1. Introduction
The Weil conjectures, proposed by André Weil in 1949, revolutionized algebraic geometry by linking the arithmetic properties of varieties over finite fields with topological concepts through cohomology theories. For a smooth projective variety X over a finite field Fq , Weil conjectured that the zeta function of X , which encodes information about the number of rational points on X over finite extensions of Fq , is a rational function. The conjectures further predict functional properties of this zeta function, analogous to the classical Riemann Hypothesis, and inspired the development of cohomology theories in characteristic zero. These cohomology theories, referred to collectively as “Weil cohomology,” adhere to certain axioms that allow the application of the Lefschetz trace formula, which expresses the count of fixed points of an endomorphism as an alternating sum of traces of the induced maps on cohomology groups. This formulation was pivotal, as it enabled Pierre Deligne’s eventual proof of the Riemann Hypothesis for varieties over finite fields. This thesis provides nothing new but an detailed introduction to the concepts underlying the Weil conjectures and the proof of Lefschetz trace formula, with some refinements in technical proofs.
2. Weil conjectures
In this paper, we shall now primarily consider X as a smooth projective variety, unless stated otherwise. While some definitions or results could be generalized to schemes, we omit these generalizations for simplicity.Also, from now on, we fix an algebraic closure Fq; thus, there exists a unique finite extension with degree r of Fq contained in Fq, namely Fqr.
To begin, we define rational points of the algebraic variety over a perfect field. We follow the definition given by Silverman in his [1, Chapter I].
Definition 1. For a perfect field k with a fixed algebraic closure k¯, an affine space over k, namely An, is defined as the set of n-tuples
.
The K-rational points of An is the set An(K) := {(x1,...,xn) ∈ An | xi ∈ K} if the field extension k ⊂ K ⊂ k¯ is an algebraic extension. An affine variety X = V(I) ⊂ An is defined over k if the defining ideal I ⊆ k¯[x1,...,xn] is prime and could be generated by polynomials in k[x1,...,xn]. The K-rational points of X are then defined by X(K) := X ∩ An(K).
In a similar manner, the projective space over k is defined as
,
where (x0,...,xn) ∼ (y0,...,yn) if and only if there exists some λ ∈ k¯× such that yi = λxi. The K-rational points of Pn is the set Pn(K) := {[x0,...,xn] ∈ Pn | xi ∈ K}. A projective variety X = V(I) ⊂ Pn is said defined over k if the defining homogeneous ideal I ⊂ k¯[x0,...,xn] is prime and could be generated by homogenous polynomials in k[x0,...,xn]. Furthermore, we define the K-rational points of X as the set X(K) := X ∩ Pn(K).
Definition 2 ([2, Section 26]). Let X be an algebraic variety over finite field Fq, denote X(Fqr) as the Fqr-rational points of X. The zeta function of X is a generating function involving the number of these rational points:
.
Note that Z(X,t) is a formal power series in Q[[t]].
Remark. Since X could be covered by finitely many affine open subsets, we can only describe X(Fqr) in affine case which is more explicit: If is defined by the ideal (f1,...,fd) ⊆
Fq[x1,...,xn], then
Example 3. Here we give an easy but concrete example to help us have an intuitive feeling about the zeta function. Suppose, it’s clear from the remark above, that #X(Fqr) is just the number of points in
. We conclude that
.
In general, for the projective case, we can obtain
via identifying .
As illustrated in the above example, zeta functions exhibit rationality for affine and projective spaces. More generally, Weil’s conjectures reveal that rationality is a property satisfied by a broad class of varieties.
Theorem 4 (Weil conjectures [3, Section 2.4]). Suppose X is a smooth projective variety over Fq with dim(X) = n, then it has the following properties
(Rationality): Z(X,t) ∈ Q(t), i.e., the zeta function of X is a rational function.
(Functional equation): if E = (∆2) is the self-intersection number of the diagonal ∆ ⊆ X × X, then
.
Here, E is also known as the Euler-Poincaré characteristic and denoted as χ in some references.
(Analogue of Riemann hypothesis): We can rewrite Z(X,t) as
where P0(t) = 1 − t, P2n(t) = 1 − qnt. Moreover, for all 1 ≤ i ≤ 2n − 1, there exists algebraic integers αi,j with absolute value qi/2 such that Pi(t) = Qj(1 − αi,jt)
(Betti number): Define the ith Betti number of X as bi(X) := deg(Pi(t)). Then we have is a reduction of a variety Y defined over the ring of integers of a number field K (so its ring of integers denoted as OK) modulo a prime ideal, then bi(X) is equal to ith Betti number of the analytification of Y ×spec(OK) spec(C).
The proof of the Weil conjectures was achieved through a series of developments, ultimately culminating in Pierre Deligne’s proof in 1974. Initially, Weil himself established rationality for specific cases such as curves and abelian varieties [4]. Dwork successfully proved both rationality (general case) and functional equation in 1960 using p-adic analysis method, see [5]. Inspired by algebraic topology and the Lefschetz fixed point theorem, Weil proposed that the conjectures may follow from an appropriate cohomology theory for varieties over finite fields, with coefficients in characteristic-zero field. However, the proof for other two statements was still open until around 1965, Grothendieck and Artin introduced ℓ-adic cohomology (a variant of étale cohomology) based on the initial ideals of Serre, in order to tackle with rationality, functional equation and Betti number. The general property of étale cohomology allowed Grothendieck to give a Lefschetz fixed-point formula for ℓ-adic cohomology. Finally, building on these developments, Deligne gave the proof for the analogue of Riemann hypothesis in 1974 [6]. As we have seen, the effort to prove these conjectures led to the development of new areas in mathematics, including étale cohomology and ℓ-adic cohomology. Although these topics go beyond the scope of this thesis, interested readers may refer to [7] (SGA) by Grothendieck for a comprehensive introduction.
3. Weil cohomology
Although we will not delve into the details of étale cohomology, this section will present a broader perspective on cohomology theories. The Weil cohomology theory is a general concept of cohomology like ℓ-adic one, it’s a certain class of cohomology theories that satisfy some axioms (resembling from singular cohomology over C). One of the most crucial properties of it is the existence of the Lefschetz trace formula (let’s use this name to replace "Lefschetz fixed-point theorem" to emphasize trace). We’ll also see how the rationality followed easily after applying the Lefschetz trace formula. First, let’s state the definition of Weil cohomology, which can be found in [8], [9] or [3] modulo some simplifications.
Definition 5. A Weil cohomology theory is a family of contravariant functors
Hi : {smooth projective variety over k} −→ {K − vector space},i ∈ Z
where k, K are fixed fields such that k is algebraically closed and char(K) = 0. Also, we denote H∗(X) as the graded K-vector space Li≥0 Hi(X), denote f∗ as the induced (grad-preserving) morphism given by f. Moreover, there are some given data (For the smooth projective varieties X, Y appear below, assume dim(X) = n, dim(Y ) = m.)
(D1) (Cup product) There exists a map ⌣: Hi(X)×Hj(X) → Hi+j(X) for any i,j makes H∗(X) a graded-commutative K-algebra, i.e., a ⌣ b = (−1)ijb ⌣ a if a ∈ Hi(X),b ∈ Hj(X). (By an abuse of the notation, we write ⌣ (a,b) as a ⌣ b.)
(D2) (Trace map) There exists a linear trace map TrX : H2n(X) → K.
(D3) (Cohomology class of cycles) For any closed subvariety Z ⊂ X of codimension c, there is a cohomology class cl(Z) ∈ H2c(X) given to Z.
Furthermore, these data should satisfy a sequence of axioms
(A1) (Finiteness) Any Hi(X) is finite-dimensional as a K-vector space. In addition, Hj(X) = 0 if j /∈ [0,2n].
(A2) (Poincaré duality) The trace map TrX is an isomorphism. For 0 ≤ i ≤ 2n, the composition
TrX◦ ⌣: Hi(X) × H2n−i(X) −→ K,(a,b) 7−→ TrX(a ⌣ b)
is K-bilinear and a perfect pairing.
(A3) (Künneth formula) Let pX,pY be the projective map from X × Y to X,Y respectively. The cup product then induces a K-algebra homomorphism
which is required to be an isomorphism.
(A4) (Case of a point) For P = spec(k), then cl(P) = 1 and TrP (1) = 1.
(A5) (Compatibility of trace map and cup product) For any a ∈ H2n(X),b ∈ H2m(Y ),
TrX×Y (p∗Xa ⌣ p∗Y b) = TrX(a) · TrY (b).
This indicates that the trace map is multiplicative with respect to cup product.
(A6) (Exterior product of cohomology classes) If Z ⊂ X,W ⊂ Y are closed subvarieties, then
cl(Z × W) = p∗X(cl(Z)) ⌣ p∗Y (cl(W)).
Here, is obtained from cl(Z)⊗ cl(W) via the map defined in (A3).
So, (A6) actually says the class map should be compatible with the Künneth formula.
(A7) (Push-forward of cohomology classes) Let f : X → Y be a morphism between X and Y . If Z ⊂ X is a closed subvariety, then for any class ω ∈ H2m(Y ) one has
TrX(cl(Z) ⌣ f∗(ω)) = deg(Z/f(Z))TrY (cl(f(Z)) ⌣ ω).
(A8) (Pull-back of cohomology classes) If morphism f : X → Y and closed subvariety W ⊂ Y satisfying :
f−(W) has pure dimension dim(W)+n−m, i.e., all irreducible components Z1,...,Zr of f−1(W) have dimension dim(W) + n − m
Either f is flat in an open neighbourhood of W, or W is generically transverse to f, i.e., f−1(W) is generically reduced.
Also, assume [f−1(W)]1 = Pri=1 miZi as a cycle of codimension m − dim(W), then
r
f∗(cl(W)) = Xmicl(Zi).
i=1
Remark. From the definition of Weil cohomology, we know f∗ acts as a pull-back map on the cohomology group when f is a morphism between smooth projective varieties X,Y with dimension n,m respectively. We can also define a push-forward map for f using Poincaré duality. For ω ∈ Hi(X), we define f∗ω ∈ H2m−2n+i(Y ) such that
TrY (f∗ω ⌣ µ) = TrX(ω ⌣ f∗µ)
for any µ ∈ H2n−i(Y ). Note this definition is well-defined due to the Poincaré duality. Indeed, as TrX(ω ⌣ f∗(·)) ∈ Hom(H2n−i(Y ),K), uniqueness of f∗ω is given by the isomorphism H2m−2n+i(Y ) ∼= Hom(H2n−i(Y ),K).
Those familiar with differential topology may notice several parallels between Weil cohomology and classical de Rham cohomology on compact manifolds. For instance, [ω] ⌣ [µ] = [ω ∧ µ] for [ω] ∈ HdRi (M),[µ] ∈ HdRj (M) for ω,µ on a compact smooth manifold M. The trace map could be interpreted as the integration map: ω 7→ RM ω. It is remarkable that if X is a smooth projective variety over some field k embedded in C, then X(C) could be regarded as a compact complex manifold [11]. Next we give some famous examples of Weil cohomology.
Example 6. The first basic example of Weil cohomology is the singular cohomology over C. Moreover, one may expect an analogy of de Rham theorem, if k = C, then the algebraic de
Rham cohomology is isomorphic to singular cohomology of analytification of X, we refer this to Grothendieck’s work [12]. In general, if char(X) = 0, the algebraic de Rham cohomology of X is always a Weil cohomology [13]. Later, we’ll define what’s an algebraic de Rham cohomology for smooth affine varieties and see how it fails to be a well-behaved cohomology when char(k) > 0.
Example 7 ([14]). For the case char(k) > 0, one classical Weil cohomology is the ℓ-adic cohomology, where ℓ is a prime different with char(k) and K = Qℓ.
3.1. Lefschetz trace formula
Now, it’s time to present our Lefschetz trace formula, which is a key result in Weil cohomology theory. Roughly speaking, it states that the number of fixed points of an endomorphism on X could be represented as the alternating sum of the traces of induced linear map on cohomology groups.
Theorem 8 (Lefschetz trace formula). Let X be a smooth projective variety of dimension n and let f : X → X be an endomorphism, then for any Weil cohomology H∗, we have
,
where ∆ = {(x,y) ∈ X × X | x = y} denote the diagonal, Γf = {(x,f(x)) ∈ X × X | x ∈ X} denote the graph of f.
Remark. Here Γf · ∆ represents the intersection number (counted with multiplicities) of the graph and the diagonal. In particular, if they intersect transversely, this number is euqal to |{x ∈ X | f(x) = x}|, the count of fixed points by f.
While the proof of the trace formula is somewhat tedious and involved, it is worthwhile to go through it as it provides a deeper understanding of Weil cohomology. We will divide the proof into several steps. First, establish some useful lemmas derived from the definition of Weil cohomology. These can also be found in de Jong’s note [9].
Lemma 9. Assume X,Y are smooth projective varieties over k with dimension n,m respectively and f : X → Y is an arbitrary morphism. H∗ is any Weil cohomology over a field K. Then the following properties valid.
The K-algebra homomorphism K → H0(X) given by K-algebra structure is an isomorphism. Hence, if α ∈ H0(X),β ∈ Hi(X), α ⌣ β = α · β by regarding α as an element in K.
cl(X) = 1 ∈ H0(X).
f∗(α ⌣ f∗γ) = f∗α ⌣ γ for any α ∈ Hi(X),γ ∈ Hj(Y ). It’s also called the projection formula.
For any closed subvariety Z ⊂ X, f∗(cl(Z)) = deg(Z/f(Z)) · cl(f(Z)).
For, and equal to 0 otherwise.
For α ∈ Hi(X), one has
where p1,p2 are the projections from X ×X to the first and the second coordinate, respectively.
Proof. (i). Using (A2) (Poincaré duality) for i = 0, we obtain H0(X) ∼= (H2n(X))∗. Since TrX is an isomorphism, we get dimK(H2n(X)) = dimK(K) = 1. Hence, the dimension of H0(X) is also 1. Moreover, the map K → H0(X) is injective since K is a field, hence bijective. (ii). Applying (A8) to the obvious map p : X → spec(k), we obtain cl(X) = p∗(cl(spec(k))) = 1 as cl(spec(k)) = 1 by (A4).
(iii). First, by the definition of push-forward of f, we know TrY (f∗(α ⌣ f∗γ) ⌣ β) = TrX((α ⌣ f∗γ) ⌣ f∗β) for any β ∈ H2n−(i+j)(Y ). Using the associativity and the definition of pushforward again, we have
TrX((α ⌣ f∗γ) ⌣ f∗β) = TrX(α ⌣ (f∗γ ⌣ f∗β)) = TrY (f∗α ⌣ (γ ⌣ β)) = TrY ((f∗α ⌣ γ) ⌣ β).
Therefore, for any β ∈ H2n−(i+j)(Y ), TrY (f∗(α ⌣ f∗γ) ⌣ β) = TrY ((f∗α ⌣ γ) ⌣ β). We conclude f∗(α ⌣ f∗γ) = f∗α ⌣ γ via (A2).
(iv) The proof is almost the same as (iii) and use (A7) for the middle step.
(v). Note pX∗(p∗Y β) ∈ Hj−2m(X). So, (A1) tells us that pX∗(p∗Y β) = 0 automatically once j ̸= 2m. Indeed, if j < 2m, Hj−2m(X) = 0; if j > 2m, Hj(Y ) = 0 and then β = 0. For the case j = 2m, pX∗(p∗Y β) ∈ H0(X) ∼= K. Therefore, for any α ∈ H2n(X), one has pX∗(p∗Y β) · TrX(α) = TrX(pX∗(p∗Y β) ⌣ α) = TrX×Y (p∗Y β ⌣ p∗Xα) = TrY (β) · TrX(α).
The first equality follows by the linearity of trace map and (i); the second one comes from the definition of push-forward of pX∗; the last one just by (A5). It follows that pX∗(p∗Y β) = TrY (β). (vi). Define φ : X ,→ X × X,x 7→ (x,f(x)) as the embedding of the graph of f. Clearly p1 ◦ φ = idX,p2 ◦ φ = f. Moreover, as φ is an isomorphism between X,Γf, deg(X/φ(X)) = 1.
It follows that φ∗(cl(X)) = cl(Γf) by (iv). Therefore,
Note the second equality given by (iii) and the last two just follow from the functoriality of H∗ and cl(X) = 1.
With these properties of Weil cohomology, we can now proceed with linear algebra manipulations.
Proposition 10 (Lemma 4.10, [3]). Following the setting in Theorem 8. Let {wj2n−r}j=1,...,kr be a basis for H2n−r(X), ∀r. Thanks to Poincaré duality (A2), we know Hr(X) ∼= (H2n−r(X))∗ via the map ω 7→ TrX((·) ⌣ ω). Therefore, there exists a dual basis {vir}i=1,...,kr for Hr(X) such that TrX(wj2n−r ⌣ vir) = δi,j. Under these assumptions, we have
cl.
Proof. Using the Künneth formula (A3), it allows us to write cl(Γf) in the following way:
cl
for some unique µtj ∈ Ht(X). Indeed, as cl(Γf) ∈ H2n(X × X), (A3) shows each term of cl(Γf) comes from µ ⊗ ω ∈ Ht(X) ⊗ H2n−t(X),∀t. As we already fix a basis {wj2n−t} for H2n−t(X), each µtj is unique clearly. On the other hand, using Lemma 9 (iii), (vi) and substitute cl(Γf), we obtain
for any fixed r and i. One remarkable is that in order to use (iii), we should commute the terms for both sides of the third equality. Now, using Lemma 9 (v), we know
if t ̸= r and. Hence, f∗vir = Pj δij(µrj) = µri , we’re done.
Before moving to the final proof of the Lefschetz trace formula, we need one last piece from algebraic geometry. We’ll just state and use it without proof.
Lemma 11 (Corollary 4.6, [3]). Let X be a smooth projective variety, if αi ∈ Ani(X) for i =
1,...,k such that= dim(X). Then, the intersection number of αi
α1 · ... · αk = TrX(cc(α1) ⌣ ... ⌣ cc(αk))
where cc is the cycle class map.
Finally, we conclude the proof of the Lefschetz trace formula with these elements.
Proof. (of Theorem 8) Recall the Proposition 10, for any t, if we replace the basis of Ht(X) by
{wjt}j=1,...,kt, then the corresponding dual basis in H2n−t(X) should be due to the graded-commutativity of H∗(X). It follows a similar result with Proposition 10, apply it to f = idX we obtain cl
.
Using Lemma 11 and substitute cl(∆), cl(Γf) we have
Γf · ∆ = TrX×X(cl(Γf) ⌣ cl(∆))
= TrX×X X (−1)t+t(2n−r)p∗1(f∗vir ⌣ wjt) ⌣ p∗2(wi2n−r ⌣ vj2n−t)
r,t,i,j
= X TrX(f∗vir ⌣ wi2n−r) · TrX(wi2n−r ⌣ vir) = X TrX(f∗vir ⌣ wi2n−r).
r,ir,i
The third equality holds by observing that the terms with t ̸= 2n−r,j ̸= i are vanishing via (A1) and TrX(wj2n−r ⌣ vir) = δi,j. Fix r, assume for some λi,s ∈ K, then
,
where the second equation follows from Tr . We’re done!
This leads us to an immediate application of the Lefschetz trace formula in proving the rationality aspect of the Weil conjecture.
Theorem 12. Assume φ : X → X is the q-power Frobenius endomorphism of smooth projective variety X with dimension n over Fq, then
.
where Pi(t) = det(1 − tHi(φ))−1. (Hi(φ) is an abbreviation for (φ∗ | Hi(X)).)
Proof. Some simple field theories tell us that Fqr could be described as the subfield of Fq that fixed by r many q-power Frobenius automorphism, i.e., Fqr = {x ∈ Fq | xqr = x}. Therefore,
#X(Fqr) = |{x ∈ X | φr(x) = x}| = Γφr · ∆,
the second equality is given by the fact that the graph of φr intersects transversely with the diagonal, see [3, Prop. 2.4]. Apply Lefschetz trace formula and substitute into the zeta function, one may expect
.
To conclude, we need a lemma from linear algebra: Let F be a linear endomorphism over a finite dimensional K-vector space V for some field K, then .
To see this, we induct on dim(V ). The case dim(V ) = 1 is trivial. For the general case, note we can always assume K is algebraically closed since the lemma doesn’t depend on the ground field. Therefore, we can assume F has an eigenvector v ∈ V and then rewrite the matrix of F as a block matrix M1 ⊕ M2, where M1 is the matrix for F|span(v) and M2 is the one for F|V/span(v). By the inductive hypothesis,
the lemma is proved. Finally, substitute F with Hi(φ), the theorem then follows by our lemma.
4. Conclusion
In this thesis, we explored the Weil conjecture and its profound implications for algebraic geometry, particularly through the application of Weil cohomology theories and the Lefschetz trace formula. By bridging the arithmetic properties of varieties over finite fields with topological concepts, the Weil conjecture has reshaped our understanding of cohomology, zeta functions, and their interconnections. The Lefschetz trace formula, in particular, provides a powerful mechanism for point-counting, revealing the intricate structure of rational points on varieties. This foundational framework not only contributed to Pierre Deligne’s proof of the Riemann Hypothesis for varieties over finite fields but also catalyzed the development of new cohomological techniques that continue to influence research in number theory and algebraic geometry.
References
[1]. J. H. Silverman. The arithmetic of elliptic curves, volume 106. Springer, 2009.
[2]. J. S. Milne. Lectures on étale cohomology, 2013.
[3]. M. Musta¸ta. Zeta functions in algebraic geometry, 2011.˘
[4]. W. André. Sur les courbes algébriques et les variétés qui s’ en déduisent. Number 1041. Actualités Sci. Ind, 1948.
[5]. B. Dwork. On the rationality of the zeta function of an algebraic variety. American Journal of Mathematics, 82(3):631–648, 1960.
[6]. P. Deligne. La conjecture de weil. i. Publications Mathématiques de l’Institut des Hautes Études Scientifiques, 43:273–307, 1974.
[7]. M. Artin, A. Grothendieck, and J.-L. Verdier. Théorie des topos et cohomologie étale des schémas, tome 1, 2, 3. Séminaire de géométrie algébrique du Bois-Marie 1963–1964 (SGA 4), pages 1972–1973, 1972.
[8]. S. L. Kleiman. Algebraic cycles and the Weil conjectures. Columbia university, Department of mathematics, 1968.
[9]. A. J. de Jong. Weil cohomology theories, 2007. Seminar notes.
[10]. R. Hartshorne. Algebraic Geometry. Number 52. Springer Science & Business Media, 1977.
[11]. C. Voisin. On the cohomology of algebraic varieties. In Proceedings of the International Congress of Mathematicians 2010 (ICM 2010) (In 4 Volumes) Vol. I: Plenary Lectures and Ceremonies Vols. II–IV: Invited Lectures, pages 476–503. World Scientific, 2010.
[12]. A. Grothendieck. On the de rham cohomology of algebraic varieties. Publications Mathématiques de l’Institut des Hautes Études Scientifiques, 29(1):95–103, 1966.
[13]. R. Hartshorne. On the de rham cohomology of algebraic varieties. Publications Mathématiques de l’IHÉS, 45:5–99, 1975.
[14]. P. Deligne. Cohomologie etale séminaire de géométrie algébrique du bois-marie sga 4 1/2. Séminaire de Géométrie Algébrique du Bois-Marie SGA, 4(1/2):1, 1977.
Cite this article
Yao,Z. (2025). From Zeta Functions to Cohomology: An Insight into the Weil Conjecture. Theoretical and Natural Science,83,193-201.
Data availability
The datasets used and/or analyzed during the current study will be available from the authors upon reasonable request.
Disclaimer/Publisher's Note
The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of EWA Publishing and/or the editor(s). EWA Publishing and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
About volume
Volume title: Proceedings of the 4th International Conference on Computing Innovation and Applied Physics
© 2024 by the author(s). Licensee EWA Publishing, Oxford, UK. This article is an open access article distributed under the terms and
conditions of the Creative Commons Attribution (CC BY) license. Authors who
publish this series agree to the following terms:
1. Authors retain copyright and grant the series right of first publication with the work simultaneously licensed under a Creative Commons
Attribution License that allows others to share the work with an acknowledgment of the work's authorship and initial publication in this
series.
2. Authors are able to enter into separate, additional contractual arrangements for the non-exclusive distribution of the series's published
version of the work (e.g., post it to an institutional repository or publish it in a book), with an acknowledgment of its initial
publication in this series.
3. Authors are permitted and encouraged to post their work online (e.g., in institutional repositories or on their website) prior to and
during the submission process, as it can lead to productive exchanges, as well as earlier and greater citation of published work (See
Open access policy for details).
References
[1]. J. H. Silverman. The arithmetic of elliptic curves, volume 106. Springer, 2009.
[2]. J. S. Milne. Lectures on étale cohomology, 2013.
[3]. M. Musta¸ta. Zeta functions in algebraic geometry, 2011.˘
[4]. W. André. Sur les courbes algébriques et les variétés qui s’ en déduisent. Number 1041. Actualités Sci. Ind, 1948.
[5]. B. Dwork. On the rationality of the zeta function of an algebraic variety. American Journal of Mathematics, 82(3):631–648, 1960.
[6]. P. Deligne. La conjecture de weil. i. Publications Mathématiques de l’Institut des Hautes Études Scientifiques, 43:273–307, 1974.
[7]. M. Artin, A. Grothendieck, and J.-L. Verdier. Théorie des topos et cohomologie étale des schémas, tome 1, 2, 3. Séminaire de géométrie algébrique du Bois-Marie 1963–1964 (SGA 4), pages 1972–1973, 1972.
[8]. S. L. Kleiman. Algebraic cycles and the Weil conjectures. Columbia university, Department of mathematics, 1968.
[9]. A. J. de Jong. Weil cohomology theories, 2007. Seminar notes.
[10]. R. Hartshorne. Algebraic Geometry. Number 52. Springer Science & Business Media, 1977.
[11]. C. Voisin. On the cohomology of algebraic varieties. In Proceedings of the International Congress of Mathematicians 2010 (ICM 2010) (In 4 Volumes) Vol. I: Plenary Lectures and Ceremonies Vols. II–IV: Invited Lectures, pages 476–503. World Scientific, 2010.
[12]. A. Grothendieck. On the de rham cohomology of algebraic varieties. Publications Mathématiques de l’Institut des Hautes Études Scientifiques, 29(1):95–103, 1966.
[13]. R. Hartshorne. On the de rham cohomology of algebraic varieties. Publications Mathématiques de l’IHÉS, 45:5–99, 1975.
[14]. P. Deligne. Cohomologie etale séminaire de géométrie algébrique du bois-marie sga 4 1/2. Séminaire de Géométrie Algébrique du Bois-Marie SGA, 4(1/2):1, 1977.