K

Table of Contents

K is a rewrite-based
executable semantic framework in which programming languages, type
systems and formal analysis tools can be defined using configurations
and rules. Configurations organize the state in units called cells,
which are labeled and can be nested. K rewrite rules make it explicit
which parts of the term are read-only, write-only, read-write, or
unused. This makes K suitable for defining truly concurrent languages
even in the presence of sharing. Computations are represented as
syntactic extensions of the original language abstract syntax, using a
nested list structure which sequentializes computational tasks, such
as program fragments. Computations are like any other terms in a
rewriting environment: they can be matched, moved from one place to
another, modified, or deleted. This makes K suitable for defining
control-intensive features such as abrupt termination, exceptions, or
call/cc.

K Tool Download

  • Install from the latest K GitHub Release.
  • Install pyk, K's scripting interface for Python. Check the API documentation for a complete reference of supported features.
  • Try our Editor Support page for links to K syntax highlighting definitions for various popular editors/IDEs. Please feel free to contribute.
  • Build or browse the code on GitHub, where you can also report bugs.

Learn K

Support

Resources

K Tutorial

The purpose of this series of lessons is to teach developers how to program in
K. While the primary use of K is in the specification of operational semantics
of programming languages, this tutorial is agnostic on how the knowledge of K
is used. For a more detailed tutorial explaining the basic principles of
programming language design, refer to the
K PL Tutorial. Note that that tutorial is somewhat
out of date presently.

This K tutorial is a work in progress. Many lessons are currently simply
placeholders for future content.

To start the K tutorial, begin with
Section 1: Basic Programming in K.

Section 1: Basic K Concepts

The goal of this first section of the K tutorial is to teach the basic
principles of K to someone with no prior experience with K as a programming
language. However, this is not written with the intended audience of someone
who is a complete beginner to programming. We are assuming that the reader
has a firm grounding in computer science broadly, as well as that they have
experience writing code in functional programming languages before.

By the end of this section, the reader ought to be able to write specifications
of simple languages in K, use these specifications to generate a fast
interpreter for their programming language, as well as write basic deductive
program verification proofs over programs in their language. This should give
them the theoretical grounding they need to begin expanding their knowledge
of K in Section 2: Intermediate K Concepts.

To begin this section, refer to
Lesson 1.1: Setting up a K Environment.

Lesson 1.1: Setting up a K Environment

The first step to learning K is to install K on your system, and configure your
editor for K development.

Installing K

You have two options for how to install K, depending on how you intend to
interact with the K codebase. If you are solely a user of K, and have no
interest in developing or making changes to K, you most likely will want to
install one of our binary releases of K. However, if you are going to be a K
developer, or simply want to build K from source, you should follow the
instructions for a source build of K.

Installing K from a binary release

K is developed as a rolling release, with each change to K that passes our
CI infrastructure being deployed on GitHub for download. The latest release of
K can be downloaded here.
This page also contains information on how to install K. It is recommended
that you fully uninstall the old version of K prior to installing the new one,
as K does not maintain entries in package manager databases, with the exception
of Homebrew on MacOS.

Installing K from source

You can clone K from GitHub with the following Git command:

git clone https://github.com/runtimeverification/k --recursive

Instructions on how to build K from source can be found
here.

Configuring your editor

K maintains a set of scripts for a variety of text editors, including vim and
emacs, in various states of maintenance. You can download these scripts with
the following Git command:

git clone https://github.com/kframework/k-editor-support

Because K allows users to define their own grammars for parsing K itself,
not all features of K can be effectively highlighted. However, at the cost of
occasionally highlighting things incorrectly, you can get some pretty good
results in many cases. With that being said, some of the editor scripts in the
above repository are pretty out of date. If you manage to improve them, we
welcome pull requests into the repository.

Troubleshooting

If you have problems installing K, we encourage you to reach out to us. If you
follow the above install instructions and run into a problem, you can
Create a bug report on GitHub

Next lesson

Once you have set up K on your system to your satisfaction, you can continue to
Lesson 1.2: Basics of Functional K.

Lesson 1.2: Basics of Functional K

The purpose of this lesson is to explain the basics of productions and
rules in K. These are two types of K sentences. A K file consists of
one or more requires or modules in K. Each module consists of one or
more imports or sentences. For more information on requires, modules, and
sentences, refer to Lesson 1.5. However, for the time
being, just think of a module as a container for sentences, and don't worry
about requires or imports just yet.

Our first K program

To start with, input the following program into your editor as file
lesson-02-a.k:

module LESSON-02-A

  syntax Color ::= Yellow() | Blue()
  syntax Fruit ::= Banana() | Blueberry()
  syntax Color ::= colorOf(Fruit) [function]

  rule colorOf(Banana()) => Yellow()
  rule colorOf(Blueberry()) => Blue()

endmodule

Save this file and then run:

kompile lesson-02-a.k

kompile is K's compiler. By default, it takes a program or specification
written in K and compiles it into an interpreter for that input. Right now we
are compiling a single file. A set of K files that are compiled together are
called a K definition. We will cover multiple file K definitions later on.
kompile will output a directory containing everything needed to execute
programs and perform proofs using that definition. In this case, kompile will
(by default) create the directory lesson-02-a-kompiled under the current
directory.

Now, save the following input file in your editor as banana.color in the same
directory as lesson-02-a.k:

colorOf(Banana())

We can now evaluate this K term by running (from the same directory):

krun banana.color

krun will use the interpreter generated by the first call to kompile to
execute this program.

You will get the following output:

<k>
  Yellow ( ) ~> .
</k>

For now, don't worry about the <k>, </k>, or ~> . portions of this
output file.

You can also execute small programs directly by specifying them on the command
line instead of putting them in a file. For example, the same program above
could also have been executed by running the following command:

krun -cPGM='colorOf(Banana())'

Now, let's look at what this definition and program did.

Productions, Constructors, and Functions

The first thing to realize is that this K definition contains 5 productions.
Productions are introduced with the syntax keyword, followed by a sort,
followed by the operator ::= followed by the definition of one or more
productions themselves, separated by the | operator. There are different
types of productions, but for now we only care about constructors and
functions. Each declaration separated by the | operator is individually
a single production, and the | symbol simply groups together productions that
have the same sort. For example, we could equally have written an identical K
definition lesson-02-b.k like so:

module LESSON-02-B

  syntax Color ::= Yellow()
  syntax Color ::= Blue()
  syntax Fruit ::= Banana()
  syntax Fruit ::= Blueberry()
  syntax Color ::= colorOf(Fruit) [function]

  rule colorOf(Banana()) => Yellow()
  rule colorOf(Blueberry()) => Blue()

endmodule

You can try compiling and running lesson-02-b.k to see that it produces the same output as lesson-02-a.k:

kompile lesson-02-b.k
krun -cPGM='colorOf(Banana())' --definition 'lesson-02-b-kompiled'

where the --definition attribute points to the directory containing a compiled version of LESSON-02-B.
Even the following definition is equivalent:

module LESSON-02-C

  syntax Color ::= Yellow()
                 | Blue()
                 | colorOf(Fruit) [function]
  syntax Fruit ::= Banana()
                 | Blueberry()

  rule colorOf(Banana()) => Yellow()
  rule colorOf(Blueberry()) => Blue()

endmodule

Each of these types of productions named above has the same underlying syntax,
but context and attributes are used to distinguish between the different
types. Tokens, brackets, lists, macros, aliases, and anywhere productions will
be covered in a later lesson, but this lesson does introduce us to constructors
and functions. Yellow(), Blue(), Banana(), and Blueberry() are
constructors. You can think of a constructor like a constructor for an
algebraic data type, if you're familiar with a functional language. The data
type itself is the sort that appears on the left of the ::= operator. Sorts
in K consist of uppercase identifiers.

Constructors can have arguments, but these ones do not. We will cover the
syntax of productions in detail in the next lesson, but for now, you can write
a production with no arguments as an uppercase or lowercase identifier followed
by the () operator.

A function is distinguished from a constructor by the presence of the
function attribute. Attributes appear in a comma separated list between
square brackets after any sentence, including both productions and rules.
Various attributes with built-in meanings exist in K and will be discussed
throughout the tutorial.

Exercise

Use krun to compute the return value of the colorOf function on a
Blueberry().

Rules, Matching, and Variables

Functions in K are given definitions using rules. A rule begins with the rule
keyword and contains at least one rewrite operator. The rewrite operator
is represented by the syntax =>. The rewrite operator is one of the built-in
productions in K, and we will discuss in more detail how it can be used in
future lessons, but for now, you can think of a rule as consisting of a
left-hand side and a right-hand side, separated by the rewrite
operator. On the left-hand side is the name of the function and zero or more
patterns corresponding to the parameters of the function. On the right-hand
side is another pattern. The meaning of the rule is relatively simple, having
defined these components. If the function is called with arguments that
match the patterns on the left-hand side, then the return value of the
function is the pattern on the right-hand side.

For example, in the above example, if the argument of the colorOf function
is Banana(), then the return value of the function is Yellow().

So far we have introduced that a constructor is a type of pattern in K. We
will introduce more complex patterns in later lessons, but there is one other
type of basic pattern: the variable. A variable, syntactically, consists
of an uppercase identifier. However, unlike a constructor, a variable will
match any pattern with one exception: Two variables with the same name
must match the same pattern.

Here is a more complex example (lesson-02-d.k):

module LESSON-02-D

  syntax Container ::= Jar(Fruit)
  syntax Fruit ::= Apple() | Pear()

  syntax Fruit ::= contentsOfJar(Container) [function]

  rule contentsOfJar(Jar(F)) => F

endmodule

Here we see that Jar is a constructor with a single argument. You can write a
production with multiple arguments by putting the sorts of the arguments in a
comma-separated list inside the parentheses.

In this example, F is a variable. It will match either Apple() or Pear().
The return value of the function is created by substituting the matched
values of all of the variables into the variables on the right-hand side of
the rule.

To demonstrate, compile this definition and execute the following program with
krun:

contentsOfJar(Jar(Apple()))

You will see when you run it that the program returns Apple(), because that
is the pattern that was matched by F.

Exercises

  1. Extend the definition in lesson-02-a.k with the addition of blackberries
    and kiwis. For simplicity, blackberries are black and kiwis are green. Then
    compile your definition and test that your additional fruits are correctly
    handled by the colorOf function.
  2. Create a new definition which defines an outfit as a multi-argument
    constructor consisting of a hat, shirt, pants, and shoes. Define a new sort,
    Boolean, with two constructors, true and false. Each of hat, shirt, pants,
    and shoes will have a single argument (a color), either black or
    white. Then define an outfitMatching function that will return true if all
    the pieces of the outfit are the same color. You do not need to define the
    case that returns false. Write some tests that your function behaves the way
    you expect.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.3: BNF Syntax and Parser Generation.

Lesson 1.3: BNF Syntax and Parser Generation

The purpose of this lesson is to explain the full syntax and semantics of
productions in K as well as how productions and other syntactic
sentences can be used to define grammars for use parsing both rules as well
as programs.

K's approach to parsing

K's grammar is divided into two components: the outer syntax of K and the
inner syntax of K. Outer syntax refers to the parsing of requires,
modules, imports, and sentences in a K definition. Inner syntax
refers to the parsing of rules and programs. Unlike the outer syntax of
K, which is predetermined, much of the inner syntax of K is defined by you, the
developer. When rules or programs are parsed, they are parsed within the
context of a module. Rules are parsed in the context of the module in which
they exist, whereas programs are parsed in the context of the
main syntax module of a K definition. The productions and other syntactic
sentences in a module are used to construct the grammar of the module, which
is then used to perform parsing.

Basic BNF productions

To illustrate how this works, we will consider a simple K definition which
defines a relatively basic calculator capable of evaluating Boolean expressions
containing and, or, not, and xor.

Input the following program into your editor as file lesson-03-a.k:

module LESSON-03-A

  syntax Boolean ::= "true" | "false"
                   | "!" Boolean [function]
                   | Boolean "&&" Boolean [function]
                   | Boolean "^" Boolean [function]
                   | Boolean "||" Boolean [function]

endmodule

You will notice that the productions in this file look a little different than
the ones from the previous lesson. In point of fact, K has two different
mechanisms for defining productions. We have previously been focused
exclusively on the first mechanism, where the ::= symbol is followed by an
alphanumeric identifier followed by a comma-separated list of sorts in
parentheses. However, this is merely a special case of a more generic mechanism
for defining the syntax of productions using a variant of
BNF Form.

For example, in the previous lesson, we had the following set of productions:

module LESSON-03-B
  syntax Color ::= Yellow() | Blue()
  syntax Fruit ::= Banana() | Blueberry()
  syntax Color ::= colorOf(Fruit) [function]
endmodule

It turns out that this is equivalent to the following definition which defines
the same grammar, but using BNF notation:

module LESSON-03-C
  syntax Color ::= "Yellow" "(" ")" | "Blue" "(" ")"
  syntax Fruit ::= "Banana" "(" ")" | "Blueberrry" "(" ")"
  syntax Color ::= "colorOf" "(" Fruit ")" [function]
endmodule

In this example, the sorts of the argument to the function are unchanged, but
everything else has been wrapped in double quotation marks. This is because
in BNF notation, we distinguish between two types of production items:
terminals and non-terminals. A terminal represents simply a literal
string of characters that is verbatim part of the syntax of that production.
A non-terminal, conversely, represents a sort name, where the syntax of that
production accepts any valid term of that sort at that position.

This is why, when we wrote the program colorOf(Banana()), krun was able to
execute that program: because it represented a term of sort Color that was
parsed and interpreted by K's interpreter. In other words, krun parses and
interprets terms according to the grammar defined by the developer. It is
automatically converted into an AST of that term, and then the colorOf
function is evaluated using the function rules provided in the definition.

You can ask yourself: How does K match the strings between the double quotes?
The answer is that K uses Flex to generate a scanner for the grammar. Flex looks
for the longest possible match of a regular expression in the input. If there
are ambiguities between 2 or more regular expressions, it will pick the one with
the highest prec attribute. You can learn more about how Flex matching works
here.

Bringing us back to the file lesson-03-a.k, we can see that this grammar
has given a simple BNF grammar for expressions over Booleans. We have defined
constructors corresponding to the Boolean values true and false, and functions
corresponding to the Boolean operators for and, or, not, and xor. We have also
given a syntax for each of these functions based on their syntax in the C
programming language. As such, we can now write programs in the simple language
we have defined.

Input the following program into your editor as and.bool in the same
directory:

true && false

We cannot interpret this program yet, because we have not given rules defining
the meaning of the && function yet, but we can parse it. To do this, you can
run (from the same directory):

kast --output kore and.bool

kast is K's just-in-time parser. It will generate a grammar from your K
definition on the fly and use it to parse the program passed on the command
line. The --output flag controls how the resulting AST is represented; don't
worry about the possible values yet, just use kore.

You ought to get the following AST printed on standard output, minus the
formatting:

inj{SortBoolean{}, SortKItem{}}(
  Lbl'UndsAnd-And-UndsUnds'LESSON-03-A'Unds'Boolean'Unds'Boolean'Unds'Boolean{}(
    Lbltrue'Unds'LESSON-03-A'Unds'Boolean{}(),
    Lblfalse'Unds'LESSON-03-A'Unds'Boolean{}()
  )
)

Don't worry about what exactly this means yet, just understand that it
represents the AST of the program that you just parsed. You ought to be able
to recognize the basic shape of it by seeing the words true, false, and
And in there. This is Kore, the intermediate representation of K, and we
will cover it in detail later.

Note that you can also tell kast to print the AST in other formats. For a
more direct representation of the original K, while still maintaining the
structure of an AST, you can say kast --output kast and.bool. This will
yield the following output:

`_&&__LESSON-03-A_Boolean_Boolean_Boolean`(
  `true_LESSON-03-A_Boolean`(.KList),
  `false_LESSON-03-A_Boolean`(.KList)
)

Note how the first output is largely a name-mangled version of the second
output. The one difference is the presence of the inj symbol in the KORE
output. We will talk more about this in later lessons.

Exercise

Parse the expression false || true with --output kast. See if you can
predict approximately what the corresponding output would be with
--output kore, then run the command yourself and compare it to your
prediction.

Ambiguities

Now let's try a slightly more advanced example. Input the following program
into your editor as and-or.bool:

true && false || false

When you try and parse this program, you ought to see the following error:

[Error] Inner Parser: Parsing ambiguity.
1: syntax Boolean ::= Boolean "||" Boolean [function]

`_||__LESSON-03-A_Boolean_Boolean_Boolean`(`_&&__LESSON-03-A_Boolean_Boolean_Boolean`(`true_LESSON-03-A_Boolean`(.KList),`false_LESSON-03-A_Boolean`(.KList)),`false_LESSON-03-A_Boolean`(.KList))
2: syntax Boolean ::= Boolean "&&" Boolean [function]

`_&&__LESSON-03-A_Boolean_Boolean_Boolean`(`true_LESSON-03-A_Boolean`(.KList),`_||__LESSON-03-A_Boolean_Boolean_Boolean`(`false_LESSON-03-A_Boolean`(.KList),`false_LESSON-03-A_Boolean`(.KList)))
        Source(./and-or.bool)
        Location(1,1,1,23)

This error is saying that kast was unable to parse this program because it is
ambiguous. K's just-in-time parser is a GLL parser, which means it can handle
the full generality of context-free grammars, including those grammars which
are ambiguous. An ambiguous grammar is one where the same string can be parsed
as multiple distinct ASTs. In this example, it can't decide whether it should
be parsed as (true && false) || false or as true && (false || false). As a
result, it reports the error to the user.

Brackets

Currently there is no way of resolving this ambiguity, making it impossible
to write complex expressions in this language. This is obviously a problem.
The standard solution in most programming languages to this problem is to
use parentheses to indicate the appropriate grouping. K generalizes this notion
into a type of production called a bracket. A bracket production in K
is any production with the bracket attribute. It is required that such a
production only have a single non-terminal, and the sort of the production
must equal the sort of that non-terminal. However, K does not otherwise
impose restrictions on the grammar the user provides for a bracket. With that
being said, the most common type of bracket is one in which a non-terminal
is surrounded by terminals representing some type of bracket such as
(), [], {}, <>, etc. For example, we can define the most common
type of bracket, the type used by the vast majority of programming languages,
quite simply.

Consider the following modified definition, which we will save to
lesson-03-d.k:

module LESSON-03-D

  syntax Boolean ::= "true" | "false"
                   | "(" Boolean ")" [bracket]
                   | "!" Boolean [function]
                   | Boolean "&&" Boolean [function]
                   | Boolean "^" Boolean [function]
                   | Boolean "||" Boolean [function]

endmodule

In this definition, if the user does not explicitly define parentheses, the
grammar remains ambiguous and K's just-in-time parser will report an error.
However, you are now able to parse more complex programs by means of explicitly
grouping subterms with the bracket we have just defined.

Consider and-or-left.bool:

(true && false) || false

Now consider and-or-right.bool:

true && (false || false)

If you parse these programs with kast, you will once again get a single
unique AST with no error. If you look, you might notice that the bracket itself
does not appear in the AST. In fact, this is a property unique to brackets:
productions with the bracket attribute are not represented in the parsed AST
of a term, and the child of the bracket is folded immediately into the parent
term. This is the reason for the requirement that a bracket production have
a single non-terminal of the same sort as the production itself.

Exercise

Write out what you expect the AST to be arising from parsing these two programs
above with --output kast, then parse them yourself and compare them to the
AST you expected. Confirm for yourself that the bracket production does not
appear in the AST.

Tokens

So far we have seen how we can define the grammar of a language. However,
the grammar is not the only relevant part of parsing a language. Also relevant
is the lexical syntax of the language. Thus far, we have implicitly been using
K's automatic lexer generation to generate a token in the scanner for each
terminal in our grammar. However, sometimes we wish to define more complex
lexical syntax. For example, consider the case of integers in C: an integer
consists of a decimal, octal, or hexadecimal number followed by an optional
suffix indicating the type of the literal.

In theory it would be possible to define this syntax via a grammar, but not
only would it be cumbersome and tedious, you would also then have to deal with
an AST generated for the literal which is not convenient to work with.

Instead of doing this, K allows you to define token productions, where
a production consists of a regular expression followed by the token
attribute, and the resulting AST consists of a typed string containing the
value recognized by the regular expression.

For example, the builtin integers in K are defined using the following
production:

syntax Int ::= r"[\\+-]?[0-9]+" [token]

Here we can see that we have defined that an integer is an optional sign
followed by a nonzero sequence of digits. The r preceding the terminal
indicates that what appears inside the double quotes is a regular expression,
and the token attribute indicates that terms which parse as this production
should be converted into a token by the parser.

It is also possible to define tokens that do not use regular expressions. This
can be useful when you wish to declare particular identifiers for use in your
semantics later. For example:

syntax Id ::= "main" [token]

Here, we declare that main is a token of sort Id. Instead of being parsed
as a symbol, it gets parsed as a token, generating a typed string in the AST.
This is useful in a semantics of C because the parser generally does not treat
the main function in C specially; only the semantics treats it specially.

Of course, languages can have more complex lexical syntax. For example, if we
wish to define the syntax of integers in C, we could use the following
production:

syntax IntConstant ::= r"(([1-9][0-9]*)|(0[0-7]*)|(0[xX][0-9a-fA-F]+))(([uU][lL]?)|([uU]((ll)|(LL)))|([lL][uU]?)|(((ll)|(LL))[uU]?))?" [token]

As you may have noted above, long and complex regular expressions
can be hard to read. They also suffer from the problem that unlike a grammar,
they are not particularly modular.

We can get around this restriction by declaring explicit regular expressions,
giving them a name, and then referring to them in productions.

Consider the following (equivalent) way to define the lexical syntax of
integers in C:

syntax IntConstant ::= r"({DecConstant}|{OctConstant}|{HexConstant})({IntSuffix}?)" [token]
syntax lexical DecConstant = r"{NonzeroDigit}({Digit}*)"
syntax lexical OctConstant = r"0({OctDigit}*)"
syntax lexical HexConstant = r"{HexPrefix}({HexDigit}+)"
syntax lexical HexPrefix = r"0x|0X"
syntax lexical NonzeroDigit = r"[1-9]"
syntax lexical Digit = r"[0-9]"
syntax lexical OctDigit = r"[0-7]"
syntax lexical HexDigit = r"[0-9a-fA-F]"
syntax lexical IntSuffix = r"{UnsignedSuffix}({LongSuffix}?)|{UnsignedSuffix}{LongLongSuffix}|{LongSuffix}({UnsignedSuffix}?)|{LongLongSuffix}({UnsignedSuffix}?)"
syntax lexical UnsignedSuffix = r"[uU]"
syntax lexical LongSuffix = r"[lL]"
syntax lexical LongLongSuffix = r"ll|LL"

As you can see, this is rather more verbose, but it has the benefit of both
being much easier to read and understand, and also increased modularity.
Note that we refer to a named regular expression by putting the name in curly
brackets. Note also that only the first sentence actually declares a new piece
of syntax in the language. When the user writes syntax lexical, they are only
declaring a regular expression. To declare an actual piece of syntax in the
grammar, you still must actually declare an explicit token production.

One final note: K uses Flex to implement
its lexical analysis. As a result, you can refer to the
Flex Manual
for a detailed description of the regular expression syntax supported. Note
that for performance reasons, Flex's regular expressions are actually a regular
language, and thus lack some of the syntactic convenience of modern
"regular expression" libraries. If you need features that are not part of the
syntax of Flex regular expressions, you are encouraged to express them via
a grammar instead.

Ahead-of-time parser generation

So far we have been entirely focused on K's support for just-in-time parsing,
where the parser is generated on the fly prior to being used. This benefits
from being faster to generate the parser, but it suffers in performance if you
have to repeatedly parse strings with the same parser. For this reason, it is
generally encouraged that when parsing programs, you use K's ahead-of-time
parser generation. K makes use of
GNU Bison to generate parsers.

By default, you can enable ahead-of-time parsing via the --gen-bison-parser
flag to kompile. This will make use of Bison's LR(1) parser generator. As
such, if your grammar is not LR(1), it may not parse exactly the same as if
you were to use the just-in-time parser, because Bison will automatically pick
one of the possible branches whenever it encounters a shift-reduce or
reduce-reduce conflict. In this case, you can either modify your grammar to be
LR(1), or you can enable use of Bison's GLR support by instead passing
--gen-glr-bison-parser to kompile. Note that if your grammar is ambiguous,
the ahead-of-time parser will not provide you with particularly readable error
messages at this time.

If you have a K definition named foo.k, and it generates a directory when
you run kompile called foo-kompiled, you can invoke the ahead-of-time
parser you generated by running foo-kompiled/parser_PGM <file> on a file.

Exercises

  1. Compile lesson-03-d.k with ahead-of-time parsing enabled. Then compare
    how long it takes to run kast --output kore and-or-left.bool with how long it
    takes to run lesson-03-d-kompiled/parser_PGM and-or-left.bool. Confirm for
    yourself that both produce the same result, but that the latter is faster.

  2. Define a simple grammar consisting of integers, brackets, addition,
    subtraction, multiplication, division, and unary negation. Integers should be
    in decimal form and lexically without a sign, whereas negative numbers can be
    represented via unary negation. Ensure that you are able to parse some basic
    arithmetic expressions using a generated ahead-of-time parser. Do not worry
    about disambiguating the grammar or about writing rules to implement the
    operations in this definition.

  3. Write a program where the meaning of the arithmetic expression based on
    the grammar you defined above is ambiguous, and then write programs that
    express each individual intended meaning using brackets.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.4: Disambiguating Parses.

Lesson 1.4: Disambiguating Parses

The purpose of this lesson is to teach how to use K's builtin features for
disambiguation to transform an ambiguous grammar into an unambiguous one that
expresses the intended ASTs.

Priority blocks

In practice, very few formal languages outside the domain of natural language
processing are ambiguous. The main reason for this is that parsing unambiguous
languages is asymptotically faster than parsing ambiguous languages.
Programming language designers instead usually use the notions of operator
precedence and associativity to make expression grammars unambiguous. These
mechanisms work by instructing the parser to reject certain ASTs in favor of
others in case of ambiguities; it is often possible to remove all ambiguities
in a grammar with these techniques.

While it is sometimes possible to explicitly rewrite the grammar to remove
these parses, because K's grammar specification and AST generation are
inextricably linked, this is generally discouraged. Instead, we use the
approach of explicitly expressing the relative precedence of different
operators in different situations in order to resolve the ambiguity.

For example, in C, && binds tighter in precedence than ||, meaning that
the expression true && false || false has only one valid AST:
(true && false) || false.

Consider, then, the third iteration on the grammar of this definition
(lesson-04-a.k):

module LESSON-04-A

  syntax Boolean ::= "true" | "false"
                   | "(" Boolean ")" [bracket]
                   > "!" Boolean [function]
                   > Boolean "&&" Boolean [function]
                   > Boolean "^" Boolean [function]
                   > Boolean "||" Boolean [function]

endmodule

In this example, some of the | symbols separating productions in a single
block have been replaced with >. This serves to describe the
priority groups associated with this block of productions.
The first priority group consists of the atoms of the
language: true, false, and the bracket operator. In general, a priority
group starts either at the ::= or > operator and extends until either the
next > operator or the end of the production block. Thus, we can see that the
second, third, fourth, and fifth priority groups in this grammar all consist
of a single production.

The meaning of these priority groups becomes apparent when parsing programs:
A symbol with a lesser priority, (i.e., one that binds looser), cannot
appear as the direct child of a symbol with a greater priority (i.e.,
one that binds tighter. In this case, the > operator can be seen as a
greater-than operator describing a transitive partial ordering on the
productions in the production block, expressing their relative priority.

To see this more concretely, let's look again at the program
true && false || false. As noted before, previously this program was
ambiguous because the parser could either choose that && was the child of ||
or vice versa. However, because a symbol with lesser priority (i.e., ||)
cannot appear as the direct child of a symbol with greater priority
(i.e., &&), the parser will reject the parse where || is under the
&& operator. As a result, we are left with the unambiguous parse
(true && false) || false. Similarly, true || false && false parses
unambiguously as true || (false && false). Conversely, if the user explicitly
wants the other parse, they can express this using brackets by explicitly
writing true && (false || false). This still parses successfully because the
|| operator is no longer the direct child of the && operator, but is
instead the direct child of the () operator, and the && operator is an
indirect parent, which is not subject to the priority restriction.

Astute readers, however, will already have noticed what seems to be a
contradiction: we have defined () as also having greater priority than ||.
One would think that this should mean that || cannot appear as a direct
child of (). This is a problem because priority groups are applied to every
possible parse separately. That is to say, even if the term is unambiguous
prior to this disambiguation rule, we still reject that parse if it violates
the rule of priority.

In fact, however, we do not reject this program as a parse error. Why is that?
Well, the rule for priority is slightly more complex than previously described.
In actual fact, it applies only conditionally. Specifically, it applies in
cases where the child is either the first or last production item in the
parent's production. For example, in the production Bool "&&" Bool, the
first Bool non-terminal is not preceded by any terminals, and the last Bool
non-terminal is not followed by any terminals. As a result of this, we apply
the priority rule to both children of &&. However, in the () operator,
the sole non-terminal is both preceded by and followed by terminals. As a
result, the priority rule is not applied when () is the parent. Because of
this, the program we mentioned above successfully parses.

Exercise

Parse the program true && false || false using kast, and confirm that the AST
places || as the top level symbol. Then modify the definition so that you
will get the alternative parse.

Associativity

Even having broken the expression grammar into priority blocks, the resulting
grammar is still ambiguous. We can see this if we try to parse the following
program (assoc.bool):

true && false && false

Priority blocks will not help us here: the problem comes between two parses
where both possible parses have a direct parent and child which is within a
single priority block (in this case, && is in the same block as itself).

This is where the notion of associativity comes into play. Associativity
applies the following additional rules to parses:

  • a left-associative symbol cannot appear as a direct rightmost child of a
    symbol with equal priority;
  • a right-associative symbol cannot appear as a direct leftmost child of a
    symbol with equal priority; and
  • a non-associative symbol cannot appear as a direct leftmost or rightmost
    child of a symbol with equal priority.

In C, binary operators are all left-associative, meaning that the expression
true && false && false parses unambiguously as (true && false) && false,
because && cannot appear as the rightmost child of itself.

Consider, then, the fourth iteration on the grammar of this definition
(lesson-04-b.k):

module LESSON-04-B

  syntax Boolean ::= "true" | "false"
                   | "(" Boolean ")" [bracket]
                   > "!" Boolean [function]
                   > left: Boolean "&&" Boolean [function]
                   > left: Boolean "^" Boolean [function]
                   > left: Boolean "||" Boolean [function]

endmodule

Here each priority group, immediately after the ::= or > operator, can
be followed by a symbol representing the associativity of that priority group:
either left: for left associativity, right: for right associativity, or
non-assoc: for non-associativity. In this example, each priority group we
apply associativity to has only a single production, but we could equally well
write a priority block with multiple productions and an associativity.

For example, consider the following, different grammar (lesson-04-c.k):

module LESSON-04-C

  syntax Boolean ::= "true" | "false"
                   | "(" Boolean ")" [bracket]
                   > "!" Boolean [function]
                   > left:
                     Boolean "&&" Boolean [function]
                   | Boolean "^" Boolean [function]
                   | Boolean "||" Boolean [function]

endmodule

In this example, unlike the one above, &&, ^, and || have the same
priority. However, viewed as a group, the entire group is left associative.
This means that none of &&, ^, and || can appear as the right child of
any of &&, ^, or ||. As a result of this, this grammar is also not
ambiguous. However, it expresses a different grammar, and you are encouraged
to think about what the differences are in practice.

Exercise

Parse the program true && false && false yourself, and confirm that the AST
places the rightmost && at the top of the expression. Then modify the
definition to generate the alternative parse.

Explicit priority and associativity declarations

Previously we have only considered the case where all of the productions
which you wish to express a priority or associativity relation over are
co-located in the same block of productions. However, in practice this is not
always feasible or desirable, especially as a definition grows in size across
multiple modules.

As a result of this, K provides a second way of declaring priority and
associativity relations.

Consider the following grammar, which we will name lesson-04-d.k and which
will express the exact same grammar as lesson-04-b.k

module LESSON-04-D

  syntax Boolean ::= "true" [group(literal)] | "false" [group(literal)]
                   | "(" Boolean ")" [group(atom), bracket]
                   | "!" Boolean [group(not), function]
                   | Boolean "&&" Boolean [group(and), function]
                   | Boolean "^" Boolean [group(xor), function]
                   | Boolean "||" Boolean [group(or), function]

  syntax priority literal atom > not > and > xor > or
  syntax left and
  syntax left xor
  syntax left or
endmodule

This introduces a couple of new features of K. First, the group(_) attribute
is used to conceptually group together sets of sentences under a common
user-defined name. For example, literal in the syntax priority sentence is
used to refer to all the productions marked with the group(literal) attribute,
i.e., true and false. A production can belong to multiple groups using
syntax such as group(myGrp1,myGrp2).

Once we understand this, it becomes relatively straightforward to understand
the meaning of this grammar. Each syntax priority sentence defines a
priority relation where > separates different priority groups. Each priority
group is defined by a list of one or more group names, and consists of all
productions which are members of at least one of those named groups.

In the same way, a syntax left, syntax right, or syntax non-assoc sentence
defines an associativity relation among left-, right-, or non-associative
groups. Specifically, this means that:

syntax left a b

is different to:

syntax left a
syntax left b

As a consequence of this, syntax [left|right|non-assoc] should not be used to
group together labels with different priority.

Prefer/avoid

Sometimes priority and associativity prove insufficient to disambiguate a
grammar. In particular, sometimes it is desirable to be able to choose between
two ambiguous parses directly while still not rejecting any parses if the term
parsed is unambiguous. A good example of this is the famous "dangling else"
problem in imperative C-like languages.

Consider the following definition (lesson-04-E.k):

module LESSON-04-E

  syntax Exp ::= "true" | "false"
  syntax Stmt ::= "if" "(" Exp ")" Stmt
                | "if" "(" Exp ")" Stmt "else" Stmt
                | "{" "}"
endmodule

We can write the following program (dangling-else.if):

if (true) if (false) {} else {}

This is ambiguous because it is unclear whether the else clause is part of
the outer if or the inner if. At first we might try to resolve this with
priorities, saying that the if without an else cannot appear as a child of
the if with an else. However, because the non-terminal in the parent symbol
is both preceded and followed by a terminal, this will not work.

Instead, we can resolve the ambiguity directly by telling the parser to
"prefer" or "avoid" certain productions when ambiguities arise. For example,
when we parse this program, we see the following ambiguity as an error message:

[Error] Inner Parser: Parsing ambiguity.
1: syntax Stmt ::= "if" "(" Exp ")" Stmt

`if(_)__LESSON-04-E_Stmt_Exp_Stmt`(`true_LESSON-04-E_Exp`(.KList),`if(_)_else__LESSON-04-E_Stmt_Exp_Stmt_Stmt`(`false_LESSON-04-E_Exp`(.KList),`;_LESSON-04-E_Stmt`(.KList),`;_LESSON-04-E_Stmt`(.KList)))
2: syntax Stmt ::= "if" "(" Exp ")" Stmt "else" Stmt

`if(_)_else__LESSON-04-E_Stmt_Exp_Stmt_Stmt`(`true_LESSON-04-E_Exp`(.KList),`if(_)__LESSON-04-E_Stmt_Exp_Stmt`(`false_LESSON-04-E_Exp`(.KList),`;_LESSON-04-E_Stmt`(.KList)),`;_LESSON-04-E_Stmt`(.KList))
        Source(./dangling-else.if)
        Location(1,1,1,30)

Roughly, we see that the ambiguity is between an if with an else or an if
without an else. Since we want to pick the first parse, we can tell K to
"avoid" the second parse with the avoid attribute. Consider the following
modified definition (lesson-04-f.k):

module LESSON-04-F

  syntax Exp ::= "true" | "false"
  syntax Stmt ::= "if" "(" Exp ")" Stmt
                | "if" "(" Exp ")" Stmt "else" Stmt [avoid]
                | "{" "}"
endmodule

Here we have added the avoid attribute to the else production. As a result,
when an ambiguity occurs and one or more of the possible parses has that symbol
at the top of the ambiguous part of the parse, we remove those parses from
consideration and consider only those remaining. The prefer attribute behaves
similarly, but instead removes all parses which do not have that attribute.
In both cases, no action is taken if the parse is not ambiguous.

Exercises

  1. Parse the program if (true) if (false) {} else {} using lesson-04-f.k
    and confirm that else clause is part of the innermost if statement. Then
    modify the definition so that you will get the alternative parse.

  2. Modify your solution from Lesson 1.3, Exercise 2 so that unary negation should
    bind tighter than multiplication and division, which should bind tighter than
    addition and subtraction, and each binary operator should be left associative.
    Write these priority and associativity declarations explicitly, and then
    try to write them inline.

  3. Write a simple grammar containing at least one ambiguity that cannot be
    resolved via priority or associativity, and then use the prefer attribute to
    resolve that ambiguity.

  4. Explain why the following grammar is not labeled ambiguous by the K parser when parsing abb, then make the parser realize the ambiguity.

module EXERCISE4

syntax Expr ::= "a" Expr "b"
              | "abb"
              | "b"

endmodule

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.5: Modules, Imports, and Requires.

Lesson 1.5: Modules, Imports, and Requires

The purpose of this lesson is to explain how K definitions can be broken into
separate modules and files and how these distinct components combine into a
complete K definition.

K's outer syntax

Recall from Lesson 1.3 that K's grammar is broken
into two components: the outer syntax of K and the inner syntax of K.
Outer syntax, as previously mentioned, consists of requires, modules,
imports, and sentences. A K semantics is expressed by the set of
sentences contained in the definition. The scope of what is considered
contained in that definition is determined both by the main semantics
module
of a K definition, as well as the requires and imports present
in the file that contains that module.

Basic module syntax

The basic unit of grouping sentences in K is the module. A module consists
of a module name, an optional list of attributes, a list of
imports, and a list of sentences.

A module name consists of one or more groups of letters, numbers, or
underscores, separated by a hyphen. Here are some valid module names: FOO,
FOO-BAR, foo0, foo0_bar-Baz9. Here are some invalid module names: -,
-FOO, BAR-, FOO--BAR. Stylistically, modules names are usually all
uppercase with hyphens separating words, but this is not strictly enforced.

Some example modules include an empty module:

module LESSON-05-A

endmodule

A module with some attributes:

module LESSON-05-B [group(attr1,attr2), private]

endmodule

A module with some sentences:

module LESSON-05-C
  syntax Boolean ::= "true" | "false"
  syntax Boolean ::= "not" Boolean [function]
  rule not true => false
  rule not false => true
endmodule

Imports

Thus far we have only discussed definitions containing a single module.
Definitions can also contain multiple modules, in which one module imports
others.

An import in K appears at the top of a module, prior to any sentences. It can
be specified with the imports keyword, followed by a module name.

For example, here is a simple definition with two modules (lesson-05-d.k):

module LESSON-05-D-1
  syntax Boolean ::= "true" | "false"
  syntax Boolean ::= "not" Boolean [function]
endmodule

module LESSON-05-D
  imports LESSON-05-D-1

  rule not true => false
  rule not false => true
endmodule

This K definition is equivalent to the definition expressed by the single module
LESSON-05-C. Essentially, by importing a module, we include all of the
sentences in the module being imported into the module that we import from.
There are a few minor differences between importing a module and simply
including its sentences in another module directly, but we will cover these
differences later. Essentially, you can think of modules as a way of
conceptually grouping sentences in a larger K definition.

Exercise

Modify lesson-05-d.k to include four modules: one containing the syntax, two
with one rule each that imports the first module, and a final module
LESSON-05-D containing no sentences that imports the second and third module.
Check to make sure the definition still compiles and that you can still evaluate
the not function.

Parsing in the presence of multiple modules

As you may have noticed, each module in a definition can express a distinct set
of syntax. When parsing the sentences in a module, we use the syntax
of that module, enriched with the basic syntax of K, in order to parse
rules in that module. For example, the following definition is a parser error
(lesson-05-e.k):

module LESSON-05-E-1
  rule not true => false
  rule not false => true
endmodule

module LESSON-05-E-2
  syntax Boolean ::= "true" | "false"
  syntax Boolean ::= "not" Boolean [function]
endmodule

This is because the syntax referenced in module LESSON-05-E-1, namely, not,
true, and false, is not imported by that module. You can solve this problem
by simply importing the modules containing the syntax you want to use in your
sentences.

Main syntax and semantics modules

When we are compiling a K definition, we need to know where to start. We
designate two specific entry point modules: the main syntax module
and the main semantics module. The main syntax module, as well as all the
modules it imports recursively, are used to create the parser for programs that
you use to parse programs that you execute with krun. The main semantics
module, as well as all the modules it imports recursively, are used to
determine the rules that can be applied at runtime in order to execute a
program. For example, in the above example, if the main semantics module is
module LESSON-05-D-1, then not is an uninterpreted function (i.e., has no
rules associated with it), and the rules in module LESSON-05-D are not
included.

While you can specify the entry point modules explicitly by passing the
--main-module and --syntax-module flags to kompile, by default, if you
type kompile foo.k, then the main semantics module will be FOO and the
main syntax module will be FOO-SYNTAX.

Splitting a definition into multiple files

So far, while we have discussed ways to break definitions into separate
conceptual components (modules), K also provides a mechanism for combining
multiple files into a single K definition, namely, the requires directive.

In K, the requires keyword has two meanings. The first, the requires
statement, appears at the top of a K file, prior to any module declarations. It
consists of the keyword requires followed by a double-quoted string. The
second meaning of the requires keyword will be covered in a later lesson,
but it is distinguished because the second case occurs only inside modules.

The string passed to the requires statement contains a filename. When you run
kompile on a file, it will look at all of the requires statements in that
file, look up those files on disk, parse them, and then recursively process all
the requires statements in those files. It then combines all the modules in all
of those files together, and uses them collectively as the set of modules to
which imports statements can refer.

Putting it all together

Putting it all together, here is one possible way in which we could break the
definition lesson-02-c.k from Lesson 1.2 into
multiple files and modules:

colors.k:

module COLORS
  syntax Color ::= Yellow()
                 | Blue()
endmodule

fruits.k:

module FRUITS
  syntax Fruit ::= Banana()
                 | Blueberry()
endmodule

colorOf.k:

requires "fruits.k"
requires "colors.k"

module COLOROF-SYNTAX
  imports COLORS
  imports FRUITS

  syntax Color ::= colorOf(Fruit) [function]
endmodule

module COLOROF
  imports COLOROF-SYNTAX

  rule colorOf(Banana()) => Yellow()
  rule colorOf(Blueberry()) => Blue()
endmodule

You would then compile this definition with kompile colorOf.k and use it the
same way as the original, single-module definition.

Exercise

Modify the name of the COLOROF module, and then recompile the definition.
Try to understand why you now get a compiler error. Then, resolve this compiler
error by passing the --main-module and --syntax-module flags to kompile.

Include path

One note can be made about how paths are resolved in requires statements.

By default, the path you specify is allowed to be an absolute or a relative
path. If the path is absolute, that exact file is imported. If the path is
relative, a matching file is looked for within all of the
include directories specified to the compiler. By default, the include
directories include the current working directory, followed by the
include/kframework/builtin directory within your installation of K. You can
also pass one or more directories to kompile via the -I command line flag,
in which case these directories are prepended to the beginning of the list.

Exercises

  1. Take the solution to Lesson 1.4, Exercise 2 which included the explicit
    priority and associativity declarations, and modify the definition so that
    the syntax of integers and brackets is in one module, the syntax of addition,
    subtraction, and unary negation is in another module, and the syntax of
    multiplication and division is in a third module. Make sure you can still parse
    the same set of expressions as before. Place priority declarations in the main
    module.

  2. Modify lesson-02-d.k from Lesson 1.2 so that the rules and syntax are in
    separate modules in separate files.

  3. Place the file containing the syntax from Exercise 2 in another directory,
    then recompile the definition. Observe why a compilation error occurs. Then
    fix the compiler error by passing -I to kompile.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.6: Integers and Booleans.

Lesson 1.6: Integers and Booleans

The purpose of this lesson is to explain the two most basic types of builtin
sorts in K, the Int sort and the Bool sort, representing
arbitrary-precision integers and Boolean algebra.

Builtin sorts in K

K provides definitions of some useful sorts in
domains.md, found in the
include/kframework/builtin directory of the K installation. This file is
defined via a
Literate programming
style that we will discuss in a future lesson. We will not cover all of the
sorts found there immediately, however, this lesson discusses some of the
details surrounding integers and Booleans, as well as providing information
about how to look up more detailed knowledge about builtin functions in K's
documentation.

Booleans in K

The most basic builtin sort K provides is the Bool sort, representing
Boolean values (i.e., true and false). You have already seen how we were
able to create this type ourselves using K's parsing and disambiguation
features. However, in the vast majority of cases, we prefer instead to import
the version of Boolean algebra defined by K itself. Most simply, you can do
this by importing the module BOOL in your definition. For example
(lesson-06-a.k):

module LESSON-06-A
  imports BOOL

  syntax Fruit ::= Blueberry() | Banana()
  syntax Bool ::= isBlue(Fruit) [function]

  rule isBlue(Blueberry()) => true
  rule isBlue(Banana()) => false
endmodule

Here we have defined a simple predicate, i.e., a function returning a
Boolean value. We are now able to perform the usual Boolean operations of
and, or, and not over these values. For example (lesson-06-b.k):"

module LESSON-06-B
  imports BOOL

  syntax Fruit ::= Blueberry() | Banana()
  syntax Bool ::= isBlue(Fruit) [function]

  rule isBlue(Blueberry()) => true
  rule isBlue(Banana()) => false

  syntax Bool ::= isYellow(Fruit) [function]
                | isBlueOrYellow(Fruit) [function]

  rule isYellow(Banana()) => true
  rule isYellow(Blueberry()) => false

  rule isBlueOrYellow(F) => isBlue(F) orBool isYellow(F)
endmodule

In the above example, Boolean inclusive or is performed via the orBool
function, which is defined in the BOOL module. As a matter of convention,
many functions over builtin sorts in K are suffixed with the name of the
primary sort over which those functions are defined. This happens so that the
syntax of K does not (generally) conflict with the syntax of any other
programming language, which would make it harder to define that programming
language in K.

Exercise

Write a function isBlueAndNotYellow which computes the appropriate Boolean
expression. If you are unsure what the appropriate syntax is to use, you
can refer to the BOOL module in
domains.md. Add a term of
sort Fruit for which isBlue and isYellow both return true, and test that
the isBlueAndNotYellow function behaves as expected on all three Fruits.

Syntax Modules

For most sorts in domains.md, K defines more than one module that can be
imported by users. For example, for the Bool sort, K defines the BOOL
module that has previously already been discussed, but also provides the
BOOL-SYNTAX module. This module, unlike the BOOL module, only declares the
values true and false, but not any of the functions that operate over the
Bool sort. The rationale is that you may want to import this module into the
main syntax module of your definition in some cases, whereas you generally do
not want to do this with the version of the module that includes all the
functions over the Bool sort. For example, if you were defining the semantics
of C++, you might import BOOL-SYNTAX into the syntax module of your
definition, because true and false are part of the grammar of C++, but
you would only import the BOOL module into the main semantics module, because
C++ defines its own syntax for and, or, and not that is different from the
syntax defined in the BOOL module.

Here, for example, is how we might redefine our Boolean expression calculator
to use the Bool sort while maintaining an idiomatic structure of modules
and imports, for the first time including the rules to calculate the values of
expressions themselves (lesson-06-c.k):

module LESSON-06-C-SYNTAX
  imports BOOL-SYNTAX

  syntax Bool ::= "(" Bool ")" [bracket]
                > "!" Bool [function]
                > left:
                  Bool "&&" Bool [function]
                | Bool "^" Bool [function]
                | Bool "||" Bool [function]
endmodule

module LESSON-06-C
  imports LESSON-06-C-SYNTAX
  imports BOOL

  rule ! B => notBool B
  rule A && B => A andBool B
  rule A ^ B => A xorBool B
  rule A || B => A orBool B
endmodule

Note the encapsulation of syntax: the LESSON-06-C-SYNTAX module contains
exactly the syntax of our Boolean expressions, and no more, whereas any other
syntax needed to implement those functions is in the LESSON-06-C module
instead.

Exercise

Add an "implies" function to the above Boolean expression calculator, using the
-> symbol to represent implication. You can look up K's builtin "implies"
function in the BOOL module in domains.md.

Integers in K

Unlike most programming languages, where the most basic integer type is a
fixed-precision integer type, the most commonly used integer sort in K is
the Int sort, which represents the mathematical integers, ie,
arbitrary-precision integers.

K provides three main modules for import when using the Int sort. The first,
containing all the syntax of integers as well as all of the functions over
integers, is the INT module. The second, which provides just the syntax
of integer literals themselves, is the INT-SYNTAX module. However, unlike
most builtin sorts in K, K also provides a third module for the Int sort:
the UNSIGNED-INT-SYNTAX module. This module provides only the syntax of
non-negative integers, i.e., natural numbers. The reasons for this involve
lexical ambiguity. Generally speaking, in most programming languages, -1 is
not a literal, but instead a literal to which the unary negation operator is
applied. K thus provides this module to ease in specifying the syntax of such
languages.

For detailed information about the functions available over the Int sort,
refer to domains.md. Note again how we append Int to the end of most of the
integer operations to ensure they do not collide with the syntax of other
programming languages.

Exercises

  1. Extend your solution from Lesson 1.4, Exercise 2 to implement the rules
    that define the behavior of addition, subtraction, multiplication, and
    division. Do not worry about the case when the user tries to divide by zero
    at this time. Use /Int to implement division. Test your new calculator
    implementation by executing the arithmetic expressions you wrote as part of
    Lesson 1.3, Exercise 2. Check to make sure each computes the value you expected.

  2. Combine the Boolean expression calculator from this lesson with your
    solution to Exercise 1, and then extend the combined calculator with the <,
    <=, >, >=, ==, and != expressions. Write some Boolean expressions
    that combine integer and Boolean operations, and test to ensure that these
    expressions return the expected truth value.

  3. Compute the following expressions using your solution from Exercise 2:
    7 / 3, 7 / -3, -7 / 3, -7 / -3. Then replace the /Int function in
    your definition with divInt instead, and observe how the value of the above
    expressions changes. Why does this occur?

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.7: Side Conditions and Rule Priority.

Lesson 1.7: Side Conditions and Rule Priority

The purpose of this lesson is to explain how to write conditional rules in K,
and to explain how to control the order in which rules are tried.

Side Conditions

So far, all of the rules we have discussed have been unconditional rules.
If the left-hand side of the rule matches the arguments to the function, the
rule applies. However, there is another type of rule, a conditional rule.
A conditional rule consists of a rule body containing the patterns to
match, and a side condition representing a Boolean expression that must
evaluate to true in order for the rule to apply.

Side conditions in K are introduced via the requires keyword immediately
following the rule body. For example, here is a rule with a side condition
(lesson-07-a.k):

module LESSON-07-A
  imports BOOL
  imports INT

  syntax Grade ::= "letter-A"
                 | "letter-B"
                 | "letter-C"
                 | "letter-D"
                 | "letter-F"
                 | gradeFromPercentile(Int) [function]

  rule gradeFromPercentile(I) => letter-A requires I >=Int 90
endmodule

In this case, the gradeFromPercentile function takes a single integer
argument. The function evaluates to letter-A if the argument passed is
greater than 90. Note that the side condition is allowed to refer to variables
that appear on the left-hand side of the rule. In the same manner as variables
appearing on the right-hand side, variables that appear in the side condition
evaluate to the value that was matched on the left-hand side. Then the
functions in the side condition are evaluated, which returns a term of sort
Bool. If the term is equal to true, then the rule applies. Bear in mind
that the side condition is only evaluated at all if the patterns on the
left-hand side of the rule match the term being evaluated.

Exercise

Write a rule that evaluates gradeFromPercentile to letter-B if the argument
to the function is in the range [80,90). Test that the function correctly
evaluates various numbers between 80 and 100.

owise Rules

So far, all the rules we have introduced have had the same priority. What
this means is that K does not necessarily enforce an order in which the rules
are tried. We have only discussed functions so far in K, so it is not
immediately clear why this choice was made, given that a function is not
considered well-defined if multiple rules for evaluating it are capable of
evaluating the same arguments to different results. However, in future lessons
we will discuss other types of rules in K, some of which can be
non-deterministic. What this means is that if more than one rule is capable
of matching, then K will explore both possible rules in parallel, and consider
each of their respective results when executing your program. Don't worry too
much about this right now, but just understand that because of the potential
later for nondeterminism, we don't enforce a total ordering on the order in
which rules are attempted to be applied.

However, sometimes this is not practical; It can be very convenient to express
that a particular rule applies if no other rules for that function are
applicable. This can be expressed by adding the owise attribute to a rule.
What this means, in practice, is that this rule has lower priority than other
rules, and will only be tried to be applied after all the other,
higher-priority rules have been tried and they have failed.

For example, in the above exercise, we had to add a side condition containing
two Boolean comparisons to the rule we wrote to handle letter-B grades.
However, in practice this meant that we compare the percentile to 90 twice. We
can more efficiently and more idiomatically write the letter-B case for the
gradeFromPercentile rule using the owise attribute (lesson-07-b.k):

module LESSON-07-B
  imports BOOL
  imports INT

  syntax Grade ::= "letter-A"
                 | "letter-B"
                 | "letter-C"
                 | "letter-D"
                 | "letter-F"
                 | gradeFromPercentile(Int) [function]

  rule gradeFromPercentile(I) => letter-A requires I >=Int 90
  rule gradeFromPercentile(I) => letter-B requires I >=Int 80 [owise]
endmodule

This rule is saying, "if all the other rules do not apply, then the grade is a
B if the percentile is greater than or equal to 80." Note here that we use both
a side condition and an owise attribute on the same rule. This is not
required (as we will see later), but it is allowed. What this means is that the
side condition is only tried if the other rules did not apply and the
left-hand side of the rule matched. You can even use more complex matching on
the left-hand side than simply a variable. More generally, you can also have
multiple higher-priority rules, or multiple owise rules. What this means in
practice is that all of the non-owise rules are tried first, in any order,
followed by all the owise rules, in any order.

Exercise

The grades D and F correspond to the percentile ranges [60, 70) and [0, 60)
respectively. Write another implementation of gradeFromPercentile which
handles only these cases, and uses the owise attribute to avoid redundant
Boolean comparisons. Test that various percentiles in the range [0, 70) are
evaluated correctly.

Rule Priority

As it happens, the owise attribute is a specific case of a more general
concept we call rule priority. In essence, each rule is assigned an integer
priority. Rules are tried in increasing order of priority, starting with a
rule with priority zero, and trying each increasing numerical value
successively.

By default, a rule is assigned a priority of 50. If the rule has the owise
attribute, it is instead given the priority 200. You can see why this will
cause owise rules to be tried after regular rules.

However, it is also possible to directly assign a numerical priority to a rule
via the priority attribute. For example, here is an alternative way
we could express the same two rules in the gradeFromPercentile function
(lesson-07-c.k):

module LESSON-07-C
  imports BOOL
  imports INT

  syntax Grade ::= "letter-A"
                 | "letter-B"
                 | "letter-C"
                 | "letter-D"
                 | "letter-F"
                 | gradeFromPercentile(Int) [function]

  rule gradeFromPercentile(I) => letter-A requires I >=Int 90 [priority(50)]
  rule gradeFromPercentile(I) => letter-B requires I >=Int 80 [priority(200)]
endmodule

We can, of course, assign a priority equal to any non-negative integer. For
example, here is a more complex example that handles the remaining grades
(lesson-07-d.k):

module LESSON-07-D
  imports BOOL
  imports INT

  syntax Grade ::= "letter-A"
                 | "letter-B"
                 | "letter-C"
                 | "letter-D"
                 | "letter-F"
                 | gradeFromPercentile(Int) [function]

  rule gradeFromPercentile(I) => letter-A requires I >=Int 90 [priority(50)]
  rule gradeFromPercentile(I) => letter-B requires I >=Int 80 [priority(51)]
  rule gradeFromPercentile(I) => letter-C requires I >=Int 70 [priority(52)]
  rule gradeFromPercentile(I) => letter-D requires I >=Int 60 [priority(53)]
  rule gradeFromPercentile(_) => letter-F                     [priority(54)]
endmodule

Note that we have introduced a new piece of syntax here: _. This is actually
just a variable. However, as a special case, when a variable is named _, it
does not bind a value that can be used on the right-hand side of the rule, or
in a side condition. Effectively, _ is a placeholder variable that means "I
don't care about this term."

In this example, we have explicitly expressed the order in which the rules of
this function are tried. Since rules are tried in increasing numerical
priority, we first try the rule with priority 50, then 51, then 52, 53, and
finally 54.

As a final note, remember that if you assign a rule a priority higher than 200,
it will be tried after a rule with the owise attribute, and if you assign
a rule a priority less than 50, it will be tried before a rule with no
explicit priority.

Exercises

  1. Write a function isEven that returns whether an integer is an even number.
    Use two rules and one side condition. The right-hand side of the rules should
    be Boolean literals. Refer back to
    domains.md for the relevant
    integer operations.

  2. Modify the calculator application from Lesson 1.6, Exercise 2, so that division
    by zero will no longer make krun crash with a "Divison by zero" exception.
    Instead, the / function should not match any of its rules if the denominator
    is zero.

  3. Write your own implementation of ==, <, <=, >, >= for integers and modify your solution from Exercise 2 to use it.
    You can use any arithmetic operations in the INT module, but do not use any built-in boolean functions for comparing integers.

    Hint: Use pattern matching and recursive definitions with rule priorities.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.8: Literate Programming with Markdown.

Lesson 1.8: Literate Programming with Markdown

The purpose of this lesson is to teach a paradigm for performing literate
programming in K, and explain how this can be used to create K definitions
that are also documentation.

Markdown and K

The K tutorial so far has been written in
Markdown. Markdown,
for those not already familiar, is a lightweight plain-text format for styling
text. From this point onward, we assume you are familiar with Markdown and how
to write Markdown code. You can refer to the above link for a tutorial if you
are not already familiar.

What you may not necessarily realize, however, is that the K tutorial is also
a sequence of K definitions written in the manner of
Literate Programming.
For detailed information about Literate Programming, you can read the linked
Wikipedia article, but the short summary is that literate programming is a way
of intertwining documentation and code together in a manner that allows
executable code to also be, simultaneously, a documented description of that
code.

K is provided with built-in support for literate programming using Markdown.
By default, if you pass a file with the .md file extension to kompile, it
will look for any code blocks containing k code in that file, extract out
that K code into pure K, and then compile it as if it were a .k file.

A K code block begins with a line of text containing the keyword ```k,
and ends when it encounters another ``` keyword.

For example, if you view the markdown source of this document, this is a K
code block:

module LESSON-08
  imports INT

Only the code inside K code blocks will actually be sent to the compiler. The
rest, while it may appear in the document when rendered by a markdown viewer,
is essentially a form of code comment.

When you have multiple K code blocks in a document, K will append each one
together into a single file before passing it off to the outer parser.

For example, the following code block contains sentences that are part of the
LESSON-08 module that we declared the beginning of above:

  syntax Int ::= Int "+" Int [function]
  rule I1 + I2 => I1 +Int I2

Exercise

Compile this file with kompile README.md --main-module LESSON-08. Confirm
that you can use the resulting compiled definition to evaluate the +
function.

Markdown Selectors

On occasion, you may want to generate multiple K definitions from a single
Markdown file. You may also wish to include a block of syntax-highlighted K
code that nonetheless does not appear as part of your K definition. It is
possible to accomplish this by means of the built-in support for syntax
highlighting in Markdown. Markdown allows a code block that was begun with
``` to be immediately followed by a string which is used to signify what
programming language the following code is written in. However, this feature
actually allows arbitrary text to appear describing that code block. Markdown
parsers are able to parse this text and render the code block differently
depending on what text appears after the backticks.

In K, you can use this functionality to specify one or more
Markdown selectors which are used to describe the code block. A Markdown
selector consists of a sequence of characters containing letters, numbers, and
underscores. A code block can be designated with a single selector by appending
the selector immediately following the backticks that open the code block.

For example, here is a code block with the foo selector:

foo bar

Note that this is not K code. By convention, K code should have the k
selector on it. You can express multiple selectors on a code block by putting
them between curly braces and prepending each with the . character. For
example, here is a code block with the foo and k selectors:

  syntax Int ::= foo(Int) [function]
  rule foo(0) => 0

Because this code block contains the k Markdown selector, by default it is
included as part of the K definition being compiled.

Exercise

Confirm this fact by using krun to evaluate foo(0).

Markdown Selector Expressions

By default, as previously stated, K includes in the definition any code block
with the k selector. However, this is merely a specific instance of a general
principle, namely, that K allows you to control which selectors get included
in your K definition. This is done by means of the --md-selector flag to
kompile. This flag accepts a Markdown selector expression, which you
can essentially think of as a kind of Boolean algebra over Markdown selectors.
Each selector becomes an atom, and you can combine these atoms via the &,
|, !, and () operators.

Here is a grammar, written in K, of the language of Markdown selector
expressions:

  syntax Selector ::= r"[0-9a-zA-Z_]+" [token]
  syntax SelectorExp ::= Selector
                       | "(" SelectorExp ")" [bracket]
                       > right:
                         "!" SelectorExp
                       > right:
                         SelectorExp "&" SelectorExp
                       > right:
                         SelectorExp "|" SelectorExp

Here is a selector expression that selects all the K code blocks in this
definition except the one immediately above:

k & (! selector)

Addendum

This code block exists in order to make the above lesson a syntactically valid
K definition. Consider why it is necessary.

endmodule

Exercises

  1. Compile this lesson with the selector expression k & (! foo) and confirm
    that you get a parser error if you try to evaluate the foo function with the
    resulting definition.

  2. Compile Lesson 1.3
    as a K definition. Identify why it fails to compile. Then pass an appropriate
    --md-selector to the compiler in order to make it compile.

  3. Modify your calculator application from Lesson 1.7, Exercise 2, to be written
    in a literate style. Consider what text might be appropriate to turn the
    resulting markdown file into documentation for your calculator.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.9: Unparsing and the format and color attributes.

Lesson 1.9: Unparsing and the format and color attributes

The purpose of this lesson is to teach the user about how terms are
pretty-printed in K, and how the user can make adjustments to the default
settings for how to print specific terms.

Parsing, Execution, and Unparsing

When you use krun to interpret a program, the tool passes through three major
phases. In the first, parsing, the program itself is parsed using either kast
or an ahead-of-time parser generated via Bison, and the resulting AST becomes
the input to the interpreter. In the second phase, execution, K evaluates
functions and (as we will discuss in depth later) performs rewrite steps to
iteratively transform the program state. The third and final phase is called
unparsing, because it consists of taking the final state of the application
after the program has been interpreted, and converting it from an AST back into
text that (in theory, anyway) could be parsed back into the same AST that was
the output of the execution phase.

In practice, parsing is not always precisely reversible. It turns out
(although we are not going to cover exactly why this is here), that
constructing a sound algorithm that takes a grammar and an AST and emits text
that could be parsed via that grammar to the original AST is an
NP-hard problem. As a result, in the interests of avoiding exponential time
algorithms when users rarely care about unparsing being completely sound, we
take certain shortcuts that provide a linear-time algorithm that approximates
a sound solution to the problem while sacrificing the notion that the result
can be parsed into the exact original term in all cases.

This is a lot of theoretical explanation, but at root, the unparsing process
is fairly simple: it takes a K term that is the output of execution and pretty
prints it according to the syntax defined by the user in their K definition.
This is useful because the original AST is not terribly user-readable, and it
is difficult to visualize the entire term or decipher information about the
final state of the program at a quick glance. Of course, in rare cases, the
pretty-printed configuration loses information of relevance, which is why K
allows you to obtain the original AST on request.

As an example of all of this, consider the following K definition
(lesson-09-a.k):

module LESSON-09-A
  imports BOOL

  syntax Exp ::= "(" Exp ")" [bracket]
               | Bool
               > "!" Exp
               > left:
                 Exp "&&" Exp
               | Exp "^" Exp
               | Exp "||" Exp

  syntax Exp ::= id(Exp) [function]
  rule id(E) => E
endmodule

This is similar to the grammar we defined in LESSON-06-C, with the difference
that the Boolean expressions are now constructors of sort Exp and we define a
trivial function over expressions that returns its argument unchanged.

We can now parse a simple program in this definition and use it to unparse some
Boolean expressions. For example (exp.bool):

id(true&&false&&!true^(false||true))

Here is a program that is not particularly legible at first glance, because all
extraneous whitespace has been removed. However, if we run krun exp.bool, we
see that the result of the unparser will pretty-print this expression rather
nicely:

<k>
  true && false && ! true ^ ( false || true ) ~> .
</k>

Notably, not only does K insert whitespace where appropriate, it is also smart
enough to insert parentheses where necessary in order to ensure the correct
parse. For example, without those parentheses, the expression above would parse
equivalent to the following one:

(((true && false) && ! true) ^ false) || true

Indeed, you can confirm this by passing that exact expression to the id
function and evaluating it, then looking at the result of the unparser:

<k>
  true && false && ! true ^ false || true ~> .
</k>

Here, because the meaning of the AST is the same both with and without
parentheses, K does not insert any parentheses when unparsing.

Exercise

Modify the grammar of LESSON-09-A above so that the binary operators are
right associative. Try unparsing exp.bool again, and note how the result is
different. Explain the reason for the difference.

Custom unparsing of terms

You may have noticed that right now, the unparsing of terms is not terribly
imaginative. All it is doing is taking each child of the term, inserting it
into the non-terminal positions of the production, then printing the production
with a space between each terminal or non-terminal. It is easy to see why this
might not be desirable in some cases. Consider the following K definition
(lesson-09-b.k):

module LESSON-09-B
  imports BOOL

  syntax Stmt ::= "{" Stmt "}" | "{" "}"
                > right:
                  Stmt Stmt
                | "if" "(" Bool ")" Stmt
                | "if" "(" Bool ")" Stmt "else" Stmt [avoid]
endmodule

This is a statement grammar, simplified to the point of meaninglessness, but
still useful as an object lesson in unparsing. Consider the following program
in this grammar (if.stmt):

if (true) {
  if (true) {}
  if (false) {}
  if (true) {
    if (false) {} else {}
  } else {
    if (false) {}
  }
}

This is how that term would be unparsed if it appeared in the output of krun:

if ( true ) { if ( true ) { } if ( false ) { } if ( true ) { if ( false ) { } else { } } else { if ( false ) { } } }

This is clearly much less legible than we started with! What are we to do?
Well, K provides an attribute, format, that can be applied to any production,
which controls how that production gets unparsed. You've seen how it gets
unparsed by default, but via this attribute, the developer has complete control
over how the term is printed. Of course, the user can trivially create ways to
print terms that would not parse back into the same term. Sometimes this is
even desirable. But in most cases, what you are interested in is controlling
the line breaking, indentation, and spacing of the production.

Here is an example of how you might choose to apply the format attribute
to improve how the above term is unparsed (lesson-09-c.k):

module LESSON-09-C
  imports BOOL

  syntax Stmt ::= "{" Stmt "}" [format(%1%i%n%2%d%n%3)] | "{" "}" [format(%1%2)]
                > right:
                  Stmt Stmt [format(%1%n%2)]
                | "if" "(" Bool ")" Stmt [format(%1 %2%3%4 %5)]
                | "if" "(" Bool ")" Stmt "else" Stmt [avoid, format(%1 %2%3%4 %5 %6 %7)]
endmodule

If we compile this new definition and unparse the same term, this is the
result we get:

if (true) {
  if (true) {}
  if (false) {}
  if (true) {
    if (false) {} else {}
  } else {
    if (false) {}
  }
}

This is the exact same text we started with! By adding the format attributes,
we were able to indent the body of code blocks, adjust the spacing of if
statements, and put each statement on a new line.

How exactly was this achieved? Well, each time the unparser reaches a term,
it looks at the format attribute of that term. That format attribute is a
mix of characters and format codes. Format codes begin with the %
character. Each character in the format attribute other than a format code is
appended verbatim to the output, and each format code is handled according to
its meaning, transformed (possibly recursively) into a string of text, and
spliced into the output at the position the format code appears in the format
string.

Provided for reference is a table with a complete list of all valid format
codes, followed by their meaning:

Format Code Meaning
n Insert '\n' followed by the current indentation level
i Increase the current indentation level by 1
d Decrease the current indentation level by 1
c Move to the next color in the list of colors for this production (see next section)
r Reset color to the default foreground color for the terminal (see next section)
an integer Print a terminal or non-terminal from the production. The integer is treated as a 1-based index into the terminals and non-terminals of the production.

If the offset refers to a terminal, move to the next color in the list of colors for this production, print the value of that terminal, then reset the color to the default foreground color for the terminal.

If the offset refers to a regular expression terminal, it is an error.

If the offset refers to a non-terminal, unparse the corresponding child of the current term (starting with the current indentation level) and print the resulting text, then set the current color and indentation level to the color and indentation level following unparsing that term.
other char Print that character verbatim

Exercise

Change the format attributes for LESSON-09-C so that if.stmt will unparse
as follows:

if (true)
{
  if (true)
  {
  }
  if (false)
  {
  }
  if (true)
  {
    if (false)
    {
    }
    else
    {
    }
  }
  else
  {
    if (false)
    {
    }
  }
}

Output coloring

When the output of unparsing is displayed on a terminal supporting colors, K
is capable of coloring the output, similar to what is possible with a syntax
highlighter. This is achieved via the color and colors attributes.

Essentially, both the color and colors attributes are used to construct a
list of colors associated with each production, and then the format attribute
is used to control how those colors are used to unparse the term. At its most
basic level, you can set the color attribute to color all the terminals in
the production a certain color, or you can use the colors attribute to
specify a comma-separated list of colors for each terminal in the production.
At a more advanced level, the %c and %r format codes control how the
formatter interacts with the list of colors specified by the colors
attribute. You can essentially think of the color attribute as a way of
specifying that you want all the colors in the list to be the same color.

Note that the %c and %r format codes are relatively primitive in nature.
The color and colors attributes merely maintain a list of colors, whereas
the %c and %r format codes merely control how to advance through that list
and how individual text is colored.

It is an error if the colors attribute does not provide all the colors needed
by the terminals and escape codes in the production. %r does not change the
position in the list of colors at all, so the next %c will advance to the
following color.

As a complete example, here is a variant of LESSON-09-A which colors the
various boolean operators:

module LESSON-09-D
  imports BOOL

  syntax Exp ::= "(" Exp ")" [bracket]
               | Bool
               > "!" Exp [color(yellow)]
               > left:
                 Exp "&&" Exp [color(red)]
               | Exp "^" Exp [color(blue)]
               | Exp "||" Exp [color(green)]

  syntax Exp ::= id(Exp) [function]
  rule id(E) => E
endmodule

For a complete list of allowed colors, see
here.

Exercises

  1. Use the color attribute on LESSON-09-C to color the keywords true and
    false one color, the keywords if and else another color, and the operators
    (, ), {, and } a third color.

  2. Use the format, color, and colors attributes to tell the unparser to
    style the expression grammar from Lesson 1.8, Exercise 3 according to your own
    personal preferences for syntax highlighting and code formatting. You can
    view the result of the unparser on a function term without evaluating that
    function by means of the command kparse <file> | kore-print -.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.10: Strings.

Lesson 1.10: Strings

The purpose of this lesson is to explain how to use the String sort in K to
represent sequences of characters, and explain where to find additional
information about builtin functions over strings.

The String Sort

In addition to the Int and Bool sorts covered in
Lesson 1.6, K provides, among others, the
String sort to represent sequences of characters. You can import this
functionality via the STRING-SYNTAX module, which contains the syntax of
string literals in K, and the STRING module, which contains all the functions
that operate over the String type.

Strings in K are double-quoted. The following list of escape sequences is
supported:

Escape Sequence Meaning
\" The literal character "
\\ The literal character \
\n The newline character (ASCII code 0x0a)
\r The carriage return character (ASCII code 0x0d)
\t The tab character (ASCII code 0x09)
\f The form feed character (ASCII code 0x0c)
\x00 \x followed by 2 hexadecimal digits indicates a code point between 0x00 and 0xFF
\u0000 \u followed by 4 hexadecimal digits indicates a code point between 0x0000 and 0xFFFF
\U00000000 \U followed by 8 hexadecimal digits indicates a code point between 0x000000 and 0x10FFFF

Please note that as of the current moment, K's unicode support is not fully
complete, so you may run into errors using code points greater than 0xff.

As an example, you can construct a string literal containing the following
block of text:

This is an example block of text.
Here is a quotation: "Hello world."
	This line is indented.
ÁÉÍÓÚ

Like so:

"This is an example block of text.\nHere is a quotation: \"Hello world.\"\n\tThis line is indented.\n\xc1\xc9\xcd\xd3\xda\n"

Basic String Functions

The full list of functions provided for the String sort can be found in
domains.md, but here we
describe a few of the more basic ones.

String concatenation

The concatenation operator for strings is +String. For example, consider
the following K rule that constructs a string from component parts
(lesson-10.k):

module LESSON-10
  imports STRING

  syntax String ::= msg(String) [function]
  rule msg(S) => "The string you provided: " +String S +String "\nHave a nice day!"
endmodule

Note that this operator is O(N), so repeated concatenations are inefficient.
For information about efficient string concatenation, refer to
Lesson 2.14.

String length

The function to return the length of a string is lengthString. For example,
lengthString("foo") will return 3, and lengthString("") will return 0.
The return value is the length of the string in code points.

Substring computation

The function to compute the substring of a string is substrString. It
takes two string indices, starting from 0, and returns the substring within the
range [start..end). It is only defined if end >= start, start >= 0, and
end <= length of string. Here, for example, we return the first 5 characters
of a string:

substrString(S, 0, 5)

Here we return all but the first 3 characters:

substrString(S, 3, lengthString(S))

Exercises

  1. Write a function that takes a paragraph of text (i.e., a sequence of
    sentences, each ending in a period), and constructs a new (nonsense) sentence
    composed of the first word of each sentence, followed by a period. Do not
    worry about capitalization or periods within the sentence which do not end the
    sentence (e.g. "Dr."). You can assume that all whitespace within the paragraph
    are spaces. For more information about the functions over strings required to
    implement such a function, refer to domains.md.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.11: Casting Terms.

Lesson 1.11: Casting Terms

The purpose of this lesson is to explain how to use cast expressions in
order to disambiguate terms using sort information. We also explain how the
variable sort inference algorithm works in K, and how to change the default
behavior by casting variables to a particular sort.

Casting in K

Sometimes the grammar you write for your rules in K can be a little bit
ambiguous on purpose. While grammars for programming languages may be
unambiguous when considered in their entirety, K allows you to write rules
involving arbitrary fragments of that grammar, and those fragments can
sometimes be ambiguous by themselves, or similar enough to other fragments
of the grammar to trigger ambiguity. As a result, in addition to the tools
covered in Lesson 1.4, K provides one
additional powerful tool for disambiguation: cast expressions.

K provides three main types of casts: the semantic cast, the strict cast, and
the projection cast. We will cover each of them, and their similarities and
differences, in turn.

Semantic casts

The most basic, and most common, type of cast in K is called the
semantic cast. For every sort S declared in a module, K provides the
following (implicit) production for use in sentences:

  syntax S ::= S ":S"

Note that S simply represents the name of the sort. For example, if we
defined a sort Exp, the actual production for that sort would be:

  syntax Exp ::= Exp ":Exp"

At runtime, this expression will not actually exist; it is merely an annotation
to the compiler describing the sort of the term inside the cast. It is telling
the compiler that the term inside the cast must be of sort Exp. For example,
if we had the following grammar:

module LESSON-11-A
  imports INT

  syntax Exp ::= Int | Exp "+" Exp
  syntax Stmt ::= "if" "(" Exp ")" Stmt | "{" "}"
endmodule

Then we would be able to write 1:Exp, or (1 + 2):Exp, but not {}:Exp.

You can also restrict the sort that a variable in a rule will match by casting
it. For example, consider the following additional module:

module LESSON-11-B
  imports LESSON-11-A
  imports BOOL

  syntax Term ::= Exp | Stmt
  syntax Bool ::= isExpression(Term) [function]

  rule isExpression(_E:Exp) => true
  rule isExpression(_) => false [owise]
endmodule

Here we have defined a very simple function that decides whether a term is
an expression or a statement. It does this by casting the variable inside the
isExpression rule to sort Exp. As a result, that variable will only match terms
of sort Exp. Thus, isExpression(1) will return true, as will isExpression(1 + 2), but
isExpression({}) will return false.

Exercise

Verify this fact for yourself by running isExpression on the above examples. Then
write an isStatement function, and test that it works as expected.

Strict casts

On occasion, a semantic cast is not strict enough. It might be that you want
to, for disambiguation purposes, say exactly what sort a term is. For
example, consider the following definition:

module LESSON-11-C
  imports INT

  syntax Exp ::= Int
               | "add[" Exp "," Exp "]"   [group(exp)]
  syntax Exp2 ::= Exp
               | "add[" Exp2 "," Exp2 "]" [group(exp2)]
endmodule

This grammar is a little ambiguous and contrived, but it serves to demonstrate
how a semantic cast might be insufficient to disambiguate a term. If we were
to write the term add[ I1:Int , I2:Int ]:Exp2, the term would be ambiguous,
because the cast is not sufficiently strict to determine whether you mean
to derive the "add" production defined in group exp or the one in group exp2.

In this situation, there is a solution: the strict cast. For every sort
S in your grammar, K also defines the following production:

  syntax S ::= S "::S"

This may at first glance seem the same as the previous cast. And indeed,
from the perspective of the grammar and from the perspective of rewriting,
they are in fact identical. However, the second variant has a unique meaning
in the type system of K: namely, the term inside the cast cannot be a
subsort, i.e., a term of another sort S2 such that the production
syntax S ::= S2 exists.

As a result, if we were to write in the above grammar the term
add[ I1:Int , I2:Int ]::Exp2, then we would know that the second derivation above
should be chosen, whereas if we want the first derivation, we could write
add[ I1:Int , I2:Int ]::Exp.

Care must be taken when using a strict cast with brackets. For example, consider a
similar grammar but using an infix "+":

module LESSON-11-D
  imports INT

  syntax Exp ::= Int
               | Exp "+" Exp   [group(exp)]
  syntax Exp2 ::= Exp
               | Exp2 "+" Exp2 [group(exp2)]
               | "(" Exp2 ")"  [bracket]
endmodule

The term I1:Int + I2:Int is ambiguous and could refer to either the production
in group exp or the one in group exp2. To differentiate, you might try to write
(I1:Int + I2:Int)::Exp2 similarly to the previous example.

Unfortunately though, this is still ambiguous. Here, the strict cast ::Exp2 applies
directly to the brackets themselves rather than the underlying term within those brackets.
As a result, it enforces that (I1:Int + I2:Int) cannot be a strict subsort of Exp2, but
it has no effect on the sort of the subterm I1:Int + I2:Int.

For cases like this, K provides an alternative syntax for strict casts:

  syntax S ::= "{" S "}::S"

The ambiguity can then be resolved with {I1:Int + I2:Int}::Exp or {I1:Int + I2:Int}::Exp2.

Projection casts

Thus far we have focused entirely on casts which exist solely to inform the
compiler about the sort of terms. However, sometimes when dealing with grammars
containing subsorts, it can be desirable to reason with the subsort production
itself, which injects one sort into another. Remember from above that such
a production looks like syntax S ::= S2. This type of production, called a
subsort production, can be thought of as a type of inheritance involving
constructors. If we have the above production in our grammar, we say that S2
is a subsort of S, or that any S2 is also an S. K implicitly maintains a
symbol at runtime which keeps track of where such subsortings occur; this
symbol is called an injection.

Sometimes, when one sort is a subsort of another, it can be the case that
a function returns one sort, but you actually want to cast the result of
calling that function to another sort which is a subsort of the first sort.
This is similar to what happens with inheritance in an object-oriented
language, where you might cast a superclass to a subclass if you know for
sure the object at runtime is in fact an instance of that class.

K provides something similar for subsorts: the projection cast.

For each pair of sorts S and S2, K provides the following production:

  syntax S ::= "{" S2 "}" ":>S"

What this means is that you take any term of sort S2 and cast it to sort
S. If the term of sort S2 consists of an injection containing a term of sort
S, then this will return that term. Otherwise, an error occurs and rewriting
fails, returning the projection function which failed to apply. The sort is
not actually checked at compilation time; rather, it is a runtime check
inserted into the code that runs when the rule applies.

For example, here is a module that makes use of projection casts:

module LESSON-11-E
  imports INT
  imports BOOL

  syntax Exp ::= Int | Bool | Exp "+" Exp | Exp "&&" Exp

  syntax Exp ::= eval(Exp) [function]
  rule eval(I:Int) => I
  rule eval(B:Bool) => B
  rule eval(E1 + E2) => {eval(E1)}:>Int +Int {eval(E2)}:>Int
  rule eval(E1 && E2) => {eval(E1)}:>Bool andBool {eval(E2)}:>Bool
endmodule

Here we have defined constructors for a simple expression language over
Booleans and integers, as well as a function eval that evaluates these
expressions to a value. Because that value could be an integer or a Boolean,
we need the casts in the last two rules in order to meet the type signature of
+Int and andBool. Of course, the user can write ill-formed expressions like
1 && true or false + true, but these will cause errors at runtime, because
the projection cast will fail.

Exercises

  1. Extend the eval function in LESSON-11-E to include Strings and add a .
    operator which concatenates them.

  2. Modify your solution from Lesson 1.9, Exercise 2 by using an Exp sort to
    express the integer and Boolean expressions that it supports, in the same style
    as LESSON-11-E. Then write an eval function that evaluates all terms of
    sort Exp to either a Bool or an Int.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.12: Syntactic Lists.

Lesson 1.12: Syntactic Lists

The purpose of this lesson is to explain how K provides support for syntactic
repetition through the use of the List{} and NeList{} constructs,
generally called syntactic lists.

The List{} construct

Sometimes, when defining a grammar in K, it is useful to define a syntactic
construct consisting of an arbitrary-length sequence of items. For example,
you might wish to define a function call construct, and need to express a way
of passing arguments to the function. You can in theory simply define these
productions using ordinary constructors, but it can be tricky to get the syntax
exactly right in K without a lot of tedious glue code.

For this reason, K provides a way of specifying that a non-terminal represents
a syntactic list (lesson-12-a.k):

module LESSON-12-A-SYNTAX
  imports INT-SYNTAX

  syntax Ints ::= List{Int,","}
endmodule

module LESSON-12-A
  imports LESSON-12-A-SYNTAX
endmodule

Note that instead of a sequence of terminals and non-terminals, the right hand
side of the Ints production contains the symbol List followed by two items
in curly braces. The first item is the non-terminal which is the element type
of the list, and the second item is a terminal representing the separator of
the list. As a special case, lists which are separated only by whitespace can
be specified with a separator of "".

This List{} construct is roughly equivalent to the following definition
(lesson-12-b.k):

module LESSON-12-B-SYNTAX
  imports INT-SYNTAX

  syntax Ints ::= Int "," Ints | ".Ints"
endmodule

module LESSON-12-B
  imports LESSON-12-B-SYNTAX
endmodule

As you can see, the List{} construct represents a cons-list with an element
at the head and another list at the tail. The empty list is represented by
a . followed by the sort of the list.

However, the List{} construct provides several key syntactic conveniences
over the above definition. First of all, when writing a list in a rule,
explicitly writing the terminator is not always required. For example, consider
the following additional module (lesson-12-c.k):

module LESSON-12-C
  imports LESSON-12-A
  imports INT

  syntax Int ::= sum(Ints) [function]
  rule sum(I:Int) => I
  rule sum(I1:Int, I2:Int, Is:Ints) => sum(I1 +Int I2, Is)
endmodule

Here we see a function that sums together a non-empty list of integers. Note in
particular the first rule. We do not explicitly mention .Ints, but in fact,
the rule in question is equivalent to the following rule:

  rule sum(I:Int, .Ints) => I

The reason for this is that K will automatically insert a list terminator
anywhere a syntactic list is expected, but an element of that list appears
instead. This works even with lists of more than one element:

  rule sum(I1:Int, I2:Int) => I1 +Int I2

This rule is redundant, but here we explicitly match a list of exactly two
elements, because the .Ints is implicitly added after I2.

Parsing Syntactic Lists in Programs

An additional syntactic convenience takes place when you want to express a
syntactic list in the input to krun. In this case, K will automatically
transform the grammar in LESSON-12-B-SYNTAX into the following
(lesson-12-d.k):

module LESSON-12-D
  imports INT-SYNTAX

  syntax Ints ::= #NonEmptyInts | #IntsTerminator
  syntax #NonEmptyInts ::= Int "," #NonEmptyInts
                         | Int #IntsTerminator
  syntax #IntsTerminator ::= ""
endmodule

This allows you to express the usual comma-separated list of arguments where
an empty list is represented by the empty string, and you don't have to
explicitly terminate the list. Because of this, we can write the syntax
of function calls in C very easily (lesson-12-e.k):

module LESSON-12-E
  syntax Id ::= r"[a-zA-Z_][a-zA-Z0-9_]*" [token]
  syntax Exp ::= Id | Exp "(" Exps ")"
  syntax Exps ::= List{Exp,","}
endmodule

Exercise

Write a function concat which takes a list of String and concatenates them
all together. Do not worry if the function is O(n^2).
Test your implementation using the syntactic sugar for lists added by the parser.

Then write some function call expressions using identifiers in C and verify with
kast that the above grammar captures the intended syntax. Make sure to test
with function calls with zero, one, and two or more arguments.

The NeList{} construct

One limitation of the List{} construct is that it is always possible to
write a list of zero elements where a List{} is expected. While this is
desirable in a number of cases, it is sometimes not what the grammar expects.

For example, in C, it is not allowable for an enum definition to have zero
members. In other words, if we were to write the grammar for enumerations like
so (lesson-12-f.k):

module LESSON-12-F
  syntax Id ::= r"[a-zA-Z_][a-zA-Z0-9_]*" [token]
  syntax Exp ::= Id

  syntax EnumSpecifier ::= "enum" Id "{" Ids "}"
  syntax Ids ::= List{Id,","}
endmodule

Then we would be syntactically allowed to write enum X {}, which instead,
ought to be a syntax error.

For this reason, we introduce the additional NeList{} construct. The syntax
is identical to List{}, except with NeList instead of List before the
curly braces. When parsing rules, it behaves identically to the List{}
construct. However, when parsing inputs to krun, the above grammar, if we
replaced syntax Ids ::= List{Id,","} with syntax Ids ::= NeList{Id,","},
would become equivalent to the following (lesson-12-g.k):

module LESSON-12-G
  syntax Id ::= r"[a-zA-Z_][a-zA-Z0-9_]*" [token]
  syntax Exp ::= Id

  syntax EnumSpecifier ::= "enum" Id "{" Ids "}"
  syntax Ids ::= Id | Id "," Ids
endmodule

In other words, only non-empty lists of Id would be allowed.

Exercises

  1. Modify the sum function in LESSON-12-C so that the Ints sort is an
    NeList{}. Verify that calling sum() with no arguments is now a syntax
    error.

  2. Write a modified sum function with the List construct that can also sum
    up an empty list of arguments. In such a case, the sum ought to be 0.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.13: Basics of K Rewriting.

Lesson 1.13: Basics of K Rewriting

The purpose of this lesson is to explain how rewrite rules that are not the
definition of a function behave, and how, using these rules, you can construct
a semantics of programs in a programming language in K.

Recap: Function rules in K

Recall from Lesson 1.2 that we have, thus far,
introduced two types of productions in K: constructors and functions.
A function is identified by the function attribute placed on the
production. As you may recall, when we write a rule with a function on the
left-hand side of the => operator, we are defining the meaning of that
function for inputs which match the patterns on the left-hand side of the rule.
If the argument to the function match the patterns, then the function is
evaluated to the value constructed by substituting the bindings for the
variables into the right-hand side of the rule.

Top-level rules

However, function rules are not the only type of rule permissible in K, nor
even the most frequently used. K also has a concept of a
top-level rewrite rule. The simplest way to ensure that a rule is treated
as a top-level rule is for the left-hand side of the rule to mention one or
more cells. We will cover how cells work and are declared in more detail
in a later lesson, but for now, what you should know is that when we ran krun
in our very first example in Lesson 1.2 and got the following output:

<k>
  Yellow ( ) ~> .
</k>

<k> is a cell, known by convention as the K cell. This cell is available
by default in any definition without needing to be explicitly declared.

The K cell contains a single term of sort K. K is a predefined sort in K
with two constructors, that can be roughly represented by the following
grammar:

  syntax K ::= KItem "~>" K
             | "."

As a syntactic convenience, K allows you to treat ~> like it is an
associative list (i.e., as if it were defined as syntax K ::= K "~>" K).
When a definition is compiled, it will automatically transform the rules you
write so that they treat the K sort as a cons-list. Another syntactic
convenience is that, for disambiguation purposes, you can write .K anywhere
you would otherwise write . and the meaning is identical.

Now, you may notice that the above grammar mentions the sort KItem. This is
another built-in sort in K. For every sort S declared in a definition (with
the exception of K and KItem), K will implicitly insert the following
production:

  syntax KItem ::= S

In other words, every sort is a subsort of the sort KItem, and thus a term
of any sort can be injected as an element of a term of sort K, also called
a K sequence.

By default, when you krun a program, the AST of the program is inserted as
the sole element of a K sequence into the <k> cell. This explains why we
saw the output we did in Lesson 1.2.

With these preliminaries in mind, we can now explain how top-level rewrite
rules work in K. Put simply, any rule where there is a cell (such as the K
cell) at the top on the left-hand side will be a top-level rewrite rule. Once
the initial program has been inserted into the K cell, the resulting term,
called the configuration, will be matched against all the top-level
rewrite rules in the definition. If only one rule matches, the substitution
generated by the matching will be applied to the right-hand side of the rule
and the resulting term is rewritten to be the new configuration. Rewriting
proceeds by iteratively applying rules, also called taking steps, until
no top-level rewrite rule can be applied. At this point the configuration
becomes the final configuration and is output by krun.

If more than one top-level rule applies, by default, K will pick just one
of those rules, apply it, and continue rewriting. However, it is
non-deterministic which rule applies. In theory, it could be any of them.
By passing the --search flag to krun, you are able to tell krun to
explore all possible non-deterministic choices, and generate a complete list of
all possible final configurations reachable by each nondeterminstic choice that
can be made. Note that the --search flag to krun only works if you pass
--enable-search to kompile first.

Unlike top-level rewrite rules, function rules are not associated with any
particular set of cells in the configuration (although they can contain cells
in their function arguments and return value). While top-level rewrite rules
apply to the entire term being rewritten, function rules apply anywhere a
function application for that function appears, and are immediately rewritten
to their return value in that position.

Another key distinction between top-level rules and function rules is that
function symbols, i.e., productions with the function attribute, are
mathematical functions rather than constructors. While a constructor is
logically distinct from any other constructor of the same sort, and can be
matched against unconditionally, a function does not necessaraily have the
same restriction unless it happens to be an injective function. Thus, two
function symbols with different arguments may still ultimately produce the
same value and thus compare equal to one another. Due to this, concrete
execution (i.e., all K definitions introduced thus far; see Lesson 1.21)
introduces the restriction that you cannot match on a function symbol on the
left-hand side of a rule, except as the top symbol on the left-hand side of
a function rule. This restriction will be later lifted when we introduce the
Haskell Backend which performs symbolic execution.

Exercise

Pass a program containing no functions to krun. You can use a term of sort
Exp from LESSON-11-E. Observe the output and try to understand why you get
the output you do. Then write two rules that rewrite that program to another.
Run krun --search on that program and observe both results. Then add a third
rule that rewrites one of those results again. Test that that rule applies as
well.

Using top-level rules to evaluate expressions

Thus far, we have focused primarily on defining functions over constructors
in K. However, now that we have a basic understanding of top-level rules,
it is possible to introduce a rewrite system to our definitions. A rewrite
system is a collection of top-level rewrite rules which performs an organized
transformation of a particular program into a result which expresses the
meaning of that program. For example, we might rewrite an expression in a
programming language into a value representing the result of evaluating that
expression.

Recall in Lesson 1.11, we wrote a simple grammar of Boolean and integer
expressions that looked roughly like this (lesson-13-a.k):

module LESSON-13-A
  imports INT

  syntax Exp ::= Int
               | Bool
               | Exp "+" Exp
               | Exp "&&" Exp
endmodule

In that lesson, we defined a function eval which evaluated such expressions
to either an integer or Boolean.

However, it is more idiomatic to evaluate such expressions using top-level
rewrite rules. Here is how one might do so in K (lesson-13-b.k):

module LESSON-13-B-SYNTAX
  imports UNSIGNED-INT-SYNTAX
  imports BOOL-SYNTAX

  syntax Val ::= Int | Bool
  syntax Exp ::= Val
               > left: Exp "+" Exp
               > left: Exp "&&" Exp
endmodule

module LESSON-13-B
  imports LESSON-13-B-SYNTAX
  imports INT
  imports BOOL

  rule <k> I1:Int + I2:Int ~> K:K </k> => <k> I1 +Int I2 ~> K </k>
  rule <k> B1:Bool && B2:Bool ~> K:K </k> => <k> B1 andBool B2 ~> K </k>

  syntax KItem ::= freezer1(Val) | freezer2(Exp)
                 | freezer3(Val) | freezer4(Exp)

  rule <k> E1:Val + E2:Exp ~> K:K </k> => <k> E2 ~> freezer1(E1) ~> K </k> [priority(51)]
  rule <k> E1:Exp + E2:Exp ~> K:K </k> => <k> E1 ~> freezer2(E2) ~> K </k> [priority(52)]
  rule <k> E1:Val && E2:Exp ~> K:K </k> => <k> E2 ~> freezer3(E1) ~> K </k> [priority(51)]
  rule <k> E1:Exp && E2:Exp ~> K:K </k> => <k> E1 ~> freezer4(E2) ~> K </k> [priority(52)]

  rule <k> E2:Val ~> freezer1(E1) ~> K:K </k> => <k> E1 + E2 ~> K </k>
  rule <k> E1:Val ~> freezer2(E2) ~> K:K </k> => <k> E1 + E2 ~> K </k>
  rule <k> E2:Val ~> freezer3(E1) ~> K:K </k> => <k> E1 && E2 ~> K </k>
  rule <k> E1:Val ~> freezer4(E2) ~> K:K </k> => <k> E1 && E2 ~> K </k>
endmodule

This is of course rather cumbersome currently, but we will soon introduce
syntactic convenience which makes writing definitions of this type considerably
easier. For now, notice that there are roughly 3 types of rules here: the first
matches a K cell in which the first element of the K sequence is an Exp whose
arguments are values, and rewrites the first element of the sequence to the
result of that expression. The second also matches a K cell with an Exp in
the first element of its K sequence, but it matches when one or both arguments
of the Exp are not values, and replaces the first element of the K sequence
with two new elements: one being an argument to evaluate, and the other being
a special constructor called a freezer. Finally, the third matches a K
sequence where a Val is first, and a freezer is second, and replaces them
with a partially evaluated expression.

This general pattern is what is known as heating an expression,
evaluating its arguments, cooling the arguments into the expression
again, and evaluating the expression itself. By repeatedly performing
this sequence of actions, we can evaluate an entire AST containing a complex
expression down into its resulting value.

Exercise

Write an addition expression with integers. Use krun --depth 1 to see the
result of rewriting after applying a single top-level rule. Gradually increase
the value of --depth to see successive states. Observe how this combination
of rules is eventually able to evaluate the entire expression.

Simplifying the evaluator: Local rewrites and cell ellipses

As you saw above, the definition we wrote is rather cumbersome. Over the
remainder of Lessons 1.13 and 1.14, we will greatly simplify it. The first step
in doing so is to teach a bit more about the rewrite operator, =>. Thus far,
all the rules we have written look like rule LHS => RHS. However, this is not
the only way the rewrite operator can be used. It is actually possible to place
a constructor or function at the very top of the rule, and place rewrite
operators inside that term. While a rewrite operator cannot appear nested
inside another rewrite operator, by doing this, we can express that some parts
of what we are matching are not changed by the rewrite operator. For
example, consider the following rule from above:

  rule <k> I1:Int + I2:Int ~> K:K </k> => <k> I1 +Int I2 ~> K </k>

We can equivalently write it like following:

  rule <k> (I1:Int + I2:Int => I1 +Int I2) ~> _:K </k>

When you put a rewrite inside a term like this, in essence, you are telling
the rule to only rewrite part of the left-hand side to the right-hand side.
In practice, this is implemented by lifting the rewrite operator to the top of
the rule by means of duplicating the surrounding context.

There is a way that the above rule can be simplified further, however. K
provides a special syntax for each cell containing a term of sort K, indicating
that we want to match only on some prefix of the K sequence. For example, the
above rule can be simplified further like so:

  rule <k> I1:Int + I2:Int => I1 +Int I2 ...</k>

Here we have placed the symbol ... immediately prior to the </k> which ends
the cell. What this tells the compiler is to take the contents of the cell,
treat it as the prefix of a K sequence, and insert an anonymous variable of
sort K at the end. Thus we can think of ... as a way of saying we
don't care about the part of the K sequence after the beginning, leaving
it unchanged.

Putting all this together, we can rewrite LESSON-13-B like so
(lesson-13-c.k):

module LESSON-13-C-SYNTAX
  imports UNSIGNED-INT-SYNTAX
  imports BOOL-SYNTAX

  syntax Val ::= Int | Bool
  syntax Exp ::= Val
               > left: Exp "+" Exp
               > left: Exp "&&" Exp
endmodule

module LESSON-13-C
  imports LESSON-13-C-SYNTAX
  imports INT
  imports BOOL

  rule <k> I1:Int + I2:Int => I1 +Int I2 ...</k>
  rule <k> B1:Bool && B2:Bool => B1 andBool B2 ...</k>

  syntax KItem ::= freezer1(Val) | freezer2(Exp)
                 | freezer3(Val) | freezer4(Exp)

  rule <k> E1:Val + E2:Exp => E2 ~> freezer1(E1) ...</k> [priority(51)]
  rule <k> E1:Exp + E2:Exp => E1 ~> freezer2(E2) ...</k> [priority(52)]
  rule <k> E1:Val && E2:Exp => E2 ~> freezer3(E1) ...</k> [priority(51)]
  rule <k> E1:Exp && E2:Exp => E1 ~> freezer4(E2) ...</k> [priority(52)]

  rule <k> E2:Val ~> freezer1(E1) => E1 + E2 ...</k>
  rule <k> E1:Val ~> freezer2(E2) => E1 + E2 ...</k>
  rule <k> E2:Val ~> freezer3(E1) => E1 && E2 ...</k>
  rule <k> E1:Val ~> freezer4(E2) => E1 && E2 ...</k>
endmodule

This is still rather cumbersome, but it is already greatly simplified. In the
next lesson, we will see how additional features of K can be used to specify
heating and cooling rules much more compactly.

Exercises

  1. Modify LESSON-13-C to add rules to evaluate integer subtraction.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.14: Defining Evaluation Order.

Lesson 1.14: Defining Evaluation Order

The purpose of this lesson is to explain how to use the heat and cool
attributes, context and context alias sentences, and the strict and
seqstrict attributes to more compactly express heating and cooling in K,
and to express more advanced evaluation strategies in K.

The heat and cool attributes

Thus far, we have been using rule priority and casts to express when to heat
an expression and when to cool it. For example, the rules for heating have
lower priority, so they do not apply if the term could be evaluated instead,
and the rules for heating are expressly written only to apply if the argument
of the expression is a value.

However, K has built-in support for deciding when to heat and when to cool.
This support comes in the form of the rule attributes heat and cool as
well as the specially named function isKResult.

Consider the following definition, which is equivalent to LESSON-13-C
(lesson-14-a.k):

module LESSON-14-A-SYNTAX
  imports UNSIGNED-INT-SYNTAX
  imports BOOL-SYNTAX

  syntax Exp ::= Int
               | Bool
               > left: Exp "+" Exp
               > left: Exp "&&" Exp
endmodule

module LESSON-14-A
  imports LESSON-14-A-SYNTAX
  imports INT
  imports BOOL

  rule <k> I1:Int + I2:Int => I1 +Int I2 ...</k>
  rule <k> B1:Bool && B2:Bool => B1 andBool B2 ...</k>

  syntax KItem ::= freezer1(Exp) | freezer2(Exp)
                 | freezer3(Exp) | freezer4(Exp)

  rule <k> E:Exp + HOLE:Exp => HOLE ~> freezer1(E) ...</k>
    requires isKResult(E) [heat]
  rule <k> HOLE:Exp + E:Exp => HOLE ~> freezer2(E) ...</k> [heat]
  rule <k> E:Exp && HOLE:Exp => HOLE ~> freezer3(E) ...</k>
    requires isKResult(E) [heat]
  rule <k> HOLE:Exp && E:Exp => HOLE ~> freezer4(E) ...</k> [heat]

  rule <k> HOLE:Exp ~> freezer1(E) => E + HOLE ...</k> [cool]
  rule <k> HOLE:Exp ~> freezer2(E) => HOLE + E ...</k> [cool]
  rule <k> HOLE:Exp ~> freezer3(E) => E && HOLE ...</k> [cool]
  rule <k> HOLE:Exp ~> freezer4(E) => HOLE && E ...</k> [cool]

  syntax Bool ::= isKResult(K) [function, symbol]
  rule isKResult(_:Int) => true
  rule isKResult(_:Bool) => true
  rule isKResult(_) => false [owise]
endmodule

We have introduced three major changes to this definition. First, we have
removed the Val sort. We replace it instead with a function isKResult.
The function in question must have the same signature and attributes as seen in
this example. It ought to return true whenever a term should not be heated
(because it is a value) and false when it should be heated (because it is not
a value). We thus also insert isKResult calls in the side condition of two
of the heating rules, where the Val sort was previously used.

Second, we have removed the rule priorities on the heating rules and the use of
the Val sort on the cooling rules, and replaced them with the heat and
cool attributes. These attributes instruct the compiler that these rules are
heating and cooling rules, and thus should implicitly apply only when certain
terms on the LHS either are or are not a KResult (i.e., isKResult returns
true versus false).

Third, we have renamed some of the variables in the heating and cooling rules
to the special variable HOLE. Syntactically, HOLE is just a special name
for a variable, but it is treated specially by the compiler. By naming a
variable HOLE, we have informed the compiler which term is being heated
or cooled. The compiler will automatically insert the side condition
requires isKResult(HOLE) to cooling rules and the side condition
requires notBool isKResult(HOLE) to heating rules.

Exercise

Modify LESSON-14-A to add rules to evaluate integer subtraction.

Simplifying further with Contexts

The above example is still rather cumbersome to write. We must explicitly write
both the heating and the cooling rule separately, even though they are
essentially inverses of one another. It would be nice to instead simply
indicate which terms should be heated and cooled, and what part of them to
operate on.

To do this, K introduces a new type of sentence, the context. Contexts
begin with the context keyword instead of the rule keyword, and usually
do not contain a rewrite operator.

Consider the following definition which is equivalent to LESSON-14-A
(lesson-14-b.k):

module LESSON-14-B-SYNTAX
  imports UNSIGNED-INT-SYNTAX
  imports BOOL-SYNTAX

  syntax Exp ::= Int
               | Bool
               > left: Exp "+" Exp
               > left: Exp "&&" Exp
endmodule

module LESSON-14-B
  imports LESSON-14-B-SYNTAX
  imports INT
  imports BOOL

  rule <k> I1:Int + I2:Int => I1 +Int I2 ...</k>
  rule <k> B1:Bool && B2:Bool => B1 andBool B2 ...</k>

  context <k> E:Exp + HOLE:Exp ...</k>
    requires isKResult(E)
  context <k> HOLE:Exp + _:Exp ...</k>
  context <k> E:Exp && HOLE:Exp ...</k>
    requires isKResult(E)
  context <k> HOLE:Exp && _:Exp ...</k>

  syntax Bool ::= isKResult(K) [function, symbol]
  rule isKResult(_:Int) => true
  rule isKResult(_:Bool) => true
  rule isKResult(_) => false [owise]
endmodule

In this example, the heat and cool rules have been removed entirely, as
have been the productions defining the freezers. Don't worry, they still exist
under the hood; the compiler is just generating them automatically. For each
context sentence like above, the compiler generates a #freezer production,
a heat rule, and a cool rule. The generated form is equivalent to the
rules we wrote manually in LESSON-14-A. However, we are now starting to
considerably simplify the definition. Instead of 3 sentences, we just have one.

context alias sentences and the strict and seqstrict attributes

Notice that the contexts we included in LESSON-14-B still seem rather
similar in form. For each expression we want to evaluate, we are declaring
one context for each operand of that expression, and they are each rather
similar to one another. We would like to be able to simplify further by
simply annotating each expression production with information about how
it is to be evaluated instead. We can do this with the seqstrict attribute.

Consider the following definition, once again equivalent to those above
(lesson-14-c.k):

module LESSON-14-C-SYNTAX
  imports UNSIGNED-INT-SYNTAX
  imports BOOL-SYNTAX

  syntax Exp ::= Int
               | Bool
               > left: Exp "+" Exp [seqstrict(exp; 1, 2)]
               > left: Exp "&&" Exp [seqstrict(exp; 1, 2)]
endmodule

module LESSON-14-C
  imports LESSON-14-C-SYNTAX
  imports INT
  imports BOOL

  rule <k> I1:Int + I2:Int => I1 +Int I2 ...</k>
  rule <k> B1:Bool && B2:Bool => B1 andBool B2 ...</k>

  context alias [exp]: <k> HERE ...</k>

  syntax Bool ::= isKResult(K) [function, symbol]
  rule isKResult(_:Int) => true
  rule isKResult(_:Bool) => true
  rule isKResult(_) => false [owise]
endmodule

This definition has two important changes from the one above. The first is
that the individual context sentences have been removed and have been
replaced with a single context alias sentence. You may notice that this
sentence begins with an identifier in square brackets followed by a colon. This
syntax is a way of naming individual sentences in K for reference by the tool
or by other sentences. The context alias sentence also has a special variable
HERE.

The second is that the productions in LESSON-14-C-SYNTAX have been given a
seqstrict attribute. The value of this attribute has two parts. The first
is the name of a context alias sentence. The second is a comma-separated list
of integers. Each integer represents an index of a non-terminal in the
production, counting from 1. For each integer present, the compiler implicitly
generates a new context sentence according to the following rules:

  1. The compiler starts by looking for the context alias sentence named. If
    there is more than one, then one context sentence is created per
    context alias sentence with that name.
  2. For each context created, the variable HERE in the context alias is
    substituted with an instance of the production the seqstrict attribute is
    attached to. Each child of that production is a variable. The non-terminal
    indicated by the integer offset of the seqstrict attribute is given the name
    HOLE.
  3. For each integer offset prior in the list to the one currently being
    processed, the predicate isKResult(E) is conjuncted together and included
    as a side condition, where E is the child of the production term with that
    offset, starting from 1. For example, if the attribute lists 1, 2, then
    the rule generated for the 2 will include isKResult(E1) where E1 is the
    first child of the production.

As you can see if you work through the process, the above code will ultimately
generate the same contexts present in LESSON-14-B.

Finally, note that there are a few minor syntactic conveniences provided by the
seqstrict attribute. First, in the special case of the context alias sentence
being <k> HERE ...</k>, you can omit both the context alias sentence
and the name from the seqstrict attribute.

Second, if the numbered list of offsets contains every non-terminal in the
production, it can be omitted from the attribute value.

Thus, we can finally produce the idiomatic K definition for this example
(lesson-14-d.k):

module LESSON-14-D-SYNTAX
  imports UNSIGNED-INT-SYNTAX
  imports BOOL-SYNTAX

  syntax Exp ::= Int
               | Bool
               > left: Exp "+" Exp [seqstrict]
               > left: Exp "&&" Exp [seqstrict]
endmodule

module LESSON-14-D
  imports LESSON-14-D-SYNTAX
  imports INT
  imports BOOL

  rule <k> I1:Int + I2:Int => I1 +Int I2 ...</k>
  rule <k> B1:Bool && B2:Bool => B1 andBool B2 ...</k>

  syntax Bool ::= isKResult(K) [function, symbol]
  rule isKResult(_:Int) => true
  rule isKResult(_:Bool) => true
  rule isKResult(_) => false [owise]
endmodule

Exercise

Modify LESSON-14-D to add a production and rule to evaluate integer
subtraction.

Nondeterministic evaluation order with the strict attribute

Thus far, we have focused entirely on deterministic evaluation order. However,
not all languages are deterministic in the order they evaluate expressions.
For example, in C, the expression a() + b() + c() is guaranteed to parse
to (a() + b()) + c(), but it is not guaranteed that a will be called before
b before c. In fact, this evaluation order is non-deterministic.

We can express non-deterministic evaluation orders with the strict attribute.
Its behavior is identical to the seqstrict attribute, except that step 3 in
the above list (with the side condition automatically added) does not take
place. In other words, if we wrote syntax Exp ::= Exp "+" Exp [strict]
instead of syntax Exp ::= Exp "+" Exp [seqstrict], it would generate the
following two contexts instead of the ones found in LESSON-14-B:

  context <k> _:Exp + HOLE:Exp ...</k>
  context <k> HOLE:Exp + _:Exp ...</k>

As you can see, these contexts will generate heating rules that can both
apply to the same term. As a result, the choice of which heating rule
applies first is non-deterministic, and as we saw in Lesson 1.13, we can
get all possible behaviors by passing --search to krun.

Exercises

  1. Add integer division to LESSON-14-D. Make division and addition strict
    instead of seqstrict, and write a rule evaluating integer division with a
    side condition that the denominator is non-zero. Run krun --search on the
    program 1 / 0 + 2 / 1 and observe all possible outputs of the program. How
    many are there total, and why?

  2. Rework your solution from Lesson 1.9, Exercise 2 to evaluate expressions from left to right using the seqstrict attribute.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.15: Configuration Declarations and Cell Nesting.

Lesson 1.15: Configuration Declarations and Cell Nesting

The purpose of this lesson is to explain how to store additional information
about the state of your interpreter by declaring cells using the
configuration sentence, as well as how to add additional inputs to your
definition.

Cells and Configuration Declarations

We have already covered the absolute basics of cells in K by looking at the
<k> cell. As explained in Lesson 1.13, the
<k> cell is available without being explicitly declared. It turns out this is
because, if the user does not explicitly specify a configuration sentence
anywhere in the main module of their definition, the configuration sentence
from the DEFAULT-CONFIGURATION module of
kast.md is imported
automatically. Here is what that sentence looks like:

  configuration <k> $PGM:K </k>

This configuration declaration declares a single cell, the <k> cell. It also
declares that at the start of rewriting, the contents of that cell should be
initialized with the value of the $PGM configuration variable.
Configuration variables function as inputs to krun. These terms are supplied
to krun in the form of ASTs parsed using a particular module. By default, the
$PGM configuration variable uses the main syntax module of the definition.

The cast on the configuration variable also specifies the sort that is used as
the entry point to the parser, in this case the K sort. It is often
useful to cast to other sorts there as well for better control over the accepted
language. The sort used for the $PGM variable is referred to as the start
symbol. During parsing, the default start symbol K subsumes all user-defined
sorts except for syntactic lists. These are excluded because they will always
produce an ambiguity error when parsing a single element.

Note that we did not explicitly specify the $PGM configuration variable when
we invoked krun on a file. This is because krun handles the $PGM variable
specially, and allows you to pass the term for that variable via a file passed
as a positional argument to krun. We did, however, specify the PGM name
explicitly when we called krun with the -cPGM command line argument in
Lesson 1.2. This is the other, explicit, way of
specifying an input to krun.

This explains the most basic use of configuration declarations in K. We can,
however, declare multiple cells and multiple configuration variables. We can
also specify the initial values of cells statically, rather than dynamically
via krun.

For example, consider the following definition (lesson-15-a.k):

module LESSON-15-A-SYNTAX
  imports INT-SYNTAX

  syntax Ints ::= List{Int,","}
endmodule

module LESSON-15-A
  imports LESSON-15-A-SYNTAX
  imports INT

  configuration <k> $PGM:Ints </k>
                <sum> 0 </sum>

  rule <k> I:Int, Is:Ints => Is ...</k>
       <sum> SUM:Int => SUM +Int I </sum>
endmodule

This simple definition takes a list of integers as input and sums them
together. Here we have declared two cells: <k> and <sum>. Unlike <k>,
<sum> does not get initialized via a configuration variable, but instead
is initialized statically with the value 0.

Note the rule in the second module: we have explicitly specified multiple
cells in a single rule. K will expect each of these cells to match in order for
the rule to apply.

Here is a second example (lesson-15-b.k):

module LESSON-15-B-SYNTAX
  imports INT-SYNTAX
endmodule

module LESSON-15-B
  imports LESSON-15-B-SYNTAX
  imports INT
  imports BOOL

  configuration <k> . </k>
                <first> $FIRST:Int </first>
                <second> $SECOND:Int </second>

  rule <k> . => FIRST >Int SECOND </k>
       <first> FIRST </first>
       <second> SECOND </second>
endmodule

This definition takes two integers as command-line arguments and populates the
<k> cell with a Boolean indicating whether the first integer is greater than
the second. Notice that we have specified no $PGM configuration variable
here. As a result, we cannot invoke krun via the syntax krun $file.
Instead, we must explicitly pass values for each configuration variable via the
-cFIRST and -cSECOND command line flags. For example, if we invoke
krun -cFIRST=0 -cSECOND=1, we will get the value false in the K cell.

You can also specify both a $PGM configuration variable and other
configuration variables in a single configuration declaration, in which case
you would be able to initialize $PGM with either a positional argument or the
-cPGM command line flag, but the other configuration variables would need
to be explicitly initialized with -c.

Exercise

Modify your solution to Lesson 1.14, Exercise 2 to add a new cell with a
configuration variable of sort Bool. This variable should determine whether
the / operator is evaluated using /Int or divInt. Test that by specifying
different values for this variable, you can change the behavior of rounding on
division of negative numbers.

Cell Nesting

It is possible to nest cells inside one another. A cell that contains other
cells must contain only other cells, but in doing this, you are able to
create a hierarchical structure to the configuration. Consider the following
definition (lesson-15-c.k), which is equivalent to the one in LESSON-15-B:

module LESSON-15-C-SYNTAX
  imports INT-SYNTAX
endmodule

module LESSON-15-C
  imports LESSON-15-C-SYNTAX
  imports INT
  imports BOOL

  configuration <T>
                  <k> . </k>
                  <state>
                    <first> $FIRST:Int </first>
                    <second> $SECOND:Int </second>
                  </state>
                </T>

  rule <k> . => FIRST >Int SECOND </k>
       <first> FIRST </first>
       <second> SECOND </second>
endmodule

Note that we have added some new cells to the configuration declaration:
the <T> cell wraps the entire configuration, and the <state> cell is
introduced around the <first> and <second> cells.

However, we have not changed the rule in this definition. This is because of
a concept in K called configuration abstraction. K allows you to specify
any number of cells in a rule (except zero) in any order you want, and K will
compile the rules into a form that matches the structure of the configuration
specified by the configuration declaration.

Here then, is how this rule would look after the configuration abstraction
has been resolved:

  rule <T>
         <k> . => FIRST >Int SECOND </k>
         <state>
           <first> FIRST </first>
           <second> SECOND </second>
         </state>
       </T>

In other words, K will complete cells to the top of the configuration by
inserting parent cells where appropriate based on the declared structure of
the configuration. This is useful because as a definition evolves, the
configuration may change, but you don't want to have to modify every single
rule each time. Thus, K follows the principle that you should only mention the
cells in a rule that are actually needed in order to accomplish its specific
goal. By following this best practice, you can significantly increase the
modularity of the definition and make it easier to maintain and modify.

Note that unlike top-level rewrite rules, cells that appear inside function
rules are not necessarily completed to the top of the configuration. They still
participate in cell ccompletion in the sense that you can mention cell
structure loosely inside a function rule and it will be completed into the
correct cell structure specified by the configuration declaration. However,
they do not complete all the way to the top, instead completing only up to
the top-most cell mentioned in the rule.

For example, if I write the following function rule in the above definition:

  rule doStuff(<first> FIRST </first>) => FIRST

The function will only match on the first cell, rather than the entire
configuration. However, if we had mentioned a parent cell in the rule, it still
would have completed the children of that parent cell as needed to ensure that
the resulting term is well formed.

Exercise

Modify your definition from the previous exercise in this lesson to wrap the
two cells you have declared in a top cell <T>. You should not have to change
any other rules in the definition.

Cell Variables

Sometimes it is desirable to explicitly match a variable against certain
fragments of the configuration. Because K's configuration is hierarchical,
we can grab subsets of the configuration as if they were just another term.
However, configuration abstraction applies here as well.
In particular, for each cell you specify in a configuration declaration, a
unique sort is assigned for that cell with a single constructor (the cell
itself). The sort name is taken by removing all special characters,
capitalizing the first letter and each letter after a hyphen, and adding the
word Cell at the end. For example, in the above example, the cell sorts are
TCell, KCell, StateCell, FirstCell, and SecondCell. If we had declared
a cell as <first-number>, then the cell sort name would be FirstNumberCell.

You can explicitly reference a variable of one of these sorts anywhere you
might instead write that cell. For example, consider the following rule:

  rule <k> true => S </k>
       (S:StateCell => <state>... .Bag ...</state>)

Here we have introduced two new concepts. The first is the variable of sort
StateCell, which matches the entire <state> part of the configuration. The
second is that we have introduced the concept of ... once again. When a cell
contains other cells, it is also possible to specify ... on either the left,
right or both sides of the cell term. Each of these three syntaxes are
equivalent in this case. When they appear on the left-hand side of a rule, they
indicate that we don't care what value any cells not explicitly named might
have. For example, we might write <state>... <first> 0 </first> ...</state> on
the left-hand side of a rule in order to indicate that we want to match the
rule when the <first> cell contains a zero, regardless of what the <second>
cell contains. If we had not included this ellipsis, it would have been a
syntax error, because K would have expected you to provide a value for each of
the child cells.

However, if, as in the example above, the ... appeared on the right-hand side
of a rule, this instead indicates that the cells not explicitly mentioned under
the cell should be initialized with their default value from the configuration
declaration. In other words, that rule will set the value of <first> and
<second> to zero.

You may note the presence of the phrase .Bag here. You can think of this as
the empty set of cells. It is used as the child of a cell when you want to
indicate that no cells should be explicitly named. We will cover other uses
of this term in later lessons.

Exercises

  1. Modify the definition from the previous exercise in this lesson so that the
    Boolean cell you created is initialized to false. Then add a production
    syntax Stmt ::= Bool ";" Exp, and a rule that uses this Stmt to set the
    value of the Boolean flag. Then add another production
    syntax Stmt ::= "reset" ";" Exp which sets the value of the Boolean flag back
    to its default value via a ... on the right-hand side. You will need to add
    an additional cell around the Boolean cell to make this work.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.16: Maps, Semantic Lists, and Sets.

Lesson 1.16: Maps, Semantic Lists, and Sets

The purpose of this lesson is to explain how to use the data structure sorts
provided by K: maps, lists, and sets.

Maps

The most frequently used type of data structure in K is the map. The sort
provided by K for this purpose is the Map sort, and it is provided in
domains.md in the MAP
module. This type is not (currently) polymorphic. All Map terms are maps that
map terms of sort KItem to other terms of sort KItem. A KItem can contain
any sort except a K sequence. If you need to store such a term in a
map, you can always use a wrapper such as syntax KItem ::= kseq(K).

A Map pattern consists of zero or more map elements (as represented by the
symbol syntax Map ::= KItem "|->" KItem), mixed in any order, separated by
whitespace, with zero or one variables of sort Map. The empty map is
represented by .Map. If all of the bindings for the variables in the keys
of the map can be deterministically chosen, these patterns can be matched in
O(1) time. If they cannot, then each map element that cannot be
deterministically constructed contributes a single dimension of polynomial
time to the cost of the matching. In other words, a single such element is
linear, two are quadratic, three are cubic, etc.

Patterns like the above are the only type of Map pattern that can appear
on the left-hand-side of a rule. In other words, you are not allowed to write
a Map pattern on the left-hand-side with more than one variable of sort Map
in it. You are, however, allowed to write such patterns on the right-hand-side
of a rule. You can also write a function pattern in the key of a map element
so long as all the variables in the function pattern can be deterministically
chosen.

Note the meaning of matching on a Map pattern: a map pattern with no
variables of sort Map will match if the map being matched has exactly as
many bindings as |-> symbols in the pattern. It will then match if each
binding in the map pattern matches exactly one distinct binding in the map
being matched. A map pattern with one Map variable will also match any map
that contains such a map as a subset. The variable of sort Map will be bound
to whatever bindings are left over (.Map if there are no bindings left over).

Here is an example of a simple definition that implements a very basic
variable declaration semantics using a Map to store the value of variables
(lesson-16-a.k):

module LESSON-16-A-SYNTAX
  imports INT-SYNTAX
  imports ID-SYNTAX

  syntax Exp ::= Id | Int
  syntax Decl ::= "int" Id "=" Exp ";" [strict(2)]
  syntax Pgm ::= List{Decl,""}
endmodule

module LESSON-16-A
  imports LESSON-16-A-SYNTAX
  imports BOOL

  configuration <T>
                  <k> $PGM:Pgm </k>
                  <state> .Map </state>
                </T>

  // declaration sequence
  rule <k> D:Decl P:Pgm => D ~> P ...</k>
  rule <k> .Pgm => . ...</k>

  // variable declaration
  rule <k> int X:Id = I:Int ; => . ...</k>
       <state> STATE => STATE [ X <- I ] </state>

  // variable lookup
  rule <k> X:Id => I ...</k>
       <state>... X |-> I ...</state>

  syntax Bool ::= isKResult(K) [symbol, function]
  rule isKResult(_:Int) => true
  rule isKResult(_) => false [owise]
endmodule

There are several new features in this definition. First, note we import
the module ID-SYNTAX. This module is defined in domains.md and provides a
basic syntax for identifiers. We are using the Id sort provided by this
module in this definition to implement the names of program variables. This
syntax is only imported when parsing programs, not when parsing rules. Later in
this lesson we will see how to reference specific concrete identifiers in a
rule.

Second, we introduce a single new function over the Map sort. This function,
which is represented by the symbol
syntax Map ::= Map "[" KItem "<-" KItem "]", represents the map update
operation. Other functions over the Map sort can be found in domains.md.

Finally, we have used the ... syntax on a cell containing a Map. In this
case, the meaning of <state>... Pattern ...</state>,
<state>... Pattern </state>, and <state> Pattern ...</state> are the same:
it is equivalent to writing <state> (Pattern) _:Map </state>.

Consider the following program (a.decl):

int x = 0;
int y = 1;
int a = x;

If we run this program with krun, we will get the following result:

<T>
  <k>
    .
  </k>
  <state>
    a |-> 0
    x |-> 0
    y |-> 1
  </state>
</T>

Note that krun has automatically sorted the collection for you. This doesn't
happen at runtime, so you still get the performance of a hash map, but it will
help make the output more readable.

Exercise

Create a sort Stmt that is a subsort of Decl. Create a production of sort
Stmt for variable assignment in addition to the variable declaration
production. Feel free to use the syntax syntax Stmt ::= Id "=" Exp ";". Write
a rule that implements variable assignment using a map update function. Then
write the same rule using a map pattern. Test your implementations with some
programs to ensure they behave as expected.

Semantic Lists

In a previous lesson, we explained how to represent lists in the AST of a
program. However, this is not the only context where lists can be used. We also
frequently use lists in the configuration of an interpreter in order to
represent certain types of program state. For this purpose, it is generally
useful to have an associative-list sort, rather than the cons-list sorts
provided in Lesson 1.12.

The type provided by K for this purpose is the List sort, and it is also
provided in domains.md, in the LIST module. This type is also not
(currently) polymorphic. Like Map, all List terms are lists of terms of the
KItem sort.

A List pattern in K consists of zero or more list elements (as represented by
the ListItem symbol), followed by zero or one variables of sort List,
followed by zero or more list elements. An empty list is represented by
.List. These patterns can be matched in O(log(N)) time. This is the only
type of List pattern that can appear on the left-hand-side of a rule. In
other words, you are not allowed to write a List pattern on the
left-hand-side with more than one variable of sort List in it. You are,
however, allowed to write such patterns on the right-hand-side of a rule.

Note the meaning of matching on a List pattern: a list pattern with no
variables of sort List will match if the list being matched has exactly as
many elements as ListItem symbols in the pattern. It will then match if each
element in sequence matches the pattern contained in the ListItem symbol. A
list pattern with one variable of sort List operates the same way, except
that it can match any list with at least as many elements as ListItem
symbols, so long as the prefix and suffix of the list match the patterns inside
the ListItem symbols. The variable of sort List will be bound to whatever
elements are left over (.List if there are no elements left over).

The ... syntax is allowed on cells containing lists as well. In this case,
the meaning of <cell>... Pattern </cell> is the same as
<cell> _:List (Pattern) </cell>, the meaning of <cell> Pattern ...</cell>
is the same as <cell> (Pattern) _:List</cell>. Because list patterns with
multiple variables of sort List are not allowed, it is an error to write
<cell>... Pattern ...</cell>.

Here is an example of a simple definition that implements a very basic
function-call semantics using a List as a function stack (lesson-16-b.k):

module LESSON-16-B-SYNTAX
  imports INT-SYNTAX
  imports ID-SYNTAX

  syntax Exp ::= Id "(" ")" | Int
  syntax Stmt ::= "return" Exp ";" [strict]
  syntax Decl ::= "fun" Id "(" ")" "{" Stmt "}"
  syntax Pgm ::= List{Decl,""}
  syntax Id ::= "main" [token]
endmodule

module LESSON-16-B
  imports LESSON-16-B-SYNTAX
  imports BOOL
  imports LIST

  configuration <T>
                  <k> $PGM:Pgm ~> main () </k>
                  <functions> .Map </functions>
                  <fstack> .List </fstack>
                </T>

  // declaration sequence
  rule <k> D:Decl P:Pgm => D ~> P ...</k>
  rule <k> .Pgm => . ...</k>

  // function definitions
  rule <k> fun X:Id () { S } => . ...</k>
       <functions>... .Map => X |-> S ...</functions>

  // function call
  syntax KItem ::= stackFrame(K)
  rule <k> X:Id () ~> K => S </k>
       <functions>... X |-> S ...</functions>
       <fstack> .List => ListItem(stackFrame(K)) ...</fstack>

  // return statement
  rule <k> return I:Int ; ~> _ => I ~> K </k>
       <fstack> ListItem(stackFrame(K)) => .List ...</fstack>

  syntax Bool ::= isKResult(K) [function, symbol]
  rule isKResult(_:Int) => true
  rule isKResult(_) => false [owise]
endmodule

Notice that we have declared the production syntax Id ::= "main" [token].
Since we use the ID-SYNTAX module, this declaration is necessary in order to
be able to refer to the main identifier directly in the configuration
declaration. Our <k> cell now contains a K sequence initially: first we
process all the declarations in the program, then we call the main function.

Consider the following program (foo.func):

fun foo() { return 5; }
fun main() { return foo(); }

When we krun this program, we should get the following output:

<T>
  <k>
    5 ~> .
  </k>
  <functions>
    foo |-> return 5 ;
    main |-> return foo ( ) ;
  </functions>
  <fstack>
    .List
  </fstack>
</T>

Note that we have successfully put on the <k> cell the value returned by the
main function.

Exercise

Add a term of sort Id to the stackFrame operator to keep track of the
name of the function in that stack frame. Then write a function
syntax String ::= printStackTrace(List) that takes the contents of the
<fstack> cell and pretty prints the current stack trace. You can concatenate
strings with +String in the STRING module in domains.md, and you can
convert an Id to a String with the Id2String function in the ID module.
Test this function by creating a new expression that returns the current stack
trace as a string. Make sure to update isKResult and the Exp sort as
appropriate to allow strings as values.

Sets

The final primary data structure sort in K is a set, i.e., an idempotent
unordered collection where elements are deduplicated. The sort provided by K
for this purpose is the Set sort and it is provided in domains.md in the
SET module. Like maps and lists, this type is not (currently) polymorphic.
Like Map and List, all Set terms are sets of terms of the KItem sort.

A Set pattern has the exact same restrictions as a Map pattern, except that
its elements are treated like keys, and there are no values. It has the same
performance characteristics as well. However, syntactically it is more similar
to the List sort: An empty Set is represented by .Set, but a set element
is represented by the SetItem symbol.

Matching behaves similarly to the Map sort: a set pattern with no variables
of sort Set will match if the set has exactly as many bindings as SetItem
symbols, and if each element pattern matches one distinct element in the set.
A set with a variable of sort Set also matches any superset of such a set.
As with map, the elements left over will be bound to the Set variable (or
.Set if no elements are left over).

Like Map, the ... syntax on a set is syntactic sugar for an anonymous
variable of sort Set.

Here is an example of a simple modification to LESSON-16-A which uses a Set
to ensure that variables are never declared more than once. In practice, you
would likely just use the in_keys symbol over maps to test for this, but
it's still useful as an example of sets in practice:

module LESSON-16-C-SYNTAX
  imports LESSON-16-A-SYNTAX
endmodule

module LESSON-16-C
  imports LESSON-16-C-SYNTAX
  imports BOOL
  imports SET

  configuration <T>
                  <k> $PGM:Pgm </k>
                  <state> .Map </state>
                  <declared> .Set </declared>
                </T>

  // declaration sequence
  rule <k> D:Decl P:Pgm => D ~> P ...</k>
  rule <k> .Pgm => . ...</k>

  // variable declaration
  rule <k> int X:Id = I:Int ; => . ...</k>
       <state> STATE => STATE [ X <- I ] </state>
       <declared> D => D SetItem(X) </declared>
    requires notBool X in D

  // variable lookup
  rule <k> X:Id => I ...</k>
       <state>... X |-> I ...</state>
       <declared>... SetItem(X) ...</declared>

  syntax Bool ::= isKResult(K) [symbol, function]
  rule isKResult(_:Int) => true
  rule isKResult(_) => false [owise]
endmodule

Now if we krun a program containing duplicate declarations, it will get
stuck on the declaration.

Exercises

  1. Modify your solution to Lesson 1.14, Exercise 2 and introduce the sorts
    Decls, Decl, and Stmt which include variable and function declaration
    (without function parameters), and return and assignment statements, as well
    as call expressions. Use List and Map to implement these operators, making
    sure to consider the interactions between components, such as saving and
    restoring the environment of variables at each call site. Don't worry about
    local function definitions or global variables for now. Make sure to test the
    resulting interpreter.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.17: Cell Multiplicity and Cell Collections.

Lesson 1.17: Cell Multiplicity and Cell Collections

The purpose of this lesson is to explain how you can create optional cells
and cells that repeat multiple times in a configuration using a feature called
cell multiplicity.

Cell Multiplicity

K allows you to specify attributes for cell productions as part of the syntax
of configuration declarations. Unlike regular productions, which use the []
syntax for attributes, configuration cells use an XML-like attribute syntax:

configuration <k color="red"> $PGM:K </k>

This configuration declaration gives the <k> cell the color red during
unparsing using the color attribute as discussed in
Lesson 1.9.

However, in addition to the usual attributes for productions, there are some
other attributes that can be applied to cells with special meaning. One such
attribute is the multiplicity attribute. By default, each cell that is
declared occurs exactly once in every configuration term. However, using the
multiplicity attribute, this default behavior can be changed. There are two
values that this attribute can have: ? and *.

Optional cells

The first cell multiplicity we will discuss is ?. Similar to a regular
expression language, this attribute tells the compiler that this cell can
appear 0 or 1 times in the configuration. In other words, it is an
optional cell. By default, K does not create optional cells in the initial
configuration, unless that optional cell has a configuration variable inside
it. However, it is possible to override the default behavior and create that
cell initially by adding the additional cell attribute initial="".

K uses the .Bag symbol to represent the absence of any cells in a particular
rule. Consider the following module:

module LESSON-17-A
  imports INT

  configuration <k> $PGM:K </k>
                <optional multiplicity="?"> 0 </optional>

  syntax KItem ::= "init" | "destroy"

  rule <k> init => . ...</k>
       (.Bag => <optional> 0 </optional>)
  rule <k> destroy => . ...</k>
       (<optional> _ </optional> => .Bag)

endmodule

In this definition, when the init symbol is executed, the <optional> cell
is added to the configuration, and when the destroy symbol is executed, it
is removed. Any rule that matches on that cell will only match if that cell is
present in the configuration.

Exercise

Create a simple definition with a Stmts sort that is a List{Stmt,""} and
a Stmt sort with the constructors
syntax Stmt ::= "enable" | "increment" | "decrement" | "disable". The
configuration should have an optional cell that contains an integer that
is created with the enable command, destroyed with the disable command,
and its value is incremented or decremented by the increment and decrement
command.

Cell collections

The second type of cell multiplicity we will discuss is *. Simlar to a
regular expression language, this attribute tells the compiler that this cell
can appear 0 or more times in the configuration. In other words, it is a
cell collection. Cells with multiplicity * must be the only child of
their parent cell. As a convention, the inner cell is usually named with the
singular form of what it contains, and the outer cell with the plural form, for
example, "thread" and "threads".

All cell collections are required to have the type attribute set to either
Set or Map. A Set cell collection is represented as a set and behaves
internally the same as the Set sort, although it actually declares a new
sort. A Map cell collection is represented as a Map in which the first
subcell of the cell collection is the key and the remaining cells are the
value.

For example, consider the following module:

module LESSON-17-B
  imports INT
  imports BOOL
  imports ID-SYNTAX

  syntax Stmt ::= Id "=" Exp ";" [strict(2)]
                | "return" Exp ";" [strict]
  syntax Stmts ::= List{Stmt,""}
  syntax Exp ::= Id
               | Int
               | Exp "+" Exp [seqstrict]
               | "spawn" "{" Stmts "}"
               | "join" Exp ";" [strict]

  configuration <threads>
                  <thread multiplicity="*" type="Map">
                    <id> 0 </id>
                    <k> $PGM:K </k>
                  </thread>
                </threads>
                <state> .Map </state>
                <next-id> 1 </next-id>

  rule <k> X:Id => I:Int ...</k>
       <state>... X |-> I ...</state>
  rule <k> X:Id = I:Int ; => . ...</k>
       <state> STATE => STATE [ X <- I ] </state>
  rule <k> S:Stmt Ss:Stmts => S ~> Ss ...</k>
  rule <k> I1:Int + I2:Int => I1 +Int I2 ...</k>

  rule <thread>...
         <k> spawn { Ss } => NEXTID ...</k>
       ...</thread>
       <next-id> NEXTID => NEXTID +Int 1 </next-id>
       (.Bag =>
       <thread>
         <id> NEXTID </id>
         <k> Ss </k>
       </thread>)

  rule <thread>...
         <k> join ID:Int ; => I ...</k>
       ...</thread>
       (<thread>
         <id> ID </id>
         <k> return I:Int ; ...</k>
       </thread> => .Bag)

  syntax Bool ::= isKResult(K) [function, symbol]
  rule isKResult(_:Int) => true
  rule isKResult(_) => false [owise]
endmodule

This module implements a very basic fork/join semantics. The spawn expression
spawns a new thread to execute a sequence of statements and returns a thread
id, and the join statement waits until a thread executes return and then
returns the return value of the thread.

Note something quite novel here: the <k> cell is inside a cell of
multiplicity *. Since the <k> cell is just a regular cell (mostly), this
is perfectly allowable. Rules that don't mention a specific thread are
automatically completed to match any thread.

When you execute programs in this language, the cells in the cell collection
get sorted and printed like any other collection, but they still display like
cells. Rules in this language also benefit from all the structural power of
cells, allowing you to omit cells you don't care about or complete the
configuration automatically. This allows you to have the power of cells while
still being a collection under the hood.

Exercises

  1. Modify the solution from Lesson 1.16, Exercise 1 so that the cell you use to
    keep track of functions in a Map is now a cell collection. Run some programs
    and compare how they get unparsed before and after this change.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.18: Term Equality and the Ternary Operator.

Lesson 1.18: Term Equality and the Ternary Operator

The purpose of this lesson is to introduce how to compare equality of terms in
K, and how to put conditional expressions directly into the right-hand side of
rules.

Term Equality

One major way you can compare whether two terms are equal in K is to simply
match both terms with a variable with the same name. This will only succeed
in matching if the two terms are equal structurally. However, sometimes this
is impractical, and it is useful to have access to a way to actually compare
whether two terms in K are equal. The operator for this is found in
domains.md in the K-EQUAL
module. The operator is ==K and takes two terms of sort K and returns a
Bool. It returns true if they are equal. This includes equality over builtin
types such as Map and Set where equality is not purely structural in
nature. However, it does not include any notion of semantic equality over
user-defined syntax. The inverse symbol for inequality is =/=K.

Ternary Operator

One way to introduce conditional logic in K is to have two separate rules,
each with a side condition (or one rule with a side condition and another with
the owise attribute). However, sometimes it is useful to explicitly write
a conditional expression directly in the right-hand side of a rule. For this
purpose, K defines one more operator in the K-EQUAL module, which corresponds
to the usual ternary operator found in many languages. Here is an example of its
usage (lesson-18.k):

module LESSON-18
  imports INT
  imports BOOL
  imports K-EQUAL

  syntax Exp ::= Int | Bool | "if" "(" Exp ")" Exp "else" Exp [strict(1)]

  syntax Bool ::= isKResult(K) [function, symbol]
  rule isKResult(_:Int) => true
  rule isKResult(_:Bool) => true

  rule if (B:Bool) E1:Exp else E2:Exp => #if B #then E1 #else E2 #fi
endmodule

Note the symbol on the right-hand side of the final rule. This symbol is
polymorphic: B must be of sort Bool, but E1 and E2 could have been
any sort so long as both were of the same sort, and the sort of the entire
expression becomes equal to that sort. K supports polymorphic built-in
operators, but does not yet allow users to write their own polymorphic
productions.

The behavior of this function is to evaluate the Boolean expression to a
Boolean, then pick one of the two children and return it based on whether the
Boolean is true or false. Please note that it is not a good idea to use this
symbol in cases where one or both of the children is potentially undefined
(for example, an integer expression that divides by zero). While the default
implementation is smart enough to only evaluate the branch that happens to be
picked, this will not be true when we begin to do program verification. If
you need short circuiting behavior, it is better to use a side condition.

Exercises

  1. Write a function in K that takes two terms of sort K and returns an
    Int: the Int should be 0 if the terms are equal and 1 if the terms are
    unequal.

  2. Modify your solution to Lesson 1.16, Exercise 1 and introduce an if
    Stmt to the syntax of the language, then implement it using the #if symbol.
    Make sure to write tests for the resulting interpreter.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.19: Debugging with GDB.

Lesson 1.19: Debugging with GDB or LLDB

The purpose of this lesson is to teach how to debug your K interpreter using
the K-language support provided in GDB or
LLDB.

Caveats

This lesson has been written with GDB support on Linux in mind. Unfortunately,
on macOS, GDB has limited support. To address this, we have introduced early
experimental support for debugging with LLDB on macOS. In some cases, the
features supported by LLDB are slightly different to those supported by GDB; the
tutorial text will make this clear where necessary. If you use a macOS with an
LLVM version older than 15, you may need to upgrade it to use the LLDB
correctly. If you encounter an issue on either operating system, please open an
issue against the K repository.

Getting started

On Linux, you will need GDB in order to complete this lesson. If you do not
already have GDB installed, then do so. Steps to install GDB are outlined in
this GDB Tutorial.

On macOS, LLDB should already have been installed with K's build dependencies
(whether you have built K from source, or installed it using kup or Homebrew).

The first thing neccessary in order to debug a K interpreter is to build the
interpreter with full debugging support enabled. This can be done relatively
simply. First, run kompile with the command line flag --enable-llvm-debug.
The resulting compiled K definition will be ready to support debugging.

Once you have a compiled K definition and a program you wish to debug, you can
start the debugger by passing the --debugger flag to krun. This will
automatically load the program you are executing into GDB and drop you into a
GDB shell ready to start executing the program.

As an example, consider the following K definition (lesson-19-a.k):

module LESSON-19-A
  imports INT

  rule I => I +Int 1
    requires I <Int 100
endmodule

If we compile this definition with kompile lesson-19-a.k --enable-llvm-debug,
and run the program 0 in the debugger with krun -cPGM=0 --debugger, we will
see the following output (roughly, and depending on which platform you are
using):

GDB / Linux

GNU gdb (Ubuntu 9.2-0ubuntu1~20.04) 9.2
Copyright (C) 2020 Free Software Foundation, Inc.
License GPLv3+: GNU GPL version 3 or later <http://gnu.org/licenses/gpl.html>
This is free software: you are free to change and redistribute it.
There is NO WARRANTY, to the extent permitted by law.
Type "show copying" and "show warranty" for details.
This GDB was configured as "x86_64-linux-gnu".
Type "show configuration" for configuration details.
For bug reporting instructions, please see:
<http://www.gnu.org/software/gdb/bugs/>.
Find the GDB manual and other documentation resources online at:
    <http://www.gnu.org/software/gdb/documentation/>.

For help, type "help".
Type "apropos word" to search for commands related to "word"...
Reading symbols from ./lesson-19-a-kompiled/interpreter...
warning: File "/home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-a-kompiled/interpreter" auto-loading has been declined by your `auto-load safe-path' set to "$debugdir:$datadir/auto-load".
To enable execution of this file add
        add-auto-load-safe-path /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-a-kompiled/interpreter
line to your configuration file "/home/dwightguth/.gdbinit".
To completely disable this security protection add
        set auto-load safe-path /
line to your configuration file "/home/dwightguth/.gdbinit".
For more information about this security protection see the
"Auto-loading safe path" section in the GDB manual.  E.g., run from the shell:
        info "(gdb)Auto-loading safe path"
(gdb)

To make full advantage of the GDB features of K, you should follow the first
command listed in this output message and add the corresponding
add-auto-load-safe-path command to your ~/.gdbinit file as prompted.
Please note that the path will be different on your machine than the one
listed above. Adding directories to the "load safe path" effectively tells GDB
to trust those directories. All content under a given directory will be recursively
trusted, so if you want to avoid having to add paths to the "load safe path" every
time you kompile a different K definition, then you can just trust a minimal
directory containing all your kompiled files; however, do not choose a top-level directory containing arbitrary files as this amounts to trusting arbitrary files and is a security risk. More info on the load safe path
can be found here.

LLDB / macOS

(lldb) target create "./lesson-19-a-kompiled/interpreter"
warning: 'interpreter' contains a debug script. To run this script in this debug session:

    command script import "/Users/brucecollie/code/scratch/lesson-19-a-kompiled/interpreter.dSYM/Contents/Resources/Python/interpreter.py"

To run all discovered debug scripts in this session:

    settings set target.load-script-from-symbol-file true

Current executable set to '/Users/brucecollie/code/scratch/lesson-19-a-kompiled/interpreter' (x86_64).
(lldb) settings set -- target.run-args  ".krun-2023-03-20-11-22-46-TcYt9ffhb2/tmp.in.RupiLwHNfn" "-1" ".krun-2023-03-20-11-22-46-TcYt9ffhb2/result.kore"
(lldb) 

LLDB applies slightly different security policies to GDB. To load K's debugging
scripts for this session only, you can run the command script import line at
the LLDB prompt. The loaded scripts will not persist across debugging sessions
if you do this. It is also possible to configure LLDB to automatically load the
K scripts when an interpreter is started in LLDB; doing so requires a slightly
less broad permission than GDB.

On macOS, the .dSYM directory that contains debugging symbols for an
executable can also contain Python scripts in Contents/Resources/Python. If
there is a Python script with a name matching the name of the current executable
(here, interpreter and interpreter.py), it will be automatically loaded if
the target.load-script-from-symbol-file setting is set). You can therefore add
the settings set command to your ~/.lldbinit without enabling full arbitrary
code execution, but you should be aware of the paths from which code can be
executed if you do so.

Basic commands

LLDB Note: the k start and k step commands are currently not
implemented in the K LLDB scripts. To work around this limitation temporarily,
you can run process launch --stop-at-entry instead of k start. To emulate
k step, first run rbreak k_step once, then continue instead of each k step. We hope to address these limitations soon.

The most basic commands you can execute in the K GDB session are to run your
program or to step through it. The first can be accomplished using GDB's
built-in run command. This will automatically start the program and begin
executing it. It will continue until the program aborts or finishes, or the
debugger is interrupted with Ctrl-C.

Sometimes you want finer-grained control over how you proceed through the
program you are debugging. To step through the rule applications in your
program, you can use the k start and k step GDB commands.

k start is similar to the built-in start command in that it starts the
program and then immediately breaks before doing any work. However, unlike
the start command which will break immediately after the main method of
a program is executed, the K start program will initialize the rewriter,
evaluate the initial configuration, and break immediately prior to applying
any rewrite steps.

In the example above, here is what we see when we run the k start command:

Temporary breakpoint 1 at 0x239210
Starting program: /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-a-kompiled/interpreter .krun-2021-08-13-14-10-50-sMwBkbRicw/tmp.in.01aQt85TaA -1 .krun-2021-08-13-14-10-50-sMwBkbRicw/result.kore
[Thread debugging using libthread_db enabled]
Using host libthread_db library "/lib/x86_64-linux-gnu/libthread_db.so.1".

Temporary breakpoint 1, 0x0000000000239210 in main ()
0x0000000000231890 in step (subject=<k>
  0 ~> .
</k>)
(gdb)

As you can see, we are stopped at the step function in the interpreter.
This function is responsible for taking top-level rewrite steps. The subject
parameter to this function is the current K configuration.

We can step through K rewrite steps one at a time by running the k step
command. By default, this takes a single rewrite step (including any function
rule applications that are part of that step).

Here is what we see when we run that command:

Continuing.

Temporary breakpoint -22, 0x0000000000231890 in step (subject=<k>
  1 ~> .
</k>)
(gdb)

As we can see, we have taken a single rewrite step. We can also pass a number
to the k step command which indicates the number of rewrite steps to take.

Here is what we see if we run k step 10:

Continuing.

Temporary breakpoint -23, 0x0000000000231890 in step (subject=<k>
  11 ~> .
</k>)
(gdb)

As we can see, ten rewrite steps were taken.

Breakpoints

The next important step in debugging an application in GDB is to be able to
set breakpoints. Generally speaking, there are three types of breakpoints we
are interested in a K semantics: Setting a breakpoint when a particular
function is called, setting a breakpoint when a particular rule is applied,
and setting a breakpoint when a side condition of a rule is evaluated.

The easiest way to do the first two things is to set a breakpoint on the
line of code containing the function or rule.

For example, consider the following K definition (lesson-19-b.k):

module LESSON-19-B
  imports BOOL

  syntax Bool ::= isBlue(Fruit) [function]
  syntax Fruit ::= Blueberry() | Banana()
  rule isBlue(Blueberry()) => true
  rule isBlue(Banana()) => false

  rule F:Fruit => isBlue(F)
endmodule

Once this program has been compiled for debugging, we can run the program
Blueberry(). We can then set a breakpoint that stops when the isBlue
function is called with the following command in GDB:

break lesson-19-b.k:4

Similarly, in LLDB, run:

breakpoint set --file lesson-19-b.k --line 4

Here is what we see if we set this breakpoint and then run the interpreter:

(gdb) break lesson-19-b.k:4
Breakpoint 1 at 0x231040: file /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-b.k, line 4.
(gdb) run
Starting program: /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-b-kompiled/interpreter .krun-2021-08-13-14-20-27-vXOQmV6lwS/tmp.in.fga98yqXlc -1 .krun-2021-08-13-14-20-27-vXOQmV6lwS/result.kore
[Thread debugging using libthread_db enabled]
Using host libthread_db library "/lib/x86_64-linux-gnu/libthread_db.so.1".

Breakpoint 1, LblisBlue'LParUndsRParUnds'LESSON-19-B'Unds'Bool'Unds'Fruit (_1=Blueberry ( )) at /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-b.k:4
4         syntax Bool ::= isBlue(Fruit) [function]
(gdb)
(lldb) breakpoint set --file lesson-19-b.k --line 4
Breakpoint 1: where = interpreter`LblisBlue'LParUndsRParUnds'LESSON-19-B'Unds'Bool'Unds'Fruit + 20 at lesson-19-b.k:4:19, address = 0x0000000100003ff4
(lldb) run
Process 50546 launched: '/Users/brucecollie/code/scratch/lesson-19-b-kompiled/interpreter' (x86_64)
Process 50546 stopped
* thread #1, queue = 'com.apple.main-thread', stop reason = breakpoint 1.1
    frame #0: 0x0000000100003ff4 interpreter`LblisBlue'LParUndsRParUnds'LESSON-19-B'Unds'Bool'Unds'Fruit(_1=Blueberry ( )) at lesson-19-b.k:4:19
   1   	module LESSON-19-B
   2   	  imports BOOL
   3   	
-> 4   	  syntax Bool ::= isBlue(Fruit) [function]
   5   	  syntax Fruit ::= Blueberry() | Banana()
   6   	  rule isBlue(Blueberry()) => true
   7   	  rule isBlue(Banana()) => false
(lldb)

As we can see, we have stopped at the point where we are evaluating that
function. The value _1 that is a parameter to that function shows the
value passed to the function by the caller.

We can also break when the isBlue(Blueberry()) => true rule applies by simply
changing the line number to the line number of that rule:

(gdb) break lesson-19-b.k:6
Breakpoint 1 at 0x2af710: file /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-b.k, line 6.
(gdb) run
Starting program: /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-b-kompiled/interpreter .krun-2021-08-13-14-32-36-7kD0ic7XwD/tmp.in.8JNH5Qtmow -1 .krun-2021-08-13-14-32-36-7kD0ic7XwD/result.kore
[Thread debugging using libthread_db enabled]
Using host libthread_db library "/lib/x86_64-linux-gnu/libthread_db.so.1".

Breakpoint 1, apply_rule_138 () at /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-b.k:6
6         rule isBlue(Blueberry()) => true
(gdb)
(lldb) breakpoint set --file lesson-19-b.k --line 6
Breakpoint 1: where = interpreter`apply_rule_140 at lesson-19-b.k:6:8, address = 0x0000000100004620
(lldb) run
Process 50681 launched: '/Users/brucecollie/code/scratch/lesson-19-b-kompiled/interpreter' (x86_64)
Process 50681 stopped
* thread #1, queue = 'com.apple.main-thread', stop reason = breakpoint 1.1
    frame #0: 0x0000000100004620 interpreter`apply_rule_140 at lesson-19-b.k:6:8
   3   	
   4   	  syntax Bool ::= isBlue(Fruit) [function]
   5   	  syntax Fruit ::= Blueberry() | Banana()
-> 6   	  rule isBlue(Blueberry()) => true
   7   	  rule isBlue(Banana()) => false
   8   	
   9   	  rule F:Fruit => isBlue(F)
(lldb) 

We can also do the same with a top-level rule:

(gdb) break lesson-19-b.k:9
Breakpoint 1 at 0x2aefa0: lesson-19-b.k:9. (2 locations)
(gdb) run
Starting program: /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-b-kompiled/interpreter .krun-2021-08-13-14-33-13-9fC8Sz4aO3/tmp.in.jih1vtxSiQ -1 .krun-2021-08-13-14-33-13-9fC8Sz4aO3/result.kore
[Thread debugging using libthread_db enabled]
Using host libthread_db library "/lib/x86_64-linux-gnu/libthread_db.so.1".

Breakpoint 1, apply_rule_107 (Var'Unds'DotVar0=<generatedCounter>
  0
</generatedCounter>, Var'Unds'DotVar1=., VarF=Blueberry ( )) at /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-b.k:9
9         rule F:Fruit => isBlue(F)
(gdb)
(lldb) breakpoint set --file lesson-19-b.k --line 9
Breakpoint 1: 2 locations.
(lldb) run
Process 50798 launched: '/Users/brucecollie/code/scratch/lesson-19-b-kompiled/interpreter' (x86_64)
Process 50798 stopped
* thread #1, queue = 'com.apple.main-thread', stop reason = breakpoint 1.1
    frame #0: 0x0000000100003f2e interpreter`apply_rule_109(Var'Unds'DotVar0=<generatedCounter>
  0
</generatedCounter>, Var'Unds'DotVar1=., VarF=Blueberry ( )) at lesson-19-b.k:9:8
   6   	  rule isBlue(Blueberry()) => true
   7   	  rule isBlue(Banana()) => false
   8   	
-> 9   	  rule F:Fruit => isBlue(F)
   10  	endmodule
(lldb)  

Unlike the function rule above, we see several parameters to this function.
These are the substitution that was matched for the function. Variables only
appear in this substitution if they are actually used on the right-hand side
of the rule.

Advanced breakpoints

Sometimes it is inconvenient to set the breakpoint based on a line number.

It is also possible to set a breakpoint based on the rule label of a particular
rule. Consider the following definition (lesson-19-c.k):

module LESSON-19-C
  imports INT
  imports BOOL

  syntax Bool ::= isEven(Int) [function]
  rule [isEven]: isEven(I) => true requires I %Int 2 ==Int 0
  rule [isOdd]: isEven(I) => false requires I %Int 2 =/=Int 0

endmodule

We will run the program isEven(4). We can set a breakpoint for when a rule
applies by means of the MODULE-NAME.label.rhs syntax:

(gdb) break LESSON-19-C.isEven.rhs
Breakpoint 1 at 0x2afda0: file /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-c.k, line 6.
(gdb) run
Starting program: /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-c-kompiled/interpreter .krun-2021-08-13-14-40-29-LNNT8YEZ61/tmp.in.ZG93vWCGGC -1 .krun-2021-08-13-14-40-29-LNNT8YEZ61/result.kore
[Thread debugging using libthread_db enabled]
Using host libthread_db library "/lib/x86_64-linux-gnu/libthread_db.so.1".

Breakpoint 1, LESSON-19-C.isEven.rhs () at /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-c.k:6
6         rule [isEven]: isEven(I) => true requires I %Int 2 ==Int 0
(gdb)
(lldb) breakpoint set --name LESSON-19-C.isEven.rhs
Breakpoint 1: where = interpreter`LESSON-19-C.isEven.rhs at lesson-19-c.k:6:18, address = 0x00000001000038e0
(lldb) run
Process 51205 launched: '/Users/brucecollie/code/scratch/lesson-19-c-kompiled/interpreter' (x86_64)
Process 51205 stopped
* thread #1, queue = 'com.apple.main-thread', stop reason = breakpoint 1.1
    frame #0: 0x00000001000038e0 interpreter`LESSON-19-C.isEven.rhs at lesson-19-c.k:6:18
   3   	  imports BOOL
   4   	
   5   	  syntax Bool ::= isEven(Int) [function]
-> 6   	  rule [isEven]: isEven(I) => true requires I %Int 2 ==Int 0
   7   	  rule [isOdd]: isEven(I) => false requires I %Int 2 =/=Int 0
   8   	
   9   	endmodule
(lldb) 

We can also set a breakpoint for when a rule's side condition is evaluated
by means of the MODULE-NAME.label.sc syntax:

(gdb) break LESSON-19-C.isEven.sc
Breakpoint 1 at 0x2afd70: file /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-c.k, line 6.
(gdb) run
Starting program: /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-c-kompiled/interpreter .krun-2021-08-13-14-41-48-1BoGfJRbYc/tmp.in.kg4F8cwfCe -1 .krun-2021-08-13-14-41-48-1BoGfJRbYc/result.kore
[Thread debugging using libthread_db enabled]
Using host libthread_db library "/lib/x86_64-linux-gnu/libthread_db.so.1".

Breakpoint 1, LESSON-19-C.isEven.sc (VarI=4) at /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-c.k:6
6         rule [isEven]: isEven(I) => true requires I %Int 2 ==Int 0
(gdb) finish
Run till exit from #0  LESSON-19-C.isEven.sc (VarI=4) at /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-c.k:6
0x00000000002b2662 in LblisEven'LParUndsRParUnds'LESSON-19-C'Unds'Bool'Unds'Int (_1=4) at /home/dwightguth/kframework-5.0.0/k-distribution/k-tutorial/1_basic/19_debugging/lesson-19-c.k:5
5         syntax Bool ::= isEven(Int) [function]
Value returned is $1 = true
(gdb)
(lldb) breakpoint set --name LESSON-19-C.isEven.sc
Breakpoint 1: where = interpreter`LESSON-19-C.isEven.sc + 1 at lesson-19-c.k:6:18, address = 0x00000001000038c1
(lldb) run
Process 52530 launched: '/Users/brucecollie/code/scratch/lesson-19-c-kompiled/interpreter' (x86_64)
Process 52530 stopped
* thread #1, queue = 'com.apple.main-thread', stop reason = breakpoint 1.1
    frame #0: 0x00000001000038c1 interpreter`LESSON-19-C.isEven.sc(VarI=0x0000000101800088) at lesson-19-c.k:6:18
   3   	  imports BOOL
   4   	
   5   	  syntax Bool ::= isEven(Int) [function]
-> 6   	  rule [isEven]: isEven(I) => true requires I %Int 2 ==Int 0
   7   	  rule [isOdd]: isEven(I) => false requires I %Int 2 =/=Int 0
   8   	
   9   	endmodule
(lldb) finish
Process 52649 stopped
* thread #1, queue = 'com.apple.main-thread', stop reason = step out
Return value: (bool) $0 = true

    frame #0: 0x00000001000069e5 interpreter`LblisEven'LParUndsRParUnds'LESSON-19-C'Unds'Bool'Unds'Int(_1=0x0000000101800088) at lesson-19-c.k:5:19
   2   	  imports INT
   3   	  imports BOOL
   4   	
-> 5   	  syntax Bool ::= isEven(Int) [function]
   6   	  rule [isEven]: isEven(I) => true requires I %Int 2 ==Int 0
   7   	  rule [isOdd]: isEven(I) => false requires I %Int 2 =/=Int 0
   8
(lldb)

Here we have used the built-in command finish to tell us whether the side
condition returned true or not. Note that once again, we see the substitution
that was matched from the left-hand side. Like before, a variable will only
appear here if it is used in the side condition.

Debugging rule matching

Sometimes it is useful to try to determine why a particular rule did or did
not apply. K provides some basic debugging commands which make it easier
to determine this.

Consider the following K definition (lesson-19-d.k):

module LESSON-19-D

  syntax Foo ::= foo(Bar)
  syntax Bar ::= bar(Baz) | bar2(Baz)
  syntax Baz ::= baz() | baz2()

  rule [baz]: foo(bar(baz())) => .K

endmodule

Suppose we try to run the program foo(bar(baz2())). It is obvious from this
example why the rule in this definition will not apply. However, in practice,
such cases are not always obvious. You might look at a rule and not immediately
spot why it didn't apply on a particular term. For this reason, it can be
useful to get the debugger to provide a log about how it tried to match that
term. You can do this with the k match command. If you are stopped after
having run k start or k step, you can obtain this log for any rule after
any step by running the command k match MODULE.label subject for a particular
top-level rule label.

For example, with the baz rule above, we get the following output:

(gdb) k match LESSON-19-D.baz subject
Subject:
baz2 ( )
does not match pattern:
baz ( )
(lldb) k match LESSON-19-D.baz subject
Subject:
baz2 ( )
does not match pattern:
baz ( )

As we can see, it provided the exact subterm which did not match against the
rule, as well as the particular subpattern it ought to have matched against.

This command does not actually take any rewrite steps. In the event that
matching actually succeeds, you will still need to run the k step command
to advance to the next step.

Final notes

In addition to the functionality provided above, you have the full power of
GDB or LLDB at your disposal when debugging. Some features are not particularly
well-adapted to K code and may require more advanced knowledge of the
term representation or implementation to use effectively, but anything that
can be done in GDB or LLDB can in theory be done using this debugging functionality.
We suggest you refer to the
GDB Documentation or
LLDB Tutorial if you
want to try to do something and are unsure as to how.

Exercises

  1. Compile your solution to Lesson 1.18, Exercise 2 with debugging support
    enabled and step through several programs you have previously used to test.
    Then set a breakpoint on the isKResult function and observe the state of the
    interpreter when stopped at that breakpoint. Set a breakpoint on the rule for
    addition and run a program that causes it to be stopped at that breakpoint.
    Finally, step through the program until the addition symbol is at the top
    of the K cell, and then use the k match command to report the reason why
    the subtraction rule does not apply. You may need to modify the definition
    to insert some rule labels.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.20: K Backends and the Haskell Backend.

Lesson 1.20: K Backends and the Haskell Backend

The purpose of this lesson is to teach about the multiple backends of K,
in particular the Haskell Backend which is the complement of the backend we
have been using so far.

K Backends

Thus far, we have not discussed the distinction between the K frontend and
the K backends at all. We have simply assumed that if you run kompile on a
K definition, there will be a compiler backend that will allow you to execute
the K definition you have compiled.

K actually has multiple different backends. The one we have been using so far
implicitly, the default backend, is called the LLVM Backend. It is
designed to support efficient, optimized concrete execution and search. It
does this by compiling your K definition to LLVM bitcode and then using LLVM
to generate machine code for it that is compiled and linked and executed.
However, K is a formal methods toolkit at the end of the day, and the primary
goal many people have when defining a programming language in K is to
ultimately be able to perform more advanced verification on programs in their
programming language.

It is for this purpose that K also provides the Haskell Backend, so called
because it is implemented in Haskell. While we will cover the features of the
Haskell Backend in more detail in the next two lessons, the important thing to
understand is that it is a separate backend which is optimized for more formal
reasoning about programming languages. While it is capable of performing
concrete execution, it does not do so as efficiently as the LLVM Backend.
In exchange, it provides more advanced features.

Choosing a backend

You can choose which backend to use to compile a K definition by means of the
--backend flag to kompile. By default, if you do not specify this flag, it
is equivalent to if you had specified --backend llvm. However, to use the
Haskell Backend instead, you can simply say kompile --backend haskell on a
particular K definition.

As an example, here is a simple K definition that we have seen before in the
previous lesson (lesson-20.k):

module LESSON-20
  imports INT

  rule I => I +Int 1
    requires I <Int 100
endmodule

Previously we compiled this definition using the LLVM Backend, but if we
instead execute the command kompile lesson-20.k --backend haskell, we
will get an interpreter for this K definition that is implemented in Haskell
instead. Unlike the default LLVM Backend, the Haskell Backend is not a
compiler per se. It does not generate new Haskell code corresponding to your
programming language and then compile and execute it. Instead, it is an
interpreter which reads the generated IR from kompile and implements in
Haskell an interpreter that is capable of interpreting any K definition.

Note that on arm64 macOS (Apple Silicon), there is a known issue with the Compact
library that causes crashes in the Haskell backend. Pass the additional flag
--no-haskell-binary to kompile to resolve this.
This flag is also needed when using krun.

Exercise

Try running the program 0 in this K definition on the Haskell Backend and
compare the final configuration to what you would get compiling the same
definition with the LLVM Backend.

Legacy backends

As a quick note, K does provide one other backend, which exists primarily as
legacy code which should be considered deprecated. This is the
Java Backend. The Java Backend is essentially a precursor to the Haskell
Backend. We will not cover this backend in any detail since it is deprecated,
but we still mention it here for the purposes of understanding.

Exercises

  1. Compile your solution to Lesson 1.18, Exercise 2 with the Haskell Backend
    and execute some programs. Compare the resulting configurations with the
    output of the same program on the LLVM Backend. Note that if you are getting
    different behaviors on the Haskell backend, you might have some luck debugging
    by passing --search to krun when using the LLVM backend.

Next lesson

Once you have completed the above exercises, you can continue to
Lesson 1.21: Unification and Symbolic Execution.

Lesson 1.21: Unification and Symbolic Execution

The purpose of this lesson is to teach the basic concepts of symbolic execution
in order to introduce the unique capabilities of the Haskell Backend at a
conceptual level.

Symbolic Execution

Thus far, all of the programs we have run using K have been concrete
configurations. What this means is that the configuration we use to initialize
the K rewrite engine is concrete; in other words, contains no logical
variables. The LLVM Backend is a concrete execution engine, meaning that
it is only capable of rewriting concrete configurations.

By contrast, the Haskell Backend performs symbolic execution, which is
capable of rewriting any configuration, including those where parts of the
configuration are symbolic, ie, contain variables or uninterpreted
functions.

Unification

Previously, we have introduced the concept that K rewrite rules operate by
means of pattern matching: the current configuration being rewritten is pattern
matched against the left-hand side of the rewrite rule, and the substitution
is used in order to construct a new term from the right-hand side. In symbolic
execution, we use
unification
instead of pattern matching. To summarize, unification behaves akin to a
two-way pattern matching where both the configuration and the left-hand side
of the rule can contain variables, and the algorithm generates a
most general unifier containing substitutions for the variables in both
which will make both terms equal.

Feasibility

Unification by itself cannot completely solve the problem of symbolic
execution. One task symbolic execution must perform is to identify whether
a particular symbolic term is feasible, that is to say, that there actually
exists a concrete instantiation of that term such that all the logical
constraints on that term can actually be satisfied. The Haskell Backend
delegates this task to Z3, an
SMT solver.
This solver is used to periodically trim configurations that are determined
to be mathematically infeasible.

Symbolic terms

The final component of symbolic execution consists of the task of introducing
symbolic terms into the configuration. This can be done one of two different
ways. First, the term being passed to krun can actually be symbolic. This
is less frequently used because it requires the user to construct an AST
that contains variables, something which our current parsing capabilities are
not well-equipped to do. The second, more common, way of introducing symbolic
terms into a configuration consists of writing rules where there exists an
existentially qualified variable on the right-hand side of the rule that does
not exist on the left-hand side of the rule.

In order to prevent users from writing such rules by accident, K requires
that such variables begin with the ? prefix. For example, here is a rule
that rewrites a constructor foo to a symbolic integer:

rule <k> foo => ?X:Int ...</k>

When this rule applies, a fresh variable is introduced to the configuration, which
then is unified against the rules that might apply in order to symbolically
execute that configuration.

ensures clauses

We also introduce here a new feature of K rules that applies when a rule
has this type of variable on the right-hand side: the ensures clause.
An ensures clause is similar to a requires clause and can appear after
a rule body, or after a requires clause. The ensures clause is used to
introduce constraints that might apply to the variable that was introduced by
that rule. For example, we could write the rule above with the additional
constraint that the symbolic integer that was introduced must be less than
five, by means of the following rule:

rule <k> foo => ?X:Int ...</k> ensures ?X <Int 5

Putting it all together

Putting all these pieces together, it is possible to use the Haskell Backend
to perform symbolic reasoning about a particular K module, determining all the
possible states that can be reached by a symbolic configuration.

For example, consider the following K definition (lesson-21.k):

module LESSON-21
    imports INT

    rule <k> 0 => ?X:Int ... </k> ensures ?X =/=Int 0
    rule <k> X:Int => 5  ... </k> requires X >=Int 10
endmodule

When we symbolically execute the program 0, we get the following output
from the Haskell Backend:

    <k>
      5 ~> .
    </k>
  #And
    {
      true
    #Equals
      ?X:Int >=Int 10
    }
  #And
    #Not ( {
      ?X:Int
    #Equals
      0
    } )
#Or
    <k>
      ?X:Int ~> .
    </k>
  #And
    #Not ( {
      true
    #Equals
      ?X:Int >=Int 10
    } )
  #And
    #Not ( {
      ?X:Int
    #Equals
      0
    } )

Note some new symbols introduced by this configuration: #And, #Or, and
#Equals. While andBool, orBool, and ==K represent functions of sort
Bool, #And, #Or, and #Equals are matching logic connectives. We
will discuss matching logic in more detail later in the tutorial, but the basic
idea is that these symbols represent Boolean operators over the domain of
configurations and constraints, as opposed to over the Bool sort.

Notice that the configuration listed above is a disjunction of conjunctions.
This is the most common form of output that can be produced by the Haskell
Backend. In this case, each conjunction consists of a configuration and a set
of constraints. What this conjunction describes, essentially, is a
configuration and a set of information that was derived to be true while
rewriting that configuration.

Similar to how we saw --search in a previous lesson, the reason we have
multiple disjuncts is because there are multiple possible output states
for this program, depending on whether or not the second rule applied. In the
first case, we see that ?X is greater than or equal to 10, so the second rule
applied, rewriting the symbolic integer to the concrete integer 5. In the
second case, we see that the second rule did not apply because ?X is less
than 10. Moreover, because of the ensures clause on the first rule, we know
that ?X is not zero, therefore the first rule will not apply a second time.
If we had omitted this constraint, we would have ended up infinitely applying
the first rule, leading to krun not terminating.

In the next lesson, we will cover how symbolic execution forms the backbone
of deductive program verification in K and how we can use K to prove programs
correct against a specification.

Exercises

  1. Create another rule in LESSON-21 that rewrites odd integers greater than
    ten to a symbolic even integer less than 10 and greater than 0. This rule will
    now apply nondeterministically along with the existing rules. Predict what the
    resulting output configuration will be from rewriting 0 after adding this
    rule. Then run the program and see whether your prediction is correct.

Once you have completed the above exercises, you can continue to
Lesson 1.22: Basics of Deductive Program Verification using K.

Lesson 1.22: Basics of Deductive Program Verification using K

In this lesson, you will familiarize yourself with the basics of using K for
deductive program verification.

1. Setup: Simple Programming Language with Function Calls

We base this lesson on a simple programming language with functions,
assignment, if conditionals, and while loops. Take your time to study its
formalization below (lesson-22.k):

module LESSON-22-SYNTAX
    imports INT-SYNTAX
    imports BOOL-SYNTAX
    imports ID-SYNTAX

    syntax Exp ::= IExp | BExp

    syntax IExp ::= Id | Int

    syntax KResult ::= Int | Bool | Ints

    // Take this sort structure:
    //
    //     IExp
    //    /    \
    // Int      Id
    //
    // Through the List{_, ","} functor.
    // Must add a `Bot`, for a common subsort for the empty list.

    syntax Bot
    syntax Bots ::= List{Bot, ","} [klabel(exps)]
    syntax Ints ::= List{Int, ","} [klabel(exps)]
                  | Bots
    syntax Ids  ::= List{Id, ","}  [klabel(exps)]
                  | Bots
    syntax Exps ::= List{Exp, ","} [klabel(exps), seqstrict]
                  | Ids | Ints

    syntax IExp ::= "(" IExp ")" [bracket]
                  | IExp "+" IExp [seqstrict]
                  | IExp "-" IExp [seqstrict]
                  > IExp "*" IExp [seqstrict]
                  | IExp "/" IExp [seqstrict]
                  > IExp "^" IExp [seqstrict]
                  | Id "(" Exps ")" [strict(2)]

    syntax BExp ::= Bool

    syntax BExp ::= "(" BExp ")" [bracket]
                  | IExp "<=" IExp [seqstrict]
                  | IExp "<"  IExp [seqstrict]
                  | IExp ">=" IExp [seqstrict]
                  | IExp ">"  IExp [seqstrict]
                  | IExp "==" IExp [seqstrict]
                  | IExp "!=" IExp [seqstrict]

    syntax BExp ::= BExp "&&" BExp
                  | BExp "||" BExp

    syntax Stmt ::=
         Id "=" IExp ";" [strict(2)]                        // Assignment
       | Stmt Stmt [left]                                   // Sequence
       | Block                                              // Block
       | "if" "(" BExp ")" Block "else" Block [strict(1)]   // If conditional
       | "while" "(" BExp ")" Block                         // While loop
       | "return" IExp ";"                    [seqstrict]   // Return statement
       | "def" Id "(" Ids ")" Block                         // Function definition

    syntax Block ::=
         "{" Stmt "}"    // Block with statement
       | "{" "}"         // Empty block
endmodule

module LESSON-22
    imports INT
    imports BOOL
    imports LIST
    imports MAP
    imports LESSON-22-SYNTAX

    configuration
      <k> $PGM:Stmt </k>
      <store> .Map </store>
      <funcs> .Map </funcs>
      <stack> .List </stack>

 // -----------------------------------------------
    rule <k> I1 + I2 => I1 +Int I2 ... </k>
    rule <k> I1 - I2 => I1 -Int I2 ... </k>
    rule <k> I1 * I2 => I1 *Int I2 ... </k>
    rule <k> I1 / I2 => I1 /Int I2 ... </k>
    rule <k> I1 ^ I2 => I1 ^Int I2 ... </k>

    rule <k> I:Id => STORE[I] ... </k>
         <store> STORE </store>

 // ------------------------------------------------
    rule <k> I1 <= I2 => I1  <=Int I2 ... </k>
    rule <k> I1  < I2 => I1   <Int I2 ... </k>
    rule <k> I1 >= I2 => I1  >=Int I2 ... </k>
    rule <k> I1  > I2 => I1   >Int I2 ... </k>
    rule <k> I1 == I2 => I1  ==Int I2 ... </k>
    rule <k> I1 != I2 => I1 =/=Int I2 ... </k>

    rule <k> B1 && B2 => B1 andBool B2 ... </k>
    rule <k> B1 || B2 => B1  orBool B2 ... </k>

    rule <k> S1:Stmt S2:Stmt => S1 ~> S2 ... </k>

    rule <k> ID = I:Int ; => . ... </k>
         <store> STORE => STORE [ ID <- I ] </store>

    rule <k> { S } => S ... </k>
    rule <k> {   } => . ... </k>

    rule <k> if (true)   THEN else _ELSE => THEN ... </k>
    rule <k> if (false) _THEN else  ELSE => ELSE ... </k>

    rule <k> while ( BE ) BODY => if ( BE ) { BODY while ( BE ) BODY } else { } ... </k>

    rule <k> def FNAME ( ARGS ) BODY => . ... </k>
         <funcs> FS => FS [ FNAME <- def FNAME ( ARGS ) BODY ] </funcs>

    rule <k> FNAME ( IS:Ints ) ~> CONT => #makeBindings(ARGS, IS) ~> BODY </k>
         <funcs> ... FNAME |-> def FNAME ( ARGS ) BODY ... </funcs>
         <store> STORE => .Map </store>
         <stack> .List => ListItem(state(CONT, STORE)) ... </stack>

    rule <k> return I:Int ; ~> _ => I ~> CONT </k>
         <stack> ListItem(state(CONT, STORE)) => .List ... </stack>
         <store> _ => STORE </store>

    rule <k> return I:Int ; ~> . => I </k>
         <stack> .List </stack>

    syntax KItem ::= #makeBindings(Ids, Ints)
                   | state(continuation: K, store: Map)
 // ----------------------------------------------------
    rule <k> #makeBindings(.Ids, .Ints) => . ... </k>
    rule <k> #makeBindings((I:Id, IDS => IDS), (IN:Int, INTS => INTS)) ... </k>
         <store> STORE => STORE [ I <- IN ] </store>
endmodule

Next, compile this example using kompile lesson-22.k --backend haskell. If
your processor is an Apple Silicon processor, add the --no-haskell-binary
flag if the compilation fails.

2. Setup: Proof Environment

Next, take the following snippet of K code and save it in lesson-22-spec.k.
This is a skeleton of the proof environment, and we will complete it as the
lesson progresses.

requires "lesson-22.k"
requires "domains.md"

module LESSON-22-SPEC-SYNTAX
    imports LESSON-22-SYNTAX

endmodule

module VERIFICATION
    imports K-EQUAL
    imports LESSON-22-SPEC-SYNTAX
    imports LESSON-22
    imports MAP-SYMBOLIC

endmodule

module LESSON-22-SPEC
    imports VERIFICATION

endmodule

3. Claims

  1. The first claim we will ask K to prove is that 3 + 4, in fact, equals 7.
    Claims are stated using the claim keyword, followed by the claim
    statement:
claim <k> 3 + 4 => 7 ... </k>

Add this claim to the LESSON-22-SPEC module and run the K prover using the
command kprove lesson-22-spec.k. You should get back the output #Top,
which denotes the Matching Logic equivalent of true and means, in this
context, that all claims have been proven correctly.

  1. The second claim reasons about the if statement that has a concrete condition:
claim <k> if ( 3 + 4 == 7 ) {
            $a = 1 ;
            } else {
            $a = 2 ;
            }
        => . ... </k>
        <store> STORE => STORE [ $a <- 1 ] </store>

stating that the given program terminates (=> .), and when it does, the value
of the variable $a is set to 1, meaning that the execution will have taken
the then branch. Add this claim to the LESSON-22-SPEC module, but also add

syntax Id ::= "$a" [token]

to the LESSON-22-SPEC-SYNTAX module in order to declare $a as a token so
that it can be used as a program variable. Re-run the K prover, which should
again return #Top.

  1. Our third claim demonstrates how to reason about both branches of an if
    statement at the same time:
claim <k> $a = A:Int ; $b = B:Int ;
          if ($a < $b) {
            $c = $b ;
          } else {
            $c = $a ;
          }
        => . ... </k>
        <store> STORE => STORE [ $a <- A ] [ $b <- B ] [ $c <- ?C:Int ] </store>
    ensures (?C ==Int A) orBool (?C ==Int B)

The program in question first assigns symbolic integers A and B to program
variables $a and $b, respectively, and then executes the given if
statement, which has a symbolic condition (A < B), updating the value of the
program variable $c in both branches. The specification we give states that
the if statement terminates, with $a and $b updated, respectively, to A
and B, and $c updated to some symbolic integer value ?C. Via the
ensures clause, which is used to specify additional constraints that hold
after execution, we also state that this existentially quantified ?C equals
either A or B.

Add the productions declaring $b and $c as tokens to the
LESSON-22-SPEC-SYNTAX module, the claim to the LESSON-22-SPEC module, run
the K prover again, and observe the output, which should not be #Top this
time. This means that K was not able to prove the claim, and we now need to
understand why. We do so by examining the output, which should look as follows:

    (InfoReachability) while checking the implication:
    The configuration's term unifies with the destination's term,
    but the implication check between the conditions has failed.

  #Not (
    #Exists ?C . {
        STORE [ $a <- A:Int ] [ $b <- B:Int ] [ $c <- ?C:Int ]
      #Equals
        STORE [ $a <- A:Int ] [ $b <- B:Int ] [ $c <- B:Int ]
    }
  #And
    {
      true
    #Equals
      ?C ==Int A orBool ?C ==Int B
    }
  )
#And
  <generatedTop>
    <k>
      _DotVar1
    </k>
    <store>
      STORE [ $a <- A:Int ] [ $b <- B:Int ] [ $c <- B:Int ]
    </store>
    <funcs>
      _Gen3
    </funcs>
    <stack>
      _Gen5
    </stack>
  </generatedTop>
#And
  {
    true
  #Equals
    A <Int B
  }

This output starts with a message telling us at which point the proof failed,
followed by the final state, which consists of three parts: some negative
Matching Logic (ML) constraints, the final configuration (<generatedTop> ... </generatedTop>), and some positive ML constraints. Generally speaking,
these positive and the negative constraints could arise from various sources,
such as (but not limited to) branches taken by the execution
(e.g. { true #Equals A <Int B } or #Not ( { true #Equals A <Int B } )),
or ensures constraints.

First, we examine the message:

(InfoReachability) while checking the implication:
The configuration's term unifies with the destination's term,
but the implication check between the conditions has failed.

which tells us that the structure of the final configuration is as expected,
but that some of the associated constraints cannot be proven. We next look at
the final configuration, in which the relevant item is the <store> ... </store> cell, because it is the only one that we are reasoning about. By
inspecting its contents:

STORE [ $a <- A:Int ] [ $b <- B:Int ] [ $c <- B:Int ]

we see that we should be within the constraints of the ensures, since the
value of $c in the store equals B in this branch. We next examine the
negative and positive constraints of the output and, more often than not, the
goal is to instruct K how to use the information from the final configuration
and the positive constraints to falsify one of the negative constraints. This
is done through simplifications.

So, the positive constraint that we have is

{ true #Equals A <Int B }

meaning that A <Int B holds. Given the analysed program, this tells us that
we are in the then branch of the if. The negative constraint is

  #Not (
    #Exists ?C . {
        STORE [ $a <- A:Int ] [ $b <- B:Int ] [ $c <- ?C:Int ]
      #Equals
        STORE [ $a <- A:Int ] [ $b <- B:Int ] [ $c <- B:Int ]
    }
  #And
    { true #Equals ?C ==Int A orBool ?C ==Int B }
  )

and we observe, from the first equality, that the existential ?C should be
instantiated with B. This would make both branches of the #And true,
falsifying the outside #Not. We just need to show K how to conclude that
?C ==Int B. We do so by introducing the following simplification into the
VERIFICATION module:

rule { M:Map [ K <- V ] #Equals M [ K <- V' ] } => { V #Equals V' } [simplification]

which formalizes our internal understanding of ?C ==Int B. The rule states
that when we update the same key in the same map with two values, and the
resulting maps are equal, then the two values must be equal as well. The
[simplification] attribute indicates to K to use this rule to simplify the
state when trying to prove claims. Like function rules, simplification rules
do not complete to the top of the configuration, but instead apply anywhere
their left-hand-side matches. Re-run the K prover, which should now return
#Top, indicating that K was able to use the simplification and prove the
required claims.

  1. Next, we show how to state and prove properties of while loops. In
    particular, we consider the following loop
claim
    <k>
        while ( 0 < $n ) {
            $s = $s + $n;
            $n = $n - 1;
            } => . ...
    </k>
    <store>
        $s |-> (S:Int => S +Int ((N +Int 1) *Int N /Int 2))
        $n |-> (N:Int => 0)
    </store>
    requires N >=Int 0

which adds the sum of the first $n integers to $s, assuming the value of $n
is non-negative to begin with. This is reflected in the store by stating that,
after the execution of the loop, the original value of $s (which is set to
equal some symbolic integer S) is incremented by ((N +Int 1) *Int N /Int 2), and the value of $n always equals 0. Add $n and $s as tokens in
the LESSON-22-SPEC-SYNTAX module, the above claim to the LESSON-22-SPEC
module, and run the K prover, which should return #Top.

  1. Finally, our last claim is about a program that uses function calls:
claim
    <k>
        def $sum($n, .Ids) {
            $s = 0 ;
            while (0 < $n) {
                $s = $s + $n;
                $n = $n - 1;
            }
            return $s;
        }

        $s = $sum(N:Int, .Ints);
    => . ... </k>
    <funcs> .Map => ?_ </funcs>
    <store> $s |-> (_ => ((N +Int 1) *Int N /Int 2)) </store>
    <stack> .List </stack>
    requires N >=Int 0

Essentially, we have wrapped the while loop from claim 3.4 into a function
$sum, and then called that function with a symbolic integer N, storing the
return value in the variable $s. The specification states that this program
ends up storing the sum of the first N integers in the variable $n. Add $sum
to the LESSON-22-SPEC-SYNTAX module, the above claim to the
LESSON-22-SPEC module, and run the K prover, which should again return
#Top.

Exercises

  1. Change the condition of the if statement in part 3.2 to take the else
    branch and adjust the claim so that the proof passes.

  2. The post-condition of the specification in part 3.3 loses some information.
    In particular, the value of ?C is in fact the maximum of A and B.
    Prove the same claim as in 3.2, but with the post-condition ensures (?C ==Int maxInt(A, B)). For this, you will need to extend the VERIFICATION
    module with two simplifications that capture the meaning of maxInt(A:Int, B:Int). Keep in mind that any rewriting rule can be used as a
    simplification; in particular, that simplifications can have requires
    clauses.

  3. Following the pattern shown in part 3.4, assuming a non-negative initial
    value of $b, specify and verify the following while loop:

while ( 0 < $b ) {
    $a = $a + $c;
    $b = $b - 1;
    $c = $c - 1;
}

Hint: You will not need additional simplifications---once you've got the
specification right, the proof will go through.

  1. Write an arbitrary yet not-too-complex function (or several functions
    interacting with each other), and try to specify and verify it (them) in K.

Section 2: Intermediate K Concepts

The goal of this second section is to supplement a beginning developer's
knowledge of K after they have gained a basic understanding of K. Each lesson
in this section can be completed independently in order to learn about a
particular facet of the K language. The lessons are written to provide basic
understanding of less commonly-used features of K to someone who is still
learning K. For more complete references of these features, the reader ought to
consult the User Manual.

The reader ought to be able to complete lessons in this section as needed in
order to learn about specific features of interest, but if desired, can also
complete the entire section in one go. Someone who has completed this entire
section ought to be able to read and understand most K specifications, as well
as write their own specifications of some complexity, and use them to perform
most common K-related tasks. They can then read about specific lessons in
Section 3: Advanced K Concepts if they want to
learn more.

Table of Contents

  1. Macros, Aliases, and Anywhere Rules
  2. Fresh Constants
  3. KLabels and Abstract Syntax
  4. Overloaded Symbols
  5. Matching Logic Connectives and #Or Patterns
  6. Function Context
  7. Record Productions and Named Nonterminals
  8. #fun and #let
  9. #as patterns
  10. The Matching Operators, :=K and :/=K
  11. Uncommon Evaluation Order Concepts
  12. IEEE 754 Floating Point and Fixed Width Integers
  13. Alpha-renaming-aware Substitution
  14. File I/O
  15. String Buffers and Byte Sequences
  16. The Intermediate Language of K, KORE
  17. Debugging Proofs using the Haskell Backend REPL

Lesson 2.1: Macros, Aliases, and Anywhere Rules

The purpose of this lesson is to explain the behavior of the macro,
macro-rec, alias, and alias-rec production attributes, as well as the
anywhere rule attribute. These attributes control the meaning of how rules
associated with them are applied.

Macros

Thus far in the K tutorial, we have described three different types of rules:

  1. Top-level rewrite rules, which rewrite a configuration composed of cells to
    another configuration;
  2. Function rules, which define the behavior of a function written over
    arbitrary input and output types; and
  3. Simplification rules, which describe ways in which the symbolic execution
    engine ought to simplify terms containing symbolic values.

This lesson introduces three more types of rules, the first of which are
macros. A production is a macro if it has the macro attribute, and all
rules whose top symbol on the left hand side is a macro are macro rules
which define the behavior of the macro. Like function rules and simplification
rules, macro rules do not participate in cell completion. However, unlike
function rules and simplification rules, macro rules are applied statically
before rewriting begins, and the macro symbol is expected to no longer appear
in the initial configuration for rewriting once all macros in that
configuration are rewritten.

The rationale behind macros is they allow you to define one piece of syntax
in terms of another piece of syntax without any runtime overhead associated
with the cost of rewriting one to the other. This process is a common one in
programming language design and specification and is referred to as
desugaring; The syntax that is transformed is typically also referred to as
syntactic sugar for another type of syntax. For example, in a language with
if statements and curly braces, you could write the following fragment
(lesson-01.k):

module LESSON-01
  imports BOOL

  syntax Stmt ::= "if" "(" Exp ")" Stmt             [macro]
                | "if" "(" Exp ")" Stmt "else" Stmt
                | "{" Stmts "}"
  syntax Stmts ::= List{Stmt,""}
  syntax Exp ::= Bool

  rule if ( E ) S => if ( E ) S else { .Stmts }
endmodule

In this example, we see that an if statement without an else clause is
defined in terms of one with an else clause. As a result, we would only
need to give a single rule for how to rewrite if statements, rather than
two separate rules for two types of if statements. This is a common pattern
for dealing with program syntax that contains an optional component to it.

It is worth noting that by default, macros are not applied recursively. To be
more precise, by default a macro that arises as a result of the expansion of
the same macro is not rewritten further. This is primarily to simplify the
macro expansion process and reduce the risk that improperly defined macros will
lead to non-terminating behavior.

It is possible, however, to tell K to expand a macro recursively. To do this,
simply replace the macro attribute with the macro-rec attribute. Note that
K does not do any kind of checking to ensure termination here, so it is
important that rules be defined correctly to always terminate, otherwise the
macro expansion phase will run forever. Fortunately, in practice it is very
simple to ensure this property for most of the types of macros that are
typically used in real-world semantics.

Exercise

Using a Nat sort containing the constructors 0 and S (i.e., a
Peano-style axiomatization of the
natural numbers where S(N) = N + 1, S(S(N)) = N + 2, etc), write a macro
that will compute the sum of two numbers.

Aliases

NOTE: This lesson introduces the concept of "aliases", which are a variant
of macros. While similar, this is different from the concept of "aliases" in
matching logic, which is introduced in Lesson 2.16.

Macros can be very useful in helping you define a programming language.
However, they can be disruptive while pretty printing a configuration. For
example, you might write a set of macros that transforms the code the user
wrote into equivalent code that is slightly harder to read. This can make it
more difficult to understand the code when it is pretty printed as part of the
output of rewriting.

K defines a relatively straightforward but novel solution to this problem,
which is known as a K alias. An alias in K is very similar to a macro,
with the exception that the rewrite rule will also be applied backwards
during the pretty-printing process.

It is very simple to make a production be an alias instead of a macro: simply
use the alias or alias-rec attributes instead of the macro or macro-rec
attributes. For example, if the example involving if statements above was
declared using an alias instead of a macro, the Stmt term if (E) {} else {}
would be pretty-printed as if (E) {}. This is because during pretty-printing,
the term participates in another macro-expansion pass. However, this macro
expansion step will only apply rules with the alias or alias-rec attribute,
and, critically, it will reverse the rule by treating the left-hand side as if
it were the right-hand side, and vice versa.

This can be very useful to allow you to define one construct in terms of
another while still being able to pretty-print the result as if it were
the original term in question. This can be especially useful for applications
of K where we are taking the output of rewriting and attempting to use it as
a code fragment that we then execute, such as with test generation.

Exercise

Modify LESSON-01 above to use an alias instead of a macro and experiment
with how various terms are pretty-printed by invoking krun on them.

anywhere rules

The last type of rule introduced in this lesson is the anywhere rule. An
anywhere rule is specified by adding the anywhere attribute to a rule. Such a
rule is similar to a function rule in that it does not participate in cell
completion, and will apply anywhere that the left-hand-side matches in the
configuration, but distinct in that the symbol in question can still be matched
against in the left-hand side of other rules, even during concrete rewriting.
The reasoning behind this is that instead of the symbol in question being a
constructor, it is a constructor modulo the axioms defined with the
anywhere attribute. Essentially, the rules with the anywhere attribute will
apply as soon as they appear in the right-hand side of a rule being applied,
but the symbol in question will still be treated as a symbol that can be
matched on if it is not completely removed by those rules.

This can be useful in certain cases to allow you to define transformations over
particular pieces of syntax while still generally giving those pieces of syntax
another meaning when the anywhere rule does not apply. For example, the ISO C
standard defines the semantics of *&x as exactly equal to x, with no
reading or writing of memory taking place, and the K semantics of C implements
this functionality using an anywhere rule that is applied at compilation time.

NOTE: the anywhere attribute is only implemented on the LLVM backend
currently. Attempting to use it in a semantics that is compiled with the
Haskell backend will result in an error being reported by the compiler. This
should be remembered when using this attribute, as it may not be suitable for
a segment of a semantics which is intended to be symbolically executed.

Exercises

  1. Write a version of the calculator from Lesson 1.14 Exercise 1, which uses
    the same syntax for evaluating expressions, but defines its arithmetic logic
    using anywhere rules rather than top-level rewrite rules.

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.2: Fresh Constants

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.3: KLabels and Abstract Syntax

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.4: Overloaded Symbols

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.5: Matching Logic Connectives and #Or Patterns

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.6: Function Context

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.7: Record Productions and Named Nonterminals

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.8: #fun and #let

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.9: #as Patterns

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.10: The Matching Operators, :=K and :/=K

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.11: Uncommon Evaluation Order Concepts

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.12: IEEE 754 Floating Point and Fixed Width Integers

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.13: Alpha-renaming-aware Substitution

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.14: File I/O

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.15: String Buffers and Byte Sequences

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.16: The Intermediate Language of K, KORE

Return to Top

Click here to return to the Table of Contents for Section 2.

Lesson 2.17: Debugging Proofs using the Haskell Backend REPL

Return to Top

Click here to return to the Table of Contents for Section 2.

K User Manual

NOTE: The K User Manual is still under construction; some features of K
may have partial or missing documentation.

Introduction

Why K?

The K Framework is a programming language and system design toolkit made for
practioners and researchers alike.

K For Practioners:
K is a framework for deriving programming languages tools from their semantic
specifications.

Typically, programming language tool development follows a similar pattern.
After a new programming language is designed, separate teams will develop
separate language tools (e.g. a compiler, interpreter, parser, symbolic
execution engine, etc). Code reuse is uncommon. The end result is that for each
new language, the same basic tools and patterns are re-implemented again and
again.

K approaches the problem differently -- it generates each of these tools from a single language specification.
The work of programming language design and tool implementation are made separate concerns.
The end result is that the exercise of
designing new languages and their associated tooling is now reduced to
developing a single language specification from which we derive our tooling for
free
.

K For Researchers:
K is a configuration- and rewrite-based executable semantic framework.

In more detail, K specifications are:

  1. Executable: compile into runnable and testable programs;
  2. Semantic: correspond to a logical theory with a sound and relatively
    complete proof system;
  3. Configuration-based: organize system states into compositional,
    hierarchical, labelled units called cells;
  4. Rewrite-based: define system transitions using rewrite rules.

K specifications are compiled into particular matching logic theories, giving
them a simple and expressive semantics. K semantic rules are implicitly defined
over the entire configuration structure, but omit unused cells, enabling a
highly modular definitional style. Furthermore, K has been used to develop
programming languages, type systems, and formal analysis tools.

Manual Objectives

As mentioned in the Why K? section above, the K Framework is designed as a
collection of language-generic command-line interface (CLI) tools which revolve
around K specifications. These tools cover a broad range of uses, but they
typically fall into one of the following categories:

  1. Transforming K Specs (e.g. compilation)
  2. Running K Specs (e.g. concrete and symbolic execution)
  3. Analyzing K Specs (e.g. theorem proving)

The main user-facing K tools include:

  • kompile - the K compiler driver
  • kparse - the stanadlone K parser and abstract syntax tree (AST)
    transformation tool
  • krun - the K interpreter and symbolic execution engine driver
  • kprove - the K theorem prover

This user manual is designed to be a tool reference.
In particular, it is not desgined to be a tutorial on how to write K
specifications or to teach the logical foundations of K. New K users should
consult our dedicated
K tutorial,
or the more language-design oriented
PL tutorial.
Researchers seeking to learn more about the logic underlying K are encouraged
to peruse the
growing literature on K and matching logic.
We will consider the manual complete when it provides a complete description of
all user-facing K tools and features.

Introduction to K

Since K specifications are the primary input into the entire system, let us
take a moment to describe them. At the highest level, K specifications describe
a programming language or system using three different pieces:

  1. the system primitives, the base datatypes used during system operation,
    e.g., numbers, lists, maps, etc;
  2. the system state, a tuple or record over system primitives which gives a
    complete snapshot of the system at any given moment;
  3. the system behavior, a set of rules which defines possible system
    evolutions.

K specifications are then defined by a collection of sentences which
correspond to the three concepts above:

  1. syntax declarations encode the system primitives;
  2. configuration declarations encode the system state;
  3. context and rule declarations encode the system behavior.

K sentences are then organized into one or modules which are stored in one or
more files. In this scheme, files may require other files and modules may
import other modules, giving rise to a hierarchy of files and modules. We
give an intuitive sketch of the two levels of grouping in the diagram below:

   example.k file
  +=======================+
  | requires ".." --------|--> File_1
  | ...                   |
  | requires ".." --------|--> File_N
  |                       |
  |  +-----------------+  |
  |  | module ..       |  |
  |  |   imports .. ---|--|--> Module_1
  |  |   ...           |  |
  |  |   imports .. ---|--|--> Module_M
  |  |                 |  |
  |  |   sentence_1    |  |
  |  |   ...           |  |
  |  |   sentence_K    |  |
  |  | endmodule       |  |
  |  +-----------------+  |
  |                       |
  +=======================+

where:

  • files and modules are denoted by double-bordered and single-borded boxes
    respectively;
  • file or module identifiers are denoted by double dots (..);
  • potential repititions are denoted by triple dots (...).

In the end, we require that the file and module hierarchies both form a
directed acyclic graph (DAG). This is, no file may recursively require itself,
and likewise, no module may recursively import itself.

We now zoom in further to discuss the various kinds of sentences contained in K
specifications:

  1. sentences that define our system's primitives, including:

    • sort declarations: define new categories of primitive datatypes
    • Backus-Naur Form (BNF) grammar declarations: define the
      operators that inhabit our primitive datatypes
    • lexical syntax declarations: define lexemes/tokens for the
      lexer/tokenizer
    • syntax associativity declarations: specify the
      associativity/grouping of our declared operators
    • syntax priority declarations: specify the priority of
      potential ambiguous operators
  2. sentences that define our system's state, including:

    • configuration declarations: define labelled, hierarchical records
      using an nested XML-like syntax
  3. sentences that define our system's behavior, including:

    • context declarations: describe how primitives and configurations
      can simplify
    • context alias declarations: define templates that can generate new
      contexts
    • rule declarations: define how the system transitions from one state
      to the next

K Process Overview

We now examine how the K tools are generally used. The main input to all of the
K tools is a K specification. For effieciency reasons, this specification is
first compiled into an intermediate representation called Kore. Once we have
obtained this intermediate representation, we can use it to do:

  1. parsing/pretty-printing, i.e., converting a K term, whose syntax is defined
    by a K specification, into a alternate representation
  2. concrete and abstract execution of a K specification
  3. theorem proving, i.e., verifying whether a set of claims about a K
    specification hold

We represent the overall process using the graphic below:

 K Compilation Process
+============================================================+
|                     +---------+                            |
|  K Specification ---| kompile |--> Kore Specification --+  |
|                     +---------+                         |  |
+=========================================================|==+
                                                          |
 K Execution Process                                      |
+=========================================================|==+
|                                                         |  |
|             +-------------------------------------------+  |
|             |                                              |
|             |       +---------+                            |
|  K Term ----+-------| kparse  |--> K Term                  |
|             |       +---------+                            |
|             |                                              |
|             |       +---------+                            |
|  K Term ----+-------|  krun   |--> K Term                  |
|             |       +---------+                            |
|             |                                              |
|             |       +---------+                            |
|  K Claims --+-------| kprove  |--> K Claims                |
|                     +---------+                            |
|                                                            |
+============================================================+

where:

  • process outlines are denoted by boxes with double-lined borders
  • executables are denoted by boxes with single-lined borders
  • inputs and outputs are denoted by words attached to lines
  • K terms typically correspond to programs defined in a particular
    language's syntax (which are either parsed using kparse or executed using
    krun)
  • K claims are a notation for describing how certain K programs should
    execute (which are checked by our theorem prover kprove)

K Compilation Process:
Let us start with a description of the compilation process. According to the
above diagram, the compiler driver is called kompile. For our purposes, it is
enough to view the K compilation process as a black box that transforms a K
specification into a lower-level Kore specification that encodes the same
information, but that is easier to work with programmatically.

K Execution Process:
We now turn our attention to the K execution process. Abstractly, we can divide
the K execution process into the following stages:

  1. the kore specification is loaded (which defines a lexer, parser, and
    unparser among other things)
  2. the input string is lexed into a token stream
  3. the token stream is parsed into K terms/claims
  4. the K term/claims are transformed according the K tool being used (e.g.
    kparse, krun, or kprove)
  5. the K term/claims are unparsed into a string form and printed

Note that all of the above steps performed in K execution process are fully
prescribed by the input K specification. Of course, there are entire languages
devoted to encoding these various stages proces individually, e.g., flex for
lexers, bison for parsers, etc. What K offers is a consistent language to
package the above concepts in a way that we believe is convenient and practical
for a wide range of uses.

Module Declaration

K modules are declared at the top level of a K file. They begin with the
module keyword and are followed by a module ID and an optional set of
attributes. They continue with zero or more imports and zero or more sentences
until the endmodule keyword is reached.

A module ID consists of an optional # at the beginning, followed by one or
more components separated by hyphens. Each component can contain letters,
numbers, or underscores.

After the module ID, attributes can be specified in square brackets. See below
for an (incomplete) list of allowed module attributes.

Following the attributes, a module can contain zero or more imports. An
import consists of the import or imports keywords followed by a module ID.
An import tells the compiler that this module should contain all the sentences
(recursively) contained by the module being imported.

Imports can be public or private. By default, they are public, which
means that all the imported syntax can be used by any module that imports the
module doing the import. However, you can explicitly override the visibility
of the import with the public or private keyword immediately prior to the
module name. A module imported privately does not export its syntax to modules
that import the module doing the import.

Following imports, a module can contain zero or more sentences. A sentence can
be a syntax declaration, a rule, a configuration declaration, a context, a
claim, or a context alias. Details on each of these can be found in subsequent
sections.

private attribute

If the module is given the private attribute, all of its imports and syntax
are private by default. Individual pieces of syntax can be made public with
the public attribute, and individual imports can be made public with the
public keyword. See relevant sections on syntax and modules for more details
on what it means for syntax and imports to be public or private.

symbolic and concrete attribute

These attributes may be placed on modules to indicate that they should only
be used by the Haskell and LLVM backends respectively. If the definition is
compiled on the opposite backend, they are implicitly removed from the
definition prior to parsing anywhere they are imported. This can be useful when
used in limited capacity in order to provide alternate semantics for certain
features on different backends. It should be used sparingly as it makes it more
difficult to trust the correctness of your semantics, even in the presence of
testing.

Syntax Declaration

Named Non-Terminals

We have added a syntax to Productions which allows non-terminals to be given a
name in productions. This significantly improves the ability to document K, by
providing a way to explicitly explain what a field in a production corresponds
to instead of having to infer it from a comment or from the rule body.

The syntax is:

name: Sort

This syntax can be used anywhere in a K definition that expects a non-terminal.

symbol(_) attribute

By default, when compiling a definition, K generates a unique "mangled" label
identifier for each syntactic production. These identifiers can be used to
reference productions externally, for example when constructing terms by hand
or programmatically via Pyk.

The symbol(_) attribute can be applied to a production to control the precise
identifier for a production that appears in a compiled definition. For example:

module SYMBOLS
    syntax Foo ::= foo() [symbol(foo)]
                 | bar()
endmodule

Here, the compiled definition will contain the following symbol declarations:

  symbol Lblfoo{}() ...
  symbol Lblbar'LParRParUnds'SYMBOLS'Unds'Foo{}() ...

The compiler enforces uniqueness[1] of symbol names specified in
this way; it would be an error to apply symbol(foo) to another production in
the module above. Additionally, symbol(_) with an argument may not co-occur
with the klabel(_) attribute (see below).

overload attribute

K supports subsort overloading[2] on symbols, whereby a
constructor can have a more specific sort for certain arguments. For example,
consider the following productions derived from a C-like language semantics:

syntax Exp  ::= LVal
              | Exp  "." Id
syntax LVal ::= LVal "." Id

Here, it is useful for the result of the dot operator to be an LVal if the
left-hand side is itself an LVal. However, there is an issue with the code
as written: if L() is a term of sort LVal, then the program L() . x has a
parsing ambiguity between the two productions for the dot operator. To resolve
this, we can mark the productions as overloads:

syntax Exp  ::= LVal
              | Exp  "." Id [overload(_._)]
syntax LVal ::= LVal "." Id [overload(_._)]

Now, the parser will select the most specific overloaded production when it
resolves ambiguities in L() . x (that is, L() . x parses to a term of sort
LVal.

Formally, the compiler organises productions into a partial order that defines
the overload relation as follows. We say that P is a more specific overload
of Q if:

  • P and Q have the same overload(_) attribute. Note that the argument
    supplied has no semantic meaning other than as a key grouping productions
    together.
  • Let S_P be the sort of P, and S_p1 etc. be the sorts of its arguments
    (c.f. for Q). The tuple (S_P, S_p1, ..., S_pN) must be elementwise
    strictly less than (S_Q, S_q1, ..., S_qN) according to the definition's
    subsorting relationship. That is, a term from production P is a restriction
    of one from production Q; when its arguments are more precise, we can give
    the result a more precise sort.

klabel(_) and symbol attributes

Note: the klabel(_), symbol approach described in this section is a legacy
feature that will be removed in the future. New code should use the symbol(_)
and overload(_) attributes to opt into explicit naming and overloading
respectively.

References here to "overloading" are explained in the section above; the use
of the klabel(_) attribute without symbol is equivalent to the new
overload(_) syntax.

By default K generates for each syntax definition a long and obfuscated klabel
string, which serves as a unique internal identifier and also is used in kast
format of that syntax. If we need to reference a certain syntax production
externally, we have to manually define the klabels using the klabel attribute.
One example of where you would want to do this is to be able to refer to a given
symbol via the syntax priority attribute, or to enable overloading of a
given symbol.

If you only provide the klabel attribute, you can use the provided klabel to
refer to that symbol anywhere in the frontend K code. However, the internal
identifier seen by the backend for that symbol will still be the long obfuscated
generated string. Sometimes you want control over the internal identifier used as
well, in which case you use the symbol attribute. This tells the frontend to
use whatever the declared klabel is directly as the internal identifier.

For example:

module MYMODULE
    syntax FooBarBaz ::= #Foo( Int, Int ) [klabel(#Foo), symbol] // symbol1
                       | #Bar( Int, Int ) [klabel(#Bar)]         // symbol2
                       | #Baz( Int, Int )                        // symbol3
endmodule

Here, we have that:

  • In frontend K, you can refer to "symbol1" as #Foo (from klabel(#Foo)),
    and the backend will see 'Hash'Foo as the symbol name.
  • In frontend K, you can refer to "symbol2" as #Bar (from klabel(#Bar)),
    and the backend will see
    'Hash'Bar'LParUndsCommUndsRParUnds'MYMODULE'Unds'FooBarBaz'Unds'Int'Unds'Int
    as the symbol name.
  • In frontend K, you can refer to "symbol3" as
    #Baz(_,_)_MYMODULE_FooBarBaz_Int_Int (from auto-generated klabel), and
    the backend will see
    'Hash'Baz'LParUndsCommUndsRParUnds'MYMODULE'Unds'FooBarBaz'Unds'Int'Unds'Int
    as the symbol name.

The symbol provided must be unique to this definition. This is enforced by
K. In general, it's recommended to use the symbol attribute whenever you use
klabel unless you explicitly have a reason not to (e.g. you want to overload
symbols, or you're using a deprecated backend). It can be very helpful use the
symbol attribute for debugging, as many debugging messages are printed in
Kast format which will be more readable with the symbol names you explicitly
declare. In addition, if you are programatically manipulating definitions via
the JSON Kast format, building terms using the user-provided pretty
symbol, klabel(...) is easier and less error-prone if the auto-generation
process for klabels changes.

Syntactic Lists

When using K's support for syntactic lists, a production like:

syntax Ints ::= List{Int, ","} [symbol(ints)]

will desugar into two productions:

syntax Ints ::= Int "," Ints [symbol(ints)]
syntax Ints ::= ".Ints"      [symbol(List{"ints"})]

Note that the symbol for the terminator of the list has been generated
automatically from the label on the original production. It is possible to
control what the terminator's label is using the terminator-symbol(_)
attribute. For example:

syntax Ints ::= List{Int, ","} [symbol(ints), terminator-symbol(.ints)]

will desugar into two productions:

syntax Ints ::= Int "," Ints [symbol(ints)]
syntax Ints ::= ".Ints"      [symbol(.ints)]

It is an error to apply terminator-symbol(_) to a non-production sentence, or
to a production that does not declare a syntactic list.

Parametric productions and bracket attributes

Some syntax productions, like the rewrite operator, the bracket operator, and
the #if #then #else #fi operator, cannot have their precise type system
expressed using only concrete sorts.

Prior versions of K solved this issue by using the K sort in this case, but
this introduces inexactness in which poorly typed terms can be created even
without having a cast operator present in the syntax, which is a design
consideration we would prefer to avoid.

It also introduces cases where terms cannot be placed in positions where they
ought to be well sorted unless their return sort is made to be KBott, which in
turn vastly complicates the grammar and makes parsing much slower.

In order to introduce this, we provide a new syntax for parametric productions
in K. This allows you to express syntax that has a sort signature based on
parametric polymorphism. We do this by means of an optional curly-brace-
enclosed list of parameters prior to the return sort of a production.

Some examples:

syntax {Sort} Sort ::= "(" Sort ")" [bracket]
syntax {Sort} KItem ::= Sort
syntax {Sort} Sort ::= KBott
syntax {Sort} Sort ::= Sort "=>" Sort
syntax {Sort} Sort ::= "#if" Bool "#then" Sort "#else" Sort "#fi"
syntax {Sort1, Sort2} Sort1 ::= "#fun" "(" Sort2 "=>" Sort1 ")" "(" Sort2 ")"

Here we have:

  1. Brackets, which can enclose any sort but should be of the same sort that was
    enclosed.
  2. Every sort is a KItem.
  3. A KBott term can appear inside any sort.
  4. Rewrites, which can rewrite a value of any sort to a value of the same sort.
    Note that this allows the lhs or rhs to be a subsort of the other.
  5. If then else, which can return any sort but which must contain that sort on
    both the true and false branches.
  6. lambda applications, in which the argument and parameter must be the same
    sort, and the return value of the application must be the same sort as the
    return value of the function.

Note the last case, in which two different parameters are specified separated
by a comma. This indicates that we have multiple independent parameters which
must be the same each place they occur, but not the same as the other
parameters.

In practice, because every sort is a subsort of K, the Sort2
parameter in #6 above does nothing during parsing. It cannot
actually reject any parse, because it can always infer that the sort of the
argument and parameter are K, and it has no effect on the resulting sort of
the term. However, it will nevertheless affect the kore generated from the term
by introducing an additional parameter to the symbol generated for the term.

function and total attributes

Many times it becomes easier to write a semantics if you have "helper"
functions written which can be used in the RHS of rules. The function
attribute tells K that a given symbol should be simplified immediately when it
appears anywhere in the configuration. Semantically, it means that evaluation
of that symbol will result in at most one return value (that is, the symbol is
a partial function).

The total attribute indicates that a symbol cannot be equal to matching logic
bottom; in other words, it has at least one value for every possible set of
arguments. It can be added to a production with the function attribute to
indicate to the symbolic reasoning engine that a given symbol is a
total function, that is it has exactly one return value for every possible
input. Other uses of the total attribute (i.e., on multi-valued symbols to
indicate they always have at least one value) are not yet implemented.

For example, here we define the _+Word_ total function and the _/Word_
partial function, which can be used to do addition/division modulo
2 ^Int 256. These functions can be used anywhere in the semantics where
integers should not grow larger than 2 ^Int 256. Notice how _/Word_ is
not defined when the denominator is 0.

syntax Int ::= Int "+Word" Int [function, total]
             | Int "/Word" Int [function]

rule I1 +Word I2 => (I1 +Int I2) modInt (2 ^Int 256)
rule I1 /Word I2 => (I1 /Int I2) modInt (2 ^Int 256) requires I2 =/=Int 0

freshGenerator attribute

In K, you can access "fresh" values in a given domain using the syntax
!VARNAME:VarSort (with the !-prefixed variable name). This is supported for
builtin sorts Int and Id already. For example, you can generate fresh
memory locations for declared identifiers as such:

rule <k> new var x ; => . ... </k>
     <env> ENV => ENV [ x <- !I:Int ] </env>
     <mem> MEM => MEM [ !I <- 0     ] </mem>

Each time a !-prefixed variable is encountered, a new integer will be used,
so each variable declared with new var _ ; will get a unique position in the
<mem>.

Sometimes you want to have generation of fresh constants in a user-defined
sort. For this, K will still generate a fresh Int, but can use a converter
function you supply to turn it into the correct sort. For example, here we can
generate fresh Foos using the freshFoo(_) function annotated with
freshGenerator.

syntax Foo ::= "a" | "b" | "c" | d ( Int )

syntax Foo ::= freshFoo ( Int ) [freshGenerator, function, total]

rule freshFoo(0) => a
rule freshFoo(1) => b
rule freshFoo(2) => c
rule freshFoo(I) => d(I) [owise]

rule <k> new var x ; => . ... </k>
     <env> ENV => ENV [ x <- !I:Int  ] </env>
     <mem> MEM => MEM [ !I <- !F:Foo ] </mem>

Now each newly allocated memory slot will have a fresh Foo placed in it.

token attribute

The token attribute signals to the Kore generator that the associated sort
will be inhabited by domain values. Sorts inhabited by domain values must not
have any constructors declared.

syntax Bytes [hook(BYTES.Bytes), token]

Converting between [token] sorts

You can convert between tokens of one sort via Strings by defining functions
implemented by builtin hooks.
The hook STRING.token2string allows conversion of any token to a string:

syntax String ::= FooToString(Foo)  [function, total, hook(STRING.token2string)]

Similarly, the hook STRING.string2Token allows the inverse:

syntax Bar ::= StringToBar(String) [function, total, hook(STRING.string2token)]

WARNING: This sort of conversion does NOT do any sort of parsing or validation.
Thus, we can create arbitary tokens of any sort:

StringToBar("The sun rises in the west.")

Composing these two functions lets us convert from Foo to Bar

syntax Bar ::= FooToBar(Foo) [function]
rule FooToBar(F) => StringToBar(FooToString(F))

Parsing comments, and the #Layout sort

Productions for the #Layout sort are used to describe tokens that are
considered "whitespace". The scanner removes tokens matching these productions
so they are not even seen by the parser. Below, we use it to define
lines begining with ; (semicolon) as comments.

syntax #Layout ::= r"(;[^\\n\\r]*)"    // Semi-colon comments
                 | r"([\\ \\n\\r\\t])" // Whitespace

prec attribute

Consider the following naive attempt at creating a language what syntax that
allows two types of variables: names that contain underbars, and names that
contain sharps/hashes/pound-signs:

syntax NameWithUnderbar ::= r"[a-zA-Z][A-Za-z0-9_]*"  [token]
syntax NameWithSharp    ::= r"[a-zA-Z][A-Za-z0-9_#]*" [token]
syntax Pgm ::= underbar(NameWithUnderbar)
             | sharp(NameWithSharp)

Although, it seems that K has enough information to parse the programs
underbar(foo) and sharp(foo) with, the lexer does not take into account
whether a token is being parsed for the sharp or for the underbar
production. It chooses an arbitary sort for the token foo (perhaps
NameWithUnderbar). Thus, during paring it is unable to construct a valid term
for one of those programs (sharp(foo)) and produces the error message:
Inner Parser: Parse error: unexpected token 'foo'.

Since calculating inclusions and intersections between regular expressions is
tricky, we must provide this information to K. We do this via the prec(N)
attribute. The lexer will always prefer longer tokens to shorter tokens.
However, when it has to choose between two different tokens of equal length,
token productions with higher precedence are tried first. Note that the default
precedence value is zero when the prec attribute is not specified.

For example, the BUILTIN-ID-TOKENS module defines #UpperId and #LowerId with
the prec(2) attribute.

  syntax #LowerId ::= r"[a-z][a-zA-Z0-9]*"                    [prec(2), token]
  syntax #UpperId ::= r"[A-Z][a-zA-Z0-9]*"                    [prec(2), token]

Furthermore, we also need to make sorts with more specific tokens subsorts of ones with more
general tokens. We add the token attribute to this production so that all
tokens of a particular sort are marked with the sort they are parsed as and not a
subsort thereof. e.g. we get underbar(#token("foo", "NameWithUnderbar"))
instead of underbar(#token("foo", "#LowerId"))

imports BUILTIN-ID-TOKENS
syntax NameWithUnderbar ::= r"[a-zA-Z][A-Za-z0-9_]*" [prec(1), token]
                          | #UpperId                [token]
                          | #LowerId                [token]
syntax NameWithSharp ::= r"[a-zA-Z][A-Za-z0-9_#]*" [prec(1), token]
                       | #UpperId                 [token]
                       | #LowerId                 [token]
syntax Pgm ::= underbar(NameWithUnderbar)
             | sharp(NameWithSharp)

unused attribute

K will warn you if you declare a symbol that is not used in any of the rules of
your definition. Sometimes this is intentional, however; in this case, you can
suppress the warning by adding the unused attribute to the production or
cell.

syntax Foo ::= foo() [unused]

configuration <foo unused=""> .K </foo>

deprecated attribute

Symbols can be marked as deprecated by adding the deprecated attribute to
their declaration. If that symbol subsequently appears in the definition (in a
rule, context, context alias or configuration), the compiler will issue a
warning.

syntax Foo ::= foo() [deprecated]
rule foo() => . // warning on this line

Symbol priority and associativity

Unlike most other parser generators, K combines the task of parsing with AST
generation. A production declared with the syntax keyword in K is both a
piece of syntax used when parsing, and a symbol that is used when rewriting.
As a result, it is generally convenient to describe expression grammars using
priority and associativity declarations rather than explicitly transforming
your grammar into a series of nonterminals, one for each level of operator
precedence. Thus, for example, a simple grammar for addition and multiplication
will look like this:

syntax Exp ::= Exp "*" Exp
             | Exp "+" Exp

However, this grammar is ambiguous. The term x+y*z might refer to x+(y*z)
or to (x+y)*z. In order to differentiate this, we introduce a partial
ordering between productions known as priority. A symbol "has tighter priority"
than another symbol if the first symbol can appear under the second, but the
second cannot appear under the first without a bracket. For example, in
traditional arithmetic, multiplication has tighter priority than addition,
which means that x+y*z cannot parse as (x+y)*z because the addition
operator would appear directly beneath the multiplication, which is forbidden
by the priority filter.

Priority is applied individually to each possible ambiguous parse of a term. It
then either accepts or rejects that parse. If there is only a single remaining
parse (after all the other disambiguation steps have happened), this is the
parse that is chosen. If all the parses were rejected, it is a parse error. If
multiple parses remain, they might be resolved by further disambiguation such
as via the prefer and avoid attributes, but if multiple parses remain after
disambiguation finishes, this is an ambiguous parse error, indicating there is
not a unique parse for that term. In the vast majority of cases, this is
an error and indicates that you ought to either change your grammar or add
brackets to the term in question.

Priority is specified in K grammars by means of one of two different
mechanisms. The first, and simplest, simply replaces the | operator in a
sequence of K productions with the > operator. This operator indicates that
everything prior to the > operator (including transitively) binds tighter
than what comes after. For example, a more complete grammar for simple
arithmetic might be:

syntax Exp ::= Exp "*" Exp
             | Exp "/" Exp
             > Exp "+" Exp
             | Exp "-" Exp

This indicates that multiplication and division bind tigher than addition
and subtraction, but that there is no relationship in priority between
multiplication and division.

As you may have noticed, this grammar is also ambiguous. x*y/z might refer to
x*(y/z) or to (x*y)/z. Indeed, if we removed division and subtraction
entirely, the grammar would still be ambiguous: x*y*z might parse as
x*(y*z), or as (x*y)*z. To resolve this, we introduce another feature:
associativity. Roughly, asssociativity tells us how symbols are allowed to nest
within other symbols with the same priority. If a set of symbols is left
associative, then symbols in that set cannot appear as the rightmost child
of other symbols in that set. If a set of symbols is right associative, then
symbols in that set cannot appear as the leftmost child of other symbols in
that set. Finally, if a set of symbols is non-associative, then symbols
in that set cannot appear as the rightmost or leftmost child of other symbols
in that set. For example, in the above example, if addition and subtraction
are left associative, then x+y+z will parse as (x+y)+z and x+y-z will
parse as (x+y)-z (because the other parse will have been rejected).

You might notice that this seems to apply only to binary infix operators. In
fact, the real behavior is slightly more complicated. Priority and
associativity (for technical reasons that go beyond the scope of this document)
really only apply when the rightmost or leftmost item in a production is a
nonterminal. If the rightmost nonterminal is followed by a terminal (or
respectively the leftmost preceded), priority and associativity do not apply.
Thus we can generalize these concepts to arbitrary context-free grammars.

Note that in some cases, this is not the behavior you want. You may actually
want to reject parses even though the leftmost and rightmost item in a
production are terminals. You can accomplish this by means of the
applyPriority attribute. When placed on a production, it tells the parser
which nonterminals of a production the priority filter ought to reject children
under, overriding the default behavior. For example, I might have a production
like syntax Exp ::= foo(Exp, Exp) [applyPriority(1)]. This tells the parser
to reject terms with looser priority binding under the first Exp, but not
the second. By default, with this production, neither position would apply
to the priority filter, because the first and last items of the production
are both terminals.

Associativity is specified in K grammars by means of one of two different
mechanisms. The first, and simplest, adds the associativity of a priority block
of symbols prior to that block. For example, we can remove the remaining
ambiguities in the above grammar like so:

syntax Exp ::= left:
               Exp "*" Exp
             | Exp "/" Exp
             > right:
               Exp "+" Exp
             | Exp "-" Exp

This indicates that multiplication and division are left-associative, ie, after
symbols with higher priority are parsed as innermost, symbols are nested with
the rightmost on top. Addition and subtraction are right associative, which
is the opposite and indicates that symbols are nested with the leftmost on top.
Note that this is similar but different from evaluation order, which also
concerns itself with the ordering of symbols, which is described in the next
section.

You may note we have not yet introduced the second syntax for priority
and associativity. In some cases, syntax for a grammar might be spread across
multiple modules, sometimes for very good reasons with respect to code
modularity. As a result, it becomes infeasible to declare priority and
associativity inline within a set of productions, because the productions
are not contiguous within a single file.

For this purpose, we introduce the equivalent syntax priority,
syntax left, syntax right, and syntax non-assoc declarations. For
example, the above grammar can be written equivalently as:

syntax Exp ::= Exp "*" Exp [group(mult)]
             | Exp "/" Exp [group(div)]
             | Exp "+" Exp [group(add)]
             | Exp "-" Exp [group(sub)]

syntax priority mult div > add sub
syntax left mult div
syntax right add sub

Here, the group(_) attribute is used to create user-defined groups of
sentences. A particular group name collectively refers to the whole set of
sentences within that group. The sets are flattened together, so we could
equivalently have written:

syntax Exp ::= Exp "*" Exp [group(mult)]
             | Exp "/" Exp [group(mult)]
             | Exp "+" Exp [group(add)]
             | Exp "-" Exp [group(add)]

syntax priority mult > add
syntax left mult
syntax right add

Note that syntax [left|right|non-assoc] should not be used to group together
productions with different priorities. For example, this code would be invalid:

syntax priority mult > add
syntax left mult add

Note that there is one other way to describe associativity, but it is
prone to a very common mistake. You can apply the attribute left, right,
or non-assoc directly to a production to indicate that it is, by itself,
left-, right-, or non-associative.

However, this often does not mean what users think it means. In particular:

syntax Exp ::= Exp "+" Exp [left]
             | Exp "-" Exp [left]

is not equivalent to:

syntax Exp ::= left:
               Exp "+" Exp
             | Exp "-" Exp

Under the first, each production is associative with itself, but not each
other. Thus, x+y+z will parse unambiguously as (x+y)+z, but x+y-z will
be ambiguous. However, in the second, x+y-z will parse unambiguously as
(x+y)-z.

Think carefully about how you want your grammar to parse. In general, if you're
not sure, it's probably best to group associativity together into the same
blocks you use for priority, rather than using left, right, or non-assoc
attributes on the productions.

Lexical identifiers

Sometimes it is convenient to be able to give a certain regular expression a
name and then refer to it in one or more regular expression terminals. This
can be done with a syntax lexical sentence in K:

syntax lexical Alphanum = r"[0-9a-zA-Z]"

This defines a lexical identifier Alphanum which can be expanded in any
regular expression terminal to the above regular expression. For example, I
might choose to then implement the syntax of identifiers as follows:

syntax Id ::= r"[a-zA-Z]{Alphanum}*" [token]

Here {Alphanum} expands to the above regular expression, making the sentence
equivalent to the following:

syntax Id ::= r"[a-zA-Z]([0-9a-zA-Z])*" [token]

This feature can be used to more modularly construct the lexical syntax of your
language. Note that K does not currently check that lexical identifiers used
in regular expressions have been defined; this will generate an error when
creating the scanner, however, and the user ought to be able to debug what
happened.

assoc, comm, idem, and unit attributes

These attributes are used to indicate whether a collection or a production
is associative, commutative, idempotent, and/or has a unit.
In general, you should not need to apply these attributes to productions
yourself, however, they do have certain special meaning to K. K will generate
axioms related to each of these concepts into your definition for you
automatically. It will also automatically sort associative-commutative
collections, and flatten the indentation of associative collections, when
unparsing.

public and private attribute

K allows users to declare certain pieces of syntax as either public or private.
All syntax is public by default. Public syntax can be used from any module that
imports that piece of syntax. A piece of syntax can be declared private with
the private attribute. This means that that syntax can only be used in the
module in which it is declared; it is not visible from modules that import
that module.

You can also change the default visibility of a module with the private
attribute, when it is placed directly on a module. A module with the private
attribute has all syntax private by default; this can be overridden on
specific sentences with the public attribute.

Note that the private module attribute also changes the default visiblity
of imports; please refer to the appropriate section elsewhere in the manual
for more details.

Here is an example usage:

module WIDGET-SYNTAX

  syntax Widget ::= foo()
  syntax WidgetHelper ::= bar() [private] // this production is not visible
                                          // outside this module
endmodule

module WIDGET [private]
  imports WIDGET-SYNTAX

  syntax Widget ::= fooImpl() // this production is not visible outside this
                              // module

  // this production is visible outside this module
  syntax KItem ::= adjustWidget(Widget) [function, public]
endmodule

Configuration Declaration

exit attribute

A single configuration cell containing an integer may have the "exit"
attribute. This integer will then be used as the return value on the console
when executing the program.

For example:

configuration <k> $PGM:Pgm </k>
              <status-code exit=""> 1 </status-code>

declares that the cell status-code should be used as the exit-code for
invocations of krun. Additionally, we state that the default exit-code is 1
(an error state). One use of this is for writing testing harnesses which assume
that the test fails until proven otherwise and only set the <status-code> cell
to 0 if the test succeeds.

Collection Cells: multiplicity and type attributes

Sometimes a semantics needs to allow multiple copies of the same cell, for
example if you are making a concurrent multi-threading programming language.
For this purpose, K supports the multiplicity and type attributes on cells
declared in the configuration.

multiplicity can take on values * and ?. Declaring multiplicity="*"
indicates that the cell may appear any number of times in a runtime
configuration. Setting multiplicity="?" indicates that the cell may only
appear exactly 0 or 1 times in a runtime configuration. If there are no
configuration variables present in the cell collection, the initial
configuration will start with exactly 0 instances of the cell collection. If
there are configuration variables present in the cell collection, the initial
configuration will start with exactly 1 instance of the cell collection.

type can take on values Set, List, and Map. For example, here we declare
several collecion cells:

configuration <k> $PGM:Pgm </k>
              <sets>  <set  multiplicity="?" type="Set">  0:Int </set>  </sets>
              <lists> <list multiplicity="*" type="List"> 0:Int </list> </lists>
              <maps>
                <map multiplicity="*" type="Map">
                  <map-key> 0:Int </map-key>
                  <map-value-1> "":String </map-value-1>
                  <map-value-2> 0:Int     </map-value-2>
                </map>
              </maps>

Declaring type="Set" indicates that duplicate occurrences of the cell should
be de-duplicated, and accesses to instances of the cell will be nondeterministic
choices (constrained by any other parts of the match and side-conditions).
Similarly, declaring type="List" means that new instances of the cell can be
added at the front or back, and elements can be accessed from the front or back,
and the order of the cells will be maintained. The following are examples of
introduction and elimination rules for these collections:

rule <k> introduce-set(I:Int) => . ... </k>
     <sets> .Bag => <set> I </set> </sets>

rule <k> eliminate-set => I ... </k>
     <sets> <set> I </set> => .Bag </sets>

rule <k> introduce-list-start(I:Int) => . ... </k>
     <lists> (.Bag => <list> I </list>) ... </lists>

rule <k> introduce-list-end(I:Int) => . ... </k>
     <lists> ... (.Bag => <list> I </list>) </lists>

rule <k> eliminate-list-start => I ... </k>
     <lists> (<list> I </list> => .Bag) ... </lists>

rule <k> eliminate-list-end => I ... </k>
     <lists> ... (<list> I </list> => .Bag) </lists>

Notice that for multiplicity="?", we only admit a single <set> instance at
a time. For the type=List cell, we can add/eliminate cells from the from or
back of the <lists> cell. Also note that we use .Bag to indicate the empty
cell collection in all cases.

Declaring type="Map" indicates that the first sub-cell will be used as a
cell-key. This means that matching on those cells will be done as a map-lookup
operation if the cell-key is mentioned in the rule (for performance). If the
cell-key is not mentioned, it will fallback to normal nondeterministic
constrained by other parts of the match and any side-conditions. Note that there
is no special meaning to the name of the cells (in this case <map>,
<map-key>, <map-value-1>, and <map-value-2>). Additionally, any number of
sub-cells are allowed, and the entire instance of the cell collection is
considered part of the cell-value, including the cell-key (<map-key> in this
case) and the surrounding collection cell (<map> in this case).

For example, the following rules introduce, set, retrieve from, and eliminate
type="Map" cells:

rule <k> introduce-map(I:Int) => . ... </k>
     <maps> ... (.Bag => <map> <map-key> I </map-key> ... </map>) ... </maps>

rule <k> set-map-value-1(I:Int, S:String) => . ... </k>
     <map> <map-key> I </map-key> <map-value-1> _ => S </map-value-1> ... </map>

rule <k> set-map-value-2(I:Int, V:Int) => . ... </k>
     <map> <map-key> I </map-key> <map-value-2> _ => V </map-value-2> ... </map>

rule <k> retrieve-map-value-1(I:Int) => S ... </k>
     <map> <map-key> I </map-key> <map-value-1> S </map-value-1> ... </map>

rule <k> retrieve-map-value-2(I:Int) => V ... </k>
     <map> <map-key> I </map-key> <map-value-2> V </map-value-2> ... </map>

rule <k> eliminate-map(I:Int) => . ... </k>
     <maps> ... (<map> <map-key> I </map-key> ... </map> => .Bag) ... </maps>

Note how each rule makes sure that <map-key> cell is mentioned, and we
continue to use .Bag to indicate the empty collection. Also note that
when introducing new map elements, you may omit any of the sub-cells which are
not the cell-key. In case you do omit sub-cells, you must use structural
framing ... to indicate the missing cells, they will receive the default
value given in the configuration ... declaration.

Rule Declaration

Rule Structure

Each K rule follows the same basic structure (given as an example here):

rule LHS => RHS requires REQ ensures ENS [ATTRS]

The portion between rule and requires is referred to as the rule body,
and may contain one or more rewrites (though not nested). Here, the rule body is
LHS => RHS, where LHS and RHS are used as placeholders for the pre- and
post- states. Note that we lose no generality referring to the LHS or the
RHS, even in the presence of multiple rewrites, as the rewrites are pulled to
the top-level anyway.

Next is the requires clause, represented here as REQ. The requires clause is
an additional predicate (function-like term of sort Bool), which is to be
evaluated before applying the rule. If the requires clause does not evaluate to
true, then the rule does not apply.

Finally is the ensures clause, represented here as ENS. The ensures clause
is to be interpreted as a post-condition, and will be automatically added to the
path condition if the rule applies. It may cause the entire term to become
undefined, but the backend will not stop itself from applying the rule in this
case. Note that concrete backends (eg. the LLVM backend) are free to ignore the
ensures clause.

Overall, the transition represented by such a rule is from a state
LHS #And REQ ending in a state RHS #And ENS. When backends apply this rule
as a transition/rewrite, they should:

  • Check if pattern LHS matches (or unifies) with the current term, giving
    substitution alpha.
  • Check if the instantiation alpha(REQ) is valid (or satisfiable).
  • Build the new term alpha(RHS #And ENS), and check if it's satisfiable.

Pattern Matching operator

Sometimes when you want to express a side condition, you want to say that a
rule matches if a particular term matches a particular pattern, or if it
instead does /not/ match a particular pattern.

The syntax in K for this is :=K and :/=K. It has similar meaning to ==K and
=/=K, except that where ==K and =/=K express equality, :=K and =/=K express
model membership. That is to say, whether or not the rhs is a member of the set
of terms expressed by the lhs pattern. Because the lhs of these operators is a
pattern, the user can use variables in the lhs of the operator. However, due to
current limitations, these variables are NOT bound in the rest of the term.
The user is thus encouraged to use anonymous variables only, although this is
not required.

This is compiled by the K frontend down to an efficient pattern matching on a
fresh function symbol.

Anonymous function applications

There are a number of cases in K where you would prefer to be able to take some
term on the RHS, bind it to a variable, and refer to it in multiple different
places in a rule.

You might also prefer to take a variable for which you know some of its
structure, and modify some of its internal structure without requiring you to
match on every single field contained inside that structure.

In order to do this, we introduce syntax to K that allows you to construct
anonymous functions in the RHS of a rule and apply them to a term.

The syntax for this is:

#fun(RuleBody)(Argument)

Note the limitations currently imposed by the implementation. These functions
are not first-order: you cannot bind them to a variable and inject them like
you can with a regular klabel for a function. You also cannot express multiple
rules or multiple parameters, or side conditions. All of these are extensions
we would like to support in the future, however.

In the following, we use three examples to illustrate the behavior of #fun.
We point out that the support for #fun is provided by the frontend, not the
backends.

The three examples are real examples borrowed or modified from existing language
semantics.

Example 1 (A Simple Self-Explained Example).

#fun(V:Val => isFoo(V) andBool isBar(V))(someFunctionReturningVal())

Example 2 (Nested #fun).

   #fun(C
=> #fun(R
=> #fun(E
=> foo1(E, R, C)
  )(foo2(C))
  )(foo3(0))
  )(foo4(1))

This example is from the beacon
semantics:https://github.com/runtimeverification/beacon-chain-spec/blob/master/b
eacon-chain.k at line 302, with some modification for simplicity. Note how
variables C, R, E are bound in the nested #fun.

Example 3 (Matching a structure).

rule foo(K, RECORD) =>
  #fun(record(... field: _ => K))(RECORD)

Unlike previous examples, the LHS of #fun in this example is no longer a
variable, but a structure. It has the same spirit as the first two examples,
but we match the RECORD with a structure record( DotVar, field: X), instead
of a standalone variable. We also use K's local rewrite syntax (i.e., the
rewriting symbol => does not occur at the top-level) to prevent writing
duplicate expressions on the LHS and RHS of the rewriting.

Macros and Aliases

A production can be tagged with the macro, alias, macro-rec, or alias-rec
attributes. In all cases, what this signifies is that this is a macro production.
Macro rules are rules where the top symbol of the left-hand-side are macro
labels. Macro rules are applied statically during compilation on all terms that
they match, and statically before program execution on the initial configuration.
Currently, macro rules are required to not have side conditions, although they
can contain sort checks.

alias rules are also applied statically in reverse prior to unparsing on the
final configuration. Note that a macro rule can have unbound variables in the
right hand side. When such a macro exists, it should be used only on the left
hand side of rules, unless the user is performing symbolic execution and expects
to introduce symbolic terms into the subject being rewritten.

However, when used on the left hand side of a rule, it functions similarly to a
pattern alias, and allows the user to concisely express a reusable pattern that
they wish to match on in multiple places.

For example, consider the following semantics:

syntax KItem ::= "foo" [alias] | "foobar"
syntax KItem ::= bar(KItem) [macro] | baz(Int, KItem)
rule foo => foobar
rule bar(I) => baz(?_, I)
rule bar(I) => I

This will rewrite baz(0, foo) to foo. First baz(0, foo) will be rewritten
statically to baz(0, foobar). Then the non-macro rule will apply (because
the rule will have been rewritten to rule baz(_, I) => I). Then foobar will
be rewritten statically after rewriting finishes to foo via the reverse form
of the alias.

Note that macros do not apply recursively within their own expansion. This is
done so as to ensure that macro expansion will always terminate. If the user
genuinely desires a recursive macro, the macro-rec and alias-rec attributes
can be used to provide this behavior.

For example, consider the following semantics:

syntax Exp ::= "int" Exp ";" | "int" Exps ";" [macro] | Exp Exp | Id
syntax Exps ::= List{Exp,","}

rule int X:Id, X':Id, Xs:Exps ; => int X ; int X', Xs ;

This will expand int x, y, z; to int x; int y, z; because the macro does
not apply the second time after applying the substitution of the first
application. However, if the macro attribute were changed to the macro-rec
attribute, it would instead expand (as the user likely intended) to
int x; int y; int z;.

The alias-rec attribute behaves with respect to the alias attribute the
same way the macro-rec attribute behaves with respect to macro.

anywhere rules

Some rules are not functional, but you want them to apply anywhere in the
configuration (similar to functional rules). You can use the anywhere
attribute on a rule to instruct the backends to make sure they apply anywhere
they match in the entire configuration.

For example, if you want to make sure that some associative operator is always
right-associated anywhere in the configuration, you can do:

syntax Stmt ::= Stmt ";" Stmt

rule (S1 ; S2) ; S3 => S1 ; (S2 ; S3) [anywhere]

Then after every step, all occurrences of _;_ will be re-associated. Note that
this allows the symbol _;_ to still be a constructor, even though it is
simplified similarly to a function.

trusted claims

You may add the trusted attribute to a given claim for the K prover to
automatically add it to the list of proven circularities, instead of trying to
discharge it separately.

Projection and Predicate functions

K automatically generates certain predicate and projection functions from the
syntax you declare. For example, if you write:

syntax Foo ::= foo(bar: Bar)

It will automatically generate the following K code:

syntax Bool ::= isFoo(K) [function]
syntax Foo ::= "{" K "}" ":>Foo" [function]
syntax Bar ::= bar(Foo) [function]

rule isFoo(F:Foo) => true
rule isFoo(_) => false [owise]

rule { F:Foo }:>Foo => F
rule bar(foo(B:Bar)) => B

The first two types of functions are generated automatically for every sort in
your K definition, and the third type of function is generated automatically
for each named nonterminal in your definition. Essentially, isFoo for some
sort Foo will tell you whether a particular term of sort K is a Foo,
{F}:>Foo will cast F to sort Foo if F is of sort Foo and will be
undefined (i.e., theoretically defined as #Bottom, the bottom symbol in
matching logic) otherwise. Finally, bar will project out the child of a foo
named bar in its production declaration.

Note that if another term of equal or smaller sort to Foo exists and has a
child named bar of equal or smaller sort to Bar, this will generate an
ambiguity during parsing, so care should be taken to ensure that named
nonterminals are sufficiently unique from one another to prevent such
ambiguities. Of course, the compiler will generate a warning in this case.

simplification attribute

The simplification attribute identifies rules outside the main semantics that
are used to simplify function patterns.

Conditions: A simplification rule is applied by matching the function
arguments, instead of unification as when applying function definition
rules. This allows function symbols to appear nested as arguments to other
functions on the left-hand side of a simplification rule, which is forbidden in
function definition rules. For example, this rule would not be accepted as a
function definition rule:

rule (X +Int Y) +Int Z => X +Int (Y +Int Z) [simplification]

A simplification rule is only applied when the current side condition implies
the requires clause of the rule, like function definition rules.

Order: The simplification attribute accepts an optional integer argument
which is the rule's simplification priority; if the optional argument is not
specified, it is equivalent to a simplification priority of 50. Backends
should attempt simplification rules in order of their simplification
priority
, but are not required to do so; in fact, the backend is free to apply
simplification rules at any time. Because of this, users must ensure that
simplification rules are sound regardless of their order of application. This
differs from the priority attribute in that rules with the priority
attribute must be applied in their priority order by the backend. It is an
error to have the priority attribute on a simplification rule.

For example, for the following definition:

    syntax WordStack ::= Int ":" WordStack | ".WordStack"
    syntax Int ::= sizeWordStack    ( WordStack       ) [function]
                 | sizeWordStackAux ( WordStack , Int ) [function]
 // --------------------------------------------------------------
    rule sizeWordStack(WS) => sizeWordStackAux(WS, 0)

    rule sizeWordStackAux(.WordStack, N) => N
    rule sizeWordStackAux(W : WS    , N) => sizeWordStackAux(WS, N +Int 1)

We might add the following simplification lemma:

    rule sizeWordStackAux(WS, N) => N +Int sizeWordStackAux(WS, 0)
      requires N =/=Int 0
      [simplification]

Then this simplification rule will only apply if the Haskell backend can prove
that notBool N =/=Int 0 is unsatisfiable. This avoids an infinite cycle of
applying this simplification lemma.

NOTE: The frontend and Haskell backend do not check that supplied
simplification rules are sound, this is the developer's responsibility. In
particular, rules with the simplification attribute must preserve definedness;
that is, if the left-hand side refers to any partial function then:

  • the right-hand side must be #Bottom when the left-hand side is #Bottom, or
  • the rule must have an ensures clause that is false when the left-hand
    side is #Bottom, or
  • the rule must have a requires clause that is false when the left-hand
    side is #Bottom.

These conditions are in order of decreasing preference: the best option is to
preserve #Bottom on the right-hand side, the next best option is to have an
ensures clause, and the least-preferred option is to have a requires clause.
The most preferred option is to write total functions and avoid the entire issue.

NOTE: The Haskell backend does not attempt to prove claims which right-hand
side is #Bottom. The reason for this is that the general case is undecidable,
and the backend might enter an infinite loop. Therefore, the backend emits a
warning if it encounters such a claim.

concrete and symbolic attributes (Haskell backend)

Users can control the application of simplification rules using the concrete
and the symbolic attributes by specifying the type of patterns the rule's
arguments are to match.

A concrete pattern is a pattern which does not contain variables or unevaluated
functions, otherwise the pattern is symbolic.

The semantics of the two attributes is defined as follows:

  • If a simplification rule is marked concrete, then all arguments must be
    concrete for the rule to match.
  • If a simplification rule is marked symbolic, then all arguments must be
    symbolic for the rule to match.
  • The following syntax concrete(<variables>) (resp. symbolic(<variables>)),
    where <variables> is a list of variable names separated by commas, can be used
    to specify the exact arguments the user expects to match concrete (resp. symbolic)
    patterns.

For example, the following will only match when all arguments
are concrete:

rule X +Int (Y +Int Z) => (X +Int Y) +Int Z [simplification, concrete]

Conversely, the following will only match when all arguments
are symbolic:

rule X +Int (Y +Int Z) => (X +Int Y) +Int Z [simplification, symbolic]

In practice, the following rules will re-associate and commute terms to combine
concrete arguments:

rule (A +Int Y) +Int Z => A +Int (Y +Int Z)
  [concrete(Y, Z), symbolic(A), simplification]

rule X +Int (B +Int Z) => B +Int (X +Int Z)
  [concrete(X, Z), symbolic(B), simplification]

The unboundVariables attribute

Normally, K rules are not allowed to contain regular (i.e., not fresh, not
existential) variables in the RHS / requires / ensures clauses which are not
bound in the LHS.

However, in certain cases this behavior might be desired, like, for example,
when specifying a macro rule which is to be used in the LHS of other rules.
To allow for such cases, but still be useful and perform the unboundness checks
in regular cases, the unboundVariables attributes allows the user to specify
a comma-separated list of names of variables which can be unbound in the rule.

For example, in the macro declaration

  rule cppEnumType => bar(_, scopedEnum() #Or unscopedEnum() ) [unboundVariables(_)]

the declaration unboundVariables(_) allows the rule to pass the unbound
variable checks, and this in turn allows for cppEnumType to be used in
the LHS of a rule to mean the pattern above:

  rule inverseConvertType(cppEnumType, foo((cppEnumType #as T::CPPType => underlyingType(T))))

The memo attribute

The memo attribute is a hint from the user to the backend to memoize a
function. Not all backends support memoization, but when the attribute is used
and the definition is compiled for a memo-supporting backend, then calls to
the function may be cached. At the time of writing, only the Haskell
backend supports memoization.

Limitations of memoization with the Haskell backend

The Haskell backend will only cache a function call if all arguments are concrete.

It is recommended not to memoize recursive functions, as each recursive call
will be stored in the cache, but only the first iteration will be retrieved from
the cache; that is, the cache will be filled with many unreachable
entries. Instead, we recommend to perform a worker-wrapper transformation on
recursive functions, and apply the memo attribute to the wrapper.

Warning: A function declared with the memo attribute must not use
uninterpreted functions in the side-condition of any rule. Memoizing such an
impure function is unsound. To see why, consider the following rules:

syntax Bool ::= impure( Int ) [function]

syntax Int ::= unsound( Int ) [function, memo]
rule unsound(X:Int) => X +Int 1 requires impure(X)
rule unsound(X:Int) => X        requires notBool impure(X)

Because the function impure is not given rules to cover all inputs, unsound
can be memoized incoherently. For example,

{unsound(0) #And {impure(0) #Equals true}} #Equals 1

but

{unsound(0) #And {impure(0) #Equals false}} #Equals 0

The memoized value of unsound(0) would be incoherently determined by which
pattern the backend encounters first.

Variable Sort Inference

In K, it is not required that users declare the sorts of variables in rules or
in the initial configuration. If the user does not explicitly declare the sort
of a variable somewhere via a cast (see below), the sort of the variable is
inferred from context based on the sort signature of every place the variable
appears in the rule.

As an example, consider the rule for addition in IMP:

    syntax Exp ::= Exp "+" Exp | Int

    rule I1 + I2 => I1 +Int I2

Here +Int is defined in the INT module with the following signature:

    syntax Int ::= Int "+Int" Int [function]

In the rule above, the sort of both I1 and I2 is inferred as Int. This is because
a variable must have the same sort every place it appears within the same rule.
While a variable appearing only on the left-hand-side of the rule could have
sort Exp instead, the same variable appears as a child of +Int, which
constriants the sorts of I1 and I2 more tightly. Since the sort must be a
subsort of Int or equal to Int, and Int has no subsorts, we infer Int
as the sorts of I1 and I2. This means that the above rule will not match
until I1 and I2 become integers (i.e., have already been evaluated).

More complex examples are possible, however:

    syntax Exp ::= Exp "+" Int | Int
    rule _ + _ => 0

Here we have two anonymous variables. They do not refer to the same variable
as one another, so they can have different sorts. The right side is constrained
by + to be of sort Int, but the left side could be either Exp or Int.
When this occurs, we have multiple solutions to the sorts of the variables in
the rule. K will only choose solutions which are maximal, however. To be
precise, if two different solutions exist, but the sorts of one solution are
all greater than or equal to the sorts of the other solution, K will discard
the smaller solution. Thus, in the case above, the variable on the left side
of the + is inferred of sort Exp, because the solution (Exp, Int) is
strictly greater than the solution (Int, Int).

It is possible, however, for terms to have multiple maximal solutions:

    syntax Exp ::= Exp "+" Int | Int "+" Exp | Int
    rule I1 + I2 => 0

In this example, there is an ambiguous parse. This could parse as either
the first + or the second. In the first case, the maximal solution chosen is
(Exp, Int). In the second, it is (Int, Exp). Neither of these solutions is
greater than the other, so both are allowed by K. As a result, this program
will emit an error because the parse is ambiguous. To pick one solution over
the other, a cast or a prefer or avoid attribute can be used.

Casting

There are three main types of casts in K: the semantic cast, the strict cast,
and the projection cast.

Semantic casts

For every sort S declared in your grammar, K will define the following
production for you for use in rules:

    syntax S ::= S ":S"

The meaning of this cast is that the term inside the cast must be less than
or equal to Sort. This can be used to resolve ambiguities, but its principle
purpose is to guide execution by telling K what sort variables must match in
order for the rule to apply. When compiled, it will generate a pattern that
matches on an injection into Sort.

Strict casts

K also introduces the strict cast:

    syntax S ::= S "::S"

The meaning at runtime is exactly the same as the semantic cast; however, it
restricts the sort of the term inside the cast to exactly Sort. That is
to say, if you use it on something that is a strictly smaller sort, it will
generate a type error. This is useful in certain circumstances to help
disambiguate terms, when a semantic cast would not have resolved the ambiguity.
As such, it is primarily used to solve ambiguities rather than to guide
execution.

Projection casts

K also introduces the projection cast:

    syntax {S2} S ::= "{" S2 "}" ":>S"

The meaning of this cast at runtime is that if the term inside is of sort
Sort, it should have it injection stripped away and the value inside is
returned as a term of static sort Sort. However, if the term is of a
different sort, it is an error and execution will get stuck. Thus the primary
usefulness of this cast is to cast the return value of a function with a
greater sort down to a strictly smaller sort that you expect the return value
of the function to have. For example:

    syntax Exp ::= foo(Exp) [function] | bar(Int) | Int
    rule foo(I:Int) => I
    rule bar(I) => bar({foo(I +Int 1)}:>Int)

Here we know that foo(I +Int 1) will return an Int, but the return sort of
foo is Exp. So we project the result into the Int sort so that it can
be placed as the child of a bar.

owise and priority attributes.

Sometimes, it is simply not convenient to explicitly describe every
single negative case under which a rule should not apply. Instead,
we simply wish to say that a rule should only apply after some other set of
rules have been tried. K introduces two different attributes that can be
added to rules which will automatically generate the necessary matching
conditions in a manner which is performant for concrete execution (indeed,
it generally outperforms during concrete execution code where the conditions
are written explicitly).

The first is the owise attribute. Very roughly, rules without an attribute
indicating their priority apply first, followed by rules with the owise
attribute only if all the other rules have been tried and failed. For example,
consider the following function:

syntax Int ::= foo(Int) [function]
rule foo(0) => 0
rule foo(_) => 1 [owise]

Here foo(0) is defined explicitly as 0. Any other integer yields the
integer 1. In particular, the second rule above will only be tried after the
first rule has been shown not to apply.

This is because the first rule has a lower number assigned for its priority
than the second rule. In practice, each rule in your semantics is implicitly
or explicitly assigned a numerical priority. Rules are tried in increasing
order of priority, starting at zero and trying each increasing numerical value
successively.

You can specify the priority of a rule with the priority attribute. For
example, I could equivalently write the second rule above as:

rule foo(_) => 1 [priority(200)]

The number 200 is not chosen at random. In fact, when you use the owise
attribute, what you are doing is implicitly setting the priority of the rule
to 200. This has a couple of implications:

  1. Multiple rules with the owise attribute all have the same priority and thus
    can apply in any order.
  2. Rules with priority higher than 200 apply after all rules with the
    owise attribute have been tried.

There is one more rule by which priorities are assigned: a rule with no
attributes indicating its priority is assigned the priority 50. Thus,
with each priority explicitly declared, the above example looks like:

syntax Int ::= foo(Int) [function]
rule foo(0) => 0 [priority(50)]
rule foo(_) => 1 [owise]

One final note: the llvm backend reserves priorities between 50 and 150
inclusive for certain specific purposes. Because of this, explicit
priorities which are given within this region may not behave precisely as
described above. This is primarily in order that it be possible where necessary
to provide guidance to the pattern matching algorithm when it would otherwise
make bad choices about which rules to try first. You generally should not
give any rule a priority within this region unless you know exactly what the
implications are with respect to how the llvm backend orders matches.

Evaluation Strategy

strict and seqstrict attributes

The strictness attributes allow defining evaluation strategies without having
to explicitly make rules which implement them. This is done by injecting
heating and cooling rules for the subterms. For this to work, you need to
define what a result is for K, by extending the KResult sort.

For example:

syntax AExp ::= Int
              | AExp "+" AExp [strict, klabel(addExp)]

This generates two heating rules (where the hole syntaxes "[]" "+" AExp and
AExp "+" "[]" is automatically added to create an evaluation context):

rule [addExp1-heat]: <k> HOLE:AExp +  AE2:AExp => HOLE ~>  [] + AE2 ... </k> [heat]
rule [addExp2-heat]: <k>  AE1:AExp + HOLE:AExp => HOLE ~> AE1 +  [] ... </k> [heat]

And two corresponding cooling rules:

rule [addExp1-cool]: <k> HOLE:AExp ~>  [] + AE2 => HOLE +  AE2 ... </k> [cool]
rule [addExp2-cool]: <k> HOLE:AExp ~> AE1 +  [] =>  AE1 + HOLE ... </k> [cool]

Note that the rules are given labels based on the klabel of the production, which
nonterminal is the hole, and whether it's the heating or the cooling rule.

You will note that these rules can apply one after another infinitely. In
practice, the KResult sort is used to break this cycle by ensuring that only
terms that are not part of the KResult sort will be heated. The heat and
cool attributes are used to tell the compiler that these are heating and
cooling rules and should be handled in the manner just described. Nothing stops
the user from writing such heating and cooling rules directly if they wish,
although we describe other more convenient syntax for most of the advanced
cases below.

One other thing to note is that in the above sentences, HOLE is just a
variable, but it has special meaning in the context of sentences with the
heat or cool attribute. In heating or cooling rules, the variable named
HOLE is considered to be the term being heated or cooled and the compiler
will generate isKResult(HOLE) and notBool isKResult(HOLE) side conditions
appropriately to ensure that the backend does not loop infinitely. The module
BOOL will also be automatically and privately included for semantic
purposes. The syntax for parsing programs will not be affected.

In order for this functionality to work, you need to define the KResult sort.
For instance, we tell K that a term is fully evaluated once it becomes an Int
here:

syntax KResult ::= Int

Note that you can also say that a given expression is only strict only in
specific argument positions. Here we use this to define "short-circuiting"
boolean operators.

syntax KResult ::= Bool

syntax BExp ::= Bool
              | BExp "||" BExp [strict(1)]
              | BExp "&&" BExp [strict(1)]

rule <k> true  || _    => true ... </k>
rule <k> false || REST => REST ... </k>

rule <k> true  && REST => REST  ... </k>
rule <k> false && _    => false ... </k>

If you want to force a specific evaluation order of the arguments, you can use
the variant seqstrict to do so. For example, this would make the boolean
operators short-circuit in their second argument first:

syntax KResult ::= Bool

syntax BExp ::= Bool
              | BExp "||" BExp [seqstrict(2,1)]
              | BExp "&&" BExp [seqstrict(2,1)]

rule <k> _    || true  => true ... </k>
rule <k> REST || false => REST ... </k>

rule <k> REST && true  => REST  ... </k>
rule <k> _    && false => false ... </k>

This will generate rules like this in the case of _||_ (note that BE1 will
not be heated unless isKResult(BE2) is true, meaning that BE2 must be
evaluated first):

rule <k>  BE1:BExp || HOLE:BExp => HOLE ~> BE1 ||  [] ... </k> [heat]
rule <k> HOLE:BExp ||  BE2:BExp => HOLE ~>  [] || BE2 ... </k> requires isKResult(BE2) [heat]

rule <k> HOLE:BExp ~>  [] || BE2 => HOLE ||  BE2 ... </k> [cool]
rule <k> HOLE:BExp ~> BE1 ||  [] =>  BE1 || HOLE ... </k> [cool]

Context Declaration

Sometimes more advanced evaluation strategies are needed. By default, the
strict and seqstrict attributes are limited in that they cannot describe
the context in which heating or cooling should occur. When this type of
control over the evaluation strategy is required, context sentences can be
used to simplify the process of declaring heating and cooling when it would be
unnecessarily verbose to write heating and cooling rules directly.

For example, if the user wants to heat a term if it exists under a foo
constructor if the term to be heated is of sort bar, one might write the
following context (with the optional label):

context [foo]: foo(HOLE:Bar)

Once again, note that HOLE is just a variable, but one that has special
meaning to the compiler indicating the position in the context that should
be heated or cooled.

This will automatically generate the following sentences:

rule [foo-heat]: <k> foo(HOLE:Bar) => HOLE ~> foo([]) ... </k> [heat]
rule [foo-cool]: <k> HOLE:Bar ~> foo([]) => foo(HOLE) ... </k> [cool]

The user may also write the K cell explicitly in the context declaration
if they want to match on another cell as well, for example:

context <k> foo(HOLE:Bar) ... </k> <state> .Map </state>

This context will now only heat or cool if the state cell is empty.

Side conditions in context declarations

The user is allowed to write a side condition in a context declaration, like
so:

context foo(HOLE:Bar) requires baz(HOLE)

This side condition will be appended verbatim to the heating rule that is
generated, however, it will not affect the cooling rule that is generated:

rule <k> foo(HOLE:Bar) => HOLE ~> foo([]) ... </k> requires baz(HOLE) [heat]
rule <k> HOLE:Bar ~> foo([]) => foo(HOLE) ... </k> [cool]

Rewrites in context declarations

The user can also include exactly one rewrite operation in a context
declaration if that rule rewrites the variable HOLE on the left hand side
to a term containing HOLE on the right hand side. For exampl;e:

context foo(HOLE:Bar => bar(HOLE))

In this case, the code generated will be as follows:

rule <k> foo(HOLE:Bar) => bar(HOLE) ~> foo([]) ... </k> [heat]
rule <k> bar(HOLE:Bar) ~> foo([]) => foo(HOLE) ... </k> [cool]

This can be useful if the user wishes to evaluate a term using a different
set of rules than normal.

result attribute

Sometimes it is necessary to be able to evaluate a term to a different sort
than KResult. This is done by means of adding the result attribute to
a strict production, a context, or an explicit heating or cooling rule:

syntax BExp ::= Bool
              | BExp "||" BExp [seqstrict(2,1), result(Bool)]

In this case, the sort check used by seqstrict and by the heat and cool
attributes will be isBool instead of isKResult. This particular example
does not really require use of the result attribute, but if the user wishes
to evaluate a term of sort KResult further, the result attribute would be
required.

hybrid attribute

In certain situations, it is desirable to treat a particular production which
has the strict attribute as a result if the term has had its arguments fully
evaluated. This can be accomplished by means of the hybrid attribute:

syntax KResult ::= Bool

syntax BExp ::= Bool
              | BExp "||" BExp [strict(1), hybrid]

This attribute is equivalent in this case to the following additional axiom
being added to the definition of isKResult:

rule isKResult(BE1:BExp || BE2:BExp) => true requires isKResult(BE1)

Sometimes you wish to declare a production hybrid with respect to a predicate
other than isKResult. You can do this by specifying a sort as the body of the
hybrid attribute, e.g.:

syntax BExp ::= BExp "||" BExp [strict(1), hybrid(Foo)]

generates the rule:

rule isFoo(BE1:BExp || BE2:BExp) => true requires isFoo(BE1)

Properly speaking, hybrid takes an optional comma-separated list of sort
names. If the list is empty, the attribute is equivalent to hybrid(KResult).
Otherwise, it generates hybrid predicates for exactly the sorts named.

Context aliases

Sometimes it is necessary to define a fairly complicated evaluation strategy
for a lot of different operators. In this case, the user could simply write
a number of complex context declarations, however, this quickly becomes
tedious. For this purpose, K has a concept called a context alias. A context
alias is a bit like a template for describing contexts. The template can then
be instantiated against particular productions using the strict and
seqstrict attributes.

Here is a (simplified) example taken from the K semantics of C++:

context alias [c]: <k> HERE:K ... </k> <evaluate> false </evaluate>
context alias [c]: <k> HERE:K ... </k> <evaluate> true </evaluate> [result(ExecResult)]

syntax Expr ::= Expr "=" Init [strict(c; 1)]

This defines the evaluation strategy during the translation phase of a C++
program for the assignment operator. It is equivalent to writing the following
context declarations:

context <k> HOLE:Expr = I:Init ... </k> <evaluate> false </evaluate>
context <k> HOLE:Expr = I:Init ... </k> <evaluate> true </evaluate> [result(ExecResult)]

What this is saying is, if the evaluate cell is false, evaluate the term
like normal to a KResult. But if the evaluate cell is true, instead
evaluate it to the ExecResult sort.

Essentially, we have given a name to this evaluation strategy in the form of
the rule label on the context alias sentences (in this case, c). We can
then say that we want to use this evaluation strategy to evaluate particular
arguments of particular productions by referring to it by name in a strict
attribute. For example, strict(c) will instantiate these contexts once for
each argument of the production, whereas strict(c; 1) will instantiate it
only for the first argument. The special variable HERE is used to tell the
compiler where you want to place the production that is to be heated or cooled.

You can also specify multiple context aliases for different parts of a production,
for example:

syntax Exp ::= foo(Exp, Exp) [strict(left; 1; right; 2)]

This says that we can evaluate the left and right arguments in either order, but to evaluate
the left using the left context alias and the right using the right context alias.

We can also say seqstrict(left; 1; right; 2), in which case we additionally must evaluate
the left argument before the right argument. Note, all strict positions are considered collectively
when determining the evaluation order of seqstrict or the hybrid predicates.

A strict attribute with no rule label associated with it is equivalent to
a strict attribute given with the following context alias:

context alias [default]: <k> HERE:K ... </k>

One syntactic convenience that is provided is that if you wish to declare the following context:

context foo(HOLE => bar(HOLE))

you can simply write the following:

syntax Foo ::= foo(Bar) [strict(alias)]

context alias [alias]: HERE [context(bar)]

Pattern Matching

As Patterns

New syntax has been added to K for matching a pattern and binding the resulting
match in its entirety to a variable.

The syntax is:

Pattern #as V::Var

In this case, Pattern, including any variables, is matched and the resulting
variables are added to the substitution if matching succeeds. Furthermore, the
term matched by Pattern is added to the substitution as V.

This code can also be used outside of any rewrite, in which case matching
occurs as if it appeared on the left hand side, and the right hand side becomes
a variable corresponding to the alias.

It is an error to use an as pattern on the right hand side of a rule.

Record-like KApply Patterns

We have added a syntax for matching on KApply terms which mimics the record
syntax in functional languages. This allows us to more easily express patterns
involving a KApply term in which we don't care about some or most of the
children, without introducing a dependency into the code on the number of
arguments which could be changed by a future refactoring.

The syntax is:

record(... field1: Pattern1, field2: Pattern2)

Note that this only applies to productions that are prefix productions.
A prefix production is considered by the implementation to be any production
whose production items match the following regular expression:

(Terminal(_)*) Terminal("(")
(NonTerminal (Terminal(",") NonTerminal)* )?
Terminal(")")

In other words, any sequence of terminals followed by an open parenthesis, an
optional comma separated list of non-terminals, and a close parenthesis.

If a prefix production has no named nonterminals, a record(...) syntax is
allowed, but in order to reference specific fields, it is necessary to give one
or more of the non-terminals in the production names.

Note: because the implementation currently creates one production per possible
set of fields to match on, and because all possible permutations of all
possible subsets of a list of n elements is a number that scales factorially
and reaches over 100 thousand productions at n=8, we currently do not allow
fields to be matched in any order like a true record, but only in the same
order as appears in the production itself.

Given that this only reduces the number of productions to the size of the power
set, this will still explode the parsing time if we create large productions of
10 or more fields that all have names. This is something that should probably
be improved, however, productions with that large of an arity are rare, and
thus it has not been viewed as a priority.

Or Patterns

Sometimes you wish to express that a rule should match if one out of multiple
patterns should match the same subterm. We can now express this in K by means
of using the #Or ML connective on the left hand side of a rule.

For example:

rule foo #Or bar #Or baz => qux

Here any of foo, bar, or baz will match this rule. Note that the behavior is
ill-defined if it is not the case that all the clauses of the or have the same
bound variables.

Matching global context in function rules

On occasion it is highly desirable to be able to look up information from the
global configuration and match against it when evaluating a function. For this
purpose, we introduce a new syntax for function rules.

This syntax allows the user to match on function context from within a
function rule:

syntax Int ::= foo(Int) [function]

rule [[ foo(0) => I ]]
     <bar> I </bar>

rule something => foo(0)

This is completely desugared by the K frontend and does not require any special
support in the backend. It is an error to have a rewrite inside function
context, as we do not currently support propagating such changes back into the
global configuration. It is also an error if the context is not at the top
level of a rule body.

Desugared code:

syntax Int ::= foo(Int, GeneratedTopCell) [function]

rule foo(0, <generatedTop>
              <bar> I </bar>
              ...
            </generatedTop> #as Configuration) => I
rule <generatedTop>
       <k> something ... </k>
       ...
     </generatedTop> #as Configuration
  => <generatedTop>
       <k> foo(0, Configuration> ... </k>
       ...
     </generatedTop>

Collection patterns

It is allowed to write patterns on the left hand side of rules which refer to
complex terms of sort Map, List, and Set, despite these patterns ostensibly
breaking the rule that terms which are functions should not appear on the left
hand side of rules. Such terms are destructured into pattern matching
operations.

The following forms are allowed:

// 0 or more elements followed by 0 or 1 variables of sort List followed by
// 0 or more elements
ListItem(E1) ListItem(E2) L:List ListItem(E3) ListItem(E4)

// the empty list
.List

// 0 or more elements in any order plus 0 or 1 variables of sort Set
// in any order
SetItem(K1) SetItem(K2) S::Set SetItem(K3) SetItem(K4)

// the empty set
.Set

// 0 or more elements in any order plus by 0 or 1 variables of sort Map
// in any order
K1 |-> E1 K2 |-> E2 M::Map K3 |-> E3 K4 |-> E4

// the empty map
.Map

Here K1, K2, K3, K4 etc can be any pattern except a pattern containing both
function symbols and unbound variables. An unbound variable is a variable whose
binding cannot be determined by means of decomposing non-set-or-map patterns or
map elements whose keys contain no unbound variables.

This is determined recursively, ie, the term K1 |-> E2 E2 |-> E3 E3 |-> E4 is
considered to contain no unbound variables.

Note that in the pattern K1 |-> E2 K3 |-> E4 E4 |-> E5, K1 and K3 are
unbound, but E4 is bound because it is bound by deconstructing the key E3, even
though E3 is itself unbound.

In the above examples, E1, E2, E3, and E4 can be any pattern that is normally
allowed on the lhs of a rule.

When a map or set key contains function symbols, we know that the variables in
that key are bound (because of the above restriction), so it is possible to
evaluate the function to a concrete term prior to performing the lookup.

Indeed, this is the precise semantics which occurs; the function is evaluated
and the result is looked up in the collection.

For example:

syntax Int ::= f(Int) [function]
rule f(I:Int) => I +Int 1
rule <k> I:Int => . ... </k> <state> ... SetItem(f(I)) ... </state>

This will rewrite I to . if and only if the state cell contains
I +Int 1.

Note that in the case of Set and Map, one guarantee is that K1, K2, K3, and K4
represent /distinct/ elements. Pattern matching fails if the correct number of
distinct elements cannot be found.

Matching on cell fragments

K allows matching fragments of the configuration and using them to construct
terms and use as function parameters.

configuration <t>
                <k> #init ~> #collectOdd ~> $PGM </k>
                <fs>
                  <f multiplicity="*" type="Set"> 1 </f>
                </fs>
              </t>

The #collectOdd construct grabs the entire content of the <fs> cell.
We may also match on only a portion of its content. Note that the fragment
must be wrapped in a <f> cell at the call site.

syntax KItem ::= "#collectOdd"
rule <k> #collectOdd => collectOdd(<fs> Fs </fs>) ... </k>
     <fs> Fs </fs>

The collectOdd function collects the items it needs

syntax Set ::= collectOdd(FsCell) [function]
rule collectOdd(<fs> <f> I </f> REST </fs>) => SetItem(I) collectOdd(<fs> REST </fs>) requires I %Int 2 ==Int 1
rule collectOdd(<fs> <f> I </f> REST </fs>) =>            collectOdd(<fs> REST </fs>) requires I %Int 2 ==Int 0
rule collectOdd(<fs> .Bag </fs>) => .Set

all-path and one-path attributes to distinguish reachability claims

As the Haskell backend can handle both one-path and all-path reachability
claims, but both these are encoded as rewrite rules in K, these attributes can
be used to clarify what kind of claim a rule is.

In addition of being able to annotate a rule with one of them
(if annotating with more at the same time, only one of them would be chosen),
one can also annotate whole modules, to give a default claim type for all rules
in that module.

Additionally, the Haskell backend introduces an extra command line option
for the K frontend, --default-claim-type, with possible values
all-path and one-path to allow choosing a default type for all
claims.

Set Variables

Motivation

Set variables were introduced as part of Matching Mu Logic, the mathematical
foundations for K. In Matching Mu Logic, terms evaluate to sets of values.
This is useful for both capturing partiality (as in 3/0) and capturing
non-determinism (as in 3 #Or 5). Consequently, symbol interpretation is
extended to have a collective interpretation over sets of input values.

Usually, K rules are given using regular variables, which expect that the term
they match is both defined and has a unique interpretation.

However, it is sometimes useful to have simplification rules which work over
any kind of pattern, be it undefined or non-deterministic. This behavior can be
achieved by using set variables to stand for any kind of pattern.

Syntax

Any variable prefixed by @ will be considered a set variable.

Example

Below is a simplification rule which motivated this extension:

  rule #Ceil(@I1:Int /Int @I2:Int) =>
    {(@I2 =/=Int 0) #Equals true} #And #Ceil(@I1) #And #Ceil(@I2)
    [anywhere]

This rule basically says that @I1:Int /Int @I2:Int is defined if @I1 and
@I2 are defined and @I2 is not 0. Using sets variables here is important as
it allows the simplification rule to apply any symbolic patterns, without
caring whether they are defined or not.

This allows simplifying the expression #Ceil((A:Int /Int B:Int) / C:Int) to:

{(C =/=Int 0) #Equals true} #And #Ceil(C) #And ({(B =/=Int 0) #Equals true}
#And #Ceil(B) #And #Ceil(A)`

See kframework/kore#729 for
more details.

SMT Translation

K makes queries to an SMT solver (Z3) to discharge proof obligations when doing
symbolic execution. You can control how these queries are made using the
attributes smtlib, smt-hook, and smt-lemma on declared productions.
These attributes guide the prover when it tries to apply rules to discharge a
proof obligation.

  • smt-hook(...) allows you to specify a term in SMTLIB2 format which should
    be used to encode that production, and assumes that all symbols appearing in
    the term are already declared by the SMT solver.
  • smtlib(...) allows you to declare a new SMT symbol to be used when that
    production is sent to Z3, and gives it uninterpreted function semantics.
  • smt-lemma can be applied to a rule to encode it as a conditional equality
    when sending queries to Z3. A rule rule LHS => RHS requires REQ will be
    encoded as the conditional equality (=> REQ (= (LHS RHS)). Every symbol
    present in the rule must have an smt-hook(...) or smtlib(...) attribute.
syntax Int ::= "~Int" Int          [function, klabel(~Int_), symbol,
                                    smtlib(notInt)]
             | Int "^%Int" Int Int [function, klabel(_^%Int__), symbol,
                                    smt-hook((mod (^ #1 #2) #3))]

In the example above, we declare two productions ~Int_ and _^%Int__, and
tell the SMT solver to:

  • use uninterpreted function semantics for ~Int_ via SMTLIB2 symbol
    notInt, and
  • use the SMTLIB2 term (mod (^ #1 #2) #3) (where #N marks the Nth
    production non-terminal argument positions) for _^%Int__, where mod and
    ^ already are declared by the SMT solver.

Caution

Set variables are currently only supported by the Haskell backend.
The use of rules with set variables should be sound for all other backends
which just execute by rewriting, however it might not be safe for backends
which want to guarantee coverage.

Variables occurring only in the RHS of a rule

This section presents possible scenarios requiring variables to only appear in
the RHS of a rule.

Summary

Except for ? variables and ! (fresh) variables, which are
required to only appear in the RHS of a rule, all other variables must
also appear in the LHS of a rule. This restriction also applies to anonymous
variables; in particular, for claims, ?_ (not _) should be used in the RHS
to indicate that something changes but we don't care to what value.

To support specifying random-like behavior, the above restriction can be relaxed
by annotating a rule with the unboundVariables attribute whenever the rule
intentionally contains regular variables only occurring in the RHS.

Introduction

K uses question mark variables of the form ?X to refer to
existential variables, and uses ensures to specify logical constraints on
those variables.
These variables are only allowed to appear in the RHS of a K rule.

If the rules represent rewrite (semantic) steps or verification claims,
then the ? variables are existentially quantified at the top of the RHS;
otherwise, if they represent equations, the ? variables are quantified at the
top of the entire rule.

Note that when both ?-variables and regular variables are present,
regular variables are (implicitly) universally quantified on top of the rule
(already containing the existential quantifications).
This essentially makes all ? variables depend on all regular variables.

All examples below are intended more for program verification /
symbolic execution, and thus concrete implementations might choose to ignore
them altogether or to provide ad-hoc implementations for them.

Example: Verification claims

Consider the following definition of a (transition) system:

module A
  rule foo => true
  rule bar => true
  rule bar => false
endmodule

Consider also, the following specification of claims about the definition above:

module A-SPEC
  rule [s1]: foo => ?X:Bool
  rule [s2]: foo =>  X:Bool  [unboundVariables(X)]
  rule [s3]: bar => ?X:Bool
  rule [s4]: bar =>  X:Bool  [unboundVariables(X)]
endmodule
One-path interpretation
  • (s1) says that there exists a path from foo to some boolean, which is
    satisfied easily using the foo => true rule
  • (s3) says the same thing about bar and can be satisfied by either of
    bar => true and bar => false rules
  • (s2) and (s4) can be better understood by replacing them with instances for
    each element of type Bool, which can be interpreted that
    both true and false are reachable from foo for (s2), or bar for (s4),
    respectively.
    • (s2) cannot be verified as we cannot find a path from foo to false.
    • (s4) can be verified by using bar => true to show true is reachable and
      bar => false to achieve the same thing for false
All-path interpretation
  • (s1) says that all paths from foo will reach some boolean, which is
    satisfied by the foo => true rule and the lack of other rules for foo

  • (s3) says the same thing about bar and can be satisfied by checking that
    both bar => true and bar => false end in a boolean, and there are no
    other rules for bar

  • (s2) and (s4) can be better understood by replacing them with instances for
    each element of type Bool, which can be interpreted that
    both true and false are reachable in all paths originating in
    foo for (s2), or bar for (s4), respectively.
    This is a very strong claim, requiring that all paths originating in
    foo (bar) pass through both true and false,
    so neither (s2) nor (s4) can be verified.

    Interestingly enough, adding a rule like false => true would make both
    (s2) and (s4) hold.

Example: Random Number Construct rand()

The random number construct rand() is a language construct which could be
easily conceived to be part of the syntax of a programming language:

Exp ::= "rand" "(" ")"

The intended semantics of rand() is that it can rewrite to any integer in
a single step. This could be expressed as the following following infinitely
many rules.

rule  rand() => 0
rule  rand() => 1
rule  rand() => 2
  ...    ...
rule rand() => (-1)
rule rand() => (-2)
  ...    ...

Since we need an instance of the rule for every integer, one could summarize
the above infinitely many rules with the rule

rule rand() => I:Int [unboundVariables(I)]

Note that I occurs only in the RHS in the rule above, and thus the rule
needs the unboundVariables(I) attribute to signal that this is intentionally.

One can define variants of rand() by further constraining the output variable
as a precondition to the rule.

Rand-like examples
  1. randBounded(M,N) can rewrite to any integer between M and N

    syntax Exp ::= randBounded(Int, Int)
    rule randBounded(M, N) => I
      requires M <=Int I andBool I <=Int N
      [unboundVariables(I)]
    
  2. randInList(Is) takes a list Is of items
    and can rewrite in one step to any item in Is.

    syntax Exp ::= randInList (List)
    rule randInList(Is) => I
      requires I inList Is
      [unboundVariables(I)]
    
  3. randNotInList(Is) takes a list Is of items
    and can rewrite in one step to any item not in Is.

    syntax Exp ::= randNotInList (List)
    rule randNotInList(Is) => I
      requires notBool(I inList Is)
      [unboundVariables(I)]
    
  4. randPrime(), can rewrite to any prime number.

    syntax Exp ::= randPrime ()
    rule randPrime() => X:Int
      requires isPrime(X)
      [unboundVariables(X)]
    

    where isPrime(_) is a predicate that can be defined in the usual way.

Note 1: all above are not function symbols, but language constructs.

Note 2: Currently the frontend does not allow rules with universally quantified
variables in the RHS which are not bound in the LHS.

Note 3. Allowing these rules in a concrete execution engine would require an
algorithm for generating concrete instances for such variables, satisfying the
given constraints; thus the unboundVariables attribute serves two purposes:

  • to allow such rules to pass the variable checks, and
  • to signal (concrete execution) backends that specialized algorithm would be
    needed to instantiate these variables.

Example: Fresh Integer Construct fresh(Is)

The fresh integer construct fresh(Is) is a language construct.

Exp ::= ... | "fresh" "(" List{Int} ")"

The intended semantics of fresh(Is) is that it can always rewrite to an
integer that in not in Is.

Note that fresh(Is) and randNotInList(Is) are different; the former
does not need to be able to rewrite to every integers not in Is,
while the latter requires so.

For example, it is correct to implement fresh(Is) so it always returns the
smallest positive integer that is not in Is, but same implementation for
randNotInList(Is) might be considered inadequate.
In other words, there exist multiple correct implementations of fresh(Is),
some of which may be deterministic, but there only exists a unique
implementation of randNotInList(Is).
Finally, note that randNotInList(Is) is a correct implementation
for fresh(Is); Hence, concrete execution engines can choose to handle
such rules accordingly.

We use the following K syntax to define fresh(Is)

syntax Exp ::= fresh (List{Int})
rule fresh(Is:List{Int}) => ?I:Int
  ensures notBool (?I inList{Int} Is)

A variant of this would be a choiceInList(Is) language construct which would
choose some number from a list:

syntax Exp ::= choiceInList (List{Int})
rule choiceInList(Is:List{Int}) => ?I:Int
  ensures ?I inList{Int} Is

Note: This definition is different from one using a ! variable to indicate
freshness because using ! is just syntactic sugar for generating globally
unique instances and relies on a special configuration cell, and cannot be
constrained, while the fresh described here is local and can be constrained.
While the first is more appropriate for concrete execution, this might be
better for symbolic execution / program verification.

Example: Arbitrary Number (Unspecific Function) arb()

The function arb() is not a PL construct, but a mathematical function.
Therefore, its definition should not be interpreted as an execution step, but
rather as an equality.

The intended semantics of arb() is that it is an unspecified nullary function.
The exact return value of arb() is unspecified in the semantics but up to the
implementations.
However, being a mathematical function, arb() must return the same value in
any one implementation.

We do not need special frontend syntax to define arb().
We only need to define it in the usual way as a function
(instead of a language construct), and provide no axioms for it.
The total attribute ensures that the function is total, i.e.,
that it evaluates to precisely one value for each input.

Variants

There are many variants of arb(). For example, arbInList(Is) is
an unspecified function whose return value must be an element from Is.

Note that arbInList(Is) is different from choiceInList(Is), because
choiceInList(Is) transitions to an integer in Is (could be a different one
each time it is used), while arbInList(Is) is equal to a (fixed)
integer not in Is.

W.r.t. the arb variants, we can use ? variables and the function
annotation to signal that we're defining a function and the value of the
function is fixed, but non-determinate.

syntax Int ::= arbInList(List{Int}) [function]
rule arbInList(Is:List{Int}) => ?I:Int
  ensures ?I inList{Int} Is

If elimination of existentials in equational rules is needed, one possible
approach would be through Skolemization,
i.e., replacing the ? variable with a new uninterpreted function depending
on the regular variables present in the function.

Example: Interval (Non-function Symbols) interval()

The symbol interval(M,N) is not a PL construct, nor a function in the
first-order sense, but a proper matching-logic symbol, whose interpretation is
in the powerset of its domain.
Its axioms will not use rewrites but equalities.

The intended semantics of interval(M,N) is that it equals the set of
integers that are larger than or equal to M and smaller than or equal to N.

Since expressing the axiom for interval requires an an existential
quantification on the right-hand-side, thus making it a non-total symbol
defined through an equation, using ? variables might be confusing since their
usage would be different from that presented in the previous sections.

Hence, the proposal to support this would be to write this as a proper ML rule.
A possible syntax for this purpose would be:

eq  interval(M,N)
    ==
    #Exists X:Int .
        (X:Int #And { X >=Int M #Equals true } #And { X <=Int N #Equals true })

Additionally, the symbol declaration would require a special attribute to
signal the fact that it is not a constructor but a defined symbol.

Since this feature is not clearly needed by K users at the moment, it is only
presented here as an example; its implementation will be postponed for such time
when its usefulness becomes apparent.

Parser Generation

In addition to on-the-fly parser generation using kast, K is capable of
ahead-of-time parser generation of LR(1) or GLR parsers using Flex and Bison.
This can be done one of two different ways.

  1. You can explicitly request for a particular parser to be generated by
    invoking kast --gen-parser <outputFile> or
    kast --gen-glr-parser <outputFile> respectively. kast will then create a
    parser based on the same command line flags that govern on-the-fly parsing,
    like -s to specify the starting sort, and -m to specify the module to
    parse under. By default, this generates a parser for the sort of the $PGM
    configuration variable in the main syntax module of the definition.
  2. You can request that a specific set of parsers be generated for all the
    configuration variables of your definition by passing the
    --gen-bison-parser or --gen-glr-bison-parser flags to kompile.
    kompile will decide the sorts to use as start symbols based on the sorts
    in the configuration declaration for the configuration variables. The $PGM
    configuration variable will be generated based on the main syntax module
    of the definition. The user must explicitly annotate the configuration
    declaration with the other modules to use to parse the other configuration
    variables as attributes. For example, if I have the following cell in the
    configuration declaration: <cell> foo($FOO:Foo, $BAR:Bar) </cell>,
    One might annotate it with the attribute pair parser="FOO, TEST; BAR, TEST2"
    to indicate that configuration variable $FOO should be parsed in the
    TEST module, and configuration variable $BAR should be parsed in the
    TEST2 module. If the user forgets to annotate the declaration with the
    parser attribute, only the $PGM parser will be generated.

Bison-generated parsers are extremely fast compared to kast, but they have
some important limitations:

  • Bison parsers will always output Kore. You can then pass the resulting AST
    directly to llvm-krun or kore-exec and bypass the krun frontend, making
    them very fast, but lower-level.
  • Bison parsers do not yet support macros. This may change in a future release.
    Note that you can use anywhere rules instead of macros in most cases to get
    around this limitation, although they will not benefit from unparsing via the
    alias attribute.
  • Obligation falls on the user to ensure that the grammar they write is LR(1)
    if they choose to use LR(1) parsing. If this does not happen, the parser
    generated will have shift/reduce or reduce/reduce conflicts and the parser
    may behave differently than kast would (kast is a GLL parser, ie, it
    is based on LL parsers and parses all unambiguous context-free grammars).
    K provides an attribute, not-lr1, which can be applied to modules known to
    not be LR(1), and will trigger a warning if the user attempts to generate an
    LR(1) parser which recursively imports that module.
  • If you are using LR(1) based parsing, the prefer and avoid attributes are
    ignored. It is only possible to implement these attributes by means of
    generalized LL or LR parsing and a postprocessing on the AST to remove the
    undesirable ambiguity.
  • Obligation falls on the user to ensure that the grammar they write has as
    few conflicts as possible if they are using GLR parsing. Bison's GLR support
    is quite primitive, and in the worst case it can use exponential space and
    time to parse a program, which generally leads the generated parser to report
    "memory exhausted", indicating that the parse could not be completed within
    the stack space allocated by Bison. It's best to ensure that the grammar is
    as close to LR(1) as possible and only utilizes conflicts where absolutely
    necessary. One tool that can be used to facilitate this is to pass
    --bison-lists to kompile. This will disable support for the List{Sort}
    syntax production, and it will make NeList{Sort} left associative, but the
    resulting productions generated for NeList{Sort} will be LR(1) and use bounded
    stack space.
  • If the grammar you are parsing is context-sensitive (for example, because
    it requires a symbol table to parse), one thing you can do to make this
    language parse in K is to implement the language as an ambiguous grammar.
    Bison's GLR parser will generate an amb production that is parametric in
    the sort of the ambiguity. You can then import the K-AMBIGUITIES module
    and use rewriting to resolve the ambiguities using whatever preprocessing
    mechanisms you prefer.

Location Information

K is able to insert file, line, and column metadata into the parse tree on a
per-sort basis when parsing using a bison-generated parser. To enable this,
mark the sort with the locations attribute.

  syntax Exp [locations]
  syntax Exp ::= Exp "/" Exp | Int

K implicitly wraps productions of these sorts in a #location term (see the
K-LOCATIONS module in kast.md). The metadata can thus be accessed with
ordinary rewrite rules:

  rule #location(_ / 0, File, StartLine, _StartColumn, _EndLine, _EndColumn) =>
  "Error: Division by zero at " +String File +String ":" Int2String(StartLine)

Sometimes it is desirable to allow code to be written in a file which
overwrites the current location information provided by the parser. This can be
done via a combination of the #LineMarker sort and the --bison-file flag to
the parser generator. If you declare a production of sort #LineMarker which
contains a regular expression terminal, this will be treated as a
line marker by the bison parser. The user will then be expected to provide
an implementation of the parser for the line marker in C. The function expected
by the parser has the signature void line_marker(char *, yyscan_t), where
yyscan_t is a
reentrant flex scanner.
The string value of the line marker token as specified by your regular
expression can be found in the first parameter of the function, and you can
set the line number used by the scanner using yyset_lineno(int, yyscan_t). If
you declare the variable extern char *filename, you can also set the current
file name by writing a malloc'd, zero-terminated string to that variable.

Unparsing

A number of factors go into how terms are unparsed in K. Here we describe some
of the features the user can use to control how unparsing happens.

Brackets

One of the phases that the unparser goes through is to insert productions
tagged with the bracket attribute where it believes this is necessary
in order to create a correct string that will be parsed back into the original
AST. The most common case of this is in expression grammars. For example,
consider the following grammar:

syntax Exp ::= Int
             | Exp "*" Exp
             > Exp "+" Exp

Here we have declared that expressions can contain integer addition and
multiplication, and that multiplication binds tighter than addition. As a
result, when writing a program, if we want to write an expression that first
applies addition, then multiplication, we must use brackets: (1 + 2) * 3.
Similarly, if we have such an AST, we must insert brackets into the AST
in order to faithfully unparse the term in a manner that will be parsed back
into the same ast, because if we do not, we end up unparsing the term as
1 + 2 * 3, which will be parsed back as 1 + (2 * 3) because of the priority
declaration in the grammar.

You can control how the unparser will insert such brackets by adding a
production with the bracket attribute and the correct sort. For example, if,
instead of parentheses, you want to use curly braces, you could write:

syntax Exp ::= "{" Exp "}" [bracket]

This would signal to the unparser how brackets should look for terms of sort
Exp, and it will use this syntax when unparsing terms of sort Exp.

Commutative collections

One thing that K will do (unless you pass the --no-sort-collections flag to
krun) is to sort associative, commutative collections (such as Set and Map)
alphanumerically. For example, if I have a collection whose keys are sort Id
and they have the values a, b, c, and d, then unparsing will always print
first the key a, then b, then c, then d, because this is the alphabetic order
of these keys when unparsed.

Furthermore, K will sort numeric keys numerically. For example, if I have a
collection whose keys are 1, 2, 5, 10, 30, it will first display 1, then 2,
then 5, then 10, then 30, because it will sort these keys numerically. Note
that this is different than an alphabetic sort, which would sort them as
1, 10, 2, 30, 5. We believe the former is more intuitive to users.

Substitution filtering

K will remove substitution terms corresponding to anonymous variables when
using the --pattern flag if those anonymous variables provide no information
about the named variables in your serach pattern. You can disable this behavior
by passing --no-substitution-filtering to krun. When this flag is not passed,
and you are using the Haskell backend, any equality in a substitution (ie, an
#Equals under an #And under an #Or), will be hidden from the user if the
left hand side is a variable that was anonymous in the --pattern passed by
the user, unless that variable appears elsewhere in the substitution. If you
want to see that variable in the substitution, you can either disable this
filtering, or give that variable a name in the original search pattern.

Variable alpha renaming

K will automatically rename variables that appear in the output configuration.
Similar to commutative collections, this is done to normalize the resulting
configuration so that equivalent configurations will be printed identically
regardless of how they happen to be reached. This pass can be disabled by
passing --no-alpha-renaming to krun.

Macro expansion

K will apply macros in reverse on the output configuration if the macro was
created with the alias or alias-rec attribute. See the section on macro
expansion for more details.

Formatting

format attribute

K allows you to control how terms are unparsed using the format attribute.
By default, a domain value is unparsed by printing its string value verbatim,
and an application pattern is unparsed by printing its terminals and children
in the sequence implied by its concrete syntax, separated by spaces. However,
K gives you complete control over how you want to unparse the symbol.

A format attribute is a string containing zero or more escape sequences that
tell K how to unparse the symbol. Escape sequences begin with a '%' and are
followed by either an integer, or a single non-digit character. Below is a
list of escape sequences recognized by the formatter:

Escape Sequence Meaning
n Insert '\n' followed by the current indentation level
i Increase the current indentation level by 1
d Decrease the current indentation level by 1
c Move to the next color in the list of colors for this production
r Reset color to the default foreground color for the terminal (See below for more information on how colors work)
an integer Print a terminal or nonterminal from the production (See below for more information)
any other char Print that character verbatim

Using the integer escape sequence

In the integer escape sequence %a, the integer a is treated as a 1-based
index into the terminals and nonterminals of the production.

  • If the offset refers to a terminal, move to the next color in the list of
    colors for this production, print the value of that terminal, then reset the
    color to the default foreground color for the terminal.

  • If the offset refers to a regular expression terminal, it is an error.

  • If the offset refers to a nonterminal, print the unparsed representation of
    the corresponding child of the current term.

color and colors attributes

K allows you to take advantage of ANSI terminal codes for foreground color
in order to colorize output pretty-printed by the unparser. This is controlled
via the color and colors attributes of productions. These attributes
combine with the format attribute to control how a term is colorized.

The first thing to understand about how colorization works is that the color
and colors attributes are used to construct a list of colors associated
with each production, and the format attribute then uses that list to choose
the color for each part of the production. For more information on how the
format attribute chooses a color from the list, see above, but essentially,
each terminal or %c in the format attribute advances the pointer in the list
by one element, and terminals and %r reset the current color to the default
foreground color of the terminal afterwards.

There are two ways you can construct a list of colors associated with a
production:

  • The color attribute creates the entire list all with the same color, as
    specified by the value of the attribute. When combined with the default format
    attribute, this will color all the terminals in that production that color, but
    more advanced techniques can be used as well.

  • The colors attribute creates the list from a manual, comma-separated list
    of colors. The attribute is invalid if the length of the list is not equal to
    the number of terminals in the production plus the number of %c substrings in
    the format attribute.

Attributes Reference

Attribute Syntax Overview

In K, many different syntactic categories accept an optional trailing list of
keywords known as attributes. Attribute lists have two different syntaxes,
depending on where they occur. Each attribute also has a type which describes
where it may occur.

The first syntax is a square-bracketed ([]) list of words. This syntax is
available for following attribute types:

  1. module attributes - may appear immediately after the module keyword
  2. sort attributes - may appear immediately after a sort declaration
  3. production attributes - may appear immediately after a BNF production
    alternative
  4. rule attributes - may appear immediately after a rule
  5. context attributes - may appear immediately after a context or context
    alias
  6. context alias attributes - may appear immediately after a context alias
  7. claim attributes - may appear immediately after a claim

The second syntax is the XML attribute syntax, i.e., a space delemited list of
key-and-quoted-value pairs appearing inside the start tag of an XML element:
<element key1="value" key2="value2" ... > </element>. This syntax is
available for the following attribute types:

  1. cell attributes - may appear inside of the cell start tag in
    configuration declarations

Unrecognized attributes are reported as an error. When we talk about
the type of an attribute, we mean a syntactic category to which an attribute
can be attached where the attribute has some semantic effect.

Attribute Index

We now provide an index of available attributes organized alphabetically with a
brief description of each. Note that the same attribute may appear in the index
multiple times to indicate its effect in different contexts or with/without
arguments. A legend describing how to interpret the index follows.

Name Type Backend Reference
alias-rec prod all Macros and Aliases
alias prod all Macros and Aliases
all-path claim haskell all-path and one-path attributes to distinguish reachability claims
anywhere rule all anywhere rules
applyPriority(_) prod all Symbol priority and associativity
avoid prod all Symbol priority and associativity
binder prod all No reference yet.
bracket prod all Parametric productions and bracket attributes
color(_) prod all color and colors attributes
colors(_) prod all color and colors attributes
concrete mod llvm symbolic and concrete attribute
concrete(_) rule haskell concrete and symbolic attributes (Haskell backend)
concrete rule haskell concrete and symbolic attributes (Haskell backend)
context(_) alias all Context aliases
deprecated prod all deprecated attribute
exit = "" cell all exit attribute
format prod all format attribute
freshGenerator prod all freshGenerator attribute
function prod all function and total attributes
group(_) all all Symbol priority and associativity
hook(_) prod all No reference yet
hybrid(_) prod all hybrid attribute
hybrid prod all hybrid attribute
klabel(_) prod all klabel(_) and symbol attributes
left prod all Symbol priority and associativity
locations sort all Location Information
macro-rec prod all Macros and Aliases
macro prod all Macros and Aliases
memo rule haskell The memo attribute
multiplicity = "_" cell all Collection Cells: multiplicity and type attributes
non-assoc prod all Symbol priority and associativity
one-path claim haskell all-path and one-path attributes to distinguish reachability claims
overload(_) prod all overload(_) attribute
owise rule all owise and priority attributes
prec(_) token all prec attribute
prefer prod all Symbol priority and associativity
priority(_) rule all owise and priority attributes
private mod all private attribute
private prod all public and private attribute
public mod all No reference yet.
public prod all public and private attribute
result(_) ctxt all result attribute
result(_) rule all result attribute
right prod all Symbol priority and associativity
seqstrict(_) prod all strict and seqstrict attributes
seqstrict prod all strict and seqstrict attributes
simplification rule haskell simplification attribute (Haskell backend)
simplification(_) rule haskell simplification attribute (Haskell backend)
smt-hook(_) prod haskell SMT Translation
smtlib(_) prod haskell SMT Translation
smt-lemma rule haskell SMT Translation
strict prod all strict and seqstrict attributes
strict(_) prod all strict and seqstrict attributes
symbolic mod haskell symbolic and concrete attribute
symbolic rule haskell concrete and symbolic attributes (Haskell backend)
symbolic(_) rule haskell concrete and symbolic attributes (Haskell backend)
symbol prod all klabel(_) and symbol attributes
terminator-symbol(_) prod all klabel(_) and symbol attributes
token prod all token attribute
token sort all token attribute
total prod all function and total attributes
trusted claim haskell trusted attribute
type = "_" cell all Collection Cells: multiplicity and type attributes
unboundVariables(_) rule all The unboundVariables attribute
unused prod all unused attribute
concrete mod all Specify that this module should only be included in concrete backends (LLVM backend).
symbolic mod all Specify that this module should only be included in symbolic backends (Haskell backend).
stream = "_" cell all Specify that this cell should be hooked up to a stream, either stdin, stdout, or stderr.

Internal Attribute Index

Some attributes should not generally appear in user code, except in some
unusual or complex examples. Such attributes are typically generated by the
compiler and used internally. We list these attributes below as a reference for
interested readers:

Name Type Backend Reference
assoc prod all assoc, comm, idem and unit attributes
comm prod all assoc, comm, idem and unit attributes
digest mod all Contains the hash of the textual contents of the module.
idem prod all assoc, comm, idem and unit attributes
unit prod all assoc, comm, idem and unit attributes
userList prod all Identifies the desugared form of Lst ::= List{Elm,"delim"}
predicate prod all Specifies the sort of a predicate label
element prod all Specifies the label of the elements in a list
bracketLabel prod all Keep track of the label of a bracket production since it can't have a klabel
injective prod all Label a given production as injective (unique output for each input)
internal prod all Production is reserved for internal use by the compiler
cool rule all strict and seqstrict attributes
heat rule all strict and seqstrict attributes

Index Legend

  • Name - the attribute's name (optionally followed by an underscore _ to indicate the attribute takes arguments)

  • Type - the syntactic categories where this attribute is not ignored;
    the possible values are the types mentioned above or shorthands:

    1. all - short for any type except cell
    2. mod - short for module
    3. sort
    4. prod - short for production
    5. rule
    6. ctxt - short for context or context alias
    7. claim
    8. cell
  • Backend - the backends that do not ignore this attribute; possible values:

    1. all - all backends
    2. llvm - the LLVM backend
    3. haskell - the Haskell backend
  • Effect - the attribute's effect (when it applies)

Pending Documentation

Backend features not yet given documentation:

  • Parser of KORE terms and definitions
  • Term representation of K terms
  • Hooked sorts and symbols
  • Substituting a substitution into the RHS of a rule
    • domain values
    • functions
    • variables
    • symbols
    • polymorphism
    • hooks
    • injection compaction
    • overload compaction
  • Pattern Matching / Unification of subject and LHS of rule
    • domain values
    • symbols
    • side conditions
    • and/or patterns
    • list patterns
    • nonlinear variables
    • map/set patterns
      • deterministic
      • nondeterministic
    • modulo injections
    • modulo overloads
  • Stepping
    • initialization
    • termination
  • Print kore terms
  • Equality/comparison of terms
  • Owise rules
  • Strategy #STUCK axiom
  • User substitution
    • binders
    • kvar

To get a complete list of hooks supported by K, you can run:

grep -P -R "(?<=[^-])hook\([^)]*\)" k-distribution/include/kframework/builtin/ \
     --include "*.k" -ho | \
sed 's/hook(//' | sed 's/)//' | sort | uniq | grep -v org.kframework

All of these hooks will also eventually need documentation.


  1. Except for in a very limited number of special cases from the
    K standard library. ↩︎

  2. The Maude documentation
    has an example in a context that's somewhat similar to K; discussion of
    ad-hoc overloading is not relevant. ↩︎

K Cheat Sheet

This is a quick reference of the most commonly used K tools.

kompile (--gen-bison-parser)? {file}                : generate parser, optionally with ahead of time
krun {file}                                         : interpret file
krun -cPGM='{string}'                               : interpret string
kast --output (kore | kast) (-e|{file})             : parse expression or file
kompile (--enable-search --backend haskell)? {file} : generate parser, enabling non-deterministic run
krun (--search-all)? {file}                         : interpret file, evaluating non-deterministic runs as well
foo-kompiled/parser_PGM {file}                      : ahead of time parse
kompile (--main-module)? (--syntax-module)? {file}  : generate parser for {file}.k {file}-syntax.k, explicitly state main modules
kparse <file> | kore-print -                        : parse and unparse a file
kompile {file} --enable-llvm-debug                  : generate debuggable output for {file}.k
krun {file} --debugger                              : debug K code
kprove {file}                                       : Verify specs in {file}

During GDB debugging session (see here for
LLDB breakpoint syntax):

break {file}:{linenum}                              : add a breakpoint to {file}'s {linenum} numbered line
k match {module}.{label} subject                    : investigate matching

K Tools

Here we document how to use some of the most commonly used K tools.

Minimizing Output

When one is working with kore-repl or the prover in general and looking at
specific configurations using config, sometimes the configurations can be huge.

One tool to help print configuration compactly is the pyk print utility:

pyk print

We are going to use --minimize option (which is actually used automatically
when printing with pyk). This will filter out many uninteresting cells for the
current config and make the result more compact.

Then, when invoking the prover, you can minimize your output by piping it into
the pyk print ... facility with arguments for controlling the output:

kprove --output json --definition DEFN ... \
    | jq .term                             \
    | pyk print DEFN /dev/stdin --omit-labels ... --keep-labels ...

You can also use this in the kore-repl more easily, by making a help script.
In your current directory, save a new script pykprint.sh:

#!/bin/bash

kast --input kore --output json --definition $1 /dev/stdin \
    | jq .term                                             \
    | pyk print $1 /dev/stdin --omit-labels $2

Now call config | bash pykprint.sh DEFN in Kore REPL to make the output
smaller.

The options you have to control the output are as follows:

  • --no-minimize: do not remove uninteresting cells.
  • --omit-cells: remove the selected cells from the output.
  • --keep-cells: keep only the selected cells in the output.

Note: Make sure that there is no whitespace around , in the omit list,
otherwise you'll get an error (, is a list separator, so this
requirement is strict).

Debugging

The LLVM Backend has support for integration with GDB. You can run the debugger
on a particular program by passing the --debugger flag to krun, or by
invoking the llvm backend interpreter directly. Below we provide a simple
tutorial to explain some of the basic commands supported by the LLVM backend.

LLDB Support

GDB is not well-supported on macOS, particularly on newer OS versions and Apple
Silicon ARM hardware. Consequently, if the --debugger option is passed to krun
on macOS, LLDB[^1] is launched instead of GDB. However, the K-specific debugger
scripts that GDB uses have not been ported to LLDB yet, and so the instructions
in the rest of this section will not work.

The K Definition

Here is a sample K definition we will use to demonstrate debugging
capabilities:

module TEST
  imports INT

  configuration <k> foo(5) </k>
  rule [test]: I:Int => I +Int 1 requires I <Int 10

  syntax Int ::= foo(Int) [function]
  rule foo(I) => 0 -Int I

endmodule

You should compile this definition with --backend llvm --enable-llvm-debug to
use the debugger most effectively.

Stepping

Important: When you first run krun with option --debugger, GDB / LLDB
will instruct you on how to modify ~/.gdbinit or ~/.lldbinit to enable
printing abstract syntax of K terms in the debugger. If you do not perform this
step, you can still use all the other features, but K terms will be printed as
their raw address in memory.

GDB will need the kompiled interpreter in its safe path in order to access the
pretty printing python script within it. A good way to do this would be to pick
a minimum top-level path that covers all of your kompiled semantics (ie. set auto-load safe-path ~/k-semantics). LLDB has slightly different security
policies that do not require fully-arbitrary code execution.

This section uses GDB syntax to demonstrate the debugging features. Please
refer to the GDB to LLDB command map on
macOS.

You can break before every step of execution is taken by setting a breakpoint
on the k_step function.

(gdb) break definition.kore:k_step
Breakpoint 1 at 0x25e340
(gdb) run
Breakpoint 1, 0x000000000025e340 in step (subject=`<generatedTop>{}`(`<k>{}`(`kseq{}`(`inj{Int{}, KItem{}}`(#token("0", "Int")),dotk{}(.KList))),`<generatedCounter>{}`(#token("0", "Int"))))
(gdb) continue
Continuing.

Breakpoint 1, 0x000000000025e340 in step (subject=`<generatedTop>{}`(`<k>{}`(`kseq{}`(`inj{Int{}, KItem{}}`(#token("1", "Int")),dotk{}(.KList))),`<generatedCounter>{}`(#token("0", "Int"))))
(gdb) continue 2
Will ignore next crossing of breakpoint 1.  Continuing.

Breakpoint 1, 0x000000000025e340 in step (subject=`<generatedTop>{}`(`<k>{}`(`kseq{}`(`inj{Int{}, KItem{}}`(#token("3", "Int")),dotk{}(.KList))),`<generatedCounter>{}`(#token("0", "Int"))))
(gdb)

Breaking on a specific rule

You can break when a rule is applied by giving the rule a rule label. If the
module name is TEST and the rule label is test, you can break when the rule
applies by setting a breakpoint on the TEST.test.rhs function:

(gdb) break TEST.test.rhs
Breakpoint 1 at 0x25e250: file /home/dwightguth/test/./test.k, line 4.
(gdb) run
Breakpoint 1, TEST.test.rhs (VarDotVar0=`<generatedCounter>{}`(#token("0", "Int")), VarDotVar1=dotk{}(.KList), VarI=#token("0", "Int")) at /home/dwightguth/test/./test.k:4
4         rule [test]: I:Int => I +Int 1 requires I <Int 10
(gdb)

Note that the substitution associated with that rule is visible in the
description of the frame.

You can also break when a side condition is applied using the TEST.test.sc
function:

(gdb) break TEST.test.sc
Breakpoint 1 at 0x25e230: file /home/dwightguth/test/./test.k, line 4.
(gdb) run
Breakpoint 1, TEST.test.sc (VarI=#token("0", "Int")) at /home/dwightguth/test/./test.k:4
4         rule [test]: I:Int => I +Int 1 requires I <Int 10
(gdb)

Note that every variable used in the side condition can have its value
inspected when stopped at this breakpoint, but other variables are not visible.

You can also break on a rule by its location:

(gdb) break test.k:4
Breakpoint 1 at 0x25e230: test.k:4. (2 locations)
(gdb) run
Breakpoint 1, TEST.test.sc (VarI=#token("0", "Int")) at /home/dwightguth/test/./test.k:4
4         rule [test]: I:Int => I +Int 1 requires I <Int 10
(gdb) continue
Continuing.

Breakpoint 1, TEST.test.rhs (VarDotVar0=`<generatedCounter>{}`(#token("0", "Int")), VarDotVar1=dotk{}(.KList), VarI=#token("0", "Int")) at /home/dwightguth/test/./test.k:4
4         rule [test]: I:Int => I +Int 1 requires I <Int 10
(gdb) continue
Continuing.

Breakpoint 1, TEST.test.sc (VarI=#token("1", "Int")) at /home/dwightguth/test/./test.k:4
4         rule [test]: I:Int => I +Int 1 requires I <Int 10
(gdb)

Note that this sets a breakpoint at two locations: one on the side condition
and one on the right hand side. If the rule had no side condition, the first
would not be set. You can also view the locations of the breakpoints and
disable them individually:

(gdb) info breakpoint
Num     Type           Disp Enb Address            What
1       breakpoint     keep y   <MULTIPLE>
        breakpoint already hit 3 times
1.1                         y     0x000000000025e230 in TEST.test.sc at /home/dwightguth/test/./test.k:4
1.2                         y     0x000000000025e250 in TEST.test.rhs at /home/dwightguth/test/./test.k:4
(gdb) disable 1.1
(gdb) continue
Continuing.

Breakpoint 1, TEST.test.rhs (VarDotVar0=`<generatedCounter>{}`(#token("0", "Int")), VarDotVar1=dotk{}(.KList), VarI=#token("1", "Int")) at /home/dwightguth/test/./test.k:4
4         rule [test]: I:Int => I +Int 1 requires I <Int 10
(gdb) continue
Continuing.

Breakpoint 1, TEST.test.rhs (VarDotVar0=`<generatedCounter>{}`(#token("0", "Int")), VarDotVar1=dotk{}(.KList), VarI=#token("2", "Int")) at /home/dwightguth/test/./test.k:4
4         rule [test]: I:Int => I +Int 1 requires I <Int 10
(gdb)

Now only the breakpoint when the rule applies is enabled.

Breaking on a function

You can also break when a particular function in your semantics is invoked:

(gdb) info functions foo
All functions matching regular expression "foo":

File /home/dwightguth/test/./test.k:
struct __mpz_struct *Lblfoo'LParUndsRParUnds'TEST'UndsUnds'Int(struct __mpz_struct *);
(gdb) break Lblfoo'LParUndsRParUnds'TEST'UndsUnds'Int
Breakpoint 1 at 0x25e640: file /home/dwightguth/test/./test.k, line 6.
(gdb) run
Breakpoint 1, Lblfoo'LParUndsRParUnds'TEST'UndsUnds'Int (_1=#token("1", "Int")) at /home/dwightguth/test/./test.k:6
6         syntax Int ::= foo(Int) [function]
(gdb)

In this case, the variables have numbers instead of names because the names of
arguments in functions in K come from rules, and we are stopped before any
specific rule has applied. For example, _1 is the first argument to the
function.

You can also set a breakpoint in this location by setting it on the line
associated with its production:

(gdb) break test.k:6
Breakpoint 1 at 0x25e640: file /home/dwightguth/test/./test.k, line 6.
(gdb) run
Breakpoint 1, Lblfoo'LParUndsRParUnds'TEST'UndsUnds'Int (_1=#token("1", "Int")) at /home/dwightguth/test/./test.k:6
6         syntax Int ::= foo(Int) [function]

These two syntaxes are equivalent; use whichever is easier for you.

You can also view the stack of function applications:

(gdb) bt
#0  Lblfoo'LParUndsRParUnds'TEST'UndsUnds'Int (_1=#token("1", "Int")) at /home/dwightguth/test/./test.k:6
#1  0x000000000025e5f8 in apply_rule_111 (VarDotVar0=`<generatedCounter>{}`(#token("0", "Int")), VarDotVar1=dotk{}(.KList)) at /home/dwightguth/test/./test.k:9
#2  0x0000000000268a52 in take_steps ()
#3  0x000000000026b7b4 in main ()
(gdb)

Here we see that foo was invoked while applying the rule on line 9 of test.k,
and we also can see the substitution of that rule. If foo was evaluated while
evaluating another function, we would also be able to see the arguments of that
function as well, unless the function was tail recursive, in which case no
stack frame would exist once the tail call was performed.

Breaking on a set of rules or functions

Using rbreak <regex> you can set breakpoints on multiple functions.

  • rbreak Lbl - sets a breakpoint on all non hooked functions

  • rbreak Lbl.*TEST - sets a breakpoint on all functions from module TEST

  • rbreak hook_INT - sets a breakpoint on all hooks from module INT

Other debugger issues

  • <optimized out> try kompiling without -O1, -O2, or -O3.
  • (gdb) break definition.kore:break -> No source file named definition.kore.
    send --enable-llvm-debug to kompile in order to generate debug info symbols.

Profiling your K semantics

The first thing to be aware of is in order to get meaningful data,
you need to build the semantics and all of its dependencies with
optimizations enabled but without the frame pointer elimination
optimization
. For example, for EVM, this means rebuilding GMP, MPFR,
JEMalloc, Crypto++, SECP256K1, etc with the following exports.

export CFLAGS="-DNDEBUG -O2 -fno-omit-frame-pointer"
export CXXFLAGS="-DNDEBUG -O2 -fno-omit-frame-pointer"

You can skip this step, but if you do, any samples within these
libraries will not have correct stack trace information, which means
you will likely not get a meaningful set of data that will tell you
where the majority of time is really being spent. Don't worry about
rebuilding literally every single dependency though. Just focus on the
ones that you expect to take a non-negligible amount of runtime. You
will be able to tell if you haven't done enough later, and you can go
back and rebuild more. Once this is done, you then build K with
optimizations and debug info enabled, like so:

mvn package -Dproject.build.type="FastBuild"

Next, you build the semantics with optimizations and debug info
enabled (i.e., kompile -ccopt -O2 --iterated -ccopt -fno-omit-frame-pointer).

Once all this is done, you should be ready to profile your
application. Essentially, you should run whatever test suite you
usually run, but with perf record -g -- prefixed to the front. For
example, for KEVM it's the following command. (For best data, don't
run this step in parallel.)

perf record -g -- make test-conformance

Finally, you want to filter out just the samples that landed within
the llvm backend and view the report. For this, you need to know the
name of the binary that was generated by your build system. Normally
it is interpreter, but e.g. if you are building the web3 client for
kevm, it would be kevm-client. You will want to run the following
command.

perf report -g -c $binary_name

If all goes well, you should see a breakdown of where CPU time has
been spent executing the application. You will know that sufficient
time was spent rebuilding dependencies with the correct flags when the
total time reported by the main method is close to 100%. If it's not
close to 100%, this is probably because a decent amount of self time
was reported in stack traces that were not built with frame pointers
enabled, meaning that perf was unable to walk the stack. You will have
to go back, rebuild the appropriate libraries, and then record your
trace again.

Your ultimate goal is to identify the hotspots that take the most
time, and make them execute faster. Entries like step and
step_1234 like functions refer to the cost of matching. An entry
like side_condition_1234 is a side condition and apply_rule_1234
is constructing the rhs of a rule. You can convert from this rule
ordinal to a location using the llvm-kompile-compute-loc script in
the bin folder of the llvm backend repo. For example,

llvm-kompile-compute-loc 5868 evm-semantics/.build/defn/llvm/driver-kompiled

spits out the following text.

Line: 18529
/home/dwightguth/evm-semantics/./.build/defn/llvm/driver.k:493:10

This is the line of definition.kore that the axiom appears on as
well as the original location of the rule in the K semantics. You can
use this information to figure out which rules and functions are
causing the most time and optimize them to be more efficient.

Running tests - kserver

The kserver is a front-end tool based on Nailgun
which helps to reduce the startup time of the JVM. Calling kserver in a terminal
window will wait for all kompile/kprove calls and force them to run in the same process
and share the same threads. This also reduces the thread contention significantly. kompile
uses all the threads available to do rule parsing. Another benefit is that it saves caches,
and each time you call kprove/kast, you can access those directly w/o extra disk usage.
Running the regression-new integration tests on a powerful machine (32 threads) takes 8m,
with the kserver active it takes 2m. You can start the kserver in two ways.

  • blocking: call kserver in the command line. Close it after you are done testing. Useful for quick testing.
  • non-blocking: call spawn-kserver <log.flie> and close it with stop-kserver - this is used for automation on CI

Because we reuse caches, you should stop and restart the server between runs.
The Nailgun implementation hasn't been updated in the last 3-5 years, and it's not compatible with Java 18 onwards.

K Builtins

The K Builtins (also referred to as the K Prelude or the K Standard Library)
consists of several files which contain definitions that make working with K
simpler. These files can be found under include/kframework/builtin in your K
installation directory, and can be imported with requires "FILENAME" (without
the path prefix).

  • domains: Basic datatypes which are universally useful.
  • kast: Representation of K internal data-structures (not to be
    included in normal definitions).
  • prelude: Automatically included into every K definition.
  • ffi: FFI interface for calling out to native C code from K.
  • json: JSON datatype and parsers/unparsers for JSON strings.
  • rat: Rational number representation.
  • substitution: Hooked implementation of capture-aware
    sustitution for K definitions.
  • unification: Hooked implementation of unification
    exposed directly to K definitions.

Basic Builtin Types in K

A major piece of the K prelude consists of a series of modules that contain
implementations of basic data types and language features in K. You do not need
to require this file yourself; it is required automatically in every K
definition unless --no-prelude is passed to kompile. K may not work correctly
if some of these modules do not exist or do not declare certain functions.

Note that some functions in the K prelude functions are not total, that is,
they are not defined on all possible input values. When you invoke such a
function on an undefined input, the behavior is undefined. In particular, when
this happens, interpreters generated by the K LLVM backend may crash.

requires "kast.md"

Default Modules

K declares certain modules that contain most of the builtins you usually want
when defining a language in K. In particular, this includes integers, booleans,
strings, identifiers, I/O, lists, maps, and sets. The DOMAINS-SYNTAX module
is designed to be imported by the syntax module of the language and contains
only the program-level syntax of identifiers, integers, booleans, and strings.
The DOMAINS module contains the rest of the syntax, including builtin
functions over those and the remaining types.

Note that not all modules are included in DOMAINS. A few less-common modules
are not, including ARRAY, COLLECTIONS, FLOAT, STRING-BUFFER, BYTES,
K-REFLECTION, MINT.

module DOMAINS-SYNTAX
  imports SORT-K
  imports ID-SYNTAX
  imports UNSIGNED-INT-SYNTAX
  imports BOOL-SYNTAX
  imports STRING-SYNTAX
endmodule

module DOMAINS
  imports DOMAINS-SYNTAX
  imports INT
  imports BOOL
  imports STRING
  imports BASIC-K
  imports LIST
  imports K-IO
  imports MAP
  imports SET
  imports ID
  imports K-EQUAL
endmodule

Arrays

Provided here is an implementation for fixed-sized, contiguous maps from Int
to KItem. In some previous versions of K, the Array type was a builtin type
backed by mutable arrays of objects. However, in modern K, the Array type is
implemented by means of the List type; users should not access this interface
directly and should instead make only of the functions listed below. Users of
this module should import only the ARRAY module.

module ARRAY-SYNTAX
  imports private LIST

  syntax Array

Array lookup

You can look up an element in an Array by its index in O(log(N)) time. Note
that the base of the logarithm is a relatively high number and thus the time is
effectively constant.

  syntax KItem ::= Array "[" Int "]" [function]

Array update

You can create a new Array with a new value for a key in O(log(N)) time, or
effectively constant.

  syntax Array ::= Array "[" key: Int "<-" value: KItem "]" [function, klabel(_[_<-_]), symbol]

Array reset

You can create a new Array where a particular key is reset to its default
value in O(log(N)) time, or effectively constant.

  syntax Array ::= Array "[" Int "<-" "undef" "]" [function]

Multiple array update

You can create a new Array from a List L of size N where the N
elements starting at index are replaced with the contents of L, in
O(N*log(K)) time (where K is the size of the array), or effectively linear.
Having index + N > K yields an exception.

  syntax Array ::= updateArray(Array, index: Int, List) [function]

Array fill

You can create a new Array where the length elements starting at index
are replaced with value, in O(length*log(N)) time, or effectively linear.

  syntax Array ::= fillArray(Array, index: Int, length: Int, value: KItem) [function]

Array range check

You can test whether an integer is within the bounds of an array in O(1) time.

  syntax Bool ::= Int "in_keys" "(" Array ")" [function, total]
endmodule

module ARRAY-IN-K [private]
  imports public ARRAY-SYNTAX
  imports private LIST
  imports private K-EQUAL
  imports private INT
  imports private BOOL

Array creation

You can create an array with length elements where each element is
initialized to value in O(1) time. Note that the array is stored in a manner
where only the highest element that is actually modified is given a value
in its internal representation, which means that subsequent array operations
may incur a one-time O(N) resizing cost, possibly amortized across multiple
operations.

  syntax Array ::= makeArray(length: Int, value: KItem) [function, public]

Implementation of Arrays

The remainder of this section consists of an implementation in K of the
operations listed above. Users of the ARRAY module should not make use
of any of the syntax defined in any of these modules.

  syntax Array ::= arr(List, Int, KItem)

  rule makeArray(I::Int, D::KItem) => arr(.List, I, D)

  rule arr(L::List, _, _       ) [ IDX::Int ] => L[IDX] requires 0 <=Int IDX andBool IDX  <Int size(L)
  rule arr(_      , _, D::KItem) [ _        ] => D      [owise]

  syntax List ::= ensureOffsetList(List, Int, KItem) [function]
  rule ensureOffsetList(L::List, IDX::Int, D::KItem) => L makeList(IDX +Int 1 -Int size(L), D) requires         IDX >=Int size(L)
  rule ensureOffsetList(L::List, IDX::Int, _::KItem) => L                                      requires notBool IDX >=Int size(L)

  rule arr(L::List, I::Int, D::KItem) [ IDX::Int <- VAL::KItem ] => arr(ensureOffsetList(L, IDX, D) [ IDX <- VAL ], I, D)

  rule arr(L::List, I::Int, D::KItem) [ IDX::Int <- undef ] => arr(L, I, D) [ IDX <- D ]

  rule updateArray(arr(L::List, I::Int, D::KItem), IDX::Int, L2::List) => arr(updateList(ensureOffsetList(L, IDX +Int size(L2) -Int 1, D), IDX, L2), I, D)

  rule fillArray(arr(L::List, I::Int, D::KItem), IDX::Int, LEN::Int, VAL::KItem) => arr(fillList(ensureOffsetList(L, IDX +Int LEN -Int 1, D), IDX, LEN, VAL), I, D)

  rule IDX::Int in_keys(arr(_, I::Int, _)) => IDX >=Int 0 andBool IDX <Int I
endmodule

module ARRAY-SYMBOLIC [symbolic]
  imports ARRAY-IN-K
endmodule

module ARRAY-KORE
  imports ARRAY-IN-K
endmodule

module ARRAY
  imports ARRAY-SYMBOLIC
  imports ARRAY-KORE
endmodule

Maps

Provided here is the syntax of an implementation of immutable, associative,
commutative maps from KItem to KItem. This type is hooked to an
implementation of maps provided by the backend. For more information on
matching on maps and allowable patterns for doing so, refer to K's
user documentation.

module MAP
  imports private BOOL-SYNTAX
  imports private INT-SYNTAX
  imports private LIST
  imports private SET

  syntax Map [hook(MAP.Map)]

Map concatenation

The Map sort represents a generalized associative array. Each key can be
paired with an arbitrary value, and can be used to reference its associated
value. Multiple bindings for the same key are not allowed.

You can construct a new Map consisting of key/value pairs of two Maps. The
result is #False if the maps have keys in common (in particular, this will
yield an exception during concrete execution). This operation is O(Nlog(M))
where N is the size of the smaller map, when it appears on the right hand side.
When it appears on the left hand side and all variables are bound, it is
O(N
log(M)) where M is the size of the map it is matching and N is the number
of elements being matched. When it appears on the left hand side containing
variables not bound elsewhere in the term, it is O(N^K) where N is the size of
the map it is matching and K is the number of unbound keys being matched. In
other words, one unbound variable is linear, two is quadratic, three is cubic,
etc.

  syntax Map ::= Map Map                        [left, function, hook(MAP.concat), klabel(_Map_), symbol, assoc, comm, unit(.Map), element(_|->_), index(0), format(%1%n%2)]

Map unit

The map with zero elements is represented by .Map.

  syntax Map ::= ".Map"                         [function, total, hook(MAP.unit), klabel(.Map), symbol]

Map elements

An element of a Map is constructed via the |-> operator. The key is on the
left and the value is on the right.

  syntax Map ::= KItem "|->" KItem                      [function, total, hook(MAP.element), klabel(_|->_), symbol, injective]

  syntax priority _|->_ > _Map_ .Map
  syntax non-assoc _|->_

Map lookup

You can look up the value associated with the key of a map in O(log(N)) time.
Note that the base of the logarithm is a relatively high number and thus the
time is effectively constant. The value is #False if the key is not in the
map (in particular, this will yield an exception during concrete execution).

  syntax KItem ::= Map "[" KItem "]"                    [function, hook(MAP.lookup), klabel(Map:lookup), symbol]

Map lookup with default

You can also look up the value associated with the key of a map using a
total function that assigns a specific default value if the key is not present
in the map. This operation is also O(log(N)), or effectively constant.

  syntax KItem ::= Map "[" KItem "]" "orDefault" KItem      [function, total, hook(MAP.lookupOrDefault), klabel(Map:lookupOrDefault)]

Map update

You can insert a key/value pair into a map in O(log(N)) time, or effectively
constant.

  syntax Map ::= Map "[" key: KItem "<-" value: KItem "]"           [function, total, klabel(Map:update), symbol, hook(MAP.update), prefer]

Map delete

You can remove a key/value pair from a map via its key in O(log(N)) time, or
effectively constant.

  syntax Map ::= Map "[" KItem "<-" "undef" "]"     [function, total, hook(MAP.remove), klabel(_[_<-undef]), symbol]

Map difference

You can remove the key/value pairs in a map that are present in another map in
O(N*log(M)) time (where M is the size of the first map and N is the size of the
second), or effectively linear. Note that only keys whose value is the same
in both maps are removed. To remove all the keys in one map from another map,
you can say removeAll(M1, keys(M2)).

  syntax Map ::= Map "-Map" Map                 [function, total, hook(MAP.difference)]

Multiple map update

You can update a map by adding all the key/value pairs in the second map in
O(N*log(M)) time (where M is the size of the first map and N is the size of the
second map), or effectively linear. If any keys are present in both maps, the
value from the second map overwrites the value in the first. This function is
total, which is distinct from map concatenation, a partial function only
defined on maps with disjoint keys.

  syntax Map ::= updateMap(Map, Map)            [function, total, hook(MAP.updateAll)]

Multiple map removal

You can remove a Set of keys from a map in O(N*log(M)) time (where M is the
size of the Map and N is the size of the Set), or effectively linear.

  syntax Map ::= removeAll(Map, Set)            [function, total, hook(MAP.removeAll)]

Map keys (as Set)

You can get a Set of all the keys in a Map in O(N) time.

  syntax Set ::= keys(Map)                      [function, total, hook(MAP.keys)]

Map keys (as List)

You can get a List of all the keys in a Map in O(N) time.

  syntax List ::= "keys_list" "(" Map ")"       [function, hook(MAP.keys_list)]

Map key membership

You can check whether a key is present in a map in O(1) time.

  syntax Bool ::= KItem "in_keys" "(" Map ")"       [function, total, hook(MAP.in_keys)]

Map values (as List)

You can get a List of all the values in a map in O(N) time.

  syntax List ::= values(Map)                   [function, hook(MAP.values)]

Map size

You can get the number of key/value pairs in a map in O(1) time.

  syntax Int ::= size(Map)                      [function, total, hook(MAP.size), klabel(sizeMap)]

Map inclusion

You can determine whether a Map is a strict subset of another Map in O(N)
time (where N is the size of the first map). Only keys that are bound to the
same value are considered equal.

  syntax Bool ::= Map "<=Map" Map               [function, total, hook(MAP.inclusion)]

Map choice

You can get an arbitrarily chosen key of a Map in O(1) time. The same key
will always be returned for the same map, but no guarantee is given that two
different maps will return the same element, even if they are similar.

  syntax KItem ::= choice(Map)                      [function, hook(MAP.choice), klabel(Map:choice)]

Implementation of Maps

The remainder of this section contains lemmas used by the Java and Haskell
backend to simplify expressions of sort Map. They do not affect the semantics
of maps, merely describing additional rules that the backend can use to
simplify terms.

endmodule

module MAP-KORE-SYMBOLIC [symbolic,haskell]
  imports MAP
  imports private K-EQUAL
  imports private BOOL

  rule #Ceil(@M:Map [@K:KItem]) => {(@K in_keys(@M)) #Equals true} #And #Ceil(@M) #And #Ceil(@K) [simplification]

  // Symbolic update

  // Adding the definedness condition `notBool (K in_keys(M))` in the ensures clause of the following rule would be redundant
  // because K also appears in the rhs, preserving the case when it's #Bottom.
  rule (K |-> _ M:Map) [ K <- V ] => (K |-> V M) [simplification]
  rule M:Map [ K <- V ] => (K |-> V M) requires notBool (K in_keys(M)) [simplification]
  rule M:Map [ K <- _ ] [ K <- V ] => M [ K <- V ] [simplification]
  // Adding the definedness condition `notBool (K1 in_keys(M))` in the ensures clause of the following rule would be redundant
  // because K1 also appears in the rhs, preserving the case when it's #Bottom.
  rule (K1 |-> V1 M:Map) [ K2 <- V2 ] => (K1 |-> V1 (M [ K2 <- V2 ])) requires K1 =/=K K2 [simplification]

  // Symbolic remove
  rule (K |-> _ M:Map) [ K <- undef ] => M ensures notBool (K in_keys(M)) [simplification]
  rule M:Map [ K <- undef ] => M requires notBool (K in_keys(M)) [simplification]
  // Adding the definedness condition `notBool (K1 in_keys(M))` in the ensures clause of the following rule would be redundant
  // because K1 also appears in the rhs, preserving the case when it's #Bottom.
  rule (K1 |-> V1 M:Map) [ K2 <- undef ] => (K1 |-> V1 (M [ K2 <- undef ])) requires K1 =/=K K2 [simplification]

  // Symbolic lookup
  rule (K  |->  V M:Map) [ K ]  => V ensures notBool (K in_keys(M)) [simplification]
  rule (K1 |-> _V M:Map) [ K2 ] => M [K2] requires K1 =/=K K2 ensures notBool (K1 in_keys(M)) [simplification]
  rule (_MAP:Map [ K  <-  V1 ]) [ K ]  => V1 [simplification]
  rule ( MAP:Map [ K1 <- _V1 ]) [ K2 ] => MAP [ K2 ] requires K1 =/=K K2 [simplification]

  rule (K  |->  V M:Map) [  K ] orDefault _ => V ensures notBool (K in_keys(M)) [simplification]
  rule (K1 |-> _V M:Map) [ K2 ] orDefault D => M [K2] orDefault D requires K1 =/=K K2 ensures notBool (K1 in_keys(M)) [simplification]
  rule (_MAP:Map [ K  <-  V1 ]) [ K ] orDefault _ => V1 [simplification]
  rule ( MAP:Map [ K1 <- _V1 ]) [ K2 ] orDefault D => MAP [ K2 ] orDefault D requires K1 =/=K K2 [simplification]
  rule .Map [ _ ] orDefault D => D [simplification]

  // Symbolic in_keys
  rule K in_keys(_M [ K <- undef ]) => false [simplification]
  rule K in_keys(_M [ K <- _ ]) => true [simplification]
  rule K1 in_keys(M [ K2 <- _ ]) => true requires K1 ==K K2 orBool K1 in_keys(M) [simplification]
  rule K1 in_keys(M [ K2 <- _ ]) => K1 in_keys(M) requires K1 =/=K K2 [simplification]

  rule {false #Equals @Key in_keys(.Map)} => #Ceil(@Key) [simplification]
  rule {@Key in_keys(.Map) #Equals false} => #Ceil(@Key) [simplification]
  rule {false #Equals @Key in_keys(Key' |-> Val @M)} => #Ceil(@Key) #And #Ceil(Key' |-> Val @M) #And #Not({@Key #Equals Key'}) #And {false #Equals @Key in_keys(@M)} [simplification]
  rule {@Key in_keys(Key' |-> Val @M) #Equals false} => #Ceil(@Key) #And #Ceil(Key' |-> Val @M) #And #Not({@Key #Equals Key'}) #And {@Key in_keys(@M) #Equals false} [simplification]

/*
// The rule below is automatically generated by the frontend for every sort
// hooked to MAP.Map. It is left here to serve as documentation.

  rule #Ceil(@M:Map (@K:KItem |-> @V:KItem)) => {(@K in_keys(@M)) #Equals false} #And #Ceil(@M) #And #Ceil(@K) #And #Ceil(@V)
    [simplification]
*/
endmodule

module MAP-SYMBOLIC
  imports MAP-KORE-SYMBOLIC
endmodule

Range Maps

Provided here is the syntax of an implementation of immutable, associative,
commutative range maps from Int to KItem. This type is hooked to an
implementation of range maps provided by the LLVM backend.
Currently, this type is not supported by other backends.
Although the underlying range map data structure supports any key sort, the
current implementation by the backend only supports Int keys due to
limitations of the underlying ordering function.

module RANGEMAP
  imports private BOOL-SYNTAX
  imports private INT-SYNTAX
  imports private LIST
  imports private SET

Range, bounded inclusively below and exclusively above.

  syntax Range ::= "[" KItem "," KItem ")"    [klabel(Rangemap:Range), symbol]

  syntax RangeMap [hook(RANGEMAP.RangeMap)]

Range map concatenation

The RangeMap sort represents a map whose keys are stored as ranges, bounded
inclusively below and exclusively above. Contiguous or overlapping ranges that
map to the same value are merged into a single range.

You can construct a new RangeMap consisting of range/value pairs of two
RangeMaps. If the RangeMaps have overlapping ranges an exception will be
thrown during concrete execution. This operation is O(N*log(M)) (where N is
the size of the smaller map and M is the size of the larger map).

  syntax RangeMap ::= RangeMap RangeMap                        [left, function, hook(RANGEMAP.concat), klabel(_RangeMap_), symbol, assoc, comm, unit(.RangeMap), element(_r|->_), index(0), format(%1%n%2)]

Range map unit

The RangeMap with zero elements is represented by .RangeMap.

  syntax RangeMap ::= ".RangeMap"                         [function, total, hook(RANGEMAP.unit), klabel(.RangeMap), symbol]

Range map elements

An element of a RangeMap is constructed via the r|-> operator. The range
of keys is on the left, and the value is on the right.

  syntax RangeMap ::= Range "r|->" KItem                      [function, hook(RANGEMAP.elementRng), klabel(_r|->_), symbol, injective]

  syntax priority _r|->_ > _RangeMap_ .RangeMap
  syntax non-assoc _r|->_

Range map lookup

You can look up the value associated with a key of a RangeMap in O(log(N))
time (where N is the size of the RangeMap). This will yield an exception
during concrete execution if the key is not in the range map.

  syntax KItem ::= RangeMap "[" KItem "]"                    [function, hook(RANGEMAP.lookup), klabel(RangeMap:lookup), symbol]

Range map lookup with default

You can also look up the value associated with a key of a RangeMap using a
total function that assigns a specific default value if the key is not present
in the RangeMap. This operation is also O(log(N)) (where N is the size of
the range map).

  syntax KItem ::= RangeMap "[" KItem "]" "orDefault" KItem      [function, total, hook(RANGEMAP.lookupOrDefault), klabel(RangeMap:lookupOrDefault)]

Range map lookup for range of key

You can look up for the range that a key of a RangeMap is stored in in
O(log(N)) time (where N is the size of the RangeMap). This will yield an
exception during concrete execution if the key is not in the range map.

  syntax Range ::= "find_range" "(" RangeMap "," KItem ")"                    [function, hook(RANGEMAP.find_range), klabel(RangeMap:find_range)]

Range map update

You can insert a range/value pair into a RangeMap in O(log(N)) time (where N
is the size of the RangeMap). Any ranges adjacent to or overlapping with the
range to be inserted will be updated accordingly.

  syntax RangeMap ::= RangeMap "[" keyRange: Range "<-" value: KItem "]"           [function, klabel(RangeMap:update), symbol, hook(RANGEMAP.updateRng), prefer]

Range map delete

You can remove a range/value pair from a RangeMap in O(log(N)) time (where N
is the size of the RangeMap). If all or any part of the range is present in
the range map, it will be removed.

  syntax RangeMap ::= RangeMap "[" Range "<-" "undef" "]"     [function, hook(RANGEMAP.removeRng), klabel(_r[_<-undef]), symbol]

Range map difference

You can remove the range/value pairs in a RangeMap that are also present in
another RangeMap in O(max{M,N}*log(M)) time (where M is the size of the
first RangeMap and N is the size of the second RangeMap). Note that only
the parts of overlapping ranges whose value is the same in both range maps
will be removed.

  syntax RangeMap ::= RangeMap "-RangeMap" RangeMap                 [function, total, hook(RANGEMAP.difference)]

Multiple range map update

You can update a RangeMap by adding all the range/value pairs in the second
RangeMap in O(N*log(M+N)) time (where M is the size of the first RangeMap
and N is the size of the second RangeMap). If any ranges are overlapping,
the value from the second range map overwrites the value in the first for the
parts where ranges are overlapping. This function is total, which is distinct
from range map concatenation, a partial function only defined on range maps
with non overlapping ranges.

  syntax RangeMap ::= updateRangeMap(RangeMap, RangeMap)            [function, total, hook(RANGEMAP.updateAll)]

Multiple range map removal

You can remove a Set of ranges from a RangeMap in O(N*log(M)) time (where
M is the size of the RangeMap and N is the size of the Set). For every
range in the set, all or any part of it that is present in the range map will
be removed.

  syntax RangeMap ::= removeAll(RangeMap, Set)            [function, hook(RANGEMAP.removeAll)]

Range map keys (as Set)

You can get a Set of all the ranges in a RangeMap in O(N) time (where N
is the size of the RangeMap).

  syntax Set ::= keys(RangeMap)                      [function, total, hook(RANGEMAP.keys)]

Range map keys (as List)

You can get a List of all the ranges in a RangeMap in O(N) time (where N
is the size of the RangeMap).

  syntax List ::= "keys_list" "(" RangeMap ")"       [function, hook(RANGEMAP.keys_list)]

Range map key membership

You can check whether a key is present in a RangeMap in O(log(N)) time (where
N is the size of the RangeMap).

  syntax Bool ::= KItem "in_keys" "(" RangeMap ")"       [function, total, hook(RANGEMAP.in_keys)]

Range map values (as List)

You can get a List of all values in a RangeMap in O(N) time (where N is the
size of the RangeMap).

  syntax List ::= values(RangeMap)                   [function, hook(RANGEMAP.values)]

Range map size

You can get the number of range/value pairs in a RangeMap in O(1) time.

  syntax Int ::= size(RangeMap)                      [function, total, hook(RANGEMAP.size), klabel(sizeRangeMap)]

Range map inclusion

You can determine whether a RangeMap is a strict subset of another RangeMap
in O(M+N) time (where M is the size of the first RangeMap and N is the size
of the second RangeMap). Only keys within equal or overlapping ranges that
are bound to the same value are considered equal.

  syntax Bool ::= RangeMap "<=RangeMap" RangeMap               [function, total, hook(RANGEMAP.inclusion)]

Range map choice

You can get an arbitrarily chosen key of a RangeMap in O(1) time. The same
key will always be returned for the same range map, but no guarantee is given
that two different range maps will return the same element, even if they are
similar.

  syntax KItem ::= choice(RangeMap)                      [function, hook(RANGEMAP.choice), klabel(RangeMap:choice)]
endmodule

Sets

Provided here is the syntax of an implementation of immutable, associative,
commutative sets of KItem. This type is hooked to an implementation of sets
provided by the backend. For more information on matching on sets and allowable
patterns for doing so, refer to K's
user documentation.

module SET
  imports private INT-SYNTAX
  imports private BASIC-K

  syntax Set [hook(SET.Set)]

Set concatenation

The Set sort represents a mathematical set (A collection of unique items).
The sets are nilpotent, i.e., the concatenation of two sets containing elements
in common is #False (note however, this may be silently allowed during
concrete execution). If you intend to add an element to a set that might
already be present in the set, use the |Set operator instead.

The concatenation operator is O(Nlog(M)) where N is the size of the smaller
set, when it appears on the right hand side. When it appears on the left hand
side and all variables are bound, it is O(N
log(M)) where M is the size of the
set it is matching and N is the number of elements being matched. When it
appears on the left hand side containing variables not bound elsewhere in the
term, it is O(N^K) where N is the size of the set it is matching and K is the
number of unbound keys being mached. In other words, one unbound variable is
linear, two is quadratic, three is cubic, etc.

  syntax Set ::= Set Set                  [left, function, hook(SET.concat), klabel(_Set_), symbol, assoc, comm, unit(.Set), idem, element(SetItem), format(%1%n%2)]

Set unit

The set with zero elements is represented by .Set.

  syntax Set ::= ".Set"                   [function, total, hook(SET.unit), klabel(.Set), symbol]

Set elements

An element of a Set is constructed via the SetItem operator.

  syntax Set ::= SetItem(KItem)               [function, total, hook(SET.element), klabel(SetItem), symbol, injective]

Set union

You can compute the union of two sets in O(N*log(M)) time (Where N is the size
of the smaller set). Note that the base of the logarithm is a relatively high
number and thus the time is effectively linear. The union consists of all the
elements present in either set.

  syntax Set ::= Set "|Set" Set              [left, function, total, hook(SET.union), comm]
  rule S1:Set |Set S2:Set => S1 (S2 -Set S1) [concrete]

Set intersection

You can compute the intersection of two sets in O(N*log(M)) time (where N
is the size of the smaller set), or effectively linear. The intersection
consists of all the elements present in both sets.

  syntax Set ::= intersectSet(Set, Set)   [function, total, hook(SET.intersection), comm]

Set complement

You can compute the relative complement of two sets in O(N*log(M)) time (where
N is the size of the second set), or effectively linear. This is the set of
elements in the first set that are not present in the second set.

  syntax Set ::= Set "-Set" Set           [function, total, hook(SET.difference), klabel(Set:difference), symbol]

Set membership

You can compute whether an element is a member of a set in O(1) time.

  syntax Bool ::= KItem "in" Set              [function, total, hook(SET.in), klabel(Set:in), symbol]

Set inclusion

You can determine whether a Set is a strict subset of another Set in O(N)
time (where N is the size of the first set).

  syntax Bool ::= Set "<=Set" Set         [function, total, hook(SET.inclusion)]

Set size

You can get the number of elements (the cardinality) of a set in O(1) time.

  syntax Int ::= size(Set)                [function, total, hook(SET.size)]

Set choice

You can get an arbitrarily chosen element of a Set in O(1) time. The same
element will always be returned for the same set, but no guarantee is given
that two different sets will return the same element, even if they are similar.

  syntax KItem ::= choice(Set)                [function, hook(SET.choice), klabel(Set:choice)]
endmodule

Implementation of Sets

The following lemmas are simplifications that the Haskell backend can
apply to simplify expressions of sort Set.

module SET-KORE-SYMBOLIC [symbolic,haskell]
  imports SET
  imports private K-EQUAL
  imports private BOOL

  //Temporarly rule for #Ceil simplification, should be generated in front-end

// Matching for this version not implemented.
  // rule #Ceil(@S1:Set @S2:Set) =>
  //        {intersectSet(@S1, @S2) #Equals .Set} #And #Ceil(@S1) #And #Ceil(@S2)
  //   [simplification]

//simpler version
  rule #Ceil(@S:Set SetItem(@E:KItem)) =>
         {(@E in @S) #Equals false} #And #Ceil(@S) #And #Ceil(@E)
    [simplification]

  // -Set simplifications
  rule S              -Set .Set           => S          [simplification]
  rule .Set           -Set  _             => .Set       [simplification]
  rule SetItem(X)     -Set (S SetItem(X)) => .Set
                               ensures notBool (X in S) [simplification]
  rule S              -Set (S SetItem(X)) => .Set
                               ensures notBool (X in S) [simplification]
  rule (S SetItem(X)) -Set S              => SetItem(X)
                               ensures notBool (X in S) [simplification]
  rule (S SetItem(X)) -Set SetItem(X)     => S
                               ensures notBool (X in S) [simplification]
  // rule SetItem(X)     -Set S              => SetItem(X)
  //                            requires notBool (X in S)  [simplification]
  // rule (S1 SetItem(X)) -Set (S2 SetItem(X))  => S1 -Set S2
  //                             ensures notBool (X in S1)
  //                             andBool notBool (X in S2) [simplification]



  // |Set simplifications
  rule S    |Set .Set => S    [simplification, comm]
  rule S    |Set S    => S    [simplification]

  rule (S SetItem(X)) |Set SetItem(X) => S SetItem(X)
                             ensures notBool (X in S) [simplification, comm]
  // Currently disabled, see runtimeverification/haskell-backend#3301
  // rule (S SetItem(X)) |Set S          => S SetItem(X)
  //                            ensures notBool (X in S) [simplification, comm]

  // intersectSet simplifications
  rule intersectSet(.Set, _   ) => .Set    [simplification, comm]
  rule intersectSet( S  , S   ) =>  S      [simplification]

  rule intersectSet( S SetItem(X), SetItem(X))     => SetItem(X)
                                                        ensures notBool (X in S)      [simplification, comm]
  // Currently disabled, see runtimeverification/haskell-backend#3294
  // rule intersectSet( S SetItem(X) , S)             => S ensures notBool (X in S)      [simplification, comm]
  rule intersectSet( S1 SetItem(X), S2 SetItem(X)) => intersectSet(S1, S2) SetItem(X)
                                                        ensures notBool (X in S1)
                                                        andBool notBool (X in S2)     [simplification]

  // membership simplifications
  rule _E in .Set           => false   [simplification]
  rule E  in (S SetItem(E)) => true
              ensures notBool (E in S) [simplification]

// These two rules would be sound but impose a giant overhead on `in` evaluation:
  // rule E1 in (S SetItem(E2)) => true requires E1 in S
  //                                 ensures notBool (E2 in S) [simplification]
  // rule E1 in (S SetItem(E2)) => E1 in S requires E1 =/=K E2
  //                                 ensures notBool (E2 in S) [simplification]

  rule X in ((SetItem(X) S) |Set  _            ) => true
                                    ensures notBool (X in S) [simplification]
  rule X in ( _             |Set (SetItem(X) S)) => true
                                    ensures notBool (X in S) [simplification]

endmodule

module SET-SYMBOLIC
  imports SET-KORE-SYMBOLIC
endmodule

Lists

Provided here is the syntax of an implementation of immutable, associative
lists of KItem. This type is hooked to an implementation of lists provided
by the backend. For more information on matching on lists and allowable
patterns for doing so, refer to K's
user documentation.

module LIST
  imports private INT-SYNTAX
  imports private BASIC-K

  syntax List [hook(LIST.List)]

List concatenation

The List sort is an ordered collection that may contain duplicate elements.
They are backed by relaxed radix balanced trees, which means that they support
efficiently adding elements to both sides of the list, concatenating two lists,
indexing, and updating elements.

The concatenation operator is O(log(N)) (where N is the size of the longer
list) when it appears on the right hand side. When it appears on the left hand
side, it is O(N), where N is the number of elements matched on the front and
back of the list.

  syntax List ::= List List               [left, function, total, hook(LIST.concat), klabel(_List_), symbol, smtlib(smt_seq_concat), assoc, unit(.List), element(ListItem), format(%1%n%2)]

List unit

The list with zero elements is represented by .List.

  syntax List ::= ".List"                 [function, total, hook(LIST.unit), klabel(.List), symbol, smtlib(smt_seq_nil)]

List elements

An element of a List is constucted via the ListItem operator.

  syntax List ::= ListItem(KItem)             [function, total, hook(LIST.element), klabel(ListItem), symbol, smtlib(smt_seq_elem)]

List prepend

An element can be added to the front of a List using the pushList operator.

  syntax List ::= pushList(KItem, List)       [function, total, hook(LIST.push), klabel(pushList), symbol]
  rule pushList(K::KItem, L1::List) => ListItem(K) L1

List indexing

You can get an element of a list by its integer offset in O(log(N)) time, or
effectively constant. Positive indices are 0-indexed from the beginning of the
list, and negative indices are -1-indexed from the end of the list. In other
words, 0 is the first element and -1 is the last element.

  syntax KItem ::= List "[" Int "]"           [function, hook(LIST.get), klabel(List:get), symbol]

List update

You can create a new List with a new value at a particular index in
O(log(N)) time, or effectively constant.

  syntax List ::= List "[" index: Int "<-" value: KItem "]" [function, hook(LIST.update), klabel(List:set)]

List of identical elements

You can create a list with length elements, each containing value, in O(N)
time.

  syntax List ::= makeList(length: Int, value: KItem) [function, hook(LIST.make)]

Multiple list update

You can create a new List which is equal to dest except the N elements
starting at index are replaced with the contents of src in O(N*log(K)) time
(where K is the size of destand N is the size of src), or effectively linear. Having index + N > K yields an exception.

  syntax List ::= updateList(dest: List, index: Int, src: List) [function, hook(LIST.updateAll)]

List fill

You can create a new List where the length elements starting at index
are replaced with value, in O(length*log(N)) time, or effectively linear.

  syntax List ::= fillList(List, index: Int, length: Int, value: KItem) [function, hook(LIST.fill)]

List slicing

You can compute a new List by removing fromFront elements from the front
of the list and fromBack elements from the back of the list in
O((fromFront+fromBack)*log(N)) time, or effectively linear.

  syntax List ::= range(List, fromFront: Int, fromBack: Int)   [function, hook(LIST.range), klabel(List:range), symbol]

List membership

You can compute whether an element is in a list in O(N) time. For repeated
comparisons, it is much better to first convert to a set using List2Set.

  syntax Bool ::= KItem "in" List             [function, total, hook(LIST.in), klabel(_inList_)]

List size

You can get the number of elements of a list in O(1) time.

  syntax Int ::= size(List)               [function, total, hook(LIST.size), klabel (sizeList), smtlib(smt_seq_len)]
endmodule

Collection Conversions

It is possible to convert from a List to a Set or from a Set to a list.
Converting from a List to a Set and back will not provide the same list;
duplicates will have been removed and the list may be reordered. Converting
from a Set to a List and back will generate the same set.

Note that because sets are unordered and lists are ordered, converting from a
Set to a List will generate some arbitrary ordering of elements, which may
be different from the natural ordering you might assume, or may not. Two
equal sets are guaranteed to generate the same ordering, but no guarantee is
otherwise provided about what the ordering will be. In particular, adding an
element to a set may completely reorder the elements already in the set, when
it is converted to a list.

module COLLECTIONS
  imports LIST
  imports SET
  imports MAP

  syntax List ::= Set2List(Set) [function, total, hook(SET.set2list)]
  syntax Set ::= List2Set(List) [function, total, hook(SET.list2set)]

endmodule

Booleans

Provided here is the syntax of an implementation of boolean algebra in K.
This type is hooked to an implementation of booleans provided by the backend.
Note that this algebra is different from the builtin truth in matching logic.
You can, however, convert from the truth of the Bool sort to the truth in
matching logic via the expression {B #Equals true}.

The boolean values are true and false.

module SORT-BOOL
  syntax Bool [hook(BOOL.Bool)]
endmodule

module BOOL-SYNTAX
  imports SORT-BOOL
  syntax Bool ::= "true"  [token]
  syntax Bool ::= "false" [token]
endmodule

module BOOL-COMMON
  imports private BASIC-K
  imports BOOL-SYNTAX

Basic boolean arithmetic

You can:

  • Negate a boolean value.
  • AND two boolean values.
  • XOR two boolean values.
  • OR two boolean values.
  • IMPLIES two boolean values (i.e., P impliesBool Q is the same as
    notBool P orBool Q)
  • Check equality of two boolean values.
  • Check inequality of two boolean values.

Note that only andThenBool and orElseBool are short-circuiting. andBool
and orBool may be short-circuited in concrete backends, but in symbolic
backends, both arguments will be evaluated.

  syntax Bool ::= "notBool" Bool          [function, total, klabel(notBool_), symbol, smt-hook(not), group(boolOperation), hook(BOOL.not)]
                > Bool "andBool" Bool     [function, total, klabel(_andBool_), symbol, left, smt-hook(and), group(boolOperation), hook(BOOL.and)]
                | Bool "andThenBool" Bool [function, total, klabel(_andThenBool_), symbol, left, smt-hook(and), group(boolOperation), hook(BOOL.andThen)]
                | Bool "xorBool" Bool     [function, total, klabel(_xorBool_), symbol, left, smt-hook(xor), group(boolOperation), hook(BOOL.xor)]
                | Bool "orBool" Bool      [function, total, klabel(_orBool_), symbol, left, smt-hook(or), group(boolOperation), hook(BOOL.or)]
                | Bool "orElseBool" Bool  [function, total, klabel(_orElseBool_), symbol, left, smt-hook(or), group(boolOperation), hook(BOOL.orElse)]
                | Bool "impliesBool" Bool [function, total, klabel(_impliesBool_), symbol, left, smt-hook(=>), group(boolOperation), hook(BOOL.implies)]
                > left:
                  Bool "==Bool" Bool      [function, total, klabel(_==Bool_), symbol, left, comm, smt-hook(=), hook(BOOL.eq)]
                | Bool "=/=Bool" Bool     [function, total, klabel(_=/=Bool_), symbol, left, comm, smt-hook(distinct), hook(BOOL.ne)]

Implementation of Booleans

The remainder of this section consists of an implementation in K of the
operations listed above.

  rule notBool true => false
  rule notBool false => true

  rule true andBool B:Bool => B:Bool
  rule B:Bool andBool true => B:Bool [simplification]
  rule false andBool _:Bool => false
  rule _:Bool andBool false => false [simplification]

  rule true andThenBool K::Bool => K
  rule K::Bool andThenBool true => K [simplification]
  rule false andThenBool _ => false
  rule _ andThenBool false => false  [simplification]

  rule false xorBool B:Bool => B:Bool
  rule B:Bool xorBool false => B:Bool [simplification]
  rule B:Bool xorBool B:Bool => false

  rule true orBool _:Bool => true
  rule _:Bool orBool true => true [simplification]
  rule false orBool B:Bool => B
  rule B:Bool orBool false => B   [simplification]

  rule true orElseBool _ => true
  rule _ orElseBool true => true     [simplification]
  rule false orElseBool K::Bool => K
  rule K::Bool orElseBool false => K [simplification]

  rule true impliesBool B:Bool => B
  rule false impliesBool _:Bool => true
  rule _:Bool impliesBool true => true       [simplification]
  rule B:Bool impliesBool false => notBool B [simplification]

  rule B1:Bool =/=Bool B2:Bool => notBool (B1 ==Bool B2)
endmodule

module BOOL-KORE [symbolic]
  imports BOOL-COMMON

  rule {true #Equals notBool @B} => {false #Equals @B} [simplification]
  rule {notBool @B #Equals true} => {@B #Equals false} [simplification]
  rule {false #Equals notBool @B} => {true #Equals @B} [simplification]
  rule {notBool @B #Equals false} => {@B #Equals true} [simplification]

  rule {true #Equals @B1 andBool @B2} => {true #Equals @B1} #And {true #Equals @B2} [simplification]
  rule {@B1 andBool @B2 #Equals true} => {@B1 #Equals true} #And {@B2 #Equals true} [simplification]
  rule {false #Equals @B1 orBool @B2} => {false #Equals @B1} #And {false #Equals @B2} [simplification]
  rule {@B1 orBool @B2 #Equals false} => {@B1 #Equals false} #And {@B2 #Equals false} [simplification]
endmodule

module BOOL
  imports BOOL-COMMON
  imports BOOL-KORE
endmodule

Integers

Provided here is the syntax of an implementation of arbitrary-precision
integer arithmetic in K. This type is hooked to an implementation of integers
provided by the backend. For a fixed-width integer type, see the MINT module
below.

The UNSIGNED-INT-SYNTAX module provides a syntax of whole numbers in K.
This is useful because often programming languages implement the sign of an
integer as a unary operator rather than part of the lexical syntax of integers.
However, you can also directly reference integers with a sign using the
INT-SYNTAX module.

module UNSIGNED-INT-SYNTAX
  syntax Int [hook(INT.Int)]
  syntax Int ::= r"[0-9]+" [prefer, token, prec(2)]
endmodule

module INT-SYNTAX
  imports UNSIGNED-INT-SYNTAX
  syntax Int ::= r"[\\+-]?[0-9]+" [prefer, token, prec(2)]
endmodule

module INT-COMMON
  imports INT-SYNTAX
  imports private BOOL

Integer arithmetic

You can:

  • Compute the bitwise complement ~Int of an integer value in twos-complement.
  • Compute the exponentiation ^Int of two integers.
  • Compute the exponentiation of two integers modulo another integer (^%Int).
    A ^%Int B C is equal in value to (A ^Int B) %Int C, but has a better
    asymptotic complexity.
  • Compute the product *Int of two integers.
  • Compute the quotient /Int or modulus %Int of two integers using
    t-division, which rounds towards zero. Division by zero is #False.
  • Compute the quotient divInt or modulus modInt of two integers using
    Euclidean division, in which the remainder is always non-negative. Division
    by zero is #False.
  • Compute the sum +Int or difference -Int of two integers.
  • Compute the arithmetic right shift >>Int of two integers. Shifting by a
    negative quantity is #False.
  • Compute the left shift of two integers. Shifting by a negative quantity is
    #False.
  • Compute the bitwise and of two integers in twos-complement.
  • Compute the bitwise xor of two integers in twos-complement.
  • Compute the bitwise inclusive-or of two integers in twos-complement.
  syntax Int ::= "~Int" Int                     [function, klabel(~Int_), symbol, total, hook(INT.not), smtlib(notInt)]
               > left:
                 Int "^Int" Int                 [function, klabel(_^Int_), symbol, left, smt-hook(^), hook(INT.pow)]
               | Int "^%Int" Int Int            [function, klabel(_^%Int__), symbol, left, smt-hook((mod (^ #1 #2) #3)), hook(INT.powmod)]
               > left:
                 Int "*Int" Int                 [function, total, klabel(_*Int_), symbol, left, comm, smt-hook(*), hook(INT.mul)]
               /* FIXME: translate /Int and %Int into smtlib */
               /* /Int and %Int implement t-division, which rounds towards 0 */
               | Int "/Int" Int                 [function, klabel(_/Int_), symbol, left, smt-hook(div), hook(INT.tdiv)]
               | Int "%Int" Int                 [function, klabel(_%Int_), symbol, left, smt-hook(mod), hook(INT.tmod)]
               /* divInt and modInt implement e-division according to the Euclidean division theorem, therefore the remainder is always positive */
               | Int "divInt" Int               [function, klabel(_divInt_), symbol, left, smt-hook(div), hook(INT.ediv)]
               | Int "modInt" Int               [function, klabel(_modInt_), symbol, left, smt-hook(mod), hook(INT.emod)]
               > left:
                 Int "+Int" Int                 [function, total, klabel(_+Int_), symbol, left, comm, smt-hook(+), hook(INT.add)]
               | Int "-Int" Int                 [function, total, klabel(_-Int_), symbol, left, smt-hook(-), hook(INT.sub)]
               > left:
                 Int ">>Int" Int                [function, klabel(_>>Int_), symbol, left, hook(INT.shr), smtlib(shrInt)]
               | Int "<<Int" Int                [function, klabel(_<<Int_), symbol, left, hook(INT.shl), smtlib(shlInt)]
               > left:
                 Int "&Int" Int                 [function, total, klabel(_&Int_), symbol, left, comm, hook(INT.and), smtlib(andInt)]
               > left:
                 Int "xorInt" Int               [function, total, klabel(_xorInt_), symbol, left, comm, hook(INT.xor), smtlib(xorInt)]
               > left:
                 Int "|Int" Int                 [function, total, klabel(_|Int_), symbol, left, comm, hook(INT.or), smtlib(orInt)]

Integer minimum and maximum

You can compute the minimum and maximum minInt and maxInt of two integers.

  syntax Int ::= "minInt" "(" Int "," Int ")"   [function, total, smt-hook((ite (< #1 #2) #1 #2)), hook(INT.min)]
               | "maxInt" "(" Int "," Int ")"   [function, total, smt-hook((ite (< #1 #2) #2 #1)), hook(INT.max)]

Absolute value

You can compute the absolute value absInt of an integer.

  syntax Int ::= absInt ( Int )                 [function, total, smt-hook((ite (< #1 0) (- 0 #1) #1)), hook(INT.abs)]

Log base 2

You can compute the log base 2, rounded towards zero, of an integer. The log
base 2 of an integer is equal to the index of the highest bit set in the
representation of a positive integer. Log base 2 of zero or a negative number
is #False.

  syntax Int ::= log2Int ( Int )                [function, hook(INT.log2)]

Bit slicing

You can compute the value of a range of bits in the twos-complement
representation of an integer, as interpeted either unsigned or signed, of an
integer. index is offset from 0 and length is the number of bits, starting
with index, that should be read. The number is assumed to be represented
in little endian notation with each byte going from least significant to
most significant. In other words, 0 is the least-significant bit, and each
successive bit is more significant than the last.

  syntax Int ::= bitRangeInt           ( Int, index: Int, length: Int ) [function, hook(INT.bitRange)]
               | signExtendBitRangeInt ( Int, index: Int, length: Int ) [function, hook(INT.signExtendBitRange)]

Integer comparisons

You can compute whether two integers are less than or equal to, less than,
greater than or equal to, greater than, equal, or unequal to another integer.

  syntax Bool ::= Int "<=Int" Int         [function, total, klabel(_<=Int_), symbol, smt-hook(<=), hook(INT.le)]
                | Int "<Int" Int          [function, total, klabel(_<Int_), symbol, smt-hook(<), hook(INT.lt)]
                | Int ">=Int" Int         [function, total, klabel(_>=Int_), symbol, smt-hook(>=), hook(INT.ge)]
                | Int ">Int" Int          [function, total, klabel(_>Int_), symbol, smt-hook(>), hook(INT.gt)]
                | Int "==Int" Int         [function, total, klabel(_==Int_), symbol, comm, smt-hook(=), hook(INT.eq)]
                | Int "=/=Int" Int        [function, total, klabel(_=/=Int_), symbol, comm, smt-hook(distinct), hook(INT.ne)]

Divides

You can compute whether one integer evenly divides another. This is the
case when the second integer modulo the first integer is equal to zero.

  syntax Bool ::= Int "dividesInt" Int    [function]

Random integers

You can, on concrete backends, compute a pseudorandom integer, or seed the
pseudorandom number generator. These operations are represented as
uninterpreted functions on symbolic backends.

  syntax Int ::= randInt(Int) [function, hook(INT.rand), impure]
  syntax K ::= srandInt(Int) [function, hook(INT.srand), impure]

Implementation of Integers

The remainder of this section consists of an implementation in K of some
of the operators above, as well as lemmas used by the Java and Haskell backend
to simplify expressions of sort Int. They do not affect the semantics of
integers, merely describing additional rules that the backend can use to
simplify terms.

endmodule

module INT-SYMBOLIC [symbolic]
  imports INT-COMMON
  imports INT-SYMBOLIC-KORE
  imports private BOOL

  // Arithmetic Normalization
  rule I +Int 0 => I [simplification]
  rule I -Int 0 => I [simplification]

  rule X modInt N => X requires 0 <=Int X andBool X <Int N [simplification]
  rule X   %Int N => X requires 0 <=Int X andBool X <Int N [simplification]

  // Bit-shifts
  rule X <<Int 0 => X [simplification]
  rule 0 <<Int _ => 0 [simplification]
  rule X >>Int 0 => X [simplification]
  rule 0 >>Int _ => 0 [simplification]
endmodule

module INT-SYMBOLIC-KORE [symbolic, haskell]
  imports INT-COMMON
  imports ML-SYNTAX
  imports private BOOL

  // Definability Conditions
  rule #Ceil(@I1:Int /Int   @I2:Int) => {(@I2 =/=Int 0) #Equals true} #And #Ceil(@I1) #And #Ceil(@I2) [simplification]
  rule #Ceil(@I1:Int %Int   @I2:Int) => {(@I2 =/=Int 0) #Equals true} #And #Ceil(@I1) #And #Ceil(@I2) [simplification]
  rule #Ceil(@I1:Int modInt @I2:Int) => {(@I2 =/=Int 0) #Equals true} #And #Ceil(@I1) #And #Ceil(@I2) [simplification]
  rule #Ceil(@I1:Int >>Int  @I2:Int) => {(@I2 >=Int 0)  #Equals true} #And #Ceil(@I1) #And #Ceil(@I2) [simplification]
  rule #Ceil(@I1:Int <<Int  @I2:Int) => {(@I2 >=Int 0)  #Equals true} #And #Ceil(@I1) #And #Ceil(@I2) [simplification]
endmodule

module INT-KORE [symbolic]
  imports private K-EQUAL
  imports private BOOL
  imports INT-COMMON

  rule I1:Int ==K I2:Int => I1 ==Int I2 [simplification]
  rule {K1 ==Int K2 #Equals true} => {K1 #Equals K2} [simplification]
  rule {true #Equals K1 ==Int K2} => {K1 #Equals K2} [simplification]
  rule {K1 ==Int K2 #Equals false} => #Not({K1 #Equals K2}) [simplification]
  rule {false #Equals K1 ==Int K2} => #Not({K1 #Equals K2}) [simplification]
  rule {K1 =/=Int K2 #Equals true} => #Not({K1 #Equals K2}) [simplification]
  rule {true #Equals K1 =/=Int K2} => #Not({K1 #Equals K2}) [simplification]
  rule {K1 =/=Int K2 #Equals false} => {K1 #Equals K2} [simplification]
  rule {false #Equals K1 =/=Int K2} => {K1 #Equals K2} [simplification]

  // Arithmetic Normalization
  rule I +Int B => B +Int I          [concrete(I), symbolic(B), simplification(51)]
  rule A -Int I => A +Int (0 -Int I) [concrete(I), symbolic(A), simplification(51)]

  rule (A +Int I2) +Int I3 => A +Int (I2 +Int I3) [concrete(I2, I3), symbolic(A), simplification]
  rule I1 +Int (B +Int I3) => B +Int (I1 +Int I3) [concrete(I1, I3), symbolic(B), simplification]
  rule I1 -Int (B +Int I3) => (I1 -Int I3) -Int B [concrete(I1, I3), symbolic(B), simplification]
  rule I1 +Int (I2 +Int C) => (I1 +Int I2) +Int C [concrete(I1, I2), symbolic(C), simplification]
  rule I1 +Int (I2 -Int C) => (I1 +Int I2) -Int C [concrete(I1, I2), symbolic(C), simplification]
  rule (I1 -Int B) +Int I3 => (I1 +Int I3) -Int B [concrete(I1, I3), symbolic(B), simplification]
  rule I1 -Int (I2 +Int C) => (I1 -Int I2) -Int C [concrete(I1, I2), symbolic(C), simplification]
  rule I1 -Int (I2 -Int C) => (I1 -Int I2) +Int C [concrete(I1, I2), symbolic(C), simplification]
  rule (C -Int I2) -Int I3 => C -Int (I2 +Int I3) [concrete(I2, I3), symbolic(C), simplification]

  rule I1 &Int (I2 &Int C) => (I1 &Int I2) &Int C [concrete(I1, I2), symbolic(C), simplification]
endmodule

module INT
  imports INT-COMMON
  imports INT-SYMBOLIC
  imports INT-KORE
  imports private K-EQUAL
  imports private BOOL

  rule bitRangeInt(I::Int, IDX::Int, LEN::Int) => (I >>Int IDX) modInt (1 <<Int LEN)

  rule signExtendBitRangeInt(I::Int, IDX::Int, LEN::Int) => (bitRangeInt(I, IDX, LEN) +Int (1 <<Int (LEN -Int 1))) modInt (1 <<Int LEN) -Int (1 <<Int (LEN -Int 1))

  rule I1:Int divInt I2:Int => (I1 -Int (I1 modInt I2)) /Int I2
  requires I2 =/=Int 0
  rule
    I1:Int modInt I2:Int
  =>
    ((I1 %Int absInt(I2)) +Int absInt(I2)) %Int absInt(I2)
  requires I2 =/=Int 0    [concrete, simplification]

  rule minInt(I1:Int, I2:Int) => I1 requires I1 <=Int I2
  rule minInt(I1:Int, I2:Int) => I2 requires I1 >=Int I2

  rule I1:Int =/=Int I2:Int => notBool (I1 ==Int I2)
  rule (I1:Int dividesInt I2:Int) => (I2 %Int I1) ==Int 0

  syntax Int ::= freshInt(Int)    [freshGenerator, function, total, private]
  rule freshInt(I:Int) => I
endmodule

IEEE 754 Floating-point Numbers

Provided here is the syntax of an implementation of arbitrary-precision
floating-point arithmetic in K based on a generalization of the IEEE 754
standard. This type is hooked to an implementation of floats provided by the
backend.

The syntax of ordinary floating-point values in K consists of an optional sign
(+ or -) followed by an optional integer part, followed by a decimal point,
followed by an optional fractional part. Either the integer part or the
fractional part must be specified. The mantissa is followed by an optional
exponent part, which consists of an e or E, an optional sign (+ or -),
and an integer. The expoennt is followed by an optional suffix, which can be
either f, F, d, D, or pNxM where N and M are positive integers.
p and x can be either upper or lowercase.

The value of a floating-point literal is computed as follows: First the
mantissa is read as a rational number. Then it is multiplied by 10 to the
power of the exponent, which is interpreted as an integer, and defaults to
zero if it is not present. Finally, it is rounded to the nearest possible
value in a floating-point type represented like an IEEE754 floating-point type,
with the number of bits of precision and exponent specified by the suffix.
A suffix of f or f represents the IEEE binary32 format. A suffix of d
or D, or no suffix, represents the IEEE binary64 format. A suffix of
pNxM (either upper or lowercase) specifies exactly N bits of precision and
M bits of exponent. The number of bits of precision is assumed to include
any optional 1 that precedes the IEEE 754 mantissa. In other words, p24x8
is equal to the IEEE binary32 format, and p53x11 is equal to the IEEE
binary64 format.

module FLOAT-SYNTAX
  syntax Float [hook(FLOAT.Float)]
  syntax Float ::= r"([\\+-]?[0-9]+(\\.[0-9]*)?|\\.[0-9]+)([eE][\\+-]?[0-9]+)?([fFdD]|([pP][0-9]+[xX][0-9]+))?" [token, prec(1)]
  syntax Float ::= r"[\\+-]?Infinity([fFdD]|([pP][0-9]+[xX][0-9]+))?" [token, prec(3)]
  syntax Float ::= r"NaN([fFdD]|([pP][0-9]+[xX][0-9]+))?" [token, prec(3)]
endmodule

module FLOAT
  imports FLOAT-SYNTAX
  imports private BOOL
  imports private INT-SYNTAX

Float precision

You can retrieve the number of bits of precision in a Float.

  syntax Int ::= precisionFloat(Float) [function, total, hook(FLOAT.precision)]

Float exponent bits

You can retrieve the number of bits of exponent range in a Float.

  syntax Int ::= exponentBitsFloat(Float) [function, total, hook(FLOAT.exponentBits)]

Float exponent

You can retrieve the value of the exponent bits of a Float as an integer.

  syntax Int ::= exponentFloat(Float) [function, total, hook(FLOAT.exponent)]

Float sign

You can retrieve the value of the sign bit of a Float as a boolean. True
means the sign bit is set.

  syntax Bool ::= signFloat(Float)      [function, total, hook(FLOAT.sign)]

Float special values

You can check whether a Float value is infinite or Not-a-Number.

  syntax Bool ::= isNaN(Float)          [function, total, smt-hook(fp.isNaN), hook(FLOAT.isNaN)]
                | isInfinite(Float)     [function, total]

Float arithmetic

You can:

  • Compute the unary negation --Float of a float. --Float X is distinct
    from 0.0 -Float X. For example, 0.0 -Float 0.0 is positive zero.
    --Float 0.0 is negative zero.
  • Compute the exponentation ^Float of two floats.
  • Compute the product *Float, quotient /Float, or remainder %Float of two
    floats. The remainder is computed based on rounding the quotient of the two
    floats to the nearest integer.
  • Compute the sum +Float or difference -Float of two floats.
  syntax Float ::= "--Float" Float             [function, total, smt-hook(fp.neg), hook(FLOAT.neg)]
                 > Float "^Float" Float        [function, left, hook(FLOAT.pow)]
                 > left:
                   Float "*Float" Float        [function, left, smt-hook((fp.mul roundNearestTiesToEven #1 #2)), hook(FLOAT.mul)]
                 | Float "/Float" Float        [function, left, smt-hook((fp.div roundNearestTiesToEven #1 #2)), hook(FLOAT.div)]
                 | Float "%Float" Float        [function, left, smt-hook((fp.rem roundNearestTiesToEven #1 #2)), hook(FLOAT.rem)]
                 > left:
                   Float "+Float" Float        [function, left, smt-hook((fp.add roundNearestTiesToEven #1 #2)), hook(FLOAT.add)]
                 | Float "-Float" Float        [function, left, smt-hook((fp.sub roundNearestTiesToEven #1 #2)), hook(FLOAT.sub)]

Floating-point mathematics

You can:

  • Compute the Nth integer root rootFloat of a float.
  • Compute the absolute value absFloat of a float.
  • Round a floating-point number to a specified precision and exponent
    range (roundFloat). The resulting Float will yield the specified values
    when calling precisionFloat and exponentBitsFloat and when performing
    further computation.
  • Round a float to the next lowest floating-point value which is an integer
    (floorFloat).
  • Round a float to the next highest floating-point value which is an integer
    (ceilFloat).
  • Round a float to the next closest floating-point value which is an integer, in
    the direction of zero (truncFloat).
  • Compute the natural exponential expFloat of a float (i.e. e^x).
  • Compute the natural logarithm logFloat of a float.
  • Compute the sine sinFloat of a float.
  • Compute the cosine cosFloat of a float.
  • Compute the tangent tanFlooat of a float.
  • Compute the arcsine asinFloat of a float.
  • Compute the arccosine acosFloat of a float.
  • Compute the arctangent atanFloat of a float.
  • Compute the arctangent atan2Float of two floats.
  • Compute the maximum maxFloat of two floats.
  • Compute the minimum minFloat of two floats.
  • Compute the square root sqrtFloat of a float.
  • Compute the largest finite value expressible in a specified precision and
    exponent range (maxValueFloat).
  • Compute the smallest positive finite value expressible in a specified
    precision and exponent range (minValueFloat).
  syntax Float ::= rootFloat(Float, Int)        [function, hook(FLOAT.root)]
                 | absFloat(Float)              [function, total, smt-hook(fp.abs), hook(FLOAT.abs)]
                 | roundFloat(Float, precision: Int, exponentBits: Int)  [function, hook(FLOAT.round)]
                 | floorFloat(Float)            [function, total, hook(FLOAT.floor)]
                 | ceilFloat(Float)             [function, total, hook(FLOAT.ceil)]
                 | truncFloat(Float)            [function, total, hook(FLOAT.trunc)]
                 | expFloat(Float)              [function, total, hook(FLOAT.exp)]
                 | logFloat(Float)              [function, hook(FLOAT.log)]
                 | sinFloat(Float)              [function, total, hook(FLOAT.sin)]
                 | cosFloat(Float)              [function, total, hook(FLOAT.cos)]
                 | tanFloat(Float)              [function, hook(FLOAT.tan)]
                 | asinFloat(Float)             [function, hook(FLOAT.asin)]
                 | acosFloat(Float)             [function, hook(FLOAT.acos)]
                 | atanFloat(Float)             [function, total, hook(FLOAT.atan)]
                 | atan2Float(Float, Float)     [function, hook(FLOAT.atan2)]
                 | maxFloat(Float, Float)       [function, smt-hook(fp.max), hook(FLOAT.max)]
                 | minFloat(Float, Float)       [function, smt-hook(fp.min), hook(FLOAT.min)]
                 | sqrtFloat(Float)             [function]
                 | maxValueFloat(precision: Int, exponentBits: Int)      [function, hook(FLOAT.maxValue)]
                 | minValueFloat(precision: Int, exponentBits: Int)      [function, hook(FLOAT.minValue)]

Floating-point comparisons

Compute whether a float is less than or equasl to, less than, greater than or
equal to, greater than, equal, or unequal to another float. Note that
X ==Float Y and X ==K Y might yield different values. The latter should be
used in cases where you want to compare whether two values of sort Float
contain the same term. The former should be used when you want to implement
the == operator of a programming language. In particular, NaN =/=Float NaN
is true, because NaN compares unequal to all values, including itself, in
IEEE 754 arithmetic. 0.0 ==Float -0.0 is also true.

  syntax Bool ::= Float "<=Float" Float       [function, smt-hook(fp.leq), hook(FLOAT.le)]
                | Float "<Float" Float        [function, smt-hook(fp.lt), hook(FLOAT.lt)]
                | Float ">=Float" Float       [function, smt-hook(fp.geq), hook(FLOAT.ge)]
                | Float ">Float" Float        [function, smt-hook(fg.gt), hook(FLOAT.gt)]
                | Float "==Float" Float       [function, comm, smt-hook(fp.eq), hook(FLOAT.eq), klabel(_==Float_)]
                | Float "=/=Float" Float      [function, comm, smt-hook((not (fp.eq #1 #2)))]

  rule F1:Float =/=Float F2:Float => notBool (F1 ==Float F2)

Conversion between integer and float

You can convert an integer to a floating-point number with the specified
precision and exponent range. You can also convert a floating-point number
to the nearest integer. This operation rounds to the nearest integer, but it
also avoids the double-rounding that is present in ceilFloat and floorFloat
if the nearest integer is not representable in the specified floating-point
type.

  syntax Float ::= Int2Float(Int, precision: Int, exponentBits: Int)    [function, hook(FLOAT.int2float)]
  syntax Int ::= Float2Int(Float)    [function, total, hook(FLOAT.float2int)]

Implementation of Floats

The remainder of this section consists of an implementation in K of some of the
operators above.

  rule sqrtFloat(F:Float) => rootFloat(F, 2)

  rule isInfinite(F:Float) => F >Float maxValueFloat(precisionFloat(F), exponentBitsFloat(F)) orBool F <Float --Float maxValueFloat(precisionFloat(F), exponentBitsFloat(F))

endmodule

Strings

Provided here is the syntax of an implementation of Unicode strings in K. This
type is hooked to an implementation of strings provided by the backend. The
implementation is currently incomplete and does not fully support encodings
and code points beyond the initial 256 code points of the Basic Latin and
Latin-1 Supplement blocks. In the future, there may be breaking changes to
the semantics of this module in order to support this functionality.

The syntax of strings in K is delineated by double quotes. Inside the double
quotes, any character can appear verbatim except double quotes, backslash,
newline, and carriage return. K also supports the following escape sequences:

  • " - the " character
  • \ - the \ character
  • \n - newline character
  • \r - carriage return character
  • \t - tab character
  • \f - form feed character
  • \xFF - \x followed by two hexadecimal characters indicates a code point
    between 0x00 and 0xff
  • \uFFFF - \u followed by four hexadecimal characters indicates a code point
    between 0x0000 and 0xffff
  • \UFFFFFFFF - \U followed by eight hexadecimal characters indicates a code
    point between 0x000000 and 0x10ffff
module STRING-SYNTAX
  syntax String [hook(STRING.String)]
  syntax String ::= r"[\\\"](([^\\\"\\n\\r\\\\])|([\\\\][nrtf\\\"\\\\])|([\\\\][x][0-9a-fA-F]{2})|([\\\\][u][0-9a-fA-F]{4})|([\\\\][U][0-9a-fA-F]{8}))*[\\\"]"      [token]
endmodule

module STRING-COMMON
  imports STRING-SYNTAX
  imports private INT
  imports private FLOAT-SYNTAX
  imports private K-EQUAL
  imports private BOOL

String concatenation

You can concatenate two strings in O(N) time. For successive concatenation
operations, it may be better to use the STRING-BUFFER module.

  syntax String ::= String "+String" String    [function, total, left, hook(STRING.concat)]

String length

You can get the length of a string in O(1) time.

  syntax Int ::= lengthString ( String ) [function, total, hook(STRING.length)]

Character and integer conversion

You can convert between a character (as represented by a string containing
a single code point) and an integer in O(1) time.

  syntax String ::= chrChar ( Int )      [function, hook(STRING.chr)]
  syntax Int ::= ordChar ( String )      [function, hook(STRING.ord)]

String substring

You can compute a substring of a string in O(N) time (where N is the
length of the substring). There are two important facts to note:

  1. the range generated includes the character at startIndex but excludes the
    character at endIndex, i.e., the range is [startIndex..endIndex).
  2. this function is only defined on valid indices (i.e., it is defined when
    startIndex < endIndex and endIndex is less than or equal to the string
    length).
  syntax String ::= substrString ( String , startIndex: Int , endIndex: Int ) [function, total, hook(STRING.substr)]

You can find the first (respectively, last) occurrence of a substring, starting
at a certain index, in another string in O(N*M) time.
Returns -1 if the substring is not found.

  syntax Int ::= findString ( haystack: String , needle: String , index: Int )   [function, hook(STRING.find)]
  syntax Int ::= rfindString ( haystack: String , needle: String , index: Int )  [function, hook(STRING.rfind)]

You can find the first (respectively, last) occurrence of one of the characters
of the search string, starting at a certain index, in another string in
O(N*M) time.

  syntax Int ::= findChar ( haystack: String , needles: String , index: Int )     [function, hook(STRING.findChar)]
  syntax Int ::= rfindChar ( haystack: String , needles: String , index: Int )    [function, hook(STRING.rfindChar)]

String and Bool conversion

  syntax String ::= Bool2String(Bool) [function, total]
  rule Bool2String(true)  => "true"
  rule Bool2String(false) => "false"
  syntax Bool ::= String2Bool(String) [function]
  rule String2Bool("true")  => true
  rule String2Bool("false") => false

String and float conversion

You can convert between a String and a Float. The String will be
represented in the syntax of the Float sort (see the section on the FLOAT
module above for details of that syntax). Which particular string is returned
by Float2String is determined by the backend, but the same Float is
guaranteed to return the same String, and converting that String back to a
Float is guaranteed to return the original Float.

You can also convert a Float to a string in a particular syntax using the
variant of Float2String with a format. In this case, the resulting string
is one which results directly from passing that format to mpfr_printf. This
functionality may not be supported on backends that do not use Gnu MPFR to
implement floating-point numbers.

  syntax String ::= Float2String ( Float )              [function, total, hook(STRING.float2string)]
  syntax String ::= Float2String ( Float , format: String )     [function, klabel(FloatFormat), hook(STRING.floatFormat)]
  syntax Float  ::= String2Float ( String )             [function, hook(STRING.string2float)]

String and integer conversions

You can convert between a String and an Int. The String will be represented
in the syntax of the INT module (i.e., a nonempty sequence of digits
optionally prefixed by a sign). When converting from an Int to a String,
the sign will not be present unless the integer is negative.

You can also convert between a String and an Int in a particular radix.
This radix can be anywhere between 2 and 36. For a radix 2 <= N <= 10, the
digits 0 to N-1 will be used. For a radix 11 <= N <= 36, the digits 0 to 9
and the first N-10 letters of the Latin alphabet will be used. Both uppercase
and lowercase letters are supported by String2Base. Whether the letters
returned by Base2String are upper or lowercase is determined by the backend,
but the backend will consistently choose one or the other.

  syntax Int    ::= String2Int   ( String )             [function, hook(STRING.string2int)]
  syntax String ::= Int2String   ( Int )                [function, total, hook(STRING.int2string)]
  syntax String ::= Base2String  ( Int , base: Int )          [function, hook(STRING.base2string)]
  syntax Int    ::= String2Base  ( String , base: Int )       [function, hook(STRING.string2base)]

String count and replace

You can replace one, some, or all occurrences of a string within another
string in O(N*M) time. The replaceAll, replace, and replaceFirst methods
are identical, except replaceFirst replaces exactly one ocurrence of the
string, the first occurrence. replace replaces the first times occurrences.
And replaceAll replaces every occurrence.

You can also count the number of times a string occurs within another string
using countAllOccurrences.

  syntax String ::= "replaceAll" "(" haystack: String "," needle: String "," replacement: String ")"      [function, total, hook(STRING.replaceAll)]
  syntax String ::= "replace" "(" haystack: String "," needle: String "," replacement: String "," times: Int ")" [function, hook(STRING.replace)]
  syntax String ::= "replaceFirst" "(" haystack: String "," needle: String "," replacement: String ")"    [function, total, hook(STRING.replaceFirst)]
  syntax Int ::= "countAllOccurrences" "(" haystack: String "," needle: String ")"            [function, total, hook(STRING.countAllOccurrences)]

String equality and lexicographic comparison

You can compare whether two strings are equal or unequal, or whether one string
is less than, less than or equal to, greater than, or greater than or equal to
another according to the natural lexicographic ordering of strings.

  syntax Bool ::= String "==String" String  [function, total, comm, hook(STRING.eq)]
                | String "=/=String" String [function, total, comm, hook(STRING.ne)]
  rule S1:String =/=String S2:String => notBool (S1 ==String S2)

  syntax Bool ::= String  "<String" String [function, total, hook(STRING.lt)]
                | String "<=String" String [function, total, hook(STRING.le)]
                | String  ">String" String [function, total, hook(STRING.gt)]
                | String ">=String" String [function, total, hook(STRING.ge)]

Implementation of Strings

What follows is a few String hooks which are deprecated and only are supported
on certain outdated backends of K, as well as an implementation of several
of the above operations in K.

  syntax String ::= categoryChar(String)       [function, hook(STRING.category)]
                  | directionalityChar(String) [function, hook(STRING.directionality)]

  syntax String ::= "newUUID" [function, hook(STRING.uuid), impure]

  rule S1:String <=String S2:String => notBool (S2 <String S1)
  rule S1:String >String S2:String => S2 <String S1
  rule S1:String >=String S2:String => notBool (S1 <String S2)

  rule findChar(S1:String, S2:String, I:Int) => #if findString(S1, substrString(S2, 0, 1), I) ==Int -1 #then findChar(S1, substrString(S2, 1, lengthString(S2)), I) #else #if findChar(S1, substrString(S2, 1, lengthString(S2)), I) ==Int -1 #then findString(S1, substrString(S2, 0, 1), I) #else minInt(findString(S1, substrString(S2, 0, 1), I), findChar(S1, substrString(S2, 1, lengthString(S2)), I)) #fi #fi requires S2 =/=String ""
  rule findChar(_, "", _) => -1
  rule rfindChar(S1:String, S2:String, I:Int) => maxInt(rfindString(S1, substrString(S2, 0, 1), I), rfindChar(S1, substrString(S2, 1, lengthString(S2)), I)) requires S2 =/=String ""
  rule rfindChar(_, "", _) => -1

  rule countAllOccurrences(Source:String, ToCount:String) => 0
            requires findString(Source, ToCount, 0) <Int 0
  rule countAllOccurrences(Source:String, ToCount:String) => 1 +Int countAllOccurrences(substrString(Source, findString(Source, ToCount, 0) +Int lengthString(ToCount), lengthString(Source)), ToCount)
            requires findString(Source, ToCount, 0) >=Int 0

  rule replaceFirst(Source:String, ToReplace:String, Replacement:String) => substrString(Source, 0, findString(Source, ToReplace, 0))
                +String Replacement +String substrString(Source, findString(Source, ToReplace, 0) +Int lengthString(ToReplace), lengthString(Source))
                requires findString(Source, ToReplace, 0) >=Int 0
  rule replaceFirst(Source:String, ToReplace:String, _:String) => Source
        requires findString(Source, ToReplace, 0) <Int 0


  // Note that the replace function is undefined when Count < 0. This allows different backends to
  // implement their own behavior without contradicting these semantics. For instance, a symbolic
  // backend can return #Bottom for that case, while a concrete backend can throw an exception.
  rule replace(Source:String, ToReplace:String, Replacement:String, Count:Int) =>
       substrString(Source, 0, findString(Source, ToReplace, 0)) +String Replacement +String
       replace(substrString(Source, findString(Source, ToReplace, 0) +Int lengthString(ToReplace), lengthString(Source)), ToReplace, Replacement, Count -Int 1)
        requires Count >Int 0 andBool findString(Source, ToReplace, 0) >=Int 0
  rule replace(Source:String, _, _, Count) => Source
        requires Count >=Int 0 [owise]
  rule replaceAll(Source:String, ToReplace:String, Replacement:String) => replace(Source, ToReplace, Replacement, countAllOccurrences(Source, ToReplace))

endmodule

module STRING-KORE [symbolic]
  imports private K-EQUAL
  imports STRING-COMMON

  rule S1:String ==K S2:String => S1 ==String S2 [simplification]

endmodule

module STRING
  imports STRING-COMMON
  imports STRING-KORE
endmodule

String Buffers

It is a well known fact that repeated string concatenations are quadratic
in performance whereas use of an efficient mutable representation of arrays
can yield linear performance. We thus provide such a sort, the StringBuffer
sort. Axiomatically, it is implemented below on symbolic backends using the
String module. However, on concrete backends it provides an efficient
implementation of string concatenation. There are three operations:

  • .StringBuffer creates a new StringBuffer with current content equal
    to the empty string.
  • +String takes a StringBuffer and a String and appends the String to
    the end of the StringBuffer
  • StringBuffer2String converts a StringBuffer to a String. This operation
    copies the string so that subsequent modifications to the StringBuffer
    will not change the value of the String returned by this function.
module STRING-BUFFER-IN-K [symbolic]
  imports private BASIC-K
  imports STRING

  syntax StringBuffer ::= ".StringBuffer" [function, total]
  syntax StringBuffer ::= StringBuffer "+String" String [function, total, avoid]
  syntax StringBuffer ::= String
  syntax String ::= StringBuffer2String ( StringBuffer ) [function, total]

  rule {SB:String +String S:String}::StringBuffer => (SB +String S)::String
  rule .StringBuffer => ""
  rule StringBuffer2String(S:String) => S
endmodule

module STRING-BUFFER-HOOKED [concrete]
  imports private BASIC-K
  imports STRING

  syntax StringBuffer [hook(BUFFER.StringBuffer)]
  syntax StringBuffer ::= ".StringBuffer" [function, total, hook(BUFFER.empty), impure]
  syntax StringBuffer ::= StringBuffer "+String" String [function, total, hook(BUFFER.concat), avoid]
  syntax String ::= StringBuffer2String ( StringBuffer ) [function, total, hook(BUFFER.toString)]
endmodule

module STRING-BUFFER
  imports STRING-BUFFER-HOOKED
  imports STRING-BUFFER-IN-K
endmodule

Byte Arrays

Provided here is the syntax of an implementation of fixed-width arrays of Bytes
in K. This type is hooked to an implementation of bytes provided by the backend.
On the LLVM backend, it is possible to opt in to a faster, mutable
representation (using the --llvm-mutable-bytes flag to kompile) where
multiple references can occur to the same Bytes object and when one is
modified, the others are also modified. Care should be taken when using this
feature, however, as it is possible to experience divergent behavior with
symbolic backends unless the Bytes type is used in a manner that preserves
consistency.

module BYTES-SYNTAX
  imports private STRING-SYNTAX

  syntax Bytes [hook(BYTES.Bytes)]
  syntax Bytes ::= r"b[\\\"](([\\x20\\x21\\x23-\\x5B\\x5D-\\x7E])|([\\\\][tnfr\\\"\\\\])|([\\\\][x][0-9a-fA-F]{2}))*[\\\"]"      [token]
endmodule
module BYTES-STRING-ENCODE [symbolic]
  imports BYTES-SYNTAX

Encoding/decoding between Bytes and String

You can encode/decode between Bytes and String using UTF-8, UTF-16LE, UTF-16BE, UTF-32LE, and UTF-32BE

    syntax String ::= decodeBytes ( encoding: String , contents: Bytes ) [function, hook(BYTES.decodeBytes)]
    syntax Bytes ::= encodeBytes ( encoding: String , contents: String ) [function, hook(BYTES.encodeBytes)]
endmodule
module BYTES-HOOKED
  imports STRING-SYNTAX
  imports BYTES-SYNTAX
  imports BYTES-STRING-ENCODE

Empty byte array

The byte array of length zero is represented by .Bytes.

  syntax Bytes ::= ".Bytes" [function, total, hook(BYTES.empty)]

Endianness

When converting to/from an integer, byte arrays can be treated as either little
endian (ie, least significant byte first) or big endian (ie, most significant
byte first).

  syntax Endianness ::= "LE" [klabel(littleEndianBytes), symbol]
                      | "BE" [klabel(bigEndianBytes), symbol]

Signedness

When converting to/from an integer, byte arrays can be treated as either signed
or unsigned.

  syntax Signedness ::= "Signed" [klabel(signedBytes), symbol]
                      | "Unsigned" [klabel(unsignedBytes), symbol]

Integer and Bytes conversion

You can convert from a Bytes to an Int. In order to do this, the endianness
and signedness of the Bytes must be provided. The resulting integer is
created by means of interpreting the Bytes as either a twos-complement
representation, or an unsigned representation, of an integer, in the specified
byte order.

You can also convert from an Int to a Bytes. This comes in two variants.
In the first, the length of the resulting Bytes in bytes is explicitly
specified. If the length is greater than the highest set bit in the magnitude
of the integer, the result is padded with 0 bits if the number is positive
and 1 bits if the number is negative. If the length is less than the highest
bit set in the magnitude of the integer, the most-significant bits of the
integer will be truncated. The endianness of the resulting Bytes object
is as specified.

In the second variant, both endianness and signedness are specified, and
the resulting Bytes object will be the smallest number of bytes necessary
for the resulting Bytes object to be convertible back to the original integer
via Bytes2Int. In other words, if the highest bit set in the magnitude of the
integer is N, then the byte array will be at least N+1 bits long, rounded up
to the nearest byte.

  syntax Int ::= Bytes2Int(Bytes, Endianness, Signedness) [function, total, hook(BYTES.bytes2int)]
  syntax Bytes ::= Int2Bytes(length: Int, Int, Endianness) [function, total, hook(BYTES.int2bytes)]
                 | Int2Bytes(Int, Endianness, Signedness) [function, total, klabel(Int2BytesNoLen)]

String and Bytes conversion

You can convert between a Bytes and a String in O(N) time. The resulting
value is a copy of the original and will not be affected by subsequent
mutations of the input or output value.

  syntax String ::= Bytes2String(Bytes) [function, total, hook(BYTES.bytes2string)]
  syntax Bytes ::= String2Bytes(String) [function, total, hook(BYTES.string2bytes)]

Bytes update

You can set the value of a particular byte in a Bytes object in O(1) time.
The result is #False if value is not in the range [0..255] or if index
is not a valid index (ie, less than zero or greater than or equal to the length
of the Bytes term).

  syntax Bytes ::= Bytes "[" index: Int "<-" value: Int "]" [function, hook(BYTES.update)]

Bytes lookup

You can get the value of a particular byte in a Bytes object in O(1) time.
The result is #False if index is not a valid index (see above).

  syntax Int ::= Bytes "[" Int "]" [function, hook(BYTES.get)]

Bytes substring

You can get a new Bytes object containing a range of bytes from the input
Bytes in O(N) time (where N is the length of the substring). The range
of bytes included is [startIndex..endIndex). The resulting Bytes is
a copy and mutations to it do not affect mutations to the original Bytes.
The result is #False if startIndex or endIndex are not valid.

  syntax Bytes ::= substrBytes(Bytes, startIndex: Int, endIndex: Int) [function, hook(BYTES.substr)]

Multiple bytes update

You can modify a Bytes to return a Bytes which is equal to dest except the
N elements starting at index are replaced with the contents of src in O(N)
time. If --llvm-mutable-bytes is active, this will not create a new Bytes
object and will instead modify the original on concrete backends. The result is
#False if index + N is not a valid index.

  syntax Bytes ::= replaceAtBytes(dest: Bytes, index: Int, src: Bytes) [function, hook(BYTES.replaceAt)]

Multiple bytes update

You can modify a Bytes to return a Bytes which is equal to dest except the
count bytes starting at index are replaced with count bytes of value
Int2Bytes(1, v, LE/BE) in O(count) time. This does not create a new Bytes
object and will instead modify the original if --llvm-mutable-bytes is active.
This will throw an exception if index + count is not a valid index. The
acceptable range of values for v is -128 to 127. This will throw an exception
if v is outside of this range. This is implemented only for the LLVM backend.

  syntax Bytes ::= memsetBytes(dest: Bytes, index: Int, count: Int, v: Int) [function, hook(BYTES.memset)]

Bytes padding

You can create a new Bytes object which is at least length bytes long by
taking the input sequence and padding it on the right (respectively, on the
left) with the specified value. If --llvm-mutable-bytes is active, this does
not create a new Bytes object if the input is already at least length bytes
long, and will instead return the input unchanged. The result is #False if
value is not in the range [0..255], or if the length is negative.

  syntax Bytes ::= padRightBytes(Bytes, length: Int, value: Int) [function, hook(BYTES.padRight)]
                 | padLeftBytes(Bytes, length: Int, value: Int) [function, hook(BYTES.padLeft)]

Bytes reverse

You can reverse a Bytes object in O(N) time. If --llvm-mutable-bytes is
active, this will not create a new Bytes object and will instead modify the
original.

  syntax Bytes ::= reverseBytes(Bytes) [function, total, hook(BYTES.reverse)]

Bytes length

You can get the length of a Bytes term in O(1) time.

  syntax Int ::= lengthBytes(Bytes) [function, total, hook(BYTES.length), smtlib(lengthBytes)]

Bytes concatenation

You can create a new Bytes object by concatenating two Bytes objects
together in O(N) time.

  syntax Bytes ::= Bytes "+Bytes" Bytes [function, total, hook(BYTES.concat), right]

endmodule

Implementation of Bytes

The remainder of this module consists of an implementation of some of the
operators listed above in K.

module BYTES-CONCRETE [concrete]
  imports BYTES-HOOKED
endmodule

module BYTES-KORE
  imports BYTES-HOOKED
  imports BYTES-SYMBOLIC-CEIL
endmodule

module BYTES-SYMBOLIC-CEIL [symbolic]
  imports BYTES-HOOKED
  imports private INT
  imports private BOOL

  rule #Ceil(padRightBytes(_, LEN, VAL)) => {(0 <=Int LEN andBool 0 <=Int VAL andBool VAL <Int 256) #Equals true} [simplification]
  rule #Ceil(padLeftBytes(_, LEN, VAL))  => {(0 <=Int LEN andBool 0 <=Int VAL andBool VAL <Int 256) #Equals true} [simplification]
endmodule

module BYTES
  imports BYTES-CONCRETE
  imports BYTES-KORE
  imports private INT

  rule Int2Bytes(I::Int, E::Endianness, Unsigned) => Int2Bytes((log2Int(I) +Int 8) /Int 8, I, E)
    requires I >Int 0
  rule Int2Bytes(0, _::Endianness, _) => .Bytes
  rule Int2Bytes(I::Int, E::Endianness, Signed) => Int2Bytes((log2Int(I) +Int 9) /Int 8, I, E)
    requires I >Int 0
  rule Int2Bytes(I::Int, E::Endianness, Signed) => Int2Bytes((log2Int(~Int I) +Int 9) /Int 8, I, E)
    requires I <Int -1
  rule Int2Bytes(-1, E::Endianness, Signed) => Int2Bytes(1, -1, E)
endmodule

Program identifiers

Provided here is an implementation for program identifiers in K. Developers
of semantics for a particular language may wish to use their own implementation
instead of the one provided here if their syntax differs from the syntax
defined below. However, this is provided for convenience for developers who
do not care about the lexical syntax of identifiers.

Provided are the following pieces of functionality:

  • Id2String - Convert an Id to a String containing its name
  • String2Id - Convert a String to an Id with the specified name
  • !X:Id - You can get a fresh identifier distinct from any previous identifier
    generated by this syntax.
module ID-SYNTAX-PROGRAM-PARSING
  imports BUILTIN-ID-TOKENS
  syntax Id ::= r"(?<![A-Za-z0-9\\_])[A-Za-z\\_][A-Za-z0-9\\_]*"     [prec(1), token]
              | #LowerId                                             [token]
              | #UpperId                                             [token]
endmodule

module ID-SYNTAX
  syntax Id [token]
endmodule

module ID-COMMON
  imports ID-SYNTAX
  imports private STRING

  syntax String ::= Id2String ( Id )    [function, total, hook(STRING.token2string)]
  syntax Id ::= String2Id (String) [function, total, hook(STRING.string2token)]
  syntax Id ::= freshId(Int)    [freshGenerator, function, total, private]

  rule freshId(I:Int) => String2Id("_" +String Int2String(I))
endmodule

module ID
  imports ID-COMMON
endmodule

Equality and conditionals

Provided here are implementations of two important primitives in K:

  • ==K - the equality between two terms. Returns true if they are equal
    and false if they are not equal.
  • #if #then #else #fi - polymorphic conditional function. If the first
    argument evaluates to true, the second argument is returned. Otherwise,
    the third argument is returned. Note that this does not short-circuit on
    symbolic backends.
module K-EQUAL-SYNTAX
  imports private BOOL
  imports private BASIC-K

  syntax Bool ::= left:
                  K "==K" K           [function, total, comm, smt-hook(=), hook(KEQUAL.eq), klabel(_==K_), symbol, group(equalEqualK)]
                | K "=/=K" K          [function, total, comm, smt-hook(distinct), hook(KEQUAL.ne), klabel(_=/=K_), symbol, group(notEqualEqualK)]

  syntax priority equalEqualK notEqualEqualK > boolOperation mlOp

  syntax {Sort} Sort ::= "#if" Bool "#then" Sort "#else" Sort "#fi" [function, total, symbol(ite), smt-hook(ite), hook(KEQUAL.ite)]

endmodule

module K-EQUAL-KORE [symbolic]
  imports private BOOL
  imports K-EQUAL-SYNTAX

  rule K1:Bool ==K K2:Bool => K1 ==Bool K2 [simplification]
  rule {K1 ==K K2 #Equals true} => {K1 #Equals K2} [simplification]
  rule {true #Equals K1 ==K K2} => {K1 #Equals K2} [simplification]
  rule {K1 ==K K2 #Equals false} => #Not({K1 #Equals K2}) [simplification]
  rule {false #Equals K1 ==K K2} => #Not({K1 #Equals K2}) [simplification]
  rule {K1 =/=K K2 #Equals true} => #Not({K1 #Equals K2}) [simplification]
  rule {true #Equals K1 =/=K K2} => #Not({K1 #Equals K2}) [simplification]
  rule {K1 =/=K K2 #Equals false} => {K1 #Equals K2} [simplification]
  rule {false #Equals K1 =/=K K2} => {K1 #Equals K2} [simplification]

endmodule

module K-EQUAL
  imports private BOOL
  imports K-EQUAL-SYNTAX
  imports K-EQUAL-KORE

  rule K1:K =/=K K2:K => notBool (K1 ==K K2)

  rule #if C:Bool #then B1::K #else _ #fi => B1 requires C
  rule #if C:Bool #then _ #else B2::K #fi => B2 requires notBool C

endmodule

Meta operations

Provided below are a few miscellaneous, mostly deprecated functions in K.
It is not recommended to use any of them directly as they are largely
unsupported in modern K. There are a few exceptions:

  • #getenv - Returns the value of an environment variable
  • #kompiledDirectory - Returns the path to the current compiled K definition
    directory.
  • #unparseKORE - Takes a K term and converts it to a string.
module K-REFLECTION
  imports BASIC-K
  imports STRING

  syntax K ::= "#configuration" [function, impure, hook(KREFLECTION.configuration)]
  syntax String ::= #sort(K) [function, hook(KREFLECTION.sort)]
  syntax KItem ::= #fresh(String)   [function, hook(KREFLECTION.fresh), impure]
  syntax KItem ::= getKLabel(K)  [function, hook(KREFLECTION.getKLabel)]

  syntax K ::= #getenv(String) [function, impure, hook(KREFLECTION.getenv)]

  syntax String ::= #kompiledDirectory() [function, hook(KREFLECTION.kompiledDir)]

  // meaningful only for the purposes of compilation to a binary, otherwise
  // undefined
  syntax List ::= #argv() [function, hook(KREFLECTION.argv)]

  syntax {Sort} String ::= #unparseKORE(Sort) [function, hook(KREFLECTION.printKORE)]
  syntax IOError ::= "#noParse" "(" String ")" [klabel(#noParse), symbol]

endmodule

I/O in K

Concrete execution in K supports I/O operations. This functionality is not
supported during symbolic execution, because symbolic execution must exist
completely free of side-effects, and I/O is an irreducible type of side effect.
However, it is useful in many cases when defining concrete execution to be able
to make reference to I/O operations.

The design of these I/O operations is based on the POSIX standard, for the most
part. For example, the #read K function maps to the read POSIX function. We
do not at this time have a higher-level API for I/O, but this may be
implemented at some point in the future.

I/O operations generally return either their result, or an IOError term
corresponding to the errno returned by the underlying system call.

module K-IO
  imports private LIST
  imports private STRING
  imports private INT

I/O errors

Aside from EOF, which is returned by #getc if the file is at end-of-file, all
of the below I/O errors correspond to possible values for errno after calling
a library function. If the errno returned is not one of the below errnos
known to K, #unknownIOError is returned along with the integer errno value.

  syntax IOError ::= "#EOF" [klabel(#EOF), symbol] 
                   | #unknownIOError(errno: Int) [klabel(#unknownIOError), symbol]
                   | "#E2BIG" [klabel(#E2BIG), symbol]
                   | "#EACCES" [klabel(#EACCES), symbol]
                   | "#EAGAIN" [klabel(#EAGAIN), symbol]
                   | "#EBADF" [klabel(#EBADF), symbol]
                   | "#EBUSY" [klabel(#EBUSY), symbol]
                   | "#ECHILD" [klabel(#ECHILD), symbol]
                   | "#EDEADLK" [klabel(#EDEADLK), symbol]
                   | "#EDOM" [klabel(#EDOM), symbol]
                   | "#EEXIST" [klabel(#EEXIST), symbol]
                   | "#EFAULT" [klabel(#EFAULT), symbol]
                   | "#EFBIG" [klabel(#EFBIG), symbol]
                   | "#EINTR" [klabel(#EINTR), symbol]
                   | "#EINVAL" [klabel(#EINVAL), symbol]
                   | "#EIO" [klabel(#EIO), symbol]
                   | "#EISDIR" [klabel(#EISDIR), symbol]
                   | "#EMFILE" [klabel(#EMFILE), symbol]
                   | "#EMLINK" [klabel(#EMLINK), symbol]
                   | "#ENAMETOOLONG" [klabel(#ENAMETOOLONG), symbol]
                   | "#ENFILE" [klabel(#ENFILE), symbol]
                   | "#ENODEV" [klabel(#ENODEV), symbol]
                   | "#ENOENT" [klabel(#ENOENT), symbol]
                   | "#ENOEXEC" [klabel(#ENOEXEC), symbol]
                   | "#ENOLCK" [klabel(#ENOLCK), symbol]
                   | "#ENOMEM" [klabel(#ENOMEM), symbol]
                   | "#ENOSPC" [klabel(#ENOSPC), symbol]
                   | "#ENOSYS" [klabel(#ENOSYS), symbol]
                   | "#ENOTDIR" [klabel(#ENOTDIR), symbol]
                   | "#ENOTEMPTY" [klabel(#ENOTEMPTY), symbol]
                   | "#ENOTTY" [klabel(#ENOTTY), symbol]
                   | "#ENXIO" [klabel(#ENXIO), symbol]
                   | "#EPERM" [klabel(#EPERM), symbol]
                   | "#EPIPE" [klabel(#EPIPE), symbol]
                   | "#ERANGE" [klabel(#ERANGE), symbol]
                   | "#EROFS" [klabel(#EROFS), symbol]
                   | "#ESPIPE" [klabel(#ESPIPE), symbol]
                   | "#ESRCH" [klabel(#ESRCH), symbol]
                   | "#EXDEV" [klabel(#EXDEV), symbol]
                   | "#EWOULDBLOCK" [klabel(#EWOULDBLOCK), symbol]
                   | "#EINPROGRESS" [klabel(#EINPROGRESS), symbol]
                   | "#EALREADY" [klabel(#EALREADY), symbol]
                   | "#ENOTSOCK" [klabel(#ENOTSOCK), symbol]
                   | "#EDESTADDRREQ" [klabel(#EDESTADDRREQ), symbol]
                   | "#EMSGSIZE" [klabel(#EMSGSIZE), symbol]
                   | "#EPROTOTYPE" [klabel(#EPROTOTYPE), symbol]
                   | "#ENOPROTOOPT" [klabel(#ENOPROTOOPT), symbol]
                   | "#EPROTONOSUPPORT" [klabel(#EPROTONOSUPPORT), symbol]
                   | "#ESOCKTNOSUPPORT" [klabel(#ESOCKTNOSUPPORT), symbol]
                   | "#EOPNOTSUPP" [klabel(#EOPNOTSUPP), symbol]
                   | "#EPFNOSUPPORT" [klabel(#EPFNOSUPPORT), symbol]
                   | "#EAFNOSUPPORT" [klabel(#EAFNOSUPPORT), symbol]
                   | "#EADDRINUSE" [klabel(#EADDRINUSE), symbol]
                   | "#EADDRNOTAVAIL" [klabel(#EADDRNOTAVAIL), symbol]
                   | "#ENETDOWN" [klabel(#ENETDOWN), symbol]
                   | "#ENETUNREACH" [klabel(#ENETUNREACH), symbol]
                   | "#ENETRESET" [klabel(#ENETRESET), symbol]
                   | "#ECONNABORTED" [klabel(#ECONNABORTED), symbol]
                   | "#ECONNRESET" [klabel(#ECONNRESET), symbol]
                   | "#ENOBUFS" [klabel(#ENOBUFS), symbol]
                   | "#EISCONN" [klabel(#EISCONN), symbol]
                   | "#ENOTCONN" [klabel(#ENOTCONN), symbol]
                   | "#ESHUTDOWN" [klabel(#ESHUTDOWN), symbol]
                   | "#ETOOMANYREFS" [klabel(#ETOOMANYREFS), symbol]
                   | "#ETIMEDOUT" [klabel(#ETIMEDOUT), symbol]
                   | "#ECONNREFUSED" [klabel(#ECONNREFUSED), symbol]
                   | "#EHOSTDOWN" [klabel(#EHOSTDOWN), symbol]
                   | "#EHOSTUNREACH" [klabel(#EHOSTUNREACH), symbol]
                   | "#ELOOP" [klabel(#ELOOP), symbol]
                   | "#EOVERFLOW" [klabel(#EOVERFLOW), symbol]

I/O result sorts

Here we see sorts defined to contain either an Int or an IOError, or
either a String or an IOError. These sorts are used to implement the
return sort of functions that may succeed, in which case they return a value,
or may fail, in which case their return value indicates an error and the
error indicated is returned via errno.

  syntax IOInt ::= Int | IOError
  syntax IOString ::= String | IOError

Opening a file

You can open a file in K using #open. An optional mode indicates the file
open mode, which can have any value allowed by the fopen function in C.
The returned value is the file descriptor that was opened, or an error.

  syntax IOInt ::= "#open" "(" path: String ")" [function]
               | "#open" "(" path: String "," mode: String ")" [function, hook(IO.open), impure]

  rule #open(S:String) => #open(S:String, "r+")

Get/set position in file

You can get the current offset in a file using #tell. You can also seek
to a particular offset using #seek or #seekEnd. #seek is implemented via
a call to lseek with the SEEK_SET whence. #seekEnd is implemented via a
call to lseek with the SEEK_END whence. You can emulate the SEEK_CUR
whence by means of #seek(FD, #tell(FD) +Int Offset).

  syntax IOInt ::= "#tell" "(" fd: Int ")" [function, hook(IO.tell), impure]
  syntax K ::= "#seek" "(" fd: Int "," index: Int ")" [function, hook(IO.seek), impure]
             | "#seekEnd" "(" fd: Int "," fromEnd: Int ")" [function, hook(IO.seekEnd), impure]

Read from file

You can read a single character from a file using #getc. #EOF is returned
if you are at end-of-fie.

You can also read up to length characters in a file using #read. The
resulting read characters are returned, which may be fewer characters than
requested. A string of zero length being returned indicates end-of-file.

  syntax IOInt ::= "#getc" "(" fd: Int ")"             [function, hook(IO.getc), impure]
  syntax IOString ::= "#read" "(" fd: Int "," length: Int ")"    [function, hook(IO.read), impure]

Write to file

You can write a single character to a file using #putc. You can also write
a string to a file using #write. The returned value on success is .K.

  syntax K ::= "#putc" "(" fd: Int "," value: Int ")"      [function, hook(IO.putc), impure]
             | "#write" "(" fd: Int "," value: String ")" [function, hook(IO.write), impure]

Closing a file

You can close a file using #close. The returned value on success is .K.

  syntax K ::= "#close" "(" fd: Int ")" [function, hook(IO.close), impure]

Locking/unlocking a file

You can lock or unlock parts of a file using the #lock and #unlock
functions. The lock starts at the beginning of the file and continues for
endIndex bytes. Note that Unix systems do not actually prevent locked files
from being read and modified; you will have to lock both sides of a concurrent
access to guarantee exclusivity.

  syntax K ::= "#lock" "(" fd: Int "," endIndex: Int ")" [function, hook(IO.lock), impure]
             | "#unlock" "(" fd: Int "," endIndex: Int ")" [function, hook(IO.unlock), impure]

Networking

You can accept a connection on a socket using #accept, or shut down the
write end of a socket with #shutdownWrite. Note that facility is not provided
for opening, binding, and listening on sockets. These functions are implemented
in order to support creating stateful request/response servers where the
request loop is implemented using rewriting in K, but the connection
initialization is written in native code and linked into the LLVM backend.

  syntax IOInt ::= "#accept" "(" fd: Int ")" [function, hook(IO.accept), impure]
  syntax K ::= "#shutdownWrite" "(" fd: Int ")" [function, hook(IO.shutdownWrite), impure]

Time

You can get the current time in seconds since midnight UTC on January 1, 1970
using #time.

  syntax Int ::= "#time" "(" ")" [function, hook(IO.time), impure]

Builtin file descriptors

Provided here are functions that return the file descriptor for standard input,
standard output, and standard error.

  syntax Int ::= "#stdin"   [function, total]
                | "#stdout" [function, total]
                | "#stderr" [function, total]

  rule #stdin => 0
  rule #stdout => 1
  rule #stderr => 2

Shell access

You can execute a command using the shell using the #system operator. Care
must be taken to sanitize inputs to this function or security issues may
result. Note that K has no facility for reasoning about logic that happens
outside its process, so any functionality that you wish to be able to formally
reason about in K should not be implemented via the #system operator.

  syntax KItem ::= #system ( String ) [function, hook(IO.system), impure]
                 | "#systemResult" "(" Int /* exit code */ "," String /* stdout */ "," String /* stderr */ ")" [klabel(#systemResult), symbol]

Temporary files

You can get a temporary file and open it atomically using the #mkstemp
operator. The resulting file will be closed and deleted when K rewriting ends.
For more info on the argument to #mkstemp, see man mkstemp.

  syntax IOFile ::= #mkstemp(template: String) [function, hook(IO.mkstemp), impure]
  syntax IOFile ::= IOError
                  | "#tempFile" "(" path: String "," fd: Int ")" [klabel(#tempFile), symbol]

Deleting a file

You can delete a file using its absolute or relative path using the #remove
operator. It returns .K on success or an IOError on failure.

  syntax K ::= #remove(path: String) [function, total, hook(IO.remove), impure]

Logging

You can log information to disk using the #logToFile operator. Semantically,
this operator returns .K. However, it has a side effect that is not reasoned
about which is that value will be written to a uniquely-identified file
containing name in its name. The file is only flushed to disk when rewriting
finishes.

  syntax K ::= #logToFile(name: String, value: String) [function, total, hook(IO.log), impure, returnsUnit, klabel(#logToFile), symbol]

Strings can also be logged via the logging mechanisms available to the backend.
On the LLVM backend, this just means logging the text to standard error. On the
Haskell backend, a log message of type InfoUserLog is created with the
specified text.

  syntax K ::= #log(value: String) [function, total, hook(IO.logString), impure, returnsUnit, klabel(#log), symbol]

Terms can also be logged to standard error in surface syntax, rather than as
KORE using #trace. This operator has similar semantics to #logToFile (i.e.
it returns .K, but prints as an impure side effect). Note that calling
#trace is equivalent to invoking the kprint tool for the first term that is
logged, which requires re-parsing the underlying K definition. Subsequent calls
do not incur this overhead again; the definition is cached.

  syntax K ::= #trace(value: KItem) [function, total, hook(IO.traceTerm), impure, returnsUnit, klabel(#trace), symbol]
             | #traceK(value: K)    [function, total, hook(IO.traceTerm), impure, returnsUnit, klabel(#traceK), symbol]

Implementation of high-level I/O streams in K

Below is an implementation of the stream="stdin" and stream="stdout"
cell attributes in K. You should not refer to these symbols or modules directly
in your definition. It is provided only so that the K compiler can make use of
it. For more information on how to use this feature, refer to IMP++ in the K
tutorial.

  syntax Stream ::= #buffer(K)
                  | #istream(Int)
                  | #parseInput(String, String)
                  | #ostream(Int)

endmodule

// NOTE: DO NOT DIRECTLY IMPORT *-STREAM MODULES
// These stream modules will be automatically instantiated and implicitly imported
// into the main module when `stream` attributes appear in configuration cells.
// Only `Stream` productions and `[stream]` rules will be imported.
// The cell name will be replaced with the one of the main configuration.

module STDIN-STREAM
  imports K-IO
  imports K-REFLECTION
  imports LIST
  imports INT
  imports BOOL

  configuration <stdin> ListItem(#buffer($STDIN:String)) ListItem($IO:String) ListItem(#istream(#stdin)) </stdin>

  // read one character at a time until we read whitespace
  rule [stdinGetc]:
       <stdin>
       ListItem(#parseInput(_:String, Delimiters:String))
       ListItem(#buffer(S:String => S +String chrChar({#getc(N)}:>Int)))
       ListItem("on")
       ListItem(#istream(N:Int))
       </stdin>
    requires findChar(S, Delimiters, 0) ==Int -1 // [stdin]
       [stream, priority(200)]

  // when we reach whitespace, if it parses create a ListItem
  rule [stdinParseString]:
       <stdin>
       (ListItem(#parseInput("String", Delimiters:String)) => ListItem(S))
       ListItem(#buffer(S:String => ""))
       _:List
       </stdin>
    requires findChar(S, Delimiters, 0) =/=Int -1 // [stdin]
       [stream]

  // a hack: handle the case when we read integers without the help of the IO server
  rule [stdinParseInt]:
       <stdin>
       (ListItem(#parseInput("Int", Delimiters:String))
       => ListItem(String2Int(substrString(S, 0, findChar(S, Delimiters, 0)))))
       ListItem(#buffer(S:String => substrString(S,findChar(S, Delimiters, 0) +Int 1, lengthString(S))))
       _:List
       </stdin>
    requires findChar(S, Delimiters, 0) =/=Int -1
       andBool lengthString(S) >Int 1 // [stdin]
       [stream]

  rule [stdinTrim]:
       <stdin>
       ListItem(#parseInput(Sort:String, Delimiters:String))
       ListItem(#buffer(S:String => substrString(S, 1, lengthString(S))))
       _:List
       </stdin>
    requires findChar(S, Delimiters, 0) =/=Int -1
       andBool Sort =/=String "String"
       andBool lengthString(S) <=Int 1 // [stdin]
       [stream]

  // NOTE: This unblocking rule will be instantiated and inserted carefully
  // when necessary according to user-defined rules, since otherwise it will
  // lead to a diverging (i.e., non-terminating) transition system definition.
  // Currently, it supports only a simple pattern matching on the top of the
  // input stream cell, e.g.,
  //   rule <k> read() => V ... </k> <in> ListItem(V:Int) => .List ...  </in>
  // Non-supported rules that refer to the input stream cell in a sophisticated
  // way will get stuck in concrete execution mode with real IO enabled (i.e.,
  // under `--io on` option), while they will still work in symbolic execution
  // mode or concrete execution mode with real IO disabled (i.e., under `--io
  // off`, `--search`, or `--debug` options).
  //
  // TODO: More patterns need to be supported as well. In that case, we need to
  // have a way to specify such patterns.
  rule [stdinUnblock]:
       <stdin>
         (.List => ListItem(#parseInput(?Sort:String, ?Delimiters:String)))
         ListItem(#buffer(_:String))
         ...
       </stdin>

  /*
  syntax Stream ::= "#noIO"

  rule ListItem(#buffer(_))
       (ListItem(#noIO) ListItem(#istream(_:Int)) => .List) [stdin]
  */

endmodule

module STDOUT-STREAM
  imports K-IO
  imports LIST
  imports STRING

  configuration <stdout> ListItem(#ostream(#stdout)) ListItem($IO:String) ListItem(#buffer("")) </stdout>
//configuration <stderr> ListItem(#ostream(#stderr)) ListItem($IO:String) ListItem(#buffer("")) </stderr>

  rule [stdoutBufferFloat]:
       <stdout>
       ListItem(#ostream(_))
       ListItem(_)
       ListItem(#buffer(Buffer:String => Buffer +String Float2String(F)))
       (ListItem(F:Float) => .List)
       _:List
       </stdout>
       // [stdout, stderr]
       [stream, priority(25)]
  rule [stdoutBufferInt]:
       <stdout>
       ListItem(#ostream(_))
       ListItem(_)
       ListItem(#buffer(Buffer:String => Buffer +String Int2String(I)))
       (ListItem(I:Int) => .List)
       _:List
       </stdout>
       // [stdout, stderr]
       [stream, priority(25)]
  rule [stdoutBufferString]:
       <stdout>
       ListItem(#ostream(_))
       ListItem(_)
       ListItem(#buffer(Buffer:String => Buffer +String S))
       (ListItem(S:String) => .List)
       _:List
       </stdout>
       // [stdout, stderr]
       [stream, priority(25)]

  // Send first char from the buffer to the server
  rule [stdoutWrite]:
       <stdout>
       ListItem(#ostream(N:Int => {#write(N, S) ~> N:Int}:>Int))
       ListItem("on")
       ListItem(#buffer(S:String => ""))
       _:List
       </stdout>
    requires S =/=String "" // [stdout, stderr]
       [stream, priority(30)]

  /*
  syntax Stream ::= "#noIO"

  rule ListItem(#buffer(Buffer:String => Buffer +String Float2String(F)))
       (ListItem(F:Float) => .List)
       _:List [stdout, stderr]
  rule ListItem(#buffer(Buffer:String => Buffer +String Int2String(I)))
       (ListItem(I:Int) => .List)
       _:List [stdout, stderr]
  rule ListItem(#buffer(Buffer:String => Buffer +String S))
       (ListItem(S:String) => .List)
       _:List [stdout, stderr]

  rule (ListItem(#ostream(_:Int)) ListItem(#noIO) => .List)
       ListItem(#buffer(_))
       _:List [stdout, stderr]
  */

endmodule

Machine Integers

Provided here is an implementation of arbitrarily large fixed-precision binary
integers in K. This type is hooked to an implementation of integers provided
by the backend, and in particular makes use of native machine integers for
certain sizes of integer. For arbitrary-precision integers, see the INT
module above.

The syntax of machine integers in K is the same as arbitrary-precision integers
(i.e., an optional sign followed by a sequence of digits) except that machine
integers always end in a suffix pN where N is an integer indicating the
width in bits of the integer. The MInt sort is parametric, and this is
reflected in the literals. For example, the sort of 0p8 is MInt{8}.

module MINT-SYNTAX
  /*@\section{Description} The MInt implements machine integers of arbitrary
   * bit width represented in 2's complement. */
  syntax {Width} MInt{Width} [hook(MINT.MInt)]

  /*@ Machine integer of bit width and value. */
  syntax {Width} MInt{Width} ::= r"[\\+-]?[0-9]+[pP][0-9]+" [token, prec(2), hook(MINT.literal)]
endmodule

module MINT
  imports MINT-SYNTAX
  imports private INT
  imports private BOOL

Bitwidth of MInt

You can get the number of bits of width in an MInt using bitwidthMInt.

  syntax {Width} Int ::= bitwidthMInt(MInt{Width})   [function, total, hook(MINT.bitwidth)]

Int and MInt conversions

You can convert from an MInt to an Int using the MInt2Signed and
MInt2Unsigned functions. an MInt does not have a sign; its sign is instead
reflected in how operators interpret its value either as a signed integer or as
an unsigned integer. Thus, you can interpret a MInt as a signed integer witth
MInt2Signed, or as an unsigned integer respectively using MInt2Unsigned.

You can also convert from an Int to an MInt using Int2MInt. Care must
be given to ensure that the sort context where the Int2MInt operator appears
has the correct bitwidth, as this will influence the width of the resulting
MInt.

  syntax {Width} Int ::= MInt2Signed(MInt{Width})     [function, total, hook(MINT.svalue)]
                       | MInt2Unsigned(MInt{Width})     [function, total, hook(MINT.uvalue), smt-hook(bv2int)]

  syntax {Width} MInt{Width} ::= Int2MInt(Int) [function, total, hook(MINT.integer), smt-hook(int2bv)]

MInt min and max values

You can get the minimum and maximum values of a signed or unsigned MInt
with az specified bit width using sminMInt, smaxMInt, uminMInt, and
umaxMInt.

  syntax Int ::= sminMInt(Int)    [function]
               | smaxMInt(Int)    [function]
               | uminMInt(Int)    [function]
               | umaxMInt(Int)    [function]
  rule sminMInt(N:Int) => 0 -Int (1 <<Int (N -Int 1))
  rule smaxMInt(N:Int) => (1 <<Int (N -Int 1)) -Int 1
  rule uminMInt(_:Int) => 0
  rule umaxMInt(N:Int) => (1 <<Int N) -Int 1

MInt bounds checking

You can check whether a specified Int will be represented in an MInt
with a specified width without any loss of precision when interpreted as
a signed or unsigned integer using soverflowMInt and uoverflowMInt.

  syntax Bool ::= soverflowMInt(width: Int, Int)   [function]
                | uoverflowMInt(width: Int, Int)   [function]
  rule
    soverflowMInt(N:Int, I:Int)
  =>
    I <Int sminMInt(N) orBool I >Int smaxMInt(N)
  rule
    uoverflowMInt(N:Int, I:Int)
  =>
    I <Int uminMInt(N) orBool I >Int umaxMInt(N)

MInt arithmetic

You can:

  • Compute the bitwise complement ~MInt of an MInt.
  • Compute the unary negation --MInt of an MInt.
  • Compute the product *MInt of two MInts.
  • Compute the quotient /sMInt of two MInts interpreted as signed integers.
  • Compute the modulus %sMInt of two MInts interpreted as signed integers.
  • Compute the quotient /uMInt of two MInts interpreted as unsigned
    integers.
  • Compute the modulus %uMInt of two MInts interpreted as unsigned integers.
  • Compute the sum +MInt of two MInts.
  • Compute the difference -MInt of two MInts.
  • Compute the left shift <<MInt of two MInts. The second MInt is always
    interpreted as positive.
  • Compute the arithmetic right shift >>aMInt of two MInts. The second
    MInt is always interpreted as positve.
  • Compute the logical right shift >>lMInt of two MInts. The second MInt
    is always interpreted as positive.
  • Compute the bitwise and &MInt of two MInts.
  • Compute the bitwise xor xorMInt of two MInts.
  • Compute the bitwise inclusive or |MInt of two MInts.
  syntax {Width} MInt{Width} ::= "~MInt" MInt{Width} [function, total, hook(MINT.not), smt-hook(bvnot)]
                               | "--MInt" MInt{Width} [function, total, hook(MINT.neg), smt-hook(bvuminus)]
                               > left:
                                 MInt{Width} "*MInt" MInt{Width} [function, total, hook(MINT.mul), smt-hook(bvmul)]
                               | MInt{Width} "/sMInt" MInt{Width} [function, hook(MINT.sdiv), smt-hook(bvsdiv)]
                               | MInt{Width} "%sMInt" MInt{Width} [function, hook(MINT.srem), smt-hook(bvsrem)]
                               | MInt{Width} "/uMInt" MInt{Width} [function, hook(MINT.udiv), smt-hook(bvudiv)]
                               | MInt{Width} "%uMInt" MInt{Width} [function, hook(MINT.urem), smt-hook(bvurem)]
                               > left:
                                 MInt{Width} "+MInt" MInt{Width} [function, total, hook(MINT.add), smt-hook(bvadd)]
                               | MInt{Width} "-MInt" MInt{Width} [function, total, hook(MINT.sub), smt-hook(bvsub)]
                               > left:
                                 MInt{Width} "<<MInt" MInt{Width} [function, hook(MINT.shl), smt-hook(bvshl)]
                               | MInt{Width} ">>aMInt" MInt{Width} [function, hook(MINT.ashr), smt-hook(bvashr)]
                               | MInt{Width} ">>lMInt" MInt{Width} [function, hook(MINT.lshr), smt-hook(bvlshr)]
                               > left:
                                 MInt{Width} "&MInt" MInt{Width} [function, total, hook(MINT.and), smt-hook(bvand)]
                               > left:
                                 MInt{Width} "xorMInt" MInt{Width} [function, total, hook(MINT.xor), smt-hook(bvxor)]
                               > left:
                                 MInt{Width} "|MInt" MInt{Width} [function, total, hook(MINT.or), smt-hook(bvor)]

MInt comparison

You can compute whether one MInt is less than, less than or equal to, greater
than, or greater than or equal to another MInt when interpreted as signed
or unsigned integers. You can also compute whether one MInt is equal to or
unequal to another MInt.

  syntax {Width} Bool ::= MInt{Width} "<sMInt" MInt{Width} [function, total, hook(MINT.slt), smt-hook(bvslt)]
                        | MInt{Width} "<uMInt" MInt{Width} [function, total, hook(MINT.ult), smt-hook(bvult)]
                        | MInt{Width} "<=sMInt" MInt{Width} [function, total, hook(MINT.sle), smt-hook(bvsle)]
                        | MInt{Width} "<=uMInt" MInt{Width} [function, total, hook(MINT.ule), smt-hook(bvule)]
                        | MInt{Width} ">sMInt" MInt{Width} [function, total, hook(MINT.sgt), smt-hook(bvsgt)]
                        | MInt{Width} ">uMInt" MInt{Width} [function, total, hook(MINT.ugt), smt-hook(bvugt)]
                        | MInt{Width} ">=sMInt" MInt{Width} [function, total, hook(MINT.sge), smt-hook(bvsge)]
                        | MInt{Width} ">=uMInt" MInt{Width} [function, total, hook(MINT.uge), smt-hook(bvuge)]
                        | MInt{Width} "==MInt" MInt{Width} [function, total, hook(MINT.eq), smt-hook(=)]
                        | MInt{Width} "=/=MInt" MInt{Width} [function, total, hook(MINT.ne), smt-hook(distinct)]

MInt min/max

You can compute the signed minimum sMinMInt, the signed maximum sMaxMInt,
the unsigned minimum uMinMInt, and the unsigned maximum uMaxMInt of two
MInts.

  syntax {Width} MInt{Width} ::= sMaxMInt(MInt{Width}, MInt{Width}) [function, total, hook(MINT.smax), smt-hook((ite (bvslt #1 #2) #2 #1))]
                               | sMinMInt(MInt{Width}, MInt{Width}) [function, total, hook(MINT.smin), smt-hook((ite (bvslt #1 #2) #1 #2))]
                               | uMaxMInt(MInt{Width}, MInt{Width}) [function, total, hook(MINT.umax), smt-hook((ite (bvult #1 #2) #2 #1))]
                               | uMinMInt(MInt{Width}, MInt{Width}) [function, total, hook(MINT.umin), smt-hook((ite (bvult #1 #2) #1 #2))]

MInt to MInt conversion

You can convert an MInt of one width to another width with roundMInt.
The resulting MInt will be truncated starting from the most significant bit
if the resulting width is smaller than the input. The resulting MInt will be
zero-extended with the same low-order bits if the resulting width is larger
than the input.

  syntax {Width1, Width2} MInt{Width1} ::= roundMInt(MInt{Width2}) [function, total, hook(MINT.round)]
endmodule

K Language Features

Defined below is a series of modules used to parse inner syntax in K (ie, the
contents of rules, configuration declarations, and contexts).

Much of this file exists in tight correspondence with the K implementation, and
K will not work correctly if it is altered without corresponding changes to the
source code of the K tools.

Users should only import a few modules from this file. In particular, this
includes SORT-K, BASIC-K, ML-SYNTAX, DEFAULT-LAYOUT,
DEFAULT-CONFIGURATION, and K-AMBIGUITIES. The remaining modules should not
be imported by the user; they are used implicitly by the implementation of K.

Basic K Sorts

The SORT-K module declares the K sort, and nothing else.

module SORT-K
  syntax K [hook(K.K)]
endmodule

The BASIC-K module declares the K, KItem, and KConfigVar sorts, and
imports the syntax of matching logic.

module BASIC-K
  imports ML-SYNTAX
  imports SORT-BOOL
  syntax KItem [hook(K.KItem)]
  syntax K     ::= KItem
  syntax KConfigVar [token]
  syntax KItem ::= KConfigVar
endmodule

KAST Abstract Syntax

Below is defined the abstract syntax of concrete terms in K, the KAST syntax.
Users should rarely if ever have to refer to this syntax; in general, it
suffices to use concrete syntax in rules, configuration declarations, contexts,
etc.

This syntax is used directly by the K implementation, and exists here as a
reference for the syntax of KAST, but it should not be imported directly by
the user.

module KSTRING
  syntax KString ::= r"[\\\"](([^\\\"\\n\\r\\\\])|([\\\\][nrtf\\\"\\\\])|([\\\\][x][0-9a-fA-F]{2})|([\\\\][u][0-9a-fA-F]{4})|([\\\\][U][0-9a-fA-F]{8}))*[\\\"]"      [token]
  // optionally qualified strings, like in Scala "abc", i"abc", r"a*bc", etc.
endmodule

module BUILTIN-ID-TOKENS
  syntax #LowerId ::= r"[a-z][a-zA-Z0-9]*"                    [prec(2), token]
  syntax #UpperId ::= r"[A-Z][a-zA-Z0-9]*"                    [prec(2), token]
endmodule

module SORT-KBOTT
  imports SORT-K
  syntax KBott
endmodule

module KAST
  imports BASIC-K
  imports SORT-KBOTT
  imports KSTRING
  imports BUILTIN-ID-TOKENS

  syntax KBott ::= "#token" "(" KString "," KString ")"  [klabel(#KToken), symbol]
                 | "#klabel" "(" KLabel ")"              [klabel(#WrappedKLabel), symbol]
                 | KLabel "(" KList ")"                  [klabel(#KApply), symbol]
  syntax KItem ::= KBott

  syntax KLabel ::= r"`(\\\\`|\\\\\\\\|[^`\\\\\\n\\r])+`" [token]
                  | #LowerId                                   [token]
                  | r"(?<![a-zA-Z0-9])[#a-z][a-zA-Z0-9]*"               [token, prec(1)]
                       // something that doesn't collide with meta-variables

  syntax KList ::= K
                 | ".KList"          [klabel(#EmptyKList), symbol]
                 | KList "," KList   [klabel(#KList), left, assoc, unit(#EmptyKList), symbol, prefer]
endmodule


// To be used when parsing/pretty-printing ground configurations
module KSEQ
  imports KAST
  imports K-TOP-SORT
  syntax K ::= ".K"      [klabel(#EmptyK), symbol]
             | "."       [klabel(#EmptyK), symbol, deprecated, unparseAvoid]
  syntax K ::= K "~>" K  [klabel(#KSequence), left, assoc, unit(#EmptyK), symbol]
  syntax left #KSequence
  syntax {Sort} Sort     ::= "(" Sort ")"    [bracket, group(defaultBracket), applyPriority(1)]
endmodule

Syntax of Matching Logic

K provides direct access to the symbols of Matching Logic, while giving them
their own concrete syntax distinct from the syntax of the KORE intermediate
representation. These symbols are primarily used during symbolic execution.
The LLVM Backend has relatively little understanding of Matching Logic directly
and use of these symbols directly in rules is likely to cause it to crash.
However, these symbols are necessary when providing lemmas and other types of
logical assistance to proofs and symbolic execution in the Haskell Backend.

The correspondance between K symbols and KORE symbols is as follows:

  • #Top - \top
  • #Bottom - \bottom
  • #Not - \not
  • #Ceil - \ceil
  • #Floor - \floor
  • #Equals - \equals
  • #And - \and
  • #Or - \or
  • #Implies - \implies
  • #Exists - \exists
  • #Forall - \forall
  • #AG - allPathGlobally
  • #wEF - weakExistsFinally
  • #wAF - weakAlwaysFinally
module ML-SYNTAX [not-lr1]
  imports SORT-K

  syntax {Sort} Sort ::= "#Top" [klabel(#Top), symbol, group(mlUnary)]
                       | "#Bottom" [klabel(#Bottom), symbol, group(mlUnary)]
                       | "#Not" "(" Sort ")" [klabel(#Not), symbol, mlOp, group(mlUnary, mlOp)]

  syntax {Sort1, Sort2} Sort2 ::= "#Ceil" "(" Sort1 ")" [klabel(#Ceil), symbol, mlOp, group(mlUnary, mlOp)]
                                | "#Floor" "(" Sort1 ")" [klabel(#Floor), symbol, mlOp, group(mlUnary, mlOp)]
                                | "{" Sort1 "#Equals" Sort1 "}" [klabel(#Equals), symbol, mlOp, group(mlEquals, mlOp), comm, format(%1%i%n%2%d%n%3%i%n%4%d%n%5)]

  syntax priority mlUnary > mlEquals > mlAnd

  syntax {Sort} Sort ::= Sort "#And" Sort [klabel(#And), symbol, assoc, left, comm, unit(#Top), mlOp, group(mlAnd, mlOp), format(%i%1%d%n%2%n%i%3%d)]
                       > Sort "#Or" Sort [klabel(#Or), symbol, assoc, left, comm, unit(#Bottom), mlOp, group(mlOp), format(%i%1%d%n%2%n%i%3%d)]
                       > Sort "#Implies" Sort [klabel(#Implies), symbol, mlOp, group(mlImplies, mlOp), format(%i%1%d%n%2%n%i%3%d)]

  syntax priority mlImplies > mlQuantifier

  syntax {Sort1, Sort2} Sort2 ::= "#Exists" Sort1 "." Sort2 [klabel(#Exists), symbol, mlOp, mlBinder, group(mlQuantifier, mlOp)]
                                | "#Forall" Sort1 "." Sort2 [klabel(#Forall), symbol, mlOp, mlBinder, group(mlQuantifier, mlOp)]

  syntax {Sort} Sort ::= "#AG" "(" Sort ")" [klabel(#AG), symbol, mlOp, group(mlOp)]
                       | "#wEF" "(" Sort ")" [klabel(weakExistsFinally), symbol, mlOp, group(mlOp)]
                       | "#wAF" "(" Sort ")" [klabel(weakAlwaysFinally), symbol, mlOp, group(mlOp)]
endmodule

Variables in K

Provided below is the syntax of variables in K. There are four types of
variables in K:

  1. Regular variables. These are denoted by variables that begin with an
    underscore or a capital letter. These variables match exactly one value
    and can be used to refer to it on the right-hand-side.
  2. Fresh constants. These are denoted by variables that begin with an !. This
    is a convenience syntax which can be used on the right-hand-side only, and
    refer to a unique value of the specified sort which is distinct from any
    other value that has been generated or will be generated by the !X syntax.
    Note that this may not be distinct from values produced via other means.
  3. Existential variables. This refers to variables that are existentially
    quantified and begin with a ?. They are not required to appear on the
    left-hand-side prior to appearing on the right-hand-side, and generally
    refer to symbolic quantities that are introduced during rewriting. Refer to
    K's documentation for more details.
  4. Set variables. These are denoted by variables that begin with a @.
    These variables refer to a set of values and are generally used when writing
    simplification rules in the Haskell Backend. For more information, refer to
    K's documentation.

There is also a fifth type of "variable", although it is not technically a
variable. This refers to configuration variables, which are used to insert
values into the initial configuration that come from outside the semantics.
The most common of these is the $PGM variable, which conventionally contains
the program being executed and is placed in the <k> cell in the configuration
declaration. These "variables" begin with a $ and their values are populated
by the frontend prior to symbolic or concrete execution of a program.

module KVARIABLE-SYNTAX
  syntax #KVariable
endmodule

// To be used when parsing/pretty-printing symbolic configurations
module KSEQ-SYMBOLIC
  imports KSEQ
  imports ML-SYNTAX
  imports KVARIABLE-SYNTAX

  syntax #KVariable ::= r"(?<![A-Za-z0-9_\\$!\\?@])(\\!|\\?|@)?([A-Z][A-Za-z0-9'_]*|_|_[A-Z][A-Za-z0-9'_]*)"   [token, prec(1)]
                      | #UpperId                                                          [token]
  syntax KConfigVar ::= r"(?<![A-Za-z0-9_\\$!\\?@])(\\$)([A-Z][A-Za-z0-9'_]*)"            [token]
  syntax KBott      ::= #KVariable
  syntax KBott      ::= KConfigVar
endmodule

Syntax of Cells

While the backend treats cells as regular productions like any other, the
frontend provides a significant amount of convenience notation for dealing with
groups of cells, in order to make writing modular definitions easier. As a
result, we need a syntax for groups of cells and for referring to cells within
rules, configuration declarations, and functions.

For historical reasons, the Bag sort is used to refer to groups of cells.
This may change in a future release. Users can combine cells in any order
by concatenating them together, and can refer to the absence of any cells with
the .Bag symbol. You can also refer to cells within a function by placing
the cell context symbol, [[ K ]] at the top of a rule, placing a function
symbol inside, and referring to cells afterwards. This implicitly inserts
a reference to the configuration at the time prior to the currently-applied
rule being applied which can be matched on within the function. Functions with
such context cannot be referred to in the initial configuration, because the
prior configuration does not yet exist.

module KCELLS
  imports KAST

  syntax Cell
  syntax Bag ::= Bag Bag  [left, assoc, klabel(#cells), symbol, unit(#cells)]
               | ".Bag"   [klabel(#cells), symbol]
               | ".::Bag" [klabel(#cells), symbol]
               | Cell
  syntax Bag ::= "(" Bag ")" [bracket]
  syntax KItem ::= Bag
  syntax #RuleBody ::= "[" "[" K "]" "]" Bag    [klabel(#withConfig), symbol, avoid]
  syntax non-assoc #withConfig
  syntax Bag ::= KBott
endmodule

Users can also refer to cells in rules. When doing so, an optional ... can
be placed immediately after the start of the cell or immediately before the
end. In a cell whose contents are commutative, these are equivalent to one
another and are also equivalent to placing ... in both places. This means
that what is placed in the cell will be combined with the cell contents'
concatenation operator with an unnamed variable. In other words, you match on
some number of elements in the collection and do not care about the rest of
the collection.

In a cell whose contents are not commutative, the ... operators correspond
to a variable on the respective side of the contents of the cell that the
... appears. For example, <foo>... L </foo>, if L is a list, means
some number of elements followed by L. Note that not all combinations are
supported. Cells whose contents are sort K can only have ... appear at the
tail of the cell, and cells whose contents are sort List can only have ...
appear on at most one side in a single rule.

module RULE-CELLS
  imports KCELLS
  imports RULE-LISTS
  // if this module is imported, the parser automatically
  // generates, for all productions that have the attribute 'cell' or 'maincell',
  // a production like below:
  //syntax Cell ::= "<top>" #OptionalDots K #OptionalDots "</top>" [klabel(<top>)]

  syntax #OptionalDots ::= "..." [klabel(#dots), symbol]
                         | ""    [klabel(#noDots), symbol]

  syntax Int
  // this production will be added by the compiler to help handle bang variables,
  // however, it is valuable to put it here because without this production, it
  // will not exist at the point in time when rules and claims are parsed, and
  // as a result it makes it very difficult to write proof claims over fragments
  // of code that exercise rules containing bang variables. We put it here because
  // this production will "vanish" after parsing finishes and not be picked up
  // by the compiler, which is the behavior we want in this case since an actual
  // production will be generated by the compiler later on.
  syntax GeneratedCounterCell ::= "<generatedCounter>" Int "</generatedCounter>" [cell, klabel(<generatedCounter>), symbol, internal]
endmodule

Users can also declare cells in a configuration declaration. This generates a
specific set of productions that is used internally to implement the cell. The
most important of these is the cell itself, and attributes on this production
can be specified in an xml-attribute-like syntax.

You can also use an xml-short-tag-like syntax to compose configuration cells
together which were defined in different modules. However, it is a requirement
that any K definition have at most one fully-composed configuration; thus, all
other configuration declarations must appear composed within another
configuration declaration.

module CONFIG-CELLS
  imports KCELLS
  imports RULE-LISTS
  syntax #CellName ::= r"[a-zA-Z][a-zA-Z0-9\\-]*"  [token, prec(1)]
                     | #LowerId            [token]
                     | #UpperId            [token]

  syntax Cell ::= "<" #CellName #CellProperties ">" K "</" #CellName ">" [klabel(#configCell), symbol]
  syntax Cell ::= "<" #CellName "/>" [klabel(#externalCell), symbol]

  syntax #CellProperties ::= #CellProperty #CellProperties [klabel(#cellPropertyList), symbol]
                           | ""                            [klabel(#cellPropertyListTerminator), symbol]
  syntax #CellProperty ::= #CellName "=" KString           [klabel(#cellProperty), symbol]
endmodule

Syntax of Rules

Rules can have an optional requires clause or an ensures clause. For backwards-
compatibility, you can refer to the requires clause with both the requires
and when keywords; The latter, however, is deprecated and may be removed in
a future release.

The requires clause specifies the preconditions that must be true in order
for the rule to apply. The ensures clause specifies the information which
becomes true after the rule has applied. It is a requirement that information
present in the ensures clause refer to existential variables only.

When doing concrete execution, you can think of the requires clause as a
side-condition. In other words, even if the rule matches, it will not apply
unless the requires clause, which must be of sort Bool, evaluates to
true.

module REQUIRES-ENSURES
  imports BASIC-K

  syntax #RuleBody ::= K

  syntax #RuleContent ::= #RuleBody                                 [klabel("#ruleNoConditions"), symbol]
                        | #RuleBody "requires" Bool                 [klabel("#ruleRequires"), symbol]
                        | #RuleBody "ensures"  Bool                 [klabel("#ruleEnsures"), symbol]
                        | #RuleBody "requires" Bool "ensures" Bool  [klabel("#ruleRequiresEnsures"), symbol]
endmodule

Miscellaneous modules

The below modules are used in various ways as indicators to the implementation
that certain automatically generated syntax should be created by the parser.
These modules should not be imported directly by the user.

module K-TOP-SORT
  imports SORT-KBOTT
  syntax KItem ::= KBott
  syntax {Sort} KItem ::= Sort
endmodule

module K-BOTTOM-SORT
  imports SORT-KBOTT
  syntax KItem ::= KBott
  syntax {Sort} Sort ::= KBott
endmodule

module K-SORT-LATTICE
  imports K-TOP-SORT
  imports K-BOTTOM-SORT
endmodule

module AUTO-CASTS
  // if this module is imported, the parser automatically
  // generates, for all sorts, productions of the form:
  // Sort  ::= Sort ":Sort"  // semantic cast - force the inner term to be `Sort` or a subsort
  // Sort  ::= Sort "::Sort" // strict cast - force the inner term to be exactly `Sort`. Useful for disambiguation
  // Sort ::= "{" Sort "}" "::Sort" // synonym for strict cast
  // Sort  ::= "{" K "}"    ":>Sort" // projection cast. Allows any term to be placed in a context that expects `Sort`
  // this is part of the mechanism that allows concrete user syntax in K
endmodule

module AUTO-FOLLOW
  // if this module is imported, the parser automatically
  // generates a follow restriction for every terminal which is a prefix
  // of another terminal. This is useful to prevent ambiguities such as:
  // syntax K ::= "a"
  // syntax K ::= "b"
  // syntax K ::= "ab"
  // syntax K ::= K K
  // #parse("ab", "K")
  // In the above example, the terminal "a" is not allowed to be followed by a "b"
  // because it would turn the terminal into the terminal "ab".
endmodule

module PROGRAM-LISTS
  imports SORT-K
  // if this module is imported, the parser automatically
  // replaces the default productions for lists:
  // Es ::= E "," Es [userList("*"), klabel('_,_)]
  //      | ".Es"    [userList("*"), klabel('.Es)]
  // into a series of productions more suitable for programs:
  // Es#Terminator ::= ""      [klabel('.Es)]
  // Ne#Es ::= E "," Ne#Es     [klabel('_,_)]
  //         | E Es#Terminator [klabel('_,_)]
  // Es ::= Ne#Es
  //      | Es#Terminator      // if the list is *
endmodule

module RULE-LISTS
  // if this module is imported, the parser automatically
  // adds the subsort production to the parsing module only:
  // Es ::= E        [userList("*")]

endmodule

module RECORD-PRODUCTIONS
  // if this module is imported, prefix productions of the form
  // syntax Sort ::= name(Args)
  // will be able to be parsed with don't-care variables according
  // to their nonterminal's names
endmodule

module SORT-PREDICATES
  // if this module is imported, the Bool sort will be annotated with
  // syntax Bool ::= isSort(K) [function]
  // and all sorts will be annotated with
  // syntax Sort ::= project:Sort(K) [function]
endmodule

Additional Syntax for K Terms in Rules

Certain additional features are available when parsing the contents of rules
and contexts. For more information on each of these, refer to K's
documentation.

module KREWRITE
  syntax {Sort} Sort ::= Sort "=>" Sort [klabel(#KRewrite), symbol]
  syntax non-assoc #KRewrite
  syntax priority #KRewrite > #withConfig
endmodule

// To be used to parse semantic rules
module K
  imports KSEQ-SYMBOLIC
  imports REQUIRES-ENSURES
  imports RECORD-PRODUCTIONS
  imports SORT-PREDICATES
  imports K-SORT-LATTICE
  imports AUTO-CASTS
  imports AUTO-FOLLOW
  imports KREWRITE

  syntax {Sort} Sort ::= Sort "#as" Sort [klabel(#KAs), symbol]
  // functions that preserve sorts and can therefore have inner rewrites
  syntax {Sort} Sort ::= "#fun" "(" Sort ")" "(" Sort ")" [klabel(#fun2), symbol, prefer]
  // functions that do not preserve sort and therefore cannot have inner rewrites
  syntax {Sort1, Sort2} Sort1 ::= "#fun" "(" Sort2 "=>" Sort1 ")" "(" Sort2 ")" [klabel(#fun3), symbol]

  syntax {Sort1, Sort2} Sort1 ::= "#let" Sort2 "=" Sort2 "#in" Sort1 [klabel(#let), symbol]

  /*@ Set membership over terms. In addition to equality over
      concrete patterns, K also supports computing equality
      between a concrete pattern and a symbolic pattern.
      This is compiled efficiently down to pattern matching,
      and can be used by putting a term with unbound variables
      in the left child of :=K or =/=K. Note that this does not
      bind variables used on the lhs however (although this may
      change in the future).*/

  syntax Bool ::= left:
                  K ":=K" K           [function, total, klabel(_:=K_), symbol, group(equalEqualK)]
                | K ":/=K" K          [function, total, klabel(_:/=K_), symbol, group(notEqualEqualK)]
endmodule

// To be used to parse terms in full K
module K-TERM
  imports KSEQ-SYMBOLIC
  imports RECORD-PRODUCTIONS
  imports SORT-PREDICATES
  imports K-SORT-LATTICE
  imports AUTO-CASTS
  imports AUTO-FOLLOW
  imports KREWRITE
endmodule

Layout Information

When constructing a scanner for use during parsing, often you wish to ignore
certain types of text, such as whitespace and comments. However, the specific
syntax which each language must ignore is a little different from language
to language, and thus you wish to specify it manually. You can do this by
defining productions of the #Layout sort. For more information, refer to
K's documentation. However, this module will be implicitly imported if no
productions are declared of sort #Layout. This module will also be used
for the purposes of parsing K rules. If you wish to declare a language with
no layout productions, simply create a sort declaration for the #Layout sort
in your code (e.g. syntax #Layout).

module DEFAULT-LAYOUT
    syntax #Layout ::= r"(\\/\\*([^\\*]|(\\*+([^\\*\\/])))*\\*+\\/)" // C-style multi-line comments
                     | r"(\\/\\/[^\\n\\r]*)"                         // C-style single-line comments
                     | r"([\\ \\n\\r\\t])"                           // Whitespace
endmodule

Default Configuration

If the user has no configuration declaration in their seamantics, the below
configuration declaration will be implicitly imported.

module DEFAULT-CONFIGURATION
  imports BASIC-K

  configuration <k> $PGM:K </k>
endmodule

Parsing Ambiguous Languages

On occasion, it may be desirable to parse a language with an ambiguous grammar
when parsing a program, and perform additional semantic analysis at a later
time in order to resolve the ambiguities. A good example of this is as a
substitute for the lexer hack in parsers of the C programming language.

The following module contains a declaration for ambiguities in K. Usually,
an ambiguous parse is an error. However, when you use the --gen-glr-parser
flag to kast, or the --gen-glr-bison-parser flag to kompile, ambiguities
instead become instances of the below parametric production, which you can use
regular K rules to disambiguate as necessary.

module K-AMBIGUITIES

  syntax {Sort} Sort ::= amb(Sort, Sort) [klabel(amb), symbol]

endmodule

Annotating Parses with Locations

Another feature of K's Bison parser is the ability to annotate terms parsed
with location information about the file and line where they occurred. For
more information about how to use this, refer to K's documentation. However,
the below module exists to provide a user syntax for the annotations that
are generated by the parser.

module K-LOCATIONS
  imports STRING-SYNTAX
  imports INT-SYNTAX

  // filename, startLine, startCol, endLine, endCol
  syntax {Sort} Sort ::= #location(Sort, String, Int, Int, Int, Int) [klabel(#location), symbol, format(%3)]

endmodule

K Prelude

The following files, integral to defining semantics in K, are automatically
required by every definition via this file. This behavior can be disabled
via kompile --no-prelude, however, semantics will likely break unless
they provide their own versions of these files, which are assumed to exist
by the compiler. There are, however, circumstances where passing this flag is
appropriate, such as if you are manually requiring these files in your
definition, if your definition was automatically condensed into a single file
with kompile -E, or if you wish to modify the inner syntax of K by providing
your own version of these files with different syntax.

requires "kast.md"
requires "domains.md"

K Foreign Function Interface

The K Foreign Function Interface (FFI) module provides a way to call native
functions directly from a K semantics using the C ABI. It also provides
utilities for allocating and deallocating byte buffers with static addresses
that are suitable for being passed to native code.

It is built off of the underlying libffi library
(https://sourceware.org/libffi/) and is subject to some of the same
limitations as that library. Bear in mind, because this library exposes
a number of unsafe C APIs directly, misuse of the library is likely to lead
to memory corruption in your interpreter and can cause segmentation faults or
corrupted term representations that lead to undefined behavior at runtime.

requires "domains.md"

module FFI-SYNTAX
  imports private LIST

The FFIType sort is used to declare the native C ABI types of operands passed
to the #ffiCall function. These types roughly correspond to the types
declared in ffi.h by libffi.

  syntax FFIType ::= "#void" [klabel(#ffi_void), symbol]
                  | "#uint8" [klabel(#ffi_uint8), symbol]
                  | "#sint8" [klabel(#ffi_sint8), symbol]
                  | "#uint16" [klabel(#ffi_uint16), symbol]
                  | "#sint16" [klabel(#ffi_sint16), symbol]
                  | "#uint32" [klabel(#ffi_uint32), symbol]
                  | "#sint32" [klabel(#ffi_sint32), symbol]
                  | "#uint64" [klabel(#ffi_uint64), symbol]
                  | "#sint64" [klabel(#ffi_sint64), symbol]
                  | "#float" [klabel(#ffi_float), symbol]
                  | "#double" [klabel(#ffi_double), symbol]
                  | "#uchar" [klabel(#ffi_uchar), symbol]
                  | "#schar" [klabel(#ffi_schar), symbol]
                  | "#ushort" [klabel(#ffi_ushort), symbol]
                  | "#sshort" [klabel(#ffi_sshort), symbol]
                  | "#uint" [klabel(#ffi_uint), symbol]
                  | "#sint" [klabel(#ffi_sint), symbol]
                  | "#ulong" [klabel(#ffi_ulong), symbol]
                  | "#slong" [klabel(#ffi_slong), symbol]
                  | "#longdouble" [klabel(#ffi_longdouble), symbol]
                  | "#pointer" [klabel(#ffi_pointer), symbol]
                  | "#complexfloat" [klabel(#ffi_complexfloat), symbol]
                  | "#complexdouble" [klabel(#ffi_complexdouble), symbol]
                  | "#complexlongdouble" [klabel(#ffi_complexlongdouble), symbol]
                  | "#struct" "(" List ")" [klabel(#ffi_struct), symbol]
endmodule

module FFI
  imports FFI-SYNTAX
  imports private BYTES
  imports private STRING
  imports private BOOL
  imports private LIST
  imports private INT

FFI Calls

The #ffiCall functions are designed to call a native C ABI function and
return a native result. They come in three variants:

Non-variadic

In the first variant, #ffiCall(Address, Args, ArgTypes, ReturnType) takes
an integer address of a function (which can be obtained from
#functionAddress), a List of Bytes containing the arguments of the
function, a List of FFITypes containing the types of the parameters of the
function, and an FFIType containing the return type of the function, and
returns the return value of the function as a Bytes.

  syntax Bytes ::= "#ffiCall" "(" Int "," List "," List "," FFIType ")" [function, hook(FFI.call)]

Variadic

In the second variant,
#ffiCall(Address, Args, FixedTypes, VariadicTypes, ReturnType takes an
integer address of a function, a List of Bytes containing the arguments
of the call, a List of FFITypes containing the types of the fixed
parameters of the function, a List of FFITypes containing the types of the
variadic parameters of the function, and an FFIType containing the return
type of the function, and returns the return value of the function as a
Bytes.

  syntax Bytes ::= "#ffiCall" "(" Int "," List "," List "," List "," FFIType ")" [function, hook(FFI.call_variadic)]

Generic

In the third variant,
#ffiCall(IsVariadic, Address, Args, ArgTypes, NFixed, ReturnType takes
a boolean indicating whether the function is variadic or not, an integer
address of a function, a List of Bytes containing the arguments of the
call, a List of FFITypes containing the parameter typess of the call
followed by the types of the variadic arguments of the call, if any, an Int
containing how many of the arguments of the call are fixed or not, and an
FFIType containing the return type of the function, and returns the return
value of the function as a Bytes.

  syntax Bytes ::= "#ffiCall" "(" Bool "," Int "," List "," List "," Int "," FFIType ")" [function]

  rule #ffiCall(false, Addr::Int, Args::List, Types::List, _, Ret::FFIType) => #ffiCall(Addr, Args, Types, Ret)
  rule #ffiCall(true, Addr::Int, Args::List, Types::List, NFixed::Int, Ret::FFIType) => #ffiCall(Addr, Args, range(Types, 0, size(Types) -Int NFixed), range(Types, NFixed, 0), Ret)

Symbol Lookup

The FFI module provides a mechanism to look up any function symbol and return
that function's address.

  syntax Int ::= "#functionAddress" "(" String ")" [function, hook(FFI.address)]

Direct Memory Management

Most memory used by the LLVM backend to represent terms is managed
automatically via garbage collection. However, a consequence of this is that
a particular term does not have a fixed address across its entire lifetime
in most cases. Sometimes this is undesirable, especially if you intend for
the address of the memory to be taken by the semantics or if you intend
to pass this memory directly to native code. As a result, the FFI module
exposes the following unsafe APIs for memory management. Note that use of
these APIs leaves the burden of memory management completely on the user,
and thus misuse of these functions can lead to things like use-after-free
and other memory corruption bugs.

Allocation

#alloc(Key, Size, Align) will allocate Size bytes with an alignment
requirement of Align (which must be a power of two), and return it as a
Bytes term. The memory is uniquely identified by its key and that key will
be used later to free the memory. The memory is not implicitly freed by garbage
collection; failure to call #free on the memory at a later date can lead to
memory leaks.

  syntax Bytes ::= "#alloc" "(" KItem "," Int "," Int ")" [function, hook(FFI.alloc)]

Addressing

#addess(B) will return an Int representing the address of the first byte of
B, which must be a Bytes. Unless the Bytes term was allocated by #alloc,
the return value is unspecified and may not be the same across multipl
invocations on the same byte buffer. However, it is guaranteed that memory
allocated by #alloc will have the same address throughout its lifetime.

  syntax Int ::= "#address" "(" Bytes ")" [function, hook(FFI.bytes_address)]

Deallocation

#free(Key) will free the memory of the Bytes object that was allocated
by a previous call to #alloc. If Key was not used in a previous call to
#alloc, or the memory was already freed, no action is taken. It will generate
undefined behavior if the Bytes term returned by the previous call to
#alloc is still referenced by any other term in the configuration or a
currently evaluating rule. The function returns .K.

  syntax K ::= "#free" "(" KItem ")" [function, hook(FFI.free)]

Reading

#nativeRead(Addr, Mem) will read native memory at address Addr into Mem,
reading exactly lengthBytes(Mem) bytes. This will generate undefined behavior
if Addr does not point to a readable segment of memory at least
lengthBytes(Mem) bytes long.

  syntax K ::= "#nativeRead" "(" Int "," Bytes ")" [function, hook(FFI.read)]

Writing

#nativeWrite(Addr, Mem) will write the contents of Mem to native memory at
address Addr. The memory will be read prior to being written, and a write
will only happen if the memory has a different value than the current value of
Mem. This will generate undefined behavior if Addr does not point to a
readable segment of memory at least lengthBytes(Mem) bytes long, or if the
memory at address Addr has a different value than currently contained in
Mem, and the memory in question is not writeable.

  syntax K ::= "#nativeWrite" "(" Int "," Bytes ")" [function, hook(FFI.write)]
endmodule

Syntax of JSON

K provides builtin support for reading/writing to JSON. While the JSON-SYNTAX
module is not precisely the syntax of JSON (utilizing K's syntax for strings,
integers, and floating point numbers rather than the syntax used by JSON),
you can still convert directly to/from the actual syntax of JSON using
the JSON2String and String2JSON hooks.

module JSON-SYNTAX
    imports INT-SYNTAX
    imports STRING-SYNTAX
    imports BOOL-SYNTAX
    imports FLOAT-SYNTAX

    syntax JSONs   ::= List{JSON,","}      [symbol(JSONs)]
    syntax JSONKey ::= String
    syntax JSON    ::= "null"              [klabel(JSONnull)   , symbol]
                     | String | Int | Float | Bool
                     | JSONKey ":" JSON    [klabel(JSONEntry)  , symbol]
                     | "{" JSONs "}"       [klabel(JSONObject) , symbol]
                     | "[" JSONs "]"       [klabel(JSONList)   , symbol]
endmodule

Conversion between JSON and String

Given a string written in valid JSON, you can convert it to the JSON
sort with the String2JSON function. Assuming the user has not extended
the syntax of the JSON sort with their own constructors, any term of sort
JSON can also be converted to a String using the JSON2String function.

module JSON
    imports JSON-SYNTAX

    syntax String ::= JSON2String(JSON) [function, symbol(JSON2String), hook(JSON.json2string)]

    syntax JSON ::= String2JSON(String) [function, symbol(String2JSON), hook(JSON.string2json)]
endmodule

Rational Numbers in K

K provides support for arbitrary-precision rational numbers represented as a
quotient between two integers. The sort representing these values is Rat.
Int is a subsort of Rat, and it is guaranteed that any integer will be
represented as an Int and can be matched as such on the left hand side
of rules. K also supports the usual arithmetic operators over rational numbers.

module RAT-SYNTAX
  imports INT-SYNTAX
  imports private BOOL

  syntax Rat

  syntax Rat ::= Int

Arithmetic

You can:

  • Raise a rational number to any negative or nonnegative integer.
  • Multiply or divide two rational numbers to obtain a product or quotient.
  • Add or subtract two rational numbers to obtain a sum or difference.
  syntax Rat ::= left:
                 Rat "^Rat" Int [function, total, klabel(_^Rat_), symbol, smtlib(ratpow), hook(RAT.pow)]
               > left:
                 Rat "*Rat" Rat [function, total, klabel(_*Rat_), symbol, left, smtlib(ratmul), hook(RAT.mul)]
               | Rat "/Rat" Rat [function,             klabel(_/Rat_), symbol, left, smtlib(ratdiv), hook(RAT.div)]
               > left:
                 Rat "+Rat" Rat [function, total, klabel(_+Rat_), symbol, left, smtlib(ratadd), hook(RAT.add)]
               | Rat "-Rat" Rat [function, total, klabel(_-Rat_), symbol, left, smtlib(ratsub), hook(RAT.sub)]

Comparison

You can determine whether two rational numbers are equal, unequal, or compare
one of less than, less than or equalto, greater than, or greater than or equal
to the other:

  syntax Bool ::= Rat  "==Rat" Rat [function, total, klabel(_==Rat_),  symbol, smtlib(rateq), hook(RAT.eq)]
                | Rat "=/=Rat" Rat [function, total, klabel(_=/=Rat_), symbol, smtlib(ratne), hook(RAT.ne)]
                | Rat   ">Rat" Rat [function, total, klabel(_>Rat_),   symbol, smtlib(ratgt), hook(RAT.gt)]
                | Rat  ">=Rat" Rat [function, total, klabel(_>=Rat_),  symbol, smtlib(ratge), hook(RAT.ge)]
                | Rat   "<Rat" Rat [function, total, klabel(_<Rat_),   symbol, smtlib(ratlt), hook(RAT.lt)]
                | Rat  "<=Rat" Rat [function, total, klabel(_<=Rat_),  symbol, smtlib(ratle), hook(RAT.le)]

Min/Max

You can compute the minimum and maximum of two rational numbers:

  syntax Rat ::= minRat(Rat, Rat) [function, total, klabel(minRat), symbol, smtlib(ratmin), hook(RAT.min)]
               | maxRat(Rat, Rat) [function, total, klabel(maxRat), symbol, smtlib(ratmax), hook(RAT.max)]

Conversion to Floating Point

You can convert a rational number to the nearest floating point number that
is representable in a Float of a specified number of precision and exponent
bits:

  syntax Float ::= Rat2Float(Rat, precision: Int, exponentBits: Int) [function]
endmodule

Implementation of Rational Numbers

The remainder of this file consists of an implementation in K of the
operations listed above. Users of the RAT module should not use any of the
syntax defined in any of these modules.

As a point of reference for users, it is worth noting that rational numbers
are normalized to a canonical form by this module,. with the canonical form
bearing the property that it is either an Int, or a pair of integers
I /Rat J such that
I =/=Int 0 andBool J >=Int 2 andBool gcdInt(I, J) ==Int 1 is always true.

module RAT-COMMON
  imports RAT-SYNTAX

  // invariant of < I , J >Rat : I =/= 0, J >= 2, and I and J are coprime
  syntax Rat ::= "<" Int "," Int ">Rat" [format(%2 /Rat %4)]
endmodule

module RAT-SYMBOLIC [symbolic]
  imports private RAT-COMMON
  imports ML-SYNTAX
  imports private BOOL

  rule
    #Ceil(@R1:Rat /Rat @R2:Rat)
  =>
    {(@R2 =/=Rat 0) #Equals true} #And #Ceil(@R1) #And #Ceil(@R2)
  [simplification]
endmodule

module RAT-KORE
  imports private RAT-COMMON
  imports private K-EQUAL

  /*
   * equalities
   */

  // NOTE: the two rules below may not work correctly in non-kore backends

  rule R ==Rat S => R ==K S

  rule R =/=Rat S => R =/=K S
endmodule

module RAT [private]
  imports private RAT-COMMON
  imports public RAT-SYMBOLIC
  imports public RAT-KORE
  imports public RAT-SYNTAX
  imports private INT
  imports private BOOL

  /*
   * arithmetic
   */

  rule < I , I' >Rat +Rat < J , J' >Rat => ((I *Int J') +Int (I' *Int J)) /Rat (I' *Int J')
  rule I:Int         +Rat < J , J' >Rat => ((I *Int J') +Int J) /Rat J'
  rule < J , J' >Rat +Rat I:Int         => I +Rat < J , J' >Rat
  rule I:Int         +Rat J:Int         => I +Int J

  rule < I , I' >Rat *Rat < J , J' >Rat => (I *Int J) /Rat (I' *Int J')
  rule I:Int         *Rat < J , J' >Rat => (I *Int J) /Rat J'
  rule < J , J' >Rat *Rat I:Int         => I *Rat < J , J' >Rat
  rule I:Int         *Rat J:Int         => I *Int J

  rule < I , I' >Rat /Rat < J , J' >Rat => (I *Int J') /Rat (I' *Int J)
  rule I:Int         /Rat < J , J' >Rat => (I *Int J') /Rat J
  rule < I , I' >Rat /Rat J:Int         => I /Rat (I' *Int J) requires J =/=Int 0
  rule I:Int         /Rat J:Int         => makeRat(I, J)      requires J =/=Int 0

  // derived

  rule R -Rat S => R +Rat (-1 *Rat S)

  // normalize

  syntax Rat ::= makeRat(Int, Int)      [function]
               | makeRat(Int, Int, Int) [function]

  rule makeRat(0, J) => 0 requires J =/=Int 0

  rule makeRat(I, J) => makeRat(I, J, gcdInt(I,J)) requires I =/=Int 0 andBool J =/=Int 0

  // makeRat(I, J, D) is defined when I =/= 0, J =/= 0, D > 0, and D = gcd(I,J)
  rule makeRat(I, J, D) => I /Int D                       requires J ==Int D // implies J > 0 since D > 0
  rule makeRat(I, J, D) => < I /Int D , J /Int D >Rat     requires J >Int 0 andBool J =/=Int D
  rule makeRat(I, J, D) => makeRat(0 -Int I, 0 -Int J, D) requires J <Int 0

  // gcdInt(a,b) computes the gcd of |a| and |b|, which is positive.
  syntax Int ::= gcdInt(Int, Int) [function, public]

  rule gcdInt(A, 0) => A        requires A >Int 0
  rule gcdInt(A, 0) => 0 -Int A requires A <Int 0
  rule gcdInt(A, B) => gcdInt(B, A %Int B) requires B =/=Int 0 // since |A %Int B| = |A| %Int |B|

  /*
   * exponentiation
   */

  rule _ ^Rat 0 => 1
  rule 0 ^Rat N => 0 requires N =/=Int 0

  rule < I , J >Rat ^Rat N => powRat(< I , J >Rat, N) requires N >Int 0
  rule X:Int        ^Rat N => X ^Int N                requires N >Int 0

  rule X ^Rat N => (1 /Rat X) ^Rat (0 -Int N) requires X =/=Rat 0 andBool N <Int 0

  // exponentiation by squaring

  syntax Rat ::= powRat(Rat, Int) [function]

  // powRat(X, N) is defined when X =/= 0 and N > 0
  rule powRat(X, 1) => X
  rule powRat(X, N) => powRat(X *Rat X, N /Int 2) requires N >Int 1 andBool N %Int 2  ==Int 0
  rule powRat(X, N) => powRat(X, N -Int 1) *Rat X requires N >Int 1 andBool N %Int 2 =/=Int 0

  /*
   * inequalities
   */

  rule R >Rat S => R -Rat S >Rat 0 requires S =/=Rat 0

  rule < I , _ >Rat >Rat 0 => I >Int 0
  rule I:Int        >Rat 0 => I >Int 0

  // derived

  rule R >=Rat S => notBool R <Rat S

  rule R <Rat S => S >Rat R

  rule R <=Rat S => S >=Rat R

  rule minRat(R, S) => R requires R <=Rat S
  rule minRat(R, S) => S requires S <=Rat R

  rule maxRat(R, S) => R requires R >=Rat S
  rule maxRat(R, S) => S requires S >=Rat R

  syntax Float ::= #Rat2Float(Int, Int, Int, Int) [function, hook(FLOAT.rat2float)]
  rule Rat2Float(Num:Int, Prec:Int, Exp:Int) => #Rat2Float(Num, 1, Prec, Exp)
  rule Rat2Float(< Num, Dem >Rat, Prec, Exp) => #Rat2Float(Num, Dem, Prec, Exp)

endmodule

Capture-Aware Substitution in K

One of the traditional ways in which functional languages are given operational
semantics is via substitution. In particular, you can view a function as
declaring a particular bound variable, the parameter of the function, as well
as the body of the function, within which both bound and free variables can
occur, and implement the process of beta-reduction (one of the axioms of the
lambda calculus) by means of a substitution operator which is aware of the
difference between free variables and bound variables and prevents variable
capture.

In K this is implemented using two mechanisms: The KVar sort, and the
binder attribute.

The KVar Sort

K introduces a new hooked sort, KVar, which the substitution operator
(defined below) understands in a particular way. The syntax of KVar is the
same as for sort Id in DOMAINS, but with a different sort name. Similarly,
some of the same operators are defined over KVar which are defined for Id,
such as conversion from String to KVar and support for the !Var:KVar
syntax.

A KVar is simply an identifier with special meaning during substitution.
KVars must begin with a letter or underscore,
and can be followed by zero or more letters, numbers, or underscores.

module KVAR-SYNTAX-PROGRAM-PARSING
  imports BUILTIN-ID-TOKENS

  syntax KVar ::= r"(?<![A-Za-z0-9\\_])[A-Za-z\\_][A-Za-z0-9\\_]*"     [prec(1), token]
                | #LowerId                                             [token]
                | #UpperId                                             [token]
endmodule

module KVAR-SYNTAX
  syntax KVar [token, hook(KVAR.KVar)]
endmodule

module KVAR-COMMON
  imports KVAR-SYNTAX
  imports private STRING

  syntax KVar ::= String2KVar (String) [function, total, hook(STRING.string2token)]
  syntax KVar ::= freshKVar(Int)    [freshGenerator, function, total, private]

  rule freshKVar(I:Int) => String2KVar("_" +String Int2String(I))
endmodule

module KVAR
  imports KVAR-COMMON
endmodule

The binder Attribute

A production can be given the attribute binder. Such a production must have
at least two nonterminals. The first nonterminal from left to right must be of
sort KVar, and contains the bound variable. The last nonterminal from left
to right contains the term that is bound. For example, I could describe lambdas
in the lambda calculus with the production
syntax Val ::= "lambda" KVar "." Exp [binder].

Substitution

K provides a hooked implementation of substitution, currently only implemented
on the Java and LLVM backends. Two variants exist: the first substitutes
a single KVar for a single KItem. The second takes a Map with KVar
keys and KItem values, and substitutes each element in the map atomically.

Internally, this is implemented in the LLVM backend by a combination of
de Bruijn indices for bound variables and names for free variables. Free
variables are also sometimes given a unique numeric identifier in order to
prevent capture, and the rewriter will automatically assign unique names to
such identifiers when rewriting finishes. The names assigned will always begin
with the original name of the variable and be followed by a unique integer
suffix. However, the names assigned after rewriting finishes might be different
from the names that would be assigned if rewriting were to halt prematurely,
for example due to krun --depth.

module SUBSTITUTION
  imports private MAP
  imports KVAR

  syntax {Sort} Sort ::= Sort "[" KItem "/" KItem "]"  [function, hook(SUBSTITUTION.substOne), impure]
  syntax {Sort} Sort ::= Sort "[" Map "]"      [function, hook(SUBSTITUTION.substMany), impure]
endmodule

K PL Tutorial

Here you will learn how to use the K tool to define languages by means of a series of screencast movies. It is recommended to do these in the indicated order, because K features already discussed in a previous language definition will likely not be rediscussed in latter definitions. The screencasts follow quite closely the structure of the files under the tutorial folder in the K tool distribution. If you'd rather follow the instructions there and do the tutorial exercises yourself, then go back to https://kframework.org and download the K tool, if you have not done it already. Or, you can first watch the screencasts below and then do the exercises, or do them in parallel.

K Overview

Make sure you watch the K overview video before you do the K tutorial:

Learning K

[34'46"] Part 1: Defining LAMBDA

Here you will learn how to define a very simple functional language in K and the basics of how to use the K tool. The language is a call-by-value variant of lambda calculus with builtins and mu, and its definition is based on substitution.

[37'07"] Part 2: Defining IMP

Here you will learn how to define a very simple, prototypical textbook C-like imperative language, called IMP, and several new features of the K tool.

[33'10"] Part 3: Defining LAMBDA++

Here you will learn how to define constructs which abruptly change the execution control, as well as how to define functional languages using environments and closures. LAMBDA++ extends the LAMBDA language above with a callcc construct.

[46'46"] Part 4: Defining IMP++

Here you will learn how to refine configurations, how to generate fresh elements, how to tag syntactic constructs and rules, how to exhaustively search the space of non-deterministic or concurrent program executions, etc. IMP++ extends the IMP language above with increment, blocks and locals, dynamic threads, input/output, and abrupt termination.

[17'03"] Part 5: Defining Type Systems

Here you will learn how to define various kinds of type systems following various approaches or styles using K.

[??'??"] Part 6: Miscellaneous Other K Features

Here you will learn a few other K features, and better understand how features that you have already seen work.

  • [??'??"] ...

Learning Language Design and Semantics using K

[??'??"] Part 7: SIMPLE: Designing Imperative Programming Languages

Here you will learn how to design imperative programming languages using K. SIMPLE is an imperative language with functions, threads, pointers, exceptions, multi-dimensional arrays, etc. We first define an untyped version of SIMPLE, then a typed version. For the typed version, we define both a static and a dynamic semantics.

[??'??"] Part 8: KOOL: Designing Object-Oriented Programming Languages

Here woul will learn how to design object-oriented programming languages using K. KOOL is an object-oriented language that extends SIMPLE with classes and objects. We first define an untyped version of KOOL, then a typed version, with both a dynamic and a static semantics.

[??'??"] Part 9: FUN: Designing Functional Programming Languages

H
ere woul will learn how to design functional programming languages using K. FUN is a higher-order functional language with general let, letrec, pattern matching, references, lists, callcc, etc. We first define an untyped version of FUN, then a let-polymorphic type inferencer.

[??'??"] Part 10: LOGIK: Designing Logic Programming Languages

Here you will learn how to design a logic programming language using K.

K overview

Go to Youtube mirror, if the above does not work.

Go back to https://kframework.org for further links, the K tool and contact information.

Learning K

We start by introducing the basic features of K by means of a series
of very simple languages. The objective here is neither to learn those
languages nor to study their underlying paradigm, but simply to learn K.

  • LAMBDA: Lambda calculus defined.
  • IMP: A simple imperative language.
  • LAMBDA++: LAMBDA extended with control flow.
  • IMP++: IMP extended with threads and IO.
  • TYPES: LAMBDA type system.

Part 1: Defining LAMBDA

Here you will learn how to define a very simple language in K and the basics
of how to use the K tool. The language is a variant of call-by-value lambda
calculus and its definition is based on substitution. Specifically, you will
learn the following:

  • How to define a module.
  • How to define a language syntax.
  • How to use the defined syntax to parse programs.
  • How to import predefined modules.
  • How to define evaluation strategies using strictness attributes.
  • How to define semantic rules.
  • How the predefined generic substitution works.
  • How to generate PDF and HTML documentation from ASCII definitions.
  • How to include builtins (integers and Booleans) into your language.
  • How to define derived language constructs.

This folder contains several lessons, each adding new features to LAMBDA.

Syntax Modules and Basic K Commands

Here we define our first K module, which contains the initial syntax of the
LAMBDA language, and learn how to use the basic K commands.

Let us create an empty working folder, and open a terminal window
(to the left) and an editor window (to the right). We will edit our K
definition in the right window in a file called lambda.k, and will call
the K tool commands in the left window.

Let us start by defining a K module, containing the syntax of LAMBDA.

K modules are introduced with the keywords module ... endmodule.

The keyword syntax adds new productions to the syntax grammar, using a
BNF-like notation.

Terminals are enclosed in double-quotes, like strings.

You can define multiple productions for the same non-terminal in the same
syntax declaration using the | separator.

Productions can have attributes, which are enclosed in square brackets.

The attribute left tells the parser that we want the lambda application to be
left associative. For example, a b c d will then parse as (((a b) c) d).

The attribute bracket tells the parser to not generate a node for the
parenthesis production in the abstract syntax trees associated to programs.
In other words, we want to allow parentheses to be used for grouping, but we
do not want to bother to give them their obvious (ignore) semantics.

In our variant of lambda calculus defined here, identifiers and lambda
abstractions are meant to be irreducible, that is, are meant to be values.
However, so far Val is just another non-terminal, just like Exp,
without any semantic meaning. It will get a semantic meaning later.

After we are done typing our definition in the file lambda.k, we can kompile
it with the command:

kompile lambda.k

If we get no errors then a parser has been generated. This parser will be
called from now on by default by the krun tool. To see whether and how the
parser works, we are going to write some LAMBDA programs and store them in
files with the extension .lambda.

Let us create a file identity.lambda, which contains the identity lambda
abstraction:

lambda x . x

Now let us call krun on identity.lambda:

krun identity.lambda

Make sure you call the krun command from the folder containing your language
definition (otherwise type krun --help to learn how to pass a language
definition as a parameter to krun). The krun command produces the output:

<k>
  lambda x . x
</k>

If you see such an output it means that your program has been parsed (and then
pretty printed) correctly. If you want to see the internal abstract syntax
tree (AST) representation of the parsed program, which we call the K AST, then
type kast in the command instead of krun:

kast identity.lambda

You should normally never need to see this internal representation in your
K definitions, so do not get scared (yes, it is ugly for humans, but it is
very convenient for tools).

Note that krun placed the program in a <k> ... </k> cell. In K, computations
happen only in cells. If you do not define a configuration in your definition,
like we did here, then a configuration will be created automatically for you
which contains only one cell, the default k cell, which holds the program.

Next, let us create a file free-variable-capture.lambda, which contains an
expression which, in order to execute correctly in a substitution-based
semantics of LAMBDA, the substitution operation needs to avoid
variable-capture:

a (((lambda x.lambda y.x) y) z)

Next, file closed-variable-capture.lambda shows an expression which also
requires a capture-free substitution, but this expression is closed (that is,
it has no free variables) and all its bound variables are distinct (I believe
this is the smallest such expression):

(lambda z.(z z)) (lambda x.lambda y.(x y))

Finally, the file omega.lambda contains the classic omega combinator
(or closed expression), which is the smallest expression which loops forever
(not now, but after we define the semantics of LAMBDA):

(lambda x.(x x)) (lambda x.(x x))

Feel free to define and parse several other LAMBDA programs to get a feel for
how the parser works. Parse also some incorrect programs, to see how the
parser generates error messages.

In the next lesson we will see how to define semantic rules that iteratively
rewrite expressions over the defined syntax until they evaluate to a result.
This way, we obtain our first programming language defined using K.

Go to Lesson 2, LAMBDA: Module Importing, Rules, Variables

MOVIE (out of date) [4'07"]

Module Importing, Rules, Variables

We here learn how to include a predefined module (SUBSTITUTION), how to
use it to define a K rule (the characteristic rule of lambda calculus),
and how to make proper use of variables in rules.

Let us continue our lambda.k definition started in the previous lesson.

The requires keyword takes a .k file containing language features that
are needed for the current definition, which can be found in the
k-distribution/include/kframework/builtin folder. Thus, the command

requires "substitution.k"

says that the subsequent definition of LAMBDA needs the generic substitution,
which is predefined in file substitution.k under the folder
k-distribution/include/kframework/builtin. Note that substitution can be defined itself in K,
although it uses advanced features that we have not discussed yet in this
tutorial, so it may not be easy to understand now.

Using the imports keyword, we can now modify LAMBDA to import the module
SUBSTITUTION, which is defined in the required substitution.k file.

Now we have all the substitution machinery available for our definition.
However, since our substitution is generic, it cannot know which language
constructs bind variables, and what counts as a variable; however, this
information is critical in order to correctly solve the variable capture
problem. Thus, you have to tell the substitution that your lambda construct
is meant to be a binder, and that your Id terms should be treated as variables
for substitution. The former is done using the attribute binder.
By default, binder binds all the variables occurring anywhere in the first
argument of the corresponding syntactic construct within its other arguments;
you can configure which arguments are bound where, but that will be discussed
in subsequent lectures. To tell K which terms are meant to act as variables
for binding and substitution, we have to explicitly subsort the desired syntactic
categories to the builtin KVariable sort.

Now we are ready to define our first K rule. Rules are introduced with the
keyword rule and make use of the rewrite symbol, =>. In our case,
the rule defines the so-called lambda calculus beta-reduction, which
makes use of substitution in its right-hand side, as shown in lambda.k.

By convention, variables that appear in rules start with a capital letter
(the current implementation of the K tool may even enforce that).

Variables may be explicitly tagged with their syntactic category (also called
sort). If tagged, the matching term will be checked at run-time for
membership to the claimed sort. If not tagged, then no check will be made.
The former is safer, but involves the generation of a side condition to the
rule, so the resulting definition may execute slightly slower overall.

In our rule in lambda.k we tagged all variables with their sorts, so we chose
the safest path. Only the V variable really needs to be tagged there,
because we can prove (using other means, not the K tool, as the K tool is not
yet concerned with proving) that the first two variables will always have the
claimed sorts whenever we execute any expression that parses within our
original grammar.

Let us compile the definition and then run some programs. For example,

krun closed-variable-capture.lambda

yields the output

<k>
  lambda y . ((lambda x . (lambda y . (x  y))) y)
</k> 

Notice that only certain programs reduce (some even yield non-termination,
such as omega.lambda), while others do not. For example,
free-variable-capture.lambda does not reduce its second argument expression
to y, as we would expect. This is because the K rewrite rules between syntactic
terms do not apply anywhere they match. They only apply where they have been
given permission to apply by means of appropriate evaluation strategies of language
constructs, which is done using strictness attributes, evaluation contexts,
heating/cooling rules, etc., as discussed in the next lessons.

The next lesson will show how to add LAMBDA the desired evaluation strategies
using strictness attributes.

Go to Lesson 3, LAMBDA: Evaluation Strategies using Strictness

MOVIE (out of date) [4'03"]

Evaluation Strategies using Strictness

Here we learn how to use the K strict attribute to define desired evaluation
strategies. We will also learn how to tell K which terms are already
evaluated, so it does not attempt to evaluate them anymore and treats them
internally as results of computations.

Recall from the previous lecture that the LAMBDA program
free-variable-capture.lambda was stuck, because K was not given permission
to evaluate the arguments of the lambda application construct.

You can use the attribute strict to tell K that the corresponding construct
has a strict evaluation strategy, that is, that its arguments need to be
evaluated before the semantics of the construct applies. The order of
argument evaluation is purposely unspecified when using strict, and indeed
the K tool allows us to detect all possible non-deterministic behaviors that
result from such intended underspecification of evaluation strategies. We will
learn how to do that when we define the IMP language later in this tutorial;
we will also learn how to enforce a particular order of evaluation.

In order for the above strictness declaration to work effectively and
efficiently, we need to tell the K tool which expressions are meant to be
results of computations, so that it will not attempt to evaluate them anymore.
One way to do it is to make Val a syntactic subcategory of the builtin
KResult syntactic category. Since we use the same K parser to also parse
the semantics, we use the same syntax keyword to define additional syntax
needed exclusively for the semantics (like KResults). See lambda.k.

Compile again and then run some programs. They should all work as expected.
In particular, free-variable-capture.lambda now evaluates to a y.

We now got a complete and working semantic definition of call-by-value
lambda-calculus. While theoretically correct, our definition is not
easy to use and disseminate. In the next lessons we will learn how to
generate formatted documentation for LAMBDA and how to extend LAMBDA
in order to write human readable and interesting programs.

Go to Lesson 4, LAMBDA: Generating Documentation; Latex Attributes.

MOVIE (out of date) [2'20"]

Generating Documentation; Latex Attributes

In this lesson we learn how to generate formatted documentation from K
language definitions. We also learn how to use Latex attributes to control
the formatting of language constructs, particularly of ones which have a
mathematical flavor and we want to display accordingly.

To enhance readability, we may want to replace the keyword lambda by the
mathematical lambda symbol in the generated documentation. We can control
the way we display language constructs in the generated documentation
by associating them Latex attributes.

This is actually quite easy. All we have to do is to associate a latex
attribute to the production defining the construct in question, following
the Latex syntax for defining new commands (or macros).

In our case, we associate the attribute latex(\lambda{#1}.{#2}) to the
production declaring the lambda abstraction (recall that in Latex, #n refers
to the n-th argument of the defined new command).

We will later see, in Lesson 9, that we can add arbitrarily complex Latex
comments and headers to our language definitions, which give us maximum
flexibility in formatting our language definitions.

Now we have a simple programming language, with a nice documentation. However,
it is not easy to write interesting programs in this language. Almost all
programming languages build upon existing data-types and libraries. The K
tool provides a few of these (and you can add more).

In the next lesson we show how we can add builtin integers and Booleans to
LAMBDA, so we can start to evaluate meaningful expressions.

Go to Lesson 5, LAMBDA: Adding Builtins; Side Conditions.

MOVIE (out of date) [3'13"]

Adding Builtins; Side Conditions

We have already added the builtin identifiers (sort Id) to LAMBDA expressions,
but those had no operations on them. In this lesson we add integers and
Booleans to LAMBDA, and extend the builtin operations on them into
corresponding operations on LAMBDA expressions. We will also learn how to add
side conditions to rules, to limit the number of instances where they can
apply.

The K tool provides several builtins, which are automatically included in all
definitions. These can be used in the languages that we define, typically by
including them in the desired syntactic categories. You can also define your
own builtins in case the provided ones are not suitable for your language
(e.g., the provided builtin integers and operations on them are arbitrary
precision).

For example, to add integers and Booleans as values to our LAMBDA, we have to
add the productions

syntax Val ::= Int | Bool

Int and Bool are the nonterminals that correspond to these builtins.

To make use of these builtins, we have to add some arithmetic operation
constructs to our language. We prefer to use the conventional infix notation
for these, and the usual precedences (i.e., multiplication and division bind
tighter than addition, which binds tighter than relational operators).
Inspired from SDF, we use > instead of
| to state that all the previous constructs bind tighter than all the
subsequent ones. See lambda.k.

The only thing left is to link the LAMBDA arithmetic operations to the
corresponding builtin operations, when their arguments are evaluated.
This can be easily done using trivial rewrite rules, as shown in lambda.k.
In general, the K tool attempts to uniformly add the corresponding builtin
name as a suffix to all the operations over builtins. For example, the
addition over integers is an infix operation named +Int.

Compile the new lambda.k definition and evaluate some simple arithmetic
expressions. For example, if arithmetic.lambda is (1+2*3)/4 <= 1, then

krun arithmetic.lambda

yields, as expected, true. Note that the parser took the desired operation
precedence into account.

Let us now try to evaluate an expression which performs a wrong computation,
namely a division by zero. Consider the expression arithmetic-div-zero.lambda
which is 1/(2/3). Since division is strict and 2/3 evaluates to 0, this
expression reduces to 1/0, which further reduces to 1 /Int 0 by the rule for
division, which is now stuck (with the current back-end to the K tool).

In fact, depending upon the back-end that we use to execute K definitions and
in particular to evaluate expressions over builtins, 1 /Int 0 can evaluate to
anything. It just happens that the current back-end keeps it as an
irreducible term. Other K back-ends may reduce it to an explicit error
element, or issue a segmentation fault followed by a core dump, or throw an
exception, etc.

To avoid requesting the back-end to perform an illegal operation, we may use a
side condition in the rule of division, to make sure it only applies when the
denominator is non-zero.

Like in other operational formalisms, the role of the K side
conditions is to filter the number of instances of the rule. The notion
of a side condition comes from logics, where a sharp distinction is made
between a side condition (cheap) and a premise (expensive). Premises are
usually resolved using further (expensive) logical derivations, while side
conditions are simple (cheap) conditions over the rule meta-variables within
the underlying mathematical domains (which in K can be extended by the user,
as we will see in future lessons). Regarded as a logic, K derives rewrite
rules from other rewrite rules; therefore, the K side conditions cannot
contain other rewrites in them (using =>). This contrasts other rewrite
engines, for example Maude, which
allow conditional rules with rewrites in conditions.
The rationale behind this deliberate restriction in K is twofold:

  • On the one hand, general conditional rules require a complex, and thus slower
    rewrite engine, which starts recursive (sometimes exhaustive) rewrite sessions
    to resolve the rewrites in conditions. In contrast, the side conditions in K
    can be evaluated efficiently by back-ends, for example by evaluating builtin
    expressions and/or by calling builtin functions.
  • On the other hand, the semantic definitional philosophy of K is that rule
    premises are unnecessary, so there is no need to provide support for them.

Having builtin arithmetic is useful, but writing programs with just lambda
and arithmetic constructs is still a pain. In the next two lessons we will
add conditional (if_then_else) and binding (let and letrec) constructs,
which will allow us to write nicer programs.

Go to Lesson 6, LAMBDA: Selective Strictness; Anonymous Variables.

MOVIE (out of date) [4'52"]

Selective Strictness; Anonymous Variables

We here show how to define selective strictness of language constructs,
that is, how to state that certain language constructs are strict only
in some arguments. We also show how to use anonymous variables.

We next define a conditional if construct, which takes three arguments,
evaluates only the first one, and then reduces to either the second or the
third, depending on whether the first one evaluated to true or to false.

K allows to define selective strictness using the same strict attribute,
but passing it a list of numbers. The numbers correspond to the arguments
in which we want the defined construct to be strict. In our case,

syntax Exp ::= "if" Exp "then" Exp "else" Exp   [strict(1)]

states that the conditional construct is strict in the first argument.

We can now assume that its first argument will eventually reduce to a value, so
we only write the following two semantic rules:

rule if true  then E else _ => E
rule if false then _ else E => E

Thus, we assume that the first argument evaluates to either true or false.

Note the use of the anonymous variable _. We use such variables purely for
structural reasons, to state that something is there but we don't care what.
An anonymous variable is therefore completely equivalent to a normal variable
which is unsorted and different from all the other variables in the rule. If
you use _ multiple times in a rule, they will all be considered distinct.

Compile lambda.k and write and execute some interesting expressions making
use of the conditional construct. For example, the expression

if 2<=1 then 3/0 else 10

evaluates to 10 and will never evaluate 3/0, thus avoiding an unwanted
division-by-zero.

In the next lesson we will introduce two new language constructs, called
let and letrec and conventionally found in functional programming
languages, which will allow us to already write interesting LAMBDA programs.

Go to Lesson 7, LAMBDA: Derived Constructs; Extending Predefined Syntax.

MOVIE (out of date) [2'14"]

Derived Constructs, Extending Predefined Syntax

In this lesson we will learn how to define derived language constructs, that
is, ones whose semantics is defined completely in terms of other language
constructs. We will also learn how to add new constructs to predefined
syntactic categories.

When defining a language, we often want certain language constructs to be
defined in terms of other constructs. For example, a let-binding construct
of the form

let x = e in e'

is nothing but syntactic sugar for

(lambda x . e') e

This can be easily achieved with a rule, as shown in lambda.k.

Compile lambda.k and write some programs using let binders.

For example, consider a lets.lambda program which takes arithmetic.lambda
and replaces each integer by a let-bound variable. It should evaluate to
true, just like the original arithmetic.lambda.

Let us now consider a more interesting program, namely one that calculates the
factorial of 10:

let f = lambda x . (
        (lambda t . lambda x . (t t x))
        (lambda f . lambda x . (if x <= 1 then 1 else (x * (f f (x + -1)))))
        x
      )
in (f 10)

This program follows a common technique to define fixed points in untyped
lambda calculus, based on passing a function to itself.

We may not like to define fixed-points following the approach above, because
it requires global changes in the body of the function meant to be recursive,
basically to pass it to itself (f f in our case above). The approach below
isolates the fixed-point aspect of the function in a so-called fixed-point
combinator
, which we call fix below, and then apply it to the function
defining the body of the factorial, without any changes to it:

let fix = lambda f . (
          (lambda x . (f (lambda y . (x x y))))
          (lambda x . (f (lambda y . (x x y))))
        )
in let f = fix (lambda f . lambda x .
                (if x <= 1 then 1 else (x * (f (x + -1)))))
   in (f 10)

Although the above techniques are interesting and powerful (indeed, untyped
lambda calculus is in fact Turing complete), programmers will probably not
like to write programs this way.

We can easily define a more complex derived construct, called letrec and
conventionally encountered in functional programming languages, whose semantics
captures the fixed-point idea above. In order to keep its definition simple
and intuitive, we define a simplified variant of letrec, namely one which only
allows to define one recursive one-argument function. See lambda.k.

There are two interesting observations here.

First, note that we have already in-lined the definition of the fix
combinator in the definition of the factorial, to save one application of the
beta reduction rule (and the involved substitution steps). We could have
in-lined the definition of the remaining let, too, but we believe that the
current definition is easier to read.

Second, note that we extended the predefined Id syntactic category with two
new constants, $x and $y. The predefined identifiers cannot start with
$, so programs that will be executed with this semantics cannot possibly
contain the identifiers xandx andy. In other words, by adding them to Id they
become indirectly reserved for the semantics. This is indeed desirable,
because any possible uses of xinthebodyofthefunctiondefinedusingletrecwouldbecapturedbythelambdax in the body of the function defined using `letrec` would be captured by the `lambdaxdeclaration in the definition ofletrec`.

Using letrec, we can now write the factorial program as elegantly as it can
be written in a functional language:

letrec f x = if x <= 1 then 1 else (x * (f (x + -1)))
in (f 10)

In the next lesson we will discuss an alternative definition of letrec, based
on another binder, mu, specifically designed to define fixed points.

Go to Lesson 8, LAMBDA: Multiple Binding Constructs.

MOVIE (out of date) [5'10"]

Multiple Binding Constructs

Here we learn how multiple language constructs that bind variables can
coexist. We will also learn about or recall another famous binder besides
lambda, namely mu, which can be used to elegantly define all kinds of
interesting fixed-point constructs.

The mu binder has the same syntax as lambda, except that it replaces
lambda with mu.

Since mu is a binder, in order for substitution to know how to deal with
variable capture in the presence of mu, we have to tell it that mu is a
binding construct, same like lambda. We take advantage of being there and
also add mu its desired latex attribute.

The intuition for

mu x . e

is that it reduces to e, but each free occurrence of x in e behaves
like a pointer that points back to mu x . e.

With that in mind, let us postpone the definition of mu and instead redefine
letrec F X = E in E' as a derived construct, assuming mu available. The
idea is to simply regard F as a fixed-point of the function

lambda X . E

that is, to first calculate

mu F . lambda X . E

and then to evaluate E' where F is bound to this fixed-point:

let F = mu F . lambda X . E in E'

This new definition of letrec may still look a bit tricky, particularly
because F is bound twice, but it is much simpler and cleaner than our
previous definition. Moreover, now it is done in a type-safe manner
(this aspect goes beyond our objective in this tutorial).

Let us now define the semantic rule of mu.

The semantics of mu is actually disarmingly simple. We just have to
substitute mu X . E for each free occurrence of X in E:

mu X . E => E[(mu X . E) / X]

Compile lambda.k and execute some recursive programs. They should be now
several times faster. Write a few more recursive programs, for example ones
for calculating the Ackermann function, for calculating the number of moves
needed to solve the Hanoi tower problem, etc.

We have defined our first programming language in K, which allows us to
write interesting functional programs. In the next lesson we will learn how
to fully document our language definition, in order to disseminate it, to ship
it to colleagues or friends, to publish it, to teach it, and so on.

Go to Lesson 9, LAMBDA: A Complete and Commented Definition.

MOVIE (out of date) [2'40"]

A Complete and Documented K Definition

In this lesson you will learn how to add formal comments to your K definition,
in order to nicely document it. The generated document can be then used for
various purposes: to ease understanding the K definition, to publish it,
to send it to others, etc.

The K tool allows a literate programming style, where the executable
language definition can be documented by means of annotations. One such
annotation is the latex(_) annotation, where you can specify how to format
the given production when producing Latex output via the --output latex
option to krun, kast, and kprove.

There are three types of comments, which we discuss next.

Ordinary comments

These use // or /* ... */, like in various programming languages. These
comments are completely ignored.

Document annotations

Use the @ symbol right after // or /* in order for the comment to be
considered an annotation and thus be processed by the K tool when it
generates documentation.

As an example, we can go ahead and add such an annotation at the beginning
of the LAMBDA module, explaining how we define the syntax of this language.

Header annotations

Use the ! symbol right after // or /* if you want the comment to be
considered a header annotation, that is, one which goes before
\begin{document} in the generated Latex. You typically need header
annotations to include macros, or to define a title, etc.

As an example, let us set a Latex length and then add a title and an
author to this K definition.

Compile the documentation and take a look at the results. Notice the title.

Feel free to now add lots of annotations to lambda.k.

Then compile and check the result. Depending on your PDF viewer, you
may also see a nice click-able table of contents, with all the sections
of your document. This could be quite convenient when you define large
languages, because it helps you jump to any part of the semantics.

Tutorial 1 is now complete. The next tutorial will take us through the
definition of a simple imperative language and will expose us to more
feature of the K framework and the K tool.

MOVIE (out of date) [6'07"]

Part 2: Defining IMP

Here you will learn how to define a very simple imperative language in K
and the basics of how to work with configurations, cells, and computations.
Specifically, you will learn the following:

  • How to define languages using multiple modules.
  • How to define sequentially strict syntactic constructs.
  • How to use K's syntactic lists.
  • How to define, initialize and configure configurations.
  • How the language syntax is swallowed by the builtin K syntactic category.
  • The additional syntax of the K syntactic category.
  • How the strictness annotations are automatically desugared into rules.
  • The first steps of the configuration abstraction mechanism.

Like in the previous tutorial, this folder contains several lessons, each
adding new features to IMP. Do them in order. Also, make sure you completed
and understood the previous tutorial.

Defining a More Complex Syntax

Here we learn how to define a more complex language syntax than LAMBDA's,
namely the C-like syntax of IMP. Also, we will learn how to define languages
using multiple modules, because we are going to separate IMP's syntax from
its semantics using modules. Finally, we will also learn how to use K's
builtin support for syntactic lists.

The K tool provides modules for grouping language features. In general, we
can organize our languages in arbitrarily complex module structures.
While there are no rigid requirements or even guidelines for how to group
language features in modules, we often separate the language syntax from the
language semantics in different modules.

In our case here, we start by defining two modules, IMP-SYNTAX and IMP, and
import the first in the second, using the keyword imports. As their names
suggest, we will place all IMP's syntax definition in IMP-SYNTAX and all its
semantics in IMP.

Note, however, that K does no more than simply includes all the
contents of the imported module in the one which imports it (making sure
that everything is only kept once, even if you import it multiple times).
In other words, there is currently nothing fancy in K tool's module system.

IMP has six syntactic categories, as shown in imp.k: AExp for arithmetic
expressions, BExp for Boolean expressions, Block for blocks, Stmt for
statements, Pgm for programs and Ids for comma-separated lists of
identifiers. Blocks are special statements, whose role is to syntactically
constrain the conditional statement and the while loop statement to only
take blocks as branches and body, respectively.

There is nothing special about arithmetic and Boolean expressions. They
are given the expected strictness attributes, except for <= and &&,
for demonstration purposes.

The <= is defined to be seqstrict, which means that it evaluates its
arguments in order, from left-to-right (recall that the strict operators
can evaluate their arguments in any, fully interleaved, orders). Like
strict, the seqstrict annotation can also be configured; for example, one
can specify in which arguments and in what order. By default, seqstrict
refers to all the arguments, in their left-to-right order. In our case here,
it is equivalent with seqstrict(1 2).

The && is only strict in its first argument, because we will give it a
short-circuited semantics (its second argument will only be evaluated when
the first evaluates to true). Recall the K tool also allows us to associate
LaTex attributes to constructs, telling the document generator how to display
them. For example, we associate <= the attribute latex({#1}\leq{#2}),
which makes it be displayed \leq everywhere in the generated LaTex
documentation.

In this tutorial we take the freedom to associate the various constructs
parsing precedences that we have already tested and we know work well, so that
we can focus on the semantics here instead of syntax. In practice, though,
you typically need to experiment with precedences until you obtain the desired
parser.

Blocks are defined using curly brackets, and they can either be empty or
hold a statement.

Nothing special about the IMP statements. Note that ; is an assignment
statement terminator, not a statement separator. Note also that blocks are
special statements.

An IMP program declares a comma-separated list of variables using the keyword
int like in C, followed by a semicolon ;, followed by a statement.
Syntactically, the idea here is that we can wrap any IMP program within a
main(){...} function and get a valid C program. IMP does not allow variable
declarations anywhere else except through this construct, at the top-level of
the program. Other languages provided with the K distribution (see, e.g., the
IMP++ language also discussed in this tutorial) remove this top-level program
construct of IMP and add instead variable declaration as a statement construct,
which can be used anywhere in the program, not only at the top level.

Note how we defined the comma-separated list of identifiers using
List{Id,","}. The K tool provides builtin support for generic syntactic
lists. In general,

syntax B ::= List{A,T}

declares a new non-terminal, B, corresponding to T-separated sequences of
elements of A, where A is a non-terminal and T is a terminal. These
lists can also be empty, that is, IMP programs declaring no variable are also
allowed (e.g., int; {} is a valid IMP program). To instantiate and use
the K builtin lists, you should alias each instance with a (typically fresh)
non-terminal in your syntax, like we do with the Ids nonterminal.

Like with other K features, there are ways to configure the syntactic lists,
but we do not discuss them here.

Recall from Tutorial 1 (LAMBDA) that in order for strictness to work well
we also need to tell K which computations are meant to be results. We do
this as well now, in the module IMP: integers and Booleans are K results.

Kompile imp.k and test the generated parser by running some programs.
Since IMP is a fragment of C, you may want to select the C mode in your
editor when writing these programs. This will also give your the feel that
you are writing programs in a real programming language.

For example, here is sum.imp, which sums in sum all numbers up to n:

int n, sum;
n = 100;
sum=0;
while (!(n <= 0)) {
  sum = sum + n;
  n = n + -1;
}

Now krun it and see how it looks parsed in the default k cell.

The program collatz.imp tests the Collatz conjecture for all numbers up to
m and accumulates the total number of steps in s:

int m, n, q, r, s;
m = 10;
while (!(m<=2)) {
  n = m;
  m = m + -1;
  while (!(n<=1)) {
    s = s+1;
    q = n/2;
    r = q+q+1;
    if (r<=n) {
      n = n+n+n+1;         // n becomes 3*n+1 if odd
    } else {n=q;}          //        of   n/2 if even
  }
}

Finally, program primes.imp counts in s all the prime numbers up to m:

int i, m, n, q, r, s, t, x, y, z;
m = 10;  n = 2;
while (n <= m) {
  // checking primality of n and writing t to 1 or 0
  i = 2;  q = n/i;  t = 1;
  while (i<=q && 1<=t) {
    x = i;
    y = q;
    // fast multiplication (base 2) algorithm
    z = 0;
    while (!(x <= 0)) {
      q = x/2;
      r = q+q+1;
      if (r <= x) { z = z+y; } else {}
      x = q;
      y = y+y;
    } // end fast multiplication
    if (n <= z) { t = 0; } else { i = i+1;  q = n/i; }
  } // end checking primality
  if (1 <= t) { s = s+1; } else {}
  n = n+1;
}

All the programs above will run once we define the semantics of IMP. If you
want to execute them now, wrap them in a main(){...} function and compile
them and run them with your favorite C compiler.

Before we move to the K semantics of IMP, we would like to make some
clarifications regarding the K builtin parser, kast. Although it is quite
powerful, you should not expect magic from it! While the K parser can parse
many non-trivial languages (see, for example, the KOOL language in
pl-tutorial/2_languages) in the K distribution), it was
never meant to be a substitute for real parsers. We often call the syntax
defined in K the syntax of the semantics, to highlight the fact that its
role is to serve as a convenient notation when writing the semantics, not
necessarily as a means to define concrete syntax of arbitrarily complex
programming languages. See the KERNELC language for an example on how to connect an external parser for concrete syntax to
the K tool.

The above being said, we strongly encourage you to strive to make the
builtin parser work with your desired language syntax! Do not give up
simply because you don't want to deal with syntactic problems. On the
contrary, fight for your syntax! If you really cannot define your desired
syntax because of tool limitations, we would like to know. Please tell us.

Until now we have only seen default configurations. In the next lesson we
will learn how to define a K custom configuration.

Go to Lesson 2, IMP: Defining a Configuration.

MOVIE (out of date) [09'15"]

Defining a Configuration

Here we learn how to define a configuration in K. We also learn how to
initialize and how to display it.

As explained in the overview presentation on K, configurations are quite
important, because all semantic rules match and apply on them.
Moreover, they are the backbone of configuration abstraction, which allows
you to only mention the relevant cells in each semantic rule, the rest of
the configuration context being inferred automatically. The importance of
configuration abstraction will become clear when we define more complex
languages (even in IMP++). IMP does not really need it. K configurations
are constructed making use of cells, which are labeled and can be arbitrarily
nested.

Configurations are defined with the keyword configuration. Cells are
defined using an XML-ish notation stating clearly where the cell starts
and where it ends.

While not enforced by the tool, we typically like to put the entire
configuration in a top-level cell, called T. So let's define it:

configuration <T>...</T>

Cells can have other cells inside. In our case of IMP, we need a cell to
hold the remaining program, cell which we typically call k, and a cell to
hold the program state. Let us add them:

configuration <T> <k>...</k> <state>...</state> </T>

K allows us to also specify how to initialize a configuration at the same
time with declaring the configuration. All we have to do is to fill in
the contents of the cells with some terms. The syntactic categories of
those terms will also indirectly define the types of the corresponding
cells.

For example, we want the k cell to initially hold the program that is passed
to krun. K provides a builtin configuration variable, called $PGM, which
is specifically designed for this purpose: krun will place its program there
(after it parses it, or course). The K tool allows users to define their own
configuration variables, too, which can be used to develop custom
initializations of program configurations with the help of krun; this can be
quite useful when defining complex languages, but we do not discuss it in
this tutorial.

configuration <T> <k> $PGM </k> <state>...</state>  </T>

Moreover, we want the program to be a proper Pgm term (because we do not
want to allow krun to take fragments of programs, for example, statements).
Therefore, we tag $PGM with the desired syntactic category, Pgm:

configuration <T> <k> $PGM:Pgm </k> <state>...</state>  </T>

Like for other variable tags in K, a run-time check will be performed and the
semantics will get stuck if the passed term is not a well-formed program.

We next tell K that the state cell should be initialized with the empty map:

configuration <T> <k> $PGM:Pgm </k> <state> .Map </state>  </T>

Recall that in K . stands for nothing. However, since there are various
types of nothing, to avoid confusion we can suffix the . with its desired
type. K has several builtin data-types, including lists, sets, bags, and
maps. .Map is the empty map.

Kompile imp.k and run several programs to see how the configuration is
initialized as desired.

When configurations get large, and they do when defining large programming
languages, you may want to color the cells in order to more easily distinguish
them. This can be easily achieved using the color cell attribute, following
again an XML-ish style:

configuration <T color="yellow">
                <k color="green"> $PGM:Pgm </k>
                <state color="red"> .Map </state>
              </T>

In the next lesson we will learn how to write rules that involve cells.

Go to Lesson 3, IMP: Computations, Results, Strictness; Rules Involving Cells.

MOVIE (out of date) [04'21"]

Computations, Results, Strictness; Rules Involving Cells

In this lesson we will learn about the syntactic category K of computations,
about how strictness attributes are in fact syntactic sugar for rewrite rules
over computations, and why it is important to tell the tool which
computations are results. We will also see a K rule that involves cells.

K Computations

Computation structures, or more simply computations, extend the abstract
syntax of your language with a list structure using ~> (read followed
by
or and then, and written \curvearrowright in Latex) as a separator.
K provides a distinguished sort, K, for computations. The extension of the
abstract syntax of your language into computations is done automatically by
the K tool when you declare constructs using the syntax keyword, so the K
semantic rules can uniformly operate only on terms of sort K. The intuition
for computation structures of the form

t1 ~> t2 ~> ... ~> tn

is that the listed tasks are to be processed in order. The initial
computation typically contains the original program as its sole task, but
rules can then modify it into task sequences, as seen shortly.

Strictness in Theory

The strictness attributes, used as annotations to language constructs,
actually correspond to rules over computations. For example, the
strict(2) attribute of the assignment statement corresponds to the
following two opposite rules (X ranges over Id and A over AExp):

X=A; => A ~> X=[];
A ~> X=[]; => X=A;

The first rule pulls A from the syntactic context X=A; and schedules it
for processing. The second rule plugs A back into its context.
Inspired from the chemical abstract machine, we call rules of the first
type above heating rules and rules of the second type cooling rules.
Similar rules are generated for other arguments in which operations are
strict. Iterative applications of heating rules eventually bring to the
top of the computation atomic tasks, such as a variable lookup, or a
builtin operation, which then make computational progress by means of other
rules. Once progress is made, cooling rules can iteratively plug the result
back into context, so that heating rules can pick another candidate for
reduction, and so on and so forth.

When operations are strict only in some of their arguments, the corresponding
positions of the arguments in which they are strict are explicitly enumerated
in the argument of the strict attribute, e.g., strict(2) like above, or
strict(2 3) for an operation strict in its second and third arguments, etc.
If an operation is simply declared strict then it means that it is strict
in all its arguments. For example, the strictness of addition yields:

A1+A2 => A1 ~> []+A2
A1 ~> []+A2 => A1+A2
A1+A2 => A2 ~> A1+[]
A2 ~> A1+[] => A1+A2

It can be seen that such heating/cooling rules can easily lead to
non-determinism, since the same term may be heated many different ways;
these different evaluation orders may lead to different behaviors in some
languages (not in IMP, because its expressions do not have side effects,
but we will experiment with non-determinism in its successor, IMP++).

A similar desugaring applies to sequential strictness, declared with the
keyword seqstrict. While the order of arguments of strict is irrelevant,
it matters in the case of seqstrict: they are to be evaluated in the
specified order; if no arguments are given, then they are assumed by default
to be evaluated from left-to-right. For example, the default heating/cooling
rules associated to the sequentially strict <= construct above are
(A1, A2 range over AExp and I1 over Int):

A1<=A2 => A1 ~> []<=A2
A1 ~> []<=A2 => A1<=A2
I1<=A2 => A2 ~> I1<=[]
A2 ~> I1<=[] => I1<=A2

In other words, A2 is only heated/cooled after A1 is already evaluated.

While the heating/cooling rules give us a nice and uniform means to define
all the various allowable ways in which a program can evaluate, all based
on rewriting, the fact that they are reversible comes with a serious practical
problem: they make the K definitions unexecutable, because they lead to
non-termination.

Strictness in Practice; K Results

To break the reversibility of the theoretical heating/cooling rules, and,
moreover, to efficiently execute K definitions, the current implementation of
the K tool relies on users giving explicit definitions of their languages'
results.

The K tool provides a predicate isKResult, which is automatically defined
as we add syntactic constructs to KResult (in fact the K tool defines such
predicates for all syntactic categories, which are used, for example, as
rule side conditions to check user-declared variable memberships, such as
V:Val stating that V belongs to Val).

The kompile tool, depending upon what it is requested to do, changes the
reversible heating/cooling rules corresponding to evaluation strategy
definitions (e.g., those corresponding to strictness attributes) to avoid
non-termination. For example, when one is interested in obtaining an
executable model of the language (which is the default compilation mode of
kompile), then heating is performed only when the to-be-pulled syntactic
fragment is not a result, and the corresponding cooling only when the
to-be-plugged fragment is a result. In this case, e.g., the heating/cooling
rules for assignment are modified as follows:

X=A; => A ~> X=[];  requires notBool isKResult(A)
A ~> X=[]; => X=A;  requires isKResult(A)

Note that non-termination of heating/cooling is avoided now. The only thing
lost is the number of possible behaviors that a program can manifest, but
this is irrelevant when all we want is one behavior.

As will be discussed in the IMP++ tutorial, the heating/cooling rules are
modified differently by kompile when we are interested in other aspects
of the language definition, such us, for example, in a search-able model that
comprises all program behaviors. This latter model is obviously more general
from a theoretical perspective, but, in practice, it is also slower to execute.
The kompile tool strives to give you the best model of the language for the
task you are interested in.

Can't Results be Inferred Automatically?

This is a long story, but the short answer is: No!. Maybe in some cases
it is possible, but we prefer to not attempt it in the K tool. For example,
you most likely do not want any stuck computation to count as a result,
since some of them can happen simply because you forgot a semantic rule that
could have further reduce it! Besides, in our experience with defining large
languages, it is quite useful to take your time and think of what the results
of your language's computations are. This fact in itself may help you improve
your overall language design. We typically do it at the same time with
defining the evaluation strategies of our languages. Although in theory K
could infer the results of your language as the stuck computations, based on
the above we have deliberately decided to not provide this feature, in spite
of requests from some users. So you currently do have to explicitly define
your K results if you want to effectively use the K tool. Note, however, that
theoretical definitions, not meant to be executed, need not worry about
defining results (that's because in theory semantic rules apply modulo the
reversible heating/cooling rules, so results are not necessary).

A K Rule Involving Cells

All our K rules so far in the tutorial were of the form

rule left => right requires condition

where left and right were syntactic, or more generally computation, terms.

Here is our first K rule explicitly involving cells:

rule <k> X:Id => I ...</k> <state>... X |-> I ...</state>

Recall that the k cell holds computations, which are sequences of tasks
separated by ~>. Also, the state cell holds a map, which is a set of
bindings, each binding being a pair of computations (currently, the
K builtin data-structures, like maps, are untyped; or, said differently,
they are all over the type of computations, K).

Therefore, the two cells mentioned in the rule above hold collections
of things, ordered or not. The ...s, which we also call cell frames,
stand for more stuff there, which we do not care about.

The rewrite relation => is allowed in K to appear anywhere in a term, its
meaning being that the corresponding subterm is rewritten as indicated in the
shown context. We say that K's rewriting is local.

The rule above says that if the identifier X is the first task in the k
cell, and if X is bound to I somewhere in the state, then X rewrites
to I locally in the k cell. Therefore, IMP variables need to be already
declared when looked up.

Of course, the K rule above can be translated into an ordinary rewrite rule
of the form

rule <k> X ~> Rest </k> <state> Before (X |-> I) After </state>
  => <k> I ~> Rest </k> <state> Before (X |-> I) After </state>

Besides being more verbose and thus tedious to write, this ordinary rule
is also more error-prone; for example, we may forget the Rest variable
in the right-hand-side, etc. Moreover, the concurrent semantics of K
allows for its rules to be interpreted as concurrent transactions, where
the context is the read-only component of the transaction, while the
subterms which are rewritten are read/write component of the transaction;
thus, K rule instances can apply concurrently if they only overlap
on read-only parts, while they cannot if regarded as ordinary rewrite logic
rules. Note: our current implementation of the K tool is not concurrent,
so K rules are in fact desugared as normal rewrite rules in the K tool.

Kompile imp.k using a documentation option and check out how the K rule
looks in the generated document. The ... frames are displayed as cell
tears, metaphorically implying that those parts of the cells that we
do not care about are torn away. The rewrite relation is replaced by a
horizontal line: specifically, the subterm which rewrites, X, is
underlined, and its replacement is written underneath the line.

In the next lesson we define the complete K semantics of IMP and
run the programs we parsed in the first lesson.

Go to Lesson 4, IMP: Configuration Abstraction, Part 1; Types of Rules.

MOVIE (out of date) [10'30"]

Configuration Abstraction, Part 1; Types of Rules

Here we will complete the K definition of IMP and, while doing so, we will
learn the very first step of what we call configuration abstraction.

The IMP Semantic Rules

Let us add the remaining rules, in the order in which the language constructs
were defined in IMP-SYNTAX.

The rules for the arithmetic and Boolean constructs are self-explanatory.
Note, however, that K will infer the correct sorts of all the variables in
these rules, because they appear as arguments of the builtin operations
(_+Int_, etc.). Moreover, the inferred sorts will be enforced dynamically.
Indeed, we do not want to apply the rule for addition, for example, when the
two arguments are not integers. In the rules for &&, although we prefer to
not do it here for simplicity, we could have eliminated the dynamic check by
replacing B (and similarly for _) with B:K. Indeed, it can be shown
that whenever any of these rules apply, B (or _) is a BExp anyway.
That's because there is no rule that can touch such a B (or _); this
will become clearer shortly, when we discuss the first step of configuration
abstraction. Therefore, since we know that B will be a BExp anyway, we
could save the time it takes to check its sort; such times may look minor,
but they accumulate, so some designers may prefer to avoid run-time checks
whenever possible.

The block rules are trivial. However, the rule for non-empty blocks is
semantically correct only because we do not have local variable declarations
in IMP. We will have to change this rule in IMP++.

The assignment rule has two =>: one in the k cell dissolving the
assignment statement, and the other in the state cell updating the value of
the assigned variable. Note that the one in the state is surrounded by
parentheses: (_ => I). That is because => is greedy: it matches as much
as it can to the left and to the right, until it reaches the cell boundaries
(closed or open). If you want to limit its scope, or for clarity, you can use
parentheses like here.

The rule for sequential composition simply desugars S1 S2 into S1 ~> S2.
Indeed, the two have exactly the same semantics. Note that statements
evaluate to nothing (.), so once S1 is processed in S1 ~> S2, then the
next task is automatically S2, without wasting any step for the transition.

The rules for the conditional and while statements are clear. One thing to
keep in mind now is that the while unrolling rule will not apply
indefinitely in the positive branch of the resulting conditional, because
of K's configuration abstraction, which will be discussed shortly.

An IMP program declares a set of variables and then executes a
statement in the state obtained after initializing all those variables
to 0. The rules for programs initialize the declared variables one by one,
checking also that there are no duplicates. We check for duplicates only for
demonstration purposes, to illustrate the keys predefined operation that
returns the set of keys of a map, and the set membership operation in.
In practice, we typically define a static type checker for our language,
which we execute before the semantics and reject inappropriate programs.

The use of the .Ids in the second rule is not necessary. We could have
written int; S instead of int .Ids; S and the K tool would parse it and
kompile the definition correctly, because it uses the same parser used for
parsing programs also to parse the semantics. However, we typically prefer to
explicitly write the nothing values in the semantics, for clarity;
the parser has been extended to accept these. Note that the first rule
matches the entire k cell, because int_;_ is the top-level program
construct in IMP, so there is nothing following it in the computation cell.
The anonymous variable stands for the second argument of this top-level program
construct, not for the rest of the computation. The second rule could have
also been put in a complete k cell, but we preferred not to, for simplicity.

Our IMP semantics is now complete, but there are a few more things that we
need to understand and do.

Configuration Abstraction, Part 1

First, let us briefly discuss the very first step of configuration abstraction.
In K, all semantic rules are in fact rules between configurations. As soon
explained in the IMP++ tutorial, the declared configuration cell structure is
used to automatically complete the missing configuration parts in rules.
However, many rules do not involve any cells, being rules between syntactic
terms (of sort K); for example, we had only three rules involving cells in our
IMP semantics. In this case, the k cell will be added automatically and the
actual rewrite will happen on top of the enclosed computation. For example,
the rule for the while loop is automatically translated into the following:

rule <k> while (B) S => if (B) {S while (B) S} else {} ...</k>

Since the first task in computations is what needs to be done next, the
intuition for this rule completion is that the syntactic transition
only happens when the term to rewrite is ready for processing. This explains,
for example, why the while loop unrolling does not indefinitely apply in the
positive branch of the conditional: the inner while loop is not ready for
evaluation yet. We call this rule completion process, as well as other
similar ones, configuration abstraction. That is because the incomplete
rule abstracts away the configuration structure, thus being easier to read.
As seen soon when we define IMP++, configuration abstraction is not only a
user convenience; it actually significantly increases the modularity of our
definitions. The k-cell-completion is only the very first step, though.

If you really want certain rewrites over syntactic terms to apply
anywhere they match, then you should tag the rule with the attribute
anywhere, which was discussed in Tutorial 1, Lesson 2.5.

Kompile and then krun the programs that you only parsed in Lesson 1. They
should all execute as expected. The state cell shows the final state
of the program. The k cell shows the final code contents, which should be
empty whenever the IMP program executes correctly.

Kompile also with the documentation option and take a look at the generated
documentation. The assignment rule should particularly be of interest,
because it contains two local rewrites.

In the next lesson we comment the IMP definition and conclude this tutorial.

Go to Lesson 5, IMP: Completing and Documenting IMP.

MOVIE (out of date) [09'16"]

Completing and Documenting IMP

We here learn no new concepts, but it is a good moment to take a break
and contemplate what we learned so far.

Let us add lots of formal annotations to imp.k.

Once we are done with the annotations, we kompile with the documentation
option and then take a look at the produced document. We often call these
documents language posters. Depending on how much information you add to
these language posters, they can serve as standalone, formal presentations
of your languages. For example, you can print them as large posters and
post them on the wall, or in poster sessions at conferences.

This completes our second tutorial. The next tutorials will teach us more
features of the K framework, such as how to define languages with complex
control constructs (like callcc), languages which are concurrent, and so on.

MOVIE (out of date) [03'45"]

Part 3: Defining LAMBDA++

Here you will learn how to define language constructs which abruptly change
the execution control flow, and how to define language semantics following
and environment/store style. Specifically, you will learn the following:

  • How to define constructs like callcc, which allow you to take snapshots of
    program executions and to go back in time at any moment.
  • How to define languages in an environment/store style.
  • Some basic notions about the use of closures and closure-like semantic
    structures to save and restore execution environments.
  • Some basic intuitions about reusing existing semantics in new languages,
    as well as some of the pitfalls in doing so.

Abrupt Changes of Control

Here we add call-with-current-continuation (callcc) to the definition of
LAMBDA completed in Tutorial 1, and call the resulting language LAMBDA++.
While doing so, we will learn how to define language constructs that
abruptly change the execution control flow.

Take over the lambda.k definition from Lesson 8 in Part 1 of this Tutorial,
which is the complete definition of the LAMBDA language, but without the
comments.

callcc is a good example for studying the capabilities of a framework to
support abrupt changes of control, because it is one of the most
control-intensive language constructs known. Scheme is probably the first
programming language that incorporated the callcc construct, although
similar constructs have been recently included in many other languages in
one form or another.

Here is a quick description: callcc e passes the remaining computation
context, packaged as a function k, to e (which is expected to be a function);
if during its evaluation e passes any value to k, then the current
execution context is discarded and replaced by the one encoded by k and
the value is passed to it; if e evaluates normally to some value v and
passes nothing to k in the process, then v is returned as a result of
callcc e and the execution continues normally. For example, we want the
program callcc-jump.lambda:

(callcc (lambda k . ((k 5) + 2))) + 10

to evaluate to 15, not 17! Indeed, the computation context [] + 10 is
passed to callcc's argument, which then sends it a 5, so the computation
resumes to 5 + 10. On the other hand, the program callcc-not-jump.lambda

(callcc (lambda k . (5 + 2))) + 10

evaluates to 17.

If you like playing games, you can metaphorically think of callcc e as
saving your game state in a file and passing it to your friend e.
Then e can decide at some moment to drop everything she was doing, load
your game and continue to play it from where you were.

The behavior of many popular control-changing constructs can be obtained
using callcc. The program callcc-return.lambda shows, for example, how to
obtain the behavior of a return statement, which exits the current execution
context inside a function and returns a value to the caller's context:

letrec f x = callcc (lambda return . (
  f (if (x <= 0) then ((return 1) / 0) else 2)
))
in (f -3)

This should evaluate to 1, in spite of the recursive call to f
and of the division by zero! Note that return is nothing but a variable
name, but one which is bound to the current continuation at the beginning of
the function execution. As soon as 1 is passed to return, the computation
jumps back in time to where callcc was defined! Change -3 to 3 and the
program will loop forever.

callcc is quite a powerful and beautiful language construct, although one
which is admittedly hard to give semantics to in some frameworks.
But not in K 😃 Here is the entire K syntax and semantics of callcc:

syntax Exp ::= "callcc" Exp  [strict]
syntax Val ::= cc(K)
rule <k> (callcc V:Val => V cc(K)) ~> K </k>
rule <k> cc(K) V ~> _ =>  V ~> K </k>

Let us first discuss the annotated syntax. We declared callcc strict,
because its argument may not necessarily be a function yet, so it may need
to be evaluated. As explained above, we need to encode the remaining
computation somehow and pass it to callcc's argument. More specifically,
since LAMBDA is call-by-value, we have to encode the remaining computation as
a value. We do not want to simply subsort computations to Val, because there
are computations which we do not want to be values. A simple solution to
achieve our goal here is to introduce a new value construct, say cc (from
current-continuation), which holds any computation.

Note that, inspired from SDF,
K allows you to define the syntax of helping semantic operations, like cc,
more compactly. Typically, we do not need a fancy syntax for such operators;
all we need is a name, followed by open parenthesis, followed by a
comma-separated list of arguments, followed by closed parenthesis. If this
is the syntax that you want for a particular construct, then K allows you to
drop all the quotes surrounding the terminals, as we did above for cc.

The semantic rules do exactly what the English semantics of callcc says.
Note that here, unlike in our definition of LAMBDA in Tutorial 1, we had
to mention the cell <k/> in our rules. This is because we need to make sure
that we match the entire remaining computation, not only a fragment of it!
For example, if we replace the two rules above with

rule (callcc V:Val => V cc(K)) ~> K
rule cc(K) V ~> _ =>  V ~> K

then we get a callcc which is allowed to non-deterministically pick a
prefix of the remaining computation and pass it to its argument, and then
when invoked within its argument, a non-deterministic prefix of the new
computation is discarded and replaced by the saved one. Wow, that would
be quite a language! Would you like to write programs in it? 😃

Consequently, in K we can abruptly change the execution control flow of a
program by simply changing the contents of the <k/> cell. This is one of
the advantages of having an explicit representation of the execution context,
like in K or in reduction semantics with evaluation contexts. Constructs like
callcc are very hard and non-elegant to define in frameworks such as SOS,
because those implicitly represent the execution context as proof context,
and the latter cannot be easily changed.

Now that we know how to handle cells in configurations and use them in rules,
in the next lesson we take a fresh look at LAMBDA and define it using
an environment-based style, which avoids the complexity of substitution
(e.g., having to deal with variable capture) and is closer in spirit to how
functional languages are implemented.

Go to Lesson 2, LAMBDA++: Semantic (Non-Syntactic) Computation Items.

MOVIE (out of date) [6'28"]

Semantic (Non-Syntactic) Computation Items

In this lesson we start another semantic definition of LAMBDA++, which
follows a style based on environments instead of substitution. In terms of
K, we will learn how easy it is to add new items to the syntactic category
of computations K, even ones which do not have a syntactic nature.

An environment binds variable names of interest to locations where their
values are stored. The idea of environment-based definitions is to maintain
a global store mapping locations to values, and then have environments
available when we evaluate expressions telling where the variables are
located in the store. Since LAMBDA++ is a relatively simple language, we
only need to maintain one global environment. Following a similar style
like in IMP, we place all cells into a top cell T:

configuration <T>
                <k> $PGM:Exp </k>
                <env> .Map </env>
                <store> .Map </store>
              </T>

Recall that $PGM is where the program is placed by krun after parsing. So
the program execution starts with an empty environment and an empty store.

In environment-based definitions of lambda-calculi, lambda abstractions
evaluate to so-called closures:

rule <k> lambda X:Id . E => closure(Rho,X,E) ...</k>
     <env> Rho </env>

A closure is like a lambda abstraction, but it also holds the environment
in which it was declared. This way, when invoked, a closure knows where to
find in the store the values of all the variables that its body expression
refers to. We will define the lookup rule shortly.

Therefore, unlike in the substitution-based definitions of LAMBDA and
LAMBDA++, neither the lambda abstractions nor the identifiers are values
anymore here, because they both evaluate further: lambda abstractions to
closures and identifiers to their values in the store. In fact, the only
values at this moment are the closures, and they are purely semantic entities,
which cannot be used explicitly in programs. That's why we modified the
original syntax of the language to include no Val syntactic category
anymore, and that's why we need to add closures as values now; same like
before, we add a Val syntactic category which is subsorted
to KResult. In general, whenever you have any strictness attributes,
your should also define some K results.

Invoking a closure is a bit more involved than the substitution-based
beta-reduction: we need to switch to the closure's environment, then create a
new, or fresh, binding for the closure's parameter to the value passed to the
closure, then evaluate the closure's body, and then switch back to the
caller's environment, which needs to be stored somewhere in the meanwhile.
We can do all these with one rule:

rule <k> closure(Rho,X,E) V:Val => E ~> Rho' ...</k>
     <env> Rho' => Rho[X <- !N] </env>
     <store>... .Map => (!N:Int |-> V) ...</store>

Therefore, we atomically do all the following:

  • switch the computation to the closure's body, E, followed by a
    caller-environment-recovery task Rho' (note that Rho' is the
    current environment),
  • generate a fresh location !N (the ! is important, we discuss it below),
    bind X to !N in closure's environment and switch the current environment
    Rho' to that one,
  • write the value passed to the closure, V, at location !N.

This was the most complex K rule we've seen so far in the tutorial. Note,
however, that this one rule achieves a lot. It is, in fact, quite compact
considering how much it does. Note also that everything that this K rule
mentions is needed also conceptually in order to achieve this task, so it
is minimal from that point of view. That would not be the case if we
used, instead, a conventional rewrite rule, because we would have had to
mention the remaining store, say Sigma, in both sides of the rule, to say
it stays unchanged. Here we just use ....

The declaration of the fresh variable above, !N, is new and needs
some explanation. First, note that !N appears only in the right-hand-side
terms in the rule, that is, it is not matched when the rule is applied.
Instead, a fresh Nat element is generated each time the rule is applied.
In K, we can define syntactic categories which have the capability to
generate fresh elements like above, using unbound variables whose name starts
with a !. The details of how to do that are beyond the scope of this
tutorial (see Tutorial 6). All we need to know here is that an arbitrary
fresh element of that syntactic category is generated each time the rule
is applied. We cannot rely on the particular name or value of the generated
element, because that can change with the next version of the K tool, or
even from execution to execution with the same version. All you can rely
on is that each newly generated element is distinct from the previously
generated elements for the same syntactic category.

Unlike in the substitution-based definition, we now also need a lookup rule:

rule <k> X => V ...</k>
     <env>... X |-> N ...</env>
     <store>... N |-> V ...</store>

This rule speaks for itself: replace X by the value V located in the store
at X's location N in the current environment.

The only thing left to define is the auxiliary environment-recovery operation:

rule _:Val ~> (Rho => .) ... _ => Rho

When the item preceding the environment recovery task Rho in the
computation becomes a value, replace the current environment with Rho
and dissolve Rho from the computation.

Let us kompile and ... fail:

kompile lambda

gives a parsing error saying that V:Val does not fit there in the closure
invocation rule. That's because Val and Exp are currently completely
disconnected, so K rightfully complains that we want to apply a value to
another one, because application was defined to work with expressions, not
values. What we forgot here was to state that Exp includes Val:

syntax Exp ::= Val

Now everything works, but it is a good time to reflect a bit.

So we added closures, which are inherently semantic entities, to the syntax
of expressions. Does that mean that we can now write LAMBDA programs with
closures in them? Interestingly, with our current definition of LAMBDA,
which purposely did not follow the nice organization of IMP into syntax and
semantic modules, and with K's default parser, kast, you can. But you are
not supposed to speculate this! In fact, if you use an external parser, that
parser will reject programs with explicit closures. Also, if we split the
LAMBDA definition into two modules, one called LAMBDA-SYNTAX containing
exclusively the desired program syntax and one called LAMBDA importing the
former and defining the syntax of the auxiliary operations and the semantics,
then even K's default parser will reject programs using auxiliary syntactic
constructs.

Indeed, when you kompile a language, say lang.k, the tool will by default
attempt to find a module LANG-SYNTAX and generate the program parser from
that. If it cannot find it, then it will use the module LANG instead. There
are also ways to tell kompile precisely which syntax module you want to use
for the program parser if you don't like the default convention.
See kompile --help.

Another insightful thought to reflect upon, is the relationship between your
language's values and other syntactic categories. It is often the case that
values form a subset of the original language syntax, like in IMP (Part 2 of
the tutorial), but sometimes that is not true, like in our case here. When
that happens, in order for the semantics to be given smoothly and uniformly
using the original syntax, you need to extend your language's original
syntactic categories with the new values. The same holds true in other
semantic approaches, not only in K, even in ones which are considered purely
syntactic. As it should be clear by now, K does not enforce you to use a
purely syntactic style in your definitions; nevertheless, K does allow you to
develop purely syntactic definitions, like LAMBDA in Part 1 of the tutorial,
if you prefer those.

krun some programs, such as those provided in Lesson 1 of the LAMBDA
tutorial (Part 1). Note the closures, both as results in the <k/> cell,
and as values in the store. Also, since variables are not values anymore,
expressions that contain free variables may get stuck with one of those on
top of their computation. See, for example, free-variable-capture.lambda,
which gets stuck on z, because z is free, so it cannot evaluate it.
If you want, you can go ahead and manually provide a configuration with
z mapped to some location in the environment and that location mapped to
some value in the store, and then you can also execute this program. The
program omega.lambda should still loop.

Although we completely changed the definitional style of LAMBDA, the semantics
of the other constructs do not need to change, as seen in the next lesson.

Go to Lesson 3, LAMBDA++: Reusing Existing Semantics.

MOVIE (out of date) [8'02"]

Reusing Existing Semantics

In this lesson we will learn that, in some cases, we can reuse existing
semantics of language features without having to make any change!

Although the definitional style of the basic LAMBDA language changed quite
radically in our previous lesson, compared to its original definition in
Part 1 of the tutorial, we fortunately can reuse a large portion of the
previous definition. For example, let us just cut-and-paste the rest of the
definition from Lesson 7 in Part 1 of the tutorial.

Let us kompile and krun all the remaining programs from Part 1 of the
tutorial. Everything should work fine, although the store contains lots of
garbage. Garbage collection is an interesting topic, but we do not do it
here. Nevertheless, much of this garbage is caused by the intricate use of
the fixed-point combinator to define recursion. In a future lesson in this
tutorial we will see that a different, environment-based definition of
fixed-points will allocate much less memory.

One interesting question at this stage is: how do we know when we can reuse
an existing semantics of a language feature? Well, I'm afraid the answer is:
we don't. In the next lesson we will learn how reuse can fail for quite subtle
reasons, which are impossible to detect statically (and some non-experts may
fail to even detect them at all).

Go to Lesson 4, LAMBDA++: Do Not Reuse Blindly!.

MOVIE (out of date) [3'21"]

Do Not Reuse Blindly!

It may be tempting to base your decision to reuse an existing semantics of
a language feature solely on syntactic considerations; for example, to reuse
whenever the parser does not complain. As seen in this lesson, this could
be quite risky.

Let's try (and fail) to reuse the definition of callcc from Lesson 1:

syntax Exp ::= "callcc" Exp  [strict]
syntax Val ::= cc(K)
rule <k> (callcc V:Val => V cc(K)) ~> K </k>
rule <k> cc(K) V ~> _ =>  V ~> K </k>

The callcc examples that we tried in Lesson 1 work, so it may look it works.

However, the problem is that cc(K) should also include an environment,
and that environment should also be restored when cc(K) is invoked.
Let's try to illustrate this bug with callcc-env1.lambda

let x = 1 in
  ((callcc lambda k . (let x = 2 in (k x))) + x)

where the second argument of +, x, should be bound to the top x, which
is 1. However, since callcc does not restore the environment, that x
should be looked up in the wrong, callcc-inner environment, so we should see
the overall result 4.

Hm, we get the right result, 3 ... (Note: you may get 4, depending on
your version of K and platform; but both 3 and 4 are possible results, as
explained below and seen in the tests). How can we get 3? Well, recall that
+ is strict, which means that it can evaluate its arguments in any order.
It just happened that in the execution that took place above its second
argument was evaluated first, to 1, and then the callcc was evaluated, but
its cc value K had already included the 1 instead of x ... In Part 4 of
the tutorial we will see how to explore all the non-deterministic behaviors of
a program; we could use that feature of K to debug semantics, too.
For example, in this case, we could search for all behaviors of this program
and we would indeed get two possible value results: 3 and 4.

One may think that the problem is the non-deterministic evaluation order
of +, and thus that all we need to do is to enforce a deterministic order
in which the arguments of + are evaluated. Let us follow this path to
see what happens. There are two simple ways to make the evaluation order
of +'s arguments deterministic. One is to make + seqstrict in the
semantics, to enforce its evaluation from left-to-right. Do it and then
run the program above again; you should get only one behavior for the
program above, 4, which therefore shows that copying-and-pasting our old
definition of callcc was incorrect. However, as seen shortly, that only
fixed the problem for the particular example above, but not in general.
Another conventional approach to enforce the desired evaluation order is to
modify the program to enforce the left-to-right evaluation order using let
binders, as we do in callcc-env2.lambda:

let x = 1 in
  let a = callcc lambda k . (let x = 2 in (k x)) in
    let b = x in
      (a + b)

With your installation of K you may get the "expected" result 4 when you
execute this program, so it may look like our non-deterministic problem is
fixed. Unfortunately, it is not. Using the K tool to search for all the
behaviors in the program above reveals that the final result 3 is still
possible. Moreover, both the 3 and the 4 behaviors are possible regardless
of whether + is declared to be seqstrict or just strict. How is that
possible? The problem is now the non-deterministic evaluation strategy of
the function application construct. Indeed, recall that the semantics of
the let-in construct is defined by desugaring to lambda application:

rule let X = E in E' => (lambda X . E') E

With this, the program above eventually reduces to

(lambda a . ((lambda b . a + b) x))
(callcc lambda k . (let x = 2 in (k x)))

in an environment where x is 1. If the first expression evaluates first,
then it does so to a closure in which x is bound to a location holding 1,
so when applied later on to the x inside the argument of callcc (which is
2), it will correctly lookup x in its enclosed environment and thus the
program will evaluate to 3. On the other hand, if the second expression
evaluates first, then the cc value will freeze the first expression as is,
breaking the relationship between its x and the current environment in which
it is bound to 1, being inadvertently captured by the environment of the
let-in construct inside the callcc and thus making the entire expression
evaluate to 4.

So the morale is: Do not reuse blindly. Think!

In the next lesson we fix the environment-based semantics of callcc by having
cc also wrap an environment, besides a computation. We will also give a more
direct semantics to recursion, based on environments instead of fixed-point
combinators.

Go to Lesson 5, LAMBDA++: More Semantic Computation Items.

MOVIE (out of date) [3'37"]

More Semantic Computation Items

In this lesson we see more examples of semantic (i.e., non-syntactic)
computational items, and how useful they can be. Specifically, we fix the
environment-based definition of callcc and give an environment-based
definition of the mu construct for recursion.

Let us first fix callcc. As discussed in Lesson 4, the problem that we
noticed there was that we only recovered the computation, but not the
environment, when a value was passed to the current continuation. This is
quite easy to fix: we modify cc to take both an environment and a
computation, and its rules to take a snapshot of the current environment with
it, and to recover it at invocation time:

syntax Val ::= cc(Map,K)
rule <k> (callcc V:Val => V cc(Rho,K)) ~> K </k> <env> Rho </env>
rule <k> cc(Rho,K) V:Val ~> _ =>  V ~> K </k> <env> _ => Rho </env>

Let us kompile and make sure it works with the callcc-env2.lambda program,
which should evaluate to 3, not to 4.

Note that the cc value, which can be used as a computation item in the <k/>
cell, is now quite semantic in nature, pretty much the same as the closures.

Let us next add one more closure-like semantic computational item, for mu.
But before that, let us reuse the semantics of letrec in terms of mu that
was defined in Lesson 8 of Part 1 of the tutorial on LAMBDA:

syntax Exp ::= "letrec" Id Id "=" Exp "in" Exp [macro]
             | "mu" Id "." Exp                 [latex(\mu{#1}.{#2})]
rule letrec F:Id X = E in E' => let F = mu F . lambda X . E in E'

We removed the binder annotation of mu, because it is not necessary
anymore (since we do not work with substitutions anymore).

To save the number of locations needed to evaluate mu X . E, let us replace
it with a special closure which already binds X to a fresh location holding
the closure itself:

syntax Exp ::= muclosure(Map,Exp)

rule <k> mu X . E => muclosure(Rho[X <- !N], E) ...</k>
     <env> Rho </env>
     <store>... .Map => (!N:Int |-> muclosure(Rho[X <- !N], E)) ...</store>

Since each time mu X . E is encountered during the evaluation it needs to
evaluate E, we conclude that muclosure cannot be a value. We can declare
it as either an expression or as a computation. Let's go with the former.

Finally, here is the rule unrolling the muclosure:

rule muclosure(Rho,E) => E ~> Rho' ...
Rho' => Rho

Note that the current environment Rho' needs to be saved before and
restored after E is executed, because the fixed point may be invoked
from a context with a completely different environment from the one
in which mu X . E was declared.

We are done. Let us now kompile and krun factorial-letrec.lambda from
Lesson 7 in Part 1 of the tutorial on LAMBDA. Recall that in the previous
lesson this program generated a lot of garbage into the store, due to the
need to allocate space for the arguments of all those lambda abstractions
needed to run the fixed-point combinator. Now we need much fewer locations,
essentially only locations for the argument of the factorial function, one at
each recursive call. Anyway, much better than before.

In the next lesson we wrap up the environment definition of LAMBDA++ and
generate its documentation.

Go to Lesson 6, LAMBDA++: Wrapping Up and Documenting LAMBDA++.

MOVIE (out of date) [5'19"]

Wrapping Up and Documenting LAMBDA++

In this lesson we wrap up and nicely document LAMBDA++. In doing so, we also
take the freedom to reorganize the semantics a bit, to make it look better.

See the lambda.k file, which is self-explanatory.

Part 3 of the tutorial is now complete. Part 4 will teach you more features
of the K framework, in particular how to exhaustively explore the behaviors
of non-deterministic or concurrent programs.

MOVIE (out of date) [6'23"]

Part 4: Defining IMP++

IMP++ extends IMP, which was discussed in Part 2 of this tutorial, with several
new syntactic constructs. Also, some existing syntax is generalized, which
requires non-modular changes of the existing IMP semantics. For example,
global variable declarations become local declarations and can occur
anywhere a statement can occur. In this tutorial we will learn the following:

  • That (and how) existing syntax/semantics may change as a language evolves.
  • How to refine configurations as a language evolves.
  • How to define and use fresh elements of desired sorts.
  • How to tag syntactic constructs and rules, and how to use such tags
    with the superheat/supercool options of kompile.
  • How the search option of krun works.
  • How to stream cells holding semantic lists to the standard input/output,
    and thus obtain interactive interpreters for the defined languages.
  • How to delete, save and restore cell contents.
  • How to add/delete cells dynamically.
  • More details on how the configuration abstraction mechanism works.

Like in the previous tutorials, this folder contains several lessons, each
adding new features to IMP++. Do them in order and make sure you completed
and understood the previous tutorials.

Extending/Changing an Existing Language Syntax

Here we learn how to extend the syntax of an existing language, both with
new syntactic constructs and with more general uses of existing constructs.
The latter, in particular, requires changes of the existing semantics.

Consider the IMP language, as defined in Lesson 4 of Part 2 of the tutorial.

Let us first add the new syntactic constructs, with their precedences:

  • variable increment, ++, which increments an integer variable and
    evaluates to the new value;
  • read, which reads and evaluates to a new integer from the input buffer;
  • print, which takes a comma-separated list of arithmetic expressions and
    evaluates and prints each of them in order, from left to right, to the
    output buffer; we therefore define a new list syntactic category, AExps,
    which we pass as an argument to print; note we do not want to declare
    print to be strict, because we do not want to first evaluate the
    arguments and then print them (for example, if the second argument performs
    an illegal operation, say division by zero, we still want to print the first
    argument); we also go ahead and add strings as arithmetic expressions,
    because we intend print to also take strings, in order to print nice
    messages to the user;
  • halt, which abruptly terminates the program; and
  • spawn, which takes a statement and creates a new concurrent thread
    executing it and sharing its environment with the parent thread.

Also, we want to allow local variable declarations, which can appear anywhere
a statement can appear. Their scope ranges from the place they are defined
until the end of the current block, and they can shadow previous declarations,
both inside and outside the current block. The simplest way to define the
syntax of the new variable declarations is as ordinary statements, at the same
time removing the previous Pgm syntactic category and its construct.
Programs are now just statements.

We are now done with adding the new syntax and modifying the old one.
Note that the old syntax was modified in a way which makes the previous IMP
programs still parse, but this time as statements. Let us then modify
the configuration variable $PGM to have the sort Stmt instead of Pgm,
and let us try to run the old IMP programs, for example sum.imp.

Note that they actually get stuck with the global declaration on the top
of their computations. This is because variable declarations are now treated
like any statements, in particular, the sequential composition rule applies.
This makes the old IMP rule for global variable declarations not match anymore.
We can easily fix it by replacing the anonymous variable _, which matched
the program's statement that now turned into the remaining computation in
the <k/> cell, with the cell frame variable ..., which matches the
remaining computation. Similarly, we have to change the rule for the case
where there are no variables left to declare into one that dissolves itself.

We can now run all the previous IMP programs, in spite of the fact that
our IMP++ semantics is incomplete and, more interestingly, in spite of the
fact that our current semantics of blocks is incorrect in what regards the
semantics of local variable declarations (note that the old IMP programs do
not declare block-local variables, which is why they still run correctly).

Let us also write some proper IMP++ programs, which we would like to execute
once we give semantics to the new constructs.

div.imp is a program manifesting non-deterministic behaviors due to the
desired non-deterministic evaluation strategy of division and the fact that
expressions will have side effects once we add variable increment. We will
be able to see all the different behaviors of this program. Challenge: can
you identify the behavior where the program performs a division-by-zero?

If we run div.imp now, it will get stuck with the variable increment
construct on top of the computation cell. Once we give it a semantics,
div.imp will execute completely (all the other constructs in div.imp
already have their semantics defined as part of IMP).

Note that some people prefer to define all their semantics in a by need
style, that is, they first write and parse lots of programs, and then they
add semantics to each language construct on which any of the programs gets
stuck, and so on and so forth until they can run all the programs.

io.imp is a program which exercises the input/output capabilities of the
language: reads two integers and prints three strings and an integer.
Note that the variable declaration is not the first statement anymore.

sum-io.imp is an interactive variant of the sum program.

spawn.imp is a program which dynamically creates two threads that interact
with the main thread via the shared variable x. Lots of behaviors will be
seen here once we give spawn the right semantics.

Finally, locals.imp tests whether variable shadowing/unshadowing works well.

In the next lesson we will prepare the configuration for the new constructs,
and will see what it takes to adapt the semantics to the new configuration.
Specifically, we will split the state cell into an environment cell and a
store cell, like in LAMBDA++ in Part 3 of the tutorial.

Go to Lesson 2, IMP++: Configuration Refinement; Freshness.

MOVIE (out of date) [07'47"]

Configuration Refinement; Freshness

To prepare for the semantics of threads and local variables, in this lesson we
split the state cell into an environment and a store. The environment and
the store will be similar to those in the definition of LAMBDA++ in Part
3 of the Tutorial. This configuration refinement will require us to change
some of IMP's rules, namely those that used the state.

To split the state map, which binds program variables to values, into an
environment mapping program variables to locations and a store mapping
locations to values, we replace in the configuration declaration the cell

<state color="red"> .Map </state>

with two cells

<env color="LightSkyBlue"> .Map </env>
<store color="red"> .Map </store>

Structurally speaking, this split of a cell into other cells is a major
semantic change, which, unfortunately, requires us to revisit the existing
rules that used the state cell. One could, of course, argue that we could
have avoided this problem if we had followed from the very beginning the
good-practice style to work with an environment and a store, instead of a
monolithic state. While that is a valid argument, highlighting the fact that
modularity is not only a feature of the framework alone, but one should also
follow good practices to achieve it, it is also true that if all we wanted
in Part 2 of the tutorial was to define IMP as is, then the split of the state
in an environment and a store is unnecessary and not really justified.

The first rule which used a state cell is the lookup rule:

rule <k> X:Id => I ...</k> <state>... X |-> I ...</state>

We modify it as follows:

rule <k> X:Id => I ...</k>
     <env>... X |-> N ...</env>
     <store>... N |-> I ...</store>

So we first match the location N of X in the environment, then the value
I at location N in the store, and finally we rewrite X to I into the
computation. This rule also shows an instance of a more complex
multiset matching, where two variables (X and N) are matched each twice.

The assignment rule is modified quite similarly.

The variable declaration rule is trickier, though, because we need to allocate
a fresh location in the store and bind the newly declared variable to it.
This is quite similar to the way we allocated space for variables in
the environment-based definition of LAMBDA++ in Part 3 of the tutorial.

rule <k> int (X,Xs => Xs); ...</k>
     <env> Rho => Rho[X <- !N:Int] </env>
     <store>... .Map => !N |-> 0 ...</store>

Note the use of the fresh (!N) variable notation above. Recall from
the LAMBDA++ tutorial that each time the rule with fresh (!) variables is
applied, fresh elements of corresponding sorts are generated for the fresh
variables, distinct from all the previously generated elements; also, we
cannot and should not assume anything about the particular element that is
being generated, except that it is different from the previous ones.

kompile and krun sum.imp to see how the fresh locations have been
generated and used. There were two fresh locations needed, for the two
variables. Note also that a cell holding the counter has been added to the
configuration.

In the next lesson we will add the semantics of variable increment, and see
how that yields non-deterministic behaviors in programs and how to explore
those behaviors using the K tool.

Go to Lesson 3, IMP++: Tagging; Superheat/Supercool Kompilation Options.

MOVIE (out of date) [04'06"]

Variable increment; Search

In this lesson we add the semantics of variable increment. We also learn
how to instruct the kompile tool to instrument the language model for
exhaustive analysis.

The variable increment rule is self-explanatory:

rule <k> ++X => I +Int 1 ...</k>
     <env>... X |-> N ...</env>
     <store>... N |-> (I => I +Int 1) ...</store>

We can now run programs like our div.imp program introduced in Lesson 1.
Do it.

The addition of increment makes the evaluation of expressions have side
effects. That, in combination with the non-determinism allowed by the
strictness attributes in how expression constructs evaluate their
arguments, makes expressions in particular and programs in general have
non-deterministic behaviors. One possible execution of the div.imp program
assigns 1 to y's location, for example, but this program manifests several
other behaviors, too.

To see all the (final-state) behaviors that a program can have, you can kompile
the semantics with --enable-search and call the krun tool with the option
--search. For example:

krun div.imp --search

In the next lesson we add input/output to our language and learn how to
generate a model of it which behaves like an interactive interpreter!

Go to Lesson 4, IMP++: Semantic Lists; Input/Output Streaming.

MOVIE (out of date) [06'56"]

Semantic Lists; Input/Output Streaming

In this lesson we add semantics to the read and print IMP++ constructs.
In doing so, we also learn how to use semantic lists and how to connect
cells holding semantic lists to the standard input and standard output.
This allows us to turn the K semantics into an interactive interpreter.

We start by adding two new cells to the configuration,

<in color="magenta"> .List </in>
<out color="Orchid"> .List </out>

each holding a semantic list, initially empty. Semantic lists are
space-separated sequences of items, each item being a term of the form
ListItem(t), where t is a term of sort K. Recall that the semantic maps,
which we use for states, environments, stores, etc., are sets of pairs
t1 |-> t2, where t1 and t2 are terms of sort K. The ListItem wrapper
is currently needed, to avoid parsing ambiguities.

Since we want the print statement to also print strings, we need to tell
K that strings are results. To make it more interesting, let us also overload
the + symbol on arithmetic expressions to also take strings and, as a
result, to concatenate them. Since + is already strict, we only need to add
a rule reducing the IMP addition of strings to the builtin operation +String
which concatenates two strings.

The semantics of read is immediate: read and consumes the first integer item
from the <in/> cell; note that our read only reads integer values (it gets
stuck if the first item in the <in/> cell is not an integer).

The semantics of print is a bit trickier. Recall that print takes an
arbitrary number of arithmetic expression arguments, and evaluates and outputs
each of them in order, from left to right. For example,
print("Hello", 3/0, "Bye"); outputs "Hello" and then gets stuck on the
illegal division by zero operation. In other words, we do not want it to
first evaluate all its arguments and then print them, because that would miss
outputting potentially valuable information. So the first step is to evaluate
the first argument of print. In some sense, what we'd like to say is that
print has the evaluation strategy strict(1). However, strictness
attributes only work with individual language constructs, while what we need
is an evaluation strategy that involves two constructs: print and the list
(comma) construct of AExps. If we naively associate print the strict(1)
evaluation strategy then its first and unique argument, an AExps list, will
be scheduled for evaluation and the execution will get stuck because we have
no rules for evaluating AExps terms. If we make the list construct of
AExps strict then we get the wrong semantics for print which first
evaluates all its arguments and then outputs them. The correct way to
tell K that print should evaluate only its first argument is by using a
context declaration:

context print(HOLE:AExp, _);

Note the HOLE of sort AExp above. Contexts allow us to define finer-grain
evaluation strategies than the strictness attributes, involving potentially
more than one language construct, like above. The HOLE indicates the
argument which is requested to be evaluated. For example, the strict
attribute of division corresponds to two contexts:

context HOLE / _
context _ / HOLE

In their full generality, contexts can be any terms with precisely one
occurrence of a HOLE, and with arbitrary side conditions on any variables
occurring in the context term as well as on the HOLE. See Part 6 of the
tutorial for more examples.

Once evaluated, the first argument of print is expected to become either an
integer or a string. Since we want to print both integers and string values,
to avoid writing two rules, one for each type of value, we instead add a new
syntactic category, Printable, which is the union of integers and strings.

Let us kompile and krun the io.imp program discussed in Lesson 1. As
expected, it gets stuck with a read construct on top of the computation and
with an empty <in/> cell. To run it, we need to provide some items in the
<in/> cell, so that the rule of read can match. Let us add

<in> ListItem(3) ListItem(5) ListItem(7) </in>

Now, if we krun io.imp, we can see that its execution completes normally
(the <k/> cell is empty), that the first two items have been removed by the
two read constructs from the <in/> cell, and that the desired strings and
numbers have been placed into the <out/> cell.

Cells holding semantic lists can be connected to the standard input and
standard output buffers, and krun knows how to handle these appropriately.
Let us connect the <in/> cell to the standard input using the cell attribute
stream="stdin" and the <out/> cell to the standard output with the
attribute stream="sdtout". A cell connected to the standard input will
take its items from the standard input and block the rewriting process when
an input is needed until an item is available in the standard input buffer.
A cell connected to the standard output buffer will send all its items, in
order, to the standard output.

Let us kompile and krun io.imp again. It prints the message and then
waits for your input numbers. Type in two numbers, then press <Enter>.
A message with their sum is then printed, followed by the final configuration.
If you do not want to see the final configuration, and thus obtain a realistic
interpreter for our language, then call krun with the option --output none:

krun io.imp --output none

Let us now krun our interactive sum program, which continuously reads numbers
from the console and prints the sum of numbers up to them:

krun sum-io.imp

Try a few numbers, then 0. Note that the program terminated, but with junk
in the <k/> cell, essentially with a halt statement on its top. Of course,
because halt has been reached and it has no semantics yet.

In the next lesson we give the semantics of halt and also fix the semantics
of blocks with local variable declarations.

Go to Lesson 5, IMP++: Deleting, Saving and Restoring Cell Contents.

MOVIE (out of date) [05'21"]

Deleting, Saving and Restoring Cell Contents

In this lesson we will see how easily we can delete, save and/or restore
contents of cells in order to achieve the desired semantics of language
constructs that involve abrupt changes of control or environments. We have
seen similar or related K features in the LAMBDA++ language in Part 3 of the
tutorial.

Let us start by adding semantics to the halt statement. As its name says,
what we want is to abruptly terminate the execution of the program. Moreover,
we want the program configuration to look as if the program terminated
normally, with an empty computation cell. The simplest way to achieve that is
to simply empty the computation cell when halt is encountered:

rule <k> halt; ~> _ => . </k>

It is important to mention the entire <k/> cell here, with both its membranes
closed, to make sure that its entire contents is discarded. Note the
anonymous variable, which matches the rest of the computation.

kompile and krun sum-io.imp. Note that unlike in Lesson 4, the program
terminates with an empty computation cell now.

As mentioned earlier, the semantics of blocks that was inherited from IMP is
wrong. Program locals.imp shows it very clearly: the environments are not
correctly restored at block exits. One way to fix the problem is to take
a snapshot of the current environment when a block is entered and save it
somewhere, and then to restore it when the block is left. There are many
ways to do this, which you can explore on your own: for example you can add
a new list cell for this task where to push/pop the environment snapshots in
a stack style; or you can use the existing environment cell for this purpose,
but then you need to change the variable access rules to search through the
stacked environments for the variable.

My preferred solution is to follow a style similar to how we saved/restored
LAMBDA++ environments in Part 3 of the Tutorial, namely to use the already
existing <k/> cell for such operations. More specifically, we place a
reminder item in the computation whenever we need to take a snapshot of
some cell contents; the item simply consists of the entire contents of the cell.
Then, when the reminder item is reached, we restore the contents of the cell:

rule <k> {S} => S ~> Rho ...</k> <env> Rho </env>

The only thing left now is to give the definition of environment restore:

rule <k> Rho => . ...</k> <env> _ => Rho </env>

Done. kompile and krun locals.imp. Everything should work correctly now.
Note that the rule above is different from the one we had for LAMBDA++ in
Part 3 of the tutorial, in that here there is no value preceding the environment
restoration item in the computation; that's because IMP++ statements,
unlike LAMBDA++'s expressions, evaluate to nothing (.).

In the next lesson we will give semantics to the spawn S construct, which
dynamically creates a concurrent shared-memory thread executing statement S.

Go to Lesson 6, IMP++: Adding/Deleting Cells Dynamically; Configuration Abstraction, Part 2.

MOVIE (out of date) [04'30"]

Adding/Deleting Cells Dynamically; Configuration Abstraction, Part 2

In this lesson we add dynamic thread creation and termination to IMP, and
while doing so we learn how to define and use configurations whose structure
can evolve dynamically.

Recall that the intended semantics of spawn S is to spawn a new concurrent
thread that executes S. The new thread is being passed at creation time
its parent's environment, so it can share with its parent the memory
locations that its parent had access to at creation time. No other locations
can be shared, and no other memory sharing mechanism is available.
The parent and the child threads can evolve unrestricted, in particular they
can change their environments by declaring new variables or shadowing existing
ones, can create other threads, and so on.

The above suggests that each thread should have its own computation and its
own environment. This can be elegantly achieved if we group the <k/> and
<env/> cells in a <thread/> cell in the configuration. Since at any given
moment during the execution of a program there could be zero, one or more
instances of such a <thread/> cell in the configuration, it is a good idea
to declare the <thread/> cell with multiplicity * (i.e., zero, one or more):

<thread multiplicity="*" color="blue">
  <k color="green"> $PGM:Stmt </k>
  <env color="LightSkyBlue"> .Map </env>
</thread>

This multiplicity declaration is not necessary, but it is a good idea to do
it for several reasons:

  1. it may help the configuration abstraction process,
    which may in turn significantly increase the compactness and modularity of
    your subsequent rules;
  2. it may help various analysis and execution tools,
    for example static analyzers to give you error messages when you create cells
    where you should not, or K compilers to improve performance by starting
    actual concurrent hardware threads or processes corresponding to each cell
    instance; and
  3. it may help you better understand and control the dynamics
    of your configuration, and thus your overall semantics.

For good encapsulation, I also prefer to put all thread cells into one cell,
<threads/>. This is technically unnecessary, though; to convince yourself
that this is indeed the case, you can remove this cell once we are done with
the semantics and everything will work without having to make any changes.

Before we continue, let us kompile an krun some programs that used to
work, say sum-io.imp. In spite of the relatively radical configuration
reorganization, those programs execute just fine! How is that possible?
In particular, why do rules like the lookup and assignment still work,
unchanged, in spite of the fact that the <k/> and <env/> cells are not at
the same level with the <store/> cell in the configuration anymore?

Welcome to configuration abstraction, part 2. Recall that the role of
configuration abstraction is to allow you to only write the relevant
information in each rule, and have the compiler fill-in the obvious and boring
details. According to the configuration that we declared for our new
language, there is only one reasonable way to complete rules like the lookup,
namely to place the <k/> and </env> cells inside a <thread/> cell,
inside a <threads/> cell:

rule <threads>...
       <thread>...
         <k> X:Id => I ...</k>
         <env>... X |-> N ...</env>
       ...</thread>
     ...<threads/>
     <store>... N |-> I ...</store>  [lookup]

This is the most direct, compact and local way to complete the configuration
context of the lookup rule. If for some reason you wanted here to match the
<k/> cell of one thread and the <env/> cell of another thread, then you
would need to explicitly tell K so, by mentioning the two thread cells,
for example:

rule <thread>...
         <k> X:Id => I ...</k>
     ...</thread>
     <thread>...
         <env>... X |-> N ...</env>
     ...</thread>
     <store>... N |-> I ...</store>  [lookup]

By default, K completes rules in a greedy style. Think this way: what is the
minimal number of changes to my rule to make it fit the declared
configuration? That's what the K tool will do.

Configuration abstraction is technically unnecessary, but once you start
using it and get a feel for how it works, it will become your best friend.
It allows you to focus on the essentials of your semantics, and at the same
time gives you flexibility in changing the configuration later on without
having to touch the rules. For example, it allows you to remove the
<threads/> cell from the configuration, if you don't like it, without
having to touch any rule.

We are now ready to give the semantics of spawn:

rule <k> spawn S => . ...</k> <env> Rho </env>
     (. => <thread>... <k> S </k> <env> Rho </env> ...</thread>)

Note configuration abstraction at work, again. Taking into account
the declared configuration, and in particular the multiplicity information
* in the <thread/> cell, the only reasonable way to complete the rule
above is to wrap the <k/> and <env/> cells on the first line within a
<thread/> cell, and to fill-in the ...s in the child thread with the
default contents of the other subcells in <thread/>. In this case there
are no other cells, so we can get rid of those ...s, but that would
decrease the modularity of this rule: indeed, we may later on add other
cells within <thread/> as the language evolves, for example a function
or an exception stack, etc.

In theory, we should be able to write the rule above even more compactly
and modularly, namely as

rule <k> spawn S => . ...</k> <env> Rho </env>
     (. => <k> S </k> <env> Rho </env>)

Unfortunately, this currently does not work in the K tool, due to some
known limitations of our current configuration abstraction algorithm.
This latter rule would be more modular, because it would not even depend
on the cell name thread. For example, we may later decide to change
thread into agent, and we would not have to touch this rule.
We hope this current limitation will be eliminated soon.

Once a thread terminates, its computation cell becomes empty. When that
happens, we can go ahead and remove the useless thread cell:

rule <thread>... <k> . </k> ...</thread> => .

Let's see what we've got. kompile and krun spawn.imp.
Note the following:

  • The <threads/> cell is empty, so all threads terminated normally;
  • The value printed is different from the value in the store; the store value
    is not even the one obtained if the threads executed sequentially.

Therefore, interesting behaviors may happen; we would like to see them all!

krun spawn.imp --search

However, the above does not work.

spawn.imp is an interactive program, which reads a number from the
standard input. When analyzing programs exhaustively using the search option,
krun has to disable the streaming capabilities (just think about it and you
will realize why). The best you can do in terms of interactivity with search
is to pipe some input to krun: krun will flush the standard input buffer
into the cells connected to it when creating the initial configuration (will
do that no matter whether you run it with or without the --search option).
For example:

echo 23 | krun spawn.imp --search

puts 23 in the standard input buffer, which is then transferred in the
<in/> cell as a list item, and then the exhaustive search procedure is
invoked.

However, even after piping some input, the spawn.imp program outputs
an error:

[Error] krun: You must pass --enable-search to kompile to be able to use krun --search with the LLVM backend

As explained in Lesson 3, by default kompile optimizes the generated
language model for execution. In particular, it does not insert any
backtracking markers where transition attempts should be made, so krun
lacks the information it needs to exhaustively search the generated language
model.

kompile with the search feature enabled:

kompile imp --enable-search

Now echo 23 | krun spawn.imp --search gives us all 12 behaviors of the
spawn.imp program.

We currently have no mechanism for thread synchronization. In the next lesson
we add a join statement, which allows a thread to wait until another completes.

Go to Lesson 7, IMP++: Everything Changes: Syntax, Configuration, Semantics.

MOVIE (out of date) [11'40"]

Everything Changes: Syntax, Configuration, Semantics

In this lesson we add thread joining, one of the simplest thread
synchronization mechanisms. In doing so, we need to add unique ids
to threads in the configuration, and to modify the syntax to allow spawn
to return the id of the newly created thread. This gives us an opportunity
to make several other small syntactic and semantics changes to the language,
which make it more powerful or more compact at a rather low cost.

Before we start, let us first copy and modify the previous spawn.imp program
from Lesson 1 to make use of thread joining. Recall from Lesson 6 that in some
runs of this program the main thread completed before the child threads,
printing a possibly undesired value of x. What we want now is to assign
unique ids to the two spawned threads, and then to modify the main thread to
join the two child threads before printing. To avoid adding a new type to
the language, let's assume that thread ids are integer numbers. So we declare
two integers, t1 and t2, and assign them the two spawn commands. In order
for this to parse, we will have to change the syntax of spawn to be an
arithmetic expression construct instead of a statement. Once we do that,
we have a slight syntactic annoyance: we need to put two consecutive ;
after the spawn assignment, one for the assignment statement inside the spawn,
and another for the outer assignment. To avoid the two consecutive semicolons,
we can syntactically enforce spawn to take a block as argument, instead of a
statement. Now it looks better. The new spawn.imp program is still
non-deterministic, because the two threads can execute in any order and even
continue to have a data-race on the shared variable x, but we should see fewer
behaviors when we use the join statements. If we want to fully synchronize
this program, we can have the second thread start with a join(t1) statement.
Then we should only see one behavior for this program.

Let us now modify the language semantics. First, we move the spawn
construct from statements to expressions, and make it take a block.
Second, we add one more sub-cell to the thread cell in the configuration,
<id/>, to hold the unique identifier of the thread. We want the main
thread to have id 0, so we initialize this cell with 0. Third, we modify
the spawn rule to generate a fresh integer identifier, which is put in the
<id/> cell of the child thread and returned as a result of spawn in the
parent thread. Fourth, let us add the join statement to the language,
both syntactically and semantically. So in order for the join(T) statement
to execute, thread T must have its computation empty. However, in order
for this to work we have to get rid of the thread termination cleanup rule.
Indeed, we need to store somewhere the information that thread T terminated;
the simplest way to do it is to not remove the terminated threads. Feel free
to experiment with other possibilities, too, here. For example, you may add
another cell, <done/>, in which you can store all the thread ids of the
terminated and garbage-collected threads.

Let us now kompile imp.k and convince ourselves that the new spawn.imp
with join statements indeed has fewer behaviors than its variant without
join statements. Also, let us convince ourselves that the fully synchronized
variant of it indeed has only one behavior.

Note that now spawn, like variable increment, makes the evaluation of
expressions to have side effects. Many programming languages in fact allow
expressions to be evaluated only for their side effects, and not for their
value. This is typically done by simply adding a ; after the expression
and thus turning it into a statement. For example, ++x;. Let as also
allow arithmetic expressions in our language to be used as statements, by
simply adding the production AExp ";" to Stmt, with evaluation strategy
strict and with the expected semantics discarding the value of the AExp.

Another simple change in syntax and semantics which gives our language more
power, is to remove the ; from the syntax of variable assignments and to make
them expression instead of statement constructs. This change, combined with
the previous one, will still allow us to parse all the programs that we could
parse before, but will also allow us to parse more programs. For example, we
can now do sequence assignments like in C: x = y = z = 0. The semantics
of assignment now has to return the assigned value also to the computation,
because we want the assignment expression to evaluate to the assigned value.

Let us also make another change, but this time one which only makes the
definition more compact. Instead of defining statement sequential
composition as a binary construct for statements, let us define a new
syntactic construct, Stmts, as whitespace-separated lists of Stmt. This
allows us to get rid of the empty blocks, because we can change the syntax of
blocks to {Stmts} and Stmts also allows the empty sequence of statements.
However, we do have to make sure that .Stmts dissolves.

In general, unless you are defining a well-established programming language,
it is quite likely that your definitions will suffer lots of changes like the
ones seen in this lecture. You add a new construct, which suggests changes
in the existing syntax making in fact your language parse more programs,
which then requires corresponding changes in the semantics, and so on.
Also, compact definitions are desirable in general, because they are easier
to read and easier to change if needed later.

In the next lesson we wrap up and document the definition of IMP++.

Go to Lesson 8, IMP++: Wrapping up Larger Languages.

Wrapping up Larger Languages

In this lesson we wrap up IMP++'s semantics and also generate its poster.
While doing so, we also learn how to display larger configurations in order
to make them easier to read and print.

Note that we rearrange a bit the semantics, to group the semantics of old
IMP's constructs together, and separate it from the new IMP++'s semantics.

You can go even further and manually edit the generated Latex document.
You typically want to do that when you want to publish your language
definition, or parts of it, and you need to finely tune it to fit the
editing requirements. For example, you may want to insert some negative
spaces, etc.

Part 4 of the tutorial is now complete. At this moment you should know most
of K framework's features and how to use the K tool. You can now define or
design your own programming languages, and then execute and analyze programs.

MOVIE (out of date) [06'26"]

Part 5: Defining Type Systems

In this part of the tutorial we will show that defining type systems for
languages is essentially no different from defining semantics. The major
difference is that programs and fragments of programs now rewrite to their
types, instead of to concrete values. In terms of K, we will learn how
to use it for a certain particular but important kind of applications.

Imperative, Environment-Based Type Systems

In this lesson you learn how to define a type system for an imperative
language (the IMP++ language defined in Part 4 of the tutorial), using a style
based on type environments.

Let us copy the imp.k file from Part 4 of the tutorial, Lesson 7, which holds
the semantics of IMP++, and modify it into a type system. The resulting type
system, when executed, yields a type checker.

We start by defining the new strictness attributes of the IMP++ syntax.
While doing so, remember that programs and fragments of programs now reduce
to their types. So types will be the new results of our new (type) semantics.
We also clean up the semantics by removing the unnecessary tags, and also
use strict instead of seqstrict wherever possible, because strict gives
implementations more freedom. Interestingly, note that spawn is strict now,
because the code of the child thread should type in the current parent's type
environment. Note that this is not always the case for threads, see for example
SIMPLE in the languages tutorial, but it works here for our simpler IMP++.

From a typing perspective, the && construct is strict in both its arguments;
its short-circuit (concrete) semantics is irrelevant for its (static) type
system. Similarly, both the conditional and the while loop are strict
constructs when regarded through the typing lenses.

Finally, the sequential composition is now sequentially strict! Indeed,
statements are now going to reduce to their type, stmt, and it is critical
for sequential composition to type its argument statements left-to-right;
for example, imagine that the second argument is a variable declaration (whose
type semantics will modify the type environment).

We continue by defining the new results of computations, that is, the actual
types. In this simple imperative language, we only have a few constant types:
int, bool, string, block and stmt.

We next define the new configuration, which is actually quite simple. Besides
the <k/> cell, all we need is a type environment cell, <tenv/>, which will
hold a map from identifiers to their types. A type environment is therefore
like a state in the abstract domain of type values.

Let us next modify the semantic rules, turning them into a type system. In
short, the idea is to reduce the basic values to their types, and then have a
rule for each language construct reducing it to its result type whenever its
arguments have the expected types.

We write the rules in the order given by the syntax declarations, to make
sure we do not forget any construct.

Integers reduce to their type, int.

So do the strings.

Variables are now looked up in the type environment and reduced to their type
there. Since we only declare integer variables in IMP++, their type in tenv
will always be int. Nevertheless, we write the rule generically, so that we
would not have to change it later if we add other type declarations to IMP++.
Note that we reject programs which lookup undeclared variables. Rejection,
in this case, means rewriting getting stuck.

Variable increment types to int, provided the variable has type int.

Read types to int, because we only allow integer input.

Division is only allowed on integers, so it rewrites to int provided that its
arguments rewrite to int. Note, however, that in order to write int / int,
we have to explicitly add int to the syntax of arithmetic expressions.
Otherwise, the K parser rightfully complains, because / was declared on
arithmetic expressions, not on types. One simple and generic way to allow
types to appear anywhere, is to define Type as a syntactic subcategory of all
the other syntactic categories. Let's do it on a by-need basis, though.

Addition is overloaded, so we add two typing rules for it: one for integers
and another for strings.

As discussed, spawn types to stmt provided that its argument types to
block.

The assignment construct was strict(2); its typing policy is that the declared
type of X should be identical to the type of the assigned value. Like for
lookup, we define this rule more generically than needed for IMP++, for any
type, not only for int.

The typing rules for Boolean expression constructs are in the same spirit.
Note that we need only one rule for &&.

The typing of blocks is a bit trickier. First, note that we still need to
recover the environment after the block is typed, because we do not want the
block-local variables to be visible in the outer type environment. We recover
the type environment only after the block-enclosed statements type; moreover,
we also opportunistically yield a block type on the computation when we
discard the type environment recovery item. To account for the fact that the
block-enclosed statement can itself be a block (e.g., {{S}}), we would need an
additional rule. Since we do not like repetition, we instead group the types
block and stmt into one syntactic category, BlockOrStmtType, and now we
can have only one rule. We also include BlockOrStmtType in Type, as a
replacement for the two basic types.

The expression statement types as expected. Recall that we only allow
arithmetic expressions, which type to int, to be used as statements in IMP++.

The conditional was declared strict in all its arguments. Its typing policy
is that its first argument types to bool and its two branches to block.
If that is the case, then it yields a stmt type.

For while, its first argument should type to bool and its second to block.

Variable declarations add new bindings to the type environment. Recall that
we can only declare variables of integer type in IMP++.

The typing policy of print is that it can only print integer or string values,
and in that case it types to stmt. Like for BlockOrStmtType, to avoid
having two similar rules, one for int and another for string, we prefer to
introduce an additional syntactic category, PrintableType, which includes both
int and string types.

halt types to stmt; so its subsequent code is also typed.

join types to stmt, provided that its argument types to int.

Sequential composition was declared as a whitespace-separated sequentially
strict list. Its typing policy is that all the statements in the list should
type to stmt or block in order for the list to type to stmt. Since
lists are maintained internally as cons-lists, this is probably the simplest
way to do it:

rule .Stmts => stmt
rule _:BlockOrStmtType Ss => Ss

Note that the first rule, which types the empty sequence of statements to stmt,
is needed anyway, to type empty blocks {} (together with the block rule).

kompile imp.k and krun all the programs in Part 4 of the tutorial. They
should all type to stmt.

In the next lesson we will define a substitution-based type system for LAMBDA.

Go to Lesson 2, Type Systems: Substitution-Based Higher-Order Type Systems.

MOVIE (out of date) [10'11"]

Substitution-Based Higher-Order Type Systems

In this lesson you learn how to define a substitution-based type system for
a higher-order language, namely the LAMBDA language defined in Part 1 of the
tutorial.

Let us copy the definition of LAMBDA from Part 1 of the tutorial, Lesson 8.
We are going to modify it into a type systems for LAMBDA.

Before we start, it is important to clarify an important detail, namely that
our type system will yield a type checker when executed, not a type
inferencer. In particular, we are going to change the LAMBDA syntax
to allow us to associate a type to each declared variable. The
constructs which declare variables are lambda, let, letrec and mu.
The syntax of all these will therefore change.

Since here we are not interested in a LAMBDA semantics anymore, we take the
freedom to eliminate the Val syntactic category, our previous results.
Our new results are going to be the types, because programs will now reduce
to their types.

As explained, the syntax of the lambda construct needs to change, to also
declare the type of the variable that it binds. We add the new syntactic
category Type, with the following constructs: int, bool, the function
type (which gives it its higher-order status), and parentheses as bracket.
Also, we make types our K results.

We are now ready to define the typing rules.

Let us start with the typing rule for lambda abstraction: lambda X : T . E
types to the function type T -> T', where T' is the type obtained by further
typing E[T/X]. This can be elegantly achieved by reducing the lambda
abstraction to T -> E[T/X], provided that we extend the function type construct
to take expressions, not only types, as arguments, and to be strict.
This can be easily achieved by redeclaring it as a strict expression construct
(strictness in the second argument would suffice in this example, but it is
more uniform to define it strict overall).

The typing rule for application is as simple as it can get: (T1->T2) T1 => T2.

Let us now give the typing rules of arithmetic and Boolean expression
constructs. First, let us get rid of Val. Second, rewrite each value to its
type, similarly to the type system for IMP++ in the previous lesson. Third,
replace each semantic rule by its typing rule. Fourth, make sure you
do not forget to subsort Type to Exp, so your rules above will parse.

The typing policy of the conditional statement is that its first argument
should type to bool and its other two arguments should type to the same type
T, which will also be the result type of the conditional. So we make the
conditional construct strict in all its three arguments and we write the
obvious rule: if bool then T:Type else T => T. We want a runtime check that
the latter arguments are actually typed, so we write T:Type.

There is nothing special about let, except that we have to make sure we
change its syntax to account for the type of the variable that it binds.
This rule is a macro, so the let is desugared statically.

Similarly, the syntax of letrec and mu needs to change to account for the
type of the variable that they bind. The typing of letrec remains based on
its desugaring to mu; we have to make sure the types are also included now.

The typing policy of mu is that its body should type to the same type T of
its variable, which is also the type of the entire mu expression. This can
be elegantly achieved by rewriting it to (T -> T) E[T/X]. Recall that
application is strict, so E[T/X] will be eventually reduced to its type.
Then the application types correctly only if that type is also T, and in
that case the result type will also be T.

kompile and krun some programs. You can, for example, take the LAMBDA
programs from the first tutorial, modify them by adding types to their
variable declarations, and then type check them using krun.

In the next lesson we will discuss an environment-based type system
for LAMBDA.

Go to Lesson 3, Type Systems: Environment-Based Higher-Order Type Systems.

MOVIE (out of date) [6'52"]

Environment-Based Higher-Order Type Systems

In this lesson you learn how to define an environment-based type system for
a higher-order language, namely the LAMBDA language defined in Part 1 of the
tutorial.

The simplest and fastest way to proceed is to copy the substitution-based
type system of LAMBDA from the previous lesson and modify it into an
environment-based one. A large portion of the substitution-based definition
will remain unchanged. We only have to modify the rules that use
substitution.

We do not need the substitution anymore, so we can remove the require and
import statements. The syntax of types and expressions stays unchanged, but
we can now remove the binder tag of lambda.

Like in the type system of IMP++ in Lesson 1, we need a configuration that
contains, besides the <k/> cell, a <tenv/> cell that will hold the type
environment.

In an environment-based definition, unlike in a substitution-based one, we
need to lookup variables in the environment. So let us start with the
type lookup rule:

rule <k> X:Id => T ...</k> <tenv>... X |-> T ...</k>

The type environment is populated by the semantic rule of lambda:

rule <k> lambda X : T . E => (T -> E) ~> Rho ...</k>
     <tenv> Rho => Rho[X <- T] </tenv>

So X is bound to its type T in the type environment, and then T -> E
is scheduled for processing. Recall that the arrow type construct has been
extended into a strict expression construct, so E will be eventually reduced
to its type. Like in other environment-based definitions, we need to make
sure that we recover the type environment after the computation in the scope
of the declared variable terminates.

The typing rule of application does not change, so it stays as elegant as it
was in the substitution-based definition:

rule (T1 -> T2) T1 => T2

So do the rules for arithmetic and Boolean constructs, and those for the
if, and let, and letrec.

The mu rule needs to change, because it was previously defined using
substitution. We modify it in the same spirit as we modified the lambda
rule: bind X to its type in the environment, schedule its body for typing
in its right context, and then recover the type environment.

Finally, we give the semantics of environment recovery, making sure
the environment is recovered only after the preceding computation is
reduced to a type:

rule _:Type ~> (Rho => .) ... _ => Rho

The changes that we applied to the substitution-based definition were
therefore quite systematic: each substitution invocation was replaced with
an appropriate type environment update/recovery.

Go to Lesson 4, Type Systems: A Naive Substitution-Based Type Inferencer.

A Naive Substitution-Based Type Inferencer

In this lesson you learn how to define a naive substitution-based type
inferencer for a higher-order language, namely the LAMBDA language
defined in Part 1 of the tutorial.

Unlike in the type checker defined in Lessons 2 and 3, where we had to
associate a type with each declared variable, a type inferencer
attempts to infer the types of all the variables from the way those
variables are used. Let us take a look at this program, say plus.lambda:

lambda x . lambda y . x + y

Since x and y are used in an integer addition context, we can infer
that they must have the type int and the result of the addition is
also an int, so the type of the entire expression is int -> int -> int.
Similarly, the program if.lambda

lambda x . lambda y . lambda z .
  if x then y else z

can only make sense when x has type bool and y and z have the same
type, say t, in which case the type of the entire expression is
bool -> t -> t -> t. Since the type t can be anything, we say that
the type of this expression is polymorphic. That means that the code
above can be used in different contexts, where t can be an int, a
bool, a function type int -> int, and so on.

In the identity.lambda program

let f = lambda x . x
in f 1

f has such a polymorphic type, which is then applied to an integer,
so this program is type-safe and its type is int.

A typical polymorphic expression is the composition

lambda f . lambda g . lambda x .
  g (f x)

which has the type (t1 -> t2) -> (t2 -> t3) -> (t1 -> t3), polymorphic
in 3 types.

Let us now define our naive type inferencer and then we discuss more
examples. The idea is quite simple: we conceptually do the same
operations like we did within the type checker defined in Lesson 2,
with two important differences:

  1. instead of declaring a type with each declared variable, we assume
    a fresh type for that variable; and
  2. instead of checking that the types of expressions satisfy the
    type properties of the context in which they are used, we impose
    those properties as type equality constraints. A general-purpose
    unification-based constraint solving mechanism is then used to solve
    the generated type constraints.

Let us start with the syntax, which is essentially identical to that
of the type checker in Lesson 2, except that bound variables are not
declared a type anymore. Also, to keep things more compact, we put
all the Exp syntax declarations in one syntax declaration this time.

Before we modify the rules, let us first define our machinery for
adding and solving constraints. First, we require and import the
unification procedure. We do not discuss unification here, but if you
are interested you can consult the unification.k files under
k-distribution/include/kframework/builtin, which contains our current generic
definition of unification, which is written also in K. The generic unification
provides a sort, Mgu, for most-general-unifier, an operation
updateMgu(Mgu,T1,T2) which updates Mgu with additional constraints
generated by forcing the terms T1 and T2 to be equal, and an operation
applyMgu(Mgu,T) which applies Mgu to term T. For our use
of unification here, we do not even need to know how Mgu terms are
represented internally.

We define a K item construct, =, which takes two Type terms and
enforces them to be equal by means of updating the current Mgu.
Once the constraints are added to the Mgu, the equality dissolves
itself. With this semantics of = in mind, we can now go ahead and
modify the rules of the type checker systematically into rules
for a type inferencer. The changes are self-explanatory and
mechanical: for example, the rule

rule int * int => int

changes into rule

rule T1:Type  * T2:Type => T1 = int ~> T2 = int ~> int

generating the constraints that the two arguments of multiplication
have the type int, and the result type is int. Recall that each type
equality on the <k/> cell updates the current Mgu appropriately and
then dissolves itself; thus, the above says that after imposing the
constraints T1=int and T2=int, multiplication yields a type int.

As mentioned above, since types of variables are not declared anymore,
but inferred, we have to generate a fresh type for each variable at its
declaration time, and then generate appropriately constraints for it.
For example, the type semantics of lambda and mu become:

rule lambda X . E => T -> E[T/X]  when fresh(T:Type)
rule mu X . E => (T -> T) E[T/X]  when fresh(T:Type)

that is, we add a condition stating that the previously declared type
is now a fresh one. This type will be further constrained by how the
variable X is being used within E.

Interestingly, the previous typing rule for lambda application is not
powerful enough anymore. Indeed, since types are not given anymore,
it may very well be the case that the inferred type of the first
argument of the application construct is not yet a function type
(remember, for example, the program composition.lambda above). What
we have to do is to enforce it to be a function type, by means of
fresh types and constraints. We can introduce a fresh type for the
result of the application, and then write the expected rule as
follows:

rule T1:Type T2:Type => T1 = (T2 -> T) ~> T  when fresh(T:Type)

The conditional requires that its first argument is a bool and its
second and third arguments have the same type, which is also the
result type.

The macros do not change, in particular let is desugared into lambda
application. We will next see that this is a significant restriction,
because it limits the polymorphism of our type system.

We are done. We have a working type inferencer for LAMBDA.

Let's kompile it and krun the programs above. They all work as
expected. Let us also try some additional programs, to push it to its
limits.

First, let us test mu by means of a letrec example:

letrec f x = 3
in f

We can also try all the programs that we had in our first tutorial, on
lambda, for example the factorial.imp program:

letrec f x = if x <= 1 then 1 else (x * (f (x + -1)))
in (f 10)

Those programs are simple enough that they should all work as
expected with our naive type inferencer here.

Let us next try to type some tricky programs, which involve more
complex and indirect type constraints.

tricky-1.lambda:

lambda f . lambda x . lambda y . (
  (f x y) + x + (let x = y in x)
)

tricky-2.lambda:

lambda x .
  let f = lambda y . if true then y else x
  in (lambda x . f 0)

tricky-3.lambda:

lambda x . let f = lambda y . if true then x 7 else x y
           in f

tricky-4.lambda:

lambda x . let f = lambda x . x
           in let d = (f x) + 1
              in x

tricky-5.lambda:

lambda x . let f = lambda y . x y
           in let z = x 0 in f

It is now time to see the limitations of this naive type inferencer.
Consider the program

let id = lambda x . x
in if (id true) then (id 1) else (id 2)

Our type inferencer fails graciously with a clash in the <mgu/> cell
between int and bool. Indeed, the desugaring macro of let turns it
into a lambda and an application, which further enforce id to have a
type of the form t -> t for some fresh type t. The first use of id
in the condition of if will then constrain t to be bool, while the
other uses in the two branches will enforce t to be int. Thus the
clash in the <mgu/> cell.

Similarly, the program

let id = lambda x . x
in id id

yields a different kind of conflict: if id has type t -> t, in order
to apply id to itself it must be the case that its argument, t, equals
t -> t. These two type terms cannot be unified because there is a
circular dependence on t, so we get a cycle in the <mgu/> cell.

Both limitations above will be solved when we change the semantics of
let later on, to account for the desired polymorphism.

Before we conclude this lesson, let us see one more interesting
example, where the lack of let-polymorphism leads not to a type error,
but to a less generic type:

let f1 = lambda x . x in
  let f2 = f1 in
    let f3 = f2 in
      let f4 = f3 in
        let f5 = f4 in
          if (f5 true) then f2 else f3

Our current type inferencer will infer the type bool -> bool for the
program above. Nevertheless, since all functions f1, f2, f3, f4, f5
are the identity function, which is polymorphic, we would expect the
entire program to type to the same polymorphic identity function type.

This limitation will be also addressed when we define our
let-polymorphic type inferencer.

Before that, in the next lesson we will show how easily we can turn
the naive substitution-based type inferencer discussed in this lesson
into a similarly naive, but environment-based type inferencer.

Go to Lesson 5, Type Systems: A Naive Environment-Based Type Inferencer.

A Naive Environment-Based Type Inferencer

In this lesson you learn how to define a naive environment-based type
inferencer for a higher-order language. Specifically, we take the
substitution-based type inferencer for LAMBDA defined in Lesson 4 and
turn it into an environment-based one.

Recall from Lesson 3, where we defined an environment-based type
checker for LAMBDA based on the substitution-based one in Lesson 2,
that the transition from a substitution-based definition to an
environment-based one was quite systematic and mechanical: each
substitution occurrence E[T/X] is replaced by E, but at the same time
the variable X is bound to type T in the type environment. One benefit
of using type environments instead of substitution is that we replace
a linear complexity operation (the substitution) with a constant
complexity one (the variable lookup).

There is not much left to say which has not been already said in
Lesson 3: we remove the unnecessary binder annotations for the
variable binding operations, then add a <tenv/> cell to the
configuration to hold the type environment, then add a new rule for
variable lookup, and finally apply the transformation of substitutions
E[T/X] into E as explained above.

The resulting type inferencer should now work exactly the same way as
the substitution-based one, except, of course, that the resulting
configurations will contain a <tenv/> cell now.

As sanity check, let us consider two more LAMBDA programs that test
the static scoping nature of the inferencer. We do that because
faulty environment-based definitions often have this problem. The
program

let x = 1
in let f = lambda a . x
   in let x = true
      in f 3

should type to int, not to bool, and so it does. Similarly, the
program

let y = 0
in letrec f x = if x <= 0
                then y
                else let y = true
                     in f (x + 1)
   in f 1

should also type to int, not bool, and so it does, too.

The type inferencer defined in this lesson has the same limitations,
in terms of polymorphism, as the one in Lesson 4. In the next
lesson we will see how it can be parallelized, and in further lessons
how to make it polymorphic.

Go to Lesson 6, Type Systems: Parallel Type Checkers/Inferencers.

Parallel Type Checkers/Inferencers

In this lesson you learn how to define parallel type checkers or
inferencers. For the sake of a choice, we will parallelize the one in
the previous lesson, but the ideas are general. We are using the same
idea to define type checkers for other languages in the K tool
distribution, such as SIMPLE and KOOL.

The idea is in fact quite simple. Instead of one monolithic typing
task, we generate many smaller tasks, which can be processed in
parallel. We use the same approach to define parallel semantics as we
used for threads in IMP++ in Part 4 of the tutorial, that is, we add a
cell holding all the parallel tasks, making sure we declare the cell
holding a task with multiplicity *. For the particular type
inferencer that we chose here, the one in Lesson 5, each task will
hold an expression to type together with a type environment (so it
knows where to lookup its free variables). We have the following
configuration then:

configuration <tasks color="yellow">
                <task color="orange" multiplicity="*">
                  <k color="green"> $PGM:Exp </k>
                  <tenv color="red"> .Map </tenv>
                </task>
              </tasks>
              <mgu color="blue"> .Mgu </mgu>

Now we have to take each typing rule we had before and change it to
yield parallel typing. For example, our rule for typing
multiplication was the following in Lesson 5:

rule T1:Type * T2:Type => T1 = int ~> T2 = int ~> int

Since * was strict, its two arguments eventually type, and once that
happens the rule above fires. Unfortunately, the strictness of
multiplication makes the typing of the two expressions sequential in
our previous definition. To avoid typing the two expressions
sequentially and instead generating two parallel tasks, we remove the
strict attribute of multiplication and replace the rule above with the
following:

rule <k> E1 * E2 => int ...</k> <tenv> Rho </tenv>
     (. => <task> <k> E1 = int </k> <tenv> Rho </tenv> </task>
           <task> <k> E2 = int </k> <tenv> Rho </tenv> </task>)

Therefore, we generate two tasks for typing E1 and E2 in the same type
environment as the current task, and let the current task continue by
simply optimistically reducing E1*E2 to its expected result type, int.
If E1 or E2 will not type to int, then either their corresponding
tasks will get stuck or the <mgu/> cell will result into a clash or cycle,
so the program will not type overall in spite of the fact that we
allowed the task containing the multiplication to continue. This is
how we get maximum of parallelism in this case.

Before we continue, note that the new tasks hold equalities in them,
where one of its arguments is an expression, while previously the
equality construct was declared to take types. What we want now is
for the equality construct to possibly take any expressions, and first
type them and then generate the type constraint like before. This can
be done very easily by just extending the equality construct to
expressions and declaring it strict:

syntax KItem ::= Exp "=" Exp  [strict]

Unlike before, where we only passed types to the equality construct,
we now need a runtime check that its arguments are indeed types before
we can generate the updateMgu command:

rule <k> T:Type = T':Type => . ...</k>
     <mgu> Theta:Mgu => updateMgu(Theta,T,T') </mgu>

Like before, an equality will therefore update the <mgu/> cell and then
it dissolves itself, letting the <k/> cell in the corresponding task
empty. Such empty tasks are unnecessary, so they can be erased:

rule <task>... <k> . </k> ...</task> => .

We can now follow the same style as for multiplication to write the
parallel typing rules of the other arithmetic constructs, and even for
the conditional.

To parallelize the typing of lambda we generate two fresh types, one
for the variable and one for the body, and make sure that we generate
the correct type constraint and environment in the body task:

rule <k> lambda X . E => Tx -> Te ...</k> <tenv> TEnv </tenv>
     (. => <task> <k> E = Te </k> <tenv> TEnv[Tx/X] </tenv> </task>)
  when fresh(Tx:Type) andBool fresh(Te:Type)

Note that the above also allows us to not need to change and then
recover the environment of the current cell.

For function application we also need to generate two fresh types:

rule <k> E1 E2 => T ...</k> <tenv> Rho </tenv>
     (. => <task> <k> E1 = T2 -> T </k> <tenv> Rho </tenv> </task>
           <task> <k> E2 = T2 </k> <tenv> Rho </tenv> </task>)
  when fresh(T2:Type) andBool fresh(T:Type)

The only rule left is that of mu X . E. In this case we only need one
fresh type, because X, E and mu X . E have all the same type:

rule <k> mu X . E => T ...</k>  <tenv> TEnv </tenv>
     (. => <task> <k> E = T </k> <tenv> TEnv[T/X] </tenv> </task>)
  when fresh(T:Type)

We do not need the type environment recovery operation, so we delete it.

We can now kompile and krun all the programs that we typed in Lesson 5.
Everything should work.

In this lesson we only aimed at parallelizing the type inferencer in
Lesson 5, not to improve its expressiveness; it still has the same
limitations in terms of polymorphism. The next lessons are dedicated
to polymorphic type inferencers.

Go to Lesson 7, Type Systems: A Naive Substitution-based Polymorphic Type Inferencer.

A Naive Substitution-based Polymorphic Type Inferencer

In this lesson you learn how little it takes to turn a naive monomorphic
type inferencer into a naive polymorphic one, basically only changing
a few characters. In terms of the K framework, you will learn that
you can have complex combinations of substitutions in K, both over
expressions and over types.

Let us start directly with the change. All we have to do is to take
the LAMBDA type inferencer in Lesson 4 and only change the macro

rule let X = E in E' => (lambda X . E') E  [macro]

as follows:

rule let X = E in E' => E'[E/X]  [macro]

In other words, we are inlining the beta-reduction rule of
lambda-calculus within the original rule. In terms of typing,
the above forces the type inferencer to type E in place for each
occurrence of X in E'. Unlike in the first rule, where X had to get
one type only which satisfied the constrains of all X's occurrences in
E', we now never associate any type to X anymore.

Let us kompile and krun some examples. Everything that worked with
the type inferencer in Lesson 4 should still work here, although the
types of some programs can now be more general. For example, reconsider
the nested-lets.lambda program

let f1 = lambda x . x in
  let f2 = f1 in
    let f3 = f2 in
      let f4 = f3 in
        let f5 = f4 in
          if (f5 true) then f2 else f3

which was previously typed to bool -> bool. With the new rule above,
the sequence of lets is iteratively eliminated and we end up with the
program

if (lambda x . x) true then (lambda x . x) else (lambda x . x)

which now types (with both type inferencers) to a type of the form
t -> t, for some type variable t, which is more general than the
previous bool -> bool type that the program typed to in Lesson 4.

We can also now type programs that were not typable before, such as

let id = lambda x . x
in if (id true) then (id 1) else (id 2)

and

let id = lambda x . x
in id id

Let us also test it on some trickier programs, also not typable
before, such as

let f = lambda x . x
in let g = lambda y . f y
   in g g

which gives us a type of the form t -> t for some type variable t,
and as

let f = let g = lambda x . x
        in let h = lambda x . lambda x . (g g g g)
           in h
in f

which types to t1 -> t2 -> t3 -> t3 for some type variables t1, t2, t3.

Here is another program which was not typable before, which is
trickier than the others above in that a lambda-bound variable appears
free in a let-bound expression:

lambda x . (
  let y = lambda z . x
  in if (y true) then (y 1) else (y (lambda x . x))
)

The above presents no problem now, because once lambda z . x gets
substituted for y we get a well-typed expression which yields that x
has the type bool, so the entire expression types to bool -> bool.

The cheap type inferencer that we obtained above therefore works as
expected. However, it has two problems which justify a more advanced
solution. First, substitution is typically considered an elegant
mathematical instrument which is not too practical in implementations,
so an implementation of this type inferencer will likely be based on
type environments anyway. Additionally, we mix two kinds of
substitutions in this definition, one where we substitute types and
another where we substitute expressions, which can only make things
harder to implement efficiently. Second, our naive substitution of E
for X in E' can yield an exponential explosion in size of the original
program. Consider, for example, the following classic example which
is known to generate a type whose size is exponential in the size of
the program (and is thus used as an argument for why let-polymorphic
type inference is exponential in the worst-case):

let f00 = lambda x . lambda y . x in
  let f01 = lambda x . f00 (f00 x) in
    let f02 = lambda x . f01 (f01 x) in
      let f03 = lambda x . f02 (f02 x) in
        let f04 = lambda x . f03 (f03 x) in
          // ... you can add more nested lets here
          f04

The particular instance of the pattern above generates a type which
has 17 type variables! The desugaring of each let doubles the size of
the program and of its resulting type. While such programs are little
likely to appear in practice, it is often the case that functions can
be quite complex and large while their type can be quite simple in the
end, so we should simply avoid retyping each function each time it is
used.

This is precisely what we will do next. Before we present the classic
let-polymorphic type inferencer in Lesson 9, which is based on
environments, we first quickly discuss in Lesson 8 an intermediate
step, namely a naive environment-based variant of the inferencer
defined here.

Go to Lesson 8, Type Systems: A Naive Environment-based Polymorphic Type Inferencer.

A Naive Environment-based Polymorphic Type Inferencer

In this short lesson we discuss how to quickly turn a naive
environment-based monomorphic type inferencer into a naive let-polymorphic
one. Like in the previous lesson, we only need to change a few
characters. In terms of the K framework, you will learn how to have
both environments and substitution in the same definition.

Like in the previous lesson, all we have to do is to take the LAMBDA
type inferencer in Lesson 5 and only change the rule

rule let X = E in E' => (lambda X . E') E

as follows:

rule let X = E in E' => E'[E/X]

The reasons why this works have already been explained in the previous
lesson, so we do not repeat them here.

Since our new let rule uses substitution, we have to require the
substitution module at the top and also import SUBSTITUTION in the
current module, besides the already existing UNIFICATION.

Everything which worked with the type inferencer in Lesson 7 should
also work now. Let us only try the exponential type example,

let f00 = lambda x . lambda y . x in
  let f01 = lambda x . f00 (f00 x) in
    let f02 = lambda x . f01 (f01 x) in
      let f03 = lambda x . f02 (f02 x) in
        let f04 = lambda x . f03 (f03 x) in
          f04

As expected, this gives us precisely the same type as in Lesson 7.

So the only difference between this type inferencer and the one in
Lesson 7 is that substitution is only used for LAMBDA-to-LAMBDA
transformations, but not for infusing types within LAMBDA programs.
Thus, the syntax of LAMBDA programs is preserved intact, which some
may prefer. Nevertheless, this type inferencer is still expensive and
wasteful, because the let-bound expression is typed over and over
again in each place where the let-bound variable occurs.

In the next lesson we will discuss a type inferencer based on the
classic Damas-Hindley-Milner type system, which maximizes the reuse of
typing work by means of parametric types.

Go to Lesson 9, Type Systems: Let-Polymorphic Type Inferencer (Damas-Hindley-Milner).

Let-Polymorphic Type Inferencer (Damas-Hindley-Milner)

In this lesson we discuss a type inferencer based on what we call today
the Damas-Hindley-Milner type system, which is at the core of many
modern functional programming languages. The first variant of it was
proposed by Hindley in 1969, then, interestingly, Milner rediscovered
it in 1978 in the context of the ML language. Damas formalized it as
a type system in his PhD thesis in 1985. More specifically, our type
inferencer here, like many others as well as many implementations of
it, follows more closely the syntax-driven variant proposed by Clement
in 1987.

In terms of K, we will see how easily we can turn one definition which
is considered naive (our previous type inferencer in Lesson 8) into a
definition which is considered advanced. All we have to do is to
change one existing rule (the rule of the let binder) and to add a new
one. We will also learn some new predefined features of K, which make
the above possible.

The main idea is to replace the rule

rule let X = E in E' => E'[E/X]

which creates potentially many copies of E within E' with a rule
which types E once and then reuses that type in each place where X
occurs free in E'. The simplest K way to type E is to declare the
let construct strict(2). Now we cannot simply bind X to the type
of E, because we would obtain a variant of the naive type inferencer
we already discussed, together with its limitations, in Lesson 5 of this
tutorial. The trick here is to parameterize the type of E in all its
unconstrained fresh types, and then create fresh copies of those
parameters in each free occurrence of X in E'.

Let us discuss some examples, before we go into the technical details.
Consider the first let-polymorphic example which failed to be typed
with our first naive type-inferencer:

let id = lambda x . x
in if (id true) then (id 1) else (id 2)

When typing lambda x . x, we get a type of the form t -> t, for some
fresh type t. Instead of assigning this type to id as we did in the
naive type inferencers, we now first parametrize this type in its
fresh variable t, written

(forall t) t -> t

and then bind id to this parametric type. The intuition for the
parameter is that it can be instantiated with any other type, so this
parametric type stands, in fact, for infinitely many non-parametric
types. This is similar to what happens in formal logic proof systems,
where rule schemas stand for infinitely many concrete instances of
them. For this reason, parametric types are also called type schemas.

Now each time id is looked up within the let-body, we create a fresh
copy of the parameter t, which can this way be independently
constrained by each local context. Let's suppose that the three id
lookups yield the types t1 -> t1, t2 -> t2, and respectively t3 -> t3.
Then t1 will be constrained to be bool, and t2 and t3 to be int,
so we can now safely type the program above to int.

Therefore, a type schema comprises a summary of all the typing work
that has been done for typing the corresponding expression, and an
instantiation of its parameters with fresh copies represents an
elegant way to reuse all that typing work.

There are some subtleties regarding what fresh types can be made
parameters. Let us consider another example, discussed as part of
Lesson 7 on naive let-polymorphism:

lambda x . (
  let y = lambda z . x
  in if (y true) then (y 1) else (y (lambda x . x))
)

This program should type to bool -> bool, as explained in Lesson 7.
The lambda construct will bind x to some fresh type tx. Then the
let-bound expression lambda z . x types to tz -> tx for some
additional fresh type tz. The question now is what should the
parameters of this type be when we generate the type schema? If we
naively parameterize in all fresh variables, that is in both tz and
tx obtaining the type schema (forall tz,tx) tz -> tx, then there will
be no way to infer that the type of x, tx, must be a bool! The
inferred type of this expression would then wrongly be tx -> t for
some fresh types tx and t. That's because the parameters are replaced
with fresh copies in each occurrence of y, and thus their relationship
to the original x is completely lost. This tells us that we cannot
parameterize in all fresh types that appear in the type of the
let-bound expression. In particular, we cannot parameterize in those
which some variables are already bound to in the current type
environment (like x is bound to tx in our example above).
In our example, the correct type schema is (forall tz) tz -> tx,
which now allows us to correctly infer that tx is bool.

Let us now discuss another example, which should fail to type:

lambda x .
  let f = lambda y . x y
  in if (f true) then (f 1) else (f 2)

This should fail to type because lambda y . x y is equivalent to x,
so the conditional imposes the conflicting constraints that x should be
a function whose argument is either a bool or an int. Let us try to
type it using our currently informal procedure. Like in the previous
example, x will be bound to a fresh type tx. Then the let-bound
expression types to ty -> tz with ty and tz fresh types, adding also
the constraint tx = ty -> tz. What should the parameters of this type
be? If we ignore the type constraint and simply make both ty and tz
parameters because no variable is bound to them in the type
environment (indeed, the only variable x in the type environment is
bound to tx), then we can wrongly type this program to tx -> tz
following a reasoning similar to the one in the example above.
In fact, in this example, none of ty and tz can be parameters, because
they are constrained by tx.

The examples above tell us two things: first, that we have to take the
type constraints into account when deciding the parameters of the
schema; second, that after applying the most-general-unifier solution
given by the type constraints everywhere, the remaining fresh types
appearing anywhere in the type environment are consequently constrained
and cannot be turned into parameters. Since the type environment can in
fact also hold type schemas, which already bind some types, we only need
to ensure that none of the fresh types appearing free anywhere in the
type environment are turned into parameters of type schemas.

Thanks to generic support offered by the K tool, we can easily achieve
all the above as follows.

First, add syntax for type schemas:

syntax TypeSchema ::= "(" "forall" Set ")" Type  [binder]

The definition below will be given in such a way that the Set argument
of a type schema will always be a set of fresh types. We also declare
this construct to be a binder, so that we can make use of the generic
free variable function provided by the K tool.

We now replace the old rule for let

rule let X = E in E' => E'[E/X]

with the following rule:

rule <k> let X = T:Type in E => E ~> tenv(TEnv) ...</k>
     <mgu> Theta:Mgu </mgu>
     <tenv> TEnv
      => TEnv[(forall freeVariables(applyMgu(Theta, T)) -Set
                      freeVariables(applyMgu(Theta, values TEnv))
              ) applyMgu(Theta, T) / X]
     </tenv>

So the type T of E is being parameterized and then bound to X in the
type environment. The current mgu Theta, which comprises all the type
constraints accumulated so far, is applied to both T and the types in
the type environment. The remaining fresh types in T which do not
appear free in the type environment are then turned into type parameters.
The function freeVariables returns, as expected, the free variables of
its argument as a Set; this is why we declared the type schema to be a
binder above.

Now a LAMBDA variable in the type environment can be bound to either a
type or a type schema. In the first case, the previous rule we had
for variable lookup can be reused, but we have to make sure we check
that T there is of sort Type (adding a sort membership, for example).
In the second case, as explained above, we have to create fresh copies
of the parameters. This can be easily achieved with another
predefined K function, as follows:

rule <k> X:Id => freshVariables(Tvs,T) ...</k>
     <tenv>... X |-> (forall Tvs) T ...</tenv>

Indeed, freshVariables takes a set of variables and a term, and returns the
same term but with each of the given variables replaced by a fresh copy.

The operations freeVariables and freshVariables are useful in many K
definitions, so they are predefined in module substitution.k.

Our definition of this let-polymorphic type inferencer is now
complete. To test it, kompile it and then krun all the LAMBDA
programs discussed since Lesson 4. They should all work as expected.

K Languages

Here we present several "real-world" language examples. These languages
demonstrate many of the features you would expect to find in a full-fledged
programming language.

  • SIMPLE: Imperative programming language with threads.
  • KOOL: SIMPLE extended with object-oriented features.
  • FUN: A functional language with algebraic data-types and pattern-matching.
  • LOGIK: A logical programming language based on clause unification.

SIMPLE — Untyped

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K semantic definition of the untyped SIMPLE language.
SIMPLE is intended to be a pedagogical and research language that captures
the essence of the imperative programming paradigm, extended with several
features often encountered in imperative programming languages.
A program consists of a set of global variable declarations and
function definitions. Like in C, function definitions cannot be
nested and each program must have one function called main,
which is invoked when the program is executed. To make it more
interesting and to highlight some of K's strengths, SIMPLE includes
the following features in addition to the conventional imperative
expression and statement constructs:

  • Multidimensional arrays and array references. An array evaluates
    to an array reference, which is a special value holding a location (where
    the elements of the array start) together with the size of the array;
    the elements of the array can be array references themselves (particularly
    when the array is multi-dimensional). Array references are ordinary values,
    so they can be assigned to variables and passed/received by functions.

  • Functions and function values. Functions can have zero or
    more parameters and can return abruptly using a return statement.
    SIMPLE follows a call-by-value parameter passing style, with static scoping.
    Function names evaluate to function abstractions, which hereby become ordinary
    values in the language, same like the array references.

  • Blocks with locals. SIMPLE variables can be declared
    anywhere, their scope being from the place where they are declared
    until the end of the most nested enclosing block.

  • Input/Output. The expression read() evaluates to the
    next value in the input buffer, and the statement write(e)
    evaluates e and outputs its value to the output buffer. The
    input and output buffers are lists of values.

  • Exceptions. SIMPLE has parametric exceptions (the value thrown as
    an exception can be caught and bound).

  • Concurrency via dynamic thread creation/termination and
    synchronization. One can spawn a thread to execute any statement.
    The spawned thread shares with its parent its environment at creation time.
    Threads can be synchronized via a join command which blocks the current thread
    until the joined thread completes, via re-entrant locks which can be acquired
    and released, as well as through rendezvous commands.

Like in many other languages, some of SIMPLE's constructs can be
desugared into a smaller set of basic constructs. We do that at the end
of the syntax module, and then we only give semantics to the core constructs.

Note: This definition is commented slightly more than others, because it is
intended to be one of the first non-trivial definitions that the new
user of K sees. We recommend the beginner user to first check the
language definitions discussed in the K tutorial.

module SIMPLE-UNTYPED-SYNTAX
  imports DOMAINS-SYNTAX

Syntax

We start by defining the SIMPLE syntax. The language constructs discussed
above have the expected syntax and evaluation strategies. Recall that in K
we annotate the syntax with appropriate strictness attributes, thus giving
each language construct the desired evaluation strategy.

Identifiers

Recall from the K tutorial that identifiers are builtin and come under the
syntactic category Id. The special identifier for the function
main belongs to all programs, and plays a special role in the semantics,
so we declare it explicitly. This would not be necessary if the identifiers
were all included automatically in semantic definitions, but that is not
possible because of parsing reasons (e.g., K variables used to match
concrete identifiers would then be ambiguously parsed as identifiers). They
are only included in the parser generated to parse programs (and used by the
kast tool). Consequently, we have to explicitly declare all the
concrete identifiers that play a special role in the semantics, like
main below.

  syntax Id ::= "main" [token]

Declarations

There are two types of declarations: for variables (including arrays) and
for functions. We are going to allow declarations of the form
var x=10, a[10,10], y=23;, which is why we allow the var
keyword to take a list of expressions. The non-terminals used in the two
productions below are defined shortly.

  syntax Stmt ::= "var" Exps ";"
                | "function" Id "(" Ids ")" Block

Expressions

The expression constructs below are standard. Increment (++) takes
an expression rather than a variable because it can also increment an array
element. Recall that the syntax we define in K is what we call the syntax
of the semantics
: while powerful enough to define non-trivial syntaxes
(thanks to the underlying SDF technology that we use), we typically refrain
from defining precise syntaxes, that is, ones which accept precisely the
well-formed programs (that would not be possible anyway in general). That job
is deferred to type systems, which can also be defined in K. In other words,
we are not making any effort to guarantee syntactically that only variables
or array elements are passed to the increment construct, we allow any
expression. Nevertheless, we will only give semantics to those, so expressions
of the form ++5, which parse (but which will be rejected by our type
system in the typed version of SIMPLE later), will get stuck when executed.
Arrays can be multidimensional and can hold other arrays, so their
lookup operation takes a list of expressions as argument and applies to an
expression (which can in particular be another array lookup), respectively.
The construct sizeOf gives the size of an array in number of elements
of its first dimension. Note that almost all constructs are strict. The only
constructs which are not strict are the increment (since its first argument
gets updated, so it cannot be evaluated), the input read which takes no
arguments so strictness is irrelevant for it, the logical and and or constructs
which are short-circuited, the thread spawning construct which creates a new
thread executing the argument expression and return its unique identifier to
the creating thread (so it cannot just evaluate its argument in place), and the
assignment which is only strict in its second argument (for the same reason as
the increment).

  syntax Exp ::= Int | Bool | String | Id
               | "(" Exp ")"             [bracket]
               | "++" Exp
               > Exp "[" Exps "]"        [strict]
               > Exp "(" Exps ")"        [strict]
               | "-" Exp                 [strict]
               | "sizeOf" "(" Exp ")"    [strict]
               | "read" "(" ")"
               > left:
                 Exp "*" Exp             [strict, left]
               | Exp "/" Exp             [strict, left]
               | Exp "%" Exp             [strict, left]
               > left:
                 Exp "+" Exp             [strict, left]
               | Exp "-" Exp             [strict, left]
               > non-assoc:
                 Exp "<" Exp             [strict, non-assoc]
               | Exp "<=" Exp            [strict, non-assoc]
               | Exp ">" Exp             [strict, non-assoc]
               | Exp ">=" Exp            [strict, non-assoc]
               | Exp "==" Exp            [strict, non-assoc]
               | Exp "!=" Exp            [strict, non-assoc]
               > "!" Exp                 [strict]
               > left:
                 Exp "&&" Exp            [strict(1), left]
               | Exp "||" Exp            [strict(1), left]
               > "spawn" Block
               > Exp "=" Exp             [strict(2), right]

We also need comma-separated lists of identifiers and of expressions.
Moreover, we want them to be strict, that is, to evaluate to lists of results
whenever requested (e.g., when they appear as strict arguments of
the constructs above).

  syntax Ids  ::= List{Id,","}           [overload(Exps)]
  syntax Exps ::= List{Exp,","}          [overload(Exps), strict]  // automatically hybrid now
  syntax Exps ::= Ids
  syntax Val
  syntax Vals ::= List{Val,","}          [overload(Exps)]
  syntax Bottom
  syntax Bottoms ::= List{Bottom,","}    [overload(Exps)]
  syntax Ids ::= Bottoms

Statements

Most of the statement constructs are standard for imperative languages.
We syntactically distinguish between empty and non-empty blocks, because we
chose Stmts not to be a (;-separated) list of
Stmt. Variables can be declared anywhere inside a block, their scope
ending with the block. Expressions are allowed to be used for their side
effects only (followed by a semicolon ;). Functions are allowed
to abruptly return. The exceptions are parametric, i.e., one can throw a value
which is bound to the variable declared by catch. Threads can be
dynamically created and terminated, and can synchronize with join,
acquire, release and rendezvous. Note that the
strictness attributes obey the intended evaluation strategy of the various
constructs. In particular, the if-then-else construct is strict only in its
first argument (the if-then construct will be desugared into if-then-else),
while the loop constructs are not strict in any arguments. The print
statement construct is variadic, that is, it takes an arbitrary number of
arguments.

  syntax Block ::= "{" "}"
                | "{" Stmt "}"

  syntax Stmt ::= Block
                | Exp ";"                               [strict]
                | "if" "(" Exp ")" Block "else" Block   [avoid, strict(1)]
                | "if" "(" Exp ")" Block                [macro]
                | "while" "(" Exp ")" Block
                | "for" "(" Stmt Exp ";" Exp ")" Block  [macro]
                | "return" Exp ";"                      [strict]
                | "return" ";"                          [macro]
                | "print" "(" Exps ")" ";"              [strict]
// NOTE: print strict allows non-deterministic evaluation of its arguments
// Either keep like this but document, or otherwise make Exps seqstrict.
// Of define and use a different expression list here, which is seqstrict.
                | "try" Block "catch" "(" Id ")" Block
                | "throw" Exp ";"                       [strict]
                | "join" Exp ";"                        [strict]
                | "acquire" Exp ";"                     [strict]
                | "release" Exp ";"                     [strict]
                | "rendezvous" Exp ";"                  [strict]

The reason we allow Stmts as the first argument of for
instead of Stmt is because we want to allow more than one statement
to be executed when the loop is initialized. Also, as seens shorly, macros
may expand one statement into more statements; for example, an initialized
variable declaration statement var x=0; desugars into two statements,
namely var x; x=0;, so if we use Stmt instead of Stmts
in the production of for above then we risk that the macro expansion
of statement var x=0; happens before the macro expansion of for,
also shown below, in which case the latter would not apply anymore because
of syntactic mismatch.

  syntax Stmt ::= Stmt Stmt                          [right]

// I wish I were able to write the following instead, but confuses the parser.
//
// syntax Stmts ::= List{Stmt,""}
// syntax Top ::= Stmt | "function" Id "(" Ids ")" Block
// syntax Pgm ::= List{Top,""}
//
// With that, I could have also eliminated the empty block

Desugared Syntax

This part desugars some of SIMPLE's language constructs into core ones.
We only want to give semantics to core constructs, so we get rid of the
derived ones before we start the semantics. All desugaring macros below are
straightforward.

  rule if (E) S => if (E) S else {}
  rule for(Start Cond; Step) {S} => {Start while (Cond) {S Step;}}
  rule for(Start Cond; Step) {} => {Start while (Cond) {Step;}}
  rule var E1:Exp, E2:Exp, Es:Exps; => var E1; var E2, Es;
  rule var X:Id = E; => var X; X = E;

For the semantics, we can therefore assume from now on that each
conditional has both branches, that there are only while loops, and
that each variable is declared alone and without any initialization as part of
the declaration.

endmodule


module SIMPLE-UNTYPED
  imports SIMPLE-UNTYPED-SYNTAX
  imports DOMAINS

Basic Semantic Infrastructure

Before one starts adding semantic rules to a K definition, one needs to
define the basic semantic infrastructure consisting of definitions for
values and configuration. As discussed in the definitions
in the K tutorial, the values are needed to know when to stop applying
the heating rules and when to start applying the cooling rules corresponding
to strictness or context declarations. The configuration serves as a backbone
for the process of configuration abstraction which allows users to only
mention the relevant cells in each semantic rule, the rest of the configuration
context being inferred automatically. Although in some cases the configuration
could be automatically inferred from the rules, we believe that it is very
useful for language designers/semanticists to actually think of and design
their configuration explicitly, so the current implementation of K requires
one to define it.

Values

We here define the values of the language that the various fragments of
programs evaluate to. First, integers and Booleans are values. As discussed,
arrays evaluate to special array reference values holding (1) a location from
where the array's elements are contiguously allocated in the store, and
(2) the size of the array. Functions evaluate to function values as
λ-abstractions (we do not need to evaluate functions to closures
because each function is executed in the fixed global environment and
function definitions cannot be nested). Like in IMP and other
languages, we finally tell the tool that values are K results.

  syntax Val ::= Int | Bool | String
               | array(Int,Int)
               | lambda(Ids,Stmt)
  syntax Exp ::= Val
  syntax Exps ::= Vals
  syntax Vals ::= Bottoms
  syntax KResult ::= Val
                   | Vals  // TODO: should not need this

The inclusion of values in expressions follows the methodology of
syntactic definitions (like, e.g., in SOS): extend the syntax of the language
to encompass all values and additional constructs needed to give semantics.
In addition to that, it allows us to write the semantic rules using the
original syntax of the language, and to parse them with the same (now extended
with additional values) parser. If writing the semantics directly on the K
AST, using the associated labels instead of the syntactic constructs, then one
would not need to include values in expressions.

Configuration

The K configuration of SIMPLE consists of a top level cell, T,
holding a threads cell, a global environment map cell genv
mapping the global variables and function names to their locations, a shared
store map cell store mapping each location to some value, a set cell
busy holding the locks which have been acquired but not yet released
by threads, a set cell terminated holding the unique identifiers of
the threads which already terminated (needed for join), input
and output list cells, and a nextLoc cell holding a natural
number indicating the next available location. Unlike in the small languages
in the K tutorial, where we used the fresh predicate to generate fresh
locations, in larger languages, like SIMPLE, we prefer to explicitly manage
memory. The location counter in nextLoc models an actual physical
location in the store; for simplicity, we assume arbitrarily large memory and
no garbage collection. The threads cell contains one thread
cell for each existing thread in the program. Note that the thread cell has
multiplicity *, which means that at any given moment there could be zero,
one or more thread cells. Each thread cell contains a
computation cell k, a control cell holding the various
control structures needed to jump to certain points of interest in the program
execution, a local environment map cell env mapping the thread local
variables to locations in the store, and finally a holds map cell
indicating what locks have been acquired by the thread and not released so far
and how many times (SIMPLE's locks are re-entrant). The control cell
currently contains only two subcells, a function stack fstack which
is a list and an exception stack xstack which is also a list.
One can add more control structures in the control cell, such as a
stack for break/continue of loops, etc., if the language is extended with more
control-changing constructs. Note that all cells except for k are
also initialized, in that they contain a ground term of their corresponding
sort. The k cell is initialized with the program that will be passed
to the K tool, as indicated by the $PGM variable, followed by the
execute task (defined shortly).

  // the syntax declarations below are required because the sorts are
  // referenced directly by a production and, because of the way KIL to KORE
  // is implemented, the configuration syntax is not available yet
  // should simply work once KIL is removed completely
  // check other definitions for this hack as well

  syntax ControlCell
  syntax ControlCellFragment

  configuration <T color="red">
                  <threads color="orange">
                    <thread multiplicity="*" type="Map" color="yellow">
                      <id color="pink"> -1 </id>
                      <k color="green"> $PGM:Stmt ~> execute </k>
                    //<br/> // TODO(KORE): support latex annotations #1799
                      <control color="cyan">
                        <fstack color="blue"> .List </fstack>
                        <xstack color="purple"> .List </xstack>
                      </control>
                    //<br/> // TODO(KORE): support latex annotations #1799
                      <env color="violet"> .Map </env>
                      <holds color="black"> .Map </holds>
                    </thread>
                  </threads>
                //<br/> // TODO(KORE): support latex annotations #1799
                  <genv color="pink"> .Map </genv>
                  <store color="white"> .Map </store>
                  <busy color="cyan"> .Set </busy>
                  <terminated color="red"> .Set </terminated>
                //<br/> // TODO(KORE): support latex annotations #1799
                  <input color="magenta" stream="stdin"> .List </input>
                  <output color="brown" stream="stdout"> .List </output>
                  <nextLoc color="gray"> 0 </nextLoc>
                </T>

Declarations and Initialization

We start by defining the semantics of declarations (for variables,
arrays and functions).

Variable Declaration

The SIMPLE syntax was desugared above so that each variable is
declared alone and its initialization is done as a separate statement.
The semantic rule below matches resulting variable declarations of the
form var X; on top of the k cell
(indeed, note that the k cell is complete, or round, to the
left, and is torn, or ruptured, to the right), allocates a fresh
location L in the store which is initialized with a special value
(indeed, the unit ., or nothing, is matched anywhere
in the map ‒note the tears at both sides‒ and replaced with the
mapping L ↦ ⊥), and binds X to L in the local
environment shadowing previous declarations of X, if any.
This possible shadowing of X requires us to therefore update the
entire environment map, which is expensive and can significantly slow
down the execution of larger programs. On the other hand, since we know
that L is not already bound in the store, we simply add the binding
L ↦ ⊥ to the store, thus avoiding a potentially complete
traversal of the the store map in order to update it. We prefer the approach
used for updating the store whenever possible, because, in addition to being
faster, it offers more true concurrency than the latter; indeed, according
to the concurrent semantics of K, the store is not frozen while
L ↦ ⊥ is added to it, while the environment is frozen during the
update operation Env[L/X]. The variable declaration command is
also removed from the top of the computation cell and the fresh location
counter is incremented. The undefined symbol added in the store
is of sort KItem, instead of Val, on purpose; this way, the
store lookup rules will get stuck when one attempts to lookup an
uninitialized location. All the above happen in one transactional step,
with the rule below. Note also how configuration abstraction allows us to
only mention the needed cells; indeed, as the configuration above states,
the k and env cells are actually located within a
thread cell within the threads cell, but one needs
not mention these: the configuration context of the rule is
automatically transformed to match the declared configuration
structure.

  syntax KItem ::= "undefined"

  rule <k> var X:Id; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> undefined ...</store>
       <nextLoc> L => L +Int 1 </nextLoc>

Array Declaration

The K semantics of the uni-dimensional array declaration is somehow similar
to the above declaration of ordinary variables. First, note the
context declaration below, which requests the evaluation of the array
dimension. Once evaluated, say to a natural number N, then
N +Int 1 locations are allocated in the store for
an array of size N, the additional location (chosen to be the first
one allocated) holding the array reference value. The array reference
value array(L,N) states that the array has size N and its
elements are located contiguously in the store starting with location
L. The operation L … L' ↦ V, defined at the end of this
file in the auxiliary operation section, initializes each location in
the list L … L' to V. Note that, since the dimensions of
array declarations can be arbitrary expressions, this virtually means
that we can dynamically allocate memory in SIMPLE by means of array
declarations.

  context var _:Id[HOLE];

  rule <k> var X:Id[N:Int]; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> array(L +Int 1, N)
                          (L +Int 1) ... (L +Int N) |-> undefined ...</store>
       <nextLoc> L => L +Int 1 +Int N </nextLoc>
    requires N >=Int 0

SIMPLE allows multi-dimensional arrays. For semantic simplicity, we
desugar them all into uni-dimensional arrays by code transformation.
This way, we only need to give semantics to uni-dimensional arrays.
First, note that the context rule above actually evaluates all the array
dimensions (that's why we defined the expression lists strict!):
Upon evaluating the array dimensions, the code generation rule below
desugars multi-dimensional array declaration to uni-dimensional declarations.
To this aim, we introduce two special unique variable identifiers,
$1 and $2. The first variable, $1, iterates
through and initializes each element of the first dimension with an array
of the remaining dimensions, declared as variable $2:

  syntax Id ::= "$1" [token] | "$2" [token]
  rule var X:Id[N1:Int, N2:Int, Vs:Vals];
    => var X[N1];
       {
         for(var $1 = 0; $1 <= N1 - 1; ++$1) {
           var $2[N2, Vs];
           X[$1] = $2;
         }
       }

Ideally, one would like to perform syntactic desugarings like the one
above before the actual semantics. Unfortunately, that was not possible in
this case because the dimension expressions of the multi-dimensional array need
to be evaluated first. Indeed, the desugaring rule above does not work if the
dimensions of the declared array are arbitrary expressions, because they can
have side effects (e.g., a[++x,++x]) and those side effects would be
propagated each time the expression is evaluated in the desugaring code (note
that both the loop condition and the nested multi-dimensional declaration
would need to evaluate the expressions given as array dimensions).

Function declaration

Functions are evaluated to λ-abstractions and stored like any other
values in the store. A binding is added into the environment for the function
name to the location holding its body. Similarly to the C language, SIMPLE
only allows function declarations at the top level of the program. More
precisely, the subsequent semantics of SIMPLE only works well when one
respects this requirement. Indeed, the simplistic context-free parser
generated by the grammar above is more generous than we may want, in that it
allows function declarations anywhere any declaration is allowed, including
inside arbitrary blocks. However, as the rule below shows, we are not
storing the declaration environment with the λ-abstraction value as
closures do. Instead, as seen shortly, we switch to the global environment
whenever functions are invoked, which is consistent with our requirement that
functions should only be declared at the top. Thus, if one declares local
functions, then one may see unexpected behaviors (e.g., when one shadows a
global variable before declaring a local function). The type checker of
SIMPLE, also defined in K (see examples/simple/typed/static),
discards programs which do not respect this requirement.

  rule <k> function F(Xs) S => .K ...</k>
       <env> Env => Env[F <- L] </env>
       <store>... .Map => L |-> lambda(Xs, S) ...</store>
       <nextLoc> L => L +Int 1 </nextLoc>

When we are done with the first pass (pre-processing), the computation
cell k contains only the token execute (see the configuration
declaration above, where the computation item execute was placed
right after the program in the k cell of the initial configuration)
and the cell genv is empty. In this case, we have to call
main() and to initialize the global environment by transferring the
contents of the local environment into it. We prefer to do it this way, as
opposed to processing all the top level declarations directly within the global
environment, because we want to avoid duplication of semantics: the syntax of
the global declarations is identical to that of their corresponding local
declarations, so the semantics of the latter suffices provided that we copy
the local environment into the global one once we are done with the
pre-processing. We want this separate pre-processing step precisely because
we want to create the global environment. All (top-level) functions end up
having their names bound in the global environment and, as seen below, they
are executed in that same global environment; all these mean, in particular,
that the functions "see" each other, allowing for mutual recursion, etc.

  syntax KItem ::= "execute"
  rule <k> execute => main(.Exps); </k>
       <env> Env </env>
       <genv> .Map => Env </genv>

Expressions

We next define the K semantics of all the expression constructs.

Variable lookup

When a variable X is the first computational task, and X is bound to some
location L in the environment, and L is mapped to some value V in the
store, then we rewrite X into V:

  rule <k> X:Id => V ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V:Val ...</store>

Note that the rule above excludes reading , because is not
a value and V is checked at runtime to be a value.

Variable/Array increment

This is tricky, because we want to allow both ++x and ++a[5].
Therefore, we need to extract the lvalue of the expression to increment.
To do that, we state that the expression to increment should be wrapped
by the auxiliary lvalue operation and then evaluated. The semantics
of this auxiliary operation is defined at the end of this file. For now, all
we need to know is that it takes an expression and evaluates to a location
value. Location values, also defined at the end of the file, are integers
wrapped with the operation loc, to distinguish them from ordinary
integers.

  context ++(HOLE => lvalue(HOLE))
  rule <k> ++loc(L) => I +Int 1 ...</k>
       <store>... L |-> (I => I +Int 1) ...</store>

Arithmetic operators

There is nothing special about the following rules. They rewrite the
language constructs to their library counterparts when their arguments
become values of expected sorts:

  rule I1 + I2 => I1 +Int I2
  rule Str1 + Str2 => Str1 +String Str2
  rule I1 - I2 => I1 -Int I2
  rule I1 * I2 => I1 *Int I2
  rule I1 / I2 => I1 /Int I2 requires I2 =/=K 0
  rule I1 % I2 => I1 %Int I2 requires I2 =/=K 0
  rule - I => 0 -Int I
  rule I1 < I2 => I1 <Int I2
  rule I1 <= I2 => I1 <=Int I2
  rule I1 > I2 => I1 >Int I2
  rule I1 >= I2 => I1 >=Int I2

The equality and inequality constructs reduce to syntactic comparison
of the two argument values (which is what the equality on K terms does).

  rule V1:Val == V2:Val => V1 ==K V2
  rule V1:Val != V2:Val => V1 =/=K V2

The logical negation is clear, but the logical conjunction and disjunction
are short-circuited:

  rule ! T => notBool(T)
  rule true  && E => E
  rule false && _ => false
  rule true  || _ => true
  rule false || E => E

Array lookup

Untyped SIMPLE does not check array bounds (the dynamically typed version of
it, in examples/simple/typed/dynamic, does check for array out of
bounds). The first rule below desugars the multi-dimensional array access to
uni-dimensional array access; recall that the array access operation was
declared strict, so all sub-expressions involved are already values at this
stage. The second rule rewrites the array access to a lookup operation at a
precise location; we prefer to do it this way to avoid locking the store.
The semantics of the auxiliary lookup operation is straightforward,
and is defined at the end of the file.

// The [anywhere] feature is underused, because it would only be used
// at the top of the computation or inside the lvalue wrapper. So it
// may not be worth, or we may need to come up with a special notation
// allowing us to enumerate contexts for [anywhere] rules.
  rule V:Val[N1:Int, N2:Int, Vs:Vals] => V[N1][N2, Vs]
    [anywhere]

  rule array(L,_)[N:Int] => lookup(L +Int N)
    [anywhere]

Size of an array

The size of the array is stored in the array reference value, and the
sizeOf construct was declared strict, so:

  rule sizeOf(array(_,N)) => N

Function call

Function application was strict in both its arguments, so we can
assume that both the function and its arguments are evaluated to
values (the former expected to be a λ-abstraction). The first
rule below matches a well-formed function application on top of the
computation and performs the following steps atomically: it switches
to the function body followed by return; (for the case in
which the function does not use an explicit return statement); it
pushes the remaining computation, the current environment, and the
current control data onto the function stack (the remaining
computation can thus also be discarded from the computation cell,
because an unavoidable subsequent return statement ‒see
above‒ will always recover it from the stack); it switches the
current environment (which is being pushed on the function stack) to
the global environment, which is where the free variables in the
function body should be looked up; it binds the formal parameters to
fresh locations in the new environment, and stores the actual
arguments to those locations in the store (this latter step is easily
done by reducing the problem to variable declarations, whose semantics
we have already defined; the auxiliary operation mkDecls is
defined at the end of the file). The second rule pops the
computation, the environment and the control data from the function
stack when a return statement is encountered as the next
computational task, passing the returned value to the popped
computation (the popped computation was the context in which the
returning function was called). Note that the pushing/popping of the
control data is crucial. Without it, one may have a function that
contains an exception block with a return statement inside, which
would put the xstack cell in an inconsistent state (since the
exception block modifies it, but that modification should be
irrelevant once the function returns). We add an artificial
nothing value to the language, which is returned by the
nulary return; statements.

  syntax KItem ::=  (Map,K,ControlCellFragment)

  rule <k> lambda(Xs,S)(Vs:Vals) ~> K => mkDecls(Xs,Vs) S return; </k>
       <control>
         <fstack> .List => ListItem((Env,K,C)) ...</fstack>
         C
       </control>
       <env> Env => GEnv </env>
       <genv> GEnv </genv>

  rule <k> return(V:Val); ~> _ => V ~> K </k>
       <control>
         <fstack> ListItem((Env,K,C)) => .List ...</fstack>
         (_ => C)
       </control>
       <env> _ => Env </env>

  syntax Val ::= "nothing"
  rule return; => return nothing;

Like for division-by-zero, it is left unspecified what happens
when the nothing value is used in domain calculations. For
example, from the the perspective of the language semantics,
7 +Int nothing can evaluate to anything, or
may not evaluate at all (be undefined). If one wants to make sure that
such artificial values are never misused, then one needs to define a static
checker (also using K, like our the type checker in
examples/simple/typed/static) and reject programs that do.
Note that, unlike the undefined symbol which had the sort K
instead of Val, we defined nothing to be a value. That
is because, as explained above, we do not want the program to get
stuck when nothing is returned by a function. Instead, we want the
behavior to be unspecified; in particular, if one is careful to never
use the returned value in domain computation, like it happens when we
call a function for its side effects (e.g., with a statement of the
form f(x);), then the program does not get stuck.

Read

The read() expression construct simply evaluates to the next
input value, at the same time discarding the input value from the
in cell.

  rule <k> read() => I ...</k> <input> ListItem(I:Int) => .List ...</input>

Assignment

In SIMPLE, like in C, assignments are expression constructs and not statement
constructs. To make it a statement all one needs to do is to follow it by a
semi-colon ; (see the semantics for expression statements below).
Like for the increment, we want to allow assignments not only to variables but
also to array elements, e.g., e1[e2] = e3 where e1 evaluates
to an array reference, e2 to a natural number, and e3 to any
value. Thus, we first compute the lvalue of the left-hand-side expression
that appears in an assignment, and then we do the actual assignment to the
resulting location:

  context (HOLE => lvalue(HOLE)) = _

  rule <k> loc(L) = V:Val => V ...</k> <store>... L |-> (_ => V) ...</store>

Statements

We next define the K semantics of statements.

Blocks

Empty blocks are simply discarded, as shown in the first rule below.
For non-empty blocks, we schedule the enclosed statement but we have to
make sure the environment is recovered after the enclosed statement executes.
Recall that we allow local variable declarations, whose scope is the block
enclosing them. That is the reason for which we have to recover the
environment after the block. This allows us to have a very simple semantics
for variable declarations, as we did above. One can make the two rules below
computational if one wants them to count as computational steps.

  rule {} => .K
  rule <k> { S } => S ~> setEnv(Env) ...</k>  <env> Env </env>

The basic definition of environment recovery is straightforward and
given in the section on auxiliary constructs at the end of the file.

There are two common alternatives to the above semantics of blocks.
One is to keep track of the variables which are declared in the block and only
recover those at the end of the block. This way one does more work for
variable declarations but conceptually less work for environment recovery; we
say conceptually because it is not clear that it is indeed the case that
one does less work when AC matching is involved. The other alternative is to
work with a stack of environments instead of a flat environment, and push the
current environment when entering a block and pop it when exiting it. This
way, one does more work when accessing variables (since one has to search the
variable in the environment stack in a top-down manner), but on the other hand
uses smaller environments and the definition gets closer to an implementation.
Based on experience with dozens of language semantics and other K definitions,
we have found that our approach above is the best trade-off between elegance
and efficiency (especially since rewrite engines have built-in techniques to
lazily copy terms, by need, thus not creating unnecessary copies),
so it is the one that we follow in general.

Sequential composition

Sequential composition is desugared into K's builtin sequentialization
operation (recall that, like in C, the semi-colon ; is not a
statement separator in SIMPLE — it is either a statement terminator or a
construct for a statement from an expression). Note that K allows
to define the semantics of SIMPLE in such a way that statements eventually
dissolve from the top of the computation when they are completed; this is in
sharp contrast to (artificially) evaluating them to a special
skip statement value and then getting rid of that special value, as
it is the case in other semantic approaches (where everything must evaluate
to something). This means that once S₁ completes in the rule below, S₂
becomes automatically the next computation item without any additional
(explicit or implicit) rules.

  rule S1:Stmt S2:Stmt => S1 ~> S2

A subtle aspect of the rule above is that S₁ is declared to have sort
Stmts and not Stmt. That is because desugaring macros can indeed
produce left associative sequential composition of statements. For example,
the code var x=0; x=1; is desugared to
(var x; x=0;) x=1;, so although originally the first term of
the sequential composition had sort Stmt, after desugaring it became
of sort Stmts. Note that the attribute [right] associated
to the sequential compositon production is an attribute of the syntax, and not
of the semantics: e.g., it tells the parser to parse
var x; x=0; x=1; as var x; (x=0; x=1;), but it
does not tell the rewrite engine to rewrite (var x; x=0;) x=1; to
var x; (x=0; x=1;).

Expression statements

Expression statements are only used for their side effects, so their result
value is simply discarded. Common examples of expression statements are ones
of the form ++x;, x=e;, e1[e2]=e3;, etc.

  rule _:Val; => .K

Conditional

Since the conditional was declared with the strict(1) attribute, we
can assume that its first argument will eventually be evaluated. The rules
below cover the only two possibilities in which the conditional is allowed to
proceed (otherwise the rewriting process gets stuck).

  rule if ( true) S else _ => S
  rule if (false) _ else S => S

While loop

The simplest way to give the semantics of the while loop is by unrolling.
Note, however, that its unrolling is only allowed when the while loop reaches
the top of the computation (to avoid non-termination of unrolling). The
simple while loop semantics below works because our while loops in SIMPLE are
indeed very basic. If we allowed break/continue of loops then we would need
a completely different semantics, which would also involve the control cell.

  rule while (E) S => if (E) {S while(E)S}

Print

The print statement was strict, so all its arguments are now
evaluated (recall that print is variadic). We append each of
its evaluated arguments to the output buffer, and discard the residual
print statement with an empty list of arguments.

  rule <k> print(V:Val, Es => Es); ...</k> <output>... .List => ListItem(V) </output>
  rule print(.Vals); => .K

Exceptions

SIMPLE allows parametric exceptions, in that one can throw and catch a
particular value. The statement try S₁ catch(X) S₂
proceeds with the evaluation of S₁. If S₁ evaluates normally, i.e.,
without any exception thrown, then S₂ is discarded and the execution
continues normally. If S₁ throws an exception with a statement of the
form throw E, then E is first evaluated to some value V
(throw was declared to be strict), then V is bound to X, then
S₂ is evaluated in the new environment while the reminder of S₁ is
discarded, then the environment is recovered and the execution continues
normally with the statement following the try S₁ catch(X) S₂ statement.
Exceptions can be nested and the statements in the
catch part (S₂ in our case) can throw exceptions to the
upper level. One should be careful with how one handles the control data
structures here, so that the abrupt changes of control due to exception
throwing and to function returns interact correctly with each other.
For example, we want to allow function calls inside the statement S₁ in
a try S₁ catch(X) S₂ block which can throw an exception
that is not caught by the function but instead is propagated to the
try S₁ catch(X) S₂ block that called the function.
Therefore, we have to make sure that the function stack as well as other
potential control structures are also properly modified when the exception
is thrown to correctly recover the execution context. This can be easily
achieved by pushing/popping the entire current control context onto the
exception stack. The three rules below modularly do precisely the above.

  syntax KItem ::= (Id,Stmt,K,Map,ControlCellFragment)

  syntax KItem ::= "popx"

  rule <k> (try S1 catch(X) {S2} => S1 ~> popx) ~> K </k>
       <control>
         <xstack> .List => ListItem((X, S2, K, Env, C)) ...</xstack>
         C
       </control>
       <env> Env </env>

  rule <k> popx => .K ...</k>
       <xstack> ListItem(_) => .List ...</xstack>

  rule <k> throw V:Val; ~> _ => { var X = V; S2 } ~> K </k>
       <control>
         <xstack> ListItem((X, S2, K, Env, C)) => .List ...</xstack>
         (_ => C)
       </control>
       <env> _ => Env </env>

The catch statement S₂ needs to be executed in the original environment,
but where the thrown value V is bound to the catch variable X. We here
chose to rely on two previously defined constructs when giving semantics to
the catch part of the statement: (1) the variable declaration with
initialization, for binding X to V; and (2) the block construct for
preventing X from shadowing variables in the original environment upon the
completion of S₂.

Threads

SIMPLE's threads can be created and terminated dynamically, and can
synchronize by acquiring and releasing re-entrant locks and by rendezvous.
We discuss the seven rules giving the semantics of these operations below.

Thread creation

Threads can be created by any other threads using the spawn S
construct. The spawn expression construct evaluates to the unique identifier
of the newly created thread and, at the same time, a new thread cell is added
into the configuration, initialized with the S statement and sharing the
same environment with the parent thread. Note that the newly created
thread cell is torn. That means that the remaining cells are added
and initialized automatically as described in the definition of SIMPLE's
configuration. This is part of K's configuration abstraction mechanism.

  rule <thread>...
         <k> spawn S => !T:Int ...</k>
         <env> Env </env>
       ...</thread>
       (.Bag => <thread>...
               <k> S </k>
               <env> Env </env>
               <id> !T </id>
             ...</thread>)

Thread termination

Dually to the above, when a thread terminates its assigned computation (the
contents of its k cell) is empty, so the thread can be dissolved.
However, since no discipline is imposed on how locks are acquired and released,
it can be the case that a terminating thread still holds locks. Those locks
must be released, so other threads attempting to acquire them do not deadlock.
We achieve that by removing all the locks held by the terminating thread in its
holds cell from the set of busy locks in the busy cell
(keys(H) returns the domain of the map H as a set, that is, only
the locks themselves ignoring their multiplicity). As seen below, a lock is
added to the busy cell as soon as it is acquired for the first time
by a thread. The unique identifier of the terminated thread is also collected
into the terminated cell, so the join construct knows which
threads have terminated.

  rule (<thread>... <k>.K</k> <holds>H</holds> <id>T</id> ...</thread> => .Bag)
       <busy> Busy => Busy -Set keys(H) </busy>
       <terminated>... .Set => SetItem(T) ...</terminated>

Thread joining

Thread joining is now straightforward: all we need to do is to check whether
the identifier of the thread to be joined is in the terminated cell.
If yes, then the join statement dissolves and the joining thread
continues normally; if not, then the joining thread gets stuck.

  rule <k> join T:Int; => .K ...</k>
       <terminated>... SetItem(T) ...</terminated>

Acquire lock

There are two cases to distinguish when a thread attempts to acquire a lock
(in SIMPLE any value can be used as a lock):
(1) The thread does not currently have the lock, in which case it has to
take it provided that the lock is not already taken by another thread (see
the side condition of the first rule).
(2) The thread already has the lock, in which case it just increments its
counter for the lock (the locks are re-entrant). These two cases are captured
by the two rules below:

  rule <k> acquire V:Val; => .K ...</k>
       <holds>... .Map => V |-> 0 ...</holds>
       <busy> Busy (.Set => SetItem(V)) </busy>
    requires (notBool(V in Busy))

  rule <k> acquire V; => .K ...</k>
       <holds>... V:Val |-> (N => N +Int 1) ...</holds>

Release lock

Similarly, there are two corresponding cases to distinguish when a thread
releases a lock:
(1) The thread holds the lock more than once, in which case all it needs to do
is to decrement the lock counter.
(2) The thread holds the lock only once, in which case it needs to remove it
from its holds cell and also from the the shared busy cell,
so other threads can acquire it if they need to.

  rule <k> release V:Val; => .K ...</k>
       <holds>... V |-> (N => N -Int 1) ...</holds>
    requires N >Int 0

  rule <k> release V; => .K ...</k> <holds>... V:Val |-> 0 => .Map ...</holds>
       <busy>... SetItem(V) => .Set ...</busy>

Rendezvous synchronization

In addition to synchronization through acquire and release of locks, SIMPLE
also provides a construct for rendezvous synchronization. A thread whose next
statement to execute is rendezvous(V) gets stuck until another
thread reaches an identical statement; when that happens, the two threads
drop their rendezvous statements and continue their executions. If three
threads happen to have an identical rendezvous statement as their next
statement, then precisely two of them will synchronize and the other will
remain blocked until another thread reaches a similar rendezvous statement.
The rule below is as simple as it can be. Note, however, that, again, it is
K's mechanism for configuration abstraction that makes it work as desired:
since the only cell which can multiply containing a k cell inside is
the thread cell, the only way to concretize the rule below to the
actual configuration of SIMPLE is to include each k cell in a
thread cell.

  rule <k> rendezvous V:Val; => .K ...</k>
       <k> rendezvous V; => .K ...</k>

Auxiliary declarations and operations

In this section we define all the auxiliary constructs used in the
above semantics.

Making declarations

The mkDecls auxiliary construct turns a list of identifiers
and a list of values in a sequence of corresponding variable
declarations.

  syntax Stmt ::= mkDecls(Ids,Vals)  [function]
  rule mkDecls((X:Id, Xs:Ids), (V:Val, Vs:Vals)) => var X=V; mkDecls(Xs,Vs)
  rule mkDecls(.Ids,.Vals) => {}

Location lookup

The operation below is straightforward.

  syntax Exp ::= lookup(Int)
  rule <k> lookup(L) => V ...</k> <store>... L |-> V:Val ...</store>

Environment recovery

We have already discussed the environment recovery auxiliary operation in the
IMP++ tutorial:

// TODO: eliminate the env wrapper, like we did in IMP++

  syntax KItem ::= setEnv(Map)
  rule <k> setEnv(Env) => .K ...</k> <env> _ => Env </env>

While theoretically sufficient, the basic definition for environment
recovery alone is suboptimal. Consider a loop while (E)S,
whose semantics (see above) was given by unrolling. S
is a block. Then the semantics of blocks above, together with the
unrolling semantics of the while loop, will yield a computation
structure in the k cell that increasingly grows, adding a new
environment recovery task right in front of the already existing sequence of
similar environment recovery tasks (this phenomenon is similar to the ``tail
recursion'' problem). Of course, when we have a sequence of environment
recovery tasks, we only need to keep the last one. The elegant rule below
does precisely that, thus avoiding the unnecessary computation explosion
problem:

  rule (setEnv(_) => .K) ~> setEnv(_)

In fact, the above follows a common convention in K for recovery
operations of cell contents: the meaning of a computation task of the form
cell(C) that reaches the top of the computation is that the current
contents of cell cell is discarded and gets replaced with C. We
did not add support for these special computation tasks in our current
implementation of K, so we need to define them as above.

lvalue and loc

For convenience in giving the semantics of constructs like the increment and
the assignment, that we want to operate the same way on variables and on
array elements, we used an auxiliary lvalue(E) construct which was
expected to evaluate to the lvalue of the expression E. This is only
defined when E has an lvalue, that is, when E is either a variable or
evaluates to an array element. lvalue(E) evaluates to a value of
the form loc(L), where L is the location where the value of E
can be found; for clarity, we use loc to structurally distinguish
natural numbers from location values. In giving semantics to lvalue
there are two cases to consider. (1) If E is a variable, then all we need
to do is to grab its location from the environment. (2) If E is an array
element, then we first evaluate the array and its index in order to identify
the exact location of the element of concern, and then return that location;
the last rule below works because its preceding context declarations ensure
that the array and its index are evaluated, and then the rule for array lookup
(defined above) rewrites the evaluated array access construct to its
corresponding store lookup operation.

// For parsing reasons, we prefer to allow lvalue to take a K

  syntax Exp ::= lvalue(K)
  syntax Val ::= loc(Int)

// Local variable

  rule <k> lvalue(X:Id => loc(L)) ...</k> <env>... X |-> L:Int ...</env>

// Array element: evaluate the array and its index;
// then the array lookup rule above applies.

  context lvalue(_::Exp[HOLE::Exps])
  context lvalue(HOLE::Exp[_::Exps])

// Finally, return the address of the desired object member

  rule lvalue(lookup(L:Int) => loc(L))

Initializing multiple locations

The following operation initializes a sequence of locations with the same
value:

  syntax Map ::= Int "..." Int "|->" K [function]
  rule N...M |-> _ => .Map  requires N >Int M
  rule N...M |-> K => N |-> K (N +Int 1)...M |-> K  requires N <=Int M

The semantics of SIMPLE is now complete. Make sure you kompile the
definition with the right options in order to generate the desired model.
No kompile options are needed if you only only want to execute the definition
(and thus get an interpreter), but if you want to search for a different
program behaviors then you need to kompile with the --enable-search option

endmodule

Go to Lesson 2, SIMPLE typed static

SIMPLE — Untyped

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K semantic definition of the untyped SIMPLE language.
SIMPLE is intended to be a pedagogical and research language that captures
the essence of the imperative programming paradigm, extended with several
features often encountered in imperative programming languages.
A program consists of a set of global variable declarations and
function definitions. Like in C, function definitions cannot be
nested and each program must have one function called main,
which is invoked when the program is executed. To make it more
interesting and to highlight some of K's strengths, SIMPLE includes
the following features in addition to the conventional imperative
expression and statement constructs:

  • Multidimensional arrays and array references. An array evaluates
    to an array reference, which is a special value holding a location (where
    the elements of the array start) together with the size of the array;
    the elements of the array can be array references themselves (particularly
    when the array is multi-dimensional). Array references are ordinary values,
    so they can be assigned to variables and passed/received by functions.

  • Functions and function values. Functions can have zero or
    more parameters and can return abruptly using a return statement.
    SIMPLE follows a call-by-value parameter passing style, with static scoping.
    Function names evaluate to function abstractions, which hereby become ordinary
    values in the language, same like the array references.

  • Blocks with locals. SIMPLE variables can be declared
    anywhere, their scope being from the place where they are declared
    until the end of the most nested enclosing block.

  • Input/Output. The expression read() evaluates to the
    next value in the input buffer, and the statement write(e)
    evaluates e and outputs its value to the output buffer. The
    input and output buffers are lists of values.

  • Exceptions. SIMPLE has parametric exceptions (the value thrown as
    an exception can be caught and bound).

  • Concurrency via dynamic thread creation/termination and
    synchronization. One can spawn a thread to execute any statement.
    The spawned thread shares with its parent its environment at creation time.
    Threads can be synchronized via a join command which blocks the current thread
    until the joined thread completes, via re-entrant locks which can be acquired
    and released, as well as through rendezvous commands.

Like in many other languages, some of SIMPLE's constructs can be
desugared into a smaller set of basic constructs. We do that at the end
of the syntax module, and then we only give semantics to the core constructs.

Note: This definition is commented slightly more than others, because it is
intended to be one of the first non-trivial definitions that the new
user of K sees. We recommend the beginner user to first check the
language definitions discussed in the K tutorial.

module SIMPLE-UNTYPED-SYNTAX
  imports DOMAINS-SYNTAX

Syntax

We start by defining the SIMPLE syntax. The language constructs discussed
above have the expected syntax and evaluation strategies. Recall that in K
we annotate the syntax with appropriate strictness attributes, thus giving
each language construct the desired evaluation strategy.

Identifiers

Recall from the K tutorial that identifiers are builtin and come under the
syntactic category Id. The special identifier for the function
main belongs to all programs, and plays a special role in the semantics,
so we declare it explicitly. This would not be necessary if the identifiers
were all included automatically in semantic definitions, but that is not
possible because of parsing reasons (e.g., K variables used to match
concrete identifiers would then be ambiguously parsed as identifiers). They
are only included in the parser generated to parse programs (and used by the
kast tool). Consequently, we have to explicitly declare all the
concrete identifiers that play a special role in the semantics, like
main below.

  syntax Id ::= "main" [token]

Declarations

There are two types of declarations: for variables (including arrays) and
for functions. We are going to allow declarations of the form
var x=10, a[10,10], y=23;, which is why we allow the var
keyword to take a list of expressions. The non-terminals used in the two
productions below are defined shortly.

  syntax Stmt ::= "var" Exps ";"
                | "function" Id "(" Ids ")" Block

Expressions

The expression constructs below are standard. Increment (++) takes
an expression rather than a variable because it can also increment an array
element. Recall that the syntax we define in K is what we call the syntax
of the semantics
: while powerful enough to define non-trivial syntaxes
(thanks to the underlying SDF technology that we use), we typically refrain
from defining precise syntaxes, that is, ones which accept precisely the
well-formed programs (that would not be possible anyway in general). That job
is deferred to type systems, which can also be defined in K. In other words,
we are not making any effort to guarantee syntactically that only variables
or array elements are passed to the increment construct, we allow any
expression. Nevertheless, we will only give semantics to those, so expressions
of the form ++5, which parse (but which will be rejected by our type
system in the typed version of SIMPLE later), will get stuck when executed.
Arrays can be multidimensional and can hold other arrays, so their
lookup operation takes a list of expressions as argument and applies to an
expression (which can in particular be another array lookup), respectively.
The construct sizeOf gives the size of an array in number of elements
of its first dimension. Note that almost all constructs are strict. The only
constructs which are not strict are the increment (since its first argument
gets updated, so it cannot be evaluated), the input read which takes no
arguments so strictness is irrelevant for it, the logical and and or constructs
which are short-circuited, the thread spawning construct which creates a new
thread executing the argument expression and return its unique identifier to
the creating thread (so it cannot just evaluate its argument in place), and the
assignment which is only strict in its second argument (for the same reason as
the increment).

  syntax Exp ::= Int | Bool | String | Id
               | "(" Exp ")"             [bracket]
               | "++" Exp
               > Exp "[" Exps "]"        [strict]
               > Exp "(" Exps ")"        [strict]
               | "-" Exp                 [strict]
               | "sizeOf" "(" Exp ")"    [strict]
               | "read" "(" ")"
               > left:
                 Exp "*" Exp             [strict, left]
               | Exp "/" Exp             [strict, left]
               | Exp "%" Exp             [strict, left]
               > left:
                 Exp "+" Exp             [strict, left]
               | Exp "-" Exp             [strict, left]
               > non-assoc:
                 Exp "<" Exp             [strict, non-assoc]
               | Exp "<=" Exp            [strict, non-assoc]
               | Exp ">" Exp             [strict, non-assoc]
               | Exp ">=" Exp            [strict, non-assoc]
               | Exp "==" Exp            [strict, non-assoc]
               | Exp "!=" Exp            [strict, non-assoc]
               > "!" Exp                 [strict]
               > left:
                 Exp "&&" Exp            [strict(1), left]
               | Exp "||" Exp            [strict(1), left]
               > "spawn" Block
               > Exp "=" Exp             [strict(2), right]

We also need comma-separated lists of identifiers and of expressions.
Moreover, we want them to be strict, that is, to evaluate to lists of results
whenever requested (e.g., when they appear as strict arguments of
the constructs above).

  syntax Ids  ::= List{Id,","}           [overload(Exps)]
  syntax Exps ::= List{Exp,","}          [overload(Exps), strict]  // automatically hybrid now
  syntax Exps ::= Ids
  syntax Val
  syntax Vals ::= List{Val,","}          [overload(Exps)]
  syntax Bottom
  syntax Bottoms ::= List{Bottom,","}    [overload(Exps)]
  syntax Ids ::= Bottoms

Statements

Most of the statement constructs are standard for imperative languages.
We syntactically distinguish between empty and non-empty blocks, because we
chose Stmts not to be a (;-separated) list of
Stmt. Variables can be declared anywhere inside a block, their scope
ending with the block. Expressions are allowed to be used for their side
effects only (followed by a semicolon ;). Functions are allowed
to abruptly return. The exceptions are parametric, i.e., one can throw a value
which is bound to the variable declared by catch. Threads can be
dynamically created and terminated, and can synchronize with join,
acquire, release and rendezvous. Note that the
strictness attributes obey the intended evaluation strategy of the various
constructs. In particular, the if-then-else construct is strict only in its
first argument (the if-then construct will be desugared into if-then-else),
while the loop constructs are not strict in any arguments. The print
statement construct is variadic, that is, it takes an arbitrary number of
arguments.

  syntax Block ::= "{" "}"
                | "{" Stmt "}"

  syntax Stmt ::= Block
                | Exp ";"                               [strict]
                | "if" "(" Exp ")" Block "else" Block   [avoid, strict(1)]
                | "if" "(" Exp ")" Block                [macro]
                | "while" "(" Exp ")" Block
                | "for" "(" Stmt Exp ";" Exp ")" Block  [macro]
                | "return" Exp ";"                      [strict]
                | "return" ";"                          [macro]
                | "print" "(" Exps ")" ";"              [strict]
// NOTE: print strict allows non-deterministic evaluation of its arguments
// Either keep like this but document, or otherwise make Exps seqstrict.
// Of define and use a different expression list here, which is seqstrict.
                | "try" Block "catch" "(" Id ")" Block
                | "throw" Exp ";"                       [strict]
                | "join" Exp ";"                        [strict]
                | "acquire" Exp ";"                     [strict]
                | "release" Exp ";"                     [strict]
                | "rendezvous" Exp ";"                  [strict]

The reason we allow Stmts as the first argument of for
instead of Stmt is because we want to allow more than one statement
to be executed when the loop is initialized. Also, as seens shorly, macros
may expand one statement into more statements; for example, an initialized
variable declaration statement var x=0; desugars into two statements,
namely var x; x=0;, so if we use Stmt instead of Stmts
in the production of for above then we risk that the macro expansion
of statement var x=0; happens before the macro expansion of for,
also shown below, in which case the latter would not apply anymore because
of syntactic mismatch.

  syntax Stmt ::= Stmt Stmt                          [right]

// I wish I were able to write the following instead, but confuses the parser.
//
// syntax Stmts ::= List{Stmt,""}
// syntax Top ::= Stmt | "function" Id "(" Ids ")" Block
// syntax Pgm ::= List{Top,""}
//
// With that, I could have also eliminated the empty block

Desugared Syntax

This part desugars some of SIMPLE's language constructs into core ones.
We only want to give semantics to core constructs, so we get rid of the
derived ones before we start the semantics. All desugaring macros below are
straightforward.

  rule if (E) S => if (E) S else {}
  rule for(Start Cond; Step) {S} => {Start while (Cond) {S Step;}}
  rule for(Start Cond; Step) {} => {Start while (Cond) {Step;}}
  rule var E1:Exp, E2:Exp, Es:Exps; => var E1; var E2, Es;
  rule var X:Id = E; => var X; X = E;

For the semantics, we can therefore assume from now on that each
conditional has both branches, that there are only while loops, and
that each variable is declared alone and without any initialization as part of
the declaration.

endmodule


module SIMPLE-UNTYPED
  imports SIMPLE-UNTYPED-SYNTAX
  imports DOMAINS

Basic Semantic Infrastructure

Before one starts adding semantic rules to a K definition, one needs to
define the basic semantic infrastructure consisting of definitions for
values and configuration. As discussed in the definitions
in the K tutorial, the values are needed to know when to stop applying
the heating rules and when to start applying the cooling rules corresponding
to strictness or context declarations. The configuration serves as a backbone
for the process of configuration abstraction which allows users to only
mention the relevant cells in each semantic rule, the rest of the configuration
context being inferred automatically. Although in some cases the configuration
could be automatically inferred from the rules, we believe that it is very
useful for language designers/semanticists to actually think of and design
their configuration explicitly, so the current implementation of K requires
one to define it.

Values

We here define the values of the language that the various fragments of
programs evaluate to. First, integers and Booleans are values. As discussed,
arrays evaluate to special array reference values holding (1) a location from
where the array's elements are contiguously allocated in the store, and
(2) the size of the array. Functions evaluate to function values as
λ-abstractions (we do not need to evaluate functions to closures
because each function is executed in the fixed global environment and
function definitions cannot be nested). Like in IMP and other
languages, we finally tell the tool that values are K results.

  syntax Val ::= Int | Bool | String
               | array(Int,Int)
               | lambda(Ids,Stmt)
  syntax Exp ::= Val
  syntax Exps ::= Vals
  syntax Vals ::= Bottoms
  syntax KResult ::= Val
                   | Vals  // TODO: should not need this

The inclusion of values in expressions follows the methodology of
syntactic definitions (like, e.g., in SOS): extend the syntax of the language
to encompass all values and additional constructs needed to give semantics.
In addition to that, it allows us to write the semantic rules using the
original syntax of the language, and to parse them with the same (now extended
with additional values) parser. If writing the semantics directly on the K
AST, using the associated labels instead of the syntactic constructs, then one
would not need to include values in expressions.

Configuration

The K configuration of SIMPLE consists of a top level cell, T,
holding a threads cell, a global environment map cell genv
mapping the global variables and function names to their locations, a shared
store map cell store mapping each location to some value, a set cell
busy holding the locks which have been acquired but not yet released
by threads, a set cell terminated holding the unique identifiers of
the threads which already terminated (needed for join), input
and output list cells, and a nextLoc cell holding a natural
number indicating the next available location. Unlike in the small languages
in the K tutorial, where we used the fresh predicate to generate fresh
locations, in larger languages, like SIMPLE, we prefer to explicitly manage
memory. The location counter in nextLoc models an actual physical
location in the store; for simplicity, we assume arbitrarily large memory and
no garbage collection. The threads cell contains one thread
cell for each existing thread in the program. Note that the thread cell has
multiplicity *, which means that at any given moment there could be zero,
one or more thread cells. Each thread cell contains a
computation cell k, a control cell holding the various
control structures needed to jump to certain points of interest in the program
execution, a local environment map cell env mapping the thread local
variables to locations in the store, and finally a holds map cell
indicating what locks have been acquired by the thread and not released so far
and how many times (SIMPLE's locks are re-entrant). The control cell
currently contains only two subcells, a function stack fstack which
is a list and an exception stack xstack which is also a list.
One can add more control structures in the control cell, such as a
stack for break/continue of loops, etc., if the language is extended with more
control-changing constructs. Note that all cells except for k are
also initialized, in that they contain a ground term of their corresponding
sort. The k cell is initialized with the program that will be passed
to the K tool, as indicated by the $PGM variable, followed by the
execute task (defined shortly).

  // the syntax declarations below are required because the sorts are
  // referenced directly by a production and, because of the way KIL to KORE
  // is implemented, the configuration syntax is not available yet
  // should simply work once KIL is removed completely
  // check other definitions for this hack as well

  syntax ControlCell
  syntax ControlCellFragment

  configuration <T color="red">
                  <threads color="orange">
                    <thread multiplicity="*" type="Map" color="yellow">
                      <id color="pink"> -1 </id>
                      <k color="green"> $PGM:Stmt ~> execute </k>
                    //<br/> // TODO(KORE): support latex annotations #1799
                      <control color="cyan">
                        <fstack color="blue"> .List </fstack>
                        <xstack color="purple"> .List </xstack>
                      </control>
                    //<br/> // TODO(KORE): support latex annotations #1799
                      <env color="violet"> .Map </env>
                      <holds color="black"> .Map </holds>
                    </thread>
                  </threads>
                //<br/> // TODO(KORE): support latex annotations #1799
                  <genv color="pink"> .Map </genv>
                  <store color="white"> .Map </store>
                  <busy color="cyan"> .Set </busy>
                  <terminated color="red"> .Set </terminated>
                //<br/> // TODO(KORE): support latex annotations #1799
                  <input color="magenta" stream="stdin"> .List </input>
                  <output color="brown" stream="stdout"> .List </output>
                  <nextLoc color="gray"> 0 </nextLoc>
                </T>

Declarations and Initialization

We start by defining the semantics of declarations (for variables,
arrays and functions).

Variable Declaration

The SIMPLE syntax was desugared above so that each variable is
declared alone and its initialization is done as a separate statement.
The semantic rule below matches resulting variable declarations of the
form var X; on top of the k cell
(indeed, note that the k cell is complete, or round, to the
left, and is torn, or ruptured, to the right), allocates a fresh
location L in the store which is initialized with a special value
(indeed, the unit ., or nothing, is matched anywhere
in the map ‒note the tears at both sides‒ and replaced with the
mapping L ↦ ⊥), and binds X to L in the local
environment shadowing previous declarations of X, if any.
This possible shadowing of X requires us to therefore update the
entire environment map, which is expensive and can significantly slow
down the execution of larger programs. On the other hand, since we know
that L is not already bound in the store, we simply add the binding
L ↦ ⊥ to the store, thus avoiding a potentially complete
traversal of the the store map in order to update it. We prefer the approach
used for updating the store whenever possible, because, in addition to being
faster, it offers more true concurrency than the latter; indeed, according
to the concurrent semantics of K, the store is not frozen while
L ↦ ⊥ is added to it, while the environment is frozen during the
update operation Env[L/X]. The variable declaration command is
also removed from the top of the computation cell and the fresh location
counter is incremented. The undefined symbol added in the store
is of sort KItem, instead of Val, on purpose; this way, the
store lookup rules will get stuck when one attempts to lookup an
uninitialized location. All the above happen in one transactional step,
with the rule below. Note also how configuration abstraction allows us to
only mention the needed cells; indeed, as the configuration above states,
the k and env cells are actually located within a
thread cell within the threads cell, but one needs
not mention these: the configuration context of the rule is
automatically transformed to match the declared configuration
structure.

  syntax KItem ::= "undefined"

  rule <k> var X:Id; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> undefined ...</store>
       <nextLoc> L => L +Int 1 </nextLoc>

Array Declaration

The K semantics of the uni-dimensional array declaration is somehow similar
to the above declaration of ordinary variables. First, note the
context declaration below, which requests the evaluation of the array
dimension. Once evaluated, say to a natural number N, then
N +Int 1 locations are allocated in the store for
an array of size N, the additional location (chosen to be the first
one allocated) holding the array reference value. The array reference
value array(L,N) states that the array has size N and its
elements are located contiguously in the store starting with location
L. The operation L … L' ↦ V, defined at the end of this
file in the auxiliary operation section, initializes each location in
the list L … L' to V. Note that, since the dimensions of
array declarations can be arbitrary expressions, this virtually means
that we can dynamically allocate memory in SIMPLE by means of array
declarations.

  context var _:Id[HOLE];

  rule <k> var X:Id[N:Int]; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> array(L +Int 1, N)
                          (L +Int 1) ... (L +Int N) |-> undefined ...</store>
       <nextLoc> L => L +Int 1 +Int N </nextLoc>
    requires N >=Int 0

SIMPLE allows multi-dimensional arrays. For semantic simplicity, we
desugar them all into uni-dimensional arrays by code transformation.
This way, we only need to give semantics to uni-dimensional arrays.
First, note that the context rule above actually evaluates all the array
dimensions (that's why we defined the expression lists strict!):
Upon evaluating the array dimensions, the code generation rule below
desugars multi-dimensional array declaration to uni-dimensional declarations.
To this aim, we introduce two special unique variable identifiers,
$1 and $2. The first variable, $1, iterates
through and initializes each element of the first dimension with an array
of the remaining dimensions, declared as variable $2:

  syntax Id ::= "$1" [token] | "$2" [token]
  rule var X:Id[N1:Int, N2:Int, Vs:Vals];
    => var X[N1];
       {
         for(var $1 = 0; $1 <= N1 - 1; ++$1) {
           var $2[N2, Vs];
           X[$1] = $2;
         }
       }

Ideally, one would like to perform syntactic desugarings like the one
above before the actual semantics. Unfortunately, that was not possible in
this case because the dimension expressions of the multi-dimensional array need
to be evaluated first. Indeed, the desugaring rule above does not work if the
dimensions of the declared array are arbitrary expressions, because they can
have side effects (e.g., a[++x,++x]) and those side effects would be
propagated each time the expression is evaluated in the desugaring code (note
that both the loop condition and the nested multi-dimensional declaration
would need to evaluate the expressions given as array dimensions).

Function declaration

Functions are evaluated to λ-abstractions and stored like any other
values in the store. A binding is added into the environment for the function
name to the location holding its body. Similarly to the C language, SIMPLE
only allows function declarations at the top level of the program. More
precisely, the subsequent semantics of SIMPLE only works well when one
respects this requirement. Indeed, the simplistic context-free parser
generated by the grammar above is more generous than we may want, in that it
allows function declarations anywhere any declaration is allowed, including
inside arbitrary blocks. However, as the rule below shows, we are not
storing the declaration environment with the λ-abstraction value as
closures do. Instead, as seen shortly, we switch to the global environment
whenever functions are invoked, which is consistent with our requirement that
functions should only be declared at the top. Thus, if one declares local
functions, then one may see unexpected behaviors (e.g., when one shadows a
global variable before declaring a local function). The type checker of
SIMPLE, also defined in K (see examples/simple/typed/static),
discards programs which do not respect this requirement.

  rule <k> function F(Xs) S => .K ...</k>
       <env> Env => Env[F <- L] </env>
       <store>... .Map => L |-> lambda(Xs, S) ...</store>
       <nextLoc> L => L +Int 1 </nextLoc>

When we are done with the first pass (pre-processing), the computation
cell k contains only the token execute (see the configuration
declaration above, where the computation item execute was placed
right after the program in the k cell of the initial configuration)
and the cell genv is empty. In this case, we have to call
main() and to initialize the global environment by transferring the
contents of the local environment into it. We prefer to do it this way, as
opposed to processing all the top level declarations directly within the global
environment, because we want to avoid duplication of semantics: the syntax of
the global declarations is identical to that of their corresponding local
declarations, so the semantics of the latter suffices provided that we copy
the local environment into the global one once we are done with the
pre-processing. We want this separate pre-processing step precisely because
we want to create the global environment. All (top-level) functions end up
having their names bound in the global environment and, as seen below, they
are executed in that same global environment; all these mean, in particular,
that the functions "see" each other, allowing for mutual recursion, etc.

  syntax KItem ::= "execute"
  rule <k> execute => main(.Exps); </k>
       <env> Env </env>
       <genv> .Map => Env </genv>

Expressions

We next define the K semantics of all the expression constructs.

Variable lookup

When a variable X is the first computational task, and X is bound to some
location L in the environment, and L is mapped to some value V in the
store, then we rewrite X into V:

  rule <k> X:Id => V ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V:Val ...</store>

Note that the rule above excludes reading , because is not
a value and V is checked at runtime to be a value.

Variable/Array increment

This is tricky, because we want to allow both ++x and ++a[5].
Therefore, we need to extract the lvalue of the expression to increment.
To do that, we state that the expression to increment should be wrapped
by the auxiliary lvalue operation and then evaluated. The semantics
of this auxiliary operation is defined at the end of this file. For now, all
we need to know is that it takes an expression and evaluates to a location
value. Location values, also defined at the end of the file, are integers
wrapped with the operation loc, to distinguish them from ordinary
integers.

  context ++(HOLE => lvalue(HOLE))
  rule <k> ++loc(L) => I +Int 1 ...</k>
       <store>... L |-> (I => I +Int 1) ...</store>

Arithmetic operators

There is nothing special about the following rules. They rewrite the
language constructs to their library counterparts when their arguments
become values of expected sorts:

  rule I1 + I2 => I1 +Int I2
  rule Str1 + Str2 => Str1 +String Str2
  rule I1 - I2 => I1 -Int I2
  rule I1 * I2 => I1 *Int I2
  rule I1 / I2 => I1 /Int I2 requires I2 =/=K 0
  rule I1 % I2 => I1 %Int I2 requires I2 =/=K 0
  rule - I => 0 -Int I
  rule I1 < I2 => I1 <Int I2
  rule I1 <= I2 => I1 <=Int I2
  rule I1 > I2 => I1 >Int I2
  rule I1 >= I2 => I1 >=Int I2

The equality and inequality constructs reduce to syntactic comparison
of the two argument values (which is what the equality on K terms does).

  rule V1:Val == V2:Val => V1 ==K V2
  rule V1:Val != V2:Val => V1 =/=K V2

The logical negation is clear, but the logical conjunction and disjunction
are short-circuited:

  rule ! T => notBool(T)
  rule true  && E => E
  rule false && _ => false
  rule true  || _ => true
  rule false || E => E

Array lookup

Untyped SIMPLE does not check array bounds (the dynamically typed version of
it, in examples/simple/typed/dynamic, does check for array out of
bounds). The first rule below desugars the multi-dimensional array access to
uni-dimensional array access; recall that the array access operation was
declared strict, so all sub-expressions involved are already values at this
stage. The second rule rewrites the array access to a lookup operation at a
precise location; we prefer to do it this way to avoid locking the store.
The semantics of the auxiliary lookup operation is straightforward,
and is defined at the end of the file.

// The [anywhere] feature is underused, because it would only be used
// at the top of the computation or inside the lvalue wrapper. So it
// may not be worth, or we may need to come up with a special notation
// allowing us to enumerate contexts for [anywhere] rules.
  rule V:Val[N1:Int, N2:Int, Vs:Vals] => V[N1][N2, Vs]
    [anywhere]

  rule array(L,_)[N:Int] => lookup(L +Int N)
    [anywhere]

Size of an array

The size of the array is stored in the array reference value, and the
sizeOf construct was declared strict, so:

  rule sizeOf(array(_,N)) => N

Function call

Function application was strict in both its arguments, so we can
assume that both the function and its arguments are evaluated to
values (the former expected to be a λ-abstraction). The first
rule below matches a well-formed function application on top of the
computation and performs the following steps atomically: it switches
to the function body followed by return; (for the case in
which the function does not use an explicit return statement); it
pushes the remaining computation, the current environment, and the
current control data onto the function stack (the remaining
computation can thus also be discarded from the computation cell,
because an unavoidable subsequent return statement ‒see
above‒ will always recover it from the stack); it switches the
current environment (which is being pushed on the function stack) to
the global environment, which is where the free variables in the
function body should be looked up; it binds the formal parameters to
fresh locations in the new environment, and stores the actual
arguments to those locations in the store (this latter step is easily
done by reducing the problem to variable declarations, whose semantics
we have already defined; the auxiliary operation mkDecls is
defined at the end of the file). The second rule pops the
computation, the environment and the control data from the function
stack when a return statement is encountered as the next
computational task, passing the returned value to the popped
computation (the popped computation was the context in which the
returning function was called). Note that the pushing/popping of the
control data is crucial. Without it, one may have a function that
contains an exception block with a return statement inside, which
would put the xstack cell in an inconsistent state (since the
exception block modifies it, but that modification should be
irrelevant once the function returns). We add an artificial
nothing value to the language, which is returned by the
nulary return; statements.

  syntax KItem ::=  (Map,K,ControlCellFragment)

  rule <k> lambda(Xs,S)(Vs:Vals) ~> K => mkDecls(Xs,Vs) S return; </k>
       <control>
         <fstack> .List => ListItem((Env,K,C)) ...</fstack>
         C
       </control>
       <env> Env => GEnv </env>
       <genv> GEnv </genv>

  rule <k> return(V:Val); ~> _ => V ~> K </k>
       <control>
         <fstack> ListItem((Env,K,C)) => .List ...</fstack>
         (_ => C)
       </control>
       <env> _ => Env </env>

  syntax Val ::= "nothing"
  rule return; => return nothing;

Like for division-by-zero, it is left unspecified what happens
when the nothing value is used in domain calculations. For
example, from the the perspective of the language semantics,
7 +Int nothing can evaluate to anything, or
may not evaluate at all (be undefined). If one wants to make sure that
such artificial values are never misused, then one needs to define a static
checker (also using K, like our the type checker in
examples/simple/typed/static) and reject programs that do.
Note that, unlike the undefined symbol which had the sort K
instead of Val, we defined nothing to be a value. That
is because, as explained above, we do not want the program to get
stuck when nothing is returned by a function. Instead, we want the
behavior to be unspecified; in particular, if one is careful to never
use the returned value in domain computation, like it happens when we
call a function for its side effects (e.g., with a statement of the
form f(x);), then the program does not get stuck.

Read

The read() expression construct simply evaluates to the next
input value, at the same time discarding the input value from the
in cell.

  rule <k> read() => I ...</k> <input> ListItem(I:Int) => .List ...</input>

Assignment

In SIMPLE, like in C, assignments are expression constructs and not statement
constructs. To make it a statement all one needs to do is to follow it by a
semi-colon ; (see the semantics for expression statements below).
Like for the increment, we want to allow assignments not only to variables but
also to array elements, e.g., e1[e2] = e3 where e1 evaluates
to an array reference, e2 to a natural number, and e3 to any
value. Thus, we first compute the lvalue of the left-hand-side expression
that appears in an assignment, and then we do the actual assignment to the
resulting location:

  context (HOLE => lvalue(HOLE)) = _

  rule <k> loc(L) = V:Val => V ...</k> <store>... L |-> (_ => V) ...</store>

Statements

We next define the K semantics of statements.

Blocks

Empty blocks are simply discarded, as shown in the first rule below.
For non-empty blocks, we schedule the enclosed statement but we have to
make sure the environment is recovered after the enclosed statement executes.
Recall that we allow local variable declarations, whose scope is the block
enclosing them. That is the reason for which we have to recover the
environment after the block. This allows us to have a very simple semantics
for variable declarations, as we did above. One can make the two rules below
computational if one wants them to count as computational steps.

  rule {} => .K
  rule <k> { S } => S ~> setEnv(Env) ...</k>  <env> Env </env>

The basic definition of environment recovery is straightforward and
given in the section on auxiliary constructs at the end of the file.

There are two common alternatives to the above semantics of blocks.
One is to keep track of the variables which are declared in the block and only
recover those at the end of the block. This way one does more work for
variable declarations but conceptually less work for environment recovery; we
say conceptually because it is not clear that it is indeed the case that
one does less work when AC matching is involved. The other alternative is to
work with a stack of environments instead of a flat environment, and push the
current environment when entering a block and pop it when exiting it. This
way, one does more work when accessing variables (since one has to search the
variable in the environment stack in a top-down manner), but on the other hand
uses smaller environments and the definition gets closer to an implementation.
Based on experience with dozens of language semantics and other K definitions,
we have found that our approach above is the best trade-off between elegance
and efficiency (especially since rewrite engines have built-in techniques to
lazily copy terms, by need, thus not creating unnecessary copies),
so it is the one that we follow in general.

Sequential composition

Sequential composition is desugared into K's builtin sequentialization
operation (recall that, like in C, the semi-colon ; is not a
statement separator in SIMPLE — it is either a statement terminator or a
construct for a statement from an expression). Note that K allows
to define the semantics of SIMPLE in such a way that statements eventually
dissolve from the top of the computation when they are completed; this is in
sharp contrast to (artificially) evaluating them to a special
skip statement value and then getting rid of that special value, as
it is the case in other semantic approaches (where everything must evaluate
to something). This means that once S₁ completes in the rule below, S₂
becomes automatically the next computation item without any additional
(explicit or implicit) rules.

  rule S1:Stmt S2:Stmt => S1 ~> S2

A subtle aspect of the rule above is that S₁ is declared to have sort
Stmts and not Stmt. That is because desugaring macros can indeed
produce left associative sequential composition of statements. For example,
the code var x=0; x=1; is desugared to
(var x; x=0;) x=1;, so although originally the first term of
the sequential composition had sort Stmt, after desugaring it became
of sort Stmts. Note that the attribute [right] associated
to the sequential compositon production is an attribute of the syntax, and not
of the semantics: e.g., it tells the parser to parse
var x; x=0; x=1; as var x; (x=0; x=1;), but it
does not tell the rewrite engine to rewrite (var x; x=0;) x=1; to
var x; (x=0; x=1;).

Expression statements

Expression statements are only used for their side effects, so their result
value is simply discarded. Common examples of expression statements are ones
of the form ++x;, x=e;, e1[e2]=e3;, etc.

  rule _:Val; => .K

Conditional

Since the conditional was declared with the strict(1) attribute, we
can assume that its first argument will eventually be evaluated. The rules
below cover the only two possibilities in which the conditional is allowed to
proceed (otherwise the rewriting process gets stuck).

  rule if ( true) S else _ => S
  rule if (false) _ else S => S

While loop

The simplest way to give the semantics of the while loop is by unrolling.
Note, however, that its unrolling is only allowed when the while loop reaches
the top of the computation (to avoid non-termination of unrolling). The
simple while loop semantics below works because our while loops in SIMPLE are
indeed very basic. If we allowed break/continue of loops then we would need
a completely different semantics, which would also involve the control cell.

  rule while (E) S => if (E) {S while(E)S}

Print

The print statement was strict, so all its arguments are now
evaluated (recall that print is variadic). We append each of
its evaluated arguments to the output buffer, and discard the residual
print statement with an empty list of arguments.

  rule <k> print(V:Val, Es => Es); ...</k> <output>... .List => ListItem(V) </output>
  rule print(.Vals); => .K

Exceptions

SIMPLE allows parametric exceptions, in that one can throw and catch a
particular value. The statement try S₁ catch(X) S₂
proceeds with the evaluation of S₁. If S₁ evaluates normally, i.e.,
without any exception thrown, then S₂ is discarded and the execution
continues normally. If S₁ throws an exception with a statement of the
form throw E, then E is first evaluated to some value V
(throw was declared to be strict), then V is bound to X, then
S₂ is evaluated in the new environment while the reminder of S₁ is
discarded, then the environment is recovered and the execution continues
normally with the statement following the try S₁ catch(X) S₂ statement.
Exceptions can be nested and the statements in the
catch part (S₂ in our case) can throw exceptions to the
upper level. One should be careful with how one handles the control data
structures here, so that the abrupt changes of control due to exception
throwing and to function returns interact correctly with each other.
For example, we want to allow function calls inside the statement S₁ in
a try S₁ catch(X) S₂ block which can throw an exception
that is not caught by the function but instead is propagated to the
try S₁ catch(X) S₂ block that called the function.
Therefore, we have to make sure that the function stack as well as other
potential control structures are also properly modified when the exception
is thrown to correctly recover the execution context. This can be easily
achieved by pushing/popping the entire current control context onto the
exception stack. The three rules below modularly do precisely the above.

  syntax KItem ::= (Id,Stmt,K,Map,ControlCellFragment)

  syntax KItem ::= "popx"

  rule <k> (try S1 catch(X) {S2} => S1 ~> popx) ~> K </k>
       <control>
         <xstack> .List => ListItem((X, S2, K, Env, C)) ...</xstack>
         C
       </control>
       <env> Env </env>

  rule <k> popx => .K ...</k>
       <xstack> ListItem(_) => .List ...</xstack>

  rule <k> throw V:Val; ~> _ => { var X = V; S2 } ~> K </k>
       <control>
         <xstack> ListItem((X, S2, K, Env, C)) => .List ...</xstack>
         (_ => C)
       </control>
       <env> _ => Env </env>

The catch statement S₂ needs to be executed in the original environment,
but where the thrown value V is bound to the catch variable X. We here
chose to rely on two previously defined constructs when giving semantics to
the catch part of the statement: (1) the variable declaration with
initialization, for binding X to V; and (2) the block construct for
preventing X from shadowing variables in the original environment upon the
completion of S₂.

Threads

SIMPLE's threads can be created and terminated dynamically, and can
synchronize by acquiring and releasing re-entrant locks and by rendezvous.
We discuss the seven rules giving the semantics of these operations below.

Thread creation

Threads can be created by any other threads using the spawn S
construct. The spawn expression construct evaluates to the unique identifier
of the newly created thread and, at the same time, a new thread cell is added
into the configuration, initialized with the S statement and sharing the
same environment with the parent thread. Note that the newly created
thread cell is torn. That means that the remaining cells are added
and initialized automatically as described in the definition of SIMPLE's
configuration. This is part of K's configuration abstraction mechanism.

  rule <thread>...
         <k> spawn S => !T:Int ...</k>
         <env> Env </env>
       ...</thread>
       (.Bag => <thread>...
               <k> S </k>
               <env> Env </env>
               <id> !T </id>
             ...</thread>)

Thread termination

Dually to the above, when a thread terminates its assigned computation (the
contents of its k cell) is empty, so the thread can be dissolved.
However, since no discipline is imposed on how locks are acquired and released,
it can be the case that a terminating thread still holds locks. Those locks
must be released, so other threads attempting to acquire them do not deadlock.
We achieve that by removing all the locks held by the terminating thread in its
holds cell from the set of busy locks in the busy cell
(keys(H) returns the domain of the map H as a set, that is, only
the locks themselves ignoring their multiplicity). As seen below, a lock is
added to the busy cell as soon as it is acquired for the first time
by a thread. The unique identifier of the terminated thread is also collected
into the terminated cell, so the join construct knows which
threads have terminated.

  rule (<thread>... <k>.K</k> <holds>H</holds> <id>T</id> ...</thread> => .Bag)
       <busy> Busy => Busy -Set keys(H) </busy>
       <terminated>... .Set => SetItem(T) ...</terminated>

Thread joining

Thread joining is now straightforward: all we need to do is to check whether
the identifier of the thread to be joined is in the terminated cell.
If yes, then the join statement dissolves and the joining thread
continues normally; if not, then the joining thread gets stuck.

  rule <k> join T:Int; => .K ...</k>
       <terminated>... SetItem(T) ...</terminated>

Acquire lock

There are two cases to distinguish when a thread attempts to acquire a lock
(in SIMPLE any value can be used as a lock):
(1) The thread does not currently have the lock, in which case it has to
take it provided that the lock is not already taken by another thread (see
the side condition of the first rule).
(2) The thread already has the lock, in which case it just increments its
counter for the lock (the locks are re-entrant). These two cases are captured
by the two rules below:

  rule <k> acquire V:Val; => .K ...</k>
       <holds>... .Map => V |-> 0 ...</holds>
       <busy> Busy (.Set => SetItem(V)) </busy>
    requires (notBool(V in Busy))

  rule <k> acquire V; => .K ...</k>
       <holds>... V:Val |-> (N => N +Int 1) ...</holds>

Release lock

Similarly, there are two corresponding cases to distinguish when a thread
releases a lock:
(1) The thread holds the lock more than once, in which case all it needs to do
is to decrement the lock counter.
(2) The thread holds the lock only once, in which case it needs to remove it
from its holds cell and also from the the shared busy cell,
so other threads can acquire it if they need to.

  rule <k> release V:Val; => .K ...</k>
       <holds>... V |-> (N => N -Int 1) ...</holds>
    requires N >Int 0

  rule <k> release V; => .K ...</k> <holds>... V:Val |-> 0 => .Map ...</holds>
       <busy>... SetItem(V) => .Set ...</busy>

Rendezvous synchronization

In addition to synchronization through acquire and release of locks, SIMPLE
also provides a construct for rendezvous synchronization. A thread whose next
statement to execute is rendezvous(V) gets stuck until another
thread reaches an identical statement; when that happens, the two threads
drop their rendezvous statements and continue their executions. If three
threads happen to have an identical rendezvous statement as their next
statement, then precisely two of them will synchronize and the other will
remain blocked until another thread reaches a similar rendezvous statement.
The rule below is as simple as it can be. Note, however, that, again, it is
K's mechanism for configuration abstraction that makes it work as desired:
since the only cell which can multiply containing a k cell inside is
the thread cell, the only way to concretize the rule below to the
actual configuration of SIMPLE is to include each k cell in a
thread cell.

  rule <k> rendezvous V:Val; => .K ...</k>
       <k> rendezvous V; => .K ...</k>

Auxiliary declarations and operations

In this section we define all the auxiliary constructs used in the
above semantics.

Making declarations

The mkDecls auxiliary construct turns a list of identifiers
and a list of values in a sequence of corresponding variable
declarations.

  syntax Stmt ::= mkDecls(Ids,Vals)  [function]
  rule mkDecls((X:Id, Xs:Ids), (V:Val, Vs:Vals)) => var X=V; mkDecls(Xs,Vs)
  rule mkDecls(.Ids,.Vals) => {}

Location lookup

The operation below is straightforward.

  syntax Exp ::= lookup(Int)
  rule <k> lookup(L) => V ...</k> <store>... L |-> V:Val ...</store>

Environment recovery

We have already discussed the environment recovery auxiliary operation in the
IMP++ tutorial:

// TODO: eliminate the env wrapper, like we did in IMP++

  syntax KItem ::= setEnv(Map)
  rule <k> setEnv(Env) => .K ...</k> <env> _ => Env </env>

While theoretically sufficient, the basic definition for environment
recovery alone is suboptimal. Consider a loop while (E)S,
whose semantics (see above) was given by unrolling. S
is a block. Then the semantics of blocks above, together with the
unrolling semantics of the while loop, will yield a computation
structure in the k cell that increasingly grows, adding a new
environment recovery task right in front of the already existing sequence of
similar environment recovery tasks (this phenomenon is similar to the ``tail
recursion'' problem). Of course, when we have a sequence of environment
recovery tasks, we only need to keep the last one. The elegant rule below
does precisely that, thus avoiding the unnecessary computation explosion
problem:

  rule (setEnv(_) => .K) ~> setEnv(_)

In fact, the above follows a common convention in K for recovery
operations of cell contents: the meaning of a computation task of the form
cell(C) that reaches the top of the computation is that the current
contents of cell cell is discarded and gets replaced with C. We
did not add support for these special computation tasks in our current
implementation of K, so we need to define them as above.

lvalue and loc

For convenience in giving the semantics of constructs like the increment and
the assignment, that we want to operate the same way on variables and on
array elements, we used an auxiliary lvalue(E) construct which was
expected to evaluate to the lvalue of the expression E. This is only
defined when E has an lvalue, that is, when E is either a variable or
evaluates to an array element. lvalue(E) evaluates to a value of
the form loc(L), where L is the location where the value of E
can be found; for clarity, we use loc to structurally distinguish
natural numbers from location values. In giving semantics to lvalue
there are two cases to consider. (1) If E is a variable, then all we need
to do is to grab its location from the environment. (2) If E is an array
element, then we first evaluate the array and its index in order to identify
the exact location of the element of concern, and then return that location;
the last rule below works because its preceding context declarations ensure
that the array and its index are evaluated, and then the rule for array lookup
(defined above) rewrites the evaluated array access construct to its
corresponding store lookup operation.

// For parsing reasons, we prefer to allow lvalue to take a K

  syntax Exp ::= lvalue(K)
  syntax Val ::= loc(Int)

// Local variable

  rule <k> lvalue(X:Id => loc(L)) ...</k> <env>... X |-> L:Int ...</env>

// Array element: evaluate the array and its index;
// then the array lookup rule above applies.

  context lvalue(_::Exp[HOLE::Exps])
  context lvalue(HOLE::Exp[_::Exps])

// Finally, return the address of the desired object member

  rule lvalue(lookup(L:Int) => loc(L))

Initializing multiple locations

The following operation initializes a sequence of locations with the same
value:

  syntax Map ::= Int "..." Int "|->" K [function]
  rule N...M |-> _ => .Map  requires N >Int M
  rule N...M |-> K => N |-> K (N +Int 1)...M |-> K  requires N <=Int M

The semantics of SIMPLE is now complete. Make sure you kompile the
definition with the right options in order to generate the desired model.
No kompile options are needed if you only only want to execute the definition
(and thus get an interpreter), but if you want to search for a different
program behaviors then you need to kompile with the --enable-search option

endmodule

Go to Lesson 2, SIMPLE typed static

SIMPLE — Typed — Static

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K definition of the static semantics of the typed SIMPLE
language, or in other words, a type system for the typed SIMPLE
language in K. We do not re-discuss the various features of the
SIMPLE language here. The reader is referred to the untyped version of
the language for such discussions. We here only focus on the new and
interesting problems raised by the addition of type declarations, and
what it takes to devise a type system/checker for the language.

When designing a type system for a language, no matter in what
paradigm, we have to decide upon the intended typing policy. Note
that we can have multiple type systems for the same language, one for
each typing policy. For example, should we accept programs which
don't have a main function? Or should we allow functions that do not
return explicitly? Or should we allow functions whose type expects
them to return a value (say an int) to use a plain
return; statement, which returns no value, like in C?
And so on and so forth. Typically, there are two opposite tensions
when designing a type system. On the one hand, you want your type
system to be as permissive as possible, that is, to accept as many
programs that do not get stuck when executed with the untyped
semantics as possible; this will keep the programmers using your
language happy. On the other hand, you want your type system to have
a reasonable performance when implemented; this will keep both the
programmers and the implementers of your language happy. For example,
a type system for rejecting programs that could perform
division-by-zero is not expected to be feasible in general. A simple
guideline when designing typing policies is to imagine how the
semantics of the untyped language may get stuck and try to prevent
those situations from happening.

Before we give the K type system of SIMPLE formally, we discuss,
informally, the intended typing policy:

  • Each program should contain a main() function. Indeed,
    the untyped SIMPLE semantics will get stuck on any program which does
    not have a main function.

  • Each primitive value has its own type, which can be int
    bool, or string. There is also a type void
    for nonexistent values, for example for the result of a function meant
    to return no value (but only be used for its side effects, like a
    procedure).

  • The syntax of untyped SIMPLE is extended to allow type
    declarations for all the variables, including array variables. This is
    done in a C/Java-style. For example, int x; or
    int x=7, y=x+3;, or int[][][] a[10,20];
    (the latter defines a 10 × 20 matrix of arrays of integers).
    Recall from untyped SIMPLE that, unlike in C/Java, our multi-dimensional
    arrays use comma-separated arguments, although they have the array-of-array
    semantics.

  • Functions are also typed in a C/Java style. However, since in SIMPLE
    we allow functions to be passed to and returned by other functions, we also
    need function types. We will use the conventional higher-order arrow-notation
    for function types, but will separate the argument types with commas. For
    example, a function returning an array of bool elements and
    taking as argument an array x of two-integer-argument functions
    returning an integer, is declared using a syntax of the form
    bool[] f(((int,int)->int)[] x) { ... }
    and has the type ((int,int)->int)[] -> bool[].

  • We allow any variable declarations at the top level. Functions
    can only be declared at the top level. Each function can only access the
    other functions and variables declared at the top level, or its own locally
    declared variables. SIMPLE has static scoping.

  • The various expression and statement constructs take only elements of
    the expected types.

  • Increment and assignment can operate both on variables and on array
    elements. For example, if f has type int->int[][] and
    function g has the type int->int, then the
    increment expression ++f(7)[g(2),g(3)] is valid.

  • Functions should only return values of their declared result
    type. To give the programmers more flexibility, we allow functions to
    use return; statements to terminate without returning an
    actual value, or to not explicitly use any return statement,
    regardless of their declared return type. This flexibility can be
    handy when writing programs using certain functions only for their
    side effects. Nevertheless, as the dynamic semantics shows, a return
    value is automatically generated when an explicit return
    statement is not encountered.

  • For simplicity, we here limit exceptions to only throw and catch
    integer values. We let it as an exercise to the reader to extend the
    semantics to allow throwing and catching arbitrary-type exceptions.
    Like in programming languages like Java, one can go even further and
    define a semantics where thrown exceptions are propagated through
    try-catch statements until one of the corresponding type is found.
    We will do this when we define the KOOL language, not here.
    To keep the definition if SIMPLE simple, here we do not attempt to
    reject programs which throw uncaught exceptions.

Like in untyped SIMPLE, some constructs can be desugared into a
smaller set of basic constructs. In general, it should be clear why a
program does not type by looking at the top of the k cells in
its stuck configuration.

module SIMPLE-TYPED-STATIC-SYNTAX
  imports DOMAINS-SYNTAX

Syntax

The syntax of typed SIMPLE extends that of untyped SIMPLE with support
for declaring types to variables and functions.

  syntax Id ::= "main" [token]

Types

Primitive, array and function types, as well as lists (or tuples) of types.
The lists of types are useful for function arguments.

  syntax Type ::= "void" | "int" | "bool" | "string"
                | Type "[" "]"
                | "(" Type ")"             [bracket]
                > Types "->" Type

  syntax Types ::= List{Type,","}          [overload(exps)]

Declarations

Variable and function declarations have the expected syntax. For variables,
we basically just replaced the var keyword of untyped SIMPLE with a
type. For functions, besides replacing the function keyword with a
type, we also introduce a new syntactic category for typed variables,
Param, and lists over it.

  syntax Param ::= Type Id
  syntax Params ::= List{Param,","}

  syntax Stmt ::= Type Exps ";"
                | Type Id "(" Params ")" Block

Expressions

The syntax of expressions is identical to that in untyped SIMPLE,
except for the logical conjunction and disjunction which have
different strictness attributes, because they now have different
evaluation strategies.

  syntax Exp ::= Int | Bool | String | Id
               | "(" Exp ")"             [bracket]
               | "++" Exp
               > Exp "[" Exps "]"        [strict]
               > Exp "(" Exps ")"        [strict]
               | "-" Exp                 [strict]
               | "sizeOf" "(" Exp ")"    [strict]
               | "read" "(" ")"
               > left:
                 Exp "*" Exp             [strict, left]
               | Exp "/" Exp             [strict, left]
               | Exp "%" Exp             [strict, left]
               > left:
                 Exp "+" Exp             [strict, left]
               | Exp "-" Exp             [strict, left]
               > non-assoc:
                 Exp "<" Exp             [strict, non-assoc]
               | Exp "<=" Exp            [strict, non-assoc]
               | Exp ">" Exp             [strict, non-assoc]
               | Exp ">=" Exp            [strict, non-assoc]
               | Exp "==" Exp            [strict, non-assoc]
               | Exp "!=" Exp            [strict, non-assoc]
               > "!" Exp                 [strict]
               > left:
                 Exp "&&" Exp            [strict, left]
               | Exp "||" Exp            [strict, left]
               > "spawn" Block
               > Exp "=" Exp             [strict(2), right]

Note that spawn has not been declared strict. This may
seem unexpected, because the child thread shares the same environment
with the parent thread, so from a typing perspective the spawned
statement makes the same sense in a child thread as it makes in the
parent thread. The reason for not declaring it strict is because we
want to disallow programs where the spawned thread calls the
return statement, because those programs would get stuck in
the dynamic semantics. The type semantics of spawn below will reject
such programs.

We still need lists of expressions, defined below, but note that we do
not need lists of identifiers anymore. They have been replaced by the lists
of parameters.

  syntax Exps ::= List{Exp,","}          [strict, overload(exps)]

Statements

The statements have the same syntax as in untyped SIMPLE, except for
the exceptions, which now type their parameter. Note that, unlike in untyped
SIMPLE, all statement constructs which have arguments and are not desugared
are strict, including the conditional and the while. Indeed, from a
typing perspective, they are all strict: first type their arguments and then
type the actual construct.

  syntax Block ::= "{" "}"
                | "{" Stmt "}"

  syntax Stmt ::= Block
                | Exp ";"                                  [strict]
                | "if" "(" Exp ")" Block "else" Block      [avoid, strict]
                | "if" "(" Exp ")" Block                   [macro]
                | "while" "(" Exp ")" Block                [strict]
                | "for" "(" Stmt Exp ";" Exp ")" Block     [macro]
                | "return" Exp ";"                         [strict]
                | "return" ";"
                | "print" "(" Exps ")" ";"                 [strict]
                | "try" Block "catch" "(" Param ")" Block  [strict(1)]
                | "throw" Exp ";"                          [strict]
                | "join" Exp ";"                           [strict]
                | "acquire" Exp ";"                        [strict]
                | "release" Exp ";"                        [strict]
                | "rendezvous" Exp ";"                     [strict]

Note that the sequential composition is now sequentially strict,
because, unlike in the dynamic semantics where statements dissolved,
they now reduce to the stmt type, which is a result.

  syntax Stmt ::= Stmt Stmt                             [seqstrict, right]

Desugaring macros

We use the same desugaring macros like in untyped SIMPLE, but, of
course, including the types of the involved variables.

  rule if (E) S => if (E) S else {}
  rule for(Start Cond; Step) {S:Stmt} => {Start while(Cond){S Step;}}
  rule for(Start Cond; Step) {} => {Start while(Cond){Step;}}
  rule T:Type E1:Exp, E2:Exp, Es:Exps; => T E1; T E2, Es;               [anywhere]
  rule T:Type X:Id = E; => T X; X = E;                                  [anywhere]

endmodule


module SIMPLE-TYPED-STATIC
  imports SIMPLE-TYPED-STATIC-SYNTAX
  imports DOMAINS

Static semantics

Here we define the type system of SIMPLE. Like concrete semantics,
type systems defined in K are also executable. However, K type
systems turn into type checkers instead of interpreters when executed.

The typing process is done in two (overlapping) phases. In the first
phase the global environment is built, which contains type bindings
for all the globally declared variables and functions. For functions,
the declared types will be ``trusted'' during the first phase and
simply bound to their corresponding function names and placed in the
global type environment. At the same time, type-checking tasks that
the function bodies indeed respect their claimed types are generated.
All these tasks are (concurrently) verified during the second phase.
This way, all the global variable and function declarations are
available in the global type environment and can be used in order to
type-check each function code. This is consistent with the semantics
of untyped SIMPLE, where functions can access all the global variables
and can call any other function declared in the same program. The
two phases may overlap because of the K concurrent semantics. For
example, a function task can be started while the first phase is still
running; moreover, it may even complete before the first phase does,
namely when all the global variables and functions that it needs have
already been processed and made available in the global environment by
the first phase task.

Extended syntax and results

The idea is to start with a configuration holding the program to type
in one of its cells, then apply rewrite rules on it mixing types and
language syntax, and eventually obtain a type instead of the original
program. In other words, the program reduces to its type using
the K rules giving the type system of the language. In doing so,
additional typing tasks for function bodies are generated and solved
the same way. If this rewriting process gets stuck, then we say that
the program is not well-typed. Otherwise the program is well-typed
(by definition). We did not need types for statements and for blocks
as part of the typed SIMPLE syntax, because programmers are not allowed
to use such types explicitly. However, we are going to need them in the
type system, because blocks and statements reduce to them.

We start by allowing types to be used inside expressions and statements in
our language. This way, types can be used together with language syntax in
subsequent K rules without any parsing errors. Like in the type system of
IMP++ in the K tutorial, we prefer to group the block and statement types
under one syntactic sub-category of types, because this allows us to more
compactly state that certain terms can be either blocks or statements. Also,
since programs and fragments of program will reduce to their types, in order
for the strictness and context declarations to be executable we state that
types are results (same like we did in the IMP++ tutorial).

  syntax Exp ::= Type
  syntax Exps ::= Types
  syntax BlockOrStmtType ::= "block" | "stmt"
  syntax Type ::= BlockOrStmtType
  syntax Block ::= BlockOrStmtType
  syntax KResult ::= Type
                   | Types    //TODO: remove this, eventually

Configuration

The configuration of our type system consists of a tasks cell
holding various typing task cells, and a global type environment.
Each task includes a k cell holding the code to type, a tenv
cell holding the local type environment, and a return cell holding
the return type of the currently checked function. The latter is needed in
order to check whether return statements return values of the expected type.
Initially, the program is placed in a k cell inside a
task cell. Since the cells with multiplicity ? are not
included in the initial configuration, the task cell holding
the original program in its k cell will contain no other
subcells.

  configuration <T color="yellow">
                  <tasks color="orange">
                    <task multiplicity="*" color="yellow" type="Set">
                      <k color="green"> $PGM:Stmt </k>
                      <tenv multiplicity="?" color="cyan"> .Map </tenv>
                      <returnType multiplicity="?" color="black"> void </returnType>
                    </task>
                  </tasks>
//                  <br/>
                  <gtenv color="blue"> .Map </gtenv>
                </T>

Variable declarations

Variable declarations type as statements, that is, they reduce to the
type stmt. There are only two cases that need to be
considered: when a simple variable is declared and when an array
variable is declared. The macros at the end of the syntax module
above take care of reducing other variable declarations, including
ones where the declared variables are initialized, to only these two
cases. The first case has two subcases: when the variable declaration
is global (i.e., the task cell contains only the k
cell), in which case it is added to the global type environment
checking at the same time that the variable has not been already
declared; and when the variable declaration is local (i.e., a
tenv cell is available), in which case it is simply added to
the local type environment, possibly shadowing previous homonymous
variables. The third case reduces to the second, incrementally moving
the array dimension into the type until the array becomes a simple
variable.

  rule <task> <k> T:Type X:Id; => stmt ...</k> </task>
       <gtenv> Rho (.Map => X |-> T) </gtenv>
    requires notBool(X in keys(Rho))
  rule <k> T:Type X:Id; => stmt ...</k> <tenv> Rho => Rho[X <- T] </tenv>

  context _:Type _::Exp[HOLE::Exps];
// The rule below may need to sort E to Exp in the future, if the
// parser gets stricter; without that information, it may not be able
// to complete the LHS into T E[int,Ts],.Exps; (and similarly for the RHS)
  rule T:Type E:Exp[int,Ts:Types]; => T[] E[Ts];
// I want to write the rule below as _:Type (E:Exp[.Types] => E),
// but the list completion seems to not work well with that.
  rule T:Type E:Exp[.Types]; => T E;

Function declarations

Functions are allowed to be declared only at the top level (the
task cell holds only its k subcell). Each function
declaration reduces to a variable declaration (a binding of its name
to its declared function type), but also adds a task into the
tasks cell. The task consists of a typing of the statement
declaring all the function parameters followed by the function body,
together with the expected return type of the function. The
getTypes and mkDecls functions, defined at the end of
the file in the section on auxiliary operations, extracts the list of
types and makes a sequence of variable declarations from a list of
function parameters, respectively. Note that, although in the dynamic
semantics we include a terminating return statement at the
end of the function body to eliminate from the analysis the case when
the function does not provide an explicit return, we do not need to
include such a similar return statement here. That's because
the return statements type to stmt anyway, and the
entire code of the function body needs to type anyway.

  rule <task> <k> T:Type F:Id(Ps:Params) S => getTypes(Ps)->T F; ...</k> </task>
       (.Bag => <task>
               <k> mkDecls(Ps) S </k> <tenv> .Map </tenv> <returnType> T </returnType>
             </task>)

Checking if main() exists}

Once the entire program is processed (generating appropriate tasks
to type check its function bodies), we can dissolve the main
task cell (the one holding only a k subcell). Since
we want to enforce that programs include a main function, we also
generate a function task executing main() to ensure that it
types (remove this task creation if you do not want your type system
to reject programs without a main function).

  rule <task> <k> stmt => main(.Exps); </k> (.Bag => <tenv> .Map </tenv>) </task>

Collecting the terminated tasks

Similarly, once a non-main task (i.e., one which contains a
tenv subcells) is completed using the subsequent rules (i.e.,
its k cell holds only the block or stmt
type), we can dissolve its corresponding cell. Note that it is
important to ensure that we only dissolve tasks containing a
tenv cell with the rule below, because the main task should
not dissolve this way! It should do what the above rule says.
In the end, there should be no task cell left in the configuration
when the program correctly type checks.

  rule <task>... <k> _:BlockOrStmtType </k> <tenv> _ </tenv> ...</task> => .Bag

Basic values

The first three rewrite rules below reduce the primitive values to
their types, as we typically do when we define type systems in K.

  rule _:Int => int
  rule _:Bool => bool
  rule _:String => string

Variable lookup

There are three cases to distinguish for variable lookup: (1) if the
variable is bound in the local type environment, then look its type up
there; (2) if a local environment exists and the variable is not bound
in it, then look its type up in the global environment; (3) finally,
if there is no local environment, meaning that we are executing the
top-level pass, then look the variable's type up in the global
environment, too.

  rule <k> X:Id => T ...</k> <tenv>... X |-> T ...</tenv>

  rule <k> X:Id => T ...</k> <tenv> Rho </tenv> <gtenv>... X |-> T ...</gtenv>
    requires notBool(X in keys(Rho))

  rule <task> <k> X:Id => T ...</k> </task> <gtenv>... X |-> T ...</gtenv>

Increment

We want the increment operation to apply to any lvalue, including
array elements, not only to variables. For that reason, we define a
special context extracting the type of the argument of the increment
operation only if that argument is an lvalue. Otherwise the rewriting
process gets stuck. The operation ltype is defined at the
end of this file, in the auxiliary operation section. It essentially
acts as a filter, getting stuck if its argument is not an lvalue and
letting it reduce otherwise. The type of the lvalue is expected to be
an integer in order to be allowed to be incremented, as seen in the
rule ++ int => int below.

  context ++(HOLE => ltype(HOLE))
  rule ++ int => int

Common expression constructs

The rules below are straightforward and self-explanatory:

  rule int + int => int
  rule string + string => string
  rule int - int => int
  rule int * int => int
  rule int / int => int
  rule int % int => int
  rule - int => int
  rule int < int => bool
  rule int <= int => bool
  rule int > int => bool
  rule int >= int => bool
  rule T:Type == T => bool
  rule T:Type != T => bool
  rule bool && bool => bool
  rule bool || bool => bool
  rule ! bool => bool

Array access and size

Array access requires each index to type to an integer, and the
array type to be at least as deep as the number of indexes:

// NOTE:
// We used to need parentheses in the RHS, to avoid capturing Ts as an attribute
// Let's hope that is not a problem anymore.

  rule (T[])[int, Ts:Types] => T[Ts]
  rule T:Type[.Types] => T

sizeOf only needs to check that its argument is an array:

  rule sizeOf(_T[]) => int

Input/Output

The read expression construct types to an integer, while print types
to a statement provided that all its arguments type to integers or
strings.

  rule read() => int

  rule print(T:Type, Ts => Ts); requires T ==K int orBool T ==K string
  rule print(.Types); => stmt

Assignment

The special context and the rule for assignment below are similar
to those for increment: the LHS of the assignment must be an lvalue
and, in that case, it must have the same type as the RHS, which then
becomes the type of the assignment.

  context (HOLE => ltype(HOLE)) = _
  rule T:Type = T => T

Function application and return

Function application requires the type of the function and the
types of the passed values to be compatible. Note that a special case
is needed to handle the no-argument case:

  rule (Ts:Types -> T)(Ts) => T requires Ts =/=K .Types
  rule (void -> T)(.Types) => T

The returned value must have the same type as the declared
function return type. If an empty return is encountered, than
we should check that we are in a function (and not a thread)
context, that is, a return cell must be available:

  rule <k> return T:Type; => stmt ...</k> <returnType> T </returnType>
  rule <k> return; => stmt ...</k> <returnType> _ </returnType>

Blocks

To avoid having to recover type environments after blocks, we prefer
to start a new task for block body, making sure that the new task
is passed the same type environment and return cells. The value
returned by return statements must have the same type as
stated in the return cell. The print variadic
function is allowed to only print integers and strings. The thrown
exceptions can only have integer type.

  rule {} => block

  rule <task> <k> {S} => block ...</k> <tenv> Rho </tenv> R </task>
       (.Bag => <task> <k> S </k> <tenv> Rho </tenv> R </task>)

Expression statement

  rule _:Type; => stmt

Conditional and while loop

  rule if (bool) block else block => stmt
  rule while (bool) block => stmt

Exceptions

We currently force the parameters of exceptions to only be integers.
Moreover, for simplicity, we assume that integer exceptions can be
thrown from anywhere, including from functions which do not define
any try-catch block (with the currently unchecked ‒also for
simplicity‒ expectation that the caller functions would catch those
exceptions).

  rule try block catch(int X:Id) {S} => {int X; S}
  rule try block catch(int X:Id) {} => {int X;}
  rule throw int; => stmt

Concurrency

Nothing special about typing the concurrency constructs, except that
we do not want the spawned thread to return, so we do not include any
return cell in the new task cell for the thread statement.
Same like with the functions above, we do not check for thrown
exceptions which are not caught.

  rule <k> spawn S => int ...</k> <tenv> Rho </tenv>
       (.Bag => <task> <k> S </k> <tenv> Rho </tenv> </task>)
  rule join int; => stmt
  rule acquire _:Type; => stmt
  rule release _:Type; => stmt
  rule rendezvous _:Type; => stmt

  rule _:BlockOrStmtType _:BlockOrStmtType => stmt

Auxiliary constructs

The function mkDecls turns a list of parameters into a
list of variable declarations.

  syntax Stmt ::= mkDecls(Params)  [function]
  rule mkDecls(T:Type X:Id, Ps:Params) => T X; mkDecls(Ps)
  rule mkDecls(.Params) => {}

The ltype context allows only expressions which have an
lvalue to evaluate.

  syntax LValue ::= Id
  rule isLValue(_:Exp[_:Exps]) => true
  syntax Exp ::= LValue  // K should be able to infer this
                         // if not added, then it gets stuck with an Id on k cell

// Instead of the second LValue production above you can use a rule:
//  rule isLValue(_:Exp[_:Exps]) => true

  syntax Exp ::= ltype(Exp)
//  context ltype(HOLE:LValue)
// The above context does not work due to some error, so we write instead
  context ltype(HOLE) requires isLValue(HOLE)

The function getTypes is the same as in SIMPLE typed dynamic.

  syntax Types ::= getTypes(Params)  [function]
  rule getTypes(T:Type _:Id) => T, .Types   // I would like to not use .Types
  rule getTypes(T:Type _:Id, P, Ps) => T, getTypes(P,Ps)
  rule getTypes(.Params) => void, .Types

endmodule

Go to Lesson 3, SIMPLE typed dynamic

SIMPLE — Typed — Dynamic

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K dynamic semantics of the typed SIMPLE language.
It is very similar to the semantics of the untyped SIMPLE, the
difference being that we now dynamically check the typing policy
described in the static semantics of typed SIMPLE. Because of the
dynamic nature of the semantics, we can also perform some additional
checks which were not possible in the static semantics, such as
memory leaks due to accessing an array out of its bounds. We will
highlight the differences between the dynamically typed and the
untyped SIMPLE as we proceed with the semantics. We recommend the
reader to consult the typing policy and the syntax of types discussed
in the static semantics of the typed SIMPLE language.

module SIMPLE-TYPED-DYNAMIC-SYNTAX
  imports DOMAINS-SYNTAX

Syntax

The syntax of typed SIMPLE extends that of untyped SIMPLE with support
for declaring types to variables and functions.

The syntax below is identical to that of the static semantics of typed
SIMPLE. However, the K strictness attributes are like those of the untyped
SIMPLE, to capture the desired evaluation strategies of the various language
constructs.

  syntax Id ::= "main" [token]

Types

  syntax Type ::= "void" | "int" | "bool" | "string"
                | Type "[" "]"
                | "(" Type ")"           [bracket]
                > Types "->" Type
  syntax Types ::= List{Type,","}        [overload(exps)]

Declarations

  syntax Param ::= Type Id
  syntax Params ::= List{Param,","}

  syntax Stmt ::= Type Exps ";"
                | Type Id "(" Params ")" Block

Expressions

  syntax Exp ::= Int | Bool | String | Id
               | "(" Exp ")"             [bracket]
               | "++" Exp
               > Exp "[" Exps "]"        [strict]
               > Exp "(" Exps ")"        [strict]
               | "-" Exp                 [strict]
               | "sizeOf" "(" Exp ")"    [strict]
               | "read" "(" ")"
               > left:
                 Exp "*" Exp             [strict, left]
               | Exp "/" Exp             [strict, left]
               | Exp "%" Exp             [strict, left]
               > left:
                 Exp "+" Exp             [strict, left]
               | Exp "-" Exp             [strict, left]
               > non-assoc:
                 Exp "<" Exp             [strict, non-assoc]
               | Exp "<=" Exp            [strict, non-assoc]
               | Exp ">" Exp             [strict, non-assoc]
               | Exp ">=" Exp            [strict, non-assoc]
               | Exp "==" Exp            [strict, non-assoc]
               | Exp "!=" Exp            [strict, non-assoc]
               > "!" Exp                 [strict]
               > left:
                 Exp "&&" Exp            [strict(1), left]
               | Exp "||" Exp            [strict(1), left]
               > "spawn" Block
               > Exp "=" Exp             [strict(2), right]

Like in the static semantics, there is no need for lists of identifiers
(because we now have lists of parameters).

  syntax Exps ::= List{Exp,","}          [strict, overload(exps)]
  syntax Val
  syntax Vals ::= List{Val,","}          [overload(exps)]

Statements

  syntax Block ::= "{" "}"
                | "{" Stmt "}"

  syntax Stmt ::= Block
                | Exp ";"                               [strict]
                | "if" "(" Exp ")" Block "else" Block   [avoid, strict(1)]
                | "if" "(" Exp ")" Block                [macro]
                | "while" "(" Exp ")" Block
                | "for" "(" Stmt Exp ";" Exp ")" Block  [macro]
                | "print" "(" Exps ")" ";"              [strict]
                | "return" Exp ";"                      [strict]
                | "return" ";"
                | "try" Block "catch" "(" Param ")" Block
            | "throw" Exp ";"                       [strict]
                | "join" Exp ";"                        [strict]
                | "acquire" Exp ";"                     [strict]
                | "release" Exp ";"                     [strict]
                | "rendezvous" Exp ";"                  [strict]

  syntax Stmt ::= Stmt Stmt                          [right]

The same desugaring macros like in the statically typed SIMPLE.

  rule if (E) S => if (E) S else {}
  rule for(Start Cond; Step) {S:Stmt} => {Start while(Cond){S Step;}}
  rule for(Start Cond; Step) {} => {Start while(Cond){Step;}}
  rule T:Type E1:Exp, E2:Exp, Es:Exps; => T E1; T E2, Es;               [anywhere]
  rule T:Type X:Id = E; => T X; X = E;                                  [anywhere]

endmodule


module SIMPLE-TYPED-DYNAMIC
  imports SIMPLE-TYPED-DYNAMIC-SYNTAX
  imports DOMAINS

Semantics

Values and results

These are similar to those of untyped SIMPLE, except that the array
references and the function abstrations now also hold their types.
These types are needed in order to easily compute the type of any
value in the language (see the auxiliary typeOf operation at
the end of this module).

  syntax Val ::= Int | Bool | String
               | array(Type,Int,Int)
               | lambda(Type,Params,Stmt)
  syntax Exp ::= Val
  syntax Exps ::= Vals
  syntax KResult ::= Val
                   | Vals  // TODO: should not need this

Configuration

The configuration is almost identical to that of untyped SIMPLE,
except for a return cell inside the control cell.
This return cell will hold, like in the static semantics of
typed SIMPLE, the expected type of the value returned by the function
being executed. The contents of this cell will be set whenever a
function is invoked and will be checked whenever the evaluation of the
function body encounters an explicit return statement.

  // the syntax declarations below are required because the sorts are
  // referenced directly by a production and, because of the way KIL to KORE
  // is implemented, the configuration syntax is not available yet
  // should simply work once KIL is removed completely
  // check other definitions for this hack as well

  syntax ControlCell
  syntax ControlCellFragment

  configuration <T color="red">
                  <threads color="orange">
                    <thread multiplicity="*" color="yellow" type="Map">
                      <id color="pink"> 0 </id>
                      <k color="green"> ($PGM:Stmt ~> execute) </k>
//                      <br/>
                      <control color="cyan">
                        <fstack color="blue"> .List </fstack>
                        <xstack color="purple"> .List </xstack>
                        <returnType color="LimeGreen"> void </returnType>
                       </control>
//                      <br/>
                      <env color="violet"> .Map </env>
                      <holds color="black"> .Map </holds>
                    </thread>
                  </threads>
//                  <br/>
                  <genv color="pink"> .Map </genv>
                  <store color="white"> .Map </store>
                  <busy color="cyan">.Set</busy>
                  <terminated color="red"> .Set </terminated>
                  <input color="magenta" stream="stdin"> .List </input>
                  <output color="brown" stream="stdout"> .List </output>
                  <nextLoc color="gray"> 0 </nextLoc>
                </T>

Declarations and Initialization

Variable Declaration

The undefined construct is now parameterized by a type.
A main difference between untyped SIMPLE and dynamically typed SIMPLE
is that the latter assigns a type to each of its locations and that
type cannot be changed during the execution of the program. We do not
do any memory management in our semantic definitions here, so
locations cannot be reclaimed, garbage collected and/or reused. Each
location corresponds precisely to an allocated variable or array
element, whose type was explicitly or implicitly declared in the
program and does not change. It is therefore safe to type each
location and then never allow that type to change. The typed
undefined values effectively assign both a type and an undefined value
to a location.

  syntax KItem ::= undefined(Type)

  rule <k> T:Type X:Id; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> undefined(T) ...</store>
       <nextLoc> L:Int => L +Int 1 </nextLoc>

Array Declaration

The dynamic semantics of typed array declarations is similar to that
in untyped SIMPLE, but we have to make sure that we associate the
right type to the allocated locations.

  rule <k> T:Type X:Id[N:Int]; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> array(T, L +Int 1, N)
                          (L +Int 1)...(L +Int N) |-> undefined(T) ...</store>
       <nextLoc> L:Int => L +Int 1 +Int N </nextLoc>
    requires N >=Int 0

  context _:Type _::Exp[HOLE::Exps];

The desugaring of multi-dimensional arrays into unidimensional
ones is also similar to that in untyped SIMPLE, although we have to
make sure that all the declared variables have the right types. The
auxiliary operation T<Vs>, defined at the end of the file,
adds the length of Vs dimensions to the type T.

// TODO: Check the desugaring below to be consistent with the one for untyped simple

  syntax Id ::= "$1" [token] | "$2" [token]
  rule T:Type X:Id[N1:Int, N2:Int, Vs:Vals];
    => T[]<Vs> X[N1];
       {
         T[][]<Vs> $1=X;
         for(int $2=0; $2 <= N1 - 1; ++$2) {
           T X[N2,Vs];
           $1[$2] = X;
         }
       }

Function declaration

Store all function parameters, as well as the return type, as part
of the lambda abstraction. In the spirit of dynamic typing, we will
make sure that parameters are well typed when the function is invoked.

  rule <k> T:Type F:Id(Ps:Params) S => .K ...</k>
       <env> Env => Env[F <- L] </env>
       <store>... .Map => L |-> lambda(T, Ps, S) ...</store>
       <nextLoc> L => L +Int 1 </nextLoc>

Calling main()

When done with the first pass, call main().

  syntax KItem ::= "execute"
  rule <k> execute => main(.Exps); </k>
       <env> Env </env>
       <genv> .Map => Env </genv>

Expressions

Variable lookup

  rule <k> X:Id => V ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V:Val ...</store>

Variable/Array increment

  context ++(HOLE => lvalue(HOLE))
  rule <k> ++loc(L) => I +Int 1 ...</k>
       <store>... L |-> (I:Int => I +Int 1) ...</store>

Arithmetic operators

  rule I1 + I2 => I1 +Int I2
  rule Str1 + Str2 => Str1 +String Str2
  rule I1 - I2 => I1 -Int I2
  rule I1 * I2 => I1 *Int I2
  rule I1 / I2 => I1 /Int I2 requires I2 =/=K 0
  rule I1 % I2 => I1 %Int I2 requires I2 =/=K 0
  rule - I => 0 -Int I
  rule I1 < I2 => I1 <Int I2
  rule I1 <= I2 => I1 <=Int I2
  rule I1 > I2 => I1 >Int I2
  rule I1 >= I2 => I1 >=Int I2
  rule V1:Val == V2:Val => V1 ==K V2
  rule V1:Val != V2:Val => V1 =/=K V2
  rule ! T => notBool(T)
  rule true  && E => E
  rule false && _ => false
  rule true  || _ => true
  rule false || E => E

Array lookup

Check array bounds, as part of the dynamic typing policy.

// Same comment as for simple untyped regarding [anywhere]
  rule V:Val[N1:Int, N2:Int, Vs:Vals] => V[N1][N2, Vs]
    [anywhere]

// Same comment as for simple untyped regarding [anywhere]
  rule array(_:Type, L:Int, M:Int)[N:Int] => lookup(L +Int N)
    requires N >=Int 0 andBool N <Int M  [anywhere]

Size of an array

  rule sizeOf(array(_,_,N)) => N

Function call

Define function call and return together, to see their relationship.
Note that the operation mkDecls now declares properly typed
instantiated variables, and that the semantics of return also
checks that that type of the returned value is expected one.

  syntax KItem ::= (Type,Map,K,ControlCellFragment)

  rule <k> lambda(T,Ps,S)(Vs:Vals) ~> K => mkDecls(Ps,Vs) S return; </k>
       <control>
         <fstack> .List => ListItem((T',Env,K,C)) ...</fstack>
         <returnType> T' => T </returnType>
         C
       </control>
       <env> Env => GEnv </env>
       <genv> GEnv </genv>

  rule <k> return V:Val; ~> _ => V ~> K </k>
       <control>
         <fstack> ListItem((T',Env,K,C)) => .List ...</fstack>
         <returnType> T => T' </returnType>
         (_ => C)
       </control>
       <env> _ => Env </env>
    requires typeOf(V) ==K T   // check the type of the returned value

Like the undefined above, nothing also gets
tagged with a type now. The empty return statement is
completed to return the nothing value tagged as expected.

  syntax Val ::= nothing(Type)
  rule <k> return; => return nothing(T); ...</k> <returnType> T </returnType>

Read

  rule <k> read() => I ...</k> <input> ListItem(I:Int) => .List ...</input>

Assignment

The assignment now checks that the type of the assigned location is
preserved:

  context (HOLE => lvalue(HOLE)) = _

  rule <k> loc(L) = V:Val => V ...</k> <store>... L |-> (V' => V) ...</store>
    requires typeOf(V) ==K typeOf(V')

Statements

Blocks

  rule {} => .K
  rule <k> { S } => S ~> setEnv(Env) ...</k>  <env> Env </env>

Sequential composition

  rule S1:Stmt S2:Stmt => S1 ~> S2

Expression statements

  rule _:Val; => .K

Conditional

  rule if ( true) S else _ => S
  rule if (false) _ else S => S

While loop

  rule while (E) S => if (E) {S while(E)S}

Print

We only allow printing integers and strings:

  rule <k> print(V:Val, Es => Es); ...</k> <output>... .List => ListItem(V) </output>
    requires typeOf(V) ==K int orBool typeOf(V) ==K string
  rule print(.Vals); => .K

Exceptions

Exception parameters are now typed, but note that the semantics below
works correctly only when the thrown exception has the same type as
the innermost try-catch paramete. To keep things simple, for the time
being we can assume that SIMPLE only throws and catches integer
values, in which case our semantics below works fine:

  syntax KItem ::= (Param,Stmt,K,Map,ControlCellFragment)  // Param instead of Id

  syntax KItem ::= "popx"

  rule <k> (try S1 catch(P) S2 => S1 ~> popx) ~> K </k>
       <control>
         <xstack> .List => ListItem((P, S2, K, Env, C)) ...</xstack>
         C
       </control>
       <env> Env </env>

  rule <k> popx => .K ...</k>
       <xstack> ListItem(_) => .List ...</xstack>

  rule <k> throw V:Val; ~> _ => { T X = V; S2 } ~> K </k>
       <control>
         <xstack> ListItem((T:Type X:Id, S2, K, Env, C)) => .List ...</xstack>
         (_ => C)
       </control>
       <env> _ => Env </env>

Threads

Thread creation

   rule <thread>...
          <k> spawn S => !T:Int +Int 1 ...</k>
          <env> Env </env>
        ...</thread>
        (.Bag => <thread>...
                <k> S </k>
                <env> Env </env>
                <id> !T +Int 1 </id>
              ...</thread>)

Thread termination

   rule (<thread>... <k>.K</k> <holds>H</holds> <id>T</id> ...</thread> => .Bag)
        <busy> Busy => Busy -Set keys(H) </busy>
        <terminated>... .Set => SetItem(T) ...</terminated>

Thread joining

   rule <k> join T:Int; => .K ...</k>
        <terminated>... SetItem(T) ...</terminated>

Acquire lock

   rule <k> acquire V:Val; => .K ...</k>
        <holds>... .Map => V |-> 0 ...</holds>
        <busy> Busy (.Set => SetItem(V)) </busy>
     requires (notBool(V in Busy:Set))

   rule <k> acquire V; => .K ...</k>
        <holds>... V:Val |-> (N:Int => N +Int 1) ...</holds>

Release lock

   rule <k> release V:Val; => .K ...</k>
        <holds>... V |-> (N => N:Int -Int 1) ...</holds>
      requires N >Int 0

   rule <k> release V; => .K ...</k> <holds>... V:Val |-> 0 => .Map ...</holds>
        <busy>... SetItem(V) => .Set ...</busy>

Rendezvous synchronization

   rule <k> rendezvous V:Val; => .K ...</k>
        <k> rendezvous V; => .K ...</k>

Auxiliary declarations and operations

Turns a list of parameters and a list of instance values for them
into a list of variable declarations.

  syntax Stmt ::= mkDecls(Params,Vals)  [function]
  rule mkDecls((T:Type X:Id, Ps:Params), (V:Val, Vs:Vals))
    => T X=V; mkDecls(Ps,Vs)
  rule mkDecls(.Params,.Vals) => {}

Location lookup.

  syntax Exp ::= lookup(Int)  // see NOTES.md for why Exp instead of KItem
  rule <k> lookup(L) => V ...</k> <store>... L |-> V:Val ...</store>

Environment recovery.

// TODO: same comment regarding setEnv(...) as for simple untyped

  syntax KItem ::= setEnv(Map)
  rule <k> setEnv(Env) => .K ...</k>  <env> _ => Env </env>
  rule (setEnv(_) => .K) ~> setEnv(_)

lvalue and loc

  syntax Exp ::= lvalue(K)
  syntax Val ::= loc(Int)

  rule <k> lvalue(X:Id => loc(L)) ...</k>  <env>... X |-> L:Int ...</env>

  //context lvalue(_[HOLE])
  //context lvalue(HOLE[_])
  context lvalue(_::Exp[HOLE::Exps])
  context lvalue(HOLE::Exp[_::Exps])

  rule lvalue(lookup(L:Int) => loc(L))

Adds the corresponding depth to an array type

  syntax Type ::= Type "<" Vals ">"  [function]
  rule T:Type<_,Vs:Vals> => T[]<Vs>
  rule T:Type<.Vals> => T

Sequences of locations.

  syntax Map ::= Int "..." Int "|->" K [function]
  rule N...M |-> _ => .Map  requires N >Int M
  rule N...M |-> K => N |-> K (N +Int 1)...M |-> K  requires N <=Int M

// Type of a value.
  syntax Type ::= typeOf(K)  [function]
  rule typeOf(_:Int) => int
  rule typeOf(_:Bool) => bool
  rule typeOf(_:String) => string
  rule typeOf(array(T,_,_)) => (T[])   // () needed! K parses [] as "no tags"
  rule typeOf(lambda(T,Ps,_)) => getTypes(Ps) -> T
  rule typeOf(undefined(T)) => T
  rule typeOf(nothing(T)) => T

List of types of a parameter.

  syntax Types ::= getTypes(Params)  [function]
  rule getTypes(T:Type _:Id) => T, .Types   // I would like to not use .Types
  rule getTypes(T:Type _:Id, P, Ps) => T, getTypes(P,Ps)
  rule getTypes(.Params) => void, .Types
endmodule

KOOL — Untyped

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K semantic definition of the untyped KOOL language. KOOL
is aimed at being a pedagogical and research language that captures
the essence of the object-oriented programming paradigm. Its untyped
variant discussed here is simpler than the typed one, ignoring several
intricate aspects of types in the presence of objects. A program
consists of a set of class declarations. Each class can extend at
most one other class (KOOL is single-inheritance). A class can
declare a set of fields and a set of methods, all public and called
the class' members. Specifically, KOOL includes the
following features:

  • Class declarations, where a class may or may not explicitly
    extend another class. In case a class does not explicitly extend
    another class, then it is assumed that it extends the default top-most
    and empty (i.e., no members) class called Object. Each class
    is required to declare precisely one homonymous method, called its
    constructor. Each valid program should contain one class
    named Main, whose constructor, Main(), takes no
    arguments. The execution of a program consists of creating an object
    instance of class Main and invoking the constructor
    Main() on it, that is, of executing new Main();.

  • All features of SIMPLE (see examples/simple/untyped),
    i.e., multidimensional arrays, function (here called "method")
    abstractions with call-by-value parameter passing style and static
    scoping, blocks with locals, input/output, parametric exceptions, and
    concurrency via dynamic thread creation/termination and synchronization.
    The only change in the syntax of SIMPLE when imported in KOOL is the
    function declaration keyword, function, which is changed into
    method. The exact same desugaring macros from SIMPLE are
    also included in KOOL. We can think of KOOL's classes as embedding
    SIMPLE programs (extended with OO constructs, as discussed next).

  • Object creation using the new C(e1,...,en)
    expression construct. An object instance of class C is first
    created and then the constructor C(e1,...,en) is implicitly
    called on that object. KOOL only allows (and requires) one
    constructor per class. The class constructor can be called either
    implicitly during a new object creation for the class, or explicitly.
    The superclass constructor is not implicitly invoked when a
    class constructor is invoked; if you want to invoke the superclass
    constructor from a subclass constructor then you have to do it
    explicitly.

  • An expression construct this, which evaluates to the
    current object.

  • An expression construct super, which is used (only) in
    combination with member lookup (see next) to refer to a superclass
    field or method.

  • A member lookup expression construct e.x, where e
    is an expression (either an expression expected to evaluate to an object
    or the super construct) and x is a class member name,
    that is, a field or a method name.

  • Expression constructs e instanceOf C and
    (C) e, where e is an expression expected
    to evaluate to an object and C a class name. The former
    tells whether the class of e is a subclass of C,
    that is, whether e can be used as an instance of C,
    and the latter changes the class of e to C. These
    operations always succeed: the former returns a Boolean value, while
    the latter changes the current class of e to C
    regardless of whether it is safe to do so or not. The typed version
    of KOOL will check the safety of casting by ensuring that the instance
    class of the object is a subclass of C. In untyped KOOL we
    do not want to perform this check because we want to allow the
    programmer maximum of flexibility: if one always accesses only
    available members, then the program can execute successfully despite
    the potentially unsafe cast.

There are some specific aspects of KOOL that need to be discussed.

First, KOOL is higher-order, allowing function abstractions to be
treated like any other values in the language. For example, if
m is a method of object e then e.m
evaluates to the corresponding function abstraction. The function
abstraction is in fact a closure, because in addition to the method
parameters and body it also encapsulates the object value (i.e., the
environment of the object together with its current class—see below)
that e evaluates to. This way, function abstractions can be
invoked anywhere and have the capability to change the state of their
object. For example, if m is a method of object e
which increments a field c of e when invoked, and if
getm is another method of e which simply returns
m when invoked, then the double application
(e.getm())() has the same effect as e.m(), that is,
increments the counter c of e. Note that the
higher-order nature of KOOL was not originally planned; it came as a
natural consequence of evaluating methods to closures and we decided
to keep it. If you do not like it then do not use it.

Second, since all the fields and methods are public in KOOL and since
they can be redeclared in subclasses, it is not immediately clear how
to lookup the member x when we write e.x and
e is different from super. We distinguish two cases,
depending on whether e.x occurs in a method invocation
context (i.e., e.x(...)) or in a field context. KOOL has
dynamic method dispatch, so if e.x is invoked as a method
then x will be searched for starting with the instance class of
the object value to which e evaluates. If e.x
occurs in a non-method-invocation context then x will be
treated as a field (although it may hold a method closure due to the
higher-order nature of KOOL) and thus will be searched starting with
the current class of the object value of e (which, because of
this and casting, may be different from its instance class).
In order to achieve the above, each object value will consist of a
pair holding the current class of the object and an environment stack
with one layer for each class in the object's instance class hierarchy.

Third, although KOOL is dynamic method dispatch, its capabilities
described above are powerful enough to allow us to mimic static
method dispatch. For example, suppose that you want to invoke method
m() statically. Then all you need to do is to declare a
local variable and bind it to m, for example var staticm = m;, and
then call staticm(). This works because
staticm is first bound to the method closure that m
evaluates to, and then looked up as any local variable when invoked.
We only enable the dynamic method dispatch when we have an object
member on an application position, e.g., m().

In what follows, we limit our comments to the new, KOOL-specific
aspects of the language. We refer the reader to the untyped SIMPLE
language for documentation on the the remaining features, because
those were all borrowed from SIMPLE.

module KOOL-UNTYPED-SYNTAX
  imports DOMAINS-SYNTAX

Syntax

The syntax of KOOL extends that of SIMPLE with object-oriented
constructs. We removed from the K annotated syntax of SIMPLE two
constructs, namely the one for function declarations (because we want
to call them methods now) and the one for function application
(because application is not strict in the first argument
anymore—needs to initiate dynamic method dispatch). The additional
syntax includes:

  • First, we need a new dedicated identifier, Object, for
    the default top-most class.
  • Second, we rename the function keyword of SIMPLE into method.
  • Third, we add syntax for class declarations together with a
    macro making classes which extend nothing to extend Object.
  • Fourth, we change the strictness attribute of application
    into strict(2).
  • Finally, we add syntax and corresponding strictness
    for the KOOL object-oriented constructs.
  syntax Id ::= "Object" [token] | "Main" [token]

  syntax Stmt ::= "var" Exps ";"
                | "method" Id "(" Ids ")" Block  // called "function" in SIMPLE
                | "class" Id Block               // KOOL
                | "class" Id "extends" Id Block  // KOOL

  syntax Exp ::= Int | Bool | String | Id
               | "this"                                 // KOOL
               | "super"                                // KOOL
               | "(" Exp ")"             [bracket]
               | "++" Exp
               | Exp "instanceOf" Id     [strict(1)]    // KOOL
               | "(" Id ")" Exp          [strict(2)]    // KOOL  cast
               | "new" Id "(" Exps ")"   [strict(2)]    // KOOL
               | Exp "." Id                             // KOOL
               > Exp "[" Exps "]"        [strict]
               > Exp "(" Exps ")"        [strict(2)]    // was strict in SIMPLE
               | "-" Exp                 [strict]
               | "sizeOf" "(" Exp ")"    [strict]
               | "read" "(" ")"
               > left:
                 Exp "*" Exp             [strict, left]
               | Exp "/" Exp             [strict, left]
               | Exp "%" Exp             [strict, left]
               > left:
                 Exp "+" Exp             [strict, left]
               | Exp "-" Exp             [strict, left]
               > non-assoc:
                 Exp "<" Exp             [strict, non-assoc]
               | Exp "<=" Exp            [strict, non-assoc]
               | Exp ">" Exp             [strict, non-assoc]
               | Exp ">=" Exp            [strict, non-assoc]
               | Exp "==" Exp            [strict, non-assoc]
               | Exp "!=" Exp            [strict, non-assoc]
               > "!" Exp                 [strict]
               > left:
                 Exp "&&" Exp            [strict(1), left]
               | Exp "||" Exp            [strict(1), left]
               > "spawn" Block
               > Exp "=" Exp             [strict(2), right]

  syntax Ids  ::= List{Id,","}

  syntax Exps ::= List{Exp,","}          [strict, overload(exps)]
  syntax Val
  syntax Vals ::= List{Val,","}          [overload(exps)]

  syntax Block ::= "{" "}"
                | "{" Stmt "}"

  syntax Stmt ::= Block
                | Exp ";"                               [strict]
                | "if" "(" Exp ")" Block "else" Block   [avoid, strict(1)]
                | "if" "(" Exp ")" Block                [macro]
                | "while" "(" Exp ")" Block
                | "for" "(" Stmt Exp ";" Exp ")" Block  [macro]
                | "return" Exp ";"                      [strict]
                | "return" ";"                          [macro]
                | "print" "(" Exps ")" ";"              [strict]
                | "try" Block "catch" "(" Id ")" Block
                | "throw" Exp ";"                       [strict]
                | "join" Exp ";"                        [strict]
                | "acquire" Exp ";"                     [strict]
                | "release" Exp ";"                     [strict]
                | "rendezvous" Exp ";"                  [strict]

  syntax Stmt ::= Stmt Stmt                          [right]

Old desugaring rules, from SIMPLE

  rule if (E) S => if (E) S else {}
  rule for(Start Cond; Step) {S} => {Start while (Cond) {S Step;}}
  rule var E1::Exp, E2::Exp, Es::Exps; => var E1; var E2, Es;       [anywhere]
  rule var X::Id = E; => var X; X = E;                              [anywhere]

New desugaring rule

  rule class C:Id S => class C extends Object S                     // KOOL

endmodule

Semantics

We first discuss the new configuration of KOOL, which extends that of
SIMPLE. Then we include the semantics of the constructs borrowed from
SIMPLE unchanged; we refrain from discussing those, because they were
already discussed in the K definition of SIMPLE. Then we discuss
changes to SIMPLE's semantics needed for the more general meaning of
the previous SIMPLE constructs (for example for thread spawning,
assignment, etc.). Finally, we discuss in detail the
semantics of the additional KOOL constructs.

module KOOL-UNTYPED
  imports KOOL-UNTYPED-SYNTAX
  imports DOMAINS

Configuration

KOOL removes one cell and adds two nested cells to the configuration
of SIMPLE. The cell which is removed is the one holding the global
environment, because a KOOL program consists of a set of classes only,
with no global declarations. In fact, since informally speaking each
KOOL class now includes a SIMPLE program, it is safe to say that the
global variables in SIMPLE became class fields in KOOL. Let us now
discuss the new cells that are added to the configuration of SIMPLE.

  • The cell crntObj holds data pertaining to the current
    object, that is, the object environment in which the code in cell
    k executes: crntClass holds the current class (which
    can change as methods of the current object are invoked);
    envStack holds the stack of environments as a list,
    each layer corresponding to one class in the objects' instance class
    hierarchy; location, which is optional, holds the location in
    the store where the current object is or has to be located (this is
    useful both for method closures and for the semantics of object
    creation).

  • The cell classes holds all the declared classes, each
    class being held in its own class cell which contains a name
    (className), a parent (extends), and the actual
    member declarations (declarations).

  // the syntax declarations below are required because the sorts are
  // referenced directly by a production and, because of the way KIL to KORE
  // is implemented, the configuration syntax is not available yet
  // should simply work once KIL is removed completely
  // check other definitions for this hack as well
  syntax EnvCell
  syntax ControlCell
  syntax EnvStackCell
  syntax CrntObjCellFragment

  configuration <T color="red">
                  <threads color="orange">
                    <thread multiplicity="*" type="Set" color="yellow">
                      <k color="green"> $PGM:Stmt ~> execute </k>
                    //<br/> // TODO(KORE): support latex annotations #1799
                      <control color="cyan">
                        <fstack color="blue"> .List </fstack>
                        <xstack color="purple"> .List </xstack>
                      //<br/> // TODO(KORE): support latex annotations #1799
                        <crntObj color="Fuchsia">  // KOOL
                           <crntClass> Object </crntClass>
                           <envStack> .List </envStack>
                           <location multiplicity="?"> .K </location>
                        </crntObj>
                      </control>
                    //<br/> // TODO(KORE): support latex annotations #1799
                      <env color="violet"> .Map </env>
                      <holds color="black"> .Map </holds>
                      <id color="pink"> 0 </id>
                    </thread>
                  </threads>
                //<br/> // TODO(KORE): support latex annotations #1799
                  <store color="white"> .Map </store>
                  <busy color="cyan">.Set </busy>
                  <terminated color="red"> .Set </terminated>
                  <input color="magenta" stream="stdin"> .List </input>
                  <output color="brown" stream="stdout"> .List </output>
                  <nextLoc color="gray"> 0 </nextLoc>
                //<br/> // TODO(KORE): support latex annotations #1799
                  <classes color="Fuchsia">        // KOOL
                     <classData multiplicity="*" type="Map" color="Fuchsia">
                        // the Map has as its key the first child of the cell,
                        // in this case the className cell.
                        <className color="Fuchsia"> Main </className>
                        <baseClass color="Fuchsia"> Object </baseClass>
                        <declarations color="Fuchsia"> .K </declarations>
                     </classData>
                  </classes>
                </T>

Unchanged Semantics from untyped SIMPLE

The semantics below is taken over from SIMPLE unchanged.
The semantics of function declaration and invocation, including the
use of the special lambda abstraction value, needs to change
in order to account for the fact that methods are now invoked into
their object's environment. The semantics of function return actually
stays unchanged. Also, the semantics of program initialization is
different: now we have to create an instance of the Main
class which also calls the constructor Main(), while in
SIMPLE we only had to invoke the function Main().
Finally, the semantics of thread spawning needs to change, too: the
parent thread needs to also share its object environment with the
spawned thread (in addition to its local environment, like in SIMPLE).
This is needed in order to be able to spawn method invokations under
dynamic method dispatch; for example, spawn { run(); }
will need to look up the method run() in the newly created
thread, operation which will most likely fail unless the child thread
sees the object environment of the parent thread. Note that the
spawn statement of KOOL is more permissive than the threads
of Java. In fact, the latter can be implemented in terms of our
spawn—see the program threads.kool for a sketch.

Below is a subset of the values of SIMPLE, which are also values
of KOOL. We will add other values later in the semantics, such as
object and method closures.

  syntax Val ::= Int | Bool | String
               | array(Int,Int)
  syntax Exp ::= Val
  syntax Exps ::= Vals
  syntax KResult ::= Val
  syntax KResult ::= Vals

The semantics below are taken verbatim from the untyped SIMPLE
definition.

  syntax KItem ::= "undefined"

  rule <k> var X:Id; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> undefined ...</store>
       <nextLoc> L:Int => L +Int 1 </nextLoc>


  context var _:Id[HOLE];

  rule <k> var X:Id[N:Int]; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> array(L +Int 1, N)
                          (L +Int 1) ... (L +Int N) |-> undefined ...</store>
       <nextLoc> L:Int => L +Int 1 +Int N </nextLoc>
    requires N >=Int 0


  syntax Id ::= "$1" [token] | "$2" [token]
  rule var X:Id[N1:Int, N2:Int, Vs:Vals];
    => var X[N1];
       {
         var $1=X;
         for(var $2=0; $2 <= N1 - 1; ++$2) {
           var X[N2,Vs];
           $1[$2] = X;
         }
       }


  rule <k> X:Id => V ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V:Val ...</store>


  context ++(HOLE => lvalue(HOLE))
  rule <k> ++loc(L) => I +Int 1 ...</k>
       <store>... L |-> (I:Int => I +Int 1) ...</store>


  rule I1 + I2 => I1 +Int I2
  rule Str1 + Str2 => Str1 +String Str2
  rule I1 - I2 => I1 -Int I2
  rule I1 * I2 => I1 *Int I2
  rule I1 / I2 => I1 /Int I2 requires I2 =/=K 0
  rule I1 % I2 => I1 %Int I2 requires I2 =/=K 0
  rule - I => 0 -Int I
  rule I1 < I2 => I1 <Int I2
  rule I1 <= I2 => I1 <=Int I2
  rule I1 > I2 => I1 >Int I2
  rule I1 >= I2 => I1 >=Int I2

  rule V1:Val == V2:Val => V1 ==K V2
  rule V1:Val != V2:Val => V1 =/=K V2
  rule ! T => notBool(T)
  rule true  && E => E
  rule false && _ => false
  rule true  || _ => true
  rule false || E => E


  rule V:Val[N1:Int, N2:Int, Vs:Vals] => V[N1][N2, Vs]
    [anywhere]

  rule array(L,_)[N:Int] => lookup(L +Int N)
    [anywhere]


  rule sizeOf(array(_,N)) => N

The semantics of function application needs to change into dynamic
method dispatch invocation, which is defined shortly. However,
interestingly, the semantics of return stays unchanged.

  rule <k> return(V:Val); ~> _ => V ~> K </k>
       <control>
         <fstack> ListItem(fstackFrame(Env,K,XS,<crntObj> CO </crntObj>)) => .List ...</fstack>
         <xstack> _ => XS </xstack>
         <crntObj> _ => CO </crntObj>
       </control>
       <env> _ => Env </env>

  syntax Val ::= "nothing"
  rule return; => return nothing;


  rule <k> read() => I ...</k> <input> ListItem(I:Int) => .List ...</input>


  context (HOLE => lvalue(HOLE)) = _

  rule <k> loc(L) = V:Val => V ...</k> <store>... L |-> (_ => V) ...</store>

  rule {} => .K
  rule <k> { S } => S ~> setEnv(Env) ...</k>  <env> Env </env>


  rule S1::Stmt S2::Stmt => S1 ~> S2

  rule _:Val; => .K

  rule if ( true) S else _ => S
  rule if (false) _ else S => S

  rule while (E) S => if (E) {S while(E)S}

  rule <k> print(V:Val, Es => Es); ...</k> <output>... .List => ListItem(V) </output>
  rule print(.Vals); => .K


  syntax KItem ::= xstackFrame(Id,Stmt,K,Map,K)
  // TODO(KORE): drop the additional production once parsing issue #1842 is fixed
                 | (Id,Stmt,K,Map,K)

  syntax KItem ::= "popx"

  rule <k> (try S1 catch(X) {S2} => S1 ~> popx) ~> K </k>
       <control>
         <xstack> .List => ListItem(xstackFrame(X, S2, K, Env, C)) ...</xstack>
         C
       </control>
       <env> Env </env>

  rule <k> popx => .K ...</k>
       <xstack> ListItem(_) => .List ...</xstack>

  rule <k> throw V:Val; ~> _ => { var X = V; S2 } ~> K </k>
       <control>
         <xstack> ListItem(xstackFrame(X, S2, K, Env, C)) => .List ...</xstack>
         (_ => C)
       </control>
       <env> _ => Env </env>

Thread spawning needs a new semantics, because we want the child
thread to also share the object environment with its parent. The new
semantics of thread spawning will be defined shortly. However,
interestingly, the other concurrency constructs keep their semantics
from SIMPLE unchanged.

  // TODO(KORE): ..Bag should be . throughout this definition #1772
  rule (<thread>... <k>.K</k> <holds>H</holds> <id>T</id> ...</thread> => .Bag)
  /*
  rule (<thread>... <k>.</k> <holds>H</holds> <id>T</id> ...</thread> => .)
  */
       <busy> Busy => Busy -Set keys(H) </busy>
       <terminated>... .Set => SetItem(T) ...</terminated>

  rule <k> join T:Int; => .K ...</k>
       <terminated>... SetItem(T) ...</terminated>

  rule <k> acquire V:Val; => .K ...</k>
       <holds>... .Map => V |-> 0 ...</holds>
       <busy> Busy (.Set => SetItem(V)) </busy>
    requires (notBool(V in Busy:Set))

  rule <k> acquire V; => .K ...</k>
       <holds>... V:Val |-> (N:Int => N +Int 1) ...</holds>

  rule <k> release V:Val; => .K ...</k>
       <holds>... V |-> (N => N:Int -Int 1) ...</holds>
    requires N >Int 0

  rule <k> release V; => .K ...</k> <holds>... V:Val |-> 0 => .Map ...</holds>
       <busy>... SetItem(V) => .Set ...</busy>

  rule <k> rendezvous V:Val; => .K ...</k>
       <k> rendezvous V; => .K ...</k>

Unchanged auxiliary operations from untyped SIMPLE

  syntax Stmt ::= mkDecls(Ids,Vals)  [function]
  rule mkDecls((X:Id, Xs:Ids), (V:Val, Vs:Vals)) => var X=V; mkDecls(Xs,Vs)
  rule mkDecls(.Ids,.Vals) => {}

  // TODO(KORE): clarify sort inferences #1803
  syntax Exp ::= lookup(Int)
  /*
  syntax KItem ::= lookup(Int)
  */
  rule <k> lookup(L) => V ...</k> <store>... L |-> V:Val ...</store>

  syntax KItem ::= setEnv(Map)
  rule <k> setEnv(Env) => .K ...</k>  <env> _ => Env </env>
  rule (setEnv(_) => .K) ~> setEnv(_)
  // TODO: How can we make sure that the second rule above applies before the first one?
  //       Probably we'll deal with this using strategies, eventually.

  syntax Exp ::= lvalue(K)
  syntax Val ::= loc(Int)

  rule <k> lvalue(X:Id => loc(L)) ...</k> <env>... X |-> L:Int ...</env>

  context lvalue(_::Exp[HOLE::Exps])
  context lvalue(HOLE::Exp[_::Exps])

  rule lvalue(lookup(L:Int) => loc(L))


  syntax Map ::= Int "..." Int "|->" K
    [function]
  rule N...M |-> _ => .Map  requires N >Int M
  rule N...M |-> K => N |-> K (N +Int 1)...M |-> K  requires N <=Int M

Changes to the existing untyped SIMPLE semantics

When we extend a language, sometimes we need to do more than just add
new language constructs and semantics for them. Sometimes we want to
also extend the semantics of existing language constructs, in order to
get more from them.

Program initialization

In SIMPLE, once all the global declarations were processed, the
function main() was invoked. In KOOL, the global
declarations are classes, and their specific semantics is given
shortly; essentially, they are pre-processed one by one and added
into the class cell structure in the configuration.
Once all the classes are processed, the computation item
execute, which was placed right after the program in the
initial configuration, is reached. In SIMPLE, the program was
initialized by calling the method main(). In KOOL, the
program is initialized by creating an object instance of class
Main. This will also implicitly call the method
Main() (the Main class constructor). The emptiness
of the env cell below is just a sanity check, to make sure
that the user has not declared anything but classes at the top level
of the program.

  syntax KItem ::= "execute"
  rule <k> execute => new Main(.Exps); </k> <env> .Map </env>

The semantics of new (defined below) requires the
execution of all the class' declarations (and also of its
superclasses').

Object and method closures

Before we can define the semantics of method application (previously
called function application in SIMPLE), we need to add two more values
to the language, namely object and method closures:

  syntax Val ::= objectClosure(Id, List)
               | methodClosure(Id,Int,Ids,Stmt)

An object value consists of an objectClosure-wrapped bag
containing the current class of the object and the environment stack
of the object. The current class of an object will always be one of
the classes mapped to an environment in the environment stack of the
object. A method closure encapsulates the method's parameters and
code (last two arguments), as well as the object context in which the
method code should execute. This object context includes the current
class of the object (the first argument of methodClosure) and
the object environment stack (located in the object stored at the
location specified as the second argument of methodClosure).

Method application

KOOL has a complex mechanism to invoke methods, because it allows both
dynamic method dispatch and methods as first-class-citizen values (the
latter making it a higher-order language). The invocation mechanism
will be defined later. What is sufficient to know for now is that
the two arguments of the application construct eventually reduce to
values, the first being a method closure and the latter a list of
values. The semantics of the method closure application is then as
expected: the local environment and control are stacked, then we
switch to method closure's class and object environment and execute
the method body. The mkDecls construct is the one that came
with the unchanged semantics of SIMPLE above.

  syntax KItem ::= fstackFrame(Map,K,List,K)
  // TODO(KORE): drop the additional production once parsing issue #1842 is fixed
                 | (Map,K,K)

  rule <k> methodClosure(Class,OL,Xs,S)(Vs:Vals) ~> K
           => mkDecls(Xs,Vs) S return; </k>
       <env> Env => .Map </env>
       <store>... OL |-> objectClosure(_, EnvStack)...</store>
     //<br/> // TODO(KORE): support latex annotations #1799
       <control>
          <xstack> XS </xstack>
          <fstack> .List => ListItem(fstackFrame(Env, K, XS, <crntObj> Obj' </crntObj>))
          ...</fstack>
          <crntObj> Obj' => <crntClass> Class </crntClass> <envStack> EnvStack </envStack> </crntObj>
       </control>

Spawn

We want to extend the semantics of spawn to also share the
current object environment with the child thread, in addition to the
current environment. This extension will allow us to also use method
invocations in the spawned statements, which will be thus looked up as
expected, using dynamic method dispatch. This lookup operation would
fail if the child thread did not have access to its parent's object
environment.

  rule <thread>...
         <k> spawn S => !T:Int ...</k>
         <env> Env </env>
         <crntObj> Obj </crntObj>
       ...</thread>
       (.Bag => <thread>...
               <k> S </k>
               <env> Env </env>
               <id> !T </id>
               <crntObj> Obj </crntObj>
             ...</thread>)

Semantics of the new KOOL constructs

Class declaration

Initially, the classes forming the program are moved into their
corresponding cells:

  rule <k> class Class1 extends Class2 { S } => .K ...</k>
       <classes>... (.Bag => <classData>
                            <className> Class1 </className>
                            <baseClass> Class2 </baseClass>
                            <declarations> S </declarations>
                        </classData>)
       ...</classes>

Method declaration

Like in SIMPLE, method names are added to the environment and bound
to their code. However, unlike in SIMPLE where each function was
executed in the same environment, namely the program global
environment, a method in KOOL needs to be executed into its object's
environment. Thus, methods evaluate to closures, which encapsulate
their object's context (i.e., the current class and environment stack
of the object) in addition to method's parameters and body. This
approach to bind method names to method closures in the environment
will also allow objects to pass their methods to other objects, to
dynamically change their methods by assigning them other method
closures, and even to allow all these to be done from other objects.
This gives the KOOL programmer a lot of power; one should use this
power wisely, though, because programs can become easily hard to
understand and reason about if one overuses these features.

  rule <k> method F:Id(Xs:Ids) S => .K ...</k>
       <crntClass> Class:Id </crntClass>
       <location> OL:Int </location>
       <env> Env => Env[F <- L] </env>
       <store>... .Map => L |-> methodClosure(Class,OL,Xs,S) ...</store>
       <nextLoc> L => L +Int 1 </nextLoc>

New

The semantics of new consists of two actions: memory
allocation for the new object and execution of the corresponding
constructor. Then the created object is returned as the result of the
new operation; the value returned by the constructor, if any,
is discarded. The current environment and object are stored onto the
stack and recovered after new (according to the semantics of
return borrowed from SIMPLE, when the statement
return this; in the rule below is reached and evaluated),
because the object creation part of new will destroy them.
The rule below also initializes the object creation process by
emptying the local environment and the current object, and allocating
a location in the store where the created object will be eventually
stored (this is what the storeObj task after the object
creation task in the rule below will do—its rule is defined
shortly). The location where the object will be stored is also made
available in the crntObj cell, so that method closures can
refer to it (see rule above).

  syntax KItem ::= "envStackFrame" "(" Id "," Map ")"

  rule <k> new Class:Id(Vs:Vals) ~> K
           => create(Class) ~> storeObj ~> Class(Vs); return this; </k>
       <env> Env => .Map </env>
       <nextLoc> L:Int => L +Int 1 </nextLoc>
     //<br/> // TODO(KORE): support latex annotations #1799
       <control> <xstack> XS </xstack>
         <crntObj> Obj
                   => <crntClass> Object </crntClass>
                      <envStack> ListItem(envStackFrame(Object, .Map)) </envStack>
                      <location> L </location>
         </crntObj>
         <fstack> .List => ListItem(fstackFrame(Env, K, XS, <crntObj> Obj </crntObj>)) ...</fstack>
       </control>

The creation of a new object (the memory allocation part only) is
a recursive process, requiring to first create an object for the
superclass. A memory object representation is a layered structure:
for each class on the path from the instance class to the root of the
hierarchy there is a layer including the memory allocated for the
members (both fields and methods) of that class.

  syntax KItem ::= create(Id)

  rule <k> create(Class:Id)
           => create(Class1) ~> setCrntClass(Class) ~> S ~> addEnvLayer ...</k>
       <className> Class </className>
       <baseClass> Class1:Id </baseClass>
       <declarations> S </declarations>

  rule <k> create(Object) => .K ...</k>

The next operation sets the current class of the current object.
This is necessary to be done at each layer, because the current class
of the object is enclosed as part of the method closures (see the
semantics of method declarations above).

  syntax KItem ::= setCrntClass(Id)

  rule <k> setCrntClass(C) => .K ...</k>
       <crntClass> _ => C </crntClass>

The next operation adds a new tagged environment layer to the
current object and gets ready for the next layer by clearing the
environment (note that create expects the environment to be
empty).

  syntax KItem ::= "addEnvLayer"

  rule <k> addEnvLayer => .K ...</k>
       <env> Env => .Map </env>
       <crntClass> Class:Id </crntClass>
       <envStack> .List => ListItem(envStackFrame(Class, Env)) ...</envStack>

The following operation stores the created object at the location
reserved by new. Note that the location reserved by
new was temporarily stored in the crntObj cell
precisely for this purpose. Now that the newly created object is
stored at its location and that all method closures are aware of it,
the location is unnecessary and thus we delete it from the
crntObj cell.

  syntax KItem ::= "storeObj"

  rule <k> storeObj => .K ...</k>
       <crntObj> <crntClass> CC </crntClass> <envStack> ES </envStack> (<location> L:Int </location> => .Bag) </crntObj>
       <store>... .Map => L |-> objectClosure(CC, ES) ...</store>

Self reference

The semantics of this is straightforward: evaluate to the
current object.

  rule <k> this => objectClosure(CC, ES) ...</k>
       <crntObj> <crntClass> CC </crntClass> <envStack> ES </envStack> </crntObj>

Object member access

We can access an object member (field or method) either explicitly,
using the construct e.x, or implicitly, using only the member
name x directly. The borrowed semantics of SIMPLE will
already lookup a sole name in the local environment. The first rule
below reduces implicit member access to explicit access when the name
cannot be found in the local environment. There are two cases to
analyze for explicit object member access, depending upon whether the
object is a proper object or it is just a redirection to the parent
class via the construct super. In the first case, we
evaluate the object expression and lookup the member starting with the
current class (static scoping). Note the use of the conditional
evaluation context. In the second case, we just lookup the member
starting with the superclass of the current class. In both cases,
the lookupMember task eventually yields a lookup(L)
task for some appropriate location L, which will be further
solved with the corresponding rule borrowed from SIMPLE. Note that the
current object is not altered by super, so future method
invocations see the entire object, as needed for dynamic method dispatch.

  rule <k> X:Id => this . X ...</k> <env> Env:Map </env>
    requires notBool(X in keys(Env))

  context HOLE._::Id requires (HOLE =/=K super)

// TODO: explain how Assoc matching has been replaced with two rules here.
// Maybe also improve it a bit.

/*  rule objectClosure(<crntClass> Class:Id </crntClass>
                     <envStack>... envStackFrame(Class,EnvC) EStack </envStack>)
       . X:Id
    => lookupMember(envStackFrame(Class,EnvC) EStack, X) */

  rule objectClosure(Class:Id, ListItem(envStackFrame(Class,Env)) EStack)
       . X:Id
    => lookupMember(ListItem(envStackFrame(Class,Env)) EStack, X)
  rule objectClosure(Class:Id, (ListItem(envStackFrame(Class':Id,_)) => .List) _)
       . _X:Id
    requires Class =/=K Class'

/*  rule <k> super . X => lookupMember(EStack, X) ...</k>
       <crntClass> Class </crntClass>
       <envStack>... envStackFrame(Class,EnvC) EStack </envStack> */
  rule <k> super . X => lookupMember(EStack, X) ...</k>
       <crntClass> Class:Id </crntClass>
       <envStack> ListItem(envStackFrame(Class,_)) EStack </envStack>
  rule <k> super . _X ...</k>
       <crntClass> Class </crntClass>
       <envStack> ListItem(envStackFrame(Class':Id,_)) => .List ...</envStack>
    requires Class =/=K Class'

Method invocation

Unlike in SIMPLE, in KOOL application was declared strict only in its
second argument. That is because we want to ensure dynamic method
dispatch when the first argument is a method access. As a
consequence, we need to consider all the cases of interest for the
first argument and to explicitly say what to do in each case. In all
cases except for method access in a proper object (i.e., not
super), we want the same behavior for the first argument as
if it was not in a method invocation position. When it is a member
access (the third rule below), we look it up starting with the
instance class of the corresponding object. This ensures dynamic
dispatch for methods; it actually dynamically dispatches field
accesses, too, which is correct in KOOL, because one can assign method
closures to fields and the field appeared in a method invocation
context. The last context declaration below says that method
applications or array accesses are also allowed as first argument to
applications; that is because methods are allowed to return methods
and arrays are allowed to hold methods in KOOL, since it is
higher-order. If that is the case, then we want to evaluate the
method call or the array access.

  rule <k> (X:Id => V)(_:Exps) ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V:Val ...</store>

  rule <k> (X:Id => this . X)(_:Exps) ...</k>
       <env> Env </env>
    requires notBool(X in keys(Env))

  context HOLE._::Id(_) requires HOLE =/=K super

  rule (objectClosure(_, EStack) . X
    => lookupMember(EStack, X:Id))(_:Exps)

/*  rule <k> (super . X
            => lookupMember(EStack,X))(_:Exps)...</k>
       <crntClass> Class </crntClass>
       <envStack>... envStackFrame(Class,_) EStack </envStack> */
  rule <k> (super . X
            => lookupMember(EStack,X))(_:Exps)...</k>
       <crntClass> Class </crntClass>
       <envStack> ListItem(envStackFrame(Class,_)) EStack </envStack>
  rule <k> (super . _X)(_:Exps) ...</k>
       <crntClass> Class </crntClass>
       <envStack> ListItem(envStackFrame(Class':Id,_)) => .List ...</envStack>
    requires Class =/=K Class'

  // TODO(KORE): fix getKLabel #1801
  rule (A:Exp(B:Exps))(C:Exps) => A(B) ~> #freezerFunCall(C)
  rule (A:Exp[B:Exps])(C:Exps) => A[B] ~> #freezerFunCall(C)
  rule V:Val ~> #freezerFunCall(C:Exps) => V(C)
  syntax KItem ::= "#freezerFunCall" "(" K ")"
  /*
  context HOLE(_:Exps)
    when getKLabel(HOLE) ==K #klabel(`_(_)`) orBool getKLabel(HOLE) ==K #klabel(`_[_]`)
  */

Eventually, each of the rules above produces a lookup(L)
task as a replacement for the method. When that happens, we just
lookup the value at location L:

  rule <k> (lookup(L) => V)(_:Exps) ...</k>  <store>... L |-> V:Val ...</store>

The value V looked up above is expected to be a method closure,
in which case the semantics of method application given above will
apply. Otherwise, the execution will get stuck.

Instance Of

It searches the object environment for a layer corresponding to the
desired class. It returns true iff it can find the class,
otherwise it returns false; it only gets stuck when its first
argument does not evaluate to an object.

  rule objectClosure(_, ListItem(envStackFrame(C,_)) _)
       instanceOf C => true

  rule objectClosure(_, (ListItem(envStackFrame(C,_)) => .List) _)
       instanceOf C'  requires C =/=K C'
//TODO: remove the sort cast ::Id of C above, when sort inference bug fixed

  rule objectClosure(_, .List) instanceOf _ => false

Cast

In untyped KOOL, we prefer to not check the validity of casting. In
other words, any cast is allowed on any object, simply changing the
current class of the object to the desired class. The execution will
get stuck later if one attempts to access a field which is not
available. Moreover, the execution may complete successfully even
in the presence of invalid casts, provided that each accessed member
during the current execution is, or happens to be, available.

  rule (C) objectClosure(_ , EnvStack) => objectClosure(C ,EnvStack)

KOOL-specific auxiliary declarations and operations

Here we define all the auxiliary constructs used in the above
KOOL-specific semantics (those used in the SIMPLE fragment
have already been defined in a corresponding section above).

Objects as lvalues

The current machinery borrowed with the semantics of SIMPLE allows us
to enrich the set of lvalues, this way allowing new means to assign
values to locations. In KOOL, we want object member names to be
lvalues, so that we can assign values to them using the already
existing machinery. The first rule below ensures that the object is
always explicit, the evaluation context enforces the object to be
evaluated, and finally the second rule initiates the lookup for the
member's location based on the current class of the object.

  rule <k> lvalue(X:Id => this . X) ...</k>  <env> Env </env>
    requires notBool(X in keys(Env))

  context lvalue((HOLE . _)::Exp)

/*  rule lvalue(objectClosure(<crntClass> C </crntClass>
                            <envStack>... envStackFrame(C,EnvC) EStack </envStack>)
              . X
              => lookupMember(<envStack> envStackFrame(C,EnvC) EStack </envStack>,
                              X))  */
  rule lvalue(objectClosure(Class, ListItem(envStackFrame(Class,Env)) EStack)
              . X
              => lookupMember(ListItem(envStackFrame(Class,Env)) EStack,
                              X))
  rule lvalue(objectClosure(Class, (ListItem(envStackFrame(Class':Id,_)) => .List) _)
              . _X)
    requires Class =/=K Class'

Lookup member

It searches for the given member in the given environment stack,
starting with the most concrete class and going up in the hierarchy.

  // TODO(KORE): clarify sort inferences #1803
  syntax Exp ::= lookupMember(List, Id)  [function]
  /*
  syntax KItem ::= lookupMember(EnvStackCell,Id)  [function]
  */

//  rule lookupMember(<envStack> envStackFrame(_, <env>... X|->L ...</env>) ...</envStack>, X)
//    => lookup(L)
  rule lookupMember(ListItem(envStackFrame(_, X|->L _)) _, X)
    => lookup(L)

//  rule lookupMember(<envStack> envStackFrame(_, <env> Env </env>) => .List ...</envStack>, X)
//    when notBool(X in keys(Env))
  rule lookupMember(ListItem(envStackFrame(_, Env)) Rest, X) =>
       lookupMember(Rest, X)
    requires notBool(X in keys(Env))
//TODO: beautify the above

endmodule

Go to Lesson 2, KOOL typed dynamic.

KOOL — Untyped

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K semantic definition of the untyped KOOL language. KOOL
is aimed at being a pedagogical and research language that captures
the essence of the object-oriented programming paradigm. Its untyped
variant discussed here is simpler than the typed one, ignoring several
intricate aspects of types in the presence of objects. A program
consists of a set of class declarations. Each class can extend at
most one other class (KOOL is single-inheritance). A class can
declare a set of fields and a set of methods, all public and called
the class' members. Specifically, KOOL includes the
following features:

  • Class declarations, where a class may or may not explicitly
    extend another class. In case a class does not explicitly extend
    another class, then it is assumed that it extends the default top-most
    and empty (i.e., no members) class called Object. Each class
    is required to declare precisely one homonymous method, called its
    constructor. Each valid program should contain one class
    named Main, whose constructor, Main(), takes no
    arguments. The execution of a program consists of creating an object
    instance of class Main and invoking the constructor
    Main() on it, that is, of executing new Main();.

  • All features of SIMPLE (see examples/simple/untyped),
    i.e., multidimensional arrays, function (here called "method")
    abstractions with call-by-value parameter passing style and static
    scoping, blocks with locals, input/output, parametric exceptions, and
    concurrency via dynamic thread creation/termination and synchronization.
    The only change in the syntax of SIMPLE when imported in KOOL is the
    function declaration keyword, function, which is changed into
    method. The exact same desugaring macros from SIMPLE are
    also included in KOOL. We can think of KOOL's classes as embedding
    SIMPLE programs (extended with OO constructs, as discussed next).

  • Object creation using the new C(e1,...,en)
    expression construct. An object instance of class C is first
    created and then the constructor C(e1,...,en) is implicitly
    called on that object. KOOL only allows (and requires) one
    constructor per class. The class constructor can be called either
    implicitly during a new object creation for the class, or explicitly.
    The superclass constructor is not implicitly invoked when a
    class constructor is invoked; if you want to invoke the superclass
    constructor from a subclass constructor then you have to do it
    explicitly.

  • An expression construct this, which evaluates to the
    current object.

  • An expression construct super, which is used (only) in
    combination with member lookup (see next) to refer to a superclass
    field or method.

  • A member lookup expression construct e.x, where e
    is an expression (either an expression expected to evaluate to an object
    or the super construct) and x is a class member name,
    that is, a field or a method name.

  • Expression constructs e instanceOf C and
    (C) e, where e is an expression expected
    to evaluate to an object and C a class name. The former
    tells whether the class of e is a subclass of C,
    that is, whether e can be used as an instance of C,
    and the latter changes the class of e to C. These
    operations always succeed: the former returns a Boolean value, while
    the latter changes the current class of e to C
    regardless of whether it is safe to do so or not. The typed version
    of KOOL will check the safety of casting by ensuring that the instance
    class of the object is a subclass of C. In untyped KOOL we
    do not want to perform this check because we want to allow the
    programmer maximum of flexibility: if one always accesses only
    available members, then the program can execute successfully despite
    the potentially unsafe cast.

There are some specific aspects of KOOL that need to be discussed.

First, KOOL is higher-order, allowing function abstractions to be
treated like any other values in the language. For example, if
m is a method of object e then e.m
evaluates to the corresponding function abstraction. The function
abstraction is in fact a closure, because in addition to the method
parameters and body it also encapsulates the object value (i.e., the
environment of the object together with its current class—see below)
that e evaluates to. This way, function abstractions can be
invoked anywhere and have the capability to change the state of their
object. For example, if m is a method of object e
which increments a field c of e when invoked, and if
getm is another method of e which simply returns
m when invoked, then the double application
(e.getm())() has the same effect as e.m(), that is,
increments the counter c of e. Note that the
higher-order nature of KOOL was not originally planned; it came as a
natural consequence of evaluating methods to closures and we decided
to keep it. If you do not like it then do not use it.

Second, since all the fields and methods are public in KOOL and since
they can be redeclared in subclasses, it is not immediately clear how
to lookup the member x when we write e.x and
e is different from super. We distinguish two cases,
depending on whether e.x occurs in a method invocation
context (i.e., e.x(...)) or in a field context. KOOL has
dynamic method dispatch, so if e.x is invoked as a method
then x will be searched for starting with the instance class of
the object value to which e evaluates. If e.x
occurs in a non-method-invocation context then x will be
treated as a field (although it may hold a method closure due to the
higher-order nature of KOOL) and thus will be searched starting with
the current class of the object value of e (which, because of
this and casting, may be different from its instance class).
In order to achieve the above, each object value will consist of a
pair holding the current class of the object and an environment stack
with one layer for each class in the object's instance class hierarchy.

Third, although KOOL is dynamic method dispatch, its capabilities
described above are powerful enough to allow us to mimic static
method dispatch. For example, suppose that you want to invoke method
m() statically. Then all you need to do is to declare a
local variable and bind it to m, for example var staticm = m;, and
then call staticm(). This works because
staticm is first bound to the method closure that m
evaluates to, and then looked up as any local variable when invoked.
We only enable the dynamic method dispatch when we have an object
member on an application position, e.g., m().

In what follows, we limit our comments to the new, KOOL-specific
aspects of the language. We refer the reader to the untyped SIMPLE
language for documentation on the the remaining features, because
those were all borrowed from SIMPLE.

module KOOL-UNTYPED-SYNTAX
  imports DOMAINS-SYNTAX

Syntax

The syntax of KOOL extends that of SIMPLE with object-oriented
constructs. We removed from the K annotated syntax of SIMPLE two
constructs, namely the one for function declarations (because we want
to call them methods now) and the one for function application
(because application is not strict in the first argument
anymore—needs to initiate dynamic method dispatch). The additional
syntax includes:

  • First, we need a new dedicated identifier, Object, for
    the default top-most class.
  • Second, we rename the function keyword of SIMPLE into method.
  • Third, we add syntax for class declarations together with a
    macro making classes which extend nothing to extend Object.
  • Fourth, we change the strictness attribute of application
    into strict(2).
  • Finally, we add syntax and corresponding strictness
    for the KOOL object-oriented constructs.
  syntax Id ::= "Object" [token] | "Main" [token]

  syntax Stmt ::= "var" Exps ";"
                | "method" Id "(" Ids ")" Block  // called "function" in SIMPLE
                | "class" Id Block               // KOOL
                | "class" Id "extends" Id Block  // KOOL

  syntax Exp ::= Int | Bool | String | Id
               | "this"                                 // KOOL
               | "super"                                // KOOL
               | "(" Exp ")"             [bracket]
               | "++" Exp
               | Exp "instanceOf" Id     [strict(1)]    // KOOL
               | "(" Id ")" Exp          [strict(2)]    // KOOL  cast
               | "new" Id "(" Exps ")"   [strict(2)]    // KOOL
               | Exp "." Id                             // KOOL
               > Exp "[" Exps "]"        [strict]
               > Exp "(" Exps ")"        [strict(2)]    // was strict in SIMPLE
               | "-" Exp                 [strict]
               | "sizeOf" "(" Exp ")"    [strict]
               | "read" "(" ")"
               > left:
                 Exp "*" Exp             [strict, left]
               | Exp "/" Exp             [strict, left]
               | Exp "%" Exp             [strict, left]
               > left:
                 Exp "+" Exp             [strict, left]
               | Exp "-" Exp             [strict, left]
               > non-assoc:
                 Exp "<" Exp             [strict, non-assoc]
               | Exp "<=" Exp            [strict, non-assoc]
               | Exp ">" Exp             [strict, non-assoc]
               | Exp ">=" Exp            [strict, non-assoc]
               | Exp "==" Exp            [strict, non-assoc]
               | Exp "!=" Exp            [strict, non-assoc]
               > "!" Exp                 [strict]
               > left:
                 Exp "&&" Exp            [strict(1), left]
               | Exp "||" Exp            [strict(1), left]
               > "spawn" Block
               > Exp "=" Exp             [strict(2), right]

  syntax Ids  ::= List{Id,","}

  syntax Exps ::= List{Exp,","}          [strict, overload(exps)]
  syntax Val
  syntax Vals ::= List{Val,","}          [overload(exps)]

  syntax Block ::= "{" "}"
                | "{" Stmt "}"

  syntax Stmt ::= Block
                | Exp ";"                               [strict]
                | "if" "(" Exp ")" Block "else" Block   [avoid, strict(1)]
                | "if" "(" Exp ")" Block                [macro]
                | "while" "(" Exp ")" Block
                | "for" "(" Stmt Exp ";" Exp ")" Block  [macro]
                | "return" Exp ";"                      [strict]
                | "return" ";"                          [macro]
                | "print" "(" Exps ")" ";"              [strict]
                | "try" Block "catch" "(" Id ")" Block
                | "throw" Exp ";"                       [strict]
                | "join" Exp ";"                        [strict]
                | "acquire" Exp ";"                     [strict]
                | "release" Exp ";"                     [strict]
                | "rendezvous" Exp ";"                  [strict]

  syntax Stmt ::= Stmt Stmt                          [right]

Old desugaring rules, from SIMPLE

  rule if (E) S => if (E) S else {}
  rule for(Start Cond; Step) {S} => {Start while (Cond) {S Step;}}
  rule var E1::Exp, E2::Exp, Es::Exps; => var E1; var E2, Es;       [anywhere]
  rule var X::Id = E; => var X; X = E;                              [anywhere]

New desugaring rule

  rule class C:Id S => class C extends Object S                     // KOOL

endmodule

Semantics

We first discuss the new configuration of KOOL, which extends that of
SIMPLE. Then we include the semantics of the constructs borrowed from
SIMPLE unchanged; we refrain from discussing those, because they were
already discussed in the K definition of SIMPLE. Then we discuss
changes to SIMPLE's semantics needed for the more general meaning of
the previous SIMPLE constructs (for example for thread spawning,
assignment, etc.). Finally, we discuss in detail the
semantics of the additional KOOL constructs.

module KOOL-UNTYPED
  imports KOOL-UNTYPED-SYNTAX
  imports DOMAINS

Configuration

KOOL removes one cell and adds two nested cells to the configuration
of SIMPLE. The cell which is removed is the one holding the global
environment, because a KOOL program consists of a set of classes only,
with no global declarations. In fact, since informally speaking each
KOOL class now includes a SIMPLE program, it is safe to say that the
global variables in SIMPLE became class fields in KOOL. Let us now
discuss the new cells that are added to the configuration of SIMPLE.

  • The cell crntObj holds data pertaining to the current
    object, that is, the object environment in which the code in cell
    k executes: crntClass holds the current class (which
    can change as methods of the current object are invoked);
    envStack holds the stack of environments as a list,
    each layer corresponding to one class in the objects' instance class
    hierarchy; location, which is optional, holds the location in
    the store where the current object is or has to be located (this is
    useful both for method closures and for the semantics of object
    creation).

  • The cell classes holds all the declared classes, each
    class being held in its own class cell which contains a name
    (className), a parent (extends), and the actual
    member declarations (declarations).

  // the syntax declarations below are required because the sorts are
  // referenced directly by a production and, because of the way KIL to KORE
  // is implemented, the configuration syntax is not available yet
  // should simply work once KIL is removed completely
  // check other definitions for this hack as well
  syntax EnvCell
  syntax ControlCell
  syntax EnvStackCell
  syntax CrntObjCellFragment

  configuration <T color="red">
                  <threads color="orange">
                    <thread multiplicity="*" type="Set" color="yellow">
                      <k color="green"> $PGM:Stmt ~> execute </k>
                    //<br/> // TODO(KORE): support latex annotations #1799
                      <control color="cyan">
                        <fstack color="blue"> .List </fstack>
                        <xstack color="purple"> .List </xstack>
                      //<br/> // TODO(KORE): support latex annotations #1799
                        <crntObj color="Fuchsia">  // KOOL
                           <crntClass> Object </crntClass>
                           <envStack> .List </envStack>
                           <location multiplicity="?"> .K </location>
                        </crntObj>
                      </control>
                    //<br/> // TODO(KORE): support latex annotations #1799
                      <env color="violet"> .Map </env>
                      <holds color="black"> .Map </holds>
                      <id color="pink"> 0 </id>
                    </thread>
                  </threads>
                //<br/> // TODO(KORE): support latex annotations #1799
                  <store color="white"> .Map </store>
                  <busy color="cyan">.Set </busy>
                  <terminated color="red"> .Set </terminated>
                  <input color="magenta" stream="stdin"> .List </input>
                  <output color="brown" stream="stdout"> .List </output>
                  <nextLoc color="gray"> 0 </nextLoc>
                //<br/> // TODO(KORE): support latex annotations #1799
                  <classes color="Fuchsia">        // KOOL
                     <classData multiplicity="*" type="Map" color="Fuchsia">
                        // the Map has as its key the first child of the cell,
                        // in this case the className cell.
                        <className color="Fuchsia"> Main </className>
                        <baseClass color="Fuchsia"> Object </baseClass>
                        <declarations color="Fuchsia"> .K </declarations>
                     </classData>
                  </classes>
                </T>

Unchanged Semantics from untyped SIMPLE

The semantics below is taken over from SIMPLE unchanged.
The semantics of function declaration and invocation, including the
use of the special lambda abstraction value, needs to change
in order to account for the fact that methods are now invoked into
their object's environment. The semantics of function return actually
stays unchanged. Also, the semantics of program initialization is
different: now we have to create an instance of the Main
class which also calls the constructor Main(), while in
SIMPLE we only had to invoke the function Main().
Finally, the semantics of thread spawning needs to change, too: the
parent thread needs to also share its object environment with the
spawned thread (in addition to its local environment, like in SIMPLE).
This is needed in order to be able to spawn method invokations under
dynamic method dispatch; for example, spawn { run(); }
will need to look up the method run() in the newly created
thread, operation which will most likely fail unless the child thread
sees the object environment of the parent thread. Note that the
spawn statement of KOOL is more permissive than the threads
of Java. In fact, the latter can be implemented in terms of our
spawn—see the program threads.kool for a sketch.

Below is a subset of the values of SIMPLE, which are also values
of KOOL. We will add other values later in the semantics, such as
object and method closures.

  syntax Val ::= Int | Bool | String
               | array(Int,Int)
  syntax Exp ::= Val
  syntax Exps ::= Vals
  syntax KResult ::= Val
  syntax KResult ::= Vals

The semantics below are taken verbatim from the untyped SIMPLE
definition.

  syntax KItem ::= "undefined"

  rule <k> var X:Id; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> undefined ...</store>
       <nextLoc> L:Int => L +Int 1 </nextLoc>


  context var _:Id[HOLE];

  rule <k> var X:Id[N:Int]; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> array(L +Int 1, N)
                          (L +Int 1) ... (L +Int N) |-> undefined ...</store>
       <nextLoc> L:Int => L +Int 1 +Int N </nextLoc>
    requires N >=Int 0


  syntax Id ::= "$1" [token] | "$2" [token]
  rule var X:Id[N1:Int, N2:Int, Vs:Vals];
    => var X[N1];
       {
         var $1=X;
         for(var $2=0; $2 <= N1 - 1; ++$2) {
           var X[N2,Vs];
           $1[$2] = X;
         }
       }


  rule <k> X:Id => V ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V:Val ...</store>


  context ++(HOLE => lvalue(HOLE))
  rule <k> ++loc(L) => I +Int 1 ...</k>
       <store>... L |-> (I:Int => I +Int 1) ...</store>


  rule I1 + I2 => I1 +Int I2
  rule Str1 + Str2 => Str1 +String Str2
  rule I1 - I2 => I1 -Int I2
  rule I1 * I2 => I1 *Int I2
  rule I1 / I2 => I1 /Int I2 requires I2 =/=K 0
  rule I1 % I2 => I1 %Int I2 requires I2 =/=K 0
  rule - I => 0 -Int I
  rule I1 < I2 => I1 <Int I2
  rule I1 <= I2 => I1 <=Int I2
  rule I1 > I2 => I1 >Int I2
  rule I1 >= I2 => I1 >=Int I2

  rule V1:Val == V2:Val => V1 ==K V2
  rule V1:Val != V2:Val => V1 =/=K V2
  rule ! T => notBool(T)
  rule true  && E => E
  rule false && _ => false
  rule true  || _ => true
  rule false || E => E


  rule V:Val[N1:Int, N2:Int, Vs:Vals] => V[N1][N2, Vs]
    [anywhere]

  rule array(L,_)[N:Int] => lookup(L +Int N)
    [anywhere]


  rule sizeOf(array(_,N)) => N

The semantics of function application needs to change into dynamic
method dispatch invocation, which is defined shortly. However,
interestingly, the semantics of return stays unchanged.

  rule <k> return(V:Val); ~> _ => V ~> K </k>
       <control>
         <fstack> ListItem(fstackFrame(Env,K,XS,<crntObj> CO </crntObj>)) => .List ...</fstack>
         <xstack> _ => XS </xstack>
         <crntObj> _ => CO </crntObj>
       </control>
       <env> _ => Env </env>

  syntax Val ::= "nothing"
  rule return; => return nothing;


  rule <k> read() => I ...</k> <input> ListItem(I:Int) => .List ...</input>


  context (HOLE => lvalue(HOLE)) = _

  rule <k> loc(L) = V:Val => V ...</k> <store>... L |-> (_ => V) ...</store>

  rule {} => .K
  rule <k> { S } => S ~> setEnv(Env) ...</k>  <env> Env </env>


  rule S1::Stmt S2::Stmt => S1 ~> S2

  rule _:Val; => .K

  rule if ( true) S else _ => S
  rule if (false) _ else S => S

  rule while (E) S => if (E) {S while(E)S}

  rule <k> print(V:Val, Es => Es); ...</k> <output>... .List => ListItem(V) </output>
  rule print(.Vals); => .K


  syntax KItem ::= xstackFrame(Id,Stmt,K,Map,K)
  // TODO(KORE): drop the additional production once parsing issue #1842 is fixed
                 | (Id,Stmt,K,Map,K)

  syntax KItem ::= "popx"

  rule <k> (try S1 catch(X) {S2} => S1 ~> popx) ~> K </k>
       <control>
         <xstack> .List => ListItem(xstackFrame(X, S2, K, Env, C)) ...</xstack>
         C
       </control>
       <env> Env </env>

  rule <k> popx => .K ...</k>
       <xstack> ListItem(_) => .List ...</xstack>

  rule <k> throw V:Val; ~> _ => { var X = V; S2 } ~> K </k>
       <control>
         <xstack> ListItem(xstackFrame(X, S2, K, Env, C)) => .List ...</xstack>
         (_ => C)
       </control>
       <env> _ => Env </env>

Thread spawning needs a new semantics, because we want the child
thread to also share the object environment with its parent. The new
semantics of thread spawning will be defined shortly. However,
interestingly, the other concurrency constructs keep their semantics
from SIMPLE unchanged.

  // TODO(KORE): ..Bag should be . throughout this definition #1772
  rule (<thread>... <k>.K</k> <holds>H</holds> <id>T</id> ...</thread> => .Bag)
  /*
  rule (<thread>... <k>.</k> <holds>H</holds> <id>T</id> ...</thread> => .)
  */
       <busy> Busy => Busy -Set keys(H) </busy>
       <terminated>... .Set => SetItem(T) ...</terminated>

  rule <k> join T:Int; => .K ...</k>
       <terminated>... SetItem(T) ...</terminated>

  rule <k> acquire V:Val; => .K ...</k>
       <holds>... .Map => V |-> 0 ...</holds>
       <busy> Busy (.Set => SetItem(V)) </busy>
    requires (notBool(V in Busy:Set))

  rule <k> acquire V; => .K ...</k>
       <holds>... V:Val |-> (N:Int => N +Int 1) ...</holds>

  rule <k> release V:Val; => .K ...</k>
       <holds>... V |-> (N => N:Int -Int 1) ...</holds>
    requires N >Int 0

  rule <k> release V; => .K ...</k> <holds>... V:Val |-> 0 => .Map ...</holds>
       <busy>... SetItem(V) => .Set ...</busy>

  rule <k> rendezvous V:Val; => .K ...</k>
       <k> rendezvous V; => .K ...</k>

Unchanged auxiliary operations from untyped SIMPLE

  syntax Stmt ::= mkDecls(Ids,Vals)  [function]
  rule mkDecls((X:Id, Xs:Ids), (V:Val, Vs:Vals)) => var X=V; mkDecls(Xs,Vs)
  rule mkDecls(.Ids,.Vals) => {}

  // TODO(KORE): clarify sort inferences #1803
  syntax Exp ::= lookup(Int)
  /*
  syntax KItem ::= lookup(Int)
  */
  rule <k> lookup(L) => V ...</k> <store>... L |-> V:Val ...</store>

  syntax KItem ::= setEnv(Map)
  rule <k> setEnv(Env) => .K ...</k>  <env> _ => Env </env>
  rule (setEnv(_) => .K) ~> setEnv(_)
  // TODO: How can we make sure that the second rule above applies before the first one?
  //       Probably we'll deal with this using strategies, eventually.

  syntax Exp ::= lvalue(K)
  syntax Val ::= loc(Int)

  rule <k> lvalue(X:Id => loc(L)) ...</k> <env>... X |-> L:Int ...</env>

  context lvalue(_::Exp[HOLE::Exps])
  context lvalue(HOLE::Exp[_::Exps])

  rule lvalue(lookup(L:Int) => loc(L))


  syntax Map ::= Int "..." Int "|->" K
    [function]
  rule N...M |-> _ => .Map  requires N >Int M
  rule N...M |-> K => N |-> K (N +Int 1)...M |-> K  requires N <=Int M

Changes to the existing untyped SIMPLE semantics

When we extend a language, sometimes we need to do more than just add
new language constructs and semantics for them. Sometimes we want to
also extend the semantics of existing language constructs, in order to
get more from them.

Program initialization

In SIMPLE, once all the global declarations were processed, the
function main() was invoked. In KOOL, the global
declarations are classes, and their specific semantics is given
shortly; essentially, they are pre-processed one by one and added
into the class cell structure in the configuration.
Once all the classes are processed, the computation item
execute, which was placed right after the program in the
initial configuration, is reached. In SIMPLE, the program was
initialized by calling the method main(). In KOOL, the
program is initialized by creating an object instance of class
Main. This will also implicitly call the method
Main() (the Main class constructor). The emptiness
of the env cell below is just a sanity check, to make sure
that the user has not declared anything but classes at the top level
of the program.

  syntax KItem ::= "execute"
  rule <k> execute => new Main(.Exps); </k> <env> .Map </env>

The semantics of new (defined below) requires the
execution of all the class' declarations (and also of its
superclasses').

Object and method closures

Before we can define the semantics of method application (previously
called function application in SIMPLE), we need to add two more values
to the language, namely object and method closures:

  syntax Val ::= objectClosure(Id, List)
               | methodClosure(Id,Int,Ids,Stmt)

An object value consists of an objectClosure-wrapped bag
containing the current class of the object and the environment stack
of the object. The current class of an object will always be one of
the classes mapped to an environment in the environment stack of the
object. A method closure encapsulates the method's parameters and
code (last two arguments), as well as the object context in which the
method code should execute. This object context includes the current
class of the object (the first argument of methodClosure) and
the object environment stack (located in the object stored at the
location specified as the second argument of methodClosure).

Method application

KOOL has a complex mechanism to invoke methods, because it allows both
dynamic method dispatch and methods as first-class-citizen values (the
latter making it a higher-order language). The invocation mechanism
will be defined later. What is sufficient to know for now is that
the two arguments of the application construct eventually reduce to
values, the first being a method closure and the latter a list of
values. The semantics of the method closure application is then as
expected: the local environment and control are stacked, then we
switch to method closure's class and object environment and execute
the method body. The mkDecls construct is the one that came
with the unchanged semantics of SIMPLE above.

  syntax KItem ::= fstackFrame(Map,K,List,K)
  // TODO(KORE): drop the additional production once parsing issue #1842 is fixed
                 | (Map,K,K)

  rule <k> methodClosure(Class,OL,Xs,S)(Vs:Vals) ~> K
           => mkDecls(Xs,Vs) S return; </k>
       <env> Env => .Map </env>
       <store>... OL |-> objectClosure(_, EnvStack)...</store>
     //<br/> // TODO(KORE): support latex annotations #1799
       <control>
          <xstack> XS </xstack>
          <fstack> .List => ListItem(fstackFrame(Env, K, XS, <crntObj> Obj' </crntObj>))
          ...</fstack>
          <crntObj> Obj' => <crntClass> Class </crntClass> <envStack> EnvStack </envStack> </crntObj>
       </control>

Spawn

We want to extend the semantics of spawn to also share the
current object environment with the child thread, in addition to the
current environment. This extension will allow us to also use method
invocations in the spawned statements, which will be thus looked up as
expected, using dynamic method dispatch. This lookup operation would
fail if the child thread did not have access to its parent's object
environment.

  rule <thread>...
         <k> spawn S => !T:Int ...</k>
         <env> Env </env>
         <crntObj> Obj </crntObj>
       ...</thread>
       (.Bag => <thread>...
               <k> S </k>
               <env> Env </env>
               <id> !T </id>
               <crntObj> Obj </crntObj>
             ...</thread>)

Semantics of the new KOOL constructs

Class declaration

Initially, the classes forming the program are moved into their
corresponding cells:

  rule <k> class Class1 extends Class2 { S } => .K ...</k>
       <classes>... (.Bag => <classData>
                            <className> Class1 </className>
                            <baseClass> Class2 </baseClass>
                            <declarations> S </declarations>
                        </classData>)
       ...</classes>

Method declaration

Like in SIMPLE, method names are added to the environment and bound
to their code. However, unlike in SIMPLE where each function was
executed in the same environment, namely the program global
environment, a method in KOOL needs to be executed into its object's
environment. Thus, methods evaluate to closures, which encapsulate
their object's context (i.e., the current class and environment stack
of the object) in addition to method's parameters and body. This
approach to bind method names to method closures in the environment
will also allow objects to pass their methods to other objects, to
dynamically change their methods by assigning them other method
closures, and even to allow all these to be done from other objects.
This gives the KOOL programmer a lot of power; one should use this
power wisely, though, because programs can become easily hard to
understand and reason about if one overuses these features.

  rule <k> method F:Id(Xs:Ids) S => .K ...</k>
       <crntClass> Class:Id </crntClass>
       <location> OL:Int </location>
       <env> Env => Env[F <- L] </env>
       <store>... .Map => L |-> methodClosure(Class,OL,Xs,S) ...</store>
       <nextLoc> L => L +Int 1 </nextLoc>

New

The semantics of new consists of two actions: memory
allocation for the new object and execution of the corresponding
constructor. Then the created object is returned as the result of the
new operation; the value returned by the constructor, if any,
is discarded. The current environment and object are stored onto the
stack and recovered after new (according to the semantics of
return borrowed from SIMPLE, when the statement
return this; in the rule below is reached and evaluated),
because the object creation part of new will destroy them.
The rule below also initializes the object creation process by
emptying the local environment and the current object, and allocating
a location in the store where the created object will be eventually
stored (this is what the storeObj task after the object
creation task in the rule below will do—its rule is defined
shortly). The location where the object will be stored is also made
available in the crntObj cell, so that method closures can
refer to it (see rule above).

  syntax KItem ::= "envStackFrame" "(" Id "," Map ")"

  rule <k> new Class:Id(Vs:Vals) ~> K
           => create(Class) ~> storeObj ~> Class(Vs); return this; </k>
       <env> Env => .Map </env>
       <nextLoc> L:Int => L +Int 1 </nextLoc>
     //<br/> // TODO(KORE): support latex annotations #1799
       <control> <xstack> XS </xstack>
         <crntObj> Obj
                   => <crntClass> Object </crntClass>
                      <envStack> ListItem(envStackFrame(Object, .Map)) </envStack>
                      <location> L </location>
         </crntObj>
         <fstack> .List => ListItem(fstackFrame(Env, K, XS, <crntObj> Obj </crntObj>)) ...</fstack>
       </control>

The creation of a new object (the memory allocation part only) is
a recursive process, requiring to first create an object for the
superclass. A memory object representation is a layered structure:
for each class on the path from the instance class to the root of the
hierarchy there is a layer including the memory allocated for the
members (both fields and methods) of that class.

  syntax KItem ::= create(Id)

  rule <k> create(Class:Id)
           => create(Class1) ~> setCrntClass(Class) ~> S ~> addEnvLayer ...</k>
       <className> Class </className>
       <baseClass> Class1:Id </baseClass>
       <declarations> S </declarations>

  rule <k> create(Object) => .K ...</k>

The next operation sets the current class of the current object.
This is necessary to be done at each layer, because the current class
of the object is enclosed as part of the method closures (see the
semantics of method declarations above).

  syntax KItem ::= setCrntClass(Id)

  rule <k> setCrntClass(C) => .K ...</k>
       <crntClass> _ => C </crntClass>

The next operation adds a new tagged environment layer to the
current object and gets ready for the next layer by clearing the
environment (note that create expects the environment to be
empty).

  syntax KItem ::= "addEnvLayer"

  rule <k> addEnvLayer => .K ...</k>
       <env> Env => .Map </env>
       <crntClass> Class:Id </crntClass>
       <envStack> .List => ListItem(envStackFrame(Class, Env)) ...</envStack>

The following operation stores the created object at the location
reserved by new. Note that the location reserved by
new was temporarily stored in the crntObj cell
precisely for this purpose. Now that the newly created object is
stored at its location and that all method closures are aware of it,
the location is unnecessary and thus we delete it from the
crntObj cell.

  syntax KItem ::= "storeObj"

  rule <k> storeObj => .K ...</k>
       <crntObj> <crntClass> CC </crntClass> <envStack> ES </envStack> (<location> L:Int </location> => .Bag) </crntObj>
       <store>... .Map => L |-> objectClosure(CC, ES) ...</store>

Self reference

The semantics of this is straightforward: evaluate to the
current object.

  rule <k> this => objectClosure(CC, ES) ...</k>
       <crntObj> <crntClass> CC </crntClass> <envStack> ES </envStack> </crntObj>

Object member access

We can access an object member (field or method) either explicitly,
using the construct e.x, or implicitly, using only the member
name x directly. The borrowed semantics of SIMPLE will
already lookup a sole name in the local environment. The first rule
below reduces implicit member access to explicit access when the name
cannot be found in the local environment. There are two cases to
analyze for explicit object member access, depending upon whether the
object is a proper object or it is just a redirection to the parent
class via the construct super. In the first case, we
evaluate the object expression and lookup the member starting with the
current class (static scoping). Note the use of the conditional
evaluation context. In the second case, we just lookup the member
starting with the superclass of the current class. In both cases,
the lookupMember task eventually yields a lookup(L)
task for some appropriate location L, which will be further
solved with the corresponding rule borrowed from SIMPLE. Note that the
current object is not altered by super, so future method
invocations see the entire object, as needed for dynamic method dispatch.

  rule <k> X:Id => this . X ...</k> <env> Env:Map </env>
    requires notBool(X in keys(Env))

  context HOLE._::Id requires (HOLE =/=K super)

// TODO: explain how Assoc matching has been replaced with two rules here.
// Maybe also improve it a bit.

/*  rule objectClosure(<crntClass> Class:Id </crntClass>
                     <envStack>... envStackFrame(Class,EnvC) EStack </envStack>)
       . X:Id
    => lookupMember(envStackFrame(Class,EnvC) EStack, X) */

  rule objectClosure(Class:Id, ListItem(envStackFrame(Class,Env)) EStack)
       . X:Id
    => lookupMember(ListItem(envStackFrame(Class,Env)) EStack, X)
  rule objectClosure(Class:Id, (ListItem(envStackFrame(Class':Id,_)) => .List) _)
       . _X:Id
    requires Class =/=K Class'

/*  rule <k> super . X => lookupMember(EStack, X) ...</k>
       <crntClass> Class </crntClass>
       <envStack>... envStackFrame(Class,EnvC) EStack </envStack> */
  rule <k> super . X => lookupMember(EStack, X) ...</k>
       <crntClass> Class:Id </crntClass>
       <envStack> ListItem(envStackFrame(Class,_)) EStack </envStack>
  rule <k> super . _X ...</k>
       <crntClass> Class </crntClass>
       <envStack> ListItem(envStackFrame(Class':Id,_)) => .List ...</envStack>
    requires Class =/=K Class'

Method invocation

Unlike in SIMPLE, in KOOL application was declared strict only in its
second argument. That is because we want to ensure dynamic method
dispatch when the first argument is a method access. As a
consequence, we need to consider all the cases of interest for the
first argument and to explicitly say what to do in each case. In all
cases except for method access in a proper object (i.e., not
super), we want the same behavior for the first argument as
if it was not in a method invocation position. When it is a member
access (the third rule below), we look it up starting with the
instance class of the corresponding object. This ensures dynamic
dispatch for methods; it actually dynamically dispatches field
accesses, too, which is correct in KOOL, because one can assign method
closures to fields and the field appeared in a method invocation
context. The last context declaration below says that method
applications or array accesses are also allowed as first argument to
applications; that is because methods are allowed to return methods
and arrays are allowed to hold methods in KOOL, since it is
higher-order. If that is the case, then we want to evaluate the
method call or the array access.

  rule <k> (X:Id => V)(_:Exps) ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V:Val ...</store>

  rule <k> (X:Id => this . X)(_:Exps) ...</k>
       <env> Env </env>
    requires notBool(X in keys(Env))

  context HOLE._::Id(_) requires HOLE =/=K super

  rule (objectClosure(_, EStack) . X
    => lookupMember(EStack, X:Id))(_:Exps)

/*  rule <k> (super . X
            => lookupMember(EStack,X))(_:Exps)...</k>
       <crntClass> Class </crntClass>
       <envStack>... envStackFrame(Class,_) EStack </envStack> */
  rule <k> (super . X
            => lookupMember(EStack,X))(_:Exps)...</k>
       <crntClass> Class </crntClass>
       <envStack> ListItem(envStackFrame(Class,_)) EStack </envStack>
  rule <k> (super . _X)(_:Exps) ...</k>
       <crntClass> Class </crntClass>
       <envStack> ListItem(envStackFrame(Class':Id,_)) => .List ...</envStack>
    requires Class =/=K Class'

  // TODO(KORE): fix getKLabel #1801
  rule (A:Exp(B:Exps))(C:Exps) => A(B) ~> #freezerFunCall(C)
  rule (A:Exp[B:Exps])(C:Exps) => A[B] ~> #freezerFunCall(C)
  rule V:Val ~> #freezerFunCall(C:Exps) => V(C)
  syntax KItem ::= "#freezerFunCall" "(" K ")"
  /*
  context HOLE(_:Exps)
    when getKLabel(HOLE) ==K #klabel(`_(_)`) orBool getKLabel(HOLE) ==K #klabel(`_[_]`)
  */

Eventually, each of the rules above produces a lookup(L)
task as a replacement for the method. When that happens, we just
lookup the value at location L:

  rule <k> (lookup(L) => V)(_:Exps) ...</k>  <store>... L |-> V:Val ...</store>

The value V looked up above is expected to be a method closure,
in which case the semantics of method application given above will
apply. Otherwise, the execution will get stuck.

Instance Of

It searches the object environment for a layer corresponding to the
desired class. It returns true iff it can find the class,
otherwise it returns false; it only gets stuck when its first
argument does not evaluate to an object.

  rule objectClosure(_, ListItem(envStackFrame(C,_)) _)
       instanceOf C => true

  rule objectClosure(_, (ListItem(envStackFrame(C,_)) => .List) _)
       instanceOf C'  requires C =/=K C'
//TODO: remove the sort cast ::Id of C above, when sort inference bug fixed

  rule objectClosure(_, .List) instanceOf _ => false

Cast

In untyped KOOL, we prefer to not check the validity of casting. In
other words, any cast is allowed on any object, simply changing the
current class of the object to the desired class. The execution will
get stuck later if one attempts to access a field which is not
available. Moreover, the execution may complete successfully even
in the presence of invalid casts, provided that each accessed member
during the current execution is, or happens to be, available.

  rule (C) objectClosure(_ , EnvStack) => objectClosure(C ,EnvStack)

KOOL-specific auxiliary declarations and operations

Here we define all the auxiliary constructs used in the above
KOOL-specific semantics (those used in the SIMPLE fragment
have already been defined in a corresponding section above).

Objects as lvalues

The current machinery borrowed with the semantics of SIMPLE allows us
to enrich the set of lvalues, this way allowing new means to assign
values to locations. In KOOL, we want object member names to be
lvalues, so that we can assign values to them using the already
existing machinery. The first rule below ensures that the object is
always explicit, the evaluation context enforces the object to be
evaluated, and finally the second rule initiates the lookup for the
member's location based on the current class of the object.

  rule <k> lvalue(X:Id => this . X) ...</k>  <env> Env </env>
    requires notBool(X in keys(Env))

  context lvalue((HOLE . _)::Exp)

/*  rule lvalue(objectClosure(<crntClass> C </crntClass>
                            <envStack>... envStackFrame(C,EnvC) EStack </envStack>)
              . X
              => lookupMember(<envStack> envStackFrame(C,EnvC) EStack </envStack>,
                              X))  */
  rule lvalue(objectClosure(Class, ListItem(envStackFrame(Class,Env)) EStack)
              . X
              => lookupMember(ListItem(envStackFrame(Class,Env)) EStack,
                              X))
  rule lvalue(objectClosure(Class, (ListItem(envStackFrame(Class':Id,_)) => .List) _)
              . _X)
    requires Class =/=K Class'

Lookup member

It searches for the given member in the given environment stack,
starting with the most concrete class and going up in the hierarchy.

  // TODO(KORE): clarify sort inferences #1803
  syntax Exp ::= lookupMember(List, Id)  [function]
  /*
  syntax KItem ::= lookupMember(EnvStackCell,Id)  [function]
  */

//  rule lookupMember(<envStack> envStackFrame(_, <env>... X|->L ...</env>) ...</envStack>, X)
//    => lookup(L)
  rule lookupMember(ListItem(envStackFrame(_, X|->L _)) _, X)
    => lookup(L)

//  rule lookupMember(<envStack> envStackFrame(_, <env> Env </env>) => .List ...</envStack>, X)
//    when notBool(X in keys(Env))
  rule lookupMember(ListItem(envStackFrame(_, Env)) Rest, X) =>
       lookupMember(Rest, X)
    requires notBool(X in keys(Env))
//TODO: beautify the above

endmodule

Go to Lesson 2, KOOL typed dynamic.

KOOL — Typed — Dynamic

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K dynamic semantics of the typed KOOL language. It is
very similar to the semantics of the untyped KOOL, the difference
being that we now check the typing policy dynamically. Since we have
to now declare the types of variables and methods, we adopt a syntax
for those which is close to Java. Like in the semantics of
untyped KOOL, where we borrowed almost all the semantics of untyped
SIMPLE, we are going to also borrow much of the semantics of
dynamically typed SIMPLE here. We will highlight the differences
between the dynamically typed and the untyped KOOL as we proceed with
the semantics. In general, the type policy of the typed KOOL language
is similar to that of Java. You may find it useful to also read
the discussion in the preamble of the static semantics of typed KOOL
before proceeding.

module KOOL-TYPED-DYNAMIC-SYNTAX
  imports DOMAINS-SYNTAX

Syntax

Like for the untyped KOOL language, the syntax of typed KOOL extends
that of typed SIMPLE with object-oriented constructs.
The syntax below was produced by copying and modifying/extending the
syntax of dynamically typed SIMPLE. In fact, the only change we made
to the existing syntax of dynamically typed SIMPLE was to change the
strictness of the application construct like in untyped KOOL, from
strict to strict(2) (because application is not
strict in the first argument anymore due to dynamic method dispatch).
The KOOL-specific syntactic extensions are identical to those in
untyped KOOL.

  syntax Id ::= "Object" [token] | "Main" [token]

Types

  syntax Type ::= "void" | "int" | "bool" | "string"
                | Id                              // KOOL class
                | Type "[" "]"
                | "(" Type ")"           [bracket]
                > Types "->" Type
  // TODO(KORE): drop klabel once issues #1913 are fixed
  syntax Types ::= List{Type,","}   [symbol(_,_::Types)]
  /*
  syntax Types ::= List{Type,","}
  */

Declarations

  syntax Param ::= Type Id
  syntax Params ::= List{Param,","}

  syntax Stmt ::= Type Exps ";" [avoid]
                | Type Id "(" Params ")" Block    // stays like in typed SIMPLE
                | "class" Id Block                // KOOL
                | "class" Id "extends" Id Block   // KOOL

Expressions

  syntax Exp ::= Int | Bool | String | Id
               | "this"                                 // KOOL
               | "super"                                // KOOL
               | "(" Exp ")"             [bracket]
               | "++" Exp
               | Exp "instanceOf" Id     [strict(1)]    // KOOL
               | "(" Id ")" Exp          [strict(2)]    // KOOL  cast
               | "new" Id "(" Exps ")"   [strict(2)]    // KOOL
               | Exp "." Id                             // KOOL
               > Exp "[" Exps "]"        [strict]
               > Exp "(" Exps ")"        [strict(2)]    // was strict in SIMPLE
               | "-" Exp                 [strict]
               | "sizeOf" "(" Exp ")"    [strict]
               | "read" "(" ")"
               > left:
                 Exp "*" Exp             [strict, left]
               | Exp "/" Exp             [strict, left]
               | Exp "%" Exp             [strict, left]
               > left:
                 Exp "+" Exp             [strict, left]
               | Exp "-" Exp             [strict, left]
               > non-assoc:
                 Exp "<" Exp             [strict, non-assoc]
               | Exp "<=" Exp            [strict, non-assoc]
               | Exp ">" Exp             [strict, non-assoc]
               | Exp ">=" Exp            [strict, non-assoc]
               | Exp "==" Exp            [strict, non-assoc]
               | Exp "!=" Exp            [strict, non-assoc]
               > "!" Exp                 [strict]
               > left:
                 Exp "&&" Exp            [strict(1), left]
               | Exp "||" Exp            [strict(1), left]
               > "spawn" Block
               > Exp "=" Exp             [strict(2), right]

  syntax Exps ::= List{Exp,","}          [strict, overload(exps)]
  syntax Val
  syntax Vals ::= List{Val,","}          [overload(exps)]

Statements

  syntax Block ::= "{" "}"
                | "{" Stmt "}"

  syntax Stmt ::= Block
                | Exp ";"                               [strict]
                | "if" "(" Exp ")" Block "else" Block   [avoid, strict(1)]
                | "if" "(" Exp ")" Block                [macro]
                | "while" "(" Exp ")" Block
                | "for" "(" Stmt Exp ";" Exp ")" Block  [macro]
                | "print" "(" Exps ")" ";"              [strict]
                | "return" Exp ";"                      [strict]
                | "return" ";"
                | "try" Block "catch" "(" Param ")" Block
                | "throw" Exp ";"                       [strict]
                | "join" Exp ";"                        [strict]
                | "acquire" Exp ";"                     [strict]
                | "release" Exp ";"                     [strict]
                | "rendezvous" Exp ";"                  [strict]

  syntax Stmt ::= Stmt Stmt                          [right]

Desugaring macros

  rule if (E) S => if (E) S else {}
  rule for(Start Cond; Step) {S::Stmt} => {Start while(Cond){S Step;}}
  rule T::Type E1::Exp, E2::Exp, Es::Exps; => T E1; T E2, Es;           [anywhere]
  rule T::Type X::Id = E; => T X; X = E;                                [anywhere]

  rule class C:Id S => class C extends Object S                     // KOOL

endmodule

Semantics

We first discuss the new configuration, then we include the semantics of
the constructs borrowed from SIMPLE which stay unchanged, then those
whose semantics had to change, and finally the semantics of the
KOOL-specific constructs.

module KOOL-TYPED-DYNAMIC
  imports KOOL-TYPED-DYNAMIC-SYNTAX
  imports DOMAINS

Configuration

The configuration of dynamically typed KOOL is almost identical to
that of its untyped variant. The only difference is the cell
return, inside the control cell, whose role is to
hold the expected return type of the invoked method. That is because
we want to dynamically check that the value that a method returns has
the expected type.

  // the syntax declarations below are required because the sorts are
  // referenced directly by a production and, because of the way KIL to KORE
  // is implemented, the configuration syntax is not available yet
  // should simply work once KIL is removed completely
  // check other definitions for this hack as well
  syntax EnvCell
  syntax ControlCellFragment
  syntax EnvStackCell
  syntax CrntObjCellFragment

  configuration <T color="red">
                  <threads color="orange">
                    <thread multiplicity="*" type="Set" color="yellow">
                      <k color="green"> ($PGM:Stmt ~> execute) </k>
                    //<br/> // TODO(KORE): support latex annotations #1799
                      <control color="cyan">
                        <fstack color="blue"> .List </fstack>
                        <xstack color="purple"> .List </xstack>
                        <returnType color="LimeGreen"> void </returnType>  // KOOL
                      //<br/> // TODO(KORE): support latex annotations #1799
                        <crntObj color="Fuchsia">  // KOOL
                           <crntClass> Object </crntClass>
                           <envStack> .List </envStack>
                           <location multiplicity="?"> .K </location>
                        </crntObj>
                      </control>
                    //<br/> // TODO(KORE): support latex annotations #1799
                      <env color="violet"> .Map </env>
                      <holds color="black"> .Map </holds>
                      <id color="pink"> 0 </id>
                    </thread>
                  </threads>
                //<br/> // TODO(KORE): support latex annotations #1799
                  <store color="white"> .Map </store>
                  <busy color="cyan">.Set </busy>
                  <terminated color="red"> .Set </terminated>
                  <input color="magenta" stream="stdin"> .List </input>
                  <output color="brown" stream="stdout"> .List </output>
                  <nextLoc color="gray"> 0 </nextLoc>
                //<br/> // TODO(KORE): support latex annotations #1799
                  <classes color="Fuchsia">        // KOOL
                     <classData multiplicity="*" type="Map" color="Fuchsia">
                        <className color="Fuchsia"> Main </className>
                        <baseClass color="Fuchsia"> Object </baseClass>
                        <declarations color="Fuchsia"> .K </declarations>
                     </classData>
                  </classes>
                </T>

Unchanged semantics from dynamically typed SIMPLE

The semantics below is taken over from dynamically typed SIMPLE
unchanged. Like for untyped KOOL, the semantics of function/method
declaration and invocation, and of program initialization needs to
change. Moreover, due to subtyping, the semantics of several imported
SIMPLE constructs can be made more general, such as that of the
return statement, that of the assignment, and that of the exceptions.
We removed all these from the imported semantics of SIMPLE below and
gave their modified semantics right after, together with the extended
semantics of thread spawning (which is identical to that of untyped
KOOL).

  syntax Val ::= Int | Bool | String
               | array(Type,Int,Int)
  syntax Exp ::= Val
  syntax Exps ::= Vals
  syntax KResult ::= Val
  syntax KResult ::= Vals


  syntax KItem ::= undefined(Type)

  rule <k> T:Type X:Id; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> undefined(T) ...</store>
       <nextLoc> L:Int => L +Int 1 </nextLoc>


  rule <k> T:Type X:Id[N:Int]; => .K ...</k>
       <env> Env => Env[X <- L] </env>
       <store>... .Map => L |-> array(T, L +Int 1, N)
                          (L +Int 1)...(L +Int N) |-> undefined(T) ...</store>
       <nextLoc> L:Int => L +Int 1 +Int N </nextLoc>
    requires N >=Int 0

  context _:Type _::Exp[HOLE::Exps];


  syntax Id ::= "$1" [token] | "$2" [token]
  rule T:Type X:Id[N1:Int, N2:Int, Vs:Vals];
    => T[]<Vs> X[N1];
       {
         T[][]<Vs> $1=X;
         for(int $2=0; $2 <= N1 - 1; ++$2) {
           T X[N2,Vs];
           $1[$2] = X;
         }
       }


  rule <k> X:Id => V ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V:Val ...</store>


  context ++(HOLE => lvalue(HOLE))
  rule <k> ++loc(L) => I +Int 1 ...</k>
       <store>... L |-> (I:Int => I +Int 1) ...</store>


  rule I1 + I2 => I1 +Int I2
  rule Str1 + Str2 => Str1 +String Str2
  rule I1 - I2 => I1 -Int I2
  rule I1 * I2 => I1 *Int I2
  rule I1 / I2 => I1 /Int I2 requires I2 =/=K 0
  rule I1 % I2 => I1 %Int I2 requires I2 =/=K 0
  rule - I => 0 -Int I
  rule I1 < I2 => I1 <Int I2
  rule I1 <= I2 => I1 <=Int I2
  rule I1 > I2 => I1 >Int I2
  rule I1 >= I2 => I1 >=Int I2
  rule V1:Val == V2:Val => V1 ==K V2
  rule V1:Val != V2:Val => V1 =/=K V2
  rule ! T => notBool(T)
  rule true  && E => E
  rule false && _ => false
  rule true  || _ => true
  rule false || E => E


  rule V:Val[N1:Int, N2:Int, Vs:Vals] => V[N1][N2, Vs]
    [anywhere]

  rule array(_:Type, L:Int, M:Int)[N:Int] => lookup(L +Int N)
    requires N >=Int 0 andBool N <Int M  [anywhere]

  rule sizeOf(array(_,_,N)) => N


  syntax Val ::= nothing(Type)
  rule <k> return; => return nothing(T); ...</k> <returnType> T </returnType>


  rule <k> read() => I ...</k> <input> ListItem(I:Int) => .List ...</input>


  context (HOLE => lvalue(HOLE)) = _


  rule {} => .K
  rule <k> { S } => S ~> setEnv(Env) ...</k>  <env> Env </env>


  rule S1:Stmt S2:Stmt => S1 ~> S2


  rule _:Val; => .K


  rule if ( true) S else _ => S
  rule if (false) _ else S => S


  rule while (E) S => if (E) {S while(E)S}


  rule <k> print(V:Val, Es => Es); ...</k> <output>... .List => ListItem(V) </output>
    requires typeOf(V) ==K int orBool typeOf(V) ==K string
  rule print(.Vals); => .K


  rule (<thread>... <k>.K</k> <holds>H</holds> <id>T</id> ...</thread> => .Bag)
       <busy> Busy => Busy -Set keys(H) </busy>
       <terminated>... .Set => SetItem(T) ...</terminated>

  rule <k> join T:Int; => .K ...</k>
       <terminated>... SetItem(T) ...</terminated>

  rule <k> acquire V:Val; => .K ...</k>
       <holds>... .Map => V |-> 0 ...</holds>
       <busy> Busy (.Set => SetItem(V)) </busy>
    requires (notBool(V in Busy:Set))

  rule <k> acquire V; => .K ...</k>
       <holds>... V:Val |-> (N:Int => N +Int 1) ...</holds>

  rule <k> release V:Val; => .K ...</k>
       <holds>... V |-> (N => N:Int -Int 1) ...</holds>
    requires N >Int 0

  rule <k> release V; => .K ...</k> <holds>... V:Val |-> 0 => .Map ...</holds>
       <busy>... SetItem(V) => .Set ...</busy>

  rule <k> rendezvous V:Val; => .K ...</k>
       <k> rendezvous V; => .K ...</k>

Unchanged auxiliary operations from dynamically typed SIMPLE

  syntax Stmt ::= mkDecls(Params,Vals)  [function]
  rule mkDecls((T:Type X:Id, Ps:Params), (V:Val, Vs:Vals))
    => T X=V; mkDecls(Ps,Vs)
  rule mkDecls(.Params,.Vals) => {}

  syntax Exp ::= lookup(Int)
  rule <k> lookup(L) => V ...</k> <store>... L |-> V:Val ...</store>

  syntax KItem ::= setEnv(Map)
  rule <k> setEnv(Env) => .K ...</k>  <env> _ => Env </env>
  rule (setEnv(_) => .K) ~> setEnv(_)

  syntax Exp ::= lvalue(K)
  syntax Val ::= loc(Int)
  rule <k> lvalue(X:Id => loc(L)) ...</k>  <env>... X |-> L:Int ...</env>

  context lvalue(_::Exp[HOLE::Exps])
  context lvalue(HOLE::Exp[_::Exps])

  rule lvalue(lookup(L:Int) => loc(L))

  syntax Type ::= Type "<" Vals ">"  [function]
  rule T:Type<_,Vs:Vals> => T[]<Vs>
  rule T:Type<.Vals> => T

  syntax Map ::= Int "..." Int "|->" K [function]
  rule N...M |-> _ => .Map  requires N >Int M
  rule N...M |-> K => N |-> K (N +Int 1)...M |-> K  requires N <=Int M

  syntax Type ::= typeOf(K)  [function]
  rule typeOf(_:Int) => int
  rule typeOf(_:Bool) => bool
  rule typeOf(_:String) => string
  rule typeOf(array(T,_,_)) => (T[])
  rule typeOf(undefined(T)) => T
  rule typeOf(nothing(T)) => T

  syntax Types ::= getTypes(Params)  [function]
  rule getTypes(T:Type _:Id) => T, .Types
  rule getTypes(T:Type _:Id, P, Ps) => T, getTypes(P,Ps)
  rule getTypes(.Params) => void, .Types

Changes to the existing dynamically typed SIMPLE semantics

We extend/change the semantics of several SIMPLE constructs in order
to take advantage of the richer KOOL semantic infrastructure and thus
get more from the existing SIMPLE constructs.

Program initialization

Like in untyped KOOL.

  syntax KItem ::= "execute"
  rule <k> execute => new Main(.Exps); </k> <env> .Map </env>

Method application

The only change to untyped KOOL's values is that method closures are
now typed (their first argument holds their type):

 syntax Val ::= objectClosure(Id,List)
              | methodClosure(Type,Id,Int,Params,Stmt)

The type held by a method clossure will be the entire type of the
method, not only its result type like the lambda-closure of typed
SIMPLE. The reason for this change comes from the the need to
dynamically upcast values when passed to contexts where values of
superclass types are expected; since we want method closures to be
first-class-citizen values in our language, we have to be able to
dynamically upcast them, and in order to do that elegantly it is
convenient to store the entire ``current type'' of the method closure
instead of just its result type. Note that this was unnecessary in
the semantics of the dynamically typed SIMPLE language.

Method closure application needs to also set a new return type in
the return cell, like in dynamically typed SIMPLE, in order
for the values returned by its body to be checked against the return
type of the method. To do this correctly, we also need to stack the
current status of the return cell and then pop it when the
method returns. We have to do the same with the current object
environment, so we group them together in the stack frame.

  syntax KItem ::= fstackFrame(Map, K, List, Type, K)

  rule <k> methodClosure(_->T,Class,OL,Ps,S)(Vs:Vals) ~> K
           => mkDecls(Ps,Vs) S return; </k>
       <env> Env => .Map </env>
       <store>... OL |-> objectClosure(_, EStack)...</store>
     //<br/> // TODO(KORE): support latex annotations #1799
       <control>
          <fstack> .List => ListItem(fstackFrame(Env, K, XS, T', <crntObj> Obj' </crntObj>)) ...</fstack>
          <xstack> XS </xstack>
          <returnType> T' => T </returnType>
          <crntObj> Obj' => <crntClass> Class </crntClass> <envStack> EStack </envStack> </crntObj>
       </control>

At method return, we have to check that the type of the returned
value is a subtype of the expected return type. Moreover, if that is
the case, then we also upcast the returned value to one of the
expected type. The computation item unsafeCast(V,T) changes
the typeof V to T without any additional checks; however, it only
does it when V is an object or a method, otherwise it returns V
unchanged.

  rule <k> return V:Val; ~> _
           => subtype(typeOf(V), T) ~> true? ~> unsafeCast(V, T) ~> K
       </k>
       <control>
         <fstack> ListItem(fstackFrame(Env, K, XS, RT, <crntObj> CO </crntObj>)) => .List ...</fstack>
         <xstack> _ => XS </xstack>
         <returnType> T:Type => RT </returnType>
         <crntObj> _ => CO </crntObj>
       </control>
       <env> _ => Env </env>

Assignment

Typed KOOL allows to assign subtype instance values to supertype
lvalues. The semantics of assignment below is similar in spirit to
dynamically typed SIMPLE's, but a check is performed that the assigned
value's type is a subtype of the location's type. If that is the
case, then the assigned value is returned as a result and stored, but
it is upcast appropriately first, so the context will continue to see
a value of the expected type of the location. Note that the type of a
location is implicit in the type of its contents and it never changes
during the execution of a program; its type is assigned when the
location is allocated and initialized, and then only type-preserving
values are allowed to be stored in each location.

  rule <k> loc(L) = V:Val
           => subtype(typeOf(V),typeOf(V')) ~> true?
              ~> unsafeCast(V, typeOf(V')) ...</k>
       <store>... L |-> (V' => unsafeCast(V, typeOf(V'))) ...</store>

Typed exceptions

Exceptions are propagated now until a catch that can handle them is
encountered.

  syntax KItem ::= xstackFrame(Param, Stmt, K, Map, K)
  syntax KItem ::= "popx"

  rule <k> (try S1 catch(P) S2 => S1 ~> popx) ~> K </k>
       <control>
         <xstack> .List => ListItem(xstackFrame(P, S2, K, Env, C)) ...</xstack>
         C
       </control>
       <env> Env </env>

  rule <k> popx => .K ...</k>
       <xstack> ListItem(_) => .List ...</xstack>

  rule <k> throw V:Val; ~> _
        => if (subtype(typeOf(V),T)) { T X = V; S2 } else { throw V; } ~> K
       </k>
       <control>
         <xstack> ListItem(xstackFrame(T:Type X:Id, S2, K, Env, C)) => .List ...</xstack>
         (_ => C)
       </control>
       <env> _ => Env </env>

Spawn

Like in untyped KOOL.

  rule <thread>...
         <k> spawn S => !T:Int ...</k>
         <env> Env </env>
         <crntObj> Obj </crntObj>
       ...</thread>
       (.Bag => <thread>...
               <k> S </k>
               <env> Env </env>
               <id> !T </id>
               <crntObj> Obj </crntObj>
             ...</thread>)

Semantics of the new KOOL constructs

Class declaration

Like in untyped KOOL.

  rule <k> class Class1 extends Class2 { S } => .K ...</k>
       <classes>... (.Bag => <classData>
                            <className> Class1 </className>
                            <baseClass> Class2 </baseClass>
                            <declarations> S </declarations>
                        </classData>)
       ...</classes>

Method declaration

Methods are now typed and we need to store their types in their
closures, so that their type contract can be checked at invocation
time. The rule below is conceptually similar to that of untyped KOOL;
the only difference is the addition of the types.

  rule <k> T:Type F:Id(Ps:Params) S => .K ...</k>
       <crntClass> C </crntClass>
       <location> OL </location>
       <env> Env => Env[F <- L] </env>
       <store>... .Map => L|->methodClosure(getTypes(Ps)->T,C,OL,Ps,S) ...</store>
       <nextLoc> L => L +Int 1 </nextLoc>

New

The semantics of new in dynamically typed KOOL is also
similar to that in untyped KOOL, the main difference being the
management of the return types. Indeed, when a new object is created
we also have to stack the current type in the return cell in
order to be recovered after the creation of the new object. Only the
first rule below needs to be changed; the others are identical to
those in untyped KOOL.

  syntax KItem ::= envStackFrame(Id, Map)

  rule <k> new Class:Id(Vs:Vals) ~> K
           => create(Class) ~> (storeObj ~> ((Class(Vs)); return this;)) </k>
       <env> Env => .Map </env>
       <nextLoc> L:Int => L +Int 1 </nextLoc>
     //<br/> // TODO(KORE): support latex annotations #1799
       <control>
         <xstack> XS </xstack>
         <crntObj> Obj
                   => <crntClass> Object </crntClass>
                      <envStack> ListItem(envStackFrame(Object, .Map)) </envStack>
                      <location> L </location>
         </crntObj>
         <returnType> T => Class </returnType>
         <fstack> .List => ListItem(fstackFrame(Env, K, XS, T, <crntObj>Obj</crntObj>)) ...</fstack>
       </control>

  syntax KItem ::= create(Id)

  rule <k> create(Class:Id)
           => create(Class1) ~> setCrntClass(Class) ~> S ~> addEnvLayer ...</k>
       <className> Class </className>
       <baseClass> Class1:Id </baseClass>
       <declarations> S </declarations>

  rule <k> create(Object) => .K ...</k>

  syntax KItem ::= setCrntClass(Id)

  rule <k> setCrntClass(C) => .K ...</k>
       <crntClass> _ => C </crntClass>

  syntax KItem ::= "addEnvLayer"

  rule <k> addEnvLayer => .K ...</k>
       <env> Env => .Map </env>
       <crntClass> Class:Id </crntClass>
       <envStack> .List => ListItem(envStackFrame(Class, Env)) ...</envStack>

  syntax KItem ::= "storeObj"

  rule <k> storeObj => .K ...</k>
       <crntObj>
         <crntClass> Class </crntClass>
         <envStack> EStack </envStack>
         (<location> L:Int </location> => .Bag)
       </crntObj>
       <store>... .Map => L |-> objectClosure(Class, EStack) ...</store>

Self reference

Like in untyped KOOL.

  rule <k> this => objectClosure(Class, EStack) ...</k>
       <crntObj>
         <crntClass> Class </crntClass>
         <envStack> EStack </envStack>
         ...
       </crntObj>

Object member access

Like in untyped KOOL.

  rule <k> X:Id => this . X ...</k> <env> Env:Map </env>
    requires notBool(X in keys(Env))

  context HOLE . _::Id requires (HOLE =/=K super)

/*  rule objectClosure(<crntObj> <crntClass> Class:Id </crntClass>
                     <envStack>... ListItem((Class,EnvC:EnvCell)) EStack </envStack> </crntObj>)
       . X:Id
    => lookupMember(<envStack> ListItem((Class,EnvC)) EStack </envStack>, X) */
  rule objectClosure(Class:Id,
                     ListItem(envStackFrame(Class,Env)) EStack)
       . X:Id
    => lookupMember(ListItem(envStackFrame(Class,Env)) EStack, X)
  rule objectClosure(Class:Id,
                     (ListItem(envStackFrame(Class':Id,_)) => .List) _EStack)
       . _X:Id
    requires Class =/=K Class'

/*  rule <k> super . X => lookupMember(<envStack>EStack</envStack>, X) ...</k>
       <crntClass> Class </crntClass>
       <envStack>... ListItem((Class,EnvC:EnvCell)) EStack </envStack> */
  rule <k> super . X => lookupMember(EStack, X) ...</k>
       <crntClass> Class:Id </crntClass>
       <envStack> ListItem(envStackFrame(Class,_)) EStack </envStack>
  rule <k> super . _X ...</k>
       <crntClass> Class:Id </crntClass>
       <envStack> (ListItem(envStackFrame(Class':Id,_)) => .List) _EStack </envStack>
    requires Class =/=K Class'

Method invocation

The method lookup is the same as in untyped KOOL.

  rule <k> (X:Id => V)(_:Exps) ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V:Val ...</store>

  rule <k> (X:Id => this . X)(_:Exps) ...</k>
       <env> Env </env>
    requires notBool(X in keys(Env))

  context HOLE._::Id(_) requires HOLE =/=K super

  rule (objectClosure(_, EStack) . X
    => lookupMember(EStack, X:Id))(_:Exps)

/*  rule <k> (super . X
            => lookupMember(<envStack>EStack</envStack>,X))(_:Exps)...</k>
       <crntClass> Class </crntClass>
       <envStack>... ListItem((Class,_)) EStack </envStack> */
  rule <k> (super . X
            => lookupMember(EStack,X))(_:Exps)...</k>
       <crntClass> Class:Id </crntClass>
       <envStack> ListItem(envStackFrame(Class,_)) EStack </envStack>
  rule <k> (super . _X)(_:Exps)...</k>
       <crntClass> Class:Id </crntClass>
       <envStack> (ListItem(envStackFrame(Class':Id,_)) => .List) _EStack </envStack>
    requires Class =/=K Class'

  // TODO(KORE): fix getKLabel #1801
  rule (A:Exp(B:Exps))(C:Exps) => A(B) ~> #freezerFunCall(C)
  rule (A:Exp[B:Exps])(C:Exps) => A[B] ~> #freezerFunCall(C)
  rule V:Val ~> #freezerFunCall(C:Exps) => V(C)
  syntax KItem ::= "#freezerFunCall" "(" K ")"
  /*
  context HOLE(_:Exps)
    requires getKLabel HOLE ==KLabel '_`(_`) orBool getKLabel HOLE ==KLabel '_`[_`]
  */

  rule <k> (lookup(L) => V)(_:Exps) ...</k>  <store>... L |-> V:Val ...</store>

Instance of

Like in untyped KOOL.

  rule objectClosure(_, ListItem(envStackFrame(C,_)) _)
       instanceOf C => true

  rule objectClosure(_, (ListItem(envStackFrame(C::Id,_)) => .List) _)
       instanceOf C'  requires C =/=K C'

  rule objectClosure(_, .List) instanceOf _ => false

Cast

Unlike in untyped KOOL, in typed KOOL we actually check that the object
can indeed be cast to the claimed type.

  rule (C:Id) objectClosure(Irrelevant, EStack)
    => objectClosure(Irrelevant, EStack) instanceOf C ~> true?
       ~> objectClosure(C, EStack)

KOOL-specific auxiliary declarations and operations

Objects as lvalues

Like in untyped KOOL.

  rule <k> lvalue(X:Id => this . X) ...</k>  <env> Env </env>
    requires notBool(X in keys(Env))

  context lvalue((HOLE . _)::Exp)

/*  rule lvalue(objectClosure(<crntObj> <crntClass> C </crntClass>
                            <envStack>... ListItem((C,EnvC:EnvCell)) EStack </envStack> </crntObj>)
              . X
              => lookupMember(<envStack> ListItem((C,EnvC)) EStack </envStack>,
                              X)) */
  rule lvalue(objectClosure(C:Id,
                            ListItem(envStackFrame(C,Env)) EStack)
              . X
              => lookupMember(ListItem(envStackFrame(C,Env)) EStack,
                              X))
  rule lvalue(objectClosure(C,
                            (ListItem(envStackFrame(C',_)) => .List) _EStack)
              . _X)
    requires C =/=K C'

Lookup member

Like in untyped KOOL.

  syntax Exp ::= lookupMember(List,Id)  [function]

  rule lookupMember(ListItem(envStackFrame(_, X |-> L _)) _, X) => lookup(L)

  // TODO: fix rule below as shown once we support functions with deep rewrites
  // rule lookupMember(<envStack> ListItem((_, <env> Env </env>)) => .List
  //                     ...</envStack>, X)
  //   requires notBool(X in keys(Env))
  rule lookupMember(ListItem(envStackFrame(_, Env)) L, X)
    => lookupMember(L, X)
    requires notBool(X in keys(Env))

typeOf for the additional values}

  rule typeOf(objectClosure(C,_)) => C
  rule typeOf(methodClosure(T:Type,_,_,_Ps:Params,_)) => T

Subtype checking

The subclass relation induces a subtyping relation.

  syntax Exp ::= subtype(Types,Types)

  rule subtype(T:Type, T) => true

  rule <k> subtype(C1:Id, C:Id) => subtype(C2, C) ...</k>
       <className> C1 </className>
       <baseClass> C2:Id </baseClass>
    requires C1 =/=K C

  rule subtype(Object,Class:Id) => false
    requires Class =/=K Object

  rule subtype(Ts1->T2,Ts1'->T2') => subtype(((T2)::Type,Ts1'),((T2')::Type,Ts1))

// Note that the following rule would be wrong!
//  rule subtype(T[],T'[]) => subtype(T,T')

  rule subtype((T:Type,Ts),(T':Type,Ts')) => subtype(T,T') && subtype(Ts,Ts')
    requires Ts =/=K .Types
  rule subtype(.Types,.Types) => true

Unsafe Casting

Performs unsafe casting. One should only use it in combination with
the subtype relation above.

  syntax Val ::= unsafeCast(Val,Type)  [function]

  rule unsafeCast(objectClosure(_,EStack), C:Id)
    => objectClosure(C,EStack)

  rule unsafeCast(methodClosure(_T',C,OL,Ps,S), T) => methodClosure(T,C,OL,Ps,S)

  rule unsafeCast(V:Val, T:Type) => V  requires typeOf(V) ==K T

Generic guard

A generic computational guard: it allows the computation to continue
only if a prefix guard evaluates to true.

  syntax KItem ::= "true?"
  rule true ~> true? => .K

endmodule

Go to Lesson 3, KOOL typed static.

KOOL — Typed — Static

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K static semantics of the typed KOOL language.
It extends the static semantics of typed SIMPLE with static semantics
for the object-oriented constructs. Also, the static semantics of
some of the existing SIMPLE constructs need to change, in order to
become more generous with regards to the set of accepted programs,
mostly due to subtyping. For example, the assignment construct
x = e required that both the variable x and the
expression e had the same type in SIMPLE. In KOOL, the type
of e can be a subtype of the type of x.
Specifically, we define the following typing policy for KOOL,
everything else not mentioned below borrowing its semantics from
SIMPLE:

  • Each class C yields a homonymous type, which can be
    explicitly used in programs to type variables and methods, possibly in
    combination with other types.

  • Since now we have user-defined types, we check that each type
    used in a KOOL program is well-formed, that is, it is constructed only
    from primitive and class types corresponding to declared classes.

  • Class members and their types form a class type
    environment
    . Each class will have such a type environment.
    Each member in a class is allowed to be declared only once. Since in
    KOOL we allow methods to be assigned to fields, we make no distinction
    between field and method members; in other words, we reject programs
    declaring both a field and a method with the same name.

  • If an identifier is not found in the local type environment, it
    will be searched for in the current class type environment. If not
    there, then it will be searched for in its superclass' type
    environment. And so on and so forth. If not found until the
    Object class is reached, a typing error is reported.

  • The assignment allows variables to be assigned values of
    more concrete types. The result type of the assignment expression
    construct will be the (more abstract) type of the assigned variable,
    and not the (more concrete) type of the expression, like in Java.

  • Exceptions are changed (from SIMPLE) to allow throwing and
    catching only objects, like in Java. Also, unlike in SIMPLE, we do
    not check whether the type of the thrown exception matches the type of
    the caught variable, because exceptions can be caught by other
    try/catch blocks, even by ones in other methods. To avoid
    having to annotate each method with what exceptions it can throw, we
    prefer to not check the type safety of exceptions (although this is an
    excellent homework!). We only check that the try block
    type-checks and that the catch block type-checks after we bind
    the caught variable to its claimed type.

  • Class declarations are not allowed to have any cycles in their
    extends relation. Such cycles would lead to non-termination of
    new, as it actually does in the dynamic semantics of KOOL
    where no such circularity checks are performed.

  • Methods overriding other methods should be in the right subtyping
    relationship with the overridden methods: co-variant in the codomain
    and contra-variant in the domain.

module KOOL-TYPED-STATIC-SYNTAX
  imports DOMAINS-SYNTAX

Syntax

The syntax of statically typed KOOL is identical to that of
dynamically typed KOOL, they both taking as input the same programs.
What differs is the K strictness attributes. Like in statically
typed SIMPLE, almost all language constructs are strict now, since we
want each to type its arguments almost all the time. Like in the
other two KOOL definitions, we prefer to copy and then modify/extend
the syntax of statically typed SIMPLE.

Note: This paragraph is old, now we can do things better. We keep
it here only for historical reasons, to see how much we used to suffer 😃

Annoying K-tool technical problem:
Currently, the K tool treats the "non-terminal" productions (i.e.,
productions consisting of just one non-terminal), also called
"subsorting" production, differently from the other productions.
Specifically, it does not insert a node in the AST for them. This may
look desirable at first, but it has a big problem: it does not allow
us to treat the subsort differently in different context. For
example, since we want Id to be both a type (a class name) and a
program variable, and since we want expressions to reduce to their
types, we are in an impossible situations in which we do not know how
to treat an identifier in the semantics: as a type, i.e., a result of
computations, or as a program variable, i.e., a non-result. Ideally,
we would like to tag the identifiers at parse-time with their local
interpretation, but that, unfortunately, is not possible with the
current parsing capabilities of the K tool, because it requires to
insert additional information in the AST for the subsort productions.
This will be fixed soon. Until then, unfortunately, we have to do the
job of the parser manually. Instead of subsorting Id directly
to Type, we "wrap" it first, say with a wrapper called
class(...), exactly how the parser should have done.
The major drawback of this is that all the typed KOOL programs
in kool/typed/programs need to also be modified to always
declare class types accordingly. The modified programs can be found
in kool/typed/static/programs. So make sure you execute the
static semantics of KOOL using the modified programs. To avoid seeing
the wrapper in the generated documentation, we associate it an
"invisibility" latex attribute below.

  syntax Id ::= "Object" [token] | "Main" [token]

Types

  syntax Type ::= "void" | "int" | "bool" | "string"
                | Id                     [klabel("class"), symbol, avoid]  // see next
                | Type "[" "]"
                | "(" Type ")"           [bracket]
                > Types "->" Type

  syntax Types ::= List{Type,","}        [overload(exps)]

Declarations

  syntax Param ::= Type Id
  syntax Params ::= List{Param,","}

  syntax Stmt ::= Type Exps ";" [avoid]
                | Type Id "(" Params ")" Block
                | "class" Id Block
                | "class" Id "extends" Id Block

Expressions

  syntax FieldReference ::= Exp "." Id          [strict(1)]
  syntax ArrayReference ::= Exp "[" Exps "]"    [strict]

  syntax Exp ::= Int | Bool | String | Id
               | "this"
               | "super"
               | "(" Exp ")"             [bracket]
               | "++" Exp
               | Exp "instanceOf" Id     [strict(1)]
               | "(" Id ")" Exp          [strict(2)]
               | "new" Id "(" Exps ")"   [strict(2)]
               > Exp "(" Exps ")"        [strict]
               | "-" Exp                 [strict]
               | "sizeOf" "(" Exp ")"    [strict]
               | "read" "(" ")"
               > left:
                 Exp "*" Exp             [strict, left]
               | Exp "/" Exp             [strict, left]
               | Exp "%" Exp             [strict, left]
               > left:
                 Exp "+" Exp             [strict, left]
               | Exp "-" Exp             [strict, left]
               > non-assoc:
                 Exp "<" Exp             [strict, non-assoc]
               | Exp "<=" Exp            [strict, non-assoc]
               | Exp ">" Exp             [strict, non-assoc]
               | Exp ">=" Exp            [strict, non-assoc]
               | Exp "==" Exp            [strict, non-assoc]
               | Exp "!=" Exp            [strict, non-assoc]
               > "!" Exp                 [strict]
               > left:
                 Exp "&&" Exp            [strict, left]
               | Exp "||" Exp            [strict, left]
               > "spawn" Block  // not strict: to check return and exceptions
               > Exp "=" Exp             [strict(2), right]

  syntax Exp ::= FieldReference | ArrayReference
  syntax priority _.__KOOL-TYPED-STATIC-SYNTAX > _[_]_KOOL-TYPED-STATIC-SYNTAX > _(_)_KOOL-TYPED-STATIC-SYNTAX

  syntax Exps ::= List{Exp,","}          [strict, overload(exps)]

Statements

  syntax Block ::= "{" "}"
                | "{" Stmt "}"

  syntax Stmt ::= Block
                | Exp ";"                                 [strict]
                | "if" "(" Exp ")" Block "else" Block     [avoid, strict]
                | "if" "(" Exp ")" Block                  [macro]
                | "while" "(" Exp ")" Block               [strict]
                | "for" "(" Stmt Exp ";" Exp ")" Block    [macro]
                | "return" Exp ";"                        [strict]
                | "return" ";"
                | "print" "(" Exps ")" ";"                [strict]
                | "try" Block "catch" "(" Param ")" Block [strict(1)]
                | "throw" Exp ";"                         [strict]
                | "join" Exp ";"                          [strict]
                | "acquire" Exp ";"                       [strict]
                | "release" Exp ";"                       [strict]
                | "rendezvous" Exp ";"                    [strict]

  syntax Stmt ::= Stmt Stmt                            [seqstrict, right]

Desugaring macros

  rule if (E) S => if (E) S else {}
  rule for(Start Cond; Step) {S:Stmt} => {Start while(Cond){S Step;}}
  rule T:Type E1:Exp, E2:Exp, Es:Exps; => T E1; T E2, Es;               [anywhere]
  rule T:Type X:Id = E; => T X; X = E;                                  [anywhere]

  rule class C:Id S => class C extends Object S

endmodule

Static semantics

We first discuss the configuration, then give the static semantics
taken over unchanged from SIMPLE, then discuss the static semantics of
SIMPLE syntactic constructs that needs to change, and in the end we
discuss the static semantics and additional checks specifically
related to the KOOL proper syntax.

module KOOL-TYPED-STATIC
  imports KOOL-TYPED-STATIC-SYNTAX
  imports DOMAINS

Configuration

The configuration of our type system consists of a tasks
cell with the same meaning like in statically typed SIMPLE, of an
out cell streamed to the standard output that will be used to
display typing error messages, and of a cell classes holding
data about each class in a separate class cell. The
task cells now have two additional optional subcells, namely
ctenvT and inClass. The former holds a temporary
class type environment; its contents will be transferred into the
ctenv cell of the corresponding class as soon as all the
fields and methods in the task are processed. In fact, there will be
three types of tasks in the subsequent semantics, each determined by
the subset of cells that it holds:

  1. Main task, holding only a k cell holding the
    original program as a set of classes. The role of this task is to
    process each class, generating a class task (see next) for each.

  2. Class task, holding k, ctenvT, and
    inClass subcells. The role of this task type is to process
    a class' contents, generating a class type environment in the
    ctenvT cell and a method task (see next) for each method in
    the class. To avoid interference with object member lookup rules
    below, it is important to add the class type environment to a class
    atomically; this is the reason for which we use ctenvT
    temporary cells within class tasks (instead of adding each member
    incrementally to the class' type environment).

  3. Method task, holding k, tenv and
    return cells. These tasks are similar to SIMPLE's function
    tasks, so we do not discuss them here any further.

Each class cell hods its name (in the className
cell) and the name of the class it extends (in the extends
cell), as well as its type environment (in the ctenv cell)
and the set of all its superclasses (in the extendsAll cell).
The later is useful for example for checking whether there are cycles
in the class extends relation.

  configuration <T multiplicity="?" color="yellow">
                  <tasks color="orange" multiplicity="?">
                    <task multiplicity="*" color="yellow" type="Set">
                      <k color="green"> $PGM:Stmt </k>
                      <tenv multiplicity="?" color="cyan"> .Map </tenv>
                      <ctenvT multiplicity="?" color="blue"> .Map </ctenvT>
                      <returnType multiplicity="?" color="black"> void </returnType>
                      <inClass multiplicity="?" color="Fuchsia"> .K </inClass>
                    </task>
                  </tasks>
//                  <br/>
                  <classes color="Fuchsia">
                    <classData multiplicity="*" type="Map">
                      <className color="Fuchsia"> Object </className>
                      <baseClass color="Fuchsia"> .K </baseClass>
                      <baseClasses color="Fuchsia"> .Set </baseClasses>
                      <ctenv multiplicity="?" color="blue"> .Map </ctenv>
                    </classData>
                  </classes>
                </T>
                <output color="brown" stream="stdout"> .List </output>

Unchanged semantics from statically typed SIMPLE

The syntax and rules below are borrowed unchanged from statically
typed SIMPLE, so we do not discuss them much here.

  syntax Exp ::= Type
  syntax Exps ::= Types
  syntax BlockOrStmtType ::= "block" | "stmt"
  syntax Type ::= BlockOrStmtType
  syntax Block ::= BlockOrStmtType
  syntax KResult ::= Type
                   | Types  // TODO: should not be needed


  context _:Type _::Exp[HOLE::Exps];

  rule T:Type E:Exp[int,Ts:Types]; => T[] E[Ts];
  rule T:Type E:Exp[.Types]; => T E;


  rule <task>... <k> _:BlockOrStmtType </k> <tenv> _ </tenv> ...</task> => .Bag


  rule _:Int => int
  rule _:Bool => bool
  rule _:String => string


  rule <k> X:Id => T ...</k> <tenv>... X |-> T ...</tenv>


  context ++(HOLE => ltype(HOLE))
  rule ++ int => int
  rule int + int => int
  rule string + string => string
  rule int - int => int
  rule int * int => int
  rule int / int => int
  rule int % int => int
  rule - int => int
  rule int < int => bool
  rule int <= int => bool
  rule int > int => bool
  rule int >= int => bool
  rule T:Type == T => bool
  rule T:Type != T => bool
  rule bool && bool => bool
  rule bool || bool => bool
  rule ! bool => bool


  rule (T[])[int, Ts:Types] => T[Ts]
  rule T:Type[.Types] => T

  rule sizeOf(_T[]) => int


  rule read() => int

  rule print(T:Type, Ts => Ts); requires T ==K int orBool T ==K string
  rule print(.Types); => stmt


  context (HOLE => ltype(HOLE)) = _


  rule <k> return; => stmt ...</k> <returnType> _ </returnType>


  rule {} => block

  rule <task> <k> {S:Stmt} => block ...</k> <tenv> Rho </tenv> R </task>
       (.Bag => <task> <k> S </k> <tenv> Rho </tenv> R </task>)

  rule _:Type; => stmt
  rule if (bool) block else block => stmt
  rule while (bool) block => stmt

  rule join int; => stmt
  rule acquire _:Type; => stmt
  rule release _:Type; => stmt
  rule rendezvous _:Type; => stmt

  syntax Stmt ::= BlockOrStmtType
  rule _:BlockOrStmtType _:BlockOrStmtType => stmt

Unchanged auxiliary operations from dynamically typed SIMPLE

  syntax Stmt ::= mkDecls(Params)  [function]
  rule mkDecls(T:Type X:Id, Ps:Params) => T X; mkDecls(Ps)
  rule mkDecls(.Params) => {}

  syntax LValue ::= Id
                  | FieldReference
                  | ArrayReference
  syntax Exp ::= LValue

  syntax Exp ::= ltype(Exp)
// We would like to say:
//  context ltype(HOLE:LValue)
// but we currently cannot type the HOLE
  context ltype(HOLE) requires isLValue(HOLE)

// OLD approach:
//  syntax Exp ::= ltype(Exp)  [function]
//  rule ltype(X:Id) => X
//  rule ltype(E:Exp [Es:Exps]) => E[Es]

  syntax Types ::= getTypes(Params)  [function]
  rule getTypes(T:Type _:Id) => T, .Types
  rule getTypes(T:Type _:Id, P, Ps) => T, getTypes(P,Ps)
  rule getTypes(.Params) => void, .Types

Changes to the existing statically typed SIMPLE semantics

Below we give the new static semantics for language constructs that
come from SIMPLE, but whose SIMPLE static semantics was too
restrictive or too permissive and thus had to change.

Local variable declaration

Since we can define new types in KOOL (corresponding to classes), the
variable declaration needs to now check that the claimed types exist.
The operation checkType, defined at the end of this module,
checks whether the argument type is correct (it actually works with
lists of types as well).

  rule <k> T:Type X:Id; => checkType(T) ~> stmt ...</k>
       <tenv> Rho => Rho[X <- T] </tenv>

Class member declaration

In class tasks, variable declarations mean class member declarations.
Since we reduce method declarations to variable declarations (see
below), a variable declaration in a class task can mean either a field
or a method declaration. Unlike local variable declarations, which
can shadow previous homonymous local or member declarations, member
declarations are regarded as a set, so we disallow multiple
declarations for the same member (one could improve upon this, like in
Java, by treating members with different types or number of arguments
as different, etc., but we do not do it here). We also issue an error
message if one attempts to redeclare the same class member. The
framed variable declaration in the second rule below should be read
"stuck". In fact, it is nothing but a unary operation called
stuck, which takes a K-term as argument and does nothing
with it; this stuck operation is displayed as a frame in this
PDF document because of its latex attribute (see the ASCII .k file,
at the end of this module).

  rule <k> T:Type X:Id; => checkType(T) ~> stmt ...</k>
       <ctenvT> Rho (.Map => X |-> T) </ctenvT>
    requires notBool(X in keys(Rho))

  rule <k> T:Type X:Id; => stuck(T X;) ...</k>
       <ctenvT>... X |-> _ ...</ctenvT>
       <inClass> C:Id </inClass>
//       <br/>
       <output>... .List => ListItem("Member \"" +String Id2String(X)
                              +String "\" declared twice in class \""
                              +String Id2String(C) +String "\"!\n") </output>

Method declaration

A method declaration requires two conceptual checks to be performed:
first, that the method's type is consistent with the type of the
homonymous method that it overrides, if any; and second, that its body
types correctly. At the same time, it should also be added to the
type environment of its class. The first conceptual task is performed
using the checkMethod operation defined below, and the second
by generating a corresponding method task. To add it to the class
type environment, we take advantage of the fact that KOOL is higher
order and reduce the problem to a field declaration problem, which we
have already defined. The role of the ctenvT cell in the
rule below is to structurally ensure that the method declaration takes
place in a class task (we do not want to allow methods to be declared,
for example, inside other methods).

  rule <k> T:Type F:Id(Ps:Params) S
        => checkMethod(F, getTypes(Ps)->T, C')
           ~> getTypes(Ps)->T F; ...</k>
//       <br/>
       <inClass> C </inClass>
       <ctenvT> _ </ctenvT> // to ensure we are in a class pass
       <className> C </className>
       <baseClass> C' </baseClass>
//       <br/>
       (.Bag => <task>
               <k> mkDecls(Ps) S </k>
               <inClass> C </inClass>
               <tenv> .Map </tenv>
               <returnType> T </returnType>
             </task>)

Assignment

A more concrete value is allowed to be assigned to a more abstract
variable. The operation checkSubtype is defined at the end
of the module and it also works with pairs of lists of types.

  rule T:Type = T':Type => checkSubtype(T', T) ~> T

Method invocation and return

Methods can be applied on values of more concrete types than their
arguments:

  rule (Ts:Types -> T:Type) (Ts':Types) => checkSubtype(Ts',Ts) ~> T

Similarly, we allow values of more concrete types to be returned by
methods:

  rule <k> return T:Type; => checkSubtype(T,T') ~> stmt ...</k>
       <returnType> T':Type </returnType>

Exceptions

Exceptions can throw and catch values of any types. Since unlike in Java
KOOL's methods do not declare the exception types that they can throw,
we cannot test the full type safety of exceptions. Instead, we
only check that the try and the catch statements
type correctly.

  rule try block catch(T:Type X:Id) S => {T X; S}
  rule throw _T:Type ; => stmt

Spawn

The spawned cell needs to also be passed the parent's class.

// explain why

  rule <k> spawn S:Block => int ...</k>
       <tenv> Rho </tenv>
       <inClass> C </inClass>
       (.Bag => <task>
               <k> S </k>
               <tenv> Rho </tenv>
               <inClass> C </inClass>
             </task>)

Semantics of the new KOOL constructs

Class declaration

We process each class in the main task, adding the corresponding data
into its class cell and also adding a class task for it. We
also perform some well-formedness checks on the class hierarchy.

Initiate class processing
We create a class cell and a class task for each task. Also, we start
the class task with a check that the class it extends is declared
(this delays the task until that class is processed using another
instance of this rule).

// There seems to be some error with the configuration concretization,
// as the rule below does not work when rewriting . to both the task
// and the class cells; I had to include two separate . rewrites

// TODO: the following fails krun; see #2117
  rule <task> <k> class C:Id extends C':Id { S:Stmt } => stmt ...</k> </task>
       (.Bag => <classData>...
               <className> C </className>
               <baseClass> C' </baseClass>
             ...</classData>)
//       <br/>
       (.Bag => <task>
                <k> checkType(`class`(C')) ~> S </k>
                <inClass> C </inClass>
                <ctenvT> .Map </ctenvT>
             </task>)

// You may want to try the thing below, but that failed, too
/*
syntax Type ::= "stmtStop"

  rule <tasks>...
       <task> <k> class C:Id extends C':Id { S:Stmt } => stmtStop ...</k> </task>
       (.Bag => <task>
                <k> checkType(`class`(C')) ~> S </k>
                <inClass> C </inClass>
                <ctenvT> .Map </ctenvT>
             </task>)
       ...</tasks>
       <classes>...
       .Bag => <classData>...
               <className> C </className>
               <baseClass> C' </baseClass>
             ...</classData>
       ...</classes>
//       <br/>
*/

Check for unique class names

  rule (<T>...
          <className> C </className>
          <className> C </className>
        ...</T> => .Bag)
       <output>... .List => ListItem("Class \"" +String Id2String(C)
                                  +String "\" declared twice!\n") </output>

Check for cycles in class hierarchy
We check for cycles in the class hierarchy by transitively closing the
class extends relation using the extendsAll cells, and
checking that a class will never appear in its own extendsAll
cell. The first rule below initiates the transitive closure of the
superclass relation, the second transitively closes it, and the third
checks for cycles.

  rule <baseClass> C </baseClass>
       <baseClasses> .Set => SetItem(C) </baseClasses>  [priority(25)]

  rule <classData>...
         <baseClasses> SetItem(C) Cs:Set (.Set => SetItem(C')) </baseClasses>
       ...</classData>
       <classData>... <className>C</className> <baseClass>C'</baseClass> ...</classData>
    requires notBool(C' in (SetItem(C) Cs))  [priority(25)]

  rule (<T>...
          <className> C </className>
          <baseClasses>... SetItem(C) ...</baseClasses>
        ...</T> => .Bag)
       <output>... .List => ListItem("Class \"" +String Id2String(C)
                                  +String "\" is in a cycle!\n") </output>
    [priority(25)]

New

To type new we only need to check that the class constructor
can be called with arguments of the given types, so we initiate a call
to the constructor method in the corresponding class. If that
succeeds, meaning that it types to stmt, then we discard the
stmt type and produce instead the corresponding class type of
the new object. The auxiliary discard operation is defined
also at the end of this module.

  rule new C:Id(Ts:Types) => `class`(C) . C (Ts) ~> discard ~> `class`(C)

Self reference

The typing rule for this is straightforward: reduce to the
current class type.

  rule <k> this => `class`(C) ...</k>
       <inClass> C:Id </inClass>

Super

Similarly, super types to the parent class type.
Note that for typing concerns, super can be considered as an object
(recall that this was not the case in the dynamic semantics).

   rule <k> super => `class`(C') ...</k>
        <inClass> C:Id </inClass>
        <className> C </className>
        <baseClass> C':Id </baseClass>

Object member access

There are several cases to consider here. First, if we are in a class
task, we should lookup the member into the temporary class type
environemnt in cell ctenvT. That is because we want to allow
initialized field declarations in classes, such as int x=10;.
This is desugared to a declaration of x, which is added to
ctenvT during the class task processing, followed by an
assignment of x to 10. In order for the assignment to type
check, we need to know that x has been declared with type
int; this information can only be found in the
ctenvT cell. Second, we should redirect non-local variable
lookups in method tasks to corresponding member accesses (the
local variables are handled by the rule borrowed from SIMPLE).
This is what the second rule below does. Third, we should allow
object member accesses as lvalues, which is done by the third rule
below. These last two rules therefore ensure that each necessary
object member access is explicitly allowed for evaluation. Recall
from the annotated syntax module above that the member access
operation is strict in the object. That means that the object is
expected to evaluate to a class type. The next two rules below define
the actual member lookup operation, moving the search to the
superclass when the member is not found in the current class. Note
that this works because we create the class type environments
atomically; thus, a class either has its complete type environment
available, in which case these rules can safely apply, or its cell
ctenv is not yet available, in which case these rules have to
wait. Finally, the sixth rule below reports an error when the
Object class is reached.

  rule <k> X:Id => T ...</k>
       <ctenvT>... X |-> T ...</ctenvT>

  rule <k> X:Id => this . X ...</k>
       <tenv> Rho </tenv>
    requires notBool(X in keys(Rho))

// OLD approach:
//  rule ltype(E:Exp . X:Id) => E . X

  rule <k> `class`(C:Id) . X:Id => T ...</k>
       <className> C </className>
       <ctenv>... X |-> T:Type ...</ctenv>

  rule <k> `class`(C1:Id => C2) . X:Id ...</k>
       <className> C1 </className>
       <baseClass> C2:Id </baseClass>
       <ctenv> Rho </ctenv>
    requires notBool(X in keys(Rho))

  rule <k> `class`(Object) . X:Id => stuck(`class`(Object) . X) ...</k>
       <inClass> C:Id </inClass>
//      <br/>
       <output>... .List => ListItem("Member \"" +String Id2String(X)
                              +String "\" not declared! (see class \""
                              +String Id2String(C) +String "\")\n") </output>

Instance of and casting

As it is hard to check statically whether casting is always safe,
the programmer is simply trusted from a typing perspective. We only
do some basic upcasting and downcasting checks, to reject casts which
will absolutely fail. However, dynamic semantics or implementations
of the language need to insert runtime checks for downcasting to be safe.

  rule `class`(_C1:Id) instanceOf _C2:Id => bool
  rule (C:Id) `class`(C) => `class`(C)
  rule <k> (C2:Id) `class`(C1:Id) => `class`(C2) ...</k>
       <className> C1 </className>
       <baseClasses>...SetItem(C2)...</baseClasses>    // upcast
  rule <k> (C2:Id) `class`(C1:Id) => `class`(C2) ...</k>
       <className> C2 </className>
       <baseClasses>...SetItem(C1)...</baseClasses>    // downcast
  rule <k> (C2) `class`(C1:Id) => stuck((C2) `class`(C1)) ...</k>
       <classData>...
         <className> C1 </className>
         <baseClasses> S1 </baseClasses>
       ...</classData>
       <classData>...
         <className> C2 </className>
         <baseClasses> S2 </baseClasses>
       ...</classData>
       <output>... .List => ListItem("Classes \"" +String Id2String(C1)
                              +String "\" and \"" +String Id2String(C2)
                              +String "\" are incompatible!\n") </output>
    requires notBool(C1 in S2) andBool notBool(C2 in S1)

Cleanup tasks

Finally, we need to clean up the terminated tasks. Each of the three
types of tasks is handled differently. The main task is replaced by a
method task holding new main();, which will ensure that a
main class with a main() method actually exists
(first rule below). A class task moves its temporary class type
environment into its class' cell, and then it dissolves itself (second
rule). A method task simply dissolves when terminated (third rule);
the presence of the tenv cell in that rule ensures that that
task is a method task.
Finally, when all the tasks are cleaned up, we can also remove the
tasks cell, issuing a corresponding message. Note that
checking for cycles or duplicate methods can still be performed after
the tasks cell has been removed.

// discard main task when done, issuing a "new main();" command to
// make sure that the class main and the method main() are declared.

  rule <task> <k> stmt => new Main(.Exps); </k>
              (.Bag => <tenv> .Map </tenv>
                    <returnType> void </returnType>
                    <inClass> Main </inClass>)
       </task>

// discard class task when done, adding a ctenv in class

  rule (<task>
          <k> stmt </k>
          <ctenvT> Rho </ctenvT>
          <inClass> C:Id </inClass>
        </task> => .Bag)
        <className> C </className>
        (.Bag => <ctenv> Rho </ctenv>)

// discard method task when done

  rule <task>...
         <k> stmt </k>
         <tenv> _ </tenv>  // only to ensure that this is a method task
       ...</task> => .Bag

// cleanup tasks and output a success message when done

  rule (<T>... <tasks> .Bag </tasks> ...</T> => .Bag)
       <output>... .List => ListItem("Type checked!\n") </output>

KOOL-specific auxiliary declarations and operations

Subtype checking

The subclass relation introduces a subtyping relation.

  syntax KItem ::= checkSubtype(Types,Types)

  rule checkSubtype(T:Type, T) => .K

  rule <k> checkSubtype(`class`(C:Id), `class`(C':Id)) => .K ...</k>
       <className> C </className>
       <baseClasses>... SetItem(C') ...</baseClasses>

  rule checkSubtype(Ts1->T2,Ts1'->T2')
    => checkSubtype(((T2)::Type,Ts1'),((T2')::Type,Ts1))

// note that the following rule would be wrong!
//  rule checkSubtype(T[],T'[]) => checkSubtype(T,T')

  rule checkSubtype((T:Type,Ts),(T':Type,Ts'))
    => checkSubtype(T,T') ~> checkSubtype(Ts,Ts')
    requires Ts =/=K .Types

  rule checkSubtype(.Types,.Types) => .K
  rule checkSubtype(.Types,void) => .K

Checking well-formedness of types

Since now any Id can be used as the type of a class, we need to
check that the types used in the program actually exists

  syntax KItem ::= checkType(Types)

  rule checkType(T:Type,Ts:Types) => checkType(T) ~> checkType(Ts)
    requires Ts =/=K .Types
  rule checkType(.Types) => .K
  rule checkType(int) => .K
  rule checkType(bool) => .K
  rule checkType(string) => .K
  rule checkType(void) => .K
  rule <k> checkType(`class`(C:Id)) => .K ...</k> <className> C </className>
  rule checkType(`class`(Object)) => .K
  rule checkType(Ts:Types -> T:Type) => checkType(T,Ts)
  rule checkType(T:Type[]) => checkType(T)

Checking correct overiding of methods

The checkMethod operation below searches to see whether
the current method overrides some other method in some superclass.
If yes, then it issues an additional check that the new method's type
is more concrete than the overridden method's. The types T and T'
below can only be function types. See the definition of
checkSubtype on function types at the end of this module (it
is co-variant in the codomain and contra-variant in the domain).

  syntax KItem ::= checkMethod(Id,Type,Id)

  rule <k> checkMethod(F:Id, T:Type, C:Id) => checkSubtype(T, T') ...</k>
       <className> C </className>
       <ctenv>... F |-> T':Type ...</ctenv>

  rule <k> checkMethod(F:Id, _T:Type, (C:Id => C')) ...</k>
       <className> C </className>
       <baseClass> C':Id </baseClass>
       <ctenv> Rho </ctenv>
    requires notBool(F in keys(Rho))

  rule checkMethod(_:Id,_,Object) => .K

Generic operations which could be part of the K framework

  syntax KItem ::= stuck(K)

  syntax KItem ::= "discard"
  rule _:KResult ~> discard => .K

endmodule

FUN — Untyped — Environment

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K semantic definition of the untyped FUN language.
FUN is a pedagogical and research language that captures the essence
of the functional programming paradigm, extended with several features
often encountered in functional programming languages.
Like many functional languages, FUN is an expression language, that
is, everything, including the main program, is an expression.
Functions can be declared anywhere and are first class values in the
language.
FUN is call-by-value here, but it has been extended (as student
homework assignments) with other parameter-passing styles.
To make it more interesting and to highlight some of K's strengths,
FUN includes the following features:

  • The basic builtin data-types of integers, booleans and strings.

  • Builtin lists, which can hold any elements, including other lists.
    Lists are enclosed in square brackets and their elements are
    comma-separated; e.g., [1,2,3].

  • User-defined data-types, by means of constructor terms.
    Constructor names start with a capital letter (while any other
    identifier in the language starts with a lowercase letter), and they
    can be followed by an arbitrary number of comma-separated arguments
    enclosed in parentheses; parentheses are not needed when the
    constructor takes no arguments.
    For example, Pair(5,7) is a constructor term holding two
    numbers, Cons(1,Cons(2,Cons(3,Nil))) is a list-like
    constructor term holding 3 elements, and
    Tree(Tree(Leaf(1), Leaf(2)), Leaf(3)) is a tree-like
    constructor term holding 3 elements.
    In the untyped version of the FUN language, no type checking or
    inference is performed to ensure that the data constructors are used
    correctly.
    The execution will simply get stuck when they are misused.
    Moreover, since no type checking is performed, the data-types are not
    even declared in the untyped version of FUN.

  • Functions and let/letrec binders can take
    multiple space-separated arguments, but these are desugared to
    ones that only take one argument, by currying. For example, the
    expressions

    fun x y -> x y
    let x y = y in x
    

    are desugared, respectively, into the following expressions:

    fun x -> fun y -> x y
    let x = fun y -> y in x
    
  • Functions can be defined using pattern matching over the
    available data-types. For example, the program

    letrec max = fun [h] -> h
                 |   [h|t] -> let x = max t
                              in  if h > x then h else x
    in max [1, 3, 5, 2, 4, 0, -1, -5]
    

    defines a function max that calculates the maximum element of
    a non-empty list, and the function

    letrec ack = fun Pair(0,n) -> n + 1
                 |   Pair(m,0) -> ack Pair(m - 1, 1)
                 |   Pair(m,n) -> ack Pair(m - 1, ack Pair(m, n - 1))
    in ack Pair(2,3)
    

    calculates the Ackermann function applied to a particular pair of numbers.
    Patterns can be nested. Patterns can currently only be used in function
    definitions, and not directly in let/letrec binders.
    For example, this is not allowed:

    letrec Pai(x,y) = Pair(1,2) in x+y
    

    But this is allowed:

    let f Pair(x,y) = x+y in f Pair(1,2)
    

    because it is first reduced to

    let f = fun Pair(x,y) -> x+y in f Pair(1,2)
    

    by uncurrying of the let binder, and pattern matching is
    allowed in function arguments.

  • We include a callcc construct, for two reasons: first,
    several functional languages support this construct; second, some
    semantic frameworks have difficulties defining it. Not K.

  • Finally, we include mutables by means of referencing an
    expression, getting the reference of a variable, dereferencing and
    assignment. We include these for the same reasons as above: there are
    languages which have them, and they are not easy to define in some
    semantic frameworks.

Like in many other languages, some of FUN's constructs can be
desugared into a smaller set of basic constructs. We do that as usual,
using macros, and then we only give semantics to the core constructs.

Note:
We recommend the reader to first consult the dynamic semantics of the
LAMBDA++ language in the first part of the K Tutorial.
To keep the comments below small and focused, we will not re-explain
functional or K features that have already been explained in there.

Syntax

//require "modules/pattern-matching.k"

module FUN-UNTYPED-COMMON
  imports DOMAINS-SYNTAX

FUN is an expression language. The constructs below fall into
several categories: names, arithmetic constructs, conventional
functional constructs, patterns and pattern matching, data constructs,
lists, references, and call-with-current-continuation (callcc).
The arithmetic constructs are standard; they are present in almost all
our K language definitions. The meaning of FUN's constructs are
discussed in more depth when we define their semantics in the next
module.

The Syntactic Constructs

We start with the syntactic definition of FUN names.
We have several categories of names: ones to be used for functions and
variables, others to be used for data constructors, others for types and
others for type variables. We will introduce them as needed, starting
with the former category. We prefer the names of variables and functions
to start with lower case letters. We take the freedom to tacitly introduce
syntactic lists/sequences for each nonterminal for which we need them:

  syntax Name                                      [token]
  syntax Names ::= List{Name,","}                  [overload(exps)]

Expression constructs will be defined throughtout the syntax module.
Below are the very basic ones, namely the builtins, the names, and the
parentheses used as brackets for grouping. Lists of expressions are
declared strict, so all expressions in the list get evaluated whenever
the list is on a position which can be evaluated:

  syntax Exp ::= Int | Bool | String | Name
               | "(" Exp ")"                       [bracket]
  syntax Exps  ::= List{Exp,","}                   [strict, overload(exps)]
  syntax Val
  syntax Exp ::= Val
  syntax Exps ::= Vals
  syntax Vals ::= List{Val,","}                    [overload(exps)]
  syntax Bottom
  syntax Bottoms ::= List{Bottom,","}              [overload(exps)]

We next define the syntax of arithmetic constructs, together with
their relative priorities and left-/non-associativities. We also
tag all these rules as members of a new group, "arith", so we can more easily
define global syntax priorities later (at the end of the syntax module).

  syntax Exp ::= left:
                 Exp "*" Exp                       [strict, group(arith)]
               | Exp "/" Exp                       [strict, group(arith)]
               | Exp "%" Exp                       [strict, group(arith)]
               > left:
                 Exp "+" Exp                       [strict, left, group(arith)]
               | Exp "^" Exp                       [strict, left, group(arith)]
// left attribute should not be necessary; currently a parsing bug
               | Exp "-" Exp                       [strict, prefer, group(arith)]
// the "prefer" attribute above is to not parse x-1 as x(-1)
// Due to some parsing problems, we currently cannot add unary minus:
               | "-" Exp                           [strict, group(arith)]
               > non-assoc:
                 Exp "<" Exp                       [strict, group(arith)]
               | Exp "<=" Exp                      [strict, group(arith)]
               | Exp ">" Exp                       [strict, group(arith)]
               | Exp ">=" Exp                      [strict, group(arith)]
               | Exp "==" Exp                      [strict, group(arith)]
               | Exp "!=" Exp                      [strict, group(arith)]
               > "!" Exp                           [strict, group(arith)]
               > Exp "&&" Exp                      [strict(1), left, group(arith)]
               > Exp "||" Exp                      [strict(1), left, group(arith)]

The conditional construct has the expected evaluation strategy,
stating that only the first argument is evaluate:

  syntax Exp ::= "if" Exp "then" Exp "else" Exp    [strict(1)]

FUN's builtin lists are formed by enclosing comma-separated
sequences of expressions (i.e., terms of sort Exps) in square
brackets. The list constructor cons adds a new element to the
top of the list, head and tail get the first element
and the tail sublist of a list if they exist, respectively, and get
stuck otherwise, and null?? tests whether a list is empty or
not; syntactically, these are just expression constants.
In function patterns, we are also going to allow patterns following the
usual head/tail notation; for example, the pattern [x_1,...,x_n|t]
binds x_1, ..., x_n to the first elements of the matched list,
and t to the list formed with the remaining elements. We define list
patterns as ordinary expression constructs, although we will make sure that
we do not give them semantics if they appear in any other place then in a
function case pattern.

  syntax Exp ::= "[" Exps "]"                             [strict, klabel(list)]
               | "head" [macro] | "tail" [macro] | "null?" [macro]
               | "[" Exps "|" Exp "]"
  syntax Val ::= "[" Vals "]"                             [klabel(list)]
  syntax Cons ::= "cons"
  syntax Val ::= Cons
  syntax Val ::= Cons Val                                 [klabel(apply)]

Data constructors start with capital letters and they may or may
not have arguments. We need to use the attribute "prefer" to make
sure that, e.g., Cons(a) parses as constructor Cons with
argument a, and not as the expression Cons (because
constructor names are also expressions) regarded as a function applied
to the expression a. Also, note that the constructor is strict
in its second argument, because we want to evaluate its arguments but
not the constuctor name itsef.

  syntax ConstructorName                         [token]
  syntax Exp ::= ConstructorName
               | ConstructorName "(" Exps ")"    [prefer, strict(2), klabel(constructor)]
  syntax Val ::= ConstructorName "(" Vals ")"    [klabel(constructor)]

A function is essentially a |-separated ordered
sequence of cases, each case of the form pattern -> expression,
preceded by the language construct fun. Patterns will be defined
shortly, both for the builtin lists and for user-defined constructors.
Recall that the syntax we define in K is not meant to serve as a
ultimate parser for the defined language, but rather as a convenient
notation for K abstract syntax trees, which we prefer when we write
the semantic rules. It is therefore often the case that we define a
more ``generous'' syntax than we want to allow programs to use.
We do it here, too. Specifically, the syntax of Cases
below allows any expressions to appear as pattern. This syntactic
relaxation permits many wrong programs to be parsed, but that is not a
problem because we are not going to give semantics to wrong combinations,
so those programs will get stuck; moreover, our type inferencer will reject
those programs anyway. Function application is just concatenation of
expressions, without worrying about type correctness. Again, the type
system will reject type-incorrect programs.

  syntax Exp ::= "fun" Cases
               | Exp Exp                              [strict, left, klabel(apply)]
// NOTE: We would like eventually to also have Exp "(" Exps ")
  syntax Case  ::= Exp "->" Exp
  syntax Cases ::= List{Case, "|"}

The let and letrec binders have the usual syntax
and functional meaning. We allow multiple and-separated bindings.
Like for the function cases above, we allow a more generous syntax for
the left-hand sides of bindings, noting that the semantics will get stuck
on incorrect bindings and that the type system will reject those programs.

  syntax Exp ::= "let" Bindings "in" Exp
               | "letrec" Bindings "in" Exp                 [prefer]
// The "prefer" attribute for letrec currently needed due to tool bug,
// to make sure that "letrec" is not parsed as "let rec".
  syntax Binding  ::= Exp "=" Exp
  syntax Bindings ::= List{Binding,"and"}

References are first class values in FUN. The construct ref
takes an expression, evaluates it, and then it stores the resulting value
at a fresh location in the store and returns that reference. Syntactically,
ref is just an expression constant. The construct &
takes a name as argument and evaluates to a reference, namely the store
reference where the variable passed as argument stores its value; this
construct is a bit controversial and is further discussed in the
environment-based semantics of the FUN language, where we desugar
ref to it. The construct @ takes a reference
and evaluates to the value stored there. The construct := takes
two expressions, the first expected to evaluate to a reference; the value
of its second argument will be stored at the location to which the first
points (the old value is thus lost). Finally, since expression evaluation
now has side effects, it makes sense to also add a sequential composition
construct, which is sequentially strict. This evaluates to the value of
its second argument; the value of the first argument is lost (which has
therefore been evaluated only for its side effects.

  syntax Exp ::= "ref"                             [macro]
               | "&" Name
               | "@" Exp                                     [strict]
               | Exp ":=" Exp                                [strict]
               | Exp ";" Exp                       [strict(1), right]

Call-with-current-continuation, named callcc in FUN, is a
powerful control operator that originated in the Scheme programming
language, but it now exists in many other functional languages. It works
by evaluating its argument, expected to evaluate to a function, and by
passing the current continuation, or evaluation context (or computation,
in K terminology), as a special value to it. When/If this special value
is invoked, the current context is discarded and replaced with the one
held by the special value and the computation continues from there.
It is like taking a snapshot of the execution context at some moment
in time and then, when desired, being able to get back in time to that
point. If you like games, it is like saving the game now (so you can
work on your homework!) and then continuing the game tomorrow or whenever
you wish. To issustrate the strength of callcc, we also
allow exceptions in FUN by means of a conventional try-catch
construct, which will desugar to callcc. We also need to
introduce the special expression contant throw, but we need to
use it as a function argument name in the desugaring macro, so we define
it as a name instead of as an expression constant:

  syntax Exp ::= "try" Exp "catch" "(" Name ")" Exp [macro]
  syntax Val ::= "callcc"
  syntax Name ::= "throw" [token]

Finally, FUN also allows polymorphic datatype declarations. These
will be useful when we define the type system later on.

  syntax Exp ::= "datatype" Type "=" TypeCases Exp [macro]
// NOTE: In a future version of K, we want the datatype declaration
// to be a construct by itself, but that is not possible currently
// because K's parser wronly identifies the __ operation allowing
// a declaration to appear in front of an expression with the function
// application construct, giving ambiguous parsing errors.

We next need to define the syntax of types and type cases that appear
in datatype declarations.

Like in many functional languages, type parameters/variables in
user-defined types are quoted identifiers.

  syntax TypeVar                        [token]
  syntax TypeVars ::= List{TypeVar,","} [overload(types)]

Types can be basic types, function types, or user-defined
parametric types. In the dynamic semantics we are going to simply ignore
all the type declations, so here the syntax of types below is only useful
for generating the desired parser. To avoid syntactic ambiguities with
the arrow construct for function cases, we use the symbol --> as
a constructor for function types:

  syntax TypeName [token]
  syntax Type ::= "int" | "bool" | "string"
                | Type "-->" Type                            [right]
                | "(" Type ")"                             [bracket]
                | TypeVar
                | TypeName             [klabel(TypeName), avoid]
                | Type TypeName   [klabel(Type-TypeName), symbol, macro]
                | "(" Types ")" TypeName                    [prefer]
  syntax Types ::= List{Type,","} [overload(types)]
  syntax Types ::= TypeVars

  syntax TypeCase ::= ConstructorName
                    | ConstructorName "(" Types ")"
  syntax TypeCases ::= List{TypeCase,"|"}     [symbol(_|TypeCase_)]

Additional Priorities

  syntax priority @__FUN-UNTYPED-COMMON
                > apply
                > arith
                > _:=__FUN-UNTYPED-COMMON
                > let_in__FUN-UNTYPED-COMMON
                  letrec_in__FUN-UNTYPED-COMMON
                  if_then_else__FUN-UNTYPED-COMMON
                > _;__FUN-UNTYPED-COMMON
                > fun__FUN-UNTYPED-COMMON
                > datatype_=___FUN-UNTYPED-COMMON
endmodule

module FUN-UNTYPED-MACROS
  imports FUN-UNTYPED-COMMON

Desugaring macros

We desugar the list non-constructor operations to functions matching
over list patterns. In order to do that we need some new variables; for
those, we follow the same convention like in the K tutorial, where we
added them as new identifier constructs starting with the character $,
so we can easily recognize them when we debug or trace the semantics.

  syntax Name ::= "$h" [token] | "$t" [token]
  rule head => fun [$h|$t] -> $h
  rule tail => fun [$h|$t] -> $t
  rule null? => fun [.Exps] -> true | [$h|$t] -> false

Multiple-head list patterns desugar into successive one-head patterns:

  rule [E1,E2,Es:Exps|T] => [E1|[E2,Es|T]]                   [anywhere]

Uncurrying of multiple arguments in functions and binders:

  rule P1 P2 -> E => P1 -> fun P2 -> E                       [anywhere]
  rule F P = E => F = fun P -> E                             [anywhere]

We desugar the try-catch construct into callcc:

  syntax Name ::= "$k" [token] | "$v" [token]
  rule try E catch(X) E'
    => callcc (fun $k -> (fun throw -> E)(fun X -> $k E'))

For uniformity, we reduce all types to their general form:

  rule `Type-TypeName`(T:Type, Tn:TypeName) => (T) Tn

The dynamic semantics ignores all the type declarations:

  rule datatype _T = _TCs E => E

endmodule


module FUN-UNTYPED-SYNTAX
  imports FUN-UNTYPED-COMMON
  imports BUILTIN-ID-TOKENS

  syntax Name ::= r"[a-z][_a-zA-Z0-9]*"           [token, prec(2)]
                | #LowerId                        [token]
  syntax ConstructorName ::= #UpperId             [token]
  syntax TypeVar  ::= r"['][a-z][_a-zA-Z0-9]*"    [token]
  syntax TypeName ::= Name                        [token]
endmodule

Semantics

The semantics below is environment-based. A substitution-based
definition of FUN is also available, but that drops the &
construct as explained above.

module FUN-UNTYPED
  imports FUN-UNTYPED-COMMON
  imports FUN-UNTYPED-MACROS
  imports DOMAINS
  //imports PATTERN-MATCHING

Configuration

The k, env, and store cells are standard
(see, for example, the definition of LAMBDA++ or IMP++ in the first
part of the K tutorial).

  configuration <T color="yellow">
                  <k color="green"> $PGM:Exp </k>
                  <env color="violet"> .Map </env>
                  <store color="white"> .Map </store>
                </T>

Values and results

We only define integers, Booleans and strings as values here, but will
add more values later.

  syntax Val ::= Int | Bool | String
  syntax Vals ::= Bottoms
  syntax KResult ::= Val

Lookup

  rule <k> X:Name => V ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V ...</store>

Arithmetic expressions

  rule I1 * I2 => I1 *Int I2
  rule I1 / I2 => I1 /Int I2 requires I2 =/=K 0
  rule I1 % I2 => I1 %Int I2 requires I2 =/=K 0
  rule I1 + I2 => I1 +Int I2
  rule S1 ^ S2 => S1 +String S2
  rule I1 - I2 => I1 -Int I2
  rule - I => 0 -Int I
  rule I1 < I2 => I1 <Int I2
  rule I1 <= I2 => I1 <=Int I2
  rule I1 > I2 => I1 >Int I2
  rule I1 >= I2 => I1 >=Int I2
  rule V1:Val == V2:Val => V1 ==K V2
  rule V1:Val != V2:Val => V1 =/=K V2
  rule ! T => notBool(T)
  rule true  && E => E
  rule false && _ => false
  rule true  || _ => true
  rule false || E => E

Conditional

  rule if  true then E else _ => E
  rule if false then _ else E => E

Lists

We have already declared the syntactic list of expressions strict, so
we can assume that all the elements that appear in a FUN list are
evaluated. The only thing left to do is to state that a list of
values is a value itself, that is, that the list square-bracket
construct is indeed a constructor, and to give the semantics of
cons. Since cons is a builtin function and is
expected to take two arguments, we have to also state that
cons itself is a value (specifically, a function/closure
value, but we do not need that level of detail here), and also that
cons applied to a value is a value (specifically, it would be
a function/closure value that expects the second, list argument):

  rule cons V:Val [Vs:Vals] => [V,Vs]

Data Constructors

Constructors take values as arguments and produce other values:

  syntax Val ::= ConstructorName

Functions and Closures

Like in the environment-based semantics of LAMBDA++ in the first part
of the K tutorial, functions evaluate to closures. A closure includes
the current environment besides the function contents; the environment
will be used at execution time to lookup all the variables that appear
free in the function body (we want static scoping in FUN).

  syntax Val ::= closure(Map,Cases)
  rule <k> fun Cases => closure(Rho,Cases) ...</k>  <env> Rho </env>

Note: The reader may want to get familiar with
how the pre-defined pattern matching works before proceeding.
The best way to do that is to consult
k/include/modules/pattern-matching.k.

We distinguish two cases when the closure is applied.
If the first pattern matches, then we pick the first case: switch to
the closed environment, get the matching map and bind all its
variables, and finally evaluate the function body of the first case,
making sure that the environment is properly recovered afterwards.
If the first pattern does not match, then we drop it and thus move on
to the next one.

  rule (.K => getMatching(P, V)) ~> closure(_, P->_ | _) V:Val
  rule <k> matchResult(M:Map) ~> closure(Rho, _->E | _) _
           => bindMap(M) ~> E ~> setEnv(Rho') ...</k>
       <env> Rho' => Rho </env>
  rule (matchFailure => .K) ~> closure(_, (_->_ | Cs:Cases => Cs)) _
//  rule <k> closure(Rho, P->E | _) V:Val
//           => bindMap(getMatching(P,V)) ~> E ~> setEnv(Rho') ...</k>
//       <env> Rho' => Rho </env>  when isMatching(P,V)
//  rule closure(_, (P->_ | Cs:Cases => Cs)) V:Val  when notBool isMatching(P,V)

Let and Letrec

To highlight the similarities and differences between let and
letrec, we prefer to give them direct semantics instead of
to desugar them like in LAMBDA. See the formal definitions of
bindTo, bind, and assignTo at the end of
this module. Informally, bindTo(Xs, Es) first
evaluates the expressions Es in Exps in the current
environment (i.e., it is strict in its second argument), then it binds
the variables in Xs in Names to new locations and adds
those bindings to the environment, and finally writes the values
previously obtained after evaluating the expressions Es to those
new locations; bind(Xs) does only the bindings of
Xs to new locations and adds those bindings to the environment;
and assignTo(Xs,Es) evaluates the expressions
Es in the current environment and then it writes the resulting
values to the locations to which the variables Xs are already
bound to in the environment.

Therefore, let Xs = Es in E first
evaluates Es in the current environment, then adds new
bindings for Xs to fresh locations in the environment, then
writes the values of Es to those locations, and finally
evaluates E in the new environment, making sure that the
environment is properly recovered after the evaluation of E.
On the other hand, letrec does the same things but in a
different order: it first adds new bindings for Xs to fresh
locations in the environment, then it evaluates Es in the new
environment, then it writes the resulting values to their
corresponding locations, and finally it evaluates E and
recovers the environment. The crucial difference is that the
expressions Es now see the locations of the variables Xs
in the environment, so if they are functions, which is typically the
case with letrec, their closures will encapsulate in their
environments the bindings of all the bound variables, including
themselves (thus, we may have a closure value stored at location
L, whose environment contains a binding of the form
F ↦ L; this way, the closure can invoke
itself).

  rule <k> let Bs in E
        => bindTo(names(Bs),exps(Bs)) ~> E ~> setEnv(Rho) ...</k>
       <env> Rho </env>

  rule <k> letrec Bs in E
        => bind(names(Bs))~>assignTo(names(Bs),exps(Bs))~>E~>setEnv(Rho)...</k>
       <env> Rho </env>

Recall that our syntax allows let and letrec to
take any expression in place of its binding. This allows us to use
the already existing function application construct to bind names to
functions, such as, e.g., let x y = y in ....
The desugaring macro in the syntax module uncurries such declarations,
and then the semantic rules above only work when the remaining
bindings are identifiers, so the semantics will get stuck on programs
that misuse the let and letrec binders.

References

The semantics of references is self-explanatory, except maybe for the
desugaring rule of ref, which is further discussed. Note
that &X grabs the location of X from the environment.
Sequential composition, which is needed only to accumulate the
side effects due to assignments, was strict in the first argument.
Once evaluated, its first argument is simply discarded:

  syntax Name ::= "$x" [token]
  rule ref => fun $x -> & $x
  rule <k> & X => L ...</k>  <env>... X |-> L ...</env>
  rule <k> @ L:Int => V:Val ...</k>  <store>... L |-> V ...</store>
  rule <k> L:Int := V:Val => V ...</k>  <store>... L |-> (_=>V) ...</store>
  rule _V:Val; E => E

The desugaring rule of ref (first rule above) works
because & takes a variable and returns its location (like in C).
Note that some ``pure'' functional programming researchers strongly dislike
the & construct, but favor ref. We refrain from having
a personal opinion on this issue here, but support & in the
environment-based definition of FUN because it is, technically speaking,
more powerful than ref. From a language design perspective, it
would be equally easy to drop & and instead give a direct
semantics to ref. In fact, this is precisely what we do in the
substitution-based definition of FUN, because there appears to be no way
to give a substitution-based definition to the & construct.

Callcc

As we know it from the LAMBDA++ tutorial, call-with-current-continuation
is quite easy to define in K. We first need to define a special
value wrapping an execution context, that is, an environment saying
where the variables should be looked up, and a computation structure
saying what is left to execute (in a substitution-based definition,
this special value would be even simpler, as it would only need to
wrap the computation structure---see, for example, the
substitution-based semantics of LAMBDA++ in the the first part of the
K tutorial, or the substitution-based definition of FUN). Then
callcc creates such a value containing the current
environment and the current remaining computation, and passes it to
its argument function. When/If invoked, the special value replaces
the current execution context with its own and continues the execution
normally.

  syntax Val ::= cc(Map,K)
  rule <k> (callcc V:Val => V cc(Rho,K)) ~> K </k>  <env> Rho </env>
  rule <k> cc(Rho,K) V:Val ~> _ => V ~> K </k>  <env> _ => Rho </env>

Auxiliary operations

Environment recovery

The environment recovery operation is the same as for the LAMBDA++
language in the K tutorial and many other languages provided with the
K distribution. The first ``anywhere'' rule below shows an elegant
way to achieve the benefits of tail recursion in K.

  syntax KItem ::= setEnv(Map)  // TODO: get rid of env
  //rule (setEnv(_) => .) ~> setEnv(_)  [anywhere]
  rule <k> _:Val ~> (setEnv(Rho) => .K) ...</k> <env> _ => Rho </env>

bindTo, bind and assignTo

The meaning of these operations has already been explained when we
discussed the let and letrec language constructs
above.

  syntax KItem ::= bindTo(Names,Exps)         [strict(2)]
                 | bindMap(Map)
                 | bind(Names)

  rule (.K => getMatchingAux(Xs,Vs)) ~> bindTo(Xs:Names,Vs:Vals)
  rule matchResult(M:Map) ~> bindTo(_:Names, _:Vals) => bindMap(M)

  rule bindMap(.Map) => .K
  rule <k> bindMap((X:Name |-> V:Val => .Map) _:Map) ...</k>
       <env> Rho => Rho[X <- !L:Int] </env>
       <store>... .Map => !L |-> V ...</store>

  rule bind(.Names) => .K
  rule <k> bind(X:Name,Xs => Xs) ...</k>
       <env> Rho => Rho[X <- !_L:Int] </env>

  syntax KItem ::= assignTo(Names,Exps)  [strict(2)]

  rule <k> assignTo(.Names,.Vals) => .K ...</k>
  rule <k> assignTo((X:Name,Xs => Xs),(V:Val,Vs:Vals => Vs)) ...</k>
       <env>... X |-> L ...</env>
       <store>... .Map => L |-> V ...</store>

Getters

The following auxiliary operations extract the list of identifiers
and of expressions in a binding, respectively.

  syntax Names ::= names(Bindings)  [function]
  rule names(.Bindings) => .Names
  rule names(X:Name=_ and Bs) => (X,names(Bs))::Names

  syntax Exps ::= exps(Bindings)  [function]
  rule exps(.Bindings) => .Exps
  rule exps(_:Name=E and Bs) => E,exps(Bs)

  /* Extra kore stuff */
  syntax KResult ::= Vals
  syntax Exps ::= Names
  syntax Names ::= Bottoms

  /* Matching */
  syntax MatchResult ::= getMatching(Exp, Val)                      [function]
                       | getMatchingAux(Exps, Vals)                 [function]
                       | mergeMatching(MatchResult, MatchResult)    [function]
                       | matchResult(Map)
                       | "matchFailure"

  rule getMatching(C:ConstructorName(Es:Exps), C(Vs:Vals)) => getMatchingAux(Es, Vs)
  rule getMatching([Es:Exps], [Vs:Vals])                   => getMatchingAux(Es, Vs)
  rule getMatching(C:ConstructorName, C) => matchResult(.Map)
  rule getMatching(B:Bool, B)            => matchResult(.Map)
  rule getMatching(I:Int, I)             => matchResult(.Map)
  rule getMatching(S:String, S)          => matchResult(.Map)
  rule getMatching(N:Name, V:Val) => matchResult(N |-> V)
  rule getMatching(_, _) => matchFailure        [owise]

  rule getMatchingAux((E:Exp, Es:Exps), (V:Val, Vs:Vals)) => mergeMatching(getMatching(E, V), getMatchingAux(Es, Vs))
  rule getMatchingAux(.Exps, .Vals)                       => matchResult(.Map)
  rule getMatchingAux(_, _) => matchFailure     [owise]

  rule mergeMatching(matchResult(M1:Map), matchResult(M2:Map)) => matchResult(M1 M2)
    requires intersectSet(keys(M1), keys(M2)) ==K .Set
  //rule mergeMatching(_, _) => matchFailure      [owsie]
  rule mergeMatching(matchResult(_:Map), matchFailure) => matchFailure
  rule mergeMatching(matchFailure, matchResult(_:Map)) => matchFailure
  rule mergeMatching(matchFailure, matchFailure)       => matchFailure

Besides the generic decomposition rules for patterns and values,
we also want to allow [head|tail] matching for lists, so we add
the following custom pattern decomposition rule:

  rule getMatching([H:Exp | T:Exp], [V:Val, Vs:Vals])
    => getMatchingAux((H, T), (V, [Vs]))
endmodule

Go to Lesson 2, FUN untyped, Substitution-Based.

FUN — Untyped — Environment

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K semantic definition of the untyped FUN language.
FUN is a pedagogical and research language that captures the essence
of the functional programming paradigm, extended with several features
often encountered in functional programming languages.
Like many functional languages, FUN is an expression language, that
is, everything, including the main program, is an expression.
Functions can be declared anywhere and are first class values in the
language.
FUN is call-by-value here, but it has been extended (as student
homework assignments) with other parameter-passing styles.
To make it more interesting and to highlight some of K's strengths,
FUN includes the following features:

  • The basic builtin data-types of integers, booleans and strings.

  • Builtin lists, which can hold any elements, including other lists.
    Lists are enclosed in square brackets and their elements are
    comma-separated; e.g., [1,2,3].

  • User-defined data-types, by means of constructor terms.
    Constructor names start with a capital letter (while any other
    identifier in the language starts with a lowercase letter), and they
    can be followed by an arbitrary number of comma-separated arguments
    enclosed in parentheses; parentheses are not needed when the
    constructor takes no arguments.
    For example, Pair(5,7) is a constructor term holding two
    numbers, Cons(1,Cons(2,Cons(3,Nil))) is a list-like
    constructor term holding 3 elements, and
    Tree(Tree(Leaf(1), Leaf(2)), Leaf(3)) is a tree-like
    constructor term holding 3 elements.
    In the untyped version of the FUN language, no type checking or
    inference is performed to ensure that the data constructors are used
    correctly.
    The execution will simply get stuck when they are misused.
    Moreover, since no type checking is performed, the data-types are not
    even declared in the untyped version of FUN.

  • Functions and let/letrec binders can take
    multiple space-separated arguments, but these are desugared to
    ones that only take one argument, by currying. For example, the
    expressions

    fun x y -> x y
    let x y = y in x
    

    are desugared, respectively, into the following expressions:

    fun x -> fun y -> x y
    let x = fun y -> y in x
    
  • Functions can be defined using pattern matching over the
    available data-types. For example, the program

    letrec max = fun [h] -> h
                 |   [h|t] -> let x = max t
                              in  if h > x then h else x
    in max [1, 3, 5, 2, 4, 0, -1, -5]
    

    defines a function max that calculates the maximum element of
    a non-empty list, and the function

    letrec ack = fun Pair(0,n) -> n + 1
                 |   Pair(m,0) -> ack Pair(m - 1, 1)
                 |   Pair(m,n) -> ack Pair(m - 1, ack Pair(m, n - 1))
    in ack Pair(2,3)
    

    calculates the Ackermann function applied to a particular pair of numbers.
    Patterns can be nested. Patterns can currently only be used in function
    definitions, and not directly in let/letrec binders.
    For example, this is not allowed:

    letrec Pai(x,y) = Pair(1,2) in x+y
    

    But this is allowed:

    let f Pair(x,y) = x+y in f Pair(1,2)
    

    because it is first reduced to

    let f = fun Pair(x,y) -> x+y in f Pair(1,2)
    

    by uncurrying of the let binder, and pattern matching is
    allowed in function arguments.

  • We include a callcc construct, for two reasons: first,
    several functional languages support this construct; second, some
    semantic frameworks have difficulties defining it. Not K.

  • Finally, we include mutables by means of referencing an
    expression, getting the reference of a variable, dereferencing and
    assignment. We include these for the same reasons as above: there are
    languages which have them, and they are not easy to define in some
    semantic frameworks.

Like in many other languages, some of FUN's constructs can be
desugared into a smaller set of basic constructs. We do that as usual,
using macros, and then we only give semantics to the core constructs.

Note:
We recommend the reader to first consult the dynamic semantics of the
LAMBDA++ language in the first part of the K Tutorial.
To keep the comments below small and focused, we will not re-explain
functional or K features that have already been explained in there.

Syntax

//require "modules/pattern-matching.k"

module FUN-UNTYPED-COMMON
  imports DOMAINS-SYNTAX

FUN is an expression language. The constructs below fall into
several categories: names, arithmetic constructs, conventional
functional constructs, patterns and pattern matching, data constructs,
lists, references, and call-with-current-continuation (callcc).
The arithmetic constructs are standard; they are present in almost all
our K language definitions. The meaning of FUN's constructs are
discussed in more depth when we define their semantics in the next
module.

The Syntactic Constructs

We start with the syntactic definition of FUN names.
We have several categories of names: ones to be used for functions and
variables, others to be used for data constructors, others for types and
others for type variables. We will introduce them as needed, starting
with the former category. We prefer the names of variables and functions
to start with lower case letters. We take the freedom to tacitly introduce
syntactic lists/sequences for each nonterminal for which we need them:

  syntax Name                                      [token]
  syntax Names ::= List{Name,","}                  [overload(exps)]

Expression constructs will be defined throughtout the syntax module.
Below are the very basic ones, namely the builtins, the names, and the
parentheses used as brackets for grouping. Lists of expressions are
declared strict, so all expressions in the list get evaluated whenever
the list is on a position which can be evaluated:

  syntax Exp ::= Int | Bool | String | Name
               | "(" Exp ")"                       [bracket]
  syntax Exps  ::= List{Exp,","}                   [strict, overload(exps)]
  syntax Val
  syntax Exp ::= Val
  syntax Exps ::= Vals
  syntax Vals ::= List{Val,","}                    [overload(exps)]
  syntax Bottom
  syntax Bottoms ::= List{Bottom,","}              [overload(exps)]

We next define the syntax of arithmetic constructs, together with
their relative priorities and left-/non-associativities. We also
tag all these rules as members of a new group, "arith", so we can more easily
define global syntax priorities later (at the end of the syntax module).

  syntax Exp ::= left:
                 Exp "*" Exp                       [strict, group(arith)]
               | Exp "/" Exp                       [strict, group(arith)]
               | Exp "%" Exp                       [strict, group(arith)]
               > left:
                 Exp "+" Exp                       [strict, left, group(arith)]
               | Exp "^" Exp                       [strict, left, group(arith)]
// left attribute should not be necessary; currently a parsing bug
               | Exp "-" Exp                       [strict, prefer, group(arith)]
// the "prefer" attribute above is to not parse x-1 as x(-1)
// Due to some parsing problems, we currently cannot add unary minus:
               | "-" Exp                           [strict, group(arith)]
               > non-assoc:
                 Exp "<" Exp                       [strict, group(arith)]
               | Exp "<=" Exp                      [strict, group(arith)]
               | Exp ">" Exp                       [strict, group(arith)]
               | Exp ">=" Exp                      [strict, group(arith)]
               | Exp "==" Exp                      [strict, group(arith)]
               | Exp "!=" Exp                      [strict, group(arith)]
               > "!" Exp                           [strict, group(arith)]
               > Exp "&&" Exp                      [strict(1), left, group(arith)]
               > Exp "||" Exp                      [strict(1), left, group(arith)]

The conditional construct has the expected evaluation strategy,
stating that only the first argument is evaluate:

  syntax Exp ::= "if" Exp "then" Exp "else" Exp    [strict(1)]

FUN's builtin lists are formed by enclosing comma-separated
sequences of expressions (i.e., terms of sort Exps) in square
brackets. The list constructor cons adds a new element to the
top of the list, head and tail get the first element
and the tail sublist of a list if they exist, respectively, and get
stuck otherwise, and null?? tests whether a list is empty or
not; syntactically, these are just expression constants.
In function patterns, we are also going to allow patterns following the
usual head/tail notation; for example, the pattern [x_1,...,x_n|t]
binds x_1, ..., x_n to the first elements of the matched list,
and t to the list formed with the remaining elements. We define list
patterns as ordinary expression constructs, although we will make sure that
we do not give them semantics if they appear in any other place then in a
function case pattern.

  syntax Exp ::= "[" Exps "]"                             [strict, klabel(list)]
               | "head" [macro] | "tail" [macro] | "null?" [macro]
               | "[" Exps "|" Exp "]"
  syntax Val ::= "[" Vals "]"                             [klabel(list)]
  syntax Cons ::= "cons"
  syntax Val ::= Cons
  syntax Val ::= Cons Val                                 [klabel(apply)]

Data constructors start with capital letters and they may or may
not have arguments. We need to use the attribute "prefer" to make
sure that, e.g., Cons(a) parses as constructor Cons with
argument a, and not as the expression Cons (because
constructor names are also expressions) regarded as a function applied
to the expression a. Also, note that the constructor is strict
in its second argument, because we want to evaluate its arguments but
not the constuctor name itsef.

  syntax ConstructorName                         [token]
  syntax Exp ::= ConstructorName
               | ConstructorName "(" Exps ")"    [prefer, strict(2), klabel(constructor)]
  syntax Val ::= ConstructorName "(" Vals ")"    [klabel(constructor)]

A function is essentially a |-separated ordered
sequence of cases, each case of the form pattern -> expression,
preceded by the language construct fun. Patterns will be defined
shortly, both for the builtin lists and for user-defined constructors.
Recall that the syntax we define in K is not meant to serve as a
ultimate parser for the defined language, but rather as a convenient
notation for K abstract syntax trees, which we prefer when we write
the semantic rules. It is therefore often the case that we define a
more ``generous'' syntax than we want to allow programs to use.
We do it here, too. Specifically, the syntax of Cases
below allows any expressions to appear as pattern. This syntactic
relaxation permits many wrong programs to be parsed, but that is not a
problem because we are not going to give semantics to wrong combinations,
so those programs will get stuck; moreover, our type inferencer will reject
those programs anyway. Function application is just concatenation of
expressions, without worrying about type correctness. Again, the type
system will reject type-incorrect programs.

  syntax Exp ::= "fun" Cases
               | Exp Exp                              [strict, left, klabel(apply)]
// NOTE: We would like eventually to also have Exp "(" Exps ")
  syntax Case  ::= Exp "->" Exp
  syntax Cases ::= List{Case, "|"}

The let and letrec binders have the usual syntax
and functional meaning. We allow multiple and-separated bindings.
Like for the function cases above, we allow a more generous syntax for
the left-hand sides of bindings, noting that the semantics will get stuck
on incorrect bindings and that the type system will reject those programs.

  syntax Exp ::= "let" Bindings "in" Exp
               | "letrec" Bindings "in" Exp                 [prefer]
// The "prefer" attribute for letrec currently needed due to tool bug,
// to make sure that "letrec" is not parsed as "let rec".
  syntax Binding  ::= Exp "=" Exp
  syntax Bindings ::= List{Binding,"and"}

References are first class values in FUN. The construct ref
takes an expression, evaluates it, and then it stores the resulting value
at a fresh location in the store and returns that reference. Syntactically,
ref is just an expression constant. The construct &
takes a name as argument and evaluates to a reference, namely the store
reference where the variable passed as argument stores its value; this
construct is a bit controversial and is further discussed in the
environment-based semantics of the FUN language, where we desugar
ref to it. The construct @ takes a reference
and evaluates to the value stored there. The construct := takes
two expressions, the first expected to evaluate to a reference; the value
of its second argument will be stored at the location to which the first
points (the old value is thus lost). Finally, since expression evaluation
now has side effects, it makes sense to also add a sequential composition
construct, which is sequentially strict. This evaluates to the value of
its second argument; the value of the first argument is lost (which has
therefore been evaluated only for its side effects.

  syntax Exp ::= "ref"                             [macro]
               | "&" Name
               | "@" Exp                                     [strict]
               | Exp ":=" Exp                                [strict]
               | Exp ";" Exp                       [strict(1), right]

Call-with-current-continuation, named callcc in FUN, is a
powerful control operator that originated in the Scheme programming
language, but it now exists in many other functional languages. It works
by evaluating its argument, expected to evaluate to a function, and by
passing the current continuation, or evaluation context (or computation,
in K terminology), as a special value to it. When/If this special value
is invoked, the current context is discarded and replaced with the one
held by the special value and the computation continues from there.
It is like taking a snapshot of the execution context at some moment
in time and then, when desired, being able to get back in time to that
point. If you like games, it is like saving the game now (so you can
work on your homework!) and then continuing the game tomorrow or whenever
you wish. To issustrate the strength of callcc, we also
allow exceptions in FUN by means of a conventional try-catch
construct, which will desugar to callcc. We also need to
introduce the special expression contant throw, but we need to
use it as a function argument name in the desugaring macro, so we define
it as a name instead of as an expression constant:

  syntax Exp ::= "try" Exp "catch" "(" Name ")" Exp [macro]
  syntax Val ::= "callcc"
  syntax Name ::= "throw" [token]

Finally, FUN also allows polymorphic datatype declarations. These
will be useful when we define the type system later on.

  syntax Exp ::= "datatype" Type "=" TypeCases Exp [macro]
// NOTE: In a future version of K, we want the datatype declaration
// to be a construct by itself, but that is not possible currently
// because K's parser wronly identifies the __ operation allowing
// a declaration to appear in front of an expression with the function
// application construct, giving ambiguous parsing errors.

We next need to define the syntax of types and type cases that appear
in datatype declarations.

Like in many functional languages, type parameters/variables in
user-defined types are quoted identifiers.

  syntax TypeVar                        [token]
  syntax TypeVars ::= List{TypeVar,","} [overload(types)]

Types can be basic types, function types, or user-defined
parametric types. In the dynamic semantics we are going to simply ignore
all the type declations, so here the syntax of types below is only useful
for generating the desired parser. To avoid syntactic ambiguities with
the arrow construct for function cases, we use the symbol --> as
a constructor for function types:

  syntax TypeName [token]
  syntax Type ::= "int" | "bool" | "string"
                | Type "-->" Type                            [right]
                | "(" Type ")"                             [bracket]
                | TypeVar
                | TypeName             [klabel(TypeName), avoid]
                | Type TypeName   [klabel(Type-TypeName), symbol, macro]
                | "(" Types ")" TypeName                    [prefer]
  syntax Types ::= List{Type,","} [overload(types)]
  syntax Types ::= TypeVars

  syntax TypeCase ::= ConstructorName
                    | ConstructorName "(" Types ")"
  syntax TypeCases ::= List{TypeCase,"|"}     [symbol(_|TypeCase_)]

Additional Priorities

  syntax priority @__FUN-UNTYPED-COMMON
                > apply
                > arith
                > _:=__FUN-UNTYPED-COMMON
                > let_in__FUN-UNTYPED-COMMON
                  letrec_in__FUN-UNTYPED-COMMON
                  if_then_else__FUN-UNTYPED-COMMON
                > _;__FUN-UNTYPED-COMMON
                > fun__FUN-UNTYPED-COMMON
                > datatype_=___FUN-UNTYPED-COMMON
endmodule

module FUN-UNTYPED-MACROS
  imports FUN-UNTYPED-COMMON

Desugaring macros

We desugar the list non-constructor operations to functions matching
over list patterns. In order to do that we need some new variables; for
those, we follow the same convention like in the K tutorial, where we
added them as new identifier constructs starting with the character $,
so we can easily recognize them when we debug or trace the semantics.

  syntax Name ::= "$h" [token] | "$t" [token]
  rule head => fun [$h|$t] -> $h
  rule tail => fun [$h|$t] -> $t
  rule null? => fun [.Exps] -> true | [$h|$t] -> false

Multiple-head list patterns desugar into successive one-head patterns:

  rule [E1,E2,Es:Exps|T] => [E1|[E2,Es|T]]                   [anywhere]

Uncurrying of multiple arguments in functions and binders:

  rule P1 P2 -> E => P1 -> fun P2 -> E                       [anywhere]
  rule F P = E => F = fun P -> E                             [anywhere]

We desugar the try-catch construct into callcc:

  syntax Name ::= "$k" [token] | "$v" [token]
  rule try E catch(X) E'
    => callcc (fun $k -> (fun throw -> E)(fun X -> $k E'))

For uniformity, we reduce all types to their general form:

  rule `Type-TypeName`(T:Type, Tn:TypeName) => (T) Tn

The dynamic semantics ignores all the type declarations:

  rule datatype _T = _TCs E => E

endmodule


module FUN-UNTYPED-SYNTAX
  imports FUN-UNTYPED-COMMON
  imports BUILTIN-ID-TOKENS

  syntax Name ::= r"[a-z][_a-zA-Z0-9]*"           [token, prec(2)]
                | #LowerId                        [token]
  syntax ConstructorName ::= #UpperId             [token]
  syntax TypeVar  ::= r"['][a-z][_a-zA-Z0-9]*"    [token]
  syntax TypeName ::= Name                        [token]
endmodule

Semantics

The semantics below is environment-based. A substitution-based
definition of FUN is also available, but that drops the &
construct as explained above.

module FUN-UNTYPED
  imports FUN-UNTYPED-COMMON
  imports FUN-UNTYPED-MACROS
  imports DOMAINS
  //imports PATTERN-MATCHING

Configuration

The k, env, and store cells are standard
(see, for example, the definition of LAMBDA++ or IMP++ in the first
part of the K tutorial).

  configuration <T color="yellow">
                  <k color="green"> $PGM:Exp </k>
                  <env color="violet"> .Map </env>
                  <store color="white"> .Map </store>
                </T>

Values and results

We only define integers, Booleans and strings as values here, but will
add more values later.

  syntax Val ::= Int | Bool | String
  syntax Vals ::= Bottoms
  syntax KResult ::= Val

Lookup

  rule <k> X:Name => V ...</k>
       <env>... X |-> L ...</env>
       <store>... L |-> V ...</store>

Arithmetic expressions

  rule I1 * I2 => I1 *Int I2
  rule I1 / I2 => I1 /Int I2 requires I2 =/=K 0
  rule I1 % I2 => I1 %Int I2 requires I2 =/=K 0
  rule I1 + I2 => I1 +Int I2
  rule S1 ^ S2 => S1 +String S2
  rule I1 - I2 => I1 -Int I2
  rule - I => 0 -Int I
  rule I1 < I2 => I1 <Int I2
  rule I1 <= I2 => I1 <=Int I2
  rule I1 > I2 => I1 >Int I2
  rule I1 >= I2 => I1 >=Int I2
  rule V1:Val == V2:Val => V1 ==K V2
  rule V1:Val != V2:Val => V1 =/=K V2
  rule ! T => notBool(T)
  rule true  && E => E
  rule false && _ => false
  rule true  || _ => true
  rule false || E => E

Conditional

  rule if  true then E else _ => E
  rule if false then _ else E => E

Lists

We have already declared the syntactic list of expressions strict, so
we can assume that all the elements that appear in a FUN list are
evaluated. The only thing left to do is to state that a list of
values is a value itself, that is, that the list square-bracket
construct is indeed a constructor, and to give the semantics of
cons. Since cons is a builtin function and is
expected to take two arguments, we have to also state that
cons itself is a value (specifically, a function/closure
value, but we do not need that level of detail here), and also that
cons applied to a value is a value (specifically, it would be
a function/closure value that expects the second, list argument):

  rule cons V:Val [Vs:Vals] => [V,Vs]

Data Constructors

Constructors take values as arguments and produce other values:

  syntax Val ::= ConstructorName

Functions and Closures

Like in the environment-based semantics of LAMBDA++ in the first part
of the K tutorial, functions evaluate to closures. A closure includes
the current environment besides the function contents; the environment
will be used at execution time to lookup all the variables that appear
free in the function body (we want static scoping in FUN).

  syntax Val ::= closure(Map,Cases)
  rule <k> fun Cases => closure(Rho,Cases) ...</k>  <env> Rho </env>

Note: The reader may want to get familiar with
how the pre-defined pattern matching works before proceeding.
The best way to do that is to consult
k/include/modules/pattern-matching.k.

We distinguish two cases when the closure is applied.
If the first pattern matches, then we pick the first case: switch to
the closed environment, get the matching map and bind all its
variables, and finally evaluate the function body of the first case,
making sure that the environment is properly recovered afterwards.
If the first pattern does not match, then we drop it and thus move on
to the next one.

  rule (.K => getMatching(P, V)) ~> closure(_, P->_ | _) V:Val
  rule <k> matchResult(M:Map) ~> closure(Rho, _->E | _) _
           => bindMap(M) ~> E ~> setEnv(Rho') ...</k>
       <env> Rho' => Rho </env>
  rule (matchFailure => .K) ~> closure(_, (_->_ | Cs:Cases => Cs)) _
//  rule <k> closure(Rho, P->E | _) V:Val
//           => bindMap(getMatching(P,V)) ~> E ~> setEnv(Rho') ...</k>
//       <env> Rho' => Rho </env>  when isMatching(P,V)
//  rule closure(_, (P->_ | Cs:Cases => Cs)) V:Val  when notBool isMatching(P,V)

Let and Letrec

To highlight the similarities and differences between let and
letrec, we prefer to give them direct semantics instead of
to desugar them like in LAMBDA. See the formal definitions of
bindTo, bind, and assignTo at the end of
this module. Informally, bindTo(Xs, Es) first
evaluates the expressions Es in Exps in the current
environment (i.e., it is strict in its second argument), then it binds
the variables in Xs in Names to new locations and adds
those bindings to the environment, and finally writes the values
previously obtained after evaluating the expressions Es to those
new locations; bind(Xs) does only the bindings of
Xs to new locations and adds those bindings to the environment;
and assignTo(Xs,Es) evaluates the expressions
Es in the current environment and then it writes the resulting
values to the locations to which the variables Xs are already
bound to in the environment.

Therefore, let Xs = Es in E first
evaluates Es in the current environment, then adds new
bindings for Xs to fresh locations in the environment, then
writes the values of Es to those locations, and finally
evaluates E in the new environment, making sure that the
environment is properly recovered after the evaluation of E.
On the other hand, letrec does the same things but in a
different order: it first adds new bindings for Xs to fresh
locations in the environment, then it evaluates Es in the new
environment, then it writes the resulting values to their
corresponding locations, and finally it evaluates E and
recovers the environment. The crucial difference is that the
expressions Es now see the locations of the variables Xs
in the environment, so if they are functions, which is typically the
case with letrec, their closures will encapsulate in their
environments the bindings of all the bound variables, including
themselves (thus, we may have a closure value stored at location
L, whose environment contains a binding of the form
F ↦ L; this way, the closure can invoke
itself).

  rule <k> let Bs in E
        => bindTo(names(Bs),exps(Bs)) ~> E ~> setEnv(Rho) ...</k>
       <env> Rho </env>

  rule <k> letrec Bs in E
        => bind(names(Bs))~>assignTo(names(Bs),exps(Bs))~>E~>setEnv(Rho)...</k>
       <env> Rho </env>

Recall that our syntax allows let and letrec to
take any expression in place of its binding. This allows us to use
the already existing function application construct to bind names to
functions, such as, e.g., let x y = y in ....
The desugaring macro in the syntax module uncurries such declarations,
and then the semantic rules above only work when the remaining
bindings are identifiers, so the semantics will get stuck on programs
that misuse the let and letrec binders.

References

The semantics of references is self-explanatory, except maybe for the
desugaring rule of ref, which is further discussed. Note
that &X grabs the location of X from the environment.
Sequential composition, which is needed only to accumulate the
side effects due to assignments, was strict in the first argument.
Once evaluated, its first argument is simply discarded:

  syntax Name ::= "$x" [token]
  rule ref => fun $x -> & $x
  rule <k> & X => L ...</k>  <env>... X |-> L ...</env>
  rule <k> @ L:Int => V:Val ...</k>  <store>... L |-> V ...</store>
  rule <k> L:Int := V:Val => V ...</k>  <store>... L |-> (_=>V) ...</store>
  rule _V:Val; E => E

The desugaring rule of ref (first rule above) works
because & takes a variable and returns its location (like in C).
Note that some ``pure'' functional programming researchers strongly dislike
the & construct, but favor ref. We refrain from having
a personal opinion on this issue here, but support & in the
environment-based definition of FUN because it is, technically speaking,
more powerful than ref. From a language design perspective, it
would be equally easy to drop & and instead give a direct
semantics to ref. In fact, this is precisely what we do in the
substitution-based definition of FUN, because there appears to be no way
to give a substitution-based definition to the & construct.

Callcc

As we know it from the LAMBDA++ tutorial, call-with-current-continuation
is quite easy to define in K. We first need to define a special
value wrapping an execution context, that is, an environment saying
where the variables should be looked up, and a computation structure
saying what is left to execute (in a substitution-based definition,
this special value would be even simpler, as it would only need to
wrap the computation structure---see, for example, the
substitution-based semantics of LAMBDA++ in the the first part of the
K tutorial, or the substitution-based definition of FUN). Then
callcc creates such a value containing the current
environment and the current remaining computation, and passes it to
its argument function. When/If invoked, the special value replaces
the current execution context with its own and continues the execution
normally.

  syntax Val ::= cc(Map,K)
  rule <k> (callcc V:Val => V cc(Rho,K)) ~> K </k>  <env> Rho </env>
  rule <k> cc(Rho,K) V:Val ~> _ => V ~> K </k>  <env> _ => Rho </env>

Auxiliary operations

Environment recovery

The environment recovery operation is the same as for the LAMBDA++
language in the K tutorial and many other languages provided with the
K distribution. The first ``anywhere'' rule below shows an elegant
way to achieve the benefits of tail recursion in K.

  syntax KItem ::= setEnv(Map)  // TODO: get rid of env
  //rule (setEnv(_) => .) ~> setEnv(_)  [anywhere]
  rule <k> _:Val ~> (setEnv(Rho) => .K) ...</k> <env> _ => Rho </env>

bindTo, bind and assignTo

The meaning of these operations has already been explained when we
discussed the let and letrec language constructs
above.

  syntax KItem ::= bindTo(Names,Exps)         [strict(2)]
                 | bindMap(Map)
                 | bind(Names)

  rule (.K => getMatchingAux(Xs,Vs)) ~> bindTo(Xs:Names,Vs:Vals)
  rule matchResult(M:Map) ~> bindTo(_:Names, _:Vals) => bindMap(M)

  rule bindMap(.Map) => .K
  rule <k> bindMap((X:Name |-> V:Val => .Map) _:Map) ...</k>
       <env> Rho => Rho[X <- !L:Int] </env>
       <store>... .Map => !L |-> V ...</store>

  rule bind(.Names) => .K
  rule <k> bind(X:Name,Xs => Xs) ...</k>
       <env> Rho => Rho[X <- !_L:Int] </env>

  syntax KItem ::= assignTo(Names,Exps)  [strict(2)]

  rule <k> assignTo(.Names,.Vals) => .K ...</k>
  rule <k> assignTo((X:Name,Xs => Xs),(V:Val,Vs:Vals => Vs)) ...</k>
       <env>... X |-> L ...</env>
       <store>... .Map => L |-> V ...</store>

Getters

The following auxiliary operations extract the list of identifiers
and of expressions in a binding, respectively.

  syntax Names ::= names(Bindings)  [function]
  rule names(.Bindings) => .Names
  rule names(X:Name=_ and Bs) => (X,names(Bs))::Names

  syntax Exps ::= exps(Bindings)  [function]
  rule exps(.Bindings) => .Exps
  rule exps(_:Name=E and Bs) => E,exps(Bs)

  /* Extra kore stuff */
  syntax KResult ::= Vals
  syntax Exps ::= Names
  syntax Names ::= Bottoms

  /* Matching */
  syntax MatchResult ::= getMatching(Exp, Val)                      [function]
                       | getMatchingAux(Exps, Vals)                 [function]
                       | mergeMatching(MatchResult, MatchResult)    [function]
                       | matchResult(Map)
                       | "matchFailure"

  rule getMatching(C:ConstructorName(Es:Exps), C(Vs:Vals)) => getMatchingAux(Es, Vs)
  rule getMatching([Es:Exps], [Vs:Vals])                   => getMatchingAux(Es, Vs)
  rule getMatching(C:ConstructorName, C) => matchResult(.Map)
  rule getMatching(B:Bool, B)            => matchResult(.Map)
  rule getMatching(I:Int, I)             => matchResult(.Map)
  rule getMatching(S:String, S)          => matchResult(.Map)
  rule getMatching(N:Name, V:Val) => matchResult(N |-> V)
  rule getMatching(_, _) => matchFailure        [owise]

  rule getMatchingAux((E:Exp, Es:Exps), (V:Val, Vs:Vals)) => mergeMatching(getMatching(E, V), getMatchingAux(Es, Vs))
  rule getMatchingAux(.Exps, .Vals)                       => matchResult(.Map)
  rule getMatchingAux(_, _) => matchFailure     [owise]

  rule mergeMatching(matchResult(M1:Map), matchResult(M2:Map)) => matchResult(M1 M2)
    requires intersectSet(keys(M1), keys(M2)) ==K .Set
  //rule mergeMatching(_, _) => matchFailure      [owsie]
  rule mergeMatching(matchResult(_:Map), matchFailure) => matchFailure
  rule mergeMatching(matchFailure, matchResult(_:Map)) => matchFailure
  rule mergeMatching(matchFailure, matchFailure)       => matchFailure

Besides the generic decomposition rules for patterns and values,
we also want to allow [head|tail] matching for lists, so we add
the following custom pattern decomposition rule:

  rule getMatching([H:Exp | T:Exp], [V:Val, Vs:Vals])
    => getMatchingAux((H, T), (V, [Vs]))
endmodule

Go to Lesson 2, FUN untyped, Substitution-Based.

// NOTE: this definition is not up to date with the latest version of K, as it
// uses both substitution and symbolic reasoning.
// It is intended for documentation and academic purposes only.

FUN — Untyped — Substitution

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the substitution-based definition of FUN. For additional
explanations regarding the semantics of the various FUN constructs,
the reader should consult the emvironment-based definition of FUN.

Syntax

requires "substitution.md"
//requires "modules/pattern-matching.k"

module FUN-UNTYPED-COMMON
  imports DOMAINS-SYNTAX

The Syntactic Constructs

  syntax Name
  syntax Names ::= List{Name,","}

  syntax Exp ::= Int | Bool | String | Name
               | "(" Exp ")"                       [bracket]
  syntax Exps  ::= List{Exp,","}                   [strict]
  syntax Val
  syntax Vals ::= List{Val,","}

  syntax Exp ::= left:
                 Exp "*" Exp                       [strict, arith]
               | Exp "/" Exp                       [strict, arith]
               | Exp "%" Exp                       [strict, arith]
               > left:
                 Exp "+" Exp                       [strict, left, arith]
               | Exp "^" Exp                       [strict, left, arith]
               | Exp "-" Exp                       [strict, prefer, arith]
               | "-" Exp                           [strict, arith]
               > non-assoc:
                 Exp "<" Exp                       [strict, arith]
               | Exp "<=" Exp                      [strict, arith]
               | Exp ">" Exp                       [strict, arith]
               | Exp ">=" Exp                      [strict, arith]
               | Exp "==" Exp                      [strict, arith]
               | Exp "!=" Exp                      [strict, arith]
               > "!" Exp                           [strict, arith]
               > Exp "&&" Exp                      [strict(1), left, arith]
               > Exp "||" Exp                      [strict(1), left, arith]

  syntax Exp ::= "if" Exp "then" Exp "else" Exp    [strict(1)]

  syntax Exp ::= "[" Exps "]"                      [strict]
               | "cons" |  "head" | "tail" | "null?"
               | "[" Exps "|" Exp "]"
  syntax Val ::= "[" Vals "]"

  syntax ConstructorName
  syntax Exp ::= ConstructorName
               | ConstructorName "(" Exps ")"      [prefer, strict(2)]
  syntax Val ::= ConstructorName "(" Vals ")"

  syntax Exp ::= "fun" Cases
               | Exp Exp                           [strict, left]
  syntax Case  ::= Exp "->" Exp                    [binder]
// NOTE: The binder attribute above is the only difference between this
// module and the syntax module of environment-based FUN.  We need
// to fix a bug in order to import modules and override the attributes
// of operations.
  syntax Cases ::= List{Case, "|"}

  syntax Exp ::= "let" Bindings "in" Exp
               | "letrec" Bindings "in" Exp                 [prefer]
  syntax Binding  ::= Exp "=" Exp
  syntax Bindings ::= List{Binding,"and"}

  syntax Exp ::= "ref"
               | "&" Name
               | "@" Exp                           [strict]
               | Exp ":=" Exp                      [strict]
               | Exp ";" Exp                       [strict(1), right]

  syntax Exp ::= "callcc"
               | "try" Exp "catch" "(" Name ")" Exp
  syntax Name ::= "throw" [token]

  syntax Exp ::= "datatype" Type "=" TypeCases Exp

  syntax TypeVar
  syntax TypeVars ::= List{TypeVar,","}

  syntax TypeName
  syntax Type ::= "int" | "bool" | "string"
                | Type "-->" Type                            [right]
                | "(" Type ")"                             [bracket]
                | TypeVar
                | TypeName             [klabel(TypeName), avoid]
                | Type TypeName   [klabel(Type-TypeName), onlyLabel]
                | "(" Types ")" TypeName                    [prefer]
  syntax Types ::= List{Type,","}
  syntax Types ::= TypeVars

  syntax TypeCase ::= ConstructorName
                    | ConstructorName "(" Types ")"
  syntax TypeCases ::= List{TypeCase,"|"}     [klabel(_|TypeCase_)]

Additional Priorities

  syntax priority @__FUN-UNTYPED-COMMON
                > ___FUN-UNTYPED-COMMON
                > arith
                > _:=__FUN-UNTYPED-COMMON
                > let_in__FUN-UNTYPED-COMMON
                  letrec_in__FUN-UNTYPED-COMMON
                  if_then_else__FUN-UNTYPED-COMMON
                > _;__FUN-UNTYPED-COMMON
                > fun__FUN-UNTYPED-COMMON
                > datatype_=___FUN-UNTYPED-COMMON
endmodule

module FUN-UNTYPED-MACROS
  imports FUN-UNTYPED-COMMON

Desugaring macros

  rule P1 P2 -> E => P1 -> fun P2 -> E                       [macro-rec]
  rule F P = E => F = fun P -> E                             [macro-rec]

  rule [E1,E2,Es:Exps|T] => [E1|[E2,Es|T]]                   [macro-rec]

//  rule 'TypeName(Tn:TypeName) => (.TypeVars) Tn              [macro]
  rule `Type-TypeName`(T:Type, Tn:TypeName) => (T) Tn          [macro]

  syntax Name ::= "$h" | "$t"
  rule head => fun [$h|$t] -> $h                             [macro]
  rule tail => fun [$h|$t] -> $t                             [macro]
  rule null? => fun [.Exps] -> true | [$h|$t] -> false       [macro]

  syntax Name ::= "$k" | "$v"
  rule try E catch(X) E'
    => callcc (fun $k -> (fun throw -> E)(fun X -> $k E'))   [macro]

  rule datatype _T = _TCs E => E                               [macro]

mu needed for letrec, but we put it here so we can also write
programs with mu in them, which is particularly useful for testing.

  syntax Exp ::= "mu" Case

endmodule


module FUN-UNTYPED-SYNTAX
  imports FUN-UNTYPED-COMMON
  imports BUILTIN-ID-TOKENS

  syntax Name ::= r"[a-z][_a-zA-Z0-9]*"            [token, prec(2)]
                | #LowerId                         [token]
  syntax ConstructorName ::= #UpperId              [token]
  syntax TypeVar  ::= r"['][a-z][_a-zA-Z0-9]*"     [token]
  syntax TypeName ::= Name                         [token]
endmodule

Semantics

module FUN-UNTYPED
  imports FUN-UNTYPED-COMMON
  imports FUN-UNTYPED-MACROS
  imports DOMAINS
  imports SUBSTITUTION
  //imports PATTERN-MATCHING

  configuration <T color="yellow">
                  <k color="green"> $PGM:Exp </k>
                  <store color="white"> .Map </store>
                </T>

Both Name and functions are values now:

  syntax Val ::= Int | Bool | String | Name
  syntax Exp ::= Val
  syntax Exps ::= Vals
  syntax KResult ::= Val
  syntax Exps ::= Names
  syntax Vals ::= Names

  rule I1 * I2 => I1 *Int I2
  rule I1 / I2 => I1 /Int I2 when I2 =/=K 0
  rule I1 % I2 => I1 %Int I2 when I2 =/=K 0
  rule I1 + I2 => I1 +Int I2
  rule S1 ^ S2 => S1 +String S2
  rule I1 - I2 => I1 -Int I2
  rule - I => 0 -Int I
  rule I1 < I2 => I1 <Int I2
  rule I1 <= I2 => I1 <=Int I2
  rule I1 > I2 => I1 >Int I2
  rule I1 >= I2 => I1 >=Int I2
  rule V1:Val == V2:Val => V1 ==K V2
  rule V1:Val != V2:Val => V1 =/=K V2
  rule ! T => notBool(T)
  rule true  && E => E
  rule false && _ => false
  rule true  || _ => true
  rule false || E => E

  rule if  true then E else _ => E
  rule if false then _ else E => E

  rule isVal(cons) => true
  rule isVal(cons _V:Val) => true
  rule cons V:Val [Vs:Vals] => [V,Vs]

  syntax Val ::= ConstructorName

  rule isVal(fun _) => true
  syntax KVar ::= Name
  syntax Name ::= freshName(Int)    [freshGenerator, function]
  rule freshName(I:Int) => {#parseToken("Name", "#" +String Int2String(I))}:>Name

  rule (. => getMatching(P, V)) ~> (fun P->_ | _) V:Val
  rule matchResult(M:Map) ~> (fun _->E | _) _ => E[M]
  rule (matchFailure => .) ~> (fun (_->_ | Cs:Cases => Cs)) _
//  rule (fun P->E | _) V:Val => E[getMatching(P,V)]  when isMatching(P,V)
//  rule (fun (P->_ | Cs:Cases => Cs)) V:Val  when notBool isMatching(P,V)

We can reduce multiple bindings to one list binding, and then
apply the usual desugaring of let into function application.
It is important that the rule below is a macro, so let is eliminated
immediately, otherwise it may interfere in ugly ways with substitution.

  rule let Bs in E => ((fun [names(Bs)] -> E) [exps(Bs)])    [macro]

We only give the semantics of one-binding letrec.
Multipe bindings are left as an exercise.

  // changed because of parsing error
  //rule mu X:Name -> E => E[(mu X -> E) / X]
  rule mu X:Name -> E => E[X |-> (mu X -> E)]
  rule letrec F:Name = E in E' => let F = (mu F -> E) in E'  [macro]

We cannot have & anymore, but we can give direct
semantics to ref. We also have to declare ref to
be a value, so that we will never heat on it.

//  rule <k> & X => L ...</k>  <env>... X |-> L </env>
  rule isVal(ref) => true
  rule <k> ref V:Val => !L:Int ...</k> <store>... .Map => !L |-> V ...</store>
  rule <k> @ L:Int => V:Val ...</k>  <store>... L |-> V ...</store>
  rule <k> L:Int := V:Val => V ...</k>  <store>... L |-> (_=>V) ...</store>
  rule _V:Val; E => E

  syntax Val ::= cc(K)
  rule isVal(callcc) => true
  rule <k> (callcc V:Val => V cc(K)) ~> K </k>
  rule <k> cc(K) V:Val ~> _ => V ~> K </k>

Auxiliary getters

  syntax Names ::= names(Bindings)  [function]
  rule names(.Bindings) => .Names
  rule names(X:Name=_ and Bs) => X,names(Bs)

  syntax Exps ::= exps(Bindings)  [function]
  rule exps(.Bindings) => .Exps
  rule exps(_:Name=E and Bs) => E,exps(Bs)

  /* Extra kore stuff */
  syntax KResult ::= Vals
  syntax Exps ::= Names

  /* Matching */
  syntax MatchResult ::= getMatching(Exp, Val)                      [function]
                       | getMatchingAux(Exps, Vals)                 [function]
                       | mergeMatching(MatchResult, MatchResult)    [function]
                       | matchResult(Map)
                       | "matchFailure"

  rule getMatching(C:ConstructorName(Es:Exps), C(Vs:Vals)) => getMatchingAux(Es, Vs)
  rule getMatching([Es:Exps], [Vs:Vals])                   => getMatchingAux(Es, Vs)
  rule getMatching(C:ConstructorName, C) => matchResult(.Map)
  rule getMatching(B:Bool, B)            => matchResult(.Map)
  rule getMatching(I:Int, I)             => matchResult(.Map)
  rule getMatching(S:String, S)          => matchResult(.Map)
  rule getMatching(N:Name, V:Val) => matchResult(N |-> V)
  rule getMatching(_, _) => matchFailure        [owise]

  rule getMatchingAux((E:Exp, Es:Exps), (V:Val, Vs:Vals)) => mergeMatching(getMatching(E, V), getMatchingAux(Es, Vs))
  rule getMatchingAux(.Exps, .Vals)                       => matchResult(.Map)
  rule getMatchingAux(_, _) => matchFailure     [owise]

  rule mergeMatching(matchResult(M1:Map), matchResult(M2:Map)) => matchResult(M1 M2)
    requires intersectSet(keys(M1), keys(M2)) ==K .Set
  //rule mergeMatching(_, _) => matchFailure      [owsie]
  rule mergeMatching(matchResult(_:Map), matchFailure) => matchFailure
  rule mergeMatching(matchFailure, matchResult(_:Map)) => matchFailure
  rule mergeMatching(matchFailure, matchFailure)       => matchFailure

Besides the generic decomposition rules for patterns and values,
we also want to allow [head|tail] matching for lists, so we add
the following custom pattern decomposition rule:

  rule getMatching([H:Exp | T:Exp], [V:Val, Vs:Vals])
    => getMatchingAux((H, T), (V, [Vs]))
endmodule

// NOTE: this definition is not runnable as is.
// It is intended for documentation and academic purposes only.

LOGIK

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K semantic definition of LOGIK, a trivial language
capturing the essence of the logic programming paradigm. In this
definition, we explicitly focus on simplicity and mathematical
clarity, not on advanced logic programming features or performance.
Those are covered in the LOGIK++ extension under examples/logik++.

Specifically, a LOGIK program consists of a sequence of Horn clauses
of the form

P :- P1, P2, ..., Pn .

followed by a query of the form

?- Q1, Q2, ..., Qm .

where P, P1, P2, ..., Pn, Q1, Q2,
..., Qm are literals. The
symbol :- is read "if". A literal has the form
p(T1,T2,...,Tk), where p is a predicate symbol
and where T1,T2,...,Tk are terms. Terms are built as
usual, with operation symbols and variables. A common
convention in logic programming languages, also adopted here, is that
variables are capitalized and operation symbols are not. Operations
with zero arguments are called constants and are written without
parentheses, that is, c instead of c(). Horn
clauses without conditions, called facts, are written
without :-, that is, P. instead of P :- ..

For example, the LOGIK program below gives a few facts about a
parent predicate, then several clauses defining some useful
predicates including an ancestor predicate, and finally a
query asking for those who both have ancestors and are ancestors
themselves in the parent relation:

parent(david,john).
parent(jim,david).
parent(steve,jim).
parent(nathan,steve).

grandparent(A,B):-
  parent(A,X),
  parent(X,B).

ancestor(A,B):-
  parent(A,X),
  parents(X,B).

parents(X,X).
parents(A,B):-
  ancestor(A,B).

both(X) :- ancestor(A,X), ancestor(X,B).

?- both(X).

Above, we only have constant operation symbols, so these and variables
are the only terms that can be used in predicates. As expected, the
LOGIK program above will give us three solutions for X:
david, steve, and jim. If we inline the
both(X) predicate in the query, that is, if we replace the
query with ?- ancestor(A,X), ancestor(X,B). then we get
10 solutions, one for for each triple A, X, and
B satisfying both predicates ancestor(A,X) and
ancestor(X,B).

As another example, the program below defines an append
predicate followed by a simple goal:

append(nil,L,L).
append(cons(H,T),L,cons(H,Z)) :- append(T,L,Z).

?- append(cons(a,nil), cons(b,nil), V).

Besides the predicate symbol append, the program above also
includes a constant symbol nil and a binary operation symbol
cons. Additionally, the query also includes two more
constants, a and b. The capitalized identifiers are
all variables. As expected, the LOGIK program above yields only one
solution, namely V = cons(a,cons(b,nil)). On the other hand,
if we change the query to:

?- append(L1, cons(a,L2), cons(a,cons(b,cons(a,nil)))).

then LOGIK yields two solutions: one where L1 is
cons(a,cons(b,nil)) and L2 is nil,
and another where L1 is nil and L2 is
cons(a,cons(b,nil)).

The programs above all generated ground solutions, that is,
solutions where the query variables are mapped to ground terms (i.e.,
terms without variables). Let us now consider the following query:

?- append(cons(a,nil), Y, Z).

There are obviously infinitely many ground solutions for the query
above, e.g.,
Y = nil and Z = cons(a,nil),
Y = cons(a,nil) and Z = cons(a,cons(a,nil)),
Y = cons(b,nil) and Z = cons(a,cons(b,nil)),
Y = cons(c,cons(b,nil)) and Z = cons(a,cons(c,cons(b,nil))),
etc. However, all the ground solutions for the query above can be
elegantly characterized by the property that Z is bound to a list
starting with a and followed by the list that Y is
bound to. This property can in fact be described as a symbolic solution
to the query: Z = cons(a,Y) or, equivalently,
Y = Symb and Z = cons(a,Symb). It is possible to
define a ``more general than'' relation on such symbolic solutions,
in the sense that the more particular solution can be obtained as a
specialization/substitution of the more general one, and then it can
be shown that the above is the most general solution to the
stated query. Logic programming languages, including our LOGIK,
attempt to always compute such most general solutions.

Logic programming languages are highly non-deterministic, in that
several Horn clauses may be used at the same time, each possibly
resulting in a different solution. Implementations of logic
programming languages consist of complex, optimized search and
indexing algorithms, which we are not concerned with here. Instead,
we here take advantage of K's builtin support for search.
Specifically, to find all the solutions of a LOGIK program, we have to
use krun with the option --search. However, note
that some programs have infinitely many solutions which cannot relate
to each other by the "more general" relation. For example, the query

?- append(L1, cons(a,L2), L3) .

To address such cases and terminate, logic programming languages allow
the user to choose how many solutions to be computed and displayed.
In LOGIK, we can use the --bound option of krun for
this purpose.

Finally, note that some queries have no solution. In some cases that
is easy to detect by exhaustive analysis, such as for the following
query:

?- append(cons(a,L1), L2, cons(b,L3)).

Logic programming languages, including LOGIK, terminate in such cases
and report a no solution answer. However, there are cases where
exhaustive analysis is not sufficient, such as for the query:

?- append(cons(a,L), nil, L).

In such cases, logic programming languages do not terminate. While
one may devise techniques to detect non-termination in some cases,
one cannot do it in general (same like for all Turing-complete
languages).

requires "unification.k"

module LOGIK-COMMON
  imports DOMAINS-SYNTAX

Syntax

The syntax of LOGIK is straightforward: a program is a sequence of
Horn clauses followed by a query:

  syntax Literal
  syntax Term ::= Literal | Literal "(" Terms ")"
  syntax Terms ::= List{Term,","}
  syntax Clause ::= Term ":-" Terms "." | Term "."
  syntax Query ::= "?-" Terms "."
  syntax Pgm ::= Query | Clause Pgm
endmodule

module LOGIK-SYNTAX
  imports LOGIK-COMMON
  imports BUILTIN-ID-TOKENS

Variables and literals are defined as tokens following the conventions
used in Prolog (variables start with _ or capital letter, while literals
start with lower case letters):

  syntax #KVariable ::= r"[A-Z_][A-Za-z0-9_]*"   [token, prec(2)]
                      | #UpperId                 [token]
  syntax Term ::= #KVariable [klabel(#SemanticCastToTerm)]
  syntax Literal ::= r"[a-z][a-zA-Z0-9_]*"                 [token]
                   | #LowerId                              [token]
endmodule

module LOGIK
  imports LOGIK-COMMON
  imports DOMAINS
  imports UNIFICATION

Unification is at the core of logic programming. Here we are
going to use the predefined unification procedure (the same one we
used in the type inferencers in Tutorial 5).

Configuration

The configuration stores each clause in its own cell for easy access,
and the most general unifier in a cell named mgu, same like
the type inferencers. The k cell holds the query and the
fresh cell holds a fresh clause instance to be attempted on
the next query item. To more easily read the solutions, we add a
second top-level cell, solution. Both top cells are
optional. Indeed, we start with the main top cell and, when a
solution is found, we move it into the solution cell and
discard the main cell.

  configuration <T color="yellow" multiplicity="?">
                  <k color="green"> $PGM:Pgm </k>
                  <fresh color="orange"> .K </fresh>
                  <clauses color="red">
                    <clause color="pink" multiplicity="*"> .K </clause>
                  </clauses>
                  <mgu> .K </mgu>
                </T>
                <solution multiplicity="?"> .K </solution>

Pre- and post-processing

Before we launch the semantics, we first scan the given program and
place each clause in its own cell, and then place the query in the
k cell and initialize the mgu with the variables from the query.

Note that we put a fresh instance of the clause to avoid interference with
the query variables. By a "fresh instance" of a clause we mean one whose
variables are renamed with fresh names; we need that in order to avoid
undesired unification conflicts due to particular names chosen for
variables in the original program, as well as conflicts due to
subsequent uses of the same clause. It is safe to rename the
variables in a clause, because clauses are universally quantified in
their variables. This process of creating a fresh instance of a
clause is similar to how we created fresh instances of type schemas in
the higher-order type inferencer discussed in Tutorial 5. Indeed, we
can safely regard clauses as "clause schemas" comprising infinitely
many instances, one for each context.

  rule <k> C:Clause Pgm => Pgm </k>
       (.Bag => <clause> #renameVariables(C) </clause>)

  rule <k> ?- Ls:Terms. => Ls ...</k>
       <mgu> _ => #variablesMap(#variables(Ls)) </mgu>

We also sequentialize the goals for easier processing:

  rule L:Term, Ls:Terms => L ~> Ls
  rule .Terms => .

When all the goals are solved, indicated by the empty k
cell, the calculated most general unifier (mgu) is in the mgu
cell. In that case, to ease reading of the final solution we move the
mgu in the solution cell and delete the rest of the
configuration:

rule <T>... <k> . </k> <mgu> Theta </mgu> ...</T>
  => <solution> Theta </solution>

Since we are not interested in seeing the failed attempts to solve
the query, we collapse all the error configurations into an empty
configuration (recall that both top-level cells in the configuration
were declared optional). This way, if we see an empty configuration
when we search for all solutions, we know that some attempts failed
(but we do not know which ones).

// this would be nice, but we need feedback from the external unifier
// for this.
//  rule <T>... <mgu> _:MguError </mgu> ...</T> => .

Semantics

Once all the infrastructure is in place, the actual semantics of LOGIK
is quite simple. All we have to do is to pick some (fresh instance of
a) clause, then unify its conclusion with the first query literal, and
then replace that literal with condition of the clause. The intuition
here is the following: to satisfy the first literal in the query, we
need to find some instance of some clause that matches it, and then to
similarly show that we can satisfy the conditions of that clause.
Mathematically, this is an instance of the proof principle called
resolution: if p ∨ q and ¬ p ∨ r hold, then so does
q ∨ r. We let it as an exercise to the reader to see how the two
relate (hint: assume the negation of the goal together with all the
clauses, and then derive false).

The following two rules are tightly connected and they together
perform the following core task: pick a fresh instance of a clause
which unifies with the first goal item, then add its conditions as new
goals.

Pick a clause and generate a fresh instance of it when the
fresh cell is empty:

  rule <fresh> . => #renameVariables(C) </fresh> <clause> C </clause>
       <k> T:Term ...</k>
  requires #unifiable(T,head(C))

  syntax Term ::= head(Clause) [function]
  rule head(L.) => L
  rule head(L:-_.) => L

If the goal is unifiable with the fresh clause's head, replace the goal
with the clause body, and empty the fresh cell (so that
another clause can be chosen using the rule above):

  rule <k> L:Term => . ...</k>
       <fresh>  L:Term . => . </fresh>

  rule <k> L:Term :KItem => Ls ...</k>
       <fresh>  L:Term :- Ls:Terms. => . </fresh>

Note that there is no problem if a clause is chosen whose
conclusion literal does not unify with the first goal literal.
The search
option of krun will systematically try all clauses, so no
solution is missed. Of course, the above is not the most efficient
way to implement a logic programming language, but recall that our
objective here was to present a simple and mathematically clean
solution. We encourage the interested reader to consult the LOGIK++
language definition for a more efficient definition of a richer logic
programming language.

endmodule

// NOTE: this definition is not runnable as is.
// It is intended for documentation and academic purposes only.

LOGIK

Author: Grigore Roșu (grosu@illinois.edu)
Organization: University of Illinois at Urbana-Champaign

Author: Traian Florin Șerbănuță (traian.serbanuta@unibuc.ro)
Organization: University of Bucharest

Abstract

This is the K semantic definition of LOGIK, a trivial language
capturing the essence of the logic programming paradigm. In this
definition, we explicitly focus on simplicity and mathematical
clarity, not on advanced logic programming features or performance.
Those are covered in the LOGIK++ extension under examples/logik++.

Specifically, a LOGIK program consists of a sequence of Horn clauses
of the form

P :- P1, P2, ..., Pn .

followed by a query of the form

?- Q1, Q2, ..., Qm .

where P, P1, P2, ..., Pn, Q1, Q2,
..., Qm are literals. The
symbol :- is read "if". A literal has the form
p(T1,T2,...,Tk), where p is a predicate symbol
and where T1,T2,...,Tk are terms. Terms are built as
usual, with operation symbols and variables. A common
convention in logic programming languages, also adopted here, is that
variables are capitalized and operation symbols are not. Operations
with zero arguments are called constants and are written without
parentheses, that is, c instead of c(). Horn
clauses without conditions, called facts, are written
without :-, that is, P. instead of P :- ..

For example, the LOGIK program below gives a few facts about a
parent predicate, then several clauses defining some useful
predicates including an ancestor predicate, and finally a
query asking for those who both have ancestors and are ancestors
themselves in the parent relation:

parent(david,john).
parent(jim,david).
parent(steve,jim).
parent(nathan,steve).

grandparent(A,B):-
  parent(A,X),
  parent(X,B).

ancestor(A,B):-
  parent(A,X),
  parents(X,B).

parents(X,X).
parents(A,B):-
  ancestor(A,B).

both(X) :- ancestor(A,X), ancestor(X,B).

?- both(X).

Above, we only have constant operation symbols, so these and variables
are the only terms that can be used in predicates. As expected, the
LOGIK program above will give us three solutions for X:
david, steve, and jim. If we inline the
both(X) predicate in the query, that is, if we replace the
query with ?- ancestor(A,X), ancestor(X,B). then we get
10 solutions, one for for each triple A, X, and
B satisfying both predicates ancestor(A,X) and
ancestor(X,B).

As another example, the program below defines an append
predicate followed by a simple goal:

append(nil,L,L).
append(cons(H,T),L,cons(H,Z)) :- append(T,L,Z).

?- append(cons(a,nil), cons(b,nil), V).

Besides the predicate symbol append, the program above also
includes a constant symbol nil and a binary operation symbol
cons. Additionally, the query also includes two more
constants, a and b. The capitalized identifiers are
all variables. As expected, the LOGIK program above yields only one
solution, namely V = cons(a,cons(b,nil)). On the other hand,
if we change the query to:

?- append(L1, cons(a,L2), cons(a,cons(b,cons(a,nil)))).

then LOGIK yields two solutions: one where L1 is
cons(a,cons(b,nil)) and L2 is nil,
and another where L1 is nil and L2 is
cons(a,cons(b,nil)).

The programs above all generated ground solutions, that is,
solutions where the query variables are mapped to ground terms (i.e.,
terms without variables). Let us now consider the following query:

?- append(cons(a,nil), Y, Z).

There are obviously infinitely many ground solutions for the query
above, e.g.,
Y = nil and Z = cons(a,nil),
Y = cons(a,nil) and Z = cons(a,cons(a,nil)),
Y = cons(b,nil) and Z = cons(a,cons(b,nil)),
Y = cons(c,cons(b,nil)) and Z = cons(a,cons(c,cons(b,nil))),
etc. However, all the ground solutions for the query above can be
elegantly characterized by the property that Z is bound to a list
starting with a and followed by the list that Y is
bound to. This property can in fact be described as a symbolic solution
to the query: Z = cons(a,Y) or, equivalently,
Y = Symb and Z = cons(a,Symb). It is possible to
define a ``more general than'' relation on such symbolic solutions,
in the sense that the more particular solution can be obtained as a
specialization/substitution of the more general one, and then it can
be shown that the above is the most general solution to the
stated query. Logic programming languages, including our LOGIK,
attempt to always compute such most general solutions.

Logic programming languages are highly non-deterministic, in that
several Horn clauses may be used at the same time, each possibly
resulting in a different solution. Implementations of logic
programming languages consist of complex, optimized search and
indexing algorithms, which we are not concerned with here. Instead,
we here take advantage of K's builtin support for search.
Specifically, to find all the solutions of a LOGIK program, we have to
use krun with the option --search. However, note
that some programs have infinitely many solutions which cannot relate
to each other by the "more general" relation. For example, the query

?- append(L1, cons(a,L2), L3) .

To address such cases and terminate, logic programming languages allow
the user to choose how many solutions to be computed and displayed.
In LOGIK, we can use the --bound option of krun for
this purpose.

Finally, note that some queries have no solution. In some cases that
is easy to detect by exhaustive analysis, such as for the following
query:

?- append(cons(a,L1), L2, cons(b,L3)).

Logic programming languages, including LOGIK, terminate in such cases
and report a no solution answer. However, there are cases where
exhaustive analysis is not sufficient, such as for the query:

?- append(cons(a,L), nil, L).

In such cases, logic programming languages do not terminate. While
one may devise techniques to detect non-termination in some cases,
one cannot do it in general (same like for all Turing-complete
languages).

requires "unification.k"

module LOGIK-COMMON
  imports DOMAINS-SYNTAX

Syntax

The syntax of LOGIK is straightforward: a program is a sequence of
Horn clauses followed by a query:

  syntax Literal
  syntax Term ::= Literal | Literal "(" Terms ")"
  syntax Terms ::= List{Term,","}
  syntax Clause ::= Term ":-" Terms "." | Term "."
  syntax Query ::= "?-" Terms "."
  syntax Pgm ::= Query | Clause Pgm
endmodule

module LOGIK-SYNTAX
  imports LOGIK-COMMON
  imports BUILTIN-ID-TOKENS

Variables and literals are defined as tokens following the conventions
used in Prolog (variables start with _ or capital letter, while literals
start with lower case letters):

  syntax #KVariable ::= r"[A-Z_][A-Za-z0-9_]*"   [token, prec(2)]
                      | #UpperId                 [token]
  syntax Term ::= #KVariable [klabel(#SemanticCastToTerm)]
  syntax Literal ::= r"[a-z][a-zA-Z0-9_]*"                 [token]
                   | #LowerId                              [token]
endmodule

module LOGIK
  imports LOGIK-COMMON
  imports DOMAINS
  imports UNIFICATION

Unification is at the core of logic programming. Here we are
going to use the predefined unification procedure (the same one we
used in the type inferencers in Tutorial 5).

Configuration

The configuration stores each clause in its own cell for easy access,
and the most general unifier in a cell named mgu, same like
the type inferencers. The k cell holds the query and the
fresh cell holds a fresh clause instance to be attempted on
the next query item. To more easily read the solutions, we add a
second top-level cell, solution. Both top cells are
optional. Indeed, we start with the main top cell and, when a
solution is found, we move it into the solution cell and
discard the main cell.

  configuration <T color="yellow" multiplicity="?">
                  <k color="green"> $PGM:Pgm </k>
                  <fresh color="orange"> .K </fresh>
                  <clauses color="red">
                    <clause color="pink" multiplicity="*"> .K </clause>
                  </clauses>
                  <mgu> .K </mgu>
                </T>
                <solution multiplicity="?"> .K </solution>

Pre- and post-processing

Before we launch the semantics, we first scan the given program and
place each clause in its own cell, and then place the query in the
k cell and initialize the mgu with the variables from the query.

Note that we put a fresh instance of the clause to avoid interference with
the query variables. By a "fresh instance" of a clause we mean one whose
variables are renamed with fresh names; we need that in order to avoid
undesired unification conflicts due to particular names chosen for
variables in the original program, as well as conflicts due to
subsequent uses of the same clause. It is safe to rename the
variables in a clause, because clauses are universally quantified in
their variables. This process of creating a fresh instance of a
clause is similar to how we created fresh instances of type schemas in
the higher-order type inferencer discussed in Tutorial 5. Indeed, we
can safely regard clauses as "clause schemas" comprising infinitely
many instances, one for each context.

  rule <k> C:Clause Pgm => Pgm </k>
       (.Bag => <clause> #renameVariables(C) </clause>)

  rule <k> ?- Ls:Terms. => Ls ...</k>
       <mgu> _ => #variablesMap(#variables(Ls)) </mgu>

We also sequentialize the goals for easier processing:

  rule L:Term, Ls:Terms => L ~> Ls
  rule .Terms => .

When all the goals are solved, indicated by the empty k
cell, the calculated most general unifier (mgu) is in the mgu
cell. In that case, to ease reading of the final solution we move the
mgu in the solution cell and delete the rest of the
configuration:

rule <T>... <k> . </k> <mgu> Theta </mgu> ...</T>
  => <solution> Theta </solution>

Since we are not interested in seeing the failed attempts to solve
the query, we collapse all the error configurations into an empty
configuration (recall that both top-level cells in the configuration
were declared optional). This way, if we see an empty configuration
when we search for all solutions, we know that some attempts failed
(but we do not know which ones).

// this would be nice, but we need feedback from the external unifier
// for this.
//  rule <T>... <mgu> _:MguError </mgu> ...</T> => .

Semantics

Once all the infrastructure is in place, the actual semantics of LOGIK
is quite simple. All we have to do is to pick some (fresh instance of
a) clause, then unify its conclusion with the first query literal, and
then replace that literal with condition of the clause. The intuition
here is the following: to satisfy the first literal in the query, we
need to find some instance of some clause that matches it, and then to
similarly show that we can satisfy the conditions of that clause.
Mathematically, this is an instance of the proof principle called
resolution: if p ∨ q and ¬ p ∨ r hold, then so does
q ∨ r. We let it as an exercise to the reader to see how the two
relate (hint: assume the negation of the goal together with all the
clauses, and then derive false).

The following two rules are tightly connected and they together
perform the following core task: pick a fresh instance of a clause
which unifies with the first goal item, then add its conditions as new
goals.

Pick a clause and generate a fresh instance of it when the
fresh cell is empty:

  rule <fresh> . => #renameVariables(C) </fresh> <clause> C </clause>
       <k> T:Term ...</k>
  requires #unifiable(T,head(C))

  syntax Term ::= head(Clause) [function]
  rule head(L.) => L
  rule head(L:-_.) => L

If the goal is unifiable with the fresh clause's head, replace the goal
with the clause body, and empty the fresh cell (so that
another clause can be chosen using the rule above):

  rule <k> L:Term => . ...</k>
       <fresh>  L:Term . => . </fresh>

  rule <k> L:Term :KItem => Ls ...</k>
       <fresh>  L:Term :- Ls:Terms. => . </fresh>

Note that there is no problem if a clause is chosen whose
conclusion literal does not unify with the first goal literal.
The search
option of krun will systematically try all clauses, so no
solution is missed. Of course, the above is not the most efficient
way to implement a logic programming language, but recall that our
objective here was to present a simple and mathematically clean
solution. We encourage the interested reader to consult the LOGIK++
language definition for a more efficient definition of a richer logic
programming language.

endmodule

Projects using K

A list of projects using the K framework. If you are working on something interesting, and you want to share it with the community,
let us know on our socials, and we will feature you on this list.



  • KAVM (Feb 2022 - Present)

    The Algorand Virtual Machine and TEAL Semantics in K
    KAVM leverages the K Framework to empower Algorand smart contracts' developers
    with property-based testing and formal verification.

  • KPlutus (2016 - Present)

    The K Semantics of Plutus-Core

  • Dedukti (Mar 2021 - Present)

    This project aims to translate real K semantics into Dedukti.

  • KWasm (Aug 2015 - Present)

    KWasm is the K semantics of WebAssembly.
    WebAssembly is a low-level (but simple and streamlined) assembly language that was originally developed to provide a fast execution engine for browser-based tools.
    More recently, it has been used in several blockchain smart-contract platforms as the underlying language for executing financial agreements.
    KWasm has been used for measuring coverage of test-suites over Wasm code and verifying programs which are compiled to Wasm.

  • KEVM (Sep 2017 - Present)

    KEVM is the K semantics of the Ethereum Virtual Machine.
    It passes all the Ethereum Test Suite, and is used for verifying EVM programs.

  • IELE (Oct 2016 - Present)

    IELE is the underlying VM integrated into the Cardano blockchain.
    IELE is a register-based VM (inspired by LLVM), which attempts to avoid many of the missteps in design present in EVM.

  • K-Michelson (Oct 2019 - Present)

    K-Michelson is the K semantics of Michelson blockchain programming language, which powers the Tezos blockchain.
    KMichelson provides additional testing tools for developers, including a unit-testing framework which is extendable to symbolic property testing.

  • C (Jul 2010 - Present)

    The K semantics of the C programming language specifies the translation, linking, and execution semantics of the C language according to the official C standard.
    It has been used to build tools like RV-Match, which detects undefined behaviors in users programs by running their test-suites through the C semantics.

Archived