Hypergeometric Motives
1 Introduction
It must have been frustrating in the early days of calculus that an integral like
(1.1) |
appeared not to be expressible in terms of known functions. This type of integral arises in computing the movement of the ideal pendulum or the length of an arc of an ellipse for example; they have remained relevant and are connected to a great deal of the mathematics of the last 200 years.
Indeed is not an elementary function. Its Maclaurin expansion
(1.2) |
is an example of a hypergeometric series. It satisfies a linear differential equation of order two of the type brilliantly analyzed by Riemann. As mentioned by Katz [Katz-RLS, p.3], Riemann was lucky. His analysis only works because any rank two differential equation on is rigid in the sense that the local behavior of solutions around the missing points uniquely determines their global behavior.
Taking a more geometric perspective, (1.1) is presenting the function as a period of the family of elliptic curves defined by
(1.3) |
This fact implies as well that satisfies an order two linear differential equation, ultimately because is two-dimensional. ††David P. Roberts is a professor at the University of Minnesota Morris. His e-mail address is [email protected] and his research is supported by grant DMS-1601350 from the National Science Foundation. ††Fernando Rodriguez Villegas is a senior research scientist at the Abdus Salam International Centre for Theoretical Physics. His email address is [email protected].
Shifting now to more arithmetic topics, if we fix a rational number then for almost all primes the number defined by
(1.4) |
is of fundamental importance. With these as the main ingredients, one builds an -function
(1.5) |
Much of the importance of the is seen through this -function. For example, the famous Birch-Swinnerton-Dyer conjecture says that the group modulo its torsion is isomorphic to , where is the order of vanishing of at . A critical advance is the result of Wiles et al. that the function
(1.6) |
on the upper half plane is a modular form. In particular, this result implies that is at least well-defined at .
The equations displayed so far represent a standard general paradigm in arithmetic geometry. One can start with any variety over , not just the varieties (1.3). There are fully developed theories of periods and point counts, and in principle one can produce analogs of the period formulas (1.1)-(1.2) and the point count formula (1.4). Interacting now with deep but widely-believed conjectures, one can break the cohomology of into irreducible motives, study -functions like (1.5), and try to find corresponding automorphic forms like (1.6).
This survey is an informal invitation to hypergeometric motives, hereafter abbreviated HGMs; see Section 4 for their definition. We write them as , with a rational function satisfying certain conditions being the family parameter and the specialization parameter. The introductory family of examples is
(1.7) |
Rigidity makes HGMs much more tractable than general motives: periods, point counts, and other invariants are given by explicit formulas in the parameters . Our broader goal in this survey is to use HGMs to gain insight into the general theory of motives; we illustrate all topics with explicit examples throughout.
Sections 2-9 are geometric in nature. The main focus is on varieties generalizing (1.3) and the discrete aspect of periods like (1.1)-(1.2), as captured in vectors of Hodge numbers, . A theme here is that HGMs form quite a broad class of irreducible motives, as very general arise. Sections 10-15 are arithmetic in nature, with the focus being on generalizations of (1.4), (1.6), and especially (1.5). Watkins has written a very useful hypergeometric motives package [Wat] in Magma and throughout this article we indicate how to use it by including small snippets of Magma code. Together these snippets are enough to let Magma beginners numerically compute with -functions using the free online Magma calculator.
2 Hypergeometric functions
We begin by generalizing (1.1)-(1.2) and explaining how this generalization leads to family parameters.
Integrals and series.
Let , and be vectors of complex numbers with and . For define, making use of the standard Gamma function ,
Via and , (1.1) is the special case and .
Expand the denominator of the integrand of (2) via the binomial theorem and use Euler’s beta integral to evaluate the individual terms. Written in terms of Pochhammer symbols , the result is
(2.2) |
In other words, the integral (2) is an alternative definition of the standard hypergeometric power series (2.2). The case and simplifies to (1.2).
Monodromy.
An excellent general reference for hypergeometric functions is [BH], and we now give a summary sufficient for this survey. The function is in the kernel of an order differential operator with singularities only at , and . This means in particular that , initially defined on the unit disk, extends to a “multivalued function” on the thrice-punctured projective line . With respect to a given basis, this multivaluedness is codified by a representation of the fundamental group into . The fundamental group is free on and , with these elements coming from counterclockwise circular paths of radius about and respectively. To emphasize the equal status of and , it is better to present this group as generated by , , and , subject to the relation . The assumption was only imposed to present the classical viewpoint (2)-(2.2) cleanly; we henceforth drop it.
A useful fact due to Levelt is the explicit description of the matrices with respect to a certain well-chosen basis. Define polynomials
Then and are companion matrices of and , while is determined by . The matrix differs minimally from the identity in that has rank . We will henceforth consider only cases where no is an integer. This ensures that the generate an irreducible subgroup of . Moreover the representation is rigid, in the following sense: suppose , , and are conjugate to , , and respectively. Then there is a single matrix such that for all three .
Family parameters.
The parameters contain information which is irrelevant for the sequel. First, the individual and are important only modulo integers. Second, the orderings of the and do not matter. To remove these irrelevancies, we will regard the degree rational function as the primary index in the sequel, calling it the family parameter. A bonus of this shift in emphasis is that an important field is made evident, the field generated by the coefficients of and . By construction, all three lie in .
The cases which naturally have underlying motives are exactly the ones with all and rational, so that is some cyclotomic field. In this survey we will substantially simplify by restricting to cases with . With this simplification, there are two natural ways to present as follows. Write and consider its factorization into irreducible polynomials, . So the factors are cyclotomic polynomials and thus have degree the Euler totient .
In our introductory example, the ways are
(2.3) |
In general, the second way is just the canonical factorization into irreducibles, while the first way is the unique “unreduction” to products of in which no factor appears in both a numerator and denominator.
To enter a family parameter into Magma, one can use either of these two ways, as in the equivalent commands
Q:=HypergeometricData([*-2,-2,1,1,1,1*]); | |||
Q:=HypergeometricData([1,1],[2,2]); | (2.4) |
In the first method, one inputs just the gamma vector formed by subscripts on the ’s, using signs to distinguish between numerator and denominator. In the second method, one inputs just the subscripts of the denominator and then numerator ’s, these being called the cyclotomic parameters. When working with underlying varieties, the gamma vectors are so useful that we often simply write rather than . After the transition to motives, the cyclotomic presentation is generally more convenient. To simplify slightly, we henceforth require that .
Orthogonal vs. symplectic.
The number is either or under our restriction . This dichotomy is strongly felt throughout this survey. It can also be expressed in terms of the fundamental bilinear form on preserved by the monodromy group ; see [BH, §4], [RV-bez, §3.5]. In the orthogonal case, is conjugate to and is symmetric. In the symplectic case, is conjugate to , and is antisymmetric.
3 Source varieties
We now describe varieties which give rise to hypergeometric motives.
Euler varieties.
We have already generalized (1.1) to (2) and (1.2) to (2.2). Assuming briefly again, a natural generalization of (1.3) is to
(3.1) |
Here is the least common denominator of the and , and the exponents are integers such that
for and . The equations (2)-(2.2) show that a specified scalar multiple of arises as a period of this variety. However the varieties (3.1) depend on how the parameters are paired: , …, . This dependence complicates the arithmetic of these varieties, so we will use an alternative collection of varieties to define hypergeometric motives.
Canonical varieties.
The alternative varieties appear under the term “circuits” in [GKZ] and are studied at greater length in [BCM]. For a gamma vector and a complex number , define by two homogeneous equations,
(3.2) |
Here and in the sequel, we systematically use the abbreviation . The canonical variety is by definition the open subvariety on which all the homogeneous coordinates are nonzero. The point is an ordinary double point on and otherwise all the are smooth. Because of this double point, we exclude the case from consideration until Section 9.
Toric models.
From a dimension-count viewpoint, the BCM equations (3.2) for canonical varieties are inefficient. They start with the variables and use two equations and projectivization to get the desired -dimensional variety . The toric models from [GKZ] start instead with variables and present by just one equation.
To obtain a toric model from a gamma vector , one proceeds as illustrated by Table 3.1. First, for each new variable choose a row vector in which is orthogonal to the given -vector . These row vectors are required to be such that is torsion-free. Second, choose a row vector which satisfies . The toric model is then
(3.3) |
So in the example of Table 3.1, the resulting equation is
(3.4) |
with . In general, the variety is the subvariety of the torus given by the equation (3.3).
Polytopes.
A toric model gives a polytope which is an aid to understanding the . The case is readily visualizable and Figure 3.1 continues our example. In general, one interprets the column vectors of the chosen matrix as points in and is their convex hull. Let be the convex hull of all the points except the one. Normalize volume so that the standard -dimensional simplex has volume , and thus has volume . Then the volume of is . The with form one triangulation of , while the with form another. The total volume of is the important number
The common topology of the with is reflected in the combinatorics of . In the case of , the genus of is the number of lattice points on the interior, while the number of punctures is the number of lattice points on the boundary. Pick’s theorem then says that the Euler characteristic of is . In the example of Figure 3.1, . For larger ambient dimension , the situation is of course much more complicated, but always .
Compactifications.
In algebraic geometry, one normally wants to compactify a given open variety such as and there are typically many natural ways of doing it.
We already saw the compactification . It is a hypersurface of degree in the projective space defined by the first equation of (3.2). On the other hand, for any choice of matrix with all entries nonnegative, homogenization of (3.3) gives a alternative compactification .
In our continuing example , the plane curve has degree seven. In contrast, the plane curve has degree just four, this number arising as the maximum column sum of the matrix in Table 3.1. Smooth curves in these degrees have genera and respectively. For , has genus so must have bad singularities while has just a single node.
Another compactification is a major focus of [BCM]. It is typically not smooth, but only has quotient singularities. These singularities are mild in the sense that looks smooth from the viewpoint of rational cohomology, and may be ignored when discussing motives as in the next section.
4 HGMs from cohomology
Here we define HGMs and explain how their behavior is simpler than other similar motives.
Motivic formalism.
Let and be subfields of ; the case of principal interest to us is . Minimally modifying Grothendieck’s original conditional definitions, André unconditionally defined a category of pure motives over with coefficients in [And]. The formal structures of this category can best be understood in terms of a huge proreductive algebraic group over , the absolute motivic Galois group of . Then is exactly the category of representations of on finite-dimensional vector spaces.
When taking cohomology, we are always implicitly working with the complex points of a variety. For a smooth projective variety over and an integer , the singular cohomology space is an object of . The image of in the general linear group of is by definition the motivic Galois group of . The purpose of , as the rest of this survey will make clear, is to group-theoretically coordinate very concrete structures on the vector space .
Two copies of the multiplicative group play important roles in the formalism of motives. One is a normal subgroup and the other a quotient: . A rank motive is said to be of weight if the representation restricted to the subgroup consists of copies the representation . The motives all have weight . The representation of on the rank one motive corresponds to the representation of the quotient group . The motive corresponding to the representation is denoted . The motives are called the Tate twists of .
Each category of pure motives is contained in a larger category of mixed motives, where an irreducible motive has a canonical weight filtration with subquotients in . We will mention mixed motives at several junctures, but our focus is sharply on pure motives.
Definition of HGMs.
Let be a gamma vector of length with negative entries and let . The hypergeometric motive is defined from the cohomology of the affine variety [RV-Mixed]. We start with the compactly supported middle cohomology space and begin by cutting out a subquotient in two steps.
First, we eliminate the contribution of the ambient -dimensional torus to obtain the primitive subspace . Second, we take any smooth compactification of , or one with at worst mild singularities as mentioned above, and consider the image of under the natural map to . As a quotient of , the space is independent of the choice of compactification.
For example, for the compactification there is a decomposition of its middle cohomology, . It is described at the level of point counts in [BCM, Thm 1.5]. Here is zero if is odd and the sum of copies of if is even.
Finally, we define the hypergeometric motive as the Hodge-normalized Tate twist , as discussed in the next section. So has weight with specified there.
More conceptually, is a mixed motive of rank and the pure motive is its weight quotient. The passage from to is closely related to the reduction of fractions as in (2.3). In particular, has rank .
It is worth stressing that the full mixed motive is itself of great interest, with its lower weight parts playing an important role in deeper studies of hypergeometric motives.
Motivic Galois groups of HGMs.
For , one likewise gets a motive . lf is transcendental then the motivic Galois group of can be cleanly expressed in terms of the monodromy group of Section 2, as follows. If , then is finite and . If , then is infinite and is the smallest algebraic group containing both and scalars; more explicitly, is the conformal symplectic group in the symplectic case of odd , and a conformal orthogonal group in the orthogonal case of even . In the case that is algebraic, including our main case that is rational, the same identification of holds almost always.
Related motives.
The toric model viewpoint is part of the program in [GKZ] to approach algebraic geometry by emphasizing the number of terms in polynomials defining -dimensional varieties. By scaling to normalize coefficients, such varieties come in -dimensional families.
For , the Newton polytope is a simplex. An abelian group of order and some exponent acts on the single associated complex variety . The essential cases here are the Fermat varieties in , defined by
(4.1) |
The group comes from scaling the variables by roots of unity and has order and exponent . Writing , the action decomposes into one-dimensional motives in . This setting of was the focus of several influential papers of Weil from around 1950, and the rank one motives appearing are Jacobi motives.
The case corresponds to general hypergeometric motives where the and can be arbitrary rational numbers. The group now has order a divisor of and some exponent . For example, for one can add the term to (4.1); then is reduced to having order but still has exponent . The action of again decomposes in and the summands include general hypergeometric motives. Our torsion-free requirement for exponent matrices is equivalent to requiring that the column vectors affine span ; in turn, this means that our HGMs constitute exactly the case .
Much of what we are describing in this article both has simpler analogs for Jacobi motives and extends to general hypergeometric motives. Indeed [BH] and [Katz-ESDE] are in the latter setting. However the associated -functions correspond to motives that have been descended to and have rank . Because of the factor , inclusion of these other settings would only modestly increase the collection of computationally accessible -functions. Also the resulting motives in have motivic Galois groups which are more complicated than the and arising ubiquitously in our setting of .
5 Hodge numbers
One of the very first things one wants to know about a motive is its Hodge numbers. Fortunately, this desire is easily satisfied for HGMs by an appealing procedure.
Background.
For a smooth projective variety over , there is a decomposition of complex vector spaces , with complex conjugation on coefficients switching and . The Hodge numbers therefore satisfy Hodge symmetry and sum to the Betti number . Classical examples are given in (6.1)-(6.2) below.
Likewise, the rank of a weight motive is decomposed into Hodge numbers . The decomposition has a simple group-theoretic reformulation: contains a subgroup which acts on by . If either or is in , as will generally be the case for us, then Hodge symmetry continues to hold.
If a motive has Hodge numbers then the Hodge numbers of its Tate twist are . The Hodge-normalization of a pure weight motive is the Tate twist for which all the nonzero Hodge numbers are in the vector and at least one of the outer ones is nonzero.
Zigzag procedure.
The procedure we are about to describe is equivalent to a formula conjectured by Corti and Golyshev [CG] and proved by different methods in Fedorov [Fed] and [RV-Mixed]. The procedure is completely combinatorial and only depends on the interlacing pattern of the roots of and in the unit circle.
To pass from a family parameter to its Hodge vector one proceeds as illustrated by Figure 5.1. One orders the parameters and , viewed as elements in say ; for more immediate readability, we associate the colors red and blue to and respectively. One draws a point at corresponding to the smallest parameter in a Cartesian plane. One then proceeds in uniform steps from left to right, drawing a point for each parameter and then moving diagonally upwards after red points and diagonally downwards after blue points. One focuses on one color or the other, counting the number of points on horizontal lines. The numbers obtained form the Hodge vector . The red and blue dots yield the same Hodge vector but contain more information. They may be used to describe the limiting mixed Hodge structure at and respectively.
The completely intertwined case.
Complete intertwining of the and gives the extreme where the resulting Hodge vector is just . Beukers and Heckman [BH] proved that complete intertwining is exactly the condition one needs for the monodromy group to be finite. They also established the complete list of such pairs . Actually they, like Schwarz who famously treated the case more than a century earlier, worked without our standing assumption . Then one needs to require complete intertwining of all the natural conjugates of and the list obtained is longer.
In our setting of , the corresponding -vectors are of odd lengths to . There are infinite collections of length and given by coprime positive integers :
(5.1) | ||||
Here and always when discussing classification, we omit consideration of whenever is listed. In case , the canonical variety consists of just points. Removing a variable, the BCM presentation takes the form
For this presentation is already trinomial; in general, one has to make a non-trivial change of variables to pass to the trinomial presentation of given by a toric model.
Beyond the closely related collections - , there are only finitely many further , all related to Weyl groups. [BH, Table 8.3] says that, modulo the quadratic twisting operation , there are just one, five, five, and fifteen respectively for the groups , , , and . One of the cases is discussed in Section 7 and the remaining cases are similarly treated in [Rob-polys].
The completely separated case.
Complete separation of the and gives the extreme where the resulting Hodge vector is . The subcase where has the simplifying feature that consists of a single Jordan block. Families in this subcase have received particular attention in the literature; the condition is sometimes verbalized as MUM, for maximal unipotent monodromy.
Classification of families in the completely separated case is easier than in the completely intertwined case. It becomes trivial in the MUM subcase because is arbitrary except for the fact that it contains no factors of . Accordingly, the number of rank families in the MUM subcase is given by a generating function
Always as under the MUM restriction multiplying by gives a bijection on parameters. Arithmetic information about the list underlying is in [RV]. For general , the case is the “mirror dual” of the Dwork case discussed after (4.1), and so motives of this family have been given special attention in the physics literature.
Signature and the Magma implementation.
A motive defined over a subfield of has a signature , which
is the trace of complex conjugation. For odd weight motives, it
is always zero. For even weight HGMs , it depends only on and
the interval , , or in which lies.
Magma’s command HodgeStructure
returns
both the Hodge vector and the signature in coded form.
To see just the Hodge vector clearly, one can
implement Q
as in (2.4) and
get the Hodge vector from say
HodgeVector(HodgeStructure(Q,2));
For example, from the gamma vector one gets the Hodge vector .
6 Projective Hypersurfaces
Here we realize some HGMs in the cohomology of the most classical varieties of all, smooth hypersurfaces in projective space.
Hodge numbers.
Let be a smooth hypersurface of degree . Let be the primitive part of its middle cohomology, meaning the part that does not come from the ambient projective space. If is odd, then this primitive part is all of . If is even, then the complementary piece that we are discarding is .
Hirzebruch gave a formula for the Hodge numbers of as a function of and . For example, the sum of the Hodge numbers and first Hodge number are respectively
(6.1) |
These special cases and Hodge symmetry are sufficient to give Hodge vectors when :
(6.2) |
For , either part of (6.1) reduces to the genus formula for smooth plane curves, .
One example for every .
Let be a desired degree and let be a desired dimension. Define
(6.3) |
The toric procedure illustrated by Table 3.1 yields the completed canonical variety:
(6.4) |
The necessary orthogonality relations on each variable’s exponents are illustrated by the case of cubic fourfolds where . Then (6.4) becomes
For , the relation is that is orthogonal to . In general, partial derivatives of (6.4) are very simple since the row vectors have just two nonzero entries, except for the case and its three nonzero entries. It is then a pleasant exercise to check via the Jacobian criterion that is smooth for .
The degree of the rational function determined by can be computed uniformly in as the cancellations to be analyzed are very structured. This degree agrees with the Betti number from (6.1). Thus is the full primitive middle cohomology of , while a priori it might have been a proper subspace. The zigzag procedure for computing Hodge numbers must agree in the end with the Hirzebruch formula. The reader might want to check the above case of cubic fourfolds, where Hirzebruch’s full formula gives .
All examples for a given .
An interesting problem is to find all which give projective smooth -folds of degree . For small parameters, this problem can be solved by direct computation. For example, consider , thus cubic fourfolds. In this case, one has the standardization of the above example, and then exactly ten more:
7 Dimension reduction
An HGM is defined in terms of a -dimensional variety but its Hodge vector raises the question of whether it also comes from a variety of dimension . The exterior zeros for low degree projective hypersurfaces as illustrated in (6.2) raise the same question. The generalized Hodge conjecture says that this dimension reduction is always possible. We illustrate here some of the appealing geometry that arises from reducing dimension.
Reduction to points.
When the reduction to dimension zero is possible in all cases. For example, corresponds to entry 45 on the Beukers-Heckman list [BH, Table 8.3]. Formula (3.3) then gives a family of cubic surfaces. An equation whose roots correspond to the famous twenty-seven lines on is
(7.1) |
The Galois group of this polynomial for generic is . It has elements and is also the monodromy group .
Reduction via splicing.
Suppose can be written as the concatenation of two lists each summing to zero. Then one can use a general splicing technique from [BCM, §6] to reduce the dimension by two. This technique is behind the scenes even of our introduction: in the family of examples there, the canonical varieties for are three-dimensional, although the more familiar source varieties are just the Legendre curves (1.3).
For an example complicated enough to be representative of the general case, take so that the canonical variety has dimension . The Hodge vector is just , so one would like to realize in the cohomology of a curve.
Splicing is possible because both and sum to zero. No further splicing is possible, but fortunately we have just treated the first sublist by other means. Splicing corresponds to taking a fiber product over the -line which in turn corresponds to just multiplying rational functions. In our case, solving (7.1) for to get the first factor, the dimension-reduced variety is given by
(7.2) |
The variable from enters only quadratically and so (7.2) defines a double cover of the -line. Taking the discriminant with respect to and removing unneeded square factors presents this hyperelliptic curve in standard form:
As the right side has degree 28, this curve has genus .
In both the new examples of this section, the middle cohomology of the dimension-reduced varieties contains not only the desired motives, with Hodge vectors and respectively, but also parasitical motives, with Hodge vectors and . In this regard, they are less attractive than the original canonical varieties. HGMs provide many illustrations like these two of the motivic principle that a motive comes from many varieties , and often no single should be viewed as the best source.
8 Distribution of Hodge vectors
In this section, we explain one of the great features of HGMs: they represent many Hodge vectors.
Completeness in ranks .
By direct computation starting from all family parameters in degrees , we have verified the following fact. Let be a vector of positive integers satisfying for all and let . Then if there exists an HGM with Hodge vector .
Many families per Hodge vector in ranks .
In ranks to , the only vectors not realized by a family of HGMs are
Table 8.1 gives a fuller sense of the situation for , where there are about family parameters. It gives the extremes of the list of possible Hodge vectors , sorted by how many families realize .
The ratio of the numbers just reported say that the number of family parameters per Hodge vector in degree is about . This ratio increases to a maximum at where it is about four million. It then decreases to zero, with some approximate sample values being two million for but only for . These numbers are computed via generating functions, similar to (5) but more complicated.
Perspective.
Section 6 offers some perspective on the general inverse problem of finding an irreducible motive with a given Hodge vector. From (6.1)-(6.2), one sees that the Hodge vectors coming from hypersurfaces are very sparse. When one looks at broader standard classes of varieties, such as complete intersections in projective spaces, more Hodge vectors arise, but they all have the same rough form: bunched in the middle. Ad hoc techniques, such as reducing Hodge numbers by imposing singularities, give many more Hodge vectors. But for many , it does not seem easy to find a corresponding motive and then prove irreducibility in this geometric way. For example, imposing ordinary double points on a sextic surface reduces the Hodge vector to . However the family of sextic surfaces is only -dimensional, and so it would it seem to be difficult to get down to e.g. . There does not seem to be even a conjectural expectation of which Hodge vectors arise from irreducible motives in .
The cases .
One could go into much more detail about the families behind any given Hodge vector. Here we say a little more about the cases , which are particularly interesting for several reasons. The giving Hodge vectors of the form typically have canonical dimension greater than two, posing instances of the dimension reduction problem. If , then the moduli theory of surfaces says that there is at least one family of surfaces also realizing . Finding such a family is a challenge.
Cases with present a greater challenge, as they cannot be realized by K3 surfaces. There are seventy-two parameters giving . None of the eleven listed in Section 6 can be spliced, underscoring the difficulty of dimension reduction. One of the four gamma vectors giving has canonical dimension eight, namely . The other three have canonical dimension ten:
In all four cases, there are many ways to splice, but no path to a surface.
9 Special and semi HGMs
We have so far been excluding the singular specialization point from consideration. Now we explain how it yields a particularly interesting motive . We also explain how other interesting motives arise when the family parameter is reflexive, in the sense of satisfying .
Interior zeros.
A Hodge-normalized motive of weight has Hodge vector with . But for the Hodge vectors explicitly considered so far, the remaining numbers are also positive. There is a reason for this restriction: Griffiths transversality says that any collection of motives moving in a family with irreducible monodromy group has Hodge vector with no interior zeros. Special and semi HGMs do not move in families, and they include cases with interior zeros.
Special HGMs.
The way to account for the double point on the canonical variety is to first of all take inertial invariants with respect to the monodromy operator . In the orthogonal case, this already give the right motive . Its Hodge vector differs from the generic Hodge vector only in that is decreased by . In the symplectic case, the motive of inertial invariants is mixed, and quotienting out by its submotive of weight and rank gives . Its Hodge vector now comes from the generic one by decreasing the two central Hodge numbers by . These drops obviously can cause interior zeros, as in or .
Semi HGMs.
For a reflexive parameter and any , the motives and are quadratic twists of one another. The interest in reflexive parameters is that nongeneric behavior is thereby forced at . The motive is a direct sum of two motives in of roughly equal rank. We call the summands semi HGMs and their Hodge vectors can have many interior zeros. For example, the summands of are studied in [Rob] and the two Hodge vectors are
(9.1) |
There is a similar decomposition of , but only after viewing it in .
10 Point counts
We now turn to arithmetic. The point counts of this section form the principal raw material from which the -functions studied in the remaining sections are built.
Background.
Let be a smooth projective variety over . Then for all primes outside a finite set , the equations defining have good reduction and so define a smooth projective variety over . For any power , one has the finite set of solutions to the defining equations. The key invariants that need to be input into the motivic formalism are the cardinalities , and famous results of Grothendieck, Deligne, and others provide the tools.
The vector spaces do not see that is defined over . The arithmetic origin of yields extra structure as follows. For any prime , one can extend coefficients to obtain vector spaces over the field of -adic numbers. Then the group acts on .
For every prime the group contains Frobenius elements , well-defined up to ambiguities that will disappear from our considerations. For any power of a prime , and any , one has the trace of the operator acting on . These -adic numbers are in fact rational and independent of . We emphasize the independence of by denoting them . The connection with point counts is the Lefschetz trace formula: . The left side for fixed and varying determines the summands on the right side in principle because the complex eigenvalues of on weight cohomology have absolute value .
Much of this transfers formally to the motivic setting. Thus for a motive and a prime , there is an action of on the corresponding -adic vector space . This action has image in . Indeed the Tate conjecture predicts that the -Zariski closure of the image of is all of .
One technical problem with André’s category is that the projectors used to define motives are not known to come from algebraic cycles. As a consequence, for a general the above compatibility of Frobenius traces is not known. However this problem does not arise for hypergeometric motives, because they are essentially the entire middle cohomology of varieties. Accordingly one has well-defined rational numbers . There are similar technical problems at the primes , but they do not affect our computations and we will ignore them.
Wild, tame, and good primes.
Returning now to very concrete considerations, we sort primes for a parameter as follows. A prime is wild if it divides a . For , a prime is tame if it is not wild but it divides either the numerator of , the denominator of , or the numerator of ; these last three conditions say that is -adically close to the special points , , and respectively. For , no primes are tame. We say that a prime is bad if it is either wild or tame, and all other primes are good.
Split powers of a good prime .
A power of a good prime is split for if , where is the least common multiple of the . One then has a collection of Jacobi sums indexed by characters of :
Here is any nonzero additive character, is any generator of the group of multiplicative characters, underlies as in Section 2, and is the standard Gauss sum. The desired quantity is then given by a sum due to Katz [Katz-ESDE, p. 258]. Renormalizing to fit our conventions, it is
(10.1) |
Here is the vertical coordinate of a lowest point on the zigzag diagram of , e.g. in Figure 5.1.
General powers of a good prime .
The Gross-Koblitz formula lets one replace the above Gauss sums by values of the -adic gamma function. This is both a computational improvement and extends the formula to all powers of any good prime. With this method, the desired integers are first approximated -adically. Errors are under control and exact values are determined from sufficiently good approximations. See [BCM] for a closely related approach to the essential numbers and references to earlier contributions.
11 Frobenius polynomials
Frobenius polynomials are a concise way of packaging the point counts of the preceding section. They play the leading role in the formula for -functions of the next section. After saying what they are, this section explains several reasons why they are useful, even before one gets to -functions.
Capturing point counts.
Consider the numbers for a fixed motive of rank , a fixed good prime , and varying . They can be captured in a single degree polynomial . The relation, which comes from summing the geometric series belonging to each of the eigenvalues, is
(11.1) |
Write
Then the for determine . Thus the for determine . But, even better, Poincaré duality on a source variety ultimately implies that one has for a sign . For HGMs, this sign is known and in fact always when is odd. So can be computed using only for .
Relation with Hodge vectors.
Indexing by weight , consider as examples the rank six family parameters
(11.2) |
The first two are the families from Section 7, with Hodge vectors respectively and ; the last one has Hodge vector . Specializing at a randomly chosen common point gives motives .
After the required initialization of a variable by _<x>:=PolynomialRing(Integers())
,
and after inputting Q
as in (2.4),
Magma quickly gives some Frobenius polynomials
via e.g. EulerFactor(Q0,3/2,5)
:
These displays illustrate a basic motivic principle: as weight increases, motives of a given rank become more complicated. A more refined principle involves Hodge numbers and can be expressed by forming a weakly increasing vector , where an entry appears times. Then the Newton-over-Hodge inequality is For , …, , these lower bounds from the Hodge vector controlling are . For the bounds would be smaller, leaving more possibilities for Frobenius polynomials. In this sense, spread out Hodge vectors correspond to more complicated motives.
Congruences.
Reduced to , the numbers for depend only on the mod Galois representation belonging to . In our examples, suppose one kills in (11.2) by replacing all by . Then and both become . This agreement implies that is independent of . This independence can be seen for the primes and in the displayed Frobenius polynomials. The analogous congruences hold for any , when one changes our Tate twist convention to make the weight of the number of integers among the and , minus one. This web of congruences, like the web corresponding to splicing considered in Section 7, makes it clear that HGMs constitute a natural collection of motives.
Finite Galois groups.
Frobenius polynomials render Galois-theoretic aspects of the situation very concrete. As a warm-up, consider as a representative of the relatively familiar case of ordinary Galois theory. Here the -adic representations all come from a single representation . Let be the partition of obtained by taking the degrees of the irreducible factors of from (7.1). Then the twenty-five possibilities for the pair correspond to the twenty-five conjugacy classes in the finite group . If one can collect enough classes, then one can conclude that the image is all of . In our example , the above primes and give and respectively. In ATLAS notation, these are the classes and . They do not quite suffice to prove . But the prime gives the class and since no maximal subgroup contains elements from , , and , indeed .
The Chebotarev density theorem says that each pair appears proportionally to the number of elements in its conjugacy class. For example, the classes , and occur with frequency , , and respectively.
Infinite Galois groups.
The cases and are beyond classical Galois theory as the motivic Galois groups have positive dimension. But the situation remains quite similar. Consider for example odd weight motives of rank so that is in the conformal symplectic group . The Weyl group of is the hyperoctahedral group of signed permutation matrices, with order . A separable , being conformally palindromic, has Galois group within . If it has Galois group all of then necessarily contains a certain twisted maximal torus. Suppose a second prime satisfies the same condition and moreover the joint Galois group of is all of . Then contains two maximal tori which are sufficiently different to force , by the classification of subgroups containing a maximal torus.
To analyze a given motive, the necessary computations can be
done using Magma’s GaloisGroup
command. The order of the Galois group of
is , , , , for , , , ,
, and the pair satisfies
the criterion. For , all have
Galois group except . Excluding
13, all pairs satisfy
the criterion. In general, it becomes easier to establish
genericity as the weight increases, a reflection of the
growth in complexity discussed above.
Applying this two-prime technique to the special and semi HGMs of Section 9 suggests that almost always their motivic Galois groups are as big as possible. In particular, the exotic Hodge vectors with interior zeros arising there indeed come from irreducible motives. Details in the case (9.1) are given in [Rob].
The Chebotarev density theorem extends to the full motivic setting if all the -functions described below have their expected analytic properties. Readers wishing to see a glimpse of this theory can compute hundreds of for for or . By all appearances, the data matches the Sato-Tate measure , meaning the pushforward of Haar measure on the compact group to via the defining character. One would have to compute thousands of before one could confidently distinguish this measure from the Gaussian measure of mean and standard deviation .
12 -functions
We now finally define -functions and illustrate how everything works by some numeric computations.
Local invariants.
Let be a motive of rank and weight , having bad reduction within a finite set of primes. We have discussed two types of local invariants associated to . Corresponding to the place of is the Hodge vector with total , and also a signature . Corresponding to a prime is the degree Frobenius polynomial . For primes , there is also a Frobenius polynomial , now of degree , and moreover a conductor exponent , both to be discussed shortly. The conductor of , which can be viewed as quantifying the severity of its bad reduction, is the integer .
Formal products.
The local invariants can be combined into a holomorphic function in the right half-plane , called the completed -function of :
(12.1) |
The product over primes alone is the -function , while the remaining factors give the completion. The infinity factor is given by an explicit formula:
(12.2) |
Here and . The factors involving only appear when is even; in the common case that , they can be replaced by , by the duplication formula.
Both the -function and the completing factor are multiplicative in so that . Another simple aspect of the formalism is that Tate twists correspond to shifts: .
Expected analytic properties.
The -function of the unital motive is just the Riemann zeta function , and the completing factor is . Riemann established that has a meromorphic continuation to the whole -plane, with poles only at and ; moreover he proved that . The product is expected to have similar analytic properties. First, for irreducible and not of the form , there should be an analytic continuation to the entire -plane, bounded in vertical strips. Second, always
(12.3) |
for some sign . For comparison with Section 14, note that most everything said in the last three sections generalizes to motives in , with Frobenius polynomials being in . However (12.3) takes the more complicated form , with the complex conjugate motive and only on the unit circle.
Determining invariants at bad primes.
One approach to the conductor exponents and Frobenius polynomials associated to bad primes is to compute them directly by studying the bad reduction of an underlying variety. For an HGM , Magma takes this approach for primes which are tame for , as sketched in Section 13.
A very different approach uses the fact that the list of possible for a given prime is finite, and the product (12.1) has the conjectured analytic properties for at most one member of the product list. The current state of HGMs for the wild primes of mixes the two approaches: we first greatly reduce the length of the lists by using proved and conjectured general facts. Then we search within the much smaller product list for the right quantities.
Our view is that numerical computations such as those that follow in this section and Section 15 admit only one plausible interpretation: the bad factors have been properly identified and the analytic properties indeed hold. However rigorous confirmation does not seem to be in sight at the moment, despite the progress described in Section 14.
A rank four example.
For of degree and , Watkins numerically identified all the bad quantities, so that the corresponding -functions are immediately accessible on Magma. For example, take the family parameter to be , implemented as always by modifying (2.4). At the specialization point , the Hodge vector is . The corresponding -function, set up so that calculations are done with 10 digits of precision, is
L := LSeries(Q,1:Precision:=10);
The bad information stored in Magma is revealed by
EulerFactor(L,
)
and Conductor(L)
to be , ,
and . The sign
is calculated numerically, with Sign(L)
returning . So the order of vanishing
of at the central point should be odd.
This order is apparently three since
Evaluate(L,2:Derivative:=1);
returns zero to ten decimal places, but the same command with replaced by returns 51.72756346.
A rank six example.
More typically, Magma does not know
and for wild primes and one needs to input this information.
As an example, take
with Hodge
vector .
The only prime bad for the data is .
A good first guess is that is just the
constant . A short search over
some possible is implemented after redefining Q
by
[CFENew(LSeries(Q,1:Precision:=10),
BadPrimes:=[<3,c,1>]): c in [6..10]];
The returned number for is ,
while the numbers for the other are all at least .
This information strongly suggests that indeed
and . After setting up L
with
[<3,9,1>]
, analytic calculations can
be done as before. For example, here
the order of central vanishing is apparently 2.
In the miraculous command CFENew
,
CFE
stands for the Magma command
CheckFunctionalEquation
, implemented
by Tim Dokchitser using his [Dok]; New
reflects subsequent improvements by Watkins.
13 Bad primes
Fix a hypergeometric motive and a prime . We now sketch how Magma computes the local invariants when is tame for , and describe some conjectural basic features for the case when is wild for .
Tame primes.
When is tame for , the conductor exponent is the codimension of the invariants of a power of a Levelt matrix from Section 2. When , the simple shape of gives a completely explicit formula: except in the orthogonal case with even, where . When ,
(13.1) |
Here , if is negative, and if is positive. So there is separate periodic behavior for and , as illustrated by the top part of Figure 13.1. The example of this table comes from the case of (5.1), so the conductor there is very simply computed as the discriminant of the octic algebra .
Because ramification is at worst tame, the degree of is . When is positive, is computed by slightly modifying the formulas for point counts sketched in Section 10. In the other cases, comes from Jacobi motives as mentioned around (4.1), extracted from how the family degenerates at the relevant cusp .
Wild primes.
To simplify the overview, we just exclude the case where . Write specialization points as with . The bottom part of Figure 13.1 shows right away that the situation is complicated.

A function is graphed in both parts of Figure 13.1 and its general definition goes as follows. For a positive integer, write
Let
Define and transition points and . Then
In the tame case, is just the constant function . In general, there are plateaus corresponding to the cusps and , and then a ramp of length between them.
We conjecture that
(13.2) |
with equality if and are relatively prime. The second statement is proved in [LNV] in the general trinomial setting of (5.1). All of (13.2) has been computationally verified in many instances. As one passes from one family to another via mod congruences as in Section 11, wild ramification at does not change. This fact and other theoretical stabilities give us confidence in (13.2). To make Magma more fully automatic, a key step would be to define a more complicated function with , and equality under broad circumstances.
At present, we understand a factor of the Frobenius polynomial as follows. For , comes from modifying the tame formulas; in particular its degree is given by replacing by in (13.1). If , corresponding to being at the bottom of the ramp, we use an erasing principle explained to us by Katz. Here one simply ignores all and that have denominator divisible by . Let and be respectively the number of ’s and ’s remaining. Then is a multiple of , so that the formulas described in Section 10 still make sense, as the choice of an auxiliary additive character on again does not matter. The resulting has degree . We conjecture that the complementary factor is whenever and are relatively prime. In practice, when it is usually also, but not always.
14 Automorphy
One of the most exciting aspects of the theory of motives is its conjectured extremely tight connection to automorphic representations of adelic groups through the Langlands program.
Background.
Let be the adele ring of ; it is a restricted product of all the completions , including . A cuspidal automorphic representation of has an -function known to have an analytic continuation and functional equation. The main conjecture is that, after incorporating Tate twists to make normalizations match, the set of -functions coming from irreducible rank motives in is exactly the subset of automorphic -functions for which the infinity factor has the form (12.2).
The case .
For a motive with nonvanishing Hodge numbers and conductor , one can switch to classical language. The desired automorphic representation is entirely given by a power series in as in (1.6), but now this newform on has weight .
To exhibit some matches between motivic and automorphic -functions, consider the four reflexive parameters yielding motives with Hodge vector :
Magma computes automatically with these reducible motives, reporting their conductors to be , , , and . However these computations do not see the decompositions analogous to (9.1), where now and respectively have Hodge vectors and . In the Frobenius polynomial
the belonging to can be distinguished from the belonging to whenever the latter is not a multiple of . The reader might enjoy searching in the LMFDB’s complete lists [lmfdb] of modular forms to see that the and for and let one identify the relevant forms and in particular determine the above-displayed factorizations . Part of the further information given by the LMFDB is that two of the forms are expressible using the Dedekind eta function , via .
Higher rank.
For a given motive with larger rank , one can usually replace by the adelic points of a smaller group determined by the motivic Galois group of . In favorable cases, the representation sought again corresponds to a holomorphic form. For rank three orthogonal motives, classical modular forms are again relevant, but a symmetric square is now involved. In rank four, Hilbert modular forms are needed for orthogonal motives and Siegel modular forms are needed for symplectic motives. Numerical and sometimes proved matches have been found in these three settings. For example, [DPVZ] treats some interesting rank four orthogonal cases.
Generally speaking, the Hodge numbers of central concern earlier in this survey continue to play a large role. In particular, motives for which all are or have theoretical advantages, and their motivic -functions at least have a meromorphic continuation with the right functional equation [PT].
15 Numerical computations
We promised in the introduction that we would equip the reader to numerically explore a large collection of motivic -functions. We conclude this survey by giving sample computations in the context of two important topics, always assuming that the expected analytic continuation and functional equation indeed hold. In both topics, we let be the center of the functional equation. The conductors in our examples are small for their Hodge vectors , allowing us to keep runtimes short and/or work to high precision.
Special values.
If is a motive in then the numbers for integers are mostly forced to be , because of poles in the infinity factor (12.2) and the functional equation. However, when is nonzero it is expected to be arithmetically significant [Del-PL]. The arithmetic interpretation involves a determinant of periods like (2). To see the significance without entering into periods, one can look at the ratio , for a positive quadratic discriminant. Then the periods cancel out so that should be rational.
For a sample computation, take and use
(2.4) and
L:=LSeries(Q,1024)
to define its -function, as usual.
While is wild for the family, it is unramified in
because because the exponent is at the bottom
of the ramp of Section 13. The erasing procedure
from the end of that section applies, yielding
Since
is squarefree, it is the conductor, by the recipe
before (13.1). Magma gets all the bad factors right automatically. As a confirmation,
CFENew(L)
quickly returns to the default
digits.
Evaluate(L,2)
gives
.
Twisting by a with makes the conductor go up by a factor
of and precision needs to be reduced.
Evaluate(LSeries(Q,1024:
QuadraticTwist:=5,Precision:=10),2);
takes six minutes to give its answer of . This ratio and then two others are apparently
The two -functions appearing in are completely different analytically, and so the apparent fact that quotients are rational is very remarkable.
Readers wanting to work out their own examples might want to begin with having odd weight. Then if , one has conjecturally rational quotients for . The lateral argument fits into the theory only in the rare case that the most central entries of the Hodge vector are . In the even weight case, one needs to have to make fit into the theory, as in our example.
Critical zeros.
For a weight motive , all the zeros of the completed -function lie in the critical strip . The Riemann hypothesis for then predicts that all the zeros lie on the critical line . We now show by examples that numerical identification of low-lying zeros is possible in modestly high rank.
For the examples, take with and . So is orthogonal with Hodge vector while is symplectic with Hodge vector .
The only bad prime in each case is . A search says that and . For , so erasing applies, yielding as a factor of . A short search says that this factor is all of and .
In general, the Hardy Z-function of a motive is
with . It is a real-valued function of the real variable , even or odd depending on whether the sign is or .

Figure 15.1 was computed via many calls to Evaluate
at points of the
form . The signs in the two cases are and , and the orders of
central vanishing are the minimum possible, and .
On both plots, all local maxima are above the axis and all local minima are
beneath the axis. Zeros off the critical line would likely cause a disruption
of this pattern; thus the plots not only identify zeros on the critical line, but suggest
a lack of zeros off the critical line.