Skip to main content

ORIGINAL RESEARCH article

Front. Energy Res., 19 April 2023
Sec. Energy Storage

Influence of fiber hollowness on the local thermo-electro-elastic field in a thermoelectric composite

Chuanbin Yu
Chuanbin Yu1*Haoxin LiuHaoxin Liu1Chaofan DuChaofan Du1Shichao Xing,,
Shichao Xing2,3,4*
  • 1College of Architectural Science and Engineering, Yangzhou University, Yangzhou, Jiangsu, China
  • 2School of Mechanical Engineering, Tongling University, Tongling, Anhui, China
  • 3Key Laboratory of Construction Hydraulic Robots of Anhui Higher Education Institutes, Tongling University, Tongling, Anhui, China
  • 4Key Laboratory of Additive Manufacturing of Tongling City, Tongling University, Tongling, Anhui, China

The fiber geometry is one of the important parameters in the effective conversion performance and local strength of thermoelectric composites. In this study, the plane problem of a hollow fiber embedded within a non-linear thermoelectric medium in the presence of a uniform remote in-plane electric current and a uniform remote energy flux is investigated based on the complex variable method. Closed-form expressions for all the potential functions characterizing the thermoelectric field and the associated thermal stress field in both the matrix and fiber are obtained by solving the corresponding boundary value problem. Numerical examples are presented to illustrate the effect of hollowness ratio of the fiber on the local energy conversion efficiency and interfacial thermal stress concentration. It is found that a higher conversion efficiency and a lower amount of thermal stress concentration around a hollow fiber than that around a solid fiber could be achieved simultaneously by appropriate selection of the hollowness ratio of the fiber. The results can be directly used for performance optimization and reliability evaluation in design of thermoelectric composites in engineering.

1 Introduction

Thermoelectric material is a novel kind of functional material, which can directly convert heat into electricity and vice versa (Bell, 2008). A thermoelectric device is free of moving parts, noises, and environmentally harmful emissions, making it extremely attractive in renewable energy fields, such as waste heat recovery, solar power generation, carbon reduction, and Freon-free refrigeration (Zhang et al., 2016; He and Tritt, 2017; Mahmoudinezhad et al., 2018). The electric transport and thermal conduction in a thermoelectric material are intrinsically coupled, making it highly challenging to improve the conversion efficiency in a single-phase thermoelectric material (Yang et al., 2013). Therefore, a lot of effort in developing high-performance thermoelectric material is devoted to engineering hybrid composites by introducing inclusions/fibers into single-phase thermoelectric material to enhance the overall conversion efficiency (Wan et al., 2015; Guo et al., 2016; Srivastava et al., 2018).

From the point of view of solid mechanics, the existence of the inclusions/fibers may cause severe concentration of the temperature field and the associated thermal stress field in thermoelectric materials under in-service conditions. Most thermoelectric materials are semiconductors with poor mechanical properties, and excessive thermal stress may lead to the premature failure of the thermoelectric devices (Li et al., 2015; Song et al., 2019). The analysis of thermoelectric materials with the inclusions/fibers can further improve our understanding of the mechanical behavior of such new materials and thus is essential for optimal design and improved reliability of thermoelectric composites. Using complex variable methods, rigorous thermo-electro-mechanical analyses of thermoelectric materials containing elastic inhomogeneities have been carried out by several authors (Zhang et al., 2017; Wang et al., 2018; Yang et al., 2020). In the corresponding discussions of the inhomogeneity problem in thermoelectric materials, it has been assumed that the inclusion or fiber is assumed to have a solid cross-section. In engineering practice, hollow fibers or particles are often introduced into thermoelectric materials due to their distinct microstructures. To the best of our knowledge, the effect of a hollow particle/fiber on the thermoelectric and thermoelastic responses of a thermoelectric composite has never been thoroughly investigated. This motivates our current work.

The paper is organized as follows: basic equations of a non-linear thermoelectric material are presented in Section 2, followed by the detailed analysis of the plane problem of an infinite thermoelectric matrix containing a hollow fiber subjected to a uniform remote thermoelectric loading by using the complex variable technique in Section 3. In Section 4, we use the explicit results from Section 3 and present numerical examples to illustrate the influence of the hollowness ratio on the local energy conversion efficiency and thermal stress concentration around a hollow fiber in a thermoelectric composite. Finally, we summarize our conclusion in Section 5.

2 Basic equations for thermoelectric materials

In a homogenous thermoelectric medium wherein both electric charges and energy are conserved, the governing equations and transport equations for the non-linearly coupled electric and heat conduction are given as follows (Yang et al., 2013):

I=0,J=0(1)

together with

I=δϕδεT(2)
J=q+ϕI=ϕ+εTIκT(3)

where is the Nabla operator, I=Ix,IyT is the electric current vector, J=Jx,JyT denotes the energy flux vector, and q=qx,qyT represents the heat flux vector. In addition, ϕ is the electric potential, T is the temperature, δ is the electric conductivity, κ is the thermal conductivity, and ε is the Seebeck coefficient. Inserting Eqs 2, 3 into the governing equations in Eq. 1 results in

2ϕ+εT=0(4)
2T+δκϕ+εTϕ+εT=0(5)

According to Eqs 4, 5, the electric potential ϕ and temperature T can be expressed in terms of two potential functions, namely, hz and gz, of the complex variable z=x+iy (i=1 is the imaginary unit) as follows (Yu et al., 2018):

ϕ=RehzεRegz+δε4κhzhz¯(6)
T=Regzδ4κhzhz¯(7)

where “Re” denotes the real part of a complex number and the overbar denotes the complex conjugate. It follows from Eqs 2, 3, 6, 7 that the electric current components Ix,Iy, the energy flux components Jx,Jy, and the heat flux components qx,qy are expressed in terms of the two complex functions as follows (Yu et al., 2018):

IxiIy=δhz(8)
JxiIy=κgzδ2hzhz(9)
qxiqy=κgz+δ2hzhz¯+εδhzδ4κhzhz¯Regz(10)

where the prime () denotes the derivative of the complex variable z. In addition, the resultant electric current and energy flux on a certain directed curve s can be expressed as follows (Yu et al., 2018):

Insds=δImhz(11)
Jnsds=κImgz+δ4κh2z(12)

which can be used as the boundary conditions of the electric current and energy flux in the integral form. Here, “Im” denotes the imaginary part of a complex number.

For a steady thermal stress problem where the elastic constants are assumed to be temperature-independent, the three in-plane stresses (σx,σy,τxy) and the two in-plane displacements (ux,uy) caused by the uneven temperature distribution Tx,y given in Eq. 7 can be expressed in terms of two elastic potential functions, namely, φz and ψz, together with the aforementioned thermoelectric potential functions hz and gz as follows (Yu et al., 2018):

σy+σx=2φz+φz¯+2ξhzhz¯(13a)
σyσx+2iτxy=2z¯φz+ψz+2ξhzHz¯(13b)
2μux+iuy=βφzzφz¯ψz¯ξhz¯Hz+2μλGz(14a)

Here, Hz and Gz are two complex functions satisfying Hz=hz and Gz=gz, μ is the shear modulus, ξ=μλvδ/4κ is a real constant, β=34v*, λ=1+v*λ*, and λv=1+v*λ*/1v* for plane strain, β=3v*/1+v*, λ=λ*, and λv=1+v*λ* for plane stress with λ* representing the thermal expansion coefficient and v* being the Poisson’s ratio. Additionally, the resultant forces (fx, fy) on a certain directed curve s can be calculated as

ifx+ifyds=φz+zφz¯+ψz¯+ξhz¯Hz(14b)

which will be used as the stress boundary condition in the integral form. To summarize, the mathematical formulation of the non-linearly coupled thermo-electro-mechanical problem is completed using the complex function theory with each unknown function determined by the corresponding boundary conditions.

3 Solutions of thermoelectric material with a hollow fiber

As illustrated in Figure 1, we consider the plane strain deformation of an infinite thermoelectric matrix enclosing a hollow fiber subjected to a uniform remote electric current Iy and a uniform remote energy flux Jy. Let S1 and S2 denote the regions occupied by the surrounding matrix (zR1) and the hollow fiber (R2zR1), respectively. The fiber and matrix are assumed to be perfectly bonded, and both the electric and heat conduction can penetrate the interface L1. The continuity conditions of the thermoelectric field across L1 can be expressed as follows (Wang et al., 2018):

h1z1+h1z1¯=h2z1+h2z1¯(15a)
h1z1h1z1¯=δ2δ1h2z1h2z1¯(15b)
g1z1+g1z1¯δ12κ1h1z1h1z1¯=g2z1+g2z1¯δ22κ2h2z1h2z1¯(16a)
g1z1g1z1¯+δ14κ1h1z12h1z1¯2=κ2κ1g2z1g2z1¯+δ24κ2h2z12h2z1¯2(16b)

where z1=R1eiθ represents those points on the interface L1 with θ denoting the angle between the outer normal vector and the positive x-axis in the z-plane as shown in Figure 1; the index 1 or 2 indicates the corresponding quantity defined in either the matrix (region S1) or the hollow fiber (region S2), respectively. Furthermore, the inner surface of the hollow fiber is assumed to be electrically impermeable and thermally insulated

h2z2h2z2¯=0(17a)
g2z2g2z2¯+δ24κ2h2z22h2z2¯2=0(17b)

where z2=R2eiθ represents those points on the inner surface L2. In ensuing analysis, we first solve the thermoelectric potentials h1z and h2z from the electric boundary conditions in Eqs 15a, 15b, 17a and then substitute them into the thermal boundary conditions in Eqs 16a, 16b, 17b to determine g1z and g2z. Considering Eq. 8 and the far-field condition of the electric current, the potential function h1z takes the following form:

h1z=iIyδ1z+U0+h1*z(18)

where U0 is a real constant related to the equilibrium thermoelectric potential and is set as zero (i.e., U0=0) without loss of generality in this paper. The analytical function h1*z satisfying h1*z=0 represents the disturbance on the electric field caused by the fiber. The region occupied by the matrix is the exterior of contour L1; thus, h1*z can be expanded into Fourier series as h1*z=n=1+anzn, where an is the complex coefficient to be determined. The region occupied by the fiber can be regarded as the intersection of the exterior of contour L2 and the interior of contour L1, and the potential function h2z then can be expanded as h2z=+bnzn with bn being the unknown coefficients to be determined from the continuity conditions in Eqs 15a, 15b, 17a. Omitting details, the thermoelectric potential functions are finally determined as follows:

h1z=A1zR1+A1zR11(19a)
h2z=C1zR1+αC1¯zR11(19b)

where A1=iIyR1/δ1 and α is the hollowness ratio defined by the hollow area over the total area of the cross-section as α=R2/R12. In addition, A1 and C1 are two complex constants calculated by

A1=α1+δ2δ1+1δ2δ11+δ2δ1+α1δ2δ1A1¯,C1=21+δ2δ1+α1δ2δ1A1

FIGURE 1
www.frontiersin.org

FIGURE 1. Hollow fiber embedded in a thermoelectric matrix under uniform remote in-plane thermoelectric loading.

Considering that the energy flux at infinity is uniform, according to Eqs 9, 18, the potential function g1z takes the form of

g1z=Iy24κ1δ1z2+iJyκ1z+T0+g1*z(20)

where the real constant T0 denotes the uniform temperature field in the matrix and the analytical function g1*z satisfying g1*z=0 represents the disturbance on the thermal field caused by the fiber. The same argument applies to Eqs 16a, 16b, 17b, we obtain

g1z=B2zR12+B1zR1+T0+B1zR11+B2zR12(21a)
g2z=D2zR12+D1zR1+D0+αD1¯zR11+α2D2¯zR12(21b)

where

B2=D2¯21+κ2κ1α2+1κ2κ1+Δ22,B1=D1¯21+κ2κ1α+1κ2κ1
D2=2B2Δ21+κ2κ1+1κ2κ1α2,D1=2B11+κ2κ1+1κ2κ1α
D0=T0δ14κ1A12+A12+δ24κ21+α2C12

Here, the two complex constants, namely, 2 and 2, are defined as

Δ2=δ14κ12A1A1¯A12+A1¯2δ24κ22ακ2κ11α2C12
Δ2=δ14κ12A1A1¯A12+A1¯2δ24κ22α+κ2κ11α2C1¯2

So far, the quantities related to the electric current and temperature fields in both the matrix and fiber can be determined from Eqs 6 to 10 together with Eqs 19, 21a. The following studies will deal with the analysis of the associated thermal stress problem. Considering the single-valued conditions of elastic displacement (Eq. 14) and the resultant force (Eq. 15), the elastic potential functions in both the matrix and fiber take the following form:

φ1z=P0lnzR1+φ1*z(22a)
ψ1z=Q0zlnzR1+ψ1*z(22b)
φ2z=Y0lnzR1+φ2*z(23a)
ψ2z=W0zlnzR1+ψ2*z(23b)

where φj*z and ψj*z (j = 1, 2) are holomorphic functions in the respective regions. Taking Eqs 22, 23 into Eqs. 14, 15 and eliminating the multi-valued terms, we can obtain

P0=2μ1λ1β1+1B1R1,Q0z=P0¯+ξ1A1¯R1h1z
Y0=2μ2λ2β2+1αD1¯R1,W0z=Y0¯+ξ2αC1R1h2z

The analytical functions φj*z and ψj*z (j = 1, 2) in Eqs 22, 23 will be determined by the following elastic boundary conditions:

φ1z1+z1φ1z1¯+ψ1z1¯+ξ1h1z1¯H1z1=φ2z1+z1φ2z1¯+ψ2z1¯+ξ2h2z1¯H2z1(24a)
1μ1β1φ1z1z1φ1z1¯ψ1z1¯ξ1h1z1¯H1z1+2μ1λ1G1z1=1μ2β2φ2z1z1φ2z1¯ψ2z1¯ξ2h2z1¯H2z1+2μ2λ2G2z1(24b)
φ2z2+z2φ2z2¯+ψ2z2¯+ξ2h2z2¯H2z2=0(25)

Equations 24a, 24b describe the continuity of elastic stresses and displacements across the perfectly bonded interface L1, and Eq. 25 indicates that the inner surface of the fiber L2 is mechanically free. By enforcing the boundary conditions in Eqs 24, 25 with the aid of Eqs 19, 21b, we arrive at

φ1z=P0lnzR1+ω2Y1+Ω1zR11(26a)
ψ1z=Q0zlnzR1+ω1Y1¯+Ω1¯zR11+ω1Y2¯+Ω2¯zR12+ω2Y1+Ω1+ω1Y3¯+Ω3¯zR13(26b)
φ2z=Y0lnzR1+Y1zR11+Y1zR11+Y2zR12+Y3zR13(27a)
ψ2z=W0zlnzR1+Σ3¯+Y1ω3Y3¯zR13+Σ2¯ω3Y2¯zR12+Σ1¯Y1ω3Y1¯zR11+ω2ω4ω3Y1¯+Σ1¯3Y3+ω4Ω1¯zR1(27b)

where the four real dimensionless constants ω1, ω2, ω3, and ω4 are given as

ω1=β2+11μ2μ1ω2=β2+11+β1μ2μ1ω3=β2+μ2μ1μ2μ11ω4=μ2μ1β1+1μ2μ11

and the four complex constants Y1, Y2, Y3 and Y1 are identified as

Y1=Σ1Λ1ω3α+1αY2=Σ2+αY0¯ω3α2
Y3=αω4Ω1¯+αΣ1¯Λ2+1αω3ω2ω4Y1¯3α1α
Y1=α3ω3Λ2αω4Ω1+Σ1+3α1αΛ3¯Σ3¯α3ω31αω3ω2ω4+3α1α2

together with

Λ3=12α3ξ2R1C12
Λ2=α2ξ2R1C1¯2
Λ1=12+lnαα2ξ2R1C12
Ω3=2μ2R131μ2μ1λ2D2λ1B2ξ1R12A1A1¯
Ω2=μ2R11μ2μ1λ2D1λ1B1P0¯
Ω1=2μ2R11μ2μ1λ2D0λ1B0ξ1R12A12
Ω1=2μ2R11+β1μ2μ1λ1B2λ2α2D2¯
Σ3=2μ2R131μ2μ1λ2D2λ1B2ξ2R12αC12
Σ2=μ2R11μ2μ1λ2D1λ1B1Y0¯
Σ1=2μ2R11μ2μ1λ2D0λ1B0ξ2R12C12
Σ1=2μ2R11μ2μ1λ1B2λ2α2D2¯

Up to here, the elastic stress components σx, σy, and τxy in both the matrix and fiber are determined from Eq. 13 with the aid of the two thermoelectric potential functions and two elastic potential functions, and the stress components in polar coordinates can be determined according to Eq. 13 and the following stress transformation formula:

σtt+σnn=σx+σy(28a)
σttσnn+2iτnt=e2iθσyσx+2iτxy(28b)

where σnn, σtt, and τnt are the normal stress, hoop stress, and shear stress, respectively.

4 Numerical examples

To illustrate the application of the analysis model, some numerical examples are presented. In the following examples, the matrix is set as a bismuth tellurium semiconductor, and the material constants are given as δ1=1.67×105Sm1, κ1=2Wm1K1, ε1=2×104VK1, μ1=60GPa, v1=0.3, and λ1=4.2×106K1 (Huang et al., 2008). The imposed far-field loadings are Iy=2×104Am2 and Jy=8×103Wm2, and the average temperature on the matrix is set as T0=350K. In addition, the Seebeck coefficient and Poisson’s ratio of the fiber are taken as the same with those of the matrix material.

In engineering practice, the thermoelectric conversion efficiency is a crucial parameter for the design of thermoelectric composites. Figure 2 shows the effect of the hollowness ratio α on the local energy conversion efficiency η around a hollow fiber with different electric and thermal conductivity properties. A finite region (10mm×10mm) bounded by ABCD around the hollow fiber is selected to calculate the local thermoelectric conversion efficiency, as shown in Figure 1. The thermoelectric conversion efficiency in the area ABCD is given as follows (Liu, 2012):

η=qinputqoutputqinput=qCDqABmaxqCD,qAB

where qAB=ABqyxdx and qCD=CDqyxdx denote the total heat flux on border AB and CD, respectively.

FIGURE 2
www.frontiersin.org

FIGURE 2. Effect of the hollowness ratio and conductivity properties of the fiber on the local energy conversion efficiency.

As shown in Figure 2, with specific fiber radius R1=4mm, when the conductivity properties and hollowness ratio of the fiber are changed, different values of the local energy conversion efficiency are obtained. It can be found that the local energy conversion efficiency increased when the electric conductivity of the fiber increased and reduced when the thermal conductivity of the fiber increases. Thus, we can conclude that fibers with higher electric conductivity and lower thermal conductivity are desirable for the design of thermoelectric composites. By comparing Figure 2D with Figure 2A, we found that the thermal conductivity of the fiber has a more dominant effect on the local thermoelectric conversion efficiency than the electric conductivity when the hollowness ratio α is relative small.

Figure 3 illustrates the variation of local energy conversion efficiency around a hollow fiber with desirable conductivity properties with respect to the hollowness ratio α. For all values of the fiber hollowness ratio, α=0.45 leads to the highest energy conversion efficiency, as shown in Figure 3A. We further compared the conversion efficiency around the hollow fiber with different α values to that around a solid fiber (namely, α=0) under the same service conditions as shown in Figure 3B. As shown in this figure, compared with a solid fiber with radius of 4.5 mm, the highest conversion efficiency around a hollow fiber with α=0.45 is increased by 22.3%, showing the tremendous value of hollow fibers in improving the conversion performance of a thermoelectric composite material.

FIGURE 3
www.frontiersin.org

FIGURE 3. (A) Local energy conversion efficiency; (B) relative energy conversion efficiency around a desirable fiber with different hollowness ratios.

In addition to the conversion performance, the mechanical reliability also deserved further consideration when designing a thermoelectric composite. Figure 4 presents the variation of the interfacial normal stress distribution around a softer hollow fiber with desirable conductivity properties. Here, the soft fiber has a lower elastic modulus but a larger linear expansion coefficient than the matrix material. Since the thermal expansion of a softer fiber is greater than that of the matrix material, the matrix will prevent the fiber from expanding outward, and the interface is under compression, as shown in Figure 4A. In addition, the maximum compressive normal stress is achieved when θ=1.5π in this case, which may cause extrusion damage at this point and must be avoided in design. For given fiber–matrix system, as the fiber hollowness ratio α increases, the maximum normal stress at the interface is reduced. The maximum normal stress in the vicinity of a hollow fiber is then compared to that of a solid fiber and the results are illustrated in Figure 4B. The relative concentration factor reaches 0.6 for a soft fiber of radius R1=3mm, which means that the normal stress concentration around the hollow fiber is 40% lower than that around a solid fiber under the same service condition. In contrast to the soft fiber, the interfacial normal stress becomes tensile around a hollow fiber with harder elastic property, as shown in Figure 5A, which implies that the interface failure mode is more likely to be interface de-bonding in this case. Moreover, as shown in Figure 5B, we found that the concentration of interfacial normal stress can still be suppressed by increasing the fiber hollowness ratio in thermoelectric composites reinforced by hard fibers.

FIGURE 4
www.frontiersin.org

FIGURE 4. (A) Distribution of normal stress; (B) relative concentration factor of normal stress around a soft fiber with different hollowness ratios.

FIGURE 5
www.frontiersin.org

FIGURE 5. (A) Distribution of normal stress; (B) relative concentration factor of normal stress around a hard fiber with different hollowness ratios.

5 Concluding remarks

In this paper, the problem of a hollow fiber embedded in an infinite thermoelectric matrix under the remote uniform electric current and energy flux is investigated. The closed-form solutions of the electric field, temperature field, and associated thermal stress field in the entire composite are derived via the use of a complex variable technique. The following conclusions are summarized for designing and optimizing thermoelectric composites reinforced by hollow fibers/particles:

1. Fibers with lower thermal conductivity are helpful to improve the energy conversion efficiency of thermoelectric composites, but it may aggravate the interfacial thermal stress concentration.

2. A higher thermoelectric conversion efficiency and a lower amount of the normal stress concentration than that around a solid fiber could be achieved simultaneously by appropriate selection of the hollowness ratio of the fiber.

Data availability statement

The original contributions presented in the study are included in the article/Supplementary Material; further inquiries can be directed to the corresponding authors.

Author contributions

CY and SX contributed to the conception and design of the study. CY performed the formal analysis, and wrote the first draft of the manuscript. HL and CD wrote sections of the manuscript. CY, HL, CD, and SX contributed to manuscript revision and supervision and read and approved the submitted version.

Funding

This work was supported by the National Natural Science Foundation of China (grant no. 11802263&11902116), Natural Science Foundation of Jiangsu Province (grant no. BK20210787), Natural Science Foundation of Guangdong Province (grant no. 2022A1515011773), Natural Science Foundation of Guangzhou City (grant no. 202201010317), Qinglan Project of Yangzhou University, Provincial Natural Science Research of University of Anhui Province (KJ 2021A1058), and Key Laboratory of Construction Hydraulic Robots of Anhui Higher Education Institutes (grant no. TLXYCHR-O-21YB04).

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Bell, L. E. (2008). Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 321 (5895), 1457–1461. doi:10.1126/science.1158899

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, H., Xin, H., Qin, X., Zhang, J., Li, D., Li, Y., et al. (2016). Enhanced thermoelectric performance of highly oriented polycrystalline SnSe based composites incorporated with SnTe nanoinclusions. J. Alloy. Compd. 689, 87–93. doi:10.1016/j.jallcom.2016.07.291

CrossRef Full Text | Google Scholar

He, J., and Tritt, T. M. (2017). Advances in thermoelectric materials research: Looking back and moving forward. Science 357 (6358), eaak9997. doi:10.1126/science.aak9997

PubMed Abstract | CrossRef Full Text | Google Scholar

Huang, M. J., Chou, P. K., and Lin, M. C. (2008). An investigation of the thermal stresses induced in a thin-film thermoelectric cooler. J. Therm. Stress. 31 (5), 438–454. doi:10.1080/01495730801912512

CrossRef Full Text | Google Scholar

Li, G., An, Q., LiGoddard, W. A., Zhai, P., Zhang, Q., Snyder, G. J., et al. (2015). Brittle failure mechanism in thermoelectric skutterudite CoSb3. Chem. Mat. 27 (18), 6329–6336. doi:10.1021/acs.chemmater.5b02268

CrossRef Full Text | Google Scholar

Liu, L. (2012). A continuum theory of thermoelectric bodies and effective properties of thermoelectric composites. Int. J. Eng. Sci. 55, 35–53. doi:10.1016/j.ijengsci.2012.02.003

CrossRef Full Text | Google Scholar

Mahmoudinezhad, S., Rezania, A., and Rosendahl, L. A. (2018). Behavior of hybrid concentrated photovoltaic-thermoelectric generator under variable solar radiation. Energy Convers. manage. 164, 443–452. doi:10.1016/j.enconman.2018.03.025

CrossRef Full Text | Google Scholar

Song, H., Song, K., and Gao, C. (2019). Temperature and thermal stress around an elliptic functional defect in a thermoelectric material. Mech. Mat. 130, 58–64. doi:10.1016/j.mechmat.2019.01.008

CrossRef Full Text | Google Scholar

Srivastava, D., Norman, C., Azough, F., Schäfer, M. C., Guilmeau, E., and Freer, R. (2018). Improving the thermoelectric properties of SrTiO3-based ceramics with metallic inclusions. J. Alloy. Compd. 731, 723–730. doi:10.1016/j.jallcom.2017.10.033

CrossRef Full Text | Google Scholar

Wan, S., Huang, X., Qiu, P., Bai, S., and Chen, L. (2015). The effect of short carbon fibers on the thermoelectric and mechanical properties of p-type CeFe4Sb12 skutterudite composites. Mat. Des. 67, 379–384. doi:10.1016/j.matdes.2014.11.050

CrossRef Full Text | Google Scholar

Wang, P., Wang, B. L., Wang, K. F., and Cui, Y. J. (2018). Analysis of inclusion in thermoelectric materials: The thermal stress field and the effect of inclusion on thermoelectric properties. Compos. Part B-Eng. 166, 130–138. doi:10.1016/j.compositesb.2018.11.120

CrossRef Full Text | Google Scholar

Yang, H., Yu, C., Tang, J., Qiu, J., and Zhang, X. (2020). Electric-current-induced thermal stress around a non-circular rigid inclusion in a two-dimensional nonlinear thermoelectric material. Acta Mech. 231 (11), 4603–4619. doi:10.1007/s00707-020-02770-z

CrossRef Full Text | Google Scholar

Yang, Y., Ma, F. Y., Lei, C. H., Liu, Y. Y., and Li, J. Y. (2013). Nonlinear asymptotic homogenization and the effective behavior of layered thermoelectric composites. J. Mech. Phys. Solids 61 (8), 1768–1783. doi:10.1016/j.jmps.2013.03.006

CrossRef Full Text | Google Scholar

Yu, C., Zou, D., Li, Y., Yang, H., and Gao, C. (2018). An arc-shaped crack in nonlinear fully coupled thermoelectric materials. Acta Mech. 229 (5), 1989–2008. doi:10.1007/s00707-017-2099-6

CrossRef Full Text | Google Scholar

Zhang, A. B., Wang, B. L., Wang, J., and Du, J. K. (2017). Two-dimensional problem of thermoelectric materials with an elliptic hole or a rigid inclusion. Int. J. Therm. Sci. 117, 184–195. doi:10.1016/j.ijthermalsci.2017.03.020

CrossRef Full Text | Google Scholar

Zhang, Q. H., Huang, X. Y., Bai, S. Q., Shi, X., Uher, C., and Chen, L. D. (2016). Thermoelectric devices for power generation: Recent progress and future challenges. Adv. Eng. Mat. 18 (2), 194–213. doi:10.1002/adem.201500333

CrossRef Full Text | Google Scholar

Keywords: thermoelectric materials, complex variable method, hollow fiber, conversion efficiency, thermal stress

Citation: Yu C, Liu H, Du C and Xing S (2023) Influence of fiber hollowness on the local thermo-electro-elastic field in a thermoelectric composite. Front. Energy Res. 11:1181191. doi: 10.3389/fenrg.2023.1181191

Received: 07 March 2023; Accepted: 05 April 2023;
Published: 19 April 2023.

Edited by:

Haopeng Song, Nanjing University of Aeronautics and Astronautics, China

Reviewed by:

Wenzhi Yang, Lanzhou University, China
Jian Hua, Tsinghua University, China
Min Li, Hohai University, China
Yongjian Wang, Nanjing Agricultural University, China

Copyright © 2023 Yu, Liu, Du and Xing. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Chuanbin Yu, cbyu@yzu.edu.cn; Shichao Xing, tluxsc@tlu.edu.cn

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.