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

Mixed Hodge Structures and Renormalization in Physics

Spencer Bloch Dept. of Mathematics, University of Chicago, Chicago, IL 60637, USA
E-mail address: bloch@math.uchicago.edu
 and  Dirk Kreimer CNRS-IHES, 91440 Bures sur Yvette, France and Center for Math. Phys.  Boston U., Boston, MA 02215
E-mail address: kreimer@ihes.fr

1. Introduction

1.1.

This paper is a collaboration between a mathematician and a physicist. It is based on the observation that renormalization of Feynman amplitudes in physics is closely related to the theory of limiting mixed Hodge structures in mathematics. Whereas classical physical renormalization methods involve manipulations with the integrand of a divergent integral, limiting Hodge theory involves moving the chain of integration so the integral becomes convergent and studying the monodromy as the chain varies.

Even methods like minimal subtraction in the context of dimensional or analytic regularization implicitly modify the integrand through the definition of a measure dDk\int d^{D}k via analytic continuation. Still, as a regulator dimensional regularization is close to our approach in so far as it leaves the rational integrand assigned to a graph unchanged. Minimal subtraction as a renormalization scheme differs though from the renormalization schemes which we consider -momentum subtractions essentially- by a finite renormalization. Many of the nice algebro-geometric structures developed below are not transparent in that scheme.

The advantages of the limiting Hodge method are firstly that it is linked to a very central and powerful program in mathematics: the study of Hodge structures and their variations. As a consequence, one gains a number of tools, like weight, Hodge, and monodromy filtrations to study and classify the Feynman amplitudes. Secondly, the method depends on the integration chain, and hence on the graph, but it is in some sense independent of the integrand. For this reason it should adapt naturally e.g. to gauge theories where the numerator of the integrand is complicated.

An important point is to analyse the nature of the poles. Limiting mixed Hodge structures demand that the divergent subintegrals have at worst log poles. This does not imply that we can not apply our approach to perturbative amplitudes which have worse than logarithmic degree of divergence. It only means that we have to correctly isolate the polynomials in masses and external momenta which accompany those divergences such that the corresponding integrands have singularities provided by log-poles. This is essentially automatic from the notion of a residue available by our very methods. As a very pleasant byproduct, we learn that physical renormalization schemes -on-shell subtractions, momentum subtractions, Weinberg’s scheme,- belong to a class of schemes for which this is indeed automatic.

Moreover, for technical reasons, it is convenient to work with projective rather than affine integrals. One of the central physics results in this paper is that the renormalization problem can be reduced to the study of logarithmically divergent, projective integrals. This is again familiar from analytic regulators. The fact that it can be achieved here by leaving the integrand completely intact will hopefully some fine day allow to understand the nature of the periods assigned to renormalized values in quantum field theory.

A remark for Mathematicians: our focus in this paper has been renormalization, which is a problem arising in physics. We suspect, however, that similar methods will apply more generally for example to period integrals whenever the domain of integration is contained in +n{\mathbb{R}}^{+n} and the integrand is a rational function with polar locus defined by a polynomial with non-negative real coefficients. The toric methods and the monodromy computations should go through in that generality.

Acknowledgments

Both authors thank Francis Brown, Hélène Esnault and Karen Yeats for helpful discussions. This work was partially supported by NSF grants DMS-0603781 and DMS-0653004. S.B. thanks the IHES for hospitality January-March 2006 and January-March 2008. D.K. thanks Chicago University for hospitality in February 2007.

1.2. Physics Introduction

This paper studies the renormalization problem in the context of parametric representations, with an emphasis on algebro-geometric properties. We will not study the nature of the periods one obtains from renormalizable quantum field theories in an even dimension of space-time. Instead, we provide the combinatorics of renormalization such that a future motivic analysis of renormalized amplitudes is feasible along the lines of [2]. Our result will in particular put renormalization in the framework of a limiting mixed Hodge structure, which hopefully provides a good starting point for an analysis of the periods in renormalized amplitudes. That these amplitudes are provided by numbers which are periods (in the sense of [11]) is an immediate consequence of the properties of parametric representations, and will also emerge naturally below (see Thm.(7.3)).

The main result of this paper is a careful study of the singularities of the first Kirchhoff–Symanzik polynomial, which carries all the short-distance singularities of the theory. The study of this polynomial can proceed via an analysis with the help of projective integrals. Along the way, we will also give useful formulas for parametric representations involving affine integrals, and clarify the role of the second Kirchhoff–Symanzik polynomial for affine and projective integrals.

Our methods are general, but in concrete examples we restrict ourselves to ϕ44\phi^{4}_{4} theory. Parametric representations are used which result from free-field propagators for propagation in flat space-time. In such circumstances, the advantages of analytic regularizations are also available in our study of parametric representations as we will see. In particular, our use of projective integrals below combines such advantages with the possibility to discuss renormalization on the level of the pairing between integration chains and de Rham classes.

In examples, special emphasis is given to the study of particular renormalization schemes, the momentum scheme (MOM-scheme, Weinberg’s scheme, on-shell subtractions).

Also, we often consider Green functions as functions of a single kinematical scale q2>0q^{2}>0. Green functions are defined throughout as the scalar coefficient functions (structure functions) for the radiative corrections to tree-level amplitudes rr. They are to be regarded as scalar quantities of the form 1+𝒪()1+{\mathcal{O}}(\hbar). Renormalized amplitudes are then, in finite order in perturbation theory, polynomial corrections in L=lnq2/μ2L=\ln q^{2}/\mu^{2} (μ2>0\mu^{2}>0) without constant term, providing the quantum corrections to the tree-level amplitudes appearing as monomials in a renormalizable Lagrangian [15]:

(1.1) ϕR(Γ)=j=1aug(Γ)pj(Γ)Lj.\phi_{R}(\Gamma)=\sum_{j=1}^{{\textrm{aug}}(\Gamma)}p_{j}(\Gamma)L^{j}.

Correspondingly, Green functions become triangular series in two variables

(1.2) Gr(α,L)=1+j=1γjr(α)Lj=1+j=1cjr(L)αj.G^{r}(\alpha,L)=1+\sum_{j=1}^{\infty}\gamma_{j}^{r}(\alpha)L^{j}=1+\sum_{j=1}^{\infty}c_{j}^{r}(L)\alpha^{j}.

The series γjr(α)\gamma_{j}^{r}(\alpha) are related by the renormalization group which leaves only the γ1r(α)\gamma_{1}^{r}(\alpha) undetermined, while the polynomials cjr(L)c_{j}^{r}(L) are bounded in degree by jj. The series γ1r\gamma_{1}^{r} fulfill ordinary differential equations driven by the primitive graphs of the theory [16].

The limiting Hodge structure A(Γ)A(\Gamma) which we consider for each Feynman graph Γ\Gamma provides contribution of a graph Γ\Gamma to the coefficients of γ1r\gamma_{1}^{r} in the limit. This limit is a period matrix (a column vector here) which has, from top to bottom, the periods provided by a renormalized graph Γ\Gamma as entries. The first entry is the contribution to γ1r\gamma_{1}^{r} of a graph with res(Γ)=r\textrm{res}(\Gamma)=r and the kk-th is a rational multiple of the contribution to γkr\gamma_{k}^{r}. In section 9.1 we determine the rational weights which connect these periods to the coefficients pj(Γ)p_{j}(\Gamma) attributed to the renormalization of a graph Γ\Gamma.

We include a discussion of the structure of renormalization which comes from an analysis of the second Kirchhoff–Symanzik polynomial. While this polynomial does not provide short-distance singularities in its own right, it leads to integrals of the form

(1.3) ωln(f)\int\omega\ln(f)

for a renormalized Feynman amplitude, with ω\omega a de Rham class determined by the first Kirchhoff–Symanzik polynomial, and ff -congruent to one along any remaining exceptional divisor- determined by the second. We do not actually do the monodromy calculation for integrals (1.3) involving a logarithm, but it will be similar to the calculation for (1.5) which we do. A full discussion of the Hodge structure of a Green function seems feasible but will be postponed to future work.

1.3. Math Introduction

Let n1{\mathbb{P}}^{n-1} be the projective space of lines in n{\mathbb{C}}^{n} which we view as an algebraic variety with homogeneous coordinates A1,,AnA_{1},\dotsc,A_{n}. Let ψ(A1,,An)\psi(A_{1},\dotsc,A_{n}) be a homogeneous polynomial of some degree dd, and let Xn1X\subset{\mathbb{P}}^{n-1} be the hypersurface defined by ψ=0\psi=0. We assume the coefficients of ψ\psi are all real and 0\geq 0. Let σ={[a1,,an]|ai0,i}\sigma=\{[a_{1},\dotsc,a_{n}]\ |\ a_{i}\geq 0,\forall i\} be the topological (n1)(n-1)-chain (simplex) in n1{\mathbb{P}}^{n-1}, where [][\ldots] refers to homogeneous coordinates. We will also use the notation σ=n1(0)\sigma={\mathbb{P}}^{n-1}({\mathbb{R}}^{\geq 0}). Our assumption about coefficients implies

(1.4) σX=LXL(0),\sigma\cap X=\bigcup_{L\subset X}L({\mathbb{R}}^{\geq 0}),

where LL runs through all coordinate coordinate linear spaces L:Ai1==Aip=0L:A_{i_{1}}=\cdots=A_{i_{p}}=0 contained in XX (see (see Fig.1)).

Refer to caption

Figure 1. Picture of XX and LL

The genesis of the renormalization problem in physics is the need to assign values to integrals

(1.5) σω\int_{\sigma}\omega

where ω\omega is an algebraic (n1)(n-1)-form on n1{\mathbb{P}}^{n-1} with poles along XX. The problem is an important one for physical applications, and there is an extensive literature (see, for example, [10, 22, 21]) focusing on practical formulae to reinterpret (1.5) in some consistent way as a polynomial in logt\log t. (Here tt parametrizes a deformation of the integration chain. As a first approximation, one can think of tω\int_{t}^{\infty}\omega when ω\omega has a logarithmic pole at t=0t=0.)

A similar problem arises in pure mathematics in the study of degenerating varieties, e.g. a family of elliptic curves degenerating to a rational curve with a node. In the classical setup, one is given a family f:𝒳Df:{\mathcal{X}}\to D, where DD is a disk with parameter tt. The map ff is proper (so the fibres XtX_{t} are compact). 𝒳{\mathcal{X}} is assumed to be non-singular, as are the fibres Xt,t0X_{t},\ t\neq 0. X0X_{0} may be singular, though one commonly invokes resolution of singularities to assume X0𝒳X_{0}\subset{\mathcal{X}} is a normal crossing divisor. Choose a basis σ1,t,,σr,t\sigma_{1,t},\dotsc,\sigma_{r,t} for the homology of the fibre Hp(Xt,)H_{p}(X_{t},{\mathbb{Q}}) in some fixed degree pp. By standard results in differential topology, the fibre space is locally topologically trivial over D=D{0}D^{*}=D-\{0\}, and we may choose the classes σi,t\sigma_{i,t} to be locally constant. If we fix a smooth fibre t00t_{0}\neq 0, the monodromy transformation m:Hp(Xt0)Hp(Xt0)m:H_{p}(X_{t_{0}})\to H_{p}(X_{t_{0}}) is obtained by winding around t=0t=0. An important theorem ([7], III,2) says this transformation is quasi-unipotent, i.e. after possibly introducing a root t=t1/nt^{\prime}=t^{1/n} (which has the effect of replacing mm by mnm^{n}), midm-id is nilpotent. The matrix

(1.6) N:=logm=[(idm)+(idm)2/2+]N:=\log m=-\Big{[}(id-m)+(id-m)^{2}/2+\ldots\Big{]}

is thus also nilpotent. This is the mathematical equivalent of locality in physics. It insures that our renormalization of (1.5) will be a polynomial in logt\log t rather than an infinite series. We take a cohomology class [ωt]Hp(Xt,)[\omega_{t}]\in H^{p}(X_{t},{\mathbb{C}}) which varies algebraically. For example, in a family of elliptic curves y2=x(x1)(xt)y^{2}=x(x-1)(x-t), the holomorphic 11-form ωt=dx/y\omega_{t}=dx/y is such a class. Note ωt\omega_{t} is single-valued over all of DD^{*}. It is not locally constant. The expression

(1.7) exp((Nlogt)/2πi)(σ1,tωtσr,tωt).\exp\Big{(}-(N\log t)/2\pi i\Big{)}\begin{pmatrix}\int_{\sigma_{1,t}}\omega_{t}\\ \vdots\\ \int_{\sigma_{r,t}}\omega_{t}\end{pmatrix}.

is then single-valued and analytic on DD^{*}. Suppose ωt\omega_{t} chosen such that the entries of the column vector in (1.7) grow at worst like powers of |log|t|||\log|t|| as |t|0|t|\to 0. A standard result in complex analysis then implies that (1.7) is analytic at t=0t=0. We can write this

(1.8) (σ1,tωtσr,tωt)exp((Nlogt)/2πi)(a1ar).\begin{pmatrix}\int_{\sigma_{1,t}}\omega_{t}\\ \vdots\\ \int_{\sigma_{r,t}}\omega_{t}\end{pmatrix}\sim\exp\Big{(}(N\log t)/2\pi i\Big{)}\begin{pmatrix}a_{1}\\ \vdots\\ a_{r}\end{pmatrix}.

Here the aja_{j} are constants which are periods of a limiting Hodge structure. The exponential on the right expands as a matrix whose entries are polynomials in logt\log t, and the equivalence relation \sim means that the difference between the two sides is a column vector of (multi-valued) analytic functions vanishing at t=0t=0.

We would like to apply this program to the integral (1.5). Let Δ:1nAj=0\Delta:\prod_{1}^{n}A_{j}=0 be the coordinate divisor in n1{\mathbb{P}}^{n-1}. Note that the chain σ\sigma has boundary in Δ\Delta, so as a first attempt to interpret (1.5) as a period, we might consider the pairing

(1.9) Hn1(n1X,ΔXΔ)×Hn1(n1X,ΔXΔ)H^{n-1}({\mathbb{P}}^{n-1}-X,\Delta-X\cap\Delta)\times H_{n-1}({\mathbb{P}}^{n-1}-X,\Delta-X\cap\Delta)\to{\mathbb{C}}

The form ω\omega is an algebraic (n1)(n-1)-form and it vanishes on Δ\Delta for degree reasons, so it does give a class in the relative cohomology group appearing in (1.9) (see the discussion (9.8)-(9.10)). On the other hand, the chain σ\sigma meets XX (1.4), so we do not get a class in homology. Instead we consider a family of coordinate divisors Δt:1nAj,t=0\Delta_{t}:\prod_{1}^{n}A_{j,t}=0 with Δ0=Δ\Delta_{0}=\Delta. (For details, see section 6.) For t=ε>0t=\varepsilon>0 there is a natural chain σε\sigma_{\varepsilon} which is what the physicists would call a cutoff. We have σεΔε\partial\sigma_{\varepsilon}\subset\Delta_{\varepsilon} and σεX=\sigma_{\varepsilon}\cap X=\emptyset, so σεω\int_{\sigma_{\varepsilon}}\omega is defined. One knows on abstract grounds that the monodromy of

Hn1(n1X,ΔtXΔt)H_{n-1}({\mathbb{P}}^{n-1}-X,\Delta_{t}-X\cap\Delta_{t})

is quasi-unipotent as above ([7], III,§2). The main mathematical work in this paper will be to compute the monodromy of σε\sigma_{\varepsilon} in the specific case of Feynman amplitudes in physics. More precisely, XX will be a graph hypersurface XΓX_{\Gamma} associated to a graph Γ\Gamma (section 5). We will write down chains τγε\tau_{\gamma}^{\varepsilon}, one for each flag of core (one particle irreducible in physics) subgraphs γ={Γ1Γp(γ)Γ}\gamma=\{\Gamma_{1}\subsetneq\cdots\Gamma_{p(\gamma)}\subsetneq\Gamma\}, representing linearly independent homology classes in Hn1(n1X,ΔεXΔε)H_{n-1}({\mathbb{P}}^{n-1}-X,\Delta_{\varepsilon}-X\cap\Delta_{\varepsilon}). (The combinatorics here is similar to that found in [1], [18].) We will show the monodromy in our case is given by

(1.10) m(σε)=σε+γ(1)p(γ)τγε.m(\sigma_{\varepsilon})=\sigma_{\varepsilon}+\sum_{\gamma}(-1)^{p(\gamma)}\tau_{\gamma}^{\varepsilon}.

We will then exhibit a nilpotent matrix NN such that

(1.11) (m(σε)m(τγε))=exp(N)(σετγε).\begin{pmatrix}m(\sigma_{\varepsilon})\\ \vdots\\ m(\tau_{\gamma}^{\varepsilon})\\ \vdots\end{pmatrix}=\exp(N)\begin{pmatrix}\sigma_{\varepsilon}\\ \vdots\\ \tau_{\gamma}^{\varepsilon}\\ \vdots\end{pmatrix}.

With this in hand, renormalization is automatic for any physical theory for which Γ\Gamma and its subgraphs are at worst logarithmically divergent after taking out suitable polynomials in masses and momenta. Namely, such a physical theory gives a differential form ωΓ\omega_{\Gamma} as in (1.5) and we may repeat the above argument:

(1.12) exp((Nlogt)/2πi)(σtωτγtω)\exp(-(N\log t)/2\pi i)\begin{pmatrix}\int_{\sigma_{t}}\omega\\ \vdots\\ \int_{\tau_{\gamma}^{t}}\omega\\ \vdots\end{pmatrix}

is single-valued on the punctured disk. The hypothesis of log divergence at worst for subgraphs of Γ\Gamma will imply that the integrals will grow at worst like a power of log\log as |t|0|t|\to 0,(lemma 9.2). Precisely as in (1.8), one gets the renormalization

(1.13) σtωΓ=k=0rbk(logt)k+O(t),\int_{\sigma_{t}}\omega_{\Gamma}=\sum_{k=0}^{r}b_{k}(\log t)^{k}+O(t),

where O(t)O(t) denotes a (multi-valued) analytic function vanishing at t=0t=0. The renormalization schemes considered here can be characterized by the condition b0=0b_{0}=0.

Of course, the requirement that a physical theory have at worst log divergences is a very strong constraint. The difficult computations in section 7 show how general divergences encountered in physics can be reduced to log divergences.

Remarks 1.1.

The renormalization scheme outlined above, and worked out in detail in the following sections, has a number of properties, some of which may seem strange to the physicist.
(i) It does not work in renormalization schemes which demand counter-terms which are not defined by subtractions at fixed values of masses and momenta of the theory. So conditions on the regulator for example, as in minimal subtraction where one defines the counterterm by projection onto a pole part, are not considered. In such schemes, and for graphs which are worse than log divergent, a topological procedure of the sort given here can not work. It is necessary instead to modify the integrand ωΓ\omega_{\Gamma} in a non-canonical way.
(ii) On the other hand, our approach is very canonical. It depends on the choice of a parameter tt, as any renormalization scheme must. Somewhat more subtle is the dependence on the monodromy associated to the choice of a family Δt\Delta_{t} of coordinate divisors deforming the given Δ=Δ0\Delta=\Delta_{0}. We have taken the most evident such monodromy, moving all the vertices of the simplex. Note that this choice is stable in the sense that a small deformation leaves the monodromy unchanged.
(iii) It would seem that our answer is much more complicated than need be, because Γ\Gamma will in general contain far more core subgraphs than divergent subgraphs. For example, in φ4\varphi^{4}-theory, the “dunce’s cap” (see Fig.2)

Refer to caption

Figure 2. Dunce’s cap. Here and in following figures, external half-edges are often not drawn and are determined by the requirement that all vertices are four-valent.

has only one divergent subgraph, given in the picture by edges 1,21,2. It has 33 core subgraphs (3,4,1),(3,4,2),(1,2)(3,4,1),(3,4,2),(1,2). From the point of view of renormalization, this problem disappears. The τγε\tau^{\varepsilon}_{\gamma} are tubes, and the integral τγεωΓ\int_{\tau^{\varepsilon}_{\gamma}}\omega_{\Gamma} is basically a residue which will vanish unless γΓ\gamma\subset\Gamma is a divergent subgraph. In (1.12), the column vector of integrals will consist mostly of 0’s and the final regularization (1.13) will involve only divergent subgraphs.
(iv) An important property of the theory is the presence of a limiting mixed Hodge structure. The constants on the right hand side of (1.8) are periods of a mixed Hodge structure called the limiting MHS for the degeneration. One may hope that the tendency for Feynman amplitudes to be multi-zeta numbers [4] will some day be understood in terms of this Hodge structure. From the point of view of this paper, the vector space WHn1(n1XΓ,ΔtXΓΔt)W\subset H_{n-1}({\mathbb{P}}^{n-1}-X_{\Gamma},\Delta_{t}-X_{\Gamma}\cap\Delta_{t}) spanned by σt\sigma_{t} and the τγt\tau^{t}_{\gamma} is invariant under the monodromy. One may ask whether the image of WW in the limiting MHS spans a sub-Hodge structure. If so, we would expect that this HS would be linked to the multi-zeta numbers. Note that WW is highly non-trivial even when Γ\Gamma has no subdivergences. This WW is an essentially new invariant which comes out of the monodromy. See section (9.2) for a final discussion of our viewpoint.
(v) There are a number of renormalization schemes in physics, some of which are not compatible with our approach. One general test is that our scheme depends only on the graph polynomials of Γ\Gamma. For example, suppose Γ=Γ1Γ2\Gamma=\Gamma_{1}\cup\Gamma_{2} where the Γi\Gamma_{i} meet at a single vertex. Then the renormalization polynomial in logt\log t our theory yields for Γ\Gamma will be the product of the renormalizations for the Γi\Gamma_{i}.

Most of the mathematical work involved concerns the calculation of monodromy for a particular topological chain. It is perhaps worth taking a minute to discuss a toy model. Suppose one wants to calculate 0ω\int_{0}^{\infty}\omega, where ω=dz(zi)z\omega=\frac{dz}{(z-i)z}. The integral diverges, so instead we consider tω\int_{t}^{\infty}\omega as a function of t=εeiθt=\varepsilon e^{i\theta} for 0θ2π0\leq\theta\leq 2\pi. If we take the path [t,][t,\infty] to be a great circle, then as tt winds around 0, the path will get tangled in the singularity of ω\omega at z=iz=i. Assuming we do not understand the singularities of our integral far from 0, this could be a problem. Instead we chose our path to follow the small circle from εeiθ\varepsilon e^{i\theta} to ε\varepsilon and then the positive real axis from ε\varepsilon to \infty. The variation of monodromy is the difference in the paths for θ=0\theta=0 and θ=2π\theta=2\pi. In this case, it is the circle {|t|=ε}\{|t|=\varepsilon\}. If we assume something (at worst superficial log divergence for the given graph and all subgraphs in the given physical theory) about the behavior of ω\omega near the pole at 0, then the behavior of our integral for |t|<<1|t|<<1 is determined by this monodromy, which is a topological invariant. This is quite different from the usual approach in physics involving complicated algebraic manipulations with ω\omega. A glance at fig.(10) suggests that our toy model is too simple. We have to work with two scales, ε<<η<<1\varepsilon<<\eta<<1. This is because in the more complicated situation, we have to deal with cylinders of small radius η\eta, but then we have further to slightly deform the boundaries of the cylinder (cf. fig.(12)).

1.4. Leitfaden

Section 2 is devoted to Hopf algebras of graphs and of trees. These have played a central role in the combinatorics of renormalization. In particular, the insight afforded by passing from graphs to trees is important. Since the combinatorics of core subgraphs is even more complicated than that of divergent subgraphs, it seemed worth going carefully through the construction. Section 3 studies the toric variety we obtain from a graph Γ\Gamma by blowing up certain coordinate linear spaces in the projective space with homogeneous coordinates labeled by the edges of Γ\Gamma. The orbits of the torus action are related to flags of core subgraphs of the given graph. In section 4, we use the {\mathbb{R}}-structure on our toric variety to construct certain topological chains which will be used to explicit the monodromy. Section 5 recalls the basic properties of the graph polynomial ψΓψ(Γ)\psi_{\Gamma}\equiv\psi(\Gamma) and the graph hypersurface XΓ:ψΓ=0X_{\Gamma}:\psi_{\Gamma}=0. The crucial point is corollary 5.3 which says that the strict transform of XΓX_{\Gamma} on our toric blowup avoids points with coordinates 0\geq 0. Any chain we construct which stays close to the locus of such points necessarily is away from XΓX_{\Gamma} and hence also away from the polar locus of our integrand. Section 6 computes the monodromy of our chain. Section 7 considers how to reduce Feynman amplitude calculations as they arise in physics, including masses and momenta as well as divergences which are worse than logarithmic, to the basic situation where limiting methods can apply. In section 8 we calculate the nilpotent matrix NN which is the log of the monodromy transformation, and in section 9 we prove the main renormalization theorem in the log divergent case, to which we have reduced the theory.

2. Hopf algebras of trees and graphs

2.1. Graphs

In this section we bring together material on graphs and the graph Hopf algebra which will be used in the sequel. We also discuss Hopf algebras related to rooted trees and prove a result (proposition 2.5) relating the Hopf algebra of core graphs to a suitable Hopf algebra of labeled trees. Strictly speaking this is not used in the paper, but it provides the best way we know to understand flags of core subgraphs, and these play a central role in the monodromy computations. Trees labeled by divergent subgraphs have a long history in renormalization theory [12], [13].

A graph Γ\Gamma is determined by giving a finite set HE(Γ)HE(\Gamma) of half-edges, together with two further sets E(Γ)E(\Gamma) (edges) and V(Γ)V(\Gamma) (vertices) and surjective maps

(2.1) pV:HE(Γ)V;pE:HE(Γ)E.p_{V}:HE(\Gamma)\to V;\quad p_{E}:HE(\Gamma)\to E.

(Note we do not allow isolated vertices.) In combinatorics, one typically assumes all fibres pE1(e)p_{E}^{-1}(e) consist of exactly two half-edges (ee an internal edge), while in physics the calculus of path integrals and correlation functions dictates that one admit external edges eEe\in E with #pE1(e)=1\#p_{E}^{-1}(e)=1. If all internal edges of Γ\Gamma are shrunk to 0, the resulting graph (with no internal edges) is called the residue res(Γ)\text{res}(\Gamma). In certain theories, the vertices are decomposed into different types V=ViV=\amalg V_{i}, and the valence of the vertices in ViV_{i}, #pV1(v)\#p_{V}^{-1}(v), is fixed independent of vViv\in V_{i}.

We will typically work with labeled graphs which are triples (Γ,A,ϕ:AE(Γ))(\Gamma,A,\phi:A\cong E(\Gamma)). We refer to AA as the set of edges.

A graph is a topological space with Betti numbers |Γ|=h1(Γ)=dimH1(Γ,)|\Gamma|=h_{1}(\Gamma)=\dim H_{1}(\Gamma,{\mathbb{Q}}) and h0(Γ)h_{0}(\Gamma). We say Γ\Gamma is connected if h0=1h_{0}=1. Sometimes h1h_{1} is referred to as the loop number.

A subgraph γΓ\gamma\subset\Gamma is determined (for us) by a subset E(γ)E(Γ)E(\gamma)\subset E(\Gamma). We write Γ//γ\Gamma/\!/\gamma for the quotient graph obtained by contracting all edges of γ\gamma to points. If γ\gamma is not connected, Γ//γ\Gamma/\!/\gamma is different from the naive quotient Γ/γ\Gamma/\gamma. If γ=Γ\gamma=\Gamma, we take Γ//Γ=\Gamma/\!/\Gamma=\emptyset to be the empty set. It will be convenient when we discuss Hopf algebras below to have the empty set as a graph.

Also, for γ=e\gamma=e a single edge, we have the contraction Γ//e=Γ/e\Gamma/\!/e=\Gamma/e. In this case we also consider the cut graph Γe\Gamma-e obtained by removing ee and also any remaining isolated vertex.

A graph Γ\Gamma is said to be core (1PI1PI in physics terminology) if for any edge ee we have |Γe|<|Γ||\Gamma-e|<|\Gamma|.

A cycle γΓ\gamma\subset\Gamma is a core subgraph such that |γ|=1|\gamma|=1. If Γ\Gamma is core, it can be written as a union of cycles (see e.g. the proof of lemma 7.4 in [2]).

2.1.1. Self-energy graphs

Special care has to be taken when the residue res(γ)\textrm{res}(\gamma) of a connected component γ\gamma of some subgraph consists of two half-edges connected to a vertex, |res(γ)|=2|\textrm{res}(\gamma)|=2. Such graphs are called self-energy graphs in physics. In such a situation, if the internal edges of γ\gamma contract to a point, we are left with two edges in Γ//γ\Gamma/\!/\gamma, which are connected at this point uu . It might happen that the theory provides more than one two-point vertex. In fact, for a massive theory, there are two two-point vertices provided by the theory corresponding to the two monomials in the Lagrangian quadratic in the fields, we call them of mass and kinetic type. Γ//γ\Gamma/\!/\gamma represents then a sum over two graphs by summing over the two types of vertices for that point uu. (see Fig.(3) for an example).

Refer to caption

Figure 3. This vertex graph has a propagator correction given by edges 4,5,64,5,6. The non-trivial part of the coproduct then delivers on the left the subgraph with internal edges 4,5,64,5,6 amongst other terms. The coproduct on the right has a co-graph on edges 1,2,31,2,3. There is a two-point vertex uu between edges 2,32,3. Choosing two labels u=m2u=m^{2} or u=u=\Box allows to distinguish between mass and wave-function renormalization. We remind the reader that the corresponding monomials in the Lagrangian are m2ϕ2/2m^{2}\phi^{2}/2 and ϕϕ/2\phi\Box\phi/2.

The edges and vertices of various types have weights. We set the weight of an edge to be two, the weight of a vertex with valence greater than two is zero, the weight of a vertex of mass type is zero, the weight of the kinetic type is +2.

Then, the superficial degree of divergence sdd(Γ)\textrm{sdd}(\Gamma) for a connected core graph Γ\Gamma is

(2.2) sdd(Γ)=4|Γ|2|Γ[1]|+2|Γ[0],kin|,\textrm{sdd}(\Gamma)=4|\Gamma|-2|\Gamma^{[1]}|+2|\Gamma^{[0],\textrm{kin}}|,

where Γ[0],kin\Gamma^{[0],\textrm{kin}} is the set of vertices of kinetic type, and Γ[1]\Gamma^{[1]} the set of internal edges. Γ[0]\Gamma^{[0]}, the set of interaction vertices (for which we assume we have only one type) does not show up as they have weight zero, nor does Γ[0],mass\Gamma^{[0],\textrm{mass}}. By |||\cdots| we denote the cardinality of these sets.

Note that a graph Γ//γ\Gamma/\!/\gamma which has one two-point vertex labeled m2m^{2} (of mass type) which appears after contracting a self-energy subgraph γ\gamma has an improved power-counting as its edge number is 2h1(Γ//γ)+12h_{1}(\Gamma/\!/\gamma)+1. If the two-point vertex is labeled by \Box (kinetic type), it has not changed though: sdd(Γ//γ)=sdd(Γ)\textrm{sdd}(\Gamma/\!/\gamma)=\textrm{sdd}(\Gamma), as the weight of the two-point vertex compensates for the weight of the extra propagator. Quite often, in massless theories, one then omits the use of these two-point vertices altogether.

2.2. Hopf algebras of graphs

Let 𝒫{\mathcal{P}} be a class of graphs. We assume 𝒫\emptyset\in{\mathcal{P}} and that Γ𝒫\Gamma\in{\mathcal{P}} and ΓΓ\Gamma^{\prime}\cong\Gamma implies Γ𝒫\Gamma^{\prime}\in{\mathcal{P}}. We say 𝒫{\mathcal{P}} is closed under extension if given γΓ\gamma\subset\Gamma we have

(2.3) γ,Γ𝒫γ,Γ//γ𝒫.\gamma,\Gamma\in{\mathcal{P}}\Leftrightarrow\gamma,\Gamma/\!/\gamma\in{\mathcal{P}}.

Easy examples of such classes of graphs are 𝒫=core graphs{\mathcal{P}}=\text{core graphs}, and 𝒫=log divergent graphs{\mathcal{P}}=\text{log divergent graphs}, where Γ\Gamma is log divergent (in ϕ44\phi^{4}_{4} theory) if it is core and if further #E(Γi)=2|Γi|\#E(\Gamma_{i})=2|\Gamma_{i}| for every connected component ΓiΓ\Gamma_{i}\subset\Gamma. (Both examples are closed under extension by virtue of the identity |γ|+|Γ//γ|=|Γ||\gamma|+|\Gamma/\!/\gamma|=|\Gamma|.) Examples which arise in physical theories are more subtle. Verification of (2.3) requires an analysis of which graphs can arise from a given Lagrangian. To verify 𝒫={Γ|sdd(Γ)0}{\mathcal{P}}=\{\Gamma\ |\ \textrm{sdd}(\Gamma)\geq 0\} satisfies (2.3) one must consider self-energy graphs and the role of vertices of kinetic type as discussed above.

In particular, in massless ϕ44\phi_{4}^{4} theory divergent graphs are closed under extension, and so is the class of graphs for which 4|Γ|2|Γ[1]|+2|Γ[0],kin|+2|Γ[0],mass|04|\Gamma|-2|\Gamma^{[1]}|+2|\Gamma^{[0],\textrm{kin}}|+2|\Gamma^{[0],\textrm{mass}}|\geq 0. Note that this may contain superficially convergent graphs if there are sufficiently many two-point vertices of mass type. It pays to include them in the class of graphs to be considered, which enables one to discuss the effect of mass in the renormalization group flow.

Associated to a class 𝒫{\mathcal{P}} which is closed under extension as above, we define a (commutative, but not cocommutative) Hopf algebra H=H𝒫H=H_{\mathcal{P}} as follows. As a vector space, HH is freely spanned by isomorphism classes of graphs in 𝒫{\mathcal{P}}. (A number of variants are possible. One may work with oriented graphs, for example. In this case, the theory of graph homology yields a (graded commutative) differential graded Hopf algebra. One may also rigidify by working with disjoint unions of subgraphs of a given labeled graph.) HH becomes a commutative algebra with 1=[]1=[\emptyset] and product given by disjoint union. Define a comultiplication Δ:HHH\Delta:H\to H\otimes H:

(2.4) Δ(Γ)=γΓγ𝒫γΓ//γ.\Delta(\Gamma)=\sum_{\begin{subarray}{c}\gamma\subset\Gamma\\ \gamma\in{\mathcal{P}}\end{subarray}}\gamma\otimes\Gamma/\!/\gamma.

One checks that (2.3) implies that (2.4) is coassociative. Since HH is graded by loop numbers and each HnH_{n} is finite dimensional, the theory of Hopf algebras guarantees the existence of an antipode, so HH is a Hopf algebra.

If 𝒫𝒫{\mathcal{P}}^{\prime}\subset{\mathcal{P}} with Hopf algebras H,HH^{\prime},H ( e.g. take 𝒫{\mathcal{P}} to be core graphs, and 𝒫𝒫{\mathcal{P}}^{\prime}\subset{\mathcal{P}} divergent core graphs) then the map HHH\twoheadrightarrow H^{\prime} obtained by sending Γ0\Gamma\mapsto 0 if Γ𝒫\Gamma\not\in{\mathcal{P}}^{\prime} is a homomorphism of Hopf algebras. For example, the divergent Hopf algebra carries the information needed for renormalization [13], while the core Hopf algebra H𝒞H_{\mathcal{C}} determines the monodromy. In terms of groupschemes, one has Spec(Hlog. div.)Spec(H𝒞){\rm Spec\,}(H_{\text{log. div.}})\hookrightarrow{\rm Spec\,}(H_{{\mathcal{C}}}) is a closed subgroupscheme, and renormalization can be viewed as a morphism from the affine line with coordinate LL to Spec(Hlog. div.){\rm Spec\,}(H_{\text{log. div.}}). Already here we use that for divergent graphs with sdd(Γ)>0\textrm{sdd}(\Gamma)>0, we can evaluate them as polynomials in masses and external momenta with coefficients determined from log divergent graphs, see below.

Let Γi,i=1,2\Gamma_{i},i=1,2 be core graphs (a similar discussion will be valid for other classes of graphs) and let viΓiv_{i}\in\Gamma_{i} be vertices. Let Γ=Γ1Γ2\Gamma=\Gamma_{1}\cup\Gamma_{2} where the two graphs are joined by identifying v1v2v_{1}\sim v_{2}. Then Γ\Gamma is core (cf. proposition 3.2). Further, core subgraphs ΓΓ\Gamma^{\prime}\subset\Gamma all arise as the image of Γ1Γ2Γ\Gamma_{1}^{\prime}\amalg\Gamma_{2}^{\prime}\to\Gamma for ΓiΓi\Gamma_{i}^{\prime}\subset\Gamma_{i} core. Thus

(2.5) Δ(Γ)=ΓΓ//Γ=(Γ1Γ1//Γ1)(Γ2Γ2//Γ2)+(ΓΓ1Γ2)(Γ1//Γ1Γ2//Γ2)+Γ(Γ//ΓΓ1//Γ1Γ2//Γ2).\Delta(\Gamma)=\sum\Gamma^{\prime}\otimes\Gamma/\!/\Gamma^{\prime}=\Big{(}\sum\Gamma_{1}^{\prime}\otimes\Gamma_{1}/\!/\Gamma_{1}^{\prime}\Big{)}\Big{(}\sum\Gamma_{2}^{\prime}\otimes\Gamma_{2}/\!/\Gamma_{2}^{\prime}\Big{)}+\\ \sum(\Gamma^{\prime}-\Gamma_{1}^{\prime}\cdot\Gamma_{2}^{\prime})\otimes(\Gamma_{1}/\!/\Gamma_{1}^{\prime}\cdot\Gamma_{2}/\!/\Gamma_{2}^{\prime})+\sum\Gamma^{\prime}\otimes\Big{(}\Gamma/\!/\Gamma^{\prime}-\Gamma_{1}/\!/\Gamma_{1}^{\prime}\cdot\Gamma_{2}/\!/\Gamma_{2}^{\prime}\Big{)}.

It follows that the vector space IH𝒞I\subset H_{\mathcal{C}} spanned by elements ΓΓ1Γ2\Gamma-\Gamma_{1}\cdot\Gamma_{2} as above satisfies Δ(I)IH𝒞+H𝒞I\Delta(I)\subset I\otimes H_{\mathcal{C}}+H_{\mathcal{C}}\otimes I. Since II is an ideal, we see that H¯𝒞:=H𝒞/I\overline{H}_{\mathcal{C}}:=H_{\mathcal{C}}/I is a commutative Hopf algebra. Roughly speaking, H¯𝒞\overline{H}_{\mathcal{C}} is obtained from H𝒞H_{\mathcal{C}} by identifying one vertex reducible graphs with products of the component pieces.

Generalization to theories with more vertex and edge types are straightforward.

Refer to caption

Figure 4. In Eq.(2.6), we give the coproduct for this wheel with three spokes in the core Hopf algebra.

Fig.(4) gives the wheel with three spokes. This graph, which in ϕ4\phi^{4} theory (external edges to be added such that each vertex is four-valent) has a residue 6ζ(3)6\zeta(3) for conceptual reasons [2], has a coproduct (we omit edge labels and identify terms which are identical under this omission, which gives the indicated multiplicities)

(2.6) Δ([Uncaptioned image])=[Uncaptioned image]\displaystyle\Delta\left(\;\raisebox{-11.38109pt}{\epsfbox{wsix.eps}}\;\right)=\;\raisebox{-11.38109pt}{\epsfbox{wsix.eps}}\; \displaystyle\otimes 𝕀\displaystyle\mathbb{I}
+𝕀\displaystyle+\mathbb{I} \displaystyle\otimes
+4[Uncaptioned image]\displaystyle+4\;\raisebox{-11.38109pt}{\epsfbox{wthr.eps}}\; \displaystyle\otimes
+3[Uncaptioned image]\displaystyle+3\;\raisebox{-11.38109pt}{\epsfbox{wfour.eps}}\; \displaystyle\otimes
6[Uncaptioned image]\displaystyle 6\;\raisebox{-11.38109pt}{\epsfbox{wfive.eps}}\; \displaystyle\otimes [Uncaptioned image].\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{wtad.eps}}\;.

For example, the three possible labelings for the four-edge cycle in the third line are 45234523, 56315631 and 64126412. While the graph has a non-trivial coproduct in the core Hopf algebra, it is a primitive element in the renormalization Hopf algebra. It is tempting to hope that the core coproduct relates to the Hodge structure underlying the period which appears in the residue of this graph.

2.3. Rooted tree Hopf algebras [12], [3]

We introduce the Hopf algebra of decorated non-planar rooted trees H𝒯H_{\mathcal{T}} using non-empty finite sets as decorations (decorations will be sets of edge labels of Feynman graphs below) to label the vertices of the rooted tree Hopf algebra H𝒯()H_{\mathcal{T}}(\emptyset). Products in H𝒯H_{\mathcal{T}} are disjoint unions of trees (forests). We write the coproduct as

(2.7) Δ(T)=T𝕀+𝕀T+admissible cuts CPC(T)RC(T).\Delta(T)=T\otimes\mathbb{I}+\mathbb{I}\otimes T+\sum_{\textrm{admissible cuts $C$}}P^{C}(T)\otimes R^{C}(T).

Edges are oriented away from the root and a vertex which has no outgoing edge we call a foot. An admissible cut is a subset of edges of a tree such that no path from the root to any vertex of TT traverses more than one element of that subset. Such a cut CC separates TT into at least 22 components. The component containing the root is denoted RC(T)R^{C}(T), and the product of the other components is PC(T)P^{C}(T).

A ladder is a tree without side branching. Decorated ladders generate a sub-Hopf algebra L𝒯H𝒯L_{{\mathcal{T}}}\subset H_{{\mathcal{T}}}. A general element in L𝒯L_{{\mathcal{T}}} is a sum of bamboo forests, that is disjoint unions of ladders. Decorated ladders have an associative shuffle product

(2.8) L1L2:=kshuffle(1,2)L(k)L_{1}\star L_{2}:=\sum_{k\in\text{shuffle}(\ell_{1},\ell_{2})}L(k)

where i\ell_{i} denotes the ordered set of decorations for LiL_{i} and shuffle(1,2)\text{shuffle}(\ell_{1},\ell_{2}) is the set of all ordered sets obtained by shuffling together 1\ell_{1} and 2\ell_{2}.

Lemma 2.1.

Let 𝒦L𝒯{\mathcal{K}}\subset L_{\mathcal{T}} be the ideal generated by elements of the form L1L2L1L2L_{1}\cdot L_{2}-L_{1}\star L_{2}. Then Δ(𝒦)𝒦L𝒯+L𝒯𝒦\Delta({\mathcal{K}})\subset{\mathcal{K}}\otimes L_{\mathcal{T}}+L_{\mathcal{T}}\otimes{\mathcal{K}}.

Proof.

Write Δ(Li)=j=0diLijLidij\Delta(L_{i})=\sum_{j=0}^{d_{i}}L_{ij}\otimes L_{i}^{d_{i}-j} where did_{i} is the length of LiL_{i} and LijL_{ij} (resp. LijL_{i}^{j}) is the bottom (resp. top) subladder of length jj. Then

(2.9) Δ(L1)Δ(L2)=j,μL1jL2μL1d1jL2d2μ\displaystyle\Delta(L_{1})\Delta(L_{2})=\sum_{j,\mu}L_{1j}L_{2\mu}\otimes L_{1}^{d_{1}-j}L_{2}^{d_{2}-\mu}
Δ(L1L2)=kΔ(L(k))=k,νL(k)νL(k)d1+d2ν.\displaystyle\Delta(L_{1}\star L_{2})=\sum_{k}\Delta(L(k))=\sum_{k,\nu}L(k)_{\nu}\otimes L(k)^{d_{1}+d_{2}-\nu}.

Consider pairs (j,μ)(j,\mu) of indices in (2.9) and write j+μ=νj+\mu=\nu. Among the pairs k,νk,\nu we consider the subset K(j,μ)K(j,\mu) for which the first ν=j+μ\nu=j+\mu elements of the ordered set consist of a shuffle of the decorations on the ladders L1j,L2μL_{1j},L_{2\mu}. It is clear that the remaining d1+d2νd_{1}+d_{2}-\nu elements of kk will then run through shuffles of the decorations of L1d1j,L2d2μL_{1}^{d_{1}-j},L_{2}^{d_{2}-\mu}, so

(2.10) Δ(L1)Δ(L2)Δ(L1L2)=j,μ((L1jL2μkK(j,μ)L(k)j+μ)L1d1jL2d2μ)+j,μ(kK(j,μ)L(k)j+μ(L1d1jL2d2μL(k)d1+d2jμ))𝒦L𝒯+L𝒯𝒦.\Delta(L_{1})\Delta(L_{2})-\Delta(L_{1}\star L_{2})=\sum_{j,\mu}\Big{(}(L_{1j}L_{2\mu}-\sum_{k\in K(j,\mu)}L(k)_{j+\mu})\otimes L_{1}^{d_{1}-j}L_{2}^{d_{2}-\mu}\Big{)}+\\ \sum_{j,\mu}\Big{(}\sum_{k\in K(j,\mu)}L(k)_{j+\mu}\otimes(L_{1}^{d_{1}-j}L_{2}^{d_{2}-\mu}-L(k)^{d_{1}+d_{2}-j-\mu})\Big{)}\in{\mathcal{K}}\otimes L_{\mathcal{T}}+L_{\mathcal{T}}\otimes{\mathcal{K}}.

Remark 2.2.

Any bamboo forest is equivalent mod 𝒦{\mathcal{K}} to a sum of stalks. Indeed, one has e.g.

(2.11) L1L2L3(L1L2)L3(L1L2)L3L1L2L3.L_{1}\cdot L_{2}\cdot L_{3}\equiv(L_{1}\star L_{2})\cdot L_{3}\equiv(L_{1}\star L_{2})\star L_{3}\equiv L_{1}\star L_{2}\star L_{3}.

For any decoration \ell, one has an operator [3]

(2.12) B+:H𝒯H𝒯B_{+}^{\ell}:H_{\mathcal{T}}\to H_{\mathcal{T}}

which carries any forest to the tree obtained by connecting a single root vertex with decoration \ell to all the roots of the forest. This operator is a Hochschild 1-cocycle, i.e.

(2.13) ΔB+=B+𝕀+(idB+)Δ.\Delta B_{+}^{\ell}=B_{+}^{\ell}\otimes\mathbb{I}+(\text{id}\otimes B_{+}^{\ell})\Delta.

Let 𝒥H𝒯{\mathcal{J}}\subset H_{\mathcal{T}} be the smallest ideal containing the ideal 𝒦{\mathcal{K}} as in lemma 2.1 and stable under all the operators B+B_{+}^{\ell}. Generators of 𝒥{\mathcal{J}} as an abelian group are obtained by starting with elements of 𝒦{\mathcal{K}} and successively applying B+B_{+}^{\ell} for various \ell and multiplying by elements of H𝒯H_{\mathcal{T}}. It follows from (2.13) that Δ𝒥𝒥H𝒯+H𝒯𝒥\Delta{\mathcal{J}}\subset{\mathcal{J}}\otimes H_{\mathcal{T}}+H_{\mathcal{T}}\otimes{\mathcal{J}}. Define

(2.14) H¯𝒯:=H𝒯/𝒥.\overline{H}_{\mathcal{T}}:=H_{\mathcal{T}}/{\mathcal{J}}.

A flag in a core graph Γ\Gamma is a chain

(2.15) f:=Γ1Γn=Γf:=\emptyset\subsetneq\Gamma_{1}\subsetneq\cdots\subsetneq\Gamma_{n}=\Gamma

of core subgraphs. Write F(Γ)F(\Gamma) for the collection of all maximal flags of Γ\Gamma. One checks easily that for a maximal flag, n=|Γ|n=|\Gamma|. Let us consider an example.

(2.16) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{abef.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.17) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{cdef.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.18) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{abef.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.19) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{abef.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.20) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{abcde.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.21) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{abcdf.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.22) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{cdef.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.23) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{cdef.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.24) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{abcde.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.25) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{abcdf.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.26) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{abcde.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,
(2.27) \displaystyle\subsetneq [Uncaptioned image][Uncaptioned image],\displaystyle\;\raisebox{-11.38109pt}{\epsfbox{abcdf.eps}}\;\subsetneq\;\raisebox{-11.38109pt}{\epsfbox{abcdef.eps}}\;,

are the twelve flags for the graph given in Fig.(5). We omitted the edge labels in the above flags. Note that only the first two , (2.16,2.17) are relevant for the renormalization Hopf algebra to be introduced below.

Refer to caption

Figure 5. A graph with overlapping subdivergences. The renormalization Hopf algebra gives Δ123456=561234+125634+345612\Delta^{\prime}{123456}=56\otimes 1234+1256\otimes 34+3456\otimes 12. Note that each edge belongs to some subgraph with sdd0\textrm{sdd}\geq 0.

To the flag ff we associate the ladder L(f)L(f) with nn vertices decorated by ΓiΓi1\Gamma_{i}-\Gamma_{i-1}. (More precisely, the foot is decorated by Γ1\Gamma_{1} and the root by ΓΓn1\Gamma-\Gamma_{n-1}.). Define

(2.28) ρL:H𝒞L𝒯;ρL(Γ):=fF(Γ)L(f)\rho_{L}:H_{\mathcal{C}}\to L_{\mathcal{T}};\quad\rho_{L}(\Gamma):=\sum_{f\in F(\Gamma)}L(f)

Here the set of labels DD will be the set of subsets of graph labels.

Lemma 2.3.

The map ρL\rho_{L} is a homomorphism of Hopf algebras.

Proof.

For a flag ff let f(p)f^{(p)} be the bottom pp vertices with the given labeling, and let f(p)f_{(p)} be the top npn-p vertices with the quotient labeling gotten by contracting the core graph associated to the bottom pp vertices. For γΓ\gamma\subset\Gamma a core subgraph, define F(Γ,γ):={fF(Γ)|γf}F(\Gamma,\gamma):=\{f\in F(\Gamma)\ |\ \gamma\in f\}. There is a natural identification

(2.29) F(Γ,γ)=F(γ)×F(Γ//γ).F(\Gamma,\gamma)=F(\gamma)\times F(\Gamma/\!/\gamma).

We have

(2.30) (ρLρL)Δ𝒞(Γ)=γρL(γ)ρL(Γ//γ)=γfF(Γ,γ)L(f|γ|)L(f|γ|).(\rho_{L}\otimes\rho_{L})\circ\Delta_{{\mathcal{C}}}(\Gamma)=\sum_{\gamma}\rho_{L}(\gamma)\otimes\rho_{L}(\Gamma/\!/\gamma)=\sum_{\gamma}\sum_{f\in F(\Gamma,\gamma)}L(f^{|\gamma|})\otimes L(f_{|\gamma|}).

On the other hand

(2.31) ΔLρL(Γ)=fF(Γ)i=1nL(f(i))L(f(i)).\Delta_{L}\circ\rho_{L}(\Gamma)=\sum_{f\in F(\Gamma)}\sum_{i=1}^{n}L(f^{(i)})\otimes L(f_{(i)}).

The assertion of the lemma is that there is a 111-1 correspondence

(2.32) {γ,max. flag of Γ containing γ}{max. flag of Γ ,in}.\{\gamma,\text{max. flag of $\Gamma$ containing }\gamma\}\leftrightarrow\{\text{max. flag of $\Gamma$ },i\leq n\}.

This is clear. ∎

In fact, the tree structure associated to a maximal flag ff of Γ\Gamma is rather more intricate than just a ladder. Though we do not use this tree structure in the sequel, we present the construction in some detail to help in understanding the difference between the core and renormalization Hopf algebra.

We want to associate a forest T(f)T(f) to the flag ff, and we proceed by induction on n=|Γ|n=|\Gamma|. We can write Γ=Γ(j)\Gamma=\bigcup\Gamma^{(j)} in such a way that all the Γ(j)\Gamma^{(j)} are core and one vertex irreducible, and such that |Γ|=|Γ(j)||\Gamma|=\sum|\Gamma^{(j)}|. This decomposition is unique. If it is nontrivial, we define T(f)=T(f(j))T(f)=\prod T(f^{(j)}) where f(j)f^{(j)} is the induced flag from ff on Γ(j)\Gamma^{(j)}. We now may assume Γ\Gamma is one vertex irreducible. If the Γi\Gamma_{i} in our flag are all one vertex irreducible, we take T(f)=L(f)T(f)=L(f) to be a ladder as above. Otherwise, let m<nm<n be maximal such that ΓmΓ\Gamma_{m}\subsetneq\Gamma is one vertex reducible. By induction, we have a forest T(f|Γm)T(f|\Gamma_{m}). To define T(f)T(f), we glue the foot of the ladder with decorations Γm+1Γm,,ΓΓn1\Gamma_{m+1}-\Gamma_{m},\dotsc,\Gamma-\Gamma_{n-1} to all the roots of T(f|Γm)T(f|\Gamma_{m}). (For an example, see figs.(6) and (7).)

Refer to caption

Figure 6. The core Hopf algebra on rooted trees. We indicate subgraphs by edge labels on the vertices of rooted trees. The dots indicate seven more such trees, corresponding to flags ΓiΓjΓk\Gamma_{i}\subsetneq\Gamma_{j}\subsetneq\Gamma_{k} with Γi\Gamma_{i} a cycle on four edges. The last tree represents a sum of two flags, 343456123456+56345612345634\subsetneq 3456\subsetneq 123456+56\subsetneq 3456\subsetneq 123456, again indicating graphs by edge labels. Hence that tree corresponds to a sum of two ladders, as it should.
Lemma 2.4.

Let Γ=Γ(j)\Gamma=\bigcup\Gamma^{(j)} where Γ\Gamma and the Γ(j)\Gamma^{(j)} are core. Assume |Γ|=j|Γ(j)||\Gamma|=\sum_{j}|\Gamma^{(j)}|. Then, viewing flags fF(Γ)f\in F(\Gamma) as sets of core subgraphs, ordered by inclusion, there is a 111-1 correspondence between F(Γ)F(\Gamma) and shuffles of the F(Γ(j))F(\Gamma^{(j)}).

Proof.

One checks easily that the Γ(j)\Gamma^{(j)} can have no edges in common. Further, there is a 111-1 correspondence between core subgraphs ΓΓ\Gamma^{\prime}\subset\Gamma and collections of core subgraphs Γ(j)Γ(j)\Gamma^{(j)}{}^{\prime}\subset\Gamma^{(j)}. Here, the dictionary is given by Γ{ΓΓ(j)}\Gamma^{\prime}\mapsto\{\Gamma^{\prime}\cap\Gamma^{(j)}\} and {Γ(j)}Γ(j)\{\Gamma^{(j)}{}^{\prime}\}\mapsto\bigcup\Gamma^{(j)}{}^{\prime}. The assertion of the lemma follows. ∎

Refer to caption

Figure 7. The two graphs differ in how the subdivergences are inserted. a,c3,4a,c\in{3,4}, and b,d5,6b,d\in{5,6}, ca,bdc\not=a,b\not=d. So there are eight such legal trees, plus the two which are identical between the two graphs. Note the permutation of labels at the feet of the trees in ρ(Γ)\rho(\Gamma): 1a2b12ab1a2b\leftrightarrow 12ab. Keeping that order, we can uniquely reconstruct each graph from the knowledge of the labels at the feet: 1a2b,34,561a2b,34,56 and 12ab,34,5612ab,34,56, which are the cycles in each graph. Note that in the difference of the two graphs, only the difference of those eight trees remains, corresponding to a primitive element in the renormalization Hopf algebra. The core Hopf algebra hence stores much more information than the renormalization Hopf algebra, which we hope to use in the future to understand the periods assigned to Feynman graphs by the Feynman rules.

As a consequence of lemma 2.4 we may partition the flags F(Γ)F(\Gamma) associated to a core Γ\Gamma as follows. Given fF(Γ)f\in F(\Gamma), Let ΓmΓ\Gamma_{m}\subset\Gamma be maximal in the flag ff such that Γm\Gamma_{m} is 11-vertex reducible. The flag ff induces a flag fmf_{m} on Γm\Gamma_{m}, and we know that it is a shuffle of flags fm(j)f_{m}^{(j)} on Γm(j)\Gamma_{m}^{(j)} where Γm=Γm(j)\Gamma_{m}=\bigcup\Gamma_{m}^{(j)} as in the lemma. We say two flags are equivalent, fff\sim f^{\prime}, if ff and ff^{\prime} agree at Γm\Gamma_{m} and above, and if they simply correspond to two different shuffles of the flags fm(j)f_{m}^{(j)}. We now have

(2.33) T(f)ffL(f)mod𝒥.T(f)\equiv\sum_{f^{\prime}\sim f}L(f^{\prime})\mod{\mathcal{J}}.

Indeed, T(f)T(f) is obtained by successive B+B_{+}^{\ell} operations applied to the forest T(f|Γm)T(f|\Gamma_{m}). The latter, by remark 2.2, coincides with the righthand side of (2.33). We conclude

Proposition 2.5.

With notation as above, there exist homomorphisms of Hopf algebras

(2.34) H𝒞ρLL𝒯H¯𝒞ρTH¯𝒯\begin{CD}H_{\mathcal{C}}@>{\rho_{L}}>{}>L_{\mathcal{T}}\\ @V{}V{}V@V{}V{}V\\ \overline{H}_{\mathcal{C}}@>{\rho_{T}}>{}>\overline{H}_{\mathcal{T}}\end{CD}

Here ρT(Γ)\rho_{T}(\Gamma) is the sum T(f)T(f) over equivalence classes of flags ff as above. We will barely use H𝒯H_{\mathcal{T}} in the following, and introduced it for completeness and the benefit of the reader used to it.

2.4. Renormalization Hopf algebras

In a similar manner, one may define homomorphisms

(2.35) ρ:HH¯𝒯\rho_{\mathcal{R}}:H_{\mathcal{R}}\to\overline{H}_{\mathcal{T}}

for any one of the renormalization Hopf algebras obtained by imposing restrictions on external leg structure. For a graph Γ\Gamma, let, as before, the residue of Γ\Gamma, res(Γ)\textrm{res}(\Gamma), be the graph with no loops obtained by shrinking all its internal edges to a point. What remains are the external half edges connected to that point (cf. section 2.1). Note that ”doubling” an edge by putting a two-point vertex in it does not change the residue.

In ϕ44\phi_{4}^{4} theory for example, graphs have 2m2m external legs, with m0m\geq 0. For a renormalizable theory, there is a finite set of external leg structures \mathcal{R} such that we obtain a renormalization Hopf algebra for that set.

For example, for massive ϕ44\phi_{4}^{4} theory, there are three such structures: the four-point vertex, and two two-point vertices, of kinetic type and mass type.

Let us now consider flags associated to core graphs. Such chains ΓiΓi+1Γ\cdots\Gamma_{i}\subsetneq\Gamma_{i+1}\subsetneq\cdots\subsetneq\Gamma correspond to decorated ladders, and the coproduct on the level of such ladders is a sum over all possibilities to cut an edge in such a ladder, splitting the chain

(2.36) [Γi][Γi+1//ΓiΓ//Γi].[\cdots\subsetneq\Gamma_{i}]\otimes[\Gamma_{i+1}/\!/\Gamma_{i}\subsetneq\cdots\subsetneq\Gamma/\!/\Gamma_{i}].

So let us call such an admissible cut renormalization-admissible, if all core graphs Γi\Gamma_{i}, Γ//Γi\Gamma/\!/\Gamma_{i} obtained by the cut have residues in {\mathcal{R}}.

The set of renormalization-admissible cuts is a subset of the admissible cuts of a core graph, and the coproduct respects this. Hence the renormalization Hopf algebra HH_{\mathcal{R}} is a quotient Hopf algebra of the core Hopf algebra.

If we enlarge the set {\mathcal{R}} to include other local field operators appearing for example in an operator product expansion we get quotient Hopf algebras between the core and the renormalization Hopf algebra.

2.5. External leg structures

External edges are usually labeled by data which characterize the amplitude under consideration. Let σ\sigma be such data. For graphs Γ\Gamma with a given residue res(Γ){\textrm{res}}(\Gamma), there is a finite set τ{σ}res(Γ)\tau\in\{\sigma\}_{{\textrm{res}}(\Gamma)} of possible data τ\tau. A choice of such data determines a labeling of the corresponding vertex to which a subgraph shrinks. Let Γ//γτ\Gamma/\!/\gamma_{\tau} be that co-graph with the corresponding vertex labeling.

One gets a Hopf algebra structure on pairs (Γ,σ)(\Gamma,\sigma) by using the renormalization coproduct Δ(Γ)=ΓΓ′′\Delta(\Gamma)=\Gamma^{\prime}\otimes\Gamma^{\prime\prime} by setting Δ(Γ,σ)=τ{σ}res(Γ)(Γ,τ)(Γτ′′,σ)\Delta(\Gamma,\sigma)=\sum_{\tau\in\{\sigma\}_{\textrm{res}(\Gamma^{\prime})}}(\Gamma^{\prime},\tau)\otimes(\Gamma^{\prime\prime}_{\tau},\sigma). We regard the decomposition into external leg structures as a partition of unity and write

(2.37) τ{σ}res(Γ)(Γ,τ)=(Γ,𝕀).\sum_{\tau\in\{\sigma\}_{\textrm{res}(\Gamma)}}(\Gamma,\tau)=(\Gamma,\mathbb{I}).

In our applications we only need this for (sub)graphs γ\gamma with |res(γ)|=2|\textrm{res}(\gamma)|=2, and the use of these notions will become clear in the applications below.

3. Combinatorics of blow-ups

We consider n1{\mathbb{P}}^{n-1} with fixed homogeneous coordinates 𝒜:={A1,,An}{\mathcal{A}}:=\{A_{1},\dotsc,A_{n}\}. Suppose given a subset S2𝒜S\subset 2^{{\mathcal{A}}}. Assume 𝒜S{\mathcal{A}}\not\in S and that SS has the property that whenever μ1,μ2S\mu_{1},\mu_{2}\in S with μ1μ2𝒜\mu_{1}\cup\mu_{2}\neq{\mathcal{A}}, then μ1μ2S\mu_{1}\cup\mu_{2}\in S. For μS\mu\in S we write Lμn1L_{\mu}\subset{\mathbb{P}}^{n-1} for the coordinate linear space defined by Ai=0,iμA_{i}=0,\ i\in\mu. Write L(S):={Lμ|μS}L(S):=\{L_{\mu}\ |\ \mu\in S\}. We see that

(3.1) LμiL(S);Lμ1Lμ2Lμ1Lμ2L(S).L_{\mu_{i}}\in L(S);\ L_{\mu_{1}}\cap L_{\mu_{2}}\neq\emptyset\Rightarrow L_{\mu_{1}}\cap L_{\mu_{2}}\in L(S).

We can stratify the set L(S)L(S) taking L(S)1L(S)_{1} to be the set of all minimal elements (under inclusion) of L(S)L(S). More generally, L(S)iL(S)_{i} will be the set of minimal elements in L(S)j=1i1L(S)jL(S)-\coprod_{j=1}^{i-1}L(S)_{j}.

Proposition 3.1.

(i) Elements in L(S)1L(S)_{1} are all disjoint, so we may define P(S)1P(S)_{1} to be the variety defined by blowing up elements in L(S)1L(S)_{1} on n1{\mathbb{P}}^{n-1}. We do not need to specify an order in which to perform the blowups.
(ii) More generally, the strict transforms of elements in L(S)i+1L(S)_{i+1} to the space P(S)iP(S)_{i} obtained by successively blowing the strict transform of L(S)j,j=1,,iL(S)_{j},\ j=1,\dotsc,i are disjoint, so we may inductively define P(S)P(S) to be the successive blowup of the L(S)iL(S)_{i}.
(iii) Let EiP(S)E_{i}\subset P(S) correspond to the blowup of Lμi,i=1,,rL_{\mu_{i}},\ i=1,\dotsc,r. (EiE_{i} is the unique exceptional divisor with image LμiL_{\mu_{i}} in P(S)P(S).)Then E1ErE_{1}\cap\cdots\cap E_{r}\neq\emptyset if and only if after possibly reordering, we have inclusions Lμ1LμrL_{\mu_{1}}\subset\cdots\subset L_{\mu_{r}}.
(iv) The total exceptional divisor EP(S)E\subset P(S) is a normal crossings divisor.
(v) Let Mn1M\subset{\mathbb{P}}^{n-1} be a coordinate linear space. Assume MLM\not\subset L for any LL(S)L\in L(S). Then ML(S):={ML|LL(S)}M\cap L(S):=\{M\cap L\ |\ L\in L(S)\} satisfies (3.1). The strict transform of MM in P(S)P(S) is obtained by blowing up elements of ML(S)M\cap L(S) on MM as in (i) and (ii) above.

Proof.

If L1L2L(S)iL_{1}\neq L_{2}\in L(S)_{i} and L1L2L_{1}\cap L_{2}\neq\emptyset, then L1L2L(S)jL_{1}\cap L_{2}\in L(S)_{j} for some j<ij<i. This means that when we get to the ii-th step, L1L2L_{1}\cap L_{2} has already been blown up, so the strict transforms of the LiL_{i} are disjoint, proving (ii). For (iii)(iii), EiLμ1Lμr\bigcap E_{i}\neq\emptyset\Leftarrow L_{\mu_{1}}\subset\cdots\subset L_{\mu_{r}} follows from the above argument. Conversely, if we have strict inclusions among the LμiL_{\mu_{i}}, we may write (abusively) Lμi/Lμi1L_{\mu_{i}}/L_{\mu_{i-1}} for the projective space with homogeneous coordinates the homogeneous coordinates on LμiL_{\mu_{i}} vanishing on Lμi1L_{\mu_{i-1}}. The exceptional divisor on the blowup of Lμi1LμiL_{\mu_{i-1}}\subset L_{\mu_{i}} is identified with Lμi1×(Lμi/Lμi1)L_{\mu_{i-1}}\times(L_{\mu_{i}}/L_{\mu_{i-1}}). A straightforward calculation identifies nonempty open sets (open toric orbits in the sense to be discussed below) in Ei\bigcap E_{i} and

(3.2) Lμ1×(Lμ2/Lμ1)××(Lμr/Lμr1)L_{\mu_{1}}\times(L_{\mu_{2}}/L_{\mu_{1}})\times\cdots\times(L_{\mu_{r}}/L_{\mu_{r-1}})

The remaining parts of the proposition follow from the algorithm in [8]. ∎

For us, sets SS as above will arise in the context of graphs. Recall in 2.1 we defined the notion of core graph.

Proposition 3.2.

Let Γ\Gamma be a graph, and let Γ1,Γ2Γ\Gamma_{1},\Gamma_{2}\subset\Gamma be core subgraphs. Then the union Γ1Γ2\Gamma_{1}\cup\Gamma_{2} is a core subgraph.

Proof.

Removing an edge increases the Euler-Poincaré characteristic by 11. If h1h_{1} doesn’t drop, then either h0h_{0} increases (the graph disconnects when ee is removed) or ee has a unary vertex so removing ee drops the number of vertices. Suppose ee is an edge of Γ1\Gamma_{1} (assumed core). Then ee cannot have a unary vertex. If, on the other hand, removing ee disconnects Γ1Γ2\Gamma_{1}\cup\Gamma_{2}, then since the Γi\Gamma_{i} are core what must happen is that each Γi\Gamma_{i} has precisely one vertex of ee. But this would imply that Γ1\Gamma_{1} is not core, a contradiction. ∎

To a graph Γ\Gamma we may associate the projective space (Γ){\mathbb{P}}(\Gamma) with homogeneous coordinates Ae,eE(Γ)A_{e},\ e\in E(\Gamma) labeled by the edges of Γ\Gamma. Let Γ\Gamma be a core graph. A coordinate linear space L(Γ)L\subset{\mathbb{P}}(\Gamma) is a non-empty linear space defined by some subset of the homogeneous coordinate functions, L:Ae1==Aep=0L:A_{e_{1}}=\cdots=A_{e_{p}}=0. Define L(Γ)L(\Gamma) to be the set of coordinate linear spaces in (Γ){\mathbb{P}}(\Gamma) such that the corresponding set of edges ei1,,eipe_{i_{1}},\dotsc,e_{i_{p}} is the edge set of a core subgraph ΓΓ\Gamma^{\prime}\subset\Gamma. It follows from proposition 3.2 that L(Γ)L(\Gamma) satisfies condition (3.1), so the iterated blowup

(3.3) π:P(Γ)(Γ)\pi:P(\Gamma)\to{\mathbb{P}}(\Gamma)

as in proposition 3.1 is defined. Define

(3.4) =LL(Γ)L(Γ);E=EL=π1.{\mathcal{L}}=\bigcup_{L\in L(\Gamma)}L\subset{\mathbb{P}}(\Gamma);\quad E=\bigcup E_{L}=\pi^{-1}{\mathcal{L}}.
Lemma 3.3.

Suppose (Γ)=n1{\mathbb{P}}(\Gamma)={\mathbb{P}}^{n-1} with coordinates A1,,AnA_{1},\dotsc,A_{n}. Let L(Γ)L\subset{\mathbb{P}}(\Gamma) be defined by A1==Ap=0A_{1}=\cdots=A_{p}=0. Let πL:PL(Γ)\pi_{L}:P_{L}\to{\mathbb{P}}(\Gamma) be the blowup of LL. Then the exceptional divisor EPLE\subset P_{L} is identified with p1×L{\mathbb{P}}^{p-1}\times L. Further A1,,ApA_{1},\dotsc,A_{p} induce coordinates on the vertical fibres p1{\mathbb{P}}^{p-1} and Ap+1,,AnA_{p+1},\dotsc,A_{n} give homogeneous coordinates on LL.

Proof.

This is standard. One way to see it is to use the map n1Lp1,[a1,,an][a1,,ap]{\mathbb{P}}^{n-1}-L\to{\mathbb{P}}^{p-1},\ [a_{1},\dotsc,a_{n}]\mapsto[a_{1},\dotsc,a_{p}]. (Here, and in the sequel, [][\cdots] denotes a point in homogeneous coordinates.) This extends to a map ff on PLP_{L}:

(3.5) EPLfp1πL|EπLLn1.\begin{CD}E@>{\hookrightarrow}>{}>P_{L}@>{f}>{}>{\mathbb{P}}^{p-1}\\ @V{}V{\pi_{L}|_{E}}V@V{}V{\pi_{L}}V\\ L@>{\hookrightarrow}>{}>{\mathbb{P}}^{n-1}.\end{CD}

The resulting map πL|E×f:EL×p1\pi_{L}|_{E}\times f:E\cong L\times{\mathbb{P}}^{p-1}. ∎

It will be helpful to better understand the geometry of P(Γ)P(\Gamma). Let 𝔾m=Spec[t,t1]{\mathbb{G}}_{m}={\rm Spec\,}{\mathbb{Q}}[t,t^{-1}] be the standard one dimensional algebraic torus. Define T=𝔾mn/𝔾mT={\mathbb{G}}_{m}^{n}/{\mathbb{G}}_{m} where the quotient is taken with respect to the diagonal embedding. For all practical purposes, it suffices to consider complex points

(3.6) T()=×n/××n1.T({\mathbb{C}})={\mathbb{C}}^{\times n}/{\mathbb{C}}^{\times}\cong{\mathbb{C}}^{\times n-1}.

A toric variety PP is an equivariant (partial) compactification of TT. In other words, TPT\subset P is an open set, and we have an extension of the natural group map mm

(3.7) T×TT×Pmm¯TP.\begin{CD}T\times T@>{\subset}>{}>T\times P\\ @V{m}V{}V@V{\bar{m}}V{}V\\ T@>{\subset}>{}>P.\end{CD}

For example, (Γ){\mathbb{P}}(\Gamma) is a toric variety for a torus T(Γ)T(\Gamma). Canonically, we may write T(Γ)=(eEdge(Γ)𝔾m)/𝔾m.T(\Gamma)=(\prod_{e\in\text{Edge}(\Gamma)}{\mathbb{G}}_{m})\big{/}{\mathbb{G}}_{m}. More important for us:

Proposition 3.4.

(i) P(Γ)P(\Gamma) is a toric variety for T=T(Γ)T=T(\Gamma).
(ii) The orbits of TT on P(Γ)P(\Gamma) are in 111-1 correspondence with pairs (F,ΓpΓ1Γ)(F,\ \Gamma_{p}\subsetneq\cdots\subsetneq\Gamma_{1}\subsetneq\Gamma). Here FΓF\subset\Gamma is a (possibly empty) subforest (subgraph with h1(F)=0h_{1}(F)=0) and the Γi\Gamma_{i} are core subgraphs of Γ\Gamma. We require that the image of Fi:=FΓiF_{i}:=F\cap\Gamma_{i} in Γi//Γi+1\Gamma_{i}/\!/\Gamma_{i+1} be a subforest for each ii. (cf. (3.2)). The orbit associated to such a pair is canonically identified with the open orbit in the toric variety (Γp//Fp)×((Γp1//Γp)//Fp1)××((Γ//Γ1)//F){\mathbb{P}}(\Gamma_{p}/\!/F_{p})\times{\mathbb{P}}((\Gamma_{p-1}/\!/\Gamma_{p})/\!/F_{p-1})\times\cdots\times{\mathbb{P}}((\Gamma/\!/\Gamma_{1})/\!/F).

Proof.

A general reference for toric varieties is [9]. The fact (i) that P(Γ)P(\Gamma) is a toric variety follows inductively from the fact that the blowup of an invariant ideal II in a toric variety is toric. Indeed, the torus acts on II and hence on the blowup Proj(I)\text{Proj}(I).

We recall some toric constructions. Let N=Edge(Γ)/N={\mathbb{Z}}^{\text{Edge}(\Gamma)}/{\mathbb{Z}}, and let M=hom(N,)M=\hom(N,{\mathbb{Z}}). We have canonically T=Spec[M]T={\rm Spec\,}{\mathbb{Q}}[M] where [M]{\mathbb{Q}}[M] is the group ring of the lattice MM. A fan (op. cit., 1.4, p. 20) {\mathcal{F}} is a finite set of convex cones in N=NN_{\mathbb{R}}=N\otimes{\mathbb{R}} satisfying certain simple axioms. To a cone CNC\subset N_{\mathbb{R}} one associates the dual cone (op. cit. p. 4)

(3.8) C={mM|m,c0,cC}C^{\vee}=\{m\in M_{\mathbb{R}}\ |\ \langle m,c\rangle\geq 0,\forall c\in C\}

(resp. the semigroup C=CMC^{\vee}_{\mathbb{Z}}=C^{\vee}\cap M). The toric variety V()V({\mathcal{F}}) associated to the fan {\mathcal{F}} is then a union of the affine sets U(C):=Spec[C]U(C):={\rm Spec\,}{\mathbb{Q}}[C^{\vee}_{\mathbb{Z}}]. For example, our NN has rank n1n-1. There are nn evident elements ee determined by the nn edges of Γ\Gamma. Let Ce={eeree|re0}C_{e}=\{\sum_{e^{\prime}\neq e}r_{e^{\prime}}e^{\prime}\ |\ r_{e^{\prime}}\geq 0\} be the cone spanned by all edges except ee. The spanning edges for CeC_{e} form a basis for NN which implies that U(Ce)𝔸n1U(C_{e})\cong{\mathbb{A}}^{n-1}. Since all the coordinate rings lie in [M]{\mathbb{Q}}[M] (i.e. T(Γ)U(Ce)T(\Gamma)\subset U(C_{e})), one is able to glue together the U(Ce)U(C_{e}). The resulting toric variety associated to the fan {Ce|eEdge(E)}\{C_{e}\ |\ e\in\text{Edge}(E)\} is canonically identified with (Γ){\mathbb{P}}(\Gamma).

Remark 3.5.

Our toric varieties will all be smooth (closures of orbits in smooth toric varieties are smooth), which is equivalent ([9], §2) to the condition that cones in the fan are all generated by subsets of bases for the lattice NN. Faces of these cones are in 111-1 correspondence with subsets of the generating set.

In general, the orbits of the torus action are in 111-1 correspondence with the cones CC in the fan (op. cit. 3.1, p.51). The subgroup of NN generated by CNC\cap N corresponds to the subgroup of TT which acts trivially on the orbit. For example, in the case of projective space n1{\mathbb{P}}^{n-1}, there are nn cones CeC_{e} of dimension n1n-1 corresponding to the nn fixed points (0,,1,,0)n1(0,\dotsc,1,\dotsc,0)\in{\mathbb{P}}^{n-1}. For any SEdge(Γ)S\subsetneq\text{Edge}(\Gamma), the cone C(S)C(S) spanned by the edges of SS corresponds to the orbit {(,xe,)|xe=0eS}n1\{(\dotsc,x_{e},\ldots)\ |\ x_{e}=0\Leftrightarrow e\in S\}\subset{\mathbb{P}}^{n-1}. Let L:Ae=0,eΓΓL:A_{e}=0,e\in\Gamma^{\prime}\subset\Gamma be a coordinate linear space in (Γ){\mathbb{P}}(\Gamma) associated to a subgraph ΓΓ\Gamma^{\prime}\subset\Gamma. It follows from lemma 3.3 that the exceptional divisor ELPLE_{L}\subset P_{L} in the blowup of LL can be identified with

(3.9) EL=(Γ)×(Γ//Γ).E_{L}={\mathbb{P}}(\Gamma^{\prime})\times{\mathbb{P}}(\Gamma/\!/\Gamma^{\prime}).

Let e(Γ)=eΓeNe(\Gamma^{\prime})=\sum_{e\in\Gamma^{\prime}}e\subset N_{\mathbb{R}}, and write τ(Γ)=0e(Γ)\tau(\Gamma^{\prime})={\mathbb{R}}^{\geq 0}\cdot e(\Gamma^{\prime}). The subgroup e(Γ)N{\mathbb{Z}}\cdot e(\Gamma^{\prime})\subset N determines a 11-parameter subgroup G(Γ)T=Spec[M]G(\Gamma^{\prime})\subset T={\rm Spec\,}{\mathbb{Q}}[M]. It follows from (3.9) that G(Γ)G(\Gamma^{\prime}) acts trivially on ELE_{L}. One has τ(Γ)CCe\tau(\Gamma^{\prime})\subset C^{\prime}\subset C_{e} for all eΓe\not\in\Gamma^{\prime}, where CC^{\prime} is the cone generated by the edges of Γ\Gamma^{\prime}. For all eΓe^{\prime}\in\Gamma^{\prime} we define a subcone Ce,eCeC_{e,e^{\prime}}\subset C_{e} to be spanned by τ(Γ)\tau(\Gamma^{\prime}) together with all edges of Γ\Gamma except e,ee,e^{\prime}. The fan for PLP_{L} is then

(3.10) {Ce,eΓ}{Ce,e,eΓ,eΓ}.\{C_{e},\ e\in\Gamma^{\prime}\}\cup\{\ C_{e,e^{\prime}},\ e\not\in\Gamma^{\prime},e^{\prime}\in\Gamma^{\prime}\}.

Note that Ce,eΓC_{e},\ e\not\in\Gamma^{\prime} is not a cone in the fan for PLP_{L}. More generally, let {\mathcal{F}} be the fan for P(Γ)P(\Gamma). Certainly, {\mathcal{F}} will contain as cones the half-lines τ(Γ)\tau(\Gamma^{\prime}) for all core subgraphs ΓΓ\Gamma^{\prime}\subset\Gamma as well as the 0e,eΓ{\mathbb{R}}^{\geq 0}e,\ e\in\Gamma. but we must make precise which subsets of this set of half-lines span higher dimensional cones in {\mathcal{F}}. By general theory, the cones correspond to the nonempty orbits. In other words,

(3.11) 0e1,,0ep,0e(Γ1),,0e(Γq){\mathbb{R}}^{\geq 0}e_{1},\dotsc,{\mathbb{R}}^{\geq 0}e_{p},{\mathbb{R}}^{\geq 0}e(\Gamma_{1}),\dotsc,{\mathbb{R}}^{\geq 0}e(\Gamma_{q})

span a cone in {\mathcal{F}} if and only if the intersection

(3.12) E1EqD1Dp,E_{1}\cap\cdots\cap E_{q}\cap D_{1}\cap\cdots\cap D_{p}\neq\emptyset,

where EiP(Γ)E_{i}\subset P(\Gamma) is the exceptional divisor corresponding to L(Γi)L(\Gamma_{i}) and DjP(Γ)D_{j}\subset P(\Gamma) is the strict transform of the coordinate divisor Aei=0A_{e_{i}}=0 in (Γ){\mathbb{P}}(\Gamma). To understand (3.12), consider the simple case E1D1E_{1}\cap D_{1}. We have a core subgraph Γ1Γ\Gamma_{1}\subset\Gamma, and an edge e1e_{1} of Γ\Gamma. We know by lemma 3.3 that E1(Γ1)×(Γ//Γ1)E_{1}\cong{\mathbb{P}}(\Gamma_{1})\times{\mathbb{P}}(\Gamma/\!/\Gamma_{1}). If e1e_{1} is an edge of Γ1\Gamma_{1}, then D1E1=(Γ1//e1)×(Γ//Γ1)D_{1}\cap E_{1}={\mathbb{P}}(\Gamma_{1}/\!/e_{1})\times{\mathbb{P}}(\Gamma/\!/\Gamma_{1}). Otherwise

D1E1=(Γ1)×((Γ//Γ1)//e1).D_{1}\cap E_{1}={\mathbb{P}}(\Gamma_{1})\times{\mathbb{P}}((\Gamma/\!/\Gamma_{1})/\!/e_{1}).

One (degenerate) possibility is that e1e_{1} is an edge of Γ1\Gamma_{1} which forms a loop (tadpole). In this case, e1e_{1} is itself a core subgraph of Γ\Gamma, and the divisor D1D_{1} should be treated as one of the exceptional divisors EiE_{i}. Thus, we omit this possibility. Another possibility is that e1Γ1e_{1}\not\in\Gamma_{1}, but that the image of e1e_{1} in Γ//Γ1\Gamma/\!/\Gamma_{1} forms a loop. In this case, Γ2:=Γ1e1\Gamma_{2}:=\Gamma_{1}\cup e_{1} is a core subgraph, so the linear space L2:Ae=0,eΓ2L_{2}:A_{e}=0,\ e\in\Gamma_{2} gets blown up in the process of constructing P(Γ)P(\Gamma). But blowing L2L_{2} separates E1E_{1} and D1D_{1}, so the intersection of the strict transforms of D1D_{1} and E1E_{1} in P(Γ)P(\Gamma) is empty. The general argument to show that (3.12) is empty if and only if the conditions of (ii) in the proposition are fulfilled is similar and is left for the reader. Note that the case where there are no divisors DiD_{i} follows from proposition 3.1(iii). ∎

We are particularly interested in orbits corresponding to filtrations by core subgraphs ΓpΓ1Γ\Gamma_{p}\subsetneq\cdots\subsetneq\Gamma_{1}\subsetneq\Gamma. Let VP(Γ)V\subset P(\Gamma) be the closure of this orbit. We want to exhibit a toric neighborhood of VV which retracts onto VV as a vector bundle of rank pp. As in the proof of proposition 3.4 we have e(Γi):=eΓiee(\Gamma_{i}):=\sum_{e\in\Gamma_{i}}e. The cone CC spanned by the e(Γi)e(\Gamma_{i}) lies in the fan {\mathcal{F}}. For cones CC^{\prime}\in{\mathcal{F}} we write C>CC^{\prime}>C if CC is a subcone of CC^{\prime}. By the general theory, this will happen if and only if CCC\subset C^{\prime} is a subcone which appears on the boundary of CC^{\prime}. The orbit corresponding to CC^{\prime} will then appear in the closure of the orbit for CC.

Proposition 3.6.

With notation as above, Let C{\mathcal{F}}_{C}\subset{\mathcal{F}} be the subset of cones CC^{\prime} such that we have CC′′CC^{\prime}\leq C^{\prime\prime}\geq C for some C′′C^{\prime\prime}\in{\mathcal{F}}. Write P0P(Γ)P^{0}\subset P(\Gamma) for the open toric subvariety corresponding to the subfan C{\mathcal{F}}_{C}\subset{\mathcal{F}}. We have VP0P(Γ)V\hookrightarrow P^{0}\subset P(\Gamma). Further there is a retraction π:P0V\pi:P^{0}\to V realizing P0P^{0} as a rank pp vector bundle over VV which is equivariant for the action of the torus TT.

Proof.

One has the following functoriality for toric varieties [9], §1.4. Suppose ϕ:NN′′\phi:N^{\prime}\to N^{\prime\prime} is a homomorphism of lattices (finitely generated free abelian groups). Let ,′′{\mathcal{F}}^{\prime},{\mathcal{F}}^{\prime\prime} be fans in N,N′′N_{\mathbb{R}}^{\prime},N_{\mathbb{R}}^{\prime\prime}. Suppose for each cone σ\sigma^{\prime}\in{\mathcal{F}}^{\prime} there exists a cone σ′′′′\sigma^{\prime\prime}\in{\mathcal{F}}^{\prime\prime} such that ϕ(σ)σ′′\phi(\sigma^{\prime})\subset\sigma^{\prime\prime}. Then there is an induced map on toric varieties V()V(′′)V({\mathcal{F}}^{\prime})\to V({\mathcal{F}}^{\prime\prime}). Let N=N=n/N^{\prime}=N={\mathbb{Z}}^{n}/{\mathbb{Z}} as above, and N′′=N/(e(Γ1)++e(Γp))N^{\prime\prime}=N^{\prime}/({\mathbb{Z}}e(\Gamma_{1})+\cdots+{\mathbb{Z}}e(\Gamma_{p})). One has the evident surjection ϕ:NN′′\phi:N^{\prime}\twoheadrightarrow N^{\prime\prime}. We take as fan =C{\mathcal{F}}^{\prime}={\mathcal{F}}_{C}\subset{\mathcal{F}}. The closure VV of the orbit corresponds to the fan ′′{\mathcal{F}}^{\prime\prime} in N′′N_{\mathbb{R}}^{\prime\prime} given by the images of all cones C′′CC^{\prime\prime}\geq C (op. cit. §3.1). Such a C′′C^{\prime\prime} is generated by e(Γ1),,e(Γp),f1,,fqe(\Gamma_{1}),\dotsc,e(\Gamma_{p}),f_{1},\dotsc,f_{q}, and there are no linear relations among these elements (remark 3.5). A subcone CC′′C^{\prime}\leq C^{\prime\prime} is generated by a subset e(Γi1),,e(Γia),f1,,fbe(\Gamma_{i_{1}}),\dotsc,e(\Gamma_{i_{a}}),f_{1},\dotsc,f_{b}. The image is simply the cone in N′′N_{\mathbb{R}}^{\prime\prime} generated by the images of the ff’s. If we have another cone C1C1′′CC_{1}^{\prime}\leq C_{1}^{\prime\prime}\geq C in {\mathcal{F}}^{\prime} with the same image in ′′{\mathcal{F}}^{\prime\prime}, it will have generators say g1,,gbg_{1},\dotsc,g_{b} together with some of the e(Γi)e(\Gamma_{i})’s. Reordering the gg’s, we find that there are relations

(3.13) fi+aije(Γj)=gi+bije(Γj)f_{i}+\sum a_{ij}e(\Gamma_{j})=g_{i}+\sum b_{ij}e(\Gamma_{j})

with aij,bij0a_{ij},b_{ij}\geq 0. It follows that the cones in {\mathcal{F}} spanned by fi,e(Γ1),,e(Γp)f_{i},e(\Gamma_{1}),\dotsc,e(\Gamma_{p}) and gi,e(Γ1),,e(Γp)g_{i},e(\Gamma_{1}),\dotsc,e(\Gamma_{p}) meet in a subset strictly larger that the cone spanned by the e(Γj)e(\Gamma_{j}). By the fan axioms, the intersection of two cones in a fan is a common face of both, so these two cones coincide, which implies fi=gif_{i}=g_{i}. In particular, for each cone in ′′{\mathcal{F}}^{\prime\prime}, there is a unique minimal cone in {\mathcal{F}}^{\prime} lying over it. This is the hypothesis for [19], p. 58, proposition 1.33. One concludes that the map π:P0V\pi:P^{0}\to V induced by the map ′′{\mathcal{F}}^{\prime}\to{\mathcal{F}}^{\prime\prime} is an equivariant fibration, with fibre the toric bundle associated to the fan generated by the e(Γi), 1ipe(\Gamma_{i}),\ 1\leq i\leq p. This toric variety is just affine pp-space, so we get an equivariant 𝔸p{\mathbb{A}}^{p}-fibration over VV. Any such fibration is necessarily a vector bundle with structure group 𝔾mp{\mathbb{G}}_{m}^{p}. Indeed, this amounts to saying that any automorphism of the polynomial ring k[x1,,xp]k[x_{1},\dotsc,x_{p}] which intertwines the diagonal action of 𝔾mp{\mathbb{G}}_{m}^{p} is necessarily of the form xicixix_{i}\mapsto c_{i}x_{i} with cik×c_{i}\in k^{\times}. ∎

Remark 3.7.

We will need to understand how these constructions are compatible. Let VV be a closed orbit corresponding to a cone CC as above, and let V1VV_{1}\subset V be a smaller closed orbit corresponding to a larger cone C1>CC_{1}>C. (The correspondence between cones and orbits is inclusion-reversing.) As above we have a toric variety V1P10P(Γ)V_{1}\subset P_{1}^{0}\subset P(\Gamma) and a retraction π1:P10V1\pi_{1}:P_{1}^{0}\to V_{1}. The fan 1{\mathcal{F}}_{1}^{\prime} for P10P_{1}^{0} is given by the set of cones C1C_{1}^{\prime} in {\mathcal{F}} such that

(3.14) C1C′′C1(>C).C_{1}^{\prime}\leq C^{\prime\prime}\geq C_{1}\ (>C).

It follows that 1=C{\mathcal{F}}_{1}^{\prime}\subset{\mathcal{F}}^{\prime}={\mathcal{F}}_{C}, so P10P0P_{1}^{0}\subset P^{0} is an open subvariety. Let V0VV^{0}\subset V be the image of the composition P10P0𝜋VP_{1}^{0}\subset P^{0}\xrightarrow{\pi}V. Then V0V^{0} is the open toric subvariety of VV corresponding as above to the closed orbit V1VV_{1}\subset V, and we have a retraction V0πVV1V^{0}\xrightarrow{\pi_{V}}V_{1}. One gets commutative diagrams

(3.15) P10=P10P0π1ππV1πVV0V\begin{CD}P_{1}^{0}=P_{1}^{0}@>{\subset}>{}>P^{0}\\ @V{\pi_{1}}V{}V@V{\pi}V{}V@V{\pi}V{}V\\ V_{1}@<{\pi_{V}}<{}<V^{0}@>{\subset}>{}>V\end{CD}

and

(3.16) P0|V1P10ππ1V1=V1.\begin{CD}P^{0}|V_{1}@>{\subset}>{}>P_{1}^{0}\\ @V{\pi}V{}V@V{\pi_{1}}V{}V\\ V_{1}=V_{1}.\end{CD}
Remark 3.8.

Using the toric structure, one can realize these vector bundles as direct sums of line bundles corresponding to characters of the tori acting on the fibres.The inclusion on the top line of (3.16) corresponds to characters which act trivially on all of VV.

Remark 3.9.

(compare proposition 3.4). Given a flag of core subgraphs

(3.17) ΓpΓp1Γ1Γ,\Gamma_{p}\subsetneq\Gamma_{p-1}\subsetneq\cdots\subsetneq\Gamma_{1}\subsetneq\Gamma,

let Li(Γ)L_{i}\subset{\mathbb{P}}(\Gamma) be defined by the edge variables for edges in Γi\Gamma_{i}, so we have L1Lp(Γ)L_{1}\subsetneq\cdots\subsetneq L_{p}\subsetneq{\mathbb{P}}(\Gamma). For L(Γ)L\subset{\mathbb{P}}(\Gamma) a coordinate linear space, let T(L)LT(L)\subset L be the subtorus where none of the coordinates vanish. Then the orbit associated to (3.17) is

(3.18) T(L1)×T(L2/L1)××T(Lp/Lp1)×T(n1/Lp)T(L_{1})\times T(L_{2}/L_{1})\times\cdots\times T(L_{p}/L_{p-1})\times T({\mathbb{P}}^{n-1}/L_{p})

(Here the notation Li+1/LiL_{i+1}/L_{i} is as in (3.2).)

4. Topological Chains on Toric Varieties

One can define the notion of non-negative real points V(0)V({\mathbb{R}}^{\geq 0}) and positive real points V(>0)V({\mathbb{R}}^{>0}). For a torus T=Spec[N]T={\rm Spec\,}{\mathbb{Q}}[N^{\vee}] for some NgN\cong{\mathbb{Z}}^{g} we take

T(>0)={ϕ:[N]|ϕ(n)>0,nN}.T({\mathbb{R}}^{>0})=\{\phi:{\mathbb{Q}}[N^{\vee}]\to{\mathbb{R}}\ |\ \phi(n)>0,\forall n\in N^{\vee}\}.

A toric variety VV can be stratified as a disjoint union of tori V=TαV=\coprod T_{\alpha}. Define

(4.1) V(0)=Tα(>0);\displaystyle V({\mathbb{R}}^{\geq 0})=\coprod T_{\alpha}({\mathbb{R}}^{>0});
V(>0)=T(>0),\displaystyle V({\mathbb{R}}^{>0})=T({\mathbb{R}}^{>0}),

where TVT\subset V is the open orbit. Let VP(Γ)V\subset P(\Gamma) be the closure of the orbit associated to a flag (3.17), and let T(V)T=Spec[N]T(V)\subset T={\rm Spec\,}{\mathbb{Q}}[N^{\vee}] be the subtorus acting trivially on VV. Let πV:PVV\pi_{V}:P_{V}\to V be the vector bundle as in proposition 3.6. We write PV=1pP_{V}={\mathcal{L}}_{1}\oplus\cdots\oplus{\mathcal{L}}_{p} as a direct sum of line bundles, where each i{\mathcal{L}}_{i} is equivariant for T(V)T(V). Let K(V)(S1)pT(V)()K(V)\cong(S^{1})^{p}\subset T(V)({\mathbb{C}}) be the maximal compact subgroup. Note that one has a canonical identification T(V)=𝔾mpT(V)={\mathbb{G}}_{m}^{p} associated to the 11-parameter subgroups of T(V)T(V) generated by e(Γi)Ne(\Gamma_{i})\in N. In particular, the identification K(V)=(S1)pK(V)=(S^{1})^{p} is canonical as well. For all closed orbits VV we may fix metrics on the i{\mathcal{L}}_{i} which are compatible under inclusions (3.16) and are (necessarily) invariant under the action of K(V)K(V). We fix also a constant η>0\eta>0. We can then define SVηPVS_{V}^{\eta}\subset P_{V} to be the product of the circle bundles of radius η\eta embedded in the i{\mathcal{L}}_{i}. SVηS_{V}^{\eta} becomes a principal bundle over VV with structure group K(V)K(V). Note that SVηPV(0)S^{\eta}_{V}\cap P_{V}({\mathbb{R}}^{\geq 0}) contains a unique point in every fibre of SVηS_{V}^{\eta} over a point of V()V({\mathbb{R}}). Let 0<ε<<η0<\varepsilon<<\eta be another constant. We need to define a chain σVη,εV(>0)\sigma_{V}^{\eta,\varepsilon}\subset V({\mathbb{R}}^{>0}). We consider closures V1VV_{1}\subset V of codimension 11 orbits in VV. For each such V1V_{1} we have an open P(V)1VP(V)_{1}\subset V and a retraction P(V)1V1P(V)_{1}\to V_{1} which is a line bundle with a metric. The fibres of P(V)1(>0)P(V)_{1}({\mathbb{R}}^{>0}) have a canonical coordinate r>0r>0. If V1V_{1} corresponds to an intersection of V=E1EpV=E_{1}\cap\cdots\cap E_{p} with another exceptional divisor Ep+1E_{p+1}, then we remove from each fibre of P(V)1(>0)P(V)_{1}({\mathbb{R}}^{>0}) over V1(>0)V_{1}({\mathbb{R}}^{>0}) the locus where r<ηr<\eta. If, on the other hand V1V_{1} corresponds to an intersection of V with one of the DiD_{i} (i.e. with a strict transform of one of the coordinate divisors), then we remove the locus r<εr<\varepsilon. Repeating this process for each V1V_{1} (i.e. for each irreducible toric divisor in VV), we obtain a compact σVη,εV(>0)\sigma_{V}^{\eta,\varepsilon}\subset V({\mathbb{R}}^{>0}) which stays away from the boundary components. (Here ”boundary components” are exceptional divisors together with strict transforms of coordinate divisors.)

Example 4.1.

Consider the case V=P(Γ)V=P(\Gamma). Let π:P(Γ)(Γ)\pi:P(\Gamma)\to{\mathbb{P}}(\Gamma), and let

σ={(A1,,An)|Ai0}(Γ)()\sigma=\{(A_{1},\dotsc,A_{n})\ |\ A_{i}\geq 0\}\subset{\mathbb{P}}(\Gamma)({\mathbb{R}})

be the original integration chain. We have σP(Γ)η,επ1(σ)\sigma_{P(\Gamma)}^{\eta,\varepsilon}\subset\pi^{-1}(\sigma) defined by excising away points within a distance of η\eta from an EiE_{i} or ε\varepsilon from the strict transform DjD_{j} of a coordinate divisor Aj=0A_{j}=0. (cf. fig.(8)). It is a manifold with corners.

Refer to caption

Figure 8. P(Γ)P(\Gamma) and the real chain σP(Γ)η,ε\sigma_{P(\Gamma)}^{\eta,\varepsilon}.

Define τVη,ε\tau_{V}^{\eta,\varepsilon} to be the inverse image of σVη,ε\sigma_{V}^{\eta,\varepsilon} in SVηS_{V}^{\eta}. The fibres of τVη,ε\tau_{V}^{\eta,\varepsilon} over σVη,ε\sigma_{V}^{\eta,\varepsilon} are products (S1)p(S^{1})^{p} with a canonical origin at the point where this fibre meets PV(0)P_{V}({\mathbb{R}}^{\geq 0}). For an angle 0θ2π0\leq\theta\leq 2\pi, we can thus define τVη,ε,θτVη,ε\tau_{V}^{\eta,\varepsilon,\theta}\subset\tau_{V}^{\eta,\varepsilon} to be swept out by the origin in each fibre under the action of [0,θ]pK(V)[0,\theta]^{p}\subset K(V). The chains τVη,ε,θ\tau_{V}^{\eta,\varepsilon,\theta} have {\mathbb{R}}-dimension n1n-1 which is equal to the complex dimension of (Γ){\mathbb{P}}(\Gamma) and P(Γ)P(\Gamma).

Example 4.2.

Here is an example which is too simple to correspond to any graph, but is sufficient to clarify the toric picture. Take

(4.2) L1:A1=A2=0;L2:A2=0L_{1}:A_{1}=A_{2}=0;\quad L_{2}:A_{2}=0

in 2{\mathbb{P}}^{2} with coordinates A1,A2,A3A_{1},A_{2},A_{3}. Take P𝜋2P\xrightarrow{\pi}{\mathbb{P}}^{2} to be the blowup of L1=(0,0,1)L_{1}=(0,0,1). Let E1PE_{1}\subset P be the exceptional divisor, and let E2PE_{2}\subset P be the strict transform of L2L_{2}. Note that E2E_{2} is already a divisor so it is not necessary to blow up again. Take V=E11V=E_{1}\cong{\mathbb{P}}^{1}. The fan {\mathcal{F}} for PP is fig.(9).

Refer to caption

Figure 9. Fan for Example 4.2.

The cone C=0(e1+e2)C={\mathbb{R}}^{\geq 0}\cdot(e_{1}+e_{2}), so the fan =C{\mathcal{F}}^{\prime}={\mathcal{F}}_{C}\subset{\mathcal{F}} is the subset of cones lying in the first quadrant. The toric variety PVP_{V} is 𝔸2{\mathbb{A}}^{2} with (0,0)(0,0) blown up. It projects down onto VV as a line bundle. SVηPV()S_{V}^{\eta}\subset P_{V}({\mathbb{C}}) is then a circle bundle over V()V({\mathbb{C}}). VV has two suborbits V2=E1E2V_{2}=E_{1}\cap E_{2} and V1=E1D1V_{1}=E_{1}\cap D_{1}, where D1D_{1} is the strict transform of the divisor A1=0A_{1}=0 in 2{\mathbb{P}}^{2}. We may interpret z:=A1/A2z:=A_{1}/A_{2} as a coordinate on VV, so V1:z=0V_{1}:z=0 and V2:z=V_{2}:z=\infty. We have P(V)1=V{z=}P(V)_{1}=V-\{z=\infty\} and P(V)2=V{z=0}P(V)_{2}=V-\{z=0\}. The real chain σVη,ε={ηz1/ε}\sigma_{V}^{\eta,\varepsilon}=\{\eta\leq z\leq 1/\varepsilon\}, and τVη,ε\tau_{V}^{\eta,\varepsilon} is the S1S^{1}-bundle of radius η\eta over σVη,ε\sigma_{V}^{\eta,\varepsilon}. On the other hand, V2V_{2} corresponds to the cone labeled C2C_{2} in fig.(9), and the fan C2{\mathcal{F}}_{C_{2}} is just C2C_{2} itself. The toric variety PV2𝔸2P_{V_{2}}\cong{\mathbb{A}}^{2} is a rank 22 vector bundle over the point V2V_{2}. We have PV2PVP_{V_{2}}\subset P_{V}. In this case σV2η,ε\sigma_{V_{2}}^{\eta,\varepsilon} is simply the point V2V_{2}, and τV2η,εS1×S1PV1()\tau_{V_{2}}^{\eta,\varepsilon}\cong S^{1}\times S^{1}\subset P_{V_{1}}({\mathbb{C}}). In local coordinates around V1V_{1} given by eigenfunctions for the torus action we have

(4.3) τVη,ε,θ={(ηeiμ,z)|ηz1/ε, 0μθ}\displaystyle\tau_{V}^{\eta,\varepsilon,\theta}=\{(\eta e^{i\mu},z)\ |\ \eta\leq z\leq 1/\varepsilon,\ 0\leq\mu\leq\theta\}
τV1η,ε,θ={(ηeiμ,ηeiν)| 0μ,νθ}\displaystyle\tau_{V_{1}}^{\eta,\varepsilon,\theta}=\{(\eta e^{i\mu},\eta e^{i\nu})\ |\ 0\leq\mu,\nu\leq\theta\}
τVη,ε,θτV1η,ε,θ={(ηeiμ,η)| 0μθ}.\displaystyle\tau_{V}^{\eta,\varepsilon,\theta}\cap\tau_{V_{1}}^{\eta,\varepsilon,\theta}=\{(\eta e^{i\mu},\eta)\ |\ 0\leq\mu\leq\theta\}.

We want now to establish a basic formula for the boundary of the chains τVη,ε,θ\tau_{V}^{\eta,\varepsilon,\theta}. Here VV runs through the closures of orbits in P(Γ)P(\Gamma) associated to flags of core subgraphs (3.17). We include the big orbit V=P(Γ)V=P(\Gamma). We write |V|:=codim(V/P(Γ))|V|:=\text{codim}(V/P(\Gamma)). We may express the boundary chains τVη,ε,θ\partial\tau_{V}^{\eta,\varepsilon,\theta} locally (in fact Zariski-locally) in coordinates which are eigenfunctions for the torus action. It is clear (cf. (4.3)) that boundary terms are obtained by setting a suitable one of these coordinates to be constant: either ηeiθ\eta e^{i\theta} or η\eta or ε\varepsilon. (The presence of 1/ε1/\varepsilon in the first line of (4.3) simply means that the appropriate coordinate near that point is 1/z1/z.)

Proposition 4.3.

For a suitable orientation, the boundary

(4.4) V(1)|V|τVη,ε,θ\partial\sum_{V}(-1)^{|V|}\tau_{V}^{\eta,\varepsilon,\theta}

will contain no chains with one coordinate constant =η=\eta.

Proof.

(Cf. fig.(10) ). For a given boundary term, we can choose local eigenfunction coordinates x1,,xn1x_{1},\dotsc,x_{n-1} such that be boundary term is given by x1=ηx_{1}=\eta. We take the chains to be oriented in some consistant way by this ordering of coordinates. (Note that these coordinates are defined on a Zariski open set. The obstruction to choosing consistent orientations for various open sets is a class in the first Zariski cohomology of P(Γ)P(\Gamma) with constant /2{\mathbb{Z}}/2{\mathbb{Z}}-coefficients. Since this cohomology group vanishes, we can choose such consistent orientations.) If τVη,ε,θ\partial\tau_{V}^{\eta,\varepsilon,\theta} contains a term with x1=ηx_{1}=\eta, there are two possibilities. Either x1x_{1} is a real coordinate on τVη,ε,θ\tau_{V}^{\eta,\varepsilon,\theta} or it is a circular coordinate. If x1x_{1} is a real coordinate, then the fact that x1=ηx_{1}=\eta appears in the boundary means that locally x1=0x_{1}=0 defines a codimension 11 orbit closure V1VV_{1}\hookrightarrow V. In τV1η,ε,θ\partial\tau_{V_{1}}^{\eta,\varepsilon,\theta}, x1x_{1} will appear as a circular coordinate. Since |V|=|V1|+1|V|=|V_{1}|+1, the same chain x1=ηx_{1}=\eta will appear in τVη,ε,θ\partial\tau_{V}^{\eta,\varepsilon,\theta} and in τV1η,ε,θ\partial\tau_{V_{1}}^{\eta,\varepsilon,\theta} and will cancel in (4.4). If, on the other hand, x1=ηeiθx_{1}=\eta e^{i\theta} is a circular coordinate, then for suitable ordering of coordinates, the chain will be an (S1)p(S^{1})^{p}-bundle over a chain σ\sigma contained in the locus where certain coordinates 0\geq 0. But then (4.4) will contain another chain which is an (S1)p1(S^{1})^{p-1}-bundle over {x1η}×σ\{x_{1}\geq\eta\}\times\sigma, and the boundary components involving x1=ηx_{1}=\eta will occur with opposite signs and will cancel. ∎

The boundary chain (4.4) is an (n2)(n-2)-chain involving two scales 0<ε<η0<\varepsilon<\eta. We want to construct an (n1)(n-1)-chain ξη,ε,θ\xi^{\eta,\varepsilon,\theta} which amounts to a scaling ηε\eta\to\varepsilon. To do this, we construct a vector field vv on P(Γ)P(\Gamma). Let E=EiE=\sum E_{i} be the exceptional divisor. vv will be 0 outside a neighborhood NN of EE. Locally, at a point on NN which is close to divisors E1,,EpE_{1},\dotsc,E_{p} we have coordinates x1,,xpx_{1},\dotsc,x_{p} which are eigenfunctions for the torus action such that locally Ei:xi=0E_{i}:x_{i}=0. Locally we will take vv to be radial and inward-pointing in each xix_{i}. We glue these local vv’s using a partition of unity. ”Flowing” the (n2)(n-2)-chain (4.4) along this vector field yields an (n1)(n-1)-chain ξη,ε,θ\xi^{\eta,\varepsilon,\theta}. If this is done with care, we can arrange

(4.5) ξη,ε,θV(1)|V|τVη,ε,θV(1)|V|τVε,ε,θ.\partial\xi^{\eta,\varepsilon,\theta}\equiv\partial\sum_{V}(-1)^{|V|}\tau_{V}^{\eta,\varepsilon,\theta}-\partial\sum_{V}(-1)^{|V|}\tau_{V}^{\varepsilon,\varepsilon,\theta}.

Here \equiv means that the two sides differ by a chain lying in an ε\varepsilon-neighborhood of the strict transform DD of the coordinate divisor Δ\Delta in P(Γ)P(\Gamma). Another important property of the chain ξη,ε,θ\xi^{\eta,\varepsilon,\theta} is

Lemma 4.4.

ξη,ε,2πξη,ε,0\xi^{\eta,\varepsilon,2\pi}\equiv\xi^{\eta,\varepsilon,0}.

Proof.

The point is that τVη,ε,2π0\partial\tau_{V}^{\eta,\varepsilon,2\pi}\equiv 0 except for the case V=P(Γ)V=P(\Gamma), and τP(Γ)η,ε,θ\tau_{P(\Gamma)}^{\eta,\varepsilon,\theta} is independent of θ\theta. (See fig.(10)).

Refer to caption

Figure 10. The monodromy chain, with angular variable θ\theta.

Define the chain cη,ε,θ=V(1)|V|τVη,ε,θξη,ε,θc^{\eta,\varepsilon,\theta}=\sum_{V}(-1)^{|V|}\tau_{V}^{\eta,\varepsilon,\theta}-\xi^{\eta,\varepsilon,\theta}. We have

(4.6) cη,ε,θ=V(1)|V|τVε,ε,θ.\partial c^{\eta,\varepsilon,\theta}=\partial\sum_{V}(-1)^{|V|}\tau_{V}^{\varepsilon,\varepsilon,\theta}.

Note that cη,ε,0=σP(Γ)η,εc^{\eta,\varepsilon,0}=\sigma_{P(\Gamma)}^{\eta,\varepsilon}, i.e. all chains involving at least one circular variable die at θ=0\theta=0. We define the variation,

(4.7) var(cη,ε,θ)=cη,ε,2πcη,ε,0VP(Γ)(1)|V|τVε,ε.var(c^{\eta,\varepsilon,\theta})=c^{\eta,\varepsilon,2\pi}-c^{\eta,\varepsilon,0}\equiv\sum_{V\subsetneq P(\Gamma)}(-1)^{|V|}\tau_{V}^{\varepsilon,\varepsilon}.

It is a sum of “(S1)p(S^{1})^{p}-tubes” over all E1EpP(Γ)E_{1}\cap\cdots\cap E_{p}\subsetneq P(\Gamma).

5. The Graph Hypersurface

Associated to a graph Γ\Gamma with nn edges, one has the graph polynomial

(5.1) ψΓ(A1,,An)=TeTAe\psi_{\Gamma}(A_{1},\dotsc,A_{n})=\sum_{T}\prod_{e\not\in T}A_{e}

where TT runs through spanning trees of Γ\Gamma. This polynomial has degree h1(Γ)h_{1}(\Gamma). For more detail, see [2] and the references cited there. Let X=XΓ:ψΓ=0X=X_{\Gamma}:\psi_{\Gamma}=0 be the graph hypersurface in n1{\mathbb{P}}^{n-1}. For μEdge(Γ)\mu\subset\text{Edge}(\Gamma), let Lμ(Γ)L_{\mu}\subset{\mathbb{P}}(\Gamma) be defined by Ae=0,eμA_{e}=0,\ e\in\mu. Let Γμ=eμeΓ\Gamma_{\mu}=\bigcup_{e\in\mu}e\subset\Gamma be the subgraph with edges in μ\mu. Note the dictionary ΓμLμ\Gamma_{\mu}\leftrightarrow L_{\mu} is inclusion reversing.

Lemma 5.1.

(i) LμXΓ(Γ)L_{\mu}\subset X_{\Gamma}\subset{\mathbb{P}}(\Gamma) if and only if h1(Γμ)>0h_{1}(\Gamma_{\mu})>0.
(ii) If h1(Γμ)>0h_{1}(\Gamma_{\mu})>0, there exists a unique νμ\nu\subseteq\mu such that h1(Γν)=h1(Γμ)h_{1}(\Gamma_{\nu})=h_{1}(\Gamma_{\mu}) and such that moreover Γν\Gamma_{\nu} is a core graph.
(iii) We have in (ii) that ν=ξ\nu=\bigcup\xi where ξ\xi runs through all minimal subsets of μ\mu such that LξXL_{\xi}\subset X.
(iv) Lμ=LνML_{\mu}=L_{\nu}\cap M, where MM is a coordinate linear space not contained in XΓX_{\Gamma}.

Proof.

These assertions are straightforward from the results in [2], section 3. Note that (iv) justifies our strategy of only blowing up core subgraphs. ∎

We have seen (remark 3.4) that our blowup P(Γ)P(\Gamma) is stratified as a union of tori indexed by pairs

(5.2) (F,{ΓpΓ1Γ//γ})(F,\{\Gamma_{p}\subsetneq\cdots\subsetneq\Gamma_{1}\subsetneq\Gamma/\!/\gamma\})

where FΓF\subset\Gamma is a suitable subforest and the Γi\Gamma_{i} are core.

Proposition 5.2.

(i) As in proposition 3.4, the torus corresponding to (5.2) is

(5.3) T(Γp//Fp)×T((Γp1//Γp)//Fp1)××T((Γ//Γ1)//F).T(\Gamma_{p}/\!/F_{p})\times T((\Gamma_{p-1}/\!/\Gamma_{p})/\!/F_{p-1})\times\cdots\times T((\Gamma/\!/\Gamma_{1})/\!/F).

Here T(Γ):=(Γ)ΔT(\Gamma):={\mathbb{P}}(\Gamma)-\Delta, where Δ:eEdge(Γ)Ae=0\Delta:\prod_{e\in\text{Edge}(\Gamma)}A_{e}=0.
(ii) The strict transform YY of XΓX_{\Gamma} in P(Γ)P(\Gamma) meets the stratum (5.3) in a union of pullbacks

(5.4) pr11(XΓp0)pr21(XΓp1//Γp0)prp1(X(Γ//γ)//Γ10).pr_{1}^{-1}(X^{0}_{\Gamma_{p}})\cup pr_{2}^{-1}(X^{0}_{\Gamma_{p-1}/\!/\Gamma_{p}})\cup\cdots\cup pr_{p}^{-1}(X^{0}_{(\Gamma/\!/\gamma)/\!/\Gamma_{1}}).

Here the pripr_{i} are the projections to the various subtori in (5.3), and X0X^{0} denotes the restriction of the corresponding graph hypersurface to the open torus in the projective space.

Proof.

Let ΓΓ\Gamma^{\prime}\subset\Gamma be a subgraph and let L:Ae=0,eEdge(Γ)L:A_{e}=0,\ e\in\text{Edge}(\Gamma^{\prime}). Assume h1(Γ)>0h_{1}(\Gamma^{\prime})>0, so LXΓL\subset X_{\Gamma}. Let PL(Γ)P_{L}\to{\mathbb{P}}(\Gamma) be the blowup of LL. Let ELPLE_{L}\subset P_{L} be the exceptional divisor, and let YLPLY_{L}\subset P_{L} be the strict transform of XΓX_{\Gamma}. The basic geometric result (op. cit. prop. 3.5) is that EL=(Γ)×(Γ//Γ)E_{L}={\mathbb{P}}(\Gamma^{\prime})\times{\mathbb{P}}(\Gamma/\!/\Gamma^{\prime}) and

(5.5) YLEL=(XΓ×(Γ//Γ))((Γ)×XΓ//Γ).Y_{L}\cap E_{L}=\Big{(}X_{\Gamma^{\prime}}\times{\mathbb{P}}(\Gamma/\!/\Gamma^{\prime})\Big{)}\cup\Big{(}{\mathbb{P}}(\Gamma^{\prime})\times X_{\Gamma/\!/\Gamma^{\prime}}\Big{)}.

The assertions of the proposition follow by an induction argument. ∎

Corollary 5.3.

The strict transform YY of XΓX_{\Gamma} in P(Γ)P(\Gamma) does not meet the non-negative points P(Γ)(0)P(\Gamma)({\mathbb{R}}^{\geq 0}) (4.1).

Proof.

It suffices by (4.1) to show that YY doesn’t meet the positive points in any stratum. By proposition 5.2, it suffices to show that for any graph Γ\Gamma, the graph hypersurface XΓX_{\Gamma} has no {\mathbb{R}}-points with coordinates all >0>0. This is immediate because ψΓ\psi_{\Gamma} is a sum of monomials with non-negative coefficients. ∎

Remark 5.4.

The Feynman amplitude is obtained by calculating an integral over σ=(Γ)(0)\sigma={\mathbb{P}}(\Gamma)({\mathbb{R}}^{\geq 0}) with an integrand which has a pole along XΓX_{\Gamma}. Again using that ψΓ\psi_{\Gamma} is a sum of monomials with non-negative coefficients, one sees from lemma 5.1 that

(5.6) σXΓ=μLμ(0)\sigma\cap X_{\Gamma}=\bigcup_{\mu}L_{\mu}({\mathbb{R}}^{\geq 0})

where LμΓμL_{\mu}\leftrightarrow\Gamma_{\mu} with ΓμΓ\Gamma_{\mu}\subset\Gamma a core subgraph. The iterated blowup P(Γ)(Γ)P(\Gamma)\to{\mathbb{P}}(\Gamma) is exactly what is necessary to separate the non-negative real points from the strict transform of XΓX_{\Gamma}.

Remark 5.5.

The points where ψΓ0\psi_{\Gamma}\neq 0 have some remarkable properties. It is shown in [20] that for any angular sector SS with angle <π<\pi, ψΓ(a1,,an)0\psi_{\Gamma}(a_{1},\dotsc,a_{n})\neq 0 at any complex projective point aa such that the ai0a_{i}\neq 0 and all the arg(ai)\arg(a_{i}) lie in SS.

6. Monodromy

Let pi=(0,,1,0,,0)np_{i}=(0,\dotsc,1,0,\dotsc,0)\in{\mathbb{C}}^{n} be the ii-th coordinate vector. Define

σaff={i=1nτipi|τi0,τi=1}n{(0,,0)}n1.\sigma^{aff}=\{\sum_{i=1}^{n}\tau_{i}p_{i}\ |\ \tau_{i}\geq 0,\ \sum\tau_{i}=1\}\subset{\mathbb{C}}^{n}-\{(0,\dotsc,0)\}\to{\mathbb{P}}^{n-1}.

Fix a positive constant ε<<1\varepsilon<<1 and choose qk=(qk1,,qkn)n, 1knq_{k}=(q_{k1},\dotsc,q_{kn})\in{\mathbb{R}}^{n},\ 1\leq k\leq n with 1ε<qkj11-\varepsilon<q_{kj}\leq 1 and |qjkq,m|ε2|q_{jk}-q_{\ell,m}|\leq\varepsilon^{2}. We assume the qkq_{k} are algebraically generic. Write rk(t)=pk+tqknr_{k}(t)=p_{k}+tq_{k}\in{\mathbb{C}}^{n}. Define (cf. fig.(11))

(6.1) σtaff={i=1nτkrk(t)|τk0,τk=1}\sigma^{aff}_{t}=\{\sum_{i=1}^{n}\tau_{k}r_{k}(t)\ |\ \tau_{k}\geq 0,\ \sum\tau_{k}=1\}

Refer to caption

Figure 11. Moving Δt\Delta_{t}.

We write σ\sigma and σ~t\widetilde{\sigma}_{t} for the images of these chains in n1{\mathbb{P}}^{n-1}. Of course, σ=σn1\sigma=\sigma_{{\mathbb{P}}^{n-1}} as above, and we know that σXΓ=LσL\sigma\cap X_{\Gamma}=\bigcup_{L\subset{\mathcal{L}}}\sigma_{L}. Here {\mathcal{L}} is as in (3.4).

Lemma 6.1.

Let N{\mathcal{L}}\subset N_{\mathcal{L}} be a neighborhood of {\mathcal{L}} in n1{\mathbb{P}}^{n-1} and let σNσ\sigma\subset N_{\sigma} be a neighborhood of σ\sigma. Then there exists ε0>0\varepsilon_{0}>0 such that εε0\varepsilon\leq\varepsilon_{0} implies that for all 0θ2π0\leq\theta\leq 2\pi, we have σ~εeiθNσ\widetilde{\sigma}_{\varepsilon e^{i\theta}}\subset N_{\sigma} and σ~εeiθXΓN\widetilde{\sigma}_{\varepsilon e^{i\theta}}\cap X_{\Gamma}\subset N_{\mathcal{L}}.

Proof.

We have σXΓ\sigma\cap X_{\Gamma}\subset{\mathcal{L}}. By compacity, σ~εeiθNσ\widetilde{\sigma}_{\varepsilon e^{i\theta}}\subset N_{\sigma} for ε<<1\varepsilon<<1. Again by compacity, if we shrink NσN_{\sigma} we will have NσXΓNN_{\sigma}\cap X_{\Gamma}\subset N_{\mathcal{L}}. ∎

Remark 6.2.

Write Hk,tH_{k,t} for the projective span of the points

r1(t),,rk(t)^,,rn(t),r_{1}(t),\dotsc,\widehat{r_{k}(t)},\dotsc,r_{n}(t),

and let Δt=k=1nHk,t\Delta_{t}=\bigcup_{k=1}^{n}H_{k,t}. Thus, Δ=Δ0\Delta=\Delta_{0} and we may consider the monodromy for Δεeiθ, 0θ2π\Delta_{\varepsilon e^{i\theta}},\ 0\leq\theta\leq 2\pi. More precisely, renormalization in physics involves an integral over the chain σ\sigma. The integrand has poles along XΓX_{\Gamma}. Since σXΓ\sigma\cap X_{\Gamma}\neq\emptyset, the integral is possibly divergent. On the other hand, by corollary 5.3, the chain σε\sigma_{\varepsilon} does not meet XΓX_{\Gamma} and so represents a singular homology class

(6.2) [σε]Hn1(n1XΓ,ΔεΔεXΓ,).[\sigma_{\varepsilon}]\in H_{n-1}({\mathbb{P}}^{n-1}-X_{\Gamma},\Delta_{\varepsilon}-\Delta_{\varepsilon}\cap X_{\Gamma},{\mathbb{Z}}).

(Since all qkj>0q_{kj}>0, it follows that σεσ\sigma_{\varepsilon}\subset\sigma, and points in σε\sigma_{\varepsilon} have all coordinates >0>0.) We consider the topological pairs (n1XΓ,ΔεeiθΔεeiθXΓ)({\mathbb{P}}^{n-1}-X_{\Gamma},\Delta_{\varepsilon e^{i\theta}}-\Delta_{\varepsilon e^{i\theta}}\cap X_{\Gamma}) as a family over the circle and we continuously deform our chain σε\sigma_{\varepsilon} to a family of chains σεeiθ\sigma_{\varepsilon e^{i\theta}} on n1XΓ{\mathbb{P}}^{n-1}-X_{\Gamma} with boundary on ΔεeiθΔεeiθXΓ\Delta_{\varepsilon e^{i\theta}}-\Delta_{\varepsilon e^{i\theta}}\cap X_{\Gamma}. (We will not be able to take σεeiθ=σ~εeiθ\sigma_{\varepsilon e^{i\theta}}=\widetilde{\sigma}_{\varepsilon e^{i\theta}} because this chain can meet XΓX_{\Gamma}.) The monodromy map mm is an automorphism of (6.2) obtained by winding around the circle: m(σε)=σεe2πim(\sigma_{\varepsilon})=\sigma_{\varepsilon e^{2\pi i}}. We will calculate m(σε)m(\sigma_{\varepsilon}) and see that it determines in a natural way the renormalization expansion we want.

Recall we have π:P(Γ)(Γ)\pi:P(\Gamma)\to{\mathbb{P}}(\Gamma), and π1(XΓ)=YΓE\pi^{-1}(X_{\Gamma})=Y_{\Gamma}\cup E, where Y=YΓY=Y_{\Gamma} is the strict transform of XΓX_{\Gamma} and E=EiE=\bigcup E_{i} is the exceptional divisor. (The EiE_{i} are closures of orbits associated to core subgraphs of Γ\Gamma.) We may transfer our monodromy problem to P(Γ)P(\Gamma). Δεeiθ\Delta_{\varepsilon e^{i\theta}} is in general position with respect to the blowups, so we obtain a family of divisors Δεeiθ=πΔεeiθ\Delta_{\varepsilon e^{i\theta}}^{\prime}=\pi^{*}\Delta_{\varepsilon e^{i\theta}} on P(Γ)P(\Gamma). Since π:P(Γ)EYΓ(Γ)XΓ\pi:P(\Gamma)-E-Y_{\Gamma}\cong{\mathbb{P}}(\Gamma)-X_{\Gamma}, we have an isomorphism of topological pairs

(6.3) (P(Γ)EYΓ,ΔεeiθΔεeiθ(EYΓ))((Γ)XΓ,ΔεeiθΔεeiθXΓ).\Big{(}P(\Gamma)-E-Y_{\Gamma},\Delta_{\varepsilon e^{i\theta}}^{\prime}-\Delta_{\varepsilon e^{i\theta}}^{\prime}\cap(E\cup Y_{\Gamma})\Big{)}\cong\Big{(}{\mathbb{P}}(\Gamma)-X_{\Gamma},\Delta_{\varepsilon e^{i\theta}}-\Delta_{\varepsilon e^{i\theta}}\cap X_{\Gamma}\Big{)}.

In section 4 we have defined chains τVη,ε,θ,ξη,ε,θ,cη,ε,θ\tau_{V}^{\eta,\varepsilon,\theta},\xi^{\eta,\varepsilon,\theta},c^{\eta,\varepsilon,\theta} on P(Γ)P(\Gamma). These chains sit on (or, in the case of ξ\xi, within) various (S1)p(S^{1})^{p}-bundles over P(Γ)(0)P(\Gamma)({\mathbb{R}}^{\geq 0}) where the S1S^{1} have radius η\eta with respect to a chosen metric. From corollary 5.3 it follows that for 0<η<<10<\eta<<1, none of these chains meets YΓY_{\Gamma}. By construction, these chains do not meet EE, so they may be identified with chains on (Γ)XΓ{\mathbb{P}}(\Gamma)-X_{\Gamma}. We claim that a small modification of the chains cη,ε,θc^{\eta,\varepsilon,\theta} will represent the monodromy chains σεeiθ\sigma_{\varepsilon e^{i\theta}}. The monodromy chains σεeiθ\sigma_{\varepsilon e^{i\theta}} should have boundary on Δεeiθ\Delta_{\varepsilon e^{i\theta}}. On the other hand, the chains cη,ε,θc^{\eta,\varepsilon,\theta} were cut off so they had boundaries on tubes a distance ε\varepsilon from the toric divisors DjD_{j} given by the strict transforms of the Aj=0A_{j}=0 (see fig.(12)).

Refer to caption

Figure 12. The chain τEη,ε\tau_{E}^{\eta,\varepsilon}.

We must “massage” these brutal cutoffs to get them into Δεeiθ\Delta_{\varepsilon e^{i\theta}}. Our chains τ\tau sit on tubes or products of tubes or products of tubes of radius η\eta which we can think of as lying on n1{\mathbb{P}}^{n-1}-{\mathcal{L}}. Since ε<<η\varepsilon<<\eta, when we deform ΔΔεeiθ\Delta\to\Delta_{\varepsilon e^{i\theta}} the homotopy type of the circles, or product of circles where these divisors intersect the tubes doesn’t change. This may seem strange because Δ{\mathcal{L}}\subset\Delta while Δεeiθ\Delta_{\varepsilon e^{i\theta}} is in general position with respect to {\mathcal{L}}, but the intersections with a hollow tubular neighborhood of {\mathcal{L}} are canonically homotopic. Indeed, we may take Δεeiθ\Delta_{\varepsilon e^{i\theta}} to correspond to a point in a small contractible disk in the moduli space for coordinate simplices around the point corresponding to Δ\Delta. The canonical path up to homotopy between the two points in moduli will induce the desired homotopy on the intersections. (See fig.(13). The two sets of four dots on the circles are canonically homotopic.).

Refer to caption

Figure 13. Homotopy invariance of Δttube over \Delta_{t}\cap\text{tube over }{\mathcal{L}}.

In more detail, by corollary 5.3, the chains τη,ε,θ\tau^{\eta,\varepsilon,\theta} are bounded away from XΓX_{\Gamma} by a bound which is independent of ε\varepsilon as ε0\varepsilon\to 0. Outside of some tubular neighborhood NN of XΓX_{\Gamma} we may find a space MM disjoint from XΓX_{\Gamma} such that MM contains open neighborhoods of both ΔNΔ\Delta-N\cap\Delta and ΔεeiθNΔεeiθ\Delta_{\varepsilon e^{i\theta}}-N\cap\Delta_{\varepsilon e^{i\theta}} and such that we have deformation retractions MΔNΔM\to\Delta-N\cap\Delta and MΔεeiθNΔεeiθM\to\Delta_{\varepsilon e^{i\theta}}-N\cap\Delta_{\varepsilon e^{i\theta}}. Shrinking ε\varepsilon, we may assume our ε\varepsilon-cutoffs lie in MM. We may then use the deformation retract to extend the chain slightly to a chain τ~Vη,ε,θ\tilde{\tau}_{V}^{\eta,\varepsilon,\theta} which bounds on Δεeiθ\Delta_{\varepsilon e^{i\theta}}. It remains to consider the chains ξη,ε,θ\xi^{\eta,\varepsilon,\theta}. Recall these were obtained by flowing the chain V(1)|V|τVη,ε,θ\partial\sum_{V}(-1)^{|V|}\tau_{V}^{\eta,\varepsilon,\theta} inward toward the exceptional divisor EE, so ηε\eta\to\varepsilon (cf. fig.(10)). We are in a small neighborhood of E(0)E({\mathbb{R}}^{\geq 0}) hence by corollary 5.3 we are away from XΓX_{\Gamma}. The point to be checked is that the term V(1)|V|τVε,ε,θ\partial\sum_{V}(-1)^{|V|}\tau_{V}^{\varepsilon,\varepsilon,\theta} is very close to Δεeiθ\Delta_{\varepsilon e^{i\theta}} so by the same deformation retraction argument as above we can extend the chain to bound on Δεeiθ\Delta_{\varepsilon e^{i\theta}}. The subtlety is that we are ε\varepsilon-close to EE as well, so we need the distance from Δεeiθ\Delta_{\varepsilon e^{i\theta}} to be o(ε)o(\varepsilon). Recall (6.1) we have the vertices rk(εeiθ)=[qk1εeiθ,,1+qkkεeiθ,,qknεeiθ]n1r_{k}(\varepsilon e^{i\theta})=[q_{k1}\varepsilon e^{i\theta},\dotsc,1+q_{kk}\varepsilon e^{i\theta},\dotsc,q_{kn}\varepsilon e^{i\theta}]\in{\mathbb{P}}^{n-1}. The coordinate divisor Δεeiθ\Delta_{\varepsilon e^{i\theta}} is determined by these projective points. The projective point does not change if we scale the coordinates by eiθe^{i\theta}, so the image in n1{\mathbb{P}}^{n-1} of the affine simplex below, parametrized by τ1,,τn0,τj=1\tau_{1},\dotsc,\tau_{n}\geq 0,\ \sum\tau_{j}=1, will have boundary in Δεeiθ\Delta_{\varepsilon e^{i\theta}}:

(6.4) eiθτ1(1+εeiθq11,,εeiθq1n)++eiθτp(εeiθqp1,,1+εeiθqpp,,εeiθqpn)+τp+1(εeiθqp+1,1,,1+εeiθqp+1,p+1,,εeiθqp+1,n)++τn(εeiθqn1,,εeiθqnn+1).e^{i\theta}\tau_{1}(1+\varepsilon e^{i\theta}q_{11},\dotsc,\varepsilon e^{i\theta}q_{1n})+\cdots\\ +e^{i\theta}\tau_{p}(\varepsilon e^{i\theta}q_{p1},\dotsc,1+\varepsilon e^{i\theta}q_{pp},\dotsc,\varepsilon e^{i\theta}q_{pn})\\ +\tau_{p+1}(\varepsilon e^{i\theta}q_{p+1,1},\dotsc,1+\varepsilon e^{i\theta}q_{p+1,p+1},\dotsc,\varepsilon e^{i\theta}q_{p+1,n})+\cdots\\ +\tau_{n}(\varepsilon e^{i\theta}q_{n1},\dotsc,\varepsilon e^{i\theta}q_{nn}+1).

Consider for example τVε,ε,θ\partial\tau_{V}^{\varepsilon,\varepsilon,\theta} where VV is the orbit closure corresponding to the blowup of A1==Ap=0A_{1}=\cdots=A_{p}=0. Take in (6.4) τ1,,τpε\tau_{1},\dotsc,\tau_{p}\leq\varepsilon so terms in τjε\tau_{j}\varepsilon may be neglected for jpj\leq p. Take uj:=τj/τku_{j}:=\tau_{j}/\tau_{k} where k>pk>p is chosen so that (say) τk1/n\tau_{k}\geq 1/n. As a consequence, u1,,upnεu_{1},\dotsc,u_{p}\leq n\varepsilon. The corresponding projective point can then be written

(6.5) [eiθ(u1+ε)+O(ε2),,eiθ(up+ε)+O(ε2),up+1+eiθε+O(ε2),,un+eiθε+O(ε2)].\Big{[}e^{i\theta}(u_{1}+\varepsilon)+O(\varepsilon^{2}),\dotsc,e^{i\theta}(u_{p}+\varepsilon)+O(\varepsilon^{2}),\\ u_{p+1}+e^{i\theta}\varepsilon+O(\varepsilon^{2}),\dotsc,u_{n}+e^{i\theta}\varepsilon+O(\varepsilon^{2})\Big{]}.

The boundary is given by setting one or more of the uj=0u_{j}=0. Points in τVε,ε,θ\partial\tau_{V}^{\varepsilon,\varepsilon,\theta} can be approximated by points (6.5) which then deform into Δεeiθ\Delta_{\varepsilon e^{i\theta}}. To see this, note that since VV is a codimension 11 orbit closure, there will locally be one coordinate on P(Γ)P(\Gamma) near VV which takes the constant value εeiθ\varepsilon e^{i\theta} on τVε,ε,θ\partial\tau_{V}^{\varepsilon,\varepsilon,\theta} (cf. fig.(10)). On the other hand, (6.5) is in homogeneous coordinates on (Γ){\mathbb{P}}(\Gamma). To transform to P(Γ)P(\Gamma) near a general point of VV, one fixes p\ell\leq p and looks at ratios

(6.6) eiθ(uj+ε)+O(ε2)eiθ(u+ε)+O(ε2)\frac{e^{i\theta}(u_{j}+\varepsilon)+O(\varepsilon^{2})}{e^{i\theta}(u_{\ell}+\varepsilon)+O(\varepsilon^{2})}

for 1jp1\leq j\neq\ell\leq p. Clearly, at the boundary u=0u_{\ell}=0 we will get p1p-1 coordinates uj/ε+O(ε)u_{j}/\varepsilon+O(\varepsilon) which are close to 0{\mathbb{R}}^{\geq 0}, and one coordinate (corresponding to the local defining equation for VV) of the form εeiθ+O(ε2)\varepsilon e^{i\theta}+O(\varepsilon^{2}). The remaining coordinates on VV are ratios of the uj+εeiθ+O(ε2),jp+1u_{j}+\varepsilon e^{i\theta}+O(\varepsilon^{2}),\ j\geq p+1. Since uk=1u_{k}=1, these ratios are again close to 0{\mathbb{R}}^{\geq 0}. The calculation for orbit closures VV of codimension 2\geq 2 in P(Γ)P(\Gamma) is similar and is left for the reader. We have proven

Proposition 6.3.

With notation as above, the monodromy of the chain σεHn1(n1XΓ,ΔεXΓΔε)\sigma_{\varepsilon}\in H_{n-1}({\mathbb{P}}^{n-1}-X_{\Gamma},\Delta_{\varepsilon}-X_{\Gamma}\cap\Delta_{\varepsilon}) is represented by the chains c~η,ε,θ\tilde{c}^{\eta,\varepsilon,\theta} given by modifying the chains cη,ε,θc^{\eta,\varepsilon,\theta} to have boundary in Δεeiθ\Delta_{\varepsilon e^{i\theta}}. In particular, the monodromy m(σε)=c~η,ε,2πm(\sigma_{\varepsilon})=\tilde{c}^{\eta,\varepsilon,2\pi} is given by

(6.7) m(σε)=V(1)|V|τ~Vε,εm(\sigma_{\varepsilon})=\sum_{V}(-1)^{|V|}\tilde{\tau}_{V}^{\varepsilon,\varepsilon}

where τ~Vε,ε\tilde{\tau}_{V}^{\varepsilon,\varepsilon} is the chain τVε,ε\tau_{V}^{\varepsilon,\varepsilon} defined in section 4 with boundary extended to Δε\Delta_{\varepsilon} as above.

It will be convenient to simplify the notation and write

(6.8) τVε:=τ~Vε,ε.\tau_{V}^{\varepsilon}:=\tilde{\tau}_{V}^{\varepsilon,\varepsilon}.

7. Parametric representations

In this section we list well-known representations of the Feynman rules and then prepare for a subsequent analysis of short-distance singularities in terms of mixed Hodge structures.

7.1. Kirchhoff–Symanzik polynomials

Let

(7.1) ψ(Γ)\displaystyle\psi(\Gamma) =\displaystyle= TeTAe,\displaystyle\sum_{T}\prod_{e\not\in T}A_{e},
(7.2) ϕ(Γ)\displaystyle\phi(\Gamma) =\displaystyle= T1T2=TQ(T1)Q(T2)eT1T2Ae,\displaystyle\sum_{T_{1}\cup T_{2}=T}Q(T_{1})\cdot Q(T_{2})\prod_{e\not\in T_{1}\cup T_{2}}A_{e},

be the two homogenous Kirchhoff–Symanzik polynomials [10, 22]. Here, TT is a spanning tree of the 1PI graph Γ\Gamma and T1,T2T_{1},T_{2} are disjoint trees which together cover all vertices of Γ\Gamma. Also, Q(Ti)Q(T_{i}) is the sum of all external momenta attached to vertices covered by TiT_{i}. Note that ϕ(Γ)\phi(\Gamma) can be written as

(7.3) kinetic invariants (qiqj)Rqiqj.\sum_{\textrm{kinetic invariants $(q_{i}\cdot q_{j})$}}R_{q_{i}\cdot q_{j}}.

Here, qiq_{i} are external momenta attached to T1T_{1} and qjq_{j} to T2T_{2}, and RqiqjR_{q_{i}\cdot q_{j}} are rational functions of the edge variables only, and the sum is over independent such kinematical invariants where momentum conservation has been taken into account. We extend the definition to the empty graph 𝕀\mathbb{I} by ψ(𝕀)=1\psi(\mathbb{I})=1, ϕ(𝕀)=0\phi(\mathbb{I})=0.

Let ||γ|\cdot|_{\gamma} denote the degree of a polynomial with regard to variables of the graph γ\gamma.

Lemma 7.1.

i) degϕ=degψ+1\deg\phi=\deg\psi+1.
ii)

(7.4) ψ(Γ)=ψ(Γ//γ)ψ(γ)+ψΓ,γ\psi(\Gamma)=\psi(\Gamma/\!/\gamma)\psi(\gamma)+\psi_{\Gamma,\gamma}

with |ψΓ,γ|γ>|ψ(γ)|γ|\psi_{\Gamma,\gamma}|_{\gamma}>|\psi(\gamma)|_{\gamma} for all core graphs Γ\Gamma and subgraphs γ\gamma.
iii)

(7.5) ϕ(Γ)=ϕ(Γ//γ)ψ(γ)+ϕΓ,γ\phi(\Gamma)=\phi(\Gamma/\!/\gamma)\psi(\gamma)+\phi_{\Gamma,\gamma}

with |ϕΓ,γ|γ>|ψ(γ)|γ|\phi_{\Gamma,\gamma}|_{\gamma}>|\psi(\gamma)|_{\gamma} for all core graphs Γ\Gamma and subgraphs γ\gamma.

Proof: i) by definition, ii) has been proved in [2], iii) follows similarly by noting that the two-trees of ϕ\phi are obtained from the spanning trees of ψ\psi by removing an edge. If that edge belongs to Γ//γ\Gamma/\!/\gamma, we get ϕ(Γ//γ)ψ(γ)\phi(\Gamma/\!/\gamma)\psi(\gamma). If it belongs to γ\gamma, we get a monomial mm with |m|γ>|ψ(γ)|γ|m|_{\gamma}>|\psi(\gamma)|_{\gamma}. \Box
Note that it might happen that ϕ(Γ//γ)=0\phi(\Gamma/\!/\gamma)=0, if the external momenta flows through subgraphs γ\gamma only. In such a case (which can lead to infrared divergences) one easily shows ϕΓ,γ=ψ(Γ//γ)ϕ(γ)\phi_{\Gamma,\gamma}=\psi(\Gamma/\!/\gamma)\phi(\gamma).

7.2. Feynman rules

¿From these polynomials one constructs the Feynman rules of a given theory. For example we have in ϕ4\phi^{4} theory for a vertex graph Γ\Gamma, sdd(Γ)=0\textrm{sdd}(\Gamma)=0,

(7.6) Φ(Γ)=>ϵkeedges eAeme2ϕ(Γ)ψ(Γ)ψ2(Γ)𝑑A1𝑑A|Γ[1]|.\Phi(\Gamma)=\int_{\mathbb{R}_{>\epsilon}^{k}}\frac{e^{-\sum_{\textrm{edges $e$}}A_{e}m_{e}^{2}-\frac{\phi(\Gamma)}{\psi(\Gamma)}}}{\psi^{2}(\Gamma)}d\!A_{1}\cdots d\!A_{|\Gamma^{[1]}|}.

We will write >ϵ𝑑AΓ\int_{>\epsilon}d\!A_{\Gamma} to abbreviate the affine chain of integration.

The integral is over the kk-dimensional hypercube of positive real coordinates in >ϵ\mathbb{R}_{>\epsilon} with a small strip of width 1ϵ>01\gg\epsilon>0 removed at each axis. We regard the integrand

(7.7) ι(Γ):=eeAeme2ϕ(Γ)ψ(Γ)ψ2(Γ),\mathbb{C}\ni\iota(\Gamma):=\frac{e^{-\sum_{e}A_{e}m_{e}^{2}-\frac{\phi(\Gamma)}{\psi(\Gamma)}}}{\psi^{2}(\Gamma)},

ι(Γ)=ι(Γ)({m2},{qiqj},{A})\iota(\Gamma)=\iota\left(\Gamma)(\{m^{2}\},\{q_{i}\cdot q_{j}\},\{A\}\right) as a function of the set of internal masses {m2}\{m^{2}\}, the set of external momenta {qiqj}\{q_{i}\cdot q_{j}\} (which can be considered as labels on external half-edges) and the set of graph coordinates {A}\{A\}, and ι\iota takes values in \mathbb{C}. We often omit the AA dependence and abbreviate P={m},{qiqj}P=\{m\},\{q_{i}\cdot q_{j}\} for all these external parameters of the integrand: ι=ι(P)\iota=\iota(P). The renormalization schemes we consider are determined by the condition that the Green function shall vanish at a particular renormalization point RR, so that renormalization becomes an iterated sequence of subtractions

(7.8) ι(P,R):=ι(P)ι(R).\iota_{-}(P,R):=\iota(P)-\iota(R).

We let sdd(Γ){\textrm{sdd}}(\Gamma) be the superficial degree of divergence of a graph Γ\Gamma given as (see also Eq.(2.2) for a refined version)

(7.9) sdd(Γ):=D|Γ|edges ewevertices vwv,{\textrm{sdd}}(\Gamma):=D|\Gamma|-\sum_{\textrm{edges $e$}}w_{e}-\sum_{\textrm{vertices $v$}}w_{v},

where |Γ||\Gamma| is the rank of the first Betti homology, DD the dimension of spacetime which we keep as an integer, wew_{e} the weights of the propagator for edge ee as prescribed by free field theory and wvw_{v} the weight of the vertex as prescribed by the interaction Lagrangian. Note that we can set the width ϵ\epsilon to zero, >ϵ𝑑AΓ>0𝑑AΓ\int_{>\epsilon}d\!A_{\Gamma}\to\int_{>0}d\!A_{\Gamma} if the integrand ι(Γ)\iota_{-}(\Gamma) is evaluated on a graph Γ\Gamma which has no divergent subgraphs.

Throughout, we assume that all all masses and external momenta are in general position so that there are no zeroes in the ϕ\phi-polynomial off the origin for positive values of the AA variables. In particular, we assume that the point PP is chosen appropriately away from all mass-shell and kinematical singularities. We remind the reader of the notation (Γ,σ)(\Gamma,\sigma) (section (2.5)) where σ\sigma stores all the necessary detail on how to evaluate the graph Γ\Gamma.

A special role is played by the evaluations (Γ,σP=0)(\Gamma,\sigma_{P=0}). They set all internal masses and momenta to zero. Note that this leads immediately to infrared divergences: the Feynman integrands ι()(P=0)\iota(\cdot)(P=0) are missing the exponential in the numerator, which provides a regulator at large values of the AA variables, and hence an infrared regulator. The ultraviolet singularities at small values of the AA variables are taken into account by the renormalization procedure itself, and hence by our limiting mixed Hodge structure. We will eliminate the case P=0P=0 below using that ι\iota_{-} evaluates to zero if there is no dependence on masses or external momenta.

7.3. General remarks on renormalization and QFT

We now consider the renormalization Hopf algebra HH_{\mathcal{R}} of 1PI Feynman graphs in section (2.4). We use the notation

(7.10) Δ(Γ)=γγΓ//γ,\Delta(\Gamma)=\sum_{\gamma}\gamma\otimes\Gamma/\!/\gamma,

for its coproduct. Also, Δ(𝕀)=𝕀𝕀\Delta(\mathbb{I})=\mathbb{I}\otimes\mathbb{I}. Projection PP into the augmentation ideal on the rhs is written as

(7.11) (idP)Δ(Γ)=Γ//γγΓ//γ,({\textrm{id}}\otimes P)\Delta(\Gamma)=\sum_{\emptyset\not=\Gamma/\!/\gamma}\gamma\otimes\Gamma/\!/\gamma,

so that for example the antipode SS is

(7.12) S(Γ)=Γ//γS(γ)Γ//γ=:Γ¯.S(\Gamma)=-\sum_{\emptyset\not=\Gamma/\!/\gamma}S(\gamma)\Gamma/\!/\gamma=:-\bar{\Gamma}.

Furthermore, we introduce a forest notation for the antipode:

(7.13) S(Γ)=[for](1)|[for]|Γ//[for]j=1|[for]|γ[for],j,S(\Gamma)=\sum_{\textrm{[for]}}(-1)^{|\textrm{[for]}|}\Gamma/\!/\textrm{[for]}\prod_{j=1}^{|\textrm{[for]}|}\gamma_{\textrm{[for]},j},

where the sum is over all forests [for] and the product is over all subgraphs which make up the forest. Here, a forest [for] is a possibly empty collection of proper superficially divergent 1PI subgraphs γ[for],j\gamma_{\textrm{[for]},j} of Γ\Gamma which are mutually disjoint or nested. We call a forest [for] maximal if Γ//[for]\Gamma/\!/\textrm{[for]} is a primitive element of the Hopf algebra. As edge sets

(7.14) Γ=(Γ//[for])(jγj).\Gamma=\left(\Gamma/\!/\mathrm{[for]}\right)\cup\left(\cup_{j}\gamma_{j}\right).

This is in one-to-one correspondence with the representation of the antipode as a sum over all cuts on rooted trees ρ𝒯(Γ)\rho_{\mathcal{T}}(\Gamma) as detailed in section (2.3) above. The integer |[for]||\textrm{[for]}| is the number of edges removed in this representation.

Let us first assume that the graph Γ\Gamma and all its core subgraphs have a non-positive superficial degree of divergence, so they are convergent or provide log-pole: sdd0{\textrm{sdd}}\leq 0 for all elements in (the complement of) the forests.

As the integrand ι(Γ)(P)\iota(\Gamma)(P) depends on P={m},{qiqj}P=\{m\},\{q_{i}\cdot q_{j}\} only through the argument of the exponential, we redefine the second Kirchhoff–Symanzik polynomial as follows:

(7.15) ϕ(Γ)({qiqj})φ(Γ)(P):=ϕ(Γ)({qiqj})+ψ(Γ)eAeme2.\phi(\Gamma)(\{q_{i}\cdot q_{j}\})\to\varphi(\Gamma)(P):=\phi(\Gamma)(\{q_{i}\cdot q_{j}\})+\psi(\Gamma)\sum_{e}A_{e}m_{e}^{2}.

Then, the unrenormalized integrand is

(7.16) ι(Γ)(P)=expφ(Γ)(P)ψ(Γ)ψ2(Γ).\iota(\Gamma)(P)=\frac{\exp^{-\frac{\varphi(\Gamma)(P)}{\psi(\Gamma)}}}{\psi^{2}(\Gamma)}.

With this notation, the renormalized integrand is (in all sums and products over jj here and in the following, jj runs from 11 to |[for]||\textrm{[for]}|)

ιR(Γ)(P,R)\displaystyle\iota_{R}(\Gamma)(P,R) =\displaystyle= [for](1)[for]exp(φ(Γ//[for])(P)ψ(Γ//[for])+jφ(γj)(R)ψ(γj))ψ2(γ//[for])jψ2(γj)\displaystyle\sum_{\textrm{[for]}}(-1)^{\textrm{[for]}}\frac{\exp{-\left(\frac{\varphi(\Gamma/\!/\textrm{[for]})(P)}{\psi(\Gamma/\!/\textrm{[for]})}+\sum_{j}\frac{\varphi(\gamma_{j})(R)}{\psi(\gamma_{j})}\right)}}{\psi^{2}(\gamma/\!/\textrm{[for]})\prod_{j}\psi^{2}(\gamma_{j})}
[for](1)[for]exp(φ(Γ//[for])(R)ψ(Γ//[for])+jφ(γj)(R)ψ(γj))ψ2(γ//[for])jψ2(γj)\displaystyle-\sum_{\textrm{[for]}}(-1)^{\textrm{[for]}}\frac{\exp{-\left(\frac{\varphi(\Gamma/\!/\textrm{[for]})(R)}{\psi(\Gamma/\!/\textrm{[for]})}+\sum_{j}\frac{\varphi(\gamma_{j})(R)}{\psi(\gamma_{j})}\right)}}{\psi^{2}(\gamma/\!/\textrm{[for]})\prod_{j}\psi^{2}(\gamma_{j})}
=:\displaystyle=: ι¯(Γ)(P,R)+Sι(Γ)(R),\displaystyle\bar{\iota}(\Gamma)(P,R)+S^{\iota}(\Gamma)(R),

where +Sι(Γ)(R)=ι¯(Γ)(R,R)+S^{\iota}(\Gamma)(R)=-\bar{\iota}(\Gamma)(R,R) is the integrand for the counterterm, and ι¯(Γ)(P,R)\bar{\iota}(\Gamma)(P,R), the integrand in the first line, delivers upon integrating Bogoliubov’s R¯\bar{R} operation. Note that this formula (7.3) is just the evaluation

(7.18) m(SRιι)Δ(Γ),m(S_{R}^{\iota}\otimes\iota)\Delta(\Gamma),

which guarantees that the corresponding Feynman integral exists in the limit ϵ0\epsilon\to 0 [13],[15].

This Feynman integral is obtained by integrating from ϵ\epsilon to \infty each edge variable. For the renormalized Feynman integral ΦR(Γ)(P)\Phi_{R}(\Gamma)(P) we can take the limit ϵ0\epsilon\to 0, while for the R¯\bar{R}-operation

(7.19) Φ¯(Γ)(P,R;ϵ)=ϵι¯(Γ)(P,R),\bar{\Phi}(\Gamma)(P,R;\epsilon)=\int_{\epsilon}\bar{\iota}(\Gamma)(P,R),

and the counterterm

(7.20) SR;ϵΦ(Γ)=Φ¯(Γ)(R,R;ϵ),S_{R;\epsilon}^{\Phi}(\Gamma)=-\bar{\Phi}(\Gamma)(R,R;\epsilon),

the lower boundary remains as a dimension-full parameter in the integral. Note that the result (7.3) above can also be written in the PRP-R form, typical for renormalization schemes which subtract by constraints on physical parameters:

(7.21) ιR(Γ)(P,R)=Γ//γ[ι(Γ//γ)(P)ι(Γ//γ)(R)]SR;ϵι(γ),\iota_{R}(\Gamma)(P,R)=\sum_{\emptyset\not=\Gamma/\!/\gamma}\left[\iota(\Gamma/\!/\gamma)(P)-\iota(\Gamma/\!/\gamma)(R)\right]S_{R;\epsilon}^{\iota}(\gamma),

and as

(7.22) ι¯(Γ)(P,R)=Γ//γSR;ϵι(γ)ι(Γ//γ)(P)ιR(Γ)(P,R)=γSR;ϵι(γ)ι(Γ//γ)(P),\bar{\iota}(\Gamma)(P,R)=\sum_{\emptyset\not=\Gamma/\!/\gamma}S_{R;\epsilon}^{\iota}(\gamma)\iota(\Gamma/\!/\gamma)(P)\Rightarrow\iota_{R}(\Gamma)(P,R)=\sum_{\gamma}S_{R;\epsilon}^{\iota}(\gamma)\iota(\Gamma/\!/\gamma)(P),

using the notation (7.11,7.10). Similarly, for Feynman integrals,

(7.23) Φ¯(Γ)(P,R;ϵ)=Γ//γSR;ϵΦ(γ)Φ(Γ//γ)(P),ΦR(Γ)(P)=limϵ0γSR;ϵΦ(γ)Φ(Γ//γ)(P).\bar{\Phi}(\Gamma)(P,R;\epsilon)=\sum_{\emptyset\not=\Gamma/\!/\gamma}S_{R;\epsilon}^{\Phi}(\gamma)\Phi(\Gamma/\!/\gamma)(P),\;\Phi_{R}(\Gamma)(P)=\lim_{\epsilon\to 0}\sum_{\gamma}S_{R;\epsilon}^{\Phi}(\gamma)\Phi(\Gamma/\!/\gamma)(P).

When it comes to actually calculating the integral (7.6) (or, in its renormalized form (7.3)), something rather remarkable happens. By lemma 7.1(i), the term in the exponential in these integrals is homogeneous of degree 11 in the edge variables AiA_{i}. The assumption sdd(Γ)=0\text{sdd}(\Gamma)=0 means dA/ψ2dA/\psi^{2} is homogeneous of degree 0. Making the change of variable Ai=taiA_{i}=ta_{i}, we find

(7.24) dA/ψ(A)2=dt/t((1)j1ajda1daj^)/ψ(a)2=dt/tΩ/ψ2.dA/\psi(A)^{2}=dt/t\wedge(\sum(-1)^{j-1}a_{j}da_{1}\wedge\cdots\wedge\widehat{da_{j}}\wedge\cdots)/\psi(a)^{2}=dt/t\wedge\Omega/\psi^{2}.

Note that Ω/ψ2\Omega/\psi^{2} is naturally a meromorphic form on the projective space (Γ){\mathbb{P}}(\Gamma) with homogeneous coordinates the aia_{i}. Writing σ={ai0}(Γ)()\sigma=\{a_{i}\geq 0\}\subset{\mathbb{P}}(\Gamma)({\mathbb{R}}), we see that the renormalized integral can be rewritten up to a term which is O(ε)O(\varepsilon) as a sum of terms of the form

(7.25) σΩ/ψj2ε(e(tfj(a))e(tgj(a)))𝑑t/t=σΩ/ψj2(E1(εfj(a))E1(εgj(a))),\int_{\sigma}\Omega/\psi_{j}^{2}\int_{\varepsilon}^{\infty}\Big{(}e^{(-tf_{j}(a))}-e^{(-tg_{j}(a))}\Big{)}dt/t=\int_{\sigma}\Omega/\psi_{j}^{2}\Big{(}E_{1}(\varepsilon f_{j}(a))-E_{1}(\varepsilon g_{j}(a))\Big{)},

where

(7.26) E1(z):=1etzdtt=γElnz+O(z);z0E_{1}(z):=\int_{1}^{\infty}e^{-tz}\frac{d\!t}{t}=-\gamma_{E}-\ln z+O(z);\quad z\to 0

is the exponential integral. (Here fj(a),gj(a)f_{j}(a),g_{j}(a) are defined by taking the locus ai0,ai=1a_{i}\geq 0,\sum a_{i}=1.) As long as fj(a),gj(a)>0f_{j}(a),g_{j}(a)>0, we may allow ε0\varepsilon\to 0 for fixed aa. The Euler constant and logε\log\varepsilon terms cancel. When the dust settles, we are left with the projective representation for the renormalized Feynman integral

(7.27) ΦR(Γ)(P)=σΩΓ[for](1)[for]ln(φ(Γ//[for])(P)jψ(γj)+jφ(γj)(R)ψ(Γ//[for])hjψ(γh)φ(Γ//[for])(R)jψ(γj)+jφ(γj)(R)ψ(Γ//[for])hjψ(γh))ψ2(Γ//[for])jψ2(γj).\Phi_{R}(\Gamma)(P)=\int_{\sigma}\Omega_{\Gamma}\sum_{\textrm{[for]}}(-1)^{\textrm{[for]}}\frac{\ln{\left(\frac{\varphi(\Gamma/\!/\textrm{[for]})(P)\prod_{j}\psi(\gamma_{j})+\sum_{j}\varphi(\gamma_{j})(R)\psi(\Gamma/\!/\textrm{[for]})\prod_{h\not=j}\psi(\gamma_{h})}{\varphi(\Gamma/\!/\textrm{[for]})(R)\prod_{j}\psi(\gamma_{j})+\sum_{j}\varphi(\gamma_{j})(R)\psi(\Gamma/\!/\textrm{[for]})\prod_{h\not=j}\psi(\gamma_{h})}\right)}}{\psi^{2}(\Gamma/\!/\textrm{[for]})\prod_{j}\psi^{2}(\gamma_{j})}.

Note that the use of σ\sigma is justified as long as the integrand has all subdivergences subtracted, so is in the ι¯\bar{\iota} form, so that lower boundaries in the aia_{i} variables can be set to zero indeed.

By (7.21), this can be equivalently written as

(7.28) ΦR(Γ)(P)=limϵ0γSR;ϵΦ(γ)>ϵdAΓ//γι(Γ//γ)(P,R),\Phi_{R}(\Gamma)(P)=\lim_{\epsilon\to 0}\sum_{\gamma}S_{R;\epsilon}^{\Phi}(\gamma)\int_{>\epsilon}d\!A_{\Gamma/\!/\gamma}\iota_{-}(\Gamma/\!/\gamma)(P,R),

in any renormalization scheme which is described by kinematical subtractions PRP\to R.

Remark 7.2.

It will be our goal to replace the affine 𝑑A\int d\!A by the projective 𝑑Ω\int d\!\Omega in the above. The presence of lower boundaries, which can not be ignored as the integrand has divergent subgraphs, allows this only upon introducing suitable chains τγϵ\tau_{\gamma}^{\epsilon} as discussed in previous sections.

Next, we relax the case of log-divergence.

7.4. Reduction of graphs with ssd(Γ)>0{\rm{ssd}}(\Gamma)>0

We start with an example. To keep things simple but not too simple, we consider the one-loop self-energy graph in ϕ63\phi^{3}_{6} theory, a scalar field theory with a cubic interaction in six dimensions of space-time. We have

(7.29) Φ(Γ)(P)=>ϵ𝑑AΓι(Γ)(P)=>ϵ𝑑AΓeφ(Γ)ψ(Γ)ψ(Γ)3ϵ𝑑A1𝑑A2em2(A1+A2)2+q2A1A2(A1+A2)(A1+A2)3.\Phi(\Gamma)(P)=\int_{>\epsilon}d\!A_{\Gamma}\iota(\Gamma)(P)=\int_{>\epsilon}d\!A_{\Gamma}\frac{e^{-\frac{\varphi(\Gamma)}{\psi(\Gamma)}}}{\psi(\Gamma)^{3}}\equiv\int_{\epsilon}^{\infty}d\!A_{1}d\!A_{2}\frac{e^{-\frac{m^{2}(A_{1}+A_{2})^{2}+q^{2}A_{1}A_{2}}{(A_{1}+A_{2})}}}{(A_{1}+A_{2})^{3}}.

We will renormalize by suitable subtractions at chosen values of masses and momenta in the φ\varphi-polynomial. We hence (with subdivergences taken care of by suitable bar-operations ιι¯\iota\to\bar{\iota} in the general case) replace ι(Γ)(P)\iota(\Gamma)(P) by ι(Γ)(P)ι(Γ)(0)\iota(\Gamma)(P)-\iota(\Gamma)(0), as this leaves ι(Γ)(P,R)\iota_{-}(\Gamma)(P,R) invariant.

Then the above can be written, with this subtraction, and by the familiar change of variables Ai=taiA_{i}=ta_{i}, and by one partial integration in tt,

(7.30) Φ(Γ)(P)\displaystyle\Phi(\Gamma)(P) =\displaystyle= σ𝑑Ωϵdtt[m2(a1+a2)2+q2a1a2]etm2(a1+a2)2+q2a1a2(a1+a2)(a1+a2)4\displaystyle\int_{\sigma}d\!\Omega\int_{\epsilon}^{\infty}\frac{d\!t}{t}\frac{[m^{2}(a_{1}+a_{2})^{2}+q^{2}a_{1}a_{2}]e^{-t\frac{m^{2}(a_{1}+a_{2})^{2}+q^{2}a_{1}a_{2}}{(a_{1}+a_{2})}}}{(a_{1}+a_{2})^{4}}
σ𝑑Ω[m2(a1+a2)2+q2a1a2](a1+a2)4,\displaystyle-\int_{\sigma}d\!\Omega\frac{[m^{2}(a_{1}+a_{2})^{2}+q^{2}a_{1}a_{2}]}{(a_{1}+a_{2})^{4}},

where we expanded the boundary term up to terms constant in ϵ\epsilon, which gave the term in the second line. We discarded already the pure pole term 1/ϵ\sim 1/\epsilon from Φ(Γ)(P=0)=>ϵdA(A1+A2)3\Phi(\Gamma)(P=0)=\int_{>\epsilon}\frac{d\!A}{(A_{1}+A_{2})^{3}} == ϵ𝑑t/t20𝑑b21/(1+b2)3\int_{\epsilon}^{\infty}d\!t/t^{2}\int_{0}^{\infty}db_{2}1/(1+b_{2})^{3}.

Note that graphs Γ\Gamma with sdd>0\textrm{sdd}>0 have res(Γ)=2\textrm{res}(\Gamma)=2. They hence depend on a single kinematical invariant q2q^{2} say, ϕ(Γ)=ϕ(Γ)(q2)\phi(\Gamma)=\phi(\Gamma)(q^{2}), for which we write ϕ(Γ)q2\phi(\Gamma)_{q^{2}}.

The result in (7.30) leads us to define two top-degree forms. (Here Ω=a1da2a2da1\Omega=a_{1}da_{2}-a_{2}da_{1} and we still write ϕ,ψ\phi,\psi for the Kirchhoff–Symanzik polynomials regarded as dependent on either aia_{i} or AiA_{i} variables below).

(7.31) ω=ω(Γ)=Ωϕ1(Γ)ψ(Γ)4=Ωa1a2(a1+a2)4,\omega_{\Box}=\omega_{\Box}(\Gamma)=\Omega\frac{\phi_{1}(\Gamma)}{\psi(\Gamma)^{4}}=\Omega\frac{a_{1}a_{2}}{(a_{1}+a_{2})^{4}},

and

(7.32) ωm2=ωm2(Γ)=Ω(a1+a2)2ψ(Γ)4=Ω1(a1+a2)2,\omega_{m^{2}}=\omega_{m^{2}}(\Gamma)=\Omega\frac{(a_{1}+a_{2})^{2}}{\psi(\Gamma)^{4}}=\Omega\frac{1}{(a_{1}+a_{2})^{2}},

so that

Φ(Γ)(P)\displaystyle\Phi(\Gamma)(P) =\displaystyle= m2σ[ω+ωm2](q2m2)σω\displaystyle-m^{2}\int_{\sigma}[\omega_{\Box}+\omega_{m^{2}}]-(q^{2}-m^{2})\int_{\sigma}\omega_{\Box}
+m2σ[ω+ωm2]ϵdttetφ(Γ)(P)ψ(Γ)\displaystyle+m^{2}\int_{\sigma}[\omega_{\Box}+\omega_{m^{2}}]\int_{\epsilon}^{\infty}\frac{d\!t}{t}e^{-t\frac{\varphi(\Gamma)(P)}{\psi(\Gamma)}}
+(q2m2)σωϵdttetφ(Γ)(P)ψ(Γ).\displaystyle+(q^{2}-m^{2})\int_{\sigma}\omega_{\Box}\int_{\epsilon}^{\infty}\frac{d\!t}{t}e^{-t\frac{\varphi(\Gamma)(P)}{\psi(\Gamma)}}.

There are corresponding affine integrands

(7.34) ι(Γ)\displaystyle\iota_{\Box}(\Gamma) =\displaystyle= ϕ1(Γ)ψ(Γ)4eφ(Γ)(P)ψ(Γ),\displaystyle\frac{\phi_{1}(\Gamma)}{\psi(\Gamma)^{4}}e^{-\frac{\varphi(\Gamma)(P)}{\psi(\Gamma)}},
(7.35) ιm2(Γ)\displaystyle\iota_{m^{2}}(\Gamma) =\displaystyle= (a1+a2)2ψ(Γ)4eφ(Γ)(P)ψ(Γ).\displaystyle\frac{(a_{1}+a_{2})^{2}}{\psi(\Gamma)^{4}}e^{-\frac{\varphi(\Gamma)(P)}{\psi(\Gamma)}}.

The graph Γ\Gamma is renormalized by a choice of a renormalization condition RR_{\Box} for the coefficient of q2m2q^{2}-m^{2} (wave function renormalization), and by the choice of a condition Rm2R_{m^{2}} for the mass renormalization. RR is often still used to denote the pair of those.

(7.36) Φ(Γ)(P)+m2δm2+q2z=ΦR,Rm2(Γ)(P).\Phi(\Gamma)(P)+m^{2}\delta_{m^{2}}+q^{2}z_{\Box}=\Phi_{R_{\Box},R_{m^{2}}}(\Gamma)(P).

The mass counterterm is then

(7.37) m2δm2=m2σ[ω+ωm2](1ϵdttetφ(Γ)(Rm2)ψ(Γ)),m^{2}\delta_{m^{2}}=-m^{2}\int_{\sigma}[\omega_{\Box}+\omega_{m^{2}}]\left(1-\int_{\epsilon}^{\infty}\frac{d\!t}{t}e^{-t\frac{\varphi(\Gamma)(R_{m^{2}})}{\psi(\Gamma)}}\right),

and the wave-function renormalization q2zq^{2}z_{\Box} is

(7.38) q2z=q2σω(1ϵdttetφ(Γ)(R)ψ(Γ)).q^{2}z_{\Box}=-q^{2}\int_{\sigma}\omega_{\Box}\left(1-\int_{\epsilon}^{\infty}\frac{d\!t}{t}e^{-t\frac{\varphi(\Gamma)(R_{\Box})}{\psi(\Gamma)}}\right).

Note the term 1 in the ()() brackets does not involve exponentials.

The corresponding renormalized contribution is

(7.39) ΦR(Γ)(P)=(q2m2)σωlnφ(P)φ(R)+m2σ[ω+ωm2]lnφ(Γ)(P)φ(Γ)(Rm2).\Phi_{R}(\Gamma)(P)=(q^{2}-m^{2})\int_{\sigma}\omega_{\Box}\ln{\frac{\varphi(P)}{\varphi(R_{\Box})}}+m^{2}\int_{\sigma}[\omega_{\Box}+\omega_{m^{2}}]\ln{\frac{\varphi(\Gamma)(P)}{\varphi(\Gamma)(R_{m^{2}})}}.

The transition from the unrenormalized contribution to the renormalized one is particularly simple upon defining Feynman rules in accordance with external leg structures:

(7.40) Φ((Γ,σ))\displaystyle\Phi((\Gamma,\sigma_{\Box})) =\displaystyle= (q2m2)>ϵ𝑑Aϕ1(Γ)eφ(Γ)(P)ψ(Γ)ψ(Γ)D/2+1,\displaystyle(q^{2}-m^{2})\int_{>\epsilon}d\!A\frac{\phi_{1}(\Gamma)e^{-\frac{\varphi(\Gamma)(P)}{\psi(\Gamma)}}}{\psi(\Gamma)^{D/2+1}},
(7.41) Φ((Γ,σm2))\displaystyle\Phi((\Gamma,\sigma_{m^{2}})) =\displaystyle= m2>ϵ𝑑A[ϕ1(Γ)+ψ(Γ)eAe]eφ(Γ)(P)ψ(Γ)ψ(Γ)D/2+1,\displaystyle m^{2}\int_{>\epsilon}d\!A\frac{[\phi_{1}(\Gamma)+\psi(\Gamma)\sum_{e}A_{e}]e^{-\frac{\varphi(\Gamma)(P)}{\psi(\Gamma)}}}{\psi(\Gamma)^{D/2+1}},

so that renormalization proceeds as before on log-divergent integrands.

This example extends straightforwardly to the case of Γ\Gamma having divergent subgraphs. Let us return to ϕ44\phi_{4}^{4} theory and define for a core graph Γ\Gamma with sdd(Γ)=2\textrm{sdd}(\Gamma)=2, (so that it is a self-energy graph and hence has only two external legs, and thus a single kinematical invariant q2q^{2}), and graph-polynomials ψ(Γ)\psi(\Gamma), ϕ(Γ)=ϕq2(Γ)\phi(\Gamma)=\phi_{q^{2}}(\Gamma), φ(Γ)=φ(Γ)(P)=ϕq2(Γ)+ψ(Γ)eAeme2,\varphi(\Gamma)=\varphi(\Gamma)(P)=\phi_{q^{2}}(\Gamma)+\psi(\Gamma)\sum_{e}A_{e}m_{e}^{2}, the forms

(7.43) ω(Γ)=ΩΓϕ1(Γ)ψ3(Γ),\omega_{\Box}(\Gamma)=\Omega_{\Gamma}\frac{\phi_{1}(\Gamma)}{\psi^{3}(\Gamma)},
(7.44) ωm2(Γ)=ΩΓϕ1(Γ)+ψ(Γ)eAeψ3(Γ).\omega_{m^{2}}(\Gamma)=\Omega_{\Gamma}\frac{\phi_{1}(\Gamma)+\psi(\Gamma)\sum_{e}A_{e}}{\psi^{3}(\Gamma)}.

The corresponding complete affine integrands ι,ιm2\iota_{\Box},\iota_{m^{2}} are immediate replacing aia_{i} by AiA_{i} variables, and multiplying by exponentials expφ(Γ)(P)/ψ(Γ)\exp{-\varphi(\Gamma)(P)/\psi(\Gamma)}, with PRP\to R for counterterms.

One finds by a straightforward computation

(7.45) ΦR((Γ,σ))(P)\displaystyle\Phi_{R_{\Box}}((\Gamma,\sigma_{\Box}))(P) =\displaystyle= γSR;ϵΦ(γ)ϵω(Γ//γ)lnφ(Γ//γ)(P)φ(Γ//γ)(R)\displaystyle\sum_{\gamma}S^{\Phi}_{R;\epsilon}(\gamma)\int_{{\epsilon}}\omega_{\Box}(\Gamma/\!/\gamma)\ln{\frac{\varphi(\Gamma/\!/\gamma)(P)}{\varphi(\Gamma/\!/\gamma)(R_{\Box})}}
(7.46) =\displaystyle= ΩΓ[for](1)[for]ω(Γ//[for])lnφ(Γ//[for])(P)φ(Γ//[for])(R),\displaystyle\int\Omega_{\Gamma}\sum_{\textrm{[for]}}(-1)^{\textrm{[for]}}\omega_{\Box}(\Gamma/\!/\textrm{[for]})\ln{\frac{\varphi(\Gamma/\!/\textrm{[for]})(P)}{\varphi(\Gamma/\!/\textrm{[for]})(R_{\Box})}},

and

(7.47) ΦRm2((Γ,σm2))(P)\displaystyle\Phi_{R_{m^{2}}}((\Gamma,\sigma_{m^{2}}))(P) =\displaystyle= γSR;ϵΦ(γ)ϵωm2(Γ//γ)lnφ(Γ//γ)(P)φ(Γ//γ)(Rm2)\displaystyle\sum_{\gamma}S^{\Phi}_{R;\epsilon}(\gamma)\int_{{\epsilon}}\omega_{m^{2}}(\Gamma/\!/\gamma)\ln{\frac{\varphi(\Gamma/\!/\gamma)(P)}{\varphi(\Gamma/\!/\gamma)(R_{m^{2}})}}
(7.48) =\displaystyle= ΩΓ[for](1)[for]ωm2(Γ//[for])lnφ(Γ//[for])(P)φ(Γ//[for])(Rm2).\displaystyle\int\Omega_{\Gamma}\sum_{\textrm{[for]}}(-1)^{\textrm{[for]}}\omega_{m^{2}}(\Gamma/\!/\textrm{[for]})\ln{\frac{\varphi(\Gamma/\!/\textrm{[for]})(P)}{\varphi(\Gamma/\!/\textrm{[for]})(R_{m^{2}})}}.

We set

(7.49) ΦR(Γ)(P)ΦR((Γ,𝕀))(P)=ϕR((Γ,σ))(P)+ΦRm2((Γ,σm2))(P),\Phi_{R}(\Gamma)(P)\equiv\Phi_{R}((\Gamma,\mathbb{I}))(P)=\phi_{R_{\Box}}((\Gamma,\sigma_{\Box}))(P)+\Phi_{R_{m^{2}}}((\Gamma,\sigma_{m^{2}}))(P),

in the external leg structure notation of section (2.5). We can combine the results for graphs Γ\Gamma for all degrees of divergence sdd(Γ)0\textrm{sdd}(\Gamma)\geq 0 by defining ω(Γ)=Ω/ψ2(Γ)\omega(\Gamma)=\Omega/\psi^{2}(\Gamma) for a log divergent graph with the results above. And that’s that. Well, we have to hasten and say a word about the Feynman rules when the subgraphs γ\gamma have sdd(γ)>0\textrm{sdd}(\gamma)>0, and hence also about SR;ϵΦ(γ)S^{\Phi}_{R;\epsilon}(\gamma) in that case.

We use, with PP the projection into the augmentation ideal, the notation

(7.50) Γ¯=Γ+m(SPP)Δ=:Γ+(Γ)1Γ′′.\bar{\Gamma}=\Gamma+m(S\circ P\otimes P)\Delta=:\Gamma+(\Gamma^{\prime})^{-1}\Gamma^{\prime\prime}.

Let us consider the quotient Hopf algebra given by quadratically divergent graphs: Δ2(Γ)=γ,sdd(γ)=2γΓ//γ\Delta_{2}(\Gamma)=\sum_{\gamma,\textrm{sdd}(\gamma)=2}\gamma\otimes\Gamma/\!/\gamma. We write

(7.51) Δ2(Γ)=:Γ𝕀+𝕀Γ+Γ2Γ′′.\Delta_{2}(\Gamma)=:\Gamma\otimes\mathbb{I}+\mathbb{I}\otimes\Gamma+\Gamma^{\prime}_{2}\otimes\Gamma^{\prime\prime}.

We add 0=+Γ21Γ′′Γ21Γ′′0=+{\Gamma_{2}^{\prime}}^{-1}\Gamma^{\prime\prime}-{\Gamma_{2}^{\prime}}^{-1}\Gamma^{\prime\prime}, so

(7.52) Γ¯\displaystyle\bar{\Gamma} =\displaystyle= Γ+Γ21Γ′′Γ21Γ′′+Γ1Γ′′\displaystyle\Gamma+{\Gamma_{2}^{\prime}}^{-1}\Gamma^{\prime\prime}-{\Gamma_{2}^{\prime}}^{-1}\Gamma^{\prime\prime}+{\Gamma^{\prime}}^{-1}\Gamma^{\prime\prime}
(7.53) =\displaystyle= (Γ+Γ21Γ′′)+(Γ1Γ21)Γ′′.\displaystyle\left(\Gamma+{\Gamma_{2}^{\prime}}^{-1}\Gamma^{\prime\prime}\right)+\left({\Gamma^{\prime}}^{-1}-{\Gamma_{2}^{\prime}}^{-1}\right)\Gamma^{\prime\prime}.

Here the sum is over all terms of the coproduct with the Γ2\Gamma_{2}^{\prime} terms being present whenever Γ\Gamma^{\prime} is quadratically divergent.

Evaluating the terms Γ2\Gamma_{2}^{\prime} by 1/ψ2(γ2)=ι(γ2)(P=0)1/\psi^{2}(\gamma_{2}^{\prime})=\iota(\gamma_{2}^{\prime})(P=0) decomposes the bar-operation on the level of integrands as follows.

(7.54) ι¯(Γ)(P)=(ι(Γ)(P)+ι(Γ21)(P=0)ι(Γ′′)(P))I+ι(Γ1)(R)ι(Γ′′)(P),\bar{\iota}(\Gamma)(P)=\overbrace{\left(\iota(\Gamma)(P)+{\iota(\Gamma_{2}^{\prime}}^{-1})(P=0)\iota(\Gamma^{\prime\prime})(P)\right)}^{I}+\iota({\Gamma^{\prime}}^{-1})(R)\iota(\Gamma^{\prime\prime})(P),

where ι(Γ1)(R)SR;ϵι(Γ)\iota({\Gamma^{\prime}}^{-1})(R)\equiv S^{\iota}_{R;\epsilon}(\Gamma^{\prime}) appears because a subtraction of a P=0P=0 term, from a quadratically divergent term, precisely delivers those counterterms by our previous analysis. Note that they contain terms which do not have an exponential, as in the example (7.38,7.37). Often, as a two-point vertex of mass type improves the powercounting of the co-graph, we might keep self-energy subgraphs massless, in which case only terms involving RR_{\Box} contribute.

We are left to decompose the terms denoted II. We find by direct computation

(7.56) I\displaystyle I =\displaystyle= [ω(Γ)+ω(Γ21)ω(Γ′′)]eφ(Γ)(P)ψ(Γ)II\displaystyle\overbrace{\left[\omega(\Gamma)+\omega({\Gamma_{2}^{\prime}}^{-1})\omega(\Gamma^{\prime\prime})\right]e^{-\frac{\varphi(\Gamma)(P)}{\psi(\Gamma)}}}^{II}
ω(Γ21)ω(Γ′′)[eφ(Γ)(P)ψ(Γ)eφ(Γ//Γ2)(P)ψ(Γ//Γ2)]III.\displaystyle-\underbrace{\omega({\Gamma_{2}^{\prime}}^{-1})\omega(\Gamma^{\prime\prime})\left[e^{-\frac{\varphi(\Gamma)(P)}{\psi(\Gamma)}}-e^{-\frac{\varphi(\Gamma/\!/\Gamma^{\prime}_{2})(P)}{\psi(\Gamma/\!/\Gamma^{\prime}_{2})}}\right]}_{III}.

The terms denoted IIII gives us the final integrand ι(Γ)(P)\iota(\Gamma)(P) with a corresponding form ωII(Γ)\omega_{II}(\Gamma). ωII(Γ)=ω(Γ)\omega_{II}(\Gamma)=\omega(\Gamma) if there are no subgraphs with sdd=2\textrm{sdd}=2. Note that IIII has the full Γ\Gamma as an argument in the common exponential,

(7.57) II=ωII(Γ)exp(φ(Γ)/ψ(Γ)),II=\omega_{II}(\Gamma)\exp(-\varphi(\Gamma)/\psi(\Gamma)),

which defines ωII\omega_{II}. The rational coefficient ωII\omega_{II} has log-poles only for all subgraphs including the ones with sdd=2\textrm{sdd}=2.

The terms IIIIII is considered in t,ait,a_{i} variables. We can integrate tt as before. As the rational part of the integrand factorizes in Γ2\Gamma^{\prime}_{2} and Γ′′\Gamma^{\prime\prime} variables, we similarly decompose the former into s,bis,b_{i}, iΓ2[1]i\in{\Gamma_{2}^{\prime}}^{[1]}, variables. We note ss only appears in the log (after the tt integration) as a coefficient of ϕΓ,Γ2\phi_{\Gamma,\Gamma^{\prime}_{2}}, using Lemma (7.1). Partial integration in ss eliminates the log and delivers a top-degree form for the bib_{i} integration. These terms precisely compensate against the constant terms mentioned above, as ϕΓ,Γ2=ϕ1(Γ2)ϕ1(ΓΓ2)\phi_{\Gamma,\Gamma_{2}^{\prime}}=\phi_{1}(\Gamma_{2}^{\prime})\phi_{1}(\Gamma-\Gamma_{2}^{\prime}), using that res(Γ2)=2\textrm{res}(\Gamma_{2}^{\prime})=2.

We hence summarize

Theorem 7.3.
(7.58) ΦR(Γ)(P)=limϵ0γSR;ϵΦ(γ)ϵωII(Γ//γ)lnφ(Γ//γ)(P)φ(Γ//γ)(R).\Phi_{R}(\Gamma)(P)=\lim_{\epsilon\to 0}\sum_{\gamma}S^{\Phi}_{R;\epsilon}(\gamma)\int_{\epsilon}\omega_{II}(\Gamma/\!/\gamma)\ln{\frac{\varphi(\Gamma/\!/\gamma)(P)}{\varphi(\Gamma/\!/\gamma)(R)}}.

It is understood that each counterterm is computed with a subtraction RR as befits its argument γ\gamma, and forms Γ\Gamma are chosen in accordance with the previous derivations. Here, ωII\omega_{II} is constructed to have log-poles only. As a projective integral this reads

(7.59) ΦR(Γ)(P)\displaystyle\Phi_{R}(\Gamma)(P) =\displaystyle= ΩΓ[for](1)[for]×\displaystyle\int\Omega_{\Gamma}\sum_{\rm{[for]}}(-1)^{\rm{[for]}}\times
×ln(φ(Γ//[for])(P)jψ(γj)+jφ(γj)(R)ψ(Γ//[for])hjψ(γh)φ(Γ//[for])(R)jψ(γj)+jφ(γj)(R)ψ(Γ//[for])hjψ(γh))\displaystyle\times\ln{\left(\frac{\varphi(\Gamma/\!/\rm{[for]})(P)\prod_{j}\psi(\gamma_{j})+\sum_{j}\varphi(\gamma_{j})(R)\psi(\Gamma/\!/\rm{[for]})\prod_{h\not=j}\psi(\gamma_{h})}{\varphi(\Gamma/\!/\rm{[for]})(R)\prod_{j}\psi(\gamma_{j})+\sum_{j}\varphi(\gamma_{j})(R)\psi(\Gamma/\!/\rm{[for]})\prod_{h\not=j}\psi(\gamma_{h})}\right)}
×ω(Γ//[for])jω(γj).\displaystyle\times\omega(\Gamma/\!/\rm{[for]})\prod_{j}\omega(\gamma_{j}).
Remark 7.4.

Similar formulas can be obtained for the bar-operations and counterterms, with the same rational functions in the integrands, and exponentials exp(φ(Γ//γ)(X)/ψ(Γ//γ))\exp(-\varphi(\Gamma/\!/\gamma)(X)/\psi(\Gamma/\!/\gamma)), with X=PX=P or X=RX=R as needed.

Remark 7.5.

We have worked with choices of renormalizations for mass and wave functions, RR,Rm2R\to R_{\Box},R_{m^{2}}. One can actually also define PP,Pm2P\to P_{\Box},P_{m^{2}}, and for example set masses to zero in all exponentials (φ()(P)ϕq2()\varphi(\cdot)(P)\to\phi_{q^{2}}(\cdot)), that’s essentially the Weinberg scheme if one then subtracts at q2=μ2q^{2}=\mu^{2}.

Remark 7.6.

This all is nicely reflected in properties of analytic regulators. For example in dimensional regularization the identity dDk[k2]ρ=0\int d^{D}k[k^{2}]^{\rho}=0, ρ\forall\rho, leads to Φ(Γ)(P=0)=0\Phi(\Gamma)(P=0)=0 immediately, where Φ\Phi now indicates unrenormalized Feynman rules using that regulator.

Remark 7.7.

We are working so far with constant lower boundaries. The chains introduced in previous sections have moving lower boundaries which respect the hierarchy in each flag. We will study that difference in section (9.1).

7.5. Specifics of the MOM-scheme

We define the MOM-scheme by setting all masses to zero in radiative corrections and keeping a single kinematical invariant q2q^{2} in the ϕ\phi-polynomial, P={0},{qiqjq2}P=\{0\},\{q_{i}\cdot q_{j}\sim q^{2}\},

(7.60) ϕ(Γ)=q2Rq2(Γ).\phi(\Gamma)=q^{2}R_{q^{2}}(\Gamma).

Such a situation arises if we set masses to zero (possibly after factorization of a polynomial part from the amplitude as in the Weinberg scheme), and for vertices if we consider the case of zero momentum transfers, or evaluate at a symmetric point qi2=q2q_{i}^{2}=q^{2}, where ii denotes the external half-edges of Γ\Gamma. If we want to emphasize the q2q^{2} dependence we write ϕq2\phi_{q^{2}}. Trivially, ϕq2=q2ϕ1\phi_{q^{2}}=q^{2}\phi_{1}. In the MOM-scheme, subtractions are done at q2=μ2q^{2}=\mu^{2}, which defines RR for all graphs. Counter-terms in the MOM-scheme become very simple when expressed in parametric integrals thanks to the homogeneity of the ϕ\phi-polynomial. Note that we hence have φ(Γ)=ϕ(Γ)\varphi(\Gamma)=\phi(\Gamma) as we have set all masses to zero.

In a MOM-scheme, renormalized diagrams are polynomials in lnq2/μ2\ln q^{2}/\mu^{2}:

Theorem 7.8.

For all Γ\Gamma,

(7.61) ΦMOM(Γ)(q2/μ2)=j=1aug(Γ)cj(Γ)lnjq2/μ2.\Phi_{\rm{MOM}}(\Gamma)(q^{2}/\mu^{2})=\sum_{j=1}^{\rm{aug}(\Gamma)}c_{j}(\Gamma)\ln^{j}{q^{2}/\mu^{2}}.

Here, aug(Γ)=max[for]|[for]|\textrm{aug}(\Gamma)=\max_{\textrm{[for]}}|\textrm{[for]}|.
Proof: Consider a sequence γ1γ2γaug(Γ)Γ\gamma_{1}\subsetneq\gamma_{2}\cdots\gamma_{\textrm{aug}(\Gamma)}\subsetneq\Gamma. This is in one-to-one correspondence with some decorated rooted tree appearing in ρ(Γ)\rho_{\mathcal{R}}(\Gamma) (2.35). Choose one edge ejγj/γj1e_{j}\in\gamma_{j}/\gamma_{j-1} in each decoration and de-homogenize with respect to that edge. We get a sequence of lower boundaries ϵ,ϵ/A2,ϵ/A2/A3,\epsilon,\epsilon/A_{2},\epsilon/A_{2}/A_{3},\cdots. Use the affine representation and integrate to obtain the result. \Box

7.5.1. MOM scheme results from residues

In such a scheme, it is particularly useful to take a derivative with respect to lnq2\ln q^{2}. We consider

(7.62) p1(Γ):=q2q2ΦMOM(Γ)(q2/μ2)|q2=μ2,p_{1}(\Gamma):=q^{2}\partial_{q^{2}}\Phi_{\textrm{MOM}}(\Gamma)(q^{2}/\mu^{2})_{|_{q^{2}=\mu^{2}}},

where we evaluate at q2=μ2q^{2}=\mu^{2} after taking the derivative. This number, which for a primitive element of the renormalization Hopf algebra is the residue of that graph in the sense of [2], is our main concern for a general graph. It will be obtained in the limit of the limiting mixed Hodge structure we construct.

Remark 7.9.

It is not that this limit would not exist for general schemes. But the limit would be a complicated function of ratios of masses and kinematical invariants, which has a constant term given by the number p1(Γ)p_{1}(\Gamma) and beyond that a dependence on these ratios which demands a much finer Hodge theoretic study than we can offer here.

But first we need to remind ourselves how coefficients of higher powers of logarithms of complicated graphs related to coefficients of lower powers of sub- and co-graphs thanks to the renormalization group.

7.5.2. The counterterm SMOMΦS_{\rm MOM}^{\Phi}

For SMOMΦ(Γ)=:j=1aug(Γ)sj(Γ)lnjμ2S_{\textrm{MOM}}^{\Phi}(\Gamma)=:\sum_{j=1}^{\textrm{aug}(\Gamma)}s_{j}(\Gamma)\ln^{j}\mu^{2}, we simply use the renormalization group or the scattering type formula. In particular, we have

(7.63) SMOMΦ(Γ)=j=1aug(Γ)1j!(1)j[p1p1]jfactorsΔj1(Γ).S_{\textrm{M}OM}^{\Phi}(\Gamma)=\sum_{j=1}^{\textrm{aug}(\Gamma)}\frac{1}{j!}(-1)^{j}\underbrace{[p_{1}\otimes\cdots\otimes p_{1}]}_{{j\textrm{factors}}}\Delta^{j-1}(\Gamma).

This is easily derived [6, 17] upon noting that p1(Γ)=Φ(SY(Γ))p_{1}(\Gamma)=\Phi(S\star Y(\Gamma)).

Note that this determines counter-terms by iteration: for a kk-loop graph, knowledge of all the lower order counterterms suffices to determine all contributions to the kk-loop counterterm but the lowest order coefficient of lnμ2\ln\mu^{2}. But then, that coefficient is given by the formula

(7.64) s1(Γ)=p1(Γ)lnμ2,s_{1}(\Gamma)=p_{1}(\Gamma)\ln\mu^{2},

which itself only involves counter-terms of less than kk loops, by the structure of the bar operation.

7.5.3. p1(Γ)p_{1}(\Gamma) from co-graphs

We can now summarize the consequences of the renormalization group and our projective representations for parametric representations of Feynman integrals. The interesting question is about the logs which we had in numerators. Thm.(7.3) becomes

Theorem 7.10.
(7.65) p1(Γ)=limϵ0γSMOM;ϵΦ(γ)q2q2ϵωII(Γ//γ)lnϕq2/μ2(Γ//γ).p_{1}(\Gamma)=\lim_{\epsilon\to 0}\sum_{\gamma}S^{\Phi}_{\rm{MOM};\epsilon}(\gamma)q^{2}\partial_{q^{2}}\int_{\epsilon}\omega_{II}(\Gamma/\!/\gamma)\ln{\phi_{q^{2}/\mu^{2}}(\Gamma/\!/\gamma)}.

This limit is

(7.66) p1(Γ)\displaystyle p_{1}(\Gamma) =\displaystyle= ΩΓ[for](1)[for]×\displaystyle\int\Omega_{\Gamma}\sum_{\rm{[for]}}(-1)^{\rm{[for]}}\times
×q2q2ln(ϕq2/μ2(Γ//[for])jψ(γj)+jϕ1(γj)ψ(Γ//[for])hjψ(γh))\displaystyle\times q^{2}\partial_{q^{2}}\ln{\left(\phi_{q^{2}/\mu^{2}}(\Gamma/\!/\rm{[for]})\prod_{j}\psi(\gamma_{j})+\sum_{j}\phi_{1}(\gamma_{j})\psi(\Gamma/\!/\rm{[for]})\prod_{h\not=j}\psi(\gamma_{h})\right)}
×ω(Γ//[for])jω(γj).\displaystyle\times\omega(\Gamma/\!/\rm{[for]})\prod_{j}\omega(\gamma_{j}).

The derivative with respect to lnq2\ln q^{2} can be taken inside the integral in (7.65) if and only if all edges carrying external momentum are in the complement C(Γ)C(\Gamma) of all edges belonging to divergent subgraphs. In that case, q2q2lnϕq2(Γ//γ)=1q^{2}\partial_{q^{2}}\ln{\phi_{q^{2}}(\Gamma/\!/\gamma)}=1 and no logs in the numerator appear.

Remark 7.11.

Note that overlapping divergent graphs can force all edges to belong to divergent subgraphs, cf. Fig.(5).

Proof: If all edges carrying external momentum are in the complement to divergent subgraphs, we bring the counter-terms under the integrand using the bar-operation. We can take the limit ϵ0\epsilon\to 0 in the integrand for all edge variables belonging to subgraphs, and this limit commutes with the derivative with respect to lnq2\ln q^{2} by assumption: each ϕq2(Γ//γ)\phi_{q}^{2}(\Gamma/\!/\gamma) is a linear combination of terms Aeψe(Γ//γ)A_{e}\psi_{e}(\Gamma/\!/\gamma), where ee is in that complement C(Γ)C(\Gamma) of subgraph edges, and ψe(Γ//γ)=ψ(Γ//γ/e)\psi_{e}(\Gamma/\!/\gamma)=\psi(\Gamma/\!/\gamma/e). Applying then the Chen-Wu theorem [21] with respect to the elements of C(Γ)C(\Gamma) disentangles the q2q^{2} dependence from the limit in ϵ\epsilon. \Box

Remark 7.12.

Note that the discussion below with respect to the limiting Hodge structure assumes that we have this situation of disentanglement of divergent subgraphs and edges carrying external momentum. We hence have no logarithms in the numerator. But note that the general case does no harm to the ensuing discussion: by Lemma (7.1), any logarithms in the numerator are congruent to one along any exceptional divisor of XΓ//[for]X_{\Gamma/\!/\rm{[for]}}. Furthermore, when external momentum interferes with subgraphs, all logs can be turned to rational functions by a partial integration. The fact that the second Kirchhoff–Symanzik polynomial is a linear combination of ψ\psi-polynomials, applied to graphs with an extra shrunken edge, in the MOM-case establishes these rational functions to have poles coming from our analysis of this ψ(Γ)\psi(\Gamma) polynomial. A full mathematical discussion of this ”ωlnf\int\omega\ln f” situation should be subject to future work.

7.5.4. Examples

¿From now on we measure q2q^{2} in units of μ2\mu^{2} so that subtractions are done at 11. This simplifies notation. Let us first consider the Dunce’s cap in detail, (14).

Refer to caption

Figure 14. The Dunce’s cap, again. We label the edges 1,2,3,41,2,3,4. Resolved in trees, we find three trees in the core Hopf algebra. We label the vertices by edge labels of the graph. The sets 123123 and 124124 correspond to a triangle graph as indicated, the sets 1212 and 3434 are one-loop vertex graphs, and tadpoles appear in the coproduct on the rhs for edges 33 or 44. The coproduct in the core Hopf algebra is, expressed in edge labels, Δ(1234)=1234+1243+3412\Delta^{\prime}(1234)=123\otimes 4+124\otimes 3+34\otimes 12. Only the last term appears in the renormalization Hopf algebra.

We have the following data (pathq(Γ)\textrm{path}_{q}(\Gamma) refers to the momentum path through the graph):

(7.67) ψ(Γ)\displaystyle\psi(\Gamma) =\displaystyle= (A1+A2)(A3+A4)+A3A4,\displaystyle(A_{1}+A_{2})(A_{3}+A_{4})+A_{3}A_{4},
(7.68) ψ(γ)\displaystyle\psi(\gamma) =\displaystyle= A3+A4,ψ(Γ//γ)=A1+A2,\displaystyle A_{3}+A_{4},\psi(\Gamma/\!/\gamma)=A_{1}+A_{2},
(7.69) pathq(Γ)\displaystyle\textrm{path}_{q}(\Gamma) =\displaystyle= e1,\displaystyle e_{1},
(7.70) ϕ1(Γ)\displaystyle\phi_{1}(\Gamma) =\displaystyle= A1(A2A3+A3A4+A4A2)=A1ψ(Γ//e1)=A1ψ1(Γ),\displaystyle A_{1}(A_{2}A_{3}+A_{3}A_{4}+A_{4}A_{2})=A_{1}\psi(\Gamma/\!/e_{1})=A_{1}\psi^{1}(\Gamma),
(7.71) ϕ1(γ)\displaystyle\phi_{1}(\gamma) =\displaystyle= A3A4,ϕ1(Γ//γ)=A1A2,\displaystyle A_{3}A_{4},\phi_{1}(\Gamma/\!/\gamma)=A_{1}A_{2},
(7.72) {[for]}\displaystyle\{\textrm{[for]}\} =\displaystyle= {,(34)}.\displaystyle\{\emptyset,(34)\}.
(7.73) Φ(Γ)ϵ(q2)=ϵi=14dAiexpq2ϕ1(Γ)ψ(Γ)ψ2(Γ).\Phi(\Gamma)_{\epsilon}(q^{2})=\int_{\epsilon}^{\infty}\prod_{i=1}^{4}d\!A_{i}\frac{\exp{-q^{2}\frac{\phi_{1}(\Gamma)}{\psi(\Gamma)}}}{\psi^{2}(\Gamma)}.

Hence we choose a function τ(ϵ)\tau(\epsilon) which goes to zero rapidly enough so that limϵ0τ(ϵ)/ϵ=0\lim_{\epsilon\to 0}\tau(\epsilon)/\epsilon=0 and compute

(7.75) Φ¯ϵ(Γ)(q2,μ2)\displaystyle\bar{\Phi}_{\epsilon}(\Gamma)(q^{2},\mu^{2}) =\displaystyle= ϵdA1dA2τ(ϵ)dA3dA4{expq2ϕ1(Γ)ψ(Γ)ψ2(Γ)\displaystyle\int_{\epsilon}^{\infty}d\!A_{1}d\!A_{2}\int_{\tau(\epsilon)}^{\infty}d\!A_{3}d\!A_{4}\left\{\frac{\exp{-q^{2}\frac{\phi_{1}(\Gamma)}{\psi(\Gamma)}}}{\psi^{2}(\Gamma)}\right.
exp[q2ϕ1(Γ//γ)ψ(Γ//γ)]ψ2(Γ//γ)exp[ϕ1(γ)ψ(γ)]ψ2(γ)}\displaystyle-\left.\frac{\exp{\left[-q^{2}\frac{\phi_{1}(\Gamma/\!/\gamma)}{\psi(\Gamma/\!/\gamma)}\right]}}{\psi^{2}(\Gamma/\!/\gamma)}\frac{\exp{\left[-\frac{\phi_{1}(\gamma)}{\psi(\gamma)}\right]}}{\psi^{2}(\gamma)}\right\}
=\displaystyle= q2ϵ𝑑A1𝑑A2τ(ϵ)q2𝑑A3𝑑A4{expϕ1(Γ)ψ(Γ)ψ2(Γ)}\displaystyle\int_{q^{2}\epsilon}^{\infty}d\!A_{1}d\!A_{2}\int_{\tau(\epsilon)q^{2}}^{\infty}d\!A_{3}d\!A_{4}\left\{\frac{\exp{-\frac{\phi_{1}(\Gamma)}{\psi(\Gamma)}}}{\psi^{2}(\Gamma)}\right\}
{q2ϵ𝑑A1𝑑A2τ(ϵ)𝑑A3𝑑A4exp[ϕ1(Γ//γ)ψ(Γ//γ)]ψ2(Γ//γ)exp[ϕ1(γ)ψ(γ)]ψ2(γ)}.\displaystyle-\left\{\int_{q^{2}\epsilon}^{\infty}d\!A_{1}d\!A_{2}\int_{\tau(\epsilon)}^{\infty}d\!A_{3}d\!A_{4}\frac{\exp{\left[-\frac{\phi_{1}(\Gamma/\!/\gamma)}{\psi(\Gamma/\!/\gamma)}\right]}}{\psi^{2}(\Gamma/\!/\gamma)}\frac{\exp{\left[-\frac{\phi_{1}(\gamma)}{\psi(\gamma)}\right]}}{\psi^{2}(\gamma)}\right\}.

Let us now re-scale to variables AiA1BiA_{i}\to A_{1}B_{i} for all variables i2,3,4i\in 2,3,4. We get

(7.76) Φ¯ϵ(Γ)(q2,μ2)\displaystyle\bar{\Phi}_{\epsilon}(\Gamma)(q^{2},\mu^{2}) =\displaystyle= q2ϵdA1A1q2ϵ𝑑B2q2τ(ϵ)/A1𝑑B3𝑑B4{expA1(B2B3+B3B4+B4B2)(1+B2)(B3+B4)+B3B4[(1+B2)(B3+B4)+B3B4]2}\displaystyle\int_{q^{2}\epsilon}^{\infty}\frac{d\!A_{1}}{A_{1}}\int_{q^{2}\epsilon}^{\infty}d\!B_{2}\int_{q^{2}\tau(\epsilon)/A_{1}}^{\infty}d\!B_{3}d\!B_{4}\left\{\frac{\exp{-A_{1}\frac{(B_{2}B_{3}+B_{3}B_{4}+B_{4}B_{2})}{(1+B_{2})(B_{3}+B_{4})+B_{3}B_{4}}}}{[(1+B_{2})(B_{3}+B_{4})+B_{3}B_{4}]^{2}}\right\}
{q2ϵdA1A1q2ϵ𝑑B2τ(ϵ)/A1𝑑B3𝑑B4expA1B21+B2(1+B2)2expA1B3B4B3+B4(B3+B4)2}.\displaystyle-\left\{\int_{q^{2}\epsilon}^{\infty}\frac{d\!A_{1}}{A_{1}}\int_{q^{2}\epsilon}^{\infty}d\!B_{2}\int_{\tau(\epsilon)/A_{1}}^{\infty}d\!B_{3}d\!B_{4}\frac{\exp{-A_{1}\frac{B_{2}}{1+B_{2}}}}{(1+B_{2})^{2}}\frac{\exp{-A_{1}\frac{B_{3}B_{4}}{B_{3}+B_{4}}}}{(B_{3}+B_{4})^{2}}\right\}.

We re-scale once more B4=B3C4B_{4}=B_{3}C_{4}. Also, we set the lower boundaries in the B2B_{2} and C4C_{4} integrations to zero. This is justified as A1A_{1} and B3B_{3} remain positive.

(7.77) Φ¯ϵ(Γ)(q2,μ2)\displaystyle\bar{\Phi}_{\epsilon}(\Gamma)(q^{2},\mu^{2}) =\displaystyle= q2ϵdA1A10𝑑B2q2τ(ϵ)/A1dB3B30𝑑C4{expA1(B2+B3C4+C4B2)(1+B2)(1+C4)+B3C4[(1+B2)(1+C4)+B3C4]2}\displaystyle\int_{q^{2}\epsilon}^{\infty}\frac{d\!A_{1}}{A_{1}}\int_{0}^{\infty}d\!B_{2}\int_{q^{2}\tau(\epsilon)/A_{1}}^{\infty}\frac{d\!B_{3}}{B_{3}}\int_{0}^{\infty}d\!C_{4}\left\{\frac{\exp{-A_{1}\frac{(B_{2}+B_{3}C_{4}+C_{4}B_{2})}{(1+B_{2})(1+C_{4})+B_{3}C_{4}}}}{[(1+B_{2})(1+C_{4})+B_{3}C_{4}]^{2}}\right\}
q2ϵdA1A10𝑑B2τ(ϵ)/A1dB3B30𝑑C4{expA1B21+B2(1+B2)2expA1B3C41+C4(1+C4)2}.\displaystyle-\int_{q^{2}\epsilon}^{\infty}\frac{d\!A_{1}}{A_{1}}\int_{0}^{\infty}d\!B_{2}\int_{\tau(\epsilon)/A_{1}}^{\infty}\frac{d\!B_{3}}{B_{3}}\int_{0}^{\infty}d\!C_{4}\left\{\frac{\exp{-A_{1}\frac{B_{2}}{1+B_{2}}}}{(1+B_{2})^{2}}\frac{\exp{-A_{1}B_{3}\frac{C_{4}}{1+C_{4}}}}{(1+C_{4})^{2}}\right\}.

Taking a derivative wrt lnq2\ln q^{2} and using that limϵ0τ(ϵ)/ϵ=0\lim_{\epsilon\to 0}\tau(\epsilon)/\epsilon=0, delivers three remaining terms

(7.78) lnq2Φ¯ϵ(Γ)q2=1\displaystyle\partial_{\ln q^{2}}\bar{\Phi}_{\epsilon}(\Gamma)_{q^{2}=1} =\displaystyle= 0𝑑B2τ(ϵ)/ϵdB3B30𝑑C4{1[(1+B2)(1+C4)+B3C4]2}\displaystyle\int_{0}^{\infty}d\!B_{2}\int_{\tau(\epsilon)/\epsilon}^{\infty}\frac{d\!B_{3}}{B_{3}}\int_{0}^{\infty}d\!C_{4}\left\{\frac{1}{[(1+B_{2})(1+C_{4})+B_{3}C_{4}]^{2}}\right\}
0𝑑B2τ(ϵ)/(q2ϵ)dB3B30𝑑C4{1(1+B2)2eϵq2B3C41+C4(1+C4)2}\displaystyle-\int_{0}^{\infty}d\!B_{2}\int_{\tau(\epsilon)/(q^{2}\epsilon)}^{\infty}\frac{d\!B_{3}}{B_{3}}\int_{0}^{\infty}d\!C_{4}\left\{\frac{1}{(1+B_{2})^{2}}\frac{e^{-\epsilon q^{2}\frac{B_{3}C_{4}}{1+C_{4}}}}{(1+C_{4})^{2}}\right\}
+q2ϵdA1A10𝑑B20𝑑C4{eA1B2(1+B2)[(1+B2)(1+C4)]2}.\displaystyle+\int_{q^{2}\epsilon}^{\infty}\frac{d\!A_{1}}{A_{1}}\int_{0}^{\infty}d\!B_{2}\int_{0}^{\infty}d\!C_{4}\left\{\frac{e^{-A_{1}\frac{B_{2}}{(1+B_{2})}}}{[(1+B_{2})(1+C_{4})]^{2}}\right\}.

Integrating B3B_{3} in the second line and A1A_{1} in the third, we find

(7.79) lnq2Φ¯ϵ(Γ)q2=1\displaystyle\partial_{\ln q^{2}}\bar{\Phi}_{\epsilon}(\Gamma)_{q^{2}=1} =\displaystyle= 0𝑑B2τ(ϵ)/ϵdB3B30𝑑C41[(1+B2)(1+C4)+B3C4]2\displaystyle\int_{0}^{\infty}d\!B_{2}\int_{\tau(\epsilon)/\epsilon}^{\infty}\frac{d\!B_{3}}{B_{3}}\int_{0}^{\infty}d\!C_{4}\frac{1}{[(1+B_{2})(1+C_{4})+B_{3}C_{4}]^{2}}
+lnτ(ϵ)/ϵΩγψ2(γ)ΩΓ//γψ2(Γ//γ).\displaystyle+\ln{\tau(\epsilon)/\epsilon}\int\frac{\Omega_{\gamma}}{\psi^{2}(\gamma)}\int\frac{\Omega_{\Gamma/\!/\gamma}}{\psi^{2}(\Gamma/\!/\gamma)}.

Using the exponential integral, those B3B_{3} and A1A_{1} integrations also deliver finite contributions

(7.80) 0𝑑B20𝑑C4{1(1+B2)2lnC41+C4(1+C4)2}+0𝑑B20𝑑C4{lnB21+B2[(1+B2)(1+C4)]2}=0.-\int_{0}^{\infty}d\!B_{2}\int_{0}^{\infty}d\!C_{4}\left\{\frac{1}{(1+B_{2})^{2}}\frac{\ln{\frac{C_{4}}{1+C_{4}}}}{(1+C_{4})^{2}}\right\}+\int_{0}^{\infty}d\!B_{2}\int_{0}^{\infty}d\!C_{4}\left\{\frac{\ln{\frac{B_{2}}{1+B_{2}}}}{[(1+B_{2})(1+C_{4})]^{2}}\right\}=0.

This cancellation of logs is no accident: while in this simple example it looks as if it originates from the fact that the co-graph and subgraph are identical, actually the cross-ratio

(7.81) lnϕ(Γ//γ)ψ(γ)ψ(Γ//γ)ϕ(γ)\ln\frac{\phi(\Gamma/\!/\gamma)\psi(\gamma)}{\psi(\Gamma/\!/\gamma)\phi(\gamma)}

vanishes identically when integrated against the de-homogenized product measure

(7.82) 0𝑑AΓ//γ𝑑Aγ1ψ2(Γ//γ)ψ2(γ).\int_{0}dA_{\Gamma/\!/\gamma}dA_{\gamma}\frac{1}{\psi^{2}(\Gamma/\!/\gamma)\psi^{2}(\gamma)}.

This is precisely because C(Γ)=e1C(\Gamma)=e_{1} has an empty intersection with γ[1]=e3,e4\gamma^{[1]}=e_{3},e_{4}.

But then, this cancelation of logs will break down if ϕ(Γ)\phi(\Gamma) is not as nicely disentangled from ϕ(γ)\phi(\gamma) for all log-poles as it is here, and will be replaced by logs congruent to 1 along subdivergences in general, in accordance with Thm.(7.10).

Let us study this in some detail. Consider the graph on the upper left in Fig.(7), and consider the finite lnϕ/ψ\ln\phi/\psi-type contributions of the exponential integral to in the vicinity of the exceptional divisor for the subspace A3=A4=0A_{3}=A_{4}=0.

Routing an external momentum through edges 1,6, we have the following graph polynomials:

ϕ1(Γ)\displaystyle\phi_{1}(\Gamma) =\displaystyle= A1[A3A4(A5+A6)+A5A6(A3+A4)+A2(A3+A4)(A5+A6)]\displaystyle A_{1}[A_{3}A_{4}(A_{5}+A_{6})+A_{5}A_{6}(A_{3}+A_{4})+A_{2}(A_{3}+A_{4})(A_{5}+A_{6})]
+A6A5[(A1+A2)(A3+A4)+A3A4]\displaystyle+A_{6}A_{5}[(A_{1}+A_{2})(A_{3}+A_{4})+A_{3}A_{4}]
(7.84) ϕ1(Γ/34)\displaystyle\phi_{1}(\Gamma/34) =\displaystyle= A1[A5A6+A2(A5+A6)]+A5A6[(A1+A2)]\displaystyle A_{1}[A_{5}A_{6}+A_{2}(A_{5}+A_{6})]+A_{5}A_{6}[(A_{1}+A_{2})]
(7.85) ϕ1(34)\displaystyle\phi_{1}(34) =\displaystyle= A3A4\displaystyle A_{3}A_{4}
(7.86) ψ(Γ/34)\displaystyle\psi(\Gamma/34) =\displaystyle= (A1+A2)(A5+A6)+A5A6\displaystyle(A_{1}+A_{2})(A_{5}+A_{6})+A_{5}A_{6}
(7.87) ψ(34)\displaystyle\psi(34) =\displaystyle= A3+A4.\displaystyle A_{3}+A_{4}.

We have C(Γ)=e1,e6C(\Gamma)=e_{1},e_{6}, and γΓ,res(γ)0γ[1]=e3,e4,e5,e6\cup_{{\gamma\subsetneq\Gamma,\textrm{res}(\gamma)\geq 0}}\gamma^{[1]}=e_{3},e_{4},e_{5},e_{6}. The intersection is e6e_{6}. We hence find, with suitable de-homogenization,

(7.88) lnB5B6(1+B2)X+B5B6+B2(B5+B6)Y(1+B2)(B5+B6)+B5B6lnC41+C4[(1+B2)(B5+B6)+B5B6]2[1+C4]2dB2dC4dB5dB6.\frac{\ln{\frac{\overbrace{B_{5}B_{6}(1+B_{2})}^{X}+\overbrace{B_{5}B_{6}+B_{2}(B_{5}+B_{6})}^{Y}}{(1+B_{2})(B_{5}+B_{6})+B_{5}B_{6}}}-\ln{\frac{C_{4}}{1+C_{4}}}}{[(1+B_{2})(B_{5}+B_{6})+B_{5}B_{6}]^{2}[1+C_{4}]^{2}}d\!B_{2}d\!C_{4}d\!B_{5}d\!B_{6}.

Here, the term XX denotes a term which would be absent if the momenta would only go through edge 1 and hence the above intersection would be empty, while YY indicates the terms from the momentum flow through edge 1.

This is of the form ln(fΓ//γ/fγ)[ωΓ//γωγ]\ln(f_{\Gamma/\!/\gamma}/f_{\gamma})[\omega_{\Gamma/\!/\gamma}\wedge\omega_{\gamma}]. If the term XX would be absent, a partial integration

(7.89) ϵlnxu+vxu+w(xu+w)2ϵ1(xu+w)2\int_{\epsilon}^{\infty}\frac{\ln{\frac{xu+v}{xu+w}}}{(xu+w)^{2}}\sim\int_{\epsilon}^{\infty}\frac{1}{(xu+w)^{2}}

would show the vanishing of this expression as above. The presence of XX leaves us with a contribution which can be written, replacing lnC4/(1+C4)\ln{C_{4}/(1+C_{4})} by lnY/ψ(Γ/34)\ln Y/\psi(\Gamma/34),

(7.90) lnB5B6(1+B2)X+B5B6+B2(B5+B6)YB5B6+B2(B5+B6)Y[(1+B2)(B5+B6)+B5B6]2[1+C4]2.\frac{\ln{\frac{\overbrace{B_{5}B_{6}(1+B_{2})}^{X}+\overbrace{B_{5}B_{6}+B_{2}(B_{5}+B_{6})}^{Y}}{\underbrace{B_{5}B_{6}+B_{2}(B_{5}+B_{6})}_{Y}}}}{[(1+B_{2})(B_{5}+B_{6})+B_{5}B_{6}]^{2}[1+C_{4}]^{2}}.

As promised, it is congruent to one along the remaining log-pole at A5=A6=0A_{5}=A_{6}=0. It has to be: the forest where the subgraph 5656 shrinks to a point looses the momentum flow through edge 6 and could not contribute any counterterm for a pole remaining in the terms discussed above.

Note that in general higher powers of logarithms can appear in the numerator as subgraphs can have substructure. Lacking a handle to notate all the log-poles which do not cancel due to partial integration identities known beyond mankind we consider it understood that all terms from the asymptotic expansion of the exponential integral up to constant terms (higher order terms in ϵ\epsilon are not needed as all poles are logarithmic only) are kept without being shown explicitly in further notation. We emphasize though that all those logarithm terms in the numerator are congruent to one along log-poles -and deserve study in their own right elsewhere-, and hence thanks to Lemma (7.1) which guarantees indeed all necessary cancelations, we have in all cases:

(7.91) p1(Γ)=limϵ0lnq2Φ¯ϵ(Γ)q2=1.p_{1}(\Gamma)=\lim_{\epsilon\to 0}\partial_{\ln q^{2}}\bar{\Phi}_{\epsilon}(\Gamma)_{q^{2}=1}.
Remark 7.13.

There is freedom in the choice of τ\tau, a natural choice comes from the rooted tree representation ρ(Γ)\rho(\Gamma) of the forest. Each forest is part of a legal tree tt and any subgraph γ\gamma corresponds to a vertex vv in that tree. If dvd_{v} is the distance of vv to the root of tt, τ(ϵ)=ϵdv+1\tau(\epsilon)=\epsilon^{d_{v}+1} is a natural choice.

8. NN

8.1. for physicists: The antipode as monodromy

Let us now come back to the core Hopf algebra and prepare for an analysis in terms of limiting mixed Hodge structures. This will be achieved in two steps: an analysis of the structure of the antipode of the renormalization Hopf algebra, which will then allow to define a matrix NN for the monodromy in question such that S(Γ)S(\Gamma) can be expressed in a particularly nice way. In fact, because of orientations, the NN which arises in the monodromy calculation is the negative of the NN computed in this section. We omit the minus sign to simplify the notation.

Let us consider the antipode first. Thanks to the above lemma we can write for the antipode S(Γ)S(\Gamma)

(8.1) S(Γ)=j=0|Γ|(1)j|C|=jtPC(t)RC(t).S(\Gamma)=-\sum_{j=0}^{|\Gamma|}(-1)^{j}\sum_{|C|=j}\sum_{t}P^{C}(t)R^{C}(t).

Here, we abuse notation in an obvious manner identifying Γ\Gamma and ρT(Γ)\rho_{T}(\Gamma), the latter being the indicated sum over trees, in accordance with Eq.(2.34).

We also define R¯(Γ)=S(Γ)\overline{R}(\Gamma)=-S(\Gamma). Let us now label the edges of each t(Γ)t(\Gamma) once and for all by 1,2,,|Γ|11,2,\cdots,|\Gamma|-1. Then, we have |Γ|1|\Gamma|-1 cuts CC with |C|=1|C|=1, and

(8.2) (|Γ|1j)\left({\genfrac{}{}{0.0pt}{}{|\Gamma|-1}{j}}\right)

cuts of cardinality |C|=j|C|=j. We hence can define a vector v(Γ)v(\Gamma) with 2|Γ|12^{|\Gamma|-1} entries in HH, ordered according to a never decreasing cardinality of cuts:

(8.3) v(Γ)=(Γ,tPC(t)RC(t)(|Γ|11) entries of cardinality 1,,tPC(t)RC(t)(|Γ|1j) entries of cardinality j,)T.v(\Gamma)=(\Gamma,\underbrace{\sum_{t}P^{C}(t)R^{C}(t)}_{\textrm{$\left({\genfrac{}{}{0.0pt}{}{|\Gamma|-1}{1}}\right)$ entries of cardinality $1$}},\cdots,\underbrace{\sum_{t}P^{C}(t)R^{C}(t)}_{\textrm{$\left({\genfrac{}{}{0.0pt}{}{|\Gamma|-1}{j}}\right)$ entries of cardinality $j$}},\cdots)^{T}.

Example: Dunce’s cap with edges 1,2,3,41,2,3,4 and divergent subgraph 3,43,4, comare Fig.(14). The core coproduct is

(8.4) Δc=1234+1243+3412.\Delta_{c}^{\prime}=123\otimes 4+124\otimes 3+34\otimes 12.

The vector vv is then

(8.5) v=(1234(123)(4)+(124)(3)+(12)(34)).v=\left({\genfrac{}{}{0.0pt}{}{1234}{(123)(4)+(124)(3)+(12)(34)}}\right).

Let N(2)N^{(2)} be the to-by-two matrix

(8.6) N(2)=(0100).N^{(2)}=\left(\begin{array}[]{cc}0&1\\ 0&0\\ \end{array}\right).

Note that

(8.7) [(1001)N(2)](1234(123)(4)+(124)(3)+(12)(34))=(R¯(1234)(123)(4)+(124)(3)+(12)(34)),\left[\left(\begin{array}[]{cc}1&0\\ 0&1\\ \end{array}\right)-N^{(2)}\right]\left(\begin{array}[]{c}1234\\ (123)(4)+(124)(3)+(12)(34)\\ \end{array}\right)=\left(\begin{array}[]{c}\overline{R}(1234)\\ (123)(4)+(124)(3)+(12)(34)\\ \end{array}\right),

with

(8.8) R¯(1234)=1234(123)(4)(124)(3)(34)(12).\overline{R}(1234)=1234-(123)(4)-(124)(3)-(34)(12).

In fact, it is our first task to find a nilpotent matrix NN, N|Γ|=0N^{|\Gamma|}=0, such that

(8.9) j=0|Γ|1(1)jNj/j!=(R¯(Γ),tR¯(PC(t))R¯(RC(t))(|Γ|11) entries of cardinality 1,,tR¯(PC(t))R¯(RC(t))(|Γ|1j) entries of cardinality j,),)T.\sum_{j=0}^{|\Gamma|-1}(-1)^{j}N^{j}/j!=(\overline{R}(\Gamma),\underbrace{\sum_{t}\overline{R}(P^{C}(t))\overline{R}(R^{C}(t))}_{\textrm{$\left({\genfrac{}{}{0.0pt}{}{|\Gamma|-1}{1}}\right)$ entries of cardinality $1$}},\cdots,\underbrace{\sum_{t}\overline{R}(P^{C}(t))\overline{R}(R^{C}(t))}_{\textrm{$\left({\genfrac{}{}{0.0pt}{}{|\Gamma|-1}{j}}\right)$ entries of cardinality $j$}},\cdots),)^{T}.

For PC(t)=itiP^{C}(t)=\prod_{i}t_{i} we here have abbreviated R¯(PC(t))\overline{R}(P^{C}(t)) for iR¯(ti)\prod_{i}\overline{R}(t_{i}).

8.2. The matrix NN

Let M(0,1)M(0,1) be the space of matrices with entries in the two point set {0,1}\{0,1\}.

Let now m+1m+1 be the number of loops m=|Γ|1m=|\Gamma|-1 in the graph and let us construct a nilpotent 2m×2m2^{m}\times 2^{m} square matrix NN(m)N\equiv N^{(m)}, Nm+1=0N^{m+1}=0, in M(0,1)M(0,1) as follows.

Consider first the m+1m+1-th row of the Pascal triangle, for example for m=3m=3 it reads 1,3,3,11,3,3,1. For this example, we will then construct blocks of sizes 1×11\times 1, 1×31\times 3, 3×33\times 3, 3×13\times 1 and 1×11\times 1, all with entries either 0 or 11.

So this gives us in general m+2m+2 blocks Mj(m)M^{(m)}_{j}, 0jm+10\leq j\leq m+1, of matrices of size M0(m):1×1M^{(m)}_{0}:1\times 1, M1(m):1×mM^{(m)}_{1}:1\times m, M2(m):m×m(m1)/2!M^{(m)}_{2}:m\times m(m-1)/2!, \cdots, Mm(m):m×1M^{(m)}_{m}:m\times 1, Mm+1(m):1×1M^{(m)}_{m+1}:1\times 1.

In the block Mj(m)M^{(m)}_{j}, 0j<(m+2)/20\leq j<(m+2)/2, fill the columns, from left to right, by never increasing sequences of binary numbers (read from top to bottom) where each such number contains jj entries 11 for the block Mj(m)M^{(m)}_{j}. Put M0(m)=(0)M^{(m)}_{0}=(0) in the left upper corner and M1(m)M^{(m)}_{1} to the left of it. For j2j\geq 2, put the block Mj(m)M^{(m)}_{j} below and to the right of the block Mj1(m)M^{(m)}_{j-1}, in NN. All entries in NN outside these blocks are zero. Determine the entries of the blocks Mj(m)M^{(m)}_{j}, m+1j(m+2)/2m+1\geq j\geq(m+2)/2, by the requirement that N=NN^{\bot}=N, where NN^{\bot} is obtained from NN by reflection along the diagonal which goes from the lower left to the upper right. We write Mi(m)=Mm+1i(m){M^{(m)}_{i}}^{\bot}=M^{(m)}_{m+1-i}. For odd integer mm, we have M(m+1)/2(m)=M(m+1)/2+1(m){M^{(m)}_{(m+1)/2}}^{\bot}=M^{(m)}_{(m+1)/2+1}, by construction. Here are Mj(3)M^{(3)}_{j} and N,N2,N3N,N^{2},N^{3} for m=3m=3:

(8.10) M0(3)=(0),M1(3)=(1,1,1),M2(3)=(110101011),M3(3)=(111),M4(3)=(0).M^{(3)}_{0}=(0),M^{(3)}_{1}=(1,1,1),M^{(3)}_{2}=\left(\begin{array}[]{ccc}1&1&0\\ 1&0&1\\ 0&1&1\\ \end{array}\right),M^{(3)}_{3}=\left(\begin{array}[]{c}1\\ 1\\ 1\\ \end{array}\right),M^{(3)}_{4}=(0).
(8.19) N(3)\displaystyle N^{(3)} =\displaystyle= (0|1¯11¯|00000000|1¯10¯|00000|101|00000|0¯11¯|00000000|1¯0000000|10000000|1¯00000000),\displaystyle\left(\begin{array}[]{cccccccc}0&|\underline{{\textbf{1}}}&{\textbf{1}}&\underline{{\textbf{1}}}|&0&0&0&0\\ 0&0&0&0&|\overline{{\textbf{1}}}&{\textbf{1}}&\overline{{\textbf{0}}}|&0\\ 0&0&0&0&|{\textbf{1}}&{\textbf{0}}&{\textbf{1}}|&0\\ 0&0&0&0&|\underline{{\textbf{0}}}&{\textbf{1}}&\underline{{\textbf{1}}}|&0\\ 0&0&0&0&0&0&0&|\overline{{\textbf{1}}}\\ 0&0&0&0&0&0&0&|{\textbf{1}}\\ 0&0&0&0&0&0&0&|\underline{{\textbf{1}}}\\ 0&0&0&0&0&0&0&0\\ \end{array}\right),
(8.28) N(3)2\displaystyle{N^{(3)}}^{2} =\displaystyle= (0000222000000002000000020000000200000000000000000000000000000000),\displaystyle\left(\begin{array}[]{cccccccc}0&0&0&0&2&2&2&0\\ 0&0&0&0&0&0&0&2\\ 0&0&0&0&0&0&0&2\\ 0&0&0&0&0&0&0&2\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0\\ \end{array}\right),
(8.37) N(3)3\displaystyle{N^{(3)}}^{3} =\displaystyle= (0000000600000000000000000000000000000000000000000000000000000000).\displaystyle\left(\begin{array}[]{cccccccc}0&0&0&0&0&0&0&6\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0\\ 0&0&0&0&0&0&0&0\\ \end{array}\right).

We can now write, for 1jm1\leq j\leq m,

(8.38) Nj=j!nj(m),N^{j}=j!n^{(m)}_{j},

where the matrix nj(m)M(0,1)n^{(m)}_{j}\in M(0,1), by construction. Hence

(8.39) exp{LN(m)}=j=0m(L)jj!N(m)j=j=0m(L)jnj(m).\exp{\left\{-LN^{(m)}\right\}}=\sum_{j=0}^{m}\frac{(-L)^{j}}{j!}{N^{(m)}}^{j}=\sum_{j=0}^{m}(-L)^{j}n^{(m)}_{j}.

This is obvious from the set-up above. Furthermore, directly from construction, nj(m)n^{(m)}_{j}, j1j\geq 1, has a block structure into blocks of size

(8.40) (1×m),, j1 middle blocks missing,,(m×1),(1\times m),\cdots,\underbrace{\cdots}_{\textrm{ $j-1$ middle blocks missing}},\cdots,(m\times 1),

located in the uppermost right corner of size 2mj+1×2mj+12^{m-j+1}\times 2^{m-j+1} as in the above example.

8.3. Math:The Matrix NN

In this section we compute the matrix NN which gives the log of the monodromy. Because of orientations, the answer we get is the negative of the physical NN computed in the previous section.

Our basic result gives the monodromy

(8.41) m(σ1)=I(1)pτI=σ1+I,p1(1)pτI.m(\sigma_{1})=\sum_{I}(-1)^{p}\tau_{I}=\sigma_{1}+\sum_{I,\ p\geq 1}(-1)^{p}\tau_{I}.

Here we have changed notation. I={i1,,ip}I=\{i_{1},\dotsc,i_{p}\} refers to a flag Γi1ΓipΓ\Gamma_{i_{1}}\subsetneq\cdots\subsetneq\Gamma_{i_{p}}\subsetneq\Gamma of core subgraphs. More generally

(8.42) m(τI)=JI(1)qpτJ.m(\tau_{I})=\sum_{J\supset I}(-1)^{q-p}\tau_{J}.

Here J={j1,,jq}IJ=\{j_{1},\dotsc,j_{q}\}\supset I. to verify (8.42), consider e.g. the case corresponding to Γ1Γ\Gamma_{1}\subsetneq\Gamma. We have seen (lemma 3.3) that the blowup of (Γ){\mathbb{P}}(\Gamma) along the linear space defined by the edge variables associated to edges of Γ1\Gamma_{1} yields as exceptional divisor E1(Γ1)×(Γ//Γ1)E_{1}\cong{\mathbb{P}}(\Gamma_{1})\times{\mathbb{P}}(\Gamma/\!/\Gamma_{1}). In fact, the strict transform of E1E_{1} in the full blowup P(Γ)P(\Gamma) can be identified with P(Γ1)×P(Γ//Γ1)P(\Gamma_{1})\times P(\Gamma/\!/\Gamma_{1}). To see this, note that by proposition 3.4, the intersection in P(Γ)P(\Gamma) of distinct exceptional divisors E1EpE_{1}\cap\cdots\cap E_{p} is non-empty if and only if after reordering, the corresponding core subgraphs of Γ\Gamma form a flag. This means, for example, that E1EIE_{1}\cap E_{I}\neq\emptyset if and only if the flag corresponding to II has a subflag of core subgraphs contained in Γ1\Gamma_{1}, and the remaining core subgraphs form a flag containing Γ1\Gamma_{1}. In this way, we blow up appropriate linear spaces in (Γ1){\mathbb{P}}(\Gamma_{1}) and in (Γ/Γ1){\mathbb{P}}(\Gamma\!/\Gamma_{1}). the result is P(Γ1)×P(Γ//Γ1)P(Γ)P(\Gamma_{1})\times P(\Gamma/\!/\Gamma_{1})\subset P(\Gamma). The chain τ1\tau_{1} is an S1S^{1}-bundle over the chain σ(Γ1)×σ(Γ//Γ1)\sigma_{{\mathbb{P}}(\Gamma_{1})}\times\sigma_{{\mathbb{P}}(\Gamma/\!/\Gamma_{1})} (slightly modified along the boundaries as above), and the monodromy map is the product of the monodromies on each factor. (The monodromy takes place on P(Γ1)×P(Γ//Γ1)P(\Gamma_{1})\times P(\Gamma/\!/\Gamma_{1}). In the end, one takes the S1S^{1}-bundle over m(σ(Γ1)×σ(Γ//Γ1))m(\sigma_{{\mathbb{P}}(\Gamma_{1})}\times\sigma_{{\mathbb{P}}(\Gamma/\!/\Gamma_{1})}).) But this yields exactly (8.42). The result for a general m(τI)m(\tau_{I}) is precisely analogous. To compute NN, suppose Γ\Gamma has exactly kk core subgraphs ΓΓ\Gamma^{\prime}\subsetneq\Gamma. (This means that P(Γ)P(\Gamma) will have kk exceptional divisors EiE_{i}.) Consider the commutative ring

(8.43) R:=[x1,,xk]/(x12,,xk2,M1,,Mr),R:={\mathbb{Q}}[x_{1},\dotsc,x_{k}]/(x_{1}^{2},\dotsc,x_{k}^{2},M_{1},\dotsc,M_{r}),

where we think of the xix_{i} as corresponding to exceptional divisors EiE_{i} on P(Γ)P(\Gamma), and the MjM_{j} are monomials corresponding to empty intersections of the EiE_{i}. The notation means that we factor the polynomial ring in the xix_{i} by the ideal generated by the indicated elements. We may if we like drop the MjM_{j} from the ideal. This will simply mean the column vector on which NN acts will have many entries equal to 0. As a vector space, we can identify RR with the free vector space on σ1\sigma_{1} and the τI\tau_{I} by mapping σ11\sigma_{1}\mapsto 1 and τIiIxi\tau_{I}\mapsto\prod_{i\in I}x_{i}. With this identification, the monodromy map mm is given (compare (8.42)) by multiplication by (1x1)(1x2)(1xk)(1-x_{1})(1-x_{2})\cdots(1-x_{k}). But the map REndvec. sp.(R)R\to\text{End}_{\text{vec. sp.}}(R) given by multiplication is a homomorphism of rings, so log(m)\log(m) is given by (note xi2=0x_{i}^{2}=0)

(8.44) log((1x1)(1xk))=xi.\log\Big{(}(1-x_{1})\cdots(1-x_{k})\Big{)}=-\sum x_{i}.

Thus NN is the matrix for the map given by multiplication by xi-\sum x_{i}. If we ignore the relations MjM_{j} and just write the matrix for the action on [x1,,xk]/(x12,,xk2){\mathbb{Q}}[x_{1},\dotsc,x_{k}]/(x_{1}^{2},\dotsc,x_{k}^{2}), it has size 2k×2k2^{k}\times 2^{k} and is strictly upper triangular. For k=3k=3, the matrix is N(3)-N^{(3)} (8.19).

9. Renormalization: the removal of log-poles

Recall we have defined sdd(Γ)\textrm{sdd}(\Gamma), the degree of superficial divergence of a graph with respect to a given physical theory, (2.2). The choice of the theory determines a differential form ωΓ\omega_{\Gamma} associated to Γ\Gamma. We will be interested in the logarithmic divergent case, when sdd(Γ)0\textrm{sdd}(\Gamma)\geq 0, but ωΓ\omega_{\Gamma} has been chosen such that it only has log-poles, see in particular section 7.4. The affine integral in this case will be overall logarithmically divergent, but this overall divergence can be eliminated by passing to the associated projective integral. If, for all core subgraphs ΓΓ\Gamma^{\prime}\subset\Gamma, we have sdd(Γ)<0\textrm{sdd}(\Gamma^{\prime})<0, then the projective integral actually converges and we are done. If Γ>0\Gamma^{\prime}>0 for some subgraph, then one is obliged to manipulate the differential form as described in section 7 above. To simplify notation, from now on we assume that all graphs and subgraphs have sdd0\textrm{sdd}\leq 0, while all following lemmas hold similarly for higher degrees of divergence with the appropriate choice of ωII\omega_{II}. Below, we spell all results out for the case ωII=Ω2n1/ψΓ2\omega_{II}=\Omega_{2n-1}/\psi_{\Gamma}^{2}, and we set ψΓψ(Γ)\psi_{\Gamma}\equiv\psi(\Gamma).

Lemma 9.1.

Let ΓΓ\Gamma^{\prime}\subsetneq\Gamma be core graphs and assume sdd(Γ)=0\textrm{sdd}(\Gamma)=0. Let LXΓ(Γ)L\subset X_{\Gamma}\subset{\mathbb{P}}(\Gamma) be the coordinate linear space defined by the edges occurring in Γ\Gamma^{\prime}. Let π:PL(Γ)\pi:P_{L}\to{\mathbb{P}}(\Gamma) be the blowup of LL. Then πωΓ\pi^{*}\omega_{\Gamma} has a logarithmic pole on EE if and only if sdd(Γ)=0\textrm{sdd}(\Gamma^{\prime})=0. Similarly, the pullback of ωΓ\omega_{\Gamma} to the full core blowup P(Γ)P(\Gamma) (cf. formula (3.3)) has a log pole of order along the exceptional divisor EΓE_{\Gamma^{\prime}} associated to Γ\Gamma^{\prime} if and only if sdd(Γ)=0\textrm{sdd}(\Gamma^{\prime})=0.

Proof.

We give the proof for ϕ4\phi^{4}-theory. Let the loop number |Γ|=m|\Gamma|=m so the graph has 2m2m edges (2.2). Let

(9.1) Ω2m1=(1)iAidA1dAi^dA2m=A2m2md(A1/A2m)d(A2m1/A2m).\Omega_{2m-1}=\sum(-1)^{i}A_{i}dA_{1}\wedge\cdots\wedge\widehat{dA_{i}}\wedge\cdots\wedge dA_{2m}=A_{2m}^{2m}d(A_{1}/A_{2m})\wedge\cdots\wedge d(A_{2m-1}/A_{2m}).

Then

(9.2) ωΓ=Ω2m1ψΓ2.\omega_{\Gamma}=\frac{\Omega_{2m-1}}{\psi_{\Gamma}^{2}}.

Suppose L:A1==Ap=0L:A_{1}=\cdots=A_{p}=0. We can write the graph polynomial ([2], prop. 3.5)

(9.3) ψΓ=ψΓ(A1,,Ap)ψΓ//Γ(Ap+1,,A2m)+R\psi_{\Gamma}=\psi_{\Gamma^{\prime}}(A_{1},\dotsc,A_{p})\psi_{\Gamma/\!/\Gamma^{\prime}}(A_{p+1},\dotsc,A_{2m})+R

where the degree of RR in A1,,ApA_{1},\dotsc,A_{p} is strictly greater than degψΓ=|Γ|\deg\psi_{\Gamma^{\prime}}=|\Gamma^{\prime}|. Let ai=Ai/A2ma_{i}=A_{i}/A_{2m}, and let bi=ai/ap,i<pb_{i}=a_{i}/a_{p},\ i<p. Locally on PP we can take b1,,bp1,ap,ap+1,,A2mb_{1},\dotsc,b_{p-1},a_{p},a_{p+1},\dotsc,A_{2m} as local coordinates and write

(9.4) ωΓ=±app2|Γ|dapapdb1da2m1F2.\omega_{\Gamma}=\pm a_{p}^{p-2|\Gamma^{\prime}|}\frac{da_{p}}{a_{p}}\wedge\frac{db_{1}\wedge\cdots\wedge da_{2m-1}}{F^{2}}.

Here FF is some polynomial in the aia_{i}’s and the bjb_{j}’s which is not divisible by apa_{p}. The assertion for the blowup of LL follows immediately. The assertion for P(Γ)P(\Gamma) is also clear because we can find a non-empty open set on (Γ){\mathbb{P}}(\Gamma) meeting LL such that the inverse images in P(Γ)P(\Gamma) and in PLP_{L} are isomorphic. ∎

We want to state the basic renormalization result coming out of our monodromy method. For this, we restrict to the case

(9.5) sdd(Γ)2,ΓΓ,\textrm{sdd}(\Gamma^{\prime})\leq 2,\ \forall\Gamma^{\prime}\subseteq\Gamma,

with an understanding that appropriate forms ωII(Γ)\omega_{II}(\Gamma^{\prime}) have been chosen so that the differential forms has log-poles only. The following lemma applies then to ϕ4\phi^{4}-theory. A physicist wishing to apply our results to another theory needs only check the lemma holds with ωΓ\omega_{\Gamma} replaced by the integrand given by Feynman rules.

Lemma 9.2.

Let τVε\tau^{\varepsilon}_{V} be the chains on (Γ){\mathbb{P}}(\Gamma) constructed above (section 4) (including the case τP(Γ)ε=σε\tau^{\varepsilon}_{P(\Gamma)}=\sigma_{\varepsilon}). Then, assuming (9.5), we will have

(9.6) |τVεωΓ|=O(|log|ε||k),|ε|0\Big{|}\int_{\tau^{\varepsilon}_{V}}\omega_{\Gamma}\Big{|}=O(|\log|\varepsilon||^{k}),\ |\varepsilon|\to 0

for some k0k\geq 0.

Proof.

We first consider the integral for the chain σε=τP(Γ)ε\sigma_{\varepsilon}=\tau_{P(\Gamma)}^{\varepsilon}. Locally on the blowup P(Γ)P(\Gamma) the integrand will look like (9.4) but there may be more than one log form; i.e. ω~dap1/ap1dapk/apk\widetilde{\omega}da_{p_{1}}/a_{p_{1}}\wedge\cdots\wedge da_{p_{k}}/a_{p_{k}}. An easy estimate for such an integral over a compact chain satisfying ajεa_{j}\geq\varepsilon gives C(|logε|)kC(|\log\varepsilon|)^{k}. The integrals over τVε,VP(Γ)\tau_{V}^{\varepsilon},\ V\subsetneq P(\Gamma) involve first integrating over one or more circles. Locally the chain is an (S1)p(S^{1})^{p}-bundle over an intersection x1==xp=0x_{1}=\cdots=x_{p}=0 in local coordinates. We may compute the integral by first taking residues. VV will be the closure of a torus orbit in P(Γ)P(\Gamma) associated to a flag ΓpΓ1Γ\Gamma_{p}\subsetneq\cdots\subsetneq\Gamma_{1}\subsetneq\Gamma (proposition 3.4). We may assume xix_{i} is a local equation for the exceptional divisor in P(Γ)P(\Gamma) associated to ΓiΓ\Gamma_{i}\subset\Gamma. By lemma 9.1, our integrand will have a pole on xi=0x_{i}=0 if and only if sdd(Γi)=0\textrm{sdd}(\Gamma_{i})=0. (Note that the integrand has no singularities on τVε\tau_{V}^{\varepsilon}, so we may integrate in any order.) The situation is confusing because sdd(Γi)<0sdd(Γ//Γi)>0\textrm{sdd}(\Gamma_{i})<0\Rightarrow\textrm{sdd}(\Gamma/\!/\Gamma_{i})>0 so one might expect non-log growth in this case. The problem does not arise, because the residue will vanish. Assuming sdd(Γi)=0,i\textrm{sdd}(\Gamma_{i})=0,\ \forall i, the residue integral is

(9.7) jτP(Γj//Γj+1)εωΓpωΓ//Γ1.\int_{\prod_{j}\tau_{P(\Gamma_{j}/\!/\Gamma_{j+1})}^{\varepsilon}}\omega_{\Gamma_{p}}\wedge\cdots\wedge\omega_{\Gamma/\!/\Gamma_{1}}.

Since sdd(Γi//Γi+1)=0\textrm{sdd}(\Gamma_{i}/\!/\Gamma_{i+1})=0, we may simply write (9.7) as a product of integrals and argue as above. ∎

We want now to apply the argument sketched in the introduction to our situation. There is one mathematical point which must be dealt with first. We want to consider σtωΓ\int_{\sigma_{t}}\omega_{\Gamma} as a function of tt. Here we must be a bit careful. For t=εeiθt=\varepsilon e^{i\theta} and |θ|<<1|\theta|<<1 we are ok, but as θ\theta grows, our chain may meet XΓX_{\Gamma}. Topologically, we have (proposition 6.3) the chains c~η,ε,θ\tilde{c}^{\eta,\varepsilon,\theta} which miss XΓX_{\Gamma} and which represent the correct homology class in H((Γ)XΓ,ΔtXΓΔt)H_{*}({\mathbb{P}}(\Gamma)-X_{\Gamma},\Delta_{t}-X_{\Gamma}\cap\Delta_{t}), but one must show our integral depends only on the class in homology relative to Δt\Delta_{t}, i.e. ωΓ\omega_{\Gamma} integrates to zero over any chain on ΔtXΓΔt\Delta_{t}-X_{\Gamma}\cap\Delta_{t}. Intuitively, this is because ωΓ|Δt=0\omega_{\Gamma}|\Delta_{t}=0, but, because Δt\Delta_{t} has singularities it is best to be more precise. Quite generally, assume UU is a smooth variety of dimension rr, and DUD\subset U is a normal crossings divisor (i.e. for any point uUu\in U there exist local coordinates x1,,xrx_{1},\dotsc,x_{r} near uu, and prp\leq r such that D:x1x2xp=0D:x_{1}x_{2}\cdots x_{p}=0 near uu). One has sheaves

(9.8) ΩUq(logD)(D)ΩUqΩUq(logD)\Omega^{q}_{U}(\log D)(-D)\subset\Omega^{q}_{U}\subset\Omega^{q}_{U}(\log D)

where ΩUq\Omega^{q}_{U} is the sheaf of algebraic (or complex analytic; in fact, either will work here) qq-forms on XX, and ΩUq(logD)\Omega^{q}_{U}(\log D) is obtained by adjoining locally wedges of differential forms dxi/xi, 1ipdx_{i}/x_{i},\ 1\leq i\leq p. Locally, ΩUq(logD)(D):=x1x2xpΩUq(logD)\Omega^{q}_{U}(\log D)(-D):=x_{1}x_{2}\cdots x_{p}\Omega^{q}_{U}(\log D). All three sheaves are easily seen to be stable under exterior differential (for varying qq). The resulting complexes calculate the de Rham cohomology for (U,D),U,(UD)(U,D),U,(U-D) respectively, [7]. Note that in the top degree r=dimUr=\dim U we have

(9.9) ΩUr(logD)(D)=𝒪Ux1x2xpdx1dxpx1x2xpdxp+1dxr=ΩUr.\Omega^{r}_{U}(\log D)(-D)={\mathcal{O}}_{U}\cdot x_{1}x_{2}\cdots x_{p}\frac{dx_{1}\wedge\cdots\wedge dx_{p}}{x_{1}x_{2}\cdots x_{p}}dx_{p+1}\wedge\cdots\wedge dx_{r}=\Omega^{r}_{U}.

It follows that we get a maps

(9.10) ΩUr[r]ΩU(logD)(D);Γ(U,ΩUr)HDRr(U,D).\Omega^{r}_{U}[-r]\to\Omega^{*}_{U}(\log D)(-D);\quad\Gamma(U,\Omega^{r}_{U})\to H^{r}_{DR}(U,D).

In particular, taking U=(Γ)XΓU={\mathbb{P}}(\Gamma)-X_{\Gamma}, we see that integrals ch.rel.ΔtωΓ\int_{\text{ch.rel.}\Delta_{t}}\omega_{\Gamma} are well-defined.

Theorem 9.3.

We suppose given a graph Γ\Gamma such that all core subgraphs ΓΓ\Gamma^{\prime}\subseteq\Gamma have superficial divergence sdd(Γ)0\textrm{sdd}(\Gamma^{\prime})\leq 0 for a given physical theory. Let ωΓ\omega_{\Gamma} be the form associated to the given theory. Let NN be the upper-triangular matrix of size K×KK\times K described in the previous section, where KK is the number of chains of core subgraphs

ΓpΓ.\Gamma_{p}\subsetneq\cdots\subsetneq\Gamma.

Then the lefthand side of the expression below is single-valued and analytic for tt in a disk about 0 so the limit

(9.11) lim|t|0exp(Nlogt2πi)(τP(Γ)tωΓτVtωΓ)=(a1ak)\lim_{|t|\to 0}\exp(-N\frac{\log t}{2\pi i})\begin{pmatrix}\int_{\tau^{t}_{P(\Gamma)}}\omega_{\Gamma}\\ \vdots\\ \int_{\tau^{t}_{V}}\omega_{\Gamma}\\ \vdots\end{pmatrix}=\begin{pmatrix}a_{1}\\ \vdots\\ a_{k}\end{pmatrix}

exists.

Proof.

The proof proceeds as outlined in section 1.3. NN is chosen to be nilpotent and such that the lefthand side has no monodromy. The lemma 9.2 assures that terms have at worst log growth. Since they are single-valued on DD^{*}, they extend to the origin. ∎

Remark 9.4.

It is time to compare what we are calculating here with what a physicist computes according to Thm.(7.3). The transition is understood upon noticing that in our constructions of chains, we pick up the residue from each exceptional divisor by computing the monodromy. In physics we iterate those residues as iterated integrals. Below the top entry a1a_{1} this gives different rational weights to them in according with the scattering type formula of [6]. We discuss this below in section (9.1).

Definition 9.5.

With notation as above, the renormalized value σωΓ\int_{\sigma}\omega_{\Gamma} is the top entry in the column vector exp(+Nlogt2πi)(a1)\exp(+N\frac{\log t}{2\pi i})\begin{pmatrix}a_{1}\\ \vdots\end{pmatrix}.

Remark 9.6.

Note that the terms τVtωΓ\int_{\tau^{t}_{V}}\omega_{\Gamma} on the lefthand side of (9.11) may be calculated recursively. As in lemma 9.2 above, VV corresponds to a flag of core subgraphs of Γ\Gamma. As in formula (9.7), the integral dies unless all the Γi//Γi+1\Gamma_{i}/\!/\Gamma_{i+1} are log divergent. In this case, one gets

(9.12) (2πi)p1τP(Γi//Γi+1)tωΓi//Γi+1.(2\pi i)^{p-1}\prod\int_{\tau^{t}_{P(\Gamma_{i}/\!/\Gamma_{i+1})}}\omega_{\Gamma_{i}/\!/\Gamma_{i+1}}.

If, in addition, the subquotients Γi//Γi+1\Gamma_{i}/\!/\Gamma_{i+1} are primitive, i.e. they are log divergent but have no divergent subgraphs, then the integrals in (9.12) will converge as |t|0|t|\to 0. Upto a term which is O(t)O(t) and can be ignored in the limit, they may be replaced by their limits as t0t\to 0. These entries in (9.11) may then be taken to be constant.

Example 9.7.

Consider the dunce’s cap fig.(2). It has 33 core subgraphs, but only the 22-edged graph γ\gamma with edges 1,21,2 is log divergent. Thus, the column vector in (9.11) has 44 entries, but only 22 are non-zero. Dropping unnecessary rows and columns, the matrix N=(0100)N=\begin{pmatrix}0&-1\\ 0&0\end{pmatrix}. The constant entry in the column vector is

(9.13) 2πiσγΩ1ψγ2σΓ//γΩ1ψΓ//γ2=2πi(0da(a+1)2)2=2πi.2\pi i\int_{\sigma_{\gamma}}\frac{\Omega_{1}}{\psi_{\gamma}^{2}}\int_{\sigma_{\Gamma/\!/\gamma}}\frac{\Omega_{1}}{\psi_{\Gamma/\!/\gamma}^{2}}=2\pi i\Big{(}\int_{0}^{\infty}\frac{da}{(a+1)^{2}}\Big{)}^{2}=2\pi i.

It remains to connect σωΓ\int_{\sigma}\omega_{\Gamma} to the physicists computation.

9.1. lMHS vs ΦR\Phi_{R}

Let us understand how the period matrix pT=(a1,a2,,ar)p^{T}=(a_{1},a_{2},\cdots,a_{r}) which we have constructed connects to the coefficients cjc_{j}

(9.14) ΦMOM(Γ)(q2/μ2)=j=1rcj(Γ)lnjq2/μ2.\Phi_{\textrm{MOM}}(\Gamma)(q^{2}/\mu^{2})=\sum_{j=1}^{r}c_{j}(\Gamma)\ln^{j}q^{2}/\mu^{2}.

Going to variables

tΓ,a1,,a|Γ[1]|,ai=1,\displaystyle t_{\Gamma},a_{1},\ldots,a_{|\Gamma^{[1]}|},\sum a_{i}=1,
t1,b1,,b|Γ1[1]|,bi=1,\displaystyle t_{1},b_{1},\ldots,b_{|{\Gamma_{1}}^{[1]}|},\sum b_{i}=1,
,\displaystyle\ldots,
tp,z1,,b|Γp[1]|,zi=1,\displaystyle t_{p},z_{1},\ldots,b_{|{\Gamma_{p}}^{[1]}|},\sum z_{i}=1,

for a chain of core graphs ΓpΓ1Γ\Gamma_{p}\subsetneq\cdots\subsetneq\Gamma_{1}\subsetneq\Gamma gives, for each such flag and constant lower boundaries ϵ\epsilon, an iterated integral over

(9.15) ϵ𝑑tϵ/t𝑑t1ϵ/t/t1/tp1𝑑tp.\int_{\epsilon}^{\infty}dt\int_{\epsilon/t}^{\infty}dt_{1}\cdots\int_{\epsilon/t/t_{1}\cdots/t_{p-1}}^{\infty}dt_{p}.

As the integral has a logarithmic pole along any tit_{i} integration, the difference between integrating against the chains, which only collect the coefficients of lnϵ\ln\epsilon for each such integral, and the iteration above is a factorial for each flag. A summation over all flags established the desired relation using tree factorials [14]:
As the entries in the vector (a1,)T(a_{1},\cdots)^{T} are in one-to-one correspondence with forests of Γ\Gamma, identifying a1a_{1} with the empty forest, we can write the top-entry defined in Defn.(9.5) as

(9.16) [for](lnt2πi)|[for]|a[for],\sum_{\textrm{[for]}}\left(\frac{\ln t}{2\pi i}\right)^{|\textrm{[for]}|}a_{\textrm{[for]}},

where

(9.17) a[for]=p1(Γ//[for])jp1(γj),a_{\textrm{[for]}}=p_{1}(\Gamma/\!/\textrm{[for]})\prod_{j}p_{1}(\gamma_{j}),

using the notation of Eqs.(7.14,7.62). Then,

(9.18) lntΦMOM(Γ)(t)=[for]aug(Γ)(lnt[for]!)|[for]|a[for].\partial_{\ln t}\Phi_{\textrm{MOM}}(\Gamma)(t)=\sum_{\textrm{[for]}}\textrm{aug}(\Gamma)\left(\frac{\ln t}{\textrm{[for]}^{*}!}\right)^{|\textrm{[for]}|}a_{\textrm{[for]}}.

Here, [for]!\textrm{[for]}^{*}! is a forest factorial defined as follows. Any forest [for] defines a tree TT and a collection of edges CC such that PC(T)P^{C}(T) and RC(T)R^{C}(T) denote the core sub- and co-graphs in question. The complement set T[1]/CT^{[1]}/C defines a forest iti\cup_{i}t_{i} say. We set [for]!=iti!\textrm{[for]}^{*}!=\prod_{i}t_{i}!, for standard tree factorials ti!t_{i}! [14]. For example, comparing the two graphs

(9.19) Γ1=[Uncaptioned image],Γ2=[Uncaptioned image],\Gamma_{1}=\;\raisebox{-8.53581pt}{\epsfbox{grapha.eps}}\;,\;\Gamma_{2}=\;\raisebox{-8.53581pt}{\epsfbox{graphb.eps}}\;,

we have the two vectors

(9.20) (p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image]))\left(\begin{array}[]{l}p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{grapha.eps}}\;\right)\\ p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\right)p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphd.eps}}\;\right)\\ p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphd.eps}}\;\right)p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\right)\\ p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\right)p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\right)p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\right)\end{array}\right)

and

(9.21) (p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])p1([Uncaptioned image])).\left(\begin{array}[]{l}p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphb.eps}}\;\right)\\ p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\right)p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphd.eps}}\;\right)\\ p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\right)p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphd.eps}}\;\right)\\ p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\right)p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\right)p_{1}\left(\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\right)\end{array}\right).

Hence, we find the same ln2t\ln^{2}t term upon computing Eq.(9.16) for the monodromy.

On the other hand, the tree factorials deliver 1/2 for that term in the case of Γ1\Gamma_{1}, and 11 for Γ2\Gamma_{2}, while we get 22 in both cases for the term lnt\sim\ln t. Indeed, the flag

(9.22) [Uncaptioned image][Uncaptioned image][Uncaptioned image]\;\raisebox{-8.53581pt}{\epsfbox{graphc.eps}}\;\subsetneq\;\raisebox{-8.53581pt}{\epsfbox{graphd.eps}}\;\subsetneq\;\raisebox{-8.53581pt}{\epsfbox{grapha.eps}}\;

corresponds to a tree with two edges. The term ln2t\sim\ln^{2}t comes from the cut CC which corresponds to both of these edges. The complement is the empty cut, whose tree factorial is 3!3! simply. As we took a derivative with respect to lnt\ln t, we get a factor of aug(Γ)=3\textrm{aug}(\Gamma)=3, which leaves us with a factor 3/3!=1/23/3!=1/2.

For  [Uncaptioned image] , we note that the tree factorial is 33 instead of 3!3! (we have two flags instead of one), which leaves us with a factor 11.

9.2. Limiting Mixed Hodge Structures

In this final paragraph at the suggestion of the referee we outline the structure of a limiting mixed Hodge structure associated to a variation of mixed Hodge structure and how it might apply to the Feynman graph amplitudes.

Let Γ\Gamma be a log divergent graph with nn loops and 2n2n edges. The graph hypersurface XΓ:ψΓ=0X_{\Gamma}:\psi_{\Gamma}=0 is a hypersurface in 2n1{\mathbb{P}}^{2n-1}, and the Feynman integrand represents a cohomology class

(9.23) [ΩψΓ2]H2n1(2n1XΓ,)=H2n1(2n1XΓ,)=H=H.\Big{[}\frac{\Omega}{\psi_{\Gamma}^{2}}\Big{]}\in H^{2n-1}({\mathbb{P}}^{2n-1}-X_{\Gamma},{\mathbb{C}})=H^{2n-1}({\mathbb{P}}^{2n-1}-X_{\Gamma},{\mathbb{Q}})\otimes{\mathbb{C}}=H_{\mathbb{C}}=H_{\mathbb{Q}}\otimes{\mathbb{C}}.

The cohomology group has a mixed Hodge structure, which means there are defined two filtrations:
(i) The weight filtration WHW_{*}H_{\mathbb{Q}} which is defined over {\mathbb{Q}} and increasing. It looks like

(9.24) 0W2nHW2n+1HW4n2H=H.0\subset W_{2n}H_{\mathbb{Q}}\subset W_{2n+1}H_{\mathbb{Q}}\subset\cdots\subset W_{4n-2}H_{\mathbb{Q}}=H_{\mathbb{Q}}.

Blowing up on XΓX_{\Gamma} so it becomes a normal crossings divisor DD_{*}, there is a spectral sequence relating the graded pieces W2n1+i/W2n1+i1W_{2n-1+i}/W_{2n-1+i-1} to the Tate twist by i-i of the cohomology in degree 2n1i2n-1-i of the codimension i1i-1 strata of DD. (So, for example, gr2nWgr^{W}_{2n} is related to jH2n2(Dj)(1)\oplus_{j}H^{2n-2}(D_{j})(-1) where D=DjD=\bigcup D_{j}.)
(ii) The Hodge filtration FHF^{*}H_{\mathbb{C}} which is defined over {\mathbb{C}} and decreasing:

(9.25) (0)F2n1F2n2F1=H.(0)\subset F^{2n-1}\subset F^{2n-2}\subset\cdots\subset F^{1}=H_{\mathbb{C}}.

The filtrations are subject to the compatibility condition that the filtration

(9.26) Fp(grqW):=FpHWq/FpHWq1F^{p}(gr^{W}_{q}\otimes{\mathbb{C}}):=F^{p}H_{\mathbb{C}}\cap W_{q}\otimes{\mathbb{C}}\Big{/}F^{p}H_{\mathbb{C}}\cap W_{q-1}\otimes{\mathbb{C}}

is the Hodge filtration of a pure Hodge structure of weight qq. (This is simply the condition that FgrqWF^{*}gr^{W}_{q}\otimes{\mathbb{C}} be qq-opposite to its complex conjugate, i.e. that grqW=FpF¯qp+1gr^{W}_{q}\otimes{\mathbb{C}}=F^{p}\oplus\overline{F}^{q-p+1} for any pp.)

Let us say that a class ωH\omega\in H_{\mathbb{C}} has Hodge level pp if ωFpHFp+1H\omega\in F^{p}H_{\mathbb{C}}-F^{p+1}H_{\mathbb{C}}. An important problem is to determine the Hodge level of the Feynman form (9.23). One may speculate that the Hodge level of the Feynman form equals the transcendental weight of the period. (The transcendental weight of a multizeta number ζ(n1,,np)\zeta(n_{1},\dotsc,n_{p}) is the sum of the nin_{i}.) For example, in [4] one finds many examples of Feynman amplitudes of the form ζ(N)*\zeta(N) where * is rational. In all known cases N=2n3N=2n-3. To estimate the Hodge level, one may use the pole order filtration [7], 3.12. One blows up on XΓ2n1X_{\Gamma}\subset{\mathbb{P}}^{2n-1} to replace XX by a normal crossings divisor D=i=1rDiD=\bigcup_{i=1}^{r}D_{i}. Let ω\omega on 2n1XΓ{\mathbb{P}}^{2n-1}-X_{\Gamma} be a (2n1)(2n-1)-form and let I{1,,r}I\subset\{1,\dotsc,r\} be the indices ii such that ω\omega has a pole along DiD_{i}. Write pi+1p_{i}+1 for the order of this pole, with pi0p_{i}\geq 0. Then the Hodge level of ω\omega is 2n1pi\geq 2n-1-\sum p_{i}. (For a more precise statement, see op. cit.) For example, if XΓX_{\Gamma} is smooth (this happens only when n=1n=1) one would get p1=1p_{1}=1 so the Hodge level would be 2n2\geq 2n-2.

Proposition 9.8.

For the Feynman form, at least 22 of the pi1p_{i}\geq 1. The pole order calculation thus suggests the Hodge level of the Feynman form above is 2n3\leq 2n-3.

Proof.

The situation for n=1n=1 is trivial, so we assume n2n\geq 2. The space of symmetric n×nn\times n-matrices has dimension d:=n(n+1)2d:=\frac{n(n+1)}{2}. Let d1{\mathbb{P}}^{d-1} be viewed as the projectivized space of such matrices, so a point corresponds to a matrix upto scale. The determinant of the universal matrix defines a hypersurface 𝒳d1{\mathcal{X}}\subset{\mathbb{P}}^{d-1}. More generally, we define 𝒳pd1{\mathcal{X}}_{p}\subset{\mathbb{P}}^{d-1} to be the locus where the rank of the corresponding symmetric matrix is np\leq n-p. We have 𝒳=𝒳1{\mathcal{X}}={\mathcal{X}}_{1}, and it is easy to see that 𝒳p{\mathcal{X}}_{p} has codimension p(p+1)2\frac{p(p+1)}{2} in d1{\mathbb{P}}^{d-1}. Points in 𝒳p{\mathcal{X}}_{p} will have multiplicity p\geq p on 𝒳{\mathcal{X}}.

There is an inclusion ρ:2n1d1\rho:{\mathbb{P}}^{2n-1}\hookrightarrow{\mathbb{P}}^{d-1} such that XΓ=𝒳2n1X_{\Gamma}={\mathcal{X}}\cap{\mathbb{P}}^{2n-1}. Points of 𝒳22n1{\mathcal{X}}_{2}\cap{\mathbb{P}}^{2n-1} will have multiplicity 2\geq 2 in XΓX_{\Gamma} and codimension 3\leq 3 in 2n1{\mathbb{P}}^{2n-1} . This means that in the local ring on 2n1{\mathbb{P}}^{2n-1} at a general point of 𝒳22n1{\mathcal{X}}_{2}\cap{\mathbb{P}}^{2n-1}, there will be functions x1,x2,x3x_{1},x_{2},x_{3} which form part of a system of coordinates on 2n1{\mathbb{P}}^{2n-1} such that a local defining equation ψ\psi for XΓX_{\Gamma} lies in (x1,x2,x3)2(x_{1},x_{2},x_{3})^{2}. We may construct our normal crossings divisor DD as above by first blowing up 𝒳22n1{\mathcal{X}}_{2}\cap{\mathbb{P}}^{2n-1} in 2n1{\mathbb{P}}^{2n-1}. Subsequent blowups will not affect the pole order, which may be computed at the generic point of the exceptional divisor EE. We have

(9.27) dx1dx2dx3ψ2=x12dx1d(x2/x1)d(x3/x1)x14ϕ(x1,x2/x1,x3/x1,).\frac{dx_{1}dx_{2}dx_{3}\cdots}{\psi^{2}}=\frac{x_{1}^{2}dx_{1}d(x_{2}/x_{1})d(x_{3}/x_{1})\cdots}{x_{1}^{4}\phi(x_{1},x_{2}/x_{1},x_{3}/x_{1},\ldots)}.

It follows that the Feynman form has a double pole on EE as well as a double pole on the strict transform of XΓX_{\Gamma} in the blowup. ∎

Remark 9.9.

(i) To give a complete proof that the Hodge level is 2n3\leq 2n-3 one would have to show the double order pole was not killed by an exact form.
(ii) It would be exciting to be able to say something about the weight filtration on H2n1(2n1XΓ)H^{2n-1}({\mathbb{P}}^{2n-1}-X_{\Gamma}).
(iii) The data in [4] suggests that double zetas which occur will have transcendental weight 2n42n-4. For example, the bipartite graph Γ\Gamma consisting of the 1212 edges joining sets of 33 and 44 vertices has Feynman amplitude a rational multiple of ζ(3,5)\zeta(3,5). In general, a calculation as above shows 𝒳32n1{\mathcal{X}}_{3}\cap{\mathbb{P}}^{2n-1} has multiplicity 3\geq 3 and codimension 6\leq 6. If one could show that for the bipartite Γ\Gamma that this codimension drops to 55, then the same argument as above would yield 33 poles with pi1p_{i}\geq 1, suggesting a Hodge level 2n42n-4.

Next we should consider the mixed Hodge structure necessary for the relative period calculation. Recall (3.3) we work in a toric blowup P=P(Γ)2n1P=P(\Gamma)\to{\mathbb{P}}^{2n-1}. Let BPB\subset P be the complement of the big toric orbit in PP. It is the union of the strict transform of the coordinate divisor Δ2n1\Delta\subset{\mathbb{P}}^{2n-1} and the exceptional divisors. Let YPY\subset P be the strict transform of XΓX_{\Gamma}. The relevant cohomology group is the middle group in the sequence

(9.28) H2n2(BYB;)H2n1(PY,BYB;)H2n1(PY,).H^{2n-2}(B-Y\cap B;{\mathbb{Q}})\to H^{2n-1}(P-Y,B-Y\cap B;{\mathbb{Q}})\to H^{2n-1}(P-Y,{\mathbb{Q}}).

If all the subgraphs ΓΓ\Gamma^{\prime}\subsetneq\Gamma have sdd(Γ)<0\text{sdd}(\Gamma^{\prime})<0, then renormalization is unnecessary. The Feynman amplitude as we have defined it is simply a period of the mixed Hodge structure (9.28). The weight filtration for the group on the left involves the cohomology of the strata of the normal crossings divisor BB. For example, we have an exact sequence

(9.29) H0(B(1)YB(1),)H0(B(0)YB(0),)W0H2n2(BYB(0),).H^{0}(B_{(1)}-Y\cap B_{(1)},{\mathbb{Q}})\to H^{0}(B_{(0)}-Y\cap B_{(0)},{\mathbb{Q}})\to W_{0}H^{2n-2}(B-Y\cap B_{(0)},{\mathbb{Q}}).

Here we write B(i)B_{(i)} for the disjoint union of the components of the strata of dimension ii. We know from corollary 5.3 that YB(0)=Y\cap B_{(0)}=\emptyset, and a bit of thought about the combinatorics of B(i),i=0,1B_{(i)},i=0,1 reveals that W0H2n2(BYB,)=(0)W_{0}H^{2n-2}(B-Y\cap B,{\mathbb{Q}})={\mathbb{Q}}(0). This gives a map of the trivial Hodge structure (0){\mathbb{Q}}(0) to our period motive:

(9.30) (0)H2n1(PY,BYB;).{\mathbb{Q}}(0)\to H^{2n-1}(P-Y,B-Y\cap B;{\mathbb{Q}}).

When the period is a rational multiple of ζ(2n3)\zeta(2n-3) we expect that there is a map of Hodge structures (32n)H2n1(2n1XΓ,){\mathbb{Q}}(3-2n)\to H^{2n-1}({\mathbb{P}}^{2n-1}-X_{\Gamma},{\mathbb{Q}}) and that the extension of (32n){\mathbb{Q}}(3-2n) by (0){\mathbb{Q}}(0) associated to ζ(2n3)\zeta(2n-3) is a subquotient of (9.28).

Finally the main focus of this paper has been the renormalization case when one or more proper subgraphs of Γ\Gamma has sdd=0\text{sdd}=0. In this case, the Feynman form will have a pole along one or more divisor in BB, so (9.28) is no longer the relevant Hodge structure. In this case, we work with the limiting mixed Hodge structure HlimH_{lim} associated to Ht:=H2n1(2n1XΓ,ΔtΔtXΓ)H_{t}:=H^{2n-1}({\mathbb{P}}^{2n-1}-X_{\Gamma},\Delta_{t}-\Delta_{t}\cap X_{\Gamma}). Let DD be a small disk around t=0t=0, and let D=D{0}D^{*}=D-\{0\}. Then HD=t0HtH_{D^{*}}=\bigcup_{t\neq 0}H_{t} becomes a local system on DD^{*}. Let D=HD𝒪D{\mathcal{H}}_{D^{*}}=H_{D^{*}}\otimes{\mathcal{O}}_{D^{*}} be the corresponding analytic bundle. If we untwist by the monodromy, we get a trivial local system (h=dimHth=\dim H_{t})

(9.31) Dhexp(Nlogt)HDD.{\mathbb{C}}^{h}_{D^{*}}\cong\exp(-N\log t)H_{D^{*}}\subset{\mathcal{H}}_{D^{*}}.

Since this local system is trivial, it extends (trivially) across t=0t=0. It also has a canonical {\mathbb{Q}}-structure defined from the {\mathbb{Q}}-structure at any point t00t_{0}\neq 0. The analytic bundle D{\mathcal{H}}_{D^{*}} has a Hodge filtration FDF^{*}{\mathcal{H}}_{D^{*}} coming from the Hodge filtrations on the HtH_{t}. (Note the Hodge filtration is not horizontal, so there is no Hodge filtration on the local system HDH_{D^{*}}.) From (9.31) we get a canonical trivialization of the analytic bundle D𝒪Dh{\mathcal{H}}_{D^{*}}\cong{\mathcal{O}}_{D^{*}}^{h} and hence a canonical extension across t=0t=0. One can show [5], 2.1(i) that the Hodge filtration extends across t=0t=0 as well.

Thus, on the fibre H0H_{0} we have a Hodge filtration and a {\mathbb{Q}}-structure. If you think in terms of periods, i.e. using the pairing H0,×H0H_{0,{\mathbb{Q}}}^{\vee}\times H_{0}\to{\mathbb{C}}, the above description of the Hodge filtration as a limit across t=0t=0 coincides with the computation (9.11). What we have not given is the weight filtration. This monodromy weight or limiting weight filtration is more subtle, essentially being determined by the endomorphism NN together with the given weight filtrations on the fibres HtH_{t}. We hope that the computation of NN in this paper will help to understand this structure, but at the moment the weight structures on the HtH_{t} are not well enough understood to say more. For the general theory, the interested reader is referred to [5] and the references cited there.

References

  • [1] Bestvina, M., and Feighn, M., The topology at infinity of Out(Fn)Out(F_{n}), Inv. Math. 140, no. 3, (2000), pp. 651-692.
  • [2] Bloch, S, Esnault, H., and Kreimer, D., On Motives Associated to Graph Polynomials, Comm. Math. Phys. 267 (2006), 181-225.
  • [3] Bergbauer, C., and Kreimer, D., Hopf algebras in renormalization theory: locality and Dyson-Schwinger equations from Hochschild cohomology, IRMA Lect.Math.Theor.Phys. 10 (2006) 133-164; hep-th/0506190.
  • [4] Broadhurst, D.J., and Kreimer, D., Association of multiple zeta values with positive knots via Feynman diagrams up to 9 loops, Phys. Lett.B 393 (1997) 403.
  • [5] Cattani, E., and Kaplan, A., Degenerating variations of Hodge structure, in Théorie de Hodge, Asterisque 179-180 (1989), pp. 67-96.
  • [6] Connes, A., and Kreimer, D., Renormalization in quantum field theory and the Riemann-Hilbert problem II: the β\beta-function, diffeomorphisms and the renormalization group, Commun. Math. Phys. 216 (2001) 215-241.
  • [7] Deligne, P., Equations Différentielles à Points singuliers Réguliers, SLN 163, Springer (1970).
  • [8] Esnault, H., Schechtman, V., Viehweg, E.: Cohomology of local systems on the complement of hyperplanes. Invent. math. 109 (1992), 557–561. Erratum: Invent.math. 112 (1993), 447.
  • [9] Fulton, W.: Introduction to Toric Varieties, Annals of Math. Studies 131, Princeton University Press (1993).
  • [10] Itzykson, J.-C., Zuber, J.-B., Quantum Field Theory. Mc-Graw-Hill, 1980.
  • [11] Kontsevich, M., Zagier, D., Periods, in Mathematics Unlimited- 2001 and Beyond, edited by W. Schmid and Björn Engquist, Springer (2001).
  • [12] Kreimer, D., On overlapping Divergences, Commun.Math.Phys. 204 (1999) 669, arXiv:hep-th/9810022.
  • [13] Kreimer, D., On the Hopf algebra structure of perturbative quantum field theories, Adv.Theor.Math.Phys.2 (1998) 303–334, arXiv:q-alg/9707029.
  • [14] Kreimer, D., Chen’s iterated integral represents the OPE, Adv.Theor.Math.Phys.3 (1999) 627-670, arXiv:hep-th/9901099.
  • [15] Kreimer, D., Dyson Schwinger equations: From Hopf algebras to number theory, Fields Inst.Commun.50 (2007) 225–248, arXiv:hep-th/0609004.
  • [16] Kreimer, D., and Yeats, K., Recursion and growth estimates in renormalizable quantum field theory, Commun.Math.Phys.279 (2008) 401–427, arXiv:hep-th/0612179.
  • [17] Kreimer, D., and Yeats, K., An Etude in non-linear Dyson-Schwinger Equations, Nucl.Phys.Proc.Suppl.160 (2006) 116–121, arXiv:hep-th/0605096.
  • [18] Looijenga, E., Cellular decompositions of compactified moduli spaces of pointed curves. The moduli space of curves (Texel Island, 1994), 369–400, Progr. Math., 129, Birkh user Boston, Boston, MA, (1995).
  • [19] Oda, T.: Convex Bodies and Algebraic Geometry, Springer Verlag (1985).
  • [20] Sokal, A., The multivariate Tutte polynomial for graphs and matroids, in ”Surveys in Combinatorics, 2005”, edited by Bridget S. Webb (Cambridge University Press, 2005), pp. 173-226, arXiv.math.CO/0503607.
  • [21] Smirnov, V., Evaluating Feynman Integrals, Springer Tracts in Modern Physics 211. Berlin, Springer. ix, 247p. (2004).
  • [22] Todorov, I., Analytic Properties of Feynman Diagrams in Quantum Field Theory, Monographs in Natural Philosophy 38, Pergamon Press (1971).