Some problems of Erdős on the sum-of-divisors function

By Paul Pollack and Carl Pomerance

For Richard Guy on his 99th birthday. May his sequence be unbounded.

Abstract

Let denote the sum of all of the positive divisors of , and let denote the sum of the proper divisors of . The functions and were favorite subjects of investigation by the late Paul Erdős. Here we revisit three themes from Erdős’s work on these functions. First, we improve the upper and lower bounds for the counting function of numbers with deficient but abundant, or vice versa. Second, we describe a heuristic argument suggesting the precise asymptotic density of not in the range of the function ; these are the so-called nonaliquot numbers. Finally, we prove new results on the distribution of friendly -sets, where a friendly -set is a collection of distinct integers which share the same value of .

1. Introduction

Let be the sum of the natural number divisors of , so that . Let be the sum of only the proper divisors of , so that . Interest in these functions dates back to the ancient Greeks, who classified numbers as deficient, perfect, or abundant according to whether , , or , respectively. For example, is deficient, is abundant, and is perfect. The ancient Greeks also found and remarked on 2-cycles, where and , calling such a pair amicable; for example, 220 and 284.

A conjecture of Catalan in 1888 Reference 8, as amended by Dickson in 1913 Reference 12, claims that for every positive integer , the aliquot sequence either terminates at 0 or enters some cycle, such as a perfect number, an amicable pair, or some longer cycle. Despite much numerical work, there are still 12 numbers for which the conjecture is in doubt, the first being 276. This numerical work uncovered long stretches where some aliquot sequences grow monotonically and exponentially, and other stretches where they decay in like manner. This behavior was taken into account in the Guy–Selfridge Reference 25 counter-conjecture that for asymptotically all even numbers , the aliquot sequence starting with is unbounded.

It was proved by Lenstra Reference 29 that for each there is an aliquot sequence that strictly increases for terms. Erdős Reference 20 showed that this is common and, in fact, if is abundant, then but for a set of numbers of asymptotic density 0, the aliquot sequence starting with strictly increases for terms. In the same paper, Erdős claimed the analogous result for deficient numbers (with “increases” replaced with “decreases”), but this claim was retracted in Reference 22. In that paper, the claim was shown for . It is conjectured that the full claim holds, but this has never been shown.

These results assert that if is abundant, it is likely that is also abundant, and similarly for deficient. Say is an aliquot reversal if this fails. More precisely, say is an up-down reversal if is nondeficient yet is deficient, and say is a down-up reversal if is deficient and is nondeficient.

Though the terminology is new to this paper, aliquot reversals have been studied for some time because of their connection with amicable pairs; in fact the lesser member of an amicable pair is an up-down reversal, while the larger member is a down-up reversal. Erdős was the first to prove, in Reference 15, that the set of numbers involved in an amicable pair has vanishing asymptotic density, and he did this by showing that there are few up-down reversals. The count of up-down reversals was improved later by Rieger Reference 40, by Erdős–Rieger Reference 21, and by the second author (see Reference 36). In that last paper it was shown that the number of giving an up-down reversal is at most

for a certain absolute constant . Here, the subscripts indicate iteration of the natural log function. The proof uses results of Erdős Reference 13 for the count of primitive nondeficient numbers. Incorporating an improvement on these results due to Avidon Reference 3, this argument gives . In Reference 33, Corollary 1.5, the same upper estimate Equation 1.1 was shown for the number of down-up reversals in , but only with the smaller constant .

Our first theorem is an improvement on these upper bounds.

Theorem 1.1.

Both the count of down-up reversals in and the count of up-down reversals in are bounded above by Equation 1.1 with .

The proof uses recent results from Reference 2 on the number of solutions to congruences of the form .

The problem of obtaining lower bounds on the number of aliquot reversals does not seem to have been previously considered. Note that the upper bounds of the last paragraph just barely give density 0. In the following results we show there is a reason for this: reversals are fairly common. We prove the following.

Theorem 1.2.

The number of up-down reversals in is .

Theorem 1.3.

The number of down-up reversals in is .

One might wonder if these results lend support to the Catalan–Dickson conjecture or to the Guy–Selfridge counter-conjecture. It is helpful to think in terms of parity. In fact, and have opposite parity precisely when is a square or a double of a square, and such comprise a very sparse set. In addition, it follows from the proof in Reference 20 that for each fixed , almost all have the same parity as (the th iterate of at ). Since an aliquot sequence terminates only if it hits a prime, a perhaps reasonable conjecture is that aliquot sequences beginning with even numbers usually either are unbounded or enter some cycle of even numbers. Numbers involved in such cycles have density 0 (see Reference 28). Numerical evidence indicates that most numbers involved in these cycles are amicable (i.e., they are in a 2-cycle), and we know that amicable numbers are quite sparsely distributed (see Reference 38). Thus, despite the fact that our results show nonmonotonicity is fairly common, the Guy–Selfridge point of view is believable. And some cases of nonmonotonicity are quite rare. For example, take numbers , . They are abundant, but the number of them to with deficient is . Further, in a recent preprint of Bosma Reference 6, it is shown that about of the even numbers to 1 million have their aliquot sequence still open beyond . This evidence seems to put a chill on the Catalan–Dickson conjecture! On the other hand, it is hard to align this numerical evidence with the recent paper of Bosma–Kane Reference 7 where it is shown that the geometric mean of for tends to a constant smaller than 1 as .

We remark that the lower bounds in Theorems 1.2, 1.3 are established for even numbers. It would be of interest to see what kind of lower bounds can be established for odd numbers.

We turn now to our second topic. Positive integers not in the range of are said to be nonaliquot, though the more colorful term untouchable (see Alanen Reference 1) has also been used. The concept dates back to at least ca. 1000 CE Reference 41. A useful initial observation is that if and are distinct primes, then . A slightly stronger form of Goldbach’s conjecture is that every even integer is a sum of two distinct primes. If so, then every odd belongs to the range of ; noting that , , and , we would have that is the only odd nonaliquot number.

While the conjecture of the last paragraph remains out of reach, it follows from (independent) work of Chudakov, van der Corput, and Estermann that almost all even numbers can be written as a sum of two distinct primes. (See Reference 46, Chapter 3 for a modern treatment.) Hence, the set of odd nonaliquot numbers has asymptotic density zero.

What about even numbers? In Reference 18, Erdős proved that a positive proportion of even numbers are nonaliquot. By an elaboration of Erdős’s methods, Chen and Zhao Reference 9 have shown that the nonaliquot even integers make up a set of lower density greater than . (This improves earlier results reported in Reference 5 and Reference 44, Corollary 9.2, p. 64.) Erdős was unable to decide whether or not a positive proportion of even numbers do belong to the image of , but this has been resolved affirmatively in very recent work of Luca and the second author Reference 30.

In §3, refining some thoughts in Reference 44, we propose a probabilistic model for that suggests a precise value for the density of nonaliquot numbers.

Conjecture 1.4.

The set of nonaliquot numbers has asymptotic density , where

At , the expression inside the limit has value . Table 1 collects counts of nonaliquot numbers up to ; these counts are consistent with a limiting density of .

Finally, we consider the distribution of the rational numbers . We are interested here in statistics for

thought of as a random variable on the set of natural numbers , equipped with the uniform measure. A theorem of Wirsing Reference 47 asserts that an equation of the form has at most solutions , as , uniformly in . Thus, we have a pointwise bound , as . The average behavior of was determined by Erdős Reference 17, Theorem 2; for a certain constant ,

Note that counts the number of ordered pairs of distinct integers with , and this is the way Erdős states his result.

Call a pair (unordered) of distinct integers and a friendly pair if . Evidently, if is coprime to , then is also a friendly pair. Call the friendly pair primitive friendly if and have no nontrivial common unitary divisor. (We say is a unitary divisor of , and write , if and .) The key ingredient in Erdős’s proof of Equation 1.2 is the convergence of , where runs over the larger members of the primitive friendly pairs (with multiplicities). We prove a substantially stronger result on the distribution of such pairs:

Theorem 1.5.

The number of primitive friendly pairs contained in is at most , as .

Erdős’s theorem on the convergence of follows immediately by partial summation. The exponent appears to be the limit of our method, but the numerical data (see Table 2) suggest that this is not sharp. Perhaps the true count of primitive friendly pairs in is .

By studying primitive friendly -sets for integers , we are able to show that has a limiting distribution. Set .

Theorem 1.6.

For each fixed nonnegative integer , we have , as , for some constant . Moreover, as .

Clearly, . Though the constants for are effectively computable (via the proof of Theorem 1.6), we have not found good rigorous numerical estimates for them. However, explicit computations of suggest that , , , and ; see Table 3.

Notation.

We use the Bachmann–Landau , and -notations, with their usual meanings, as well as the associated Vinogradov and notations. denotes the largest prime factor of , with the convention that , and denotes the largest squarefree divisor of . We reserve the letter , with or without subscripts or dashes, for primes. We write for the exponent on the highest power of dividing . The notation stands for the greatest common unitary divisor of and . We use for Euler’s totient function, for the number of prime divisors, for the number of prime factors counted with multiplicity, and for the total number of divisors.

2. Aliquot reversals

2.1. Upper bounds

2.1.1. A structural lemma on primitive nondeficient numbers

The proof of Theorem 1.1 runs along similar lines as the proof of the main theorem of Reference 36. In particular, the primitive nondeficient numbers play a critical role. (These are numbers with and for each proper divisor of , .) The main new ingredient is a lemma that is perhaps of independent interest: Almost all primitive nondeficient numbers have .

We will deduce the lemma from the results of Reference 2 concerning solutions to the congruence . In that paper, is called a regular solution if

(It is straightforward to check that these numbers are solutions of the congruence.) In the remaining case, is called sporadic. Note that regular solutions can exist only when . The following is a slightly less precise form of Reference 2, Theorem 1.

Proposition 2.1.

The number of sporadic solutions to is at most , as , uniformly for .

The restriction is inconvenient for our purposes. We can ameliorate this as follows. Call a sporadic solution to convenient if all of the following hold:

(i)

is an integer at least ,

(ii)

if is a unitary divisor of with , then ,

(iii)

there is a prime with .

Proposition 2.2.

The number of convenient solutions to is at most , as , uniformly in with .

Sketch of the proof.

Hornfeck and Wirsing Reference 27, Satz 2 have shown that there are multiply perfect numbers ; hence, we may assume that .

While not phrased this way in Reference 2, the proof of Reference 2, Theorem 1 when is handled there by reduction to the convenient case. Indeed, Lemma 5 in Reference 2 asserts that if is a sporadic solution to and , then is convenient. In our setup here, we start by assuming is convenient. Thus, the conclusion of Reference 2, Lemma 5 holds trivially. Moreover, only the convenience hypothesis is needed in the proofs of Lemmas 6 and 7 from Reference 2. That is, both of those lemmas hold with the assumption of convenience replacing the hypothesis is a sporadic solution satisfying ”.

With these lemmas in place, we can again run the proof of Reference 2, Theorem 1. It proceeds exactly as before, except that now there is no need to assume . That assumption was responsible for the restriction , which can now also be dispensed with. Following the proof through, we deduce that there are at most convenient solutions , uniformly for .

We now deduce the key lemma.

Lemma 2.3.

Fix . The number of primitive nondeficient with is , as .

Proof.

After a dyadic subdivision argument, it is enough to prove the claimed upper bound for the number of primitive nondeficient with . By Reference 27 or Reference 47, the number of perfect is , as . Thus, we can assume that is abundant. We can also assume , since otherwise belongs to a set of size at most . Write , where . Since is primitive nondeficient,

Hence,

and

The integer is a sporadic solution to . Moreover, conditions (i)–(iii) in the definition of a convenient solution are easily verified for , with and .

Fixing and , Proposition 2.2 shows that the number of corresponding values of is . Summing on , the number of possibilities for given is

Summing on gives a final upper bound of , as desired.

Remark.

The proof can be improved to achieve the upper bound , as . For our purposes, any power-savings upper bound is sufficient.

It would be desirable to prove Lemma 2.3 with a smaller exponent than . If the lemma holds with (say), then the arguments below show that Theorem 1.1 holds with . In particular, if can be taken arbitrarily large, we would have a qualitative improvement over any bound of the shape Equation 1.1. Such an improvement would follow from a conjecture posed in Reference 2:

Conjecture 2.4.

Let , and let be an integer with . Let be an integer with . The number of solutions to is

where both the implied constant and are absolute constants.

Remark.

This is a corrected form of the conjecture as it appeared in Reference 2. The original formulation allowed and ; these cases must be excluded to account for popular values of or .

2.1.2. Proof of Theorem 1.1

Let

We begin with a lemma.

Lemma 2.5.

The number of integers for which is not divisible by every integer is .

Proof.

We use Reference 36, Theorem 2, which asserts that for coprime integers with , the number of integers such that no prime has is , uniformly. Applied with the result implies that the number of counted in the lemma is

For future reference we state the result of Avidon Reference 3 mentioned in the introduction.

Lemma 2.6.

The number of primitive nondeficient numbers is at most

We shall also use a result of Toulmonde Reference 45, Théorème 1 (see the discussion in §10) on the modulus of continuity of the distribution function for .

Lemma 2.7.

If is in the range of , then uniformly for , the number of integers with

is .

Proof of Theorem 1.1.

Let be fixed. Suppose is nondeficient with deficient, and let be the least nondeficient number that divides . If each prime power dividing is at most , then Lemma 2.5 implies that but for choices for , we may assume that . This in turn implies that , contradicting that is deficient. Thus, has a prime power divisor exceeding . If , then Lemma 2.6 implies that the number of divisible by such a primitive nondeficient number is at most as . Hence we may assume that and is divisible by a prime power . If is a proper power, the number of divisible by such a large power is , so we may assume that is prime and . But then, by Lemma 2.3, the number of divisible by such a primitive nondeficient is . Since may be taken arbitrarily close to 0, our result follows.

The argument for down-up reversals is similar, but more difficult. Suppose is deficient and is not deficient, and let be the least nondeficient number dividing . Again, by Lemma 2.5, if each prime power dividing is at most , then we may assume that . This implies that , contradicting deficient. Thus, we may assume has some prime power divisor greater than .

Write where . By standard results on the distribution of integers all of whose prime factors are small (see Reference 10), we may assume that and . We have . Let . The congruence implies that , so that . The congruence also implies that is in a residue class modulo .

We first assume that

Given , by the Brun–Titchmarsh inequality, the number of choices for is

where for the last step we use that is primitive nondeficient, so that . We have , so summing the displayed upper bound over and then over , we get that the number of choices is

We can restrict to which have no proper prime power divisor in . Indeed, by an argument analogous to that just seen, the number of as above for which is divisible by a given prime power is . Summing over proper prime powers , we find that only a negligible set of arise in this manner. Since , we may assume that is divisible by a prime .

We now wish to find a good upper bound for the sum of as runs over primitive nondeficient numbers divisible by a prime greater than . We first assume that . Then . Using as and Lemma 2.3 we have , which is sufficient for our needs. So now assume that . If also , a brief calculation with Lemma 2.6 suffices.

Now assume that and . We forget at this point that is running over primitive nondeficient integers. For each fixed positive integer ,

Let and consider values of with

We have

Thus, letting range over numbers of the form , for ,

If we choose , and let be arbitrarily small, this estimate completes the argument in the case .

Now we assume . Since and , we have . Consider the -prefix” divisor of . This is the largest divisor of in with each prime dividing bounded above by each prime dividing . We may assume that

Indeed, if , then every prime factor of is greater than , so that

since has fewer than distinct prime factors. Also, since is deficient, we have

The last two displays imply that for sufficiently large, we have deficient, a contradiction. Thus, we have Equation 2.2.

Let denote the set of integers which are the -prefix for some primitive nondeficient . We will show that there are very few integers with divisible by some member of . To do this, we show that the set is rather sparse.

Let , and assume that and . Then, as above,

so for to be nondeficient, there is some absolute constant such that

By Lemma 2.7, for , the number of integers which satisfy the last display is at most for all sufficiently large . (Here could be replaced with any number exceeding .)

We also wish to count the number of , with . For this we ignore the condition and use only a standard result in Reference 10: the number of such is at most for all sufficiently large .

As in the case , we find that the number of integers with for some (and assuming with and ) is

Again as before we treat the cases of small and large separately. If , then our estimates for the number of suffice to give an acceptable estimate. If , we use Equation 2.1 with , getting

This too is sufficiently small, so completing our proof.

2.2. Lower bounds

Here we prove Theorems 1.2, 1.3.

Proof of Theorem 1.2.

Let be large, let be the least integer with , and let be a prime with . We shall consider numbers where and is prime with . Since is abundant, such numbers are abundant. With no further restrictions on the parameters , the number of ’s created in this way is . We will show that with some mild restrictions on we can be assured that is deficient, and these restrictions will only reduce our count by a factor of .

We shall want the numbers to satisfy:

(i)

,

(ii)

,

(iii)

has no prime factors below .

We can ensure satisfies (ii) and (iii) by eliminating those divisible by any of

a proper prime power ,

a prime ,

a prime .

Let . By Brun’s sieve Reference 26, Theorem 2.2, there is an absolute constant such that the number of divisible by no prime in the second or third bullet points is

(The sum on here is , by the Brun–Titchmarsh inequality and the bound .) The number of divisible by some prime with is , which is , as . Finally, the number of divisible by some proper prime power is , and this is also . Thus, there are values of satisfying (ii) and (iii). In fact, this is the correct order of magnitude for this count, since (iii) by itself already implies an upper bound of the same form.

We now turn to (i). We apply the Turán–Kubilius inequality Reference 43, Theorem 1, p. 302 to investigate the normal order of the additive function . Uniformly for ,

and

Hence, the number of positive integers where is . Since , this is also an upper bound on the number of failures of (i).

We conclude that there are values of satisfying all of (i)–(iii). By partial summation,

where the sum is over satisfying (i)–(iii).

For each , let , where runs over primes. Using that is free of prime factors up to gives

which is . In what follows, we will assume that

Note that discarding contrary values of changes only by and so preserves the validity of Equation 2.3.

We now complete the construction of a number by choosing a prime . Suppose an odd prime divides . Then

If , then , so that . Such a prime cannot be 2 by assumption, and it cannot be , since and . Thus, , and so . We conclude that if and , then . We now restrict the ’s that are available to us by removing for each odd prime a single coprime residue class that will ensure that . By another application of Brun’s sieve, this cuts down the number of choices for by a factor proportional to . For a surviving prime , let

If and , then as we have seen, is in a fixed residue class mod . Using that also avoids one residue class modulo each prime not exceeding , the Brun–Titchmarsh inequality implies that the average value of , over surviving primes , is . Keeping Equation 2.4 in mind, we conclude that there are values of for which

We finally note that for chosen as above, we have , so that . Hence, has only odd prime factors. From Equation 2.5, we conclude that is deficient. The theorem then follows with the estimate:

where we have used Equation 2.3.

Proof of Theorem 1.3.

We take as the least integer with . We consider numbers with primes, , , and . As in the last proof, we impose some conditions on . Namely, we insist that satisfy all of

(i)

has no prime factors below ,

(ii)

,

(iii)

Equation 2.4 holds,

(iv)

for , we have .

We will show we can construct a collection of such that satisfies

It will turn out that the most stringent restriction is the first one.

By an elementary sieve, those satisfying (i) have

as . By arguments seen above, restricting further to satisfy (ii) changes the value of by , and restricting to satisfy (iii) affects the sum only by . Hence, Equation 2.7 still holds if is restricted to satisfy all of (i)–(iii).

We now turn to (iv). Changing by , we can assume that is not divisible by any proper prime power exceeding . If and , then there is a prime power with ; since , it follows that . Let

Using that and that is free of prime factors up to , we see that as ,

From Reference 36, Theorem 1 and Remark 1, the final sum on is at most , as , uniformly for . Inserting this estimate above,

Hence, the contribution to the left-hand side of Equation 2.7 from those values of with is at most . Since , after discarding these we may still assert Equation 2.6. Moreover, since and , we have

for all remaining (and all large ).

Note that for any choice of distinct primes greater than , the number is deficient; indeed, for large ,

and this is . We shall show that with some mild restrictions on , we will have abundant. For satisfying the above restrictions, let be the set of primes with and let be the product of the members of .

We next consider primes . They are chosen with the requirements that both and are coprime to . We have . By our choices of , we have coprime to . So for a prime , is in a particular residue class mod . Similarly this holds too for . Thus, the density of the primes which avoid these classes is at least

So, almost all of the primes may be used, and for these primes , we have and .

We now choose primes with . For such a prime , is abundant. Let . Note that for any prime , we have . Further, we have and . By the prime number theorem for progressions, the fraction of with is asymptotically . Similarly, if , the fraction of choices for with is asymptotically . Since

we have that the proportion of primes with is .

Putting these thoughts together, we have that the number of choices for is

Since , this completes the proof of the theorem.

3. The density of nonaliquot numbers

3.1. A heuristic argument

For a positive integer , let

For a positive integer , let

and let denote the set of members of . We assume that maps to for each . This is asymptotically true if and . Further, we assume that for we have . This is asymptotically true up to sets of vanishing density as . We assume in addition that is a random map subject to the prior assumptions.

Now suppose is even. If , then and , so that . The probability that an integer is not in the range of is about

Now , so the above probability is asymptotically

We thus would have that the density of even numbers not in the range of is

where the indicates that runs over even numbers. Since almost all odd numbers are in the image of , is also the density of nonaliquot numbers.

We now prove that the limit Equation 3.1 exists. Let

Note that . The difference comes from those with and . Such a number is divisible by for some prime power and . Writing , we have

Since, by Mertens’ theorem, as , we have

Thus the limit in Equation 3.1 exists if and only if exists, in which case this latter limit is also .

For a prime , we have and, for ,

Thus,

Since is always at most , it follows that the limit in Equation 3.1 exists and

We would like to compute a numerical approximation to , and it may seem that the first expression in Equation 3.3 is more easily computed than the second, since it involves a finite sum while the latter involves an infinite sum. However, we claim the infinite sum can be replaced with the sum over , while changing to 1. That is,

The equation Equation 3.4 may be proved as follows: Introduce a parameter , which will eventually be chosen as a slow-growing function of . Define . For even ,

Thus,

A nearly identical computation gives

Note that , and that when is even, is also even. Hence, for all large we have

Again almost identically, we have

By Mertens’ theorem (with the elementary error estimate shown in Reference 43, Theorem 10, p. 17),

Making this substitution, we find that

By choosing , say, from Equation 3.2, Equation 3.5, Equation 3.6, and Equation 3.7, we see that

Thus, Equation 3.4 follows. This completes the heuristic justification of Conjecture 1.4.

3.2. Computations

The following strategy was used to construct Table 1. Explicit computations on Goldbach’s conjecture have shown that every even number in is a sum of two distinct primes and (see Reference 31). Hence, is the only odd nonaliquot number up to . In Reference 39, §5, an algorithm is described that enumerates the elements of in time . We ran the algorithm to . We made one modification, necessitated by memory constraints: Instead of using a 0-1 array to record which numbers had appeared so far, the even numbers in the range of were written to disk as they were found by the algorithm, and sorted later with the GNU sort utility. The final data file required 36.1 GB.

On the computation of an approximation for , this is straightforward using Equation 3.4 for large values of . Note that is effectively computable, since the above calculations show that for each fixed integral value of ,

is within of . We have not used this approach, but we record some intermediate values for the expression in Equation 3.4 at and for :

3.3. A analogue

Let . As with , partial results on Goldbach’s conjecture imply that the range of contains all odd numbers but for a set of asymptotic density 0. Little is known about even values. In Reference 19 Erdős asks if a positive proportion of even numbers are of the form and if a positive proportion of even numbers are not of the form . The proof in Reference 30 does give a positive proportion of evens that are of the form , but the even numbers not of this form remain an enigma. However, as with the problem is amenable to calculation and heuristics. Let denote the number of integers in not of the form . By an analogous algorithm from Reference 39, we have extended the calculations of that paper 100-fold; see Table 4. The data file had 42.1 GB. (This file is larger than the one with nonaliquots since there are more even numbers with than there are with .)

The conjectural density is

At this expression rounds to . It is difficult to say if the actual counts are supportive of the conjectural density ; the evidence is somewhat less compelling than with the nonaliquots. At least the count densities seem to have peaked and are now slowly descending, perhaps to the limit . With both the nonaliquots and the analogues, the heuristic is very much an “at infinity” type of argument, while the counts are definitely very finite. It might be of some interest to try to do a statistical sample of random numbers that are much larger than our exhaustive counts, but it is not clear how one would decide if a given very large number is of the required form.

Here are the intermediate calculations of the expression in Equation 3.8 for various values of :

4. Primitive friendly pairs

We begin by collecting some results needed for the proof of Theorem 1.5. The following lemma says that we have a strong bound on the number of solutions to , even if we specify only the denominator of . For the proof of this result, in a more precise form, see Reference 32, Theorem 4.1.

Lemma 4.1.

The number of for which can be written as a fraction with a given denominator is at most , as , uniformly in .

The next lemma is implicit in the proof of Reference 23, Theorem 11; for an explicit statement and proof, see Reference 32, Lemma 4.2.

Lemma 4.2.

Let be an integer in . The number of integers for which is , as , uniformly in .

The next result, which was inspired by Erdős’s method of proof in Reference 16, appears as Reference 35, Lemma 3.1.

Lemma 4.3.

Given a natural number , the following algorithm outputs a unitary divisor of with . Moreover, at most inputs correspond to the same output , as .

Algorithm A.
\begin{algorithm} \DontPrintSemicolon\KwIn{A natural number $m$} \KwOut{A divisor $a$ of $m$ for which $a \parallel m$ and $\gcd(a,\sigma(a))=1$} Factor $m=p_1^{e_1} p_2^{e_2}\cdots p_k^{e_k}$, where $p_1 > p_2 > \dots> p_k$.\; $a \leftarrow1$\tcp*{Initialize} \For{$i=1$ \emph{\KwTo}~$k$} {\If{$\gcd(\sigma(p_i^{e_i}a), p_i^{e_i} a)=1$}{$a\leftarrow p_i^{e_i} a$}} \KwRet$a$\; \end{algorithm}
Proof of Theorem 1.5.

Suppose that and form a primitive friendly pair contained in . Write , where is squarefree, is squarefull, and . Define and by the equation , where is the output of Algorithm A when . Note that and are coprime squarefull integers, since is a unitary divisor of . Write

and let . Define analogously, but now with reference to instead of . Since , it follows that . With , choose for which

Define analogously. Finally, choose with

So, running over all of the possibilities for the , the , , , and , we cover all possibilities for . Thus, it is enough to prove that the number of pairs corresponding to a given choice of these variables is bounded by , as . We will establish four upper bounds on the number of these pairs . The theorem will follow by combining these upper bounds.

Our first upper bound uses that

To prove this, write , where is squarefree, is squarefull, and . Since divides and , and , so that . Now take a prime . Since and have no common unitary divisor, cannot exactly divide both and . So either or . Hence, and Equation 4.1 holds. By Lemma 4.3, given there are only possibilities for . Similarly, determines in at most ways. Thus, and determine the product up to possibilities. Now Equation 4.1, along with the maximal order of the divisor function, shows that and determine up to possibilities. Invoking Lemma 4.1, determine the pair in at most ways. Since the number of squarefull numbers in is , the number of pairs is at most

as .

For a second upper bound, expanding the equation , we find that . Since ,

By Lemma 4.2, the number of possibilities for is at most

By Wirsing’s theorem (or Lemma 4.1), this is also a bound on the number of possibilities for the pair . A symmetric argument gives our third upper bound,

To prove the fourth upper bound, note that since , we have . Consequently, the number of possibilities for is

By Lemma 4.1, the number of possibilities for the pair is at most

Taking the geometric mean of the four upper bounds Equation 4.2, Equation 4.3, Equation 4.4, and Equation 4.5, the number of possibilities for the pair does not exceed

Since , we have ; similarly, . From Equation 4.1, . Substituting these bounds for and into Equation 4.6 gives a final upper bound of , as .

4.1. Computations

To compile Table 2, for each , we found all solutions with , recording only those solutions with . To find the values of , we used the gp script solveBA of Michel Marcus (see Reference 42), which was based on an earlier program by Jan Munch Pedersen. The script solveBA uses a recursive algorithm to find all solutions up to a given limit satisfying an equation of the form . The values of were computed using Algorithm B3 below.

Algorithm B1.
\begin{algorithm} \DontPrintSemicolon\KwIn{A natural number $n$ and a divisor $d$ of $n$} \KwOut{The smallest $D\parallel n$ with $d\mid D$} $D \leftarrow d$.\\ Iterate $D \mapsto D\cdot(D,n/D)$ until $D$ stabilizes.\\ \KwRet$D$\; \end{algorithm}
Algorithm B2.
\begin{algorithm} \DontPrintSemicolon\KwIn{A natural number $n$ and a divisor $d$ of $n$} \KwOut{The largest $D\parallel n$ with $D\mid d$} $D \leftarrow d$.\\ Iterate $D \mapsto D/(D,n/D)$ until $D$ stabilizes.\\ \KwRet$D$\; \end{algorithm}
Algorithm B3.
\begin{algorithm} \DontPrintSemicolon\KwIn{Natural numbers $m$ and $n$} \KwOut{$\gcud(m,n)$} $d \leftarrow\gcd(m,n)$ \\ $d_1 \leftarrow$ output of $B1$ with input $d,m$\\ $d_2 \leftarrow$ output of $B1$ with input $d,n$\\ $d_3 \leftarrow d_1d_2/d$ \\ \KwRet the output of B2 with the pair $d,d_3$ \end{algorithm}

5. The distribution function for

Define a friendly -set as a set of distinct integers all of which share the same value of , and call such a set primitive if there is no that is a unitary divisor of each element. The proof of Theorem 1.6 will use the following crude upper bound on the count of primitive friendly -sets.

Lemma 5.1.

Fix . The number of primitive friendly -sets contained in is at most , as .

When , Lemma 5.1 gives an upper bound of , which is inferior to the result of Theorem 1.5. However, the proof of Lemma 5.1 is substantially simpler.

Proof.

List the elements of the set as . Write the common value of as in lowest terms. If , then there are at most possible values of , by Lemma 4.2, and so also at most possibilities for the -set, by Lemma 4.1. So we suppose that .

Let be a prime dividing . Then divides each . Since is primitive friendly, it is impossible for to exactly divide each . Thus, if , then . Hence, the number of possibilities for the integer is at most

From the maximal order of the -fold divisor function, determines the set in at most ways. The result follows.

Proof of Theorem 1.6.

We may assume , since when the asymptotic claimed in the theorem holds with . We begin by placing each primitive -set in increasing order, . We then list the primitive -sets in increasing order of , breaking ties arbitrarily. Say the th tuple is labeled as . Put

and observe that a natural number has at least friends in if and only if belongs to one of the sets . For each , put

The existence of the limit follows from inclusion–exclusion. (Each finite intersection of the sets can be described as the set of numbers not exceeding a certain constant multiple of that satisfy certain congruence conditions.) The are weakly increasing in and bounded by ; thus, it makes sense to set

Clearly, . We now show that .

If has at least friends in , but , then for some . Since , the number of these is at most

We now study the sum on . Without the restriction to , the sum converges (or is a finite sum). Indeed, from Lemma 5.1, the contribution to the unrestricted sum from those with is bounded by , as ; the desired result now follows upon summing along the powers of . So with the restriction to , the series decays to as . Letting and then , in that order, we get that .

It remains to prove that . Since , Erdős’s estimate Equation 1.2 shows that .

Remark.
(i)

A similar argument to the one just seen will show that each positive integral moment tends to a finite limit as . Here one uses that can be viewed as the count of (ordered) friendly -tuples in having final element and all other elements distinct from . As a consequence, the tend to zero more rapidly than any power of .

(ii)

The computations of , , reported in Table 3 were made with the aforementioned solveBA script.

We do not know how to show that each is nonzero. This is equivalent to the assertion that the count of natural numbers with is unbounded in , a conjecture mentioned already in Reference 19. Note that this conjecture would follow from the infinitude of perfect numbers. Since the may eventually vanish, we cannot show that the are strictly decreasing. However, we can prove that the sequence is strictly decreasing until it hits zero.

Theorem 5.2.

If , then .

Proof.

Since , there are examples of friendly -sets. (When , we consider every singleton subset of the natural numbers to be a friendly -set.) Fix such a set once and for all, chosen to have as small as possible. This set is automatically primitive. Every integer of the form , with and , has at least friends in . It suffices to show that values of give rise to an with exactly friends in .

Sort all of the primitive friendly -sets in increasing order, and then list the sets in order of their largest entry. Say the th set consists of . With a large, fixed integer to be chosen later, we restrict attention to satisfying

Let be the modulus of this congruence. For large , there are values of satisfying Equation 5.1. (The implied constant also depends on and the set , but since these parameters are fixed, we ignore this in our notation.) We now show that for most of these , the integer has exactly friends in .

If has at least friends in , then for some ,

In particular, . Suppose to begin with that . Then from Equation 5.1, and so . Putting ,

Since are in increasing order and make up a friendly -set, we have , by the initial choice of . Hence, , and so this last upper bound on contradicts Equation 5.1. So we must have . From and , we see that . In particular, . Using and Equation 5.1, we see that belongs to a fixed residue class modulo . The number of these is at most

Hence, the number of with satisfying Equation 5.1 that have at least friends in is

Since the sum on is only over , and since we can take arbitrarily large, the main term here can be made smaller than any constant multiple of , by the same argument used in the proof of Theorem 1.6. The error term is bounded by a power of less than , by Lemma 5.1. So fixing large enough, we have that for large , there are values of satisfying Equation 5.1 for which has exactly friends in , as desired.

6. Some final remarks

In the paper Reference 17, which inspired our work above on friendly pairs, Erdős writes “I would like to call attention to three simple problems which as far as I know are still unsolved”. The problems are:

Are there infinitely many solutions to the equation ?

For each real number , is there an infinite sequence of pairs with and as ?

If is the number of coprime pairs with and , do we have as ?

We note that the first two problems have been recently solved in the affirmative (see Reference 24, Reference 34), and the third was “almost” solved in the affirmative (solved without the coprimality condition) in 1980 (see Reference 37). It seems fitting to close this paper with a complete proof of the third problem.

Following an idea of Erdős Reference 14, say is such that there are primes with . Let be large, let , let be the set of primes with , let , and let . Consider integers which are the product of distinct primes in . Since , as , there are of these integers , each with . But, as was shown by Erdős (see Reference 11), the number of integers to not divisible by any prime is . Hence, there is one value such that has solutions , as . It follows that there are pairs with .

This would solve the third problem above if we had and if the pairs created were coprime. We know from Reference 37 that , the current record being a number slightly larger than ; see Reference 4. Let be fixed but arbitrarily small, and let . The number of pairs which reduce to a coprime pair with is at most

Summing this for gives . Thus, almost all of our pairs reduce to a pair where . In fact, there is some where pairs reduce to coprime pairs with . Each of these pairs comes from at most pairs , so there must be at least coprime pairs with . If is sufficiently small, we see that for all sufficiently large values of .

Figures

Table 1.

Counts of nonaliquots to various heights from through , along with the corresponding densities .

100000000 16246940 0.1625 1000000000 165826606 0.1658
200000000 32721193 0.1636 2000000000 333261274 0.1666
300000000 49265355 0.1642 3000000000 501171681 0.1671
400000000 65855060 0.1646 4000000000 669372486 0.1673
500000000 82468000 0.1649 5000000000 837801755 0.1676
600000000 99107582 0.1652 6000000000 1006383348 0.1677
700000000 115764316 0.1654 7000000000 1175094232 0.1679
800000000 132438792 0.1655 8000000000 1343935989 0.1680
900000000 149128373 0.1657 9000000000 1512867678 0.1681
10000000000 1681871718 0.1682
Table 2.

Counts of primitive friendly pairs. Here “Limit” is with the “Count” being the number of pairs to and “Exponent” being that such that Count = .

Limit 3 4 5 6 7 8 9 10 11
Count 9 19 38 66 148 292 548 966 1688
Exponent .318 .320 .316 .303 .310 .308 .304 .298 .293
Table 3.

, , and for , with .

Table 4.

Counts of numbers not in the range of to various heights from through , along with the corresponding densities .

100000000 11355049 0.1136 1000000000 113482572 0.1135
200000000 22718595 0.1136 2000000000 226723208 0.1134
300000000 34079552 0.1136 3000000000 339811983 0.1133
400000000 45433178 0.1136 4000000000 452815609 0.1132
500000000 56783449 0.1136 5000000000 565739195 0.1131
600000000 68129835 0.1135 6000000000 678597689 0.1131
700000000 79472816 0.1135 7000000000 791405364 0.1131
800000000 90811869 0.1135 8000000000 904168542 0.1130
900000000 102146368 0.1135 9000000000 1016908189 0.1130
10000000000 1129598504 0.1130

Mathematical Fragments

Equation (1.1)
Theorem 1.1.

Both the count of down-up reversals in and the count of up-down reversals in are bounded above by Equation 1.1 with .

Theorem 1.2.

The number of up-down reversals in is .

Theorem 1.3.

The number of down-up reversals in is .

Conjecture 1.4.

The set of nonaliquot numbers has asymptotic density , where

Equation (1.2)
Theorem 1.5.

The number of primitive friendly pairs contained in is at most , as .

Theorem 1.6.

For each fixed nonnegative integer , we have , as , for some constant . Moreover, as .

Proposition 2.2.

The number of convenient solutions to is at most , as , uniformly in with .

Lemma 2.3.

Fix . The number of primitive nondeficient with is , as .

Lemma 2.5.

The number of integers for which is not divisible by every integer is .

Lemma 2.6.

The number of primitive nondeficient numbers is at most

Lemma 2.7.

If is in the range of , then uniformly for , the number of integers with

is .

Equation (2.1)
Equation (2.2)
Equation (2.3)
Equation (2.4)
Equation (2.5)
Equation (2.6)
Equation (2.7)
Equation (3.1)
Equation (3.2)
Equation (3.3)
Equation (3.4)
Equation (3.5)
Equation (3.6)
Equation (3.7)
Equation (3.8)
Lemma 4.1.

The number of for which can be written as a fraction with a given denominator is at most , as , uniformly in .

Lemma 4.2.

Let be an integer in . The number of integers for which is , as , uniformly in .

Lemma 4.3.

Given a natural number , the following algorithm outputs a unitary divisor of with . Moreover, at most inputs correspond to the same output , as .

Equation (4.1)
Equation (4.2)
Equation (4.3)
Equation (4.4)
Equation (4.5)
Equation (4.6)
Lemma 5.1.

Fix . The number of primitive friendly -sets contained in is at most , as .

Equation (5.1)

References

Reference [1]
J. D. Alanen, Empirical study of aliquot sequences, Ph.D. thesis, Yale University, 1972.
Reference [2]
Aria Anavi, Paul Pollack, and Carl Pomerance, On congruences of the form , Int. J. Number Theory 9 (2013), no. 1, 115–124, DOI 10.1142/S1793042112501266. MR2997493,
Show rawAMSref \bib{APP13}{article}{ author={Anavi, Aria}, author={Pollack, Paul}, author={Pomerance, Carl}, title={On congruences of the form $\sigma (n)\equiv a\pmod n$}, journal={Int. J. Number Theory}, volume={9}, date={2013}, number={1}, pages={115--124}, issn={1793-0421}, review={\MR {2997493}}, doi={10.1142/S1793042112501266}, }
Reference [3]
Michael R. Avidon, On the distribution of primitive abundant numbers, Acta Arith. 77 (1996), no. 2, 195–205. MR1411032,
Show rawAMSref \bib{avidon96}{article}{ author={Avidon, Michael R.}, title={On the distribution of primitive abundant numbers}, journal={Acta Arith.}, volume={77}, date={1996}, number={2}, pages={195--205}, issn={0065-1036}, review={\MR {1411032}}, }
Reference [4]
R. C. Baker and G. Harman, Shifted primes without large prime factors, Acta Arith. 83 (1998), no. 4, 331–361. MR1610553,
Show rawAMSref \bib{BH98}{article}{ author={Baker, R. C.}, author={Harman, G.}, title={Shifted primes without large prime factors}, journal={Acta Arith.}, volume={83}, date={1998}, number={4}, pages={331--361}, issn={0065-1036}, review={\MR {1610553}}, }
Reference [5]
William D. Banks and Florian Luca, Nonaliquots and Robbins numbers, Colloq. Math. 103 (2005), no. 1, 27–32, DOI 10.4064/cm103-1-4. MR2148946,
Show rawAMSref \bib{BL05}{article}{ author={Banks, William D.}, author={Luca, Florian}, title={Nonaliquots and Robbins numbers}, journal={Colloq. Math.}, volume={103}, date={2005}, number={1}, pages={27--32}, issn={0010-1354}, review={\MR {2148946}}, doi={10.4064/cm103-1-4}, }
Reference [6]
W. Bosma, Aliquot sequences with small starting values, preprint available from http://www.math.ru.nl/~bosma/.
Reference [7]
Wieb Bosma and Ben Kane, The aliquot constant, Q. J. Math. 63 (2012), no. 2, 309–323, DOI 10.1093/qmath/haq050. MR2925292,
Show rawAMSref \bib{bk}{article}{ author={Bosma, Wieb}, author={Kane, Ben}, title={The aliquot constant}, journal={Q. J. Math.}, volume={63}, date={2012}, number={2}, pages={309--323}, issn={0033-5606}, review={\MR {2925292}}, doi={10.1093/qmath/haq050}, }
Reference [8]
E. Catalan, Propositions et questions diverses (French), Bull. Soc. Math. France 16 (1888), 128–129. MR1504036,
Show rawAMSref \bib{catalan88}{article}{ author={Catalan, E.}, title={Propositions et questions diverses}, language={French}, journal={Bull. Soc. Math. France}, volume={16}, date={1888}, pages={128--129}, issn={0037-9484}, review={\MR {1504036}}, }
Reference [9]
Yong-Gao Chen and Qing-Qing Zhao, Nonaliquot numbers, Publ. Math. Debrecen 78 (2011), no. 2, 439–442, DOI 10.5486/PMD.2011.4820. MR2796778,
Show rawAMSref \bib{CZ11}{article}{ author={Chen, Yong-Gao}, author={Zhao, Qing-Qing}, title={Nonaliquot numbers}, journal={Publ. Math. Debrecen}, volume={78}, date={2011}, number={2}, pages={439--442}, issn={0033-3883}, review={\MR {2796778}}, doi={10.5486/PMD.2011.4820}, }
Reference [10]
N. G. de Bruijn, On the number of positive integers and free of prime factors , Nederl. Acad. Wetensch. Proc. Ser. A. 54 (1951), 50–60. MR0046375,
Show rawAMSref \bib{deB}{article}{ author={de Bruijn, N. G.}, title={On the number of positive integers $\leq x$ and free of prime factors $>y$}, journal={Nederl. Acad. Wetensch. Proc. Ser. A.}, volume={54}, date={1951}, pages={50--60}, review={\MR {0046375}}, }
Reference [11]
N. G. de Bruijn, On the number of positive integers and free prime factors . II, Nederl. Akad. Wetensch. Proc. Ser. A 69=Indag. Math. 28 (1966), 239–247. MR0205945,
Show rawAMSref \bib{deB-2}{article}{ author={de Bruijn, N. G.}, title={On the number of positive integers $\leq x$ and free prime factors $>y$. II}, journal={Nederl. Akad. Wetensch. Proc. Ser. A 69=Indag. Math.}, volume={28}, date={1966}, pages={239--247}, review={\MR {0205945}}, }
Reference [12]
L. E. Dickson, Theorems and tables on the sum of the divisors of a number, Q. J. Math. 44 (1913), 264–296.
Reference [13]
P. Erdős, On primitive abundant numbers, J. London Math. Soc. 10 (1935), 49–58.
Reference [14]
P. Erdős, On the normal number of prime factors of and some related problems concerning Euler’s -function, Q. J. Math., Oxford Ser. 6 (1935), 205–213.
Reference [15]
P. Erdös, On amicable numbers, Publ. Math. Debrecen 4 (1955), 108–111. MR0069198,
Show rawAMSref \bib{erdos55}{article}{ author={Erd{\"o}s, P.}, title={On amicable numbers}, journal={Publ. Math. Debrecen}, volume={4}, date={1955}, pages={108--111}, issn={0033-3883}, review={\MR {0069198}}, }
Reference [16]
P. Erdös, On perfect and multiply perfect numbers, Ann. Mat. Pura Appl. (4) 42 (1956), 253–258. MR0082516,
Show rawAMSref \bib{erdos56}{article}{ author={Erd{\"o}s, P.}, title={On perfect and multiply perfect numbers}, journal={Ann. Mat. Pura Appl. (4)}, volume={42}, date={1956}, pages={253--258}, issn={0003-4622}, review={\MR {0082516}}, }
Reference [17]
P. Erdős, Remarks on number theory. II. Some problems on the function, Acta Arith. 5 (1959), 171–177. MR0107623,
Show rawAMSref \bib{erdos59}{article}{ author={Erd{\H {o}}s, P.}, title={Remarks on number theory. II. Some problems on the $\sigma $\ function}, journal={Acta Arith.}, volume={5}, date={1959}, pages={171--177}, issn={0065-1036}, review={\MR {0107623}}, }
Reference [18]
P. Erdős, Über die Zahlen der Form und , Elem. Math. 28 (1973), 83–86.
Reference [19]
P. Erdős, On the distribution of numbers of the form and on some related questions, Pacific J. Math. 52 (1974), 59–65. MR0354601,
Show rawAMSref \bib{erdos74}{article}{ author={Erd{\H {o}}s, P.}, title={On the distribution of numbers of the form $\sigma (n)/n$ and on some related questions}, journal={Pacific J. Math.}, volume={52}, date={1974}, pages={59--65}, issn={0030-8730}, review={\MR {0354601}}, }
Reference [20]
P. Erdős, On asymptotic properties of aliquot sequences, Math. Comp. 30 (1976), no. 135, 641–645. MR0404115,
Show rawAMSref \bib{erdos76}{article}{ author={Erd{\H {o}}s, P.}, title={On asymptotic properties of aliquot sequences}, journal={Math. Comp.}, volume={30}, date={1976}, number={135}, pages={641--645}, issn={0025-5718}, review={\MR {0404115}}, }
Reference [21]
P. Erdős and G. J. Rieger, Ein Nachtrag über befreundete Zahlen (German), J. Reine Angew. Math. 273 (1975), 220. MR0364157,
Show rawAMSref \bib{ER75}{article}{ author={Erd{\H {o}}s, P.}, author={Rieger, G. J.}, title={Ein Nachtrag \"uber befreundete Zahlen}, language={German}, journal={J. Reine Angew. Math.}, volume={273}, date={1975}, pages={220}, issn={0075-4102}, review={\MR {0364157}}, }
Reference [22]
P. Erdős, A. Granville, C. Pomerance, and C. Spiro, On the normal behavior of the iterates of some arithmetic functions, Analytic number theory (Allerton Park, IL, 1989), Progr. Math., vol. 85, Birkhäuser Boston, Boston, MA, 1990, pp. 165–204, DOI 10.1007/978-1-4612-3464-7_13. MR1084181,
Show rawAMSref \bib{erdos90}{article}{ author={Erd{\H {o}}s, P.}, author={Granville, A.}, author={Pomerance, C.}, author={Spiro, C.}, title={On the normal behavior of the iterates of some arithmetic functions}, conference={ title={Analytic number theory}, address={Allerton Park, IL}, date={1989}, }, book={ series={Progr. Math.}, volume={85}, publisher={Birkh\"auser Boston, Boston, MA}, }, date={1990}, pages={165--204}, review={\MR {1084181}}, doi={10.1007/978-1-4612-3464-7\_13}, }
Reference [23]
Paul Erdős, Florian Luca, and Carl Pomerance, On the proportion of numbers coprime to a given integer, Anatomy of integers, CRM Proc. Lecture Notes, vol. 46, Amer. Math. Soc., Providence, RI, 2008, pp. 47–64. MR2437964,
Show rawAMSref \bib{ELP08}{article}{ author={Erd{\H {o}}s, Paul}, author={Luca, Florian}, author={Pomerance, Carl}, title={On the proportion of numbers coprime to a given integer}, conference={ title={Anatomy of integers}, }, book={ series={CRM Proc. Lecture Notes}, volume={46}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2008}, pages={47--64}, review={\MR {2437964}}, }
Reference [24]
Kevin Ford, Florian Luca, and Carl Pomerance, Common values of the arithmetic functions and , Bull. Lond. Math. Soc. 42 (2010), no. 3, 478–488, DOI 10.1112/blms/bdq014. MR2651943,
Show rawAMSref \bib{FLP}{article}{ author={Ford, Kevin}, author={Luca, Florian}, author={Pomerance, Carl}, title={Common values of the arithmetic functions $\phi $ and $\sigma $}, journal={Bull. Lond. Math. Soc.}, volume={42}, date={2010}, number={3}, pages={478--488}, issn={0024-6093}, review={\MR {2651943}}, doi={10.1112/blms/bdq014}, }
Reference [25]
Richard K. Guy and J. L. Selfridge, What drives an aliquot sequence?, Math. Comput. 29 (1975), 101–107. Collection of articles dedicated to Derrick Henry Lehmer on the occasion of his seventieth birthday. MR0384669,
Show rawAMSref \bib{GS}{article}{ author={Guy, Richard K.}, author={Selfridge, J. L.}, title={What drives an aliquot sequence?}, note={Collection of articles dedicated to Derrick Henry Lehmer on the occasion of his seventieth birthday}, journal={Math. Comput.}, volume={29}, date={1975}, pages={101--107}, issn={0378-4754}, review={\MR {0384669}}, }
Reference [26]
H. Halberstam and H.-E. Richert, Sieve methods, Academic Press [A subsidiary of Harcourt Brace Jovanovich, Publishers], London-New York, 1974. London Mathematical Society Monographs, No. 4. MR0424730,
Show rawAMSref \bib{HR}{book}{ author={Halberstam, H.}, author={Richert, H.-E.}, title={Sieve methods}, note={London Mathematical Society Monographs, No. 4}, publisher={Academic Press [A subsidiary of Harcourt Brace Jovanovich, Publishers], London-New York}, date={1974}, pages={xiv+364 pp. (loose errata)}, review={\MR {0424730}}, }
Reference [27]
Bernhard Hornfeck and Eduard Wirsing, Über die Häufigkeit vollkommener Zahlen (German), Math. Ann. 133 (1957), 431–438. MR0090600,
Show rawAMSref \bib{HW57}{article}{ author={Hornfeck, Bernhard}, author={Wirsing, Eduard}, title={\"Uber die H\"aufigkeit vollkommener Zahlen}, language={German}, journal={Math. Ann.}, volume={133}, date={1957}, pages={431--438}, issn={0025-5831}, review={\MR {0090600}}, }
Reference [28]
Mitsuo Kobayashi, Paul Pollack, and Carl Pomerance, On the distribution of sociable numbers, J. Number Theory 129 (2009), no. 8, 1990–2009, DOI 10.1016/j.jnt.2008.10.011. MR2522719,
Show rawAMSref \bib{kpp}{article}{ author={Kobayashi, Mitsuo}, author={Pollack, Paul}, author={Pomerance, Carl}, title={On the distribution of sociable numbers}, journal={J. Number Theory}, volume={129}, date={2009}, number={8}, pages={1990--2009}, issn={0022-314X}, review={\MR {2522719}}, doi={10.1016/j.jnt.2008.10.011}, }
Reference [29]
H. W. Lenstra, Jr., Problem 6064, Amer. Math. Monthly 82 (1975), 1016. Solution by the proposer, op. cit. 84 (1977), 580.
Reference [30]
Florian Luca and Carl Pomerance, The range of the sum-of-proper-divisors function, Acta Arith. 168 (2015), no. 2, 187–199, DOI 10.4064/aa168-2-6. MR3339454,
Show rawAMSref \bib{LP14}{article}{ author={Luca, Florian}, author={Pomerance, Carl}, title={The range of the sum-of-proper-divisors function}, journal={Acta Arith.}, volume={168}, date={2015}, number={2}, pages={187--199}, issn={0065-1036}, review={\MR {3339454}}, doi={10.4064/aa168-2-6}, }
Reference [31]
Tomás Oliveira e Silva, Siegfried Herzog, and Silvio Pardi, Empirical verification of the even Goldbach conjecture and computation of prime gaps up to , Math. Comp. 83 (2014), no. 288, 2033–2060, DOI 10.1090/S0025-5718-2013-02787-1. MR3194140,
Show rawAMSref \bib{silva14}{article}{ author={Oliveira e Silva, Tom{\'a}s}, author={Herzog, Siegfried}, author={Pardi, Silvio}, title={Empirical verification of the even Goldbach conjecture and computation of prime gaps up to $4\cdot 10^{18}$}, journal={Math. Comp.}, volume={83}, date={2014}, number={288}, pages={2033--2060}, issn={0025-5718}, review={\MR {3194140}}, doi={10.1090/S0025-5718-2013-02787-1}, }
Reference [32]
Paul Pollack, On the greatest common divisor of a number and its sum of divisors, Michigan Math. J. 60 (2011), no. 1, 199–214, DOI 10.1307/mmj/1301586311. MR2785871,
Show rawAMSref \bib{pollack11}{article}{ author={Pollack, Paul}, title={On the greatest common divisor of a number and its sum of divisors}, journal={Michigan Math. J.}, volume={60}, date={2011}, number={1}, pages={199--214}, issn={0026-2285}, review={\MR {2785871}}, doi={10.1307/mmj/1301586311}, }
Reference [33]
Paul Pollack, Some arithmetic properties of the sum of proper divisors and the sum of prime divisors, Illinois J. Math. 58 (2014), no. 1, 125–147. MR3331844,
Show rawAMSref \bib{pollack14}{article}{ author={Pollack, Paul}, title={Some arithmetic properties of the sum of proper divisors and the sum of prime divisors}, journal={Illinois J. Math.}, volume={58}, date={2014}, number={1}, pages={125--147}, issn={0019-2082}, review={\MR {3331844}}, }
Reference [34]
Paul Pollack, Remarks on fibers of the sum-of-divisors function, in Analytic number theory (in honor of Helmut Maier’s 60th birthday), M. Rassias and C. Pomerance, eds., Springer, Cham, Switzerland, 2015, pp. 305–320.
Reference [35]
Paul Pollack and Carl Pomerance, Prime-perfect numbers, Integers 12 (2012), no. 6, 1417–1437, DOI 10.1515/integers-2012-0044. MR3011565,
Show rawAMSref \bib{PP12}{article}{ author={Pollack, Paul}, author={Pomerance, Carl}, title={Prime-perfect numbers}, journal={Integers}, volume={12}, date={2012}, number={6}, pages={1417--1437}, issn={1867-0652}, review={\MR {3011565}}, doi={10.1515/integers-2012-0044}, }
Reference [36]
Carl Pomerance, On the distribution of amicable numbers, J. Reine Angew. Math. 293/294 (1977), 217–222. MR0447087,
Show rawAMSref \bib{pomerance77}{article}{ author={Pomerance, Carl}, title={On the distribution of amicable numbers}, journal={J. Reine Angew. Math.}, volume={293/294}, date={1977}, pages={217--222}, issn={0075-4102}, review={\MR {0447087}}, }
Reference [37]
Carl Pomerance, Popular values of Euler’s function, Mathematika 27 (1980), no. 1, 84–89, DOI 10.1112/S0025579300009967. MR581999,
Show rawAMSref \bib{pomerance80}{article}{ author={Pomerance, Carl}, title={Popular values of Euler's function}, journal={Mathematika}, volume={27}, date={1980}, number={1}, pages={84--89}, issn={0025-5793}, review={\MR {581999}}, doi={10.1112/S0025579300009967}, }
Reference [38]
Carl Pomerance, On amicable numbers, in Analytic number theory (in honor of Helmut Maier’s 60th birthday), M. Rassias and C. Pomerance, eds., Springer, Cham, Switzerland, 2015, pp. 321–327.
Reference [39]
Carl Pomerance and Hee-Sung Yang, Variant of a theorem of Erdős on the sum-of-proper-divisors function, Math. Comp. 83 (2014), no. 288, 1903–1913, DOI 10.1090/S0025-5718-2013-02775-5. MR3194134,
Show rawAMSref \bib{PY14}{article}{ author={Pomerance, Carl}, author={Yang, Hee-Sung}, title={Variant of a theorem of Erd\H os on the sum-of-proper-divisors function}, journal={Math. Comp.}, volume={83}, date={2014}, number={288}, pages={1903--1913}, issn={0025-5718}, review={\MR {3194134}}, doi={10.1090/S0025-5718-2013-02775-5}, }
Reference [40]
G. J. Rieger, Bemerkung zu einem Ergebnis von Erdős über befreundete Zahlen (German), J. Reine Angew. Math. 261 (1973), 157–163. MR0319882,
Show rawAMSref \bib{rieger73}{article}{ author={Rieger, G. J.}, title={Bemerkung zu einem Ergebnis von Erd\H os \"uber befreundete Zahlen}, language={German}, journal={J. Reine Angew. Math.}, volume={261}, date={1973}, pages={157--163}, issn={0075-4102}, review={\MR {0319882}}, }
Reference [41]
J. Sesiano, Two problems of number theory in Islamic times, Arch. Hist. Exact Sci. 41 (1991), no. 3, 235–238, DOI 10.1007/BF00348408. MR1107382,
Show rawAMSref \bib{sesiano91}{article}{ author={Sesiano, J.}, title={Two problems of number theory in Islamic times}, journal={Arch. Hist. Exact Sci.}, volume={41}, date={1991}, number={3}, pages={235--238}, issn={0003-9519}, review={\MR {1107382}}, doi={10.1007/BF00348408}, }
Reference [42]
N. J. A. Sloane, On-line encyclopedia of integer sequences, http://oeis.org. Sequence A001970; gp script by Michael Marcus online at http://oeis.org/A246827/a246827.gp.txt.
Reference [43]
Gérald Tenenbaum, Introduction to analytic and probabilistic number theory, Cambridge Studies in Advanced Mathematics, vol. 46, Cambridge University Press, Cambridge, 1995. Translated from the second French edition (1995) by C. B. Thomas. MR1342300,
Show rawAMSref \bib{tenenbaum}{book}{ author={Tenenbaum, G{\'e}rald}, title={Introduction to analytic and probabilistic number theory}, series={Cambridge Studies in Advanced Mathematics}, volume={46}, note={Translated from the second French edition (1995) by C. B. Thomas}, publisher={Cambridge University Press, Cambridge}, date={1995}, pages={xvi+448}, isbn={0-521-41261-7}, review={\MR {1342300}}, }
Reference [44]
H. J. J. te Riele, A theoretical and computational study of generalized aliquot sequences, Mathematisch Centrum, Amsterdam, 1976. Mathematical Centre Tracts, No. 74. MR0556033,
Show rawAMSref \bib{tR76}{book}{ author={te Riele, H. J. J.}, title={A theoretical and computational study of generalized aliquot sequences}, note={Mathematical Centre Tracts, No. 74}, publisher={Mathematisch Centrum, Amsterdam}, date={1976}, pages={x+76}, isbn={90-6196-131-9}, review={\MR {0556033}}, }
Reference [45]
Vincent Toulmonde, Module de continuité de la fonction de répartition de (French), Acta Arith. 121 (2006), no. 4, 367–402, DOI 10.4064/aa121-4-6. MR2224402,
Show rawAMSref \bib{Toul}{article}{ author={Toulmonde, Vincent}, title={Module de continuit\'e de la fonction de r\'epartition de $\phi (n)/n$}, language={French}, journal={Acta Arith.}, volume={121}, date={2006}, number={4}, pages={367--402}, issn={0065-1036}, review={\MR {2224402}}, doi={10.4064/aa121-4-6}, }
Reference [46]
R. C. Vaughan, The Hardy-Littlewood method, 2nd ed., Cambridge Tracts in Mathematics, vol. 125, Cambridge University Press, Cambridge, 1997, DOI 10.1017/CBO9780511470929. MR1435742,
Show rawAMSref \bib{vaughan97}{book}{ author={Vaughan, R. C.}, title={The Hardy-Littlewood method}, series={Cambridge Tracts in Mathematics}, volume={125}, edition={2}, publisher={Cambridge University Press, Cambridge}, date={1997}, pages={xiv+232}, isbn={0-521-57347-5}, review={\MR {1435742}}, doi={10.1017/CBO9780511470929}, }
Reference [47]
Eduard Wirsing, Bemerkung zu der Arbeit über vollkommene Zahlen (German), Math. Ann. 137 (1959), 316–318. MR0104636,
Show rawAMSref \bib{wirsing59}{article}{ author={Wirsing, Eduard}, title={Bemerkung zu der Arbeit \"uber vollkommene Zahlen}, language={German}, journal={Math. Ann.}, volume={137}, date={1959}, pages={316--318}, issn={0025-5831}, review={\MR {0104636}}, }

Article Information

MSC 2010
Primary: 11N37 (Asymptotic results on arithmetic functions)
Secondary: 11N64 (Other results on the distribution of values or the characterization of arithmetic functions)
Author Information
Paul Pollack
Department of Mathematics, University of Georgia, Athens, Georgia 30602
pollack@uga.edu
MathSciNet
Carl Pomerance
Department of Mathematics, Dartmouth College, Hanover, New Hampshire 03755
carl.pomerance@dartmouth.edu
MathSciNet
Additional Notes

The research of the first named author was supported in part by NSF grant DMS-1402268.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 3, Issue 1, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2016 by the authors under Creative Commons Attribution-Noncommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/10
  • MathSciNet Review: 3481968
  • Show rawAMSref \bib{3481968}{article}{ author={Pollack, Paul}, author={Pomerance, Carl}, title={Some problems of Erd\H{o}s on the sum-of-divisors function}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={3}, number={1}, date={2016}, pages={1-26}, issn={2330-0000}, review={3481968}, doi={10.1090/btran/10}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.