return

Celebratio Mathematica

Walter D. Neumann

Hyperbolic 3-manifolds, the Bloch group,
and the work of Walter Neumann

by Stavros Garoufalidis and Don Zagier

1. Ideal triangulations and the gluing equations

The start­ing point of the pa­per [1] was Thur­ston’s amaz­ing in­sight in the 1980’s that all 3-di­men­sion­al man­i­folds should be ca­non­ic­ally di­vis­ible in­to pieces hav­ing a well-defined geo­met­ric struc­ture of one of eight types, the most im­port­ant of which is the hy­per­bol­ic one. In con­junc­tion with the fam­ous Mostow ri­gid­ity the­or­em, this means that 3-di­men­sion­al to­po­logy be­comes a part, first of dif­fer­en­tial geo­metry, and then of al­geb­ra­ic num­ber the­ory, something that is not at all the case in oth­er di­men­sions. The main class is that of ori­ented hy­per­bol­ic 3-man­i­folds, which have a rieman­ni­an met­ric with con­stant neg­at­ive curvature that can be nor­mal­ized to \( -1 \) and hence are loc­ally iso­met­ric to hy­per­bol­ic 3-space \( \mathbb{H}^3 \). Of par­tic­u­lar in­terest is the volume spec­trum, the set of volumes of all com­plete hy­per­bol­ic 3-man­i­folds of fi­nite volume (in which case they are either com­pact or the uni­on of a com­pact part and a fi­nite num­ber of “cusps” dif­feo­morph­ic to the product of a half-line and a tor­us). These volumes have both strik­ing num­ber-the­or­et­ic­al prop­er­ties (they be­long to the im­age un­der the reg­u­lat­or map of the Bloch group, as dis­cussed in Sec­tion 3) and strik­ing met­ric prop­er­ties (they form a count­able well-ordered sub­set of \( \mathbb{R}_{ > 0} \), as dis­cussed in Sec­tion 2), and the primary goal of the pa­per [1] was to un­der­stand them as thor­oughly as pos­sible.

In this sec­tion we dis­cuss ideal tri­an­gu­la­tions and their NZ-equa­tions in some de­tail. Ideal tri­an­gu­la­tions of 3-man­i­folds with tor­us bound­ary com­pon­ents were in­tro­duced by Thur­ston [e7] as a con­veni­ent way to de­scribe and ef­fect­ively com­pute [e24] com­plete hy­per­bol­ic struc­tures on 3-man­i­folds. Re­call that in hy­per­bol­ic geo­metry an ideal tet­ra­hed­ron is the con­vex hull of four points in the bound­ary \( \partial (\mathbb{H}^3)\cong\mathbb{P}^1(\mathbb{C}) \) of 3-di­men­sion­al hy­per­bol­ic space \( \mathbb{H}^3 \). The (ori­ent­a­tion-pre­serving) iso­metry group of \( \mathbb{H}^3 \) is the group \( \operatorname{PSL}_2(\mathbb{C}) \), act­ing on the bound­ary by frac­tion­al lin­ear trans­form­a­tions of \( \mathbb{P}^1(\mathbb{C}) \), and since un­der this group any four dis­tinct points can be put in stand­ard po­s­i­tion \( (0,1,\infty,z) \) for some \( z\in\mathbb{C}\smallsetminus\{0,1\} \) (the cross-ra­tio of the four points), any (ori­ented) ideal tet­ra­hed­ron is the con­vex hull \( \Delta(z) \) for some com­plex num­ber \( z~\in \mathbb{C}\smallsetminus\{0,1\} \), called the shape para­met­er of the tet­ra­hed­ron. This num­ber is not quite unique be­cause of the choice of which three ver­tices we send to 0, 1 and \( \infty \), mean­ing that the ori­ented tet­ra­hedra \( \Delta(z) \), \( \Delta(z^{\prime}) \) and \( \Delta(z^{\prime\prime}) \), where \( z^{\prime}=1/(1-z) \) and \( z^{\prime\prime}=1-1/z \), are iso­met­ric. In this way one at­taches a shape para­met­er \( z \), \( z^{\prime} \) or \( z^{\prime\prime} \) to each pair of op­pos­ite edges of a giv­en ori­ented ideal tet­ra­hed­ron. Note that the shape of a tet­ra­hed­ron is an ar­bit­rary com­plex num­ber not equal to 0 or 1, and wheth­er it has pos­it­ive, zero or neg­at­ive ima­gin­ary part is ir­rel­ev­ant to our dis­cus­sion.

When \( \mathcal{T} \) is an ideal tri­an­gu­la­tion of a hy­per­bol­ic 3-man­i­fold \( M \) with \( N \) tet­ra­hedra with shape para­met­ers \( z_1,\dots,z_N \), then we get one poly­no­mi­al equa­tion (“glu­ing equa­tion”) for each edge. Spe­cific­ally, the shape para­met­ers at that edge of all tet­ra­hedra that are in­cid­ent with it must clearly have ar­gu­ments that add up to \( 2\pi \) (be­cause oth­er­wise the met­ric would not be smooth along that edge), but in fact the shape para­met­ers them­selves have product \( +1 \) by an easy ar­gu­ment. Since each of the pos­sible shape para­met­ers \( z \), \( z^{\prime} \), \( z^{\prime\prime} \) of \( \Delta(z) \) be­longs to the mul­ti­plic­at­ive group \( \langle z,1-z,-1\rangle \), this equa­tion for each edge \( e_i \) has the form \begin{equation} \pm\prod_{j=1}^N z_j^{R^{\prime}_{ij}} (1-z_j)^{R^{\prime\prime}_{ij}} = 1. \end{equation}

Since it is eas­ily seen that the num­ber of edges in the tri­an­gu­la­tion is the same as the num­ber \( N \) of sim­plices, this gives us \( N \) poly­no­mi­al equa­tions among the \( N \) com­plex num­bers \( z_1,\dots,z_N \). The ob­vi­ous thought is that this leads to a 0-di­men­sion­al mod­uli space and ex­plains the ri­gid­ity, but this is wrong since ri­gid­ity only ap­plies to com­plete hy­per­bol­ic struc­tures, and in fact the \( N \) edge re­la­tions are nev­er mul­ti­plic­at­ively in­de­pend­ent. An ob­vi­ous ex­ample is that their product is al­ways 1, and this is the only de­pend­ence if the bound­ary of the 3-man­i­fold is a tor­us (of­ten called a cusp), but in gen­er­al there are \( h \) mul­ti­plic­at­ively in­de­pend­ent re­la­tions among the \( N \) equa­tions (1), where \( h \) is the num­ber of cusps of the 3-man­i­fold (as­sum­ing that all bound­ary com­pon­ents are tori). Thus the true ex­pec­ted di­men­sion of the mod­uli space of hy­per­bol­ic struc­tures is in fact \( h \). That the di­men­sion really is \( h \) is an im­port­ant the­or­em of Thur­ston (§5 of [e7]) and was giv­en a new and sim­pler proof in [1] us­ing the al­geb­ra­ic struc­ture of the glu­ing equa­tions. The 0-di­men­sion­al mod­uli space (ri­gid­ity) arises when we re­quire the hy­per­bol­ic struc­ture giv­en by the shape para­met­ers \( z_i \) to be com­plete, be­cause this leads to two fur­ther in­de­pend­ent re­la­tions at each cusp. Spe­cific­ally, each “peri­pher­al curve” (mean­ing an iso­topy class of curves on the tor­us cross-sec­tion of one of the cusps) gives a re­la­tion, so if one chooses a me­ridi­an and a lon­git­ude \( (\mu_i,\lambda_i) \) at each cusp \( i=1,\dots,h \), one ob­tains \( 2h \) fur­ther equa­tions \begin{equation} \label{ML} \pm\prod_{j=1}^N z_j^{M^{\prime}_{ij}} (1-z_j)^{M^{\prime\prime}_{ij}} = 1, \quad \pm\prod_{j=1}^N z_j^{L^{\prime}_{ij}} (1-z_j)^{L^{\prime\prime}_{ij}} = 1 \quad (i=1,\dots,h). \end{equation} Thus the full set of glu­ing equa­tions is de­scribed by an \( (N+2h) \times 2N \) mat­rix \[ U= \begin{pmatrix} R^{\prime}&R^{\prime\prime}\\M^{\prime}&M^{\prime\prime}\\ L^{\prime}&L^{\prime\prime} \end{pmatrix}. \]

It was shown in [1] that the mat­rix \( U \) has some key sym­plect­ic prop­er­ties, of which the most im­port­ant (The­or­em 2.2) says that \begin{equation} U J_{2N}U^t = \biggl(\,\begin{matrix} 0_{N,N}& 0_{2h,N} \\ 0_{N,2h} & 2 J_{2h} \end{matrix}\biggr), \end{equation} where \[ J_{2n}:= \biggl(\,\begin{matrix} 0\phantom{_n}&-1_n\\ \smash[b]{1_n}&\phantom{-}0\phantom{_n} \end{matrix}\biggr), \] while the oth­ers say that the \( N\times2N \) mat­rix \( R=(R^{\prime}\,R^{\prime\prime}) \) has rank \( N-h \), the full mat­rix \( U \) has rank \( N+h \), and the space \( [U]\subset\mathbb{R}^{2N} \) spanned by the rows of \( U \) is the or­tho­gon­al com­ple­ment of \( [R] \) with re­spect to the sym­plect­ic struc­ture \( J_{2N} \) (Pro­pos­i­tion 2.3). The rank state­ment was then used to prove that the di­men­sion of the above-men­tioned de­form­a­tion space of (non­com­plete) hy­per­bol­ic struc­tures is at least \( h \), and then a fur­ther ar­gu­ment us­ing Mostow ri­gid­ity showed that it is ex­actly \( h \). This gives rise in the 1-cusp case to a poly­no­mi­al \( P(m,\ell)=0 \), where \( m^2 \) and \( \ell^2 \) de­note the left-hand sides of the two equa­tions (2) and \( P \) is a cer­tain poly­no­mi­al (a factor of what is now called the \( A \)-poly­no­mi­al) that was cal­cu­lated ex­pli­citly in [1] for the simplest hy­per­bol­ic knot \( 4_1 \) (fig­ure-8). In the gen­er­al case the de­form­a­tion space is a com­pon­ent of the “char­ac­ter vari­ety” as in­tro­duced and stud­ied in [e4] and stud­ied later by Wal­ter and his stu­dents Ab­hijit Cham­pan­erkar [e11] and Stefan Till­mann [e12].

To get a com­plete man­i­fold with \( h \) cusps, we im­pose \( h \) fur­ther glu­ing con­di­tions, namely the equa­tions (1) and the first of each of the equa­tions (2). (But in fact these \( h \) me­ridi­an equa­tions to­geth­er with the edge equa­tions im­ply the lon­git­ude equa­tions, so in the end all re­la­tions (1) and (2) hold.) Now ri­gid­ity ap­plies and there are no de­form­a­tions. But there is an­oth­er pro­cess, with a lot more free­dom, to ob­tain ri­gid com­plete hy­per­bol­ic struc­tures from the ori­gin­al cusped man­i­fold, namely to do a Dehn sur­gery at some or all of the cusps. Spe­cific­ally, when we do a \( (p,q) \)-sur­gery at a cusp (mean­ing that we trun­cate the 3-man­i­fold at the tor­us bound­ary and glue on a sol­id tor­us in such a way as to kill the ho­mo­topy class of \( p \) times a chosen me­ridi­an times \( q \) times a chosen lon­git­ude), then we im­pose the glu­ing equa­tion \( m^{2p}\ell^{2q}=1 \), where \( m^2 \) and \( \ell^2 \) as above are the left-hand sides of one of the pairs of equa­tions (2). In the next sec­tion, still fol­low­ing [1], we dis­cuss how the volumes be­have un­der this pro­cess.

2. Volumes and Dehn surgeries

As already stated, the ori­gin­al main pur­pose of [1] was to study the arith­met­ic of the set of volumes of all hy­per­bol­ic 3-man­i­folds. It had been ob­served by Thur­ston, us­ing earli­er work of Jørgensen, that this volume spec­trum is a well-ordered sub­set of the pos­it­ive reals. In oth­er words, there is a smal­lest volume (which is known), a second smal­lest, a third smal­lest, …, then a smal­lest lim­it point, a second smal­lest lim­it point, …, then lim­its of these, and so on. The proof of well-ordered­ness shows that these simple and high­er-or­der lim­it points arise by “clos­ing up” one or more of the cusps of a non­com­pact hy­per­bol­ic 3-man­i­fold \( M \) by Dehn sur­ger­ies to ob­tain a count­able col­lec­tion of man­i­folds with few­er cusps whose volumes tend from be­low to that of \( M \). The ob­ject of [1] was to study the speed with which these volumes con­verge.

Be­fore study­ing the ef­fect of sur­ger­ies, we must un­der­stand the volume of a single hy­per­bol­ic 3-man­i­fold \( M \). Clearly it can be giv­en as the sum of the volumes of the tet­ra­hedra of any ideal tri­an­gu­la­tion, so the first step is to un­der­stand these. An (ori­ented and nonde­gen­er­ate) ideal tet­ra­hed­ron can be para­met­rized either by a shape para­met­er \( z \) in the com­plex up­per half-plane, as ex­plained in the pre­ced­ing sec­tion, or, in case \( z \) is in the up­per half-plane, by the three angles \( \alpha \), \( \beta \), \( \gamma \) (pos­it­ive and with \( \alpha+\beta+\gamma=\pi \)) of the Eu­c­lidean tri­angle that one “sees” by look­ing at the tet­ra­hed­ron from any of its four cusps. When \( \operatorname{Im}(z) > 0 \), the angles and the shapes are re­lated by \( (\alpha,\beta,\gamma)=(\arg(z),\arg(z^{\prime}),\arg(z^{\prime\prime})) \), with \( z^{\prime} \) and \( z^{\prime\prime} \) as above.

Ac­cord­ing to Chapter 7 of Thur­ston’s notes, writ­ten by Mil­nor, the volume of this tet­ra­hed­ron when \( \operatorname{Im}(z) > 0 \) is giv­en in terms of these two para­met­riz­a­tions by the two for­mu­las \begin{equation} \label{Kummer} \operatorname{Vol}(\Delta(z)) = D(z) = \text{Л}(\alpha)+\text{Л}(\beta)+\text{Л}(\gamma), \end{equation} the equal­ity of the two be­ing an iden­tity due to Kum­mer. (Both for­mu­las are ac­tu­ally true for all \( z\ne0,\,1 \), but the sum of \( \alpha \), \( \beta \) and \( \gamma \) is \( -\pi \) if \( z \) is in the lower half-plane.) Here \( D(z) \) and \( \text{Л}(\theta) \) are the Bloch–Wign­er di­log­ar­ithm (see Sec­tion 3) and the Lob­achevsky func­tion, defined re­spect­ively by \begin{equation} \begin{aligned} D(z) &=\operatorname{Im}\bigl(\operatorname{Li}_2(z)+\log|z|\log(1-z)\bigr), \\ \text{Л}(\theta) &= -\int_0^{\theta} \log|2\sin t|\,dt =\frac12\sum_{n=1}^\infty\frac{\sin(2n\theta)}{n^2} = \frac12 D(e^{2i\theta}). \end{aligned} \end{equation} Thus the volume of a hy­per­bol­ic man­i­fold \( M \) tri­an­gu­lated by \( N \) ideal tet­ra­hedra with shape para­met­ers \( z_i \) is giv­en in terms of the di­log­ar­ithm func­tion by \begin{equation} \operatorname{Vol}(M) = \sum_{j=1}^N D(z_j). \end{equation}

Now let \( M \) be a 3-man­i­fold with \( h \) cusps. It has a unique com­plete hy­per­bol­ic struc­ture as a quo­tient \( \mathbb{H}^3/\Gamma \) for some lat­tice \( \Gamma\subset\operatorname{PSL}_2(\mathbb{C}) \) iso­morph­ic to \( \pi_1(M) \), the cor­res­pond­ing shape para­met­ers \( z^0=(z^0_1,\dots,z^0_N) \) be­ing a solu­tion of all \( N+2h \) equa­tions (1) and (2). A small de­form­a­tion of this hy­per­bol­ic struc­ture will have a nearby shape para­met­er vec­tor \( z=(z_1,\dots,z_N) \) sat­is­fy­ing only (1). By the res­ult of Thur­ston men­tioned earli­er, the space of all such de­form­a­tions has the struc­ture of a smooth com­plex man­i­fold of di­men­sion \( h \), so it is iso­morph­ic to a small neigh­bor­hood \( \mathfrak{U} \) of 0 in \( \mathbb{C}^h \). Thus each \( z_i \) de­pends holo­morph­ic­ally on \( \mathfrak{u}\in\mathfrak{U} \) and \( z_i(0)=z_i^0 \). To make this more vis­ible, and to choose nice co­ordin­ates \( \mathfrak{u}_i \) on \( \mathfrak{U} \), we have to look more closely at the struc­ture of the cusps on \( M \) and at Dehn sur­ger­ies.

Each cusp has the struc­ture \( [0,\infty)\times T^2 \), where \( T^2 \) is a totally geodesic­ally em­bed­ded tor­us in \( M \) and has a flat Eu­c­lidean met­ric, unique up to ho­mothety, in­duced by the hy­per­bol­ic met­ric on \( M \). More con­cretely, the cusps of \( M=\mathbb{H}^3/\Gamma \) are in­dexed by \( P\in\mathbb{P}^1(\mathbb{C}) \), whose sta­bil­izer \( \Gamma_P \) is free abeli­an of rank 2. After con­jug­a­tion we can place \( P \) at \( \infty \), in which case \( \Gamma_\infty \) has the form \[ \pm\biggl(\,\begin{matrix} 1&\Lambda\\0&1 \end{matrix}\biggr) \] for some lat­tice \( \Lambda\subset\mathbb{C} \), unique up to ho­mothety, and then we can identi­fy \( \Lambda \) with \( H_1(T^2) \) and \( T^2 \) with \( \mathbb{C}/\Lambda \). There is a ca­non­ic­al quad­rat­ic form on \( \Lambda \) defined as the square of the length of a vec­tor di­vided by the volume of \( T^2 \). If we choose an ori­ented basis \( (\mu,\lambda) \) of \( H_1(T^2) \) (“me­ridi­an” and “lon­git­ude”) to identi­fy \( \Lambda \) with \( \mathbb{Z}^2 \), then this quad­rat­ic form is giv­en by \[ Q(p,q)=\frac{|p\tau+q|^2}{\operatorname{Im}(\tau) } \] if we res­cale \( \Lambda \) after ho­mothety to be \( \mathbb{Z}\tau+\mathbb{Z} \) for some \( \tau \) in the com­plex up­per half-plane.

On the oth­er hand, each ele­ment of \( \Lambda=H_1(T^2) \) can be iden­ti­fied with an iso­topy class of closed curves on \( T^2 \). Do­ing a \( (p,q) \) Dehn sur­gery at a cusp, once the basis of \( H_1(T^2) \) has been chosen, means re­mov­ing \( T^2\times[0,\infty) \) from \( M \) and re­pla­cing it by a sol­id tor­us in such way that the curve on \( T^2 \) cor­res­pond­ing to the class \( p\mu+q\lambda \) bounds in the sol­id tor­us. Here we as­sume that the in­tegers \( p \) and \( q \) are coprime, but we also al­low “\( \infty \)” as a value for \( (p,q) \), mean­ing that we leave this cusp un­touched. The Dehn sur­ger­ies on \( M \) are then de­scribed by tuples \[ \kappa=((p_1,q_1),\dots,(p_h,q_h))\in(\mathbb{Z}^2\cup\{\infty\})^h. \] Thur­ston’s the­or­em [e7] tells us that if all \( (p_i,q_i) \) are near enough to \( \infty \) (mean­ing that \( (p_i,q_i)=\infty \) or \( p_i^2+q_i^2 \) is large), the surgered man­i­fold \( M_\kappa \) is hy­per­bol­ic, in which case both the shape para­met­ers \( z_1,\dots,z_N \) and the volume \( \operatorname{Vol}(M_\kappa) \) be­come func­tions of \( \kappa \). Ex­pli­citly, the de­formed shape para­met­ers \( \boldsymbol{z}(\kappa)=\boldsymbol{z}(\boldsymbol{p},\boldsymbol{bq}) \) that tend to \( \boldsymbol{z}^0 \) as all the pairs \( (p_i,q_i) \) tend to in­fin­ity are giv­en by adding to the ori­gin­al glu­ing equa­tions (1) the \( h \) new equa­tions giv­en by the product of the \( p_i \)-th power of the first ex­pres­sion by the \( q_i \)-th power of the second one in (2), and then \( \operatorname{Vol}(M_\kappa) \) is giv­en by (6) with \( z_j \) re­placed by \( z_j(\kappa) \).

We now get two real num­bers at each cusp: the value \( Q_i(p_i,q_i) \) of the quad­rat­ic form cor­res­pond­ing to that cusp at the pair \( (p_i,q_i) \) (with the con­ven­tion \( Q_i(\infty)=\infty \)) and the length \( L_i=L_i(\kappa) \) of the short geodes­ic on \( M_\kappa \), which is the core of the sol­id tor­us ad­ded by the Dehn sur­gery (or 0 if the \( i \)-th cusp has not been surgered). They are re­lated by \begin{equation} \label{systole} L_i = \frac{2\pi}{Q_i(p_i,q_i)} + \mathrm{O}\biggl(\sum_{i=1}^h\frac1{p_i^4+q_i^4}\biggr) \end{equation} ([1], Pro­pos­i­tion 4.3) as all \( \kappa_i=(p_i,q_i) \) tend to in­fin­ity. The main volume res­ult of [1], proved by us­ing (6) and ana­lyz­ing the changes of the di­log­ar­ithms un­der small changes of the \( z \)’s, is then giv­en by the pair of asymp­tot­ic for­mu­las \begin{align}\label{VolDehn1} \operatorname{Vol}(M_\kappa) &= \operatorname{Vol}(M) -\sum_{i=1}^h\biggl(\frac{\pi^2}{Q_i(p_i,q_i)} + \mathrm{O}\biggl(\frac1{p_i^4+q_i^4}\biggr)\!\biggr) \tag{8a}\\ \label{VolDehn2} &= \operatorname{Vol}(M) -\sum_{i=1}^h\biggl(\frac{\pi L_i}2 + \mathrm{O}(L_i^2)\!\biggr) \tag{8b} \end{align} (The­or­ems 1A and 1B in [1]), which are equi­val­ent to one an­oth­er by vir­tue of (7). These volumes, as \( \kappa=(\boldsymbol{p},\boldsymbol{bq}) \) ranges over all \( h \)-tuples of suf­fi­ciently large pairs of coprime in­tegers or the sym­bol \( \infty \) mean­ing un­surgered, all be­long to the hy­per­bol­ic volume spec­trum, and equa­tion (8a) has as an im­me­di­ate co­rol­lary a de­scrip­tion of the loc­al struc­ture of this volume spec­trum near its lim­it point, be­cause the asymp­tot­ics of the num­ber of lat­tice points, or of prim­it­ive lat­tice points, in a large el­lipse, is well known. The pre­cise asymp­tot­ic state­ment, which we will not re­peat here, is for­mu­lated ex­pli­citly as a co­rol­lary to The­or­em 1A in [1].

All of this is only for in­teg­ral (and coprime) val­ues of \( p_i \) and \( q_i \). However, as is ex­plained in [1] in de­tail, the shape para­met­ers \( z_j(\kappa) \) and the lengths \( L_i(\kappa) \) are defined for \( p \) and \( q \) real rather than just in­teg­ral and coprime in pairs, and equa­tions (7) and (8) still re­main true. The only point is that in the defin­i­tion of the \( z_j \)’s we had to take the \( p_i \)-th and \( q_i \)-th powers of the equa­tions in (2), and one can­not in gen­er­al take real powers of com­plex num­bers in a well-defined way, but since the left-hand sides of the ex­pres­sions in (2) are near to 1 for small de­form­a­tions of the ori­gin­al value \( \boldsymbol{z}=\boldsymbol{z}^0 \) and since a com­plex num­ber near 1 has a well-defined log­ar­ithm near 0, there is no prob­lem. When the \( p_i \) and \( q_i \) are not in­teg­ral, we are no longer “filling in” the cusp by glu­ing on a sol­id tor­us, but are simply chan­ging the hy­per­bol­ic struc­ture on the ori­gin­al open to­po­lo­gic­al man­i­fold \( M \), with the new hy­per­bol­ic struc­tures in gen­er­al be­ing in­com­plete. If we now define \( 2h \) com­plex num­bers \( \mathfrak{u}=(\mathfrak{u}_1,\dots,\mathfrak{u}_h) \) and \( \mathfrak{v}=(\mathfrak{v}_1,\dots,\mathfrak{v}_h) \) by \[ \mathfrak{u}_i=\sum_{j=1}^N\biggl(M^{\prime}_{ij}\log\frac{z_j}{z_j^0} +M^{\prime\prime}_{ij}\log\frac{1-z_j}{1-z_j^0}\biggr),\quad \mathfrak{v}_i=\sum_{j=1}^N\biggl(L^{\prime}_{ij} \log\frac{z_j}{z_j^0}+L^{\prime\prime}_{ij}\log\frac{1-z_j}{1-z_j^0}\biggr), \] then from the sym­plect­ic prop­er­ties of the glu­ing equa­tions it fol­lows that ([2], Lemma 4.1) \[ p_i\mathfrak{u}_i + q_i\mathfrak{v}_i= 2\pi i\quad(i=1,\dots,h), \] and we can take \( \mathfrak{u} \) as ca­non­ic­al co­ordin­ates for the above-men­tioned neigh­bor­hood \( \mathfrak{U} \) of \( 0\in\mathbb{C}^h \), in which case each \( \mathfrak{v}_i \) be­comes an odd power series in the \( \mathfrak{u} \)’s with lin­ear term \( \tau_i\mathfrak{u}_i \) and we can write \( L_i(\mathfrak{u}) \) and \( \operatorname{Vol}(\mathfrak{u}) \) in­stead of \( L_i(\kappa) \) and \( \operatorname{Vol}(M_\kappa) \). We should men­tion that \( \mathfrak{u} \) and \( \mathfrak{v} \) can be defined in­vari­antly, without us­ing any tri­an­gu­la­tion, as fol­lows: the de­formed hy­per­bol­ic struc­ture on \( M \) cor­res­ponds to a ho­mo­morph­ism \( \rho:\Gamma\to\operatorname{PSL}_2(\mathbb{C}) \) near to the in­clu­sion map, and then \( \mathfrak{u} \) and \( \mathfrak{v} \) are simply the log­ar­ithms of the ra­tios of the ei­gen­val­ues of the im­ages un­der \( \rho \) of the me­ridi­ans and lon­git­udes, re­spect­ively.

In terms of the new co­ordin­ates, equa­tion (8b) be­comes \begin{equation} \operatorname{Vol}(\mathfrak{u}) = \operatorname{Vol}(M) -\frac\pi 2\sum_{i=1}^h L_i(\mathfrak{u}) + \varepsilon(\mathfrak{u}), \tag{9} \end{equation} with \( \varepsilon(\mathfrak{u})=\mathrm{O}(\|\mathfrak{u}\|^4) \) as \( \mathfrak{u} \) tends to 0 in \( \mathbb{C}^h \). The­or­em 2 of [1] was the state­ment that the func­tion \( \varepsilon(\mathfrak{u}) \) defined by (9) is har­mon­ic, and hence is the real part of a holo­morph­ic func­tion \( f(\mathfrak{u}) \) near 0 (uniquely de­term­ined if we fix \( f(0)=0 \)). The­or­em 3, proved us­ing equa­tion (3), said that \( \partial\mathfrak{v}_i/\partial\mathfrak{u}_j \) is sym­met­ric in \( i \) and \( j \), which im­plies that there is a single func­tion \( \Phi(\mathfrak{u}) \) with \( \partial\Phi/\partial\mathfrak{u}_i=2\mathfrak{v}_i \) for all \( i \), and that \( f \) is giv­en in terms of \( \Phi \) by \( 4f=\Phi-\mathfrak{u}\cdot\mathfrak{v} \) (or equi­val­ently by \( -8f=(E-2)\Phi \), where \( E=\sum\mathfrak{u}_i\, \partial/\partial\mathfrak{u}_i \) is the Euler op­er­at­or), so that the volume cor­rec­tion \( \varepsilon(\mathfrak{u}) \) in (9) is giv­en by \begin{equation} \label{VolDehn4} \varepsilon(\mathfrak{u}) = \operatorname{Im}(f(\mathfrak{u})), \quad f(\mathfrak{u}):=\frac14\int_0^\mathfrak{u} \biggl(\sum_{i=1}^h(\mathfrak{v}_i\,d\mathfrak{u}_i-\mathfrak{u}_i\,d\mathfrak{v}_i)\!\biggr). \tag{10} \end{equation} The func­tion \( f \) is now of­ten called the Neu­mann–Za­gi­er po­ten­tial func­tion, al­though this name was used in the ori­gin­al pa­per for \( \Phi \) in­stead. It should per­haps also be men­tioned that sim­pler proofs of the last res­ults de­scribed could prob­ably have been ob­tained by us­ing the second rather than the first volume for­mula in (4).

There is one more im­port­ant point about volumes. An­oth­er in­sight by Thur­ston was that the volume of a hy­per­bol­ic 3-man­i­fold, which is a pos­it­ive real num­ber, is ac­tu­ally in a nat­ur­al way the ima­gin­ary part of a com­plexi­fied volume whose real part is the Chern–Si­mons in­vari­ant, an im­port­ant to­po­lo­gic­al in­vari­ant tak­ing val­ues in \( \mathbb{R}/(4\pi^2\mathbb{Z}) \) whose defin­i­tion we omit here. It was con­jec­tured in [1], and proved soon af­ter­wards by Yoshida [e2], that the above for­mu­las re­main true with the volumes re­placed by their com­plexi­fied ver­sions, the func­tions \( L_i(\mathfrak{u}) \) also lif­ted suit­ably from \( \mathbb{R} \) to \( \mathbb{C} \), and \( \varepsilon(\mathfrak{u}) \) re­placed by \( f(\mathfrak{u}) \). Later, in [2], Wal­ter showed how to lift (6) to an ex­pli­cit and com­put­able ex­pres­sion for the com­plexi­fied volume of \( M_{\boldsymbol{p},\boldsymbol{bq}} \) in terms of the com­plex di­log­ar­ithm.

3. The Bloch group and the extended Bloch group

The Bloch group of a field is an ana­logue of its mul­ti­plic­at­ive group, but with the re­la­tion \( [xy]=[x]+[y] \) sat­is­fied by the log­ar­ithm func­tion re­placed by the func­tion­al equa­tion of the di­log­ar­ithm. In this sec­tion we re­call its defin­i­tion and the defin­i­tion of the “ex­ten­ded Bloch group” that was in­tro­duced by Wal­ter [4] and fur­ther de­veloped by Zick­ert and Goette [e15], [e26], and ex­plain their con­nec­tions with the volume and com­plexi­fied volume. The next sec­tion tells how these things re­late to the sym­plect­ic struc­ture. We should men­tion that parts of both sec­tions have been trans­ferred here from the arX­iv ver­sion of [e33] and also ed­ited some­what for the pur­pose of the present ex­pos­i­tion.

The di­log­ar­ithm func­tion \( \operatorname{Li}_2(z) \), defined for \( |z| < 1 \) as \( \sum_{n=1}^\infty z^n/n^2 \) and then ex­ten­ded ana­lyt­ic­ally to either the cut plane \( \mathbb{C}\smallsetminus[1,\infty) \) or to the uni­ver­sal cov­er of \( \mathbb{P}^1(\mathbb{C})\smallsetminus\{0,1,\infty\} \), sat­is­fies a fam­ous func­tion­al equa­tion called the five-term re­la­tion. This func­tion­al equa­tion was dis­covered re­peatedly dur­ing the 19th cen­tury and can be writ­ten in many equi­val­ent forms, each say­ing that a sum of five di­log­ar­ithm val­ues is a lin­ear com­bin­a­tion of products of simple log­ar­ithms. The func­tion \( \operatorname{Li}_2 \) is many-val­ued, but the mod­i­fied di­log­ar­ithm (5) is a single-val­ued real ana­lyt­ic func­tion from \( \mathbb{P}^1(\mathbb{C})\smallsetminus\{0,1,\infty\} \) to \( \mathbb{R} \) that ex­tends con­tinu­ously to all of \( \mathbb{P}^1(\mathbb{C}) \) and sat­is­fies “clean” ver­sions of the five-term re­la­tions with no log­ar­ithmic cor­rec­tion terms. Since \( D(z) \) also sat­is­fies the two func­tion­al equa­tions \( D(1-z)=-D(z)=D(1/z) \) (im­ply­ing that \( D(z)=D(z^{\prime})=D(z^{\prime\prime}) \) for the three shape para­met­ers of an ori­ented ideal hy­per­bol­ic tet­ra­hed­ron), this “clean” func­tion­al equa­tion still can be writ­ten in many dif­fer­ent forms, one stand­ard one be­ing \[ D(x) + D(y) + D\biggl(\frac{1-x}{1-xy}\biggr) + D(xy) + D\biggl(\frac{1-y}{1-xy}\biggr) = 0 \] for \( (x,y)\ne(1,1) \) in \( \mathbb{C}^2 \). An­oth­er nice ver­sion is the cyc­lic one \( \sum_{i\!\pmod5} D(z_i)=0 \) if \( \{z_i\}_{i\in\mathbb{Z}} \) is a se­quence of com­plex num­bers sat­is­fy­ing \( 1-z_i=z_{i-1}z_{i+1} \) for all \( i \) (which im­plies by a short cal­cu­la­tion that they have peri­od 5). Yet an­oth­er, with a clear in­ter­pret­a­tion in terms of 3-di­men­sion­al hy­per­bol­ic geo­metry, says that the signed sum of \( D(r_i) \) is 0 if \( r_1,\dots,r_5 \) are the cross-ra­tios of the five sub­sets of car­din­al­ity 4 of a set of five dis­tinct points in \( \mathbb{P}^1(\mathbb{C}) \).

The five ar­gu­ments \( z_i \) of any ver­sion of the five-term re­la­tion sat­is­fy \( \sum(z_i)\wedge(1-z_i)=0 \), where the sum is taken in the second ex­ter­i­or power of the mul­ti­plic­at­ive group of \( \mathbb{C} \). For in­stance, for the “cyc­lic ver­sion” above we have \[ \sum_i(z_i)\wedge(1-z_i) = \sum_i(z_i)\wedge(z_{i-1}z_{i+1})=\sum_i((z_i)\wedge(z_{i-1})+(z_{i-1})\wedge(z_i))=0. \] The Bloch group \( \mathcal{B}(F) \) of an ar­bit­rary field \( F \), in­tro­duced by Bloch [e1] in 1978, is mo­tiv­ated by this ob­ser­va­tion and is defined as the quo­tient of the ker­nel of the map \( d:\mathbb{Z}[F]\to\mathsf{Li}^2(F^\times) \) send­ing \( [x] \) to \( x\wedge(1-x) \) for \( x\ne0,1 \) (and to 0 for \( x=0,1 \)) by the sub­group gen­er­ated by the five-term re­la­tion of the di­log­ar­ithm. The pre­cise defin­i­tion var­ies slightly in the lit­er­at­ure be­cause of del­ic­ate 2- and 3-tor­sion is­sues arising from the par­tic­u­lar defin­i­tion of the ex­ter­i­or square (for in­stance, does one re­quire \( x\wedge x=0 \) for all \( x \) or just \( x\wedge y=-y\wedge x \)?), wheth­er one re­quires \( d([0]) \) and \( d([1]) \) to van­ish or merely to be tor­sion, and the par­tic­u­lar ver­sion of the five-term re­la­tion used. We will gloss over this point for now, but will come back to it in con­nec­tion with the ex­ten­ded Bloch group.

From our point of view, the clearest mo­tiv­a­tion for the defin­i­tion of the Bloch group is the fact that the shape para­met­ers \( \{z_i\} \) for any ideal tri­an­gu­la­tion \( \bigcup_i\Delta(z_i) \) of a com­plete hy­per­bol­ic 3-man­i­fold \( M \) sat­is­fy \[ \sum_i(z_i)\wedge(1-z_i)=0. \] (This is a con­sequence of the sym­plect­ic nature of the NZ re­la­tions, as we will ex­plain in more de­tail in the next sec­tion.) Thus to any such tri­an­gu­la­tion we can as­so­ci­ate a class \( \sum_i[z_i] \) in the Bloch group. But this class is in fact in­de­pend­ent of the tri­an­gu­la­tion, since (mod­ulo some tech­nic­al points con­cern­ing the fact that the shapes can de­gen­er­ate to 0 or 1 un­der 2-3 Pach­ner moves) any two tri­an­gu­la­tions are linked by a series of “2-3 Pach­ner moves” in which two tet­ra­hedra shar­ing a com­mon face are re­placed by the three tet­ra­hedra defined by their two non­shared and two of their three shared ver­tices, and the (signed) sum of the shape para­met­ers of these five tet­ra­hedra is pre­cisely the five-term re­la­tion and does not af­fect the class of \( \sum_i[z_i] \) in the Bloch group. Thus one has a class \( [M]\in\mathcal{B}(\mathbb{C}) \). Moreover, from the very defin­i­tion of the Bloch group it fol­lows that the func­tion \( D \) ex­tends to a lin­ear map from \( \mathcal{B}(\mathbb{C}) \) to \( \mathbb{R} \), and from the dis­cus­sion in the last sec­tion we see that the value of \( D \) on the class \( [M] \) is equal to the volume of \( M \). Al­though we will not use it, we men­tion that by a res­ult of Suslin the Bloch group \( \mathcal{B}(F) \) of any field \( F \) is iso­morph­ic up to tor­sion to the al­geb­ra­ic \( K \)-group \( K_3(F) \), with \( D \) cor­res­pond­ing to the Borel reg­u­lat­or map from \( K_3(\mathbb{C}) \) to \( \mathbb{R} \) in the case \( F=\mathbb{C} \).

On the oth­er hand, as de­scribed at the end of the last sec­tion, the hy­per­bol­ic volume should ac­tu­ally be seen as the ima­gin­ary part of a com­plexi­fied volume tak­ing val­ues in \( \mathbb{C}/4\pi^2\mathbb{Z} \), so we would like to re­place the func­tion \( D(z) \) by some com­plex-val­ued ver­sion of the di­log­ar­ithm which, even though it may be many-val­ued at in­di­vidu­al ar­gu­ments \( z \), be­comes one-val­ued mod­ulo \( 4\pi^2 \) if we take a lin­ear com­bin­a­tion of its val­ues with ar­gu­ments be­long­ing to the Bloch group. This is the idea be­hind the pas­sage from the ori­gin­al Bloch group to the ex­ten­ded one. The first ob­ser­va­tion (see [e9]) is that the func­tion \( L(v):=\operatorname{Li}_2(1-e^v) \) has the de­riv­at­ive \( v/(e^{-v}-1) \), which is mero­morph­ic and has residues in \( 2\pi i\mathbb{Z} \), so that \( L \) it­self lifts to a well-defined func­tion from \( \mathbb{C}\smallsetminus2\pi i\mathbb{Z} \) to \( \mathbb{C}/4\pi^2\mathbb{Z} \) and sat­is­fies the func­tion­al equa­tion \[ L(v+2\pi in)=L(v)-2\pi in\log(1-e^v) \quad\text{for }n\in\mathbb{Z}. \] We now in­tro­duce the com­plex 1-man­i­fold \[ \widehat{\mathbb{C}} = \{(u,v)\in\mathbb{C}^2\mid e^u+e^v=1\}. \] This is an abeli­an cov­er of \( \mathbb{C}^\times\smallsetminus\{0,1\} \) via \( z=e^u=1-e^v \), with Galois group iso­morph­ic to \( \mathbb{Z}^2 \). The ex­ten­ded Bloch group \( \widehat{\mathcal{B}}(\mathbb{C}) \) as defined in [e15] or [e26] is the ker­nel of the map \( \hat d:\mathbb{Z}[\widehat{\mathbb{C}}]\to\mathsf{Li}^2(\mathbb{C}) \), where \( \mathsf{Li}^2(\mathbb{C}) \) is defined by re­quir­ing only \( x\wedge y+y\wedge x=0 \) (rather than \( x\wedge x=0 \), which is stronger by 2-tor­sion) and where \( \hat d \) maps \( [u,v]:=[(u,v)]\in\mathbb{Z}[\widehat{\mathbb{C}}] \) to \( u\wedge v \), di­vided by an ap­pro­pri­ate lif­ted ver­sion of the five-term re­la­tion, namely, the \( \mathbb{Z} \)-span of the set of ele­ments \( \sum_{j=1}^5(-1)^j[u_j,v_j] \) of \( \mathbb{Z}(\widehat{\mathbb{C}}) \) sat­is­fy­ing \[ (u_2,u_4)=(u_1+u_3,u_3+u_5) \quad\text{and}\quad (v_1,v_3,v_5)=(u_5+v_2,v_2+v_4,u_1+v_4). \] There is an ex­ten­ded reg­u­lat­or map from \( \widehat{\mathcal{B}}(\mathbb{C}) \) to \( \mathbb{C}/4\pi^2\mathbb{Z} \) giv­en by map­ping \( \sum[u_j,v_j] \) to \( \sum\mathcal{L}(u_j,v_j) \), where \[ \textstyle\mathcal{L}(u,v)=L(v)+\frac12uv-\frac{\pi^2}6, \] which one can check van­ishes mod­ulo \( 4\pi^2 \) on the lif­ted five-term re­la­tion. One can also define \( \widehat{\mathcal{B}}(F) \) for any sub­field \( F \) of \( \mathbb{C} \), such as an em­bed­ded num­ber field, by re­pla­cing \( \widehat{\mathbb{C}} \) by the sub­set \( \widehat F \) con­sist­ing of pairs \( (u,v) \) with \( e^u=1-e^v\in F \).

As a fi­nal re­mark, one can won­der to what ex­tent study­ing just hy­per­bol­ic 3-man­i­folds lets one un­der­stand the full Bloch group of \( \overline{\mathbb{Q}} \). For in­stance, does every ele­ment of \( \mathcal{B}(\overline{\mathbb{Q}}) \) oc­cur as a ra­tion­al lin­ear com­bin­a­tion of the Bloch group in­vari­ants of some hy­per­bol­ic 3-man­i­folds? Even more ba­sic­ally, does every num­ber field with at least one non­real em­bed­ding oc­cur as the trace field of some hy­per­bol­ic 3-man­i­fold? The lat­ter ques­tion was posed ex­pli­citly by Wal­ter in [3].

4. Symplectic properties

In ret­ro­spect, the sym­plect­ic prop­er­ties as de­scribed in equa­tion (3) and the fol­low­ing text, and their re­fine­ment from \( \mathbb{Q} \) to \( \mathbb{Z} \) as giv­en in the fol­low-up pa­per [2], turned out to be the most im­port­ant as­pects of these pa­pers. They are re­spons­ible both for the ex­ist­ence of the po­ten­tial func­tion and for all of the ap­plic­a­tions to quant­iz­a­tion that we will de­scribe in the next sec­tion, as well as many of the con­nec­tions to num­ber the­ory de­scribed in Sec­tion 6.

Define an \( N\times2N \) mat­rix \( H=(A\,B) \) whose rows form a \( \mathbb{Z} \)-basis for the lat­tice spanned by the edge equa­tions (1) to­geth­er with one “peri­pher­al” equa­tion (a coprime lin­ear com­bin­a­tion of the me­ridi­an and lon­git­ude equa­tions in (2)) at each cusp. Then the above cited res­ults in [1] im­ply that \( AB^t \) is sym­met­ric and that \( H \) has rank \( N \), mean­ing that its \( 2N \) columns gen­er­ate \( \mathbb{Q}^N \). To­geth­er, these two state­ments are equi­val­ent to say­ing that \( H \) can be ex­ten­ded to a \( 2N\times2N \) mat­rix \[ \biggl(\,\begin{matrix} A & B\\C & D\end{matrix}\biggr) \] in \( \operatorname{Sp}_{2N}(\mathbb{Q}) \), mean­ing that \[ \biggl(\,\begin{matrix} A & B\\C & D\end{matrix}\biggr)^{-1}= \biggl(\,\begin{matrix}\phantom{-}{D^t}&{-B^t}\\{-C^t}&\phantom{-}{A^t}\end{matrix}\biggr). \]

But in fact the \( 2N \) columns of \( H \) span the lat­tice \( \mathbb{Z}^N \), which is equi­val­ent to say­ing that \( H \) can be com­pleted to a \( 2N\times2N \) sym­plect­ic mat­rix over \( \mathbb{Z} \). (We will call such a mat­rix half-sym­plect­ic.) This fol­lows from the chain com­plex defined by Wal­ter in [2]. Ex­pli­citly, for any sim­plex \( \Delta \), let \( J_\Delta \) be the abeli­an group gen­er­ated by \( e_1,e_2,e_3 \) (cor­res­pond­ing to the pairs of op­pos­ite edges) sub­ject to the re­la­tion \( e_1+e_2+e_3=0 \). This is a free abeli­an group of rank 2, with a ca­non­ic­al nonsin­gu­lar, skew-sym­met­ric bi­lin­ear form giv­en by ([2], Sec­tion 4) \begin{equation} \label{omega} \langle e_1, e_2 \rangle = \langle e_2, e_3 \rangle = \langle e_3, e_1 \rangle = -\langle e_2, e_1 \rangle = - \langle e_3, e_2 \rangle =- \langle e_1, e_3 \rangle = 1. \tag{11} \end{equation} The Neu­mann chain com­plex as­so­ci­ated to an ideal tri­an­gu­la­tion is then defined by \begin{equation} \label{ChainComplex} 0 \longrightarrow C_0\stackrel\alpha\longrightarrow C_1\stackrel\beta\longrightarrow J\stackrel{\beta^*}\longrightarrow C_1\stackrel{\alpha^*}\longrightarrow C_0\longrightarrow 0. \tag{12} \end{equation} Here \( C_0 \) and \( C_1 \) are the free abeli­an groups on the un­ori­ented 0- and 1-sim­plices (cusps and edges), re­spect­ively, and \[ J=\bigoplus_{i=1}^NJ_{\Delta_i} \] (sum over the 3-sim­plices or tet­ra­hedra), while \( \alpha \) maps any cusp to the sum of its in­cid­ent edges, the \( J_\Delta \)-com­pon­ent of \( \beta \) of any edge is the sum of the edges of \( \Delta \) that are iden­ti­fied with it, and \( \alpha^* \) and \( \beta^* \) are the du­als of \( \alpha \) and \( \mathcal{B} \) with re­spect to the ob­vi­ous scal­ar products on \( C_i \) and the sym­plect­ic form on \( J \). Wal­ter shows ([2], The­or­em 4.1) that the se­quence (12) is a chain com­plex and, at least after tensor­ing with \( \mathbb{Z}\bigl[\frac12\bigr] \), is ex­act ex­cept in the middle, where the ho­mo­logy is the sum of \( h \) rank-2 mod­ules iso­morph­ic to \( H_1(T^2_i) \) \( (i=1,\dots,h) \). Note that the map \( \beta \) is giv­en pre­cisely by the mat­rix \( R=(R^{\prime}\,R^{\prime\prime}) \) as defined in (1) if we choose the ob­vi­ous basis for \( C_1 \) and the basis of \( J \) giv­en by choos­ing the basis \( (e_1,e_2) \) for every \( J_{\Delta_j} \). The rest of the proof that \( H \) is half-sym­plect­ic fol­lows eas­ily from the the­or­em just quoted and will be left to the read­er.

We make two re­marks about this. The first is that both the con­struc­tion of the chain com­plex and the state­ment about its ho­mo­logy were done in [2] also for 3-man­i­folds with bound­ary com­pon­ents of ar­bit­rary genus (so the ver­tices of \( C_0 \) need not be cusps), and of course also do not re­quire any hy­per­bol­ic struc­ture. The oth­er is that the glu­ing equa­tions of [1] and the sym­plect­ic res­ults of [2] were ex­ten­ded to ar­bit­rary \( \operatorname{PGL}_n \)-rep­res­ent­a­tions in [e29].

Half-sym­plect­ic matrices oc­cur in oth­er con­texts, e.g., in con­nec­tion with Nahm’s con­jec­ture on the mod­u­lar­ity of cer­tain \( q \)-hy­per­geo­met­ric series, and also lead to a new de­scrip­tion of the Bloch group. Both top­ics will be dis­cussed in more de­tail in Sec­tion 6.

5. Quantization

Per­haps the most far-reach­ing con­sequences of Wal­ter’s work on the com­bin­at­or­ics of 3-di­men­sion­al tri­an­gu­la­tions have been the ap­plic­a­tions of the sym­plect­ic struc­ture to quant­iz­a­tion.

Re­call the defin­i­tion of \( J_\Delta \) for a single tet­ra­hed­ron \( \Delta \) as the abeli­an group \[ \langle e_1,e_2,e_3\mid e_1+e_2+e_3=0\rangle \] with the sym­plect­ic struc­ture (11). This sym­plect­ic struc­ture on each space \( J_\Delta \otimes_\mathbb{Z} \mathbb{Q} \) for any ideal tet­ra­hed­ron \( \Delta \) leads to an in­teg­ral Lag­rangi­an sub­space of the 10-di­men­sion­al sym­plect­ic space \[ \bigoplus_{j=1}^5 J_{\Delta_j} \otimes_\mathbb{Z}\mathbb{Q} \] as­so­ci­ated to five tet­ra­hedra \( \Delta_1,\dots,\Delta_5 \) that par­ti­cip­ate in a 2-3 Pach­ner move. Roughly speak­ing, the Lag­rangi­an sub­space re­cords the lin­ear re­la­tions among the angles of the five tet­ra­hedra, where the signed sum of the angles around each in­teri­or edge of the Pach­ner move is zero.

The quant­iz­a­tion of this Lag­rangi­an sub­space has ap­peared nu­mer­ous times in the math­em­at­ics and phys­ics lit­er­at­ure, un­der dif­fer­ent names, and has led to in­ter­est­ing quantum in­vari­ants in di­men­sions 2, 3 and 4. We briefly dis­cuss this now. In di­men­sion 2, Kashaev and in­de­pend­ently Fo­ckGon­char­ov [e8], [e17], [e13] used the above NZ-sym­plect­ic struc­ture to study the change of co­ordin­ates of ideally tri­an­gu­lated sur­faces un­der a 2-2 Pach­ner move. They found that the cor­res­pond­ing iso­morph­ism of com­mut­at­ive al­geb­ras can be de­scribed in terms of cluster al­geb­ras, lead­ing to two dual sets of co­ordin­ates (the so-called \( \mathcal{X} \)-co­ordin­ates and the \( \mathcal{A} \)-co­ordin­ates) whose quant­iz­a­tion leads to a rep­res­ent­a­tion of the so-called Ptolemy group­oid, and in par­tic­u­lar of the map­ping class group of a punc­tured sur­face, and also of braid groups. These rep­res­ent­a­tions are al­ways in­fin­ite-di­men­sion­al (be­cause there are no fi­nite square matrices \( A \) and \( B \) sat­is­fy­ing the re­la­tion \( AB-BA=I \)), the Hil­bert spaces are typ­ic­ally \( L^2(\mathbb{R}^n) \) for some \( n \), and the cor­res­pond­ing the­ory is usu­ally known as quantum Teichmüller the­ory.

Go­ing one di­men­sion high­er, the \( \mathcal{X} \)-co­ordin­ates of a 3-di­men­sion­al ideal tri­an­gu­la­tion are noth­ing but the shapes of the ideal tet­ra­hedra, where­as the \( \mathcal{A} \)-co­ordin­ates are the Ptolemy vari­ables of the ideal tet­ra­hedra. The lat­ter are as­sign­ments of nonzero com­plex num­bers to the edges of the ideal tri­an­gu­la­tion (where iden­ti­fied edges are giv­en the same vari­able) that sat­is­fy a sys­tem of quad­rat­ic equa­tions: a (suit­ably) signed sum \( ab+cd=ef \), where \( (a,b) \), \( (b,d) \) and \( (e,f) \) are the Ptolemy vari­ables of the three pairs of op­pos­ite edges. It turns out that the NZ glu­ing equa­tions for shapes are equi­val­ent to the Ptolemy equa­tions (see for in­stance [e25]), and this is not only the­or­et­ic­ally in­ter­est­ing, but prac­tic­ally, too. The quant­iz­a­tion of the shape and Ptolemy vari­ables of an ideal tri­an­gu­la­tion uses two in­gredi­ents, the kin­emat­ic­al ker­nel of Kashaev [e30] and a spe­cial func­tion, the Fad­deev quantum di­log­ar­ithm that sat­is­fies an in­teg­ral pentagon iden­tity. Ac­cord­ing to Kashaev, the kin­emat­ic­al ker­nel is noth­ing but the quant­iz­a­tion of the NZ Lag­rangi­an men­tioned above. The out­come of this quant­iz­a­tion is the ex­ist­ence of to­po­lo­gic­al in­vari­ants of ideally tri­an­gu­lated 3-man­i­folds, the in­vari­ants be­ing ana­lyt­ic func­tions in a cut place \( \mathbb{C}^{\prime}=\mathbb{C}\setminus (-\infty,0] \), ex­pressed in terms of fi­nite-di­men­sion­al state in­teg­rals whose in­teg­rand is of­ten de­term­ined by the com­bin­at­or­i­al data of an ideal tri­an­gu­la­tion, namely its Neu­mann–Za­gi­er matrices. This con­struc­tion, which is of­ten known as quantum hy­per­bol­ic geo­metry, has been ax­io­mat­ized by Kashaev, and uses as in­put the com­bin­at­or­i­al data of an ideal tri­an­gu­la­tion to­geth­er with a self-dual loc­ally com­pact abeli­an group with fixed Gaus­si­an, Four­i­er ker­nel and quantum di­log­ar­ithm. This then leads to fur­ther ana­lyt­ic in­vari­ants of 3-man­i­folds, two ex­amples of which are the Kashaev–Luo–Vartan­ov in­vari­ants [e28] and the mero­morph­ic 3D-in­dex [e32], for which the LCA groups are \( \mathbb{R} \times \mathbb{R} \) and \( S^1 \times \mathbb{Z} \), re­spect­ively. It is worth not­ing that the An­der­sen–Kashaev state in­teg­rals are con­jec­tured to be the par­ti­tion func­tion of com­plex Chern–Si­mons the­ory (i.e., Chern–Si­mons the­ory with com­plex gauge group). The lat­ter is not known to sat­is­fy the cut-and-paste ar­gu­ments that the \( \operatorname{SU}(2) \) Chern–Si­mons the­ory does, and as a res­ult, one does not have an a pri­ori defin­i­tion of com­plex Chern–Si­mons the­ory oth­er than the state in­teg­rals, nor a clear reas­on why the in­fin­ite-di­men­sion­al path in­teg­ral loc­al­izes to a fi­nite-di­men­sion­al one.

Fi­nally, go­ing yet one di­men­sion high­er, the five ideal tet­ra­hedra that par­ti­cip­ate in a 2-3 Pach­ner move form the bound­ary of a single 4-di­men­sion­al sim­plex, a penta­choron. (Ex­cuse our Greek.) This gives a 4-di­men­sion­al in­ter­pret­a­tion of the NZ-struc­ture and of the kin­emat­ic­al ker­nel, and us­ing a com­plex root of unity, Kashaev was able to give a tensor in­vari­ant un­der 4-di­men­sion­al Pach­ner moves and thus con­struct cor­res­pond­ing to­po­lo­gic­al in­vari­ants of closed, tri­an­gu­lated 4-man­i­folds at roots of unity [e30]. This con­cludes our dis­cus­sion of the kin­emat­ic­al ker­nel in di­men­sions 2, 3 and 4.

In a dif­fer­ent dir­ec­tion, math­em­at­ic­al phys­i­cists, us­ing cor­res­pond­ence prin­ciples among su­per­sym­met­ric the­or­ies, have came up with un­ex­pec­ted con­struc­tions of vari­ous col­lec­tions of \( q \)-series with in­teger coef­fi­cients as­so­ci­ated to 3-man­i­folds. Per­haps the most re­mark­able of these is the 3D-in­dex of Dimofte, Gai­otto and Gukov [e23], [e22], where the \( q \)-series in ques­tion, which in this case are in­dexed by pairs of in­tegers, were defined ex­pli­citly in terms of the NZ-matrices of a suit­able ideal tri­an­gu­la­tion, with their coef­fi­cients count­ing the num­ber of BPS states of a su­per­sym­met­ric the­ory. This DGG 3D-in­dex was sub­sequently shown [e27] to be a to­po­lo­gic­al in­vari­ant of cusped hy­per­bol­ic 3-man­i­folds, and was also ex­ten­ded to a mero­morph­ic func­tion of two vari­ables (in case the bound­ary of the 3-man­i­fold is a single tor­us) whose Laurent coef­fi­cients are the DGG in­dex [e32].

A quite dif­fer­ent place where the NZ equa­tions ap­pear in quantum to­po­logy is in con­nec­tion with the Kashaev in­vari­ant and Kashaev’s fam­ous Volume Con­jec­ture. The Kashaev in­vari­ant \( \langle K\rangle_n \) is a com­put­able al­geb­ra­ic num­ber that was defined for any knot \( K \) and any pos­it­ive in­teger \( n \) by Kashaev [e5] in 1995 us­ing ideas of quantum to­po­logy sim­il­ar to those dis­cussed above, and of which an al­tern­at­ive defin­i­tion in terms of the so-called colored Jones poly­no­mi­al was later found by H. Murakami and J. Murakami. The Volume Con­jec­ture [e6] says that the log­ar­ithm of \( |\langle K\rangle_n| \) is asymp­tot­ic­ally equal to \( n/(2\pi) \) times the hy­per­bol­ic volume of the knot com­ple­ment \( M=S^3\smallsetminus K \) whenev­er \( M \) is hy­per­bol­ic, a very sur­pris­ing con­nec­tion between hy­per­bol­ic geo­metry and 3-di­men­sion­al quantum to­po­logy that has giv­en rise to a great deal of sub­sequent re­search and has been re­fined in many ways and by sev­er­al au­thors in con­nec­tion with com­plex Chern–Si­mons the­ory. In par­tic­u­lar, one has the con­jec­tur­al sharpen­ing [e16], [e18] \begin{equation} \label{arithmVolConj} \langle K\rangle_n \sim n^{3/2} e^{\operatorname{V}_{\mathbb{C}}(K)n/(2\pi i)}\Phi^K\biggl(\frac{2\pi i}n\biggr) \tag{13} \end{equation} to all or­ders in \( 1/n \) as \( n\to\infty \), where \( \Phi^K(h) \) is a power series in \( h \) with al­geb­ra­ic coef­fi­cients that can be com­puted to any or­der in any ex­pli­cit ex­ample, e.g., \begin{equation} \label{as41} \Phi^{4_1}(h)= \frac1{\sqrt[4]{3}} \biggl(1 + \frac{11}{72\sqrt{-3}}h + \frac{697}{2(72\sqrt{-3})^2}h^2 + \frac{724351}{30(72\sqrt{-3})^3}h^3 +\cdots\biggr) \tag{14} \end{equation} for the \( 4_1 \) (fig­ure-8) knot. In [e20], an ex­pli­cit can­did­ate for this power series is con­struc­ted for any knot as a form­al Gaus­si­an in­teg­ral whose in­teg­rand is defined in terms of the NZ data of an ideal tri­an­gu­la­tion of \( M \). It is not yet known bey­ond the lead­ing term that the series con­struc­ted there is a to­po­lo­gic­al in­vari­ant (i.e., in­de­pend­ent of the choice of tri­an­gu­la­tion), al­though this would of course fol­low from the con­jec­ture that the asymp­tot­ic for­mula (13) holds with this series. In a fol­low-up pa­per [e31], the con­struc­tion was ex­ten­ded, still us­ing NZ data in an es­sen­tial way, to give an ex­pli­citly com­put­able power series \( \Phi_\alpha^K(h) \) for any \( \alpha\in\mathbb{Q}/\mathbb{Z} \), with \( \Phi_0^K=\Phi^K \), that is ex­pec­ted to be the power series pre­dicted by the quantum mod­u­lar­ity con­jec­ture for knots that we will dis­cuss in the next sec­tion.

Fi­nally, it is worth not­ing that the NZ-equa­tions and their sym­plect­ic prop­er­ties lead to an ex­pli­cit quant­iz­a­tion of the shape vari­ables, where one re­places each \( z \), \( z^{\prime} \) and \( z^{\prime\prime} \) by op­er­at­ors that suit­ably com­mute. This was car­ried out by Dimofte [e21], who defined a quant­ized ver­sion of the glu­ing equa­tions, a so-called quantum curve, which is ex­pec­ted to an­ni­hil­ate the par­ti­tion func­tion of com­plex Chern–Si­mons the­ory and to be ul­ti­mately re­lated to the asymp­tot­ics of quantum in­vari­ants.

6. Connections to number theory

The pa­per [1] and its se­quel [2] sug­ges­ted or led to sev­er­al in­ter­est­ing de­vel­op­ments in pure num­ber the­ory as well as in to­po­logy. In this fi­nal sec­tion we de­scribe of a few of these.

Values of Dedekind zeta functions and higher Bloch groups

An im­port­ant sub­class of hy­per­bol­ic man­i­folds \( M=\mathbb{H}^3/\Gamma \) are the arith­met­ic ones, where \( \Gamma \) is either the Bi­an­chi group \( \operatorname{SL}_2(\mathcal{O}_F) \) for some ima­gin­ary quad­rat­ic field \( F \) or more gen­er­ally a group of units in a qua­ternion al­gebra over a num­ber field \( F \) of high­er de­gree \( n=r_1+2 \) hav­ing only one com­plex em­bed­ding up to com­plex con­jug­a­tion. In both cases, clas­sic­al res­ults (proved in the first case by Hum­bert already in 1919) say that the volume of  \( M \) is a simple mul­tiple (a power of \( \pi \) times the square-root of the dis­crim­in­ant) of the value at \( s=2 \) of the Dede­kind zeta func­tion \( \zeta_F(s) \) of the field \( F \). An im­me­di­ate con­sequence of this and of the volume for­mu­las dis­cussed in Sec­tion 2 is that this zeta value is a mul­tiple of a lin­ear com­bin­a­tion of val­ues of the Bloch–Wign­er di­log­ar­ithm at al­geb­ra­ic ar­gu­ments. This con­sequence was ob­served in [e3] and was also gen­er­al­ized there to the value of \( \zeta_F(2) \) for ar­bit­rary num­ber fields \( F \), with \( [F:\mathbb{Q}]=r_1+2r_2 \) for any value of \( r_2\ge1 \). (If \( r_2=0 \), then the well-known Klin­gen–Siegel the­or­em as­serts that \( \zeta_F(2) \) is a ra­tion­al mul­tiple of \( \pi^{2r_1}\sqrt{D_F} \).) Now the group \( \operatorname{SL}_2(\mathcal{O}_F) \) acts as a dis­crete group of iso­met­ries of \( (\mathbb{H}^2)^{r_1}\times(\mathbb{H}^3)^{r_2} \) with a quo­tient of fi­nite volume, and there are also qua­ternion­ic groups \( \Gamma \) over \( F \) that act freely and dis­cretely on \( (\mathbb{H}^3)^{r_2} \), the volume of the quo­tient in both cases be­ing an ele­ment­ary mul­tiple of \( \zeta_F(2) \). This gives a “poly-3-hy­per­bol­ic” man­i­fold \( M=(\mathbb{H}^3)^{r_2}/\Gamma \) with volume pro­por­tion­al to \( \zeta_F(2) \). A rather amus­ing lemma says that any such man­i­fold has a de­com­pos­i­tion (dis­joint ex­cept for the bound­ar­ies) in­to fi­nitely many \( r_2 \)-fold products of hy­per­bol­ic tet­ra­hedra, and it fol­lows that \( \zeta_F(2) \) for any num­ber field has an ex­pres­sion as a lin­ear com­bin­a­tion of \( r_2 \)-fold products of val­ues of \( D(z) \) at al­geb­ra­ic ar­gu­ments, gen­er­al­iz­ing the Klin­gen–Siegel the­or­em in an un­ex­pec­ted way.

The con­nec­tion with 3-di­men­sion­al hy­per­bol­ic geo­metry ap­plies only to the val­ues of Dede­kind zeta func­tions at \( s=2 \), but sug­ges­ted that there might be sim­il­ar state­ments for \( \zeta_F(m) \) with \( m > 2 \) in terms of the \( m \)-th poly­log­ar­ithm func­tion \( \operatorname{Li}_m(z) \). Ex­tens­ive nu­mer­ic­al ex­per­i­ments led to a con­crete con­jec­ture say­ing that this is the case and also to a defin­i­tion (ori­gin­ally highly spec­u­lat­ive, but now sup­por­ted by more the­ory) of “high­er Bloch groups” \( \mathcal{B}_m(F) \) that should be iso­morph­ic after tensor­ing with \( \mathbb{Q} \) to the high­er al­geb­ra­ic \( K \)-groups \( K_{2m-1}(F) \) and should ex­press the Borel reg­u­lat­or in terms of poly­log­ar­ithms. (For a sur­vey, see [e9].) This con­jec­ture, now over 30 years old, has been stud­ied ex­tens­ively by Beil­in­son, De­ligne, de Jeu, Gon­char­ov, Ruden­ko and oth­ers, with the cases \( m=3 \) and \( m=4 \) now be­ing es­sen­tially settled.

Units in cyclotomic extensions of number fields

As already men­tioned in the last sec­tion, the ana­lys­is of the Kashaev in­vari­ant and the mod­u­lar gen­er­al­iz­a­tion of the Volume Con­jec­ture dis­cussed be­low led to the defin­i­tion of cer­tain power series \( \Phi_\alpha^K(h) \) as­so­ci­ated to a hy­per­bol­ic knot and a num­ber \( \alpha\in\mathbb{Q}/\mathbb{Z} \) that can be com­puted nu­mer­ic­ally in any giv­en case. Ex­tens­ive nu­mer­ic­al com­pu­ta­tions for simple knots and simple ra­tion­al num­bers \( \alpha \) sug­ges­ted that this power series not only has al­geb­ra­ic coef­fi­cients, but that (up to a root of a unity and the square-root of a num­ber in the trace field \( F_K \) of the knot in­de­pend­ent of \( \alpha \)) its \( n \)-th power be­longs to \( F_{K,n}[\![h]\!] \), where \( n \) is the de­nom­in­at­or of \( \alpha \) and \( F_{K,n}=F_K(e^{2\pi i\alpha}) \) the \( n \)-th cyc­lo­tom­ic ex­ten­sion of \( F_K \). Equi­val­ently, \( \Phi_\alpha^K(h) \) it­self is the product of a power series in \( F_{K,n}[\![h]\!] \) with the \( n \)-th root of an ele­ment of \( F_{K,n}^\times \). Moreover, in each case stud­ied the lat­ter factor turned out to be the \( n \)-th root of a unit, and not just a nonzero num­ber, of \( F_{K,n} \), and the “sis­ter knots” (like \( 5_2 \) and the \( (-2,3,7) \)-pret­zel knot) hav­ing the same Bloch group class were the same for both knots, even though the rest of the power series were com­pletely dif­fer­ent. This led us, to­geth­er with Frank Calegari, to con­jec­ture and later to prove [e35] that there was a ca­non­ic­al class of ele­ments in cyc­lo­tom­ic ex­ten­sions of ar­bit­rary num­ber fields as­so­ci­ated to ele­ments of their Bloch groups, wheth­er or not the fields arise from to­po­logy. Ex­pli­citly, to any num­ber field \( F \) and any ele­ment \( \xi \) of the Bloch group of \( F \) one can as­so­ci­ate ca­non­ic­ally defined ele­ments of \( U(F_n)/U(F_n)^n \) for every \( n \), where \( U(F_n) \) de­notes the group of units (more pre­cisely, of \( S \)-units for some \( S \) de­pend­ing on \( \xi \) but in­de­pend­ent of \( n \)) of the \( n \)-th cyc­lo­tom­ic ex­ten­sion \( F_n \) of \( F \). Ac­tu­ally, two quite dif­fer­ent con­struc­tions were giv­en, one in terms of an ele­ment of \( \mathcal{B}(F) \) and one in terms of an ele­ment in \( K_3(F) \), and work in pro­gress an­nounced in [e35] sug­gests that there will be a gen­er­al­iz­a­tion to \( \mathcal{B}_m(F) \) and \( K_{2m-1}(F) \) for any \( m > 2 \). The con­struc­tion in terms of the Bloch group is quite simple, al­though the proof that it gives units and is in­de­pend­ent (up to \( n \)-th powers) of all choices is long: if \( \xi \) is rep­res­en­ted by \( \sum[z_i]\in\mathbb{Z}[F] \), then the num­ber \( \prod D_{\zeta}(z_i^{1/n}) \) is the product of an \( n \)-th power in \( F_n \) with an \( S \)-unit, and this is the unit we are look­ing for. Here \( \zeta \) is a prim­it­ive \( n \)-th root of unity and \[ D_{\zeta}(x)=\prod_{k=1}^{n-1}(1-\zeta^kx)^k \] is the “cyc­lic quantum di­log­ar­ithm” func­tion, which by a res­ult of Kashaev, Mangaz­eev and Strogan­ov sat­is­fies an ana­logue of the five-term re­la­tion of the clas­sic­al di­log­ar­ithm.

\( q \)-series and Nahm’s conjecture

An un­ex­pec­ted con­sequence of the work on units just de­scribed was a proof of one dir­ec­tion of a con­jec­ture by the math­em­at­ic­al phys­i­cist Wern­er Nahm that had pre­dicted an ex­tremely sur­pris­ing con­nec­tion between the Bloch group and the mod­u­lar­ity of cer­tain \( q \)-hy­per­geo­met­ric series. The simplest case of such a “Nahm sum” is the in­fin­ite series \begin{equation} \label{Nahm0} F_{a,b,c}(q)= \sum_{n=0}^\infty\frac{q^{\frac12an^2 +bn +c}}{(q)_n} \quad(a,b,c\in\mathbb{Q},\, a > 0), \tag{15} \end{equation} where \[ (q)_n=(1-q)(1-q^2)\cdots(1-q^n) \] is the so-called quantum factori­al. This func­tion is known to be mod­u­lar (in \( \tau \), where \( q=e^{2\pi i\tau} \)) when \( (a,b,c) \) is \( \bigl(2,0,-\frac1{60}\bigr) \) or \( \bigl(2,1,\frac{11}{60}\bigr) \) by the fam­ous Ro­gers–Ramanu­jan iden­tit­ies and in a hand­ful of oth­er cases by clas­sic­al res­ults of Euler, Gauss and oth­ers. It is very rare that a \( q \)-hy­per­geo­met­ric series (mean­ing an in­fin­ite sum whose ad­ja­cent terms dif­fer by fixed ra­tion­al func­tions of \( q \) and \( q^n \)) is at the same time a mod­u­lar func­tion, and in fact for the spe­cial series (15) this hap­pens only for sev­en triples \( (a,b,c) \), as pre­dicted by Nahm’s con­jec­ture and proved in [e14]. Nahm raised the gen­er­al ques­tion when a (mul­ti­di­men­sion­al) \( q \)-hy­per­geo­met­ric series can be mod­u­lar and, mo­tiv­ated by ex­amples com­ing from char­ac­ters of ver­tex op­er­at­or al­geb­ras, dis­covered a pos­sible an­swer in terms of the Bloch group. Con­cretely, he gen­er­al­ized (16) to \begin{equation} \label{Nahm1} F_{a,b,c}(q)= \sum_{n_1,\dots,n_N\ge0} \frac{q^{\frac12n^tan+b^tn+c}}{(q)_{n_1}\cdots(q)_{n_N}} \tag{16} \end{equation} for any \( N\ge1 \), where \( a \) is now a pos­it­ive def­in­ite sym­met­ric \( N\times N \) mat­rix with ra­tion­al coef­fi­cients, \( b \) a vec­tor in \( \mathbb{Q}^N \), and \( c \) a ra­tion­al num­ber. There is still no com­plete “if and only if” con­jec­ture pre­dict­ing ex­actly when these Nahm sums are mod­u­lar func­tions, but Nahm gave a pre­cise con­jec­ture for a ne­ces­sary con­di­tion and a par­tial con­jec­ture for the suf­fi­ciency. It is the first part that was proved in [e35] while the cor­rect for­mu­la­tion and proof of the con­verse dir­ec­tion is still an act­ive re­search sub­ject.

The mod­u­lar­ity cri­terion that Nahm found de­pended on his ob­ser­va­tion that for any solu­tion \( (z_1,\dots,z_N) \) of the sys­tem of equa­tions \begin{equation} \label{NahmEq} 1- z_i = \prod_{j=1}^N z_j^{a_{ij}} \quad(i=1,\dots,N) \tag{17} \end{equation} the ele­ment \( [z_1]+\cdots+[z_N] \) of \( \mathbb{Z}[\mathbb{C}] \) be­longs to \( \mathcal{B}(\mathbb{C}) \). This is a dir­ect con­sequence of the sym­metry of \( a \), be­cause \[ \sum_i(z_i)\wedge(1-z_i)=\sum_{i,j}a_{ij}(z_i)\wedge(z_j)=0. \] (This is the same ar­gu­ment as was used in Sec­tion 4 for the cor­res­pond­ing state­ment for the Neu­mann–Za­gi­er equa­tions, and ap­plies more gen­er­ally to all half-sym­plect­ic matrices, as dis­cussed be­low.) The “only if” dir­ec­tion of Nahm’s con­jec­ture then says that \( F_{a,b,c}(q) \) can be mod­u­lar only if this ele­ment of the Bloch group van­ishes for the unique solu­tion of the Nahm equa­tion hav­ing all \( z_i\in(0,1) \). The proof, giv­en in [e35] uses both the res­ults there about the units com­ing from Bloch ele­ments as de­scribed above and an asymp­tot­ic ana­lys­is of Nahm sums near roots of unity pub­lished sep­ar­ately by the two of us.

Half-symplectic matrices and the Bloch group

At the end of Sec­tion 4, we saw how the NZ equa­tions lead to a “half-sym­plect­ic mat­rix,” mean­ing the up­per half \( H=(A\,B) \) of a \( 2N\times2N \) sym­plect­ic mat­rix \[ \biggl(\,\begin{matrix} A & B\\ C & D\end{matrix}\biggr) \] over \( \mathbb{Z} \). To any such mat­rix we as­so­ci­ate the sys­tem of poly­no­mi­al equa­tions \begin{equation} \label{NZeq} \prod_{j=1}^N z_j^{A_{ij}} = (-1)^{(AB^t)_{ii}\vphantom{y_{y_y}}}\prod_{j=1}^N (1-z_j)^{B_{ij}} \quad(i=1,\dots,N). \tag{18} \end{equation} This is a straight gen­er­al­iz­a­tion of the ori­gin­al NZ equa­tions in the to­po­lo­gic­al set­ting ex­cept pos­sibly for the sign, but we checked in thou­sands of ex­amples us­ing SnapPy [e24] that this is the right sign, and this could pre­sum­ably be proved us­ing the “par­ity con­di­tion” in [2]. An­oth­er spe­cial case of (18) is the Nahm equa­tion (17), at least when the mat­rix \( a \) is in­teg­ral and even, the half-sym­plect­ic mat­rix then be­ing \( (1\,a) \) and the full sym­plect­ic mat­rix \[ \biggl(\,\begin{matrix}1 & a \\ 0 & 1\end{matrix}\biggr). \]

For any solu­tion \( z \) of (18), the ele­ment \( 2\sum_{j=1}^N[z_j] \) be­longs to the usu­al Bloch group by the same ar­gu­ment as was used for Nahm sums. We can in fact di­vide by 2 and lift to an ele­ment \( [H,z] \) of the ex­ten­ded Bloch group by set­ting \begin{equation} \label{Hz} [H,z] = \sum_{j=1}^N [u_j,v_j] + [\xi,\xi^{\prime}] + [-\xi,\xi^{\prime}-\xi+\pi i], \tag{19} \end{equation} where \( u_j \) and \( v_j \) are ar­bit­rary choices of log­ar­ithms of \( z_j \) and \( 1-z_j \), re­spect­ively, \( \xi \) is defined as \[ \textstyle\frac1\pi(Au-Bv)^t(Cu-Dv) \] for any com­ple­tion \( (C\,D) \)of \( H \) to a full sym­plect­ic mat­rix, and \( \xi^{\prime} \) is any choice of log­ar­ithm of \( 1-e^\xi \). The facts that this is in the ker­nel of \( d:\mathbb{Z}[\widehat{\mathbb{C}}]\to\mathsf{Li}^2(\mathbb{C}) \) and that its im­age mod­ulo ex­ten­ded five-term re­la­tions is in­de­pend­ent of all choices (at least mod­ulo 8-tor­sion in the Bloch group of the field gen­er­ated by the \( z \)’s) can be checked by dir­ect com­pu­ta­tions which are sketched in Sec­tion 6.1 of [e33], to­geth­er with a more pre­cise form that elim­in­ates the tor­sion am­bi­gu­ity.

We have the five fol­low­ing equi­val­ence re­la­tions among pairs, each mo­tiv­ated by a change of the choices made in the to­po­lo­gic­al situ­ation, that do not change this class:

  • Sta­bil­ity: in­crease \( N \) by \( N+1 \), re­place \( A \) and \( B \) in \( H=(A\,B) \) by their dir­ect sums with \( (1) \) and \( (0) \), re­spect­ively, and set \( z_{N+1}=1 \), cor­res­pond­ing in the to­po­lo­gic­al case to adding a de­gen­er­ate sim­plex to a 3-man­i­fold tri­an­gu­la­tion.
  • Chan­ging the equa­tions: mul­tiply \( H \) on the left by an ele­ment of \( \operatorname{GL}_N(\mathbb{Z}) \) without chan­ging the \( z \)’s. This cor­res­ponds to re­pla­cing the \( N \) re­la­tions (18) by mul­ti­plic­at­ive com­bin­a­tions of them in an in­vert­ible way.
  • Re­num­ber­ing: mul­tiply \( H \) on the right by an \( N\times N \) per­muta­tion mat­rix, and per­mute the \( z_j \)’s by the same mat­rix, cor­res­pond­ing to a re­num­ber­ing of the sim­plices of a tri­an­gu­la­tion.
  • New shape para­met­ers: for each \( j=1,\dots,N \), mul­tiply the \( N\times2 \) mat­rix \( (A_j\,B_j) \) (where \( A_j \) and \( B_j \) de­note the \( j \)-th columns of \( A \) and \( B \), re­spect­ively) by a power of the ele­ment \[ \biggl(\,\begin{matrix} 0 & -1\\ 1 & -1 \end{matrix}\biggr) \] of or­der 3 in \( \operatorname{SL}_2(\mathbb{Z}) \), and re­place \( z_j \) cor­res­pond­ingly by \( z_j \), \( z_j^{\prime}=1/(1-z_j) \), or \( z_j^{\prime\prime}=1/(1-z_j) \).
  • Al­geb­ra­ic 2-3 Pach­ner moves: re­move two columns of \( A \) and the same two columns of \( B \) and re­place them by three new columns that are spe­cif­ic \( \mathbb{Z} \)-lin­ear com­bin­a­tions, sim­ul­tan­eously re­pla­cing \( N \) by \( N+1 \) and chan­ging two of the \( z_j \) to three oth­ers in such a way that the cor­res­pond­ing change of \( [H,z] \) is a five-term re­la­tion. The ex­pli­cit for­mu­las were first writ­ten down in the spe­cial case cor­res­pond­ing to the Nahm sums (16) by Sander Zwegers in an un­pub­lished 2011 con­fer­ence talk and were then giv­en for ar­bit­rary sym­plect­ic matrices in equa­tion (3-27) of [e20]. This cor­res­ponds to sta­bil­iz­ing \( H \) three times and then mul­tiply­ing it on the left by a spe­cif­ic ele­ment of \( \operatorname{Sp}_{10}(\mathbb{Z}) \), and then un­stabil­iz­ing three times.

This gives us a new abeli­an group that maps to the ex­ten­ded Bloch group, namely the set of all pairs \( (H,z) \) as above mod­ulo these equi­val­ence re­la­tions, with ad­di­tion giv­en by dir­ect sum. In fact this map is an iso­morph­ism, mean­ing that any ele­ment of the Bloch group can be real­ized by some half-sym­plect­ic mat­rix and solu­tion of the cor­res­pond­ing gen­er­al­ized Neu­mann–Za­gi­er equa­tions and that any five-term re­la­tion can be lif­ted to one com­ing from an al­geb­ra­ic 2-3 Pach­ner move. A more com­plete dis­cus­sion is giv­en in Sec­tion 6.1 of [e33] (full de­tails will be giv­en later), while Sec­tion 6.3 of the same pa­per shows how to at­tach Nahm sum-like \( q \)-series to ar­bit­rary half-sym­plect­ic matrices.

From the Kashaev invariant to quantum modular forms

Nahm’s con­jec­ture already high­lighted a con­nec­tion between half-sym­plect­ic matrices and ques­tions of mod­u­lar­ity, but there are oth­er and more dir­ect con­nec­tions between hy­per­bol­ic 3-man­i­folds and the mod­u­lar group \( \operatorname{SL}_2(\mathbb{Z}) \) that we now de­scribe.

At the end of Sec­tion 5 we dis­cussed Kashaev’s volume con­jec­ture and its re­fine­ment (13). That state­ment in turn was gen­er­al­ized in [e19] on the basis of nu­mer­ic­al com­pu­ta­tions to a con­jec­tur­al asymp­tot­ic for­mula hav­ing a strong mod­u­lar fla­vor. To state it, we first note that the Kashaev in­vari­ant \( \langle K\rangle_n \) of a knot \( K \) can be gen­er­al­ized to a func­tion \( J^K:\mathbb{Q}/\mathbb{Z}\to\overline{\mathbb{Q}} \) whose value at \( -1/n \) for any \( n\ge1 \) is \( \langle K\rangle_n \) and which is \( \operatorname{Gal}(\overline{\mathbb{Q}}/\mathbb{Q}) \)-equivari­ant. Then the con­jec­tur­al gen­er­al­iz­a­tion of (13) is the state­ment that \begin{equation} \label{QMCb} J^K\biggl(\frac{an+b}{cn+d}\biggr) \sim (cn+d)^{3/2} e^{\operatorname{Vol}_\mathbb{C}(K)(n+d/c)/2\pi i} \Phi^K_{a/c}\biggl(\frac{2 \pi i}{c(cn+d)}\biggr) J^K(n) \tag{20} \end{equation} for every mat­rix \[ \biggl(\,\begin{matrix} a & b\\ c & d\end{matrix}\biggr) \in\operatorname{SL}_2(\mathbb{Z}) \] as \( n \) tends to in­fin­ity through either in­tegers or ra­tion­al num­bers with bounded de­nom­in­at­or, with (13) be­ing the spe­cial case when \[ \biggl(\,\begin{matrix} a & b\\ c & d\end{matrix}\biggr) = \biggl(\,\begin{matrix} 0 & -1\\ 1 & \phantom{-}0\end{matrix}\biggr) \] and \( n \) is in­teg­ral. Here \( \Phi^K_\alpha(h) \) for \( \alpha\in\mathbb{Q}/\mathbb{Z} \) is a power series that is con­jec­tured to be the one con­struc­ted in [e31] and dis­cussed at the end of Sec­tion 5.

In [e33], this mod­u­lar­ity con­jec­ture was veri­fied ex­per­i­ment­ally for a few knots to many terms and to a high de­gree of pre­ci­sion, and was also suc­cess­ively re­fined in sev­er­al dif­fer­ent dir­ec­tions, the fi­nal state­ment be­ing the ex­ist­ence of a whole mat­rix \( \boldsymbol{J}^K \) of \( \overline{\mathbb{Q}} \)-val­ued func­tions on \( \mathbb{Q}/\mathbb{Z} \) (gen­er­al­ized Kashaev in­vari­ants), which con­jec­tur­ally has much bet­ter mod­u­lar­ity prop­er­ties than the ori­gin­al scal­ar func­tion \( J^K \). Ex­pli­citly, (20) lifts to a sim­il­ar state­ment with \( J^K \) re­placed by the mat­rix \( \boldsymbol{J}^K \) and the com­pleted form­al power series \[ \widehat\Phi^K_{a/c}(h)=e^{\operatorname{Vol}_\mathbb{C}(K)/2\pi ic^2h}\Phi^K_{a/c}(h) \] by a mat­rix \( \widehat{\boldsymbol{\Phi}}^K_{a/c}(h) \) of com­pleted form­al power series act­ing by right mul­ti­plic­a­tion. The rows and columns of these matrices are in­dexed by the bounded para­bol­ic flat con­nec­tions, or equi­val­ently by an in­dex 0 (trivi­al con­nec­tion) and in­dices \( 1,\dots,r \) cor­res­pond­ing to the solu­tions of the NZ equa­tion for a tri­an­gu­la­tion of the knot com­ple­ment, with the ori­gin­al scal­ar-val­ued func­tions \( J^K \) and \( \Phi^K \) be­ing the \( (0,1) \) and \( (1,1) \) entries of \( \boldsymbol{J}^K \) and \( \boldsymbol{\Phi}^K \), re­spect­ively. The really new as­pect is that, by re­pla­cing the ori­gin­al scal­ar func­tions by matrices, we ob­tain a mat­rix of com­pleted form­al power series in \( h \) that (con­jec­tur­ally, like everything else in this story) ex­tend to real-ana­lyt­ic func­tions on the pos­it­ive and neg­at­ive real line and in fact to holo­morph­ic func­tions on the two cut planes \( \mathbb{C}\smallsetminus(-\infty,0] \) and \( \mathbb{C}\smallsetminus[0,\infty) \). This dis­cov­ery, which arises through the pos­sib­il­ity of as­so­ci­at­ing mat­rix-val­ued \( q \)-series to the knot [e34], gives rise to the new concept of “holo­morph­ic quantum mod­u­lar form” that then turned out to ap­pear also in many oth­er situ­ations, in­clud­ing vari­ous known mod­u­lar ob­jects like mock mod­u­lar forms or Ei­s­en­stein series of odd weight on the full mod­u­lar group.

Fi­nally, we men­tion that the new gen­er­al­ized Kashaev in­vari­ants have beau­ti­ful arith­met­ic prop­er­ties gen­er­al­iz­ing the known prop­erty [e10] that the ori­gin­al Kashaev in­vari­ant be­longs to the so-called Habiro ring \[ \mathcal{H}=\varprojlim\mathbb{Z}[q]/(q)_n. \] The res­ults of [e33] and [e34] sug­gest that there should be a Habiro ring \( \mathcal{H}_F \) as­so­ci­ated to any num­ber field \( F \) in which the gen­er­al­ized Kashaev in­vari­ants take their val­ues and which is graded by the Bloch group of \( F \). (The lat­ter prop­erty is in­vis­ible in the clas­sic­al case since \( \mathcal{B}(\mathbb{Q})\otimes\mathbb{Q}=\{0\} \).) We are cur­rently work­ing on this jointly with Peter Scholze, and already have a can­did­ate for \( \mathcal{H}_F \), as well as a par­tial lift­ing of the al­geb­ra­ic units of [e35] to form­al power series with Habiro-like prop­er­ties.

Works

[1] W. D. Neu­mann and D. Za­gi­er: “Volumes of hy­per­bol­ic three-man­i­folds,” To­po­logy 24 : 3 (1985), pp. 307–​332. MR 815482 Zbl 0589.​57015 article

[2] W. D. Neu­mann: “Com­bin­at­or­ics of tri­an­gu­la­tions and the Chern–Si­mons in­vari­ant for hy­per­bol­ic 3-man­i­folds,” pp. 243–​271 in To­po­logy ’90: Pa­pers from the re­search semester in low-di­men­sion­al to­po­logy held at Ohio State Uni­versity (Colum­bus, OH, Feb­ru­ary–June 1990). Edi­ted by B. Apanasov, W. D. Neu­mann, A. W. Re­id, and L. Sieben­mann. Ohio State Uni­versity Math­em­at­ics Re­search In­sti­tute Pub­lic­a­tions 1. de Gruyter (Ber­lin), 1992. MR 1184415 Zbl 0768.​57006 incollection

[3] W. D. Neu­mann: “Hil­bert’s 3rd prob­lem and in­vari­ants of 3-man­i­folds,” pp. 383–​411 in The Ep­stein birth­day schrift. Edi­ted by I. Riv­in, C. Rourke, and C. Series. Geo­metry and To­po­logy Mono­graphs 1. Geo­metry and To­po­logy Pub­lish­ers (Cov­entry, UK), 1998. Ded­ic­ated to Dav­id Ep­stein on the oc­ca­sion of his 60th birth­day. MR 1668316 Zbl 0902.​57013 ArXiv math/​9712226 incollection

[4] W. D. Neu­mann: “Ex­ten­ded Bloch group and the Chee­ger–Chern–Si­mons class,” Geom. To­pol. 8 : 1 (2004), pp. 413–​474. MR 2033484 Zbl 1053.​57010 ArXiv math/​0307092 article