Sunday, October 08, 2006

Bit Twiddling Hacks

Bit Twiddling Hacks
By Sean Eron Anderson
seander@cs.stanford.edu
Individually, the code snippets here are in the public domain (unless otherwise noted) — feel free to use them however you please. The aggregate collection and descriptions are © 1997-2005 Sean Eron Anderson. The code and descriptions are distributed in the hope that they will be useful, but WITHOUT ANY WARRANTY and without even the implied warranty of merchantability or fitness for a particular purpose. As of May 5, 2005, all the code has been tested thoroughly. Thousands of people have read it. Moreover, Professor Randal Bryant, the Dean of Computer Science at Carnegie Mellon University, has personally tested almost everything with his Uclid code verification system. What he hasn't tested, I have checked against all possible inputs. To the first person to inform me of a legitimate bug in the code, I'll pay a bounty of US$10 (by check or Paypal).


Contents
About the operation counting methodology
Compute the sign of an integer
Compute the integer absolute value (abs) without branching
Compute the minimum (min) or maximum (max) of two integers without branching
Determining if an integer is a power of 2
Sign extending
Sign extending from a constant bit width
Sign extending from a variable bit-width
Sign extending from a variable bit-width in 3 operations
Conditionally set or clear bits without branching
Merge bits from two values according to a mask
Counting bits set
Counting bits set, naive way
Counting bits set by lookup table
Counting bits set, Brian Kernighan's way
Counting bits set in 12, 24, or 32-bit words using 64-bit instructions
Counting bits set, in parallel
Computing parity (1 if an odd number of bits set, 0 otherwise)
Compute parity of a word the naive way
Compute parity by lookup table
Compute parity of a byte using 64-bit multiply and modulus division
Compute parity in parallel
Swapping Values
Swapping values with XOR
Swapping individual bits with XOR
Reversing bit sequences
Reverse bits the obvious way
Reverse bits in word by lookup table
Reverse the bits in a byte with 3 operations (64-bit muliply and modulus division)
Reverse the bits in a byte with 4 operations (64-bit multiply, no division)
Reverse the bits in a byte with 7 operations (no 64-bit, only 32)
Reverse an N-bit quantity in parallel with 5 * lg(N) operations
Modulus division (aka computing remainders)
Computing modulus division by 1 << s without a division operation (obvious)
Computing modulus division by (1 << s) - 1 without a division operation
Computing modulus division by (1 << s) - 1 in parallel without a division operation
Finding integer log base 2 of an integer (aka the position of the highest bit set)
Find the log base 2 of an integer with the MSB N set in O(N) operations (the obvious way)
Find the integer log base 2 of an integer with an 64-bit IEEE float
Find the log base 2 of an integer with a lookup table
Find the log base 2 of an N-bit integer in O(lg(N)) operations
Find the log base 2 of an N-bit integer in O(lg(N)) operations with multiply and lookup
Find integer log base 10 of an integer
Find integer log base 2 of a 32-bit IEEE float
Find integer log base 2 of the pow(2, r)-root of a 32-bit IEEE float (for unsigned integer r)
Counting consecutive trailing zero bits (or finding bit indices)
Count the consecutive zero bits (trailing) on the right in parallel
Count the consecutive zero bits (trailing) on the right by binary search
Count the consecutive zero bits (trailing) on the right by casting to a float
Count the consecutive zero bits (trailing) on the right with modulus division and lookup
Count the consecutive zero bits (trailing) on the right with multiply and lookup
Round up to the next highest power of 2 by float casting
Round up to the next highest power of 2
Interleaving bits (aka computing Morton Numbers)
Interleave bits the obvious way
Interleave bits by table lookup
Interleave bits with 64-bit multiply
Interleave bits by Binary Magic Numbers
Testing for ranges of bytes in a word (and counting occurances found)
Determine if a word has a zero byte
Determine if a word has byte less than n
Determine if a word has a byte greater than n
Determine if a word has a byte between m and n

--------------------------------------------------------------------------------

About the operation counting methodology
When totaling the number of operations for algorithms here, any C operator is counted as one operation. Intermediate assignments, which need not be written to RAM, are not counted. Of course, this operation counting approach only serves as an approximation of the actual number of machine instructions and CPU time. All operations are assumed to take the same amount of time, which is not true in reality, but CPUs have been heading increasingly in this direction over time. There are many nuances that determine how fast a system will run a given sample of code, such as cache sizes, memory bandwidths, instruction sets, etc. In the end, benchmarking is the best way to determine whether one method is really faster than another, so consider the techniques below as possibilities to test on your target architecture.


--------------------------------------------------------------------------------

Compute the sign of an integer
int v; // we want to find the sign of v
int sign; // the result goes here

// CHAR_BIT is the number of bits per byte (normally 8).
sign = -(v < 0); // if v < 0 then -1, else 0.
// or, to avoid branching on CPUs with flag registers (IA32):
sign = -(int)((unsigned int)((int)v) >> (sizeof(int) * CHAR_BIT - 1));
// or, for one less instruction (but not portable):
sign = v >> (sizeof(int) * CHAR_BIT - 1);

The last expression above evaluates to sign = v >> 31 for 32-bit integers. This is one operation faster than the obvious way, sign = -(v < 0). This trick works because when signed integers are shifted right, the value of the far left bit is copied to the other bits. The far left bit is 1 when the value is negative and 0 otherwise; all 1 bits gives -1. Unfortunately, this behavior is architecture-specific.
Alternatively, if you prefer the result be either -1 or +1, then use:

sign = +1 | (v >> (sizeof(int) * CHAR_BIT - 1)); // if v < 0 then -1, else +1

Alternatively, if you prefer the result be either -1, 0, or +1, then use:

sign = (v != 0) | -(int)((unsigned int)((int)v) >> (sizeof(int) * CHAR_BIT - 1));
// Or, for more speed but less portability:
sign = (v != 0) | (v >> (sizeof(int) * CHAR_BIT - 1)); // -1, 0, or +1
// Or, for portability, brevity, and (perhaps) speed:
sign = (v > 0) - (v < 0); // -1, 0, or +1

Caveat: On March 7, 2003, Angus Duggan pointed out that the 1989 ANSI C specification leaves the result of signed right-shift implementation-defined, so on some systems this hack might not work. For greater portability, Toby Speight suggested on September 28, 2005 that CHAR_BIT be used here and throughout rather than assuming bytes were 8 bits long. Angus recommended the more portable versions above, involving casting on March 4, 2006.


--------------------------------------------------------------------------------

Compute the integer absolute value (abs) without branching
int v; // we want to find the absolute value of v
int r; // the result goes here

r = (v ^ (v >> (sizeof(int) * CHAR_BIT - 1))) -
(v >> (sizeof(int) * CHAR_BIT - 1));

Some CPUs don't have an integer absolute value instruction (or the compiler fails to use them). On machines where branching is expensive, the above expression can be faster than the obvious approach, r = (v < 0) ? -v : v, even though the number of operations is the same.
Caveats: On March 7, 2003, Angus Duggan pointed out that the 1989 ANSI C specification leaves the result of signed right-shift implementation-defined, so on some systems this hack might not work. I've read that ANSI C does not require values to be represented as two's complement, so it may not work for that reason as well (on a diminishingly small number of old machines that still use one's complement). On March 14, 2004, Keith H. Duggar sent me the solution above; it is superior to the one I initially came up with, r=(+1|(v>>(sizeof(int)*CHAR_BIT-1)))*v, because a multiply is not used. Unfortunately, this method has been patented in the USA on June 6, 2000 by Vladimir Yu Volkonsky and assigned to Sun Microsystems. (For an unpatented non-multiplying alternative, consider (v ^ (v >> (sizeof(int) * CHAR_BIT - 1))) + ((unsigned) v >> (sizeof(int) * CHAR_BIT - 1)), suggested by Thorbjørn Willoch on June 21, 2004 or my variation, (v ^ (v >> (sizeof(int) * CHAR_BIT - 1))) + (v < 0), both of which take one more operation than the patented one, though it may be computed in parallel on modern CPUs.) On August 13, 2006, Yuriy Kaminskiy told me that the patent is likely invalid because the method was published well before the patent was even filed, such as in How to Optimize for the Pentium Processor by Agner Fog, dated November, 9, 1996. Yuriy also mentioned that this document was translated to Russian in 1997, which Vladimir could have read. Moreover, the Internet Archive also has an old link to it.



--------------------------------------------------------------------------------

Compute the minimum (min) or maximum (max) of two integers without branching
int x; // we want to find the minimum of x and y
int y;
int r; // the result goes here

r = y + ((x - y) & -(x < y)); // min(x, y)

On machines where branching is expensive, the above expression can be faster than the obvious approach, r = (x < y) ? x : y, even though it involves two more instructions. It works because if x < y, then -(x < y) will be all ones, so r = y + (x - y) & ~0 = y + x - y = x. Otherwise, if x >= y, then -(x < y) will be all zeros, so r = y + (x - y) & 0 = y. On some machines, evaluating (x < y) as 0 or 1 requires a branch instruction, so there may be no advantage.
To find the maximum, use:

r = x - ((x - y) & -(x < y)); // max(x, y)

Quick and dirty versions:
If you know that INT_MIN <= x - y <= INT_MAX, then you can use the following, which are faster because (x - y) only needs to be evaluated once. (Note that the 1989 ANSI C specification doesn't specify the result of signed right-shift, so these aren't portable.)
r = y + ((x - y) & ((x - y) >> (sizeof(int) * CHAR_BIT - 1))); // min(x, y)
r = x - ((x - y) & ((x - y) >> (sizeof(int) * CHAR_BIT - 1))); // max(x, y)

On March 7, 2003, Angus Duggan pointed out the right-shift portability issue. On May 3, 2005, Randal E. Bryant alerted me to the need for the precondition, INT_MIN <= x - y <= INT_MAX, and suggested the non-quick and dirty version as a fix. Both of these issues concern only the quick and dirty version. Nigel Horspoon pointed out on July 6, 2005 that gcc produced the same code on a Pentium as the obvious solution because of how it evaluates (x < y).


--------------------------------------------------------------------------------

Determining if an integer is a power of 2
unsigned int v; // we want to see if v is a power of 2
bool f; // the result goes here

f = (v & (v - 1)) == 0;

Note that 0 is incorrectly considered a power of 2 here. To remedy this, use:
f = !(v & (v - 1)) && v;


--------------------------------------------------------------------------------

Sign extending from a constant bit width
Sign extension is automatic for built-in types, such as chars and ints. But suppose you have a signed two's complement number, x, that is stored using only b bits. Moreover, suppose you want to convert x to an int, which has more than b bits. A simple copy will work if x is positive, but if negative, the sign must be extended. For example, if we have only 4 bits to store a number, then -3 is represented as 1101 in binary. If we have 8 bits, then -3 is 11111101. The most significant bit of the 4-bit representation is replicated sinistrally to fill in the destination when we convert to a representation with more bits; this is sign extending. In C, sign extension from a constant bit width is trivial, since bit fields may be specified in structs or unions. For example, to convert from 5 bits to an full integer:
int x; // convert this from using 5 bits to a full int
int r; // resulting sign extended number goes here
struct {signed int x:5;} s;
r = s.x = x;

The following is a C++ template function that uses the same language feature to convert from B bits in one operation (though the compiler is generating more, of course).
template
inline T signextend(const T x)
{
struct {T x:B;} s;
return s.x = x;
}

int r = signextend(x); // sign extend 5 bit number x to r

John Byrd caught a typo in the code (attributed to html formatting) on May 2, 2005. On March 4, 2006, Pat Wood pointed out that the ANSI C standard requires that the bitfield have the keyword "signed" to be signed; otherwise, the sign is undefined.


--------------------------------------------------------------------------------

Sign extending from a variable bit-width
Sometimes we need to extend the sign of a number but we don't know a priori the number of bits, b, in which it is represented. (Or we could be programming in Java, which lacks bitfields.)
unsigned b; // number of bits representing the number in x
int x; // sign extend this b-bit number to r
int r; // resulting sign-extended number
int const m = 1 << (b - 1); // mask can be pre-computed if b is fixed
r = -(x & m) | x;

The code above requires five operations. Sean A. Irvine suggested that I add sign extension methods to this page on June 13, 2004, and he provided m = (1 << (b - 1)) - 1; r = -(x & ~m) | x; as a starting point from which I optimized.


--------------------------------------------------------------------------------

Sign extending from a variable bit-width in 3 operations
The following may be slow on some machines, due to the effort required for multiplication and division. This version is 4 operations. If you know that your initial bit width, b, is greater than 1, you might do this type of sign extension in 3 operations by using r = (x * multipliers[b]) / multipliers[b], which requires only one array lookup.
unsigned b; // number of bits representing the number in x
int x; // sign extend this b-bit number to r
int r; // resulting sign-extended number
#define M(B) (1 << ((sizeof(x) * CHAR_BIT) - B)) // CHAR_BIT=bits/byte
int const multipliers[] =
{
0, M(1), M(2), M(3), M(4), M(5), M(6), M(7),
M(8), M(9), M(10), M(11), M(12), M(13), M(14), M(15),
M(16), M(17), M(18), M(19), M(20), M(21), M(22), M(23),
M(24), M(25), M(26), M(27), M(28), M(29), M(30), M(31),
M(32)
}; // (add more if using more than 64 bits)
int const divisors[] =
{
1, ~M(1), M(2), M(3), M(4), M(5), M(6), M(7),
M(8), M(9), M(10), M(11), M(12), M(13), M(14), M(15),
M(16), M(17), M(18), M(19), M(20), M(21), M(22), M(23),
M(24), M(25), M(26), M(27), M(28), M(29), M(30), M(31),
M(32)
}; // (add more for 64 bits)
#undef M
r = (x * multipliers[b]) / divisors[b];

The following variation is not portable, but on architectures that employ an arithmetic right-shift, maintaining the sign, it should be fast.
const int s = -b; // OR: sizeof(x) * CHAR_BIT - b;
r = (x << s) >> s;

Randal E. Bryant pointed out a bug on May 3, 2005 in an earlier version (that used multipliers[] for divisors[]), where it failed on the case of x=1 and b=1.


--------------------------------------------------------------------------------

Conditionally set or clear bits without branching
bool f; // conditional flag
unsigned int m; // the bit mask
unsigned int w; // the word to modify: if (f) w |= m; else w &= ~m;

w ^= (-f ^ w) & m;

On some architectures, the lack of branching can more than make up for what appears to be twice as many operations. For instance, informal speed tests on an AMD Athlon™ XP 2100+ indicated it was 5-10% faster. Glenn Slayden informed me of this expression on December 11, 2003.


--------------------------------------------------------------------------------

Merge bits from two values according to a mask
unsigned int a; // value to merge in non-masked bits
unsigned int b; // value to merge in masked bits
unsigned int mask; // 1 where bits from b should be selected; 0 where from a.
unsigned int r; // result of (a & ~mask) | (b & mask) goes here

r = a ^ ((a ^ b) & mask);

This shaves one operation from the obvious way of combining two sets of bits according to a bit mask. If the mask is a constant, then there may be no advantage.
Ron Jeffery sent this to me on February 9, 2006.



--------------------------------------------------------------------------------

Counting bits set (naive way)
unsigned int v; // count the number of bits set in v
unsigned int c; // c accumulates the total bits set in v

for (c = 0; v; v >>= 1)
{
c += v & 1;
}

The naive approach requires one iteration per bit, until no more bits are set. So on a 32-bit word with only the high set, it will go through 32 iterations.


--------------------------------------------------------------------------------

Counting bits set by lookup table
const unsigned char BitsSetTable256[] =
{
0, 1, 1, 2, 1, 2, 2, 3, 1, 2, 2, 3, 2, 3, 3, 4,
1, 2, 2, 3, 2, 3, 3, 4, 2, 3, 3, 4, 3, 4, 4, 5,
1, 2, 2, 3, 2, 3, 3, 4, 2, 3, 3, 4, 3, 4, 4, 5,
2, 3, 3, 4, 3, 4, 4, 5, 3, 4, 4, 5, 4, 5, 5, 6,
1, 2, 2, 3, 2, 3, 3, 4, 2, 3, 3, 4, 3, 4, 4, 5,
2, 3, 3, 4, 3, 4, 4, 5, 3, 4, 4, 5, 4, 5, 5, 6,
2, 3, 3, 4, 3, 4, 4, 5, 3, 4, 4, 5, 4, 5, 5, 6,
3, 4, 4, 5, 4, 5, 5, 6, 4, 5, 5, 6, 5, 6, 6, 7,
1, 2, 2, 3, 2, 3, 3, 4, 2, 3, 3, 4, 3, 4, 4, 5,
2, 3, 3, 4, 3, 4, 4, 5, 3, 4, 4, 5, 4, 5, 5, 6,
2, 3, 3, 4, 3, 4, 4, 5, 3, 4, 4, 5, 4, 5, 5, 6,
3, 4, 4, 5, 4, 5, 5, 6, 4, 5, 5, 6, 5, 6, 6, 7,
2, 3, 3, 4, 3, 4, 4, 5, 3, 4, 4, 5, 4, 5, 5, 6,
3, 4, 4, 5, 4, 5, 5, 6, 4, 5, 5, 6, 5, 6, 6, 7,
3, 4, 4, 5, 4, 5, 5, 6, 4, 5, 5, 6, 5, 6, 6, 7,
4, 5, 5, 6, 5, 6, 6, 7, 5, 6, 6, 7, 6, 7, 7, 8
};

unsigned int v; // count the number of bits set in 32-bit value v
unsigned int c; // c is the total bits set in v

// Option 1:
c = BitsSetTable256[v & 0xff] +
BitsSetTable256[(v >> 8) & 0xff] +
BitsSetTable256[(v >> 16) & 0xff] +
BitsSetTable256[v >> 24];

// Option 2:
unsigned char * p = (unsigned char *) &v;
c = BitsSetTable256[p[0]] +
BitsSetTable256[p[1]] +
BitsSetTable256[p[2]] +
BitsSetTable256[p[3]];


// To initially generate the table algorithmically:
BitsSetTable256[0] = 0;
for (int i = 0; i < 256; i++)
{
BitsSetTable256[i] = (i & 1) + BitsSetTable256[i / 2];
}



--------------------------------------------------------------------------------

Counting bits set, Brian Kernighan's way
unsigned int v; // count the number of bits set in v
unsigned int c; // c accumulates the total bits set in v
for (c = 0; v; c++)
{
v &= v - 1; // clear the least significant bit set
}

Brian Kernighan's method goes through as many iterations as there are set bits. So if we have a 32-bit word with only the high bit set, then it will only go once through the loop.
Published in 1988, the C Programming Language 2nd Ed. (by Brian W. Kernighan and Dennis M. Ritchie) mentions this in exercise 2-9. On April 19, 2006 Don Knuth pointed out to me that this method "was first published by Peter Wegner in CACM 3 (1960), 322. (Also discovered independently by Derrick Lehmer and published in 1964 in a book edited by Beckenbach.)"



--------------------------------------------------------------------------------

Counting bits set in 12, 24, or 32-bit words using 64-bit instructions
unsigned int v; // count the number of bits set in v
unsigned int c; // c accumulates the total bits set in v

// option 1, for at most 12-bit values in v:
c = (v * 0x1001001001001ULL & 0x84210842108421ULL) % 0x1f;

// option 2, for at most 24-bit values in v:
c = ((v & 0xfff) * 0x1001001001001ULL & 0x84210842108421ULL) % 0x1f;
c += (((v & 0xfff000) >> 12) * 0x1001001001001ULL & 0x84210842108421ULL)
% 0x1f;

// option 3, for at most 32-bit values in v:
c = ((v & 0xfff) * 0x1001001001001ULL & 0x84210842108421ULL) % 0x1f;
c += (((v & 0xfff000) >> 12) * 0x1001001001001ULL & 0x84210842108421ULL) %
0x1f;
c += ((v >> 24) * 0x1001001001001ULL & 0x84210842108421ULL) % 0x1f;

This method requires a 64-bit CPU with fast modulus division to be efficient. The first option takes only 3 operations; the second option takes 10; and the third option takes 15.
Rich Schroeppel originally created a 9-bit version, similiar to option 1; see the Programming Hacks section of Beeler, M., Gosper, R. W., and Schroeppel, R. HAKMEM. MIT AI Memo 239, Feb. 29, 1972. His method was the inspiration for the variants above, devised by Sean Anderson. Randal E. Bryant offered a couple bug fixes on May 3, 2005.




--------------------------------------------------------------------------------

Counting bits set, in parallel
unsigned int v; // count bits set in this (32-bit value)
unsigned int c; // store the total here
const int S[] = {1, 2, 4, 8, 16}; // Magic Binary Numbers
const int B[] = {0x55555555, 0x33333333, 0x0F0F0F0F, 0x00FF00FF, 0x0000FFFF};

c = v;
c = v - ((v >> 1) & B[0]);
c = ((c >> S[1]) & B[1]) + (c & B[1]);
c = ((c >> S[2]) + c) & B[2];
c = ((c >> S[3]) + c) & B[3];
c = ((c >> S[4]) + c) & B[4];

The B array, expressed as binary, is:
B[0] = 0x55555555 = 01010101 01010101 01010101 01010101
B[1] = 0x33333333 = 00110011 00110011 00110011 00110011
B[2] = 0x0F0F0F0F = 00001111 00001111 00001111 00001111
B[3] = 0x00FF00FF = 00000000 11111111 00000000 11111111
B[4] = 0x0000FFFF = 00000000 00000000 11111111 11111111

We can adjust the method for larger integer sizes by continuing with the patterns for the Binary Magic Numbers, B and S. If there are k bits, then we need the arrays S and B to be ceil(lg(k)) elements long, and we must compute the same number of expressions for c as S or B are long. For a 32-bit v, 16 operations are used.
The best method for counting bits in a 32-bit integer v is the following:

unsigned int const w = v - ((v >> 1) & 0x55555555); // temp
unsigned int const x = (w & 0x33333333) + ((w >> 2) & 0x33333333); // temp
unsigned int const c = ((x + (x >> 4) & 0xF0F0F0F) * 0x1010101) >> 24; // count

The best bit counting method takes only 12 operations, which is the same as the lookup-table method, but avoids the memory and potential cache misses of a table. It is a hybrid between the purely parallel method above and the earlier methods using multiplies (in the section on counting bits with 64-bit instructions), though it doesn't use 64-bit instructions. The counts of bits set in the bytes is done in parallel, and the sum total of the bits set in the bytes is computed by multiplying by 0x1010101 and shifting right 24 bits.

See Ian Ashdown's nice newsgroup post for more information on counting the number of bits set (also known as sideways addition). The best bit counting method was brought to my attention on October 5, 2005 by Andrew Shapira; he found it in pages 187-188 of Software Optimization Guide for AMD Athlon™ 64 and Opteron™ Processors. Charlie Gordon suggested a way to shave off one operation from the purely parallel version on December 14, 2005, and Don Clugston trimmed three more from it on December 30, 2005. He also pointed out that if rotate operations were available, changing the shifts to rotates would eliminate the need for the AND masks because the overcounted sum total could instead be right-shifted 3 at the end. I made a typo with Don's suggestion that Eric Cole spotted on January 8, 2006.



--------------------------------------------------------------------------------

Computing parity the naive way
unsigned int v; // word value to compute the parity of
bool parity = false; // parity will be the parity of b

while (v)
{
parity = !parity;
v = v & (v - 1);
}


The above code uses an approach like Brian Kernigan's bit counting, above. The time it takes is proportional to the number of bits set.


--------------------------------------------------------------------------------

Compute parity of a byte using 64-bit multiply and modulus division
unsigned char b; // byte value to compute the parity of
bool parity =
(((b * 0x0101010101010101ULL) & 0x8040201008040201ULL) % 0x1FF) & 1;

The method above takes around 4 operations, but only works on bytes.


--------------------------------------------------------------------------------

Compute parity in parallel
unsigned int v; // word value to compute the parity of
v ^= v >> 16;
v ^= v >> 8;
v ^= v >> 4;
v &= 0xf;
return (0x6996 >> v) & 1;

The method above takes around 9 operations, and works for 32-bit words. It may be optimized to work just on bytes in 5 operations by removing the two lines immediately following "unsigned int v;". The method first shifts and XORs the eight nibbles of the 32-bit value together, leaving the result in the lowest nibble of v. Next, the binary number 0110 1001 1001 0110 (0x6996 in hex) is shifted to the right by the value represented in the lowest nibble of v. This number is like a miniature 16-bit parity-table indexed by the low four bits in v. The result has the parity of v in bit 1, which is masked and returned.
Thanks to Mathew Hendry for pointing out the shift-lookup idea at the end on Dec. 15, 2002. That optimization shaves two operations off using only shifting and XORing to find the parity.



--------------------------------------------------------------------------------

Compute parity by lookup table
const bool ParityTable[] =
{
0, 1, 1, 0, 1, 0, 0, 1, 1, 0, 0, 1, 0, 1, 1, 0,
1, 0, 0, 1, 0, 1, 1, 0, 0, 1, 1, 0, 1, 0, 0, 1,
1, 0, 0, 1, 0, 1, 1, 0, 0, 1, 1, 0, 1, 0, 0, 1,
0, 1, 1, 0, 1, 0, 0, 1, 1, 0, 0, 1, 0, 1, 1, 0,
1, 0, 0, 1, 0, 1, 1, 0, 0, 1, 1, 0, 1, 0, 0, 1,
0, 1, 1, 0, 1, 0, 0, 1, 1, 0, 0, 1, 0, 1, 1, 0,
0, 1, 1, 0, 1, 0, 0, 1, 1, 0, 0, 1, 0, 1, 1, 0,
1, 0, 0, 1, 0, 1, 1, 0, 0, 1, 1, 0, 1, 0, 0, 1,
1, 0, 0, 1, 0, 1, 1, 0, 0, 1, 1, 0, 1, 0, 0, 1,
0, 1, 1, 0, 1, 0, 0, 1, 1, 0, 0, 1, 0, 1, 1, 0,
0, 1, 1, 0, 1, 0, 0, 1, 1, 0, 0, 1, 0, 1, 1, 0,
1, 0, 0, 1, 0, 1, 1, 0, 0, 1, 1, 0, 1, 0, 0, 1,
0, 1, 1, 0, 1, 0, 0, 1, 1, 0, 0, 1, 0, 1, 1, 0,
1, 0, 0, 1, 0, 1, 1, 0, 0, 1, 1, 0, 1, 0, 0, 1,
1, 0, 0, 1, 0, 1, 1, 0, 0, 1, 1, 0, 1, 0, 0, 1,
0, 1, 1, 0, 1, 0, 0, 1, 1, 0, 0, 1, 0, 1, 1, 0
};

unsigned char b; // byte value to compute the parity of
bool parity = ParityTable[b];

// OR, for 32-bit words:

unsigned int v; // word value to compute the parity of
bool parity = ParityTable[(v & 0x000000ff)] ^
ParityTable[(v & 0x0000ff00) >> 8] ^
ParityTable[(v & 0x00ff0000) >> 16] ^
ParityTable[(v & 0xff000000) >> 24];

// Variation:
unsigned char * p = (unsigned char *) &v;
bool parity = ParityTable[p[0]] ^
ParityTable[p[1]] ^
ParityTable[p[2]] ^
ParityTable[p[3]];

Randal E. Byrant encouraged the addition of the (admittedly) obvious last variation with variable p on May 3, 2005. Bruce Rawles found a typo in an instance of the table variable's name on September 27, 2005, and he received a $10 bug bounty.


--------------------------------------------------------------------------------

Swapping values with XOR
#define SWAP(a, b) (((a) ^= (b)), ((b) ^= (a)), ((a) ^= (b)))

This is an old trick to exchange the values of the variables a and b without using extra space for a temporary variable.
On January 20, 2005, Iain A. Fleming pointed out that the macro above doesn't work when you swap with the same memory location, such as SWAP(a[i], a[j]) with i == j. So if that may occur, consider defining the macro as (((a) == (b)) || (((a) ^= (b)), ((b) ^= (a)), ((a) ^= (b)))).



--------------------------------------------------------------------------------

Swapping individual bits with XOR
unsigned int i, j; // positons of bit sequences to swap
unsigned int n; // number of consecutive bits in each sequence
unsigned int b; // bits to swap reside in b
unsigned int r; // bit-swapped result goes here

int x = ((b >> i) ^ (b >> j)) & ((1 << n) - 1); // XOR temporary
r = b ^ ((x << i) | (x << j));

As an example of swapping ranges of bits suppose we have have b = 00101111 (expressed in binary) and we want to swap the n = 3 consecutive bits starting at i = 1 (the second bit from the right) with the 3 consecutive bits starting at j = 5; the result would be r = 11100011 (binary).
This method of swapping is similar to the general purpose XOR swap trick, but intended for operating on individual bits. The variable x stores the result of XORing the pairs of bit values we want to swap, and then the bits are set to the result of themselves XORed with x. Of course, the result is undefined if the sequences overlap.



--------------------------------------------------------------------------------

Reverse bits the obvious way
unsigned int v; // reverse the bits in this
unsigned int t = v << 1; // t will have the reversed bits of v
int i;

v >>= 1;
for (i = sizeof(v) * CHAR_BIT - 2; i; i--)
{
t |= v & 1;
t <<= 1;
v >>= 1;
}
t |= v;

On October 15, 2004, Michael Hoisie pointed out a bug in the original version. Randal E. Byrant suggested removing an extra operation on May 3, 2005. Behdad Esfabod suggested a slight change that eliminated one iteration of the loop on May 18, 2005.



--------------------------------------------------------------------------------

Reverse bits in word by lookup table
const unsigned char BitReverseTable256[] =
{
0x00, 0x80, 0x40, 0xC0, 0x20, 0xA0, 0x60, 0xE0, 0x10, 0x90, 0x50, 0xD0, 0x30, 0xB0, 0x70, 0xF0,
0x08, 0x88, 0x48, 0xC8, 0x28, 0xA8, 0x68, 0xE8, 0x18, 0x98, 0x58, 0xD8, 0x38, 0xB8, 0x78, 0xF8,
0x04, 0x84, 0x44, 0xC4, 0x24, 0xA4, 0x64, 0xE4, 0x14, 0x94, 0x54, 0xD4, 0x34, 0xB4, 0x74, 0xF4,
0x0C, 0x8C, 0x4C, 0xCC, 0x2C, 0xAC, 0x6C, 0xEC, 0x1C, 0x9C, 0x5C, 0xDC, 0x3C, 0xBC, 0x7C, 0xFC,
0x02, 0x82, 0x42, 0xC2, 0x22, 0xA2, 0x62, 0xE2, 0x12, 0x92, 0x52, 0xD2, 0x32, 0xB2, 0x72, 0xF2,
0x0A, 0x8A, 0x4A, 0xCA, 0x2A, 0xAA, 0x6A, 0xEA, 0x1A, 0x9A, 0x5A, 0xDA, 0x3A, 0xBA, 0x7A, 0xFA,
0x06, 0x86, 0x46, 0xC6, 0x26, 0xA6, 0x66, 0xE6, 0x16, 0x96, 0x56, 0xD6, 0x36, 0xB6, 0x76, 0xF6,
0x0E, 0x8E, 0x4E, 0xCE, 0x2E, 0xAE, 0x6E, 0xEE, 0x1E, 0x9E, 0x5E, 0xDE, 0x3E, 0xBE, 0x7E, 0xFE,
0x01, 0x81, 0x41, 0xC1, 0x21, 0xA1, 0x61, 0xE1, 0x11, 0x91, 0x51, 0xD1, 0x31, 0xB1, 0x71, 0xF1,
0x09, 0x89, 0x49, 0xC9, 0x29, 0xA9, 0x69, 0xE9, 0x19, 0x99, 0x59, 0xD9, 0x39, 0xB9, 0x79, 0xF9,
0x05, 0x85, 0x45, 0xC5, 0x25, 0xA5, 0x65, 0xE5, 0x15, 0x95, 0x55, 0xD5, 0x35, 0xB5, 0x75, 0xF5,
0x0D, 0x8D, 0x4D, 0xCD, 0x2D, 0xAD, 0x6D, 0xED, 0x1D, 0x9D, 0x5D, 0xDD, 0x3D, 0xBD, 0x7D, 0xFD,
0x03, 0x83, 0x43, 0xC3, 0x23, 0xA3, 0x63, 0xE3, 0x13, 0x93, 0x53, 0xD3, 0x33, 0xB3, 0x73, 0xF3,
0x0B, 0x8B, 0x4B, 0xCB, 0x2B, 0xAB, 0x6B, 0xEB, 0x1B, 0x9B, 0x5B, 0xDB, 0x3B, 0xBB, 0x7B, 0xFB,
0x07, 0x87, 0x47, 0xC7, 0x27, 0xA7, 0x67, 0xE7, 0x17, 0x97, 0x57, 0xD7, 0x37, 0xB7, 0x77, 0xF7,
0x0F, 0x8F, 0x4F, 0xCF, 0x2F, 0xAF, 0x6F, 0xEF, 0x1F, 0x9F, 0x5F, 0xDF, 0x3F, 0xBF, 0x7F, 0xFF
};

unsigned int v; // reverse 32-bit value, 8 bits at time
unsigned int c; // c will get v reversed

// Option 1:
c = (BitReverseTable256[v & 0xff] << 24) |
(BitReverseTable256[(v >> 8) & 0xff] << 16) |
(BitReverseTable256[(v >> 16) & 0xff] << 8) |
(BitReverseTable256[(v >> 24) & 0xff]);

// Option 2:
unsigned char * p = (unsigned char *) &v;
unsigned char * q = (unsigned char *) &c;
q[3] = BitReverseTable256[p[0]];
q[2] = BitReverseTable256[p[1]];
q[1] = BitReverseTable256[p[2]];
q[0] = BitReverseTable256[p[3]];

The first method takes about 17 operations, and the second takes about 12, assuming your CPU can load and store bytes easily.


--------------------------------------------------------------------------------

Reverse the bits in a byte with 3 operations (64-bit multiply and modulus division):
unsigned char b; // reverse this (8-bit) byte

b = (b * 0x0202020202ULL & 0x010884422010ULL) % 1023;

The multiply operation creates five separate copies of the 8-bit byte pattern to fan-out into a 64-bit value. The AND operation selects the bits that are in the correct (reversed) positions, relative to each 10-bit groups of bits. The multiply and the AND operations copy the bits from the original byte so they each appear in only one of the 10-bit sets. The reversed positions of the bits from the original byte coincide with their relative positions within any 10-bit set. The last step, which involves modulus division by 2^10 - 1, has the effect of merging together each set of 10 bits (from positions 0-9, 10-19, 20-29, ...) in the 64-bit value. They do not overlap, so the addition steps underlying the modulus division behave like or operations.
This method was attributed to Rich Schroeppel in the Programming Hacks section of Beeler, M., Gosper, R. W., and Schroeppel, R. HAKMEM. MIT AI Memo 239, Feb. 29, 1972.



--------------------------------------------------------------------------------

Reverse the bits in a byte with 4 operations (64-bit multiply, no division):
unsigned char b; // reverse this byte

b = ((b * 0x80200802ULL) & 0x0884422110ULL) * 0x0101010101ULL >> 32;

The following shows the flow of the bit values with the boolean variables a, b, c, d, e, f, g, and h, which comprise an 8-bit byte. Notice how the first multiply fans out the bit pattern to multiple copies, while the last multiply combines them in the fifth byte from the right.
abcd efgh (-> hgfe dcba)
* 1000 0000 0010 0000 0000 1000 0000 0010 (0x80200802)
-------------------------------------------------------------------------------------------------
0abc defg h00a bcde fgh0 0abc defg h00a bcde fgh0
& 0000 1000 1000 0100 0100 0010 0010 0001 0001 0000 (0x0884422110)
-------------------------------------------------------------------------------------------------
0000 d000 h000 0c00 0g00 00b0 00f0 000a 000e 0000
* 0000 0001 0000 0001 0000 0001 0000 0001 0000 0001 (0x0101010101)
-------------------------------------------------------------------------------------------------
0000 d000 h000 0c00 0g00 00b0 00f0 000a 000e 0000
0000 d000 h000 0c00 0g00 00b0 00f0 000a 000e 0000
0000 d000 h000 0c00 0g00 00b0 00f0 000a 000e 0000
0000 d000 h000 0c00 0g00 00b0 00f0 000a 000e 0000
0000 d000 h000 0c00 0g00 00b0 00f0 000a 000e 0000
-------------------------------------------------------------------------------------------------
0000 d000 h000 dc00 hg00 dcb0 hgf0 dcba hgfe dcba hgfe 0cba 0gfe 00ba 00fe 000a 000e 0000
>> 32
-------------------------------------------------------------------------------------------------
0000 d000 h000 dc00 hg00 dcb0 hgf0 dcba hgfe dcba
& 1111 1111
-------------------------------------------------------------------------------------------------
hgfe dcba

Note that the last two steps can be combined on some processors because the registers can be accessed as bytes; just multiply so that a register stores the upper 32 bits of the result and the take the low byte. Thus, it may take only 6 operations.
Devised by Sean Anderson, July 13, 2001.



--------------------------------------------------------------------------------

Reverse the bits in a byte with 7 operations (no 64-bit):
b = ((b * 0x0802LU & 0x22110LU) | (b * 0x8020LU & 0x88440LU)) * 0x10101LU >> 16

Devised by Sean Anderson, July 13, 2001. Typo spotted and correction supplied by Mike Keith, January 3, 2002.


--------------------------------------------------------------------------------

Reverse an N-bit quantity in parallel in 5 * lg(N) operations:
unsigned int v; // 32 bit word to reverse bit order

// swap odd and even bits
v = ((v >> 1) & 0x55555555) | ((v & 0x55555555) << 1);
// swap consecutive pairs
v = ((v >> 2) & 0x33333333) | ((v & 0x33333333) << 2);
// swap nibbles ...
v = ((v >> 4) & 0x0F0F0F0F) | ((v & 0x0F0F0F0F) << 4);
// swap bytes
v = ((v >> 8) & 0x00FF00FF) | ((v & 0x00FF00FF) << 8);
// swap 2-byte long pairs
v = ( v >> 16 ) | ( v << 16);

The following variation is also O(lg(N)), however it requires more operations to reverse v. Its virtue is in taking less slightly memory by computing the constants on the fly.
unsigned int s = sizeof(v) * CHAR_BIT; // bit size; must be power of 2
unsigned int mask = ~0;
while ((s >>= 1) > 0)
{
mask ^= (mask << s);
v = ((v >> s) & mask) | ((v << s) & ~mask);
}

These methods above are best suited to situations where N is large.
See Dr. Dobb's Journal 1983, Edwin Freed's article on Binary Magic Numbers for more information. The second variation was suggested by Ken Raeburn on September 13, 2005. Veldmeijer mentioned that the first version could do without ANDS in the last line on March 19, 2006.



--------------------------------------------------------------------------------

Compute modulus division by 1 << s without a division operator
const unsigned int n; // numerator
const unsigned int s;
const unsigned int d = 1 << s; // So d will be one of: 1, 2, 4, 8, 16, 32, ...
unsigned int m; // m will be n % d
m = n & (d - 1);

Most programmers learn this trick early, but it was included for the sake of completeness.


--------------------------------------------------------------------------------


Compute modulus division by (1 << s) - 1 without a division operator
unsigned int n; // numerator
const unsigned int s; // s > 0
const unsigned int d = (1 << s) - 1; // so d is either 1, 3, 7, 15, 31, ...).
unsigned int m; // n % d goes here.

for (m = n; n > d; n = m)
{
for (m = 0; n; n >>= s)
{
m += n & d;
}
}
// Now m is a value from 0 to d, but since with modulus division
// we want m to be 0 when it is d.
m = m == d ? 0 : m;

This method of modulus division by an integer that is one less than a power of 2 takes at most 5 + (4 + 5 * ceil(N / s)) * ceil(lg(N / s)) operations, where N is the number of bits in the numerator. In other words, it takes at most O(N * lg(N)) time.
Devised by Sean Anderson, August 15, 2001. Before Sean A. Irvine corrected me on June 17, 2004, I mistakenly commented that we could alternatively assign m = ((m + 1) & d) - 1; at the end. Michael Miller spotted a typo in the code April 25, 2005.



--------------------------------------------------------------------------------

Compute modulus division by (1 << s) - 1 in parallel without a division operator

// The following is for a word size of 32 bits!

const unsigned int M[] =
{
0x00000000, 0x55555555, 0x33333333, 0xc71c71c7,
0x0f0f0f0f, 0xc1f07c1f, 0x3f03f03f, 0xf01fc07f,
0x00ff00ff, 0x07fc01ff, 0x3ff003ff, 0xffc007ff,
0xff000fff, 0xfc001fff, 0xf0003fff, 0xc0007fff,
0x0000ffff, 0x0001ffff, 0x0003ffff, 0x0007ffff,
0x000fffff, 0x001fffff, 0x003fffff, 0x007fffff,
0x00ffffff, 0x01ffffff, 0x03ffffff, 0x07ffffff,
0x0fffffff, 0x1fffffff, 0x3fffffff, 0x7fffffff
};

const unsigned int Q[][6] =
{
{ 0, 0, 0, 0, 0, 0}, {16, 8, 4, 2, 1, 1}, {16, 8, 4, 2, 2, 2},
{15, 6, 3, 3, 3, 3}, {16, 8, 4, 4, 4, 4}, {15, 5, 5, 5, 5, 5},
{12, 6, 6, 6 , 6, 6}, {14, 7, 7, 7, 7, 7}, {16, 8, 8, 8, 8, 8},
{ 9, 9, 9, 9, 9, 9}, {10, 10, 10, 10, 10, 10}, {11, 11, 11, 11, 11, 11},
{12, 12, 12, 12, 12, 12}, {13, 13, 13, 13, 13, 13}, {14, 14, 14, 14, 14, 14},
{15, 15, 15, 15, 15, 15}, {16, 16, 16, 16, 16, 16}, {17, 17, 17, 17, 17, 17},
{18, 18, 18, 18, 18, 18}, {19, 19, 19, 19, 19, 19}, {20, 20, 20, 20, 20, 20},
{21, 21, 21, 21, 21, 21}, {22, 22, 22, 22, 22, 22}, {23, 23, 23, 23, 23, 23},
{24, 24, 24, 24, 24, 24}, {25, 25, 25, 25, 25, 25}, {26, 26, 26, 26, 26, 26},
{27, 27, 27, 27, 27, 27}, {28, 28, 28, 28, 28, 28}, {29, 29, 29, 29, 29, 29},
{30, 30, 30, 30, 30, 30}, {31, 31, 31, 31, 31, 31}
};

const unsigned int R[][6] =
{
{0x00000000, 0x00000000, 0x00000000, 0x00000000, 0x00000000, 0x00000000},
{0x0000ffff, 0x000000ff, 0x0000000f, 0x00000003, 0x00000001, 0x00000001},
{0x0000ffff, 0x000000ff, 0x0000000f, 0x00000003, 0x00000003, 0x00000003},
{0x00007fff, 0x0000003f, 0x00000007, 0x00000007, 0x00000007, 0x00000007},
{0x0000ffff, 0x000000ff, 0x0000000f, 0x0000000f, 0x0000000f, 0x0000000f},
{0x00007fff, 0x0000001f, 0x0000001f, 0x0000001f, 0x0000001f, 0x0000001f},
{0x00000fff, 0x0000003f, 0x0000003f, 0x0000003f, 0x0000003f, 0x0000003f},
{0x00003fff, 0x0000007f, 0x0000007f, 0x0000007f, 0x0000007f, 0x0000007f},
{0x0000ffff, 0x000000ff, 0x000000ff, 0x000000ff, 0x000000ff, 0x000000ff},
{0x000001ff, 0x000001ff, 0x000001ff, 0x000001ff, 0x000001ff, 0x000001ff},
{0x000003ff, 0x000003ff, 0x000003ff, 0x000003ff, 0x000003ff, 0x000003ff},
{0x000007ff, 0x000007ff, 0x000007ff, 0x000007ff, 0x000007ff, 0x000007ff},
{0x00000fff, 0x00000fff, 0x00000fff, 0x00000fff, 0x00000fff, 0x00000fff},
{0x00001fff, 0x00001fff, 0x00001fff, 0x00001fff, 0x00001fff, 0x00001fff},
{0x00003fff, 0x00003fff, 0x00003fff, 0x00003fff, 0x00003fff, 0x00003fff},
{0x00007fff, 0x00007fff, 0x00007fff, 0x00007fff, 0x00007fff, 0x00007fff},
{0x0000ffff, 0x0000ffff, 0x0000ffff, 0x0000ffff, 0x0000ffff, 0x0000ffff},
{0x0001ffff, 0x0001ffff, 0x0001ffff, 0x0001ffff, 0x0001ffff, 0x0001ffff},
{0x0003ffff, 0x0003ffff, 0x0003ffff, 0x0003ffff, 0x0003ffff, 0x0003ffff},
{0x0007ffff, 0x0007ffff, 0x0007ffff, 0x0007ffff, 0x0007ffff, 0x0007ffff},
{0x000fffff, 0x000fffff, 0x000fffff, 0x000fffff, 0x000fffff, 0x000fffff},
{0x001fffff, 0x001fffff, 0x001fffff, 0x001fffff, 0x001fffff, 0x001fffff},
{0x003fffff, 0x003fffff, 0x003fffff, 0x003fffff, 0x003fffff, 0x003fffff},
{0x007fffff, 0x007fffff, 0x007fffff, 0x007fffff, 0x007fffff, 0x007fffff},
{0x00ffffff, 0x00ffffff, 0x00ffffff, 0x00ffffff, 0x00ffffff, 0x00ffffff},
{0x01ffffff, 0x01ffffff, 0x01ffffff, 0x01ffffff, 0x01ffffff, 0x01ffffff},
{0x03ffffff, 0x03ffffff, 0x03ffffff, 0x03ffffff, 0x03ffffff, 0x03ffffff},
{0x07ffffff, 0x07ffffff, 0x07ffffff, 0x07ffffff, 0x07ffffff, 0x07ffffff},
{0x0fffffff, 0x0fffffff, 0x0fffffff, 0x0fffffff, 0x0fffffff, 0x0fffffff},
{0x1fffffff, 0x1fffffff, 0x1fffffff, 0x1fffffff, 0x1fffffff, 0x1fffffff},
{0x3fffffff, 0x3fffffff, 0x3fffffff, 0x3fffffff, 0x3fffffff, 0x3fffffff},
{0x7fffffff, 0x7fffffff, 0x7fffffff, 0x7fffffff, 0x7fffffff, 0x7fffffff}
};

unsigned int n; // numerator
const unsigned int s; // s > 0
const unsigned int d = (1 << s) - 1; // so d is either 1, 3, 7, 15, 31, ...).
unsigned int m; // n % d goes here.

m = (n & M[s]) + ((n & ~M[s]) >> s);

for (const unsigned int * q = &Q[s][0], * r = &R[s][0]; m > d; q++, r++)
{
m = (m >> *q) + (m & *r);
}
m = m == d ? 0 : m; // OR, less portably: m = m & -((signed)(m - d) >> s);

This method of finding modulus division by an integer that is one less than a power of 2 takes at most O(lg(N)) time, where N is the number of bits in the numerator (32 bits, for the code above). The number of operations is at most 12 + 9 * ceil(lg(N)). The tables may be removed if you know the denominator at compile time; just extract the few relevent entries and unroll the loop. It may be easily extended to more bits.
It finds the result by summing the values in base (1 << s) in parallel. First every other base (1 << s) value is added to the previous one. Imagine that the result is written on a piece of paper. Cut the paper in half, so that half the values are on each cut piece. Align the values and sum them onto a new piece of paper. Repeat by cutting this paper in half (which will be a quarter of the size of the previous one) and summing, until you cannot cut further. After performing lg(N/s/2) cuts, we cut no more; just continue to add the values and put the result onto a new piece of paper as before, while there are at least two s-bit values.

Devised by Sean Anderson, August 20, 2001. A typo was spotted by Randy E. Bryant on May 3, 2005 (after pasting the code, I had later added "unsinged" to a variable declaration). As in the previous hack, I mistakenly commented that we could alternatively assign m = ((m + 1) & d) - 1; at the end, and Don Knuth corrected me on April 19, 2006 and suggested m = m & -((signed)(m - d) >> s).



--------------------------------------------------------------------------------

Find the log base 2 of an integer with the MSB N set in O(N) operations (the obvious way)
unsigned int v; // 32-bit word to find the log base 2 of
unsigned r = 0; // r will be lg(v)

while (v >>= 1) // unroll for more speed...
{
r++;
}

The log base 2 of an integer is the same as the position of the highest bit set (or most significant bit set, MSB). The following log base 2 methods are faster than this one.


--------------------------------------------------------------------------------

Find the integer log base 2 of an integer with an 64-bit IEEE float
int v; // 32-bit integer to find the log base 2 of
int r; // result of log_2(v) goes here
union { unsigned int u[2]; double d; } t; // temp

t.u[BYTE_ORDER==LITTLE_ENDIAN] = 0x43300000;
t.u[BYTE_ORDER!=LITTLE_ENDIAN] = v;
t.d -= 4503599627370496.0;
r = (t.u[BYTE_ORDER==LITTLE_ENDIAN] >> 20) - 0x3FF;

The code above loads a 64-bit (IEEE-754 floating-point) double with a 32 bit integer by storing the integer in the mantissa while the exponent is set to 252. From this newly minted double, 252 (expressed as a double) is subtracted, which sets the resulting exponent to the log base 2 of the input value, v. All that is left is shifting the exponent bits into position (20 bits right) and subtracting the bias, 0x3FF (which is 1023 decimal). This technique only takes 5 operations, but many CPUs are slow at manipulating doubles, and the endianess of the architecture must be accommodated.
Eric Cole sent me this on January 15, 2006. Evan Felix pointed out a typo on April 4, 2006.



--------------------------------------------------------------------------------

Find the log base 2 of an integer with a lookup table
static const char LogTable256[] =
{
0, 0, 1, 1, 2, 2, 2, 2, 3, 3, 3, 3, 3, 3, 3, 3,
4, 4, 4, 4, 4, 4, 4, 4, 4, 4, 4, 4, 4, 4, 4, 4,
5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5,
5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5, 5,
6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6,
6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6,
6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6,
6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6, 6,
7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7,
7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7,
7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7,
7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7,
7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7,
7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7,
7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7,
7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7, 7
};

unsigned int v; // 32-bit word to find the log of
unsigned r = 0; // r will be lg(v)
register unsigned int t, tt; // temporaries

if (tt = v >> 16)
{
r = (t = v >> 24) ? 24 + LogTable256[t] : 16 + LogTable256[tt & 0xFF];
}
else
{
r = (t = v >> 8) ? 8 + LogTable256[t] : LogTable256[v];
}


The lookup table method takes only about 7 operations to find the log of a 32-bit value. If extended for 64-bit quantities, it would take roughly 9 operations. Another operation can be trimmed off by using four tables, with the possible additions incorporated into each. Using int table elements may be faster, depending on your architecture.

// To initially generate the log table algorithmically:
LogTable256[0] = LogTable256[1] = 0;
for (int i = 2; i < 256; i++)
{
LogTable256[i] = 1 + LogTable256[i / 2];
}

Behdad Esfahbod and I shaved off a fraction of an operation (on average) on May 18, 2005.
--------------------------------------------------------------------------------

Find the log base 2 of an N-bit integer in O(lg(N)) operations
unsigned int v; // 32-bit value to find the log2 of
const unsigned int b[] = {0x2, 0xC, 0xF0, 0xFF00, 0xFFFF0000};
const unsigned int S[] = {1, 2, 4, 8, 16};
int i;

register unsigned int r = 0; // result of log2(v) will go here
for (i = 4; i >= 0; i--) // unroll for speed...
{
if (v & b[i])
{
v >>= S[i];
r |= S[i];
}
}

// OR (IF YOUR CPU BRANCHES SLOWLY):

register unsigned int shift;
shift = ( ( v & 0xFFFF0000 ) != 0 ) << 4; v >>= shift; r |= shift;
shift = ( ( v & 0xFF00 ) != 0 ) << 3; v >>= shift; r |= shift;
shift = ( ( v & 0xF0 ) != 0 ) << 2; v >>= shift; r |= shift;
shift = ( ( v & 0xC ) != 0 ) << 1; v >>= shift; r |= shift;
shift = ( ( v & 0x2 ) != 0 ) << 0; v >>= shift; r |= shift;

// OR (IF YOU KNOW v IS A POWER OF 2):

const unsigned int b[] = {0xAAAAAAAA, 0xCCCCCCCC, 0xF0F0F0F0, 0xFF00FF00,
0xFFFF0000};
register unsigned int r = (v & b[0]) != 0;
for (i = 4; i > 0; i--) // unroll for speed...
{
r |= ((v & b[i]) != 0) << i;
}

Of course, to extend the code to find the log of a 33- to 64-bit number, we would append another element, 0xFFFFFFFF00000000, to b, append 32 to S, and loop from 5 to 0. This method is much slower than the earlier table-lookup version, but if you don't want big table or your architecture is slow to access memory, it's a good choice. The second variation involves more operations, but it may be faster on machines with high branch costs (e.g. PowerPC), and it was sent to me by Eric Cole on January 7, 2006. The third variation was suggested to me by John Owens on April 24, 2002; it's faster, but it is only suitable when the input is known to be a power of 2. On May 25, 2003, Ken Raeburn suggested improving the general case by using smaller numbers for b[], which load faster on some architectures (for instance if the word size is 16 bits, then only one load instruction may be needed). These values work for the general version, but not for the special-case version below it, where v is a power of 2; Glenn Slayden brought this oversight to my attention on December 12, 2003.


--------------------------------------------------------------------------------

Find the log base 2 of an N-bit integer in O(lg(N)) operations with multiply and lookup
unsigned int v; // find the log base 2 of 32-bit v
int r; // result goes here

static const int MultiplyDeBruijnBitPosition[32] =
{
0, 1, 28, 2, 29, 14, 24, 3, 30, 22, 20, 15, 25, 17, 4, 8,
31, 27, 13, 23, 21, 19, 16, 7, 26, 12, 18, 6, 11, 5, 10, 9
};

v |= v >> 1; // first round down to power of 2
v |= v >> 2;
v |= v >> 4;
v |= v >> 8;
v |= v >> 16;
v = (v >> 1) + 1;

r = MultiplyDeBruijnBitPosition[(v * 0x077CB531UL) >> 27];

The code above computes the log base 2 of a 32-bit integer with a small table lookup and multiply. It requires only 15 operations, compared to (up to) 20 for the previous method. The purely table-based method requires the fewest operations, but this offers a reasonable compromise between table size and speed. If v is known to be a power of 2, then only the last line is needed (3 operations).
Eric Cole devised this January 8, 2006 after reading about the entry below to round up to a power of 2 and the method below for computing the number of trailing bits with a multiply and lookup using a DeBruijn sequence.



--------------------------------------------------------------------------------

Find integer log base 10 of an integer
unsigned int v; // non-zero 32-bit integer value to compute the log base 10 of
int r; // result goes here
int t; // temporary

static unsigned int const PowersOf10[] =
{1, 10, 100, 1000, 10000, 100000,
1000000, 10000000, 100000000, 1000000000};

t = (IntegerLogBase2(v) + 1) * 1233 >> 12; // (use a lg2 method from above)
r = t - (v < PowersOf10[t]);

The integer log base 10 is computed by first using one of the techniques above for finding the log base 2. By the relationship log10(v) = log2(v) / log2(10), we need to multiply it by 1/log2(10), which is approximately 1233/4096, or 1233 followed by a right shift of 12. Adding one is needed because the IntegerLogBase2 rounds down. Finally, since the value t is only an approximation that may be off by one, the exact value is found by subtracting the result of v < PowersOf10[t].
This method takes 6 more operations than IntegerLogBase2. It may be sped up (on machines with fast memory access) by modifying the log base 2 table-lookup method above so that the entries hold what is computed for t (that is, pre-add, -mulitply, and -shift). Doing so would require a total of only 9 operations to find the log base 10, assuming 4 tables were used (one for each byte of v).

Eric Cole suggested I add a version of this on January 7, 2006.



--------------------------------------------------------------------------------

Find integer log base 2 of a 32-bit IEEE float
const float v; // find int(log2(v)), where v > 0.0 && finite(v) && isnormal(v)
int c; // 32-bit int c gets the result;

c = *(const int *) &v; // OR, for portability: memcpy(&c, &v, sizeof c);
c = (c >> 23) - 127;

The above is fast, but IEEE 754-compliant architectures utilize subnormal (also called denormal) floating point numbers. These have the exponent bits set to zero (signifying pow(2,-127)), and the mantissa is not normalized, so it contains leading zeros and thus the log2 must be computed from the mantissa. To accomodate for subnormal numbers, use the following:
const float v; // find int(log2(v)), where v > 0.0 && finite(v)
int c; // 32-bit int c gets the result;
int x = *(const int *) &v; // OR, for portability: memcpy(&x, &v, sizeof x);

c = x >> 23;

if (c)
{
c -= 127;
}
else
{ // subnormal, so recompute using mantissa: c = intlog2(x) - 149;
register unsigned int t; // temporary
// Note that LogTable256 was defined earlier
if (t = x >> 16)
{
c = LogTable256[t] - 133;
}
else
{
c = (t = x >> 8) ? LogTable256[t] - 141 : LogTable256[x] - 149;
}
}

On June 20, 2004, Sean A. Irvine suggested that I include code to handle subnormal numbers. On June 11, 2005, Falk Hüffner pointed out that ISO C99 6.5/7 specified undefined behavior for the common type punning idiom *(int *)&, though it has worked on 99.9% of C compilers. He proposed using memcpy for maximum portability or a union with a float and an int for better code generation than memcpy on some compilers.


--------------------------------------------------------------------------------

Find integer log base 2 of the pow(2, r)-root of a 32-bit IEEE float (for unsigned integer r)
const int r;
const float v; // find int(log2(pow((double) v, 1. / pow(2, r)))),
// where isnormal(v) and v > 0
int c; // 32-bit int c gets the result;

c = *(const int *) &v; // OR, for portability: memcpy(&c, &v, sizeof c);
c = ((((c - 0x3f800000) >> r) + 0x3f800000) >> 23) - 127;

So, if r is 0, for example, we have c = int(log2((double) v)). If r is 1, then we have c = int(log2(sqrt((double) v))). If r is 2, then we have c = int(log2(pow((double) v, 1./4))).
On June 11, 2005, Falk Hüffner pointed out that ISO C99 6.5/7 left the type punning idiom *(int *)& undefined, and he suggested using memcpy.



--------------------------------------------------------------------------------

Count the consecutive zero bits (trailing) on the right in parallel
unsigned int v; // 32-bit word input to count zero bits on right
unsigned int c = 32; // c will be the number of zero bits on the right,
// so if v is 1101000 (base 2), then c will be 3
const unsigned int B[] = {0x55555555, 0x33333333, 0x0F0F0F0F, 0x00FF00FF, 0x0000FFFF};
const unsigned int S[] = {1, 2, 4, 8, 16}; // Our Magic Binary Numbers

for (int i = 4; i >= 0; --i) // unroll for more speed
{
if (v & B[i])
{
v <<= S[i];
c -= S[i];
}
}

if (v)
{
c--;
}

Here, we are basically doing the same operations as finding the log base 2 in parallel, but the values of b are inverted (in order to count from the right rather than the left), we shift v up rather than down, and c starts at the maximum and is decreased. We also have the additional step at the end, decrementing c if there is anything left in v. The number of operations is at most 4 * lg(N) + 2, roughly, for N bit words.


--------------------------------------------------------------------------------

Count the consecutive zero bits (trailing) on the right by binary search
unsigned int v; // 32-bit word input to count zero bits on right
unsigned int c = 0; // c will be the number of zero bits on the right,
// so if v is 1101000 (base 2), then c will be 3
// NOTE: if 0 == v, then c = 31.
if (v & 0x1)
{
// special case for odd v (assumed to happen half of the time)
}
else
{
if ((v & 0xffff) == 0)
{
v >>= 16;
c += 16;
}
if ((v & 0xff) == 0)
{
v >>= 8;
c += 8;
}
if ((v & 0xf) == 0)
{
v >>= 4;
c += 4;
}
if ((v & 0x3) == 0)
{
v >>= 2;
c += 2;
}
if ((v & 0x1) == 0)
{
c++;
}
}

The code above is similar to the previous method, but it computes the number of trailing zeros by accumulating c in a manner akin to binary search. In the first step, it checks if the bottom 16 bits of v are zeros, and if so, shifts v right 16 bits and adds 16 to c, which reduces the number of bits in v to consider by half. Each of the subsequent conditional steps likewise halves the number of bits until there is only 1. This method is faster than the last one (by about 33%) because the bodies of the if statements are executed less often.
Matt Whitlock suggested this on January 25, 2006.



--------------------------------------------------------------------------------

Count the consecutive zero bits (trailing) on the right by casting to a float
unsigned int v; // find the number of trailing zeros in v
int r; // the result goes here
float f = (float)(v & -v); // cast the least significant bit in v to a float
r = (*(unsigned int *)&f >> 23) - 0x7f;

Although this only takes about 6 operations, the time to convert an integer to a float can be high on some machines. The exponent of the 32-bit IEEE floating point representation is shifted down, and the bias is subtracted to give the position of the least significant 1 bit set in v. If v is zero, then the result is -127.


--------------------------------------------------------------------------------

Count the consecutive zero bits (trailing) on the right with modulus division and lookup
unsigned int v; // find the number of trailing zeros in v
int r; // put the result in r
const int Mod37BitPosition[] = // maps a bit value mod 37 to its position
{
32, 0, 1, 26, 2, 23, 27, 0, 3, 16, 24, 30, 28, 11, 0, 13, 4,
7, 17, 0, 25, 22, 31, 15, 29, 10, 12, 6, 0, 21, 14, 9, 5,
20, 8, 19, 18
};
r = Mod37BitPosition[(-v & v) % 37];

The code above finds the number of zeros that are trailing on the right, so binary 0100 would produce 2. It makes use of the fact that the first 32 bit position values are relatively prime with 37, so performing a modulus division with 37 gives a unique number from 0 to 36 for each. These numbers may then be mapped to the number of zeros using a small lookup table. It uses only 4 operations, however indexing into a table and performing modulus division may make it unsuitable for some situations. I came up with this independently and then searched for a subsequence of the table values, and found it was invented earlier by Reiser, according to Hacker's Delight.


--------------------------------------------------------------------------------

Count the consecutive zero bits (trailing) on the right with multiply and lookup
unsigned int v; // find the number of trailing zeros in 32-bit v
int r; // result goes here
const int MultiplyDeBruijnBitPosition[32] =
{
0, 1, 28, 2, 29, 14, 24, 3, 30, 22, 20, 15, 25, 17, 4, 8,
31, 27, 13, 23, 21, 19, 16, 7, 26, 12, 18, 6, 11, 5, 10, 9
};
r = MultiplyDeBruijnBitPosition[((v & -v) * 0x077CB531UL) >> 27];

Converting bit vectors to indices of set bits is an example use for this. It requires one more operation than the earlier one involving modulus division, but the multiply may be faster. The expression (v & -v) extracts the least significant 1 bit from v. The constant 0x077CB531UL is a de Bruijn sequence, which produces a unique pattern of bits into the high 5 bits for each possible bit position that it is multiplied against. When there are no bits set, it returns 0. More information can be found by reading the paper Using de Bruijn Sequences to Index 1 in a Computer Word by Charles E. Leiserson, Harald Prokof, and Keith H. Randall.
On October 8, 2005 Andrew Shapira suggested I add this.



--------------------------------------------------------------------------------

Round up to the next highest power of 2 by float casting
unsigned int const v; // Round this 32-bit value to the next highest power of 2
unsigned int r; // Put the result here. (So v=3 -> r=4; v=8 -> r=8)

if (v > 1)
{
float f = (float)v;
unsigned int const t = 1 << ((*(unsigned int *)&f >> 23) - 0x7f);
r = t << (t < v);
}
else
{
r = 1;
}

The code above uses 8 operations, but works on all v <= (1<<31).
Quick and dirty version, for domain of 1 < v < (1<<25):

float f = (float)(v - 1);
r = 1 << ((*(unsigned int*)(&f) >> 23) - 126);

Although the quick and dirty version only uses around 6 operations, it is roughly three times slower than the technique below (which involves 12 operations) when benchmarked on an Athlon™ XP 2100+ CPU. Some CPUs will fare better with it, though.
On September 27, 2005 Andi Smithers suggested I include a technique for casting to floats to find the lg of a number for rounding up to a power of 2. Similar to the quick and dirty version here, his version worked with values less than (1<<25), due to mantissa rounding, but it used one more operation.



--------------------------------------------------------------------------------

Round up to the next highest power of 2
unsigned int v; // compute the next highest power of 2 of 32-bit v

v--;
v |= v >> 1;
v |= v >> 2;
v |= v >> 4;
v |= v >> 8;
v |= v >> 16;
v++;

In 12 operations, this code computes the next highest power of 2 for a 32-bit integer. The result may be expressed by the formula 1 << (lg(v - 1) + 1). Note that in the edge case where v is 0, it returns 0, which isn't a power of 2; you might append the expression v += (v == 0) to remedy this if it matters. It would be faster by 2 operations to use the formula and the log base 2 methed that uses a lookup table, but in some situations, lookup tables are not suitable, so the above code may be best. (On a Athlon™ XP 2100+ I've found the above shift-left and then OR code is as fast as using a single BSR assembly language instruction, which scans in reverse to find the highest set bit.) It works by copying the highest set bit to all of the lower bits, and then adding one, which results in carries that set all of the lower bits to 0 and one bit beyond the highest set bit to 1. If the original number was a power of 2, then the decrement will reduce it to one less, so that we round up to the same original value.
Devised by Sean Anderson, Sepember 14, 2001.



--------------------------------------------------------------------------------

Interleave bits the obvious way
unsigned short x; // Interleave bits of x and y, so that all of the
unsigned short y; // bits of x are in the even positions and y in the odd;
unsigned int z = 0; // z gets the resulting Morton Number.

for (int i = 0; i < sizeof(x) * CHAR_BIT; i++) // unroll for more speed...
{
z |= (x & 1 << i) << i | (y & 1 << i) << (i + 1);
}

Interleaved bits (aka Morton numbers) are useful for linearizing 2D integer coordinates, so x and y are combined into a single number that can be compared easily and has the property that a number is usually close to another if their x and y values are close.


--------------------------------------------------------------------------------

Interleave bits by table lookup
const unsigned short MortonTable256[] =
{
0x0000, 0x0001, 0x0004, 0x0005, 0x0010, 0x0011, 0x0014, 0x0015,
0x0040, 0x0041, 0x0044, 0x0045, 0x0050, 0x0051, 0x0054, 0x0055,
0x0100, 0x0101, 0x0104, 0x0105, 0x0110, 0x0111, 0x0114, 0x0115,
0x0140, 0x0141, 0x0144, 0x0145, 0x0150, 0x0151, 0x0154, 0x0155,
0x0400, 0x0401, 0x0404, 0x0405, 0x0410, 0x0411, 0x0414, 0x0415,
0x0440, 0x0441, 0x0444, 0x0445, 0x0450, 0x0451, 0x0454, 0x0455,
0x0500, 0x0501, 0x0504, 0x0505, 0x0510, 0x0511, 0x0514, 0x0515,
0x0540, 0x0541, 0x0544, 0x0545, 0x0550, 0x0551, 0x0554, 0x0555,
0x1000, 0x1001, 0x1004, 0x1005, 0x1010, 0x1011, 0x1014, 0x1015,
0x1040, 0x1041, 0x1044, 0x1045, 0x1050, 0x1051, 0x1054, 0x1055,
0x1100, 0x1101, 0x1104, 0x1105, 0x1110, 0x1111, 0x1114, 0x1115,
0x1140, 0x1141, 0x1144, 0x1145, 0x1150, 0x1151, 0x1154, 0x1155,
0x1400, 0x1401, 0x1404, 0x1405, 0x1410, 0x1411, 0x1414, 0x1415,
0x1440, 0x1441, 0x1444, 0x1445, 0x1450, 0x1451, 0x1454, 0x1455,
0x1500, 0x1501, 0x1504, 0x1505, 0x1510, 0x1511, 0x1514, 0x1515,
0x1540, 0x1541, 0x1544, 0x1545, 0x1550, 0x1551, 0x1554, 0x1555,
0x4000, 0x4001, 0x4004, 0x4005, 0x4010, 0x4011, 0x4014, 0x4015,
0x4040, 0x4041, 0x4044, 0x4045, 0x4050, 0x4051, 0x4054, 0x4055,
0x4100, 0x4101, 0x4104, 0x4105, 0x4110, 0x4111, 0x4114, 0x4115,
0x4140, 0x4141, 0x4144, 0x4145, 0x4150, 0x4151, 0x4154, 0x4155,
0x4400, 0x4401, 0x4404, 0x4405, 0x4410, 0x4411, 0x4414, 0x4415,
0x4440, 0x4441, 0x4444, 0x4445, 0x4450, 0x4451, 0x4454, 0x4455,
0x4500, 0x4501, 0x4504, 0x4505, 0x4510, 0x4511, 0x4514, 0x4515,
0x4540, 0x4541, 0x4544, 0x4545, 0x4550, 0x4551, 0x4554, 0x4555,
0x5000, 0x5001, 0x5004, 0x5005, 0x5010, 0x5011, 0x5014, 0x5015,
0x5040, 0x5041, 0x5044, 0x5045, 0x5050, 0x5051, 0x5054, 0x5055,
0x5100, 0x5101, 0x5104, 0x5105, 0x5110, 0x5111, 0x5114, 0x5115,
0x5140, 0x5141, 0x5144, 0x5145, 0x5150, 0x5151, 0x5154, 0x5155,
0x5400, 0x5401, 0x5404, 0x5405, 0x5410, 0x5411, 0x5414, 0x5415,
0x5440, 0x5441, 0x5444, 0x5445, 0x5450, 0x5451, 0x5454, 0x5455,
0x5500, 0x5501, 0x5504, 0x5505, 0x5510, 0x5511, 0x5514, 0x5515,
0x5540, 0x5541, 0x5544, 0x5545, 0x5550, 0x5551, 0x5554, 0x5555
};

unsigned short x; // Interleave bits of x and y, so that all of the
unsigned short y; // bits of x are in the even positions and y in the odd;
unsigned int z; // z gets the resulting 32-bit Morton Number.

z = MortonTable256[y >> 8] << 17 |
MortonTable256[x >> 8] << 16 |
MortonTable256[y & 0xFF] << 1 |
MortonTable256[x & 0xFF];


For more speed, use an additional table with values that are MortonTable256 pre-shifted one bit to the left. This second table could then be used for the y lookups, thus reducing the operations by two, but almost doubling the memory required. Extending this same idea, four tables could be used, with two of them pre-shifted by 16 to the left of the previous two, so that we would only need 11 operations total.
--------------------------------------------------------------------------------

Interleave bits with 64-bit multiply
In 11 operations, this version interleaves bits of two bytes (rather than shorts, as in the other versions), but many of the operations are 64-bit multiplies so it isn't appropriate for all machines.
unsigned char x; // Interleave bits of (8-bit) x and y, so that all of the
unsigned char y; // bits of x are in the even positions and y in the odd;
unsigned short z; // z gets the resulting 16-bit Morton Number.

z = ((x * 0x0101010101010101ULL & 0x8040201008040201ULL) *
0x0102040810204081ULL >> 49) & 0x5555 |
((y * 0x0101010101010101ULL & 0x8040201008040201ULL) *
0x0102040810204081ULL >> 48) & 0xAAAA;

Holger Bettag was inspired to suggest this technique on October 10, 2004 after reading the multiply-based bit reversals here.


--------------------------------------------------------------------------------

Interleave bits by Binary Magic Numbers
const unsigned int B[] = {0x55555555, 0x33333333, 0x0F0F0F0F, 0x00FF00FF};
const unsigned int S[] = {1, 2, 4, 8};

unsigned int x; // Interleave lower 16 bits of x and y, so the bits of x
unsigned int y; // are in the even positions and bits from y in the odd;
unsigned int z; // z gets the resulting 32-bit Morton Number.

x = (x | (x << S[3])) & B[3];
x = (x | (x << S[2])) & B[2];
x = (x | (x << S[1])) & B[1];
x = (x | (x << S[0])) & B[0];

y = (y | (y << S[3])) & B[3];
y = (y | (y << S[2])) & B[2];
y = (y | (y << S[1])) & B[1];
y = (y | (y << S[0])) & B[0];

z = x | (y << 1);


--------------------------------------------------------------------------------

Determine if a word has a zero byte
// Fewer operations:
unsigned int v; // 32-bit word to check if any 8-bit byte in it is 0
bool hasZeroByte = ~((((v & 0x7F7F7F7F) + 0x7F7F7F7F) | v) | 0x7F7F7F7F);

The code above may be useful when doing a fast string copy in which a word is copied at a time; it uses 5 operations. On the other hand, testing for a null byte in the obvious ways (which follow) have at least 7 operations (when counted in the most sparing way), and at most 12.
// More operations:
bool hasNoZeroByte = ((v & 0xff) && (v & 0xff00) && (v & 0xff0000) && (v & 0xff000000))
// OR:
unsigned char * p = (unsigned char *) &v;
bool hasNoZeroByte = *p && *(p + 1) && *(p + 2) && *(p + 3);

The code at the beginning of this section (labeled "Fewer operations") works by first zeroing the high bits of the 4 bytes in the word. Subsequently, it adds a number that will result in an overflow to the high bit of a byte if any of the low bits were initialy set. Next the high bits of the original word are ORed with these values; thus, the high bit of a byte is set iff any bit in the byte was set. Finally, we determine if any of these high bits are zero by ORing with ones everywhere except the high bits and inverting the result. Extending to 64 bits is trivial; simply increase the constants to be 0x7F7F7F7F7F7F7F7F.
For an additional improvement, a fast pretest that requires only 4 operations may be performed to determine if the word may have a zero byte. The test also returns true if the high byte is 0x80, so there are occasional false positives, but the slower and more reliable version above may then be used on candidates for an overall increase in speed with correct output.


bool hasZeroByte = ((v + 0x7efefeff) ^ ~v) & 0x81010100;
if (hasZeroByte) // or may just have 0x80 in the high byte
{
hasZeroByte = ~((((v & 0x7F7F7F7F) + 0x7F7F7F7F) | v) | 0x7F7F7F7F);
}

There is yet a faster method — use hasless(v, 1), which is defined below; it works in 4 operations and requires no subsquent verification. It simplifies to

bool hasZeroByte = (v - 0x01010101UL) & ~v & 0x80808080UL;
The subexpression (v - 0x01010101UL), evaluates to a high bit set in any byte whenever the corresponding byte in v is zero or greater than 0x80. The sub-expression ~v & 0x80808080UL evaluates to high bits set in bytes where the byte of v doesn't have its high bit set (so the byte was less than 0x80). Finally, by ANDing these two sub-expressions the result is the high bits set where the bytes in v were zero, since the high bits set due to a value greater than 0x80 in the first sub-expression are masked off by the second.
Paul Messmer suggested the fast pretest improvement on October 2, 2004. Juha Järvi later suggested hasless(v, 1) on April 6, 2005, which he found on Paul Hsieh's Assembly Lab; previously it was written in a newsgroup post on April 27, 1987 by Alan Mycroft.



--------------------------------------------------------------------------------

Determine if a word has a byte less than n
Test if a word x contains an unsigned byte with value < n. Specifically for n=1, it can be used to find a 0-byte by examining one long at a time, or any byte by XORing x with a mask first. Uses 4 arithmetic/logical operations when n is constant.
Requirements: x>=0; 0<=n<=128

#define hasless(x,n) (((x)-~0UL/255*(n))&~(x)&~0UL/255*128)

To count the number of bytes in x that are less than n in 7 operations, use
#define countless(x,n) (((~0UL/255*(127+(n))-((x)&~0UL/255*127))&~(x)&~0UL/255*128)/128%255)

Juha Järvi sent this clever technique to me on April 6, 2005. The countless macro was added by Sean Anderson on April 10, 2005, inspired by Juha's countmore, below.


--------------------------------------------------------------------------------

Determine if a word has a byte greater than n
Test if a word x contains an unsigned byte with value > n. Uses 3 arithmetic/logical operations when n is constant.
Requirements: x>=0; 0<=n<=127

#define hasmore(x,n) (((x)+~0UL/255*(127-(n))|(x))&~0UL/255*128)

To count the number of bytes in x that are more than n in 6 operations, use:
#define countmore(x,n) (((((x)&~0UL/255*127)+~0UL/255*(127-(n))|(x))&~0UL/255*128)/128%255)

The macro hasmore was suggested by Juha Järvi on April 6, 2005, and he added countmore on April 8, 2005.


--------------------------------------------------------------------------------

Determine if a word has a byte between m and n
When m < n, this technique tests if a word x contains an unsigned byte value, such that m < value < n. It uses 7 arithmetic/logical operations when n and m are constant.
Note: Bytes that equal n can be reported by likelyhasbetween as false positives, so this should be checked by character if a certain result is needed.

Requirements: x>=0; 0<=m<=127; 0<=n<=128

#define likelyhasbetween(x,m,n) \ ((((x)-~0UL/255*(n))&~(x)&((x)&~0UL/255*127)+~0UL/255*(127-(m)))&~0UL/255*128)


This technique would be suitable for a fast pretest. A variation that takes one more operation (8 total for constant m and n) but provides the exact answer is:
#define hasbetween(x,m,n) ((~0UL/255*(127+(n))-((x)&~0UL/255*127)&~(x)&((x)&~0UL/255*127)+~0UL/255*(127-m))&~0UL/255*128)

To count the number of bytes in x that are between m and n (exclusive) in 10 operations, use:
#define countbetween(x,m,n) (hasbetween(x,m,n)/128%255)

Juha Järvi suggested likelyhasbetween on April 6, 2005. From there, Sean Anderson created hasbetween and countbetween on April 10, 2005.



--------------------------------------------------------------------------------
Hits:

Fast Bit Counting Routines

Fast Bit Counting Routines
Compiled from various sources by Gurmeet Singh Manku
A common problem asked in job interviews is to count the number of bits that are on in an unsigned integer. Here are seven solutions to this problem. Source code in C is available.


Iterated Count

int bitcount (unsigned int n)
{
int count=0;
while (n)
{
count += n & 0x1u ;
n >>= 1 ;
}
return count ;
}

Sparse Ones

int bitcount (unsigned int n)
{
int count=0 ;
while (n)
{
count++ ;
n &= (n - 1) ;
}
return count ;
}


Dense Ones

int bitcount (unsigned int n)
{
int count = 8 * sizeof(int) ;
n ^= (unsigned int) -1 ;
while (n)
{
count-- ;
n &= (n - 1) ;
}
return count ;
}

Precompute_8bit

// static int bits_in_char [256] ;

int bitcount (unsigned int n)
{
// works only for 32-bit ints

return bits_in_char [n & 0xffu]
+ bits_in_char [(n >> 8) & 0xffu]
+ bits_in_char [(n >> 16) & 0xffu]
+ bits_in_char [(n >> 24) & 0xffu] ;
}



Iterated Count runs in time proportional to the total number of bits. It simply loops through all the bits, terminating slightly earlier because of the while condition. Useful if 1's are sparse and among the least significant bits. Sparse Ones runs in time proportional to the number of 1 bits. The line n &= (n - 1) simply sets the rightmost 1 bit in n to 0. Dense Ones runs in time proportional to the number of 0 bits. It is the same as Sparse Ones, except that it first toggles all bits (n ~= -1), and continually subtracts the number of 1 bits from sizeof(int). Precompute_8bit assumes an array bits_in_char such that bits_in_char[i] contains the number of 1 bits in the binary representation for i. It repeatedly updates count by masking out the last eight bits in n, and indexing into bits_in_char.
Precompute_16bit

// static char bits_in_16bits [0x1u << 16] ;

int bitcount (unsigned int n)
{
// works only for 32-bit ints

return bits_in_16bits [n & 0xffffu]
+ bits_in_16bits [(n >> 16) & 0xffffu] ;
}


Precompute_16bit is a variant of Precompute_8bit in that an array bits_in_16bits[] stores the number of 1 bits in successive 16 bit numbers (shorts).

Parallel Count

#define TWO(c) (0x1u << (c))
#define MASK(c) (((unsigned int)(-1)) / (TWO(TWO(c)) + 1u))
#define COUNT(x,c) ((x) & MASK(c)) + (((x) >> (TWO(c))) & MASK(c))

int bitcount (unsigned int n)
{
n = COUNT(n, 0) ;
n = COUNT(n, 1) ;
n = COUNT(n, 2) ;
n = COUNT(n, 3) ;
n = COUNT(n, 4) ;
/* n = COUNT(n, 5) ; for 64-bit integers */
return n ;
}

Parallel Count carries out bit counting in a parallel fashion. Consider n after the first line has finished executing. Imagine splitting n into pairs of bits. Each pair contains the number of ones in those two bit positions in the original n. After the second line has finished executing, each nibble contains the number of ones in those four bits positions in the original n. Continuing this for five iterations, the 64 bits contain the number of ones among these sixty-four bit positions in the original n. That is what we wanted to compute.

Nifty Parallel Count

#define MASK_01010101 (((unsigned int)(-1))/3)
#define MASK_00110011 (((unsigned int)(-1))/5)
#define MASK_00001111 (((unsigned int)(-1))/17)

int bitcount (unsigned int n)
{
n = (n & MASK_01010101) + ((n >> 1) & MASK_01010101) ;
n = (n & MASK_00110011) + ((n >> 2) & MASK_00110011) ;
n = (n & MASK_00001111) + ((n >> 4) & MASK_00001111) ;
return n % 255 ;
}

Nifty Parallel Count works the same way as Parallel Count for the first three iterations. At the end of the third line (just before the return), each byte of n contains the number of ones in those eight bit positions in the original n. A little thought then explains why the remainder modulo 255 works.

MIT HACKMEM Count

int bitcount(unsigned int n)
{
/* works for 32-bit numbers only */
/* fix last line for 64-bit numbers */

register unsigned int tmp;

tmp = n - ((n >> 1) & 033333333333)
- ((n >> 2) & 011111111111);
return ((tmp + (tmp >> 3)) & 030707070707) % 63;
}

MIT Hackmem Count is funky. Consider a 3 bit number as being 4a+2b+c. If we shift it right 1 bit, we have 2a+b. Subtracting this from the original gives 2a+b+c. If we right-shift the original 3-bit number by two bits, we get a, and so with another subtraction we have a+b+c, which is the number of bits in the original number. How is this insight employed? The first assignment statement in the routine computes tmp. Consider the octal representation of tmp. Each digit in the octal representation is simply the number of 1's in the corresponding three bit positions in n. The last return statement sums these octal digits to produce the final answer. The key idea is to add adjacent pairs of octal digits together and then compute the remainder modulus 63. This is accomplished by right-shifting tmp by three bits, adding it to tmp itself and ANDing with a suitable mask. This yields a number in which groups of six adjacent bits (starting from the LSB) contain the number of 1's among those six positions in n. This number modulo 63 yields the final answer. For 64-bit numbers, we would have to add triples of octal digits and use modulus 1023. This is HACKMEM 169, as used in X11 sources. Source: MIT AI Lab memo, late 1970's.

No Optimization Some Optimization Heavy Optimization

Precomp_16 52.94 Mcps Precomp_16 76.22 Mcps Precomp_16 80.58 Mcps
Precomp_8 29.74 Mcps Precomp_8 49.83 Mcps Precomp_8 51.65 Mcps
Parallel 19.30 Mcps Parallel 36.00 Mcps Parallel 38.55 Mcps
MIT 16.93 Mcps MIT 17.10 Mcps Nifty 31.82 Mcps
Nifty 12.78 Mcps Nifty 16.07 Mcps MIT 29.71 Mcps
Sparse 5.70 Mcps Sparse 15.01 Mcps Sparse 14.62 Mcps
Dense 5.30 Mcps Dense 14.11 Mcps Dense 14.56 Mcps
Iterated 3.60 Mcps Iterated 3.84 Mcps Iterated 9.24 Mcps

Mcps = Million counts per second

Which of the several bit counting routines is the fastest? Results of speed trials on an i686 are summarized in the table on left. "No Optimization" was compiled with plain gcc. "Some Optimizations" was gcc -O3. "Heavy Optimizations" corresponds to gcc -O3 -mcpu=i686 -march=i686 -fforce-addr -funroll-loops -frerun-cse-after-loop -frerun-loop-opt -malign-functions=4.


Thanks to Seth Robertson who suggested performing speed trials by extending bitcount.c. Seth also pointed me to MIT_Hackmem routine. Thanks to Denny Gursky who suggested the idea of Precompute_11bit. That would require three sums (11-bit, 11-bit and 10-bit precomputed counts). I then tried Precompute_16bit which turned out to be even faster.

If you have niftier solutions up your sleeves, please send me an e-mail


--------------------------------------------------------------------------------

Saturday, October 07, 2006

Binary Magic Numbers for counting bits...

f...@cityscape.co.uk (Tony Yule) wrote:
>Twenty five years ago I attended a Lecture on aspects of programming. During the lecture an
>algorithm for counting the number of bits set in a word was described. As I remember the
>algorithm only involved logical operations and shifts; i.e. no loops were involved.

>At the time the site operated a CDC 6600 computer which had an instruction which performed this
>function. The operation was supported by the divide unit. (The 6600 had a number of units
>that supported different types of operations.)


>I have never seen this algorithm described and have been unable to reinvent it.


>I apologise for posting this request here but it is the only newsgroup that I have been able
>to find with "algorithm" in the title.


>If anyone knows the algorithm I would be grateful of a description or a reference.


>Otherwise if the newsgroup for such a request exists I would appreciate its name.


>Replies to apy...@bcs.org.uk please.



What you want is called "sideways addition." The naive approach is:

int NaiveSideSum( int val )
{
int temp = value;
int result = 0;


while (temp != 0)
{
result += temp & 0x01;
temp >> 1;
}
return result;



}


Nothing too exciting here. The worst-case performance of this
algorithm is N iterations of the WHILE loop (where N is the number of
bit positions in an integer). Of course, the maximum number for any
instance is determined by the most significant bit of the input value.

Fortunately, we can operate on multiple fields within the input value
simultaneously by first adding adjacent bits, then bit pairs, and so
on. This process, which involves "binary magic numbers," takes V
iterations in all cases, where V is the base-2 logarithm of N rounded
to the next highest integer.


Here's the function in C (with following explanations):


int FastSideSum( int value )
{
int i;
int result = value;


for (i = 0; i < V; i++)
result = ((result >> S[i]) & B[i]) + (result & B[i]);


return result;



}


Yes, there is still a loop involved. However, you can always unroll it
for any reasonable value of V. Thus for V = 4 (i.e, N = 16):

int SuperFastSideSum_16( int value )
{
int result = value;


result = ((result >> S[1]) & B[1]) + (result & B[1]);
result = ((result >> S[2]) & B[2]) + (result & B[2]);
result = ((result >> S[3]) & B[3]) + (result & B[3]);
result = ((result >> S[4]) & B[4]) + (result & B[4]);


return result;



}


If you look at the compiler output for these two functions and
calculate the number of machine cycles required for each, it becomes
evident that the binary magic number approach is preferable for large
N's.

Now, to explain S[] and B[]. The B array consists of binary magic
numbers, which are thoroughly discussed in:


Freed, Edwin E. 1983. "Binary Magic Numbers," Dr. Dobb's Journal
Vol. 78 (April), pp. 24-37.


Magic numbers in mathematics are numbers which have special or unusual
properties when used in certain calculations. Freed examined a class
of binary magic numbers that can be used to:


(a) Determine the positions of bits within words;
(b) Reverse, permute and map bits within words;
(c) Compute sideways (unweighted and weighted) sums and parity; and
(d) Convert Gray code values to their binary equivalents.


Binary magic numbers offer improved (i.e., faster) algorithms over
those that can otherwise be developed, typically by a factor of N /
log-2(N).


Freed's numbers are members of a sequence, where the Nth number of the
sequence is itself an infinite sequence from right to left of 2**N 1's
followed by 2**N 0's, followed 2**N 1's, and so on. The initial
numbers are:


...0101010101010101
...0011001100110011
...0000111100001111
...0000000011111111
...


For a word size of 16 bits then, we have four "B-constants":


B[1] = 0101010101010101
B[2] = 0011001100110011
B[3] = 0000111100001111
B[4] = 0000000011111111


There are three B-constants for words sizes (N) of 8 or fewer bits,
five B-constants for words sizes of 17 to 32 bits, and so on. As
mentioned above, N is the word size in bits, and V is the base-2
logarithm of N rounded to the next highest integer. B[1] through B[V]
are thus the binary magic numbers truncated to the N least-significant
bits, and V is:


N V
----------
8 3
9-16 4
17-32 5
33-64 6
...


The constants S[1] through S[N] are a table of the powers of 2. That
is:


S[1] = 2
S[2] = 4
S[3] = 8
S[4] = 16
...


About those CDC computers: According to Freed, the CDC 6400, 6500, and
6600 series processors had a "CXi Xk" instruction that counted the
number of nonzero bits in register Xk and stored the result in
register Xi. These were clearly CISC processors!


The binary magic numbers and associated algorithms that Freed
discusses are a godsend for anyone whose algorithms involve bit
twiddling. (One example of particular interest to computer graphics
mavericks is the bit reversal process required by the Fast Fourier
Transform and related algorithms.) Unfortunately, old back issues of
Dr. Dobb's Journal are *very* hard to find, and the DDJ Source Code
CD-ROM only goes back to 1986.


Freed (who was at the Mathematics Department of the Harvey Mudd
College in Claremont, California when he wrote his article)
acknowledged the assistance of Donald Knuth, and noted that many of
his algorithms were derived from the preprint version of Knuth's "The
Art of Computer Programming - Combinatorial Algorithms (Volume 4)."
Unless I missed it, however, this book was never published.


If anyone knows of another reference (one that is easier to find in a
good mathematics or computer science research lbrary, that is)
regarding binary magic numbers and their applications, I for one would
like to know about it :+)


(One paper of interest -- I wasn't aware of it myself until I began
writing this response -- may be:


Guibas, L., and J. Stolfi. 1981. "A Language for Bitmap
Manipulation," ACM Transactions on Graphics 1(3):192-214


which deals with transposing bit matrices. The paper likely refers to
previous papers on the topic.)


Dr. Dobb's Journal began life as "Dr. Dobb's Journal of Computer
Calisthenics & Orthdontia (Running Light Without Overbyte)." In
between discussions of CP/M games and Small C, it had some timeless
articles that are still worth their weight in gold. I'm glad I kept my
bsck issues!


Ian Ashdown, P. Eng. | READ THE BOOK!
Research & Development Manager | Radiosity: A Programmer's Perspective
Ledalite Architectural Products Inc. | by Ian Ashdown
| John Wiley & Sons, 1994
Visit http://www.ledalite.com | (Sneaky Internet Advertising)