It is immediate that J0 = imageM. If δ > 0, then π−1(δ)< β. Then 〈Jβ, ∈〉 ⊨ (∃x) (∃s)φ(x, π−1(δ), s) and so 〈M, ∈〉 ⊨ (∃x)(∃s)φ(x, δ, s). Let xM and let sM such that 〈M, ∈〉 ⊨ φ(x, δ, s). By the absoluteness of φ(x, δ, s), any admissible structure image containing x, s and δ satisfies φ(x, δ, s). By Proposition 3.3.3, Jδ = xM.

Furthermore, for any xM, there exists a δ < β such that π−1(x)∈ Jδ. Hence 〈Jβ, ∈〉 ⊨ (∃δ)(∃y)(∃s)(π−1(x)∈ yφ(y, δ, s)). Then 〈M, ∈〉 ⊨ (∃δ)(∃y)(∃s)(xyφ(y, δ, s)). By Proposition 3.3.3 again, xJδʹ for some δʹ < γ. We conclude that JγM

Suppose x <Jβ π(x) for some xX. Take x0 to be the <Jβ-least such set. Then x0Jγ since π(x0)∈ Jγ and so there is an x1X such that π(x1)= x0. If x0 <Jβ x1, then π(x0)<Jα π(x1)= x0 which contradicts the assumption that x0 <Jβ π(x0). So x1 <Jβ π(x1)= x0 which contradicts the choice of x0 to be the Jβ-least counterexample.

An immediate consequence of the Condensation Theorem (3.3.7) is that V = L implies the Continuum Hypothesis. We leave the proof as an exercise.

Corollary 3.3.8. 2ωL = 2ωLω1.

Another important tool in the theory of constructibility is the use of Skolem functions. A relation or function is Σn(image)-definable if it is Σn-definable over image (possibly with parameters).

Definition 3.3.9. Given a structure image, a Σn-Skolem function for image is a Σn(image)-definable function h : ω × MM with parameter pM such that for any AM which is Σn-definable over image with parameters p, xM, there is an iω satisfying h(i, x)∈ A.

Thus a Σn-Skolem function is a Σn-definable choice function for Σn-sets.

Proposition 3.3.10. Suppose that α > 0 is admissible and h is a Σn-Skolem function for 〈Jα, ∈〉. Then

(i)for any xJα, x ∈ Range(h(ω ×{x})) and 〈Range(h(ω ×{x})), ∈〉 ≺ΣnJα, ∈〉;

(ii)if qJα and XJα is closed under ordered pairs, then X ∪{q} ⊆ Range(h(ω ×(X ∪ {q})) and 〈Range(h(ω ×(X ∪{q})), ∈〉 ≺ΣnJα, ∈〉.

Proof. Note that (i) is straightforward to verify and is a standard fact in model theory.

For (ii), since X is closed under ordered pairs, any finite sequence from X used to define a finite sequence in Range(h(ω ×(X ∪{q})) may be coded by a single parameter. Hence it is routine to show 〈Range(h(ω ×(X ∪{q})), ∈〉 ≺ΣnJα, ∈〉.

We remark that the existence of a Σn-Skolem function is not atrivial fact. Inmodel theory, defining a Skolem function for a structure image requires a well-ordering of image. In the constructible universe L, however, there is a uniformly Σ1-definable Skolem function.

Proposition 3.3.11. There is a Σ1-formula φ(x, y, i, z) such that for any ordinal α, the function

image

is a Σ1-Skolem function forJα, ∈〉.

Proof. By Proposition 3.3.3, there is a Σ1-formula φ<(u, v) defining a well-ordering <J of Jα for any admissible α. Clearly there is a Σ1-enumeration {φi(u, v, w)} of all Σ0-formulas. In other words, there is a Σ1-formula φ(x, y, i, z) so that 〈Jα, ∈〉 ⊨ φ(x, y, i, z) ⇔ 〈Jα, ∈〉 ⊨ φi(x, y, z). Define hα(i, x) to be y0 such that (y0, y1) is the <J-least pair with 〈Jα, ∈〉 ⊨ φi(y0, y1, x). By Proposition 3.3.3, hα(i, x) is Σ1-definable.

By the admissibility of α, using some tedious coding argument which may be found in [22], one can show that the relation “〈Jα, ∈〉 ⊨ φi(y0, y1, x)” is uniformly Δ1. Let hα(i, x)= y0 if and only if there exists a y1 such that 〈Jα, ∈〉 ⊨ φi(y0, y1, x) and there exists a z ={(z0, z1)|(z0, z1)<J (y0, y1)} such that for any pair (z0, z1)∈ z, 〈Jα, ∈〉 ⊨ ¬φi(y0, y1, x). Since z is uniformly Δ1-definable for any admissible α, we have hα(i, x) to be uniformly Σ1-definable for any admissible α.

If α is admissible, we call the Σ1-Skolem function hα the canonical Σ1-Skolem function for α.

Proposition 3.3.12. If α > ω is admissible, then αimage.

Proof. Suppose α is admissible and ω < image. By Theorem 3.3.6, Lα is admissible. By Theorem 1.3.4, there is a recursive function f such that for each nimage, Wf(n) = {〈i, j〉| i <o j <o n}. Fix a number nimage with |n|= α. Then the order type of Wf(n) is α and Wf(n)Lω+1Lα by Theorem 3.2.5. By Σ-induction there is a Σ1 order-preserving bijection g : Wf(n)α. Since Wf(n)Lα, αLα by the admissibility of α. This is a contradiction.

3.3.3 Exercises

Exercise 3.3.1. Prove thatLω1, ∈〉 ⊨ KP + “V = L”.

Exercise 3.3.2. Prove Corollary 3.3.8.

Exercise 3.3.3 (Putnam [115]). Prove that for any countable ordinal α, there is a countable ordinal β such that Lβ+α Lβ ∩ 2ω =0.

3.4 Projecta and master codes

In the interest of keeping the chapter within reasonable length and the proofs simpler, we present weaker versions of results in Jensen [55]. Nevertheless, the key ideas are extracted from [55] and the results are sufficient for recursion-theoretic applications.

3.4.1 Projecta

Definition 3.4.1. Let α > 0 and n ≥ 1. The Σn-projectum image, of α is the least ordinal ρ for which there is a Σn(Jα)-definable function f mapping a subset of Jρ onto Jα

Clearly image exists for any n ≥ 1. One may replace Jρ with ωρ in the above definition. To prove this, we need the following lemma whose proof is left to the reader:

Lemma 3.4.2. Let α > 0. For each 1 ≤ γωα, there is a Σ1(Jα)-definable bijection between ωα and γ × ωα.

Proposition 3.4.3. Let α > 0. There Is A Σ1(Jα)-definable function from ωα onto Jα.

Proof. Let f0 = (f 0, f 1) be a bijection between ωα and (ωα)2 guaranteed by Lemma 3.4.2 (setting γ = ωα) defined via the < -least parameter pJα). Define fn+1(γ)= (f 0(γ), fn(f 1(γ))) for γ < ωα. It is easy to verify that fn is Σ1(Jα)-definable with parameter p and is a bijection between ωα and (ωα)n+2. Let hα be the canonical Σ1-Skolem function for Jα. Set X = hα(ω ×(ωα ×{p})). Note that for each n < ω and γ < ωα, there is an in such that fn(γ)= hα(in, (γ, p)). Thus γX for every γ < ωα.

By Proposition 3.3.10, 〈X, ∈〉 ≺Σ1Jα, ∈〉. By Theorem 3.3.7, there is a βα and an isomorphism π : 〈X, ∈〉 ≅ 〈Jβ, ∈〉. Moreover, π(p) ≤ p. Since γX for every γ < ωα, we have β = α and π(γ)= γ for every γ < ωα. Hence π−1 : 〈Jα, ∈〉 ≺Σ1Jα, ∈〉 and π−1(f0)= f0. Thus f0 is also defined by π(p) in 〈Jα, ∈〉. By the assumption on p, p π(p) and so p = π(p). Then for any xX, there are i < ω and γ < ωα such that

x = hα(i, (γ, p)) = hα(π(i), (π(γ), π(p))) = π(x),

implying that X = Jα.

Note that hα is not necessarily total. Let φ(u, v, w, s) be a Σ0-formula such that hα(i, γ, p)= x if and only if 〈Jα, ∈〉 ⊨ (∃y)φ(i, γ, p, y). Then by the proof of Proposition 3.3.11, there is a total Σ1-function

image

Let h(γ)= hʹ(f1(γ)) for γ < ωα.

Applying Proposition 3.4.3, we have:

Proposition 3.4.4. α is admissible if and only if there is no Σ1(Jα)-definable function mapping a γ < ωα onto ωα.

The next lemma is a characterization of image.

Lemma 3.4.5. Suppose that α > 0 and ρ = image. Then AJρ for any γ < ρ and any Σ1(Jα)-definable AJγ.

Proof. If ρ = 1 then the conclusion is immediate. Suppose ρ > 1. Then ρ is a limit ordinal. Let γ < ρ and AJγ. Let hα be the canonical Σ1-Skolem function for Jα. Fix pJα such that A is Σ1(Jα)-definable with parameter p. Let X = hα(ω×(Jγ×{p})). Then 〈X, ∈〉 ≺Σ1Jα, ∈〉. By Theorem 3.3.7, there is a βα such that π : 〈X, ∈〉 ≅ 〈Jβ, ∈〉 is an isomorphism. Note that π(x)= x for all xJγ since Jγ is transitive. Hence π(x)= x for xA. Thus A is also Σ1-definable in Jβ with parameter π(p)∈ Jβ. We claim that β < ρ. Otherwise, let f be Σ1(Jα) mapping a subset of Jρ onto Jα. Then fπhα is Σ1(Jα) and maps a subset of ω ×(Jγ ×{p}) onto Jα. But ω × JγJγ+2Jρ, and so it maps a subset of Jγ+2 onto Jα, contradicting the choice of ρ.

Lemma 3.4.6. Let α > 0 and ρ = image. Then there is an AJρ which is Σ1(Jα)-definable and A ∉ Jα.

Proof. Let f be Σ1(Jα)-definable mapping a subset of Jρ onto Jα. Then the domain of f restricted to Jρ is Σ1(Jα) and not an element of Jα by admissibility.

Lemma 3.4.5 and Lemma 3.4.6 yield the following characterization of image:

Theorem 3.4.7. Let α > 0. Then image is the least ordinal ρ for which there is a Σ1(Jα) set AJρ not in Jα.

Now suppose that xω is Σ1(Jα)-definable with parameter p and x ∉ Jα. Let X = hα(ω ×(J1 ×{p}))). We may identify ω × J1 with J1. Hence X = h(J1 ×{p})) for some Σ1-function h. Then 〈X, ∈〉 ≺Σ1Jα, ∈〉. By transitive collapse there is a βα such that π : 〈X, ∈〉 ≅ 〈Jβ, ∈〉. Since xJ1, π(n)= n for all nω. Thus x is Σ1(Jβ)-definable with parameter π(p). If β < α, then xJα which is a contradiction. Thus β = α. Now π is Σ1(Jα)-definable and so πh is Σ1(Jα)-definable and maps a set yJ1 onto Jα. Since J1 = HF, by Proposition 3.2.6, we may assume that y is a real. Hence for any admissible α, if Jα+1 Jα contains a Σ1(Jα)-definable real, then it contains a real “coding Jα”.

Remark 3.4.8. Theorem 3.4.7 remains valid for n ≥ 1. However, the proof requires more than a straightforward generalization of what was presented above. The major difficulty is that there is no canonical Σn-Skolem function for n ≥ 2. To overcome this difficulty, a finer analysis of Jα is introduced. See Jensen [55] for details.

3.4.2 Master codes

Definition 3.4.9. Given an ordinal α and n ≥ 0, a Σn-master code for Jα is a Σn(Jα)-definable set Aimage such that for any m ≥ 1, any set Bimage, that is Σn+m(Jα)-definable is Σm-definable in 〈image, ∈, A〉.

There is an interpretation of the notion of a master code in the context of recursion theory. First suppose α = 1 and hence image = 1. Then Jα = Jρ = HF. Let A ⊆ HF be a Σ1-master code. By Proposition 3.2.6, there is a real x such that x is Δ1-definable in 〈HF, ∈, A〉 and A is Δ1-definable in 〈HF, ∈, x〉. By Theorem 3.2.5, every Σ1-definable real in 〈HF, ∈, x〉 is r.e. in x. Hence every real Σ2-definable in 〈HF, ∈〉 is Σ1-definable in 〈HF, ∈, A〉 and so r.e. in x. Thus the Σ1-master codes are precisely the reals that are r.e. and complete (hence Turing equivalent to imageʹ). In general, a real which is a Σn-master code in 〈J1, ∈〉 is Σn-complete and hence Turing equivalent to image(n). We may therefore consider master codes to be Turing jumps in the set-theoretic sense. Second, suppose that image = 1. A Σ1-master code for 〈Jα, ∈〉 is a set A ⊆ HF such that every Σ2(Jα)-definable real is Σ1 in 〈HF, ∈, A〉. Again by Proposition 3.2.6 and Theorem 3.2.5, every Σ2(Jα)-definable real is r.e. in some real x which is Σ1(Jα)-definable. In particular, every real zJα is recursive in x. Thus if α is much larger than image, the real x must be remarkably strong in computational power. Historically, the notion of a master code for ρ = 1 was first introduced by Boolos and Putnam [8] when they studied the Turing degrees of reals in the constructible universe. We now show that Σ1-master codes exist for any admissible ordinal α.

Theorem 3.4.10. Let α > 0. There is a Σ1-master code forJα, ∈〉, i.e. a Σ1(Jα)-definable Aimage such that for any Bimage if B is Σm+1(Jα)-definable, then it is Σm-definable inimage, ∈, A〉.

Proof. We only prove the case m = 1 for simplicity. Suppose that f is Σ1(Jα)-definable with parameter p mapping a subset of image onto α. Fix a Δ1-enumeration {φi(x, p)}i<ω of Σ1-formulas with parameter p. Let

image

It is obvious that A is Σ1-definable over 〈Jα, ∈〉. Now suppose that Bimage is Σ2-definable over 〈Jα, ∈〉. Then there is a Σ1(Jα)-formula φ(u, v, w) such that

image

where s is a parameter in Jα. Since f is onto, there is an s0image such that f(s0)= s.

Then

image

Since f is Σ1(Jα)-definable with parameter p, there is an index i such that for x, yimage

image

Thus

image

Hence B is Σ1-definable over 〈image, ∈, A〉.

As an example, imageʹ is a Σ1(J1)-master code since it satisfies the property described in Theorem 3.4.10. Jensen [55] proved the following general result:

Theorem 3.4.11. For any ordinal α > 0 and n > 0, there is a Σn(Jα)-master code Aimage. Moreover, image is the least ordinal γ for which there is a Σn(Jα) Aωγ not in Jα.

A Σn-master code corresponds to image(n) in recursion theory. The equality image(n+m) =(imagen)(m) has a natural analog in fine structure theory. See [55] for details.

3.4.3 Exercises

Exercise 3.4.1. Prove Lemma 3.4.2.

Exercise 3.4.2. Prove Proposition 3.4.4.

Exercise 3.4.3. Prove that there is a countable admissible ordinal α > ωsuch that image = α.

3.5 ω-models

Definition 3.5.1. A structure image is an ω-model if ωM = ω.

ω-models are widely used in recursion theory and descriptive set theory. A major point of interest in an ω-model is its standard part.

Definition 3.5.2. Let image =〈M, E〉 be a structure and M1M. Then M1 is E-closed if for every xM1, yEx and yM implies yM1. The standard part of image is the maximal E-closed wellfounded subset of M.

Obviously if M1M is E-closed, then for any Σ0-formula φ(v) and xM1, 〈M, E〉 ⊨ φ(x) if and only if 〈M1, E〉 ⊨ φ(x).

Lemma 3.5.3. Suppose that image ⊨ KP. Then x belongs to the standard part of M if and only if rk(x), the rank of x, belongs to the standard part of image.

Proof. Since image ⊨ KP, by the Σ-induction Theorem (3.1.8) the rank function rk is Δ1-definable.

If x belongs to the standard part of M, then by induction on the rank of x, it is straightforward to verify that rk(x) also belongs to the standard part.

If x does not belong to the standard part of M, then there is a descending sequence {xn}n<ω in M so that xn+1Exn for every n and x0Ex. Then rk(xn+1)Erk(xn) for every n and rk(x0)Erk(x). So rk(x) does not belong to the standard part of M either.

Lemma 3.5.4. Suppose that image ⊨ KP and M1M is the standard part of M. Then 〈M1, E〉 ⊨ KP.

Proof. Since M1M is E-closed, the extensionality axiom holds in image1 =〈M1, E〉. image1 satisfies the foundation axiom since M1 is the standard part of M. Similarly it is straightforward to verify that pairing, union and the Δ0-comprehension axioms hold in image1. We prove image1Δ0-collection.

Suppose that φ(u, v) is a Δ0-formula and aM1 such that

M1, E〉⊨(∀xEa)(∃y)φ(x, y).

Since φ(u, v) is Δ0 and M1 is an E-closed subset of M, we have

M, E〉⊨(∀xEa)(∃y)φ(x, y).

Now 〈M, E〉 ⊨ KP implies that there is a bM such that

M, E〉⊨(∀xEa)(∃yEb)φ(x, y).

By Lemma 3.5.3, there is a b0 such that rk(b0)= α0 is the least ordinal in image to satisfy the collection axiom for φ. If α0M1, then by Lemma 3.5.3, b0M1 and we are done. Otherwise, α0M M1. But then there is a nonstandard α10 such that for every xEa, there is a yEb0 in M1 with rk(y)1 and 〈M, E〉 ⊨ φ(x, y). Note that φ(x, y)∧ rk(y)1 is a Σ-formula and

M, E〉⊨(∀xEa)(∃y)(φ(x, y)∧ rk(y)1).

By the Σ-collection Theorem (3.1.6), there is a b1M such that

M, E〉⊨(∀xEa)(∃yEb1)φ(x, y)∧ rk(y)1).

Define

c ={y | yEb1 ∧〈M, E〉 ⊨ rk(y)1 ∧∃xEa(φ(x, y))}.

By Theorem 3.1.6, cM.

Then rk(c)0 and α0 = rk(b0). Since α0 is the least ordinal in image to satisfy the collection axiom for φ, we have a contradiction.

Proposition 3.5.5. Let image ⊨ KP be an ω-model. Then image.

Proof. Let M1M be the standard part of M. By Lemma 3.5.4, 〈M1, E〉 ⊨ KP. Hence 〈M1, E〉 is also an ω-model of KP. Since ω∪{ω}⊆ M1, by Δ0-separation every recursive set belongs to 〈M1, E〉. By Σ-induction, imageM1. Let π : 〈M1, E〉→〈N, ∈〉 be the collapsing map. Then 〈N, ∈〉 is admissible so that imageN. By Theorem 3.3.4, imageN. We prove by induction on β < image that π−1image is the identity map.

Note that by Proposition 3.3.3 there exists a Δ1-formula φ(u, v) such that for any ordinal β < image, 〈N, ∈〉 ⊨ φ(β, x) if and only if x = Jβ. Moreover,

N, ∈〉 ⊨ φ(β, Jβ)∧(∀xJβ)(∃γβ)(∃y)(xyφ(γ, y)).

We prove by induction that π−1(Jγ)= Jγ for every γ < image.

Since imageM1, we have π(β)= β for any β < image. Thus π(x)= x for xJ1. Since 〈N, ∈〉 ⊨ φ(1, x) if and only if x = J1, we have 〈M1, E〉 ⊨ φ(1, x) if and only if x = J1. So π−1(J1)= J1.

For a successor ordinal β + 1. π clearly preserves the Gödel primitive recursive operators. Hence Range(π−1Jβ+1)= Jβ+1. Since 〈N, ∈〉 ⊨ φ(β + 1, x) if and only if x = Jβ+1, we have 〈M1, E〉 ⊨ φ(β + 1, x) if and only if x = Jβ+1. So π−1(Jβ+1)= Jβ+1.

Now suppose β < image is a limit ordinal. By induction, if γ < β then 〈M1, E〉 ⊨ φ(γ, x) if and only if x = Jγ. Let z = π−1(Jβ). Since π−1(β)= β, we have

M1, E〉 ⊨ φ(β, z)∧(∀xEz)(∃γEβ)(∃y(xEy)∧ φ(γ, y)).

Thus z = Jβ. By the transitivity of image, we conclude that π−1image is the identity map so that imageM

3.5.1 Exercise

Exercise 3.5.1. Prove that ifM, E〉 ⊨ KP is a nonstandard ω-model, then for any nonstandard ordinal αM, there is a set A ⊆ {β | 〈M, E〉 ⊨ βEα} such that (image, <) is isomorphic toA, E, where image is the set of rational numbers.

3.6 Coding structures

The method of coding a language or a structure by a set of natural numbers was introduced by Gödel in the proof of the Incompleteness Theorem. Coding is widely used today in mathematical logic. In this section, we will not discuss the formalism of coding. Instead, we present a general description of how the method works.

3.6.1 Coding a language

Given a countable language image, there are many ways of coding it by a set of numbers. What we want is a “nice” coding scheme. In many cases, one may code a language image by a recursive set imageω. A language that may be thus coded is called a recursive language. For example, both the first order language of arithmetic and the language of Zermelo–Fraenkel set theory are recursive. Moreover, in a recursive language, the set of well-formed formulas and the set of sentences are recursive. We may effectively associate each formula φ with its Gödel numberφ⌝. We say that a theory T is recursive if its corresponding set of Gödel numbers image is recursive. The theories considered in this book, such as KP, ZFC, are recursive.

Under coding with Gödel numbers, a proof sequence is identified with a finite sequence of numbers, and may therefore be considered as a finite subset of ω. Moreover, if both image and T are recursive, then so is the set of proofs in T. It follows that the set of sentences which are theorems of T is r.e.

3.6.2 Coding a structure

Given a recursive language image, one may associate with it a recursive set image and code a countable structure in image by the set of natural numbers ω with relations in image defined over ω, so that the resulting structure is isomorphic to the original structure. If image = 〈M, R0, …, Rn〉 is countable, a coding image of image is a structure such that imageimage. If all the relations image are recursive, then the structure image is said to be recursive. However, very few structures are recursive.

Let image be a recursive language and image a structure coded by image. For a formula ̂φ(u) in image corresponding to the formula φ(u) in image, we may compute the complexity of the set image. The procedure to calculate the complexity goes as follows: Recursively decode image to obtain the formula φ(u). Then the complexity of A is the complexity of φ(u). For example, suppose image is the language of set theory and image corresponds to the formula φ(w): (∀u)(∀v)(uvvwuw) (“w is transitive”). Then the set

image

is image.

In general, given a language image with relation symbols R0,…, Rn, a structure and a formula φimage, the set {mω |imagêφ(m)} is arithmetical in ̂R0 ⊕ … ⊕ ̂Rn. Moreover, this calculation is uniform. Hence if {φi}i<ω is recursive, then the set {(i, m)| imagêφi(m)} is Turing computable in ( ̂R0 ⊕ … ⊕ ̂Rn)(ω) via an algorithm with index e. In particular, there is an e such that if {φi}i<ω is a recursive set of sentences, then imagêφi if and only if Φ(̂R0⊕…⊕ ̂Rn)(ω) (i)= 1. As a consequence, we have the following:

Proposition 3.6.1. Given a recursive theory T, the set {image|(∀φ)(φTimagêφ)} is image.

For example, {image| image ⊨ KP + “V = L”} is image (where for simplicity we make no distinction between KP+“V = L” and the corresponding sentences interpreted in image).

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset