Solid oxide fuel cell

Last updated

Scheme of a solid-oxide fuel cell Solid oxide fuel cell.svg
Scheme of a solid-oxide fuel cell

A solid oxide fuel cell (or SOFC) is an electrochemical conversion device that produces electricity directly from oxidizing a fuel. Fuel cells are characterized by their electrolyte material; the SOFC has a solid oxide or ceramic electrolyte.

Contents

Advantages of this class of fuel cells include high combined heat and power efficiency, long-term stability, fuel flexibility, low emissions, and relatively low cost. The largest disadvantage is the high operating temperature which results in longer start-up times and mechanical and chemical compatibility issues. [1]

Introduction

Solid oxide fuel cells are a class of fuel cells characterized by the use of a solid oxide material as the electrolyte. SOFCs use a solid oxide electrolyte to conduct negative oxygen ions from the cathode to the anode. The electrochemical oxidation of the hydrogen, carbon monoxide or other organic intermediates by oxygen ions thus occurs on the anode side. [2] [3] More recently, proton-conducting SOFCs (PC-SOFC) are being developed which transport protons instead of oxygen ions through the electrolyte with the advantage of being able to be run at lower temperatures than traditional SOFCs. [4] [5]

They operate at very high temperatures, typically between 600 and 1,000 °C. [2] [3] At these temperatures, SOFCs do not require expensive platinum group metals catalyst, [3] [6] as is currently necessary for lower temperature fuel cells such as PEMFCs, and are not vulnerable to carbon monoxide catalyst poisoning. However, vulnerability to sulfur poisoning [7] [3] [8] has been widely observed and the sulfur must be removed before entering the cell. For fuels that are of lower quality, such as gasified biomass, coal, or biogas, the fuel processing becomes increasingly complex and, consequently, more expensive. The gasification process, which transforms the raw material into a gaseous state suitable for fuel cells, can generate significant quantities of aromatic compounds. These compounds include smaller molecules like methane and toluene, as well as larger polyaromatic and short-chain hydrocarbon compounds. These substances can lead to carbon buildup in SOFCs. Moreover, the expenses associated with reforming and desulfurization are comparable in magnitude to the cost of the fuel cell itself. These factors become especially critical for systems with lower power output or greater portability requirements. [9]

Solid oxide fuel cells have a wide variety of applications, from use as auxiliary power units in vehicles to stationary power generation with outputs from 100 W to 2 MW. In 2009, Australian company, Ceramic Fuel Cells successfully achieved an efficiency of an SOFC device up to the previously theoretical mark of 60%. [10] [11] The higher operating temperature make SOFCs suitable candidates for application with heat engine energy recovery devices or combined heat and power, which further increases overall fuel efficiency. [12]

Because of these high temperatures, light hydrocarbon fuels, such as methane, propane, and butane can be internally reformed within the anode. [13] SOFCs can also be fueled by externally reforming heavier hydrocarbons, such as gasoline, diesel, jet fuel (JP-8) or biofuels. Such reformates are mixtures of hydrogen, carbon monoxide, carbon dioxide, steam and methane, formed by reacting the hydrocarbon fuels with air or steam in a device upstream of the SOFC anode. SOFC power systems can increase efficiency by using the heat given off by the exothermic electrochemical oxidation within the fuel cell for endothermic steam reforming process. Additionally, solid fuels such as coal and biomass may be gasified to form syngas which is suitable for fueling SOFCs in integrated gasification fuel cell power cycles.

Thermal expansion demands a uniform and well-regulated heating process at startup. SOFC stacks with planar geometry require on the order of an hour to be heated to operating temperature. Micro-tubular fuel cell design [14] [15] geometries promise much faster start up times, typically in the order of minutes.

Unlike most other types of fuel cells, SOFCs can have multiple geometries. The planar fuel cell design geometry is the typical sandwich type geometry employed by most types of fuel cells, where the electrolyte is sandwiched in between the electrodes. SOFCs can also be made in tubular geometries where either air or fuel is passed through the inside of the tube and the other gas is passed along the outside of the tube. The tubular design is advantageous because it is much easier to seal air from the fuel. The performance of the planar design is currently better than the performance of the tubular design, however, because the planar design has a lower resistance comparatively. Other geometries of SOFCs include modified planar fuel cell designs (MPC or MPSOFC), where a wave-like structure replaces the traditional flat configuration of the planar cell. Such designs are highly promising because they share the advantages of both planar cells (low resistance) and tubular cells.[ citation needed ]

Operation

Cross section of three ceramic layers of a tubular SOFC. From inner to outer: porous cathode, dense electrolyte, porous anode SOFC-en.svg
Cross section of three ceramic layers of a tubular SOFC. From inner to outer: porous cathode, dense electrolyte, porous anode

A solid oxide fuel cell is made up of four layers, three of which are ceramics (hence the name). A single cell consisting of these four layers stacked together is typically only a few millimeters thick. Hundreds of these cells are then connected in series to form what most people refer to as an "SOFC stack". The ceramics used in SOFCs do not become electrically and ionically active until they reach very high temperature and as a consequence, the stacks have to run at temperatures ranging from 500 to 1,000 °C. Reduction of oxygen into oxygen ions occurs at the cathode. These ions can then diffuse through the solid oxide electrolyte to the anode where they can electrochemically oxidize the fuel. In this reaction, a water byproduct is given off as well as two electrons. These electrons then flow through an external circuit where they can do work. The cycle then repeats as those electrons enter the cathode material again.

Balance of plant

Most of the downtime of a SOFC stems from the mechanical balance of plant, the air preheater, prereformer, afterburner, water heat exchanger, anode tail gas oxidizer, and electrical balance of plant, power electronics, hydrogen sulfide sensor and fans. Internal reforming leads to a large decrease in the balance of plant costs in designing a full system. [11]

Anode

The ceramic anode layer must be very porous to allow the fuel to flow towards the electrolyte. Consequently, granular matter is often selected for anode fabrication procedures. [16] Like the cathode, it must conduct electrons, with ionic conductivity a definite asset. The anode is commonly the thickest and strongest layer in each individual cell, because it has the smallest polarization losses, and is often the layer that provides the mechanical support. Electrochemically speaking, the anode's job is to use the oxygen ions that diffuse through the electrolyte to oxidize the hydrogen fuel. The oxidation reaction between the oxygen ions and the hydrogen produces heat as well as water and electricity. If the fuel is a light hydrocarbon, for example, methane, another function of the anode is to act as a catalyst for steam reforming the fuel into hydrogen. This provides another operational benefit to the fuel cell stack because the reforming reaction is endothermic, which cools the stack internally. The most common material used is a cermet made up of nickel mixed with the ceramic material that is used for the electrolyte in that particular cell, typically YSZ (yttria stabilized zirconia). These nanomaterial-based catalysts, help stop the grain growth of nickel. Larger grains of nickel would reduce the contact area that ions can be conducted through, which would lower the cells efficiency. Perovskite materials (mixed ionic/electronic conducting ceramics) have been shown to produce a power density of 0.6 W/cm2 at 0.7 V at 800 °C which is possible because they have the ability to overcome a larger activation energy. [17]

Chemical Reaction:

H2 +O2- ——> H2O+2e

However, there are a few disadvantages associated with YSZ as anode material. Ni coarsening, carbon deposition, reduction-oxidation instability, and sulfur poisoning are the main obstacles limiting the long-term stability of Ni-YSZ. Ni coarsening refers to the evolution of Ni particles in doped in YSZ grows larger in grain size, which decreases the surface area for the catalytic reaction. Carbon deposition occurs when carbon atoms, formed by hydrocarbon pyrolysis or CO disproportionation, deposit on the Ni catalytic surface. [18] Carbon deposition becomes important especially when hydrocarbon fuels are used, i.e. methane, syngas. The high operating temperature of SOFC and the oxidizing environment facilitate the oxidation of Ni catalyst through reaction Ni + 12 O2 = NiO. The oxidation reaction of Ni reduces the electrocatalytic activity and conductivity. Moreover, the density difference between Ni and NiO causes volume change on the anode surface, which could potentially lead to mechanical failure. Sulfur poisoning arises when fuel such as natural gas, gasoline, or diesel is used. Again, due to the high affinity between sulfur compounds (H2S, (CH3)2S) and the metal catalyst, even the smallest impurities of sulfur compounds in the feed stream could deactivate the Ni catalyst on the YSZ surface. [19]

Current research is focused on reducing or replacing Ni content in the anode to improve long-term performance. The modified Ni-YSZ containing other materials including CeO2, Y2O3, La2O3, MgO, TiO2, Ru, Co, etc. are invented to resist sulfur poisoning, but the improvement is limited due to the rapid initial degradation. [20] Copper-based cerement anode is considered as a solution to carbon deposition because it is inert to carbon and stable under typical SOFC oxygen partial pressures (pO2). Cu-Co bimetallic anodes in particular show a great resistivity of carbon deposition after the exposure to pure CH4 at 800C. [21] And Cu-CeO2-YSZ exhibits a higher electrochemical oxidation rate over Ni-YSZ when running on CO and syngas, and can achieve even higher performance using CO than H2, after adding a cobalt co-catalyst. [22] Oxide anodes including zirconia-based fluorite and perovskites are also used to replace Ni-ceramic anodes for carbon resistance. Chromite i.e. La0.8Sr0.2Cr0.5Mn0.5O3 (LSCM) is used as anodes and exhibited comparable performance against Ni–YSZ cermet anodes. LSCM is further improved by impregnating Cu and sputtering Pt as the current collector. [21]

Electrolyte

The electrolyte is a dense layer of ceramic that conducts oxygen ions. Its electronic conductivity must be kept as low as possible to prevent losses from leakage currents. The high operating temperatures of SOFCs allow the kinetics of oxygen ion transport to be sufficient for good performance. However, as the operating temperature approaches the lower limit for SOFCs at around 600 °C, the electrolyte begins to have large ionic transport resistances and affect the performance. Popular electrolyte materials include yttria-stabilized zirconia (YSZ) (often the 8% form 8YSZ), scandia stabilized zirconia (ScSZ) (usually 9 mol% Sc2O3 – 9ScSZ) and gadolinium doped ceria (GDC). [23] The electrolyte material has crucial influence on the cell performances. [24] Detrimental reactions between YSZ electrolytes and modern cathodes such as lanthanum strontium cobalt ferrite (LSCF) have been found, and can be prevented by thin (<100 nm) ceria diffusion barriers. [25]

If the conductivity for oxygen ions in SOFC can remain high even at lower temperatures (current target in research ~500 °C), material choices for SOFC will broaden and many existing problems can potentially be solved. Certain processing techniques such as thin film deposition [26] can help solve this problem with existing materials by:

Cathode

The cathode, or air electrode, is a thin porous layer on the electrolyte where oxygen reduction takes place. The overall reaction is written in Kröger-Vink Notation as follows:

Cathode materials must be, at a minimum, electrically conductive. Currently, lanthanum strontium manganite (LSM) is the cathode material of choice for commercial use because of its compatibility with doped zirconia electrolytes. Mechanically, it has a similar coefficient of thermal expansion to YSZ and thus limits stress buildup because of CTE mismatch. Also, LSM has low levels of chemical reactivity with YSZ which extends the lifetime of the materials. Unfortunately, LSM is a poor ionic conductor, and so the electrochemically active reaction is limited to the triple phase boundary (TPB) where the electrolyte, air and electrode meet. LSM works well as a cathode at high temperatures, but its performance quickly falls as the operating temperature is lowered below 800 °C. In order to increase the reaction zone beyond the TPB, a potential cathode material must be able to conduct both electrons and oxygen ions. Composite cathodes consisting of LSM YSZ have been used to increase this triple phase boundary length. Mixed ionic/electronic conducting (MIEC) ceramics, such as perovskite LSCF, are also being researched for use in intermediate temperature SOFCs as they are more active and can make up for the increase in the activation energy of the reaction. [27]

Interconnect

The interconnect can be either a metallic or ceramic layer that sits between each individual cell. Its purpose is to connect each cell in series, so that the electricity each cell generates can be combined. Because the interconnect is exposed to both the oxidizing and reducing side of the cell at high temperatures, it must be extremely stable. For this reason, ceramics have been more successful in the long term than metals as interconnect materials. However, these ceramic interconnect materials are very expensive when compared to metals. Nickel- and steel-based alloys are becoming more promising as lower temperature (600–800 °C) SOFCs are developed. The material of choice for an interconnect in contact with Y8SZ is a metallic 95Cr-5Fe alloy. Ceramic-metal composites called "cermet" are also under consideration, as they have demonstrated thermal stability at high temperatures and excellent electrical conductivity.

Polarizations

Polarizations, or overpotentials, are losses in voltage due to imperfections in materials, microstructure, and design of the fuel cell. Polarizations result from ohmic resistance of oxygen ions conducting through the electrolyte (iRΩ), electrochemical activation barriers at the anode and cathode, and finally concentration polarizations due to inability of gases to diffuse at high rates through the porous anode and cathode (shown as ηA for the anode and ηC for cathode). [28] The cell voltage can be calculated using the following equation:

where:

In SOFCs, it is often important to focus on the ohmic and concentration polarizations since high operating temperatures experience little activation polarization. However, as the lower limit of SOFC operating temperature is approached (~600 °C), these polarizations do become important. [29]

Above mentioned equation is used for determining the SOFC voltage (in fact for fuel cell voltage in general). This approach results in good agreement with particular experimental data (for which adequate factors were obtained) and poor agreement for other than original experimental working parameters. Moreover, most of the equations used require the addition of numerous factors which are difficult or impossible to determine. It makes very difficult any optimizing process of the SOFC working parameters as well as design architecture configuration selection. Because of those circumstances a few other equations were proposed: [30]

where:

This method was validated and found to be suitable for optimization and sensitivity studies in plant-level modelling of various systems with solid oxide fuel cells. [32] With this mathematical description it is possible to account for different properties of the SOFC. There are many parameters which impact cell working conditions, e.g. electrolyte material, electrolyte thickness, cell temperature, inlet and outlet gas compositions at anode and cathode, and electrode porosity, just to name some. The flow in these systems is often calculated using the Navier–Stokes equations.

Ohmic polarization

Ohmic losses in an SOFC result from ionic conductivity through the electrolyte and electrical resistance offered to the flow of electrons in the external electrical circuit. This is inherently a materials property of the crystal structure and atoms involved. However, to maximize the ionic conductivity, several methods can be done. Firstly, operating at higher temperatures can significantly decrease these ohmic losses. Substitutional doping methods to further refine the crystal structure and control defect concentrations can also play a significant role in increasing the conductivity. Another way to decrease ohmic resistance is to decrease the thickness of the electrolyte layer.

Ionic conductivity

An ionic specific resistance of the electrolyte as a function of temperature can be described by the following relationship: [30]

where: – electrolyte thickness, and – ionic conductivity.

The ionic conductivity of the solid oxide is defined as follows: [30]

where: and – factors depended on electrolyte materials, – electrolyte temperature, and – ideal gas constant.

Concentration polarization

The concentration polarization is the result of practical limitations on mass transport within the cell and represents the voltage loss due to spatial variations in reactant concentration at the chemically active sites. This situation can be caused when the reactants are consumed by the electrochemical reaction faster than they can diffuse into the porous electrode, and can also be caused by variation in bulk flow composition. The latter is due to the fact that the consumption of reacting species in the reactant flows causes a drop in reactant concentration as it travels along the cell, which causes a drop in the local potential near the tail end of the cell.

The concentration polarization occurs in both the anode and cathode. The anode can be particularly problematic, as the oxidation of the hydrogen produces steam, which further dilutes the fuel stream as it travels along the length of the cell. This polarization can be mitigated by reducing the reactant utilization fraction or increasing the electrode porosity, but these approaches each have significant design trade-offs.

Activation polarization

The activation polarization is the result of the kinetics involved with the electrochemical reactions. Each reaction has a certain activation barrier that must be overcome in order to proceed and this barrier leads to the polarization. The activation barrier is the result of many complex electrochemical reaction steps where typically the rate limiting step is responsible for the polarization. The polarization equation shown below is found by solving the Butler–Volmer equation in the high current density regime (where the cell typically operates), and can be used to estimate the activation polarization:

where:

The polarization can be modified by microstructural optimization. The Triple Phase Boundary (TPB) length, which is the length where porous, ionic and electronically conducting pathways all meet, directly relates to the electrochemically active length in the cell. The larger the length, the more reactions can occur and thus the less the activation polarization. Optimization of TPB length can be done by processing conditions to affect microstructure or by materials selection to use a mixed ionic/electronic conductor to further increase TPB length.

Mechanical Properties

Current SOFC research focuses heavily on optimizing cell performance while maintaining acceptable mechanical properties because optimized performance often compromises mechanical properties. Nevertheless, mechanical failure represents a significant problem to SOFC operation. The presence of various kinds of load and Thermal stress during operation requires high mechanical strength. Additional stresses associated with changes in gas atmosphere, leading to reduction or oxidation also cannot be avoided in prolonged operation. [33] When electrode layers delaminate or crack, conduction pathways are lost, leading to a redistribution of current density and local changes in temperature. These local temperature deviations, in turn, lead to increased thermal strains, which propagate cracks and Delamination. Additionally, when electrolytes crack, separation of fuel and air is no longer guaranteed, which further endangers the continuous operation of the cell. [34]

Since SOFCs require materials with high oxygen conductivity, thermal stresses provide a significant problem. The Coefficient of thermal expansion in mixed ionic-electronic perovskites can be directly related to oxygen vacancy concentration, which is also related to ionic conductivity. [35] Thus, thermal stresses increase in direct correlation with improved cell performance. Additionally, however, the temperature dependence of oxygen vacancy concentration means that the CTE is not a linear property, which further complicates measurements and predictions.

Just as thermal stresses increase as cell performance improves through improved ionic conductivity, the fracture toughness of the material also decreases as cell performance increases. This is because, to increase reaction sites, porous ceramics are preferable. However, as shown in the equation below, fracture toughness decreases as porosity increases. [36]

Where:

= fracture toughness

= fracture toughness of the non-porous structure

= experimentally determined constant

= porosity

Thus, porosity must be carefully engineered to maximize reaction kinetics while maintaining an acceptable fracture toughness. Since fracture toughness represents the ability of pre-existing cracks or pores to propagate, a potentially more useful metric is the failure stress of a material, as this depends on sample dimensions instead of crack diameter. Failure stresses in SOFCs can also be evaluated using a ring-on ring biaxial stress test. This type of test is generally preferred, as sample edge quality does not significantly impact measurements. The determination of the sample's failure stress is shown in the equation below. [37]

Where:

= failure stress of the small deformation

= critical applied force

= height of the sample

= Poisson's ratio

= diameter (sup = support ring, load = loading ring, s = sample)

However, this equation is not valid for deflections exceeding 1/2h, [38] making it less applicable for thin samples, which are of great interest in SOFCs. Therefore, while this method does not require knowledge of crack or pore size, it must be used with great caution and is more applicable to support layers in SOFCs than active layers. In addition to failure stresses and fracture toughness, modern fuel cell designs that favor mixed ionic electronic conductors (MIECs), Creep (deformation) pose another great problem, as MIEC electrodes often operate at temperatures exceeding half of the melting temperature. As a result, diffusion creep must also be considered. [39]

Where:

= equivalent creep strain

= Diffusion coefficient

= temperature

= kinetic constant

= equivalent stress (e.g. von Mises)

= creep stress exponential factor

= particle size exponent (2 for Nabarro–Herring creep, 3 for Coble creep)

To properly model creep strain rates, knowledge of Microstructure is therefore of significant importance. Due to the difficulty in mechanically testing SOFCs at high temperatures, and due to the microstructural evolution of SOFCs over the lifetime of operation resulting from Grain growth and coarsening, the actual creep behavior of SOFCs is currently not completely understood

Target

DOE target requirements are 40,000 hours of service for stationary fuel cell applications and greater than 5,000 hours for transportation systems (fuel cell vehicles) at a factory cost of $40/kW for a 10 kW coal-based system [40] without additional requirements. Lifetime effects (phase stability, thermal expansion compatibility, element migration, conductivity and aging) must be addressed. The Solid State Energy Conversion Alliance 2008 (interim) target for overall degradation per 1,000 hours is 4.0%. [41]

Research

Research is going now in the direction of lower-temperature SOFCs (600 °C). Low temperature systems can reduce costs by reducing insulation, materials, start-up and degradation-related costs. With higher operating temperatures, the temperature gradient increases the severity of thermal stresses, which affects materials cost and life of the system. [42] An intermediate temperature system (650-800 °C) would enable the use of cheaper metallic materials with better mechanical properties and thermal conductivity. New developments in nano-scale electrolyte structures have been shown to bring down operating temperatures to around 350 °C, which would enable the use of even cheaper steel and elastomeric/polymeric components. [43]

Lowering operating temperatures has the added benefit of increased efficiency. Theoretical fuel cell efficiency increases with decreasing temperature. For example, the efficiency of a SOFC using CO as fuel increases from 63% to 81% when decreasing the system temperature from 900 °C to 350 °C. [43]

Research is also under way to improve the fuel flexibility of SOFCs. While stable operation has been achieved on a variety of hydrocarbon fuels, these cells typically rely on external fuel processing. In the case of natural gas, the fuel is either externally or internally reformed and the sulfur compounds are removed. These processes add to the cost and complexity of SOFC systems. Work is under way at a number of institutions to improve the stability of anode materials for hydrocarbon oxidation and, therefore, relax the requirements for fuel processing and decrease SOFC balance of plant costs.

Research is also going on in reducing start-up time to be able to implement SOFCs in mobile applications. [44] This can be partially achieved by lowering operating temperatures, which is the case for proton-exchange membrane fuel cell (PEMFCs). [45] Due to their fuel flexibility, they may run on partially reformed diesel, and this makes SOFCs interesting as auxiliary power units (APU) in refrigerated trucks.

Specifically, Delphi Automotive Systems are developing an SOFC that will power auxiliary units in automobiles and tractor-trailers, while BMW has recently stopped a similar project. A high-temperature SOFC will generate all of the needed electricity to allow the engine to be smaller and more efficient. The SOFC would run on the same gasoline or diesel as the engine and would keep the air conditioning unit and other necessary electrical systems running while the engine shuts off when not needed (e.g., at a stop light or truck stop). [46]

Rolls-Royce is developing solid-oxide fuel cells produced by screen printing onto inexpensive ceramic materials. [47] Rolls-Royce Fuel Cell Systems Ltd is developing an SOFC gas turbine hybrid system fueled by natural gas for power generation applications in the order of a megawatt (e.g. Futuregen).[ citation needed ]

3D printing is being explored as a possible manufacturing technique that could be used to make SOFC manufacturing easier by the Shah Lab at Northwestern University. This manufacturing technique would allow SOFC cell structure to be more flexible, which could lead to more efficient designs. This process could work in the production of any part of the cell. The 3D printing process works by combining about 80% ceramic particles with 20% binders and solvents, and then converting that slurry into an ink that can be fed into a 3D printer. Some of the solvent is very volatile, so the ceramic ink solidifies almost immediately. Not all of the solvent evaporates, so the ink maintains some flexibility before it is fired at high temperature to densify it. This flexibility allows the cells to be fired in a circular shape that would increase the surface area over which electrochemical reactions can occur, which increases the efficiency of the cell. Also, the 3D printing technique allows the cell layers to be printed on top of each other instead of having to go through separate manufacturing and stacking steps. The thickness is easy to control, and layers can be made in the exact size and shape that is needed, so waste is minimized. [48]

Ceres Power Ltd. has developed a low cost and low temperature (500–600 degrees) SOFC stack using cerium gadolinium oxide (CGO) in place of current industry standard ceramic, yttria stabilized zirconia (YSZ), which allows the use of stainless steel to support the ceramic. [49]

Solid Cell Inc. has developed a unique, low-cost cell architecture that combines properties of planar and tubular designs, along with a Cr-free cermet interconnect.[ citation needed ]

The high temperature electrochemistry center (HITEC) at the University of Florida, Gainesville is focused on studying ionic transport, electrocatalytic phenomena and microstructural characterization of ion conducting materials. [50]

SiEnergy Systems, a Harvard spin-off company, has demonstrated the first macro-scale thin-film solid-oxide fuel cell that can operate at 500 degrees. [51]

SOEC

A solid oxide electrolyser cell (SOEC) is a solid oxide fuel cell set in regenerative mode for the electrolysis of water with a solid oxide, or ceramic, electrolyte to produce oxygen and hydrogen gas. [52]

SOECs can also be used to do electrolysis of CO2 to produce CO and oxygen [53] or even co-electrolysis of water and CO2 to produce syngas and oxygen.

ITSOFC

SOFCs that operate in an intermediate temperature (IT) range, meaning between 600 and 800 °C, are named ITSOFCs. Because of the high degradation rates and materials costs incurred at temperatures in excess of 900 °C, it is economically more favorable to operate SOFCs at lower temperatures. The push for high-performance ITSOFCs is currently the topic of much research and development. One area of focus is the cathode material. It is thought that the oxygen reduction reaction is responsible for much of the loss in performance so the catalytic activity of the cathode is being studied and enhanced through various techniques, including catalyst impregnation. The research on NdCrO3 proves it to be a potential cathode material for the cathode of ITSOFC since it is thermochemically stable within the temperature range. [54]

Another area of focus is electrolyte materials. To make SOFCs competitive in the market, ITSOFCs are pushing towards lower operational temperature by use of alternative new materials. However, efficiency and stability of the materials limit their feasibility. One choice for the electrolyte new materials is the ceria-salt ceramic composites (CSCs). The two-phase CSC electrolytes GDC (gadolinium-doped ceria) and SDC (samaria-doped ceria)-MCO3 (M=Li, Na, K, single or mixture of carbonates) can reach the power density of 300-800 mW*cm−2. [55]

LT-SOFC

Low-temperature solid oxide fuel cells (LT-SOFCs), operating lower than 650 °C, are of great interest for future research because the high operating temperature is currently what restricts the development and deployment of SOFCs. A low-temperature SOFC is more reliable due to smaller thermal mismatch and easier sealing. Additionally, a lower temperature requires less insulation and therefore has a lower cost. Cost is further lowered due to wider material choices for interconnects and compressive nonglass/ceramic seals. Perhaps most importantly, at a lower temperature, SOFCs can be started more rapidly and with less energy, which lends itself to uses in portable and transportable applications.[ citation needed ]

As temperature decreases, the maximum theoretical fuel cell efficiency increases, in contrast to the Carnot cycle. For example, the maximum theoretical efficiency of an SOFC using CO as a fuel increases from 63% at 900 °C to 81% at 350 °C. [56]

This is a materials issue, particularly for the electrolyte in the SOFC. YSZ is the most commonly used electrolyte because of its superior stability, despite not having the highest conductivity. Currently, the thickness of YSZ electrolytes is a minimum of ~10 μm due to deposition methods, and this requires a temperature above 700 °C. Therefore, low-temperature SOFCs are only possible with higher conductivity electrolytes. Various alternatives that could be successful at low temperature include gadolinium-doped ceria (GDC) and erbia-cation-stabilized bismuth (ERB). They have superior ionic conductivity at lower temperatures, but this comes at the expense of lower thermodynamic stability. CeO2 electrolytes become electronically conductive and Bi2O3 electrolytes decompose to metallic Bi under the reducing fuel environment. [57]

To combat this, researchers created a functionally graded ceria/bismuth-oxide bilayered electrolyte where the GDC layer on the anode side protects the ESB layer from decomposing while the ESB on the cathode side blocks the leakage current through the GDC layer. This leads to near-theoretical open-circuit potential (OPC) with two highly conductive electrolytes, that by themselves would not have been sufficiently stable for the application. This bilayer proved to be stable for 1400 hours of testing at 500 °C and showed no indication of interfacial phase formation or thermal mismatch. While this makes strides towards lowering the operating temperature of SOFCs, it also opens doors for future research to try and understand this mechanism. [58]

Comparison of ionic conductivity of various solid oxide electrolytes Ion Conductivity.png
Comparison of ionic conductivity of various solid oxide electrolytes

Researchers at the Georgia Institute of Technology dealt with the instability of BaCeO3 differently. They replaced a desired fraction of Ce in BaCeO3 with Zr to form a solid solution that exhibits proton conductivity, but also chemical and thermal stability over the range of conditions relevant to fuel cell operation. A new specific composition, Ba(Zr0.1Ce0.7Y0.2)O3-δ (BZCY7) that displays the highest ionic conductivity of all known electrolyte materials for SOFC applications. This electrolyte was fabricated by dry-pressing powders, which allowed for the production of crack free films thinner than 15 μm. The implementation of this simple and cost-effective fabrication method may enable significant cost reductions in SOFC fabrication. [59] However, this electrolyte operates at higher temperatures than the bilayered electrolyte model, closer to 600 °C rather than 500 °C.

Currently, given the state of the field for LT-SOFCs, progress in the electrolyte would reap the most benefits, but research into potential anode and cathode materials would also lead to useful results, and has started to be discussed more frequently in literature.

SOFC-GT

An SOFC-GT system is one which comprises a solid oxide fuel cell combined with a gas turbine. Such systems have been evaluated by Siemens Westinghouse and Rolls-Royce as a means to achieve higher operating efficiencies by running the SOFC under pressure. SOFC-GT systems typically include anodic and/or cathodic atmosphere recirculation, thus increasing efficiency.[ citation needed ]

Theoretically, the combination of the SOFC and gas turbine can give result in high overall (electrical and thermal) efficiency. [60] Further combination of the SOFC-GT in a combined cooling, heat and power (or trigeneration) configuration (via HVAC) also has the potential to yield even higher thermal efficiencies in some cases. [61]

Another feature of the introduced hybrid system is on the gain of 100% CO2 capturing at comparable high energy efficiency. These features like zero CO2 emission and high energy efficiency make the power plant performance noteworthy. [62]

DCFC

For the direct use of solid coal fuel without additional gasification and reforming processes, a direct carbon fuel cell (DCFC) has been developed as a promising novel concept of a high-temperature energy conversion system. The underlying progress in the development of a coal-based DCFC has been categorized mainly according to the electrolyte materials used, such as solid oxide, molten carbonate, and molten hydroxide, as well as hybrid systems consisting of solid oxide and molten carbonate binary electrolyte or of liquid anode (Fe, Ag, In, Sn, Sb, Pb, Bi, and its alloying and its metal/metal oxide) solid oxide electrolyte. [63] People's research on DCFC with GDC-Li/Na2CO3 as the electrolyte, Sm0.5Sr0.5CoO3 as cathode shows good performance. The highest power density of 48 mW*cm−2 can be reached at 500 °C with O2 and CO2 as oxidant and the whole system is stable within the temperature range of 500 °C to 600 °C. [64]

SOFC operated on landfill gas

Every household produces waste/garbage on a daily basis. In 2009, Americans produced about 243 million tons of municipal solid waste, which is 4.3 pounds of waste per person per day. All that waste is sent to landfill sites. Landfill gas which is produced from the decomposition of waste that gets accumulated at the landfills has the potential to be a valuable source of energy since methane is a major constituent. Currently, the majority of the landfills either burn away their gas in flares or combust it in mechanical engines to produce electricity. The issue with mechanical engines is that incomplete combustion of gasses can lead to pollution of the atmosphere and is also highly inefficient.[ citation needed ]

The issue with using landfill gas to fuel an SOFC system is that landfill gas contains hydrogen sulfide. Any landfill accepting biological waste will contain about 50-60 ppm of hydrogen sulfide and around 1-2 ppm mercaptans. However, construction materials containing reducible sulfur species, principally sulfates found in gypsum-based wallboard, can cause considerably higher levels of sulfides in the hundreds of ppm. At operating temperatures of 750 °C hydrogen sulfide concentrations of around 0.05 ppm begin to affect the performance of the SOFCs.[ citation needed ]

Ni + H2S → NiS + H2

The above reaction controls the effect of sulfur on the anode.

This can be prevented by having background hydrogen which is calculated below.

At 453 K the equilibrium constant is 7.39 x 10−5

ΔG calculated at 453 K was 35.833 kJ/mol

Using the standard heat of formation and entropy ΔG at room temperature (298 K) came out to be 45.904 kJ/mol

On extrapolation to 1023 K, ΔG is -1.229 kJ/mol

On substitution, Keq at 1023 K is 1.44 x 10−4. Hence theoretically we need 3.4% hydrogen to prevent the formation of NiS at 5 ppm H2S. [65]

See also

Related Research Articles

<span class="mw-page-title-main">Fuel cell</span> Device that converts the chemical energy from a fuel into electricity

A fuel cell is an electrochemical cell that converts the chemical energy of a fuel and an oxidizing agent into electricity through a pair of redox reactions. Fuel cells are different from most batteries in requiring a continuous source of fuel and oxygen to sustain the chemical reaction, whereas in a battery the chemical energy usually comes from substances that are already present in the battery. Fuel cells can produce electricity continuously for as long as fuel and oxygen are supplied.

<span class="mw-page-title-main">Strontium titanate</span> Chemical compound

Strontium titanate is an oxide of strontium and titanium with the chemical formula SrTiO3. At room temperature, it is a centrosymmetric paraelectric material with a perovskite structure. At low temperatures it approaches a ferroelectric phase transition with a very large dielectric constant ~104 but remains paraelectric down to the lowest temperatures measured as a result of quantum fluctuations, making it a quantum paraelectric. It was long thought to be a wholly artificial material, until 1982 when its natural counterpart—discovered in Siberia and named tausonite—was recognised by the IMA. Tausonite remains an extremely rare mineral in nature, occurring as very tiny crystals. Its most important application has been in its synthesized form wherein it is occasionally encountered as a diamond simulant, in precision optics, in varistors, and in advanced ceramics.

A regenerative fuel cell or reverse fuel cell (RFC) is a fuel cell run in reverse mode, which consumes electricity and chemical B to produce chemical A. By definition, the process of any fuel cell could be reversed. However, a given device is usually optimized for operating in one mode and may not be built in such a way that it can be operated backwards. Standard fuel cells operated backwards generally do not make very efficient systems unless they are purpose-built to do so as with high-pressure electrolysers, regenerative fuel cells, solid-oxide electrolyser cells and unitized regenerative fuel cells.

<span class="mw-page-title-main">Proton-exchange membrane fuel cell</span> Power generation technology

Proton-exchange membrane fuel cells (PEMFC), also known as polymer electrolyte membrane (PEM) fuel cells, are a type of fuel cell being developed mainly for transport applications, as well as for stationary fuel-cell applications and portable fuel-cell applications. Their distinguishing features include lower temperature/pressure ranges and a special proton-conducting polymer electrolyte membrane. PEMFCs generate electricity and operate on the opposite principle to PEM electrolysis, which consumes electricity. They are a leading candidate to replace the aging alkaline fuel-cell technology, which was used in the Space Shuttle.

<span class="mw-page-title-main">Molten carbonate fuel cell</span>

Molten-carbonate fuel cells (MCFCs) are high-temperature fuel cells that operate at temperatures of 600 °C and above.

<span class="mw-page-title-main">Alkaline fuel cell</span> Type of fuel cell

The alkaline fuel cell (AFC), also known as the Bacon fuel cell after its British inventor, Francis Thomas Bacon, is one of the most developed fuel cell technologies. Alkaline fuel cells consume hydrogen and pure oxygen, to produce potable water, heat, and electricity. They are among the most efficient fuel cells, having the potential to reach 70%.

A proton-exchange membrane, or polymer-electrolyte membrane (PEM), is a semipermeable membrane generally made from ionomers and designed to conduct protons while acting as an electronic insulator and reactant barrier, e.g. to oxygen and hydrogen gas. This is their essential function when incorporated into a membrane electrode assembly (MEA) of a proton-exchange membrane fuel cell or of a proton-exchange membrane electrolyser: separation of reactants and transport of protons while blocking a direct electronic pathway through the membrane.

<span class="mw-page-title-main">Protonic ceramic fuel cell</span>

A protonic ceramic fuel cell or PCFC is a fuel cell based around a ceramic, solid, electrolyte material as the proton conductor from anode to cathode. These fuel cells produce electricity by removing an electron from a hydrogen atom, pushing the charged hydrogen atom through the ceramic membrane, and returning the electron to the hydrogen on the other side of the ceramic membrane during a reaction with oxygen. The reaction of many proposed fuels in PCFCs produce electricity and heat, the latter keeping the device at a suitable temperature. Efficient proton conductivity through most discovered ceramic electrolyte materials require elevated operational temperatures around 600-700 degrees Celsius, however intermediate temperature ceramic fuel cells and lower temperature alternative are an active area of research. In addition to hydrogen gas, the ability to operate at intermediate and high temperatures enables the use of a variety of liquid hydrogen carrier fuels, including: ammonia, and methane. The technology shares the thermal and kinetic advantages of high temperature molten carbonate and solid oxide fuel cells, while exhibiting all of the intrinsic benefits of proton conduction in proton-exchange membrane fuel cells (PEMFC) and phosphoric acid fuel cells (PAFC). PCFCs exhaust water at the cathode and unused fuel, fuel reactant products and fuel impurities at the anode. Common chemical compositions of the ceramic membranes are barium zirconate (BaZrO3), cesium dihydrogen phosphate (CsH2PO4), and complex solid solutions of those materials with other ceramic oxides. The acidic oxide ceramics are sometimes broken into their own class of protonic ceramic fuel cells termed "solid acid fuel cells".

<span class="mw-page-title-main">Aluminium smelting</span> Process of extracting aluminium from its oxide alumina

Aluminium smelting is the process of extracting aluminium from its oxide, alumina, generally by the Hall-Héroult process. Alumina is extracted from the ore bauxite by means of the Bayer process at an alumina refinery.

Genoa Joint Laboratories (GJL) is a scientific research activity founded in 2002, combining expertise in electroceramics and electrochemistry of three facilities: National Research Council - Institute for Energetics and Interphases (CNR-IENI), Department of Chemical and Process Engineering with University of Genova (DICHeP), and the Department of Chemistry and Industrial Chemistry with University of Genova (DCCI), all located in Genoa, Italy.

<span class="mw-page-title-main">Yttria-stabilized zirconia</span> Ceramic with room temperature stable cubic crystal structure

Yttria-stabilized zirconia (YSZ) is a ceramic in which the cubic crystal structure of zirconium dioxide is made stable at room temperature by an addition of yttrium oxide. These oxides are commonly called "zirconia" (ZrO2) and "yttria" (Y2O3), hence the name.

<span class="mw-page-title-main">Solid oxide electrolyzer cell</span> Type of fuel cell

A solid oxide electrolyzer cell (SOEC) is a solid oxide fuel cell that runs in regenerative mode to achieve the electrolysis of water by using a solid oxide, or ceramic, electrolyte to produce hydrogen gas and oxygen. The production of pure hydrogen is compelling because it is a clean fuel that can be stored, making it a potential alternative to batteries, methane, and other energy sources. Electrolysis is currently the most promising method of hydrogen production from water due to high efficiency of conversion and relatively low required energy input when compared to thermochemical and photocatalytic methods.

The lithium–air battery (Li–air) is a metal–air electrochemical cell or battery chemistry that uses oxidation of lithium at the anode and reduction of oxygen at the cathode to induce a current flow.

A metal–air electrochemical cell is an electrochemical cell that uses an anode made from pure metal and an external cathode of ambient air, typically with an aqueous or aprotic electrolyte.

Gadolinium-doped ceria (GDC) (known alternatively as gadolinia-doped ceria, gadolinium-doped cerium oxide (GCO), cerium-gadolinium oxide (CGO), or cerium(IV) oxide, gadolinium-doped, formula Gd:CeO2) is a ceramic electrolyte used in solid oxide fuel cells (SOFCs). It has a cubic structure and a density of around 7.2 g/cm3 in its oxidised form. It is one of a class of ceria-doped electrolytes with higher ionic conductivity and lower operating temperatures (<700 °C) than those of yttria-stabilized zirconia, the material most commonly used in SOFCs. Because YSZ requires operating temperatures of 800–1000 °C to achieve maximal ionic conductivity, the associated energy and costs make GDC a more optimal (even "irreplaceable", according to researchers from the Fraunhofer Society) material for commercially viable SOFCs.

A triple phase boundary (TPB) is a geometrical class of phase boundary and the location of contact between three different phases. A simple example of a TPB is a coastline where land, air and sea meet to create an energetic location driven by solar, wind and wave energy capable of supporting a high level of biodiversity. This concept is particularly important in the description of electrodes in fuel cells and batteries. For example for fuel cells, the three phases are an ion conductor (electrolyte), an electron conductor, and a virtual "porosity" phase for transporting gaseous or liquid fuel molecules. The electrochemical reactions that fuel cells use to produce electricity occur in the presence of these three phases. Triple phase boundaries are thus the electrochemically active sites within electrodes.

<span class="mw-page-title-main">Mixed conductor</span>

Mixed conductors, also known as mixed ion-electron conductors(MIEC), are a single-phase material that has significant conduction ionically and electronically. Due to the mixed conduction, a formally neutral species can transport in a solid and therefore mass storage and redistribution are enabled. Mixed conductors are well known in conjugation with high-temperature superconductivity and are able to capacitate rapid solid-state reactions.

Solid acid fuel cells (SAFCs) are a class of fuel cells characterized by the use of a solid acid material as the electrolyte. Similar to proton exchange membrane fuel cells and solid oxide fuel cells, they extract electricity from the electrochemical conversion of hydrogen- and oxygen-containing gases, leaving only water as a byproduct. Current SAFC systems use hydrogen gas obtained from a range of different fuels, such as industrial-grade propane and diesel. They operate at mid-range temperatures, from 200 to 300 °C.

<span class="mw-page-title-main">Reversible solid oxide cell</span>

A reversible solid oxide cell (rSOC) is a solid-state electrochemical device that is operated alternatively as a solid oxide fuel cell (SOFC) and a solid oxide electrolysis cell (SOEC). Similarly to SOFCs, rSOCs are made of a dense electrolyte sandwiched between two porous electrodes. Their operating temperature ranges from 600°C to 900°C, hence they benefit from enhanced kinetics of the reactions and increased efficiency with respect to low-temperature electrochemical technologies.

<span class="mw-page-title-main">Lithium aluminium germanium phosphate</span> Chemical compound

Lithium aluminium germanium phosphate, typically known with the acronyms LAGP or LAGPO, is an inorganic ceramic solid material whose general formula is Li
1+x
Al
x
Ge
2-x
(PO
4
)
3
. LAGP belongs to the NASICON family of solid conductors and has been applied as a solid electrolyte in all-solid-state lithium-ion batteries. Typical values of ionic conductivity in LAGP at room temperature are in the range of 10–5 - 10–4 S/cm, even if the actual value of conductivity is strongly affected by stoichiometry, microstructure, and synthesis conditions. Compared to lithium aluminium titanium phosphate (LATP), which is another phosphate-based lithium solid conductor, the absence of titanium in LAGP improves its stability towards lithium metal. In addition, phosphate-based solid electrolytes have superior stability against moisture and oxygen compared to sulfide-based electrolytes like Li
10
GeP
2
S
12
(LGPS) and can be handled safely in air, thus simplifying the manufacture process. Since the best performances are encountered when the stoichiometric value of x is 0.5, the acronym LAGP usually indicates the particular composition of Li
1.5
Al
0.5
Ge
1.5
(PO
4
)
3
, which is also the typically used material in battery applications.

References

  1. Badwal, SPS. "Review of Progress in High Temperature Solid Oxide Fuel Cells". Journal of the Australian Ceramics Society. 50 (1). Archived from the original on 29 November 2014.
  2. 1 2 Singh, Mandeep; Zappa, Dario; Comini, Elisabetta (August 2021). "Solid oxide fuel cell: Decade of progress, future perspectives and challenges". International Journal of Hydrogen Energy. 46 (54): 27643–27674. doi:10.1016/j.ijhydene.2021.06.020. S2CID   237909427.
  3. 1 2 3 4 Boldrin, Paul; Brandon, Nigel P. (11 July 2019). "Progress and outlook for solid oxide fuel cells for transportation applications". Nature Catalysis. 2 (7): 571–577. doi:10.1038/s41929-019-0310-y. hdl: 10044/1/73325 . S2CID   199179410.
  4. Gao, Yang; Zhang, Mingming; Fu, Min; Hu, Wenjing; Tong, Hua; Tao, Zetian (September 2023). "A comprehensive review of recent progresses in cathode materials for Proton-conducting SOFCs". Energy Reviews. 2 (3): 100038. doi: 10.1016/j.enrev.2023.100038 . S2CID   259652830.
  5. Vignesh, D.; Rout, Ela (2 March 2023). "Technological Challenges and Advancement in Proton Conductors: A Review". Energy & Fuels. 37 (5): 3428–3469. doi: 10.1021/acs.energyfuels.2c03926 . S2CID   256964689.
  6. Wang, Qi; Fan, Hui; Xiao, Yanfei; Zhang, Yihe (November 2022). "Applications and recent advances of rare earth in solid oxide fuel cells". Journal of Rare Earths. 40 (11): 1668–1681. doi:10.1016/j.jre.2021.09.003. S2CID   240563264.
  7. Hagen, Anke; Rasmussen, Jens F.B.; Thydén, Karl (September 2011). "Durability of solid oxide fuel cells using sulfur containing fuels". Journal of Power Sources. 196 (17): 7271–7276. Bibcode:2011JPS...196.7271H. doi:10.1016/j.jpowsour.2011.02.053.
  8. Kim, Jun Hyuk; Liu, Mingfei; Chen, Yu; Murphy, Ryan; Choi, YongMan; Liu, Ying; Liu, Meilin (5 November 2021). "Understanding the Impact of Sulfur Poisoning on the Methane-Reforming Activity of a Solid Oxide Fuel Cell Anode". ACS Catalysis. 11 (21): 13556–13566. doi:10.1021/acscatal.1c02470.
  9. Boldrin, Paul; Ruiz-Trejo, Enrique; Mermelstein, Joshua; Bermúdez Menéndez, José Miguel; Ramı́rez Reina, Tomás; Brandon, Nigel P. (23 November 2016). "Strategies for Carbon and Sulfur Tolerant Solid Oxide Fuel Cell Materials, Incorporating Lessons from Heterogeneous Catalysis". Chemical Reviews. 116 (22): 13633–13684. doi: 10.1021/acs.chemrev.6b00284 . hdl: 10044/1/41491 . PMID   27933769.
  10. Ceramic fuel cells achieves world-best 60% efficiency for its electricity generator units Archived 3 June 2014 at the Wayback Machine . Ceramic Fuel Cells Limited. 19 February 2009
  11. 1 2 Electricity from wood through the combination of gasification and solid oxide fuel cells, Ph.D. Thesis by Florian Nagel, Swiss Federal Institute of Technology Zurich, 2008
  12. Radenahmad, Nikdalila; Azad, Atia Tasfiah; Saghir, Muhammad; Taweekun, Juntakan; Bakar, Muhammad Saifullah Abu; Reza, Md Sumon; Azad, Abul Kalam (March 2020). "A review on biomass derived syngas for SOFC based combined heat and power application". Renewable and Sustainable Energy Reviews. 119: 109560. doi:10.1016/j.rser.2019.109560.
  13. Xu, Qidong; Guo, Zengjia; Xia, Lingchao; He, Qijiao; Li, Zheng; Temitope Bello, Idris; Zheng, Keqing; Ni, Meng (February 2022). "A comprehensive review of solid oxide fuel cells operating on various promising alternative fuels". Energy Conversion and Management. 253: 115175. doi:10.1016/j.enconman.2021.115175.
  14. Sammes, N.M.; et al. (2005). "Design and fabrication of a 100 W anode supported micro-tubular SOFC stack". Journal of Power Sources. 145 (2): 428–434. Bibcode:2005JPS...145..428S. doi:10.1016/j.jpowsour.2005.01.079.
  15. Panthi, D.; et al. (2014). "Micro-tubular solid oxide fuel cell based on a porous yttria-stabilized zirconia support". Scientific Reports. 4: 5754. Bibcode:2014NatSR...4E5754P. doi:10.1038/srep05754. PMC   4148670 . PMID   25169166.
  16. Ott, J; Gan, Y; McMeeking, R; Kamlah, M (2013). "A micromechanical model for effective conductivity in granular electrode structures" (PDF). Acta Mechanica Sinica. 29 (5): 682–698. Bibcode:2013AcMSn..29..682O. doi:10.1007/s10409-013-0070-x. S2CID   51915676.
  17. Zhu, Tenglong; Fowler, Daniel E.; Poeppelmeier, Kenneth R.; Han, Minfang; Barnett, Scott A. (2016). "Hydrogen Oxidation Mechanisms on Perovskite Solid Oxide Fuel Cell Anodes". Journal of the Electrochemical Society. 163 (8): F952–F961. doi:10.1149/2.1321608jes.
  18. Bao, Zhenghong; Yu, Fei (1 January 2018), Li, Yebo; Ge, Xumeng (eds.), "Chapter Two - Catalytic Conversion of Biogas to Syngas via Dry Reforming Process", Advances in Bioenergy, Elsevier, vol. 3, pp. 43–76, doi:10.1016/bs.aibe.2018.02.002 , retrieved 14 November 2020
  19. Rostrup-Nielsen, J. R. (1982). "Sulfur Poisoning". In Figueiredo, José Luís (ed.). Progress in Catalyst Deactivation. NATO Advanced Study Institutes Series. Dordrecht: Springer Netherlands. pp. 209–227. doi:10.1007/978-94-009-7597-2_11. ISBN   978-94-009-7597-2.
  20. Sasaki, K.; Susuki, K. (2006). "H2S Poisoning of Solid Oxide Fuel Cells". Journal of the Electrochemical Society. 153 (11): 11. Bibcode:2006JElS..153A2023S. doi:10.1149/1.2336075.
  21. 1 2 Ge, Xiao-Ming; Chan, Siew-Hwa; Liu, Qing-Lin; Sun, Qiang (2012). "Solid Oxide Fuel Cell Anode Materials for Direct Hydrocarbon Utilization". Advanced Energy Materials. 2 (10): 1156–1181. doi:10.1002/aenm.201200342. ISSN   1614-6840. S2CID   95175720.
  22. Costa-Nunes, Olga; Gorte, Raymond J.; Vohs, John M. (1 March 2005). "Comparison of the performance of Cu–CeO2–YSZ and Ni–YSZ composite SOFC anodes with H2, CO, and syngas". Journal of Power Sources. 141 (2): 241–249. Bibcode:2005JPS...141..241C. doi:10.1016/j.jpowsour.2004.09.022. ISSN   0378-7753.
  23. Nigel Sammes; Alevtina Smirnova; Oleksandr Vasylyev (2005). "Fuel Cell Technologies: State and Perspectives". NATO Science Series, Mathematics, Physics and Chemistry. 202: 19–34. Bibcode:2005fcts.conf.....S. doi:10.1007/1-4020-3498-9_3.
  24. Steele, B.C.H., Heinzel, A. (2001). "Materials for fuel-cell technologies". Nature. 414 (15 November): 345–352. Bibcode:2001Natur.414..345S. doi:10.1038/35104620. PMID   11713541. S2CID   4405856.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  25. Mohan Menon; Kent Kammer; et al. (2007). "Processing of Ce1-xGdxO2-δ (GDC) thin films from precursors for application in solid oxide fuel cells". Advanced Materials Engineering. 15–17: 293–298. doi:10.4028/www.scientific.net/AMR.15-17.293. S2CID   98044813.
  26. Charpentier, P (2000). "Preparation of thin film SOFCs working at reduced temperature". Solid State Ionics. 135 (1–4): 373–380. doi:10.1016/S0167-2738(00)00472-0. ISSN   0167-2738. S2CID   95598314.
  27. Shen, F.; Lu, K. (August 2018). "Comparison of Different Perovskite Cathodes in Solid Oxide Fuel Cells". Fuel Cells. 18 (4): 457–465. doi:10.1002/fuce.201800044. ISSN   1615-6846. S2CID   104669264.
  28. Shimada, Hiroyuki; Suzuki, Toshio; Yamaguchi, Toshiaki; Sumi, Hirofumi; Hamamoto, Koichi; Fujishiro, Yoshinobu (January 2016). "Challenge for lowering concentration polarization in solid oxide fuel cells". Journal of Power Sources. 302: 53–60. doi:10.1016/j.jpowsour.2015.10.024.
  29. Hai-Bo Huo; Xin-Jian Zhu; Guang-Yi Cao (2006). "Nonlinear modeling of a SOFC stack based on a least squares support vector machine". Journal of Power Sources. 162 (2): 1220–1225. Bibcode:2006JPS...162.1220H. doi:10.1016/j.jpowsour.2006.07.031.
  30. 1 2 3 4 Milewski J, Miller A (2006). "Influences of the Type and Thickness of Electrolyte on Solid Oxide Fuel Cell Hybrid System Performance". Journal of Fuel Cell Science and Technology. 3 (4): 396–402. doi:10.1115/1.2349519.
  31. M. Santarelli; P. Leone; M. Calì; G. Orsello (2007). "Experimental evaluation of the sensitivity to fuel utilization and air management on a 100 kW SOFC system". Journal of Power Sources. 171 (2): 155–168. Bibcode:2007JPS...171..155S. doi:10.1016/j.jpowsour.2006.12.032.
  32. Kupecki J.; Milewski J.; Jewulski J. (2013). "Investigation of SOFC material properties for plant-level modeling". Central European Journal of Chemistry. 11 (5): 664–671. doi: 10.2478/s11532-013-0211-x .
  33. Mahato, N; Banerjee, A; Gupta, A; Omar, S; Balani, K (1 July 2015). "Progress in material selection for solid oxide fuel cell technology: A review". Progress in Materials Science. 72: 141–337. doi:10.1016/j.pmatsci.2015.01.001.
  34. Nakajo, Arata; Kuebler, Jakob; Faes, Antonin; Vogt, Ulrich; Schindler, Hansjürgen; Chiang, Lieh-Kwang; Modena, Stefano; Van Herle, Jan (25 January 2012). "Compilation of mechanical properties for the structural analysis of solid oxide fuel cell stacks. Part I. Constitutive materials of anode-supported cells". Ceramics International. 38: 3907–3927. doi:10.1016/j.ceramint.2012.01.043.
  35. Ullmann, H.; Trofimenko, N.; Tietz, F.; Stöver, D.; Ahmad-Khanlou, A. (1 December 2000). "Correlation between thermal expansion and oxide ion transport in mixed conducting perovskite-type oxides for SOFC cathodes". Solid State Ionics. 138 (1–2): 79–90. doi:10.1016/S0167-2738(00)00770-0.
  36. Radovic, M.; Lara-Curzio, E. (December 2004). "Mechanical properties of tape cast nickel-based anode materials for solid oxide fuel cells before and after reduction in hydrogen". Acta Materialia. 52 (20): 5747–5756. Bibcode:2004AcMat..52.5747R. doi:10.1016/j.actamat.2004.08.023.
  37. ASTM. "Standard Test Method for Monotonic Equibiaxial Flexural Strength of Advanced Ceramics at Ambient Temperature, ASTM Standard C1499-04".
  38. Kao, Robert; Perrone, Nicholas; Capps, Webster (1971). "Large-Deflection Solution of the Coaxial-Ring-Circular-Glass-Plate Flexure Problem". Journal of the American Ceramic Society. 54 (11): 566–571. doi:10.1111/j.1151-2916.1971.tb12209.x. ISSN   0002-7820.
  39. Nakajo, Arata; Kuebler, Jakob; Faes, Antonin; Vogt, Ulrich F.; Schindler, Hans Jürgen; Chiang, Lieh-Kwang; Modena, Stefano; Van herle, Jan; Hocker, Thomas (25 January 2012). "Compilation of mechanical properties for the structural analysis of solid oxide fuel cell stacks. Constitutive materials of anode-supported cells". Ceramics International. 38 (5): 3907–3927. doi:10.1016/j.ceramint.2012.01.043.
  40. SECA Coal-Based Systems – LGFCS. www.osti.gov. Retrieved 19 February 2019.
  41. Fuel Cell Stacks Still Going Strong After 5,000 Hours. www.energy.gov (24 March 2009). Retrieved 27 November 2011. Archived 8 October 2009 at the Wayback Machine
  42. Ishihara, Tatsumi (2009). Perovskite Oxide for Solid Oxide Fuel Cells . Springer. p.  19. ISBN   978-0-387-77708-5.
  43. 1 2 Wachsman, Eric; Lee, Kang (18 November 2011). "Lowering the Temperature of Solid Oxide Fuel Cells". Science. 334 (6058): 935–9. Bibcode:2011Sci...334..935W. doi:10.1126/science.1204090. PMID   22096189. S2CID   206533328.
  44. Spivey, B. (2012). "Dynamic modeling, simulation, and MIMO predictive control of a tubular solid oxide fuel cell". Journal of Process Control. 22 (8): 1502–1520. doi:10.1016/j.jprocont.2012.01.015.
  45. "Fuel Cell Comparison". Nedstack. Retrieved 6 November 2016.
  46. Lamp, P.; Tachtler, J.; Finkenwirth, O.; Mukerjee, S.; Shaffer, S. (November 2003). "Development of an Auxiliary Power Unit with Solid Oxide Fuel Cells for Automotive Applications". Fuel Cells. 3 (3): 146–152. doi:10.1002/fuce.200332107.
  47. Gardner, F.J; Day, M.J; Brandon, N.P; Pashley, M.N; Cassidy, M (March 2000). "SOFC technology development at Rolls-Royce". Journal of Power Sources. 86 (1–2): 122–129. doi:10.1016/S0378-7753(99)00428-0.
  48. "Northwestern group invent inks to make SOFCs by 3D printing". Fuel Cells Bulletin. 2015: 11. 2015. doi:10.1016/S1464-2859(15)70024-6.
  49. "The Ceres Cell". Company Website. Ceres Power. Archived from the original on 13 December 2013. Retrieved 30 November 2009.
  50. "HITEC". Hitec.mse.ufl.edu. Archived from the original on 12 December 2013. Retrieved 8 December 2013.
  51. Cooling Down Solid-Oxide Fuel Cells. Technologyreview.com. 20 April 2011. Retrieved 27 November 2011.
  52. Anne Hauch; Søren Højgaard Jensen; Sune Dalgaard Ebbesen; Mogens Mogensen (2009). "Durability of solid oxide electrolysis cells for hydrogen production" (PDF). Risoe Reports. 1608: 327–338. Archived from the original (PDF) on 11 July 2009.
  53. Rainer Küngas; Peter Blennow; Thomas Heiredal-Clausen; Tobias Holt; Jeppe Rass-Hansen; Søren Primdahl; John Bøgild Hansen (2017). "eCOs - A Commercial CO2 Electrolysis System Developed by Haldor Topsoe". ECS Trans. 78 (1): 2879–2884. Bibcode:2017ECSTr..78a2879K. doi:10.1149/07801.2879ecst.
  54. Nithya, M., and M. Rajasekhar. "Preparation and Characterization of NdCrO3 Cathode for Intermediate Temperature Fuel Cell Application." International Journal of Applied Chemistry 13, no. 4 (2017): 879-886.
  55. Zhu, Bin (2003). "Functional ceria–salt-composite materials for advanced ITSOFC applications". Journal of Power Sources. 114 (1): 1–9. Bibcode:2003JPS...114....1Z. doi:10.1016/s0378-7753(02)00592-x.
  56. Choi, S.; Yoo, S.; Park, S.; Jun, A.; Sengodan, S.; Kim, J.; Shin, J. Highly efficient and robust cathode materials for low-temperature solid oxide fuel cells: PrBa0.5Sr0.5Co(2-x)Fe(x)O(5+δ). Sci. Rep. 2013, 3, 2426-2428.
  57. Hibini, T.; Hashimoto, A.; Inoue, T.; Tokuno, J.; Yoshida, S.; Sano, M. A Low-Operating-Temperature Solid Oxide Fuel Cell in Hydrocarbon-Air Mixtures. Science. 2000. 288, 2031-2033.
  58. Wachsman, E.; Lee, Kang T. (2011). "Lowering the Temperature of Solid Oxide Fuel Cells". Science. 334 (6058): 935–939. Bibcode:2011Sci...334..935W. doi:10.1126/science.1204090. PMID   22096189. S2CID   206533328.
  59. Zuo, C.; Zha, S.; Liu, M.; Hatano, M.; Uchiyama, M. Ba(Zr0.1Ce0.7Y0.2)O3-δ as an Electrolyte for Low-Temperature Solid-Oxide Fuel Cells. Advanced Materials. 2006, 18, 3318-3320
  60. S.H. Chan; H.K. Ho; Y. Tian (2003). "Multi-level modeling of SOFC-gas turbine hybrid system". International Journal of Hydrogen Energy. 28 (8): 889–900. doi:10.1016/S0360-3199(02)00160-X.
  61. L. K. C. Tse; S. Wilkins; N. McGlashan; B. Urban; R. Martinez-Botas (2011). "Solid oxide fuel cell/gas turbine trigeneration system for marine applications". Journal of Power Sources. 196 (6): 3149–3162. Bibcode:2011JPS...196.3149T. doi:10.1016/j.jpowsour.2010.11.099.
  62. Isfahani, SNR; Sedaghat, Ahmad (15 June 2016). "A hybrid micro gas turbine and solid state fuel cell power plant with hydrogen production and CO2 capture". International Journal of Hydrogen Energy. 41 (22): 9490–9499. doi:10.1016/j.ijhydene.2016.04.065. S2CID   100859434.
  63. Giddey, S; Badwal, SPS; Kulkarni, A; Munnings, C (2012). "A comprehensive review of direct carbon fuel cell technology". Progress in Energy and Combustion Science. 38 (3): 360–399. doi:10.1016/j.pecs.2012.01.003.
  64. Wu, Wei; Ding, Dong; Fan, Maohong; He, Ting (30 May 2017). "A High Performance Low Temperature Direct Carbon Fuel Cell". ECS Transactions. 78 (1): 2519–2526. Bibcode:2017ECSTr..78a2519W. doi:10.1149/07801.2519ecst. ISSN   1938-6737. OSTI   1414432.
  65. Khan, Feroze (1 January 2012). Effect of Hydrogen Sulfide in Landfill Gas on Anode Poisoning of Solid Oxide Fuel Cells (Thesis). Youngstown State University.