Function (computer programming)

Last updated

In computer programming, a function, subprogram, procedure, method, routine or subroutine is a callable unit [1] that has a well-defined behavior and can be invoked by other software units to exhibit that behavior.

Contents

Callable units provide a powerful programming tool. [2] The primary purpose is to allow for the decomposition of a large and/or complicated problem into chunks that have relatively low cognitive load and to assign the chunks meaningful names (unless they are anonymous). Judicious application can reduce the cost of developing and maintaining software, while increasing its quality and reliability. [3]

Callable units are present at multiple levels of abstraction in the programming environment. For example, a programmer may write a function in source code that is compiled to machine code that implements similar semantics. There is a callable unit in the source code and an associated one in the machine code, but they are different kinds of callable units with different implications and features.

History

The idea of a callable unit was initially conceived by John Mauchly and Kathleen Antonelli during their work on ENIAC and recorded in a January 1947 Harvard symposium on "Preparation of Problems for EDVAC-type Machines". [4] Maurice Wilkes, David Wheeler, and Stanley Gill are generally credited with the formal invention of this concept, which they termed a closed sub-routine, [5] [6] contrasted with an open subroutine or macro. [7] However, Alan Turing had discussed subroutines in a paper of 1945 on design proposals for the NPL ACE, going so far as to invent the concept of a return address stack. [8]

The idea of a subroutine was worked out after computing machines had already existed for some time. The arithmetic and conditional jump instructions were planned ahead of time and have changed relatively little, but the special instructions used for procedure calls have changed greatly over the years. The earliest computers and microprocessors, such as the Manchester Baby and the RCA 1802, did not have a single subroutine call instruction. Subroutines could be implemented, but they required programmers to use the call sequence—a series of instructions—at each call site.

Subroutines were implemented in Konrad Zuse's Z4 in 1945.

In 1945, Alan M. Turing used the terms "bury" and "unbury" as a means of calling and returning from subroutines. [9] [10]

In January 1947 John Mauchly presented general notes at 'A Symposium of Large Scale Digital Calculating Machinery' under the joint sponsorship of Harvard University and the Bureau of Ordnance, United States Navy. Here he discusses serial and parallel operation suggesting

...the structure of the machine need not be complicated one bit. It is possible, since all the logical characteristics essential to this procedure are available, to evolve a coding instruction for placing the subroutines in the memory at places known to the machine, and in such a way that they may easily be called into use.

In other words, one can designate subroutine A as division and subroutine B as complex multiplication and subroutine C as the evaluation of a standard error of a sequence of numbers, and so on through the list of subroutines needed for a particular problem. ... All these subroutines will then be stored in the machine, and all one needs to do is make a brief reference to them by number, as they are indicated in the coding. [4]

Kay McNulty had worked closely with John Mauchly on the ENIAC team and developed an idea for subroutines for the ENIAC computer she was programming during World War II. [11] She and the other ENIAC programmers used the subroutines to help calculate missile trajectories. [11]

Goldstine and von Neumann wrote a paper dated 16 August 1948 discussing the use of subroutines. [12]

Some very early computers and microprocessors, such as the IBM 1620, the Intel 4004 and Intel 8008, and the PIC microcontrollers, have a single-instruction subroutine call that uses a dedicated hardware stack to store return addresses—such hardware supports only a few levels of subroutine nesting, but can support recursive subroutines. Machines before the mid-1960s—such as the UNIVAC I, the PDP-1, and the IBM 1130—typically use a calling convention which saved the instruction counter in the first memory location of the called subroutine. This allows arbitrarily deep levels of subroutine nesting but does not support recursive subroutines. The IBM System/360 had a subroutine call instruction that placed the saved instruction counter value into a general-purpose register; this can be used to support arbitrarily deep subroutine nesting and recursive subroutines. The Burroughs B5000 [13] (1961) is one of the first computers to store subroutine return data on a stack.

The DEC PDP-6 [14] (1964) is one of the first accumulator-based machines to have a subroutine call instruction that saved the return address in a stack addressed by an accumulator or index register. The later PDP-10 (1966), PDP-11 (1970) and VAX-11 (1976) lines followed suit; this feature also supports both arbitrarily deep subroutine nesting and recursive subroutines. [15]

Language support

In the very early assemblers, subroutine support was limited. Subroutines were not explicitly separated from each other or from the main program, and indeed the source code of a subroutine could be interspersed with that of other subprograms. Some assemblers would offer predefined macros to generate the call and return sequences. By the 1960s, assemblers usually had much more sophisticated support for both inline and separately assembled subroutines that could be linked together.

One of the first programming languages to support user-written subroutines and functions was FORTRAN II. The IBM FORTRAN II compiler was released in 1958. ALGOL 58 and other early programming languages also supported procedural programming.

Libraries

Even with this cumbersome approach, subroutines proved very useful. They allowed the use of the same code in many different programs. Memory was a very scarce resource on early computers, and subroutines allowed significant savings in the size of programs.

Many early computers loaded the program instructions into memory from a punched paper tape. Each subroutine could then be provided by a separate piece of tape, loaded or spliced before or after the main program (or "mainline" [16] ); and the same subroutine tape could then be used by many different programs. A similar approach was used in computers that loaded program instructions from punched cards. The name subroutine library originally meant a library, in the literal sense, which kept indexed collections of tapes or decks of cards for collective use.

Return by indirect jump

To remove the need for self-modifying code, computer designers eventually provided an indirect jump instruction, whose operand, instead of being the return address itself, was the location of a variable or processor register containing the return address.

On those computers, instead of modifying the function's return jump, the calling program would store the return address in a variable so that when the function completed, it would execute an indirect jump that would direct execution to the location given by the predefined variable.

Jump to subroutine

Another advance was the jump to subroutine instruction, which combined the saving of the return address with the calling jump, thereby minimizing overhead significantly.

In the IBM System/360, for example, the branch instructions BAL or BALR, designed for procedure calling, would save the return address in a processor register specified in the instruction, by convention register 14. To return, the subroutine had only to execute an indirect branch instruction (BR) through that register. If the subroutine needed that register for some other purpose (such as calling another subroutine), it would save the register's contents to a private memory location or a register stack.

In systems such as the HP 2100, the JSB instruction would perform a similar task, except that the return address was stored in the memory location that was the target of the branch. Execution of the procedure would actually begin at the next memory location. In the HP 2100 assembly language, one would write, for example

... JSB MYSUB    (Calls subroutine MYSUB.) BB    ...          (Will return here after MYSUB is done.) 

to call a subroutine called MYSUB from the main program. The subroutine would be coded as

MYSUB NOP          (Storage for MYSUB's return address.) AA    ...          (Start of MYSUB's body.) ... JMP MYSUB,I  (Returns to the calling program.) 

The JSB instruction placed the address of the NEXT instruction (namely, BB) into the location specified as its operand (namely, MYSUB), and then branched to the NEXT location after that (namely, AA = MYSUB + 1). The subroutine could then return to the main program by executing the indirect jump JMP MYSUB, I which branched to the location stored at location MYSUB.

Compilers for Fortran and other languages could easily make use of these instructions when available. This approach supported multiple levels of calls; however, since the return address, parameters, and return values of a subroutine were assigned fixed memory locations, it did not allow for recursive calls.

Incidentally, a similar method was used by Lotus 1-2-3, in the early 1980s, to discover the recalculation dependencies in a spreadsheet. Namely, a location was reserved in each cell to store the return address. Since circular references are not allowed for natural recalculation order, this allows a tree walk without reserving space for a stack in memory, which was very limited on small computers such as the IBM PC.

Call stack

Most modern implementations of a function call use a call stack, a special case of the stack data structure, to implement function calls and returns. Each procedure call creates a new entry, called a stack frame, at the top of the stack; when the procedure returns, its stack frame is deleted from the stack, and its space may be used for other procedure calls. Each stack frame contains the private data of the corresponding call, which typically includes the procedure's parameters and internal variables, and the return address.

The call sequence can be implemented by a sequence of ordinary instructions (an approach still used in reduced instruction set computing (RISC) and very long instruction word (VLIW) architectures), but many traditional machines designed since the late 1960s have included special instructions for that purpose.

The call stack is usually implemented as a contiguous area of memory. It is an arbitrary design choice whether the bottom of the stack is the lowest or highest address within this area, so that the stack may grow forwards or backwards in memory; however, many architectures chose the latter.[ citation needed ]

Some designs, notably some Forth implementations, used two separate stacks, one mainly for control information (like return addresses and loop counters) and the other for data. The former was, or worked like, a call stack and was only indirectly accessible to the programmer through other language constructs while the latter was more directly accessible.

When stack-based procedure calls were first introduced, an important motivation was to save precious memory.[ citation needed ] With this scheme, the compiler does not have to reserve separate space in memory for the private data (parameters, return address, and local variables) of each procedure. At any moment, the stack contains only the private data of the calls that are currently active (namely, which have been called but haven't returned yet). Because of the ways in which programs were usually assembled from libraries, it was (and still is) not uncommon to find programs that include thousands of functions, of which only a handful are active at any given moment.[ citation needed ] For such programs, the call stack mechanism could save significant amounts of memory. Indeed, the call stack mechanism can be viewed as the earliest and simplest method for automatic memory management.

However, another advantage of the call stack method is that it allows recursive function calls, since each nested call to the same procedure gets a separate instance of its private data.

In a multi-threaded environment, there is generally more than one stack. [17] An environment that fully supports coroutines or lazy evaluation may use data structures other than stacks to store their activation records.

Delayed stacking

One disadvantage of the call stack mechanism is the increased cost of a procedure call and its matching return.[ clarification needed ] The extra cost includes incrementing and decrementing the stack pointer (and, in some architectures, checking for stack overflow), and accessing the local variables and parameters by frame-relative addresses, instead of absolute addresses. The cost may be realized in increased execution time, or increased processor complexity, or both.

This overhead is most obvious and objectionable in leaf procedures or leaf functions, which return without making any procedure calls themselves. [18] [19] [20] To reduce that overhead, many modern compilers try to delay the use of a call stack until it is really needed.[ citation needed ] For example, the call of a procedure P may store the return address and parameters of the called procedure in certain processor registers, and transfer control to the procedure's body by a simple jump. If the procedure P returns without making any other call, the call stack is not used at all. If P needs to call another procedure Q, it will then use the call stack to save the contents of any registers (such as the return address) that will be needed after Q returns.

Features

In general, a callable unit is a list of instructions that, starting at the first instruction, executes sequentially except as directed via its internal logic. It can be invoked (called) many times during the execution of a program. Execution continues at the next instruction after the call instruction when it returns control.

Implementations

The features of implementations of callable units evolved over time and varies by context. This section describes features of the various common implementations.

General characteristics

Most modern programming languages provide features to define and call functions, including syntax for accessing such features, including:

Naming

Some languages, such as Pascal, Fortran, Ada and many dialects of BASIC, use a different name for a callable unit that returns a value (function or subprogram) vs. one that does not (subroutine or procedure). Other languages, such as C, C++, C# and Lisp, use only one name for a callable unit, function. The C-family languages use the keyword void to indicate no return value.

Call syntax

If declared to return a value, a call can be embedded in an expression in order to consume the return value. For example, a square root callable unit might be called like y = sqrt(x).

A callable unit that does not return a value is called as a stand-alone statement like print("hello"). This syntax can also be used for a callable unit that returns a value, but the return value will be ignored.

Some older languages require a keyword for calls that do not consume a return value, like CALL print("hello").

Parameters

Most implementations, especially in modern languages, support parameters which the callable declares as formal parameters. A caller passes actual parameters, a.k.a. arguments, to match. Different programming languages provide different conventions for passing arguments.

ConventionDescriptionUsed in
by value A copy of the argment is passedDefault in most Algol-like languages after Algol 60, such as Pascal, Delphi, Simula, CPL, PL/M, Modula, Oberon, Ada, and many others including C, C++ and Java
by reference A reference to the argument is passed; typically its addressSelectable in most Algol-like languages after Algol 60, such as Algol 68, Pascal, Delphi, Simula, CPL, PL/M, Modula, Oberon, Ada, and many others including C++, Fortran, PL/I
by result The value computed during the call is copied to the argument on returnAda OUT parameters
by value-result A copy of the argument is passed in and the value computed during the call is copied to the argument on returnAlgol, Swift in-out parameters
by name Like a macro replace the parameters with the unevaluated argument expressions, then evaluate the argument in the context of the caller every time that the callable uses the parameterAlgol, Scala
by constant valueLike by-value except that the parameter is treated as a constantPL/I NONASSIGNABLE parameters, Ada IN parameters

Return value

In some languages, such as BASIC, a callable has different syntax (i.e. keyword) for a callable that returns a value vs. one that does not. In other languages, the syntax is the same regardless. In some of these languages an extra keyword is used to declare no return value; for example void in C, C++ and C#. In some languages, such as Python, the difference is whether the body contains a return statement with a value, and a particular callable may return with or without a value based on control flow.

Side effects

In many contexts, a callable may have side effect behavior such as modifying passed or global data, reading from or writing to a peripheral device, accessing a file, halting the program or the machine, or temporarily pausing program execution.

In strictly functional programming languages such as Haskell, a function can have no side effects, which means it cannot change the state of the program. Functions always return the same result for the same input. Such languages typically only support functions that return a value, since there is no value in a function that has neither return value nor side effect.

Local variables

Most contexts support local variables memory owned by a callable to hold intermediate values. These variables are typically stored in the call's activation record on the call stack along with other information such as the return address.

Nested call recursion

If supported by the language, a callable may call itself, causing its execution to suspend while another nested execution of the same callable executes. Recursion is a useful means to simplify some complex algorithms and break down complex problems. Recursive languages provide a new copy of local variables on each call. If the programmer desires the recursive callable to use the same variables instead of using locals, they typically declare them in a shared context such static or global.

Languages going back to ALGOL, PL/I and C and modern languages, almost invariably use a call stack, usually supported by the instruction sets to provide an activation record for each call. That way, a nested call can modify its local variables without effecting any of the suspended calls variables.

Recursion allows direct implementation of functionality defined by mathematical induction and recursive divide and conquer algorithms. Here is an example of a recursive function in C/C++ to find Fibonacci numbers:

intFib(intn){if(n<=1){returnn;}returnFib(n-1)+Fib(n-2);}

Early languages like Fortran did not initially support recursion because only one set of variables and return address were allocated for each callable. [21] Early computer instruction sets made storing return addresses and variables on a stack difficult. Machines with index registers or general-purpose registers, e.g., CDC 6000 series, PDP-6, GE 635, System/360, UNIVAC 1100 series, could use one of those registers as a stack pointer.

Nested scope

Some languages, e.g., Ada, Pascal, PL/I, Python, support declaring and defining a function inside, e.g., a function body, such that the name of the inner is only visible within the body of the outer.

Reentrancy

If a callable can be executed properly even when another execution of the same callable is already in progress, that callable is said to be reentrant . A reentrant callable is also useful in multi-threaded situations since multiple threads can call the same callable without fear of interfering with each other. In the IBM CICS transaction processing system, quasi-reentrant was a slightly less restrictive, but similar, requirement for application programs that were shared by many threads.

Overloading

Some languages support overloading allow multiple callables with the same name in the same scope, but operating on different types of input. Consider the square root function applied to real number, complex number and matrix input. The algorithm for each type of input is different, and the return value may have a different type. By writing three separate callables with the same name. i.e. sqrt, the resulting code may be easier to write and to maintain since each one has a name that is relatively easy to understand and to remember instead of giving longer and more complicated names like sqrt_real, sqrt_complex, qrt_matrix.

Overloading is supported in many languages that support strong typing. Often the compiler selects the overload to call based on the type of the input arguments or it fails if the input arguments do not select an overload. Older and weakly-typed languages generally do not support overloading.

Here is an example of overloading in C++, two functions Area that accept different types:

// returns the area of a rectangle defined by height and widthdoubleArea(doubleh,doublew){returnh*w;}// returns the area of a circle defined by radiusdoubleArea(doubler){returnr*r*3.14;}intmain(){doublerectangle_area=Area(3,4);doublecircle_area=Area(5);}

PL/I has the GENERIC attribute to define a generic name for a set of entry references called with different types of arguments. Example:

DECLARE gen_name GENERIC(     name  WHEN(FIXED BINARY),     flame  WHEN(FLOAT),     pathname OTHERWISE);

Multiple argument definitions may be specified for each entry. A call to "gen_name" will result in a call to "name" when the argument is FIXED BINARY, "flame" when FLOAT", etc. If the argument matches none of the choices "pathname" will be called.

Closure

A closure is a callable plus values of some of its variables captured from the environment in which it was created. Closures were a notable feature of the Lisp programming language, introduced by John McCarthy. Depending on the implementation, closures can serve as a mechanism for side-effects.

Exception reporting

Besides its happy path behavior, a callable may need to inform the caller about an exceptional condition that occurred during its execution.

Most modern languages support exceptions which allows for exceptional control flow that pops the call stack until an exception handler is found to handle the condition.

Languages that do not support exceptions can use the return value to indicate success or failure of a call. Another approach is to use a well-known location like a global variable for success indication. A callable writes the value and the caller reads it after a call.

In the IBM System/360, where return code was expected from a subroutine, the return value was often designed to be a multiple of 4—so that it could be used as a direct branch table index into a branch table often located immediately after the call instruction to avoid extra conditional tests, further improving efficiency. In the System/360 assembly language, one would write, for example:

           BAL  14, SUBRTN01    go to a subroutine, storing return address in R14            B    TABLE(15)      use returned value in reg 15 to index the branch table, *                              branching to the appropriate branch instr. TABLE      B    OK             return code =00   GOOD                  }            B    BAD            return code =04   Invalid input         } Branch table            B    ERROR          return code =08   Unexpected condition  } 

Call overhead

A call has runtime overhead, which may include but is not limited to:

Various techniques are employed to minimize the runtime cost of calls.

Compiler optimization

Some optimizations for minimizing call overhead may seem straight forward, but cannot be used if the callable has side effects. For example, in the expression (f(x)-1)/(f(x)+1), the function f cannot be called only once with its value used two times since the two calls may return different results. Moreover, in the few languages which define the order of evaluation of the division operator's operands, the value of x must be fetched again before the second call, since the first call may have changed it. Determining whether a callable has a side effect is difficult indeed, undecidable by virtue of Rice's theorem. So, while this optimization is safe in a purely functional programming language, a compiler for an language not limited to functional typically assumes the worst case, that every callable may have side effects.

Inlining

Inlining eliminates calls for particular callables. The compiler replaces each call with the compiled code of the callable. Not only does this avoid the call overhead, but it also allows the compiler to optimize code of the caller more effectively by taking into account the context and arguments at that call. Inlining, however, usually increases the compiled code size, except when only called once or the body is very short, like one line.

Sharing

Callables can be defined within a program, or separately in a library that can be used by multiple programs.

Inter-operability

A compiler translates call and return statements into machine instructions according to a well-defined calling convention. For code compiled by the same or a compatible compiler, functions can be compiled separately from the programs that call them. The instruction sequences corresponding to call and return statements are called the procedure's prologue and epilogue.

Built-in functions

A built-in function, or builtin function, or intrinsic function, is a function for which the compiler generates code at compile time or provides in a way other than for other functions. [22] A built-in function does not need to be defined like other functions since it is built in to the programming language. [23]

Programming

Trade-offs

Advantages

Advantages of breaking a program into functions include:

  • Decomposing a complex programming task into simpler steps: this is one of the two main tools of structured programming, along with data structures
  • Reducing duplicate code within a program
  • Enabling reuse of code across multiple programs
  • Dividing a large programming task among various programmers or various stages of a project
  • Hiding implementation details from users of the function
  • Improving readability of code by replacing a block of code with a function call where a descriptive function name serves to describe the block of code. This makes the calling code concise and readable even if the function is not meant to be reused.
  • Improving traceability (i.e. most languages offer ways to obtain the call trace which includes the names of the involved functions and perhaps even more information such as file names and line numbers); by not decomposing the code into functions, debugging would be severely impaired

Disadvantages

Compared to using in-line code, invoking a function imposes some computational overhead in the call mechanism.[ citation needed ]

A function typically requires standard housekeeping code – both at the entry to, and exit from, the function (function prologue and epilogue – usually saving general purpose registers and return address as a minimum).

Conventions

Many programming conventions have been developed regarding callables.

With respect to naming, many developers name a callable with a phrase starting with a verb when it does a certain task, with an adjective when it makes an inquiry, and with a noun when it is used to substitute variables.

Some programmers suggest that a callable should perform exactly one task, and if it performs more than one task, it should be split up into multiple callables. They argue that callables are key components in software maintenance, and their roles in the program must remain distinct.

Proponents of modular programming advocate that each callable should have minimal dependency on the rest of the codebase. For example, the use of global variables is generally deemed unwise, because it adds coupling between all callables that use the global variables. If such coupling is not necessary, they advise to refactor callables to accept passed parameters instead.

Examples

Early BASIC

Early BASIC variants require each line to have a unique number (line number) that orders the lines for execution, provides no separation of the code that is callable, no mechanism for passing arguments or to return a value and all variables are global. It provides the command GOSUB where sub is short for sub procedure, subprocedure or subroutine. Control jumps to the specified line number and then continues on the next line on return.

10REM A BASIC PROGRAM20GOSUB10030GOTO20100INPUTGIVEMEANUMBER;N110PRINTTHESQUAREROOTOF;N;120PRINTIS;SQRT(N)130RETURN

This code repeatedly asks the user to enter a number and reports the square root of the value. Lines 100-130 are the callable.

Small Basic

In Microsoft Small Basic, targeted to the student first leaning how to program in a text-based language, a callable unit is called a subroutine. The Sub keyword denotes the start of a subroutine and is followed by a name identifier. Subsequent lines are the body which ends with the EndSub keyword. [24]

SubSayHelloTextWindow.WriteLine("Hello!")EndSub

This can be called as SayHello(). [25]

Visual Basic

In later versions of Visual Basic (VB), including the latest product line and VB6, the term procedure is used for the callable unit concept. The keyword Sub is used to return no value and Function to return a value. When used in the context of a class, a procedure is a method. [26]

Each parameter has a data type that can be specified, but if not, defaults to Object for later versions based on .NET and variant for VB6. [27]

VB supports parameter passing conventions by value and by reference via the keywords ByVal and ByRef, respectively. Unless ByRef is specified, an argument is passed ByVal. Therefore, ByVal is rarely explicitly specified.

For a simple type like a number these conventions are relatively clear. Passing ByRef allows the procedure to modify the passed variable whereas passing ByVal does not. For an object, semantics can confuse programmers since an object is always treated as a reference. Passing an object ByVal copies the reference; not the state of the object. The called procedure can modify the state of the object via its methods yet cannot modify the object reference of the actual parameter.

SubDoSomething()' Some Code HereEndSub

The does not return a value and has to be called stand-alone, like DoSomething

FunctionGiveMeFive()asIntegerGiveMeFive=5EndFunction

This returns the value 5, and a call can be part of an expression like y = x + GiveMeFive()

SubAddTwo(ByRefintValueasInteger)intValue=intValue+2EndSub

This has a side-effect modifies the variable passed by reference and could be called for variable v like AddTwo(v). Giving v is 5 before the call, it will be 7 after.

C and C++

In C and C++, a callable unit is called a function. A function definition starts with the name of the type of value that it returns or void to indicate that it does not return a value. This is followed by the function name, formal arguments in parentheses, and body lines in braces.

In C++, a function declared in a class (as non-static) is called a member function or method. A function outside of a class can be called a free function to distinguish it from a member function. [28]

voiddoSomething(){/* some code */}

This function does not return a value and is always called stand-alone, like doSomething()

intgiveMeFive(){return5;}

This function returns the integer value 5. The call can be stand-alone or in an expression like y = x + giveMeFive()

voidaddTwo(int*pi){*pi+=2;}

This function has a side-effect modifies the value passed by address to the input value plus 2. It could be called for variable v as addTwo(&v) where the ampersand (&) tells the compiler to pass the address of a variable. Giving v is 5 before the call, it will be 7 after.

voidaddTwo(int&i){i+=2;}

This function requires C++ would not compile as C. It has the same behavior as the preceding example but passes the actual parameter by reference rather than passing its address. A call such as addTwo(v) does not include an ampersand since the compiler handles passing by reference without syntax in the call.

PL/I

In PL/I a called procedure may be passed a descriptor providing information about the argument, such as string lengths and array bounds. This allows the procedure to be more general and eliminates the need for the programmer to pass such information. By default PL/I passes arguments by reference. A (trivial) function to change the sign of each element of a two-dimensional array might look like:

change_sign: procedure(array);   declare array(*,*) float;   array = -array; end change_sign; 

This could be called with various arrays as follows:

/* first array bounds from -5 to +10 and 3 to 9 */ declare array1 (-5:10, 3:9)float; /* second array bounds from 1 to 16 and 1 to 16 */ declare array2 (16,16) float; call change_sign(array1); call change_sign(array2); 

Python

In Python, the keyword def denotes the start of a function definition. The statements of the function body follow as indented on subsequent lines and end at the line that is indented the same as the first line or end of file. [29]

defformat_greeting(name):return"Welcome "+namedefgreet_martin():print(format_greeting("Martin"))

The first function returns greeting text that includes the name passed by the caller. The second function calls the first and is called like greet_martin() to write "Welcome Martin" to the console.

Prolog

In the procedural interpretation of logic programs, logical implications behave as goal-reduction procedures. A rule (or clause) of the form:

A :- B

which has the logical reading:

A if B

behaves as a procedure that reduces goals that unify with A to subgoals that are instances ofB.

Consider, for example, the Prolog program:

mother_child(elizabeth,charles).father_child(charles,william).father_child(charles,harry).parent_child(X,Y):-mother_child(X,Y).parent_child(X,Y):-father_child(X,Y).

Notice that the motherhood function, X=mother(Y) is represented by a relation, as in a relational database. However, relations in Prolog function as callable units.

For example, the procedure call ?-parent_child(X,charles) produces the output X=elizabeth. But the same procedure can be called with other input-output patterns. For example:

?-parent_child(elizabeth,Y).Y=charles.?-parent_child(X,Y).X=elizabeth,Y=charles.X=charles,Y=harry.X=charles,Y=william.?-parent_child(william,harry).no.?-parent_child(elizabeth,charles).yes.

See also

Related Research Articles

Forth is a procedural, concatenative, stack-oriented programming language and interactive integrated development environment designed by Charles H. "Chuck" Moore and first used by other programmers in 1970. Although not an acronym, the language's name in its early years was often spelled in all capital letters as FORTH. The FORTH-79 and FORTH-83 implementations, which were not written by Moore, became de facto standards, and an official technical standard of the language was published in 1994 as ANS Forth. A wide range of Forth derivatives existed before and after ANS Forth. The free and open-source software Gforth implementation is actively maintained, as are several commercially supported systems.

In computer science, threaded code is a programming technique where the code has a form that essentially consists entirely of calls to subroutines. It is often used in compilers, which may generate code in that form or be implemented in that form themselves. The code may be processed by an interpreter or it may simply be a sequence of machine code call instructions.

In computer science, imperative programming is a programming paradigm of software that uses statements that change a program's state. In much the same way that the imperative mood in natural languages expresses commands, an imperative program consists of commands for the computer to perform. Imperative programming focuses on describing how a program operates step by step, rather than on high-level descriptions of its expected results.

In computing, inline expansion, or inlining, is a manual or compiler optimization that replaces a function call site with the body of the called function. Inline expansion is similar to macro expansion, but occurs during compilation, without changing the source code, while macro expansion occurs prior to compilation, and results in different text that is then processed by the compiler.

<span class="mw-page-title-main">Stack (abstract data type)</span> Abstract data type

In computer science, a stack is an abstract data type that serves as a collection of elements with two main operations:

In computer programming, a parameter or a formal argument is a special kind of variable used in a subroutine to refer to one of the pieces of data provided as input to the subroutine. These pieces of data are the values of the arguments with which the subroutine is going to be called/invoked. An ordered list of parameters is usually included in the definition of a subroutine, so that, each time the subroutine is called, its arguments for that call are evaluated, and the resulting values can be assigned to the corresponding parameters.

The Burroughs Large Systems Group produced a family of large 48-bit mainframes using stack machine instruction sets with dense syllables. The first machine in the family was the B5000 in 1961, which was optimized for compiling ALGOL 60 programs extremely well, using single-pass compilers. The B5000 evolved into the B5500 and the B5700. Subsequent major redesigns include the B6500/B6700 line and its successors, as well as the separate B8500 line.

<span class="mw-page-title-main">Pointer (computer programming)</span> Object which stores memory addresses in a computer program

In computer science, a pointer is an object in many programming languages that stores a memory address. This can be that of another value located in computer memory, or in some cases, that of memory-mapped computer hardware. A pointer references a location in memory, and obtaining the value stored at that location is known as dereferencing the pointer. As an analogy, a page number in a book's index could be considered a pointer to the corresponding page; dereferencing such a pointer would be done by flipping to the page with the given page number and reading the text found on that page. The actual format and content of a pointer variable is dependent on the underlying computer architecture.

In computer science, computer engineering and programming language implementations, a stack machine is a computer processor or a virtual machine in which the primary interaction is moving short-lived temporary values to and from a push down stack. In the case of a hardware processor, a hardware stack is used. The use of a stack significantly reduces the required number of processor registers. Stack machines extend push-down automata with additional load/store operations or multiple stacks and hence are Turing-complete.

In computer programming, a thunk is a subroutine used to inject a calculation into another subroutine. Thunks are primarily used to delay a calculation until its result is needed, or to insert operations at the beginning or end of the other subroutine. They have many other applications in compiler code generation and modular programming.

In computer programming, a return statement causes execution to leave the current subroutine and resume at the point in the code immediately after the instruction which called the subroutine, known as its return address. The return address is saved by the calling routine, today usually on the process's call stack or in a register. Return statements in many programming languages allow a function to specify a return value to be passed back to the code that called the function.

In computer science, a tail call is a subroutine call performed as the final action of a procedure. If the target of a tail is the same subroutine, the subroutine is said to be tail recursive, which is a special case of direct recursion. Tail recursion is particularly useful, and is often easy to optimize in implementations.

In computer programming, a nested function is a named function that is defined within another, enclosing, block and is lexically scoped within the enclosing block – meaning it is only callable by name within the body of the enclosing block and can use identifiers declared in outer blocks, including outer functions. The enclosing block is typically, but not always, another function.

In computer science, a call stack is a stack data structure that stores information about the active subroutines of a computer program. This type of stack is also known as an execution stack, program stack, control stack, run-time stack, or machine stack, and is often shortened to simply "the stack". Although maintenance of the call stack is important for the proper functioning of most software, the details are normally hidden and automatic in high-level programming languages. Many computer instruction sets provide special instructions for manipulating stacks.

In computer science, a calling convention is an implementation-level (low-level) scheme for how subroutines or functions receive parameters from their caller and how they return a result. When some code calls a function, design choices have been taken for where and how parameters are passed to that function, and where and how results are returned from that function, with these transfers typically done via certain registers or within a stack frame on the call stack. There are design choices for how the tasks of preparing for a function call and restoring the environment after the function has completed are divided between the caller and the callee. Some calling convention specifies the way every function should get called. The correct calling convention should be used for every function call, to allow the correct and reliable execution of the whole program using these functions.

<span class="mw-page-title-main">Recursion (computer science)</span> Use of functions that call themselves

In computer science, recursion is a method of solving a computational problem where the solution depends on solutions to smaller instances of the same problem. Recursion solves such recursive problems by using functions that call themselves from within their own code. The approach can be applied to many types of problems, and recursion is one of the central ideas of computer science.

The power of recursion evidently lies in the possibility of defining an infinite set of objects by a finite statement. In the same manner, an infinite number of computations can be described by a finite recursive program, even if this program contains no explicit repetitions.

This article describes the calling conventions used when programming x86 architecture microprocessors.

Little Computer 3, or LC-3, is a type of computer educational programming language, an assembly language, which is a type of low-level programming language.

This article attempts to set out the various similarities and differences between the various programming paradigms as a summary in both graphical and tabular format with links to the separate discussions concerning these similarities and differences in extant Wikipedia articles.

Toi is an imperative, type-sensitive language that provides the basic functionality of a programming language. The language was designed and developed from the ground-up by Paul Longtine. Written in C, Toi was created with the intent to be an educational experience and serves as a learning tool for those looking to familiarize themselves with the inner-workings of a programming language.

References

  1. "Terminology Glossary". nist.gov. NIST. Retrieved 9 February 2024. Callable unit: (Of a software program or logical design) Function, method, operation, subroutine, procedure, or analogous structural unit that appears within a module.
  2. Donald E. Knuth (1997). The Art of Computer Programming, Volume I: Fundamental Algorithms. Addison-Wesley. ISBN   0-201-89683-4.
  3. O.-J. Dahl; E. W. Dijkstra; C. A. R. Hoare (1972). Structured Programming. Academic Press. ISBN   0-12-200550-3.
  4. 1 2 Mauchly, J.W. (1982). "Preparation of Problems for EDVAC-Type Machines". In Randell, Brian (ed.). The Origins of Digital Computers. Springer. pp. 393–397. doi:10.1007/978-3-642-61812-3_31. ISBN   978-3-642-61814-7.
  5. Wheeler, D. J. (1952). "The use of sub-routines in programmes" (PDF). Proceedings of the 1952 ACM national meeting (Pittsburgh) on - ACM '52. p. 235. doi: 10.1145/609784.609816 .
  6. Wilkes, M. V.; Wheeler, D. J.; Gill, S. (1951). Preparation of Programs for an Electronic Digital Computer. Addison-Wesley.
  7. Dainith, John (2004). ""open subroutine." A Dictionary of Computing". Encyclopedia.com. Retrieved 14 January 2013.
  8. Turing, Alan M. (1945), Report by Dr. A.M. Turing on proposals for the development of an Automatic Computing Engine (ACE): Submitted to the Executive Committee of the NPL in February 1946 reprinted in Copeland, B. J., ed. (2005). Alan Turing's Automatic Computing Engine . Oxford: Oxford University Press. p. 383. ISBN   0-19-856593-3.
  9. Turing, Alan Mathison (19 March 1946) [1945], Proposals for Development in the Mathematics Division of an Automatic Computing Engine (ACE) (NB. Presented on 1946-03-19 before the Executive Committee of the National Physical Laboratory (Great Britain).)
  10. Carpenter, Brian Edward; Doran, Robert William (1 January 1977) [October 1975]. "The other Turing machine". The Computer Journal . 20 (3): 269–279. doi: 10.1093/comjnl/20.3.269 . (11 pages)
  11. 1 2 Isaacson, Walter (18 September 2014). "Walter Isaacson on the Women of ENIAC". Fortune. Archived from the original on 12 December 2018. Retrieved 14 December 2018.
  12. Herman H. Goldstine; John von Neumann (1947). "Part II, Volume I-3, Planning and Coding of Problems for an Electronic Computing Instrument" (PDF). Report on the Mathematical and Logical aspects of an Electronic Computing Instrument (Technical report). (see p. 163 of the pdf for the relevant page)
  13. The Operational Characteristics of the Processors for the Burroughs B5000 (PDF). Revision A. Burroughs Corporation. 1963. 5000-21005. Retrieved 8 February 2024.
  14. "Push-Down Instructions" (PDF). Programmed Data Processor 6 - Handbook (PDF). p. 37. Retrieved 8 February 2024.
  15. Guy Lewis Steele Jr. AI Memo 443. 'Debunking the "Expensive Procedure Call" Myth; or, Procedure call implementations considered harmful". Section "C. Why Procedure Calls Have a Bad Reputation".
  16. Frank, Thomas S. (1983). Introduction to the PDP-11 and Its Assembly Language. Prentice-Hall software series. Prentice-Hall. p. 195. ISBN   9780134917047 . Retrieved 6 July 2016. We could supply our assembling clerk with copies of the source code for all of our useful subroutines and then when presenting him with a mainline program for assembly, tell him which subroutines will be called in the mainline [...]
  17. Buttlar, Dick; Farrell, Jacqueline; Nichols, Bradford (1996). PThreads Programming: A POSIX Standard for Better Multiprocessing. "O'Reilly Media, Inc.". pp. 2–5. ISBN   978-1-4493-6475-5. OCLC   1036778036.
  18. "ARM Information Center". Infocenter.arm.com. Retrieved 29 September 2013.
  19. "x64 stack usage". Microsoft Docs. Microsoft. Retrieved 5 August 2019.
  20. "Function Types". Msdn.microsoft.com. Retrieved 29 September 2013.
  21. Verhoeff, Tom (2018). "A Master Class on Recursion". In Böckenhauer, Hans-Joachim; Komm, Dennis; Unger, Walter (eds.). Adventures Between Lower Bounds and Higher Altitudes: Essays Dedicated to Juraj Hromkovič on the Occasion of His 60th Birthday. Springer. p. 616. ISBN   978-3-319-98355-4. OCLC   1050567095.
  22. "Built-in functions". ibm.com. 9 March 2017. Retrieved 25 December 2023.
  23. Study Material Python. April 2023. p. 87. Retrieved 25 December 2023.
  24. "Small Basic". Small Basic. Retrieved 8 February 2024.
  25. "Small Basic Getting Started Guide: Chapter 9: Subroutines". Microsoft. 17 January 2024.
  26. "Procedures in Visual Basic". Microsoft Learn. 15 September 2021. Retrieved 8 February 2024.
  27. "Dim statement (Visual Basic)". Microsoft Learn. 15 September 2021. Retrieved 8 February 2024.
  28. "what is meant by a free function".
  29. "4. More Control Flow Tools — Python 3.9.7 documentation".