AMMONIA-ASSISTED CO2 CAPTURING AND UPGRADING TO VALUABLE CHEMICALS

Information

  • Patent Application
  • 20240175144
  • Publication Number
    20240175144
  • Date Filed
    October 11, 2023
    a year ago
  • Date Published
    May 30, 2024
    6 months ago
  • CPC
    • C25B3/07
    • C25B3/26
    • C25B11/031
    • C25B11/065
    • C25B11/089
  • International Classifications
    • C25B3/07
    • C25B3/26
    • C25B11/031
    • C25B11/065
    • C25B11/089
Abstract
The present disclosure relates to an electrochemical method for converting captured CO2 into formate (HCOO−). This method involves capturing waste CO2 by co-absorption of the waste CO2 with green ammonia (NH3) to form ammonium bicarbonate (NH4HCO3) and converting the ammonium bicarbonate (NH4HCO3) into formate (HCOO−), wherein said converting is carried out in an integrated flow electrolyzer system. Another aspect of the present disclosure relates to an integrated flow electrolyzer system comprising an alkaline electrolyzer for producing green NH3 from NO3−, an NH3—CO2 absorbing unit whereby waste CO2 is co-absorbed with ammonia (NH3) to form ammonium bicarbonate (NH4HCO3), and a bicarbonate electrolyzer for converting the ammonium bicarbonate (NH4HCO3) into formate (HCOO−).
Description

This application claims the benefit of U.S. Provisional Patent Application Ser. No. 63/415,154, filed Oct. 11, 2022, which is hereby incorporated by reference in its entirety.


This invention was made with government support under grant number 2021-67021-34650 awarded by USDA/NIFA and grant number CHE2036944 awarded by National Science Foundation. The government has certain rights in the invention.


FIELD OF THE INVENTION

Disclosed herein are methods for ammonia-assisted capturing and upgrading of CO2 to valuable chemicals and electrolyzers for carrying out such methods.


BACKGROUND OF THE INVENTION

Reduction of the net CO2 emissions to zero by 2025 is an urgent need for limiting the global warming to a safe level (Nitopi et al., “Progress and Perspectives of Electrochemical CO2 Reduction on Copper in Aqueous Electrolyte,” Chem. Rev. 119(12): 7610-7672 (2019); Pachauri et al., “Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change,” Ipcc: (2014); Sullivan et al., “Coupling Electrochemical CO2 Conversion with CO2 Capture,” Nat. Catal. 4(11):952-958 (2021); Rogelj et al., “A New Scenario Logic for the Paris Agreement Long-term Temperature Goal,” Nature 573(7774):357-363 (2019)). Powered by renewable electricity from solar or wind sources, CO2 can be electrochemically converted into valuable chemicals and fuels (i.e., the CO2 reduction reaction, or CO2RR) (Nitopi et al., “Progress and Perspectives of Electrochemical CO2 Reduction on Copper in Aqueous Electrolyte,” Chem. Rev. 119(12): 7610-7672 (2019); De Luna et al., “What Would it Take for Renewably Powered Electrosynthesis to Displace Petrochemical Processes?,” Science 364(6438): eaav3506 (2019)). However, CO2 electrolyzers in most studies are fed with pressurized and purified CO2 gas, production of which requires energy- and capital-intensive regeneration processes from the captured CO2 (Keith et al., “A Process for Capturing CO2 From the Atmosphere.” Joule 2(8): 1573-1594 (2018); Welch et al., “Bicarbonate or Carbonate Processes for Coupling Carbon Dioxide Capture and Electrochemical Conversion,” ACS Energy Lett. 5(3):940-945 (2020)). Specifically, after capturing CO2 from air or flue gases, the release of CO2 is usually accomplished by heating the capturing media at 120-150° C. (Boot-Handford et al., “Carbon Capture and Storage Update,” Energy Environ. Sci 7(1): 130-189 (2014); Haszeldine, R. S., “Carbon Capture and Storage: How Green Can Black Be?,” Science 325(5948): 1647-1652 (2009)). Then, the collected CO2 must be compressed into a pressurized container for storage and transportation before its utilization. Therefore, integrating CO2 capture and conversion steps is vital to decreasing the energy costs and making the overall process sustainable (Welch et al., “Bicarbonate or Carbonate Processes for Coupling Carbon Dioxide Capture and Electrochemical Conversion,” ACS Energy Lett. 5(3):940-945 (2020); Pérez-Gallent et al., “Integrating CO2 Capture with Electrochemical Conversion Using Amine-Based Capture Solvents as Electrolytes,” Ind. Eng. Chem. Res. 60(11):4269-4278 (2021)).





KOH(aq)+CO2(g)→KHCO3(aq)   (RXN 1)





2RNH2(aq)+CO2(g)→RNHCOO(aq)+RNH3 (aq)   (RXN 2)





CO2(g)+H2O(l)+NH3(g)→NH4HCO3(aq)   (RXN 3)


One strategy is to convert CO2 directly in its capture solutions upon its in situ release (Welch et al., “Bicarbonate or Carbonate Processes for Coupling Carbon Dioxide Capture and Electrochemical Conversion,” ACS Energy Lett. 5(3):940-945 (2020)). In bipolar membrane (“BPM”)-based electrolyzers, aqueous bicarbonate—generated by absorbing CO2 in KOH capture solution (RXN 1)—can react with H+ produced by the BPM to form in situ CO2 (“i-CO2”) at the BPM-electrode interface (RXN 4), which can be subsequently converted into value-added products such as CO (Lees et al., “Electrodes Designed for Converting Bicarbonate Into CO,” ACS Energy Lett. 5(7):2165-2173 (2020); Zhang et al., “Porous Metal Electrodes Enable Efficient Electrolysis of Carbon Capture Solutions,” Energy Environ. Sci. 15:705-713 (2022); Lees et al., “Continuum Model to Define the Chemistry and Mass Transfer in a Bicarbonate Electrolyzer,” ACS Energy Lett. 7:834-842 (2022); Yang et al., “Cation-Driven Increases of CO2 Utilization in a Bipolar Membrane Electrode Assembly for CO2 Electrolysis,” ACS Energy Lett. 6(12):4291-4298 (2021); Blommaert et al., “Insights and Challenges for Applying Bipolar Membranes in Advanced Electrochemical Energy Systems,” ACS Energy Lett. 6(7):2539-2548 (2021)), formate (Li et al., “Conversion of Bicarbonate to Formate in an Electrochemical Flow Reactor,” ACS Energy Lett. 5(8):2624-2630 (2020)), and CH4 (Lees et al., “Electrolytic Methane Production from Reactive Carbon Solutions,” ACS Energy Lett. 7:1712-1718 (2022)). However, the performance of such bicarbonate electrolyzers is inferior to direct CO2 electrolyzers fed with gaseous CO2, largely due to the inadequate local i-CO2 concentration (Kas et al., “Modeling the Local Environment within Porous Electrode during Electrochemical Reduction of Bicarbonate,” Ind. Eng. Chem. Res., 61(29): 10461-10473 (2022); Kim et al., “Electrocatalytic Reduction of Low Concentrations of CO2 Gas in a Membrane Electrode Assembly Electrolyzer,” ACS Energy Lett. 6(10):3488-3495 (2021)). Besides, the metal cation bicarbonate electrolyzer has significantly higher electrical energy consumption because the BPM requires an additional potential of 0.828 V (under standard conditions) for H+ generation by water dissociation upon large current densities (Blommaert et al., “Insights and Challenges for Applying Bipolar Membranes in Advanced Electrochemical Energy Systems,” ACS Energy Lett. 6(7): 2539-2548 (2021)), and BPM is thicker than conventional cation exchange membranes (“CEM”) and anion exchange membranes (“AEM”).


Apart from the KOH solution, amines such as monoethanolamine (“MEA”) solution are commonly used for capturing CO2 due to the facile kinetics of the formation of amine-CO2 adducts (RXN 2) (Lee et al., “Electrochemical Upgrade of CO2 From Amine Capture Solution,” Nat. Energy 6(1):46-53 (2021); Chen et al., “Electrochemical Reduction of Carbon Dioxide in a Monoethanolamine Capture Medium,” (hemSusChem 10(20):4109-4118 (2017); Rochelle, G. T., “Amine Scrubbing for CO2 Capture,” Science 325(5948): 1652-1654 (2009)). Conversion of MEA-CO2 adduct to CO in the BPM-based electrolyzers has been reported in previous studies but only at low current densities (<50 mA cm−2) (Lee et al., “Electrochemical Upgrade of CO2 From Amine Capture Solution,” Nat. Energy 6(1):46-53 (2021); Chen et al., “Electrochemical Reduction of Carbon Dioxide in a Monoethanolamine Capture Medium,” ChemSusChem 10(20):4109-4118 (2017); Kim et al., “Insensitive Cation Effect on Single-atom Ni Catalyst Allows Selective Electrochemical Conversion of Captured CO2 in Universal Media,” Energy & Environmental Science 15(10):DOI:10.1039/D2EE01825J (2022)). In addition to the insufficient i-CO2 concentration, the bulky carbamate (RNHCOO) and ethanolammonium (RNH3+) ions may hinder the mass transport at the electrode double layer, which limits the i-CO2RR performance.


As a suitable alternative, CO2 can be captured by ammonia (NH3) solution to form ammonium bicarbonate (NH4HCO3) (RXN 3) (Kim et al., “Insensitive Cation Effect on Single-atom Ni Catalyst Allows Selective Electrochemical Conversion of Captured CO2 in Universal Media,” Energy & Environmental Science 15(10):DOI:10.1039/D2EE01825J (2022); Zhao et al., “Post-combustion CO2 Capture by Aqueous Ammonia: A State-of-the-art Review,” Int. J. Greenh. Gas Control. 9:355-371 (2012)). Compared to MEA, NH3 has higher CO2-capturing capacity because it doubles the stoichiometric ratio of CO2 to the capturing agent (see RXN 2 and 3) (Shakerian et al., “A Comparative Review Between Amines and Ammonia as Sorptive Media for Post-combustion CO2 Capture,” Appl. Energy 148:10-22 (2015)). More importantly, release of CO2 from NH4HCO3 requires much lower energy than that from MEA-CO2 or KHCO3, as illustrated by their decomposition temperatures: 36° C., 120° C., and 150° C. for NH4HCO3, MEA-CO2, and KHCO3, respectively (Wang et al., “Current Status and Challenges of the Ammonia Escape Inhibition Technologies in Ammonia-based CO2 Capture Process,” Appl. Energy 230:734-749 (2018)). The ease of CO2 release from NH4HCO3 (RXN 5) is expected to provide sufficient i-CO2 in the electrolyzer and facilitate i-CO2RR with reduced energy input. Besides, the cost of NH3 (13.5 USD per kmol) is much lower than that of KOH (86 USD per kmol) or MEA (92 USD per kmol), and NH3 as the CO2 capture agent is less corrosive and less toxic as compared to KOH (Wang et al., “Current Status and Challenges of the Ammonia Escape Inhibition Technologies in Ammonia-based CO2 Capture Process,” Appl. Energy 230:734-749 (2018)). Moreover, NH3 can be readily produced by the electro-reduction of NOx or NOx that are abundant in agricultural or industrial wastes (Liu et al., “Electrocatalytic Nitrate Reduction on Oxide-derived Silver with Tunable Selectivity to Nitrite and Ammonia,” ACS Catal. 11(14):8431-8442 (2021); Daiyan et al., “Nitrate Reduction to Ammonium: From CuO Defect Engineering to Waste NOx-to-NH3 Economic Feasibility,” Energy Environ. Sci. 14(6):3588-3598 (2021); Kwon et al., “Nitric Oxide Utilization for Ammonia Production Using Solid Electrolysis Cell at Atmospheric Pressure,” ACS Energy Lett. 6(12):4165-4172 (2021); Kim et al., “Unveiling Electrode-Electrolyte Design-Based NO Reduction for NH3 Synthesis,” ACS Energy Lett. 5(11):3647-3656 (2020)). Using waste-derived NH3 for CO2 capture and conversion will simultaneously alleviate the environmental burdens of NOx/NOx and CO2 by “fixing” them into stable and value-added chemical products (FIG. 1) (van Langevelde et al., “Electrocatalytic Nitrate Reduction for Sustainable Ammonia Production,” Joule 5(2):290-294 (2021); Wang et al., “Nitrate Electroreduction: Mechanism Insight, In situ Characterization, Performance Evaluation, and Challenges,” Chem. Soc. Rev. 50(12):6720-6733 (2021)).


Conventional work: water dissociation driven in BPM electrolyzer





HCO3(aq)+H+(aq)→H2O(l)+i-CO2(g)   (RXN 4)


One known obstacle to using NH3 as the capture agent is its volatility, and it may escape from NH3 solution during capturing operation and handling (Wang et al., “Current Status and Challenges of the Ammonia Escape Inhibition Technologies in Ammonia-based CO2 Capture Process,” Appl. Energy 230:734-749 (2018)). Through a series of system modeling and optimization, such as the advanced flash stripper process (Jiang et al., “Advancement of Ammonia Based Post-combustion CO2 Capture Using the Advanced Flash Stripper Process,” Appl. Energy 202:496-506 (2017); Jiang et al., “Advancement of Ammonia-based Post-combustion CO2 Capture Technology: Process Modifications,” Fuel Processing Technology 210:106544 (2020)), both operational and economic feasibility have been demonstrated for NH3-based CO2 capture (chilled NH3 process) (Hanak et al., “Efficiency Improvements for the Coal-fired Power Plant Retrofit With CO2 Capture Plant Using Chilled Ammonia Process,” Appl. Energy 151:258-272 (2015); Li et al., “Technoeconomic Assessment of an Advanced Aqueous Ammonia-based Postcombustion Capture Process Integrated With a 650-MW Coal-fired Power Station,” Environ. Sci. Technol. 50(19): 10746-10755 (2016)) from post-combustion streams. For example, a proposed two-stage adsorption process has reduced NH3 slip by more than 50%, leading to the recovery of >99% of vaporized NH3 (Li et al., “Technical and Energy Performance of an Advanced, Aqueous Ammonia-based CO2 Capture Technology for a 500 MW Coal-fired Power Station,” Environ. Sci. Technol. 49(16): 10243-10252 (2015)). The CO2-avoided cost has been reduced to US$ 40.7/ton CO2 for NH3-based CO2 capture, which is 44% less than the MEA-based process (US$ 75.1/ton CO2) (Jiang et al., “Advancement of Ammonia Based Post-combustion CO2 Capture Using the Advanced Flash Stripper Process,” Appl. Energy 202:496-506 (2017)). Although modeling has shown the success of the capability of using NH3 as the capture agent, the utilization of CO2 captured media—the direct conversion of NH4HCO3 solution in the electrolyzers—has never been developed. Besides, the previous studies were limited to theoretically modeling the technical and economic capability of NH3-based CO2 capture. Experimentally developing a combined process for NH3-mediated CO2 capture (with NH3 derived waste resources) and its direct utilization (conversion in electrolyzers) is still lacking.


The present disclosure is directed to overcoming limitations in the art.


SUMMARY OF THE INVENTION

One aspect of the present disclosure relates to an electrochemical method for converting captured CO2 into formate (HCOO). This method involves capturing waste CO2 by co-absorption of the waste CO2 with green ammonia (NH3) to form ammonium bicarbonate (NH4HCO3) and converting the ammonium bicarbonate (NH4HCO3) into formate (HCOO), wherein said converting is carried out in an integrated flow electrolyzer system.


Another aspect of the present disclosure relates to an integrated flow electrolyzer system comprising an alkaline electrolyzer for producing green NH3 from NO3, an NH3—CO2 absorbing unit whereby waste CO2 is co-absorbed with ammonia (NH3) to form ammonium bicarbonate (NH4HCO3), and a bicarbonate electrolyzer for converting the ammonium bicarbonate (NH4HCO3) into formate (HCOO).


The present disclosure involves an electrochemical process and integrated flow system for converting captured CO2 into formate (HCOO), as illustrated in FIG. 1 and FIG. 2, involving thermal decomposition driven in an AEM/CEM electrolyzer, according to the following reaction:





NH4HCO3(aq)→NH4+(aq)+i-CO2(g)+OH(aq)





i-CO2(g)+H2O+2e→HCOO+HO  (RXN 5).


Direct electrochemical conversion of CO2 capture solutions (instead of gaseous CO2) into valuable chemicals can circumvent the energy-intensive CO2 regeneration and pressurization steps, but the performance of such processes is limited by the sluggish release of CO2 and the use of energy-consuming bipolar membranes (BPMs). It has been unexpectedly discovered that an ammonium bicarbonate (NH4HCO3)-fed electrolyzer outperforms the state-of-the-art KHCO3 electrolyzer owing to its favorable thermal decomposition property, which allows for a 3-fold increase of the in-situ CO2 concentration, a 23% increase in formate faradaic efficiency, and a 35% reduction in cell voltage by substituting BPM with an anion exchange membrane. An integrated process of combining NH4HCO3 electrolysis with CO2 capturing by on-site generated ammonia from the electro-reduction of nitrate is demonstrated, which features a remarkable 99.8% utilization of CO2 capturing agent. Such a multi-purpose process offers a sustainable route for the simultaneous removal of N wastes and streamlined CO2 capturing-upgrading.





BRIEF DESCRIPTION OF THE DRAWINGS


FIG. 1 shows a schematic illustration of one embodiment of an integrated process of CO2 capture and utilization disclosed herein. This process combines green NH3 production from NO3, waste CO2 capture by co-absorption with NH3, and production of value-added commodity chemicals from the bicarbonate electrolyzers. In the NO3 electrolyzer, the air was used as the carrier gas to purge NH3-containing gases out of the reactor for its further capture of CO2 in water: acid-base reaction between NH3 and CO2 (NH3+CO2+H2O→NH4HCO3).



FIG. 2 shows a schematic illustration of one embodiment of an integrated process of CO2 capture and utilization disclosed herein. This process combines green NH3 production from NO3, waste CO2 capture by co-absorption with NH3, and production of value-added commodity chemicals from the bicarbonate electrolyzers.



FIG. 3 is a photograph showing an experimental system for the 5-hour electrolysis of NH4HCO3 in an AEM-based electrolyzer.



FIGS. 4A-C show the characterization of ED-Bi. FIG. 4A is an SEM image (scale bar=50 μm), FIG. 4B is an SEM image with higher magnification (scale bar=20 μm), and FIG. 4C shows the XRD pattern of ED-Bi.



FIGS. 5A-C show electrochemical conversion of CO2 capture solutions to formate. FIG. 5A is a graph showing Faradaic efficiency (left Y-axis) of products and the amount of in situ generated CO2 (right Y-axis) during the electrolysis of different CO2 capture solutions (2.5 M for all cases): KHCO3, NH4HCO3, and MEA-CO2. The electrolysis was performed at 40° C.for 30 min at a constant current density of 100 mA cm−2. The MEA-CO2 adduct was prepared by bubbling CO2 for 30 min into the solution of MEA, then purging argon to remove dissolved CO2. 2.5 M KCl was added to the MEA-CO2 solution to increase its conductivity (Lee et al., “Electrochemical Upgrade of CO2 From Amine Capture Solution,” Nat. Energy 6(1):46-53 (2021), which is hereby incorporated by reference in its entirety). FIGS. 5B-C are graphs providing a comparison of formate faradaic efficiency with KHCO3 and NH4HCO3 at different current densities (FIG. 5B) and different cell temperatures (FIG. 5C) during 30-min electrolysis. The catholyte volume was 40 mL for 100-150 mA cm−2 and 120 mL for 200-300 mA cm−2.



FIGS. 6A-B are graphs showing a comparison of H2 faradaic efficiency between KHCO3 and NH4HCO3 at different current densities (FIG. 6A) and different cell temperatures (FIG. 6B) for BPM-based electrolysis for half-hour operation. The catholyte volume was 120 mL for 200-300 mA cm−2 and 40 mL for 100-150 mA cm−2.



FIGS. 7A-F show conversion of NH4HCO3 in the electrolyzers with different membranes. FIG. 7A is a graph showing the cell voltage profile, and FIG. 7B is a graph showing faradaic efficiency (left Y-axis) of products and the amount of in situ generated CO2 (right Y-axis) for the electrolysis of KHCO3 and NH4HCO3 with different membranes: BPM, CEM, and AEM. The electrolysis was performed at 40° C.for 30 min at a constant current density of 100 mA cm−2. FIG. 7C provides a graphical comparison of the cell voltage and the FE towards product (formate) in this work with the reported performances (Lees et al., “Electrodes Designed for Converting Bicarbonate Into CO,” ACS Energy Lett. 5(7):2165-2173 (2020); Zhang et al., “Porous Metal Electrodes Enable Efficient Electrolysis of Carbon Capture Solutions,” Energy Environ. Sci. 15:705-713 (2022); Li et al., “Conversion of Bicarbonate to Formate in an Electrochemical Flow Reactor,” ACS Energy Lett. 5(8):2624-2630 (2020); Li et al., “CO2 Electroreduction From Carbonate Electrolyte,” ACS Energy Lett. 4(6): 1427-1431 (2019); which are hereby incorporated by reference in their entirety) using KHCO3 and BPM at 100 mA cm−2. FIG. 7D is a schematic illustration of the local environment at the cathode for the electrolyzers with NH4HCO3 feed when using different membranes. FIG. 7E is a graph showing cell voltage—time profile and formate faradaic efficiency (both catholyte and anolyte) for 5-hour electrolysis in AEM-based system at 40° C. FIG. 7F is a graph showing the amount of NH4+ monitored in the NH4HCO3 solution and the NH3-absorbing solution during 5-hour electrolysis.



FIG. 8 is an SEM image of ED-Bi after electrolysis.



FIG. 9 is a graph showing an analysis of the energy consumption for formic acid production from CO2 electro-reduction with different cell configurations. Case 1-i and 1-ii correspond to the electrolyzers with gaseous CO2 feed, and case 2-i to 2-v correspond to the electrolyzers fed with CO2 capture solutions.



FIG. 10 is a photograph showing an experimental system for the production of NH4HCO3 from CO2 and NO3-derived NH3.



FIG. 11 is a graph showing the cell voltage profile of NH3 production by NO3 reduction in the alkaline electrolyzer for 6-hour electrolysis.



FIGS. 12A-C show an integrated process combining CO2 capturing by NO3 derived NH3 and formate production from NH4HCO3. FIG. 12A is a graph showing utilization of the reaction ingredients for each individual process, including 1) electron utilization (faradaic efficiency) of the electrochemical NO3-to-NH3 reaction; 2) utilization of the on-site generated CO2 capturing agent (NH3) for NH4HCO3 production; 3) electron utilization (faradaic efficiency) of the electrochemical formate production from the as-prepared NH4HCO3. The NH4HCO3 solution for electrolysis was prepared by dissolving 7.9 g of the NH4HCO3 (obtained by the reaction between NO3-derived NH3 and CO2) in 40 mL of deionized water. FIGS. 12B-C provide a graphical comparison of XRD patterns (FIG. 12B), and 1H NMR spectrum (FIG. 12C) of the as-prepared NH4HCO3 and the commercial product.





DETAILED DESCRIPTION OF THE INVENTION

The present disclosure relates to electrolysis methods and systems for converting captured CO2 into formate (HCOO). Electrolysis is a process where electrical current is used to drive a non-spontaneous redox reaction.


One aspect of the present disclosure relates to an electrochemical method for converting captured CO2 into formate (HCOO). This method involves capturing waste CO2 by co-absorption of the waste CO2 with ammonia (NH3) to form ammonium bicarbonate (NH4HCO3) and converting the ammonium bicarbonate (NH4HCO3) into formate (HCOO), wherein said converting is carried out in an integrated flow electrolyzer system.


Another aspect of the present disclosure relates to an integrated flow electrolyzer system comprising an alkaline electrolyzer for producing NH3 from NO3, an NH3—CO2 absorbing unit whereby waste CO2 is co-absorbed with ammonia (NH3) to form ammonium bicarbonate (NH4HCO3), and a bicarbonate electrolyzer for converting the ammonium bicarbonate (NH4HCO3) into formate (HCOO).


In some embodiments, the ammonia (NH3) co-absorbed with waste CO2 in the methods and systems of the present disclosure is green ammonia (NH3). As used herein, “green” ammonia (NH3) is NH3 produced from waste reactive nitrogen, or from reduction of N2 by green H2. In some embodiments, the NH3 can be readily produced by the electro-reduction of NOx or NOx that are abundant in agricultural or industrial wastes.


In the electrochemical methods and systems disclosed herein, the capture of waste CO2 by co-absorption may be carried out, for example and without limitation, through an integrated electricity-driven process for economically upcycling waste nitrogen enabled by low-concentration NO3 electrodialysis and high-performance NH3 electrosynthesis from various reactive nitrogen (Nr) forms. For example, an integrated electricity-driven process may comprise the following three core components: (i) NO3 recovery from low-concentration waste streams by electrodialysis, (ii) Nr-to-NH3 conversion by electrolysis, and (iii) formation of NH3 and NH3-based chemicals; and two optional components as the logical extension: (iv) direct NH3 fuel cell, and (v) NH3-mediated bicarbonate electrolysis, which is discussed in a U.S. provisional patent application entitled “A Membrane-Free Alkaline Electrolyzer for Upcycling Waste Into Ammonia,” filed the same day as the present disclosure, which is hereby incorporated by reference in its entirety. In a membrane-free alkaline electrolyzer (MFAEL) with NaOH/KOH/H2O as the robust electrolyte, a record-high NH3 partial current density of 4.22±0.25 A cm−2 from NO3 reduction with a faradaic efficiency (FE) of 84.5±4.9% may be achieved on a simple commercial nickel foam as the cathode material. Continuous production of pure NH3-based chemicals (NH3 solution and solid NH4HCO3) was realized by collecting NH3 by water and CO2-saturated solutions, respectively, without the need for additional separation procedures. Low energy consumption can recover NO3 from low concentration (100 ppm NO3—N) by efficient electrodialysis. Such an integrated process offers an all-sustainable and economically viable route for upcycling waste Nr into the highest-demanded N-based chemical product —NH3, so that the growing trend of Nr buildup could be largely decelerated and reversed.


In some embodiments, the capturing is carried out according to the following formula:





CO2+H2O+NH3→NH4HCO3.


In some embodiments, the converting the ammonium bicarbonate (NH4HCO3) into formate (HCOO) is carried out in an integrated flow electrolyzer system with an anion exchange membrane to avoid using a bipolar membrane.


The term “electrolyzer,” as used herein, refers to an apparatus for performing electrolysis. An electrolyzer typically has a pair of electrodes (e.g., an anode and a cathode), a reaction medium (e.g., an electrolyte solution), and a power supply, which is typically an external source of power to add electrical energy to a reaction taking place in the reaction medium. The electrolyzer may or may not have a membrane dividing or separating the anode and the cathode. The electrodes facilitate the transfer of electrical energy into the reaction medium by extending into the reaction medium at one end and connecting to an external power supply at the other end. In the integrated flow electrolyzer system described herein, two or more electrolyzer units may be combined together with other units (e.g., a co-absorption unit to capture waste CO2 by co-absorption of the waste CO2 with ammonia (NH3) to form ammonium bicarbonate (NH4HCO3)) to provide an integrated system for carrying out a series of chemical reactions.


One embodiment of an integrated flow electrolyzer system comprising multiple units is illustrated in FIG. 1. In the particular embodiment illustrated in FIG. 1, an integrated process and system for CO2 capture and utilization combines NH3 production from NO3 in a membrane-free alkaline electrolyzer, waste CO2 capture by co-absorption with NH3 in a co-absorber unit, and production of value-added commodity chemicals (formate) from a membrane-containing bicarbonate electrolyzer.


As illustrated in FIG. 1, integrated flow electrolyzer system 10 comprises alkaline electrolyzer 12 for producing NH3 from NO3, which comprises reaction chamber 14 formed by walls 16 and lid 18, pair of electrodes 20A and 20B, which extend into reaction medium or electrolyte solution 22, and each is shown connected to power supply 24 (i.e., anode, positive electrode 20A connected to the positive terminal “+” of the power supply and cathode, negative electrode 20B connected to the negative terminal “−” of the power supply) and extending into reaction medium 22. Current from power supply 24 enters reaction medium 22 through anode 20A and leaves reaction medium 22 through cathode 20B. Air may enter electrolyzer 12 through intake 32, and NH3 mixed gases may exit electrolyzer 12 through outtake 34.


In some embodiments, anode 20A and cathode 20B comprise nickel wire mesh, although other conductive materials and/or metals in other forms (e.g., metal foam) may also be used for the electrodes of the alkaline electrolyzer.


In some embodiments, the alkaline electrolyzer is a membrane-free alkaline electrolyzer. In some embodiments, the alkaline electrolyzer comprises an ion exchange membrane or other permeable membrane barrier.


In some embodiments, the integrated flow electrolyzer system comprises an alkaline electrolyzer, such as membrane-free alkaline electrolyzer 12 shown in FIG. 1 for producing NH3 from NO3. In an alkaline electrolyzer, the reaction medium or electrolyte solution typically includes sodium hydroxide and/or potassium hydroxide and water. In some embodiments, the reaction medium comprises a concentrated NaOH—KOH solution (as an electrolyte) and electrodes made of commercial nickel wire mesh, although other reaction mediums and electrode materials may also be suitable, such as concentration of NaOH and KOH solution, and porous Raney nickel.


In some embodiments, chemical reactions occurring in the processes and systems described herein are carried out at elevated temperatures. For example, reaction mediums may be heated to a temperature above room temperature, or to temperatures of between 35-80° C., or any temperature or range of temperatures therein. In some embodiments, heating is carried out by applying a heat source to a particular reaction or reaction unit, such as an electrolyzer. With reference again to FIG. 1, heat source 128 is shown underneath electrolyzer 12 to heat reaction chamber 14 and reaction medium 22. In some embodiments, the reaction medium is heated to a temperature of 35-40° C., 35-45° C., 35-50° C., 35-55° C., 35-60° C., 35-65° C., 35-70° C., 35-75° C., 35-80° C., 40-45° C., 40-50° C., 40-55° C., 40-60° C., 40-65° C., 40-70° C., 40-75° C., 40-80° C., 45-50° C., 45-55° C., 45-60° C., 45-65° C., 45-70° C., 45-75° C., 45-80° C., 50-55° C., 50-60° C., 50-65° C., 50-70° C., 50-75° C., 50-80° C., 55-60° C., 55-65° C., 55-70° C., 55-75° C., 55-80° C., 60-65° C., 60-70° C., 60-75° C., 60-80° C., 65-70° C., 65-75° C., 65-80° C., 70-75° C., 70-80° C., or 75-80° C. In some embodiments, the reaction medium is heated to a temperature of 80° C. or more.


In some embodiments, the use of high alkalinity reaction medium and elevated temperature (e.g., 80° C.) in an alkaline electrolyzer in the integrated flow electrolyzer system of the present disclosure facilitates simultaneous NH3 production and separation. Other conditions and reaction mediums can also be used to achieve the production of ammonium (NH3).


In some embodiments, NO3 reduction is carried out in a membrane-free alkaline electrolyzer of the integrated flow electrolyzer system of the present disclosure and is performed under a continuous flow of air. In some embodiments, the air in the continuous flow of air is N2. In some embodiments, a continuous flow of air carries the produced NH3 into a CO2-saturated water solution, illustrated in FIG. 1 and discussed below as absorbing unit 200. In some embodiments, produced NH3 is cooled at absorbing unit 200 to a temperature of 5° C., or 0-10° C., or any temperature or range therein suitable for NH4HCO3 formation.


Referring again to FIG. 1, integrated flow electrolyzer system 10 also includes a means of waste CO2 capture by co-absorption with NH3 from electrolyzer 12. This occurs in absorbing unit 200, where NH3 from electrolyzer 12 flows from electrolyzer 12 through conduit 30 and is combined with waste CO2 in co-absorber medium 232. As used herein, a “co-absorber” absorbs acidic CO2 gas and basic NH3 gas. In some embodiments, the co-absorber is water. In some embodiments, the co-absorber is saline water.


In the embodiment of integrated flow electrolyzer system 10 illustrated in FIG. 1, value-added commodity chemicals, such as formate, are derived from bicarbonate electrolyzer 112. In electrolyzer 12 of integrated flow electrolyzer system 10, air is used as a carrier gas to purge NH3-containing gases out of reaction chamber 22 for its further capture of CO2 in water in absorbing unit 200. Acid-base reaction between NH3 and CO2 (NH3+CO2+H2O→NH4HCO3) occurs in absorbing unit 200.


A third unit shown in integrated system 10 of FIG. 1 is bicarbonate electrolyzer 112, which is a membrane-containing electrolyzer comprising anode 120A, cathode 120B, and exchange membrane 126. Exchange membrane 126 is shown separating the two electrodes 120A and 120B. In some embodiments, the exchange membrane is an anion exchange membrane. An anion exchange membrane is a semipermeable membrane generally made from ionomers and designed to conduct anions while being impermeable to cations and gases such as oxygen or hydrogen. Anion exchange membranes are used in electrolytic cells to separate reactants present around the two electrodes while transporting the anions essential for the electrolyzer operation. In some embodiments, the exchange membrane is a cation exchange membrane. A cation exchange membrane is a semipermeable membrane generally made from ionomers and designed to conduct cations while being impermeable to anions and gasses such as oxygen or hydrogen. Cation exchange membranes are used in electrolytic cells and fuel cells to separate reactants present around the two electrodes while transporting the cations essential for the cell operation from the cathode to the anode.


In the embodiment illustrated in FIG. 2, integrated flow electrolyzer system 10 (FIG. 1) comprises AEM-based bicarbonate electrolyzer 112, which comprises flow field plates 114A and 114B with serpentine flow channels 122 (also the reaction medium) embedded within field plates 114A and 114B allowing the flow of reaction medium 122, a pair of electrodes 120A and 120B, placed over flow channels 122 in direct contact with field plates 114A and 114B. Each electrode 120A and 120B of the pair is shown connected to power supply 124 (i.e., anode, positive electrode 120A connected to the positive terminal “+” of the power supply and cathode, negative electrode 120B connected to the negative terminal “−” of the power supply) and in contact with reaction medium 122, where electrodes 120A and 120B are separated by anion exchange membrane 126. Current from power supply 124 enters reaction medium 122 through anode 120A and exits reaction medium 122 through cathode 120B.


The above disclosure is general. A more specific description is provided below in the following examples. The examples are described solely for the purpose of illustration and are not intended to limit the scope of the present disclosure. Changes in the form and substitution of equivalents are contemplated as circumstances suggest or render expedient. Although specific terms have been employed herein, such terms are intended in a descriptive sense and not for the purposes of limitation.


EXAMPLES

The following examples are provided to illustrate embodiments of the present application but are by no means intended to limit scope.


Example 1—Ammonia-Mediated CO2 Capture and Direct Electroreduction into Formate Materials and Methods
Preparation of Electrodes

The electrodeposited-Bi (ED-Bi) was prepared in a two-electrode system in a one-compartment electrochemical cell by a modified method from literature (Li et al., “Conversion of Bicarbonate to Formate in an Electrochemical Flow Reactor.” ACS Energy Lett. 5(8):2624-2630 (2020), which is hereby incorporated by reference in its entirety). The aqueous Bi3+ precursor was prepared by adding 1.5 mmol of Bi(NO3)3·5H2O into 40 mL of deionized water. Concentrated HNO3 (5 mL) was added to the solution in order to fully dissolve the Bi precursor. The electrodeposition was conducted at a constant current of 72 mA for 5 min. A piece of carbon paper (3×3 cm2, Freudenberg H23) and Pt foil were immersed in the electrolyte as cathode and anode, respectively.


Materials Characterization

X-ray diffraction (XRD) crystallography was collected with a Siemens D500 diffractometer operated with a Copper K-α source (λ=1.5418 Å) at 45 kV and 30 mA and equipped with a diffracted beam monochromator (carbon). Scanning Electron Microscopy (SEM) was conducted on a field-emission scanning electron microscope (FEI Quanta-250) equipped with a light-element X-ray detector and an Oxford Aztec energy-dispersive X-ray analysis system.


Electrochemical Measurements
Electrochemical Conversion of CO2 Captured Solutions in the Flow Cell

The flow electrolyzer contains two flow-field plates with serpentine channels, PTFE and silicone gaskets, and the membrane electrode assembly, which contains two electrodes and a membrane, and was formed after assembling the cell hardware. The anode (2.5×2.5 cm2) and cathode (2.0×2.0 cm2) flow plates were made from titanium and stainless steel, respectively. The catholyte and anolyte were circulated by a peristaltic pump (Masterflex® L/S®) at 50 mL min−1. A piece of Ni foam (MIT corporation, 80-110 pores per Inch, average hole diameters about 0.25 mm) with geometric area of 6.25 cm2 (2.5×2.5 cm2) and 40 mL of 1.0 M KOH were used as the anode and anolyte, respectively. The prepared ED-Bi on carbon paper and a 2.5 M of CO2 capturing solution (i.e., NH4HCO3, KHCO3, or MEA-CO2) were used as the cathode and catholyte, respectively. The volume of catholyte was 40 mL and 120 mL for the current density of 100-150 mA cm−2 and 200-300 mA cm−2, respectively. A piece of bipolar membrane (Fumatech FBM), anion-exchange membrane (Tokuyama A201), or cation exchange membrane (Nafion115) was used as the ion exchange membrane. Argon was purged into the headspace of the catholyte for the on-line collection and off-line quantification of gaseous products (CO, H2, and CO2). The temperature of the flow cell was controlled by a 50-watt 110 V heater (Dioxide Materials).


The 5-hour electrolysis was performed in a similar flow cell set-up with an anion-exchange membrane (Tokuyama A201). The photo of the experimental set-up is shown in FIG. 3. The volumes of catholyte and anolyte were 500 and 200 mL, respectively. The catholyte was directly connected to an NH3 absorbing solution (0.2 M H2SO4, 200 mL). The formate concentration in both catholyte and anolyte was quantified by NMR spectroscopy at each hour interval, and the total absorbed NH4+ in the absorbing solution was quantified after electrolysis.


Production of NH4HCO3 from CO2 and nitrate (NO3)-derived NH3


The experimental setup for NH4HCO3 production comprises two major components connected in tandem (FIG. 3): an electrolyzer for NH3 production by NO3 electro-reduction, and an NH4HCO3 formation unit for the reaction between NH3 and CO2. The configuration of the one-compartment NH3-producing electrolyzer was modified from our previous work (Chen et al., “Revealing Nitrogen-containing Species in Commercial Catalysts Used for Ammonia Electrosynthesis.” Nat. Catal. 3(12): 1055-1061 (2020), which is hereby incorporated by reference in its entirety). In brief, the cell body or reaction chamber comprised a 100 mL screw-cap polytetrafluoroethylene (PTFE) bottle and a custom-made stainless-steel lid. The electrolyte contained 29.7 g of NaOH, 48.1 g of KOH, and 38.9 g of deionized water (with 60 wt. % base with equimolar NaOH and KOH, and 40 wt. % of H2O). 139.9 mmol of KNO3 was added as the reactant. The cell was kept at 80° C.in an oil bath. Two 10 cm2 nickel mesh electrodes (3.3×3 cm2, 200 mesh) attached to nickel wires (0.04″ diameter) were used as the cathode and anode, and a flow of N2 (200 mL min−1) was bubbled into the electrolyte to carry the produced NH3 into the NH4HCO3 formation unit, which comprised 100 mL of CO2-saturated deionized water cooled to 5° C.and magnetically stirred at 400 r.p.m. To maintain the saturation of CO2, 500 mL min−1 of CO2 was bubbled into the NH4HCO3 formation unit during the experiment. Electrolysis was carried out at a constant current density of 500 mA cm−2 for 6 hours; under these conditions, the applied charge was equal to the theoretical amount of charge required to fully reduce NO3 in the system to NH3. To determine the utilization of NH3 in the NH4HCO3 formation unit, the outlet gas from the NH4HCO3 formation unit was bubbled into an acidic solution (100 mL of 0.5 M H2SO4) to collect any unused NH3.


To obtain solid NH4HCO3 sample for further i-CO2RR experiments, a similar configuration with larger electrolyzer volume (2.5 L) was used with 25 times the quantity of all chemicals for electrolyte preparation. Other conditions include: 2.8 mol of added KNO3 as the reactant, 100 cm2 of the electrode area, 500 mL min−1 of the N2 flow rate, 250 mA cm−2 of the applied current density, and 24 hours of the electrolysis duration. Due to the low solubility of NH4HCO3 (1.81 mol per liter of water at 5° C.), solid was precipitated in the NH4HCO3 formation unit, which was separated by vacuum filtration. The effective NH4HCO3 content of the collected sample was determined by dissolving a certain amount of the sample in deionized water, followed by measuring its NH4+ concentration (detailed in the following section).


Product Analysis
Quantification of Soluble Products in the Electrolyte

Formate was quantified by ion chromatography (IC, Thermo Scientific Dionex Easion). 50 or 100 μL of the sample solution was diluted with deionized water and injected into IC for its quantification.


NO3 and NO2 were analyzed by High-Performance Liquid Chromatography (HPLC (Chou et al., “A High Performance Liquid Chromatography Method for Determining Nitrate and Nitrite Levels in Vegetables,” Journal of Food and Drug Analysis 11(3):233-238 (2003), which is hereby incorporated by reference in its entirety)) (Agilent Technologies, 1260 Infinity II LC System) equipped with a variable wavelength detector (Agilent 1260 Infinity Variable Wavelength Detector VL). The wavelength of 213 nm was used for detection. A C18 HPLC column (Gemini® 3 μm, 110Å, 100×3 mm) was used for analysis at 25° C. with a binary gradient pumping method to drive mobile phase at 0.4 mL min−1. The mobile phase consisted of 0.01 M n-octylamine (for ion pairing) in a mixed solution containing 30 vol % of methanol and 70 vol % of DI water, and the pH of the mobile phase was adjusted to 7.0 with H3PO4. The running time was 30 min for each sample, and the retention time for NO3 and NO2 was around 18 and 16 min, respectively. The calibration solutions for NO3 or NO2 were prepared with KNO3 and KNO2 in the concentration range of 0.0625-2 mM.


For the electrochemical NO3 -to-NH3 reaction, the conversion of NO3(X) and faradaic efficiency of product i (FEi for NH3 and NO2) were calculated by






X
=




n
0

-
n


n
0


×
100

%









FE


i

=




n
i

·

z
i

·
F

Q

×
1

0

0

%





where n0 is the initial amount of NO3 (mol); n is the amount of NO3 after electrolysis (mol); ni is the amount of product i (mol); zi is the number of electrons transferred to product i; F is the Faraday constant (96,485 C mol−1); Q is the total charge passed through the electrolytic cell (C).


NH4+ content was determined by 1H Nuclear Magnetic Resonance (NMR) spectroscopy on a Bruker Avance NEO 400 MHZ NMR spectrometer. The sample solution was first diluted with 0.1 M H2SO4 to the proper range of NH3 concentration. 800 μL of the diluted sample solution was then mixed with 200 μL of DMSO-d6 and 200 μL of 32 μM maleic acid (internal standard) in DMSO-d6. The scan number was 1,024 with a water suppression method. Standard NH3 solutions were prepared for calibration with concentrations ranging from 0 to 5 mg L−1 (in N).


Quantification of Gas Products

The gas products were analyzed by an off-line gas chromatography (GC, SRI Instruments, 8610C, Multiple Gas #3) equipped with HayeSep D and MolSieve 5Å columns. A thermal conductivity detector was used to detect H2, and a flame ionization detector was used to detect CO and CO2. The calibration curves for H2 (10-10,000 ppm, Cal Gas Direct), CO (110-8,000 ppm, Cal Gas Direct), and CO2 (5,000 -50,000 ppm, Cal Gas Direct) were established by analyzing the calibration gases.


In the electrolysis of CO2-capturing solutions at 5 min, 15 min, and 25 min, the outlet of the electrolyzer was connected to a standard FlexFoil sample bag (1 L, SKC INC) for on-line collection of gas products. Then, the collected gases were injected into GC for their off-line analysis. A 12-min GC program was applied. The rate of gas generation (r, mol s−1) was calculated by






r
=

c
×
1


0

-
6


×


p


V
.

×

10

-
6


/
60


RT









where c is the gas content (ppm); {dot over (V)} is the volumetric flow rate of the inlet gas to the sample bags (300 mL min−1); p is the atmospheric pressure (1.013×105 Pa); R is the gas constant (8.314 J mol−1 K−1); T is the room temperature (20° C. or 293.15 K). The total amount of gas production (mol) was calculated by integrating the plot of gas production rate (mol s−1) vs. reaction time(s) with a linear fitting.


Results and Discussion

The i-CO2RR performance in a BPM-based electrolyzer with NH4HCO3 as the electrolyte and reactant was first investigated. Formate was chosen as the target i-CO2RR product in this study due to its known economic feasibility (De Luna et al., “What Would it Take for Renewably Powered Electrosynthesis to Displace Petrochemical Processes?,” Science 364(6438): eaav3506 (2019; Shin et al., “Techno-economic Assessment of Low-temperature Carbon Dioxide Electrolysis,” Nat. Sustain. 4(10):911-919 (2021); which are hereby incorporated by reference in their entirety). Besides, recent work has proposed that ammonium formate is a safe, and energy-dense electrochemical fuel ionic liquid with an energy density of 3.2 kWh/L, higher than that of formic acid and hydrogen (Schiffer et al., “Ammonium Formate as a Safe, Energy-Dense Electrochemical Fuel Ionic Liquid,” ACS Energy Lett. 7:3260-3267 (2022), which is hereby incorporated by reference in its entirety). Electro-deposited bismuth (ED-Bi) on a carbon paper substrate was used as the i-CO2RR catalyst (Li et al., “Conversion of Bicarbonate to Formate in an Electrochemical Flow Reactor,” ACS Energy Lett. 5(8):2624-2630 (2020), which is hereby incorporated by reference in its entirety). Scanning electron microscopy (SEM) images and X-ray diffraction (XRD) pattern show the uniform deposition of metallic Bi on carbon paper (FIGS. 4A-C). The electrolyzer with a zero-gap configuration included the ED-Bi cathode, a BPM, and a piece of nickel foam as the anode. The catholyte and anolyte were supplied to their respective electrode compartment at an identical flow rate of 50 mL min−1.



FIG. 5A shows the comparison of i-CO2RR performance at 40° C.with three kinds of CO2-capturing solutions (2.5 M for all cases): NH4HCO3, KHCO3, and MEA-CO2. At 100 mA cm−2, the faradaic efficiency (FE) of formate production was 75+3% for NH4HCO3, much higher than that of KHCO3 (59±7%) and MEA-CO2 (18±3%). The main side reaction was the hydrogen evolution reaction (HER), and the FE towards CO production was less than 2%. Interestingly, the trend of FE towards formate is consistent with the amount of generated i-CO2, which was determined by summing the amounts of CO2 (evolved from the catholyte and detected by gas chromatography) and i-CO2RR products generated during the electrolysis. The total i-CO2 produced from NH4HCO3 was 3.4 or 10.8 times that of KHCO3 or MEA-CO2, respectively. The poor i-CO2RR performance with MEA-CO2 could be attributed to the bulky ions of MEA-CO2 that inhibit i-CO2 formation and transportation (Kim et al., “Insensitive Cation Effect on Single-atom Ni Catalyst Allows Selective Electrochemical Conversion of Captured CO2 in Universal Media,” Energy & Environmental Science 15(10):DOI:10.1039/D2EE01825J (2022), which is hereby incorporated by reference in its entirety). Furthermore, by comparing the results with the corresponding control cells without applying current (Table 1), the portion of produced i-CO2 from thermal decomposition was found to be 93% for NH4HCO3, much higher than that for KHCO3 (52%), highlighting the ease of CO2 release from NH4HCO3.









TABLE 1







The source of i-CO2 in the BPM-based electrolyzers with KHCO3 or


NH4HCO3 at cell temperature of 40° C. for half-hour operation.












i-CO2 from
i-CO2 from BPM-




thermo-
induced chemical



Total i-CO2
decomposition
decomposition


System
(mmol)a
(mmol)b
(mmol)c













KHCO3
4.2
2.2
2.0


NH4HCO3
14.1
12.5
1.6






aTotal i-CO2 values were determined by summing the amounts of both CO2 and i-CO2RR products generated from the cathode during the half-hour electrolysis.




bThe values of i-CO2 from thermo-decomposition were obtained from the same BPM-based electrolyzer without applying any current




cThe i-CO2 values from BPM-induced chemical decomposition were obtained from subtracting the “i-CO2 from thermo-decomposition” from the “total i-CO2”.







The i-CO2RR performances of the systems with NH4HCO3 and KHCO3 were further compared at different current densities ranging from 100 to 300 mA cm−2 (FIG. 5B) and cell temperatures from 25 to 60° C.(FIG. 5C). The system with NH4HCO3 shows 10-20% higher formate-oriented FE than KHCO3 under all tested conditions. Limited by the mass transport limit of i-CO2, the remaining FE attributed to the competing H2 (major) and CO (minor, <2%) evolutions, which exhibited an opposite trend as compared to formate formation (FIGS. 6A-B and Table 2). At 100 mA cm−2, the highest formate FE was 90% with NH4HCO3 at 60° C., which is higher than the reported values in the bicarbonate electrolyzers for formate production (Li et al., “Conversion of Bicarbonate to Formate in an Electrochemical Flow Reactor,” ACS Energy Lett. 5(8):2624-2630 (2020); Min and Kanan, “Pd-catalyzed Electrohydrogenation of Carbon Dioxide to Formate: High Mass Activity at Low Overpotential and Identification of the Deactivation Pathway,” J. Am. Chem. Soc. 137(14):4701-4708 (2015); which are hereby incorporated by reference in their entirety).









TABLE 2







Summary of the faradaic efficiencies of i-CO2 reduction


products in the BPM-based electrolyzers with KHCO3 or NH4


HCO3 at different reaction conditions.
















FE





FE towards
FE towards
towards
Total



Conditions
H2 (%)
formate (%)
CO (%)
(%)
















KHCO3
100 mA/cm2
37.11
59.50
1.96
98.57


(40° C.)
150 mA/cm2
36.02
60.00
1.74
97.76



200 mA/cm2
43.64
56.64
1.81
102.09



250 mA/cm2
48.04
51.88
1.57
101.49



300 mA/cm2
52.32
46.93
1.50
100.75



100 mA/cm2
18.98
79.49
3.45
101.92


NH4HCO3
150 mA/cm2
29.96
69.33
2.76
102.05


(40° C.)
200 mA/cm2
31.84
69.83
3.00
104.67



250 mA/cm2
41.00
59.16
2.06
102.22



300 mA/cm2
44.50
54.81
2.39
101.70


KHCO3
RT (25° C.)
48.18
54.99
2.06
105.23


(100 mA
50° C.
22.10
74.50
2.79
99.39


cm−2)
60° C.
11.20
82.11
3.36
96.67


NH4HCO3
RT (25° C.)
18.37
76.92
2.60
97.89


(100 mA
50° C.
15.26
82.58
3.72
101.56


cm−2)
60° C.
9.81
89.00
4.08
102.89









Since the thermal decomposition dominates the production of i-CO2 from NH4HCO3 (compared to the minor contribution from BPM), it was further sought to remove the energy-consuming BPM in the electrolyzer to reduce the cell voltage. As shown in FIG. 7A, substituting BPM with CEM or AEM indeed resulted in a decrease in the cell voltage from 3.7 V to 2.4 V. FIG. 7B compares the i-CO2RR performances with NH4HCO3 feed using BPM, CEM, and AEM as the membrane. The i-CO2 production only decreased slightly as the BPM was replaced by CEM or AEM due to the absence of H+ from BPM by water dissociation. In particular, when AEM is used, HER can be largely suppressed to 13% under a relatively high local pH (FIG. 7D), resulting in a maximum formate FE of 82±4% among the systems with three membranes. It is worth noting that the possible crossover of formate in three membrane cases was taken into account, and the total FE was calculated by summing the formate concentration in both catholyte and anolyte. No formate crossover in the CEM- and BPM-based eletrolyzers was observed, and the percentage of its crossover in the AEM-based system was 4.7% (Table 3).









TABLE 3







The detection of crossover in the


NH4HCO3 systems with three membranes.[a]











Formate in Catholyte
Formate in anolyte
Crossover


System
(mM)
(mM)
percentage (%)













BPM
74.13
None
0


CEM
61.99
None
0


AEM
75.21
3.75
4.7





[a]The electrolysis was performed at 100 mA cm−2 for half-hour.






The above results demonstrate that replacing BPM with AEM for the NH4HCO3 electrolyzer not only reduces the energy consumption significantly, but also increases the FE towards formate due to the favorable microenvironments at the electrode-membrane interface. The results are in stark contrast to the reported bicarbonate electrolyzer with KHCO3 feed (Li et al., “Conversion of Bicarbonate to Formate in an Electrochemical Flow Reactor,” ACS Energy Lett. 5(8):2624-2630 (2020), which is hereby incorporated by reference in its entirety), in which AEM showed a lower performance (<20% formate FE) compared to BPM (62% formate FE): in that case, generation of i-CO2 almost solely relies on the H+ supply from the membrane owing to the sluggishness of the thermal decomposition of KHCO3. As shown in FIG. 7C, the performance of the NH4HCO3 electrolyzer used here is superior to the state-of-the-art bicarbonate electrolyzers in terms of cell voltage and FE.


To investigate the volatility of NH3 in the NH4HCO3-based electrolyzer, a 5-hour electrolysis was conducted in the AEM-based system and NH3 loss during electrolysis was quantified (FIG. 3). In theory, in the aqueous electrolyte with buffered NH4HCO3 solution, NH4+ (aq) is the dominating species (detailed analysis in Example 2, infra). Indeed, at 100 mA cm−2, negligible NH3 loss was observed in the electrolyzer, since the amount of NH4+ (aq) was virtually unchanged in the NH4HCO3 solution at each hour interval and the escaped NH3 was merely 0.07% (in NH4+ equivalent) among the total NH4HCO3 detected in the NH3-absorbing solution (FIG. 7F). A very stable cell voltage was observed at 2.2 V throughout the entire electrolysis, and the total formate FE (catholyte+anolyte) varied in a narrow range between 74% and 79% (FIG. 7E), both of which indicate the stable operation of the NH4HCO3 electrolysis with AEM. A slight increase in the cross-over of formate to anolyte was observed during 5-hour electrolysis. This was attributed to the gradual increase in the concentration gradient of formate between catholyte and anolyte. Even so, the total crossover was still <5% after 5 hours. SEM image (FIG. 8) showed that Bi catalyst maintained a nano-sized structure, where the morphology of nanosheets were likely formed from the in-situ transformation from Bi nanoparticles under the electrolysis conditions (Lee et al., “Bismuth Nanosheets Derived by In Situ Morphology Transformation of Bismuth Oxides for Selective Electrochemical CO2 Reduction to Formate,” ACS Appl. Mater. Interfaces. 14(12): 14210-14217 (2022); Han et al., “Ultrathin Bismuth Nanosheets from In situ Topotactic Transformation for Selective Electrocatalytic CO2 Reduction to Formate,” Nat. Comm. 9(1): 1-8 (2018); which are hereby incorporated by reference in their entirety).


To quantitatively compare the energy consumption of CO2RR in different cell configurations, the breakdown of energy consumption for formic acid production at the current density of 100 mA cm−2 was analyzed (detailed calculation methods are described in Example 3, infra). As shown in FIG. 9, for the systems with gaseous CO2 feed (cases 1-i and 1-ii), the majority of energy consumption arises from the regeneration of CO2, leading to higher total energy consumption than the systems fed with CO2 capture solutions. Compared with other cases, conversion of MEA-CO2 solution (case 2-ii) requires additional input of electrical energy owing to its low FE towards the desired product. The lowest energy consumption was corresponding to the systems with NH4HCO3 feed following the order of AEM<CEM<BPM (cases 2-iii to 2-v), in accordance with the experimental data provided here.


Apart from the ease of CO2 release, another key advantage of NH3-based CO2 capturing is that NH3 can be sustainably produced from wastes. As NO3 —N is a major form of pollutants in wastewater (Temkin et al., “Exposure-based Assessment and Economic Valuation of Adverse Birth Outcomes and Cancer Risk Due to Nitrate in United States Drinking Water,” Environ. Res. 176:108442 (2019), which is hereby incorporated by reference in its entirety), its electrochemical reduction offers a sustainable pathway to NH3 as a waste-derived CO2 capturing agent, while alleviating the environmental impact of NO3 itself. For this purpose, an integrated system comprising an electrolyzer for NH3 production and an absorbing unit for capturing CO2 was developed (FIG. 10). The NH3-producing electrolyzer was modified from the configuration in previous work (Chen et al., “Revealing Nitrogen-containing Species in Commercial Catalysts Used for Ammonia Electrosynthesis,” Nat. Catal. 3(12):1055-1061 (2020), which is hereby incorporated by reference in its entirety), which contains a concentrated NaOH—KOH solution as the electrolyte and commercial nickel wire mesh as the electrodes. The use of high alkalinity and elevated temperature (80° C.) facilitated simultaneous NH3 production and separation. NO3 reduction was performed under a continuous flow of N2, which carries the produced NH3 into a CO2-saturated water solution cooled at 5° C. for NH4HCO3 formation.


Electrolysis was performed at 500 mA cm−2 for 6 hours with the supply of theoretical charge of NO3 to-NH3 reaction (FIG. 11), yielding a NH4HCO3 solution with a concentration of 1.21 M in the integrated NO3-to-NH4HCO3 system. For NO3 electrolysis, FIG. 12A showed the FE of NH3 production was 99.5% with the nearly complete conversion of NO3+ (98.8%), while the FE of the side-product NO2 was only 0.06%. For the CO2 capturing unit, only a trace amount of NH3 evolution (0.2% of the total NH3 generated) was detected from its outlet, suggesting its high utilization of NH3 as the on-site generated CO2 capturing agent. The low boiling point (−33.34° C.) of ammonia and its high vapor pressure in the alkaline environment, as well as its high pKa (acid-base reaction between NH3 and CO2 in water: NH3+ CO2+H2O→NH4HCO3) have guaranteed the simultaneous NH3 generation, separation, and CO2 capturing.


Further extending the electrolysis duration in a scaled-up reactor led to the precipitation of solid NH4HCO3. The crystal phase of the separated solid was confirmed by comparing the XRD pattern with the commercial NH4HCO3 product (FIG. 12B). In the BPM-based bicarbonate electrolyzer, electrolysis with the as-prepared NH4HCO3 showed a formate FE of 70%, which is close to the result with commercial NH4HCO3 (FE: 79%) under identical conditions. The slightly lower performance is possibly because of its lower effective NH4HCO3 content (91.4%, determined by 1H NMR spectroscopy as shown in FIG. 12C) due to the incorporation of water during its precipitation. Such a tandem process can effectively “seal” the waste nitrogen and waste carbon into NH4HCO3 as a stable, pure, and easy-to-handle chemical product. Benefiting from the highly reversible nature of NH4HCO3 formation and decomposition, waste N-derived NH3 can “shuttle” the waste CO2 for its electrochemical conversion by the CO2 capture-release cycling process.


In summary, the study described herein demonstrated that NH4HCO3 can serve as a unique, highly reactive platform that bridges CO2 capturing and its electrochemical conversion via its facile in situ release. In this regard, NH4HCO3 has shown more desirable i-CO2RR performance compared to KHCO3 and MEA-CO2 owing to its much lower energy requirement for releasing i-CO2 within the electrolyzer. A mildly elevated temperature was proven sufficient for generating adequate i-CO2 for its efficient electro-reduction with negligible ammonia loss, lifting the requirement of the energy-consuming BPM in the cell system, and thus providing a proper solution to the high cell voltage of the prevailing electrolyzers for CO2 capture solutions. The highly selective produced ammonium formate can be used as an appealing candidate as energy carrier (Schiffer et al., “Ammonium Formate as a Safe, Energy-Dense Electrochemical Fuel Ionic Liquid,” ACS Energy Lett. 7:3260-3267 (2022), which is hereby incorporated by reference in its entirety).


Example 2—Ammonia Loss Analysis

The pH of the cathode electrolyte the dominance of NH4+: The system (2.5 M NH4HCO3) has a mild and stable alkalinity, pH=7.8, which can be seen in the Handbooks in Separation Science-Capillary Electromigration Separation Methods (Poole, C. F., Capillary Electromigration Separation Methods. Elsevier: 2018, which is hereby incorporated by reference in its entirety). The pH of NH4HCO3 aqueous solutions is in the middle between pKa of NH4+ (9.26) and pKa of HCO3 (6.35). Therefore, NH4+ indeed is the dominating species (96.6% of all nitrogen at 25° C.) in the NH4+(aq)/NH3(aq) equilibrium.


The two equilibria: One is the NH3(g)/NH3(aq) equilibrium, and the other is the NH4+(aq)/NH3(aq) equilibrium. The two equilibria are independent and subject to different equilibrium constants: The first is governed by the Henry's law, and the second is controlled by the solution pH. In our cathode electrolyte (2.5 M NH4HCO3), the maximum partial pressure of NH3(gas) on the immediate surface of the solution is merely 98.6 Pa (or ˜0.1vol. %), based on the Henry's constant at 25° C. (69 mol/(kg bar)). NH3 loss was not observed in the experiments using 2.5 M NH4HCO3.


Example 3—Energy Consumption Estimation

the energy consumption for CO2 reduction toward formate was calculated in different well-known cell configurations:

    • (1) Feed pure CO2 gas:
    • (1-i) gas and liquid feed-alkaline flow electrolyzer
    • (1-ii) gas-feed membrane electrode assembly flow electrolyzer
    • (2) Feed CO2 capture solutions:
    • (2-i) KHCO3 feed into bipolar membrane (BPM)-based electrolyzer
    • (2-ii) MEA-CO2 feed into BPM-based electrolyzer
    • (2-iii) NH4HCO3 feed into BPM-based electrolyzer
    • (2-iv) NH4HCO3 feed into cation exchange membrane (CEM)-based electrolyzer
    • (2-v) NH4HCO3 feed into anion exchange membrane (AEM)-based electrolyzer HER was assumed as the only side reaction. The current density was assumed at 100 mA cm−2.


CO2 regeneration. In the conventional cases by feeding purified CO2, significant energy input is required to regenerate CO2 through a few thermal and compression steps (Keith et al., “A Process for Capturing CO2 From the Atmosphere,” Joule 2(8): 1573-1594 (2018); Welch et al., “Bicarbonate or Carbonate Processes for Coupling Carbon Dioxide Capture and Electrochemical Conversion,” ACS Energy Lett. 5(3):940-945 (2020); which are hereby incorporated by reference in their entirety). From a typical calcium caustic recovery loop, the energy consumption for CO2 regeneration (CaCO3→CaO+CO2) is 178.3 KJ/molCO2 (Keith et al., “A Process for Capturing CO2 From the Atmosphere,” Joule 2(8): 1573-1594 (2018), which is hereby incorporated by reference in its entirety).


In the alkaline electrolyzer (i-1), the major loss of CO2 could be due to the combination (or alkaline hydration) reaction (CO2+OH→HCO3) and the crossover of bicarbonate from the cathode to the anode. Based on the previous literature (Dinh et al., “CO2 Electroreduction to Ethylene Via Hydroxide-mediated Copper Catalysis at an Abrupt Interface,” Science 360(6390): 783-787 (2018); Gu et al., “Modulating Electric Field Distribution by Alkali Cations for CO2 Electroreduction in Strongly Acidic Medium,” Nat. Catal. 5(4): 1-9 (2022); which are hereby incorporated by reference in their entirety), a model can be built to estimate the CO2 consumption. At the steady state, the rate of CO2 supply (through diffusion) is equal to that of CO2 combination reaction in the nearby region of the cathode:








D

CO

2






d
2


c



dx


2



=

k
·
c
·

[

OH
-

]






where DCO2 is the diffusion coefficient of CO2 in water (1.91×10−9 m2·s−1), c is the CO2 concentration in electrolyte (M), x (m) is the distance between an electrolyte location and the gas-solution interface, k is the rate constant for the combination reaction (CO2+OH→HCO3, 2.23 mol−1·m3·s−1), and [OH] is the concentration of OH (a constant 1 M assumed in this work). Solving the equation by integration and considering the boundary conditions (c=c0, or interfacial CO2 concentration, when x=0; and c=0, when x=∞), the following expression can be obtained:






c
=


c
0

·

exp

(


-



k
[

OH
-

]


D

CO

2





·

x

)






where c0 is the interfacial CO2 concentration in the electrolyte at the gas-solution interface, which is assumed as the CO2 solubility (0.038 M understand the standard conditions) in water. The CO2 consumption rate due to the combination reaction (Jh), which is the flux of CO2 across the gas-solution interface, is calculated, as follows:







J
h

=




-

D

CO

2



·


dc



dx





|

x
=
0



=



c
0

·


k
·


D

CO

2


[

OH
-

]




=


7
.
8


4
×
1


0

-
6




mol

·

s

-
1


·

cm

-
2










The CO2 consumption due to its reduction reaction at 100 mA cm−2 is calculated, as follows:







J
r

=



n
·

j

CO

2




z
·
F


=



1
·
0.1


2
·
96485


=


5
.
1


8
×
1


0

-
7




mol

·

s

-
1


·

cm

-
2










where jCO2 is the partial current density of CO2 reduction, n is the number of required CO2 molecules for one molecular formate produced (n=1 here), z is the number of electrons involved in the CO2-to-formate reaction, and F is the Faraday constant (96,485 C·mol−1). Then, the CO2 utilization in the alkaline electrolyzer (1-i) is calculated, as follows:





CO2 utilization=Jr/(Jr+Jh)=6.2%


Membrane electrode assembly-based flow electrolyzers by using AEM suffer from a ˜30% CO2 loss due to the bicarbonate crossover from cathode to anode (Weng et al., “Towards Membrane-electrode Assembly Systems for CO2 Reduction: A Modeling Study,” Energy Environ. Sci. 12(6): 1950-1968 (2019); Lee et al., “Electrochemical Upgrade of CO2 From Amine Capture Solution,” Nat. Energy 6(1):46-53 (2021); which are hereby incorporated by reference in their entirety), regardless of feeding with either a humidified pure CO2 gas (1-ii) or circulating aqueous NH4HCO3 (2-v). In addition, when feeding pure CO2 into the MEA cell, there is an additional ˜35% CO2 loss due to the escape of unreacted CO2 (Lee et al., “Electrochemical Upgrade of CO2 From Amine Capture Solution,” Nat. Energy 6(1):46-53 (2021), which is hereby incorporated by reference in its entirety). As such, the CO2 utilization in the MEA-based electrolyzers with AEM can be estimated as 35% and 70% when feeding gaseous CO2 and aqueous NH4HCO3, respectively.


By contrast, 90% of CO2 utilization can be assumed for the cases of using BPM or CEM when feeding CO2 capture solutions (2-i, 2-ii, 2-iii, and 2-iv), by avoiding the bicarbonate crossover between electrodes and the gas escape at cathode (Lee et al., “Electrochemical Upgrade of CO2 From Amine Capture Solution,” Nat. Energy 6(1):46-53 (2021); Li et al., “CO2 Electroreduction From Carbonate Electrolyte,” ACS Energy Lett. 4(6): 1427-1431 (2019); which are hereby incorporated by reference in their entirety).









TABLE 4







Summary of CO2 consumption and utilization for all cases.










CO2 consumption














Combination
HCO3
Unreacted

Total CO2


Case
reaction
crossover
escape
Total
utilization





1-i
93.8%


93.8%  
6.2% 


1-ii

30%
35%
65% 
35%


2-i



0%
90%


2-ii



0%
90%


2-iii



0%
90%


2-iv



0%
90%


2-v
  30%


30% 
70%









Thus, the energy consumption for CO2 regeneration is shown as follows: CO2 regeneration energy=(178.3 KJ/mol)/(CO2 utilization)


Thermodynamic Energy

In the AEM-based eletrolyzer through NH4HCO3 feed (case 2-v), the cathodic electrolyte is 2.5 M NH4HCO3, and the anodic electrolyte is 1 M KOH. An AEM separates the cathodic and anodic electrolytes.


1. Estimation of Catholyte pH During Electrolysis

The catholyte is 40 mL of 2.5 M NH4HCO3 and acts as a buffer. The total amount of NH4HCO3 before thermodynamic equilibrium is 40×2.5=100 mmol. At the equilibrium, the pH of the 2.5 M NH4HCO3 is ˜7.80 (i.e., the middle point between the two pKa values: 6.35 of H2CO3 and 9.25 of NH4+, CRC handbook). Therefore, the amounts of NH4+ and NH3 are 96.45 mmol (96.45% of N) and 3.55 mmol (3.55% of N), respectively, based on the Henderson-Hasselbalch equation: pH=pKa(NH4+)+log10([NH3]/[NH4+]) and the mass conservation ([NH3]+[NH4+]=100 mmol), where [NH3] and [NH4+] are concentrations of NH3 and NH4+, respectively.


The cathodic reaction is





CO2+2e+H2O→HCOO+OH


After electrolysis at 400 mA (100 mA cm−2) for 30 min, the cathodic reaction generates 3.7 mmol of OH (assuming 100% FE to HCOO), and this OH should react with and thus turn 3.7 mmol NH4+ to NH3. Based on the measurement of i-CO2, the total decomposed NH4HCO3 should be 14.1 mmol. So, the remaining NH4+ is 96.45−3.7−14.1×96.45%=79.15 mmol. The amount of remaining NH3 is 3.55+3.7−14.1×3.55%=6.75 mmol.


Using the Henderson-Hasselbalch equation, the solution pH after electrolysis reaction is





pH=pKa(NH4+)+log10([NH3]/NH4+])=9.25+log10(6.75/79.15)=8.18


So, it may be assumed that the electrolyte pH is around 8.0 during the reaction. Note that the transport of OH through the AEM was not considered, because the concentration of HCO3 is much higher than OH.


2. Obtaining the Standard Reduction Potential of CO2/HCOO at pH=8.0


From CRC handbook, the following ΔG+ values can be found:





ΔG0 (HCOO)=−351.0 KJ/mol, ΔG0 (OH)=−157.2 KJ/mol, ΔG0 (CO2, gas)=−394.4 KJ/mol, and ΔG0 (H2O)=−237.1 KJ/mol.


The cathodic reaction is





CO2+2eH2O→HCOO+OH


For this reaction, ΔG0=351.0−157.2−(−394.4)−(−237.1)=123.3 KJ/mol φrev(CO2/HCOO)=φ0 (CO2/HCOO)=−123.3×1000/(96485×2)=0.639 VSHE


This corresponds to the standard condition at pH=14 ([OH]=1 M). So, the reversible reduction potential at pH=8.0 is





φrev(CO2/HCOO, pH=8.0)=−0.639+0.0592/(2×log(1/[OH]))=−0.639+0.05916/(2×log[1/(10−14/10−8.0)])=0.462 VSHE


3. Obtaining Thermodynamic Cell Voltage for the pH-Asymmetric Configuration. For Clarity, φrev and Erev are Used to Stand for the “Reversible Electrode Potential” and the “Reversible Cell Voltage”, Respectively, Throughout this Document.


For the cell with AEM, the cathodic reaction is





CO2+2e+H2O→HCOO+OH  (RXN 1, pH=8.0)


The reversible reduction potential for RXN 1 is −0.462 VSHE as was calculated. And the anodic reaction is





4OH→O2+2H2O+4e  (RXN 2, pH =14)


The standard reduction potential for RXN 2 is 0.401 VSHE from CRC handbook.


To calculate the thermodynamic cell voltage for the overall reaction, RXN 1 is multiplied by 2 and RXN 2 is added to eliminate the electrons:





2CO2+4OH (pH 14)→2HCOO+O2+2OH (pH 8.0)   (RXN 3)


Note that OH cannot be simply eliminated because of the difference in pH. The thermodynamic cell potential for RXN 3, Erev(RXN 3), is −0.462−0.401=−0.863 V. In the actual cell configuration, because AEM can transport OH, the following equation should apply:





2OH(pH 8.0)→2OH(pH 14)   (RXN 4)


To calculate the reversible potential for this equation, the standard reduction potential of H2O/H2 may be used:





2H2O+2e→H2+2OH (pH 14)   (RXN 5)





φrev(H2O/H2, pH 14)=−0.828 VSHE (from CRC handbook)





2H2O+2e→H2+2OH (pH 8.0)   (RXN 6)





φrev(H2O/H2, pH 8.0)=−0.828 VSHE+0.05916/(2×log10(1/[OH]2))=−0.473 VSHE.


RXN 4 corresponds to a cell constructed by using RXN 5 as the cathode, and RXN 6 as the anode.


Therefore, Erev for RXN 4 is −0.828−(−0.473)=−0.355 V


By adding RXN 3 and RXN 4 the equation is:





2CO2+2OH (pH 14)→2HCOO+O2   (RXN 7)





ΔG(RXN 7)=ΔG(RXN 3)+ΔG(RXN 4).


Because ΔG=−z·F·φrev for reduction electrode reaction or −z·F·Erev for cell reaction, the equation is:





−4×F×Erev(RXN 7)=−4×F×φrev (RXN 3)−2×F×Erev (RXN 4)





Erev(RXN 7)=φrev(RXN 3)+½×Erev(RXN 4)=−0.863+½×(−0.355)=−1.040 V.


So, the thermodynamic cell voltage for RXN 7, Erev(RXN 7), is 1.040 V, which is the thermodynamic cell voltage for our AEM-based electrolyzer.


Note that this value is exactly equal to the cell voltage when we directly calculate the Erev of RXN 7 at standard conditions (pH=14):





φ0 (CO2/HCOO)=−0.639 VSHE (we calculated this value before)





φ0 (O2/OH)=0.401 VSHE (from CRC handbook)


The thermodynamic cell potential E0=−0.639−0.401=−1.040 V, corresponding to 1.040 V of thermodynamic cell voltage.


For the CEM electrolyzer (K+ transport) with NH3HCO3 feed (case 2-iv) and BPM electrolyzer with CO2 capture solutions feed (case 2-i, 2-ii, and 2-iii), the catholyte pH was assumed at 8.0 and anolyte pH was 14.


The thermodynamic cell potential is Erev=−0.462−0.401=−0.863 V, corresponding to thermodynamic cell voltage of 0.863 V.


Considering the different FE of cathodic product (FEc) in different cell configurations, the actual thermal energy to produce 1 mole of formate product can be calculated as follows:








Δ


G
minimum


=


-
z

·
F
·

(


E

cell
-
formate


rev





FE


c


)



,




where, z is number of electrons required per mole of formate product; ΔGminimum is the required minimum electrical energy applied to the electrolytic cell per unit of formate product, Ecell-formateo is the standard reversible cell voltage for formate production.


From the literature, in the feeding of pure CO2 (cases 1-i and 1-ii), the FEc toward target products can attain ˜90% (Kirner et al., “Exploring Electrochemical Flow-Cell Designs and Parameters for CO2 Reduction to Formate under Industrially Relevant Condition,” J. Electrochem. Soc. 169(5): (2022); Lee et al., “Bismuth Nanosheets Derived by In Situ Morphology Transformation of Bismuth Oxides for Selective Electrochemical CO2 Reduction to Formate,” ACS Appl. Mater. Interfaces. 14(12): 14210-14217 (2022); which are hereby incorporated by reference in their entirety). In the feeding of KHCO3 into the BPM-based electrolyzer (case 2-i), as was observed from the experimental results and the literature (Li et al., “Conversion of Bicarbonate to Formate in an Electrochemical Flow Reactor.” ACS Energy Lett. 5(8):2624-2630 (2020), which is hereby incorporated by reference in its entirety), the FEc is assumed as 55%. In the feeding of MEA-CO2 in the BPM-based electrolyzer (2-ii), the FEc is assumed as 20% at the current density of 100 mA cm−2. Based on the experimental results, in the feeding of NH4HCO3 to AEM (2-v), BPM (2-iii), and CEM (2-iv), the FEcs are assumed as 90%, 80%, and 70%, respectively.









TABLE 5







Summary of thermodynamic energy for all cases.













Ecell-formateo
FEc
ΔGminimum



Case
(V)
(%)
(kJ mol−1)
















1-i
−1.040
90
223.0



1-ii
−1.040
90
223.0



2-i
−0.863
55
302.8



2-ii
−0.863
20
832.7



2-iii
−0.863
80
208.2



2-iv
−0.863
70
237.9



2-v
−1.040
90
223.0










Cathode Energy Loss

Cathode energy loss is due to the overpotential for CO2 reduction toward target products. Based on the literature, the overpotential of CO2-to-formate on Bi-based catalysts was assumed at 100 mA cm−2 is 0.7 V (Lee et al., “Bismuth Nanosheets Derived by In Situ Morphology Transformation of Bismuth Oxides for Selective Electrochemical CO2 Reduction to Formate,” ACS Appl. Mater. Interfaces. 14(12): 14210-14217 (2022), which is hereby incorporated by reference in its entirety).


Then, the cathodic energy loss can be calculated, as follows:





Cathode energy loss=z·F·η/FEc,


where z is the number of electrons involved in the CO2-to-formate reaction, and F is the Faraday constant (96,485 C·mol−1).


Anode Energy Loss

In the alkaline medium (e.g., 1 M KOH), the overpotential for OER on Ni-based catalysts can be as low as <400 mV at current density of 100 mA cm−2. For example, the literature reported a nanostructured NiCo alloy that showed an overpotential of 326 mV (Wu et al., “A Nanostructured Nickel-cobalt Alloy With an Oxide Layer for an Efficient Oxygen Evolution Reaction,” J. Mater. Chem. 5(21): 10669-10677 (2017), which is hereby incorporated by reference in its entirety). So, this value was used to calculate anode energy loss, as follows:





Anode energy loss=z·F·η/FEc=2×96485×0.326/FEc


where z is the number of electrons involved in the CO2-to-formate reaction, and F is the Faraday constant (96,485 C·mol−1).


Ohmic Loss

The Ohmic loss is mainly caused by the membrane resistance.


The potential drop (Δϕ) across the membrane is calculated as follows:







Δ

ϕ

=


i

L

κ





where i is the current density and L is the thickness of the membrane.


Based on the literature, the κ value for AEM (A201 membrane) is around 20 mS cm−1 in OH (Duan et al., “Water Uptake, Ionic Conductivity and Swelling Properties of Anion-exchange Membrane,” J. Power Sources 243:773-778 (2013), which is hereby incorporated by reference in its entirety). With its thickness of 28 μm and at the current density of 100 mA cm−2. the calculated Δϕ is 14 mV.


The κ value for Nafion 115 membrane is 21 mS cm−1 in K+, so the Δϕ is 60 mV.


The BPM is thicker than AEM and CEM, because it is sandwiched by a cation exchange layer (CEL) and an anion exchange layer (AEL). The Δϕ value for BPM (Fumasep FBM) is assumed to be 280 mV at 100 mA cm−2, based on the specification document. In addition, 0.828 V potential is required under the standard condition to dissociate water into H+ and OH by using BPM, which should also be included in the ohmic loss.


Therefore, the energy loss due to membrane resistance is calculated as follows:





For AEM and CEM cases: Ohmic loss=z·F·Δϕ/FEc





For BPM cases: Ohmic loss=z·F·(Δϕ+0.828 V)/FEc


Separation of Product

In the near neutral media, an electrodialysis (ED) process is used to convert formate to formic acid before the separation of formic acid (Li et al., “CO2 Electroreduction From Carbonate Electrolyte,” ACS Energy Lett. 4(6): 1427-1431 (2019), which is hereby incorporated by reference in its entirety). Then, a pressure-swing distillation (PSD) method is used for the separation of formic acid from the mixed solutions (Mahida et al., “Process Analysis of Pressure-swing Distillation for the Separation of Formic Acid-water Mixture,” Chem. Pap. 75(2):599-609 (2021), which is hereby incorporated by reference in its entirety). The reported energy consumption to separate formic acid is about 265 KJ/mol (Gu et al., “Modulating Electric Field Distribution by Alkali Cations for CO2 Electroreduction in Strongly Acidic Medium,” Nat. Catal. 5(4): 1-9 (2022); Mahida et al., “Process Analysis of Pressure-swing Distillation for the Separation of Formic Acid-water Mixture,” Chem. Pap. 75(2): 599-609 (2021); which are hereby incorporated by reference in their entirety). The energy consumption in the ED process is ignored in our modeling, because of the much lower energy consumption than the PSD process (Gu et al., “Modulating Electric Field Distribution by Alkali Cations for CO2 Electroreduction in Strongly Acidic Medium,” Nat. Catal. 5(4): 1-9 (2022), which is hereby incorporated by reference in its entirety).


In summary, the total energy consumption can be obtained by summing all components identified above: thermodynamic energy consumption, cathode energy loss, anode energy loss, ohmic energy loss, and separation energy use.


Although preferred embodiments have been depicted and described in detail herein, it will be apparent to those skilled in the relevant art that various modifications, additions, substitutions, and the like can be made without departing from the spirit of the invention and these are therefore considered to be within the scope of the invention as defined in the claims which follow.

Claims
  • 1. An electrochemical method for converting captured CO2 into formate (HCOO−), said method comprising: capturing waste CO2 by co-absorption of the waste CO2 with green ammonia (NH3) to form ammonium bicarbonate (NH4HCO3) and converting the ammonium bicarbonate (NH4HCO3) into formate (HCOO−), wherein said converting is carried out in an integrated flow electrolyzer system.
  • 2. The method according to claim 1, wherein said capturing is carried out according to the following formula: CO2+H2O+NH3→NH4HCO3.
  • 3. The method according to claim 1 or claim 2, wherein the integrated flow electrolyzer system comprises an alkaline electrolyzer for producing NH3 from NO3−.
  • 4. The method according to claim 3, wherein the alkaline electrolyzer comprises an anode, a cathode, and a reaction medium. 5 The method according to any one of the preceding claims, wherein the reaction medium comprises a concentrated NaOH—KOH solution.
  • 6. The method according to claim 4 or claim 5, wherein the anode and the cathode comprise nickel wire mesh or nickel alloys.
  • 7. The method according to any one of claims 4-6, wherein the alkaline electrolyzer has no membrane between the anode and the cathode.
  • 8. The method according to any one of the preceding claims, wherein the integrated flow electrolyzer system comprises an NH3—CO2 absorbing unit wherein said capturing is carried out.
  • 9. The method according to any one of the preceding claims, wherein the integrated flow electrolyzer comprises a bicarbonate electrolyzer for carrying out said converting.
  • 10. The method according to claim 9, wherein the bicarbonate electrolyzer comprises an anode, a cathode, and an anion exchange membrane.
  • 11. The method according to claim 10, wherein the bicarbonate electrolyzer further comprises: a reaction medium comprising ammonium bicarbonate (NH4HCO3); anda power supply operably connected to the anode and the cathode.
  • 12. The method according to claim 9 or claim 10, wherein the anode comprises nickel.
  • 13. The method according to any one of claims 9-11, wherein the anode comprises nickel foam.
  • 14. The method according to any one of claims 9-13, wherein the cathode comprises electrodeposited-Bi (ED-Bi).
  • 15. The method according to claim 14, wherein the cathode is formed on carbon paper.
  • 16. The method according to any one of the preceding claims, wherein said converting is carried out at a temperature of between 35-80° C.
  • 17. The method according to any one of the preceding claims, wherein said converting is carried out at a temperature of between 40-60° C.
  • 18. An integrated flow electrolyzer system comprising: an alkaline electrolyzer for producing NH3 from NO3;an NH3—CO2 absorbing unit whereby waste CO2 is co-absorbed with ammonia (NH3) to form ammonium bicarbonate (NH4HCO3); anda bicarbonate electrolyzer for converting the ammonium bicarbonate (NH4HCO3) into formate (HCOO−).
  • 19. The electrolyzer system according to claim 18, wherein the alkaline electrolyzer comprises a first anode, a first cathode, a first reaction medium, and a first power supply operably connected to the first anode and the first cathode.
  • 20. The electrolyzer system according to claim 19, wherein the first reaction medium comprises a concentrated NaOH—KOH solution.
  • 21. The electrolyzer system according to claim 19 or claim 20, wherein the first anode and the first cathode comprise nickel wire mesh.
  • 22. The electrolyzer system according to any one of claims 18-21, wherein the alkaline electrolyzer has no membrane between the first anode and the first cathode.
  • 23. The electrolyzer system according to any one of claims 18-22, wherein the bicarbonate electrolyzer comprises a second anode, a second cathode, and an anion exchange membrane.
  • 24. The electrolyzer system according to claim 10, wherein the bicarbonate electrolyzer further comprises: a second reaction medium comprising ammonium bicarbonate (NH4HCO3); anda second power supply operably connected to the second anode and the second cathode.
  • 25. The electrolyzer system according to claim 23 or claim 24, wherein the second anode comprises nickel.
  • 26. The electrolyzer system according to any one of claims 23-25, wherein the second anode comprises nickel foam.
  • 27. The electrolyzer system according to any one of claims 23-26, wherein the second cathode comprises electrodeposited-Bi (ED-Bi).
  • 28. The electrolyzer system according to any one of claims 23-27, wherein the second cathode is formed on carbon paper.
Provisional Applications (1)
Number Date Country
63415154 Oct 2022 US