3.1. Thermoelectric Measurements
NaxCoO2 exhibits an anisotropic magnetoresistive effect or simply magnetoresistance (MR); when the magnetic field is applied axially (parallel), the MR remains negative. For the radial or perpendicular magnetic field, the ferromagnetism is present in the first case and is suppressed so that the MR changes from negative to positive. When the material is no more ferromagnetic and has no Curie temperature or hysteresis effect with the changing temperature, becoming paramagnetic.
The strong electron to electron influence and the triangular lattice frustration, which induct a narrow band when a magnetic field is applied, are linked to the large thermoelectric power and to the superconductive state behavior when water is added. This ceramic material can be considered a Mott insulator or a strongly correlated material when
. A phase transition in the material occurs around 350 K. Blangero et al. suggest that Na
+ interlayer redistribution between CoO
2 layers should be the cause of this phase transition [
22]. The Na
+ interlayer redistribution could also lead to a large thermoelectric power when a magnetic field is applied.
Other dopants can lead to a large thermoelectric power. Polycrystalline samples with a base sodium content of 0.5–0.87, doped with elements such as calcium (Ca), bismuth (Bi), strontium (Sr), and (Cu) up to levels of 0.1 or 10%, exhibit a higher Seebeck coefficient and power factor than undoped sodium cobaltite [
16,
23,
24,
25,
26]. When dopants such as Sr or Ca are incorporated into the Na layer, they decrease disorder and increase the overall crystallinity of the system. As a result, the mean free path of the free carriers increases, which enhances both their mobility and the probe’s conductivity [
27], ultimately raising the figure of merit (ZT) [
23]. The same effect is expected when the sample is hydrated and exposed to a magnetic field.
The 30 mm-diameter probe exhibited a resistance of 1.9 MΩ at 330 K without a magnetic field, dropping to 380 KΩ when magnets were positioned above and below the pill. Correspondingly, the sample current increased from 0.14 µA to 0.68 µA. At 340 K, resistance was 1.65 MΩ without the magnetic field, falling to 240 KΩ with the magnets, and the current rose from 0.28 µA to 1.9 µA. See
Figure 5. Measured voltage was 0.26 V at 330 K and 0.45 V at 340 K, showing a 0.19 V increase over a 10 K temperature rise [
28]. The 30 mm-diameter probe magnetoresistance was five to seven times lower than the resistance measured without magnets.
A maximum current of 1 µA was observed for the 10 mm-diameter probe, while a current of up to 3.5 µA was measured for the 30 mm-diameter sample, both at a hot-side temperature of 375 K without a magnet attached. When the same NdFeB magnets were used, an average magnetic field of 0.1 T was applied to the 10 mm diameter, 10 mm thick sample, and an average field of 0.2 T was applied to the 30 mm diameter, 2.5 mm thick sample. The 10 mm diameter probe exhibited a much smaller decrease in resistance and a smaller voltage increase than the 30 mm sample (see
Figure 6 and
Figure 7).
Furthermore, the 10 mm probe showed a voltage of 0.1 V at 380 K, while the voltage for the 30 mm sample approached 0.3 V at the same hot-side temperature without the application of a magnetic field. See
Figure 6.
From
Figure 6 and
Figure 7, it can be concluded that the 10 mm sample is not suitable for the transducer design; thus, the 30 mm-diameter sample was selected for further analysis.
Applying a magnetic field inside the pill results in a substantial reduction in dynamic drain resistance compared to the non-magnetized sample. Consequently, thermoelectric power and efficiency are observed to improve markedly due to lower dynamic drain resistance for the identical sample, suggesting that applying a magnetic field should be a viable method to enhance the performance of Na
xCoO
2 based thermoelectric devices. Under an average axial magnetic field of 0.2 T inside the 30 mm-in-diameter sample, applied in the direction of temperature gradient, there is a significant enhancement in voltage, beginning at 0.13 V at 300 K, rising slowly to 0.2 V at 320 K, and soaring to approximately 0.45 V as the temperature approaches 340 K. See
Figure 7. This notable rise in voltage in the presence of a magnetic field is accompanied by the anisotropic magnetoresistance (MR) effect displayed by Na
xCoO
2. Axially applied fields sustain a negative MR, while a perpendicular application shifts MR from negative to positive, indicating the suppression of ferromagnetism at lower than 300 K temperatures.
Under an average axial magnetic field of 0.1 T applied along the temperature gradient in the 10 mm-diameter sample, the voltage increased from 0.025 V at 320 K to 0.08 V at 360 K. In comparison, when no magnets were attached, the voltage was 0.015 V at 320 K and 0.05 V at 360 K. Also, it can be noted that at 380 K, while the magnets are demagnetizing, the generated voltage remains the same, 0.1 V. Above 380 K, the NdFeB magnets are completely demagnetized, so no difference in voltage is observed. Still, this transducer’s operating temperature range of 300–380 K is preferred, as it encompasses the water boiling point (373 K), and NdFeB magnets, which are the strongest in the world, offer a maximum remanent magnetic field of 1.5 T (for N55 grade).
The observed electrical behavior, in conjunction with strong electron correlations and lattice frustrations present within the triangular lattice, contribute to substantial thermomagnetic and thermoelectric power, 1.75 µW at 360 K and 13.7 µW at 380 K. See
Figure 8. There is a significant power gain with just a 20 K temperature difference.
At 300 K or at room temperature, the sample of sodium cobaltite has a monoclinic cell arrangement; the sodium ion crystallographic position is shifted out from the center of the prism. At room temperature, the crystallographic structure has a lower symmetry and the Na
+ ions’ motion is also low. As the phase transition occurs near 350 K [
22], the crystallographic cells change into a rhombohedral structure, mainly manifesting as a Na ion rearrangement in the interslab to form a higher symmetrical position. The Na mobility will increase as the temperature increases. The motion of Na
+ in the interslab due to phase shifting creates an additional electrical field that can alter the probe’s voltage response. See
Figure 7. Therefore, the presence of an axial magnetic field, the sample hydration, and the phase shifting near 350 K, all allow energy to be harnessed in a more efficient way. The presence of water inside sodium cobaltite leads to its partial autoionization, generating additional ionic charge carriers that are guided by the electric and magnetic fields, along with the usual electronic carriers in the material. In hydrated thermoelectric material, the viscosity of the solution tends to decrease as temperature increases. Lower viscosity allows ions to diffuse more easily, further enhancing ionic and electronic conductivity. In this case, the applied magnetic field also contributes to increased electronic conductivity due to negative magnetoresistance.
In the next subsection, the humidity and water effect is presented. It is expected that the partial autoionization of water in the presence of sodium cobaltite would further improve free carriers’ mobility and conductivity, ultimately leading to an increased power factor.
3.2. Water or Moisture Effect on NaxCoO2
Water or moisture from the air increased the CoO
2 layer thickness, because the hydrogen atoms of hydronium ions H
3O
+ were attached to the oxygen atoms inside the CoO
2 layers. This bond was creating a tensile strain of the interlayer distance, from which Na atoms can go through a critical transition, meaning that from this point CoO
2 lattices can become a set of non-interacting monolayers [
21]. When hydration time reaches or exceeds 10 days, two other phases are formed, one with four water molecules and another with six water molecules. While one hydrogen bonds to an oxygen atom in the CoO
2, each oxygen atom and second hydrogen lies in a plane between the Na and Co layers. Such intercalation of four and six water molecules largely separates the CoO
2 layers and expands the c-axis parameters. To be noted that the distance between crystal layers “c” increases from c = 11.15 Å, for non-hydrated sodium cobaltite to the longest value c = 22.38 Å, for the hydrated phase with six water molecules.
However, these configurations with four and six water molecules are not stable for long at room temperature [
29]. The water insertion between alternately stacked CoO
2 layers and Na layers creates a thermo-chemical battery-like structure. This thermo-chemical battery is based on the deintercalation or intercalation process of Na
+ ions [
30]. One major characteristic of this thermo-chemical battery is that it can store thermal energy as chemical energy, and it can generate ionic currents under the temperature difference. This combination of the accumulation and the generation of electricity could be very useful to integrate into small power sources. The same process can be used to create fuel cells and supercapacitors [
31,
32], where the sodium cobaltite is selected as electrode material. The materials showed a significantly enhanced surface capacitance of 5.69 F cm
2. At or above room temperature, the chemical structure of hydrated Na
xCoO
2 evolves into Na
x(H
3O)
z(H
2O)
nCoO
2, featuring a mix of Na
+ and H
3O
+ ions interspersed with water molecules, within the CoO
2 layers. Here z is the count of the total number of hydronium atoms and n is the count of the total number of water atoms. The water is partially ionized. This unique configuration significantly bolsters electrical conductivity due to the increased ion count. To be noted is that the configuration from
Figure 1 and
Figure 3 was used in all measurements regarding the sodium cobaltite sample hydration. Also, the 30 mm-in-diameter sodium cobaltite sample was manually hydrated with three dosages, 1 mL, 1.5 mL, and 2 mL, by using a syringe. Experiments showed that an optimal hydration dosage of 2 mL increased water evaporation time without damaging the probe, and this dosage was used in all subsequent experiments.
The electrical resistivity variation with the temperature of the hydrated sample, placed in a constant magnetic field (see
Figure 1 and
Figure 3), was recorded with an HMC8012 Bench Digital Multimeter, produced by Rohde & Schwarz GmbH, Munich, Germany.
At 312 K, the electrical resistivity of the hydrated probe is
, at 322 K is
, and it then decreases to
at higher temperatures, 360–370 K. See
Figure 9. The hydrate formula
is calculated by dividing the number of moles of water by the number of moles of
.
The incorporation of water and the application of a magnetic field are projected to escalate these values further. In pristine NaxCoO2 samples with x > 0.8, the addition of water markedly surges the current from 0.5 μA to a peak of 320 μA at 340 K, enhancing power output by 15 times. However, at temperatures reaching 370 K, rapid water evaporation diminishes power output sharply within three minutes. A viable strategy to mitigate this involves operating at temperatures below 340 K and leveraging the temperature change on the cold side to maintain an optimum temperature gradient.
Although voltage increases with temperature from 0.13 to 0.24 V over a 310–380 K temperature range (see
Figure 10), the observed peak idle current was between 1.1–1.8 mA for temperatures around 350–360 K. When the water starts to evaporate over 370 K, the measured current drops rapidly down to
. The estimated power of 200 µW at 380 K can be maintained for several seconds only. See
Table 1.
The formation of hydrogen bonds within sodium cobaltite is explained in the next paragraph. Hydrogen bonds are a type of dipole–dipole interaction that occurs between an electronegative atom (such as oxygen) and a hydrogen atom bonded to another electronegative atom. In the context of sodium cobalt oxide, the lattice oxygen atoms are highly electronegative, making them prime sites for hydrogen bonding with water molecules. When water molecules intercalate into the sodium cobalt oxide lattice, they form hydrogen bonds with these lattice oxygen atoms. This can be represented by
which represents the hydrogen bond formed between the water molecule’s hydrogen and the lattice oxygen. The water molecule intercalation has an impact on sodium cobaltite material properties. Hydrogen bonding can alter the electronic environment of lattice oxygen atoms, affecting the material’s electronic band structure and electronic conductivity. Intercalated water molecules can scatter phonons, which are heat-carrying particles, thereby reducing thermal conductivity and improving the thermoelectric figure of merit (ZT). Additionally, the introduction of water molecules and the formation of hydrogen bonds can influence charge carrier concentration and mobility, leading to large variations in the Seebeck and thermomagnetic (if magnets are added) coefficients, even for small temperature differences between the hot and cold sides. Water molecule intercalation increases the electrochemical reactions of NaCoO
2. The insertion of water molecules enhances the mobility of Na
+ ions within the layers, thereby improving ionic conductivity, which can be quantitatively assessed using electrochemical impedance spectroscopy (EIS) by measuring the decrease in charge-transfer resistance in the presence of water [
15]. Water’s role in modifying the electrochemical stability and cycling performance of NaCoO
2 electrodes in sodium-ion batteries is critical; it can either facilitate sodium ion transport, improving charge/discharge rates, or induce structural instability, leading to capacity fade, with cyclic voltammetry (CV) and galvanostatic charge–discharge tests being instrumental in evaluating these effects [
33,
34].
Thermal and chemical stability of hydrated NaCoO
2 can be monitored by various techniques. Differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA) can be used to study the thermal behavior of hydrated NaCoO
2, helping determine the temperatures at which water is released from the structure, indicating the stability of the hydrated phases [
35]. Long-term exposure to moisture can lead to the leaching of sodium ions and oxidation or reduction of cobalt in the lattice, which can be monitored by using chemical analysis techniques such as inductively coupled plasma mass spectrometry (ICP-MS) to track elemental composition changes over time [
18]. Sodium cobaltites have been used as electrolytes for fuel cell applications. A fuel cell can be fabricated from a Pd/NaCo
2O
4/NaCo
2O
4 disk and placed between two sheets of platinum mesh current collectors at anode and carbon paper current collectors coated with Teflon and Nafion [
36]. An increase in thermomagnetic and thermoelectric power is observed in
Figure 11, with values of 50 µW at 340 K and 100 µW at 357 K, when a quantity of 2 mL of water is injected into the 30 mm-diameter sample. The power doubles with a temperature difference of just 17 K.
To be mentioned again is that the 30 mm-in-diameter sample was manually hydrated with the optimum dosage of 2 mL by using a syringe, and placed in the electrical heat-press, for all experiments presented above. A peak voltage of 0.24 V was obtained at 380 K and 0.17 V at 360 K as the probe cooled to ambient temperature. The current dropped from 1.8 mA to 0.82 mA in just 2 min, when the temperature was increased from 350 K to 360 K but, after that, the current gradually decreased from 0.82 mA to 0.2 mA in 4 min, until the injected water evaporated. Upon sample cooling from 360 K (current was 0.2 mA) to 350 K, the current slowly reached 100
A after another 10 min. All graphs from
Figure 9,
Figure 10 and
Figure 11 were realized by letting the sample cool to ambient temperature. An average current of 0.82 mA that lasts about 6 min can be considered at 360 K to be a hot-side temperature. To demonstrate the reliability and durability of the proposed device under hydration conditions, the effect of adding 1 mL, 1.5 mL, and 2 mL of water was monitored for several hours, with 42 mm-diameter NdFeB magnets attached to the sample. These experiments were repeated several days per week for over three months to determine whether the resulting voltage and current data were conclusive. In
Figure 12, voltage variation following the addition of 1 mL of water was recorded over several hours. Each time water was added, the voltage spiked to 1.4–1.5 V for a brief period (seconds) before dropping to 1.1–1.2 V and gradually decreasing to 0.3 V over a period of 12 to 15 min. It took approximately 23–30 min for the voltage to stabilize at its lowest value, 0.2–0.3 V, at a constant temperature of 350 K, remaining steady until the injected water evaporated and only ambient humidity remained.
Here, y is the number of moles of water per moles of sodium cobaltite. Experimental observations indicate that an optimal sodium concentration of approximately x ≈ 0.88 significantly enhances the thermoelectric performance of NaxCoO2 over a hot-side temperature of 370 K. This enhancement could be further amplified by substituting sodium ions with holes, where holes will be replaced by H3O+ ions obtained from water autoxidation in the cobalt oxide layers, providing a route to increase the power factor. It should be noted that a small fraction (1–10%) of the H3O+ ions replace Na+ ions, while the majority of hydronium ions attach to the oxygen atoms within the CoO2 layers. For temperatures higher than 370 K, the figure of merit ZT increases from 0.01 to 0.06.
3.3. The Electronic Voltage Booster
The electronic scheme shown below is used to capture useful electrical energy from five or more thermoelectric modules made of Na
xCoO
2 pills stacked and connected parallel. For this parallel connection, it is assumed that all thermoelectric modules are identical, resulting in no variation in output voltages. The electronic device of energy harvesting type is based on the properties of hydrated sodium cobaltite, placed in a magnetic field. A single Na
xCoO
2 thermoelectric pill generates 15–20 mV with small temperature differences of 10–12 K between the hot and cold side. The maximum current generated by a single sodium cobaltite sample at a temperature of 340 K is 350 µA. The TJ1 and TJ2 JFET 2SK117 (see
Figure 13) transistors used in the circuit can initially conduct and open with a gate-source voltage (VGS) of −0.1 V and a current of 1.5 mA; with only −0.05 V gate-source voltage, the current required to open the transistor is at least 2–3 mA.
It is obvious that the thermoelectric generator (TEG) can be used also in cold weather, when the air temperature is only 280 K or less. The most important fact is that, in any situation, we should have at least a 30 K difference between the hot and cold side to obtain a voltage of 50 mV. In addition, from two to four TEGs could be connected in series to obtain more voltage. As can be observed, each TEG will need no more than 10 pills to make the voltage booster or the Meissner circuit work.
From the datasheet of the 2SK117 JFET transistor, all parameters can be found: variation of drain-source VDS voltage, drain current ID, and all the gate-source voltage (VGS) conditions. In case of higher input voltage of VGS being around −0.3V…−0.4 V, the input current can be less than 0.2–0.1 mA.
In the schematic, there are two 2SK117 TJ1 and TJ2 transistors in parallel to share the input current (the JFET transistor is not overloaded in current) and to halve the dynamic drain resistance. Lower resistance also proportionally reduces conduction losses. In conduction, the 2SK117 JFET transistors have a dynamic drain resistance of 50 Ω, while for the 2SK170 JFET, the dynamic drain resistance is between 50 and 150 Ω. For any transistor used as a switch, it is preferable that dynamic drain resistance is minimal to reduce conduction losses. The BS170 MOSFET TM1 transistor has exactly this role of taking over the current in the circuit (it closes) and significantly reducing conduction losses (1.8–5 Ω dynamic drain resistance) when the inductive voltage pulse initially produced by the opening of the JFET transistors is sufficiently high at the gate, starting from 0.8 V and up.
The capacitor between the two gates prevents the simultaneous opening and closing of the transistors, creating a slight delay that allows the inductive pulse to propagate for a determined duration in the transformer’s secondary. When the gate voltage is sufficiently high (>0.8 V), the TM1 BS170 MOSFET transistor takes over the Meissner circuit oscillation up to a maximum voltage of 18 V. The oscillation frequency of the Meissner circuit is mainly determined by the inductance or reactance of the secondary and the capacitance of the capacitor between the TJ1 and TJ2 JFET gate and the TM1 BS170 MOSFET gate, if this is greater than the internal parasitic capacitance of the BS170 and 2SK117 transistors.
The internal parasitic capacitance of the transistors (C
GS) is sufficiently large, 62 pF. Because the capacitance
between TJ1 and TM1 gates is summed with C
GS and TJ1 input capacitance of 13 pF at saturation, resulting in 97 pF, the oscillating frequency will be around 35 kHz. At this oscillation frequency and with a transformer ratio of 1:48, a maximum gate voltage of 18 V can be reached with an input voltage of 0.4–0.5 V (see
Figure 13 and
Figure 14). Proportionally, with a voltage of 0.2 V, the maximum output voltage obtained on the secondary is 8 V. Also, for an input voltage of 0.1 V, the maximum output voltage will be 4 V.
The 1–2 µF filtering capacitor placed between the secondary of the transformer and ground also serves to limit the voltage to a convenient value for the voltage regulator part.
Since the 1–2 µF filtering capacitor is connected in a series with a combined capacitance of 97 pF, the smaller capacitance (97 pF) will be considered in the Meissner circuit and will be used to determine the oscillating frequency.
It is known that a voltage regulator is more efficient when the input voltage is closer to the output voltage. The capacitor must also be chosen based on the current provided by the thermoelectric generator (TEG) at the transformer’s output; the lower the current, the smaller the capacitor can be and, conversely, the higher the current provided by the generator, the larger the capacitor required for voltage limitation at the output.
For a thermoelectric generator (TEG) with 10 sodium cobaltite pills of 30 mm in diameter and thickness of 2.5 mm, the current at the output of the transformer secondary is 3.5 mA/48 = 73 µA. So, the current drops under 100 microamperes when it reaches the voltage regulator part, with the voltage being limited to 4 V by the respective capacitor (at 0.1 V TEG input). For such small currents, a classic Zener diode would not function correctly and, moreover, all generated power would be consumed by the diode. To prevent a Zener diode from consuming too much of the power generated by the thermoelectrics, the operating current of the diode needs to be as low as possible. Current technology has achieved a 50 microamperes operating current (MMSZ4689-TP SMD type Zener diode used for 5.1 V in all electronic schematics, a component produced by Vishay Semiconductors Company, Shelton, United States).
The diode can operate even below that operating current, down to 20 µA, but in that case, the voltage at the output of the regulator drops below 4 V. The TM2 BS170 MOSFET transistor, a component produced by Infineon Technologies AG, Neubiberg, Germany, used also in the voltage regulation part requires very low gate currents (10…50 µA; the gate of a MOSFET is voltage-controlled) and can support maximum currents of 500 mA between the drain and source.
The 100 × 30 mm electronic board can be further reduced geometrically if all the components are available in the SMD variant. See
Figure 14. The 2SK117 JFET transistor from the electronic scheme, a component produced by Toshiba Corporation, Tokyo, Japan, can be replaced by the equivalent 2SK209 SMD, but the transformers are not available in sizes smaller than 10 × 10 × 10 mm. Even if one can be found, very small transformers are wound with very thin winding wire of 0.05mm at a 1:48 ratio, which implicitly increases the dynamic drain resistance of the coil and total conduction losses. A larger transformer with a 15 × 25 mm magnetic core is preferred because it is wound with 0.2–0.3 mm wire; this way the conduction losses are minimized. The same issue occurs with the coil used in switching voltage regulators; it cannot be significantly reduced in size because two essential conditions would no longer be met: inductance greater than 1–10 mH and dynamic drain resistance below 1 Ω.
The electronic board also contains an astable circuit that signals with a red (or yellow) LED every 5–10 s. The high resistance of 4.7 MΩ and 500 KΩ used together with a Schottky diode 1N5817 on the opposite side of the LED minimize energy consumption and allow monitoring of the supercapacitor’s charge level. The red (or yellow, or white) LED can also light up when the voltage on the supercapacitor reaches 1.8 V, 2.2 V, or 3 V. A red LED, without a current-limiting resistor, connected directly to the supercapacitor (the circuit also has a parallel socket for connecting two pins to the load) at a voltage of 3 V, stays lit for at least 1 min.
The estimated rate of water evaporation at 360 K was approximately 4 mL/h, given a specific surface area of 7 cm
2 and a relative humidity of 40%. At 360 K, an average current of 0.82 mA was recorded (see
Table 1) over a 6 min period, with a voltage reaching 0.165 mV. However, water evaporated more quickly, completing in 16 min (compared to the estimated rate of 2 mL per 30 min). Therefore, the current of 0.82 mA was sustained for 6 min, followed by a lower average current of 0.14 mA for an additional 10 min. At 350 K, the current and voltage were sustained for 23 min after 1 mL of water was injected into the sample. The voltage spikes suggest that a sudden current drop occurs simultaneously, keeping thermoelectric power within the same range. However, the crystallographic phase transition at 350 K requires further analysis.
At 340 K, the estimated rate of water evaporation under the same conditions was lower, at 1.7 mL/h. Since the 30 mm-diameter pill was hydrated with 2 mL of water, the estimated time until evaporation was 70 min. Experiments recorded an average current of 0.35 mA (see
Table 1), sustained for 35 min. Each pill will require an additional 1.7 mL every half hour to function properly.