information...this article is a more complete, improved version of the Preprint Nr.604/11/93 Bielefeld Univesity (BiBoS). The LaTeX file is compatible with: EmTex version 3.0 [3a], LaTeX version 2.09, TCI LaTeX. The actual LaTeX file (placed in the "BODY" case) has the checksum value = 4246 130 (obtained by the checksum command 'sum' from UNIX System V/386 Release 3.2) --------------- BODY \documentstyle{article} \author{C.P. Gr\"unfeld\thanks{Permantent adress: IGSS, Institute of Atomic Physics, Bucharest-Magurele, P. O. Box MG-6, RO-76900, Romania, E-mail grunfeld@ifa.ro}\\ Equipe de physique math\'ematique et g\'eom\'etrie, Institut de \\ Math\'ematique de Paris Jussieu, CNRS, Universit\'e Paris VII, \\ case 7012, couloir 45-55, 5-\`eme \'etage, 2 pl. Jussieu, \\ Paris, 75251, France \and E. Georgescu\\ IGSS, Institute of Atomic Physics, Bucharest-Magurele,\\ P. O. Box MG-6, RO-76900, Romania} \title{ON A CLASS OF KINETIC EQUATIONS FOR REACTING GAS MIXTURES } \date{ } \begin{document} \maketitle \begin{abstract} We consider a general class of kinetic equations for real gases with (possibly) multiple inelastic collisions and chemical reactions. We prove the existence, uniqueness and positivity of solutions for the Cauchy problem and obtain the conservation relations for mass, momentum and energy, the H-Theorem as well as the law of the mass action. \end{abstract} \section{Introduction} We investigate the mathematical properties of a class of Boltzmann-type kinetic equations for a model of reacting gas composed of several species of mass points with well-defined, unique internal energy state and multi-particle (in)elastic collisions (reactions). The number of species and the multiplicity of the collisions may be arbitrary. The gas particles move freely between collisions. The gas collisions occur with energy and momentum conservation according to the laws of the classical mechanics. For the one-component gas, with elastic binary collisions, the model kinetic equations can be reduced to the classical Boltzmann equation. Our interest in this model is due to the following thing. Certain kinetic equations for the real gas, which are important for applications, but less understood mathematically, appear to belong to our class of Boltzmann-type kinetic equations, as soon as they are written in convenient form. The main example refers to the Wang Chang and Uhlenbeck \cite{wa} as well as the Ludwig and Heil \cite{lu} equations, describing the real gas with inelastic collisions and chemical reactions, respectively. The fact that the equations introduced in Ref.\cite{wa}, \cite{lu} belong to the class examined in this paper is the consequence of the point of view, implicitly adopted in Ref.% \cite{wa}, \cite{lu} (see also \cite{ce}, \cite{ku}): in certain situations a real gas particle (molecule, atom, etc.) with internal structure can be considered as a mechanical system that differs from a mass point by a succession of internal states; each internal state has a well-defined value of the energy. It becomes convenient to treat different internal states of the gas particle with internal structure, as distinct, structureless point-objects, belonging to different species, of given mass and unique internal energy state, and described by different distribution functions. Consequently we can think of the gas of particles with internal structure as a gas mixture of different mass-points, with unique internal energy, and re-write the original kinetic equations in a suitable form by re-labeling the original distribution functions (each original distribution function, describing a succession of internal states of a particle with internal structure, is replaced by a sequence of distribution functions associated to each internal state). The aim of the present paper is to solve the Cauchy problem for the aforementioned class of reactive Boltzmann-type equations and to prove the basic global conservation relations, the H-Theorem, as well as the law of the mass action. The analysis reveals new mathematical difficulties, in comparison with the classical Boltzmann equation and other rigorous models $% \left[ 5-9\right] $. The difficulties are essentially due to the presence of the internal energy. They are introduced by the reaction thresholds, and are already visible in the case of the gas model with three-body collisions (reactions). The situations with more than three-body collisions (reactions ) do not introduce additional conceptual problems. However, the mathematical difficulties are better understood by investigating the general model than particular cases that might contain irrelevant details. The plan of the paper is as follows. In the next section we introduce the class of reactive Boltzmann-type kinetic equations. The main result, Theorem 1, obtained in Section 3, proves the existence, uniqueness and positivity of solutions (with small initial data) for the Cauchy problem associated to this class of equations. The solutions are global (in time) in the case in which the endo-energetic reactions are not present at the gas processes. In the case of the simple gas with elastic binary collisions, Theorem 1 reduces to known existence results on the classical Boltzmann equation \cite{be}. The argument of Theorem 1 follows by fixed point techniques, due to estimations based on the (local) conservation relations for mass, momentum and energy. The key estimation is given in Lemma 1. In Section 4 we prove the bulk conservation relations for mass, momentum and energy as well as the H-Theorem. Finally, the following fact should be remarked. The probability of multi-particle Collisions is zero, in some sense (\cite{ai}), in the dynamics of the classical hard sphere gas with elastic collisions (which plays an essential role in the validation of the classical Boltzmann equation). The situation seems being different in the case of the reacting gas: reaction processes {\it producing more that two} {\it particles} could be important to the gas evolution. {\it \ }Some of the results presented here, have been announced in \cite{gr}. \section{The frame} Consider a model of reacting gas without external fields, composed of $N$ $% \geq 1$, distinct species of mass-points, with one-state internal energy. Each species of gas constituents will be labeled by some simple index $% k=1,...,N$. The gas particles have a free classical motion, in the whole space, between (in)elastic, instant collisions. By hypothesis, at most, $% M\geq 2$ identical partners may participate in some in (out) collision (reaction) channel. During the gas processes, the particles may change their chemical nature (in particular, mass and internal energy) and velocity. It is supposed that the collisions occur with the conservation of mass, momentum and energy, respectively, according to the laws of classical mechanics. The particles internal energies enter in the energy balance. Let ${\cal M}:=\left\{ \gamma =(\gamma _n)_{1\le n\le N}\mid \gamma _n\in \left\{ 0,1,\ldots ,M\right\} \right\} $ be a multi-index set. A certain gas collision (reaction) process can be specified by a couple $(\alpha ,\beta )\in {\cal M\times M}$. Here $\alpha =(\alpha _1,\ldots ,\alpha _N)$ is the ''in'' channel. It designates the pre-collision configuration, with $\alpha _n\in \left\{ 0,1,\ldots ,M\right\} $ participants of the species $n$, $1\le n\le N$. Further, $\beta =(\beta _1,\ldots ,\beta _N)$ denotes the ''out'' channel. It refers to the post-collision configuration, with $\beta _n\in \left\{ 0,1,\ldots ,M\right\} $ participants of the species $n$, $1\le n\le N $. For some $\gamma \in {\cal M}$, the total numbers of the particles in channel $\gamma $ is $\mid \gamma \mid :=\sum_{n=1}^N\gamma _n$. The family of those species present in the channel $\gamma \in {\cal M}$ can be identified by ${\cal N}(\gamma ):=\left\{ n\mid 1\le n\le N,\;\gamma _n\ge 1\right\} $. Consequently, if $\gamma \in {\cal M}$, with $\mid \gamma \mid \geq 1$, for each $n\in {\cal N}(\gamma )$, there are exactly $\gamma _n$ identical particles of the species $n$, participating in $\gamma $. Their velocities will be denoted by ${\bf w}_{n,1},...,{\bf w}_{n,\gamma _n}\in {\bf R}^3$. Also set ${\bf w}=(({\bf w}_{n,i})_{1\leq i\leq {\gamma _n}% })_{n\in {\cal N}(\gamma )}$, understanding that ${\bf w\in R}^{3\mid \gamma \mid }$. By $m_n>0$ and $E_n\in {\bf R}$, denote the mass and the internal energy, respectively of a mass-point of the species $n=1,...,N$. Let $V_\gamma ({\bf w})$ and $W_\gamma ({\bf w})$ be the classical mass center velocity and the total energy, respectively, for the particles in channel $\gamma $, i.e.,% $$ V_\gamma ({\bf w}):=(\sum_{n=1}^N\gamma _nm_n)^{-1}\sum_{n\in {\cal N}% (\gamma )}\sum_{i=1}^{\gamma _n}m_n{\bf w}_{n,i}, $$ $$ W_\gamma ({\bf w}):=\sum_{n\in {\cal N}(\gamma )}\sum_{i=1}^{\gamma _n}(2^{-1}m_n{\bf w}_{n,i}^2+E_n). $$ According to the previous conservation assumptions we are interested in those gas processes $(\alpha ,\beta )\in {\cal M\times M}$, where \begin{equation} \label{15}\sum_{n=1}^Nm_n(\alpha _n-\beta _n)=0,\qquad V_\alpha ({\bf w}% )=V_\beta ({\bf u}),\qquad W_\alpha ({\bf w})=W_\beta ({\bf u}), \end{equation} with ${\bf w}=(({\bf w}_{n,i})_{1\leq i\leq {\alpha _n}})_{n\in {\cal N}% (\alpha )}$ and ${\bf u}=(({\bf u}_{n,i})_{1\leq i\leq {\beta _n}})_{n\in {\cal N}(\beta )}$ defining the velocities of the particles in the channels $% \alpha $ and $\beta $, respectively. Suppose that one knows the transition law (\cite{wa}, \cite{lu}) $K_{\alpha ,\beta }$ of each reaction process $(\alpha ,\beta )$. Following the standard Boltzmann procedure, we can formally write equations similar to those introduced in \cite{wa}, \cite{lu} \begin{equation} \label{1}\partial _tf_k+{\bf v}\cdot \nabla f_k=P_k(f)-S_k(f),\qquad 1\le k\le N. \end{equation} The unknowns are the functions $f_k:{\bf R}_{+}\times {\bf R}^3\times {\bf R}% ^3\rightarrow {\bf R}_{+}$,$\;1\le k\le N$, where ${\bf R}_{+}:=[0,\infty )$% . Here $f_k=f_k(t,{\bf v},{\bf x})$ ($t$-time, ${\bf v}$ -velocity, ${\bf x}$% -position) is the distribution function for species $k$ of mass-points and $% f:=(f_1,\ldots ,f_N)$. The collision processes are described by the nonlinear terms $P_k(f)$ and $S_k(f)$% $$ P_k(f)(t,{\bf v},{\bf x}):=\sum_{\alpha ,\beta \in {\cal M}}\ \alpha _k\ \int_{{\bf R}^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid }}\,f_\beta (t,{\bf u},{\bf x})\,K_{\beta ,\alpha }({\bf u},{\bf w})\ $$ \begin{equation} \label{2}\times \ \delta ({\bf w}_{k,\alpha _k}-{\bf v})\ \delta (V_\beta (% {\bf u})-V_\alpha ({\bf w}))\ \delta (W_\beta ({\bf u})-W_\alpha ({\bf w}))\ d{\bf u\otimes }d{\bf w}, \end{equation} \smallskip\ $$ S_k(f)(t,{\bf v},{\bf x}):=\sum_{\alpha ,\beta \in {\cal M}}\alpha _k\int_{% {\bf R}^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid }}\ \ f_\alpha (t,{\bf w},{\bf x})K_{\alpha ,\beta }({\bf w},{\bf u}) $$ \begin{equation} \label{2'}\times \ \delta ({\bf w}_{k,\alpha _k}-{\bf v})\ \delta (V_\beta (% {\bf u})-V_\alpha ({\bf w}))\ \delta (W_\beta ({\bf u})-W_\alpha ({\bf w}))\ d{\bf u\otimes }d{\bf w}, \end{equation} for all $t\ge 0$, ${\bf v}$, ${\bf x}\in {\bf R}^3$; $1\le k\le N$. Here, $% (K_{\alpha ,\beta })_{(\alpha ,\beta )\in {\cal M}\times {\cal M}}$ is the family of transition functions $K_{\alpha ,\beta }:{\bf R}^{3\mid \alpha \mid }\times {\bf R}^{3\mid \beta \mid }\rightarrow {\bf R}_{+}$, $\alpha $, $\beta \in {\cal M}$ and \ $$ f_\gamma (t,{\bf w},{\bf x})=\Pi _{n\in {\cal N}(\gamma )}\,\Pi _{i=1}^{\gamma _n}\,f_n(t,{\bf w}_{n,i},{\bf x}), $$ We introduce the following general assumptions: a) $K_{\alpha ,\beta }\equiv 0\,\,$if $\mid \alpha \mid \le 1\,$or $\mid \beta \mid \leq 1$. b) If for some $\alpha $, $\beta \in {\cal M}$,$\;\sum_{n=1}^N\alpha _nm_n% \not =\sum_{n=1}^N\beta _nm_n$, then $K_{\alpha ,\beta }\equiv 0$. c) For each ${\bf w}$, ${\bf u}$ and $n\in {\cal N}(\alpha )$ fixed, $% K_{\alpha ,\beta }({\bf w},{\bf u})$ is invariant at the interchange of components ${\bf w}_{n,1},...,{\bf w}_{n,\alpha _n}$ of ${\bf w}$; a similar statement is true with respect to the interchange of the components of ${\bf % u}$. d) For each $a\in {\bf R}^3$, define the map ${\bf w\rightarrow }T(a){\bf w}$ by setting $(T(a){\bf w})_{n,i}:={\bf w}_{n,i}+a$, for all $n,i$ ; then $% K_{\alpha ,\beta }({\bf w},{\bf u})\equiv K_{\alpha ,\beta }(T(a){\bf w},% {\bf u})\equiv K_{\alpha ,\beta }({\bf w},T(a){\bf u})$, for all $a\in {\bf R% }^3$, $({\bf w,u)}\in {\bf R}^{3\left| \alpha \right| }\times {\bf R}^{3\mid \beta \mid }$ and $\alpha $, $\beta \in {\cal M}.$ Assumption a) excludes the ''spontaneous decay'' $(\mid \alpha \mid \le 1)$ and the ''total fusion'' $(\mid \beta \mid \le 1)$. Condition b) states the mass conservation during the gas processes. Moreover, c) expresses the ''indistinguishability'' of identical collision partners. Finally, d) claims absence of external fields. The presence of the Dirac $\delta $-''functions'' in (\ref{2}) and (\ref{2'}% ) expresses the conservation of the total energy and momentum, respectively, during collisions. It can be easily seen that the kinetic equations introduced in \cite{wa}, \cite{lu} can be written in the form (\ref{1}), by redefining the distribution functions according to the remarks in the previous section. For some channel $\gamma \in {\cal M}$, let% $$ W_{r,\gamma }({\bf w}):=W_\gamma ({\bf w})-2^{-1}(\sum_{n=1}^N\gamma _nm_n)V_\gamma ({\bf w})^2-\sum_{n=1}^N\gamma _nE_n,\qquad {\bf w\in {R}}% ^{3\mid \gamma \mid }, $$ be the corresponding mass center kinetic energy. Obviously, $W_{r,\gamma }(% {\bf w})\ge 0$. We suppose that, $\forall \alpha $, $\beta \in {\cal M}$, the transition law $K_{\alpha ,\beta }$ is continuous on the set $\{({\bf w},{\bf u})\in {\bf R}% ^{3\mid \alpha \mid }\times {\bf R}^{3\mid \beta \mid }\mid W_{r,\alpha }(% {\bf w})>0,W_{r,\beta }({\bf u})>0\}$. We introduce the following hypothesis, extending a class of cut-off conditions for elastic binary collisions \cite{be}. {\bf ASSUMPTION.}{\it - There are some constants} $C>0$,$\;0\leq q\leq 1$% {\it , such that} {\it for all}$\;\alpha $, $\beta $, ${\bf w\in {R}}^{3\mid \alpha \mid }$,$\;{\bf u\in {R}}^{3\mid \beta \mid }$, {\it we have} \begin{equation} \label{3}K_{\alpha ,\beta }({\bf w},{\bf u})\le C\ \frac{1+W_{r,\alpha }(% {\bf w})^{q/2}+W_{r,\beta }({\bf u})^{q/2}}{W_{r,\alpha }({\bf w})^{(3\mid \alpha \mid -5)/2}+W_{r,\beta }({\bf u})^{(3\mid \beta \mid -5)/2}}. \end{equation} \ In the rest of this section we give a meaning to (\ref{2}), (\ref{2'}). Let $% C_c({\bf R}{^3}\times {\bf R}^3)$ denote the space of continuous functions with compact support on ${\bf R}{^3}\times {\bf R}^3$. For each $\tau \geq 0$% , $n=1,...,N$, fixed, let $\dot C_{n,\tau }$ be the closure of $C_c({\bf R}{% ^3}\times {\bf R}^3)$-real in the norm% $$ \left| h\right| _{n\tau }=\sup \left\{ \exp \left[ \tau m_n({\bf x}^2+{\bf v}% ^2)\right] \left| h({\bf v},{\bf x})\right| :{\bf v},{\bf x}\in {\bf R}% ^3\right\} ,\;h\in C_c({\bf R}{^3}\times {\bf R}^3). $$ Set $\dot C_\tau =\Pi $$_{1\le n\le N}\,\dot C_{n,\tau }$, with norm $\mid h\mid _\tau :=\max \limits_{1\le n\le N}\left| h_n\right| _{n,\tau }$, for $% h=(h_1,\ldots ,h_N)\in \dot C_\tau $. Let $\delta _\epsilon :{\bf R}% \rightarrow {\bf R}_{+}$, $\epsilon >0$, be an even mollifier with supp $% \delta _\epsilon =[-\epsilon ,\epsilon ]$ (i.e. $\delta _\epsilon (t)=:\epsilon ^{-1}J(t/\epsilon )$, for some even function $J\in C_c({\bf R};% {\bf R}_{+})$, with $supp$ $J=[-1,1]$ and $\left\| J\right\| _{L^1}=1$). Set $\delta _\epsilon ^3(y):=\delta _\epsilon (y_1)\cdot \,\delta _\epsilon (y_2)\cdot \,\delta _\epsilon (y_3)$, with $y=(y_1,y_2,y_3)\in {\bf R}^3$. For some $\tau >0$ and $f=(f_1,\ldots ,f_N)\in \dot C_\tau $, define% $$ P_{k\epsilon \eta }(f)({\bf v},{\bf x}):=\sum_{\alpha ,\beta \in {\cal M}% }\alpha _k\int_{{\bf R}^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid -3}}\ d{\bf u\otimes }d{\bf \tilde w}_{(k)}\ \left[ f_\beta ({\bf u},{\bf x}% )\right. $$ \begin{equation} \label{4}\,\left. \times \,K_{\beta ,\alpha }({\bf u},{\bf w})\,\delta _\epsilon ^3(V_\beta ({\bf u})-V_\alpha ({\bf w}))\,\delta _\eta (W_\beta (% {\bf u}))-W_\alpha ({\bf w}))\right] _{{\bf w}_{k,\alpha _k}={\bf v}}, \end{equation} \smallskip\ \ $$ R_{k\epsilon \eta }(f)({\bf v},{\bf x}):=\sum_{\alpha ,\beta \in {\cal M}% }\alpha _k\int_{{\bf R}^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid -3}}\ d{\bf u\otimes }d{\bf \tilde w}_{(k)}\,\ \left[ f_{\alpha ,k}({\bf w},% {\bf x})\right. $$ \begin{equation} \label{4'}\,\left. \times \,K_{\alpha ,\beta }({\bf w},{\bf u)\,}\delta _\epsilon ^3(V_\beta ({\bf u})-V_\alpha ({\bf w}))\,\delta _\eta (W_\beta (% {\bf u}))-W_\alpha ({\bf w}))\right] _{{\bf w}_{k,\alpha _k}={\bf v}}\;, \end{equation} for all ${\bf v}$, ${\bf x}\in {\bf R}^3$, $1\le k\le N$. Here by definition, the terms with $\alpha _k=0$, vanish identically, $d{\bf \tilde w% }_{(k)}$ is the Euclidean element of area induced by $d{\bf w}$ on the manifold $\left\{ {\bf w\in R^{3\mid \alpha \mid }:w}_{k,\alpha _k}={\bf v}% \right\} ${\it ,} while% $$ f_\beta ({\bf u},{\bf x}):=\Pi _{n\in {\cal N}(\beta )}\Pi _{i=1}^{\beta _n}f_n({\bf u}_{n,i},{\bf x}), $$ $$ f_{\alpha ,k}({\bf w},{\bf x})=\Pi _{n\in {\cal N}(\alpha )\setminus \left\{ k\right\} }\Pi _{i=1}^{\alpha _n}f_n({\bf w}_{n,i},{\bf x})\cdot \Pi _{p=1}^{\alpha _k-1}f_k({\bf w}_{k,p},{\bf x}). $$ {\bf PROPOSITION 1}. Let $\tau >0$ and $f\in \dot C_\tau $. {\it a) For each} $k=1,...,N$, {\it there exist the limits}% $$ P_k(f)({\bf v},{\bf x})=\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}P_{k\epsilon \eta }(f)({\bf v},{\bf x}% )\;and\;R_k(f)({\bf v},{\bf x})=\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}R_{k\epsilon \eta }(f)({\bf v},{\bf x}), $$ $\forall $$({\bf v},{\bf x})\in {\bf R}^3\times {\bf R}^3$. {\it Also,} $% P_k(f)\in C_{k,\mu }({\bf R}^3\times {\bf R}^3)$ {\it for all $\mu \in \left[ 0,\tau \right) $}, {\it while} $\sup \limits_{{\bf v},{\bf x}}\left\{ (1+{\bf v}^2)^{-q/2}\left| R_k(f)({\bf v},{\bf x})\right| \right\} <\infty $. {\it b) Let }$h\in C({\bf R}^3)$ {\it with} $\sup \limits_{{\bf v}}\left\{ (1+{\bf v}^2)^{-1}\left| h({\bf v})\right| \right\} <\infty ${\it . Then, }$% \forall ${\it \ }${\bf x}\in {\bf R}^3$, $$ \int_{{\bf R}^3}h({\bf v})\ P_k(f)({\bf v},{\bf x})\ d{\bf v}=\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}\int_{{\bf R% }{^3}}h({\bf v})\ P_{k\epsilon \eta }(f)({\bf v},{\bf x})\ d{\bf v}, $$ $$ \int_{{\bf R}^3}h({\bf v})\ f_k({\bf v},{\bf x})R_k(f)({\bf v},{\bf x})\ d% {\bf v}=\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}\int_{{\bf R}{^3}}h({\bf v})\,f_k({\bf v},{\bf x})R_{k\epsilon \eta }(f)(% {\bf v},{\bf x})\ d{\bf v}, $$ {\it for each} $k=1,...,N$ \ {\bf Proof.} Set $T_\beta ({\bf u})=W_\beta ({\bf u})-\sum_{n=1}^N\beta _nE_n $. We associate Jacobi coordinates $(\underline{{\sl V}},\xi )\in {\bf % R}^3\times {\bf R}^{3\mid \beta \mid -3}$ to the form $T_\beta ({\bf u})$ on ${\bf R}^{3\mid \beta \mid }$, with $\xi :=(\xi _1,\ldots ,\xi _{\mid \beta \mid -1})$, $\,\xi _i\in {\bf R}{^3}$,${\ \;}i=1,\ldots ,\mid \beta \mid -1$ $\,$(see (A.2) Appendix A). Consider a representation of $\xi $ in spherical coordinates on ${\bf R}^{3\mid \beta \mid -3}$,$\;\xi =r{\bf n}$ , with $(r,% {\bf n})\in \left[ 0,\infty \right) \times \Omega _{3\mid \beta \mid -4}$, where $\Omega _{3\mid \beta \mid -4}$ is the unit sphere in ${\bf R}^{3\mid \beta \mid -3}$. In (\ref{4}) and (\ref{4'}) we choose $(\underline{V},r,% {\bf n})$ as new integration variables such that ${\bf u}={\bf u}(\underline{% V},r,{\bf n})$. Then the limits of Prop.1 follow by repeated application of Lebesgue's dominated convergence theorem, using the properties of $K_{\alpha ,\beta }$, $\delta _\epsilon ^3$ and $\delta _\eta $. The continuity of $% P_k(f)$ and $R_k(f)$ is a consequence of the continuity of $K_{\alpha ,\beta }$.$\Box $\ The proof of Prop.1 provides the limits (\ref{4}), (\ref{4'}) in explicit form. Define% $$ t_{\beta ,\alpha }({\bf w})=\left\{ \begin{array}{l} \left[ W_{r,\alpha }( {\bf w})+\sum\limits_{n=1}^N(\alpha _n-\beta _n)E_n\right] ^{1/2}\,if\;W_{r,\alpha }({\bf w})+\sum\limits_{n=1}^N(\alpha _n-\beta _n)E_n\geq 0,\medskip \\ 0,\;\;otherwise. \end{array} \right. $$ If \begin{equation} \label{5}W_{r,\alpha }({\bf w})+\sum_{n=1}^N(\alpha _n-\beta _n)E_n\ge 0, \end{equation} then set \begin{equation} \label{6}{\bf u}_{\beta \alpha }({\bf w},{\bf n}):={\bf u}(\underline{V},r,% {\bf n})_{\mid \underline{V}=V_\alpha ({\bf w}),\;r=t_{\beta ,\alpha }({\bf w% })}. \end{equation} For the sake of simplicity, ${\bf u}_{\beta \alpha }$ will replace the notation ${\bf u}_{\beta \alpha }({\bf w},{\bf n})$. Define \begin{equation} \label{7}p_{\beta \alpha }({\bf w},{\bf n}):=2^{-1}\Delta _\beta \cdot t_{\beta ,\alpha }({\bf w})^{3\mid \beta \mid -5}K_{\beta ,\alpha }({\bf u}% _{\beta \alpha },{\bf w}), \end{equation} \begin{equation} \label{7'}r_{\beta \alpha }({\bf w},{\bf n}):=2^{-1}\Delta _\beta \cdot t_{\beta ,\alpha }({\bf w})^{3\mid \beta \mid -5}K_{\alpha ,\beta }({\bf w},% {\bf u}_{\beta \alpha }), \end{equation} where the constant $\Delta _\beta $ is introduced by the Jacobian of ${\bf % w\rightarrow }(\underline{V},r,{\bf n})$. With the definitions (\ref{4}), (% \ref{4'}), we can write% $$ P_k(f)({\bf v},{\bf x}) $$ \begin{equation} \label{8}=\,\sum_{\alpha ,\beta \in {\cal M}}\alpha _k\int_{{\bf R}^{3\mid \alpha \mid -3}\times \Omega _{3\mid \beta \mid -4}}d{\bf \tilde w}% _{(k)}\otimes d{\bf n\,}\left[ p_{\beta \alpha }({\bf w},{\bf n})f_\beta (% {\bf u}_{\beta \alpha },{\bf x})\right] _{{\bf w}_{k,\alpha _k}={\bf v}}, \end{equation} \smallskip\ $$ R_k(f)({\bf v},{\bf x}) $$ \begin{equation} \label{8'}\,=\,\sum_{\alpha ,\beta \in {\cal M}}\alpha _k\int_{{\bf R}% ^{3\mid \alpha \mid -3}\times \Omega _{3\mid \beta \mid -4}}d{\bf \tilde w}% _{(k)}\,\otimes d{\bf n\,}\left[ r_{\beta \alpha }({\bf w},{\bf n})f_{\alpha ,k}({\bf w},{\bf x})\right] _{{\bf w}_{k,\alpha _k}={\bf v}}. \end{equation} For $f$ as in Prop. 1, we define $S_k(f)({\bf v},{\bf x})=f_k({\bf v},{\bf x}% )R_k(f)({\bf v},{\bf x})$ with $R_k(f)$ given by (\ref{8'}). We point out the following simple relations resulting from the definition of ${\bf u}_{\beta \alpha }$, provided that, condition (\ref{5}), is fulfilled: \begin{equation} \label{9} \begin{array}{c} V_\beta ( {\bf u}_{\beta \alpha })=V_\alpha ({\bf w}),\quad W_\beta ({\bf u}_{\beta \alpha })=W_\alpha ({\bf w}),\medskip \\ \;W_{r,\beta }({\bf u}_{\beta \alpha })=W_{r,\alpha }({\bf w})+\sum_{n=1}^N(\alpha _n-\beta _n)E_n. \end{array} \end{equation} By (\ref{3}) and (\ref{9}), there exists some constant $K>0$ such that if condition (\ref{5}) is fulfilled, then (for $q\in \left[ 0,1\right] $ introduced in (\ref{3})), \begin{equation} \label{10}p_{\beta \alpha }({\bf w},{\bf n})\le K\left[ 1+W_{r,\alpha }({\bf % w})^{q/2}\right] ,\quad r_{\beta \alpha }({\bf w},{\bf n})\le K\left[ 1+W_{r,\alpha }({\bf w})^{q/2}\right] . \end{equation} {\bf REMARK:} In the definitions of $p_{\beta \alpha }$ and $r_{\beta \alpha }$ , the presence of $t_\beta $ exhibits the contributions of the reaction thresholds. \ \ \section{Existence theory} In this paper we are interested to solve Eq.(\ref{1}) in $\dot C_0$ (the space of continuous distribution functions, vanishing at infinity in the velocity and position variables). With the notations of Prop.1, for some $% \tau >0$ fixed, set $P(f)=(P_1(f),...,P_N(f))$ and $S(f)=(S_1(f),\ldots ,S_N(f))$, $\forall \,\,f\in \dot C_\tau \subset \dot C_0$. Then $% f\rightarrow P(f)$ and $f\rightarrow S(f)$, considered as maps in $\dot C_0$% , have extensions (also denoted $P$ and $S)$ to their natural domains in $% \dot C_0$. The Cauchy problem for Eq.(\ref{1}) formulated in $\dot C_0$ is \begin{equation} \label{11}d_tf=Af+P(f)-S(f),\quad f(t=0)=f_0, \end{equation} with $A$ the infinitesimal generator of the positivity preserving, continuous group $\{U^t\}_{t\in {\bf R}}$ of isometries of $\dot C_0$, given by its $\dot C_{n,0}$ components, $1\le n\le N,$% \begin{equation} \label{u}(U^tf)_n({\bf v},{\bf x}):=U_n^tf_n({\bf v},{\bf x})=f_n({\bf v},% {\bf x}-t{\bf v}),\quad (t,{\bf v},{\bf x})\in {\bf R}\times {\bf R}^3\times {\bf R}^3. \end{equation} We call $f\in C(0,T;\dot C_0)$ a mild solution, on $\left[ 0,T\right] $, of Eq.(\ref{11}) (in $\dot C_0$ ) if $P(f)$, $S(f)\in C(0,T;\dot C_0)$, and $% \,f $ satisfies \begin{equation} \label{12}f(t)=U^tf_0+\int_0^tU^{t-s}P(f(s))ds-\int_0^tU^{t-s}S(f(s))ds, \end{equation} (the integral being in $\dot C_0$ in the sense of Riemann ). Our main result states the existence, uniqueness and positivity of mild solutions, for initial data close to the vacuum state. These solutions are (time) global in the case of the gas with purely exo-energetic reactions and/or elastic (multiple) collisions. For $T>0$ fixed, consider $C(0,T;\dot C_\tau )$ with the usual sup norm, denoted $\left\| \circ \right\| _\tau $. If $g=(g_1,...,g_N)\in C(0,T;\;\dot C_\tau )$ and $t\in \left[ 0,T\right] $, by $g_n(t,{\bf v},{\bf x})$, denote the value of $g_n(t)\in \dot C_{n,\tau }$ , $1\leq n\leq N$, at $({\bf v},% {\bf x})\in {\bf R}^3\times {\bf R}^3$. Let $\dot C_\tau ^{+}:=\left\{ g=(g_1,...,g_N)\in \dot C_\tau :\;g_n({\bf v},{\bf x})\geq 0,\forall \,\;(% {\bf v},{\bf x})\in {\bf R}^3\times {\bf R}^3\,;n=1,...,N\right\} $. Finally, for some $R>0$, put ${\cal H}_\tau (R)=\left\{ h\mid h\in C(0,T;\dot C_\tau ^{+}),\left\| \;h\right\| _\tau \le R\right\} $. {\bf THEOREM 1 }{\it Let} $\tau >0$ {\it and} $f_0\in \dot C_\tau ^{+}$. {\bf \ }{\it a) For each} $T>0,\;\exists \,R_T$, $R_T^{*}>0$ {\it such that if } $\mid f_0\mid _\tau \le R_T$, {\it then Eq.(\ref{11}) in $\dot C_0$}, {\it \ has a unique mild solution} $f$ {\it on $\left[ 0,T\right] $}, {\it % satisfying} $U^{-t}f\in {\cal H}_\tau (R_T^{*})$.{\it \ The map }$% f_0\rightarrow f$ {\it is continuous from }$\left\{ h\in \dot C_\tau ^{+}:\left| h\right| _\tau \,\leq R_T\right\} $ {\it to } $C(0,T;\;\dot C_0)$% . {\it b)} {\it Assume} {\it that} $K_{\alpha ,\beta }\equiv 0$ {\it for each couple} $(\alpha ,\beta )$ {\it that yields} $\sum_{n=1}^N(\alpha _n-\beta _n)E_n<0$ ({\it exo-energetic reactions). In this case}, $\exists \,R,R^{*}>0 $ {\it such that if } $\mid f_0\mid _\tau \le R${\it , then for each} $T>0,$ {\it Eq. (\ref{11}) in $\dot C_0$},{\it \ has a unique mild solution} $f$, {\it on }$\left[ 0,T\right] $, {\it satisfying} $U^{-t}f\in {\cal H}_\tau (R^{*})$. {\it The map} $f_0\rightarrow f$ {\it is continuous from} $\left\{ h\in \dot C_\tau ^{+}:\left| h\right| _\tau \,\leq R_T\right\} $ {\it to} $C(0,T;\;\dot C_0)$. {\it c) In each situation, $U^{-t}f\in C^1(0,T;\dot C_0)$}. \ {\bf REMARK} - In the case of the simple gas, with binary elastic collisions, the statements of Theorem 1.b) reduce to known results on the classical Boltzmann equation. The proof of Theorem 1 will be given in several steps. We would like to apply the Banach fixed point theorem to Eq. (\ref{12}) in $C(0,T;\dot C_0)$. This is not, directly, possible since, $P$ and $S$ may be unbounded. However, writing Eq. (\ref{12}) more conveniently, the smothering properties of the time integrals appear to play a compensating role. \ The argument uses the following key estimation, extending certain energetic inequalities, obtained for the classical Boltzmann equation in \cite{be}. For some $\gamma \in {\cal M}$, define \begin{equation} \label{13}\Phi _\gamma (t,{\bf w},{\bf x},{\bf v}):=\sum_{n\in {\cal N}% (\gamma )}\sum_{i=1}^{\gamma _n}m_n\{[{\bf x}-t({\bf w}_{n,i}-{\bf v})]^2+% {\bf w}_{n,i}^2\}, \end{equation} for all ${\bf w\in {R}}^{3\mid \gamma \mid },\;{\bf x},{\bf v}\in {\bf R}^3$% , $t>0$. Also set% $$ \Gamma _{k\gamma }(t,{\bf v},{\bf x}):=\gamma _k\exp [\tau m_k({\bf v}^2+% {\bf x}^2){\bf ]} $$ \begin{equation} \label{14}\times \int_{{\bf R}^{3\mid \gamma \mid -3}}d{\bf \tilde w}% _{(k)}\int_0^tds\left[ (1+W_{r,\gamma }({\bf w})^{q/2})\exp (-\tau \Phi _\gamma (s,{\bf w},{\bf x},{\bf v}))\right] _{{\bf w}_{k,\gamma _k}={\bf v}% }, \end{equation} for all ${\bf v}$, ${\bf x}\in {\bf R}^3$,$\ t>0$; $q\in \left[ 0,1\right] $. {\bf LEMMA 1}.{\it a) Under conditions (\ref{15}),} \begin{equation} \label{16}\Phi _\beta (t,{\bf u},{\bf x},{\bf v})=\Phi _\alpha (t,{\bf w},% {\bf x},{\bf v})+2(1+t^2)\sum_{n=1}^N(\alpha _n-\beta _n)E_n. \end{equation} {\it b) }$\max \limits_{\gamma \in {\cal M,}1\leq k\leq N}$ $\sup \{\Gamma _{k\gamma }(t,{\bf v},{\bf x})\mid (t,{\bf v},{\bf x})\in {\bf R}_{+}\times {\bf R}^3\times {\bf R}^3\}=const<\infty $.\ The proof is given in Appendix B. Let $T,\tau >0$ and $f_0\in \dot C_\tau ^{+}$. With the substitution $% g(t):=U^{-t}f(t)$, Eq.(\ref{12}) becomes, \begin{equation} \label{17}g(t)=f_0+\int_0^tP^{\#}(g)(s)ds-\int_0^tS^{\#}(g)(s)ds,\;0\le t\le T, \end{equation} Here $P^{\#}$ and $S^{\#}$ are considered as operators in $C(0,T;\dot C_0)$, defined on their natural domains {\em D}$(P^{\#})$ and {\em D}$(S^{\#})$, respectively, by% $$ P^{\#}(g)(t):=U^{-t}P(U^tg(t)),\quad \;S^{\#}(g)(t):=U^{-t}S(U^tg(t)). $$ It follows that, we can prove Theorem 1, by looking for those $g\in C(0,T;\dot C_0^{+})$ solving Eq.(\ref{17}) in $\dot C_0$. Since $U^t$ leaves $C(0,T;\dot C_0^{+})$ invariant, we may equivalently look for those $% g=(g_1,...,g_N)\in {\em D}(P^{\#})\cap {\em D}(S^{\#})\cap C(0,T;\;\dot C_0^{+})$, solving the system \begin{equation} \label{18}g_k(t,{\bf v},{\bf x})=I_k(g)(t,{\bf v},{\bf x}),\;k=1,...,N, \end{equation} $(t,{\bf v},{\bf x})\in \left[ 0,T\right] \times {\bf R}^3\times {\bf R}^3$. Here $I_k(g)\in C(\left[ 0,T\right] \times {\bf R}^3\times {\bf R}^3)$ is given by% $$ I_k(g)(t,{\bf v},{\bf x})=f_{k,0}(t,{\bf v},{\bf x})\exp [-\int_0^tR_k^{\#}(g)(\lambda ,{\bf v},{\bf x})d\lambda ]\ $$ \begin{equation} \label{19}+\int_0^t\exp [-\int_s^tR_k^{\#}(g)(\lambda ,{\bf v},{\bf x}% )d\lambda ]P_k^{\#}(g)(s,{\bf v},{\bf x})ds,\, \end{equation} with $R_k^{\#}(g)(t,{\bf v},{\bf x}):=U_k^{-t}R_k(U^tg(t))({\bf v},{\bf x})$% , $k=1,...,N$ (the integrals being in the classical sense). Obviously, the system (\ref{18}) represents a weak form of Eq.(\ref{17}). Due to the assumptions on $f_0$, it will appear that Eq.(\ref{18}) has solutions given by elements of $C(0,T;\dot C_\tau ^{+})$. Let $I(g):=(I_1(g),...,I_N(g))$. We show that $g\rightarrow I(g)$ fulfills the conditions for applying the Banach fixed point theorem in ${\cal H}_\tau (R)$, for $R$ small enough.\ {\bf PROPOSITION 2.}a){\it \ If }$g\in C(0,T;\dot C_\tau ^{+})$, {\it then also }$I(g)\in C(0,T;\;\dot C_\tau ^{+})$. {\it b)} {\it For each} $T>0$, {\it there exist }$R_T,R_T^{*}>0$, {\it with }% $R_T^{*}\rightarrow 0$, {\it as }$R_T\rightarrow 0$, {\it such that if }$% \mid f_0\mid _\tau \le R_T$, {\it then} $g\rightarrow I(g)$ {\it leaves }$% {\cal H}_\tau (R_T^{*})$ {\it invariant. Moreover, if} $K_{\beta ,\alpha }\equiv 0,$ {\it whenever} $\sum_{n=1}^N(\alpha _n-\beta _n)E_n<0,$ {\it % then there exist }$R,R^{*}>0$,{\it \ independent of} $T$, {\it with }$% R^{*}\rightarrow 0$, {\it as $R$} {\it $\rightarrow 0$}, {\it \ such that if }$\mid f_0\mid _\tau \le R$, {\it then } $g\rightarrow I(g)$ {\it leaves }$\;% {\cal H}_\tau (R^{*})$ {\it invariant.}\ {\bf Proof. }a) First, remark that $C(0,T;\dot C_{k,\tau })$, $k=1,...,N$, can be identified with the set of those $h\in C([0,T]\times {\bf R}^3\times {\bf R}^3)$ (real) with the property \begin{equation} \label{20}\sup \limits_{\left| {\bf x}\right| +\left| {\bf v}\right| \geq r}\{\exp [\tau m_k({\bf x}^2+{\bf v}^2)]\left| h(t,{\bf v},{\bf x})\right| \}\rightarrow 0\quad as\;r\rightarrow \infty , \end{equation} uniformly in $t\in [0,T]$. We verify (\ref{20}). If $g\in C(0,T;\dot C_\tau ^{+}),$ $\gamma \in {\cal M}$, denote% $$ G_\gamma ^{\#}(t,{\bf w,\,x},{\bf v})=\Pi _{n\in {\cal N}(\gamma )}\Pi _{i=1}^{\gamma _n}g_n(t,\,{\bf w}_{n,i},\;{\bf x}-t({\bf w}_{n,i}-{\bf v}% ))\exp \left[ \tau \Phi _\gamma (t,{\bf w},{\bf x},{\bf v})\right] . $$ Using the definitions of $P^{\#}$ and $P$, as well as Rel.(\ref{10}) and Lemma 1.a), we estimate (\ref{19}): since $R_k^{\#}(g)(t,{\bf v},{\bf x})\ge 0$, $1\le k\le N$, the exponents are negative in (\ref{19}) ; moreover, $% P_k^{\#}(g)(t,{\bf v},{\bf x})\geq 0$; then, with the notations of Rel. (\ref {8}), (\ref{8'}), for some constant $K>0$, $$ 0\le I_k(g)(t,{\bf v},{\bf x}) $$ $$ \leq f_{k,0}(t,{\bf v},{\bf x})+K\sum_{\alpha ,\beta \in {\cal M}}\alpha _k\int_{{R}^{3\mid \alpha \mid -3}\times \Omega _{3\mid \beta \mid -4}}d{\bf % \tilde w}_{(k)}\otimes d{\bf n\;}\left[ (1+W_{r,\alpha }({\bf w}% )^{q/2})\right. \ $$ \begin{equation} \label{21}\left. \times \int_0^tds\;\Lambda _{\beta \alpha }(s)\,G_\beta ^{\#}(s,{\bf u}_{\beta \alpha },{\bf x},{\bf v})\ \exp [-\tau \Phi _\alpha (s,{\bf w},{\bf x},{\bf v})]\right] _{_{{\bf w}_{k,\alpha _k}={\bf v}}}, \end{equation} for all $\;(t,{\bf v},{\bf x})\in \left[ 0,T\right] \times {\bf R}^3\times {\bf R}^3$. Here, \begin{equation} \label{22}\Lambda _{\beta \alpha }(t)=\left\{ \begin{array}{l} \exp \left[ -2\tau (1+t^2)\sum_{n=1}^N(\alpha _n-\beta _n)E_n\right] ,\;if\;K_{\beta ,\alpha } \not\equiv 0,\medskip\ \\ 0,\;if\;K_{\beta ,\alpha }\equiv 0. \end{array} \right. \end{equation} Since $g\in C(0,T;\,\dot C_\tau ^{+})$, by (\ref{20}), (\ref{9}), there exists $r\ge 0$ such that $G_\beta ^{\#}(t,{\bf u,x},{\bf v})$ $\le \epsilon \left\| g\right\| _\tau ^{\mid \beta \mid -1}$, provided that $\Phi _\beta (t,{\bf u},{\bf v},{\bf x})\ge r$. Observe that $m_k({\bf x}^2+{\bf v}^2)$ $% \leq \Phi $$_\alpha (t,{\bf w},{\bf v},{\bf x})\mid _{_{{\bf w}_{k,\alpha _k}={\bf v}}}$. Consequently, by Lemma 1a), for each$\;\epsilon >0$, there exists $r_0>0$ (possibly depending on $T$) such that if $({\bf x}^2+{\bf v}% ^2)\ge r_0$, then% $$ 0\le G_\beta ^{\#}(t,\,{\bf u}_{\beta \alpha },\,{\bf x},\,{\bf v})_{\mid _{% {\bf w}_{k,\alpha _k}={\bf v}}}\le \epsilon \left\| g\right\| _\tau ^{\mid \beta \mid -1}, $$ uniformly in the rest of variables. We introduce the last inequality in (\ref {21}). There exist two constants $K_1>0$ and $r_1>0$, such that, if ${\bf x}% ^2+{\bf v}^2\ge r_1$, then% $$ 0\le I_k(g)(t,{\bf v},{\bf x})\le f_{k,0}({\bf v},{\bf x})\ + $$ \begin{equation} \label{24}+\epsilon \,K_2\,\Lambda (T)\exp [-\tau m_k({\bf x}^2+{\bf v}% ^2)]\sum_{\mid \alpha \mid \ge 2,\,,\mid \beta \mid \ge 2}\Gamma _{k\alpha }(t,{\bf v},{\bf x})\left\| g\right\| _\tau ^{\mid \beta \mid -1}, \end{equation} for all $\;(t,{\bf v},{\bf x})\in \left[ 0,T\right] \times {\bf R}^3\times {\bf R}^3$. Here, \begin{equation} \label{23}\Lambda (T):=\sup \nolimits_{\alpha ,\beta \in {\cal M}}\;[\sup \nolimits_{0\le t\le T}\Lambda _{\beta \alpha }(t)]. \end{equation} Since $f_{k,0}$ is in $\dot C_{k,\tau }$, it is now sufficient to apply Lemma 1.b) to obtain that (\ref{20}) is satisfied. This concludes the proof of a). b). Since $0\le G_\beta ^{\#}\le \left\| g\right\| _\tau ^{\mid \beta \mid }$% , the same procedure as before implies% $$ \mid I_k(g)(t)\mid _{k,\tau }\le \mid f_0\mid _\tau +K_2\Lambda (T)\sum_{\mid \beta \mid \ge 2}\left\| g\right\| _\tau ^{\mid \beta \mid }, $$ for some constant $K_2>0$. Now the argument can be easily concluded. $\Box $% \ Let $I$ denote the map $g\rightarrow I(g)$, according to Prop 2.b). Clearly, Eq.(\ref{18}) can be formulated in $C(0,T;\dot C_\tau ^{+})$ as \begin{equation} \label{nou}g=I(g). \end{equation} \ {\bf PROPOSITION 3.} {\it For each} $T>0$ {\it there exist} $R_T,R_T^{*}>0$, {\it with} $R_T^{*}\rightarrow 0$, {\it as }$R_T\rightarrow 0$, {\it such that if } $\mid f_0\mid _\tau \le R_T$, {\it then} $I$ {\it is a strict contraction on} ${\cal H}_\tau (R_T^{*})${\it . Assume that} $K_{\beta ,\alpha }\equiv 0,$ {\it whenever} $\sum_{n=1}^N(\alpha _n-\beta _n)E_n\le 0$ {\it . In this case, there exist} $R$, $R^{*}>0${\it , with $% R^{*}\rightarrow 0$},{\it \ as $R\rightarrow 0$, independent of }$T$, {\it \ such that if} $\mid f_0\mid _\tau \le R$, {\it then} $I$ {\it is a strict contraction on} ${\cal H}_\tau (R^{*})${\it .}\ {\bf Proof.} {\it -} By (\ref{19}) , for $g,h\in C(0,T;\dot C_\tau ^{+})$, we can write% $$ \mid I_k(g)(t,{\bf v},{\bf x})-I_k(h)(t,{\bf v},{\bf x})\mid $$ \begin{equation} \label{25}\le Q_k^A(g,h)(t,{\bf v},{\bf x})+Q_k^B(g,h)(t,{\bf v},{\bf x}% )+Q_k^C(g,h)(t,{\bf v},{\bf x}), \end{equation} with% $$ Q_k^A(t,{\bf v},{\bf x}):=f_{k,0}({\bf v},{\bf x})\mid \exp [-\int_0^tR_k^{\#}(g)(\lambda ,{\bf v},{\bf x})d\lambda ]-\exp [-\int_0^tR_k^{\#}(h)(\lambda ,{\bf v},{\bf x})d\lambda ]\mid , $$ $$ Q_k^B(t,{\bf v},{\bf x}) $$ $$ :=\int_0^tds\,P_k^{\#}(g)(s,{\bf v},{\bf x})\left| \exp [-\int_s^tR_k^{\#}(g)(\lambda ,{\bf v},{\bf x})d\lambda ]\ -\exp [-\int_s^tR_k^{\#}(h)(\lambda ,{\bf v},{\bf x})d\lambda ]\right| , $$ $$ Q_k^C(t,{\bf v},{\bf x}):=\int_0^t\exp [-\int_s^tR_k^{\#}(h)(\lambda ,{\bf v}% ,{\bf x})d\lambda ]\mid P_k^{\#}(g)(s,{\bf v},{\bf x})-P_k^{\#}(h)(s,{\bf v},% {\bf x})\mid ds. $$ for all $(t,{\bf v},{\bf x})\in \left[ 0,T\right] \times {\bf R}^3\times {\bf R}^3$; $k=1,\ldots ,N$. First we estimate $Q_k^A(t,{\bf v},{\bf x})$ . Since $g_k(t,{\bf v},{\bf x})$% , $h_k(t,{\bf v},{\bf x})\ge 0$, then $R_k^{\#}(g)(t,{\bf v},{\bf x})\ge 0$ and $R_k^{\#}(h)(t,{\bf v},{\bf x})\ge 0$, hence we can write% $$ Q_k^A(g,h)(t,{\bf v},{\bf x})\le f_{k,0}({\bf v},{\bf x})\int_0^t\mid R_k^{\#}(g)(\lambda ,{\bf v},{\bf x})-R_k^{\#}(h)(\lambda ,{\bf v},{\bf x}% )\mid d\lambda , $$ for all$\;(t,{\bf v},{\bf x})\in \left[ 0,T\right] \times {\bf R}^3\times {\bf R}^3$. Using the definitions of $R_k^{\#},\;$by arguments similar to those in the proof of Prop 2, there exists a constant $C_1>0$ such that% $$ Q_k^A(g,h)(t,{\bf v},{\bf x}) $$ \begin{equation} \label{26}\leq C_1f_{k,0}({\bf v},{\bf x})\left\| g-h\right\| _\tau \left( \sum_{\mid \alpha \mid \ge 2}\sum_{n=0}^{\mid \alpha \mid -2}\left\| g\right\| _\tau ^{\mid \alpha \mid -n-2}\left\| h\right\| _\tau ^n\Gamma _{k\alpha }(t,{\bf v},{\bf x})\right) . \end{equation} for all$\;(t,{\bf v},{\bf x})\in \left[ 0,T\right] \times {\bf R}^3\times {\bf R}^3$. Since% $$ Q_k^B(g,h)(t,{\bf v},{\bf x)} $$ $$ \leq \int_0^tP_k^{\#}(g)(s,{\bf v},{\bf x})\,ds\left( \int_0^t\mid (R_k^{\#}(g)(\lambda ,{\bf v},{\bf x})-R_k^{\#}(h)(\lambda ,{\bf v},{\bf x}% )\mid d\lambda \right) , $$ similar estimations give (for some constant $C_2>0$)% $$ Q_k^B(g,h)(t,{\bf v},{\bf x)} $$ $$ \leq C_2\Lambda (T)\exp [-\tau m_k({\bf x}^2+{\bf v}^2)]\left\| g-h\right\| _\tau \left( \sum_{\mid \beta \mid \ge 2}\left\| g\right\| _\tau ^{\mid \beta \mid }\right) $$ \begin{equation} \label{27}\times \left( \sum_{\mid \alpha \mid \ge 2}\Gamma _{k\alpha }(t,% {\bf v},{\bf x})\right) \left( \sum_{\mid \alpha \mid \ge 2}\sum_{n=0}^{\mid \alpha \mid -2}\left\| g\right\| _\tau ^{\mid \alpha \mid -n-2}\left\| h\right\| _\tau ^n\Gamma _{k\alpha }(t,{\bf v},{\bf x})\right) , \end{equation} with $\Lambda (T)$ defined by (\ref{23}); $(t,{\bf v},{\bf x})\in \left[ 0,T\right] \times {\bf R}^3\times {\bf R}^3$. In the same way, for some constant $C_3>0,$we obtain% $$ Q_k^C(g,h)(t,{\bf v},{\bf x})\ \le C_3\Lambda (T)\exp [-\tau m_k({\bf x}^2+% {\bf v}^2)] $$ \begin{equation} \label{28}\times \left\| g-h\right\| _\tau \left( \sum_{\mid \alpha \mid \ge 2}\Gamma _{k\alpha }(t,{\bf v},{\bf x})\right) \left( \sum_{\mid \beta \mid \ge 2}\sum_{n=1}^{\mid \beta \mid -1}\left\| g\right\| _\tau ^{\mid \beta \mid -n-1}\left\| h\right\| _\tau ^n\right) , \end{equation} $(t,{\bf v},{\bf x})\in \left[ 0,T\right] \times {\bf R}^3\times {\bf R}^3$. By inequalities (\ref{26})-(\ref{28}), applying Lemma 1.b), one can find a constant $C_0>0$ and a polynomial $p(\cdot )$ with positive coefficients, such that $\forall \,r>0$, \begin{equation} \label{29}\left\| I(g)-I(h)\right\| _\tau \le C_0[\mid f_0\mid _\tau +r\Lambda (T)]\,p(r)\left\| g-h\right\| _\tau , \end{equation} provided that $g,h\in {\cal H}_\tau (r)$. Now, by Prop 2 and Rel. (\ref{29}), we can choose $R_T,R_T^{*}>0$ such that if $\mid f_0\mid _\tau \le R_T$ then $I$ is a strict contraction on ${\cal H}% _\tau (R_T^{*})$. By (\ref{23}), if $K_{\beta ,\alpha }\equiv 0$, whenever $% \sum_{n=1}^N(\alpha _n-\beta _n)E_n\le 0,$ then $\Lambda (T)=1$, $\forall T>0.$ Consequently there exist $R,\;R^{*}>0$, independent of $T$, such that if $\mid f_0\mid _\tau \le R$, then $I$ is a strict contraction on ${\cal H}% _\tau (R^{*})$. This concludes the proof of Prop. 3.$\Box $ The existence and uniqueness part in Theorem 1a) follows by Prop.2, Prop.3 and the Banach fixed point theorem: for $R_T,R_T^{*}>0$, small enough, Eq.(% \ref{nou}), with $\left| f_0\right| \leq R_T$, can be uniquely solved in $% {\cal H}_\tau (R_T^{*})$. To conclude the argument it is sufficient to remark that ${\cal H}_\tau (R_T^{*})\subset {\em D}(P^{\#})\cap {\em D}% (S^{\#})\subset $ $C(0,T,\dot C_0)$. To prove the rest of Theorem 1a), namely the continuity of the solution in the initial datum, first remark by Prop.2, that, for each $g\in {\cal H}_\tau (R_T^{*})$, fixed, the map $% f_0\rightarrow I(g)$ is continuous from $\left\{ h\in \dot C_\tau ^{+}:\,\left| h\right| _\tau \leq R_T\right\} $ to $C(0,T,\dot C_\tau )$. Then the proof follows by means of the inequality (\ref{29}). Part b) of Theorem 1 can be similarly proved. \ Part c) of Theorem 1 is immediate: $g\in C^1(0,T;\,\dot C_\tau )$, by Eq.(% \ref{17}), while the solution $f$ of Eq.(\ref{12}) is related to $g$ by $% f=U^tg$. $\;\Box $ \section{Conservation relations, H-Theorem and the law of the mass action} In this section we prove the global conservations relations for mass, momentum and energy and a H-Theorem analogous to the results obtained in the case of the classical Boltzmann equation with elastic binary collisions. Set $\Psi _k^0({\bf v})=m_k,\Psi _k^4({\bf v})=2^{-1}m_k{\bf v}^2+E_k$ and $% \Psi _k^i({\bf v})=m_k{\bf v}_i$ for all ${\bf v}=({\bf v}_1,{\bf v}_2,{\bf v% }_3)\in {\bf R}^3$, with ${\bf v}_i\in {\bf R},\;i=1,2,3,\;1\le k\le N$. The following result states the bulk momentum and energy conservation relations.\ \ {\bf THEOREM 2}.{\it a) Let }$\tau >0$ {\it and $f\in \dot C_\tau $. Then for% } $i=0,1,\ldots ,4$, $$ \sum_{k=1}^N\int_{{\bf R}^3}\Psi _k^i({\bf v})\left( P_k(f)({\bf v},{\bf x}% )-S_k(f)({\bf v},{\bf x})\right) d{\bf v}\equiv 0. $$ $\forall {\bf x\in R}^3$. b) {\it If} $f_0$ {\it and} $f$ {\it are as in Theorem 1, then, for each }$% t\in \left[ O,T\right] $, $$ \sum_{k=1}^N\int_{{\bf R}^3\times {\bf R}^3}\Psi _k^i({\bf v})f_k(t,{\bf v},% {\bf x})\,d{\bf v\otimes }d{\bf x}\equiv \sum_{k=1}^N\int_{{\bf R}^3\times {\bf R}^3}\Psi _k^i({\bf v})f_{k,0}({\bf v},{\bf x})\,d{\bf v\otimes }d{\bf % x.} $$ \ {\bf Proof.} - a) We give the argument for $\Psi _k^4$, the other cases being similar. By Prop.1.b),% $$ \sum_{k=1}^N\int_{{\bf R}^3}\Psi _k^4({\bf v})P_k(f)({\bf v},{\bf x})\,d{\bf % v}=\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}\sum_{\alpha ,\beta \in {\cal M}}\sum_{k=1}^N\sum_{i=1}^{\alpha _k}\int_{% {\bf R}^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid }}\ (2^{-1}m_k% {\bf w}_{k,i}^2+E_k) $$ $$ \times \ f_\beta ({\bf u},{\bf x})K_{\beta ,\alpha }({\bf u},{\bf w}% )\,\delta _\epsilon ^3(V_\beta ({\bf u})-V_\alpha ({\bf w}))\ \,\delta _\eta (W_\beta ({\bf u})-W_\alpha ({\bf w}))\ d{\bf u\otimes }d{\bf w\,} $$ $$ =\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}\sum_{\alpha ,\beta \in {\cal M}}\sum_{k=1}^N\sum_{i=1}^{\beta _k}\int_{% {\bf R}^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid }}(2^{-1}m_k% {\bf u}_{k,i}^2+E_k) $$ $$ \times \ \,f_\alpha ({\bf w},{\bf x})K_{\alpha ,\beta }({\bf w},{\bf u})\ \delta _\epsilon ^3(V_\beta ({\bf u})-V_\alpha ({\bf w}))\,\delta _\eta (W_\beta ({\bf u})-W_\alpha ({\bf w}))\ d{\bf u\otimes }d{\bf w\,,} $$ where the last equality results by interchanging $\alpha $ and ${\bf w}$ with $\beta $ and ${\bf u}$ respectively, and using the symmetry of $\delta _\epsilon ^3$ and $\delta _\eta $, respectively as well as the invariance of $K_{\beta ,\alpha }$ at permutations. Then it is sufficient to remark that% $$ \sum_{k=1}^N\int_{{\bf R}^3}\Psi _k^4({\bf v})\left( P_k(f)({\bf v},{\bf x}% )-S_k(f)({\bf v},{\bf x})\right) \,d{\bf v} $$ $$ =\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}\sum_{\alpha ,\beta \in {\cal M}}\int_{{\bf R}^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid }}\left( W_\beta ({\bf u})-W_\alpha ({\bf w}% )\right) $$ $$ \times \,f_\alpha ({\bf w},{\bf x})K_{\alpha ,\beta }({\bf w},{\bf u}% )\,\delta _\epsilon ^3(V_\beta ({\bf u})-V_\alpha ({\bf w}))\ \delta _\eta (W_\beta ({\bf u})-W_\alpha ({\bf w}_\alpha ))\ d{\bf u\otimes }d{\bf w\,}% \equiv 0. $$ b) Let $f$ be as in Theorem 1. Note that for each $t$ fixed, $U_k^t$, introduced in Rel.(\ref{u}) is a positivity preserving, linear isometry on $% L^1({\bf R}^3\times {\bf R}^3,\,d{\bf v\otimes }d{\bf x})$; $k=1,...,N.$ Further, by Theorem 1c), $g=U^{-t}f$ is of class $C^1$ and verifies Eq.(\ref {17}). Then for each $i=1,...,4$,% $$ {\frac d{dt}}\sum_{k=1}^N\int_{{\bf R}^3\times {\bf R}^3}\Psi _k^i\,({\bf v}% )f_k(t,{\bf v},{\bf x})\,d{\bf v\otimes }d{\bf x=}{\frac d{dt}}% \sum_{k=1}^N\int_{{\bf R}^3\times {\bf R}^3}\Psi _k^i({\bf v})\,g_k(t,{\bf v}% ,{\bf x})\,d{\bf v\otimes }d{\bf x}\ $$ $$ =\sum_{k=1}^N\int_{{\bf R}^3\times {\bf R}^3}\Psi _k^i({\bf v})\left( P_k^{\#}(g)(t,{\bf v},{\bf x})-S_k^{\#}(g)(t,{\bf v},{\bf x})\right) \,d{\bf % v\otimes }d{\bf x} $$ $$ =\sum_{k=1}^N\int_{{\bf R}^3\times {\bf R}^3}\Psi _k^i({\bf v})\left( P_k(f)(t,{\bf v},{\bf x})-S_k(f)(t,{\bf v},{\bf x})\right) \,d{\bf v\otimes }% d{\bf x}\equiv 0, $$ using again the $L^1$- properties of $U_k^t$ and Part a). This concludes the proof.$\Box $\ \ Let $C_1,\ldots ,C_N>0$ be constants and $C^\alpha :=C^{\alpha _1}\times \ldots \times C^{\alpha _N},\;\alpha \in {\cal M}$. In the rest of this section, we suppose the following detailed balance condition \begin{equation} \label{33b}C^\beta K_{\alpha ,\beta }({\bf w,u})\equiv C^\alpha K_{\beta ,\alpha }({\bf u,w)},\quad \forall \;{\bf w},\;{\bf u,\;}\alpha ,\;\beta , \end{equation} First remark that if $f=(f_1,..,f_N)\in \dot C_\tau ^{+}$, $\tau >0$, with $% f_n>0$, then $f_n\log f_n\in L^1({\bf R}^3\times {\bf R}^3,d{\bf v\otimes }d% {\bf x})$, for all $n=1,...,N$. The argument is standard (\cite{be}): let $% \log {}^{+}$ ($\log {}^{-}$) denote the positive (negative) part of the function $\log $;\ clearly $f_n\log {}^{+}f_n\in L^1$; it is sufficient to prove the same for $f_n\log {}^{-}f_n$; to this end, in the inequality $\xi \log {}^{-}\xi \leq \eta -\xi \log \eta $, valid for $\xi >0$ and $0<\eta \leq 1$ (\cite{be}), we take $\xi =f_n({\bf v},{\bf x})$ and $\eta =\exp (-% {\bf v}^2-{\bf x}^2)$, obtaining. \begin{equation} \label{tos}(f_n\log {}^{-}f_n)({\bf v},{\bf x})\leq \exp (-{\bf v}^2-{\bf x}% ^2)+({\bf v}^2+{\bf x}^2)f_n({\bf v},{\bf x}). \end{equation} Therefore we can define the H-function \begin{equation} \label{31}H(f)=\sum_{k=1}^N\int_{{\bf R}^3\times {\bf R}^3}\,f_k({\bf v},% {\bf x})\,[\log \left( C_kf_k({\bf v},{\bf x})\right) -1]\,d{\bf v\otimes }d% {\bf x.} \end{equation} {\bf PROPOSITION} 4.{\it a)}\ {\it Let }$\tau >0$, $f=(f_1,..,f_N)\in \dot C_\tau ^{+}$, {\it such that} $f_n>0$ {\it and} $\sup (1+{\bf x}^2+{\bf v}% ^2)^{-1}\left| \log f_n({\bf v},{\bf x})\right| <\infty $ {\it for all} $% n=1,...,N$. {\it Then} $$ D(f):=\sum_{k=1}^N\int_{{\bf R}^3\times {\bf R}^3}\left( P_k(f)({\bf v},{\bf % x})-S_k(f)({\bf v},{\bf x})\right) \log \left( C_kf_k({\bf v},{\bf x}% )\,\right) d{\bf v\otimes }d{\bf x}\leq 0. $$ {\it Moreover }$D(f)\equiv 0$ {\it iff for each} couple $(\alpha ,\beta )$ {\it such that} $K_{\alpha ,\beta }\not \equiv 0,$ {\it it follows that} \begin{equation} \label{34}C^\alpha f_\alpha ({\bf w},{\bf x})\equiv C^\beta f_\beta ({\bf u},% {\bf x}), \end{equation} $\forall \,{\bf x}\in {\bf R}^3$, {\it provided that} ${\bf w,u}$ {\it % satisfy (}\ref{15}{\it )}. {\it b) Let $f_0$} {\it and $f$} be as in{\it \ Theorem 1. In addition, suppose that } $f_{n,0}>0$ {\it and} $\sup (1+{\bf x}^2+{\bf v}% ^2)^{-1}\left| \log f_{n,0}({\bf v},{\bf x})\right| <\infty ${\it \ for all $% n=1,...,N$. Then }$t\rightarrow H(f)$ {\it is of class }$C^1${\it and}% $$ {\frac d{dt}}H(f)(t)=D(f(t)){\bf \,.} $$ \ {\bf Proof.}a) In our case, $D(f)$ is well defined. Then, by Prop.1b), taking $h=\log f_k$ and $f=(f_1,\ldots ,f_N)\in \dot C_\tau ^{+}$, with $% f_n>0,$ for all $n=1,\ldots ,N$, $$ \sum_{k=1}^N\int_{{\bf R}^3}P_k(f)({\bf v},{\bf x})\log \left( C_kf_k({\bf v}% ,{\bf x})\right) \,d{\bf v}=\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}\sum_{\alpha ,\beta \in {\cal M}}\int_{{\bf R% }^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid }}\,f_\beta ({\bf u},% {\bf x}) $$ $$ \times \log \left( C^\alpha f_\alpha ({\bf w},{\bf x})\right) K_{\beta ,\alpha }({\bf u},{\bf w})\,\delta _\epsilon ^3(V_\beta ({\bf u})-V_\alpha (% {\bf w}))\,\delta _\eta (W_\beta ({\bf u})-W_\alpha ({\bf w}))\,d{\bf % u\otimes }d{\bf w\,} $$ $$ =\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}\sum_{\alpha ,\beta \in {\cal M}}\int_{{\bf R}^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid }}\,f_\alpha ({\bf w},{\bf x})\log \left( C^\beta f_\beta ({\bf u},{\bf x})\right) $$ \begin{equation} \label{36}\times K_{\alpha ,\beta }({\bf w},{\bf u})\,\delta _\epsilon ^3(V_\beta ({\bf u})-V_\alpha ({\bf w}))\,\delta _\eta (W_\beta ({\bf u}% )-W_\alpha ({\bf w}))d{\bf u\otimes }d{\bf w\,} \end{equation} where the last equality results by interchanging $\alpha $ and ${\bf w}$ with $\beta $ and ${\bf u}$ respectively, and using the symmetry of $\delta _\epsilon ^3$ and $\delta _\eta $, respectively as well as the invariance of $K_{\beta ,\alpha }$ at permutations Similarly% $$ \sum_{k=1}^N\int_{{\bf R}^3}S_k(f)({\bf v},{\bf x})\log \left( C_kf_k({\bf v}% ,{\bf x})\right) \,d{\bf v}=\lim _{\eta \rightarrow 0}\lim _{\epsilon \rightarrow 0}\sum_{\alpha ,\beta \in {\cal M}}\int_{{\bf R}^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid }}\,f_\alpha ({\bf w},{\bf x}) $$ $$ \times \log \left( C^\alpha f_\alpha ({\bf w},{\bf x})\right) K_{\alpha ,\beta }({\bf w},{\bf u})\,\delta _\epsilon ^3\,(V_\beta ({\bf u})-V_\alpha (% {\bf w}))\,\delta _\eta (W_\beta ({\bf u})-W_\alpha ({\bf w}))\,d{\bf % u\otimes }d{\bf w\,} $$ $$ =\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}\sum_{\alpha ,\beta \in {\cal M}}\int_{{\bf R}^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid }}f_\beta ({\bf u},{\bf x})\log \left( C^\beta f_\beta ({\bf u},{\bf x})\right) $$ \begin{equation} \label{38}\times K_{\beta ,\alpha }({\bf u},{\bf w})\,\,\delta _\epsilon ^3(V_\beta ({\bf u})-V_\alpha ({\bf w}))\,\delta _\eta (W_\beta ({\bf u}% )-W_\alpha ({\bf w}))\,d{\bf u\otimes }d{\bf w\,} \end{equation} Then a few algebraic manipulations involving Rel. (\ref{36}), (\ref{38}) imply that% $$ D(f)=2^{-1}\int_{{\bf R}^3}d{\bf x\,}\lim \limits_{\eta \rightarrow 0}\lim \limits_{\epsilon \rightarrow 0}\sum_{\alpha ,\beta \in {\cal M}}\int_{{\bf R% }^{3\mid \beta \mid }\times {\bf R}^{3\mid \alpha \mid }}H_{\alpha ,\beta }(% {\bf w},{\bf u},{\bf x}) $$ \begin{equation} \label{40}\times \delta _\epsilon ^3(V_\beta ({\bf u})-V_\alpha ({\bf w}% ))\,\delta _\eta (W_\beta ({\bf u})-W_\alpha ({\bf w}))\,d{\bf u\otimes }d% {\bf w\,}, \end{equation} with $$ H_{\alpha ,\beta }({\bf w},{\bf u},{\bf x})=\left[ K_{\beta ,\alpha }({\bf u}% ,{\bf w})\,f_\beta ({\bf u},{\bf x})\,-K_{\alpha ,\beta }({\bf w},{\bf u}% )\,f_\alpha ({\bf w},{\bf x})\right] \log {\frac{C^\alpha f_\alpha ({\bf w},% {\bf x})}{C^\beta f_\beta ({\bf w},{\bf x})}.} $$ Assuming that condition (\ref{33b}) is fulfilled, it follows that $$ H_{\alpha ,\beta }({\bf w},{\bf u},{\bf x})=-K_{\beta ,\alpha }({\bf u},{\bf % w})f_\beta ({\bf u},{\bf x})\left( {\frac{C^\alpha f_\alpha ({\bf w},{\bf x}% ) }{C^\beta f_\beta ({\bf w},{\bf x})}-1}\right) \log {\frac{C^\alpha f_\alpha ({\bf w},{\bf x})}{C^\beta f_\beta ({\bf w},{\bf x})}}\le 0. $$ hence $D(f)\le 0$. Rel.(\ref{34}) is now obvious. b) Since $U_k^t$,$1\le k\le N$, is a positivity preserving isometry on $L^1(% {\bf R}^3\times {\bf R}^3,\,d{\bf v\otimes }d{\bf x})$, using Rel.(\ref{31}% ), we obtain $H(f(t))=H(g(t))$. But $g(t)=U^{-t}f(t)$ is the unique solution of Eq.(\ref{17}) and $g=U^{-t}f$ is of class $C^1$. From the definition of $% I $ and the fact that $f_0$ satisfies the conditions of Prop.4 a) it follows that $\forall \,t\in \left[ 0,T\right] $, $f(t)$ satisfies the conditions of Prop.4 a), so that $D(f(t))$ is well defined. Moreover, using again the $L^1 $-properties of $U_k^t$, $$ \sum_{k=1}^N\int_{{\bf R}^3\times {\bf R}^3}d{\bf v\otimes }d{\bf x}\left( P_k^{\#}(g)(t,{\bf v},{\bf x})-S_k^{\#}(g)(t,{\bf v},{\bf x})\right) \log \left( C_kg_k(t,{\bf v},{\bf x})\right) $$ $$ =\sum_{k=1}^N\int_{{\bf R}^3\times {\bf R}^3}d{\bf v\otimes }d{\bf x}\left( P_k(f)(t,{\bf v},{\bf x})-S_k(f)(t,{\bf v},{\bf x})\right) \log C_kf_k(t,% {\bf v},{\bf x}). $$ Putting all these together and then using part a), it follows that $\frac d{dt}H(f)(t)=\frac d{dt}H(g)(t)=$ $D(f)(t)$ $\leq 0$, concluding the proof.$% \Box $ The main result of this section follows from Prop 4 using techniques similar to those of \cite{be}.\ \ {\bf THEOREM 3.} {\it Let} $f_0=(f_{1,0},...,f_{N,0})$ and $% \;f=(f_1,...,f_N) $ {\it be} {\it as in Theorem 1 with }$f_{n,0}>0$,$\;1\le n\le N$. {\it Then under condition (}\ref{33b}),{\it \ $H(f(t_2))\leq H(f(t_1))$} {\it for all} $t_2\geq t_1$. {\bf Proof.} Define the sequence $% f_0^{(l)}=(f_{1,0}^{(l)},...,f_{N,0}^{(l)}) $, $l=1,2,...$ by setting for each $k=1,...,N$,% $$ f_{k,0}^{(l)}({\bf v},{\bf x})=\max \left\{ f_{k,0}({\bf v},{\bf x}),\;\frac{% \left| f_0\right| _\tau }{l(1+{\bf v}^2+{\bf x}^2)}\exp \left[ -\tau m_k(% {\bf v}^2+{\bf x}^2)\right] \right\} . $$ Let $f^{(l)}$and $f$ be the mild solutions of Eq.(\ref{11}), provided by Theorem 1, for initial data $f_0^{(l)}$ and $f_0$, respectively. Obviously, $% H(f(t))$ and $H(f^{(l)}(t))$ are well defined. Clearly $\left| f_0^{(l)}-f_0\right| _\tau \rightarrow 0$, as $l\rightarrow \infty $. Then $% \left\| f^{(l)}-f\right\| _0\rightarrow 0$, as $l\rightarrow \infty $, by the continuity in initial datum, stated in Theorem 1. Moreover, for each $% t\in \left[ 0,T\right] $, the sequence $\left( U^{-t}f^{(l)}(t)\right) _{l\in {\bf N}}$ is bounded in $\dot C_\tau $. Using (\ref{tos}), it follows that for each $t$, the sequence$(f^{(l)}(t)\log f^{(l)}(t))_{l\in {\bf N}}$ is bounded by some function in $L^1({\bf R}^3\times {\bf R}^3,\,d{\bf % v\otimes }d{\bf x})$, hence $H(f^{(l)}(t))\rightarrow H(f(t))$, as $% l\rightarrow \infty $, by the dominated convergence theorem. By (\ref{19}), and the definition of $f_{k,0}^{(l)}$, the conditions of Prop.4b) are fulfilled by $f_{k,0}^{(l)}$, for each $l$. Then the function $t\rightarrow H(f^{(l)}(t))$ is non decreasing. Consequently, the same is true for $H(f)$, concluding the proof. $\Box $.\ {\bf REMARK} Rel. (\ref{34}) is satisfied by local maxwellians (\cite{be}) and it provides a generalization of the law of the mass action. Indeed, under condition{\it \ }(\ref{33b}), let the local maxwellians solving Eq.(\ref{1}) be given by \begin{equation} \label{44}\omega _n=\omega _n(q_n,u,T):=q_n(m_n/2\pi kT)^{3/2}\exp [-m_n(% {\bf v}-{\bf u})^2/2kT], \end{equation} for all $\;\;n=1,\ldots ,N$. Here $q_n=q_n(t,{\bf x})$ is the concentration of species $n$, while ${\bf u}={\bf u}(t,{\bf x})$ and $T=T(t,{\bf x})$ are the notations for the gas bulk velocity and the equilibrium temperature, respectively ( $k$ denotes the Boltzmann constant).\ \ {\bf COROLLARY 1.}{\it \ The concentrations } $q_n$, $n=1,\ldots ,N$, {\it % satisfy non trivially the law of the mass action, i.e., for all} $\alpha $, $% \beta $, \begin{equation} \label{45}\sum_{n=1}^N(\alpha _n-\beta _n)[{\frac 32}\log (m_n/2kT)+\log (C_nq_n)+E_n/kT]\equiv 0. \end{equation} \ {\bf Proof.} The result{\bf \ }is immediate from (\ref{34}), by the mass conservation condition, namely $K_{\alpha ,\beta }=0$ when $\sum_n(\alpha _n-\beta _n)m_n=0)$.$\Box $\ \section{Final remarks} Theorem 1 is applicable to the reactive, expanding gas. Our global existence results do not cover the case in which endo-energetic chemical reactions (collisions) are present in the gas processes. In the latter case, one should avoid possible pathologies, introduced by the particles which may loose completely their relative kinetic energy during the endo-energetic reactions. The balance conditions (\ref{33b}) play no role in Theorems 1, 2, but it is essential for the validity of the results stated in Theorem 3. Theorem 1 can be analogously proved considering instead of $\dot C_0$ the space of those $f=(f_1,...,f_N)$ with $f_n\in L^1({\bf R}^3\times {\bf R}^3;d% {\bf v}\otimes d{\bf x)}$, equipped with the norm $\max _{1\leq k\leq N}\left| f_k\right| _{L^1}$. In the latter case, the continuity condition on $K_{\alpha ,\beta }$ can be replaced by measurability.\ Then $\dot C_\tau $ (with $\tau >0$) need be replaced by a space of measurable functions (e.g. $% \dot C_{n,\tau }$ can be replaced by the space of those $h\in L^\infty ({\bf % R}^3\times {\bf R}^3;d{\bf v}\otimes d{\bf x)}$, with $ess\sup \exp \left[ \tau ({\bf x}^2+m_k{\bf v}^2)\right] \left| h({\bf v},{\bf x})\right| <\infty $ ). Some of the results of this paper have been announced in \cite{gr}, (where the main theorem is actually valid for $\tau =0$, since $\left\{ U^t\right\} _{t\in {\bf R}}$ is not a continuous group of isometries on $\dot C_\tau $ for $\tau >0$). \smallskip\ {\bf Acknowledgements:} One of us (C.P. Grunfeld) performed part of this work at University Paris VII. He would like to thank the Laboratory of Mathematical Physics and Geometry, particularly Prof. A. Boutet de Monvel for the warm hospitality. It is a privilege to him to thank the French C.N.R.S. for financial support. \medskip\ {\Large {\bf Appendix A}} \smallskip Let $n$ be non-negative integer and $a_1,\ldots ,a_n>0$, constants. Consider a positive quadratic form $T:=T({\bf v}_1,\ldots ,{\bf v}_n)=\sum_{i=1}^na_i% {\bf v}_i^2$ on ${\bf R}^{3n}$ , $r_i\in {\bf R}^3,1\le i\le n$. Consider the transformation% $$ \qquad {\bf R}^{3n}\ni ({\bf v}_1,\ldots ,{\bf v}_n)\rightarrow (\underline{V% },\zeta )\in {\bf R}^3\times {\bf R}^{3n-3}\;,\eqno (A.1) $$ defined by% $$ \underline{V}:=(\sum_{i=1}^na_i)^{-1}\sum_{i=1}^na_i{\bf v}_i, $$ $$ \zeta :=(\zeta _1,\ldots ,\zeta _{n-1}), $$ $$ \zeta _i:={\bf v}_{i+1}-(\sum_{j=1}^ia_j)^{-1}\sum_{j=1}^ia_j{\bf v}_j,\quad \;i=1,\ldots ,n-1. $$ By transformation (A.1), the form $T$ becomes% $$ T=T(\underline{{\sl V}},\zeta )=(\sum_{i=1}^na_i)\underline{{\sl V}}% ^2+\sum_{i=1}^{n-1}\mu _i\zeta _i^2, $$ with% $$ \mu _i^{-1}=a_{i+1}^{-1}+(\sum_{j=1}^ia_j)^{-1},\qquad i=1,\ldots ,n-1. $$ The system of coordinates on ${\bf R}^{3n}$ resulting from (A.1) will be called a Jacobi system of coordinates associated to $T({\bf v}_1,\ldots ,% {\bf v}_n)$. Moreover, the same term will designate the system of coordinates obtained by the transformation% $$ {\bf R}^{3n}\ni ({\bf v}_1,\ldots ,{\bf v}_n)\rightarrow (\underline{V},\xi )\in {\bf R}^3\times {\bf R}^{3n-3},\eqno (A.2) $$ where $\xi :=(\xi _1,\ldots ,\xi _{n-1})$ and $\xi _i:=\mu _i^{1/2}\zeta _i$ , with \underline{${\sl V}$} and $\zeta _i$ as in (A.1) ; $1\le i\le n-1$. Obviously, by (A.2), $T=T(\underline{{\sl V}},\xi )=(\sum_{i=1}^na_i) \underline{{\sl V}}^2+\xi ^2$. \ \medskip\ {\Large {\bf Appendix B}} \smallskip Part a) of Lemma 1 is straightforward. To prove b), we first consider $\mid \gamma \mid =2$ (with $\gamma _k\ge 1$). Then, by (\ref{10}) clearly, there exist two non-negative constants $c_0$ and $c$ such that% $$ \Gamma _{k\gamma }(t,{\bf v},{\bf x}) $$ $$ \;\le c_0\int_{{\bf R}^3}d{\bf y}(1+\mid {\bf y}-{\bf v}\mid ^q)\exp (-c{\bf % y}^2)\int_0^tds\exp \left\{ -c[{\bf x}-({\bf y}-{\bf v})s]^2\right\} \eqno % (B.1) $$ for all $t\ge 0,{\bf x},{\bf v}\in {\bf R}^3,q\in [0,1]$. By (B.1), the $\sup $ estimation on $\Gamma $$_{k\gamma }(t,{\bf v},{\bf x})$ reduces to known inequalities of Lemma 2.5 in \cite{be}. Briefly, in this case, for each $t\ge 0$,% $$ \;\Gamma _0(t,{\bf v},{\bf x}):=\int_0^tds\exp \left\{ -c[{\bf x}-({\bf y}-% {\bf v})s]^2\right\} $$ $$ \leq \int_0^tds\exp \left\{ -c[\left| {\bf x}\right| -\left| {\bf y}-{\bf v}% \right| s]^2\right\} \le \left| {\bf y}-{\bf v}\right| ^{-1}.\eqno (B.2) $$ Introducing (B.2) in the right side of (B.1), and integrating with respect to the reference frame with the{\bf \ }${\bf y}_3$ axis oriented in the direction of ${\bf v}$, we get% $$ \Gamma _{k\gamma }(t,{\bf v},{\bf x})\le \int_{-\infty }^{+\infty }d{\bf y}% _3\exp (-c{\bf y}_3^2)\int_0^\infty \rho {\frac{1+[\rho ^2+({\bf y}_3-{\bf v}% )^2]^{q/2}}{[\rho ^2+({{\bf y}_3-{\bf v}})^2]^{1/2}}}\exp [-c\rho ^2]\,d\rho $$ $$ \le const.\int_{-\infty }^{+\infty }d{\bf y}_3\exp (-c{\bf y}% _3^2)\int_0^\infty (1+\rho ^q)\exp [-c\rho ^2]d\rho \le const. $$ The case $\mid \gamma \mid >2$ (with $\gamma _k\ge 1$) can be reduced to $% \mid \gamma \mid =2$ as follows. With the notations of (\ref{14}), consider the form $T_\gamma ({\bf \tilde w}_{(k)}):=W_\gamma ({\bf w}% )-\sum_{n=1}^N\gamma _nE_n-2^{-1}m_k{\bf w}_{k,\gamma _k}^2$, representing the kinetic energy of $\left| \gamma \right| -1$ particles in the channel $% \gamma $ (more precisely, the kinetic energy of all the particles in channel $\gamma $, except the particle with velocity ${\bf w}_{k,\gamma _k}$). To $% T_\gamma ({\bf \tilde w}_{(k)})$, we associate a Jacobi system of coordinates ${\bf R}^{3\mid \gamma \mid -3}\ni {\bf \tilde w}_{(k)}{\bf % \rightarrow }(\underline{V},\xi )\in {\bf R}^3\times {\bf R}^{3\mid \gamma \mid -6}$, of type (A.2) in Appendix A. Then, in the new variables, $% m^{-1}(\sum_{n\in {\cal N}(\gamma )}\sum_{i=1}^{\gamma _n}m_n{\bf w}% _{n,i}-m_k{\bf w}_{k,\gamma _k})=\underline{V}$, and $T_\gamma ({\bf \tilde w% }_{(k)})=\frac m2\underline{V}^2+\xi ^2$ with $m:=\sum_{n=1}^N\gamma _nm_n-m_k$. A few simple manipulations show that on $\{{\bf w\in {R}}^{3\mid \gamma \mid }\mid {\bf w}_{k,\gamma _k}={\bf v}\}$, both $\Phi _\gamma $, given by (\ref{13}), and the relative energy $W_{r,\gamma }$ of {\it all} the particles in channel $\gamma $ (see Section 2), can be written in terms of $(\underline{V},\xi )\in {\bf R}^3\times {\bf R}^{3(\mid \gamma \mid -2)}$ as $$ \;\Phi _\gamma (t,{\bf w},{\bf x},{\bf v})_{\mid {\bf w}_{k,\gamma _k}={\bf v% }}=m_k({\bf x}^2+{\bf v}^2)+m\underline{V}^2 $$ $$ +2(1+t^2)\xi ^2+m[{\bf x}-(\underline{V}-{\bf v})t]^2,\eqno (B.3) $$ and% $$ W_{r,\gamma }({\bf w})_{\mid {\bf w}_{k,\gamma _k}={\bf v}}=\xi ^2+2^{-1}(m_k^{-1}+m^{-1})^{-1}(\underline{V}-{\bf v})^2.\eqno (B.4) $$ Then we choose $(\underline{V},\xi )$ as new integration variables for the integral upon $d{\bf \tilde w}_{(k)}$ in (\ref{14}) and we introduce (B.3) and (B.4) in (\ref{14}). One obtains that there exist two constants $c_0$ and $c$ such that% $$ \;\quad \Gamma _{k,\gamma }(t,{\bf v},{\bf x})\le c_0\int_{{\bf R}^3}d \underline{V}\exp (-c\underline{V}^2)\int_{{\bf R}^{3(\mid \gamma \mid -2)}}d\xi \exp (-c\xi ^2) $$ $$ \times \left\{ 1+[\xi ^2+(\underline{V}-{\bf v})^2]^{q/2}\right\} \int_0^tds\exp \left[ -c[x-(\underline{V}-{\bf v})s]^2\right] ,\eqno (B.5) $$ $t\ge 0,{\bf v},{\bf x}\in {\bf R}^3,q\in [0,1].$ Since for some constant $c_1>0$,% $$ \quad 1+[\xi ^2+(\underline{V}-{\bf v})^2]^{q/2}\le c_1.\,(1+\mid \xi \mid ^q)\,(1+\mid \underline{V}-{\bf v}\mid ^q),\eqno (B.6) $$ we introduce (B.6) in (B.5 ) and integrating with respect to $\xi $ we obtain% $$ \Gamma _{k,\gamma }(t,{\bf v},{\bf x})\le c_2\int_{{\bf R}^3}d\underline{V}% (1+\mid \underline{V}-{\bf v}\mid ^q)\exp (-c\underline{V}^2)\int_0^tds\exp \{-c[{\bf x}-(\underline{V}-{\bf v})s]^2\}. $$ with $c,c_2>0$, constants. This is exactly (B.1), concluding the proof. $% \Box $.\ \begin{thebibliography}{99} \bibitem{wa} C.S.G. Wang Chang, E. Uhlenbeck, and J. de Boer, in J. de Boer and G. E. Uhlenbeck Eds., Studies in Statistical Mechanics{\it ,} vol. 2, Part C, North Holland, Amsterdam (1964). \bibitem{lu} G. Ludwig and M. Heil, {\it Boundary layer theory with dissociation and ionization,} in Advances in Applied Mechanics vol. 6, Academic Press, New York (1960). \bibitem{ce} C. Cercignani, {\it The Boltzmann equation and its applications,} in Applied Mathematical Sciences, No. 67, Springer Verlag, Berlin (1988). \bibitem{ku} I. Kuscer, {\it Dissociation and recombination in an inhomogeneous gas,} Physica A (1991) v.{\bf 176,} p.542-556. \bibitem{be1} N. Bellomo and G. Toscani,{\it \ The Cauchy Problem for discrete Boltzmann equation with multiple collisions. Existence, uniqueness and stability}, Rapporto interno n. 5, Dipartimento di Matematica Politecnico di Torino, (1990). \bibitem{be2} N. Bellomo and S. Kawashima, {\it The discrete Boltzmann equation with multiple collisions. Global existence and stability in the initial value problem,} J. Math. Phys.(1990) ,v.{\bf 31, } p.245-253. \bibitem{ga} R. Gatignol and F. Coulouvrat, {\it Description hydrodynamique d'un gaz en th\'eorie cin\'etique discr\`ete: le mod\`ele g\'en\'eral,} C.R. Acad Sci. Paris S\'erie II, (1988) v.{\bf 306,} p.169-174. \bibitem{ga2} E. Gabetta, R. Monaco, {\it The discrete Boltzmann equations with bi-molecular chemical reactions,} Coll. Discrete models in fluid dynamics, Ser. Adv. Math. Appl. Sci., 2, World Scientific (1991). \bibitem{sp} G. Spiga {\it On extended kinetic theory with chemical reactions, }Coll. Nonlinear kinetic theory and mathematical aspects of hyperbolic systems, Ser. Addv. Math. Appl.Sci., 9, World Scientific, (1992.) \bibitem{be} N. Bellomo, A. Palczewski and G. Toscani , {\it Mathematical topics in nonlinear kinetic theory,} World Scientific, Singapore (1988). \bibitem{ai} I.\ Aizenmann, {\it A sufficient condition for avoidance of sets by measure preserving flows, }Duke Math.\ J., (1978) v. {\bf 45}., p. 809-814. \bibitem{gr} C. P. Gr\"unfeld, {\it On a class of kinetic equations for reacting gas mixtures with multiple collisions,} C.R. Acad. Sci. Paris Tome I, (1993) v.{\bf 316,} p.953 - 958. \end{thebibliography} \end{document}