
Citation: Massimiliano Giona, Luigi Pucci. Hyperbolic heat/mass transport and stochastic modelling - Three simple problems[J]. Mathematics in Engineering, 2019, 1(2): 224-251. doi: 10.3934/mine.2019.2.224
[1] | La-Su Mai, Suriguga . Local well-posedness of 1D degenerate drift diffusion equation. Mathematics in Engineering, 2024, 6(1): 155-172. doi: 10.3934/mine.2024007 |
[2] | Diogo Gomes, Julian Gutierrez, Ricardo Ribeiro . A mean field game price model with noise. Mathematics in Engineering, 2021, 3(4): 1-14. doi: 10.3934/mine.2021028 |
[3] | Franco Flandoli, Eliseo Luongo . Heat diffusion in a channel under white noise modeling of turbulence. Mathematics in Engineering, 2022, 4(4): 1-21. doi: 10.3934/mine.2022034 |
[4] | Andrea Lo Giudice, Roberto Nuca, Luigi Preziosi, Nicolas Coste . Wind-blown particulate transport: A review of computational fluid dynamics models. Mathematics in Engineering, 2019, 1(3): 508-547. doi: 10.3934/mine.2019.3.508 |
[5] | Gabriele Grillo, Giulia Meglioli, Fabio Punzo . Global existence for reaction-diffusion evolution equations driven by the $ {\text{p}} $-Laplacian on manifolds. Mathematics in Engineering, 2023, 5(3): 1-38. doi: 10.3934/mine.2023070 |
[6] | Camilla Nobili . The role of boundary conditions in scaling laws for turbulent heat transport. Mathematics in Engineering, 2023, 5(1): 1-41. doi: 10.3934/mine.2023013 |
[7] | Connor Mooney, Arghya Rakshit . Singular structures in solutions to the Monge-Ampère equation with point masses. Mathematics in Engineering, 2023, 5(5): 1-11. doi: 10.3934/mine.2023083 |
[8] | Luca Formaggia, Alessio Fumagalli, Anna Scotti . A multi-layer reactive transport model for fractured porous media. Mathematics in Engineering, 2022, 4(1): 1-32. doi: 10.3934/mine.2022008 |
[9] | Alessia Nota, Juan J. L. Velázquez . Homoenergetic solutions of the Boltzmann equation: the case of simple-shear deformations. Mathematics in Engineering, 2023, 5(1): 1-25. doi: 10.3934/mine.2023019 |
[10] | Siddhartha Mishra . A machine learning framework for data driven acceleration of computations of differential equations. Mathematics in Engineering, 2019, 1(1): 118-146. doi: 10.3934/Mine.2018.1.118 |
Thermal and heat transport problems have recently become a fertile and exciting field of theoretical and applied research due to advances in phononics and in the miniaturization of microelectromechanical systems [1,2]. New effects have been discovered (e.g. thermal rectification) and new devices tested (thermal diodes and transistors) [3]. On equal footing, the analysis of microscale thermal problem has questioned the application of classical phenomenological laws, such as the Fourier law, and more generally the formulation of transport equations in out-of-equilibrium conditions. There is enough experimental and theoretical work indicating that the classical Fourier constitutive equation, leading to parabolic heat transport models, is not completely adequate for processes evolving at small length and short time scales [4,5].
The natural physical candidate for a generalization of the Fourier law is represented by the hyperbolic models defined by constitutive equations with memory for the heat flux. The archetype of this class of models is the Cattaneo heat transfer equation [6,7],
Following parallel pathways, mass (solute) transport in 17ic systems displays phenomenologically clear deviations for a pure Fickian behavior, referred to as "case Ⅱ diffusion" in the mass transport literature [16,17]. The origin of this anomalous behavior stems from the swelling of the 17ic matrix, causing the coupling between mass (solute) transport and viscoelastic stress relaxation [18,19]. The use of hyperbolic and Cattaneo-based models encompasses also biomedical applications of controlled release of active principles [20].
In a more general perspective, the mathematical structure of the Cattaneo equation has been embedded in a consistent thermodynamic theory of non-equilibrium phenomena thanks to the seminal work by Müller and Ruggeri [21] that is commonly referred to as "extended theory of thermodynamics". Extended thermodynamics generalizes the classical approach of irreversible thermodynamics [22] by including, in the definition of the thermodynamic variables, the explicit contribution of the fluxes that vanish in equilibrium conditions. This approach has been further elaborated and extended by the contribution of Jou, Casas-Vazquez, Lebon, any many others [23,24,25].
There is however a formal bottleneck in all the hyperbolic formulations of thermodynamic and transport models grounded on the Cattaneo equation deriving from the fact that the latter does not preserve positivity (of the local concentrations) in space dimensions greater than one. This has been reported in [26] by considering the Green function for the two-dimensional Cattaneo heat equation.
It should be clearly stated that this does not inficiate neither the reach of extended thermodynamic theories nor the validity of many of the results found within this approach. Similar problems involving the deprecated occurrence of negative density values arise in the application of Grad's 13 moment expansion, and are ultimately a consequence of the approximations (essentially a truncated power-series expansion) underlying these approaches [27]. However, the positivity issue remains as a standing problem requiring a formal solution in the development of the theory.
Recently, in order to overcome the positivity issue, a stochastic approach to hyperbolic transport problems has been proposed, applicable in any space dimension, via the concept of Generalized Poisson-Kac (GPK) processes [28,29,30,31]. These developments originate from the work by Mark Kac that in a series of lectures, (dated 1956 and subsequently reprinted in 1974), showed that the one-dimensional Cattaneo equation represents the evolution equation for the probability density function associated with a simple stochastic differential equation driven by Poissonian fluctuations [32,33]. The trajectories of the process introduced by Kac are almost everywhere smooth, contrarily to Wiener fluctuations, usually considered in statistical physics, that are characterized by almost nowhere differential trajectories possessing fractal character [34,35]. The transport equations originating for GPK theory are indeed hyperbolic and involve only the first-order derivatives with respect to time and space coordinates, but require a vector-valued system of partial probability density functions [28,29,30,31]. This vector-valued description of the concentration, parametrized with respect to the state of the stochastic perturbation, physically corresponds to a spinorial description of the concentration fields [36], and is conceptually similar to the 4-wave formulation of the Dirac's equation compared to its non-relativistic counterpart (the Schrödinger equation). It is convenient to refer to it as the "partial-wave formulation" of a hyperbolic transport problem possessing finite propagation velocity (see Section 2). The importance of this kind of models has been envisaged by Rosenau in an illuminating article on kinetic theory [37].
Albeit the stochastic theory indicates that the original Cattaneo equation and the stochastic partial-wave formulation are equivalent in one spatial dimension (see Section 2 for details), there are indeed subtle issues that arise whenever transport problems in bounded domains (intervals) are considered, associated with the setting of proper boundary conditions accounting for the wave-like propagation of the partial concentration waves [38]. While the Cattaneo model admits a stochastic explanation, at least for one-dimensional spatial problems, the Guyer-Krumhansl model does not correspond to any underlting stochastic dynamics, and even in one-dimensional problems the associated Green function may attain negative values [39,40]. For this reason this model and its generalizations are not considered in the present work.
Due to the practical interest in the development of hyperbolic transport models at nanoscale, and their theoretical relevance (as only a hyperbolic model can be relativistically covariant, i.e., consistent with the space-time structure emerging from special relativity [21,41]), this article analyzes the mathematical description and some physical implications of the partial wave representation of hyperbolic transport models of heat and mass transport, focusing on the physical advantages of this approach compared with the classical one based exclusively on the evolution of the overall concentration/temperature and on the connection between finite propagation and the flux regularity. One-dimensional paradigmatic problems are considered, since only in this case the partial-wave representation can be compared with the corresponding Cattaneo equation for the overall concentration/temperature field (see Section 5).
It should be mentioned that some of the results obtained may have also some interest in the formulation of transport equations (continuous hydrodynamic limit) of anomalous transport processes, e.g. derived from the Continuous Time Random Walk paradigm [42,43,44], as they suggest that the stochastic spinorial (partial-wave) description of a process may capture finer details that cannot be enucleated when solely the evolution equation for the overall concentration is considered.
The article is organized as follows. Section 2 defines the setting of the problem and the different representations of a linear hyperbolic transport equation in one spatial dimension, based either on the Cattaneo equation for the overall concentration or on the partial-wave formulation derived from the undelying stochastic dynamics. Section 3 develops the closed-form expression for the spinorial Green function in the partial-wave representation. Section 4 analyzes heat transport problems arising from boundary layer theory in order to show how the occurrence of a finite propagation velocity regularizes the behavior of interfacial fluxes still maintaining the long-term large-distance properties that can be derived from the corresponding parabolic models. Finally, Section 5 addresses the simplest conceivable problem in heat transport: the stationary temperature distribution in a material slab kept at two different temperatures at the endpoints. Albeit its apparent "triviality", the analysis of this problem is physically enlightning, as it indicates the relevance of the partial-wave formulation, and the possible occurrence of negative temperature values in particular (and extreme) situations, that cannot be revealed using an approach exclusively based on the overall concentration (temperature) fields. The case of boundary conditions on the fluxes, corresponding in parabolic schemes to the Neumann boundary conditions, has been analyzed in [38] within the partial-wave formulation of hyperbolic transport models. The analysis is not repeated here and the reader is referred to [38] for details.
Consider the simplest stochastic motion
dx(t)=b(−1)χ(t,λ)dt | (2.1) |
where
Its statistical description involves two partial probability densities
∂p+(x,t)∂t=−b∂p+(x,t)∂x−λ[p+(x,t)−p−(x,t)]∂p−(x,t)∂t=b∂p−(x,t)∂x+λ[p+(x,t)−p−(x,t)] | (2.2) |
By analogy, it is possible to infer from this class of stochastic dynamics undulatory transport models for mass, momentum and energy [31]. Throughout this article we refer mainly to the heat transfer case, but the analysis of mass transport problems would be completely analogous.
Eq. (2.2) suggests that a one-dimensional heat-transfer model possessing finite propagation velocity, and fulfilling the positivity requirement (the temperature, viewed as absolute temperature, cannot attain negative values) would require two scalar fields
∂T+(x,t)∂t=−b∂T+(x,t)∂x−λ[T+(x,t)−T−(x,t)]∂T−(x,t)∂t=b∂T−(x,t)∂x+λ[T+(x,t)−T−(x,t)] | (2.3) |
with the same meaning for
T(x,t)=T+(x,t)+T−(x,t) | (2.4) |
It should be stressed that the physical meaning of the two temperatures
Considering absolute temperatures expressed in Kelvin, it follows that
T±(x,t)≥0 | (2.5) |
that henceforth will be referred to as the positivity requirement for the two partial heat waves. With reference to the waves
Jq(x,t)=b[T+(x,t)−T−(x,t)] | (2.6) |
Alternatively, given
T±(x,t)=12[T(x,t)±Jq(x,t)b] | (2.7) |
The temperature
∂T(x,t)∂t=−∂Jq(x,t)∂x | (2.8) |
where the heat flux is defined by the constitutive equation
12λ∂Jq(x,t)∂t+Jq(x,t)=−D0∂T(x,t)∂x | (2.9) |
where
12λ∂2T(x,t)∂t2+∂T(x,t)∂t=D0∂2T(x,t)∂x2 | (2.10) |
Alternatively, one can explicit the heat flux from eq. (2.9), namely
Jq(x,t)=e−t/τcJq(x,0)−D0∫t0e−(t−θ)/τc∂T(x,θ)∂xdθ=e−t/τcJq(x,0)−D(t)∗∂T(x,t)∂x | (2.11) |
where
∂T(x,t)∂t=D(t)∗∂2T(x,t)∂x2−e−t/τc∂Jq(x,0)∂x | (2.12) |
that is formally analogous to the convolution-type equation used for describing anomalous transport problems using fractional operators (Riemann-Liouville derivatives) [43,44,45].
Therefore,
In the Poisson-Kac formulation of the problem, the basic (primitive) quantities are the two partial heat waves
T±(x,0)=T±,0(x)≥0 | (2.13) |
where
In the Cattaneo formulation, it is natural to express the initial condition in two ways: either (ⅰ) by fixing the initial temperature profile
Jq(x,0)=Jq(x∗,0)−∫xx∗θ0(ξ)dξ | (2.14) |
where
Jq(x,0)=−∫x−∞θ0(ξ)dξ | (2.15) |
However, the situation is different in bounded domains
As regards the boundary conditions, each transport formulation implies some natural setting of the boundary conditions. For example, in the Poisson-Kac formulation, this implies to fix the values either of
With respect to all the other formulations, the Poisson-Kac description, involving the system of two partial waves
T+(x,t)+T−(x,t)≥0 | (2.16) |
as in the Cattaneo or in the memory-kernel formulation. The effect of this more restrictive requirement will become clear in Section 5, in connection with the simplest problem of stationary heat conduction in a finite domain.
In this Section, we consider the propagation of the Poisson-Kac equations (2.2) over the real line,
In point of fact, the latter (partial-wave) representation is physically interesting from several points of view:
● it enables a fully local description of the transport problem, in that solely the values of
● imbedding Poisson-Kac processes in the Minkowski space-time of special relativity (in the present case when only a single spatial coordinate is considered), the partial probability waves transform as probabilistic spinorial quantities [36], and the associated Green function can be therefore referred to as the spinorial Green functions of the problem.
By considering that the spinorial Green function has never been explicited and physically discussed, it justifies the content of the present Section.
In deriving the analytical expression for the spinorial Green function
Consider the Cattaneo formulation associated with the Poisson-Kac process (2.1)-(2.2) for
p(x,0)=f(x),∂p(x,t)∂t|t=0=g(x) | (3.1) |
and with regularity conditions at infinity. The general solution of this problem admits a closed-form solution [48]
p(x,t)=K1(x,t)∗f(x)+K2(x,t)∗g(x) | (3.2) |
where "*" indicates here convolution with respect to the spatial coordinate,
Since
f(x)=p0+(x)+p0−(x),g(x)=b[dp0−(x)dx−dp0+(x)dx] | (3.3) |
In deriving the second eq. (3.3), the extension of the balance equation at
J(x,t)=−∫x−∞∂p(η,t)∂tdη=−∫x−∞dη∫∞−∞∂K1(η−ξ,t)∂tf(ξ)dξ−∫x−∞dη∫∞−∞∂K2(η−ξ,t)∂tg(ξ)dξ | (3.4) |
and consequently the flux
J(x,t)=H1(x,t)∗f(x)+H2(x,t)∗g(x) | (3.5) |
where
Hh(x,t)=−∫x−∞∂Kh(η,t)∂tdηh=1,2 | (3.6) |
Substituting the expression for
p(x,t)=K1(x,t)∗[p0+(x)+p0−(x)]+bK2(x,t)∗[dp0−(x)dx−dp0+(x)dx] | (3.7) |
and analogously for
K2(x,t)∗dp0−(x)dx=∫∞−∞K2(x−ξ,t)dp0−(ξ)dξdξ=∫∞−∞∂[K2(x−ξ)p0−(ξ)]∂ξdξ⏟=0−∫∞−∞∂K2(x−ξ,t)∂ξp0−(ξ)dξ=∫∞−∞∂K2(x−ξ,t)∂xp0−(ξ)dξ=∂K2(x,t)∂x∗p0−(x) | (3.8) |
and the evolution equations for
(p(x,t)J(x,t))=(K1(x,t)−∂K2(x,t)/∂xH1(x,t)−∂H2(x,t)/∂x)∗(p0(x)J0(x)) | (3.9) |
The expression for
−∂H2(x,t)∂x=∂∂x∫x−∞∂K2(η,t)∂tdη=∂K2(x,t)∂t | (3.10) |
and is explicitly reported in the Appendix. In turn, it is much more cumbersome to derive the expression for
To begin with, observe that the flux
∂2J(x,t)∂t2+2λ∂J(x,t)∂t=D0∂2J(x,t)∂x2 | (3.11) |
Next, consider two transport problems, differing from each other in the initial conditions, as depicted in Figure 1:
● Case (1): Symmetric initial conditions for the partial probability waves, i.e.,
p0+(x)=p0−(x)=δ(x)2 | (3.12) |
which implies
p0(x)=δ(x),J0(x)=0 | (3.13) |
● Case (2): Antisymmetric initial conditions for the partial probability waves, i.e.,
p0+(x)=−p0−(x)=δ(x)2 | (3.14) |
and therefore,
p0(x)=0,J0(x)b=δ(x) | (3.15) |
Let
p(1)(x,t)=K1(x,t) | (3.16) |
In the antisymmetric case, the solution
p(2)(x,t)=J(1)(x,t)b=−b∂K2(x,t)∂x∗[p0+(x)+p0−(x)]=−b∂K2(x,t)∂x | (3.17) |
But from eq. (3.9), substituting the initial conditions for Case (1), one obtains that
J(1)(x,t)=H1(x,t) | (3.18) |
From eqs. (3.17)-(3.18) the expression for
H1(x,t)=−b2∂K2(x,t)∂x | (3.19) |
Therefore, expressing
(p+(x,t)+p−(x,t)p+(x,t)−p−(x,t))=(K1−b∂K2/∂x−b∂K2/∂x∂K2/∂t)∗(p0+(x)+p0−(x)p0+(x)−p0−(x)) | (3.20) |
in which the matrix-valued kernel is expressed in terms of
p(x,t)=G(x,t)∗p0(x),p(x,t)=(p+(x,t)p−(x,t)),p0(x)=(p0+(x)p0−(x)) | (3.21) |
follows
G(x,t)=12(K1−2b∂xK2+∂tK2K1−∂tK2K1−∂tK2K1+2b∂xK2+∂tK2) | (3.22) |
where
G(x,t)=G(δ)(x,t)+G(c)(x,t) | (3.23) |
where
G(δ)(x,t)=e−λt(δ(x−bt)00δ(x+bt)) | (3.24) |
and
G(c)(x,t)=λe−λt2bη(bt−|x|)((t+x/b)I1(λz)/zI0(λz)I0(λz)(t−x/b)I1(λz)/z) | (3.25) |
where
Expressed with respect to the partial probability waves, the spinorial Green function
G(c)1,1(x,t)=A(x,t)G(x,t;b,λ),G(c)2,2(x,t)=A(x,t)G(x,t;−b,λ) | (3.26) |
where only the sign of the velocity
The analytical expressions (3.21)-(3.25) obtained for
To begin with, consider balanced initial conditions, namely
p±(x,t)=e−λt2δ(x∓bt)+λe−λt4bη(bt−|x|)[(t±xb)I1(λz)z+I0(λz)] | (3.27) |
Let
Figure 3 reports the comparison of the continuous (smooth) part of the partial-wave profiles and of the overall density
In both cases the agreement between stochastic simulations and the analytical expression is excellent, and the small fluctuations at short time scales (panels (a) in Figures 3-4) derive from the relatively small number
In this Section we consider simple boundary-layer problems approached within the hyperbolic model described in Section 2, in order to show that the constraint of a finite propagation velocity, characteristic of eq. (2.3), regularizes the singularities often appearing in boundary-layer analysis associated with the interfacial fluxes.
Consider the conduction problem eq. (2.3) on the semi-infinite line
T(x,0)=T+(x,0)=T−(x,0)=0 | (4.1) |
while the wall temperature at
T(0,t)=T+(0,t)+T−(0,t)=T0 | (4.2) |
This problem can be conveniently approached in the Laplace domain, indicating with
dˆJq(x,s)dx=−sˆT(x,s)dˆT(x,s)dx=−(sb2+2λb2)ˆJq(x,s) | (4.3) |
where
d2ˆT(x,s)dx2=μ2(s)ˆT(x,s),μ(s)=(s2b2+sD0)1/2 | (4.4) |
where
ˆq(x,s)=1se−μ(s)x=1sexp[−xb√(s+b22D0)2−b44D20] | (4.5) |
Considering that [50]
L−1[e−α√(s+c)2−c2]=e−αcδ(t−α)+αce−ctI1(c√t2−α2)√t2−α2η(t−α) | (4.6) |
where, as in the previous Section,
q(x,t)=e−xλ/bη(t−x/b)+xλbη(t−x/b)∫tx/be−λτI1(λ√τ2−x2/b2)√τ2−x2/b2dτ | (4.7) |
Figure 5 panels (a) and (b) depict the normalized thermal front
Observe that the Poisson-Kac (Cattaneo) model is characterized by a discontinuity of temperature profile at the moving front edge
∂T(x,t)∂t=D0∂2T(x,t)∂x2 | (4.8) |
equipped with the same initial and boundary conditions, that in the present case is given by
Next, consider the interfacial flux. From the second equation (4.3) and from (4.5) it follows that
ˆJq(x,s)=T0√s2b2+sD0e−μ(s)x | (4.9) |
so that the Laplace transform of the normalized interfacial flux
ˆj0(s)=1√s2b2+sD0=b√(s+b22D0)2−b44D20 | (4.10) |
the inverse Laplace transform of which takes the expression [50]
j0(t)=be−b2t/2D0I0(b2t2D0) | (4.11) |
where
● Since
J0(t=0)=Jq(0,0)=bT0 | (4.12) |
i.e., the interfacial flux in a transport model possessing finite propagation velocity is always a smooth and bounded function of time, even at
j0(t)=√D0πt | (4.13) |
and displays the characteristic power-law singularity
● From the asymptotic expansion of the Bessel function for large arguments
j0(t)=√D0πt[1+O(2D0b2t)] | (4.14) |
implying that, for any values of
● In a heat transport model possessing finite propagation velocity the invariant rescaling,
The regularization effects induced by a finite propagation velocity, shown above with the aid of a simple example, arise also for more general boundary-layer problems. Indeed this is a generic property, dictated by the fact that the fluxes in this class of models are necessarily bounded due to the constitutive equation (2.6) whenever the the initial temperature field
In higher dimensional problems, the analogy between the Cattaneo equation and stochastic models of transport possessing finite propagation velocity breaks down [29], but the qualitative result of a flux regularization applies also to the Generalized Poisson-Kac (GPK) formulation of the (heat) transport equations. For details on the GPK formulation of transport models see [31].
Below, we address another classical two-dimensional problem that, under some simplifying assumptions, involves a one-dimensional formulation of the stochastic perturbation. Consider a two-dimensional straight channel of width
∂T(x,y,t)∂t=−vx(y)∂T(x,y,t)∂x=D0∂2T(x,y,t)∂y2 | (4.15) |
equipped with some initial condition
T|x=0=T0,T|y=0,W=Tw | (4.16) |
In the Leveque boundary-layer problem one is mainly interested to the stationary heat flux at the solid walls, that is related to the width of the thermal boundary layer that develops along the channel for axial lengthscales shorther than the width of the channel itself.
The corresponding Poisson-Kac (Cattaneo) formulation of the problem involves two partial temperature waves
∂T+(x,y,t)∂t=−vx(y)∂T+(x,y,t)∂x−b∂T+(x,y,t)∂y−λ[T+(x,y,t)−T−(x,y,t)]∂T−(x,y,t)∂t=−vx(y)∂T−(x,y,t)∂x+b∂T−(x,y,t)∂y+λ[T+(x,y,t)−T−(x,y,t)] | (4.17) |
We consider a shear flow,
j0(x)=Ax1/3,A=[∫∞0e−η3/9dη]−1 | (4.18) |
Also in this case, the Poisson-Kac solution possesses the same qualitative features envisaged in the previous boundary-layer problem, namely: (ⅰ) the finite propagation velocity regularizes the behaviour of the flux at the interface, and (ⅱ) the asymptotics (in the present case, the large-distance properties) of the Poisson-Kac solution converges to that of the corresponding (limit) parabolic transport model.
Consider the simplest heat transfer model on the interval
T(x)|x=0=T0,T(x)|x=L=TL | (5.1) |
Without loss of generality, suppose
T(x)=T0+(TL−T0)xL | (5.2) |
as one would obtain by enforcing the Fourier law with a constant heat conductivity.
Let us analyze the same problem within the Poisson-Kac formulation. This formulation differs from the other two in that, from physical reasons, a detailed positivity requirement should be enforced on each
dT+(x)dx=dT−(x)dx=−λb2J0=−B | (5.3) |
where
T+(x)=A−Bx,T−(x)=C−Bx | (5.4) |
Imposing the Dirichlet boundary conditions (5.1) to
A+C=T0,A+C−2BL=TL | (5.5) |
which implies that
T+(x)=T0−C+(TL−T0)2Lx,T−(x)=C+(TL−T0)2Lx | (5.6) |
where
0≤C≤T0 | (5.7) |
Condition (5.7) can be interpreted as follows: there is a
The constant
B=T0−TL2L=J02D0[T+(x)−T−(x)]=b2D0(T0−2C) | (5.8) |
which returns
C=12[(1−D0bL)T0+D0bLTL] | (5.9) |
Condition
TLT0≤1+bLD0=r∗ | (5.10) |
This means that for any positive value of
This result can be easily derived within the Poisson-Kac formulation of the transport problem, in which the basic quantities are the partial temperature waves
T(x)±J0b≥0 | (5.11) |
that corresponds to eq. (2.7). But, the positivity requirement of the quantities in eq. (5.11) cannot be easily derived within the Cattaneo approach, as it is an intrinsic by-product of the stochastic formulation of the process in terms of partial waves.
In order to avoid misunderstandings, it is important to stress out that the critical ratio
Nevertheless, the conceptual and theoretical importance of this result is rather clear: it indicates a criticality in the heat transport model based on the Cattaneo equation (breakdown of positivity of the partial temperature waves) that can be discovered only within the stochastic (partial wave) formulation of the problem.
To give a numerical example, set
The time evolution of
From this analysis it follows that, in the setting of a linear hyperbolic transport scheme on an interval, the Dirichlet boundary conditions (5.1) should be modified in order to ensure the fulfillment of positivity requirements. The simplest way to modify eq. (5.1) is to consider
T+(x,t)|x=0=[T0−T−(x,t)|x=0] | (5.12) |
and
T−(x,t)|x=L=[TL−T+(x,t)|x=L] | (5.13) |
where
From eqs. (5.3) and (5.12)-(5.13) it follows that at steady state
T+(x)=T+0−Bx,T−(x)=T−L−B(x−L) | (5.14) |
where
From
This means that positivity issues can arise solely in the neighborhood of
● either
● or
T+(x)=−Bx,T−(x)=TL+2BL−Bx | (5.15) |
and the value of the constant
dT+(x)dx=−B=−λb[T+(x)−T−(x)]=λb(TL+2BL) | (5.16) |
that implies
B=−bTL2D01+bLD0 | (5.17) |
so that the temperature value at
T(0+)=TL1+bLD0 | (5.18) |
In this case, the difference between the actual value of
T(0+)−T0=TL−T0(1+bLD0)1+bLD0 | (5.19) |
It is rather clear the physical phenomenology originating the possible occurrence of negative values at
In this article, several prototypical problems involving heat/mass transport models possessing finite propagation velocity (hyperbolic models) have been solved in closed form, and their physical implications thoroughly discussed. Two qualitative results are worth of special attention.
The finite propagation velocity induces as a by-product the regularization of the associated fluxes (in the meaning that it eliminates the singularities that may characterize the initial temporal/spatial scaling of the fluxes associated with a classical parabolic transport model involving Fourier/Fick constitutive equations), still maintaining a perfectly analogous long-term behavior observed in their parabolic counterparts. This is a nice and physically consistent result, as it indicates that the occurrence of singularities in boundary-layer problems analyzed within the paradigm of parabolic transport models, are essentially - to quote Müller [21] - an artifact of the oversimplified assumption of strictly Fickian/Fourier constitutive models at very short time scales. In principle, it would be possible to verify experimentally this claim, although it is rather clear that any experimental falsification of this property, would automatically question the basic assumptions of special relativity, as with reference to the problem treated in Section 4,
The second general observation that follows from the present analysis is that, although formally equivalent, the Cattaneo and the Poisson-Kac (stochastic) formulation of a heat/mass transport problem in one spatial dimension display conceptual differences when boundary value problems over finite domains (intervals) are considered. This depends on the physical information that is implicitly imbedded into the decomposition of a concentration field (say
These observation can be conceptually extended to non-linear transport models and to higher dimensional systems. In the latter case, it is sufficient to consider for the stochastic-based formulation the generalization of the Poisson-Kac models developed in [29,30,31]. Of course, the connection with the Cattaneo model in the latter case cannot be performed as the Cattaneo equation in higher dimensions violates the positivity requirements and cannot be recovered from any stochastic model.
The extension of the analysis developed in this article for the spinorial Green function of the one-dimensional Poisson-Kac model can be in principle extended to spatial dimensions higher than one, starting from the mathematical works by Kolesnik [54,55,56] for some classes of Generalized Poisson-Kac processes.
In this Appendix, the analytical expressions for the kernels necessary to explicit in closed form the spinorial Green function
z=√t2−x2/b2 | (6.1) |
the kernels
K1(x,t)=e−λt2[δ(x+bt)+δ(x−bt)]+λte−λt2b[η(x+bt)−η(x−bt)]I1(λz)z+λe−λt2b[η(x+bt)−η(x−bt)]I0(λz) | (6.2) |
K2(x,t)=e−λt2b[η(x+bt)−η(x−bt)]I0(λz) | (6.3) |
where
∂K2(x,t)∂t=e−λt2[δ(x+bt)+δ(x−bt)]+λte−λt2b[η(x+bt)−η(x−bt)]I1(λz)z−λe−λt2b[η(x+bt)−η(x−bt)]I0(λz) | (6.4) |
and
∂K2(x,t)∂x=e−λt2b[δ(x+bt)−δ(x−bt)]−λxe−λt2b3[η(x+bt)−η(x−bt)]I1(λz)z | (6.5) |
where the properties
One of the author (M.G.) thanks Alexander Kolesnik for useful and stimulating discussions during his stay at the Institute of Mathematics and Computer Science of the Kishinev University.
The authors declare no conflict of interest.
[1] | Wang L, Li B (2008) Phononics gets hot. Phys World 21: 27-29. |
[2] |
Maldovan M (2013) Sound and heat revolutions in phononics. Nature 503: 209-217. doi: 10.1038/nature12608
![]() |
[3] |
Li N, Ren J, Wang L, et al. (2012) Colloquium: Phononics: manipulating heat flow with electronic analogs and beyond. Rev Mod Phys 84: 1045-1066. doi: 10.1103/RevModPhys.84.1045
![]() |
[4] | Straughan B (2011) Heat waves, Springer, New York. |
[5] | Sellitto A, Cimmelli VA, Jou D (2016) Mesoscopic Theories of Heat Transport in Nanosystems, Springer, Heidelberg. |
[6] | Cattaneo C (1948) Sulla conduzione del calore. Atti Sem Mat Fis Univ Modena 3: 83-101. |
[7] | Cattaneo C (1958) Sur une forme de l'equation de la chaleur eliminant le paradoxe s'une propagation instantanee. Compt Rendu Acad Sci 247: 431-433. |
[8] |
Bubnov VA (1976) Wave concepts in the theory of heat. Int J Heat Mass Tran 19: 175-184. doi: 10.1016/0017-9310(76)90110-1
![]() |
[9] |
Glass DE, Özisik MN, Vick B (1985) Hyperbolic heat conduction with surface radiation. Int J Heat mass Tran 28: 1823-1830. doi: 10.1016/0017-9310(85)90204-2
![]() |
[10] | Joseph DD, Preziosi L (1989) Heat waves, Rev Mod Phys 61: 41-72. |
[11] |
Guyer RA, Krumhansl JA (1966) Solution of the Linearized Phonon Boltzmann Equation. Phys Rev 148: 766-778. doi: 10.1103/PhysRev.148.766
![]() |
[12] |
Guyer RA, Krumhansl JA (1966) Thermal Conductivity, Second Sound, and Phonon Hydrodynamic phenomena in Nonmetallic Crystals. Phys Rev 148: 778-788. doi: 10.1103/PhysRev.148.778
![]() |
[13] |
Jou D, Cimmelli VA (2016) Constitutive equations for heat conduction in nanosystems and nonequilibrium processes: an overview. Communications in Applied and Industrial Mathematics 7: 196-222. doi: 10.1515/caim-2016-0014
![]() |
[14] |
Jou D, Cimmelli VA, Sellitto A (2012) Nonlocal heat transport with phonons and electrons: Application to metallic nanowires. Int J Heat Mass Tran 55: 2338-2344. doi: 10.1016/j.ijheatmasstransfer.2012.01.033
![]() |
[15] |
Guo Y, Jou D, Wang M (2017) Macroscopic heat transport equations and heat waves in nonequilibrium states. Physica D: Nonlinear Phenomena 342: 24-31. doi: 10.1016/j.physd.2016.10.005
![]() |
[16] |
Camera-Roda G, Sarti GC (1990) Mass Transport with Relaxation in Polymers. AIChE J 36: 851-860. doi: 10.1002/aic.690360606
![]() |
[17] |
Edwards DA, Cohen DS (1995) A Mathematical Model for a Dissolving Polymer. AIChE J 41: 2345-2355. doi: 10.1002/aic.690411102
![]() |
[18] |
Ferreira JA, de Oliveira P, da Silva PM, et al. (2014) Molecular Transport in Viscoelastic Materials: Mechanistic Properties and Chemical Affinities. SIAM J Appl Math 74: 1598-1614. doi: 10.1137/140954027
![]() |
[19] |
Ferreira JA, Grassi M, Gudino E, et al. (2015) A new look to non-Fickian diffusion. Appl Math Model 39: 194-204. doi: 10.1016/j.apm.2014.05.030
![]() |
[20] |
Ferreira JA, Naghipoor J, de Oliveira P (2016) A coupled non-Fickian model of a cardiovascular drug delivery system. Math Med Biol 33: 329-357. doi: 10.1093/imammb/dqv023
![]() |
[21] | Müller I, Ruggeri T (1993) Extended Thermodynamics, Springer-Verlag, New York. |
[22] | de Groot SR, Mazur P (1984) Non-Equilibrium Thermodynamics, Dover Publ. Inc. New York. |
[23] | Jou D, Casas-Vazquez J, Lebon G (2010) Extended Irreversible Thermodynamics, Springer, New York. |
[24] | Jou D, Casas-Vazquez J, Lebon G (1988) Extended Irreversible Thermodynamics, Rep Prog Phys 51: 1105-1179. |
[25] |
Jou D, Casas-Vazquez J, Lebon G (1999) Extended Irreversible Thermodynamics revisited (1988-1998). Rep Prog Phys 62: 1035-1142. doi: 10.1088/0034-4885/62/7/201
![]() |
[26] |
Körner C, Bergmann HW (1998) The physical defects of the hyperbolic heat conduction equation. Appl Phys A 67: 397-401. doi: 10.1007/s003390050792
![]() |
[27] | Struchtrup H (2005) Macroscopic Transport Equations for Rarefied Gas Flows, Springer, Berlin. |
[28] | Giona M, Brasiello A, Crescitelli S (2016) Generalized Poisson-Kac processes: basic properties and implications in extended thermodynamics and transport. J Non-Equil Thermody 41: 107-114. |
[29] |
Giona M, Brasiello A, Crescitelli S (2017) Stochastic foundations of undulatory transport phenomena: generalized Poisson-Kac processes - part I basic theory. J Phys A 50: 335002. doi: 10.1088/1751-8121/aa79d4
![]() |
[30] |
Giona M, Brasiello A, Crescitelli S (2017) Stochastic foundations of undulatory transport phenomena: generalized Poisson-Kac processes - part II Irreversibility, norms and entropies. J Phys A 50: 335003. doi: 10.1088/1751-8121/aa79c5
![]() |
[31] |
Giona M, Brasiello A, Crescitelli S (2017) Stochastic foundations of undulatory transport phenomena: generalized Poisson-Kac processes - part III extensions and applications to kinetic theory and transport. J Phys A 50: 335004. doi: 10.1088/1751-8121/aa79d6
![]() |
[32] | Kac M (1956) Some Stochastic Problems in Physics and Mathematics, Socony Mobil Oil Company Colloquium Lectures. |
[33] | Kac M (1974) A stochastic model related to the telegrapher's equation. Rocky MT J Math 4: 497-510. |
[34] | Gardiner CW (1997) Handbook of Stochastic Methods, Springer-Verlag, Berlin. |
[35] | Falconer K (2003) Fractal Geometry, J Wiley & Sons, Chichister. |
[36] | Giona M (2018) Covariance and spinorial statistical description of simple relativistic stochastic kinematics, in preparation. |
[37] |
Rosenau P (1993) Random walker and the telegrapher's equation: A paradigm of a generalized hydrodynamics. Phys Rev E 48: R655-R657. doi: 10.1103/PhysRevE.48.R655
![]() |
[38] |
Brasiello A, Crescitelli S, Giona M (2016) One-dimensional hyperbolic transport: positivity and admissible boundary conditions derived from the wave formulation. Physica A 449: 176-191. doi: 10.1016/j.physa.2015.12.111
![]() |
[39] |
Zhukovsky K (2016) Violation of the maximum principle and negative solutions with pulse propagation in Guyer-Krumhansl model. Int J Heat Mass Tran 98: 523-529. doi: 10.1016/j.ijheatmasstransfer.2016.03.021
![]() |
[40] |
Zhukovsky K (2017) Exact negative solutions for Guyer-Krumhansl type equation and the maximum principle violation. Entropy 19: 440. doi: 10.3390/e19090440
![]() |
[41] |
Giona M (2017) Relativistic analysis of stochastic kinematics. Phys Rev E 96: 042133. doi: 10.1103/PhysRevE.96.042133
![]() |
[42] | Weiss G (1994) Aspects and Applications of the Random Walk, North-Holland, Amsterdam. |
[43] |
Metzler R, Klafter J (2000) The random walk's guide to anomalous diffusion: a fractional dynamics approach. Phys Rep 339: 1-77. doi: 10.1016/S0370-1573(00)00070-3
![]() |
[44] |
Metzler R, Klafter J (2004) The restaurant at the end of the random walk: recent developments in the description of anomalous diffusion by fractional dynamics. J Phys A 37: R161-R208. doi: 10.1088/0305-4470/37/31/R01
![]() |
[45] | Oldham KB, Spanier J (2006) The Fractional Calculus, Dover Publ. Inc., Mineola. |
[46] |
Criado-Sancho M, Lebot JE (1993) On the admissible values of the heat flux in hyperbolic heat transport. Phys Lett A 177: 323-326. doi: 10.1016/0375-9601(93)90008-N
![]() |
[47] |
Camacho J (1995) Third law of thermodynamics and maximum heat flux. Phys Lett A 202: 88-90. doi: 10.1016/0375-9601(95)00325-W
![]() |
[48] | Polyanin AD (2002) Handbook of Linear Partial Differential Equations for Engineers and Scientists, Chapman & Hall/CRC, Boca Raton. |
[49] | Kolesnik AD, Ratanov N (2013) Telegraph Processes and Option Pricing, Springer, Heidelberg. |
[50] | Ghizzetti A, Ossicini A (1971) Trasformate di Laplace e Calcolo Simbolico, UTET, Torino. |
[51] | Leveque A (1928) Les lois de la transmission de chaleur par convection. Annales des Mines 13: 201-299. |
[52] | Giona M, Adrover A, Cerbelli S, et al. (2009) Laminar dispersion at high éclet numbers in finite-length channels: Effects of the near-wall velocity profile and connection |
[53] | Marin E, Vaca-Oyola LS, Delgado-Vasallo O (2016) On thermal waves' velocity: some open questions in thermal waves' physics. Rev Mex Fis E 62: 1-4. |
[54] | Kolesnik AD, Turbin AF (1998) The equation of symmetric Markovian random evolution in a plane, Stoch Proc Appl 75: 67-87. |
[55] |
Kolesnik AD (2001) Weak Convergence of a Planar Random Evolution to the Wiener process. J Theor Probab 14: 485-494. doi: 10.1023/A:1011167815206
![]() |
[56] |
Kolesnik AD, Pinsky MA (2011) Random Evolutions are Driven by the Hyperparabolic Operators. J Stat Phys 142: 828-846. doi: 10.1007/s10955-011-0131-0
![]() |
1. | Massimiliano Giona, Mirko D’Ovidio, Davide Cocco, Andrea Cairoli, Rainer Klages, Age representation of Lévy walks: partial density waves, relaxation and first passage time statistics, 2019, 52, 1751-8113, 384001, 10.1088/1751-8121/ab38eb | |
2. | Massimiliano Giona, Andrea Cairoli, Davide Cocco, Rainer Klages, Spectral Properties of Stochastic Processes Possessing Finite Propagation Velocity, 2022, 24, 1099-4300, 201, 10.3390/e24020201 | |
3. | Massimiliano Giona, Andrea Cairoli, Rainer Klages, Extended Poisson-Kac Theory: A Unifying Framework for Stochastic Processes with Finite Propagation Velocity, 2022, 12, 2160-3308, 10.1103/PhysRevX.12.021004 | |
4. | Moh Shahid Khan, Ravi Kumar Mandava, Design of dynamically balanced gait for the biped robot while crossing the obstacle, 2024, 238, 0954-4062, 9125, 10.1177/09544062241246878 |