Abstract
The development of non-Fourier heat conduction models is encouraged by the invalidity of Fourier’s law to explain heat conduction in ultrafast or ultrasmall systems. The production of negative entropy will result from the combination of traditional nonequlibrium thermodynamics and non-Fourier heat conduction models. To resolve this paradox, extended irreversible thermodynamics (EIT) introduces a new state variable. However, real dynamics variables like force and momentum are still missing from nonequilibrium thermodynamics and EIT’s generalized force and generalized flux. Heat has both mass and energy, according to thermomass theory and Einstein’s mass-energy relation. The generalized heat conduction model containing non-Fourier effects was established by thermomass gas model. The thermomass theory reshapes the concept of the generalized force and flux, temperature, and entropy production in nonequilibrium thermodynamics and revisits the assumption for the linear regression of the fluctuations in Onsager reciprocal relation. The generalized heat conduction model based on thermomass theory has been used to study thermal conductivity, thermoelectric effect, and thermal rectification effect in nanosystems.
1 Introduction
1.1 Local equilibrium thermodynamics
Equilibrium thermodynamics, developed in the 19th century, is concerned with the macroscopic properties of matter at near-equilibrium, but it only involves the initial and final equilibrium states and cannot describe the process itself. In 1931, Onsager [1] used fluctuation linear regression and the microscopic reversibility assumption to derive general reciprocity relationships that apply to transport processes like heat conduction, electrical conduction, and diffusion. In 1961, Prigogine [2], [3] discovered a dissipative system that operates in an environment with which it exchanges energy and matter, often far from thermodynamic equilibrium. These two pioneering works, along with a series of other developments [4], gave rise to local equilibrium thermodynamics, also known as classical irreversible thermodynamics (CIT).
In CIT, the entropy production obtains
which has a structure of a bilinear form in the generalized flux, J, and the generalized force, X. For general irreversible energy transfer processes, Eq. (1) can be written as
where q is the heat flux, P v is the stress tensor, u f is the fluid velocity, J k is the mass component diffusion flux, μ k is the chemical potential, and i is the electric potential. Each term represents the entropy production induced by heat conduction, momentum diffusion, mass component diffusion, and electrical conduction, respectively.
1.2 Extended irreversible thermodynamics (EIT)
The local equilibrium hypothesis implies large time and space scales, so high-frequency processes and non-Fourier heat conduction cannot be adequately described by CIT, such as nanomaterials, neutron scattering, and superfluids [5]. Therefore, the proposal of extended irreversible thermodynamics (EIT) goes beyond the local equilibrium assumption and introduces dissipative fluxes (heat flux, diffusion flux, electric flux, and viscous pressure) as independent variables to describe these phenomena [6]. The key to constructing EIT is the establishment of flux evolution equation and the reconstruction of non-equilibrium entropy.
Inserting the CV model [7]–[9],
where τ is the relaxation time and κ is the thermal conductivity, into the classical expression of the entropy production, Eq. (2), it can be found that
which no longer remains positive definite due to the presence of linear terms in the equation. When Fourier’s law holds, the entropy not only depends on the temperature, T, also depends on the heat flux, q. To eliminate this paradox, EIT considers the heat flux, q, as an independent variable. Combined the CV model the extended entropy production is
To obtain an evolution equation for q compatible with the positive value of σ S by assuming that the force X is linearly related to q, i.e. the corresponding positive extended entropy production is
By altering entropy and entropy production as illustrated in Figure 1, EIT eliminates the paradox of negative entropy production in non-Fourier heat conduction. Therefore, it is necessary to redefine the thermodynamics temperature so that the nonequilibrium temperature, θ, in EIT is
where T eq is the temperature at equilibrium state, q is the local heat flux, κ is the thermal conductivity, C is the specific heat capacity, ρ is the density of material, and τ is the relaxation time. Eq. (7) indicates that θ is lower than T eq . However, EIT does not give an explicit physical definition of generalized forces and fluxes. The nonequilibrium temperature, θ, is based on a mathematical derivation of the extended entropy production and requires more physical discussions. Thus, more fundamental research on nonequilibrium thermodynamics is required to address non-Fourier heat conduction.
![Figure 1:
Entropy evolutions in an isolated 1D system of the extended entropy (solid line) and the conventional equilibrium entropy (dash line) [10].](/document/doi/10.1515/jnet-2023-0094/asset/graphic/j_jnet-2023-0094_fig_001.jpg)
Entropy evolutions in an isolated 1D system of the extended entropy (solid line) and the conventional equilibrium entropy (dash line) [10].
In nanosystems, the limited velocity of energy carriers (phonons) is responsible for the limiting heat propagation speed. The time required for energy carriers to accelerate is disregarded by Fourier’s law. For example, Tzou [11] experimentally observed the lagging behavior of heat transfer in small-scale and fast-transient conditions, which further proves that heat has inertia [1], [12]. Therefore, Fourier’s law fails because the phonon transport is only partially ballistic in certain temporal and geographical scales. Some transport equations were derived macroscopically by EIT and applied to study the heat conduction in nanowires [13]–[18]. Meanwhile, several simulations and theoretical studies revealed the one-dimensional nanosystem’s anomalous length dependence of thermal conductivity [19]–[26]. Although low-dimensional materials have longer phonon mean free paths than conventional materials, which causes their thermal conductivity to increase with length over a larger size range. However, Fourier’s law is still the foundation for these computations and measurements of thermal conductivity. This method of approximating thermal conductivity may not be appropriate when Fourier’s law fails. Hence, it is highly desirable to have a more comprehensive heat conduction model that is appropriate for non-Fourier heat conduction and has explicit macroscopic physical meaning.
1.3 Thermomass theory
Before the 20th century, it was generally accepted that the heat transfer process was simply an energy transfer without the transport of substance [27]. In contemporary thermodynamics, via Einstein’s special relativity, Guo et al. [28]–[30] determined the equivalent mass (also known as the thermomass) of the thermal energy as a component of the system’s invariant mass. The inertia effect of thermomass is attributed to the nonlocal and nonlinear heat conduction. The variation of effective thermal conductivity with system size and temperature at steady state is investigated using the governing generalized heat conduction equation based on the thermomass gas model [31], [32], which is consistent with the experimental results [33]. Recently, Nie et al. [34] clarified the mathematical rationality of generalized heat conduction equation in the general equation for nonequilibrium reversible-irreversible coupling (GENERIC) framework and revealed that thermomass theory gives the physical significance of Hamiltonian energy and dissipation potential in heat conduction under GENERIC framework. The mass balance equation of a thermomass gas:
The momentum balance equation of a thermomass gas:
with
where ρ T is the density of the thermomass gas, u T is the drift velocity of the thermomass gas, P T is the thermomass pressure, and f T is the resistance force per unit volume of the thermomass gas flowing through the medium. c is the velocity of light and α is a proportional parameter depending on the state of medium. Eq. (8) and (9) describe the motion of the thermomass gas without any other artificial hypothesis. Refer to the detailed derivation process [31], [32], the generalized heat conduction equation can be written as
where
τ T is a characteristic time between the temperature gradient and heat flux. l is a characteristic length measuring the spatial inertia of thermomass. b is a dimensionless parameter for characterizing flow compressibility of thermomass gas. Eq. (11) shows the transient inertia effect in the 1st term. The 2nd and 3rd terms represent the spatial inertia effects. The driving force is represented by the 4th term, and the resistance force by the final term. The Fourier’s law is recovered by Eq. (11) if all the effects of inertia are disregarded. The CV model is what remains when the spatial inertia effect alone is negligible.
In this review, we first briefly introduce the extended irreversible thermodynamics (EIT) and the generalized heat conduction model based on thermomass theory. Then, we reshape the concepts such as entropy production, temperature and Onsager reciprocal relation (ORR) based on EIT and thermomass theory. Finally, we review recent applications of thermomass theory for thermal conductivity, thermoelectric effect, and thermal rectification effect in nanosystems.
2 Entropy production for heat conduction
The entropy production is the key quantity in CIT. It serves as the foundation for the minimum entropy production principle and the derivation of the Onsager reciprocal relation. It was noted in the introduction that in the case of transient non-Fourier heat conduction, the conventional expression [Eq. (4)] in traditional nonequilibrium thermodynamics will encounter the paradox of negative entropy production. In order to avoid the paradox of violating the second law of thermodynamics, the extended entropy production expression [Eq. (6)] for heat conduction based on EIT introduces new state variables to maintain semi-positive definite in non-Fourier heat conductions.
A generalized expression of entropy production can be reshaped based on thermomass theory,
with
where ξ is the dimensionless coefficient from the ratio of the mechanical energy of thermomass, E h , to the internal energy u,
where E T is made up of potential energy and kinetic energy. In general, the internal energy corresponding potential part is predominant. Only in extreme situations, like high-frequency processes or extremely large heat fluxes, does the kinetic component have any significance; in these cases, the temperature distribution is insufficient to adequately describe the transport process. Eq. (6), which states that the system entropy depends on both the heat flux and the classical variables, is consistent with the modified expression for entropy production based on the thermomass theory.
In the expression for entropy production of EIT, the generalized force is
where θ is the nonequilibrium temperature and expression is shown in Eq. (7). If the characteristic time, τ T , Eq. (12a) is brought into Eq. (16), the generalized fore based on thermomass theory can be written as [35]
As a result, the generalized fluxes can be broken down into the density and drift velocity, and the generalized force in EIT matches the real resistance force in the thermomass model. The 1st term on the right side of Eq. (16) denotes the driving force and the 2nd term represents the inertia effect. From the perspective of thermomass theory, the resistance force has the same units of Nm−3 as body force in continuum mechanics. Therefore, the reason why the entropy production formula in CIT fails in non-Fourier heat conduction is that it is the resistance force rather than the driving force (thermodynamics forces in CIT) that determines the dissipation or irreversibility of the transmission process.
3 Temperature in nonequilibrium thermal dynamics
The nonequilibrium temperature in EIT is the result of a modification in entropy production, which alters the expression of temperature [Eq. (7)]. The reason for this discrepancy is the influence of local nonequilibrium. However, in the derivation, the expression for the nonequilibrium temperature still uses the equilibrium temperature. The application of the equilibrium temperature in non-equilibrium circumstances is not sufficiently understood. The nonequilibrium molecular dynamics simulations [36] and the forced oscillator studies [37] reveal that in the nonequilibrium steady states, the kinetic temperature may be significantly different from thermodynamic equilibrium criteria in the measurement, at which point the entropy also needs to be reconsidered. Therefore, the non-equilibrium temperature is independent of whether the transport process is transient or not should be revisited.
3.1 The zeroth law
The zeroth law gives the definition of thermal equilibrium and allows the temperature to become the measurable quantity. According to Fourier’s law, heat flux can only occur between two systems with different temperatures, T eq . EIT proposes the nonequilibrium temperature, θ, as the criterion of thermal equilibrium [5], [6]. According to Eq. (7), θ will reduce due to the presence of heat flux, q, which means that heat flux can occur among systems with the same T eq . based on the thermomass model, the heat flux is driven by the static pressure gradient of thermomass gas, ∇p T . Thus, we can obtain a relation between the total temperature of thermomass gas, T t and the static temperature, T,
Equation (18) shows that the static temperature, T, is smaller than the total temperature of thermomass gas, T t , which means that the heat conduction occurs as long as there is a difference of static temperature between two systems. Analogous to EIT, the static temperature, T, corresponds to the nonequilibrium temperature, θ, while the total temperature of thermomass gas, T t , corresponds to the equilibrium temperature, T eq , which carries all the energy of system. When the heat flux is stagnant, θ is equal to T eq and T is equal to T t .
3.2 The second law
In Section 2, thermomass theory reshapes the expression of entropy production by the resistance force and the drift velocity. Hence, we can also reshape the temperature by using the total temperature of thermomass gas, T t , to measure the internal energy of system. From the classical relation among temperature, entropy, and internal energy, T −1 = ds/de, we can obtain
We can find that Eq. (19) is similar to the EIT’s derivation, Eq. (7). However, there is a small distinction between the relationship between the static temperature, T, and the stagnant temperature, T t , and that between the nonequilibrium temperature, θ, and the equilibrium temperature, T eq . There is no difference between entropy of the nonequilibrium and that of static temperatures. It is the internal energy expression that differs. In EIT, the definition of internal energy is based on the equilibrium temperature, T eq , which includes all molecular energies of phonon gas. But in thermomass theory, the internal energy is defined as the static temperature, T without the drift kinetic energy. Therefore, the temperature produced by both the thermomass theory and the EIT is lower than the total or equilibrium temperature; the reason for this discrepancy is the way internal energy is described differently by the two theories.
4 Onsager reciprocal relations (ORR)
The foundation of nonequilibrium thermodynamics, which elucidates the coupling relationship between generalized forces and flux, is the Onsager reciprocal relations (ORR). In ORR, two assumptions are made: the first is linear regression of the fluctuations, and the second is microscopic reversibility. ORR indicates that the generalized forces, X and flux, J need to satisfy
where σ is the local entropy production rate, J represents the time derivative of a state variable, α, and X denotes the derivative of the entropy deviation with respect to α, and ΔS is the entropy change in the equilibrium state. We can get the linear law by combining the linear regression of the fluctuations and Eq. (21),
where L is the phenomenological coefficient of the linear transport law. The generalized heat conduction equation based on thermomass theory provides a new understanding of generalized forces and flux in heat conduction, which is here defined by real forces, namely the driving and resistance forces. Eq. (20) can be rewritten as
where entropy production is the production of the driving and resistance forces in Eq. (13), divided by T. The detailed derivation process has been discussed in Ref. [38]. Hence, the generalized forces and flux satisfy the linear phenomenological law in CIT, we can obtain
where Λ ij = L ij /(Tρ i ρ j ) and its unit is m3s/kg. The symmetry of Λ is the same as L in Eq. (22). Combining with Eqs. (24), (21) can be rewritten as
where δ are variables that measure fluctuations at near-equilibrium [4]. Thus, Eq. (25) represents the linear regression of fluctuations due to the resistance force. If the reduced limit leading to a stable non-equilibrium state is suddenly eliminated, the drift speed will decay exponentially,
where R i ρ i denotes the inertial effect and R i = 1/v s 2 for heat conduction. According to Eq. (26), the characteristic relaxation time of Eq. (24) can be extracted
When i = j, the resistance force causes the flux to decay proportionate to the drift velocity; when i ≠ j, other types of fluxes are induced and they will be impeded by resistance force. For heat conduction, the heat flux is resulted from the temperature gradient. The temperature gradient produces the heat flux in the case of heat conduction. The heat flux will be zero when the temperature gradient is removed. Thermal energy has the inertial effect according to thermomass theory. The heat flux will therefore degrade freely when the temperature gradient abruptly disappears. The relationship in free decay of heat flow is the same as that in steady heat conduction, which is the same as the assumption of linear regression of fluctuations in ORR. This is because the thermomass theory states that the relationship between resistance force and velocity does not depend on driving or inertial forces.
To put it succinctly, thermomass theory demonstrates that the linear regression of fluctuations that Onsager’s proof relied upon is actually a balance between resistance force and inertial effect. The characteristic displacement’s time derivative is the inertial effect. The product of relaxation time and drift velocity yields characteristic displacement. These state variables’ time derivatives yield the flux, satisfying Onsager’s initial demand to demonstrate the reciprocal relationship. The ORR can be macroscopically derived from the symmetry of the matrix, Λ, based on the thermomass theory, as proved by the third law of Newtonian dynamics or the Galilean invariance. This proof can be extended to analyze thermoelectric effects by relating the reversible thermoelectric reciprocal relationship between the reversible Seebeck coefficient and the reversible Peltier coefficient to be equivalent to the conventional Kelvin relation [39].
5 Applications of the generalized heat conduction model
5.1 Thermal conductivity of nanomaterials
As mentioned above, when the thermal inertia effect cannot be ignored, the generalized heat conduction equation can be used to study non-Fourier heat conduction, such as thermal conductivity of nanomaterials, size dependence of thermal conductivity, and thermal rectification.
Cao et al. [28] investigated the thermal conductivity of a single carbon nanotube (CNT) with different lengths based on the thermomass gas model. They used a CNT with an intrinsic thermal conductivity of 5000 W/mK as an example, and the temperature difference between the two ends was assumed to be 20 K. They predicted the apparent thermal conductivity of CNT by
where κ app is the apparent thermal conductivity and κ is the intrinsic thermal conductivity. The difference between the apparent and intrinsic thermal conductivities is due to the spatial inertia of thermomass.
Figure 2 indicates that the apparent thermal conductivity of CNT increases with increasing length. Eq. (28) reveals that the apparent thermal conductivity is always smaller than intrinsic thermal conductivity of materials. The length dependence of thermal conductivity of CNT agrees with the MD simulations [22], [40], which shows an exponentially increasing trend. Recent several theoretical studies [19], [20], [24] and experimental observations [25], [26] reveal the anomalous heat conduction in one-dimensional system that there is a presence of divergent thermal conductivity with length. The thermomass model could be used as a mesoscopic theory to explore the physical mechanism of anomalous heat conduction from micro to macro.
![Figure 2:
The predicted apparent thermal conductivity of CNT with incremental length. Reprinted with permissions from AIP Publishing [28].](/document/doi/10.1515/jnet-2023-0094/asset/graphic/j_jnet-2023-0094_fig_002.jpg)
The predicted apparent thermal conductivity of CNT with incremental length. Reprinted with permissions from AIP Publishing [28].
At nanoscales, the thermal conductivity of materials is limited by phonon mean free paths (MFP). In this case, the rarefication effect of phonon gas should be considered to calculate the effective thermal conductivity. To address the rarefication effect of phonon gas, Dong et al. [41] proposed a macroscopic heat conduction based on the phonon gas model, to which a second order resistance term is added in the generalized heat conduction equation. They were able to derive an explicit expression for effective thermal conductivity of Si nanomaterials.
For a nanofilm, the effective thermal conductivity is
For a nanowire, the effective thermal conductivity is
where λ E is the characteristic length depending on the material properties and ambient temperature. Figure (3) demonstrates that the effective thermal conductivities of Si nanofilm and nanowire increase with length and approach the bulk thermal conductivities. The size dependence of thermal conductivity predicted by the generalized heat conduction model agrees with the experimental results, where the additional second order resistance term is applied to assess the rarefication effect of phonon gas caused by the enhanced viscous friction on nanoscale boundaries.
![Figure 3:
The length dependence of the effective thermal conductivity (κ
eff
) of Si nanofilm [dash line, Eq. (29)] and nanowire [solid line, Eq. (30)] predicted by the modified generalized heat conduction model. The solid circles and triangles are the experimental results of nanofilm [42], [43] and nanowire [44], respectively. Reprinted with permissions from Elsevier [41].](/document/doi/10.1515/jnet-2023-0094/asset/graphic/j_jnet-2023-0094_fig_003.jpg)
The length dependence of the effective thermal conductivity (κ eff ) of Si nanofilm [dash line, Eq. (29)] and nanowire [solid line, Eq. (30)] predicted by the modified generalized heat conduction model. The solid circles and triangles are the experimental results of nanofilm [42], [43] and nanowire [44], respectively. Reprinted with permissions from Elsevier [41].
5.2 Thermoelectric effect of nanomaterials
The development of nanotechnology enhances the thermoelectric figure of merit (ZT) of nanomaterials [45]–[48]. This enhanced mechanism may be due to the fact that while the nanostructure maintains the electric conductivity, the phonon boundary scattering also reduces the thermal conductivity of the material. The thermomass theory gives a new perspective to analyze the thermoelectric effect. Energy exchange occurs between electrons and phonons [49],
where I is the electric current and E is the electric field. Combining the generalized heat conduction model without the inertia effect and the dimensionless figure of merit, ZT = S 2 ρ −1 κ −1 T. When IE > 0, electrical energy is converted to thermal energy, and the system is a typical thermoelectric cooler. The heat flux is also hindered, and the thermoelectric ZT decreases. When IE < 0, thermal energy is converted into electrical energy, which is a thermoelectric generator. The heat flux is further increased, and the effective thermoelectric ZT increases. In addition, the inertial effect may favor the ZT of the device.
Rogolino et al. [50], proposed a nonlinear equation of thermoelectric coupling based on thermomass theory and used it to evaluate the efficiency of the thermoelectric generator and investigate the effect of current on thermoelectric efficiency of device.
Cimmelli and Rogolino [51] examined the efficiency of a thermoelectric energy generator composed of a silicon-germanium alloy using the generalized heat conduction model and talked about nonlocal and nonlinear heat transport in nanosystems. They looked into the dependence of the thermoelectric efficiency on composition and temperature differential in addition to experimentally determining the thermal conductivity of a nanowire as a function of temperature and composition.
Youssef and AI-Lehaibi [52] drew on thermomass theory to limit the propagated speed of thermal and mechanical waves to set up an extended thermoelasticity model. It brings thermoelastic materials closer to their true physical behavior by weakening the coupling relationship between mechanical waves and thermal energy.
5.3 Thermal rectification in nanosystems
Recently thermal rectification effect in nanosystems has attracted wide interesting. Researchers have constructed a variety of thermal rectification devices through molecular dynamics simulation and experiments [53]–[55], among which the most important forms are asymmetric nanostructures. For example, mass graded carbon nanotube [56], carbon nanocone [57], asymmetric graphene ribbon [58], [59], low-dimensional heterojunction [60], asymmetric molecular bridge [61] and so on. The mechanism of thermal rectification phenomenon is attributed to the asymmetry of phonon frequency-energy spectrum coupling [62], but there is no definite quantitative studies. A potential mechanism of thermal rectification is proposed by combining thermomass theory and hydrodynamics. The convective acceleration term in the Navier–Stokes equation represents when the fluid accelerates or decelerates. Therefore, when the cross-sectional area of the flow channel is changed, the flow at the same pressure difference is different in the anteroposterior direction. For a medium with uniformly varying cross section, the equivalent thermal resistance, ϒ, can be expressed as [49]
where the first term on the right side of Eq. (32) is the inertial effect of thermomass. When the heat flow direction is opposite, this part will change positively or negatively. The second term on the right side of Eq. (32) is resistance force of thermomass, which is mainly determined by the shape of medium but not by the direction of heat flux. When the heat flux is from the narrow to wide (NTW), A 2 > A 1 , inertial thermal resistance is less than 0; When the heat flux is from the wide to narrow (WTN), A 1 > A 2 , inertial thermal resistance is greater than 0. Figure 4(a) shows a trapezoidal silicon nanofilm with thickness D of 50 nm, width of narrow end, L N = 50 nm, width of narrow end, L W = 300 nm, and length of nanofilm, L = 500 nm. The thermal rectification ratio, η, and wall angle, θ, are defined by

Schematic of thermal rectification model based on thermomass theory: (a) Thermal rectification in the trapezoidal nanoribbon. (b) Thermal rectification ratio changes with wall angle and temperature.
Figure 4(b) shows that the rectification ratio decreases with the increasing temperature. The larger the wall angle, the larger the rectification ratio. When the wall angle is 18.9°, the rectification ratio can reach nearly 40 % at 220 K. However, when the wall angle is large, the flow separation phenomenon at the wall may occur, and the rectification ratio may sharply reduce or even become negative.
When the thermal inertia force is considered, the total equivalent thermal resistance of an asymmetric conductor contains inertial thermal resistance that changes with the direction of heat flow, resulting in thermal rectification. This is a steady-state non-Fourier heat conduction phenomenon caused by the inertia force of the thermomass. The equivalent thermal resistance in NTW mode is smaller than that in WTN mode, which is consistent with the experimental results in silicon nanofilms [63]. The rectification effect is more significant at lower temperatures and increases with the increase of wall angle. If flow separation occurs, the equivalent thermal resistance in NTW mode may be higher than that in WTN mode. Flow separation is more likely to occur when the wall angle and the heat flow are large, which is consistent with the results of MD simulation [58]. However, in the real and complex heterogeneous nanosystems, more factors will affect the rectification effect, such as phonon MFP, phonon boundary scattering, ballistic-diffusive phonon coupling, and negative differential thermal resistance, etc.
6 Conclusions
Heat conduction that does not follow Fourier’s law in nature cannot be explained by the conventional nonequilibrium thermodynamics. To overcome the paradox of negative entropy production in combination of non-Fourier heat conduction and conventional nonequilibrium thermodynamics, EIT adds the new state variable to the expression of entropy production. To make the generalized force, generalized flux, and temperature compatible with non-Fourier heat conduction in EIT, they are modified.
The mass-energy relation found in Einstein’s special relativity is where the idea of thermomass theory originates. It implies that there is mass and energy duality in heat. A generalized heat conduction equation is presented based on thermomass gas model, which defines the real force and real flux. Analogy to EIT, the generalized force corresponds to the resistance forces of thermomass and the generalized flux corresponds to the density and the drift velocity of thermomass. From the viewpoint of thermomass theory, the controversy in Onsager’s proof of his reciprocal relationship is revisited. It shows that the linear decay of fluctuations is a balance between inertial forces and resistance.
The size-dependent thermal conductivity of nanomaterials is calculated by further extending the generalized heat conduction model. Effective thermal conductivities in nanofilms and nanowires are predicted by analyzing the viscous and rarefied effects of thermomass gas, which agrees well with experiments. Analogy to electrical transport and fluid flow, thermomass model can be used to explore the physical mechanisms of thermoelectric effect and thermal rectification in nanosystems.
Acknowledgments
The authors acknowledge the Beijing Super Cloud Center (BSCC) for providing HPC resources that have contributed to the research results reported within this paper. URL: http://www.blsc.cn/.
-
Research ethics: Not applicable.
-
Author contributions: The authors have accepted responsibility for the entire content of this manuscript and approved its submission.
-
Competing interests: The authors state no competing interests.
-
Research funding: This research is supported by National Natural Science Foundation of China (Grant No. 52006050), the Fundamental Research Funds for the Provincial Universities of Zhejiang (Grant No. GK219909299001-005). Prof. Yuan Dong also acknowledges the previous support by National Natural Science Foundation of China (Grant Nos. 51136001, 51356001).
-
Data availability: Not applicable.
References
[1] L. Onsager, “Reciprocal relations in irreversible processes. I,” Phys. Rev., vol. 37, no. 4, p. 405, 1931. https://doi.org/10.1103/physrev.37.405.Suche in Google Scholar
[2] I. Prigogine, “Time, structure, and fluctuations,” Science, vol. 201, no. 4358, pp. 777–785, 1978. https://doi.org/10.1126/science.201.4358.777.Suche in Google Scholar PubMed
[3] L. Onsager, “Reciprocal relations in irreversible processes. II,” Phys. Rev., vol. 38, no. 12, p. 2265, 1931. https://doi.org/10.1103/physrev.38.2265.Suche in Google Scholar
[4] L. Onsager and S. Machlup, “Fluctuations and irreversible processes,” Phys. Rev., vol. 91, no. 6, p. 1505, 1953. https://doi.org/10.1103/physrev.91.1505.Suche in Google Scholar
[5] G. Lebon, D. Jou, and J. Casas-Vázquez, Understanding non-Equilibrium Thermodynamics, vol. 295, Berlin, Springer, 2008.10.1007/978-3-540-74252-4Suche in Google Scholar
[6] D. Jou, J. Casas-Vazquez, and G. Lebon, “Extended irreversible thermodynamics revisited (1988-98),” Rep. Prog. Phys., vol. 62, no. 7, p. 1035, 1999. https://doi.org/10.1088/0034-4885/62/7/201.Suche in Google Scholar
[7] C. Cattaneo, “Sulla conduzione del calore,” Atti Semin. Mat. Fis. Univ. Modena, vol. 3, pp. 83–101, 1948.Suche in Google Scholar
[8] P. Vernotte, “Paradoxes in the continuous theory of the heat equation,” CR Acad. Sci., vol. 246, no. 3, pp. 153–154, 1958.Suche in Google Scholar
[9] P. M. Morse and H. Feshbach, Methods of theoretical physics, New York, McGraw Hill Company, 1953, p. 1407.Suche in Google Scholar
[10] M. Criado-Sancho and J. Llebot, “Behavior of entropy in hyperbolic heat conduction,” Phys. Rev. E, vol. 47, no. 6, p. 4104, 1993. https://doi.org/10.1103/physreve.47.4104.Suche in Google Scholar PubMed
[11] D. Y. Tzou, “Experimental support for the lagging behavior in heat propagation,” J. Thermophys. Heat Transfer, vol. 9, no. 4, pp. 686–693, 1995. https://doi.org/10.2514/3.725.Suche in Google Scholar
[12] W. Nernst, Die theoretischen grundlagen des neuen wärmesatzes. Thermodynamical Papers of the Physico-Chemical Institute of the University of Berlin, H. W. Knapp, Ed., Halle, 1918.Suche in Google Scholar
[13] V. Cimmelli, A. Sellitto, and D. Jou, “Nonlocal effects and second sound in a nonequilibrium steady state,” Phys. Rev. B, vol. 79, no. 1, p. 014303, 2009. https://doi.org/10.1103/physrevb.79.014303.Suche in Google Scholar
[14] V. A. Cimmelli, A. Sellitto, and D. Jou, “Nonequilibrium temperatures, heat waves, and nonlinear heat transport equations,” Phys. Rev. B, vol. 81, no. 5, p. 054301, 2010. https://doi.org/10.1103/physrevb.81.054301.Suche in Google Scholar
[15] V. Cimmelli, A. Sellitto, and D. Jou, “Nonlinear evolution and stability of the heat flow in nanosystems: beyond linear phonon hydrodynamics,” Phys. Rev. B, vol. 82, no. 18, p. 184302, 2010. https://doi.org/10.1103/physrevb.82.184302.Suche in Google Scholar
[16] A. Sellitto, F. X. Alvarez, and D. Jou, “Second law of thermodynamics and phonon-boundary conditions in nanowires,” J. Appl. Phys., vol. 107, no. 6, p. 064302, 2010. https://doi.org/10.1063/1.3309477.Suche in Google Scholar
[17] D. Jou, G. Lebon, and M. Criado-Sancho, “Variational principles for thermal transport in nanosystems with heat slip flow,” Phys. Rev. E, vol. 82, no. 3, p. 031128, 2010. https://doi.org/10.1103/physreve.82.031128.Suche in Google Scholar
[18] D. Jou, M. Criado-Sancho, and J. Casas-Vázquez, “Heat fluctuations and phonon hydrodynamics in nanowires,” J. Appl. Phys., vol. 107, no. 8, p. 084302, 2010. https://doi.org/10.1063/1.3380842.Suche in Google Scholar
[19] R. Livi and S. Lepri, “Heat in one dimension,” Nature, vol. 421, no. 6921, p. 327, 2003. https://doi.org/10.1038/421327a.Suche in Google Scholar PubMed
[20] B. Li and J. Wang, “Anomalous heat conduction and anomalous diffusion in one-dimensional systems,” Phys. Rev. Lett., vol. 91, no. 4, p. 044301, 2003. https://doi.org/10.1103/physrevlett.91.044301.Suche in Google Scholar PubMed
[21] C.-W. Chang, D. Okawa, H. Garcia, A. Majumdar, and A. Zettl, “Breakdown of Fourier’s law in nanotube thermal conductors,” Phys. Rev. Lett., vol. 101, no. 7, p. 075903, 2008. https://doi.org/10.1103/physrevlett.101.075903.Suche in Google Scholar
[22] S. Maruyama, “A molecular dynamics simulation of heat conduction in finite length SWNTs,” Phys. B, vol. 323, nos. 1–4, pp. 193–195, 2002. https://doi.org/10.1016/s0921-4526(02)00898-0.Suche in Google Scholar
[23] M. Fujii, X. Zhang, H. Xie, H. Ago, K. Takahashi, T. Ikuta, H. Abe, and T. Shimizu, “Measuring the thermal conductivity of a single carbon nanotube,” Phys. Rev. Lett., vol. 95, no. 6, p. 065502, 2005. https://doi.org/10.1103/physrevlett.95.065502.Suche in Google Scholar
[24] S.-N. Li and B.-Y. Cao, “Fractional Boltzmann transport equation for anomalous heat transport and divergent thermal conductivity,” Int. J. Heat Mass Transfer, vol. 137, pp. 84–89, 2019. https://doi.org/10.1016/j.ijheatmasstransfer.2019.03.120.Suche in Google Scholar
[25] F. Yao, et al.., “Experimental evidence of superdiffusive thermal transport in Si0.4Ge0.6 thin films,” Nano Lett., vol. 22, no. 17, pp. 6888–6894, 2022. https://doi.org/10.1021/acs.nanolett.2c01050.Suche in Google Scholar PubMed
[26] L. Yang, et al.., “Observation of superdiffusive phonon transport in aligned atomic chains,” Nat. Nanotechnol., vol. 16, no. 7, pp. 764–768, 2021. https://doi.org/10.1038/s41565-021-00884-6.Suche in Google Scholar PubMed
[27] E. Mendoza, “A sketch for a history of early thermodynamics,” Phys. Today, vol. 14, no. 2, pp. 32–42, 1961. https://doi.org/10.1063/1.3057388.Suche in Google Scholar
[28] B.-Y. Cao and Z.-Y. Guo, “Equation of motion of a phonon gas and non-Fourier heat conduction,” J. Appl. Phys., vol. 102, no. 5, p. 53503, 2007. https://doi.org/10.1063/1.2775215.Suche in Google Scholar
[29] Z. Y. Guo, K. Cheng, J. W. Li, and H. Pao, “Motion and transfer of thermal mass-thermal mass and thermon gas,” J. Eng. Thermophys., vol. 27, no. 4, pp. 631–634, 2006.Suche in Google Scholar
[30] Z. Y. Guo, B. Y. Cao, H. Y. Zhu, and Q. G. Zhang, “State equation of phonon gas and conservation equations for phonon gas motion,” Acta Phys. Sin., vol. 56, no. 6, pp. 3306–3312, 2007. https://doi.org/10.7498/aps.56.3306.Suche in Google Scholar
[31] M. Wang, B.-Y. Cao, and Z.-Y. Guo, “General heat conduction equations based on the thermomass theory,” Front. Heat Mass Transfer, vol. 1, no. 1, 2010. https://doi.org/10.5098/hmt.v1.1.3004.Suche in Google Scholar
[32] Y. Dong, B.-Y. Cao, and Z.-Y. Guo, “Generalized heat conduction laws based on thermomass theory and phonon hydrodynamics,” J. Appl. Phys., vol. 110, no. 6, p. 063504, 2011.10.1063/1.3634113Suche in Google Scholar
[33] M. Wang and Z. Y. Guo, “Understanding of temperature and size dependences of effective thermal conductivity of nanotubes,” Phys. Lett. A, vol. 374, no. 42, pp. 4312–4315, 2010. https://doi.org/10.1016/j.physleta.2010.08.058.Suche in Google Scholar
[34] B.-D. Nie, B.-Y. Cao, Z.-Y. Guo, and Y.-C. Hua, “Thermomass theory in the framework of GENERIC,” Entropy, vol. 22, no. 2, p. 227, 2020. https://doi.org/10.3390/e22020227.Suche in Google Scholar PubMed PubMed Central
[35] Y. Dong, B.-Y. Cao, and Z.-Y. Guo, “General expression for entropy production in transport processes based on the thermomass model,” Phys. Rev. E, vol. 85, no. 6, p. 061107, 2012. https://doi.org/10.1103/physreve.85.061107.Suche in Google Scholar PubMed
[36] A. Baranyai, “Temperature of nonequilibrium steady-state systems,” Phys. Rev. E, vol. 62, no. 5, p. 5989, 2000. https://doi.org/10.1103/physreve.62.5989.Suche in Google Scholar PubMed
[37] T. Hatano and D. Jou, “Measuring nonequilibrium temperature of forced oscillators,” Phys. Rev. E, vol. 67, no. 2, p. 026121, 2003. https://doi.org/10.1103/physreve.67.026121.Suche in Google Scholar
[38] Y. Dong, “Clarification of Onsager reciprocal relations based on thermomass theory,” Phys. Rev. E, vol. 86, no. 6, p. 062101, 2012. https://doi.org/10.1103/physreve.86.062101.Suche in Google Scholar PubMed
[39] Y.-C. Hua, T.-W. Xue, and Z.-Y. Guo, “Reversible reciprocal relation of thermoelectricity,” Phys. Rev. E, vol. 103, no. 1, p. 012107, 2021. https://doi.org/10.1103/physreve.103.012107.Suche in Google Scholar
[40] N. Mingo and D. Broido, “Length dependence of carbon nanotube thermal conductivity and the “problem of long waves”,” Nano Lett., vol. 5, no. 7, pp. 1221–1225, 2005. https://doi.org/10.1021/nl050714d.Suche in Google Scholar PubMed
[41] Y. Dong, B.-Y. Cao, and Z.-Y. Guo, “Size dependent thermal conductivity of Si nanosystems based on phonon gas dynamics,” Physica E Low Dimens. Syst. Nanostruct., vol. 56, pp. 256–262, 2014. https://doi.org/10.1016/j.physe.2013.10.006.Suche in Google Scholar
[42] Y. Ju and K. Goodson, “Phonon scattering in silicon films with thickness of order 100 nm,” Appl. Phys. Lett., vol. 74, no. 20, pp. 3005–3007, 1999. https://doi.org/10.1063/1.123994.Suche in Google Scholar
[43] Y. S. Ju, “Phonon heat transport in silicon nanostructures,” Appl. Phys. Lett., vol. 87, no. 15, p. 153402, 2005. https://doi.org/10.1063/1.2089178.Suche in Google Scholar
[44] D. Li, Y. Wu, P. Kim, L. Shi, P. Yang, and A. Majumdar, “Thermal conductivity of individual silicon nanowires,” Appl. Phys. Lett., vol. 83, no. 14, pp. 2934–2936, 2003. https://doi.org/10.1063/1.1616981.Suche in Google Scholar
[45] A. I. Hochbaum, R. Chen, R. D. Delgado, W. Liang, E. C. Garnett, M. Najarian, A. Majumdar, and P. Yang, “Enhanced thermoelectric performance of rough silicon nanowires,” Nature, vol. 451, no. 7175, pp. 163–167, 2008. https://doi.org/10.1038/nature06381.Suche in Google Scholar PubMed
[46] A. I. Boukai, Y. Bunimovich, J. Tahir-Kheli, J.-K. Yu, W. A. GoddardIii, and J. R. Heath, “Silicon nanowires as efficient thermoelectric materials,” Nature, vol. 451, no. 7175, pp. 168–171, 2008. https://doi.org/10.1038/nature06458.Suche in Google Scholar PubMed
[47] M. S. Dresselhaus, G. Chen, M. Y. Tang, R. Yang, H. Lee, D. Wang, Z. Ren, J. P. Fleurial, and P. Gogna, “New directions for low‐dimensional thermoelectric materials,” Adv. Mater., vol. 19, no. 8, pp. 1043–1053, 2007. https://doi.org/10.1002/adma.200600527.Suche in Google Scholar
[48] G. Chen, M. Dresselhaus, G. Dresselhaus, J.-P. Fleurial, and T. Caillat, “Recent developments in thermoelectric materials,” Int. Mater. Rev., vol. 48, no. 1, pp. 45–66, 2003. https://doi.org/10.1179/095066003225010182.Suche in Google Scholar
[49] Y. Dong, B. Cao, and Z. Guo, “Thermomass theory: a mechanical pathway to analyze anomalous heat conduction in nanomaterials,” in Nanomaterials, Rijeka, IntechOpen, 2017.10.5772/67780Suche in Google Scholar
[50] P. Rogolino, A. Sellitto, and V. Cimmelli, “Influence of nonlinear effects on the efficiency of a thermoelectric generator,” Z. Angew. Math. Phys., vol. 66, pp. 2829–2842, 2015. https://doi.org/10.1007/s00033-015-0516-z.Suche in Google Scholar
[51] V. A. Cimmelli and P. Rogolino, “New and recent results for thermoelectric energy conversion in graded alloys at nanoscale,” Nanomaterials, vol. 12, no. 14, p. 2378, 2022. https://doi.org/10.3390/nano12142378.Suche in Google Scholar PubMed PubMed Central
[52] H. M. Youssef and E. A. Al-Lehaibi, “General generalized thermoelasticity theory (GGTT),” J. Therm. Anal. Calorim., vol. 1, pp. 5917–5926, 2023. https://doi.org/10.1007/s10973-023-12144-x.Suche in Google Scholar
[53] M. Wong, C. Tso, T. Ho, and H. Lee, “A review of state of the art thermal diodes and their potential applications,” Int. J. Heat Mass Transfer, vol. 164, p. 120607, 2021. https://doi.org/10.1016/j.ijheatmasstransfer.2020.120607.Suche in Google Scholar
[54] H. Liu, H. Wang, and X. Zhang, “A brief review on the recent experimental advances in thermal rectification at the nanoscale,” Appl. Sci., vol. 9, no. 2, p. 344, 2019. https://doi.org/10.3390/app9020344.Suche in Google Scholar
[55] S. Zhao, Y. Zhou, and H. Wang, “Review of thermal rectification experiments and theoretical calculations in 2D materials,” Int. J. Heat Mass Transfer, vol. 195, p. 123218, 2022. https://doi.org/10.1016/j.ijheatmasstransfer.2022.123218.Suche in Google Scholar
[56] C. W. Chang, D. Okawa, A. Majumdar, and A. Zettl, “Solid-state thermal rectifier,” Science, vol. 314, no. 5802, pp. 1121–1124, 2006. https://doi.org/10.1126/science.1132898.Suche in Google Scholar PubMed
[57] N. Yang, G. Zhang, and B. Li, “Carbon nanocone: a promising thermal rectifier,” Appl. Phys. Lett., vol. 93, no. 24, p. 243111, 2008. https://doi.org/10.1063/1.3049603.Suche in Google Scholar
[58] N. Yang, G. Zhang, and B. Li, “Thermal rectification in asymmetric graphene ribbons,” Appl. Phys. Lett., vol. 95, no. 3, p. 033107, 2009. https://doi.org/10.1063/1.3183587.Suche in Google Scholar
[59] H. Wang, S. Hu, K. Takahashi, X. Zhang, H. Takamatsu, and J. Chen, “Experimental study of thermal rectification in suspended monolayer graphene,” Nat. Commun., vol. 8, no. 1, p. 15843, 2017. https://doi.org/10.1038/ncomms15843.Suche in Google Scholar PubMed PubMed Central
[60] Y. Zhang, Q. Lv, H. Wang, S. Zhao, Q. Xiong, R. Lv, and X. Zhang, “Simultaneous electrical and thermal rectification in a monolayer lateral heterojunction,” Science, vol. 378, no. 6616, pp. 169–175, 2022. https://doi.org/10.1126/science.abq0883.Suche in Google Scholar PubMed
[61] Y. Dong, C. Diao, Y. Song, H. Chi, D. J. Singh, and J. Lin, “Molecular bridge thermal diode enabled by vibrational mismatch,” Phys. Rev. Appl., vol. 11, no. 2, p. 024043, 2019. https://doi.org/10.1103/physrevapplied.11.024043.Suche in Google Scholar
[62] B. Li, J. Lan, and L. Wang, “Interface thermal resistance between dissimilar anharmonic lattices,” Phys. Rev. Lett., vol. 95, no. 10, p. 104302, 2005. https://doi.org/10.1103/physrevlett.95.104302.Suche in Google Scholar PubMed
[63] M. Schmotz, J. Maier, E. Scheer, and P. Leiderer, “A thermal diode using phonon rectification,” New J. Phys., vol. 13, no. 11, p. 113027, 2011. https://doi.org/10.1088/1367-2630/13/11/113027.Suche in Google Scholar
© 2024 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Artikel in diesem Heft
- Frontmatter
- Editorial
- Thermodynamic costs of temperature stabilization in logically irreversible computation
- Hydrodynamic, electronic and optic analogies with heat transport in extended thermodynamics
- Variations on the models of Carnot irreversible thermomechanical engine
- Revisit nonequilibrium thermodynamics based on thermomass theory and its applications in nanosystems
- On the dynamic thermal conductivity and diffusivity observed in heat pulse experiments
- On the influence of the fourth order orientation tensor on the dynamics of the second order one
- Lack-of-fit reduction in non-equilibrium thermodynamics applied to the Kac–Zwanzig model
- Optimized quantum drift diffusion model for a resonant tunneling diode
- The wall effect in a plane counterflow channel
- Buoyancy driven convection with a Cattaneo flux model
- Thermodynamics of micro- and nano-scale flow and heat transfer: a mini-review
- Poroacoustic front propagation under the linearized Eringen–Cattaneo–Christov–Straughan model
Artikel in diesem Heft
- Frontmatter
- Editorial
- Thermodynamic costs of temperature stabilization in logically irreversible computation
- Hydrodynamic, electronic and optic analogies with heat transport in extended thermodynamics
- Variations on the models of Carnot irreversible thermomechanical engine
- Revisit nonequilibrium thermodynamics based on thermomass theory and its applications in nanosystems
- On the dynamic thermal conductivity and diffusivity observed in heat pulse experiments
- On the influence of the fourth order orientation tensor on the dynamics of the second order one
- Lack-of-fit reduction in non-equilibrium thermodynamics applied to the Kac–Zwanzig model
- Optimized quantum drift diffusion model for a resonant tunneling diode
- The wall effect in a plane counterflow channel
- Buoyancy driven convection with a Cattaneo flux model
- Thermodynamics of micro- and nano-scale flow and heat transfer: a mini-review
- Poroacoustic front propagation under the linearized Eringen–Cattaneo–Christov–Straughan model