Ordinal arithmetic

Last updated

In the mathematical field of set theory, ordinal arithmetic describes the three usual operations on ordinal numbers: addition, multiplication, and exponentiation. Each can be defined in essentially two different ways: either by constructing an explicit well-ordered set that represents the result of the operation or by using transfinite recursion. Cantor normal form provides a standardized way of writing ordinals. In addition to these usual ordinal operations, there are also the "natural" arithmetic of ordinals and the nimber operations.

Contents

Addition

The union of two disjoint well-ordered sets S and T can be well-ordered. The order-type of that union is the ordinal that results from adding the order-types of S and T. If two well-ordered sets are not already disjoint, then they can be replaced by order-isomorphic disjoint sets, e.g. replace S by {0} × S and T by {1} × T. This way, the well-ordered set S is written "to the left" of the well-ordered set T, meaning one defines an order on ST in which every element of S is smaller than every element of T. The sets S and T themselves keep the ordering they already have.

The definition of addition α + β can also be given by transfinite recursion on β:

Ordinal addition on the natural numbers is the same as standard addition. The first transfinite ordinal is ω, the set of all natural numbers, followed by ω + 1, ω + 2, etc. The ordinal ω + ω is obtained by two copies of the natural numbers ordered in the usual fashion and the second copy completely to the right of the first. Writing 0' < 1' < 2' < ... for the second copy, ω + ω looks like

0 < 1 < 2 < 3 < ... < 0' < 1' < 2' < ...

This is different from ω because in ω only 0 does not have a direct predecessor while in ω + ω the two elements 0 and 0' do not have direct predecessors.

Properties

Ordinal addition is, in general, not commutative. For example, 3 + ω = ω since the order relation for 3 + ω is 0 < 1 < 2 < 0' < 1' < 2' < ..., which can be relabeled to ω. In contrast ω + 3 is not equal to ω since the order relation 0 < 1 < 2 < ... < 0' < 1' < 2' has a largest element (namely, 2') and ω does not (ω and ω + 3 are equipotent, but not order isomorphic).

Ordinal addition is still associative; one can see for example that (ω + 4) + ω = ω + (4 + ω) = ω + ω.

Addition is strictly increasing and continuous in the right argument:

but the analogous relation does not hold for the left argument; instead we only have:

Ordinal addition is left-cancellative: if α + β = α + γ, then β = γ. Furthermore, one can define left subtraction for ordinals βα: there is a unique γ such that α = β + γ. On the other hand, right cancellation does not work:

but

Nor does right subtraction, even when βα: for example, there does not exist any γ such that γ + 42 = ω.

If the ordinals less than α are closed under addition and contain 0, then α is occasionally called a γ-number (see additively indecomposable ordinal). These are exactly the ordinals of the form ωβ.

Multiplication

The disjoint union
html.skin-theme-clientpref-night .mw-parser-output div:not(.notheme)>.tmp-color,html.skin-theme-clientpref-night .mw-parser-output p>.tmp-color,html.skin-theme-clientpref-night .mw-parser-output table:not(.notheme) .tmp-color{color:inherit!important}@media(prefers-color-scheme:dark){html.skin-theme-clientpref-os .mw-parser-output div:not(.notheme)>.tmp-color,html.skin-theme-clientpref-os .mw-parser-output p>.tmp-color,html.skin-theme-clientpref-os .mw-parser-output table:not(.notheme) .tmp-color{color:inherit!important}}
{ (0,n) : n [?] N } [?]
{ (1,n) : n [?] N } , using lexicographic order, has order type o * 2. This is different from o. OmegaPlusOmega svg.svg
The disjoint union { (0,n) : nN } { (1,n) : nN } , using lexicographic order, has order type ω • 2. This is different from ω.
The set {
(n,0),
(n,1) : n [?] N } , under lexicographic order, has order type 2 * o, which is equal to o. TwoTimesOmega svg.svg
The set { (n,0), (n,1) : nN } , under lexicographic order, has order type 2 • ω, which is equal to ω.

The Cartesian product, S×T, of two well-ordered sets S and T can be well-ordered by a variant of lexicographical order that puts the least significant position first. Effectively, each element of T is replaced by a disjoint copy of S. The order-type of the Cartesian product is the ordinal that results from multiplying the order-types of S and T.

The definition of multiplication can also be given inductively (the following induction is on β):

As an example, here is the order relation for ω · 2:

00 < 10 < 20 < 30 < ... < 01 < 11 < 21 < 31 < ...,

which has the same order type as ω + ω. In contrast, 2 · ω looks like this:

00 < 10 < 01 < 11 < 02 < 12 < 03 < 13 < ...

and after relabeling, this looks just like ω. Thus, ω · 2 = ω+ωω = 2 · ω, showing that multiplication of ordinals is not in general commutative, c.f. pictures.

As is the case with addition, ordinal multiplication on the natural numbers is the same as standard multiplication.

Properties

α · 0 = 0 · α = 0, and the zero-product property holds: α · β = 0 → α = 0 or β = 0. The ordinal 1 is a multiplicative identity, α · 1 = 1 · α = α. Multiplication is associative, (α · β) · γ = α · (β · γ). Multiplication is strictly increasing and continuous in the right argument: (α < β and γ > 0) → γ·α < γ·β. Multiplication is not strictly increasing in the left argument, for example, 1 < 2 but 1 · ω = 2 · ω = ω. However, it is (non-strictly) increasing, i.e. αβα·γβ·γ.

Multiplication of ordinals is not in general commutative. Specifically, a natural number greater than 1 never commutes with any infinite ordinal, and two infinite ordinals α and β commute if and only if αm = βn for some nonzero natural numbers m and n. The relation "α commutes with β" is an equivalence relation on the ordinals greater than 1, and all equivalence classes are countably infinite.

Distributivity holds, on the left: α(β + γ) = αβ + αγ. However, the distributive law on the right (β + γ)α = βα+γα is not generally true: (1 + 1) · ω = 2 · ω = ω while 1 · ω + 1 · ω = ω + ω, which is different. There is a left cancellation law: If α > 0 and α · β = α · γ, then β = γ. Right cancellation does not work, e.g. 1 · ω = 2 · ω = ω, but 1 and 2 are different. A left division with remainder property holds: for all α and β, if β > 0, then there are unique γ and δ such that α = β · γ + δ and δ < β. Right division does not work: there is no α such that α · ωωω ≤ (α + 1) · ω.

The ordinal numbers form a left near-semiring, but do not form a ring. Hence the ordinals are not a Euclidean domain, since they are not even a ring; furthermore the Euclidean "norm" would be ordinal-valued using the left division here.

A δ-number (see Multiplicatively indecomposable ordinal) is an ordinal β greater than 1 such that αβ = β whenever 0 < α < β. These consist of the ordinal 2 and the ordinals of the form β = ωωγ.

Exponentiation

The definition of exponentiation via order types is most easily explained using Von Neumann's definition of an ordinal as the set of all smaller ordinals. Then, to construct a set of order type αβ consider the set of all functions f : βα such that f(x) = 0 for all but finitely many elements xβ (essentially, we consider the functions with finite support). This set is ordered lexicographically with the least significant position first: we write f < g if and only if there exists xβ with f(x) < g(x) and f(y) = g(y) for all yβ with x < y. This is a well-ordering and hence gives an ordinal number.

The definition of exponentiation can also be given inductively (the following induction is on β, the exponent):

Both definitions simplify considerably if the exponent β is a finite number: αβ is then just the product of β copies of α; e.g. ω3 = ω·ω·ω, and the elements of ω3 can be viewed as triples of natural numbers, ordered lexicographically. This agrees with the ordinary exponentiation of natural numbers.

But for infinite exponents, the definition may not be obvious. For example, αω can be identified with a set of finite sequences of elements of α, properly ordered. The equation 2ω = ω expresses the fact that finite sequences of zeros and ones can be identified with natural numbers, using the binary number system. The ordinal ωω can be viewed as the order type of finite sequences of natural numbers; every element of ωω (i.e. every ordinal smaller than ωω) can be uniquely written in the form where k, n1, ..., nk are natural numbers, c1, ..., ck are nonzero natural numbers, and n1 > ... > nk.

The same is true in general: every element of αβ (i.e. every ordinal smaller than αβ) can be uniquely written in the form where k is a natural number, b1, ..., bk are ordinals smaller than β with b1 > ... > bk, and a1, ..., ak are nonzero ordinals smaller than α. This expression corresponds to the function f : βα which sends bi to ai for i = 1, ..., k and sends all other elements of β to 0.

While the same exponent-notation is used for ordinal exponentiation and cardinal exponentiation, the two operations are quite different and should not be confused. The cardinal exponentiation AB is defined to be the cardinal number of the set of all functions BA, while the ordinal exponentiation αβ only contains the functions βα with finite support, typically a set of much smaller cardinality. To avoid confusing ordinal exponentiation with cardinal exponentiation, one can use symbols for ordinals (e.g. ω) in the former and symbols for cardinals (e.g. ) in the latter.

Properties

Jacobsthal showed that the only solutions of αβ = βα with αβ are given by α = β, or α = 2 and β = 4, or α is any limit ordinal and β = εα where ε is an ε-number larger than α. [1]

Beyond exponentiation

There are ordinal operations that continue the sequence begun by addition, multiplication, and exponentiation, including ordinal versions of tetration, pentation, and hexation. See also Veblen function.

Cantor normal form

Every ordinal number α can be uniquely written as , where k is a natural number, are nonzero natural numbers, and are ordinal numbers. The degenerate case α = 0 occurs when k = 0 and there are no βs nor cs. This decomposition of α is called the Cantor normal form of α, and can be considered the base-ω positional numeral system. The highest exponent is called the degree of , and satisfies . The equality applies if and only if . In that case Cantor normal form does not express the ordinal in terms of smaller ones; this can happen as explained below.

A minor variation of Cantor normal form, which is usually slightly easier to work with, is to set all the numbers ci equal to 1 and allow the exponents to be equal. In other words, every ordinal number α can be uniquely written as , where k is a natural number, and are ordinal numbers.

Another variation of the Cantor normal form is the "base δ expansion", where ω is replaced by any ordinal δ > 1, and the numbers ci are nonzero ordinals less than δ.

The Cantor normal form allows us to uniquely expressand orderthe ordinals α that are built from the natural numbers by a finite number of arithmetical operations of addition, multiplication and exponentiation base-: in other words, assuming in the Cantor normal form, we can also express the exponents in Cantor normal form, and making the same assumption for the as for α and so on recursively, we get a system of notation for these ordinals (for example,

denotes an ordinal).

The ordinal ε0 (epsilon nought) is the set of ordinal values α of the finite-length arithmetical expressions of Cantor normal form that are hereditarily non-trivial where non-trivial means β1<α when 0<α. It is the smallest ordinal that does not have a finite arithmetical expression in terms of ω, and the smallest ordinal such that , i.e. in Cantor normal form the exponent is not smaller than the ordinal itself. It is the limit of the sequence

The ordinal ε0 is important for various reasons in arithmetic (essentially because it measures the proof-theoretic strength of the first-order Peano arithmetic: that is, Peano's axioms can show transfinite induction up to any ordinal less than ε0 but not up to ε0 itself).

The Cantor normal form also allows us to compute sums and products of ordinals: to compute the sum, for example, one need merely know (see the properties listed in § Addition and § Multiplication) that

if (if one can apply the distributive law on the left and rewrite this as , and if the expression is already in Cantor normal form); and to compute products, the essential facts are that when is in Cantor normal form and , then

and

if n is a non-zero natural number.

To compare two ordinals written in Cantor normal form, first compare , then , then , then , and so on. At the first occurrence of inequality, the ordinal that has the larger component is the larger ordinal. If they are the same until one terminates before the other, then the one that terminates first is smaller.

Factorization into primes

Ernst Jacobsthal showed that the ordinals satisfy a form of the unique factorization theorem: every nonzero ordinal can be written as a product of a finite number of prime ordinals. This factorization into prime ordinals is in general not unique, but there is a "minimal" factorization into primes that is unique up to changing the order of finite prime factors ( Sierpiński 1958 ).

A prime ordinal is an ordinal greater than 1 that cannot be written as a product of two smaller ordinals. Some of the first primes are 2, 3, 5, ... , ω, ω + 1, ω2 + 1, ω3 + 1, ..., ωω, ωω + 1, ωω + 1 + 1, ... There are three sorts of prime ordinals:

Factorization into primes is not unique: for example, 2×3 = 3×2, ω = ω, (ω+1)×ω = ω×ω and ω×ωω = ωω. However, there is a unique factorization into primes satisfying the following additional conditions:

This prime factorization can easily be read off using the Cantor normal form as follows:

So the factorization of the Cantor normal form ordinal

ωα1n1 + ⋯ + ωαknk (with α1 > ⋯ > αk)

into a minimal product of infinite primes and natural numbers is

(ωωβ1ωωβm)nk (ωαk−1−αk + 1)nk−1 ⋯ (ωα1α2 + 1)n1

where each ni should be replaced by its factorization into a non-increasing sequence of finite primes and

αk = ωβ1 + ⋯ + ωβm with β1 ≥ ⋯ ≥ βm.

Large countable ordinals

As discussed above, the Cantor normal form of ordinals below ε0 can be expressed in an alphabet containing only the function symbols for addition, multiplication and exponentiation, as well as constant symbols for each natural number and for ω. We can do away with the infinitely many numerals by using just the constant symbol 0 and the operation of successor, S (for example, the natural number 4 may be expressed as S(S(S(S(0))))). This describes an ordinal notation : a system for naming ordinals over a finite alphabet. This particular system of ordinal notation is called the collection of arithmetical ordinal expressions, and can express all ordinals below ε0, but cannot express ε0. There are other ordinal notations capable of capturing ordinals well past ε0, but because there are only countably many finite-length strings over any finite alphabet, for any given ordinal notation there will be ordinals below ω1 (the first uncountable ordinal) that are not expressible. Such ordinals are known as large countable ordinals.

The operations of addition, multiplication and exponentiation are all examples of primitive recursive ordinal functions, and more general primitive recursive ordinal functions can be used to describe larger ordinals.

Natural operations

The natural sum and natural product operations on ordinals were defined in 1906 by Gerhard Hessenberg, and are sometimes called the Hessenberg sum (or product) ( Sierpiński 1958 ). The natural sum of α and β is often denoted by αβ or α # β, and the natural product by αβ or αβ.

The natural sum and product are defined as follows. Let and be in Cantor normal form (i.e. and ). Let be the exponents sorted in nonincreasing order. Then is defined as

The natural product of and is defined as

For example, suppose and . Then , whereas . And , whereas .

The natural operations come up in the theory of well partial orders; given two well partial orders and , of ordinal types and , the ordinal type of the disjoint union is , while the ordinal type of the direct product is . [2]

The natural sum and product are commutative and associative, and natural product distributes over natural sum. The operations are also monotonic, in the sense that if then ; if then ; and if and then .

We have .

We always have and . If both and then . If both and then .

Natural sum and product are not continuous in the right argument, since, for example , and not ; and , and not .

The natural sum and product are the same as the addition and multiplication (restricted to ordinals) of John Conway's field of surreal numbers.

The natural operations come up in the theory of well partial orders; given two well partial orders S and T, of types (maximum linearizations) o(S) and o(T), the type of the disjoint union is o(S) ⊕ o(T), while the type of the direct product is o(S) ⊗ o(T). [3] One may take this relation as a definition of the natural operations by choosing S and T to be ordinals α and β; so αβ is the maximum order type of a total order extending the disjoint union (as a partial order) of α and β; while αβ is the maximum order type of a total order extending the direct product (as a partial order) of α and β. [4] A useful application of this is when α and β are both subsets of some larger total order; then their union has order type at most αβ. If they are both subsets of some ordered abelian group, then their sum has order type at most αβ.

We can also define the natural sum of α and β inductively (by simultaneous induction on α and β) as the smallest ordinal greater than the natural sum of α and γ for all γ < β and of γ and β for all γ < α. There is also an inductive definition of the natural product (by mutual induction), but it is somewhat tedious to write down and we shall not do so (see the article on surreal numbers for the definition in that context, which, however, uses surreal subtraction, something that obviously cannot be defined on ordinals).

The natural sum is associative and commutative. It is always greater or equal to the usual sum, but it may be strictly greater. For example, the natural sum of ω and 1 is ω + 1 (the usual sum), but this is also the natural sum of 1 and ω. The natural product is associative and commutative and distributes over the natural sum. The natural product is always greater or equal to the usual product, but it may be strictly greater. For example, the natural product of ω and 2 is ω · 2 (the usual product), but this is also the natural product of 2 and ω.

Under natural addition, the ordinals can be identified with the elements of the free commutative monoid generated by the gamma numbers ωα. Under natural addition and multiplication, the ordinals can be identified with the elements of the free commutative semiring generated by the delta numbers ωωα. The ordinals do not have unique factorization into primes under the natural product. While the full polynomial ring does have unique factorization, the subset of polynomials with non-negative coefficients does not: for example, if x is any delta number, then

x5 + x4 + x3 + x2 + x + 1 = (x + 1) (x4 + x2 + 1) = (x2 + x + 1) (x3 + 1)

has two incompatible expressions as a natural product of polynomials with non-negative coefficients that cannot be decomposed further.

Nimber arithmetic

There are arithmetic operations on ordinals by virtue of the one-to-one correspondence between ordinals and nimbers. Three common operations on nimbers are nimber addition, nimber multiplication, and minimum excludance (mex). Nimber addition is a generalization of the bitwise exclusive or operation on natural numbers. The mex of a set of ordinals is the smallest ordinal not present in the set.

Notes

  1. Ernst Jacobsthal, Vertauschbarkeit transfiniter Ordnungszahlen, Mathematische Annalen, Bd 64 (1907), 475-488. Available here
  2. D. H. J. De Jongh and R. Parikh, Well-partial orderings and hierarchies, Indag. Math. 39 (1977), 195–206. Available here
  3. D. H. J. De Jongh and R. Parikh, Well-partial orderings and hierarchies, Indag. Math. 39 (1977), 195–206. Available here
  4. Philip W. Carruth, Arithmetic of ordinals with applications to the theory of ordered Abelian groups, Bull. Amer. Math. Soc. 48 (1942), 262–271. See Theorem 1. Available here

Related Research Articles

The propagation constant of a sinusoidal electromagnetic wave is a measure of the change undergone by the amplitude and phase of the wave as it propagates in a given direction. The quantity being measured can be the voltage, the current in a circuit, or a field vector such as electric field strength or flux density. The propagation constant itself measures the dimensionless change in magnitude or phase per unit length. In the context of two-port networks and their cascades, propagation constant measures the change undergone by the source quantity as it propagates from one port to the next.

In mathematics, the nimbers, also called Grundy numbers, are introduced in combinatorial game theory, where they are defined as the values of heaps in the game Nim. The nimbers are the ordinal numbers endowed with nimber addition and nimber multiplication, which are distinct from ordinal addition and ordinal multiplication.

<span class="mw-page-title-main">Limit ordinal</span> Infinite ordinal number class

In set theory, a limit ordinal is an ordinal number that is neither zero nor a successor ordinal. Alternatively, an ordinal λ is a limit ordinal if there is an ordinal less than λ, and whenever β is an ordinal less than λ, then there exists an ordinal γ such that β < γ < λ. Every ordinal number is either zero, or a successor ordinal, or a limit ordinal.

In mathematics, the epsilon numbers are a collection of transfinite numbers whose defining property is that they are fixed points of an exponential map. Consequently, they are not reachable from 0 via a finite series of applications of the chosen exponential map and of "weaker" operations like addition and multiplication. The original epsilon numbers were introduced by Georg Cantor in the context of ordinal arithmetic; they are the ordinal numbers ε that satisfy the equation

In the mathematical discipline of set theory, there are many ways of describing specific countable ordinals. The smallest ones can be usefully and non-circularly expressed in terms of their Cantor normal forms. Beyond that, many ordinals of relevance to proof theory still have computable ordinal notations. However, it is not possible to decide effectively whether a given putative ordinal notation is a notation or not ; various more-concrete ways of defining ordinals that definitely have notations are available.

In mathematical logic and set theory, an ordinal notation is a partial function mapping the set of all finite sequences of symbols, themselves members of a finite alphabet, to a countable set of ordinals. A Gödel numbering is a function mapping the set of well-formed formulae of some formal language to the natural numbers. This associates each well-formed formula with a unique natural number, called its Gödel number. If a Gödel numbering is fixed, then the subset relation on the ordinals induces an ordering on well-formed formulae which in turn induces a well-ordering on the subset of natural numbers. A recursive ordinal notation must satisfy the following two additional properties:

  1. the subset of natural numbers is a recursive set
  2. the induced well-ordering on the subset of natural numbers is a recursive relation

In mathematics, particularly set theory, non-recursive ordinals are large countable ordinals greater than all the recursive ordinals, and therefore can not be expressed using recursive ordinal notations.

In mathematics, the Veblen functions are a hierarchy of normal functions, introduced by Oswald Veblen in Veblen (1908). If φ0 is any normal function, then for any non-zero ordinal α, φα is the function enumerating the common fixed points of φβ for β<α. These functions are all normal.

In mathematical logic and set theory, an ordinal collapsing function is a technique for defining certain recursive large countable ordinals, whose principle is to give names to certain ordinals much larger than the one being defined, perhaps even large cardinals, and then "collapse" them down to a system of notations for the sought-after ordinal. For this reason, ordinal collapsing functions are described as an impredicative manner of naming ordinals.

In set theory, a branch of mathematics, an additively indecomposable ordinalα is any ordinal number that is not 0 such that for any , we have Additively indecomposable ordinals were named the gamma numbers by Cantor,p.20 and are also called additive principal numbers. The class of additively indecomposable ordinals may be denoted , from the German "Hauptzahl". The additively indecomposable ordinals are precisely those ordinals of the form for some ordinal .

A synchronous frame is a reference frame in which the time coordinate defines proper time for all co-moving observers. It is built by choosing some constant time hypersurface as an origin, such that has in every point a normal along the time line and a light cone with an apex in that point can be constructed; all interval elements on this hypersurface are space-like. A family of geodesics normal to this hypersurface are drawn and defined as the time coordinates with a beginning at the hypersurface. In terms of metric-tensor components , a synchronous frame is defined such that

In continuum mechanics, a compatible deformation tensor field in a body is that unique tensor field that is obtained when the body is subjected to a continuous, single-valued, displacement field. Compatibility is the study of the conditions under which such a displacement field can be guaranteed. Compatibility conditions are particular cases of integrability conditions and were first derived for linear elasticity by Barré de Saint-Venant in 1864 and proved rigorously by Beltrami in 1886.

<span class="mw-page-title-main">Ordinal number</span> Generalization of "n-th" to infinite cases

In set theory, an ordinal number, or ordinal, is a generalization of ordinal numerals aimed to extend enumeration to infinite sets.

In mathematics, Ricci calculus constitutes the rules of index notation and manipulation for tensors and tensor fields on a differentiable manifold, with or without a metric tensor or connection. It is also the modern name for what used to be called the absolute differential calculus, developed by Gregorio Ricci-Curbastro in 1887–1896, and subsequently popularized in a paper written with his pupil Tullio Levi-Civita in 1900. Jan Arnoldus Schouten developed the modern notation and formalism for this mathematical framework, and made contributions to the theory, during its applications to general relativity and differential geometry in the early twentieth century.

In mathematics, differential forms on a Riemann surface are an important special case of the general theory of differential forms on smooth manifolds, distinguished by the fact that the conformal structure on the Riemann surface intrinsically defines a Hodge star operator on 1-forms without specifying a Riemannian metric. This allows the use of Hilbert space techniques for studying function theory on the Riemann surface and in particular for the construction of harmonic and holomorphic differentials with prescribed singularities. These methods were first used by Hilbert (1909) in his variational approach to the Dirichlet principle, making rigorous the arguments proposed by Riemann. Later Weyl (1940) found a direct approach using his method of orthogonal projection, a precursor of the modern theory of elliptic differential operators and Sobolev spaces. These techniques were originally applied to prove the uniformization theorem and its generalization to planar Riemann surfaces. Later they supplied the analytic foundations for the harmonic integrals of Hodge (1941). This article covers general results on differential forms on a Riemann surface that do not rely on any choice of Riemannian structure.

<span class="mw-page-title-main">Joos–Weinberg equation</span> Equation for arbitrary spin particles

In relativistic quantum mechanics and quantum field theory, the Joos–Weinberg equation is a relativistic wave equation applicable to free particles of arbitrary spin j, an integer for bosons or half-integer for fermions. The solutions to the equations are wavefunctions, mathematically in the form of multi-component spinor fields. The spin quantum number is usually denoted by s in quantum mechanics, however in this context j is more typical in the literature.

Buchholz's psi-functions are a hierarchy of single-argument ordinal functions introduced by German mathematician Wilfried Buchholz in 1986. These functions are a simplified version of the -functions, but nevertheless have the same strength as those. Later on this approach was extended by Jäger and Schütte.

In mathematics, Rathjen's  psi function is an ordinal collapsing function developed by Michael Rathjen. It collapses weakly Mahlo cardinals to generate large countable ordinals. A weakly Mahlo cardinal is a cardinal such that the set of regular cardinals below is closed under . Rathjen uses this to diagonalise over the weakly inaccessible hierarchy.

In complex geometry, the lemma is a mathematical lemma about the de Rham cohomology class of a complex differential form. The -lemma is a result of Hodge theory and the Kähler identities on a compact Kähler manifold. Sometimes it is also known as the -lemma, due to the use of a related operator , with the relation between the two operators being and so .

References