Enhancement of Ammonia Synthesis on a Co3Mo3N-Ag

Aug 16, 2017 - The electrochemical promotion of ammonia synthesis by potassium ions (K+) on a Co3Mo3N-Ag electrocatalyst was studied in a K-β″-Al2O...
3 downloads 9 Views 1MB Size
Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES

Article

Enhancement of Ammonia Synthesis on a Co3Mo3NAg Electrocatalyst in a K-#Al2O3 Solid Electrolyte Cell Javier Díez-Ramírez, Vasileios Kyriakou, Ioannis Garagounis, Anastasios Vourros, Eirini Vasileiou, Paula Sanchez, Fernando Dorado, and Michael Stoukides ACS Sustainable Chem. Eng., Just Accepted Manuscript • DOI: 10.1021/ acssuschemeng.7b01618 • Publication Date (Web): 16 Aug 2017 Downloaded from http://pubs.acs.org on August 17, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Sustainable Chemistry & Engineering is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Enhancement of Ammonia Synthesis on a Co3Mo3N-Ag Electrocatalyst in a K-βAl2O3 Solid Electrolyte Cell J. Díez-Ramírez1, V. Kyriakou2,3, I. Garagounis2,3, A. Vourros2,3, E. Vasileiou2,3, P. Sánchez1, F. Dorado1, M. Stoukides2,3,*

1

Departamento de Ingeniería Química, Facultad de Ciencias y Tecnologías Químicas,

Avenida Camilo José Cela 12, 13071 Ciudad Real, Spain. 2

Department of Chemical Engineering, Aristotle University, Thessaloniki, 54124,

Greece 3

Chemical Processes & Energy Resources Institute, CERTH, Thessaloniki, 56071,

Greece

*Corresponding author. Tel.: +30 2310 996165; Fax:+30 2310 996145; E-mail address: [email protected]

1 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract The electrochemical promotion of ammonia synthesis by potassium ions (K+) on a Co3Mo3N-Ag electrocatalyst was studied in a K-β"-Al2O3 solid electrolyte cell. The effect of temperature (400-550 oC), PH2/PN2 feed composition (1.0, 3.0 and 6.0) and applied voltage was explored in detail. The catalyst was prepared by ammonolysis of the mixed oxide and was characterized by XRD and SEM. A volcano-type behavior was found, i.e., around 1% of potassium per total moles of Co3Mo3N improved the ammonia formation rate by as much as 48%. However, high percentages of potassium act as poison for the reaction, possibly due to the formation of K-N-H compounds which block the active sites. Faradaic efficiency (Λ) values close to 300 are for the first time reported in NH3 synthesis.

Keywords Ammonia synthesis, nitride catalysts, Electrochemical Promotion, K-β"-Al2O3, cation EPOC

Research highlights - Potassium ions promoted the Co3Mo3N-Ag electrocatalyst during ammonia synthesis. - Volcano-type behavior with voltage is observed for the ammonia formation rate. - Faradaic efficiency values up to 300 are reported for the first time in NH3 synthesis.

2 ACS Paragon Plus Environment

Page 2 of 26

Page 3 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Introduction Ammonia is one of the most important chemicals with a worldwide production of 146 million tons in 2015 [1]. Its uses cover a broad range of different industrial sectors, such as fertilizers, refrigeration, explosives and pharmaceuticals [2]. In the past years, the quest for a clean energy future has also made ammonia an ideal candidate as an energy carrier, due to its high hydrogen density and zero carbon emissions during oxidation [3]. Nowadays, ammonia is produced via the well-known Haber-Bosh process, which was developed one hundred years ago and is considered one of the most significant scientific discoveries of mankind [4]. In this energy-intensive process, ammonia is produced from its elements at temperatures between 400-500 ºC and at pressures of 100-150 bars [2,4,5],  + 3 ↔ 2

(1)

The high pressure is dictated by the reaction stoichiometry according to the Le Chatelier principle. On the other hand, the operating temperature is a trade-off solution between faster reaction rates and higher equilibrium conversion. The catalyst usually employed, is a potassium promoted iron oxide (magnetite) [4]. Aside from the widely-used Febased catalyst, Ru-based materials are considered the most effective for the process [6,7]. However, their prohibitive cost and the limited availability of Ru prevent their widespread application in industry.

More recently, nitrides and specifically Co3Mo3N have been identified as promising catalytic composites for ammonia synthesis [8]. The good performance of Co3Mo3N is mainly attributed to the ability of molybdenum to dissociate the dinitrogen molecule [9,10]. The activity of molybdenum increases significantly when it is alloyed with iron, nickel or cobalt (which have excellent hydrogen adsorption properties), with the latter exhibiting the optimum behavior [10]. Furthermore, if an alkali metal, such as potassium [10-12], rubidium [11] or cesium [8,11] is added to the nitride, the reaction rate is further enhanced. As to the reaction mechanism of these promoted nitrides, it is considered that initially an electron is transferred to the active center from the alkali 3 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

promoter and then to the adsorbed dinitrogen, which then dissociates to atomic nitrogen [10]. Kojima et al. reported that 2% of cesium and 5% of potassium were the optimum amounts of alkali addition. The use of higher quantities of the alkali promoter led to the decomposition of the Co3Mo3N active phase into Co and Mo2N and the loss of catalytic activity [10].

On the other hand, chemical reactions can also be promoted electrochemically by employing the catalysts as working electrodes in a solid electrolyte cell reactor (SECR) [13]. Numerous studies of Electrochemical Promotion of Catalysis (EPOC) with different electrochemical promoters (Oδ-, Hδ+, Naδ+, Kδ+) have been conducted in the last three and half decades [14-16]. Specifically, in ammonia synthesis, EPOC, also called the Non-Faradaic Electrochemical Modification of Catalytic Activity (NEMCA) effect, has been reported in proton conducting cells [17, 18]. An impressive 1300% increase in ammonia formation and a maximum faradaic efficiency of 600% (Λ = 6) has been reported by Yiokari et al. [17] in a high pressure (operating at 50 bars) electrochemical reactor, when protons were pumped to the catalyst surface [17]. The results were attributed to the supply of the electropositive protons to the catalyst surface, which enhanced the chemisorptive binding energy of the electron acceptor, nitrogen [17]. It should be pointed out, however, that Yiokari et al. observed this significantly high NEMCA effect when the conversion to ammonia was much lower than that corresponding to thermodynamic equilibrium. In the opposite case, a very weak effect was observed [18]. This is because the electrochemically supplied protons not only modify the catalytic properties of the electrode, but also serve as the carriers of the electrical power required for the nonspontaneous ammonia synthesis reaction [19].

The effect of electrochemical promotion is more clearly observed when the promoting species are neither reactants nor products of the catalytic reaction. In the last two decades, several catalytic reactions have been shown to be electrochemically promoted in alkaline SECRs [15,16,20]. In these studies, some unique advantages of the electrochemical vs the chemical promotion of alkali metals have been identified, which include among others, the in situ optimization of the alkali promoter coverage and tuning of product selectivities, as well as the permanent promotion effect [21].

4 ACS Paragon Plus Environment

Page 4 of 26

Page 5 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Despite the aforementioned advantages, to this day, no alkaline cationic conductor has been employed for the study of EPOC in ammonia synthesis. Hence, the aim of this work is to investigate the electrochemical enhancement of the ammonia synthesis rate on a Co3Mo3N nitride catalyst by using a K-β"-Al2O3 potassium conductor under atmospheric total pressure. Potassium is used because it constitutes a common promoter for ammonia synthesis. The Co3Mo3N was mixed with Ag in order to improve its adherence with the solid electrolyte and its electronic conductivity so it can be used as an electrode. The effect of various parameters, such as the H2/N2 feed ratio, the operating temperature and the applied potential is examined and discussed.

Experimental Catalyst preparation The catalyst was prepared according to Hargreaves and McKay [22]. The intermediate oxide was obtained through the co-precipitation method. An aqueous solution of cobalt nitrate (Co(NO3)2, Merck) was mixed with an aqueous solution of ammonium dimolybdate ((NH4)2Mo2O7, Alfa Aesar) in the proper molar ratio (Co/Mo = 1). The solution was heated under stirring until all the water was evaporated and a dark purple powder was received. Then, the powder was dried at 100 oC overnight and calcined in static air at 500 oC for 6 h to obtain the desired oxide. In order to obtain the Co3Mo3N nitride, the oxide powder was milled in a roll-mill for 24 h (HDPE vessel, YSZ balls of 5 mm diameter, ethanol) followed by high-temperature ammonolysis [22] (100 mL·min1

, 99.99% NH3-Air Liquide Hellas). At room temperature, the flow was switched from

ammonia to pure nitrogen (100 mL·min-1) for 1 h. The bulk oxidation was prevented by overnight treatment with 0.1% O2 in N2, prior to its exposure to the atmosphere. Cell construction The electrochemical single chamber cell (Figure 1) consisted of a 19-mm-diameter, 1mm-thick K-β"-Al2O3 pellet (Ionotec) as the solid electrolyte (SE). In the electrochemical cell the Au counter (CE) and reference (RE) electrodes were first deposited on the one side of the electrolyte by firing a gold organometallic paste (Fuel Cell Materials, ref-231001) in two consecutive stages, firstly at 300 ºC for 1 h and secondly at 800 ºC for 2 h (heating and cooling ramps of 5 ºC/min). The working electrode (WE) was prepared by mixing the as-prepared Co3Mo3N catalyst with Ag (at a weight ratio of Co3Mo3N/Ag = 60/40). The Co3Mo3N-Ag powder was mixed with an 5 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ink vehicle (terpineol based, Fuel Cell Materials) until a homogenous paste was formed and was screen printed (40 mg) on the K-β"-Al2O3. The electrode was heated at 650 oC in nitrogen for 45 minutes. Gold wires were used for current collection and connection of the cell with the external circuit.

Figure 1. Schematic diagram of the single-chamber K+-conducting cell reactor used for the electrochemical promotion of ammonia synthesis.

Characterization techniques A Phenom ProX scanning electron microscope was used to acquire images of the Co3Mo3N-Ag electrode. This instrument was also equipped with an energy dispersive X-ray spectroscopy (EDX) analyzer to determine the average composition of the samples. The X-ray diffractograms were obtained using a Philips X’Pert instrument using nickel-filtered Cu Kα beam operating at 40 kV, 40 mA and a scan rate of 0.02° in the 2θ range of 30°-60°.

All the electrochemical measurements were carried out using a VersaSTAT 4 electrochemical workstation (Princeton Applied Research) and the corresponding VersaStudio software for data processing. Cyclic voltammograms were obtained in the range of UWR = -2.0 to + 2.0 V with a scan rate of 50 mV/s. Three to five cycles are necessary before two consecutive ones are identical. Before any electrochemical promotion measurement, the WE was “cleaned” of potassium ions by applying a UWR of 2.0 V for 2 h (the current reached zero).

Ammonia measurements The experiments were carried out under atmospheric pressure, at temperatures between 400 and 550 oC and with a total flowrate of 75 mL·min-1. Under these operating

6 ACS Paragon Plus Environment

Page 6 of 26

Page 7 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

conditions N2 conversions were differential (typically less than 0.15%). The reactant gasses were H2 and N2 (99.999% purity, Air Liquide Hellas) and their flows were controlled with mass-flow controllers (Bronkhorst High Tech). The ammonia concentration in the outlet stream of the reactor was measured by two independent methods; one photometric and one with a real-time analyzer. The main method of measuring the ammonia was via an EAA-24r-EP online, real-time analyzer (Los Gatos Research), with a Cavity Ringdown Spectroscopy (CRDS) technique. These measurements were confirmed by a photometric method in which the products were bubbled through a H3BO3 solution (10 cm3 and 4.5 pH) to collect the produced ammonia. The solution was then measured in a Pcompact photometric analyzer (Aqualytic), at a wavelength of 660 nm.

EPOC parameters In the electrochemical promotion of catalysis there is a correspondence between the different potentials applied and the amount of ions electrochemically transferred to the metal film. Several parameters are commonly used to quantify the EPOC effect:

The rate enhancement ratio, ρ, defined by Eq. 1:

ρ = r/r0

(2)

where r is the electropromoted catalytic rate and r0 is the unpromoted rate. The apparent Faradaic efficiency, Λ, defined by Eq. 3:

Λi = ∆rcatalytic/(I/F)

(3)

where ∆rcatalytic is the current- or potential-induced observed change in catalytic rate and I is the average current. In this work, the latter is calculated by dividing the total charge transported at each potensiostatic step by the duration of the step. For this reaction system, it implies:

Λ NH3 = 3∆r NH3/(I/F) where 3∆r NH3 is given in mol/s of H+. 7 ACS Paragon Plus Environment

(4)

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 26

The promotion index (PIK+), calculated by the equation: ∆ 

=



(5)

where θK+ is the potassium to catalyst molar ratio calculated from the integration of the current (I) versus time (t) curves according to Faraday’s law:  ││

θK+= #

 !"

(6)

where n is the potassium ion charge, i.e., +1, F is the Faraday constant (96484.6 C/mol) and NG is the moles of Co3Mo3N.

Results and Discussion Cell characterization SEM micrographs and EDX analysis of the Co3Mo3N-Ag catalyst-electrode are shown in Figures 2a and 2b, respectively. The images were obtained after exposing the sample to a PH2/PN2 = 3.0 mixture at 500 oC for 24 h. It can be seen that the electrode shows a relatively homogeneous structure with particle diameters between 70 and 150 nm. The electrode also has a high porosity, which facilitates the diffusion of the reactants and products.

8 ACS Paragon Plus Environment

Page 9 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

Figure 2. SEM images (a) and elemental mapping (b) of the Co3Mo3N-Ag electrode after exposure to reaction conditions for 24 h. (T = 500 oC, PH2/PN2 = 3.0). Moreover, a uniform distribution of the catalyst elements without significant particle agglomeration was verified by means of EDX analysis (Figure 2b). The crystallography of the Co3Mo3N-Ag electrode was examined by XRD analysis (Figure 3). The main peaks observed correspond to the cubic phase of Co3Mo3N (JCPDS No. 04-008-1301) [23] and the cubic phase of metallic silver (JCPDS No. 87-0720). The average particle size for the Co3Mo3N catalyst on the K-β”-Al2O3 solid electrolyte was 103 nm, while that of Ag is 185 nm. These values were calculated according to the Scherrer method [24] and are reasonably close to those observed in the SEM images of Figure 2.

9 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering



* Co3Mo3N

Intensity (a.u.)

♦ Ag

*

* 30

* 35



*

*

* 40

45

θ

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 26

50

55

60

2 (º) Figure 3. X-ray diffraction pattern of the Co3Mo3N-Ag electrode after exposure to reaction conditions for 24 h. (T = 500 oC, PH2/PN2 = 3.0). Electrochemical promotion studies The electrochemical studies with the Co3Mo3N-Ag/β-Al2O3/Au cell were carried out at temperatures between 400 and 550 ºC for PH2/PN2 = 1.0, 3.0 and 6.0. First, the fresh sample was treated with reactant gas (PH2/PN2 = 3.0, T = 500 oC) for 6 h to eliminate oxygen compounds originating from the passivation step and to activate the catalyst surface [10]. Moreover, prior to all studies, a constant cell voltage of UWR = 2.0 V was applied for 2 h in order to clean the catalyst surface from potassium ions. Then, a new lower voltage was applied and usually after 30-40 min a new steady-state was achieved and ammonia was measured again. Figure 4 shows the effect of the reaction temperature on the rate of NH3 formation (rNH3) under unpromoted (UWR = 2.0 V) and promoted conditions (UWR = 1.0 V). The Co3Mo3N-Ag electrode was very active for ammonia synthesis, with rates as high as 8 x 10-4 mol·g-1·h-1 observed at 500 oC. Under unpromoted conditions, i.e. UWR = 2.0 V, the optimum ratio for ammonia synthesis was PH2/PN2 = 1. This is not unusual since the ammonia rates, while high, are still far from those corresponding to the equilibrium conversions and thus the optimum feed ratio can be different from the stoichiometric. Upon application of a promoting potential (UWR = 1.0 V), the formation rates of ammonia are improved for ratios of PH2/PN2 = 1 and 3 and at temperatures higher than 450 ºC. A maximum in rNH3 was observed at 500 oC both for the promoted and unpromoted catalyst-electrode. The latter observation is due to the 10 ACS Paragon Plus Environment

Page 11 of 26

fact that the reaction rate is controlled by kinetics up to 500 oC, but at higher temperatures the reverse reaction, i.e. NH3 decomposition, dominates, according to the Le Chatelier principle [25].

1000

PH2/PN2= 1 PH2/PN2= 3

800

μ

r N H 3 , m ol g -1 h -1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

PH2/PN2= 6 600

400

200

1V (Promoted)

2V (Unpromoted) 0 400

425

450

475

500

525

550

o

425

450

475

500

525

550

o

Temperature, C

Figure 4.

400

Temperature, C

Effect of reaction temperature on the rate of NH3 synthesis under

unpromoted (UWR = 2.0 V) and promoted (UWR = 1.0 V) conditions. The electrocatalytic activity of the working electrode was studied by varying the applied voltage from UWR = + 2.0 to -0.4 V. The effect of the cell voltage (bottom axis) and potassium to catalyst ratio (top axis) on the ammonia formation rate expressed in mol·g1

·h-1, is shown in Figure 5 for PH2/PN2 = 3.0 and for temperatures between 400 and 500

o

C. It can be seen that the catalyst exhibits a volcano-type behavior with the cell

potential. The highest activity is achieved between 1.25 and 0.95 V vs RE, which correspond to a K+ to Co3Mo3N ratio between 0.01 and 0.015. In other words, by pumping a small amount of potassium ions to the Co3Mo3N-Ag electrode, its catalytic activity toward ammonia is significantly enhanced. Specifically, at 450 oC, rNH3 increased from 214 to 317 µmol·g-1·h-1 and at 500 oC it increased from 640 to 817 µmol·g-1·h-1. The observed increase is attributed to the decrease of the catalyst work function which weakens the chemical bond between the catalyst active sites and the electron donor adsorbate (H2) and at the same time strengthens the bond with electron acceptor (N2) [17]. However, when the cell voltage decreases further, a decrease in rNH3 is observed.

There are two possible reasons for this "poisoning" effect: (a) a

considerable fraction of the active surface area is occupied by the alkali species, which inhibits adsorption and/or reaction of the gaseous species, (b) “over-promotion” of the 11 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

catalyst-electrode, i.e. an excessive strengthening of the N-catalyst site bond which then prevents H2 adsorption. While the former is supported by the cyclic voltammograms of Figure 10, there is as yet no evidence that the latter does not also contribute to the observed decline in reaction rate.

θK+ 0.00

900

0.01

0.02

0.03

400 oC 450 oC 500 oC

600 450 300

No K+ coverage

750

rNH3, µmol g-1 h-1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 26

150 0 2.0

1.5

1.0

0.5

0.0

-0.5

UWR, V Figure 5. Effect of the cell voltage (bottom axis) and potassium to catalyst ratio (top axis) on the ammonia formation rate in µmol·h-1·g-1 for PH2/PN2 = 3.0. The volcano-type behavior for alkali addition has also been observed in catalytic studies of ammonia synthesis. Kojima et al. [10,26,27] studied cobalt molybdenum bimetallic nitride catalysts in great detail and showed that the addition of up to 5 mol% of potassium led to a significant increase in the ammonia production [10]. Higher quantities of impregnated potassium, however, led to surface structural changes detrimental to ammonia synthesis. Yunusov et al. [12], also reported a volcano-type behavior when varying the alkali content of potassium carbonyl ruthenate catalysts. The authors demonstrated that the increased electron density on Ru atoms due to potassium addition was beneficial to dinitrogen activation [12].

Figure 6 shows the effect of the cell voltage (bottom axis) and potassium to Co3Mo3N (top axis) on rNH3 at 450 oC, for various PH2/PN2 feed ratios (PH2/PN2 = 1.0, 3.0 and 6.0). When the catalyst surface was clear of K+ (UWR = 2.0 V) the highest rate of 220 µmol·g1

·h-1 is observed for an equimolar feed composition. Under potassium pumping

12 ACS Paragon Plus Environment

Page 13 of 26

conditions, the maximum rate for ammonia synthesis increases to 317 µmol·g-1·h-1, but this maximum rate is attained at a feed ratio of PH2/PN2 = 3.0.

θK+ 0.00

350

0.01

0.02

0.03

PH2/PN2= 1 300

250

PH2/PN2= 3

No K + coverage

rNH3, µmol g-1 h-1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

PH2/PN2= 6

200

150 2.0

1.5

1.0

0.5

0.0

-0.5

UWR, V Figure 6. Dependence of rNH3 on the applied voltage and potassium to catalyst ratio at 450 ºC, for various PH2/PN2 ratios. In general, the voltage at which the maximum ammonia rates are obtained, is shifted to lower values (1.19 V, 0.98 V and 0.82 V for PH2/PN2 = 1.0, 3.0 and 6.0, respectively) when the inlet H2 concentration increases. As was described above, by pumping potassium ions to the working electrode, the adsorption of the N2 molecule (electron acceptor) is enhanced at the expense of H2 (electron donor). Thus, on a surface where adsorbed hydrogen is the dominant species, more K+ must be pumped in order to remove these species from the surface (higher potassium to catalyst ratio, or more negative voltages) and facilitate the (dissociative) adsorption of N2. The effect of electrochemical promotion presented here can be compared to conventional catalytic alkali promotion. Kojima et al. [10] improved their Co3Mo3N performance by 33% (from 650 to 869 µmol·h-1·g-1 at 400 ºC for PH2/PN2 = 3.0) with an impregnation of 5 mol% of potassium in the nitride. In this work, the maximum formation rate obtained was slightly higher, reaching 915 µmol·h-1·g-1, and was achieved at 500 ºC for PH2/PN2 = 1.0, corresponding, however, to a rate enhancement of less than 20%. On the other hand, an up to 48% improvement of the formation rate was

13 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

achieved at 450 ºC where the unpromoted ammonia formation was quite lower (reached 317 from 214 µmol·h-1·g-1 at PH2/PN2 = 3.0). The main advantage of the electrochemical system, however, resides is the fact that the promoter addition can be controlled in situ and reversibly. Furthermore, the quantity of promoter can be adjusted depending on the conditions used. In the present work, the different conditions (temperature and PH2/PN2 ratio) strongly affected the amount of potassium needed to reach the highest values of ammonia production. The ability to actively control the percentage of promoter on the catalyst surface can be very important in decentralized applications where feed conditions may vary widely and the catalyst will have to be adapted to them.

350 2V

1.3V 1.15V 1V 0.8V 0.6V 0.4V 0V -0.4V

2V

300

250

3

rNH , µmol g-1 h-1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 26

200

150 0

200

400

600

800

Time on stream, min Figure 7. Dynamic response of rNH3 to step changes in the applied voltage, given at the top of the graph, at 450 ºC and inlet PH2/PN2 = 3. The response of the NH3 formation rate with time, for a typical experiment is presented in Figure 7). The data in this case were obtained at a temperature of 450 oC and under a stoichiometric hydrogen to nitrogen feed ratio (3/1). For the first two hours, 2 V are applied across the cell to completely remove K+ from the working electrode, and a steady rNH3 of 214 µmol·h-1·g-1 is obtained. The voltage is then decreased to 1.3 V and a substantial increase in NH3 formation is observed. After about 1 h, the voltage is again reduced and a further increase in rNH3 is observed. This continues until the applied potential reaches 1 V, where rNH3 = 317 µmol·h-1·g-1 (an increase of 48%), after which 14 ACS Paragon Plus Environment

Page 15 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

any further reduction in the voltage leads to lower ammonia rates. At a voltage somewhere between 0.4 and 0 V, the NH3 formation rate drops below the initial steady value of 214 µmol·h-1·g-1, and it becomes even lower at -0.4 V. However, when a potential of 2 V is once again applied to the cell the ammonia rate quickly returns its original value, confirming the reversibility of the process.

EPOC parameters In order to quantify the EPOC effect in studies that were conducted at different conditions, the rate enhancement ratio (ρ), the promotion index (PIK+) and the apparent faradaic efficiency (ΛNH3) were evaluated. Figures 8 and 9 show the effect of applied potential (UWR) and potassium to catalyst ratio (θK+) on ρ, PIK+ and ΛNH3, at various temperatures (400-500 ºC) and reactant compositions (1 < PH2/PN2 < 6). The highest ρ value was 1.5 and it was obtained at 450 ºC for PH2/PN2 = 3.0. Under these conditions, the PIK+ values were around 60. The promotion index is positive for promoters and negative for poisons [28]. In this sense, as was discussed above, when the potential applied decreases beyond the optimum, the potassium ions begin to act as a poison for the catalyst. The highest ΛNH3 was 310 and was obtained at 500 ºC for a feed ratio of 1. Of these three parameters, the promotion index, PIK+, is the most representative in studies like this one, i.e. when a cationic promoter (K+) is used. The physicochemical meaning of this parameter is the increase in the reaction rate upon a differential increase in the surface coverage of the promoting species (in our case of the potassium to catalyst molar ratio) [14]. Figures 8 and 9 show that an increase in θK+ from 0 to about 0.01, has a positive effect on the reaction rate at which point a maximum rate is attained. When further increasing θK+, a decrease in the reaction rate is observed because the catalyst surface is "poisoned" by the addition of excessive amounts of the promoting species. On the other hand, in studies such as this where the promoting ion does not take part in the electrode reaction Λ is a less meaningful parameter than PIK+ [14,29,30]. Although, in this work it has been defined somewhat differently and essentially represents the efficiency of the moles of K+ transported to the catalyst, in promoting NH3 formation. It also helps in the comparison with other works in electrochemical ammonia synthesis.

15 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

θK

+

0.01

1.8

0.02

0.03

(a) 1.6

ρ

1.4 1.2 1.0 0.8 70

(b)

60

400 oC 450 oC 500 oC

PIK

+

50 40 30 20 10 0 300

(c)

250 200

ΛNH3

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

150 100 50 0 1.5

1.0

0.5

0.0

-0.5

UWR, V

Figure 8. Dependence of (a) the rate enhancement ratio (ρ), (b) the promotion index (PIK+) and (c) the apparent faradaic efficiency (ΛNH3) on the applied potential (UWR) and potassium to catalyst ratio (θK+) for PH2/PN2 = 3.0 at T = 400, 450 and 500 oC. A large number of papers on the electrochemical synthesis of ammonia have been published in the past two decades [2, 5, 7]. Nevertheless, in only a few of them a gaseous mixture of hydrogen and nitrogen, rather than pure nitrogen, was introduced at the cathode. Therefore, very few studies focused on the EPOC effect either for the 16 ACS Paragon Plus Environment

Page 16 of 26

Page 17 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

forward reaction of NH3 synthesis [17, 18, 31,32] or for the reverse reaction of NH3 decomposition [33-35]. Yiokari et al. [17], used a single-chamber reactor and studied the reaction on a commercial Fe catalyst. Upon “pumping” H+ to the catalyst surface, the reaction rate could increase by as much as 1300% (ρ = 14). When H+ were removed from the catalyst, a complete, though reversible, deactivation of the catalyst was observed [17]. A very weak effect (Λ < 3, ρ < 2) was observed by Marnellos et al [18] and by Ouzounidou et al [32], who studied the reaction on Pd and Fe electrodes, respectively. It should be pointed out, however, that in the work of Yiokari et al, the conversions both with and without applying the voltage, were well below those predicted by thermodynamic equilibrium. An explanation for the weak NEMCA effect in reactions with limited conversion was proposed by Garagounis et al. [19]. The role of protons on equilibrium limited reactions is both, electrochemical and catalytic. The pumped protons not only modify the catalytic properties of the working electrode, but also carry the electrical power required for ammonia synthesis [19]. That is likely one of the reasons which makes the Λ values reported herein the highest found in literature for ammonia synthesis.

The EPOC effect on catalytic decomposition of ΝΗ3 has been studied on Pd [18], Ru [34], Ag [34,35] using proton (H+) conducting cells and on Fe using both, H+ and K+ conductors [33]. On a Ag electrode, although Λ remained near unity, reaction rate enhancements (ρ) as high as 57 were achieved [35]. Pitselis et al. [33] studied the reaction using Fe films interfaced either with CaZr0.9In0.1O3-α (a H+ conductor) or with K2YZr(PO4)3 (K+ conductor). The rate of NH3 decomposition decreased significantly upon decreasing the catalyst potential, i.e. upon pumping H+ or K+ to the Fe surface. The effect of K+ was more pronounced than that of H+ and could result in a complete poisoning of the reaction [33].

17 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

θK 0.01

1.8

+

0.02

0.03

(a) 1.6

ρ

1.4 1.2 1.0 0.8 70 (b) 60

PH2/PN2= 1 PH2/PN2= 3

PIK+

50 40

PH2/PN2= 6 30 20 10 0 180

(c)

150 120

ΛNH3

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

90 60 30 0 1.5

1.0

0.5

0.0

-0.5

UWR, V

Figure 9. Effect of the applied potential (UWR) and potassium to catalyst ratio (θK+) on (a) the rate enhancement ratio (ρ), (b) the promotion index (PIK+) and (c) the apparent faradaic efficiency (ΛNH3) at T = 450 oC for PH2/PN2 = 1.0, 3.0 and 6.0. Cyclic voltammetry The results of the cyclic voltammetry study were obtained at 450 ºC with different PH2/PN2 ratios and are presented in Figure 10. The cyclic voltammograms display cathodic peaks when the scan goes to lower potentials (UWR = + 2.0 to - 2.0 V) and 18 ACS Paragon Plus Environment

Page 18 of 26

Page 19 of 26

anodic peaks when the scan goes to higher potentials (UWR = -2.0 to + 2.0 V). It has been widely established [36-41] that the cathodic peaks are related to species formed on the surface of the catalyst when the ions of the electroactive support migrate to the catalyst surface and interact with the various absorbed reactants. The anodic peaks are related to the decomposition of these species. Hence, in the present study, if the cathodic and anodic peaks have the same area, the formed compounds have undergone decomposition upon application of positive polarization. In other words, the applied potential of UWR = 2.0 V was high enough to electrochemically decompose or remove the surface promoter compounds, and that is why it was selected as a reference unpromoted state.

100

5 H2 - 95 N2

80 Current, 10-5 A

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

50 H2 - 50 N2

60

75 H2 - 25 N2

40

85 H2 - 15 N2

20 0 -20 -40 -60 -80 -2

-1

0 UWR, V

1

2

Figure 10. Cyclic voltammograms recorded at 450 oC for various PH2/PN2 ratios. The Λ values reported herein are the highest found in literature for ammonia synthesis. This can be largely attributed to the size of the conducting ion, as in reference [33], where, because a K+ cation is 30 times the size of a H+ ion, it was expected to have a much more pronounced EPOC effect. Another reason is that K+ do not take part in the reaction being studied [19]. However, compared to other reactions studied in alkali conducting cells [29,30] these values are far from impressive. One reason for this is the slight change in the definition of Λ, while another is the very nature of the reaction being studied [19]. Another important reason, as evidenced by Figure 10 is the formation of compounds between the potassium promoter species and the adsorbed

19 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 26

reactant species. In fact, the cathodic peaks begin at about 1.0 V, that is, the voltage at which the maximum NH3 formation rates were obtained. In other words, the formed compounds block the active sites on the catalyst surface and gradually negate the promotion effect observed at more positive voltages. The formation of these compounds also generates higher currents than would be observed if they were not formed, leading to reduced Λ values (Eq. 4).

The areas of the peaks in the cyclic voltammograms depend on the PH2/PN2 ratio. The ratio which gives the highest anodic and cathodic area was 1.0 (50% H2 - 50% N2). An increase in the H2 feed concentration decreases the cathodic and anodic area, as does a higher inlet concentration of N2. This means that although the formation of K-N or K-H compounds is possible, the formation of potassium compounds of both reactants, i.e. KN-H, is more likely. However, the identity of these compounds is purely speculative. According to previous studies, three K-N-H compounds are possible candidates: (i) K+ NH3 molecular complexes [42] with a weak interaction between the potassium ion and the ammonia molecule, which could explain the decomposition at UWR = 2 V, (ii) K2NH2 [42], which has a strong ionic character and is formed between two negatively charged lone pair electrons of a nitrogen atom and the positively-charged potassium ions, and (iii) KNH2 compounds [43], which can be formed via the reaction: K + NH3 → KNH2 + ½ H2

(7)

These compounds, whose formation is obviously enhanced by the increased presence of K (or K+), gradually block the active sites of the catalyst, leading to a decline in the ammonia formation rate, as observed in the electrochemical experiments.

Conclusions The following conclusions can be drawn from this study: - The catalytic activity of nitride catalysts, such as Co3Mo3N was electrochemically enhanced in a K+ conducting cell. - The highest values of ammonia synthesis under unpromoted conditions (UWR = 2.0 V) were obtained, at PH2/PN2 = 1. The optimum feed ratio shifts to PH2/PN2 = 3 under promoted conditions (UWR = 1.0 V).

20 ACS Paragon Plus Environment

Page 21 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

- The catalyst exhibited a volcano-type behavior with applied voltage, indicating that a small quantity of potassium ions improves the activity of Co3Mo3N-Ag toward ammonia synthesis. However, when the voltage decreases further and more K+ are pumped to the catalyst (working electrode), the rate of ammonia formation decreases due to the high surface concentration of potassium ions. The poisoning effect is attributed to blocking of active sites by K+ and the formation of K-N-H poisoning compounds at high potassium to catalyst ratios (lower voltages). - By pumping small amounts of potassium ions to the Co3Mo3N-Ag electrode the formation rate could be enhanced by as much as 48% at 450 ºC. Moreover, the in-situ control of the amount of potassium pumping can help in adapting the system to different operation parameters (temperature, feed ratio). - Faradaic efficiencies up to 300 were obtained and they are the highest found in the literature for ammonia synthesis.

References [1] Apodaca, L.E. Nitrogen (Fixed)-Ammonia in Mineral Commodity Summaries, US Geological Survey, Reston VA, January 2016, pp. 118-119. [2] Amar, I.A.; Lan, R.; Petit, C.T.G.; Tao, S. Solid-state electrochemical synthesis of ammonia: A review. J. Solid State Electrochem. 2011, 15, 1845-1860. [3] Zamfirescu, C.; Dincer, I. Using ammonia as a sustainable fuel. J. Power Sources 2008, 185, 459-465, DOI: 10.1016/j.jpowsour.2008.02.097 [4] Liu, H. Ammonia synthesis catalyst 100 years: Practice, enlightenment and challenge. Chinese J. Catal. 2014, 35, 1619-1640. [5] Garagounis, I.; Kyriakou, V.; Skodra, A.; Vasileiou, E.; Stoukides, M. Electrochemical Synthesis of Ammonia in Solid Electrolyte Cells. Frontiers in Energy Research 2014, 2, DOI: 10.3389/fenrg.2014.00001. [6] Brown, D.E.; Edmonds, T.; Joyner, R.W.; McCarroll, J.J.; Tennison, S.R. The genesis and development of the commercial BP doubly promoted catalyst for ammonia synthesis. Catal. Lett. 2014, 144, 545-552. [7] Kyriakou, V.; Garagounis, I.; Vasileiou, E.; Vourros, A.; Stoukides, M. Progress in the Electrochemical Synthesis of Ammonia. Catal. Today 2016, 286, 2-13. [8] Jacobsen, C.J.H.; Dahl, S.; Clausen, B.G.S.; Bahn, S.; Logadottir, A.; Nørskov, J.K. Catalyst design by interpolation in the periodic table: Bimetallic ammonia synthesis catalysts. J. Am. Chem. Soc. 2001, 123, 8404-8405. 21 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

[9] Aika, K.I.; Ozaki, A. Mechanism and isotope effect in ammonia synthesis over molybdenum nitride. J. Catal. 1969, 14, 311-321, DOI: 10.1016/0021-9517(69)903212. [10] Kojima, R.; Aika, K.I. Cobalt molybdenum bimetallic nitride catalysts for ammonia synthesis: Part 1. Preparation and characterization. Appl. Catal. A: Gen. 2001, 215, 149-160, DOI: 10.1016/S0926-860X(01)00529-4. [11] Murata, S.; Aika, K.I. Preparation and characterization of chlorine-free ruthenium catalysts and the promoter effect in ammonia synthesis. 1. An alumina-supported ruthenium catalyst. J. Catal. 1992, 136, 110-117. [12] Yunusov, S.M.; Kalyuzhnaya, E.S.; Moroz, B.L.; Ivanova, A.S.; Reshetenko, T.V.; Avdeeva, L.B.; Likholobov, V.A.; Shur, V.B. New ammonia synthesis catalysts based on supported potassium carbonyl ruthenates as precursors of catalytically active ruthenium particles and potassium promoter. J. Mol. Catal. A - Chem. 2004, 219, 149153. [13] Garagounis, I.; Kyriakou, V.; Anagnostou, C.; Bourganis, V.; Papachristou, I.; Stoukides, M. Solid electrolytes: Applications in heterogeneous catalysis and chemical cogeneration. Ind. Eng. Chem. Res. 2011, 50, 431-472. [14] Vayenas, C.G.; Bebelis, S.; Pliangos, C.; Brosda, S.; Tsiplakides, D. Electrochemical Activation of Catalysis: Promotion, Electrochemical Promotion and Metal-Support Interactions; Kluwer Academic/Plenum, New York, 2001. [15] Vayenas, C.G.; Koutsodontis, C.G. Non-Faradaic Electrochemical Activation of Catalysis. J. Chem. Phys. 2008, 128, 182506-182513. [16] Vayenas, C.G. Bridging Electrochemistry and Heterogeneous Catalysis, J. Solid State Electrochem. 2011, 15, 1425-1435. [17] Yiokari, C.G.; Pitselis, G.E.; Polydoros, D.G.; Katsaounis, A.D.; Vayenas, C.G. High-pressure electrochemical promotion of ammonia synthesis over an industrial iron catalyst. J. Phys. Chem. A 2000, 104, 10600-10602. [18] Marnellos, G.; Zisekas, S.; Stoukides M. Synthesis of ammonia at atmospheric pressure with the use of solid state proton conductors. J. Catal. 2000, 193, 80-87. [19] Garagounis, I.; Kyriakou, V.; Stoukides, M. Electrochemical promotion of catalytic reactions: Thermodynamic analysis and calculation of the limits in Faradaic Efficiency. Solid State Ionics 2013, 231, 58-62.

22 ACS Paragon Plus Environment

Page 22 of 26

Page 23 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

[20] De Lucas-Consuegra, A. New trends of alkali promotion in heterogeneous catalysis: Electrochemical promotion with alkaline ionic conductors. Catal. Surv. Asia 2015, 19, 25-37. [21] Gutiérrez-Guerra, N.; González-Cobos, J.; Serrano-Ruiz, J.C.; Valverde, J.L.; De Lucas-Consuegra, A. Electrochemical Activation of Ni Catalysts with Potassium Ionic Conductors for CO2 Hydrogenation. Top. Catal. 2015, 58, 1256-1269. [22] Hargreaves, J.S.J.; McKay, D. A comparison of the reactivity of lattice nitrogen in Co3Mo3N and Ni2Mo3N catalysts. J. Mol. Catal. A - Chem. 2009, 305, 125-129, DOI: 10.1016/j.molcata.2008.08.006. [23] Guskos, N.; Żołnierkiewicz, G.; Typek, J.; Guskos, A.; Adamski, P.; Moszyński, D. Structure and magnetic properties of chromium doped cobalt molybdenum nitrides. J. Solid State Chem. 2016, 241, 205-211. [24] Patterson, A.L. The scherrer formula for X-ray particle size determination. Phys. Rev. 1939, 56, 978-982. [25] Marnellos, G.; Stoukides, M. Ammonia Synthesis at Atmospheric Pressure. Science 1998, 282, 98-100. [26] Kojima, R.; Aika, K.I. Cobalt molybdenum bimetallic nitride catalysts for ammonia synthesis: Part 2. Kinetic study. Appl. Catal. A - Gen. 2001, 218, 121-128. [27] Kojima, R.; Aika, K.I. Cobalt molybdenum bimetallic nitride catalysts for ammonia synthesis - Part 3. Reactant gas treatment. Appl. Catal. A – Gen. 2001, 219, 157-170. [28] Theleritis, D.; Makri, M.; Souentie, S.; Caravaca, A.; Katsaounis, A.; Vayenas, C.G. Comparative Study of the Electrochemical Promotion of CO2 Hydrogenation over Ru-Supported Catalysts using Electronegative and Electropositive Promoters. ChemElectroChem 2014, 1, 254-262. [29] Vayenas, C.G.; Bebelis, S.; Despotopoulou, M. Non-Faradaic Electrochemical Modification of Catalytic Activity, J. Catal. 1991, 128, 415-435. [30] Yentekakis, I.V.; Moggridge, G.; Vayenas, C.G.; Lambert, R.M. In situ controlled promotion of catalyst surfaces via NEMCA: The effect of Na on the Pt-catalyzed CO oxidation, J. Catal. 1994, 146, 292-305. [31] Vasileiou, E.; Kyriakou, V.; Garagounis, I.; Vourros, A.; Manerbino, A.; Coors, W.G.; Stoukides, M. Electrochemical Enhancement of Ammonia Synthesis in a BaZr0.7Ce0.2Y0.1O2.9 Solid Electrolyte Cell. Solid State Ionics 2016, 288, 357-362.

23 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

[32] Ouzounidou, M.; Skodra, A.; Kokkofitis, C.; Stoukides, M. Catalytic and Electrocatalytic Synthesis of ΝΗ3 in a H+ Conducting Cell by Using an Industrial Fe Catalyst. Solid State Ionics 2007, 178, 153-159. [33] Pitselis, G.E.; Petrolekas, P.D.; Vayenas, C.G. Electrochemical promotion of ammonia decomposition over Fe catalyst films interfaced with K+- & H+- conductors, Ionics 1997, 3, 110-116. [34] Skodra, A.; Ouzounidou, M.; Stoukides, M. ΝΗ3 Decomposition in a Single Chamber Proton-Conducting Cell. Solid State Ionics 2006, 177, 2217-2220. [35] Zisekas, S.; Karagiannakis, G.; Kokkofitis, C.; Stoukides, M. ΝΗ3 Decomposition in a Proton Conducting Solid Electrolyte Cell. J. Appl. Electrochem. 2008, 38, 11431149. [36] Díez-Ramírez, J.; Sánchez, P.; Valverde, J.L.; Dorado, F. Electrochemical promotion and characterization of PdZn alloy catalysts with K and Na ionic conductors for pure gaseous CO2 hydrogenation, J. CO2 Util. 2016, 16, 375-383. [37] Ruiz, E.; Cillero, D.; Morales, Á.; Vicente, G.S.; De Diego, G.; Martínez, P.J.; Sánchez, J.M. Bench scale study of electrochemically promoted CO2 capture on Pt/KβAl2O3. Electrochim. Acta 2013, 112, 967-975. [38] Lucas-Consuegra, A.; González-Cobos, J.; Carcelén, V.; Magén, C.; Endrino, J.L.; Valverde, J.L. Electrochemical promotion of Pt nanoparticles dispersed on a diamondlike carbon matrix: A novel electrocatalytic system for H2 production. J. Catal. 2013, 307, 18-26. [39] De Lucas-Consuegra, A.; Dorado, F.; Valverde, J.L.; Karoum, R.; Vernoux, P. Low-temperature propene combustion over Pt/K-βAl2O3 electrochemical catalyst: Characterization, catalytic activity measurements, and investigation of the NEMCA effect. J. Catal. 2007, 251, 474-48, 10.1016/j.jcat.2007.06.031. [40] Kotsionopoulos, N.; Bebelis, S. In situ electrochemical modification of catalytic activity for propane combustion of Pt/β-Al2O3 catalyst-electrodes. Top. Catalysis 2007, 44, 379-389. [41] Kotsionopoulos, N.; Bebelis, S. Electrochemical characterization of the Pt/β″-Al 2O3 system under conditions of in situ electrochemical modification of catalytic activity for propane combustion. J. Appl. Electrochem. 2010, 40, 1883-1891. [42] Hsiao, Y.W.; Chang, K.M.; Su, T.M. Theoretical study of the neutral and ionic states of hypermetalated potassium compounds K2OH and K2NH2 and potassium complexes KH2O and KNH3. Chem. Phys. 1992, 162, 335-348. 24 ACS Paragon Plus Environment

Page 24 of 26

Page 25 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

[43] Van Ommen, J.G.; Bolink, W.J.; Prasad, J.; Mars, P. The nature of the potassium compound acting as a promoter in iron-alumina catalysts for ammonia synthesis, J. Catal. 1975, 38, 120-127.

For Table of Contents Use Only.

Ammonia production accounts for 1% of global energy. This work investigates ways of enhancing ambient pressure production which would lower the energy demand.

For Table of Contents Use Only.

For Table of Contents Use Only.

For Table of Contents Use Only.

25 ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

For Table of Contents Use Only.

26 ACS Paragon Plus Environment

Page 26 of 26