In this article an abstract discrete system is considered, consisting of an arbitrary, finite numer of particles modelled as mathematical points to which analytic funcitons of time are attached. We prove that a space-time average of these analytic functions can be defined, satisfying a relation of the same form with the balance equations in continuum mechanics.
Authors
C. Vamos Tiberiu Popoviciu, Institute of Numerical Analysis, Romanian Academy
A. Georgescu
N. Suciu Tiberiu Popoviciu, Institute of Numerical Analysis, Romanian Academy
Keywords
Paper coordinates
C. Vamoş, A. Georgescu, N. Suciu (1996), Balance equations for a finite number of particles, Stud. Cerc. Mat., 48(1-2), 115-127.
References
see the expansion block below.
PDF
soon
About this paper
Journal
Stud. Cerc. Mat.
Publisher Name
DOI
Not available yet.
Print ISSN
Not available yet.
Online ISSN
Not available yet.
[1] A. Akhiezer and P. Peletminski, Les methodes de la physique statistique, Mir. Moscou, 1980.
[2] R. Balescu, Equilibium and Noneequilibrium Statistical Mechanics, Wiley, New York, 1975.
[3] C. Cercignani, Mathematical Metods in Kinetic Theory, Plenum, New York, 1990.
[4] S. Chapman and T. G. Cowling, The Mathematical Theory of Non-uniform Gases. University Press, Cambridge, 1953.
[5] E. G. D. Cohen, Fifty years of kinetic theory. Physica A 194 (1993), 229-257.
[6] P. Glansdorff and I. Prigogine, Thermodynamic Theory of Structure, Stability and Fluctuations. Wiley, London, 1971.
[7] J. K. Irfinb and J. G. Kirkwood, The statistical mechanical theory of transport processes, IV. The equations of hydrodynamics. J. Chem. Phys. 18 (1950) 817-40.
[8] A. Kolmogorov and S. Fomine, Elements de la theorie des fonctions et de l’analyse fonctionelle. Mir. Moscou, 1974.
[9] J. Muller, Thermodynamics. Pitman, London, 1985.
[10] A. Sveshnikov and A. Tikhonov, The Theory of Functions of a Complex Variable, Mir. Moscou, 1978.
[11] C. Truesdell and R. A. Toupin, The Classical Field Theories. In: Handbuch der Physik, III, Part. I, ed. S. Flugge, Springer, Berlin, 1960.
[12] C. Vamoș, A. Georgescu, N. Suciu and I. Turcu, Balance equations for physical systems with corpuscular structure. In press at Physica A.
Paper in HTML form
1996b Suciu-Vamos-Georgescu - Balance equations for a finite number of particles
ACADEMIA ROMÂNĂ
STUDII SI CERCETARI MATEMATICE
(MATHEMATICAL REPORTS)
EXTRAS1-2
TOMUL 48
1996
ianuarie - aprilie
BALANCE EQUATIONS FOR A FINITE NUMBER OF PARTICLES
C. VAMOŞ, A. GEORGESCU and N. SUCIU
Abstract
In this article an abstract discrete system is considered, consisting of an arbitrary, finite number of particles modelled as mathematical points to which analytic functions of time are attached. We prove that a space-time average of these analytic functions can be defined, satisfying a relation of the same form with the balance equations in continuum mechanics.
1. INTRODUCTION
The balance equations are postulated relations for fundamental physical quantities (mass, momentum, energy, entropy etc.) valid for all continuous media [11]. We take over the differential expression of the balance equations from [9]. Let Psi\Psi be a physical quantity additive with respect to space, associated to a continuous medium. That is, there exists a function psi\psi of space ( vec(r))(\vec{r}) and time (t)(t), called the volume density of Psi\Psi, such that, for any volume VV, the integral int_(V)psid bar(r)\int_{V} \psi \mathrm{~d} \bar{r} represents the amount of Psi\Psi contained in VV. The differential form of the balance equation at a regular point (i.e. without shocks or other discontinuities) is
where del_(t)\partial_{t} is the temporal derivative, del_(alpha)\partial_{\alpha} is the derivative with respect to the alpha\alpha component of bar(r),Phi_(alpha)\bar{r}, \Phi_{\alpha} is the alpha\alpha component of the flux density of Psi,nu_(alpha)\Psi, \nu_{\alpha} is the alpha\alpha component of the velocity, pp is the production density of Psi\Psi due to interior processes and ss is the supply density of Psi\Psi controlled from the exterior of VV. The quantities bar(Phi),p\bar{\Phi}, p and ss are expressed by the constitutive equations characterizing the considered material.
The statistical method of derivation of the balance equations for a macroscopic physical system from its microscopic structure was initiated by Boltzmann (e.g. [1] - [3]). This method relies on the evolution equation of the probability density in the phase space for the system consisting of all the microscopic components of the physical system (Liouville equation). Even for the simple case of the ideal gas, because of mathematical difficulties, the derivation of the balance equations and constitutive equations is possible only using certain hypotheses, approximations and simplifications [4]. So far, these results have been extended to hard sphere fluids, also a very idealized molecular model [5].
For a more complicated microscopic structure, the existence of the balance equations is implicity postulated and the problem of the statistical mechanics is reduced to the calculation of the constitutive equations.
In this article we show that the existence of mathematical relations of the form (1) can be proved in the more general framework of the kinematic description for the microscopic evolution of an arbitrary corpuscular physical system. The position and motion of the microscopic particles can be modelled as mathematical points, and the physical quantities characterizing the particles (mass, momentum, acceleration, kinetic momentum etc.) can be modelled as functions of time. Thus, the differential equations (the dynamic system) corresponding to the microscopic evolution of the corpuscular physical system are not known, but the existence of its solution is postulated.
In the following we shall consider an abstract mathematical model for the 'kinematic microscopic description discussed above. This mathematical discrete system consists of an arbitrary, finite number of mathematical points to which arbitrary functions of time are associated and it has an abstract nature because there is no physical specification for these time functions. Certainly, for a given corpuscular physical system, the functions of time can represent the mass of the particles, their momentum or any other physical quantity, but in general no specific physical quatity is assigned to the mathematical points. To be more concise, we shall use the name of "particles" or "material points" for these mathematical points together with the assigned abstract functions of time.
The particles can appear or disappear as a result of certain instantaneous processes. Every particle has a temporal interval of existence which can be different from the temporal interval over which we study the discrete system. We assume that a single existence interval corresponds to a given particle, i.e. its disappearance at one time precludes its reappearance. Even if a particle of the same type appears later on, it is considered as a new particle. For the abstract discrete system we do not impose any connection between the disappearance of some particles and the appearance of others. Certainly, in the case of processes like chemical reactions, such connections exist and mass, momentum and energy must be conserved.
The only dynamical requirement is that the evolution of the abstract discrete system (both the variation of the particles position and the associated functions) should be given by analitic functions of time. Under these circumstances we shall prove that a space-time average of the arbitrary functions of time has a.e. continuous partial derivatives. We emphasize that the averaging is an ordinary mathematical one, not a statistical average on an infinite ensemble of identical copies of the physical system. Although these averages preserve the discontinuities associated to the particles as discontinuity surfaces, they satisfy a relation of the form (1). To eliminate and posibility of confusion, we shall call the relations obtained for the abstract discrete system, the "discrete analogue" of the relations of continuum mechanics having a similar form.
We shall apply these results to a Hamiltonian system formed by a single type of particles which neither disappear, nor appear. Every particle is characterized by
a constant mass. For this case we shall write the discrete analogue of the balance equations for mass, momentum and energy. We shall briefly discuss the relation with the approach in nonequilibrium statistical mechanics.
2. THE DISCRETE ANALOGUE OF CONTINUOUS FIELDS
We study the evolution during the temporal interval I=[0,T]subRI=[0, T] \subset \mathbb{R} of an abstract discrete system consisting of NN particles (i.e. mathematical points with the assigned functions of time). We denote by I_(i)=[t_(i)^(+),t_(i)^(-)]sub II_{i}=\left[t_{i}^{+}, t_{i}^{-}\right] \subset I the existence inteval of the ii-th particle ( 1 <= i <= N1 \leq i \leq N ). Obviously 0 <= t_(i)^(+) < T0 \leq t_{i}^{+}<T and 0 < t_(i)^(-) <= T0<t_{i}^{-} \leq T. If I_(i)=II_{i}=I, then the ii-th particle exists over the whole interval II. If I_(i)!=II_{i} \neq I for one particle at least, then there are moments when the number of particles is smaller than the total number of particles NN. Denoting by n(t)n(t) the number of the particles existing at the moment t in It \in I, we have n(t) <= Nn(t) \leq N for each t in It \in I. The equality holds only if I_(i)=II_{i}=I for all i <= Ni \leq N, i.e. if no particles are generated or destroyed over the interval II.
Let varphi_(i):I rarrR\varphi_{i}: I \rightarrow \mathbb{R} be an arbitrary function of time characterizing the ii-th particle. If I_(i)!=II_{i} \neq I, then varphi_(i)(t)=0\varphi_{i}(t)=0 for all t in I\\I_(i)t \in I \backslash I_{i}. We assume that the restriction varphi_(i)∣I_(i)\varphi_{i} \mid I_{i} can be represented as a Taylor series, i.e. it is an analytic function. In the interval I_(i)I_{i} the function varphi_(i)\varphi_{i} may take any real value, including zero (e.g. the velocity components of a motionless particle). Hence varphi_(i)\varphi_{i} is discontinuous at t_(i)^(+)t_{i}^{+}and t_(i)^(-)t_{i}^{-}if varphi_(i)(t_(i)^(+))!=0\varphi_{i}\left(t_{i}^{+}\right) \neq 0 and varphi_(i)(t_(i)^(-))!=0\varphi_{i}\left(t_{i}^{-}\right) \neq 0, respectively. Similarly, the derivatives of varphi_(i)\varphi_{i} at t_(i)^(+)t_{i}^{+}and t_(i)^(-)t_{i}^{-}may be continuous or discontinuous. The alpha\alpha components of the radius vector vec(r)_(i),x_(alpha i):I rarrR(alpha=1,2,3)\vec{r}_{i}, x_{\alpha i}: I \rightarrow \mathbb{R}(\alpha=1,2,3), and of the velocity bar(xi)_(i),xi_(alpha i):I rarrR(alpha=1,2,3)\bar{\xi}_{i}, \xi_{\alpha i}: I \rightarrow \mathbb{R}(\alpha=1,2,3) may be treated as particular cases of functions varphi_(i)\varphi_{i}. The functions x_(alpha i)x_{\alpha i} and xi_(alpha i)\xi_{\alpha i} supply a kinematic description of the motion of the discrete system.
Definition. For two arbitrary positive real parameters tau < T//2\tau<T / 2 and aa, we define a function D_(varphi):R^(3)xx(tau,T-tau)rarrRD_{\varphi}: \mathbb{R}^{3} \times(\tau, T-\tau) \rightarrow \mathbb{R} as
where V=4pia^(3)//3V=4 \pi a^{3} / 3 is the volume of the open sphere of center bar(r)\bar{r} and radius aa denoted by S( vec(r),a)S(\vec{r}, a) and H^(+)H^{+}is the left continuous Heaviside function.
Since H^(+)(a^(2)-( bar(r)_(i)(t^('))-( bar(r)))^(2))H^{+}\left(a^{2}-\left(\bar{r}_{i}\left(t^{\prime}\right)-\bar{r}\right)^{2}\right) vanishes if the ii-th particle is located outside the sphere S( vec(r),a)S(\vec{r}, a) and varphi_(i)(t^('))\varphi_{i}\left(t^{\prime}\right) vanishes if t^(')in II_(i)t^{\prime} \in I I_{i}, then a nonvanishing contribution to D_(varphi)D_{\varphi} is due only to particles which lie in S( vec(r),a)S(\vec{r}, a) over the interval (t-tau,t+tau)(t-\tau, t+\tau). Therefore D_(varphi)( vec(r),t)\mathrm{D}_{\varphi}(\vec{r}, t) characterizes the mean distribution of varphi\varphi about the point of radius vector vec(r)\vec{r} at the moment tt, and it is a spatial average on the sphere S( vec(r),a)S(\vec{r}, a) and a temporal
one over the interval (t-tau,t+tau)(t-\tau, t+\tau). Obviously, it also depends on the parameters tau\tau and aa, but we are not interesed in this dependence.
The function H^(+)(a^(2)-( vec(r)_(i)(t^('))-( vec(r)))^(2))H^{+}\left(a^{2}-\left(\vec{r}_{i}\left(t^{\prime}\right)-\vec{r}\right)^{2}\right) in (2) takes only the values 0 and 1 . The jumps occur when the ii-th particle enters or leaves the open sphere S( vec(r),a)S(\vec{r}, a). These moments are among the solutions u_(i)u_{i} of the equation
where |h_(i)(( vec(r)),t)|^(1//2)\left|h_{i}(\vec{r}, t)\right|^{1 / 2} is the distance at the moment tt between the i -th particle and the surface del S( vec(r),a)\partial S(\vec{r}, a) of the sphere S( vec(r),a)S(\vec{r}, a). Since x_(alpha i)x_{\alpha i}, and hence h_(i)h_{i}, are analytic functions with respect to u_(i)u_{i}, and I_(i)I_{i} is a closed interval, then either equation (3) has a finite number of solutions or h_(i)h_{i} vanishes identically ([10], p 78 ). In the latter case the particle moves along the surface of S( vec(r),a)S(\vec{r}, a) and does not enter the sphere, hence H^(+)(a^(2)-( bar(r)_(i)(t)-( bar(r)))^(2))H^{+}\left(a^{2}-\left(\bar{r}_{i}(t)-\bar{r}\right)^{2}\right) is identically zero and has no jumps. Since bar(r)_(i)(u_(i))\bar{r}_{i}\left(u_{i}\right) is a known function, then the isolated zeros of (3) are implicit functions u_(i)( bar(r))u_{i}(\bar{r}). The implicit function theorem can be applied only at interior points and it does not ensure the existence of u_(i)( vec(r))u_{i}(\vec{r}) for u_(i)=t_(i)^(+-)u_{i}=t_{i}^{ \pm}, i.e. vec(r)in del S( vec(r)_(i)(t_(i)^(+-)),a)\vec{r} \in \partial S\left(\vec{r}_{i}\left(t_{i}^{ \pm}\right), a\right). This case will be discussed separately. For u_(i)in(t_(1)^(+),t_(1)^(-))u_{i} \in\left(t_{1}^{+}, t_{1}^{-}\right), if
where x_(alpha)x_{\alpha} are the components of vec(r)\vec{r}. According to (4), the function u_(i)( vec(r))u_{i}(\vec{r}) is not differentiable at the points of the discriminant surface of the family {del S( vec(r)_(i)(t),a);t inI_(i)}\left\{\partial S\left(\vec{r}_{i}(t), a\right) ; t \in I_{i}\right\}.
We denote the moments when the ii-th particle enters (leaves) the sphere S( bar(r),a)S(\bar{r}, a) by t_(i)^(+) < u_(i1)^(') < u_(i2)^(') < dots < u_(in^('))^(') < t_(i)^(-)(t_(i)^(+) < u_(i1)^('') < u_(i2)^('') < dots < u_(in^(''))^('') < t_(i)^(-))t_{i}^{+}<u_{i 1}^{\prime}<u_{i 2}^{\prime}<\ldots<u_{i n^{\prime}}^{\prime}<t_{i}^{-}\left(t_{i}^{+}<u_{i 1}^{\prime \prime}<u_{i 2}^{\prime \prime}<\ldots<u_{i n^{\prime \prime}}^{\prime \prime}<t_{i}^{-}\right). Since the sphere S( bar(r),a)S(\bar{r}, a) is open, H^(+)(a^(2)-( vec(r)_(i)(t)-( bar(r)))^(2))H^{+}\left(a^{2}-\left(\vec{r}_{i}(t)-\bar{r}\right)^{2}\right) as a function of tt is left (right) continuous when the particle enters (leaves). Hence for t inI_(i)t \in I_{i}, we have
where H^(-)H^{-}is the right continuous Heaviside jump function. The first term in the right-hand side of (6) vanishes if the ii-th particle is generated in the exterior of S( vec(r),a)S(\vec{r}, a) and equals 1 otherwise. If u_(i)=t_(i)^(+)u_{i}=t_{i}^{+}, then the particle can only enter the sphere after its generation, hence u_(i1)^(')=t_(i)^(+)u_{i 1}^{\prime}=t_{i}^{+}and relation (6) holds. If u_(i)=t_(i)^(-)u_{i}=t_{i}^{-}, then the particle could only leave the sphere before its disappearance, hence u_(in^(''))^('')=t_(i)^(-)u_{i n^{\prime \prime}}^{\prime \prime}=t_{i}^{-} and relation (6) holds again. So (6) contains all the possible situations if we take t_(i)^(+) <= u_(i1)^(')t_{i}^{+} \leq u_{i 1}^{\prime} and u_(in^(''))^('') <= t_(i)^(-)u_{i n^{\prime \prime}}^{\prime \prime} \leq t_{i}^{-}. The following notation will be used:
Proposition 1. The function D_(varphi)D_{\varphi} defined by (2) has partial derivatives continuous a.e. in R^(3)xx(tau,T-tau)\mathbb{R}^{3} \times(\tau, T-\tau).
Proof. To study the differentiability of D_(varphi)D_{\varphi}, we consider the function g_(i):R^(3)xx(tau,T-tau)rarrRg_{i}: \mathbb{R}^{3} \times(\tau, T-\tau) \rightarrow \mathbb{R}
is a continuous function, except a finite number of jump discontinuities {t_(i)^(+),t_(i)^(-)}uuU_(i)^(')uuU_(i)^('')\left\{t_{i}^{+}, t_{i}^{-}\right\} \cup U_{i}^{\prime} \cup U_{i}^{\prime \prime}. Hence G_(i)G_{i} is Riemann integrable and g_(i)g_{i} has partial derivative with respect to tt a.e. in (tau,T-tau)(\tau, T-\tau), equal to
The discontinuities of del_(t)g_(i)\partial_{t} g_{i} with respect to ( vec(r),t\vec{r}, t ) are related to those of G_(i)G_{i}. From (9) it follows that G_(i)G_{i} is discontinuous when varphi_(i)\varphi_{i} is discontinuous and H^(+)H^{+} nonvanishing, or conversely, when H^(+)H^{+}is discontinuous and varphi_(i)\varphi_{i} nonvanishing. In the first case the i -th particle appears or disappears in S( vec(r),a)S(\vec{r}, a), i.e. t=t_(i)^(+-)t=t_{i}^{ \pm}and vec(r)in S( vec(r)_(i)(t_(i)^(+-)),a)\vec{r} \in S\left(\vec{r}_{i}\left(t_{i}^{ \pm}\right), a\right), and in the second case the ii-th particle lies in the surface of S( vec(r),a)S(\vec{r}, a), i.e. t inI_(i)t \in I_{i} and vec(r)in del S( vec(r)_(i)(t),a)\vec{r} \in \partial S\left(\vec{r}_{i}(t), a\right).
Hence the derivative (10) is not continuous over
The set Omega_(i)^(')\Omega_{i}^{\prime} has null Lebesgue measure in R^(3)xx(tau,T-tau)\mathbb{R}^{3} \times(\tau, T-\tau), hence del_(t)g_(i)\partial_{t} g_{i} is a.e. continuous.
In the appendix we show that the derivative of g_(i)g_{i} with respect to x_(alpha)x_{\alpha} exists and is continuous a.e. in R^(3)xx(tau,T-tau)\mathbb{R}^{3} \times(\tau, T-\tau). Using definition (2) and relations (10), (11), (A3) and (A4), it follows the a.e. continuity of the partial derivatives of D_(varphi)D_{\varphi}, given by
where (del_(t)D_(varphi))_(g)\left(\partial_{t} D_{\varphi}\right)_{g} is determined by the particles generation.
Proof. We use a theorem stating that every function with bounded variation may be uniquely split into a sum of two functions: one continuous and a jump function ([8], p331). We apply this theorem to G_(i)G_{i} given by (9) considered as a function of tt. But except a finite number of jump discontinuities, G_(i)G_{i} is analytic on II and then its continuous part G_(i)^(')G_{i}^{\prime} is also absolutely continuous. Hence we may write G_(i)=G_(i)^(')+G_(i)^('')G_{i}=G_{i}^{\prime}+G_{i}^{\prime \prime}, where G_(i)^('')G_{i}^{\prime \prime} is the jump function. Replacing this relation in (12), it follows that del_(t)D_(varphi)\partial_{t} D_{\varphi} can also be written as a two term sum
It contains the discontinuous variations of G_(i)G_{i} during the temporal interval [t-tau,t+tau][t-\tau, t+\tau]. As proved in the preceding section, del_(t)D_(varphi)\partial_{t} D_{\varphi} exists if G_(i)G_{i} is not discontinuous at t+taut+\tau and t-taut-\tau (see expression (11)), therefore we consider only the jumps occuring at the interior points of [t-tau,t+tau][t-\tau, t+\tau], i.e. in ( t-tau,t+taut-\tau, t+\tau ). From ( 9 ) it follows that such a variation can take place if the particle is generated inside the sphere S( bar(r),a)S(\bar{r}, a) during the temporal interval ( t-tau,t+taut-\tau, t+\tau ). Hence the jump of G_(i)G_{i} is equal to
where W_(i)^(')=U_(i)^(')nn(t-tau,t+tau)W_{i}^{\prime}=U_{i}^{\prime} \cap(t-\tau, t+\tau) and W_(i)^('')=U_(i)^('')nn(t-tau,t+tau)W_{i}^{\prime \prime}=U_{i}^{\prime \prime} \cap(t-\tau, t+\tau). The sign of varphi_(i)(u)\varphi_{i}(u) is positive (negative) if the particle enters (leaves) the sphere S( bar(r),a)S(\bar{r}, a), and it is given by the sign of the expression -( bar(r)_(i)(u)-( bar(r)))* bar(xi)_(i)(u)-\left(\bar{r}_{i}(u)-\bar{r}\right) \cdot \bar{\xi}_{i}(u) which is proportional to the interior normal component of vec(xi)_(i)\vec{\xi}_{i} to the surface of S( bar(r),a)S(\bar{r}, a) at the moment uu. Hence we may use a single sum in (19) if we denote U_(i)=W_(i)^(')uuW_(i)^('')U_{i}=W_{i}^{\prime} \cup W_{i}^{\prime \prime}. Replacing (19) in (18) we obtain
Since varphi_(i)(t)\varphi_{i}(t) and xi_(alpha i)(t)\xi_{\alpha i}(t) are analytic functions with respect to time, then their product is also analytic and relation (20) can be written as
The physical quantity varphi_(i) vec(xi)\varphi_{i} \vec{\xi} represents the transport of varphi\varphi by the ii-th particle, and the space-time average D_( vec(varphi) vec(xi))D_{\vec{\varphi} \vec{\xi}} represents the mean flux of varphi\varphi.
Relation (14) follows from (15), (17) and (22).
In contrast to balance equation (1), relation (14) does not contain a quantity equivalent to the velocity vec(v)\vec{v}. The velocity is not a volume density, but an average quantity. To define a discrete analogue, we must divide D_(varphi)D_{\varphi} by the number of the particles contributing to D_(varphi)D_{\varphi}. Let D_(1)D_{1} be the density D_(varphi)D_{\varphi} corresponding to varphi_(i)(t)=1\varphi_{i}(t)=1 for all t inI_(i)t \in I_{i} and i <= Ni \leq N. Since D_(1)D_{1} characterizes the average number of particles per unit volume, the discrete average of varphi\varphi is defined as
if D_(1)( vec(r),t)!=0D_{1}(\vec{r}, t) \neq 0 and it vanishes if D_(1)( vec(r),t)=0D_{1}(\vec{r}, t)=0. It is easy to show that bar(bar(varphi))= bar(varphi), bar(varphi_(1)+varphi_(2))= bar(varphi_(1))+ bar(varphi_(2))\overline{\bar{\varphi}}=\bar{\varphi}, \overline{\varphi_{1}+\varphi_{2}}=\overline{\varphi_{1}}+\overline{\varphi_{2}} and bar(lambda varphi)=lambda bar(varphi)\overline{\lambda \varphi}=\lambda \bar{\varphi}, where lambda\lambda is a real function of bar(r)\bar{r} and tt.
The mean motion of the particle is given by the discrete average of the velocity bar(bar(xi))\overline{\bar{\xi}} with the components bar(xi)_(alpha)\bar{\xi}_{\alpha}. To introduce bar(xi)_(alpha)\bar{\xi}_{\alpha} in (14), we write
This is the discrete analogue of the balance equation (1).
4. THE HAMILTONIAN SYSTEMS
In this section we consider a Hamiltonian system consisting of a single type of particles. The abstract particles considered till now become real particles with mass mm, satisfying the principles of classical mechanics. Obviously, the mass mm is constant in time and the same for all the particles. Since we have a single type of particles which are not generated or destroyed, the generating term (21) vanishes.
The relation (24) for mass is obtained if varphi_(i)(t)=m\varphi_{\mathrm{i}}(t)=m for all t inI_(i)t \in I_{i} and i <= Ni \leq N. Then D_(varphi)=D_(m)=mD_(1)D_{\varphi}=D_{m}=m D_{1} is the discrete analogue of the mass density. Moreover, bar(varphi)=m\bar{\varphi}=m, bar(Phi)_(m)^(')=0,varphi^(˙)_(i)=0\bar{\Phi}_{m}^{\prime}=0, \dot{\varphi}_{i}=0 and (24) becomes the discrete analogue of the continuity equation
For the alpha\alpha component of momentum we have varphi_(i)=p_(alpha i)=mxi_(alpha i)\varphi_{i}=p_{\alpha i}=m \xi_{\alpha i} and D_(p_(alpha))=D_(mxi_(alpha))=D_(m) bar(xi)_(alpha)D_{p_{\alpha}}=D_{m \xi_{\alpha}}=D_{m} \bar{\xi}_{\alpha}. The discrete analogue of the kinetic part of the flux density
takes the form of a symmetric tensor
The derivative varphi^(˙)_(i)\dot{\varphi}_{i} is the a component of the force vec(f)_(i)\vec{f}_{i} acting on the ii-th particle and relation (24) becomes
Making additional hypotheses on the interaction between particles, one can prove that equation (27) is the discrete analogue of the momentum equation in continuum mechanics [7].
Choosing as physical quantity the kinetic energy of the particles varphi_(i)=E_(i)=(1)/(2)mxi_(i)^(2)\varphi_{i}=E_{i}=\frac{1}{2} m \xi_{i}^{2}, we obtain
D_(E)=(1)/(2)mD_([ vec(xi)+( vec(xi)- vec(xi))]^(2))=(1)/(2)mD_( vec(xi)^(2))+(1)/(2)mD_( bar(vec(xi))( bar(xi)- bar(xi)))+(1)/(2)mD_(( vec(xi)- bar(xi))^(2))D_{E}=\frac{1}{2} m D_{[\vec{\xi}+(\vec{\xi}-\vec{\xi})]^{2}}=\frac{1}{2} m D_{\vec{\xi}^{2}}+\frac{1}{2} m D_{\overline{\vec{\xi}}(\bar{\xi}-\bar{\xi})}+\frac{1}{2} m D_{(\vec{\xi}-\bar{\xi})^{2}}
where we used the linearity of D_(varphi)D_{\varphi} with respect to varphi\varphi, i.e. D_(varphi_(1)+varphi_(2))=D_(varphi_(1))+D_(varphi_(2))D_{\varphi_{1}+\varphi_{2}}=D_{\varphi_{1}}+D_{\varphi_{2}}. The second term of the expression vanishes D_( vec(xi),( vec(xi)- vec(xi)))= bar(bar(xi))*D_( vec(xi)- vec(xi))= bar(bar(xi))*(D_( vec(xi))-D_( vec(xi)))=0D_{\vec{\xi},(\vec{\xi}-\vec{\xi})}=\overline{\bar{\xi}} \cdot D_{\vec{\xi}-\vec{\xi}}=\overline{\bar{\xi}} \cdot\left(D_{\vec{\xi}}-D_{\vec{\xi}}\right)=0. The last term is the discrete analogue of the kinetic energy density of the thermal motion, since (1)/(2)m( bar(xi)- bar(bar(xi)))^(2)\frac{1}{2} m(\bar{\xi}-\overline{\bar{\xi}})^{2} represents the kinetic energy of the relative motion of the ii-th particle with respect to the mean motion of the particles in the sphere S( bar(r),a)S(\bar{r}, a), over (t-tau,t+tau)(t-\tau, t+\tau). We denote this term by
epsi=(1)/(2)mD_(( vec(xi)- vec(xi))^(2))=(1)/(2)sum_(alpha=1)^(3)sigma_(alpha alpha)^(')\varepsilon=\frac{1}{2} m D_{(\vec{\xi}-\vec{\xi})^{2}}=\frac{1}{2} \sum_{\alpha=1}^{3} \sigma_{\alpha \alpha}^{\prime}
the last equality following directly from (26). Similarly
In addition to relations (25), (27) and (28), discrete analogues of balance equations for any physical quantity are possible.
5. CONCLUSION
Relation (24) is not a balance equation, but an identity of the same form with a balance equation. It has been derived under general conditions, for an arbitrary, finite number of mathematical points to which analytic functions of time were attached. Due to this very general approach, the results can be applied to a large number of corpuscular physical systems. For example, the discrete analogue of the balance equation (24) is valid for an arbitrary physical quantity, and for an arbitrary number of particles (even very small). Also, since the dynamical equations for the microscopic evolution have not been used explicity, relation (24) holds for any microscopic interaction forces satisfying the analycity condition. In this article we have considered the case of the Hamiltonian systems consisting of a single type of particles which can not be generated or destroyed.
To transform the function D_(varphi)D_{\varphi} defined by (2) into a continuous field, and relation (24) into a balance equation, a statistical average on an ensemble formed by a very large number of identical copies of the considered corpuscular system is needed. Although, if the number of particles contributing to the value of D_(varphi)D_{\varphi} is large enough, then D_(varphi)D_{\varphi} approximates closely the continuous field corresponding to the physical quantity varphi\varphi. That is, if the physical system satisfies the local equilibrium principle [6], then the parameters aa and tau\tau can be chosen so that the particles lying in the sphere S( vec(r),a)S(\vec{r}, a) over the interval (t-tau,t+tau)(t-\tau, t+\tau) should form a near-equilibrium thermodynamical system [12]. Obviously, in this case the total number of particles NN can no longer be arbitrary, but it must be large enough to ensure the validity of the thermodynamical limit.
The balance equation in continuum mechanics can also be obtained as the limit for a rarr0a \rightarrow 0 and tau rarr0\tau \rightarrow 0 of the statistical average of relation (24). Both methods to obtain the balance equation (1) from (24) will be discuss in another article.
APPENDIX
Here we study the differentiability with respect to vec(r)\vec{r} of the function g_(i)g_{i} defined by (8). Although (6) holds only for t inI_(i)t \in I_{i}, it may be substituted in (8) because varphi_(i)\varphi_{i}
vanishes for t in I\\I_(i)t \in I \backslash I_{i}, and we obtain *
where U_(i)^(')U_{i}^{\prime} and U_(i)^('')U_{i}^{\prime \prime} are defined in (7). First, we consider 2tau < t_(i)^(-)-t_(i)^(+)2 \tau<t_{i}^{-}-t_{i}^{+}. The following cases are possible:
(a) t <= t_(i)^(+)-taut \leq t_{i}^{+}-\tau. Then (t-tau,t+tau)nnI_(i)=O/(t-\tau, t+\tau) \cap I_{i}=\varnothing and varphi_(i)\varphi_{i} vanishes in the integration intervals in (A1), such that g_(i)( vec(r),t)=0g_{i}(\vec{r}, t)=0 for all vec(r)inR^(3)\vec{r} \in \mathbb{R}^{3}.
(b) t in(t_(i)^(+)-tau,t_(i)^(+)+tau]t \in\left(t_{i}^{+}-\tau, t_{i}^{+}+\tau\right]. Then (t-tau,t+tau)nnI_(i)=[t_(i)^(+),t+tau)(t-\tau, t+\tau) \cap I_{i}=\left[t_{i}^{+}, t+\tau\right) and the integral in (A1) have the same limits. The first term in (A1) depends on vec(r)\vec{r} through the function H^(+)(a^(2)-( bar(r)_(i)(t_(i)^(+))-( bar(r)))^(2))H^{+}\left(a^{2}-\left(\bar{r}_{i}\left(t_{i}^{+}\right)-\bar{r}\right)^{2}\right) which can take only the values 0 and 1 . Hence, when this function is continuous with respect to bar(r)\bar{r}, its derivative exists and equals zero. Then the first term in (A1) is not differentiable if H^(+)(a^(2)-( bar(r)_(i)(t_(i)^(+))-( bar(r)))^(2))H^{+}\left(a^{2}-\left(\bar{r}_{i}\left(t_{i}^{+}\right)-\bar{r}\right)^{2}\right) is discontinuous, i.e. vec(r)in del S( vec(r)_(i)(t_(i)^(+)),a)\vec{r} \in \partial S\left(\vec{r}_{i}\left(t_{i}^{+}\right), a\right). The others terms in (A1) depend on vec(r)\vec{r} through the moments uu defined by (3). These terms are not differentiable either if uu is not differentiable (i.e. relation (4) is not satisfied) of if the moments uu coincide with the integration limits (i.e. vec(r)in del S( vec(r)_(i)(t_(i)^(+)),a)\vec{r} \in \partial S\left(\vec{r}_{i}\left(t_{i}^{+}\right), a\right) or vec(r)in del S( vec(r)_(i)(t+tau),a)\vec{r} \in \partial S\left(\vec{r}_{i}(t+\tau), a\right) ). In this case the integration intervals have discontinuous variations with respect to vec(r)\vec{r}.
(c) t in(t_(i)^(+)+tau,t_(i)^(-)-tau)t \in\left(t_{i}^{+}+\tau, t_{i}^{-}-\tau\right). Then (t-tau,t+tau)nnI_(i)=(t-tau,t+tau)(t-\tau, t+\tau) \cap I_{i}=(t-\tau, t+\tau) and the integrals with u <= t-tauu \leq t-\tau are equal to the integral of the first term. Using the expression
where W_(2)^(')=U_(i)^(')nn(t-tau,t_(i)^(-))W_{2}^{\prime}=U_{i}^{\prime} \cap\left(t-\tau, t_{i}^{-}\right)and W_(2)^('')=U_(i)^('')nn(t-tau,t_(i)^(-)]W_{2}^{\prime \prime}=U_{i}^{\prime \prime} \cap\left(t-\tau, t_{i}^{-}\right]. As for (b), the first term is not differentiable if vec(r)in del S( vec(r)_(i)(t-tau),a)\vec{r} \in \partial S\left(\vec{r}_{i}(t-\tau), a\right) and the other terms if vec(r)in del S( vec(r)_(i)(t+tau),a)\vec{r} \in \partial S\left(\vec{r}_{i}(t+\tau), a\right), or if relation (4) is not satisfied.
(d) t in[t_(i)^(-)tau,t_(i)^(-)+tau)t \in\left[t_{i}^{-} \tau, t_{i}^{-}+\tau\right). Then (t-tau,t+tau)nnI_(i)=(t-tau,t_(i)^(-)](t-\tau, t+\tau) \cap I_{i}=\left(t-\tau, t_{i}^{-}\right]and the expression for g_(i)( vec(r),t)g_{i}(\vec{r}, t) is identic with (A2), except the upper integration limit is t_(i)^(-)t_{i}^{-}. So g_(i)( vec(r),t)g_{i}(\vec{r}, t) is not differentiable if vec(r)in del S( vec(r)_(i)(t-tau),a)\vec{r} \in \partial S\left(\vec{r}_{i}(t-\tau), a\right) or vec(r)in del S( vec(r)_(i)(t_(i)^(-)),(a^(˙)))\vec{r} \in \partial S\left(\vec{r}_{i}\left(t_{i}^{-}\right), \dot{a}\right) or (4) is not satisfied.
(e) t >= t_(i)^(-)+taut \geq t_{i}^{-}+\tau. Then (t-tau,t+tau)nnI_(i)=O/(t-\tau, t+\tau) \cap I_{i}=\varnothing and g_(i)( vec(r),t)=0g_{i}(\vec{r}, t)=0.
If tau < t_(i)^(-)-t_(i)^(+) < 2tau\tau<t_{i}^{-}-t_{i}^{+}<2 \tau, then the possible cases for(A1) are t <= t_(i)^(+)-tau,t in(t_(i)^(+)-tau,t_(i)^(-)-tau)t \leq t_{i}^{+}-\tau, t \in\left(t_{i}^{+}-\tau, t_{i}^{-}-\tau\right), t in[t_(i)^(-)-tau,t_(i)^(+)+tau],t in(t_(i)^(+)+tau,t_(i)^(-)+tau)t \in\left[t_{i}^{-}-\tau, t_{i}^{+}+\tau\right], t \in\left(t_{i}^{+}+\tau, t_{i}^{-}+\tau\right) and t >= t_(i)^(-)+taut \geq t_{i}^{-}+\tau and the discussion is similar. Finally, if t_(i)^(-)-t_(i)^(+) <= taut_{i}^{-}-t_{i}^{+} \leq \tau, then other five intervals for tt exist. Taking into account that t in(tau,T-tau)t \in(\tau, T-\tau), the set where the function g_(i)g_{i} is not differentiable with respect to vec(r)\vec{r} is
{:[(A3)Omega_(i)^('')={( bar(r)_(,)t)∣t in(t_(i)^(+)-tau,t_(i)^(+)+tau]nn(tau,T-tau)" and "( vec(r))in del S( bar(r)_(i)(t_(i)^(+)),a)}uu],[ uu{(( vec(r)),t)∣t in[t_(i)^(-)-tau,t_(i)^(-)+tau)nn(tau,T-tau)" and "( vec(r))in del S( bar(r)_(i)(t_(i)^(-)),a)}uu],[ uu{( vec(r)_(,)t)∣t in(t_(i)^(+)+-tau,t_(i)^(-)+-tau)nn(tau,T-tau)" and "( vec(r))in del S( bar(r)_(i)(t∓tau),a)}uu],[ uu{(( vec(r)),t)∣t in(tau,T-tau)," exists "t^(')in(t-tau,t+tau)nnI_(i):}" such that "],[{:( vec(r))in del S( bar(r)_(i)(t^(')),a)" and "( vec(r)_(i)(t^('))-( vec(r)))* vec(xi)_(i)(t^('))=0}]:}\begin{align*}
\Omega_{i}^{\prime \prime} & =\left\{\left(\bar{r}_{,} t\right) \mid t \in\left(t_{i}^{+}-\tau, t_{i}^{+}+\tau\right] \cap(\tau, T-\tau) \text { and } \vec{r} \in \partial S\left(\bar{r}_{i}\left(t_{i}^{+}\right), a\right)\right\} \cup \tag{A3}\\
& \cup\left\{(\vec{r}, t) \mid t \in\left[t_{i}^{-}-\tau, t_{i}^{-}+\tau\right) \cap(\tau, T-\tau) \text { and } \vec{r} \in \partial S\left(\bar{r}_{i}\left(t_{i}^{-}\right), a\right)\right\} \cup \\
& \cup\left\{\left(\vec{r}_{,} t\right) \mid t \in\left(t_{i}^{+} \pm \tau, t_{i}^{-} \pm \tau\right) \cap(\tau, T-\tau) \text { and } \vec{r} \in \partial S\left(\bar{r}_{i}(t \mp \tau), a\right)\right\} \cup \\
& \cup\left\{(\vec{r}, t) \mid t \in(\tau, T-\tau), \text { exists } t^{\prime} \in(t-\tau, t+\tau) \cap I_{i}\right. \text { such that } \\
& \left.\vec{r} \in \partial S\left(\bar{r}_{i}\left(t^{\prime}\right), a\right) \text { and }\left(\vec{r}_{i}\left(t^{\prime}\right)-\vec{r}\right) \cdot \vec{\xi}_{i}\left(t^{\prime}\right)=0\right\}
\end{align*}
This is a set of null Lebesgue measure in R^(3)xx(tau,T-tau)\mathbb{R}^{3} \times(\tau, T-\tau).
Only the terms in (A1) which contain uu in the integration interval have a nonvanishing contribution to the derivative of g_(i)g_{i} with respect to x_(alpha)x_{\alpha}, denoted by del_(alpha)g_(i^(**))\partial_{\alpha} g_{i^{*}}. Using (5) and taking into account that the sign of the terms in (A1) coincides with the sign of the expression -( bar(r)_(i)(u)-( bar(r)))* bar(xi)_(i)(u)-\left(\bar{r}_{i}(u)-\bar{r}\right) \cdot \bar{\xi}_{i}(u), we obtain
where U_(i)=(U_(i)^(')uuU_(i)^(''))nn(t-tau,t+tau)U_{i}=\left(U_{i}^{\prime} \cup U_{i}^{\prime \prime}\right) \cap(t-\tau, t+\tau). It is obvious that del_(alpha)g_(i)\partial_{\alpha} g_{i} is continuous over R^(3)xx(tau,T-tau)\\Omega_(i)^('')\mathbb{R}^{3} \times(\tau, T-\tau) \backslash \Omega_{i}^{\prime \prime}.
Acknowledgement. We are grateful to Dr. I. Turcu for several fruitful discussions.
REFERENCES
A. Akhiezer and P. Péletminski, Les méthodes de la physique statistique. Mir, Moscou, 1980.
R. Balescu, Equilibrium and Nonequilibrium Statistical Mechanics. Wiley, New York, 1975.
C. Cercignani, Mathematical Methods in Kinetic Theory. Plenum, New York, 1990.
S. Chapman and T. G. Cowling, The Mathematical Theory of Non-uniform Gases. University Press, Cambridge, 1953.
E. G. D. Cohen, Fifty years of kinetic theory. Physica A. 194 (1993) 229-257.
P. Glansdorff and I. Prigogine, Thermodynamic Theory of Structure, Stability and Fluctuations. Wiley, London, 1971.
J. K. Irving and J. G. Kirkwood, The statistical mechanical theory of transport processes, IV. The equations of hydrodynamics. J. Chem. Phys. 18 (1950) 817-40.
A. Kolmogorov and S. Fomine, Éléments de la théorie des fonctions et de l'analyse fonctionelle. Mir, Moscou, 1974.
J. Müller, Thermodynamics. Pitman, London, 1985.
A. Sveshnikov and A. Tikhonov, The Theory of Functions of a Complex Variable. Mir, Moscou, 1978.
C. Truesdell and R. A. Toupin, The Classical Field Theories. In: Handbuch der Physik, III, Part 1, ed. S. Flugge, Springer, Berlin, 1960.
C. Vamoş, A. Georgescu, N. Suciu and I. Turcu, Balance equations for physical systems with corpuscular structure. In press at Physica A.
Received March 1, 1993
Revised form March 13, 1995
Romanian Academy
Institute of Applied Mathematics P. O. Box 1-24, 70700 Bucharest I