Between Turing and Kleene
Between Turing and Kleene
Sam Sanders
1 Between Turing and Kleene computability . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Short summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Extending the scope of Turing computability . . . . . . . . . . . . . . . . . . 2
1.2.1 A new notion of reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.2 Continuous and discontinuous functionals . . . . . . . . . . . . . . . 5
1.3 The need for an extension of Turing computation . . . . . . . . . . . . . . 7
1.3.1 Computing with second-order representations . . . . . . . . . . . . 8
1.3.2 On higher-order computation . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2 Some results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1 Nets and computability theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.1 Nets, a very short introduction . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.2 Nets and convergence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.3 Nets and compactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 On the uncountability of R . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Discontinuous functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3.1 Grilliot’s trick . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3.2 Discontinuous functions and N -reduction . . . . . . . . . . . . . . . 17
Between Turing and Kleene
Sam Sanders
where NN is the Baire space and N is the set of natural numbers. Indeed, as
discussed in Remark 1, some basic primitive recursive operations relegate the
coding of real numbers (via elements of NN ) to the background. Moreover, a list
Between Turing and Kleene 3
of theorems that can be brought in the form (1) can be found in Example 2
below, while we discuss the scope of theorems that can be brought in this form
at the end of this section and in Section 1.2.2.
Secondly, to improve readability, one often uses type theoretic notation in
(1), i.e. n0 for type 0 objects n ∈ N, x1 for type 1 objects x ∈ NN , and Y 2
for type 2 objects Y : NN → N. We will only occasionally need type 3 objects,
which map type 2 objects to natural numbers. We generally use Greek capitals
Θ3 , Λ3 , . . . for such objects.
Thirdly, to compare the logical strength of theorems of the form (1), one
establishes results of the following form over weak systems:
Next, we list some theorems that have been studied via the above paradigm
based on (3) and S1-S9.
Example 2 (Some representative theorems)
– The Lindelöf, Heine-Borel, and Vitali covering theorems involving uncount-
able coverings ([29, 30, 34]),
– The Lebesgue number lemma ([33, 34]),
– The Baire category theorem ([33]),
– Convergence theorems for nets ([39, 40, 42]),
– Local-global principles like Pincherle’s theorem ([34]),
– The uncountability of R and the Bolzano-Weierstrass theorem for countable
sets in Cantor space ([35, 37]),
– Weak fragments of the Axiom of (countable) Choice ([36]).
– Basic properties of functions of bounded variation, like the Jordan decom-
position theorem ([38]).
Many more theorems are equivalent -in the sense of higher-order RM as in [20]-
to the theorems in the above list, as can be found in the associated references.
Fourth, for all the reasons discussed in Section 1.3, we formulate a version of (3)
based on Turing computability as follows:
where s2→1 , t2→2 are terms of Gödel’s T and ‘{e}X ’ is the e-th Turing machine
with oracle X ⊂ N. We note that (4) readily1 implies (3); we discuss the gener-
ality of (4) at the end of this section.
In line with the nomenclature of computability theory, we call the antecedent
and consequent of (2) ‘problems’ and say that
solving the problem (∀Z 2 )(∃y 1 )B(Z, x) N -reduces to solving the problem
(∀Y 2 )(∃x1 )A(Y, x)
in case (4) holds for the parameters mentioned. We view the N -reduction relation
as ‘neutral’ between the Turing and Kleene framework and the reader readily
verifies that N -reduction is transitive. In case the term s(Z, x) can be replaced
by a term u(x), i.e. the latter has no access to Z, we refer to (4) as strong
N -reduction.
Finally, the critical reader may wonder about the generality of (4). The latter
is quite general, for the following two reasons.
– It is an empirical observation based on [29–38] that positive results in S1-
S9 computability theory can be witnessed by terms of Gödel’s T of low
complexity. In this light, there is no real loss of generality if we use terms of
Gödel’s T as in (4).
– A theorem of (third-order) ordinary mathematics generally has the form
(1), unless the former implies the existence of a discontinuous function on
R. In the latter case, an ‘indirect’ treatment is still possible via the so-called
Grilliot’s trick, which we sketch in Section 2.3.1.
Like the reader, we feel that the second item deserves a more detailed expla-
nation, which is in Section 1.2.2. Regarding the first item, intellectual honesty
compels us to admit that many of our S1-S9 results are witnessed by terms of
Gödel’s T additionally involving Feferman’s search operator (already found in
Hilbert-Bernays [16]) defined for any f 1 as:
(
the least n0 such that f (n) = 0 (∃m0 )(f (m) = 0)
µ(f ) := . (5)
0 otherwise
Now, NFP proves the Lindelöf lemma ([48]), inspiring Theorem 12.
Finally, we can combine the above as follows: the case distinction from the
beginning of this section distinguishes between whether a given theorem T in
the language of third-order arithmetic implies the existence of a discontinuous
function (on R or 2N ), or not. It is then an empirical observation based on [29–38]
that for this theorem T, either the theorem T implies the existence of ∃2 via the
aforementioned Grilliot’s trick, or T is provable from a fragment of NFP where
A may include third-order parameters.
In the former case, the theorem T can be analysed ‘indirectly’ using N -
reduction, namely via Grilliot’s trick, as discussed in Section 2.3. In case the the-
orem T is provable from a fragment NFP (with third-order parameters), we can
generally bring T in the form (1) and hence analyse it directly via N -reduction.
In light of (8), fragments of NFP can always be analysed via N -reduction.
In conclusion, second-order comprehension has been generalised to higher
types in two (more-or-less-known ways) ways, namely as follows.
– Formulate ‘characteristic functionals’ like ∃2 from (6) that decide the truth
of certain formulas.
– Formulate NFP as in Definition 4 for formulas involving higher-order param-
eters and variables.
Between Turing and Kleene 7
Finally, we show that NFP classically follows from comprehension and vice
versa, assuming a fragment of the induction axiom.
Regarding the second item, a Jordan realiser is therefore not computable from
(finite iterations of) ∃2 , the higher-order counterpart of the Turing jump. The
same holds for S2k , which is a type two functional that can decide Πk1 -formulas
(involving first- and second-order parameters). The usual proof of the Jordan
decomposition theorem implies that Kleene’s ∃3 computes a Jordan realiser. But
∃3 implies full second-order arithmetic, and the same holds for the combination
of all S2k .
In conclusion, there is a huge difference in the computational hardness of the
Jordan decomposition theorem depending on whether we use representations or
not. However, this theorem deals with functions of bounded variation, a class
‘very close’ to the class of continuous functions. Hence, (Turing) computing with
3
Examples of such frameworks are: reverse mathematics ([43, 46]), constructive analy-
sis ([3, I.13], [5]), predicative analysis ([13]), and computable analysis ([52]). Bishop’s
constructive analysis is not based on Turing computability directly, but one of its
‘intended models’ is (constructive) recursive mathematics, as discussed in [7]. One
aim of Feferman’s predicative analysis is to capture Bishop’s approach.
4
The fan functional constitutes an early natural example of this difference: it has a
computable code but is not S1-S9 computable (but S1-S9 computable in Kleene’s ∃2
from Section 1.2.2). The fan functional computes a modulus of uniform continuity
for continuous functions on Cantor space; details may be found in [24].
Between Turing and Kleene 9
due to Vitali, Heine-Borel, and Lindelöf. However, all these functionals are not
S1-S9 computable in any type two functional, i.e. the former are ‘hard to com-
pute’ (see [29,31,34]). As a result, ITTMs yield ‘too strong’ a baseline framework
for our purposes.
2 Some results
2.1.1 Nets, a very short introduction Nets are the generalisation of the
concept of sequence to possibly uncountable index sets, nowadays called nets
or Moore-Smith sequences. These were first described in [25] and then formally
introduced by Moore and Smith in [26] and by Vietoris in [51]. These authors also
established the generalisation to nets of various basic theorems due to Bolzano-
Weierstrass, Dini, and Arzelà ([26, §8-9] and [51, §4]).
One well-know application is the formulation of fundamental topological no-
tions like compactness in terms of nets, as pioneered in [4], while Kelley’s text-
book [17] is standard. Tukey’s monograph [49] builds a similar framework, based
on very specific nets, called phalanxes, where the index sets consist of finite
subsets ordered by inclusion. We now list some basic definitions.
Now, we shall mostly use nets where the index set consists of finite sets of real
numbers ordered by inclusion, i.e. Tukey’s ‘phalanxes’ from [49]. As noted in
Remark 1, real numbers can readily be represented via elements of Baire space
using primitive recursive operations. Thus, such phalanxes are essentially nets
indexed by NN . The notion of ‘sub-sequence’ of course generalises to ‘sub-net’
(see e.g. [42]), but we do not need this (slightly technical) notion here.
Finally, we discuss an alternative to nets and why it is not suitable here.
Remark 9 (Nets and filters) For completeness, we discuss the intimate con-
nection between filters and nets. Now, a topological space X is compact if and
only if every filter base has a refinement that converges to some point of X,
which follows by [2, Prop. 3.4].
Whatever the meaning of the previous italicised notions, the similarity to the
Bolzano-Weierstrass theorem for nets is obvious, and not a coincidence: for every
net r, there is an associated filter base B(r) such that if the erstwhile converges,
so does the latter to the same point; one similarly associates a net r(B) to a
given filter base B with the same convergence properties (see [2, §2]).
Hence, filters provide an alternative to nets, but we have chosen to work with
nets for the following reasons, where the second one is the most pressing.
– Nets have a greater intuitive clarity compared to filters, in our opinion, due
to the similarity between nets and sequences.
– Nets are ‘more economical’ in terms of ontology: consider the aforementioned
filter base B(r) associated to the net r. By [2, Prop. 2.1], the base has strictly
higher type than the net. The same holds for r(B) versus B.
– The notion of refinement mirrors the notion of sub-net ([2, §2]). The former
is studied in [41] in the context of paracompactness; the associated results
suggest that the notion of sub-net works better in weak systems.
12 S. Sanders
We now have Theorem 10 where C is Cantor space ordered via the lexicographic
ordering ≤lex , i.e. the notion of ‘increasing net in C’ is obvious following Defi-
nition 8. We note that subsets of NN or R are given by characteristic functions,
well-known from measure and probability theory and going back one hundred
plus of years ([12]).
Proof. To show that the second item strongly N -reduces to the first one, let
fd : D → C be an increasing net in C indexed by Baire space and consider the
∗
formula (∃d ∈ D)(fd ≥lex σ ∗ 00 . . . ), where σ 0 is a finite binary sequence. The
latter formula is equivalent to a formula of the form (∃g 1 )(Y (g, n) = 0) where
Y has the form t(λd.fd , n) for a term t of Gödel’s T . Now use J(Y ) to define
the limit f = limd fd , as follows: f (0) is 1 if (∃d ∈ D)(fd ≥lex 100 . . . ) and zero
otherwise. One then defines f (n + 1) in terms of f n in the same way. Note that
we only used J(Y ) to define f , i.e. we have a strong N -reduction.
∗
For the remaining case, fix some Y 2 and let w1 be a sequence of ele-
ments in NN . Define fw : D → C as fw := λk.F (w, k) where F (w, k) is 1 if
∗
(∃i < |w|)(Y (w(i), k) = 0), and zero otherwise. Then λw1 .fw is a monotone net
(phalanx) in C indexed by Baire space (modulo coding). In case limw fw = f ,
then it is readily verified that:
The reader is warned that not all N -reduction results are as elegant.
Between Turing and Kleene 13
Here, B(w) is the left-most end-point in [0, 1] of the intervals of the form
B(w(i), Ψ (B(w(i)))) for i ≤ k that is not covered by the union. Note that B(w)
∗
and xw are readily defined using µ2 . Modulo coding of reals, λw1 .xw can be
viewed as a monotone net (phalanx) indexed by Baire space and we must have
limw xw = 1. If (wk )k∈N is a modulus of convergence, then |xw2 − 1| < 41 by
definition, implying xw2 = 1. Hence, ∪i<|w2 | B(w(i), Ψ (w(i))) covers [0, 1] ∩ Q.
Now adjoin to w2 all the points w2 (i) ± Ψ (w2 (i)) for i < |w2 |, to obtain a cover-
ing of [0, 1]. This ‘adjoining’ takes the form of s(Ψ, w2 ) while xw takes the form
t(Ψ, µ2 )(w) for terms s, t of Gödel’s T , using the notation from (4).
For the second part, replace the output 1 by 34 + 2N1+3 in the first case of xw ,
where N is as follows: adjoin to w all the points w(i) ± Ψ (w(i)) for i < |w|, to
5
The notion of Lebesgue number is familiar from topology (see e.g. [27, p. 175]) and
amounts to the following: for a metric space (X, d) and an open covering O of X,
the real number δ > 0 is a Lebesgue number for O if every subset Y of X with
diam(Y ) := supx,y∈Y d(x, y) < δ is contained in some member of the covering.
6
A modulus of convergence for a net xd : D → R with limd xd = x is a sequence
(dk )k∈N with (∀k ∈ N)(∀d dn )(|xd − x| < 21k ).
14 S. Sanders
obtain a covering of [0, 1]. Now use µ2 to find N ∈ N such that 21N is a Lebesgue
number for the latter covering. Note that the modified net is still monotone
as extending w can only increase the associated Lebesgue number. Clearly, any
cluster point of the modfied net is found in ( 34 , 1). A straightforward unbounded
search can now recover a Lebesgue number from the cluster point of the net
without access to Ψ , i.e. we have a strong µN -reduction. ⊓
⊔
In light of the first part of the previous proof, the ‘post-processing’ term s in (4)
seems necessary as a Turing machine cannot evaluate a third-order functional
at a given point due to type restrictions.
(∀Y 2 ) (∀n0 )(∃f 1 )(Y (f, n) = 0) → (∃Z 0→1 )(∀n0 )(Y (Z(n), n) = 0) ,
where we exclude the trivial case (∃f 1 )(∀n0 )(Y (f, n) = 0).
Theorem 12 The Lebesgue number lemma strongly N -reduces to NFP for A(n) ≡
(∃f 1 )(Y (f, n) = 0) for any Y 2 .
which merely expresses that for every x ∈ [0, 1], there is n ∈ N and y ∈ [0, 1]
such that B(x, 21n ) ⊂ B(y, Ψ (y)). Applying NFP with parameter Ψ , we obtain
γ ∈ K0 such that
1
(∀f ∈ NN ) (∃g ∈ NN )[B(r(f ),
2γ(f )
) ⊂ B(r(g), Ψ (r(g)))] .
Now compute an upper bound for γ on 2N , using the Kleene associate for the fan
functional ([24, §8.3.2]). This upper bound yields the required Lebesgue number,
which only depends on γ 1 , not on Ψ , i.e. we have obtained a strong N -reduction.
⊓
⊔
We conjecture that HBU does not strongly µN -reduce to the fragment of NFP
from Theorem 12.
Between Turing and Kleene 15
A trivial manipulation of definitions shows that NIN and Cantor’s theorem are
logicially equivalent. We however have the following theorem and associated
Conjecture 14.
Theorem 13
Indeed, in case there are no i, j ≤ k as in (12), then the measure of ∪i≤k B(xi , t(Z)(xi ))
Pk 1
is at most n=0 2i+1 < 1, contradicting the fact that ∪i≤k B(xi , t(Z)(xi )) cov-
ers [0, 1]. In light of (12), given the finite sequence s(Z, x0 , . . . , xk ) defined as
x0 , t(Z)(x0 ), . . . , xk , t(Z)(xk ), we can perform an unbounded search (on a Tur-
ing machine) to find i, j ≤ k and k ∈ N such that t(Z)(xi ) =Q t(Z)(xj ) and
[|xi − xj |](k) >Q 21k , where [z](m) is the approximation of z ∈ R up to 2m+1 1
.
Hence, we also obtain the consequent of (4) for the case at hand.
For the second part, fix A ⊂ [0, 1] and Y : [0, 1] → N such that Y is injective
on A. Now consider the following:
(
1
Y (x)+5 x∈A
t(Y, A)(x) := 12 .
8 x 6∈ A
One readily shows that t(Y, A) : R → R is Baire class 2, as it only has countably
many points of discontinuity by definition. For a finite sub-covering x0 , . . . , xk ∈
[0, 1] of ∪x∈[0,1] B(x, t(Y, A)(x)), there must be j ≤ k, with xj 6∈ A. Indeed, as
in the previous paragraph, the measure of ∪i≤k B(xi , t(Y, A)(xi )) is otherwise
Pk 1
at most n=0 2i+5 < 1, a contradiction. One can effectively decide whether
1 1
t(Y, A)(xi ) < 8 or t(Y, A)(xi ) > 16 for i ≤ k, i.e. one readily finds a j ≤ k with
xj 6∈ A. ⊓
⊔
16 S. Sanders
In light of the previous proof, the ‘post-processing’ term s in (4) again seems
necessary as a Turing machine cannot evaluate a third-order functional at a
point due to type restrictions.
Based on the previous proof, we conjecture the following.
Conjecture 14 The problem NIN does not N -reduce to the Heine-Borel theo-
rem HBU restricted to Baire class 2 functions, nor to the (full ) Lebesgue number
lemma.
be brought in the form7 ‘(∃Y 2 )(∀α1 )’, which can also be obtained by representing
continuous functions via second-order codes.
for some index e ∈ N and term s of Gödel’s T , which is exactly (4). The details
are somewhat tedious, but we nonetheless can say that the negation of (13)
N -reduces to the negation of (∃2 ).
Finally, the usual ‘interval-halving’ proof of the existence of a maximum of a
continuous function on [0, 1], can be done using ∃2 , yielding a term t of Gödel’s
T such that:
(∀E 2 ) (∀f 1 )B(E, f ) → (∀g)A(g, t(E)(g)) .
(17)
8
The construction of a discontinuous function on R in the proof of [20, Prop. 3.14]
depends on whether ε(g0 ) ∈ [0, 21 ] or ε(g0 ) ∈ [ 12 , 1], where g0 is the constant 0
function and ε as in (13). This non-effective case distinction can be replaced by an
effective case distinction whether ε(g0 ) < 34 or ε(g0 ) > 14 . The proof in the first case
goes through unmodified, while one replaces yx in the second case by −yx.
9
The join of the sequences (qx − q)q∈Q∩[0,1] and (−qx)q∈Q∩[0,1] suffices.
18 S. Sanders
The contraposition of (17) then has the same form as (15). One readily obtains
an index e ∈ N and term s of Gödel’s T with
as one only needs to decide g(r) ≥ g(q) for r, q ∈ [0, 1] ∩ Q to find a maximum
of a (Lipschitz) continuous function g : [0, 1] → R. Hence, an unbounded search
on a Turing machine will find f 1 with ¬B(E, f ). We note that (18) is a case of
N -reduction of the negation of (∃2 ) to the negation of (13).
In conclusion, we observe that the negation of (∃2 ) will N -reduce to the
negation of (13), and vice versa. Thus, it perhaps makes sense to drop the
‘negation of’ here and distinguish between (1) and its negation in the definition
of N -reduction.
Bibliography
[1] Sanjeev Arora and Boaz Barak, Computational complexity. A modern approach.,
Cambridge University Press, 2009.
[2] Robert G. Bartle, Nets and filters in topology, Amer. Math. Monthly 62 (1955),
551–557.
[3] Michael J. Beeson, Foundations of constructive mathematics, Ergebnisse der Math-
ematik und ihrer Grenzgebiete, vol. 6, Springer, 1985.
[4] Garrett Birkhoff, Moore-Smith convergence in general topology, Ann. of Math. (2)
38 (1937), no. 1, 39–56.
[5] Errett Bishop, Foundations of constructive analysis, McGraw-Hill, 1967.
[6] E. Borel, Leçons sur la théorie des fonctions, Gauthier-Villars, Paris, 1898.
[7] Douglas Bridges and Fred Richman, Varieties of constructive mathematics, Lon-
don Mathematical Society Lecture Note Series, vol. 97, Cambridge University
Press, Cambridge, 1987.
[8] L. E. J. Brouwer, Collected works. Vol. 1, North-Holland Publishing Co., Amster-
dam, 1975. Philosophy and foundations of mathematics; Edited by A. Heyting.
[9] Wilfried Buchholz, Solomon Feferman, Wolfram Pohlers, and Wilfried Sieg, Iter-
ated inductive definitions and subsystems of analysis, LNM 897, Springer, 1981.
[10] Georg Cantor, Ueber eine Eigenschaft des Inbegriffs aller reellen algebraischen
Zahlen, J. Reine Angew. Math. 77 (1874), 258–262.
[11] Pierre Cousin, Sur les fonctions de n variables complexes, Acta Math. 19 (1895),
1–61.
[12] Lejeune P. G. Dirichlet, Über die Darstellung ganz willkürlicher Funktionen durch
Sinus- und Cosinusreihen, Repertorium der physik, von H.W. Dove und L. Moser,
bd. 1, 1837.
[13] Solomon Feferman, How a Little Bit goes a Long Way: Predicative Foundations
of Analysis, 2013. unpublished notes from 1977-1981 with updated introduction,
https://math.stanford.edu/∼feferman/papers/pfa(1).pdf .
Between Turing and Kleene 19
[39] Sam Sanders, Nets and Reverse Mathematics: initial results, Lecture notes in Com-
puter Science 11558, Proceedings of CiE19, Springer (2019), 253-264.
[40] , Reverse Mathematics and computability theory of domain theory, Lecture
notes in Computer Science 11541, Proceedings of WoLLIC19, Springer (2019),
550-568.
[41] , Reverse Mathematics of topology: dimension, paracompactness, and split-
tings, Notre Dame Journal for Formal Logic 61 (2020), no. 4, 537-559.
[42] , Nets and Reverse Mathematics: a pilot study, Computability 10 (2021),
no. 1, 31-62.
[43] Stephen G. Simpson, Subsystems of second order arithmetic, 2nd ed., Perspectives
in Logic, CUP, 2009.
[44] Robert I. Soare, Recursively enumerable sets and degrees, Perspectives in Mathe-
matical Logic, Springer, 1987.
[45] Ernst Specker, Nicht konstruktiv beweisbare Sätze der Analysis, J. Symbolic Logic
14 (1949), 145–158 (German).
[46] J. Stillwell, Reverse mathematics, proofs from the inside out, Princeton Univ.
Press, 2018.
[47] A. S. Troelstra, Choice sequences, Clarendon Press, Oxford, 1977. A chapter of
intuitionistic mathematics; Oxford Logic Guides.
[48] Anne Sjerp Troelstra and Dirk van Dalen, Constructivism in mathematics. Vol.
I, Studies in Logic and the Foundations of Mathematics, vol. 121, North-Holland,
1988.
[49] John W. Tukey, Convergence and Uniformity in Topology, Annals of Mathematics
Studies, no. 2, Princeton University Press, Princeton, N. J., 1940.
[50] Alan Turing, On computable numbers, with an application to the Entscheidungs-
problem, Proceedings of the London Mathematical Society 42 (1936), 230-265.
[51] Leopold Vietoris, Stetige Mengen, Monatsh. Math. Phys. 31 (1921), no. 1, 173–204
(German).
[52] Klaus Weihrauch, Computable analysis, Springer-Verlag, Berlin, 2000.
[53] P. D. Welch, Transfinite machine models, Turing’s legacy: developments from Tur-
ing’s ideas in logic, Lect. Notes Log., vol. 42, Assoc. Symbol. Logic, La Jolla, CA,
2014, pp. 493–529.