1 Introduction and Main Results

Let \(C_n\) denote the elliptic curve \(y^2 =x(x-n)(x+3n)\), where n is an integer. The group of rational points of \(C_n\) will be denoted by \(\Gamma _n\). We say that \((x,y)\in C_n\) is a nontrivial rational point of \(C_n\) if xy are nonzero rational numbers; that is, if the order of (xy) as an element of the group \(\Gamma _n\) is greater than two. Our first result shows that the existence of nontrivial rational points of \(C_n\) is closely related to the number of pieces in certain tilings of the equilateral triangle.

Theorem 1.1

For every positive and squarefree integer n the following are equivalent.

  1. (i)

    There is a positive integer k such that the equilateral triangle can be dissected into \(n\cdot k^2\) congruent triangles.

  2. (ii)

    Either \(n\le 3\), or at least one of the curves \(C_n\) and \(C_{-n}\) has a nontrivial rational point.

The proof of Theorem 1.1 is based on the fact that the congruent copies of a triangle with sides abc and corresponding angles \(\alpha ,\beta ,\gamma \) tile an equilateral triangle if and only if either \(\alpha , \beta , \gamma \) are multiples of \(\pi /6\), or \(\gamma \in \{ \pi /3 ,2\pi /3\}\) and abc are pairwise commensurable (see [4, Thm. 3.3]). By the law of cosines, we have \(\gamma =\pi /3\) or \(2\pi /3\) if and only if \(c^2 = a^2 +b^2 \pm ab\). Such triples are, e.g., \((a,b,c)=(7,8,13)\) or \((a,b,c)=(3,5,7)\).

Suppose that abc are positive integers with \(c^2 = a^2 +b^2 \pm ab\). Then the triangle with sides abc tiles an equilateral triangle T. If the side length of T is m and the tiling has N pieces, then, comparing the areas we get \(m^2 =N\cdot ab\), and thus the square free part of N is the same as that of ab. For example, if \((a,b,c)=(7,8,13)\), then the construction described in [3, Thm. 3.1] produces a tiling with \(2,469,600=14\cdot 420^2\) pieces. For the triangle with sides 3, 5, 7, a tiling with \(10,935 = 15\cdot 27^2\) pieces was found by Michael Beeson (see [2, Fig. 22, p. 28]).

As we shall see, a simple transformation maps these triples into nontrivial rational points of one of the corresponding curves \(C_n\) or \(C_{-n}\). Thus the triple (7, 8, 13) gives the point \((-6,48)\) of \(C_{-14}\), and (3, 5, 7) gives the point \((-5,50)\) of \(C_{-15}\).

In the other direction, every nontrivial rational point of \(C_n\) or \(C_{-n}\) determines a triple (abc) as above. For example, from the point \((-1,8)\) of \(C_{-5}\) we obtain the triple (5, 16, 19), and the from the point \((-1,30)\) of \(C_{17}\) we get (17, 225, 217). The proof of Theorem 1.1 will be given in the next section.

Remarks 1.2

  1. 1.

    Since every triangle \(\Delta \) can be dissected into \(m^2\) congruent triangles similar to \(\Delta \) for every m, it is clear that (i) of Theorem 1.1 is equivalent to the following statement.

    (i\('\)) There are infinitely many positive integers k such that the equilateral triangle can be dissected into \(n\cdot k^2\)congruent triangles.

  2. 2.

    We shall prove in Lemma 3.1 that if p is a positive prime, then the only torsion points of \(\Gamma _p\) and \(\Gamma _{-p}\) are the points having zero y-coordinates. Therefore, if n is a positive prime, then (ii) of Theorem 1.1 is equivalent to the following statement.

    (ii\('\)) Either \(n\le 3\), or at least one of the groups \(\Gamma _n\)and \(\Gamma _{-n}\)has positive rank.

It is easy to see that if nk are nonzero integers then \(C_n\) has a nontrivial rational point if and only if \(C_{nk^2}\) has one. Therefore, we have the following corollary of Theorem 1.1.

Corollary 1.3

If the equilateral triangle can be dissected into N congruent triangles, then either \(N=k^2\), \(N=2k^2\) or \(N=3k^2\) for some k, or at least one of the curves \(C_N\) and \(C_{-N}\) has a nontrivial rational point.

We remark that the converse is not true. For example, \((-1,8)\) is a nontrivial rational point of \(C_{-5}\), but the equilateral triangle cannot be dissected into 5 congruent triangles. This follows from a result of Beeson stating that the equilateral triangle cannot be dissected into p congruent triangles for any prime \(p>3\) (see [1]). On the other hand, the equilateral triangle can be dissected into \(5k^2\) congruent triangles for infinitely many positive integer k by Theorem 1.1.

In Sect. 3 we shall prove that if p is a positive prime and \(p\equiv 7\) (mod 24), then the curves \(C_p\) and \(C_{-p}\) have no nontrivial rational points (see Corollary 3.6). Comparing with Theorem 1.1 we obtain the following.

Corollary 1.4

If p is a positive prime such that \(p\equiv 7\) (mod 24), then the equilateral triangle cannot be dissected into \(p\cdot k^2\) congruent triangles for any k.

2 Proof of Theorem 1.1

(i) \(\Rightarrow \) (ii): Suppose that the equilateral triangle T can be tiled with \(n\cdot k^2\) congruent triangles having angles \(\alpha ,\beta ,\gamma \) and corresponding sides abc. We may assume that the sides of T equal 1.

By [4, Thm. 3.3], one of the following cases holds: \(\alpha =\beta =\pi /6\) and \(\gamma =2\pi /3\); \(\alpha =\pi /6 ,\)\(\beta =\pi /2,\)\(\gamma =\pi /3\); \(\gamma \in \{ \pi /3 ,2\pi /3\}\) and abc are pairwise commensurable.

Comparing the areas of T and the tiles we obtain \(nk^2 \cdot ab\cdot \frac{\sqrt{3}}{4}=\frac{\sqrt{3}}{4}\); that is,

$$\begin{aligned} nk^2 \cdot ab=1. \end{aligned}$$
(1)

If \(\alpha =\beta =\pi /6\), then \(a=b\) and thus, by (1), \(a=b=1/(k\cdot \sqrt{n} )\). By \(c/a=\sqrt{3}\) we have \(c=\sqrt{3} /(k\cdot \sqrt{n} )\). Since the side of the equilateral triangle is tiled with segments of length a and c, we obtain \(1=ra+sc\) with suitable nonnegative integers rs. Thus \(r+s\sqrt{3} =k\cdot \sqrt{n}\). Since n is squarefree, this implies \(n=1\) or \(n=3\).

If \(\alpha =\pi /6\), \(\beta =\pi /2\) and \(\gamma =\pi /3\), then \(b=2a\) and thus, by (1), \(a=1/(k\cdot \sqrt{2n})\). By \(c/a=\sqrt{3}\) we have \(c=\sqrt{3} /(k\cdot \sqrt{2n} )\). The side of the equilateral triangle is tiled with segments of length a, 2a and c, hence \(1=ra+sc\) with suitable nonnegative integers rs. Thus \(r+s\sqrt{3} =k\cdot \sqrt{2n}\). Since n is squarefree, this implies \(n=2\) or \(n=6\). Now (9, 27) is a point of \(C_6 :y^2 =x(x-6)(x+18)\), and thus the statement of (ii) is true in these cases.

In the remaining cases abc are pairwise commensurable, and \(\gamma =\pi /3\) or \(\gamma =2\pi /3\). Then we have \(c^2=a^2 +b^2 \pm ab\) by the law of cosines. Since \(qa+rb+sc=1\) with nonnegative integers qrs, it follows that abc are rational. Replacing a by \(-a\) if necessary, we may assume \(c^2=a^2 +b^2 +ab\). Under this change (1) becomes \(\pm nk^2 \cdot ab=1\). We put \(t=(c-b)/a\); then t is rational, and \(b=c-ta\). We have

$$\begin{aligned} c^2= & {} a^2 +b^2 +ab=a^2 +(c-ta)^2 +ac-ta^2 \\= & {} a^2 (t^2 -t+1) -2act + ac+c^2 , \end{aligned}$$

\(a^2 (t^2 -t+1) =ac(2t-1)\), and \(a/c=(2t-1)/d\), where \(d=t^2 -t+1\). Note that \(d\ne 0\), as the polynomial \(X^2 -X+1\) has no rational roots. Then we have \(b/c=1-(ta/c)=(1-t^2)/d\). From (1) we get

$$\begin{aligned} 1=\pm nk^2 ab=\pm n \cdot (2t-1)(1-t^2 )\cdot (ck/d)^2 \end{aligned}$$

and \((2t-1)(t^2 -1)=\mp nv^2\), where \(v=d/(nkc)\) is a nonzero rational number.

Putting \(x=n(2t-1)\) we get \(t=(x+n)/(2n)\), \(t-1=(x-n)/(2n)\), \(t+1=(x+3n)/(2n)\), and

$$\begin{aligned} x(x-n)(x+3n)= (2t-1)(t^2 -1) \cdot 4n^3 = \mp nv^2 \cdot 4n^3 = \mp y^2, \end{aligned}$$

where \(y=2n^2 v\). Therefore, either (xy) is a point of \(C_{n}\) or \((-x,y)\) is a point of \(C_{-n}\).

(ii) \(\Rightarrow \) (i): It is clear that if \(n\le 3\) then the equilateral triangle can be dissected into n congruent triangles.

Suppose that xy are rational numbers, \(y\ne 0\), and (xy) is a rational point of either \(C_n\) or \(C_{-n}\). Then one of \(t=x/n\) and \(t=-x/n\) satisfies \(t(t+1)(t-3)=\pm y^2 /n^3\). Fix such a t. Note that \(t\ne 0, -1,3\). Putting \(a=4t\), \(b=t^2 -2t-3\) and \(c=t^2 +3\) we have \(ab\ne 0\) and \(a^2 +b^2 +ab=c^2\). Then |a|, |b|, c are the sides of a rational triangle \(\Delta \) such that \(a^2 +b^2 \pm |a|\cdot |b|=c^2\), and thus, by the law of cosines, the angle between the sides of length |a| and |b| equals \(\pi /3\) or \(2\pi /3\). By [3, Thm. 3.1], there is an equilateral triangle T that can be dissected into triangles congruent to \(\Delta \). Let m be the length of the side of T, and let N be the number of pieces of the decomposition. Then \(N|ab|=m^2\), hence

$$\begin{aligned} m^2 /N= |ab|=4|t(t^2 -2t-3)|=4|t(t+1)(t-3)|=4y^2 /n^3 \end{aligned}$$

and \(N=n^3 m^2 /(4y^2 ) =nk^2\), where \(k=nm/(2y)\). Now k is rational and n is squarefree by assumption, so \(N=nk^2\) implies that k must be an integer. We have found a dissection of T into \(n\cdot k^2\) congruent triangles, proving (i). \(\square \)

3 Rational Points of \(C_{\pm p}\)

In this section we show that if p is a positive prime and \(p\equiv 7\) (mod 24), then \(C_p\) and \(C_{-p}\) have no nontrivial rational points (see Corollary 3.6). Recall that the group of rational points of \(C_n\) is denoted by \(\Gamma _n\).

Lemma 3.1

Let p be a positive prime. Then the torsion points of the group \(\Gamma _p\) are the points (0, 0), (p, 0), \((-3p ,0)\) and \({{\mathcal {O}}}\) (the point at infinity). The torsion points of \(\Gamma _{-p}\) are the points (0, 0), \((-p,0)\), (3p, 0) and \({{\mathcal {O}}}\).

Proof

The points listed above, being of order two and one, are torsion points. Suppose there exists another torsion point (xy). Since the discriminant of the curves equals \(p^2 \cdot (3p)^2 \cdot (4p)^2 =3^2 \cdot 2^4 \cdot p^6\), it follows from the Nagell–Lutz theorem that \(x,y\in {{\mathbb {Z}}}\), \(y\ne 0\) and \(y \mid 3 \cdot 2^2 \cdot p^3\). We distinguish between two cases.

Case I: \(p\mid y\). Then \(p\mid x\), \(x=pz\), \(p^2 \mid y\), \(y=p^2 u\), \(u\ne 0\), and

$$\begin{aligned} pu^2 =z(z\mp 1)(z\pm 3). \end{aligned}$$
(2)

Clearly, \(z\ge -2\). It is easy to check that if \(-2 \le z\le 13\) then \(z(z\mp 1)(z\pm 3)\) is not of the form \(qu^2\), where q is prime and \(u\ne 0\), except when \(z=4\) and \(z(z+ 1)(z- 3)=5\cdot 2^2\). This gives the point \(P_1 =(20,50)\) of \(\Gamma _{-5}\). One can easily check that the x-coordinate of \(2P_1\) is not an integer, hence \(P_1\) is not a torsion point. (Thus \(\Gamma _{-5}\) has positive rank.) Therefore, we may assume \(z\ge 14\).

If \(p=2\) or \(p=3\) then \(y=p^2 u\mid 3 \cdot 2^2 \cdot p^3\) implies that all prime factors of z and \(z\pm 1\) are 2 and 3. Thus \(z=2^\alpha \), \(z\pm 1=3^\beta \) or the other way around. Then \(z\le 10\) which is impossible.

Therefore, we may assume \(p>3\). Then at most one of the terms z, \(z\mp 1\), \(z\pm 3\) is divisible by p. Since \(u\mid 3\cdot 2^2 \cdot p^3\), it follows from (2) that the product of two of the terms z, \(z\mp 1\)\(z\pm 3\) is a divisor of \(3^2 \cdot 2^4 = 144\). By \(z\ge 4\) this implies \(z(z-3)\le 144\), hence \(z\le 13\) which is impossible.

Case II: \(p\not \mid y\). Then \(y\mid 12\). Replacing x by \(-x\) if necessary, we have \(x(x+p)(x-3p)=\pm y^2\), and thus

$$\begin{aligned} |x(x+p)(x-3p)|=y^2 \mid 144. \end{aligned}$$
(3)

It is easy to see that if a is a positive integer and x is an integer different from 0 and a, then \(|x(a-x)|\ge a-1\). Therefore, \(|x(x+p)|\ge p-1\), \(|x(x-3p)|\ge 3p-1\), \(|(x+p)(x-3p)|\ge 4p -1\),

$$\begin{aligned} (p-1)(3p-1)(4p-1)\le |x(x+p)(x-3p)|^2 \le 144 ^2 , \end{aligned}$$

and thus \(p\le 11\).

It follows from (3) that there are (positive or negative) divisors \(d_1 ,d_2\) of 144 such that \(d_2 -d_1 =4p\), \(|x\cdot d_1 \cdot d_2 |\) is a square and is a divisor of 144, where \(x=d_2 -p\). Checking the cases \(p=2,3,5,7,11\), we find that the only possibility is \(p=5\), \((d_1 , d_2 )=(-16,4)\) and \(x=-1\). This gives the point \(P_2 =(-1,8)\) of \(\Gamma _{-5}\). One can easily check that \(P_2 =P_1 +P_0\), where \(P_0 =(-5,0)\) and \(P_1 =(20,50)\). Since \(P_0\) is a torsion point of \(\Gamma _{-5}\) and \(P_1\) is not, it follows that \(P_2\) is not a torsion point either.

\(\square \)

Theorem 3.2

 

  1. (i)

    The rank of \(\Gamma _p\) is at most two for every positive prime p.

  2. (ii)

    If \(p\not \equiv 1\) (mod 24), then the rank of \(\Gamma _p\) is at most one.

  3. (iii)

    If \(p=2\), \(p=3\) or \(p\equiv 5\), 7 or 19 (mod 24), then the rank of \(\Gamma _p\) is zero.

In the proof of Theorem 3.2 we apply the method described in [5, §5, Chap. III, pp. 92–94]. Consider the curves

$$\begin{aligned} C_p : y^2 =x^3 +2px^2 -3p^2 x \quad \text {and} \quad \overline{C} _p : y^2 =x^3 -4px^2 +16p^2 x \end{aligned}$$

with groups of rational points \(\Gamma _p =C_p ({{\mathbb {Q}}})\) and \(\overline{\Gamma }_p =\overline{C} _p ({{\mathbb {Q}}})\). We define \(\alpha :\Gamma _p \rightarrow {{\mathbb {Q}}}^* /{{\mathbb {Q}}}^{* 2}\) by \(\alpha ({{{\mathcal {O}}}}) =1\), \(\alpha (0,0)=-3p^2 \equiv -3\) and, for \(x\ne 0\), \(\alpha (x,y)=x\) (mod \({{\mathbb {Q}}}^{* 2}\)). Then \(\alpha \) is a homomorphism from \(\Gamma _p\) into \({{\mathbb {Q}}}^* /{{\mathbb {Q}}}^{* 2}\).

We also define \(\overline{\alpha }:\overline{\Gamma }_p \rightarrow {{\mathbb {Q}}}^* /{{\mathbb {Q}}}^{* 2}\) by \(\overline{\alpha }({{{\mathcal {O}}}}) =1\), \(\overline{\alpha }(0,0)=16p^2 \equiv 1\) and, for \(x\ne 0\), \(\alpha (x,y)=x\) (mod \({{\mathbb {Q}}}^{* 2}\)). Then \(\overline{\alpha }\) is a homomorphism from \(\overline{\Gamma }_p\) into \({{\mathbb {Q}}}^* /{{\mathbb {Q}}}^{* 2}\). The rank r of \(\Gamma _p\) satisfies

$$\begin{aligned} 2^r = \frac{\#\, \alpha (\Gamma _p )\cdot \#\, \overline{\alpha }( \overline{\Gamma }_p )}{4} \end{aligned}$$
(4)

(see [5, p. 91]). Here \(\alpha (\Gamma _p )\) equals the set of divisors \(b_1\) of \(b=-3p^2\) (mod \({{\mathbb {Q}}}^{* 2}\)) such that the equation

$$\begin{aligned} N^2 =b_1 M^4 +2pM^2 e^2 +(-3p^2 /b_1 )\,e^4 \end{aligned}$$
(5)

is solvable in pairwise coprime integers NMe satisfying \(M\ne 0\) and \(\gcd (e,b_1 )=\gcd (M,-3p^2 /b_1 )=1\) (see [5, pp. 92–93]). Similarly, \(\alpha (\overline{\Gamma }_p )\) equals the set of divisors \(b_1\) of \(\overline{b} =16p^2\) (mod \({{\mathbb {Q}}}^{* 2}\)) such that the equation

$$\begin{aligned} N^2 =b_1 M^4 -4 pM^2 e^2 +(16p^2 /b_1 )\,e^4 \end{aligned}$$
(6)

is solvable in pairwise coprime integers NMe satisfying \(M\ne 0\) and \(\gcd (e,b_1 )=\gcd (M, 16p^2 /b_1 )=1\).

The statement of Theorem 3.2 is an immediate consequence of (4) and of the following lemma.

Lemma 3.3

 

  1. (i)

    \(\#\, \alpha (\Gamma _p ) \le 8\) for every positive prime p.

  2. (ii)

    If \(p=2\), \(p=3\) or \(p\equiv 5\), 7, 13 or 19 (mod 24), then \(\#\, \alpha (\Gamma _p ) \le 4\).

  3. (iii)

    \(\#\, \alpha (\overline{\Gamma }_p ) \le 2\) for every positive prime p.

  4. (iv)

    If \(p\not \equiv 1\) (mod 12), then \(\#\, \alpha (\overline{\Gamma }_p ) =1\).

Proof

 

  1. (i)

    is obvious from \(b_1 \in \{ \pm 1, \pm 3, \pm p, \pm 3p\}\) (mod \({{\mathbb {Q}}}^{* 2}\)).

  2. (ii)

    If \(p=3\) then \(b_1 \in \{ \pm 1, \pm 3\}\) (mod \({{\mathbb {Q}}}^{* 2}\)), and \(\#\, \alpha (\Gamma _p ) \le 4\). Therefore, we may assume \(p\ne 3\). We have \((p,0), (-3p,0)\in \Gamma _p\) and \(\alpha (0,0)=-3p^2 \equiv -3\), and thus \(1,p, -3,-3p\in \alpha (\Gamma _p )\). Since \(\alpha (\Gamma _p )\) is a subgroup of \({{\mathbb {Q}}}^* /{{\mathbb {Q}}}^{* 2}\), it follows that \(\#\, \alpha (\Gamma _p )\) equals 4 or 8, and it equals 8 if and only if \(-1\in \alpha (\Gamma _p )\).

Suppose that \(\#\, \alpha (\Gamma _p )=8\). Then \(-1\in \alpha (\Gamma _p )\) and thus, by \(b_1 \mid 3p^2\), (5) is solvable for at least one of \(b_1 =-1\) and \(b_1 =-p^2\).

Suppose that \(N^2 =-M^4 +2pM^2 e^2 +3p^2 e^4\) is solvable. If \(p=2\), then M is odd by \(\gcd (M, 3p^2 )=1\), and \(N^2 \equiv -M^4\) (mod 4), which is impossible. If \(p>3\), then \(p\not \mid M\) by \(\gcd (M, 3p^2 )=1\), and thus we have \(\big (\frac{-1}{p}\big )=1\) and \(p\equiv 1\) (mod 4).

We have \(N^2 =(3pe^2 -M^2 )(pe^2 +M^2 )=A\cdot B\). Since \(p\not \mid M\) and \(\gcd (M,e)=1\), it follows that \(\gcd (A,B) \mid 4\). If \(\gcd (A,B)=1\) or 4, then A and B are squares. Thus \(3pe^2 -M^2 =n^2\), hence \(-M^2 \equiv n^2\) (mod 3), which is impossible, as \(3\not \mid M\).

If \(\gcd (A,B)=2\), then A / 2 and B / 2 are squares. Thus \(3pe^2 -M^2 =2n^2\), hence \(-M^2 \equiv 2n^2\) (mod p). Since \(p\not \mid M\) and \(p\equiv 1\) (mod 4), we get \(\big (\frac{2}{p}\big )=1\) and \(p\equiv 1\) (mod 8).

Next suppose that \(N^2 =-p^2 M^4 +2pM^2 e^2 +3 e^4\) is solvable. Then we have \(\gcd (M,3)=1\). If \(p=2\), then e is odd (since otherwise both N and e would be even), and \(N^2 \equiv 3e^4\) (mod 4), which is impossible. Suppose \(p>3\). Then \(p\not \mid e\) (since otherwise both e and N would be divisible by p), and thus \(\big (\frac{3}{p}\big )=1\) and \(p\equiv \pm 1\) (mod 12).

We have \(N^2 =(3e^2 -pM^2 )(e^2 +pM^2 )=C\cdot D\). Since \(p\not \mid e\) and \(\gcd (M,e)=1\), it follows that \(\gcd (C,D) \mid 4\). If \(\gcd (C,D)=1\) or 4, then C and D are squares. Thus \(3e^2 -pM^2 =n^2\), \(-pM^2 \equiv n^2\) (mod 3), \(p\equiv -1\) (mod 3) and \(p\equiv - 1\) (mod 12).

If \(\gcd (C,D)=2\), then C / 2 and D / 2 are squares. Thus \(e^2 + pM^2 =2n^2\), hence \(e^2 \equiv 2n^2\) (mod p), \(\big (\frac{2}{p}\big )=1\), \(p\equiv \pm 1\) (mod 8).

We proved that if \(\#\, \alpha (\Gamma _p )=8\), then \(p>3\) and either \(p\equiv 1\) (mod 8), or \(p\equiv - 1\) (mod 12). This proves (ii).

(iii) We have to estimate \(\#\, \alpha (\overline{\Gamma }_p )\). It is clear that if \(b_1 <0\) then (6) has no solutions, and thus, by \(b_1 \mid 16p^2\), we have \(b_1 \in \{ 2^\alpha p^\beta :0\le \alpha \le 4, 0\le \beta \le 2\}\). If \(p=2\), then we obtain \(\alpha (\overline{\Gamma }_p )\subset \{ 1,2\}\) (mod \({{\mathbb {Q}}}^{*2}\)). Therefore, we may assume \(p>2\).

Let \(b_1 =2p^\beta \), and suppose that (6) is solvable. Then M is odd by \(\gcd (M,16p^2 /b_1 )=1\), and thus the left hand side of (6) is divisible by 4, while the right hand side is not, which is impossible.

Next let \(b_1 =8p^\beta \), and suppose that (6) is solvable. Then N is even and, consequently, e is odd. Thus the left hand side of (6) is divisible by 4, while the right hand side is not, which is impossible. We obtain that \(b_1 \in \{ 1,p,p^2, 4, 4p, 4p^2 ,16, 16p, 16p^2\}\) and \(b_1 \in \{ 1,p\}\) (mod \({{\mathbb {Q}}}^{*2}\)). This proves (iii).

(iv) Suppose that \(\#\, \alpha (\overline{\Gamma }_p ) =2\). Then \(p\in \alpha (\overline{\Gamma }_p )\), and (6) is solvable for at least one of \(b_1 =p\), \(b_1 =4p\) and \(b_1 =16p\).

Let \(b_1 =p\), and suppose that \(N^2 =p M^4 -4 pM^2 e^2 +16pe^4\) is solvable. Then M is odd by \(\gcd (M,16p)=1\), and \(N^2 \equiv pM^4\) (mod 4). Hence \(p>2\) and \(p\equiv 1\) (mod 4). We have \(N=pN_1\) and

$$\begin{aligned} pN_1^2 =M^4 -4M^2 e^2 +16e^4 =(M^2 -2e^2 )^2 +12e^4 . \end{aligned}$$

Now \(p\not \mid e\) by \(\gcd (e,b_1 )=1\), and we get \(\big (\frac{-12}{p}\big )=1\). Since \(p\equiv 1\) (mod 4), we obtain \(\big (\frac{3}{p}\big )=1\), \(p\equiv \pm 1\) (mod 12) and \(p\equiv 1\) (mod 12).

The case \(b_1 =16p\) is similar with the roles of M and e exchanged. Therefore, if (6) is solvable for \(b_1 =16p\), then \(p\equiv 1\) (mod 12).

Finally, let \(b_1 =4p\), and suppose that \(N^2 =4p M^4 -4 pM^2 e^2 +4pe^4\) is solvable. Then \(2\not \mid M\) by \(\gcd (M,4p)=1\), and \(2p\mid N\). Let \(N=2pN_1\), then \(pN_1 ^2 =M^4 -M^2 e^2 +e^4\). Since M is odd, we have \(M^4 -M^2 e^2 +e^4 \equiv 1\) (mod 4), and thus \(p\equiv 1\) (mod 4). We have

$$\begin{aligned} 4pN_1^2 =4M^4 -4M^2 e^2 +4e^4 =(2M^2 -e^2 )^2 +3e^4 . \end{aligned}$$

Now \(p\not \mid e\) by \(\gcd (e,b_1 )=1\), and we get \(\big (\frac{-3}{p}\big )=1\). Since \(p\equiv 1\) (mod 4), we obtain \(\big (\frac{3}{p}\big )=1\), \(p\equiv \pm 1\) (mod 12) and \(p\equiv 1\) (mod 12).

We proved that if \(\#\, \alpha (\Gamma _p )=2\), then \(p\equiv 1\) (mod 12). This proves (iv). \(\square \)

Our next aim is to prove

Theorem 3.4

 

  1. (i)

    The rank of \(\Gamma _{-p}\) is at most two for every positive prime p.

  2. (ii)

    If \(p\not \equiv 1\) (mod 12), then the rank of \(\Gamma _{-p}\) is at most one.

  3. (iii)

    If \(p=2\), \(p=3\) or \(p\equiv 7\) (mod 24), then the rank of \(\Gamma _{-p}\) is zero.

We consider the curves

$$\begin{aligned} C_{-p} : y^2 =x^3 -2px^2 -3p^2 x \quad \text {and} \quad \overline{C} _{-p} : y^2 =x^3 +4px^2 +16p^2 x. \end{aligned}$$

First we prove the following lemma.

Lemma 3.5

 

  1. (i)

    \(\#\, \alpha (\Gamma _{-p} ) \le 8\) for every prime p.

  2. (ii)

    If \(p=2\), \(p=3\) or \(p\equiv 7\) (mod 12), then \(\#\, \alpha (\Gamma _{-p} ) \le 4\).

  3. (iii)

    \(\#\, \alpha (\overline{\Gamma }_{-p} ) \le 2\) for every prime p.

  4. (iv)

    If \(p\ne 3\) and \(p\not \equiv 1\), 13 or 19 (mod 24), then \(\#\, \alpha (\overline{\Gamma }_{-p} ) =1\).

Proof

The proof of the statement (i) is the same as in the case of Lemma 3.3.

(ii) Suppose \(\#\, \alpha (\Gamma _{-p} ) =8\). As in the proof of (ii) of Lemma 3.3, this implies \(p\ne 3\) and \(-1\in \alpha (\Gamma _{-p} )\). Therefore, by \(b_1 \mid -3p^2\), \(N^2 =b_1 M^4 -2pM^2 e^2 +(-3p^2 /b_1 )e^4\) is solvable for at least one of \(b_1 =-1\) and \(b_1 =-p^2\).

Suppose that \(N^2 =-M^4 -2pM^2 e^2 +3p^2 e^4\) is solvable. If \(p=2\), then M is odd by \(\gcd (M, 3p^2 )=1\), and \(N^2 \equiv -M^4\) (mod 4), which is impossible. If \(p>3\), then \(p\not \mid M\) by \(\gcd (M, 3p^2 )=1\), and thus we have \(\big (\frac{-1}{p}\big )=1\) and \(p\equiv 1\) (mod 4).

Next suppose that \(N^2 =-p^2 M^4 -2pM^2 e^2 +3 e^4\) is solvable; then \(\gcd (M,3)=1\). If \(p=2\), then e is odd (since otherwise both N and e would be even), and \(N^2 \equiv 3e^4\) (mod 4), which is impossible. Suppose \(p>3\). Then \(p\not \mid e\) by \(\gcd (e, b_1 )=1\), and thus \(\big (\frac{3}{p}\big )=1\) and \(p\equiv \pm 1\) (mod 12).

We have \(N^2 =(3e^2 +pM^2 )(e^2 -pM^2 )=C\cdot D\). Since \(p\not \mid e\) and \(\gcd (M,e)=1\), it follows that \(\gcd (C,D) \mid 4\). If \(\gcd (C,D)=1\) or 4, then C and D are squares. Thus \(3e^2 +pM^2 =n^2\), \(pM^2 \equiv n^2\) (mod 3), \(p\equiv 1\) (mod 3) and \(p\equiv 1\) (mod 12).

If \(\gcd (C,D)=2\), then C / 2 and D / 2 are squares. Thus \(3e^2 +pM^2 =2n^2\), hence \(p\equiv pM^2 \equiv 2n^2 \equiv 2\) (mod 3). Since \(p\equiv \pm 1\) (mod 12), we get \(p\equiv -1\) (mod 12).

We proved that if \(\#\, \alpha (\Gamma _{-p} )=8\), then \(p\equiv 1\) (mod 4) or \(p\equiv -1\) (mod 12). This proves (ii).

(iii) The argument proving (iii) of Lemma 3.3 shows that \(\alpha (\overline{\Gamma }_{-p} ) \subset \{ 1,p\}\) (mod \({{\mathbb {Q}}}^{*2}\)).

(iv) Suppose \(\#\, \alpha (\overline{\Gamma }_{-p} ) =2\). Then \(p\in \alpha (\overline{\Gamma }_{-p} )\), and

$$\begin{aligned} N^2 =b_1 M^4 +4pM^2 e^2 +(16p^2 /b_1 )\,e^4 \end{aligned}$$

is solvable for at least one of \(b_1 =p\), \(b_1 =4p\) and \(b_1 =16p\).

Let \(b_1 =p\), and suppose that \(N^2 =p M^4 +4 pM^2 e^2 +16pe^4\) is solvable. Then M is odd by \(\gcd (M,16p)=1\), and \(N^2 \equiv pM^4\) (mod 4). Hence \(p>2\) and \(p\equiv 1\) (mod 4). We have \(N=pN_1\) and

$$\begin{aligned} pN_1^2 =M^4 +4M^2 e^2 +16e^4 =(M^2 +2e^2 )^2 +12e^4 . \end{aligned}$$

Now \(p\not \mid e\) by \(\gcd (e,b_1 )=1\), and we get \(\big (\frac{-12}{p}\big )=1\). Since \(p\equiv 1\) (mod 4), we obtain \(\big (\frac{3}{p}\big )=1\), \(p\equiv \pm 1\) (mod 12) and \(p\equiv 1\) (mod 12).

The case \(b_1 =16p\) is similar with the roles of M and e exchanged. Therefore, if (6) is solvable for \(b_1 =16p\), then \(p\equiv 1\) (mod 12).

Finally, let \(b_1 =4p\), and suppose that \(N^2 =4p M^4 +4 pM^2 e^2 +4pe^4\) is solvable. Then M is odd by \(\gcd (M,4p)=1\). Also, \(2p\mid N\), and thus e is odd. Let \(N=2pN_1\), then \(pN_1 ^2 =M^4 +M^2 e^2 +e^4\). Thus \(pN_1^2 \equiv 3\) (mod 8), hence \(p\equiv 3\) (mod 8). We have

$$\begin{aligned} 4pN_1^2 =4M^4 +4M^2 e^2 +4e^4 =(2M^2 +e^2 )^2 +3e^4 . \end{aligned}$$

Now \(p\not \mid e\) by \(\gcd (e,b_1 )=1\), and we get \(p=3\) or \(\big (\frac{-3}{p}\big )=1\). Suppose \(p\ne 3\). Since \(p\equiv 3\) (mod 4), we obtain \(\big (\frac{3}{p}\big )=-1\), \(p\equiv 5\) or 7 (mod 12). Since \(p\equiv 3\) (mod 8), we get \(p\equiv 19\) (mod 24).

We proved that if \(\#\, \alpha ({\overline{\Gamma }}_{-p} )=2\), then \(p=3\) or \(p\equiv 1\) (mod 12) or \(p\equiv 19\) (mod 24), This proves (iv). \(\square \)

Proof of Theorem 3.4

Statements (i) and (ii) of the theorem follow from Lemma 3.5 and from (4). If \(p=2\) or \(p\equiv 7\) (mod 24), then the rank of \(\Gamma _{-p}\) is zero by Lemma 3.5 and (4).

What remains to prove is that the rank of \(\Gamma _{-3}\) is zero. Since \(\#\, \alpha (\overline{\Gamma }_{-3} ) \le 2\) by Lemma 3.5, it is enough to show that \(\#\, \alpha (\Gamma _{-3} ) \le 2\).

Consider the curve \(C_{-3} :y^2 =x^3 -6x^2 -27x\). Then \(b_1 \in \{ \pm 1, \pm 3, \pm 9, \pm 27\}\), and thus \(\alpha (\Gamma _{-3} ) \subset \{ \pm 1, \pm 3\}\) (mod \({{\mathbb {Q}}}^{* 2}\)). We show that \(3 \notin \alpha (\Gamma _{-3} )\). Suppose \(3 \in \alpha (\Gamma _{-3} )\). Then the equation \(N^2 =b_1 M^4 -6M^2 e^2 -(27/b_1 )e^4\) is solvable for at least one of \(b_1 =3\) and \(b_1 =27\).

Suppose that \(N^2 =3M^4 -6M^2 e^2 -9e^4\) is solvable. Then \(3\not \mid M\) by \(\gcd (M,9) =1\), and \(3\not \mid e\) since \(3\mid N\). Let \(N=3N_1\). Then \(3N_1^2 =M^4 -2M^2 e^2 -3e^4\), hence \(M^4 \equiv 2M^2 e^2\) (mod 3), which is impossible.

Finally, suppose that \(N^2 =27M^4 -6M^2 e^2 -e^4\) is solvable. Then \(3\not \mid e\) by \(\gcd (e,b_1 )=1\). Thus \(N^2 \equiv - e^2\) (mod 3), which is impossible. \(\square \)

Corollary 3.6

If \(p=2\), \(p=3\) or \(p\equiv 7\) (mod 24), then the curves \(C_p\) and \(C_{-p}\) have no nontrivial rational points. \(\square \)

4 Numerical Examples

As the following table shows, for all primes \(3<p<100\), if \(p\not \equiv 7\) (mod 24), then at least one of the curves \(C_p\) and \(C_{-p}\) has nontrivial rational points and, consequently, \(\Gamma _p\) or \(\Gamma _{-p}\) has positive rank. Note that the point (75, 210) belongs to both \(C_{-23}\) and \(C_{73}\).

The points below were found by searching for integer solutions of \(N^2 =b_1 M^4 \pm 2p M^2 e^2 +b_2 e^4\) with \(b_1 b_2 =-3p^2\), and putting \(x=b_1 M^2 /e^2\), \(y=b_1 MN/e^3\). The solutions for \(p \ne 83\) were found by using GNU Octave (https://www.gnu.org/software/octave/). I am grateful to Peter Salvi for finding a solution for \(p=83\); he used Julia 1.0 (https://julialang.org/blog/2018/08/one-point-zero).

$$\begin{aligned} p= & {} 5: (-1, 8)\in \Gamma _{-5}, \\ p= & {} 11: (75, 720)\in \Gamma _{11},\\ p= & {} 13: (-12, 90)\in \Gamma _{13},\\ p= & {} 17: (-1,30)\in \Gamma _{17},\\ p= & {} 19: \Big ( \displaystyle {\frac{17689}{225}, \frac{1374688}{3375}}\Big )\in \Gamma _{-19},\\ p= & {} 23: (75, 210)\in \Gamma _{-23},\\ p= & {} 29: \Big ( \displaystyle {-\frac{529}{25}, \frac{16744}{125}}\Big ) \in \Gamma _{-29}, \\ p= & {} 37: \Big ( \displaystyle {\frac{231361}{324}, \frac{116481365}{5832}} \Big )\in \Gamma _{37}, \\ p= & {} 41: ( -121, 198)\in \Gamma _{41}, \\ p= & {} 43: \Big ( \displaystyle {\frac{4165798849}{21538881}, \frac{171543655606240}{99961946721}} \Big )\in \Gamma _{-43}, \\ p= & {} 47: (1875, 79050)\in \Gamma _{-47}, \\ p= & {} 53: \Big ( -\displaystyle {\frac{167281}{4225}, \frac{89165272}{274625}} \Big ) \in \Gamma _{-53}, \\ p= & {} 59: \Big ( -\displaystyle {\frac{930433009}{6076225}, \frac{13189530387264}{14977894625}} \Big )\in \Gamma _{59}, \\ p= & {} 61: ( -108, 1170)\in \Gamma _{61}, \\ p= & {} 67: \Big ( \displaystyle {\frac{909373939321}{51279921}, \frac{863887766632341760}{367215514281} } \Big )\in \Gamma _{-67}, \\ p= & {} 71: (507, 9282)\in \Gamma _{-71}, \\ p= & {} 73: (75, 210)\in \Gamma _{73}, \\ p= & {} 83: \Big ( -\displaystyle {\frac{2140232721200}{59682001401}, \frac{13897116923228469980}{14580253260262899}} \Big )\in \Gamma _{83}, \\ p= & {} 89: \Big ( -\displaystyle {\frac{121}{289}, \frac{489280}{4913}} \Big )\in \Gamma _{-89}, \\ p= & {} 97: \Big ( -\displaystyle {\frac{121}{25}, \frac{45408}{125}} \Big )\in \Gamma _{-97}. \\ \end{aligned}$$