
To see that and let
. By Remark 8.5, we have
Let . Since
is
-truncated, we have
. This gives
for
. If
, then
is also
-truncated by Lemma 7.13,
so
since
.
This gives
if
.
If
, then
since
is strictly increasing. Since
is a monomial by Corollary 7.21, this gives
. In all three cases, we have
, so
as desired.
If is a successor, then either
or
. In both cases,
so . Since
, we have
,
so Remark 8.8 with
and
gives
Lemma
satisfies
Proof. Let and let
. We want to show that
Note that . We claim that
. If
, then this follows from the fact that
satisfies
and
,
then we have
. If
and
, then
by Remark 8.8 and likewise, if
and
, then
.
Now suppose towards contradiction that .
We will show that
. As
is the unique
-truncated
element in
and
is
-truncated, this is a
contradiction.
Since by
,
we have
, so
By , so
If , then Lemma 7.5
gives
so . Suppose
and let
with
.
If
, then
so by Remark 8.8. As
, this too gives
.
Finally, if
, then
so by Remark 8.8. As
, we have
so once again.
Lemma
satisfies
Proof. Let and let
be a sequence of real numbers. Consider the sum
. If
,
then
since
satisfies
for some
. If
is a limit, then
and
where . If
is a successor, then we may write
If for all
,
then
where . If
for some
, then let
be minimal with
.
We have
, where
Note that
so it remains to show that .
Since
by Lemma 7.13 and Corollary
7.21, this follows from the fact that
satisfies
For , we have
if and only if
since
. This proves that
satisfies
. Let
. We have
so is satisfied. As for
, it suffices to show that
since
for all
by
, we have
Now for , let
. Since
by
. Since
it is enough to show that .
This holds because
satisfies
and
.
Lemma
satisfies
.
Proof. Let with
and let
. We
want to show that
. Since
and likewise for
,
it is enough to show that
We proceed by case distinction:
If , then the result
follows from
for
.
If and
,
then
Since and
satisfies
, we have
If and
,
then
. Since
and
satisfies
, we have
Thus,
If and
,
then the argument is similar to the previous case.
Proof. Let and let
be a sequence of real numbers. We need to show that the
sum
is in
.
If
, then
. If
for some
, then
We have , since
satisfies
since and
satisfies
. We conclude by noting that
is closed under addition.
Remark , the argument that
is satisfied gives
and the proof of Lemma 8.22 also tells us that satisfies
.
Lemma satisfies
for all
.
Proof. It suffices [ satisfies
.
Suppose towards contradiction that there is some
with
. Take
with
. Since
, Lemma 4.8 gives
Since and
are both
monomials, they must be equal. The axiom
gives
, a contradiction.
For all with
,
the axioms
,
,
,
and
automatically hold
in
since they hold in
, as does the axiom
if
is an ordinal.
We have completed the proof of the following:
Proposition is a hyperserial skeleton of force
.
Let us finally examine the confluence and universality of .
Proposition . Then
is
-confluent. In particular,
is
-confluent.
Proof. Clearly, is
-confluent. Consider
and write
. By our definition
of
, we either have have
or
. If
, then
and, more generally,
for all
with
, since
by Lemma 3.7. Assume from now on that
.
We set and
.
For
, let us first show by
induction that
If , then
and
so we indeed have . Let
and suppose that
for
. If
and
is a successor, then our induction hypothesis
yields
Writing , we have
so
If and
is a limit, then
there is
such that
.
For this
, we have
so . Finally, if
and
is a successor, then
, where
if
is a limit. This gives
In both cases, we have , so
, since
.
Let us now show that exists. Let
, so
by our assumption
that
. Take
with
. We have
, so
Since is an infinite monomial, it is
-truncated, so
so long as it is defined. Thus,
is either equal
to
or
.
If , then
for
with
.
On the other hand, if
, then
Take with
.
Then
so and, more generally,
when
and
.
Propositions 8.25 and 8.26 yield:
Corollary , then
is a
confluent hyperserial skeleton of force
.
Remark . Then
Given and
,
we have
. Given
, we have
.
Let us now show that satisfies a universal
property. We start with a lemma.
Lemma with
and any
, we have
.
Proof. Choose and
such that
.
Then
and
so it suffices
to show that
. Since
are monomials and
,
we have
The monotonicity of gives
. We conclude that
,
since
and
are
monomials.
Proposition be a confluent hyperserial skeleton of force
and let
be an embedding
of force
. Let
be a subclass (we no longer require that
). If
,
then there is a unique embedding
of force that extends
.
Proof. Since is confluent, we
have an external composition
.
We first claim that
. If
is a limit, then
so there is
nothing to show. If
is a successor, then for
, there is
with
. We have
so . Having shown our claim,
we may assume without loss of generality that
.
Given , the series
is
-truncated,
so
is
-atomic,
by Remark 7.21. We set
.
Note that for
and
,
the series
exists in by
. Let
.
If
is a limit, then
is
finite and we define
If is a successor, let
and
be as in Remark 8.6. We define
Note that in both the limit and successor case, we have
Given and
,
we have
by Lemma 8.29 and, if
, then
. Thus,
for
. In particular,
is
order preserving, since
Next, we extend to all of
by setting
for
.
Note that
extends
.
It is straightforward to check that
is an
embedding of monomial groups which respects real powers. We need to show
that
preserves the order. Let
. If both
,
then
. This leaves us two
cases to consider:
Suppose ,
, and
.
Set
. We claim that
. If
, then
,
so this follows from Lemma 7.25. If
, then
,
so this follows from Lemmas 7.25 and 7.27.
In either case, we have
,
so
. From this, we see
that
so Since
,
this gives
. Thus
so .
Suppose ,
, and
.
Set
. As before, Lemmas
7.25 and 7.27 give
, so
Since , this gives
, so
and .
By Proposition 2.3, the function
extends uniquely into a strongly linear strictly increasing embedding
, which we still denote by
.
We claim that is an embedding of force
. By Lemma 8.3, we
need only show that
commutes with logarithms and
hyperlogarithms. We begin with logarithms. Let
and
. If
, then
for some
and
If , then
If , then
and
If , then
In all cases, we have, . For
, we have
Now, let and let
.
Note that
, so we need to
show that
. Write
. If
,
then
If , then
. If
,
then
If , then
If , then
and
If , then
and
Since for
and since
for
, this
completes the proof of our claim that
is an
embedding of force
.
We finish with the uniqueness of .
Let
be another embedding of force
that extends
.
To see that
, we only need to
show that
for all
.
For
, we have
so . For
, we deduce that
Since is strongly linear, this gives
for
, so
by the injectivity of
.
In this subsection, we prove Theorem 7.4. Recall that is a confluent hyperserial skeleton of force
.
Definition . For
and
, we define
as follows:
.
if
is a successor.
if
is a non-zero
limit.
if
.
We set , so
and we have the force
inclusion
whenever
or
and
. We
set
Note that by Lemma 2.1. Note also
that
and
.
Theorem 7.4 is a consequence of the next two propositions:
Proposition is a confluent hyperserial skeleton of force
.
Proof. By Corollary 7.24, it suffices to show that
for . Fix
and fix
. Fix also a limit
ordinal
with
.
Then
so either
exists in
or
exists in
. In either case,
.
Proposition be a confluent hyperserial skeleton of force
and let
be a force
embedding. Then there is a unique force
embedding
extending
.
Proof. We will show for each
and each
that there is a unique force
embedding
extending
. We have
,
so suppose that we have defined this unique embedding
when
,
and when
,
. If
is a successor,
then
, so by Proposition 8.4 (if
) or
Proposition 8.30 (if
),
the embedding
extends uniquely to an embedding
Since uniquely extends
and since
uniquely extends
, we see that
uniquely
extends
. If
is a limit, then we set
. The
map
is only defined on
, which may not equal
,
but
is defined on all of
and so
extends uniquely to a force
embedding
,
which we also denote by
. As
each
uniquely extends
, we see that
uniquely
extends
as well. Likewise, we define
to be the unique force
embedding
extending
.
Theorem 7.4 follows from Propositions 8.32 and 8.33.
M. Aschenbrenner, L. van den Dries, and J. van der Hoeven. On numbers, germs, and transseries. In Proc. Int. Cong. of Math. 2018, volume 1, pages 1–24. Rio de Janeiro, 2018.
M. Aschenbrenner, L. van den Dries, and J. van der
Hoeven. The surreal numbers as a universal -field. J. Eur. Math. Soc.,
21(4):1179–1199, 2019.
V. Bagayoko, J. van der Hoeven, and V. Mantova. Defining a surreal hyperexponential. Technical Report, HAL, 2020. http://hal.archives-ouvertes.fr/hal-02861485.
A. Berarducci and V. Mantova. Surreal numbers, derivations and transseries. Technical Report http://arxiv.org/abs/1503.00315, Arxiv, 2015.
É. Borel. Leçons sur les séries divergentes. Gauthier-Villars, 2-nd edition, 1928. Reprinted by Jacques Gabay.
M. Boshernitzan. Hardy fields, existence of transexponential functions. Æquationes Math., 30:258–280, 1986.
N. Bourbaki. Fonctions d'une variable réelle. Éléments de Mathématiques (Chap. 5). Hermann, 2-nd edition, 1961.
B. I. Dahn and P. Göring. Notes on exponential-logarithmic terms. Fundamenta Mathematicae, 127:45–50, 1986.
L. van den Dries. Tame topology and o-minimal structures, volume 248 of London Math. Soc. Lect. Note. Cambridge University Press, 1998.
L. van den Dries and Ph. Ehrlich. Fields of surreal numbers and exponentiation. Fundamenta Mathematicae, 167(2):173–188, 2001.
L. van den Dries, J. van der Hoeven, and E. Kaplan. Logarithmic hyperseries. Trans. of the AMS, 372(7):5199–5241, 2019.
L. van den Dries, A. Macintyre, and D. Marker. Logarithmic-exponential power series. Journal of the London Math. Soc., 56(2):417–434, 1997.
P. du Bois-Reymond. Sur la grandeur relative des infinis des fonctions. Annali di Matematica Pura ed Applicata (1867-1897), 4(1):338–353, 1870.
P. du Bois-Reymond. Über asymptotische Werte, infinitäre Approximationen und infinitäre Auflösung von Gleichungen. Math. Ann., 8:363–414, 1875.
P. du Bois-Reymond. Über die Paradoxen des Infinitärscalcüls. Math. Ann., 11:149–167, 1877.
J. Écalle. Introduction aux fonctions analysables et preuve constructive de la conjecture de Dulac. Hermann, collection: Actualités mathématiques, 1992.
H. Gonshor. An Introduction to the Theory of Surreal Numbers. Cambridge Univ. Press, 1986.
H. Hahn. Über die nichtarchimedischen Größensysteme. Sitz. Akad. Wiss. Wien, 116:601–655, 1907.
G. H. Hardy. Orders of infinity. Cambridge Univ. Press, 1910.
G. H. Hardy. Properties of logarithmico-exponential functions. Proceedings of the London Mathematical Society, 10(2):54–90, 1911.
D. Harvey and J. van der Hoeven. Faster polynomial multiplication over finite fields using cyclotomic coefficient rings. Accepted for publication in J. of Complexity, 2019.
J. van der Hoeven. Automatic asymptotics. PhD thesis, École polytechnique, Palaiseau, France, 1997.
J. van der Hoeven. Transséries fortement monotones. Chapter 1 of unpublished CNRS activity report, https://www.texmacs.org/joris/schmeling/rap1-2000.pdf, 2000.
J. van der Hoeven. Operators on generalized power series. Journal of the Univ. of Illinois, 45(4):1161–1190, 2001.
J. van der Hoeven. Transseries and real differential algebra, volume 1888 of Lecture Notes in Mathematics. Springer-Verlag, 2006.
J. van der Hoeven. Transserial Hardy fields. In Felipe Cano, Frank Loray, Juan José Moralez-Ruiz, Paulo Sad, and Mark Spivakovsky, editors, Differential Equations and Singularities. 60 years of J. M. Aroca, volume 323 of Astérisque, pages 453–487. SMF, 2009.
H. Kneser. Reelle analytische Lösungen der
Gleichung und verwandter
Funktionalgleichungen. Jour. f. d. reine und angewandte
Math., 187(1/2):56–67, 1950.
M. C. Schmeling. Corps de transséries. PhD thesis, Université Paris-VII, 2001.
class of ordinal numbers
7
or
7
if
is a successor,
if
is a limit 7
for each exponent
of
7
for each exponent
of
7
unique ordinal with
and
7
field of well-based series over
8
set of monomials
with
8
dominant monomial
of
8
is a truncation of
9
operator support
of
10
relative support of
10
to the power
10
field of logarithmic hyperseries 11
formal hyperlogarithm
of strength
11
group of logarithmic monomials of strength
11
field of logarithmic series of strength
12
group of logarithmic (hyper)monomials 12
derivation on
12
-th
derivative
of
12
for
,
unique series with
13
hyperlogarithm of force
14
axiom of domain definition
14
class of
-atomic
elements 14
axiom of functional equations
15
axiom of asymptotics 15
axiom of monotonicity
15
axiom of regularity 15
axiom of transfinite products
15
projection
16
class of series
with
16
class of series
with
16
Taylor expansion of
at
20
Taylor expansion of
at
20
axioms for hyperserial fields
36
minimal extension of
of force
40
(partially defined) functional inverse of
41
class of
-truncated
series 44
unique
-truncated
series in
47
minimal extension of
with
50
group of monomials for
50
class of series
53
formal adjunction of
to a field 54
group of formal products
54
-atomic element 14
Cantor normal form 7
coefficient of an ordinal 7
-confluent field 16
confluent hyperserial field 36
confluent hyperserial skeleton 17
confluent hyperserial skeleton of force 40
dominant monomial 8
embedding of force 16
exponent of an ordinal 7
extension of force 16
field of well-based series 8
hyperserial field 36
hypersupport 54
infinitesimal series 9
logarithmic hyperseries 11
monomial group 8
operator support 10
positive infinite series 9
real powering operation 10
relatively well-based function 10
strongly linear function 9
support of a series 8
-truncated series 44
truncation of a series 9
well-based family of series 9
well-based function 9
well-based series 8
well-based set 8