METALLIC MESH-BASED GAS DIFFUSION ELECTRODES FOR UTILIZATION OF SPARINGLY SOLUBLE GASES IN ELECTROCHEMICAL REACTIONS WITH NONAQUEOUS ELECTROLYTES
A system and method for supplying a gas to an electrochemical system is described.
Latest Massachusetts Institute of Technology Patents:
- AN IMPLANTABLE CLOSED-LOOP SYSTEM FOR MONITORING AND TREATMENT
- THREE-DIMENSIONAL MACHINE KNITTING OF ELECTRONIC TEXTILE FOR ACTIVITY RECOGNITION AND BIOMECHANICAL MONITORING
- Skills and tasks demand forecasting
- Self-righting systems, methods, and related components
- Barocaloric heat transfer systems and methods of use
This application claims priority to U.S. Provisional Application No. 63/000,458, filed Mar. 26, 2020, which is incorporated by reference in its entirety.
FEDERALLY SPONSORED RESEARCH OR DEVELOPMENTThis invention was made with Government support under Grant No. CBET1944007 awarded by the National Science Foundation. The Government has certain rights in the invention.
FIELD OF THE INVENTIONThe invention relates to gas diffusion electrodes.
BACKGROUNDElectrochemical transformations in nonaqueous solvents are important for synthetic and energy storage applications. Use of nonpolar gaseous reactants such as nitrogen and hydrogen in nonaqueous solvents can be limited by their low solubility and slow transport. Conventional gas diffusion electrodes can improve transport of gaseous species in aqueous electrolytes by facilitating efficient gas-liquid contacting in the vicinity of the electrode. Conventional gas diffusion electrodes cannot improve the transport in many nonaqueous electrolytes, however, as the hydrophobic interactions necessary for creating gas-liquid contacting are not present in nonaqueous electrolytes. This can lead to flooding of the electrode and low rates for gas utilization when using nonaqueous electrolytes.
SUMMARYIn one aspect, an electrochemical system can include a housing including a chamber, an electrode within the housing, and a gas permeable metal on a surface of the electrode in contact with the chamber.
In another aspect, a method of supplying a gas to an electrochemical system can include contacting a gas with a gas permeable metal on a surface of an electrode in a chamber of a housing. The gas can be a precursor that is converted to a reactive gas by the electrode.
In another aspect, a method of oxidizing or reducing a gas can include contacting a gas with a gas permeable metal on a surface of an electrode. The gas can be a sparingly soluble gas. The sparingly soluble gas can be nitrogen or hydrogen. In certain circumstances, the ammonia can be produced at a Faradaic yield of at least 30% or at least 40%. In certain circumstances, the method can include supplying a pressure of the gas in the chamber to create a pressure differential at the electrode. The method can allow for the use of a sparingly soluble gas as a reagent in chemical reactions.
In another aspect, an electrochemical system can include a first electrode including a housing including a chamber, an electrode within the housing, and a gas permeable metal on a surface of the electrode in contact with the chamber, and a second electrode including a gas inlet to a housing including a gas permeable metal on a surface of an electrode and a first outlet to release a product from the system.
In certain circumstances, the system can include a gas inlet to the housing.
In certain circumstances, the system can include a first outlet of the housing to release a product from the housing.
In certain circumstances, the gas permeable metal can include a porous metal or a metal mesh system. In certain circumstances, each gas permeable metal can include a metal mesh system. In certain circumstances, the metal mesh can include 100, 200, 300, 400 or 500 fibers per inch. In other circumstances, the metal mesh can be asymmetric and include 100, 200, 300, 400 or 500 fibers per inch in one direction and 500, 1000, 1500, or 2000 fibers per inch in a second direction.
In certain circumstances, the gas permeable metal can include openings of between 1 and 200 micrometers, preferably between 2 and 100 micrometers.
In certain circumstances, each gas permeable metal can include openings of between 1 and 200 micrometers, preferably between 2 and 100 micrometers.
In certain circumstances, the gas permeable metal can include metal fibers or a porous metal.
In certain circumstances, at least one gas permeable metal can include metal fibers or a porous metal.
In certain circumstances, the gas permeable metal can include stainless steel, steel, nickel, iron, copper, silver, gold, or platinum.
In certain circumstances, the gas permeable metal can include a catalyst on a surface of the gas permeable metal. For example, the catalyst can include a surface treated with catalytic nanoparticles or catalytic nanoparticles deposited on the surface. In certain circumstances, the gas permeable metal can include a catalyst, for example, a catalytic metal, metal oxide, metal sulfide, or metal phosphide.
In certain circumstances, the gas permeable metal can be exposed to a pressure gradient. In certain circumstances, at least one gas permeable metal can be exposed to a pressure gradient. In certain circumstances, the method can include supplying a pressure of the gas in the chamber to create a pressure differential in the housing.
In certain circumstances, the method can include applying a voltage to the electrode.
Other aspects, embodiments, and features will be apparent from the following description, the drawings, and the claims.
A method to utilize sparingly soluble gases in electrochemical reactions at high rates in nonaqueous solvents is described. The method can be relevant for electroorganic synthesis and fuel production where control of proton activity is important. The method relies on metallic supports and a pressure gradient applied across the electrode. The method can be used in a variety of electrochemical systems, for example, as applied herein to hydrogen oxidation in two nonaqueous solvents and nitrogen reduction in one solvent; the two chemistries are coupled to produce ammonia from nitrogen and hydrogen at high rates.
A sparingly soluble gas, generally, is a non-polar gas that does not react or interact favorably with a solvents. For example, N2, H2, CO, and CH4 are sparingly soluble. Gasses with solubilities less than 50-100 mM at 1 atm can be considered to be a sparingly soluble gas. A gas with a Henry's constant <0.05 M/atm can be considered to be sparingly soluble gas.
An electrochemical system can include a housing including a chamber, an electrode within the housing, and a gas permeable metal on a surface of the electrode in contact with the chamber. The system can be used in a method of supplying a gas to an electrochemical system can include contacting a gas with a gas permeable metal on a surface of an electrode in a chamber of a housing. The method can include applying a voltage to the electrode. The system can be a gas diffusion electrochemical system, in which metallic supports can be used in the GDEs combined with a pressure gradient across the GDE. The gas can be at a higher pressure relative to the liquid, to obtain effective gas-liquid contacting at the electrode surface for high reaction rates in nonaqueous solvents. Metallic supports can avoid flooding of the electrodes in the absence of hydrophobic repulsion, while the pressure gradient helps maintain the gas-liquid interface in the desired location. As the method relies on a physical effect (a pressure gradient) to establish the gas-liquid boundary, it can be used with any solvent, including nonaqueous solvents.
The gas permeable metal can be a metal support, which can include metal fibers or a porous metal. The metallic support can be stainless steel, which was be woven into a fine cloths with very thin fibers. In one example, the cloth can be a 400×400 mesh, which contains 400 fibers 25 micrometers in diameter per inch of length, with a spacing of approximately 25-40 micrometers between fibers. The gas permeable metal can include openings of between 1 and 200 micrometers, preferably between 2 and 100 micrometers. The metallic supports can be made from any metal that is amenable to forming, including, but not limited to stainless steel (304 and 316), steel, nickel, iron, copper, silver, gold, or platinum. The metals can be formed into porous materials which are gas permeable, such as metal cloths and meshes, but also metal filters and sponges. The characteristic pore size of the material can be at least as large as 200 micrometers, and down to 2 micrometers; the pore size can be smaller if larger operating pressure gradients are desired. In certain circumstances, the gas permeable metal can include a metal mesh system. The metal mesh can be symmetric or asymmetric. In certain circumstances, the metal mesh can include 100, 200, 300, 400 or 500 fibers per inch. In other circumstances, the metal mesh can be asymmetric and include 100, 200, 300, 400 or 500 fibers per inch in one direction and 500, 1000, 1500, or 2000 fibers per inch in a second direction. In certain circumstances, the gas permeable metal can be exposed to a pressure gradient. In certain circumstances, the method can include supplying a pressure of the gas in the chamber to create a pressure differential in the housing. The pressure gradient applied across the cloth can be 0.5 to 10 kilopascals, for example, 1, 2, 3, 4 or 5 kilopascals. The pressure gradient can depend on the electrolyte used and the pore size of the support. At high pressure gradients, above the Laplace pressure of the material, gas may cross the support and enter the electrolyte; the invention still works even under these operating conditions.
The gas permeable metal can include additional catalysts grown, placed, or deposited on a surface of the metal. Potential catalysts can include: metals, such as silver, gold, platinum, nickel, lithium, zinc, or titanium; metal oxides, such as iridium oxide, cobalt oxide, iron oxide, copper oxide, titanium oxide, or silver oxide; metal sulfides, such as molybdenum sulfide, or cadmium sulfide; metal nitrides, such as lithium nitride, cobalt nitride, nickel nitride, and mixtures thereof; metal phosphides, such as cobalt phosphide, nickel phosphide, or mixtures thereof; molecular catalysts, such as metal phthalocyanines, such as cobalt phthalocyanine or metal porphyrins.
Catalysts can be deposited onto the metal substrate or synthesized on its surface. Methods for deposition include electroplating of metals, electroless plating of metals, electrophoretic deposition, sputtering, pulse laser deposition, chemical vapor deposition, spin-coating, or application of catalyst inks. Methods for in-situ manufacture include oxidation (for making oxides), treatment with nitrogen and ammonia (for making nitrides), heating with sulfur (for making sulfides), heating with phosphorus (for making phosphides), or thermal decomposition of complex materials.
Referring to
The gas species can be a gas species that can be oxidized or reduced, for example, N2, O2, H2, CO or CO2.
The solvent can be an inert organic solvent that in which the electrolyte salt, substrate, and proton carrier can be dissolved. In certain circumstances, a carbonylation reaction or reductive amination reaction can involve a substrate dissolved in the solvent. The concentration of the proton carrier can be 5 mM, 10 mM, 15 mM, 20 mM, 25 mM, 30 mM, 35 mM, 40 mM, 50 mM, 60 mM, 70 mM, 80 mM, 90 mM, 100 mM, 200 mM, 250 mM, 300 mM, 350 mM, 400 mM, 450 mM, or 500 mM. For example, for nitrogen reduction, the concentration of the proton carrier can be 50 mM or higher. The concentration of the substrate can be 5 mM, 10 mM, 15 mM, 20 mM, 25 mM, 30 mM, 35 mM, 40 mM, 50 mM, 60 mM, 70 mM, 80 mM, 90 mM, 100 mM, 200 mM, 250 mM, 300 mM, 350 mM, 400 mM, 450 mM, or 500 mM. For example, for nitrogen reduction, the concentration of the substrate can be 50 mM or higher. The concentration of the electrolyte can be 0.1 M, 0.25 M, 0.5 M, 1 M, 2 M, 4 M, 5 M, 6 M, 7 M, 8 M, 9 M, or 10 M.
The temperature and pressure can be ambient temperature and pressure. There can be a pressure gradient between housing 12 and electrolyte fluid 25 and housing 18 and electrolyte fluid 25.
The reaction product can be produced in a gas phase.
The voltage can be between about 0.2V and 40.0V, between about 0.4V and 35.0V, or between about 1V and 30.0V. For example, the voltage can be about 1.0V, 5.0V, 10.0V, 15.0V, 20.0V, 25.0V, 30.0V, 35.0V, or 40.0V.
Each of the first electrode and the second electrode can be or can include a noble metal, for example, platinum or palladium.
A variety of reactor designs can implement the method. The method described herein can be performed under various different electrochemical cell geometries and configurations, which include an anode and cathode, connected to an external power source, with an ionically conductive medium between the two electrodes. A third reference electrode may be incorporated if necessary for control of the potential at the electrodes. Resistive losses can be reduced by decreasing the distance between electrodes. The process may be conducted under batch or continuous conditions. An ionically conductive membrane, such as Nafion or Selemion, or a separator, such as Celgard or Daramic, can be used in the structure, but is not required.
The overall reaction may be tuned by choice of the cathode and the reactor conditions. For instance, if nitrogen is flowed to the cathode, then nitrogen will be reduced to generate ammonia; if hydrogen is flowed to the anode, then hydrogen will be reduced to generate protons.
The nonaqueous solvent can include acetonitrile, DMSO (dimethyl sulfoxide), DMF (dimethylformamide), THF (tetrahydrofuran), DCM (dichloromethane), and propionitrile. The electrolyte can contain a conductive salt such as TBABF4 (Tetrabutylammonium tetrafluoroborate), TBAPF6 (Tetrabutylammonium hexafluorophosphate), NaClO4(Sodium perchlorate), LiClO4(Lithium perchlorate), or TEAP(tetraethylammonium perchlorate), or a combination thereof. The non-aqueous solvent can assist substrate solubility.
The electrodes can include a catalyst. Catalysts that may be used in these GDEs include metals such as alkali metals such as lithium, sodium, potassium, alkali-earth metals such as magnesium, transition metals such nickel, platinum, copper, gold, silver. Metallic catalysts can be deposited onto the supports electrochemically from solution, via electroless plating, or sputtered onto the supports ex situ. Catalyst nanoparticles such as metal oxides, metal nitrides, and metal sulfides can be deposited onto the supports via drop casting, sputtering, or pulse laser deposition. For nitrogen reduction, lithium metal can be deposited in situ electrochemically. For hydrogen oxidation, nickel can be deposited electrochemically onto stainless steel cloths, onto which platinum is then electrochemically deposited.
Electrochemical transformations in nonaqueous solvents are important for synthetic and energy storage applications. Use of nonpolar gaseous reactants such as nitrogen and hydrogen in nonaqueous solvents is limited by their low solubility and slow transport. Conventional gas diffusion electrodes improve transport of gaseous species in aqueous electrolytes by facilitating efficient gas-liquid contacting in the vicinity of the electrode. Their use with nonaqueous solvents is hampered by the absence of hydrophobic repulsion between the liquid phase and carbon fiber support. Herein, a method to overcome transport limitations in tetrahydrofuran is reported using a stainless steel cloth-based support for ammonia synthesis paired with hydrogen oxidation. An ammonia partial current density of 8.7±1.5 mA cm−2 and a Faradaic efficiency of 35±6% are obtained using a lithium-mediated approach. Hydrogen oxidation current densities up to 25 mA cm−2 are obtained in two nonaqueous solvents with nearly unity Faradaic efficiency. The approach can be applied to produce ammonia from nitrogen and water splitting-derived hydrogen.
Electrochemical synthesis of chemicals is an attractive alternative approach to traditional thermochemical methods. In some reactions, electric potential can act as a thermodynamic driving force instead of high temperatures and pressures, which may allow for operation at milder conditions and in a modular fashion. See, for example, Schiffer, Z. J. & Manthiram, K. Electrification and Decarbonization of the Chemical Industry. Joule 1, 10-14 (2017); and Yan, M., Kawamata, Y. & Baran, P. S. Synthetic Organic Electrochemistry: Calling All Engineers. Angew. Chemie Int. Ed. 57, 4149-4155 (2018), each of which is incorporated by reference in its entirety. Ammonia (NH3) production is an example of a reaction that may benefit from being operated electrochemically. See, for example, Chen, J. G. et al. Beyond fossil fuel-driven nitrogen transformations. Science (80). 360, eaar6611 (2018); and Soloveichik, G. Electrochemical synthesis of ammonia as a potential alternative to the Haber-Bosch process. Nat. Catal. 2, 377-380 (2019), each of which is incorporated by reference in its entirety. NH3 is currently produced predominantly via the Haber-Bosch process, which operates at high temperatures (300-500° C.) and pressures (200-300 bar) and requires a coupled steam reforming plant for hydrogen (H2) production. See, for example, Shipman, M. A. & Symes, M. D. Recent progress towards the electrosynthesis of ammonia from sustainable resources. Catal. Today 286, 57-68 (2017), which is incorporated by reference in its entirety. This leads to high capital costs for the process and centralization of production, a situation that is poorly matched with the distributed nature of ammonia utilization. See, for example, Comer, B. M. et al. Prospects and Challenges for Solar Fertilizers. Joule 3, 1578-1605 (2019) which is incorporated by reference in its entirety. Alternative methods for producing hydrogen, such as water splitting, may overcome some of the issues associated with the traditional Haber-Bosch process, such as the large amount of CO2 emissions and high capital cost associated with steam reforming. See, for example, Suryanto, B. H. R. et al. Challenges and prospects in the catalysis of electroreduction of nitrogen to ammonia. Nat. Catal. 2, 290-296 (2019), which is incorporated by reference in its entirety. However, these methods do not overcome the need for large scales for the ammonia synthesis reactor itself, as it must still be run at high temperatures and pressures. An electrochemical process—even one which utilizes multiple reactors—that produces ammonia from nitrogen and water requires a thermodynamic minimum potential of 1.17 V at standard conditions. See, for example, Soloveichik, G. Electrochemical synthesis of ammonia as a potential alternative to the Haber-Bosch process. Nat. Catal. 2, 377-380 (2019), which is incorporated by reference in its entirety. Potential is a potent thermodynamic driver, providing mild conditions conducive to modular and small-scale operation of electrochemical processes. See, for example, Foster, S. L. et al. Catalysts for nitrogen reduction to ammonia. Nat. Catal. 1, 490-500 (2018), which is incorporated by reference in its entirety.
Despite the attractiveness of electrochemistry in synthetic applications, several challenges must be overcome to allow for efficient scale-up of the technology. One of the most important issues is the use of non-renewable reactants at the counter electrode. For reductive chemistries, which are important for energy storage and certain synthetic applications, the counter reaction is often oxidation of solvent or sacrificial anodes made of active metals. See, for example, Jiao, F. & Xu, B. Electrochemical Ammonia Synthesis and Ammonia Fuel Cells. Adv. Mater. 31, 1805173 (2019); Davis, S. J. et al. Net-zero emissions energy systems. Science (80). 360, eaas9793 (2018); Liu, X., Jiao, Y., Zheng, Y., Jaroniec, M. & Qiao, S.-Z. Building Up a Picture of the Electrocatalytic Nitrogen Reduction Activity of Transition Metal Single-Atom Catalysts. J. Am. Chem. Soc. 141, 9664-9672 (2019); Peters, B. K. et al. Scalable and safe synthetic organic electroreduction inspired by Li-ion battery chemistry. Science (80-.). 363, 838-845 (2019); Matthessen, R., Fransaer, J., Binnemans, K. & De Vos, D. E. Electrocarboxylation: towards sustainable and efficient synthesis of valuable carboxylic acids. Beilstein J. Org. Chem. 10, 2484-2500 (2014); and Motile, S. et al. Modern Electrochemical Aspects for the Synthesis of Value-Added Organic Products. Angew. Chemie-Int. Ed. 57, 6018-6041 (2018), each of which is incorporated by reference in its entirety. In aqueous systems, solvent oxidation is permissible and often desired. However, oxidation of organic solvents or sacrificial anodes decreases the atom economy of reactions greatly and makes processes utilizing these reactions poorly amenable to continuous operation. See, for example, Matthessen, R., Fransaer, J., Binnemans, K. & De Vos, D. E. Electrocarboxylation: towards sustainable and efficient synthesis of valuable carboxylic acids. Beilstein J. Org. Chem. 10, 2484-2500 (2014), which is incorporated by reference in its entirety.
In the context of nitrogen reduction, one method that produces ammonia at high rates and Faradaic efficiencies is the lithium-mediated approach. The approach involves reacting lithium metal with nitrogen to form lithium nitride, a reaction which is spontaneous at ambient conditions. The lithium nitride is then protonated to make ammonia and a lithium salt. The lithium salt is electrochemically reduced to lithium metal to close the catalytic cycle (
Oxidizing H2 at the anode to produce protons of a controlled thermodynamic activity avoids the aforementioned issues. See, for example, Singh, A. R. et al. Strategies toward Selective Electrochemical Ammonia Synthesis. ACS Catal. 9, 8316-8324 (2019), which is incorporated by reference in its entirety. As an added benefit, hydrogen oxidation can be used as a renewable anode reaction for synthetic applications in which sacrificial anodes are used, allowing for continuous production of useful chemicals. See, for example, Matthessen, R., Fransaer, J., Binnemans, K. & De Vos, D. E. Electrocarboxylation: towards sustainable and efficient synthesis of valuable carboxylic acids. Beilstein J. Org. Chem. 10, 2484-2500 (2014), which is incorporated by reference in its entirety. However, the rate of hydrogen oxidation in nonaqueous solvents at flooded electrodes is limited by the solubility of hydrogen and its corresponding diffusion-limited oxidation rate, equal to several milliamperes per square centimeter (mA cm−2) (
One way to overcome diffusion limitations for gaseous reactants in electrochemical reactions is to use gas diffusion electrodes (GDEs), in which intimate contact between the gas, electrolyte, and catalyst is generated. See, for example, Mathur, V. & Crawford, J. Fundamentals of Gas Diffusion Layers in PEM Fuel Cells. Recent Trends Fuel Cell Sci. Technol. 400, 116-128 (2007), which is incorporated by reference in its entirety. This contacting minimizes the distance that gas molecules have to travel through the electrolyte to react at the catalyst (
In the aforementioned applications, the electrolytes are typically aqueous solutions or water-saturated polymeric materials, while the GDE support is hydrophobized to control wetting. See, for example, Mathur, V. & Crawford, J. Fundamentals of Gas Diffusion Layers in PEM Fuel Cells. Recent Trends Fuel Cell Sci. Technol. 400, 116-128 (2007), which is incorporated by reference in its entirety. The hydrophobic interactions between the electrolyte and support, as well as the small pore size in the support prevent electrolyte penetration and flooding into the fibrous structure of the GDE. Instead, a thin layer of electrolyte is in contact with the catalyst through which reactant gas molecules must diffuse (
An electrochemical Haber-Bosch reactor is described below, where hydrogen and nitrogen are utilized to produce ammonia at ambient conditions using electrical potential. The electrolyte contains one molar tetrafluoroborate (1 M LiBF4) and 0.11 molar ethanol (EtOH) in tetrahydrofuran (THF); it is nonaqueous, as water is not used as the bulk solvent for the electrolyte. Nitrogen gas is reduced at the cathode on a stainless steel cloth by lithium metal which is electrochemically plated onto the mesh in situ. The steel cloth is set up to act as gas diffusion electrode, with the electrolyte and gas well-separated by the sloth, which generates efficient gas-liquid contacting at the electrode. The steel mesh acts as the support onto which the catalyst, lithium metal, is deposited in situ. Nitrogen gas was flowed past the electrode, with single pass conversions reaching ˜10%.
At the anode, the electrode was a platinum-coated stainless steel cloth that was also set up as a gas diffusion electrode. Hydrogen gas was oxidized at anode with nearly unity (>99%) Faradaic efficiency and rates at least an order of magnitude higher than possible at flooded electrodes (˜2.5 mA cm−2 at flooded electrodes vs 25 mA cm−2 at the GDE). High single pass conversions (>80%) of the feed gas were demonstrated.
In addition to demonstrating the ability to oxidize hydrogen and reduce nitrogen in THF-based electrolytes, the work also demonstrates that platinum coated stainless steel cloths can be used to oxidize hydrogen in other nonaqueous solvents. The solvent used to demonstrate this was 1 M LiBF4 in 9:1 propylene carbonate/ethylene carbonate. In this solvent, the transport-limited current density for hydrogen oxidation at conventional flooded platinum electrodes is ˜0.25 mA cm−2, fartoo low for practical applications. Oxidation was demonstrated with unity selectivity at rates two orders of magnitude higher than previously achieved, at 25 mA cm−2.
The architecture described herein can also be used to improve the selectivities and rates of fuel production from gaseous feedstocks such as N2, CO, and CO2 by alleviating diffusion limitations and allowing for precise control of proton activity in a nonaqueous solvent. For example, the hydrogen oxidation anode described herein can also be coupled to a wide range of cathodic hydrogenations, making it broadly useful in electroorganic synthesis, an emerging area that is finding commercial interest in energy and pharmaceutical companies. The methods and systems described herein currently holds the record for the highest rates of ammonia synthesis at ambient conditions (
As electrochemical reactions involving poorly soluble gases at flooded electrodes are limited by the diffusion rate of gas molecules from the bulk electrolyte to the electrode surface, methods to decrease the diffusion distance have been proposed. In gas diffusion electrodes, the interface between the gas and liquid phases is positioned close to the electroactive surface to significantly reduce the distance that the gas must travel, thus increasing rates. In traditional carbon-fiber gas diffusion electrodes, which are used with aqueous electrolytes, the carbon fibers are hydrophobic; the hydrophobic repulsion between the electrolyte and carbon fibers, as well as the small spacing (pore size) between the fibers prevents catalyst flooding by the electrolyte. Thus, a well-defined boundary between the gas and liquid close to the catalyst is obtained. Hydrophobic coatings of carbon fibers with nonaqueous electrolytes do not allow for development of a well-defined gas-liquid interface as the interactions between the carbon fiber and electrolyte and no longer unfavorable, and sometimes favorable, leading to flooding of the catalyst and increases in the distance that molecules need to diffuse.
To understand where the liquid-gas interface is relative to the solid surface, one can look at the relative energies of the gaseous and liquid phases under various conditions. First, assume that the electrode is parallel to the gravitational vector; some modification of the analysis can be made to accommodate angled electrodes. Next, find the configurations of gas and liquid that are energetically stable or meta-stable to determine the location of the gas-liquid interface. For these analyses, assume that the system is closed and that the amount of gas or liquid in it is constant. Transitions between states require the gas and liquid to move around, i.e. exchange places. The total energy in the system can be quantified as follows:
Here, P is the pressure at a given location, ρ is the density of the phase at a given location, 9 is the gravitational acceleration constant, y is the vertical location, v is the velocity of the phase at a given location. The below analysis will assume stagnant phases, so v=0. One can note from the outset that the liquid phase is more affected by the gravitational force than the gas phase, while both phases are affected by the pressure. These heuristics can be used to predict qualitatively the configurations of gas-liquid boundaries and their stability. For ease of analysis, one can also plot an average energy as a function of position along the x-axis:
In the simplest case, take a box only half-filled with liquid. The stable configuration is one where the liquid is all below the gas, as it is denser, irrespective of initial configuration, shown in
A traditional hydrophobic gas diffusion electrode with an aqueous electrolyte. In this case, the barrier (wall) is porous, so the water and gas could, in theory, reach the equilibrium state in
The gas similarly cannot enter the liquid phase, as forming the first bubble is also limited by the Laplace pressure:
Therefore there are two energy barriers, one for the liquid, one for the gas, which maintains a meta-stable boundary between the gas and liquid at the GDE, which allows for increased rates. When nonaqueous electrolyte is used, the energy barrier for the liquid phase penetration is broken, which leads to electrolyte penetration, catalyst flooding, and lower rates of gas utilization. In some cases, the capillary action can make the energy inside the electrode lower than outside of it (carbon fibers+THF, for instance), which exacerbates flooding. See,
Pgas−Pliquid>ρgH
This requirement imposes a maximum height of GDE that can be used. The height can be increased by decreasing the pore size of the GDE, which increases the Laplace pressure of the GDE, which in turn increases the highest allowable pressure gradient. Through this analysis, surprisingly, both traditional carbon fiber GDEs and metal mesh GDEs with a pressure gradient can both utilize local energy minima for phase distributions to obtain gas-liquid interfaces close to the electrode surface for increased rates. In carbon fiber GDEs, stability is obtained by a “phantom” pressure gradient from hydrophobic interactions, while in metal mesh GDEs, the pressure gradient can be explicitly applied.
Results
Carbon fiber electrodes for hydrogen oxidation
First, SSCs were examined to improve the rates of the hydrogen oxidation reaction (HOR) in a THF-based electrolyte in order to utilize it as a renewable anode chemistry. Initially, commercially available platinum on carbon fiber (Pt/C) GDEs were used, which are capable of greatly increasing the rate of hydrogen oxidation in aqueous electrolytes (
Stainless steel cloth electrodes for hydrogen oxidation
The fibrous structure of Pt/C GDEs and favorable interactions between the electrolyte and carbon are responsible for flooding of the electrode (
Hydrogen oxidation was confirmed by cyclic voltammetry experiments (
The HOR FE demonstrates robustness to changes in non-zero pressure gradients across the Pt/SSC (
In order to demonstrate the generality of the approach of using SSC as a GDE support for nonaqueous solvents, oxidation of hydrogen at high rates in a 1 M LiBF4 in 9:1 propylene carbonate/dimethyl carbonate electrolyte was attempted. The electrolyte is similar to electrolytes used in some batch lithium-mediated nitrogen reduction approaches. See, for example, Kim, K., Chen, Y., Han, J.-I., Yoon, H. C. & Li, W. Lithium-mediated ammonia synthesis from water and nitrogen: a membrane-free approach enabled by an immiscible aqueous/organic hybrid electrolyte system. Green Chem. (2019) doi:10.1039/C9GC01338E; and Kim, K. et al. Electrochemical Synthesis of Ammonia from Water and Nitrogen: A Lithium-Mediated Approach Using Lithium-Ion Conducting Glass Ceramics. ChemSusChem 11, 120-124 (2018), each of which is incorporated by reference in its entirety. See, for example, each of which is incorporated by reference in its entirety. It was found that the transport-limited hydrogen oxidation current density at flooded Pt foil electrodes is ˜0.25 mA cm−2 (
Stainless steel cloth electrodes for nitrogen reduction
Having overcome transport limitations for hydrogen oxidation by using SSCs as an anode, SSCs for the cathodic reaction of nitrogen reduction was implemented. The lithium metal-mediated approach for N2 reduction has been reported to be diffusion limited in THF on flooded copper and steel foils (
To overcome diffusion limitations, a stainless steel cloth (SSC) was used as the GDE substrate onto which lithium metal was plated in situ. It was found that the rate of the nitrogen reduction reaction (NRR) is significantly enhanced compared to the flooded case (
Nitrogen reduction control experiments
NH3 was confirmed to be produced via N2 reduction by performing control experiments in which argon and isotopically labelled N2 were used as feed gases (
The isotope labeling experiments were performed at two different operating conditions and architectures, which is necessary for conclusive proof of nitrogen reduction. See, for example, Kibsgaard, J., Norskov, J. K. & Chorkendorff, I. The Difficulty of Proving Electrochemical Ammonia Synthesis. ACS Energy Lett. 4, 2986-2988 (2019), which is incorporated by reference in its entirety. The produced NH3 can be found in both the solution and the gas phases (
In the absence of a pressure gradient across the SSC, the system reverts to a flooded state and generally shows poor efficiency for N2 reduction (
Demonstration of electrochemical Haber-Bosch
After obtaining efficient chemistries using SSCs independently at both the cathode and anode, the two reactions were coupled into an electrochemical Haber-Bosch (eHB) reactor, which produces NH3 from N2 and H2 at ambient conditions. Using SSC-based GDEs for both electrodes in a single reactor (
The eHB reactor operates at ambient conditions, which allows it to be operated at smaller scales than traditional Haber-Bosch. However, H2 is usually sourced from steam-methane reforming, which utilizes fossil fuels and is not readily modularized. See, for example, Inc., N. Equipment design and cost estimation for small modular biomass systems, synthesis gas cleanup and oxygen separation equipment. NREL Subcontract report http://www.nrel.gov/docs/fy06osti/39946.pdf (2006), which is incorporated by reference in its entirety. Water electrolysis is a modular alternative H2 source. By coupling an electrochemical Haber-Bosch reactor and a water electrolyzer (
While development of GDEs capable of operating in nonaqueous solvents is imperative for practical electrochemical synthesis, many other aspects of system design require further development. Physical methods to recycle volatile organic solvents and separate the products when using gaseous feedstocks may be necessary in practical systems. Alternatively, bulk volatile solvents may be replaced with non-volatile analogs with similar properties, such as specially tailored polymers or ionic liquids. Polymeric electrolytes may open up avenues for manufacturing species-selective membranes for use in nonaqueous systems, analogous to Nafion in aqueous systems, and for manufacture of membrane-electrode assemblies (MEAs) for gas phase reactions. Electrolyte engineering can also decrease the ionic resistance in the cell, which is important for energy efficiency at high currents.
In the current work, for instance, the energy efficiency for NH3 production ranges from 1.4 to 2.8% at an applied cell potential of 20-30 V, with 70-80% of the energy losses coming from large electrolyte resistance (
This work demonstrates the possibility of utilizing metal cloth-based supports for high rate electrochemical reactions of sparingly soluble gaseous reactants in a nonaqueous solvent. These SSCs were used to produce ammonia from nitrogen and water-derived hydrogen at the highest reported rates at ambient conditions, 30.4±3.5 nmol cm−2 s−1(
Methods
Electrolyte preparation
Electrolyte solutions were prepared by dissolving 1 M of LiBF4 (Sigma-Aldrich, 98+%) in molecular sieve-dried tetrahydrofuran (Acros Organics, 99+%, stabilized with BHT) to which ethanol (VWR International Koptek, anhydrous, 200 proof) was added to yield an ethanol concentration of 0.11 M. The obtained solution was centrifuged at 6000 rpm for 10 minutes to precipitate insoluble impurities. The clear solution was transferred to oven-dried glass vials and used within 12 hours of preparation. In experiments in which hydrogen oxidation is quantified, ferrocene (Alfa Aesar, 99%) is added to the solutions to yield a ferrocene concentration of ˜0.25 M.
Preparation of platinum-coated steel cloths
Stainless steel cloths (McMaster-Carr, 304 stainless steel, 400×400 mesh) were electroplated with nickel followed by platinum (
After striking the cloth with nickel, platinum can be deposited. The platinum plating solution used was a citrate-ammonium bath, containing 35 mM (NH4)2PtCl6 (Alfa Aesar), 400 mM trisodium citrate (anhydrous, Beantown Chemical), and 75 mM of NH4Cl (Alfa Aesar). The nickel-stricken cloth was used as the working electrode; a piece of platinum foil was used as a soluble counter electrode in a beaker cell, which was kept over a water bath at 90° C. A constant reductive current of 10 mA (˜5 mA cmgeom−2) was applied to the cloth for 5 minutes. The platinum-coated cloths (
Gas diffusion electrode experiments
Experiments were performed in 3-compartment cells (
1.75 mL of electrolyte was added to each electrode compartment, for a total electrolyte volume of 3.5 mL. Note that this is the volume of electrolyte added to each compartment and may not be the final electrolyte volume due to electrolyte evaporation. At this point, a pressure gradient was established across the working electrode due to the fact that the electrolyte compartment is at atmospheric pressure while the gas compartment is at positive pressure due to the water column at the compartment outlet; this pressure gradient prevents electrolyte flow into the gas compartment and establishes a robust gas-electrolyte interface. Initially, the height of the water column was chosen to redirect gas flow into the electrolyte compartment. Gas was flowed into the electrolyte for 10 minutes at 10 standard cubic centimeters per minute (sccm) to saturate the solution with the gas. The height of the water column was then lowered to obtain the desired pressure gradient across the SSC while maintaining flow past the SSC. In certain experiments, the flowrate of the gas past the electrode was also decreased at this stage (
For nitrogen reduction experiments, an additional vial containing 2 mL of 0.1 M H3BO3 (Alfa Aesar, 99.99%) was added between the gas compartment and the water column to capture any gas phase ammonia (
For some hydrogen oxidation experiments, the electrolyte contained ferrocene at a concentration of ˜0.25 M, the oxidation of which was used to estimate the FE toward hydrogen oxidation (see Supplementary Discussion). In mass-balance closure and eHB cell experiments, the electrolyte was unchanged from the one used in nitrogen reduction experiments.
Nitrogen reduction control experiments
Nitrogen reduction to ammonia at SSC cathodes was confirmed by varying the feed gas in NRR experiments (
Two architectures were used to confirm nitrogen reduction. In one set of experiments, a 3-compartment cell (
The catholyte from the 3-compartment cell experiments, all the electrolyte from the 4-compartment cell experiments, and the acid trapped were acidified with 0.05 M H2SO4 in water to convert all NH3 to NH4+. NMR spectra of the obtained solutions were taken with solvent suppression on a Bruker Avance Neo 500.18 MHz spectrometer (
Hydrogen oxidation quantification
To quantify the HOR FE, an excess of ferrocene (˜0.25 M) was added to the electrolyte prior to electrolysis. As ferrocene is thermodynamically more difficult to oxidize than H2, but easier than THF (
For experiments at low flowrates and high currents (i.e. at high conversions), the HOR FE was also computed by estimating the hydrogen flowrate out of the gas compartment and by using a hydrogen mass balance over the gas compartment (
Other examples follow. The procedures and materials for nitrogen reduction outlined below have been optimized for reliable ammonia production. However, certain deviations from the procedure and material vendors, which are specifically called out in-text and in prior work, may lead to poorer ammonia production. See, for example, Lazouski, N., Schiffer, Z. J., Williams, K. & Manthiram, K. Understanding Continuous Lithium-Mediated Electrochemical Nitrogen Reduction. Joule 3, 1127-1139 (2019), which is incorporated by reference in its entirety. Follow the procedure closely for reproducibility and high yields.
Electrolyte solution preparation
Dry THF was used as the solvent in THF-based experiments described below. It was obtained by drying as-purchased THF over 20% v/v of freshly dried molecular sieves for at least 48 hours in a round-bottom flask sealed with a rubber septum stopper. The sieves were prepared by washing with acetone and heating at 300° C. for 5 hours in a muffle furnace. The water content of dry THF was found to be 7.1±0.3 ppm (n=3) via Karl-Fischer titration. As-purchased LiBF4, stored in an Ar glovebox, was dissolved in dry THF to obtain electrolyte solutions containing 1 M LiBF4. As discussed in Lazouski et al., it is imperative for the LiBF4 to be pure; LiBF4 purchased from Sigma-Aldrich was found to be sufficiently pure for these experiments, while other vendors' may require purification. Ethanol was added to the solution to obtain a concentration of EtOH of 0.11 M. Insoluble residue was removed from the solutions by centrifugation at 6000 rpm (4430 rcf) for 10 minutes. Clear electrolyte solutions were stored in oven-dried glass vials in a desiccator and used within 12 hours of preparation. While the solutions are somewhat water sensitive, performing centrifugation and solution transfer operations in atmospheric air is permissible, as long as the operations do not expose the solutions to air for long periods of time (hours, i.e. during storage). Oxygen from the atmosphere is typically purged from the electrolyte during saturation by gas flow (see below).
In propylene carbonate-based experiments, solvents and the electrolyte salt were used as received. LiBF4 was dissolved in a 9:1 by volume mixture of propylene carbonate and dimethyl carbonate to produce a 1 M LiBF4 in PC/DMC electrolyte. Dimethyl carbonate was added to the electrolyte in order to reduce the viscosity of the solution, as opposed to using pure propylene carbonate as the solvent. The resulting solution was centrifuged at 6000 rpm (4430 rcf) for 20 minutes to remove insoluble impurities. Clear electrolyte was transferred to oven-dried vials and used within 12 hours of preparation.
Nitrogen reduction experiments—steel foils
Flooded steel foil experiments (
Assembly of gas diffusion cell
Gas diffusion experiments were performed using 3- and 4-compartment cells (
All cell parts, separators, and electrodes were dried in an oven at 80° C. prior to use. Aluminum current collectors for the GDE were made from 0.016″ thick 6061 aluminum; the hole was made with a 7/16″ wood drill bit and sanded to smooth the edges. O-rings between the compartments (
Once the cell is assembled, the inlet of the gas compartment is attached to a bubbler containing THF, through which the gas is bubbled. The outlet is attached to a tube that enters a tall (50 cm) burette containing a water column to control the gauge pressure inside the gas compartment and the pressure gradient across the GDE (
After the cell is set up and gas is flowing through the bubbler at 10 sccm, 1.75 mL of LiBF4/EtOH/THF electrolyte is added to the counter compartment, followed by 1.75 mL of electrolyte to the working compartment. Note that this is the volume of electrolyte added to each compartment and may not be the final electrolyte volume due to electrolyte evaporation. Initially, some of the electrolyte in the working compartment may enter the gas compartment; however, after the entire GDE is contacted by the electrolyte, the pressure in the gas compartment increases, as evidenced by motion of the gas level in the burette, and the electrolyte returns to the working compartment, after which gas begins to flow through the GDE (
In a 4-compartment cell used in combined electrochemical Haber-Bosch (eHB) experiments, the setup is similar to the aforementioned 3-compartment cell. The main change is that the counter electrode is replaced by a second gas compartment and gas diffusion electrode. In summary, the stack becomes:
Nitrogen reduction experiments—steel cloth experiments
Nitrogen reduction reaction (NRR) experiments focused on studying the NRR on steel cloths were performed in 3-compartment cells, the setup of which is detailed in Assembly of a gas diffusion cell. A platinum foil anode and a circular stainless steel cloth (SSC) cathode (diameter=14 mm) were used as the electrodes. The steel cloth electrode was rinsed with DI water and dried at 80° C. prior to use; it was used only for one experiment before discarding. The platinum foil was reused indefinitely. The method of applying current and preparing quantification samples was analogous to the method described in Nitrogen reduction experiments—steel foil experiments, with one change: the catholyte was diluted in a 100 mL volumetric flask due to higher amounts of produced NH3.
One modification that was made in NRR experiments was the addition of a 2 mL 0.1 M boric acid trap between the gas compartment and the pressure-controlling burette to capture any gas phase ammonia. In experiments where gas was fed through the SSC, the trap was inserted at the gas outlet in the catholyte chamber, so all fed gas still went through it. When no pressure gradient across the SSC was applied (
Nitrogen reduction experiments—time evolution of ammonia
Experiments were performed in which the concentration of ammonia in the electrolyte was measured as a function of operating time (
nNH
After the experiment, the cathode was diluted in a 100 mL volumetric flask to quantify the remaining ammonia.
Ammonia quantification—calibration solution preparation A fresh calibration curve was made during each batch of quantifications. The calibration solutions contained a known amount of NH4Cl in Milli-Q water (
15N2 isotope labeling experiments
Isotope labeling experiments were used as a control to confirm N2 reduction to NH3. In order to remove any NH3 and NOx potentially found in 15N2 stock (and house 14N2), the gases were successively passed through solutions of 0.1 M NaOH in water (to capture NOx), 0.05 M H2SO4 in water (to capture NH3), and THF containing activated molecular sieves (to capture water and to saturate the gas with THF) before being fed to the cell. Two isotope labeling experiments were performed: one utilized a typical 3-compartment cells with a steel cloth cathode, Daramic separator, and platinum foil anode with pressure control, while the other utilized an eHB reactor with a steel cloth cathode, Pt/SSC anode, and no separator.
Prior to any experiments, 10 sccm of Ar were bubbled through the entire setup (traps, bubbler, and cell) for 15 minutes to remove trace amounts of 14N2 and other impurities. The desired gas (Ar, 14N2, or 15N2) was fed to the cell and through the SSC at 5 sccm for 10 minutes. In the eHB experiment, H2 was fed through the Pt/SSC at 5 sccm for 10 minutes as well.
In the 3-compartment cell experiment, the pressure gradient across the SSC was decreased to 1 kPa, after which 25 mA of current was applied for 4.8 minutes. The catholyte was removed from the cell into a glass vial. The cathode compartment was rinsed with 0.05 M H2SO4 twice and the resulting fractions were added to the catholyte. The entire mixture was diluted to ˜4 mL with 0.05 M H2SO4. Similarly, the boric acid trap was diluted to a total volume of ˜4 mL with 0.05 M H2SO4. The solution was acidified to convert all the ammonia to ammonium (NH4+) for NMR analysis. See, for example, Nielander, A. C. et al. A Versatile Method for Ammonia Detection in a Range of Relevant Electrolytes via Direct Nuclear Magnetic Resonance Techniques. ACS Catal. 9, 5797-5802 (2019), which is incorporated by reference in its entirety.
In the eHB experiment, the pressure gradient across both electrodes was decreased to 1 kPa, after which the gases were flowed past the electrodes for an additional 2 minutes. 20 mA of current were passed for 6 minutes. The electrolyte was removed from the cell into a 10 mL volumetric flask. The cell was rinsed with 0.05 M H2SO4 and the resulting solutions were added to the volumetric flask and diluted to the mark (10 mL). The boric acid trap was diluted to a total volume of ˜4 mL.
The ammonia content in the resulting solutions was quantified using the salicylate method described above. The electrolyte solutions for quantification were neutralized and diluted 50-fold (40 μL of sample, 10 μL of 0.4 M NaOH, 1950 μL of water) in 3 compartment case or 20-fold (100 μL of sample, 25 μL of 0.4 M NaOH, 1875 μL of water) in the eHB case, while the trap solutions were diluted 10-fold (200 μL of sample, 50 μL of 0.4 M NaOH, 1750 μL of water). NMR spectra of the undiluted solutions were measured on a three-channel Bruker Avance Neo 500.18 MHz spectrometer. Solvent suppression of the largest one (H2O) or three (THF+H2O) peaks was used to increase the signal-to-noise ratio for ammonium peaks. Locking and shimming was done on 1H from water in the solution; no additional compounds were added to the solution prior to NMR. 64 scans were measured for all spectra. The N—H coupling in the NMR spectra confirms ammonia formation from feed N2 (
Preparation of platinum-coated steel cloths
In order to prepare an SSC for hydrogen oxidation, platinum metal, an effective HOR catalyst, was electrodeposited onto the steel cloths. It was found that platinum metal has poor adhesion to stainless steel, which has also been observed in the literature. See, for example, Stoychev, D., Papoutsis, A., Kelaidopoulou, A., Kokkinidis, G. & Milchev, A. Electrodeposition of platinum on metallic and nonmetallic substrates—selection of experimental conditions. Mater. Chem. Phys. 72, 360-365 (2001), which is incorporated by reference in its entirety. In view of this, the steel cloths were first treated by “striking” with nickel. A Wood's nickel strike solution, which consists of 1 M NiCl2 and 1 M HCl in water, was used. Typically, a piece of steel cloth that is 3 cm by 5 cm is taken and submerged to have 2.5 cm by 5 cm in the nickel strike solution. The cloth was used as the working electrode while a piece of nickel foil was used as the soluble counter electrode in an undivided beaker cell. The cloth was pretreated by applying an oxidative current of 15 mA cmgeom−2 for 30 seconds, immediately after which a reductive current of 30 mA cmgeom−2 was applied for 5 minutes obtain a nickel-plated stainless steel cloth. The cloth was thoroughly rinsed with DI water and dried in air at 80° C. The cell potential required for nickel plating was typically ˜1 V. Some of the current went toward hydrogen evolution, evidenced by gas evolution on the cloth; assuming 90% FE toward nickel plating, the resulting nickel layer is approximately 3 μm thick. The cloths visibly change colors after striking with nickel (
After striking the cloth with nickel, platinum can be deposited. The nickel-stricken cloth was cut into smaller pieces to submerge ˜1.5 cm by 1.5 cm into 10 mL of a platinum plating solution. The platinum plating solution used was a citrate-ammonium bath, chosen for its high current efficiency toward platinum plating, low current density required, and the non-hygroscopic nature of the platinum precursor. See, for example, Rao, C. R. K. & Trivedi, D. C. Chemical and electrochemical depositions of platinum group metals and their applications. Coord. Chem. Rev. 249, 613-631 (2005); and Baumgartner & Raub. The Electrodeposition of Platinum and Platinum Alloys. Platin. Met. Rev. 32, 188-197 (1988), each of which is incorporated by reference in its entirety. The bath contained 35 mM (NH4)2PtCl6, 400 mM trisodium citrate, and 75 mM of NH4Cl. The nickel-stricken cloth is used as the working electrode; a piece of platinum foil is used as a soluble counter electrode in a beaker cell, which is kept over a water bath at 90° C. It is possible that the (NH4)2PtCl6 will not fully dissolve until the solution reaches 90° C. A constant reductive current of 10 mA (˜5 mA cmgeom−2) was applied to the cloth for 5 minutes. The cell potential required for platinum plating is typically ˜1.7-1.8 V. If the potential was lower than ˜1.7 V, then a higher current, up to 20 mA, was applied; this is common for fresh baths. The cloth should turn darker after platinum plating (
Ammonia contamination in the cathode compartment from the Pt/SSC anode is unlikely as using Pt foils (in 3-compartment experiments as opposed to Pt/SSC in 4-compartment and undivided experiments) which have not been in contact with the ammonium bath as the anode also leads to production of ammonia in NRR experiments (
Hydrogen oxidation experiments
Hydrogen oxidation experiments focused on studying HOR were performed in 3-compartment cells, the setup of which is detailed in Assembly of a gas diffusion cell. The cell parts and separators were washed with acetone (to remove residual ferrocene) and water and dried in an oven at 80° C. between experiments. A steel foil was used as the cathode, while Pt-coated steel cloths or Pt/C carbon paper disks were used as the anode; both the anode and cathode materials were fresh in every experiment and used only once. 10 sccm of H2 was flowed past the anode after saturating the electrolyte with H2. Regardless of current applied, 7.2 coloumbs of charge were passed though the cell, after which hydrogen oxidation Faradaic efficiency was quantified as described below in Quantification of hydrogen oxidation. In certain experiments (
Quantification of hydrogen oxidation via mass balance
In cases where the flowrate of gas past GDEs is low and the applied current is high, the conversion of the gas may be high (
In Equation 3, Iapplied is the applied current density, FE is the Faradaic efficiency towards the reaction of interest, n is the number of electrons involved in the electrochemical reaction (n=2 for HOR, n=6 for NRR), F is Faraday's constant, Qapplied is the flowrate of gas set using the flow controller, R is the universal gas constant, T0 and P0 are the standard temperature and pressure (273K, 1 bar), respectively, used if the Qapplied is given in units standard volumetric flow units (e.g. sccm).
The conversion can be estimated by measuring the outlet flowrate of gas and by using a mass balance over the gas in the gas compartment, given in Equation 4.
Qin−Qout=ξQin #(4)
From Equations 3 and 4, the Faradaic efficiency of gas conversion can be computed. Implicitly, one can assume that the gas compartment does not have leaks, and that gas dissolution into the electrolyte is negligible. Practically, this process may be used to quantify hydrogen oxidation, as the conversions can be very high.
The flowrate of gas leaving the compartment is difficult to quantify using a flow controller as it is directed to a pressure-controlling water column which releases the gas to the ambient environment and adds additional back pressure which may be difficult to control (
The HOR FE quantification experiments were performed using 3-compartment cells with a Pt/SSC anode, Daramic separator, and steel cathode. The electrolyte used was either 1 M LiBF4/0.11 M EtOH/THF (
It was found that when N2 is fed to the cell, the interval between bubbles does not significantly change for either tested electrolyte when current is applied (
A limitation with the method of mass balancing was found, however. When hydrogen is fed a rate of 0.2 sccm to a 3-compartment cell with a Pt/SSC anode and THF-based electrolyte and 25 mA of current are applied, there should be some gas leaving the gas compartment, as 25 mA corresponds to a hydrogen oxidation rate of at most 0.176 sccm at 100% FE. However, it was observed that no bubbles are evolved and the level of the total amount of gas decreases with time; a constant amount of gas with 25 mA of applied current is obtained when H2 is fed at a rate of 0.22-0.23 sccm. While this demonstrates the limitation of the mass balance method and possible unaccounted for sources of hydrogen depletion, hydrogen is oxidized at high rates in this system with close to unity FE.
eHB experiments and coupling to water splitting
A 4-compartment cell with a steel cloth cathode and Pt/SSC anode was assembled as described in Assembly of a gas diffusion cell. The operation of the 4-compartment cell was similar to operation of the 3-comparment cells in NRR and HOR experiments. 3.5 mL of 1 M LiBF4, 0.11 M EtOH in THF was added to the 4-compartment cell (1.75 mL to each compartment), while 10 sccm of THF-rich N2 and 10 sccm of THF-rich H2 were fed to the cathode and anode compartments, respectively. The solutions were saturated with their respective gases for 10 minutes by flowing gas through the SSCs. The pressure gradient across the SSCs was lowered using a water column to 1 kPa, at which point the gas flowed past the SSC. 25 mA were applied to the cell for 4.8 minutes, after which the ammonia content of the cathode chamber was analyzed as described in Nitrogen reduction experiments.
In long term experiments (
To couple the eHB to water splitting, a commercially available water-splitting cell (Fuel Cell Technologies) was assembled. The electrodes were part of a membrane electrode assembly (MEA) purchased from FuelCellStore with an electrode area of 5 cm2. The cathode side was platinum black with a loading of 3 mg cm−2; the anode side was iridium ruthenium oxide with a loading of 3 mg cm−2; both electrodes were on a Nafion 115 membrane. The bolts on the electrolyzer were tightened with a torque wrench with a torque of 40 lb-in.
Milli-Q water was fed continuously to the anode of the water splitting cell at ˜70 mL/min with a peristaltic pump. A constant current of 200 mA was applied across the electrolyzer; the voltage required was 1.59 V. This corresponds to an output H2 flowrate of 1.5 sccm. The cell was slightly angled to help oxygen bubbles to leave the anode compartment. The cathode compartment was sealed off at one end to force hydrogen to flow in a single direction. The hydrogen was first fed to a vial containing magnesium sulfate (MgSO4) to capture some of the moisture in the gas stream, after which it was fed to a vial with THF and molecular sieves to saturate the gas with THF. The cell was then operated analogously to the way an eHB cell was, with the difference that the feed rate of H2 was 1.5 sccm, as defined by the water splitting current.
DISCUSSIONComputing the diffusion-limited current density for H2 oxidation
One can estimate the diffusion-limited current density for H2 oxidation and find that it is fairly close to the value one can obtain via direct measurement (
The solubility of hydrogen in pure THF is 3.3-3.4 mM.9,10 See, for example, Gibanel, F., López, M. C., Royo, F. M., Santafé, J. & Urieta, J. S. Solubility of nonpolar gases in tetrahydrofuran at 0 to 30° C. and 101.33 kPa partial pressure of gas. J. Solution Chem. 22, 211-217 (1993); and Brunner, E. Solubility of Hydrogen in 10 Organic Solvents at 298.15, 323.15, and 373.15 K. J. Chem. Eng. Data 30, 269-273 (1985), each of which is incorporated by reference in its entirety. However, the properties of the solvent change with addition of large amounts of electrolyte, leading to a “salting-out” effect, decreasing the solubility of the gas. See, for example, Weisenberger, S. & Schumpe, A. Estimation of Gas Solubilities in Salt Solutions at Temperatures from 273 K to 363 K. AIChE J. 42, 298-300 (1996), which is incorporated by reference in its entirety. One can estimate that the solubility of hydrogen in the electrolyte is close to half of its pure solvent solubility, approximately 1.7±0.8 mM, following Lazouski et al. The diffusivity of hydrogen in the electrolyte is also an estimate, computed by using an approximate value of the viscosity of solution, and assumed to be 3.8±0.8-10−9 m2 s−1. The diffusion boundary layer thickness was previously measured and found to be 50±15 μm;1 corrections to the diffusion boundary layer thickness due to differences in diffusion coefficients of various species are not used due to the already large uncertainties in estimates of other parameters. Combining these assumptions, the estimated diffusion-limited current density for H2 oxidation is 2.5±1.5 mA cm−2, which is fairly close to the experimentally measured value (˜2.75 mA cm−2,
Development of the gas-liquid interface across the SSC
Initially, the gas compartment and electrolyte compartments contain no electrolyte and are separated by a vertically standing SSC or CC GDE. When gas is fed to the gas compartment, it predominantly leaves through the GDE if any resistance to flow (by means of water column or otherwise) is applied to the outlet of the gas compartment. When a small volume of electrolyte (880 uL) is added, the SSC and CC GDEs behave differently.
In the case of the CC GDE, the electrolyte completely stays within the electrolyte compartment and wets the GDE, and gas begins to flow out through the gas compartment outlet if there is not sufficient pressure to force the gas through the GDE. Adding additional electrolyte (for a total of 1.75 mL) does not qualitatively change the picture.
In the case of a SSC GDE, the electrolyte actually goes through the vertically standing GDE into the gas compartment (
Computing the diffusion-limited current density of N2 reduction
The diffusion-limited current density for nitrogen reduction in a 1 M LiBF4 in THF electrolyte has been estimated in Lazouski et al. By following a procedure similar to the one outlined above and in the aforementioned work, one can estimate the diffusion-limited current density in an aqueous electrolyte at a flooded electrode by using Equation 7.
The diffusion coefficient and solubility of nitrogen in pure water are well known. At 25° C. and 1 bar of N2 partial pressure, the diffusivity of N2 in water is 2.01±0.1·10−9 m2 s−1, while the solubility is 0.66 mM. See, for example, Ferrell, R. T. & Himmelblau, D. M. Diffusion coefficients of nitrogen and oxygen in water. J. Chem. Eng. Data 12, 111-115 (1967); and Battino, R., Rettich, T. R. & Tominaga, T. The Solubility of Nitrogen and Air in Liquids. J. Phys. Chem. Ref Data 13, 563-600 (1984), each of which is incorporated by reference in its entirety. The diffusion boundary layer thickness depends heavily on the hydrodynamics of the electrolyte; typical values for CO2 reduction in an aqueous electrolyte are 60-160 μm; the boundary layer thickness may be thinner is well-defined and vigorous hydrodynamics are observed, such as in systems utilizing rotating disk electrodes (RDEs). See, for example, Weng, L. C., Bell, A. T. & Weber, A. Z. Modeling gas-diffusion electrodes for CO2 reduction. Phys. Chem. Chem. Phys. 20, 16973-16984 (2018), which is incorporated by reference in its entirety. From these values, it was found that the diffusion limited current density for nitrogen reduction in water is a mere 0.48-1.27 mA cm−2.
Source of ammonia in two phases
It was found that produced NH3 can be found in both the solution and the gas phases (
The lack of stripping of ammonia from the bulk solution is likely due to strong interactions between Li+ and NH3, stronger than between Li+ and THF. See, for example, Kaufmann, E., Gose, J. & Schleyer, P. v. R. Thermodynamics of solvation of lithium compounds. A combined MNDO and ab initio study. Organometallics 8, 2577-2584 (1989), which is incorporated by reference in its entirety. The formation Li—NH3 complexes in solution increases the solubility of ammonia in the electrolyte (
Estimating energy efficiency and consumption
A detailed description for estimating the energy efficiency of the process and sources of energy losses has been described in Lazouski et al. Briefly, the energy efficiency is computed as follows by the formula given in Equation 8.
In Equation 8, UNH
In Equation 9, F is Faraday's constant (96485 C/mol), Vtotal is the total applied cell voltage, FE is the ammonia Faradaic efficiency, and MNH
References, each of which is incorporated by reference in its entirety.
- 1. Lazouski, N., Schiffer, Z. J., Williams, K. & Manthiram, K. Understanding Continuous Lithium-Mediated Electrochemical Nitrogen Reduction. Joule 3, 1127-1139 (2019).
- 2. Nielander, A. C. et al. A Versatile Method for Ammonia Detection in a Range of Relevant Electrolytes via Direct Nuclear Magnetic Resonance Techniques. ACS Catal. 9, 5797-5802 (2019).
- 3. Stoychev, D., Papoutsis, A., Kelaidopoulou, A., Kokkinidis, G. & Milchev, A. Electrodeposition of platinum on metallic and nonmetallic substrates—selection of experimental conditions. Mater. Chem. Phys. 72, 360-365 (2001).
- 4. Rao, C. R. K. & Trivedi, D. C. Chemical and electrochemical depositions of platinum group metals and their applications. Coord. Chem. Rev. 249, 613-631 (2005).
- 5. Baumgartner & Raub. The Electrodeposition of Platinum and Platinum Alloys. Platin. Met. Rev. 32, 188-197 (1988).
- 6. Gagne, R. R., Koval, C. A. & Lisensky, G. C. Ferrocene as an Internal Standard for Electrochemical Measurements. Inorg. Chem. 19, 2854-2855 (1980).
- 7. Singh, A., Chowdhury, D. R. & Paul, A. A kinetic study of ferrocenium cation decomposition utilizing an integrated electrochemical methodology composed of cyclic voltammetry and amperometry. Analyst 139, 5747-5754 (2014).
- 8. Bard, A. J. & Faulkner, L. R. Electrochemical Methods. Fundamentals and Applications. (John Wiley & Sons, Inc, 2001).
- 9. Gibanel, F., López, M. C., Royo, F. M., Santafé, J. & Urieta, J. S. Solubility of nonpolar gases in tetrahydrofuran at 0 to 30° C. and 101.33 kPa partial pressure of gas. J. Solution Chem. 22, 211-217 (1993).
- 10. Brunner, E. Solubility of Hydrogen in 10 Organic Solvents at 298.15, 323.15, and 373.15 K. J. Chem. Eng. Data 30, 269-273 (1985).
- 11. Weisenberger, S. & Schumpe, A. Estimation of Gas Solubilities in Salt Solutions at Temperatures from 273 K to 363 K. AIChE J. 42, 298-300 (1996).
- 12. Ferrell, R. T. & Himmelblau, D. M. Diffusion coefficients of nitrogen and oxygen in water. J. Chem. Eng. Data 12, 111-115 (1967).
- 13. Battino, R., Rettich, T. R. & Tominaga, T. The Solubility of Nitrogen and Air in Liquids. J. Phys. Chem. Ref Data 13, 563-600 (1984).
- 14. Weng, L. C., Bell, A. T. & Weber, A. Z. Modeling gas-diffusion electrodes for CO2 reduction. Phys. Chem. Chem. Phys. 20, 16973-16984 (2018).
- 15. Kaufmann, E., Gose, J. & Schleyer, P. v. R. Thermodynamics of solvation of lithium compounds. A combined MNDO and ab initio study. Organometallics 8, 2577-2584 (1989).
- 16. Lobaccaro, P. et al. Effects of temperature and gas-liquid mass transfer on the operation of small electrochemical cells for the quantitative evaluation of CO2 reduction electrocatalysts. Phys. Chem. Chem. Phys. 18, 26777-26785 (2016).
17. Lee, H. K. et al. Favoring the unfavored: Selective electrochemical nitrogen fixation using a reticular chemistry approach. Sci. Adv. 4, eaar3208 (2018).
- 18. Zhou, F. et al. Electro-synthesis of ammonia from nitrogen at ambient temperature and pressure in ionic liquids. Energy Environ. Sci. 10, 2516-2520 (2017).
- 19. Tsuneto, A., Kudo, A. & Sakata, T. Lithium-mediated electrochemical reduction of high pressure N2 to NH3. J. Electroanal. Chem. 367, 183-188 (1994).
- 20. Andersen, S. Z. et al. A rigorous electrochemical ammonia synthesis protocol with quantitative isotope measurements. Nature 570, 504-508 (2019).
- 21. Kim, K., Yoo, C., Kim, J., Yoon, H. C. & Han, J. Electrochemical Synthesis of Ammonia from Water and Nitrogen in Ethylenediamine under Ambient Temperature and Pressure. J. Electrochem. Soc. 163, F1523-F1526 (2016).
- 22. Kim, K. et al. Communication—Electrochemical Reduction of Nitrogen to Ammonia in 2-Propanol under Ambient Temperature and Pressure. J. Electrochem. Soc. 163, F610-F612 (2016).
- 23. Köleli, F. & Röpke, T. Electrochemical hydrogenation of dinitrogen to ammonia on a polyaniline electrode. Appl. Catal. B Environ. 62, 306-310 (2006).
- 24. Pappenfus, T. M., Lee, K., Thoma, L. M. & Dukart, C. R. Wind to Ammonia: Electrochemical Processes in Room Temperature Ionic Liquids. in ECS Transactions vol. 16 89-93 (ECS, 2009).
Another example of a standalone electrode architecture is depicted in
Referring to
Referring to
Referring to
An example is further described here.
Stand-alone electrode for utilizing sparingly soluble gases
Metallic meshes and other porous materials as gas diffusion electrodes can generally utilize sparingly soluble gases in electrochemical reactions in aqueous and nonaqueous electrolytes. However, in some applications, the gas diffusion electrodes are utilized in a custom, parallel-electrode architecture. While the architecture is efficient and convenient for both testing and synthetic applications, there may be applications that require utilization of sparingly soluble gases in other architectures. Examples include rapid resting of reactions in easy-to-setup beaker cells, synthetic reactions which require large volumes of solvent, and those which use gases in counter and balancing reactions. For these and other applications, a stand-alone electrode is described herein that uses the nonaqueous GDEs for utilizing sparingly soluble gases in electrochemical reactions.
The basic standalone architecture can be seen in
There are a number of possible configurations of the standalone electrode. In particular, the electrode may be in a “flow in” configuration, where the gas is fed through one inlet and used up by the reactions occurring at the electrode (
One of the key requirements for invention operation is controlling the pressure gradient across the GDE. The location and configuration of the pressure control differ somewhat in the three gas flow configurations. In the “flow-through” configuration, the GDE itself establishes the necessary pressure gradients, and simply flowing gas into the gas compartment at a high enough rate and/or positive pressure is sufficient.
In the “flow-past” configuration, the pressure in the gas compartment is controlled after the gas leaves standalone electrode (
In the “flow-in” configuration, the pressure in the gas compartment is controlled prior to entering electrode (
Demonstration of standalone electrode use
A flow-in standalone electrode with a platinum-coated stainless-steel cloth GDE was assembled to demonstrate the capability of the electrode to utilize gases in electrochemical reactions in various solvents. The standalone electrode was used as the anode, to which either nitrogen or hydrogen gas was fed. A platinum foil was used as the cathode and an Ag/AgCl electrode was used as a reference. Two electrolyte compositions were used sequentially with the same electrode to demonstrate the solvent-agnostic nature of the electrode: 0.1 M tetrabutylammonium tetrafluoroborate with 0.05 M hexafluoroisopropyl alcohol in acetonitrile was used as a nonaqueous electrolyte, while a 0.05 M Na2SO4 in water solution was used as an aqueous electrolyte. Either nitrogen or hydrogen gas was first flowed into the electrode in a flow-through configuration to fill the gas compartment of the electrode and saturate the electrolyte solution, after which the pressure in the electrode gas compartment was decreased to put the electrode in a flow-in configuration. A linear-sweep voltammogram was measured by sweeping the potential from open circuit voltage to a high oxidation potential, first with nitrogen in the gas compartment, followed by having hydrogen in the compartment. A significantly higher current was obtained when using hydrogen as the feed gas when compared to using nitrogen, which demonstrates that hydrogen oxidation is occurring. The currents obtained significantly exceeded the diffusion limited hydrogen oxidation current, demonstrating the use of the standalone electrode as a gas diffusion electrode. Results of the exemplary reaction are shown in
Details of one or more embodiments are set forth in the accompanying drawings and description. Other features, objects, and advantages will be apparent from the description, drawings, and claims. Although a number of embodiments of the invention have been described, it will be understood that various modifications may be made without departing from the spirit and scope of the invention. It should also be understood that the appended drawings are not necessarily to scale, presenting a somewhat simplified representation of various features and basic principles of the invention.
Claims
1. An electrochemical system comprising:
- a housing including a chamber;
- an electrode within the housing; and
- a gas permeable metal on a surface of the electrode in contact with the chamber.
2. The system of claim 1, further comprising a gas inlet to the housing.
3. The system of claim 1, further comprising a first outlet of the housing to release a product from the housing.
4. The system of claim 1, wherein the gas permeable metal includes a metal mesh.
5. The system of claim 4, wherein the metal mesh includes 100, 200, 300, 400, 500, 1000, 1500, or 2000 fibers per inch.
6. The system of claim 1, wherein the gas permeable metal includes openings of between 1 and 200 micrometers, preferably between 2 and 100 micrometers.
7. The system of claim 1, wherein the gas permeable metal includes metal fibers or a porous metal.
8. The system of claim 1, wherein the gas permeable metal includes stainless steel, steel, nickel, iron, copper, silver, gold, or platinum.
9. The system of claim 1, wherein the gas permeable metal includes a catalytic metal, metal oxide, metal sulfide, or metal phosphide.
10. The system of claim 1, wherein gas permeable metal is exposed to a pressure gradient.
11. A method of supplying a gas to an electrochemical system comprising:
- contacting a gas with a gas permeable metal on a surface of an electrode in a chamber of a housing.
12. The method of claim 11, wherein the gas is a sparingly soluble gas.
13. The method of claim 11, further comprising supplying a pressure of the gas in the chamber to create a pressure differential at the electrode.
14. The method of claim 11, further comprising applying a voltage to the electrode.
15. The method of claim 11, wherein the gas permeable metal includes a metal mesh.
16. The method of claim 15, wherein the metal mesh includes 100, 200, 300, 400, 500, 1000, 1500, or 2000 fibers per inch.
17. The method of claim 11, wherein the gas permeable metal includes openings of between 1 and 200 micrometers, preferably between 2 and 100 micrometers.
18. The method of claim 11, wherein the gas permeable metal includes metal fibers or a porous metal.
19. The method of claim 11, wherein the gas permeable metal includes stainless steel, steel, nickel, iron, copper, silver, gold, or platinum.
20. A method of oxidizing or reducing a gas comprising:
- contacting a gas with a gas permeable metal on a surface of an electrode.
21. The method of claim 20, wherein the gas is a sparingly soluble gas.
22. The method of claim 20, wherein the gas is hydrogen.
23. The method of claim 20, wherein the gas is nitrogen.
24. The method of claim 23, wherein the ammonia is produced at a Faradaic yield of at least 30% or at least 40%.
25. The method of claim 20, wherein supplying a pressure of the gas in the chamber to create a pressure differential at the electrode.
26. An electrochemical system comprising:
- a first electrode including: a housing including a chamber; an electrode within the housing; and a gas permeable metal on a surface of the electrode in contact with the chamber; and
- a second electrode including a gas inlet to a housing including a gas permeable metal on a surface of an electrode and a first outlet to release a product from the system.
27. The system of claim 26, wherein each gas permeable metal includes a metal mesh.
28. The system of claim 27, wherein each metal mesh includes 100, 200, 300, 400, 500, 1000, 1500, or 2000 fibers per inch.
29. The system of claim 26, wherein at least one gas permeable metal includes openings of between 1 and 200 micrometers, preferably between 2 and 100 micrometers.
30. The system of claim 26, wherein at least one gas permeable metal includes metal fibers or a porous metal.
31. The system of claim 26, wherein each gas permeable metal is exposed to a pressure gradient.
Type: Application
Filed: Mar 24, 2021
Publication Date: Sep 30, 2021
Applicant: Massachusetts Institute of Technology (Cambridge, MA)
Inventors: Karthish Manthiram (Cambridge, MA), Nikifar Lazouski (Boston, MA)
Application Number: 17/211,235