SciPostPhysLectNotes 75
SciPostPhysLectNotes 75
Notes 75 (2023)
Abstract
These lecture notes provide a self-contained introduction to Euler integrals, which are
frequently encountered in applications. In particle physics, they arise as Feynman inte-
grals or string amplitudes. Our four selected topics demonstrate the diverse mathemat-
ical techniques involved in the study of Euler integrals, including polyhedral geometry,
very affine varieties, differential equations, and computational algebra.
Contents
Introduction 2
3 Twisted (co)homology 19
3.1 Twisted chains 20
3.2 Twisted cochains 23
3.3 Back to Euler integrals 25
5 Open problems 38
References 39
1
SciPost Phys. Lect. Notes 75 (2023)
Introduction
Consider ℓ Laurent polynomials f1 , . . . , fℓ in n variables x = (x 1 , . . . , x n ) with complex coeffi-
cients. By an Euler integral, we mean any integral of the following form:
ν ν
x 1 1 · · · x n n dx 1
Z Z
dx n dx
s1 sℓ ∧ ··· ∧ = f −s x ν . (1)
Γ f1 · · · fℓ x1 xn Γ x
The right-hand side is our shorthand notation. The first example is the Euler beta function
Z1 Z∞
xν dx Γ (ν)Γ (1 − s)
B(ν, 1 − s) = = , where Γ (u) = t u−1 e−t dt , (2)
0 (1 − x)s x Γ (ν + 1 − s) 0
is the gamma function. The equality in blue will be derived below. Such integrals have been
called many different names, depending on the context in which they are studied. They were
called generalized Euler integrals by Gelfand, Kapranov and Zelevinsky [31]. This was moti-
vated by Euler’s integral representation of Gauss’ hypergeometric function. In fact, the integral
(1) represents a generalized hypergeometric function, and the name hypergeometric integral has
appeared in the literature as well [5]. When s1 = · · · = sℓ = 1 and Γ = R+ n
, our integral is a
−1
function of ν called the Mellin transform of ( f1 · · · fℓ ) [53]. This lead the authors of [9] to use
the name Euler-Mellin integrals for general s and Γ = R+ n
. Seminal contributions like [4, 30]
justify the name Aomoto-Gelfand integrals. In physics, Feynman integrals in quantum field the-
ory and string amplitudes in superstring theory take the form (1) for particular choices of f i .
We elaborate on these specific polynomials below. In Bayesian statistics, Euler integrals appear
as marginal likelihood integrals [13]. In our title, we chose to use Euler integrals as an umbrella
term for all these instances of (1).
In different sections, we will view the integral (1) as a function of different sets of param-
eters. For instance, in Section 1, we will fix Γ = R+ n
and think of (1) as a function of s and
ν. On the other hand, in Section 3, we think of the integrand as an element of a cohomology
vector space. Hence, the integral gives a linear function which sends Γ to (1). We will also
consider the case where the coefficients of f i depend on some parameters z. In this case our
integral is a function of z satisfying some interesting differential equations, see Section 4.
As mentioned above, Euler integrals appear in particle physics. The first important exam-
ple comes from quantum field theory, where Feynman integrals are used to describe particle
scattering processes. For a complete introduction to the subject, we refer to the recent book by
Weinzierl [61]. In the Lee-Pomeransky representation [43], up to a prefactor involving gamma
functions in s, ν, the Feynman integral of a graph G takes the form
xν
Z
dx
IG = , (3)
Rn (UG + FG ) x
s
+
where n is the number of internal edges of G, and UG , FG are the first and second Symanzik
polynomials associated to the graph. We illustrate this with one of our running examples.
Example 0.1. Consider the triangle diagram G in (4) with three massless internal edges. The
internal edges carry variables (x 1 , x 2 , x 3 ). The three external (open) edges attached to each
vertex carry the kinematic parameters (t 1 , t 2 , t 3 ):
t2
x3 x1 (4)
t1 x2 t3
2
SciPost Phys. Lect. Notes 75 (2023)
The polynomial UG is the sum over spanning trees of G, with each term given by the x i ’s not
present in the tree:
UG = + + = x1 + x2 + x3 . (5)
The FG polynomial is given similarly as a sum of spanning two-forests (disjoint unions of two
trees), each weighted with minus the corresponding kinematic variable:
FG = + + = −t 1 · x 2 x 3 − t 2 · x 3 x 1 − t 3 · x 1 x 2 . (6)
The exponents νi are typically taken to be non-negative integers and s = D/2 is half the space-
time dimension D. It is often convenient to think of (ν1 , ν2 , ν3 ) and s as generic parameters,
which is referred to as analytic and dimensional regularization respectively. ⋄
The second application of Euler integrals in physics comes from scattering amplitudes in
string theory. Instead of particles, one computes the probability of strings interacting with
each other. See [44] for a comprehensive review. This offers a nice immediate connection to
algebraic geometry: The integration is on the moduli space M0,m of genus zero curves with m
marked points. Equivalently, this is the space of configurations of m distinct points on P1 up
to its automorphisms PSL(2). We can represent these points as the columns of a 2 × m matrix
with nonzero 2 × 2 minors. Two such matrices M1 , M2 represent equivalent configurations
if there is an invertible 2 × 2 matrix T and an n × n invertible diagonal matrix D such that
T · M1 · D = M2 . We can use the action of T and D to fix 3 out of m points, leaving n = m − 3
degrees of freedom. Following [8, Eq. (1.5)], we write a point of M0,m as
1 1 1 1 ··· 1 0
M = , (8)
0 1 1 + x1 1 + x1 + x2 · · · 1 + x1 + · · · + x n 1
where n = m − 3 and the 2 × 2 minors f i j = M1i M2 j − M1 j M2i , i < j are nonzero. The genus
zero contribution to the m-point string amplitude is given by an Euler integral depending on
an extra parameter α′ :
Z α′ ν1 α′ νn
′ n
x1 · · · xn dx
Im = (α ) · α′ si j
. (9)
M+
Q x
0,m 1<i+1< j<m f i j
The pairs (i, j) excluded in the product in the denominator are those for which the minor f i j
is either constant or one of the x-variables. The integration is over the positive part M+ 0,m of
M0,m , which is the subset of points M satisfying f i j > 0, for all 1 ≤ i < j ≤ m. Using the
n
parameterization (8), one checks that this is R+ .
There are two physically interesting limits: α′ → 0 and α′ → ∞. The first one is called
the field theory limit in which strings become particles, and the second is the high-energy limit.
We will see in Section 2 that both of them admit an elegant geometric description.
Example 0.2 (m = 4). The moduli space M0,4 has dimension 1. The four-point string ampli-
tude (9) is I4 = α′ · B(α′ ν, −α′ ν + α′ s13 ), where B is the beta function from (2). ⋄
3
SciPost Phys. Lect. Notes 75 (2023)
The minors f23 and f34 are not included in the integrand of (9), since they would only shift
α′ ν j
the exponents of x j . The five-point string amplitude is given by
Z α′ ν1 α′ ν2
x1 x2 dx 1 dx 2
I5 = (α′ )2 · . (12)
(1 + x 1 )α′ s13 (1 + x1 + x2 )α s14 (x + x2 )α s24
′ ′
R2+ 1 x1 x2
The parameters (ν1 , ν2 , s13 , s14 , s24 ) describe momenta and angles of the 5 strings involved in
the scattering process. ⋄
Euler integrals have many other applications, including marginal likelihood integrals [13],
wave functions in cosmology [7], and correlation functions of conformal field theories [15,24].
These notes present the basics on Euler integrals from different points of view. They pro-
vide a roadmap through the literature for a reader who is new to the subject. At the same
time, we hope they serve as a helpful overview of important results for experts. Section 1
discusses convergence and meromorphic continuation, which leads us to study convex poly-
topes and polyhedral cones. Section 2 is about certain limits which are meaningful in physics
applications. This brings in algebraic equations, very affine varieties and Euler characteris-
tics. Section 3 develops the theory of (algebraic) twisted (co)homology on these very affine
varieties. Section 4 identifies difference and differential equations satisfied by Euler integrals.
Finally, Section 5 contains a list of open problems.
n
To ensure that the integrand is finite on R+ , we make the following assumption.
α α
where ci,α ∈ R+ , supp( f i ) ⊂ Zn is the support of f i (see Definition 1.1) and x α = x 1 1 · · · x n n .
4
SciPost Phys. Lect. Notes 75 (2023)
(0,0,1)
(1,0,1) (0,1,1)
(1,0,0) (0,1,0)
(1,1,0)
p
and log is only defined up to translates
p by integer multiples of 2π −1. When s is not an
integer, exp(si F ) ̸= exp(si (F + 2π −1k)) for some integer k, i.e., there are multiple branches
of f i (x)si . Assumption 1 ensures that f i takes positive values on R+ n
, so that there is precisely
one positive branch of log f i and log x j . In this section, our integrand is
where the unique positive branches of log f i and log x j are intended.
As it turns out, statements about convergence of (13) involve convex polytopes and poly-
hedral cones. We start by introducing these objects, and then switch to convergence results
from [8, 9, 53]. In [9, 53], (13) was called an Euler-Mellin integral and weaker assumptions on
f i are used. In this text, we stick with Assumption 1 for simplicity.
s · P = {s · p : p ∈ P} .
Here s · p is the usual scalar multiplication for vectors in Rn . It is easy to check that s · P is
indeed a convex polytope. The Minkowski sum of two polytopes P, Q is a new polytope
P + Q = {p + q : p ∈ P, q ∈ Q} .
This binary operation is commutative and associative. An example is shown in Figure 1 (left),
where we take the sum of three polytopes in R2 . Each is the convex hull of the points repre-
sented by black bullets. The dimension of a polytope is the dimension of the smallest affine
space containing it. Figure 1 (left) shows two polytopes of dimension two (these are also
called polygons), and two polytopes of dimension one (i.e., line segments). In the right part
of that figure, we show a three-dimensional convex polytope in R3 . The polytopes we will
encounter in this text arise as the Newton polytope of a Laurent polynomial.
5
SciPost Phys. Lect. Notes 75 (2023)
Definition 1.1 (Newton polytope). Let f = α∈Zn cα x α ∈ C[x 1±1 , . . . , x n±1 ] be a Laurent poly-
P
nomial. The support of f is the set supp( f ) = {α ∈ Zn : cα ̸= 0}. The Newton polytope
∆( f ) ⊂ Rn is defined as the convex hull of the support, i.e., ∆( f ) = conv(supp( f )).
Example 1.2. The polytopes in Figure 1, from left to right, are the Newton polytopes of
1 + x1 , 1 + x1 + x2 , x1 + x2 , (1 + x 1 )(1 + x 1 + x 2 )(x 1 + x 2 ) ,
P y = {p ∈ P : y · p = min y · q} .
q∈P
CQ = { y ∈ Rn : Q ⊆ P y } .
For any face Q ⊆ P, CQ is a polyhedral cone. I.e., there is a finite set A ⊂ Rn such that
¨ «
X
CQ = pos(A) = c r r : c r ∈ R≥0 . (15)
r∈A
All our cones are polyhedral, so we will sometimes just refer to them as cones. The dimension
of a cone is the dimension of the smallest linear space containing it. For our cones CQ , we have
dim CQ = n − dim Q. E.g., if v ∈ P is a vertex, we have dim C v = n. If dim P = n and Q is a
facet, then CQ is a one-dimensional cone. These are called rays. When Q runs over all faces,
the cones CQ tile up Rn . The same is true for the vertices v. In symbols:
[ [
Rn = CQ = Cv . (16)
Q v
Example 1.3. A pentagon P in R2 has five vertices. These give five pointed full-dimensional
cones in its normal fan. The ray separating two neighboring cones C v1 and C v2 is the cone
C v1 v2 corresponding to the edge containing v1 and v2 . This is illustrated in Figure 2. The cone
C P = {0} is the only zero-dimensional one. The normal fan Σ P is invariant under translations
of P. I.e., Σ P = Σ P+w for w ∈ Rn . We encourage the reader to check this. ⋄
6
SciPost Phys. Lect. Notes 75 (2023)
Figure 3: The polar dual of P and its subdivision induced by the normal fan Σ P .
If P is full-dimensional and it contains the origin in its interior int(P), then P ◦ is again a
polytope. Its vertices lie on the rays of the normal fan Σ P . Hence, the normal fan induces a
subdivision [
P◦ = B v , where B v = C v ∩ P ◦ . (17)
v
Example 1.4. The polar dual and its subdivision are illustrated in Figure 3. To satisfy 0∈int(P),
we translated our polytope so that (0, 0) is an interior lattice point. ⋄
Lemma 1.5. Let P be a full-dimensional polytope in Rn . We have 0 ∈ int(P) if and only if for all
vertices v of P, y · v < 0 for all y ∈ C v \ {0}.
Here rQ is any ray generator of the ray CQ , and vQ is any vertex contained in Q. If and only if
P p = 0 belongs to this set. For a given vertex v of P, a
all right-hand sides rQ · vQ are negative,
vector y ∈ C v can be written as y = v∈Q cQ rQ , with cQ ≥ 0. The lemma follows.
7
SciPost Phys. Lect. Notes 75 (2023)
Suppose (s13 , s14 , s34 ) = (1, 1, 1). By Theorem 1.6, the integral converges if (ν1 , ν2 ) ∈ int(P),
where P is the pentagon in the middle of Figure 1. For (ν1 , ν2 ) = (1, 1), we find using
in Mathematica that I5 = π2 /6. Multiplying the integrand with x1*x2, i.e., using
(ν1 , ν2 ) = (2, 2), the program prints a message saying that the integral does not converge. ⋄
R
The normalized volume of a compact set B ⊂ Rn is defined as Vol(B) = n! · B 1dx. Our
proof of Theorem 1.6 bounds the Euler integral (13) in terms of the normalized volume of
polytopes, see Equation (23) below. It uses Lemma 1.5, as well as the following lemma.
Lemma 1.8. Let C ⊂ Rn be an n-dimensional polyhedral cone and let v ∈ Rn be such that y ·v < 0
for all y ∈ C \ {0}. Then B = { y ∈ C : y · v ≥ −1} is a polytope with volume
Z
Vol(B) = exp( y · v) d y . (19)
C
If, instead, y · v ≥ 0 for some y ∈ C \ {0}, then the integral above diverges.
Proof. It suffices to show this in the case where C is simplicial, with n ray generators r1 , . . . , rn .
This is because if C is not simplicial, it can be subdivided into finitely many simplicial cones
C1 , . . . , Ck , and we would conclude
X k X k Z Z
Vol(B) = Vol(B ∩ Ci ) = exp( y · v) = exp( y · v) .
i=1 i=1 Ci C
Since y · v < 0 for all y ∈ C, we may also assume that the ray generators ri are scaled so
that ri · v = −1. This means that Vol(B) = |det(A)|, where A = (r1 , . . . , rn ) is a matrix whose
columns are the ray generators. Since C is simplicial, a point y = ( y1 , . . . , yn ) ∈ C can be
written uniquely as y = Az, where z = (z1 , . . . , zn ) are new nonnegative coordinates. Hence
Z Z Z Y n
exp( y · v) = |det(A)| exp(v ⊤ Az)dz = |det(A)| exp(zi (ri · v))dz .
n n
C R+ R+ i=1
The integral is finite if and only if ri · v < 0 for all rays, which is equivalent to y · v < 0 for all
y ∈ C \ {0}. In this case, it equals | det(A)| = Vol(B), as desired.
8
SciPost Phys. Lect. Notes 75 (2023)
exp( y · ν)
Z Z
−s ν dx
I (s, ν) = f x = dy .
Rn f (exp( y))
x s
Rn +
p p
The equality in this display uses |r a+ −1b | = |exp(log(r)(a + −1b))| = |exp(a log(r))| = r a
for a positive real number r and real numbers a, b. Notice that this means we may assume s
and ν are real. We first consider the case where ℓ = 1. Let P be the Newton polytope ∆( f ) of
f = f1 . We have seen in (16) that its normal fan subdivides Rn into n-dimensional polyhedral
cones C v , where v runs over the vertices of P. Therefore
X XZ exp( y · ν)
I (s, ν) = I v (s, ν) = dy . (20)
−C f (exp( y))
s
v v v
Notice that we use the cones −C v instead of C v for this decomposition,Pbecause these are the
domains on which we can find easy bounds for the integrand. Let f = Pα cα · x α with cα > 0,
as in Assumption 1. With our change of variables, this becomes f (e y ) = α cα · e y·α . Since v is
a vertex of ∆ f , one of the exponents α equals v. For y ∈ −C v , y · v ≥ y ·α for α ∈ supp( f )\{v}.
This gives the following chain of inequalities:
X
c v · exp( y · v) ≤ f (exp( y)) ≤ cα · exp( y · v) . (21)
α
Z Z
M −s · exp( y · (ν − sv)) d y ≤ I v (s, ν) ≤ c v−s · exp( y · (ν − sv)) d y .
−C v −C v
The integral appearing on the left and right of this expression can be written as
Z
exp( y · w) d y , with w = sv − ν .
Cv
By Lemma 1.8, this integral converges if and only if y · w < 0 for all y ∈ C v \ {0}. Notice that w
is a vertex of the polytope s · P − ν, and the cone Csv−ν in its normal fan equals C v . By Lemma
1.5, y · w < 0 for all y ∈ Cw \ {0} = C v \ {0} for all vertices w if and only if 0 ∈ int(s · P − ν).
This is equivalent to ν ∈ int(s · P), which proves the theorem for ℓ = 1.
When ℓ > 1, the formula (20) generalizes to a sum over the vertices of P(s):
X XZ exp( y · ν)
I (s, ν) = I v (s, ν) = dy .
v v
f
−C 1 (exp( y)) s1 · · · f (exp( y))sℓ
ℓ
v
Again, the integral in these expressions converges if and only if 0 ∈ int(P(s) − ν).
9
SciPost Phys. Lect. Notes 75 (2023)
Remark 1.9. The decomposition of the integral I (s, ν) in (20) is referred to as sector decom-
position in the physics literature [36, Section 3.4]. This is used in state-of-the-art algorithms
for evaluating Feynman integrals numerically, which use Monte Carlo sampling and tropical
geometry [11, 12].
By Lemma 1.8, if Re(si ) > 0 and ν ∈ int(P(s)), the bounds in (22) can be written as
ℓ ℓ
−Re(si ) −Re(si )
Y Y
Mi · Vol(B v−ν ) ≤ |I v (s, ν)| ≤ ci,v · Vol(B v−ν ) . (23)
i
i=1 i=1
1 ∞
xν yν
Z Z
dx dy
= , with s̃ = ν + 1 − s . (24)
0 (1 − x) x
s
0 (1 + y) y
s̃
Theorem 1.6 predicts convergence when Re(s̃) > 0 and ν ∈ int(P(s̃)), where
To justify the blue equality in (2), we observe that when these convergence conditions hold,
Z∞ Z∞
ν−s −t yν dy
Γ (ν + 1 − s) · B(ν, 1 − s) = t e dt ν+1−s
0 0 (1 + y) y
Z ∞Z ∞ ν−1 −s
ty t t d y dt
= e−t .
0 0 1+ y 1+ y (1 + y)2
ty t td ydt
With the coordinate change u = 1+ y , w= 1+ y we have u + w = t and (1+ y)2
= dudw. Hence
Z ∞ Z ∞
ν−1 −u
Γ (ν + 1 − s) · B(ν, 1 − s) = u e du w−s e−w dw = Γ (ν)Γ (1 − s) .
0 0
While the integrals in these equalities only make sense in their respective convergence regions,
we may use the definition of the gamma function to extend the beta function to a meromorphic
function on C2 :
Γ (ν)Γ (1 − s)
B(ν, 1 − s) = .
Γ (ν + 1 − s)
Its poles are countably many lines in (s, ν)-space, given by ν, 1 − s ∈ Z≤0 . ⋄
The fact that the beta function extends to a meromorphic function whose poles are given
by some gamma functions is an example of a general phenomenon, proved in [9]. We include
the statement, but omit the proof. We refer the reader to [9, Theorem 2.4] for full details.
10
SciPost Phys. Lect. Notes 75 (2023)
Theorem 1.11. Suppose the Minkowski sum ∆( f1 )+· · ·+∆( fℓ ) has dimension n and the polytope
Pℓ
P(s) = i=1 Re(si ) · ∆( f i ) is given by N < ∞ inequalities:
P(s) = {p ∈ Rn : ri · p ≥ w i · Re(s) , i = 1, . . . , N } , r i ∈ R n , w i ∈ Rℓ .
Under Assumption 1, I (s, ν) from (13) admits a meromorphic continuation of the form
N
Y
Φ f (s, ν) · Γ (ri · ν − w i · s) , (26)
i=1
It is worth noting that only the entire factor Φ f (s, ν) in Theorem 1.11 depends on the spe-
cific positive coefficients of f . The gamma factors only depend on the polyhedral data coming
from P(s). For fixed, positive s, the poles of this meromorphic continuation are hyperplanes
emanating from the boundary of P(s). This is illustrated in Figure 4 for (18).
Example 1.12. In Example 1.10, we read off from (25) that N = 2 and r1 = 1, w1 = 0,
r2 = −1, w2 = −1. The gamma factors in Theorem 1.11 are Γ (r1 · ν − w1 · s̃) = Γ (ν) and
Γ (r2 · ν − w2 · s̃) = Γ (1 − s). The entire function Φ1+ y (s, ν) equals Γ (ν + 1 − s)−1 . ⋄
Notice that δ−1 plays the role of the inverse string tension α′ in the string amplitude (9). We
are interested in the opposite limits limδ→∞ I (δ) and limδ→0+ I (δ). Motivated by the physics
11
SciPost Phys. Lect. Notes 75 (2023)
application, these are called field theory limit and high energy limit respectively [8, 52]. They
are the leading terms in the series expansions of I (δ) around δ = ∞ and δ = 0.
As it turns out, both the field theory limit and the high energy limit can be expressed in
terms of complex critical points of the potential function or log-likelihood function
These critical points are the complex solutions to the n rational function equations
∂ f1 ∂ fℓ
∂ (log f −s x ν ) νj ∂ xj ∂ xj
= − s1 − · · · − sℓ = 0, j = 1, . . . , n . (28)
∂ xj xj f1 fℓ
These rational functions are defined where neither x j nor f i (x) are zero. We define
This is an example of a very affine variety, see [41, page 6]. The set of complex critical points
of log L is Crit(log L) = {x ∈ X : x satisfies (28)}. Since (28) consists of n equations in n
unknowns, we expect Crit(log L) to be finite. A solution x ∈ Crit(log L) is degenerate if
∂ ∂
Hlog L = det x j xk log L(x) = 0. (30)
∂ xj ∂ xk j,k
This determinant is called the toric Hessian of log L. It is much like the usual Hessian determi-
nant, but with ∂ /∂ x j replaced by the Euler operator x j (∂ /∂ x j ). Using the toric version will
be convenient later in the section. The following result is Theorem 1 in [40].
Theorem 2.1. There is a dense open subset U ⊂ Cℓ+n such that for (s, ν) ∈ U, the number of
solutions to (28) equals the signed Euler characteristic (−1)n · χ(X ) of the very affine variety X ,
and all solutions are non-degenerate, meaning that Hlog L (x) ̸= 0 for all x ∈ Crit(log L).
Theorem 2.1 says that the number of points in Crit(log L) depends only on the topology of
the space X . We will see how to compute the Euler characteristic χ(X ) below.
In Section 2.1 we discuss how to compute Crit(log L) using numerical homotopy continu-
ation. Sections 2.2 and 2.3 explain how these critical points are used to compute field theory
and high energy limits respectively.
Its partial derivatives give two rational function equations g1 = g2 = 0 in the unknowns x 1 , x 2 :
s13 s14 s34 ν1 s14 s34 ν
g1 = − − − + , g2 = − − + 2 . (31)
1 + x1 1 + x1 + x2 x1 + x2 x1 1 + x1 + x2 x1 + x2 x2
Importantly, we think of s and ν as parameters at this stage. We will emphasize the dependence
of g i on these parameters by writing g i (x; s, ν). The fixed complex parameters we want to solve
for are denoted by (s∗ , ν∗ ) ∈ U. Here is how to code this up in Julia:
12
SciPost Phys. Lect. Notes 75 (2023)
Here we chose s∗ = (1, 1, 1) and ν∗ = (1, 1), like in Example 1.7. The strategy for solving
g1 (x; s∗ , ν∗ ) = g2 (x; s∗ , ν∗ ) = 0 consists of two steps:
1. Solve g1 (x; s̃, ν̃) = g2 (x; s̃, ν̃) = 0 for a different set of parameters (s̃, ν̃) ∈ U.
2. Deform the start parameters (s̃, ν̃) continuously into the target parameters (s∗ , ν∗ ) and,
along the way, keep track of the solutions to g1 (x; s, ν) = g2 (x; s, ν) = 0.
Both these steps require numerically tracking solution paths as we vary the parameters. This
can be phrased as numerically solving an ordinary differential equation called the Davidenko
equation. For details, we refer to the standard textbook [58].
Step 1 is done using the monodromy method [25]. We explain how this works in a nut-
shell, using Figure 5 as an illustration. In that cartoon, the solutions for a fixed point (s, ν)
are represented by the points on the blue surface lying directly above it. This surface rep-
resents the incidence space {(x, s, ν) ∈ X × Cℓ+n : g1 (x; s, ν) = g2 (x; s, ν) = 0}. Choose
a random point x̃ ∈ X , and let (s̃, ν̃) be any solution to the linear system of equations
g1 (x̃; s, ν) = g2 (x̃; s, ν) = 0. Clearly, x̃ is a point lying above (s̃, ν̃). This is called the seed
solution. Now walk a loop in (s, ν)-space while keeping track of the seed solution x̃ along the
way. When we arrive back at (s̃, ν̃), there is a good chance we picked up a new solution x̃ new
to the system g1 (x; s̃, ν̃) = g2 (x; s̃, ν̃) = 0. Now repeat this procedure to populate the solution
set. In practice, this technique is extremely effective. In Julia, all this happens via
R1 = monodromy_solve(g_sys) 1
The variable start_pars stands for start parameters. It contains ℓ + n = 5 complex numbers,
the first ℓ = 3 of which give s̃, and the last n = 2 give ν̃. If all went well, the variable
start_sols contains all solutions to g1 (x; s̃, ν̃) = g2 (x; s̃, ν̃) = 0. Hence, step 1 is completed.
13
SciPost Phys. Lect. Notes 75 (2023)
By Theorem 2.1, the number of solutions in start_solutions equals (−1)n · χ(X ). This
gives a way of computing χ(X ), which has been applied in some challenging cases [2, 59]. In
our example, the number of solutions is 2. Here is a way to verify that χ(X ) = 2. The real
part X R of the very affine variety X is the complement of an arrangement of five lines in R2 .
These lines are given by {x 1 = 0}, {x 2 = 0}, {1 + x 1 = 0}, {1 + x 1 + x 2 = 0} and {x 1 + x 2 = 0}.
By [60, Theorem 1.2.1], the signed Euler characteristic of X is the number of bounded cells of
X R . In our case, the bounded cells are two triangles.
In step 2, we use our start solutions as initial conditions for path tracking from (s̃, ν̃) to
∗ ∗
(s , ν ). That is, we use the solutions start_sols for parameters start_pars to compute
the solutions solutions(R2) for the target parameters [s_star; ν_star]:
target_parameters = [s_star;ν_star]) 2
solutions(R2) 3
The last line prints an accurate numerical approximation of the two critical points:
p p
5−1 − 5−1
,1 , ,1 . (32)
2 2
Proof. We prove the first equality. The second equality uses [8, Section 7.1, Claim 4]. See
also [59, Theorem 13]. For any fixed δ ∈ R+ , we have ν/δ ∈ int(P(s/δ)), and a vertex v of
P(s) gives a vertex (v − ν)/δ of P(s/δ) − ν/δ. We can use (23) to obtain the estimate
ℓ ℓ
−Re(si )/δ −Re(si )/δ
XY XY
Mi · Vol(B v−ν ) ≤ δ n · I (δ) ≤ ci,v · Vol(B v−ν ) . (34)
δ i δ
v i=1 v i=1
The sums are over vertices of P(s). The factor δ n in the middle comes from
−n
I (δ) = δ · I (s/δ, ν/δ), with I (s, ν) as in (13). Using the scaling property of the volume
Vol(B(v−ν)/δ ) = δ n · Vol(B v−ν ), we can cancel δ n from (34). Taking the limit δ → ∞ gives
X
lim I (δ) = Vol(B v−ν ) = Vol((P(s) − ν)◦ ) ,
δ→∞
v
as desired. For the last equality, see (17) and Figure 3.
Example 2.3. Let us verify the formula (33) for the integral representation (24) of the beta
function. The Newton polytope P(s̃) is a segment given by (25). The dual polytope (P(s̃) − ν)◦
1 1
is given by ( ν−s̃ , ν ). Its volume Vol((P(s̃) − ν)◦ ) is ν1 + s̃−ν
1 s̃
= ν(s̃−ν) . Checking that this equals
our field theory limit can be done in one line of Mathematica code:
14
SciPost Phys. Lect. Notes 75 (2023)
Finally, we compute the sum of H− log L ( y) evaluated at the critical points of log L. We solve
d ν s̃
log L( y) = − = 0.
dy y 1+ y
ν
The unique solution is y = s̃−ν . The toric Hessian determinant of − log L is
d d d s̃ y s̃ y
H− log L ( y) = − y y log L( y) = − y ν− = . (35)
dy dy dy 1+ y (1 + y)2
ν ν(s̃−ν)
The value at y = s̃−ν is s̃ . We have now confirmed (33). ⋄
Example 2.4. Consider the integral I (δ) from (12) with α′ = δ1 . The Newton polytope
P(s) = P(s13 , s14 , s24 ) of f1 (x 1 , x 2 ) = 1 + x 1 , f2 (x 1 , x 2 ) = 1 + x 1 + x 2 , f3 (x 1 , x 2 ) = x 1 + x 2
is two dimensional for positive s. When s13 = s14 = s24 = 1, it is as in Figure 1. We take
ν1 = ν2 = 1. The values of H− log L at the two critical points (32) are
1 p 1 p
25 + 11 5 , 25 − 11 5 .
2 2
The sum of the reciprocals of these two numbers is 5. This is the area of (P(s)−ν)◦ in the right
part of Figure 3, normalized by a factor 2! = 2 (recall our definition of Vol in the discussion
preceding Lemma 1.8). Let us illustrate the computation of the dual volume using the Julia
package OSCAR.jl (v0.12.0) [54]. It calls polymake for polytope computations [29]. The
Newton polytope P = P(s13 , s14 , s24 ) of Example 2.4 is computed as follows:
The dual polytope (P(s) − ν)◦ and its volume are computed by the commands polarize
and volume respectively. The vertices of the dual polytope (P − ν)◦ are (1, 1), (1, 0), (0, −1),
(−1, −1), (0, 1) as in Figure 3. The normalized volume is 5, as expected.
The reader is encouraged to repeat this example for the Feynman integral (7). ⋄
We point out that the field theory limit α′ → 0 of the string amplitude (9) is the scattering
amplitude for a physical model called bi-adjoint scalar φ 3 theory. The expression in terms of
critical points was first discovered in [18]. The critical point equations (28) are called the
scattering equations in this context. For a connection to algebraic statistics, see [59].
As a final remark on field theory limits, note that the coordinates of the individual critical
points in Crit(log L) are algebraic functions of s, ν. They are usually not rational functions, like
in Example 2.3. For instance, eliminating x 2 from (31) gives a quadratic equation in x 1 , result-
ing in the square roots in (32) via the quadratic formula. However, the sum over Crit(log L)
15
SciPost Phys. Lect. Notes 75 (2023)
in Theorem 2.2 is a rational function in s and ν by Galois theory. This is called the canonical
function of P(s). On the domain Re(si ) > 0, ν ∈ int(P(s)), it evaluates to Vol((P(s) − ν)◦ ).
When s with positive real part is fixed and P(s) is viewed as a polytope in n-dimensional ν-
space Rn , the canonical function defines a meromorphic top form on Rn . That form is called
the canonical form of P(s) in the theory of positive geometries [6].
We used the following conventions in (36). As above, for a positive real number r, we take
1 π 1
the branch of the logarithm for which log r ∈ R. For the square root, we set (−r) 2 = e i 2 r 2 .
Notice that, rather than requiring Re(si ) > 0, here weponly allow real values for si . The reader
can check that, using the values s13 = s14 = s34 − −1 = 1 in the example of Section 2.1
instead, there are no positive critical points.
To prove Theorem 2.5, we use two lemmas. The first is on the concavity of log L.
∂ ∂
xj xk log L(x) , (37)
∂ xj ∂ xk j,k
n
of log L is negative definite for any x ∈ R+ .
Proof. Substituting x j = exp(z j ), the toric Hessian matrix (37) is the usual Hessian matrix
Pℓ
of − i=1 si log f i (ez1 , . . . , ezn ). Each summand is strictly concave for z ∈ Rn by [17, Theorem
1.13], so the toric Hessian is indeed a negative definite matrix.
To state the next lemma, we introduce a version of the algebraic moment map µC : X −→ Cn .
This name comes from toric geometry, see Fulton’s book [28, Section 4.2]. Our moment map
is slightly different from the one used by Fulton. It is given by
ℓ ℓ
X
−1 ∂ fi X
−1 ∂ f i
µC (x) = x1 si f i (x) (x), . . . , x n si f i (x) (x) . (38)
i=1
∂ x1 i=1
∂ xn
This map is closely related to our critical points. From (28), it is clear that x ∈ Crit(log L) if
and only if µC (x) = ν. The crucial properties of µC are summarized in the following Lemma.
∂ µk n
of µC to R+
n
is a diffeomorphism, and the Jacobian matrix ∂ x j j,k=1 is positive definite.
16
SciPost Phys. Lect. Notes 75 (2023)
Proof. This theorem follows from [28, Section 4.2] if ℓ = 1. The statement for ℓ ≥ 1 has
appeared in [8, Claim 4]. For the readers’ convenience, we provide a concise proof. Let
Ai = supp( f i ) ⊂ Zn be the set of exponents appearing in f i (x) = α∈Ai ci,α x α . We construct
P
where ei is the i-th standard basis vector of Zℓ . Consider the Laurent polynomial
fˆ = s1 y1 f1 + · · · + sℓ yℓ fℓ ∈ R+ [ y1 , . . . , yℓ , x 1±1 , . . . , x n±1 ]. We define the map
∂ fˆ ∂ fˆ ∂ fˆ ∂ fˆ
µ̂ : Rℓ+n
+ → int(pos(A)) , ( y, x) 7−→ y1 , . . . , yℓ , x1 , . . . , xn .
∂ y1 ∂ yℓ ∂ x1 ∂ xn
Here, pos(A) is the positive hull seen in (15). The reader who is unfamiliar with moment
maps should check that the image of µ̂ indeed lies in the interior of the cone pos(A). By the
statement labeled (An ) in [28, page 83], µ̂ is a diffeomorphism.
We consider the polytope P̂(s) consisting of all points in pos(A) whose first ℓ coordinates
are (s1 , . . . , sℓ ). The preimage of P̂(s) under µ̂ is given by
X̂ + = {( y, x) ∈ Rℓ+n
+ : y1 f1 (x) = · · · = yℓ fℓ (x) = 1} .
In fact, µ̂|X̂ + : X̂ + → int( P̂(s)) is a diffeomorphism. We now relate this to the moment map
µ in (39). To identify the domains of µ and µ̂, we introduce the map κ : R+ n
→ X̂ + with
−1 −1
κ(x) = ( f1 (x) , . . . , fℓ (x) , x 1 , . . . , x n ). For the co-domains, note that ι : int(P(s)) → int( P̂(s))
with ι(v) = (s1 , . . . , sℓ , v) is an isomorphism. We obtain a diagram of diffeomorphisms
µ̂
X̂ + / int( P̂(s))
O O
κ ι
µ
n
R+ / int(P(s))
ℓ
∂ µk ∂ ∂ ∂ ∂
X
xj = −x j xk log L(x) = si x j xk log f i .
∂ xj j,k
∂ xj ∂ xk j,k i=1
∂ xj ∂ xk j,k
n
is positive definite on R+ . The positivity follows from Lemma 2.6.
Note that Lemma 2.7 implies that, under our assumptions, Crit(log L) ∩ R+ n
= µ−1 (ν) con-
sists of a single point {a}. This is the first claim in Theorem 2.5. While Theorem 2.5 uses the
fiber µ−1 (ν), Theorem 2.2 sums over the fiber µ−1 C (ν) of the complexified map µC .
17
SciPost Phys. Lect. Notes 75 (2023)
Figure 6). There exists a positive number ϵ such that the inequality log L(x) − log L(a) ≤ −ϵ
n
is true for R+ \ U. We obtain the following inequality for 0 < δ < 1:
Z Z
1 dx dx
− δ1 − δϵ
L(a) L(x) δ =e exp δ−1 (log L(x) − log L(a) + ϵ) (41)
n
R+ \U x n
R+ \U x
Z
− δϵ dx
≤e exp (log L(x) − log L(a) + ϵ) . (42)
Rn \U x
+
Here, we used the fact that log L(x) − log L(a) + ϵ is negative for any x ∈ R+n
\ U, and the
final integral in (42) is bounded because of Theorem 1.6. The inequality (42) shows that the
integral over R+ n
\ U converges to zero as δ → 0+ .
Let P be an orthogonal matrix which diagonalizes the Hessian matrix of log L:
∂ log L
2
⊤
P · (a) · P = D.
∂ x j∂ xk j,k
Here D is an n×n diagonal matrix, with negative diagonal entries. We perform a linear change
of coordinates y = P T (x − a). The Taylor expansion of log L(x) − log L(a) around y = 0 looks
like 12 ( y T D y + r( y)). Plugging this into our integral gives
y D y + r( y)
Z Z T
− δ1
1 dx dy
L(a) L(x) δ = exp .
i (P y + a)i
Q
U x P T (U−a) 2δ
The last denominator is the product of the entries of P y + a. Without loss of generality,pwe
may assume that P T (U − a) is a product of small intervals (−ε, ε)n . Replacing yi with yi / δ,
the last integral becomes
p
r( δ y)
Z
n 1 T dy
δ2 exp y Dy + Q p .
p p
(−ε/ δ,ε/ δ)n 2 δ i ( δP y + a) i
p p p
The function r( δ y)/δ is bounded for 0 < δ < 1 and y ∈ (−ε/ δ, ε/ δ)n and it converges
to 0 when δ tends to 0. Therefore, Lebesgue’s dominance convergence theorem proves
Z Z
− 2n − δ1
1 dx 1 T dy
lim δ L(a) L(x) δ = e2 y Dy .
δ→0 +
U x Rn a1 · · · an
This leaves us with a Gaussian integral. To finish the proof, recall that
Z∞ v
t 2π
λ 2
z
e 2 dz = , for λ < 0 ,
−∞ −λ
and use the fact that H− log L (a) = (a1 · · · an )2 · i (−Dii ).
Q
Example 2.8. Let us verify the formula (36) for the integral representation (24) of the beta
function. The integral (24) is expressed by Gamma functions as in (2):
ν
s̃ − ν ν Γ ( s̃−ν
δ )Γ ( δ )
I (δ) = B , = .
δ δ Γ ( δs̃ )
1 ν
The function L(a)− δ , where a is the unique critical point s̃−ν from (2.3), is given by
−s̃ − δ1
ν ν
s̃
.
s̃ − ν s̃ − ν
p 1
formula Γ (x) ∼
Stirling’s Ç 2πe−x x x− 2 (x → +∞) shows that the left-hand side of (36) is
1
s̃
given by ν(s̃−ν) . We have seen in Example 2.3 that this equals H− log L (a)− 2 . ⋄
18
SciPost Phys. Lect. Notes 75 (2023)
3 Twisted (co)homology
n
In this section, we abandon the concrete integration contour R+ , and we drop Assumption 1.
±1 ±1 ℓ ℓ
We fix f = ( f1 , . . . , fℓ ) ∈ C[x 1 , . . . , x n ] , s = (s1 , . . . , sℓ ) ∈ C and ν = (ν1 , . . . , νn ) ∈ Cn .
The perspective we take is that the Euler integral (1) is the result of a pairing between the
integration contour Γ and the differential n-form dx x . More generally, an n-form φ gives
Z
〈Γ , φ〉 = f −s x ν φ . (43)
Γ
This works nicely when Γ is a twisted n-cycle and φ is a twisted n-cocycle. We will introduce
these concepts, and see that the pairing (43) is a perfect pairing of finite dimensional C-vector
spaces. In particular, the integral (43) always evaluates to a (finite) complex number.
This story is reminiscent of the classical duality between singular homology and de Rham
cohomology, where one pairsRan integration contour ∆ on a complex manifold X with a differ-
ential form φ by evaluating ∆ φ [34, Chapter 0]. In our setting, X is the very affine variety
seen in (29). The material in this section is like that standard theory, but with a twist. For
instance, recall from the beginning of Section 1 that our integrand f −s x ν is multi-valued. To
make sense of the integral (43), we need to specify a branch. We will see that this information
is carried by our twisted cycle Γ . Also, a central role in de Rham’s cohomology theory is played
by Stokes’ theorem, which says that for an (n − 1)-form ψ,
Z Z
dψ = ψ. (44)
∆ ∂∆
In our setting, the twisted boundary operator ∂ω takes the choice of branch into account, and
the twisted differential ∇ω replaces the ordinary differential d to accommodate our integrals:
Z Z
f −s x ν ∇ω ψ = f −s x ν ψ . (45)
Γ ∂ω Γ
The meaning of the index ω will become clear soon. For now, it simply indicates the twist.
The theory of twisted (co)homology goes back to the seminal work of Deligne and
Grothendieck [23]. It has been investigated in the context of Euler integrals and hyperge-
ometric functions by several authors, among which we mention Aomoto, Gelfand, Iwasaki,
19
SciPost Phys. Lect. Notes 75 (2023)
Kapranov, Kita, Matsumoto and Zelevinsky. See [5, 31] and references therein. The relevance
of this theory in particle physics was first realized by Mastrolia and Mizera [46].
The section is organized as follows. We start by discussing twisted chains and cycles,
leading to a twisted version of the usual chain complex of X . Next, we switch to the dual
complex, called the twisted de Rham complex of X . We discuss properties of the (co)homology
of these complexes, ultimately leading to the perfect pairing in (43).
Here, a is 0 or 1 and θ is 0 or π. The remaining simplex is the line segment [ϵ, 1 − ϵ],
parameterized by t 7→ (1−t)ϵ+t(1−ϵ). These parameterizations fix the orientations visualized
by the arrows in Figure 7. ⋄
One checks that each branch τ of f −s x ν indeed satisfies the equation dτ − dlog( f −s x ν ) τ = 0.
The one-form dlog( f −s x ν ) is of crucial importance in this section, so we introduce the notation
ω = dlog( f −s x ν ). We have ω = g1 dx 1 + · · · + g n dx n , where g j are the rational functions in
(28). The symbol L−ω stresses the term −ω in the operator applied to τ in (47).
Picking a (linear combination of) branch(es) of f −s x ν on a singular k-simplex ∆ means
picking a section τ of L−ω on a sufficiently small open subset U ⊃ ∆. We formalize this
intuition by considering the direct limit
Here the open sets U containing ∆ are ordered by inclusion, and when U ⊂ U ′ , the map
L−ω (U ′ ) → L−ω (U) is given by restriction. A reader who is unfamiliar with direct limits can
simply think of elements in L−ω (∆) as branches of f −s x ν , restricted to ∆.
20
SciPost Phys. Lect. Notes 75 (2023)
Example 3.2. We continue Example 3.1. On each of the five simplices we define a section of
L−ω , where ω = (−s(1− x)−1 +ν x −1 ) dx is the logarithmic differential of f −s x ν = (1− x)−s x ν .
Here s and ν are fixed complex numbers. Notice that f −s x ν is the integrand of the beta function
in (2). At x = ϵ, both (1 − x) and x are positive. Let ζ = exp(−s log(1 − ϵ) + ν log ϵ) ∈ C,
where we evaluate the positive branch of the logarithm. The conditions
uniquely define τ0,0 ∈ L−ω (∆0,0 (ϵ)). Constraints like τ0,0 (ϵ) = ζ are called initial conditions.
On our other simplices, we choose τa,θ ∈ L−ω (∆a,θ (ϵ)), τ− ∈ L−ω ([ϵ, 1 − ϵ]), with
τ0,π (−ϵ) = τ0,0 (−ϵ) , τ− (ϵ) = τ0,0 (ϵ) , τ1,π (1−ϵ) = τ− (1−ϵ) , τ1,0 (1+ϵ) = τ1,π (1+ϵ) .
This gives us all ingredients to define the space of twisted k-chains on X , with twist ω:
M
Ck (X , −ω) = ∆ ⊗ L−ω (∆) . (50)
∆⊂X , k-simplex
Comparing (50) with (46) motivates why this construction is sometimes called the space of
k-chains with coefficients in L−ω . In words, Ck (X , −ω) consists of finite C-linear combinations
of elements of the form ∆ ⊗ τ, where τ : ∆ → C is an element of L−ω (∆). We say that ∆ is
loaded with the branch τ. We note that Ck (X , −ω) is non-zero for any k ≥ 0.
Let us now clarify the meaning of (43) for a twisted k-chain Γ ∈ Ck (X , −ω).
The integrals on the right are the familiar integrals of single valued k-forms on k-simplices.
We now explain how to take boundaries in this twisted setting. A k-simplex ∆ on X has
boundary ∂ ∆ = ∂ ∆0 + · · · + ∂ ∆k ∈ Ck−1 (X ). That is, if ∆ is given by the parameterization
ϕ : ∆ → X with ∆ the standard k-simplex in Rn , then ∂ ∆i is the (k − 1)-simplex in X coming
from the restriction of ϕ to the i-th boundary component of ∆. Let Γ = ∆ ⊗ τ ∈ Ck (X , −ω)
be a k-simplex ∆ on X , loaded with τ. Observe that there is a natural restriction map
ρ∆,∆′ : L−ω (∆) → L−ω (∆′ ) whenever ∆′ ⊂ ∆. The twisted boundary ∂ω (Γ ) of Γ is
Morally, this boundary operator simply keeps track of the branch of the twisted chain.
21
SciPost Phys. Lect. Notes 75 (2023)
Example 3.4. Consider again the five simplices illustrated in Figure 7, loaded with the
branches specified in Example 3.2. The twisted boundaries are
The orientation of the boundary components of a 1-simplex is like in standard singular ho-
mology: End point minus starting point. The restrictions of our sections to these boundary
points are simply given by their value at the point. It is instructive to reduce the number of
parameters in (53) by using our definitions and findings from p Example 3.2. For instance, we
have τ0,π (−ϵ) = τ0,0 (−ϵ), τ− (ϵ) = τ0,0 (ϵ), τ0,π (ϵ) = exp( −1ν2π) τ0,0 (ϵ), and so on. ⋄
Definition 3.5. Let X be the very affine variety from (29). Let ω = dlog( f −s x ν ), and let
Ck (X , −ω) be the space (50) of twisted k-chains on X . The twisted chain complex is
∂ω ∂ω ∂ω
(C• (X , −ω), ∂ω ) : · · · −→ Ck (X , −ω) −→ Ck−1 (X , −ω) −→ · · · −→ C0 (X , −ω) −→ 0 . (54)
The homology of this complex is obtained by considering all twisted chains whose twisted
boundary is zero, modulo those that are twisted boundaries themselves.
Definition 3.6. The k-th homology vector space of (C• (X , −ω), ∂ω ) is the quotient space
{Γ ∈ Ck (X , −ω) : ∂ω (Γ ) = 0}
H k (X , −ω) = . (55)
∂ω Ck+1 (X , −ω)
Elements of H k (X , −ω) are called twisted k-cycles (or sometimes loaded k-cycles). We will
primarily be interested in the n-th homology space H n (X , −ω), because these are the cycles on
which we can integrate n-forms. While it is easy to construct cycles in the usual (non-twisted)
singular homology, this is a bit more complicated in our twisted setting. The running example
of this section illustrates a standard construction [5, Section 3.2.4].
Example 3.7. None of the five twisted chains in Example 3.4 are twisted cycles, since they
have non-zero twisted boundaries. However, we can use the expressions (53) to find a linear
combination of these chains whose twisted boundary is zero. For ease of notation, let us write
Γa,θ = ∆a,θ (ϵ) ⊗ τa,θ ∈ C1 (X , −ω), and Γ− = [ϵ, 1 − ϵ] ⊗ τ− ∈ C1 (X , −ω). Consider
Note that (56) only makes sense when ν and s are non-integer. This genericity assumption will
appear in our theorems below. One checks that ∂ω (Γ ) = 0 by expanding it using (53), and
applying identities like at the end of Example 3.4. The class [Γ ] ∈ H1 (X , −ω) of Γ is non-zero.
We will show this in Example 3.22. Hence, Γ is not a boundary of a 2-chain. ⋄
22
SciPost Phys. Lect. Notes 75 (2023)
de Rham complex, we want to regard regular k-forms φ modulo those that integrate to zero
in (51) on a twisted cycle Γ = ∆ ⊗ τ. A first step towards formalizing this is the following
important observation.
Lemma 3.8. For any ψ ∈ Ωk−1 (X ) and any twisted k-chain Γ ∈ Ck (X , −ω), we have
Z Z Z
−s ν −s ν
f −s x ν ψ .
f x (d + ω∧)ψ = d f x ψ = (58)
Γ Γ ∂ω (Γ )
Equation (58) will be our twisted version of Stokes’ theorem (45), where the twisted dif-
ferential ∇ω is given by d + ω∧. That is, for any 0 ≤ k ≤ n we define
A regular k-form φ is closed if its twisted differential is zero, i.e., ∇ω (φ) = 0. In particular, all
n-forms are closed, since Ωn+1 (X ) = 0. A regular k-form φ is called exact if it is the twisted
differential of some (k − 1)-form: φ = ∇ω (ψ). Here is a consequence of Lemma 3.8.
Lemma 3.9. Let Γ ∈ Ck (X , −ω) be a twisted cycle and let φ ∈ Ωk (X ) be a closed k-form, i.e.,
∂ω (Γ ) = 0 and ∇ω (φ) = 0. If Γ is a twisted boundary or φ is exact, i.e., Γ = ∂ω (Γ ′ ) for some
Γ ′ ∈ Ck+1 (X , −ω) or φ = ∇ω (ψ) for some ψ ∈ Ωk−1 (X ), we have Γ f −s x ν φ = 0.
R
23
SciPost Phys. Lect. Notes 75 (2023)
Definition 3.10. Let X be the very affine variety from (29). Let ω = dlog( f −s x ν ), and let
Ωk (X ) be the space (57) of regular k-forms. The (algebraic) twisted de Rham complex is
∇ω ∇ω ∇ω
(Ω• (X ), ∇ω ) : 0 −→ Ω0 (X ) −→ Ω1 (X ) −→ · · · −→ Ωn (X ) −→ 0 . (60)
This complex is also called the twisted cochain complex, to emphasize its duality with (54)
(see below). As said above, since exact forms integrate to zero, we pass to cohomology.
Definition 3.11. The k-th twisted cohomology vector space of (60) is the quotient space
We regard φ in (43) as a twisted cocycle, i.e., an element of the n-th twisted cohomology
Ωn (X )
H n (X , ω) = . (61)
∇ω Ωn−1 (X )
The twisted de Rham complex in Definition 3.10 is called algebraic because we work with
the regular k-forms Ωk (X ) in the sense of algebraic geometry. One can build an analogous
complex using holomorphic k-forms, for which the coefficients g j1 ,..., jk in (57) can be any holo-
morphic functions on X . By the Grothendieck-Deligne comparison theorem [23, Corollaire
6.3], the cohomology of this holomorphic twisted de Rham complex is isomorphic to that of
(60). Since our cocycles φ will be regarded as elements in this cohomology, it suffices to work
with the algebraic complex (60). This is also the preferred setting for doing computations,
because the regular k-forms (57) have a very concrete description.
Here is an example of how to compute relations in twisted cohomology.
Example 3.12. Consider again the Euler beta integral (2). The twisted differential is
s ν
∇ω = d + + dx ∧ . (62)
1− x x
dx −ν dx
= .
1− x s x
More generally, we shall derive in Section 4 that for a, b ∈ Z, we have the relation
(1 − s)−a (ν) b dx
xb dx
= . (63)
(1 − x)a x (1 + ν − s) b−a x
Here, for a complex number γ and an integer a, we used the following notation:
γ(γ + 1) · · · (γ + a − 1)
(a > 0) ,
(γ)a := 1 (a = 0) , ⋄
(γ − 1)−1 (γ − 2)−1 · · · (γ + a)−1 (a < 0) .
24
SciPost Phys. Lect. Notes 75 (2023)
Theorem 3.13 (Vanishing theorem). Let X be the very affine variety from (29). There exists a
dense open subset U ⊂ Cℓ+n such that, for each (s, ν) ∈ U, we have
H k (X , ω) = 0 , for all k ̸= n , with ω = dlog( f −s x ν ) . (64)
A description of the open subset U ⊂ Cℓ+n follows from the proof in [3, Theorem A.1]. In
our running example, one can take U = {(s, ν) ∈ Cℓ+n : s, ν, s − ν ∈
/ Z}.
One of the consequences of this vanishing theorem is a geometric description of the di-
mension of H n (X , ω). By [5, Theorem 2.2], the topological Euler characteristic χ(X ) of the
very affine variety X is given by the alternating sum of cohomology dimensions:
n
X
χ(X ) = (−1)k dimC H k (X , ω) . (65)
k=0
25
SciPost Phys. Lect. Notes 75 (2023)
Proof. Well-definedness follows from Lemma 3.9. The pairing is perfect by [5, Lemma 2.9(1)],
using the Deligne-Grothendieck comparison theorem [23, Corollaire 6.3].
The pairing (68) is called the period pairing between twisted homology and cohomology.
In the theorem, V ∨ = HomC (V, C) denotes the dual vector space of a C-vector space V . The
maps H k (X , −ω) → H k (X , ω)∨ and H k (X , ω) → H k (X , −ω)∨ are given by
respectively. Notice that Theorem 3.17 makes no assumptions on s and ν. We spell out three
important implications (Corollaries 3.18, 3.19 and 3.21).
Corollary 3.18. Let X , ω be as above. For any k, dimC H k (X , −ω) = dimC H k (X , ω).
Corollary 3.19. If (s, ν) lies in the open subset U from Theorem 3.13, the vanishing theorem
extends to twisted homology: H k (X , −ω) = 0 when k ̸= n, and dimC H n (X , −ω) = |χ(X )|.
This means that, when (s, ν) ∈ U , we can find a set of χ = |χ(X )| basis elements [φ1 ],. . . ,[φχ ]
for H n (X , ω), and a set of χ basis elements [Γ1 ], . . . , [Γχ ] for H n (X , −ω). In particular, for any
(a, b) ∈ Zℓ+n , there exist coefficients c1 , . . . , cχa,b ∈ C such that
a,b
x b dx a,b
= c1 [φ1 ] + · · · + cχa,b [φχ ] , in H n (X , ω) . (69)
fa x
Example 3.20. For the beta integral, we have seen in Example 3.12 that for φ1 = [dx/x],
Corollary 3.21 says that relations in cohomology like (63) are equivalent to relations be-
tween Euler integrals which hold for any twisted cycle. That is, Equation (63) implies
for any [Γ ] ∈ H1 (C \ {0, 1}, −ω). More generally, the expansion (69) in terms of a basis gives
Z ν+b Z ν Z ν
x dx a,b x x
s+a
= c1 s
a,b
φ1 + · · · + cχ s
φχ , for all [Γ ] ∈ H n (X , −ω) .
Γ f x Γ f Γ f
The integrals on the right-hand side are called a set of master integrals in physics, see [61].
Example 3.22. In Example 3.7, we promised to show that Γ from (56) is nonzero in twisted
homology. For this, let us fix values of s, ν such that 0 < ν < ν − s + 1. These are precisely
the convergence conditions derived in Example 1.10. We will come back to this later. The
twisted cycle Γ in (56) depends on ϵ, but by the Cauchy-Goursat theorem, the value of the
26
SciPost Phys. Lect. Notes 75 (2023)
integral 〈Γ , dx
x 〉 is independent of ϵ ∈ (0, 1/2). According to the three terms in (56), we split
the integral up into three parts: 〈Γ , dx
x 〉 = I 0 (ϵ) + I − (ϵ) + I 1 (ϵ). The first summand is
Z 2π p p
1 −1θ −s ν −1θ ν
I0 (ϵ) = p (1 − ϵe ) ϵ e dθ .
e2π −1ν −1 0
Because of the assumption ν > 0, we have limϵ→0+ I0 (ϵ) = 0. Analogously, one shows
limϵ→0+ I1 (ϵ) = 0. Finally, by (2), we have
Z 1−ϵ
dx xν dx
·
Γ, = lim+ I− (ϵ) = lim+ = B(ν, 1 − s) .
x ϵ→0 ϵ→0
ϵ (1 − x) x
s
restricted to ∆. Such a branch is fixed after fixing the branches of the logarithms log f i , log x j ,
which are independent of s and ν. Our integral is the following function of s, ν:
Z Z
(s, ν) 7−→ f −s x ν φ = τ(s, ν)(x) φ = 〈∆ ⊗ τ(s, ν), φ〉 . (70)
∆⊗τ(s,ν) ∆
Taking derivatives of (70) in s, ν can be done under the integration sign [42, Chapter XVII,
Theorem 8.2]. This implies the following Proposition.
Proposition 3.23. The function (s, ν) 7→ 〈∆ ⊗ τ(s, ν), φ〉 from (70) is holomorphic.
It is straightforward to extend this to the case where Γ (s, ν) = i d p (s, ν) · ∆ p ⊗ τ p (s, ν),
P
where the coefficients d p (s, ν) are meromorphic functions. Similar to Definition 3.3, we set
X Z
〈Γ (s, ν), φ〉 = d p (s, ν) · τ p (s, ν)(x) φ . (71)
p ∆p
By Proposition 3.23, this is meromorphic in s, ν. Notice that we can also view this as the sum
p 〈∆ p ⊗ τ p (s, ν), d p (s, ν)φ〉, allowing meromorphic coefficients in cohomology. This will be
P
27
SciPost Phys. Lect. Notes 75 (2023)
Sketch of proof of Theorem* 3.24. Consider the algebraic moment map in (39) with si = 1. By
Pℓ
Lemma 2.7, µ : R+ n
→ int(P) is a diffeomorphism, with P = i=1 ∆( f i ). This gives
Z Z
dy
ν dx
−s
f ( y) y = f (µ−1 (x))−s µ−1 (x)ν Jµ−1 Q ,
j µ (x) j
n y −1
R+ P
where Jµ−1 is the Jacobian determinant of µ−1 , and the last denominator is the product of
the entries of µ−1 . Like in Example 3.22, for values of s, ν where the integral on the right
converges, we replace P by a twisted cycle Γ (P), called the regularization of P, such that
Z Z ® ¸
d y dx dx
f ( y)−s y ν = f (µ−1 (x))−s µ−1 (x)ν Jµ−1 Q = Γ (P), Jµ−1 Q .
j µ (x) j j µ (x) j
y −1 −1
R
+
n
Γ (P)
The integral on the right side is the pairing of a holomorphic n-form with the twisted cy-
cle Γ (P). Here we need to use the analytic version of twisted (co)homology. The mani-
fold is X̃ = {x ∈ Cn : f1 (µ−1 (x)) · · · fℓ (µ−1 (x))µ−1 (x)1 · · · µ−1 (x)n ̸= 0}, and the twist is
ω̃ = dlog( f (µ−1 (x))−s µ−1 (x)ν ). The construction of the regularization Γ (P) is that in [5, Sec-
tions 3.2.4 and 3.2.5]. For this, when P is not smooth, one needs to replace the moment map
µ by that of a different toric variety, obtained by blowing up the toric variety of P in its singular
locus. The cycle Γ (s, ν) in the Theorem* is the pullback µ∗ (Γ (P)) of Γ (P) under µ.
Example 3.25. We have seen two integral formulas for B(ν, 1 − s) in Example 1.10:
1 1
xν xν yν
Z Z Z
dx dx dy
= = . (72)
0 (1 − x) x
s
0 (1 − x)s−1 x(1 − x) R+ (1 + y) y
s̃
The coordinate transformation is µ : R+ → (0, 1), with x = µ( y) = y(1 + y)−1 . This is the
moment map from Lemma 2.7 up to scaling by s̃. Its complexification µC is an isomorphism
µC
X y = C \ {0, −1} −→ C \ {0, 1} = X x .
The cocycle dx/(x(1− x)) pulls back to d y/ y under µ. A twisted chain Γ =∆⊗τ∈ C1 (X x , −ω x )
is naturally pulled back to a cycle µ∗ (Γ ) ∈ C1 (X y , −ω y ) via
µ∗ (Γ ) = µ−1 (∆) ⊗ (τ ◦ µ) .
As usual, this definition for simplices is extended linearly to C1 (X x , −ω x ). With this notation
in place, it is easy to check that for any Γ ∈ C1 (X x , −ω x ), we have
xν yν
Z Z
dx dx dy dy
· ·
∗
Γ, = = = µ (Γ ), .
Γ (1 − x) x µ∗ (Γ ) (1 + y)
x(1 − x) s s̃ y y
By Example 3.22, if we pick Γ as in (56) (with s replaced by s − 1), the left integral is a
meromorphic function in (s, ν) which agrees with B(ν, 1 − s) if 0 < ν < ν − s + 1. Hence, the
regularization of R+ is µ∗ (Γ ). It depends on s̃, ν as explained above. ⋄
28
SciPost Phys. Lect. Notes 75 (2023)
Example 4.1. We have seen in Example 3.22 that the beta function B(ν, 1 − s) is given by
xν
Z
dx
I (s, ν) = B(ν, 1 − s) = , (73)
Γ (1 − x) x
s
where Γ is the twisted 1-cycle from (56). This agrees with the integral over (0, 1) in (2) when
Re(s) > 0 and 0 < Re(ν) < Re(ν) − Re(s) + 1. The shift operator in s, denoted σs , acts by
σs • I (s, ν) = I (s + 1, ν) .
Similarly, the action of σν is σν • I (s, ν) = I (s, ν + 1). We claim that the shift operators
S1 = 1 − σs (1 − σν ) , and S2 = ν + s σν σs , (74)
xν x · x ν dx (1 − x) · x ν dx
Z Z Z
dx
σs (1 − σν ) • I = σs • − = .
Γ (1 − x) x Γ (1 − x) x Γ (1 − x)
s s s+1 x
xν ν xν
Z Z
s
S2 • I = + dx = ∇ω (1) = 0 .
Γ (1 − x) Γ (1 − x)
s x 1− x s
The goal of Section 4.1 is to derive operators like (74) for general Euler integrals. These
operators appeared in [8, Section 3.1], [3, Section 3] and [47]. Shift operators for Feynman
integrals were studied in [10]. Section 4.2 discusses differential operators. For that, we must
view our Euler integrals as functions of a new set of parameters: The coefficients of f i .
Example 4.2. Fix two generic complex numbers s, ν ∈ C. We modify the beta integral (2) by
introducing complex valued parameters z1 , z2 for the coefficients of the denominator f :
xν
Z
dx
I (z1 , z2 ) = .
Γ (z1 + z2 x) x
s
The dependence on z = (z1 , z2 ) is subtle. For instance, the very affine variety X from (29)
depends on z and, necessarily, so does Γ . In this example, one can think about I (z1 , z2 ) as a
function on a small neighborhood of (z1 , z2 ) = (1, −1), which corresponds to our original beta
integral. In that neighborhood, one can modify Γ by keeping the 1-simplices in Example 3.1
fixed, and varying the sections in Example 3.2 with z1 , z2 . For instance, the initial condition
29
SciPost Phys. Lect. Notes 75 (2023)
τ0,0 (ϵ; z1 , z2 ) = ζ(z1 , z2 ) from (49) is given by ζ(z1 , z2 ) = exp(−s log(z1 + z2 ϵ) + ν log ϵ), and
for the first log we use the analytic continuation of the positive branch near 1 − ϵ.
The differential operator ∂zi acts by partial derivation in zi , for i = 1, 2, and rational func-
tions in z act by multiplication. Here are two annihilating operators for I (z1 , z2 ):
−sz1 · x ν dx −sz2 x · x ν dx
Z Z
(z1 ∂z1 + z2 ∂z2 ) • I (z) = + = −s · I (z) .
Γ (z1 + z2 x) Γ (z1 + z2 x)
s+1 x s+1 x
xν ν
Z
sz2
0 = P2 • I (z) = ∇ω(z) (1) , with ω(z) = − dx . ⋄
Γ (z1 + z2 x) x z1 + z2 x
s
The differential operators (75) form an A-hypergeometric system or GKZ system of linear par-
tial differential equations. Such systems were introduced by Gelfand, Kapranov and Zelevinsky
to study A-hypergeometric functions [30,31]. We will introduce these systems and recall their
relation to Euler integrals in Section 4.2. For a recent survey, see [56].
Here e j is the j-th standard basis vector. The reader should check the passages from (76) to
(77) and (78) to (79) carefully. The expressions (77) and (79) show that the action of the
shift operators can be viewed as an action on cohomology H n (X , ω(s, ν)). For instance,
σsi • [φ(s, ν)] = [ f i−1 φ(s + ei , ν)] , σν j • [φ(s, ν)] = [x j φ(s, ν + e j )] . (80)
In [47], this action is defined on a cohomology vector space with coefficients in C(s, ν). We
will also use the inverses σs−1 , σν−1 of these shift operators. Their action is straightforward to
i j
define. The variables si and ν j act on I (s, ν) by multiplication: si • I (s, ν) = si I (s, ν), and
ν j • I (s, ν) = ν j I (s, ν). Notice that the operators si and σsi do not commute:
30
SciPost Phys. Lect. Notes 75 (2023)
Definition 4.3. The ring of difference operators R = C(s, ν)〈σs±1 , . . . , σs±1 , σν±1 , . . . , σν±1 〉 is the
1 ℓ 1 n
C(s, ν)-vector space with basis σsa σνb , where (a, b) ∈ Zℓ+n . I.e., it consists of finite sums
X X
g a,b (s, ν) σsa1 · · · σsaℓ σνb1 · · · σνbn = g a,b (s, ν) σsa σνb .
1 ℓ 1 n
(a,b)∈Zℓ+n (a,b)∈Zℓ+n
The product is subject to the following relations. For any rational function g(s, ν) ∈ C(s, ν),
[σs±1 , g(s, ν)] = (g(s ± ei , ν) − g(s, ν))σsi , and [σν±1 , g(s, ν)] = (g(s, ν ± e j ) − g(s, ν))σν j .
i j
Proof. WeP need to show that the operators (82) and (83) annihilate IΓ (s, ν). Let
f i (x) = α ci,α · x α , where ci,α ∈ C for i = 1, 2, . . . , ℓ. Hence, using (79), we find
X
α dx dx
·
f i (σν ) • IΓ (s, ν) = Γ , ci,α · x = Γ , f i (x) . (84)
α
x x
To show that this is zero for any cycle Γ , we need to show that the n-form on the right is exact
(Lemma 3.9). That is, we need to find ψ such that it equals ∇ω (ψ). The solution is
∂ fi
νj − 1 Xℓ
∂ xj (x) dx
dx ĵ
− si = ∇ω (−1) j−1
,
xj i=1
f i (x) x x1 · · · x n
31
SciPost Phys. Lect. Notes 75 (2023)
Example 4.5. The operators S1 and σν−1 S2 from (74) are (82) and (83) for n = ℓ = 1 and
f = 1 − x. We can use these operators to obtain the relation (63). Observe that
s · (1 − σs + σs σν ) − (ν + sσν σs ) = s − sσs − ν ∈ J .
dx dx ν dx
σν • = (1 − σs−1 ) • = • .
x x 1+ν−s x
Now, we use these identities to write σsa σνb •[dx/x] = [x b /(1− x)a ·dx/x] in terms of [dx/x]:
dx dx dx
ν ν+1 ν
σsa σνb • = σsa σνb−1 1+ν−s • = σsa σνb−2 2+ν−s 1+ν−s • = ...
x x x
Here the second equality uses the commutation rules. Suppose b, a ∈ Z>0 are positive. After
repeating this step b times, we begin expanding σsa :
dx (ν) b dx (ν) b dx
s−ν
σsa σνb • = σsa • = σsa−1 •
x (1 + ν − s) b x (ν − s)(1 + ν − s) b−1 s x
(ν) b dx
s−ν+1
s−ν
= σsa−2 •
(ν − s − 1)(ν − s)(1 + ν − s) b−2 s+1 s
x
(ν) b 1 dx
= ··· = • .
(1 + ν − s) b−a (−s)(−s − 1) · · · (−s − a + 1) x
The rational function we obtain is precisely that of (63). Notice that the numerator factors
(s − ν), (s − ν + 1), . . . cancel with the denominators (ν − s), (ν − s − 1), . . . , and the minus
signs are absorbed in the denominator factors s, s + 1, . . .. The reader should check that if a, b
satisfy different sign conditions, we arrive at the same formula. ⋄
Example 4.5 illustrates the concept of contiguity relations for the Euler beta integral. In gen-
eral, the action of shift operators can be captured by contiguity matrices. Let [φ1 ], . . . , [φχ ] be
a basis for the twisted cohomology H n (X , ω(s, ν)) (for generic s, ν). There are χ = (−1)n ·χ(X )
such basis elements by Corollary 3.19. We assume the [φk ] have coefficients that are rational
functions in s, ν. There are χ × χ-matrices Csi , Cν j such that
〈Γ , φ1 〉 〈Γ , φ1 〉 〈Γ , φ1 〉 〈Γ , φ1 〉
. . . .
σsi • .. = Csi · .. , σν j • .. = Cν j · .. .
〈Γ , φχ 〉 〈Γ , φχ 〉 〈Γ , φχ 〉 〈Γ , φχ 〉
Here σsi and σν j act entry-wise on vectors, and the contiguity matrices Csi , Cν j have entries
in C(s, ν). For more on these matrices and how to compute them, see [47, Section 5].
Remark 4.6. In the context of Feynman integrals, special choices of contiguity relations are
known as dimension-shift identities, because they relate Feynman integrals evaluated in differ-
ent space-time dimensions. We refer to [61, Section 6.2] for more details.
32
SciPost Phys. Lect. Notes 75 (2023)
Remark 4.7. Shift relations (80) give a practical way of enlarging the Nilsson-Passare domain
of convergence of Euler integrals in (s, ν), as explained
R ∞ in Section 1.2. For instance, the shift
operator σν−1 ν − s̃σs̃ annihilates the beta integral 0 y ν /(1 + y)s̃ d y/ y. Hence
y ν−1 d y yν
Z Z
dy
(ν − 1) · = s̃ · . (85)
R+ (1 + y)s̃ y R+ (1 + y)s̃+1 y
This is an equality of meromorphic functions after replacing R+ with its regularizer from The-
orem* 3.24. Suppose s̃ ∈ R+ . The convergence region from Theorem 1.6 for the left hand side
is Re(ν) ∈ (1, s̃ + 1), while that of the right hand side is Re(ν) ∈ (0, s̃ + 1). Hence, (85) is used
to evaluate the meromorphic continuation from Theorem 1.11 for Re(ν) ∈ (0, 1).
Here Ai = supp( f i ) ∈ Zn is the support of f i , in the sense of Definition 1.1. We fix complex
parameters s = (s1 , . . . , sℓ ) ∈ Cℓ and ν = (ν1 , . . . , νn ) ∈ Cn . The variety X defined in (29) is
also dependent on z. For this reason, we write X z instead of X . The Euler integral (1), seen as
a function of the coefficients z = (zi,α )i,α , defines a holomorphic function on an open subset
U of coefficient space CA = CA1 × · · · × CAℓ . To define this function, we need to specify how
the twisted integration cycle Γ varies with z. For a fixed set of coefficients z ∗ = (zi,α ∗
)i,α ∈ CA,
let [Γ (z ∗ )] ∈ H n (X z ∗ ; −ω(z ∗ )) be a twisted cycle. The open set U is a sufficiently small neigh-
borhood U of z ∗ . The choice [Γ (z ∗ )] gives rise to a family of cycles [Γ (z)] ∈ H n (X z ; −ω(z))
defined for z ∈ U. For this, we fix the singular n-simplices, and vary the sections of L−ω(z) in
the only sensible way. That is, Γ (z ∗ ) = p d p · ∆ p ⊗ τ p (z ∗ ), where τ p (z ∗ ) ∈ L−ω(z ∗ ) is a branch
P
Here d p ∈ C are constants. This was illustrated for the beta integral in Example 4.2. It is
crucial that, with this construction, the twisted boundary ∂ω(z) (Γ (z)) is zero for all z ∈ U.
is holomorphic on U.
The theorem follows from the definition of a holomorphic function and differentiation under
the integral sign [42, Chapter XVII, Theorem 8.2].
33
SciPost Phys. Lect. Notes 75 (2023)
Our goal is to derive a system of differential equations satisfied by IΓ (z). This is an example
of a class of such systems, called GKZ systems (after Gelfand, Kapranov and Zelevinsky) or A-
hypergeometric systems. We introduce these in general, and then specialize to our integrals.
The proof of the main theorem in this section (Theorem 4.12) uses the theory of D-modules.
Here the ring D is the Weyl algebra, which plays an analogous role as that of the ring R of
difference operators (see Section 4.1). To state Theorem 4.12, it is unnecessary to introduce
D-modules. We refer the interested reader to [22] for a nice introduction.
Let d be a positive integer and let A ⊂ Zd be a finite subset. To each α ∈ A, we associate
a complex variable zα and a partial derivative operator ∂α = ∂∂z . For a function f (z) of
α
∂f
z = (zα )α∈A, its partial derivative ∂ z (z) with respect to zα is denoted by ∂α f (z). The toric
α
ideal IA ⊂ C[∂α , α ∈ A] is an ideal generated by all binomials
Y Y
∂αuα − ∂αvα ,
α∈A α∈A
where u = (uα )α∈A, v = (vα )α∈A ∈ NA are such that A · (u − v) = 0. That is,
X X
uα α = vα α . (89)
α∈A α∈A
Of course, there are infinitely many integer vectors u − v satisfying A· (u − v), but finitely many
suffice to generate the ideal IA. Fix a vector β ∈ Cd . The GKZ system HA(β) associated to A
and β is the following system of partial differential equations in f (z):
X
HA(β) : zα ∂α • f (z)α − f (z)β = 0 , and P(∂ ) • f (z) = 0 , for P(∂ ) ∈ IA . (90)
α∈A
Note that the first equation of (90) is an identity of vectors in Cd , and it is enough to check
that P(∂ ) f = 0 for a finite set of generators of IA. On an open subset U ⊂ CA, we define the
space of solutions SolHA(β) (U) of HA(β) as the complex vector space
Our Euler integral (88) satisfies the GKZ system specified by the following parameters. Set
d = ℓ + n. The Cayley configuration A ⊂ Zd of A1 , . . . , Aℓ is
A = {(ei , α) : α ∈ Ai , i = 1, . . . , ℓ} ⊂ Zd . (91)
Proposition 4.9. For A, β as above, and IΓ (z) as in (88), we have IΓ (z) ∈ SolHA(β) (U).
Before proving Proposition 4.9, we encourage the reader to check that (75) is the GKZ
system for the beta integral. In that example, the toric ideal IA is 0. Notice that Proposition
4.9 is independent of the choice of cycle Γ (z).
Proof of Proposition
4.9. By the definition of the Cayley configuration (91), the constraint (89)
for u = ui,α i=1,...,ℓ , v = (vi,α )i=1,...,ℓ takes the following form :
α∈Ai α∈Ai
X X ℓ X
X ℓ X
X
ui = ui,α = vi,α (i = 1, . . . , ℓ) , and ui,α α = vi,α α . (92)
α∈Ai α∈Ai i=1 α∈Ai i=1 α∈Ai
34
SciPost Phys. Lect. Notes 75 (2023)
xα .
−s1 −si −1 −sℓ
∂(ei ,α) • f −s = −si f1 · · · fi · · · fℓ (93)
Qℓ u Qℓ v
∂(ei,α,α) − ∂(ei,α,α) . We calculate
Q Q
The operators generating IA are i=1 α∈Ai i=1 α∈Ai
i i
ℓ Y ℓ uY
i −1
Z
ui,α α dx
Pℓ
ν+
P
u
Y Y
−s1 −u1 −sℓ −uℓ
∂(ei,α,α) • IΓ (z) = (−si − j) f1 · · · fℓ x i=1 α∈Ai
.
i=1 α∈Ai
i
i=1 j=0 Γ (z) x
Qℓ Q v
Doing the same for i=1 α∈Ai ∂(ei,α,α) and applying (92) we see that IA annihilates IΓ (z).
i
It remains to verify that the other operators in the GKZ system annihilate IΓ (z) as well:
X
zα ∂α • IΓ (z)α − IΓ (z)β = 0 . (94)
α∈A
For any j = 1, . . . , n, the (ℓ + j)-th entry of (94) is derived by differentiating under the integral
sign and observing that the result is the pairing of Γ (z) with the exact n-form
ℓ
!
∧ · · · ∧ · · · ∧
dx 1 dx j−1 dx j+1 dx n
X X dx
∇ω (−1) j−1 =− si f i−1 zi,α x α α j + ν j .
x 1 · · · x j−1 x j+1 · · · x n i=1 α∈A
x
i
Here, α j is the j-th entry of α. This generalizes what happened for P2 in Example 4.2.
Proposition 4.9 implies the following homogeneity property for the function IΓ (z).
Proof. Suppose f (z) satisfies (95). We fix any 1 ≤ i ≤ d. Taking the derivative of (95)
with respect to ui and substituting u = (1, . . . , 1), we obtain the i-th entry of the identity
(Aθ − β) • f (z) = 0. For the other direction, suppose f (z) is annihilated by Aθ − β. To prove
(95), it is enough to prove it for u(i) = ui · ei with ui ∈ C∗ , and ei the standard basis vector.
For any z ∈ U, the functions φ1 : ui 7→ f (u(i)α1 z1 , . . . , u(i)αN zN ) and φ2 : ui 7→ u(i)β f (z) are
both annihilated by ui ∂∂u − βi . We also have φ1 (1) = φ2 (1). Unique solvability of the initial
i
value problem of an O.D.E. implies φ1 = φ2 , so (95) holds for u = u(i).
35
SciPost Phys. Lect. Notes 75 (2023)
Example 4.11. We revisit the Euler integral (7) of Example 0.1. The differential equations
satisfied by this integral as a function of t 1 , t 2 , t 3 are the running example of [39]. In this case,
they can be derived in an easy way from the GKZ differential equations for
ν ν ν
x1 1 x2 2 x3 3
Z
dx 1 dx 2 dx 3
IΓ (z) = .
Γ (z1 x 1 + z2 x 2 + z3 x 3 + z4 x 2 x 3 + z5 x 1 x 3 + z6 x 1 x 2 )s x 1 x 2 x 3
1 1 1 1 1 1
1 0 0 0 1 1
.
0 1 0 1 0 1
0 0 1 1 1 0
We denote these columns by α1 , . . . , α6 ∈ Z4 , and set ∂i = ∂αi for brevity. The toric ideal IA is
generated by P̃1 = ∂1 ∂4 − ∂3 ∂6 and P̃2 = ∂2 ∂5 − ∂3 ∂6 . These operators can be found using any
computer algebra software. For instance, in Macaulay2 [33], the commands are
needsPackage "Quasidegrees" 1
A = matrix{1,1,1,1,1,1},{1,0,0,0,1,1},{0,1,0,1,0,1},{0,0,1,1,1,0} 2
D = QQ[d_1..d_6] 3
T = toricIdeal(A,D) 4
z4 z5 z6
ν ν ν
IΓ 1, 1, 1, , , = z1 1 z2 2 z3 3 IΓ (z) . (97)
z2 z3 z1 z3 z1 z2
36
SciPost Phys. Lect. Notes 75 (2023)
Here, we used (98) when passing from (99) to (100). Repeating this for the monomials in P̃1 ,
P̃2 , we find that IΓ (1, 1, 1, −t 1 , −t 2 , −t 3 ) is annihilated by the following operators:
P1 = t 1 ∂ t2 − t 3 ∂ t2 + (1 − s + ν2 + ν3 )∂ t 1 − (1 − s + ν1 + ν2 )∂ t 3 ,
1 3
P2 = t 2 ∂ t2 − t 3 ∂ t2 + (1 − s + ν1 + ν3 )∂ t 2 − (1 − s + ν1 + ν2 )∂ t 3 ,
2 3
P3 = t 1 ∂ t 1 + t 2 ∂ t 2 + t 3 ∂ t 3 + ν1 + ν2 + ν3 − s .
These operators agree with the ones in [39, Equation (2.6)] after setting s = D/2. ⋄
The local solutions of HA(β) at a point z ∗ ∈ CA are given by the direct limit
Theorem 4.12. Let β = −(s, ν) ∈ Cℓ+n be non-resonant. For any z ∗ ∈ CA, the map
H n (X z ∗ , −ω(z ∗ )) → SolHA(β),z ∗ given by [Γ (z ∗ )] 7→ IΓ (z) is a vector space isomorphism.
Remark 4.13. Theorem 4.12 implies that, for generic s, ν (in the sense of the Vanishing The-
orem 3.13), the dimension of the local solution space SolHA(β),z ∗ equals the signed Euler char-
acteristic of X z ∗ . For z ∗ outside an algebraic hypersurface {EA = 0} ⊂ CA, this number equals
the normalized volume of the convex hull of A [1, Theorem 5.15]. The polynomial EA is the
principal A-determinant, as introduced by Gelfand, Kapranov and Zelevinsky [32, Chapter 6].
While the function IΓ is an integral over Γ (z) against a particular cohomology class [ dxx ],
the pairing 〈Γ (z), φ〉 is well-defined for any twisted cocycle [φ] ∈ H n (X z , ω). Let [φ1 ],. . . ,[φχ ]
be a basis for the twisted cohomology H n (X z , ω) for z ∈ U. Again, by Corollary 3.19, there
are χ = (−1)n · χ(X z ∗ ) basis elements. We assume the [φk ] have coefficients that are rational
functions in z. There exist χ × χ-matrices Pα (α ∈ A) such that
〈Γ (z), φ1 〉 〈Γ (z), φ1 〉
.. ..
∂α • . = Pα · . . (101)
〈Γ (z), φχ 〉 〈Γ (z), φχ 〉
Here ∂α acts entry-wise on vectors. These expressions form the so-called Pfaffian system. The
Pfaffian system can be derived from a system of differential operators, like a GKZ system. The
general procedure is explained in [20, Section 3].
Remark 4.14. Pfaffian systems lead to one of the most efficient ways of evaluating Feynman
integrals [37]. In practice, (101) can be solved by providing boundary conditions 〈Γ (z ∗ ), φi 〉
for i = 1, . . . , χ at some z = z ∗ and using path-ordered exponentiation of the matrices Cα to
evaluate the Pfaffian system at other values of z. See [38] for a pedagogical introduction.
37
SciPost Phys. Lect. Notes 75 (2023)
5 Open problems
The previous sections provide an overview of the basics of Euler integrals. While this is a
classical topic, the theory is currently still very much in development. We conclude with a list
of open research problems, hoping that the reader will join this effort.
1. Evaluating integrals numerically. When the integral (13) converges, it can be evalu-
ated numerically using sector decomposition and Monte Carlo integration, see Remark 1.9.
When the real part of the exponents is large, it becomes increasingly important to con-
centrate the Monte Carlo samples in the close neighborhood of the critical point a from
Section 2.3. This could lead to effective numerical algorithms for evaluating convergent
Euler integrals, with applications in Bayesian statistics [13]. Likewise, in physics appli-
cations one often has to analytically continue in the parameters (s, ν) before numerical
evaluation, which can be achieved using the results of Sections 3.3 and 4.1.
2. Generic parameters. Theorems 2.1 and 3.13 make genericity assumptions on the expo-
nents s, ν. In Theorem 3.13 that means outside a closed algebraic subvariety, in Theorem
3.13 it means outside a countable union of hyperplanes. The former can be seen as a limit
of the latter, by driving the parameter δ from Section 2 to zero [47]. It is interesting to
describe these hyperplanes explicitly, and investigate this problem in more detail.
5. Nice bases of cohomology. There are several reasons for which it is favorable to use basis
element for cohomology which are represented by dlog forms [57]. These are regular n-
forms obtained as dlog of a rational function. Another notion of a nice basis is related
to so-called canonical differentials equations for Feynman integrals [37]. In both cases, it
would be interesting to find criteria for such bases to exist.
6. Beyond Euler integrals. While our framework deals with Euler integral defined by (1),
there are other types of integrals that resemble it. The list includes exponential integrals
[27, 45], matrix hypergeometric integrals [35], and integrals over M g,n [26, 62]. Theorems
stated in this article mostly remain unsolved for these integrals.
8. χ-Stratification. In Section 4.2, the very affine variety X z depends on the coefficients z of
the Laurent polynomials. We propose to study the loci in coefficient space on which the
Euler characteristic χ(X z ) is constant. E.g., for which z ∈ (C∗ )A is |χ(X z )| minimal?
38
SciPost Phys. Lect. Notes 75 (2023)
Acknowledgements
This article served as accompanying notes for a series of four lectures given by the third named
author at the Max Planck Institute for Mathematics in the Sciences in Leipzig. We thank Jörg
Lehnert for organizing these lectures, and Raluca Vlad for pointing out several typos in a
previous version.
Funding information This material is based upon work supported by the Sivian Fund and
the U.S. Department of Energy, Office of Science, Office of High Energy Physics under Award
Number DE-SC0009988. It is also supported by JSPS KAKENHI Grant Number 22K13930 and
partially supported by JST CREST Grant Number JP19209317.
References
[1] A. Adolphson, Hypergeometric functions and rings generated by monomials, Duke Math.
J. 73, 269 (1994), doi:10.1215/S0012-7094-94-07313-4.
[2] D. Agostini, T. Brysiewicz, C. Fevola, L. Kühne, B. Sturmfels, S. Telen and T. Lam, Likeli-
hood degenerations, Adv. Math. 414, 108863 (2023), doi:10.1016/j.aim.2023.108863.
[3] D. Agostini, C. Fevola, A.-L. Sattelberger and S. Telen, Vector spaces of generalized Euler
integrals, (arXiv preprint) doi:10.48550/arXiv.2208.08967.
[4] K. Aomoto, Les équations aux différences linéaires et les intégrales des fonctions multiformes,
J. Fac. Sci. Univ. Tokyo 22, 271 (1975).
[5] K. Aomoto, M. Kita, T. Kohno and K. Iohara, Theory of hypergeometric functions, Springer,
Tokyo, Japan, ISBN 9784431539124 (2011), doi:10.1007/978-4-431-53938-4.
[6] N. Arkani-Hamed, Y. Bai and T. Lam, Positive geometries and canonical forms, J. High
Energy Phys. 11, 039 (2017), doi:10.1007/JHEP11(2017)039.
[7] N. Arkani-Hamed, P. Benincasa and A. Postnikov, Cosmological polytopes and the wave-
function of the Universe, (arXiv preprint) doi:10.48550/arXiv.1709.02813.
[8] N. Arkani-Hamed, S. He and T. Lam, Stringy canonical forms, J. High Energy Phys. 02,
069 (2021), doi:10.1007/JHEP02(2021)069.
[9] C. Berkesch, J. Forsgård and M. Passare, Euler-Mellin integrals and A-hypergeometric func-
tions, Michigan Math. J. 63, 101 (2014), doi:10.1307/MMJ/1395234361.
[10] T. Bitoun, C. Bogner, R. P. Klausen and E. Panzer, Feynman integral relations from paramet-
ric annihilators, Lett. Math. Phys. 109, 497 (2018), doi:10.1007/s11005-018-1114-8.
[11] M. Borinsky, Tropical Monte Carlo quadrature for Feynman integrals, Ann. Inst. Henri
Poincaré Comb. Phys. Interact. (2023), doi:10.4171/AIHPD/158.
[12] M. Borinsky, H. J. Munch and F. Tellander, Tropical Feynman integration in the Minkowski
regime, Comput. Phys. Commun. 292, 108874 (2023), doi:10.1016/j.cpc.2023.108874.
[13] M. Borinsky, A.-L. Sattelberger, B. Sturmfels and S. Telen, Bayesian integrals on toric
varieties, SIAM J. Appl. Algebra Geometry 7, 77 (2023), doi:10.1137/22M1490569.
39
SciPost Phys. Lect. Notes 75 (2023)
[15] R. Britto, S. Mizera, C. Rodriguez and O. Schlotterer, Coaction and double-copy proper-
ties of configuration-space integrals at genus zero, J. High Energy Phys. 05, 053 (2021),
doi:10.1007/JHEP05(2021)053.
[16] F. Brown and C. Dupont, Single-valued integration and superstring amplitudes in genus
zero, Commun. Math. Phys. 382, 815 (2021), doi:10.1007/s00220-021-03969-4.
[19] S. Caron-Huot and A. Pokraka, Duals of Feynman integrals. Part I. Differential equations,
J. High Energy Phys. 12, 045 (2021), doi:10.1007/JHEP12(2021)045.
[21] K. Cho and K. Matsumoto, Intersection theory for twisted cohomologies and
twisted Riemann’s period relations I, Nagoya Math. J. 139, 67 (1995),
doi:10.1017/S0027763000005304.
[23] P. Deligne, Équations différentielles à points singuliers réguliers, Springer, Berlin, Heidel-
berg, Germany, ISBN 9783540363736 (2006), doi:10.1007/BFb0061194.
[24] V. S. Dotsenko and V. A. Fateev, Conformal algebra and multipoint correlation functions in
2D statistical models, Nucl. Phys. B 240, 312 (1984), doi:10.1016/0550-3213(84)90269-
4.
[25] T. Duff, C. Hill, A. Jensen, K. Lee, A. Leykin and J. Sommars, Solving polynomial sys-
tems via homotopy continuation and monodromy, IMA J. Numer. Anal. 39, 1421 (2019),
doi:10.1093/IMANUM/DRY017.
[27] J. Fresán, C. Sabbah and J.-D. Yu, Quadratic relations between periods of connections,
Tohoku Math. J. 75, 175 (2023), doi:10.2748/tmj.20211209.
[28] W. Fulton, Introduction to toric varieties, Princeton University Press, Princeton, USA, ISBN
9780691000497 (1993).
[29] E. Gawrilow and M. Joswig, polymake: A framework for analyzing convex poly-
topes, in Polytopes–combinaorics and computation, Birkhäuser, Basel, Switzerland, ISBN
9783764363512 (2000).
40
SciPost Phys. Lect. Notes 75 (2023)
[30] I. M. Gelfand, General theory of hypergeometric functions, Sov. Math. Doklady 33, 537
(1986).
[33] D. R. Grayson and M. E. Stillman, Macaulay2, a software system for research in algebraic
geometry, https://macaulay2.com/.
[34] P. Griffiths and J. Harris, Principles of algebraic geometry, Wiley & Sons, Hoboken, USA,
ISBN 9780471050599 (2014), doi:10.1002/9781118032527.
[36] G. Heinrich, Collider physics at the precision frontier, Phys. Rep. 922, 1 (2021),
doi:10.1016/j.physrep.2021.03.006.
[37] J. M. Henn, Multiloop integrals in dimensional regularization made simple, Phys. Rev. Lett.
110, 251601 (2013), doi:10.1103/PhysRevLett.110.251601.
[38] J. M. Henn, Lectures on differential equations for Feynman integrals, J. Phys. A: Math.
Theor. 48, 153001 (2015), doi:10.1088/1751-8113/48/15/153001.
[39] J. Henn, E. Pratt, A.-L. Sattelberger and S. Zoia, D-Module techniques for solv-
ing differential equations in the context of Feynman integrals, (arXiv preprint)
doi:10.48550/arXiv.2303.11105.
[40] J. Huh, The maximum likelihood degree of a very affine variety, Compositio Math. 149,
1245 (2013), doi:10.1112/S0010437X13007057.
[41] J. Huh and B. Sturmfels, Likelihood geometry, Springer, Cham, Switzerland, ISBN
9783319048697 (2014), doi:10.1007/978-3-319-04870-3_3.
[42] S. Lang, Undergraduate analysis, Springer, New York, USA, ISBN 9780387948416
(1996), doi:10.1007/978-1-4757-2698-5.
[43] R. N. Lee and A. A. Pomeransky, Critical points and number of master integrals, J. High
Energy Phys. 11, 165 (2013), doi:10.1007/JHEP11(2013)165.
[44] C. R. Mafra and O. Schlotterer, Tree-level amplitudes from the pure spinor superstring,
Phys. Rep. 1020, 1 (2023), doi:10.1016/j.physrep.2023.04.001.
[45] H. Majima, K. Matsumoto and N. Takayama, Quadratic relations for confluent hypergeo-
metric functions, Tohoku Math. J 52, 489 (2000).
[46] P. Mastrolia and S. Mizera, Feynman integrals and intersection theory, J. High Energy Phys.
02, 139 (2019), doi:10.1007/JHEP02(2019)139.
[47] S.-J. Matsubara-Heo and S. Telen, Twisted cohomology and likelihood ideals, (arXiv
preprint) doi:10.48550/arXiv.2301.13579.
41
SciPost Phys. Lect. Notes 75 (2023)
[48] K. Matsumoto, Intersection numbers for logarithmic k-forms, Osaka J. Math. 35, 873
(1998).
[49] K. Matsumoto, Relative twisted homology and cohomology groups associated with Lauri-
cella’s F D , (arXiv preprint) doi:10.48550/arXiv.1804.00366.
[51] S. Mizera, Scattering amplitudes from intersection theory, Phys. Rev. Lett. 120, 141602
(2018), doi:10.1103/PhysRevLett.120.141602.
[52] S. Mizera, Aspects of scattering amplitudes and moduli space localization, Springer, Cham,
Switzerland, ISBN 9783030530099 (2020), doi:10.1007/978-3-030-53010-5.
[53] L. Nilsson and M. Passare, Mellin transforms of multivariate rational functions, J. Geom.
Anal. 23, 24 (2013), doi:10.1007/S12220-011-9235-7.
[54] Oscar – open source computer algebra research system, version 0.12.1-dev (2023), https:
//juliapackages.com/p/oscar.
[55] E. Panzer, Hepp’s bound for Feynman graphs and matroids, Ann. Inst. Henri Poincaré
Comb. Phys. Interact. 10, 31 (2022), doi:10.4171/AIHPD/126.
[57] K. Saito, Theory of logarithmic differential forms and logarithmic vector fields, J. Fac. Sci.
Univ. Tokyo Sect. IA Math. 27, 265 (1980).
[59] B. Sturmfels and S. Telen, Likelihood equations and scattering amplitudes, Alg. Stat. 12,
167 (2021), doi:10.2140/astat.2021.12.167.
[60] A. Varchenko, Critical points of the product of powers of linear functions and families of
bases of singular vectors, Compos. Math. 97, 385 (1995).
[63] G. M. Ziegler, Lectures on polytopes, Springer, New York, USA, ISBN 9780387943299
(1994), doi:10.1007/978-1-4613-8431-1.
42