10

Functions

And that is also the way the human mind worksby the compounding of old ideas into new structures that become new ideas that can themselves be used in compounds, and round and round endlessly, growing ever more remote from the basic earthbound imagery that is each language’s soil.

Douglas R. Hofstadter, I Am a Strange Loop

This chapter marks the culmination of a lot of hard work. The previous chapters add useful functionality in their own right, but each also supplies a piece of a puzzle. We’ll take those piecesexpressions, statements, variables, control flow, and lexical scopeadd a couple more, and assemble them all into support for real user-defined functions and function calls.

10 . 1Function Calls

You’re certainly familiar with C-style function call syntax, but the grammar is more subtle than you may realize. Calls are typically to named functions like:

average(1, 2);

But the name of the function being called isn’t actually part of the call syntax. The thing being calledthe calleecan be any expression that evaluates to a function. (Well, it does have to be a pretty high precedence expression, but parentheses take care of that.) For example:

getCallback()();

There are two call expressions here. The first pair of parentheses has getCallback as its callee. But the second call has the entire getCallback() expression as its callee. It is the parentheses following an expression that indicate a function call. You can think of a call as sort of like a postfix operator that starts with (.

This “operator” has higher precedence than any other operator, even the unary ones. So we slot it into the grammar by having the unary rule bubble up to a new call rule.

unary          → ( "!" | "-" ) unary | call ;
callprimary ( "(" arguments? ")" )* ;

This rule matches a primary expression followed by zero or more function calls. If there are no parentheses, this parses a bare primary expression. Otherwise, each call is recognized by a pair of parentheses with an optional list of arguments inside. The argument list grammar is:

argumentsexpression ( "," expression )* ;

This rule requires at least one argument expression, followed by zero or more other expressions, each preceded by a comma. To handle zero-argument calls, the call rule itself considers the entire arguments production to be optional.

I admit, this seems more grammatically awkward than you’d expect for the incredibly common “zero or more comma-separated things” pattern. There are some sophisticated metasyntaxes that handle this better, but in our BNF and in many language specs I’ve seen, it is this cumbersome.

Over in our syntax tree generator, we add a new node.

      "Binary   : Expr left, Token operator, Expr right",
tool/GenerateAst.java
in main()
      "Call     : Expr callee, Token paren, List<Expr> arguments",
      "Grouping : Expr expression",
tool/GenerateAst.java, in main()

It stores the callee expression and a list of expressions for the arguments. It also stores the token for the closing parenthesis. We’ll use that token’s location when we report a runtime error caused by a function call.

Crack open the parser. Where unary() used to jump straight to primary(), change it to call, well, call().

      return new Expr.Unary(operator, right);
    }

lox/Parser.java
in unary()
replace 1 line
    return call();
  }
lox/Parser.java, in unary(), replace 1 line

Its definition is:

lox/Parser.java
add after unary()
  private Expr call() {
    Expr expr = primary();

    while (true) { 
      if (match(LEFT_PAREN)) {
        expr = finishCall(expr);
      } else {
        break;
      }
    }

    return expr;
  }
lox/Parser.java, add after unary()

The code here doesn’t quite line up with the grammar rules. I moved a few things around to make the code cleanerone of the luxuries we have with a handwritten parser. But it’s roughly similar to how we parse infix operators. First, we parse a primary expression, the “left operand” to the call. Then, each time we see a (, we call finishCall() to parse the call expression using the previously parsed expression as the callee. The returned expression becomes the new expr and we loop to see if the result is itself called.

The code to parse the argument list is in this helper:

lox/Parser.java
add after unary()
  private Expr finishCall(Expr callee) {
    List<Expr> arguments = new ArrayList<>();
    if (!check(RIGHT_PAREN)) {
      do {
        arguments.add(expression());
      } while (match(COMMA));
    }

    Token paren = consume(RIGHT_PAREN,
                          "Expect ')' after arguments.");

    return new Expr.Call(callee, paren, arguments);
  }
lox/Parser.java, add after unary()

This is more or less the arguments grammar rule translated to code, except that we also handle the zero-argument case. We check for that case first by seeing if the next token is ). If it is, we don’t try to parse any arguments.

Otherwise, we parse an expression, then look for a comma indicating that there is another argument after that. We keep doing that as long as we find commas after each expression. When we don’t find a comma, then the argument list must be done and we consume the expected closing parenthesis. Finally, we wrap the callee and those arguments up into a call AST node.

10 . 1 . 1Maximum argument counts

Right now, the loop where we parse arguments has no bound. If you want to call a function and pass a million arguments to it, the parser would have no problem with it. Do we want to limit that?

Other languages have various approaches. The C standard says a conforming implementation has to support at least 127 arguments to a function, but doesn’t say there’s any upper limit. The Java specification says a method can accept no more than 255 arguments.

Our Java interpreter for Lox doesn’t really need a limit, but having a maximum number of arguments will simplify our bytecode interpreter in Part III. We want our two interpreters to be compatible with each other, even in weird corner cases like this, so we’ll add the same limit to jlox.

      do {
lox/Parser.java
in finishCall()
        if (arguments.size() >= 255) {
          error(peek(), "Can't have more than 255 arguments.");
        }
        arguments.add(expression());
lox/Parser.java, in finishCall()

Note that the code here reports an error if it encounters too many arguments, but it doesn’t throw the error. Throwing is how we kick into panic mode which is what we want if the parser is in a confused state and doesn’t know where it is in the grammar anymore. But here, the parser is still in a perfectly valid stateit just found too many arguments. So it reports the error and keeps on keepin’ on.

10 . 1 . 2Interpreting function calls

We don’t have any functions we can call, so it seems weird to start implementing calls first, but we’ll worry about that when we get there. First, our interpreter needs a new import.

lox/Interpreter.java
import java.util.ArrayList;
import java.util.List;
lox/Interpreter.java

As always, interpretation starts with a new visit method for our new call expression node.

lox/Interpreter.java
add after visitBinaryExpr()
  @Override
  public Object visitCallExpr(Expr.Call expr) {
    Object callee = evaluate(expr.callee);

    List<Object> arguments = new ArrayList<>();
    for (Expr argument : expr.arguments) { 
      arguments.add(evaluate(argument));
    }

    LoxCallable function = (LoxCallable)callee;
    return function.call(this, arguments);
  }
lox/Interpreter.java, add after visitBinaryExpr()

First, we evaluate the expression for the callee. Typically, this expression is just an identifier that looks up the function by its name, but it could be anything. Then we evaluate each of the argument expressions in order and store the resulting values in a list.

Once we’ve got the callee and the arguments ready, all that remains is to perform the call. We do that by casting the callee to a LoxCallable and then invoking a call() method on it. The Java representation of any Lox object that can be called like a function will implement this interface. That includes user-defined functions, naturally, but also class objects since classes are “called” to construct new instances. We’ll also use it for one more purpose shortly.

There isn’t too much to this new interface.

lox/LoxCallable.java
create new file
package com.craftinginterpreters.lox;

import java.util.List;

interface LoxCallable {
  Object call(Interpreter interpreter, List<Object> arguments);
}
lox/LoxCallable.java, create new file

We pass in the interpreter in case the class implementing call() needs it. We also give it the list of evaluated argument values. The implementer’s job is then to return the value that the call expression produces.

10 . 1 . 3Call type errors

Before we get to implementing LoxCallable, we need to make the visit method a little more robust. It currently ignores a couple of failure modes that we can’t pretend won’t occur. First, what happens if the callee isn’t actually something you can call? What if you try to do this:

"totally not a function"();

Strings aren’t callable in Lox. The runtime representation of a Lox string is a Java string, so when we cast that to LoxCallable, the JVM will throw a ClassCastException. We don’t want our interpreter to vomit out some nasty Java stack trace and die. Instead, we need to check the type ourselves first.

    }

lox/Interpreter.java
in visitCallExpr()
    if (!(callee instanceof LoxCallable)) {
      throw new RuntimeError(expr.paren,
          "Can only call functions and classes.");
    }

    LoxCallable function = (LoxCallable)callee;
lox/Interpreter.java, in visitCallExpr()

We still throw an exception, but now we’re throwing our own exception type, one that the interpreter knows to catch and report gracefully.

10 . 1 . 4Checking arity

The other problem relates to the function’s arity. Arity is the fancy term for the number of arguments a function or operation expects. Unary operators have arity one, binary operators two, etc. With functions, the arity is determined by the number of parameters it declares.

fun add(a, b, c) {
  print a + b + c;
}

This function defines three parameters, a, b, and c, so its arity is three and it expects three arguments. So what if you try to call it like this:

add(1, 2, 3, 4); // Too many.
add(1, 2);       // Too few.

Different languages take different approaches to this problem. Of course, most statically typed languages check this at compile time and refuse to compile the code if the argument count doesn’t match the function’s arity. JavaScript discards any extra arguments you pass. If you don’t pass enough, it fills in the missing parameters with the magic sort-of-like-null-but-not-really value undefined. Python is stricter. It raises a runtime error if the argument list is too short or too long.

I think the latter is a better approach. Passing the wrong number of arguments is almost always a bug, and it’s a mistake I do make in practice. Given that, the sooner the implementation draws my attention to it, the better. So for Lox, we’ll take Python’s approach. Before invoking the callable, we check to see if the argument list’s length matches the callable’s arity.

    LoxCallable function = (LoxCallable)callee;
lox/Interpreter.java
in visitCallExpr()
    if (arguments.size() != function.arity()) {
      throw new RuntimeError(expr.paren, "Expected " +
          function.arity() + " arguments but got " +
          arguments.size() + ".");
    }

    return function.call(this, arguments);
lox/Interpreter.java, in visitCallExpr()

That requires a new method on the LoxCallable interface to ask it its arity.

interface LoxCallable {
lox/LoxCallable.java
in interface LoxCallable
  int arity();
  Object call(Interpreter interpreter, List<Object> arguments);
lox/LoxCallable.java, in interface LoxCallable

We could push the arity checking into the concrete implementation of call(). But, since we’ll have multiple classes implementing LoxCallable, that would end up with redundant validation spread across a few classes. Hoisting it up into the visit method lets us do it in one place.

10 . 2Native Functions

We can theoretically call functions, but we have no functions to call yet. Before we get to user-defined functions, now is a good time to introduce a vital but often overlooked facet of language implementationsnative functions. These are functions that the interpreter exposes to user code but that are implemented in the host language (in our case Java), not the language being implemented (Lox).

Sometimes these are called primitives, external functions, or foreign functions. Since these functions can be called while the user’s program is running, they form part of the implementation’s runtime. A lot of programming language books gloss over these because they aren’t conceptually interesting. They’re mostly grunt work.

But when it comes to making your language actually good at doing useful stuff, the native functions your implementation provides are key. They provide access to the fundamental services that all programs are defined in terms of. If you don’t provide native functions to access the file system, a user’s going to have a hell of a time writing a program that reads and displays a file.

Many languages also allow users to provide their own native functions. The mechanism for doing so is called a foreign function interface (FFI), native extension, native interface, or something along those lines. These are nice because they free the language implementer from providing access to every single capability the underlying platform supports. We won’t define an FFI for jlox, but we will add one native function to give you an idea of what it looks like.

10 . 2 . 1Telling time

When we get to Part III and start working on a much more efficient implementation of Lox, we’re going to care deeply about performance. Performance work requires measurement, and that in turn means benchmarks. These are programs that measure the time it takes to exercise some corner of the interpreter.

We could measure the time it takes to start up the interpreter, run the benchmark, and exit, but that adds a lot of overheadJVM startup time, OS shenanigans, etc. That stuff does matter, of course, but if you’re just trying to validate an optimization to some piece of the interpreter, you don’t want that overhead obscuring your results.

A nicer solution is to have the benchmark script itself measure the time elapsed between two points in the code. To do that, a Lox program needs to be able to tell time. There’s no way to do that nowyou can’t implement a useful clock “from scratch” without access to the underlying clock on the computer.

So we’ll add clock(), a native function that returns the number of seconds that have passed since some fixed point in time. The difference between two successive invocations tells you how much time elapsed between the two calls. This function is defined in the global scope, so let’s ensure the interpreter has access to that.

class Interpreter implements Expr.Visitor<Object>,
                             Stmt.Visitor<Void> {
lox/Interpreter.java
in class Interpreter
replace 1 line
  final Environment globals = new Environment();
  private Environment environment = globals;

  void interpret(List<Stmt> statements) {
lox/Interpreter.java, in class Interpreter, replace 1 line

The environment field in the interpreter changes as we enter and exit local scopes. It tracks the current environment. This new globals field holds a fixed reference to the outermost global environment.

When we instantiate an Interpreter, we stuff the native function in that global scope.

  private Environment environment = globals;

lox/Interpreter.java
in class Interpreter
  Interpreter() {
    globals.define("clock", new LoxCallable() {
      @Override
      public int arity() { return 0; }

      @Override
      public Object call(Interpreter interpreter,
                         List<Object> arguments) {
        return (double)System.currentTimeMillis() / 1000.0;
      }

      @Override
      public String toString() { return "<native fn>"; }
    });
  }

  void interpret(List<Stmt> statements) {
lox/Interpreter.java, in class Interpreter

This defines a variable named “clock”. Its value is a Java anonymous class that implements LoxCallable. The clock() function takes no arguments, so its arity is zero. The implementation of call() calls the corresponding Java function and converts the result to a double value in seconds.

If we wanted to add other native functionsreading input from the user, working with files, etc.we could add them each as their own anonymous class that implements LoxCallable. But for the book, this one is really all we need.

Let’s get ourselves out of the function-defining business and let our users take over . . . 

10 . 3Function Declarations

We finally get to add a new production to the declaration rule we introduced back when we added variables. Function declarations, like variables, bind a new name. That means they are allowed only in places where a declaration is permitted.

declarationfunDecl
               | varDecl
               | statement ;

The updated declaration rule references this new rule:

funDecl"fun" function ;
functionIDENTIFIER "(" parameters? ")" block ;

The main funDecl rule uses a separate helper rule function. A function declaration statement is the fun keyword followed by the actual function-y stuff. When we get to classes, we’ll reuse that function rule for declaring methods. Those look similar to function declarations, but aren’t preceded by fun.

The function itself is a name followed by the parenthesized parameter list and the body. The body is always a braced block, using the same grammar rule that block statements use. The parameter list uses this rule:

parametersIDENTIFIER ( "," IDENTIFIER )* ;

It’s like the earlier arguments rule, except that each parameter is an identifier, not an expression. That’s a lot of new syntax for the parser to chew through, but the resulting AST node isn’t too bad.

      "Expression : Expr expression",
tool/GenerateAst.java
in main()
      "Function   : Token name, List<Token> params," +
                  " List<Stmt> body",
      "If         : Expr condition, Stmt thenBranch," +
tool/GenerateAst.java, in main()

A function node has a name, a list of parameters (their names), and then the body. We store the body as the list of statements contained inside the curly braces.

Over in the parser, we weave in the new declaration.

    try {
lox/Parser.java
in declaration()
      if (match(FUN)) return function("function");
      if (match(VAR)) return varDeclaration();
lox/Parser.java, in declaration()

Like other statements, a function is recognized by the leading keyword. When we encounter fun, we call function. That corresponds to the function grammar rule since we already matched and consumed the fun keyword. We’ll build the method up a piece at a time, starting with this:

lox/Parser.java
add after expressionStatement()
  private Stmt.Function function(String kind) {
    Token name = consume(IDENTIFIER, "Expect " + kind + " name.");
  }
lox/Parser.java, add after expressionStatement()

Right now, it only consumes the identifier token for the function’s name. You might be wondering about that funny little kind parameter. Just like we reuse the grammar rule, we’ll reuse the function() method later to parse methods inside classes. When we do that, we’ll pass in “method” for kind so that the error messages are specific to the kind of declaration being parsed.

Next, we parse the parameter list and the pair of parentheses wrapped around it.

    Token name = consume(IDENTIFIER, "Expect " + kind + " name.");
lox/Parser.java
in function()
    consume(LEFT_PAREN, "Expect '(' after " + kind + " name.");
    List<Token> parameters = new ArrayList<>();
    if (!check(RIGHT_PAREN)) {
      do {
        if (parameters.size() >= 255) {
          error(peek(), "Can't have more than 255 parameters.");
        }

        parameters.add(
            consume(IDENTIFIER, "Expect parameter name."));
      } while (match(COMMA));
    }
    consume(RIGHT_PAREN, "Expect ')' after parameters.");
  }
lox/Parser.java, in function()

This is like the code for handling arguments in a call, except not split out into a helper method. The outer if statement handles the zero parameter case, and the inner while loop parses parameters as long as we find commas to separate them. The result is the list of tokens for each parameter’s name.

Just like we do with arguments to function calls, we validate at parse time that you don’t exceed the maximum number of parameters a function is allowed to have.

Finally, we parse the body and wrap it all up in a function node.

    consume(RIGHT_PAREN, "Expect ')' after parameters.");
lox/Parser.java
in function()

    consume(LEFT_BRACE, "Expect '{' before " + kind + " body.");
    List<Stmt> body = block();
    return new Stmt.Function(name, parameters, body);
  }
lox/Parser.java, in function()

Note that we consume the { at the beginning of the body here before calling block(). That’s because block() assumes the brace token has already been matched. Consuming it here lets us report a more precise error message if the { isn’t found since we know it’s in the context of a function declaration.

10 . 4Function Objects

We’ve got some syntax parsed so usually we’re ready to interpret, but first we need to think about how to represent a Lox function in Java. We need to keep track of the parameters so that we can bind them to argument values when the function is called. And, of course, we need to keep the code for the body of the function so that we can execute it.

That’s basically what the Stmt.Function class is. Could we just use that? Almost, but not quite. We also need a class that implements LoxCallable so that we can call it. We don’t want the runtime phase of the interpreter to bleed into the front end’s syntax classes so we don’t want Stmt.Function itself to implement that. Instead, we wrap it in a new class.

lox/LoxFunction.java
create new file
package com.craftinginterpreters.lox;

import java.util.List;

class LoxFunction implements LoxCallable {
  private final Stmt.Function declaration;
  LoxFunction(Stmt.Function declaration) {
    this.declaration = declaration;
  }
}
lox/LoxFunction.java, create new file

We implement the call() of LoxCallable like so:

lox/LoxFunction.java
add after LoxFunction()
  @Override
  public Object call(Interpreter interpreter,
                     List<Object> arguments) {
    Environment environment = new Environment(interpreter.globals);
    for (int i = 0; i < declaration.params.size(); i++) {
      environment.define(declaration.params.get(i).lexeme,
          arguments.get(i));
    }

    interpreter.executeBlock(declaration.body, environment);
    return null;
  }
lox/LoxFunction.java, add after LoxFunction()

This handful of lines of code is one of the most fundamental, powerful pieces of our interpreter. As we saw in the chapter on statements and state, managing name environments is a core part of a language implementation. Functions are deeply tied to that.

Parameters are core to functions, especially the fact that a function encapsulates its parametersno other code outside of the function can see them. This means each function gets its own environment where it stores those variables.

Further, this environment must be created dynamically. Each function call gets its own environment. Otherwise, recursion would break. If there are multiple calls to the same function in play at the same time, each needs its own environment, even though they are all calls to the same function.

For example, here’s a convoluted way to count to three:

fun count(n) {
  if (n > 1) count(n - 1);
  print n;
}

count(3);

Imagine we pause the interpreter right at the point where it’s about to print 1 in the innermost nested call. The outer calls to print 2 and 3 haven’t printed their values yet, so there must be environments somewhere in memory that still store the fact that n is bound to 3 in one context, 2 in another, and 1 in the innermost, like:

A separate environment for each recursive call.

That’s why we create a new environment at each call, not at the function declaration. The call() method we saw earlier does that. At the beginning of the call, it creates a new environment. Then it walks the parameter and argument lists in lockstep. For each pair, it creates a new variable with the parameter’s name and binds it to the argument’s value.

So, for a program like this:

fun add(a, b, c) {
  print a + b + c;
}

add(1, 2, 3);

At the point of the call to add(), the interpreter creates something like this:

Binding arguments to their parameters.

Then call() tells the interpreter to execute the body of the function in this new function-local environment. Up until now, the current environment was the environment where the function was being called. Now, we teleport from there inside the new parameter space we’ve created for the function.

This is all that’s required to pass data into the function. By using different environments when we execute the body, calls to the same function with the same code can produce different results.

Once the body of the function has finished executing, executeBlock() discards that function-local environment and restores the previous one that was active back at the callsite. Finally, call() returns null, which returns nil to the caller. (We’ll add return values later.)

Mechanically, the code is pretty simple. Walk a couple of lists. Bind some new variables. Call a method. But this is where the crystalline code of the function declaration becomes a living, breathing invocation. This is one of my favorite snippets in this entire book. Feel free to take a moment to meditate on it if you’re so inclined.

Done? OK. Note when we bind the parameters, we assume the parameter and argument lists have the same length. This is safe because visitCallExpr() checks the arity before calling call(). It relies on the function reporting its arity to do that.

lox/LoxFunction.java
add after LoxFunction()
  @Override
  public int arity() {
    return declaration.params.size();
  }
lox/LoxFunction.java, add after LoxFunction()

That’s most of our object representation. While we’re in here, we may as well implement toString().

lox/LoxFunction.java
add after LoxFunction()
  @Override
  public String toString() {
    return "<fn " + declaration.name.lexeme + ">";
  }
lox/LoxFunction.java, add after LoxFunction()

This gives nicer output if a user decides to print a function value.

fun add(a, b) {
  print a + b;
}

print add; // "<fn add>".

10 . 4 . 1Interpreting function declarations

We’ll come back and refine LoxFunction soon, but that’s enough to get started. Now we can visit a function declaration.

lox/Interpreter.java
add after visitExpressionStmt()
  @Override
  public Void visitFunctionStmt(Stmt.Function stmt) {
    LoxFunction function = new LoxFunction(stmt);
    environment.define(stmt.name.lexeme, function);
    return null;
  }
lox/Interpreter.java, add after visitExpressionStmt()

This is similar to how we interpret other literal expressions. We take a function syntax nodea compile-time representation of the functionand convert it to its runtime representation. Here, that’s a LoxFunction that wraps the syntax node.

Function declarations are different from other literal nodes in that the declaration also binds the resulting object to a new variable. So, after creating the LoxFunction, we create a new binding in the current environment and store a reference to it there.

With that, we can define and call our own functions all within Lox. Give it a try:

fun sayHi(first, last) {
  print "Hi, " + first + " " + last + "!";
}

sayHi("Dear", "Reader");

I don’t know about you, but that looks like an honest-to-God programming language to me.

10 . 5Return Statements

We can get data into functions by passing parameters, but we’ve got no way to get results back out. If Lox were an expression-oriented language like Ruby or Scheme, the body would be an expression whose value is implicitly the function’s result. But in Lox, the body of a function is a list of statements which don’t produce values, so we need dedicated syntax for emitting a result. In other words, return statements. I’m sure you can guess the grammar already.

statementexprStmt
               | forStmt
               | ifStmt
               | printStmt
               | returnStmt
               | whileStmt
               | block ;

returnStmt"return" expression? ";" ;

We’ve got one morethe final, in factproduction under the venerable statement rule. A return statement is the return keyword followed by an optional expression and terminated with a semicolon.

The return value is optional to support exiting early from a function that doesn’t return a useful value. In statically typed languages, “void” functions don’t return a value and non-void ones do. Since Lox is dynamically typed, there are no true void functions. The compiler has no way of preventing you from taking the result value of a call to a function that doesn’t contain a return statement.

fun procedure() {
  print "don't return anything";
}

var result = procedure();
print result; // ?

This means every Lox function must return something, even if it contains no return statements at all. We use nil for this, which is why LoxFunction’s implementation of call() returns null at the end. In that same vein, if you omit the value in a return statement, we simply treat it as equivalent to:

return nil;

Over in our AST generator, we add a new node.

      "Print      : Expr expression",
tool/GenerateAst.java
in main()
      "Return     : Token keyword, Expr value",
      "Var        : Token name, Expr initializer",
tool/GenerateAst.java, in main()

It keeps the return keyword token so we can use its location for error reporting, and the value being returned, if any. We parse it like other statements, first by recognizing the initial keyword.

    if (match(PRINT)) return printStatement();
lox/Parser.java
in statement()
    if (match(RETURN)) return returnStatement();
    if (match(WHILE)) return whileStatement();
lox/Parser.java, in statement()

That branches out to:

lox/Parser.java
add after printStatement()
  private Stmt returnStatement() {
    Token keyword = previous();
    Expr value = null;
    if (!check(SEMICOLON)) {
      value = expression();
    }

    consume(SEMICOLON, "Expect ';' after return value.");
    return new Stmt.Return(keyword, value);
  }
lox/Parser.java, add after printStatement()

After snagging the previously consumed return keyword, we look for a value expression. Since many different tokens can potentially start an expression, it’s hard to tell if a return value is present. Instead, we check if it’s absent. Since a semicolon can’t begin an expression, if the next token is that, we know there must not be a value.

10 . 5 . 1Returning from calls

Interpreting a return statement is tricky. You can return from anywhere within the body of a function, even deeply nested inside other statements. When the return is executed, the interpreter needs to jump all the way out of whatever context it’s currently in and cause the function call to complete, like some kind of jacked up control flow construct.

For example, say we’re running this program and we’re about to execute the return statement:

fun count(n) {
  while (n < 100) {
    if (n == 3) return n; // <--
    print n;
    n = n + 1;
  }
}

count(1);

The Java call stack currently looks roughly like this:

Interpreter.visitReturnStmt()
Interpreter.visitIfStmt()
Interpreter.executeBlock()
Interpreter.visitBlockStmt()
Interpreter.visitWhileStmt()
Interpreter.executeBlock()
LoxFunction.call()
Interpreter.visitCallExpr()

We need to get from the top of the stack all the way back to call(). I don’t know about you, but to me that sounds like exceptions. When we execute a return statement, we’ll use an exception to unwind the interpreter past the visit methods of all of the containing statements back to the code that began executing the body.

The visit method for our new AST node looks like this:

lox/Interpreter.java
add after visitPrintStmt()
  @Override
  public Void visitReturnStmt(Stmt.Return stmt) {
    Object value = null;
    if (stmt.value != null) value = evaluate(stmt.value);

    throw new Return(value);
  }
lox/Interpreter.java, add after visitPrintStmt()

If we have a return value, we evaluate it, otherwise, we use nil. Then we take that value and wrap it in a custom exception class and throw it.

lox/Return.java
create new file
package com.craftinginterpreters.lox;

class Return extends RuntimeException {
  final Object value;

  Return(Object value) {
    super(null, null, false, false);
    this.value = value;
  }
}
lox/Return.java, create new file

This class wraps the return value with the accoutrements Java requires for a runtime exception class. The weird super constructor call with those null and false arguments disables some JVM machinery that we don’t need. Since we’re using our exception class for control flow and not actual error handling, we don’t need overhead like stack traces.

We want this to unwind all the way to where the function call began, the call() method in LoxFunction.

          arguments.get(i));
    }

lox/LoxFunction.java
in call()
replace 1 line
    try {
      interpreter.executeBlock(declaration.body, environment);
    } catch (Return returnValue) {
      return returnValue.value;
    }
    return null;
lox/LoxFunction.java, in call(), replace 1 line

We wrap the call to executeBlock() in a try-catch block. When it catches a return exception, it pulls out the value and makes that the return value from call(). If it never catches one of these exceptions, it means the function reached the end of its body without hitting a return statement. In that case, it implicitly returns nil.

Let’s try it out. We finally have enough power to support this classic examplea recursive function to calculate Fibonacci numbers:

fun fib(n) {
  if (n <= 1) return n;
  return fib(n - 2) + fib(n - 1);
}

for (var i = 0; i < 20; i = i + 1) {
  print fib(i);
}

This tiny program exercises almost every language feature we have spent the past several chapters implementingexpressions, arithmetic, branching, looping, variables, functions, function calls, parameter binding, and returns.

10 . 6Local Functions and Closures

Our functions are pretty full featured, but there is one hole to patch. In fact, it’s a big enough gap that we’ll spend most of the next chapter sealing it up, but we can get started here.

LoxFunction’s implementation of call() creates a new environment where it binds the function’s parameters. When I showed you that code, I glossed over one important point: What is the parent of that environment?

Right now, it is always globals, the top-level global environment. That way, if an identifier isn’t defined inside the function body itself, the interpreter can look outside the function in the global scope to find it. In the Fibonacci example, that’s how the interpreter is able to look up the recursive call to fib inside the function’s own bodyfib is a global variable.

But recall that in Lox, function declarations are allowed anywhere a name can be bound. That includes the top level of a Lox script, but also the inside of blocks or other functions. Lox supports local functions that are defined inside another function, or nested inside a block.

Consider this classic example:

fun makeCounter() {
  var i = 0;
  fun count() {
    i = i + 1;
    print i;
  }

  return count;
}

var counter = makeCounter();
counter(); // "1".
counter(); // "2".

Here, count() uses i, which is declared outside of itself in the containing function makeCounter(). makeCounter() returns a reference to the count() function and then its own body finishes executing completely.

Meanwhile, the top-level code invokes the returned count() function. That executes the body of count(), which assigns to and reads i, even though the function where i was defined has already exited.

If you’ve never encountered a language with nested functions before, this might seem crazy, but users do expect it to work. Alas, if you run it now, you get an undefined variable error in the call to counter() when the body of count() tries to look up i. That’s because the environment chain in effect looks like this:

The environment chain from count()'s body to the global scope.

When we call count() (through the reference to it stored in counter), we create a new empty environment for the function body. The parent of that is the global environment. We lost the environment for makeCounter() where i is bound.

Let’s go back in time a bit. Here’s what the environment chain looked like right when we declared count() inside the body of makeCounter():

The environment chain inside the body of makeCounter().

So at the point where the function is declared, we can see i. But when we return from makeCounter() and exit its body, the interpreter discards that environment. Since the interpreter doesn’t keep the environment surrounding count() around, it’s up to the function object itself to hang on to it.

This data structure is called a closure because it “closes over” and holds on to the surrounding variables where the function is declared. Closures have been around since the early Lisp days, and language hackers have come up with all manner of ways to implement them. For jlox, we’ll do the simplest thing that works. In LoxFunction, we add a field to store an environment.

  private final Stmt.Function declaration;
lox/LoxFunction.java
in class LoxFunction
  private final Environment closure;

  LoxFunction(Stmt.Function declaration) {
lox/LoxFunction.java, in class LoxFunction

We initialize that in the constructor.

lox/LoxFunction.java
constructor LoxFunction()
replace 1 line
  LoxFunction(Stmt.Function declaration, Environment closure) {
    this.closure = closure;
    this.declaration = declaration;
lox/LoxFunction.java, constructor LoxFunction(), replace 1 line

When we create a LoxFunction, we capture the current environment.

  public Void visitFunctionStmt(Stmt.Function stmt) {
lox/Interpreter.java
in visitFunctionStmt()
replace 1 line
    LoxFunction function = new LoxFunction(stmt, environment);
    environment.define(stmt.name.lexeme, function);
lox/Interpreter.java, in visitFunctionStmt(), replace 1 line

This is the environment that is active when the function is declared not when it’s called, which is what we want. It represents the lexical scope surrounding the function declaration. Finally, when we call the function, we use that environment as the call’s parent instead of going straight to globals.

                     List<Object> arguments) {
lox/LoxFunction.java
in call()
replace 1 line
    Environment environment = new Environment(closure);
    for (int i = 0; i < declaration.params.size(); i++) {
lox/LoxFunction.java, in call(), replace 1 line

This creates an environment chain that goes from the function’s body out through the environments where the function is declared, all the way out to the global scope. The runtime environment chain matches the textual nesting of the source code like we want. The end result when we call that function looks like this:

The environment chain with the closure.

Now, as you can see, the interpreter can still find i when it needs to because it’s in the middle of the environment chain. Try running that makeCounter() example now. It works!

Functions let us abstract over, reuse, and compose code. Lox is much more powerful than the rudimentary arithmetic calculator it used to be. Alas, in our rush to cram closures in, we have let a tiny bit of dynamic scoping leak into the interpreter. In the next chapter, we will explore deeper into lexical scope and close that hole.

Challenges

  1. Our interpreter carefully checks that the number of arguments passed to a function matches the number of parameters it expects. Since this check is done at runtime on every call, it has a performance cost. Smalltalk implementations don’t have that problem. Why not?

  2. Lox’s function declaration syntax performs two independent operations. It creates a function and also binds it to a name. This improves usability for the common case where you do want to associate a name with the function. But in functional-styled code, you often want to create a function to immediately pass it to some other function or return it. In that case, it doesn’t need a name.

    Languages that encourage a functional style usually support anonymous functions or lambdasan expression syntax that creates a function without binding it to a name. Add anonymous function syntax to Lox so that this works:

    fun thrice(fn) {
      for (var i = 1; i <= 3; i = i + 1) {
        fn(i);
      }
    }
    
    thrice(fun (a) {
      print a;
    });
    // "1".
    // "2".
    // "3".
    

    How do you handle the tricky case of an anonymous function expression occurring in an expression statement:

    fun () {};
    
  3. Is this program valid?

    fun scope(a) {
      var a = "local";
    }
    

    In other words, are a function’s parameters in the same scope as its local variables, or in an outer scope? What does Lox do? What about other languages you are familiar with? What do you think a language should do?