Cousin’s lemma in second-order arithmetic

By Jordan Mitchell Barrett, Rodney G. Downey, and Noam Greenberg

Abstract

Cousin’s lemma is a compactness principle that naturally arises when studying the gauge integral, a generalisation of the Lebesgue integral. We study the axiomatic strength of Cousin’s lemma for various classes of functions, using Friedman and Simpson’s reverse mathematics in second-order arithmetic. We prove that, over :

(i)

Cousin’s lemma for continuous functions is equivalent to ;

(ii)

Cousin’s lemma for Baire class 1 functions is equivalent to ;

(iii)

Cousin’s lemma for Baire class 2 functions, or for Borel functions, is equivalent to (modulo some induction).

1. Introduction

For real-valued functions, the most general version of the fundamental theorem of calculus relies on a generalisation of both the Lebesgue and (improper) Riemann integrals, called, among other names, the Denjoy integral Reference Den12Reference Gor94. This integral can integrate all derivatives, which the Riemann and Lebesgue integrals fail to do Reference Gor94.

Denjoy Reference Den12 first defined this integral in 1912, and shortly after, Luzin Reference Luz12 and Perron Reference Per14 gave equivalent characterisations. However, all these definitions were complex and highly non-constructive, making Denjoy’s integral impractical for applications Reference Gor97. In 1957, Kurzweil Reference Kur57 discovered an equivalent, “elementary” definition, in the style of the Riemann integral. This re-formulation is widely known as the gauge integral, or Henstock–Kurzweil integral.

For bounded functions with compact support, the gauge integral coincides with the Lebesgue integral, and hence the gauge integral gives a Riemann-like definition of the Lebesgue integral for a wide class of functions. We also remark that the gauge integral is suitable for integrating classes of highly discontinuous functions, and can be viewed as a mathematically rigorous formalisation of Feynman’s path integral Reference Mul87.

A gauge is simply a function from the unit interval to the positive real numbers. A -fine partition is a finite tagged partition for which . This generalises the mesh size of a partition: a partition has mesh size if it is -fine for the constant map with value . The gauge integral is defined exactly like the Riemann integral, except that it allows arbitrary gauges rather than only constant ones: the integral is if for every there is a gauge  such that for every -fine partition, the associated Riemann sum is within  of . So every Riemann integrable function is gauge integrable, by choosing constant gauges. But also the function is gauge integrable on , by taking, given , a gauge that approaches 0 sufficiently rapidly so that the contribution of a leftmost interval to a Riemann sum cannot be unbounded. Similarly, Dirichlet’s function (the indicator function of the rationals) is integrable by taking , where lists the rationals in .

The key fact that enables the theory of the gauge integral is Cousin’s lemma Reference Cou95, which states that for any gauge , a -fine partition exists. Without Cousin’s lemma, Kurzweil’s definition of the gauge integral would be vacuously satisfied by all functions and all values, trivialising the notion.

The gauge integral and Cousin’s lemma have had metamathematical exploration before; notably in the setting of descriptive set theory by Dougherty and Kechris Reference DK91, Becker Reference Bec92, and Walsh Reference Wal17, for example. Recently, it was also explored in higher-order reverse mathematics by Normann and Sanders Reference NS19Reference NS20. These studies found that the gauge integral requires powerful axioms to prove. For example, working in third-order arithmetic, Normann and Sanders showed that Cousin’s lemma and the existence of the gauge integral were each equivalent to full second-order arithmetic.

The point is that Cousin’s lemma is a statement about all functions from to ; no definability assumptions are involved. It is thus a statement which is not expressible in second-order arithmetic, which only allows quantification over real numbers (but not sets of real numbers). Second-order arithmetic is the study of countable mathematics, including objects which can be coded by countable objects, for instance separable metric spaces, continuous functions, and Borel sets.

Reverse mathematics of second-order arithmetic, as developed by H. Friedman and S. Simpson, is the project of understanding the proof-theoretic strength of the theorems of mathematics in terms of comprehension and induction strength required to prove them. What axioms of mathematics are actually required to prove a particular statement? The aim is to find optimal proofs, ones which require the least “amount” of axioms; and then show that these proofs are indeed the best that can be found, by finding a “reversal”: a proof of the axioms used, starting with the investigated theorem.

The main insight of the project of reverse mathematics is that almost all theorems of mainstream mathematics are equivalent to one of five axiomatic systems, and that these systems are linearly ordered by logical implication; see, for example, Reference Sim09. One advantage of working within second-order arithmetic is that the proof-theoretic strength is often aligned with complexity, as defined using the tools of computability theory. Very informally speaking, a theorem which asserts the existence of complicated objects requires strong axioms to prove, and vice versa. This connection between computability and proof theory has resulted in a rich body of research.

It is thus natural to investigate the strength of Cousin’s lemma within second-order arithmetic, but in order to do so, we must restrict ourselves to classes of countably-coded functions. In this paper we consider Borel functions, and sub-classes of these.

Below we will observe that Cousin’s lemma is a form of a compactness principle. We would thus naturally start by looking at the system (weak König’s lemma), which informally is understood to be equivalent to the compactness of the unit interval. But more specifically, it is equivalent to countable compactness, for example, to the statement that any countable open cover of has a finite subcover Reference Sim09, Thm.IV.1.2. For continuous functions, Cousin’s lemma should be equivalent to its version for constant functions. Despite the fact that this equivalence seems to use a form of compactness, namely that continuous functions on obtain a minimum, we are able to show:

Cousin’s lemma for continuous gauges is equivalent to .

However, once we go up the hierarchy of Borel functions (as measured say by the Baire class of a function), it turns out that this form of uncountable compactness is significantly stronger:

Cousin’s lemma for Baire class 1 gauges is equivalent to .

Cousin’s lemma for all Borel gauges is provable in -induction.

Cousin’s lemma for Baire class 2 gauges implies .

We leave open the question of whether the extra induction is required. For background on reverse mathematics, and the definitions of these three subsystems of second-order arithmetic, see Reference Sim09.

2. Cousin’s lemma and compactness

Let be a gauge. We can view  as assigning, to each , the open interval . Cousin’s lemma is very close to saying that this open cover of the unit interval has a finite subcover. It will be easier to work with this simplified version, rather than with tagged partitions, partly because this notion can be extended to other compact metric spaces; we will be working with Cantor space as well as the unit interval.

The metamathematical twist is that for our reversals, we will be working over . Some of the gauges that we will work with may be definable in some models of , but not have values in these models. For example, we will work with Baire class 1 functions, coded in a model by a sequence of continuous functions, where the values of the functions are Cauchy sequences, which need not converge. Nonetheless, relations such as will be still definable in the model, and this will be enough to formalise Cousin’s lemma.

In light of this, it is important to note that a -fine tagged partition , with , gives us not only the finite set of points giving us a cover, say , but also the distances , bounded by ; and while may not be in the model, the value is. Thus we define:

Definition 2.1.

Let be a compact metric space, and let be a gauge on . A -fine cover of  is a finite set and a function from to  satisfying:

(i)

; and

(ii)

For all , .

Assuming that is a function coded in the model for which the relations (for an open or closed ball ) are defined, we have:

Lemma 2.2 ().

Let be a gauge.

(a)

If there is a -fine tagged partition then there is a -fine cover.

(b)

If there is a -fine cover then there is a -fine tagged partition.

Proof.

For (a), suppose that is a -fine partition: for all , . Let , and for let . Then for all , and trivially , so is a -fine cover.

For (b), suppose that equipped with is a -fine cover. For brevity, for let . By applying a reverse greedy algorithm, we may assume that  is minimal in that for no distinct in  do we have . This implies that for all , ; otherwise, a point between  and  would not be covered by or by ; if then either (if ) or (if ). So we choose

as well as and ; then is a -fine partition, as .

In light of Theorem 2.2, we state:

Definition 2.3 (Cousin’s lemma).

Let  be a compact metric space, and let  be a class of functions. Cousin’s lemma for  on  states that every in  has a -fine cover.

If not stated otherwise, we work with . For many of our arguments, though, it will be more convenient to work in Cantor space, under the metric for the largest  satisfying . We state one connection between Cousin’s lemma on Cantor space and on the unit interval informally. Once we work with particular classes of functions, we will see that the following argument, in each particular case, can be made to hold in .

Lemma 2.4.

For reasonable classes of functions, Cousin’s lemma for  on Cantor space implies Cousin’s lemma for  on the unit interval.

Proof.

We use the standard map defined by

Then  is continuous and for all ,

Hence, given in , we let , which is a gauge on Cantor space. Suppose that is a -fine cover. Let , and let ; this is a -fine cover.

3. The Borel case

The “classical” proof of Cousin’s lemma can be carried out in the system of -induction. Note that while we can formalise Borel codes and functions in , in the system we can prove that for every Borel function  and every , there is a point (this follows from Reference Sim09, Lem.V.3.7).

We need the following:

Lemma 3.1 (-induction).

If is an infinite -definable tree, then  has a path.

Proof.

Let

Then  is (this uses -choice, which is provable in ; see Reference Sim09, VIII.3.21). By assumption, the empty sequence is in ; and every has a child in . So we can apply -dependent choice on numbers, which is provable in -induction Reference Sim09, VIII.4.10.

Proposition 3.2 (-induction).

Cousin’s lemma holds for all Borel functions.

Proof.

The proof of Theorem 2.4 holds in  for the class  of Borel functions. Hence, we work with Cantor space. Let be a Borel function. Let

where is the clopen set determined by . Note that if and only if , so this is a Borel condition. However, the quantification over all means that  is (relative to a code for ).

Now let

Then  is a tree. We claim that it is finite. If not, then by Theorem 3.1, has a path . Let  be sufficiently large so that ; this contradicts .

So we know that there is some  which bounds the length of all the strings in . By induction, exists, as a finite set: by induction on the length-lexicographic order of strings of length , we show that for all , is a finite set in the model. The formula over which we induct is of the form , so -induction certainly suffices.

It follows that the set of leaves of  exists (again as a finite set). By taking all immediate extensions of the leaves of , we obtain a finite antichain of strings with .

Finally, there is a sequence witnessing that each is in , i.e., and . This follows from -choice (or by -induction, as  is finite). Letting and gives a -fine cover.

4. The continuous case

In this section we show that over the base system  of recursive comprehension, Cousin’s lemma for continuous functions is equivalent to weak König’s lemma . If is continuous then for all , exists, and so for a -fine cover , we can always choose .

In one direction, the argument is quick.

Proposition 4.1 ().

Cousin’s lemma holds for continuous functions.

Proof.

Let be continuous. Weak König’s lemma implies that  obtains a minimum  Reference Sim09, Th.IV.2.2, which must be positive. A partition of  into intervals of length , and choosing any points as tags, gives a -fine partition.

Note that the argument works for Cantor space as well.

In the other direction, we could argue directly by coding a class by an effectively closed subset of ; but the argument is combinatorially simpler by passing via the Heine-Borel theorem.

Theorem 4.2 ().

Cousin’s lemma for continuous functions implies .

Proof.

is equivalent to the Heine-Borel theorem, which states the countable compactness of the unit interval: if is a sequence of open intervals and , then there is finite subcover: there is some such that (see Reference Sim09, Lem IV.1.1).

Let be such an open cover of .

For , let be the distance from a point  to the complement of . A code for  as a continuous function can be obtained uniformly from the (end-points of) the interval . Let

Then  is a continuous function on  (a code can be obtained effectively from the codes for the ’s). Since covers the unit interval, for all  there is some such that , so is positive on .

We claim: for all and ,

(*)

if then .

Suppose not: then for all , . As a result,

which is impossible; of course we use for all .

Now let  be a -fine cover (as mentioned we can choose ). For each find some with ; by  bounding (see for example the proof of Reference Sim09, Lem.IX.1.8), there is some  such that for all , . Hence covers , as required.

Remark 4.3.

The proof above may obscure the main intuition behind it. Consider a model in which fails, witnessed by a “cover” which is not really a cover: outside the model, there are points not covered by any , but there are none such inside it. Let . This is an effectively closed set (relative to an oracle in the model), and the rough idea is to let . Then no can be covered by a -fine partition. The fact that for any finite , is not covered by reflects down to the original model. The function  is not really a gauge, but the model doesn’t see this, as it thinks that  is empty.

The only difficulty is that , while continuous, is not computable relative to an oracle in the model; it is only lower semi-computable. In other words, the model does not contain a code for this function. The definition of  in the proof above gives a continuous lower bound to with a code in the model.

5. Baire class 1

A function is Baire class 1 if it is the pointwise limit of continuous functions. The Baire classes were introduced by Baire in his PhD thesis Reference Bai99, as a natural generalisation of the continuous functions. One motivation for Baire functions is that many functions arising in analysis are not continuous, such as step functions Reference Hea93, Walsh functions Reference Wal23, or Dirichlet’s function Reference Dir29. However, all such “natural” functions generally have low Baire class; for example, the derivative of any differentiable function is Baire class 1, as are functions arising from Fourier series Reference KL90.

The Baire classes have previously been studied with respect to computability Reference KT14Reference PDD17. In particular, Kuyper and Terwijn showed that a real number is 1-generic if and only if every effective Baire 1 function is continuous at Reference KT14. Indeed, the Turing jump function is in some sense a universal Baire class 1 function on Baire space.

In second-order arithmetic, we can formalise this notion as follows:

Definition 5.1.

A code for a Baire class 1 function is a sequence of (codes of) continuous functions such that for all , is a Cauchy sequence.

Note that we do not require the limit of this Cauchy sequence to exist; this would follow from arithmetic comprehension, but the definition still makes sense in . If is a code in the model for a Baire class 1 function , the relation for an open or closed ball  is definable, even if is not an object in the model; for example, we say that if for some , for all but finitely many , . Thus, it is meaningful to say that is a code for a gauge.

In this section we show that Cousin’s lemma for Baire class 1 functions is equivalent to . We start with:

Lemma 5.2 ().

If is Baire class 1, then for open sets , is , uniformly.

Proof.

What this means: if is a code for a Baire class 1 function, and is a sequence of (codes of) open subsets of , then there is a sequence of (codes of) subsets of  with

where again is a definable relation on  and .

The lemma is well-known. The -preimage of a closed ball by a continuous function is closed, and an open ball is the union of closed balls. Hence

Remark 5.3.

In fact, classically, a function is Baire class 1 if and only if the pull back of open sets is . The other direction requires a little bit of work; it is like Shoenfield’s limit lemma, but for arbitrary spaces requires some topological considerations.

In second-order arithmetic, to make sense of the other direction, we would need to define codes for functions which pull back open sets to  sets, similar to how continuous functions are coded by collection of pairs indicating that the pull back of an open set contains some open set. For such a code, we can computably construct a Baire class 1 code for the function, but the argument seems to require induction.

However, even stating the consistency of such a “Borel class 2” code seems complicated, as the containment relation between sets is complete.

Proposition 5.4 ().

Cousin’s lemma holds for Baire class 1 gauges.

Proof.

The argument of Theorem 2.4 holds for Baire class 1 functions, and so it suffices to work in Cantor space. Let be a Baire class 1 gauge. By , for all , exists.

Now the main two points are:

(i)

In , we can tell which subsets of  are empty;

(ii)

In , given a sequence of non-empty subsets of , there is a choice sequence so that for all .

The point is that whether a tree has a path is an arithmetic question, and so whether the union of a countable sequence of closed subsets of Cantor space is empty or not is also arithmetic. For (ii), using  we can obtain a sequence of binary trees, each of which has a path, such that (where is the set of paths of ), and then apply (or in , just take the leftmost path of each ).

We can therefore repeat the proof of Theorem 3.2. Let be a Baire class 1 gauge on Cantor space. We define the set  and the tree  in exactly the same way; and  exist since they are arithmetically definable: is  by Theorem 5.2. The same argument shows that  cannot have a path, and so by , is finite. As above, from  we obtain a finite subset  of  which covers all of Cantor space. By (ii), we can find a sequence as required.

In the other direction:

Theorem 5.5 ().

Cousin’s lemma for Baire class 1 functions implies .

The idea of the proof is as follows. Suppose we have a Cauchy sequence with no limit. The sequence of functions determines a Baire class 1 function , which is a gauge since has no limit. But there cannot be any -fine partition, since no such partition can cover the gap where should be. Here are the details.

Proof of Theorem 5.5.

If fails then there is an increasing Cauchy sequence of real numbers in the unit interval that has no limit Reference Sim09, Thm.III.2.2. For each , define

This is a sequence of continuous functions. This sequence is pointwise Cauchy: for all , and ,

so we use the fact that is Cauchy. Note that in the “real world”, converges uniformly to the continuous function where , but since does not exist in the model, this function does not have a code in the model. However in we have just shown that is a code for a Baire class 1 function, which we call .

We claim that  is a gauge. For all , since , and since is Cauchy, there is some such that for all but finitely many , . Then .

Now by Theorem 2.2 and Cousin’s lemma applied to the gauge , there is a -fine tagged partition ; recall that this means that for all .

We derive a contradiction. By -induction, there is a least  such that for all , . Since is increasing, for all but finitely many , . Since none of , or are the limit of , there is some such that one of the following happens:

for all but finitely many , ; or

for all but finitely many , .

Without loss of generality, assume the former. Then for all but finitely many ,

whence , contradicting our assumption. The argument in the other case is the same.

6. Higher Baire classes

Baire class 1 functions aren’t closed under taking pointwise limits. Indeed, by iterating the operation of taking pointwise limits, we obtain a transfinite hierarchy of functions, which exhausts all Borel functions.

In second-order arithmetic, recall that in , for a Baire class 1 function , the relation for open or closed balls  is definable even if does not exist. We can therefore iterate. By (external) induction on standard  we define:

Definition 6.1.

For each , a code for a Baire class  function  is a sequence of codes for Baire class  functions such that for all , is Cauchy, in the sense that for all  there is an open ball  of radius such that for all but finitely many , .

If  is such a code, then for any open ball , we say that if for some , for all but finitely many , .

Of course by taking constant sequences, we see that every Baire class  function is also Baire class , so the classes are increasing.

implies, for each , that if is Baire class  then exists for all . Similarly to Theorem 5.2, the inverse images of open sets by Baire class  functions are . We can further extend the definition to Baire class  functions for ordinals ; a code in this case is an -ranked well-founded tree, where each non-leaf has full splitting, the leaves are labeled by continuous functions, and each non-leaf node represents the pointwise limit of the functions represented by its children. The relation is then defined by transfinite recursion on the rank of a node, and so it makes most sense to use as a base system for this kind of development; we do not pursue it here. We remark that implies that every Borel function is Baire class  for some .⁠Footnote1

1

This follows from the fact that every well-founded tree is ranked; see Reference Hir00.

Now the proof of Theorem 5.4 cannot be replicated for Baire class  functions for any for . While suffices to determine if a given subset of Cantor space is empty or not, it is -complete to determine whether a given  set is empty.⁠Footnote2 And indeed we show that Cousin’s lemma for Baire class 2 functions is much stronger than : it implies .

2

Every set is the projection of a set in Cantor space (see for example Reference Sac90, Thm.I.1.5), and the question of whether a set is empty or not is equivalent to a question about well-foundedness of a tree, which is -complete.

It would again be more convenient to work in Cantor space, and so we need a converse of Theorem 2.4.

Lemma 6.2 ().

Cousin’s lemma for Baire class 2 functions on the unit interval implies Cousin’s lemma for Baire class 2 functions on Cantor space.

Of course there is nothing special for the case , the lemma holds for all Baire classes.

Proof.

We use the standard embedding of Cantor space into the unit interval: for let

and let be the ternary Cantor set. We use the following facts: is continuous; the function is computable on ; and: for , if then , i.e.,

for all .

Let be a Baire class 2 gauge on Cantor space. Define by letting, for ,

Then is positive on , and it is Baire class 2; we use the fact that  (the characteristic function of ) is Baire class 1, indeed with a very nice, uniformly computable approximation of piecewise linear functions with for all  when , and for all but finitely many  when .

Suppose that is a -fine cover. Define and for let . Then is a -fine cover. The main point is that cannot be covered by any , as for such  we have .

Before we give the reversal, we motivate our technique by showing something weaker: that does not imply Cousin’s lemma for Baire class 2 functions. We will show that every -model of Cousin’s lemma for Baire class 2 functions must contain (in fact for any computable ordinal ), in particular Cousin’s lemma for Baire class 2 functions fails in the model consisting of the arithmetic sets.

We review some definitions. The Turing jump of is the complete set, identified with an element of . As mentioned above, the function is Baire class 1, but more importantly, its graph (the relation ) is . We will use the following well-known fact, which is provable in :

Lemma 6.3 ().

Let be a function. The following are equivalent:

(1)

is -definable;

(2)

There is a computable function such that for all , ;

(3)

is effective Baire class 2.

Proof.

This is a double application of Shoenfield’s limit lemma and Post’s hierarchy theorem: a function is -definable if and only if it is -computable if and only if for some computable (see for example Reference Soa87, Lem.III.3.3;Thm.IV.2.2). The lemma now follows from a uniform relativisation to oracles .

As a result we get:

Lemma 6.4 ().

Suppose that is . Suppose that Cousin’s lemma for Baire class 2 functions holds. Then there is a finite such that

Proof.

Define . Then is . By Theorem 6.3 (note that we already know that Cousin’s lemma for Baire class 2 functions implies ), is a Baire class 2 gauge. A -fine cover is as required.

Recall the definition of a transfinite iteration of the Turing jump. Suppose that is a linear ordering of a subset of , which will usually be well-founded. An iteration of the Turing jump along  is a set satisfying, for all ,

where and . Again note that we view the domain of  as a subset of , so both and are identified with elements of Cantor space, and so taking the Turing jump makes sense.

If  is indeed an ordinal (is well-founded), then implies there is at most one iteration of the Turing jump along , and if  is such, we write . When this is relativised to an oracle , we write . The relation is an iteration of the Turing jump along  is . is equivalent to the statement that for all ordinals  and all , exists Reference Sim09, Thm.VIII.3.15.

Now let be an -model of , let be a computable ordinal, and suppose that ; we show that Cousin’s lemma for Baire class 2 functions fails in . Let . If , there is some such that

let be the least such . The function (as well as its domain) is . Further, for such , we let be the least  such that

the function is as well. By their definition, for all ,

as . Hence, we let ; for all ,

Since , is total on , and since  is an -model, the same definition holds in , so believes that is a Baire class 2 gauge on Cantor space. Let be finite. Since , we have

and so

But since  is finite, the latter fact is absolute for  (the set is a clopen set and so its complement is a non-empty clopen set;  contains an element of any non-empty clopen set). Hence Theorem 6.4 fails in .

Exactly the same argument shows that for any computable ordinal , is an element of every -model of Cousin’s lemma for Baire class 2 functions. The more general argument below builds on this technique. In the meantime, we note that this argument gives a purely computability-theoretic result:

Theorem 6.5.

For every computable ordinal , there is an effective Baire class 2 gauge  such that every -fine partition must contain .

Proof.

Extend the function  above to a function on all of  by defining , and transform to a gauge as in the proof of Theorem 6.4. This is still a -definable function since the relation is (indeed it is ).

Theorem 6.6 ().

Cousin’s lemma for Baire class 2 functions implies .

Proof.

By Theorem 5.5, we may work in .

Let be an ordinal and let ; and suppose that does not exist, i.e., there is no iteration of the Turing jump (relativised to ) along . For simplicity of notation, suppose .

Let

The set  does not exist (as a set in the model), but is definable. Again note that since holds, for all there is exactly one iteration of the Turing jump along , which is the set we call . We observe that  cannot have a greatest element; if exists then by , so does .

For any , since is not an iteration of the Turing jump along , by arithmetic comprehension, there is a least such that ; call this . Then , so . Similar to the argument above, we also let be the least  such that

From and  we can compute some such that for all in ,

Let be finite. Then

is an element of . Since does not have a last element, take any in ; then

Therefore, Theorem 6.4 fails, and so Cousin’s lemma for Baire class 2 functions does not hold.

Mathematical Fragments

Lemma 2.2 ().

Let be a gauge.

(a)

If there is a -fine tagged partition then there is a -fine cover.

(b)

If there is a -fine cover then there is a -fine tagged partition.

Lemma 2.4.

For reasonable classes of functions, Cousin’s lemma for  on Cantor space implies Cousin’s lemma for  on the unit interval.

Lemma 3.1 (-induction).

If is an infinite -definable tree, then  has a path.

Proposition 3.2 (-induction).

Cousin’s lemma holds for all Borel functions.

Lemma 5.2 ().

If is Baire class 1, then for open sets , is , uniformly.

Proposition 5.4 ().

Cousin’s lemma holds for Baire class 1 gauges.

Theorem 5.5 ().

Cousin’s lemma for Baire class 1 functions implies .

Lemma 6.3 ().

Let be a function. The following are equivalent:

(1)

is -definable;

(2)

There is a computable function such that for all , ;

(3)

is effective Baire class 2.

Lemma 6.4 ().

Suppose that is . Suppose that Cousin’s lemma for Baire class 2 functions holds. Then there is a finite such that

References

Reference [Bai99]
René-Louis Baire, Sur les fonctions de variables réelles, Ann. Mat. Pura Appl. (4) 3 (1899), no. 1, 1–123.
[Bar20]
Jordan Mitchell Barrett, The reverse mathematics of Cousin’s lemma, Honours Thesis, Victoria University of Wellington, 2020.
Reference [Bec92]
Howard Becker, Descriptive set-theoretic phenomena in analysis and topology, Set theory of the continuum (Berkeley, CA, 1989), Math. Sci. Res. Inst. Publ., vol. 26, Springer, New York, 1992, pp. 1–25, DOI 10.1007/978-1-4613-9754-0_1. MR1233807,
Show rawAMSref \bib{Bec92}{article}{ label={Bec92}, author={Becker, Howard}, title={Descriptive set-theoretic phenomena in analysis and topology}, conference={ title={Set theory of the continuum}, address={Berkeley, CA}, date={1989}, }, book={ series={Math. Sci. Res. Inst. Publ.}, volume={26}, publisher={Springer, New York}, }, date={1992}, pages={1--25}, review={\MR {1233807}}, doi={10.1007/978-1-4613-9754-0\_1}, }
Reference [Cou95]
Pierre Cousin, Sur les fonctions de variables complexes (French), Acta Math. 19 (1895), no. 1, 1–61, DOI 10.1007/BF02402869. MR1554861,
Show rawAMSref \bib{cousinFonctionsVariablesComplexes1895}{article}{ label={Cou95}, author={Cousin, Pierre}, title={Sur les fonctions de $n$ variables complexes}, language={French}, journal={Acta Math.}, volume={19}, date={1895}, number={1}, pages={1--61}, issn={0001-5962}, review={\MR {1554861}}, doi={10.1007/BF02402869}, }
Reference [Den12]
Arnaud Denjoy, Une extension de l’intégrale de M. Lebesgue, C. R. Acad. Sci. Paris 154 (1912), 859–862.
Reference [Dir29]
G. Lejeune Dirichlet, Sur la convergence des séries trigonométriques qui servent à représenter une fonction arbitraire entre des limites données (French), J. Reine Angew. Math. 4 (1829), 157–169, DOI 10.1515/crll.1829.4.157. MR1577724,
Show rawAMSref \bib{dirichletConvergenceSeriesTrigonometriques1829}{article}{ label={Dir29}, author={Lejeune Dirichlet, G.}, title={Sur la convergence des s\'{e}ries trigonom\'{e}triques qui servent \`a repr\'{e}senter une fonction arbitraire entre des limites donn\'{e}es}, language={French}, journal={J. Reine Angew. Math.}, volume={4}, date={1829}, pages={157--169}, issn={0075-4102}, review={\MR {1577724}}, doi={10.1515/crll.1829.4.157}, }
Reference [DK91]
Randall Dougherty and Alexander S. Kechris, The complexity of antidifferentiation, Adv. Math. 88 (1991), no. 2, 145–169, DOI 10.1016/0001-8708(91)90006-S. MR1120611,
Show rawAMSref \bib{DK91}{article}{ label={DK91}, author={Dougherty, Randall}, author={Kechris, Alexander S.}, title={The complexity of antidifferentiation}, journal={Adv. Math.}, volume={88}, date={1991}, number={2}, pages={145--169}, issn={0001-8708}, review={\MR {1120611}}, doi={10.1016/0001-8708(91)90006-S}, }
Reference [Gor94]
Russell A. Gordon, The integrals of Lebesgue, Denjoy, Perron, and Henstock, Graduate Studies in Mathematics, vol. 4, American Mathematical Society, Providence, RI, 1994, DOI 10.1090/gsm/004. MR1288751,
Show rawAMSref \bib{gordonIntegralsLebesgueDenjoy1994}{book}{ label={Gor94}, author={Gordon, Russell A.}, title={The integrals of Lebesgue, Denjoy, Perron, and Henstock}, series={Graduate Studies in Mathematics}, volume={4}, publisher={American Mathematical Society, Providence, RI}, date={1994}, pages={xii+395}, isbn={0-8218-3805-9}, review={\MR {1288751}}, doi={10.1090/gsm/004}, }
Reference [Gor97]
Russell A. Gordon, Is nonabsolute integration worth doing?, Real Anal. Exchange 22 (1996/97), no. 1, 23–33. Summer Symposium in Real Analysis, XX (Windsor, ON, 1996). MR1433598,
Show rawAMSref \bib{gordonNonabsoluteIntegrationWorth1996}{article}{ label={Gor97}, author={Gordon, Russell A.}, title={Is nonabsolute integration worth doing?}, note={Summer Symposium in Real Analysis, XX (Windsor, ON, 1996)}, journal={Real Anal. Exchange}, volume={22}, date={1996/97}, number={1}, pages={23--33}, issn={0147-1937}, review={\MR {1433598}}, }
Reference [Hea93]
Oliver Heaviside, Electromagnetic Theory, volume 1. The Electrician Printing and Publishing Company, London, 1893.
Reference [Hir00]
Jeffry L. Hirst, Reverse mathematics and rank functions for directed graphs, Arch. Math. Logic 39 (2000), no. 8, 569–579, DOI 10.1007/s001530050165. MR1797809,
Show rawAMSref \bib{Hirst}{article}{ label={Hir00}, author={Hirst, Jeffry L.}, title={Reverse mathematics and rank functions for directed graphs}, journal={Arch. Math. Logic}, volume={39}, date={2000}, number={8}, pages={569--579}, issn={0933-5846}, review={\MR {1797809}}, doi={10.1007/s001530050165}, }
Reference [KL90]
A. S. Kechris and A. Louveau, A classification of Baire class functions, Trans. Amer. Math. Soc. 318 (1990), no. 1, 209–236, DOI 10.2307/2001236. MR946424,
Show rawAMSref \bib{kechrisClassificationBaireClass1990}{article}{ label={KL90}, author={Kechris, A. S.}, author={Louveau, A.}, title={A classification of Baire class $1$ functions}, journal={Trans. Amer. Math. Soc.}, volume={318}, date={1990}, number={1}, pages={209--236}, issn={0002-9947}, review={\MR {946424}}, doi={10.2307/2001236}, }
Reference [KT14]
Rutger Kuyper and Sebastiaan A. Terwijn, Effective genericity and differentiability, J. Log. Anal. 6 (2014), Paper 4, 14, DOI 10.4115/jla.2014.6.4. MR3257036,
Show rawAMSref \bib{kuyperEffectiveGenericityDifferentiability2014}{article}{ label={KT14}, author={Kuyper, Rutger}, author={Terwijn, Sebastiaan A.}, title={Effective genericity and differentiability}, journal={J. Log. Anal.}, volume={6}, date={2014}, pages={Paper 4, 14}, review={\MR {3257036}}, doi={10.4115/jla.2014.6.4}, }
Reference [Kur57]
Jaroslav Kurzweil, Generalized ordinary differential equations and continuous dependence on a parameter (Russian), Czechoslovak Math. J. 7(82) (1957), 418–449. MR111875,
Show rawAMSref \bib{kurzweilGeneralizedOrdinaryDifferential1}{article}{ label={Kur57}, author={Kurzweil, Jaroslav}, title={Generalized ordinary differential equations and continuous dependence on a parameter}, language={Russian}, journal={Czechoslovak Math. J.}, volume={7(82)}, date={1957}, pages={418--449}, issn={0011-4642}, review={\MR {111875}}, }
Reference [Luz12]
Nikolai N. Luzin, Sur les propriétés de l’intégrale de M. Denjoy, C. R. Acad. Sci. Paris 155 (1912), 1475–1478.
Reference [Mul87]
P. Muldowney, A general theory of integration in function spaces, including Wiener and Feynman integration, Pitman Research Notes in Mathematics Series, vol. 153, Longman Scientific & Technical, Harlow; John Wiley & Sons, Inc., New York, 1987. MR887535,
Show rawAMSref \bib{muldowney}{book}{ label={Mul87}, author={Muldowney, P.}, title={A general theory of integration in function spaces, including Wiener and Feynman integration}, series={Pitman Research Notes in Mathematics Series}, volume={153}, publisher={Longman Scientific \& Technical, Harlow; John Wiley \& Sons, Inc., New York}, date={1987}, pages={viii+115}, isbn={0-582-99465-9}, review={\MR {887535}}, }
Reference [NS19]
Dag Normann and Sam Sanders, On the mathematical and foundational significance of the uncountable, J. Math. Log. 19 (2019), no. 1, 1950001, 40, DOI 10.1142/S0219061319500016. MR3960896,
Show rawAMSref \bib{sandersNormannCousin}{article}{ label={NS19}, author={Normann, Dag}, author={Sanders, Sam}, title={On the mathematical and foundational significance of the uncountable}, journal={J. Math. Log.}, volume={19}, date={2019}, number={1}, pages={1950001, 40}, issn={0219-0613}, review={\MR {3960896}}, doi={10.1142/S0219061319500016}, }
Reference [NS20]
Dag Normann and Sam Sanders, Open sets in computability theory and reverse mathematics, J. Logic Comput. 30 (2020), no. 8, 1639–1679, DOI 10.1093/logcom/exaa049. MR4182786,
Show rawAMSref \bib{MR4182786}{article}{ label={NS20}, author={Normann, Dag}, author={Sanders, Sam}, title={Open sets in computability theory and reverse mathematics}, journal={J. Logic Comput.}, volume={30}, date={2020}, number={8}, pages={1639--1679}, issn={0955-792X}, review={\MR {4182786}}, doi={10.1093/logcom/exaa049}, }
Reference [PDD17]
Michelle Porter, Adam Day, and Rodney Downey, Notes on computable analysis, Theory Comput. Syst. 60 (2017), no. 1, 53–111, DOI 10.1007/s00224-016-9732-y. MR3596781,
Show rawAMSref \bib{porterNotesComputableAnalysis2017}{article}{ label={PDD17}, author={Porter, Michelle}, author={Day, Adam}, author={Downey, Rodney}, title={Notes on computable analysis}, journal={Theory Comput. Syst.}, volume={60}, date={2017}, number={1}, pages={53--111}, issn={1432-4350}, review={\MR {3596781}}, doi={10.1007/s00224-016-9732-y}, }
Reference [Per14]
Oskar Perron, Über den Integralbegriff, Sitzungsberichte, Heidelberger Akademie der Wissenschaften, 1914.
Reference [Sac90]
Gerald E. Sacks, Higher recursion theory, Perspectives in Mathematical Logic, Springer-Verlag, Berlin, 1990, DOI 10.1007/BFb0086109. MR1080970,
Show rawAMSref \bib{Sacks}{book}{ label={Sac90}, author={Sacks, Gerald E.}, title={Higher recursion theory}, series={Perspectives in Mathematical Logic}, publisher={Springer-Verlag, Berlin}, date={1990}, pages={xvi+344}, isbn={3-540-19305-7}, review={\MR {1080970}}, doi={10.1007/BFb0086109}, }
Reference [Sim09]
Stephen G. Simpson, Subsystems of second order arithmetic, 2nd ed., Perspectives in Logic, Cambridge University Press, Cambridge; Association for Symbolic Logic, Poughkeepsie, NY, 2009, DOI 10.1017/CBO9780511581007. MR2517689,
Show rawAMSref \bib{simpsonSubsystemsSecondOrder2009}{book}{ label={Sim09}, author={Simpson, Stephen G.}, title={Subsystems of second order arithmetic}, series={Perspectives in Logic}, edition={2}, publisher={Cambridge University Press, Cambridge; Association for Symbolic Logic, Poughkeepsie, NY}, date={2009}, pages={xvi+444}, isbn={978-0-521-88439-6}, review={\MR {2517689}}, doi={10.1017/CBO9780511581007}, }
Reference [Soa87]
Robert I. Soare, Recursively enumerable sets and degrees, Perspectives in Mathematical Logic, Springer-Verlag, Berlin, 1987. A study of computable functions and computably generated sets, DOI 10.1007/978-3-662-02460-7. MR882921,
Show rawAMSref \bib{Soare}{book}{ label={Soa87}, author={Soare, Robert I.}, title={Recursively enumerable sets and degrees}, series={Perspectives in Mathematical Logic}, note={A study of computable functions and computably generated sets}, publisher={Springer-Verlag, Berlin}, date={1987}, pages={xviii+437}, isbn={3-540-15299-7}, review={\MR {882921}}, doi={10.1007/978-3-662-02460-7}, }
Reference [Wal23]
J. L. Walsh, A closed set of normal orthogonal functions, Amer. J. Math. 45 (1923), no. 1, 5–24, DOI 10.2307/2387224. MR1506485,
Show rawAMSref \bib{walshClosedSetNormal1923}{article}{ label={Wal23}, author={Walsh, J. L.}, title={A closed set of normal orthogonal functions}, journal={Amer. J. Math.}, volume={45}, date={1923}, number={1}, pages={5--24}, issn={0002-9327}, review={\MR {1506485}}, doi={10.2307/2387224}, }
Reference [Wal17]
Sean Walsh, Definability aspects of the Denjoy integral, Fund. Math. 237 (2017), no. 1, 1–29, DOI 10.4064/fm773-9-2016. MR3606396,
Show rawAMSref \bib{WalshDefin}{article}{ label={Wal17}, author={Walsh, Sean}, title={Definability aspects of the Denjoy integral}, journal={Fund. Math.}, volume={237}, date={2017}, number={1}, pages={1--29}, issn={0016-2736}, review={\MR {3606396}}, doi={10.4064/fm773-9-2016}, }

Article Information

MSC 2020
Primary: 03B30 (Foundations of classical theories (including reverse mathematics)), 03F35 (Second- and higher-order arithmetic and fragments), 03D78 (Computation over the reals, computable analysis), 26A39 (Denjoy and Perron integrals, other special integrals)
Author Information
Jordan Mitchell Barrett
Victoria University of Wellington, P.O. Box 600, Wellington, 6140, New Zealand
math@jmbarrett.nz
Homepage
MathSciNet
Rodney G. Downey
Victoria University of Wellington, P.O. Box 600, Wellington, 6140, New Zealand
rod.downey@vuw.ac.nz
Homepage
MathSciNet
Noam Greenberg
Victoria University of Wellington, P.O. Box 600, Wellington, 6140, New Zealand
greenberg@msor.vuw.ac.nz
Homepage
ORCID
MathSciNet
Additional Notes

The second and third authors were partially supported by the Marsden Fund of New Zealand. Many of the results in this paper are also contained in Barrett’s honours thesis.

Communicated by
Vera Fischer
Journal Information
Proceedings of the American Mathematical Society, Series B, Volume 9, Issue 12, ISSN 2330-1511, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2022 by the authors under Creative Commons Attribution-NonCommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/bproc/111
  • MathSciNet Review: 4405506
  • Show rawAMSref \bib{4405506}{article}{ author={Barrett, Jordan}, author={Downey, Rodney}, author={Greenberg, Noam}, title={Cousin's lemma in second-order arithmetic}, journal={Proc. Amer. Math. Soc. Ser. B}, volume={9}, number={12}, date={2022}, pages={111-124}, issn={2330-1511}, review={4405506}, doi={10.1090/bproc/111}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.