Download as pdf or txt
Download as pdf or txt
You are on page 1of 2

COMMENTARY

COMMENTARY

Homogeneous catalysis for the nitrogen fuel cycle


Yanming Liua,b and Thomas J. Meyera,1

The nitrogen cycle is an important biogeochemical dinitrogen and dihydrogen at ambient pressures
cycle in which nitrogen, a building block of proteins and temperature. The results from Habibzadeh et al.
and nucleic acids, is chemically activated. In the cycle, (3) are an important extension of the remarkable high-
atmospheric dinitrogen is converted into ammonia as oxidation-state chemistry of these complexes with RuII
our primary source of bioavailable nitrogen. Ammonia and OsII. Known examples in this chemistry include cat-
is a carbon-free energy carrier and widely used as alysts for water oxidation, catalytic organic oxidations,
a chemical feedstock. Ammonia oxidation is also a and a remarkable series of inorganic reactions such as
critical step in global nitrogen cycling. Harnessing the the oxidation of NH3 to NO3−, all in polypyridyl coordi-
energy stored in ammonia involves oxidization to nation environments.
dinitrogen and hydrogen, NH3 → 1/2N2 + 3/2H2, with The discovery reported here is an important ex-
the hydrogen produced utilized in hydrogen-based tension. It exploits the reactivity of these complexes to
fuel cells (Fig. 1). Ammonia can be also easily lique- a reaction of potential interest in energy application:
fied. As a hydrogen carrier, it provides a strategy for homogenous electrochemical oxidation of ammonia
the efficient storage, transport, and utilization of re- to N2. Exploitation of the reaction could provide a
newable hydrogen on large scales, and it is not sur- basis for efficient molecular catalysis in the oxidation
prising that ammonia oxidation has attracted great of ammonia to dinitrogen under mild conditions, as
interest in clean energy and environmental pollution shown in Fig. 1.
control. Electrochemical oxidation of ammonia occurs Ammonia oxidation has been extensively studied in
at room temperature under ambient pressures (1, 2) biological systems and in heterogeneous systems based
with an interest in active electrocatalysts with low on thermal conversion, photocatalysis, or electrocatalysis
overpotentials and high faradaic efficiencies for its oxi- (4, 5). Biological ammonia oxidation is driven by bacteria
dization to hydrogen to dinitrogen. and archaea, and they transform ammonia to dinitrogen
A study in PNAS by Habibzadeh et al. (3) reports by anaerobic ammonium oxidation, or oxidize ammonia
that ruthenium polypyridyl complexes can be used to nitrite or nitrate by nitrification (5, 6) or, as in Fig. 1, to
as electrocatalysts for the oxidation of ammonia to dinitrogen as a product.
Electrolysis of ammonia to dinitrogen and dihy-
drogen should be thermodynamically accessible un-
der mild conditions with renewable energy sources.
An ideal electrocatalyst for ammonia oxidation should
have appropriate binding energies for ammonia and
for the key oxidation intermediates that interconvert
the two (7). There are obvious problems, including
stable coordination of ammonia in the first place and
accessing the higher oxidation states at the metal
while avoiding competing decomposition pathways.
Metals such as Pt, Ir, Ru, and Au and their alloys
Fig. 1. Simplified reaction scheme illustrating an ammonia/N2 fuel cycle in which are active heterogeneous catalysts for electrochem-
NH3 is electrochemically oxidized to N2 and H2, NH3 →1/2N2 + 3/2H2, with H2 ical oxidation of ammonia to dinitrogen (1, 2, 8).
utilized in a hydrogen/O2 fuel cell. In a reverse cycle, N2 is electrochemically
reduced to NH3 by nitrogen fixation (1/2N2 +3/2H2O → NH3 + 3/4O2). In a
Among these electrocatalysts, Pt and Ir alloys have
complete ammonia fuel cycle, both electrochemical ammonia oxidation and been considered as the most effective for ammonia
nitrogen fixation would be driven by renewable energy sources. oxidation, with their low overpotentials and high
Downloaded at EXPERIMENTALES/CAMPUS UNIV on November 20, 2019

a
Department of Chemistry, The University of North Carolina at Chapel Hill, Chapel Hill, NC 27599; and bSchool of Environmental Science and
Technology, Dalian University of Technology, Dalian 116024, China
Author contributions: Y.L. and T.J.M. wrote the paper.
The authors declare no conflict of interest.
Published under the PNAS license.
See companion article on page 2849.
1
To whom correspondence should be addressed. Email: tjmeyer@unc.edu.
Published online February 8, 2019.

2794–2795 | PNAS | February 19, 2019 | vol. 116 | no. 8 www.pnas.org/cgi/doi/10.1073/pnas.1822090116


current densities (2). The results of experiments on catalyst structure the reaction of ammonia splitting to dinitrogen and dihydrogen (1:3).
and composition have been used to lower overpotential and en- The results of 15NH3 labeling experiments also show that dinitrogen is
hance the kinetics and efficiency for ammonia electrooxidation. In generated from ammonia oxidation. There are molybdenum and iron
heterogeneous catalysis, a major concern for metal heterogeneous complex catalysts that are active for fixing dinitrogen to ammonia in
catalysis is electrode-surface poisoning by adsorbed intermediates. nitrogen fuel cycles (13, 14), but this is a molecular catalyst for the
One answer is homogeneous electrocatalysis at inert, conducting electrochemical oxidation ammonia to dinitrogen at low overpotential.
electrodes to overcome activity loss of the catalyst.
As noted above, given their background redox chemistry, A study in PNAS by Habibzadeh et al. reports
polypyridyl complexes like [Ru(trpy)(bpy)(NH3)]2+ (trpy = 2,2′:6′,2′′- that ruthenium polypyridyl complexes can be
terpyridine, bpy = 2,2′-bipyridine) are of particular interest in this
chemistry. Mechanistic studies have shown that they undergo used as electrocatalysts for the oxidation of
single-electron oxidation from RuII to RuIII, followed by oxidation ammonia to dinitrogen and dihydrogen at
and proton loss to RuIV = NH, with the latter undergoing nucleo- ambient pressures and temperature.
philic attack on the imido N to give RuII-hydroxylamines (9). There
are related results on the oxidation of coordinated ammonia li- Mechanistic details were also investigated by cyclic voltam-
gands to N2 in dinuclear ruthenium complexes, mononuclear os- metry and additional studies in solution. In THF, the rate of
mium complexes, and on a series of molybdenum complexes (10– ammonia oxidation was first order in 1a, suggesting that single-
12). These observations provide a basis for dinitrogen generation electron oxidation of RuII to RuIII occurs before the rate-limiting
by oxidation of ammonia in complexes like [Ru(trpy)(bpy)(NH3)]2+, step. Based on the results of previous studies on a series of transition
but E1/2 for the RuIII/RuII couple was 0.15 V more positive than the metal complexes, including examples from ruthenium, osmium, and
onset potential for ammonia oxidation at glassy carbon electrodes. molybdenum, the overall reaction presumably proceeds through
A key to driving the reaction has come from the results of Habib-
hydrazine or nitride pathways (11, 12, 15). In the hydrazine pathway,
zadeh et al. (3) by exploiting the background synthetic chemistry
deprotonation is followed by N–N bond formation to give the in-
to prepare the modified complex [Ru(trpy)(dmabpy)(NH3)]2+ [trpy =
termediate hydrazine complex, RuII(H2N-NH2)2+, with the latter gen-
2,2′:6′,2′′-terpyridine, dmabpy = 4,4´-bis(dimethylamino)-2,2′-
erated by nucleophilic attack of the amine on a transient imido
bipyridine] (catalyst 1a in ref. 3). They demonstrate 1a to be an
intermediate. In the nitride pathway, N–N coupling occurs between
efficient homogenous catalyst for electrochemical oxidation of
a nitride, RuV(N), and NH3.
ammonia to dinitrogen at room temperature and ambient pres-
Experimental evidence was found for the hydrazine inter-
sure. The introduction of the dmabpy ligand reduces the RuIII/RuII
mediate [(trpy)(dmabpy)RuIII(N2H4)]2+ (5a in ref. 3) by heating
oxidation potential from 1.03 to 0.68 V vs. normal hydrogen elec-
trode (NHE), which is ∼0.19 V below the potential for direct oxi- [(trpy)(dmabpy)Ru(Cl)][Cl] in a hydrazine hydrate solution satu-
dation at a glassy carbon electrode. rated with [NH4][PF6]. The solid was characterized by 1 H NMR
The reactivity of catalyst 1a toward ammonia oxidation was and cyclic voltammetry. 1H NMR spectra of reaction from 1a, 5a,
first shown by the appearance of a significantly enhanced current or [(trpy)(dmabpy)RuIII(N2H4)][PF6]2, with and without added NH3,
density with added ammonia after oxidation of Ru II to Ru III in are all consistent with N–N bond formation through the hydrazine
[Ru(trpy)(dmabpy)(NH3)]3+. Compared with uncatalyzed ammonia complex as an intermediate.
oxidation on glassy carbon electrodes, catalyst 1a lowers the over- The mechanistic details revealed by Habibzadeh et al. (3) will be
potential for ammonia oxidation by ∼0.3 V in THF solution. Of useful for defining conditions for maximizing the oxidation of ammo-
significant interest here is that dinitrogen and dihydrogen are pro- nia to dinitrogen in catalytic systems in the future. The key features
duced from ammonia electrolysis in the solvent with high faradaic emerging here show that RuII oxidation to RuIII is followed by depro-
efficiencies of 86% and 78% at 0.73 V vs. NHE, showing that the tonation and N–N bond formation through a hydrazine intermediate.
catalyst converts ammonia to dinitrogen and dihydrogen. The ratio The challenge now will be to identify conditions that enhance re-
of dinitrogen to dihydrogen is 1:2.74, close to the stoichiometry for activity further at larger scales for the design of ammonia fuel cycles.

1 Little DJ, Smith MR, III, Hamann TW (2015) Electrolysis of liquid ammonia for hydrogen generation. Energy Environ Sci 8:2775–2781.
2 Sacré N, et al. (2018) Tuning Pt–Ir interactions for NH3 electrocatalysis. ACS Catal 8:2508–2518.
3 Habibzadeh F, Miller SL, Hamann TW, Smith MR, 3rd (2019) Homogeneous electrocatalytic oxidation of ammonia to N2 under mild conditions. Proc Natl Acad Sci
USA 116:2849–2853.
4 Chen JG, et al. (2018) Beyond fossil fuel-driven nitrogen transformations. Science 360:eaar6611.
5 Pratscher J, Dumont MG, Conrad R (2011) Ammonia oxidation coupled to CO2 fixation by archaea and bacteria in an agricultural soil. Proc Natl Acad Sci USA
108:4170–4175.
6 Canfield DE, Glazer AN, Falkowski PG (2010) The evolution and future of Earth’s nitrogen cycle. Science 330:192–196.
7 Katsounaros I, et al. (2018) On the mechanism of the electrochemical conversion of ammonia to dinitrogen on Pt(100) in alkaline environment. J Catal 359:82–91.
8 Xu W, et al. (2018) Electrodeposited NiCu bimetal on carbon paper as stable non-noble anode for efficient electrooxidation of ammonia. Appl Catal B
237:1101–1109.
9 Thompson MS, Meyer TJ (1981) Oxidation of coordinated ammonia to nitrate. J Am Chem Soc 103:5577–5579.
Downloaded at EXPERIMENTALES/CAMPUS UNIV on November 20, 2019

10 Ishitani O, Ando E, Meyer TJ (2003) Dinitrogen formation by oxidative intramolecular N—N coupling in cis,cis-[(bpy)2(NH3)RuORu(NH3)(bpy)2]4+. Inorg Chem
42:1707–1710.
11 Coia GM, et al. (1994) Preparation of osmium hydrazido complexes by interception of an osmium(IV) imido intermediate. J Am Chem Soc 116:3649–3650.
12 Bezdek MJ, Guo S, Chirik PJ (2016) Coordination-induced weakening of ammonia, water, and hydrazine X-H bonds in a molybdenum complex. Science 354:730–733.
13 Arashiba K, Miyake Y, Nishibayashi Y (2011) A molybdenum complex bearing PNP-type pincer ligands leads to the catalytic reduction of dinitrogen into ammonia.
Nat Chem 3:120–125.
14 Macleod KC, Holland PL (2013) Recent developments in the homogeneous reduction of dinitrogen by molybdenum and iron. Nat Chem 5:559–565.
15 Bhattacharya P, et al. (2017) Ammonia oxidation by abstraction of three hydrogen atoms from a Mo-NH3 complex. J Am Chem Soc 139:2916–2919.

Liu and Meyer PNAS | February 19, 2019 | vol. 116 | no. 8 | 2795

You might also like