The present invention relates generally to devices and techniques that convert the heat energy stored in the entropy of a hot gas into electrical work, without mechanical motion.
Almost 90% of the world's primary energy consumption occurs through the use of heat, and approximately 60% of this thermal energy is lost as rejected waste heat. Given the large magnitude of energy in waste heat, its efficient conversion to electrical power offers a significant opportunity to lower greenhouse gas emissions across the energy sector, transportation, and manufacturing. However, it has been difficult to optimize the performance of new direct energy conversion approaches because of the inability to decouple entropy change from thermal and electrical transport in materials and continuously operating devices.
Thermoelectric (TE) heat engines are widely considered to be the most promising candidates for converting distributed heat sources to electricity, with their electrochemical analogues, thermogalvanic (TG) heat engines, also receiving attention. For both, the inability to completely decouple carrier entropy (thermopower α), heat transport (thermal conductivity κ) and charge transport (electrical resistivity ρ) in a single material has so far kept efficiencies well below those of traditional thermofluid cycles. Alternatively, thermally regenerative electrochemical cycles separate charge and heat transport in the time domain, but only produce electricity intermittently. Electrochemical sodium heat engines, which have demonstrated high efficiency and reasonable scalability, have stringent temperature requirements and stability issues. Thus, it remains an unsolved challenge in direct heat-to-electricity conversion to realize continuous thermodynamic cycles that completely decouple entropy, heat and charge transport and operate across a broad range of temperatures.
In one aspect, the present invention provides a thermoelectrochemical heat converter that generates electrical power from hot gases without mechanical motion. In operation, gas flows through solid-state ion-conducting membranes and generates current across an external load. A pump is not necessary to maintain operational pressures. Multiple membrane assemblies can be connected in series and in parallel to control the current and voltage output of the device.
The device is the first to directly convert the entropy of a hot gas to electrical work without the need of overpressures. The device can be operated as an open system with net flow of gas, or as a closed-loop system with recirculation. In the open-system case, an external pump is not necessary to maintain operational pressures. The device offers potential improvements in heat conversion efficiency, temperature range, and power per unit weight over a thermoelectric.
In one aspect, a device according to an embodiment of the invention includes two ion-transporting, gas-impermeable, and electron-blocking membranes, one at the hot side, and one at the cold side. Each membrane is contacted on both sides by porous electrodes connected to external loads.
In some embodiments, membranes can be connected in series or in parallel to increase voltage or current output, respectively. A pump can be attached to recirculate gas in the system. The device can be connected with others, such as thermoelectric generators, for more efficient use of waste heat.
The heat converters according to the invention have many applications, including directly generating electrical power from hot gases or waste heat without mechanical motion, separating mixtures of gases, purifying hot gases, scrubbing carbon dioxide from smoke or exhaust, converting solar heat energy to electrical work.
In one aspect, the invention provides a device for direct thermoelectrochemical heat-to-electricity conversion. The device includes a first electrochemical cell having a first gas-impermeable, electron-blocking membrane capable of transporting an ion I at a first temperature, and a first pair of electrodes on opposite sides of the first membrane; and a second electrochemical cell having a second gas-impermeable, electron-blocking membrane capable of transporting the ion I at a second temperature lower than the first temperature, and a second pair of electrodes on opposite sides of the second membrane. The device also includes a closed-circuit chamber A having a working fluid A, a pump A, and a counter-flow heat exchanger A, wherein the closed-circuit chamber A is connected to the first electrochemical cell on a side A of the first membrane and to the second electrochemical cell on a side A of the second membrane; and a closed-circuit chamber B comprising a working fluid B, a pump B, and a counter-flow heat exchanger B, wherein the closed-circuit chamber B is connected to the first electrochemical cell on a side B of the first membrane and to the second electrochemical cell on a side B of the second membrane. The working fluid A is capable of undergoing a reversible redox half-reaction of the general form RA→OA+I+e− and wherein the working fluid B is capable of undergoing a reversible redox half-reaction of the general form RB→OB+I+e−, where RA and RB are reduced forms of active species in working fluid A and working fluid B, respectively, OA and OB are the oxidized forms of active species in working fluid A and working fluid B, respectively. The first electrochemical cell is connected electrically in series with the second electrochemical cell. One of the first pair of electrodes is porous to the working fluid A, and another one of the first pair of electrodes is porous to the working fluid B. Similarly, one of the second pair of electrodes is porous to the working fluid A, and another one of the second pair of electrodes is porous to the working fluid B. One of the first pair of electrodes is electrically connected to one the second pair of electrodes via an electrical load to produce electricity. The device thereby operates such that the first electrochemical cell runs a forward redox reaction, gaining entropy, and the second electrochemical cell runs a reverse redox reaction, expelling entropy.
Embodiments of the present invention provide continuous electrochemical heat engines based on two redox-active working fluids separated by ion-selective membranes. As shown in
The working fluids may be, for example, oxygen, hydrogen, water, carbon monoxide, carbon dioxide, or mixtures thereof. More generally, the working fluids may be liquids, gases, dissolved species or slurries, supporting redox processes with different entropies of reduction and containing a species that crosses the ion-transporting membrane as the ion I. The species that undergo redox reactions within the working fluids are distinct from the atom or ion that traverses the ion-conducting membrane; e.g. while the redox-active species could be complexes of transition metals undergoing outer-sphere electron transfer, the ion crossing the membrane (such as a proton, or a hydroxide, or others) does not have to participate in those reactions. Examples of working fluids are: (a) an aqueous or non-aqueous solution of redox couples, with supporting ions such as H+, OFF, Cl−, or others crossing the cell membranes. Redox couples could be complexes of transition metals (Fe, Cu, V, Co, or others), organic molecules (quinones, pyridines or others), polyelectrolytes, or some others. (b) Slurries of redox-active solid materials (such as lithium iron phosphate LixFePO4, lithium titanate, or others) and supporting ions, such as Li+ or others, (c) molten phases, e.g. metals, (d) gaseous phases as described below.
The first membrane or second membrane may be an ion-conducting ceramic, an ion-conducting polymer, or a molten salt. Examples include: yttria-stabilized zirconia (YSZ) or doped ceria (CeO2) for oxygen gas (as O2− in the membrane), yttrium-doped barium zirconates (BYZ) for hydrogen or water vapour (as H+ or OH− in the membrane), a molten carbonate salt (NaCO3, LiCO3, KCO3, their mixtures, or others) on porous (LiAlO2, beta-alumina, or others) support for CO and CO2 (as CO32− in the membrane). Further examples include: (a) beta- or beta”-alumina, other ion-conducting ceramics, or molten salts, for Na+, K+, or other metal ions crossing the membrane, (b) ion-conducting polymers, such as Nafion, PET or others for protons or hydroxide ions crossing in liquid solvents.
The porous electrodes may be alloys of W, Mo, Ni, other metals, or ceramics that could further be supported on an electronically conducting or mixed ion-electron-conducting framework, e.g., Pt on carbon cloth.
In some embodiments, the first and/or second electrochemical cell may include multiple electrochemical cells connected in series.
The principles of the present invention are highly generalizable: a wide range of species, including liquids, gases, dissolved species and slurries, supporting redox processes with different entropies of reduction, ΔS, can serve as the working fluids. Expressed per coulomb of charge transferred, ΔS manifests as the electrochemical thermopower α; the difference of the thermopowers in the two reactions α1-α2 determines the open-circuit voltage (OCV) output of the device as ΔVOC=|(α1-α2) ΔT|, where ΔT=(TH−TC). Table 1 lists the thermopower of individual liquid redox couples measured experimentally in this work, which allow for combined α1-α2 in excess of −3 mV K−1 (also shown in
Embodiments of the invention include two types of continuous electrochemical heat engines that operate at room temperature and up to ˜900° C., respectively. The ability to fully decouple entropy conversion, thermal transport, and electrical transport enables system efficiencies over 30% of the Carnot limit. Simulations suggest even higher performance at maximum power in scaled systems, making continuous electrochemical heat engines a promising new approach.
For the low-temperature embodiment, the aqueous V2+/3+∥Fe(CN)63−/4− couples were chosen on the merits of their high charge capacity, facile redox behavior, and large α1-α2. For the high-temperature system, oxygen gas was used as an entropy carrier via the H2/H2O∥O2 couples, mediated by solid-oxide electrochemical cells.
The high and low temperature embodiments are examples that represent select points in the wide space open to the materials and system design of continuous electrochemical heat engines. To further explore this parameter space, we developed a modeling framework to estimate the practical performance of continuous electrochemical heat engines. Based on a simple device configuration, the overall heat engine efficiency is given as:
where RL is the resistance loss in the leads and Paux is any auxiliary power input, such as a pump driving circulation. The thermodynamic heat input is I TH (α1-α2), QL reflects all conductive leaks in the system and heat leaks from the mass transport of reactants are (1−εHX) {dot over (m)} cp ΔT, where εHX is the effectiveness of the recuperative heat exchanger. The simulation includes mass transport and solution resistance as well as conductive heat leaks along the current collecting leads and along the length of a counter-flow heat exchanger. The relevant performance metrics for continuous electrochemical heat engines are the maximum power and efficiency at the maximum power point.
Maximum power density is shown in
An analysis of a generalized liquid-phase heat engine also points to two distinct operating regimes. For active species concentrations and temperatures corresponding to those in
Our modeling shows that the doubled number of interfaces relative to TE and TG systems does not necessarily limit performance, as the added irreversibility is compensated by the increased thermopower and counterflow heat exchange. Furthermore, the ability to form stacks of cells in series at each temperature to increase voltage without the coupling to heat losses is fundamentally different from that in TE systems: it enables further minimization of heat leaks while independently optimizing the electrical performance. Even after accounting for practical losses in a simple device configuration, the continuous electrochemical heat engine can scalably reach maximum power point efficiencies well over 30% of ηc under diverse operating conditions. It is also worth noting that stacks of multiple electrodes can achieve much higher areal power densities than individual cells. For example, with 100 cells per stack (the geometry simulated in
By decoupling thermal and electrical entropy generation pathways, the principles of the present invention enable effective energy conversion in regimes heretofore inaccessible to thermoelectric, thermogalvanic, regenerative, or other thermal-fluid heat engines. While electricity generation is described in this work, operating these systems in reverse could in principle enable electrochemical refrigeration as well. In addition to the significant flexibility in size and form, a vast parameter space exists for the optimization of working fluids: redox transformations of pure substances, ion-transporting liquids, and gas-phase reactants could all be used. With the development of suitable redox chemistries and flow systems, continuous electrochemical heat engines could fill a vital missing space in the existing landscape of energy harvesting technologies.
The V2+/3+∥Fe(CN)63−/4− energy harvester was tested under constant N2 purge in both electrode compartments. The negative electrode was fabricated using carbon cloth (ELAT hydrophilic, 400 μm thickness), heat treated in air at 400° C. for 30 hrs prior to the experiment to functionalize the electrode surface, as described previously. The positive electrode was fabricated with a 0.5 mg/cm2 Pt loading on carbon paper (Spectracarb 2050a, 252 μm thickness), as described above. The thicker (127 μm) Nafion membrane was to prevent crossover between the vanadium and ferrocyanide electrolytes. Prior to testing both electrode compartments were filled with 3MVOSO4 (Aldrich) in 6M HCl (Aldrich) in degassed water, 1 mL in the negative electrode compartment and 3 mL in the positive electrode compartment. A potential of 1.4V was then applied across the cell until 430 coulombs of charge had passed through the cell, such that the negative electrode compartment then contained a 1:1 mixture of V2+ and V3+, and the positive electrode compartment contained a 1:1 mixture of V4+ and V5+ species. The positive electrode compartment was emptied with a syringe, rinsed with degassed water, and re-filled with 375 mM K4Fe(CN)6 and 375 mM K4Fe(CN)6 in degassed pH 7.2 phosphate buffer prior to testing. The positive electrode compartment was covered in aluminum foil due to the known sensitivity of ferrocyanide compounds to light. The vanadium compartment was left open to monitor the blue-green-purple color change that confirms a successful reduction procedure.
In a scaled-up energy harvesting system, the circulation of electrolyte may result in a reduction of mass-transfer related overpotential, and correspondingly more favorable polarization behavior is expected. To gain insight into to the magnitude of the improvement that could be realized with fast flow outside of the membrane-electrode assembly, peristaltic pumps (ZJchao, 12V) were used to provide jet-impingement mass transfer enhancement inside both working fluid chambers. The improvement in mass transfer behavior is shown in
We used anode-supported solid-oxide button fuel cells available commercially from Fuel Cell Materials (ASC2.0) and used without modification. The cells with membranes 500, 502 were sealed with molten Ag at approximately 920° C. in two Probostat testing rigs (Norwegian Electroceramics) shown in
We controlled for microscopic gas leaks across the fuel cells by ensuring they have the same open-circuit potentials at the same temperatures. Measured open-circuit potentials were within 10 mV of the Nernstian limit. Over the course of the two-day experiment, we adjusted TH down as the open-circuit potential degraded by ˜3 mV to avoid artificially inflating the voltage and power density of the system. This resulted in a slight underestimation of the power density of the system.
The gas compositions supplied to the system were 5% H2 balance Ar, humidified at a room temperature of 18° C., versus dry 21% O2 in Ar, both at flow rates of approximately 80 sccm as measured by mass flow controllers (MKS) calibrated with Ar.
The J-V curve in
In one embodiment of the invention, V2+/3+∥Fe(CN)63−/4− liquid flow cells were constructed as shown in
A counterflow heat exchanger was constructed in a similar manner to the two flow cells, as shown in
All energy harvesting experiments were conducted in a N2-purged glove box (MTI VGB-4 with Instru-Tech Stinger pressure regulator, MTI O2 sensor and no H2O regulation). For these experiments, two of the liquid flow cells were connected in a single fluidic circuit as shown in
The effective power input to the energy harvesting system was estimated adding the thermodynamic heat input required by the electrode process I*TH (α1-α2) to the sensible heat leaked through the heat exchanger. This sensible heat leak was estimated by measuring the temperature increase of the electrolyte solutions as they traveled from the hot to cold cell, and multiplying the temperature increase ΔT of both solutions from the cold cell by the mass flow rate m and heat capacity cp of each electrolyte solution. Ideally, the power input into the system would be calculated as Pin=I*TH (α1-α2)+[({dot over (m)} cp)FCN(II/III)+({dot over (m)}cp)V(II/III)] (THot−TInlet), where TInlet is the measured temperature of both electrolyte solutions between the exits from the hot side of the heat exchanger and the inlets of the hot cell. However, temperature measurements at different points in the flow system indicated that the electrolyte circulation was slow enough that the electrolyte emerging from the cold cell nearly equilibrated with the glove box environmental temperature (˜28° C.) before entering the heat exchanger. Since this heat transfer from the environment constituted an additional energy input, it was considered improper to measure the energy input in this way. As a result, for the purposes of the efficiency calculation, the energy input was calculated conservatively as (I*VOC)+[({dot over (m)} cp)FCN(II/III)+({dot over (m)} cp)V(II/III)] (THot−TCold). This energy input is much less than the total energy input from the heater, as the heat leaks through the cell leads and other heat loss to the environment. However, since it is equivalent to the energy input required in the complete absence of the heat exchanger, it is likely an upper bound on the energy input that would be required in a scaled up, insulated system. The cp and density ρ values of both electrolytes were measured as described in the next section. The mass flow rate {dot over (m)} was based on the measured densities ρ and the volumetric flow rates Q through the pumps, which had previously been calibrated using a graduated cylinder.
The power output from the energy harvesting system was obtained by measuring the system's current-voltage curve with a potentiostat (Biologic SP-240). The resistance of the long (>2 m) leads and contacts between the energy harvesting system inside the glove box and the potentiostat in the laboratory was measured between 1 and 1.5Ω, but this resistance was not compensated in the electrical measurement or the reported polarization curves because it varied slightly each time the leads were connected to the cell. The efficiency of energy conversion η was then calculated as:
Here the subscript V denotes the V2+/3+ electrolyte, and FCN denotes the Fe(CN)63−/4− electrolyte. This yielded the reported values of η=0.042 (0.34 ηc) at 0.25 mA cm−2 and η=0.018 (0.15 ηc) at the maximum power point of 1.8 mA cm−2.
These efficiency values do not include the impact of the heat exchanger, out of consideration for the leak of heat into the system from the ambient discussed above. However, it is interesting to project the efficiency of a scaled-up system, in which the flow rate of electrolyte could be increased such that the heat loss to the ambient due to long residence times in the heat exchanger could be minimized. In this case, the effective efficiency of the energy harvesting system could be estimated as:
Here εHX (Q) denotes the efficiency of the heat exchanger as a function of flow rate Q, which is equivalent to the heat exchanger effectiveness in this case because the {dot over (m)} cp are matched in both electrolyte streams. For example, based on the performance of the heat exchanger given in Extended Data
The open-circuit voltage of the system is VOC=(α1-α2)(TH−TC)=αΔT
Voltage is solved as a function of current density: V(I)=V(local cO, cR)−2ηact−IRΩ
Here RΩ is the Ohmic resistance of both the cell and the solution together. The voltage is added in series for the cells in the stack. The open-circuit potential for one cell was taken as Nernstian, including concentration terms, and a temperature-dependent reference potential:
For the heat engine operating with two cells using the same redox couples, the non-equilibrium Nernst voltage simplifies to
Here, concentrations are squared to account for the two concentration ratios on the two sides of the membrane, and referenced to 1M. The symmetric nature of each cell, and only equal concentrations considered, warrant this simplification. The total temperature coefficient α of the system was used as a parameter. Notably, as the current density approaches the mass transport limit, the concentration term becomes large, and dominates the resulting voltage loss.
Activation overpotential is given by the Butler-Volmer equation:
For the liquid cell, we used a symmetry factor of 0.5, activation energy of 50 kJ/mol, and referenced the values of k0 to 1M concentrations at 273 K. Notably, the concentrations used are local at the electrode. The activation overpotential diverges as the current density approaches the limiting current density, and one of the concentrations approaches zero. For the gas cell, the exchange current density formalism was used, with reference values given below.
Ohmic resistance was taken as ⅓ of resistance values for 1M HBr, 6 and Nafion resistance 7 was used for a membrane of thickness 25 microns, independent of temperature. The thickness of the acid solution was taken as the minimum of hc and 0.15 mm. The conductivity of an acid solution was modeled to increase with temperature as diffusion is enhanced with decreasing viscosity of the fluid.
In calculating the local concentrations, a plug flow was assumed in the cell. The limiting current is given analytically with a Taylor series solution. The dimensionless measure of mass transport giving the maximum reagent utilization is calculated as:
Note that the Peclet number varies with temperature for a constant volumetric flow rate, due to the temperature enhancement of diffusion. The factor γ was calculated individually for the hot and cold cells. The Taylor series was evaluated to 30 terms, giving a compromise between underestimating the limiting current density and computational complexity.
Limiting current was calculated from the total inlet flux and the factor γ:
J
lim
=γh
c
w
c
c
O,R,inlet
F
The local concentrations at the electrodes are given as
depending on whether the species is consumed or produced at the electrodes. As the current density approaches the calculated limit, one of the local concentrations becomes fully depleted (even though the flow of the reagents to the cell may not be completely consumed). This affects both the Nernstian potential term, and the activation overpotential.
For connecting two cells in series, outlet reagent fluxes are calculated trivially at the first cell (hot cell in this simulation), and are used as inlet fluxes for the other cell. Since the cells are always current-matched and operating in reverse of each other, the inlet fluxes to the first cell are recovered from the outlet fluxes of the second. This assumes complete mixing of the electrolyte in between the cells, so that the concentrations of active species at the inlets of all cells are homogeneous.
The heat conductivity, specific heat, and dynamic viscosity for the solution in the liquid cells were assumed identical to water and taken from tables for liquid water at atmospheric pressure. For gases, respective temperature-dependent values were taken for O2, H2O, and H2.
For binary mixtures of gases, e.g. H2 and H2O, the specific heat and density were taken as linear combinations of the respective constituent values, while the heat conductivity and viscosity were recalculated 13,14 for the mixtures. For gases, the partial pressures were used as proxies for the composition fractions (x1, x2).
Laminar flow regime was used for the majority of calculations, and the assumption verified by checking the Reynolds number. The heat exchanger was assumed to have a counter-flow configuration with straight concentric circular pipes. Dimensionless quantities were calculated at the mean temperature between hot and cold cells for each working fluid in a circular pipe:
The average Nusselt numbers were calculated separately for the thermal entrance region and fully developed flows under the assumption of laminar flow. The length of the entrance region for establishing laminar flow is given as LER=0.06×Re×2 r. The Nusselt number was calculated for the entrance region using the Sieder and Tate correlation 15 modifying the traditional Graetz solution, neglecting the temperature dependence of viscosity:
The length LER varied widely and was in general not negligible compared to the simulated heat exchanger lengths (0.5-10 m). The Nusselt number for the fully developed laminar flow regions outside of the entrance lengths was taken as 48/11.
The convective heat transfer coefficient was calculated as h=κNu. This is equivalent to making 2 r the assumption that the convective “depletion” width is comparable to the radius of the pipe, which is reasonable for long pipes.
The heat exchanger is modeled as a counter-flow heat exchanger. The thermal resistance of the heat exchanger wall is given analytically:
In general, the heat conductivity of the exchanger is given as:
The five terms in the denominator correspond to heat transfer across the fluid layers, the fouling resistances, and across the pipe wall in each heat exchanger. UA was first taken as an input parameter, together with N, ri, and t, for the particular temperature and working fluids of the simulation. Assuming fully developed flows, LHX was calculated. UA was then re-calculated, accounting for entrance regions in the heat exchanger. For example, if the two entrance lengths were calculated to be 10% of LHX each, then the final UA was comprised of 80% the input value for fully developed flows, and 20% using equation (2.5) with one of the coefficients h re-calculated as above for an entrance region. When varying the input UA value parametrically, the parameter N was varied conjointly, so the total length LHX remained constant (
In the number of thermal units formalism, the heat exchanger efficiency is given as:
This expression is simplified for the constraint of matching heat flows in the two pipes of the heat exchanger, which was enforced in simulations. The conductive heat leak along the cross-sectional area of the walls of each of 2N heat exchanger tubes in the system is given as:
The head pressure in each annular tube is given analytically as:
The head pressure in each cell chamber is
The head pressure scales directly with the total area of cells, and independent of the number of cells in a stack of a given total area. Overpressures built up in the pipe junctions and bends were ignored. Note that the flow rate vvol in each tube or cell depends inversely on number of identical heat exchanger tubes N. Since fluid utilization rates were never close to unity at maximum power points, the performance of one heat exchanger was calculated, and then the result was doubled for the system. The total head pressure to be pumped is given as:
P
head=2(ΔPHX,i+ΔPHX,o+2ΔPcell)
For the liquid cell, the pumping was assumed to be mechanical:
For the gas cell, pumping was assumed to be electrical at 20% efficiency:
The operating pressure of the cells was taken as 1 atm. The density was calculated from STP values via the ideal gas law at the midpoint temperature of the system. The power dissipated to the resistance of the electrical leads is given as:
This term is the main origin of the scaling behavior of the system upon stacking. The resistance of mechanical components holding the stack together (i.e. bipolar plates) is ignored.
The power output of the system is
P
system
=IV−P
lead
−P
pump
The reversible entropy change for the electrochemical reaction at the hot side is:
ΔS=ΣSprod−ΣSreact
This has two components: the configurational concentration term, equivalent to the Nernstian concentration ratio, and the thermodynamic term. In the case of the liquid cell, the thermodynamic term is the total effective Seebeck coefficient α divided by the electron charge q.
Phenomenologically, the heat input to the system is given as for a thermoelectric with a heat exchanger:
{dot over (Q)}=αIT+2{dot over (Q)}HX+(1−εHX)({dot over (m)}cp)totalΔT−0.5Plead
The efficiency of the system is
equivalent to equation (1) in the main text.
For each set of design parameters (heat exchanger size, cell dimensions, stack size), and materials parameters (ohmic resistances, exchange current densities), the current density was swept to find the maximum power density. For the liquid system, the circulation flow rate was also left free during the optimization via the Pe number.
Constants and Parameters—Gas Cells
Total area 1 m2, 100 cells, each 10 cm long and 10 cm wide, with chamber height 1 cm. For the redox couples, a mixture of 10% H2 and 90% H2Oversus 21% O2 were used, for a thermopower of −0.42 mV/K. Using this mixture of gases in our experiment would have increased the power densities in
Electrolyte ohmic resistance: modeled as doped ceria with area-specific resistance (ASR) 0.1 Ωcm2 at 500° C., and activation energy 57 kJ/mol. Additionally, a 100 nm layer of YSZ was modeled.
Activation overpotentials: hot cell at TH with j00=500 mA cm−2 at 700° C. and activation energy 100 kJ/mol for the cathode and the anode. For the cold cell at TC, j00=150 mA cm−2 at 500° C., and activation energy 96.65 kJ/mol. Reference pressures pH2=0.97 atm, pH2O=0.03 atm, and pO2=0.21 atm were used, with unity pressure dependences for the anode, and square-root pressure dependences for the cathode.
Heat exchangers were modeled with silica heat conductivity, number of tubes N=50, each with wall thickness 2 mm, inner radius 2 cm, and outer radius 4 cm, and conductivity for fully developed flows UA=20 W K−1. Leads were modeled as molybdenum, radius 1 cm, and with the same length as the heat exchanger.
Total area 1 m2, 100 cells, each 10 cm long and 10 cm wide, with chamber height 0.2 mm. All thermohydraulic parameters were taken as for liquid water. The diffusion coefficient of active species in the fluid at room temperature was taken as D=10−5 cm2 sec−1. Heat exchangers were modeled with titanium heat conductivity, conductivity for fully developed flows UA=400 kW K−1, wall thickness 0.25 mm, inner radius 0.25 cm, and outer radius 0.5 cm, with a number of tubes N=10000. Leads were modeled as molybdenum, with cross-section area 50 mm2, and with the same length as the heat exchanger.
This application claims priority from U.S. Provisional Patent Application 62/314,856 filed Mar. 29, 2016, which is incorporated herein by reference.
Number | Date | Country | |
---|---|---|---|
62314856 | Mar 2016 | US |