2 1.2 (a) Model of a thermoelectric couple including all electronics and thermoelectrics. b) The thermoelectric effect is the result of thermodiffusion of charged carriers from the hot end to the cold end. The sample is shown in yellow. heater block in green and thermocouple in blue. a) two-point linear design.
Summary of Introduction
Motivation of Thermoelectric Research
Thermoelectric conversion of high-exergy industrial waste heat could provide approx. 10 TeraBTU/year of recovered energy in the USA alone [76]. Conversion of waste heat from cars can improve fuel economy by 5%, thereby saving a much larger 1 Quadrillion BTU/year [63], but unit integration is more complicated due to the different operating conditions of cars [104].
Thermoelectric Energy Conversion
To paraphrase Max Planck's formulation [158], "the rate of entropy production of a heat engine is always equal to or greater than zero." At the Carnot efficiency, the entropy production is equal to zero. It will transport through the Peltier effect an amount of heat equivalent to the power produced.
Entropy Co-Transport
It can be re-expressed in terms of the entropy density (s) and the carrier concentration (n) as presence =−qdnds. Here we consider coupling the carrier transport to degrees of freedom associated with the entropy associated with an order–disorder phase transition.
Super-ionic Thermoelectrics
In type I superionics such as AgI [97], the ionic conductivity increases abruptly at a phase transition temperature. For type I superionic conductors, there is a sudden enthalpy release at the phase transition temperature (i.e., a first-order phase transition) and a simultaneous discontinuous increase in ionic conductivity.
Key Challenges and Results
In Figure 2.2, the mangetic field will be applied through the sample, while the current is applied from wire 1 to wire 3 and voltage measured from wire 2 to 4. Finally, the Seebeck measurement is path independent [49] — the voltage depends only on the temperatures at the measurement points without regard to temperature.
Synthesis
An encapsulated sample of indium was measured under thermal diffusivity, and the difference between its observed melting temperature and its literature value (156.6 ◦C) was used to confirm the calibration. Electrical conductivity could not be checked in the same way and its calibration could not be confirmed.
Chemical Characterization
This is done by pressing a thermocouple on a hotter point in the sample and a thermocouple on a colder point on the sample, see figure 3.2. The Seebeck coefficient, as defined in Equation 1.1, is actually the ratio of the gradients of V and T. Single-point Seebeck measurements are therefore an approximation to. The thermocouples contact the sample without the mediation of the heating block as in the four-point linear design.
Apparatus and Protocols
The wing nuts are located below the top plate and are used to separate the two boron nitride blocks without external mechanical support (eg the user's arm). On occasion—for example, immediately after installing new thermocouples—the Seebeck should be tested at slightly above room temperature (40 to 50◦C) in both air and vacuum to ensure that the pressure is sufficient to overcome errors. of thermal contact described above. .
Challenges of Phase Transition Seebeck
While this error is small for a typical thermoelectric material, for phase change materials it can be significant. Raw Seebeck plots in which some of the ∆T is above and below the phase transition are difficult to analyze.
The Multi-Ramp Seebeck Technique
Quantitatively, the fraction of loss rate per hour is a function of the loss rate (LR), the geometric density (ρ) and the leg radius (r) as: This provides a limiting condition on the thickness of baffles that can be effectively added to the module. The second major problem was the chemical degradation of the material and device, especially under applied current conditions.
It is also possible that these cracks are related to the approximately 1.4% volume expansion at the 410K phase transition of the material [191].
Order-Disorder Phase Transitions Type
Therefore, the temperature of the phase transition is ≈60,000 K. and monatomic oxygen is actually very rarely observed. If ∆S and ∆U are themselves independent of temperature, there will be a first-order phase transition. Therefore, a continuous transition requires that the entropy of the ordered phase increases continuously, and that its. internal energy increases continuously to the phase transition temperature.
In the vicinity of an order-disorder phase transition, the free energy is described in terms of an order parameter.
Ag 2 Se
During the heating and cooling process, diffractograms were measured from 2θ= 20◦ to 2θ= 52◦ to determine the nature of the phase transition. The discrepancy in the phase transition during heating and cooling is characteristic of a first-order process and a consequence of the non-adiabatic heating. The DSC curve for a first order transition should start to rise at the phase transition temperature.
The phase transition is between the β−Cu2Se(RT) phase and the α−Cu2Se(ht) phase.
Cu 2 Se
Diffractometry
To investigate this effect, we plot the peaks that disappear at the phase transition temperature (Figure 5.13(a)). These peaks show critical power law behavior as the phase transition temperature is approached (Figure 5.13(b)). The phase diagram of Vucic (Figure 5.9) anticipates a (non-lambda) second order transition at a lower temperature.
By studying the high-temperature structure of Cu2Se, it is clear that the peak at 4.1 ˚A (Figure 5.15(a)) is a superposition of the shortest Cu-Cu and Se-Se distances in the [110] direction.
Cu 2 Se Calorimetry
However, the low thermal diffusivity and nonlinearity of the heat capacity can still lead to errors in the DSC-derived heat capacity. A fixed temperature rate causes an error in the heat capacity resolution due to the thermal diffusivity of the sample. The heat capacity measured at these two facilities showed clearly different calorimetry curves (Figure 5.16), which can be explained by equation 5.7.
At 355 K there is a noticeable change in the slope of the heat capacity. (Figure 5.17) This is consistent with a second-order (non-lambda) phase transition and Vucic's predictions.
From the crystallographic data, it appears that Cu1.97Ag0.03Se has a broken and distorted version of the second-order transition of Cu2Se, that is, it appears to be a weak first-order transition. The deformation occurs with dissolution of the secondary CuAgSe phase near 380 K. The heat capacity of Cu1.97Ag0.03Se shows a doubled peak, see Figure 5.22 The temperature of the first peak corresponds to the temperature dissolution of the observed CuAgSe phase by crystallography. Some of what is labeled as specific heat in Figure 5.22 is surely enthalpy of formation due to the first order component of the phase transition.
The solid heat capacity is due to the contribution of kinetic and potential energy to the heat capacity from the equivalence theorem.
Band Structure Modeling
The carrier concentration is the number of carriers between the band edge and the Fermi level. These models are successful because only band states within 3 kbT of the electronic chemical potential (i.e. the Fermi level) contribute significantly to electron transport [128]. Thermoelectrics are heavily but not metallically doped, so that the dominant conductor band dominates, but the Fermi level is not far from the band edge.
3qh2 n−2/3T m∗(1 +λ) (6.5) Again, the inverse Seebeck dependence on n, which is generally defended in the introduction, is present.
Transport Measurements
Fivos Drymiotis was measured by the oscillation method and showed a comparable change in the Seebeck coefficient through the phase transition. The lattice thermal conductivity of Ag2Se decreases only slightly as the temperature increases through the phase transition temperature, indicating that most of the change in overall thermal conductivity is due to the decrease in electrical conductivity, see Figure 6.7(a). The drop in zT is 30%, which is consistent with the anomalous 15% drop in the Seebeck coefficient at the phase transition temperature as zT varies with Seebeck in second.
The first term represents the entropy transport due to the change in the number of carriers present.
Entropy Co-Transport
Here we consider coupling the carrier transport to degrees of freedom related to the structural changes of a phase transition. A phase transition is always associated with an entropy change because there is always a simultaneous transformation in system symmetries [162]. In chapter 5 we found that the phase transition of Cu2Se has a significantly increased heat capacity over a wide temperature range.
If part of the entropy associated with the phase transition is transport-related, a large Seebeck enhancement is possible.
Cu 2 Se Transport near the Phase Transition
Analysis
The square of the Seebeck excess over the band structure prediction (b) explains the size and width of the zT peaks. When the square of the measured Seebeck divided by the predicted Seebeck band structure is compared to zT, as in Figure 7.7b, it is seen that the anomalous increase in Seebeck almost explains the observed width and height of the zT peak. The Seebeck excess (Δα) compared to the band structure slightly overestimates the height of the zT peak. α are significantly increased over the same temperature range of 393 K to 410 K.
This suggests that some aspect of the lambda-type phase transition increases the zT of Cu2Se.
Comparison of the thermal properties of Cu1.97Ag0.03Se and Cu2Se is particularly illustrative. Between 390 K and 410 K there is a steady decrease in the diffusivity, indicating that the solution of the secondary phases widens the phase transition region. The anomalous range corresponds to the temperatures between the first order phase transition and the decomposition of the CuAgSe secondary.
zT and Seebeck for the mixed phase transition of Cu1.97Ag0.03Se is similar to that of its Cu2Se main phase.
Ion-mediated Enhancement
For type I superionic conductors, there is a sudden enthalpy release at the phase transition temperature (i.e. a first-order phase transition) and a concomitant discontinuous increase in ionic conductivity [21]. Literature results show that Ag2Se ion conductivity increases by four orders of magnitude at 412 K, see Figure 8.2 [136] Disorder of the ions requires absorption of enthalpy. In a type II superionic phase transition, the ion disorder propagates continuously (as in a second-order transition) until the phase transition temperature is reached.
Mechanistically, it may function through a dependence of occupation of soft modes (e.g. the Zn4Sb3 ratchet) [173] on the concentration of Ag+ and Cu+.
Future work
Relationships between the crystal structure and electronic properties of the compounds Ag2S, Ag2Se, Cu2Se. The Hubbard Model Thermoforce: Effects of Multiple Orbitals and Magnetic Fields in the Atomic Boundary. Single-crystal neutron diffraction analysis of the cation distribution in the high-temperature phases Cu2−xS, Cu2−xSe and Ag2Se.
Phase transitions in solids: an approach to the study of chemistry and physics of solids.