Algorithm

From Wikipedia, the free encyclopedia
(Redirected from Algorithm design)

In a loop, subtract the larger number against the smaller number. Halt the loop when the subtraction will make a number negative. Assess two numbers whether one of them equal to zero or not. If yes, take the other number as the greatest common divisor. If no, put the two number in the subtraction loop again.
Flowchart of using successive subtractions to find the greatest common divisor of number r and s

In mathematics and computer science, an algorithm (/ˈælɡərɪðəm/ ) is a finite sequence of rigorous instructions, typically used to solve a class of specific problems or to perform a computation.[1] Algorithms are used as specifications for performing calculations and data processing. More advanced algorithms can use conditionals to divert the code execution through various routes (referred to as automated decision-making) and deduce valid inferences (referred to as automated reasoning), achieving automation eventually. Using human characteristics as descriptors of machines in metaphorical ways was already practiced by Alan Turing with terms such as "memory", "search" and "stimulus".[2]

In contrast, a heuristic is an approach to problem solving that may not be fully specified or may not guarantee correct or optimal results, especially in problem domains where there is no well-defined correct or optimal result.[3] For example, social media recommender systems rely on heuristics in such a way that, although widely characterized as "algorithms" in 21st century popular media, cannot deliver correct results due to the nature of the problem.

As an effective method, an algorithm can be expressed within a finite amount of space and time[4] and in a well-defined formal language[5] for calculating a function.[6] Starting from an initial state and initial input (perhaps empty),[7] the instructions describe a computation that, when executed, proceeds through a finite[8] number of well-defined successive states, eventually producing "output"[9] and terminating at a final ending state. The transition from one state to the next is not necessarily deterministic; some algorithms, known as randomized algorithms, incorporate random input.[10]

History[edit]

Ancient algorithms[edit]

Since antiquity, step-by-step procedures for solving mathematical problems have been attested. This includes Babylonian mathematics (around 2500 BC),[11] Egyptian mathematics (around 1550 BC),[11] Indian mathematics (around 800 BC and later; e.g. Shulba Sutras, Kerala School, and Brāhmasphuṭasiddhānta),[12][13] The Ifa Oracle (around 500 BC), Greek mathematics (around 240 BC, e.g. sieve of Eratosthenes and Euclidean algorithm),[14] and Arabic mathematics (9th century, e.g. cryptographic algorithms for code-breaking based on frequency analysis).[15]

Al-Khwārizmī and the term algorithm[edit]

Around 825, Iranian scientist and polymath Muḥammad ibn Mūsā al-Khwārizmī wrote kitāb al-ḥisāb al-hindī ("Book of Indian computation") and kitab al-jam' wa'l-tafriq al-ḥisāb al-hindī ("Addition and subtraction in Indian arithmetic"). Both of these texts are lost in the original Arabic at this time. (However, his other book on algebra remains.)[16]

In the early 12th century, Latin translations of said al-Khwarizmi texts involving the Hindu–Arabic numeral system and arithmetic appeared: Liber Alghoarismi de practica arismetrice (attributed to John of Seville) and Liber Algorismi de numero Indorum (attributed to Adelard of Bath).[17] Hereby, alghoarismi or algorismi is the Latinization of Al-Khwarizmi's name; the text starts with the phrase Dixit Algorismi ("Thus spoke Al-Khwarizmi").[18]

In 1240, Alexander of Villedieu writes a Latin text titled Carmen de Algorismo. It begins with:

Haec algorismus ars praesens dicitur, in qua / Talibus Indorum fruimur bis quinque figuris.

which translates to:

Algorism is the art by which at present we use those Indian figures, which number two times five.

The poem is a few hundred lines long and summarizes the art of calculating with the new styled Indian dice (Tali Indorum), or Hindu numerals.[19]

English evolution of the word[edit]

Around 1230, the English word algorism is attested and then by Chaucer in 1391. English adopted the French term.[20][21]

In the 15th century, under the influence of the Greek word ἀριθμός (arithmos, "number"; cf. "arithmetic"), the Latin word was altered to algorithmus.

In 1656, in the English dictionary Glossographia, it says:[22]

Algorism ([Latin] algorismus) the Art or use of Cyphers, or of numbering by Cyphers; skill in accounting.

Augrime ([Latin] algorithmus) skil in accounting or numbring.

In 1658, in the first edition of The New World of English Words, it says:[23]

Algorithme, (a word compounded of Arabick and Spanish,) the art of reckoning by Cyphers.

In 1706, in the sixth edition of The New World of English Words, it says:[24]

Algorithm, the Art of computing or reckoning by numbers, which contains the five principle Rules of Arithmetick, viz. Numeration, Addition, Subtraction, Multiplication and Division; to which may be added Extraction of Roots: It is also call'd Logistica Numeralis.

Algorism, the practical Operation in the several Parts of Specious Arithmetick or Algebra; sometimes it is taken for the Practice of Common Arithmetick by the ten Numeral Figures.

In 1751, in the Young Algebraist's Companion, Daniel Fenning contrasts the terms algorism and algorithm as follows:[25]

Algorithm signifies the first Principles, and Algorism the practical Part, or knowing how to put the Algorithm in Practice.

Since at least 1811, the term algorithm is attested to mean a "step-by-step procedure" in English.[26][27]

In 1842, in the Dictionary of Science, Literature and Art, it says:

ALGORITHM, signifies the art of computing in reference to some particular subject, or in some particular way; as the algorithm of numbers; the algorithm of the differential calculus.[28]

Machine usage[edit]

Ada Lovelace's diagram from "Note G", the first published computer algorithm

In 1928, a partial formalization of the modern concept of algorithms began with attempts to solve the Entscheidungsproblem (decision problem) posed by David Hilbert. Later formalizations were framed as attempts to define "effective calculability"[29] or "effective method".[30] Those formalizations included the GödelHerbrandKleene recursive functions of 1930, 1934 and 1935, Alonzo Church's lambda calculus of 1936, Emil Post's Formulation 1 of 1936, and Alan Turing's Turing machines of 1936–37 and 1939.

Informal definition[edit]

One informal definition is "a set of rules that precisely defines a sequence of operations",[31][need quotation to verify] which would include all computer programs (including programs that do not perform numeric calculations), and (for example) any prescribed bureaucratic procedure[32] or cook-book recipe.[33]

In general, a program is an algorithm only if it stops eventually[34]—even though infinite loops may sometimes prove desirable.

A prototypical example of an algorithm is the Euclidean algorithm, which is used to determine the maximum common divisor of two integers; an example (there are others) is described by the flowchart above and as an example in a later section.

Boolos, Jeffrey & 1974, 1999 offer an informal meaning of the word "algorithm" in the following quotation:

No human being can write fast enough, or long enough, or small enough† ( †"smaller and smaller without limit ... you'd be trying to write on molecules, on atoms, on electrons") to list all members of an enumerably infinite set by writing out their names, one after another, in some notation. But humans can do something equally useful, in the case of certain enumerably infinite sets: They can give explicit instructions for determining the nth member of the set, for arbitrary finite n. Such instructions are to be given quite explicitly, in a form in which they could be followed by a computing machine, or by a human who is capable of carrying out only very elementary operations on symbols.[35]

An "enumerably infinite set" is one whose elements can be put into one-to-one correspondence with the integers. Thus Boolos and Jeffrey are saying that an algorithm implies instructions for a process that "creates" output integers from an arbitrary "input" integer or integers that, in theory, can be arbitrarily large. For example, an algorithm can be an algebraic equation such as y = m + n (i.e., two arbitrary "input variables" m and n that produce an output y), but various authors' attempts to define the notion indicate that the word implies much more than this, something on the order of (for the addition example):

Precise instructions (in a language understood by "the computer")[36] for a fast, efficient, "good"[37] process that specifies the "moves" of "the computer" (machine or human, equipped with the necessary internally contained information and capabilities)[38] to find, decode, and then process arbitrary input integers/symbols m and n, symbols + and = ... and "effectively"[39] produce, in a "reasonable" time,[40] output-integer y at a specified place and in a specified format.

The concept of algorithm is also used to define the notion of decidability—a notion that is central for explaining how formal systems come into being starting from a small set of axioms and rules. In logic, the time that an algorithm requires to complete cannot be measured, as it is not apparently related to the customary physical dimension. From such uncertainties, that characterize ongoing work, stems the unavailability of a definition of algorithm that suits both concrete (in some sense) and abstract usage of the term.

Most algorithms are intended to be implemented as computer programs. However, algorithms are also implemented by other means, such as in a biological neural network (for example, the human brain implementing arithmetic or an insect looking for food), in an electrical circuit, or in a mechanical device.

Formalization[edit]

Algorithms are essential to the way computers process data. Many computer programs contain algorithms that detail the specific instructions a computer should perform—in a specific order—to carry out a specified task, such as calculating employees' paychecks or printing students' report cards. Thus, an algorithm can be considered to be any sequence of operations that can be simulated by a Turing-complete system. Authors who assert this thesis include Minsky (1967), Savage (1987), and Gurevich (2000):

Minsky: "But we will also maintain, with Turing ... that any procedure which could "naturally" be called effective, can, in fact, be realized by a (simple) machine. Although this may seem extreme, the arguments ... in its favor are hard to refute".[41] Gurevich: "… Turing's informal argument in favor of his thesis justifies a stronger thesis: every algorithm can be simulated by a Turing machine … according to Savage [1987], an algorithm is a computational process defined by a Turing machine".[42]

Turing machines can define computational processes that do not terminate. The informal definitions of algorithms generally require that the algorithm always terminates. This requirement renders the task of deciding whether a formal procedure is an algorithm impossible in the general case—due to a major theorem of computability theory known as the halting problem.

Typically, when an algorithm is associated with processing information, data can be read from an input source, written to an output device and stored for further processing. Stored data is regarded as part of the internal state of the entity performing the algorithm. In practice, the state is stored in one or more data structures.

For some of these computational processes, the algorithm must be rigorously defined and specified in the way it applies in all possible circumstances that could arise. This means that any conditional steps must be systematically dealt with, case by case; the criteria for each case must be clear (and computable).

Because an algorithm is a precise list of precise steps, the order of computation is always crucial to the functioning of the algorithm. Instructions are usually assumed to be listed explicitly, and are described as starting "from the top" and going "down to the bottom"—an idea that is described more formally by flow of control.

So far, the discussion on the formalization of an algorithm has assumed the premises of imperative programming. This is the most common conception—one that attempts to describe a task in discrete, "mechanical" terms. Associated with this conception of formalized algorithms is the assignment operation, which sets the value of a variable. It derives from the intuition of "memory" as a scratchpad. An example of such an assignment can be found below.

For some alternate conceptions of what constitutes an algorithm, see functional programming and logic programming.

Expressing algorithms[edit]

Algorithms can be expressed in many kinds of notation, including natural languages, pseudocode, flowcharts, drakon-charts, programming languages or control tables (processed by interpreters). Natural language expressions of algorithms tend to be verbose and ambiguous and are rarely used for complex or technical algorithms. Pseudocode, flowcharts, drakon-charts and control tables are structured ways to express algorithms that avoid many of the ambiguities common in statements based on natural language. Programming languages are primarily intended for expressing algorithms in a form that can be executed by a computer, but they are also often used as a way to define or document algorithms.

There is a wide variety of representations possible and one can express a given Turing machine program as a sequence of machine tables (see finite-state machine, state-transition table and control table for more), as flowcharts and drakon-charts (see state diagram for more), or as a form of rudimentary machine code or assembly code called "sets of quadruples" (see Turing machine for more).

Representations of algorithms can be classified into three accepted levels of Turing machine description, as follows:[43]

1 High-level description
"...prose to describe an algorithm, ignoring the implementation details. At this level, we do not need to mention how the machine manages its tape or head."
2 Implementation description
"...prose used to define the way the Turing machine uses its head and the way that it stores data on its tape. At this level, we do not give details of states or transition function."
3 Formal description
Most detailed, "lowest level", gives the Turing machine's "state table".

For an example of the simple algorithm "Add m+n" described in all three levels, see Examples.

Design[edit]

Algorithm design refers to a method or a mathematical process for problem-solving and engineering algorithms. The design of algorithms is part of many solution theories, such as divide-and-conquer or dynamic programming within operation research. Techniques for designing and implementing algorithm designs are also called algorithm design patterns,[44] with examples including the template method pattern and the decorator pattern.

One of the most important aspects of algorithm design is resource (run-time, memory usage) efficiency; the big O notation is used to describe e.g., an algorithm's run-time growth as the size of its input increases.

Typical steps in the development of algorithms:

  1. Problem definition
  2. Development of a model
  3. Specification of the algorithm
  4. Designing an algorithm
  5. Checking the correctness of the algorithm
  6. Analysis of algorithm
  7. Implementation of algorithm
  8. Program testing
  9. Documentation preparation[clarification needed]

Computer algorithms[edit]

Logical NAND algorithm implemented electronically in 7400 chip
Flowchart examples of the canonical Böhm-Jacopini structures: the SEQUENCE (rectangles descending the page), the WHILE-DO and the IF-THEN-ELSE. The three structures are made of the primitive conditional GOTO (IF test THEN GOTO step xxx, shown as diamond), the unconditional GOTO (rectangle), various assignment operators (rectangle), and HALT (rectangle). Nesting of these structures inside assignment-blocks results in complex diagrams (cf. Tausworthe 1977:100, 114).

"Elegant" (compact) programs, "good" (fast) programs: The notion of "simplicity and elegance" appears informally in Knuth and precisely in Chaitin:

Knuth: " ... we want good algorithms in some loosely defined aesthetic sense. One criterion ... is the length of time taken to perform the algorithm .... Other criteria are adaptability of the algorithm to computers, its simplicity, and elegance, etc."[45]
Chaitin: " ... a program is 'elegant,' by which I mean that it's the smallest possible program for producing the output that it does"[46]

Chaitin prefaces his definition with: "I'll show you can't prove that a program is 'elegant'"—such a proof would solve the Halting problem (ibid).

Algorithm versus function computable by an algorithm: For a given function multiple algorithms may exist. This is true even without expanding the available instruction set available to the programmer. Rogers observes that "It is ... important to distinguish between the notion of algorithm, i.e. procedure and the notion of function computable by algorithm, i.e. mapping yielded by procedure. The same function may have several different algorithms".[47]

Unfortunately, there may be a tradeoff between goodness (speed) and elegance (compactness)—an elegant program may take more steps to complete a computation than one that is less elegant. An example that uses Euclid's algorithm appears below.

Computers (and computors), models of computation: A computer (or human "computer"[48]) is a restricted type of machine, a "discrete deterministic mechanical device"[49] that blindly follows its instructions.[50] Melzak's and Lambek's primitive models[51] reduced this notion to four elements: (i) discrete, distinguishable locations, (ii) discrete, indistinguishable counters[52] (iii) an agent, and (iv) a list of instructions that are effective relative to the capability of the agent.[53]

Minsky describes a more congenial variation of Lambek's "abacus" model in his "Very Simple Bases for Computability".[54] Minsky's machine proceeds sequentially through its five (or six, depending on how one counts) instructions unless either a conditional IF-THEN GOTO or an unconditional GOTO changes program flow out of sequence. Besides HALT, Minsky's machine includes three assignment (replacement, substitution)[55] operations: ZERO (e.g. the contents of the location replaced by 0: L ← 0), SUCCESSOR (e.g. L ← L+1), and DECREMENT (e.g. L ← L − 1).[56] Rarely must a programmer write "code" with such a limited instruction set. But Minsky shows (as do Melzak and Lambek) that his machine is Turing complete with only four general types of instructions: conditional GOTO, unconditional GOTO, assignment/replacement/substitution, and HALT. However, a few different assignment instructions (e.g. DECREMENT, INCREMENT, and ZERO/CLEAR/EMPTY for a Minsky machine) are also required for Turing-completeness; their exact specification is somewhat up to the designer. The unconditional GOTO is convenient; it can be constructed by initializing a dedicated location to zero e.g. the instruction " Z ← 0 "; thereafter the instruction IF Z=0 THEN GOTO xxx is unconditional.

Simulation of an algorithm: computer (computor) language: Knuth advises the reader that "the best way to learn an algorithm is to try it . . . immediately take pen and paper and work through an example".[57] But what about a simulation or execution of the real thing? The programmer must translate the algorithm into a language that the simulator/computer/computor can effectively execute. Stone gives an example of this: when computing the roots of a quadratic equation the computer must know how to take a square root. If they do not, then the algorithm, to be effective, must provide a set of rules for extracting a square root.[58]

This means that the programmer must know a "language" that is effective relative to the target computing agent (computer/computor).

But what model should be used for the simulation? Van Emde Boas observes "even if we base complexity theory on abstract instead of concrete machines, the arbitrariness of the choice of a model remains. It is at this point that the notion of simulation enters".[59] When speed is being measured, the instruction set matters. For example, the subprogram in Euclid's algorithm to compute the remainder would execute much faster if the programmer had a "modulus" instruction available rather than just subtraction (or worse: just Minsky's "decrement").

Structured programming, canonical structures: Per the Church–Turing thesis, any algorithm can be computed by a model known to be Turing complete, and per Minsky's demonstrations, Turing completeness requires only four instruction types—conditional GOTO, unconditional GOTO, assignment, HALT. Kemeny and Kurtz observe that, while "undisciplined" use of unconditional GOTOs and conditional IF-THEN GOTOs can result in "spaghetti code", a programmer can write structured programs using only these instructions; on the other hand "it is also possible, and not too hard, to write badly structured programs in a structured language".[60] Tausworthe augments the three Böhm-Jacopini canonical structures:[61] SEQUENCE, IF-THEN-ELSE, and WHILE-DO, with two more: DO-WHILE and CASE.[62] An additional benefit of a structured program is that it lends itself to proofs of correctness using mathematical induction.[63]

Canonical flowchart symbols[64]: The graphical aid called a flowchart offers a way to describe and document an algorithm (and a computer program corresponding to it). Like the program flow of a Minsky machine, a flowchart always starts at the top of a page and proceeds down. Its primary symbols are only four: the directed arrow showing program flow, the rectangle (SEQUENCE, GOTO), the diamond (IF-THEN-ELSE), and the dot (OR-tie). The Böhm–Jacopini canonical structures are made of these primitive shapes. Sub-structures can "nest" in rectangles, but only if a single exit occurs from the superstructure. The symbols and their use to build the canonical structures are shown in the diagram.

Examples[edit]

Algorithm example[edit]

One of the simplest algorithms is to find the largest number in a list of numbers of random order. Finding the solution requires looking at every number in the list. From this follows a simple algorithm, which can be stated in a high-level description in English prose, as:

High-level description:

  1. If there are no numbers in the set, then there is no highest number.
  2. Assume the first number in the set is the largest number in the set.
  3. For each remaining number in the set: if this number is larger than the current largest number, consider this number to be the largest number in the set.
  4. When there are no numbers left in the set to iterate over, consider the current largest number to be the largest number of the set.

(Quasi-)formal description: Written in prose but much closer to the high-level language of a computer program, the following is the more formal coding of the algorithm in pseudocode or pidgin code:

Algorithm LargestNumber
Input: A list of numbers L.
Output: The largest number in the list L.
if L.size = 0 return null
largestL[0]
for each item in L, do
    if item > largest, then
        largestitem
return largest
  • "←" denotes assignment. For instance, "largestitem" means that the value of largest changes to the value of item.
  • "return" terminates the algorithm and outputs the following value.

Euclid's algorithm[edit]

In mathematics, the Euclidean algorithm or Euclid's algorithm, is an efficient method for computing the greatest common divisor (GCD) of two integers (numbers), the largest number that divides them both without a remainder. It is named after the ancient Greek mathematician Euclid, who first described it in his Elements (c. 300 BC).[65] It is one of the oldest algorithms in common use. It can be used to reduce fractions to their simplest form, and is a part of many other number-theoretic and cryptographic calculations.

The example-diagram of Euclid's algorithm from T.L. Heath (1908), with more detail added. Euclid does not go beyond a third measuring and gives no numerical examples. Nicomachus gives the example of 49 and 21: "I subtract the less from the greater; 28 is left; then again I subtract from this the same 21 (for this is possible); 7 is left; I subtract this from 21, 14 is left; from which I again subtract 7 (for this is possible); 7 is left, but 7 cannot be subtracted from 7." Heath comments that "The last phrase is curious, but the meaning of it is obvious enough, as also the meaning of the phrase about ending 'at one and the same number'."(Heath 1908:300).

Euclid poses the problem thus: "Given two numbers not prime to one another, to find their greatest common measure". He defines "A number [to be] a multitude composed of units": a counting number, a positive integer not including zero. To "measure" is to place a shorter measuring length s successively (q times) along longer length l until the remaining portion r is less than the shorter length s.[66] In modern words, remainder r = lq×s, q being the quotient, or remainder r is the "modulus", the integer-fractional part left over after the division.[67]

For Euclid's method to succeed, the starting lengths must satisfy two requirements: (i) the lengths must not be zero, AND (ii) the subtraction must be "proper"; i.e., a test must guarantee that the smaller of the two numbers is subtracted from the larger (or the two can be equal so their subtraction yields zero).

Euclid's original proof adds a third requirement: the two lengths must not be prime to one another. Euclid stipulated this so that he could construct a reductio ad absurdum proof that the two numbers' common measure is in fact the greatest.[68] While Nicomachus' algorithm is the same as Euclid's, when the numbers are prime to one another, it yields the number "1" for their common measure. So, to be precise, the following is really Nicomachus' algorithm.

A graphical expression of Euclid's algorithm to find the greatest common divisor for 1599 and 650
 1599 = 650×2 + 299
 650 = 299×2 + 52
 299 = 52×5 + 39
 52 = 39×1 + 13
 39 = 13×3 + 0

Computer language for Euclid's algorithm[edit]

Only a few instruction types are required to execute Euclid's algorithm—some logical tests (conditional GOTO), unconditional GOTO, assignment (replacement), and subtraction.

  • A location is symbolized by upper case letter(s), e.g. S, A, etc.
  • The varying quantity (number) in a location is written in lower case letter(s) and (usually) associated with the location's name. For example, location L at the start might contain the number l = 3009.

An inelegant program for Euclid's algorithm[edit]

"Inelegant" is a translation of Knuth's version of the algorithm with a subtraction-based remainder-loop replacing his use of division (or a "modulus" instruction). Derived from Knuth 1973:2–4. Depending on the two numbers "Inelegant" may compute the g.c.d. in fewer steps than "Elegant".

The following algorithm is framed as Knuth's four-step version of Euclid's and Nicomachus', but, rather than using division to find the remainder, it uses successive subtractions of the shorter length s from the remaining length r until r is less than s. The high-level description, shown in boldface, is adapted from Knuth 1973:2–4:

INPUT:

1 [Into two locations L and S put the numbers l and s that represent the two lengths]:
INPUT L, S
2 [Initialize R: make the remaining length r equal to the starting/initial/input length l]:
R ← L

E0: [Ensure rs.]

3 [Ensure the smaller of the two numbers is in S and the larger in R]:
IF R > S THEN
the contents of L is the larger number so skip over the exchange-steps 4, 5 and 6:
GOTO step 7
ELSE
swap the contents of R and S.
4 L ← R (this first step is redundant, but is useful for later discussion).
5 R ← S
6 S ← L

E1: [Find remainder]: Until the remaining length r in R is less than the shorter length s in S, repeatedly subtract the measuring number s in S from the remaining length r in R.

7 IF S > R THEN
done measuring so
GOTO 10
ELSE
measure again,
8 R ← R − S
9 [Remainder-loop]:
GOTO 7.

E2: [Is the remainder zero?]: EITHER (i) the last measure was exact, the remainder in R is zero, and the program can halt, OR (ii) the algorithm must continue: the last measure left a remainder in R less than measuring number in S.

10 IF R = 0 THEN
done so
GOTO step 15
ELSE
CONTINUE TO step 11,

E3: [Interchange s and r]: The nut of Euclid's algorithm. Use remainder r to measure what was previously smaller number s; L serves as a temporary location.

11 L ← R
12 R ← S
13 S ← L
14 [Repeat the measuring process]:
GOTO 7

OUTPUT:

15 [Done. S contains the greatest common divisor]:
PRINT S

DONE:

16 HALT, END, STOP.

An elegant program for Euclid's algorithm[edit]

The following version of Euclid's algorithm requires only six core instructions to do what thirteen are required to do by "Inelegant"; worse, "Inelegant" requires more types of instructions.[clarify] The flowchart of "Elegant" can be found at the top of this article. In the (unstructured) Basic language, the steps are numbered, and the instruction LET [] = [] is the assignment instruction symbolized by ←.

5 REM Euclid's algorithm for greatest common divisor
6 PRINT "Type two integers greater than 0"
10 INPUT A,B
20 IF B=0 THEN GOTO 80
30 IF A > B THEN GOTO 60
40 LET B=B-A
50 GOTO 20
60 LET A=A-B
70 GOTO 20
80 PRINT A
90 END

How "Elegant" works: instead of an outer "Euclid loop", "Elegant" calculates the remainder of a division using modulo and shifts the variables A and B in each iteration. The following algorithm can be used with programming languages from the C-family:

  • The standard function abs changes negative integer to positive integer
  • When input A or B has the value 0, the algorithm stops and the result is 0
  • If input A is greater than input B, the algorithm will automatically swap variables A and B during the first iteration via modulo
  • The iteration (a do while loop) starts and only stops when the variable B is set to 0:
    • % calculates the modulo of division A and B, which reduces the number (e.g.: 23 = 3 × 6 + remainder 5)
    • A is equated with B
    • B is equated with the modulo-result
    • The iteration continues as long as B is greater than 0
  • When the iteration stops, variable A always contains the greatest common divisor
// Euclid's algorithm for greatest common divisor
int euclidAlgorithm (int A, int B) {
    A = abs(A);
    B = abs(B);
    if (A == 0 || B == 0) {
       return 0;
    }
    do { 
       int res = A % B; 
       A = B;
       B = res;
    } while (B > 0);
  return A;

Testing the Euclid algorithms[edit]

Does an algorithm do what its author wants it to do? A few test cases usually give some confidence in the core functionality. But tests are not enough. For test cases, one source[69] uses 3009 and 884. Knuth suggested 40902, 24140. Another interesting case is the two relatively prime numbers 14157 and 5950.

But "exceptional cases"[70] must be identified and tested. Will "Inelegant" perform properly when R > S, S > R, R = S? Ditto for "Elegant": B > A, A > B, A = B? (Yes to all). What happens when one number is zero, both numbers are zero? ("Inelegant" computes forever in all cases; "Elegant" computes forever when A = 0.) What happens if negative numbers are entered? Fractional numbers? If the input numbers, i.e. the domain of the function computed by the algorithm/program, is to include only positive integers including zero, then the failures at zero indicate that the algorithm (and the program that instantiates it) is a partial function rather than a total function. A notable failure due to exceptions is the Ariane 5 Flight 501 rocket failure (June 4, 1996).

Proof of program correctness by use of mathematical induction: Knuth demonstrates the application of mathematical induction to an "extended" version of Euclid's algorithm, and he proposes "a general method applicable to proving the validity of any algorithm".[71] Tausworthe proposes that a measure of the complexity of a program be the length of its correctness proof.[72]

Measuring and improving the Euclid algorithms[edit]

Elegance (compactness) versus goodness (speed): With only six core instructions, "Elegant" is the clear winner, compared to "Inelegant" at thirteen instructions. However, "Inelegant" is faster (it arrives at HALT in fewer steps). Algorithm analysis[73] indicates why this is the case: "Elegant" does two conditional tests in every subtraction loop, whereas "Inelegant" only does one. As the algorithm (usually) requires many loop-throughs, on average much time is wasted doing a "B = 0?" test that is needed only after the remainder is computed.

Can the algorithms be improved?: Once the programmer judges a program "fit" and "effective"—that is, it computes the function intended by its author—then the question becomes, can it be improved?

The compactness of "Inelegant" can be improved by the elimination of five steps. But Chaitin proved that compacting an algorithm cannot be automated by a generalized algorithm;[74] rather, it can only be done heuristically; i.e., by exhaustive search (examples to be found at Busy beaver), trial and error, cleverness, insight, application of inductive reasoning, etc. Observe that steps 4, 5 and 6 are repeated in steps 11, 12 and 13. Comparison with "Elegant" provides a hint that these steps, together with steps 2 and 3, can be eliminated. This reduces the number of core instructions from thirteen to eight, which makes it "more elegant" than "Elegant", at nine steps.

The speed of "Elegant" can be improved by moving the "B=0?" test outside of the two subtraction loops. This change calls for the addition of three instructions (B = 0?, A = 0?, GOTO). Now "Elegant" computes the example-numbers faster; whether this is always the case for any given A, B, and R, S would require a detailed analysis.

Algorithmic analysis[edit]

It is frequently important to know how much of a particular resource (such as time or storage) is theoretically required for a given algorithm. Methods have been developed for the analysis of algorithms to obtain such quantitative answers (estimates); for example, an algorithm which adds up the elements of a list of n numbers would have a time requirement of , using big O notation. At all times the algorithm only needs to remember two values: the sum of all the elements so far, and its current position in the input list. Therefore, it is said to have a space requirement of , if the space required to store the input numbers is not counted, or if it is counted.

Different algorithms may complete the same task with a different set of instructions in less or more time, space, or 'effort' than others. For example, a binary search algorithm (with cost ) outperforms a sequential search (cost ) when used for table lookups on sorted lists or arrays.

Formal versus empirical[edit]

The analysis, and study of algorithms is a discipline of computer science, and is often practiced abstractly without the use of a specific programming language or implementation. In this sense, algorithm analysis resembles other mathematical disciplines in that it focuses on the underlying properties of the algorithm and not on the specifics of any particular implementation. Usually pseudocode is used for analysis as it is the simplest and most general representation. However, ultimately, most algorithms are usually implemented on particular hardware/software platforms and their algorithmic efficiency is eventually put to the test using real code. For the solution of a "one off" problem, the efficiency of a particular algorithm may not have significant consequences (unless n is extremely large) but for algorithms designed for fast interactive, commercial or long life scientific usage it may be critical. Scaling from small n to large n frequently exposes inefficient algorithms that are otherwise benign.

Empirical testing is useful because it may uncover unexpected interactions that affect performance. Benchmarks may be used to compare before/after potential improvements to an algorithm after program optimization. Empirical tests cannot replace formal analysis, though, and are not trivial to perform in a fair manner.[75]

Execution efficiency[edit]

To illustrate the potential improvements possible even in well-established algorithms, a recent significant innovation, relating to FFT algorithms (used heavily in the field of image processing), can decrease processing time up to 1,000 times for applications like medical imaging.[76] In general, speed improvements depend on special properties of the problem, which are very common in practical applications.[77] Speedups of this magnitude enable computing devices that make extensive use of image processing (like digital cameras and medical equipment) to consume less power.

Classification[edit]

There are various ways to classify algorithms, each with its own merits.

By implementation[edit]

One way to classify algorithms is by implementation means.

int gcd(int A, int B) {
    if (B == 0)
        return A;
    else if (A > B)
        return gcd(A-B,B);
    else
        return gcd(A,B-A);
}
Recursive C implementation of Euclid's algorithm from the above flowchart
Recursion
A recursive algorithm is one that invokes (makes reference to) itself repeatedly until a certain condition (also known as termination condition) matches, which is a method common to functional programming. Iterative algorithms use repetitive constructs like loops and sometimes additional data structures like stacks to solve the given problems. Some problems are naturally suited for one implementation or the other. For example, towers of Hanoi is well understood using recursive implementation. Every recursive version has an equivalent (but possibly more or less complex) iterative version, and vice versa.
Serial, parallel or distributed
Algorithms are usually discussed with the assumption that computers execute one instruction of an algorithm at a time. Those computers are sometimes called serial computers. An algorithm designed for such an environment is called a serial algorithm, as opposed to parallel algorithms or distributed algorithms. Parallel algorithms are algorithms that take advantage of computer architectures where multiple processors can work on a problem at the same time. Distributed algorithms are algorithms that use multiple machines connected with a computer network. Parallel and distributed algorithms divide the problem into more symmetrical or asymmetrical subproblems and collect the results back together. For example, a CPU would be an example of a parallel algorithm. The resource consumption in such algorithms is not only processor cycles on each processor but also the communication overhead between the processors. Some sorting algorithms can be parallelized efficiently, but their communication overhead is expensive. Iterative algorithms are generally parallelizable, but some problems have no parallel algorithms and are called inherently serial problems.
Deterministic or non-deterministic
Deterministic algorithms solve the problem with exact decision at every step of the algorithm whereas non-deterministic algorithms solve problems via guessing although typical guesses are made more accurate through the use of heuristics.
Exact or approximate
While many algorithms reach an exact solution, approximation algorithms seek an approximation that is closer to the true solution. The approximation can be reached by either using a deterministic or a random strategy. Such algorithms have practical value for many hard problems. One of the examples of an approximate algorithm is the Knapsack problem, where there is a set of given items. Its goal is to pack the knapsack to get the maximum total value. Each item has some weight and some value. Total weight that can be carried is no more than some fixed number X. So, the solution must consider weights of items as well as their value.[78]
Quantum algorithm
Quantum algorithms run on a realistic model of quantum computation. The term is usually used for those algorithms which seem inherently quantum, or use some essential feature of Quantum computing such as quantum superposition or quantum entanglement.

By design paradigm[edit]

Another way of classifying algorithms is by their design methodology or paradigm. There is a certain number of paradigms, each different from the other. Furthermore, each of these categories includes many different types of algorithms. Some common paradigms are:

Brute-force or exhaustive search
Brute force is a method of problem-solving that involves systematically trying every possible option until the optimal solution is found. This approach can be very time consuming, as it requires going through every possible combination of variables. However, it is often used when other methods are not available or too complex. Brute force can be used to solve a variety of problems, including finding the shortest path between two points and cracking passwords.
Divide and conquer
A divide-and-conquer algorithm repeatedly reduces an instance of a problem to one or more smaller instances of the same problem (usually recursively) until the instances are small enough to solve easily. One such example of divide and conquer is merge sorting. Sorting can be done on each segment of data after dividing data into segments and sorting of entire data can be obtained in the conquer phase by merging the segments. A simpler variant of divide and conquer is called a decrease-and-conquer algorithm, which solves an identical subproblem and uses the solution of this subproblem to solve the bigger problem. Divide and conquer divides the problem into multiple subproblems and so the conquer stage is more complex than decrease and conquer algorithms. An example of a decrease and conquer algorithm is the binary search algorithm.
Search and enumeration
Many problems (such as playing chess) can be modeled as problems on graphs. A graph exploration algorithm specifies rules for moving around a graph and is useful for such problems. This category also includes search algorithms, branch and bound enumeration and backtracking.
Randomized algorithm
Such algorithms make some choices randomly (or pseudo-randomly). They can be very useful in finding approximate solutions for problems where finding exact solutions can be impractical (see heuristic method below). For some of these problems, it is known that the fastest approximations must involve some randomness.[79] Whether randomized algorithms with polynomial time complexity can be the fastest algorithms for some problems is an open question known as the P versus NP problem. There are two large classes of such algorithms:
  1. Monte Carlo algorithms return a correct answer with high-probability. E.g. RP is the subclass of these that run in polynomial time.
  2. Las Vegas algorithms always return the correct answer, but their running time is only probabilistically bound, e.g. ZPP.
Reduction of complexity
This technique involves solving a difficult problem by transforming it into a better-known problem for which we have (hopefully) asymptotically optimal algorithms. The goal is to find a reducing algorithm whose complexity is not dominated by the resulting reduced algorithm's. For example, one selection algorithm for finding the median in an unsorted list involves first sorting the list (the expensive portion) and then pulling out the middle element in the sorted list (the cheap portion). This technique is also known as transform and conquer.
Back tracking
In this approach, multiple solutions are built incrementally and abandoned when it is determined that they cannot lead to a valid full solution.

Optimization problems[edit]

For optimization problems there is a more specific classification of algorithms; an algorithm for such problems may fall into one or more of the general categories described above as well as into one of the following:

Linear programming
When searching for optimal solutions to a linear function bound to linear equality and inequality constraints, the constraints of the problem can be used directly in producing the optimal solutions. There are algorithms that can solve any problem in this category, such as the popular simplex algorithm.[80] Problems that can be solved with linear programming include the maximum flow problem for directed graphs. If a problem additionally requires that one or more of the unknowns must be an integer then it is classified in integer programming. A linear programming algorithm can solve such a problem if it can be proved that all restrictions for integer values are superficial, i.e., the solutions satisfy these restrictions anyway. In the general case, a specialized algorithm or an algorithm that finds approximate solutions is used, depending on the difficulty of the problem.
Dynamic programming
When a problem shows optimal substructures—meaning the optimal solution to a problem can be constructed from optimal solutions to subproblems—and overlapping subproblems, meaning the same subproblems are used to solve many different problem instances, a quicker approach called dynamic programming avoids recomputing solutions that have already been computed. For example, Floyd–Warshall algorithm, the shortest path to a goal from a vertex in a weighted graph can be found by using the shortest path to the goal from all adjacent vertices. Dynamic programming and memoization go together. The main difference between dynamic programming and divide and conquer is that subproblems are more or less independent in divide and conquer, whereas subproblems overlap in dynamic programming. The difference between dynamic programming and straightforward recursion is in caching or memoization of recursive calls. When subproblems are independent and there is no repetition, memoization does not help; hence dynamic programming is not a solution for all complex problems. By using memoization or maintaining a table of subproblems already solved, dynamic programming reduces the exponential nature of many problems to polynomial complexity.
The greedy method
A greedy algorithm is similar to a dynamic programming algorithm in that it works by examining substructures, in this case not of the problem but of a given solution. Such algorithms start with some solution, which may be given or have been constructed in some way, and improve it by making small modifications. For some problems they can find the optimal solution while for others they stop at local optima, that is, at solutions that cannot be improved by the algorithm but are not optimum. The most popular use of greedy algorithms is for finding the minimal spanning tree where finding the optimal solution is possible with this method. Huffman Tree, Kruskal, Prim, Sollin are greedy algorithms that can solve this optimization problem.
The heuristic method
In optimization problems, heuristic algorithms can be used to find a solution close to the optimal solution in cases where finding the optimal solution is impractical. These algorithms work by getting closer and closer to the optimal solution as they progress. In principle, if run for an infinite amount of time, they will find the optimal solution. Their merit is that they can find a solution very close to the optimal solution in a relatively short time. Such algorithms include local search, tabu search, simulated annealing, and genetic algorithms. Some of them, like simulated annealing, are non-deterministic algorithms while others, like tabu search, are deterministic. When a bound on the error of the non-optimal solution is known, the algorithm is further categorized as an approximation algorithm.

By field of study[edit]

Every field of science has its own problems and needs efficient algorithms. Related problems in one field are often studied together. Some example classes are search algorithms, sorting algorithms, merge algorithms, numerical algorithms, graph algorithms, string algorithms, computational geometric algorithms, combinatorial algorithms, medical algorithms, machine learning, cryptography, data compression algorithms and parsing techniques.

Fields tend to overlap with each other, and algorithm advances in one field may improve those of other, sometimes completely unrelated, fields. For example, dynamic programming was invented for optimization of resource consumption in industry but is now used in solving a broad range of problems in many fields.

By complexity[edit]

Algorithms can be classified by the amount of time they need to complete compared to their input size:

  • Constant time: if the time needed by the algorithm is the same, regardless of the input size. E.g. an access to an array element.
  • Logarithmic time: if the time is a logarithmic function of the input size. E.g. binary search algorithm.
  • Linear time: if the time is proportional to the input size. E.g. the traverse of a list.
  • Polynomial time: if the time is a power of the input size. E.g. the bubble sort algorithm has quadratic time complexity.
  • Exponential time: if the time is an exponential function of the input size. E.g. Brute-force search.

Some problems may have multiple algorithms of differing complexity, while other problems might have no algorithms or no known efficient algorithms. There are also mappings from some problems to other problems. Owing to this, it was found to be more suitable to classify the problems themselves instead of the algorithms into equivalence classes based on the complexity of the best possible algorithms for them.

Continuous algorithms[edit]

The adjective "continuous" when applied to the word "algorithm" can mean:

  • An algorithm operating on data that represents continuous quantities, even though this data is represented by discrete approximations—such algorithms are studied in numerical analysis; or
  • An algorithm in the form of a differential equation that operates continuously on the data, running on an analog computer.[81]

Algorithm = Logic + Control[edit]

In logic programming, algorithms are viewed as having both "a logic component, which specifies the knowledge to be used in solving problems, and a control component, which determines the problem-solving strategies by means of which that knowledge is used."[82]

The Euclidean algorithm illustrates this view of an algorithm.[83][84] Here is a logic programming representation, using :- to represent "if", and the relation gcd(A, B, C) to represent the function gcd(A, B) = C:

gcd(A, A, A).
gcd(A, B, C) :- A > B, gcd(A-B, B, C).
gcd(A, B, C) :- B > A, gcd(A, B-A, C).

In the logic programming language Ciao the gcd relation can be represented directly in functional notation:

gcd(A, A) := A.
gcd(A, B) := gcd(A-B, B) :- A > B.
gcd(A, B) := gcd(A, B-A) :- B > A.

The Ciao implementation translates the functional notation into a relational representation in Prolog, extracting the embedded subtractions, A-B and B-A, as separate conditions:

gcd(A, A, A).
gcd(A, B, C) :- A > B, A' is A-B, gcd(A', B, C).
gcd(A, B, C) :- B > A, B' is B-A, gcd(A, B, C).

The resulting program has a purely logical (and "declarative") reading, as a recursive (or inductive) definition, which is independent of how the logic is used to solve problems:

The gcd of A and A is A.
The gcd of A and B is C, if A > B and A' is A-B and the gcd of A' and B is C.
The gcd of A and B is C, if B > A and B' is B-A and the gcd of A and B' is C.

Different problem-solving strategies turn the logic into different algorithms. In theory, given a pair of integers A and B, forward (or "bottom-up") reasoning could be used to generate all instances of the gcd relation, terminating when the desired gcd of A and B is generated. Of course, forward reasoning is entirely useless in this case. But in other cases, such as the definition of the Fibonacci sequence[82] and Datalog, forward reasoning can be an efficient problem solving strategy. (See for example the logic program for computing fibonacci numbers in Algorithm = Logic + Control).

In contrast with the inefficiency of forward reasoning in this example, backward (or "top-down") reasoning using SLD resolution turns the logic into the Euclidean algorithm:

To find the gcd C of two given numbers A and B:
If A = B, then C = A.
If A > B, then find the gcd of  A-B and B, which is C.
If B > A, then find the gcd of A and  B-A, which is C.

One of the advantages of the logic programming representation of the algorithm is that its purely logical reading makes it easier to verify that the algorithm is correct relative to the standard non-recursive definition of gcd.[83] Here is the standard definition written in Prolog:

gcd(A, B, C) :- divides(C, A), divides(C, B),
    forall((divides(D, A), divides(D, B)), D =< C).
           
 divides(C, Number) :- 
   between(1, Number, C), 0 is Number mod C.

This definition, which is the specification of the Euclidean algorithm, is also executable in Prolog: Backward reasoning treats the specification as the brute-force algorithm that iterates through all of the integers C between 1 and A, checking whether C divides both A and B, and then for each such C iterates again through all of the integers D between 1 and A, until it finds a C such that C is greater than or equal to all of the D that also divide both A and B. Although this algorithm is hopelessly inefficient, it shows that formal specifications can often be written in logic programming form, and they can be executed by Prolog, to check that they correctly represent informal requirements.

Legal issues[edit]

Algorithms, by themselves, are not usually patentable. In the United States, a claim consisting solely of simple manipulations of abstract concepts, numbers, or signals does not constitute "processes" (USPTO 2006), so algorithms are not patentable (as in Gottschalk v. Benson). However practical applications of algorithms are sometimes patentable. For example, in Diamond v. Diehr, the application of a simple feedback algorithm to aid in the curing of synthetic rubber was deemed patentable. The patenting of software is controversial,[85] and there are criticized patents involving algorithms, especially data compression algorithms, such as Unisys's LZW patent.

Additionally, some cryptographic algorithms have export restrictions (see export of cryptography).

History: Development of the notion of "algorithm"[edit]

Ancient Near East[edit]

The earliest evidence of algorithms is found in the Babylonian mathematics of ancient Mesopotamia (modern Iraq). A Sumerian clay tablet found in Shuruppak near Baghdad and dated to c. 2500 BC described the earliest division algorithm.[11] During the Hammurabi dynasty c. 1800 – c. 1600 BC, Babylonian clay tablets described algorithms for computing formulas.[86] Algorithms were also used in Babylonian astronomy. Babylonian clay tablets describe and employ algorithmic procedures to compute the time and place of significant astronomical events.[87]

Algorithms for arithmetic are also found in ancient Egyptian mathematics, dating back to the Rhind Mathematical Papyrus c. 1550 BC.[11] Algorithms were later used in ancient Hellenistic mathematics. Two examples are the Sieve of Eratosthenes, which was described in the Introduction to Arithmetic by Nicomachus,[88][14]: Ch 9.2  and the Euclidean algorithm, which was first described in Euclid's Elements (c. 300 BC).[14]: Ch 9.1 

Discrete and distinguishable symbols[edit]

Tally-marks: To keep track of their flocks, their sacks of grain and their money the ancients used tallying: accumulating stones or marks scratched on sticks or making discrete symbols in clay. Through the Babylonian and Egyptian use of marks and symbols, eventually Roman numerals and the abacus evolved (Dilson, p. 16–41). Tally marks appear prominently in unary numeral system arithmetic used in Turing machine and Post–Turing machine computations.

Manipulation of symbols as "place holders" for numbers: algebra[edit]

Muhammad ibn Mūsā al-Khwārizmī, a Persian mathematician, wrote the Al-jabr in the 9th century. The terms "algorism" and "algorithm" are derived from the name al-Khwārizmī, while the term "algebra" is derived from the book Al-jabr. In Europe, the word "algorithm" was originally used to refer to the sets of rules and techniques used by Al-Khwarizmi to solve algebraic equations, before later being generalized to refer to any set of rules or techniques.[89] This eventually culminated in Leibniz's notion of the calculus ratiocinator (c. 1680):

A good century and a half ahead of his time, Leibniz proposed an algebra of logic, an algebra that would specify the rules for manipulating logical concepts in the manner that ordinary algebra specifies the rules for manipulating numbers.[90]

Cryptographic algorithms[edit]

The first cryptographic algorithm for deciphering encrypted code was developed by Al-Kindi, a 9th-century Arab mathematician, in A Manuscript On Deciphering Cryptographic Messages. He gave the first description of cryptanalysis by frequency analysis, the earliest codebreaking algorithm.[15]

Mechanical contrivances with discrete states[edit]

The clock: Bolter credits the invention of the weight-driven clock as "The key invention [of Europe in the Middle Ages]", in particular, the verge escapement[91] that provides us with the tick and tock of a mechanical clock. "The accurate automatic machine"[92] led immediately to "mechanical automata" beginning in the 13th century and finally to "computational machines"—the difference engine and analytical engines of Charles Babbage and Countess Ada Lovelace, mid-19th century.[93] Lovelace is credited with the first creation of an algorithm intended for processing on a computer—Babbage's analytical engine, the first device considered a real Turing-complete computer instead of just a calculator—and is sometimes called "history's first programmer" as a result, though a full implementation of Babbage's second device would not be realized until decades after her lifetime.

Logical machines 1870 – Stanley Jevons' "logical abacus" and "logical machine": The technical problem was to reduce Boolean equations when presented in a form similar to what is now known as Karnaugh maps. Jevons (1880) describes first a simple "abacus" of "slips of wood furnished with pins, contrived so that any part or class of the [logical] combinations can be picked out mechanically ... More recently, however, I have reduced the system to a completely mechanical form, and have thus embodied the whole of the indirect process of inference in what may be called a Logical Machine" His machine came equipped with "certain moveable wooden rods" and "at the foot are 21 keys like those of a piano [etc.] ...". With this machine he could analyze a "syllogism or any other simple logical argument".[94]

This machine he displayed in 1870 before the Fellows of the Royal Society.[95] Another logician John Venn, however, in his 1881 Symbolic Logic, turned a jaundiced eye to this effort: "I have no high estimate myself of the interest or importance of what are sometimes called logical machines ... it does not seem to me that any contrivances at present known or likely to be discovered really deserve the name of logical machines"; see more at Algorithm characterizations. But not to be outdone he too presented "a plan somewhat analogous, I apprehend, to Prof. Jevon's abacus ... [And] [a]gain, corresponding to Prof. Jevons's logical machine, the following contrivance may be described. I prefer to call it merely a logical-diagram machine ... but I suppose that it could do very completely all that can be rationally expected of any logical machine".[96]

Jacquard loom, Hollerith punch cards, telegraphy and telephony – the electromechanical relay: Bell and Newell (1971) indicate that the Jacquard loom (1801), precursor to Hollerith cards (punch cards, 1887), and "telephone switching technologies" were the roots of a tree leading to the development of the first computers.[97] By the mid-19th century the telegraph, the precursor of the telephone, was in use throughout the world, its discrete and distinguishable encoding of letters as "dots and dashes" a common sound. By the late 19th century the ticker tape (c. 1870s) was in use, as was the use of Hollerith cards in the 1890 U.S. census. Then came the teleprinter (c. 1910) with its punched-paper use of Baudot code on tape.

Telephone-switching networks of electromechanical relays (invented 1835) was behind the work of George Stibitz (1937), the inventor of the digital adding device. As he worked in Bell Laboratories, he observed the "burdensome' use of mechanical calculators with gears. "He went home one evening in 1937 intending to test his idea... When the tinkering was over, Stibitz had constructed a binary adding device".[98]

The mathematician Martin Davis observes the particular importance of the electromechanical relay (with its two "binary states" open and closed):

It was only with the development, beginning in the 1930s, of electromechanical calculators using electrical relays, that machines were built having the scope Babbage had envisioned."[99]

Mathematics during the 19th century up to the mid-20th century[edit]

Symbols and rules: In rapid succession, the mathematics of George Boole (1847, 1854), Gottlob Frege (1879), and Giuseppe Peano (1888–1889) reduced arithmetic to a sequence of symbols manipulated by rules. Peano's The principles of arithmetic, presented by a new method (1888) was "the first attempt at an axiomatization of mathematics in a symbolic language".[100]

But Heijenoort gives Frege (1879) this kudos: Frege's is "perhaps the most important single work ever written in logic. ... in which we see a "'formula language', that is a lingua characterica, a language written with special symbols, "for pure thought", that is, free from rhetorical embellishments ... constructed from specific symbols that are manipulated according to definite rules".[101] The work of Frege was further simplified and amplified by Alfred North Whitehead and Bertrand Russell in their Principia Mathematica (1910–1913).

The paradoxes: At the same time a number of disturbing paradoxes appeared in the literature, in particular, the Burali-Forti paradox (1897), the Russell paradox (1902–03), and the Richard Paradox.[102] The resultant considerations led to Kurt Gödel's paper (1931)—he specifically cites the paradox of the liar—that completely reduces rules of recursion to numbers.

Effective calculability: In an effort to solve the Entscheidungsproblem defined precisely by Hilbert in 1928, mathematicians first set about to define what was meant by an "effective method" or "effective calculation" or "effective calculability" (i.e., a calculation that would succeed). In rapid succession the following appeared: Alonzo Church, Stephen Kleene and J.B. Rosser's λ-calculus[103] a finely honed definition of "general recursion" from the work of Gödel acting on suggestions of Jacques Herbrand (cf. Gödel's Princeton lectures of 1934) and subsequent simplifications by Kleene.[104] Church's proof[105] that the Entscheidungsproblem was unsolvable, Emil Post's definition of effective calculability as a worker mindlessly following a list of instructions to move left or right through a sequence of rooms and while there either mark or erase a paper or observe the paper and make a yes-no decision about the next instruction.[106] Alan Turing's proof of that the Entscheidungsproblem was unsolvable by use of his "a- [automatic-] machine"[107]—in effect almost identical to Post's "formulation", J. Barkley Rosser's definition of "effective method" in terms of "a machine".[108] Kleene's proposal of a precursor to "Church thesis" that he called "Thesis I",[109] and a few years later Kleene's renaming his Thesis "Church's Thesis"[110] and proposing "Turing's Thesis".[111]

Emil Post (1936) and Alan Turing (1936–37, 1939)[edit]

Emil Post (1936) described the actions of a "computer" (human being) as follows:

"...two concepts are involved: that of a symbol space in which the work leading from problem to answer is to be carried out, and a fixed unalterable set of directions.

His symbol space would be

"a two-way infinite sequence of spaces or boxes ... The problem solver or worker is to move and work in this symbol space, being capable of being in, and operating in but one box at a time. ... a box is to admit of but two possible conditions, i.e., being empty or unmarked, and having a single mark in it, say a vertical stroke.
"One box is to be singled out and called the starting point. ... a specific problem is to be given in symbolic form by a finite number of boxes [i.e., INPUT] being marked with a stroke. Likewise, the answer [i.e., OUTPUT] is to be given in symbolic form by such a configuration of marked boxes...
"A set of directions applicable to a general problem sets up a deterministic process when applied to each specific problem. This process terminates only when it comes to the direction of type (C ) [i.e., STOP]".[112] See more at Post–Turing machine
Alan Turing's statue at Bletchley Park

Alan Turing's work[113] preceded that of Stibitz (1937); it is unknown whether Stibitz knew of the work of Turing. Turing's biographer believed that Turing's use of a typewriter-like model derived from a youthful interest: "Alan had dreamt of inventing typewriters as a boy; Mrs. Turing had a typewriter, and he could well have begun by asking himself what was meant by calling a typewriter 'mechanical'".[114] Given the prevalence at the time of Morse code, telegraphy, ticker tape machines, and teletypewriters, it is quite possible that all were influences on Turing during his youth.

Turing—his model of computation is now called a Turing machine—begins, as did Post, with an analysis of a human computer that he whittles down to a simple set of basic motions and "states of mind". But he continues a step further and creates a machine as a model of computation of numbers.[115]

"Computing is normally done by writing certain symbols on paper. We may suppose this paper is divided into squares like a child's arithmetic book...I assume then that the computation is carried out on one-dimensional paper, i.e., on a tape divided into squares. I shall also suppose that the number of symbols which may be printed is finite...
"The behavior of the computer at any moment is determined by the symbols which he is observing, and his "state of mind" at that moment. We may suppose that there is a bound B to the number of symbols or squares that the computer can observe at one moment. If he wishes to observe more, he must use successive observations. We will also suppose that the number of states of mind which need be taken into account is finite...
"Let us imagine that the operations performed by the computer to be split up into 'simple operations' which are so elementary that it is not easy to imagine them further divided."[116]

Turing's reduction yields the following:

"The simple operations must therefore include:
"(a) Changes of the symbol on one of the observed squares
"(b) Changes of one of the squares observed to another square within L squares of one of the previously observed squares.

"It may be that some of these change necessarily invoke a change of state of mind. The most general single operation must, therefore, be taken to be one of the following:

"(A) A possible change (a) of symbol together with a possible change of state of mind.
"(B) A possible change (b) of observed squares, together with a possible change of state of mind"
"We may now construct a machine to do the work of this computer."[116]

A few years later, Turing expanded his analysis (thesis, definition) with this forceful expression of it:

"A function is said to be "effectively calculable" if its values can be found by some purely mechanical process. Though it is fairly easy to get an intuitive grasp of this idea, it is nevertheless desirable to have some more definite, mathematical expressible definition ... [he discusses the history of the definition pretty much as presented above with respect to Gödel, Herbrand, Kleene, Church, Turing, and Post] ... We may take this statement literally, understanding by a purely mechanical process one which could be carried out by a machine. It is possible to give a mathematical description, in a certain normal form, of the structures of these machines. The development of these ideas leads to the author's definition of a computable function, and to an identification of computability † with effective calculability...
"† We shall use the expression "computable function" to mean a function calculable by a machine, and we let "effectively calculable" refer to the intuitive idea without particular identification with any one of these definitions".[117]

J. B. Rosser (1939) and S. C. Kleene (1943)[edit]

J. Barkley Rosser defined an "effective [mathematical] method" in the following manner (italicization added):

"'Effective method' is used here in the rather special sense of a method each step of which is precisely determined and which is certain to produce the answer in a finite number of steps. With this special meaning, three different precise definitions have been given to date. [his footnote #5; see discussion immediately below]. The simplest of these to state (due to Post and Turing) says essentially that an effective method of solving certain sets of problems exists if one can build a machine which will then solve any problem of the set with no human intervention beyond inserting the question and (later) reading the answer. All three definitions are equivalent, so it doesn't matter which one is used. Moreover, the fact that all three are equivalent is a very strong argument for the correctness of any one." (Rosser 1939:225–226)

Rosser's footnote No. 5 references the work of (1) Church and Kleene and their definition of λ-definability, in particular, Church's use of it in his An Unsolvable Problem of Elementary Number Theory (1936); (2) Herbrand and Gödel and their use of recursion, in particular, Gödel's use in his famous paper On Formally Undecidable Propositions of Principia Mathematica and Related Systems I (1931); and (3) Post (1936) and Turing (1936–37) in their mechanism-models of computation.

Stephen C. Kleene defined as his now-famous "Thesis I" known as the Church–Turing thesis. But he did this in the following context (boldface in original):

"12. Algorithmic theories... In setting up a complete algorithmic theory, what we do is to describe a procedure, performable for each set of values of the independent variables, which procedure necessarily terminates and in such manner that from the outcome we can read a definite answer, "yes" or "no," to the question, "is the predicate value true?"" (Kleene 1943:273)

History after 1950[edit]

A number of efforts have been directed toward further refinement of the definition of "algorithm", and activity is on-going because of issues surrounding, in particular, foundations of mathematics (especially the Church–Turing thesis) and philosophy of mind (especially arguments about artificial intelligence). For more, see Algorithm characterizations.

See also[edit]

Notes[edit]

  1. ^ "Definition of ALGORITHM". Merriam-Webster Online Dictionary. Archived from the original on February 14, 2020. Retrieved November 14, 2019.
  2. ^ Blair, Ann, Duguid, Paul, Goeing, Anja-Silvia and Grafton, Anthony. Information: A Historical Companion, Princeton: Princeton University Press, 2021. p. 247
  3. ^ David A. Grossman, Ophir Frieder, Information Retrieval: Algorithms and Heuristics, 2nd edition, 2004, ISBN 1402030045
  4. ^ "Any classical mathematical algorithm, for example, can be described in a finite number of English words" (Rogers 1987:2).
  5. ^ Well defined with respect to the agent that executes the algorithm: "There is a computing agent, usually human, which can react to the instructions and carry out the computations" (Rogers 1987:2).
  6. ^ "an algorithm is a procedure for computing a function (with respect to some chosen notation for integers) ... this limitation (to numerical functions) results in no loss of generality", (Rogers 1987:1).
  7. ^ "An algorithm has zero or more inputs, i.e., quantities which are given to it initially before the algorithm begins" (Knuth 1973:5).
  8. ^ "A procedure which has all the characteristics of an algorithm except that it possibly lacks finiteness may be called a 'computational method'" (Knuth 1973:5).
  9. ^ "An algorithm has one or more outputs, i.e. quantities which have a specified relation to the inputs" (Knuth 1973:5).
  10. ^ Whether or not a process with random interior processes (not including the input) is an algorithm is debatable. Rogers opines that: "a computation is carried out in a discrete stepwise fashion, without the use of continuous methods or analogue devices ... carried forward deterministically, without resort to random methods or devices, e.g., dice" (Rogers 1987:2).
  11. ^ a b c d Chabert, Jean-Luc (2012). A History of Algorithms: From the Pebble to the Microchip. Springer Science & Business Media. pp. 7–8. ISBN 9783642181924.
  12. ^ Sriram, M. S. (2005). "Algorithms in Indian Mathematics". In Emch, Gerard G.; Sridharan, R.; Srinivas, M. D. (eds.). Contributions to the History of Indian Mathematics. Springer. p. 153. ISBN 978-93-86279-25-5.
  13. ^ Hayashi, T. (2023, January 1). Brahmagupta. Encyclopedia Britannica.
  14. ^ a b c Cooke, Roger L. (2005). The History of Mathematics: A Brief Course. John Wiley & Sons. ISBN 978-1-118-46029-0.
  15. ^ a b Dooley, John F. (2013). A Brief History of Cryptology and Cryptographic Algorithms. Springer Science & Business Media. pp. 12–3. ISBN 9783319016283.
  16. ^ Burnett, Charles (2017). "Arabic Numerals". In Thomas F. Glick (ed.). Routledge Revivals: Medieval Science, Technology and Medicine (2006): An Encyclopedia. Taylor & Francis. p. 39. ISBN 978-1-351-67617-5. Archived from the original on March 28, 2023. Retrieved May 5, 2019.
  17. ^ "algorism". Oxford English Dictionary (Online ed.). Oxford University Press. (Subscription or participating institution membership required.)
  18. ^ Brezina, Corona (2006). Al-Khwarizmi: The Inventor Of Algebra. The Rosen Publishing Group. ISBN 978-1-4042-0513-0.
  19. ^ "Abu Jafar Muhammad ibn Musa al-Khwarizmi". members.peak.org. Archived from the original on August 21, 2019. Retrieved November 14, 2019.
  20. ^ Mehri, Bahman (2017). "From Al-Khwarizmi to Algorithm". Olympiads in Informatics. 11 (2): 71–74. doi:10.15388/ioi.2017.special.11.
  21. ^ "algorismic". The Free Dictionary. Archived from the original on December 21, 2019. Retrieved November 14, 2019.
  22. ^ Blount, Thomas (1656). Glossographia or a Dictionary... London: Humphrey Moseley and George Sawbridge.
  23. ^ Phillips, Edward (1658). The new world of English words, or, A general dictionary containing the interpretations of such hard words as are derived from other languages...
  24. ^ Phillips, Edward; Kersey, John (1706). The new world of words: or, Universal English dictionary. Containing an account of the original or proper sense, and various significations of all hard words derived from other languages ... Together with a brief and plain explication of all terms relating to any of the arts and sciences ... to which is added, the interpretation of proper names. Printed for J. Phillips etc.
  25. ^ Fenning, Daniel (1751). The young algebraist's companion, or, A new & easy guide to algebra; introduced by the doctrine of vulgar fractions: designed for the use of schools ... illustrated with variety of numerical & literal examples ... Printed for G. Keith & J. Robinson. p. xi.
  26. ^ The Electric Review 1811-07: Vol 7. Open Court Publishing Co. July 1811. p. [1]. Yet it wants a new algorithm, a compendious method by which the theorems may be established without ambiguity and circumlocution, [...]
  27. ^ "algorithm". Oxford English Dictionary (Online ed.). Oxford University Press. (Subscription or participating institution membership required.)
  28. ^ Already 1684, in Nova Methodus pro Maximis et Minimis, Leibnitz used the Latin term "algorithmo".
  29. ^ Kleene 1943 in Davis 1965:274
  30. ^ Rosser 1939 in Davis 1965:225
  31. ^ Stone 1973:4
  32. ^ Simanowski, Roberto (2018). The Death Algorithm and Other Digital Dilemmas. Untimely Meditations. Vol. 14. Translated by Chase, Jefferson. Cambridge, Massachusetts: MIT Press. p. 147. ISBN 9780262536370. Archived from the original on December 22, 2019. Retrieved May 27, 2019. [...] the next level of abstraction of central bureaucracy: globally operating algorithms.
  33. ^ Dietrich, Eric (1999). "Algorithm". In Wilson, Robert Andrew; Keil, Frank C. (eds.). The MIT Encyclopedia of the Cognitive Sciences. MIT Cognet library. Cambridge, Massachusetts: MIT Press (published 2001). p. 11. ISBN 9780262731447. Retrieved July 22, 2020. An algorithm is a recipe, method, or technique for doing something.
  34. ^ Stone requires that "it must terminate in a finite number of steps" (Stone 1973:7–8).
  35. ^ Boolos and Jeffrey 1974,1999:19
  36. ^ cf Stone 1972:5
  37. ^ Knuth 1973:7 states: "In practice, we not only want algorithms, but we also want good algorithms ... one criterion of goodness is the length of time taken to perform the algorithm ... other criteria are the adaptability of the algorithm to computers, its simplicity, and elegance, etc."
  38. ^ cf Stone 1973:6
  39. ^ Stone 1973:7–8 states that there must be, "...a procedure that a robot [i.e., computer] can follow in order to determine precisely how to obey the instruction". Stone adds finiteness of the process, and definiteness (having no ambiguity in the instructions) to this definition.
  40. ^ Knuth, loc. cit
  41. ^ Minsky 1967, p. 105
  42. ^ Gurevich 2000:1, 3
  43. ^ Sipser 2006:157
  44. ^ Goodrich, Michael T.; Tamassia, Roberto (2002). Algorithm Design: Foundations, Analysis, and Internet Examples. John Wiley & Sons, Inc. ISBN 978-0-471-38365-9. Archived from the original on April 28, 2015. Retrieved June 14, 2018.
  45. ^ Knuth 1973:7
  46. ^ Chaitin 2005:32
  47. ^ Rogers 1987:1–2
  48. ^ In his essay "Calculations by Man and Machine: Conceptual Analysis" Seig 2002:390 credits this distinction to Robin Gandy, cf Wilfred Seig, et al., 2002 Reflections on the foundations of mathematics: Essays in honor of Solomon Feferman, Association for Symbolic Logic, A.K. Peters Ltd, Natick, MA.
  49. ^ cf Gandy 1980:126, Robin Gandy Church's Thesis and Principles for Mechanisms appearing on pp. 123–148 in J. Barwise et al. 1980 The Kleene Symposium, North-Holland Publishing Company.
  50. ^ A "robot": "A computer is a robot that performs any task that can be described as a sequence of instructions." cf Stone 1972:3
  51. ^ Lambek's "abacus" is a "countably infinite number of locations (holes, wires, etc.) together with an unlimited supply of counters (pebbles, beads, etc.). The locations are distinguishable, the counters are not". The holes have unlimited capacity, and standing by is an agent who understands and is able to carry out the list of instructions" (Lambek 1961:295). Lambek references Melzak who defines his Q-machine as "an indefinitely large number of locations ... an indefinitely large supply of counters distributed among these locations, a program, and an operator whose sole purpose is to carry out the program" (Melzak 1961:283). B-B-J (loc. cit.) add the stipulation that the holes are "capable of holding any number of stones" (p. 46). Both Melzak and Lambek appear in The Canadian Mathematical Bulletin, vol. 4, no. 3, September 1961.
  52. ^ If no confusion results, the word "counters" can be dropped, and a location can be said to contain a single "number".
  53. ^ "We say that an instruction is effective if there is a procedure that the robot can follow in order to determine precisely how to obey the instruction." (Stone 1972:6)
  54. ^ cf Minsky 1967: Chapter 11 "Computer models" and Chapter 14 "Very Simple Bases for Computability" pp. 255–281, in particular,
  55. ^ cf Knuth 1973:3.
  56. ^ But always preceded by IF-THEN to avoid improper subtraction.
  57. ^ Knuth 1973:4
  58. ^ Stone 1972:5. Methods for extracting roots are not trivial: see Methods of computing square roots.
  59. ^ Leeuwen, Jan (1990). Handbook of Theoretical Computer Science: Algorithms and complexity. Volume A. Elsevier. p. 85. ISBN 978-0-444-88071-0.
  60. ^ John G. Kemeny and Thomas E. Kurtz 1985 Back to Basic: The History, Corruption, and Future of the Language, Addison-Wesley Publishing Company, Inc. Reading, MA, ISBN 0-201-13433-0.
  61. ^ Tausworthe 1977:101
  62. ^ Tausworthe 1977:142
  63. ^ Knuth 1973 section 1.2.1, expanded by Tausworthe 1977 at pages 100ff and Chapter 9.1
  64. ^ cf Tausworthe 1977
  65. ^ Heath 1908:300; Hawking's Dover 2005 edition derives from Heath.
  66. ^ " 'Let CD, measuring BF, leave FA less than itself.' This is a neat abbreviation for saying, measure along BA successive lengths equal to CD until a point F is reached such that the length FA remaining is less than CD; in other words, let BF be the largest exact multiple of CD contained in BA" (Heath 1908:297)
  67. ^ For modern treatments using division in the algorithm, see Hardy and Wright 1979:180, Knuth 1973:2 (Volume 1), plus more discussion of Euclid's algorithm in Knuth 1969:293–297 (Volume 2).
  68. ^ Euclid covers this question in his Proposition 1.
  69. ^ "Euclid's Elements, Book VII, Proposition 2". Aleph0.clarku.edu. Archived from the original on May 24, 2012. Retrieved May 20, 2012.
  70. ^ While this notion is in widespread use, it cannot be defined precisely.
  71. ^ Knuth 1973:13–18. He credits "the formulation of algorithm-proving in terms of assertions and induction" to R W. Floyd, Peter Naur, C.A.R. Hoare, H.H. Goldstine and J. von Neumann. Tausworth 1977 borrows Knuth's Euclid example and extends Knuth's method in section 9.1 Formal Proofs (pp. 288–298).
  72. ^ Tausworthe 1997:294
  73. ^ cf Knuth 1973:7 (Vol. I), and his more-detailed analyses on pp. 1969:294–313 (Vol II).
  74. ^ Breakdown occurs when an algorithm tries to compact itself. Success would solve the Halting problem.
  75. ^ Kriegel, Hans-Peter; Schubert, Erich; Zimek, Arthur (2016). "The (black) art of run-time evaluation: Are we comparing algorithms or implementations?". Knowledge and Information Systems. 52 (2): 341–378. doi:10.1007/s10115-016-1004-2. ISSN 0219-1377. S2CID 40772241.
  76. ^ Gillian Conahan (January 2013). "Better Math Makes Faster Data Networks". discovermagazine.com. Archived from the original on May 13, 2014. Retrieved May 13, 2014.
  77. ^ Haitham Hassanieh, Piotr Indyk, Dina Katabi, and Eric Price, "ACM-SIAM Symposium On Discrete Algorithms (SODA) Archived July 4, 2013, at the Wayback Machine, Kyoto, January 2012. See also the sFFT Web Page Archived February 21, 2012, at the Wayback Machine.
  78. ^ Kellerer, Hans; Pferschy, Ulrich; Pisinger, David (2004). Knapsack Problems | Hans Kellerer | Springer. Springer. doi:10.1007/978-3-540-24777-7. ISBN 978-3-540-40286-2. S2CID 28836720. Archived from the original on October 18, 2017. Retrieved September 19, 2017.
  79. ^ For instance, the volume of a convex polytope (described using a membership oracle) can be approximated to high accuracy by a randomized polynomial time algorithm, but not by a deterministic one: see Dyer, Martin; Frieze, Alan; Kannan, Ravi (January 1991). "A Random Polynomial-time Algorithm for Approximating the Volume of Convex Bodies". J. ACM. 38 (1): 1–17. CiteSeerX 10.1.1.145.4600. doi:10.1145/102782.102783. S2CID 13268711.
  80. ^ George B. Dantzig and Mukund N. Thapa. 2003. Linear Programming 2: Theory and Extensions. Springer-Verlag.
  81. ^ Tsypkin (1971). Adaptation and learning in automatic systems. Academic Press. p. 54. ISBN 978-0-08-095582-7.
  82. ^ a b Kowalski, Robert (1979). "Algorithm=Logic+Control". Communications of the ACM. 22 (7): 424–436. doi:10.1145/359131.359136. S2CID 2509896.
  83. ^ a b Warren, D.S., 2023. Writing correct prolog programs. In Prolog: The Next 50 Years (pp. 62-70). Cham: Springer Nature Switzerland.
  84. ^ Kowalski, R., Dávila, J., Sartor, G. and Calejo, M., 2023. Logical English for law and education. In Prolog: The Next 50 Years (pp. 287–299). Cham: Springer Nature Switzerland.
  85. ^ "The Experts: Does the Patent System Encourage Innovation?". The Wall Street Journal. May 16, 2013. ISSN 0099-9660. Retrieved March 29, 2017.
  86. ^ Knuth, Donald E. (1972). "Ancient Babylonian Algorithms" (PDF). Commun. ACM. 15 (7): 671–677. doi:10.1145/361454.361514. ISSN 0001-0782. S2CID 7829945. Archived from the original (PDF) on December 24, 2012.
  87. ^ Aaboe, Asger (2001). Episodes from the Early History of Astronomy. New York: Springer. pp. 40–62. ISBN 978-0-387-95136-2.
  88. ^ Ast, Courtney. "Eratosthenes". Wichita State University: Department of Mathematics and Statistics. Archived from the original on February 27, 2015. Retrieved February 27, 2015.
  89. ^ Chabert, Jean-Luc (2012). A History of Algorithms: From the Pebble to the Microchip. Springer Science & Business Media. p. 2. ISBN 9783642181924.
  90. ^ Davis 2000:18
  91. ^ Bolter 1984:24
  92. ^ Bolter 1984:26
  93. ^ Bolter 1984:33–34, 204–206.
  94. ^ All quotes from W. Stanley Jevons 1880 Elementary Lessons in Logic: Deductive and Inductive, Macmillan and Co., London and New York. Republished as a googlebook; cf Jevons 1880:199–201. Louis Couturat 1914 the Algebra of Logic, The Open Court Publishing Company, Chicago and London. Republished as a googlebook; cf Couturat 1914 in:75–76 gives a few more details; he compares this to a typewriter as well as a piano. Jevons states that the account is to be found at January 20, 1870 The Proceedings of the Royal Society.
  95. ^ Jevons 1880:199–200
  96. ^ All quotes from John Venn 1881 Symbolic Logic, Macmillan and Co., London. Republished as a googlebook. cf Venn 1881:120–125. The interested reader can find a deeper explanation in those pages.
  97. ^ Bell and Newell diagram 1971:39, cf. Davis 2000
  98. ^ * Melina Hill, Valley News Correspondent, A Tinkerer Gets a Place in History, Valley News West Lebanon NH, Thursday, March 31, 1983, p. 13.
  99. ^ Davis 2000:14
  100. ^ van Heijenoort 1967:81ff
  101. ^ van Heijenoort's commentary on Frege's Begriffsschrift, a formula language, modeled upon that of arithmetic, for pure thought in van Heijenoort 1967:1
  102. ^ Dixon 1906, cf. Kleene 1952:36–40
  103. ^ cf. footnote in Alonzo Church 1936a in Davis 1965:90 and 1936b in Davis 1965:110
  104. ^ Kleene 1935–6 in Davis 1965:237ff, Kleene 1943 in Davis 1965:255ff
  105. ^ Church 1936 in Davis 1965:88ff
  106. ^ cf. "Finite Combinatory Processes – formulation 1", Post 1936 in Davis 1965:289–290
  107. ^ Turing 1936–37 in Davis 1965:116ff
  108. ^ Rosser 1939 in Davis 1965:226
  109. ^ Kleene 1943 in Davis 1965:273–274
  110. ^ Kleene 1952:300, 317
  111. ^ Kleene 1952:376
  112. ^ Turing 1936–37 in Davis 1965:289–290
  113. ^ Turing 1936 in Davis 1965, Turing 1939 in Davis 1965:160
  114. ^ Hodges, p. 96
  115. ^ Turing 1936–37:116
  116. ^ a b Turing 1936–37 in Davis 1965:136
  117. ^ Turing 1939 in Davis 1965:160

Bibliography[edit]

  • Zaslavsky, C. (1970). Mathematics of the Yoruba People and of Their Neighbors in Southern Nigeria. The Two-Year College Mathematics Journal, 1(2), 76–99. https://doi.org/10.2307/3027363

Further reading[edit]

External links[edit]

Algorithm repositories