Imaginary hyperelliptic curve

Summary

A hyperelliptic curve is a particular kind of algebraic curve. There exist hyperelliptic curves of every genus . If the genus of a hyperelliptic curve equals 1, we simply call the curve an elliptic curve. Hence we can see hyperelliptic curves as generalizations of elliptic curves. There is a well-known group structure on the set of points lying on an elliptic curve over some field , which we can describe geometrically with chords and tangents. Generalizing this group structure to the hyperelliptic case is not straightforward. We cannot define the same group law on the set of points lying on a hyperelliptic curve, instead a group structure can be defined on the so-called Jacobian of a hyperelliptic curve. The computations differ depending on the number of points at infinity. Imaginary hyperelliptic curves are hyperelliptic curves with exactly 1 point at infinity: real hyperelliptic curves have two points at infinity.

Formal definition edit

Hyperelliptic curves can be defined over fields of any characteristic. Hence we consider an arbitrary field   and its algebraic closure  . An (imaginary) hyperelliptic curve of genus   over   is given by an equation of the form

 
where   is a polynomial of degree not larger than   and   is a monic polynomial of degree  . Furthermore, we require the curve to have no singular points. In our setting, this entails that no point   satisfies both   and the equations   and  . This definition differs from the definition of a general hyperelliptic curve in the fact that   can also have degree   in the general case. From now on we drop the adjective imaginary and simply talk about hyperelliptic curves, as is often done in literature. Note that the case   corresponds to   being a cubic polynomial, agreeing with the definition of an elliptic curve. If we view the curve as lying in the projective plane   with coordinates  , we see that there is a particular point lying on the curve, namely the point at infinity   denoted by  . So we could write  .

Suppose the point   not equal to   lies on the curve and consider  . As   can be simplified to  , we see that   is also a point on the curve.   is called the opposite of   and   is called a Weierstrass point if  , i.e.  . Furthermore, the opposite of   is simply defined as  .

Alternative definition edit

The definition of a hyperelliptic curve can be slightly simplified if we require that the characteristic of   is not equal to 2. To see this we consider the change of variables   and  , which makes sense if char . Under this change of variables we rewrite   to   which, in turn, can be rewritten to  . As   we know that   and hence   is a monic polynomial of degree  . This means that over a field   with char  every hyperelliptic curve of genus   is isomorphic to one given by an equation of the form   where   is a monic polynomial of degree   and the curve has no singular points. Note that for curves of this form it is easy to check whether the non-singularity criterion is met. A point   on the curve is singular if and only if   and  . As   and  , it must be the case that   and thus   is a multiple root of  . We conclude that the curve   has no singular points if and only if   has no multiple roots. Even though the definition of a hyperelliptic curve is quite easy when char , we should not forget about fields of characteristic 2 as hyperelliptic curve cryptography makes extensive use of such fields.

Example edit

 
Figure 1: Example of a hyperelliptic curve

As an example consider   where   over  . As   has degree 5 and the roots are all distinct,   is a curve of genus  . Its graph is depicted in Figure 1.

From this picture it is immediately clear that we cannot use the chords and tangents method to define a group law on the set of points of a hyperelliptic curve. The group law on elliptic curves is based on the fact that a straight line through two points lying on an elliptic curve has a unique third intersection point with the curve. Note that this is always true since   lies on the curve. From the graph of   it is clear that this does not need to hold for an arbitrary hyperelliptic curve. Actually, Bézout's theorem states that a straight line and a hyperelliptic curve of genus 2 intersect in 5 points. So, a straight line through two point lying on   does not have a unique third intersection point, it has three other intersection points.

Coordinate ring edit

The coordinate ring of C over K is defined as

 

The polynomial   is irreducible over  , so

 

is an integral domain.

Proof

If r (x,y) were reducible over  , it would factor as (yu(x))⋅(yv(x)) for some  . But then u(x)⋅v(x) = f(x) so it has degree 2g + 1, and u(x) + v(x) = h(x) so it has degree smaller than g, which is impossible.

Note that any polynomial function   can be written uniquely as

    with  

Norm and degree edit

The conjugate of a polynomial function G(x, y) = u(x) − v(x)y in   is defined to be

 

The norm of G is the polynomial function  . Note that N(G) = u(x)2 + u(x)v(x)h(x) − v(x)2f(x), so N(G) is a polynomial in only one variable.

If G(x, y) = u(x) − v(x) ⋅ y, then the degree of G is defined as

 

Properties:

 
 
 

Function field edit

The function field K(C) of C over K is the field of fractions of K[C], and the function field   of C over   is the field of fractions of  . The elements of   are called rational functions on C. For R such a rational function, and P a finite point on C, R is said to be defined at P if there exist polynomial functions G, H such that R = G/H and H(P) ≠ 0, and then the value of R at P is

 

For P a point on C that is not finite, i.e. P =  , we define R(P) as:

If    then  , i.e. R has a zero at O.
If    then    is not defined, i.e. R has a pole at O.
If    then    is the ratio of the leading coefficients of G and H.

For   and  ,

If   then R is said to have a zero at P,
If R is not defined at P then R is said to have a pole at P, and we write  .

Order of a polynomial function at a point edit

For   and  , the order of G at P is defined as:

  if P = (a,b) is a finite point which is not Weierstrass. Here r is the highest power of (xa) which divides both u(x) and v(x). Write G(x,y) = (xa)r(u0(x) − v0(x)y) and if u0(a) − v0(a)b = 0, then s is the highest power of (xa) which divides N(u0(x) − v0(x)y) = u02 + u0v0hv02f, otherwise, s = 0.
  if P = (a,b) is a finite Weierstrass point, with r and s as above.
  if P = O.

The divisor and the Jacobian edit

In order to define the Jacobian, we first need the notion of a divisor. Consider a hyperelliptic curve   over some field  . Then we define a divisor   to be a formal sum of points in  , i.e.   where   and furthermore   is a finite set. This means that a divisor is a finite formal sum of scalar multiples of points. Note that there is no simplification of   given by a single point (as one might expect from the analogy with elliptic curves). Furthermore, we define the degree of   as  . The set of all divisors   of the curve   forms an Abelian group where the addition is defined pointwise as follows  . It is easy to see that   acts as the identity element and that the inverse of   equals  . The set   of all divisors of degree 0 can easily be checked to be a subgroup of  .
Proof. Consider the map   defined by  , note that   forms a group under the usual addition. Then   and hence   is a group homomorphism. Now,   is the kernel of this homomorphism and thus it is a subgroup of  .

Consider a function  , then we can look at the formal sum div . Here   denotes the order of   at  . We have that ord  if   has a pole of order   at  ,   if   is defined and non-zero at   and   if   has a zero of order   at  .[1] It can be shown that   has only a finite number of zeroes and poles,[2] and thus only finitely many of the ord  are non-zero. This implies that div  is a divisor. Moreover, as  ,[2] it is the case that   is a divisor of degree 0. Such divisors, i.e. divisors coming from some rational function  , are called principal divisors and the set of all principal divisors   is a subgroup of  .
Proof. The identity element   comes from a constant function which is non-zero. Suppose   are two principal divisors coming from   and   respectively. Then   comes from the function  , and thus   is a principal divisor, too. We conclude that   is closed under addition and inverses, making it into a subgroup.

We can now define the quotient group   which is called the Jacobian or the Picard group of  . Two divisors   are called equivalent if they belong to the same element of  , this is the case if and only if   is a principal divisor. Consider for example a hyperelliptic curve   over a field   and a point   on  . For   the rational function   has a zero of order   at both   and   and it has a pole of order   at  . Therefore, we find   and we can simplify this to   if   is a Weierstrass point.

Example: the Jacobian of an elliptic curve edit

For elliptic curves the Jacobian turns out to simply be isomorphic to the usual group on the set of points on this curve, this is basically a corollary of the Abel-Jacobi theorem. To see this consider an elliptic curve   over a field  . The first step is to relate a divisor   to every point   on the curve. To a point   on   we associate the divisor  , in particular   in linked to the identity element  . In a straightforward fashion we can now relate an element of   to each point   by linking   to the class of  , denoted by  . Then the map   from the group of points on   to the Jacobian of   defined by   is a group homomorphism. This can be shown by looking at three points on   adding up to  , i.e. we take   with   or  . We now relate the addition law on the Jacobian to the geometric group law on elliptic curves. Adding   and   geometrically means drawing a straight line through   and  , this line intersects the curve in one other point. We then define   as the opposite of this point. Hence in the case   we have that these three points are collinear, thus there is some linear   such that  ,   and   satisfy  . Now,   which is the identity element of   as   is the divisor on the rational function   and thus it is a principal divisor. We conclude that  .

The Abel-Jacobi theorem states that a divisor   is principal if and only if   has degree 0 and   under the usual addition law for points on cubic curves. As two divisors   are equivalent if and only if   is principal, we conclude that   and   are equivalent if and only if  . Now, every nontrivial divisor of degree 0 is equivalent to a divisor of the form  , this implies that we have found a way to ascribe a point on   to every class  . Namely, to   we ascribe the point  . This maps extends to the neutral element 0 which is maped to  . As such the map   defined by   is the inverse of  . So   is in fact a group isomorphism, proving that   and   are isomorphic.

The Jacobian of a hyperelliptic curve edit

The general hyperelliptic case is a bit more complicated. Consider a hyperelliptic curve   of genus   over a field  . A divisor   of   is called reduced if it has the form   where  ,   for all   and   for  . Note that a reduced divisor always has degree 0, also it is possible that   if  , but only if   is not a Weierstrass point. It can be proven that for each divisor   there is a unique reduced divisor   such that   is equivalent to  .[3] Hence every class of the quotient group   has precisely one reduced divisor. Instead of looking at   we can thus look at the set of all reduced divisors.

Reduced divisors and their Mumford representation edit

A convenient way to look at reduced divisors is via their Mumford representation. A divisor in this representation consists of a pair of polynomials   such that   is monic,   and  . Every non-trivial reduced divisor can be represented by a unique pair of such polynomials. This can be seen by factoring   in   which can be done as such as   is monic. The last condition on   and   then implies that the point   lies on   for every  . Thus   is a divisor and in fact it can be shown to be a reduced divisor. For example, the condition   ensures that  . This gives the 1-1 correspondence between reduced divisors and divisors in Mumford representation. As an example,   is the unique reduced divisor belonging to the identity element of  . Its Mumford representation is   and  . Going back and forth between reduced divisors and their Mumford representation is now an easy task. For example, consider the hyperelliptic curve   of genus 2 over the real numbers. We can find the following points on the curve  ,   and  . Then we can define reduced divisors   and  . The Mumford representation of   consists of polynomials   and   with   and we know that the first coordinates of   and  , i.e. 1 and 3, must be zeroes of  . Hence we have  . As   and   it must be the case that   and   and thus   has degree 1. There is exactly one polynomial of degree 1 with these properties, namely  . Thus the Mumford representation of   is   and  . In a similar fashion we can find the Mumford representation   of  , we have   and  . If a point   appears with multiplicity n, the polynomial v needs to satisfy   for  .

Cantor's algorithm edit

There is an algorithm which takes two reduced divisors   and   in their Mumford representation and produces the unique reduced divisor  , again in its Mumford representation, such that   is equivalent to  .[4] As every element of the Jacobian can be represented by the one reduced divisor it contains, the algorithm allows to perform the group operation on these reduced divisors given in their Mumford representation. The algorithm was originally developed by David G. Cantor (not to be confused with Georg Cantor), explaining the name of the algorithm. Cantor only looked at the case  , the general case is due to Koblitz. The input is two reduced divisors   and   in their Mumford representation of the hyperelliptic curve   of genus   over the field  . The algorithm works as follows

  1. Using the extended Euclidean algorithm compute the polynomials   such that   and  .
  2. Again with the use of the extended Euclidean algorithm compute the polynomials   with   and  .
  3. Put  ,   and  , which gives  .
  4. Set   and  .
  5. Set   and  .
  6. If  , then set   and   and repeat step 5 until  .
  7. Make   monic by dividing through its leading coefficient.
  8. Output  .

The proof that the algorithm is correct can be found in.[5]

Example edit

As an example consider the curve

 

of genus 2 over the real numbers. For the points

 ,   and  

and the reduced divisors

  and  

we know that

 , and
 

are the Mumford representations of   and   respectively.

We can compute their sum using Cantor's algorithm. We begin by computing

 , and
 

for  , and  .

In the second step we find

  and
 

for   and  .

Now we can compute

 ,
  and
 .

So

  and
 

Lastly we find

  and
 .

After making   monic we conclude that

 

is equivalent to  .

More on Cantor's algorithm edit

Cantor's algorithm as presented here has a general form, it holds for hyperelliptic curves of any genus and over any field. However, the algorithm is not very efficient. For example, it requires the use of the extended Euclidean algorithm. If we fix the genus of the curve or the characteristic of the field (or both), we can make the algorithm more efficient. For some special cases we even get explicit addition and doubling formulas which are very fast. For example, there are explicit formulas for hyperelliptic curves of genus 2[6] [7] and genus 3.

For hyperelliptic curves it is also fairly easy to visualize the adding of two reduced divisors. Suppose we have a hyperelliptic curve of genus 2 over the real numbers of the form

 

and two reduced divisors

  and
 .

Assume that

 ,

this case has to be treated separately. There is exactly 1 cubic polynomial

 

going through the four points

 .

Note here that it could be possible that for example  , hence we must take multiplicities into account. Putting   we find that

 

and hence

 .

As   is a polynomial of degree 6, we have that   has six zeroes and hence   has besides   two more intersection points with  , call them   and  , with  . Now,   are intersection points of   with an algebraic curve. As such we know that the divisor

 

is principal which implies that the divisor

 

is equivalent to the divisor

 .

Furthermore, the divisor

 

is principal for every point   on   as it comes from the rational function  . This gives that   and   are equivalent. Combining these two properties we conclude that

 

is equivalent to the reduced divisor

 .

In a picture this looks like Figure 2. It is possible to explicitly compute the coefficients of  , in this way we can arrive at explicit formulas for adding two reduced divisors.

 
Figure 2: Example of adding two elements of the Jacobian

References edit

  1. ^ Isabelle Déchène, The Picard Group, or how to build a group from a set
  2. ^ a b Alfred J. Menezes, Yi-Hong Wu, Robert J. Zuccherato, An elementary introduction to hyperelliptic curves[permanent dead link], page 15
  3. ^ Alfred J. Menezes, Yi-Hong Wu, Robert J. Zuccherato, An elementary introduction to hyperelliptic curves[permanent dead link], page 20
  4. ^ Alfred J. Menezes, Yi-Hong Wu, Robert J. Zuccherato, An elementary introduction to hyperelliptic curves[permanent dead link], page 22–27
  5. ^ Cantor, David G. (1987). "Computing in the Jacobian of a hyperelliptic curve". Mathematics of Computation. 48 (177): 95–101. doi:10.1090/S0025-5718-1987-0866101-0.
  6. ^ Frank Leitenberger, About the Group Law for the Jacobi Variety of a Hyperelliptic Curve
  7. ^ T. Lange (2005). "Formulae for Arithmetic on Genus $2$ Hyperelliptic Curves". Applicable Algebra in Engineering, Communication and Computing. 15 (5): 295–328. CiteSeerX 10.1.1.109.578. doi:10.1007/s00200-004-0154-8.