- Commit
- 8a163c1759003a8381dca2cfe35b6780280dafbe
- Parent
- 44628b7f88a599c910e33a7a169ef30d986fe91a
- Author
- Pablo <pablo-escobar@riseup.net>
- Date
Standardized the word "nonzero"/"non-zero"
Source code for my notes on representations of semisimple Lie algebras and Olivier Mathieu's classification of simple weight modules
Standardized the word "nonzero"/"non-zero"
5 files changed, 22 insertions, 22 deletions
Status | File Name | N° Changes | Insertions | Deletions |
Modified | sections/complete-reducibility.tex | 24 | 12 | 12 |
Modified | sections/introduction.tex | 2 | 1 | 1 |
Modified | sections/mathieu.tex | 4 | 2 | 2 |
Modified | sections/semisimple-algebras.tex | 2 | 1 | 1 |
Modified | sections/sl2-sl3.tex | 12 | 6 | 6 |
diff --git a/sections/complete-reducibility.tex b/sections/complete-reducibility.tex @@ -27,18 +27,18 @@ definitions. \begin{definition} A representation of \(\mathfrak{g}\) is called \emph{indecomposable} if it is - not isomorphic to the direct sum of two non-zero representations. + not isomorphic to the direct sum of two nonzero representations. \end{definition} \begin{definition} A representation of \(\mathfrak{g}\) is called \emph{irreducible} if it has - no non-zero subrepresentations. + no nonzero subrepresentations. \end{definition} \begin{example} The trivial representation \(K\) is an example of an irreducible representations. In fact, every \(1\)-dimensional representation \(V\) of a - Lie algebra \(\mathfrak{g}\) is irreducible: \(V\) has no non-zero proper + Lie algebra \(\mathfrak{g}\) is irreducible: \(V\) has no nonzero proper subspaces, let alone \(\mathfrak{g}\)-invariant subspaces. \end{example} @@ -57,7 +57,7 @@ representations. The existence of the decomposition should be clear from the definitions. Indeed, if \(V\) is representation of \(\mathfrak{g}\) a simple argument via induction in \(\dim V\) suffices to prove the existence: if \(V\) is indecomposable then there is nothing to prove, and if \(V\) is not -indecomposable then \(V = W \oplus U\) for some \(W, U \subsetneq V\) non-zero +indecomposable then \(V = W \oplus U\) for some \(W, U \subsetneq V\) nonzero subrepresentations, so that their dimensions are both strictly smaller than \(\dim V\) and the existence follows from the induction hypothesis. For a proof of uniqueness please refer to \cite{etingof}. @@ -78,7 +78,7 @@ this is not always the case. For instance\dots x & \mapsto \begin{pmatrix} 1 & 1 \\ 0 & 1 \end{pmatrix} \end{align*} is a representation of the Lie algebra \(K[x]\). Notice \(V\) has a single - non-zero proper subrepresentation, which is spanned by the vector \((1, 0)\). + nonzero proper subrepresentation, which is spanned by the vector \((1, 0)\). This is because if \((a + b, b) = \rho(x) \ (a, b) = \lambda (a, b)\) for some \(\lambda \in \CC\) then \(\lambda = 1\) and \(b = 0\). Hence \(V\) is indecomposable -- it cannot be broken into a direct sum of \(1\)-dimensional @@ -194,11 +194,11 @@ clear things up. Suppose without any loss in generality that \(i_k = k\) for all \(k\) and let \(j > n\). By the maximality of our set of indexes, there is some - non-zero \(w \in (V_j \oplus U) \cap W\). Say \(w = v_j + v_1 + \cdots + + nonzero \(w \in (V_j \oplus U) \cap W\). Say \(w = v_j + v_1 + \cdots + v_n\) with each \(v_i \in V_i\). Then \(v_j = w - v_1 - \cdots - v_n \in V_j - \cap (W \oplus U)\) is non-zero. Indeed, if this is not the case we find \(0 + \cap (W \oplus U)\) is nonzero. Indeed, if this is not the case we find \(0 \ne w = v_1 + \cdots + v_n \in \left( \bigoplus_{i = 1}^n V_i \right) \cap - W\), a contradiction. This implies \(V_j \cap (W \oplus U)\) is a non-zero + W\), a contradiction. This implies \(V_j \cap (W \oplus U)\) is a nonzero subrepresentation of \(V_j\). Since \(V_j\) is irreducible, \(V_j = V_j \cap (W \oplus U)\) and therefore \(V_j \subset W \oplus U\). Given the arbitrary choice of \(j\), it then follows \(V = W \oplus U\). @@ -609,7 +609,7 @@ a representation}. \begin{proposition} The Casimir element \(C_V \in \mathcal{U}(\mathfrak{g})\) is central, so that \(C_V : W \to W\) is an intertwining operator for any \(\mathfrak{g}\)-module - \(W\). Furthermore, \(C_V\) acts in \(V\) as a non-zero scalar operator + \(W\). Furthermore, \(C_V\) acts in \(V\) as a nonzero scalar operator whenever \(V\) is a non-trivial finite-dimensional irreducible representation of \(\mathfrak{g}\). \end{proposition} @@ -683,7 +683,7 @@ establish\dots 0 \arrow{r} & K \arrow{r} & W \arrow{r}{\pi} & K \arrow{r} & 0 \end{tikzcd} \end{equation} - implies \(W\) is 2-dimensional. Take any non-zero \(w \in W\) outside of the + implies \(W\) is 2-dimensional. Take any nonzero \(w \in W\) outside of the image of the inclusion \(K \to W\). % TODOOOOOOOOO: Fix this @@ -696,7 +696,7 @@ establish\dots Since \(\dim W = 2\), the irreducible component \(\mathcal{U}(\mathfrak{g}) \cdot w\) of \(w\) in \(W\) is either \(K w\) or \(W\) itself. But this component cannot be \(W\), since the image the inclusion \(K \to W\) is a - 1-dimensional representation -- i.e. a proper non-zero subrepresentation. + 1-dimensional representation -- i.e. a proper nonzero subrepresentation. Hence \(K w\) is invariant under the action of \(\mathfrak{g}\). In particular, \(X w = 0\) for all \(X \in \mathfrak{g}\). Since \(w\) lies outside the image of the inclusion \(K \to W\), \(\pi(w) \ne 0\) -- which is @@ -739,7 +739,7 @@ establish\dots Finally, we consider the case where \(V\) is not irreducible. Suppose \(H^1(\mathfrak{g}, W) = 0\) for all \(\mathfrak{g}\)-modules with \(\dim W < - \dim V\) and let \(W \subset V\) be a proper non-zero subrepresentation. Then + \dim V\) and let \(W \subset V\) be a proper nonzero subrepresentation. Then the exact sequence \begin{center} \begin{tikzcd}
diff --git a/sections/introduction.tex b/sections/introduction.tex @@ -443,7 +443,7 @@ Other interesting classes of Lie algebras are the so called \emph{simple} and \begin{definition}\label{thm:sesimple-algebra} A Lie algebra \(\mathfrak{g}\) is called \emph{semisimple} if it is the direct sum of simple Lie algebras. Equivalently, a Lie algebra - \(\mathfrak{g}\) is called \emph{semisimple} if it has no non-zero solvable + \(\mathfrak{g}\) is called \emph{semisimple} if it has no nonzero solvable ideals. \end{definition}
diff --git a/sections/mathieu.tex b/sections/mathieu.tex @@ -128,7 +128,7 @@ A particularly well behaved class of examples are the so called \notin 2 \mathbb{Z}\), so that \(K[x, x^{-1}] = \bigoplus_{k \in \mathbb{Z}} K x^k\) is a degree \(1\) admissible weight \(\mathfrak{sl}_2(K)\)-module. It follows from the remark at the end of example~\ref{ex:submod-is-weight-mod} - that any non-zero subrepresentation \(W \subset K[x, x^{-1}]\) must contain a + that any nonzero subrepresentation \(W \subset K[x, x^{-1}]\) must contain a monomial \(x^k\). But since the operators \(-\frac{\mathrm{d}}{\mathrm{d}x} + \frac{x^{-1}}{2}, x^2 \frac{\mathrm{d}}{\mathrm{d}x} + \frac{x}{2} : K[x, x^{-1}] \to K[x, x^{-1}]\) are both injective, this implies all other @@ -882,7 +882,7 @@ This wasn't an issue an example~\ref{ex:laurent-polynomial-mod} because we verified that the action of \(f \in \mathfrak{sl}_2(K)\) in \(K[x, x^{-1}]\) is injective. Since all weight spaces of \(K[x, x^{-1}]\) are \(1\)-dimensional, this implies the action of \(f\) is actually bijective, so we can obtain a -non-zero vector in \(K[x, x^{-1}]_{2 k} = K x^k\) for any \(k \in \mathbb{Z}\) +nonzero vector in \(K[x, x^{-1}]_{2 k} = K x^k\) for any \(k \in \mathbb{Z}\) by translating between weight spaced using \(f\) and \(f^{-1}\) -- here \(f^{-1}\) denote the differential operator \((- \sfrac{\mathrm{d}}{\mathrm{d}x} + \sfrac{x^{-1}}{2})^{-1}\), which is the
diff --git a/sections/semisimple-algebras.tex b/sections/semisimple-algebras.tex @@ -335,7 +335,7 @@ restriction of the Killing form to the Cartan subalgebra. For the second statement, note that if the eigenvalues of \(\mathfrak{h}\) do not span all of \(\mathfrak{h}^*\) then there is some \(H \in \mathfrak{h}\) - non-zero such that \(\alpha(H) = 0\) for all eigenvalues \(\alpha\), which is + nonzero such that \(\alpha(H) = 0\) for all eigenvalues \(\alpha\), which is to say, \(\operatorname{ad}(H) X = [H, X] = 0\) for all \(X \in \mathfrak{g}\). Another way of putting it is to say \(H\) is an element of the center \(\mathfrak{z}\) of \(\mathfrak{g}\), which is zero by the
diff --git a/sections/sl2-sl3.tex b/sections/sl2-sl3.tex @@ -77,7 +77,7 @@ around \(\lambda\). Our main objective is to show \(V\) is determined by this string of eigenvalues. To do so, we suppose without any loss in generality that -\(\lambda\) is the right-most eigenvalue of \(h\), fix some non-zero \(v \in +\(\lambda\) is the right-most eigenvalue of \(h\), fix some nonzero \(v \in V_\lambda\) and consider the set \(\{v, f v, f^2, v, \ldots\}\). \begin{theorem}\label{thm:basis-of-irr-rep} @@ -141,7 +141,7 @@ words\dots \begin{proof} If \(W\) is an irreducible representation of \(\mathfrak{sl}_2(K)\) whose right-most eigenvalue of \(h\) is \(\lambda\) and \(w \in W_\lambda\) is - non-zero, consider the linear isomorphism + nonzero, consider the linear isomorphism \begin{align*} T : V & \to W \\ f^k v & \mapsto f^k w @@ -323,7 +323,7 @@ half of the remaining elements of \(\mathfrak{sl}_3(K)\). This is exactly analogous to the situation we found in \(\mathfrak{sl}_2(K)\): \(h\) corresponds to the subalgebra \(\mathfrak{h}\), and the eigenvalues of \(h\) in turn correspond to linear functions \(\lambda : \mathfrak{h} \to k\) such that -\(H v = \lambda(H) \cdot v\) for each \(H \in \mathfrak{h}\) and some non-zero +\(H v = \lambda(H) \cdot v\) for each \(H \in \mathfrak{h}\) and some nonzero \(v \in V\). We call such functionals \(\lambda\) \emph{eigenvalues of \(\mathfrak{h}\)}, and we say \emph{\(v\) is an eigenvector of \(\mathfrak{h}\)}. @@ -464,7 +464,7 @@ adjoint action of \(\mathfrak{h}\). \begin{definition} Given a representation \(V\) of \(\mathfrak{sl}_3(K)\), we'll call the - non-zero eigenvalues of the action of \(\mathfrak{h}\) in \(V\) \emph{weights + nonzero eigenvalues of the action of \(\mathfrak{h}\) in \(V\) \emph{weights of \(V\)}. As you might have guessed, we'll correspondingly refer to eigenvectors and eigenspaces of a given weight by \emph{weight vectors} and \emph{weight spaces}. @@ -738,7 +738,7 @@ root spaces of \(\mathfrak{sl}_3(K)\) act on the weight spaces of \(V\) via translation, this implies that \(E_{1 2}\), \(E_{1 3}\) and \(E_{2 3}\) all annihilate \(V_\lambda\), or otherwise one of \(V_{\lambda + \alpha_1 - \alpha_2}\), \(V_{\lambda + \alpha_1 - \alpha_3}\) and \(V_{\lambda + \alpha_2 -- \alpha_3}\) would be non-zero -- which contradicts the hypothesis that +- \alpha_3}\) would be nonzero -- which contradicts the hypothesis that \(\lambda\) lies the furthest along the direction we chose. In other words\dots \begin{theorem} @@ -1104,7 +1104,7 @@ simpler than that. \oplus W\) generated by \(v + w\). Since \(v + w\) is a highest weight of \(V \oplus W\), it follows from corollary~\ref{thm:irr-component-of-high-vec} that \(U\) is irreducible. The projection maps \(\pi_1 : U \to V\), \(\pi_2 : - U \to W\), being non-zero homomorphism between irreducible representations of + U \to W\), being nonzero homomorphism between irreducible representations of \(\mathfrak{sl}_3(K)\) must be isomorphism. Finally, \[ V \cong U \cong W