The Stacks project

46.4 Higher exts of adequate functors

Let $A$ be a ring. In Lemma 46.3.16 we have seen that any extension of adequate functors in the category of module-valued functors on $\textit{Alg}_ A$ is adequate. In this section we show that the same remains true for higher ext groups.

Lemma 46.4.1. Let $A$ be a ring. For every module-valued functor $F$ on $\textit{Alg}_ A$ there exists a morphism $Q(F) \to F$ of module-valued functors on $\textit{Alg}_ A$ such that (1) $Q(F)$ is adequate and (2) for every adequate functor $G$ the map $\mathop{\mathrm{Hom}}\nolimits (G, Q(F)) \to \mathop{\mathrm{Hom}}\nolimits (G, F)$ is a bijection.

Proof. Choose a set $\{ L_ i\} _{i \in I}$ of linearly adequate functors such that every linearly adequate functor is isomorphic to one of the $L_ i$. This is possible. Suppose that we can find $Q(F) \to F$ with (1) and (2)' or every $i \in I$ the map $\mathop{\mathrm{Hom}}\nolimits (L_ i, Q(F)) \to \mathop{\mathrm{Hom}}\nolimits (L_ i, F)$ is a bijection. Then (2) holds. Namely, combining Lemmas 46.3.6 and 46.3.11 we see that every adequate functor $G$ sits in an exact sequence

\[ K \to L \to G \to 0 \]

with $K$ and $L$ direct sums of linearly adequate functors. Hence (2)' implies that $\mathop{\mathrm{Hom}}\nolimits (L, Q(F)) \to \mathop{\mathrm{Hom}}\nolimits (L, F)$ and $\mathop{\mathrm{Hom}}\nolimits (K, Q(F)) \to \mathop{\mathrm{Hom}}\nolimits (K, F)$ are bijections, whence the same thing for $G$.

Consider the category $\mathcal{I}$ whose objects are pairs $(i, \varphi )$ where $i \in I$ and $\varphi : L_ i \to F$ is a morphism. A morphism $(i, \varphi ) \to (i', \varphi ')$ is a map $\psi : L_ i \to L_{i'}$ such that $\varphi ' \circ \psi = \varphi $. Set

\[ Q(F) = \mathop{\mathrm{colim}}\nolimits _{(i, \varphi ) \in \mathop{\mathrm{Ob}}\nolimits (\mathcal{I})} L_ i \]

There is a natural map $Q(F) \to F$, by Lemma 46.3.12 it is adequate, and by construction it has property (2)'. $\square$

Lemma 46.4.2. Let $A$ be a ring. Denote $\mathcal{P}$ the category of module-valued functors on $\textit{Alg}_ A$ and $\mathcal{A}$ the category of adequate functors on $\textit{Alg}_ A$. Denote $i : \mathcal{A} \to \mathcal{P}$ the inclusion functor. Denote $Q : \mathcal{P} \to \mathcal{A}$ the construction of Lemma 46.4.1. Then

  1. $i$ is fully faithful, exact, and its image is a weak Serre subcategory,

  2. $\mathcal{P}$ has enough injectives,

  3. the functor $Q$ is a right adjoint to $i$ hence left exact,

  4. $Q$ transforms injectives into injectives,

  5. $\mathcal{A}$ has enough injectives.

Proof. This lemma just collects some facts we have already seen so far. Part (1) is clear from the definitions, the characterization of weak Serre subcategories (see Homology, Lemma 12.10.3), and Lemmas 46.3.10, 46.3.11, and 46.3.16. Recall that $\mathcal{P}$ is equivalent to the category $\textit{PMod}((\textit{Aff}/\mathop{\mathrm{Spec}}(A))_\tau , \mathcal{O})$. Hence (2) by Injectives, Proposition 19.8.5. Part (3) follows from Lemma 46.4.1 and Categories, Lemma 4.24.5. Parts (4) and (5) follow from Homology, Lemmas 12.29.1 and 12.29.3. $\square$

Let $A$ be a ring. As in Formal Deformation Theory, Section 90.11 given an $A$-algebra $B$ and an $B$-module $N$ we set $B[N]$ equal to the $R$-algebra with underlying $B$-module $B \oplus N$ with multiplication given by $(b, m)(b', m ') = (bb', bm' + b'm)$. Note that this construction is functorial in the pair $(B, N)$ where morphism $(B, N) \to (B', N')$ is given by an $A$-algebra map $B \to B'$ and an $B$-module map $N \to N'$. In some sense the functor $TF$ of pairs defined in the following lemma is the tangent space of $F$. Below we will only consider pairs $(B, N)$ such that $B[N]$ is an object of $\textit{Alg}_ A$.

Lemma 46.4.3. Let $A$ be a ring. Let $F$ be a module valued functor. For every $B \in \mathop{\mathrm{Ob}}\nolimits (\textit{Alg}_ A)$ and $B$-module $N$ there is a canonical decomposition

\[ F(B[N]) = F(B) \oplus TF(B, N) \]

characterized by the following properties

  1. $TF(B, N) = \mathop{\mathrm{Ker}}(F(B[N]) \to F(B))$,

  2. there is a $B$-module structure $TF(B, N)$ compatible with $B[N]$-module structure on $F(B[N])$,

  3. $TF$ is a functor from the category of pairs $(B, N)$,

  4. there are canonical maps $N \otimes _ B F(B) \to TF(B, N)$ inducing a transformation between functors defined on the category of pairs $(B, N)$,

  5. $TF(B, 0) = 0$ and the map $TF(B, N) \to TF(B, N')$ is zero when $N \to N'$ is the zero map.

Proof. Since $B \to B[N] \to B$ is the identity we see that $F(B) \to F(B[N])$ is a direct summand whose complement is $TF(N, B)$ as defined in (1). This construction is functorial in the pair $(B, N)$ simply because given a morphism of pairs $(B, N) \to (B', N')$ we obtain a commutative diagram

\[ \xymatrix{ B' \ar[r] & B'[N'] \ar[r] & B' \\ B \ar[r] \ar[u] & B[N] \ar[r] \ar[u] & B \ar[u] } \]

in $\textit{Alg}_ A$. The $B$-module structure comes from the $B[N]$-module structure and the ring map $B \to B[N]$. The map in (4) is the composition

\[ N \otimes _ B F(B) \longrightarrow B[N] \otimes _{B[N]} F(B[N]) \longrightarrow F(B[N]) \]

whose image is contained in $TF(B, N)$. (The first arrow uses the inclusions $N \to B[N]$ and $F(B) \to F(B[N])$ and the second arrow is the multiplication map.) If $N = 0$, then $B = B[N]$ hence $TF(B, 0) = 0$. If $N \to N'$ is zero then it factors as $N \to 0 \to N'$ hence the induced map is zero since $TF(B, 0) = 0$. $\square$

Let $A$ be a ring. Let $M$ be an $A$-module. Then the module-valued functor $\underline{M}$ has tangent space $T\underline{M}$ given by the rule $T\underline{M}(B, N) = N \otimes _ A M$. In particular, for $B$ given, the functor $N \mapsto T\underline{M}(B, N)$ is additive and right exact. It turns out this also holds for injective module-valued functors.

Lemma 46.4.4. Let $A$ be a ring. Let $I$ be an injective object of the category of module-valued functors. Then for any $B \in \mathop{\mathrm{Ob}}\nolimits (\textit{Alg}_ A)$ and short exact sequence $0 \to N_1 \to N \to N_2 \to 0$ of $B$-modules the sequence

\[ TI(B, N_1) \to TI(B, N) \to TI(B, N_2) \to 0 \]

is exact.

Proof. We will use the results of Lemma 46.4.3 without further mention. Denote $h : \textit{Alg}_ A \to \textit{Sets}$ the functor given by $h(C) = \mathop{\mathrm{Mor}}\nolimits _ A(B[N], C)$. Similarly for $h_1$ and $h_2$. The map $B[N] \to B[N_2]$ corresponding to the surjection $N \to N_2$ is surjective. It corresponds to a map $h_2 \to h$ such that $h_2(C) \to h(C)$ is injective for all $A$-algebras $C$. On the other hand, there are two maps $p, q : h \to h_1$, corresponding to the zero map $N_1 \to N$ and the injection $N_1 \to N$. Note that

\[ \xymatrix{ h_2 \ar[r] & h \ar@<1ex>[r] \ar@<-1ex>[r] & h_1 } \]

is an equalizer diagram. Denote $\mathcal{O}_ h$ the module-valued functor $C \mapsto \bigoplus _{h(C)} C$. Similarly for $\mathcal{O}_{h_1}$ and $\mathcal{O}_{h_2}$. Note that

\[ \mathop{\mathrm{Hom}}\nolimits _\mathcal {P}(\mathcal{O}_ h, F) = F(B[N]) \]

where $\mathcal{P}$ is the category of module-valued functors on $\textit{Alg}_ A$. We claim there is an equalizer diagram

\[ \xymatrix{ \mathcal{O}_{h_2} \ar[r] & \mathcal{O}_ h \ar@<1ex>[r] \ar@<-1ex>[r] & \mathcal{O}_{h_1} } \]

in $\mathcal{P}$. Namely, suppose that $C \in \mathop{\mathrm{Ob}}\nolimits (\textit{Alg}_ A)$ and $\xi = \sum _{i = 1, \ldots , n} c_ i \cdot f_ i$ where $c_ i \in C$ and $f_ i : B[N] \to C$ is an element of $\mathcal{O}_ h(C)$. If $p(\xi ) = q(\xi )$, then we see that

\[ \sum c_ i \cdot f_ i \circ z = \sum c_ i \cdot f_ i \circ y \]

where $z, y : B[N_1] \to B[N]$ are the maps $z : (b, m_1) \mapsto (b, 0)$ and $y : (b, m_1) \mapsto (b, m_1)$. This means that for every $i$ there exists a $j$ such that $f_ j \circ z = f_ i \circ y$. Clearly, this implies that $f_ i(N_1) = 0$, i.e., $f_ i$ factors through a unique map $\overline{f}_ i : B[N_2] \to C$. Hence $\xi $ is the image of $\overline{\xi } = \sum c_ i \cdot \overline{f}_ i$. Since $I$ is injective, it transforms this equalizer diagram into a coequalizer diagram

\[ \xymatrix{ I(B[N_1]) \ar@<1ex>[r] \ar@<-1ex>[r] & I(B[N]) \ar[r] & I(B[N_2]) } \]

This diagram is compatible with the direct sum decompositions $I(B[N]) = I(B) \oplus TI(B, N)$ and $I(B[N_ i]) = I(B) \oplus TI(B, N_ i)$. The zero map $N \to N_1$ induces the zero map $TI(B, N) \to TI(B, N_1)$. Thus we see that the coequalizer property above means we have an exact sequence $TI(B, N_1) \to TI(B, N) \to TI(B, N_2) \to 0$ as desired. $\square$

Lemma 46.4.5. Let $A$ be a ring. Let $F$ be a module-valued functor such that for any $B \in \mathop{\mathrm{Ob}}\nolimits (\textit{Alg}_ A)$ the functor $TF(B, -)$ on $B$-modules transforms a short exact sequence of $B$-modules into a right exact sequence. Then

  1. $TF(B, N_1 \oplus N_2) = TF(B, N_1) \oplus TF(B, N_2)$,

  2. there is a second functorial $B$-module structure on $TF(B, N)$ defined by setting $x \cdot b = TF(B, b\cdot 1_ N)(x)$ for $x \in TF(B, N)$ and $b \in B$,

  3. the canonical map $N \otimes _ B F(B) \to TF(B, N)$ of Lemma 46.4.3 is $B$-linear also with respect to the second $B$-module structure,

  4. given a finitely presented $B$-module $N$ there is a canonical isomorphism $TF(B, B) \otimes _ B N \to TF(B, N)$ where the tensor product uses the second $B$-module structure on $TF(B, B)$.

Proof. We will use the results of Lemma 46.4.3 without further mention. The maps $N_1 \to N_1 \oplus N_2$ and $N_2 \to N_1 \oplus N_2$ give a map $TF(B, N_1) \oplus TF(B, N_2) \to TF(B, N_1 \oplus N_2)$ which is injective since the maps $N_1 \oplus N_2 \to N_1$ and $N_1 \oplus N_2 \to N_2$ induce an inverse. Since $TF$ is right exact we see that $TF(B, N_1) \to TF(B, N_1 \oplus N_2) \to TF(B, N_2) \to 0$ is exact. Hence $TF(B, N_1) \oplus TF(B, N_2) \to TF(B, N_1 \oplus N_2)$ is an isomorphism. This proves (1).

To see (2) the only thing we need to show is that $x \cdot (b_1 + b_2) = x \cdot b_1 + x \cdot b_2$. (Associativity and additivity are clear.) To see this consider

\[ N \xrightarrow {(b_1, b_2)} N \oplus N \xrightarrow {+} N \]

and apply $TF(B, -)$.

Part (3) follows immediately from the fact that $N \otimes _ B F(B) \to TF(B, N)$ is functorial in the pair $(B, N)$.

Suppose $N$ is a finitely presented $B$-module. Choose a presentation $B^{\oplus m} \to B^{\oplus n} \to N \to 0$. This gives an exact sequence

\[ TF(B, B^{\oplus m}) \to TF(B, B^{\oplus n}) \to TF(B, N) \to 0 \]

by right exactness of $TF(B, -)$. By part (1) we can write $TF(B, B^{\oplus m}) = TF(B, B)^{\oplus m}$ and $TF(B, B^{\oplus n}) = TF(B, B)^{\oplus n}$. Next, suppose that $B^{\oplus m} \to B^{\oplus n}$ is given by the matrix $T = (b_{ij})$. Then the induced map $TF(B, B)^{\oplus m} \to TF(B, B)^{\oplus n}$ is given by the matrix with entries $TF(B, b_{ij} \cdot 1_ B)$. This combined with right exactness of $\otimes $ proves (4). $\square$

Example 46.4.6. Let $F$ be a module-valued functor as in Lemma 46.4.5. It is not always the case that the two module structures on $TF(B, N)$ agree. Here is an example. Suppose $A = \mathbf{F}_ p$ where $p$ is a prime. Set $F(B) = B$ but with $B$-module structure given by $b \cdot x = b^ px$. Then $TF(B, N) = N$ with $B$-module structure given by $b \cdot x = b^ px$ for $x \in N$. However, the second $B$-module structure is given by $x \cdot b = bx$. Note that in this case the canonical map $N \otimes _ B F(B) \to TF(B, N)$ is zero as raising an element $n \in B[N]$ to the $p$th power is zero.

In the following lemma we will frequently use the observation that if $0 \to F \to G \to H \to 0$ is an exact sequence of module-valued functors on $\textit{Alg}_ A$, then for any pair $(B, N)$ the sequence $0 \to TF(B, N) \to TG(B, N) \to TH(B, N) \to 0$ is exact. This follows from the fact that $0 \to F(B[N]) \to G(B[N]) \to H(B[N]) \to 0$ is exact.

Lemma 46.4.7. Let $A$ be a ring. For $F$ a module-valued functor on $\textit{Alg}_ A$ say $(*)$ holds if for all $B \in \mathop{\mathrm{Ob}}\nolimits (\textit{Alg}_ A)$ the functor $TF(B, -)$ on $B$-modules transforms a short exact sequence of $B$-modules into a right exact sequence. Let $0 \to F \to G \to H \to 0$ be a short exact sequence of module-valued functors on $\textit{Alg}_ A$.

  1. If $(*)$ holds for $F, G$ then $(*)$ holds for $H$.

  2. If $(*)$ holds for $F, H$ then $(*)$ holds for $G$.

  3. If $H' \to H$ is morphism of module-valued functors on $\textit{Alg}_ A$ and $(*)$ holds for $F$, $G$, $H$, and $H'$, then $(*)$ holds for $G \times _ H H'$.

Proof. Let $B$ be given. Let $0 \to N_1 \to N_2 \to N_3 \to 0$ be a short exact sequence of $B$-modules. Part (1) follows from a diagram chase in the diagram

\[ \xymatrix{ 0 \ar[r] & TF(B, N_1) \ar[r] \ar[d] & TG(B, N_1) \ar[r] \ar[d] & TH(B, N_1) \ar[r] \ar[d] & 0 \\ 0 \ar[r] & TF(B, N_2) \ar[r] \ar[d] & TG(B, N_2) \ar[r] \ar[d] & TH(B, N_2) \ar[r] \ar[d] & 0 \\ 0 \ar[r] & TF(B, N_3) \ar[r] \ar[d] & TG(B, N_3) \ar[r] \ar[d] & TH(B, N_3) \ar[r] & 0 \\ & 0 & 0 } \]

with exact horizontal rows and exact columns involving $TF$ and $TG$. To prove part (2) we do a diagram chase in the diagram

\[ \xymatrix{ 0 \ar[r] & TF(B, N_1) \ar[r] \ar[d] & TG(B, N_1) \ar[r] \ar[d] & TH(B, N_1) \ar[r] \ar[d] & 0 \\ 0 \ar[r] & TF(B, N_2) \ar[r] \ar[d] & TG(B, N_2) \ar[r] \ar[d] & TH(B, N_2) \ar[r] \ar[d] & 0 \\ 0 \ar[r] & TF(B, N_3) \ar[r] \ar[d] & TG(B, N_3) \ar[r] & TH(B, N_3) \ar[r] \ar[d] & 0 \\ & 0 & & 0 } \]

with exact horizontal rows and exact columns involving $TF$ and $TH$. Part (3) follows from part (2) as $G \times _ H H'$ sits in the exact sequence $0 \to F \to G \times _ H H' \to H' \to 0$. $\square$

Most of the work in this section was done in order to prove the following key vanishing result.

Lemma 46.4.8. Let $A$ be a ring. Let $M$, $P$ be $A$-modules with $P$ of finite presentation. Then $\mathop{\mathrm{Ext}}\nolimits ^ i_\mathcal {P}(\underline{P}, \underline{M}) = 0$ for $i > 0$ where $\mathcal{P}$ is the category of module-valued functors on $\textit{Alg}_ A$.

Proof. Choose an injective resolution $\underline{M} \to I^\bullet $ in $\mathcal{P}$, see Lemma 46.4.2. By Derived Categories, Lemma 13.27.2 any element of $\mathop{\mathrm{Ext}}\nolimits ^ i_\mathcal {P}(\underline{P}, \underline{M})$ comes from a morphism $\varphi : \underline{P} \to I^ i$ with $d^ i \circ \varphi = 0$. We will prove that the Yoneda extension

\[ E : 0 \to \underline{M} \to I^0 \to \ldots \to I^{i - 1} \times _{\mathop{\mathrm{Ker}}(d^ i)} \underline{P} \to \underline{P} \to 0 \]

of $\underline{P}$ by $\underline{M}$ associated to $\varphi $ is trivial, which will prove the lemma by Derived Categories, Lemma 13.27.5.

For $F$ a module-valued functor on $\textit{Alg}_ A$ say $(*)$ holds if for all $B \in \mathop{\mathrm{Ob}}\nolimits (\textit{Alg}_ A)$ the functor $TF(B, -)$ on $B$-modules transforms a short exact sequence of $B$-modules into a right exact sequence. Recall that the module-valued functors $\underline{M}, I^ n, \underline{P}$ each have property $(*)$, see Lemma 46.4.4 and the remarks preceding it. By splitting $0 \to \underline{M} \to I^\bullet $ into short exact sequences we find that each of the functors $\mathop{\mathrm{Im}}(d^{n - 1}) = \mathop{\mathrm{Ker}}(d^ n) \subset I^ n$ has property $(*)$ by Lemma 46.4.7 and also that $I^{i - 1} \times _{\mathop{\mathrm{Ker}}(d^ i)} \underline{P}$ has property $(*)$.

Thus we may assume the Yoneda extension is given as

\[ E : 0 \to \underline{M} \to F_{i - 1} \to \ldots \to F_0 \to \underline{P} \to 0 \]

where each of the module-valued functors $F_ j$ has property $(*)$. Set $G_ j(B) = TF_ j(B, B)$ viewed as a $B$-module via the second $B$-module structure defined in Lemma 46.4.5. Since $TF_ j$ is a functor on pairs we see that $G_ j$ is a module-valued functor on $\textit{Alg}_ A$. Moreover, since $E$ is an exact sequence the sequence $G_{j + 1} \to G_ j \to G_{j - 1}$ is exact (see remark preceding Lemma 46.4.7). Observe that $T\underline{M}(B, B) = M \otimes _ A B = \underline{M}(B)$ and that the two $B$-module structures agree on this. Thus we obtain a Yoneda extension

\[ E' : 0 \to \underline{M} \to G_{i - 1} \to \ldots \to G_0 \to \underline{P} \to 0 \]

Moreover, the canonical maps

\[ F_ j(B) = B \otimes _ B F_ j(B) \longrightarrow TF_ j(B, B) = G_ j(B) \]

of Lemma 46.4.3 (4) are $B$-linear by Lemma 46.4.5 (3) and functorial in $B$. Hence a map

\[ \xymatrix{ 0 \ar[r] & \underline{M} \ar[r] \ar[d]^1 & F_{i - 1} \ar[r] \ar[d] & \ldots \ar[r] & F_0 \ar[r] \ar[d] & \underline{P} \ar[r] \ar[d]^1 & 0 \\ 0 \ar[r] & \underline{M} \ar[r] & G_{i - 1} \ar[r] & \ldots \ar[r] & G_0 \ar[r] & \underline{P} \ar[r] & 0 } \]

of Yoneda extensions. In particular we see that $E$ and $E'$ have the same class in $\mathop{\mathrm{Ext}}\nolimits ^ i_\mathcal {P}(\underline{P}, \underline{M})$ by the lemma on Yoneda Exts mentioned above. Finally, let $N$ be a $A$-module of finite presentation. Then we see that

\[ 0 \to T\underline{M}(A, N) \to TF_{i - 1}(A, N) \to \ldots \to TF_0(A, N) \to T\underline{P}(A, N) \to 0 \]

is exact. By Lemma 46.4.5 (4) with $B = A$ this translates into the exactness of the sequence of $A$-modules

\[ 0 \to M \otimes _ A N \to G_{i - 1}(A) \otimes _ A N \to \ldots \to G_0(A) \otimes _ A N \to P \otimes _ A N \to 0 \]

Hence the sequence of $A$-modules $0 \to M \to G_{i - 1}(A) \to \ldots \to G_0(A) \to P \to 0$ is universally exact, in the sense that it remains exact on tensoring with any finitely presented $A$-module $N$. Let $K = \mathop{\mathrm{Ker}}(G_0(A) \to P)$ so that we have exact sequences

\[ 0 \to K \to G_0(A) \to P \to 0 \quad \text{and}\quad G_2(A) \to G_1(A) \to K \to 0 \]

Tensoring the second sequence with $N$ we obtain that $K \otimes _ A N = \mathop{\mathrm{Coker}}(G_2(A) \otimes _ A N \to G_1(A) \otimes _ A N)$. Exactness of $G_2(A) \otimes _ A N \to G_1(A) \otimes _ A N \to G_0(A) \otimes _ A N$ then implies that $K \otimes _ A N \to G_0(A) \otimes _ A N$ is injective. By Algebra, Theorem 10.82.3 this means that the $A$-module extension $0 \to K \to G_0(A) \to P \to 0$ is exact, and because $P$ is assumed of finite presentation this means the sequence is split, see Algebra, Lemma 10.82.4. Any splitting $P \to G_0(A)$ defines a map $\underline{P} \to G_0$ which splits the surjection $G_0 \to \underline{P}$. Thus the Yoneda extension $E'$ is equivalent to the trivial Yoneda extension and we win. $\square$

Lemma 46.4.9. Let $A$ be a ring. Let $M$ be an $A$-module. Let $L$ be a linearly adequate functor on $\textit{Alg}_ A$. Then $\mathop{\mathrm{Ext}}\nolimits ^ i_\mathcal {P}(L, \underline{M}) = 0$ for $i > 0$ where $\mathcal{P}$ is the category of module-valued functors on $\textit{Alg}_ A$.

Proof. Since $L$ is linearly adequate there exists an exact sequence

\[ 0 \to L \to \underline{A^{\oplus m}} \to \underline{A^{\oplus n}} \to \underline{P} \to 0 \]

Here $P = \mathop{\mathrm{Coker}}(A^{\oplus m} \to A^{\oplus n})$ is the cokernel of the map of finite free $A$-modules which is given by the definition of linearly adequate functors. By Lemma 46.4.8 we have the vanishing of $\mathop{\mathrm{Ext}}\nolimits ^ i_\mathcal {P}(\underline{P}, \underline{M})$ and $\mathop{\mathrm{Ext}}\nolimits ^ i_\mathcal {P}(\underline{A}, \underline{M})$ for $i > 0$. Let $K = \mathop{\mathrm{Ker}}(\underline{A^{\oplus n}} \to \underline{P})$. By the long exact sequence of Ext groups associated to the exact sequence $0 \to K \to \underline{A^{\oplus n}} \to \underline{P} \to 0$ we conclude that $\mathop{\mathrm{Ext}}\nolimits ^ i_\mathcal {P}(K, \underline{M}) = 0$ for $i > 0$. Repeating with the sequence $0 \to L \to \underline{A^{\oplus m}} \to K \to 0$ we win. $\square$

Lemma 46.4.10. With notation as in Lemma 46.4.2 we have $R^ pQ(F) = 0$ for all $p > 0$ and any adequate functor $F$.

Proof. Choose an exact sequence $0 \to F \to \underline{M^0} \to \underline{M^1}$. Set $M^2 = \mathop{\mathrm{Coker}}(M^0 \to M^1)$ so that $0 \to F \to \underline{M^0} \to \underline{M^1} \to \underline{M^2} \to 0$ is a resolution. By Derived Categories, Lemma 13.21.3 we obtain a spectral sequence

\[ R^ pQ(\underline{M^ q}) \Rightarrow R^{p + q}Q(F) \]

Since $Q(\underline{M^ q}) = \underline{M^ q}$ it suffices to prove $R^ pQ(\underline{M}) = 0$, $p > 0$ for any $A$-module $M$.

Choose an injective resolution $\underline{M} \to I^\bullet $ in the category $\mathcal{P}$. Suppose that $R^ iQ(\underline{M})$ is nonzero. Then $\mathop{\mathrm{Ker}}(Q(I^ i) \to Q(I^{i + 1}))$ is strictly bigger than the image of $Q(I^{i - 1}) \to Q(I^ i)$. Hence by Lemma 46.3.6 there exists a linearly adequate functor $L$ and a map $\varphi : L \to Q(I^ i)$ mapping into the kernel of $Q(I^ i) \to Q(I^{i + 1})$ which does not factor through the image of $Q(I^{i - 1}) \to Q(I^ i)$. Because $Q$ is a left adjoint to the inclusion functor the map $\varphi $ corresponds to a map $\varphi ' : L \to I^ i$ with the same properties. Thus $\varphi '$ gives a nonzero element of $\mathop{\mathrm{Ext}}\nolimits ^ i_\mathcal {P}(L, \underline{M})$ contradicting Lemma 46.4.9. $\square$


Comments (0)


Post a comment

Your email address will not be published. Required fields are marked.

In your comment you can use Markdown and LaTeX style mathematics (enclose it like $\pi$). A preview option is available if you wish to see how it works out (just click on the eye in the toolbar).

Unfortunately JavaScript is disabled in your browser, so the comment preview function will not work.

All contributions are licensed under the GNU Free Documentation License.




In order to prevent bots from posting comments, we would like you to prove that you are human. You can do this by filling in the name of the current tag in the following input field. As a reminder, this is tag 06Z5. Beware of the difference between the letter 'O' and the digit '0'.