This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Generic solutions of equations involving the modular jj function

Sebastian Eterović School of Mathematics, University of Leeds, Leeds, UK s.eterovic@leeds.ac.uk
Abstract.

Assuming a modular version of Schanuel’s conjecture and the modular Zilber–Pink conjecture, we show that the existence of generic solutions of certain families of equations involving the modular jj function can be reduced to the problem of finding a Zariski dense set of solutions. By imposing some conditions on the field of definition of the variety, we are also able to obtain versions of this result without relying on these conjectures, and even a result including the derivatives of jj.

Key words and phrases:
Strong existential closedness, jj function, modular Schanuel conjecture, Ax–Schanuel, Zilber–Pink
2020 Mathematics Subject Classification:
11F03, 11J89, 03C60
Supported by NSF RTG grant DMS-1646385 and EPSRC fellowship EP/T018461/1. I would like to thank Vahagn Aslanyan, Sebastián Herrero, Vincenzo Mantova, Adele Padgett, Thomas Scanlon and Roy Zhao for helpful discussions on some of the details presented here
ORCiD: 0000-0001-6724-5887

1. Introduction

In this paper we study the strong part of the Existential Closedness Problem (strong EC for short) for the modular jj function. The strong EC problem asks to find minimal geometric conditions that an algebraic variety V2nV\subset\mathbb{C}^{2n} should satisfy to ensure that for every finitely generated field KK over which VV is defined, there exists a point (z1,,zn)(z_{1},\ldots,z_{n}) in n\mathbb{H}^{n} such that (z1,,zn,j(z1),,j(zn))(z_{1},\ldots,z_{n},j(z_{1}),\ldots,j(z_{n})) is a point of VV which is generic over KK. The results of [AEK21] and [AK22] inform what the conditions of strong EC should be: in technical terms, it is expected that broadness and freeness is the minimal set of conditions (see §3.1 for definitions and Conjecture 3.5 for a precise statement).

When approaching the strong EC problem, the first immediate obstacle is the EC problem, which simply asks to find geometric conditions that an algebraic variety V2nV\subset\mathbb{C}^{2n} should satisfy to ensure that VV has a Zariski dense set of points in the graph of the jj function. When VV has such a Zariski dense set of points, we say that VV satisfies (EC). Following from the previous paragraph, it is expected that if VV is free and broad, then VV satisfies (EC) (see Conjecture 3.5). It was proven in [EH21, Theorem 1.1] that if V2nV\subset\mathbb{C}^{2n} is a variety such that the projection π:Vn\pi:V\rightarrow\mathbb{C}^{n} onto the first nn coordinates is dominant, then VV satisfies (EC). This is a partial solution to the EC problem, since asking that the projection π\pi is dominant is a stronger hypothesis than the notion of broadness.

Regarding strong EC, it was shown in [EH21, Theorem 1.2] that if one assumes a modular version of Schanuel’s conjecture, then any plane irreducible curve V2V\subset\mathbb{C}^{2} which is not a horizontal or a vertical line has a generic point over any given finitely generated field over which it is defined. On the other hand, [AEK23, Theorem 1.1] provides a version of this without assuming the modular Schanuel conjecture, but instead assuming that VV is in some precise sense “generic” with respect to the jj function.

The main results of this paper extend [EH21, Theorem 1.2] and [AEK23, Theorem 1.1] to higher dimensions. Our first main result shows that under a modular version of Schanuel’s conjecture (which we call MSCD, see Conjecture 2.3) and the modular Zilber–Pink conjecture (MZP, see Conjecture 3.10) the strong EC problem can be reduced to the EC problem.

Theorem 1.1.

Let KK\subseteq\mathbb{C} be a finitely generated field, and let V2nV\subseteq\mathbb{C}^{2n} be a broad and free variety defined over KK satisfying (EC). Then MSCD and MZP imply that VV has a point of the form (z1,,zn,j(z1),,j(zn))(z_{1},\ldots,z_{n},j(z_{1}),\ldots,j(z_{n})), with (z1,,zn)n(z_{1},\ldots,z_{n})\in\mathbb{H}^{n}, which is generic over KK.

This addresses the second question posed in [EH21, §1], and as such, this paper can be seen as a continuation of the work done there. We will also show some general cases in which we can remove the dependence on MZP, see Theorem 4.24. In particular, this has consequences on the dynamical behaviour of the jj function, see Corollary 4.26.

Using that the Ax–Schanuel theorem for jj ([PT16]) implies weak forms of both MSCD and MZP, we are able to remove the dependency on these conjectures from Theorem 1.1 by instead imposing conditions on the field of definition of VV, conditions we are calling “having no CjC_{j}-factors” (see §5 for the definition). Here CjC_{j} is a specific countable algebraically closed subfield of \mathbb{C} which is built by solving systems of equations involving the jj function, and the condition of “having no CjC_{j}-factors” ensures that VV is sufficiently generic so that the weak forms of MSCD and MZP suffice. A version of this condition already appeared in the hypotheses of [AEK23, Theorem 1.1]. We now state our second main result.

Theorem 1.2.

Let V2nV\subseteq\mathbb{C}^{2n} be a broad and free variety with no CjC_{j}-factors and satisfying (EC). Then for every finitely generated field KK containing the field of definition of VV, there is a point in VV of the form (z1,,zn,j(z1),,j(zn))(z_{1},\ldots,z_{n},j(z_{1}),\ldots,j(z_{n})), with (z1,,zn)n(z_{1},\ldots,z_{n})\in\mathbb{H}^{n}, which is generic over KK.

We will also prove a version of Theorem 1.2 which includes the derivatives of jj, see Theorem 6.5.

The EC problem for jj is still open, but certain partial results have been achieved (we already mentioned [EH21, Theorem 1.1]). One of the main results of [AK22] proves an approximate solution to the EC problem. Their approach is referred to as a blurring of the jj-function (definition can be found in §4.4). Using this, we will deduce the following in §5.2.

Theorem 1.3.

Let V2nV\subseteq\mathbb{C}^{2n} be a broad and free variety with no CjC_{j}-factors. Then for every finitely generated field KK\subset\mathbb{C} containing the field of definition of VV, there are matrices g1,,gng_{1},\ldots,g_{n} in GL2+()\mathrm{GL}_{2}^{+}(\mathbb{Q}) such that VV has a point of the form

(z1,,zn,j(g1z1),,j(gnzn)),(z_{1},\ldots,z_{n},j(g_{1}z_{1}),\ldots,j(g_{n}z_{n})),

with (z1,,zn)n(z_{1},\ldots,z_{n})\in\mathbb{H}^{n}, which is generic over KK.

1.1. Summary of the proof of Theorem 1.1

In order to reduce Strong EC to EC under MSCD and MZP, there are two main technical steps.

  1. (a)

    MSCD gives a lower bound for transcendence degree, but this inequality is only measured over \mathbb{Q}, whereas strong EC requires one to measure transcendence degree over an arbitrary finitely generated field, that is, we need a modular Schanuel-inequality with parameters. This is resolved by using the results on the existence of “convenient generators” of [AEK23, §5], which in turn is based upon the results of [AEK21] on differential existential closedness. See §4.2 for details.

  2. (b)

    Under MSCD, a point in VV of the form (z1,,zn,j(z1),,j(zn))(z_{1},\ldots,z_{n},j(z_{1}),\ldots,j(z_{n})) which is not generic in VV will produce what is known as an atypical intersection. MZP speaks precisely about such atypical components, giving us sufficient control over their behaviour.

1.2. Strong EC for exp\exp

The motivation for the EC and strong EC problems for jj originates in the study of analogous problems stated for the complex exponential function exp:×\exp:\mathbb{C}\to\mathbb{C}^{\times}, which where first considered in Zilber’s work on pseudo-exponentiation [Zil05], with further details and results in [KZ14] and [BK18]. Zilber’s work gives a model-theoretic approach to the study of the algebraic properties of the complex exponential function, and his ideas have since been expanded to many other settings.

With this in mind, the motivation for Theorem 1.1 is not simply a restating of Zilber’s conjectures on exp\exp for the case of jj, but our aim is also to give a general strategy for reducing strong EC problems to EC problems, even though there is no nice model-theoretic “pseudo-jj” structure like pseudo-exponentiation (yet). The methods presented here are expected to work in more general situations, such as in the case of Shimura varieties for which some cases of (EC) have been shown in [EZ21].

The original name of the EC problem in the case of exp\exp was Exponential Algebraic Closedness (EAC). Instead of requiring the variety VV to be broad and free, EAC requires the varieties Vn×(×)nV\subseteq\mathbb{C}^{n}\times\left(\mathbb{C}^{\times}\right)^{n} to be rotund, and additively and multiplicatively free, see [KZ14, §3.7] for definitions. The original name for the Zilber–Pink conjecture in the context of the exponential function was the conjecture on the intersection with tori (CIT).

The result below is a direct analogue of Theorem 1.1 for exp\exp, and we reference [KZ14, Theorem 1.5] as a source. We point out however that the original formulation of [KZ14, Theorem 1.5] does not require the variety VV to be additively and multiplicatively free, although these conditions are necessary for the theorem to hold. Without them, we cannot even expect the variety to intersect the graph of exp\exp. For example, if we define a variety V2×(×)2V\subset\mathbb{C}^{2}\times\left(\mathbb{C}^{\times}\right)^{2} by the following two equations: X1X2=0X_{1}-X_{2}=0 on 2\mathbb{C}^{2} (which prevents VV from being additively free) and Y1Y2=1Y_{1}-Y_{2}=1 on (×)2\left(\mathbb{C}^{\times}\right)^{2}, then VV cannot have point in the graph of any function. Not only that, if we slightly modify VV to be defined by the equations X1X2=0X_{1}-X_{2}=0 and X1=Y1X_{1}=Y_{1}, it can be checked that VV has an infinite intersection with the graph of exp\exp (one for every fixed point of exp\exp), but every point in this intersection has transcendence degree at most 1, so they are not generic in VV.

Theorem 1.4 ([KZ14, Theorem 1.5]).

Let KK\subseteq\mathbb{C} be a finitely generated field, and let Vn×(×)nV\subseteq\mathbb{C}^{n}\times\left(\mathbb{C}^{\times}\right)^{n} be an algebraic variety which is rotund, additively free, multiplicatively free, defined over KK, and satisfying (EAC). Then Schanuel’s conjecture and CIT imply that VV has a point of the form (𝐳,exp(𝐳))(\mathbf{z},\exp(\mathbf{z})) which is generic over KK.

Since our methods for proving Theorem 1.1 are expected to easily generalise to give a proof of Theorem 1.4, and many aspects of the such a proof can already be found in [BK18] and [KZ14], we will not present a proof of this result here. We remark that Theorem 1.4 only considers complex algebraic varieties, not a general model of the first-order theory of pseudo-exponentiation, and as such it falls short of the full ambition of [KZ14, Theorem 1.5].

On the other hand, we expect that by imposing conditions on the base field of VV (analogous to our notion of “having no CjC_{j}-factors”) one can proceed like in the proof of Theorem 1.2 to remove the dependence on Schanuel’s conjecture and CIT from Theorem 1.4. Although not phrased in this way, many aspects of such a result can be recovered from [BK18, Proposition 11.5], with some extra details provided by [AEK23, Theorem 5.6].

Naturally, given the similarities between the results for exp\exp and jj one should ask: when can one expect to obtain analogous results with other functions? The key ingredient in obtaining the existence of the “convenient generators” mentioned in §1.1 is the Ax–Schanuel theorem. As we will see, this theorem is also essential for obtaining uniform weak forms of Zilber–Pink, which are then used in proving Theorem 1.2. We therefore expect that the methods used here can be extended to other situations where one has such a result. Ax–Schanuel theorems have been been obtained in many situations: the exponential map of a (semi-) abelian variety in [Ax72] and [Kir09], uniformizers of any Fuchsian group of the first kind in [BSCFN21] (and more), the uniformisation map of a Shimura variety in [MPT19], variations of mixed Hodge structures in [GK23, Chi21], among many others.

1.3. Structure of the paper

  1. §2:

    We set up the basic notation and give some background on the algebraic aspects of the modular jj function.

  2. §3:

    We go over the technical details of broadness, freeness, the (EC) condition, and the uniform and weak versions of Zilber–Pink.

  3. §4:

    We first lay down the necessary groundwork for handling transcendence inequalities over different fields (using the convenient generators alluded to earlier), we then deal with the possible presence of special solutions (using uniform André–Oort), and we prove Theorem 1.1. After that, we give a few cases of Theorem 1.1 were we can remove the dependence on MZP, without imposing conditions on the field of definition of VV. We also prove Theorem 4.15 in §4.4, which is a version of Theorem 1.1 for a blurring of the jj function.

  4. §5:

    We prove Theorem 1.2. This proof, combined with the technicalities explained in the proof of Theorem 4.15, produce Theorem 1.3 in §5.2.

  5. §6:

    We give an analogue of Theorem 1.2 including the derivatives of jj.

2. Background

2.1. Basic notation

  • If LL is any subfield of \mathbb{C}, then L¯\overline{L} denotes the algebraic closure of LL in \mathbb{C}.

  • Given sets A,BA,B we define AB:={aA:aB}A\setminus B:=\{a\in A:a\not\in B\}.

  • Tuples of elements will be denoted with boldface letters; that is, if x1,,xmx_{1},\ldots,x_{m} are elements of a set XX, then we write 𝐱:=(x1,,xm)\mathbf{x}:=(x_{1},\ldots,x_{m}) for the ordered tuple. We will also sometimes use 𝐱\mathbf{x} to denote the (unordered) set {x1,,xm}\left\{x_{1},\ldots,x_{m}\right\}, which should not lead to confusion.

  • Suppose XX is a non-empty subset of m\mathbb{C}^{m}. If ff denotes a function defined on XX and 𝐱\mathbf{x} is an element of XX, then we write f(𝐱)f(\mathbf{x}) to mean (f(x1),,f(xm))(f(x_{1}),\ldots,f(x_{m})).

  • Given a subfield LL of \mathbb{C} and a subset AnA\subseteq\mathbb{C}^{n}, we say that a subset ZnZ\subseteq\mathbb{C}^{n} is the LL-Zariski closure of AA if ZZ is the smallest Zariski closed set containing AA that is defined over LL.

  • The term algebraic variety for us will just mean a Zariski closed set, not necessarily irreducible. We also identify complex algebraic varieties VV with the set of their \mathbb{C}-points V()V(\mathbb{C}).

2.2. The jj-function

We denote by \mathbb{H} the complex upper-half plane {z:Im(z)>0}\{z\in\mathbb{C}:\mathrm{Im}(z)>0\}. The group GL2+()\mathrm{GL}_{2}^{+}(\mathbb{R}) of 22 by 22 matrices with coefficients in \mathbb{R} and positive determinant, acts on \mathbb{H} via the formula

gz:=az+bcz+d for g=(abcd).gz:=\frac{az+b}{cz+d}\ \text{ for }\ g=\left(\begin{array}[]{cc}a&b\\ c&d\end{array}\right).

This action can be extended to a continuous action of GL2+()\mathrm{GL}_{2}^{+}(\mathbb{R}) on the Riemann sphere ^:={}\widehat{\mathbb{C}}:=\mathbb{C}\cup\{\infty\}. Given a subring RR of \mathbb{R} we define M2+(R)M_{2}^{+}(R) as the set of 22 by 22 matrices with coefficients in RR and positive determinant. We put

G:=GL2+()=M2+(),G:=\mathrm{GL}_{2}^{+}(\mathbb{Q})=M_{2}^{+}(\mathbb{Q}),

which is a subgroup of GL2+()\mathrm{GL}_{2}^{+}(\mathbb{R}). The modular group is defined as

Γ:=SL2()={gM2+():det(g)=1}.\Gamma:=\mathrm{SL}_{2}(\mathbb{Z})=\{g\in M_{2}^{+}(\mathbb{Z}):\det(g)=1\}.

The modular jj function is defined as the unique holomorphic function j:j:\mathbb{H}\to\mathbb{C} that satisfies

j(gz)=j(z) for every g in Γ and every z in ,j(gz)=j(z)\text{ for every }g\text{ in }\Gamma\text{ and every }z\text{ in }\mathbb{H},

and has a Fourier expansion of the form

(2.1) j(z)=q1+744+k=1akqk with q:=exp(2πiz) and ak.j(z)=q^{-1}+744+\sum_{k=1}^{\infty}a_{k}q^{k}\text{ with }q:=\exp(2\pi iz)\text{ and }a_{k}\in\mathbb{C}.

This function allows us to identify Γ\\Gamma\backslash\mathbb{H}\simeq\mathbb{C}. The quotient space Y(1):=Γ\Y(1):=\Gamma\backslash\mathbb{H} is known to be a (coarse) moduli space for one-dimensional complex tori, or equivalently, elliptic curves over \mathbb{C}. If Γz\Gamma z is a point in Y(1)Y(1) and EzE_{z} denotes an elliptic curve in the corresponding isomorphism class, then j(z)j(z) is simply the jj-invariant of the curve EzE_{z}.

It is well-known ([Mah69]) that jj satisfies the following algebraic differential equation (and none of lower order):

(2.2) 0=j′′′j32(j′′j)2+j21968j+2654208j2(j1728)2(j)2.0=\frac{j^{\prime\prime\prime}}{j^{\prime}}-\frac{3}{2}\left(\frac{j^{\prime\prime}}{j^{\prime}}\right)^{2}+\frac{j^{2}-1968j+2654208}{j^{2}(j-1728)^{2}}\left(j^{\prime}\right)^{2}.

2.3. Modular polynomials

Let {ΦN(X,Y)}N=1[X,Y]\left\{\Phi_{N}(X,Y)\right\}_{N=1}^{\infty}\subseteq\mathbb{Z}[X,Y] denote the family of modular polynomials associated with jj (see [Lan87, Chapter 5, Section 2] for the definition and main properties of this family). We recall that ΦN(X,Y)\Phi_{N}(X,Y) is irreducible in [X,Y]\mathbb{C}[X,Y], Φ1(X,Y)=XY\Phi_{1}(X,Y)=X-Y, and for N2N\geq 2, ΦN(X,Y)\Phi_{N}(X,Y) is symmetric of total degree 2N\geq 2N. Also, the action of GG on \mathbb{H} can be traced by using modular polynomials in the following way: for every gg in GG we define g~\widetilde{g} as the unique matrix of the form rgrg with rr\in\mathbb{Q} and r>0r>0, so that the entries of g~\widetilde{g} are all integers and relatively prime. Then, for every xx and yy in \mathbb{H} the following statements are equivalent:

  • (M1):

    ΦN(j(x),j(y))=0\Phi_{N}(j(x),j(y))=0;

  • (M2):

    There exists gg in GG with gx=ygx=y and det(g~)=N\det\left(\widetilde{g}\right)=N.

Definition.

A finite set AA\subset\mathbb{C} is said to be modularly independent if for every pair of distinct numbers a,ba,b in AA and every positive integer NN, we have that ΦN(a,b)0\Phi_{N}(a,b)\neq 0. Otherwise, we say that AA is modularly dependent.

An element ww is said to be modularly dependent over AA if there is aAa\in A such that the set {a,w}\{a,w\} is modularly dependent.

Definition.

Given a subset AA\subseteq\mathbb{C}, the Hecke orbit of AA is defined as

He(A):={z:aAN(ΦN(z,a)=0)}.\mathrm{He}(A):=\left\{z\in\mathbb{C}:\exists a\in A\exists N\in\mathbb{N}(\Phi_{N}(z,a)=0)\right\}.
Remark 2.1.

Let dd be a positive integer. As explained in [EH21, §7.3], combining isogeny estimates of Masser–Wüstholz and Pellarin, with gonality estimates for modular curves, we obtain that for every zz\in\mathbb{C} we have that the set

Hed(z):={wHe(z):[(z,w):(z)]d}\mathrm{He}_{d}(z):=\{w\in\mathrm{He}(z):[\mathbb{Q}(z,w):\mathbb{Q}(z)]\leq d\}

is finite. This immediately gives that if AA\subset\mathbb{C} is a finite set, then

Hed(A):={wHe(A):[(A,w):(A)]d}\mathrm{He}_{d}(A):=\{w\in\mathrm{He}(A):[\mathbb{Q}(A,w):\mathbb{Q}(A)]\leq d\}

is also finite.

2.4. Special points

A point ww in \mathbb{C} is said to be special (also known as a singular modulus) if there is zz in \mathbb{H} such that [(z):]=2[\mathbb{Q}(z):\mathbb{Q}]=2 and j(z)=wj(z)=w. Under the moduli interpretation of jj, special points are those that correspond to elliptic curves endowed with complex multiplication. We set

Σ:={z:[(z):]=2}.\Sigma:=\left\{z\in\mathbb{H}:[\mathbb{Q}(z):\mathbb{Q}]=2\right\}.

A theorem of Schneider [Sch37] says that tr.deg.(z,j(z))=0\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(z,j(z))=0 if and only if zz is in Σ\Sigma. We say that a point 𝐰\mathbf{w} in n\mathbb{C}^{n} is special if every coordinate of 𝐰\mathbf{w} is special.

2.5. The modular Schanuel conjecture

Given a subset AA of \mathbb{H}, we define dimG(A)\dim_{G}(A) as the number of distinct GG-orbits in

GA={ga:gG,aA}G\cdot A=\{ga:g\in G,a\in A\}

(which may be infinite). Equivalently, dimG(A)\dim_{G}(A) is the cardinality of the quotient set G\(GA)G\backslash(G\cdot A). Given another subset CC of \mathbb{C}, we define dimG(A|C)\dim_{G}(A|C) as the number of distinct GG-orbits in (GA)(GC)(G\cdot A)\setminus(G\cdot C). In plain words, dimG(A|C)\dim_{G}(A|C) counts the number of orbits generated by elements of AA that do not contain elements of CC.

We now present two modular versions of Schanuel’s conjecture. Both follow from the Grothendieck–André generalised period conjecture applied to powers of the modular curve (see [AEK23, §6.3] and references therein).

Conjecture 2.2 (Modular Schanuel conjecture (MSC)).

For every z1,,znz_{1},\ldots,z_{n} in \mathbb{H}:

tr.deg.(𝐳,j(𝐳))dimG(𝐳|Σ).\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}\left(\mathbf{z},j(\mathbf{z})\right)\geq\dim_{G}(\mathbf{z}|\Sigma).
Conjecture 2.3 (Modular Schanuel conjecture with derivatives (MSCD)).

For every z1,,znz_{1},\ldots,z_{n} in \mathbb{H}:

tr.deg.(𝐳,j(𝐳),j(𝐳),j′′(𝐳))3dimG(𝐳|Σ).\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}\left(\mathbf{z},j(\mathbf{z}),j^{\prime}(\mathbf{z}),j^{\prime\prime}(\mathbf{z})\right)\geq 3\dim_{G}(\mathbf{z}|\Sigma).

Clearly MSC follows from MSCD. Also, in view of the differential equation 2.2, MSCD only needs to involve up to the second derivative of jj. We also point out that MSCD is somewhat incomplete as it does not say anything about the transcendence degree of (z,j(z),j(z),j′′(z))\mathbb{Q}(z,j(z),j^{\prime}(z),j^{\prime\prime}(z)) over \mathbb{Q} when zΣz\in\Sigma. In fact, a stronger conjecture is implied by the generalised period conjecture (see [AEK23, Conjecture 6.13]) which does account for points in Σ\Sigma. However, in the proofs of our main results we will show that we can move away from points that have coordinates in Σ\Sigma, and for this reason, MSCD will suffice for us.

Remark 2.4.

Using the equivalence between (M1) and (M2) in §2.3, it is easy to see that MSC can be restated equivalently using a more geometric language (see also [Pil15, §15]): For every algebraic variety V2nV\subseteq\mathbb{C}^{2n} defined over ¯\overline{\mathbb{Q}} for which there exists (𝐳,j(𝐳))V(\mathbf{z},j(\mathbf{z}))\in V, if dimV<n\dim V<n then dimG(𝐳|Σ)<n\dim_{G}(\mathbf{z}|\Sigma)<n.

This way of thinking about MSC leads one naturally to a counterpart problem for MSC, that is, the question of determining which subvarieties V2nV\subseteq\mathbb{C}^{2n} intersect the graph of the jj-function. This gives rise to the existential closedness problem, which we discuss next.

3. Existential Closedness and the Zilber–Pink Conjecture

In this section we define the notions of broad and free, we define the existential closedness conjecture for the jj function, we recall the modular Zilber–Pink conjecture, and we give some partial results.

To do all this, we need to distinguish a family of subvarieties that will be called special. For this, it will be important to differentiate the first nn coordinates of 2n\mathbb{C}^{2n} from the last nn coordinates because we want to think of 2n\mathbb{C}^{2n} as a product n×n\mathbb{C}^{n}\times\mathbb{C}^{n} where the left factor has special subvarieties coming from the action of GG, whereas on the second factor the special subvarieties are defined by modular polynomials. So as to avoid confusion, we prefer to change the name of the second factor to Y(1)n\mathrm{Y}(1)^{n}, as in this case we are thinking of \mathbb{C} as the modular curve Y(1)\mathrm{Y}(1). This way, we view 2n\mathbb{C}^{2n} as n×Y(1)n\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}. This is purely done for purposes of notation.

We will always think of the subvarieties of n×Y(1)n\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} as being defined over the ring of polynomials [X1,,Xn,Y1,,Yn]\mathbb{C}[X_{1},\ldots,X_{n},Y_{1},\ldots,Y_{n}].

We will denote by π:n×Y(1)nn\pi_{\mathbb{C}}:\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}\to\mathbb{C}^{n} and πY:n×Y(1)nY(1)n\pi_{\mathrm{Y}}:\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}\to\mathrm{Y}(1)^{n} the corresponding projections.

3.1. Broad and free varieties

Given a matrix g=(abcd)g=\left(\begin{array}[]{cc}a&b\\ c&d\end{array}\right) in GL2()\mathrm{GL}_{2}(\mathbb{C}), we define the polynomial Mg(X,Y):=Y(cX+d)(aX+b)M_{g}(X,Y):=Y(cX+d)-(aX+b) and the rational function gX:=aX+bcX+dgX:=\frac{aX+b}{cX+d}. Note that Mg(X,gX)=0M_{g}(X,gX)=0.

Definition.

A Möbius subvariety of n\mathbb{C}^{n} is a variety defined by finitely many equations of the form Mgi,k(Xi,Xk)M_{g_{i,k}}(X_{i},X_{k}) with gi,kg_{i,k} in GL2()\mathrm{GL}_{2}(\mathbb{C}) non-scalar and i,ki,k in {1,,n}\{1,\ldots,n\} not necessarily different.

Example 3.1.
111We thanks Sebastián Herrero for providing this example.

The variety V={(x1,x2,x3)3:x12x2+3x3=0}V=\{(x_{1},x_{2},x_{3})\in\mathbb{C}^{3}:x_{1}-2x_{2}+3x_{3}=0\} is not a Möbius subvariety of 3\mathbb{C}^{3} but it contains infinitely many Möbius subvarieties. Indeed, for every integer mm define gm=(3m1001)g_{m}=\left(\begin{array}[]{cc}3m-1&0\\ 0&1\end{array}\right) and hm=(2m1001)h_{m}=\left(\begin{array}[]{cc}2m-1&0\\ 0&1\end{array}\right). Then VV contains Mgm(X1,X2)Mhm(X1,X3)M_{g_{m}}(X_{1},X_{2})\cap M_{h_{m}}(X_{1},X_{3}), which is a Möbius subvariety when m2m\geq 2.

Definition.

A special subvariety of Y(1)n\mathrm{Y}(1)^{n} is an irreducible component of an algebraic set defined by equations of the following forms:

  1. (a)

    ΦN(Yi,Yk)=0\Phi_{N}(Y_{i},Y_{k})=0, for some NN\in\mathbb{N}, and

  2. (b)

    Yi=τY_{i}=\tau, where τ\tau\in\mathbb{C} is a special point.

We allow the set of equations to be empty, so Y(1)n\mathrm{Y}(1)^{n} is itself a special variety. A special subvariety without constant coordinates is called a basic special subvariety.

Since every special point can be obtained as the solutions of ΦN(X,X)=0\Phi_{N}(X,X)=0 for some NN\in\mathbb{N}, the condition Yi=τY_{i}=\tau in the previous definition follows from the condition ΦN(Yi,Yk)=0\Phi_{N}(Y_{i},Y_{k})=0 by choosing i=ki=k, but we have opted for this presentation as it makes it easy to compare it with the definition of weakly special variety in the following paragraph.

A subvariety of Y(1)n\mathrm{Y}(1)^{n} is called weakly special if it is an irreducible component of an algebraic set defined by equations of the following forms:

  1. (i)

    ΦN(Yi,Yk)=0\Phi_{N}(Y_{i},Y_{k})=0, for some NN\in\mathbb{N},

  2. (ii)

    Y=dY_{\ell}=d, for some constant dd\in\mathbb{C}.

As it turns out, every special variety has a Zariski dense set of special points (see e.g. [Pil11, 1.4 Aside]), and so a weakly special subvariety is special if and only if it contains a special point.

If SS is a special subvariety of Y(1)n\mathrm{Y}(1)^{n}, then a (weakly) special subvariety of SS is a (weakly) special subvariety of Y(1)n\mathrm{Y}(1)^{n} that is contained in SS.

If TT is a proper positive dimensional weakly special subvariety of Y(1)n\mathrm{Y}(1)^{n}, there are m{1,,n}m\in\{1,\ldots,n\}, a basic special subvariety SY(1)mS\subseteq\mathrm{Y}(1)^{m}, and a point pY(1)nmp\in\mathrm{Y}(1)^{n-m} such that (up to re-indexing of the coordinates) TT can be written as S×{p}S\times\{p\}. We define the basic complexity of TT to be the maximal positive integer NN for which the modular polynomial ΦN\Phi_{N} is required to define SS. We denote this number by Δb(T)\Delta_{b}(T) (which represents the complexity of the basic special part of T)T). This definition differs slightly from [HP16, Definition 3.8], where the complexity is only defined for special varieties, and the complexity also depends on the constant special coordinates that the variety may have.

We remark that if SS and TT are (weakly) special subvarieties of Y(1)n\mathrm{Y}(1)^{n}, then the irreducible components of STS\cap T (if there are any) are again (weakly) special suvbarieties of Y(1)n\mathrm{Y}(1)^{n}. In fact, by Hilbert’s basis theorem the irreducible components of any non-empty intersection of (weakly) special subvarieties is again a (weakly) special subvariety.

Definition.

Given an irreducible constructible set XX, we denote by spcl(X)\mathrm{spcl}(X) the special closure of XX, that is, the smallest special subvariety containing XX. Similarly, we denote by wspcl(X)\mathrm{wspcl}(X) the weakly special closure of XX.

Following [Asl21], we also make the following definition.

Definition.

Given a tuple 𝐜=(c1,,cm)\mathbf{c}=(c_{1},\ldots,c_{m}) of complex numbers, we say that a subvariety SS of Y(1)n\mathrm{Y}(1)^{n} is 𝐜\mathbf{c}-special if SS is weakly special and the values of each of the constant coordinates of SS (if there are any) is either an element of He(𝐜)\mathrm{He}(\mathbf{c}), or a special point.

Definition.

We will say that an irreducible constructible set Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} is modularly free if πY(V)\pi_{\mathrm{Y}}(V) is not contained in a proper special subvariety of Y(1)n\mathrm{Y}(1)^{n}.

We will also say that VV is free if VV is modularly free, no coordinate of VV is constant, and π(V)\pi_{\mathbb{C}}(V) is not contained in any Möbius subvariety of n\mathbb{C}^{n} which is only defined by elements of GG.

We say that a constructible subset Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} is free if every irreducible component of VV is free.

Now we introduce some notation for coordinate projections. Let nn and \ell denote positive integers with n\ell\leq n, and let 𝐢=(i1,,i)\mathbf{i}=(i_{1},\ldots,i_{\ell}) denote a point in \mathbb{N}^{\ell} with 1i1<<in1\leq i_{1}<\ldots<i_{\ell}\leq n. Define the projection map pr𝐢:n\mathrm{pr}_{\mathbf{i}}:\mathbb{C}^{n}\rightarrow\mathbb{C}^{\ell} by

pr𝐢:(x1,,xn)(xi1,,xi).\mathrm{pr}_{\mathbf{i}}:(x_{1},\ldots,x_{n})\mapsto(x_{i_{1}},\ldots,x_{i_{\ell}}).

In particular we distinguish between the natural number nn and the tuple 𝐧=(1,,n)\mathbf{n}=(1,\ldots,n). We also use the notation 𝐧𝐢\mathbf{n}\setminus\mathbf{i} to denote the tuple of entries of 𝐧\mathbf{n} that do not appear in 𝐢\mathbf{i}.

Remark 3.2.

If TT is a (weakly) special subvariety of Y(1)n\mathrm{Y}(1)^{n}, then for any choice of indices 1i1<<in1\leq i_{1}<\cdots<i_{\ell}\leq n we have that pr𝐢(T)\mathrm{pr}_{\mathbf{i}}(T) is a (weakly) special subvariety of Y(1)\mathrm{Y}(1)^{\ell}.

Define Pr𝐢:n×Y(1)n×Y(1)\mathrm{Pr}_{\mathbf{i}}:\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}\rightarrow\mathbb{C}^{\ell}\times\mathrm{Y}(1)^{\ell} by

Pr𝐢(𝐱,𝐲):=(pr𝐢(𝐱),pr𝐢(𝐲)).\mathrm{Pr}_{\mathbf{i}}(\mathbf{x},\mathbf{y}):=(\mathrm{pr}_{\mathbf{i}}(\mathbf{x}),\mathrm{pr}_{\mathbf{i}}(\mathbf{y})).
Definition.

An algebraic set Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} is said to be broad if for any 𝐢=(i1,,i)\mathbf{i}=(i_{1},\ldots,i_{\ell}) in \mathbb{N}^{\ell} with 1i1<<in1\leq i_{1}<\ldots<i_{\ell}\leq n we have dimPr𝐢(V)\dim\mathrm{Pr}_{\mathbf{i}}(V)\geq\ell. In particular, if VV is broad then dimVn\dim V\geq n.

We say VV is strongly broad if the strict inequality dimPr𝐢(V)>\dim\mathrm{Pr}_{\mathbf{i}}(V)>\ell holds for every 𝐢\mathbf{i}.

For example, if π(V)\pi_{\mathbb{C}}(V) is Zariski dense in n\mathbb{C}^{n}, then VV is broad.

3.2. Existential Closedness

Here we will define the (EC) condition, give a conjecture for the EC problem for jj (see Conjecture 3.5), and review some known results.

Given an integer n1n\geq 1 we denote the graph of j:nnj:\mathbb{H}^{n}\rightarrow\mathbb{C}^{n} as

Ejn:={(z1,,zn,j(z1),,j(zn)):z1,,zn}.\mathrm{E}_{j}^{n}:=\{(z_{1},\ldots,z_{n},j(z_{1}),\ldots,j(z_{n})):z_{1},\ldots,z_{n}\in\mathbb{H}\}.
Definition.

We say that an algebraic variety Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} satisfies the Existential Closedness condition (EC) if the set VEjnV\cap\mathrm{E}_{j}^{n} is Zariski dense in VV.

We recall the following result which gives examples of varieties satisfying (EC).

Theorem 3.3 ([EH21, Theorem 1.1]).

Let Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be an irreducible algebraic variety. If π(V)\pi_{\mathbb{C}}(V) is Zariski dense in n\mathbb{C}^{n}, then VV satisfies (EC).

We also recall the following statement (although it falls short from giving examples of varieties satisfying (EC) as it does not prove Zariski density).

Theorem 3.4 ([Gal21, Theorem 3.31]).

Let LnL\subseteq\mathbb{C}^{n} be a subvariety defined by equations of the form Mg(X,Y)=0M_{g}(X,Y)=0, with gGL2()g\in\mathrm{GL}_{2}(\mathbb{R}). Let WY(1)nW\subseteq\mathrm{Y}(1)^{n} be a subvariety, and assume that L×WL\times W is a free and broad subvariety of n×Y(1)n\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}. Then j(L)j(L) is Euclidean dense in WW.

These theorems and the main results of [AEK21] give evidence for the following conjecture (cf [AK22, Conjecture 1.2]).

Conjecture 3.5.

Let nn be a positive integer and let Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be a broad and free variety such that V(n×Y(1)n)V\cap\left(\mathbb{H}^{n}\times\mathrm{Y}(1)^{n}\right) is Zariski dense in VV. Then VEjnV\cap\mathrm{E}_{j}^{n}\neq\emptyset.

The reader may be wondering why, in light of Theorem 3.3, Conjecture 3.5 does not ask for the stronger condition of VEjnV\cap\mathrm{E}_{j}^{n} being Zariski dense in VV. In fact, as we will see in Lemma 3.8, Conjecture 3.5 already implies the this stronger condition.

Conjecture 3.5 can be thought of as a version of Hilbert’s Nullstellensatz for systems of algebraic equations involving the jj-function (cf [DMT10] for the discussion in the case of the exponential function). The examples of [EH21, §3] show that if VV has constant coordinates or is contained in a proper subvariety of the form M×Y(1)nM\times\mathrm{Y}(1)^{n}, with MM a Möbius subvariety, then it can happen that VEjnV\cap\mathrm{E}_{j}^{n} is empty.

We also point out that the condition of V(n×Y(1)n)V\cap\left(\mathbb{H}^{n}\times\mathrm{Y}(1)^{n}\right) being Zariski dense in VV is needed, as otherwise the subvariety of 2×Y(1)2\mathbb{C}^{2}\times\mathrm{Y}(1)^{2} defined by the single equation X1=iX2X_{1}=iX_{2} is free and broad, but cannot have points in the graph of the jj function since there is no point zz in \mathbb{H} for which iziz is also in \mathbb{H}. One could alternatively resolve this issue by extending the domain of jj to the lower half-plane using Schwarz reflection.

Remark 3.6.

Let Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be a broad and free variety. If dimV=n\dim V=n, then by [PT16, Theorem 1.1] if UU is a component of VEjnV\cap\mathrm{E}_{j}^{n} (assuming there are any), then UU has dimension zero unless πY(U)T\pi_{\mathrm{Y}}(U)\subseteq T, for some proper special subvariety TY(1)nT\subset\mathrm{Y}(1)^{n}. So for a generic VV, we can only expect the intersection VEjnV\cap\mathrm{E}_{j}^{n} to be at most countable.

The following result is an approximate version of Conjecture 3.5 using what is known as a blurring of the jj function.

Theorem 3.7 ([AK22, Theorem 1.9]).

Let Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be an irreducible variety which is broad and free. Then the set of points in VV of the form

(z1,,zn,j(g1z1),,j(gnzn)),(z_{1},\ldots,z_{n},j(g_{1}z_{1}),\ldots,j(g_{n}z_{n})),

such that z1,,znz_{1},\ldots,z_{n}\in\mathbb{H} and g1,,gnGg_{1},\ldots,g_{n}\in G, is Euclidean dense in V(n×Y(1)n)V\cap\left(\mathbb{H}^{n}\times\mathrm{Y}(1)^{n}\right).

We finish this subsection by recalling a useful trick.

Lemma 3.8 (cf [Asl22a, Proposition 4.34]).

Suppose Conjecture 3.5 holds. Let Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be a broad and free irreducible variety. Then VV satisfies (EC).

Proof.

Let p1,,pmp_{1},\ldots,p_{m} be polynomials in [X1,,Xn,Y1,,Yn]\mathbb{C}[X_{1},\ldots,X_{n},Y_{1},\ldots,Y_{n}] defining VV. Let ff be a polynomial in [X1,,Xn,Y1,,Yn]\mathbb{C}[X_{1},\ldots,X_{n},Y_{1},\ldots,Y_{n}] that does not vanish identically on VV and let

W={(𝐱,𝐲)V:f(𝐱,𝐲)=0}.W=\left\{(\mathbf{x},\mathbf{y})\in V:f(\mathbf{x},\mathbf{y})=0\right\}.

To prove the lemma it suffices to show that Ejn(VW)\mathrm{E}_{j}^{n}\cap(V\setminus W)\neq\emptyset. By Conjecture 3.5 this is clear if W=W=\emptyset, hence we can assume that ff vanishes at least at one point of VV. We now define the following subvariety222This is the standard Rabinowitsch trick used in the proof of Nullstellensatz. of n+1×Y(1)n+1\mathbb{C}^{n+1}\times\mathrm{Y}(1)^{n+1}:

V:={p1(x1,,xn,y1,,yn)=0pm(x1,,xn,y1,,yn)=0yn+1f(x1,,xn,y1,,yn)1=0}.V^{\prime}:=\left\{\begin{array}[]{ccc}p_{1}(x_{1},\ldots,x_{n},y_{1},\ldots,y_{n})&=&0\\ &\vdots&\\ p_{m}(x_{1},\ldots,x_{n},y_{1},\ldots,y_{n})&=&0\\ y_{n+1}f(x_{1},\ldots,x_{n},y_{1},\ldots,y_{n})-1&=&0\end{array}\right\}.

Choose 1i1<<in+11\leq i_{1}<\cdots<i_{\ell}\leq n+1 and set 𝐢=(i1,,i)\mathbf{i}=(i_{1},\ldots,i_{\ell}). Then

dimPr𝐢(V)={dimPr𝐢(VW) if in+1,dimPr(i1,,i1)(VW)+1 if 2 and i=n+1,2 if =1 and i1=n+1.\dim\mathrm{Pr}_{\mathbf{i}}(V^{\prime})=\left\{\begin{array}[]{ll}\dim\mathrm{Pr}_{\mathbf{i}}(V\setminus W)&\text{ if }i_{\ell}\neq n+1,\\ \dim\mathrm{Pr}_{(i_{1},\ldots,i_{\ell-1})}(V\setminus W)+1&\text{ if }\ell\geq 2\text{ and }i_{\ell}=n+1,\\ 2&\text{ if }\ell=1\text{ and }i_{1}=n+1.\end{array}\right.

Now, since VV is irreducible, the set VWV\setminus W is dense in VV. Hence, for any subtuple 𝐤\mathbf{k} of 𝐧\mathbf{n} we have, by continuity of Pr𝐢\mathrm{Pr}_{\mathbf{i}}, that dimPr𝐤(VW)=dimPr𝐤(V)\dim\mathrm{Pr}_{\mathbf{k}}(V\setminus W)=\dim\mathrm{Pr}_{\mathbf{k}}(V). Since VV is broad, we conclude that VV^{\prime} is also broad.

We will now prove that VV^{\prime} is free. Since VV has no constant coordinates, and ff is non-constant, we see that no coordinate is constant on VV^{\prime}. Also, since VV is free, it is clear that VV^{\prime} is not contained in a variety of the form M×Y(1)n+1M\times\mathrm{Y}(1)^{n+1} where MM is a proper Möbius subvariety of n+1\mathbb{C}^{n+1}. Moreover, if VV is contained in a variety of the form n+1×T\mathbb{C}^{n+1}\times T where TT is a proper special subvariety of Y(1)n+1\mathrm{Y}(1)^{n+1}, then at least one of the polynomials defining TT must be of the from ΦN(yi,yn+1)\Phi_{N}(y_{i},y_{n+1}) with i{1,,n}i\in\{1,\ldots,n\} and N1N\geq 1. This implies that ΦN(yi,1/f)=0\Phi_{N}(y_{i},1/f)=0 on VWV\setminus W, hence VWV\setminus W is contained in the variety ZZ defined by the polynomial fdΦN(yi,1/f)f^{d}\Phi_{N}(y_{i},1/f) where dd is the degree of ΦN\Phi_{N} in the YY variable. This implies that VZV\subseteq Z. Since ΦN(X,Y)\Phi_{N}(X,Y) has leading term ±1\pm 1 as a polynomial in YY, it follows that ff has no zeroes on ZZ. But this implies that ff has no zeroes on VV, which is a contradiction. This proves that VV^{\prime} is free.

Since VV^{\prime} is broad and free, Conjecture 3.5 implies that there exists a point of the form (𝐳,j(𝐳))(\mathbf{z},j(\mathbf{z})) in VV^{\prime} with 𝐳n+1\mathbf{z}\in\mathbb{H}^{n+1}, and so Pr𝐧(𝐳,j(𝐳))(VW)\mathrm{Pr}_{\mathbf{n}}(\mathbf{z},j(\mathbf{z}))\in(V\setminus W). This completes the proof of the lemma. ∎

3.3. Atypical intersections and Zilber–Pink

Definition.

Suppose that VV and WW are subvarieties of a smooth algebraic variety ZZ. Let XX be an irreducible component of the intersection VWV\cap W. We say that XX is an atypical component of VWV\cap W (in ZZ) if

dimX>dimV+dimWdimZ.\dim X>\dim V+\dim W-\dim Z.

We say that the intersection VWV\cap W is atypical (in ZZ) if it has an atypical component. Otherwise, we say that VWV\cap W is typical (in ZZ), i.e. VWV\cap W is typical in ZZ if dimVW=dimV+dimWdimZ\dim V\cap W=\dim V+\dim W-\dim Z.

If Z=Y(1)nZ=\mathrm{Y}(1)^{n}, we say that XX is an atypical component of VV if there exists a special subvariety TT of Y(1)n\mathrm{Y}(1)^{n} such that XX is an atypical component of VTV\cap T. We remark that in this case, since dimTdimspcl(X)\dim T\geq\dim\mathrm{spcl}(X), it is also true that XX is an atypical component of Vspcl(X)V\cap\mathrm{spcl}(X).

We say that XX is a strongly atypical component of VV if XX is an atypical component of VV and no coordinate is constant on XX.

An atypical (resp. strongly atypical) component of VV is said to be maximal (in VV) if it is not properly contained in another atypical (resp. strongly atypical) component of VV.

Given a tuple 𝐜=(c1,,cm)\mathbf{c}=(c_{1},\ldots,c_{m}) of complex numbers, an atypical component XX of VV is said to be 𝐜\mathbf{c}-atypical is XX is an atypical component of the intersection VTV\cap T, where TT is a 𝐜\mathbf{c}-special subvariety.

Example 3.9.

Let TT be a proper special subvariety of Y(1)n\mathrm{Y}(1)^{n}. Then TT is an atypical component of itself, since

dimT=dimTT>dimT+dimTn.\dim T=\dim T\cap T>\dim T+\dim T-n.

On the other hand, although Y(1)n\mathrm{Y}(1)^{n} is a special variety, it is not atypical in itself.

Conjecture 3.10 (Modular Zilber–Pink).

For every positive integer nn, any subvariety of Y(1)n\mathrm{Y}(1)^{n} has only finitely many maximal atypical components.

From now on, we will abbreviate this conjecture as MZP. This conjecture is sometimes presented in terms of optimal varieties, which we discuss next.

Definition.

Let VV be a subvariety of Y(1)n\mathrm{Y}(1)^{n}. Given a subvariety XVX\subseteq V, we define the defect of XX to be

def(X):=dimspcl(X)dimX.\mathrm{def}(X):=\dim\mathrm{spcl}(X)-\dim X.

We say that XX is optimal in VV is for every subvariety WVW\subseteq V satisfying XWX\subsetneq W we have that def(X)<def(W)\mathrm{def}(X)<\mathrm{def}(W). We let Opt(V)\mathrm{Opt}(V) denote the set of all optimal subvarieties of VV. Observe that always VOpt(V)V\in\mathrm{Opt}(V). We think of Opt(V)\mathrm{Opt}(V) as a cycle in Y(1)n\mathrm{Y}(1)^{n}.

Remark 3.11.

A maximal atypical component of VV is optimal in VV. However, optimal subvarieties need not be maximal atypical.

On the other hand, if XX is a proper subvariety of VV which is optimal in VV, then def(X)<def(V)\mathrm{def}(X)<\mathrm{def}(V) which implies that dimVspcl(X)dimX>dimV+dimspcl(X)dimspcl(V)\dim V\cap\mathrm{spcl}(X)\geq\dim X>\dim V+\dim\mathrm{spcl}(X)-\dim\mathrm{spcl}(V), so the intersection Vspcl(X)V\cap\mathrm{spcl}(X) is atypical.

As shown in [HP16, Lemma 2.7], MZP is equivalent to the statement that any subvariety of Y(1)n\mathrm{Y}(1)^{n} contains only finitely many optimal subvarieties, i.e. Opt(V)\mathrm{Opt}(V) is a finite set.

Definition.

Let SS be a constructible set (resp. an algebraic variety) in N\mathbb{C}^{N}, where NN is some positive integer. A parametric family of constructible subsets (resp. subvarieties) of SS is a constructible set VS×QV\subseteq S\times Q, where QmQ\subseteq\mathbb{C}^{m} is another constructible set, which we denote as an indexed collection V=(V𝐪)𝐪QV=(V_{\mathbf{q}})_{\mathbf{q}\in Q}, where for each qQq\in Q the set

V𝐪:={𝐬S:(𝐬,𝐪)V}V_{\mathbf{q}}:=\left\{\mathbf{s}\in S:(\mathbf{s},\mathbf{q})\in V\right\}

is a constructible subset (resp. subvariety) of SS.333In the terminology of model-theory, we can equivalently say that VS×QV\subseteq S\times Q is a definable family of definable subsets of SS, in the language of rings.

Example 3.12.

An important example for us of a parametric family is given by the following construction. Let WW be an algebraic subvariety of n×Y(1)n\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}; we want to define the family of subvarieties of WW obtained by intersecting WW with all Möbius subvareities of n\mathbb{C}^{n} defined by elements of GL2()\mathrm{GL}_{2}(\mathbb{C}).

Given a function f:DGL2()f:D\to\mathrm{GL}_{2}(\mathbb{C}) defined on a non-empty subset DD of {1,,n}×{1,,n}\{1,\ldots,n\}\times\{1,\ldots,n\}, set

Wf:={(x1,,xn,y1,,yn)W:f(i,j)xi=xj for all (i,j)D}.W_{f}:=\{(x_{1},\ldots,x_{n},y_{1},\ldots,y_{n})\in W:f(i,j)x_{i}=x_{j}\text{ for all }(i,j)\in D\}.

Then, the collection of all such WfW_{f} forms a parametic family of subvarieties of WW. Indeed, put

Q:=D{1,,n}2GL2()D.Q:=\bigsqcup_{\emptyset\neq D\subseteq\{1,\ldots,n\}^{2}}\mathrm{GL}_{2}(\mathbb{C})^{D}.

Every function f:DGL2()f:D\to\mathrm{GL}_{2}(\mathbb{C}) can be represented as an element of GL2()D\mathrm{GL}_{2}(\mathbb{C})^{D}. Since for every finite non-empty set AA we have that GL2()A\mathrm{GL}_{2}(\mathbb{C})^{A} is a constructible subset of 4m\mathbb{C}^{4m}, where m=#Am=\#A, we have that QQ is a finite union of constructible sets, hence it is constructible. Then choosing S=WS=W and

V={((𝐱,𝐲),f)W×Q:(𝐱,𝐲)Wf}V=\{((\mathbf{x},\mathbf{y}),f)\in W\times Q:(\mathbf{x},\mathbf{y})\in W_{f}\}

we have that (Wf)f(W_{f})_{f} is just the parametric family (Vf)fQ(V_{f})_{f\in Q} associated to VV.

Pila showed that MZP implies the following uniform version of itself.

Theorem 3.13 (Uniform MZP, see [Pil22, §24.2]).

Suppose that for every positive integer nn, MZP holds for all subvarieties of Y(1)n\mathrm{Y}(1)^{n}. Let (V𝐪)𝐪Q(V_{\mathbf{q}})_{\mathbf{q}\in Q} be a parametric family of subvarieties of Y(1)n\mathrm{Y}(1)^{n}. Then there is a parametric family (W𝐩)𝐩P(W_{\mathbf{p}})_{\mathbf{p}\in P} of closed algebraic subsets of Y(1)n\mathrm{Y}(1)^{n} such that for every 𝐪Q\mathbf{q}\in Q there is 𝐩P\mathbf{p}\in P such that Opt(V𝐪)=W𝐩\mathrm{Opt}(V_{\mathbf{q}})=W_{\mathbf{p}}.

Corollary 3.14.

Let (V𝐪)𝐪Q(V_{\mathbf{q}})_{\mathbf{q}\in Q} be a parametric family of constructible subsets of Y(1)n\mathrm{Y}(1)^{n}. Then MZP implies that there is a finite collection 𝒮\mathscr{S} of proper special subvarieties of Y(1)n\mathrm{Y}(1)^{n} such that for all 𝐪Q\mathbf{q}\in Q and all special subvarieties TT, if XX is an atypical component of V𝐪TV_{\mathbf{q}}\cap T, then there is T0𝒮T_{0}\in\mathscr{S} such that XT0X\subseteq T_{0}.

Proof.

Let (W𝐩)𝐩P(W_{\mathbf{p}})_{\mathbf{p}\in P} be the parametric family given by applying Theorem 3.13 to (V𝐪)𝐪Q(V_{\mathbf{q}})_{\mathbf{q}\in Q}. Without loss of generality we may assume that for all 𝐩P\mathbf{p}\in P there is 𝐪Q\mathbf{q}\in Q such that W𝐩V𝐪W_{\mathbf{p}}\subseteq V_{\mathbf{q}}. Since V𝐪V_{\mathbf{q}} is always in Opt(V𝐪)\mathrm{Opt}(V_{\mathbf{q}}), we may remove this trivial optimal subvariety, and so we can assume that for every 𝐩P\mathbf{p}\in P, if W𝐩V𝐪W_{\mathbf{p}}\subseteq V_{\mathbf{q}}, then V𝐪V_{\mathbf{q}} is not in W𝐪W_{\mathbf{q}}.

Let TT be a special subvariety of Y(1)n\mathrm{Y}(1)^{n} and suppose that 𝐪Q\mathbf{q}\in Q is such that V𝐪TV_{\mathbf{q}}\cap T contains an atypical component XX. Then TT must be a proper subvariety of Y(1)n\mathrm{Y}(1)^{n}. XX is contained in a maximal atypical component of V𝐪V_{\mathbf{q}} which, by Remark 3.11, is contained in W𝐩W_{\mathbf{p}} for some 𝐩P\mathbf{p}\in P.

For every 𝐩P\mathbf{p}\in P, if ZZ is in W𝐩W_{\mathbf{p}}, then ZZ is an optimal proper subvariety of V𝐪V_{\mathbf{q}}, for some 𝐪Q\mathbf{q}\in Q. By Remark 3.11 we know that

dimVspcl(Z)>dimV+dimspcl(Z)n,\dim V\cap\mathrm{spcl}(Z)>\dim V+\dim\mathrm{spcl}(Z)-n,

and so in particular spcl(Z)\mathrm{spcl}(Z) is a proper special subvariety of Y(1)n\mathrm{Y}(1)^{n}. We conclude then that for every pPp\in P there are finitely many proper special subvarieties T1,𝐩,,Tm,𝐩T_{1,\mathbf{p}},\ldots,T_{m,\mathbf{p}} such that

(3.1) W𝐩i=1mTi,𝐩.W_{\mathbf{p}}\subseteq\bigcup_{i=1}^{m}T_{i,\mathbf{p}}.

We now use the compactness theorem from model theory. Let {Ti}i\left\{T_{i}\right\}_{i\in\mathbb{N}} be an enumeration of all the proper special subvarieties of Y(1)n\mathrm{Y}(1)^{n}. If the conclusion of the corollary were not true, then the following set of formulas (in the variables 𝐩\mathbf{p} and xx)

(3.2) {𝐩PxW𝐩xi=0mTi}m\left\{\mathbf{p}\in P\wedge x\in W_{\mathbf{p}}\wedge x\notin\bigcup_{i=0}^{m}T_{i}\right\}_{m\in\mathbb{N}}

would be finitely satisfiable and hence it would form a type in the language of rings with some extra constant symbols. As the family (W𝐩)𝐩P(W_{\mathbf{p}})_{\mathbf{p}\in P} only requires finitely many parameters to be defined and every TiT_{i} in definable over ¯\overline{\mathbb{Q}}, then (3.2) is a type in the language of rings with only countably many added constant symbols. As \mathbb{C} is 0\aleph_{0}-saturated, this type must be realised over \mathbb{C}, but that would mean that there is 𝐩P\mathbf{p}^{\star}\in P such that W𝐩W_{\mathbf{p}^{\star}} is not contained in the union of all proper special subvarieties, which contradicts (3.1). ∎

We will need the following “two-sorted” version of this result.

Corollary 3.15.

Let (U𝐪)𝐪Q(U_{\mathbf{q}})_{\mathbf{q}\in Q} be a parametric family of subvarieties of n×Y(1)n\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}. Then MZP implies that there is a finite collection 𝒮\mathscr{S} of proper special subvarieties of Y(1)n\mathrm{Y}(1)^{n} such that for all 𝐪Q\mathbf{q}\in Q and all special subvarieties TT, if XX is an atypical component of V𝐪(n×T)V_{\mathbf{q}}\cap(\mathbb{C}^{n}\times T), then there is T0𝒮T_{0}\in\mathscr{S} such that Xn×T0X\subseteq\mathbb{C}^{n}\times T_{0}.

Proof.

We follow the proof of [BK18, Theorem 11.4]. Let UQ×n×Y(1)nU\subseteq Q\times\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be the definable set such that for each 𝐪Q\mathbf{q}\in Q, U𝐪={(𝐱,𝐲)n×Y(1)n:(𝐪,𝐱,𝐲)U}U_{\mathbf{q}}=\left\{(\mathbf{x},\mathbf{y})\in\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}:(\mathbf{q},\mathbf{x},\mathbf{y})\in U\right\}. Given k{0,,dimU}k\in\left\{0,\ldots,\dim U\right\} define

U(k):={(𝐪,𝐱,𝐲)U:dim(U𝐪πY1(𝐲))=k}.U^{(k)}:=\left\{(\mathbf{q},\mathbf{x},\mathbf{y})\in U:\dim(U_{\mathbf{q}}\cap\pi_{\mathrm{Y}}^{-1}(\mathbf{y}))=k\right\}.

By definability of dimensions, the U(k)U^{(k)} are all constructible subsets of UU. Furthermore, define

V(k):=U(k)U(k+1)U(dimU).V^{(k)}:=U^{(k)}\cup U^{(k+1)}\cup\cdots\cup U^{(\dim U)}.

Given 𝐪Q\mathbf{q}\in Q let V𝐪(k):={(𝐱,𝐲)n×Y(1)n:(𝐪,𝐱,𝐲)V(k)}V_{\mathbf{q}}^{(k)}:=\left\{(\mathbf{x},\mathbf{y})\in\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}:(\mathbf{q},\mathbf{x},\mathbf{y})\in V^{(k)}\right\}.

For each k{0,,dimU}k\in\left\{0,\ldots,\dim U\right\} let 𝒮k\mathscr{S}_{k} be the finite collection of special subvarieties obtained by applying Corollary 3.14 to the family (πY(V𝐪(k)))𝐪Q\left(\pi_{\mathrm{Y}}\left(V_{\mathbf{q}}^{(k)}\right)\right)_{\mathbf{q}\in Q}. Set

𝒮:=k=0dimU𝒮k.\mathscr{S}:=\bigcup_{k=0}^{\dim U}\mathscr{S}_{k}.

Let TY(1)nT\subset\mathrm{Y(1)}^{n} be a special subvariety, and suppose that XX is an atypical component of U𝐪(n×T)U_{\mathbf{q}}\cap(\mathbb{C}^{n}\times T). Let k0=dimXdimπY(X)k_{0}=\dim X-\dim\pi_{\mathrm{Y}}(X), by the fibre-dimension theorem there is a Zariski open subset XXX^{\prime}\subset X such that for every (𝐱,𝐲)X(\mathbf{x},\mathbf{y})\in X^{\prime} we have that dimXπY1(𝐲)=k0\dim X^{\prime}\cap\pi_{\mathrm{Y}}^{-1}(\mathbf{y})=k_{0}. In particular dimX=dimX\dim X^{\prime}=\dim X. Also V𝐪(k0)U𝐪V_{\mathbf{q}}^{(k_{0})}\subseteq U_{\mathbf{q}}, so

dimX>dimU𝐪+dimTndimV𝐪(k0)+dimTn.\dim X>\dim U_{\mathbf{q}}+\dim T-n\geq\dim V_{\mathbf{q}}^{(k_{0})}+\dim T-n.

Let 𝐲πY(X)\mathbf{y}\in\pi_{\mathrm{Y}}(X^{\prime}), then dimXπY1(𝐲)=k0\dim X^{\prime}\cap\pi_{\mathrm{Y}}^{-1}(\mathbf{y})=k_{0}. By construction dimV𝐪(k0)πY1(𝐲)k0\dim V_{\mathbf{q}}^{(k_{0})}\cap\pi_{\mathrm{Y}}^{-1}(\mathbf{y})\geq k_{0}, so by the fibre-dimension theorem we get

dimπY(X)=dimπY(X)>dimπY(V𝐪(k0))+dimTn.\dim\pi_{\mathrm{Y}}(X)=\dim\pi_{\mathrm{Y}}(X^{\prime})>\dim\pi_{\mathrm{Y}}\left(V_{\mathbf{q}}^{(k_{0})}\right)+\dim T-n.

Then there is a special subvariety T0Y(1)nT_{0}\subset\mathrm{Y}(1)^{n} with Δ(T0)N\Delta(T_{0})\leq N such that πY(X)T0\pi_{\mathrm{Y}}(X^{\prime})\subseteq T_{0} and

dimπY(X)dimπY(V𝐪(k0)T0)+dimTT0dimT0.\dim\pi_{\mathrm{Y}}(X^{\prime})\leq\dim\pi_{\mathrm{Y}}\left(V_{\mathbf{q}}^{(k_{0})}\cap T_{0}\right)+\dim T\cap T_{0}-\dim T_{0}.

This shows that Xn×T0X^{\prime}\subset\mathbb{C}^{n}\times T_{0}, and since n×T0\mathbb{C}^{n}\times T_{0} is Zariski closed, then Xn×T0X\subset\mathbb{C}^{n}\times T_{0}. By the definition of V𝐪(k0)V_{\mathbf{q}}^{(k_{0})}, the dimension of the fibres of the restriction of πY\pi_{\mathrm{Y}} to V𝐪(k0)(n×T0)V_{\mathbf{q}}^{(k_{0})}\cap(\mathbb{C}^{n}\times T_{0}) is at least k0k_{0}. By fibre-dimension theorem we get:

dimXk0dimV𝐪(k0)(n×T0)k0+dimTT0dimT0.\dim X-k_{0}\leq\dim V_{\mathbf{q}}^{(k_{0})}\cap(\mathbb{C}^{n}\times T_{0})-k_{0}+\dim T\cap T_{0}-\dim T_{0}.

Since V𝐪(k0)U𝐪V_{\mathbf{q}}^{(k_{0})}\subseteq U_{\mathbf{q}}, this completes the proof. ∎

3.4. Weak MZP

Although MZP is open, Pila and Tsimerman ([PT16, §7]) showed that, as a consequence of the Ax–Schanuel theorem for jj, one can obtain a weak form of MZP which states that the atypical components of an algebraic subvariety of Y(1)n\mathrm{Y}(1)^{n}, are contained in finitely many parametric families of proper weakly special subvarieties. For the proofs of our main results, we will need the following version of the weak form of MZP which allows for parametric families of algebraic varieties and is “two-sorted” like Corollary 3.15.

Proposition 3.16.

Given a parametric family (U𝐪)𝐪Q(U_{\mathbf{q}})_{\mathbf{q}\in Q} of constructible subsets of n×Y(1)n\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}, there is a positive integer NN such that for every qQq\in Q, for every weakly special subvariety TY(1)nT\subset\mathrm{Y}(1)^{n} and for every atypical component XX of U𝐪(n×T)U_{\mathbf{q}}\cap(\mathbb{C}^{n}\times T), there is a proper weakly special subvariety T0Y(1)nT_{0}\subset\mathrm{Y}(1)^{n} with Δb(T0)N\Delta_{b}(T_{0})\leq N such that Xn×T0X\subseteq\mathbb{C}^{n}\times T_{0} and

dimXdimU𝐪(n×T0)+dimTT0dimT0.\dim X\leq\dim U_{\mathbf{q}}\cap(\mathbb{C}^{n}\times T_{0})+\dim T\cap T_{0}-\dim T_{0}.

In this section we will present a few technical results centred around the weak form of MZP, which build up to prove Proposition 3.16. We begin with a result of Aslanyan showing the following uniform version of weak MZP.444That Ax–Schanuel implies a uniform version of a weak form of the Zilber–Pink conjecture holds in very general contexts, see [ES22, Propositon 2.20] and [PS21].

Theorem 3.17 (Weak MZP, [Asl22b, Theorem 5.2]).

Let SS be a special subvariety of Y(1)n\mathrm{Y}(1)^{n}. Given a parametric family (U𝐪)𝐪Q(U_{\mathbf{q}})_{\mathbf{q}\in Q} of constructible subsets of SS, there is a finite collection 𝒮\mathscr{S} of proper special subvarieties of SS such that for every 𝐪\mathbf{q} in QQ and for every strongly atypical component XX of U𝐪U_{\mathbf{q}} in SS, there is T𝒮T\in\mathscr{S} such that XTX\subseteq T.

Corollary 3.18.

Let SS be a special subvariety of Y(1)n\mathrm{Y}(1)^{n}. Given a parametric family (U𝐪)𝐪Q(U_{\mathbf{q}})_{\mathbf{q}\in Q} of constructible subsets of SS, there is a finite collection 𝒮\mathscr{S} of proper special subvarieties of SS such that for every 𝐪\mathbf{q} in QQ and for every strongly atypical component XX of U𝐪U_{\mathbf{q}}, there is T0𝒮T_{0}\in\mathscr{S} satisfying the following conditions:

  1. (a)

    XT0X\subseteq T_{0},

  2. (b)

    XX is a typical component of Uspcl(X)=(U𝐪T0)(spcl(X)T0)U\cap\mathrm{spcl}(X)=(U_{\mathbf{q}}\cap T_{0})\cap(\mathrm{spcl}(X)\cap T_{0}) in T0T_{0}, that is:

    dimXdimU𝐪T0+dimspcl(X)dimT0,\dim X\leq\dim U_{\mathbf{q}}\cap T_{0}+\dim\mathrm{spcl}(X)-\dim T_{0},

    and

  3. (c)

    the intersection U𝐪T0U_{\mathbf{q}}\cap T_{0} is atypical in SS:

    dimU𝐪T0>dimUq+dimT0dimS.\dim U_{\mathbf{q}}\cap T_{0}>\dim U_{q}+\dim T_{0}-\dim S.
Proof.

We will first show that there is a family 𝒮\mathscr{S} satisfying conditions (a) and (b). We proceed by induction on the dimension of SS. When dimS=0\dim S=0, then SS is a just a point and there is nothing to prove as U𝐪=SU_{\mathbf{q}}=S for all 𝐪Q\mathbf{q}\in Q, so no U𝐪U_{\mathbf{q}} contains an atypical component. Now assume that dimS>0\dim S>0. Let 𝒮1\mathscr{S}_{1} be the finite collection of proper special subvarieties of SS obtained by applying Theorem 3.17 to (U𝐪)𝐪Q(U_{\mathbf{q}})_{\mathbf{q}\in Q}.

Suppose that XX is a strongly atypical component of U𝐪U_{\mathbf{q}}, then

dimX>dimU𝐪+dimspcl(X)dimS.\dim X>\dim U_{\mathbf{q}}+\dim\mathrm{spcl}(X)-\dim S.

Choose T1𝒮1T_{1}\in\mathscr{S}_{1} such that XT1X\subseteq T_{1}. If

dimX>dimU𝐪T1+dimspcl(X)T1dimT1,\dim X>\dim U_{\mathbf{q}}\cap T_{1}+\dim\mathrm{spcl}(X)\cap T_{1}-\dim T_{1},

then XX is a strongly atypical component of (U𝐪T1)(TT1)(U_{\mathbf{q}}\cap T_{1})\cap(T\cap T_{1}) in T1T_{1}. Since dimT1<dimS\dim T_{1}<\dim S we can apply the induction hypothesis on T1T_{1} to the family (U𝐪T1)𝐪Q\left(U_{\mathbf{q}}\cap T_{1}\right)_{\mathbf{q}\in Q} to obtain a finite collection 𝒮2\mathscr{S}_{2} of proper special subvarieties of T1T_{1} (which in turn are special subvarieties of SS) such that there is T0𝒮2T_{0}\in\mathscr{S}_{2} satisfying that XT0X\subseteq T_{0} and the intersection (U𝐪T0)(spcl(X)T0)(U_{\mathbf{q}}\cap T_{0})\cap(\mathrm{spcl}(X)\cap T_{0}) is typical in T0T_{0}.

Therefore, the collection 𝒮\mathscr{S} obtained by taking the union of 𝒮1\mathscr{S}_{1} with the finite collections obtained by the induction hypothesis applied to (U𝐪T0)𝐪Q\left(U_{\mathbf{q}}\cap T_{0}\right)_{\mathbf{q}\in Q} for every T0𝒮T_{0}\in\mathscr{S} satisfies conditions (a) and (b).

We now check that 𝒮\mathscr{S} satisfies (c). Suppose that XX is a strongly atypical component of U𝐪U_{\mathbf{q}}, and let T0𝒮T_{0}\in\mathscr{S} be an element satisfying (a) and (b). Observe that spcl(X)T0=spcl(X)\mathrm{spcl}(X)\cap T_{0}=\mathrm{spcl}(X). From (b) we get that

dimXdimspcl(X)dimU𝐪T0dimT0.\dim X-\dim\mathrm{spcl}(X)\leq\dim U_{\mathbf{q}}\cap T_{0}-\dim T_{0}.

Combining this with the fact that XX is an atypical component of U𝐪U_{\mathbf{q}} gives

dimU𝐪dimS<dimXdimspcl(X)dimU𝐪T0dimT0,\dim U_{\mathbf{q}}-\dim S<\dim X-\dim\mathrm{spcl}(X)\leq\dim U_{\mathbf{q}}\cap T_{0}-\dim T_{0},

from which we get

dimU𝐪T0>dimU𝐪+dimT0dimS,\dim U_{\mathbf{q}}\cap T_{0}>\dim U_{\mathbf{q}}+\dim T_{0}-\dim S,

thus confirming (c). ∎

Corollary 3.19.

Let (S𝐪)𝐪Q(S_{\mathbf{q}})_{\mathbf{q}\in Q} be a parametric family of proper weakly special subvarieties of a special subvariety SS of Y(1)n\mathrm{Y}(1)^{n}. Then there are only finitely many members of this family which are basic special subvarieties.

Proof.

Let 𝒮\mathscr{S} be the finite family of proper special subvarieties of SS obtained by applying Corollary 3.18 to the family (S𝐪)𝐪Q(S_{\mathbf{q}})_{\mathbf{q}\in Q}. Suppose 𝐪Q\mathbf{q}\in Q is such that S𝐪S_{\mathbf{q}} is a proper basic special subvariety of SS. Then

dimS𝐪S𝐪=dimS𝐪>dimS𝐪+dimS𝐪dimS,\dim S_{\mathbf{q}}\cap S_{\mathbf{q}}=\dim S_{\mathbf{q}}>\dim S_{\mathbf{q}}+\dim S_{\mathbf{q}}-\dim S,

which shows that S𝐪S_{\mathbf{q}} is a strongly atypical component of itself in SS. There is T0𝒮T_{0}\in\mathscr{S} such that SqT0S_{q}\subseteq T_{0} and

dimS𝐪dimS𝐪T0+dimS𝐪dimT0.\dim S_{\mathbf{q}}\leq\dim S_{\mathbf{q}}\cap T_{0}+\dim S_{\mathbf{q}}-\dim T_{0}.

From this we obtain that 0dimS𝐪T0dimT00\leq\dim S_{\mathbf{q}}\cap T_{0}-\dim T_{0}, and so we obtain that T0=S𝐪T_{0}=S_{\mathbf{q}}. Since 𝒮\mathscr{S} is a finite set, the result is proven. ∎

We can now get the following result with the same proof used in Corollary 3.15.

Theorem 3.20 (Two-sorted weak MZP).

Given a parametric family (U𝐪)𝐪Q(U_{\mathbf{q}})_{\mathbf{q}\in Q} of algebraic subvarieties of n×Y(1)n\mathbb{C}^{n}\times\mathrm{Y}(1)^{n}, there is a finite collection 𝒮\mathscr{S} of proper special subvarieties of Y(1)n\mathrm{Y}(1)^{n} such that for every 𝐪Q\mathbf{q}\in Q, every proper special subvariety TT of Y(1)n\mathrm{Y}(1)^{n} and every atypical irreducible component XX of U𝐪(n×T)U_{\mathbf{q}}\cap(\mathbb{C}^{n}\times T) satisfying that πY(X)\pi_{\mathrm{Y}}(X) has no constant coordinates, there is T0T_{0} in 𝒮\mathscr{S} such that Xn×T0X\subseteq\mathbb{C}^{n}\times T_{0}.

Although Theorem 3.17 and Corollaries 3.18 and 3.19 only speak about strongly atypical intersections, one can get the following result about general atypical intersections.

Proposition 3.21 ([Asl21, Proposition 3.4]).

Let SS be a special subvariety of Y(1)n\mathrm{Y}(1)^{n}. Given a parametric family (U𝐪)𝐪Q(U_{\mathbf{q}})_{\mathbf{q}\in Q} of constructible subsets of SS, there is a positive integer NN such that for every 𝐪Q\mathbf{q}\in Q and for every weakly special subvariety TT of SS, if XX is an atypical component of U𝐪TU_{\mathbf{q}}\cap T in SS, then there is a proper weakly special subvariety T0T_{0} of SS satisfying the following conditions:

  1. (a)

    Δb(T0)N\Delta_{b}(T_{0})\leq N,

  2. (b)

    XT0X\subseteq T_{0},

  3. (c)

    XX is a typical component of (U𝐪T0)(TT0)(U_{\mathbf{q}}\cap T_{0})\cap(T\cap T_{0}) in T0T_{0}:

    dimXdimU𝐪T0+dimwspcl(X)dimT0,\dim X\leq\dim U_{\mathbf{q}}\cap T_{0}+\dim\mathrm{wspcl}(X)-\dim T_{0},

    and

  4. (d)

    the intersection U𝐪T0U_{\mathbf{q}}\cap T_{0} is atypical in SS:

    dimU𝐪T0>dimU𝐪+dimT0dimS.\dim U_{\mathbf{q}}\cap T_{0}>\dim U_{\mathbf{q}}+\dim T_{0}-\dim S.
Proof.

The proof requires a few more calculations than the proof given in [Asl21]. We proceed by induction on nn, the case n=0n=0 being trivial. Let 𝒮\mathscr{S} be the finite family of special subvarieties given by Corollary 3.18 applied to (U𝐪)𝐪Q(U_{\mathbf{q}})_{\mathbf{q}\in Q}. Let N𝒮N_{\mathscr{S}} be the maximal complexity of the elements of 𝒮\mathscr{S}.

Suppose that for some 𝐪Q\mathbf{q}\in Q and some weakly special subvariety TST\subset S, the intersection U𝐪TU_{\mathbf{q}}\cap T contains an atypical component XX. We may assume that T=wspcl(X)T=\mathrm{wspcl}(X).

If XX is strongly atypical (i.e. has no constant coordinates), then spcl(X)=wspcl(X)\mathrm{spcl}(X)=\mathrm{wspcl}(X) and so we know that there is T0𝒮T_{0}\in\mathscr{S} satisfying (b), (c) and (d). By construction, Δb(T0)N𝒮\Delta_{b}(T_{0})\leq N_{\mathscr{S}}, which verifies (a).

Assume now that XX is not strongly atypical. Let 𝐢=(i1,,im)\mathbf{i}=(i_{1},\ldots,i_{m}) be the tuple of all of the coordinates which are constant on XX, let 𝐜Y(1)m\mathbf{c}\in\mathrm{Y}(1)^{m} be such that pr𝐢(X)={𝐜}\mathrm{pr}_{\mathbf{i}}(X)=\left\{\mathbf{c}\right\}, and define

S𝐜:=Spr𝐢1(𝐜),S_{\mathbf{c}}:=S\cap\mathrm{pr}_{\mathbf{i}}^{-1}(\mathbf{c}),

that is, S𝐜S_{\mathbf{c}} is the fibre in SS over 𝐜\mathbf{c}. As explained in [Asl21, Proposition 3.4], S𝐜S_{\mathbf{c}} is an irreducible variety, so it is weakly special. Observe that as TT is the weakly special closure of XX, then TS𝐜T\subset S_{\mathbf{c}}. When 𝐢\mathbf{i} is not all of 𝐧\mathbf{n}, let 𝐤=𝐧𝐢\mathbf{k}=\mathbf{n}\setminus\mathbf{i}. We consider now the possible cases.

Suppose that

dimU𝐪S𝐜>dimU𝐪+dimS𝐜dimS.\dim U_{\mathbf{q}}\cap S_{\mathbf{c}}>\dim U_{\mathbf{q}}+\dim S_{\mathbf{c}}-\dim S.

If dimXdimU𝐪S𝐜+dimTdimS𝐜\dim X\leq\dim U_{\mathbf{q}}\cap S_{\mathbf{c}}+\dim T-\dim S_{\mathbf{c}}, then we can let T0T_{0} be S𝐜S_{\mathbf{c}}, since Δb(S𝐜)=0\Delta_{b}(S_{\mathbf{c}})=0. In particular, this would be the case if XX happens to be a single point, in which case 𝐢=𝐧\mathbf{i}=\mathbf{n}.

If instead dimX>dimU𝐪S𝐜+dimTdimS𝐜\dim X>\dim U_{\mathbf{q}}\cap S_{\mathbf{c}}+\dim T-\dim S_{\mathbf{c}}, then 𝐢𝐧\mathbf{i}\neq\mathbf{n}. Consider the projection pr𝐤:Y(1)nY(1)nm\mathrm{pr}_{\mathbf{k}}:\mathrm{Y}(1)^{n}\to\mathrm{Y}(1)^{n-m}. Since we have that XTS𝐜X\subseteq T\subseteq S_{\mathbf{c}}, then dimX=dimpr𝐤(X)\dim X=\dim\mathrm{pr}_{\mathbf{k}}(X), dimT=dimpr𝐤(T)\dim T=\dim\mathrm{pr}_{\mathbf{k}}(T) and dimU𝐪S𝐜=dimpr𝐤(U𝐪S𝐜)\dim U_{\mathbf{q}}\cap S_{\mathbf{c}}=\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}}). Also, since S𝐜SS_{\mathbf{c}}\subseteq S, then dimS𝐜=dimpr𝐤(S𝐜)dimpr𝐤(S)\dim S_{\mathbf{c}}=\dim\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}})\leq\dim\mathrm{pr}_{\mathbf{k}}(S). Let X1X_{1} be the irreducible component of pr𝐤(U𝐪S𝐜)pr𝐤(T)\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap\mathrm{pr}_{\mathbf{k}}(T) containing pr𝐤(X)\mathrm{pr}_{\mathbf{k}}(X), and observe that pr𝐤(T)\mathrm{pr}_{\mathbf{k}}(T) is a weakly special subvariety of pr𝐤(S)\mathrm{pr}_{\mathbf{k}}(S). Thus

dimX1dimpr𝐤(X)\displaystyle\dim X_{1}\geq\dim\mathrm{pr}_{\mathbf{k}}(X) =dimX\displaystyle=\dim X
>dimU𝐪S𝐜+dimTdimS𝐜\displaystyle>\dim U_{\mathbf{q}}\cap S_{\mathbf{c}}+\dim T-\dim S_{\mathbf{c}}
=dimpr𝐤(U𝐪S𝐜)+dimpr𝐤(T)dimpr𝐤(S𝐜),\displaystyle=\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})+\dim\mathrm{pr}_{\mathbf{k}}(T)-\dim\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}}),

showing that X1X_{1} is atypical in pr𝐤(U𝐪S𝐜)pr𝐤(T)\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap\mathrm{pr}_{\mathbf{k}}(T). We also remark that by construction X1X_{1} is in fact strongly atypical, and so pr𝐤(S𝐜)\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}}) must in fact be a basic special subvariety of pr𝐤(S)\mathrm{pr}_{\mathbf{k}}(S). The collection (pr𝐤(S𝐜))𝐜Y(1)m\left(\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}})\right)_{\mathbf{c}\in\mathrm{Y}(1)^{m}} is a parametric family of weakly special subvarieties of pr𝐤(S)\mathrm{pr}_{\mathbf{k}}(S). By definability of dimensions, we may consider the parametric subfamily of those elements which are properly contained in pr𝐤(S)\mathrm{pr}_{\mathbf{k}}(S) (if there are any), and so by Corollary 3.19 among the whole family (pr𝐤(S𝐜))𝐜Y(1)m\left(\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}})\right)_{\mathbf{c}\in\mathrm{Y}(1)^{m}} we will find only finitely many basic special subvarieties. Specifically we can find a finite set CY(1)mC\subset\mathrm{Y}(1)^{m} such that every basic special subvariety found in (pr𝐤(S𝐜))𝐜Y(1)m\left(\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}})\right)_{\mathbf{c}\in\mathrm{Y}(1)^{m}} can be realised as S𝐜S_{\mathbf{c}} for some 𝐜C\mathbf{c}\in C.

We can apply the induction hypothesis to the family (pr𝐤(U𝐪S𝐜))𝐪Q\left(\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\right)_{\mathbf{q}\in Q} of constructible subsets of pr𝐤(S𝐜)\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}}). In this way we find a natural number N𝐤,𝐜N_{\mathbf{k},\mathbf{c}} and a proper weakly special subvariety T1T_{1} of pr𝐤(S𝐜)\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}}) such that

  1. (i)

    Δb(T1)N𝐤\Delta_{b}(T_{1})\leq N_{\mathbf{k}},

  2. (ii)

    X1T1X_{1}\subseteq T_{1},

  3. (iii)

    dimX1dimpr𝐤(U𝐪S𝐜)T1+dimwspcl(X1)dimT1\dim X_{1}\leq\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap T_{1}+\dim\mathrm{wspcl}(X_{1})-\dim T_{1}, and

  4. (iv)

    dimpr𝐤(U𝐪S𝐜)T1>dimpr𝐤(U𝐪S𝐜)+dimT1dimpr𝐤(S𝐜)\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap T_{1}>\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})+\dim T_{1}-\dim\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}}).

Let T0:=pr𝐤1(T1)S𝐜T_{0}:=\mathrm{pr}_{\mathbf{k}}^{-1}(T_{1})\cap S_{\mathbf{c}}. Then T0T_{0} is a weakly special subvariety of SS with the property that Δb(T0)=Δb(T1)N𝐤\Delta_{b}(T_{0})=\Delta_{b}(T_{1})\leq N_{\mathbf{k}}, dimT0=dimT1\dim T_{0}=\dim T_{1} and XT0X\subseteq T_{0}. Observe that since X1pr𝐤(T)X_{1}\subseteq\mathrm{pr}_{\mathbf{k}}(T) and pr𝐤(T)\mathrm{pr}_{\mathbf{k}}(T) is weakly special, then wspcl(X1)pr𝐤(T)\mathrm{wspcl}(X_{1})\subseteq\mathrm{pr}_{\mathbf{k}}(T). Also, one can readily check that pr𝐤(U𝐪T0)=pr𝐤(U𝐪S𝐜)T1\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap T_{0})=\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap T_{1}. So using (iii) we get:

dimX=dimpr𝐤(X)\displaystyle\dim X=\dim\mathrm{pr}_{\mathbf{k}}(X) dimX1\displaystyle\leq\dim X_{1}
dimpr𝐤(U𝐪S𝐜)T1+dimwspcl(X1)dimT1\displaystyle\leq\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap T_{1}+\dim\mathrm{wspcl}(X_{1})-\dim T_{1}
dimpr𝐤(U𝐪T0)+dimTdimT0\displaystyle\leq\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap T_{0})+\dim T-\dim T_{0}
=dimU𝐪T0+dimTdimT0,\displaystyle=\dim U_{\mathbf{q}}\cap T_{0}+\dim T-\dim T_{0},

and using (iv) we get

dimU𝐪T0\displaystyle\dim U_{\mathbf{q}}\cap T_{0} =dimpr𝐤(U𝐪T0)\displaystyle=\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap T_{0})
=dimpr𝐤(U𝐪S𝐜)T1\displaystyle=\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap T_{1}
>dimpr𝐤(U𝐪S𝐜)+dimT1dimpr𝐤(S𝐜)\displaystyle>\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})+\dim T_{1}-\dim\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}})
dimU𝐪S𝐜+dimT0dimpr𝐤(S𝐜)\displaystyle\geq\dim U_{\mathbf{q}}\cap S_{\mathbf{c}}+\dim T_{0}-\dim\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}})
>dimU𝐪+dimT0dimS.\displaystyle>\dim U_{\mathbf{q}}+\dim T_{0}-\dim S.

Thus, we can set N𝐤=max{N𝐤,𝐜𝐜C}N_{\mathbf{k}}=\max\left\{N_{\mathbf{k},\mathbf{c}}\mid\mathbf{c}\in C\right\}, and in this case T0T_{0} satisfies the conditions of the proposition.

Suppose now that

dim(U𝐪S𝐜)=dimU𝐪+dimS𝐜dimS.\dim(U_{\mathbf{q}}\cap S_{\mathbf{c}})=\dim U_{\mathbf{q}}+\dim S_{\mathbf{c}}-\dim S.

As TS𝐜T\subseteq S_{\mathbf{c}}, dimT=dimpr𝐤(T)\dim T=\dim\mathrm{pr}_{\mathbf{k}}(T) and dim(U𝐪S𝐜)=dimpr𝐤(U𝐪S𝐜)\dim(U_{\mathbf{q}}\cap S_{\mathbf{c}})=\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}}). Then

dimpr𝐤(X)=dimX\displaystyle\dim\mathrm{pr}_{\mathbf{k}}(X)=\dim X >dimU𝐪+dimTdimS\displaystyle>\dim U_{\mathbf{q}}+\dim T-\dim S
=dimpr𝐤(U𝐪S𝐜)+dimpr𝐤(T)dimpr𝐤(S𝐜).\displaystyle=\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})+\dim\mathrm{pr}_{\mathbf{k}}(T)-\dim\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}}).

This seems to puts us right back in the previous case, and the argument can still be carried through with a few subtleties, so we will just focus on those. Let X1X_{1} be the irreducible component of pr𝐤(U𝐪S𝐜)pr𝐤(T)\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap\mathrm{pr}_{\mathbf{k}}(T) containing pr𝐤(X)\mathrm{pr}_{\mathbf{k}}(X), and observe that X1X_{1} is strongly atypical. As before, we get a finite set CY(1)mC\subset\mathrm{Y}(1)^{m} such that every basic special subvariety found in (pr𝐤(S𝐜))𝐜Y(1)m\left(\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}})\right)_{\mathbf{c}\in\mathrm{Y}(1)^{m}} can be realised as S𝐜S_{\mathbf{c}} for some 𝐜C\mathbf{c}\in C. Let 𝒮𝐤,𝐜\mathscr{S}_{\mathbf{k},\mathbf{c}} be the finite family of special subvarieties of pr𝐤(S)\mathrm{pr}_{\mathbf{k}}(S) given by Corollary 3.18 applied to (pr𝐤(U𝐪S𝐜))𝐪Q\left(\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\right)_{\mathbf{q}\in Q} as a family in S𝐜S_{\mathbf{c}}. Then there is T1𝒮𝐤,𝐜T_{1}\in\mathscr{S}_{\mathbf{k},\mathbf{c}} such that

  1. (i)

    X1T1X_{1}\subseteq T_{1}

  2. (ii)

    dimX1dimpr𝐤(U𝐪S𝐜)T1+dimwspcl(X1)dimT1\dim X_{1}\leq\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap T_{1}+\dim\mathrm{wspcl}(X_{1})-\dim T_{1}, and

  3. (iii)

    dimpr𝐤(U𝐪S𝐜)T1>dimpr𝐤(U𝐪S𝐜)+dimT1dimpr𝐤(S𝐜)\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap T_{1}>\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})+\dim T_{1}-\dim\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}}).

Let N𝒮𝐤N_{\mathscr{S}_{\mathbf{k}}} be the maximal basic complexity of elements of 𝒮𝐤,𝐜\mathscr{S}_{\mathbf{k},\mathbf{c}} for all 𝐜C\mathbf{c}\in C.

Define T0:=pr𝐤1(T1)S𝐜T_{0}:=\mathrm{pr}_{\mathbf{k}}^{-1}(T_{1})\cap S_{\mathbf{c}}, which is a proper weakly special subvariety of SS satisfying Δb(T0)=Δb(T1)N𝒮𝐤\Delta_{b}\left(T_{0}\right)=\Delta_{b}(T_{1})\leq N_{\mathscr{S}_{\mathbf{k}}}, XT0X\subseteq T_{0},

dimX=dimpr𝐤(X)\displaystyle\dim X=\dim\mathrm{pr}_{\mathbf{k}}(X) dimX1\displaystyle\leq\dim X_{1}
dimpr𝐤(U𝐪T0)+dimpr𝐤(T)dimT0\displaystyle\leq\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap T_{0})+\dim\mathrm{pr}_{\mathbf{k}}(T)-\dim T_{0}
=dimU𝐪T0+dimTdimT0,\displaystyle=\dim U_{\mathbf{q}}\cap T_{0}+\dim T-\dim T_{0},

and

dimU𝐪T0\displaystyle\dim U_{\mathbf{q}}\cap T_{0} =dimU𝐪S𝐜T0\displaystyle=\dim U_{\mathbf{q}}\cap S_{\mathbf{c}}\cap T_{0}
=dimpr𝐤(U𝐪S𝐜)T1\displaystyle=\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})\cap T_{1}
>dimpr𝐤(U𝐪S𝐜)+dimT1dimpr𝐤(S𝐜)\displaystyle>\dim\mathrm{pr}_{\mathbf{k}}(U_{\mathbf{q}}\cap S_{\mathbf{c}})+\dim T_{1}-\dim\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}})
=dimU𝐪S𝐜+dimT0dimpr𝐤(S𝐜)\displaystyle=\dim U_{\mathbf{q}}\cap S_{\mathbf{c}}+\dim T_{0}-\dim\mathrm{pr}_{\mathbf{k}}(S_{\mathbf{c}})
dimU𝐪+dimT0dimS.\displaystyle\geq\dim U_{\mathbf{q}}+\dim T_{0}-\dim S.

Since there are finitely many tuples 𝐤\mathbf{k} to consider, by taking the maximum of the N𝒮𝐤N_{\mathscr{S}_{\mathbf{k}}}, the N𝐤N_{\mathbf{k}} and N𝒮N_{\mathscr{S}}, we get the desired NN. This completes the proof. ∎

Now one can prove Proposition 3.16 using the same method of proof as in Corollary 3.15. We leave the details to the reader.

4. Strong Existential Closedness

Let VmV\subseteq\mathbb{C}^{m} be an irreducible constructible set and KK\subset\mathbb{C} a subfield over which VV is defined. We say that a point 𝐯V\mathbf{v}\in V is generic in VV over KK if

tr.deg.KK(𝐯)=dimV.\mathrm{tr.deg.}_{K}K(\mathbf{v})=\dim V.
Lemma 4.1.

Let VmV\subseteq\mathbb{C}^{m} be an irreducible subvariety, and let KK\subset\mathbb{C} be a subfield over which VV is defined. Suppose that no coordinate is constant on VV. If 𝐯V\mathbf{v}\in V is generic in VV over KK, then every coordinate of 𝐯\mathbf{v} is transcendental over KK.

Proof.

Say that VV defined in the ring of polynomials [X1,,Xm]\mathbb{C}[X_{1},\ldots,X_{m}]. Since no coordinate is constant on VV, then for every cc\in\mathbb{C} and for every i{1,,m}i\in\left\{1,\ldots,m\right\} we have that V{Xi=c}V\cap\left\{X_{i}=c\right\} is a proper subvariety of VV. So, given 𝐯V\mathbf{v}\in V, if cK¯c\in\overline{K} and vi=cv_{i}=c, for some i{1,,m}i\in\left\{1,\ldots,m\right\}, then we have that

tr.deg.KK(𝐯)dimV{Xi=c}<dimV,\mathrm{tr.deg.}_{K}K(\mathbf{v})\leq\dim V\cap\left\{X_{i}=c\right\}<\dim V,

which prevents 𝐯\mathbf{v} from being generic in VV over KK. ∎

Definition.

We say that a variety Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} satisfies the strong existential closedness property (SEC for short) if for every finitely generated field KK\subset\mathbb{C} over which VV can be defined, there exists (𝐳,j(𝐳))V(\mathbf{z},j(\mathbf{z}))\in V such that (𝐳,j(𝐳))(\mathbf{z},j(\mathbf{z})) is generic in VV over KK.

Conjecture 4.2.

For every positive integer nn, every algebraic variety VV contained in n×Y(1)n\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} which is broad and free, if V(n×Y(1)n)V\cap\left(\mathbb{H}^{n}\times\mathrm{Y}(1)^{n}\right) is Zariski dense in VV, then VV satisfies (SEC).

The condition in Theorem 3.3 that π(V)\pi_{\mathbb{C}}(V) be Zariski dense in Y(1)n\mathrm{Y}(1)^{n} implies broadness but not freeness, and yet, (EC) still holds on VV. That is, even if VV is contained in a subvariety of the form n×T\mathbb{C}^{n}\times T, with TT a special subvariety of Y(1)n\mathrm{Y}(1)^{n}, VV will satisfy (EC) as long as π(V)\pi_{\mathbb{C}}(V) is Zariski dense in n\mathbb{C}^{n}. However, if VV is not free then it may not have points in VEjnV\cap\mathrm{E}_{j}^{n} which are generic, as shown in the next example.

Example 4.3.

Let DD denote the diagonal of Y(1)2\mathrm{Y}(1)^{2}, and let V=2×DV=\mathbb{C}^{2}\times D (so VV is defined over \mathbb{Q}). By Theorem 3.3 VEj2V\cap\mathrm{E}_{j}^{2} is Zariski dense in VV. However, every point (z1,z2,j(z1),j(z2))V(z_{1},z_{2},j(z_{1}),j(z_{2}))\in V satisfies that j(z1)=j(z2)j(z_{1})=j(z_{2}), and so for this point there exists γSL2()\gamma\in\mathrm{SL}_{2}(\mathbb{Z}) such that γz1=z2\gamma z_{1}=z_{2}. Therefore

tr.deg.(z1,z2,j(z1),j(z2))2<3=dimV.\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(z_{1},z_{2},j(z_{1}),j(z_{2}))\leq 2<3=\dim V.

The next proposition shows that in Conjecture 4.2 we may assume dimV=n\dim V=n.

Proposition 4.4 (cf [Asl22a, Lemma 4.30]).

Fix a positive integer nn. Assume that every subvariety of n×Y(1)n\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} which is free, broad and has dimension nn satisfies (SEC). Then every subvariety Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} which is free and broad satisfies (SEC).

Proof.

Let Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be an subvariety which is free and broad. Let k:=dimVnk:=\dim V-n. We proceed by induction on kk. The case k=0k=0 is given by the assumption, so we assume that k>0k>0.

Let KK be a finitely generated field over which VV is definable. Choose (𝐚,𝐛)V(\mathbf{a},\mathbf{b})\in V to be generic over KK. Let p1,,pn,q1,,qn1p_{1},\ldots,p_{n},q_{1},\ldots,q_{n-1}\in\mathbb{C} be algebraically independent over K(𝐚,𝐛)K(\mathbf{a},\mathbf{b}) and choose qnq_{n}\in\mathbb{C} such that

(4.1) i=1npiai+i=1nqibi=1.\sum_{i=1}^{n}p_{i}a_{i}+\sum_{i=1}^{n}q_{i}b_{i}=1.

Set K1:=K(𝐩,𝐪)K_{1}:=K(\mathbf{p},\mathbf{q}) (so K1K_{1} is also finitely generated) and let V1V_{1} be the K1¯\overline{K_{1}}-Zariski closure of (𝐚,𝐛)(\mathbf{a},\mathbf{b}). Clearly V1V_{1} is and dimV1<dimV\dim V_{1}<\dim V. We now check the remaining conditions to apply the induction hypothesis on V1V_{1}.

  1. (a)

    We first show that dimV1=dimV1\dim V_{1}=\dim V-1. Observe that

    tr.deg.KK(𝐚,𝐛)+tr.deg.K(𝐚,𝐛)K(𝐚,𝐛,𝐩,𝐪)\displaystyle\mathrm{tr.deg.}_{K}K(\mathbf{a},\mathbf{b})+\mathrm{tr.deg.}_{K(\mathbf{a},\mathbf{b})}K(\mathbf{a},\mathbf{b},\mathbf{p},\mathbf{q}) =tr.deg.KK(𝐚,𝐛,𝐩,𝐪)\displaystyle=\mathrm{tr.deg.}_{K}K(\mathbf{a},\mathbf{b},\mathbf{p},\mathbf{q})
    =tr.deg.KK1+tr.deg.K1K1(𝐚,𝐛),\displaystyle=\mathrm{tr.deg.}_{K}K_{1}+\mathrm{tr.deg.}_{K_{1}}K_{1}(\mathbf{a},\mathbf{b}),

    and since by construction

    tr.deg.KK(𝐚,𝐛)\displaystyle\mathrm{tr.deg.}_{K}K(\mathbf{a},\mathbf{b}) =dimV\displaystyle=\dim V
    tr.deg.K(𝐚,𝐛)K(𝐚,𝐛,𝐩,𝐪)\displaystyle\mathrm{tr.deg.}_{K(\mathbf{a},\mathbf{b})}K(\mathbf{a},\mathbf{b},\mathbf{p},\mathbf{q}) =2n1\displaystyle=2n-1
    tr.deg.KK1\displaystyle\mathrm{tr.deg.}_{K}K_{1} 2n1\displaystyle\geq 2n-1

    we conclude that dimV1=tr.deg.K1K1(𝐚,𝐛)=dimV1\dim V_{1}=\mathrm{tr.deg.}_{K_{1}}K_{1}(\mathbf{a},\mathbf{b})=\dim V-1.

  2. (b)

    We now show that V1V_{1} is free. Since VV is free, we know that (𝐚,𝐛)(\mathbf{a},\mathbf{b}) is not contained in any variety of the form M×TM\times T, where MM is a Möbius subvariety of n\mathbb{C}^{n} defined over \mathbb{Q} and TT is a proper special subvariety of Y(1)n\mathrm{Y}(1)^{n}. If V1V_{1} had a constant coordinate, then that would imply that some coordinate of (𝐚,𝐛)(\mathbf{a},\mathbf{b}) is in K1¯\overline{K_{1}}, but since (𝐚,𝐛)(\mathbf{a},\mathbf{b}) already satisfies (4.1) and is generic in V1V_{1} over K1K_{1}, then we would have that dimV1<dimV1\dim V_{1}<\dim V-1. Therefore V1V_{1} is free.

  3. (c)

    Next we show that V1V_{1} is broad. Let 𝐢=(i1,,i)\mathbf{i}=(i_{1},\ldots,i_{\ell}) be such that 1i1<<in1\leq i_{1}<\cdots<i_{\ell}\leq n and <n\ell<n. Consider the projection Pr𝐢(V1)\mathrm{Pr}_{\mathbf{i}}(V_{1}). We have that

    dimPr𝐢(V1)=tr.deg.K1K1(pr𝐢(𝐚),pr𝐢(𝐛)).\dim\mathrm{Pr}_{\mathbf{i}}(V_{1})=\mathrm{tr.deg.}_{K_{1}}K_{1}(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b})).

    Proceeding as we did in (a), we have that

    tr.deg.KK(pr𝐢(𝐚),pr𝐢(𝐛))+tr.deg.K(pr𝐢(𝐚),pr𝐢(𝐛))K(pr𝐢(𝐚),pr𝐢(𝐛),𝐩,𝐪)=tr.deg.KK1+tr.deg.K1K1(pr𝐢(𝐚),pr𝐢(𝐛)),\mathrm{tr.deg.}_{K}K(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b}))+\mathrm{tr.deg.}_{K(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b}))}K(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b}),\mathbf{p},\mathbf{q})=\\ \mathrm{tr.deg.}_{K}K_{1}+\mathrm{tr.deg.}_{K_{1}}K_{1}(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b})),

    and

    tr.deg.KK(pr𝐢(𝐚),pr𝐢(𝐛))\displaystyle\mathrm{tr.deg.}_{K}K(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b})) =dimPr𝐢(V)\displaystyle=\dim\mathrm{Pr}_{\mathbf{i}}(V)
    tr.deg.K(pr𝐢(𝐚),pr𝐢(𝐛))K(pr𝐢(𝐚),pr𝐢(𝐛),𝐩,𝐪)\displaystyle\mathrm{tr.deg.}_{K(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b}))}K(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b}),\mathbf{p},\mathbf{q}) 2n1\displaystyle\geq 2n-1
    tr.deg.KK1\displaystyle\mathrm{tr.deg.}_{K}K_{1} =2n.\displaystyle=2n.

    By [Asl22a, Lemma 4.31] we have that one of the two following cases must hold:

    1. (i)

      a1,,an,b1,,bnK(pr𝐢(𝐚),pr𝐢(𝐛))¯a_{1},\ldots,a_{n},b_{1},\ldots,b_{n}\in\overline{K(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b}))}. Then dimPr𝐢(V)=dimV\dim\mathrm{Pr}_{\mathbf{i}}(V)=\dim V, so

      dimPr𝐢(V1)n1.\dim\mathrm{Pr}_{\mathbf{i}}(V_{1})\geq n-1\geq\ell.
    2. (ii)

      tr.deg.K1K1(pr𝐢(𝐚),pr𝐢(𝐛))=tr.deg.KK(pr𝐢(𝐚),pr𝐢(𝐛))\mathrm{tr.deg.}_{K_{1}}K_{1}(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b}))=\mathrm{tr.deg.}_{K}K(\mathrm{pr}_{\mathbf{i}}(\mathbf{a}),\mathrm{pr}_{\mathbf{i}}(\mathbf{b})). Since VV is broad,

      dimPr𝐢(V1)=dimPr𝐢(V).\dim\mathrm{Pr}_{\mathbf{i}}(V_{1})=\dim\mathrm{Pr}_{\mathbf{i}}(V)\geq\ell.

We can now apply the induction hypothesis and deduce that there is (𝐳,j(𝐳))V1Ejn(\mathbf{z},j(\mathbf{z}))\in V_{1}\cap\mathrm{E}_{j}^{n} such that (𝐳,j(𝐳))(\mathbf{z},j(\mathbf{z})) is generic in V1V_{1} over K1K_{1}. In particular (𝐳,j(𝐳))(\mathbf{z},j(\mathbf{z})) satisfies (4.1), but since the coefficients of this equation are transcendental over KK, then

dimVtr.deg.KK(𝐳,j(𝐳))>tr.deg.K1K1(𝐳,j(𝐳))=dimV1,\dim V\geq\mathrm{tr.deg.}_{K}K(\mathbf{z},j(\mathbf{z}))>\mathrm{tr.deg.}_{K_{1}}K_{1}(\mathbf{z},j(\mathbf{z}))=\dim V_{1},

so (𝐳,j(𝐳))(\mathbf{z},j(\mathbf{z})) is generic in VV over KK. ∎

4.1. jj-derivations

As mentioned in §1.1, an obstacle towards proving Theorem 1.1 is that MSCD only gives a lower bound for the transcendence degree over \mathbb{Q}. Since we are looking for generic points over an arbitrary finitely generated field KK, we would like to have an inequality for the transcendence degree over KK. This is done in §4.2. Before that, we need to give a quick review of jj-derivations (see [AEK23, §3.1] and [Ete18, §5] for details).

A function :\partial:\mathbb{C}\rightarrow\mathbb{C} is called a jj-derivation if it satisfies the following axioms:

  1. (1)

    For all a,ba,b\in\mathbb{C}, (a+b)=(a)+(b)\partial(a+b)=\partial(a)+\partial(b).

  2. (2)

    For all a,ba,b\in\mathbb{C}, (ab)=a(b)+b(a)\partial(ab)=a\partial(b)+b\partial(a).

  3. (3)

    For all zz\in\mathbb{H} and all nn\in\mathbb{N}, (j(n)(z))=j(n+1)(z)(z)\partial\left(j^{(n)}(z)\right)=j^{(n+1)}(z)\partial(z), where j(n)j^{(n)} denotes the nn-th derivative of jj.

\mathbb{C} has non-trivial jj-derivations (in fact, it has continuum many \mathbb{C}-linearly independent jj-derivations).

Definition.

Let AA\subseteq\mathbb{C} be any set. We define the set jcl(A)j\mathrm{cl}(A) by the property: xjcl(A)x\in j\mathrm{cl}(A) if and only if (x)=0\partial(x)=0 for every jj-derivation \partial with AkerA\subseteq\ker\partial. If A=jcl(A)A=j\mathrm{cl}(A), then we say that AA is jclj\mathrm{cl}-closed.

Every jclj\mathrm{cl}-closed subset of \mathbb{C} is an algebraically closed subfield. Furthermore, jclj\mathrm{cl} has a corresponding well-defined notion of dimension555In technical terms, jclj\mathrm{cl} is a pregeometry., which we denote dimj\dim^{j}, defined in the following way. For any subsets A,BA,B\subseteq\mathbb{C}, dimj(A|B)n\dim^{j}(A|B)\geq n if and only if there exist a1,,anjcl(A)a_{1},\ldots,a_{n}\in j\mathrm{cl}(A) and jj-derivations 1,,n\partial_{1},\ldots,\partial_{n} such that BkeriB\subseteq\ker\partial_{i} for i=1,,ni=1,\ldots,n and

i(ak)={1 if i=k0 otherwise\partial_{i}\left(a_{k}\right)=\left\{\begin{array}[]{cl}1&\mbox{ if }i=k\\ 0&\mbox{ otherwise}\end{array}\right.

for every i,k=1,,ni,k=1,\ldots,n.

We define Cj:=jcl()C_{j}:=j\mathrm{cl}(\emptyset); this is a countable algebraically closed subfield of \mathbb{C} which is jclj\mathrm{cl}-closed.

Remark 4.5.

By definition, CjC_{j} is contained in the kernel of every jj-derivation, so for every finite set AA\subseteq\mathbb{C} we have that dimj(A)=dimj(A|Cj)\dim^{j}(A)=\dim^{j}(A|C_{j}).

Lemma 4.6.

Let CC\subseteq\mathbb{C} be jclj\mathrm{cl}-closed. For every zz\in\mathbb{H}, the following statements hold:

  1. (a)

    zCz\in C implies j(z),j(z),j′′(z)Cj(z),j^{\prime}(z),j^{\prime\prime}(z)\in C.666Due to the differential equation of jj (2.2), it is enough to only consider the derivatives up to j′′j^{\prime\prime}.

  2. (b)

    j(z)Cj(z)\in C implies zCz\in C.

Proof.

Follows from [Ete18, Proposition 5.8] and the fact that j(z)=0j^{\prime}(z)=0 implies that zz is algebraic ∎

The Ax–Schanuel theorem for jj [PT16, Theorem 1.3] has the following consequence:

Proposition 4.7 ([Ete18, Proposition 6.2]).

Let CC be a jclj\mathrm{cl}-closed subfield of \mathbb{C}, then for every z1,,znz_{1},\ldots,z_{n}\in\mathbb{H} we have that

tr.deg.CC(𝐳,j(𝐳),j(𝐳),j′′(𝐳))3dimG(𝐳|C)+dimj(𝐳|C).\mathrm{tr.deg.}_{C}C\left(\mathbf{z},j(\mathbf{z}),j^{\prime}(\mathbf{z}),j^{\prime\prime}(\mathbf{z})\right)\geq 3\dim_{G}\left(\mathbf{z}|C\right)+\dim^{j}\left(\mathbf{z}|C\right).

4.2. Convenient tuples

In this section we use the results of [AEK23] to show that MSCD implies the existence of “convenient generators” for any finitely generated field. We will use JJ to denote the triple of function (j,j,j′′)(j,j^{\prime},j^{\prime\prime}), so that if z1,,znz_{1},\ldots,z_{n} are elements of \mathbb{H}, then:

J(𝐳):=(j(z1),,j(zn),j(z1),,j(zn),j′′(z1),,j′′(zn)).J(\mathbf{z}):=(j(z_{1}),\ldots,j(z_{n}),j^{\prime}(z_{1}),\ldots,j^{\prime}(z_{n}),j^{\prime\prime}(z_{1}),\ldots,j^{\prime\prime}(z_{n})).
Definition.

We will say that a tuple 𝐭=(t1,,tm)\mathbf{t}=(t_{1},\ldots,t_{m}) of elements of Σ\mathbb{H}\setminus\Sigma is convenient for jj if

tr.deg.(𝐭,J(𝐭))=3dimG(𝐭)+dimj(𝐭).\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{t},J(\mathbf{t}))=3\dim_{G}(\mathbf{t})+\dim^{j}(\mathbf{t}).

We remark that since the coordinates of 𝐭\mathbf{t} are not in Σ\Sigma, then dimG(𝐭)=dimG(𝐭|Σ)\dim_{G}(\mathbf{t})=\dim_{G}(\mathbf{t}|\Sigma).

The convenience of such a tuple of elements is manifested in the following two lemmas, which show that we can obtain MSCD-type inequalities over fields generated by a convenient tuple.

Lemma 4.8.

Assume MSCD holds. Suppose that t1,,tmΣt_{1},\ldots,t_{m}\in\mathbb{H}\setminus\Sigma is a convenient tuple for jj and set F=(𝐭,J(𝐭))F=\mathbb{Q}\left(\mathbf{t},J\left(\mathbf{t}\right)\right). Then for any z1,,znz_{1},\ldots,z_{n}\in\mathbb{H} we have

(4.2) tr.deg.FF(𝐳,J(𝐳))\displaystyle\mathrm{tr.deg.}_{F}F\left(\mathbf{z},J(\mathbf{z})\right) 3dimG(𝐳|Σ,𝐭)+dimj(𝐳|𝐭), and\displaystyle\geq 3\dim_{G}\left(\mathbf{z}|\Sigma,\mathbf{t}\right)+\dim^{j}\left(\mathbf{z}|\mathbf{t}\right),\mbox{ and }
(4.3) tr.deg.FF(𝐳,j(𝐳))\displaystyle\mathrm{tr.deg.}_{F}F\left(\mathbf{z},j(\mathbf{z})\right) dimG(𝐳|Σ,𝐭)+dimj(𝐳|𝐭).\displaystyle\geq\dim_{G}\left(\mathbf{z}|\Sigma,\mathbf{t}\right)+\dim^{j}\left(\mathbf{z}|\mathbf{t}\right).
Proof.

Using the addition formula, we first get that:

(4.4) tr.deg.(𝐳,𝐭,J(𝐳),J(𝐭))=tr.deg.F+tr.deg.FF(𝐳,J(𝐳)).\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}\left(\mathbf{z},\mathbf{t},J(\mathbf{z}),J\left(\mathbf{t}\right)\right)=\mathrm{tr.deg.}_{\mathbb{Q}}F+\mathrm{tr.deg.}_{F}F(\mathbf{z},J(\mathbf{z})).

Similarly we get

(4.5) dimG(𝐳,𝐭|Σ)=dimG(𝐭|Σ)+dimG(𝐳|𝐭,Σ)\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)=\dim_{G}(\mathbf{t}|\Sigma)+\dim_{G}(\mathbf{z}|\mathbf{t},\Sigma)

and also

(4.6) dimj(𝐳,𝐭)=dimj(𝐭)+dimj(𝐳|𝐭).\dim^{j}(\mathbf{z},\mathbf{t})=\dim^{j}(\mathbf{t})+\dim^{j}(\mathbf{z}|\mathbf{t}).

Combining MSCD with (4.4), (4.5) and (4.6), we obtain (4.2). Inequality (4.3) now follows directly from (4.2). ∎

Remark 4.9.

Suppose that t1,,tmΣt_{1},\ldots,t_{m}\in\mathbb{H}\setminus\Sigma are such that j(t1),,j(tm)¯j(t_{1}),\ldots,j(t_{m})\in\overline{\mathbb{Q}}. Then under MSC we get tr.deg.(𝐭,j(𝐭))=tr.deg.(𝐭)=dimG(𝐭|Σ)\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{t},j(\mathbf{t}))=\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{t})=\dim_{G}(\mathbf{t}|\Sigma). So for any z1,,znz_{1},\ldots,z_{n}\in\mathbb{H} we can repeat the arguments in the proof of Lemma 4.8 to get:

tr.deg.(𝐭)(𝐳,𝐭,j(𝐳),j(𝐭))dimG(𝐳|Σ,𝐭).\mathrm{tr.deg.}_{\mathbb{Q}(\mathbf{t})}\mathbb{Q}(\mathbf{z},\mathbf{t},j(\mathbf{z}),j(\mathbf{t}))\geq\dim_{G}(\mathbf{z}|\Sigma,\mathbf{t}).
Lemma 4.10.

Assume MSCD holds. Let t1,,tmΣt_{1},\ldots,t_{m}\in\mathbb{H}\setminus\Sigma be a convenient tuple for jj. Then for any z1,,znz_{1},\ldots,z_{n}\in\mathbb{H} we have

tr.deg.(𝐳,𝐭,J(𝐳),J(𝐭))\displaystyle\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{z},\mathbf{t},J(\mathbf{z}),J(\mathbf{t})) 3dimG(𝐳,𝐭|Σ)+dimj(𝐭), and\displaystyle\geq 3\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)+\dim^{j}(\mathbf{t}),\mbox{ and }
tr.deg.(𝐳,𝐭,j(𝐳),j(𝐭))\displaystyle\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{z},\mathbf{t},j(\mathbf{z}),j(\mathbf{t})) dimG(𝐳,𝐭|Σ)+dimj(𝐭).\displaystyle\geq\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)+\dim^{j}(\mathbf{t}).
Proof.

Set F=(𝐭,J(𝐭))F=\mathbb{Q}\left(\mathbf{t},J\left(\mathbf{t}\right)\right). We proceed by contradiction, if z1,,znz_{1},\ldots,z_{n}\in\mathbb{H} are such

tr.deg.(𝐳,𝐭,j(𝐳),j(𝐭))<dimG(𝐳,𝐭|Σ)+dimj(𝐭),\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{z},\mathbf{t},j(\mathbf{z}),j(\mathbf{t}))<\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)+\dim^{j}(\mathbf{t}),

then we also get that

tr.deg.(𝐳,𝐭,J(𝐳),J(𝐭))<3dimG(𝐳,𝐭|Σ)+dimj(𝐭).\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{z},\mathbf{t},J(\mathbf{z}),J(\mathbf{t}))<3\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)+\dim^{j}(\mathbf{t}).

Then using that 𝐭\mathbf{t} is convenient for jj, Lemma 4.8, (4.4) and (4.5), we get:

3dimG(𝐳|Σ,𝐭)+tr.deg.F\displaystyle 3\dim_{G}(\mathbf{z}|\Sigma,\mathbf{t})+\mathrm{tr.deg.}_{\mathbb{Q}}F =3dimG(𝐳,𝐭|Σ)+dimj(𝐭)\displaystyle=3\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)+\dim^{j}(\mathbf{t})
>tr.deg.(𝐳,𝐭,J(𝐳),J(𝐭))\displaystyle>\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{z},\mathbf{t},J(\mathbf{z}),J(\mathbf{t}))
=tr.deg.F+tr.deg.FF(𝐳,J(𝐳))\displaystyle=\mathrm{tr.deg.}_{\mathbb{Q}}F+\mathrm{tr.deg.}_{F}F(\mathbf{z},J(\mathbf{z}))
tr.deg.F+3dim(𝐳|Σ,𝐭)+dimj(𝐳|𝐭).\displaystyle\geq\mathrm{tr.deg.}_{\mathbb{Q}}F+3\dim(\mathbf{z}|\Sigma,\mathbf{t})+\dim^{j}(\mathbf{z}|\mathbf{t}).

As dimj(𝐳|𝐭)0\dim^{j}(\mathbf{z}|\mathbf{t})\geq 0, this gives a contradiction. ∎

In the rest of the subsection we address the question of existence of convenient tuples, and furthermore finding a convenient tuple which contains the generators of a specific finitely generated field. We start by recalling the following result.

Theorem 4.11.

[AEK23, Theorem 5.1] Let FF be a subfield of \mathbb{C} such that tr.deg.CjF\mathrm{tr.deg.}_{C_{j}}F is finite. Then there exist t1,,tmCjt_{1},\ldots,t_{m}\in\mathbb{H}\setminus C_{j} such that:

  1. (A1):

    FCj(𝐭,J(𝐭))¯F\subseteq\overline{C_{j}\left(\mathbf{t},J\left(\mathbf{t}\right)\right)}, and

  2. (A2):

    tr.deg.CjCj(𝐭,J(𝐭))=3dimG(𝐭|Cj)+dimj(𝐭)\mathrm{tr.deg.}_{C_{j}}C_{j}\left(\mathbf{t},J\left(\mathbf{t}\right)\right)=3\dim_{G}\left(\mathbf{t}|C_{j}\right)+\dim^{j}\left(\mathbf{t}\right).

Assuming MSCD, we now refine this result to show that convenient tuples exist.

Lemma 4.12.

Let KK\subset\mathbb{C} be a finitely generated subfield. Then MSCD implies that there exist t1,,tmΣt_{1},\ldots,t_{m}\in\mathbb{H}\setminus\Sigma such that:

  1. (c1):

    K¯(𝐭,J(𝐭))¯\overline{K}\subseteq\overline{\mathbb{Q}\left(\mathbf{t},J\left(\mathbf{t}\right)\right)},

  2. (c2):

    𝐭\mathbf{t} is convenient for jj.

Furthermore, we can assume that dimG(𝐭|Σ)=m\dim_{G}(\mathbf{t}|\Sigma)=m.

Proof.

By Theorem 4.11 there exist 𝐭1:=(t1,,tk)(Cj)k\mathbf{t}_{1}:=(t_{1},\ldots,t_{k})\in(\mathbb{H}\setminus C_{j})^{k} such that

  1. (a)

    KCj(𝐭1,J(𝐭1))¯K\subseteq\overline{C_{j}\left(\mathbf{t}_{1},J\left(\mathbf{t}_{1}\right)\right)},

  2. (b)

    tr.deg.CjCj(𝐭1,J(𝐭1))=3dimG(𝐭1|Cj)+dimj(𝐭1)\mathrm{tr.deg.}_{C_{j}}C_{j}\left(\mathbf{t}_{1},J\left(\mathbf{t}_{1}\right)\right)=3\dim_{G}\left(\mathbf{t}_{1}|C_{j}\right)+\dim^{j}\left(\mathbf{t}_{1}\right).

As tr.deg.CjCj(𝐭1,J(𝐭1))\mathrm{tr.deg.}_{C_{j}}C_{j}\left(\mathbf{t}_{1},J\left(\mathbf{t}_{1}\right)\right) is finite, then there is a finitely generated field FCF\subseteq C such that tr.deg.CjCj(𝐭1,J(𝐭1))=tr.deg.FF(𝐭1,J(𝐭1))\mathrm{tr.deg.}_{C_{j}}C_{j}\left(\mathbf{t}_{1},J\left(\mathbf{t}_{1}\right)\right)=\mathrm{tr.deg.}_{F}F\left(\mathbf{t}_{1},J\left(\mathbf{t}_{1}\right)\right). As KK is finitely generated, if LL denotes the compositum of FF and KCjK\cap C_{j}, then LL has finite transcendence degree over \mathbb{Q}, so by [AEK23, Theorem 6.18], MSCD implies that there exist 𝐭2=(tk+1,,tm)(Cj)mk\mathbf{t}_{2}=(t_{k+1},\ldots,t_{m})\in(\mathbb{H}\cap C_{j})^{m-k} such that

  1. (i)

    L(𝐭2,J(𝐭2))¯L\subseteq\overline{\mathbb{Q}\left(\mathbf{t}_{2},J\left(\mathbf{t}_{2}\right)\right)},

  2. (ii)

    tr.deg.(𝐭2,J(𝐭2))=3dimG(𝐭2|Σ)\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}\left(\mathbf{t}_{2},J\left(\mathbf{t}_{2}\right)\right)=3\dim_{G}\left(\mathbf{t}_{2}\right|\Sigma).

The paragraph immediately following [AEK23, Theorem 6.18] shows that the coordinates of 𝐭2\mathbf{t}_{2} may be chosen outside of Σ\Sigma. Let 𝐭=(𝐭1,𝐭2)\mathbf{t}=(\mathbf{t}_{1},\mathbf{t}_{2}). By construction, the elements of 𝐭1\mathbf{t}_{1} share no GG-orbits with element of 𝐭2\mathbf{t}_{2}. Condition (c1) is satisfied by (a) and (i). As FF is contained in LL, then condition (c2) is satisfied by (ii) and (b).

The “furthermore” part follows from Schneider’s theorem and the equivalence between (M1) and (M2) in §2.3. ∎

4.3. Proof of Theorem 1.1

We first set up some notation that will be kept through the rest of this subsection.

Let Vn×Y(1)nV\subset\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be a broad and free variety. Let KK\subset\mathbb{C} be a finitely generated subfield such that VV is defined over KK and let t1,,tmΣt_{1},\ldots,t_{m}\in\mathbb{H}\setminus\Sigma be given by Lemma 4.12 (which assumes MSCD) applied to KK (mm may be zero, which happens when K¯K\subseteq\overline{\mathbb{Q}}). We assume that dimG(𝐭|Σ)=m\dim_{G}(\mathbf{t}|\Sigma)=m. Set F:=(𝐭,j(𝐭))F:=\mathbb{Q}(\mathbf{t},j(\mathbf{t})).

Choose (𝐱,𝐲)V(\mathbf{x},\mathbf{y})\in V generic over FF. By Lemma 4.1 and the freeness of VV we know that no coordinate of (𝐱,𝐲)(\mathbf{x},\mathbf{y}) is in F¯\overline{F}. Let Wn+m×Y(1)n+mW\subseteq\mathbb{C}^{n+m}\times\mathrm{Y}(1)^{n+m} be the ¯\overline{\mathbb{Q}}-Zariski closure of the point (𝐱,𝐭,𝐲,j(𝐭))(\mathbf{x},\mathbf{t},\mathbf{y},j(\mathbf{t})).

Lemma 4.13.

dimW=dimV+m+dimj(𝐭)\dim W=\dim V+m+\dim^{j}(\mathbf{t}).

Proof.
dimW=tr.deg.(𝐱,𝐭,𝐲,j(𝐭))\displaystyle\dim W=\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{x},\mathbf{t},\mathbf{y},j(\mathbf{t})) =tr.deg.(𝐭,j(𝐭))+tr.deg.(𝐭,j(𝐭))(𝐱,𝐭,𝐲,j(𝐭))\displaystyle=\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{t},j(\mathbf{t}))+\mathrm{tr.deg.}_{\mathbb{Q}(\mathbf{t},j(\mathbf{t}))}\mathbb{Q}(\mathbf{x},\mathbf{t},\mathbf{y},j(\mathbf{t}))
=dimG(𝐭|Σ)+dimj(𝐭)+dimV.\displaystyle=\dim_{G}(\mathbf{t}|\Sigma)+\dim^{j}(\mathbf{t})+\dim V.\qed

We are now ready to prove the first main result.

Proof of Theorem 1.1.

By Proposition 4.4, we can reduce to the case dimV=n\dim V=n. The case n=1n=1 was proven in [EH21, Theorem 1.2], so now we assume that n>1n>1.

Consider the parametric family (W𝐪)𝐪Q(W_{\mathbf{q}})_{\mathbf{q}\in Q} of subvarieties of WW defined in Example 3.12. Let 𝒮\mathscr{S} be the finite collection of proper special subvarieties of Y(1)n\mathrm{Y}(1)^{n} given by Corollary 3.15 applied to the parametric family (πY(W𝐪))𝐪Q\left(\pi_{\mathrm{Y}}(W_{\mathbf{q}})\right)_{\mathbf{q}\in Q}. Let NN be the integer given by applying Proposition 3.16 to the family (W𝐪)𝐪Q(W_{\mathbf{q}})_{\mathbf{q}\in Q}. Without loss of generality we may assume that Δb(T)N\Delta_{b}(T)\leq N for all T𝒮T\in\mathscr{S}.

Let W1WW_{1}\subseteq W be a Zariski open subset such that for all (𝐚,𝐛)W1(\mathbf{a},\mathbf{b})\in W_{1} we have that 𝐛T0\mathbf{b}\notin T_{0} for all T0𝒮T_{0}\in\mathscr{S}.

Observe that we can choose (𝐳,j(𝐳))V(\mathbf{z},j(\mathbf{z}))\in V such that (j(𝐳),j(𝐭))T0(j(\mathbf{z}),j(\mathbf{t}))\notin T_{0} for all T0𝒮T_{0}\in\mathscr{S}. This is because every equation defining T0T_{0} either gives a modular dependence between two coordinates of j(𝐳)j(\mathbf{z}), or it gives a modular dependence between a coordinate of j(𝐳)j(\mathbf{z}) and a coordinate of j(𝐭)j(\mathbf{t}). As VV is free, satisfies (EC), and 𝒮\mathscr{S} is finite, we can find the desired point. This way we get that (𝐳,𝐭,j(𝐳),j(𝐭))W1(\mathbf{z},\mathbf{t},j(\mathbf{z}),j(\mathbf{t}))\in W_{1}.

We will show that dimG(𝐳,𝐭|Σ)=n+m\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)=n+m. For this we proceed by contradiction, so suppose that dimG(𝐳,𝐭|Σ)<n+m\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)<n+m. Let TT be the special closure of the point (j(𝐳),j(𝐭))(j(\mathbf{z}),j(\mathbf{t})). Observe that dimT=dimG(𝐳,𝐭|Σ)\dim T=\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma).

Since dimG(𝐳,𝐭|Σ)<n+m\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)<n+m and dimG(𝐭|Σ)=m\dim_{G}(\mathbf{t}|\Sigma)=m, then that at least one of the following happen:

  1. (a)

    There is i{1,,n}i\in\left\{1,\ldots,n\right\}, k{1,,m}k\in\left\{1,\ldots,m\right\}, and gGg\in G such that gzi=tkgz_{i}=t_{k}.

  2. (b)

    There are i,k{1,,n}i,k\in\left\{1,\ldots,n\right\} (possibly equal) and gGg\in G such that gzi=zkgz_{i}=z_{k}.

Let Mn+mM\subset\mathbb{C}^{n+m} be a proper subvariety defined by Möbius relations defined over \mathbb{Q} and/or setting some coordinates to be a constant in Σ\Sigma, satisfying (𝐳,𝐭)M(\mathbf{z},\mathbf{t})\in M. In other words, MM is a Möbius variety witnessing the relations found in (a) and (b). Thus W(M×Y(1)n+m)W\cap(M\times\mathrm{Y}(1)^{n+m}) is an element of the family (W𝐪)𝐪Q(W_{\mathbf{q}})_{\mathbf{q}\in Q}, call it WMW_{M}. We remark that WMW_{M} is defined over ¯\overline{\mathbb{Q}}.

We now claim that dimWM<dimW\dim W_{M}<\dim W. Indeed, if the dimensions were equal, then WM=WW_{M}=W and so VPr𝐢(WM)V\subset\mathrm{Pr}_{\mathbf{i}}(W_{M}). An equation defining MM coming from either (a) or (b) would then immediately contradicts freeness of VV.

Now let XX be the irreducible component of WM(n+m×T)W_{M}\cap(\mathbb{C}^{n+m}\times T) containing (𝐳,𝐭,j(𝐳),j(𝐭))(\mathbf{z},\mathbf{t},j(\mathbf{z}),j(\mathbf{t})). Observe that XX is defined over ¯\overline{\mathbb{Q}}. Then by Lemma 4.10 (which assumes MSCD) we get

dimXtr.deg.(𝐳,𝐭,j(𝐳),j(𝐭))dimG(𝐳,𝐭|Σ)+dimj(𝐭).\dim X\geq\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{z},\mathbf{t},j(\mathbf{z}),j(\mathbf{t}))\geq\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)+\dim^{j}(\mathbf{t}).

On the other hand, using that dimWM<dimW\dim W_{M}<\dim W and Lemma 4.13 we get:

dimWM+dimn+m×Tdimn+m×Y(1)n+m\displaystyle\dim W_{M}+\dim\mathbb{C}^{n+m}\times T-\dim\mathbb{C}^{n+m}\times\mathrm{Y}(1)^{n+m} <dimW+dimTn\displaystyle<\dim W+\dim T-n
=dimT+dimj(𝐭)\displaystyle=\dim T+\dim^{j}(\mathbf{t})
dimG(𝐳,𝐭|Σ)+dimj(𝐭)\displaystyle\leq\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)+\dim^{j}(\mathbf{t})
dimX.\displaystyle\leq\dim X.

This shows that XX is an atypical component of WM(n+m×T)W_{M}\cap(\mathbb{C}^{n+m}\times T). But then by Corollary 3.15 there must be T0𝒮T_{0}\in\mathscr{S} such that Xn+m×T0X\subseteq\mathbb{C}^{n+m}\times T_{0}, which is a contradiction since (j(𝐳),j(𝐭))T0(j(\mathbf{z}),j(\mathbf{t}))\notin T_{0}.

We have thus shown that dimG(𝐳,𝐭|Σ)=n+m\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)=n+m, which proves the theorem by Lemma 4.8. ∎

Remark 4.14.

In the proof of Theorem 1.1, if we somehow knew beforehand that the coordinates of j(𝐳)j(\mathbf{z}) are all transcendental over \mathbb{Q}, then we would not need to invoke MZP, because as XX is defined over ¯\overline{\mathbb{Q}}, then any coordinate that is constant on XX must equal an element of ¯\overline{\mathbb{Q}}. So if we know that the coordinates of j(𝐳)j(\mathbf{z}) are all transcendental over \mathbb{Q}, then we would know that XX does not have constant coordinates, hence XX is strongly atypical and we can apply Theorem 3.20.

4.4. Blurring

We will now give an analogue of Theorem 1.1 for the the so-called blurring of jj by GG (see [AK22] for more details).

Theorem 4.15.

Let V2nV\subseteq\mathbb{C}^{2n} be a broad and free irreducible variety with. Then MSCD and MZP imply that for every finitely generated field KK\subset\mathbb{C}, there are g1,,gnGg_{1},\ldots,g_{n}\in G such that VV has a point of the form

(z1,,zn,j(g1z1),,j(gnzn)),(z_{1},\ldots,z_{n},j(g_{1}z_{1}),\ldots,j(g_{n}z_{n})),

with (z1,,zn)n(z_{1},\ldots,z_{n})\in\mathbb{H}^{n}, which is generic over KK.777In fact, using [AK22, Theorem 3.1], g1,,gng_{1},\ldots,g_{n} may be chosen to be upper-triangular.

To prove Theorem 4.15, proceed in exactly the same way as in the proof of Theorem 1.1 (recall that Theorem 3.7 already ensures the existence of a Zariski dense set of solutions). The key (trivial) observation is that for all z,wz,w\in\mathbb{H} and all g,hGg,h\in G we have that j(z)j(z) and j(w)j(w) are modularly dependent if and only if j(gz)j(gz) and j(hw)j(hw) are modularly dependent. This manifests in the different ingredients of the proof as follows.

  1. (a)

    MSCD implies that for every g1,,gnGg_{1},\ldots,g_{n}\in G and every z1,,znnz_{1},\ldots,z_{n}\in\mathbb{H}^{n} we have

    tr.deg.(z1,,zn,j(g1z1),,j(gnzn))dimG(z1,,zn|Σ).\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(z_{1},\ldots,z_{n},j(g_{1}z_{1}),\ldots,j(g_{n}z_{n}))\geq\dim_{G}(z_{1},\ldots,z_{n}|\Sigma).
  2. (b)

    One can restate the inequalities of §4.2 in a straightforward way. This is because for any field FF, for all z1,,znz_{1},\ldots,z_{n}\in\mathbb{H} and all g1,,gnGg_{1},\ldots,g_{n}\in G we have that

    tr.deg.FF(z1,,zn,j(z1),,j(zn))=tr.deg.FF(z1,,zn,j(g1z1),,j(gnzn)).\mathrm{tr.deg.}_{F}F(z_{1},\ldots,z_{n},j(z_{1}),\ldots,j(z_{n}))=\mathrm{tr.deg.}_{F}F(z_{1},\ldots,z_{n},j(g_{1}z_{1}),\ldots,j(g_{n}z_{n})).

    So, for example, we can restate Lemma 4.10 as: let t1,,tmΣt_{1},\ldots,t_{m}\in\mathbb{H}\setminus\Sigma be a convenient tuple for jj, then for all z1,,znz_{1},\ldots,z_{n}\in\mathbb{H} and all g1,,gnGg_{1},\ldots,g_{n}\in G we have that:

    tr.deg.(𝐳,𝐭,j(g1z1),,j(gnzn),j(𝐭))dimG(𝐳,𝐭|Σ)+dimj(𝐭).\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(\mathbf{z},\mathbf{t},j(g_{1}z_{1}),\ldots,j(g_{n}z_{n}),j(\mathbf{t}))\geq\dim_{G}(\mathbf{z},\mathbf{t}|\Sigma)+\dim^{j}(\mathbf{t}).
  3. (c)

    Corollary 3.15 still applies as stated.

4.5. Special solutions

In this subsection we show that for a variety VV as in Theorem 1.1, the set

VΣ:={(𝜶,j(𝜷))V:j(𝜶),j(𝜷) are special points}V_{\Sigma}:=\left\{(\boldsymbol{\alpha},j(\boldsymbol{\beta}))\in V:j(\boldsymbol{\alpha}),j(\boldsymbol{\beta})\mbox{ are special points}\right\}

cannot be Zariski dense in VV. We show this in Proposition 4.16. The proof relies on the results of [Pil11] (Pila’s proof of the André–Oort conjecture for powers of the modular curve).

Let XX denote a subvariety of Y(1)n\mathrm{Y}(1)^{n} defined over a finitely generated field FF. Let 𝒦X\mathcal{K}_{X} denote the set of all subfields of F¯\overline{F} over which XX is definable. Define

δF(X):=min{[K:F]K𝒦X}.\delta_{F}(X):=\min\left\{[K:F]\mid K\in\mathcal{K}_{X}\right\}.
Definition.

Given a special point τΣ\tau\in\Sigma, there is a unique quadratic polynomial ax2+bx+cax^{2}+bx+c, with a,b,ca,b,c\in\mathbb{Z}, gcd(a,b,c)=1\mathrm{gcd}(a,b,c)=1 and a>0a>0, such that aτ2+bτ+c=0a\tau^{2}+b\tau+c=0. Let Dτ=b24acD_{\tau}=b^{2}-4ac. If 𝝉=(τ1,,τn)Σn\boldsymbol{\tau}=(\tau_{1},\ldots,\tau_{n})\in\Sigma^{n}, we define

disc(𝝉):=max{Dτ1,,Dτn}.\mathrm{disc}(\boldsymbol{\tau}):=\max\left\{D_{\tau_{1}},\ldots,D_{\tau_{n}}\right\}.

For every γSL2()\gamma\in\mathrm{SL}_{2}(\mathbb{Z}) we have that disc(γτ)=disc(τ)\mathrm{disc}(\gamma\tau)=\mathrm{disc}(\tau), so it makes sense to define:

disc(j(𝝉)):=disc(𝝉).\mathrm{disc}(j(\boldsymbol{\tau})):=\mathrm{disc}(\boldsymbol{\tau}).
Definition.

A basic special subvariety TY(1)nT\subset\mathrm{Y}(1)^{n} is defined by certain modular polynomials in the ring [Y1,,Yn]\mathbb{C}[Y_{1},\ldots,Y_{n}]. Suppose that 1i1<<in1\leq i_{1}<\cdots<i_{\ell}\leq n denote some indices such that, for every s{1,,s}s\in\left\{1,\ldots,s\right\}, the variable YisY_{i_{s}} does not appear in any the polynomials defining TT. In particular this implies that pr𝐢(T)=Y(1)\mathrm{pr}_{\mathbf{i}}(T)=\mathrm{Y}(1)^{\ell}. Following [Pil11], given 𝐲Y(1)\mathbf{y}\in\mathrm{Y}(1)^{\ell} we call Tpr1(𝐲)T\cap\mathrm{pr}^{-1}(\mathbf{y}) the translate of TT by 𝐲\mathbf{y}, and denote it as tr(T,𝐲)\mathrm{tr}(T,\mathbf{y}).

We remark that tr(T,𝐲)\mathrm{tr}(T,\mathbf{y}) is a weakly special subvariety, and if every coordinate of 𝐲\mathbf{y} is a special point, then tr(T,𝐲)\mathrm{tr}(T,\mathbf{y}) is a special subvariety.

Proposition 4.16.

Let Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be an irreducible variety which is modularly free and defined over ¯\overline{\mathbb{Q}}. Suppose that the dimension of the generic fibre of πV\pi_{\mathbb{C}}\upharpoonright_{V} is less than nn. Then VΣV_{\Sigma} is not Zariski dense in VV.

Proof.

Given 𝐚n\mathbf{a}\in\mathbb{C}^{n}, let

V𝐚:={𝐛Y(1)n:(𝐚,𝐛)V}.V_{\mathbf{a}}:=\left\{\mathbf{b}\in\mathrm{Y}(1)^{n}:(\mathbf{a},\mathbf{b})\in V\right\}.

In this way we obtain a parametric family (V𝐚)𝐚π(V)(V_{\mathbf{a}})_{\mathbf{a}\in\pi_{\mathbb{C}}(V)} of subvarieties of Y(1)n\mathrm{Y}(1)^{n}.

Let F¯F\subset\overline{\mathbb{Q}} be a finitely generated field such that VV is definable over FF. For every (𝐚,𝐛)V(\mathbf{a},\mathbf{b})\in V, V𝐚V_{\mathbf{a}} is definable over the field F(𝐚)F(\mathbf{a}). If (𝐚,𝐛)VΣ(\mathbf{a},\mathbf{b})\in V_{\Sigma}, then since every coordinate of 𝐚\mathbf{a} defines a degree 2 extension of \mathbb{Q}, we get that δ(V𝐚)2n[F:]\delta_{\mathbb{Q}}(V_{\mathbf{a}})\leq 2^{n}[F:\mathbb{Q}].

By [Pil11, Theorem 13.2] (uniform André-Oort) there is a finite collection 𝒮\mathscr{S} of basic special subvarieties of Y(1)n\mathrm{Y}(1)^{n} such that for every T𝒮T\in\mathscr{S} there is a constant C>0C>0 (depending only on nn, VV and TT) such that, if 𝐢=(i1,,i)\mathbf{i}=(i_{1},\ldots,i_{\ell}) denotes indices as in the definition above, then for every 𝐚n\mathbf{a}\in\mathbb{C}^{n} and for every 𝝉\boldsymbol{\tau}\in\mathbb{H}^{\ell}, if δ(V𝐚)2n[F:]\delta_{\mathbb{Q}}(V_{\mathbf{a}})\leq 2^{n}[F:\mathbb{Q}] and tr(T,j(𝝉))\mathrm{tr}(T,j(\boldsymbol{\tau})) is a maximal special subvariety of V𝐚V_{\mathbf{a}}, then

disc(𝝉)C and δ(tr(T,j(𝝉)))C.\mathrm{disc}(\boldsymbol{\tau})\leq C\quad\mbox{ and }\quad\delta_{\mathbb{Q}}(\mathrm{tr}(T,j(\boldsymbol{\tau})))\leq C.

We remark that by [Pil11, Proposition 13.1], all the maximal special subvarieties of V𝐚V_{\mathbf{a}} are of the form tr(T,j(𝝉))\mathrm{tr}(T,j(\boldsymbol{\tau})).

As explained in [Pil11, §5.6], there are finitely many values of j(𝝉)j(\boldsymbol{\tau}) subject to disc(𝝉)C\mathrm{disc}(\boldsymbol{\tau})\leq C. Therefore, there is a finite collection 𝒮\mathscr{S}^{\star} of special subvarieties of Y(1)n\mathrm{Y}(1)^{n} such that for every (𝐚,𝐛)VΣ(\mathbf{a},\mathbf{b})\in V_{\Sigma}, the maximal special subvarieties of V𝐚V_{\mathbf{a}} are elements of 𝒮\mathscr{S}^{\star}.

By hypothesis and the fibre-dimension theorem there is a Zariski open subset VVV^{\prime}\subset V such that for all (𝐚,𝐛)V(\mathbf{a},\mathbf{b})\in V^{\prime} we have

dimVπ1(𝐚)=dimVdimπ(V)<dimV=n.\dim V\cap\pi_{\mathbb{C}}^{-1}(\mathbf{a})=\dim V-\dim\pi_{\mathbb{C}}(V)<\dim V=n.

This means that (V𝐚)𝐚π(V)(V_{\mathbf{a}})_{\mathbf{a}\in\pi_{\mathbb{C}}(V^{\prime})} is a family of proper subvarieties of Y(1)n\mathrm{Y}(1)^{n}. Therefore the elements of 𝒮\mathscr{S}^{\star} corresponding to the family (V𝐚)𝐚π(V)(V_{\mathbf{a}})_{\mathbf{a}\in\pi_{\mathbb{C}}(V^{\prime})} are proper special subvarieties of Y(1)n\mathrm{Y}(1)^{n}. As VV is free, the intersection

VT𝒮×TV\cap\bigcup_{T\in\mathscr{S}^{\star}}\mathbb{C}\times T

is contained in a proper subvariety of VV. ∎

Remark 4.17.

If Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} is not definable over ¯\overline{\mathbb{Q}}, then it is immediate by Galois-theoretic reasons that V(¯)V\left(\overline{\mathbb{Q}}\right) cannot be Zariski dense in VV. Since the elements of VΣV_{\Sigma} are all elements of V(¯)V\left(\overline{\mathbb{Q}}\right), then VΣV_{\Sigma} is also not Zariski dense in VV in this case.

Example 4.18.

The condition on the dimension of the fibres in Proposition 4.16 is necessary. For example, consider the variety V2×Y(1)2V\subseteq\mathbb{C}^{2}\times\mathrm{Y}(1)^{2} of dimension 3 defined by the single equation X1=X22X_{1}=X_{2}^{2}. It is clear that VV is modularly free, dimπ(V)=1>0\dim\pi_{\mathbb{C}}(V)=1>0, and VV is defined over ¯\overline{\mathbb{Q}}.

Let τ\tau be any element of Σ\Sigma with positive real part (of which there are infinitely many), and observe that then τ2\tau^{2}\in\mathbb{H}. Since τ2\tau^{2}\notin\mathbb{R} and (τ2)(τ)\mathbb{Q}(\tau^{2})\subseteq\mathbb{Q}(\tau), then 1<[(τ2):][(τ):]=21<[\mathbb{Q}(\tau^{2}):\mathbb{Q}]\leq[\mathbb{Q}(\tau):\mathbb{Q}]=2, thus showing that τ2Σ\tau^{2}\in\Sigma. Therefore, given any two singular moduli b1,b2Y(1)b_{1},b_{2}\in\mathrm{Y}(1), we have that (τ2,τ,b1,b2)VΣ(\tau^{2},\tau,b_{1},b_{2})\in V_{\Sigma}. Since special points are Zariski dense in Y(1)2\mathrm{Y}(1)^{2} and X1=X22X_{1}=X_{2}^{2} defines a curve in 2\mathbb{C}^{2}, the Zariski closure of VΣV_{\Sigma} is equal to VV.

Lemma 4.19.

Let VY(1)n×QV\subseteq\mathrm{Y}(1)^{n}\times Q be a parametric family of irreducible subvarieties of Y(1)n\mathrm{Y}(1)^{n}, and let dd be a positive integer. Let 𝐭(Σ)n\mathbf{t}\in(\mathbb{H}\setminus\Sigma)^{n} be a convenient tuple for jj, and suppose that VV is definable over F=(𝐭,J(𝐭))F=\mathbb{Q}(\mathbf{t},J(\mathbf{t})). Then there is a finite collection of j(𝐭)j(\mathbf{t})-special subvarieties 𝒮\mathscr{S} such that for every 𝐪Q\mathbf{q}\in Q with [F(𝐪):F]d[F(\mathbf{q}):F]\leq d, if V𝐪V_{\mathbf{q}} is a j(𝐭)j(\mathbf{t})-special subvariety, then V𝐪𝒮V_{\mathbf{q}}\in\mathscr{S}.

Proof.

Using definability of dimensions as we did in the proof of Corollary 3.15, we may first restrict to the subfamily QQQ^{\prime}\subseteq Q such that for all 𝐪Q\mathbf{q}\in Q^{\prime} we have that V𝐪V_{\mathbf{q}} is a proper subvariety of Y(1)n\mathrm{Y}(1)^{n}. So we will assume now that VY(1)n×QV\subseteq\mathrm{Y}(1)^{n}\times Q is a parametric family of proper irreducible subvarieties of Y(1)n\mathrm{Y}(1)^{n}.

Let TT be a basic special subvariety of Y(1)n\mathrm{Y}(1)^{n}, and suppose that some translate tr(T,𝐲)\mathrm{tr}(T,\mathbf{y}) is equal to V𝐪V_{\mathbf{q}}, for some 𝐪Q\mathbf{q}\in Q. Then since tr(T,𝐲)\mathrm{tr}(T,\mathbf{y}) is definable over (𝐲)\mathbb{Q}(\mathbf{y}), then so is V𝐪V_{\mathbf{q}}. On the other hand, V𝐪V_{\mathbf{q}} is definable over F(𝐪)F(\mathbf{q}), and so by [EH21, Lemma 7.1] there is a positive integer dd^{\prime} depending only on dd and FF such that if [F(𝐪):F]d[F(\mathbf{q}):F]\leq d and 𝐪\mathbf{q} is algebraic over (j(𝐭))\mathbb{Q}(j(\mathbf{t})), then [(j(𝐭),𝐪):(j(𝐭))]d[\mathbb{Q}(j(\mathbf{t}),\mathbf{q}):\mathbb{Q}(j(\mathbf{t}))]\leq d^{\prime}. So if tr(T,𝐲)=V𝐪\mathrm{tr}(T,\mathbf{y})=V_{\mathbf{q}} and [F(𝐪):F]d[F(\mathbf{q}):F]\leq d, then [(j(𝐭),𝐲):(j(𝐭))]d[\mathbb{Q}(j(\mathbf{t}),\mathbf{y}):\mathbb{Q}(j(\mathbf{t}))]\leq d^{\prime}.

By Remark 2.1 and [Pil11, Theorem 13.2] we have that the set

B={cHe(j(𝐭))j(Σ):[(j(𝐭),c):(j(𝐭))]d}B=\left\{c\in\mathrm{He}(j(\mathbf{t}))\cup j(\Sigma):[\mathbb{Q}(j(\mathbf{t}),c):\mathbb{Q}(j(\mathbf{t}))]\leq d^{\prime}\right\}

is finite. Indeed, by [EH21, Lemma 7.1] we can find a positive integer d′′d^{\prime\prime} depending only on dd^{\prime} and j(𝐭)j(\mathbf{t}) such that if cBj(Σ)c\in B\cap j(\Sigma), then [(c):]d′′[\mathbb{Q}(c):\mathbb{Q}]\leq d^{\prime\prime}, and so [Pil11, Theorem 13.2] applies. Otherwise cHed′′(j(𝐭))c\in\mathrm{He}_{d^{\prime\prime}}(j(\mathbf{t})), and so Remark 2.1 applies directly.

This shows that, for a given basic special variety TT, there are only finitely many tuples 𝐲\mathbf{y} which satisfy the following three conditions

  1. (i)

    tr(T,𝐲)\mathrm{tr}(T,\mathbf{y}) is a j(𝐭)j(\mathbf{t})-special subvariety,

  2. (ii)

    tr(T,𝐲)=V𝐪\mathrm{tr}(T,\mathbf{y})=V_{\mathbf{q}} for some 𝐪Q\mathbf{q}\in Q, and

  3. (iii)

    [F(𝐪):F]d[F(\mathbf{q}):F]\leq d.

We also have that when tr(T,𝐲)=V𝐪\mathrm{tr}(T,\mathbf{y})=V_{\mathbf{q}},

dimV𝐪tr(T,𝐲)>dimV𝐪+tr(T,𝐲)n,\dim V_{\mathbf{q}}\cap\mathrm{tr}(T,\mathbf{y})>\dim V_{\mathbf{q}}+\mathrm{tr}(T,\mathbf{y})-n,

so the intersection is atypical. By [Asl21, Thoerem 4.2] there is a finite collection 𝒮\mathscr{S} of basic special subvarieties such that for every 𝐪Q\mathbf{q}\in Q, every maximal j(𝐭)j(\mathbf{t})-atypical component of V𝐪V_{\mathbf{q}} is a translate of some T𝒮T\in\mathscr{S}. This finishes the proof. ∎

A straightforward adaptation of the proof of Proposition 4.16 gives now the following.

Proposition 4.20.

Suppose that t1,,tmΣt_{1},\ldots,t_{m}\in\mathbb{H}\setminus\Sigma define a convenient tuple for jj, and set F:=(𝐭,J(𝐭))F:=\mathbb{Q}(\mathbf{t},J(\mathbf{t})). Let Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be an irreducible variety which is modularly free and defined over F¯\overline{F}. Suppose that the dimension of the generic fibre of πV\pi_{\mathbb{C}}\upharpoonright_{V} is less than nn. Then the set

VΣ,F:={(𝜶,j(𝜷)):dimG(𝜶|Σ𝐭)=dimG(𝜷|Σ𝐭)=0}V_{\Sigma,F}:=\left\{(\boldsymbol{\alpha},j(\boldsymbol{\beta})):\dim_{G}(\boldsymbol{\alpha}|\Sigma\cup\mathbf{t})=\dim_{G}(\boldsymbol{\beta}|\Sigma\cup\mathbf{t})=0\right\}

is not Zariski dense in VV.

4.6. Results which do not require MZP

In this subsection we review some cases of Theorem 1.1 which do not depend on MZP. We recall that the case n=1n=1 (i.e. plane curves) was proven in [EH21, Theorem 1.2], and does not rely on MZP. We will show that the same is true for n=2n=2. First we need the following lemma.

Lemma 4.21.

Let V×Y(1)V\subseteq\mathbb{C}\times\mathrm{Y}(1) be a free irreducible curve. Let t1,,tmt_{1},\ldots,t_{m}\in\mathbb{H} be a convenient tuple for jj and set F=(𝐭,J(𝐭))F=\mathbb{Q}(\mathbf{t},J(\mathbf{t})). If VV is definable over F¯\overline{F}, then MSCD implies that the set

A(V):={(z,j(z))V:j(z)F¯}A(V):=\left\{(z,j(z))\in V:j(z)\in\overline{F}\right\}

is not Zariski dense in VV.

Proof.

By freeness and irreducibility of VV we have that dimπY(V)=1\dim\pi_{\mathrm{Y}}(V)=1, and so the standard fibres of the restriction πYV\pi_{\mathrm{Y}}\upharpoonright_{V} have dimension 0, so there is a non-empty Zariski open subset UVU\subset V such that if (z,j(z))UA(V)(z,j(z))\in U\cap A(V), then zF¯z\in\overline{F}.

Similarly we have that dimπ(V)=1\dim\pi_{\mathbb{C}}(V)=1. Let dd\in\mathbb{N} be the degree of the restriction πV\pi_{\mathbb{C}}\upharpoonright_{V}. Suppose (z,j(z))UA(V)(z,j(z))\in U\cap A(V), then as zF¯z\in\overline{F}, Lemma 4.8 implies that zz is in the same GG-orbit of some element of Σ𝐭\Sigma\cup\mathbf{t}. Using Proposition 4.16 we can shrink UU if necessary, and so without loss of generality we may assume that zΣz\notin\Sigma, therefore it must be that zz is in the GG-orbit of some element of 𝐭\mathbf{t}. This implies that zFz\in F, and so we get that [F(j(z)):F]d[F(j(z)):F]\leq d. Since j(z)j(z) is modularly dependent over j(𝐭)j(\mathbf{t}), this means that there are only finitely many possible values for j(z)j(z) by Remark 2.1. This finishes the proof. ∎

Theorem 4.22.

Let V2×Y(1)2V\subseteq\mathbb{C}^{2}\times\mathrm{Y}(1)^{2} be a broad and free variety. If VV satisfies (EC), then MSCD implies that VV satisfies (SEC).

Proof.

By Proposition 4.4 we may assume that dimV=2\dim V=2. Let t1,,tmΣt_{1},\ldots,t_{m}\in\mathbb{H}\setminus\Sigma be a convenient tuple for jj, set F=(𝐭,J(𝐭))F=\mathbb{Q}(\mathbf{t},J(\mathbf{t})), and assume that VV is definable over F¯\overline{F}.

By the fibre-dimension theorem, Proposition 4.20 and Lemma 4.21 we can first find a non-empty Zariski open subset UVU\subset V satisfying the following three conditions:

  1. (a)

    If 𝐢=(1)\mathbf{i}=(1) or (2)(2), then for all (a,b)Pr𝐢(U)(a,b)\in\mathrm{Pr}_{\mathbf{i}}(U) we have that dimPr𝐢1((a,b))V=dimVdimPr𝐢(V)\dim\mathrm{Pr}_{\mathbf{i}}^{-1}((a,b))\cap V=\dim V-\dim\mathrm{Pr}_{\mathbf{i}}(V).

  2. (b)

    If (z1,z2,j(z1),j(z2))UEj2(z_{1},z_{2},j(z_{1}),j(z_{2}))\in U\cap\mathrm{E}_{j}^{2}, then there is i{1,2}i\in\{1,2\} such that ziΣG𝐭z_{i}\notin\Sigma\cup G\mathbf{t}.

  3. (c)

    If 𝐢=(1)\mathbf{i}=(1) or (2)(2) and dimPr𝐢(V)=1\dim\mathrm{Pr}_{\mathbf{i}}(V)=1, then for all (z1,z2,j(z1),j(z2))UEj2(z_{1},z_{2},j(z_{1}),j(z_{2}))\in U\cap\mathrm{E}_{j}^{2} we have that pr𝐢(j(z1),j(z2))F¯\mathrm{pr}_{\mathbf{i}}(j(z_{1}),j(z_{2}))\notin\overline{F}.

Claim 4.23.

The set

A(V):={(z1,z2,j(z1),j(z2))VEj2:j(z1)F¯ or j(z2)F¯}A(V):=\left\{(z_{1},z_{2},j(z_{1}),j(z_{2}))\in V\cap\mathrm{E}_{j}^{2}:j(z_{1})\in\overline{F}\mbox{ or }j(z_{2})\in\overline{F}\right\}

is not Zariski dense in VV.

Proof.

Proceed by contradiction, so suppose that A(V)A(V) is Zariski dense in VV. Then for some i{1,2}i\in\{1,2\} we have that the subset

Ai(V):={(z1,z2,j(z1),j(z2))A(V):j(zi)F¯}A_{i}(V):=\{(z_{1},z_{2},j(z_{1}),j(z_{2}))\in A(V):j(z_{i})\in\overline{F}\}

is also Zariski dense in VV. Without loss of generality, we assume that A2(V)A_{2}(V) is Zariski dense in VV. Then we can choose (z1,z2,j(z1),j(z2))UA2(V)(z_{1},z_{2},j(z_{1}),j(z_{2}))\in U\cap A_{2}(V). Also, since A2(V)A_{2}(V) is Zariski dense in VV, then Pi(2)(A2(V))\mathrm{Pi}_{(2)}(A_{2}(V)) is Zariski dense in Pr(2)(V)\mathrm{Pr}_{(2)}(V). By broadness of VV we know that dimPr(2)(V)1\dim\mathrm{Pr}_{(2)}(V)\geq 1, and using condition (c) in the definition of UU we conclude that dimPr(2)(V)=2\dim\mathrm{Pr}_{(2)}(V)=2. Thus generic fibres of the restriction Pr(2)V\mathrm{Pr}_{(2)}\upharpoonright_{V} have dimension 0.

By condition (b) in the definition of UU, we have that either z1ΣG𝐭z_{1}\notin\Sigma\cup G\mathbf{t} or z2ΣG𝐭z_{2}\notin\Sigma\cup G\mathbf{t} (or both). We now show that both cases imply that z2z_{2} is transcendental over FF. If z2ΣG𝐭z_{2}\notin\Sigma\cup G\mathbf{t}, then by Lemma 4.8 we get that tr.deg.FF(z2,j(z2))1\mathrm{tr.deg.}_{F}F(z_{2},j(z_{2}))\geq 1, and since we are assuming that j(z2)F¯j(z_{2})\in\overline{F}, we conclude that z2F¯z_{2}\notin\overline{F}. On the other hand, if z1ΣG𝐭z_{1}\notin\Sigma\cup G\mathbf{t}, then by Lemma 4.8 (which assumes MSCD) we get that tr.deg.FF(z1,j(z1))1\mathrm{tr.deg.}_{F}F(z_{1},j(z_{1}))\geq 1. By condition (a) in the definition of UU, we know that the fibre in VV above (z2,j(z2))(z_{2},j(z_{2})) has dimension 0, and so tr.deg.FF(z2,j(z2))0\mathrm{tr.deg.}_{F}F(z_{2},j(z_{2}))\neq 0, which as before gives that z2F¯z_{2}\notin\overline{F}.

We will now show that we can reduce to the case where both z1z_{1} and z2z_{2} are transcendental over FF. For this we will separate into cases depending on the value of dimπ(V)\dim\pi_{\mathbb{C}}(V) (which is at least 1, by freeness). We first consider the case dimπ(V)=1\dim\pi_{\mathbb{C}}(V)=1. By the freeness of VV there is a Zariski open subset 𝒪π(V)\mathcal{O}\subseteq\pi_{\mathbb{C}}(V) such that for every (x1,x2)𝒪(x_{1},x_{2})\in\mathcal{O} we have that if x2F¯x_{2}\notin\overline{F}, then x1x_{1}\notin\mathcal{F}. Therefore, we may shrink UU if necessary to ensure that both z1,z2F¯z_{1},z_{2}\notin\overline{F}.

Now suppose that dimπ(V)=2=dimV\dim\pi_{\mathbb{C}}(V)=2=\dim V (the following paragraph is just an adapation of an argument present in the proof of [EH21, Proposition 7.4]). We may then shrink UU so that for all (x1,x2,y1,y2)U(x_{1},x_{2},y_{1},y_{2})\in U we have that (see e.g. the definition of triangular varieties in [EH21, §5])

tr.deg.FF(x1,x2,y1,y2)=tr.deg.FF(x1,x2).\mathrm{tr.deg.}_{F}F(x_{1},x_{2},y_{1},y_{2})=\mathrm{tr.deg.}_{F}F(x_{1},x_{2}).

By Lemma 4.8 we get that

tr.deg.FF(z1,z2)=tr.deg.FF(z1,z2,j(z1),j(z2))dimG(z1,z2|Σ,𝐭).\mathrm{tr.deg.}_{F}F(z_{1},z_{2})=\mathrm{tr.deg.}_{F}F(z_{1},z_{2},j(z_{1}),j(z_{2}))\geq\dim_{G}(z_{1},z_{2}|\Sigma,\mathbf{t}).

From this we deduce that

tr.deg.FF(z1,z2)=dimG(z1,z2|Σ,𝐭).\mathrm{tr.deg.}_{F}F(z_{1},z_{2})=\dim_{G}(z_{1},z_{2}|\Sigma,\mathbf{t}).

If z1F¯z_{1}\in\overline{F}, then this equality implies that z1ΣG𝐭z_{1}\in\Sigma\cup G\mathbf{t}. There is a positive integer d1d_{1}, which depends only on the equations defining VV, such that for all (x1,x2,y1,y2)U(x_{1},x_{2},y_{1},y_{2})\in U we have [F(x1,x2,y1):F(x1,x2)]d1[F(x_{1},x_{2},y_{1}):F(x_{1},x_{2})]\leq d_{1}. Then in particular [F(z1,z2,j(z1)):F(z1,z2)]d1[F(z_{1},z_{2},j(z_{1})):F(z_{1},z_{2})]\leq d_{1}, but since z2z_{2} is transcendental over FF, we get [F(z1,j(z1)):F(z1)]d1[F(z_{1},j(z_{1})):F(z_{1})]\leq d_{1}. We can now find another positive integer D1D_{1} depending only on VV and FF such that, if z1Σz_{1}\in\Sigma, then [(z1,j(z1)):(z1)]D1[\mathbb{Q}(z_{1},j(z_{1})):\mathbb{Q}(z_{1})]\leq D_{1} (observe that F¯F\cap\overline{\mathbb{Q}} is a number field). On the other hand, if z1G𝐭z_{1}\in G\mathbf{t}, let us say that zGtiz\in Gt_{i}, then F(ti,j(ti))¯F\cap\overline{\mathbb{Q}(t_{i},j(t_{i}))} is a finite extension of (ti,j(ti))\mathbb{Q}(t_{i},j(t_{i})). So, similarly as to the case z1Σz_{1}\in\Sigma, we get a positive integer D2D_{2} depending only on VV and FF such that, z1Gtiz_{1}\in Gt_{i} for some i{1,,,m}i\in\{1,,\ldots,m\}, then [(z1,ti,j(z1),j(ti)):(z1,ti,j(ti))]D2[\mathbb{Q}(z_{1},t_{i},j(z_{1}),j(t_{i})):\mathbb{Q}(z_{1},t_{i},j(t_{i}))]\leq D_{2}. In either case, we may use Remark 2.1 and [EH21, Lemma 2.1] to conclude that there are only finitely many possible values for j(z1)j(z_{1}). By shrinking UU, we may avoid these values.

So from now on, we assume that z1,z2F¯z_{1},z_{2}\notin\overline{F}. Since j(z2)F¯j(z_{2})\in\overline{F} and VV is free, then (z1,z2,j(z1),j(z2))(z_{1},z_{2},j(z_{1}),j(z_{2})) is not generic in VV over FF. By Lemma 4.8 we have that

2=dimV>tr.deg.FF(z1,z2,j(z1),j(z2))dimG(z1,z2|Σ,𝐭),2=\dim V>\mathrm{tr.deg.}_{F}F(z_{1},z_{2},j(z_{1}),j(z_{2}))\geq\dim_{G}(z_{1},z_{2}|\Sigma,\mathbf{t}),

so we must have that z1Gz2z_{1}\in G\cdot z_{2}. Let dd\in\mathbb{N} be the degree of the restriction Pr(2)V\mathrm{Pr}_{(2)}\upharpoonright_{V}, then

[F(z1,z2,j(z1),j(z2)):F(z2,j(z2))]d.[F(z_{1},z_{2},j(z_{1}),j(z_{2})):F(z_{2},j(z_{2}))]\leq d.

Since z1Gz2z_{1}\in G\cdot z_{2} and z2z_{2} is transcendental over FF, then by [EH21, Lemma 7.1]

[F(z1,z2,j(z1),j(z2)):F(z2,j(z2))]\displaystyle[F(z_{1},z_{2},j(z_{1}),j(z_{2})):F(z_{2},j(z_{2}))] =[F(z2,j(z1),j(z2)):F(j(z2))]\displaystyle=[F(z_{2},j(z_{1}),j(z_{2})):F(j(z_{2}))]
=[F(j(z1),j(z2)):F(j(z2))]\displaystyle=[F(j(z_{1}),j(z_{2})):F(j(z_{2}))]
d.\displaystyle\leq d.

By Remark 2.1 we conclude that there are only finitely many modular polynomials Φ1,,ΦN\Phi_{1},\ldots,\Phi_{N} such that if (z1,z2,j(z1),j(z2))UA2(V)(z_{1},z_{2},j(z_{1}),j(z_{2}))\in U\cap A_{2}(V), then k=1NΦk(j(z1),j(z2))=0\prod_{k=1}^{N}\Phi_{k}(j(z_{1}),j(z_{2}))=0. Since VV is free, we may shrink UU to avoid these finitely many modular polynomials, while still preserving Zariski openness. This contradicts the density of A(V)A(V). ∎

By the Claim and (EC) we can find (z1,z2,j(z1),j(z2))U(z_{1},z_{2},j(z_{1}),j(z_{2}))\in U satisfying j(z1),j(z2)F¯j(z_{1}),j(z_{2})\notin\overline{F} are Zariski dense in VV. By Remark 4.14, this finishes the proof. ∎

In higher dimensions, inspired by the results in [DFT21], we get the following theorem.

Theorem 4.24.

Let Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be variety such that VV projects dominantly both to n\mathbb{C}^{n} and Y(1)n\mathrm{Y}(1)^{n}. Then MSCD implies that VV has (SEC).

Proof.

The proof below is a small adaptation of the one given in [DFT21]. The domination conditions on VV imply that VV is free, broad and satisfies (EC) (by Theorem 3.3). By (the proof of) Proposition 4.4, it suffices to consider the case where dimV=n\dim V=n.

We keep the notation used earlier, so KK, 𝐭\mathbf{t} and FF are as in §4.3. Using the fibre dimension theorem, let VV^{\star} be the Zariski open subset of VV such that for all (𝐱,𝐲)V(\mathbf{x},\mathbf{y})\in V we have that

tr.deg.FF(𝐱,𝐲)=tr.deg.FF(𝐲)=tr.deg.FF(𝐱).\mathrm{tr.deg.}_{F}F(\mathbf{x},\mathbf{y})=\mathrm{tr.deg.}_{F}F(\mathbf{y})=\mathrm{tr.deg.}_{F}F(\mathbf{x}).

Take (𝐳,j(𝐳))V0(\mathbf{z},j(\mathbf{z}))\in V_{0} and assume that

tr.deg.FF(𝐳,j(𝐳))<n.\mathrm{tr.deg.}_{F}F(\mathbf{z},j(\mathbf{z}))<n.

Then by MSCD we know that dimG(𝐳|Σ,𝐭)<n\dim_{G}(\mathbf{z}|\Sigma,\mathbf{t})<n, and so we can choose a Möbius variety MnM\subset\mathbb{C}^{n} defined over (𝐭)¯\overline{\mathbb{Q}(\mathbf{t})}, and a proper j(𝐭)j(\mathbf{t})-special subvariety TT of Y(1)n\mathrm{Y}(1)^{n} such that (𝐳,j(𝐳))M×T(\mathbf{z},j(\mathbf{z}))\in M\times T and dimM=dimT=dimG(𝐳|Σ,𝐭)\dim M=\dim T=\dim_{G}(\mathbf{z}|\Sigma,\mathbf{t}).

As usual, consider the definable family (Vq)qQ(V_{q})_{q\in Q} of Example 3.12 and let VM:=V(M×Y(1)n)V_{M}:=V\cap(M\times\mathrm{Y}(1)^{n}). Observe that by MSCD and the double domination assumption we get

(4.7) tr.deg.FF(𝐳,j(𝐳))=tr.deg.FF(j(𝐳))=tr.deg.FF(j(𝐳))=dimG(𝐳|Σ,𝐭).\mathrm{tr.deg.}_{F}F(\mathbf{z},j(\mathbf{z}))=\mathrm{tr.deg.}_{F}F(j(\mathbf{z}))=\mathrm{tr.deg.}_{F}F(j(\mathbf{z}))=\dim_{G}(\mathbf{z}|\Sigma,\mathbf{t}).

As (𝐳,j(𝐳))VM(\mathbf{z},j(\mathbf{z}))\in V_{M}, then dimVMdimM\dim V_{M}\geq\dim M.

By (4.7), TT is in fact the FF-Zariski closure of j(𝐳)j(\mathbf{z}). So TT is contained in the irreducible component of Zariski closure of πY(VM)\pi_{\mathrm{Y}}(V_{M}) containing j(𝐳)j(\mathbf{z}).

Similarly, (4.7) shows that MM is the FF-Zariski closure of 𝐳\mathbf{z}. So MM is contained in the irreducible component of the Zariski closure of π(VM)\pi_{\mathbb{C}}(V_{M}) containing 𝐳\mathbf{z}. But by definition of VMV_{M} we also have that π(VM)\pi_{\mathbb{C}}(V_{M}) must be contained in MM, so MM is in fact the Zariski closure of π(VM)\pi_{\mathbb{C}}(V_{M}). Since 𝐳\mathbf{z} is generic in MM over FF, it is also generic in π(VM)\pi_{\mathbb{C}}(V_{M}), so by the fibre dimension theorem the dimension of VMπ1(𝐳)V_{M}\cap\pi_{\mathbb{C}}^{-1}(\mathbf{z}) must be equal to dimVMdimπ1(VM)\dim V_{M}-\dim\pi_{\mathbb{C}}^{-1}(V_{M}). By hypothesis dimπ1(𝐳)=0\dim\pi_{\mathbb{C}}^{-1}(\mathbf{z})=0, so dimVM=π1(VM)=dimM\dim V_{M}=\pi_{\mathbb{C}}^{-1}(V_{M})=\dim M.

This shows that dimπY(VM)dimM=dimT\dim\pi_{\mathrm{Y}}(V_{M})\leq\dim M=\dim T, and as we already showed that TT is contained in πY(VM)\pi_{\mathrm{Y}}(V_{M}), then dimπ(VM)=dimT\dim\pi_{\mathbb{C}}(V_{M})=\dim T, and TT must be equal to the irreducible component of Zariski closure of πY(VM)\pi_{\mathrm{Y}}(V_{M}) containing j(𝐳)j(\mathbf{z}). As VV is free, then dimVM<dimV=n\dim V_{M}<\dim V=n, so

dimπY(VM)T=dimT>dimπY(VM)+dimTn.\dim\pi_{\mathrm{Y}}(V_{M})\cap T=\dim T>\dim\pi_{\mathrm{Y}}(V_{M})+\dim T-n.

This shows that TT is a j(𝐭)j(\mathbf{t})-atypical component of πY(VM)T\pi_{\mathrm{Y}}(V_{M})\cap T. As VV is free and satisfies (EC), to complete the proof it suffices to show that there are only finitely many possible options for TT.

Consider the parametric family {Ui}iI\left\{U_{i}\right\}_{i\in I} of the Zariski closures of irreducible components of the members of (πY(V𝐪))𝐪Q\left(\pi_{\mathrm{Y}}(V_{\mathbf{q}})\right)_{\mathbf{q}\in Q}. Let dd be a positive integer so that for every 𝐪Q\mathbf{q}\in Q which is defined over (j(𝐭))\mathbb{Q}(j(\mathbf{t})), we have that the irreducible components of πY(V𝐪)\pi_{\mathrm{Y}}(V_{\mathbf{q}}) are defined over a field LL satisfying [L:F]d[L:F]\leq d. Then there is an integer dd^{\prime}, which only depends on dd and the family {Ui}iI\left\{U_{i}\right\}_{i\in I} with the following property: for every j(𝐭)j(\mathbf{t})-special subvariety SS of Y(1)n\mathrm{Y}(1)^{n}, if SS equals UiU_{i}, where UiU_{i} is an irreducible component of some πY(V𝐪)\pi_{\mathrm{Y}}(V_{\mathbf{q}}) with 𝐪Q\mathbf{q}\in Q definable over (j(𝐭))\mathbb{Q}(j(\mathbf{t})), then for any value cc of a constant coordinate of SS we have that

[(j(𝐭),j(Σ),c):(j(𝐭),j(Σ))]d.[\mathbb{Q}(j(\mathbf{t}),j(\Sigma),c):\mathbb{Q}(j(\mathbf{t}),j(\Sigma))]\leq d^{\prime}.

So by Lemma 4.19 this shows that there are only finitely many possible values that can appear as constant coordinates in SS. Therefore there are only finitely many options for TT. ∎

Remark 4.25.

We observe that one can also get a notion of “convenient tuples for exp\exp” in analogy to our notion of convenient tuples for jj (see [AEK23, §5.2]). Using this, one can then strengthen the main result of [DFT21] by removing the requirement of VV being definable over ¯\overline{\mathbb{Q}}, and instead allowing any finitely generated field of definition like we have done in Theorem 4.24.

In [EH21] it was shown that despite the fact that the composition of jj with itself is not defined on all of \mathbb{H}, we are still able to find solutions to certain equations involving the iterates of jj. More precisely, given a positive integer nn we define inductively

j1:=j and jn+1:=jjn.j_{1}:=j\quad\mbox{ and }\quad j_{n+1}:=j\circ j_{n}.

The domains of these iterates are defined as

1:= and n+1:={zn:j(z)},\mathbb{H}_{1}:=\mathbb{H}\quad\mbox{ and }\quad\mathbb{H}_{n+1}:=\left\{z\in\mathbb{H}_{n}:j(z)\in\mathbb{H}\right\},

so that the natural domain of jnj_{n} is n\mathbb{H}_{n}.

If perhaps unnatural at first, the motivation behind the results of [EH21] concerning iterates of jj was to show that there is some ground to stand on if one wants to consider a dynamical system using jj. While there is a lot of work on the dynamical aspects of meromorphic functions on \mathbb{C}, the same is not true about holomorphic functions whose natural domain is (under the Riemann mapping theorem) the unit disc.

The following result is a generalisation of [EH21, Theorem 1.4], and in particular, it shows that (under MSCD) for every positive integer nn there is znz\in\mathbb{H}_{n} satisfying z=jn(z)z=j_{n}(z) and

tr.deg.(z,j(z),j2(z),,jn1(z))=n.\mathrm{tr.deg.}_{\mathbb{Q}}\mathbb{Q}(z,j(z),j_{2}(z),\ldots,j_{n-1}(z))=n.
Corollary 4.26.

Let Zn+1Z\subset\mathbb{C}^{n+1} be an irreducible hypersurface defined by an irreducible polynomial p[X,Y1,,Yn]p\in\mathbb{C}[X,Y_{1},\ldots,Y_{n}] satisfying pX,pYn0\frac{\partial p}{\partial X},\frac{\partial p}{\partial Y_{n}}\neq 0. Then MSCD implies that for every every finitely generated field KK over which ZZ can be defined, there is znz\in\mathbb{H}_{n} such that (z,j(z),j2(z),,jn(z))(z,j(z),j_{2}(z),\ldots,j_{n}(z)) is generic in ZZ over KK.

Proof.

As explained in [EH21, §8], finding a point of the form (z,j(z),j2(z),,jn(z))(z,j(z),j_{2}(z),\ldots,j_{n}(z)) in ZZ can be done by finding points in VEjnV\cap\mathrm{E}_{j}^{n}, where Vn×Y(1)nV\subset\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} is defined as

V:={X2=Y1X3=Y2Xn=Yn1p(X1,,Xn,Yn)=0}.V:=\left\{\begin{array}[]{rcc}X_{2}&=&Y_{1}\\ X_{3}&=&Y_{2}\\ &\vdots&\\ X_{n}&=&Y_{n-1}\\ p(X_{1},\ldots,X_{n},Y_{n})&=&0\end{array}\right\}.

Under the conditions pYn0\frac{\partial p}{\partial Y_{n}}\neq 0 and pX0\frac{\partial p}{\partial X}\neq 0 we have that dimπ(V)=dimπY(V)=n\dim\pi_{\mathbb{C}}(V)=\dim\pi_{\mathrm{Y}}(V)=n. So by Theorem 4.24 we get that VV has a point (𝐳,j(𝐳))(\mathbf{z},j(\mathbf{z})) which is generic over KK. Since z2=j(z1),,zn=j(zn1)z_{2}=j(z_{1}),\ldots,z_{n}=j(z_{n-1}), then we also obtain a point in ZZ of the desired form which is generic over KK. ∎

5. Unconditional Results

One would like to find varieties Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} having generic points in the graph of the jj-function without having to rely on MSCD or MZP. As we mentioned in the introduction, a few unconditional results have already been obtained: see [AEK23, Theorem 1.1 and §6.2].

Definition.

A broad algebraic variety Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} is said to have no CjC_{j}-factors if for every choice of indices 1i1<<ikn1\leq i_{1}<\ldots<i_{k}\leq n we have that either dimPr𝐢(V)>k\dim\mathrm{Pr}_{\mathbf{i}}(V)>k, or dimPr𝐢(V)=k\dim\mathrm{Pr}_{\mathbf{i}}(V)=k and Pr𝐢(V)\mathrm{Pr}_{\mathbf{i}}(V) is not definable over CjC_{j}.

In particular, if Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} is an irreducible variety of dimension nn with no CjC_{j}-factors, then VV is not definable over CjC_{j}.

The condition of having no CjC_{j}-factors aims at giving a precise notion of what it would mean for a variety to be sufficiently generic with respect to the jj-function. In particular, if (Vq)qQ(V_{q})_{q\in Q} is an algebraic family of free and broad subvarieties of n×Y(1)n\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} such that for every choice of indices 1i1<<ikn1\leq i_{1}<\ldots<i_{k}\leq n we have that the family of projections (Pr𝐢(Vq))qQ(\mathrm{Pr}_{\mathbf{i}}(V_{q}))_{q\in Q} is either generically of dimension larger than kk or non-constant in qq, then this family contains varieties with no CjC_{j}-factors (because CjC_{j} is countable).

It should be observed that the condition of having no CjC_{j}-factors is stronger than simply saying that VV is not defined over CjC_{j}. Indeed, we cannot expect to prove Theorem 1.2 unconditionally under this weaker assumption. For example, suppose that V1,V2×Y(1)V_{1},V_{2}\subset\mathbb{C}\times\mathrm{Y}(1) are two different free plane curves, with V1V_{1} defined over CjC_{j} and V2V_{2} not definable over CjC_{j}. Then the variety V=V1×V22×Y(1)2V=V_{1}\times V_{2}\subset\mathbb{C}^{2}\times\mathrm{Y}(1)^{2} can be easily checked to be free, broad, and not definable over CjC_{j}. If somehow we could prove Theorem 1.2 for this VV, then in particular we would have proven Theorem 1.1 for V1V_{1} unconditionally, thus eliminating the need for MSCD.

5.1. Proof of Theorem 1.2

The proof is inspired by [BK18, Proposition 11.5]. We will set up very similar notation to the one used in §4.3. Let Vn×Y(1)nV\subset\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} be a broad and free variety. Let KK\subset\mathbb{C} be a finitely generated subfield such that VV is defined over KK. Let t1,,tmCjt_{1},\ldots,t_{m}\in\mathbb{H}\setminus C_{j} be given by Theorem 4.11 applied to KK. We assume that dimG(𝐭|Cj)=m\dim_{G}(\mathbf{t}|C_{j})=m.

Choose (𝐱,𝐲)V(\mathbf{x},\mathbf{y})\in V generic over Cj(𝐭,j(𝐭))C_{j}(\mathbf{t},j(\mathbf{t})) (which is possible since Cj(𝐭,j(𝐭))C_{j}(\mathbf{t},j(\mathbf{t})) is a countable field). By the freeness of VV and Lemma 4.1 we know that no coordinate of (𝐱,𝐲)(\mathbf{x},\mathbf{y}) is in Cj(𝐭,j(𝐭))C_{j}(\mathbf{t},j(\mathbf{t})). Let Wn+m×Y(1)n+mW\subseteq\mathbb{C}^{n+m}\times\mathrm{Y}(1)^{n+m} be the CjC_{j}-Zariski closure of the point (𝐱,𝐭,𝐲,j(𝐭))(\mathbf{x},\mathbf{t},\mathbf{y},j(\mathbf{t})).

Lemma 5.1.

WW is broad, free and dimW=dimV+m+dimj(𝐭)\dim W=\dim V+m+\dim^{j}(\mathbf{t}). Furthermore, if dimj(𝐭)>0\dim^{j}(\mathbf{t})>0, then WW is strongly broad.

Proof.

The calculation of the dimension of WW is done in the same way as in Lemma 4.13.

As VV is free, then the coordinates of 𝐱\mathbf{x} are all in distinct GG-orbits and the coordinates of 𝐲\mathbf{y} are modularly independent. On the other hand, as (𝐱,𝐲)(\mathbf{x},\mathbf{y}) is generic over Cj(𝐭,j(𝐭))C_{j}(\mathbf{t},j(\mathbf{t})), then the coordinates of 𝐱\mathbf{x} define different GG-orbits than the coordinates of 𝐭\mathbf{t}. Similarly, every coordinates of 𝐲\mathbf{y} is modularly independent from every coordinate of j(𝐭)j(\mathbf{t}). Since WW is defined over CjC_{j}, then WW cannot have constant coordinates as every coordinate of (𝐱,𝐭,𝐲,j(𝐭))(\mathbf{x},\mathbf{t},\mathbf{y},j(\mathbf{t})) is transcendental over CjC_{j}. By construction of 𝐭\mathbf{t}, the coordinates of 𝐭\mathbf{t} are all in distinct GG-orbits which implies that the coordinates of j(𝐭)j(\mathbf{t}) are modularly independent. Therefore WW is free.

Choose 1i1<<in<k1<<ksn+m1\leq i_{1}<\cdots<i_{\ell}\leq n<k_{1}<\cdots<k_{s}\leq n+m and set

(𝐢,𝐤)=(i1,,i,k1,,ks).(\mathbf{i},\mathbf{k})=(i_{1},\ldots,i_{\ell},k_{1},\ldots,k_{s}).

Write

Pr(𝐢,𝐤)(𝐱,𝐭,𝐲,j(𝐭))=(𝐱𝐢,𝐭𝐤,𝐲𝐢,j(𝐭𝐤)).\mathrm{Pr}_{(\mathbf{i},\mathbf{k})}(\mathbf{x},\mathbf{t},\mathbf{y},j(\mathbf{t}))=(\mathbf{x}_{\mathbf{i}},\mathbf{t}_{\mathbf{k}},\mathbf{y}_{\mathbf{i}},j(\mathbf{t}_{\mathbf{k}})).

Observe that Pr(𝐢,𝐤)(W)\mathrm{Pr}_{(\mathbf{i},\mathbf{k})}(W) is defined over CjC_{j}, and since (𝐱,𝐭,𝐲,j(𝐭))(\mathbf{x},\mathbf{t},\mathbf{y},j(\mathbf{t})) is (by construction) generic in WW over CjC_{j}, then by Proposition 4.7 we get

dimPr𝐢(W)\displaystyle\dim\mathrm{Pr}_{\mathbf{i}}(W) =tr.deg.CjCj(𝐱𝐢,𝐭𝐤,𝐲𝐢,j(𝐭𝐤))\displaystyle=\mathrm{tr.deg.}_{C_{j}}C_{j}(\mathbf{x}_{\mathbf{i}},\mathbf{t}_{\mathbf{k}},\mathbf{y}_{\mathbf{i}},j(\mathbf{t}_{\mathbf{k}}))
=tr.deg.CjCj(𝐭𝐤,j(𝐭𝐤))+tr.deg.Cj(𝐭𝐤,j(𝐭𝐤))Cj(𝐱𝐢,𝐭𝐤,𝐲𝐢,j(𝐭𝐤))\displaystyle=\mathrm{tr.deg.}_{C_{j}}C_{j}(\mathbf{t}_{\mathbf{k}},j(\mathbf{t}_{\mathbf{k}}))+\mathrm{tr.deg.}_{C_{j}(\mathbf{t}_{\mathbf{k}},j(\mathbf{t}_{\mathbf{k}}))}C_{j}(\mathbf{x}_{\mathbf{i}},\mathbf{t}_{\mathbf{k}},\mathbf{y}_{\mathbf{i}},j(\mathbf{t}_{\mathbf{k}}))
dimG(𝐭𝐤|Cj)+dimj(𝐭)+dimPr𝐢(V)\displaystyle\geq\dim_{G}(\mathbf{t}_{\mathbf{k}}|C_{j})+\dim^{j}(\mathbf{t})+\dim\mathrm{Pr}_{\mathbf{i}}(V)
s+.\displaystyle\geq s+\ell.

Therefore WW is broad. From the last inequality we also get that if dimj(𝐭)>0\dim^{j}(\mathbf{t})>0, then WW is strongly broad. ∎

Proof of Theorem 1.2.

By Proposition 4.4 we will assume that dimV=n\dim V=n. We then know that VV is not definable over CjC_{j}, so V(Cj)V(C_{j}) is contained in a proper subvariety of VV. In particular, there is a Zariski open subset V0VV_{0}\subseteq V such that if (𝐳,j(𝐳))V0Ejn(\mathbf{z},j(\mathbf{z}))\in V_{0}\cap\mathrm{E}_{j}^{n}, then some of the coordinates of j(𝐳)j(\mathbf{z}) are not in CjC_{j}. Also, KCjK\not\subset C_{j}, so dimj(𝐭)1\dim^{j}(\mathbf{t})\geq 1.

The case n=1n=1 was proven in [AEK23, Theorem 1.1], so now we assume that n>1n>1. Consider the parametric family of subvarieties (W𝐪)𝐪Q(W_{\mathbf{q}})_{\mathbf{q}\in Q} of WW defined in Example 3.12 and let 𝒮\mathscr{S} be the finite collection of special subvarieties of Y(1)n+m\mathrm{Y}(1)^{n+m} given by Theorem 3.20 applied to this family. Let NN be the integer given by Proposition 3.16 applied to (W𝐪)𝐪Q(W_{\mathbf{q}})_{\mathbf{q}\in Q}.

Let W0WW_{0}\subseteq W be a Zariski open subset defined over CjC_{j} such that the points (𝐚,𝐛)(\mathbf{a},\mathbf{b}) of W0W_{0} satisfy all of the following conditions:

  1. (a)

    The point 𝐛\mathbf{b} does not lie in any T𝒮T\in\mathscr{S}. As WW is free, this condition defines a Zariski open subset of WW.

  2. (b)

    The coordinate of 𝐛\mathbf{b} do not satisfy any of the modular relations Φ1,,ΦN\Phi_{1},\ldots,\Phi_{N}. As WW is free, this condition defines a Zariski open subset of WW.

  3. (c)

    For every 1i1<<in1\leq i_{1}<\cdots<i_{\ell}\leq n, and letting 𝐢=(i1,,i)\mathbf{i}=(i_{1},\ldots,i_{\ell}), we have that

    dim(WPr𝐢1((𝐚,𝐛)))=dimWdimPr𝐢(W).\dim\left(W\cap\mathrm{Pr}_{\mathbf{i}}^{-1}((\mathbf{a},\mathbf{b}))\right)=\dim W-\dim\mathrm{Pr}_{\mathbf{i}}(W).

    By the fibre-dimension theorem and the fact there are only finitely many tuples 𝐢\mathbf{i} to consider, this defines a Zariski open subset of WW.

By the construction of WW, there is a Zariski opens subset V1VV_{1}\subseteq V such that if (𝐱,𝐲)(\mathbf{x},\mathbf{y}) is any point of V1V_{1}, then (𝐱,𝐭,𝐲,j(𝐭))(\mathbf{x},\mathbf{t},\mathbf{y},j(\mathbf{t})) is a point in W0W_{0}.

Choose (𝐳,j(𝐳))V0V1(\mathbf{z},j(\mathbf{z}))\in V_{0}\cap V_{1}, so that (𝐳,𝐭,j(𝐳),j(𝐭))W0(\mathbf{z},\mathbf{t},j(\mathbf{z}),j(\mathbf{t}))\in W_{0}. We will show that dimG(𝐳,𝐭|Cj)=n+m\dim_{G}(\mathbf{z},\mathbf{t}|C_{j})=n+m. For this we proceed by contradiction, so suppose that dimG(𝐳,𝐭|Cj)<n+m\dim_{G}(\mathbf{z},\mathbf{t}|C_{j})<n+m. Let TT be the weakly special subvariety of Y(1)n+m\mathrm{Y}(1)^{n+m} of minimal dimension defined over CjC_{j} for which (j(𝐳),j(𝐭))T(j(\mathbf{z}),j(\mathbf{t}))\in T. Observe that dim(T)=dimG(𝐳,𝐭|Cj)\dim(T)=\dim_{G}(\mathbf{z},\mathbf{t}|C_{j}).

Let Mn+mM\subset\mathbb{C}^{n+m} be a subvariety of minimal dimension defined by Möbius relations defined over \mathbb{Q} and/or setting some coordinates to be a constant in CjC_{j} satisfying (𝐳,𝐭)M(\mathbf{z},\mathbf{t})\in M. Then W(M×Y(1)n+m)W\cap(M\times\mathrm{Y}(1)^{n+m}) is an element of the family (W𝐪)𝐪Q(W_{\mathbf{q}})_{\mathbf{q}\in Q}, call it WMW_{M}. We remark that WMW_{M} is defined over CjC_{j}.

Now let XX be the irreducible component of WM(n+m×T)W_{M}\cap(\mathbb{C}^{n+m}\times T) containing (𝐳,𝐭,j(𝐳),j(𝐭))(\mathbf{z},\mathbf{t},j(\mathbf{z}),j(\mathbf{t})). Observe that XX is defined over CjC_{j}. Then by Proposition 4.7 we get

(5.1) dimXtr.deg.CjCj(𝐳,𝐭,j(𝐳),j(𝐭))dimG(𝐳,𝐭|Cj)+dimj(𝐳𝐭).\dim X\geq\mathrm{tr.deg.}_{C_{j}}C_{j}(\mathbf{z},\mathbf{t},j(\mathbf{z}),j(\mathbf{t}))\geq\dim_{G}(\mathbf{z},\mathbf{t}|C_{j})+\dim^{j}(\mathbf{z}\cup\mathbf{t}).

On the other hand, as WW is free, then dimWM<dimW\dim W_{M}<\dim W, so using Lemma 5.1 we have

(5.2) dimWM+dimn+m×Tdimn+m×Y(1)n+m<dimW+dimTnm=dimj(𝐭)+dimT=dimj(𝐭)+dimG(𝐳,𝐭|Cj)dimX.\begin{array}[]{ccl}\dim W_{M}+\dim\mathbb{C}^{n+m}\times T-\dim\mathbb{C}^{n+m}\times\mathrm{Y}(1)^{n+m}&<&\dim W+\dim T-n-m\\ &=&\dim^{j}(\mathbf{t})+\dim T\\ &=&\dim^{j}(\mathbf{t})+\dim_{G}(\mathbf{z},\mathbf{t}|C_{j})\\ &\leq&\dim X.\end{array}

This shows that XX is an atypical component of WM(n+m×T)W_{M}\cap(\mathbb{C}^{n+m}\times T) in n+m×Y(1)n+m\mathbb{C}^{n+m}\times\mathrm{Y}(1)^{n+m}. If πY(X)\pi_{\mathrm{Y}}(X) has no constant coordinates, then there exists T0𝒮T_{0}\in\mathscr{S} such that XWM(n+m×T0)X\subset W_{M}\cap(\mathbb{C}^{n+m}\times T_{0}). However, this would contradict condition (a) in the definition of W0W_{0}.

So πY(X)\pi_{\mathrm{Y}}(X) has some constant coordinates. Then, as XX is defined over CjC_{j}, those constant coordinates must be given by elements of CjC_{j}. Since no element of j(𝐭)j(\mathbf{t}) is in CjC_{j}, the constant coordinates of πY(X)\pi_{\mathrm{Y}}(X) must be found among the coordinates of j(𝐳)j(\mathbf{z}). Let 1i1<<in1\leq i_{1}<\cdots<i_{\ell}\leq n denote all the coordinates of j(𝐳)j(\mathbf{z}) which are in CjC_{j}. Since (𝐳,j(𝐳))V0(\mathbf{z},j(\mathbf{z}))\in V_{0}, then <n\ell<n.

By Proposition 3.16 we know that there is a weakly special subvariety T0Y(1)n+mT_{0}\subset\mathrm{Y}(1)^{n+m} such that Δb(T0)N\Delta_{b}(T_{0})\leq N, Xn+m×T0X\subseteq\mathbb{C}^{n+m}\times T_{0}, and

(5.3) dimXdimWM(n+m×T0)+dimTT0dimT0.\dim X\leq\dim W_{M}\cap(\mathbb{C}^{n+m}\times T_{0})+\dim T\cap T_{0}-\dim T_{0}.

By condition (b) in the definition of W0W_{0} we know that Δ(T0)=0\Delta(T_{0})=0, which means that T0T_{0} is completely defined by setting certain coordinates to be constant. As the constant coordinates of πY(X)\pi_{Y}(X) are in CjC_{j}, then T0T_{0} is defined over CjC_{j}, and dimT0n+m\dim T_{0}\geq n+m-\ell. But TT is, by definition, the smallest weakly special subvariety of Y(1)n+m\mathrm{Y}(1)^{n+m} which is defined over CjC_{j} and contains the point (j(𝐳),j(𝐭))(j(\mathbf{z}),j(\mathbf{t})). So TT0=TT\cap T_{0}=T. Combining (5.1) and (5.3) we get

(5.4) dimj(𝐭)+dimT0dimWM(n+m×T0).\dim^{j}(\mathbf{t})+\dim T_{0}\leq\dim W_{M}\cap(\mathbb{C}^{n+m}\times T_{0}).

Set Θ:=WPr𝐢1(Pr𝐢(𝐳,𝐭,j(𝐳,𝐭)))\Theta:=W\cap\mathrm{Pr}_{\mathbf{i}}^{-1}\left(\mathrm{Pr}_{\mathbf{i}}(\mathbf{z},\mathbf{t},j(\mathbf{z},\mathbf{t}))\right). By the fibre dimension theorem, condition (c) of the definition of W0W_{0} and the fact that WW is strongly broad (Lemma 5.1) we know that

(5.5) dimΘdimWdimPr𝐢(W)<n+m+dimj(𝐭).\dim\Theta\leq\dim W-\dim\mathrm{Pr}_{\mathbf{i}}(W)<n+m+\dim^{j}(\mathbf{t})-\ell.

Observe that WM(n+m×T0)=W(M×T0)ΘW_{M}\cap(\mathbb{C}^{n+m}\times T_{0})=W\cap(M\times T_{0})\subseteq\Theta, so combining (5.4) and (5.5) gives

dimT0<n+m\dim T_{0}<n+m-\ell

which is a contradiction.

We deduce from this that dimG(𝐳,𝐭|Cj)=n+m\dim_{G}(\mathbf{z},\mathbf{t}|C_{j})=n+m. By [AEK23, Lemma 5.2] this implies that (𝐳,j(𝐳))(\mathbf{z},j(\mathbf{z})) is generic in VV over Cj(𝐭,j(𝐭))C_{j}(\mathbf{t},j(\mathbf{t})). ∎

With this we can get the following “generic” version of Corollary 4.26.

Corollary 5.2.

Let Zn+1Z\subset\mathbb{C}^{n+1} be an irreducible hypersurface defined by an irreducible polynomial p[X,Y1,,Yn]p\in\mathbb{C}[X,Y_{1},\ldots,Y_{n}] satisfying pX,pYn0\frac{\partial p}{\partial X},\frac{\partial p}{\partial Y_{n}}\neq 0. Suppose that ZZ is not definable over CjC_{j}. Then for every every finitely generated field KK over which ZZ can be defined, there is znz\in\mathbb{H}_{n} such that (z,j(z),j2(z),,jn(z))(z,j(z),j_{2}(z),\ldots,j_{n}(z)) is generic in ZZ over KK.

Proof.

Proceed just like in the proof of Corollary 4.26. As we are assuming that ZZ is not definable over CjC_{j}, this will imply that the corresponding variety VV has no CjC_{j}-factors, and so Theorem 1.2 applies. ∎

5.2. Proof of Theorem 1.3

The proof of Theorem 1.3 is done by a straightforward repetition of the proof of Theorem 1.2, and taking into consideration the comments in §4.4, which manifest in the proof of Theorem 1.3 as follows.

  1. (a)

    Proposition 4.7 takes the following form: for every g1,,gnGg_{1},\ldots,g_{n}\in G and every z1,,znnz_{1},\ldots,z_{n}\in\mathbb{H}^{n} we have

    tr.deg.CjCj(z1,,zn,j(g1z1),,j(gnzn))dimG(z1,,zn|Cj)+dimj(z1,,zn).\mathrm{tr.deg.}_{C_{j}}C_{j}(z_{1},\ldots,z_{n},j(g_{1}z_{1}),\ldots,j(g_{n}z_{n}))\geq\dim_{G}(z_{1},\ldots,z_{n}|C_{j})+\dim^{j}(z_{1},\ldots,z_{n}).
  2. (b)

    We can use [AEK23, Corollary 5.4] to obtain the following: if t1,,tmCjt_{1},\ldots,t_{m}\in\mathbb{H}\setminus C_{j} satisfies

    tr.deg.CjCj(𝐭,J(𝐭))=3dimG(𝐭|Cj)+dimj(𝐭)\mathrm{tr.deg.}_{C_{j}}C_{j}(\mathbf{t},J(\mathbf{t}))=3\dim_{G}(\mathbf{t}|C_{j})+\dim^{j}(\mathbf{t})

    (the existence of such tuples is guaranteed by Theorem 4.11), then setting F:=Cj(𝐭,j(𝐭))F:=C_{j}(\mathbf{t},j(\mathbf{t})) we have that for all z1,,znz_{1},\ldots,z_{n}\in\mathbb{H} and all g1,,gnGg_{1},\ldots,g_{n}\in G:

    tr.deg.FF(z1,,zn,j(g1z1),,j(gnzn))dimG(𝐳|Cj𝐭)+dimj(𝐳|𝐭).\mathrm{tr.deg.}_{F}F(z_{1},\ldots,z_{n},j(g_{1}z_{1}),\ldots,j(g_{n}z_{n}))\geq\dim_{G}(\mathbf{z}|C_{j}\cup\mathbf{t})+\dim^{j}(\mathbf{z}|\mathbf{t}).
  3. (c)

    Theorem 3.20 still applies as stated.

6. Results With Derivatives

As explained in [AK22] and [Asl22b] (among other sources), in conjunction with the EC problem for jj, one should also consider the EC problem for jj and its derivatives. In this section we will explain how the methods we have used can be adapted to study the strong EC problem for jj and its derivatives.

6.1. Definitions

We start by setting up some notation. Define Y2(1):=Y(1)×2\mathrm{Y}_{2}(1):=\mathrm{Y}(1)\times\mathbb{C}^{2}. Let EJn:={(𝐳,J(𝐳)):𝐳n}n×Y2(1)n\mathrm{E}_{J}^{n}:=\left\{(\mathbf{z},J(\mathbf{z})):\mathbf{z}\in\mathbb{H}^{n}\right\}\subseteq\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n}.

Let n,n,\ell be positive integers with n\ell\leq n and 𝐢=(i1,,i)\mathbf{i}=(i_{1},\ldots,i_{\ell}) in \mathbb{N}^{\ell} with 1i1<<in1\leq i_{1}<\ldots<i_{\ell}\leq n. Define PR𝐢:n×Y2(1)n×Y2(1)\mathrm{PR}_{\mathbf{i}}:\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n}\rightarrow\mathbb{C}^{\ell}\times\mathrm{Y}_{2}(1)^{\ell} by

PR𝐢:(𝐱,𝐲0,𝐲1,𝐲2)(pr𝐢(𝐱),pr𝐢(𝐲0),pr𝐢(𝐲1),pr𝐢(𝐲2)).\mathrm{PR}_{\mathbf{i}}:(\mathbf{x},\mathbf{y}_{0},\mathbf{y}_{1},\mathbf{y}_{2})\mapsto(\mathrm{pr}_{\mathbf{i}}(\mathbf{x}),\mathrm{pr}_{\mathbf{i}}(\mathbf{y}_{0}),\mathrm{pr}_{\mathbf{i}}(\mathbf{y}_{1}),\mathrm{pr}_{\mathbf{i}}(\mathbf{y}_{2})).

We will abuse slightly some notation we have already introduced define the maps π:n×Y2(1)nn\pi_{\mathbb{C}}:\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n}\to\mathbb{C}^{n} and πY:n×Y2(1)nY(1)n\pi_{\mathrm{Y}}:\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n}\to\mathrm{Y}(1)^{n} as the coordinate projections. Notice that πY\pi_{\mathrm{Y}} still maps onto Y(1)n\mathrm{Y}(1)^{n}, not to Y2(1)n\mathrm{Y}_{2}(1)^{n}. These projections will be used in a very similar way as to how π\pi_{\mathbb{C}} and πY\pi_{\mathrm{Y}} have been used in the previous sections, which is why we have decided to keep the names.

Definition.

An algebraic set Vn×Y2(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n} is said to be JJ-broad if for any 𝐢=(i1,,i)\mathbf{i}=(i_{1},\ldots,i_{\ell}) in \mathbb{N}^{\ell} with 1i1<<in1\leq i_{1}<\ldots<i_{\ell}\leq n we have dimPR𝐢(V)3\dim\mathrm{PR}_{\mathbf{i}}(V)\geq 3\ell. In particular, if VV is JJ-broad then dimV3n\dim V\geq 3n.

We say VV is strongly JJ-broad if the strict inequality dimPR𝐢(V)>3\dim\mathrm{PR}_{\mathbf{i}}(V)>3\ell holds for every 𝐢\mathbf{i}.

Definition.

A subvariety TY2(1)nT\subseteq\mathrm{Y}_{2}(1)^{n} is called a special subvariety of Y2(1)n\mathrm{Y}_{2}(1)^{n} if there is a Möbius subvariety MnM\subseteq\mathbb{C}^{n} defined over \mathbb{Q} such that TT is the Zariski closure over ¯\overline{\mathbb{Q}} of the set J(Mn)J(M\cap\mathbb{H}^{n}). We will say that TT is weakly special if there is a Möbius subvariety MnM\subseteq\mathbb{C}^{n} such that TT is the Zariski closure over \mathbb{C} of the set J(Mn)J(M\cap\mathbb{H}^{n}).

Definition.

We will say that an irreducible constructible set Vn×Y2(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n} is JJ-free if no coordinate of VV is constant, and VV is not contained in any subvariety of the form M×Y2(1)nM\times\mathrm{Y}_{2}(1)^{n} or n×T\mathbb{C}^{n}\times T, where MnM\subset\mathbb{C}^{n} is a proper Möbius subvariety defined over \mathbb{Q}, and TY(1)nT\subset\mathrm{Y}(1)^{n} is a proper special subvariety.

A constructible subset Vn×Y(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}(1)^{n} is JJ-free if every irreducible component of VV is free.

Definition.

We say that an algebraic variety Vn×Y2(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n} satisfies the Existential Closedness condition for JJ, or (EC)J\mathrm{(EC)}_{J}, if the set VEJnV\cap\mathrm{E}_{J}^{n} is Zariski dense in VV.

We say that VV satisfies the Strong Existential Closedness condition for JJ, or (SEC)J\mathrm{(SEC)}_{J}, if for every finitely generated field KK\subset\mathbb{C} over which VV can be defined, there exists (𝐳,J(𝐳))V(\mathbf{z},J(\mathbf{z}))\in V such that (𝐳,J(𝐳))(\mathbf{z},J(\mathbf{z})) is generic in VV over KK.

Conjecture 6.1.

For every positive integer nn, every algebraic variety Vn×Y2(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n} which is JJ-broad and JJ-free, if V(n×Y2(1)n)V\cap\left(\mathbb{H}^{n}\times\mathrm{Y}_{2}(1)^{n}\right) is Zariski dense in VV, then VV satisfies (EC)J\mathrm{(EC)}_{J}. Even more, such VV satisfy (SEC)J\mathrm{(SEC)}_{J}.

As explained in [Asl22a, §5.2], Conjecture 6.1 can be reduced to the case where dimV=3n\dim V=3n by doing an obvious adaptation of Proposition 4.4.

Before continuing to the results with derivatives, in the next few sections we will go over the key ingredients that we need.

6.2. Convenient tuples for JJ

We could start this section by giving a natural definition of convenient generators for JJ, following what we did in §4.2. However this definition would be exactly the same as the definition of convenient generators for jj. To see this we recall that [AEK21, Theorem 1.2] and the results in [AEK23, §5] already include the derivatives of jj. This is manifested in the fact that the various inequalities for jj we showed in §4.2 were all proven by first proving the statement with derivatives. So we already have all the transcendence inequalities we need.

6.3. Weak Zilber–Pink with derivatives

Here we recall one of the main results of [Asl22b].

Definition.

Given a special subvariety SS of Y2(1)n\mathrm{Y}_{2}(1)^{n} and a special subvariety TT of Y(1)n\mathrm{Y}(1)^{n}, we say that SS is associated with TT if πY(S)=T\pi_{\mathrm{Y}}(S)=T.

Definition.

Let VV be an algebraic subvariety of Y2(1)n\mathrm{Y}_{2}(1)^{n}. An atypical component of VV is an irreducible component XX of the intersection between VV and a special subvariety TT of Y2(1)n\mathrm{Y}_{2}(1)^{n} such that

dimX>dimV+dimT3n.\dim X>\dim V+\dim T-3n.

Furthermore, we say that XX is a strongly atypical component of VV if XX is an atypical component of VV and no coordinate is constant on πY(X)\pi_{\mathrm{Y}}(X).

For the definition of upper triangular DD-special subvariety used in the following theorem, see [Asl22b, §6.1]. In particular, we can choose S=Y2(1)nS=\mathrm{Y}_{2}(1)^{n}.

Theorem 6.2 (Uniform weak MZP with derivatives, see [Asl22b, Theorem 7.9]).

Let SS be an upper-triangular DD-special subvariety of Y2(1)n\mathrm{Y}_{2}(1)^{n}. Given a parametric family (V𝐪)𝐪Q(V_{\mathbf{q}})_{\mathbf{q}\in Q} of algebraic subvarieties of Y2(1)n\mathrm{Y}_{2}(1)^{n}, there is a finite collection 𝒮\mathscr{S} of proper special subvarieties of Y(1)n\mathrm{Y}(1)^{n} such that for every 𝐪Q\mathbf{q}\in Q we have that for every strongly atypical component XX of VV there is a special subvariety SS of Y2(1)n\mathrm{Y}_{2}(1)^{n} such that XSX\subseteq S and πY(S)=T\pi_{\mathrm{Y}}(S)=T for some T𝒮T\in\mathscr{S}.

Theorem 6.3.

Given a parametric family (U𝐪)𝐪Q(U_{\mathbf{q}})_{\mathbf{q}\in Q} of constructible subsets of n×Y2(1)n\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n}, there is a finite collection 𝒮\mathscr{S} of of proper special subvarieties of Y(1)n\mathrm{Y}(1)^{n} such that for every 𝐪Q\mathbf{q}\in Q we have that for every strongly atypical component XX of V𝐪V_{\mathbf{q}} there is a special subvariety SS of Y2(1)n\mathrm{Y}_{2}(1)^{n} such that Xn×SX\subseteq\mathbb{C}^{n}\times S and πY(S)=T\pi_{\mathrm{Y}}(S)=T for some T𝒮T\in\mathscr{S}.

We can now go through the same sequence of steps as in §3.4 to obtain the following analogue of Proposition 3.16.

Proposition 6.4.

Given a parametric family (U𝐪)𝐪Q(U_{\mathbf{q}})_{\mathbf{q}\in Q} of constructible subsets of n×Y2(1)n\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n}, there is a positive integer NN such that for every 𝐪Q\mathbf{q}\in Q, for every weakly special subvariety SY2(1)nS\subset\mathrm{Y}_{2}(1)^{n} and for every atypical component XX of U𝐪(n×S)U_{\mathbf{q}}\cap(\mathbb{C}^{n}\times S), there is a proper weakly special subvariety S0Y2(1)nS_{0}\subset\mathrm{Y}_{2}(1)^{n} with Δb(πY(S0))N\Delta_{b}(\pi_{\mathrm{Y}}(S_{0}))\leq N such that Xn×S0X\subseteq\mathbb{C}^{n}\times S_{0}.

6.4. Main Result

Here we present and prove an analogue of Theorem 1.2 which includes derivatives.

Definition.

Let LL be an algebraically closed subfield of \mathbb{C}. A JJ-broad algebraic variety Vn×Y2(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n} is said to have no LL-factors if for every choice of indices 1i1<<ikn1\leq i_{1}<\ldots<i_{k}\leq n we have that either dimPR𝐢(V)>3k\dim\mathrm{PR}_{\mathbf{i}}(V)>3k, or dimPR𝐢(V)=3k\dim\mathrm{PR}_{\mathbf{i}}(V)=3k and PR𝐢(V)\mathrm{PR}_{\mathbf{i}}(V) is not definable over LL.

Theorem 6.5.

Let Vn×Y2(1)nV\subseteq\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n} be a JJ-broad and JJ-free variety with no CjC_{j}-factors satisfying (EC)J\mathrm{(EC)}_{J}. Then for every finitely generated field KK over which VV can be defined, there exists 𝐳n\mathbf{z}\in\mathbb{H}^{n} such that (𝐳,J(𝐳))V(\mathbf{z},J(\mathbf{z}))\in V is generic over KK.

We will set up very similar notation to the one used in §4.3. Let Vn×Y2(1)nV\subset\mathbb{C}^{n}\times\mathrm{Y}_{2}(1)^{n} be a JJ-broad and JJ-free variety with no CjC_{j}-factors. Let KK\subset\mathbb{C} be a finitely generated subfield such that VV is defined over KK. Let t1,,tmCjt_{1},\ldots,t_{m}\in\mathbb{H}\setminus C_{j} be given by Theorem 4.11 applied to KK. By Lemma 4.6, we also assume that dimG(𝐭|Cj)=m\dim_{G}(\mathbf{t}|C_{j})=m.

Choose (𝐱,𝐲0,𝐲1,𝐲2)V(\mathbf{x},\mathbf{y}_{0},\mathbf{y}_{1},\mathbf{y}_{2})\in V generic over Cj(𝐭,J(𝐭))C_{j}(\mathbf{t},J(\mathbf{t})). Let Wn+m×Y2(1)n+mW\subseteq\mathbb{C}^{n+m}\times\mathrm{Y}_{2}(1)^{n+m} be the CjC_{j}-Zariski closure of the point (𝐱,𝐭,𝐲0,𝐲1,𝐲2,J(𝐭))(\mathbf{x},\mathbf{t},\mathbf{y}_{0},\mathbf{y}_{1},\mathbf{y}_{2},J(\mathbf{t})).

Lemma 6.6.

WW is JJ-broad, JJ-free and dimW=3dimV+3m+dimj(𝐭)\dim W=3\dim V+3m+\dim^{j}(\mathbf{t}). Furthermore, if dimj(𝐭)>0\dim^{j}(\mathbf{t})>0, then WW is strongly JJ-broad.

Proof.

Repeat the proof of Lemma 5.1. ∎

Proof of Theorem 6.5.

We assume that dimV=3n\dim V=3n. We then know that VV is not definable over CjC_{j}, so V(Cj)V(C_{j}) is contained in a proper subvariety of VV. In particular, there is a Zariski open subset V0VV_{0}\subseteq V such that if (𝐳,j(𝐳))V0(\mathbf{z},j(\mathbf{z}))\in V_{0}, then some of the coordinates of j(𝐳)j(\mathbf{z}) are not in CjC_{j}. Also, in this case KCjK\not\subset C_{j} so dimj(𝐭)1\dim^{j}(\mathbf{t})\geq 1.

Consider the parametric family of subvarieties (W𝐪)𝐪Q(W_{\mathbf{q}})_{\mathbf{q}\in Q} of WW such that for every 𝐪Q\mathbf{q}\in Q there is a Möbius subvariety M𝐪nM_{\mathbf{q}}\subseteq\mathbb{C}^{n} such that W𝐪:=W(M𝐪×Y2(1)2)W_{\mathbf{q}}:=W\cap(M_{\mathbf{q}}\times\mathrm{Y}_{2}(1)^{2}). Let 𝒮\mathscr{S} be the finite collection of special subvarieties of Y2(1)n+m\mathrm{Y}_{2}(1)^{n+m} given by Theorem 6.3. Let NN be given by Proposition 6.4 applied to (W𝐪)𝐪Q(W_{\mathbf{q}})_{\mathbf{q}\in Q}.

Let W0WW_{0}\subseteq W be a Zariski open subset defined over CjC_{j} such that the points (𝐚,𝐛0,𝐛1,𝐛2)(\mathbf{a},\mathbf{b}_{0},\mathbf{b}_{1},\mathbf{b}_{2}) of W0W_{0} satisfy all of the following conditions:

  1. (a)

    The point 𝐛0\mathbf{b}_{0} does not lie in any T𝒮T\in\mathscr{S}. As WW is free, this condition defines a Zariski open subset of WW.

  2. (b)

    The coordinate of 𝐛0\mathbf{b}_{0} do not satisfy any of the modular relations Φ1,,ΦN\Phi_{1},\ldots,\Phi_{N}. As WW is free, this condition defines a Zariski open subset of WW.

  3. (c)

    For every 1i1<<in1\leq i_{1}<\cdots<i_{\ell}\leq n, and letting 𝐢=(i1,,i)\mathbf{i}=(i_{1},\ldots,i_{\ell}), we have that

    dim(WPR𝐢1((𝐚,𝐛0,𝐛1,𝐛2)))=dimWdimPR𝐢(W).\dim\left(W\cap\mathrm{PR}_{\mathbf{i}}^{-1}((\mathbf{a},\mathbf{b}_{0},\mathbf{b}_{1},\mathbf{b}_{2}))\right)=\dim W-\dim\mathrm{PR}_{\mathbf{i}}(W).

    By the fibre-dimension theorem and the fact there are only finitely many tuples 𝐢\mathbf{i} to consider, this defines a Zariski open subset of WW.

By the construction of WW, there is a Zariski open subset V1VV_{1}\subseteq V such that if (𝐱,𝐲0,𝐲1,𝐲2)(\mathbf{x},\mathbf{y}_{0},\mathbf{y}_{1},\mathbf{y}_{2}) is any point of V1V_{1}, then (𝐱,𝐭,𝐲0,𝐲1,𝐲2,J(𝐭))(\mathbf{x},\mathbf{t},\mathbf{y}_{0},\mathbf{y}_{1},\mathbf{y}_{2},J(\mathbf{t})) is a point in W0W_{0}.

Choose (𝐳,J(𝐳))V0V1(\mathbf{z},J(\mathbf{z}))\in V_{0}\cap V_{1}, so that (𝐳,𝐭,J(𝐳),J(𝐭))W0(\mathbf{z},\mathbf{t},J(\mathbf{z}),J(\mathbf{t}))\in W_{0}. We will show that dimG(𝐳,𝐭|Cj)=n+m\dim_{G}(\mathbf{z},\mathbf{t}|C_{j})=n+m. For this we proceed by contradiction, so suppose that dimG(𝐳,𝐭|Cj)<n+m\dim_{G}(\mathbf{z},\mathbf{t}|C_{j})<n+m. Let SS be the weakly special subvariety of Y2(1)n+m\mathrm{Y}_{2}(1)^{n+m} of minimal dimension defined over CjC_{j} for which (J(𝐳),J(𝐭))T(J(\mathbf{z}),J(\mathbf{t}))\in T. Observe that dim(T)=3dimG(𝐳,𝐭|Cj)\dim(T)=3\dim_{G}(\mathbf{z},\mathbf{t}|C_{j}).

Let Mn+mM\subset\mathbb{C}^{n+m} be the Möbius subvariety of smallest dimension defined by Möbius relation over \mathbb{Q} and/or setting some coordinates to be a constant in CjC_{j}, which satisfies (𝐳,𝐭)M(\mathbf{z},\mathbf{t})\in M. Then W(M×Y2(1)n+m)W\cap(M\times\mathrm{Y}_{2}(1)^{n+m}) is an element of the family (W𝐪)𝐪Q(W_{\mathbf{q}})_{\mathbf{q}\in Q}, call it WMW_{M}. We remark that WMW_{M} is defined over CjC_{j}.

Now let XX be the irreducible component of WM(n+m×S)W_{M}\cap(\mathbb{C}^{n+m}\times S) containing (𝐳,𝐭,J(𝐳),J(𝐭))(\mathbf{z},\mathbf{t},J(\mathbf{z}),J(\mathbf{t})). Observe that XX is defined over CjC_{j}. Then by Proposition 4.7 we get

(6.1) dimXtr.deg.CjCj(𝐳,𝐭,J(𝐳),J(𝐭))3dimG(𝐳,𝐭|Cj)+dimj(𝐳,𝐭).\dim X\geq\mathrm{tr.deg.}_{C_{j}}C_{j}(\mathbf{z},\mathbf{t},J(\mathbf{z}),J(\mathbf{t}))\geq 3\dim_{G}(\mathbf{z},\mathbf{t}|C_{j})+\dim^{j}(\mathbf{z},\mathbf{t}).

On the other hand, as WW is free, then dimWM<dimW\dim W_{M}<\dim W, so using Lemma 5.1 we have

(6.2) dimWM+dimn+m×Sdimn+m×Y2(1)n+m<dimW+dimTnm=dimj(𝐭)+dimS=dimj(𝐭)+3dimG(𝐳,𝐭|Cj)dimX.\begin{array}[]{ccl}\dim W_{M}+\dim\mathbb{C}^{n+m}\times S-\dim\mathbb{C}^{n+m}\times\mathrm{Y}_{2}(1)^{n+m}&<&\dim W+\dim T-n-m\\ &=&\dim^{j}(\mathbf{t})+\dim S\\ &=&\dim^{j}(\mathbf{t})+3\dim_{G}(\mathbf{z},\mathbf{t}|C_{j})\\ &\leq&\dim X.\end{array}

This shows that XX is an atypical component of WM(n+m×S)W_{M}\cap(\mathbb{C}^{n+m}\times S) in n+m×Y2(1)n+m\mathbb{C}^{n+m}\times\mathrm{Y}_{2}(1)^{n+m}. If πY(X)\pi_{\mathrm{Y}}(X) has no constant coordinates, then by 6.3 there exists a special subvariety SY2(1)n+mS\subseteq\mathrm{Y}_{2}(1)^{n+m} and T0𝒮T_{0}\in\mathscr{S} such that XWM(n+m×S)X\subset W_{M}\cap(\mathbb{C}^{n+m}\times S) and πY(S)=T0\pi_{\mathrm{Y}}(S)=T_{0}. However, this would contradict condition (a) in the definition of W0W_{0}.

So πY(X)\pi_{\mathrm{Y}}(X) has some constant coordinates. Then, as XX is defined over CjC_{j}, those constant coordinates must be given by elements of CjC_{j}. Since no element of J(𝐭)J(\mathbf{t}) is in CjC_{j} (by Lemma 4.6), the constant coordinates of πY(X)\pi_{\mathrm{Y}}(X) must be found among the coordinates of J(𝐳)J(\mathbf{z}). Let 1i1<<in1\leq i_{1}<\cdots<i_{\ell}\leq n denote all the coordinates of j(𝐳)j(\mathbf{z}) which are in CjC_{j}. Recall that by Lemma 4.6, if some coordinate of (z,j(z),j(z),j′′(z))(z,j(z),j^{\prime}(z),j^{\prime\prime}(z)) is in CjC_{j}, then they all are. Since (𝐳,J(𝐳))V0(\mathbf{z},J(\mathbf{z}))\in V_{0}, then <n\ell<n. We also remark that at this point we have already proven the theorem for the case n=1n=1.

By Proposition 3.16 we know that there is a weakly special subvariety S0Y2(1)n+mS_{0}\subset\mathrm{Y}_{2}(1)^{n+m} such that Δb(πY(S0))N\Delta_{b}(\pi_{\mathrm{Y}}(S_{0}))\leq N, Xn+m×S0X\subseteq\mathbb{C}^{n+m}\times S_{0}, and

(6.3) dimXdimWM(n+m×S0)+dimTS0dimS0.\dim X\leq\dim W_{M}\cap(\mathbb{C}^{n+m}\times S_{0})+\dim T\cap S_{0}-\dim S_{0}.

By condition (b) in the definition of W0W_{0} we know that Δb(πY(S0))=0\Delta_{b}(\pi_{\mathrm{Y}}(S_{0}))=0, which means that πY(S0)\pi_{\mathrm{Y}}(S_{0}) is completely defined by setting certain coordinates to be constant. As the constant coordinates of πY(X)\pi_{\mathrm{Y}}(X) are in CjC_{j}, then S0S_{0} is defined over CjC_{j}, and dimS03(n+m)\dim S_{0}\geq 3(n+m-\ell) (again by Lemma 4.6). But SS is, by definition, the smallest weakly special subvariety of Y2(1)n+m\mathrm{Y}_{2}(1)^{n+m} which is defined over CjC_{j} and contains the point (J(𝐳),J(𝐭))(J(\mathbf{z}),J(\mathbf{t})). So SS0=SS\cap S_{0}=S. Combining (6.1) and (6.3) we get

(6.4) dimj(𝐭)+dimS0dimWM(n+m×S0).\dim^{j}(\mathbf{t})+\dim S_{0}\leq\dim W_{M}\cap(\mathbb{C}^{n+m}\times S_{0}).

Set Θ:=WPR𝐢1(PR𝐢(𝐳,𝐭,j(𝐳,𝐭)))\Theta:=W\cap\mathrm{PR}_{\mathbf{i}}^{-1}\left(\mathrm{PR}_{\mathbf{i}}(\mathbf{z},\mathbf{t},j(\mathbf{z},\mathbf{t}))\right). By the fibre dimension theorem, condition (c) of the definition of W0W_{0} and the fact that WW is strongly broad (Lemma 6.6) we know that

(6.5) dimΘdimWdimPR𝐢(W)<3(n+m).\dim\Theta\leq\dim W-\dim\mathrm{PR}_{\mathbf{i}}(W)<3(n+m-\ell).

Observe that WM(n+m×S0)=W(M×S0)ΘW_{M}\cap(\mathbb{C}^{n+m}\times S_{0})=W\cap(M\times S_{0})\subseteq\Theta, so combining (6.4) and (6.5) gives

dimS0<3(n+m)\dim S_{0}<3(n+m-\ell)

which is a contradiction.

We deduce from this that dimG(𝐳,𝐭|Cj)=3(n+m)\dim_{G}(\mathbf{z},\mathbf{t}|C_{j})=3(n+m). By [AEK23, Lemma 5.2] this implies that (𝐳,J(𝐳))(\mathbf{z},J(\mathbf{z})) is generic in VV over Cj(𝐭,J(𝐭))C_{j}(\mathbf{t},J(\mathbf{t})). ∎

We have not added here an analogue of Theorem 1.3 including derivatives because [AK22, Theorem 1.8], which is the available result on EC for the blurring of JJ, requires one to use a group larger than GG to define the blurring. So the result we can obtain would not be as close an approximation to Conjecture 6.1 as Theorem 1.3 is to Conjecture 4.2.

Availability of data and material

Not applicable.

Declaration

The author states that there is no conflict of interest.

References

  • [AEK21] Vahagn Aslanyan, Sebastian Eterović, and Jonathan Kirby. Differential existential closedness for the jj-function. Proc. Amer. Math. Soc., 149(4):1417–1429, 2021. https://doi.org/10.1090/proc/15333.
  • [AEK23] Vahagn Aslanyan, Sebastian Eterović, and Jonathan Kirby. A closure operator respecting the modular jj-function. Israel J. Math., 253(1):321–357, 2023. https://doi.org/10.1007/s11856-022-2362-y.
  • [AK22] Vahagn Aslanyan and Jonathan Kirby. Blurrings of the JJ-function. Q. J. Math., 73(2):461–475, 2022. https://doi.org/10.1093/qmath/haab037.
  • [Asl21] Vahagn Aslanyan. Some remarks on atypical intersections. Proc. Amer. Math. Soc., 149(11):4649–4660, 2021. https://doi.org/10.1090/proc/15611.
  • [Asl22a] Vahagn Aslanyan. Adequate predimension inequalities in differential fields. Ann. Pure Appl. Logic, 173(1):Paper No. 103030, 42, 2022. https://doi.org/10.1016/j.apal.2021.103030.
  • [Asl22b] Vahagn Aslanyan. Weak modular Zilber-Pink with derivatives. Math. Ann., 383(1-2):433–474, 2022.
  • [Ax72] James Ax. Some topics in differential algebraic geometry. I. Analytic subgroups of algebraic groups. Amer. J. Math., 94:1195–1204, 1972.
  • [BK18] Martin Bays and Jonathan Kirby. Pseudo-exponential maps, variants, and quasiminimality. Algebra Number Theory, 12(3):493–549, 2018. https://doi.org/10.2140/ant.2018.12.493.
  • [BSCFN21] D. Blázquez-Sanz, C. Casale, J. Freitag, and J. Nagloo. A differential approach to the Ax-Schanuel, I, 2021. Preprint: https://arxiv.org/abs/2102.03384.
  • [Chi21] Kenneth Chung Tak Chiu. Ax-Schanuel with derivatives for mixed period mappings, 2021. Preprint: https://arxiv.org/abs/2110.03489.
  • [DFT21] Paola D’Aquino, Antongiulio Fornasiero, and Giuseppina Terzo. A weak version of the strong exponential closure. Israel J. Math., 242(2):697–705, 2021. https://doi.org/10.1007/s11856-021-2141-1.
  • [DMT10] P. D’Aquino, A. Macintyre, and G. Terzo. Schanuel Nullstellensatz for Zilber fields. Fund. Math., 207(2):123–143, 2010. http://eudml.org/doc/283098.
  • [EH21] Sebastian Eterović and Sebastián Herrero. Solutions of equations involving the modular jj function. Trans. Amer. Math. Soc., 374(6):3971–3998, 2021. https://doi.org/10.1090/tran/8244.
  • [ES22] Sebastian Eterović and Thomas Scanlon. Likely intersections. https://math.berkeley.edu/~scanlon/papers/li-16-nov-22.pdf, 2022.
  • [Ete18] Sebastian Eterović. A Schanuel property for jj. Bull. Lond. Math. Soc., 50(2):293–315, 2018. https://doi.org/10.1112/blms.12135.
  • [EZ21] Sebastian Eterović and Roy Zhao. Algebraic varieties and automorphic functions. https://arxiv.org/abs/2107.10392, 2021.
  • [Gal21] Francesco Paolo Gallinaro. Solving systems of equations of raising-to-powers type. To appear in Israel J. Math. Preprint: https://arxiv.org/abs/2103.15675, 2021.
  • [GK23] Ziyang Gao and Bruno Klingler. The Ax-Schanuel conjecture for variations of mixed Hodge structures. Math. Ann., 2023. https://doi.org/10.1007/s00208-023-02614-w.
  • [HP16] Philipp Habegger and Jonathan Pila. O-minimality and certain atypical intersections. Ann. Sci. Éc. Norm. Supér. (4), 49(4):813–858, 2016. https://doi.org/10.24033/asens.2296.
  • [Kir09] Jonathan Kirby. The theory of the exponential differential equations of semiabelian varieties. Selecta Math. (N.S.), 15(3):445–486, 2009. https://doi.org/10.1007/s00029-009-0001-7.
  • [KZ14] Jonathan Kirby and Boris Zilber. Exponentially closed fields and the conjecture on intersections with tori. Ann. Pure Appl. Logic, 165(11):1680–1706, 2014. https://doi.org/10.1016/j.apal.2014.06.002.
  • [Lan87] Serge Lang. Elliptic functions, volume 112 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1987. With an appendix by J. Tate, https://doi.org/10.1007/978-1-4612-4752-4.
  • [Mah69] Kurt Mahler. On algebraic differential equations satisfied by automorphic functions. J. Austral. Math. Soc., 10:445–450, 1969.
  • [MPT19] Ngaiming Mok, Jonathan Pila, and Jacob Tsimerman. Ax-Schanuel for Shimura varieties. Ann. of Math. (2), 189(3):945–978, 2019. https://doi.org/10.4007/annals.2019.189.3.7.
  • [Pil11] Jonathan Pila. O-minimality and the André-Oort conjecture for n\mathbb{C}^{n}. Ann. of Math. (2), 173(3):1779–1840, 2011. https://doi.org/10.4007/annals.2011.173.3.11.
  • [Pil15] Jonathan Pila. Functional transcendence via o-minimality. In O-minimality and diophantine geometry, volume 421 of London Math. Soc. Lecture Note Ser., pages 66–99. Cambridge Univ. Press, Cambridge, 2015.
  • [Pil22] Jonathan Pila. Point-counting and the Zilber-Pink conjecture, volume 228 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2022. https://doi.org/10.1017/9781009170314.
  • [PS21] Jonathan Pila and Thomas Scanlon. Effective transcendental Zilber-Pink for variations of Hodge structures. https://arxiv.org/abs/2105.05845, 2021.
  • [PT16] Jonathan Pila and Jacob Tsimerman. Ax-Schanuel for the jj-function. Duke Math. J., 165(13):2587–2605, 2016. https://doi.org/10.1215/00127094-3620005.
  • [Sch37] Theodor Schneider. Arithmetische Untersuchungen elliptischer Integrale. Math. Ann., 113(1):1–13, 1937. https://doi.org/10.1007/BF01571618.
  • [Zil05] B. Zilber. Pseudo-exponentiation on algebraically closed fields of characteristic zero. Ann. Pure Appl. Logic, 132(1):67–95, 2005. https://doi.org/10.1016/j.apal.2004.07.001.