Exercise sessions for the Set theory course, spring semester 2021 Giulio Lo Monaco Week 1 Definition Given some symbols p, q, . . . and a (non-quantified) formula containing the given symbols and →, ¬, =, giving a truth table means to specify the truth values of the whole formula for each possible assignment of truth values on the symbols p, q, . . .. We say that two formulas are equivalent if they have the same truth tables. Example The formula p → q has the following truth table: p q p → q T T T T F F F T T F F T Recall that we can shorten ¬p → q into p ∨ q and ¬(p → ¬q) into p ∧ q. Let us compute their truth tables and see that they correspond with the expected ones, that is, p ∨ q is false if and only if both p and q are false, and p ∧ q is true if and only if both p and q are true. p q ¬p ¬p → q T T F T T F F T F T T T F F T F p q ¬q p → ¬q ¬(p → ¬q) T T F F T T F T T F F T F T F F F T T F Exercise Find the truth table of the formula ¬(p ∨ q) ∨ (¬p ∧ q). Solution We can build it one step at a time in the following table p q p ∨ q ¬(p ∨ q) ¬p ¬p ∧ q ¬(p ∨ q) ∨ (¬p ∧ q) T T T F F F F T F T F F F F F T T F T T T F F F T T F T 1 Exercise Verify if the two following formulas are equivalent: ¬(p ∧ q) ∧ (¬q ∧ ¬r) ¬p ∧ ¬q ∧ ¬r. Solution We need to calculate the truth tables of both formulas: p q r ¬q p ∧ ¬q ¬(p ∧ ¬q) ¬r ¬q ∧ ¬r ¬(p ∧ ¬q) ∧ (¬q ∧ ¬r) T T T F F T F F F T T F F F T T F F T F T T T F F F F T F F T T F T T F F T T F F T F F F F T F F F T T F F F F T T F T F F F F F F T F T T T T p q r ¬p ¬q ¬r ¬p ∧ ¬q ∧ ¬r T T T F F F F T T F F F T F T F T F T F F T F F F T T F F T T T F F F F T F T F T F F F T T T F F F F F T T T T so they are equivalent, because their truth values always coincide. Exercise Suppose we are given a language with a binary relation symbol R. Express in formulas the following properties: 1. R is a function; 2. R is a bijection; 3. R is an equivalence relation; 4. R is a linear order on a set with three elements. Solution We will have to be specifically careful to all instances of the phrase “it exists a unique x such that φ(x)”. The usual way to express this is ∃!xφ(x) but using the symbols of predicate logic this has to be written ∃xφ(x) ∧ ∀y(φ(y) → y = x). Now let us proceed with the solution 2 1. ∀x(∃yRxy ∧ ∀z(Rxz → z = y)); 2. ∀x(∃yRxy ∧ ∀z(Rxz → z = y)) ∧ ∀y(∃xRxy ∧ ∀w(Rwy → w = x)); 3. ∀x∀y∀z(Rxx) ∧ (Rxy → Ryx) ∧ ((Rxy ∧ Ryz) → Rxz); 4. ∃x∃y∃z(∀w(w = x) ∨ (w = y) ∨ (w = z))∧ (¬x = y ∧ ¬y = z ∧ ¬x = z)∧ (Rxx ∧ Rxy ∧ Ryy ∧ Ryz ∧ Rzz ∧ Rxz) ∀u∀v(Ruv → (u = v ∨ (u = x ∧ v = y) ∨ (u = x ∧ v = z) ∨ (u = y ∧ v = z))). Exercise Let us consider a language with one binary operation symbol +. Express the following properties of a realization M in formulas: 1. M is associative; 2. M contains a unit element; 3. M is a group; 4. M is an abelian group. Solution As is usual practice, the symbol + goes between variables instead of preceding them. If we want to stick to our definition of terms and formulas, let us for example use the symbol f for the sum operation, and write it in prefix notation. 1. ∀a∀b∀cffabc = fafbc; 2. ∃e∀a(fae = a) ∧ (fea = a); 3. This includes the two previous formulas plus invertibility of all ele- ments: ∀a∃b(fab = e) ∧ (fba = e); 4. This includes all of the previous plus ∀a∀bfab = fba. Note that, if we are describing an abelian group, the existence of a unit and invertibility of elements can be simplified to the following forms: ∃e∀afae = a; ∀a∃bfab = e. Week 2 Definition Fixing a language, a sequent is a string of the form Θ ⇒ φ, where Θ is a finite set of formulas in the given language, φ is a single formula and ⇒ is a new symbol, not comprised in the language. Remark The way to think of sequents is: if all the premises in the antecedent Θ are true, then we conclude that φ is also true. We can also take Θ to be an empty set, in which case we often abbreviate the sequent to the single formula φ. 3 Definition A theory is a pair (L, A), where L is a formal language and A is a set of sequents, called the axioms of set theory. Informal definition A model for a theory is a realization that makes all its axioms true. Example Given the theory whose language contains one binary operation symbol f, and having the two axioms ∀a∀b∀cffabc = fafbc; ∃e∀a(fae = a) ∧ (fea = a) a model for this theory is precisely a monoid. Example Given the theory whose language contains one binary operation symbol f and one constant symbol e, and having the three axioms ∀a∀b∀cffabc = fafbc; ∀a(fae = a) ∧ (fea = a); ∀a∃b(fab = e) ∧ (fba = e) then a model for this theory precisely a group. Example Axiomatic set theory can be formalized as a theory in the sense above. Unfortunately, we will leave its description incomplete for now, concentrating on its parts a little at a time. The language of set theory only comprises one binary relation symbol ∈. So far, we have seen only some of its axioms, namely: Extensionality ∀x∀y(x = y ↔ (∀z(z ∈ x ↔ z ∈ y))); Union ∀x∃y∀z(z ∈ y ↔ (∃t(t ∈ x ∧ z ∈ t))); Power ∀x∃y∀z(z ∈ y ↔ z ⊆ x); Pairing ∀x∀y∃z(x ∈ z ∧ y ∈ z); Infinity ∃x(∅ ∈ x ∧ ∀y(y ∈ x → y ∪ {y} ∈ x)); note that when we write ∅, we can either define it through separation (see below) or add it to the language as a constant symbol, but in that case we need an additional axiom giving it the desired property. The separation schema, rather than an axiom, is a family of axioms (in fact, an infinite family) indexed by all possible formulas in the given language. In order to define it precisely, we need a preliminary definition. Definition Let x be a variable symbol and φ a formula in which x appears. Then x is said to be free in φ if its first occurrence is not quantified, i.e. it is not preceded by ∀ or ∃. Given a formula φ(x1, . . . , xn) where the variables xi’s are free, and given other variable symbols a1, . . . , an, then φ(a1, . . . , an) will denote the formula φ where all the occurrences of the variable symbols xi’s have been replaced by the corresponding symbols ai’s. Continuation We can now define the axioms of separation. Given any formula φ(x) in the language of set theory, where x is free, then there is an axiom 4 ∀z∃y(a ∈ y ↔ (a ∈ z ∧ φ(a)). Exercise Given a set A, how do we define the intersection of all its elements using the axioms of set theory? Solution We choose one of the sets a ∈ A, and then we observe that the expression ∀b ∈ A(x ∈ b) is a formula in which x is free. The intersection is now defined using separation, as ∩b∈Ab := {x ∈ a|∀b ∈ A(x ∈ b)}. Remark We might want to define union similarly. The string ∃b ∈ A(x ∈ b) is a formula in which x is free, and it certainly expresses the desired property of the union of all sets b ∈ A. However, there is a priori no set containing all the b’s, which is precisely why we need the axiom of union. • We know Cantor-Bernstein theorem, which states that if A and B are two sets and there are two injective functions f : A → B and g : B → A, then there is a bijection A ∼= B. The proof is non-trivial, but it is presented in the course notes, except for Tarski’s fixed point thorem, which is used to find a subset C ⊆ A such that C = A − g(B − f(C)). Definition A complete lattice is a partially ordered set such that every set of points in it admits a join. Tarski’s theorem Let X be a complete lattice and f : X → X an order-preserving function. Then there is a fixed point for f, i.e. ∃xf(x) = x. Proof By the axiom of separation, we can define Y = {y ∈ X|y ≤ f(y)}. Now take x := sup Y , which exists because X is complete. We have then that ∀y ∈ Y, y ≤ f(y) ≤ f(x) because f is order-preserving, therefore by definition of supremum we obtain x ≤ f(x). Moreover, this relation implies, again since f is order-preserving, that f(x) ≤ f2 (x), which means that f(x) ∈ Y . Consequently, f(x) ≤ x. Now we have that x ≤ f(x) and f(x) ≤ x which immediately gives x = f(x). Remark Tarski’s theorem can be applied in the proof of Cantor-Bernstein theorem, because for every set A, the power set P(A) is a complete lattice, ordered by inclusion. Week 3 Definition Recall that the product of two cardinals α = |A| and β = |B| is defined as α · β = |A × B|, and their exponential is αβ = |AB |, where the latter is the set of all functions B → A. Exercise Put the following cardinals in increasing order: ℵ0, 2·ℵ0, ℵ0·2, 2ℵ0 , ℵℵ0 0 , cℵ0 , ℵc 0, ℵ0· ℵ0. Solution We know that ℵ0 ≤ 2 · ℵ0 ≤ ℵ0 · ℵ0 = ℵ0, and moreover multiplication of cardinals is commutative, because for any two sets A and B the products A × B and B × A are naturally in bijection. Therefore we have 5 ℵ0 = 2 · ℵ0 = ℵ0 · 2 = ℵ0 · ℵ0. We also know, by Cantor theorem, that 2ℵ0 > ℵ0, and certainly ℵℵ0 0 ≥ 2ℵ0 so ℵℵ0 0 > ℵ0. Next, we can show that ℵℵ0 0 = 2ℵ0 = c. To see this, compute 2ℵ0 ≤ ℵℵ0 0 ≤ (2ℵ0 )ℵ0 = 2ℵ0·ℵ0 = 2ℵ0 . Similarly we compute 2ℵ0 ≤ cℵ0 = (2ℵ0 )ℵ0 = 2ℵ0·ℵ0 = 2ℵ0 so that cℵ0 = c. Finally, we compute ℵc 0 ≤ 2ℵ0·c = 2c ≤ ℵc 0 so ℵc 0 = 2c > c. In conclusion, we can order ℵ0 = 2 · ℵ0 = ℵ0 · 2 = ℵ0 · ℵ0 < 2ℵ0 = ℵℵ0 0 = cℵ0 < ℵc 0. Exercise Show that, for two infinite cardinals α, β, we have α + β = α · β = max{α, β}. Solution Let us first consider the sum, assuming without loss of generality that α ≤ β, so that max{α, β} = β. We compute β ≤ α + β ≤ β + β = 2 · β = β. The computation for the product is entirely analogous: β ≤ α · β ≤ β · β = β2 = β. Exercise Show that, for two infinite cardinals α ≤ β, we have αβ = 2β . Solution We have to show that 2β ≤ αβ and αβ ≤ 2β . We can proceed as in the previous exercise. 2β ≤ αβ ≤ (2α )β = 2α·β = 2β Exercise Given infinite cardinals α, β, γ such that α < β, determine if it is true that αγ < βγ and that γα < γβ . Solution Neither of the two inequalities is true in general. For example, it is true that ℵ < c but 6 ℵℵ0 0 = cℵ0 = c as we have seen above, and (2c )ℵ0 = (2c )c because they both equal (2c ), as it is easily proven with an argument just like the ones seen above. In general, we can only say that αγ ≤ βγ and γα ≤ γβ . Exercise Show that |Q| = |N|. Solution We know that |Z| = |N|, so it suffices to show that Q+ is in bijection with N. By the tabular argument we build a bijection between N and N × N, that is, we enumerate N as in the following diagram (0, 0) (0, 1) (0, 2) . . . (1, 0) (1, 1) (1, 2) . . . (2, 0) (2, 1) (2, 2) . . . ... ... ... ... Now, we have a surjection N × N → Q+ given by (p, q) → p q . Choosing p and q to be coprime gives us a left inverse function, which is clearly an injection Q+ → N × N ∼= N. Moreover, there is an injection N → Q+ given simply by n → n 1 . We conclude using Cantor-Bernstein’s theorem, which provides a bijection N ∼= Q+ . Exercise Show that |R| > |N|. Solution There is an inclusion N → R, so |N| ≤ |R|. We need to prove that there is no bijection R → N. By a geometrical projection, we know that the open interval (0, 1) is in bijection with the whole R, so we are reduced to prove that (0, 1) is not in bijection with N. Assume by contradiction that there is such a bijection. Then, if we express real numbers through their decimal representation, we have an enumeration 0.h1 1h1 2h1 3 . . . 0.h2 1h2 2h2 3 . . . 0.h3 1h3 2h3 3 . . . ... Now define a new real number 0.k1k2k3 . . . where kn = 1 whenever hn n is even and kn = 2 whenever hn n is odd. This number can’t be part of this enumeration by construction, so the claimed bijection is not surjective, which is a contradiction. 7 Week 4 Exercise Characterize all well-ordered sets A such that Aop is also well-ordered. Solution These sets are precisely the finite ones. Indeed, if A is finite then each of its subsets has a maximum element, which becomes a minimum in Aop . On the other hand, if A is infinite, then we know that |ω| ≤ |A|, so that ω is an initial segment of A. Now take a non-empty subset of ω, and therefore of A, which has no maximum. Its corresponding subset in ωop ⊆ Aop has no minimum. Exercise Give an example of a well-ordering on the set ω × ω. Solution A possible solution is the lexicographic ordering: for (a, b), (a , b ) ∈ ω×ω, set (a, b) < (a , b ) if either a < a or a = a and b < b . It is immediate to check that it is a total order. Moreover, suppose S ⊆ ω × ω is non-empty. Then take a0 := min{a ∈ ω|∃b such that (a, b) ∈ S} which exists because ω is well-ordered, and b0 := min{b ∈ ω|(a0, b) ∈ S} which exists for the same reason. Then (a0, b0) is the minimum of S. Another possible well order is: (a, b) < (a , b ) if either a + b < a + b or a + b = a + b and a < a . (0, 0) (0, 1) (0, 2) (0, 3) . . . (1, 0) (1, 1) (1, 2) (1, 3) . . . (2, 0) (2, 1) (2, 2) (2, 3) . . . (3, 0) (3, 1) (3, 2) (3, 3) . . . ... ... ... ... ... This is a total order because ω is, and a + b, a + b ∈ ω. Moreover, given a non-empty subset S ⊆ ω × ω, define 8 n0 := min{n ∈ ω|∃a∃b such that (a, b) ∈ S and a + b = n} which exists because ω is well-ordered. Then define a0 := min{a ∈ ω|∃b such that (a0, b) ∈ S and a + b = n0} which again exists for the same reason. Then (a0, n0 −a0) is the minimum of S. We can give yet another solution: calling c = max{a, b} and c = {a , b }, (a, b) < (a , b ) if c < c or c = c and a < a or c = c , a = a and b < b . In other words, it restricts to the lexicographic order whenever the maxima coincide. (0, 0) (0, 1) (0, 2) (0, 3) . . . (1, 0) (1, 1) (1, 2) (1, 3) . . . (2, 0) (2, 1) (2, 2) (2, 3) . . . (3, 0) (3, 1) (3, 2) (3, 3) . . . ... ... ... ... ... This is a linear order because ω is linearly ordered and c, c ∈ ω. To see that it is a well-ordering, let us consider a non-empty subset S ⊆ ω × ω, and define c0 := min{c ∈ ω|∃(a, b) ∈ S such that max{a, b} = c} a0 := min{a ∈ ω|∃b such that (a, b) ∈ S and max{a, b} = c0} b0 := min{b ∈ ω|(a0, b) ∈ S}. Then the pair (a0, b0) is the minimum of S. Exercise Prove that the set of algebraic numbers is countable. Solution Remember that a real number r is called algebraic if there is a polynomial a0 + a1x + a2x2 + . . . + anxn with integers coefficients such that a0 + a1r . . . + anrn = 0. The set of polynomials of degree n is in bijection with Zn , because every polynomial is determined by its coefficients. Now, we know from algebra that a polynomial of degree n has at most n roots, so the set of real numbers that are roots of polynomials of degree n is contained in n × 9 Zn . As a consequence, the set Q of all algebraic numbers is contained in n∈N n × Zn . Now observe that Z is countable, as we already know, and a finite product of countable sets is easily shown to be countable, so Zn is countable as well. A fortiori, we know that n × Zn is countable, because it is a finite product of sets that are at most countable. Finally, a countable union of countable sets is countable, because n∈N N × {n} = N × N. It follows that n∈N n × Zn is countable, so that there is an injective function Q → n∈N n × Zn ∼= N and |Q| ≤ |N|. For the other direction, observe that every natural number n is algebraic, just taking the polynomial −n + x, so the inclusion N ⊆ Q implies that |N| ≤ |Q|. Axiom of choice The axiom of choice states that for every family of sets (Xi)i∈I there is a function c : I → i∈I Xi such that ∀i ∈ I we have that c(i) ∈ Xi. Well-ordering principle The well-ordering principle states that for every set X there can be defined a well-ordering on X. Exercise Show that the well-ordering principle implies the axiom of choice. Solution Assume the well-ordering principle. Then make every Xi into a wellordered set, and define the function c : i → min Xi. Remark In fact, the converse implication is also true, that is, assuming the axiom of choice we can define a well-ordering on every set, but the proof of this is much more involved. Week 5 Exercise Show that ω · 2 ω Solution If there was an isomorphism of ordered sets, each subset of ω · 2 would correspond uniquely to a subset of ω with the same order-theoretic properties. But ω · 2 has an infinite subset with a maximum, namely ω + 1, which ω doesn’t have. Therefore they can’t be isomorphic. Definition We denote ω0 := ω. Then inductively for every ordinal α we denote as ωα+1 the smallest ordinal number whose cardinality is bigger than |ωα|. For example, the symbol ω1 will denote the smallest uncountable ordinal. Exercise Find, in the class of ordinal numbers: 1. The third smallest infinite ordinal number; 2. some ordinal α such that ω2 < α < ω3 ; 3. the largest countable ordinal number; 10 4. the ω-th uncountable ordinal number. Solution We separate the four requests: 1. The smallest infinite ordinal number is ω, so we only have to take its successor twice, in the class of ordinals. We obtain that ω + 2 is what we are looking for. 2. Any ordinal of the form ω2 · n with n finite, or ω2 + β, with β < ω3 , satisfies the requirement. 3. No such ordinal exists. Indeed, for each countable ordinal α we have α < α + 1, and α + 1 is clearly also countable. 4. By ω-th, we mean that the linearly ordered set of all uncountable ordinals smaller than the one we wish to find is isomorphic to ω. Now, since the smallest uncontable ordinal is ω1, the second smallest uncountable ordinal will be ω1 + 1, so by induction we obtain that the (n + 1)-th smallest uncountable ordinal is ω1 + n. This implies that the ω-th smallest will be ω1 + ω. Exercise Write the following ordinals in the simplest possible way: 1ω , 2ω , . . . , nω , . . . , ωω . Determine which ones, if any, are countable. Solution By definition, 1ω is the supremum of the set {1m |m < ω}, which is constant on 1. Therefore, 1ω = 1. For each 1 < n < ω, the set nω is the supremum of the set {nm |m < ω}. All the elements of this set are finite ordinals, and moreover they are unbounded, because for every m < ω we have m < nm . Therefore, nω is equal to ω itself. ωω is the supremum of the set {ωn |n < ω}. Since all these ordinals are distinct, there is no simpler way of writing ωω . Moreover, it is a countable union of countable sets, therefore it is itself countable. Observation The examples above are instances of one important fact: the arithmetic of cardinals does not correspond with the arithmetic of ordinals. For example, addition and multiplication are commutative in the former, but not in the latter. Moreover, an operation like αβ can have very different results depending on whether we regard α and β as cardinals or as ordinals. However, it is still true that αβ+γ = αβ · αγ and (αβ )γ = αβ·γ , although a little more complicated to prove (it will be done later on in the course). Exercise Show that the equality α · ω1 = ω1 holds for each ordinal 1 ≤ α < ω1. Solution α · ω1 is the supremum of the set {α · β|β is countable }. Each element α · β is countable because it is a product of countable sets, so we have α · β < ω1, which implies α · ω1 ≤ ω1. On the other hand, for each countable ordinal β we have that β ≤ α · β < α · ω1, so that α · ω1 is bigger than all countable ordinals, and therefore ω1 ≤ α · ω1. We have proven both the desired inequalities, so ω1 = α · ω1. 11 Week 6 Exercise We saw in an earlier exercise that the cardinals ℵ0, 2·ℵ0, ℵ0·2, 2ℵ0 , ℵℵ0 0 , cℵ0 , ℵc 0, ℵ0· ℵ0 are ordered as ℵ0 = 2 · ℵ0 = ℵ0 · 2 = ℵ0 · ℵ0 < 2ℵ0 = ℵℵ0 0 = cℵ0 < ℵc 0. We want to do an analogous exercise, but regarding the given numbers as ordinals instead. In other words, find the increasing ordinal order of ω, 2 · ω, ω · 2, 2ω , ωω , ωω 1 , ωω1 , ω · ω. Solution First, recall from the lecture that 2·ω = ω. Moreover, we have seen above that ω·2 ω, so we obviously conclude that ω < ω·2. In another exercise, we have seen that 2ω = ω, while ωω is bigger than all ωn ’s and therefore all ω ·n’s, but it is still countable, and it behaves very differently in regard to the respective order relations. Next we compute ωω1 = (2ω )ω1 = 2ω·ω1 = 2ω1 = ω1. As for ωω 1 , this is the supremum of the set {ωn 1 |n < ω}, all of whose elements are strictly bigger than ω1, so ωω 1 > ω1. Finally, we know that ω · ω is countable, and it is obviously bigger than ω · 2 but smaller than ωω . In conclusion, the order of the given ordinals is ω = 2 · ω = 2ω < ω · 2 < ω · ω < ωω < ωω1 < ωω 1 . Exercise Find the smallest ordinal number α such that the power αω is 1. a finite ordinal number; 2. a countable ordinal number; 3. an uncountable ordinal number. Solution The solution to (1) is immediate, because 0ω = 0, and 0 is the smallest ordinal number. For (2), observe that 1ω = 1 and 2ω = ω. Since 2 = 1+ , our solution is exactly 2. For (3), observe that for every countable ordinal number β we have βω = sup{βn |n ∈ N}. Now, βn is the product of a finite number of countable sets, therefore it is countable itself. It follows that βω , being a countable union of countable sets, is itself countable. So the ordinal we are looking for must be uncountable, therefore at least ω1. Since ωω 1 is uncountable, our solution is precisely ω1. Exercise Define ωω as the smallest ordinal whose cardinality is strictly bigger than the cardinality of all ωn’s. In other words, ωω := min{α, ∀n ∈ N|ωn| < |α|}. 12 Show that ωω = supn ωn. Solution We will show both inequalities. First, observe that |ωn| < |ωω| implies ωn < ωω for each n ∈ N, so that, by definition of supremum, we obtain supn ωn ≤ ωω. Conversely, since for each n the inequality ωn < supn ωn holds, it must be true that |ωn| ≤ | supn ωn| (with the weak inequality sign, for the moment). Thus we obtain |ωn| < |ωn+1| ≤ | supn ωn| now with a strict inequality sign. By definition of ωω, this means that ωω ≤ supn ωn. Definition Remember that for every ordinal number α > 0 there are non-null natural numbers k, m0, . . . , mk and ordinal numbers γ0 > . . . > γk such that α = ωγ0 · m0 + . . . + ωγk · mk. This expression is known as the Cantor normal form of α. Exercise Find the Cantor normal form of ωn, or each n, and ωω. Solution Following the first steps of the proof of the existence theorem for Cantor normal forms, we start by looking at the set X := {γ|ωγ ≤ ωn} and then we study the ordinal ωsup X . Now, if n = 0, then X = {0, 1}, so that the above ordinal becomes ω1 = ω = ω0, and we are done. If 0 < n < ω, then X is the set of all ordinals of cardinality strictly smaller than ωn, so that sup X = ωn. Thus we obtain that ωωn ≤ ωn. We only need to show the other inequality. For that, observe that by previous exercises we know that 2ωn = ωn, so compute ωn = 2ωn ≤ ωωn . Finally, turning to ωω, our X will be the set of all ordinals of cardinality strictly smaller than ωω, therefore the same argument shows that the Cantor normal form turns out to be ωωω . Remark The same line of reasoning proves in fact a more general result: if α is an ordinal which is the smallest with its cardinality, then its Cantor normal form is ωα . 13 Proposition The Cantor normal form is unique. Proof Consider an ordinal α, and take two Cantor normal forms for it ωγ0 · m0 + . . . + ωγk · mk and ωδ0 · n0 + . . . + ωδl · nl assuming without loss of generality that k ≤ l. Since γ0 > . . . > γk and δ0 > . . . > δl, if γ0 < δ0 we would have that the first normal form is smaller than the second, which is absurd. Therefore, γ0 = δ0. For the same reason, we must have m0 = n0. Inductively, we obtain that for 0 ≤ i ≤ k we have γi = δi and mi = ni. It only remains to prove that k = l. Suppose by contradiction that k < l. Then the remaining terms of the second normal form sum up to an ordinal β = 0. Therefore, we obtain that first normal form gives α, while the second gives α + β > α, which is absurd. This concludes the proof. Week 7 Exercise Order the following ordinals increasingly: ω + 1 + ω, ω · (ω + 1), (1 + ω) · ω, (1 + ω) · (ω + 1), ω + 1 + ω2 + ω. Solution Remember that sum of ordinal numbers is associative, so the first is equal to ω + (1 + ω) = ω + ω. Also remember that multiplication distributes with respect to addition, so we have ω · (ω + 1) = ω · ω + ω = ω2 + ω. Next, we have again (1 + ω) · ω = ω · ω = ω2 . Use the same principles to compute (1 + ω) · (ω + 1) = ω · (ω + 1) = ω2 + ω. Finally, we have ω + (1 + ω2 + ω) = ω + (ω2 + ω) = (ω + ω2 ) + ω = ω · (1 + ω) + ω = ω · ω + ω = ω2 + ω. These ordinals are then ordered as ω + 1 + ω < (1 + ω) · ω < ω · (ω + 1) = (1 + ω) · (ω + 1) = ω + 1 + ω2 + ω. 14 Exercise Decide whether ωω = ω or ω(ωω ) = ωω . Solution Both are false. The proof of the first is analogous to the proof of ω ·2 = ω, which extends to all ω · n’s and all ωn ’s. As for the second, consider in both sets the subset of all elements γ such that the set {δ|δ < γ} has no maximum. For the left-hand side, this is isomorphic (as an ordered set) to ωω , for the right-hand side it is isomorphic to ω, so the result follows from the first part. Exercise Define 0 to be the smallest ordinal α such that ωα = α. Prove that 0 = ωωω··· . Solution The ordinal ωω··· is defined to be the supremum of the sequence of ordinals defined by α0 = ω and for each n ∈ ω, αn+1 = ωαn . First, observe that, this limit obviously has the required property. Therefore, by minimality of 0, we have 0 ≤ ωω··· . To prove the opposite inequality, let us start by observing that 0 cannot be finite, so α0 = ω ≤ 0. Recursively, exponentiating ω with the n-th obtained inequality, we get αn+1 = ωαn ≤ ω 0 = 0, so αn ≤ 0, for each αn of the sequence defined above. Therefore, by the property of the supremum, we obtain ωω··· ≤ 0. Exercise Show that the axiom of choice is equivalent to the statement: every surjective function f : X → Y has a right inverse. Solution Assuming the axiom of choice, then we can choose a point xy ∈ Xy for every fiber of f, and define a function by y → xy. This is clearly a right inverse to f. Conversely, consider a family of sets (Xi)i∈I, and the projection function i∈I Xi → I. This is surjective, then by assumption it has a right inverse c : I → i∈I Xi. Therefore, for every i ∈ I we have c(i) ∈ pi = Xi, so this is a choice function. Curiosities The axiom of choice has been proven to be independent of ZF set theory. Therefore, both ZF +AC and ZF +¬AC are consistent theories (provided ZF is). Both AC and its negation ¬AC have weird and counterintuitive consequences, the most famous of which, assuming AC is Banach-Tarski paradox: given a 3-dimensional full ball, there is a way of cutting it into a finite number of pieces, and then reassemble these to form two spheres identical in size to the original one. Another statement which is equivalent to AC is: – Every vector space has a basis. On the contrary, assuming ¬AC, we have these fun facts: 15 – There is an infinite set X with no injection N → X; – there is a set X and a partition of X such that there are more elements in the partition than elements in the set; – there are families of non-empty sets whose Cartesian product is empty; – there are vector spaces with no basis; – it is possible to express real numbers as a countable union of countable sets; – it is furthermore possible to express real numbers as a union of two subsets of strictly smaller cardinality. Week 8 Definition The cofinality of a cardinal κ is the minimum size of a set of cardinals smaller than κ whose sum is κ, i.e. cf(κ) = min{|I|, κ = i∈I λi with ∀i ∈ I, λi < κ}. Definition Also remember, a cardinal κ is called regular if for every family of cardinals (λi)i∈I such that λi < κ and |I| < κ, then i∈I λi < κ. Remark Obviously, for every cardinal we have cf(κ) ≤ κ, because κ 1 = κ. The statement that κ is a regular cardinal is then equivalent to the statement cf(κ) = κ. Definition The sequence of ℵ numbers is recursively defined on ordinals: – ℵ0 = ω; – for an ordinal α, ℵα+1 = ℵ+ α ; – for a limit ordinal γ, ℵγ = β<γ ℵβ. Definition The sequence of numbers is recursively defined on ordinals: – 0 = ω; – for an ordinal α, α+1 = 2 α ; – for a limit ordinal γ, γ = β<γ β. Remark Cantor’s theorem states that we always have κ < 2κ . Therefore, we also always have ℵα ≤ α. Remark With this notation, the continuum hypothesis states that ℵ1 = 1. In particular, it is consistent with ZFC that 1 = ℵ+ 0 . It is also consistent that 1 = ℵ++ 0 . In fact, for any finite number n, the statement 1 = ℵ+n 0 = ℵn 16 is consistent with ZFC. Therefore, the statement that for every ordinal α and every natural number n the equality α = ℵα+n holds is consistent with ZFC. Remark Even more, given any cardinal κ, the following statement is consistent with ZFC: there is a possibly infinite ordinal λ such that 2κ = κ+λ . We can choose such an ordinal λ for each κ with some degree of liberty. However, it has been proven that if we want to choose a single λ such that the equality 2κ = κ+λ is true simultaneously for every cardinal κ, then λ needs to be finite. Remark The generalized continuum hypothesis can be reformulated as ∀α, ℵα = α. Definition A cardinal κ is called a strong limit if whenever λ < κ then 2λ < κ. It is called a weak limit if whenever λ < κ then λ+ < κ. Furthermore, κ is strongly inaccessible if it is regular and a strong limit, it is weakly inaccessible if it is regular and a weak limit. Remark Every, strong limit is clearly also a weak limit, and every strong inaccessible is a weak inaccessible. Under GCH, the definitions of strong limit and weak limit are equivalent, and the definitions of strongly inaccessible and weakly inaccessible are likewise equivalent. Exercise Show that, for each finite n, the cardinal ℵn is regular but, if n ≥ 1, it is not a strong or weak limit. Further show that ℵω is a weak limit but not regular. Decide if ℵω is a strong limit. Solution We know that a finite sum of finite numbers is finite, which means that ℵ0 is regular. For any n, assume that we have a family of cardinals (λi)i∈I with |I| ≤ ℵn and λi ≤ ℵn, therefore we can compute i∈I λi ≤ i∈I ℵn = |I| × ℵn ≤ ℵn · ℵn = ℵn which proves that ℵn+1 is regular. For each n we have that ℵn < ℵn+1 but 2ℵn ≥ ℵ+ n = ℵn+1 so ℵn+1 is not a weak or a strong limit. ℵω is not regular because by its very definition n∈ω ℵn = ℵω. Moreover, it is a weak limit because every cardinal smaller than ℵω is either finite or of the form ℵn. It is obvious that both n + 1 < ℵω and that ℵ+ n = ℵn+1 < ℵω. In ZFC, it is not possible to prove that ℵω either is a strong limit or is not such. Under GCH, for instance, it is obviously a strong limit because strong limits coincide with weak limits. On the other hand, it is consistent with ZFC that there is an n and and infinite λ such that 2ℵn = ℵ+λ n . In this case, ℵn < ℵω, but 2ℵn ≥ ℵω. 17 Remark By transfinite induction, we can prove that it is always the case that κ ≤ ℵκ, for an arbitrary cardinal κ. Remark Whenever α is an initial ordinal (least with its cardinality) then we have cf(ℵα) = α. Exercise Show that an uncountable cardinal κ is weakly inaccessible if and only if κ = ℵκ. Solution Assume that κ = ℵκ. Then we have cf(κ) = cf(ℵκ) = κ so that κ is regular. Take now a cardinal µ < κ, therefore µ < ℵκ, so that there is an ordinal α < κ such that µ < ℵα. Therefore µ+ < ℵα+1 < ℵκ = κ. Conversely, if κ is weakly inaccessible. Since κ ≤ ℵκ, we need to prove the opposite inequality. We have ℵ0 < κ by hypothesis. Since κ is a weak limit, for each ordinal α we also have that ℵα < κ ⇒ ℵα+1 < κ. Finally, for a limit ordinal λ < κ, by the previous step and regularity of κ we deduce that ℵλ < κ. By definition of ℵ number associated to a limit ordinal, this implies that ℵκ ≤ κ. Remark The same reasoning yields that an uncountable cardinal κ is strongly inaccessible if and only if κ = κ. Remark There is a function, called the ℵ function, from the class of cardinals to itself, defined by κ → ℵκ. A weakly inaccessible cardinal can be defined as a fixed point of the ℵ function. Similarly we can define strongly inaccessible cardinals as fixed points of an analogously defined function. Remark The existence of one inaccessible cardinal is independent of ZFC. Moreover, the existence of λ+1 inaccessible cardinals is independent of ZFC + “there are λ inaccessible cardinals”. Finally, the existence of a proper class of inaccessible cardinals is independent of ZFC + “there are λ inaccessible cardinals” for every λ. Week 9 Exercise Show that, for a set x, we have rk(x) = sup{rk(z) + 1|z ∈ x}. Solution Let α = sup{rk(z) + 1|z ∈ x}. This means in particular that for every z ∈ x we have z ⊆ Vβ for some β such that β+1 ≤ α, which implies z ∈ Vα. As a consequence, x ⊆ Vα, which means that rk(x) ≤ α. We want to show that this is an equality. Suppose by contradition that rk(x) = γ < α. In other words, x ⊆ Vγ. Then for each element z ∈ x we have z ∈ Vγ, which contradicts the minimality of α. This proves that rk(x) = α, as desired. 18 Remark Another way to express this is: if ∃ max{rk(z)|z ∈ x} then rk(x) = max{rk(z)|z ∈ x} + 1. If that maximum doesn’t exist, then rk(x) = sup{rk(z)|z ∈ x}. Exercise Prove by induction that |Vω+α| = α. Solution For α = 0, notice that Vω is a countable union of finite sets, which is bigger than any finite set, therefore |Vω| = ω = 0. Now consider the isolated step, and assume by induction that |Vω+α| = α. Then |Vω+α+1| = |P(Vω+α)| = 2|Vω+α| = 2 α = α+1. Finally, if α is a limit ordinal, then |Vω+α| = | β<α Vω+β| = β<α |Vω+β| = β<α β = α. Exercise Conclude that if κ is an inaccessible cardinal, then |Vκ| = κ. Solution We can regard a cardinal as an initial ordinal. Since κ is uncountable, we have ω + κ = κ, thus the exercise above says |Vκ| = |Vω+κ| = κ but we know from last week that κ = κ, so we are done. Exercise Show that, if x and y have rank ≤ α then {x, y}, x ∪ y, x, P(x), (x, y) and xy have rank < α + ω. Solution {x, y} has rank ≤ α + 1 by the exercise above. Next, for each element z ∈ x or z ∈ y we have z ∈ Vβ for some β ≤ α, therefore, again by the same exercise above, we conclude rk(x∪y) ≤ α+1. By x we mean the union of all sets y such that y ∈ x. Since rk(x) ≤ α, we also have a fortiori rk(y) ≤ α and, for each element z ∈ y, rk(z) ≤ α. Therefore rk( x) ≤ α. In fact, we even have rk( x) < rk(x). If rk(x) = β ≤ α, then for each z ∈ x we have rk(z) ≤ β, so for each subset y ⊆ x, by the exercise above we deduce rk(y) ≤ β + 1. Thus, rk(P(x)) ≤ β + 2 ≤ α + 2. We use the definition (x, y) = {x, {x, y}}. By the above reasoning, rk({x, y}) ≤ α + 1, so by the exercise above we have rk((x, y)) ≤ max{α + 1, α + 2} = α + 2. A function y → x is defined as a specific subset of y × x. Now, if y and x have rank ≤ α, every element of them also has rank ≤ α, therefore by the previous part a pair (y0, x0) where y0 ∈ y and x0 ∈ y has rank ≤ α + 2. The above exercise implies that a function y → x, when regarded as a set, has rank ≤ α+3. The set xy of all functions y → x now has rank ≤ α+4. 19 Week 10 Definition The axiom schema of replacement says that for each formula φ containing the variables x, t1, . . . , tn, y there is an axiom as follows ∀x∀y∀z (φ (x, t1, . . . , tn, y) ∧ φ (x, t1, . . . , tn, z) → y = z) → ∀X∃Y ∀y (y ∈ Y ↔ ∃x (x ∈ X ∧ φ (x, t1, . . . , tn, y))) Intuition We should think of φ as a relation which depends on parameters t1, . . . , tn, so for the sake of intuition we now abbreviate φ(x, t1, . . . , tn, y) by xRy. Now what the first part of the above axiom is saying is that R is a binary relation such that to each x there is at most one y which is in relation with it. The second half of it says that if X is a set, then the elements y that are associated to some element x ∈ X form themselves a set Y One more layer of simplification is the following: if R is a function (wherever it is defined) then the images of elements that range over a set X form themselves a set Y . The non-triviality of this axiom lies in the fact that, if R is a function, its codomain is a priori the class of all sets, which is not a set. What the axiom says is, in these terms, that the image of a set along a class function is a set. Definition A Grothendieck universe is a set U satisfying the following axioms: 1. X ∈ U and Y ∈ X implies Y ∈ U; 2. If X ∈ U, then also P(X) ∈ U; 3. If I ∈ U and (Xi)i∈I is an I-indexed family of sets in U, then i∈I Xi ∈ U; 4. N ∈ U. Remark If U is a Grothendieck universe and X ∈ U, then |X| < |U|. To see this, simply use axiom 2 to see that P(X) ∈ U, then observe that by axiom 1 P(X) ⊆ U, therefore |X| < |P(X)| ≤ |U|. Exercise Show that if X ∈ U and Y ⊆ X then Y ∈ U. Solution We have Y ∈ P(X) ∈ U by axiom 2. Then we conclude by axiom 2. Exercise Show that if µ < |U|, then there is a set X ∈ U such that |X| = µ. Solution By the previous exercise and the fact that µ ≤ µ, it suffices to show that there is a set Y ∈ U such that |Y | = µ. Let us proceed by induction on µ. If µ = 0, then the result follows by axiom 4. Assume by induction that we have proven the result up to α, then there is a set Yα ∈ U with cardinality α. By axiom 2, its power set belongs to U and, moreover, it has cardinality α+1 by definition. Now assume that λ is a limit ordinal, and that for each α < λ there is a set 20 Yα ∈ U with cardinality α. Assume that λ < λ (the proof when λ = λ is much more complicated), so there is some α < λ such that λ < α. By inductive hypothesis, there is a set Yα ∈ U with cardinality α, so by the previous exercise there must be also a set I ∈ U with cardinality λ, so that we can index a family (Yi)i∈I including all the sets of the form Yα with α < λ. Now the union i∈I Yi belongs to U by axiom 3, and it is equal to α<λ Yα, which has cardinality λ. This concludes the proof. Exercise Show that, if U is a Grothendieck universe, then its cardinality is uncountable and strongly inaccessible. Solution By axiom 4, we have that N ∈ U, so the above remark implies that |N| < |U|, which means that U is uncountable. Let |I| < |U| and (Xi)i∈I a family of sets such that |Xi| < |U|. By the previous exercise, we know that I, Xi ∈ U, therefore i∈I Xi ∈ U. Again, we conclude by the same remark that the cardinality of this union is strictly smaller than that of U. Thus, |U| is a regular cardinal. Finally, if µ < |U|, choosing a set X such that |X| = µ, we have again by the previous exercise that X ∈ U, so axiom 2 implies that P(X) ∈ U. Once more, we use the same remark to compute 2µ = |P(X)| < |U|. Remark In particular, this implies that the existence of a Grothendieck universe is not provable in ZFC. Theorem [ZFC] If U is a Grothendieck universe, then it is a model of ZFC. Proof This is essentially the same proof seen in the lecture to see that Vκ is a model of ZFC when κ is inaccessible. In fact, we have the following stronger result. Proposition κ is an inaccessible cardinal if and only if Vκ is a Grothendieck universe. Proof Observe that axioms 1 Tarski axiom Letting Univ(y) be an abbreviation for all five axioms that say that y is a Grothendieck universe, then Tarski axiom can be formulate as follows ∀x∃y(x ∈ y ∧ Univ(y)) Definition Naturally, Tarski axiom is not provable in ZFC. We call TG the theory having all axioms of ZFC + Tarski axiom. Theorem [TG] ZFC is consistent. Proof By Tarski axiom, there is a Grothendieck universe U. Then this is a model of ZFC. Week 11 Definition Recall that a set X is called transitive if ∀x we have x ∈ X ⇒ x ⊂ X. Remark Observe that the ⊂ sign in the definition above can never become an = sign, because that would contradict the axiom of regularity. 21 Definition Also recall that the transitive closure of a set X is the smallest set TC(X) ⊇ X which is transitive. It can be explicitly defined as follows: fix X0 = X and then inductively Xn+1 = Xn. Then we have TC(X) = n∈N Xn. • We recall one last definition: Definition A set x is called symmetric is there is a finite set of atoms N(x) = {a1, . . . , an} such that whenever we have a permutation f : A → A such that ∀i = 1, . . . n, f(ai) = ai then the extension ˆf : WA → WA satisfies ˆf(x) = x. We say that a set x is hereditarily symmetric if it is symmetric and every element of TC(x) is symmetric. Remark Observe that the definition of carrier of a set is ambiguous, because it does not fix a single set N(x), rather it is based on the fact that such a set exists. So we will slightly modify this definition for the moment. Definition For a set x ∈ WA , define Carr(x) ⊆ Pfin(A) to be the collection of all carriers of x. They naturally form a poset via the inclusion relation. We will say that a set is symmetric if it satisfies Carr(x) = ∅. Exercise Prove the following: 1. Every set x ∈ V is hereditarily symmetric with carrier N(x) = ∅; 2. Every atom a ∈ A is hereditarily symmetric with carrier N(a) = {a}; Solution 1. Observe that if x ∈ V then we also have TC(x) ∈ V . Therefore, this amounts to prove that for every set x ∈ V and every permutation f : A → A we have ˆf(x) = x. Proceed inductively on bounded stages of the universe of discourse V . For V0 = there is nothing to prove, so we start at height 1. The restriction of f1 to V1 is defined as follows: for x ∈ V1, we necessarily have x = ∅, therefore f1(x) = {f0(y)|y ∈ x} = ∅ = x. Now assume that the for all ordinals up to α we have fα(x) = x whenever x ∈ V . Take x ∈ Vα+1 = P(Vα). For all y ∈ x, we have y ∈ Vα, so by inductive hypothesis fα(y) = y. Therefore, we have, by definition, fα+1(x) = {fα(y)|y ∈ x} = {y|y ∈ x} = x. Finally, if α is a limit ordinal and x ∈ Vα, then x ∈ Vβ for some β < α, so the result is true by directly applying the inductive hypothesis. 2. Since atoms have no elements by definition, we have TC(a) = a. Therefore, we only need to show that a is symmetric with carrier N(a) = {a}. But this is obvious, since whenever f(a) = a, then ˆf(a) = a by definition. 22 Exercise Give an example of a set that has two carriers C1 and C2 such that C1 ∩C2 is not a carrier. Solution For example, consider the set of atoms A = {a, b, c, d}, and take x = {a, b}. Now, if C1 = {a, b}, then it is obviously a carrier of x. On the other hand, taking C2 = {c, d} also gives a carrier of x, because the only two permutation that fix C2 are the identity and the transposition (a, b), and in both cases ˆf({a, b}) = {a, b}. Finally, we have C1 ∩ C2 = ∅ and this is not a carrier of x, because for example if f is the cycle (a, b, c, d) then ˆf(x) = {b, c} = x. Exercise Give an example of a set that is symmetric but not hereditarily symmetric. Solution Let A be an infinite set containing a subset C such that both C and A\C are infinite. Consider now P(A) ∈ WA 2 . This is clearly symmetric with carrier ∅, because if f : A → A is a permutation, then the images along ˆf of all subsets of A exhaust all subsets of A. On the other hand C is an element of TC(P(A)), but this is not symmetric, because both itself and its complement are infinite. Exercise Give an example of a set that is not symmetric but all its elements are symmetric. Solution It suffices to take the set C from the exercise above. Exercise Give an example of a set x such that TC(x) contains all the atoms, every element of x is symmetric but x is not. Solution Consider an infinite set A, and take as set of atoms the union of two distinct copies A A. So we denote every element either as a1 or as a2. Now, consider the set x = {{a1, a2}}a∈A. Every element of x is a finite set of atoms, so it is symmetric having itself as a carrier. Moreover, we have TC(x) = x ∪ (A A), which contains all the atoms. We need to show that x is not symmetric. Let C ⊆ A A be any finite subset, and select two elements a1, b1 /∈ C. Define the permutation f : A A → A A as follows: on the first copy of A, it is the identity; on the second copy of it, it is the transposition (a, b). Now we have ˆf({b1, b2}) = {b1, a2}. This set is an element of ˆf(x), but not an element of x, therefore ˆf(x) = x. Week 12 Exercise Given a filter F ⊆ P(X), prove that F is an ultrafilter if and only if for every finite union A = n i=1 Ai ⊆ X such that A ∈ F then we have Ai ∈ F for at least one of the indices i. Solution Assume that F is an ultrafilter, and by contradiction assume that there is a finite union A ∈ F such that for every index i we have Ai /∈ F. By the ultrafilter property, we have X \ Ai ∈ F. Therefore, since elements of a filter are closed under finite intersection, we have X \A = n i=0(X \Ai) ∈ F. Thus, we also have ∅ = A ∩ (X \ A) ∈ F which is impossible because every ultrafilter is proper. 23 Conversely, assume that F is a filter but not an ultrafilter. Then there is a set A ⊆ X such that A /∈ F and X \ A /∈ F, but obviously we have A ∪ (X \ A) = X ∈ F. Definition The cofinite filter is that defined as {A ⊆ X|X \ A is finite }. Exercise Show that any non-principal ultrafilter contains the cofinite filter. Solution Let F be a non-principal ultrafilter, and let A ⊆ X be such that X \ A is finite. Since F is an ultrafilter, it will suffice to show that X \ A /∈ F. In other words, we need to show that F does not contain any finite set. Since elements of F are closed under finite intersection, we are reduced to show that for every x ∈ X, we have {x} /∈ F. Suppose by contradiction that there is x ∈ X such that {x} ∈ F. Then every set containing {x} is in F, and every set not containing it is not in F, because otherwise we would have ∅ ∈ F. In other words, F is the principal ideal generated by x, which contradicts our hypothesis. Definition A two-valued measure µ on a set X is called λ-additive if for every set I such that |I| < λ and every family (Ai)i∈I of pairwise disjoint subsets of X, we have µ( i∈I Ai) = i∈I µ(Ai). Example For a point x ∈ X, we can always define a two-valued λ-additive measure on X given by µ(A) = 1 if and only if x ∈ A. Definition An uncountable cardinal λ is called measurable if for every set of cardinality λ it is possible to define a non-trivial two-valued λ-additive measure on it. • The next definitions and propositions give an idea of how large these cardinals are. Definition Let us give a definition which is recursive on ordinals. We call a cardinal 0-inaccessible if it is inaccessible. For an ordinal α > 0, we say that a cardinal κ is α-inaccessible if it is inaccessible and for every ordinal β < α the set of β-inaccessible cardinals < κ is unbounded. We will say that a cardinal is hyperinaccessible if it is α-inaccessible for every ordinal α. Proposition Every measurable cardinal is hyperinaccessible. Proposition Vopˇenka’s principle implies the existence of arbitrarily large measurable cardinals. Remark Vopˇenka’s principle is provably independent of ZFC + existence of measurable cardinals. In fact, it implies the existence of supercompact cardinals, which are a stronger version of strongly compact cardinals, but still much weaker than Vopˇenka’s principle. • Our final point will be to establish conditions under which Vopˇenka’s principle can be satisfied. Definition A model of a theory is transitive if it is transitive as a set, i.e. x ∈ y ∈ M implies x ∈ M. 24 Definition Given two models of ZFC M1 and M2, a map h : M1 → M2 is called an elementary embedding if for each formula φ(x1, . . . , xn) and any assignment xi → ai with ai ∈ M1, we have that φ(a1, . . . , an) is true in M1 if and only if φ(ha1, . . . , han) is true in M2. Definition A cardinal λ is called huge if there exists a transitive model M of ZFC and an elementary embedding h : V → M such that: 1. λ is the smallest ordinal such that h(λ) = λ; 2. M contains every function h(λ) → M. • This definition is quite technical. We won’t be interested in all the details that come with it, rather we only want to focus on the following point: Proposition Suppose λ is a huge cardinal, then Wλ is a model of ZFC in which Vopˇenka’s principle is true. Remark This does not mean that the truthfulness of Vopˇenka’s principle follows from the existence of huge cardinals. It only means that if there is a huge cardinal, then we can find at least one model of ZFC in which the principle holds true, i.e. it is consistent with ZFC. 25