Find a file
2023-03-24 22:05:49 -07:00
build.sh Fixes 2022-05-21 23:05:40 -07:00
frustration.4th clarification 2022-05-22 10:53:45 -07:00
frustration.rs Fixed an incorrect word name (thanks /u/lalalalalalala71) 2023-03-24 22:05:49 -07:00
frustration_minimal.rs Demonstrate that the Rust file fits in 500 lines after stripping comments 2022-05-29 11:45:30 -07:00
LICENSE Add project URL and license 2022-05-23 21:46:58 -07:00
README.md Fixed an incorrect stack comment where an "a" just vanished (thanks /u/bagtowneast) 2022-05-30 12:44:28 -07:00

Project URL: https://gitlab.cs.washington.edu/fidelp/frustration

Frustration - Escaping a Turing Tar Pit with Forth

What is this file?

This is a tutorial that will show you how to bootstrap an interactive programming environment from a small amount of code.

First we will design a virtual computer.

Then we will design software to run on that computer, to enable REPL-style interactive programming.

A REPL is a "Read, Evaluate, Print loop". A REPL lets you type code at the keyboard and immediately get a result back. You can also define functions, including functions that change how the environment works in fundamental ways.

What is Forth?

Forth is the programming language we will use with our computer.

Forth was invented by Chuck Moore in the 1960s as a tool for quickly coming to grips with new computer systems.

"Let us imagine a situation in which you have access to your computer. I mean sole user sitting at the board with all the lights, for some hours at a time. This is admittedly an atypical situation, but one that can always be arranged if you are competent, press hard, and will work odd hours. Can you and the computer write a program? Can you write a program that didn't descend from a pre-existing program? You can learn a bit and have a lot of fun trying."

-- Chuck Moore, "Programming a Problem-Oriented Language", 1970

As you will see, it does not take much work to get Forth running on a new machine, including a machine with a completely unfamiliar instruction set.

But before we can do any of that we will need a machine. Let's make one.

Table of Contents

  • Part 1 - The Computer
    • 1.0 - Designing the CPU
      • Defining a stack
      • Designing a stack CPU
    • 1.1 - The instruction set
      • Memory access
      • Designing the instruction set
        • The CALL instruction
        • Data processing instructions
        • The LITERAL instruction
      • Making the CPU run
        • Return-stack instructions
        • Memory instructions
        • Stack shuffling instructions
        • Conditional skip instruction
        • Arithmetic and logic
        • Input/output
  • Part 2 - The Program
    • Designing the Forth dictionary
    • Tools for building the Forth dictionary
    • Building the Forth dictionary
      • Subroutine threading
      • key
      • emit
      • subtraction
      • 0= (compare-to-zero)
      • = (equals)
    • 2.1 - The lexer
      • Skipping whitespace
      • Reading characters into a buffer
        • over
        • 2dup
        • The input buffer
        • min
        • c@ and c! (byte-by-byte memory access)
        • Filling the input buffer
        • word
    • 2.2 - Dictionary lookup
      • latest
      • find
      • ' (quote)
    • 2.3 - The outer interpreter
      • here
      • Achieving interactivity
      • immediate
      • [ and ]
      • smudge and unsmudge
      • , (comma)
      • number
      • literal
    • 2.4 - Defining subroutines
      • create
      • : (define word)
      • ; (end of definition)
    • Miscellanea
  • Part 3 - Using the interactive programming environment

Part 1 - The Computer

1.0 - Designing the CPU

This computer will have a 16-bit CPU. It will be able to access 2^16 (65536) memory locations, numbered 0 to 65535. Each of these locations, 0 to 65535, is called a "memory address".

const ADDRESS_SPACE: usize = 65536;

The job of a CPU is to load numbers from memory, do math or logic on them, then write the resulting number back into memory.

The CPU needs a temporary place to hold numbers while it is working with them.

In most CPUs, this place is called a "register". Registers work like variables in a programming language but there are only a few of them (most CPUs have between 1 and 32).

  • On 64-bit ARM the registers are named r0, r1, ..., r15.
  • On 64-bit Intel they are instead named rax, rbx, ....

Just in case those names ring any bells.

Having immediate access to dozens of registers is quite handy, but it means many choices are available to the programmer, or more likely, to the compiler. And making good choices is Hard. A lot of work goes into deciding what variable to store in what register ("register allocation") and when to dump register contents back into memory ("spilling").

Our CPU avoids these problems by not having registers; instead we store numbers in a stack.

  • Putting a number onto the top of the stack is called "push".
  • Taking the most recent number off the top of the stack is called "pop".

The CPU can only access the value that was most recently pushed onto the stack. This may seem like a big limitation right now but you will see ways of dealing with it.

The choice to use a stack instead of registers makes our CPU a "stack machine" as opposed to a "register machine".

Defining a stack

This stack is fixed-size and can hold N values.

#[derive(Debug)]
struct Stack<const N: usize> {
    mem: [u16; N],
    tos: usize  /* top-of-stack */
}

First we'll need a function to add a number to the stack.

When a fixed-size stack fills up, there is a failure case (stack overflow) that must be handled somehow.

This particular stack is a circular stack, meaning that if it ever fills up, it will discard the oldest entry instead of signaling an error. The lack of error handling makes the CPU simpler.

impl<const N: usize> Stack<N> {
    fn push(&mut self, val: u16) {
        self.tos = (self.tos.wrapping_add(1)) & (N - 1);
        self.mem[self.tos] = val;
    }

We'll also need a function to remove & return the most recently pushed number.

    fn pop(&mut self) -> u16 {
        let val = self.mem[self.tos];
        self.mem[self.tos] = 0;

You don't have to set the value back to zero. I am only doing this because it makes makes the stack look nicer when dumped out with print!().

        self.tos = (self.tos.wrapping_sub(1)) & (N - 1);
        return val;
    }

Finally, here is a function that creates a new stack. Because these are circular stacks it doesn't matter where top-of-stack (tos) starts off pointing. I arbitrarily set it to the highest index so the first value pushed will wind up at index 0, again because this makes the stack look nicer when printed out.

    fn new() -> Stack<N> {
        return Stack {tos: N-1, mem: [0; N]};
    }
}

Designing a stack CPU

Now that we have a stack let's use one in our CPU! Or two?

Why two stacks?

The first stack is called the "data stack" and is used instead of registers, as already described.

The second stack will be called the "return stack". This one holds subroutine return addresses. Don't worry if you don't know what that means; we'll get to it later when we talk about the instruction set.

In addition to stacks we are going to give the CPU a couple more things:

  1. An "instruction pointer", which holds the memory address of the next instruction that the CPU will execute.

  2. To make life simpler we put main memory straight on "the CPU" even though in a real computer, RAM would be off-chip and accessed through a data bus.

In our memory, each of the 65536 possible memory addresses will store one 8-bit byte (u8 data type in Rust). This makes it a 65536 byte (64 KB) memory. We could have chosen to make each memory address store 16-bits instead. That would make this a "word-addressed memory". Instead we are going with the "byte-addressed memory" that is more conventional in today's computers. This choice is arbitrary.

Let's add those things to the CPU.

struct Core {
    ram: [u8; ADDRESS_SPACE],
    ip: u16,  /* instruction pointer */
    dstack: Stack<16>, /* data stack */
    rstack: Stack<32>  /* return stack */
}

Finally, let's write a function that creates and returns a CPU for us to use.

use std::convert::TryInto;

impl Core {
    fn new() -> Core {
        return Core {
            ram: [0; ADDRESS_SPACE],
            ip: 0,
            dstack: Stack::new(),
            rstack: Stack::new()}
    }

1.1 - The instruction set

Now we have a CPU sitting there but it does nothing.

A working CPU would execute a list of instructions. An instruction is a number that is a command for the CPU. For example:

  • 65522 might mean "add the top two values on the data stack".
  • 65524 might mean "invert the bits of the top value on the data stack".

The map of instruction-to-behavior comes from the CPU's "instruction set" i.e. the set of all possible instructions and their behaviors.

Normally you program a CPU by putting instructions into memory and then telling the CPU the memory address where it can find the first instruction.

The CPU will:

  1. Fetch the instruction (load it from memory)
  2. Decode the instruction (look it up in the instruction set)
  3. Execute that instruction (do the thing the instruction set said to do)
  4. Move on to the next instruction and repeat.

So now we will make the CPU do those things. We'll start off by teaching it how to access memory, and then we will define the instruction set.

Memory access

Now let's write a function to read a number from the specified memory address.

    fn load(&self, addr: u16) -> u16 {

We immediately run into trouble because we are using byte-addressed memory as mentioned earlier.

Each memory location stores 8 bits (a byte)

Our CPU operates on 16 bit values and we want each memory operation to read/write 16 bits at a time for efficiency reasons.

What do we do?

This CPU chooses to do the following:

  • Read the low byte of the 16-bit number from address a
  • Read the high byte of the 16-bit number from address a+1
16 bit number in CPU: [00000000 00000001]        = 1
                       |        |
                       |        memory address a = 1
                       |
                       memory address a+1        = 0

This is called "little endian" because the low byte comes first.

We could have just as easily done the opposite:

  • Read the high byte of the 16-bit number from address a
  • Read the low byte of the 16-bit number from address a+1
16 bit number in CPU: [00000000 00000001]          = 1
                       |        |
                       |        memory address a+1 = 1
                       |
                       memory address a            = 0

This is called "big endian" because the high byte comes first.

The "le" in the function call below stands for little-endian.

        let a = addr as usize;
        return u16::from_le_bytes(self.ram[a..=a+1].try_into().unwrap());
    }

Writing to memory is very similar, it just works in the opposite direction.

    fn store(&mut self, addr: u16, val: u16) {
        let a = addr as usize;
        self.ram[a..=a+1].copy_from_slice(&val.to_le_bytes());
    }

With that taken care of, we can get around to defining the CPU's instruction set.

Designing the instruction set

Each instruction on this CPU will be the same size, 16 bits, for the following reasons:

  1. Instruction fetch always completes in 1 read. You never have to go back and fetch more bytes.

  2. If you put the first instruction at an even numbered address then you know all the rest of the instructions will also be at even numbered addresses. I will take advantage of this later.

  3. A variable length encoding would save space but 2 bytes per instruction is already pretty small so it doesn't matter very much.

Here are the instructions I picked.

The CALL instruction

CALL
------------------------------------------------------------+----
| n   n   n   n   n   n   n   n   n   n   n   n   n   n   n | 0 |
------------------------------------------------------------+----
What CALL does
  • Push instruction pointer onto the return stack.
  • Set instruction pointer to address nnnnnnnnnnnnnnn0.

This lets you call a subroutine at any even numbered address from 0 to 65534.

Why this is useful

Together with the return stack, CALL lets you call subroutines.

A subroutine is a list of instructions that does something useful and then returns control to the caller.

For example:

Address   Instruction   Meaning
100 ->           200    Call 200
102 ->           ???    Add the top two values on the data stack.
...
200 ->           ???    Push the value 3 onto the data stack
202 ->           ???    Push the value 4 onto the data stack
204 ->           ???    Return to caller

Don't worry about the other instructions I am using here. I will define them later.

I mostly want to point out the three instructions that I put at address 200 because they are a subroutine, a small self contained piece of code (6 bytes) that performs a specific task.

Do you think it's cool that you can count exactly how many bytes it took? I think it's cool.

Here is what happens when the CPU begins execution at address 100.

Address   Data stack   Return stack
100       []           []    <--- About to call subroutine...
200       []           [102]
202       [3]          [102]
204       [3 4]        [102] <--- About to return from subroutine...
102       [3 4]        []
104       [7]          []

The return stack is there to make sure that returning from a subroutine goes back to where it came from. We will talk more about the return stack later when we talk about the RET instruction.

Limitations of CALL:

This CPU cannot call an instruction that starts at an odd address.

At first this seems like a limitation, but it really isn't. If you put the first instruction at an even numbered address then all the rest of the instructions will also be at even numbered addresses. So this works fine.

Of course if you intersperse instructions and data in memory...

           _________
 ________ |_________| _____________
|________|    Data   |_____________|
Instructions         More instructions

...then you will have to be careful to make sure the second block of instructions also starts at an even numbered address. You might need to include an extra byte of data as "padding".

Data processing instructions

Data processing instructions
--------------------------------------------+---------------+----
| 1   1   1   1   1   1   1   1   1   1   1 | x   x   x   x | 0 |
--------------------------------------------+---------------+----

Sixteen of the even numbers are reserved for additional instructions that will be be described later.

The even numbers 1111111111100000 to 1111111111111110 (65504 to 65534) are reserved for these instructions. This means that CALL 65504 through CALL 65534 are not possible. Put another way, it is not possible to call a subroutine living in the top 32 bytes of memory. This is not a very severe limitation.

The LITERAL instruction

LITERAL
------------------------------------------------------------+----
| n   n   n   n   n   n   n   n   n   n   n   n   n   n   n | 1 |
------------------------------------------------------------+----
What LITERAL does
  • Place the value 0nnnnnnnnnnnnnnn on the data stack.
Why this is useful:

Programs will often need to deal with constant numbers. For example, you might want to add 2 to a memory address (to move on to the next even-numbered address) or add 32 to a character code (to convert an uppercase letter to lowercase). These constants have to come from somewhere.

Limitations of LITERAL:

To differentiate it from a call, this instruction is always an odd number. The trailing 1 is discarded before placing the number on the data stack. This missing bit means that only 2^15 values can be represented (0 to 32767). 32768 on up cannot be stored directly. You would need to do some follow-up math to get these numbers. The most direct way is to use the INV instruction, described later.

Making the CPU run

Now that the instruction set is generally described let's look at the code that implements it.

    fn step(&mut self) {

        /* 1. Fetch the instruction.
         * Also advance ip to point at the next instruction for next time. */

        let opcode = self.load(self.ip);
        self.ip = self.ip.wrapping_add(2);

        /* 2. Decode and execute the instruction */

        if (opcode >= 0xffe0) && (opcode & 1 == 0) {

            /* Data processing instruction */

            PRIMITIVES[((opcode - 0xffe0) >> 1) as usize](self);

            /* These instructions get looked up in a table.  The bit
             * math converts the instruction code into an index in the
             * table as follows:
             *
             * 0xffe0 --> 0
             * 0xffe2 --> 1
             * ...
             * 0xfffe --> 15
             *
             * The table will be described below, and these instructions
             * explained.
             */

        }
        else if (opcode & 1) == 1 {
            /* Literal */
            self.dstack.push(opcode >> 1);
        }
        else {
            /* Call */
            self.rstack.push(self.ip);
            self.ip = opcode;
        }
    }
}

The CALL and LITERAL instructions are directly handled above.

The 16 data processing instructions are each assigned a number in the appropriate range that we carved out for them...

enum Op {
    RET = 0xffe0, TOR = 0xffe2, RTO = 0xffe4, LD  = 0xffe6,
    ST  = 0xffe8, DUP = 0xffea, SWP = 0xffec, DRP = 0xffee,
    Q   = 0xfff0, ADD = 0xfff2, SFT = 0xfff4, OR  = 0xfff6,
    AND = 0xfff8, INV = 0xfffa, GEQ = 0xfffc, IO  = 0xfffe,
}

...which is then looked up in the table below. This table gives each instruction its unique behavior.

type Primitive = fn(&mut Core);

const PRIMITIVES: [Primitive; 16] = [

Return-stack instructions

    | x | {
        /* RET - Return from subroutine */
        x.ip = x.rstack.pop()
    },
    | x | {
        /* TOR - Transfer number from data stack to return stack */
        x.rstack.push(x.dstack.pop())
    },
    | x | {
        /* RTO - Transfer number from return stack to data stack */
        x.dstack.push(x.rstack.pop())
    },

Memory instructions

    | x | {
        /* LD - Load number from memory address specified on the data stack */
        let a = x.dstack.pop();
        x.dstack.push(x.load(a));
    },
    | x | {
        /* ST - Store number to memory address specified on the data stack */
        let a = x.dstack.pop();
        let v = x.dstack.pop();
        x.store(a, v);
    },

Stack shuffling instructions

Remember the problem of "register allocation" mentioned earlier, and how stack machines are supposed to avoid that problem? Well, nothing comes for free. Stack machines can only process the top value(s) on the stack. So sometimes you will have to do some work to "unbury" a crucial value and move it to the top of the stack. That's what these instructions are for.

Their use will become more obvious when we start programming the machine, soon.

    | x | {
        /* DUP - Duplicate the top number on the data stack */
        let v = x.dstack.pop();
        x.dstack.push(v);
        x.dstack.push(v);
    },
    | x | {
        /* SWP - Exchange the top two numbers on the data stack */
        let v1 = x.dstack.pop();
        let v2 = x.dstack.pop();
        x.dstack.push(v1);
        x.dstack.push(v2);
    },
    | x | {
        /* DRP - Discard the top number on the data stack */
        let _ = x.dstack.pop();
    },

Conditional skip instruction

We only have one of these: "Q". This is the only "decision-making" instruction that our CPU has. This means that all "if-then" logic, counted loops, etc. will be built using Q.

    | x | {
        /* Q - If the top number on the data stack is zero, skip the next
         * instruction. */

        let f = x.dstack.pop();
        if f == 0 {
            x.ip = x.ip.wrapping_add(2)
        }
    },

Because all of our instructions are two bytes, adding two to the instruction pointer skips the next instruction.

Arithmetic and logic

    | x | {
        /* ADD - Sum the top two numbers on the data stack. */
        let v1 = x.dstack.pop();
        let v2 = x.dstack.pop();
        x.dstack.push(v1.wrapping_add(v2));
    },
    | x | {
        /* SFT - Bit shift number left or right by the specified amount.
         * A positive shift amount will shift left, negative will shift right.
         */

        let amt = x.dstack.pop();
        let val = x.dstack.pop();
        x.dstack.push(
            if amt <= 0xf {
                val << amt
            } else if amt >= 0xfff0 {
                val >> (0xffff - amt + 1)
            } else {
                0
            }
        );
    },
    | x | { // OR - Bitwise-or the top two numbers on the data stack.
        let v1 = x.dstack.pop();
        let v2 = x.dstack.pop();
        x.dstack.push(v1 | v2);
    },
    | x | { // AND - Bitwise-and the top two numbers on the data stack.
        let v1 = x.dstack.pop();
        let v2 = x.dstack.pop();
        x.dstack.push(v1 & v2);
    },
    | x | { // INV - Bitwise-invert the top number on the data stack.
        let v1 = x.dstack.pop();
        x.dstack.push(!v1);
    },

You can use the INV instruction to compensate for the LITERAL instruction's inability to encode constants 32768 to 65535, a.k.a. the signed negative numbers.

Use two instructions instead:

  • LITERAL the complement of your desired constant
  • INV

For example,

  • LITERAL(0) INV yields 65535 (signed -1)
  • LITERAL(1) INV yields 65534 (signed -2)
  • etc.
    | x | { // GEQ - Unsigned-compare the top two items on the data stack.
        let v2 = x.dstack.pop();
        let v1 = x.dstack.pop();
        x.dstack.push(if v1 >= v2 { 0xffff } else { 0 });
    },

Input/output

The CPU needs some way to communicate with the outside world.

Some machines use memory mapped IO where certain memory addresses are routed to hardware devices instead of main memory. This machine already has the full 64K of memory connected so no address space is readily available for hardware devices. Instead we define a separate input-output space of 65536 possible locations. Each of these possible locations is called an IO "port".

    | x | { // IO - Write/read a number from/to input/output port.
        let port = x.dstack.pop();

For a real CPU you could hook up hardware such as a serial transmitter that sends data to a computer terminal, or just an output pin controller that is wired to a light bulb.

This is a fake software CPU so I am going to hook it up to stdin and stdout.

        use std::io;
        use std::io::Read;
        use std::io::Write;

I'm loosely following a pattern in which even ports are inputs and odd ports are outputs. But each port acts different. In a hardware CPU this would not be suitable but it is fine for a software emulation.

        match port {
            0 => {
                /* Push a character from stdin onto the data stack */
                let mut buf: [u8; 1] = [0];
                let _ = io::stdin().read(&mut buf);
                x.dstack.push(buf[0] as u16);
                /* You are welcome to make your own computer that supports
                 * utf-8, but this one does not. */
            }
            1 => {
                /* Pop a character from the data stack to stdout */
                let val  = x.dstack.pop();
                print!("{}", ((val & 0xff) as u8) as char);
                let _ = io::stdout().flush();
            }
            2 => {
                /* Dump CPU status.
                 * Like the front panel with the blinking lights that Chuck
                 * talked about. */
                println!("{:?} {:?}", x.ip, x.dstack);
                let _ = io::stdout().flush();
            }
            _ => {}
        }
    }

That's all the CPU instructions we'll need.

];

Part 2 - The Program

You now have an unfamiliar computer with no software. Can you and the computer write a program?

The first program is the hardest to write because you don't have any tools to help write it. The computer itself is going to be no help. Without any program it will sit there doing nothing.

What should the first program be? A natural choice would be a tool that helps you program more easily.

An interactive programming environment needs to let you do 2 things:

  1. Call subroutines by typing their name at the keyboard
  2. Define new subroutines in terms of existing ones

Begin with step 1: Call subroutines by typing their name at the keyboard

This is where we will meet Forth.

Our interactive programming environment will be a small language in the Forth family. If you want to learn how to implement a full featured Forth, please read Jonesforth, and Brad Rodriguez' series of articles "Moving Forth". The small Forth I write below will probably help you understand those Forths a little better.

Forth organizes all the computer's memory as a "dictionary" of subroutines. The point of the dictionary is to give each subroutine a name so you can run a subroutine by typing its name. The computer will look up its address for you and call it.

Designing the Forth dictionary

The dictionary starts at a low address and grows towards high addresses. It is organized as a linked list, like this:

[Link field][Name][Code .......... ]
 ^
 |
[Link field][Name][Code ...... ]
 ^
 |
[Link field][Name][Code ............... ]

The reason it is a linked list is to allow each list entry to be a different length.

Each dictionary entry contains three things:

  • "Link field": The address of the previous dictionary entry. For the first dictionary entry this field is 0.

  • "Name": A few letters to name this dictionary entry. Later you will type this name at the keyboard to call up this dictionary entry.

  • "Code": A subroutine to execute when you call up this dictionary entry. This is a list of CPU instructions. Note that one of the CPU instructions is "call". So you can have a subroutine that call other subroutines, or calls itself. This code should end with a return (RET) instruction. Here is an example subroutine:

Number Instruction  Meaning
------ -----------  -------
7      Literal(3)   Push the value 3 onto the data stack
9      Literal(4)   Push the value 4 onto the data stack
65504  RET          Return to caller

A linked list is not a very fast data structure but this doesn't really matter because dictionary lookup doesn't need to be fast. Lookups are for converting text you typed at the keyboard to subroutine addresses. You can't type very fast compared to a computer so this lookup doesn't need to be fast.

In addition to the linked list itself, you will need a couple of variables to keep track of where the dictionary is in memory:

  • Dictionary pointer: The address of the newest dictionary entry.
  • Here: The address of the first unused memory location, which comes just after the newest dictionary entry.
[Link field][Name][Code .......... ]
 ^
 |
[Link field][Name][Code ...... ]
 ^
 |
[Link field][Name][Code ............... ]
 ^                                       ^
 |                                       |
[Dictionary pointer]                    [Here]

Tools for building the Forth dictionary

If you were sitting in front of a minicomputer in 196x you would need to create the dictionary with pencil and paper, but in 20xx we will write a Rust program to help create the dictionary.

First we need to keep track of where the dictionary is:

struct Dict<'a> {
    dp: u16,   // The dictionary pointer
    here: u16, // The "here" variable
    c: &'a mut Core  // The dictionary lives in memory.  We are going to
                     // hang on to a mutable reference to the core to give
                     // us easy access to the memory.
}

Now we can write functions in Rust to help us build the dictionary.

enum Item {
    Literal(u16),
    Call(u16),
    Opcode(Op)
}
impl From<u16> for Item { fn from(a: u16) -> Self { Item::Call(a) } }
impl From<Op>  for Item { fn from(o: Op)  -> Self { Item::Opcode(o) } }

impl Dict<'_> {

    /* Helper to reserve space in the dictionary by advancing the "here"
     * pointer */

    fn allot(&mut self, n: u16) {
        self.here = self.here.wrapping_add(n);
    }

    /* Helper to append a 16 bit integer to the dictionary */

    fn comma(&mut self, val: u16) {
        self.c.store(self.here, val);
        self.allot(2);
    }

    /* Helper to append a CPU instruction to the dictionary */

    fn emit<T: Into<Item>>(&mut self, val: T) {
        match val.into() {
            Item::Call(val)    => { self.comma(val) }
            Item::Opcode(val)  => { self.comma(val as u16) }
            Item::Literal(val) => { assert!(val <= 0x7fff);
                                    self.comma((val << 1) | 1) }
        }
    }

    /* Helper to append a "name" field to the dictionary. */

The "name" field bears a closer look. To make each dictionary header a consistent size, I am choosing to not store every letter of the name. Instead I am storing only the length of the name and then the first three letters of the name.

That means these two names will compare equal:

  • ALLOW (-> 5ALL)
  • ALLOT (-> 5ALL)

Even though their first three letters are the same, these two names will compare unequal because they are different lengths:

  • FORTH (-> 5FOR)
  • FORGET (-> 6FOR)

If a name is shorter than 3 letters it is padded out with spaces.

  • X (-> 1X )

You can see that the name field is always four bytes regardless of how many letters are in the name, and the link field is two bytes. This means a dictionary header in this Forth is always six bytes.

    fn name(&mut self, n: u8, val: [u8; 3]) {
        /* Store the length and the first character */
        self.comma(n as u16 | ((val[0] as u16) << 8));

        /* Store the next two characters */
        self.comma(val[1] as u16 | ((val[2] as u16) << 8));
    }

Finally, a function that appends a new link field to the dictionary, pointing to the previous dictionary entry.

    /* Helper to append a new link field to the dictionary and update the
     * dictionary pointer appropriately. */

    fn entry(&mut self) {
        let here = self.here;
        self.comma(self.dp);
        self.dp = here;
    }
}

Now we can start building the dictionary.

To create our Forth interactive programmming environment, we will start by defining subroutines that:

  • read names from the keyboard
  • look up and execute dictionary entries by name

We will put these subroutines themselves in the dictionary so they are available for use once our interactive environment is up and running!

Building the Forth dictionary

fn build_dictionary(c: &mut Core) {
    use Op::*;
    use Item::*;

    let mut d = Dict {
        dp: 0,  /* Nothing in the dictionary yet */

        here: 2,  /* Reserve address 0 as the "reset vector", i.e. where the
                     CPU will jump to start running Forth.  We don't have a
                     Forth interpreter yet so we'll leave address 0 alone for
                     now and start the dictionary at address 2 instead. */
        c: c
    };

Subroutine threading

Consider the following facts:

  • The CPU knows how to execute a bunch of instructions strung together.
  • Forth consists of a bunch of subroutine calls strung together.
  • Subroutine CALL is a valid instruction of our CPU.

This means that we can immediately begin programming our machine in a language resembling Forth, just by writing a list of subroutine calls into the dictionary.

The line between "machine code program" and "Forth program" is very blurry. To illustrate:

Here is a subroutine consisting of a few instructions strung together.

Instruction Number  Meaning
----------- ------  -------
Literal(3)  7       Push the value 3 onto the data stack
Literal(4)  9       Push the value 4 onto the data stack
RET         65504   Return to caller

Here is a Forth subroutine consisting of a few subroutine calls strung together.

Call        Number  Meaning
----------- ------  -------
S1          1230    Call subroutine S1 which happens to live
                    at address 1230
S2          1250    Call subroutine S2 which happens to live
                    at address 1250
RET         65504   Return to caller

Both of these are valid machine code programs (list of numbers that our CPU can directly execute).

This duality between CPU instructions and Forth code comes from an idea called "subroutine threading". It is a refinement of an idea called "threaded code". This has no relation to the kind of threading that lets you run programs in parallel. You can read more about threaded code on Wikipedia or in the other Forth resources I mentioned earlier (Jonesforth, and Moving Forth by Brad Rodriguez).

Our new language starts out with the sixteen (well, eighteen) instructions built into the CPU. We can string those instructions together into a new subroutine. Each new subroutine adds to the toolbox we have available for making the next new subroutine. Repeat until you have built what you wanted to build, via function composition. This is the idea behind Forth.

We are going to be writing many series of instructions so let's start out by making a Rust macro that makes them easier to type and lets us specify a CPU instruction vs. a subroutine call with equal ease.

The macro below will convert:

  • forth!(Literal(2), ADD, RET);

to:

  • d.emit(Literal(2));
  • d.emit(ADD);
  • d.emit(RET);

which you probably recognize as code that will add a new subroutine to the dictionary.

    macro_rules! forth {
        ($x:expr) => (d.emit($x));
        ($x:expr, $($y:expr),+) => (d.emit($x); forth!($($y),+))
    }

Now we can add the first subroutine to the dictionary!

key

"key" reads a character from the keyboard and places its character code on the stack.

There is a tradition of writing stack comments for Forth subroutines to describe the stack effect of executing the subroutine. They look like this:

key ( -- n )

Read as: key does not take any parameters off the stack, and leaves one new number pushed onto the stack.

Also remember that a dictionary entry looks like this:

[Link field][Name][Code .......... ]

Given all of the above, we are now ready to define "key" and add it to the dictionary.

    // key ( -- n )
    d.entry();           /* Compile the link field into the dictionary */
    d.name(3, *b"key");  /* Compile the name field into the dictionary */
    let key = d.here;    /* (Save off the start address of the code so we
                            can call it later) */
    forth!(
        Literal(0),      /* Compile a LITERAL instruction that pushes
                            0 to the stack */

        IO,              /* Compile an IO instruction.
                          *
                          * Remember from the CPU code that IO takes a
                          * parameter on the stack to specify which port
                          * to use.
                          *
                          * Also remember that IO port 0 reads
                          * a character from standard input.
                          */

        RET              /* Compile a RET instruction */
    );

We have now compiled the "key" subroutine into the dictionary. It takes twelve bytes of memory and is laid out as shown below:

[Link field][Name][Code .......... ]
       0000  3key  1, 65534, 65504

emit

The next subroutine we will make is "emit". This is a companion to "key" that works in the opposite direction.

  • key ( -- n ) reads a character from stdin and pushes it to the stack.
  • emit ( n -- ) pops a character from the stack and writes it to stdout.
    // emit ( n -- )
    d.entry(); d.name(4, *b"emi");  let emit = d.here;
    forth!(
        Literal(1),
        IO,
        RET);

I am tired of saying "subroutine" so many times, so I am going to introduce a new term. Remember the goal our language is working towards -- we want to be able to type a word at the keyboard, and let the computer look it up in the dictionary and execute the appropriate code.

So far we have two named items in the dictionary, call and emit.

We are going to term a named dictionary item a "word". This is a Forth tradition. So call and emit are "words", or "dictionary words" if you want to be precise about it. So far these are the only words we've defined.

Let's define some more words.

- (subtraction)

Our CPU does not have subtraction so let's make subtraction by adding the two's complement.

To get the two's complement, do a bitwise invert and add 1.

This will be the most complicated Forth that we've written so far so let's walk through step by step.

    // - ( a b -- a-b )
    d.entry(); d.name(1, *b"-  ");  let sub = d.here;
    forth!(         /* Stack contents:  a b, to start off with.
                     * We want to compute a minus b */

        INV,        /* Bitwise invert the top item on the stack.
                     * Stack contents: a ~b */

        Literal(1), /* Push 1 onto the stack.
                     * Stack contents: a ~b 1 */

        ADD,        /* Add the top two items on the stack.
                     * Stack contents: a ~b+1
                     * Note that ~b+1 is the two's complement of b. */

        ADD,        /* Add the top two items on the stack.
                     * Stack contents: n
                     * Note that n = (a + ~b+1) = a - b */

        RET         /* Done, return to caller, leaving n on the data stack. */
    );

Writing it out like that takes a lot of space. Normally Forth code is written on a single line, like this:

INV 1 ADD ADD RET

Looking at it this way, it's easy to see the new word we just created (-) is made from 5 instructions. It's pretty typical for a Forth word to be made of 2-7 of them. Beyond that length, things get successively harder to understand, and it becomes a good idea to split some work off into helper words.

We will see an example of this below.

0= (compare-to-zero)

Our next word will be useful for Boolean logic.

0= ( n -- f )

In a stack comment, "f" means "flag", a.k.a. Boolean value. By Forth convention, zero is false and any nonzero value is true. However the "best" value to use for a true flag is 65535 (all ones) so the bitwise logical operations can double as Boolean logical operations.

So what 0= does is:

  • if n=0, leave on the stack f=65535
  • otherwise, leave on the stack f=0

It is like C's ! operator.

In Rust this could be implemented as:

// example code, not part of our program
fn zero_eq(n: u16) -> u16 {
    if (n == 0) {
        return 65535;
    } else {
        return 0;
    }
}

Rust has an if-then and block scope, so this is easy to write.

The literal translation to a typical register-machine assembly language would look something like this:

zero_eq:     compare r0, 0
             jump_eq is_zero
             move    r0, 0
             ret
is_zero:     move    r0, 65535
             ret

It looks simple but I want to point out a couple things about it that are not so simple.

The conditional jump instruction, jump_eq.

Our CPU doesn't have this. The only decision-making instruction we have is Q which is a conditional skip.

Q - If the top number on the data stack is zero, skip the next instruction.

A conditional jump can go anywhere. A conditional skip can only decide whether or not to skip the next instruction (i.e., it is a fixed forward jump of 2 bytes). You cannot give Q a specific address to jump to, the way jump_eq worked.

So our CPU does not make it easy to jump around in a long block of instructions -- our CPU prefers that you use subroutine calls.

The forward reference

This is another problem. Think of the job of an assembler which is converting an assembly language program to machine code. We are currently writing our code in a tiny assembler that we made in Rust! It is very simple but so far it has worked for us. The assembler of our hypothetical register-machine below has a rather nasty problem to solve.

zero_eq:     compare r0, 0
             jump_eq is_zero  <----- On this line.
             move    r0, 0
             ret
is_zero:     move    r0, 65535
             ret

It wants to emit a jump to is_zero, but that symbol has not been seen yet and is unrecognized. On top of that, the assembler also doesn't yet know what address is_zero will have, so doesn't know what jump target to emit. To successfully assemble that kind of program you would need an assembler smarter than the assembler we made for ourselves in Rust.

There are ways to solve this but let's NOT solve it.

Our CPU has no jump instruction (only call) and our assembler only lets us call things we already defined. Instead of removing these constraints, find a way to write 0= within the constraints.

Here is a start at solving the problem

is_nonzero ( -- 0 )
    Literal(0)
    RET

0= ( n -- f )
    Q            <-- pop n, if n=0 skip next instruction
    is_nonzero   <-- f=0 is now pushed to stack
    Literal(0)
    INV          <-- f=65535 is now pushed to stack
    RET          <-- Return

We got rid of the forward reference by defining is_nonzero before it was used.

We got rid of the jump instruction by using a subroutine call instead.

This code is close to working but it doesn't quite work. The problem is that is_nonzero gives control back to 0= when done, just like a subroutine call normally does, and then 0= runs as normal until it hits the return instruction at the end. So we wind up executing both the f=0 branch and the f=65535 branch, instead of just executing the f=0 branch like we wanted in this case.

It is possible to fix this last problem by adding the instructions RTO DRP to is_nonzero.

is_nonzero ( -- 0 )
    RTO          <-- Pop the return address, push to data stack
    DRP          <-- Discard it
    Literal(0)   <-- Put 0 on the data stack
    RET          <-- Return

Because we popped off and discarded one item from the return stack, the final RET instruction will not return to 0= any more. Instead it will skip one level and return to whoever called 0=. This has the result of ending 0= early, which is what we wanted to do.

0= ( n -- f )
    Q            <-- pop n, if n=0 skip next instruction
    is_nonzero   <-- this word puts f=0 on the stack then ends 0= early
    Literal(0)
    INV          <-- f=65535 is now pushed to stack
    RET          <-- Return

I call this pattern "return-from-caller". It is used occasionally in real Forth systems. My dialect of Forth will use it extensively to work around my CPU's lack of conditional jump.

Now we've explained how 0= is going to work, let's write it.

0= (compare-to-zero), for real this time

First we define the helper. It won't be reused, so I am not going to bother giving it a dictionary header and name for easy lookup later. Think of it as a private function.

    let zero = d.here;
    forth!(Literal(0), RTO, DRP, RET);

Now define 0= using the helper.

    // 0= ( n -- f )
    d.entry(); d.name(2, *b"0= ");  let zero_eq = d.here;
    forth!(Q, zero, Literal(0), INV, RET);

= (equals)

Next let's make a = equality comparison operator, using 0= and subtract. I call it an "operator" because that's what other languages would call it, but Forth has no special idea of an "operator". Everything is just words.

    // = ( a b -- a=b )
    d.entry(); d.name(1, *b"=  ");  let eq = d.here;
    forth!(sub, zero_eq, RET);

Note that 0= and subtract are both words, not CPU instructions. This makes = the first "pure" Forth word we have defined, with no direct dependency on the machine's instruction set. We could define = as - 0= on a real standards-compliant Forth system and it would still work. So Forth gets you to the point of writing "portable" code really quickly. Often you can reuse routines early in bootstrapping even though they were written and tested on a different machine. Many languages offer portability but few offer it so quickly.

2.1 - The lexer

Now that we've got some basics in place let's go back to solving the real problem of getting our language to read words from the keyboard. The first problem we have is that we need some way to separate words from each other so we know where one word ends and the next begins. This problem is called "lexing". Forth has about the simplest lexer ever, it just splits on whitespace. Anything with character code <=32 is considered whitespace. Words are delimited by whitespace. And that is all the syntax Forth has.

To read a word from the keyboard you will need to:

  1. Advance past any leading whitespace
  2. Read characters into a buffer until whitespace is seen again.

Skipping whitespace

Let's start with the "advance past leading whitespace" part

The "key" word gives us the latest keystroke as an ASCII code. (Really it is reading utf-8 characters one byte at a time but let's not get into that right now, pretend the year is 196x, we're sitting in front of a minicomputer and and utf-8 hasn't been invented yet.)

ASCII codes 0 to 32 are whitespace or control characters. Codes 33 and up are letters, numbers and symbols. So to skip whitespace all you need to do is read keys until you get an ASCII code >= 33, then return that to tell the rest of the program what key code you saw.

In Rust this could be implemented as:

// example code, not part of our program
fn skipws() -> u16 {
    loop {
        let c = key();
        if c >= 33 {
            return c;
        }
    }
}

Rust has a loop keyword, so this is easy to write. (Alarm bells should be ringing in your head at this point because we haven't put any looping constructs in our CPU or language.)

The literal translation to a typical register-machine assembly language would look something like this:

skipws:      call key
             compare r0, 32
             jump_le skipws
             ret

(More alarm bells should be ringing in your head because this is using conditional jump, which our CPU doesn't have.)

Like last time, is there a way to solve this without conditional jump?

Here is a start at solving the problem:

skipws ( -- c )
    key          <-- Put keycode on the stack:           ( c )
    DUP          <-- Duplicate top value on the stack:   ( c c )
    Literal(33)  <-- Put 33 on the stack:                ( c c 33 )
    GEQ          <-- Is c >= 33?                         ( c f )
    Q            <-- If so...
    RET          <-- ... return, leaving c on the stack. ( c )
    DRP          <-- Discard c from the stack.           ( )
    skipws       <-- Call skipws again

You will notice there is no RET statement at the end of skipws. At the end of skipws we call skipws again. This makes an infinite loop. The only way out of the loop is the RET instruction in the middle. This works similarly to the Rust code that uses a loop { } and breaks out when it sees the condition it's looking for.

Writing a word that calls itself is called "recursion".

This code almost works but there is still something wrong with it. Youll notice we were careful to make sure "skipws" removed all items it added to the data stack, before it called itself. Its last two lines were:

DRP    <-- Discard c from the stack
skipws <-- Call skipws again

If we didn't do that, skipws would leave each whitespace character it saw, on the data stack, as it looped again and again. So instead of returning the first nonwhitespace character it would return EVERY character it saw.

1st recursion: data stack: ( c1 )
2nd recursion: data stack: ( c1 c2 )
3rd recursion: data stack: ( c1 c2 c3 )

There are problems with this. It's messy. The caller has no idea how many values we are going to leave on the stack, so has no idea how many to pop off. Also, we might see more than 16 whitespace characters in a row, which would make weird things happen because our CPU's data stack only has room for 16 numbers.

For these reasons it's better to leave the data stack as we found it, when we do a recursive call. That is the reason the last two lines are DRP, skipws -- it's to stop items building up on the data stack. The final pass through this function goes down a different path that does not DRP, so it leaves something on the data stack -- the last key read.

The problem skipws still has, is that we haven't taken the same care with its return stack.

At the first line of skipws the return stack looks like this:

( caller )

That's because skipws must have been called by our CPU's CALL instruction (we have no other way of calling subroutines!), and the CALL instruction leaves a return address on the top of the return stack so RET knows where to return to at the end of the subroutine.

But we are also using CALL for a different purpose: to repeat skipws. Every time we repeat skipws, the CALL instruction will push another return address to the call stack.

DRP                                 return stack:( caller )
skipws       <-- Call skipws again. return stack:( caller x )
<-- This location has address x.

first call:    return stack: ( caller )
1st recursion: return stack: ( caller x )
2nd recursion: return stack: ( caller x x )
3rd recursion: return stack: ( caller x x x )

Clearly all these x's are garbage. When we are done with skipws we want to return to our caller, not to x.

We could patch over the problem somewhat by putting a RET instruction at x.

DRP                                 return stack:( caller )
skipws       <-- Call skipws again. return stack:( caller x )
RET          <-- x

This yields working recursive code.

Each time we loop, a useless return address x is left on the return stack. When skipws wants to quit, skipws runs a RET instruction, which transfers control to x. x is the address of a RET instruction, left on the stack earler. So we wind up running RET RET RET ... until we burn through all x's on the return stack and finally transfer control back to caller.

first call:    return stack: ( caller )         data stack: ( )
1st recursion: return stack: ( caller x )       data stack: ( )
2nd recursion: return stack: ( caller x x )     data stack: ( )
3rd recursion: return stack: ( caller x x x )   data stack: ( c )
RET:         : return stack: ( caller x x )     data stack: ( c )
RET:         : return stack: ( caller x )       data stack: ( c )
RET:         : return stack: ( caller )         data stack: ( c )
RET:         < control is passed back to our caller,
               and now they can do stuff with the "c" on the data
               stack, yay >

This works. It isn't very fast but we don't care about speed right now, just about getting our computer to work.

But there is still a problem.

Our CPU has a fixed-size circular return stack that can hold 32 numbers. What happens if you loop 32 times or more? The return stack fills up completely with the useless "x" addresses, and the address of caller is lost.

recursive call N  :  return stack: ( caller x x x ... x )
recursive call N+1:  return stack: (      x x x x ... x )  :-(

So skipping 32 or more whitespace characters in a row wouldn't work. To fix that problem we need to find a way to stop the useless "x" addresses from building up on the return stack.

1st loop: return stack: ( caller )   data stack: ( )
2nd loop: return stack: ( caller )   data stack: ( )
3rd loop: return stack: ( caller )   data stack: ( c )
RET:      < control is passed back to our caller >

The most common solution is "tail call optimization". If a function's last instruction is a recursive call, that call can be replaced with a jump. On paper this doesn't work very well on our computer, for two reasons:

  1. Our CPU has no jump, only call.

  2. Our assembler, and eventually our interactive environment, would need to be smart enough to emit a call sometimes and a jump other times. This is the same "look-ahead" problem that we saw with forward references -- you don't know that a given CALL will be followed by a RET, unless you can see the future.

    Earlier we decided to keep our assembler very dumb so it would be weird to start making it smart now.

So what are we going to do?

It is possible to get a very, very dumb caveman version of tail call optimization, by manually using the "return-from-caller" trick, RTO DRP, to "get rid of" the x that is pushed on by the skipws CALL.

skipws ( -- c ) RTO DRP ... Q RET ... skipws

1st loop: return stack: ( caller )   data stack: ( )
2nd loop: return stack: ( )          data stack: ( )
3rd loop: return stack: ( )          data stack: ( )

So now recursive calls will leave the return-stack as they found it, which is good! We don't have the useless-x problem any more. Unfortunately, the first pass through skipws discards the original caller's return address, which we wanted to keep. There is a quick hack around that problem: wrap skipws in another subroutine, and always call it through that wrapper.

skipws ( -- c ) RTO DRP ... Q RET ... skipws

wrapper ( -- c ) skipws RET

The RET in skipws returns from wrapper, but that's ok.

Finally we are able to write loops, and we did not even need to add anything to our language or CPU to get that to work, we just needed to look at things differently. Learning to look at things differently is a big part of the Forth philosophy.

We'll see a better way of solving this problem later, in the file frustration.4th, but for now this is good enough and we can get back to solving our original problem, skipping whitespace.

Skipping whitespace (for real this time)

You should now understand what the next two functions are doing because we just talked about them at length. In the real program I swapped the names of the two functions because I wanted to let the wrapper have the friendly "skipws" name.

    let skip_helper = d.here;
    forth!(RTO, DRP, key, DUP, Literal(33), GEQ, Q, RET, DRP, skip_helper);

    // skipws ( -- c )
    d.entry(); d.name(6, *b"ski");  let skipws = d.here;
    forth!(skip_helper);

Step 1 of the lexer is now working! We can now discard whitespace characters typed at the keyboard, i.e. advance to the first character of a word.

Reading characters into a buffer

The next stage of lexing is once again going to be more complicated than any code we've written before, so we are going to need some more helper words.

Until now, we have been able to structure our code in such a way that the next value we need is conveniently stored at the top of the stack. The most we've had to do is either DUPlicate this value or DRP it because it's no longer needed. In more complicated code, sometimes we will need to "dig through" the values on the stack to surface the one we want to use next. This is inefficient and ugly so we will do it as little as possible, but it will soon be necessary.

The CPU instruction SWP does stack shuffling by swapping the first two values on the data stack. We already have SWP (it's built into the CPU) but I will write out its stack effect below as a recap of what it does.

SWP ( a b -- b a ).

The problem with SWP is that it can only reach the top two values on the stack. If you wanted to dig further, you couldn't do it with SWP.

One way of digging further is by using the RTO and TOR instructions as demonstrated below in the "over" word.

over

    // over ( a b -- a b a )
    d.entry(); d.name(4, *b"ove");  let over = d.here;
    forth!(TOR,  /* data stack: ( a )      return stack: ( caller b ) */
           DUP,  /* data stack: ( a a )    return stack: ( caller b ) */
           RTO,  /* data stack: ( a a b )  return stack: ( caller ) */
           SWP,  /* data stack: ( a b a )  return stack: ( caller ) */
           RET);

"over" is a good building block for further stack shuffling words.

2dup

    // 2dup ( a b -- a b a b )
    d.entry(); d.name(4, *b"2du");  let twodup = d.here;
    forth!(over, over, RET);

The input buffer

Now we can get back to writing the lexer. Step 2 of lexing is "Read characters into a buffer until whitespace is seen again", and once that works we will be done writing the lexer!

Start by setting aside the word input buffer. We'll format it as Nabcde where N is the number of characters stored.

    let word_buf = d.here;
    d.allot(6);

It may seem strange to be plopping this down in the middle of the dictionary but it will work fine, just as long as we're setting aside an even number of bytes. As mentioned earlier, if you intersperse instructions and data in memory...

           _________
 ________ |_________| _____________
|________|    Data   |_____________|
Instructions         More instructions

...then you will have to be careful to make sure the second block of instructions also starts at an even numbered address. You might need to include an extra byte of data as "padding".

In this case we set aside one byte for length and five bytes for characters, which is a total of six bytes, so no padding is needed.

We are about to do some buffer handling so we want bounds checking. Let's write a min-value word. It will look at the top two items on the stack and return whichever is less.

This word is simple enough that I'm not going to walk through it like I did with some of the earlier words. If you want to understand how it works I recommend walking through it on paper or in your head. With a little practice this will become as natural as walking through code in any other language.

min

    // min ( a b -- n )
    d.entry(); d.name(3, *b"min");  let min = d.here;
    forth!(twodup, GEQ, Q, SWP, DRP, RET);

c@ and c! (byte-by-byte memory access)

We want to access the buffer byte-by-byte, but our machine only accesses memory 16 bits at a time.

Reading one byte at a time is pretty easy, just do a 16-bit read and discard the high byte with Literal(0xFF) AND.

    // c@ ( a -- n )
    d.entry(); d.name(2, *b"c@ ");  let cld = d.here;
    forth!(LD, Literal(0xff), AND, RET);

To write one byte at a time, we'll take the approach of reading two bytes, editing just the low byte, and then writing the full two-byte value back to memory. The high byte gets unnecessarily rewritten but we are writing back its old value so no one will know the difference.

If our CPU was multi-core, or had interrupts, there could be some problems with this approach (search the Internet for "non-atomic read-modify-write"), but ours isn't, so we are fine.

    // c! ( n a -- )
    d.entry(); d.name(2, *b"c! ");  let cst = d.here;
    forth!(DUP,                 /* ( n a a )              r: ( caller )   */
           LD,                  /* ( n a old-n )          r: ( caller )   */
           Literal(0xff), INV,  /* ( n a old-n 0xff00 )   r: ( caller )   */
           AND,                 /* ( n a old-highbyte )   r: ( caller )   */
           SWP, TOR,            /* ( n old-highbyte )     r: ( caller a ) */
           OR,                  /* ( new-n )              r: ( caller a ) */
           RTO,                 /* ( new-n a )            r: ( caller )   */
           ST,                  /* ( )                    r: ( caller )   */
           RET);

Filling the input buffer

Now we have everything we need to fill the input buffer one byte at a time:

    /* Load 1 letter into the buffer. */
    let stchar = d.here;
    forth!(Literal(word_buf), cld,  /* Retrieve the first byte of the buffer,
                                       i.e. its current length. */

           Literal(1), ADD,         /* Increment the length. */

           DUP, Literal(word_buf), cst,  /* Write-back the incremented length
                                            to the first byte of the buffer */

           /* Decide where to store the letter in the buffer.
            *
            * The 1st letter should be stored 1 byte past the buffer start
            *   (to leave room for the length).
            *
            * The 2nd letter should be stored 2 bytes past the buffer start
            * ...
            * The 5th letter should be stored 5 bytes past the buffer start.
            *
            * Any letters beyond the 5th will also be stored in the 5th slot
            * overwriting whatever letter was seen there previously.  This
            * is fine because only the first 3 letters of the word are
            * significant anyway.  What's important is that we not overrun
            * the buffer and corrupt adjacent parts of the dictionary.
            */
           Literal(5), min, Literal(word_buf), ADD,

           cst,  /* Store the letter in the buffer */
           RET);

Parsing a whole word is not much harder. Just tail-recursively call the function we just wrote, until whitespace is seen again (a character code that is <= 32).

    let getcs_helper = d.here;
    forth!(RTO, DRP, /* The "return-from-caller" trick */
           stchar,
           key, DUP, Literal(32), SWP, GEQ, Q, RET,
           getcs_helper);

This also returns the whitespace character that was seen, although we won't do much with it.

    // getcs ( -- c )
    d.entry(); d.name(5, *b"get");  let getcs = d.here;
    forth!(getcs_helper, RET);

word

The lexer is almost done, now we'll write the word that the rest of the program will use to call it.

This word is named "word".

First, it clears word_buf by setting its length byte to 0 and padding out the first three name bytes by setting them to 32 (space). Then, reads a word from the keyboard into the word_buf.

    // word ( -- )
    d.entry(); d.name(4, *b"wor");  let word = d.here;
    forth!(
        Literal(word_buf),    /* Address of word_buf */

        DUP, Literal(2), ADD, /* Address of word_buf + 2 */

        Literal(0x2020), SWP, ST, /* Set name bytes 2 and 1 to space */

        Literal(0x2000), SWP, ST, /* Set name byte 0 to space and
                                        set length to zero */

        skipws, /* Lexer step 1, skip leading whitespace */

        getcs,  /* Lexer step 2, read letters into buffer until whitespace
                      is seen again */

        DRP,    /* We don't care what whitespace character was last seen
                      so drop it */
        RET);

The lexer is now complete: we can read space-delimited words from the keyboard.

This took a long while, because we had to figure out how to do things like branching and looping, while also figuring out how to write the lexer itself. But now our dictionary is filled with useful helper words so our next steps will be faster to write.

2.2 - Dictionary lookup

Let's move on to dictionary lookup, so we can do something useful with the space-delimited words we now know how to read from the keyboard.

latest

To do dictionary lookup we first need to keep track of where the dictionary is, so let's teach Forth about the dictionary pointer (dp) variable that we've so far been tracking in Rust.

The traditional Forth name for this variable is "latest".

    // latest ( -- a )
    /* Address of "latest" variable.  This variable stores the address of
     * the latest word in the dictionary. */
    let latest_ptr = d.here; d.allot(2);
    d.entry(); d.name(6, *b"lat");  let latest = d.here;
    forth!(Literal(latest_ptr), RET);

find

Now we will write "find" which is the word that does dictionary lookup. Dictionary lookup is a linked list traversal starting at latest (the end of the dictionary). For each dictionary entry, we compare its name against the name that "word" placed in the input buffer. If it matches, we return the address of this dictionary entry's code field. Otherwise we advance to the previous dictionary entry and try again. If we don't match anything before we hit address 0 (the start of the dictionary) that means the name in the input buffer was not found in the dictionary.

The stack effect of find will be:

find ( -- xt|0 )

It's time to explain a couple more conventions often used in stack effect comments:

  • xt is "execution token". In our Forth, "execution token" just means the address of some code.

  • A vertical bar | means "or". So find will return either an execution token, or 0 if no execution token is found.

    /* Helper word ( a -- f )
     */
    let matches = d.here;
    forth!(
        /* Stash the address of the name field by putting it on the
         * return stack
         */
        Literal(2), ADD, TOR,

        /* Load the 4 bytes at word_buf */
        Literal(word_buf), DUP, Literal(2), ADD, LD, SWP, LD,

        /* Load the first 2 bytes of the name field */
        RTO, DUP, TOR, LD,

        /* Compare to the first 2 bytes at word_buf.
         * Don't worry about that bitwise AND: it will be explained later
         * when we are adding "immediate" words to the outer interpreter.
         */
        Literal(0x0080), INV, AND, eq,

        /* Compare the second 2 bytes of the name field to the second
         * 2 bytes at word_buf
         */
        SWP, RTO, Literal(2), ADD, LD, eq,

        /* If both comparisons were true, return true, else return false */
        AND, RET);

    /* Helper word ( a -- a' )
     */
    let matched = d.here;
    forth!(
        Literal(6), ADD,  /* Advance six bytes (the length of the dictionary
                             header).  This advances from the start of the
                             header to the address of the code field. */

         RTO, DRP,         /* Return-from-caller */
         RET);

    let find_helper = d.here;
    forth!(
        RTO, DRP,
        DUP, Literal(0), eq, Q, RET, /* No match - return 0 */
        DUP, matches, Q, matched,    /* Match - return the code address */
        LD, find_helper);            /* Try the next one */

And find itself is just a wrapper around the tail-recursive find_helper word.

    // find ( -- xt|0 )
    d.entry(); d.name(4, *b"fin");  let find = d.here;
    forth!(latest, LD, find_helper);

' (quote)

The ' (quote) word reads the next word from the keyboard and then looks it up in the dictionary. It works very similarly to the "address-of" operator in C. ' fn in Forth is like &fn in C.

    // ' ( -- xt|0 )
    d.entry(); d.name(1, *b"'  ");  let quote = d.here;
    forth!(word, find, RET);

2.3 - The outer interpreter

We can now look up a subroutine in the dictionary by typing its name at the keyboard.

Remember that an interactive programming environment needs to let you do two things:

  1. Call subroutines by typing their name at the keyboard
  2. Define new subroutines in terms of existing ones

We're also going to succumb to temptation at this point and add a third feature to our language.

    1. Push numbers onto the data stack by typing them at the keyboard

We haven't achieved any of these three goals yet, but we now have all of the building blocks we need to do so.

To add words to the dictionary we'll need to keep track of where the end of the dictionary is, so let's teach Forth about the "here" variable that we've so far been tracking in Rust.

here

    // here ( -- a )
    /* Address of "here" variable.  This variable stores the address of
       the first free space in the dictionary */
    let here_ptr = d.here; d.allot(2);
    d.entry(); d.name(4, *b"her");  let here = d.here;
    forth!(Literal(here_ptr), RET);

Achieving interactivity

Let's talk a little bit about how we are going to make our Forth interactive. We want to do one of two things:

  1. Call subroutines by typing their name at the keyboard
  2. Define new subroutines in terms of existing ones

Both of these things are structurally similar. We can solve either problem by reading a list of words from the keyboard and doing something with each word.

First we look up the word in the dictionary, then we either:

  1. Execute it right now (if we are in interpreting mode).
  2. Append it to the dictionary (if we are in compiling mode).

Numbers can be handled in a similar way. If we encounter a number in interpreting mode, we'll put it on the stack. If we encounter a number in compiling mode, we'll compile a LITERAL instruction that will put the number on the stack when executed.

It seems a pretty good bet that we'll be able to solve our problem with an interpreting/compiling mode flag, so let's make one.

    // state ( -- a )
    /* Address of "state" variable.  This variable stores -1 if
     * interpreting or 0 if compiling. */
    let state_ptr = d.here; d.allot(2);
    d.entry(); d.name(5, *b"sta");  let state = d.here;
    forth!(Literal(state_ptr), RET);

We need a way of switching between interpreting and compiling mode.

If you are interpreting, this is easy -- just write 0 to state.

If you are compiling, it is not so easy to go back into interpreting mode, because everything you type gets compiled. There is no way to execute a word when you are in compiling mode, so you are stuck compiling forever.

What if there was a way to execute a word in compiling mode?

We will define a special category of words called "immediate" words that are executed whenever they are seen, even if you are in compiling mode.

We will mark a word as "immediate" by setting the high bit of the length byte, in the name field of its dictionary entry.

----+---+---+---+---+---+---+---+
| i | n | n | n | n | n | n | n |
----+---+---+---+---+---+---+---+
  • nnnnnnn = length (0 to 127)
  • i = "immediate" bit (1 = immediate, 0 = ordinary)

Do you remember the bit math in "find" that I told you to not worry about just yet?

Literal(0x0080), INV, AND

This math was masking out the "immediate" flag so it would not interfere with dictionary lookup.

immediate

    /* Helper function to get the address of the latest dictionary entry */
    let word_addr = d.here;
    forth!(Literal(latest_ptr), LD, Literal(2), ADD, RET);

    // immediate ( -- )
    /* Set the "immediate" flag on the latest dictionary entry */
    d.entry(); d.name(9, *b"imm");
    forth!(word_addr, DUP, LD, Literal(0x0080), OR, SWP, ST, RET);

[ and ]

Now we can define words to switch between interpreting and compiling mode. The names [ and ] are traditional Forth names.

    // [ ( -- )
    d.entry();

    d.name(
        1 | 0x80,  /* In Rust we do not have access to the handy "immediate"
                      function, but we can make a word "immediate" by setting
                      the high bit in its length field, as is done here. */
        *b"[  ");

    let lbracket = d.here;
    forth!(Literal(0), INV, state, ST, RET);

    // ] ( -- )
    d.entry(); d.name(1 | 0x80, *b"]  ");  let rbracket = d.here;
    forth!(Literal(0), state, ST, RET);

smudge and unsmudge

By setting a different bit of the name field we can temporarily hide a word from name lookups. We will talk more about this later.

    // smudge ( -- )
    d.entry(); d.name(6 | 0x80, *b"smu");  let smudge = d.here;
    forth!(word_addr, DUP, LD, Literal(0x0040), OR, SWP, ST, RET);

    // unsmudge ( -- )
    d.entry(); d.name(8 | 0x80, *b"uns");  let unsmudge = d.here;
    forth!(word_addr, DUP, LD, Literal(0x0040), INV, AND, SWP, ST, RET);

, (comma)

Now let's make a word that appends to the dictionary. We have had a Rust helper function for this for a long time. The word below is the same thing but callable from Forth.

    // , ( n -- )
    d.entry(); d.name(1, *b",  "); let comma = d.here;
    forth!(here, LD, ST,
           here, LD, Literal(2), ADD, here, ST, RET);

number

We will read numbers the same way we read words: from the input buffer. This, incidentally, is why we chose to reserve space for five characters in the input buffer, even though we only needed to store three for word lookup. The largest 16-bit number will fit in five decimal digits.

Our numbers will be base-10. To build up a base-10 number digit by digit, we'll need to be able to multiply by 10. Our CPU has no multiply but it does have bit shift, which can be used to multiply or divide an unsigned value by any power of two.

    // x10 ( n -- n*10 )
    d.entry(); d.name(3, *b"x10");  let x10 = d.here;
    forth!(
        DUP, DUP, Literal(3), SFT, /* Find n*8 */
        ADD, ADD,                  /* (n*8) + n + n = (n*10) */
        RET);

Now we can write a word that goes through the input buffer character by character and converts it to an integer on the stack.

    /* Helper function to clear junk off the stack. */
    let end_num = d.here;
    forth!(DRP, RTO, DRP, RET);

    /* Helper function to clear junk off the stack and return -1. */
    let bad_num = d.here;
    forth!(DRP, DRP, DRP, Literal(0), INV, RTO, DRP, RET);

    // Helper function ( 0 1 -- n|-1 )
    let number_helper = d.here;
    forth!(
        RTO, DRP,
        /* Load the next character */
        DUP, Literal(word_buf), ADD, cld,

        /* If the character is not in the range 48 to 57
         * (which are the character codes for '0' to '9')
         * then this is not a number, so return the error code -1 (65535)
         */
        Literal(48), sub, DUP, Literal(10), GEQ, Q, bad_num,
        SWP, TOR, SWP, x10, ADD, RTO,

        /* If we've come to the end of the input buffer then end. */
        DUP, Literal(word_buf), cld, GEQ, Q, end_num,

        /* Move on to the next digit */
        Literal(1), ADD, number_helper);

    // number ( -- n|-1 )
    d.entry(); d.name(6, *b"num");  let number = d.here;
    forth!(Literal(0), Literal(1), number_helper);

literal

To compile an integer, we'll want to convert it to a LITERAL instruction in the dictionary. Bear in mind that only numbers 0-32767 can be directly stored in a LITERAL instruction. This code makes no attempt to automatically perform the LITERAL INV trick -- that's left up to the programmer.

    /* Compile a number */
    d.entry(); d.name(3, *b"lit");  let lit = d.here;
    forth!(DUP, ADD, Literal(1), ADD, comma, RET);

    // Helper function to compile a number ( n -- n? )
    let try_compile_lit = d.here;
    forth!(
        /* If we are in interpreting mode, */
        state, LD,

        /* then exit immediately, leaving this number on the stack. */
        Q, RET,

        /* Otherwise, turn it into a LITERAL instruction and append that
         * to the dictionary, */
        lit,

        /* and then return-from-caller. */
        RTO, DRP, RET);

Similarly, to compile a word, we'll want to convert from an execution token (xt) on the stack to a CALL instruction in the dictionary. Unless it's an immediate word, which we need to execute right now.

    // Helper function to compile a call ( xt -- xt? )
    let try_compile_call = d.here;
    forth!(
        /* If this is an immediate word, */
        DUP, Literal(4), sub, LD, Literal(0x0080), AND,

        /* or if we are in interpreting mode, */
        state, LD, OR,

        /* then we should execute this word, not compile it. */
        Q, RET,

        /* Otherwise, compile it by appending its address to the dictionary, */
        comma,

        /* and then return-from-caller. */
        RTO, DRP, RET);

    /* Given the address of a word, execute that word. */
    // execute ( xt -- )
    d.entry(); d.name(7, *b"exe");  let execute = d.here;
    forth!(TOR, RET);

    // Helper function to compile or execute a word ( xt -- )
    let do_word = d.here;
    forth!(
        /* When this function concludes, return-from-caller. */
        RTO, DRP,

        /* If this word should be compiled, compile it, */
        try_compile_call,

        /* otherwise, execute it. */
        execute, RET);

Forth can have very good error handling. This Forth does not. If we try to look up a word in the dictionary and can't find it, and if the word also can't be parsed as an number, then we print out a ? and move on to the next word.

This helper function does some stack cleanup, prints the ?, then uses the return-from-caller trick to move on to the next word.

    let bad = d.here;
    forth!(DRP, Literal(63), emit, RTO, DRP, RET);

Given all that, here's an all-in-one subroutine that figures out what to do with the contents of the input buffer.

    // dispatch ( xt -- )
    d.entry(); d.name(9, *b"int");  let dispatch = d.here;
    forth!(
        /* If the word was found in the dictionary, treat it as a word. */
        DUP, Q, do_word,

        /* If it wasn't found in the dictionary, try to parse it as a number.
         * If it isn't a number, flag it as an error. */
        DRP, number, DUP, Literal(1), ADD, zero_eq, Q, bad,

        /* If it is a number, treat it as a number. */
        try_compile_lit, RET);

And now we can write the main interpreter/compiler loop. This is the top-level code for our entire Forth system! Forth names this "quit", because you'd expect putting the word "quit" in the middle of a compiled program to bring you back to top-level.

"quit" is called the "outer interpreter" because it is the outermost interpreter loop that Forth uses. Some Forth implementations also use an "inner interpreter" to execute their threaded code. Our Forth does not have an inner interpreter because we used subroutine threading, making our threaded code a list of subroutine calls that can be directly executed by the CPU.

Let's look at what "quit" does. We've already done all the hard work so it can be quite short.

    // quit ( -- )
    d.entry(); d.name(4, *b"qui");  let quit = d.here;
    forth!(
        quote,    /* Read a word from the keyboard and look it up in
                   * the dictionary */

        dispatch, /* Figure out what to do with the word */

        quit      /* Repeat forever */
    );

You might have noticed that "quit" isn't tail-recursive -- it just calls itself normally. "quit" is never supposed to return so it doesn't matter for it to properly maintain the return stack. It will just fill up the circular stack and wrap around. That's fine.

We now have an interpreter that can compile or execute code!!!

We have now succeeded at:

    1. Call subroutines by typing their name at the keyboard
    1. Push numbers onto the data stack by typing them at the keyboard

2.4 - Defining subroutines

There are still a few more words we'll need if we want to:

    1. Define new subroutines in terms of existing ones

Let's take care of that now.

create

Here is a word to create a new dictionary header.

    // create ( -- )
    d.entry(); d.name(6, *b"cre");  let create = d.here;
    forth!(
        here, LD,
        latest, LD, comma,  /* emit the link field */
        latest, ST,         /* point "latest" at us */
        word,               /* read a word from the keyboard */

        /* emit the name field (by copying it from the input buffer) */
        Literal(word_buf), DUP, LD, comma, Literal(2), ADD, LD, comma,

        RET);

: (define word)

Here is the word to compile a new Forth word.

    // : ( -- )
    d.entry(); d.name(1, *b":  ");
    forth!(
        /* Read name from keyboard, create dictionary header */
        create,

        /* Hide the word until we are done defining it.  This lets us
         * redefine a word in terms of a previous incarnation of itself. */
        smudge,

        /* Switch to compiling mode */
        rbracket,
        RET);

; (end of definition)

Finally, here is semicolon, the "end" marker that ends the Forth word. Note that ; is immediate, as it has to switch us from compiling mode back into interpreting mode.

    // ; ( -- )
    d.entry(); d.name(1 | 0x80, *b";  ");
    forth!(
        /* Emit a RET instruction.  RET = 65504 which is outside of the
         * LITERAL instruction's 0 to 32767 range, so you have to store the
         * inverse and use INV to swap it back. */
        Literal(!(RET as u16)), INV, comma,

        /* The word is now done, so unhide it. */
        unsmudge,

        /* Switch back to interpreting mode */
        lbracket,

        RET);

Miscellanea

Wrap up the CPU instructions into dictionary words so we can call them interactively from Forth. Instructions that modify the return stack need special care, because otherwise they will mess up the wrapper we created for them, instead of acting on the caller the way they are supposed to.

    d.entry(); d.name(3, *b"ret"); forth!(RTO, DRP, RET);
    d.entry(); d.name(2, *b">r "); forth!(RTO, SWP, TOR, TOR, RET);
    d.entry(); d.name(2, *b"r> "); forth!(RTO, RTO, SWP, TOR, RET);
    d.entry(); d.name(1, *b"@  "); forth!(LD, RET);
    d.entry(); d.name(1, *b"!  "); forth!(ST, RET);
    d.entry(); d.name(3, *b"dup"); forth!(DUP, RET);
    d.entry(); d.name(4, *b"swa"); forth!(SWP, RET);
    d.entry(); d.name(4, *b"dro"); forth!(DRP, RET);

    d.entry(); d.name(1 | 0x80, *b"?  "); /* This one only works in-line. */
    forth!(Literal(!(Q as u16)), INV, comma, RET);

    d.entry(); d.name(1, *b"+  "); forth!(ADD, RET);
    d.entry(); d.name(5, *b"shi"); forth!(SFT, RET);
    d.entry(); d.name(2, *b"or "); forth!(OR, RET);
    d.entry(); d.name(3, *b"and"); forth!(AND, RET);
    d.entry(); d.name(3, *b"inv"); forth!(INV, RET);
    d.entry(); d.name(3, *b"u>="); forth!(GEQ, RET);
    d.entry(); d.name(2, *b"io "); forth!(IO, RET);

Update Forth's "latest" and "here" variables to match the ones we've been tracking in Rust.

    d.c.store(latest_ptr, d.dp);
    d.c.store(here_ptr, d.here);

Start out in interpreting mode.

    d.c.store(state_ptr, 0xffff);

Put a call to the outer interpreter at the CPU's reset vector.

    d.c.store(0, quit);
}

Finally, start the machine.

fn main() {
    /* Create the machine */
    let mut c = Core::new();

    /* Put the dictionary into memory */
    build_dictionary(&mut c);

    /* Start running the CPU from the reset vector */
    c.ip = 0;
    loop {
        c.step();
    }
}

Part 3 - Using the interactive programming environment

"The next step is a problem-oriented-language. By permitting the program to dynamically modify its control language, we mark a qualitative change in capability. We also change our attention from the program to the language it implements. This is an important, and dangerous, diversion. For it's easy to lose sight of the problem amidst the beauty of the solution."

-- Chuck Moore, "Programming a Problem-Oriented Language", 1970

Now we can start programming in "real" Forth, not a weird macro language inside Rust.

You can compile our Forth computer with:

    rustc frustration.rs

You can run our Forth computer with:

    ./frustration

However, I recommend loading a Forth program (frustration.4th, provided) which does a few more setup steps before letting you loose.

    cat frustration.4th - | ./frustration

The line above is a good way to run Frustration if you're using Linux. It concatenates together frustration.4th and - (stdin). This means you can type commands once frustration.4th has been executed.

There is a shell script supplied that will do all of the above for you.

    bash build.sh

Please read frustration.4th if you want to learn more about how to use Forth.