% File: vertex.tex % Section: CFT2 % Title: Vertex algebras % Last modified: 05.11.2003 % \documentclass[a4paper,12pt]{article} \usepackage{amsmath,amssymb} \textwidth 16cm \textheight 25cm \oddsidemargin 0cm \topmargin -1.5cm \pagestyle{empty} \begin{document} \begin{center} \large\textbf{VERTEX ALGEBRAS} \end{center} \vspace{.0cm} \begin{center} \large\textbf{Formal distributions and locality} \end{center} \vspace{.1cm} Vertex algebras encode mathematical content of 2-dimensional conformal quantum field theory (CFT${}_2$). Roughly speaking, chiral quantum fields depending on a single complex variable $z$ are replaced by operator-valued formal power series in an abstract variable $z$\,. We begin with a brief exposition of the corresponding formalism. Formal power series of the following type ($m,\,n\in\mathbb{Z}\,, \ a_n,\,a_{m,n}\in A$), \begin{equation} \label{distrib} a(z) = \sum_{n}\frac{a_n}{z^{n+1}}\,, \ \ \ \ a(z,\,w) = \sum_{m,\,n}\frac{a_{m,n}}{w^{m+1}\,z^{n+1}}\,, \end{equation} are called $A$-valued formal distributions in one, two, or more indeterminates $z,\,w,\,...$\,. We can freely multiply formal distributions by polynomials, but a product of two distributions in the same variable(s) is not generally defined. Also, there is a problem in treating negative powers of $(z-w)$ etc. as formal distributions. To do it unambiguously, we will use (when needed) the notation like \begin{equation} \label{expanzw} (z-w)_{w}^{n} \doteq \sum_{k\geqslant0}(-)^k\binom{n}{k}w^k z^{n-k}\,, \ \ (z-w)_{z}^{n} \doteq \sum_{k\geqslant0}(-)^{n+k}\binom{n}{k}w^{n-k}z^k \end{equation} which indicates, in positive powers of what variable ($w$ or $z$ in this case) the expansion is performed. So, e.g., $(z-w)_{w}^{n}$ is analogous to $(z-w)_{|z|>|w|}^{n}$ in terms of complex plane. Of crucial importance in the present formalism is the $\delta$-function: \begin{equation} \label{delta} \delta(z-w) = \delta(w-z) = \sum_{n}w^n z^{-n-1} = \ldots +\frac{z}{w^2}+\frac{1}{w}+\frac{1}{z}+\frac{w}{z^2}+\ldots\,. \end{equation} It exhibits standard-looking properties \begin{equation} \label{} f(z)\,\delta(z-w) = f(w)\,\delta(z-w)\,, \ \ (z-w)\,\delta(z-w) = 0\,, \ \ \text{Res}_z f(z)\,\delta(z-w) = f(w) \end{equation} where $\text{Res}_z$ stands for a coefficient of $z^{-1}$\,, being an analog of $(2\pi i)^{-1}\!\oint\!dz$ on a complex plane. Note $\text{Res}_z\!\circ\partial_z = 0$\,. Evidently, the definition (\ref{delta}) mimics the well-known formulas like $\delta(x-y) = \sum_n e^{in(x-y)} = \sum_n (e^{ix})^n (e^{iy})^{-n}$\,. Useful relations with the derivatives of the $\delta$-function are listed below: \begin{gather} \label{deriv} \frac{1}{n!}\,\partial_{w}^{n}\,\delta(z-w) = \frac{(-)^n}{n!}\,\partial_{z}^{n}\,\delta(z-w) = \sum_{m}\binom{m}{n}w^{m-n}z^{-m-1} \ \ \ \ \ \ \ (n\geqslant0) \\ (z-w)^m\,\partial_{w}^{n}\,\delta(z-w) = 0 \ \ \ \ \ \ (m>n\geqslant0) \\ (z-w)^m\frac{1}{n!}\,\partial_{w}^{n}\,\delta(z-w) = \frac{1}{(n\!-\!m)!}\,\partial_{w}^{n-m}\delta(z-w) \ \ \ \ \ \ (0\leqslant m\leqslant n) \label{dd} \\ \text{Res}_z\,f(z)\,\partial_{w}^{n}\,\delta(z-w) = \partial_{w}^{n}\,f(w)\,, \ \ \ \text{Res}_z\,\frac{(z-w)^m}{n!}\,\partial_{w}^{n}\,\delta(z-w) = \delta_{m,n} \ \ \ (m\,,n\geqslant0) \label{rd} \end{gather} The following formula may also be of some use: \begin{equation} \label{grenze} (z-w)_w^{-n}-(z-w)_z^{-n} = \frac{1}{(n-1)!}\,\partial_{w}^{n-1}\delta(z-w) \ \ \ \ \ \ (n>0)\,. \end{equation} Another principal notion here is locality. For the (most important) case of two variables the locality condition is \begin{equation} \label{local} (z-w)^n a(z,w)=0 \ \ \ \ \ \ \ (n\geqslant N\geqslant0)\,. \end{equation} It imposes severe restrictions on $a(z,w)$\,. Consider first the simplest case \begin{gather} (z-w)\,b(z,w)=0 \ \ \ \ \ \ \ \Rightarrow \ \ \ \ \ \ \ b_{m+1,n}=b_{m,n+1} \notag\\ \ \ \ \ \ \ \ \Rightarrow \ \ \ \ \ b(z,w) = c(w)\,\delta(z-w)\,, \ \ c(w) = \text{Res}_z b(z,w)\,. \label{loc1} \end{gather} If now only $(z-w)^2 a(z,w)=0$\,, we consecutively find that \begin{gather*} (z-w)\,a(z,w)=c_1(w)\,\delta(z-w) \equiv c_1(w)\,(z-w)\,\partial_w\delta(z-w)\,, \\ c_1(w) = \text{Res}_z (z-w)\,a(z,w)\,, \\ a(z,w) - c_1(w)\,\partial_w\delta(z-w) = c_0(w)\,\delta(z-w)\,, \\ c_0(w) = \text{Res}_z [a(z,w)-c_1(w)\,\partial_w\delta(z-w)] = \text{Res}_z a(z,w) \end{gather*} and, therefore, \begin{equation} \label{loc2} a(z,w)=c_0(w)\,\delta(z-w)+c_1(w)\,\partial_w\delta(z-w)\,. \end{equation} By induction, using (\ref{dd}) and (\ref{rd})\,, we show that $(z-w)^n a(z,w)=0$ entails \begin{equation} \label{locn} a(z,w)=\sum_{m=0}^{n-1}c_m(w)\frac{1}{m!}\partial_{w}^{m}\delta(z-w)\,, \ \ \ \ c_m(w)=\text{Res}_z (z-w)^m a(z,w)\,. \end{equation} It is easily seen from (\ref{loc1}) or (\ref{locn}) that a local formal distribution should contain infinitely many nonzero terms of positive and negative degree in both variables, otherwise it is identically zero. It is also clear that $a(w,z),\, \partial_z a(z,w), \,\partial_w a(z,w),\, f(z)\,a(z,w)$ and $f(w)\,a(z,w)$ are local if $a(z,w)$ is. One interesting example of a local distribution is provided by the product of two series with all unity coefficients: \begin{equation} \label{delta1} (\sum_n z^{-n-1})(\sum_m w^{-m-1}) = \delta(z-1)\,\delta(w-1) = \delta(z-w)\,\delta(w-1) \end{equation} (setting one argument of $\delta$-function to a nonzero number makes sense)\,. One can verify (directly, or by induction) the following analog of the Taylor formula: \begin{equation} \label{Taylor} [f(z) - \sum_{n=0}^{N}\,\frac{(z-w)^n}{n!}\,\partial_{w}^{n}\,f(w)\,] \,\partial_{w}^{N}\,\delta(z-w) = 0\,. \end{equation} It is precisely in this way we may consider an expression inside square brackets in (\ref{Taylor}) as something of the order $(z-w)^{N+1}$\,. The following formulas with binomial coefficients are widely used in the calculations performed (or implicitly meant) in the present text: \begin{gather} \binom{-n-1}{m} = (-)^m \binom{n+m}{m} \ \ \ \ \ (m\geqslant0)\,, \ \ \ \ \ \ \ \ \binom{n}{m}=0 \ \ \ \ \ (m>n\geqslant0)\,, \\ \sum_{k=m}^{n}(-)^k \binom{n}{k}\binom{k}{m}=(-)^n\delta_{m,n}\,, \ \ \ \ \sum_{k=0}^{n}(-)^k k^m\binom{n}{k}=(-)^n\,n!\,\delta_{m,n}\,, \ \ \ \ (n\geqslant m\geqslant0) \\ \sum_{k=0}^{n}(-)^k \binom{n}{k}\binom{p-k}{m}=\binom{p-n}{m-n}\,, \ \ \ \ \ \ \ \sum_{k=0}^{n}(-)^k \binom{n}{k}\binom{p+k}{m}=(-)^n\binom{p}{m-n} \end{gather} \newpage \begin{center} \large\textbf{Operator product expansion} \end{center} \vspace{.1cm} A single-variable formal distribution $a(z)$ is called a field on a vector space $V$ if $a_n$ are operators on $V$\,, and for any $v\in V$ a series $a(z)\,v$ contains only finitely many negative powers of $z$\,. In other words, $a_n v = 0$ for $n\geqslant N(v)$\,. Fields $a$ and $b$ are called (mutually) local if for some $N(a,b)\geqslant0$ \begin{equation} \label{locf} (z-w)^N [a(z),b(w)]=0 \ . \end{equation} In what follows, we assume all fields to be pairwise local. For each $n\in\mathbb{Z}$\,, an important notion of $n$-product of two fields $a$ and $b$ is introduced as follows: \begin{equation} \label{nprod} (a_n b)(w) = \sum_{m}\frac{(a_n b)_m}{w^{m+1}}\,, \ \ \ \ (a_n b)_m \doteq \sum_{k\geqslant0}(-)^k\binom{n}{k} (a_{n-k}b_{m+k} -(-)^n b_{m+n-k}a_k)\,. \end{equation} It is seen from (\ref{nprod}) that $(a_n b)(w)$ is also a local field, because for given $n$ and $v$ (or $c(z)$) we are able to find such $N$ that $(a_n b)_m v = 0$ (resp. $(a_n b)_m c = 0$) for $m\geqslant N$\,. Let us discuss the definition of the $n$-product in more details. For $n\geqslant0$ it is equal to \begin{equation} \label{n>0} (a_n b)(w)=\text{Res}_z (z-w)^n [a(z),b(w)]\,. \end{equation} Locality condition (\ref{locf}) implies $(a_n b)(w)=0 \ (n\geqslant N)$ and, due to (\ref{locn}), \begin{equation} \label{commf} [a(z),b(w)]=\sum_{n=0}^{N-1}(a_n b)(w)\frac{1}{n!} \partial_{w}^{n}\delta(z-w)\,. \end{equation} Therefore, a commutator of local fields and (a finite number of) their $n$-products with non-negative $n$ are closely related. This is also seen at the component level ($n\geqslant0$)\,: \begin{equation} \label{commm} (a_n b)_m = \sum_{k=0}^{n}(-)^k\binom{n}{k}[a_{n-k},b_{m+k}]\,, \ \ \ \ \ [a_p,\,b_q] = \sum_{k\geqslant0}\binom{p}{k}(a_k b)_{p+q-k} \end{equation} (the second sum is also finite due to locality)\,. A special case $n=-1$ corresponds to the normal (or normally ordered) product, \begin{equation} \label{normw} (a_{-1}b)_m = \sum_{k<0}a_k b_{m-k-1} + \sum_{k\geqslant0}b_{m-k-1}a_k \ \ \ \Rightarrow \ \ \ (a_{-1}b)(w) = \ :\!a(w)b(w)\!:\, \ \equiv \,\ :\!ab\!:\!(w)\,, \end{equation} where \begin{equation} \label{normzw} :\!a(z)b(w)\!:\, \doteq a_+(z)b(w)+b(w)a_-(z)\,, \ \ \ a_+(z) = \sum_{n<0}\frac{a_n}{z^{n+1}}\,, \ \ \ a_-(z) = \sum_{n\geqslant0}\frac{a_n}{z^{n+1}}\ . \end{equation} Usually $a_+(z)$ is said to contain creation and $a_-(z)$ annihilation operators. We can now verify that the following general formula for negative $n$ agrees with the definition (\ref{nprod})\,: \begin{equation} \label{n<0} (a_n b)(w) = \ \frac{1}{(-n-1)!}:\!\partial^{-n-1}a(w)\cdot b(w)\!: \ \ \ \ \ \ \ \ (n<0) \end{equation} There also exists a universal form of the $n$-product definition: \begin{equation} \label{resprod} (a_n b)(w)=\text{Res}_z\bigl(a(z)b(w)(z-w)^n_w-b(w)a(z)(z-w)^n_z\bigr)\,. \end{equation} Remarkably, $n$-products play the role of coefficient functions of the operator product expansion (OPE). In CFT${}_2$\,, the latter is formulated for the radially ordered products \begin{equation} \label{rad} \mathcal{R}\,a(z)\,b(w) \doteq \begin{cases} a(z)\,b(w) & \qquad |z|>|w| \\ b(w)\,a(z) & \qquad |w|>|z| \end{cases} \end{equation} Within the present formalism, OPE relations are expressed in terms of the definitions (\ref{expanzw})\,: \begin{equation} \label{ope} a(z)\,b(w) = \sum_n \frac{(a_n b)(w)}{(z-w)_{w}^{n+1}}\ , \ \ \ \ b(w)\,a(z) = \sum_n \frac{(a_n b)(w)}{(z-w)_{z}^{n+1}} \end{equation} or, equivalently, \begin{gather} \label{opew} a(z)\,b(w) = \sum_{n=0}^{N-1} \frac{(a_n b)(w)}{(z-w)_{w}^{n+1}} \ + :\!a(z)\,b(w)\!:\, \\ \label{opez} b(w)\,a(z) = \sum_{n=0}^{N-1} \frac{(a_n b)(w)}{(z-w)_{z}^{n+1}} \ + :\!a(z)\,b(w)\!:\, \\ \label{ope<0} :\!a(z)\,b(w)\!:\, = \sum_{n\geqslant0}(a_{-n-1}b)(w)\ (z-w)^n \end{gather} Eqs. (\ref{ope<0}) and (\ref{ope}) are to be read not literally, but in the sense of Taylor's expansion (\ref{Taylor}), its coefficients given by (\ref{n<0})\,. At the same time, (\ref{opew}) and (\ref{opez}) are derived from (\ref{grenze}) and \begin{gather} a(z)\,b(w) \ - :\!a(z)\,b(w)\!:\ =\, [a_-(z),b(w)]\,, \\ b(w)\,a(z) \ - :\!a(z)\,b(w)\!:\ =\, [b(w),a_+(z)] \end{gather} by extracting negative and non-negative powers, respectively, from the $z$-expansion of (\ref{commf})\,, and are identically true. We see that a singular part of OPE is given by $n$-products with non-negative $n$\,, or by the commutator of fields. It seems to be instructive to rewrite this derivation of OPE using the CFT${}_2$ language. Let $\mathcal{R}\,a(z)\,b(w) = \sum_n \varphi_n(w)(z-w)^{-n-1}$\,. Then \begin{multline} \label{derivope} \varphi_n(w) = \text{Res}_{z-w}\bigl((z-w)^n\,\mathcal{R}\,a(z)\,b(w)\bigr)\\ = \frac{1}{2\pi i}\oint_w dz (z-w)^n\,\mathcal{R}\,a(z)\,b(w) = \frac{1}{2\pi i}\left(\oint_{|z|>|w|}dz - \oint_{|z|<|w|}dz\right) (z-w)^n\,\mathcal{R}\,a(z)\,b(w) \\ = \frac{1}{2\pi i}\oint dz \left(a(z)\,b(w)\,(z-w)^n_{|z|>|w|}-b(w)\,a(z)\,(z-w)^n_{|z|<|w|}\right) \\ = \text{Res}_z \left(a(z)\,b(w)\,(z-w)^n_{|z|>|w|}-b(w)\,a(z)\,(z-w)^n_{|z|<|w|}\right) = (a_n b)(w)\,. \end{multline} It should be noted that from (\ref{nprod}) and (\ref{commm}) (i.e., from the definition of the $n$-product plus the locality condition), the following useful formula (Borcherds identity) can be deduced. Namely, for $l,m,n\in\mathbb{Z}$\,, and $a,b,c$ being fields, the following relations hold: \begin{equation} \label{Borch} \sum_{k\geqslant0}\binom{m}{k}(a_{l+k}b)_{m+n-k}c = \sum_{k\geqslant0}(-)^k\binom{l}{k} \bigl(a_{m+l-k}(b_{n+k}c) -(-)^l b_{n+l-k}(a_{m+k}c)\bigr)\,. \end{equation} In CFT${}_2$\,, it's common practice to construct composite operators as normal products of two (or more) fields. To find then the OPE relations with these composite fields, the so-called Wick formulas are used. In the present formalism the latter are obtained from (\ref{Borch})\,, or (\ref{commm})\,, and look like \begin{equation} \label{Wick} a_n(b_{-1}c) = b_{-1}(a_n c) + \sum_{k\geqslant0}\binom{n}{k}(a_k b)_{n-k-1}c\ . \end{equation} \newpage \begin{center} \large\textbf{Vertex algebras} \end{center} \vspace{.1cm} A vertex algebra consists of a vector space $V$ (whose elements are called `states'), a vacuum state $\Omega\in V$\,, and a `state-field correspondence' linear map $Y$ which associates a field $a(z)\equiv Y(a,z)=\sum_n a_n z^{-n-1}$ (vertex operator) to each $a\in V$ in such a way that $Y(a_n b,z)=(a_nb)(z)$ are given by (\ref{nprod}), all fields are mutually local ($a_nb$\,=\,0 for $n\geqslant N(a,b)$)\,, and $Y(\Omega,z)$ is equal to unity ($\Omega_n=\delta_{n,-1}\,\mathbf{1}$)\,. In addition, a shift operator $D$ is defined on $V$ by $(Da)(z) \doteq \partial_z a(z)$\,. Note that an expression $a_nb$ is used here in two meanings: it denotes a state (an element in $V$ obtained by the action of the operator $a_n$ upon $b\in V$), and serves simultaneously as a name of the field $(a_nb)(z)$ (associated to this state via the map $Y$)\,, which is exactly the $n$-product of two fields $a(z)$ and $b(z)$\,. Now let us study the consequences of the above definitions. Due to locality of vertex operators, we may use not only (\ref{nprod})\,, but also Borcherds' relation (\ref{Borch})\,. Evidently, \begin{equation} \label{vac} a_{-1}\,\Omega=a\,, \ \ \ \ \ \ a_n\,\Omega=0 \ \ \ (n\geqslant0)\,, \ \ \ \ \ \ \ D\,\Omega=0\,, \ \ \ \ \ \ (Da)_n=-na_{n-1}\,. \end{equation} Further, for $k\geqslant0$\,, \begin{equation} \label{exp} (D^k a)_n = (-)^k\,k!\,\binom{n}{k}\,a_{n-k}\,, \ \ \ \ a_{-k-1} = \frac{1}{k!}\,(D^k a)_{-1} \ \ \ \ \ \ \Rightarrow \ \ \ \ Y(a,z)\,\Omega = e^{zD}a\,. \end{equation} Using (\ref{Borch}) with $m=0, \,n=-2, \, c=\Omega$\,, we obtain $D(a_lb)=(Da)_lb+a_l(Db)\,,$ whence \begin{equation} \label{Leib} [D,\,a_n]=(Da)_n=-na_{n-1} \ \ \ \ \ \ \Rightarrow \ \ \ \ [D,\,Y(a,z)]=Y(Da,z)=\partial_z Y(a,z)\,. \end{equation} As an immediate generalization we find \begin{equation} \label{transl} e^{zD}Y(a,w)e^{-zD} = Y(e^{zD}\!a,w) = Y(a,w+z)_z\ . \end{equation} For $m=-1, \,n=0, \, c=\Omega$ \ formula (\ref{Borch}) yields a so-called skew symmetry property \begin{equation} \label{skew} b_la = \sum_{k\geqslant0}\,\frac{(-)^{l+k+1}}{k!}\,D^k(a_{l+k}b) \ \ \ \ \ \ \Rightarrow \ \ \ \ Y(b,z)\,a = e^{zD}\,Y(a,-z)\,b\,. \end{equation} The OPE relation assumes here the following form: \begin{equation} \label{opey} Y(a,z)Y(b,w) = Y(Y(a,z-w)_w\,b,w)\,, \ \ \ \ Y(b,w)Y(a,z) = Y(Y(a,z-w)_z b,w)\,. \end{equation} If $V={\scriptstyle\bigoplus_{k\geqslant0}}V_k$ is a graded space equipped with a vertex algebra structure (so that $\Omega\in V_0\,, \ D:V_k\rightarrow V_{k+1}$\,, \,all $a_n$ homogeneous), one usually renumbers the components of the field $a(z)$ with $a\in V_\Delta$ in such a way (shifting index $n\rightarrow n-\Delta+1$) that \begin{equation} \label{grad} a(z)=\sum_n\,a_n\,z^{-n-\Delta}, \ \ \ a_{-n}:V_k\rightarrow V_{k+n}\,, \ \ \ a_{-\Delta}\Omega= a\,, \ \ \ a_n\Omega=0 \ \ (n>-\Delta)\,. \end{equation} As the famous example, the Virasoro algebra with central charge $c$ can be reconstructed in this way from a single `conformal' vector $\omega\in\!V_2$\ : \begin{gather} \label{Vir} \omega(z)\,\equiv\, Y(\omega,z)=\sum_n\,\omega_n\,z^{-n-1} \ \equiv \,T(z) =\sum_n\,L_n\,z^{-n-2} \ \ \ \ \ (\omega_n=L_{n-1})\ , \\ \label{Vir1} L_{-2}\,\Omega=\omega\,, \ \ \ L_{-1}=D\,, \ \ \ L_0\,a = \Delta\,a \ \ \,(a\in V_{\Delta})\,, \ \ \ L_1\,\omega=0\,, \ \ \ L_2\,\omega=\frac{c}{2}\ \Omega\,. \end{gather} These definitions prove to be consistent, agree with (\ref{Borch})\,, and ultimately result in \begin{gather} [L_m,L_n] = (m-n)L_{m+n}+\frac{c}{12}\,m(m^2-1)\,\delta_{n,-m} \ , \\ \mathcal{R}\,T(z)\,T(w) = \frac{c}{2}\,(z-w)^{-4}+2\,(z-w)^{-2}\,T(w) +(z-w)^{-1}\,\partial\, T(w)+\,\ldots\ . \end{gather} \newpage \begin{center} \large\textbf{Free fermions} \end{center} \vspace{.1cm} Below we describe the (Sugawara-like) construction of a conformal vector as a bilinear combination of `more elementary' objects: in this (simplest) case, neutral fermions. The corresponding vertex algebra is generated by the vacuum state $\Omega$ and one more vector $\phi$ subject to \begin{equation} \label{neutral} \phi_0\phi = \Omega\,, \ \ \ \ \ \ \phi_n\phi = 0 \ \ \ \ \ (n>0)\ . \end{equation} This exactly corresponds to \begin{equation} \label{phiz} \phi(z) = \sum_n\,\phi_n\,z^{-n-1}\,, \ \ \ \ \ [\phi_m,\phi_n]_+ = \delta_{m+n,-1}\ , \ \ \ \ \ \phi(z)\,\phi(w) \sim \frac{1}{(z-w)_w}\ . \end{equation} Since $\phi(z)$ is a fermionic field, we deal with anticommutators instead of commutators, and all formulas like (\ref{nprod}), (\ref{commm}), and so on, should be modified accordingly. Let us now introduce a vector \begin{equation} \label{omff} \omega \doteq \frac{1}{2}\,\phi_{-2}\phi \equiv \frac{1}{2}\,(D\phi)_{-1}\phi \equiv \frac{1}{2}\,\phi_{-2}\phi_{-1}\Omega \end{equation} which is to be recognized as a conformal one. Eq.\;(\ref{Wick}) yields \begin{equation} \label{} \phi_0\omega = -\frac{1}{2}\,D\phi\,, \ \ \ \ \ \phi_1\omega = \frac{1}{2}\,\phi\,, \ \ \ \ \ \phi_n\omega = 0 \ \ \ \ \ (n>1) \end{equation} and (\ref{skew}) then gives \begin{equation} \label{omphi} \omega_0\phi = D\phi\,, \ \ \ \ \ \omega_1\phi = \frac{1}{2}\,\phi\,, \ \ \ \ \ \omega_n\phi = 0 \ \ \ \ \ (n>1) \end{equation} Analogously, we come to \begin{equation} \label{} \omega_0 D\phi = D^2\phi\,, \ \ \ \ \ \omega_1 D\phi = \frac{3}{2}\,D\phi\,, \ \ \ \ \ \omega_2 D\phi = \phi\,, \end{equation} and, finally, to \begin{equation} \label{omom} \omega_0\omega = D\omega\,, \ \ \ \ \ \omega_1\omega = 2\omega\,, \ \ \ \ \ \omega_3\omega = \frac{1}{4}\,\Omega \end{equation} (only nonzero terms are displayed). So $\omega$ (\ref{omff}) is a genuine conformal vector, with central charge $c=1/2$\,, and $\phi$ a \textit{primary} element of conformal weight $\Delta=1/2$ (that means literally (\ref{omphi}))\,, with the OPE as follows: \begin{gather} \label{OPEomf} \omega(z)\,\phi(w) = \frac{\phi(w)}{2(z-w)^2_w} + \frac{\partial\phi(w)}{(z-w)_w} + \ldots\ , \\ \label{OPEomom} \omega(z)\,\omega(w) = \frac{1}{4(z-w)^4_w} + \frac{2\omega(w)}{(z-w)^2_w} + \frac{\partial\omega(w)}{(z-w)_w}\,+\,\ldots\ . \end{gather} Note that in this case, unlike (\ref{grad}) and (\ref{Vir}), we prefer not to shift indices of the field components in (\ref{phiz}). As for the vertex algebra as a whole, it is spanned, according to the Pauli principle, by the (finite) monomials of the form \begin{equation} \label{span} \ldots \phi_{-m}^{i_m} \ldots \phi_{-1}^{i_1}\Omega\ \ \ \ \ \ (m>0, \ \ i_m = 0,1)\ . \end{equation} The second equality in (\ref{commm}) gives \begin{equation} \label{} [\omega_0,\phi_n] = (D\phi)_n = -n\phi_{n-1}\,, \ \ \ \ \ [\omega_1,\phi_n] = (-\frac{1}{2}-n)\phi_n\,, \end{equation} and one readily verifies that \begin{equation} \label{omgen} \omega_0 a = Da\,, \ \ \ \ \ \ \ \omega_1 a = \Delta a \end{equation} for any $a$ of the type (\ref{span})\,, in accordance with the definition (\ref{Vir1}) of conformal vector. \newpage Let us now proceed with a slightly more involved example related to charged fermions. Here the vertex algebra is generated by the vacuum state $\Omega$ and two fermionic (odd) states $\psi^+\ (\equiv \psi)$ and $\psi^-\ (\equiv \bar\psi)$ subject to \begin{equation} \label{charged} \psi^+_0\psi^- = \psi^-_0\psi^+ = \Omega\,, \ \ \ \ \ \ \ \psi^+_0\psi^+ = \psi^-_0\psi^- = \psi^{\pm}_n\psi^{\pm} = \psi^{\pm}_n\psi^{\mp} = 0 \ \ \ \ (n>0) \end{equation} or, in other words, \begin{gather} \label{psiz} \psi^{\pm}(z) = \sum_n\,\psi^{\pm}_n\,z^{-n-1}\,, \ \ \ \ \ [\psi^{\pm}_m,\psi^{\pm}_n]_+ = 0\,, \ \ \ \ \ [\psi^{\pm}_m,\psi^{\mp}_n]_+ = \delta_{m+n,-1}\ , \\ \label{psiOPE} \psi^{\pm}(z)\,\psi^{\pm}(w) \sim 0\,, \ \ \ \ \ \ \psi^+(z)\,\psi^-(w) \sim \psi^-(z)\,\psi^+(w) \sim \frac{1}{(z-w)_w}\ . \end{gather} Now we can construct a bilinear (bosonic) current \begin{equation} \label{J} J = \psi^+_{-1}\psi^- = \psi^+_{-1}\psi^-_{-1}\Omega \end{equation} and, for arbitrary complex number $\lambda$\,, a conformal vector \begin{equation} \label{omlam} \omega_\lambda = \frac{1}{2}J_{-1}J + (\frac{1}{2}-\lambda)DJ = (1-\lambda)\,\psi^+_{-2}\psi^- + \lambda\,\psi^-_{-2}\psi^+ = [(1-\lambda)\,\psi^+_{-2}\psi^-_{-1} + \lambda\,\psi^-_{-2}\psi^+_{-1}]\Omega\,. \end{equation} We will see that $\psi^+$ and $\psi^-$ are primary states of conformal weights $\lambda$ and $1-\lambda$\,, respectively, whereas $J$ has $\Delta=1$ and becomes primary only for $\lambda=1/2$ (when $c=1$). First, we check that two definitions of $\omega_\lambda$ (in terms of $J$ and $\psi^{\pm}$) do agree. Using (\ref{nprod}), (\ref{psiz}), and properties of $D$\,, we find $$ J_{-1}J = (\psi^+_{-1}\psi^-)_{-1}J = \sum_{k\geqslant 0}(\psi^+_{-1-k}\psi^-_{-1+k} - \psi^-_{-2-k}\psi^+_k) \,\psi^+_{-1}\psi^-_{-1}\Omega = (\psi^+_{-2}\psi^-_{-1} + \psi^-_{-2}\psi^+_{-1})\,\Omega\,, $$ $$ DJ = (D\psi^+)_{-1}\psi^- + \psi^+_{-1}D\psi^- = (\psi^+_{-2}\psi^-_{-1} + \psi^+_{-1}\psi^-_{-2})\,\Omega = (\psi^+_{-2}\psi^-_{-1} - \psi^-_{-2}\psi^+_{-1})\,\Omega\,, $$ which verifies (\ref{omlam})\,. Now we are ready to obtain, quite straightforwardly, the $n$-products (with $n\geqslant0$), (anti)commutators, and singular parts of the OPE for $\psi^{\pm}, J$ and $\omega_\lambda$ in all the relevant combinations (as usual, only nonzero contributions are listed)\,: \begin{gather} \label{Jnpsi} \psi_0^{\pm}J = \mp \psi^{\pm}\,, \ \ \ \ \ J_0\psi^{\pm} = \pm \psi^{\pm}\,, \ \ \ \ \ [J_m,\psi_n^{\pm}] = \pm\psi^{\pm}_{m+n}\,, \\ \label{Jzpsiw} \psi^{\pm}(z)J(w) \sim \mp\frac{\psi^{\pm}(w)}{(z-w)_w}\,, \ \ \ \ \ J(z)\,\psi^{\pm}(w) \sim \pm\frac{\psi^{\pm}(w)}{(z-w)_w}\,, \\ \label{JJ} J_1 J = \Omega\,, \ \ \ \ [J_m,J_n] = m\delta_{m,-n}\,, \ \ \ \ J(z)J(w) \sim \frac{1}{(z-w)^2_w}\,, \\ \label{ompsi} \omega_\lambda{}_0 \psi^{\pm} = D\psi^{\pm}\,, \ \ \ \ \ \omega_\lambda{}_1 \psi^+ = \lambda\psi^+\,, \ \ \ \ \ \omega_\lambda{}_1 \psi^- = (1-\lambda)\,\psi^-\,, \\ \label{ompsiOPE} \omega_\lambda(z)\,\psi^{\pm}(w) \sim \frac{[1\pm(2\lambda-1)]\psi^{\pm}(w)}{2(z-w)^2_w} + \frac{\partial\psi^{\pm}(w)}{(z-w)_w}\,, \\ \label{Jom} J_1\omega_\lambda = J\,, \ \ \ \ J_2\omega_\lambda = (1-2\lambda)\,\Omega\,, \ \ \ \ J(z)\,\omega_\lambda(w) \sim \frac{1-2\lambda}{(z-w)^3_w} + \frac{J(w)}{(z-w)^2_w}\,, \\ \label{omJ} \omega_\lambda{}_0 J = DJ\,, \ \ \ \ \ \omega_\lambda{}_1 J = J\,, \ \ \ \ \ \omega_\lambda{}_2 J = (2\lambda-1)\,\Omega\,, \\ \label{omJOPE} \omega_\lambda(z)J(w) \sim \frac{2\lambda-1}{(z-w)^3_w} + \frac{J(w)}{(z-w)^2_w} + \frac{\partial J(w)}{(z-w)_w}\,, \\ \label{omlaomla} \omega_\lambda{}_0\omega_\lambda = D\omega_\lambda\,, \ \ \ \ \ \omega_\lambda{}_1\omega_\lambda = 2\omega_\lambda\,, \ \ \ \ \ \omega_\lambda{}_3\omega_\lambda = (-6\lambda^2+6\lambda-1)\,\Omega\,, \\ \label{omomOPE} \omega_\lambda(z)\,\omega_\lambda(w) \sim \frac{-6\lambda^2+6\lambda-1}{(z-w)^4_w} + \frac{2\omega_\lambda(w)}{(z-w)^2_w} + \frac{\partial\omega_\lambda(w)}{(z-w)_w}\,, \end{gather} the central charge here being \,$c=-12\lambda^2+12\lambda-2 =1-3(2\lambda-1)^2$\,. \pagebreak[1] Let us now consider an important set of primary states $V^m$\,: \begin{equation} \label{mvac} V^m = \psi_{-m}^{+}\psi_{-m+1}^{+}\ldots\psi_{-1}^{+}\Omega\,, \ \ \ \ \ V^{-m} = \psi_{-m}^{-}\psi_{-m+1}^{-}\ldots\psi_{-1}^{-}\Omega \ \ \ \ \ \ (m\geqslant 0)\,. \end{equation} Of course, \,$V^0 = \Omega\,, \ V^{\pm 1}=\psi^{\pm}$\,. Using (\ref{psiz}) and (\ref{Jnpsi}), one easily checks that \begin{equation} \label{Jmvac} J_{n>0}V^m = 0\,, \ \ \ \ \ J_0 V^m = m V^m\,, \ \ \ \ \ m J_{-1}\!V^m = D V^m\,. \end{equation} The last equality is verified by induction (say, for $m>0$)\,: \begin{gather} \notag J_{-1}V^1 = J_{-1}\psi_{-1}^+\Omega = \psi_{-2}^+\Omega + \psi_{-1}^+ J = (D\psi^+)_{-1}\Omega + \psi_{-1}^+\psi_{-1}^+\psi^- = DV^1, \\ \notag \psi_{-m-1}^+ J_{-1}V^m = 0\,, \ \ \ \ J_{-1}V^{m+1} = \psi_{-m-2}^+ V^m = \frac{1}{m+1}[D,\psi_{-m-1}^+]V^m = \frac{D V^{m+1}}{m+1}\,. \end{gather} Now, from \begin{equation} \label{} \omega_{\lambda n} = \frac{1}{2}\,\left(\,\sum_{k<0} J_k J_{n-k-1} + \sum_{k\geqslant 0} J_{n-k-1}J_k\right) + (\lambda - \frac{1}{2})\,n J_{n-1} \end{equation} and, in particular, \begin{equation} \label{} \omega_{\lambda 0} = \sum_{k\geqslant 0} J_{-k-1}J_k\,, \ \ \ \ \ \omega_{\lambda 1} = \frac{1}{2}\,J_0 J_0 + (\lambda - \frac{1}{2})J_0 + \sum_{k>0} J_{-k}J_k\,, \end{equation} we see that \begin{equation} \label{primV} \omega_{\lambda 0}V^m = J_{-1}J_0 V^m = m J_{-1}V^m = DV^m\,, \ \ \ \ \ \omega_{\lambda 1}V^m = \left(\frac{m^2}{2} + m\lambda - \frac{m}{2}\right)V^m\,. \end{equation} These relations, together with $\omega_{\lambda n}V^m = 0$ for $n\!>\!1$, characterize $V^m$ as a primary state of conformal weight $(\frac{m^2}{2}+m\lambda-\frac{m}{2})$\,. There exists a remarkable representation for the field $V^m(z)$ in terms of $J_n$ and an additional invertible operator $Y$ (not to be confused with the state-field map itself!) defined by \begin{equation} \label{Y} Y\psi_n^+ = \psi_{n-1}^+Y\,, \ \ \ Y\psi_n^- = \psi_{n+1}^-Y\,, \ \ \ Y\Omega = V^1 \ \ \Longrightarrow \ \ \ YV^m = V^{m+1}, \ \ \ V^m = Y^m\Omega\,. \end{equation} From \begin{equation} \label{} J_{n\neq0} = \sum_k \psi_k^+ \psi_{n-k-1}^-\,, \ \ \ \ \ \ J_0 = \sum_{k\geqslant0} (\psi_{-k-1}^+\psi_k^- - \psi_{-k-1}^-\psi_k^+) \end{equation} one immediately concludes that \begin{equation} \label{YJ} J_n Y = Y J_n \ \ (n\neq0)\,, \ \ \ \ \ \ [J_0,Y] = Y \ \ \ \ \ \Longrightarrow \ \ \ \ \ [J_0,Y^m] = m Y^m\,. \end{equation} A lengthy argument based on the Schur lemma shows that \begin{equation} \label{Vz} V^m(z) = \ :\!e^{m\varphi(z)}\!: \ = Y^m \exp(-m\sum_{n<0}\frac{J_n}{nz^n}) \ z^{mJ_0}\, \exp(-m\sum_{n>0}\frac{J_n}{nz^n}) \end{equation} where \begin{equation} \label{phi} \varphi(z) = \ln Y + J_0\ln z - \sum_{n\neq0}\frac{J_n}{nz^n}\,, \ \ \ \ \ \ \varphi\,'(z) = J(z)\,. \end{equation} We treat $\ln Y$ as a creation operator in accordance with $[J_0,\ln Y]=1$\,. Differentiating $V^m(z)$ is in agreement with (\ref{Jmvac})\,: \ $\partial_z V^m(z)= \ m:\!J(z)V^m(z)\!: \ \equiv m(J_{-1}V^m)(z)$\,. Another identity stemming from (\ref{Vz})\,, $V^m(z)\,\Omega=e^{zD}V^m=\exp(-m\sum_{n<0}\frac{J_n}{nz^n})V^m$, is also valid due to the properties of Schur polynomials. \end{document}