1 Introduction

In the following, \(p\) is a prime number, \(n\) a positive integer, \(q=p^n\) and \(\mathbb {F}_q\) is a finite field with \(q\) elements.

To resist differential cryptanalysis, a function \(\phi \) from \(\mathbb {F}_q\) to \(\mathbb {F}_q\) used in a bloc cypher like DES has to have a low uniformity, that is to say for all \(a\in \mathbb {F}_q^*\) and \(b\in \mathbb {F}_q\), the equation \(\phi (x+a)-\phi (x)=b\) must have few solutions.

In characteristic 2, if \(\phi (x+a)+\phi (x)=b\) then \(\phi (x+a+a)+\phi (x+a)=b\). So, the functions which resist differential cryptanalysis the most are the following

Definition 1

We say that the function \(\phi :\mathbb {F}_q\rightarrow \mathbb {F}_q\) is almost perfectly nonlinear (APN) over \(\mathbb {F}_{q}\) if:

$$\begin{aligned} \forall a,b\in \mathbb {F}_{q},a\ne 0,|\{x\in \mathbb {F}_{q},\phi (x+a)-\phi (x)=b\}|\le 2 \end{aligned}$$

and if, furthermore, there exists a pair \((a,b)\) such that we have equality.

On the contrary, in odd characteristic, \(x\mapsto \phi (x+a)-\phi (x)\) can be one to one. So, the functions which resist differential cryptanalysis the most are the following:

Definition 2

If \(q\) is odd, a function \(\phi :\mathbb {F}_q\rightarrow \mathbb {F}_q\) is perfectly nonlinear (PN) over \(\mathbb {F}_{q}\) if for all \(b\in \mathbb {F}_{q}\) and all \(a\in \mathbb {F}_{q}^*\)

$$\begin{aligned} |\{x\in \mathbb {F}_{q},\phi (x+a)-\phi (x)=b\}|=1. \end{aligned}$$

In [5, 8], Jedlicka, Hernando and McGuire are interested in integers \(m\) such that the function \(x\mapsto x^m\) is APN on infinitely many extensions of \(\mathbb {F}_2\). They prove that the only integers \(m\) such that \(x\mapsto x^m\) is APN on infinitely many extensions of \(\mathbb {F}_2\) are \(m=2^k+1\) (Gold) and \(m=4^k-2^k+1\) (Kasami). They use the fact that a function \(x\mapsto x^m\) is APN over \(\mathbb {F}_{2^n}\) if and only if the rational points in \(\mathbb {F}_{2^n}\) of \((x+1)^m+x^m+(y+1)^m+y^m=0\) are points such that \(x=y\) or \(x=y+1\). This can happen on infinitely many extensions of \(\mathbb {F}_p\) only if \(\frac{(x+1)^m+x^m+(y+1)^m+y^m}{(x+y)(x+y+1)}\) has no absolutely irreducible factor over \(\mathbb {F}_2\).

In this paper, we investigate the case of monomial functions which are PN on infinitely many extensions of \(\mathbb {F}_p\) in odd characteristic. From now, we assume that the prime number \(p\) is odd. The only known PN power mappings are the following:

Proposition 1

Let \(\phi : x\mapsto x^m\) a power mapping. Then \(\phi \) is PN on \(\mathbb {F}_{p^n}\) for

  1. 1.

    \(m=2\),

  2. 2.

    \(m=p^l+1\) where l is an integer such that \(\frac{n}{\gcd (n,l)}\) is odd [1, 2],

  3. 3.

    \(m=\frac{3^l+1}{2}\) where \(p=3\) and l is an odd integer such that \(\gcd (l,n)=1\) [1].

All these monomials are PN on infinitely many extensions of \(\mathbb {F}_p\). In this paper, using the same methods as Jedlicka, Hernando and McGuire, we prove the following theorem:

Theorem 1

The only \(m\equiv 1\mod p\) such that \(x\mapsto x^m\) is PN on infinitely many extensions of \(\mathbb {F}_p\) are \(m=1+p^l\).

In the case where \(m\not \equiv 1\mod p\), using similar methods, Hernando, McGuire and Monserrat give partial results in [6]. Zieve completes the proof in [11] using a completely different method. However this method does not seem to apply in this case.

In Part 2 of this article, we give some background on algebraic curves and explain how we will prove Theorem 1. In the following parts, we prove Theorem 1.

2 Preliminaries

The notations set in this section hold for the remainder of the article.

2.1 Some background on algebraic curves

A reference for the following results is [4].

Definition 3

A polynomial \(f\in \mathbb {F}_q[x,y]\) is said absolutely irreducible if it is irreducible on an algebraic closure of \(\mathbb {F}_q\).

If \(f\in \mathbb {F}_q[x,y]\) is irreducible over \(\mathbb {F}_q\), then its absolutely irreducible factors are conjugate (see [10]).

Definition 4

For \(f\in \mathbb {F}_q[x,y]\), we denote by \(\widehat{f}\) the homogenized form of \(f\) and \(\widetilde{f}\) the dehomogenized form of \(\widehat{f}\) relatively to \(y\).

Definition 5

Let \(t=(x_0,y_0)\) be a point. We write \(f(x+x_0,y+y_0)=f_0+f_1+\cdots \) where if \(f_i\) is non zero then, it is an homogeneous polynomial of degree \(i\). Then, the multiplicity of \(f\) in \(t\), denoted by \(m_t(f)\) is the smallest \(i\) such that \(f_i\ne 0\). and the factors of \(f_{m_t(f)}\) on an algebraic closure of \(\mathbb {F}_q\) are called the tangent lines of \(f\) in \(t\).

A singular point of \(f\in \mathbb {F}_q[x,y]\) is a point \(t\) such that \(m_t(f)\ge 2\). In this case, we have \(f(t)=0\) and \(\frac{\partial f}{\partial x}(t)=0=\frac{\partial f}{\partial y}(t)\).

The intersection number of two plane curves \(u=0\) and \(v=0\) is a number indicating the multiplicity of intersection of these two curves. The intersection number of two plane curves over \(\mathbb {F}_q,\,u=0\) and \(v=0\) at a point \(t\) is \(\dim _{\mathbb {F}_q}(O_t(\mathbb {A}^2)/(u,v))\) where \(O_t(\mathbb {A}^2)\) is the ring of rational functions over the affine plane defined at \(t\). The intersection number of two plane curves \(u=0\) and \(v=0\) at a point \(t\) is denoted by \(I_t(u,v)\). However, we will not calculate this intersection number using the definition but rather using its properties:

  • \(I_t(u,v)=0\) if and only if \(m_t(u)=0\) or \(m_t(v)=0\).

  • \(I_t(u,v)=m_t(u)m_t(v)\) if and only if \(u\) and \(v\) have no common tangent lines.

  • If \(m_t(u)=1\), then \(I_t(u,v)=ord_t^u(v)\) where \(ord_t^u\) is the order on the discrete valuation ring \(O_t(\mathbb {A}^2)/(u)O_t(\mathbb {A}^2)\).

For more information on intersection numbers, we can read [4, pp. 74–81]

The following lemma is proved in [7]:

Lemma 1

Let \(J(x,y)=0\) be an affine curve over \(\mathbb {F}_q\) and \(t=(x_0,y_0)\) be a point of \(J\) of multiplicity \(m_t\). Then

$$\begin{aligned} J(x+x_0,y+y_0)=J_{m_t}+J_{m_t+1}+\cdots \end{aligned}$$

where if \(J_i\) is non zero then, it is an homogeneous polynomial of degree \(i\). We write

$$\begin{aligned} J(x,y)=u(x,y)\cdot v(x,y); \end{aligned}$$

if \(J_{m_t}\) and \(J_{m_t+1}\) are relatively prime then \(I_t(u,v)=m_t(u)\cdot m_t(v)\). In this case, if \(J\) has only one tangent line at \(t\), then \(I_t(u,v)=0\).

Theorem 2

(Bézout) Let \(u=0\) and \(v=0\) be two projective plane curves of degree \(n\) and \(m\) respectively without any common component then

$$\begin{aligned} \sum _{t}I_t(u,v)=n\cdot m. \end{aligned}$$

For a proof of this theorem see [4, p. 112]. From this theorem, we deduce the two following lemmas. The second one is proved in [5] for \(p=2\) but it is the exact same proof for \(p\ne 2\).

Lemma 2

If \(f\in \mathbb {F}_p[x,y]\) has no absolutely irreducible factor over \(\mathbb {F}_p\) then there exists a factorization \(f=uv\) such that

$$\begin{aligned} \sum _{t}I_t(u,v)\ge 2\frac{\mathrm {deg}(f)^2}{9}. \end{aligned}$$

Equivalently, if \(I_{tot}\) is any upper bound on the global intersection number \(\sum _tI_t(u,v)\) of \(u\) and \(v\) for all factorizations of \(f=u\cdot v\) over an algebraic closure of \(\mathbb {F}_p\), then

$$\begin{aligned} \frac{I_{tot}}{\frac{\mathrm {deg}(f)^2}{4}}\ge \frac{8}{9}. \end{aligned}$$

Proof

Suppose that \(f\) has no absolutely irreducible factor then, we write \(f=e_1\ldots e_r\), where each \(e_i\) is irreducible over \(\mathbb {F}_p\), but not absolutely irreducible. Then each \(e_i\) factors into \(c_i\ge 2\) factors on an algebraic closure of \(\mathbb {F}_p\) and its factors are all of degree \(\frac{\mathrm {deg}(e_i)}{c_i}\). Now, we factor each \(e_i\) into two factors \(u_i\) and \(v_i\) such that \(\mathrm {deg}(u_i)=\mathrm {deg}(v_i)+\frac{\mathrm {deg}(e_i)}{c_i}\) if \(c_i\) is odd (thus \(c_i\ge 3\)) or \(\mathrm {deg}(u_i)=\mathrm {deg}(v_i)\) if \(c_i\) is even. We set \(u=\prod _{i=1}^ru_i\) and \(v=\prod _{i=1}^rv_i\). Then \(\deg (u)-\deg (v)\le \frac{\deg (f)}{3}\). Since \(\deg (u)+\deg (v)=\deg (f)\),

$$\begin{aligned} \deg (u)\deg (v)\ge \frac{8}{9}\frac{\deg (f)^2}{4}. \end{aligned}$$

Let \(I_{tot}\) be an upper bound on the global intersection number of \(u\) and \(v\) for all factorizations of \(f\) into two factors over the algebraic closure of \(\mathbb {F}_p\). Then by Bézout’s theorem,

$$\begin{aligned} I_{tot}\ge \sum _tI_t(u,v)=\deg (u)\deg (v)\ge \frac{8}{9}\frac{\deg {(f)}^2}{4}=2\frac{\mathrm {deg}(f)^2}{9}. \end{aligned}$$

\(\square \)

Lemma 3

Let \(f\in \mathbb {F}_p[x,y],\,f_k,\,1\le k\le r\), the irreducible factors of \(f\) over \(\mathbb {F}_p\) and for all \(1\le k\le r\), we write \(f_k=f_{k,1}\ldots f_{k,c_k}\) the factorization of \(f_k\) into \(c_k\) absolutely irreducible factors. Then,

  1. 1.

    \(\deg (f_k)^2\le \sum _{t \in Sing(f)}m_t(f_k)^2\) where \(Sing(f)\) is the set of singular points of \(f\).

  2. 2.

    If \(t\) is a singular point of \(f,\,\sum _{1\le i< j\le c_k}m_t(f_{k,i})m_t(f_{k,j})\le m_t(f_k)^2\frac{c_k-1}{2c_k}\).

2.2 Strategy of proof

An equivalent definition for a PN function is that a function \(\phi \) is PN over \(\mathbb {F}_{q}\) if for all \(a\in \mathbb {F}_{q}^*\), the only rational points in \(\mathbb {F}_{q}\) of

$$\begin{aligned} \phi (x+a)-\phi (x)-\phi (y+a)+\phi (y)=0 \end{aligned}$$

are points such that \(x=y\).

In this article, we are only interested in monomial functions, \(\phi :x\mapsto x^m,\,m\ge 3\). We only have to consider the case where \(a=1\) in the definition of PN functions (see [3]).

Remark 1

If \(m\) is odd then, 0 and \(-1\) are solutions of \((x+1)^m-x^m=1\). So, in this case, \(x \mapsto x^m\) is not PN over \(\mathbb {F}_{p^n}\) for any \(n\).

We set \(f(x,y)=(x+1)^m-x^m-(y+1)^m+y^m\). Since \((x-y)\) divides \(f(x,y)\), we define \(h(x,y)=\frac{f(x,y)}{(x-y)}\).

We can assume that \(m\not \equiv 0\mod p\). Indeed, if \(x\mapsto x^m\) is PN over \(\mathbb {F}_q\) and \(m\equiv 0\mod p\) then \(x\mapsto x^{\frac{m}{p}}\) is also PN over \(\mathbb {F}_q\).

Then, the proof of Theorem 1 follow from Proposition 2 and Theorem 3 below.

Proposition 2

If \(h\) has an absolutely irreducible factor over \(\mathbb {F}_p\) then, for \(n\) sufficiently large, \(x\mapsto x^m\) is not PN on \(\mathbb {F}_{p^n}\).

Proof

Assume that \(h\) has an absolutely irreducible factor over \(\mathbb {F}_p\), denoted by \(Q\). If \(Q(x,y)=c(x-y)\) with \(c\in \mathbb {F}_p^*\), then \(f(x,y)=(y-x)^2\widetilde{Q}(x,y),\,\widetilde{Q}\in \mathbb {F}_p[x,y]\). Hence,

$$\begin{aligned} -m(y+1)^{m-1}+my^{m-1}=\frac{\partial f}{\partial y}(x,y)=2(y-x)\widetilde{Q}(x,y)+(y-x)^2\frac{\partial \widetilde{Q}}{\partial y}(x,y). \end{aligned}$$

So, we get that for all \(x\in \mathbb {F}_{p^n},\,-m(x+1)^{m-1}+mx^{m-1}=0\) which is impossible since \(m\not \equiv 0\mod p\). Let \(s\) be the degree of \(Q\). Since \(Q\ne c(x-y),\,Q(x,x)\) is not the null polynomial. So, there are at most \(s\) rational points of \(Q\) such that \(x=y\).

On the other hand, if we denote by P the number of affine rational points of \(Q\) on \(\mathbb {F}_{p^n}\), we have (see [9, p. 331]):

$$\begin{aligned} |P-p^n|\le (s-1)(s-2)\sqrt{p^n}+s^2. \end{aligned}$$

Hence, for \(n\) sufficiently large, \(Q\) has a rational point in \(\mathbb {F}_{p^n}\) such that \(x\ne y\) and \(x\mapsto x^m\) is not PN over \(\mathbb {F}_{p^n}\). \(\square \)

Theorem 3

Let \(m\) be an integer such that \(m\ge 3,\,m\equiv 1\mod p\) and \(m\ne 1+p^l\). Assume that \(\frac{m-1}{p^l}\ne p^l-1\). Then \(h\) has an absolutely irreducible factor over \(\mathbb {F}_p\).

From now, we are interested in the case where \(m\equiv 1\mod p\). We denote by l the greatest integer such that \(p^l\) divides \(m-1\) and we set

$$\begin{aligned} d:=\mathrm {gcd}(m-1,p^l-1)=\mathrm {gcd}\left( \frac{m-1}{p^l},p^l-1\right) . \end{aligned}$$

Then, by Theorem 3 and Proposition 2, we only have to treat the case where \(d=\frac{m-1}{p^l}=p^l-1\) in Theorem 1. We have \(m=p^l(p^l-1)+1\) which is odd; so \(x\mapsto x^m\) is not PN on all extensions of \(\mathbb {F}_p\).

Now, we only have to prove Theorem 3. The method of Jedlicka, Hernando and McGuire is, using Bézout’s theorem, to prove that \(h\) has an absolutely irreducible factor over \(\mathbb {F}_p\) because it has not enough singular points. In Part 3, we study singular points of \(h\) and their multiplicity. In Part 4, we bound the intersection number \(I_t(u,v)\) where \(t\) is a singular point of \(h\) and \(u,\,v\) are such that \(h=uv\). In Part 5, we prove Theorem 3.

We set \(F=(x+z)^m-x^m-(y+z)^m+y^m=z\widehat{f}\) and \(\widetilde{F}=(x+z)^m-x^m-(z+1)^m+1\) the dehomogenized form of \(F\) relatively to \(y\).

3 Singularities of \(h\)

Proposition 3

The singular points of \(h\) are described in Table 1.

Table 1 Singularities of \(h\) for \(m=1+\sum _{j=1}^b m_j p^{i_j}\) with \(1\le m_j\le p-1,\,i_j>i_{j-1},\,i_1=l\)

The proof of this theorem follows from Lemmas 4 to 11 and their corollaries (more precisions are given in the last column of the Table 1).

3.1 Singular points at infinity

We have

$$\begin{aligned} \left\{ \begin{array}{lllll} F_x&{}=&{}\dfrac{\partial F}{\partial x}&{}=&{}m(x+z)^{m-1}-mx^{m-1}\\ F_y&{}=&{}\dfrac{\partial F}{\partial y}&{}=&{}-m(y+z)^{m-1}+my^{m-1}\\ F_z&{}=&{}\dfrac{\partial F}{\partial z}&{}=&{}m(x+z)^{m-1}-m(y+z)^{m-1} \end{array}\right. . \end{aligned}$$

At infinity (\(z=0\)), \(F_x(x,y,0)=F_y(x,y,0)=0\) and

$$\begin{aligned} F_z(x,y,0)=m\big (x^{m-1}-y^{m-1}\big ). \end{aligned}$$

So \((x_0,y_0,0)\) is a singular point of \(F\) if and only if \(x_0^{m-1}=y_0^{m-1}\). If \(y_0=0\) then \(x_0=0\); so \(y_0\ne 0\) and we have to study the solutions of

$$\begin{aligned} x_0^{m-1}=1. \end{aligned}$$
(1)

Equation 1 is equivalent to \(x_0^{\frac{m-1}{p^l}}=1\). Since \(\mathrm {gcd}\big (\frac{m-1}{p^l},p\big )=1\), there are \(\frac{m-1}{p^l}\) solutions at (1) and \(x_0=1\) is the only one such that \(x_0=y_0\).

Now, we want to find the multiplicity of these singularities:

$$\begin{aligned} \widetilde{F}(x+x_0,z)&= (x+x_0+z)^m-(x+x_0)^m-(z+1)^m+1\\&= \sum _{k=2}^m\left( {\begin{array}{c}m\\ k\end{array}}\right) (x+z)^kx_0^{m-k}-\sum _{k=2}^m\left( {\begin{array}{c}m\\ k\end{array}}\right) x^kx_0^{m-k}-\sum _{k=2}^m\left( {\begin{array}{c}m\\ k\end{array}}\right) z^k. \end{aligned}$$

Since \(m-1\equiv 0\mod p^l\), for all \(2\le k<p^l,\,\left( {\begin{array}{c}m\\ k\end{array}}\right) =0\). Consider the terms of degree \(p^l-1\) of \(\widetilde{f}\):

$$\begin{aligned} \frac{1}{z}\left( {\begin{array}{c}m\\ p^l\end{array}}\right) \left( x_0^{m-p^l}(x+z)^{p^l}-x_0^{m-p^l}x^{p^l}-z^{p^l}\right) =\left( {\begin{array}{c}m\\ p^l\end{array}}\right) \left( x_0^{m-p^l}-1\right) z^{p^l-1}. \end{aligned}$$
(2)

This term vanishes (which means that \((x_0,y_0,0)\) is a singular point of multiplicity greater than \(p^l-1\)) if and only if

$$\begin{aligned} x_0^{m-p^l}=1 \end{aligned}$$

that is to say if and only if

$$\begin{aligned} x_0^d=1. \end{aligned}$$

Now, consider the terms of degree \(p^l\) of \(\widetilde{f}\):

$$\begin{aligned}&\frac{1}{z}\left( {\begin{array}{c}m\\ p^l+1\end{array}}\right) \left( x_0^{m-p^l-1}(x+z)^{p^l+1}-x_0^{m-p^l-1}x^{p^l+1}-z^{p^l+1}\right) \nonumber \\&\quad =\left( {\begin{array}{c}m\\ p^l+1\end{array}}\right) \left( x_0^{m-p^l-1}x^{p^l}+x_0^{m-p^l-1}xz^{p^l-1}+\left( x_0^{m-p^l-1}-1\right) z^{p^l}\right) . \end{aligned}$$
(3)

Since \(x_0^{m-p^l-1}\ne 0\), singular points of \(\widehat{f}\) of multiplicity greater than \(p^l-1\) have multiplicity \(p^l\).

We have just proved the following lemma:

Lemma 4

Let \(\omega \) such that \(\omega ^{\frac{m-1}{p^l}}=1\). The point \((\omega :1:0)\) is a singular point of \(\widehat{h}\) with multiplicity

$$\begin{aligned} \left\{ \begin{array}{l@{\quad }l} p^l &{} { if}\, \omega ^d=1, \omega \ne 1\\ p^l-1 &{} { otherwise} \end{array}\right. . \end{aligned}$$

Furthermore, \(\widehat{h}\) has \(\frac{m-1}{p^l}\) singular points at infinity.

3.2 Affine singular points

We have:

$$\begin{aligned} \left\{ \begin{array}{lll} f_x&{}=&{}m(x+1)^{m-1}-mx^{m-1}\\ f_y&{}=&{}-m(y+1)^{m-1}+my^{m-1} \end{array}\right. . \end{aligned}$$

So,

$$\begin{aligned} (x_0,y_0) \,\mathrm{singular \,point \,of} \,f&\Leftrightarrow \left\{ \begin{array}{l} f(x_0,y_0)=0\\ (x_0+1)^{m-1}=x_0^{m-1}\\ (y_0+1)^{m-1}=y_0^{m-1}\end{array}\right. \\&\Leftrightarrow \left\{ \begin{array}{l} x_0^{m-1}(x_0+1)-x_0^m-y_0^{m-1}(y_0+1)+y_0^m=0\\ (x_0+1)^{m-1}=x_0^{m-1}\\ (y_0+1)^{m-1}=y_0^{m-1}\end{array}\right. \\&\Leftrightarrow \left\{ \begin{array}{l} x_0^{m-1}=y_0^{m-1}\\ (x_0+1)^{m-1}=x_0^{m-1}\\ (y_0+1)^{m-1}=y_0^{m-1}\end{array}\right. . \end{aligned}$$

Finally, we have

Lemma 5

Affine singular points of \(f\) are points satisfying

$$\begin{aligned} (x_0+1)^{m-1}=x_0^{m-1}=y_0^{m-1}=(y_0+1)^{m-1}. \end{aligned}$$

From Lemma 5, we get that \(x_0,\,y_0\ne 0,-1\). Since \(p^l\) divides \(m-1\),

$$\begin{aligned} (x_0,y_0) \,\hbox {singular point of}\, f \Leftrightarrow \left\{ \begin{array}{l} x_0^{\frac{m-1}{p^l}}=y_0^{\frac{m-1}{p^l}}\\ (x_0+1)^{\frac{m-1}{p^l}}=x_0^{\frac{m-1}{p^l}}\\ (y_0+1)^{\frac{m-1}{p^l}}=y_0^{\frac{m-1}{p^l}} \end{array}\right. . \end{aligned}$$
(4)

There are at most \(\frac{m-1}{p^l}-1\) solutions to the second equation of (4). Let \(x_0\) be one of these solutions, we want to know the number of \(y_0\) such that \((x_0,y_0)\) is a singular point of \(f\).

We write \(m=1+\sum _{j=1}^b m_j p^{i_j}\) with \(1\le m_j\le p-1,\,i_j>i_{j-1},\,i_1=l\). Then,

$$\begin{aligned} (y_0+1)^{\frac{m-1}{p^l}}=y_0^{\frac{m-1}{p^l}}&\Leftrightarrow \prod _{j=1}^b(y_0+1)^{m_jp^{i_j-l}}=y_0^{\frac{m-1}{p^l}}\\&\Leftrightarrow \sum _{(k_1,\ldots ,k_b)\in \mathcal {I}}\left( \prod _{j=1}^b\left( {\begin{array}{c}m_j\\ k_j\end{array}}\right) \right) y_0^{\sum _{j=1}^bk_jp^{i_j-l}}=0, \end{aligned}$$

where \(\mathcal {I}=\{(k_1,\ldots ,k_b)\in \mathbb {Z}^b:\forall j=1\ldots b, \ 0\le k_j\le m_j\}\setminus \{(m_1,\ldots , m_b)\}\). We multiply by \(y_0^{\frac{m-1}{p^l}-m_bp^{i_b-l}}\) and we set \(\alpha =y_0^{\frac{m-1}{p^l}}\):

$$\begin{aligned}&\sum _{(k_1,\ldots ,k_{b-1})\in \mathcal {I'}}\left( \,\,\prod _{j=1}^{b-1}\left( {\begin{array}{c}m_j\\ k_j\end{array}}\right) \right) \alpha y_0^{\sum _{j=1}^{b-1}k_jp^{i_j-l}}\\&\qquad +\sum _{k_b=0}^{m_b-1}\sum _{\begin{array}{l}0 \le k_j\le m_j\\ j\ne b \end{array}} \left( \,\,\prod _{j=1}^{b}\left( {\begin{array}{c}m_j\\ k_j\end{array}}\right) \right) y_0^{\frac{m-1}{p^l}-(m_b-k_b)p^{i_b-l}+\sum _{j=1}^{b-1}k_jp^{i_j-l}}=0, \end{aligned}$$

where \(\mathcal {I'}=\{(k_1,\ldots ,k_{b-1})\in \mathbb {Z}^{b-1}:\forall j=1\ldots b-1, \ 0\le k_j\le m_j\}{\setminus }\{(m_1,\ldots , m_{b-1})\}\).

The degree of this polynomial in \(y_0\) is

$$\begin{aligned} \frac{m-1}{p^l}-p^{i_b-l}+\displaystyle \sum _{j=1}^{b-1}m_jp^{i_j-l}=2\frac{m-1}{p^l}-(m_b+1)p^{i_b-l}. \end{aligned}$$

Then, we obtain

Lemma 6

The number of affine singularities of \(h\) is at most:

$$\begin{aligned} \left( \frac{m-1}{p^l}-1\right) \left( 2\frac{m-1}{p^l}-(m_b+1)p^{i_b-l}\right) , \end{aligned}$$

where \(m=1+\sum _{j=1}^b m_j p^{i_j}\) with \(1\le m_j\le p-1,\,i_j>i_{j-1},\,i_1=l\).

Now, we study the multiplicity of affine singularities:

$$\begin{aligned} f(x+x_0,y+y_0)&= (x+x_0+1)^m-(x+x_0)^m-(y+y_0+1)^m+(y+y_0)^m\\&= \sum _{k=2}^m\left( {\begin{array}{c}m\\ k\end{array}}\right) x^k(x_0+1)^{m-k}-\sum _{k=2}^m\left( {\begin{array}{c}m\\ k\end{array}}\right) x^kx_0^{m-k}\\&-\sum _{k=2}^m\left( {\begin{array}{c}m\\ k\end{array}}\right) y^k(y_0+1)^{m-k}+\sum _{k=2}^m\left( {\begin{array}{c}m\\ k\end{array}}\right) y^ky_0^{m-k}. \end{aligned}$$

Since \(m-1\equiv 0\mod p^l\), for all \(2\le k<p^l,\,\left( {\begin{array}{c}m\\ k\end{array}}\right) =0\). So \((x_0,y_0)\) is a singularity of multiplicity at least \(p^l\). Consider the terms of degree \(p^l+1\):

$$\begin{aligned} \left( {\begin{array}{c}m\\ p^l+1\end{array}}\right) \left( \left( (x_0+1)^{m-p^l-1}-x_0^{m-p^l-1}\right) x^{p^l+1}-\left( (y_0+1)^{m-p^l-1}-y_0^{m-p^l-1}\right) y^{p^l+1}\right) . \end{aligned}$$

Since \((x_0,y_0)\) is a singular point, \((x_0+1)^{m-1}=x_0^{m-1}\) and \(x_0\ne -1,0\). So,

$$\begin{aligned} (x_0+1)^{m-p^l-1}-x_0^{m-p^l-1}=0&\Leftrightarrow (x_0+1)^{p^l}\left( (x_0+1)^{m-p^l-1}-x_0^{m-p^l-1}\right) =0\\&\Leftrightarrow -x_0^{m-p^l-1}=0. \end{aligned}$$

Hence, affine singularities have multiplicity at most \(p^l+1\). Then, we look at the terms of degree \(p^l\):

$$\begin{aligned} \left( {\begin{array}{c}m\\ p^l\end{array}}\right) \left( \left( (x_0+1)^{m-p^l}-x_0^{m-p^l}\right) x^{p^l}-\left( (y_0+1)^{m-p^l}-y_0^{m-p^l}\right) y^{p^l}\right) . \end{aligned}$$

However,

$$\begin{aligned} (x_0+1)^{m-p^l}-x_0^{m-p^l}=0&\Leftrightarrow (x_0+1)^{p^l}\left( (x_0+1)^{m-p^l}-x_0^{m-p^l}\right) =0\\&\Leftrightarrow (x_0+1)^{m-1}(x_0+1)-x_0^m-x_0^{m-p^l}=0\\&\Leftrightarrow x_0^{m-p^l}\left( x_0^{p^l-1}-1\right) =0\\&\Leftrightarrow x_0\in \mathbb {F}_{p^l}^*. \end{aligned}$$

We can do the same for \(y_0\).

We have just proved the following lemma.

Lemma 7

There are at most:

  • \(d-1\) affine singularities of \(h\) such that \(x_0=y_0\in \mathbb {F}_{p^l}^*\). Their multiplicity is \(p^l\) (\(p^l+1\) for \(f\));

  • \(\frac{m-1}{p^l}-d\) affine singularities of \(h\) such that \(x_0=y_0\not \in \mathbb {F}_{p^l}^*\). Their multiplicity is \(p^l-1\) (\(p^l\) for \(f\));

  • \((d-1)(d-2)\) affine singularities of \(h\) such that \(x_0\ne y_0\) and \(x_0,\,y_0\in \mathbb {F}_{p^l}^*\). Their multiplicity is \(p^l+1\) (for \(h\) and \(f\));

  • \(\big (\frac{m-1}{p^l}-1\big )\big (2\frac{m-1}{p^l}-(m_b+1)p^{i_b-l}-1\big )-(d-1)(d-2)\) affine singularities of \(h\) such that \(x_0\ne y_0\) and \(x_0\) or \(y_0\not \in \mathbb {F}_{p^l}^*\) (\(m=1+\sum _{j=1}^b m_j p^{i_j}\) with \(1\le m_j\le p-1,\,i_j>i_{j-1},\,i_1=l\)). Their multiplicity is \(p^l\) (for \(h\) and \(f\)).

4 Intersection number bounds

We write \(h=uv\); we want to bound the intersection number \(I_t(u,v)\) for \(t\) a singularity of \(h\).

4.1 Singularities at infinity

Let \(t=(\omega :1:0)\) be a singular point of \(h\) at infinity (\(\omega ^{\frac{m-1}{p^l}}=1\)) of multiplicity \(m_t\). We write \(\widetilde{h}(x+\omega ,z)=\widetilde{H}_{m_t}+\widetilde{H}_{m_t+1}+\cdots \) where \(\widetilde{H}_i\) is the homogeneous polynomial composed of the terms of degree \(i\) of \(\widetilde{h}(x+\omega ,z)\) and \(\widetilde{f}(x+\omega ,z)=\widetilde{F}_{m_t}+\widetilde{F}_{m_t+1}+\cdots \) where \(\widetilde{F}_i\) is the homogeneous polynomial composed of the terms of degree \(i\) of \(\widetilde{f}(x+\omega ,z)\). Then,

$$\begin{aligned} \widetilde{f}(x+\omega ,z)&= \widetilde{h}(x+\omega ,z)(x+\omega -1)\\&= \left( R+\widetilde{H}_{m_t+1}+\widetilde{H}_{m_t}\right) (x+\omega -1)\\&\text {where if} \,R \,\text {is non zero then, it is a polynomial of degree greater than} \,m_t+1\\&= x R+(\omega -1)R+x\widetilde{H}_{m_t+1}+x\widetilde{H}_{m_t}+(\omega -1)\widetilde{H}_{m_t+1}+(\omega -1)\widetilde{H}_{m_t}. \end{aligned}$$

So,

  • if \(\omega \ne 1\), then \(\widetilde{F}_{m_t}=(\omega -1)\widetilde{H}_{m_t}\) and \(\widetilde{F}_{m_t+1}=x\widetilde{H}_{m_t}+(\omega -1)\widetilde{H}_{m_t+1}\);

  • if \(\omega =1\), then \(\widetilde{F}_{m_t+1}=x\widetilde{H}_{m_t}\).

Then, we have

Lemma 8

If \(t=(\omega :1:0),\,\omega ^{\frac{m-1}{p^l}}=1\), is a singular point at infinity of \(h\) with multiplicity \(m_t\) then

  • \(\widetilde{F}_{m_t}=(\omega -1)\widetilde{H}_{m_t}\) and \(\widetilde{F}_{m_t+1}=x\widetilde{H}_{m_t}+(\omega -1)\widetilde{H}_{m_t+1}\) if \(\omega \ne 1\);

  • \(\widetilde{F}_{m_t+1}=x\widetilde{H}_{m_t}\) if \(\omega =1\).

Corollary 1

If \(t=(1:1:0)\) then

$$\begin{aligned} I_t(u,v)\le \left( \frac{p^l-1}{2}\right) ^2. \end{aligned}$$

Proof

If \(t=(1:1:0)\) then its multiplicity is \(p^l-1\). By Lemma 8 and Eq. 3, there exists \(a \in \mathbb {F}_q^*\) such that

$$\begin{aligned} \widetilde{H}_{m_t}=a\left( x^{p^l-1}+z^{p^l-1}\right) . \end{aligned}$$

Since the factors of \(\widetilde{H}_{m_t}\) are different, \(I_t(u,v)=m_t(u)m_t(v)\). We get the result since \(m_t(u)+m_t(v)=p^l-1\). \(\square \)

Corollary 2

If \(t=(\omega :1:0)\) such that \(\omega ^d=1,\,\omega \ne 1\) then

$$\begin{aligned} I_t(u,v)\le \frac{p^{2l}-1}{4}. \end{aligned}$$

Proof

Suppose that \(t=(\omega :1:0)\) such that \(\omega ^d=1\) and \(\omega \ne 1\) then, the multiplicity of \(t\) is \(p^l\). By Lemma 8 and Eq. 3, there exists \(a\in \mathbb {F}_q^*\) such that

$$\begin{aligned} (\omega -1)\widetilde{H}_{p^l}=\widetilde{F}_{p^l}=a\left( x^{p^l}\omega ^{m-p^l-1}+xz^{p^l-1}\omega ^{m-p^l-1}+\left( \omega ^{m-p^l-1}-1\right) z^{p^l}\right) . \end{aligned}$$

So all factors of \(\widetilde{H}_{p^l}\) are simple and \(I_t(u,v)=m_t(u)m_t(v)\). We get the result since \(m_t(u)+m_t(v)=p^l\). \(\square \)

Corollary 3

If \(t=(\omega :1:0)\) with \(\omega ^{\frac{m-1}{p^l}}=1,\,\omega ^d\ne 1\), then

$$\begin{aligned} I_t(u,v)=0. \end{aligned}$$

Proof

Suppose that \(t=(\omega :1:0)\) with \(\omega ^{\frac{m-1}{p^l}}=1\) and \(\omega ^d\ne 1\) then, the multiplicity of \(t\) is \(p^l-1\). By Lemma 8 and Eq. 2, there exists \(a\in \mathbb {F}_q^*\) and \(b\in \mathbb {F}_q^*\) such that

$$\begin{aligned} (\omega -1)\widetilde{H}_{p^l-1}=\widetilde{F}_{p^l-1}=a z^{p^l-1} \end{aligned}$$

and

$$\begin{aligned} \widetilde{F}_{p^l}\!=\!x\widetilde{H}_{p^l-1}\!+\!(\omega -1)\widetilde{H}_{p^l}\!=\!b\left( x^{p^l}\omega ^{m-p^l-1}+xz^{p^l-1}\omega ^{m-p^l-1}+z^{p^l}\left( \omega ^{m-p^l-1}-1\right) \right) . \end{aligned}$$

So, \(\mathrm {gcd}(\widetilde{H}_{p^l},\widetilde{H}_{p^l-1})=\mathrm {gcd}(\widetilde{F}_{p^l},\widetilde{F}_{p^l-1})=1\). Since \(\widetilde{H}_{p^l-1}\) has only one tangent line, by Lemma 1, \(I_t(u,v)=0\). \(\square \)

4.2 Affine singularities

Let \(t=(x_0,y_0)\) be an affine singular point of \(h\) of multiplicity \(m_t\).

We write \(h(x+x_0,y+y_0)=H_{m_t}+H_{m_t+1}+\cdots \) where \(H_i\) is the homogeneous polynomial composed of the terms of degree \(i\) of \(h(x+x_0,y+y_0)\).

Assume \(x_0=y_0\). Then, we write \(f(x+x_0,y+y_0)=F_{m_t+1}+F_{m_t+2}+\cdots \) where \(F_i\) is the homogeneous polynomial composed of the terms of degree \(i\) of \(f(x+x_0,y+y_0)\) and

$$\begin{aligned} f(x+x_0,y+y_0)&= h(x+x_0,y+y_0)(x+x_0-y-y_0)\\&= (R+H_{m_t+1}+H_{m_t})(x-y)\\&\text {where if } R \text { is non zero then, it is a polynomial of degree}\\&\text {greater than } m_t+1\\&= (x-y)R+(x-y)H_{m_t+1}+(x-y)H_{m_t}. \end{aligned}$$

So, \(F_{m_t+2}=(x-y)H_{m_t+1}\) and \(F_{m_t+1}=(x-y)H_{m_t}\). Furthermore, for some \(a,\,F_{m_t+1}=a(x^{m_t+1}-y^{m_t+1})\) (see proof of Lemma 7).

So, we get

Lemma 9

If \(t=(x_0,y_0)\) is an affine singular point of \(h\) with multiplicity \(m_t\) such that \(x_0=y_0\), then \(F_{m_t+2}=(x-y)H_{m_t+1}\) and \(F_{m_t+1}=(x-y)H_{m_t}\).

Furthermore, tangent lines to \(h\) at \(t\) are the factors of \(\frac{x^{m_t+1}-y^{m_t+1}}{x-y}\).

Corollary 4

If \(t=(x_0,y_0)\) is an affine singular point of \(h\) such that \(x_0=y_0\in \mathbb {F}_{p^l}^*\) then

$$\begin{aligned} I_t(u,v)\le \frac{p^{2l}-1}{4}. \end{aligned}$$

Proof

Suppose that \(t=(x_0,y_0)\) is an affine singular point of \(h\) such that \(x_0=y_0\in \mathbb {F}_{p^l}^*\) then, the multiplicity of \(t\) is \(p^l\). The factors of \(\frac{x^{p^l+1}-y^{p^l+1}}{x-y}\) are all distinct. So, by Lemma 9, tangent lines to \(u\) or \(v\) are all distinct and

$$\begin{aligned} I_t(u,v)=m_t(u)m_t(v). \end{aligned}$$

Since \(m_t(u)+m_t(v)=p^l\), we get the result. \(\square \)

Corollary 5

If \(t=(x_0,y_0)\) is an affine singular point of \(h\) such that \(x_0=y_0\not \in \mathbb {F}_{p^l}^*\) then,

$$\begin{aligned} I_t(u,v)=0. \end{aligned}$$

Proof

Suppose that \(t=(x_0,y_0)\) is an affine singular point of \(h\) such that \(x_0=y_0\not \in \mathbb {F}_{p^l}^*\) then, the multiplicity of \(t\) is \(p^l-1\). By Lemma 9,

$$\begin{aligned} H_{p^l-1}=a(x-y)^{p^l-1}\quad \hbox {and}\quad H_{p^l}=\frac{b\left( x^{p^l+1}-y^{p^l+1}\right) }{x-y}. \end{aligned}$$

Hence, \(\mathrm {gcd}(H_{p^l-1},H_{p^l})=1\). Since \(H_{p^l-1}\) has only one tangent line, by Lemma 1, \(I_t(u,v)=0\). \(\square \)

Assume now \(x_0\ne y_0\). Then, we write \(f(x+x_0,y+y_0)=F_{m_t}+F_{m_t+1}+\cdots \) where \(F_i\) is the homogeneous polynomial composed of the terms of degree \(i\) of \(f(x+x_0,y+y_0)\) and

$$\begin{aligned} f(x+x_0,y+y_0)&= h(x+x_0,y+y_0)(x+x_0-y-y_0)\\&= \left( R+H_{m_t+1}+H_{m_t}\right) (x+x_0-y-y_0)\\&\text {where if } R \text { is non zero then, it is a polynomial of degree}\\&\text {greater than } m_t+1\\&= (x_0-y_0)H_{m_t}+\left( (x-y)H_{m_t}+(x_0-y_0)H_{m_t+1}\right) \\&+\left( (x-y+x_0-y_0)R+(x-y)H_{m_t+1}\right) . \end{aligned}$$

So, \(F_{m_t}=(x_0-y_0)H_{m_t}\) and \(F_{m_t+1}=(x_0-y_0)H_{m_t+1}+(x-y)H_{m_t}\).

Then, we obtain the following lemma.

Lemma 10

If \(t=(x_0,y_0)\) is an affine singular point of \(h\) with multiplicity \(m_t\) such that \(x_0\ne y_0\) then

$$\begin{aligned} \mathbb {F}_{m_t}=(x_0-y_0)H_{m_t}\quad \hbox {and}\quad F_{m_t+1}=(x-y)H_{m_t}+(x_0-y_0)H_{m_t+1}. \end{aligned}$$

Corollary 6

If \(t=(x_0,y_0)\) is an affine singular point of \(h\) such that \(x_0\ne y_0,\,x_0,\,y_0\in \mathbb {F}_{p^l}^*\) then

$$\begin{aligned} I_t(u,v)\le \left( \frac{p^l+1}{2}\right) ^2. \end{aligned}$$

Proof

Suppose that \(t=(x_0,y_0)\) is an affine singular point of \(h\) such that \(x_0\ne y_0,\,x_0,\,y_0\in \mathbb {F}_{p^l}^*\) then, the multiplicity of \(t\) is \(p^l+1\). By Lemma 10,

$$\begin{aligned} (x_0-y_0)H_{m_t}=F_{m_t}=c_1x^{p^l+1}-c_2y^{p^l+1}\quad \hbox {with } c_1, c_2\ne 0. \end{aligned}$$

Hence, all factors of \(H_{m_t}\) are simple and then \(I_t(u,v)=m_t(u)m_t(v)\). Since \(m_t(u)+m_t(v)=p^l+1\), we get the result. \(\square \)

Corollary 7

If \(t=(x_0,y_0)\) is an affine singular point of \(h\) such that \(x_0\ne y_0\) and \(x_0\in \mathbb {F}_{p^l}^*\) and \(y_0\not \in \mathbb {F}_{p^l}^*\) or \(x_0\not \in \mathbb {F}_{p^l}^*\) and \(y_0\in \mathbb {F}_{p^l}^*\) then

$$\begin{aligned} I_t(u,v)=0. \end{aligned}$$

Proof

Suppose that \(t=(x_0,y_0)\) is an affine singular point of \(h\) such that \(x_0\ne y_0\) and \(x_0\in \mathbb {F}_{p^l}^*\) and \(y_0\not \in \mathbb {F}_{p^l}^*\) or \(x_0\not \in \mathbb {F}_{p^l}^*\) and \(y_0\in \mathbb {F}_{p^l}^*\) then, the multiplicity of \(t\) is \(p^l\). Then

$$\begin{aligned} F_{p^l}=\left\{ \begin{array}{ll} c_1x^{p^l}&{} {\text {if}}\, y_0\in \mathbb {F}_{p^l}^*, c_1 \ne 0\\ c_2y^{p^l}&{} {\text {if}}\, x_0\in \mathbb {F}_{p^l}^*, c_2 \ne 0 \end{array}\right. \quad \hbox {and}\quad F_{p^l+1}=c_1'x^{p^l+1}-c_2'y^{p^l+1}, c_1', c_2'\ne 0. \end{aligned}$$

So, by Lemma 10, \(1=\mathrm {gcd}(F_{p^l},F_{p^l+1})=\mathrm {gcd}(H_{p^l},H_{p^l+1})\) and \(H_{p^l}\) has only one tangent line. Hence, by Lemma 1, \(I_t(u,v)=0\). \(\square \)

Assume \(x_0\ne y_0\) and \(x_0,\,y_0\not \in \mathbb {F}_{p^l}\). Then, \(t\) has multiplicity \(p^l\). We have \(F_{p^l}=c_1x^{p^l}-c_2y^{p^l}=(c_3x-c_4y)^{p^l}\), where \(c_1=(x_0+1)^{m-p^l}-x_0^{p^l}\) and \(c_2=(y_0+1)^{m-p^l}-y_0^{m-p^l}\). Since \(x_0,\,y_0\not \in \mathbb {F}_{p^l}^*,\,c_1\ne 0\) and \(c_2\ne 0\). By Lemma 10,

$$\begin{aligned} F_{p^l}=(x_0-y_0)H_{p^l}\quad \hbox {and}\quad F_{p^l+1}=(x_0-y_0)H_{p^l+1}+(x-y)H_{p^l}. \end{aligned}$$

So, \(H_{p^l}\) has only one factor and \(\mathrm {gcd}(F_{p^l},F_{p^l+1})=\mathrm {gcd}(H_{p^l},H_{p^l+1})\). Furthermore, \(F_{p^l+1}=d_1x^{p^l+1}-d_2y^{p^l+1}\) with \(d_1=(x_0+1)^{m-p^l-1}-x_0^{m-p^l-1}\ne 0\) and \(d_2=(y_0+1)^{m-p^l-1}-y_0^{m-p^l-1}\ne 0\). The polynomials \(F_{p^l}\) and \(F_{p^l+1}\) have a common factor if and only if \(c_3x-c_4y\) divides \(F_{p^l+1}\). So, \(F_{p^l}\) and \(F_{p^l+1}\) have a common factor if and only if

$$\begin{aligned} \left( \frac{c_1}{c_2}\right) ^{p^l+1}=\left( \frac{d_1}{d_2}\right) ^{p^l}. \end{aligned}$$

If \((x_0,y_0)\) is a singular point of \(f\), then

$$\begin{aligned} \left\{ \begin{array}{l} x_0^{m-1}=y_0^{m-1}\\ (x_0+1)^{m-1}=x_0^{m-1}\\ (y_0+1)^{m-1}=y_0^{m-1} \end{array}\right. . \end{aligned}$$

We have:

$$\begin{aligned} d_1=(x_0+1)^{m-p^l-1}-x_0^{m-p^l-1}&= \frac{(x_0+1)^{m-1}-x_0^{m-p^l-1}(x_0+1)^{p^l}}{(x_0+1)^{p^l}}\\&= \frac{x_0^{m-1}-x_0^{m-1}-x_0^{m-p^l-1}}{(x_0+1)^{p^l}}\\&= \frac{-x_0^{m-p^l-1}}{(x_0+1)^{p^l}}. \end{aligned}$$

Similarly, \(d_2=\frac{-y_0^{m-p^l-1}}{(y_0+1)^{p^l}}\). Hence,

$$\begin{aligned} \frac{d_1}{d_2}=\frac{x_0^{m-p^l-1}(y_0+1)^{p^l}}{y_0^{m-p^l-1}(x_0+1)^{p^l}} = \frac{x_0^{m-1}y_0^{p^l}(y_0+1)^{p^l}}{y_0^{m-1}x_0^{p^l}(x_0+1)^{p^l}}=\frac{y_0^{p^l}(y_0+1)^{p^l}}{x_0^{p^l}(x_0+1)^{p^l}}. \end{aligned}$$

On the other hand, we have:

$$\begin{aligned} c_1=(x_0+1)^{m-p^l}-x_0^{m-p^l}&= \frac{(x_0+1)(x_0+1)^{m-1}-x_0^{m-p^l}(x_0+1)^{p^l}}{(x_0+1)^{p^l}}\\&= \frac{x_0^{m}+x_0^{m-1}-x_0^{m}-x_0^{m-p^l}}{(x_0+1)^{p^l}}\\&= \frac{x_0^{m-p^l}\left( x_0^{p^l-1}-1\right) }{(x_0+1)^{p^l}}. \end{aligned}$$

Similarly, \(c_2=\frac{y_0^{m-p^l}(y_0^{p^l-1}-1)}{(y_0+1)^{p^l}}\). Hence,

$$\begin{aligned} \frac{c_1}{c_2}=\frac{x_0^{m-p^l}\left( x_0^{p^l-1}-1\right) (y_0+1)^{p^l}}{y_0^{m-p^l}\left( y_0^{p^l-1}-1\right) (x_0+1)^{p^l}} = \frac{y_0^{p^l-1}(y_0+1)^{p^l}\left( x_0^{p^l-1}-1\right) }{x_0^{p^l-1}(x_0+1)^{p^l}\left( y_0^{p^l-1}-1\right) }. \end{aligned}$$

After simplification, we get that \(F_{p^l}\) and \(F_{p^l+1}\) have a common factor if and only if

$$\begin{aligned} y_0(x_0+1)^{p^l}\left( y_0^{p^l-1}-1\right) ^{p^l+1}=x_0(y_0+1)^{p^l}\left( x_0^{p^l-1}-1\right) ^{p^l+1}. \end{aligned}$$
(5)

If \((x_0,y_0)\) is not a solution of (5), then \(\mathrm {gcd}(H_{p^l},H_{p^l+1})=1\) and by Lemma 1, \(I_t(u,v)=0\).

Otherwise, we write \(u(x+x_0,y+y_0)=U_r+U_{r+1}+\cdots \), where \(U_i\) is the homogeneous polynomial composed of the terms of degree \(i\) of \(u(x+x_0,y+y_0)\) and \(U_r\ne 0\) and \(v(x+x_0,y+y_0)=V_s+V_{s+1}+\cdots \), where \(V_i\) is the homogeneous polynomial composed of the terms of degree \(i\) of \(v(x+x_0,y+y_0)\) and \(V_s\ne 0\). If \(r=0\) or \(s=0\) then \(t\) is not a point of \(u\) or \(v\) and \(I_t(u,v)=0\). Assume that \(r,\,s>0\). Since \((x_0,y_0)\) satisfies (5), \(F_{p^l}\) and \(F_{p^l+1}\) have a common factor that we denote by \(e\). We have \(H_{p^l}=U_rV_s=e^{p^l}\) and \(H_{p^l+1}=U_rV_{s+1}+U_{r+1}V_s\). Furthermore, \(\mathrm {gcd}(F_{p^l},F_{p^l+1})=e\) and thus \(\mathrm {gcd}(H_{p^l},H_{p^l+1})=e\). Since \(r\ge 1\) and \(s\ge 1,\,e\) divides \(U_r\) and \(V_s\) and consequently \(\mathrm {gcd}(U_r,V_s)\). If \(\mathrm {gcd}(U_r,V_s)=e^k,\,e^k\) divides \(\mathrm {gcd}(H_{p^l},H_{p^l+1})\) thus \(\mathrm {gcd}(U_r,V_s)=e\). We can assume without loss of generality that \(U_r=e^{p^l-1}\) and \(V_s=e\). Since \(m_t(v)=1,\,I_t(u,v)=\mathrm {ord}_t^v(u)\). Since \(e^2\) does not divide \(H_{p^l+1},\,e\) does not divide \(U_{p^l}\) and we can write \(U_{p^l}\) as the product of \(p^l\) linear factors distinct from \(e\). Each factor is not tangent to \(v\), so the order of each factor is 1 (see [4, p. 70]). Thus the order of \(U_{p^l}\) is \(p^l\) and \(\mathrm {ord}_t^v(u)\le p^l\).

Finally, we get

Lemma 11

If \(t=(x_0,y_0)\) is an affine singular point of \(h\) such that \(x_0\) and \(y_0\not \in \mathbb {F}_{p^l}^*\) and \(x_0\ne y_0\) then

  • \(I_t(u,v)=0\) if \(y_0(x_0+1)^{p^l}\big (y_0^{p^l-1}-1\big )^{p^l+1}\ne x_0(y_0+1)^{p^l}\big (x_0^{p^l-1}-1\big )^{p^l+1}\)

  • otherwise, \(I_t(u,v)\le p^l\); and there are at most \(((p^l-2)(p^l+1)+1)\big (\frac{m-1}{p^l}-1\big )\) such singular points.

5 Proof of Theorem 3

The following theorems prove Theorem 3. From now, assume \(m\ne 1+p^l\). We write \(m=1+\sum _{j=1}^b m_j p^{i_j}\) with \(1\le m_j\le p-1,\,i_j>i_{j-1},\,i_1=l\).

Theorem 4

If \(d=1\) then \(h\) has an absolutely irreducible factor over \(\mathbb {F}_p\).

Proof

Suppose that \(d=1\). Assume \(h\) has no absolutely irreducible factor over \(\mathbb {F}_p\), then by Lemma 2 we have \(e=\frac{I_{tot}}{\frac{(m-2)^2}{4}}\ge \frac{8}{9}\) where \(I_{tot}\) is an upper bound on the global intersection number for any factorization \(h=u\cdot v\). Since \(d=1\), we only have singularities of type Ib, IIc, IIIa and IIIc (see Table 1). So, by Table 1, we can take

$$\begin{aligned} I_{tot}=p^l\left( \frac{m-1}{p^l}-1\right) \left( 2\frac{m-1}{p^l}-(m_b+1)p^{i_b-l}-1\right) +\left( \frac{p^l-1}{2}\right) ^2. \end{aligned}$$
(6)

Since \(m=1+p^lk\) and \(m\ne 1+p^l,\,k\ge 2\); thus \(\frac{m-3}{4}=\frac{p^lk-2}{4}\ge \frac{p^l-1}{2}\). Hence

$$\begin{aligned} e&\le \frac{1}{\frac{(m-2)^2}{4}}\left( \frac{(m-3)^2}{16}+p^l\left( \frac{m-1}{p^l}-1\right) ^2\right) \\&\le \frac{1}{4}+\frac{4}{p^l}. \end{aligned}$$

For \(p^l\ne 3\) or 5, we have \(e<\frac{8}{9}\) which is a contradiction.

First, consider the case where \(p^l=3\). We have \(1=d=\mathrm {gcd}(2,k)\) so \(k\) is odd and 3 does not divide \(k\) by definition of \(l\). Hence \(k\ge 5\), thus, by Lemma 11

$$\begin{aligned} e&\le \frac{p^l((p^l-2)(p^l+1)+1)\left( \frac{m-1}{p^l}-1\right) +\left( \frac{p^l-1}{2}\right) ^2}{\frac{(m-2)^2}{4}}=\frac{15(k-1)+1}{\frac{(3k-1)^2}{4}}. \end{aligned}$$

However, for \(k\ge 5,\,k\mapsto \frac{15(k-1)+1}{\frac{(3k-1)^2}{4}}\) is a decreasing function. So, for \(k\ge 11,\,e<\frac{8}{9}\). Now we have to consider the case where \(k=5\) and \(k=7\). Using Eq. 6, we have

In all cases we get a contradiction since \(e<\frac{8}{9}\).

If \(p^l=5\), then \(1=d=\mathrm {gcd}(4,k)\) and \(k\) is odd. Hence, \(k=3\) or \(k\ge 7\). As in the case where \(p^l=3,\,e\le \frac{95(k-1)+4}{\frac{(5k-1)^2}{4}}\). However \(k\mapsto \frac{95(k-1)+4}{\frac{(5k-1)^2}{4}}\) is a decreasing function for \(k\ge 3\). so, for \(k\ge 17,\,e<\frac{8}{9}\) which is a contradiction. We now have to consider the case where \(k=3\), 7, 9, 11, 13. Using Eq. 6, we have

In all case, \(e<\frac{8}{9}\) which is a contradiction. \(\square \)

Theorem 5

If \(1<d<\frac{m-1}{p^l},\,h\) has an absolutely irreducible factor over \(\mathbb {F}_p\).

Proof

Suppose that \(1<d<\frac{m-1}{p^l}\). Assume \(h\) has no absolutely irreducible factor over \(\mathbb {F}_p\), then by Lemma 2, we have \(e=\frac{I_{tot}}{\frac{(m-2)^2}{4}}\ge \frac{8}{9}\) where \(I_{tot}\) is an upper bound on the global intersection number for any factorization of \(h=u\cdot v\). By Table 1, we can take:

$$\begin{aligned} I_{tot}&= \frac{p^{2l}-1}{4}(d-1)+\left( \frac{p^l-1}{2}\right) ^2\\&+\,\,p^l\left( \left( \frac{m-1}{p^l}-1\right) \left( 2\frac{m-1}{p^l}-(m_b+1)p^{i_b-l}-1\right) -(d-1)(d-2)\right) \\&+\,\,\left( \frac{p^l+1}{2}\right) ^2(d-1)(d-2)+(d-1)\frac{p^{2l}-1}{4}\\&\le \frac{p^{2l}-1}{2}(d-1)+\left( \frac{p^l-1}{2}\right) ^2(d-1)(d-2)\\&+\,\,p^l\left( \frac{m-1}{p^l}-1\right) ^2+\left( \frac{p^l-1}{2}\right) ^2. \end{aligned}$$

However, \(m=1+kp^l\) with \(k\ne 1\). Since \(d\) divides \(k\) and \(d<k\), we have \(d\le \frac{m-1}{2p^l}\). Hence,

$$\begin{aligned} e&\le \frac{2(p^{2l}-1)\left( \frac{k}{2}-1\right) +(p^l-1)^2\left( \frac{k}{2}-1\right) \left( \frac{k}{2}-2\right) +4p^l(k-1)^2+(p^l-1)^2}{(p^lk-1)^2}\\&\le \frac{1}{\left( k-\frac{1}{p^l}\right) ^2}\left( \left( 1-\frac{1}{p^{2l}}\right) (k-2)+\frac{1}{4}\left( 1-\frac{1}{p^l}\right) ^2(k-2)(k-4)\right. \\&+\left. \frac{4}{p^l}(k-1)^2+\left( 1-\frac{1}{p^l}\right) ^2\right) \\ e&\le \frac{1}{k-\frac{1}{p^l}}+\frac{1}{4}+\frac{4}{p^l}+\frac{1}{\left( k-\frac{1}{p^l}\right) ^2}. \end{aligned}$$

Since \(e\ge \frac{8}{9},\,1<d<k\) and \(\mathrm {gcd}(k,p)=1\), the only possibilities are:

On one hand, we have

$$\begin{aligned} e&\le \frac{2(p^{2l}-1)(d-1)+(p^l+1)^2(d-1)(d-2)}{(p^lk-1)^2}\nonumber \\&+\frac{4p^l(k-1)((p^l-2)(p^l+1)+1)+(p^l-1)^2}{(p^lk-1)^2}. \end{aligned}$$
(7)

On the other hand, we have:

$$\begin{aligned} e&\le \frac{2(p^{2l}-1)(d-1)+(p^l+1)^2(d-1)(d-2)}{(p^lk-1)^2}\nonumber \\&+\frac{4p^l(k-1)(2k-(m_b+1)p^{i_b-l}-1)+(p^l-1)^2}{(p^lk-1)^2}. \end{aligned}$$
(8)

First, consider the case where \(k\ge 16\). In inequality (7), e is bounded by a decreasing function of \(k\). Furthermore, if \(p^l=3\) and \(k=16\) or if \(k=17\) and \(p^l=5\) the upper bound in (7) is less than \(\frac{8}{9}\) which leaves only the case \(k=16\) and \(p^l=5\). But replacing in Eq. 8, we also get a contradiction. In the other cases, using inequality (7) or inequality (8), we have \(e<\frac{8}{9}\) which is a contradiction. \(\square \)

Theorem 6

If \(d=\frac{m-1}{p^l}\ne p^l-1\) then \(h\) has an absolutely irreducible factor over \(\mathbb {F}_p\).

Proof

Suppose that \(d=\frac{m-1}{p^l}\ne p^l-1\). First, we make some remarks. Since \(d=\frac{m-1}{p^l}\), there are only singularities of type Ia, IIa, IIIa, IIIb (see Table 1). In all these cases, the tangent lines of \(h\) in any singular point are simple. So, for all factorization \(h=uv,\,I_t(u,v)=m_t(u)m_t(v)\). Furthermore, since \(\frac{m-1}{p^l}\ne p^l-1,\,\frac{m-1}{p^l}\le \frac{p^l-1}{2}\). Assume that \(h\) has no absolutely irreducible factor over \(\mathbb {F}_p\). We write \(h=h_1\ldots h_r\) where each \(h_i\) factorizes into \(c_i\ge 2\) factors on an algebraic closure of \(\mathbb {F}_p\) and its factors are all of degree \(\frac{\mathrm {deg}(h_i)}{c_i}\). We write \(h_i=h_{i,1}\ldots h_{i,c_i}\). Then

$$\begin{aligned} A&= \sum _{k=1}^r\sum _{1\le i<j\le c_k}\sum _t I_t(h_{k,i},h_{k,j})+\sum _{1\le k<l\le r}\sum _{\begin{array}{c}1\le i\le c_k\\ 1\le j\le c_l\end{array}}\sum _t I_t(h_{k,i},h_{l,j})\\&= \sum _{k=1}^r\sum _{1\le i<j\le c_k}\sum _t m_t(h_{k,i})m_t(h_{k,j})+\sum _{1\le k<l\le r}\sum _{\begin{array}{c}1\le i\le c_k\\ 1\le j\le c_l\end{array}}\sum _t m_t(h_{k,i})m_t(h_{l,j}). \end{aligned}$$

However,

$$\begin{aligned} (m_t(h))^2&= \left( \sum _{k=1}^rm_t(h_k)\right) ^2\\&= \sum _{k=1}^rm_t(h_k)^2+2\sum _{1\le k<l\le r}m_t(h_k)m_t(h_l)\\&= \sum _{k=1}^rm_t(h_k)^2+2\sum _{1\le k<l\le r}\sum _{\begin{array}{c}1\le i\le c_k\\ 1\le j\le c_l\end{array}} m_t(h_{k,i})m_t(h_{l,j}). \end{aligned}$$

So, by Lemma 3,

$$\begin{aligned} A\le \sum _t\left( \sum _{k=1}^rm_t(h_k)^2\frac{c_k-1}{2c_k}+\frac{1}{2}\left( m_t(h)^2-\sum _{k=1}^rm_t(h_k)^2\right) \right) , \end{aligned}$$

thus

$$\begin{aligned} A\le \frac{1}{2}\sum _t\left( m_t(h)^2-\sum _{k=1}^r\frac{m_t(h_k)^2}{c_k}\right) . \end{aligned}$$

On the other hand, by Bézout’s theorem,

$$\begin{aligned} A&= \sum _{k=1}^r\sum _{1\le i<j\le c_k}\deg (h_{k,i})\deg (h_{k,j})+\sum _{1\le k<l\le r}\sum _{\begin{array}{c}1\le i\le c_k\\ 1\le j\le c_l\end{array}}\deg (h_{k,i})\deg (h_{l,j})\\&= \sum _{k=1}^r\frac{\deg (h_k)^2}{c_k^2}\frac{c_k(c_k-1)}{2}+\sum _{1\le k<l\le r}\deg (h_k)\deg (h_l)\\&= \sum _{k=1}^r\deg (h_k)^2\frac{c_k-1}{2c_k}+\frac{1}{2}\left( \deg (h)^2-\sum _{k=1}^r\deg (h_k)^2\right) \\&= \frac{1}{2}\left( \deg (h)^2-\sum _{k=1}^r\frac{\deg (h_k)^2}{c_k}\right) . \end{aligned}$$

Hence,

$$\begin{aligned} \deg (h)^2-\sum _{k=1}^r\frac{\deg (h_k)^2}{c_k}\le \sum _t\left( m_t(h)^2-\sum _{k=1}^r\frac{m_t(h_k)^2}{c_k}\right) . \end{aligned}$$

Then, by Lemma 3,

$$\begin{aligned} \deg (h)^2-\sum _tm_t(h)^2\le \sum _{k=1}^r\frac{1}{c_k}\left( \deg (h_k)^2-\sum _tm_t(h_k)^2\right) \le 0. \end{aligned}$$

We set \(k=\frac{m-1}{p^l}\). Then

$$\begin{aligned} \deg (h)^2\le \sum _tm_t(h)^2&\Leftrightarrow (m-2)^2\le 2(k-1)p^{2l}\\&+\,\,(k-1)(k-2)(1+p^l)^2+(p^l-1)^2\\&\Leftrightarrow -(2p^l+1)k^2+(p^{2l}+4p^l+3)k-(p^{2l}+2p^l+2)\le 0\\&\Leftrightarrow k\le 1\hbox { or }k\ge \frac{p^{2l}+2p^l+2}{2p^l+1}. \end{aligned}$$

However, \(k\ge 2\) (\(m\ne 1+p^l\)) and \(k\le \frac{p^l-1}{2}<\frac{p^{2l}+2p^l+2}{2p^l+1}\) which is a contradiction. \(\square \)