Next Article in Journal
Hierarchical Blocking Control for Mitigating Cascading Failures in Power Systems with Wind Power Integration
Next Article in Special Issue
Exploring Hydrogen-Enriched Fuels and the Promise of HCNG in Industrial Dual-Fuel Engines
Previous Article in Journal
Evaluation of a Simplified Model for Three-Phase Equilibrium Calculations of Mixed Gas Hydrates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Electrochemical Synthesis of Ammonia via Nitrogen Reduction and Oxygen Evolution Reactions—A Comprehensive Review on Electrolyte-Supported Cells

by
Hizkia Manuel Vieri
1,
Moo-Chang Kim
1,2,
Arash Badakhsh
3 and
Sun Hee Choi
1,4,*
1
Hydrogen·Fuel Cell Research Center, Korea Institute of Science and Technology (KIST), Seoul 02792, Republic of Korea
2
Department of Mechanical and Automotive Engineering, Seoul National University of Science and Technology, Seoul 01811, Republic of Korea
3
PNDC, University of Strathclyde, Glasgow G68 0EF, UK
4
Energy & Environment Technology, KIST School, University of Science and Technology (UST), Seoul 02792, Republic of Korea
*
Author to whom correspondence should be addressed.
Energies 2024, 17(2), 441; https://doi.org/10.3390/en17020441
Submission received: 8 December 2023 / Revised: 5 January 2024 / Accepted: 14 January 2024 / Published: 16 January 2024
(This article belongs to the Special Issue Unconventional Hydrogen Applications and Systems)

Abstract

:
The application of protonic ceramic electrolysis cells (PCECs) for ammonia (NH3) synthesis has been evaluated over the past 14 years. While nitrogen (N2) is the conventional fuel on the cathode side, various fuels such as methane (CH4), hydrogen (H2), and steam (H2O) have been investigated for the oxygen evolution reaction (OER) on the anode side. Because H2 is predominantly produced through CO2-emitting methane reforming, H2O has been the conventional carbon-free option thus far. Although the potential of utilizing H2O and N2 as fuels is considerable, studies exploring this specific combination remain limited. PCEC fabrication technologies are being developed extensively, thus necessitating a comprehensive review. Several strategies for electrode fabrication, deposition, and electrolyte design are discussed herein. The progress in electrode development for PCECs has also been delineated. Finally, the existing challenges and prospective outlook of PCEC for NH3 synthesis are analyzed and discussed. The most significant finding is the lack of past research involving PCEC with H2O and N2 as fuel configurations and the diversity of nitrogen reduction reaction catalysts. This review indicates that the maximum NH3 synthesis rate is 14 × 10−9 mol cm−2 s−1, and the maximum current density for the OER catalyst is 1.241 A cm−2. Moreover, the pellet electrolyte thickness must be maintained at approximately 0.8–1.5 mm, and the stability of thin-film electrolytes must be improved.

1. Introduction

Hydrogen (H2) has considerable potential for energy storage. However, its low energy density poses challenges to storage and transport. One solution is to convert H2 to ammonia (NH3), where H2 that is obtained from water electrolysis (i.e., green H2) reacts with nitrogen (N2) in the air to produce NH3, which can be reversibly converted into H2 and N2 after transportation [1]. Currently, NH3 is industrially produced using the Haber–Bosch process, which requires high pressure (100–200 bar) and temperature (300–400 °C) to activate the Fe-based catalysts [2,3]. Recent advancements include the hydrogenation of N2 using Ru catalysts, which require milder reaction conditions [4].
However, the thermal approach for NH3 synthesis remains expensive, as it requires a large centralized infrastructure and is energy-intensive, consuming approximately 485 kJ mol−1 (approximately 2% of the global energy use per annum) [5]. Fertilizers Europe made conjectures about the decarbonization of the European NH3 industry [6]. They proposed that approximately 10% of H2 derived for NH3 production in 2030 could possibly originate from renewable sources. Electrochemical NH3 synthesis via N2 reduction provides a relatively high efficiency of up to 20%, environmental compatibility with renewable sources (solar, tidal, and wind), on-site H+ generation from water oxidation, and adaptable reaction control [7]. This process converts sustainable electricity from sources such as wind power into NH3 for use as a synthetic fuel or chemical feedstock [8,9].
The operating conditions are classified into three types based on the temperature, namely, low (<100 °C), intermediate (200–500 °C), and high (>400 °C) temperature. High-temperature conditions enhance the catalytic activity and substantially increase the Faradaic efficiency in NH3 synthesis. High-temperature NH3 synthesis can be performed using proton-conducting electrolytes (PCEs) or oxygen-conducting electrolytes (OCEs). OCEs, while efficient, generally present slower rates of NH3 production than PCEs [10]. This review focuses on NH3 production in protonic ceramic electrolysis cells (PCECs), which operate at high temperatures (400–600 °C). As shown in Figure 1, the PCEC combines the nitrogen reduction reaction (NRR) with other electrochemical reactions that yield protons (H+), facilitating NH3 production. Liu et al. briefly reviewed three primary PCEC configurations involving H2, CH4, and H2O as proton (H+) sources. Although H2O and N2 have considerable potential as fuels, this specific combination has not been sufficiently investigated. Despite consuming substantially more electrical energy than the other configurations, using H2O and N2 directly to produce NH3 is carbon-free and abundant in the feedstock [11]. This configuration couples the NRR with the oxygen evolution reaction (OER). High-temperature NH3 synthesis has been extensively reviewed [12,13,14,15,16]. However, reviews of the specific PCEC configurations for NH3 synthesis are limited. The unavailability of a comprehensive strategy for PCEC fabrication in a single study poses a considerable challenge for researchers aiming to develop PCECs customized for NH3 synthesis.
The three essential components of a PCEC are an electrolyte, an anode, and a cathode (Figure 1). The electrolyte serves as an ion conductor that separates the electrodes and facilitates ion transport. The electrochemical reactions within the PCEC are non-spontaneous, implying that they do not occur naturally; an external influence, typically electricity, is necessary to drive the reactions [17].
  • Anode reaction
At the anode, an electric current passes through water, thereby splitting water molecules into hydrogen (protons) and oxygen ions.
Reaction: 2H2O → O2 + 4H+ + 4e
  • Electron and proton transport:
Electrons generated at the anode during water splitting are transported through an external electric circuit.
Protons generated at the anode during water splitting are transported through the electrolyte [18].
  • Cathode reaction
At the cathode, protons (H+) from the anode and nitrogen atoms react to produce NH3
Reaction: N2 + 3H+ + 3e → NH3
Based on their supporting materials, high-temperature fuel cells can be categorized into three types: (i) cathode-, (ii) electrolyte-, and (iii) anode-supported. They are named based on the component that is fabricated first, which is generally the thickest component [19]. In this review, we focus only on electrolyte-supported cells. Despite the lower ohmic losses and power densities, anode-supported cells are difficult to fabricate on a large scale owing to the highly porous support and high manufacturing costs [20,21,22,23].
A seminal review by Giddey et al. provided a comprehensive overview of material construction, major technical challenges, and the technological landscape in the field [12]. However, the reaction conditions were not broadly defined. Garagounis et al. extensively reviewed solid-state NH3 synthesis, particularly in PCECs [13]. Another review by Medvedev et al. in 2019 focused on NH3 and H2 producing PCECs, emphasizing design parameters such as thickness, partial pressure of H2O, and other operational conditions such as polarization loss, ohmic loss, thermoneutral voltage, and open-circuit voltage [16].
We observed a gap in research involving N2 at the cathode and only H2O at the anode, except for the study by Yun et al. [24], which will be discussed subsequently. In another study, although 26.8 nmol NH3 s−1cm−2 was achieved using N2 at the cathode and H2O at the anode, expensive plasma-assisted technology was employed [25]. Most studies focused on mixtures of H2O and H2. Among them, the highest NH3 production rate of 14 mol−9 cm−2 s−1 was reported by Chien et al. [26]. Figure 2 depicts the questions that we aim to answer through this review.
In this review, we summarize the existing design strategies in terms of electrolyte fabrication and properties, in addition to the electrocatalytic performance based on the NH3 synthesis rate, Faradaic efficiency, and current density. Furthermore, the NRR and OER mechanisms are discussed to address prevalent challenges. Finally, we highlight the challenges encountered in the development of PCEC devices and provide directions for further advancements in this technology.

2. Reaction Mechanism

For further advancements in the OER and NRR, the underlying mechanisms must be elucidated. Figure 3 provides a visual representation of the four primary pathways involved in N2 reduction and NH3 production in the PCECs.
Pathway A illustrates the electrochemical NH3 synthesis through a dissociative mechanism, wherein protons directly combine with electrons to reduce the adsorbed N2 on the catalyst, thereby producing NH3. In contrast to conventional methods, this direct electrochemical NH3 synthesis method does not require a stoichiometric amount of gaseous H2, resulting in a considerably higher Faradaic selectivity. To enhance Pathway A, the NRR electrode should inhibit the hydrogen evolution reaction (HER) while maintaining sufficient electrocatalytic activity for N2 reduction.
In contrast, Pathway B represents the conventional Haber–Bosch (HB) reaction, wherein N2 is reduced by gaseous H2 produced through the HER. Here, thermochemical catalytic NH3 synthesis primarily governs NH3 production. Therefore, the H2/N2 ratio is instrumental in determining the NH3 production rate because it follows the thermodynamics of thermochemical NH3 synthesis. Increasing the NH3 yield requires the PCECs to operate at a high current density to ensure that the H2/N2 ratio is approximately 3/1. However, under such conditions, the NH3 Faradaic selectivity is low because a substantial amount of H2 remains unused.
In the N2 adsorption depicted by Pathway C, the N≡N bond remains intact after N2 adsorption until a specific step in the hydrogenation process [27]. In contrast, in Pathway D, hydrogenation occurs alternately on two nitrogen atoms, and the N≡N bond breaks only in the final step, thus forming the first NH3 molecule; another NH3 molecule remains bonded to the surface of the catalyst. These pathways help to clarify the complex mechanisms governing NH3 production in PCECs [28,29,30,31,32,33].
The OER mechanism must be investigated in addition to the NRR mechanism. Liu et al. conducted a comprehensive study based on density functional theory calculations to elucidate the OER on La0.6Sr0.4Ce0.2Fe0.8O3−δ catalysts, which are prominent catalysts wherein cobalt is doped in high-entropy perovskites. Figure 4 illustrates the potential OER pathway in PCECs. Their study revealed the following steps in the OER mechanism [34]:
(i)
The reaction is initiated when a water molecule is adsorbed on the catalyst surface.
(ii)
Consequently, surface-bound hydroxyl species (HO*) are formed.
(iii)
The generated HO* decomposes into hydrogen (H*) and oxygen (O*) species.
(iv)
Protons (H+) are transferred to the cathode through the electrolyte.
(v)
Finally, gaseous oxygen materializes through desorption [35].
The analysis of the OER catalytic mechanism elucidates the intricate processes that govern efficient OERs. By investigating the intricate interplay between surface sites, reaction intermediates, and electron transfer pathways, we can acquire a deeper understanding of the underlying principles driving the catalytic activity. This can facilitate the optimal design of catalysts and contribute to the broader field of sustainable energy conversion.
Continued research can potentially reveal more nuances in the OER mechanism. Our models and predictions can be refined by integrating advanced experimental techniques and computational approaches, ultimately promoting the development of catalysts with enhanced performance and durability. Addressing the challenges associated with the OER mechanism will enable renewable energy technologies to be utilized more efficiently.

3. PCEC Design Strategies

High-temperature cells such as solid oxide fuel cells (SOFCs) have been extensively designed [14]. However, selecting an appropriate design for each application is important because the cell performance strongly depends on the cathode, anode, and electrolyte materials. Medvedev discussed the influence of the design strategies on the performance of H2-producing PCECs [16]. However, PCEC design strategies for NH3 synthesis are limited. This section describes the fabrication and design strategies for the electrolyte and electrodes.

3.1. Electrolyte Design Strategies

The electrolyte is considered the most important component of a PCEC, particularly in electrolyte-supported systems because it occupies the largest volume. BaCeO3 and BaZrO3-based materials have been reported to be good protonic conductors [20]. Although BaCeO3 exhibits high conductivity, BaZrO3 provides better stability. Furthermore, Ce and Zr have been mixed with other dopants, such as Y and Yb, resulting in the well-known BaZr0.4Ce0.4Y0.1Yb0.1O3−δ (BZCYYB 4411) and BaZr0.1Ce0.7Y0.1Yb0.1O3−δ (BZCYYb 1711) electrolytes [21,22]. In general, a higher Ce content increases the proton conductivity, whereas a higher Zr content increases the stability [23,36]. In a PCEC, the applied electrical current serves two potential pathways: proton and electron transfers. The primary objective of a PCEC is to facilitate the movement of protons (H+) from the OER to the NRR sides, where they participate in hydrogen production. However, electron transfer, in which electrons move instead of protons, can be detrimental to the cell performance. This is because electron transfer reduces the Faradaic efficiency of the cell; particularly, a portion of the electrical energy is redirected to unintended reactions, leading to energy losses and potentially reducing the overall hydrogen production efficiency. Increasing the Ce content increases the protonic transference number and decreases the electron transference number. Therefore, BZCYYb 1711 is more suitable for PCECs than BZCYYb 4411 [37].
This discussion pertains to three distinct electrolyte forms: pellets, thin films, and columnar structures. Pellet-type electrolytes have been extensively used in fuel cells and batteries [38,39]. They are relatively easy to manufacture and integrate into PCEC stacks [40]. This method involves placing a few grams of electrolyte powder into a mold, which is typically cylindrical, and ultimately shaping it into a coin-like pellet. This conventional approach has resulted in considerable advancements in terms of electrolyte composition. The details of various pellet electrolytes developed thus far are listed in Table 1.
While the solid-state reaction is common for powder preparation, sol–gel or glycine nitrate processes are commonly employed to obtain finer particles and higher peak power densities [42]. The pellet electrolyte thickness is in the range of 0.8–1.6 mm (Table 1). Medvedev suggested that thin-film technology must be developed for the electrochemical synthesis of NH3 [16]. With reference to the electrolyte properties, the thickness plays a crucial role in determining both the ohmic resistance and electron transport characteristics. Thicker electrolytes increase the ohmic resistance, whereas thinner electrolytes increase contact resistance at the electrolyte/electrode interface. In conventional electrolytes, the ohmic resistance tends to increase. This limitation can be effectively addressed using thin-film electrolytes [44]. Recent advancements in thin-film electrolytes for PCECs have been comprehensively presented in Table 2.
The technological advancements in thin-film electrolytes are attributed to their reduced dimensions. Various deposition techniques, such as inkjet printing, pulsed layer deposition (PLD), tape casting, spin coating, and screen printing, are involved [49,50,51,52,53]. Screen printing and spin coating are known to be simple. However, considerable material wastage occurs, with a minimum thickness of approximately 10 µm. Furthermore, PLD, which is recognized for its effectiveness, requires complex operation and energy-intensive vacuum conditions. Inkjet printing offers the advantage of producing highly uniform surfaces with a minimum thickness of just 0.83 µm [49].
BaZr0.1Ce0.7Y0.1Yb0.1 is an extensively investigated proton-conducting electrolyte, although some researchers have diverged, advocating other compositions as the most studied [22,45,47]. Some researchers have explored doping with elements such as Nd, Sc, In, and Hf, in addition to Y and Yb, to enhance the stability and sinterability of the electrolyte [22,43,44,46,48,54]. Ding et al. introduced a novel approach by ball milling, pelletizing, calcining, and crushing the pellets to produce a pure-phase powder on a relatively large scale (up to 4 kg per batch) [55]. Another approach involves combining all BZCYYb1711 nitrate precursors in deionized water, followed by the addition of a specific quantity of NaOH. NaOH reacts with nitrates and forms mixed metal hydroxides before calcination, which produce mixed metal oxides (perovskite). This mixture is washed and then subjected to a high-temperature solid-state reaction [54]. The conventional box furnace method at 1000–1500 °C with a ramp of 1–5 °C min−1 is typically used for sintering. Spark plasma and microwave sintering have been examined for rapid high-temperature results, with the aim of matching or surpassing the performance of conventional sintering [47,56].
The previous discussion encompassed two distinct electrolyte types (Figure 5), namely, the planar type (pellet and thin film), which necessitates sealing agents for the reactor connection, and the tubular configuration, which operates seamlessly without requiring sealing. Columnar electrolytes can be fabricated by templating using plaster molds or by rolling a thin-film electrolyte. Notably, columnar electrolytes have been employed in SOFCs [57,58,59]. Table 3 lists examples of columnar electrolyte utilization.

3.2. Electrode Design Strategies

In contrast to thermochemical catalysts, electrocatalysts must be pretreated to ensure that they can mechanically bond to the electrolyte while remaining catalytically active. In electrolyte-supported cells, the cathode and anode are deposited on opposite sides of the cell. The difference between the anode and cathode materials makes it necessary to specify the method that should be used to deposit them onto the electrolyte.
Recently, various conventional and ambient-condition deposition methods, such as the doctor blade method, drop coating, screen printing, tape casting, and spray coating, have been used to develop electrode materials for PCECs (Table 4). Although these methods are simple, the electrode thickness is controlled only through a randomized parameter, such as printing passes or number of drops. Therefore, many researchers have recently employed PLD to produce a smooth electrode surface and ensure good interfacial contact between the electrolyte and electrode [62].

4. Current Progress

The NRR and OER have been studied extensively; the catalysts used for these reactions are summarized in Table 5 and Table 6, respectively.
Based on Table 5, metallic catalysts evidently yield higher reaction rates. This could be attributed to two main factors. First, perovskite-based electrocatalysts may possibly be affected by degradation at the interface and thermal mismatch with the electrolyte, thus decreasing the performance [24]. Second, the electrochemical promotion of catalysis is more pronounced in pure metal catalysts with a higher effective double layer (S*eff) on their surfaces than in supported electrocatalysts [26]. However, pure metallic catalysts are typically expensive.
Ru is regarded as a suitable catalyst for thermochemical NH3 synthesis because of its peak position on Skulason’s volcano diagram, which shows that it requires a minimum potential for electrochemical NH3 synthesis. It has also been reported to be an ultra-efficient electrocatalyst for the NRR, with a lower reduction potential than that of Fe [71,72,73,74,75]. However, Ag is a more cost-effective option because of its natural abundance. Although noble-metal-based electrocatalysts exhibit favorable activity, efficiency, and selectivity, their practical application is inhibited by their high cost and scarcity [27]. Consequently, extensive research has been conducted on transition-metal-based electrocatalysts for the NRR. The NH3 synthesis rate of Pt catalysts can be primarily attributed to their strong HER activity [76,77]. At negative potentials, the surface of Pt nanoparticles tends to adsorb hydrogen atoms rather than nitrogen atoms, thus affecting the overall performance [78].
Considering NH3-producing PCECs, most studies have only focused on the NRR, whereas the OER has been overlooked. The NH3 synthesis reaction is typically performed at 475–600 °C [24]. Pei et al. briefly summarized the OER performance at a cell voltage of 1.3 V and operating temperature of 550 °C [67].
Currently, transition metals, particularly compounds based on Fe, Co, and Ni, have demonstrated remarkable catalytic activity for the OER [79]. A successful method to enhance the OER activity involves altering the surface electronic structure through the addition of supporting materials to the active metal (Fe, Co, or Ni). This strategy has attracted attention, particularly with reference to multi-metal materials such as high-entropy perovskites, because they provide numerous possibilities for modifying the characteristics and improving the catalytic performance [80].
Among multi-metal materials, Co-based double perovskite oxides are notable for their rapid ion diffusion and enhanced surface catalysis, resulting in high electrochemical performance in single cells [35,69,81]. Various studies have explored the application of OER catalysts in PCECs (Table 6). For example, Gd0.3Ca2.7Co3.82Cu0.18O9−δ exhibits the highest current density owing to various factors, including abundant oxygen vacancies in the central Co–O layer of the Ca3Co2O3 rock–salt subsystem, which alters the electronic charge carrier concentration. The needlelike grain morphology aids in the complex flow of the reaction components via triple conduction and open diffusion paths [70].
Despite the promising characteristics of Co-based double perovskite oxides, these high-entropy perovskite oxides have disadvantages such as instability, thermal mismatch, and high cost, which limit their widespread implementation [35,69,81,82]. Thermal mismatch occurs when the OER catalyst and electrolyte materials have different coefficients of thermal expansion (CTE). The CTE indicates the extent to which a material expands when exposed to changes in temperature. If the OER catalyst and electrolyte have significantly different CTE, they may expand at different rates as the temperature changes [83]. Therefore, catalysts with compatible mechanical properties need to be used.

5. Conclusions

This review presented a comprehensive outline of the design strategies for PCECs aimed at enhancing electrochemical NH3 synthesis. The mechanisms of the reactions involved were delineated, and design strategies for PCECs were investigated. This review provides novel insights into catalyst development for the NRR and OER. The following points summarize our findings and recommendations to further develop this technology.
Electrolytes must be further developed in terms of their architecture and thickness. For electrodes, understanding the underlying reaction mechanisms is essential. Co-based double perovskite oxides display rapid ion diffusion and improved surface catalysis, resulting in excellent electrochemical performance in individual cells. However, their application is challenging owing to high-entropy perovskite instability, thermal mismatch, and high cost. Computational analysis is indispensable when investigating the reaction mechanisms, particularly in the context of employing high-entropy perovskites as OER catalysts. Based on this review, we recommend the following research directions:
  • A more scalable approach must be investigated to deposit Fe- and Co-based perovskite electrodes to reduce catalyst wastage.
  • A more complex catalyst must be developed for the NRR because existing materials are not as advanced as OER catalysts.
  • Stability and thermal mismatch issues for the OER must be addressed to decrease wastage and increase cell stability.
We believe that investigating more scalable methods, such as atomic layer deposition, will enhance the positive environmental impact of electrochemical catalysts used for NH2 synthesis. Similar to thermochemical NH3 synthesis catalysts, investigating various support materials for Fe- or Ru-based catalysts can be beneficial for the NRR. For further advancements in OER catalysts, the stability of the catalyst must be improved, and the thermal mismatch must be eliminated to enhance the overall efficiency of the PCEC.

Author Contributions

Conceptualization, H.M.V.; investigation, M.-C.K.; writing—original draft preparation, H.M.V.; writing—review and editing, A.B.; visualization, H.M.V.; supervision, S.H.C.; funding acquisition, S.H.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (NRF), funded by the Korean Government (Ministry of ICT) [Grant No. 2021R1A2C2008662].

Data Availability Statement

Data are contained within the article.

Acknowledgments

We would like to thank the Korea Institute of Science and Technology for facilitating this work.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Teichmann, D.; Stark, K.; Müller, K.; Zöttl, G.; Wasserscheid, P.; Arlt, W. Energy Storage in Residential and Commercial Buildings via Liquid Organic Hydrogen Carriers (LOHC). Energy Environ. Sci. 2012, 5, 9044–9054. [Google Scholar] [CrossRef]
  2. Kandemir, T.; Schuster, M.E.; Senyshyn, A.; Behrens, M.; Schlögl, R. The Haber–Bosch Process Revisited: On the Real Structure and Stability of “Ammonia Iron” under Working Conditions. Angew. Chem. Int. Ed. 2013, 52, 12723–12726. [Google Scholar] [CrossRef] [PubMed]
  3. Haber, F.; Rossignol, R. Le Über Die Technische Darstellung von Ammoniak Aus Den Elementen. Zeitschrift für Elektrochemie und Angew. Phys. Chemie 1913, 19, 53–72. [Google Scholar] [CrossRef]
  4. Aika, K.i.; Hori, H.; Ozaki, A. Activation of Nitrogen by Alkali Metal Promoted Transition Metal I. Ammonia Synthesis over Ruthenium Promoted by Alkali Metal. J. Catal. 1972, 27, 424–431. [Google Scholar] [CrossRef]
  5. IEA Ammonia Technology Roadmap. Available online: https://www.iea.org/reports/ammonia-technology-roadmap (accessed on 4 October 2023).
  6. Brown, T. Feeding Life 2030: The Vision of Fertilizers Europe. Available online: https://www.ammoniaenergy.org/articles/feeding-life-2030-the-vision-of-fertilizers-europe/ (accessed on 4 October 2023).
  7. Shipman, M.A.; Symes, M.D. Recent Progress towards the Electrosynthesis of Ammonia from Sustainable Resources. Catal. Today 2017, 286, 57–68. [Google Scholar] [CrossRef]
  8. Foster, S.L.; Bakovic, S.I.P.; Duda, R.D.; Maheshwari, S.; Milton, R.D.; Minteer, S.D.; Janik, M.J.; Renner, J.N.; Greenlee, L.F. Catalysts for Nitrogen Reduction to Ammonia. Nat. Catal. 2018, 1, 490–500. [Google Scholar] [CrossRef]
  9. Jeerh, G.; Zhang, M.; Tao, S. Recent Progress in Ammonia Fuel Cells and Their Potential Applications. J. Mater. Chem. A 2021, 9, 727–752. [Google Scholar] [CrossRef]
  10. Gunduz, S.; Deka, D.J.; Ozkan, U.S. A Review of the Current Trends in High-Temperature Electrocatalytic Ammonia Production Using Solid Electrolytes. J. Catal. 2020, 387, 207–216. [Google Scholar] [CrossRef]
  11. Liu, F.; Ding, D.; Duan, C. Protonic Ceramic Electrochemical Cells for Synthesizing Sustainable Chemicals and Fuels. Adv. Sci. 2023, 10, e2206478. [Google Scholar] [CrossRef]
  12. Giddey, S.; Badwal, S.P.S.; Kulkarni, A. Review of Electrochemical Ammonia Production Technologies and Materials. Int. J. Hydrogen Energy 2013, 38, 14576–14594. [Google Scholar] [CrossRef]
  13. Garagounis, I.; Kyriakou, V.; Skodra, A.; Vasileiou, E.; Stoukides, M. Electrochemical Synthesis of Ammonia in Solid Electrolyte Cells. Front. Energy Res. 2014, 2, 1. [Google Scholar] [CrossRef]
  14. Wang, B.; Li, T.; Gong, F.; Othman, M.H.D.; Xiao, R. Ammonia as a Green Energy Carrier: Electrochemical Synthesis and Direct Ammonia Fuel Cell—A Comprehensive Review. Fuel Process. Technol. 2022, 235, 107380. [Google Scholar] [CrossRef]
  15. Juangsa, F.B.; Irhamna, A.R.; Aziz, M. Production of Ammonia as Potential Hydrogen Carrier: Review on Thermochemical and Electrochemical Processes. Int. J. Hydrogen Energy 2021, 46, 14455–14477. [Google Scholar] [CrossRef]
  16. Medvedev, D. Trends in Research and Development of Protonic Ceramic Electrolysis Cells. Int. J. Hydrogen Energy 2019, 44, 26711–26740. [Google Scholar] [CrossRef]
  17. Shen, H.; Choi, C.; Masa, J.; Li, X.; Qiu, J.; Jung, Y.; Sun, Z. Electrochemical Ammonia Synthesis: Mechanistic Understanding and Catalyst Design. Chem 2021, 7, 1708–1754. [Google Scholar] [CrossRef]
  18. Trivinho-Strixino, F.; Santos, J.S.; Souza Sikora, M. 3—Electrochemical Synthesis of Nanostructured Materials. In Nanostructures; Da Róz, A.L., Ferreira, M., de Lima Leite, F., Oliveira, O.N.B.T.-N., Eds.; William Andrew Publishing: Norwich, NY, USA, 2017; pp. 53–103. ISBN 978-0-323-49782-4. [Google Scholar]
  19. Droushiotis, N.; Grande, F.D.; Dzarfan Othman, M.H.; Kanawka, K.; Doraswami, U.; Metcalfe, I.S.; Li, K.; Kelsall, G. Comparison between Anode-supported and Electrolyte-supported Ni-CGO-LSCF Micro-tubular Solid Oxide Fuel Cells. Fuel Cells 2014, 14, 200–211. [Google Scholar] [CrossRef]
  20. Iwahara, H. Oxide-Ionic and Protonic Conductors Based on Perovskite-Type Oxides and Their Possible Applications. Solid State Ionics 1992, 52, 99–104. [Google Scholar] [CrossRef]
  21. Choi, S.; Kucharczyk, C.J.; Liang, Y.; Zhang, X.; Takeuchi, I.; Ji, H.I.; Haile, S.M. Exceptional Power Density and Stability at Intermediate Temperatures in Protonic Ceramic Fuel Cells. Nat. Energy 2018, 3, 202–210. [Google Scholar] [CrossRef]
  22. Liu, M.; Yang, L.; Wang, S.; Blinn, K.; Liu, M.; Liu, Z.; Cheng, Z. Enhanced Sulfur and Coking Tolerance of a Mixed Ion Conductor for SOFCs: BaZr0.1Ce0.7y0.2−XYbXO3−δ. Science 2009, 326, 126–129. [Google Scholar] [CrossRef]
  23. Zhu, H.; Kee, R.J. Membrane Polarization in Mixed-Conducting Ceramic Fuel Cells and Electrolyzers. Int. J. Hydrogen Energy 2016, 41, 2931–2943. [Google Scholar] [CrossRef]
  24. Yun, D.S.; Joo, J.H.; Yu, J.H.; Yoon, H.C.; Kim, J.N.; Yoo, C.Y. Electrochemical Ammonia Synthesis from Steam and Nitrogen Using Proton Conducting Yttrium Doped Barium Zirconate Electrolyte with Silver, Platinum, and Lanthanum Strontium Cobalt Ferrite Electrocatalyst. J. Power Sources 2015, 284, 245–251. [Google Scholar] [CrossRef]
  25. Sharma, R.K.; Patel, H.; Mushtaq, U.; Kyriakou, V.; Zafeiropoulos, G.; Peeters, F.; Welzel, S.; Van De Sanden, M.C.M.; Tsampas, M.N. Plasma Activated Electrochemical Ammonia Synthesis from Nitrogen and Water. ACS Energy Lett. 2021, 6, 313–319. [Google Scholar] [CrossRef]
  26. Li, C.I.; Matsuo, H.; Otomo, J. Effective Electrode Design and the Reaction Mechanism for Electrochemical Promotion of Ammonia Synthesis Using Fe-Based Electrode Catalysts. Sustain. Energy Fuels 2021, 5, 188–198. [Google Scholar] [CrossRef]
  27. Liu, D.; Chen, M.; Du, X.; Ai, H.; Lo, K.H.; Wang, S.; Chen, S.; Xing, G.; Wang, X.; Pan, H. Development of Electrocatalysts for Efficient Nitrogen Reduction Reaction under Ambient Condition. Adv. Funct. Mater. 2021, 31, 2008983. [Google Scholar] [CrossRef]
  28. Huang, Y.; Babu, D.D.; Peng, Z.; Wang, Y. Atomic Modulation, Structural Design, and Systematic Optimization for Efficient Electrochemical Nitrogen Reduction. Adv. Sci. 2020, 7, 1902390. [Google Scholar] [CrossRef] [PubMed]
  29. MacLeod, K.C.; Holland, P.L. Recent Developments in the Homogeneous Reduction of Dinitrogen by Molybdenum and Iron. Nat. Chem. 2013, 5, 559–565. [Google Scholar] [CrossRef]
  30. Wang, Y.; Zhou, W.; Jia, R.; Yu, Y.; Zhang, B. Unveiling the Activity Origin of a Copper-Based Electrocatalyst for Selective Nitrate Reduction to Ammonia. Angew. Chemie Int. Ed. 2020, 59, 5350–5354. [Google Scholar] [CrossRef]
  31. Han, J.; Liu, Z.; Ma, Y.; Cui, G.; Xie, F.; Wang, F.; Wu, Y.; Gao, S.; Xu, Y.; Sun, X. Ambient N2 Fixation to NH3 at Ambient Conditions: Using Nb2O5 Nanofiber as a High-Performance Electrocatalyst. Nano Energy 2018, 52, 264–270. [Google Scholar] [CrossRef]
  32. Mars, P.; van Krevelen, D.W. Oxidations Carried out by Means of Vanadium Oxide Catalysts. Chem. Eng. Sci. 1954, 3, 41–59. [Google Scholar] [CrossRef]
  33. Abghoui, Y.; Garden, A.L.; Howalt, J.G.; Vegge, T.; Skúlason, E. Electroreduction of N2 to Ammonia at Ambient Conditions on Mononitrides of Zr, Nb, Cr, and V: A DFT Guide for Experiments. ACS Catal. 2016, 6, 635–646. [Google Scholar] [CrossRef]
  34. Zhou, Y.; Zhang, W.; Kane, N.; Luo, Z.; Pei, K.; Sasaki, K.; Choi, Y.; Chen, Y.; Ding, D.; Liu, M. An Efficient Bifunctional Air Electrode for Reversible Protonic Ceramic Electrochemical Cells. Adv. Funct. Mater. 2021, 31, 2105386. [Google Scholar] [CrossRef]
  35. Ding, D.; Zhang, Y.; Wu, W.; Chen, D.; Liu, M.; He, T. A Novel Low-Thermal-Budget Approach for the Co-Production of Ethylene and Hydrogen via the Electrochemical Non-Oxidative Deprotonation of Ethane. Energy Environ. Sci. 2018, 11, 1710–1716. [Google Scholar] [CrossRef]
  36. Duan, C.; Kee, R.; Zhu, H.; Sullivan, N.; Zhu, L.; Bian, L.; Jennings, D.; O’Hayre, R. Highly Efficient Reversible Protonic Ceramic Electrochemical Cells for Power Generation and Fuel Production. Nat. Energy 2019, 4, 230–240. [Google Scholar] [CrossRef]
  37. Han, D.; Liu, X.; Bjørheim, T.S.; Uda, T. Yttrium-Doped Barium Zirconate-Cerate Solid Solution as Proton Conducting Electrolyte: Why Higher Cerium Concentration Leads to Better Performance for Fuel Cells and Electrolysis Cells. Adv. Energy Mater. 2021, 11, 2003149. [Google Scholar] [CrossRef]
  38. Zhu, J.; Li, X.L.; Wu, C.; Gao, J.; Xu, H.; Li, Y.; Guo, X.; Li, H.; Zhou, W. A Multilayer Ceramic Electrolyte for All-Solid-State Li Batteries. Angew. Chemie Int. Ed. 2021, 60, 3781–3790. [Google Scholar] [CrossRef] [PubMed]
  39. Brett, D.J.L.; Aguiar, P.; Clague, R.; Marquis, A.J.; Schöttl, S.; Simpson, R.; Brandon, N.P. Application of Infrared Thermal Imaging to the Study of Pellet Solid Oxide Fuel Cells. J. Power Sources 2007, 166, 112–119. [Google Scholar] [CrossRef]
  40. Feng, W.; Wu, W.; Jin, C.; Zhou, M.; Bian, W.; Tang, W.; Gomez, J.Y.; Boardman, R.; Ding, D. Exploring the Structural Uniformity and Integrity of Protonic Ceramic Thin Film Electrolyte Using Wet Powder Spraying. J. Power Sources Adv. 2021, 11, 100067. [Google Scholar] [CrossRef]
  41. Skodra, A.; Stoukides, M. Electrocatalytic Synthesis of Ammonia from Steam and Nitrogen at Atmospheric Pressure. Solid State Ionics 2009, 180, 1332–1336. [Google Scholar] [CrossRef]
  42. Yilmaz, S.; Kavici, B.; Ramakrishnan, P.; Celen, C.; Horri, B.A. Highly Conductive Cerium- and Neodymium-Doped Barium Zirconate Perovskites for Protonic Ceramic Fuel Cells. Energies 2023, 16, 4318. [Google Scholar] [CrossRef]
  43. François, M.; Lescure, V.; Heintz, O.; Combemale, L.; Demoisson, F.; Caboche, G. Synthesis of Y-Doped BaZrO3 Proton Conducting Electrolyte Material by a Continuous Hydrothermal Process in Supercritical Conditions: Investigation of the Formation Mechanism and Electrochemical Performance. Ceram. Int. 2023, 49, 25344–25352. [Google Scholar] [CrossRef]
  44. Konwar, D.; Yoon, H.H. A Methane-Fueled SOFC Based on a Thin BaZr0.1Ce0.7Y0.1Yb0.1O3−δ Electrolyte Film and a LaNi0.6Co0.4O3 Anode Functional Layer. J. Mater. Chem. A 2016, 4, 5102–5106. [Google Scholar] [CrossRef]
  45. Fan, Z.; Cao, D.; Zhou, M.; Zhu, Z.; Chen, M.; Liu, J. Barium Cerate-Zirconate Electrolyte Powder Prepared by Carbonate Coprecipitation for High Performance Protonic Ceramic Fuel Cells. Ceram. Int. 2023, 49, 8524–8532. [Google Scholar] [CrossRef]
  46. Wang, S.; Shen, J.; Zhu, Z.; Wang, Z.; Cao, Y.; Guan, X.; Wang, Y.; Wei, Z.; Chen, M. Further Optimization of Barium Cerate Properties via Co-Doping Strategy for Potential Application as Proton-Conducting Solid Oxide Fuel Cell Electrolyte. J. Power Sources 2018, 387, 24–32. [Google Scholar] [CrossRef]
  47. Malešević, A.; Radojković, A.; Žunić, M.; Dapčević, A.; Perać, S.; Branković, Z.; Branković, G. Evaluation of Stability and Functionality of BaCe1−xInxO3−δ Electrolyte in a Wider Range of Indium Concentration. J. Adv. Ceram. 2022, 11, 443–453. [Google Scholar] [CrossRef]
  48. Kane, N.; Luo, Z.; Zhou, Y.; Ding, Y.; Weidenbach, A.; Zhang, W.; Liu, M. Durable and High-Performance Thin-Film BHYb-Coated BZCYYb Bilayer Electrolytes for Proton-Conducting Reversible Solid Oxide Cells. ACS Appl. Mater. Interfaces 2023, 15, 32395–32403. [Google Scholar] [CrossRef] [PubMed]
  49. Kim, Y.S.; Chang, W.; Jeong, H.J.; Kim, K.H.; Park, H.S.; Shim, J.H. High Performance of Protonic Ceramic Fuel Cells with 1-Μm-Thick Electrolytes Fabricated by Inkjet Printing. Addit. Manuf. 2023, 71, 103590. [Google Scholar] [CrossRef]
  50. Bae, K.; Lee, S.; Jang, D.Y.; Kim, H.J.; Lee, H.; Shin, D.; Son, J.W.; Shim, J.H. High-Performance Protonic Ceramic Fuel Cells with Thin-Film Yttrium-Doped Barium Cerate-Zirconate Electrolytes on Compositionally Gradient Anodes. ACS Appl. Mater. Interfaces 2016, 8, 9097–9103. [Google Scholar] [CrossRef]
  51. Kim, D.; Bae, K.T.; Kim, K.J.; Im, H.N.; Jang, S.; Oh, S.; Lee, S.W.; Shin, T.H.; Lee, K.T. High-Performance Protonic Ceramic Electrochemical Cells. ACS Energy Lett. 2022, 7, 2393–2400. [Google Scholar] [CrossRef]
  52. Lee, K.R.; Tseng, C.J.; Jang, S.C.; Lin, J.C.; Wang, K.W.; Chang, J.K.; Chen, T.C.; Lee, S.W. Fabrication of Anode-Supported Thin BCZY Electrolyte Protonic Fuel Cells Using NiO Sintering Aid. Int. J. Hydrogen Energy 2019, 44, 23784–23792. [Google Scholar] [CrossRef]
  53. Choi, S.M.; An, H.; Yoon, K.J.; Kim, B.K.; Lee, H.W.; Son, J.W.; Kim, H.; Shin, D.; Ji, H.I.; Lee, J.H. Electrochemical Analysis of High-Performance Protonic Ceramic Fuel Cells Based on a Columnar-Structured Thin Electrolyte. Appl. Energy 2019, 233–234, 29–36. [Google Scholar] [CrossRef]
  54. Zhong, Z.; Li, Z.; Li, J.; Guo, X.; Hu, Q.; Feng, Y.; Sun, H. A Facile Method to Synthesize BaZr0.1Ce0.7Y0.1Yb0.1O3−δ (BZCYYb) Nanopowders for the Application on Highly Conductive Proton-Conducting Electrolytes. Int. J. Hydrogen Energy 2022, 47, 40054–40066. [Google Scholar] [CrossRef]
  55. Wang, M.; Wu, W.; Lin, Y.; Tang, W.; Gao, G.; Li, H.; Stewart, F.F.; Wang, L.; Yang, Y.; Ding, D. Improved Solid-State Reaction Method for Scaled-Up Synthesis of Ceramic Proton-Conducting Electrolyte Materials. ACS Appl. Energy Mater. 2023, 6, 8316–8326. [Google Scholar] [CrossRef]
  56. Sari, S.N.; Nieroda, P.; Pasierb, P. The BaCeO3-Based Composite Protonic Conductors Prepared by Spark Plasma Sintering (SPS) and Free-Sintering Methods. Int. J. Hydrogen Energy 2023, 48, 29748–29758. [Google Scholar] [CrossRef]
  57. Timurkutluk, C.; Timurkutluk, B.; Kaplan, Y. Experimental Optimization of the Fabrication Parameters for Anode-Supported Micro-Tubular Solid Oxide Fuel Cells. Int. J. Hydrogen Energy 2020, 45, 23294–23309. [Google Scholar] [CrossRef]
  58. Ren, C.; Xu, P.; Zhang, Y.; Liu, T. Understanding the Polymer Binder Effect on the Microstructure and Performance of Micro-Tubular Solid Oxide Fuel Cells with Continuously Graded Pores Fabricated by the Phase Inversion Method. Appl. Surf. Sci. 2023, 612, 155928. [Google Scholar] [CrossRef]
  59. Sun, H.; Zhang, S.; Li, C.; Rainwater, B.; Liu, Y.; Zhang, L.; Zhang, Y.; Li, C.; Liu, M. Atmospheric Plasma-Sprayed BaZr0.1Ce0.7Y0.1Yb0.1O3−δ (BZCYYb) Electrolyte Membranes for Intermediate-Temperature Solid Oxide Fuel Cells. Ceram. Int. 2016, 42, 19231–19236. [Google Scholar] [CrossRef]
  60. Xiao, Y.; Wang, M.; Bao, D.; Wang, Z.; Jin, F.; Wang, Y.; He, T. Performance of Fuel Electrode-Supported Tubular Protonic Ceramic Cells Prepared through Slip Casting and Dip-Coating Methods. Catalysts 2023, 13, 182. [Google Scholar] [CrossRef]
  61. Hou, M.; Zhu, F.; Liu, Y.; Chen, Y. A High-Performance Fuel Electrode-Supported Tubular Protonic Ceramic Electrochemical Cell. J. Eur. Ceram. Soc. 2023, 43, 6200–6207. [Google Scholar] [CrossRef]
  62. Oneill, B.J.; Jackson, D.H.K.; Lee, J.; Canlas, C.; Stair, P.C.; Marshall, C.L.; Elam, J.W.; Kuech, T.F.; Dumesic, J.A.; Huber, G.W. Catalyst Design with Atomic Layer Deposition. ACS Catal. 2015, 5, 1804–1825. [Google Scholar] [CrossRef]
  63. Kosaka, F.; Nakamura, T.; Otomo, J. Electrochemical Ammonia Synthesis Using Mixed Protonic-Electronic Conducting Cathodes with Exsolved Ru-Nanoparticles in Proton Conducting Electrolysis Cells. J. Electrochem. Soc. 2017, 164, F1323–F1330. [Google Scholar] [CrossRef]
  64. Kim, J.; Jun, A.; Gwon, O.; Yoo, S.; Liu, M.; Shin, J.; Lim, T.H.; Kim, G. Hybrid-Solid Oxide Electrolysis Cell: A New Strategy for Efficient Hydrogen Production. Nano Energy 2018, 44, 121–126. [Google Scholar] [CrossRef]
  65. Li, W.; Guan, B.; Ma, L.; Hu, S.; Zhang, N.; Liu, X. High Performing Triple-Conductive Pr2NiO4+δ Anode for Proton-Conducting Steam Solid Oxide Electrolysis Cell. J. Mater. Chem. A 2018, 6, 18057–18066. [Google Scholar] [CrossRef]
  66. Ding, H.; Wu, W.; Jiang, C.; Ding, Y.; Bian, W.; Hu, B.; Singh, P.; Orme, C.J.; Wang, L.; Zhang, Y.; et al. Self-Sustainable Protonic Ceramic Electrochemical Cells Using a Triple Conducting Electrode for Hydrogen and Power Production. Nat. Commun. 2020, 11, 1907. [Google Scholar] [CrossRef]
  67. Pei, K.; Zhou, Y.; Xu, K.; Zhang, H.; Ding, Y.; Zhao, B.; Yuan, W.; Sasaki, K.; Choi, Y.M.; Chen, Y.; et al. Surface Restructuring of a Perovskite-Type Air Electrode for Reversible Protonic Ceramic Electrochemical Cells. Nat. Commun. 2022, 13, 2207. [Google Scholar] [CrossRef] [PubMed]
  68. He, F.; Zhou, Y.; Hu, T.; Xu, Y.; Hou, M.; Zhu, F.; Liu, D.; Zhang, H.; Xu, K.; Liu, M.; et al. An Efficient High-Entropy Perovskite-Type Air Electrode for Reversible Oxygen Reduction and Water Splitting in Protonic Ceramic Cells. Adv. Mater. 2023, 35, e2209469. [Google Scholar] [CrossRef]
  69. Choi, S.; Davenport, T.C.; Haile, S.M. Protonic Ceramic Electrochemical Cells for Hydrogen Production and Electricity Generation: Exceptional Reversibility, Stability, and Demonstrated Faradaic Efficiency. Energy Environ. Sci. 2019, 12, 206–215. [Google Scholar] [CrossRef]
  70. Saqib, M.; Choi, I.G.; Bae, H.; Park, K.; Shin, J.S.; Kim, Y.D.; Lee, J.I.; Jo, M.; Kim, Y.C.; Lee, K.S.; et al. Transition from Perovskite to Misfit-Layered Structure Materials: A Highly Oxygen Deficient and Stable Oxygen Electrode Catalyst. Energy Environ. Sci. 2021, 14, 2472–2484. [Google Scholar] [CrossRef]
  71. Zhao, R.; Liu, C.; Zhang, X.; Zhu, X.; Wei, P.; Ji, L.; Guo, Y.; Gao, S.; Luo, Y.; Wang, Z.; et al. An Ultrasmall Ru2P Nanoparticles-Reduced Graphene Oxide Hybrid: An Efficient Electrocatalyst for NH3 Synthesis under Ambient Conditions. J. Mater. Chem. A 2020, 8, 77–81. [Google Scholar] [CrossRef]
  72. Skúlason, E.; Bligaard, T.; Gudmundsdóttir, S.; Studt, F.; Rossmeisl, J.; Abild-Pedersen, F.; Vegge, T.; Jónsson, H.; Nørskov, J.K. A Theoretical Evaluation of Possible Transition Metal Electro-Catalysts for N2 Reduction. Phys. Chem. Chem. Phys. 2012, 14, 1235–1245. [Google Scholar] [CrossRef]
  73. Tao, H.; Choi, C.; Ding, L.X.; Jiang, Z.; Han, Z.; Jia, M.; Fan, Q.; Gao, Y.; Wang, H.; Robertson, A.W.; et al. Nitrogen Fixation by Ru Single-Atom Electrocatalytic Reduction. Chem 2019, 5, 204–214. [Google Scholar] [CrossRef]
  74. Back, S.; Jung, Y. On the Mechanism of Electrochemical Ammonia Synthesis on the Ru Catalyst. Phys. Chem. Chem. Phys. 2016, 18, 9161–9166. [Google Scholar] [CrossRef]
  75. Montoya, J.H.; Tsai, C.; Vojvodic, A.; Nørskov, J.K. The Challenge of Electrochemical Ammonia Synthesis: A New Perspective on the Role of Nitrogen Scaling Relations. ChemSusChem 2015, 8, 2180–2186. [Google Scholar] [CrossRef] [PubMed]
  76. Mao, Y.J.; Wei, L.; Zhao, X.S.; Wei, Y.S.; Li, J.W.; Sheng, T.; Zhu, F.C.; Tian, N.; Zhou, Z.Y.; Sun, S.G. Excavated Cubic Platinum-Iridium Alloy Nanocrystals with High-Index Facets as Highly Efficient Electrocatalysts in N2 Fixation to NH3. Chem. Commun. 2019, 55, 9335–9338. [Google Scholar] [CrossRef] [PubMed]
  77. Hao, R.; Sun, W.; Liu, Q.; Liu, X.; Chen, J.; Lv, X.; Li, W.; Liu, Y.p.; Shen, Z. Efficient Electrochemical Nitrogen Fixation over Isolated Pt Sites. Small 2020, 16, e2000015. [Google Scholar] [CrossRef] [PubMed]
  78. Singh, A.R.; Rohr, B.A.; Schwalbe, J.A.; Cargnello, M.; Chan, K.; Jaramillo, T.F.; Chorkendorff, I.; Nørskov, J.K. Electrochemical Ammonia Synthesis—The Selectivity Challenge. ACS Catal. 2017, 7, 706–709. [Google Scholar] [CrossRef]
  79. Xia, J.; Huang, K.; Yao, Z.; Zhang, B.; Li, S.; Chen, Z.; Wu, F.; Wu, J.; Huang, Y. Ternary Duplex FeCoNi Alloy Prepared by Cathode Plasma Electrolytic Deposition as a High-Efficient Electrocatalyst for Oxygen Evolution Reaction. J. Alloy. Compd. 2022, 891, 161934. [Google Scholar] [CrossRef]
  80. Li, S.Y.; Nguyen, T.X.; Su, Y.H.; Lin, C.C.; Huang, Y.J.; Shen, Y.H.; Liu, C.P.; Ruan, J.J.; Chang, K.S.; Ting, J.M. Sputter-Deposited High Entropy Alloy Thin Film Electrocatalyst for Enhanced Oxygen Evolution Reaction Performance. Small 2022, 18, e2106127. [Google Scholar] [CrossRef]
  81. Vøllestad, E.; Strandbakke, R.; Tarach, M.; Catalán-Martínez, D.; Fontaine, M.L.; Beeaff, D.; Clark, D.R.; Serra, J.M.; Norby, T. Mixed Proton and Electron Conducting Double Perovskite Anodes for Stable and Efficient Tubular Proton Ceramic Electrolysers. Nat. Mater. 2019, 18, 752–759. [Google Scholar] [CrossRef]
  82. Hu, T.; Zhu, F.; Xia, J.; He, F.; Du, Z.; Zhou, Y.; Liu, Y.; Wang, H.; Chen, Y. In Situ Engineering of a Cobalt-Free Perovskite Air Electrode Enabling Efficient Reversible Oxygen Reduction/Evolution Reactions. Adv. Funct. Mater. 2023, 33, 2305567. [Google Scholar] [CrossRef]
  83. Wan, Y.; Xing, Y.; Li, Y.; Huan, D.; Xia, C. Thermal Cycling Durability Improved by Doping Fluorine to PrBaCo2O5+Δ as Oxygen Reduction Reaction Electrocatalyst in Intermediate-Temperature Solid Oxide Fuel Cells. J. Power Sources 2018, 402, 363–372. [Google Scholar] [CrossRef]
Figure 1. Schematic of the PCEC setup designed for NH3 synthesis, wherein H2O and N2 are transformed into NH3.
Figure 1. Schematic of the PCEC setup designed for NH3 synthesis, wherein H2O and N2 are transformed into NH3.
Energies 17 00441 g001
Figure 2. Questions to be answered through this review.
Figure 2. Questions to be answered through this review.
Energies 17 00441 g002
Figure 3. Potential NRR pathways in PCECs. Adapted from Duan et al. [11]. The solid one is for adsorbed atom while the dotted line is for desorbed atom.
Figure 3. Potential NRR pathways in PCECs. Adapted from Duan et al. [11]. The solid one is for adsorbed atom while the dotted line is for desorbed atom.
Energies 17 00441 g003
Figure 4. Potential OER pathway in PCECs. The solid one is for adsorbed atom while the dotted line is for desorbed atom.
Figure 4. Potential OER pathway in PCECs. The solid one is for adsorbed atom while the dotted line is for desorbed atom.
Energies 17 00441 g004
Figure 5. Illustration of (a) pellet, (b) thin-film, and (c) columnar electrolytes.
Figure 5. Illustration of (a) pellet, (b) thin-film, and (c) columnar electrolytes.
Energies 17 00441 g005
Table 1. Conductivities of proton-conducting electrolytes and their synthesis methods.
Table 1. Conductivities of proton-conducting electrolytes and their synthesis methods.
ElectrolyteMethodConductivity (S cm−1)Thickness (mm)Reference
SrCe0.95Yb0.05O3−δsol–gelUnknown1.5[41]
BaZr0.8−x−yCexNdyY0.1Yb0.1O3−δPechini method500 °C: 3.77 × 10−40.8–1.5[42]
BaZr0.85Y0.15O3−δhydrothermal process600 °C: 2.5 × 10−31.6[43]
Table 2. Conductivities of thin-film electrolytes and their deposition/synthesis methods.
Table 2. Conductivities of thin-film electrolytes and their deposition/synthesis methods.
ElectrolyteMethodConductivity (S cm−1)Thickness (µm)Reference
BaCe0.7Zr0.1Y0.2co-precipitation
solid-state reaction
dip-coating
650 °C: 2.8 × 10−2~20[45]
BaCe0.8Y0.2−xNdxO3−δcitrate–nitrate combustion350 °C: 8.5 × 10−3~20[46]
BaCe1−xInxO3−δauto-combustion reaction700 °C: 5 × 10−320–25[47]
BaZr0.1Ce0.7Y0.1Yb0.1solid-state reaction500 °C: 1.2 × 10−210[48]
BaHf0.8Yb0.2O3−δpulsed laser deposition (PLD)500 °C: 2.5 × 10−3110[48]
BaZr0.1Ce0.7Y0.1Yb0.1solid-state reaction500 °C: 1.3 × 10−2~10[22]
BaZr0.4Ce0.4Y0.1Yb0.1solid-state reaction500 °C: 5.6 × 10−3~15[21]
BaZr0.2Ce0.6Y0.1Yb0.1O3−δPechini method
inkjet printing
600 °C: 24.391[49]
BaCe0.5Zr0.35Y0.15O3−δcitric nitrate method
PLD
Unknown2–4[50]
BaZr1−x−yCexYyO3ultrafast microwave-assisted sintering
tape casting
Unknown~12[51]
BaZr0.2Ce0.6Y0.2O3solid-state reaction
spin coating
800 °C: 1 × 10−2~7[52]
BaCe0.55Zr0.3Y0.15O3−δscreen printingUnknown~2.5[53]
Table 3. Conductivities of columnar electrolytes and their deposition/synthesis method.
Table 3. Conductivities of columnar electrolytes and their deposition/synthesis method.
ElectrolyteMethodConductivityReference
BaZr0.4Ce0.4Y0.15Zn0.05O3solid-state reactionUnknown[60]
BaZr0.1Ce0.7Y0.1Yb0.1solid-state reactionUnknown[61]
Table 4. Deposition methods and electrode thickness in key research on OER and NRR catalysts.
Table 4. Deposition methods and electrode thickness in key research on OER and NRR catalysts.
CathodeDeposition MethodThickness
(µm)
Reference
La0.6Sr0.4Co0.2Fe0.8O3−δUnknown44[24]
Ag-4
Pt-8
Fedoctor blade15–25[26]
10-Fe-BCYdoctor blade15–25
0.5W-10Fe-BCYdoctor blade15–25
PrBa0.5Sr0.5Co1.5Fe0.5O5+δ-10–20[36]
Ru–Ag/MgOUnknown-[41]
Ni-BCYR--[63]
NdBa0.5Sr0.5Co1.5Fe0.5O5+δ (NBSCF)-BZCYYbdrop coating15[64]
Pr2NiO4-BZCYscreen printing13[65]
PrCo0.05Ni0.5O3−δtape casting29[66]
Ba0.9Co0.7Fe0.2Nb0.1O3−δscreen printing15[67]
Pr0.2Ba0.2Sr0.2La0.2Ca0.2CoO3−δspray coating20[68]
PrBa0.5Sr0.5Co1.5Fe0.5O5+δPLD20[69]
Gd0.3Ca2.7Co3.82Cu0.18O9−δscreen printing30[70]
Table 5. Notable NRR catalysts in PCECs for NH3 synthesis.
Table 5. Notable NRR catalysts in PCECs for NH3 synthesis.
Cathode *ElectrolyteNH3 Production Rate [mol cm−2 s−1] × 10−9 Thickness
(µm)
Reference
La0.6Sr0.4Co0.2Fe0.8O3−δBaZr0.8Y0.2O3−δ0.085044[24]
AgBaZr0.8Y0.2O3−δ0.04904
PtBaZr0.8Y0.2O3−δ<0.00108
FeBaCe0.9Y0.1O3−δ14.00015–25[26]
10-Fe-BCYBaCe0.9Y0.1O3−δ0.420015–25
0.5W-10Fe-BCYBaCe0.9Y0.1O3−δ0.570015–25
Ru–Ag/MgOSrCe0.95Yb0.05O3−δ0.0003-[41]
Ni-BCYRBaCe0.9Y0.1O3−δ0.0110-[63]
* only for OER coupled with NRR configuration.
Table 6. Notable OER catalysts in PCECs for NH3 synthesis.
Table 6. Notable OER catalysts in PCECs for NH3 synthesis.
AnodeElectrolyteCurrent Density @1.3 V and 550 °C
[A cm−2]
Thickness
(µm)
Reference
Pr0.2Ba0.2Sr0.2La0.2Ca0.2CoO3−δBaZr0.1Ce0.7Y0.1Yb0.1O3−δ−0.80020[68]
PrBa0.5Sr0.5Co1.5Fe0.5O5+δBaZr0.4Ce0.4Y0.1Yb0.1O3−δ−1.05920[69]
Gd0.3Ca2.7Co3.82Cu0.18O9−δBaZr0.1Ce0.7Y0.1Yb0.1O3−δ−1.24130[70]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vieri, H.M.; Kim, M.-C.; Badakhsh, A.; Choi, S.H. Electrochemical Synthesis of Ammonia via Nitrogen Reduction and Oxygen Evolution Reactions—A Comprehensive Review on Electrolyte-Supported Cells. Energies 2024, 17, 441. https://doi.org/10.3390/en17020441

AMA Style

Vieri HM, Kim M-C, Badakhsh A, Choi SH. Electrochemical Synthesis of Ammonia via Nitrogen Reduction and Oxygen Evolution Reactions—A Comprehensive Review on Electrolyte-Supported Cells. Energies. 2024; 17(2):441. https://doi.org/10.3390/en17020441

Chicago/Turabian Style

Vieri, Hizkia Manuel, Moo-Chang Kim, Arash Badakhsh, and Sun Hee Choi. 2024. "Electrochemical Synthesis of Ammonia via Nitrogen Reduction and Oxygen Evolution Reactions—A Comprehensive Review on Electrolyte-Supported Cells" Energies 17, no. 2: 441. https://doi.org/10.3390/en17020441

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop