To meet the cooling demands of high heat flow density hotspots in scenarios such as electronic chips, a novel three-dimensional stacked T-shaped thermoelectric cooler (STTEC) is designed in this study. Under steady-state conditions, a finite element method with coupled thermal–electrical–mechanical physical fields is utilized, and the temperature dependence of thermoelectric (TE) materials is considered. First, the cooling flux, coefficient of performance (COP), and minimum cooling temperature of STTEC under different input-current and thermal boundary conditions are investigated and compared to the traditional π-shaped thermoelectric cooler (π-TEC). Second, the effects of geometrical parameter variations under optimal currents on the cooling performance and reliability of STTEC are studied. Finally, the structural parameters are optimized. The results show that the STTEC altered the path of TE conversion and transfer, which significantly improved the optimal current. The STTEC has a remarkable advantage in cooling performance under low temperature differences or high cooling loads. Compared to the π-TEC, STTEC enhances cooling flux by 101.6%, rises COP by 358.5%, and lowers the cold-end temperature by 46.6 K. At optimal current conditions, by optimizing the thickness of the T-shaped copper slice and the height difference between the TE leg and the T-shaped copper slice, the thermal stress decreased by 18.4%. The STTEC’s novel design could inspire the manufacturing and commercialization of high-performance thermoelectric coolers.

At the present time, the miniaturization and integration of electronic chips have increased significantly and excessive heat flow density has emerged as a key constraint on the stability and performance of chips.1,2 To effectively address the thermal challenges posed by these advancements, the thermoelectric cooler (TEC) acts as a solid-state heat pump that is environmentally friendly, can remove heat using the Peltier effect, and has shown great application potential3,4 in laser communication devices and computer processors. TEC can achieve excellent localized cooling density within a limited chip package volume, thus significantly improving the chip’s performance and lifecycle.5–7 

The most effective approach to strengthening the performance of thermoelectric (TE) devices is through the exploration of high-efficiency TE materials.8–11 Nonetheless, finding more promising TE materials that can meet the demands of industrial production is still challenging, and optimization of the configuration and structure of TE devices represents another viable approach to enhance performance.12–15 

A proper device design can fully exert the material performance potential, which is well-established in the field of thermoelectric generators. For example, the beneficial geometric design of the TE legs includes the different width ratios of the P-type and N-type TE legs,16 I-shaped and Y-shaped TE legs,17 asymmetric pyramidal legs,18 cylindrical TE legs,19 and optimization of TE legs based on the genetic algorithm.20,21 Similarly, a structural design is also essential for TEC. Optimizing the geometric parameters of rectangular TE legs is a straightforward and effective approach.22,23 Due to the geometric temperature distribution that can alter TE material properties, the design of a variable cross-sectional trapezoidal-shaped leg is investigated;24 similarly, the structural shapes of TE legs, such as trapezoidal, tapered, and X-shaped legs, have been comparatively analyzed.25 Additionally, the segmented design26 and the use of algorithms can optimize the single-stage27 and two-stage28 TEC. For two-stage TEC, optimizing the geometrical parameter configurations of TE legs for the upper and lower stages29 or omitting the intermediate ceramic substrate of the two-stage TEC with the trapezoidal TE legs also are effective structural optimization methods.30 

From above, the optimal design of TE devices primarily focuses on the shape and size of TE legs based on π-shaped connections. Although previous studies have explored different connection forms, such as linear connection form31 and U-type single-arm TE element connection form design32 in the field of thermoelectric generators, the TEC differs from the thermoelectric generator; altering the TE element connection form, the path of operational current and heat conduction of the TEC will be significantly changed, and refrigeration performance and mechanical properties are also different.

As a result, the connection of TE elements and optimized configuration is also a critical structural parameter; there is still limited research in this area. In this paper, a new connection form of TE elements called stacked T-shaped thermoelectric cooler (STTEC) is explored, and the influence of different TE boundary conditions and altered structural parameters on the performance of STTEC was successively analyzed. The research results offer insights into the optimization of TEC structures.

The three-dimensional model of STTEC is shown in Fig. 1. To ensure the rationality of STTEC’s size design and facilitate optimization, the geometry dimensions of STTEC in this work were adopted from dimensions of current commercially available micro-TECs. The overall dimension is 2 × 2 × H (depending on the study) mm3 of STTEC. It includes 18 pairs of P-type and N-type TE legs, which have the same dimensions. Considering the dimensions of current commercially produced TE legs, the reasonable TE leg width of STTEC is 0.24 mm and the TE leg height is 0.3 mm. Furthermore, the T-shaped copper slices have a uniform thickness of WCu, and the height difference between the TE leg and the T-shaped copper slice is ΔH, measured in μm. The ceramic plate has a fixed thickness of 0.20 mm.

FIG. 1.

Three-dimensional model of STTEC: (a) 18 pairs and (b) single pair.

FIG. 1.

Three-dimensional model of STTEC: (a) 18 pairs and (b) single pair.

Close modal

The principle of operation is shown in Fig. 1(a). STTEC starts to operate when current is supplied; the process involves the absorption and release of heat, denoted as Qc and Qh, respectively, measured in W. The temperature of the cold and hot ends is represented by Tc and Th, respectively, measured in K.

The materials used were carefully selected, and the TE legs were made of Bi2Te3,33 a promising commercial TE material commonly utilized at low temperatures. Copper slices were chosen as excellent conductors for the electrodes. As for the substrate material, aluminum nitride was chosen since it has excellent thermal conductivity and electrical insulation. Additionally, the TE properties and thermodynamics properties of materials can be found in Fig. 2 and Table I. Poisson’s ratio of the two types of Bi2Te3 is 0.23.

FIG. 2.

(a)–(c) Thermoelectric properties of P–Bi2Te3 and N–Bi2Te3. (d) Coefficient of thermal expansion and Young’s modulus of two types of Bi2Te3.

FIG. 2.

(a)–(c) Thermoelectric properties of P–Bi2Te3 and N–Bi2Te3. (d) Coefficient of thermal expansion and Young’s modulus of two types of Bi2Te3.

Close modal
TABLE I.

Thermoelectric and thermodynamic properties of materials.23 Reproduced with permission from Gong et al., “Thermo-mechanical analysis on a compact thermoelectric cooler,” Energy 127, 1211 (2019). Copyright 2019 Elsevier.

MaterialSeebeck factor (V/K)Resistivity (S/m)Thermal conductivity [W/(m K)]Coefficient of thermal expansion (K−1)Young’s modulus (GPa)Poisson’s ratio
Copper slices ⋯ 1.68 × 10−8 401 1.77 × 10−5 119 0.326 
Ceramic substrates ⋯ 1 × 1012 170 0.68 × 10−5 310 0.33 
MaterialSeebeck factor (V/K)Resistivity (S/m)Thermal conductivity [W/(m K)]Coefficient of thermal expansion (K−1)Young’s modulus (GPa)Poisson’s ratio
Copper slices ⋯ 1.68 × 10−8 401 1.77 × 10−5 119 0.326 
Ceramic substrates ⋯ 1 × 1012 170 0.68 × 10−5 310 0.33 

The contact effects between heterogeneous interfaces of the STTEC cannot be ignored. In this study, the contact resistance value was determined using the research of Chowdhury et al.34 The specific contact effects are shown in Table II. To simplify the model, reduce computational costs, and maintain a level of accuracy, the solder layer is not considered.

TABLE II.

Modeling of contact effects.

Contact layerContact electrical resistance (Ω m2)Contact thermal resistance (m2 K/W)
TE leg-copper 1 × 10−11 1 × 10−6 
Copper-ceramic ⋯ 8 × 10−6 
substrates 
Contact layerContact electrical resistance (Ω m2)Contact thermal resistance (m2 K/W)
TE leg-copper 1 × 10−11 1 × 10−6 
Copper-ceramic ⋯ 8 × 10−6 
substrates 
At a steady state, the heat flow and the charge continuity equation are expressed, respectively, as
(1)
(2)
where q is the heat production rate per unit volume and q, C, t, D, and J denote the heat flux vector, density, specific heat capacity, time, potential shift vector, and current density vector, respectively.
The above equations are coupled through the thermoelectric eigen-structure equation as
(3)
(4)
The principal constitutive equation is expressed as
(5)
where [π], [σ], [α], [κ], [ɛ], and φ denote the Peltier coefficient matrix, conductivity coefficient matrix, Seebeck coefficient matrix, thermal conductivity coefficient matrix, dielectric constant matrix, and potential, respectively.
This can be obtained by substituting Eqs. (3)(5) into Eqs. (1) and (2),
(6)
(7)
The thermodynamic control equation is
(8)
where k is the thermal conductivity, T=fx,y,z.
The displacement–strain relationship can be obtained through the dimensionless equation,
(9)
(10)
where ε̄ij denotes the dimensionless strain component and ū, v̄, and w̄ denote the dimensionless displacement components.
The dimensionless stress–strain relationship is expressed by the Jacobi matrix as follows:
(11)
where[A] = 1vvv000v1vv000vv1v00000012v00000012v00000012v.
The three principal stresses are σ1, σ2, and σ3. The von Mises equivalent stresses are described as
(12)
Cold side cooling capacity (Qc) and hot side heating capacity (Qh) can be calculated from the heat balance equation at the cold end,
(13)
(14)
where αNP represents the TEC total Seebeck coefficients, R is the total resistance, I is the current input in the loop, and the unit is A.
The per unit area of Qc, also known as the cooling flux (qc), is expressed in W/cm2,
(15)
where S is the surface area of the TEC’s cold side in cm2.
The voltage across the TEC is determined by the sum of the voltage generated by the internal resistance and the voltage required to resist the Seebeck voltage,
(16)
From Eqs. (13) and (14), the power consumption (P) of the TEC in W and the cooling coefficient COP, dimensionless, are obtained,
(17)
(18)
where U is the electromagnetism in V, based on the conservation of energy under adiabatic boundary conditions, P = QhQc.

By introducing reasonable assumptions, avoiding over-analysis and omitting the factors that have little impact on the results, and simplifying calculations, we have the following:

  1. The electrical boundary conditions of TEC are the positive input constant current and the negative input constant current set to zero potential (as shown Fig. 1).

  2. The impact of environmental factors is ignored; thus, all exterior surfaces of TEC are assumed to be adiabatic except for the cold and hot surfaces.

  3. When calculating the cooling flux and COP, a constant temperature boundary condition is applied to two sides of TEC. The temperature of the hot side is 330.15 K, and the cold side has a common testing temperature that varies between 240.15 and 330.15 K.

  4. When analyzing the minimum cold end temperature of TEC at optimal current, the temperature of the hot side is varied between 300.15 and 330.15 K, and the cold side has cooling loads that vary between 0 and 40 W/cm2.

  5. When conducting thermal stress analysis, the fixed constraint boundary condition is applied on the hot surface of TEC, while the other boundaries are free. Strain in the height direction and all shear deformations of the hot side surface are zero.

  6. Anisotropic material properties are not taken into account.

The finite element method was utilized in this study, and the numerical results were computed by COMSOL Multi-physics 5.6 simulation software, which couples the three modules of current, solid heat transfer, and solid mechanics by three physical fields of electromagnetic heat and thermoelectric effects and thermal expansion; constructs the geometric model; sets the boundary conditions and delineates the mesh sequentially; and solves the multi-physics field by a steady-state study.

To verify the grid independence, three grid systems of STTEC were examined. As depicted in Fig. 3, the values of the cooling capacity calculated from the three grid cells were essentially the same, with the maximum error of numerical results less than 1%, thus confirming the grid’s independence. To save computational resources, a grid number of 35 048 was chosen.

FIG. 3.

Verification of grid independence (Th = Tc = 300.15 K).

FIG. 3.

Verification of grid independence (Th = Tc = 300.15 K).

Close modal

According to Fig. 4, by comparing and analyzing the power consumption obtained from the two methods, the power consumption of the STTEC is P1 = U × I. Moreover, according to the first law of thermodynamics, the electric power consumption can be solved as P2 = QhQc. The maximum deviation is (P2P1)/P1 = 0.06% < 1%; thus, self-consistency is confirmed.

FIG. 4.

Validation of model self-consistency.

FIG. 4.

Validation of model self-consistency.

Close modal

The simulation model of the commercial π-TEC (MS02702010L2) is depicted in Fig. 5. To confirm the model’s accuracy, a comparison should be made with the actual performance of the π-TEC. The dimensions parameters are shown in Table III, and the performance parameters obtained from the testing platform are shown in Table IV. The testing process for key performance parameters is as follows: first, the TEC is placed in a vacuum hood environment for extracting air, in which absolute pressure is less than 30 pa, and a constant temperature of 300.15 K is maintained at the TEC hot end by using the thermostatic water cooling system. Second, a controllable DC power supply is used to gradually increase the input current from 0.2 A at intervals of 0.1 A; by using the data detection and recording system, the maximum temperature difference (ΔTmax) and corresponding optimum current (Iop) are obtained. Finally, when measuring the maximum cooling capacity (QCmax), the TEC cold end is additionally connected with a heating resistor block with adjustable power to maintain a constant temperature of 300.15 K and QCmax and Iop are obtained by gradually adjusting the input current and analyzing the data.

FIG. 5.

Three-dimensional model of π-TEC.

FIG. 5.

Three-dimensional model of π-TEC.

Close modal
TABLE III.

Size parameters of π-TEC.

π-TECPair of legsLeg width (mm2)Leg height (mm)Leg distances (mm)Copper slices thickness (μm)Ceramic plate thickness (mm)
Sizes 18 0.24 × 0.24 0.30 0.1 0.04 0.20 
π-TECPair of legsLeg width (mm2)Leg height (mm)Leg distances (mm)Copper slices thickness (μm)Ceramic plate thickness (mm)
Sizes 18 0.24 × 0.24 0.30 0.1 0.04 0.20 
TABLE IV.

Tested key performance parameters of π-TEC (Th = 300.15 K).

π-TECIop (A)ΔTmax (K)QCmax (W)
Test data 68 1.3 
π-TECIop (A)ΔTmax (K)QCmax (W)
Test data 68 1.3 

The numerical results curve is depicted in Fig. 6, Th is set to 300.15 K alone, and the maximum temperature difference (∆Tmax) can be obtained. By setting both Th and Tc to 300.15 K, the maximum cooling capacity (QCmax) can be found, the optimum current (Iop) of the model is 1 A, QCmax is 1.28 W, and ∆Tmax is 68.3 K. The maximum error is 1.5% < 5% between the numerical results and test data. Considering acceptable experimental errors, the numerical model is valid.

FIG. 6.

Comparison of numerical results with tested data.

FIG. 6.

Comparison of numerical results with tested data.

Close modal

To illustrate the design principle of STTEC, Figs. 7 and 8 show the temperature clouds (ΔT = 30 K, I = 0.5 A). In addition, according to the application scenario high-temperature conditions, Th is set to a reasonable value of 330.15 K. From Fig. 7, we see that the π-TEC temperature distribution mainly varies along the direction of the leg height, while the temperature distribution of STTEC changes significantly along the direction of leg height and leg width.

FIG. 7.

Temperature clouds of two types of TECs. (a) π-TEC and (b) STTEC.

FIG. 7.

Temperature clouds of two types of TECs. (a) π-TEC and (b) STTEC.

Close modal
FIG. 8.

Temperature cloud of the horizontal center section. (a) π-TEC and (b) STTEC.

FIG. 8.

Temperature cloud of the horizontal center section. (a) π-TEC and (b) STTEC.

Close modal

From Fig. 8, the two TE legs of π-TEC are distinguished in temperature by 0.2 K at the same horizontal plane since the TE properties of the P-type and N-type legs are different, whereas the temperature of the horizontal section of the center of the STTEC varies alternately with a large temperature gradient and the temperature difference is 32 K. The principle that causes this phenomenon is because the STTEC has peculiar TE elements connection forms, it consists of T-shaped copper slices and two types of thermoelectric legs stacked alternatively in contact with each other, which significantly changes the current and heat transportation and facilitate the thermoelectric conversion, thus altered the temperature distribution of STTEC.

The cooling flux can be used as one of the indicators of TEC, which is usually used to evaluate the cooling applicability for hotspot cooling; it is dependent on input current (I) and temperature difference (ΔT). From Fig. 9, under different operating conditions, the cooling flux increases with I but decreases once I exceeds the optimal current (Iop). In other words, the cooling flux is influenced by three terms representing different factors: the Peltier effect’s heat extraction of the TEC, the generation of Joule heat from the device’s internal resistance, and heat conduction due to temperature difference. When I is less than Iop, the increment of Peltier refrigeration is greater than Joule heat production. When I is at its optimum value (i.e., the peak of the cooling flux), the increments of Peltier refrigeration and Joule heat production are equal. As I is greater than Iop, the rate of increase of Joule heat production is more rapid than the Peltier refrigeration, with reduced cooling flux as a result. Furthermore, as ΔT increases, the Fourier heat conduction also enhances; therefore, the cooling flux shows a decreasing trend.

FIG. 9.

Effect of ΔT and I on cooling flux of two types of TEC. (a) π-TEC and (b) STTEC.

FIG. 9.

Effect of ΔT and I on cooling flux of two types of TEC. (a) π-TEC and (b) STTEC.

Close modal

At the optimal current, the cooling fluxes of the two TECs are depicted in Fig. 10. From a temperature difference (ΔT) of 50–0 K, the increase in cooling flux of STTEC ranges from 17.2 to 101.6%. TEC is a device that pumps energy from an input current; due to the configuration of STTEC, which consists of alternating stacks with copper slices and two different TE legs, it allows for high current inputs and the Peltier effect is positively correlated with the input current, thus increasing cooling ability with better TE conversion efficiency of STTEC, resulting in better refrigeration performance. However, the stacked structure enhances the Fourier heat conduction effect and accelerates the decay of cooling flux, with the reduced cooling flux at ΔT greater than 50 K. Consequently, with actual low temperature difference cooling needs, STTEC can pump and move large amounts of heat.

FIG. 10.

At the optimal current, the influence of ΔT on the cooling fluxes of the two TECs.

FIG. 10.

At the optimal current, the influence of ΔT on the cooling fluxes of the two TECs.

Close modal

The COP is also a performance index, which is shown in Fig. 11; the behavior of the COP is different from that of cooling flux. It decreases with the increasing temperature difference (ΔT) and input current (I) for two types of TECs; on the one hand, with the increase of ΔT, the enhancement of the cooling effect also increases the Fourier heat conduction and leads to the increase of heat loss so that the COP is reduced; on the other hand, with the increase of I, the increase of the cooling production of the Peltier is made and the COP tends to decrease due to the additional increase of the Joule heat. Moreover, due to the different ways of connecting the two kinds of TECs’ thermoelectric elements, their COPs under the same I and under the same ΔT are also different. Furthermore, the same cooling flux of two TECs is depicted in Fig. 12, ΔT decreased from 40 to 0 K, and the COP of STTEC increased from 5.1% to 358.5%. The STTEC has better cooling efficiency and energy utilization, but as the temperature difference increases to 50 or 60 K, the COP of STTEC has deteriorated. Due to the special characteristics of the STTEC structure, the current transfer is improved, and the conversion efficiency of TE elements is greatly enhanced at low temperature differences, consuming less power to achieve excellent cooling effects. For hotspot cooling of the chip, which does not require to be operated under extremely high ΔT, there is a significant advantage for STTEC.

FIG. 11.

Effect of ΔT and I on COP of two types of TEC. (a) π-TEC and (b) STTEC.

FIG. 11.

Effect of ΔT and I on COP of two types of TEC. (a) π-TEC and (b) STTEC.

Close modal
FIG. 12.

At the same cooling flux, the variation of COP of two types of TEC with ΔT.

FIG. 12.

At the same cooling flux, the variation of COP of two types of TEC with ΔT.

Close modal

In practice, the TECs are subjected to cooling loads depicted in Fig. 13, inputting optimal current (Iop); the cold surface temperature (Tc) keeps increasing linearly with the increase of cooling load; and STTEC has a lower temperature growth rate. Thoroughly, two types of TECs maintain low Tc at cooling loads below 11.5 W/cm2, and the π-TEC has relatively lower Tc. Conversely, when the cooling load exceeds 12.9 W/cm2, the STTEC outperforms the π-TEC in sustaining a relatively low cold-end temperature. When the cooling load is 40 W/cm2, Tc of the STTEC is reduced by 46.6 K. This phenomenon can be explained as the T-shape structure optimized the path of current flow between the copper slice and the thermoelectric leg, which significantly reduces the internal resistance, and electric resistance is positively correlated with Joule heat. According to the working principle of TEC, its cooling performance is mainly determined by the combination of the unfavorable effect of Joule heat production and the favorable effect of Peltier cooling. Therefore, weakening the adverse effects of Joule heat production allows the Peltier cooling to dominate; the current is positively correlated with the Peltier cooling, and the increased current can be input to increase the cooling flux of the STTEC to effectively cool the loads.

FIG. 13.

Effect of Th and cooling load on Tc of two types of TEC.

FIG. 13.

Effect of Th and cooling load on Tc of two types of TEC.

Close modal

As the hot surface temperature (Th) is reduced from 330.15 to 300.15 K, the cooling load corresponding to equal Tc of π-TEC and STTEC is reduced from 12.9 W/cm2 to 11.5 W/cm2. Because the working principle of STTEC is to pump heat, improving the heat dissipation effect of the hot surface is more conducive to unleashing its refrigeration potential and achieving good cooling effects. To sum up, the STTEC proves to be effective in cooling hotspots with high heat flux density.

T-shaped copper slices are a key element for conducting heat and current, and the geometrical variation has an important effect on the STTEC’s cooling performance. In addition, changes in the geometric dimensions could alter the thermodynamic properties of the STTEC, which may reduce the reliability of the device.23 To accurately identify the location where the maximum thermal stress occurs, the maximum von Mises thermal stress on the surface of the thermoelectric leg at the contact interface between the copper slices and the thermoelectric leg characterizes the thermal stress level as an output.

As shown in Fig. 14(a), optimal current (Iop) decreases with the increasing height difference between the TE leg and the T-shaped copper slice (ΔH) and Iop increases slightly with the increase of copper slice thickness (WCu), reaching its maximum value at WCu = 50 µm, and then slowly decreases. Figure 14(b) depicts that at the optimal current, the thermal stress decreases with the increase of ΔH. When ΔH < 50 µm, the STTEC has high-level thermal stresses, and the thermal stress is beyond yield stress (112 Mpa) for thermoelectric legs of Bi2Te3 materials;35 it has low reliability. That is, when ΔH < 50 µm, the STTEC has an enhanced optimal current for thermoelectric conversion and a boost in the local temperature gradient and thermal stresses; the STTEC has easily localized fractures and cracks, which cause failure. However, as ΔH > 80 µm, the thermal stress is at a low level and effectively decreases with the increase of WCu because increasing WCu can ease thermal stresses at heterogeneous interfaces due to mismatched coefficients of thermal expansion.

FIG. 14.

Effect of WCu and ΔH on STTEC performance (Th = 330.15 K; cooling load is 20 W/cm2). (a) Optimal current. (b) Thermal stress. (c) Cold end temperature.

FIG. 14.

Effect of WCu and ΔH on STTEC performance (Th = 330.15 K; cooling load is 20 W/cm2). (a) Optimal current. (b) Thermal stress. (c) Cold end temperature.

Close modal

As shown in Fig. 14(c), increasing the height difference (ΔH) or the copper slice thickness (WCu) can decrease and then increase the cold end temperature (Tc) at the optimum current (Iop). This phenomenon can be explained as reducing ΔH increases Iop and not only promotes the Peltier effect cooling but also promotes the Fourier conduction of hot and cold sides caused by the temperature difference, reducing cooling performance as a consequence. Moreover, large ΔH leads to a significant reduction of Iop, so it is inhibited by the Peltier effect cooling. Therefore, the cold-end temperature is increased. In addition, WCu of appropriate size can improve thermoelectric conversion and heat transfer on the respective cold and hot parts, both too thin and too thick, which can reduce cooling performance. Consequently, there exists an optimal minimum cold end temperature of 283.4 K, and the size parameters of the copper slice are WCu = 50 µm and ΔH = 80 µm, and the corresponding thermal stress is 93.74 Mpa. Considering only cooling performance, it is an optimal T-shaped copper sheet size parameter. Furthermore, when the copper sizes are WCu = 60 µm and ΔH = 100 µm, Tc is 284.9 K, and the thermal stress is 76.53 Mpa and Tc is decreased by 1.5 K, but the thermal stress is reduced by 18.4%, the cooling performance was maintained at a high level, and the reliability is greatly improved of STTEC. The parametric design methods are of significant theoretical and practical meaning.

In this paper, we present a new Bi2Te3-based stacked T-shaped thermoelectric cooler (STTEC), composed of T-shaped copper slices and two types of thermoelectric legs stacked alternatively in contact with each other. The influence of performance is investigated using the method of coupling multiple physical fields. The key findings are as follows:

  1. The STTEC significantly increases optimal current by changes in the heat and current transportation, and its cooling performance advantage over π-TEC increases with the decreasing temperature difference. STTEC’s cooling flux increases by 101.6% at the temperature difference of 0 K. At the same cooling flux, the STTEC’s COP is 3.58 times higher and STTEC is suitable for application scenarios where the temperature differences are not extreme.

  2. The STTEC is less sensitive to cooling loads and advantageous for cooling high loads compared to π-TEC and maintains a lower cold-end temperature by 46.6 K; improving the heat dissipation condition can lead to greater improved cooling performance of STTEC. In practice, STTEC cools high heat flow density hot spots with great advantages.

  3. The medium size of the T-shaped copper slice can improve the cooling performance of STTEC, and increasing the thickness of copper slices and height differences is beneficial to improve reliability. A combination of cooling performance, thermal stress levels, thicker T-shaped copper slices of 60 µm, and the larger-sized height difference of 100 µm is the superior choice. In summary, while developing the STTEC to meet actual demands, the design factors must be carefully evaluated and chosen to get a synthesized performance.

This research was funded by the Zhengzhou Collaborative Innovation major Project: development and industrialization of new high-performance semiconductor refrigeration materials and components (Grant No. 18XTZX12005).

The authors have no conflicts to disclose.

Shengchao Yin: Conceptualization (equal); Data curation (equal); Methodology (equal); Writing – original draft (equal). Huadong Zhao: Funding acquisition (equal); Supervision (equal). Jingshuang Zhang: Investigation (equal); Methodology (equal). Cheng Li: Funding acquisition (equal); Visualization (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
W. M.
Mitchell
, “
The chips are down for Moore’s law
,”
Nature
530
(
7589
),
144
(
2016
).
2.
Y.
Cai
,
Y.
Wang
,
D.
Liu
, and
F. Y.
Zhao
, “
Thermoelectric cooling technology applied in the field of electronic devices: Updated review on the parametric investigations and model developments
,”
Appl. Therm. Eng.
148
(
148
),
238
(
2019
).
3.
J.
Mao
,
G.
Chen
, and
Z.
Ren
, “
Thermoelectric cooling materials
,”
Nat. Mater.
20
,
454
(
2020
).
4.
S.
Wiriyasart
,
C.
Hommalee
, and
P.
Naphon
, “
Thermal cooling enhancement of dual processors computer with thermoelectric air cooler module
,”
Case Stud. Therm. Eng.
14
(
14
),
100445
(
2019
).
5.
L.
Zhu
,
H.
Tan
, and
J.
Yu
, “
Analysis on optimal heat exchanger size of thermoelectric cooler for electronic cooling applications
,”
Energy Convers. Manage.
76
(
76
),
685
(
2013
).
6.
L. A.
Nimmagadda
and
S.
Sinha
, “
Thermoelectric property requirements for on-chip cooling of device transients
,”
IEEE Trans. Electron Devices
67
(
9
),
3716
(
2020
).
7.
C.
Selvam
,
S.
Manikandan
,
S. C.
Kaushik
et al, “
Transient performance of a Peltier super cooler under varied electric pulse conditions with phase change material
,”
Energy Convers. Manage
198
,
111822
(
2019
).
8.
B.
Poudel
,
Q.
Hao
,
Y.
Ma
,
Y.
Lan
,
A.
Minnich
,
B.
Yu
,
X.
Yan
,
D.
Wang
,
A.
Muto
,
D.
Vashaee
,
X.
Chen
,
J.
Liu
,
M. S.
Dresselhaus
,
G.
Chen
, and
Z.
Ren
, “
High-thermoelectric performance of nanostructured bismuth antimony telluride bulk alloys
,”
Science
320
,
634
(
2008
).
9.
M.
Zebarjadi
,
K.
Esfarjani
,
M. S.
Dresselhaus
,
Z. F.
Ren
, and
G.
Chen
, “
Perspectives on thermoelectrics: From fundamentals to device applications
,”
Energy Environ. Sci.
5
,
5147
(
2012
).
10.
J.
He
and
T. M.
Tritt
, “
Advances in thermoelectric materials research: Looking back and moving forward
,”
Science
357
(
6358
),
eaak9997
(
2017
).
11.
S. I.
Kim
,
K. H.
Lee
,
H. A.
Mun
,
H. S.
Kim
,
S. W.
Hwang
,
J. W.
Roh
,
D. J.
Yang
,
W. H.
Shin
,
X. S.
Li
,
Y. H.
Lee
,
G. J.
Snyder
, and
S. W.
Kim
, “
Dense dislocation arrays embedded in grain boundaries for high-performance bulk thermoelectrics
,”
Science
348
(
6230
),
109
(
2015
).
12.
Q.
Zhang
,
K.
Deng
,
L.
Wilkens
,
H.
Reith
, and
K.
Nielsch
, “
Micro-thermoelectric devices
,”
Nat. Electron.
5
(
6
),
333
(
2022
).
13.
S.
Shittu
,
G.
Li
,
X.
Zhao
, and
X.
Ma
, “
Review of thermoelectric geometry and structure optimization for performance enhancement
,”
Applied Energy
268
(
C
),
115075
(
2020
).
14.
J.
Yan
,
X.
Liao
,
D.
Yan
, and
Y.
Chen
, “
Review of micro thermoelectric generator
,”
J. Microelectromech S.
27
(
1
),
1
(
2018
).
15.
L. E.
Bell
, “
Cooling, heating, generating power, and recovering waste heat with thermoelectric systems
,”
Science
5895
,
1457
(
2008
).
16.
T.
Amit
,
S.
Lal
, and
M.
Razeeb Kafil
, “
Structural design optimization of micro-thermoelectric generator for wearable biomedical devices
,”
Energies
14
(
8
),
2339
(
2021
).
17.
K. A.
Lkhadher
,
E.
Ahmed
,
Y.
Said
et al, “
Performance comparison of TEGs for diverse variable leg geometry with the same leg volume
,”
Energy
224
,
119967
(
2021
).
18.
A.
Fabián-Mijangos
,
G.
Min
, and
J.
Alvarez-Quintana
, “
Enhanced performance thermoelectric module having asymmetrical legs
,”
Energy Convers. Manage.
148
(
148
),
1372
(
2017
).
19.
S.
Ferreira-Teixeira
and
A. M.
Pereira
, “
Geometrical optimization of a thermoelectric device: Numerical simulations
,”
Energy Convers. Manage.
169
(
169
),
217
(
2018
).
20.
Y.
Ge
,
Y.
Lin
,
Q.
He
,
W.
Wang
,
J.
Chen
, and
S. M.
Huang
, “
Geometric optimization of segmented thermoelectric generators for waste heat recovery systems using genetic algorithm
,”
Energy
233
,
121220
(
2021
).
21.
H.
Sun
,
Y.
Ge
,
W.
Liu
et al, “
Geometric optimization of two-stage thermoelectric generator using genetic algorithms and thermodynamic analysis
,”
Energy
171
,
37
(
2019
).
22.
F.
Shifa
,
R.
Alireza
, and
Y.
Gao
, “
Thermal-electric and stress analysis of thermoelectric coolers under continuous pulse input current
,”
Appl. Therm. Eng.
214
,
118910
(
2022
).
23.
T.
Gong
,
Y.
Wu
,
L.
Gao
et al, “
Thermo-mechanical analysis on a compact thermoelectric cooler
,”
Energy
172
,
1211
(
2019
).
24.
T.
Lu
,
Y.
Li
,
J.
Zhang
,
P.
Ning
, and
P.
Niu
, “
Cooling and mechanical performance analysis of a trapezoidal thermoelectric cooler with variable cross-section
,”
Energies
13
(
22
),
6070
(
2020
).
25.
H.
Liu
,
G.
Li
,
X.
Zhao
et al, “
Investigation of the impact of the thermoelectric geometry on the cooling performance and thermal—Mechanic characteristics in a thermoelectric cooler
,”
Energy
267
,
126471
(
2023
).
26.
L.
Shen
,
W.
Zhang
,
G.
Liu
,
Z.
Tu
,
Q.
Lu
,
H.
Chen
, and
Q.
Huang
, “
Performance enhancement investigation of thermoelectric cooler with segmented configuration
,”
Appl. Therm. Eng.
168
,
114852
(
2020
).
27.
S.
Manikandan
and
S. C.
Kaushik
, “
Energy and exergy analysis of an annular thermoelectric cooler
,”
Energy Convers. Manage.
106
(
106
),
804
(
2015
).
28.
D. Vi K.
Khanh
,
P. M.
Vasant
,
I.
Elamvazuthi
, and
V. N.
Dieu
, “
Geometric optimisation of thermo-electric coolers using simulated annealing
,”
Int. J. Comput. Sci. Eng.
14
(
3
),
279
(
2017
).
29.
T.-H.
Wang
,
Q.-H.
Wang
,
C.
Leng
, and
X. D.
Wang
, “
Parameter analysis and optimal design for two-stage thermoelectric cooler
,”
Applied Energy
154
(
154
),
1
(
2015
).
30.
S.
Lin
and
J.
Yu
, “
Optimization of a trapezoid-type two-stage Peltier couples for thermoelectric cooling applications
,”
Int. J. Refrig.
65
,
103
(
2016
).
31.
X.
Jia
and
Y.
Gao
, “
Optimal design of a novel thermoelectric generator with linear-shaped structure under different operating temperature conditions
,”
Appl. Therm. Eng.
78
(
78
),
533
(
2015
).
32.
X.
Wang
,
H.
Wang
,
W.
Su
,
T.
Chen
,
C.
Tan
,
M. A.
Madre
,
A.
Sotelo
, and
C.
Wang
, “
U-Type unileg thermoelectric module: A novel structure for high-temperature application with long lifespan
,”
Energy
238
(
Part B
),
121771
(
2022
).
33.
M. S.
Pourkiaei
,
H. M.
Ahmadi
,
M.
Sadeghzadeh
et al,
Thermoelectric cooler and thermoelectric generator devices: A review of present and potential applications, modeling and materials
,”
Energy
186
,
115849
(
2019
).
34.
I.
Chowdhury
,
R.
Prasher
,
K.
Lofgreen
,
G.
Chrysler
,
S.
Narasimhan
,
R.
Mahajan
,
D.
Koester
,
R.
Alley
, and
R.
Venkatasubramanian
, “
On-chip cooling by superlattice-based thin-film thermoelectrics
,”
Nat. Nanotechnol.
4
(
4
),
235
(
2009
).
35.
A. S.
Al-Merbati
,
B. S.
Yilbas
, and
A. Z.
Sahin
, “
Thermodynamics and thermal stress analysis of thermoelectric power generator: Influence of pin geometry on device performance
,”
Appl. Therm. Eng.
50
(
1
),
683
692
(
2013
).