This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Proof of the Ballantine-Merca Conjecture and theta function identities modulo 22

Letong Hong, Shengtong Zhang Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139 [email protected], [email protected]
(Date: January 2021)
Abstract.

For positive integers mm we consider the theta functions fm(z):=mk+1 square qkf_{m}(z):=\sum_{mk+1\text{ square }}q^{k}. Due to classical identities of Jacobi, it is known that

f4f6f12(mod2).f_{4}\equiv f_{6}f_{12}\pmod{2}.

Here we prove that the only triples (a,b,c)(a,b,c) for which fafbfc(mod2)f_{a}\equiv f_{b}f_{c}\pmod{2} are of the form (2q,4q,4q)(2q,4q,4q) or (4q,4q,8q)(4q,4q,8q), where qq is any positive odd number, or belong to the following finite list

{(4,6,12),(6,8,24),(8,12,24),(10,12,60),(15,24,40),(16,24,48),(20,24,120),(21,24,168)}.\{(4,6,12),(6,8,24),(8,12,24),(10,12,60),(15,24,40),(16,24,48),(20,24,120),(21,24,168)\}.

The result is inspired by the Ballantine-Merca Conjecture on recurrence relations for the parity of the partition function p(n)p(n), which we also prove here.

2010 Mathematics Subject Classification. 05A17, 05A19, 11P81, 11P83, 11F27
Keywords: Theta functions, Partitions, Parity

1. Introduction and Statement of Results

We recall the famous theta functions defined by Euler and Jacobi:

(1.1) (q;q)=n=(1)nqn(3n+1)/2,(q;q)_{\infty}=\sum_{n=-\infty}^{\infty}(-1)^{n}q^{n(3n+1)/2},
(q;q)3=n=0(1)n(2n+1)qn(n+1)/2,(q;q)_{\infty}^{3}=\sum_{n=0}^{\infty}(-1)^{n}(2n+1)q^{n(n+1)/2},

where (q;q):=n=1(1qn).(q;q)_{\infty}:=\prod_{n=1}^{\infty}(1-q^{n}). These identities motivate the definition of the following class of theta functions:

fa(q):=an+1 is a squareqn.f_{a}(q):=\sum_{an+1\text{ is a square}}q^{n}.

Observe that Euler and Jacobi’s identities give us

(q;q)af24/a(q)(mod2)(q;q)_{\infty}^{a}\equiv f_{24/a}(q)\pmod{2}

for a{1,2,3,4,6}a\in\{1,2,3,4,6\}. The goal of our paper is to classify such identities modulo 22. Our main result is the following classification theorem.

Theorem 1.1.

The only triples of positive integers (a,b,c)(a,b,c) such that fafbfc(mod2)f_{a}\equiv f_{b}f_{c}\pmod{2} are of the form

(2q,4q,4q),(4q,8q,8q),(2q,4q,4q),(4q,8q,8q),

where qq is any positive odd number, or are members of the finite set:

{(4,6,12),(6,8,24),(8,12,24),(10,12,60),(15,24,40),(16,24,48),(20,24,120),(21,24,168)}.\{(4,6,12),(6,8,24),(8,12,24),(10,12,60),(15,24,40),(16,24,48),(20,24,120),(21,24,168)\}.

This theorem is related to recent work of Ballantine and Merca [1] on the integer partition function. Recall that a partition of any non-negative integer nn is any non-increasing sequence of positive integers which sum to nn, and p(n)p(n) denotes their total number (by convention, p(0)=1p(0)=1). The generating function for p(n)p(n) is

P(q):=n=1p(n)qn=n=111qn=1(q;q).P(q):=\sum_{n=1}^{\infty}p(n)q^{n}=\prod_{n=1}^{\infty}\frac{1}{1-q^{n}}=\frac{1}{(q;q)_{\infty}}.

Thanks to the Euler-Jacobi identity (1.1), this gives the recurrence

p(n)=j=1n(1)j+1[p(n12j(3j1))+p(n12j(3j+1))].p(n)=\sum_{j=1}^{n}(-1)^{j+1}\left[p\left(n-\frac{1}{2}j(3j-1)\right)+p\left(n-\frac{1}{2}j(3j+1)\right)\right].

Ballantine and Merca demonstrated further recurrences for the parity of p(n)p(n). They proved that for (a,b)(a,b) in a certain set (listed below), the quantity

ak+1 squarep(nk)\sum_{ak+1\text{ square}}p(n-k)

is odd if and only if bn+1bn+1 is a square. They asked whether the converse is true.

Conjecture ([1]).

We have that

(1.2) ak+1 is squarep(nk)1(mod2) if and only if bn+1 is a square.\sum_{\text{ak+1 is square}}p(n-k)\equiv 1\pmod{2}\ \ \text{ if and only if }bn+1\text{ is a square}.

is true only for (a,b){(6,8),(8,12),(12,24),(15,40),(16,48),(20,120),(21,168)}.(a,b)\in\{(6,8),(8,12),(12,24),(15,40),(16,48),(20,120),(21,168)\}.

We note that statement 1.2 is equivalent to the modulo 22 theta function identity

fafbf24(mod2).f_{a}\equiv f_{b}f_{24}\pmod{2}.

Thus Theorem 1.1 resolves Ballantine-Merca’s conjecture.

Corollary 1.2.

The Ballantine-Merca Conjecture is true.

The paper is organized as follows. In Section 2 we give an important proposition that leads directly to the proof of the Ballantine-Merca Conjecture, and in Section 3 we prove Theorem 1.1 by discovering some further constraints.

Acknowledgements

We would like to thank Professor Ken Ono for his guidance and many insightful discussions. We also thank Dr Mircea Merca for pointing out a typographical error in an earlier version of the paper. We are grateful to the anonymous referee for detailed and helpful suggestions. The research was supported by the generosity of the National Science Foundation under grant DMS-2002265, the National Security Agency under grant H98230-20-1-0012, the Templeton World Charity Foundation, and the Thomas Jefferson Fund at the University of Virginia.

2. Proof of Corollary 1.2

In this section we prove the following important proposition that resolves Corollary 1.2.

Proposition 2.1.

If a,b,ca,b,c are positive integers with fafbfc(mod2)f_{a}\equiv f_{b}f_{c}\pmod{2}, then 1a=1b+1c.\frac{1}{a}=\frac{1}{b}+\frac{1}{c}.

Proof.

If fafbfc(mod2)f_{a}\equiv f_{b}f_{c}\pmod{2} then for all k1k_{1}\in\mathbb{N}, ak1+1=x2ak_{1}+1=x^{2} for some integer xx if and only if there exists an odd number of pairs (y,z)2(y,z)\in\mathbb{N}^{2} such that

(2.1) k1=y21b+z21c,b(y21),c(z21).k_{1}=\frac{y^{2}-1}{b}+\frac{z^{2}-1}{c},\quad b\mid(y^{2}-1),\quad c\mid(z^{2}-1).

We rewrite the equation as

(2.2) bz2+cy2=bck1+b+c,b(y21),c(z21).bz^{2}+cy^{2}=bck_{1}+b+c,\quad b\mid(y^{2}-1),\quad c\mid(z^{2}-1).

Notice that if we take a prime number pp such that the Legendre symbol (bcp)=1(\frac{-bc}{p})=-1, the integer solutions for the equation bm2+cn2=Nbm^{2}+cn^{2}=N are in bijection with the integer solutions for bm2+cn2=p2tNbm^{2}+cn^{2}=p^{2t}N by considering (m,n)(ptm,ptn)(m,n)\to(p^{t}m,p^{t}n). Thus, if we take tt such that pt1(modbc)p^{t}\equiv 1\pmod{bc}, then the number of positive integer solutions to

bz2+cy2=bck1+b+c,b(y21),c(z21).bz^{2}+cy^{2}=bck_{1}+b+c,\quad b\mid(y^{2}-1),\quad c\mid(z^{2}-1).

is the same as the number of positive integer solutions to

bz2+cy2=p2t(bck1+b+c),b(y21),c(z21).bz^{2}+cy^{2}=p^{2t}(bck_{1}+b+c),\quad b\mid(y^{2}-1),\quad c\mid(z^{2}-1).

Now note that

p2t(bck1+b+c)=bc(p2tk1+(p2t1)(b+c)bc)+b+c.p^{2t}(bck_{1}+b+c)=bc\left(p^{2t}k_{1}+\frac{(p^{2t}-1)(b+c)}{bc}\right)+b+c.

Therefore, we conclude that ak1+1ak_{1}+1 is a perfect square if and only if

a(p2tk1+(p2t1)(b+c)bc)+1a\left(p^{2t}k_{1}+\frac{(p^{2t}-1)(b+c)}{bc}\right)+1

is a perfect square. There is an infinite number of k1k_{1} such that ak1+1ak_{1}+1 is square. For each such k1k_{1}, we have

a(p2tk1+(p2t1)(b+c)bc)+1=p2t(ak1+1)+(p2t1)(b1+c1a1).a\left(p^{2t}k_{1}+\frac{(p^{2t}-1)(b+c)}{bc}\right)+1=p^{2t}(ak_{1}+1)+(p^{2t}-1)\left(b^{-1}+c^{-1}-a^{-1}\right).

Thus the perfect square

a(p2tk1+(p2t1)(b+c)bc)+1a\left(p^{2t}k_{1}+\frac{(p^{2t}-1)(b+c)}{bc}\right)+1

and the other perfect square

p2t(ak1+1)p^{2t}(ak_{1}+1)

differ by a constant

(p2t1)(b1+c1a1).(p^{2t}-1)\left(b^{-1}+c^{-1}-a^{-1}\right).

For infinitely many k1k_{1} to satisfy this property above, the only possibility is that

(p2t1)(b1+c1a1)=0,(p^{2t}-1)\left(b^{-1}+c^{-1}-a^{-1}\right)=0,

which implies 1/a=1/b+1/c,1/a=1/b+1/c, the desired result. ∎

As we stated in the introduction, the modulo constraint 1.2 in Ballantine-Merca Conjecture is true if and only if

fafbf24(mod2).f_{a}\equiv f_{b}f_{24}\pmod{2}.

Applying Proposition 2.1 we obtain

1a=1b+124\frac{1}{a}=\frac{1}{b}+\frac{1}{24}

and the listed pairs in Ballantine and Merca’s paper plus (22,264),(23,552)(22,264),(23,552) are all solutions. Since they already verified all a,b100000a,b\leq 100000, the conjecture is solved.

3. Proof of Theorem 1.1

In this section we apply more sophisticated analysis to prove Theorem 1.1. Throughout the section, we assume a,b,ca,b,c are positive integers with fafbfc(mod2)f_{a}\equiv f_{b}f_{c}\pmod{2}. We denote d=gcd(b,c)d=\gcd(b,c), b=bdb^{\prime}=\frac{b}{d}, and c=cdc^{\prime}=\frac{c}{d}. For a prime pp and integer nn, denote vp(n)=max{a0:pa|n}.v_{p}(n)=\max\{a\geq 0:p^{a}|n\}.

3.1. A bound on b+cb^{\prime}+c^{\prime}.

Proposition 3.1.

If positive integers a,b,ca,b,c satisfy fafbfc(mod2)f_{a}\equiv f_{b}f_{c}\pmod{2}, then we must have

(b+c)24.(b^{\prime}+c^{\prime})\mid 24.

Furthermore, if b+cb^{\prime}+c^{\prime} is divisible by 22 then we have v2(d)4v_{2}(d)\leq 4, and if b+cb^{\prime}+c^{\prime} is divisible by 33 then we have v3(d)1v_{3}(d)\leq 1.

Proof.

By Proposition 2.1 we have 1a=1b+1c\frac{1}{a}=\frac{1}{b}+\frac{1}{c}. Our condition fafbfc(mod2)f_{a}\equiv f_{b}f_{c}\pmod{2} corresponds to the statement that the number of solutions (y,z)2(y,z)\in\mathbb{N}^{2} to

(3.1) cy2+bz2=bcx+b+c=(b+c)(ax+1),b(y21),c(z21)cy^{2}+bz^{2}=bcx+b+c=(b+c)(ax+1),\quad b\mid(y^{2}-1),\quad c\mid(z^{2}-1)

is odd if and only ax+1ax+1 is a square.

First note that (b+c)d(b^{\prime}+c^{\prime})\mid d. This is because a(b+c)=dbca(b^{\prime}+c^{\prime})=db^{\prime}c^{\prime} and gcd(bc,b+c)=1\gcd(b^{\prime}c^{\prime},b^{\prime}+c^{\prime})=1.

We consider any prime number p2,3p\neq 2,3 satisfying the following modulo constraints

(3.2) a(p21),(bcp)=1.a\mid(p^{2}-1),\quad\left(\frac{-b^{\prime}c^{\prime}}{p}\right)=-1.

Letting δ=p21a\delta=\frac{p^{2}-1}{a}, we substitute x=δx=\delta to Equation (3.1) and obtain cy2+bz2=p2(b+c),cy^{2}+bz^{2}=p^{2}(b+c), or cy2+bz2=p2(b+c)c^{\prime}y^{2}+b^{\prime}z^{2}=p^{2}(b^{\prime}+c^{\prime}). Since (bcp)=1,\big{(}\frac{-b^{\prime}c^{\prime}}{p}\big{)}=-1, it follows pp divides y,zy,z. Furthermore, any solution (y,z)(y,z) of 3.1 must have y>0y>0, as otherwise b|y21=1b|y^{2}-1=1, so b=1b=1, which contradicts the identity 1a=1b+1c\frac{1}{a}=\frac{1}{b}+\frac{1}{c}. Symmetrically, we must have z>0z>0. Thus the only possible solution to 3.1 is (y,z)=(p,p)(y,z)=(p,p). Since ax+1=aδ+1=p2ax+1=a\delta+1=p^{2} is a square, 3.1 must have an odd number of solutions, thus (y,z)=(p,p)(y,z)=(p,p) must be a solution to 3.1. Therefore, we obtain b(p21),c(p21),b\mid(p^{2}-1),\ c\mid(p^{2}-1), or equivalently

(3.3) db(p21),dc(p21).db^{\prime}\mid(p^{2}-1),\quad dc^{\prime}\mid(p^{2}-1).

We have thus shown that for any prime pp, 3.2 implies 3.3. We now attempt to construct some pp that satisfies 3.2 but not 3.3. We next introduce a technical lemma to establish the existence of a desirable pp.

Lemma 3.2.

Assume uu and vv are positive integers with vv square free. Let Q=lcm(8,u,v)Q=\text{lcm}(8,u,v). Then there exists a residue class qq mod QQ with (q,Q)=1(q,Q)=1, such that for any prime number pq(modQ)p\equiv q\pmod{Q} we have

p210(modu),(vp)=1.p^{2}-1\equiv 0\pmod{u},\quad\left(\frac{-v}{p}\right)=-1.

Furthermore, if vv contains a prime divisor that is 3 mod 4 and uu is divisible by 88, we could further take qq mod QQ such that v2(p21)=v2(Q)v_{2}(p^{2}-1)=v_{2}(Q).

Before showing the lemma, we see how it finishes the proof of Proposition 3.1. We take u=a=db+cbcu=a=\frac{d}{b^{\prime}+c^{\prime}}b^{\prime}c^{\prime} and vv be the square-free reduction of bcb^{\prime}c^{\prime}. For every prime q(b+c)q\mid(b^{\prime}+c^{\prime}), we write b+c=qαs,d=qβdb^{\prime}+c^{\prime}=q^{\alpha}s,\ d=q^{\beta}d^{\prime} where gcd(s,q)=gcd(d,q)=1\gcd(s,q)=\gcd(d^{\prime},q)=1. As b+c|db^{\prime}+c^{\prime}|d, we have αβ\alpha\leq\beta. Note that

vq(u)=βα,vq(v)=0.v_{q}(u)=\beta-\alpha,\quad v_{q}(v)=0.

If q3q\geq 3, then Lemma 3.2 asserts that fixing p(modqβα)p\pmod{q^{\beta-\alpha}} suffices to ensure that pp satisfies (3.2). But for all such pp, we must have qβ(p21)q^{\beta}\mid(p^{2}-1) by 3.3. This implication holds for all pp only if q=3q=3 and β=1\beta=1.

If q=2q=2, we show that β3\beta\leq 3. If β4\beta\geq 4, then both bb^{\prime} and cc^{\prime} are odd, and b+c0(mod16)b^{\prime}+c^{\prime}\equiv 0\pmod{16}, so vv must have a prime divisor of 3 mod 4. By Lemma 3.2, we could take pp such that v2(p21)=v2(Q)v_{2}(p^{2}-1)=v_{2}(Q) and pp satisfies (3.2). Now

v2(Q)=max(βα,3)<βv_{2}(Q)=\max(\beta-\alpha,3)<\beta

contradiction the consequence 3.3 that qβ(p21)q^{\beta}\mid(p^{2}-1). Therefore we must have β3\beta\leq 3 and αβ3\alpha\leq\beta\leq 3, finishing the proof. ∎

Proof of Lemma 3.2.

We assume the prime factorization of vv is v=q1q2qmv=q_{1}q_{2}\cdots q_{m}. Let the prime factorization of uu be u=p1a1pnanq1l1q2l2qmlmu=p_{1}^{a_{1}}\cdots p_{n}^{a_{n}}q_{1}^{l_{1}}q_{2}^{l_{2}}\cdots q_{m}^{l_{m}}, where p1,,pnp_{1},\cdots,p_{n} are prime factors distinct from q1,,qmq_{1},\cdots,q_{m}. Then we want to satisfy the congruence constraints

(3.4) p21(modpiai),p21(modqili),p^{2}\equiv 1\pmod{p_{i}^{a_{i}}},\quad p^{2}\equiv 1\pmod{q_{i}^{l_{i}}},

and (q1q2qmp)=1\big{(}\frac{-q_{1}q_{2}\cdots q_{m}}{p}\big{)}=-1. Now we do case work. In each case, we describe a number of congruence conditions for pp, and show that they guarantee that pp satisfy the above constraints. Then the Chinese remainder theorem shows that a modulo class mod QQ can be taken such that all the congruence conditions are satisfied.

Case 1: If there exists an 1im1\leq i\leq m such that qiq_{i} is 3 mod 4. Then we could first satisfy 3.4 for all other prime divisors of QQ arbitrarily. In particular, we could choose p(mod2v2(Q))p\pmod{2^{v_{2}(Q)}} such that v2(p21)=v2(Q)v_{2}(p^{2}-1)=v_{2}(Q). Then by quadratic reciprocity,

(q1q2qmp)=c(pqi),\left(\frac{-q_{1}q_{2}\cdots q_{m}}{p}\right)=c\left(\frac{p}{q_{i}}\right),

where c{±1}c\in\{\pm 1\} is fixed. So we could adjust the sign of p(modqi)p\pmod{q_{i}} to ensure that p±1(modqili)p\equiv\pm 1\pmod{q_{i}^{l_{i}}} and (q1q2qmp)=1\big{(}\frac{-q_{1}q_{2}\cdots q_{m}}{p}\big{)}=-1.

Case 2: If vv is odd and all qiq_{i} are 1 mod 4. Then by quadratic reciprocity we have

(q1qmp)=(1p)(pq1)(pqm).\left(\frac{-q_{1}\cdots q_{m}}{p}\right)=\left(\frac{-1}{p}\right)\left(\frac{p}{q_{1}}\right)\cdots\left(\frac{p}{q_{m}}\right).

Now note that to ensure p21(mod2v2(u))p^{2}\equiv 1\pmod{2^{v_{2}(u)}}, we could take either of p±1(mod2v2(Q))p\equiv\pm 1\pmod{2^{v_{2}(Q)}}. Thus we could first choose modulo classes that satisfy 3.4 for all odd prime divisors of QQ, then choose the sign in p±1(mod2v2(Q))p\equiv\pm 1\pmod{2^{v_{2}(Q)}} to ensure the quadratic character (q1q2qmp)=1\big{(}\frac{-q_{1}q_{2}\cdots q_{m}}{p}\big{)}=-1.

Case 3: If q1=2q_{1}=2 and all other qiq_{i} is 1 mod 4. We observe

(2q2qmp)=(2p)(pq1)(pqm).\left(\frac{-2q_{2}\cdots q_{m}}{p}\right)=\left(\frac{-2}{p}\right)\left(\frac{p}{q_{1}}\right)\cdots\left(\frac{p}{q_{m}}\right).

Now note that to ensure p21(mod2v2(u))p^{2}\equiv 1\pmod{2^{v_{2}(u)}}, we could take either p±1(mod2v2(Q))p\equiv\pm 1\pmod{2^{v_{2}(Q)}}. If p1(mod2v2(Q))p\equiv 1\pmod{2^{v_{2}(Q)}} then (2p)=1\big{(}\frac{-2}{p}\big{)}=1, and if p1(mod2v2(Q))p\equiv-1\pmod{2^{v_{2}(Q)}} then (2p)=1\big{(}\frac{-2}{p}\big{)}=-1. Thus we could first choose modulo classes that satisfy 3.4 for all odd prime divisors of QQ, then choose the ±1\pm 1 to ensure the quadratic character is 1-1. ∎

3.2. A bound on dd.

The final piece of our proof is the following proposition.

Proposition 3.3.

Suppose positive integers a,b,ca,b,c satisfy fafbfc(mod2)f_{a}\equiv f_{b}f_{c}\pmod{2} and bcb^{\prime}\neq c^{\prime}. Then d24(bc)d\mid 24(b^{\prime}-c^{\prime}).

We first prove a lemma.

Lemma 3.4.

Suppose pp is a prime coprime to bcb^{\prime}c^{\prime} that can be expressed in the form u2+bcv2u^{2}+b^{\prime}c^{\prime}v^{2}, where u,vu,v are integers. Then the only integer solutions to cy2+bz2=(b+c)pc^{\prime}y^{2}+b^{\prime}z^{2}=(b^{\prime}+c^{\prime})p with yy coprime to bb^{\prime} and zz coprime to cc^{\prime} are (up to sign) (y,z)=(ubv,u+cv)(y,z)=(u-b^{\prime}v,u+c^{\prime}v) and (y,z)=(u+bv,ucv)(y,z)=(u+b^{\prime}v,u^{\prime}-c^{\prime}v).

Proof.

If cy2+bz2=(b+c)pc^{\prime}y^{2}+b^{\prime}z^{2}=(b^{\prime}+c^{\prime})p, then by congruence modulo pp we must have

yz1±bvu1(modp).yz^{-1}\equiv\pm b^{\prime}vu^{-1}\pmod{p}.

By swapping the sign of y,zy,z, we can, without loss of generality, assume that

yz1bvu1(modp).yz^{-1}\equiv b^{\prime}vu^{-1}\pmod{p}.

Then we have

c(uybvzp)2+b(uz+cvyp)2=b+c.c^{\prime}\left(\frac{uy-b^{\prime}vz}{p}\right)^{2}+b^{\prime}\left(\frac{uz+c^{\prime}vy}{p}\right)^{2}=b^{\prime}+c^{\prime}.

Now that the quantities in the bracket are both integers, either both are ±1\pm 1 or one of them is 0. The latter is impossible by the imposed coprime condition, and the former gives the two pairs of solutions. ∎

Proof of Proposition 3.3.

Let qhq^{h} be any maximal power of a prime dividing dd. It suffices to show qh24(bc)q^{h}\mid 24(b^{\prime}-c^{\prime}). The case of b+cb^{\prime}+c^{\prime} not coprime to qq is handled in Proposition 3.1, so we assume b+cb^{\prime}+c^{\prime} is coprime to qq.

Assume lcm(a,b,c)=qwr\text{lcm}(a,b,c)=q^{w}r, where (r,q)=1(r,q)=1. We consider the congruence classes s,ts,t modulo rqwrq^{w} defined by

s1(modr),sbcb+c(modqw),s\equiv 1\pmod{r},\quad s\equiv\frac{b^{\prime}-c^{\prime}}{b^{\prime}+c^{\prime}}\pmod{q^{w}},

and

t0(modr),t2b+c(modqw),t\equiv 0\pmod{r},\quad t\equiv\frac{2}{b^{\prime}+c^{\prime}}\pmod{q^{w}},

which are well-defined as (b+c,q)=1(b^{\prime}+c^{\prime},q)=1. Then we have

s2+bct21(modrqw).s^{2}+b^{\prime}c^{\prime}t^{2}\equiv 1\pmod{rq^{w}}.

Therefore, by a well-known theorem of Weber([2]), there exist infinitely many primes pp such that p=u2+bcv2p=u^{2}+b^{\prime}c^{\prime}v^{2} with us(modrqw)u\equiv s\pmod{rq^{w}}, vt(modrqw)v\equiv t\pmod{rq^{w}}. Furthermore, for all such pp we have p1(moda)p\equiv 1\pmod{a}. We now study the target set

{(y,z):cy2+bz2=(b+c)p,b(y21),c(z21)}.\{(y,z):\ c^{\prime}y^{2}+b^{\prime}z^{2}=(b^{\prime}+c^{\prime})p,\ b\mid(y^{2}-1),\ c\mid(z^{2}-1)\}.

By Lemma 3.4, the only possible integer solutions up to sign are (y,z)=(ubv,u+cv)(y,z)=(u-b^{\prime}v,u+c^{\prime}v) and (y,z)=(u+bv,ucv)(y,z)=(u+b^{\prime}v,u^{\prime}-c^{\prime}v). The first pair is actually a solution since

y2(ubv)21(modrqw),y^{2}\equiv(u-b^{\prime}v)^{2}\equiv 1\pmod{rq^{w}},

and

z2(u+cv)21(modrqw).z^{2}\equiv(u+c^{\prime}v)^{2}\equiv 1\pmod{rq^{w}}.

As pp is not a square, the second pair must be a solution as well. This implies

(u+bv)21(modbd).(u+b^{\prime}v)^{2}\equiv 1\pmod{b^{\prime}d}.

However, we have that

(u+bv)2(3bcb+c)2(modqw).(u+b^{\prime}v)^{2}\equiv\left(\frac{3b^{\prime}-c^{\prime}}{b^{\prime}+c^{\prime}}\right)^{2}\pmod{q^{w}}.

As wvq(b)+vq(d)w\geq v_{q}(b^{\prime})+v_{q}(d) we conclude that

(3bcb+c)21(modqvq(b)+h)\left(\frac{3b^{\prime}-c^{\prime}}{b^{\prime}+c^{\prime}}\right)^{2}\equiv 1\pmod{q^{v_{q}(b^{\prime})+h}}

which implies

8(bc)0(modqh).8(b^{\prime}-c^{\prime})\equiv 0\pmod{q^{h}}.

Thus we have qh24(bc)q^{h}\mid 24(b^{\prime}-c^{\prime}) as desired. ∎

Corollary 3.5 (Theorem 1.1).

The only triples of positive integers (a,b,c)(a,b,c) such that fafbfc(mod2)f_{a}\equiv f_{b}f_{c}\pmod{2} are of the form

(2q,4q,4q),(4q,8q,8q),(2q,4q,4q),(4q,8q,8q),

where qq is any positive odd number, or are members of the finite set:

{(4,6,12),(6,8,24),(8,12,24),(10,12,60),(15,24,40),(16,24,48),(20,24,120),(21,24,168)}.\{(4,6,12),(6,8,24),(8,12,24),(10,12,60),(15,24,40),(16,24,48),(20,24,120),(21,24,168)\}.
Proof.

By Proposition 3.1 and Proposition 3.3, we have 1a=1b+1c\frac{1}{a}=\frac{1}{b}+\frac{1}{c}, (b+c)24(b^{\prime}+c^{\prime})\mid 24, and d24(bc)d\mid 24(b^{\prime}-c^{\prime}). Thus we have reduced all cases where bcb^{\prime}\neq c^{\prime} to a finite computation, which we carried out to conclude the sporadic tuples above.

For the case where b=cb^{\prime}=c^{\prime}, since bb^{\prime} and cc^{\prime} are coprime, we have b=c=1b^{\prime}=c^{\prime}=1. The statement fafbfc(mod2)f_{a}\equiv f_{b}f_{c}\pmod{2} now reduces to

|{(y,z):y2+z2=2(d2x+1),d(y21),d(z21)}|1(mod2)\left|\{(y,z):\ y^{2}+z^{2}=2\big{(}\frac{d}{2}x+1\big{)},\ d\mid(y^{2}-1),\ d\mid(z^{2}-1)\}\right|\equiv 1\pmod{2}

if and only if d2x+1\frac{d}{2}x+1 is a square. There is an involution on this set (y,z)(z,y)(y,z)\to(z,y). Thus the parity of this set is equal to the number of fixed point, that is

(3.5) |{y:y2=d2x+1,d(y21)}|\left|\{y:\ y^{2}=\frac{d}{2}x+1,\ d\mid(y^{2}-1)\}\right|
.

If d2x+1\frac{d}{2}x+1 is not a square then 3.5 is obviously zero. If d2x+1\frac{d}{2}x+1 is a square, then 3.5 is non-zero if and only if d2x+1\frac{d}{2}x+1 is equal to 11 modulo dd. This implies that under our assumptions, a square is equal to 11 modulo dd if and only if it is equal to 11 modulo d/2d/2. This is equivalent to v2(d){2,3}v_{2}(d)\in\{2,3\}, giving the tuples (a,b,c)=(2q,4q,4q)(a,b,c)=(2q,4q,4q) and (a,b,c)=(4q,8q,8q)(a,b,c)=(4q,8q,8q). ∎

References

  • [1] C. Ballantine and M. Merca, Parity of sums of partition numbers and squares in arithmetic progressions, Ramanujan J. (2017) 44: 617–630. https://doi.org/10.1007/s11139-016-9845-6
  • [2] H. Weber, Beweis des Satzes, dass jede eigentlich primitive quadratische Form unendlich viele Primzahlen darzustellen fähig ist, Math. Ann. 20 (1882), 301-329.