Chapman–Enskog theory
   HOME

TheInfoList



OR:

Chapman–Enskog theory provides a framework in which equations of hydrodynamics for a gas can be derived from the
Boltzmann equation The Boltzmann equation or Boltzmann transport equation (BTE) describes the statistical behaviour of a thermodynamic system not in a state of equilibrium, devised by Ludwig Boltzmann in 1872.Encyclopaedia of Physics (2nd Edition), R. G. Lerne ...
. The technique justifies the otherwise phenomenological
constitutive relations In physics and engineering, a constitutive equation or constitutive relation is a relation between two physical quantities (especially kinetic quantities as related to kinematic quantities) that is specific to a material or substance, and app ...
appearing in hydrodynamical descriptions such as the
Navier–Stokes equations In physics, the Navier–Stokes equations ( ) are partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Anglo-Irish physicist and mathematician Geo ...
. In doing so, expressions for various transport coefficients such as
thermal conductivity The thermal conductivity of a material is a measure of its ability to conduct heat. It is commonly denoted by k, \lambda, or \kappa. Heat transfer occurs at a lower rate in materials of low thermal conductivity than in materials of high thermal ...
and
viscosity The viscosity of a fluid is a measure of its resistance to deformation at a given rate. For liquids, it corresponds to the informal concept of "thickness": for example, syrup has a higher viscosity than water. Viscosity quantifies the inte ...
are obtained in terms of molecular parameters. Thus, Chapman–Enskog theory constitutes an important step in the passage from a microscopic, particle-based description to a continuum hydrodynamical one. The theory is named for Sydney Chapman and David Enskog, who introduced it independently in 1916 and 1917.


Description

The starting point of Chapman–Enskog theory is the Boltzmann equation for the 1-particle distribution function f(\mathbf,\mathbf,t): : \frac+\mathbf\frac+\frac \cdot\frac=\hat f, where \hat is a nonlinear integral operator which models the evolution of f under interparticle collisions. This nonlinearity makes solving the full Boltzmann equation difficult, and motivates the development of approximate techniques such as the one provided by Chapman–Enskog theory. Given this starting point, the various assumptions underlying the Boltzmann equation carry over to Chapman–Enskog theory as well. The most basic of these requires a separation of scale between the collision duration \tau_ and the mean free time between collisions \tau_: \tau_ \ll \tau_. This condition ensures that collisions are well-defined events in space and time, and holds if the dimensionless parameter \gamma \equiv r_^3 n is small, where r_ is the range of interparticle interactions and n is the number density. In addition to this assumption, Chapman–Enskog theory also requires that \tau_ is much smaller than any ''extrinsic'' timescales \tau_. These are the timescales associated with the terms on the left hand side of the Boltzmann equation, which describe variations of the gas state over macroscopic lengths. Typically, their values are determined by initial/boundary conditions and/or external fields. This separation of scales implies that the collisional term on the right hand side of the Boltzmann equation is much smaller than the streaming terms on the left hand side. Thus, an approximate solution can be found from : \hat f = 0. It can be shown that the solution to this equation is a
Gaussian Carl Friedrich Gauss (1777–1855) is the eponym of all of the topics listed below. There are over 100 topics all named after this German mathematician and scientist, all in the fields of mathematics, physics, and astronomy. The English eponymo ...
: : f=n(\mathbf,t)\left( \frac\right)^ \exp \left -\frac \right where m is the molecule mass and k_B is the
Boltzmann constant The Boltzmann constant ( or ) is the proportionality factor that relates the average relative kinetic energy of particles in a gas with the thermodynamic temperature of the gas. It occurs in the definitions of the kelvin and the gas constant, ...
. A gas is said to be in ''local equilibrium'' if it satisfies this equation. The assumption of local equilibrium leads directly to the
Euler equations 200px, Leonhard Euler (1707–1783) In mathematics and physics, many topics are named in honor of Swiss mathematician Leonhard Euler (1707–1783), who made many important discoveries and innovations. Many of these items named after Euler include ...
, which describe fluids without dissipation, i.e. with thermal conductivity and viscosity equal to 0. The primary goal of Chapman–Enskog theory is to systematically obtain generalizations of the Euler equations which incorporate dissipation. This is achieved by expressing deviations from local equilibrium as a perturbative series in
Knudsen number The Knudsen number (Kn) is a dimensionless number defined as the ratio of the molecular mean free path length to a representative physical length scale. This length scale could be, for example, the radius of a body in a fluid. The number is name ...
\text, which is small if \tau_ \ll \tau_. Conceptually, the resulting hydrodynamic equations describe the dynamical interplay between free streaming and interparticle collisions. The latter tend to drive the gas ''towards'' local equilibrium, while the former acts across spatial inhomogeneities to drive the gas ''away'' from local equilibrium. When the Knudsen number is of the order of 1 or greater, the gas in the system being considered cannot be described as a fluid. To first order in \text one obtains the
Navier–Stokes equations In physics, the Navier–Stokes equations ( ) are partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Anglo-Irish physicist and mathematician Geo ...
. Second and third orders give rise, respectively, to the Burnett equations and super-Burnett equations.


Mathematical formulation

Since the Knudsen number does not appear explicitly in the Boltzmann equation, but rather implicitly in terms of the distribution function and boundary conditions, a dummy variable \varepsilon is introduced to keep track of the appropriate orders in the Chapman–Enskog expansion: : \frac+\mathbf\frac+\frac\cdot \frac=\frac \hat f. Small \varepsilon implies the collisional term \hat f dominates the streaming term \mathbf\frac+\frac\cdot\frac, which is the same as saying the Knudsen number is small. Thus, the appropriate form for the Chapman–Enskog expansion is : f=f^+\varepsilon f^+\varepsilon^2 f^+\cdots \ . Solutions that can be formally expanded in this way are known as ''normal'' solutions to the Boltzmann equation. This class of solutions excludes non-perturbative contributions (such as e^), which appear in boundary layers or near internal shock layers. Thus, Chapman–Enskog theory is restricted to situations in which such solutions are negligible. Substituting this expansion and equating orders of \varepsilon leads to the hierarchy : \begin J(f^,f^) &=0 \\ 2J(f^,f^) &=\left(\frac+\mathbf\frac+\frac\cdot \frac \right) f^ -\sum_^J(f^,f^), \qquad n > 0, \end where J is an integral operator, linear in both its arguments, which satisfies J(f,g) = J(g,f) and J(f,f) = \hatf. The solution to the first equation is a Gaussian: : f^=n'(\mathbf,t)\left( \frac\right)^\exp \left -\frac\right for some functions n'(\mathbf,t), \mathbf'_(\mathbf,t), and T'(\mathbf,t). The expression for f^ suggests a connection between these functions and the physical hydrodynamic fields defined as moments of f(\mathbf,\mathbf,t): : \begin n(\mathbf,t) &= \int f \, d\mathbf \\ n(\mathbf,t)v_0 (\mathbf,t) &= \int \mathbf f \, d\mathbf \\ n(\mathbf,t)T(\mathbf,t) &= \int \frac\mathbf^2 f \, d\mathbf. \end From a purely mathematical point of view, however, the two sets of functions are not necessarily the same for \varepsilon > 0 (for \varepsilon = 0 they are equal by definition). Indeed, proceeding systematically in the hierarchy, one finds that similarly to f^, each f^ also contains arbitrary functions of \mathbf and t whose relation to the physical hydrodynamic fields is ''a priori'' unknown. One of the key simplifying assumptions of Chapman–Enskog theory is to assume that these otherwise arbitrary functions can be written in terms of the exact hydrodynamic fields and their spatial gradients. In other words, the space and time dependence of f enters only implicitly through the hydrodynamic fields. This statement is physically plausible because small Knudsen numbers correspond to the hydrodynamic regime, in which the state of the gas is determined solely by the hydrodynamic fields. In the case of f^, the functions n'(\mathbf,t), \mathbf'_(\mathbf,t), and T'(\mathbf,t) are assumed exactly equal to the physical hydrodynamic fields. While these assumptions are physically plausible, there is the question of whether solutions which satisfy these properties actually exist. More precisely, one must show that solutions exist satisfying : \begin \int \sum_^\infty \varepsilon^n f^ \, d\mathbf= 0 = \int \sum_^\infty \varepsilon^f^\mathbf^2 \, d\mathbf \\ \int \sum_^\infty \varepsilon^n f^ v_i \, d\mathbf = 0, \qquad i \in \. \end Moreover, even if such solutions exist, there remains the additional question of whether they span the complete set of normal solutions to the Boltzmann equation, i.e. do not represent an artificial restriction of the original expansion in \varepsilon. One of the key technical achievements of Chapman–Enskog theory is to answer both of these questions in the positive. Thus, at least at the formal level, there is no loss of generality in the Chapman–Enskog approach. With these formal considerations established, one can proceed to calculate f^. The result is : f^=\left -\frac\left( \frac\right)^ \mathbf(\mathbf) \cdot \nabla \ln T - \frac \mathbb\mathbf_ \rightf^, where \mathbf(\mathbf) is a vector and \mathbb(\mathbf) a
tensor In mathematics, a tensor is an algebraic object that describes a multilinear relationship between sets of algebraic objects related to a vector space. Tensors may map between different objects such as vectors, scalars, and even other tensor ...
, each a solution of a linear inhomogeneous
integral equation In mathematics, integral equations are equations in which an unknown function appears under an integral sign. In mathematical notation, integral equations may thus be expressed as being of the form: f(x_1,x_2,x_3,...,x_n ; u(x_1,x_2,x_3,...,x_n) ...
that can be solved explicitly by a polynomial expansion. Here, the colon denotes the double dot product, \mathbb : \mathbb = \sum_i \sum_j T_T'_ for tensors \mathbb, \mathbb.


Predictions

To first order in the Knudsen number, the heat flux \mathbf = \frac \int f \mathbf^2 \mathbf \, d\mathbf is found to obey
Fourier's law of heat conduction Conduction is the process by which heat is transferred from the hotter end to the colder end of an object. The ability of the object to conduct heat is known as its ''thermal conductivity'', and is denoted . Heat spontaneously flows along a tem ...
, : \mathbf = -\lambda \nabla T, and the momentum-flux tensor \mathbf = m \int (\mathbf - \mathbf_0) (\mathbf - \mathbf_0)^\mathsf f \, \mathrm\mathbf is that of a
Newtonian fluid A Newtonian fluid is a fluid in which the viscous stresses arising from its flow are at every point linearly correlated to the local strain rate — the rate of change of its deformation over time. Stresses are proportional to the rate of chang ...
, : \mathbf = p \mathbb - \mu \left( \nabla \mathbf + \nabla \mathbf^T \right) + \frac\mu (\nabla \cdot \mathbf) \mathbb, with \mathbb the identity tensor. Here, \lambda and \mu are the thermal conductivity and viscosity. They can be calculated explicitly in terms of molecular parameters by solving a linear integral equation; the table below summarizes the results for a few important molecular models (m is the molecule mass and k_B is the Boltzmann constant). With these results, it is straightforward to obtain the Navier–Stokes equations. Taking velocity moments of the Boltzmann equation leads to the ''exact'' balance equations for the hydrodynamic fields n(\mathbf,t), \mathbf_0(\mathbf,t), and T(\mathbf,t): : \begin \frac+\nabla \cdot\left( n\mathbf_0\right) &= 0 \\ \frac+ \mathbf_0\cdot \nabla \mathbf_0-\frac+\frac\nabla \cdot \mathbf &= 0 \\ \frac+\mathbf_0\cdot \nabla T+\frac\left( \mathbf\nabla \mathbf_0+\nabla \cdot \mathbf\right) &= 0. \end As in the previous section the colon denotes the double dot product, \mathbb : \mathbb = \sum_i \sum_j T_T'_. Substituting the Chapman–Enskog expressions for \mathbf and \sigma, one arrives at the Navier–Stokes equations.


Comparison with experiment

An important prediction of Chapman–Enskog theory is that viscosity is independent of density (this can be seen for each molecular model in table 1, but is actually model-independent). This counterintuitive result traces back to
James Clerk Maxwell James Clerk Maxwell (13 June 1831 – 5 November 1879) was a Scottish mathematician and scientist responsible for the classical theory of electromagnetic radiation, which was the first theory to describe electricity, magnetism and li ...
, who inferred it in 1860 on the basis of more elementary kinetic arguments. It is well-verified experimentally for gases at ordinary densities. On the other hand, the theory predicts that \mu does depend on temperature. For rigid elastic spheres, the predicted scaling is \mu \propto T^, while other models typically show greater variation with temperature. For instance, for molecules repelling each other with force \propto r^ the predicted scaling is \mu \propto T^s, where s = 1/2 + 2/(\nu - 1). Taking s = 0.668, corresponding to \nu \approx 12.9, shows reasonable agreement with the experimentally observed scaling for helium. For more complex gases the agreement is not as good, most likely due to the neglect of attractive forces. Indeed, the Lennard-Jones model, which does incorporate attractions, can be brought into closer agreement with experiment (albeit at the cost of a more opaque T dependence; see the Lennard-Jones entry in table 1). Chapman–Enskog theory also predicts a simple relation between \lambda and \mu in the form \lambda = f \mu c_v, where c_v is the specific heat at constant volume and f is a purely numerical factor. For spherically symmetric molecules, its value is predicted to be very close to 2.5 in a slightly model-dependent way. For instance, rigid elastic spheres have f \approx 2.522, and molecules with repulsive force \propto r^ have f \approx 2.511 (the latter deviation is ignored in table 1). The special case of Maxwell molecules (repulsive force \propto r^) has f = 2.5 exactly.Chapman & Cowling, pp. 247 Since \lambda, \mu, and c_v can be measured directly in experiments, a simple experimental test of Chapman–Enskog theory is to measure f for the spherically symmetric
noble gases The noble gases (historically also the inert gases; sometimes referred to as aerogens) make up a class of chemical elements with similar properties; under standard conditions, they are all odorless, colorless, monatomic gases with very low ch ...
. Table 2 shows that there is reasonable agreement between theory and experiment.Chapman & Cowling p. 249


Extensions

The basic principles of Chapman–Enskog theory can be extended to more diverse physical models, including gas mixtures and molecules with internal degrees of freedom. In the high-density regime, the theory can be adapted to account for collisional transport of momentum and energy, i.e. transport over a molecular diameter ''during'' a collision, rather than over a mean free path (''in between'' collisions). Including this mechanism predicts a density dependence of the viscosity at high enough density, which is also observed experimentally. One can also carry out the theory to higher order in the Knudsen number. In particular, the second-order contribution f^ has been calculated by Burnett. In general circumstances, however, these higher-order corrections may not give reliable improvements to the first-order theory, due to the fact that the Chapman–Enskog expansion does not always converge. (On the other hand, the expansion is thought to be at least asymptotic to solutions of the Boltzmann equation, in which case truncating at low order still gives accurate results.) Even if the higher order corrections do afford improvement in a given system, the interpretation of the corresponding hydrodynamical equations is still debated.


See also

*
Transport phenomena In engineering, physics, and chemistry, the study of transport phenomena concerns the exchange of mass, energy, charge, momentum and angular momentum between observed and studied systems. While it draws from fields as diverse as continuum mecha ...
* Kinetic theory of gases *
Boltzmann equation The Boltzmann equation or Boltzmann transport equation (BTE) describes the statistical behaviour of a thermodynamic system not in a state of equilibrium, devised by Ludwig Boltzmann in 1872.Encyclopaedia of Physics (2nd Edition), R. G. Lerne ...
*
Navier–Stokes equations In physics, the Navier–Stokes equations ( ) are partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Anglo-Irish physicist and mathematician Geo ...
*
Viscosity The viscosity of a fluid is a measure of its resistance to deformation at a given rate. For liquids, it corresponds to the informal concept of "thickness": for example, syrup has a higher viscosity than water. Viscosity quantifies the inte ...
*
Thermal conductivity The thermal conductivity of a material is a measure of its ability to conduct heat. It is commonly denoted by k, \lambda, or \kappa. Heat transfer occurs at a lower rate in materials of low thermal conductivity than in materials of high thermal ...


Notes


References

The classic monograph on the topic: * Contains a technical introduction to normal solutions of the Boltzmann equation: * {{DEFAULTSORT:Chapman-Enskog theory Statistical mechanics