Namespaces
Variants
Actions

Difference between revisions of "Sheaf theory"

From Encyclopedia of Mathematics
Jump to: navigation, search
(Importing text file)
 
m
 
(3 intermediate revisions by 2 users not shown)
Line 1: Line 1:
 +
{{MSC|14}}
 +
{{TEX|done}}
 +
 
A special mathematical tool which provides a unified approach to establishing connections between local and global properties of topological spaces (in particular geometric objects) and which is a powerful method for studying many problems in contemporary algebra, geometry, topology, and analysis.
 
A special mathematical tool which provides a unified approach to establishing connections between local and global properties of topological spaces (in particular geometric objects) and which is a powerful method for studying many problems in contemporary algebra, geometry, topology, and analysis.
  
A pre-sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s0848401.png" /> on a topological space <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s0848402.png" /> assigns to each open subset <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s0848403.png" /> an Abelian group <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s0848404.png" /> (a ring, a module over a ring, etc.) and to every pair of open sets <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s0848405.png" /> a homomorphism <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s0848406.png" />, such that <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s0848407.png" /> is the identity isomorphism and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s0848408.png" /> for every triple <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s0848409.png" />. In other words, a pre-sheaf is a contravariant functor from the category of open subsets of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484010.png" /> and their inclusions into the category of groups (rings, etc.) and their homomorphisms. The mappings <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484011.png" /> are called restriction homomorphisms (for example, if the elements of the stalk <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484012.png" /> are functions of some type or other defined on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484013.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484014.png" /> is the restriction of these to the smaller subset). A topology on the set <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484015.png" />, where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484016.png" /> is, by definition, the direct limit <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484017.png" />, is defined in the following way: For each <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484018.png" /> and any <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484019.png" />, the set <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484020.png" /> consisting of those points of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484021.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484022.png" />, which are images of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484023.png" /> in the definition of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484024.png" /> is declared to be open in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484025.png" />. In this topology the stalks <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484026.png" /> are discrete, the stalk-wise algebraic operations defined on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484027.png" /> by taking direct limits are continuous and the natural projection <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484028.png" />, where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484029.png" />, is a local homeomorphism. The space <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484030.png" /> together with the stalk-wise algebraic operations and the projection <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484031.png" /> is called the sheaf of Abelian groups (rings, etc.) over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484032.png" /> associated with the pre-sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484033.png" />.
+
A pre-sheaf $F$ on a topological space $X$ assigns to each open subset $U\subseteq X$ an Abelian group $F(U)$ (a ring, a module over a ring, etc.) and to every pair of open sets $V\subseteq U$ a homomorphism $F_V^U:F(U)\to F(V)$, such that $F_U^U$ is the identity isomorphism and $F_W^U=F_W^V F_V^U$ for every triple $W\subseteq V\subseteq U$. In other words, a pre-sheaf is a contravariant functor from the category of open subsets of $X$ and their inclusions into the category of groups (rings, etc.) and their homomorphisms. The mappings $F_V^U$ are called restriction homomorphisms (for example, if the elements of the stalk $F(U)$ are functions of some type or other defined on $U$, $F_V^U$ is the restriction of these to the smaller subset). A topology on the set $\def\cF{ {\mathcal F}}\cF = \bigcup_{x\in X}\cF_x$, where $\cF_x$ is, by definition, the direct limit $\lim_{x\in U} F(U)$, defined in the following way: For each $U\subseteq X$ and any $\def\s{\sigma}\s\in F(U)$, the set $S$ consisting of those points of $\cF_x$, $x\in U$, which are images of $\s$ in the definition of $\cF_x$ is declared to be open in $\cF$. In this topology the stalks $\cF_x$ are discrete, the stalk-wise algebraic operations defined on $\cF$ by taking direct limits are continuous and the natural projection $p:\cF\to X$, where $\cF = p^{-1}(x)$, is a local homeomorphism. The space $\cF$ together with the stalk-wise algebraic operations and the projection $p$ is called the sheaf of Abelian groups (rings, etc.) over $X$ associated with the pre-sheaf $F$.
  
Every continuous mapping <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484034.png" /> for which <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484035.png" /> is called a section of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484036.png" /> over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484037.png" />. The section of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484038.png" /> over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484039.png" /> defined by the zeros in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484040.png" /> is called the zero section. If a section <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484041.png" /> is zero at a point <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484042.png" />, then <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484043.png" /> coincides with the zero section in some neighbourhood of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484044.png" />; therefore the set of points at which <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484045.png" /> is not zero (the support of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484046.png" />) is closed in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484047.png" />.
+
Every continuous mapping $s:U\to \cF$ for which $x=ps(x)$ is called a section of $\cF$ over $U$. The section of $\cF$ over $X$ defined by the zeros in $\cF_x$ is called the zero section. If a section $s$ is zero at a point $x$, then $s$ coincides with the zero section in some neighbourhood of $x$; therefore the set of points at which $s$ is not zero (the support of $s$) is closed in $U$.
  
Let <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484048.png" /> (respectively, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484049.png" />, where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484050.png" /> is a certain family of closed sets in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484051.png" />; in particular, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484052.png" />) be the group (ring, module, etc.) of all sections of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484053.png" /> over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484054.png" /> (respectively, all sections over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484055.png" /> with supports in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484056.png" />; in particular, sections with compact support). The assignment <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484057.png" /> is a pre-sheaf over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484058.png" />, called the pre-sheaf of sections of the sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484059.png" />. The assignment <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484060.png" /> used in defining the topology on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484061.png" /> also defines homomorphisms <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484062.png" /> which commute with the restrictions to <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484063.png" />, that is, it defines a homomorphism of pre-sheaves. This homomorphism is an isomorphism provided that the original pre-sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484064.png" /> satisfies the requirements: a) if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484065.png" /> and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484066.png" />, then <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484067.png" /> if the restrictions of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484068.png" /> and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484069.png" /> to each <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484070.png" /> are equal; and b) if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484071.png" /> and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484072.png" /> is a collection of elements such that the restrictions of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484073.png" /> and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484074.png" /> to <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484075.png" /> coincide, then there exists a <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484076.png" /> which has restriction to each <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484077.png" /> coinciding with <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484078.png" />. The concept of a pre-sheaf satisfying these requirements is equivalent to the concept of the sheaf associated with it, therefore such pre-sheaves are quite often called sheaves also.
+
Let $\def\G{\Gamma}\G(U,\cF)$ (respectively, $\G_\Phi(X,\cF)$, where
 +
$\Phi$ is a certain family of closed sets in $X$; in particular,
 +
$\G_c(X,\cF)$) be the group (ring, module, etc.) of all sections of
 +
$\cF$ over $U$ (respectively, all sections over $X$ with supports in
 +
$\Phi$; in particular, sections with compact support). The assignment
 +
$U\mapsto\G(U,\cF)$ is a pre-sheaf over $X$, called the pre-sheaf of
 +
sections of the sheaf $\cF$. The assignment $\s\mapsto s$ used in
 +
defining the topology on $\cF$ also defines homomorphisms $F(U) \to
 +
\G(U,\cF)$ which commute with the restrictions to $V\subseteq U$, that
 +
is, it defines a homomorphism of pre-sheaves. This homomorphism is an
 +
isomorphism provided that the original pre-sheaf $F$ satisfies the
 +
requirements:  
  
A sheaf of the form <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484079.png" /> (with the evident projection to <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484080.png" />), where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484081.png" /> is a fixed group (ring, etc.), is called a constant sheaf, and is denoted by <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484082.png" />. A sheaf which is constant in sufficiently small neighbourhoods of every <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484083.png" /> is called locally constant. The topology of such sheaves is separated (i.e. Hausdorff) if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484084.png" /> is a separated space. In more typical situations the topology of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484085.png" /> can be non-separated even if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484086.png" /> is separated (such is the case, for example, for the sheaf of germs of continuous (or differentiable) functions which is generated by the pre-sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484087.png" />, where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484088.png" /> is the set of continuous (differentiable) functions on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484089.png" />; however, the sheaf of germs of analytic functions on a manifold is separated).
+
a) if $\def\l{\lambda}U=\bigcup_\l U_\l$ and $\s, \s'\in F(U)$, then $\s=\s'$ if the restrictions of $\s$ and $\s'$ to each
 +
$U_\l$ are equal; and
  
Every homomorphism of pre-sheaves <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484090.png" /> induces a mapping of the associated sheaves <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484091.png" />, which is a local homeomorphism and maps stalks homomorphically to stalks; such a mapping of sheaves is called a sheaf homomorphism. Mono- and epimorphisms are defined in the standard way. For any sheaf homomorphism <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484092.png" /> the image <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484093.png" /> is an open subset of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484094.png" />, closed with respect to the stalk-wise algebraic operations. Every subset of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484095.png" /> which satisfies these requirements is called a subsheaf of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484096.png" />. The quotient sheaf of the sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484097.png" /> by a subsheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484098.png" /> is defined as the sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s08484099.png" /> associated with the pre-sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840100.png" />; moreover, there is an epimorphism <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840101.png" />, and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840102.png" />. For every open <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840103.png" /> there is a subsheaf in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840104.png" />, denoted by <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840105.png" />, which is the union of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840106.png" /> with the zero section of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840107.png" /> over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840108.png" />; <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840109.png" /> denotes the corresponding quotient sheaf (whose restriction to <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840110.png" /> coincides with the restriction thereto of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840111.png" />).
+
b) if $U=\bigcup_\l U_\l$ and $\s_\l\in F(U_\l)$ is a collection of elements such that the restrictions of $\s_\l$ and $s_\mu$ to $U_\l\cap U_\mu$ coincide, then there exists a $\s\in F(U)$ which has restriction to each $U_\l$ coinciding with $\s_\l$. The concept of a pre-sheaf satisfying these requirements is equivalent to the concept of the sheaf associated with it, therefore such pre-sheaves are quite often called sheaves also.
  
Since it is possible to interpret such ordinary terms as homomorphism, kernel, image, subsheaf, quotient sheaf, etc. for sheaves over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840112.png" /> in such a way that these concepts have essentially the same meaning as in algebra, one can consider them from a categorical point of view and apply to sheaf theory the constructions of [[Homological algebra|homological algebra]]. The resulting category of sheaves over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840113.png" /> has the same classical properties as the category of Abelian groups or the category of modules; in particular, one can define for sheaves direct sums, infinite direct products, inductive limits, and other concepts.
+
A sheaf of the form $X\times G$ (with the evident projection to $X$), where $G$ is a fixed group (ring, etc.), is called a constant sheaf, and is denoted by $G$. A sheaf which is constant in sufficiently small neighbourhoods of every $x\in X$ is called locally constant. The topology of such sheaves is separated (i.e. Hausdorff) if $X$ is a separated space. In more typical situations the topology of $\cF$ can be non-separated even if $X$ is separated (such is the case, for example, for the sheaf of germs of continuous (or differentiable) functions which is generated by the pre-sheaf $F$, where $F(U)$ is the set of continuous (differentiable) functions on $U$; however, the sheaf of germs of analytic functions on a manifold is separated).
  
The apparatus of sheaf theory has penetrated into various fields of mathematics thanks to the fact that there is a natural definition of the cohomology <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840114.png" /> of a space <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840115.png" /> with coefficients in a sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840116.png" />, and this without any kind of restrictions on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840117.png" /> (this is essential, for example, in algebraic geometry, where the spaces arising are, as a rule, non-separated) and to the fact that other cohomologies (under certain specific conditions) reduce to a sheaf cohomology, at least in those situations where their application is justified.
+
Every homomorphism of pre-sheaves $F\to F'$ induces a mapping of the associated sheaves $\cF\to\cF'$, which is a local homeomorphism and maps stalks homomorphically to stalks; such a mapping of sheaves is called a sheaf homomorphism. Mono- and epimorphisms are defined in the standard way. For any sheaf homomorphism $f:\cF'\to\cF$ the image $f(\cF')$ is an open subset of $\cF$, closed with respect to the stalk-wise algebraic operations. Every subset of $\cF$ which satisfies these requirements is called a subsheaf of $\cF$. The quotient sheaf of the sheaf $\cF$ by a subsheaf $\cF'$ is defined as the sheaf $\cF''$ associated with the pre-sheaf $U\mapsto \G(U,\cF)/\G(U,\cF')$; moreover, there is an epimorphism $\cF\to\cF''$, and $\cF_X''\simeq \cF_X/\cF_X'$. For every open $U\subseteq X$ there is a subsheaf in $\cF$, denoted by $\cF_U$, which is the union of $p^{-1}(U)$ with the zero section of $\cF$ over $X$; $\cF_{X\setminus U}$ denotes the corresponding quotient sheaf (whose restriction to $X\setminus U$ coincides with the restriction thereto of $\cF$).
  
To define <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840118.png" /> one first constructs the canonical resolution
+
Since it is possible to interpret such ordinary terms as homomorphism, kernel, image, subsheaf, quotient sheaf, etc. for sheaves over $X$ in such a way that these concepts have essentially the same meaning as in algebra, one can consider them from a categorical point of view and apply to sheaf theory the constructions of
 +
[[Homological algebra|homological algebra]]. The resulting category of sheaves over $X$ has the same classical properties as the category of Abelian groups or the category of modules; in particular, one can define for sheaves direct sums, infinite direct products, inductive limits, and other concepts.
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840119.png" /></td> </tr></table>
+
The apparatus of sheaf theory has penetrated into various fields of mathematics thanks to the fact that there is a natural definition of the cohomology $\def\H{ {\rm H}}\H^*(X,\cF)$ of a space $X$ with coefficients in a sheaf $\cF$, and this without any kind of restrictions on $X$ (this is essential, for example, in algebraic geometry, where the spaces arising are, as a rule, non-separated) and to the fact that other cohomologies (under certain specific conditions) reduce to a sheaf cohomology, at least in those situations where their application is justified.
  
where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840120.png" /> is the sheaf defined by the pre-sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840121.png" /> for which <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840122.png" /> is the group of all (possibly discontinuous) sections of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840123.png" /> over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840124.png" />, so that <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840125.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840126.png" /> <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840127.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840128.png" />. By definition, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840129.png" /> (<img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840130.png" /> is obtained by replacing the symbol <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840131.png" /> by <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840132.png" />). The sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840133.png" /> itself can be obtained from <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840134.png" /> so that <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840135.png" /> (in classical cohomology <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840136.png" /> is the group of locally constant functions on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840137.png" /> with values in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840138.png" />). The resolution <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840139.png" /> is an exact covariant functor of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840140.png" />: there is a short exact sequence of resolutions corresponding to a short exact  "coefficient"  sequence <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840141.png" />. The functor <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840142.png" /> turns out to be exact on the terms <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840143.png" />, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840144.png" />, of the resolution, therefore there is an exact cohomology sequence
+
To define $\H^*(X,\cF)$ one first constructs the canonical resolution
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840145.png" /></td> </tr></table>
+
$$\def\C{ {\rm C}}\C^*(\cF) : 0\to \cF\to\C^0(\cF)\to \C^1(\cF)\to \cdots,$$
 +
where $\C^0(\cF)$ is the sheaf defined by the pre-sheaf $F$ for which
 +
$F(U)$ is the group of all (possibly discontinuous) sections of $\cF$
 +
over $U$, so that $\G(U,\C^0(\cF))=F(U),\; \C^1(\cF) =
 +
\C^0(\C^0(\cF)/\cF)),\dots,\C^{p+1}(\cF)=\C^0(\C^p(\cF)/{\rm
 +
Im}\;\C^{p-1}(\cF)),\dots$. By definition, $\H^p(X,\cF) = \H^p(\G(X,\C^*(\cF)))$ ($\H_\Phi^p(X,\cF))$ is obtained by replacing the symbol $\G$ by $\G_\Phi$). The sheaf $\cF$ itself can be obtained from $C^*(\cF)$ so that $H_\Phi^0(X,\cF)=\G_\Phi(X,\cF)$ (in classical cohomology $\H^0(X,G)$ is the group of locally constant functions on $X$ with values in $G$). The resolution $C^*(\cF) $ is an exact covariant functor of $\cF$: there is a short exact sequence of resolutions corresponding to a short exact "coefficient" sequence $0\to\cF'\to\cF\to\cF''\to 0$. The functor $\G_\Phi$ turns out to be exact on the terms $C^p$, $p\ge0$, of the resolution, therefore there is an exact cohomology sequence
  
<table class="eq" style="width:100%;"> <tr><td valign="top" style="width:94%;text-align:center;"><img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840146.png" /></td> </tr></table>
+
$$\cdots \to \H_\Phi^{p-1}(X,\cF'')\to \H_\Phi^{p}(X,\cF')
 +
\to\H_\Phi^p(X,\cF)\to\H_\Phi^{p}(X,\cF'')\to \cdots,$$
  
corresponding to the indicated coefficient sequence, beginning with <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840147.png" />. The cohomology sequence of a pair <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840148.png" /> corresponds to the short exact sequence, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840149.png" /> (<img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840150.png" /> is a closed set).
+
corresponding to the indicated coefficient sequence, beginning with $0\to\G_\Phi(X,\cF')\to\G_\Phi(X,\cF)\to\cdots$. The cohomology sequence of a pair $(X,A)$ corresponds to the short exact sequence, $0\to\cF_{X\setminus A}\to\cF\to\cF_A\to 0$ ($A$ is a closed set).
  
The cohomology groups <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840151.png" /> have the following "universality" property, which casts light on their meaning: For any other resolution <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840152.png" /> (that is, an exact sequence of sheaves <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840153.png" /> beginning with <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840154.png" />) there is a natural "comparison" homomorphism <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840155.png" />, which is described in terms of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840156.png" /> by using spectral sequences. An important case is when the sheaves of the resolution are <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840158.png" />-acyclic, that is, when <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840159.png" /> for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840160.png" />; in this case the above homomorphism is an isomorphism. The basic examples of acyclic sheaves are flabby sheaves (for all <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840161.png" /> the mappings <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840162.png" /> are epimorphic) and soft sheaves (any section over a closed set extends to a section over the whole of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840163.png" />). The canonical resolution consists of flabby sheaves. If <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840164.png" /> is a paracompact space, then every flabby sheaf is also soft.
+
The cohomology groups $\H_\Phi^*(X,\cF)$ have the following "universality" property, which casts light on their meaning: For any other resolution $\def\cL{ {\mathcal L}}\cL^*$ (that is, an exact sequence of sheaves $\cL^q$ beginning with $\cF$) there is a natural "comparison" homomorphism $\H^p(\G_\Phi(X,\cL^*))\to H_\Phi^p(X,\cF)$, which is described in terms of $H_\Phi^p(X,\cL^q)$ by using [[spectral sequence]]s. An important case is when the sheaves of the resolution are $\Phi$-acyclic, that is, when $H_\Phi^p(X,\cL^q)=0$ for $p\ge 1$; in this case the above homomorphism is an isomorphism. The basic examples of acyclic sheaves are [[flabby sheaf|flabby sheaves]] (for all $U\subseteq X$ the mappings $\G(X,\cL)\to \G(U,\cL)$ are epimorphic) and soft sheaves (any section over a closed set extends to a section over the whole of $X$). The canonical resolution consists of flabby sheaves. If $X$ is a paracompact space, then every flabby sheaf is also soft.
  
The universality property enables one to compare cohomologies arising in concrete situations with sheaf cohomology (and consequently also with each other), to discern for them the natural bounds within which their application is effective, and also to apply sheaf-theoretic methods to the solution of concrete problems. For example, Aleksandrov–Čech cohomology (cf. [[Aleksandrov–Čech homology and cohomology|Aleksandrov–Čech homology and cohomology]]) can be defined using cochains obtained from the cochains of a specially selected system of open coverings by taking the direct limit. These cochains turn out to be sections of the sheaves of germs of cochains (defined analogously to the sheaves of germs of functions) constituting a resolution of the group (or even the sheaf) of coefficients, and this sheaf turns out to be soft if the space is paracompact. Thus, for paracompact spaces Aleksandrov–Čech cohomology coincides with a sheaf cohomology. An analogous conclusion holds for Zariski spaces (in particular, for algebraic varieties). Alexander–Spanier cochains also turn out to be sections of the sheaves of a resolution and, moreover, the resolution consists of soft sheaves if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840165.png" /> is paracompact, so in this case, in particular, Alexander–Spanier and Aleksandrov–Čech cohomology are naturally isomorphic. In the case of singular cohomology, identification of cochains which coincide on the "small" singular simplices, i.e. subordinated to (arbitrary) open coverings, leads to the so-called localized cochains (giving the same cohomology), which are sections of the sheaves determined by the pre-sheaves of the usual singular cochains. These sheaves turn out to be soft if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840166.png" /> is paracompact (if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840167.png" /> is hereditarily paracompact, then they are also flabby), but they form a resolution only under the additional requirement that <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840168.png" /> is weakly locally contractible (in every neighbourhood <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840169.png" /> of each point <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840170.png" /> there is a smaller neighbourhood which is contractible to a point inside <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840171.png" />). A classic example is de Rham's theorem: The cohomology of the complex of differential forms of a differentiable manifold coincides with the usual cohomology with coefficients in the field <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840172.png" /> of real numbers (the sheaves of germs of differential forms are soft and form a resolution of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840173.png" />: Sufficiently near to each point each closed differential form is exact).
+
The universality property enables one to compare cohomologies arising in concrete situations with sheaf cohomology (and consequently also with each other), to discern for them the natural bounds within which their application is effective, and also to apply sheaf-theoretic methods to the solution of concrete problems. For example, Aleksandrov–Čech cohomology (cf.
 +
[[Aleksandrov–Čech homology and cohomology|Aleksandrov–Čech homology and cohomology]]) can be defined using cochains obtained from the cochains of a specially selected system of open coverings by taking the direct limit. These cochains turn out to be sections of the sheaves of germs of cochains (defined analogously to the sheaves of germs of functions) constituting a resolution of the group (or even the sheaf) of coefficients, and this sheaf turns out to be soft if the space is paracompact. Thus, for paracompact spaces Aleksandrov–Čech cohomology coincides with a sheaf cohomology. An analogous conclusion holds for Zariski spaces (in particular, for algebraic varieties). Alexander–Spanier cochains also turn out to be sections of the sheaves of a resolution and, moreover, the resolution consists of soft sheaves if $X$ is paracompact, so in this case, in particular, Alexander–Spanier and Aleksandrov–Čech cohomology are naturally isomorphic. In the case of singular cohomology, identification of cochains which coincide on the "small" singular simplices, i.e. subordinated to (arbitrary) open coverings, leads to the so-called localized cochains (giving the same cohomology), which are sections of the sheaves determined by the pre-sheaves of the usual singular cochains. These sheaves turn out to be soft if $X$ is paracompact (if $X$ is hereditarily paracompact, then they are also flabby), but they form a resolution only under the additional requirement that $X$ is weakly locally contractible (in every neighbourhood $U$ of each point $x\in X$ there is a smaller neighbourhood which is contractible to a point inside $U$). A classic example is [[De Rham theorem|de Rham's theorem]]: The cohomology of the complex of differential forms of a differentiable manifold coincides with the usual cohomology with coefficients in the field $\R$ of real numbers (the sheaves of germs of differential forms are soft and form a resolution of $\R$: Sufficiently near to each point each closed differential form is exact).
  
There are also resolutions corresponding to any open or locally finite closed covering and these enable one to compare the cohomology of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840174.png" /> with the cohomology of the coverings (the spectral sequences for coverings). In particular, this gives an isomorphism if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840175.png" /> for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840176.png" /> for all elements of the covering and their finite intersections (Leray's theorem). Taking the direct limit with respect to open coverings gives an isomorphism between Aleksandrov–Čech cohomology <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840177.png" /> and sheaf cohomology, even for non-paracompact <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840178.png" />, provided that there are sufficiently many small open sets <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840179.png" /> in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840180.png" /> for which <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840181.png" /> when <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840182.png" /> (Cartan's theorem). This means that the cohomologies <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840183.png" />, used in algebraic geometry, with coefficients in coherent sheaves, are also isomorphic to the standard sheaf cohomology <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840184.png" />.
+
There are also resolutions corresponding to any open or locally finite
 +
closed covering and these enable one to compare the cohomology of $X$
 +
with the cohomology of the coverings (the spectral sequences for
 +
coverings). In particular, this gives an isomorphism if $\H^q =0$ for
 +
$q\ge 1$ for all elements of the covering and their finite
 +
intersections (Leray's theorem). Taking the direct limit with respect
 +
to open coverings gives an isomorphism between Aleksandrov–Čech
 +
cohomology $\def\cH{ {\mathcal H}}\check{\cH}^*$ and sheaf cohomology, even for non-paracompact $X$, provided that there are sufficiently many small open sets $U$ in $X$ for which $\check{\cH}^q(U,\cF)=0$ when $q\ge 1$ (Cartan's theorem). This means that the cohomologies $\check{\cH}^*$, used in algebraic geometry, with coefficients in [[Coherent sheaf|coherent sheaves]], are also isomorphic to the standard sheaf cohomology $\H^*$.
  
General constructions ensuring the existence of a comparison homomorphism enable one to compare also the cohomology <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840185.png" /> with the hypercohomology <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840186.png" /> (analogously, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840187.png" /> with <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840188.png" />) in case <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840189.png" /> is any differential sheaf (that is, a sheaf in which for any <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840190.png" /> the composition <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840191.png" /> is zero) with <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840192.png" /> acyclic, where the <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840193.png" /> are the derived sheaves of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840194.png" /> (these are the quotient sheaves of the kernel by the image in each dimension <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840195.png" />). The corresponding spectral sequences have many applications. Moreover, if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840196.png" /> when <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840197.png" />, then <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840198.png" />. For example, if in place of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840199.png" /> a sheaf of chains <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840200.png" /> is taken (the boundary operator lowers the dimension by one, the elements of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840201.png" /> are the chains of the pair <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840202.png" />, and the stalk <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840203.png" />), then one obtains the way the homology <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840204.png" /> depends on all the possible <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840205.png" />. For a manifold, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840206.png" /> when <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840207.png" />, and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840208.png" />, that is, [[Poincaré duality|Poincaré duality]] holds. If <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840209.png" /> is an open or closed subset of a locally compact space <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840210.png" />, then the homology of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840211.png" /> is determined by the sections of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840212.png" /> with supports in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840213.png" />, and the homology of the pair <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840214.png" /> is determined by the sections of the restriction of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840215.png" /> to <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840216.png" />. Conversely (and this is also one of the manifestations of Poincaré duality), if <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840217.png" /> is any flabby resolution for the cohomology, then the restriction of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840218.png" /> to <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840219.png" /> determines the cohomology of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840220.png" /> and the sections of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840221.png" /> with supports in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840222.png" /> determine the cohomology of the pair <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840223.png" />. Since the sheaves <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840224.png" /> are flabby for manifolds, the homology sequence of the pair <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840225.png" /> coincides up to an inversion of the numbering with the cohomology sequence of the pair <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840226.png" />. This means that dualities for manifolds, such as [[Lefschetz duality|Lefschetz duality]] <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840227.png" />, are particular cases of Poincaré duality. It turns out that the duality relations which are not covered by this scheme are corollaries of Poincaré duality and the acyclicity of the manifold in certain dimensions.
+
General constructions ensuring the existence of a comparison
 +
homomorphism enable one to compare also the cohomology $\def\cH{
 +
{\mathcal H}}\H^p(X,\cH^q)$ with the hypercohomology
 +
$\H^*(\G(X,\cL^*))$ (analogously, $\H_\Phi^p(X,\cH^q)$ with
 +
$\H^*(\G_\Phi(X,\cL^*))$) in case $\cL^*$ is any differential sheaf
 +
(that is, a sheaf in which for any $q$ the composition
 +
$\cL^q\to\cL^{q+2}$ is zero) with $\cL^q$ acyclic, where the $\cH^q$
 +
are the derived sheaves of $\cL^*$ (these are the quotient sheaves of
 +
the kernel by the image in each dimension $q$). The corresponding
 +
spectral sequences have many applications. Moreover, if $\cH^q=0$ when
 +
$q\ge 1$, then $\H^*(\G(X,\cL^*))=\H^*(X,\cH^0)$. For example, if in
 +
place of $\cL^*$ a sheaf of chains $\def\cC{ {\mathcal C}}\cC_*$ is taken (the boundary operator lowers the dimension by one, the elements of $\G(U,\cC_*)$ are the chains of the pair $(X,X\setminus U)$, and the stalk $\cH_q^x = \lim_{x\in U}\H_q(X,X\setminus U)=H_q(X,X\setminus x)$), then one obtains the way the homology $\H_*^\Phi(X,G)$ depends on all the possible $\H_\Phi^p(X,\cH_q)$. For a manifold, $\cH_q=0$ when $q>n=\dim X$, and $\H_p^\Phi(X,G)=\H_\Phi^{n-p}(X,\cH_n)$, that is,
 +
[[Poincaré duality|Poincaré duality]] holds. If $A$ is an open or closed subset of a locally compact space $X$, then the homology of $A$ is determined by the sections of $\cC_*$ with supports in $A$, and the homology of the pair $(X,A)$ is determined by the sections of the restriction of $\cC_*$ to $X\setminus A$. Conversely (and this is also one of the manifestations of Poincaré duality), if $X\setminus A$ is any flabby resolution for the cohomology, then the restriction of $\cC^*$ to $X\setminus A$ determines the cohomology of $X\setminus A$ and the sections of $\cC^*$ with supports in $A$ determine the cohomology of the pair $(X,X\setminus A)$. Since the sheaves $\cC_*$ are flabby for manifolds, the homology sequence of the pair $(X,A)$ coincides up to an inversion of the numbering with the cohomology sequence of the pair $(X,X\setminus A)$. This means that dualities for manifolds, such as
 +
[[Lefschetz duality|Lefschetz duality]] $\H_p(X,U,G) = \H^{n-p}(X\setminus U,\cH_n)$, are particular cases of Poincaré duality. It turns out that the duality relations which are not covered by this scheme are corollaries of Poincaré duality and the acyclicity of the manifold in certain dimensions.
  
Just such a situation arises in the case of a continuous mapping <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840228.png" />. A resolution for the cohomology of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840229.png" /> determines a certain differential sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840230.png" /> on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840231.png" /> for which the stalks <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840232.png" /> are the direct limits of the cohomology groups <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840233.png" /> with respect to neighbourhoods <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840234.png" /> of the points <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840235.png" /> (and for closed mappings <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840236.png" />), where <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840237.png" />. The way <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840238.png" /> depends on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840239.png" /> is described by the Leray spectral sequence of the mapping <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840240.png" /> (a particular case of this is the spectral sequence of a [[Serre fibration|Serre fibration]]). Acyclic mappings correspond to the case when <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840241.png" /> vanishes, thus ensuring that the cohomologies of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840242.png" /> and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840243.png" /> with corresponding coefficients are isomorphic (Vietoris' theorem and its generalizations). The general constructions referred to above also give the spectral sequence of a mapping and take into account (along with their cohomological structure) the degree of disconnectedness of pre-images of points; this is especially effective for zero-dimensional or finite-to-one mappings (in the case of coverings it becomes the Cartan spectral sequence). There are also special spectral sequences in categories of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840244.png" />-spaces (spaces on which a group <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840245.png" /> acts).
+
Just such a situation arises in the case of a continuous mapping $f:X\to Y$. A resolution for the cohomology of $X$ determines a certain differential sheaf $\cH^*$ on $Y$ for which the stalks $\cH_y^q$ are the direct limits of the cohomology groups $\H^q(f^{-1}(U),\cF)$ with respect to neighbourhoods $U$ of the points $y$ (and for closed mappings $\H_y^q = \H^q(f^{-1}(y),\cF)$), where $\H^*(X,\cF) = \H^*(\G(Y,\cH^*))$. The way $\H^*(X,\cF)$ depends on $\H^p(Y,\cH^q)$ is described by the Leray spectral sequence of the mapping $f$ (a particular case of this is the spectral sequence of a
 +
[[Serre fibration|Serre fibration]]). Acyclic mappings correspond to the case when $\cH^q$ vanishes, thus ensuring that the cohomologies of $X$ and $Y$ with corresponding coefficients are isomorphic (Vietoris' theorem and its generalizations). The general constructions referred to above also give the spectral sequence of a mapping and take into account (along with their cohomological structure) the degree of disconnectedness of pre-images of points; this is especially effective for zero-dimensional or finite-to-one mappings (in the case of coverings it becomes the Cartan spectral sequence). There are also special spectral sequences in categories of $G$-spaces (spaces on which a group $G$ acts).
  
 
In sheaf cohomology there is a natural way of defining a multiplicative structure. The existence of special flabby resolutions, of which the mappings are determined by a certain semi-simplicial structure, enables one to give explicit formulas for the products of cochains, analogous to the usual ones. At the same time this also makes it possible to define other cohomology operations in sheaf theory.
 
In sheaf cohomology there is a natural way of defining a multiplicative structure. The existence of special flabby resolutions, of which the mappings are determined by a certain semi-simplicial structure, enables one to give explicit formulas for the products of cochains, analogous to the usual ones. At the same time this also makes it possible to define other cohomology operations in sheaf theory.
Line 41: Line 85:
 
The apparatus of sheaf theory has many applications wherever abstract homological methods are essential: in topology (homological and cohomological dimension, local homology and duality, the structure of various classes of continuous mappings, including imbeddings onto dense subsets, and, in particular, compactifications, etc.), in the theory of analytic manifolds (homology and cohomology with coefficients in coherent analytic sheaves and their applications, cohomology and analytic differential forms, homology and analytic flows (the analogue of de Rham's theorem), etc.), and also in abstract algebraic geometry (the cohomology of affine, projective and complete algebraic varieties with coefficients in coherent algebraic sheaves, algebraic Serre duality, algebraic (combinatorial) dimension, etc.).
 
The apparatus of sheaf theory has many applications wherever abstract homological methods are essential: in topology (homological and cohomological dimension, local homology and duality, the structure of various classes of continuous mappings, including imbeddings onto dense subsets, and, in particular, compactifications, etc.), in the theory of analytic manifolds (homology and cohomology with coefficients in coherent analytic sheaves and their applications, cohomology and analytic differential forms, homology and analytic flows (the analogue of de Rham's theorem), etc.), and also in abstract algebraic geometry (the cohomology of affine, projective and complete algebraic varieties with coefficients in coherent algebraic sheaves, algebraic Serre duality, algebraic (combinatorial) dimension, etc.).
  
Some basic notions of sheaf theory and spectral sequences appeared in the work of J. Leray (1945 and later) in connection with the study of homological properties of continuous mappings of locally compact spaces, and he also gave the definition of cohomology (with compact support) with coefficients in a sheaf. A fairly complete account of sheaf theory using resolutions was later given by H. Cartan. The proof of the de Rham theorem given by A. Weil (1947) and the work of J.-P. Serre (in the early 1950's) on algebraic varieties greatly influenced the development of sheaf theory. Cohomology with coefficients in a sheaf was first defined by the Aleksandrov–Čech method. A mature view of sheaf theory could be found by the end of the 1950's in the work of A. Grothendieck [[#References|[3]]] and R. Godement [[#References|[2]]], where great generality was achieved and the methods were considerably simplified. E.g., it was shown that the category of sheaves over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840246.png" /> has a generator (that is, a sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840247.png" /> admitting non-zero homomorphisms into any non-zero sheaf; for sheaves of Abelian groups, <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840248.png" />), and hence that each sheaf can be imbedded in an injective sheaf (Grothendieck's theorem). This is the reason for the formal analogy between cohomology theory with coefficients in sheaves and the theory of derived functors in the module categories: In the category of sheaves over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840249.png" /> there are "enough" injective objects (although, as a rule, there are few projective objects), and therefore one can freely apply all the corresponding techniques of homological algebra; in particular, one can define the cohomology <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840250.png" /> (without any restriction on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840251.png" />) as derived functors of the left exact functor <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840252.png" /> (or even as <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840253.png" />). This also sheds light, for example, on the general nature of such concepts as the cohomological dimension (over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840254.png" />) of a space, the algebraic dimension of a variety and the global dimension of a ring. The description given by Grothendieck of the spectral sequence for the functor <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840255.png" /> is essential in algebraic geometry. A much simpler method of constructing injective sheaves was found by Godement. He also showed that to construct a cohomology theory it is entirely sufficient to use his canonical flabby resolution, which, from the point of view of homological algebra, turns out to be simply one of the acyclic resolutions of a sheaf. Godement was the first to apply flabby and soft sheaves (soft sheaves are acyclic only for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840256.png" /> paracompact, which explains their use primarily in topology).
+
Some basic notions of sheaf theory and spectral sequences appeared in the work of J. Leray (1945 and later) in connection with the study of homological properties of continuous mappings of locally compact spaces, and he also gave the definition of cohomology (with compact support) with coefficients in a sheaf. A fairly complete account of sheaf theory using resolutions was later given by H. Cartan. The proof of the de Rham theorem given by A. Weil (1947) and the work of J.-P. Serre (in the early 1950's) on algebraic varieties greatly influenced the development of sheaf theory. Cohomology with coefficients in a sheaf was first defined by the Aleksandrov–Čech method. A mature view of sheaf theory could be found by the end of the 1950's in the work of A. Grothendieck
 
+
{{Cite|Gr}} and R. Godement
====References====
+
{{Cite|Go}}, where great generality was achieved and the methods were considerably simplified. E.g., it was shown that the category of sheaves over $X$ has a generator (that is, a sheaf $J$ admitting non-zero homomorphisms into any non-zero sheaf; for sheaves of Abelian groups, $J=\sum_{U\subseteq X}\Z_U$), and hence that each sheaf can be imbedded in an injective sheaf (Grothendieck's theorem). This is the reason for the formal analogy between cohomology theory with coefficients in sheaves and the theory of derived functors in the module categories: In the category of sheaves over $X$ there are "enough" injective objects (although, as a rule, there are few projective objects), and therefore one can freely apply all the corresponding techniques of homological algebra; in particular, one can define the cohomology $\H_\Phi^*(X,\cF)$ (without any restriction on $X$) as derived functors of the left exact functor $\G_\Phi(X,\cF)$ (or even as $\def\Ext{ {\rm Ext}}\Ext^*(\Z_X,\cF)$). This also sheds light, for example, on the general nature of such concepts as the cohomological dimension (over $\Z$) of a space, the algebraic dimension of a variety and the global dimension of a ring. The description given by Grothendieck of the spectral sequence for the functor $\Ext$ is essential in algebraic geometry. A much simpler method of constructing injective sheaves was found by Godement. He also showed that to construct a cohomology theory it is entirely sufficient to use his canonical flabby resolution, which, from the point of view of homological algebra, turns out to be simply one of the acyclic resolutions of a sheaf. Godement was the first to apply flabby and soft sheaves (soft sheaves are acyclic only for $X$ paracompact, which explains their use primarily in topology).
<table><TR><TD valign="top">[1]</TD> <TD valign="top">  G.E. Bredon,  "Sheaf theory" , McGraw-Hill  (1967)</TD></TR><TR><TD valign="top">[2]</TD> <TD valign="top">  R. Godement,  "Topologie algébrique et théorie des faisceaux" , Hermann  (1958)</TD></TR><TR><TD valign="top">[3]</TD> <TD valign="top">  A. Grothendieck,  "Sur quelques points d'algèbre homologique"  ''Tohoku Math. J.'' , '''9'''  (1957)  pp. 119–221</TD></TR><TR><TD valign="top">[4]</TD> <TD valign="top">  R. Swan,  "The theory of sheaves" , Chicago Univ. Press  (1964)</TD></TR><TR><TD valign="top">[5]</TD> <TD valign="top">  E.G. Sklyarenko,  "Homology and cohomology of general spaces" , Springer  (Forthcoming)  (Translated from Russian)</TD></TR></table>
 
  
 +
====Comments====
 +
Let $f:X\to Y$ be a continuous mapping between topological spaces and $\cF$ a sheaf on $X$. Then the direct image of $\cF$ under $f$ is the sheaf $f_*\cF$ defined by $(f_*\cF)(U) = \cF(f^{-1}(U))$ for $U$ open in $Y$. The functor $\cF\mapsto f_*(\cF)$ is left-exact on sheaves of modules or Abelian groups. The higher direct images $f_*^q\cF$ of $\cF$ are the right derived functors of $f_*$. The sheaf $f_*^q\cF$ is the sheaf associated to the pre-sheaf $U\mapsto \H^q(f^{-1}(U),\cF)$.
  
 +
Given a sheaf $\def\cG{ {\mathcal G}}\cG$ over $Y$, let $\pi:\cG\to Y$ be the associated "sheaf space" (so that $\cG(U)=$ sections of $\pi$ over $U\subseteq Y$). Now consider the fibre product (pull-back) $f^*\cG=X\times_Y\cG$. The inverse image sheaf $f^*\cG$ is the sheaf of sections of $f^*\cG\to X$.
  
====Comments====
+
The functor $f^*$ is exact (on sheafs of modules or Abelian groups). The two functors are adjoint to one another: ${\rm Hom}(f^*\cG,\cF) \simeq {\rm Hom}(\cG,f_*\cF)$.
Let <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840257.png" /> be a continuous mapping between topological spaces and <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840258.png" /> a sheaf on <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840259.png" />. Then the direct image of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840260.png" /> under <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840261.png" /> is the sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840262.png" /> defined by <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840263.png" /> for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840264.png" /> open in <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840265.png" />. The functor <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840266.png" /> is left-exact on sheaves of modules or Abelian groups. The higher direct images <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840267.png" /> of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840268.png" /> are the right derived functors of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840269.png" />. The sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840270.png" /> is the sheaf associated to the pre-sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840271.png" />.
 
 
 
Given a sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840272.png" /> over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840273.png" />, let <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840274.png" /> be the associated  "sheaf space"  (so that <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840275.png" /> sections of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840276.png" /> over <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840277.png" />). Now consider the fibre product (pull-back) <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840278.png" />. The inverse image sheaf <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840279.png" /> is the sheaf of sections of <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840280.png" />.
 
  
The functor <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840281.png" /> is exact (on sheafs of modules or Abelian groups). The two functors are adjoint to one another: <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840282.png" />.
+
A short sequence of sheaves of Abelian groups or modules $0\to\cF'\to\cF\to\cF''\to 0$ over a space $X$ is exact if and only if the corresponding sequence of stalks $0\to\cF_x'\to\cF_x\to\cF_x''\to 0$ is exact for every $x\in X$.
  
A short sequence of sheaves of Abelian groups or modules <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840283.png" /> over a space <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840284.png" /> is exact if and only if the corresponding sequence of stalks <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840285.png" /> is exact for every <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840286.png" />.
+
Grothendieck and his school vastly generalized sheaf theory, from sheaves on a space to the more general notion of sheaves on a
 +
[[Site|site]] and that of a
 +
[[Topos|topos]] (cf.
 +
{{Cite|Ar}}–{{Cite|Gr3}}).
  
Grothendieck and his school vastly generalized sheaf theory, from sheaves on a space to the more general notion of sheaves on a [[Site|site]] and that of a [[Topos|topos]] (cf. [[#References|[a6]]][[#References|[a8]]]).
+
For the theory of coherent sheaves on algebraic varieties see also the fundamental paper
 +
{{Cite|Gr2}}. For the theory of sheaves in the étale topology and for $l$-adic sheaves see
 +
{{Cite|Gr3}},
 +
{{Cite|De}}{{Cite|FrKi}}.
  
For the theory of coherent sheaves on algebraic varieties see also the fundamental paper [[#References|[a1]]]. For the theory of sheaves in the étale topology and for <img align="absmiddle" border="0" src="https://www.encyclopediaofmath.org/legacyimages/s/s084/s084840/s084840287.png" />-adic sheaves see [[#References|[a8]]], [[#References|[a12]]]–[[#References|[a14]]].
+
For a detailed history of the origins of sheaf theory, see
 +
{{Cite|Gr2}}. The article above focuses almost entirely on the use of sheaf theory in cohomology theory, but it does have many other applications: for example, in the representation theory of rings and other algebraic systems (see, for example,
 +
{{Cite|Mu}},
 +
{{Cite|BoBo}}), and in logic, in the provision of models for constructive mathematics (see
 +
{{Cite|FoSc}}, and also
 +
{{Cite|MaRe}},
 +
{{Cite|Ti}}).
  
For a detailed history of the origins of sheaf theory, see [[#References|[a1]]]. The article above focuses almost entirely on the use of sheaf theory in cohomology theory, but it does have many other applications: for example, in the representation theory of rings and other algebraic systems (see, for example, [[#References|[a3]]], [[#References|[a4]]]), and in logic, in the provision of models for constructive mathematics (see [[#References|[a2]]], and also [[#References|[a9]]], [[#References|[a10]]]).
 
  
 
====References====
 
====References====
<table><TR><TD valign="top">[a1]</TD> <TD valign="top">  J.W. Gray,   "Fragments of the history of sheaf theory" M.P. Fourman (ed.)  C.J. Mulvey (ed.) D.S. Scott (ed.) , ''Applications of sheaves'' , ''Lect. notes in math.'' , '''753''' , Springer (1979) pp. 1–79</TD></TR><TR><TD valign="top">[a2]</TD> <TD valign="top"M.P. Fourman,   D.S. Scott,   "Sheaves and logic" M.P. Fourman (ed.) C.J. Mulvey (ed.) D.S. Scott (ed.) , ''Applications of sheaves'' , ''Lect. notes in math.'' , '''753''' , Springer (1979) pp. 302–401</TD></TR><TR><TD valign="top">[a3]</TD> <TD valign="top">  C.J. Mulvey,   "Representations of rings and modules" M.P. Fourman (ed.) C.J. Mulvey (ed.) D.S. Scott (ed.) , ''Applications of sheaves'' , ''Lect. notes in math.'' , '''753''' , Springer (1979) pp. 542–585</TD></TR><TR><TD valign="top">[a4]</TD> <TD valign="top">  F. Borceux,   G. van den Bossche,   "Algebra in a localic topos with applications to ring theory" , ''Lect. notes in math.'' , '''1038''' , Springer (1983)</TD></TR><TR><TD valign="top">[a5]</TD> <TD valign="top">  B.R. Tennison,   "Sheaf theory" , Cambridge Univ. Press (1975)</TD></TR><TR><TD valign="top">[a6]</TD> <TD valign="top">  M. Artin,   "Grothendieck topologies" , Harvard Univ. Press  (1962)</TD></TR><TR><TD valign="top">[a7]</TD> <TD valign="top">  P.T. Johnstone,   "Topos theory" , Acad. Press  (1977)</TD></TR><TR><TD valign="top">[a8]</TD> <TD valign="top">  A. Grothendieck,   et al.,  "Théorie de topos et cohomologie des schémas (SGA 4–5)" , ''Lect. notes in math.'' , '''269, 270, 305, 589''' , Springer  (1972–1977)</TD></TR><TR><TD valign="top">[a9]</TD> <TD valign="top"M. Makkai,   G.E. Reyes,   "First order categorical logic" , ''Lect. notes in math.'' , '''611''' , Springer (1977)</TD></TR><TR><TD valign="top">[a10]</TD> <TD valign="top">  M. Tierney,   "Sheaf theory and the continuum hypothesis" F.W. Lawvere (ed.) , ''Toposes, algebraic geometry and logic (Dalhousic Univ., Jan. 1971)'' , ''Lect. notes in math.'' , '''274''' , Springer (1972) pp. 13–42</TD></TR><TR><TD valign="top">[a11]</TD> <TD valign="top"J.-P. Serre,   "Faisceaux algébriques cohérents" ''Ann. of Math.'' , '''61''' (1955) pp. 197–278</TD></TR><TR><TD valign="top">[a12]</TD> <TD valign="top">  P. Deligne,   "Cohomologie étale (SGA 4 1/2)" , ''Lect. notes in math.'' , '''569''' , Springer (1977)</TD></TR><TR><TD valign="top">[a13]</TD> <TD valign="top">  J.S. Milne,   "Etale cohomology" , Princeton Univ. Press (1980)</TD></TR><TR><TD valign="top">[a14]</TD> <TD valign="top">  E. Freitag,  R. Kiehl,   "Étale cohomology and the Weil conjecture" , Springer  (1988)</TD></TR><TR><TD valign="top">[a15]</TD> <TD valign="top">  J. Leray,   "Sur la forme des espaces topologiques et sur les points fixes des répresentations" ''J. Math. Pures Appl.'' , '''24'''  (1945)  pp. 95–167</TD></TR><TR><TD valign="top">[a16]</TD> <TD valign="top">  J. Leray,  "L'anneau spectral et l'anneau fibré d'homologie d'un espace localement compact et d'une application continue"  ''J. Math. Pures Appl.'' , '''29''' (1950) pp. 1–139</TD></TR><TR><TD valign="top">[a17]</TD> <TD valign="top"A. Weil,   "Foundations of algebraic geometry" , Amer. Math. Soc. (1946)</TD></TR><TR><TD valign="top">[a18]</TD> <TD valign="top"> , ''Sem. H. Cartan'' , '''1-''' , Secr. Math. Inst. H. Poincaré  (1948-)</TD></TR></table>
+
{|
 +
|-
 +
|valign="top"|{{Ref|Ar}}||valign="top"| M. Artin, "Grothendieck topologies", Harvard Univ. Press (1962) {{MR|}} {{ZBL|0208.48701}}
 +
|-
 +
|valign="top"|{{Ref|BoBo}}||valign="top"| F. Borceux, G. van den Bossche, "Algebra in a localic topos with applications to ring theory", ''Lect. notes in math.'', '''1038''', Springer (1983) {{MR|0724431}} {{ZBL|0522.18001}}
 +
|-
 +
|valign="top"|{{Ref|Br}}||valign="top"| G.E. Bredon, "Sheaf theory", McGraw-Hill (1967) {{MR|0221500}} {{ZBL|0158.20505}}
 +
|-
 +
|valign="top"|{{Ref|De}}||valign="top"| P. Deligne, "Cohomologie étale (SGA 4 1/2)", ''Lect. notes in math.'', '''569''', Springer (1977)
 +
|-
 +
|valign="top"|{{Ref|FoSc}}||valign="top"| M.P. Fourman, D.S. Scott, "Sheaves and logic" M.P. Fourman (ed.) C.J. Mulvey (ed.) D.S. Scott (ed.), ''Applications of sheaves'', ''Lect. notes in math.'', '''753''', Springer (1979) pp. 302–401 {{MR|0555551}} {{ZBL|0415.03053}} {{ZBL|0407.00001}}
 +
|-
 +
|valign="top"|{{Ref|FrKi}}||valign="top"| E. Freitag, R. Kiehl, "Étale cohomology and the Weil conjecture", Springer (1988) {{MR|0926276}} {{ZBL|0643.14012}}
 +
|-
 +
|valign="top"|{{Ref|Go}}||valign="top"| R. Godement, "Topologie algébrique et théorie des faisceaux", Hermann (1958) {{MR|0102797}} {{ZBL|0080.16201}}
 +
|-
 +
|valign="top"|{{Ref|Gr}}||valign="top"| A. Grothendieck, "Sur quelques points d'algèbre homologique" ''Tohoku Math. J.'', '''9''' (1957) pp. 119–221 {{MR|0102537}} {{ZBL|}}
 +
|-
 +
|valign="top"|{{Ref|Gr2}}||valign="top"| J.W. Gray, "Fragments of the history of sheaf theory" M.P. Fourman (ed.) C.J. Mulvey (ed.) D.S. Scott (ed.), ''Applications of sheaves'', ''Lect. notes in math.'', '''753''', Springer (1979) pp. 1–79 {{MR|0555539}} {{ZBL|0436.55002}}
 +
|-
 +
|valign="top"|{{Ref|Gr3}}||valign="top"| A. Grothendieck, et al., "Théorie de topos et cohomologie des schémas (SGA 4–5)", ''Lect. notes in math.'', '''269, 270, 305, 589''', Springer (1972–1977)
 +
|-
 +
|valign="top"|{{Ref|Jo}}||valign="top"| P.T. Johnstone, "Topos theory", Acad. Press (1977) {{MR|0470019}} {{ZBL|0368.18001}}
 +
|-
 +
|valign="top"|{{Ref|Le}}||valign="top"| J. Leray, "Sur la forme des espaces topologiques et sur les points fixes des répresentations" ''J. Math. Pures Appl.'', '''24''' (1945) pp. 95–167 {{MR|0015786}} {{ZBL|0060.40703}}
 +
|-
 +
|valign="top"|{{Ref|Le2}}||valign="top"| J. Leray, "L'anneau spectral et l'anneau fibré d'homologie d'un espace localement compact et d'une application continue" ''J. Math. Pures Appl.'', '''29''' (1950) pp. 1–139
 +
|-
 +
|valign="top"|{{Ref|MaRe}}||valign="top"| M. Makkai, G.E. Reyes, "First order categorical logic", ''Lect. notes in math.'', '''611''', Springer (1977) {{MR|0505486}} {{ZBL|0357.18002}}
 +
|-
 +
|valign="top"|{{Ref|Mi}}||valign="top"| J.S. Milne, "Etale cohomology", Princeton Univ. Press (1980) {{MR|0559531}} {{ZBL|0433.14012}}
 +
|-
 +
|valign="top"|{{Ref|Mu}}||valign="top"| C.J. Mulvey, "Representations of rings and modules" M.P. Fourman (ed.) C.J. Mulvey (ed.) D.S. Scott (ed.), ''Applications of sheaves'', ''Lect. notes in math.'', '''753''', Springer (1979) pp. 542–585 {{MR|0555561}} {{ZBL|0432.16015}}
 +
|-
 +
|valign="top"|{{Ref|Se}}||valign="top"| J.-P. Serre, "Faisceaux algébriques cohérents" ''Ann. of Math.'', '''61''' (1955) pp. 197–278 {{MR|0068874}} {{ZBL|0067.16201}}
 +
|-
 +
|valign="top"|{{Ref|Sk}}||valign="top"| E.G. Sklyarenko, "Homology and cohomology of general spaces", Springer (Forthcoming) (Translated from Russian) {{MR|1004026}} {{ZBL|0693.55004}} {{ZBL|0831.55003}}
 +
|-
 +
|valign="top"|{{Ref|Sw}}||valign="top"| R. Swan, "The theory of sheaves", Chicago Univ. Press (1964) {{MR|}} {{ZBL|0119.25801}}
 +
|-
 +
|valign="top"|{{Ref|Te}}||valign="top"| B.R. Tennison, "Sheaf theory", Cambridge Univ. Press (1975) {{MR|0404390}} {{ZBL|0313.18010}}
 +
|-
 +
|valign="top"|{{Ref|Ti}}||valign="top"| M. Tierney, "Sheaf theory and the continuum hypothesis" F.W. Lawvere (ed.), ''Toposes, algebraic geometry and logic (Dalhousic Univ., Jan. 1971)'', ''Lect. notes in math.'', '''274''', Springer (1972) pp. 13–42 {{MR|0373888}} {{ZBL|0244.18005}}
 +
|-
 +
|valign="top"|{{Ref|We}}||valign="top"| A. Weil, "Foundations of algebraic geometry", Amer. Math. Soc. (1946) {{MR|0023093}} {{ZBL|0063.08198}}
 +
|-
 +
|}

Latest revision as of 15:06, 15 July 2014

2020 Mathematics Subject Classification: Primary: 14-XX [MSN][ZBL]

A special mathematical tool which provides a unified approach to establishing connections between local and global properties of topological spaces (in particular geometric objects) and which is a powerful method for studying many problems in contemporary algebra, geometry, topology, and analysis.

A pre-sheaf $F$ on a topological space $X$ assigns to each open subset $U\subseteq X$ an Abelian group $F(U)$ (a ring, a module over a ring, etc.) and to every pair of open sets $V\subseteq U$ a homomorphism $F_V^U:F(U)\to F(V)$, such that $F_U^U$ is the identity isomorphism and $F_W^U=F_W^V F_V^U$ for every triple $W\subseteq V\subseteq U$. In other words, a pre-sheaf is a contravariant functor from the category of open subsets of $X$ and their inclusions into the category of groups (rings, etc.) and their homomorphisms. The mappings $F_V^U$ are called restriction homomorphisms (for example, if the elements of the stalk $F(U)$ are functions of some type or other defined on $U$, $F_V^U$ is the restriction of these to the smaller subset). A topology on the set $\def\cF{ {\mathcal F}}\cF = \bigcup_{x\in X}\cF_x$, where $\cF_x$ is, by definition, the direct limit $\lim_{x\in U} F(U)$, defined in the following way: For each $U\subseteq X$ and any $\def\s{\sigma}\s\in F(U)$, the set $S$ consisting of those points of $\cF_x$, $x\in U$, which are images of $\s$ in the definition of $\cF_x$ is declared to be open in $\cF$. In this topology the stalks $\cF_x$ are discrete, the stalk-wise algebraic operations defined on $\cF$ by taking direct limits are continuous and the natural projection $p:\cF\to X$, where $\cF = p^{-1}(x)$, is a local homeomorphism. The space $\cF$ together with the stalk-wise algebraic operations and the projection $p$ is called the sheaf of Abelian groups (rings, etc.) over $X$ associated with the pre-sheaf $F$.

Every continuous mapping $s:U\to \cF$ for which $x=ps(x)$ is called a section of $\cF$ over $U$. The section of $\cF$ over $X$ defined by the zeros in $\cF_x$ is called the zero section. If a section $s$ is zero at a point $x$, then $s$ coincides with the zero section in some neighbourhood of $x$; therefore the set of points at which $s$ is not zero (the support of $s$) is closed in $U$.

Let $\def\G{\Gamma}\G(U,\cF)$ (respectively, $\G_\Phi(X,\cF)$, where $\Phi$ is a certain family of closed sets in $X$; in particular, $\G_c(X,\cF)$) be the group (ring, module, etc.) of all sections of $\cF$ over $U$ (respectively, all sections over $X$ with supports in $\Phi$; in particular, sections with compact support). The assignment $U\mapsto\G(U,\cF)$ is a pre-sheaf over $X$, called the pre-sheaf of sections of the sheaf $\cF$. The assignment $\s\mapsto s$ used in defining the topology on $\cF$ also defines homomorphisms $F(U) \to \G(U,\cF)$ which commute with the restrictions to $V\subseteq U$, that is, it defines a homomorphism of pre-sheaves. This homomorphism is an isomorphism provided that the original pre-sheaf $F$ satisfies the requirements:

a) if $\def\l{\lambda}U=\bigcup_\l U_\l$ and $\s, \s'\in F(U)$, then $\s=\s'$ if the restrictions of $\s$ and $\s'$ to each $U_\l$ are equal; and

b) if $U=\bigcup_\l U_\l$ and $\s_\l\in F(U_\l)$ is a collection of elements such that the restrictions of $\s_\l$ and $s_\mu$ to $U_\l\cap U_\mu$ coincide, then there exists a $\s\in F(U)$ which has restriction to each $U_\l$ coinciding with $\s_\l$. The concept of a pre-sheaf satisfying these requirements is equivalent to the concept of the sheaf associated with it, therefore such pre-sheaves are quite often called sheaves also.

A sheaf of the form $X\times G$ (with the evident projection to $X$), where $G$ is a fixed group (ring, etc.), is called a constant sheaf, and is denoted by $G$. A sheaf which is constant in sufficiently small neighbourhoods of every $x\in X$ is called locally constant. The topology of such sheaves is separated (i.e. Hausdorff) if $X$ is a separated space. In more typical situations the topology of $\cF$ can be non-separated even if $X$ is separated (such is the case, for example, for the sheaf of germs of continuous (or differentiable) functions which is generated by the pre-sheaf $F$, where $F(U)$ is the set of continuous (differentiable) functions on $U$; however, the sheaf of germs of analytic functions on a manifold is separated).

Every homomorphism of pre-sheaves $F\to F'$ induces a mapping of the associated sheaves $\cF\to\cF'$, which is a local homeomorphism and maps stalks homomorphically to stalks; such a mapping of sheaves is called a sheaf homomorphism. Mono- and epimorphisms are defined in the standard way. For any sheaf homomorphism $f:\cF'\to\cF$ the image $f(\cF')$ is an open subset of $\cF$, closed with respect to the stalk-wise algebraic operations. Every subset of $\cF$ which satisfies these requirements is called a subsheaf of $\cF$. The quotient sheaf of the sheaf $\cF$ by a subsheaf $\cF'$ is defined as the sheaf $\cF''$ associated with the pre-sheaf $U\mapsto \G(U,\cF)/\G(U,\cF')$; moreover, there is an epimorphism $\cF\to\cF''$, and $\cF_X''\simeq \cF_X/\cF_X'$. For every open $U\subseteq X$ there is a subsheaf in $\cF$, denoted by $\cF_U$, which is the union of $p^{-1}(U)$ with the zero section of $\cF$ over $X$; $\cF_{X\setminus U}$ denotes the corresponding quotient sheaf (whose restriction to $X\setminus U$ coincides with the restriction thereto of $\cF$).

Since it is possible to interpret such ordinary terms as homomorphism, kernel, image, subsheaf, quotient sheaf, etc. for sheaves over $X$ in such a way that these concepts have essentially the same meaning as in algebra, one can consider them from a categorical point of view and apply to sheaf theory the constructions of homological algebra. The resulting category of sheaves over $X$ has the same classical properties as the category of Abelian groups or the category of modules; in particular, one can define for sheaves direct sums, infinite direct products, inductive limits, and other concepts.

The apparatus of sheaf theory has penetrated into various fields of mathematics thanks to the fact that there is a natural definition of the cohomology $\def\H{ {\rm H}}\H^*(X,\cF)$ of a space $X$ with coefficients in a sheaf $\cF$, and this without any kind of restrictions on $X$ (this is essential, for example, in algebraic geometry, where the spaces arising are, as a rule, non-separated) and to the fact that other cohomologies (under certain specific conditions) reduce to a sheaf cohomology, at least in those situations where their application is justified.

To define $\H^*(X,\cF)$ one first constructs the canonical resolution

$$\def\C{ {\rm C}}\C^*(\cF) : 0\to \cF\to\C^0(\cF)\to \C^1(\cF)\to \cdots,$$ where $\C^0(\cF)$ is the sheaf defined by the pre-sheaf $F$ for which $F(U)$ is the group of all (possibly discontinuous) sections of $\cF$ over $U$, so that $\G(U,\C^0(\cF))=F(U),\; \C^1(\cF) = \C^0(\C^0(\cF)/\cF)),\dots,\C^{p+1}(\cF)=\C^0(\C^p(\cF)/{\rm Im}\;\C^{p-1}(\cF)),\dots$. By definition, $\H^p(X,\cF) = \H^p(\G(X,\C^*(\cF)))$ ($\H_\Phi^p(X,\cF))$ is obtained by replacing the symbol $\G$ by $\G_\Phi$). The sheaf $\cF$ itself can be obtained from $C^*(\cF)$ so that $H_\Phi^0(X,\cF)=\G_\Phi(X,\cF)$ (in classical cohomology $\H^0(X,G)$ is the group of locally constant functions on $X$ with values in $G$). The resolution $C^*(\cF) $ is an exact covariant functor of $\cF$: there is a short exact sequence of resolutions corresponding to a short exact "coefficient" sequence $0\to\cF'\to\cF\to\cF''\to 0$. The functor $\G_\Phi$ turns out to be exact on the terms $C^p$, $p\ge0$, of the resolution, therefore there is an exact cohomology sequence

$$\cdots \to \H_\Phi^{p-1}(X,\cF'')\to \H_\Phi^{p}(X,\cF') \to\H_\Phi^p(X,\cF)\to\H_\Phi^{p}(X,\cF'')\to \cdots,$$

corresponding to the indicated coefficient sequence, beginning with $0\to\G_\Phi(X,\cF')\to\G_\Phi(X,\cF)\to\cdots$. The cohomology sequence of a pair $(X,A)$ corresponds to the short exact sequence, $0\to\cF_{X\setminus A}\to\cF\to\cF_A\to 0$ ($A$ is a closed set).

The cohomology groups $\H_\Phi^*(X,\cF)$ have the following "universality" property, which casts light on their meaning: For any other resolution $\def\cL{ {\mathcal L}}\cL^*$ (that is, an exact sequence of sheaves $\cL^q$ beginning with $\cF$) there is a natural "comparison" homomorphism $\H^p(\G_\Phi(X,\cL^*))\to H_\Phi^p(X,\cF)$, which is described in terms of $H_\Phi^p(X,\cL^q)$ by using spectral sequences. An important case is when the sheaves of the resolution are $\Phi$-acyclic, that is, when $H_\Phi^p(X,\cL^q)=0$ for $p\ge 1$; in this case the above homomorphism is an isomorphism. The basic examples of acyclic sheaves are flabby sheaves (for all $U\subseteq X$ the mappings $\G(X,\cL)\to \G(U,\cL)$ are epimorphic) and soft sheaves (any section over a closed set extends to a section over the whole of $X$). The canonical resolution consists of flabby sheaves. If $X$ is a paracompact space, then every flabby sheaf is also soft.

The universality property enables one to compare cohomologies arising in concrete situations with sheaf cohomology (and consequently also with each other), to discern for them the natural bounds within which their application is effective, and also to apply sheaf-theoretic methods to the solution of concrete problems. For example, Aleksandrov–Čech cohomology (cf. Aleksandrov–Čech homology and cohomology) can be defined using cochains obtained from the cochains of a specially selected system of open coverings by taking the direct limit. These cochains turn out to be sections of the sheaves of germs of cochains (defined analogously to the sheaves of germs of functions) constituting a resolution of the group (or even the sheaf) of coefficients, and this sheaf turns out to be soft if the space is paracompact. Thus, for paracompact spaces Aleksandrov–Čech cohomology coincides with a sheaf cohomology. An analogous conclusion holds for Zariski spaces (in particular, for algebraic varieties). Alexander–Spanier cochains also turn out to be sections of the sheaves of a resolution and, moreover, the resolution consists of soft sheaves if $X$ is paracompact, so in this case, in particular, Alexander–Spanier and Aleksandrov–Čech cohomology are naturally isomorphic. In the case of singular cohomology, identification of cochains which coincide on the "small" singular simplices, i.e. subordinated to (arbitrary) open coverings, leads to the so-called localized cochains (giving the same cohomology), which are sections of the sheaves determined by the pre-sheaves of the usual singular cochains. These sheaves turn out to be soft if $X$ is paracompact (if $X$ is hereditarily paracompact, then they are also flabby), but they form a resolution only under the additional requirement that $X$ is weakly locally contractible (in every neighbourhood $U$ of each point $x\in X$ there is a smaller neighbourhood which is contractible to a point inside $U$). A classic example is de Rham's theorem: The cohomology of the complex of differential forms of a differentiable manifold coincides with the usual cohomology with coefficients in the field $\R$ of real numbers (the sheaves of germs of differential forms are soft and form a resolution of $\R$: Sufficiently near to each point each closed differential form is exact).

There are also resolutions corresponding to any open or locally finite closed covering and these enable one to compare the cohomology of $X$ with the cohomology of the coverings (the spectral sequences for coverings). In particular, this gives an isomorphism if $\H^q =0$ for $q\ge 1$ for all elements of the covering and their finite intersections (Leray's theorem). Taking the direct limit with respect to open coverings gives an isomorphism between Aleksandrov–Čech cohomology $\def\cH{ {\mathcal H}}\check{\cH}^*$ and sheaf cohomology, even for non-paracompact $X$, provided that there are sufficiently many small open sets $U$ in $X$ for which $\check{\cH}^q(U,\cF)=0$ when $q\ge 1$ (Cartan's theorem). This means that the cohomologies $\check{\cH}^*$, used in algebraic geometry, with coefficients in coherent sheaves, are also isomorphic to the standard sheaf cohomology $\H^*$.

General constructions ensuring the existence of a comparison homomorphism enable one to compare also the cohomology $\def\cH{ {\mathcal H}}\H^p(X,\cH^q)$ with the hypercohomology $\H^*(\G(X,\cL^*))$ (analogously, $\H_\Phi^p(X,\cH^q)$ with $\H^*(\G_\Phi(X,\cL^*))$) in case $\cL^*$ is any differential sheaf (that is, a sheaf in which for any $q$ the composition $\cL^q\to\cL^{q+2}$ is zero) with $\cL^q$ acyclic, where the $\cH^q$ are the derived sheaves of $\cL^*$ (these are the quotient sheaves of the kernel by the image in each dimension $q$). The corresponding spectral sequences have many applications. Moreover, if $\cH^q=0$ when $q\ge 1$, then $\H^*(\G(X,\cL^*))=\H^*(X,\cH^0)$. For example, if in place of $\cL^*$ a sheaf of chains $\def\cC{ {\mathcal C}}\cC_*$ is taken (the boundary operator lowers the dimension by one, the elements of $\G(U,\cC_*)$ are the chains of the pair $(X,X\setminus U)$, and the stalk $\cH_q^x = \lim_{x\in U}\H_q(X,X\setminus U)=H_q(X,X\setminus x)$), then one obtains the way the homology $\H_*^\Phi(X,G)$ depends on all the possible $\H_\Phi^p(X,\cH_q)$. For a manifold, $\cH_q=0$ when $q>n=\dim X$, and $\H_p^\Phi(X,G)=\H_\Phi^{n-p}(X,\cH_n)$, that is, Poincaré duality holds. If $A$ is an open or closed subset of a locally compact space $X$, then the homology of $A$ is determined by the sections of $\cC_*$ with supports in $A$, and the homology of the pair $(X,A)$ is determined by the sections of the restriction of $\cC_*$ to $X\setminus A$. Conversely (and this is also one of the manifestations of Poincaré duality), if $X\setminus A$ is any flabby resolution for the cohomology, then the restriction of $\cC^*$ to $X\setminus A$ determines the cohomology of $X\setminus A$ and the sections of $\cC^*$ with supports in $A$ determine the cohomology of the pair $(X,X\setminus A)$. Since the sheaves $\cC_*$ are flabby for manifolds, the homology sequence of the pair $(X,A)$ coincides up to an inversion of the numbering with the cohomology sequence of the pair $(X,X\setminus A)$. This means that dualities for manifolds, such as Lefschetz duality $\H_p(X,U,G) = \H^{n-p}(X\setminus U,\cH_n)$, are particular cases of Poincaré duality. It turns out that the duality relations which are not covered by this scheme are corollaries of Poincaré duality and the acyclicity of the manifold in certain dimensions.

Just such a situation arises in the case of a continuous mapping $f:X\to Y$. A resolution for the cohomology of $X$ determines a certain differential sheaf $\cH^*$ on $Y$ for which the stalks $\cH_y^q$ are the direct limits of the cohomology groups $\H^q(f^{-1}(U),\cF)$ with respect to neighbourhoods $U$ of the points $y$ (and for closed mappings $\H_y^q = \H^q(f^{-1}(y),\cF)$), where $\H^*(X,\cF) = \H^*(\G(Y,\cH^*))$. The way $\H^*(X,\cF)$ depends on $\H^p(Y,\cH^q)$ is described by the Leray spectral sequence of the mapping $f$ (a particular case of this is the spectral sequence of a Serre fibration). Acyclic mappings correspond to the case when $\cH^q$ vanishes, thus ensuring that the cohomologies of $X$ and $Y$ with corresponding coefficients are isomorphic (Vietoris' theorem and its generalizations). The general constructions referred to above also give the spectral sequence of a mapping and take into account (along with their cohomological structure) the degree of disconnectedness of pre-images of points; this is especially effective for zero-dimensional or finite-to-one mappings (in the case of coverings it becomes the Cartan spectral sequence). There are also special spectral sequences in categories of $G$-spaces (spaces on which a group $G$ acts).

In sheaf cohomology there is a natural way of defining a multiplicative structure. The existence of special flabby resolutions, of which the mappings are determined by a certain semi-simplicial structure, enables one to give explicit formulas for the products of cochains, analogous to the usual ones. At the same time this also makes it possible to define other cohomology operations in sheaf theory.

The apparatus of sheaf theory has many applications wherever abstract homological methods are essential: in topology (homological and cohomological dimension, local homology and duality, the structure of various classes of continuous mappings, including imbeddings onto dense subsets, and, in particular, compactifications, etc.), in the theory of analytic manifolds (homology and cohomology with coefficients in coherent analytic sheaves and their applications, cohomology and analytic differential forms, homology and analytic flows (the analogue of de Rham's theorem), etc.), and also in abstract algebraic geometry (the cohomology of affine, projective and complete algebraic varieties with coefficients in coherent algebraic sheaves, algebraic Serre duality, algebraic (combinatorial) dimension, etc.).

Some basic notions of sheaf theory and spectral sequences appeared in the work of J. Leray (1945 and later) in connection with the study of homological properties of continuous mappings of locally compact spaces, and he also gave the definition of cohomology (with compact support) with coefficients in a sheaf. A fairly complete account of sheaf theory using resolutions was later given by H. Cartan. The proof of the de Rham theorem given by A. Weil (1947) and the work of J.-P. Serre (in the early 1950's) on algebraic varieties greatly influenced the development of sheaf theory. Cohomology with coefficients in a sheaf was first defined by the Aleksandrov–Čech method. A mature view of sheaf theory could be found by the end of the 1950's in the work of A. Grothendieck [Gr] and R. Godement [Go], where great generality was achieved and the methods were considerably simplified. E.g., it was shown that the category of sheaves over $X$ has a generator (that is, a sheaf $J$ admitting non-zero homomorphisms into any non-zero sheaf; for sheaves of Abelian groups, $J=\sum_{U\subseteq X}\Z_U$), and hence that each sheaf can be imbedded in an injective sheaf (Grothendieck's theorem). This is the reason for the formal analogy between cohomology theory with coefficients in sheaves and the theory of derived functors in the module categories: In the category of sheaves over $X$ there are "enough" injective objects (although, as a rule, there are few projective objects), and therefore one can freely apply all the corresponding techniques of homological algebra; in particular, one can define the cohomology $\H_\Phi^*(X,\cF)$ (without any restriction on $X$) as derived functors of the left exact functor $\G_\Phi(X,\cF)$ (or even as $\def\Ext{ {\rm Ext}}\Ext^*(\Z_X,\cF)$). This also sheds light, for example, on the general nature of such concepts as the cohomological dimension (over $\Z$) of a space, the algebraic dimension of a variety and the global dimension of a ring. The description given by Grothendieck of the spectral sequence for the functor $\Ext$ is essential in algebraic geometry. A much simpler method of constructing injective sheaves was found by Godement. He also showed that to construct a cohomology theory it is entirely sufficient to use his canonical flabby resolution, which, from the point of view of homological algebra, turns out to be simply one of the acyclic resolutions of a sheaf. Godement was the first to apply flabby and soft sheaves (soft sheaves are acyclic only for $X$ paracompact, which explains their use primarily in topology).

Comments

Let $f:X\to Y$ be a continuous mapping between topological spaces and $\cF$ a sheaf on $X$. Then the direct image of $\cF$ under $f$ is the sheaf $f_*\cF$ defined by $(f_*\cF)(U) = \cF(f^{-1}(U))$ for $U$ open in $Y$. The functor $\cF\mapsto f_*(\cF)$ is left-exact on sheaves of modules or Abelian groups. The higher direct images $f_*^q\cF$ of $\cF$ are the right derived functors of $f_*$. The sheaf $f_*^q\cF$ is the sheaf associated to the pre-sheaf $U\mapsto \H^q(f^{-1}(U),\cF)$.

Given a sheaf $\def\cG{ {\mathcal G}}\cG$ over $Y$, let $\pi:\cG\to Y$ be the associated "sheaf space" (so that $\cG(U)=$ sections of $\pi$ over $U\subseteq Y$). Now consider the fibre product (pull-back) $f^*\cG=X\times_Y\cG$. The inverse image sheaf $f^*\cG$ is the sheaf of sections of $f^*\cG\to X$.

The functor $f^*$ is exact (on sheafs of modules or Abelian groups). The two functors are adjoint to one another: ${\rm Hom}(f^*\cG,\cF) \simeq {\rm Hom}(\cG,f_*\cF)$.

A short sequence of sheaves of Abelian groups or modules $0\to\cF'\to\cF\to\cF''\to 0$ over a space $X$ is exact if and only if the corresponding sequence of stalks $0\to\cF_x'\to\cF_x\to\cF_x''\to 0$ is exact for every $x\in X$.

Grothendieck and his school vastly generalized sheaf theory, from sheaves on a space to the more general notion of sheaves on a site and that of a topos (cf. [Ar][Gr3]).

For the theory of coherent sheaves on algebraic varieties see also the fundamental paper [Gr2]. For the theory of sheaves in the étale topology and for $l$-adic sheaves see [Gr3], [De][FrKi].

For a detailed history of the origins of sheaf theory, see [Gr2]. The article above focuses almost entirely on the use of sheaf theory in cohomology theory, but it does have many other applications: for example, in the representation theory of rings and other algebraic systems (see, for example, [Mu], [BoBo]), and in logic, in the provision of models for constructive mathematics (see [FoSc], and also [MaRe], [Ti]).


References

[Ar] M. Artin, "Grothendieck topologies", Harvard Univ. Press (1962) Zbl 0208.48701
[BoBo] F. Borceux, G. van den Bossche, "Algebra in a localic topos with applications to ring theory", Lect. notes in math., 1038, Springer (1983) MR0724431 Zbl 0522.18001
[Br] G.E. Bredon, "Sheaf theory", McGraw-Hill (1967) MR0221500 Zbl 0158.20505
[De] P. Deligne, "Cohomologie étale (SGA 4 1/2)", Lect. notes in math., 569, Springer (1977)
[FoSc] M.P. Fourman, D.S. Scott, "Sheaves and logic" M.P. Fourman (ed.) C.J. Mulvey (ed.) D.S. Scott (ed.), Applications of sheaves, Lect. notes in math., 753, Springer (1979) pp. 302–401 MR0555551 Zbl 0415.03053 Zbl 0407.00001
[FrKi] E. Freitag, R. Kiehl, "Étale cohomology and the Weil conjecture", Springer (1988) MR0926276 Zbl 0643.14012
[Go] R. Godement, "Topologie algébrique et théorie des faisceaux", Hermann (1958) MR0102797 Zbl 0080.16201
[Gr] A. Grothendieck, "Sur quelques points d'algèbre homologique" Tohoku Math. J., 9 (1957) pp. 119–221 MR0102537
[Gr2] J.W. Gray, "Fragments of the history of sheaf theory" M.P. Fourman (ed.) C.J. Mulvey (ed.) D.S. Scott (ed.), Applications of sheaves, Lect. notes in math., 753, Springer (1979) pp. 1–79 MR0555539 Zbl 0436.55002
[Gr3] A. Grothendieck, et al., "Théorie de topos et cohomologie des schémas (SGA 4–5)", Lect. notes in math., 269, 270, 305, 589, Springer (1972–1977)
[Jo] P.T. Johnstone, "Topos theory", Acad. Press (1977) MR0470019 Zbl 0368.18001
[Le] J. Leray, "Sur la forme des espaces topologiques et sur les points fixes des répresentations" J. Math. Pures Appl., 24 (1945) pp. 95–167 MR0015786 Zbl 0060.40703
[Le2] J. Leray, "L'anneau spectral et l'anneau fibré d'homologie d'un espace localement compact et d'une application continue" J. Math. Pures Appl., 29 (1950) pp. 1–139
[MaRe] M. Makkai, G.E. Reyes, "First order categorical logic", Lect. notes in math., 611, Springer (1977) MR0505486 Zbl 0357.18002
[Mi] J.S. Milne, "Etale cohomology", Princeton Univ. Press (1980) MR0559531 Zbl 0433.14012
[Mu] C.J. Mulvey, "Representations of rings and modules" M.P. Fourman (ed.) C.J. Mulvey (ed.) D.S. Scott (ed.), Applications of sheaves, Lect. notes in math., 753, Springer (1979) pp. 542–585 MR0555561 Zbl 0432.16015
[Se] J.-P. Serre, "Faisceaux algébriques cohérents" Ann. of Math., 61 (1955) pp. 197–278 MR0068874 Zbl 0067.16201
[Sk] E.G. Sklyarenko, "Homology and cohomology of general spaces", Springer (Forthcoming) (Translated from Russian) MR1004026 Zbl 0693.55004 Zbl 0831.55003
[Sw] R. Swan, "The theory of sheaves", Chicago Univ. Press (1964) Zbl 0119.25801
[Te] B.R. Tennison, "Sheaf theory", Cambridge Univ. Press (1975) MR0404390 Zbl 0313.18010
[Ti] M. Tierney, "Sheaf theory and the continuum hypothesis" F.W. Lawvere (ed.), Toposes, algebraic geometry and logic (Dalhousic Univ., Jan. 1971), Lect. notes in math., 274, Springer (1972) pp. 13–42 MR0373888 Zbl 0244.18005
[We] A. Weil, "Foundations of algebraic geometry", Amer. Math. Soc. (1946) MR0023093 Zbl 0063.08198
How to Cite This Entry:
Sheaf theory. Encyclopedia of Mathematics. URL: http://encyclopediaofmath.org/index.php?title=Sheaf_theory&oldid=14202
This article was adapted from an original article by E.G. Sklyarenko (originator), which appeared in Encyclopedia of Mathematics - ISBN 1402006098. See original article