Dubrovin conjecture and the second structure connection
Abstract.
We give a reformulation of the Dubrovin conjecture about the semisimplicity of quantum cohomology in terms of the so-called second structure connection of quantum cohomology. The key ingredient in our work is the notion of a twisted reflection vector which allows us to give an elegant description of the monodromy data of the quantum connection in terms of the monodromy data of its Laplace transform.
1. Introduction
1.1. Semi-simple quantum cohomology
Let be a smooth projective variety of complex dimension . The Gromov–Witten (GW) invariants of are defined via the intersection theory on the moduli space of stable maps (see Section 3.1 for more details). The structure of the GW invariants is best understood in genus . Namely, the entire information is contained in a certain deformation of the classical cup product, known as the quantum cup product. Let us fix a homogeneous basis () of . Then the quantum cup product is defined by
where the second sum is over all effective curve classes , are the so-called Novikov variables, and . The Novikov variables correspond to a choice of an ample basis of and . Let us assume that and . Then the structure constants of the quantum cup product belong to the following ring of formal power series
where the fact that there is a dependence on is a consequence of the so-called divisor equation (see [25]). The quantum cup product turns into a commutative associative algebra known as the big quantum cohomology of . We say that the quantum cohomology is semi-simple if is a semi-simple algebra, or equivalently the operators of quantum multiplication by are not nilpotent. In the limit , the quantum product becomes the classical cup product and since is always nilpotent (for ), we see that semi-simplicity is an indication that the manifold has sufficiently many rational curves. From that point of view it is very interesting to classify manifolds with semi-simple quantum cohomology. It is an observation of Alexey Bondal that semi-simplicity can be characterized using the language of derived categories. In his ICM talk in 1998, following Bondal’s ideas, Dubrovin was able to formulate a precise conjecture which is now known as the Dubrovin conjecture (see [9] and Conjecture 3.3.1). The goal of this paper is to answer a question which was raised by the second author in his joint work [26]. Namely, there is a conjectural description (see [26], Conjecture 1.6) of the so-called reflection vectors in quantum cohomology in terms of exceptional collections in the derived category. More precisely, Milanov–Xia were able to construct reflection vectors in the quantum cohomology of the blowup of a manifold relying only on certain vanishing results for GW invariants. The resulting formulas were very similar to the formulas for the central connection matrix conjectured by Galkin–Golyshev–Iritani in [13] and later on by Cotti–Dubrovin–Guzzetti [6]. The question is whether one can reformulate Dubrovin’s conjecture in terms of reflection vectors. As expected, the answer is yes and in this paper we would like to work out the precise relation between the reflection vectors and the monodromy data which enters the Dubrovin conjecture. Our main message is that by constructing a basis of reflection vectors in quantum cohomology one can obtain a proof of the -Conjecture II of Galkin–Golyshev–Iritani (see [13]) or equivalently the refined Dubrovin conjecture of Cotti–Dubrovin–Guzzetti (see [6], Conjecture 5.2). In the rest of this introduction we would like to formulate our results.
1.2. Quantum differential equations
From now on we are going to assume that quantum cohomology is semi-simple and convergent. The latter means that there exists a domain , such that, the formal power series representing the structure constants are convergent. Let us introduce the following two linear operators:
and
where for and is the classical cup product. The quantum differential equations are by definition the differential equations of the following system of ODEs:
(1) | |||||
(2) |
where , is the operator of quantum multiplication by , and
is the Euler vector field. The differential equations (1)–(2) can be viewed also as the equations defining the horizontal sections of a connection on the vector bundle which is sometimes called quantum connection or Dubrovin connection.
The differential equation with respect to has two singularities: regular at and irregular at . Near there is a geometric way to construct a fundamental solution. Namely, let us define where are linear operators defined by
The singularity at has an interesting Stokes phenomenon. Namely, suppose that are the canonical coordinates defined in a neighborhood of some semi-simple point . By definition, the quantum product and the Poincaré pairing become diagonal:
where are some holomorphic functions. Let us define the linear map
(3) |
We may think of as a matrix with entries . Let be the diagonal matrix. There exists a unique formal asymptotic solution to (1)–(2) of the form where , are matrices. The matrices are determined uniquely by plugging in the ansatz into (1)–(2) and comparing the coefficients in front of the powers of . This gives us a recursion relation for the coefficients which turns out to have a unique solution. Moreover, the solution automatically satisfies where t denotes the standard transposition of matrices. Suppose that the semi-simple point is generic and that the coordinate neighbourhood is so small that for . The rays in the -plane of the form where and , are called the Stokes rays. Let be a line not parallel to any of the Stokes rays. Let us fix an orientation of by choosing a unit vector , . Following Dubrovin (see [10]) we will say that is an admissible oriented line. The line splits the -plane into right and left half-planes. There exists unique solutions and to (1)–(2) holomorphic respectively for and which are asymptotic to as . These solutions extend analytically in along the positive part of . In particular, we obtain 3 holomorphic solutions to (1)–(2) along the positive half of which must be related as follows:
where and are some constant matrices called respectively, the Stokes matrix and the central connection matrix. The refined Dubrovin conjecture (see [6], Conjecture 5.2) consists of 3 parts. First, the big quantum cohomology of is semi-simple if and only if the bounded derived category has a full exceptional collection. The second part says that every admissible line determines a full exceptional collection which determines uniquely the Stokes matrix and the central connection matrix . Finally, the 3rd part of the conjecture gives very precise formulas for both and , that is, the -entry of is
and the -th column of is
where is the remainder of the division of by , is the so-called gamma class of , and is the Chern character of . Here the products are over the Chern roots and of respectively the holomorphic tangent bundle and the complex vector bundle . Just like in the case of , we think of as a linear map . We refer to Section 3.4 for more details.
The refined version of the conjecture still requires that the manifold is Fano. However, as it was pointed out by Arend Bayer in [3] (see also the recent work by Hiroshi Iritani [22]), by using the blowup operation we can construct many examples of non-Fano manifolds for which the first part of the Dubrovin conjecture holds. Moreover, the recent work by Milanov–Xia (see [26]) gives an indication that the blowup operation preserves the remaining two parts of the Dubrovin conjecture. Therefore, it is quite plausible that the Fano condition is redundant.
1.3. Reflection vectors
Suppose that the big quantum cohomology is semi-simple and convergent. The solutions to the quantum differential equations can be represented by complex oscillatory integrals of the following form:
where is a complex number and the semi-infinite integration cycle is chosen in such a way that the integral is convergent. It is easy to check that the above integral solves the quantum differential equations (1)–(2) iff the integrand satisfies the following system of ODEs:
(4) | ||||
(5) |
This is a system of differential equations for the horizontal sections of a connection on the trivial bundle
where
The hypersurface in is called the discriminant. The connection is known as the second structure connection. In the case when or , the connection was used by Dubrovin to define the monodromy group of a Frobenius manifold (see [8]). However, it became clear shortly afterwards that it is important to study the entire family, that is, allow to be any complex number (see [24] and [11]).
The space of solutions to (4)–(5) is quite interesting. In the examples of mirror symmetry the second structure connection of quantum cohomology can be identified with a Gauss–Manin connection. Therefore, the solutions to (4)–(5) should be thought as period integrals. In particular, by using we can introduce many of the ingreidents of Picard–Lefschetz theory. This was done by Dubrovin (see [11], Section 4). He called the solutions to (4)–(5) twisted periods because their properties are very similar to the period integrals in Givental’s twisted Picard–Lefschetz theory [15]. Motivated by the work of Givental in [17], the second author introduced in [27] the following fundamental solution to (4)–(5):
(6) |
where
(7) |
Note that both and take values in . The second structure connection has a Fuchsian singularity at infinity, therefore the series is convergent and it defines a multi-valued analytic function in the complement to the discriminant. There are many ways to choose a fundamental solution but what makes the above choice special is the specific choice of building blocks, that is, the calibrated periods (7) while the standard approach would be monomials in . The existence of such decomposition follows from Givental’s formalism of quantized symplectic transformations and their actions on vertex operators (see [17], Section 5, especially Theorem 2). Although Dubrovin already knew that one can import concepts from singularity theory to quantum cohomology, somehow the above choice of a fundamental solution makes the parallel with singularity theory much more visible (at least to the authors).
Let us choose a base point , such that, for . Then for for all sufficiently close to . Let be a positive real number, such that, for all . We define the -twisted period vectors where and the value depends on the choice of a reference path avoiding the discriminant from to . Note that at the only ambiguity is in the choice of the value for the calibrated periods, that is, we need to specify a branch of when is close to . Since is a positive real number we simply take the principal branch of the logarithm.
Let us introduce the following pairings
(8) |
Using the differential equations of it is easy to check that is independent of and . It turns out that there is an explicit formula for in terms of the Hodge grading operator and the nilpotent operator . Let us recall the so-called Euler pairing
(9) |
As a byproduct of the proof of Theorem 2.6.1 we will get the following simple formula:
where . The above formula shows that is the analogue of the -bilinear intersection form in twisted Picard–Lefschetz theory (see [15], Section 3). Furthermore, let us fix a reference path from to a point sufficiently close to a generic point on the discriminant. The local equation of the discriminant near has the form where is an eigenvalue of . It turns out that the set of all , such that, is analytic at is a codimension 1 subspace of . Suppose that , then there is a 1-dimensional subspace of vectors , such that, is analytic at and the value at belongs to where is the map (3). Therefore, for the given reference path and an arbitrary choice of there is a uniquely defined vector , such that,
(10) |
where the coefficient in front of is such that . Note that the choice of a reference path and a branch of determines for all , that is, we have a map
Using that we get that this map is periodic: It will follow from our results that is a trigonometric polynomial, that is, . If we change the value of the logarithmic branch , then . Therefore, for a fixed reference path the value of is fixed up to a factor in the spiral . We will say that is a twisted reflection vector corresponding to the given reference path. We usually supress the dependence on the logarithmic branch if the choice is irrelevant or it is clear from the context. The following formula for the local monodromy of justifies our terminology:
that is, the local monodromy is a complex reflection whose fixed points locus is the hyperplane orthogonal to . We refer to Section 2.3 for more details and for more general settings, i.e., we can introduce twisted reflection vectors for any semi-simple Frobenius manifold.
1.4. Monodromy data and reflection vectors
We continue to work in the settings from the previous two sections. Let be an admissible oriented line (see Section 1.2) with orientation . By definition is not parallel to any of the differences for . We will refer to as an admissible direction. Our choice of a reference point is such that the real line with its standard orientation is an admissible oriented line. The corresponding admissible direction is . Any other admissible direction will be equipped with a reference path to or equivalently, we fix an analytic branch of in a neighbourhood of . Finally, we consider only sufficiently close to , such that, is an admissible direction for , that is, is not parallel to for .
Let us construct a system of reference paths corresponding to . Each starts at , approaches the circle in the direction of , after hitting the circle at some point the path continues clockwise along the circle arc from to , and finally by continuously deforming the direction to the path connects and – see Figure 1. Note that if is sufficiently close to , then for some positive real number and we have a natural choice of a logarithmic branch: . Therefore, as it was explained in Section 1.3, we may choose a twisted reflection vector . In other words, each admissible direction determines a set of twisted reflection vectors . Furthermore, the admissible direction determines the following order of the eigenvalues of : we say that if is on the RHS of the line through parallel to where RHS means that we have to stand at and look in the direction . For example, for the standard admissible direction , would mean that . We will refer to the order as the lexicographical order determined by . Let us assume that the enumeration of the eigenvalues is according to the lexicographical order, that is, iff . Our main result can be stated as follows.
Theorem 1.4.1.
Let be an admissible direction and assume that the eigenvalues of the operator are enumerated according to the lexicographical order corresponding to . Then the following statements hold.
-
a)
The reflection vectors () are independent of and the Gram matrix of the Euler pairing (9) is upper-triangular
with ’s on the diagonal: .
- b)
-
c)
The inverse Stokes matrix coincides with the Gram matrix of the Euler pairing (9) in the basis ().
-
d)
The -entry of the central connection matrix is related to the components of the reflection vectors by the following formula:
In fact our result is more general. The above theorem can be formulated in the settings of semi-simple Frobenius manifolds. Under an additional technical assumption, i.e., we assume that the Frobenius manifold has a calibration for which the grading operator is a Hodge grading operator (see Definition 2.1.2), we prove that the conclusions of the above theorem remain true (see Theorem 2.6.1).
Using Theorem 1.4.1 we can answer the question raised in [26]. Following the analogy with singularity theory (see [1, 12]), we introduce the concept of a distinguished system of reference paths (see Definition 3.5.1). Let us recall the Iritani’s integral structure map (see [21]) defined by
where are the Novikov variables corresponding to an ample basis of . We have the following relation:
(11) |
which justifies why we refer to (9) as the Euler pairing.
Theorem 1.4.2.
Let be the reflection vectors corresponding to a distinguished system of reference paths. Parts (2) and (3) of the refined Dubrovin conjecture, i.e., Conjecture 3.4.1, hold if and only if there exists a full exceptional collection , such that, for all .
Theorem 1.4.2 is proved in Section 3.5 (see Theorem 3.5.1). The definition of a reflection vector is the same as the definition of a twisted reflection vector except that we require . Since the fundamental group of is generated by simple loops corresponding to the reference paths where is an admissible direction, we have the following interesting corollary.
Corollary 1.4.1.
The set of all twisted reflection vectors is a subset of
where are the reflection vectors corresponding to a distinguished system of reference paths. In addition, if the manifold satisfies the refined Dubrovin conjecture, then the set of twisted reflection vectors is a subset in the -lattice .
For the proof, thanks to the braid group action on the set of distinguished systems of reference paths (see Section 3.5), we may assume that are the reflection vectors corresponding to the reference paths for some admissible direction . Note that every twisted reflection vector is obtained from some by a sequence of local monodromy transformations () where is the monodromy transformation corresponding to the simple loop associated with . According to Theorem 1.4.1, part a), the matrix of (resp. ) in the basis is upper-triangular and the only entry depending on is in position , that is, it is equal to (resp. ). In addition, if the refined Dubrovin conjecture holds, then since , we get that the entries of belong to .
Let us point out that the full exceptional collection in Theorem 1.4.2 is not the same as the full exceptional collection in the refined Dubrovin conjecture. The reason is that the objects correspond to oscillatory integrals in which the integration paths are rays with direction while correspond to reference paths going in the opposite direction . Changing the admissible direction from to means that one has to perform a certain sequence of mutations in order to get from one exceptional collection to the other one. It turns out that the sequence of mutations that we need is well known in the theory of derived categories, i.e., this is the same sequence used to define the left Koszul dual of an exceptional collection. The precise statement is that up to a shift by the exceptional collection is the left Koszul dual to . The moral is that although there is some freedom in defining a Stokes matrix and a central connection matrix of the quantum connection, the statement of the refined Dubrovin conjecture is independent of the choices that we make. The different full exceptional collection that one might get due to the discrepancy of the definitions are related by mutations in the derived category. Let us point out that even Dubrovin himself did not use the definitions consistently – the central connection matrix in [10, 11] is the inverse of the central connection matrix in [6].
Finally, let us comment on the proofs. The relation between the quantum connection and its Laplace transform was studied by many people. In particular, the relation between Stokes multipliers and the monodromy data of the Laplace transform of the quantum connection is well known thanks to the work of Balsar–Jurkat–Lutz [2]. There is also a recent paper by Guzzetti (see [18]) who was able to remove some technical conditions from the main result in [2]. Although, we do not directly use any results from [2], the ideas for all proofs come from there, except for the formula for the connection matrix (see Theorem 2.6.1, part a) whose proof follows the ideas of Dubrovin (see [11], Theorem 4.19). Our results should not be very surprising to the experts. Especially, the work of Galkin–Golyshev–Iritani [13] and Dubrovin [11] contain almost all ideas and results necessary to prove Theorems 1.4.1 and 1.4.2. In some sense we could have written much shorter text. Nevertheless, in order to avoid gaps in the arguments due to misquoting results, we decided to have a self contained text independent of the results in [2, 11, 13].
1.5. Acknowledgements
We are thankful to Alexey Bondal for many useful discussions on the Dubrovin conjecture and especially for pointing out to us the notion of left and right Koszul dual of an exceptional collection. The first author also thanks Jin Chen and Mauricio Romo for many interesting discussions on the -conjectures and Dubrovin conjecture. This work is supported by the World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan and by JSPS Kakenhi Grant Number JP22K03265. The first author is also supported by the President’s International Fellowship Initiative of the Chinese Academy of Sciences.
2. Twisted periods of a Frobenius manifold
As promised in the introduction, we will formulate and prove Theorem 1.4.1 more abstractly in the settings of a semi-simple Frobenius manifold. Compared to Dubrovin’s theory of twisted periods (see [11], Section 4), we introduce a fundamental solution to the second structure connection which allows us to see better the analogy with singularity theory.
2.1. Frobenius manifolds
Let be a complex manifold and be the sheaf of holomorphic vector fields on . Suppose that is equipped with the following structures:
-
(F1)
A non-degenerate symmetric bilinear pairing
-
(F2)
A Frobenius multiplication: commutative associative multiplication
such that .
-
(F3)
A unit vector field: global vector field , such that,
where is the Levi–Civita connection of the pairing .
-
(F4)
An Euler vector field: global vector field , such that, there exists a constant , called conformal dimension, and
for all .
Note that the complex manifold has a structure of a holomorphic vector bundle with base : the fiber over is which has a natural structure of a vector space. Given the data (F1)-(F4), we define the so called Dubrovin connection on the vector bundle
where is the standard coordinate on , where is an endomorphism of defined by the Frobenius multiplication by the vector field , and where is an -modules morphism defined by
Definition 2.1.1.
The data , satisfying the properties , is said to be a Frobenius structure of conformal dimension if the corresponding Dubrovin connection is flat, that is, if are holomorphic local coordinates on , then the set of differential operators , pairwise commute. ∎
Near the Dubrovin connection has a fundamental solution of the following form:
(12) |
where is a diagonalizable operator, is a nilpotent operator, and the operator-valued series , satisfies the symplectic condition , where T is transposition with respect to the Frobenius pairing. It can be proved that coincides with the semi-simple part of the grading operator in the Jordan–Chevalley decomposition , where the operators and are uniquely determined by the following 3 conditions:
-
(i)
Commutativity: .
-
(ii)
The operator is diagonalizable.
-
(iii)
The operator is nilpotent.
Moreover, the operator where for finitely many and . For more details we refer to [28], Section 1.3.1. Following Givental [16] we will refer to the pair as calibration of . Sometimes we will drop from the notation and say that is the calibration.
Remark 2.1.1.
The pair is not uniquely determined from the Frobenius structure and in general there is no canonical choice. More precisely, one can prove that there exists a unipotent Lie group acting faithfully and transitively on the set of such pairs – see [28], Section 1.3.1. ∎
Definition 2.1.2.
Let be a calibration of and be the corresponding nilpotent operator. The grading operator is said to be a Hodge grading operator for the calibration if
-
(i)
The operator is diagonalizable.
-
(ii)
The following commutation relation holds: ∎
Note that if is a Hodge grading operator, then and for all . The fundamental solution takes the form . From now on we will consider only Frobenius manifolds with a fixed calibration, such that, is a Hodge grading operator. The problem that we will be interested in is local, so we will further assume that has a global flat coordinate system .
Let us fix a base point . Put , then is a basis of the reference tangent space . The flat vector fields () provide a trivialization of the tangent bundle . This allows us to identify the Frobenius multiplication with a family of associative commutative multiplications depending analytically on . The operator defined above preserves the subspace of flat vector fields. It induces a linear operator on , known to be skew symmetric with respect to the Frobenius pairing .
There are two flat connections that one can associate with the Frobenius structure. The first one is the Dubrovin connection – defined above. The Dubrovin connection in flat coordinates takes the following form:
where is the standard coordinate on and for we denote by the linear operator of Frobenius multiplication by .
We will be interested also in the second structure connection
(13) | ||||
(14) |
where is a complex parameter. This is a connection on the trivial bundle
where
The hypersurface in is called the discriminant.
2.2. Twisted periods
Let be any complex number.
Definition 2.2.1.
By a -twisted period of the Frobenius manifold we mean a sequence satisfying the following two properties:
-
(i)
Flatness: is a horizontal section for .
-
(ii)
Translation invariance: .∎
The set of all -twisted periods has a natural structure of a vector space. Note that if is sufficiently large, then the -twisted period sequence is uniquely determined from only. Indeed, by translation invariance, we have for all . Using (14)
We get that as long as is invertible we can express in terms of . Let us choose so large that is smaller than the real parts of all eigenvalues of . Then, it is clear that all can be expressed in terms of .
Suppose now that is a calibration. We will construct an isomorphism between and the space of all -twisted periods. Let us fix a reference point such that is a sufficiently large positive real number. It is easy to check that the following function is a solution to the second structure connection
(15) |
where
(16) |
Note that both and take values in . The second structure connection has a Fuchsian singularity at infinity, therefore the series is convergent for all sufficiently close to . Using the differential equations (13)–(14) we extend to a multi-valued analytic function on taking values in . We define the following multi-valued functions taking values in :
(17) |
Clearly, for each fixed , the sequence () is a period vector in the sence of Definition 2.2.1. Moreover, if is sufficiently negative, then is an invertible operator. Therefore, all -twisted period vectors of have the form for some . Note that the analytic continuation along a closed loop around the discriminant leaves the space of -twisted periods invariant. Therefore, for each we have a monodromy representation
(18) |
2.3. Local monodromy
Recall that a point is said to be semi-simple if there are local coordinates near , called canonical coordinates, such that, the multiplication and the Frobenius pairing take the following form:
where () are holomorphic functions that do not vanish at . The Frobenius manifold is said to be semi-simple if it has at least one semi-simple point. The subset of points that are not semi-simple is called the caustic. If is semi-simple, then the caustic is either the empty set or an analytic hypersurface.
From now on we will assume that is a semi-simple Frobenius manifold. Let us choose the base point , such that, for . Then for for all sufficiently close to . We would like to describe the space of horizontal sections of locally in a neighbourhood of . There is a distinguished solution which can be constructed similarly to (15) but by using Givental’s R-matrix instead of the calibration . Let us recall the definition of Givental’s R-matrix (see [16]). Let and let be the matrix whose entry is . In other words, is the matrix of the linear isomorphism , with respect to the standard basis of and the flat basis of . According to Givental, there exists a unique operator-valued formal series , such that, the Dubrovin connection has a formal asymptotic solution at of the form . Moreover, the matrix series satisfies the symplectic condition where t is the standard transposition operation for matrices. It is straightforward to check that the following Laurent series is a solution to :
(19) |
The following proposition is straightforward to prove (see [28], Section 3.2.2)
Proposition 2.3.1.
Suppose that . In a neighbourhood of , the space of holomorphic solutions of is a subspace of co-dimension 1 in the space of all solutions.
From now on we will assume that . Under this condition every solution to , locally near is a sum of a holomorphic solution and for some constant . In particular, we can easily describe the local monodromy of near . The analytic continuation along a simple counter-clockwise loop around transforms where . Note that locally near the holomorphic solutions are precisely the monodromy invariant ones.
Remark 2.3.1.
In the case when there might be solutions involving .
Let us introduce the following pairings
(20) |
Using the differential equations of it is easy to check that is independent of and . In particular, it is a monodromy invariant pairing between the space of -twisted and -twisted periods. Note that we also have the following symmetry:
It turns out that there is an explicit formula for in terms of the Hodge grading operator and the nilpotent operator . Let us recall the so-called Euler pairing
(21) |
As a byproduct of the proof of Theorem 2.6.1 we will get the following simple formula:
where . Given a reference path (avoiding the discriminant) from to , there exists a vector , such that, the period vector . Since the series in (19) involves fractional powers of , the value of depends not only on the reference path but also on the choice of a branch for . In other words, the value of is unique up to a factor in the spiral . Note that fixing the reference path and the branch of determines for all . We will refer to as the -twisted reflection vector corresponding to the reference path.
Lemma 2.3.1.
a) We have .
b) If is such that is holomorphic at , then .
Proof.
The proof is obtained by substituting formula (19) into the definition (20) and extracting the leading order term in the Laurent series expansion at . If we do this for the pairing in part a) we will get
This proves a). The proof of b) is similar.∎
Proposition 2.3.2.
Let be the local monodromy transformation of corresponding to a simple counter-clockwise loop around . Then
where are -twisted reflection vectors corresponding to the simple loop.
2.4. Asymptotic expansions and Stokes matrices
Let us assume that the base point is such that for for all sufficiently close to . We write for for simplicity.
Let be the unit circle. A point is said to be an admissible direction if the ray does not pass through for . A direction which is not admissible is said to be critical. If is a critical direction, then is also critical. Therefore, the number of all critical directions is even, say for some . Following [2] we order the critical directions in a clockwise order in such a way that
Let us assume that for all . Our assumption for implies that is an admissible direction. This is going to be our default admissible direction. It will be convenient to introduce an auxiliary reference point . Note that if we continuously change the admissible direction from to , then we will obtain a path connecting and .
Suppose that is an admissible direction. Let us consider the following oscillatory integrals:
(22) |
where is a complex number, such that, . The integral is absolutely convergent for all in the half-plane
where is the complex conjugate of . Note that if (), then is the standard Euclidean pairing. Therefore, is the half-plane in whose boundary is the line orthogonal to and which does not contain . The definition of (22) requires also a choice of and in order to be able to define fractional powers of and . Since the admissible direction is obtained after a continuous deformation from , we may fix by continuity and by choosing . Since the choice of at uniquely determines a holomorphic branch of defined on the entire half-plane . Note that when we have . Moreover, for the convergence of the integral (22) we have to require that . Therefore, both and belong to and we have a natural choice of the value of . We get that is a holomorphic function for . Put and note that for we have
Recalling the stationary phase asymptotic method and using the expansion (19) we get that
(23) |
Suppose that and are two admissible directions, such that, and belong to the same clockwise arc from to ,i.e., the arc bounded by two adjacent critical directions. By definition, the sector between the rays and does not contain for . This implies that extends to a holomorphic function in that sector. Using the Cauchy residue theorem, it is easy to prove that for all . We get that for every admissible direction , the oscillatory integral (22) extends analytically in for all where and are the two critical directions adjacent to .
Figure (2) might help visualize the domains of analyticity. Let us denote by the matrix of size whose -th column is . Since both and are solutions to the Dubrovin connection for , there exists a matrix , such that,
(24) |
Similarly, there exists a matrix , such that,
(25) |
In both formulas (24) and (25) we define by continuously rotating on in clockwise direction. The matrices and are called Stokes matrices. There is a simple relation between and (see [8], Proposition 3.10).
Proposition 2.4.1.
We have where t is the usual transposition operation of matrices.
Proof.
Let be the matrix of the Frobenius pairing. Note that . We claim that . First of all, using that is a solution to the Dubrovin connection we get that is a constant independent of and . Let us recall the asymptotic expansions and where and . In particular, we have that both and have limit when and which must be (in both cases). Therefore, when in the half-plane . This implies that the diagonal entries of must be 1. For , since is an admissible direction, we can find , such that, . If , then the entry of , that is, has an exponential growth as in the direction of – contradiction. This completes the proof of our claim.
Suppose that . Then we have and because . We get
The following proposition is well known (see [8], Proposition 3.10).
Proposition 2.4.2.
Let be the -entry of the Stokes matrix . Suppose that . Then
-
(a)
The diagonal entries for all .
-
(b)
If , then .
Proof.
Using the asymptotic expansion (23) and the identity (24) we get that , where and , has a limit when which must be the identity matrix. Part (a) follows immediately from this observation. For part (b) we need only to notice that if , then has an exponential growth as . Therefore, the limit of exists only if . ∎
Remark 2.4.1.
Note that the condition in (b) is independent of the choice of , otherwise the direction of or must belong to the cone spanned by and contradicting the fact that there are no critical directions between and . ∎
Recalling Proposition 2.4.1 we get the following corollary.
Corollary 2.4.1.
Let be the -entry of the Stokes matrix . Suppose that . Then
-
(a)
The diagonal entries for all .
-
(b)
If , then . ∎
2.5. Stokes matrices and the intersection pairing
Let us continue to work in the settings of the previous subsection. We would like to construct a system of reference paths and express the entries of the Stokes matrix in terms of the Euler pairing and the reflection vectors corresponding to the reference paths. Let be the intersection of the ray and the circle (see Figure 1). Recall that we fixed an auxiliary reference point which is connected to by contuniously deforming the direction from to . We define the path to be the composition of the counter-clockwise oriented arc from to and the line segment from to (see the blue paths on Figure 1). Let be the reflection vector corresponding to the path , that is, . Following [2], let us introduce the vectors (), such that, . The pairing is non-degenerate except for finitely many . The definition of makes sense except for finitely many values of . The properties of the corresponding period vectors can be summarized as follows (compare with [2], Proposition 1 and Theorem 2’).
Proposition 2.5.1.
a) The period vector is analytic at for , where the value of the period is specified via the reference path .
b) The following formula holds:
c) Let be a contour starting at , approaching along the ray , making a small loop around , and finally returning back to along . Put
where the value of is determined by the reference path as follows: first we fix the value at the intersection of and , then we extend by continuity to the remaining points of . Then coincides with for all where in order to specify the value of and of we take a clockwise rotation from to .
Proof.
a) Using Proposition 2.3.2 we get that where we use that for the pairing . Therefore, the period is single-valued in a neighborhood of which is possible only if it is holomorphic.
b) This is obvious from the definition of .
c) According to parts a) and b), the difference is holomorphic at . Moreover, the periods being solutions to a Fuchsian differential equation, have at most polynomial growth at . Recalling the Cauchy residue theorem we get
(26) |
Using integration by parts, it is easy to check that is invariant under the shift . Therefore, we may assume that . Note that
(27) |
Indeed, let us split the integration contour into 3 pieces: going from to , an -loop around starting and ending at , and the ray from to . Recalling the Laurent series expansion (19), we get that under the analytic continuation along the -loop, the integrand gains a factor of . Since the orientations of the first and the third contours are opposite, we get that the corresponding integrals differ by a factor of . Furthermore, since , the integral along the loop has a contribution which vanishes in the limit . This completes the proof of formula (27). Using formulas (26) and (27) we get
The statement in part c) follows from the above formula. ∎
Suppose now that is another admissible direction, such that, . In other words, is obtained from by crossing the critical direction . Let and be the vectors corresponding to the reference paths . We would like to express and in terms of and . Let us split the points into groups, such that, each group belongs to a ray with direction and the rays of different groups are different. Let be one such group whose elements are ordered in such a way that for some real numbers . We will refer to such a sequence as -sequence. Clearly this splitting is uniquely determined by the critical direction .
Proposition 2.5.2.
Suppose that is a -sequence. Then
and
Proof.
For part a), let us look at figure (3). By definition, the period defined in a neighbourhood of is obtained from via the analytic continuation along the reference path . On the other hand, the analytic continuation of along the inverse of the reference path yields . The conclusion is that the cycle is obtained from after a monodromy transformation along a small modification of the loop consisting of the following 3 pieces: the line segment from to , the line segment from to , and the arc from to . The small modification, necessary to avoid the singularity , is taken as follows: when we approach along we have to stop slightly before hitting , make an anti-clockwise rotation along until we hit and then continue along the old contour. The reason why we have to make anti-clockwise rotation, and not clockwise, is that the value of , needed to define , is determined by (resp. ) when (resp. ). Since is obtained from by anti-clockwise rotation, we have to go around anti-clockwise. Clearly, the loop decomposes into simple loops going successively clockwise around the points in the given order, i.e., first around , then , etc., finally . After this discussion the first formula that we have to prove should be clear.
For the second formula, let us argue by induction. The fact that follows immediately from the first part of the proposition which implies that is a sum of and a linear combination of . Suppose that the formula is proved for all . Let us find coefficients , such that
satisfies the defining equations of . Note that for all . Therefore, we have to solve the equations () for . For , we get
(28) |
where we used that the pairing is monodromy invariant and we dropped from all terms that do not contribute. Note that by Proposition 2.3.2 we have
Substituting this formula in (28) we get . Suppose that we proved that for . In order to determine , let us consider the equation . We get
Recalling the ansatz for we get that in only the following terms will contribute:
We have for . Note that is fixed by and that . Therefore, we may replace the above expression with
(29) |
Let us add the first two terms. After pulling out the common expression we are left with
where we used the formula for . Therefore, after adding up the first two terms in (29) we get
Clearly we can continue adding up the first two terms until we reach
The -pairing of the above expression with must be 0. We get . This completes the proof. ∎
Let us denote by the matrix whose -entry is
Proposition 2.5.3.
Suppose that and are admissible directions separated by a single critical direction . Then
Proof.
The 2nd formula in Proposition 2.5.2 implies that
(30) |
Recalling Proposition 2.5.1, c), we get that in order to complete the proof it would be sufficient to deform the integration contours and () to a common contour without changing the values of the corresponding oscillatory integrals. This would be possible thanks to our special choice of reference paths.
Let be the last entry of the -sequence containing . We pick a contour consisting of 3 parts: the ray with orientation from to , the counter-clockwise arc from to , and finally the ray (see Figure 4).
Let be a point in the -sequence preceding . We claim that
(31) |
and
(32) |
Let us justify the first identity. The argument for the second one is similar. To begin with, note that is invariant under the monodromy transformations (the monodromy transformation corresponding to the simple loop ) for . Thanks to our special choice of the reference paths, i.e., the reference paths do not intersect the ray , the fundamental group of is generated by the simple loops corresponding to the paths with . Therefore, the period integral extends to a holomorphic function in for all . In particular, extends to a holomorphic function in the domain bounded by the contours and . Furthermore, for the integrand will have an exponential decay at infinity in . Therefore, the identity follows from the Cauchy residue theorem. Finally, the formula that we have to prove follows from Proposition 2.5.1, c) and formulas (30), (31), and (32). ∎
Now we can express the Stokes matrix in terms of the reflection vectors. Let us first extend the definition of the critical directions () by allowing arbitrary so that . More precisely, is obtained from by clockwise rotation on angle Such an extension is clearly unique. Note that we have the following symmetry:
(33) |
Recall that is the matrix with columns . We proved that where is an admissible direction obtained from by crossing the critical direction and . Note that is obtained from by a clockwise rotation. If is rotated across to , then we get
where again is obtained from via a clockwise rotation. Continuing in this way, i.e., rotating clockwise until it crosses all the critical directions , we get
where we may take because we can start with sufficiently close to and at the end cross and stay sufficiently close to . Recalling the definition of the Stokes matrix we get
(34) |
where is an admissible direction whose adjacent critical directions are and . Recalling the relation (see Proposition 2.4.1) we get
Slightly modifying the above argument we will obtain a simpler formula for the Stokes matrices (see [2], Proposition 5). To begin with, we need an analogue of Proposition 2.5.2. If necessary let us change the enumeration of the points so that the following property holds: if we draw a line at parallel to and we stand at looking towards infinity in the direction , then all points with (resp. ) will be in the RHS (resp. LHS) half-plane. Note that is equivalent to for all . Therefore, recalling Proposition 2.4.2 we get that for , that is, is an upper-triangular matrix with ones on the diagonal.
Let be the reflection vector corresponding to the reference path obtained by composing and the counter-clockwise oriented arc from to (see Figure 1). Note that thanks to our choice of the indexes of we have
(35) |
Note that the formulas about () in Proposition 2.5.2 were derived in a purely algebraic way from the relations between () and (). Therefore, in the current settings we must have
(36) |
Let us define the matrix of size whose entry is
(37) |
Arguing in the same way as in the proof of Proposition 2.5.3 we get for all . Recalling the definition of the Stokes matrix we get that and hence , that is, formulas (37) give the entries of the inverse Stokes matrix .
Finally, we will finish this section by proving that the Stokes matrices and are independent of . More precisely, we will express the pairings in terms of the intersection pairing . We follow the idea from the proof of Lemma 2’ in [2].
Lemma 2.5.1.
Let be a complex number with . Let us fix a negative integer , such that, the real part of is positive. Then
Proof.
It is sufficient to prove the formula locally near . We have a Laurent series expansion
Using the substitution and the standard formulas for the Euler -integral we get
It remains only to note that , that is, substituting the Laurent series expansion of and termwise integrating in yields precisely the Laurent series of . ∎
Proposition 2.5.4.
The pairing takes the following form in the basis of reflection vectors () corresponding to the reference paths () :
Proof.
Let be a point sufficiently close to . Let us choose , such that, its real part is sufficiently negative. Let us consider the following difference
(38) |
where (resp. ) means that the value of the period is obtained from via analytic continuation along a path which approaches along , makes a small counter-clockwise (resp. clockwise) arc around , and continues towards along . We would like to compute (38) in two different ways. First, by definition where we put a sign to emphasize that the reference path should contain the clockwise arc around . We get that the difference (38) coincides with
(39) |
On the other hand, the analytic continuation can be computed using the integral formula from Lemma 2.5.1. Namely,
(40) |
where the integration path is from to (see Figure 1), the arc from to (clockwise for and anti-clockwise for ), the line segment approaching along , a small clockwise (for ) or anti-clockwise (for ) arc around , and finally a straight line segment to . Note that the integral splits into two
where the first integral does not depend on the choice of an arc around . Since is an integer, we have Therefore, the difference (38) takes the following form:
(41) |
Note that in the above formula is obtained by continuously varying a small line segment along the integration path in (40). The starting value of the argument is . If , then the segment will be rotated anti-clockwise on angle , so the final value of will be . If , then the segment will be rotated clockwise and the value of will eventually become . Recalling Lemma 2.5.1 we have
where should be . The conclusion is that the expression (41), that is the difference (38), coincides with for and with for . Note that . Comparing with our previous formula (39) we get the statement of the proposition for the case when . The case was already considered (see Lemma 2.3.1). ∎
2.6. The central connection matrix
Let us start by introducing Figure 5 which might be helpful in visualizing the constructions and following the arguments in this section. Let us identify in the standard way. There are two kinds of objects on Figure 5: points and vectors. We think of the points as elements of the -plane and of vectors as elements of the -plane. For example, when we talk about the sector then one can think about the vectors in the shaded region on Figure 5.
Suppose that is an admissible direction and let be the matrix whose columns are defined by the oscillatory integrals (22). Similarly, let us introduce also where the admissible direction is obtained from by a clock-wise rotation. Then we have as and and as and .
Recall that the definition of requires a choice of , that is we need to define an analytic branch of on both and . Since and by definition is continuously connected to the default admissible direction , we have a natural choice of a branch of on . Using the clock-wise arc from to we can extend analytically across the ray (see Figure 5) to the entire half-plane . Recall that both and extend analytically in by deforming the integration contour of (see Section 2.4). We get that , , and are 3 fundamental solutions to the Dubrovin connection analytic inside the sector . Therefore, there exist matrices and , such that,
for all . The matrix is the Stokes matrix introduced earlier. Following Dubrovin (see [10]) we will refer to as the central connection matrix. The main goal in this section is to find a formula for in terms of the reflection vectors.
It is more convenient to work with the matrix whose columns are defined in Proposition 2.5.1, c). According to Proposition 2.5.1
where . The key formula will be proved in the following proposition (see [10], Theorem 4.19).
Proposition 2.6.1.
Let be the th column of the matrix . Then
Proof.
By definition (see Proposition 2.5.1)
where . Let us analytically extend the above identity with respect to to the boundary of . On the RHS we use that . Suppose that is on the right (compared to the -direction) part of the boundary of . The analytic continuation of the LHS is given by deforming the contour to the line where is a real number (see the blue contours on Figure 5). We get
(42) |
where Let us recall that , where is determined from the branch of in and the value . Since we have restricted we get . On the other hand, in formula (42) the analytic branch of comes from the branch of in induced from . In other words, . Therefore,
Let us rewrite (42) as follows
Let us substitute , . Recalling the Laplace inversion formula we get that is the Laplace transform of the RHS, that is,
(43) |
where the integration is along the -ray, that is, . Let us comment on the convergence of the above integral. Firstly, since the calibration is analytic at , we need to choose , such that, . When is close to , since the integrand is proportional to , the integrand has at most exponential growth of order . Therefore, the integral defines an analytic function for all in the half-plane . The integral (43) is straightforward to compute because the calibration is an entire function. In other words, we may use the Taylor series expansion at . Let us also make the substitution and restrict to be such that . The RHS of (43) takes the following form
Note that and that can be produced by acting with the differential operator . The formula transforms into
where the arrow over is to denote right action of the matrix differential operator . We have to distinguish left and right action here because and do not commute. The above integral is just the definition of the -function. We get
Using the product formula
with we get
(44) |
Note that
where we used that thanks to the commutation relation . The left action can be transformed into right action thanks to the following formula:
The above formula is proved by expanding , using the commutation relation , and finally noting that the shift is equivalent to differentiation by . After all these remarks, we can easily transform (44) into
The infinite sum over is precisely our definition of the fundamental solution of the second structure connection . The formula for follows. ∎
Now we are in position to derive the precise formulas relating the monodromy data of the 1st and the 2nd structure connections. Let us first state the following simple but very useful formula:
(45) |
where is the matrix of the Frobenius pairing, that is, , T is transposition with respect to the Frobenius pairing, and t is the usual transposition of matrices. In the above identity, we use a fixed basis of flat vector fields to identify the space of linear operators in with the space of matrices of size . Let us also introduce the following convenient notation:
Let us recall the well known relations between the Stokes matrices and the central connection matrix (see [10]).
Proposition 2.6.2.
The following formulas hold:
Proof.
The 2nd formula follows from the first one and the relation . Let us prove the first formula. Suppose that . We have
(46) |
where we used the quadratic relation and the relation – see Proposition 2.4.1 and its proof. On the other hand, by definition, we have . Let us analytically extend this identity in from to along the anti-clockwise arc. We get a second identity of the form . Substituting these two formulas in (46) we get
where we used repeatedly formula (45), the symplectic condition , and the relation . ∎
Let us introduce the matrix of the pairing , that is, . Let be the matrix with columns the reflection vectors , that is, the entries of are defined by . Similarly, let be the matrix whose columns are given by the dual vectors . Using formula (37) for the entries of and Propositions 2.5.4, we get that coincides with the -entry of . Therefore,
(47) |
On the other hand, since by definition , we get , that is,
Recalling Proposition 2.6.2 we get
Recalling Proposition 2.6.1 we get
(48) |
Theorem 2.6.1.
a) The -entry of the central connection matrix is related to the components of the reflection vectors by the following formula:
c) The reflection vectors () are independent of and the Gram matrix of the Euler pairing is upper-triangular:
with ’s on the diagonal: .
d) The Gram matrix of the Euler pairing in the basis () coincides with the inverse Stokes matrix .
Proof.
a) According to (48) we have . Comparing the entries in this matrix identity we get the formula stated in part a).
b) According to (47) we have
Recalling Proposition 2.6.2 we get
Finally, since we get
The above formula implies that
c) The fact that is independent of follows immediately from part a) because the central connection matrix is independent of . The rest of the statement is an immediate consequence of Proposition 2.5.4 and part b). Indeed, if , then we have
On the other hand, recalling part b) with we get . The above identity is possible if an only if . Similarly, if , then we have
which implies that .
d) This part is an immediate consequence of formula (37) and parts b) and c). ∎
Remark 2.6.1.
Let us compare our notation to Dubrovin’s one in [10]. If is an admissible direction, then is the positive part of an admissible line in the sense of Dubrovin. Then , , and . Note that the degree operator in Dubrovin is while the nilpotent operators coincide . It follows that the inverse Stokes matrix coincides with Dubrovin’s Stokes matrix . Finally, the central connection matrix in our notation coincides with the Dubrovin’s one. ∎
3. Dubrovin conjecture
Following [9] (see also [6]) we present the so-called Dubrovin conjecture. Roughly speaking, Dubrovin conjecture relates the big quantum cohomology of a variety , as a Frobenius manifold, with its bounded derived category of coherent sheaves. The main goal is to give a reformulation of the conjecture in terms of the language introduced in the previous section.
3.1. Quantum cohomology
We recall some basic aspects about quantum cohomology, for a more detail account, for instance see [7]. Let be a smooth projective algebraic variety, with vanishing odd cohomology, and be the Deligne-Mumford stack of -pointed stable maps of genus representing a class . Let us consider the evaluation maps , , and the map . Then, the descendant Gromov-Witten invariant is defined by the following formula:
where is the virtual fundamental class in the Chow ring constructed in [4] and is the first Chern class of the tautological line bundle formed by the cotangent line at the -th marked point. In particular, genus-0 Gromov-Witten invariants (with no descendants) can be used to define a deformation of the cup product in cohomology.
Definition 3.1.1.
Let be a complexified Kähler class on a smooth projective variety . Let be a basis of and . Then, the Gromov-Witten potential is defined by the following formula:
where (with taken times) and . ∎
If is a Fano variety, then there are only a finite number of ’s such that , so , where the ’s are the formal variables associated to the basis . Therefore can be considered as a function on a formal neighbourhood of . In general, we fix an ample basis of and introduce the so called Novikov variables . The expression is interpreted as an element in the ring of formal power series and the potential is considered as a formal power series in the ring
where we identified () and we used the divisor equation to express as a function of . It is believed that the Gromov–Witten potential is convergent (see below for a more precise statement). In case of convergence, the complexified Kähler class is related to the Novikov variables via .
Definition 3.1.2.
The big quantum cohomology of is the ring , with the product given on generators by , where form a Poincaré dual basis to . We will denote this ring by .∎
Remark 3.1.1.
If we set and , then
In addition, if we set , then we get that
This restriction is known as the small quantum product and the corresponding ring is called small quantum cohomology ring. In fact, since the big quantum product is defined formally on in terms of and , the small quantum cup product is obtained by restricting to , that is, setting in the formula for . This is equivalent to setting .∎
Suppose now that the Novikov variables and that there exists a non-empty open subset where the Gromow-Witten potential converges. More precisely, we assume that contains , such that, () and () are complex numbers with sufficiently small length ( could be arbitrary because is polynomial in ). Let
be the Poincaré pairing which will be taken as a Frobenius pairing. Put
where is half of the standard cohomology degree. Then, equipped with the big quantum cup product defined above, the Poincaré pairing, and the Euler vector field is a Frobenius manifold of conformal dimension . The semi-simplicity of the quantum cup product is an indication that the target smooth algebraic variety has many rational curves. Therefore, from the point of view of birational geometry, it is very important to understand when is the Frobenius manifold underlying the big quantum cohomology semisimple? Before trying to approach an answer to this question, we need to recall some background on the bounded derived category of coherent sheaves of .
3.2. Derived categories
For more details about derived categories we refer to [14]. Let be a -linear triangulated category. Let us recall the following notation. Given an object put where is the translation functor of the triangulated category. Furthermore, denotes the complex vector space of morphisms in from to . Let us introduce also the complex of vector spaces with a trivial differential whose component in degree is .
Definition 3.2.1.
An object in is called exceptional if it satisfies the following conditions:
Definition 3.2.2.
A sequence of objects is called an exceptional collection if every object is exceptional and for . An exceptional collection is said to be full if it generates as a triangulated category. ∎
Following Bondal (see [5]) we would like to recall the mutation operations. An exceptional collection consisting of two objects is said to be an exceptional pair. Let be an exceptional pair. We define objects and , such that, the following sequences are distinguished triangles
where for a complex of vector spaces we denote by the direct sum of copies of and by the direct sum of all . The map is induced from the tautological maps , that is, fix a basis of , then is a morphism . Similarly, the map is induced from the tautological maps , that is, fix a basis of and a dual basis of , then is a morphism from . Note that taking the dual changes the sign of the grading: is in degree while is in degree . It is easy to check that both and are exceptional pairs. More generally, given an exceptional collection we define left mutation and right mutation by mutating the adjacent objects and , that is,
It turns out that these operations define an action of the braid group of strings on the set of exceptional collections, that is, the following commutation relations hold (see [5], Assertion 2.3):
and
where . Using mutations we can construct the so-called left Koszul dual of the exceptional sequence , that is, the exceptional sequence defined by
is called the left Koszul dual of . More explicitly,
In other words, using left translations (see [5], Section 2), we are moving first to the left through , then in the resulting collection we move to the left through etc.. We are not going to use it but let us point out that one can define a right Koszul dual in a similar way, that is, . As we will see now, the left (resp. right) Koszul dual corresponds to changing the admissible direction to by an anti-clockwise (resp. clockwise) rotation.
Definition 3.2.3.
Let be a full exceptional collection. The helix generated by is the infinite collection defined by the iterated mutations
A foundation of a helix is any family of consecutive objects . The collection is called the marked foundation. ∎
Definition 3.2.4.
Let be the set of isomorphism classes of objects of . The Grothendieck group of is defined as the quotient of the free abelian group generated by and the Euler relations: whenever there exist a triangle in ∎
We can define the so-called Grothendieck-Euler-Poincaré pairing as
for any pair of objects and in . Note that on the level of -theoretic groups the left and right mutations take the form
Lemma 3.2.1.
Let be a full exceptional collection and be its left Koszul dual. Then .
Proof.
Let and . Let us introduce the reflection where and is generic, such that, is a non-degenerate pairing. Using that for , it is easy to check that
Note that the relations in the above formula are identical to the relations in (35). Therefore, if we define and , such that, and , then we have a relation corresponding to (36)
Note that because for . Let be the Gram matrix of in the basis , that is, . Similarly, let be the Gram matrix of in the basis . Suppose that is the matrix describing the transition between the two bases: . Then we have
which implies that
Definition 3.2.5.
A unimodular Mukai lattice is a finitely generated free -module with a unimodular bilinear form (not necessarily symmetric) .
An element is called exceptional if . A -basis of the Mukai lattice is called exceptional if , and for .∎
Remark 3.2.1.
The projection on of a full exceptional collection in is an exceptional basis. The pair (, ), where is the Grothendieck-Euler-Poincaré pairing defined above, is a unimodular Mukai lattice. The matrix whose -entry is is called the Gram matrix associated to the exceptional collection ∎
We are interested in the case where the triangulated category is the bounded derived category of coherent sheaves of a smooth algebraic variety , which we denote by . It is interesting to know in which cases has a full exceptional collection.
3.3. Original formulation of Dubrovin conjecture
For a smooth Fano variety we have formulated two questions. The first one is about the semisimplicity of the big quantum cohomology and the second one is about the existence of a full exceptional collections in . At first sight, the questions seem unrelated but this is not the case. In fact, the main content of Dubrovin conjecture is precisely that the answer of these two questions should be linked. In his ICM talk in 1998, following a proposal of Alexey Bondal, Dubrovin proposed the following conjecture (see [9] and also [6]).
Conjecture 3.3.1 (Dubrovin 1998).
Let be a Fano variety.
-
(1)
The big quantum cohomology is semisimple if and only if admits a full exceptional collection , where .
-
(2)
The Stokes matrix for the first structure connection is equal to the Gram matrix for , i.e., .
-
(3)
The central connection matrix hast the form , where the columns of are the components of and is some operator satisfying for any .
There are two important developments that lead to a modification of the conjecture. First of all, it was suggested by Arend Bayer (see [3]) that the Fano condition is not important, so it should be dropped. Second, a precise statement about the central connection matrix, was proposed independently in [13] by Galkin-Golyshev-Iritani and in [6] by Cotti-Dubrovin-Guzzetti. An important role in this refinement is played by the so-called Gamma class.
3.4. Refined version of the conjecture
Let be a smooth projective variety of complex dimension . The cohomology class (see [13]) , where are the Chern roots of and is the Gamma function, is called the Gamma class. Following [6], we introduce also the class .
Remark 3.4.1.
Remark 3.4.2.
The Gamma class can be expended as
where is the Euler constant. This is obtained from the Taylor expansion for the Gamma function. ∎
The other ingredient introduced in [6] is given by two morphisms which are defined as follows. Let . Since is smooth, the object is isomorphic to a bounded complex of locally free sheaves , therefore a graded version of the Chern character can be defined as where where the sum is over the Chern roots of . Note that the standard Chern character is where . In other words, the difference between and is in re-scaling each Chern root by . The morphisms are defined as follows:
(50) |
where is the remainder of the division of by In order to define the monodromy data of a Frobenius manifold, Cotti–Dubrovin–Guzzetti have introduced chambers for every oriented line with orientation specified by a unit vector , . In our notation, is an open subset of the Frobenius manifold consisting of semi-simple points , such that,
-
(i)
The canonical coordinates are pairwise distinct.
-
(ii)
The vector is an admissible direction, that is, is not parallel to for all .
In every chamber , the canonical coordinates are enumerated according to the so-called lexicographical order: if then , or equivalently if we stand at and look in the admissible direction , then is on our right. The Frobenius manifold underlying quantum cohomology has a natural calibration given by the -matrix where are defined by
The nilpotent operator . The monodromy data is defined as explained in Sections 2.4 and 2.6. Namely, there are unique solutions as holomorphic for where (resp. ) is the right (resp. left) half-plane bounded by the oriented line . The 3 solutions to the quantum connection , , and are analytic in in a sector containing the positive part of the line and hence we can define the Stokes matrix and the central connection matrix by
We refer to Remark 2.6.1 where we explained the correspondence between our notation and the Dubrovin’s notation. Let us point out that the notation for the central connection matrix in [10] and [6] are different: one is the inverse of the other. We stick to the notation from [10].
Conjecture 3.4.1 (Refined Dubrovin conjecture 2018, see conjecture 5.2 in [6]).
Let be a smooth Fano variety of Hodge-Tate type, then
-
(1)
The big quantum cohomology is semisimple if and only if there exists a full exceptional collection in .
-
(2)
If is semisimple and convergent, then for any oriented line (of slope ) in the complex plane, there is a correspondence between -chambers and helices with a marked foundation in .
-
(3)
The monodromy data computed in an -chamber , in the lexicographical order, is related to the following geometric data of the corresponding exceptional collection (the marked foundation):
-
(3a)
The Stokes matrix is equal to the Gram matrix of the Grothendieck-Poincaré-Euler product on , computed with respect to the exceptional basis , that is, .
-
(3b)
The inverse central connection matrix coincides with the matrix associated with the -linear morphism defined above – see (50). The matrix is computed with respect to the exceptional basis and any pre-fixed cohomological basis .
-
(3a)
Some parts of the Dubrovin conjecture, in its original or refined forms, have been verified for several Fano varieties, see [6] for a detailed account about the cases where the conjecture has been proved.
3.5. Exceptional collections, reflection vectors, and Dubrovin conjecture
Motivated by the definition of a distinguished basis in singularity theory (see [1, 12]), let us define a distinguished system of reference paths. Recall that we have fixed a reference point , such that, for all and for all where are the canonical coordinates of . Let be the disk with center and radius (recall that is a positive real number).
Definition 3.5.1.
A system of paths inside is said to be a distinguished system of reference paths if
-
(i)
The path has no self-intersections and it connects with one of the points .
-
(ii)
For each pair of paths and with , the only common point is .
-
(iii)
The paths exit the point in an anti-clockwise order counted from the boundary of the disk .∎
Two distinguished systems of reference paths and will be considered homotopy equivalent if there exists a continuous family , , such that, is a distinguished system of reference paths and and (for more details see [12], Section 5.7). The braid group on strings acts naturally on the set of homotopy equivalence classes of distinguished reference paths. Namely, we have the operations
where is a small perturbation of the composition of and the anti-clockwise simple loop corresponding to . Similarly, we have the operation
where is a small perturbation of the composition of and the clockwise simple loop corresponding to . The operation is inverse to and the following braid group relations are satisfied:
The braid group on strings acts transitively on the set of homotopy equivalence classes of distinguished reference paths. This is almost an immediate consequence of the definition of a braid (see [12], Proposition 5.15).
Remark 3.5.1.
In singularity theory, one usually requires the order of the distinguished system of reference paths to be clockwise. We change it to anti-clockwise in order to have an agreement between the lexicographical order defined by an admissible direction. ∎
Remark 3.5.2.
The operations and are usually denoted by and . Our notation is motivated by the corresponding notation for left and right mutations in the case of exceptional collections. ∎
Partially motivated by the work of Milanov–Xia (see [26], Conjecture 1.6), we would like to propose the following conjecture.
Conjecture 3.5.1.
If is a set of reflection vectors corresponding to a system of distinguished reference paths, then there exists a full exceptional collection in , such that, for all .
Conjecture 3.5.1 is weaker than Conjecture 1.6 in [26]. Namely, the proposal in [26] is that every full exceptional collection determines a set of reflection vectors. This is more difficult to prove. Nevertheless, it is expected that any two full exceptional collections are related by a sequence of mutations. If this expectation is correct then Conjecture 1.6 in [26] is equivalent to Conjecture 3.5.1.
Theorem 3.5.1.
Conjecture 3.5.1 is equivalent to the refined Dubrovin conjecture.
Proof.
Let us consider the case when all Novikov variable . The general case can be obtained by applying the divisor equation. Suppose that the distinguished system of reference paths is given by () where is an admissible direction. Assuming that the refined Dubrovin conjecture holds, let us derive the formulas for the reflection vectors in Conjecture 3.5.1. Note that the exceptional collection will be different from . The inverse central connection matrix defines a map which according to the refined Dubrovin conjecture is given by
If is one of the basis vectors, then we have
where we used the formula for the -entry of from Theorem 2.6.1, part a). We get
Let us multiply the above identity, using the classical cup product, by where will be specified later on. Recall that
where the product is over all Chern roots of and is the Todd class of . Note also that where is the canonical bundle of . We get
Let us integrate the above identity over . For dimensional reasons, we can replace the expression with Recalling the Hierzerbruch–Riemann–Roch formula we get
Since the Poincare pairing is non-degenerate, the above formula implies
(51) |
Let us recall that by Serre duality, that is, , we have Therefore,
Clearly, we can choose -theoretic vector bundles , such that, . Once this choice is made, we get from (51) that where . We claim that are the K-theoretic classes of a full exceptional collection. Indeed, recalling Lemma 3.2.1, it is sufficient to choose a full exceptional collection , such that, its left Koszul dual is . Once we do this, we can simply put . In order to define , we simply have to invert the sequence of mutation operations that define the left Koszul dual, that is,
In order to complete the proof that the refined Dubrovin conjecture implies Conjecture 3.5.1 we need only to recall the braid group action. More precisely, note that after a small perturbation the system of reference paths corresponding to an admissible direction becomes a distinguished system of reference paths. Therefore, we need only to check the following statement. Suppose that Conjecture 3.5.1 holds for a distinguished system of reference paths satisfying the extra condition for where are the reflection vectors corresponding to . Then we calim that the conjecture holds for and moreover satisfies the extra condition: for where are the reflection vectors corresponding to . Note that
We have
It is straightforward to check that for . Thanks to formula (11), it is easy to check that if for some exceptional collection , then where , that is, Conjecture 3.5.1 holds for .
Finally, for the inverse statement, that is, Conjecture 3.5.1 implies the refined Dubrovin conjecture, one just has to go backwards. We leave the details as an exercise. ∎
Let us point out the following important property of a distinguished bases which was obtained as a byproduct of the proof of Theorem 3.5.1.
Proposition 3.5.1.
Let be a set of reflection vectors corresponding to a distinguished system of reference paths. Then the Gram matrix of the Euler pairing is upper triangular: for all and .
Indeed, if the reference paths correspond to an admissible direction, then the statement was already proved in Theorem 2.6.1, part c). In the proof of Theorem 3.5.1 we proved that the statement is invariant under the action of the braid group. Therefore, since the braid group acts transitively on the set of distinguished system of reference paths, the statement of Proposition 3.5.1 is clear.
Remark 3.5.3.
Galkin–Golyshev–Iritani proposed in [13] the so-called -Conjecture II. Roughly this conjecture says that the columns of the central connection matrix are the components of , where , for an exceptional collection . It was proved in [6] that -conjecture II is equivalent to part (3b) of the refined Dubrovin conjecture. Moreover, it was proved in [13] that -conjecture II implies part (3a) of the refined Dubrovin conjecture. Therefore, Conjecture 3.5.1 is also equivalent to -conjecture II. ∎
Remark 3.5.4.
Halpern-Leistner proposed in [19] the so-called noncommutative minimal model program. This is a set of conjectures about canonical paths on the space of stability conditions that imply previous conjectures about . In particular, it implies one direction of Dubrovin conjecture regarding the existence of exceptional collections. It would be interesting to see the relations between the canonical paths in the noncommutative minimal model program and the reflection vectors corresponding to a system of distinguished reference paths in Conjecture 3.5.1. ∎
References
- [1] V. Arnold, S. Gusein-Zade, and A. Varchenko. Singularities of differentiable maps II. Birkhauser, Basel–Boston, 1985.
- [2] W. Balser, W.B. Jurkat, and D.A. Lutz. On the reduction of connection problems for differential equations with an irregular singular point to ones with only regular singularities, I. SIAM J. Math. Anal., 12(5):691–721, 1981.
- [3] Arend Bayer. Semisimple quantum cohomology and blowups. IMRN, 2004(40):2069–2083, 2004.
- [4] Kai Behrend and Barbara Fantechi. The intrinsic normal cone. Invent. Math., 128(1):45–88, 1997.
- [5] Alexey Bondal. Representation of associative algebras and coherent sheaves. Mathematics of the USSR-Izvestiya, 34(1):23, feb 1990.
- [6] G. Cotti, B. Dubrovin, and D. Guzzetti. Helix Structures in quantum cohomology of Fano varieties. arXiv:1811.09235v2, pages 1–149, 2018.
- [7] David Cox and Sheldon Katz. Mirror Symmetry and Algebraic Geometry. Math. Surveys and Monographs 68. Amer. Math. Soc, 1999.
- [8] Boris Dubrovin. Geometry of D topological field theories. In Integrable systems and quantum groups (Montecatini Terme, 1993), volume 1620 of Lecture Notes in Math., pages 120–348. Springer, Berlin, 1996.
- [9] Boris Dubrovin. Geometry and Analytic theory of Frobenius manifolds. Documenta Mathematica, pages 315–326, 1998.
- [10] Boris Dubrovin. Painlevé Transcendents in Two-Dimensional Topological Field Theory, pages 287–412. Springer New York, New York, NY, 1999.
- [11] Boris Dubrovin. On almost duality for Frobenius manifolds. In Geometry, Topology, and Mathematical Physics, volume 212 of Amer. Math. Soc. Transl. Ser. 2, pages 75–132. Providence, RI, 2004.
- [12] Wolfgang Ebeling. Functions of several complex variables and their singularities, volume 83 of Graduate Studies in Mathematics. AMS, Providence, Rhode Island, 2007.
- [13] Sergey Galkin, Vasily Golyshev, and Hiroshi Iritani. Gamma classes and quantum cohomology of Fano manifolds: Gamma conjectures. Duke Mathematical Journal, 165(11):2005 – 2077, 2016.
- [14] Sergei Gelfand and Yuri Manin. Methods of homological algebra, volume Second Edition of Springer Monographs in Mathematics. Springer–Verlag Berlin Heidelberg, 2003.
- [15] Alexander Givental. Twisted Picard–Lefschetz formulas. Funktsional. Anal. i Prilozhen., 22(1):10–18, 1988.
- [16] Alexander Givental. Gromov–Witten invariants and quantization of quadratic hamiltonians. Mosc. Math. J., 1:551–568, 2001.
- [17] Alexander Givental. singularities and KdV hierarchies. Mosc. Math. J., 3(2):475–505, 2003.
- [18] Davide Guzzetti. On Stokes Matrices in terms of Connection Coefficients. Funkcialaj Ekvacioj, 59(3):383–433, 2016.
- [19] D. Helpern-Leistner. The noncommutative minimal model program. arXiv:2301.13168v2, pages 1–34, 2023.
- [20] K. Hori and M. Romo. Exact results in two-dimensional (2,2) supersymmetric gauge theory with boundary. arXiv:1308.2438, 2013.
- [21] H. Iritani. An integral structure in quantum cohomology and mirror symmetry for toric orbifolds. Adv. Math, 222(3):1016–1079, 2009.
- [22] Hiroshi Iritani. Quantum cohomology of blowups. arXiv:2307.13555, 2023.
- [23] M. Kontsevich L. Katzarkov and T. Pantev. Hodge theoretic aspects of mirror symmetry. In From Hodge Theory to Integrability and TQFT: -geometry, volume 78 of Proceedings of Symposia in Pure mathematics, pages 87–174. Providence, RI, 2008.
- [24] Yu. Manin and S. Merkulov. Semisimple Frobenius (super)manifolds and quantum cohomology of . Topological Methods in Nonlinear Analysis, 9(1):107 – 161, 1997.
- [25] Yuri Manin. Frobenius manifolds, quantum cohomology, and moduli spaces, volume 47 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 1999.
- [26] T. Milanov and X. Xia. Reflection vectors and quantum cohomology of blow ups. SIGMA, 20(29):60, 2024.
- [27] Todor Milanov. The period map for quantum cohomology of . Adv. in Math., 351:804–869, 2019.
- [28] Todor Milanov and Kyoji Saito. Primitive forms and vertex operators. Book in preparation.