% File: integr2.tex % Section: INT % Title: Integrable Hierarchies II % Last modified: 08.07.2002 % \documentclass[a4paper,12pt]{article} \usepackage{amsmath,amssymb} \textwidth 16cm \textheight 25cm \oddsidemargin 0cm \topmargin -1.5cm \pagestyle{empty} \begin{document} \begin{center} \large\textbf{INTEGRABLE HIERARCHIES \ II} \end{center} \vspace{.0cm} \begin{center} \large\textbf{Adler-Kostant-Symes scheme} \end{center} \vspace{.1cm} The Lax representation, which underlies classical integrability, can in many cases be naturally derived by the use of two other fundamental concepts: classical $r$-matrix and Kirillov bracket. When taken together, these two tools form the Adler-Kostant-Symes scheme. Let $\mathfrak{g}$ be a Lie algebra, $C^k_{ij}$ its structure constants, $\mathfrak{g}^*$ its dual. A linear operator $R\in\text{End}\mathfrak{g}$ \, defines a second Lie structure on $\mathfrak{g}$ if a new bracket \begin{equation} \label{R} [A,B]_R \doteq [RA,B] + [A,RB]\,, \ \ \ \ \ \ \tilde{C}^k_{ij} = R^m_i\,C^k_{mj} - R^m_j\,C^k_{mi} \end{equation} satisfies the Jacobi identity. This requirement is equivalent to \begin{equation} \label{GYBE} [A\,,\,[RB,RC] - R\,[B,C]_R\,] + \text{cycle}\,(A,B,C) = 0 \ \ \ \ \ \ \ \ \forall \,A,B,C\in\mathfrak{g}\,. \end{equation} Thus, $R$ plays here the role of a classical $r$-matrix. The following two subcases, called \emph{classical Yang-Baxter equation}, \begin{equation} \label{CYBE} [RA,RB] - R\,[A,B]_R = 0\,, \end{equation} and \emph{modified Yang-Baxter equation}, \begin{equation} \label{MYBE} [RA,RB] - R\,[A,B]_R \,\sim\, [A,B]\,, \end{equation} evidently obey the general condition (\ref{GYBE})\,. The very existence of such $R$ results in the Lax-type dynamical equations on $\mathfrak{g}^*$. Clearly, the two Lie structures on $\mathfrak{g}$ induce two versions of the Kirillov brackets on $\mathfrak{g}^*$. We are interested in the Hamiltonian dynamics on $\mathfrak{g}^*$ generated by the second bracket. To begin with, let us express $(\text{ad}_R)^*_A$ \, in terms of $\text{ad}^*$\,: \begin{multline} \label{adR} <\!(\text{ad}_R)^*_A\,\omega\,,B\!> \ \doteq \ <\!\omega\,,[A,B]_R\!> \ = \ <\!\omega\,,[RA,B] + [A,RB]\!> \ = \ <\!\text{ad}^*_{RA}\,\omega\,,B\!> \\ + <\!\text{ad}^*_A\,\omega\,,RB\!> \ = \ <\!(\text{ad}^*_{RA} + R\,{}^*\!\circ\text{ad}^*_A)\,\omega\,,B\!> \ \ \ \Longrightarrow \ \ \ (\text{ad}_R)^*_A = \text{ad}^*_{RA} + R\,{}^*\!\circ\text{ad}^*_A \end{multline} Our next observation is that $\text{Ad}^*$-\,invariant functions $H(\omega)$ on $\mathfrak{g}^*$ (they are defined by $H(\text{Ad}^*_g\,\omega)=H(\omega)$ and obey $\text{ad}^*_{H'(\omega)}\omega=0$) are central under the first Kirillov bracket, \begin{equation} \label{main1} \{H,F\}(\omega) \doteq \ <\!\omega\,,[H'(\omega),F'(\omega)]\!> \ = \ <\!\text{ad}^*_{H'(\omega)}\omega\,,F'\!> \ = 0\,, \end{equation} and are in involution under the second: \begin{equation} \label{main2} \{H_1,H_2\}_R(\omega) = \ <\!\omega\,,[H'_1,H'_2]_R\!> \ = \ <\!\text{ad}^*_{H'_1(\omega)}\omega\,,RH'_2\!> - <\!\text{ad}^*_{H'_2(\omega)}\omega\,,RH'_1\!> \ = 0\,. \end{equation} Recalling a general formula for Hamiltonian vector fields produced by the Kirillov bracket ($\bar{F}(\omega)=\text{ad}^*_{F'(\omega)}\omega$) and using (\ref{adR}), we obtain the (mutually commuting) dynamical flows \begin{equation} \label{dyn} \dot{\omega} \ \doteq \ (\text{ad}_R)^*_{H'(\omega)}\,\omega \ = \ \text{ad}^*_{RH'(\omega)}\,\omega \end{equation} associated with $\text{Ad}^*$-\,invariant functions on $\mathfrak{g}^*$. In the matrix group case, where both $\text{ad}$ and $\text{ad}^*$ are commutators, this is the Lax representation. With $\text{Ad}^*$-\,invariant $H(\omega)$\,, one can examine $H'(\omega)$ in more detail. Let $\omega=\text{Ad}^*_g\,\omega_0$\,: \begin{multline} \notag <\!\lambda\,,\,H'(\omega)\!> \ \doteq \ <\!dH\,,\,\lambda\!>_\omega \ = (\lambda H)_\omega = \frac{d}{dt}H(\omega + t\lambda)|_{t=0} = \frac{d}{dt}H(\text{Ad}^*_g\,\omega_0 + t\lambda)|_{t=0} \\ = \frac{d}{dt}H(\omega_0 + t\text{Ad}^*_{g^{-1}}\lambda)|_{t=0} = \ <\!\text{Ad}^*_{g^{-1}}\lambda\,,\,H'(\omega_0)\!> \ = \ <\!\lambda\,,\,\text{Ad}_{g^{-1}}H'(\omega_0)\!>\,, \end{multline} and we conclude that \begin{equation} \label{H'} H'(\text{Ad}^*_g\,\omega) = \text{Ad}_{g^{-1}}H'(\omega)\,. \end{equation} In the matrix case, when \,$<\!\omega\,,A\!>\,\simeq\,\text{tr}(\omega A)\,, \ \text{Ad}_gA=gAg^{-1}, \ \text{Ad}^*_g\omega=g^{-1}\omega g$\,, one has \begin{equation} \label{H'2} H'(g^{-1}\omega\,g) = g^{-1}H'(\omega)\,g\,. \end{equation} Surprisingly many integrable systems belong to the above scheme within the following setting. Let $\mathfrak{g}=\mathfrak{g}_+ + \mathfrak{g}_-$ admit a decomposition (as a linear space, not as a Lie algebra) into a direct sum of two subalgebras, and $P_+,P_-$ be the corresponding projectors. Then \begin{equation} \label{RA} R = \frac{1}{2}\,(P_+ - P_-) \ \ \ \ \ \ \ \Longrightarrow \ \ \ \ \ \ [A,B]_R = [A_+,B_+] - [A_-,B_-] \end{equation} obeys the modified Yang-Baxter equation (\ref{MYBE})\,: \begin{multline} \label{} [RA,RB] - R\,[A,B]_R = \frac{1}{4}\,[A_+ - A_-\,,\,B_+ - B_-] - R\,([A_+,B_+] - [A_-,B_-]) \\ =\frac{1}{4}\,[A_+ - A_-\,,\,B_+ - B_-] - \frac{1}{2}\,([A_+,B_+]+[A_-,B_-]) = -\frac{1}{4}\,[A,B] \end{multline} In this case, dynamical equations (\ref{dyn}) read \begin{equation} \label{dyn1} \dot{\omega} = \text{ad}^*_{H'_+(\omega)}\,\omega = -\text{ad}^*_{H'_-(\omega)}\,\omega\ . \end{equation} The operator $R$ given by (\ref{RA}) takes on a more familiar form of the simplest rational $r$-matrix $r(z-w)\sim(z-w)^{-1}$ when dealing with formal distributions. Let \begin{equation} \label{distr} \xi(z) = \sum_n\xi_n z^n\,, \ \ \ \ P_+\xi(z) = \sum_{n\geqslant0}\xi_n z^n\,, \ \ \ \ P_-\xi(z) = \sum_{n<0}\xi_n z^n\,, \ \ \ \ \text{Res}_z\,\xi(z) = \xi_{-1} \end{equation} It is readily verified that \begin{equation} \label{} P_-\xi(z) = \text{Res}_w\,\frac{\xi(w)}{(z-w)_w}\,, \ \ \ \ \ P_+\xi(z) = -\text{Res}_w\,\frac{\xi(w)}{(z-w)_z}\,, \end{equation} and, of course, \begin{equation} \notag (P_+ + P_-)\,\xi(z) = \text{Res}_w\,(\frac{1}{(z-w)_w} - \frac{1}{(z-w)_z})\,\xi(w) = \text{Res}_w\,\delta(z-w)\,\xi(w) = \xi(z)\,. \end{equation} In this case, \begin{equation} \label{Yang} R\,\xi(z) = -\frac{1}{2}\,\text{Res}_w\,(\frac{1}{(z-w)_z} + \frac{1}{(z-w)_w})\,\xi(w) \ \simeq \ \text{Res}_w\,\frac{\xi(w)}{w-z} \end{equation} because in the corresponding Poisson brackets, the way of expanding the denominator usually does not matter due to the $\mathcal{O}(z-w)$ behavior of the numerator. \vspace{.3cm} \begin{center} \large\textbf{Matrix Ansatz} \end{center} \vspace{.1cm} Here we illustrate the general scheme described above in terms of the matrix Ansatz related to AKNS hierarchy. The phase space is formed by matrices $A_i$ which are allowed to depend on infinitely many variables $t_n$\,. Let \begin{equation} \label{Ansatz} A = \sigma + \sum_{i=1}^{\infty}z^{-i} A_i\,, \ \ \ \ \ \ \ \ B_n \doteq (z^n A)_+ \ \ \ \ \ \ \ \ \ (n = 1,2,\ldots) \end{equation} Here $\sigma$ is a constant matrix, and $z$ explicitly counts the grade. For instance, $B_1=z\sigma+A_1$\,, \ $B_2=z^2\sigma+zA_1+A_2$\,. Imposing the dynamical equations as follows, \begin{equation} \label{eqs} \partial_n A = [B_n,A] \ \ \ \ \ \ \ \text{or} \ \ \ \ \ \ \ [\partial_n - B_n\,,\,A] = 0 \ \ \ \ \ \ \ \ \ \ \ \ (\,\partial_n = \partial_{ t_n}) \end{equation} we can prove their zero-curvature property (commutativity of the corresponding flows) \begin{equation} \label{zero} [\partial_m - B_m\,,\,\partial_n - B_n] = \partial_n B_m - \partial_m B_n + [B_m,B_n] = 0 \end{equation} and thus show that (\ref{eqs}) is an integrable hierarchy: \begin{multline} \notag \partial_n(z^m A) - \partial_m(z^n A) = [B_n,z^m A] - [B_m,z^n A] = [(z^n A)_+,z^m A] + [(z^m A)_-,z^n A] \\ = [(z^n A)_+,(z^m A)_+] + [(z^m A)_-,(z^n A)_-] = [B_n,B_m] + [(z^m A)_-,(z^n A)_-]\,, \end{multline} and now taking the positive parts of both sides produces (\ref{zero})\,. In addition, combining (\ref{zero}) with (\ref{eqs}) one obtains \begin{equation} \label{} \partial_m B_n = [\partial_n - B_n\,,\,(z^m A)_+] = -[\partial_n - B_n\,,\,(z^m A)_-]\,. \end{equation} This implies that all flows respect grading: $B_n$ remains a polynomial in $z$ of degree $\leqslant n$\,. To connect this with the general geometric scheme of the preceding section, we use \begin{equation} \label{id} <\,,>\,=\text{Tr}(\ldots)\doteq\text{tr}(\ldots)_0\,, \ \ \ \ \text{Ad}_g\,\xi=g\,\xi g^{-1}\!, \ \ \ \text{Ad}^*_g\,\omega = g^{-1}\omega g\,, \ \ \ \text{ad}^*_\xi \omega = -[\xi,\omega] \end{equation} and introduce an infinite family of Hamiltonians \begin{equation} \label{Ham} H_n(A) = -\frac{1}{2}\text{Tr}(A\,z^n A) \ \ \ \ \ \ \ \Longrightarrow \ \ \ \ \ \ H'_n(A) = -z^n A\,, \ \ \ \ \ (H'_n)_+ = -B_n\,. \end{equation} They are invariant under $A\rightarrow g^{-1}\!Ag$\,, agree with (\ref{H'2})\,, and do generate (\ref{eqs}) from (\ref{dyn1})\,. In terms of $A_i$\,, the main system (\ref{eqs}) reads \begin{equation} \label{Ai} \partial_n A_i = [\sigma\,,A_{n+i}] \ + \sum_{j=1}^{\min(n,\,i-1)}[A_j\,,A_{n+i-j}]\,. \end{equation} Note that so far all $t_n$ (as well as the corresponding flows) are treated on equal footing, which makes the dynamics on our phase space fairly symmetric. However, it is a general practice to choose one `non-evolutional' variable (say, $t_1\equiv x$) leaving all other $t_{n>1}$ as `times'. In this case, the following equations from (\ref{Ai})\,, \begin{equation} \label{} \partial_1 A_i \equiv \partial_x A_i = [\sigma\,,A_{i+1}] + [A_1\,,A_i]\,, \end{equation} do not introduce any time evolution, but rather express $A_{i>1}$ (actually, certain components of it, depending on the properties of $\sigma$) in terms of $A_1$ and its $x$-derivatives. In fact, we modify the phase space by including $\partial^m_x A_1$ instead of $A_{i>1}$\,, without changing eqs.\,(\ref{Ai}) themselves. In what follows, we adopt the uniform view (with all $t_n$ treated equally) when possible, keeping, however, in mind a possibility (or necessity) of changing the phase space. \vspace{.3cm} \begin{center} \large\textbf{Dressing} \end{center} \vspace{.1cm} Now we explore the origin of commuting flows (\ref{zero}) from a slightly different point of view. Consider a family $\{b_n\} \ (n>0)$ of elements of a graded algebra (usually, $b_n$ span the positive part of a Heisenberg subalgebra of an affine Kac-Moody algebra) such that \begin{equation} \label{bn} [b_m,b_n] = 0\,, \ \ \ \ \ \ (b_n)_+ = b_n\,, \ \ \ \ \ \ (b_n)_- = 0\,, \ \ \ \ \ \ \partial_m b_n = 0\,. \end{equation} Then the operators $\partial_n-b_n$ commute for any $n$, as well as their `dressed' counterparts \begin{equation} \label{dress} \partial_n - B_n \doteq W(\partial_n-b_n)\circ W^{-1} \ \ \ \ \ \ \Longleftrightarrow \ \ \ \ \ \ B_n = W b_n W^{-1} + \partial_n W\cdot W^{-1}\,. \end{equation} Formally, zero-curvature condition (\ref{zero}) appears here as a result of dressing the family of trivially commuting flows. So far these relations are useless: they do not constrain $W$ at all. But our aim is to generalize the matrix Ansatz of the preceding section like this: $z^n\sigma\rightarrow b_n\,, \ A\rightarrow W\sigma W^{-1}$. It is achieved by requiring \begin{equation} \label{constr} W = \mathbf{1} + (\ldots)_-\ , \ \ \ \ \ \ \ \ B_n = (W b_n W^{-1})_+ \end{equation} that, together with (\ref{dress})\,, entails the Sato equations \begin{equation} \label{Sato} \partial_n W \cdot W^{-1} = -(W b_n W^{-1})_- \end{equation} which can be viewed as dynamical ones if a phase space is generated by $W$. To present a generic solution of (\ref{Sato})\,, let us introduce \begin{equation} \label{Gamma} \Gamma \doteq \exp\sum_i t_i\,b_i\,, \ \ \ \ \ \ \partial_n \Gamma = b_n \Gamma\,, \ \ \ \ \ \ \theta = \theta_-^{-1}\theta_+ \doteq \Gamma g \Gamma^{-1}\,, \ \ \ \ \ \ \partial_n g = 0 \end{equation} and show that $\theta_-$ obeys (\ref{Sato}) for any (constant) element $g$ of the Lie group corresponding to our original algebra. We suppose this group to be factorizable into a product of positive and negative part, $G=G_-^{-1}G_+$\,, in accordance with the $R$-bracket recipe (\ref{RA})\,. Indeed, \begin{gather*} \partial_n\theta = \partial_n(\theta_-^{-1}\theta_+) = \partial_n\theta_-^{-1}\cdot\theta_+ + \theta_-^{-1}\partial_n\theta_+ = -\theta_-^{-1}\partial_n\theta_-\cdot\theta_-^{-1}\theta_+ + \theta_-^{-1}\partial_n\theta_+\,, \\ \partial_n\theta = \partial_n(\Gamma g \Gamma^{-1}) = \partial_n \Gamma\cdot g\Gamma^{-1} - \Gamma g \Gamma^{-1}\partial_n\Gamma\cdot\Gamma^{-1} = b_n \theta - \theta b_n = b_n\theta_-^{-1}\theta_+ - \theta_-^{-1}\theta_+ b_n \\ \Longrightarrow \ \ \ \ \ \ \partial_n\theta_-\cdot\theta_-^{-1} + \theta_- b_n\theta_-^{-1} = \partial_n\theta_+\cdot\theta_+^{-1} + \theta_+ b_n\theta_+^{-1} \end{gather*} whence the Sato relation for $\theta_-$ follows by taking a negative part. Thus, evaluation of $(\Gamma g \Gamma^{-1})_-$ which requires solving the factorization problem in the group $G$ provides an efficient method for finding solutions $B_n$ of the integrable hierarchy (\ref{zero})\,, or, equivalently, solutions $\Psi = W\,\Gamma$ of the linear problem $(\partial_n - B_n)\Psi = 0$ (with $\Gamma$ being a `vacuum' solution)\,. Moreover, we could start from any solution $\Psi$\,, define $\phi=\Psi g\Psi^{-1}$ and show in complete analogy with the above calculation that $\Phi=\phi_-\Psi$ is also a solution: \begin{multline} \notag \partial_n\phi = \partial_n(\phi_-^{-1}\phi_+) = B_n\phi_-^{-1}\phi_+ - \phi_-^{-1}\phi_+ B_n \ \ \ \ \ \Longrightarrow \ \ \ \ \ \partial_n\phi_-\cdot\phi_-^{-1} = -(\phi_-B_n\phi_-^{-1})_- \\ \Longrightarrow \ \ \ \phi_-(\partial_n - B_n)\circ\phi_-^{-1} = \partial_n - (\phi_-B_n\phi_-^{-1})_+ \ \ \ \Longrightarrow \ \ \ (\partial_n - (\phi_-B_n\phi_-^{-1})_+)\Phi = 0\,. \end{multline} Using \ $\phi_-\,(\theta_-b_n\theta_-^{-1})_-\,\phi_-^{-1} =(\phi_-\,(\theta_-b_n\theta_-^{-1})_-\,\phi_-^{-1})_-$ \,we can derive as a simple corollary that in the case $\Psi=\theta_-\Gamma$ \ a product $\phi_-\theta_-$ satisfies (\ref{Sato})\,. This illustrates the action of the dressing group, which can be also related with the orbit interpretation of $\tau$ function. Assuming \,$G_+\,|0\!> \ \sim |0\!>$ we find, schematically, \begin{equation} \label{tau} |\tau_g(t)\!> \ = \,\Gamma(t)\,g\,|0\!> \ = \Gamma g\Gamma^{-1}\,|0\!> \ = \theta\,|0\!> \ \sim \ \theta_-^{-1}\,|0\!> \ = W^{-1}\,|0\!>\,. \end{equation} \vspace{.3cm} \begin{center} \large\textbf{Conservation laws} \end{center} \vspace{.1cm} Coadjoint-invariant Hamiltonians of the type (\ref{Ham}) do not directly apply within the dressing formalism because they are identically constant on the corresponding orbits. To develop an appropriate technique, let us adopt conventions (\ref{id}) and introduce a grading operator $\mathcal{D}$ generalizing $z\frac{d}{dz}$\,: \begin{equation} \label{D} \mathcal{D}(XY) = \mathcal{D}X\,Y + X\,\mathcal{D}Y, \ \ \ \ \ \mathcal{D}b_m = m\,b_m\,, \ \ \ \ \ \text{gr}(X) = m \ \ \ \Longrightarrow \ \ \ \mathcal{D}X = mX\,. \end{equation} Then the following holds: \begin{equation} \label{Jn} \mathcal{J}_n \doteq \text{Tr}(\mathcal{D}W\,b_nW^{-1}) \ \ \ \ \ \Longrightarrow \ \ \ \ \ \partial_m\mathcal{J}_n = \text{Tr}(\mathcal{D}B_m\,W b_n W^{-1}) = \partial_n\mathcal{J}_m\,. \end{equation} This is deduced with the use of (\ref{dress}) and $(b_m b_n)_0=(\mathcal{D}X)_0=(X_-Y_-)_0=(\mathcal{D}X_+\,Y_+)_0=0$\,: \begin{multline} \notag \partial_m\mathcal{J}_n = \partial_m \text{Tr}(\mathcal{D} W b_nW^{-1}) = \text{Tr}(\mathcal{D}(B_m W - W b_m)\,b_n W^{-1} - \mathcal{D}W b_n (W^{-1}B_m - b_m W^{-1})) \\ = \text{Tr}(\mathcal{D}B_m W b_n W^{-1} + B_m\mathcal{D}W b_n W^{-1} - \mathcal{D}W b_m b_n W^{-1} - m W b_m b_n W^{-1} - \mathcal{D}W b_n W^{-1} B_m \\ + \mathcal{D}W b_n b_m W^{-1}) = \text{Tr}(\mathcal{D}B_m W b_n W^{-1} - m b_m b_n) = \text{Tr}(\mathcal{D}B_m W b_n W^{-1}) \\ =\text{Tr}((\mathcal{D}(W b_m W^{-1})-\mathcal{D}(W b_m W^{-1})_-)W b_n W^{-1}) = \text{Tr}(\mathcal{D}W b_m b_n W^{-1} + m b_m b_n \\ - b_m W^{-1}\mathcal{D}W b_n - \mathcal{D}(W b_m W^{-1})_-(W b_n W^{-1})_+) = \text{Tr}(\mathcal{D}B_n W b_m W^{-1}) = \partial_n \mathcal{J}_m\,. \end{multline} It is now natural to introduce a generating $\tau$-function as \begin{equation} \label{tau2} \mathcal{J}_n \doteq \partial_n \log \tau \ \ \ \ \ \ \ \Longrightarrow \ \ \ \ \ \ \partial_m\mathcal{J}_n = \partial_n\mathcal{J}_m = \partial_m\partial_n \log \tau\,, \end{equation} which very likely is somehow connected with the definition (\ref{tau})\,. To proceed further, we single out $t_1=x$ and assume (omitting details) that our graded algebra admits a decomposition of the form $\mathfrak{b}\!\oplus \ldots$ \,with $\mathfrak{b}$ Abelian ($b_n\in\mathfrak{b}_+$)\,, and that an equation \begin{equation} \label{DS} L \equiv \partial_x - B_1 = U(\partial_x - b_1 - Q_1)\circ U^{-1} \equiv U L_0 U^{-1} \end{equation} is solved iteratively to produce $Q_1\in\mathfrak{b}_-$ and $U=\mathbf{1}+(\ldots)_-$ as functions of $B_1$ and its $x$-derivatives. In other words, $U$ and $Q_1$ are local. They are also unique if we demand that the $\mathfrak{b}$-part of $U$ be zero. Then we set $W=UT, \ \,T\in\exp(\mathfrak{b}_-)$\,, \,and find \begin{equation} \label{loc} Q_1 = \partial_x T\cdot T^{-1}, \ \ \ \ T b_n T^{-1} = b_n\,, \ \ \ \ B_n = (U b_n U^{-1})_+\,, \ \ \ \ \partial_m\mathcal{J}_n = \text{Tr}(\mathcal{D}B_m U b_n U^{-1})\,. \end{equation} Note that $T,W$ and $\mathcal{J}_n=\partial_n\log\tau$ are nonlocal whereas $\partial_m\mathcal{J}_n$ prove to be local. Now, defining a set of local quantities \begin{equation} \label{Hn} \mathcal{H}_n \doteq \partial_1\mathcal{J}_n = \text{Tr}(b_1 U b_n U^{-1}) \ \ \ \ \ \Longrightarrow \ \ \ \ \ \ \partial_m\mathcal{H}_n = \partial_x(\partial_m\mathcal{J}_n) \end{equation} ($\mathcal{D}B_1=b_1$ is used) we see that $H_n=\int\!dx\mathcal{H}_n$ are conserved with respect to all flows (\ref{dress})\,. The Hamiltonians of these flows are usually constructed as an alternative set of conserved quantities, generated by $h\equiv Q_1$\,. Namely, \begin{equation} \label{Qn} \partial_n - b_n - Q_n \doteq U^{-1}(\partial_n - B_n)\circ U = T(\partial_n - b_n)\circ T^{-1} \end{equation} implies that all $Q_n=\partial_n\log T\in\mathfrak{b}_-$\, are local and \begin{equation} \label{h} \partial_m Q_n = \partial_n Q_m\,, \ \ \ \ \ \partial_m h = \partial_x Q_m\,, \ \ \ \ \ \partial_m\int\!dx\,h = 0\,, \ \ \ \ \ \partial_m h_n = 0 \end{equation} where $h_n\doteq\int\!dx\,\text{Tr}(b_n h)$\,. Let us express $h'_n$ in terms of the Lax operator (\ref{DS})\,. From (\ref{H'2}) we know that $h'_n(L)=Uh'_n(L_0)U^{-1}$\,. To find $h'_n(L_0)$ we consider a variation \begin{equation} \label{} L_0 \ \rightarrow \ L_0+t\lambda \doteq U(t)\, (\partial_x-b_1-h(t))\circ U^{-1}(t) \end{equation} and do the following calculation: \begin{multline} \notag \int\!dx\,\text{Tr}(\lambda h'_n(L_0)) \equiv \ <\!\lambda\,,\,h'_n(L_0)\!> \ = \frac{d}{dt}\,h_n(L_0 + t\lambda)|_{t=0} \\ = \frac{d}{dt}\int\!dx \text{Tr}(b_n h(t))|_{t=0} = -\frac{d}{dt}\int\!dx \text{Tr}(b_n U^{-1}(t)(L_0 + t\lambda)U(t))|_{t=0} \\ = -\frac{d}{dt}\int\!dx \text{Tr}(b_n U^{-1}(t)L_0 U(t) + b_n(L_0 + t\lambda))|_{t=0} = -\int\!dx\,\text{Tr}(\lambda\,b_n)\,, \end{multline} because of \begin{equation} \label{} \frac{d}{dt}\text{tr}(b_n U^{-1}(t)L_0 U(t))|_{t=0} = \text{tr}(b_n[\ldots,L_0]) = \text{tr}([L_0,b_n]\ldots) = 0\,. \end{equation} We arrive at the desired result: \begin{equation} \label{h'n} h'_n(L_0) = -b_n\,, \ \ \ \ \ \ h'_n(L) = -U b_n U^{-1}\,, \ \ \ \ \ \ (h'_n)_+ = -B_n\,. \end{equation} The formalism with distinguished $\partial_x$ used above is naturally interpreted in terms of coadjoint group action underlying the whole scheme. Let us centrally extend, via a well-known 2-cocycle, our graded algebra and its dual as follows ($a,\alpha\in\mathbb{C}$)\,: \begin{align} A&\rightarrow(A,a)\,, & [(A,a),(B,b)] \ &\doteq \ ([A,B]\,,\int\!dx\,\text{Tr}(A\,\partial_x B)\,)\,, \\ \omega&\rightarrow(\omega,\alpha)\,, & <\!(\omega,\alpha),(A,a)\!> \ &\doteq \ \alpha a + \int\!dx\,\text{Tr}(\omega A)\,. \end{align} Starting from the very definition of \,$\text{ad}^*$, \begin{equation} \notag <\!\text{ad}^*_{(A,a)}(\omega,\alpha),(B,b)\!> \ \doteq \ <\!(\omega,\alpha)\,,\,[(A,a),(B,b)]\!> \ = \int\!dx\,\text{Tr}(\text{ad}^*_A\omega\,B - \alpha\partial_x A\,B)\,, \end{equation} we obtain \begin{gather} \label{adext} \text{ad}^*_{(A,a)}(\omega,\alpha) \ = \ (\text{ad}^*_A\omega - \alpha\partial_x A\,,\,0) \ = \ (-[A\,,\,\omega - \alpha\partial_x]\,,\,0)\,, \\ \label{ident} \text{Ad}^*_g\,(\omega\,,\alpha) \ = \ g^{-1}(\omega - \alpha\partial_x)g = g^{-1}\omega g - \alpha g^{-1}\partial_x g - \alpha\partial_x = (g^{-1}\omega g - \alpha g^{-1}\partial_x g\,,\alpha)\,, \end{gather} thus identifying $(\omega\,,\alpha) \ \simeq \ \omega - \alpha\partial_x$ and recognizing the Lax operator (\ref{DS}) as an element of centrally extended phase space, with $\alpha=1$ and $\omega=B_1$\,. \end{document}