Probability of commuting

A couple years ago I wrote a blog post looking at how close the quaternions come to commuting. That is, the post looked at the average norm of xyyx.

A related question would be to ask how often quaternions do commute, i.e. the probability that xyyx = 0 for randomly chosen x and y.

There’s a general theorem for this [1]. For a discrete non-abelian group, the probability that two elements commute, chosen uniformly at random, is never more than 5/8 for any group.

To put it another way, in a finite group either all pairs of elements commute with each other or no more than 5/8 of all pairs commute, with no possibilities in between. You can’t have a group, for example, in which exactly 3 out of 4 pairs commute.

What if we have an infinite group like the quaternions?

Before we can answer that, we’ve got to say how we’d compute probabilities. With a finite group, the natural thing to do is make every point have equal probability. For a (locally compact) infinite group the natural choice is Haar measure.

Subject to some technical conditions, Haar measure is the only measure that interacts as expected with the group structure. It’s unique up to a constant multiple, and so it’s unique when we specify that the measure of the whole group has to be 1.

For compact non-abelian groups with Haar measure, we again get the result that no more than 5/8 of pairs commute.

[1] W. H. Gustafson. What is the Probability that Two Group Elements Commute? The American Mathematical Monthly, Nov., 1973, Vol. 80, No. 9, pp. 1031-1034.

The word problem

Most people have heard of word problems, but not as many have heard of the word problem. If you’re imagining that the word problem is some superlatively awful word problem, I can assure you it’s not. It’s both simpler and weirder than that.

The word problem is essentially about whether you can always apply algebraic rules in an automated way. The reason it is called the word problem is that you start by a description of your algebraic system in terms of symbols (“letters”) and concatenations of symbols (“words”) subject to certain rules, also called relations.

The word problem for groups

For example, you can describe a group by saying it contains a and b, and it satisfies the relations

a² = b²

and

a-1ba = b-1.

A couple things are implicit here. We’ve said this a group, and since every element in a group has an inverse, we’ve implied that a-1 and b-1 are in the group as well. Also from the definition of a group comes the assumption that multiplication is associative, that there’s an identity element, and that inverses work like they’re supposed to.

In the example above, you could derive everything about the group from the information given. In particular, someone could give you two words—strings made up of a, b, a-1, and b-1—and you could determine whether they are equal by applying the rules. But in general, this is not possible for groups.

Undecidable

The bad news is that in general this isn’t possible. In computer science terminology, the word problem is undecidable. There is no algorithm that can tell whether two words are equal given a list of relations, at least not in general. There are special cases where the word problem is solvable, but a general algorithm is not possible.

The word problem for semigroups

I presented the word problem above in the context of groups, but you could look at the word problem in more general contexts, such as semigroups. A semigroup is closed under some associative binary operation, and that’s it. There need not be any inverses or even an identity element.

Here’s a concrete example of a semigroup whose word problem has been proven to be undecidable. As before we have two symbols, a and b. And because we are in a semigroup, not a group, there are no inverses. Our semigroup consists of all finite sequences make out of a‘s and b‘s, subject to these five relations.

aba2b2 = b2a2ba

a2bab2a = b2a3ba

aba3b2 = ab2aba2

b3a2b2a2ba = b3a2b2a4

a4b2a2ba = b2a4

Source: Term Rewriting and All That

Experience

When I first saw groups presented this as symbols and relations, I got my hopes up that a large swath of group theory could be automated. A few minutes later my naive hopes were dashed. So in my mind I thought “Well, then this is hopeless.”

But that is not true. Sometimes the word problem is solvable. It’s like many other impossibility theorems. There’s no fifth degree analog of the quadratic equation in general, but there are fifth degree polynomials whose roots can be found in closed form. There’s no program that can tell whether any arbitrary program will halt, but that doesn’t mean you can’t tell whether some programs halt.

It didn’t occur to me at the time that it would be worthwhile to explore the boundaries, learning which word problems can or cannot be solved. It also didn’t occur to me that I would run into things like the word problem in practical applications, such as simplifying symbolic expressions and optimizing their evaluation. Undecidable problems lurk everywhere, but you can often step around them.

Nonlinear mod 5

This post is a follow-on to the previous post on perfectly nonlinear functions. In that post we defined a way to measure the degree of nonlinearity of a function between two Abelian groups. We looked at functions that take sequences of four bits to a single bit. In formal terms, our groups were GF(24) and GF(2).

In this post we’ll start out by looking at integers mod 5 to show how the content of the previous post takes a different form in different groups. Sometimes the simplicity of working in binary makes things harder to see. For example, we noted in the previous post that the distinction between addition and subtraction didn’t matter. It matters in general!

Let B be the group of integers mod 5 and A the group of pairs of integers mod 5. That is, A is the direct product B × B. We will compute the uniformity of several functions from A to B. (Recall that less uniformity means more nonlinearity.) Then we’ll conclude by looking what happens if we work modulo other integers.

Python code

Here’s the code we’ll use to compute uniformity.

    from itertools import product
    from numpy import array

    m = 5
    r = range(m)

    def deriv(f, x, a):
        return (f(x + a) - f(x))%m

    def delta(f, a, b):
        return {x for x in product(r, r) if all(deriv(f, array(x), a) == b)}

    def uniformity(f):
        u = 0
        for a in product(r, r):
            if a == (0, 0):
                continue
            for b in product(r, r):
                t = len(delta(f, array(a), array(b)))
                if t > u:
                    u = t
        return u

We didn’t call attention to it last time, but the function f(x + a) – f(x) is called a derivative of f, and that’s why we named the function above deriv.

Here are a few functions whose uniformity we’d like to compute.

    # (u, v) -> u^2 + v^2 
    def F2(x): return sum(x**2)%m

    # (u, b) -> u^3 + v^3 
    def F3(x): return sum(x**3)%m

    # (u, b) -> u^2 + v
    def G(x):  return (x[0]**2 + x[1])%m

    # (u, v) -> uv
    def H(x):  return (x[0]*x[1])%m

Now let’s look at their uniformity values.

    print( uniformity(F2) )
    print( uniformity(F3) )
    print( uniformity(G) )
    print( uniformity(H) )

Results mod 5

This prints out 5, 10, 25, and 5. This says that the functions F2 and H are the most nonlinear, in fact “perfectly nonlinear.” The function G, although nonlinear, has the same uniformity as a linear function.

The function F3 may be the most interesting, having intermediate uniformity. In some sense the sum of cubes is closer to being a linear function than the sum of squares is.

Other moduli

We can easily look at other groups by simply changing the value of m at the top of the code.

If we set the modulus to 7, we get analogous results. The uniformities of the four functions are 7, 14, 49, and 7. They’re ranked exactly as they were when the modulus was 5.

But that’s not always the case for other moduli as you can see in the table below. Working mod 8, for example, the sum of squares is more uniform than the sum of cubes, the opposite of the case mod 5 or mod 7.

    |----+-----+-----+-----+----|
    |  m |  F2 |  F3 |   G |  H |
    |----+-----+-----+-----+----|
    |  2 |   4 |   4 |   4 |  2 |
    |  3 |   3 |   9 |   9 |  3 |
    |  4 |  16 |   8 |  16 |  8 |
    |  5 |   5 |  10 |  25 |  5 |
    |  6 |  36 |  36 |  36 | 18 |
    |  7 |   7 |  14 |  49 |  7 |
    |  8 |  64 |  32 |  64 | 32 |
    |  9 |  27 |  81 |  81 | 27 |
    | 10 | 100 | 100 | 100 | 50 |
    | 11 |  11 |  22 | 121 | 11 |
    | 12 | 144 | 144 | 144 | 72 |
    |----+-----+-----+-----+----|

In every case, G is the most uniform function and H is the least. Also, G is strictly more uniform than H. But there are many different ways F2 and F3 can fit between G and H.

How big is the monster?

Voyager 1 photo of Jupiter. Jan. 6, 1979

Symmetries are captured by mathematical groups. And just as you can combine kinds symmetry to form new symmetries, you can combine groups to form new groups.

So-called simple groups are the building blocks of groups as prime numbers are the building blocks of integers [1].

Finite simple groups have been fully classified, and they fall into several families, with 26 exceptions that fall outside any of these families. The largest of these exceptional groups is called the monster.

The monster is very large, containing approximately 8 × 1053 elements. I saw a video by Richard Boucherds where he mentioned in passing that the number of elements in the group is a few orders of magnitude larger than the number of atoms in earth.

I tried to find a comparison that is closer to 8 × 1053 and settled on the number of atoms in Jupiter.

The mass of Jupiter is about 2 × 1027 kg. The planet is roughly 3/4 hydrogen and 1/4 helium by mass, and from that you can calculate that Jupiter contains about 1054 atoms.

Before doing my calculation with Jupiter, I first looked at lengths such as the diameter of the Milky Way in angstroms. But even the diameter of the observable universe in angstroms is far too small, only about 9 × 1036.

More on finite simple groups

[1] The way you build up groups from simple groups is more complicated than the way you build integers out of primes, but there’s still an analogy there.

The photo at the top of the post was taken by Voyager 1 on January 6, 1979.

Group symmetry of copula operations

You don’t often see references to group theory in a statistics book. Not that there aren’t symmetries in statistics that could be described in terms of groups, but this isn’t often pointed out.

Here’s an example from An Introduction to Copulas by Roger Nelsen.

Show that under composition the set of operations of forming the survival copula, the dual of a copula, and the co-copula of a given copula, along with the identity (i.e., ^, ~, *, and i) yields the dihedral group.

Nelsen gives the following multiplication table for copula operations.

    o | i ^ ~ *
    -----------
    i | i ^ ~ *
    ^ | ^ i * ~
    ~ | ~ * i ^
    * | * ~ ^ i

The rest of this post explains what a copula is and what the operations above are.

What is a copula?

At a high level, a copula is a mathematical device for modeling the dependence between random variables. Sklar’s theorem says you can express the joint distribution of a set of random variables in terms of their marginal distributions and a copula. If the distribution functions are continuous, the copula is unique.

The precise definition of a copula is technical. We’ll limit ourselves to copulas in two dimensions to make things a little simpler.

Let I be the unit interval [0, 1]. Then a (two-dimensional) copula is a function from I × I to I  that satisfies

\begin{align*} C(0, v) &= 0\\ C(u, 0) &= 0\\ C(u, 1) &= u\\ C(1, v) &= v \end{align*}

and is 2-increasing.

The idea of a 2-increasing function is that “gradients point northeast.” Specifically, for all points (x1, y1) and (x2, y2) with x1x2 and y1y2, we have

C(x_2, y_2) - C(x_2, y_1) - C(x_1, y_2) + C(x_1, y_1) \,\geq\, 0

The definition of copula makes no mention of probability, but the 2-increasing condition says that C acts like the joint CDF of two random variables.

Survival copula, dual copula, co-copula

For a given copula C, the corresponding survival copula, dual copula, and co-copula are defined by

\begin{align*} \hat{C}(u, v) &= u + v - 1 + C(1-u, 1-v) \\ \tilde{C}(u, v) &= u + v - C(u,v) \\ C^*(u,v) &= 1 - C(1-u, 1-v) \end{align*}

respectively.

The reason for the name “survival” has to do with a survival function, i.e. complementary CDF of a random variable. The survival copula is another copula, but the dual copula and co-copulas aren’t actually copulas.

This post hasn’t said much too about motivation or application—that would take a lot more than a short blog post—but it has included enough that you could verify that the operations do compose as advertised.

Update: See this post for more algebraic structure for copulas, a sort of convolution product.

Sharing secrets using polynomials

This post will present a couple ways to share secrets using polynomials. We have a group of n people who want to share a secret between them so that k of them will have to cooperate in order to unlock the secret. For example, maybe a committee of n = 5 wants to require the cooperation of at least k = 3 members.

Shamir’s method

Adi Shamir came up with the idea of using polynomials to share secrets as follows. First, encode the secret you want to share as an integer a0. Next, generate m = k-1 other random integers a1 through am and use these as coefficients of a polynomial f of degree m:

f(x) = a_0 + a_1x + a_2x^2 + \cdots + a_mx^m

A trusted party generates n random integers values of x and gives each person an x and the corresponding value of f(x). Since m+1 points completely determine a mth degree polynomial, if k = m+1 people share their data, they can recover f, and thus recover the secret number a0. This can be done efficiently, for example, by using Lagrange interpolation. But with fewer than k data points, the polynomial remains undetermined.

In practice we’d work over the integer modulo a large prime p. While fewer than k data points will not let someone completely determine the polynomial f, it will narrow down the possible coefficients if we’re working over the integers. Working modulo a large prime instead reveals less information.

Verifiable secret sharing

There’s a possible problem with Shamir’s method. Maybe the trusted party made a mistake. Or maybe the trusted party was dishonest and shouldn’t have been trusted. How can the parties verify that they have been given valid data without unlocking the secret? Seems we’re at a logical impasse since you’d have to recover the polynomial to know if your points are on the polynomial.

Paul Feldman came up with a way to assure the participants that the secret can be unlocked without giving them the information to unlock it. The trick is to give everyone data that in principle would let them determine the polynomial, but in practice would not.

We choose a large prime p such that p-1 has a large prime factor q [1]. Then the multiplicative group of non-zero integers mod p has a subgroup of order q. Let g be a generator of that group. The idea is to let everyone verify that

y_i = f(x_i)

for their given (xi, yi) by letting them verify that

g^{y_i} = g^{f(x_i)}

where all calculations are carried out mod p. Our trusted party does this by computing

A_i \equiv g^{a_i}\pmod{p}

for each coefficient ai and letting everyone know g and each of the Ai‘s.

In principle, anyone could solve for a0 if they know A0. But in practice, provided q is large enough, this would not be possible because doing so would require solving the discrete logarithm problem, which is computationally difficult. It’s possible to compute discrete logarithms for small q, but the difficulty goes up quickly as q gets larger.

How do the Ai‘s let everyone verify that their (xi, yi) data is correct?

Each person can verify that

g^{y_i} = \prod_{j=0}^m A_j^{x_i^j} 

using the public data and their personal data, and so they can verify that

g^{y_i} = \prod_{j=0}^m A_j^{x_i^j} = \prod_{j=0}^m g^{a_j x_i^j} = g^{f(x_i)} 

Related posts

[1] Conceptually you pick p‘s until you find one so that p-1 has a large prime factor q. In practice, you’d do it the other way around: search for large primes q until you find one such that, say, 2q + 1 is also prime.

Addition on Curve1174

I’ve written about elliptic curve and alluded to the fact that there’s a special kind of addition for points on the curve. But I haven’t gone into details because it’s more complicated than I wanted to get into.

However, there’s a special case where the details are not complicated, the so called Edwards curves. I’ll look briefly at Edwards curves in general, then focus on Curve1174, a particular Edwards curve used in cryptography.

The example here could be used in an introductory group theory course with no reference to elliptic curves. Just think of it as a funny way to add pairs of integers.

Addition on Edwards curves

For a particular class of elliptic curve, Edwards curves, the addition formula is simpler than usual. As mentioned a few days ago, an Edwards curve has the form

x² + y² = 1 + d x² y²

where d is not 0 or 1 in the underlying finite field. Then addition on the curve is given by

(x_1,y_1) + (x_2,y_2) = \left( \frac{x_1 y_2 + x_2 y_1}{1 + dx_1 x_2 y_1 y_2}, \frac{y_1 y_2 - x_1 x_2}{1 - dx_1 x_2 y_1 y_2} \right)

When d is a square, there are some exceptions. When d is not a square, as will be the case in our application, the denominators are never zero, and so the formula above is all there is to the addition rule.

Note that the division in the formula above is division in the underlying finite field, i.e. multiplication by the multiplicative inverse.

Curve1174

We’re interested in Curve1174, a particular elliptic curve used in cryptography. The underlying field is GF(p), the integers modulo the prime p = 2251 – 9. Also, d = -1174, from whence the curve takes its name.

Plotting Curve1174 over the reals

The plot above shows what Curve1174 looks like over the real numbers, though we’re interested in the curve over the integers mod p. (By the way, if d > 0 you get a curve that looks like a squircle.)

We consider the pairs of integers (x, y) that lie on the curve, i.e. those that satisfy

x² + y² = 1 + d x² y² mod p.

You can show that the sum of two points on the curve is another point on the curve, if you define addition with the formula above. The identity element for addition is the pair (0, 1). The additive inverse of a point (xy) is the point (-xy). So we have a group. Addition is commutative, and so in fact we have an Abelian group.

Python code

We can implement addition on Curve1174 in a few lines of Python.

from sympy import mod_inverse

def divide(a, b, p):
    "Compute a/b in GF(p)"
    return (a*mod_inverse(b, p))%p

def group_add(x1, y1, x2, y2, p, d):
    x3 = divide(x1*y2 + x2*y1, 1 + d*x1*x2*y1*y2, p)
    y3 = divide(y1*y2 - x1*x2, 1 - d*x1*x2*y1*y2, p)
    return (x3, y3)

The only thing we needed SymPy for was the mod_inverse function. It wouldn’t take much work to write your own mod_inverse function from scratch using the method outlined here using a variation on the Euclidean algorithm.

It’s clear that (1, 0) is a point on the curve, and so we can add it to itself with the code

p = 2**251 - 9
d = -1174
print(group_add(1, 0, 1, 0, p, d))

and find that it equals

(0, 3618502788666131106986593281521497120414687020801267626233049500247285301238),

which may come as a bit of a surprise. Arithmetic here is not intuitive; it scrambles up points well, which hints at why the curve is useful in cryptography.

Let’s find another point on the curve. Let’s set x = 2019 and see what y is. When we come up with the equation y must satisfy, the Jacobi symbol shows there is no solution.

When x = 2025 there is a solution, and we can compute it using sqrt_mod from sympy.ntheory.

x = 2025
k = divide(1 - x*x, 1 - d*x*x, p)
y = sqrt_mod(k, p)

This says the point

(2025, 588747530266665079407582947937120321357732884331117971504880828350684014295)

is on Curve1174. And since x and y only appear as squares in the equation defining the curve, once we find an (x, y) pair on the curve, the points (±x, ±y) are also on the curve.

Just for grins, let’s double the point (xy) above, i.e. add it to itself. This works out to

(2795920935947049934301363619759082573282734750667150130900931245990107374027,  2577351770662637935098262284237063829290291047539093190165388036658162531660).

Number of points on Curve1174

In general it can be hard to compute how many point lie on an elliptic curve, but in the case of Curve 1174 the number of points is known. Bernstein et al computed that the number of points on Curve1174 is p + 1 – t where t is a big number, but much smaller than p, on the order of the square root of p. Specifically,

t = 45330879683285730139092453152713398836.

Why not just absorb the 1 into t? This was done to match the notation in Hasse’s theorem. See the footnote here.

Elliptic curve cryptography (ECC)

What does all this have to do with cryptography? Cryptographers like to find problems that can be computed easily but that are hard to reverse. Most public key cryptography methods depend on the difficulty of undoing one of three things:

  • multiplication,
  • modular exponentiation, or
  • multiplication over an elliptic curve.

RSA encryption, for example, depends on the difficulty of factoring the product of two large primes.

The elliptic curve discrete logarithm problem (ECDLP) is the problem of undoing multiplication over an elliptic curve. If n is an integer and P is a point on the curve, we can compute QnP easily. If n is large, we don’t just add P to itself n times. Instead we double it log2n times and add the necessary intermediate results, analogous to fast exponentiation.

It’s easy to compute Q given n and P, but it’s hard to compute n given P and Q. This is the elliptic curve discrete logarithm problem that EEC protocols rely on for their security.

More elliptic curve cryptography posts

Groups of order 2019

How many groups have 2019 elements? What are these groups?

2019 is a semiprime, i.e. the product of two primes, 3 and 673. For every semiprime s, there are either one or two distinct groups of order s.

As explained here, if spq with pq, all groups of order s are isomorphic if q is not a factor of p-1. If q does divide p-1 then there are exactly two non-isomorphic groups of order s, one abelian and one non-abelian. In our case, 3 does divide 672, so there are two groups of order 2019. The first is easy: the cyclic group of order 2019. The second is more complex.

You could take the direct product of the cyclic groups of order 3 and 673, but that turns out to be isomorphic to the cyclic group of order 2019. But if you take the semidirect product you get the other group of order 2019, the non-abelian one.

Semidirect products

Starting with two groups G and H, the direct product G × H is the Cartesian product of G and H with multiplication defined componentwise. The semidirect product of G and H, written [1]

G \rtimes H

also starts with the Cartesian product of G and H but defines multiplication differently.

In our application, G will be the integers mod 673 with addition and H will be a three-element subgroup of the integers mod 673 with multiplication [2]. Let H be the set {1, 255, 417} with respect to multiplication [3]. Note that 1 is its own inverse and 255 and 417 are inverses of each other.

Product

Now the group product of (g1, h1) and (g2, h2) is defined to be

(g1 + h1-1g2, h1 h2)

So, for example, the product of (5, 255) and (334, 417) is (5 + 417*334, 255*417) which reduces to (645, 1) working mod 673.

(We haven’t defined the semidirect product in general, but the procedure above suffices to define the semidirect product for any two groups of prime order, and so it is sufficient to find all groups of semiprime order.)

Note that our group is non-abelian. For example, if we reverse the order of multiplication above we get (263, 1).

Identity

The identity element is just the pair consisting of the identity elements from G and H. In our case, this is (0, 1) because 0 is the additive identity and 1 is the multiplicative identity.

Inverse

The inverse of an element (gh) is given by

(-ghh-1).

So, for example, the inverse of (600, 255) is (444, 417).

Python code

The goal of this post is to be concrete rather than general.

So to make everything perfectly explicit, we’ll write a little Python code implementing the product and inverse.

    
    def hinv(h):
        if h == 255:
            return 417
        if h == 417:
            return 255
        if h == 1:
            return 1
    
    def prod(x, y):
        g1, h1 = x
        g2, h2 = y
        g3 = (g1 + hinv(h1)*g2) % 673
        h3 = (h1*h2) % 673
        return (g3, h3)
    
    def group_inv(x):
        g, h = x
        return ((-g*h)%673, hinv(h))
    
    x = (5, 255)
    y = (334, 417)
    print(prod(x, y))
    print(prod(y, x))
    print(group_inv((600, 255)))

The following code verifies that our group satisfies the axioms of a group.

    from itertools import product
    
    identity = (0, 1)
    h_list = [1, 255, 417]
    
    def elem(x):
        g, h = x
        g_ok = (0 <= g <= 672)
        h_ok = (h in h_list)
        return (g_ok and h_ok)
    
    group = product(range(673), h_list)
    assert(len([g for g in group]) == 2019)
    
    # closed under multiplication    
    for x in group:
        for y in group:
            assert( elem(prod(x, y)) )
    
    # multiplication is associative
    for x in group:
        for y in group:
            for z in group:
                xy_z = prod(prod(x, y),z)
                x_yz = prod(x, prod(y,z))
                assert(xy_z == x_yz)
    
    # identity acts like it's supposed to
    for x in group:
        assert( prod(x, identity) == x )
        assert( prod(identity, x) == x )
    
    # every element has an inverse
    for x in group:
        ginv = group_inv(x)
        assert( elem(ginv) )
        assert( prod(x, ginv) == identity )
        assert( prod(ginv, x) == identity )

Related posts

[1] The symbol for semidirect product is ⋊. It’s U+22CA in Unicode and \rtimes in LaTeX.

[2] In general, the semidirect product depends on a choice of an action of the group H on the group G. Here the action is multiplication by an element of H. Different actions can result in different groups. Sometimes the particular choice of action is made explicit as a subscript on the ⋊ symbol.

[3] How did I find these numbers? There are 672 non-zero numbers mod 673, so I picked a number, it happened to be 5, and raised it to the powers 672/3 and 2*672/3.

 

Groups of semiprime order

For each prime p, there is only one group with p elements, the cyclic group with that many elements. It would be plausible to think there is only one group of order n if and only if n is prime, but this isn’t the case.

If p and q are primes, then there are ostensibly at least two groups of order pq: the cyclic group Zpq, and Zp + Zq, the direct sum of the cyclic groups of orders p and q. However, there may just be one group of order pq after all because the two groups above could be isomorphic.

If pq = 2, then Z4 and Z2 + Z2 are not isomorphic. But the groups Z15 and Z3 + Z5 are isomorphic. That is, there is only one group of order 15, even though 15 is composite. This is the smallest such example.

Let p and q be primes with pq. If q does not divide p-1, then there is only one group of order pq. That is, all groups of order pq are isomorphic to the cyclic group Zpq. So when p = 5 and q = 3, there is only one group of order 15 because 3 does not evenly divide 5-1 = 4. The same reasoning shows, for example, that there must only be one group with 77 elements because 7 does not divide 10.

Now if q does divide p-1, then there are two distinct groups of order pq. One is the cyclic group with pq elements. But the other is non-Abelian, and so it cannot be Zp + Zq. So once again Zpq is isomorphic to Zp + Zq, but there’s a new possibility, a non-Abelian group.

Note that this does not contradict our earlier statement that Z4 and Z2 + Z2 are different groups, because we assumed p > q. If pq, then Zpq is not isomorphic to Zp + Zq.

Related posts

 

Number of groups of prime power order

John Baez left a comment on my post on group statistics saying

It’s known that the number of groups of order p^n for prime p is p^{2n^3/27+O(n^(8/3))}. It might be fun to compare this to what Mathematica says.

Here goes. First let’s let p = 2. Mathematica’s FiniteGroupCount function tops out at n = 10. And for the range of values Mathematica does support, 2^{2n^3/27} grossly overestimates the number of groups.

    Table[{FiniteGroupCount[2^n], 2^((2./27) n^3)}, {n, 1, 11}]

    {{1, 1.05269}, {2, 1.50795}, {5, 4.}, {14, 26.7365}, 
    {51, 612.794}, {267, 65536.}, {2328, 4.45033*10^7}, 
    {56092, 2.61121*10^11}, {10494213, 1.80144*10^16}, 
    {49487365422, 1.98847*10^22}, {FiniteGroupCount[2048], 4.7789*10^29}}

OEIS has entries for the sizes of various groups, and the entry for powers of 2 confirms the asymptotic formula above. Maybe it’s not a good approximation until n is very large. (Update: Here’s a reference for the asymptotic expression for all p.)

The OEIS entry for number of groups of order powers of 5 has an interesting result:

For a prime p >= 5, the number of groups of order p^n begins 1, 1, 2, 5, 15,
61 + 2*p + 2*gcd (p – 1, 3) + gcd (p – 1, 4),
3*p^2 + 39*p + 344 + 24*gcd(p – 1, 3) + 11*gcd(p – 1, 4) + 2*gcd(p – 1, 5), …

We can duplicate this with Mathematica.

    Table[FiniteGroupCount[5^n], {n, 0, 6}]

    {1, 1, 2, 5, 15, 77, 684}

and the last two numbers match the calculations given in OEIS.

There’s something interesting going on with Mathematica. It doesn’t seem to know, or agree with, the formula above for groups of order p4. For example,

    Table[FiniteGroupCount[7^n], {n, 0, 6}]

    {1, 1, 2, 5, FiniteGroupCount[2401], 83, 860}

I get similar results when I use larger primes: it can’t handle the fourth power.

    Table[FiniteGroupCount[389^n], {n, 0, 6}]

    {1, 1, 2, 5, FiniteGroupCount[22898045041], 845, 469548}

The results for n = 5 and 6 agree with OEIS.

Is OEIS wrong about the number of groups of order  p4 or should Mathematica simply return 15 but there’s a gap in the software?

Also, does anybody know why the agreement with the asymptotic formula above is so bad? It’s just as bad or worse for other primes that I tried.

More algebra posts