lie-algebras-and-their-representations

Source code for my notes on representations of semisimple Lie algebras and Olivier Mathieu's classification of simple weight modules

sl2-sl3.tex (55845B)

   1 \chapter{Representations of \(\mathfrak{sl}_2(K)\) \& \(\mathfrak{sl}_3(K)\)}\label{ch:sl3}
   2 
   3 We are, once again, faced with the daunting task of classifying the
   4 finite-dimensional modules of a given (semisimple) algebra \(\mathfrak{g}\).
   5 Having reduced the problem a great deal, all its left is classifying the simple
   6 \(\mathfrak{g}\)-modules. We have encountered numerous examples of simple
   7 \(\mathfrak{g}\)-modules over the previous chapter, but we have yet to subject
   8 them to any serious scrutiny. In this chapter we begin a systematic
   9 investigation of simple modules by looking at concrete examples. Specifically,
  10 we will classify the simple finite-dimensional modules of certain
  11 low-dimensional semisimple Lie algebras: \(\mathfrak{sl}_2(K)\) and
  12 \(\mathfrak{sl}_3(K)\).
  13 
  14 The reason why we chose \(\mathfrak{sl}_2(K)\) is a simple one: throughout the
  15 previous chapters \(\mathfrak{sl}_2(K)\) has afforded us surprisingly
  16 illuminating examples. We begin our analysis by recalling that the elements
  17 \begin{align*}
  18   e & = \begin{pmatrix} 0 & 1 \\ 0 &  0 \end{pmatrix} &
  19   f & = \begin{pmatrix} 0 & 0 \\ 1 &  0 \end{pmatrix} &
  20   h & = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}
  21 \end{align*}
  22 form a basis for \(\mathfrak{sl}_2(K)\) and satisfy
  23 \begin{align*}
  24   [e, f] & = h & [h, f] & = -2 f & [h, e] = 2 e
  25 \end{align*}
  26 
  27 Let \(M\) be a finite-dimensional simple \(\mathfrak{sl}_2(K)\)-module. We now
  28 turn our attention to the action of \(h\) on \(M\), in particular, we
  29 investigate the subspace \(\bigoplus_{\lambda} M_\lambda \subset M\) -- where
  30 \(\lambda\) ranges over the eigenvalues of \(h\!\restriction_M\) and
  31 \(M_\lambda\) is the corresponding eigenspace.
  32 
  33 At this point, this is nothing short of a gamble: why look at the eigenvalues
  34 of \(h\)? The short answer is that, as we shall see, this will pay off. We will
  35 postpone the discussion about the real reason of why we chose \(h\), but for
  36 now we may notice that, perhaps surprisingly, the action \(h\!\restriction_M\)
  37 of \(h\) on a finite-dimensional simple \(\mathfrak{sl}_2(K)\)-module \(M\)
  38 is always a diagonalizable operator.
  39 
  40 Let \(\lambda\) be any eigenvalue of \(h\!\restriction_M\). Notice
  41 \(M_\lambda\) is in general not a \(\mathfrak{sl}_2(K)\)-submodule of \(M\).
  42 Indeed, if \(m \in M_\lambda\) then the identities
  43 \begin{align*}
  44   h \cdot (e \cdot m) &=  2e \cdot m + e h \cdot m = (\lambda + 2) e \cdot m \\
  45   h \cdot (f \cdot m) &= -2f \cdot m + f h \cdot m = (\lambda - 2) f \cdot m
  46 \end{align*}
  47 follow. In other words, \(e\) sends an element of \(M_\lambda\) to an element
  48 of \(M_{\lambda + 2}\), while \(f\) sends it to an element of \(M_{\lambda -
  49 2}\). Visually, we may draw
  50 \begin{center}
  51   \begin{tikzcd}
  52     \cdots          \rar[bend left=60]                          &
  53     M_{\lambda - 2} \rar[bend left=60]{e} \lar[bend left=60]    &
  54     M_{\lambda}     \rar[bend left=60]{e} \lar[bend left=60]{f} &
  55     M_{\lambda + 2} \rar[bend left=60]    \lar[bend left=60]{f} &
  56     \cdots                                \lar[bend left=60]
  57   \end{tikzcd}
  58 \end{center}
  59 
  60 This implies \(\bigoplus_\lambda M_\lambda\) is a
  61 \(\mathfrak{sl}_2(K)\)-submodule, so that \(\bigoplus_\lambda M_\lambda\) is
  62 either \(0\) or the entirety of \(M\) -- recall that \(M\) is simple. Since
  63 \(M\) is finite dimensional, \(h\!\restriction_M\) has at least one eigenvalue
  64 and therefore
  65 \[
  66   M = \bigoplus_\lambda M_\lambda
  67 \]
  68 
  69 Even more so, we have seen that for any eigenvalue \(\lambda \in K\) of
  70 \(h\!\restriction_M\), \(\bigoplus_{k \in \mathbb{Z}} M_{\lambda - 2 k}\) is a
  71 \(\mathfrak{sl}_2(K)\)-invariant subspace, which goes to show
  72 \[
  73   M = \bigoplus_{k \in \mathbb{Z}} M_{\lambda - 2 k},
  74 \]
  75 and the eigenvalues of \(h\) all have the form \(\lambda - 2 k\) for some
  76 \(k\). By the same token, if \(a\) is the greatest \(k \in \mathbb{Z}\) such
  77 that \(V_{\lambda - 2 k} \ne 0\) and, likewise, \(b\) is the smallest \(k \in
  78 \mathbb{Z}\) such that \(V_{\lambda - 2 k} \ne 0\) then
  79 \[
  80   M = \bigoplus_{\substack{k \in \mathbb{Z} \\ a \le k \le b}}
  81       M_{\lambda - 2 k}
  82 \]
  83 
  84 The eigenvalues of \(h\) thus form an unbroken string
  85 \[
  86   \ldots, \lambda - 4, \lambda - 2, \lambda, \lambda + 2, \lambda + 4, \ldots
  87 \]
  88 around \(\lambda\). Our main objective is to show \(M\) is determined by this
  89 string of eigenvalues. To do so, we suppose without any loss in generality that
  90 \(\lambda\) is the right-most eigenvalue of \(h\), fix some nonzero \(m \in
  91 M_\lambda\) and consider the set \(\{m, f \cdot m, f^2 \cdot m, \ldots\}\).
  92 
  93 \begin{proposition}\label{thm:basis-of-irr-rep}
  94   The set \(\{m, f \cdot m, f^2 \cdot m, \ldots\}\) is a basis for \(M\). In
  95   addition, the action of \(\mathfrak{sl}_2(K)\) on \(M\) is given by the
  96   formulas
  97   \begin{equation}\label{eq:irr-rep-of-sl2}
  98     \begin{aligned}
  99         f^k \cdot m & \overset{e}{\mapsto} k(\lambda + 1 - k) f^{k - 1} \cdot m
 100       & f^k \cdot m & \overset{f}{\mapsto} f^{k + 1} \cdot m
 101       & f^k \cdot m & \overset{h}{\mapsto} (\lambda - 2 k) f^k \cdot m
 102     \end{aligned}
 103   \end{equation}
 104 \end{proposition}
 105 
 106 \begin{proof}
 107   First of all, notice \(f^k \cdot m\) lies in \(M_{\lambda - 2 k}\), so that
 108   \(\{m, f \cdot m, f^2 \cdot m, \ldots\}\) is a set of linearly independent
 109   vectors. Hence it suffices to show \(M = K \langle m, f \cdot m, f^2 \cdot m,
 110   \ldots \rangle\), which in light of the fact that \(M\) is simple is the same
 111   as showing \(K \langle m, f \cdot m, f^2 \cdot m, \ldots \rangle\) is
 112   invariant under the action of \(\mathfrak{sl}_2(K)\).
 113 
 114   The fact that \(h \cdot (f^k \cdot m) \in K \langle m, f \cdot m, f^2 \cdot
 115   m, \ldots \rangle\) follows immediately from our previous assertion that
 116   \(f^k \cdot m \in M_{\lambda - 2 k}\) -- indeed, \(h \cdot (f^k \cdot m) =
 117   (\lambda - 2 k) f^k \cdot m \in K \langle m, f \cdot m, f^2 \cdot m, \ldots
 118   \rangle\), which also goes to show one of the formulas in
 119   (\ref{eq:irr-rep-of-sl2}). Seeing \(e \cdot (f^k \cdot m) \in K \langle m, f
 120   \cdot m, f^2 \cdot m, \ldots \rangle\) is a bit more complex. Clearly,
 121   \[
 122     \begin{split}
 123       e \cdot (f \cdot m)
 124       & = h \cdot m + f \cdot (e \cdot m) \\
 125       \text{(since \(\lambda\) is the right-most eigenvalue)}
 126       & = h \cdot m + f \cdot 0 \\
 127       & = \lambda m
 128     \end{split}
 129   \]
 130 
 131   Next we compute
 132   \[
 133     \begin{split}
 134       e \cdot (f^2 \cdot m)
 135       & = (h + fe) \cdot (f \cdot m) \\
 136       & = h \cdot (f \cdot m) + f \cdot (\lambda m) \\
 137       & = 2 (\lambda - 1) f \cdot m
 138     \end{split}
 139   \]
 140 
 141   The pattern is starting to become clear: \(e\) sends \(f^k \cdot m\) to a
 142   multiple of \(f^{k - 1} \cdot m\). Explicitly, it is not hard to check by
 143   induction that
 144   \[
 145     e \cdot (f^k \cdot m) = k (\lambda + 1 - k) \cdot f^{k - 1} m,
 146   \]
 147   which which is the first formula of (\ref{eq:irr-rep-of-sl2}).
 148 \end{proof}
 149 
 150 The significance of Proposition~\ref{thm:basis-of-irr-rep} should be
 151 self-evident: we have just provided a complete description of the action of
 152 \(\mathfrak{sl}_2(K)\) on \(M\). In particular, this goes to show\dots
 153 
 154 \begin{corollary}
 155   Every eigenspace of the action of \(h\) on \(M\) is \(1\)-dimensional.
 156 \end{corollary}
 157 
 158 \begin{proof}
 159   It suffices to note \(\{m, f \cdot m, f^2 \cdot m, \ldots \}\) is a basis for
 160   \(M\) consisting of eigenvalues of \(h\) and whose only element in
 161   \(M_{\lambda - 2 k}\) is \(f^k \cdot m\).
 162 \end{proof}
 163 
 164 \begin{corollary}\label{thm:sl2-find-weights}
 165   The eigenvalues of \(h\) in \(M\) form a symmetric, unbroken string of
 166   integers separated by intervals of length \(2\) whose right-most value is
 167   \(\dim M - 1\).
 168 \end{corollary}
 169 
 170 \begin{proof}
 171   If \(f^r\) is the lowest power of \(f\) that annihilates \(m\), it follows
 172   from the formulas in (\ref{eq:irr-rep-of-sl2}) that
 173   \[
 174     0 = e \cdot 0 = e \cdot (f^r \cdot m)
 175     = r (\lambda + 1 - r) f^{r - 1} \cdot m
 176   \]
 177 
 178   This implies \(\lambda + 1 - r = 0\) -- i.e. \(\lambda = r - 1 \in
 179   \mathbb{Z}\). Now since \(\{m, f \cdot m, f^2 \cdot m, \ldots, f^{r - 1}
 180   \cdot m\}\) is a basis for \(M\), \(r = \dim V\). Hence if \(\lambda =
 181   \dim V - 1\) then the eigenvalues of \(h\) are
 182   \[
 183     \ldots, \lambda - 6, \lambda - 4, \lambda - 2, \lambda
 184   \]
 185 
 186   To see that this string is symmetric around \(0\), simply note that the
 187   left-most eigenvalue of \(h\) is precisely \(\lambda - 2 (r - 1) =
 188   -\lambda\).
 189 \end{proof}
 190 
 191 Visually, the situation it thus
 192 \begin{center}
 193   \begin{tikzcd}
 194     M_{-\lambda}      \rar[bend left=60]{e}                       &
 195     M_{- \lambda + 2} \rar[bend left=60]{e} \lar[bend left=60]{f} &
 196     M_{- \lambda + 4} \rar[bend left=60]    \lar[bend left=60]{f} &
 197     \cdots            \rar[bend left=60]    \lar[bend left=60]    &
 198     M_{\lambda - 4}   \rar[bend left=60]{e} \lar[bend left=60]    &
 199     M_{\lambda - 2}   \rar[bend left=60]{e} \lar[bend left=60]{f} &
 200     M_\lambda                               \lar[bend left=60]{f}
 201   \end{tikzcd}
 202 \end{center}
 203 
 204 Corollary~\ref{thm:sl2-find-weights} can be used to find the eigenvalues of the
 205 action of \(h\) on an arbitrary finite-dimensional
 206 \(\mathfrak{sl}_2(K)\)-module. Namely, if \(M\) and \(N\) are
 207 \(\mathfrak{sl}_2(K)\)-modules, \(m \in M_\mu\) and \(n \in N_\mu\) then by
 208 computing
 209 \[
 210   h \cdot (m + n) = h \cdot m + h \cdot n = \mu (m + n)
 211 \]
 212 we can see that \((M \oplus N)_\mu = M_\mu + N_\mu\). Hence the set of
 213 eigenvalues of \(h\) in a \(\mathfrak{sl}_2(K)\)-module \(M\) is the union of
 214 the sets of eigenvalues in its simple components, and the corresponding
 215 eigenspaces are the direct sums of the eigenspaces of such simple components.
 216 
 217 In particular, if the eigenvalues of \(M\) all have the same parity -- i.e.
 218 they are either all even integers or all odd integers -- and the dimension of
 219 each eigenspace is no greater than \(1\) then \(M\) must be simple, for if \(N,
 220 L \subset M\) are submodules with \(M = N \oplus L\) then either \(N_\lambda =
 221 0\) for all \(\lambda\) or \(L_\lambda = 0\) for all \(\lambda \in
 222 \mathfrak{h}^*\). To conclude our analysis all it is left is to show that for
 223 each \(\lambda \in \mathbb{Z}\) with \(\lambda \ge 0\) there is some
 224 finite-dimensional simple \(M\) whose highest weight is \(\lambda\).
 225 Surprisingly, we have already encountered such a \(M\).
 226 
 227 \begin{theorem}\label{thm:sl2-exist-unique}
 228   For each \(\lambda \ge 0\), \(\lambda \in \mathbb{Z}\), there exists a unique
 229   simple \(\mathfrak{sl}_2(K)\)-module whose left-most eigenvalue of \(h\) is
 230   \(\lambda\).
 231 \end{theorem}
 232 
 233 \begin{proof}
 234   Let \(M = K[x, y]^{(\lambda)}\) be the \(\mathfrak{sl}_2(K)\)-module of
 235   homogeneous polynomials of degree \(\lambda\) in two variables, as in
 236   Example~\ref{ex:sl2-polynomial-subrep}. A simple calculation shows \(M_{n - 2
 237   k} = K x^{\lambda - k} y^k\) for \(k = 0, \ldots, \lambda\) and \(M_\mu = 0\)
 238   otherwise. In particular, the right-most eigenvalue of \(M\) is \(\lambda\).
 239   Alternatively, one can readily check that if \(K^2\) is the natural
 240   \(\mathfrak{sl}_2(K)\)-module, then \(M = \operatorname{Sym}^\lambda K^2\)
 241   satisfies the relations of (\ref{eq:irr-rep-of-sl2}). Indeed, the map
 242   \begin{align*}
 243     K[x, y]^{(\lambda)} & \to     \operatorname{Sym}^\lambda K^2 \\
 244              x^k y^\ell & \mapsto e_1^k \cdot e_2^\ell
 245   \end{align*}
 246   is an isomorphism.
 247 
 248   Either way, by the previous observation that a finite-dimensional
 249   \(\mathfrak{sl}_2(K)\)-module whose eigenvalues all have the same parity and
 250   whose corresponding eigenspace are all \(1\)-dimensional must be simple,
 251   \(M\) is simple. As for the uniqueness of \(M\), it suffices to notice that
 252   if \(N\) is a finite-dimensional simple \(\mathfrak{sl}_2(K)\)-module with
 253   right-most eigenvalue \(\lambda\) and \(n \in N_\lambda\) is nonzero then
 254   relations (\ref{eq:irr-rep-of-sl2}) imply the map
 255   \begin{align*}
 256               M & \to     N           \\
 257     f^k \cdot m & \mapsto f^k \cdot n
 258   \end{align*}
 259   is an isomorphism -- this is, in effect, precisely how the isomorphism \(K[x,
 260   y]^{(\lambda)} \isoto \operatorname{Sym}^\lambda K^2\) was constructed.
 261 \end{proof}
 262 
 263 Our initial gamble of studying the eigenvalues of \(h\) may have seemed
 264 arbitrary at first, but it payed off: we have \emph{completely} described
 265 \emph{all} simple \(\mathfrak{sl}_2(K)\)-modules. It is not yet clear, however,
 266 if any of this can be adapted to a general setting. In the following section we
 267 shall double down on our gamble by trying to reproduce some of these results
 268 for \(\mathfrak{sl}_3(K)\), hoping this will somehow lead us to a general
 269 solution. In the process of doing so we will find some important clues on why
 270 \(h\) was a sure bet and the race was fixed all along.
 271 
 272 \section{Representations of \(\mathfrak{sl}_{2 + 1}(K)\)}\label{sec:sl3-reps}
 273 
 274 The study of representations of \(\mathfrak{sl}_2(K)\) reminds me of the
 275 difference between the derivative of a function \(\mathbb{R} \to \mathbb{R}\)
 276 and that of a smooth map between manifolds: it is a simpler case of something
 277 greater, but in some sense it is too simple of a case, and the intuition we
 278 acquire from it can be a bit misleading in regards to the general setting. For
 279 instance, I distinctly remember my Calculus I teacher telling the class ``the
 280 derivative of the composition of two functions is not the composition of their
 281 derivatives'' -- which is, of course, the \emph{correct} formulation of the
 282 chain rule in the context of smooth manifolds.
 283 
 284 The same applies to \(\mathfrak{sl}_2(K)\). It is a simple and beautiful
 285 example, but unfortunately the general picture, modules of arbitrary semisimple
 286 algebras, lacks its simplicity. The general purpose of this section is to
 287 investigate to which extent the framework we developed for
 288 \(\mathfrak{sl}_2(K)\) can be generalized to other semisimple Lie algebras. Of
 289 course, the algebra \(\mathfrak{sl}_3(K)\) stands as a natural candidate for
 290 potential generalizations: \(\mathfrak{sl}_3(K) = \mathfrak{sl}_{2 + 1}(K)\)
 291 after all.
 292 
 293 Our approach is very straightforward: we will fix some simple
 294 \(\mathfrak{sl}_3(K)\)-module \(M\) and proceed step by step, at each point
 295 asking ourselves how we could possibly adapt the framework we laid out for
 296 \(\mathfrak{sl}_2(K)\). The first obvious question is one we have already asked
 297 ourselves: why \(h\)?  More specifically, why did we choose to study its
 298 eigenvalues and is there an analogue of \(h\) in \(\mathfrak{sl}_3(K)\)?
 299 
 300 The answer to the former question is one we will discuss at length in the next
 301 chapter, but for now we note that perhaps the most fundamental property of
 302 \(h\) is that \emph{there exists an eigenvector \(m\) of \(h\) that is
 303 annihilated by \(e\)} -- that being the generator of the right-most eigenspace
 304 of \(h\). This was instrumental to our explicit description of the simple
 305 \(\mathfrak{sl}_2(K)\)-modules culminating in
 306 Theorem~\ref{thm:sl2-exist-unique}.
 307 
 308 Our first task is to find some analogue of \(h\) in \(\mathfrak{sl}_3(K)\), but
 309 it is still unclear what exactly we are looking for. We could say we are
 310 looking for an element of \(M\) that is annihilated by some analogue of \(e\),
 311 but the meaning of \emph{some analogue of \(e\)} is again unclear. In fact, as
 312 we shall see, no such analogue exists and neither does such element. Instead,
 313 the actual way to proceed is to consider the subalgebra
 314 \[
 315   \mathfrak{h}
 316   = \left\{
 317     X \in
 318     \begin{pmatrix} K & 0 & 0 \\ 0 & K & 0 \\ 0 & 0 & K \end{pmatrix}
 319     : \operatorname{Tr}(X) = 0
 320     \right\}
 321 \]
 322 
 323 The choice of \(\mathfrak{h}\) may seem like an odd choice at the moment, but
 324 the point is we will later show that there exists some \(m \in M\) that is
 325 simultaneously an eigenvector of each \(H \in \mathfrak{h}\) and annihilated by
 326 half of the remaining elements of \(\mathfrak{sl}_3(K)\). This is exactly
 327 analogous to the situation we found in \(\mathfrak{sl}_2(K)\): \(h\)
 328 corresponds to the subalgebra \(\mathfrak{h}\), and the eigenvalues of \(h\) in
 329 turn correspond to linear functions \(\lambda : \mathfrak{h} \to k\) such that
 330 \(H \cdot m = \lambda(H) m\) for each \(H \in \mathfrak{h}\) and some nonzero
 331 \(m \in M\). We call such functionals \(\lambda\) \emph{eigenvalues of
 332 \(\mathfrak{h}\)}, and we say \emph{\(m\) is an eigenvector of
 333 \(\mathfrak{h}\)}.
 334 
 335 Once again, we will pay special attention to the eigenvalue decomposition
 336 \begin{equation}\label{eq:weight-module}
 337   M = \bigoplus_\lambda M_\lambda
 338 \end{equation}
 339 where \(\lambda\) ranges over all eigenvalues of \(\mathfrak{h}\) and
 340 \(M_\lambda = \{ m \in M : H \cdot m = \lambda(H) m, \forall H \in \mathfrak{h}
 341 \}\). We should note that the fact that (\ref{eq:weight-module}) holds is not
 342 at all obvious. This is because in general \(M_\lambda\) is not the eigenspace
 343 associated with an eigenvalue of any particular operator \(H \in
 344 \mathfrak{h}\), but instead the eigenspace of the action of the entire algebra
 345 \(\mathfrak{h}\). Fortunately for us, (\ref{eq:weight-module}) always holds,
 346 but we will postpone its proof to the next chapter.
 347 
 348 Next we turn our attention to the remaining elements of \(\mathfrak{sl}_3(K)\).
 349 In our analysis of \(\mathfrak{sl}_2(K)\) we saw that the eigenvalues of \(h\)
 350 differed from one another by multiples of \(2\). A possible way to interpret
 351 this is to say \emph{the eigenvalues of \(h\) differ from one another by
 352 integral linear combinations of the eigenvalues of the adjoint action of
 353 \(h\)}. In English, since
 354 \begin{align*}
 355   \operatorname{ad}(h) e & = 2 e  &
 356   \operatorname{ad}(h) f & = -2 f &
 357   \operatorname{ad}(h) h & = 0,
 358 \end{align*}
 359 the eigenvalues of the adjoint actions of \(h\) are \(0\) and \(\pm 2\), and
 360 the eigenvalues of the action of \(h\) on a simple
 361 \(\mathfrak{sl}_2(K)\)-module differ from one another by integral multiples of
 362 \(2\).
 363 
 364 In the case of \(\mathfrak{sl}_3(K)\), a simple calculation shows that if \([H,
 365 X]\) is scalar multiple of \(X\) for all \(H \in \mathfrak{h}\) then all but
 366 one entry of \(X\) are zero. Hence the eigenvectors of the adjoint action of
 367 \(\mathfrak{h}\) are \(E_{i j}\) and its eigenvalues are \(\epsilon_i -
 368 \epsilon_j\), where
 369 \[
 370   \epsilon_i
 371   \begin{pmatrix}
 372     a_1 &   0 &   0 \\
 373       0 & a_2 &   0 \\
 374       0 &   0 & a_3
 375   \end{pmatrix}
 376   = a_i
 377 \]
 378 
 379 Visually we may draw
 380 
 381 \begin{figure}[h]
 382   \centering
 383   \begin{tikzpicture}[scale=2]
 384     \begin{rootSystem}{A}
 385       \filldraw[black] \weight{0}{0} circle (.5pt);
 386       \node[black, above right] at \weight{0}{0} {\small$0$};
 387       \wt[black]{-1}{2}
 388       \wt[black]{-2}{1}
 389       \wt[black]{1}{1}
 390       \wt[black]{-1}{-1}
 391       \wt[black]{2}{-1}
 392       \wt[black]{1}{-2}
 393       \node[above] at \weight{-1}{2}  {$\epsilon_2 - \epsilon_3$};
 394       \node[left]  at \weight{-2}{1}  {$\epsilon_2 - \epsilon_1$};
 395       \node[right] at \weight{1}{1}   {$\epsilon_1 - \epsilon_3$};
 396       \node[left]  at \weight{-1}{-1} {$\epsilon_3 - \epsilon_1$};
 397       \node[right] at \weight{2}{-1}  {$\epsilon_1 - \epsilon_2$};
 398       \node[below] at \weight{1}{-2}  {$\epsilon_3 - \epsilon_1$};
 399       \node[black, above] at \weight{1}{0}  {$\epsilon_1$};
 400       \node[black, above] at \weight{-1}{1} {$\epsilon_2$};
 401       \node[black, above] at \weight{0}{-1} {$\epsilon_3$};
 402       \filldraw[black] \weight{1}{0}  circle (.5pt);
 403       \filldraw[black] \weight{-1}{1} circle (.5pt);
 404       \filldraw[black] \weight{0}{-1} circle (.5pt);
 405     \end{rootSystem}
 406   \end{tikzpicture}
 407 \end{figure}
 408 
 409 If we denote the eigenspace of the adjoint action of \(\mathfrak{h}\) on
 410 \(\mathfrak{sl}_3(K)\) associated to \(\alpha\) by
 411 \(\mathfrak{sl}_3(K)_\alpha\) and fix some \(X \in \mathfrak{sl}_3(K)_\alpha\),
 412 \(H \in \mathfrak{h}\) and \(m \in M_\lambda\) then
 413 \[
 414   \begin{split}
 415     H \cdot (X \cdot m)
 416     & = X \cdot (H \cdot m) + [H, X] \cdot m \\
 417     & = X \cdot (\lambda(H) m) + \alpha(H) X \cdot m \\
 418     & = (\lambda + \alpha)(H) X \cdot m
 419   \end{split}
 420 \]
 421 so that \(X\) carries \(m\) to \(M_{\lambda + \alpha}\). In other words,
 422 \(\mathfrak{sl}_3(k)_\alpha\) \emph{acts on \(M\) by translating vectors
 423 between eigenspaces}.
 424 
 425 For instance \(\mathfrak{sl}_3(K)_{\epsilon_1 - \epsilon_3}\) will act on the
 426 adjoint \(\mathfrak{sl}_3(K)\)-modules via
 427 \begin{figure}[h]
 428   \centering
 429   \begin{tikzpicture}[scale=2]
 430     \begin{rootSystem}{A}
 431       \wt[black]{0}{0}
 432       \wt[black]{-1}{2}
 433       \wt[black]{-2}{1}
 434       \wt[black]{1}{1}
 435       \wt[black]{-1}{-1}
 436       \wt[black]{2}{-1}
 437       \wt[black]{1}{-2}
 438       \draw[-latex, black] \weight{-1.9}{1.1} -- \weight{-1.1}{1.9};
 439       \draw[-latex, black] \weight{-.9}{-.9} -- \weight{-.1}{-.1};
 440       \draw[-latex, black] \weight{0.1}{0.1} -- \weight{.9}{.9};
 441       \draw[-latex, black] \weight{1.1}{-1.9} -- \weight{1.9}{-1.1};
 442     \end{rootSystem}
 443   \end{tikzpicture}
 444 \end{figure}
 445 
 446 This is again entirely analogous to the situation we observed in
 447 \(\mathfrak{sl}_2(K)\). In fact, we may once more conclude\dots
 448 
 449 \begin{theorem}\label{thm:sl3-weights-congruent-mod-root}
 450   The eigenvalues of the action of \(\mathfrak{h}\) on a simple
 451   \(\mathfrak{sl}_3(K)\)-module \(M\) differ from one another by integral
 452   linear combinations of the eigenvalues \(\epsilon_i - \epsilon_j\) of the adjoint
 453   action of \(\mathfrak{h}\) on \(\mathfrak{sl}_3(K)\).
 454 \end{theorem}
 455 
 456 \begin{proof}
 457   This proof goes exactly as that of the analogous statement for
 458   \(\mathfrak{sl}_2(K)\): it suffices to note that if we fix some eigenvalue
 459   \(\lambda\) of \(\mathfrak{h}\) and let \(i\) and \(j\) vary then
 460   \[
 461     \bigoplus_{i j} M_{\lambda + \epsilon_i - \epsilon_j}
 462   \]
 463   is an invariant subspace of \(M\).
 464 \end{proof}
 465 
 466 To avoid confusion we better introduce some notation to differentiate between
 467 eigenvalues of the action of \(\mathfrak{h}\) on \(M\) and eigenvalues of the
 468 adjoint action of \(\mathfrak{h}\).
 469 
 470 \begin{definition}\index{weights}
 471   Given a \(\mathfrak{sl}_3(K)\)-module \(M\), we will call the \emph{nonzero}
 472   eigenvalues of the action of \(\mathfrak{h}\) on \(M\) \emph{weights of
 473   \(M\)}. As you might have guessed, we will correspondingly refer to
 474   eigenvectors and eigenspaces of a given weight by \emph{weight vectors} and
 475   \emph{weight spaces}.
 476 \end{definition}
 477 
 478 It is clear from our previous discussion that the weights of the adjoint
 479 \(\mathfrak{sl}_3(K)\)-module deserve some special attention.
 480 
 481 \begin{definition}\index{weights!roots}
 482   The weights of the adjoint \(\mathfrak{sl}_3(K)\)-module are called
 483   \emph{roots of \(\mathfrak{sl}_3(K)\)}. Once again, the expressions
 484   \emph{root vector} and \emph{root space} are self-explanatory.
 485 \end{definition}
 486 
 487 Theorem~\ref{thm:sl3-weights-congruent-mod-root} can thus be restated as\dots
 488 
 489 \begin{definition}\index{weights!root lattice}
 490   The lattice \(Q = \mathbb{Z} \langle \epsilon_i - \epsilon_j : i, j = 1, 2, 3
 491   \rangle\) is called \emph{the root lattice of \(\mathfrak{sl}_3(K)\)}.
 492 \end{definition}
 493 
 494 \begin{corollary}
 495   The weights of a simple \(\mathfrak{sl}_3(K)\)-module \(M\) are all congruent
 496   modulo the root lattice \(Q\). In other words, the weights of \(M\) all lie
 497   in a single \(Q\)-coset \(\xi \in \mfrac{\mathfrak{h}^*}{Q}\).
 498 \end{corollary}
 499 
 500 At this point we could keep playing the tedious game of reproducing the
 501 arguments from the previous section in the context of \(\mathfrak{sl}_3(K)\).
 502 However, it is more profitable to use our knowledge of
 503 \(\mathfrak{sl}_2(K)\)-modules instead. Notice that the canonical inclusion
 504 \(\mathfrak{gl}_2(K) \to \mathfrak{gl}_3(K)\) -- as described in
 505 Example~\ref{ex:gln-inclusions} -- restricts to an injective homomorphism
 506 \(\mathfrak{sl}_2(K) \to \mathfrak{sl}_3(K)\). In other words,
 507 \(\mathfrak{sl}_2(K)\) is isomorphic to the image \(\mathfrak{s}_{1 2} = K
 508 \langle E_{1 2}, E_{2 1}, [E_{1 2}, E_{2 1}] \rangle \subset
 509 \mathfrak{sl}_3(K)\) of the inclusion \(\mathfrak{sl}_2(K) \to
 510 \mathfrak{sl}_3(K)\). We may thus regard \(M\) as a
 511 \(\mathfrak{sl}_2(K)\)-module by restricting to \(\mathfrak{s}_{1 2}\).
 512 
 513 Our first observation is that, since the root spaces act by translation, the
 514 subspace
 515 \[
 516   \bigoplus_{k \in \mathbb{Z}} M_{\lambda - k (\epsilon_1 - \epsilon_2)},
 517 \]
 518 must be invariant under the action of \(E_{1 2}\) and \(E_{2 1}\) for all
 519 \(\lambda \in \mathfrak{h}^*\). This goes to show \(\bigoplus_k M_{\lambda - k
 520 (\epsilon_1 - \epsilon_2)}\) is a \(\mathfrak{sl}_2(K)\)-submodule of \(M\) for all
 521 weights \(\lambda\) of \(M\). Furthermore, one can easily see that the
 522 eigenspace of the action of \(h\) on \(\bigoplus_{k \in \mathbb{Z}} M_{\lambda
 523 - k (\epsilon_1 - \epsilon_2)}\) associated with the eigenvalue \(\lambda(H) - 2k\)
 524 is precisely the weight space \(M_{\lambda - k (\epsilon_2 - \epsilon_1)}\).
 525 
 526 Visually,
 527 \begin{center}
 528   \begin{tikzpicture}
 529     \begin{rootSystem}{A}
 530       \node at \weight{-4}{2} (l) {};
 531       \node at \weight{-2}{1} (a) {};
 532       \node at \weight{0}{0}  (b) {};
 533       \node at \weight{2}{-1} (c) {};
 534       \node at \weight{4}{-2} (r) {};
 535       \draw \weight{-3}{1.5} -- \weight{3}{-1.5};
 536       \draw[dotted] \weight{-3}{1.5} -- (l);
 537       \draw[dotted] \weight{3}{-1.5} -- (r);
 538       \foreach \i in {-1, 0, 1}{\wt[black]{-2*\i}{\i}}
 539       \draw[-latex] (l) to[bend left=40] (a);
 540       \draw[-latex] (a) to[bend left=40] (b);
 541       \draw[-latex] (b) to[bend left=40] (c);
 542       \draw[-latex] (c) to[bend left=40] (r);
 543       \draw[-latex] (r) to[bend left=40] (c);
 544       \draw[-latex] (c) to[bend left=40] (b);
 545       \draw[-latex] (b) to[bend left=40] (a);
 546       \draw[-latex] (a) to[bend left=40] (l);
 547       \node[above right] at (b) {\small\(\lambda\)};
 548       \node[above right=2pt] at \weight{-3}{1.5} {\small\(E_{1 2}\)};
 549       \node[below left=2pt]  at \weight{-3}{1.5} {\small\(E_{2 1}\)};
 550     \end{rootSystem}
 551   \end{tikzpicture}
 552 \end{center}
 553 
 554 In general, we find\dots
 555 
 556 \begin{proposition}
 557   Given \(i < j\), the subalgebra \(\mathfrak{s}_{i j} = K \langle E_{i j},
 558   E_{j i}, [E_{i j}, E_{j i}] \rangle\) is isomorphic to
 559   \(\mathfrak{sl}_2(K)\). In addition, given a weight \(\lambda \in
 560   \mathfrak{h}^*\) of \(M\), the space
 561   \[
 562     N = \bigoplus_{k \in \mathbb{Z}} M_{\lambda - k (\epsilon_i - \epsilon_j)}
 563   \]
 564   is invariant under the action of \(\mathfrak{s}_{i j}\) and
 565   \[
 566     M_{\lambda - k (\epsilon_i - \epsilon_j)}
 567     = N_{\lambda([E_{i j}, E_{j i}]) - 2k}
 568   \]
 569 \end{proposition}
 570 
 571 \begin{proof}
 572   In effect, if \(i \ne k \ne j\) then \(\mathfrak{s}_{i j}\) is the subalgebra
 573   of matrices whose \(k\)-th row and \(k\)-th column are nil. For instance, if
 574   \(i = 1\) and \(j = 3\) then
 575   \[
 576     \mathfrak{s}_{1 3}
 577     = \begin{pmatrix} K & 0 & K \\ 0 & 0 & 0 \\ K & 0 & K \end{pmatrix}
 578       \cap \mathfrak{sl}_3(K)
 579   \]
 580 
 581   In this case, the map
 582   \begin{align*}
 583     \mathfrak{s}_{1 3} & \to \mathfrak{sl}_2(K) \\
 584     \begin{pmatrix} a & 0 & b \\ 0 & 0 & 0 \\ c & 0 & -a \end{pmatrix}
 585     & \mapsto
 586     \begin{pmatrix}
 587                   a &    \tm{topA}{0} &              b \\
 588       \tm{leftA}{0} &               0 & \tm{rightA}{0} \\
 589                   c & \tm{bottomA}{0} &             -a
 590     \end{pmatrix}
 591     = \begin{pmatrix} a & b \\ c & -a \end{pmatrix}
 592     \DrawVLine[black]{topA}{bottomA}
 593     \DrawHLine[black]{leftA}{rightA}
 594   \end{align*}
 595   is an isomorphism of Lie algebras. In general, the map
 596   \begin{align*}
 597     \mathfrak{s}_{i j} & \to     \mathfrak{sl}_2(K) \\
 598     E_{i j}            & \mapsto e                  \\
 599     E_{j i}            & \mapsto f                  \\
 600     [E_{i j}, E_{j i}] & \mapsto h
 601   \end{align*}
 602   which ``erases the \(k\)-th row and the \(k\)-th column'' of a matrix is an
 603   isomorphism.
 604 
 605   To see that \(N\) is invariant under the action of \(\mathfrak{s}_{i j}\), it
 606   suffices to notice \(E_{i j}\) and \(E_{j i}\) map \(m \in M_{\lambda - k
 607   (\epsilon_i - \epsilon_j)}\) to \(E_{i j} \cdot m \in M_{\lambda - (k - 1) (\epsilon_i -
 608   \epsilon_j)}\) and \(E_{j i} \cdot m \in M_{\lambda - (k + 1) (\epsilon_i -
 609   \epsilon_j)}\), respectively. Moreover,
 610   \[
 611     (\lambda - k (\epsilon_i - \epsilon_j))([E_{i j}, E_{j i}])
 612     = \lambda([E_{i j}, E_{j i}]) - k (1 - (-1))
 613     = \lambda([E_{i j}, E_{j i}]) - 2 k,
 614   \]
 615   which goes to show \(M_{\lambda - k (\epsilon_i - \epsilon_j)} \subset
 616   N_{\lambda([E_{i j}, E_{j i}]) - 2k}\). On the other hand, if we suppose \(0
 617   < \dim M_{\lambda - k (\epsilon_i - \epsilon_j)} < \dim N_{\lambda([E_{i j}, E_{j
 618   i}]) - 2 k}\) for some \(k\) we arrive at
 619   \[
 620     \dim N
 621     = \sum_k \dim M_{\lambda - k (\epsilon_i - \epsilon_j)}
 622     < \sum_k \dim N_{\lambda([E_{i j}, E_{j i}]) - 2k}
 623     = \dim N,
 624   \]
 625   a contradiction.
 626 \end{proof}
 627 
 628 As a first consequence of this, we show\dots
 629 
 630 \begin{definition}\index{weights!weight lattice}
 631   The lattice \(P = \mathbb{Z} \langle \epsilon_1, \epsilon_2, \epsilon_3 \rangle\)
 632   is called \emph{the weight lattice of \(\mathfrak{sl}_3(K)\)}.
 633 \end{definition}
 634 
 635 \begin{corollary}\label{thm:sl3-weights-fit-in-weight-lattice}
 636   Every weight \(\lambda\) of \(M\) lies in the weight lattice \(P\).
 637 \end{corollary}
 638 
 639 \begin{proof}
 640   It suffices to note \(\lambda([E_{i j}, E_{j i}])\) is an eigenvalue of \(h\)
 641   in a finite-dimensional \(\mathfrak{sl}_2(K)\)-module, so it must be an
 642   integer. Now since
 643   \[
 644     \lambda
 645     \begin{pmatrix}
 646       a & 0 & 0     \\
 647       0 & b & 0     \\
 648       0 & 0 & -a -b
 649     \end{pmatrix}
 650     =
 651     \lambda
 652     \begin{pmatrix}
 653       a & 0 & 0  \\
 654       0 & 0 & 0  \\
 655       0 & 0 & -a
 656     \end{pmatrix}
 657     +
 658     \lambda
 659     \begin{pmatrix}
 660       0 & 0 & 0  \\
 661       0 & b & 0  \\
 662       0 & 0 & -b
 663     \end{pmatrix}
 664     =
 665     a \lambda([E_{1 3}, E_{3 1}]) + b \lambda([E_{2 3}, E_{3 2}]),
 666   \]
 667   which is to say \(\lambda = \lambda([E_{1 3}, E_{3 1}]) \epsilon_1 +
 668   \lambda([E_{2 3}, E_{3 2}]) \epsilon_2 \in P\).
 669 \end{proof}
 670 
 671 There is a clear parallel between the case of \(\mathfrak{sl}_3(K)\) and that
 672 of \(\mathfrak{sl}_2(K)\), where we observed that the eigenvalues of the action
 673 of \(h\) all lied in the lattice \(P = \mathbb{Z}\) and were congruent modulo
 674 the sublattice \(Q = 2 \mathbb{Z}\).
 675 
 676 Among other things, this last result goes to show that the diagrams we have
 677 been drawing are in fact consistent with the theory we have developed. Namely,
 678 since all weights lie in the rational span of \(\{\epsilon_1, \epsilon_2,
 679 \epsilon_3\}\), we may as well draw them in the Cartesian plane. In fact, the
 680 attentive reader may notice that \(\kappa(E_{1 2}, E_{2 3}) = - \sfrac{1}{2}\),
 681 so that the angle -- with respect to the Killing form \(\kappa\) -- between the
 682 root vectors \(E_{1 2}\) and \(E_{2 3}\) is precisely the same as the angle
 683 between the points representing their roots \(\epsilon_1 - \epsilon_2\) and
 684 \(\epsilon_2 - \epsilon_3\) in the Cartesian plane. Since \(\epsilon_1 - \epsilon_2\)
 685 and \(\epsilon_2 - \epsilon_3\) span \(\mathfrak{h}^*\), this implies the diagrams
 686 we've been drawing are given by an isometry \(\mathbb{Q} P \isoto
 687 \mathbb{Q}^2\), where \(\mathbb{Q} P\) is endowed with the bilinear form
 688 defined by \((\epsilon_i - \epsilon_j, \epsilon_k - \alpha_\ell) \mapsto \kappa(E_{i
 689 j}, E_{k \ell})\) -- which we denote by \(\kappa\) as well.
 690 
 691 To proceed we once more refer to the previously established framework: next we
 692 saw that the eigenvalues of \(h\) form an unbroken string of integers symmetric
 693 around \(0\). To prove this we analyzed the right-most eigenvalues of \(h\) and
 694 their eigenvectors, providing an explicit description of the simple
 695 \(\mathfrak{sl}_2(K)\)-modules in terms of these vectors. We may
 696 reproduce these steps in the context of \(\mathfrak{sl}_3(K)\) by fixing a
 697 direction in the plane an considering the weight lying the furthest in that
 698 direction. 
 699 
 700 For instance, let's say we fix the direction
 701 \begin{center}
 702   \begin{tikzpicture}[scale=2]
 703     \begin{rootSystem}{A}
 704       \wt[black]{0}{0}
 705       \wt[black]{-1}{2}
 706       \wt[black]{-2}{1}
 707       \wt[black]{1}{1}
 708       \wt[black]{-1}{-1}
 709       \wt[black]{2}{-1}
 710       \wt[black]{1}{-2}
 711       \draw[-latex, black, thick] \weight{-1.5}{-.5} -- \weight{1.5}{.5};
 712     \end{rootSystem}
 713   \end{tikzpicture}
 714 \end{center}
 715 and let \(\lambda\) be the weight lying the furthest in this direction.
 716 
 717 Its easy to see what we mean intuitively by looking at the previous picture,
 718 but its precise meaning is still allusive. Formally this means we will choose a
 719 linear functional \(f : \mathbb{Q} P \to \mathbb{Q}\) and pick the weight that
 720 maximizes \(f\). To avoid any ambiguity we should choose the direction of a
 721 line irrational with respect to the root lattice \(Q\) -- for if \(f\) is not
 722 irrational there may be multiple choices the ``weight lying the furthest''
 723 along this direction.
 724 
 725 \begin{definition}
 726   We say that a root \(\alpha\) is positive if \(f(\alpha) > 0\) -- i.e. if it
 727   lies to the right of the direction we chose. Otherwise we say \(\alpha\) is
 728   negative. Notice that \(f(\alpha) \ne 0\) since by definition \(\alpha \ne
 729   0\) and \(f\) is irrational with respect to the lattice \(Q\).
 730 \end{definition}
 731 
 732 The next observation we make is that all others weights of \(M\) must lie in a
 733 sort of \(\frac{1}{3}\)-cone with apex at \(\lambda\), as shown in
 734 \begin{center}
 735   \begin{tikzpicture}
 736     \AutoSizeWeightLatticefalse
 737     \begin{rootSystem}{A}
 738       \weightLattice{3}
 739       \fill[gray!50,opacity=.2] (hex cs:x=5,y=-7) -- (hex cs:x=1,y=1) --
 740       (hex cs:x=-7,y=5) arc (150:270:{7*\weightLength});
 741       \draw[black, thick] (hex cs:x=5,y=-7) -- (hex cs:x=1,y=1) --
 742       (hex cs:x=-7,y=5);
 743       \filldraw[black] (hex cs:x=1,y=1) circle (1pt);
 744       \node[above right=-2pt] at (hex cs:x=1,y=1) {\small\(\lambda\)};
 745     \end{rootSystem}
 746   \end{tikzpicture}
 747 \end{center}
 748 
 749 Indeed, if this is not the case then, by definition, \(\lambda\) is not the
 750 weight placed the furthest in the direction we chose. Given our previous
 751 assertion that the root spaces of \(\mathfrak{sl}_3(K)\) act on the weight
 752 spaces of \(M\) via translation, this implies that \(E_{1 2}\), \(E_{1 3}\) and
 753 \(E_{2 3}\) all annihilate \(M_\lambda\), or otherwise one of \(M_{\lambda +
 754 \epsilon_1 - \epsilon_2}\), \(M_{\lambda + \epsilon_1 - \epsilon_3}\) and \(M_{\lambda
 755 + \epsilon_2 - \epsilon_3}\) would be nonzero -- which contradicts the hypothesis
 756 that \(\lambda\) lies the furthest in the direction we chose. In other
 757 words\dots
 758 
 759 \begin{proposition}\label{thm:sl3-mod-is-highest-weight}
 760   There is a weight vector \(m \in M\) that is annihilated by all positive root
 761   spaces of \(\mathfrak{sl}_3(K)\).
 762 \end{proposition}
 763 
 764 \begin{proof}
 765   It suffices to note that the positive roots of \(\mathfrak{sl}_3(K)\) are
 766   precisely \(\epsilon_1 - \epsilon_2\), \(\epsilon_1 - \epsilon_3\) and \(\epsilon_2 -
 767   \epsilon_3\), with root vectors \(E_{1 2}\), \(E_{1 3}\) and \(E_{2 3}\),
 768   respectively.
 769 \end{proof}
 770 
 771 \index{weights!highest weight}
 772 We call \(\lambda\) \emph{the highest weight of \(M\)}, and we call any nonzero
 773 \(m \in M_\lambda\) \emph{a highest weight vector}. Going back to the case of
 774 \(\mathfrak{sl}_2(K)\), we then constructed an explicit basis for our simple
 775 module in terms of a highest weight vector, which allowed us to provide an
 776 explicit description of the action of \(\mathfrak{sl}_2(K)\) in terms of its
 777 standard basis, and finally we concluded that the eigenvalues of \(h\) must be
 778 symmetrical around \(0\). An analogous procedure could be implemented for
 779 \(\mathfrak{sl}_3(K)\) -- and indeed that's what we will do later down the line
 780 -- but instead we would like to focus on the problem of finding the weights of
 781 \(M\) in the first place.
 782 
 783 We will start out by trying to understand the weights in the boundary of
 784 previously drawn cone. As we have just seen, we can get to other weight spaces
 785 from \(M_\lambda\) by successively applying \(E_{2 1}\).
 786 \begin{center}
 787   \begin{tikzpicture}
 788     \begin{rootSystem}{A}
 789       \node at \weight{3}{1}  (a) {};
 790       \node at \weight{1}{2}  (b) {};
 791       \node at \weight{-1}{3} (c) {};
 792       \node at \weight{-3}{4} (d) {};
 793       \node at \weight{-5}{5} (e) {};
 794       \draw \weight{3}{1} -- \weight{-4}{4.5};
 795       \draw[dotted] \weight{-4}{4.5} -- \weight{-5}{5};
 796       \foreach \i in {1,...,4}{\wt[black]{5-2*\i}{\i}}
 797       \node[above right=-2pt] at (hex cs:x=3,y=1){\small\(\lambda\)};
 798       \draw[-latex] (a) to[bend left=40] (b);
 799       \draw[-latex] (b) to[bend left=40] (c);
 800       \draw[-latex] (c) to[bend left=40] (d);
 801       \draw[-latex] (d) to[bend left=40] (e);
 802     \end{rootSystem}
 803   \end{tikzpicture}
 804 \end{center}
 805 
 806 Notice that \(\lambda([E_{1 2}, E_{2 1}]) \in \mathbb{Z}\) is the right-most
 807 eigenvalue of the \(\mathfrak{sl}_2(K)\)-module \(\bigoplus_{k \in \mathbb{Z}}
 808 M_{\lambda - k (\epsilon_1 - \epsilon_2)}\). In particular, \(\lambda([E_{1 2},
 809 E_{2 1}])\) must be positive. In addition, since the eigenspace of the
 810 eigenvalue \(\lambda([E_{1 2}, E_{2 1}]) - 2k\) of the action of \(h\) on
 811 \(\bigoplus_{k \in \mathbb{N}} M_{\lambda - k (\epsilon_1 - \epsilon_2)}\) is
 812 \(M_{\lambda - k (\epsilon_1 - \epsilon_2)}\), the weights of \(M\) appearing the
 813 string \(\lambda, \lambda + (\epsilon_1 - \epsilon_2), \ldots, \lambda + k
 814 (\epsilon_1 - \epsilon_2), \ldots\) must be symmetric with respect to the line
 815 \(\kappa(\epsilon_1 - \epsilon_2, \alpha) =  0\). The picture is thus
 816 \begin{center}
 817   \begin{tikzpicture}
 818     \AutoSizeWeightLatticefalse
 819     \begin{rootSystem}{A}
 820       \setlength{\weightRadius}{2pt}
 821       \weightLattice{4}
 822       \draw[thick] \weight{3}{1} -- \weight{-3}{4};
 823       \wt[black]{0}{0}
 824       \node[above left] at \weight{0}{0} {\small\(0\)};
 825       \foreach \i in {1,...,4}{\wt[black]{5-2*\i}{\i}}
 826       \node[above right=-2pt] at (hex cs:x=3,y=1){\small\(\lambda\)};
 827       \draw[very thick] \weight{0}{-4} -- \weight{0}{4}
 828       node[above]{\small\(\kappa(\epsilon_1 - \epsilon_2, \alpha) = 0\)};
 829     \end{rootSystem}
 830   \end{tikzpicture}
 831 \end{center}
 832 
 833 We could apply this same argument to the subspace \(\bigoplus_k M_{\lambda - k
 834 (\epsilon_2 - \epsilon_3)}\), so that the weights in this subspace must be
 835 symmetric with respect to the line \(\kappa(\epsilon_2 - \epsilon_3, \alpha) = 0\).
 836 The picture is now
 837 \begin{center}
 838   \begin{tikzpicture}
 839     \AutoSizeWeightLatticefalse
 840     \begin{rootSystem}{A}
 841       \setlength{\weightRadius}{2pt}
 842       \weightLattice{4}
 843       \draw[thick] \weight{3}{1} -- \weight{-3}{4};
 844       \draw[thick] \weight{3}{1} -- \weight{4}{-1};
 845       \wt[black]{0}{0}
 846       \wt[black]{4}{-1}
 847       \node[above left] at \weight{0}{0} {\small\(0\)};
 848       \foreach \i in {1,...,4}{\wt[black]{5-2*\i}{\i}}
 849       \node[above right=-2pt] at (hex cs:x=3,y=1){\small\(\lambda\)};
 850       \draw[very thick] \weight{0}{-4} -- \weight{0}{4}
 851       node[above]{\small\(\kappa(\epsilon_1 - \epsilon_2, \alpha) = 0\)};
 852       \draw[very thick] \weight{-4}{0} -- \weight{4}{0}
 853       node[right]{\small\(\kappa(\epsilon_2 - \epsilon_3, \alpha) = 0\)};
 854     \end{rootSystem}
 855   \end{tikzpicture}
 856 \end{center}
 857 
 858 Needless to say, we could keep applying this method to the weights at the ends
 859 of our string, arriving at
 860 \begin{center}
 861   \begin{tikzpicture}
 862     \AutoSizeWeightLatticefalse
 863     \begin{rootSystem}{A}
 864       \setlength{\weightRadius}{2pt}
 865       \weightLattice{5}
 866       \draw[thick] \weight{3}{1} -- \weight{-3}{4};
 867       \draw[thick] \weight{3}{1} -- \weight{4}{-1};
 868       \draw[thick] \weight{-3}{4} -- \weight{-4}{3};
 869       \draw[thick] \weight{-4}{3} -- \weight{-1}{-3};
 870       \draw[thick] \weight{1}{-4} -- \weight{4}{-1};
 871       \draw[thick] \weight{-1}{-3} -- \weight{1}{-4};
 872       \wt[black]{-4}{3}
 873       \wt[black]{-3}{1}
 874       \wt[black]{-2}{-1}
 875       \wt[black]{-1}{-3}
 876       \wt[black]{1}{-4}
 877       \wt[black]{2}{-3}
 878       \wt[black]{3}{-2}
 879       \wt[black]{4}{-1}
 880       \foreach \i in {1,...,4}{\wt[black]{5-2*\i}{\i}}
 881       \node[above right=-2pt] at (hex cs:x=3,y=1){\small\(\lambda\)};
 882       \draw[very thick] \weight{-5}{5} -- \weight{5}{-5};
 883       \draw[very thick] \weight{0}{-5} -- \weight{0}{5};
 884       \draw[very thick] \weight{-5}{0} -- \weight{5}{0};
 885     \end{rootSystem}
 886   \end{tikzpicture}
 887 \end{center}
 888 
 889 We claim all dots \(\mu\) lying inside the hexagon we have drawn must also be
 890 weights -- i.e. \(M_\mu \ne 0\). Indeed, by applying the same argument to an
 891 arbitrary weight \(\nu\) in the boundary of the hexagon we get a
 892 \(\mathfrak{sl}_2(K)\)-module whose weights correspond to weights of \(M\)
 893 lying in a string inside the hexagon, and whose right-most weight is precisely
 894 the weight of \(M\) we started with.
 895 \begin{center}
 896   \begin{tikzpicture}
 897     \AutoSizeWeightLatticefalse
 898     \begin{rootSystem}{A}
 899       \setlength{\weightRadius}{2pt}
 900       \weightLattice{5}
 901       \draw[thick] \weight{3}{1}   -- \weight{-3}{4};
 902       \draw[thick] \weight{3}{1}   -- \weight{4}{-1};
 903       \draw[thick] \weight{-3}{4}  -- \weight{-4}{3};
 904       \draw[thick] \weight{-4}{3}  -- \weight{-1}{-3};
 905       \draw[thick] \weight{1}{-4}  -- \weight{4}{-1};
 906       \draw[thick] \weight{-1}{-3} -- \weight{1}{-4};
 907       \wt[black]{-4}{3}
 908       \wt[black]{-3}{1}
 909       \wt[black]{-2}{-1}
 910       \wt[black]{-1}{-3}
 911       \wt[black]{1}{-4}
 912       \wt[black]{2}{-3}
 913       \wt[black]{3}{-2}
 914       \wt[black]{4}{-1}
 915       \foreach \i in {1,...,4}{\wt[black]{5-2*\i}{\i}}
 916       \node[above right=-2pt] at \weight{1}{2} {\small\(\nu\)};
 917       \node[above right=-2pt] at (hex cs:x=3,y=1){\small\(\lambda\)};
 918       \draw[very thick] \weight{-5}{5} -- \weight{5}{-5};
 919       \draw[very thick] \weight{0}{-5} -- \weight{0}{5};
 920       \draw[very thick] \weight{-5}{0} -- \weight{5}{0};
 921       \draw[dashed, thick] \weight{1}{2} -- \weight{-2}{-1};
 922       \wt[black]{1}{2}
 923       \wt[black]{-2}{-1}
 924       \wt[black]{0}{1}
 925       \wt[black]{-1}{0}
 926     \end{rootSystem}
 927   \end{tikzpicture}
 928 \end{center}
 929 
 930 By construction, \(\nu\) corresponds to the right-most weight of a
 931 \(\mathfrak{sl}_2(K)\)-module, so that all dots lying on the dashed string must
 932 occur in \(\mathfrak{sl}_2(K)\)-module. Hence they must also be weights of
 933 \(M\). The final picture is thus
 934 \begin{center}
 935   \begin{tikzpicture}
 936     \AutoSizeWeightLatticefalse
 937     \begin{rootSystem}{A}
 938       \setlength{\weightRadius}{2pt}
 939       \weightLattice{5}
 940       \draw[thick] \weight{3}{1} -- \weight{-3}{4};
 941       \draw[thick] \weight{3}{1} -- \weight{4}{-1};
 942       \draw[thick] \weight{-3}{4} -- \weight{-4}{3};
 943       \draw[thick] \weight{-4}{3} -- \weight{-1}{-3};
 944       \draw[thick] \weight{1}{-4} -- \weight{4}{-1};
 945       \draw[thick] \weight{-1}{-3} -- \weight{1}{-4};
 946       \wt[black]{-4}{3}
 947       \wt[black]{-3}{1}
 948       \wt[black]{-2}{-1}
 949       \wt[black]{-1}{-3}
 950       \wt[black]{1}{-4}
 951       \wt[black]{2}{-3}
 952       \wt[black]{3}{-2}
 953       \wt[black]{4}{-1}
 954       \foreach \i in {1,...,4}{\wt[black]{5-2*\i}{\i}}
 955       \node[above right=-2pt] at (hex cs:x=3,y=1){\small\(\lambda\)};
 956       \draw[very thick] \weight{-5}{5} -- \weight{5}{-5};
 957       \draw[very thick] \weight{0}{-5} -- \weight{0}{5};
 958       \draw[very thick] \weight{-5}{0} -- \weight{5}{0};
 959       \wt[black]{-2}{2}
 960       \wt[black]{0}{1}
 961       \wt[black]{-1}{0}
 962       \wt[black]{0}{-2}
 963       \wt[black]{1}{-1}
 964       \wt[black]{2}{0}
 965     \end{rootSystem}
 966   \end{tikzpicture}
 967 \end{center}
 968 
 969 \index{weights!weight diagrams}
 970 This final picture is known as \emph{the weight diagram of \(M\)}. Finally\dots
 971 
 972 \begin{theorem}\label{thm:sl3-irr-weights-class}
 973   The weights of \(M\) are precisely the elements of the weight lattice \(P\)
 974   congruent to \(\lambda\) module the sublattice \(Q\) and lying inside hexagon
 975   with vertices the images of \(\lambda\) under the group generated by
 976   reflections across the lines \(\kappa(\epsilon_i - \epsilon_j, \alpha) = 0\).
 977 \end{theorem}
 978 
 979 Having found all of the weights of \(M\), the only thing we are missing is an
 980 existence and uniqueness theorem analogous to
 981 Theorem~\ref{thm:sl2-exist-unique}. It is clear from the symmetries of the
 982 locus of weights found in Theorem~\ref{thm:sl3-irr-weights-class} that if
 983 \(\lambda \in P\) is the highest weight of some finite-dimensional simple
 984 \(\mathfrak{sl}_3(K)\)-module \(M\) then \(\lambda\) lies in the cone
 985 \(\mathbb{N} \langle \epsilon_1, - \epsilon_3 \rangle\). What's perhaps more
 986 surprising is the fact that this condition is sufficient for the existence of
 987 such a \(M\). In other words, our next goal is establishing\dots
 988 
 989 \begin{definition}\index{weights!dominant weight}
 990   An element \(\lambda \in P\) is called \emph{dominant} if it lies in the cone
 991   \(\mathbb{N} \langle \epsilon_1, - \epsilon_3 \rangle\).
 992 \end{definition}
 993 
 994 \begin{theorem}\label{thm:sl3-existence-uniqueness}
 995   For each dominant \(\lambda \in P\), there exists precisely one
 996   finite-dimensional simple \(\mathfrak{sl}_3(K)\)-module \(M\) whose highest
 997   weight is \(\lambda\).
 998 \end{theorem}
 999 
1000 To proceed further we once again refer to the approach we employed in the case
1001 of \(\mathfrak{sl}_2(K)\): next we showed in
1002 Proposition~\ref{thm:basis-of-irr-rep} that any simple
1003 \(\mathfrak{sl}_2(K)\)-module is spanned by the images of its highest weight
1004 vector under \(f\). A more abstract way of putting it is to say that a simple
1005 \(\mathfrak{sl}_2(K)\)-module \(M\) of is spanned by the images of its highest
1006 weight vector under successive applications of the action of half of the root
1007 spaces of \(\mathfrak{sl}_2(K)\). The advantage of this alternative formulation
1008 is, of course, that the same holds for \(\mathfrak{sl}_3(K)\).
1009 Specifically\dots
1010 
1011 \begin{proposition}\label{thm:sl3-positive-roots-span-all-irr-rep}
1012   Given a simple \(\mathfrak{sl}_3(K)\)-module \(M\) and a highest weight
1013   vector \(m \in M\), \(M\) is spanned by the images of \(m\) under successive
1014   applications of \(E_{2 1}\), \(E_{3 1}\) and \(E_{3 2}\).
1015 \end{proposition}
1016 
1017 \begin{proof}
1018   Given the fact \(M\) is simple, it suffices to show that the subspace \(N\)
1019   spanned by successive applications of \(E_{2 1}\), \(E_{3 1}\) and \(E_{3
1020   2}\) to \(m\) is stable under the action of \(\mathfrak{sl}_3(K)\). In
1021   addition, since \([E_{2 1}, E_{3 1}] = [E_{3 1}, E_{3 2}] = 0\) and \([E_{2
1022   1}, E_{3 2}] = - E_{3 1}\), all successive product of \(E_{2 1}\), \(E_{3
1023   1}\) and \(E_{3 2}\) in \(\mathcal{U}(\mathfrak{sl}_3(K))\) can be written as
1024   \(E_{2 1}^a E_{3 1}^b E_{3 1}^c\) for some \(a\), \(b\) and \(c\), so that
1025   \(N\) is spanned by the elements \(E_{2 1}^a E_{3 1}^b E_{3 1}^c \cdot m\).
1026 
1027   Recall that \(E_{i j}\) maps \(M_\mu\) to \(M_{\mu + \epsilon_i - \epsilon_j}\).
1028   In particular, \(E_{2 1}^a E_{3 1}^b E_{3 1}^c \cdot m \in M_{\lambda - a
1029   (\epsilon_1 - \epsilon_2) - b (\epsilon_1 - \epsilon_3) - c (\epsilon_2 - \epsilon_3)}\).
1030   In other words,
1031   \[
1032     H E_{2 1}^a E_{3 1}^b E_{3 1}^c \cdot m
1033     = (\lambda - a (\epsilon_1 - \epsilon_2)
1034                - b (\epsilon_1 - \epsilon_3)
1035                - c (\epsilon_2 - \epsilon_3))(H)
1036       E_{2 1}^a E_{3 1}^b E_{3 1}^c \cdot m
1037       \in N
1038   \]
1039   for all \(H \in \mathfrak{h}\) and \(N\) is stable under the action of
1040   \(\mathfrak{h}\). On the other hand, \(N\) is clearly stable under the action
1041   of \(E_{2 1}\), \(E_{3 1}\) and \(E_{3 2}\). All it is left is to show \(N\)
1042   is stable under the action of \(E_{1 2}\), \(E_{1 3}\) and \(E_{2 3}\).
1043 
1044   We begin by analyzing the case of \(E_{1 2}\). We have
1045   \[
1046     \begin{split}
1047       E_{1 2} E_{2 1}^a E_{3 1}^b E_{3 2}^c \cdot m
1048       & = ([E_{1 2}, E_{2 1}] + E_{2 1} E_{1 2})
1049           E_{2 1}^{a - 1} E_{3 1}^b E_{3 2}^c \cdot m \\
1050       & = E_{2 1} ([E_{1 2}, E_{2 1}] + E_{2 1} E_{1 2})
1051           E_{2 1}^{a - 2} E_{3 1}^b E_{3 2}^c \cdot m \\
1052       & \phantom{=} \; +
1053           (\lambda - (a - 1) (\epsilon_1 - \epsilon_2)
1054                    - b (\epsilon_1 - \epsilon_3)
1055                    - c (\epsilon_2 - \epsilon_3)) ([E_{1 2}, E_{2 1}])
1056           E_{2 1}^{a - 1} E_{3 1}^b E_{3 2}^c \cdot m \\
1057       & = E_{2 1}^2 ([E_{1 2}, E_{2 1}] + E_{2 1} E_{1 2})
1058           E_{2 1}^{a - 3} E_{3 1}^b E_{3 2}^c \cdot m \\
1059       & \phantom{=} \; +
1060           (\lambda - (a - 1) (\epsilon_1 - \epsilon_2)
1061                    - b (\epsilon_1 - \epsilon_3)
1062                    - c (\epsilon_2 - \epsilon_3)) ([E_{1 2}, E_{2 1}])
1063           E_{2 1}^{a - 1} E_{3 1}^b E_{3 2}^c \cdot m \\
1064       & \phantom{=} \; +
1065         (\lambda - (a - 2) (\epsilon_1 - \epsilon_2)
1066                    - b (\epsilon_1 - \epsilon_3)
1067                    - c (\epsilon_2 - \epsilon_3)) ([E_{1 2}, E_{2 1}])
1068           E_{2 1}^{a - 2} E_{3 1}^b E_{3 2}^c \cdot m \\
1069       & \; \; \vdots \\
1070       & = E_{2 1}^a E_{1 2} E_{3 1}^b E_{3 2}^c \cdot m \\
1071       & \phantom{=} \; +
1072           (\lambda - (a - 1) (\epsilon_1 - \epsilon_2)
1073                    - b (\epsilon_1 - \epsilon_3)
1074                    - c (\epsilon_2 - \epsilon_3)) ([E_{1 2}, E_{2 1}])
1075           E_{2 1}^{a - 1} E_{3 1}^b E_{3 2}^c \cdot m \\
1076       & \phantom{=} \; +
1077         (\lambda - (a - 2) (\epsilon_1 - \epsilon_2)
1078                    - b (\epsilon_1 - \epsilon_3)
1079                    - c (\epsilon_2 - \epsilon_3)) ([E_{1 2}, E_{2 1}])
1080           E_{2 1}^{a - 2} E_{3 1}^b E_{3 2}^c \cdot m \\
1081       & \phantom{=} \; \; \, \vdots \\
1082       & \phantom{=} \; +
1083         (\lambda - (a - a) (\epsilon_1 - \epsilon_2)
1084                    - b (\epsilon_1 - \epsilon_3)
1085                    - c (\epsilon_2 - \epsilon_3)) ([E_{1 2}, E_{2 1}])
1086           E_{2 1}^{a - a} E_{3 1}^b E_{3 2}^c \cdot m \\
1087     \end{split}
1088   \]
1089 
1090   Since \((\lambda - (a - k) (\epsilon_1 - \epsilon_2) - b (\epsilon_1 - \epsilon_3) -
1091   c (\epsilon_2 - \epsilon_3)) ([E_{1 2}, E_{2 1}]) E_{2 1}^{a - k} E_{3 1}^b
1092   E_{3 2}^c \cdot m \in N\) for all \(k\), it suffices to show \(E_{2 1}^a E_{1
1093   2} E_{3 1}^b E_{3 2}^c \cdot m \in N\). But
1094   \[
1095     \begin{split}
1096       E_{1 2} E_{3 1}^b
1097       & = (E_{3 1} E_{1 2} - E_{3 2}) E_{3 1}^{b - 1} \\
1098       & = E_{3 1} E_{1 2} E_{3 1}^{b - 1}
1099         - E_{3 1} E_{3 2} E_{3 1}^{b - 1} \\
1100       & = E_{3 1} (E_{3 1} E_{1 2} - E_{3 2}) E_{3 1}^{b - 2}
1101         - E_{3 2} E_{3 1}^b \\
1102       & \; \; \vdots \\
1103       & = E_{3 1}^b  E_{1 2} - b E_{3 2} E_{3 1}^b \\
1104     \end{split},
1105   \]
1106   given \([E_{1 2}, E_{3 1}] = - E_{3 2}\) and \([E_{3 2}, E_{3 1}] = 0\).
1107   It then follows from the fact \(E_{1 2} \cdot m = 0\) that
1108   \[
1109     E_{2 1}^a E_{1 2} E_{3 1}^b E_{3 2}^c \cdot m
1110     = E_{2 1}^a E_{3 1}^b E_{3 2}^c E_{1 2} \cdot m
1111     - b E_{2 1}^a E_{3 1}^b E_{3 2}^{c + 1} \cdot m
1112     = - b E_{2 1}^a E_{3 1}^b E_{3 2}^{c + 1} \cdot m \in N,
1113   \]
1114   given that \(E_{1 2}\) and \(E_{3 2}\) commute. Hence \(E_{1 2} \cdot (E_{2
1115   1}^a E_{3 1}^b E_{3 2}^c \cdot m) \in N\). Similarly,
1116   \[
1117     E_{1 3} \cdot (E_{2 1}^a E_{3 1}^b E_{3 2}^c \cdot m),
1118     E_{2 3} \cdot (E_{2 1}^a E_{3 1}^b E_{3 2}^c \cdot m) \in N
1119   \]
1120 \end{proof}
1121 
1122 The same argument also goes to show\dots
1123 
1124 \begin{corollary}\label{thm:irr-component-of-high-vec}
1125   Given a finite-dimensional \(\mathfrak{sl}_3(K)\)-module \(M\) with highest
1126   weight \(\lambda\) and \(m \in M_\lambda\), the subspace spanned by
1127   successive applications of \(E_{2 1}\), \(E_{3 1}\) and \(E_{3 2}\) to \(m\)
1128   is a simple submodule whose highest weight is \(\lambda\).
1129 \end{corollary}
1130 
1131 This is very interesting to us since it implies that finding \emph{any}
1132 finite-dimensional module whose highest weight is \(\lambda\) is enough for
1133 establishing the ``existence'' part of
1134 Theorem~\ref{thm:sl3-existence-uniqueness}. Moreover, constructing such a
1135 module turns out to be quite simple.
1136 
1137 \begin{proof}[Proof of existence]
1138   Take \(\lambda = k \epsilon_1 - \ell \epsilon_3 \in P\) with \(k, \ell \ge 0\),
1139   so that \(\lambda\) is dominant. Consider the natural
1140   \(\mathfrak{sl}_3(K)\)-module \(K^3\). We claim that the highest weight of
1141   \(\operatorname{Sym}^k K^3 \otimes \operatorname{Sym}^\ell (K^3)^*\) is
1142   \(\lambda\).
1143 
1144   First of all, notice that the weight vector of \(K^3\) are the canonical
1145   basis elements \(e_1\), \(e_2\) and \(e_3\), whose corresponding weights are
1146   \(\epsilon_1\), \(\epsilon_2\) and \(\epsilon_3\) respectively. Hence the weight
1147   diagram of \(K^3\) is
1148   \begin{center}
1149     \begin{tikzpicture}[scale=2]
1150       \AutoSizeWeightLatticefalse
1151       \begin{rootSystem}{A}
1152         \weightLattice{2}
1153         \wt[black]{1}{0}
1154         \wt[black]{-1}{1}
1155         \wt[black]{0}{-1}
1156         \node[right] at \weight{1}{0}  {$\epsilon_1$};
1157         \node[above left] at \weight{-1}{1} {$\epsilon_2$};
1158         \node[below left] at \weight{0}{-1} {$\epsilon_3$};
1159       \end{rootSystem}
1160     \end{tikzpicture}
1161   \end{center}
1162   and \(\epsilon_1\) is the highest weight of \(K^3\).
1163 
1164   On the one hand, if \(\{f_1, f_2, f_3\}\) is the dual basis for \(\{e_1, e_2,
1165   e_3\}\) then \(H \cdot f_i = - \epsilon_i(H) f_i\) for each \(H \in
1166   \mathfrak{h}\), so that the weights of \((K^3)^*\) are precisely the
1167   opposites of the weights of \(K^3\). In other words,
1168   \begin{center}
1169     \begin{tikzpicture}[scale=2]
1170       \AutoSizeWeightLatticefalse
1171       \begin{rootSystem}{A}
1172         \weightLattice{2}
1173         \wt[black]{-1}{0}
1174         \wt[black]{1}{-1}
1175         \wt[black]{0}{1}
1176         \node[left]        at \weight{-1}{0} {$-\epsilon_1$};
1177         \node[below right] at \weight{1}{-1} {$-\epsilon_2$};
1178         \node[above right] at \weight{0}{1}  {$-\epsilon_3$};
1179       \end{rootSystem}
1180     \end{tikzpicture}
1181   \end{center}
1182   is the weight diagram of \((K^3)^*\) and \(\epsilon_3\) is the highest weight
1183   of \((K^3)^*\).
1184 
1185   On the other hand if we fix two \(\mathfrak{sl}_3(K)\)-modules \(N\) and
1186   \(L\), by computing
1187   \[
1188     \begin{split}
1189       H \cdot (n \otimes l)
1190       & = H \cdot n \otimes l + n \otimes H \cdot l \\
1191       & = \lambda(H) n \otimes l + n \otimes \mu(H) l \\
1192       & = (\lambda + \mu)(H) \, (n \otimes l)
1193     \end{split}
1194   \]
1195   for each \(H \in \mathfrak{h}\), \(n \in N_\lambda\) and \(l \in L_\mu\) we
1196   can see that the weights of \(N \otimes L\) are precisely the sums of the
1197   weights of \(N\) with the weights of \(L\).
1198 
1199   This implies that the highest weights of \(\operatorname{Sym}^k K^3\) and
1200   \(\operatorname{Sym}^\ell (K^3)^*\) are \(k \epsilon_1\) and \(- \ell
1201   \epsilon_3\) respectively -- with highest weight vectors \(e_1^k\) and
1202   \(f_3^\ell\). Furthermore, by the same token the highest weight of
1203   \(\operatorname{Sym}^k K^3 \otimes \operatorname{Sym}^\ell (K^3)^*\) must be
1204   \(\lambda = k e_1 - \ell e_3\) -- with highest weight vector \(e_1^k \otimes
1205   f_3^\ell\).
1206 \end{proof}
1207 
1208 The ``uniqueness'' part of Theorem~\ref{thm:sl3-existence-uniqueness} is even
1209 simpler than that.
1210 
1211 \begin{proof}[Proof of uniqueness]
1212   Let \(M\) and \(N\) be two simple \(\mathfrak{sl}_3(K)\)-modules with highest
1213   weight \(\lambda\). By Theorem~\ref{thm:sl3-irr-weights-class}, the weights
1214   of \(M\) are precisely the same as those of \(N\).
1215 
1216   Now by computing
1217   \[
1218     H \cdot (m + n)
1219     = H \cdot m + H \cdot n
1220     = \mu(H) m + \mu(H) n
1221     = \mu(H) (m + n)
1222   \]
1223   for each \(H \in \mathfrak{h}\), \(m \in M_\mu\) and \(n \in N_\mu\), we can
1224   see that the weights of \(M \oplus N\) are same as those of \(M\) and \(N\).
1225   Hence the highest weight of \(M \oplus N\) is \(\lambda\) -- with highest
1226   weight vectors given by the sum of highest weight vectors of \(M\) and \(N\).
1227 
1228   Fix some \(m \in M_\lambda\) and \(n \in N_\lambda\) and consider the
1229   submodule \(L = \mathcal{U}(\mathfrak{sl}_3(K)) \cdot m + n \subset M \oplus
1230   N\) generated by \(m + n\). Since \(m + n\) is a highest weight of \(M \oplus
1231   N\), it follows from corollary~\ref{thm:irr-component-of-high-vec} that \(L\)
1232   is simple. The projection maps \(\pi_1 : L \to M\), \(\pi_2 : L \to N\),
1233   being nonzero homomorphism between simple \(\mathfrak{sl}_3(K)\)-modules,
1234   must be isomorphism. Finally,
1235   \[
1236     M \cong L \cong N
1237   \]
1238 \end{proof}
1239 
1240 We have been very successful in our pursue for a classification of the simple
1241 modules of \(\mathfrak{sl}_2(K)\) and \(\mathfrak{sl}_3(K)\), but so far we
1242 have mostly postponed the discussion on the motivation behind our methods. In
1243 particular, we did not explain why we chose \(h\) and \(\mathfrak{h}\), and
1244 neither why we chose to look at their eigenvalues. Apart from the obvious fact
1245 we already knew it would work a priory, why did we do all that? In the
1246 following chapter we will attempt to answer this question by looking at what we
1247 did in the last chapter through more abstract lenses and studying the
1248 representations of an arbitrary finite-dimensional semisimple Lie algebra
1249 \(\mathfrak{g}\).