Primitive recursive function

Last updated

In computability theory, a primitive recursive function is, roughly speaking, a function that can be computed by a computer program whose loops are all "for" loops (that is, an upper bound of the number of iterations of every loop can be determined before entering the loop). Primitive recursive functions form a strict subset of those general recursive functions that are also total functions.

Contents

The importance of primitive recursive functions lies in the fact that most computable functions that are studied in number theory (and more generally in mathematics) are primitive recursive. For example, addition and division, the factorial and exponential function, and the function which returns the nth prime are all primitive recursive. [1] In fact, for showing that a computable function is primitive recursive, it suffices to show that its time complexity is bounded above by a primitive recursive function of the input size. [2] It is hence not particularly easy to devise a computable function that is not primitive recursive; some examples are shown in section § Limitations below.

The set of primitive recursive functions is known as PR in computational complexity theory.

Definition

A primitive recursive function takes a fixed number of arguments, each a natural number (nonnegative integer: {0, 1, 2, ...}), and returns a natural number. If it takes n arguments it is called n-ary.

The basic primitive recursive functions are given by these axioms:

  1. Constant functions Ck
    n
    : For each natural number n and every k, the k-ary constant function, defined by , is primitive recursive.
  2. Successor function: The 1-ary successor function S, which returns the successor of its argument (see Peano postulates), that is, , is primitive recursive.
  3. Projection function: For all natural numbers such that , the k-ary function defined by is primitive recursive.

More complex primitive recursive functions can be obtained by applying the operations given by these axioms:

  1. Composition operator (also called the substitution operator): Given an m-ary function and mk-ary functions :
  2. Primitive recursion operatorρ: Given the k-ary function and the (k + 2)-ary function :

    Interpretation:

    The function f acts as a for-loop from 0 up to the value of its first argument. The rest of the arguments for f, denoted here with x1, ..., xk, are a set of initial conditions for the for-loop which may be used by it during calculations but which are immutable by it. The functions g and h on the right-hand side of the equations that define f represent the body of the loop, which performs calculations. The function g is used only once to perform initial calculations. Calculations for subsequent steps of the loop are performed by h. The first parameter of h is fed the "current" value of the for-loop's index. The second parameter of h is fed the result of the for-loop's previous calculations, from previous steps. The rest of the parameters for h are those immutable initial conditions for the for-loop mentioned earlier. They may be used by h to perform calculations but they will not themselves be altered by h.

The primitive recursive functions are the basic functions and those obtained from the basic functions by applying these operations a finite number of times.

Examples

Addition

A definition of the 2-ary function , to compute the sum of its arguments, can be obtained using the primitive recursion operator . To this end, the well-known equations

are "rephrased in primitive recursive function terminology": In the definition of , the first equation suggests to choose to obtain ; the second equation suggests to choose to obtain . Therefore, the addition function can be defined as . As a computation example,

Doubling

Given , the 1-ary function doubles its argument, .

Multiplication

In a similar way as addition, multiplication can be defined by . This reproduces the well-known multiplication equations:

and

Predecessor

The predecessor function acts as the "opposite" of the successor function and is recursively defined by the rules and . A primitive recursive definition is . As a computation example,

Truncated subtraction

The limited subtraction function (also called "monus", and denoted "") is definable from the predecessor function. It satisfies the equations

Since the recursion runs over the second argument, we begin with a primitive recursive definition of the reversed subtraction, . Its recursion then runs over the first argument, so its primitive recursive definition can be obtained, similar to addition, as . To get rid of the reversed argument order, then define . As a computation example,

Converting predicates to numeric functions

In some settings it is natural to consider primitive recursive functions that take as inputs tuples that mix numbers with truth values (that is t for true and f for false), or that produce truth values as outputs. [3] This can be accomplished by identifying the truth values with numbers in any fixed manner. For example, it is common to identify the truth value t with the number 1 and the truth value f with the number 0. Once this identification has been made, the characteristic function of a set A, which always returns 1 or 0, can be viewed as a predicate that tells whether a number is in the set A. Such an identification of predicates with numeric functions will be assumed for the remainder of this article.

Predicate "Is zero"

As an example for a primitive recursive predicate, the 1-ary function shall be defined such that if , and , otherwise. This can be achieved by defining . Then, and e.g. .

Predicate "Less or equal"

Using the property , the 2-ary function can be defined by . Then if , and , otherwise. As a computation example,

Predicate "Greater or equal"

Once a definition of is obtained, the converse predicate can be defined as . Then, is true (more precisely: has value 1) if, and only if, .

If-then-else

The 3-ary if-then-else operator known from programming languages can be defined by . Then, for arbitrary ,

and

.

That is, returns the then-part, , if the if-part, , is true, and the else-part, , otherwise.

Junctors

Based on the function, it is easy to define logical junctors. For example, defining , one obtains , that is, is true if, and only if, both and are true (logical conjunction of and ).

Similarly, and lead to appropriate definitions of disjunction and negation: and .

Equality predicate

Using the above functions , and , the definition implements the equality predicate. In fact, is true if, and only if, equals .

Similarly, the definition implements the predicate "less-than", and implements "greater-than".

Other operations on natural numbers

Exponentiation and primality testing are primitive recursive. Given primitive recursive functions e, f, g, and h, a function that returns the value of g when ef and the value of h otherwise is primitive recursive.

Operations on integers and rational numbers

By using Gödel numberings, the primitive recursive functions can be extended to operate on other objects such as integers and rational numbers. If integers are encoded by Gödel numbers in a standard way, the arithmetic operations including addition, subtraction, and multiplication are all primitive recursive. Similarly, if the rationals are represented by Gödel numbers then the field operations are all primitive recursive.

Some common primitive recursive functions

The following examples and definitions are from Kleene (1952) pp. 223–231. Many appear with proofs. Most also appear with similar names, either as proofs or as examples, in Boolos-Burgess-Jeffrey 2002 pp. 63–70; they add the logarithm lo(x, y) or lg(x, y) depending on the exact derivation.

In the following the mark " ' ", e.g. a', is the primitive mark meaning "the successor of", usually thought of as " +1", e.g. a +1 =def a'. The functions 16–20 and #G are of particular interest with respect to converting primitive recursive predicates to, and extracting them from, their "arithmetical" form expressed as Gödel numbers.

  1. Addition: a+b
  2. Multiplication: a×b
  3. Exponentiation: ab
  4. Factorial a! : 0! = 1, a'! = a!×a'
  5. pred(a): (Predecessor or decrement): If a > 0 then a−1 else 0
  6. Proper subtraction a ∸ b: If a ≥ b then a−b else 0
  7. Minimum(a1, ... an)
  8. Maximum(a1, ... an)
  9. Absolute difference: | a−b | =def (a ∸ b) + (b ∸ a)
  10. ~sg(a): NOT[signum(a)]: If a=0 then 1 else 0
  11. sg(a): signum(a): If a=0 then 0 else 1
  12. a | b: (a divides b): If b=k×a for some k then 0 else 1
  13. Remainder(a, b): the leftover if b does not divide a "evenly". Also called MOD(a, b)
  14. a = b: sg | a − b | (Kleene's convention was to represent true by 0 and false by 1; presently, especially in computers, the most common convention is the reverse, namely to represent true by 1 and false by 0, which amounts to changing sg into ~sg here and in the next item)
  15. a < b: sg( a' ∸ b )
  16. Pr(a): a is a prime number Pr(a) =def a>1 & NOT(Exists c)1<c<a [ c|a ]
  17. pi: the i+1th prime number
  18. (a)i: exponent of pi in a: the unique x such that pix|a & NOT(pix'|a)
  19. lh(a): the "length" or number of non-vanishing exponents in a
  20. lo(a, b): (logarithm of a to base b): If a, b > 1 then the greatest x such that bx | a else 0
In the following, the abbreviation x =def x1, ... xn; subscripts may be applied if the meaning requires.
  • NOT_Q(x) .
  • Q OR R: Q(x) V R(x),
  • Q AND R: Q(x) & R(x),
  • Q IMPLIES R: Q(x) → R(x)
  • Q is equivalent to R: Q(x) ≡ R(x)
  • (Ey)y<z R(x, y) where (Ey)y<z denotes "there exists at least one y that is less than z such that"
  • (y)y<z R(x, y) where (y)y<z denotes "for all y less than z it is true that"
  • μyy<z R(x, y). The operator μyy<z R(x, y) is a bounded form of the so-called minimization- or mu-operator: Defined as "the least value of y less than z such that R(x, y) is true; or z if there is no such value."
φ(x) =
  • φ1(x) if Q1(x) is true,
  • . . . . . . . . . . . . . . . . . . .
  • φm(x) if Qm(x) is true
  • φm+1(x) otherwise
φ(y,x) = χ(y, COURSE-φ(y; x2, ... xn ), x2, ... xn then φ is primitive recursive in χ. The value COURSE-φ(y; x2 to n ) of the course-of-values function encodes the sequence of values φ(0,x2 to n), ..., φ(y-1,x2 to n) of the original function.

Relationship to recursive functions

The broader class of partial recursive functions is defined by introducing an unbounded search operator. The use of this operator may result in a partial function, that is, a relation with at most one value for each argument, but does not necessarily have any value for any argument (see domain). An equivalent definition states that a partial recursive function is one that can be computed by a Turing machine. A total recursive function is a partial recursive function that is defined for every input.

Every primitive recursive function is total recursive, but not all total recursive functions are primitive recursive. The Ackermann function A(m,n) is a well-known example of a total recursive function (in fact, provable total), that is not primitive recursive. There is a characterization of the primitive recursive functions as a subset of the total recursive functions using the Ackermann function. This characterization states that a function is primitive recursive if and only if there is a natural number m such that the function can be computed by a Turing machine that always halts within A(m,n) or fewer steps, where n is the sum of the arguments of the primitive recursive function. [4]

An important property of the primitive recursive functions is that they are a recursively enumerable subset of the set of all total recursive functions (which is not itself recursively enumerable). This means that there is a single computable function f(m,n) that enumerates the primitive recursive functions, namely:

f can be explicitly constructed by iteratively repeating all possible ways of creating primitive recursive functions. Thus, it is provably total. One can use a diagonalization argument to show that f is not recursive primitive in itself: had it been such, so would be h(n) = f(n,n)+1. But if this equals some primitive recursive function, there is an m such that h(n) = f(m,n) for all n, and then h(m) = f(m,m), leading to contradiction.

However, the set of primitive recursive functions is not the largest recursively enumerable subset of the set of all total recursive functions. For example, the set of provably total functions (in Peano arithmetic) is also recursively enumerable, as one can enumerate all the proofs of the theory. While all primitive recursive functions are provably total, the converse is not true.

Limitations

Primitive recursive functions tend to correspond very closely with our intuition of what a computable function must be. Certainly the initial functions are intuitively computable (in their very simplicity), and the two operations by which one can create new primitive recursive functions are also very straightforward. However, the set of primitive recursive functions does not include every possible total computable function—this can be seen with a variant of Cantor's diagonal argument. This argument provides a total computable function that is not primitive recursive. A sketch of the proof is as follows:

The primitive recursive functions of one argument (i.e., unary functions) can be computably enumerated. This enumeration uses the definitions of the primitive recursive functions (which are essentially just expressions with the composition and primitive recursion operations as operators and the basic primitive recursive functions as atoms), and can be assumed to contain every definition once, even though a same function will occur many times on the list (since many definitions define the same function; indeed simply composing by the identity function generates infinitely many definitions of any one primitive recursive function). This means that the n-th definition of a primitive recursive function in this enumeration can be effectively determined from n. Indeed if one uses some Gödel numbering to encode definitions as numbers, then this n-th definition in the list is computed by a primitive recursive function of n. Let fn denote the unary primitive recursive function given by this definition.

Now define the "evaluator function" ev with two arguments, by ev(i,j) = fi(j). Clearly ev is total and computable, since one can effectively determine the definition of fi, and being a primitive recursive function fi is itself total and computable, so fi(j) is always defined and effectively computable. However a diagonal argument will show that the function ev of two arguments is not primitive recursive.

Suppose ev were primitive recursive, then the unary function g defined by g(i) = S(ev(i,i)) would also be primitive recursive, as it is defined by composition from the successor function and ev. But then g occurs in the enumeration, so there is some number n such that g = fn. But now g(n) = S(ev(n,n)) = S(fn(n)) = S(g(n)) gives a contradiction.

This argument can be applied to any class of computable (total) functions that can be enumerated in this way, as explained in the article Machine that always halts. Note however that the partial computable functions (those that need not be defined for all arguments) can be explicitly enumerated, for instance by enumerating Turing machine encodings.

Other examples of total recursive but not primitive recursive functions are known:

Variants

Constant functions

Instead of , alternative definitions use just one 0-ary zero function as a primitive function that always returns zero, and built the constant functions from the zero function, the successor function and the composition operator.

Weak primitive recursion

The 1-place predecessor function is primitive recursive, see section #Predecessor. Fischer, Fischer & Beigel [5] removed the implicit predecessor from the recursion rule, replacing it by the weaker rule

They proved that the predecessor function still could be defined, and hence that "weak" primitive recursion also defines the primitive recursive functions.

Iterative functions

Weakening this even further by using functions of arity k+1, removing and from the arguments of completely, we get the iteration rule:

The class of iterative functions is defined the same way as the class of primitive recursive functions except with this weaker rule. These are conjectured to be a proper subset of the primitive recursive functions. [5]

Additional primitive recursive forms

Some additional forms of recursion also define functions that are in fact primitive recursive. Definitions in these forms may be easier to find or more natural for reading or writing. Course-of-values recursion defines primitive recursive functions. Some forms of mutual recursion also define primitive recursive functions.

The functions that can be programmed in the LOOP programming language are exactly the primitive recursive functions. This gives a different characterization of the power of these functions. The main limitation of the LOOP language, compared to a Turing-complete language, is that in the LOOP language the number of times that each loop will run is specified before the loop begins to run.

Computer language definition

An example of a primitive recursive programming language is one that contains basic arithmetic operators (e.g. + and −, or ADD and SUBTRACT), conditionals and comparison (IF-THEN, EQUALS, LESS-THAN), and bounded loops, such as the basic for loop, where there is a known or calculable upper bound to all loops (FOR i FROM 1 TO n, with neither i nor n modifiable by the loop body). No control structures of greater generality, such as while loops or IF-THEN plus GOTO, are admitted in a primitive recursive language.

The LOOP language, introduced in a 1967 paper by Albert R. Meyer and Dennis M. Ritchie, [6] is such a language. Its computing power coincides with the primitive recursive functions. A variant of the LOOP language is Douglas Hofstadter's BlooP in Gödel, Escher, Bach . Adding unbounded loops (WHILE, GOTO) makes the language general recursive and Turing-complete, as are all real-world computer programming languages.

The definition of primitive recursive functions implies that their computation halts on every input (after a finite number of steps). On the other hand, the halting problem is undecidable for general recursive functions.

Finitism and consistency results

The primitive recursive functions are closely related to mathematical finitism, and are used in several contexts in mathematical logic where a particularly constructive system is desired. Primitive recursive arithmetic (PRA), a formal axiom system for the natural numbers and the primitive recursive functions on them, is often used for this purpose.

PRA is much weaker than Peano arithmetic, which is not a finitistic system. Nevertheless, many results in number theory and in proof theory can be proved in PRA. For example, Gödel's incompleteness theorem can be formalized into PRA, giving the following theorem:

If T is a theory of arithmetic satisfying certain hypotheses, with Gödel sentence GT, then PRA proves the implication Con(T)→GT.

Similarly, many of the syntactic results in proof theory can be proved in PRA, which implies that there are primitive recursive functions that carry out the corresponding syntactic transformations of proofs.

In proof theory and set theory, there is an interest in finitistic consistency proofs, that is, consistency proofs that themselves are finitistically acceptable. Such a proof establishes that the consistency of a theory T implies the consistency of a theory S by producing a primitive recursive function that can transform any proof of an inconsistency from S into a proof of an inconsistency from T. One sufficient condition for a consistency proof to be finitistic is the ability to formalize it in PRA. For example, many consistency results in set theory that are obtained by forcing can be recast as syntactic proofs that can be formalized in PRA.

History

Recursive definitions had been used more or less formally in mathematics before, but the construction of primitive recursion is traced back to Richard Dedekind's theorem 126 of his Was sind und was sollen die Zahlen? (1888). This work was the first to give a proof that a certain recursive construction defines a unique function. [7] [8] [9]

Primitive recursive arithmetic was first proposed by Thoralf Skolem [10] in 1923.

The current terminology was coined by Rózsa Péter (1934) after Ackermann had proved in 1928 that the function which today is named after him was not primitive recursive, an event which prompted the need to rename what until then were simply called recursive functions. [8] [9]

See also

Notes

  1. Brainerd and Landweber, 1974
  2. Hartmanis, 1989
  3. Kleene [1952 pp. 226–227]
  4. This follows from the facts that the functions of this form are the most quickly growing primitive recursive functions, and that a function is primitive recursive if and only if its time complexity is bounded by a primitive recursive function. For the former, see Linz, Peter (2011), An Introduction to Formal Languages and Automata, Jones & Bartlett Publishers, p. 332, ISBN   9781449615529 . For the latter, see Moore, Cristopher; Mertens, Stephan (2011), The Nature of Computation, Oxford University Press, p. 287, ISBN   9780191620805
  5. 1 2 Fischer, Fischer & Beigel 1989.
  6. Meyer, Albert R.; Ritchie, Dennis M. (1967). The complexity of loop programs. ACM '67: Proceedings of the 1967 22nd national conference. doi: 10.1145/800196.806014 .
  7. Peter Smith (2013). An Introduction to Gödel's Theorems (2nd ed.). Cambridge University Press. pp. 98–99. ISBN   978-1-107-02284-3.
  8. 1 2 George Tourlakis (2003). Lectures in Logic and Set Theory: Volume 1, Mathematical Logic. Cambridge University Press. p. 129. ISBN   978-1-139-43942-8.
  9. 1 2 Rod Downey, ed. (2014). Turing's Legacy: Developments from Turing's Ideas in Logic. Cambridge University Press. p. 474. ISBN   978-1-107-04348-0.
  10. Thoralf Skolem (1923) "The foundations of elementary arithmetic" in Jean van Heijenoort, translator and ed. (1967) From Frege to Gödel: A Source Book in Mathematical Logic, 1879-1931. Harvard Univ. Press: 302-33.

Related Research Articles

In mathematical logic and computer science, a general recursive function, partial recursive function, or μ-recursive function is a partial function from natural numbers to natural numbers that is "computable" in an intuitive sense – as well as in a formal one. If the function is total, it is also called a total recursive function. In computability theory, it is shown that the μ-recursive functions are precisely the functions that can be computed by Turing machines. The μ-recursive functions are closely related to primitive recursive functions, and their inductive definition (below) builds upon that of the primitive recursive functions. However, not every total recursive function is a primitive recursive function—the most famous example is the Ackermann function.

<span class="mw-page-title-main">Exponentiation</span> Arithmetic operation

In mathematics, exponentiation is an operation involving two numbers: the base and the exponent or power. Exponentiation is written as bn, where b is the base and n is the power; this is pronounced as "b (raised) to the n". When n is a positive integer, exponentiation corresponds to repeated multiplication of the base: that is, bn is the product of multiplying n bases:

In the mathematical field of differential geometry, the Riemann curvature tensor or Riemann–Christoffel tensor is the most common way used to express the curvature of Riemannian manifolds. It assigns a tensor to each point of a Riemannian manifold. It is a local invariant of Riemannian metrics which measures the failure of the second covariant derivatives to commute. A Riemannian manifold has zero curvature if and only if it is flat, i.e. locally isometric to the Euclidean space. The curvature tensor can also be defined for any pseudo-Riemannian manifold, or indeed any manifold equipped with an affine connection.

In computational complexity theory, the complexity class ELEMENTARY of elementary recursive functions is the union of the classes

<span class="mw-page-title-main">Fermi's interaction</span> Mechanism of beta decay proposed in 1933

In particle physics, Fermi's interaction is an explanation of the beta decay, proposed by Enrico Fermi in 1933. The theory posits four fermions directly interacting with one another. This interaction explains beta decay of a neutron by direct coupling of a neutron with an electron, a neutrino and a proton.

The representation theory of groups is a part of mathematics which examines how groups act on given structures.

<span class="mw-page-title-main">Dyadic transformation</span> Doubling map on the unit interval

The dyadic transformation is the mapping

In mathematics, the jet is an operation that takes a differentiable function f and produces a polynomial, the truncated Taylor polynomial of f, at each point of its domain. Although this is the definition of a jet, the theory of jets regards these polynomials as being abstract polynomials rather than polynomial functions.

In mathematics, an operad is a structure that consists of abstract operations, each one having a fixed finite number of inputs (arguments) and one output, as well as a specification of how to compose these operations. Given an operad , one defines an algebra over to be a set together with concrete operations on this set which behave just like the abstract operations of . For instance, there is a Lie operad such that the algebras over are precisely the Lie algebras; in a sense abstractly encodes the operations that are common to all Lie algebras. An operad is to its algebras as a group is to its group representations.

Limited-memory BFGS is an optimization algorithm in the family of quasi-Newton methods that approximates the Broyden–Fletcher–Goldfarb–Shanno algorithm (BFGS) using a limited amount of computer memory. It is a popular algorithm for parameter estimation in machine learning. The algorithm's target problem is to minimize over unconstrained values of the real-vector where is a differentiable scalar function.

Primitive recursive arithmetic (PRA) is a quantifier-free formalization of the natural numbers. It was first proposed by Norwegian mathematician Skolem (1923), as a formalization of his finitistic conception of the foundations of arithmetic, and it is widely agreed that all reasoning of PRA is finitistic. Many also believe that all of finitism is captured by PRA, but others believe finitism can be extended to forms of recursion beyond primitive recursion, up to ε0, which is the proof-theoretic ordinal of Peano arithmetic. PRA's proof theoretic ordinal is ωω, where ω is the smallest transfinite ordinal. PRA is sometimes called Skolem arithmetic, although that has another meaning, see Skolem arithmetic.

In the field of mathematical modeling, a radial basis function network is an artificial neural network that uses radial basis functions as activation functions. The output of the network is a linear combination of radial basis functions of the inputs and neuron parameters. Radial basis function networks have many uses, including function approximation, time series prediction, classification, and system control. They were first formulated in a 1988 paper by Broomhead and Lowe, both researchers at the Royal Signals and Radar Establishment.

In set theory, a mathematical discipline, the Jensen hierarchy or J-hierarchy is a modification of Gödel's constructible hierarchy, L, that circumvents certain technical difficulties that exist in the constructible hierarchy. The J-Hierarchy figures prominently in fine structure theory, a field pioneered by Ronald Jensen, for whom the Jensen hierarchy is named. Rudimentary functions describe a method for iterating through the Jensen hierarchy.

In mathematics, a determinantal point process is a stochastic point process, the probability distribution of which is characterized as a determinant of some function. Such processes arise as important tools in random matrix theory, combinatorics, physics, and wireless network modeling.

In optics, the Fraunhofer diffraction equation is used to model the diffraction of waves when the diffraction pattern is viewed at a long distance from the diffracting object, and also when it is viewed at the focal plane of an imaging lens.

In mathematics, infinite compositions of analytic functions (ICAF) offer alternative formulations of analytic continued fractions, series, products and other infinite expansions, and the theory evolving from such compositions may shed light on the convergence/divergence of these expansions. Some functions can actually be expanded directly as infinite compositions. In addition, it is possible to use ICAF to evaluate solutions of fixed point equations involving infinite expansions. Complex dynamics offers another venue for iteration of systems of functions rather than a single function. For infinite compositions of a single function see Iterated function. For compositions of a finite number of functions, useful in fractal theory, see Iterated function system.

<span class="mw-page-title-main">Causal fermion systems</span> Candidate unified theory of physics

The theory of causal fermion systems is an approach to describe fundamental physics. It provides a unification of the weak, the strong and the electromagnetic forces with gravity at the level of classical field theory. Moreover, it gives quantum mechanics as a limiting case and has revealed close connections to quantum field theory. Therefore, it is a candidate for a unified physical theory. Instead of introducing physical objects on a preexisting spacetime manifold, the general concept is to derive spacetime as well as all the objects therein as secondary objects from the structures of an underlying causal fermion system. This concept also makes it possible to generalize notions of differential geometry to the non-smooth setting. In particular, one can describe situations when spacetime no longer has a manifold structure on the microscopic scale. As a result, the theory of causal fermion systems is a proposal for quantum geometry and an approach to quantum gravity.

Implicit computational complexity (ICC) is a subfield of computational complexity theory that characterizes algorithms by constraints on the way in which they are constructed, without reference to a specific underlying machine model or to explicit bounds on computational resources unlike conventional complexity theory. ICC was developed in the 1990s and employs the techniques of proof theory, substructural logic, model theory and recursion theory to prove bounds on the expressive power of high-level formal languages. ICC is also concerned with the practical realization of functional programming languages, language tools and type theory that can control the resource usage of programs in a formally verifiable sense.

In mathematics and theoretical computer science, analysis of Boolean functions is the study of real-valued functions on or from a spectral perspective. The functions studied are often, but not always, Boolean-valued, making them Boolean functions. The area has found many applications in combinatorics, social choice theory, random graphs, and theoretical computer science, especially in hardness of approximation, property testing, and PAC learning.

The unique homomorphic extension theorem is a result in mathematical logic which formalizes the intuition that the truth or falsity of a statement can be deduced from the truth values of its parts.

References