Next Article in Journal
Biophysical Characterization of Membrane Interactions of 3-Hydroxy-4-Pyridinone Vanadium Complexes: Insights for Antidiabetic Applications
Previous Article in Journal
Fibroblast Response to Cyclo- and Organic Phosphate Solutions: A Cytotoxicity Study
Previous Article in Special Issue
Bismuth-Doped Indium Oxide as a Promising Thermoelectric Material
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermoelectric Figure of Merit in a One-Dimensional Model with k4-Dispersion: An Extension of the Theory by Hicks and Dresselhaus

1
National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba 305-8568, Ibaraki, Japan
2
Department of Physics, University of Tokyo, Bunkyo, Tokyo 113-0033, Japan
*
Author to whom correspondence should be addressed.
Inorganics 2025, 13(9), 310; https://doi.org/10.3390/inorganics13090310
Submission received: 31 July 2025 / Revised: 3 September 2025 / Accepted: 11 September 2025 / Published: 22 September 2025
(This article belongs to the Special Issue Advances in Thermoelectric Materials, 2nd Edition)

Abstract

Motivated by the strategy developed by Hicks and Dresselhaus in a quantum wire corresponding to a single-chain model with k 2 -dispersion, we study a one-dimensional double-chain model with two carriers of electrons and holes, characterized by k 4 -dispersion. To understand the role of the enhancement of the density of state derived from k 4 -dispersion, we calculate an optimized dimensionless thermoelectric figure of merit ( Z T ) depending on the side length of the cross section, a, in the same way as discussed by Hicks and Dresselhaus. We find that Z T enhances as a decreases similarly to the results obtained in the single-chain model, while the enhancement of Z T is smaller than that of single-chain model. We discuss the reason in connection with the difference of electronic state between the single- and double-chain models.

Graphical Abstract

1. Introduction

The search for materials with a high thermoelectric figure of merit, Z T , is one of the most important tasks in thermoelectric research. In 1993, Hicks and Dresselhaus discussed the promising possibility of high Z T in low-dimensional systems [1,2]. Guided by their proposal, many experimental and theoretical studies have been carried out to look for the appropriate materials [3,4,5]. In particular, in one-dimensional systems such as quantum wires and carbon nanotubes, it was discussed that phonon scattering from the surfaces of one-dimensional wires reduces lattice thermal conductivity, which enhances Z T . Furthermore, in one-dimensional electron systems, the density of state at the band edge has a divergence, which also gives a chance to enhance the Seebeck coefficient that is related to the energy derivative of the density of state. In the present paper, we extend this strategy of Hicks and Dresselhaus to electron systems with k 4 -dispersion, which leads to increased divergence of the density of state at the band edge. We investigate theoretically Z T in comparison with the results in the k 2 -dispersion.

2. Inorganic Materials as a Candidate of k 4 -Dispersion

As discussed in the introduction, there are several candidate materials of the Hicks–Dresselhaus theory based on the k 2 dispersion. Firstly, Hicks and Dresselhaus predicted the enhancement of ZT in the Bi3Te2 nanowire. After that, the theory was extended to not only the organic materials such as carbon nanotube, but also the inorganic materials of silicon nanowire [6], bismuth nanowire [7], etc. For example, the bismuth nanowire has been studied from both the experimental and theoretical side, because bismuth has peculiar properties of long mean free path and small effective mass [7,8,9,10,11]. Extending the Hicks–Dresselhaus theory to the bismuth nanowire, it was found that ZT improves in the realistic parameters as the diameter of nanowire decreases [8]. In contrast to the theoretical expectation, the experimental result obtained in the bismuth nanowire has a peak of Z T 0.3 at 100 nm diameter, and decreases drastically as the diameter decreases [9]. This is due to the classical and quantum size effect. Although accurate measurements of electrical and thermoelectric transports and the size (boundary) effect in the bismuth nanowire were studied [10,11], the size of the nanowire is several hundred nm and a thinner nanowire is needed to compare the theory in detail. In addition, the optimization of the carrier concentration and size effect are suggested to play an important role in realizing the prediction of Hicks–Dresselhaus.
On the other hand, there are candidate materials with k 4 -dispersion. Here, we introduce the excitonic insulator with a one-dimensional chain structure. The excitonic insulator is a condensed state of electrons and holes by the Coulomb interaction [12]. As discussed in the inorganic materials Ta2NiSe5, the excitonic insulator is suggested to appear at room temperature [13,14,15]. In Ta2NiSe5, there are three one-dimensional chains of electron and hole bands with coulomb interaction between chains [13,14]. In the excitonic insulator, the bottom of the conduction band and the top of the hole band become flat, similar to the k 4 -dispersion shown in Figure 1.
Due to the analogy with the excitonic insulator, it is possible to have the k 4 -dispersion in one-dimensional materials with the hybridization between two chains. Therefore, in this paper, we introduce the two-chain model as a minimum model with k 4 -dispersion and study the thermoelectric properties of this model using the same method as the Hicks–Dresselhaus model.

3. Model Hamiltonian

Figure 2 shows schematic pictures of (a) the single-chain model corresponding to the Hicks–Dresselhaus theory and (b) the double-chain model, consisting of two carriers of electrons and holes depending on the chain. We consider that the conductor is square in cross section with a side of length a and the one-dimensional chains are packed in the cross section of a 2 . In the following, we investigate the double-chain model shown in Figure 2b. The effective Hamiltonian is given by
H = i t c A , i + 1 c A , i + t c B , i + 1 c B , i + Δ c A , i c B , i + H . C . + i ε 0 c A , i c A , i ,
where c A , i ( c B , i ) represents the electron annihilation operator on the i-th site of the one-dimensional chain A (B), t is the hopping integral along the chains, Δ is the hopping between A and B chains, ε 0 represents the energy difference between the fermions on the chains A and B, and H.C. means the hermite conjugate terms. After the Fourier transform, this Hamiltonian is represented by a matrix form
H = ε 0 2 t cos k b Δ Δ 2 t cos k b ,
where b is the lattice constant along the chain. Here, the A (B) chain plays the role of the electron (hole) band. If we choose ε 0 = 4 t , then the two bands (without Δ ) touch at k = 0 , leading to the k 4 -dispersion when Δ is nonzero. In this case, the energy dispersion in the vicinity of k = 0 becomes
E = 2 t ± Δ 2 + t 2 ( k b ) 4 2 t ± Δ 1 + t 2 2 Δ 2 ( k b ) 4 ,
which gives k 4 -dispersion near the bottom of the conduction band and at the top of the valence band. In the following, we consider this specific case near the bottom of the conduction band, which has the dispersion, ε k = t eff ( k b ) 4 , with t eff = t 2 / 2 Δ .
The density of state in the present model diverges as ε 3 / 4 near the bottom of the band, which is stronger than ε 1 / 2 realized in the one-dimensional k 2 -dispersion. We expect that this strong divergence of the density of state can enhance the thermoelectric properties in the present model.

4. Formula of Thermoelectric Figure of Merit (ZT)

In the standard linear-response theory, the electric conductivity in one-dimensional electron system is given by
σ = 2 e 2 L a 2 k v k 2 τ f ( ε k ) ,
where v k is the velocity defined as v k = ε k / k and f ( ε ) is the Fermi distribution function f ( ε ) = 1 / [ e ( ε μ ) / k B T + 1 ] with μ and k B being the chemical potential and the Boltzmann constant, respectively. We have used the assumption of a constant relaxation time τ and the prefactor 2 comes from the spin degrees of freedom.
In the present case of ε k = t eff ( k b ) 4 , it is straightforward to obtain
σ = A F 1 / 4 ,
where
A = 6 e 2 τ t eff b π a 2 2 k B T t eff 3 / 4 ,
and F i is defined as [2]
F i = F i ( η ) = 0 x i d x e x η + 1 ,
with η = μ / k B T . Here, we have changed the integral variable from k to x = ε k / k B T . Note that in the k 2 -dispersion case, we have σ = A F 1 / 2 .
The other linear response coefficients, L i j ( i , j = 1 , 2 ), are defined by [16]
j = L 11 E + L 12 T T , j Q = L 21 E + L 22 T T ,
where j and j Q are electric and heat currents, respectively, and L 11 is the electric conductivity σ . When we express the conductivity as
σ = d ε σ ( ε , T ) f ( ε ) ,
we can see that the following relations hold in the present case:
L 12 = L 21 = d ε ( ε μ ) σ ( ε , T ) f ( ε ) , L 22 = d ε ( ε μ ) 2 σ ( ε , T ) f ( ε ) ,
which we call the Sommerfeld–Bethe (SB) relation [17]. The SB relation is derived from the Kubo–Luttinger theory [18,19]. The validity of SB relation was studied in detail [17] and it was found that the origin of the thermoelectric effect is classified into inside of SB relation and out of SB relation. Phonon drag effect, which is an origin of huge Seebeck effects, is out of SB relation [17,20]. Recently, based on this classification, the microscopic theories of thermoelectrics have been extensively studied [21].
Using this SB relation, it is straightforward to obtain
L 12 = L 21 = A k B T e 7 3 F 3 / 4 η F 1 / 4 , L 22 = A ( k B T ) 2 e 2 11 3 F 7 / 4 14 3 η F 3 / 4 + η 2 F 1 / 4 .
Then, the Seebeck coefficient S and the electronic thermal conductivity κ e are given by
S = L 12 T L 11 = k B e 7 3 F 3 / 4 F 1 / 4 η , κ e = 1 T L 22 L 12 L 21 L 11 = A k B 2 T e 2 11 3 F 7 / 4 49 9 ( F 3 / 4 ) 2 F 1 / 4 .
Finally, introducing the lattice thermal conductivity κ L , we find the thermoelectric figure of merit
Z T = S 2 σ T κ e + κ L = 7 3 F 3 / 4 F 1 / 4 η 2 F 1 / 4 1 B + 11 3 F 7 / 4 49 9 ( F 3 / 4 ) 2 F 1 / 4 ,
with
B = A k B 2 T e 2 κ L .

5. Numerical Results

The result in Equation (13) shows that the parameter dependence of Z T comes only from the values of η = μ / k B T and B. Microscopic parameters such as τ , t eff , and the lattice parameters a , b are included in B. Therefore, when we fix the temperature T and the value of B, we can find an optimal value of η that maximizes Z T . Experimentally, η can be adjusted by, for example, doping, as discussed by Hicks and Dresselhaus [2].
To evaluate the parameter B, Hicks and Dresselhaus assumed that the mean free path of phonons are limited by surface scattering. Then, the lattice thermal conductivity was estimated as
κ L = 1 3 C v v ,
with being the length of the cross section ( = a ). We use the same parameters as used by Hicks and Dresselhaus: C v = 1.2 × 10 6   J   K 1 m 3 , v = 3 × 10 3   m   s 1 , = 10 Å. In this case, the lattice thermal conductivity becomes κ L = 1.2   W   K 1   m 1 . For the prefactor of the conductivity, A, we use τ = 1 × 10 14 s, Δ = 1 eV and b = 3 Å.
Figure 3 shows η -dependence of ZT for the Hicks–Dresselhaus theory (black lines) and the double-chain model (red lines). Here we fixed the parameters as m * = 0.01 m 0 with m 0 being the electron mass for Hicks-Dresselhaus, and t = 0.5 eV ( t eff 0.13 eV) for the double-chain model, T = 300 K and the length of the side of the cross section, a is a = 5 Å and 10 Å. The optimal value of η ( η = η * ) is determined as the value that ZT is maximum.
Figure 4 shows the optimal value of η ( η = η * ) as a function of the length of the side of the cross section, a. The black and gray lines indicate the results of Hicks–Dresselhaus for m * / m 0 = 0.01 and 1, respectively. In this paper, we use m * / m 0 = 0.01 ~1 to compare the dependence of the mass in Hicks–Dresselhaus with the dependence of hopping (t) in the present model. The pink, red, and blue lines show the numerical results of double-chain model for t = 1 eV ( t eff 0.5 eV), 0.5 eV ( t eff 0.13 eV) and 0.1 eV ( t eff 0.005 eV), respectively. The dependence of the double-chain model on a is the same as that of the single-chain model, in the view that η * decreases as a decreases.
In our actual numerical calculations, we use another universal function T i ( η ) instead of F i ( η ) , because the linear response coefficients are expressed by the energy derivative of the Fermi distribution function f ( ε ) . T i ( η ) is related to f ( ε ) . Some details are shown in Appendix A.
Using η * , the obtained maximum of Z T is shown in Figure 5. The black and gray lines are the result of the single-chain model (Hicks–Dresselhaus [2]), and the pink, red, and blue lines indicate the results of the double-chain model. As the length of the side, a, decreases, ZT increases drastically in both models, while we find that the enhancement of ZT in the double-chain model is smaller than that of the single-chain model in the realistic parameter region. This difference between the single- and double-chain models is mainly due to the difference of B. We can rewrite the prefactor of the conductivity obtained by Hicks and Dresselhaus, σ = A F 1 / 2 , as follows:
A = 1 π a 2 2 k B T 2 1 / 2 ( m * ) 1 / 2 e μ e = 2 e 2 τ t eff b π a 2 2 k B T t eff 1 / 2 ,
where μ e = e τ / m * is the electron mobility and t eff in the single-band model is defined by
t eff = 2 2 m * b 2 ,
because the energy dispersion is written as ε k = 2 k 2 / 2 m * = t eff ( k b ) 2 . This prefactor A has the same parameter dependence as that in the present model in Equation (6). However, t eff in the single-chain model (Hicks–Dresselhaus) is very large: t eff = 0.6 eV for m * = m e and t eff = 60 eV for m * = 0.01 m e . This large A causes small 1 / B in the expression of Z T , or the relative amplitude of the lattice thermal conductivity compared to κ e is very small. As a result, the figure of merit Z T becomes large for the single-chain model. In fact, when we compare the result of Z T in Figure 5 for the single-chain model with m * = m e (corresponding to t eff = 0.6 eV) and that for the double-chain model with t = 0.5 eV (corresponding to t eff = 0.13 eV), we find that the values of Z T for both cases are comparable, but Z T for the latter is slightly larger than the former.
Finally, we show the temperature (T) dependence of ZT for the single- and double-chain models in Figure 6. Here, we fixed the parameters as shown in Figure 3. We find that ZT gradually increases as the temperature increases in both cases. This will be because the relative weight of the lattice thermal conductivity decreases as the temperature increases. Note that the difference between the magnitudes of Z T for the double-chain model with t = 0.5 and that in the single-chain model with m * = m 0 slightly enhanced as the temperature increases.

6. Conclusions

In conclusion, the figure of merit of the present model (double-chain model) is not so large compared with the results obtained by Hicks and Dresselhaus. Nevertheless, the obtained value of Z T is promising to be realized in actual materials. The reason why Z T is not so large is that the effective hopping energy t eff is not so large compared to that for the single-chain model. As shown in the preceding section, if the effective mass is m * = 0.01 m e , t eff in the single-chain model becomes as large as 60 eV. Thus, if we can choose a material which has a small effective mass as m * = 0.01 m e in the double-chain model, we can have a large value of Z T comparable to that in the single-chain model.

Author Contributions

All authors have equally contributed to the conceptualization, methodology, formal analysis, investigation, writing original draft preparation, and writing review and editing of the present paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by Grants-in-Aid for Scientific Research from the Japan Society for the Promotion of Science (No. 22KK0228, No. 23K03274, and No. 25K07199), and JST-Mirai Program Grant (No. JPMJMI19A1).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

As an alternative expression, we can use
T i = T i ( η ) = 0 x i e x η d x ( e x η + 1 ) 2 ,
instead of F i ( η ) in Equation (7). T i ( η ) is directly related to the energy derivative of the Fermi distribution function, f ( ε ) . Numerically, the convergence of integral in T i ( η ) is better than F i ( η ) because T i ( η ) has a delta function-like peak near x = η . Using T i ( η ) , we obtain
L 11 = 4 3 A T 3 / 4 , L 12 = L 21 = 4 3 A k B T e T 7 / 4 η T 3 / 4 , L 22 = 4 3 A ( k B T ) 2 e 2 T 11 / 4 2 η T 7 / 4 + η 2 T 3 / 4 .
Then, the Seebeck coefficient and the electronic thermal conductivity become
S = k B e T 7 / 4 T 3 / 4 η , κ e = 4 3 A k B 2 T e 2 T 11 / 4 ( T 7 / 4 ) 2 T 3 / 4 .
Finally, the figure of merit Z T is written as
Z T = T 7 / 4 T 3 / 4 η 2 T 3 / 4 3 4 B + T 11 / 4 ( T 7 / 4 ) 2 T 3 / 4 .
The actual numerical calculation is carried out using these formulas.

References

  1. Hicks, L.D.; Dresselhaus, M.S. Effect of quantum-well structures on the thermoelectric figure of merit. Phys. Rev. B 1993, 47, 12727–12731. [Google Scholar] [CrossRef] [PubMed]
  2. Hicks, L.D.; Dresselhaus, M.S. Thermoelectric figure of merit of a one-dimensional conductor. Phys. Rev. B 1993, 47, 16631–16634. [Google Scholar] [CrossRef] [PubMed]
  3. Nolas, G.S.; Sharp, J.; Goldsmid, H.J. Low-Dimensional Thermoelectric Materials. In Thermoelectrics; Springer Series in MATERIALS SCIENCE; Springer: Berlin/Heidelberg, Germany, 2001; Volume 45. [Google Scholar]
  4. Heremans, J.P.; Dresselhaus, M.S.; Bell, L.E.; Morelli, D.T. When thermoelectrics reached the nanoscale. Nat. Nanotechnol. 2013, 8, 471. [Google Scholar] [CrossRef] [PubMed]
  5. Yamamoto, T.; Fukuyama, H. Possible High Thermoelectric Power in Semiconducting Carbon Nanotubes A Case Study of Doped One-Dimensional Semiconductors. J. Phys. Soc. Jpn. 2018, 87, 024707. [Google Scholar] [CrossRef]
  6. Latronico, G.; Eivari, H.A.; Mele, P.; Assadi, M.H.N. Insights into One-Dimensional Thermoelectric Materials: A Concise Review of Nanowires and Nanotubes. Nanomaterials 2024, 14, 1272. [Google Scholar] [CrossRef] [PubMed]
  7. Kim, J.; Shim, W.; Lee, W. Bismuth nanowire thermoelectrics. J. Mater. Chem. C 2015, 3, 11999. [Google Scholar] [CrossRef]
  8. Lin, Y.-M.; Sun, X.; Dresselhaus, M.S. Theoretical investigation of thermoelectric transport properties of cylindrical Bi nanowires. Phys. Rev. B 2000, 62, 4610. [Google Scholar] [CrossRef]
  9. Kim, J.; Lee, S.; Brovmen, Y.M.; Kim, P.; Lee, W. Diameter-dependent thermoelectric figure of merit in single-crystalline Bi nanowires. Nanoscale 2015, 7, 5053. [Google Scholar] [CrossRef] [PubMed]
  10. Murata, M.; Nakamura, D.; Hasegawa, Y.; Komine, T.; Taguchi, T.; Nakamura, S.; Jovovic, V.; Heremans, J.P. Thermoelectric properties of Bismuth nanowires in a quartz template. Appl. Phys. Lett. 2009, 94, 192104. [Google Scholar] [CrossRef]
  11. Murata, M.; Yamamoto, A.; Hasegawa, Y.; Komine, T. Experimental and Theoretical Evaluations of the Galvanomagnetic Effect in an individual Bismuth Nanowire. Nanoletters 2017, 17, 110. [Google Scholar] [CrossRef] [PubMed]
  12. Halperin, B.I.; Rice, T.M. Possible Anomalies at a Semimetal-Semiconductor Transistion. Rev. Mod. Phys. 1968, 40, 755. [Google Scholar] [CrossRef]
  13. Wakisaka, Y.; Sudayama, T.; Takubo, K.; Mizokawa, T.; Arita, M.; Namatame, H.; Taniguchi, M.; Katayama, N.; Nohara, M.; Takagi, H. Excitonic Insulator State in Ta2NiSe5 Probed by Photoemission Spectroscopy. Phys. Rev. Lett. 2009, 103, 026402. [Google Scholar] [CrossRef] [PubMed]
  14. Kaneko, T.; Ohta, Y. A New Era of Excitonic Insulators. J. Phys. Soc. Jpn. 2025, 94, 012001. [Google Scholar] [CrossRef]
  15. Takarada, S.; Ogata, M.; Matsuura, H. Theory of thermal conductivity of excitonic insulators. Phys. Rev. B 2021, 104, 165122. [Google Scholar] [CrossRef]
  16. Mahan, G.D. Many-Particle Physics; Plenum: New York, NY, USA, 1990. [Google Scholar]
  17. Ogata, M.; Fukuyama, H. Range of Validity of Sommerfeld-Bethe Relation Associated with Seebeck Coefficient and Phonon Drag Contribution. J. Phys. Soc. Jpn. 2019, 88, 074703. [Google Scholar] [CrossRef]
  18. Kubo, R. Statistical-Mechanical Theory of Irreversible Processes. I. General Theory and Simple Applications to Magnetic and Conduction Problems. J. Phys. Soc. Jpn. 1957, 12, 570. [Google Scholar] [CrossRef]
  19. Luttinger, J.M. Theory of Thermal Transport Coefficients. Phys. Rev. 1964, 135, A1505. [Google Scholar] [CrossRef]
  20. Matsuura, H.; Maebashi, H.; Ogata, M.; Fukuyama, H. Effect of Phonon Drag on Seebeck Coefficient Based on Linear Response Theory: Application to FeSb2. J. Phys. Soc. Jpn. 2019, 88, 074703. [Google Scholar] [CrossRef]
  21. Matsubara, M.; Yamamoto, T.; Fukuyama, H. Two-band Model with High Thermoelectric Power Factor and Its Application to FeSe Thin Film. J. Phys. Soc. Jpn. 2023, 92, 104704. [Google Scholar] [CrossRef]
Figure 1. Schematic pictures of normal state (black line) and excitonic insulator (red line).
Figure 1. Schematic pictures of normal state (black line) and excitonic insulator (red line).
Inorganics 13 00310 g001
Figure 2. Schematic pictures of (a) the single-chain model corresponding to the Hicks–Dresselhaus theory and (b) the double-chain model, consisting of two carriers of electrons and holes depending on the chain. t and Δ represent the hopping integral along the chain and between the chains, respectively. We consider that the conductor is square in cross section with a side of length a. The one-dimensional chains are packed in the cross section of a 2 .
Figure 2. Schematic pictures of (a) the single-chain model corresponding to the Hicks–Dresselhaus theory and (b) the double-chain model, consisting of two carriers of electrons and holes depending on the chain. t and Δ represent the hopping integral along the chain and between the chains, respectively. We consider that the conductor is square in cross section with a side of length a. The one-dimensional chains are packed in the cross section of a 2 .
Inorganics 13 00310 g002
Figure 3. η -dependence of ZT. The black and red lines indicate the results obtained by the Hicks–Dresselhaus theory and double-chain model, respectively. Here we fixed the parameters as m * = 0.01 m 0 with m 0 being the electron mass for Hicks–Dresselhaus, and t = 0.5 eV for the double-chain model, and T = 300   K . η * indicates the optimized value of η .
Figure 3. η -dependence of ZT. The black and red lines indicate the results obtained by the Hicks–Dresselhaus theory and double-chain model, respectively. Here we fixed the parameters as m * = 0.01 m 0 with m 0 being the electron mass for Hicks–Dresselhaus, and t = 0.5 eV for the double-chain model, and T = 300   K . η * indicates the optimized value of η .
Inorganics 13 00310 g003
Figure 4. Optimal value of η ( η = η * ) at which the figure of merit Z T has a maximum as a function of the length of the side of the cross section a.
Figure 4. Optimal value of η ( η = η * ) at which the figure of merit Z T has a maximum as a function of the length of the side of the cross section a.
Inorganics 13 00310 g004
Figure 5. Optimal value of the figure of merit Z T as a function of the length of the side of the cross section a. The blue, red, and pink lines show the numerical result of double-chain model. The black and gray line indicate the results which were recalculated based on Hicks–Dresselhaus [2]. The gray line overlaps with the red line.
Figure 5. Optimal value of the figure of merit Z T as a function of the length of the side of the cross section a. The blue, red, and pink lines show the numerical result of double-chain model. The black and gray line indicate the results which were recalculated based on Hicks–Dresselhaus [2]. The gray line overlaps with the red line.
Inorganics 13 00310 g005
Figure 6. Temperature dependence of ZT for the single- and double-chain models.
Figure 6. Temperature dependence of ZT for the single- and double-chain models.
Inorganics 13 00310 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Matsuura, H.; Ogata, M. Thermoelectric Figure of Merit in a One-Dimensional Model with k4-Dispersion: An Extension of the Theory by Hicks and Dresselhaus. Inorganics 2025, 13, 310. https://doi.org/10.3390/inorganics13090310

AMA Style

Matsuura H, Ogata M. Thermoelectric Figure of Merit in a One-Dimensional Model with k4-Dispersion: An Extension of the Theory by Hicks and Dresselhaus. Inorganics. 2025; 13(9):310. https://doi.org/10.3390/inorganics13090310

Chicago/Turabian Style

Matsuura, Hiroyasu, and Masao Ogata. 2025. "Thermoelectric Figure of Merit in a One-Dimensional Model with k4-Dispersion: An Extension of the Theory by Hicks and Dresselhaus" Inorganics 13, no. 9: 310. https://doi.org/10.3390/inorganics13090310

APA Style

Matsuura, H., & Ogata, M. (2025). Thermoelectric Figure of Merit in a One-Dimensional Model with k4-Dispersion: An Extension of the Theory by Hicks and Dresselhaus. Inorganics, 13(9), 310. https://doi.org/10.3390/inorganics13090310

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop