Skip to content

aucker/jlox

Folders and files

NameName
Last commit message
Last commit date

Latest commit

 

History

44 Commits
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 
 

Repository files navigation

jlox

Learning how to implement an interpreter🔨 from scratch.

There are several parts in Jlox, I don't know how to write the makefile to run Java project right now, so I use Android Studio.

Design NOTE: Spoonfuls of Syntactic Sugar

On the extreme acrid end are those with ruthlessly minimal syntax like Lisp, Forth, and Smalltalk. Lispers famously claim their language "has no syntax", while Smalltalkers proudly show that you can fit the entire grammar on an index card. This tribe has the philosophy that the language doesn't need syntactic sugar. Instead, the minimal syntax and semantics it provides are powerful enough to let library code be as expressive as if it were part of the language itself.

Near these are languages like C, Lua, and Go. They aim for simplicity and clarity over minimalism. Some, like Go, deliberately eschew both syntactic sugar and the kind of syntactic extensibility of the previous category. They want the syntax to get out of the way of the semantics, so they focus on keeping both the grammar and libraries simple. Code should be obvious more than beautiful.

Somewhere in the middle you have languages like Java, C#, and Python. Eventually you reach Ruby, C++, Perl, and D - Languages which have stuffed so much syntax into their grammar, they are running out of punctuation characters on the keyboard.

To some degree, location on the spectrum correlates with age. It's relatively easy to add bits of syntactic sugar in later releases. New syntax is a crowd pleaser, and it's less likely to break existing programs than mucking with the semantics. Once added, you can never take it away, so languages tend to sweeten with time. One of the main benefits of creating a new language from scratch is it gives you an opportunity to scrape off those accumulated layers of frosting and start over.

Syntactic sugar has a bad rap among the PL intelligentsia. There's a real fetish for minimalism in that crowd. There is some justification for that. Poorly designed, unneeded syntax raises the cognitive load w/o adding enough expressiveness to carry its weight. Since there is always pressure to cram new features into the language, it takes discipline and a focus on simplicity to avoid bloat. Once you add some syntax, you're stuck with it, so it's smart to be parsimonious.

At the same time, the most successful languages do have fairly complex grammars, at least by the time they are widely used. Programmers spend a ton of time in their language of choice, and a few niceties here and there really can improve the comfort and efficiency of their work.

Striking the right balance -- choose the right level of sweetness for your language -- relies on your own sense of taste.


When running the scan section code, the interpreter can recognize certain characters. Like !, ( etc.

Just like the following:

> 
+ - * !
PLUS + null
MINUS - null
STAR * null
BANG ! null
EOF  null
> 
can you do this () ? int main() { println() };
IDENTIFIER can null
IDENTIFIER you null
IDENTIFIER do null
THIS this null
LEFT_PAREN ( null
RIGHT_PAREN ) null
IDENTIFIER int null
IDENTIFIER main null
LEFT_PAREN ( null
RIGHT_PAREN ) null
LEFT_BRACE { null
IDENTIFIER println null
LEFT_PAREN ( null
RIGHT_PAREN ) null
RIGHT_BRACE } null
SEMICOLON ; null
EOF  null

Representing Code

Before we do all that, let's focus on the main goal - a representation for code. It should be simple for the parser to produce and easy for the interpreter to consume. 1 + 2 * 3 - 4, you know the multiplication is evaluated before the addition or subtraction. One way to visualize that precedence is using a tree. Leaf nodes are numbers, and interior nodes are operators with the branches for each of their operands.

In order to evaluate an arithmetic node, you need to know the numeric values of its subtrees, so you have to evaluate those first. That means working your way from the leaves up to the root - post-order traversal: post-order traversal

  1. Starting with the full tree, evaluate the bottom-most operation, 2 * 3
  2. Now we can evaluate the +
  3. Next, the -
  4. The final answer.

We need to get more precise about what that grammar is then. Like lexical grammars in the last chap. , there is a long ton of theory around syntactic grammars. We start by moving one level up the Chomsky hierarchy...

Context-Free Grammars

In the last chapter, the formalism we used for defining the lexical grammar - the rules for how characters get grouped into tokens - was called a regular language. Which is fine with our scanner , but not powerful enough to handle expressions which can nest arbitrarily deeply.

We need a bigger hammer - context-free grammar (CFG). It's the next heaviest tool in the toolbox of formal grammars.

Lexical-Syntactic

Rules for grammars

If you start with the rules, you can use them to generate strings that are in the grammar. Strings created this way are called derivations because each is derived from the rules of the grammar. In each step of the game, you pick a rule and follow what it tells you to do. Most of the lingo around formal grammars comes from playing them in this direction. Rules are called productions because they produce strings in the grammar.

Each production in a context-free grammar has a head - its name -- and a body, which describes what it generates. In its pure form, the body is simply a list of symbols. Symbols come in two delectable flavors:

  • A terminal is a letter from the grammar's alphabet. You can think of it like a literal value. In the syntactic grammar we're defining, the terminals are individual lexemes - tokens coming from scanner like if or 1234.
  • A nonterminal is a named reference to another rule in the grammar. It means "play that rule and insert whatever it produces here". In this way, grammar composes.

To make this concrete, we need a way to write down these production rules. People have been trying to crystallize grammar all the way back to Panini's Ashtadhyayi, which codified Sanskrit grammar a mere couple thousand years ago. Not much progress happened until John Backus and company needed a notation for specifying ALGOL 58 and came up with Backus-Naur form (BNF). Since then, nearly everyone uses some flavor of BNF, tweaked to their own tastes.

Something clean, each rule is a name, followed by an arrow (->), followed by a sequence of symbols, and finally ending with a semicolon(;). Terminals are quoted strings, and nonterminals are lowercase words.

Here is a grammar for breakfast menus:

breakfast  -> protein "with" breakfast "on the side" ;
breakfast  -> protein ;
breakfast  -> bread ;

protein    -> crispiness "crispy" "bacon" ;
protein    -> "sausage" ;
protein    -> cooked "eggs" ;

crispiness -> "really" ;
crispiness -> "really" crispiness ;

cooked     -> "scrambled" ;
cooked     -> "poached" ;
cooked     -> "fried" ;

bread      -> "toast" ;
bread      -> "biscuits" ;
bread      -> "English muffin" ;

We can use this grammar to generate random breakfasts. Let's play a round and see how it works.

With that, every nonterminal in the string has been expanded until it finally contains only terminals and we're left with: breakfast expand

Enhancing our notation

Parsing Expressions

The cover of Compilers: Principles, Techniques, and Tools literally has a dragon labeled "complexity of compiler design" being slain by a knight bearing a sword and shield branched "LAIR parser generator" and "syntax directed translation". They laid it on thick.

A little self-congratulation is well-deserved, but the truth is you don't need to know most of that stuff to bang out a high quality parser for a modern machine. As always, I encourage you to broaden your education and take it in later, but this book omits the trophy case.

Evaluating Expressions

Lox doesn't do implicit conversions in equality and Java does not either. We do have to handle nil/null specially so that we don't throw a NullPointerException if we try to call equals() on null. Otherwise, we're fine. Java's equals() method on Boolean, Double, and String have the behavior we want for Lox.

What do you expect this to evaluate to:

(0 / 0) == (0 / 0)

According to IEEE 754, which specifies the behaviors of double-precision numbers, dividing a zero by zero gives you the special NaN ("not a number") value. Strangely enough, NaN is not equal to itself.

In Java, the == operator on primitive doubles preserves that behavior, but the equals() method on the Double class does not. Lox uses the latter, so doesn't follow IEEE. These kinds of subtle incompatibilities occupy a dismaying fraction of language implementers' lives.

We could simply not detect or report a type error at all. This is what C does if you cast a pointer to some type that doesn't match the data that is actually being pointed to. C gains flexibility and speed by allowing that, but is also famously dangerous. Once you misinterpret bits in memory, all bets are off.

Few modern languages accept unsafe operations like that. Instead, most are memory safe and ensure - through a combination of static and runtime checks - that a program can never incorrectly interpret the value stored in a piece of memory.

Statements and State

Executing statements

We're running through the previous couple of chapter in microcosm, working our way through the front end. Our parser can now produce statement syntax trees, so the next and final step is to interpret them. As in expressions, we use the Visitor pattern, but we have a new visitor interface, Stmt.Visitor, to implement since statements have their own base class.

By the way, from the Interpreter.java file.

The class is default PROTECTED in java.

Global Variables

Now that we have statements, we can start working on state. Before we get into all the complexity of lexical scoping, we'll start off with the easiest kind of variables - globals. We need two new constructs.

  1. A variable declaration statement brings a new variable into the world.
    var beverage = "espresso";
    This creates a new binding that associates a name (here "beverage") with a value.
  2. Once that's done, a variable expression accesses that binding. When the identifier "beverage" is used as an expression, it looks up the value bound to that name and returns it.
    print beverage;  // "espresso"
    

Variable syntax

The clauses in control flow statements - think the then and else branches of an if statement or the body of a while - are each a single statement. But that statement is not allowed to be one that declares a name. This is OK:

if (monday) print "Ugh, already?";

But this is not:

if (monday) var beverage = "espresso";

We could allow the latter, but it's confusing. What is the scope of that beverage variable? Does it persist after the if statement? If so, what is its value on days other than Monday? Does the variable exist at all on those days?

Code like this is weird, so C, Java, and friends all disallow it. It's as if there are two levels of "precedence" for statements. Some places where a statement is allowed - like inside a block or at the top level - allow any kind of statement, including declarations. Other allow only the "higher" precedence statements that don't declare names.

To accommodate the distinction, we add another rule for kinds of statements that declare names.

program       -> declaration* EOF ;

declaration   -> varDecl
               | statement ;
               
statement     -> exprStmt
               | printStmt ;

Declaration statements go under the new declaration rule. Right now, it's only variables, but later it will include functions and classes. Any place where a declaration is allowed also allows non-declaring statements, so the declaration rule falls through to statement. Obviously, you can declare stuff at the top level of a script, so program routes to the new rule.

The rule for declaring a variable looks like:

varDecl         ->  "var" IDENTIFIER ( "=" expression  )? ";" ;

Like most statements, it starts with a leading keyword. In this case, var. Then an identifier token for the name of the variable being declared, followed by an optional initializer expression. Finally, we put a bow on it with the semicolon.

To access a variable, we define a new kind of primary expression.

primary          -> "true" | "false" | "nil"
                  | NUMBER | STRING
                  | "(" expression ")"
                  | IDENTIFIER ;

That IDENTIFIER clause matches a single identifier token, which is understood to be the name of the variable being accessed.

These new grammar rules get their corresponding syntax trees. Over in the AST generator, we add a new statement node for a variable declaration.

Environments

The bindings that associate variables to values need to be stored somewhere. Ever since the Lisp folks invented parentheses, this data structure has been called an environment. environment

You can think of it like a map where the keys are variable names and the values are the variable's, uh, values. In fact, that's how we'll implement it in Java. We could stuff that map and the code to manage it right into Interpreter, but since it forms a nicely delineated concept, we'll pull it out into its own class.

Rule about variables and scoping is, "When in double, do what Scheme does". The Scheme folks have probably spent more time thinking about variable scope than we ever will - one of the main goals of Scheme was to introduce lexical scoping lexical scoping to the world - so it's hard to go wrong if you follow in their footsteps.

Scheme allows redefining variables at the top level.

Assignment

It's possible to create a language that has variables but does not let you reassign - or mutate -- them. Haskell is one example. SML supports only mutable references and arrays - variables cannot be reassigned. Rust steers you away from mutation by requiring a mut modifier to enable assignment.

Mutating a variable is a side effect and, as the name suggests, some language folks think side effects are dirty or inelegant. Code should be pure math that produces values - crystalline, unchanging ones - like an act of divine creation. Not some grubby automaton that beats blobs of data into shape, one imperative grunt at a time.

Lox is not so austere. Lox is an imperative language, and mutation comes with the territory. Adding support for assignment doesn't require much work. Global variables already support redefinition, so most of the machinery is there now. Mainly, we're missing an explicit assignment notation.

Assignment syntax

That little = syntax is more complex than it might seem. Like most C-derived languages, assignment is an expression and not a statement. As in C, it is the lowest precedence expression form. That means the rule slots between expression and equality (the next lowest precedence expression).

expression      -> assignment ;
assignment      -> IDENTIFIER "=" assignment
                 | equality ;

This says an assignment is either an identifier followed by an = and an expression for the value, or an equality (and thus any other) expression. Later, assignment will get more complex when we add property setters on objects, like:

instance.field = "value";

Consider:

var a = "before";
a = "value";

On the second line, we don't evaluate a (which would return the string "before"). We figure out what variable a refers to, so we know where to store the right-hand side expression's value. The classic terms for these two constructs are l-value and r-value. All the expressions that we've seen so far that produce values are r-values. An l-values "evaluates" to a storage location that you can assign into.

We want the syntax tree to reflect that an l-value isn't evaluated like a normal expression. That's why the Expr.Assign node has a Token for the left-hand side, not an Expr. The problem is that the parser doesn't know it's parsing an l-value until it hits the =. In a complex l-value, that may occur many tokens later.

makeList().head.next = node;

Assignment semantics

The key difference between assignment and definition is that assignment is not allowed to create a new variable. In terms of our implementation, that means it's a runtime error if the key doesn't already exist in the environment's variable map.

The last thing the visit() method does is return the assigned value. That's because assignment is an expression that can be nested inside other expressions, like so:

var a = 1;
print a = 2;  // "2"

Our interpreter can now create, read, and modify variables. It's about as sophisticated as early BASICs. Global variables are simple, but writing a large program when any two chunks of code can accidentally step on each other's state is no fun. We want local variables, which means it's time for scope.

Note: How to fix idea go to editor with escape when editing in terminal vim: intellij support

Scope

A scope defines a region where a name maps to a certain entity. Multiple scopes enable the same name to refer to different things in different contexts. In my house, "Bob" usually refers to me. But maybe in your town you know a different Bob. Same name, but different dudes on where you say it.

Lexical scope (or the less commonly heard static scope) is a specific style of scoping where the text of the program itself shows where a scope begins and ends. In Lox, as in most modern languages, variables are lexically scoped. When you see an expression that uses some variable, you can figure out which variable declaration it refers to just by statically reading the code.

E.g.:

{
  var a = "first";
  print a;  // "first"
}

{
  var a = "second";
  print a; // "second"
}

Here, we have two blocks with a variable a declared in each of them. You and I can tell just from looking at the code that the use of a in the first print statement refers to the first a, and the second one refers to the second. scope

This is in contrast to dynamic scope where you don't know what a name refers to until you execute the code. Lox doesn't have dynamically scoped variables, but methods and fields on objects are dynamically scoped.

class Saxophone {
  play() {
    print "Careless Whisper";
  }
}

class GolfClub {
  play() {
    print "Fore!";
  }
}

fun palyIt(thing) {
  thing.play();
}

When playIt() calls thing.play(), we don't know if we're about to hear "Careless Whisper" or "Fore!". It depends on whether you pass a Saxophone or a GolfClub to the function, and we don't know that until runtime.

Scope and environments are close cousins. The former is the theoretical concept, and the latter is the machinery that implements it. As our interpreter works its way through code, syntax tree nodes that affect scope will change the environment. In a C-ish syntax like Lox's, scope is controlled by curly-braced blocks. (That's why we call it block scope).

{
  var a = "in block";
}
print a;  // Error! No more "a".

The beginning of a block introduces a new local scope, and that scope ends when execution passes the closing }. Any variables declared inside the block disappear.

Nesting and shadowing

A first cut at implementing block scope might work like this:

  1. As we visit each statement inside the block, keep track of any variables declared.
  2. After the last statement is executed, tell the environment to delete all of those variables.

That would work for the previous example. But remember, one motivation for local scope is encapsulation — a block of code in one corner of the program shouldn't interfere with some other block. Check this out:

// How loud?
var volume = 11;

// Silence.
volume = 0;

// Calculate size of 3x4x5 cuboid.
{
  var volume = 3 * 4 * 5;
  print volume;
}

Look at the block where we calculate the volume of the cuboid using a local declaration of volume. After the block exits, the interpreter will delete the global volume variable. That ain't right. When we exit the block, we should remove any variables declared inside the block, but if there is a variable with the same name declared outside the block, that's a different variable. It shouldn't get touched.

When a local variable has the same name as a variable in an enclosing scope, it shadows the outer one. Code inside the block can't see it any more - it is hidden in the "shadow" cast by the inner one - but it's still there.

When we enter a new block scope, we need to preserve variable defined in outer scopes, so they are still around when we exit the inner block. We do that by defining a fresh environment for each block containing only the variables defined in that scope. When we exit the block, we discard its environment and restore the previous one.

We also need to handle enclosing variables that are not shadowed.

var global = "outside";
{
  var local = "inside";
  print global + local;
}

Here, global lives in the outer global environment and local is defined inside the block's environment. In that print statement, both of those variables are in scope. In order to find them, the interpreter must search not only the current innermost environment, but also any enclosing ones.

We implement this by chaining the environments together. Each environment has a reference to the environment of the immediately enclosing scope. When we look up a variable, we walk that chain from innermost out until we find the variable. Starting at the inner scope is how we make local variable shadow outer ones. scope-level Before we add block syntax to the grammar, we'll beef up our Environment class with support for this nesting. First, we give each environment a reference to its enclosing one.

Block syntax and semantics

Now the Environments nest, we're ready to add blocks to the language. Behold the grammar:

statement       -> exprStmt
                 | printStmt
                 | block ;
                 
block           -> "{" declaration* "}" ;

A block is a (possible empty) series of statements or declarations surrounded by curly braces. A block is itself a statement and can appear anywhere a statement is allowed.

About

A Java version of Lox interpreter

Resources

Stars

Watchers

Forks

Releases

No releases published

Packages

No packages published