You are on page 1of 4

Communication

Cite This: J. Am. Chem. Soc. 2019, 141, 19969−19972 pubs.acs.org/JACS

Transition Metal-Free Alkyne Hydrogenation Catalysis with BaGa2, a


Hydrogen Absorbing Layered Zintl Phase
Kelsey L. Hodge and Joshua E. Goldberger*
Department of Chemistry and Biochemistry, The Ohio State University, Columbus, Ohio 43210, United States
*
S Supporting Information

tion with NiGa and NiSn12 led to the observation of catalysis


ABSTRACT: Inexpensive, transition metal-free interme- in Ni3Ga and Ni3Sn2 nanocrystals.13 A final approach is to
tallic compounds have received almost no attention as identify solids with particular electronic structures that are
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via KING ABDULLAH UNIV SCI TECHLGY on April 28, 2022 at 12:23:42 (UTC).

heterogeneous catalysts. Here, we show that BaGa2, a considered to exhibit high catalytic activity.14,15 Examples of
Zintl−Klemm compound composed of honeycomb sheets this include solid-state electrides that have low work functions
of Ga− anions separated by Ba2+ cations and known to and readily dissociate H2.16−18 For instance, LaCu0.67Si1.33 was
react with H2 under moderate conditions to form a reported to be a stable, solid-state electride with activity for
layered polyanionic hydride BaGa2H2, effectively catalyzes nitroarene hydrogenation.18 Still, to the best of our knowledge,
the hydrogenation of phenylacetylene into styrene and there are no known reports of transition metal-free
ethylbenzene under modest conditions (1−50 bar H2, heterogeneous hydrogenation catalysts.
40−100 °C). Remarkably, the catalytic activity of BaGa2 We sought a different strategy to discover unexpected
(surface specific activities up to 8390 h−1) is on the same families of heterogeneous hydrogenation catalysts, on the basis
order of magnitude as commercial Pd-based catalysts. In of the propensity of materials to cleave and absorb H2 into
contrast, BaGa2H2 shows negligible catalytic activity, their lattice under moderate conditions (<200 °C, 50 bar
thereby indicating that the unsaturated Ga− framework H2).19 One of the key steps in most hydrogenation
is necessary for phenylacetylene and styrene adsorption. mechanisms is the dissociative adsorption of H2 onto a
These findings open up future explorations of utilizing and surface. We hypothesize that solids that readily form hydride
optimizing the long-term stability of transition metal-free phases likely have low barriers for this dissociative surface
intermetallic hydrogen absorbing compounds for hydro- adsorption. Some of the aforementioned catalysts readily form
gen-based catalysis. bulk hydrides. Whether the internally absorbed hydrogen plays
a role in catalysis depends on the catalyst and the reaction. Pd,
one of the best hydrogen absorbing metals, incorporates 0.7

T he discovery of new heterogeneous catalysts has come a


long way from the random screening of solid-state
compounds. The most common platforms for heterogeneous
eqiv of H at 1 atm.20 This bulk hydrogen promotes alkyne
overhydrogenation to the alkane.21−24 In contrast, La-
Cu0.67Si1.33 also forms a hydride (LaCu0.67Si1.33H0.3) with
hydrogenations are supported transition metal nanostructures suppressed catalytic activity.18
such as Lindlar’s Pb-coated Pd catalysts on CaCO3 for alkyne Here, we report that BaGa2, a transition metal-free, layered
semihydrogenation,1 reduced Fe3O4 on calcium alumina Zintl phase that absorbs hydrogen at elevated temperatures
silicates,2 or Ru on graphite for NH3 production.3 In almost and pressures to form BaGa2H2,25 is a highly efficient catalyst
all catalysts, the transition metals on the surface are thought to for alkyne hydrogenation. Using phenylacetylene as a substrate,
serve as the active sites, enabling the binding of different we observed semihydrogenation surface specific activities
reactants and intermediates because of their ability to (SSAs) of >400 h−1 with mild conditions (40 °C, 1 bar H2,
accommodate various redox states and coordination environ- 4 mol % catalyst, dimethylformamide (DMF)) as well as a
ments. The support has a variety of roles including limiting much higher SSA (>8390 h−1) at elevated H2 pressures (51 bar
catalyst deactivation from sintering and influencing reactivity H2). These values represent the lower bounds of the turnover
via spillover.4 The need to limit precious metal usage has frequency, as they assume every surface Ga atom is catalytically
resulted in strategies to improve the catalytic activity and active. BaGa2H2 shows no appreciable catalytic activity,
selectivity per mass of active metal. These include diluting the indicating that hydrogen inside the Zintl phase does not
active metal and tuning the electronic structure via alloying participate in catalysis. Despite the high activity, the BaGa2
with a second element5−7 or achieving higher surface area to surface is prone to oxidation in H2O, suppressing catalytic
mass ratios via nanostructuring.8−11 For instance, unsupported activity. This work shows that outstanding catalytic activity can
PdGa, Pd3Ga7, Pd2Ga, and even Al13Fe4 all exhibit much be achieved in main group Zintl phases, which are a completely
higher selectivity in acetylene semihydrogenation while unexplored class of compounds for catalysis.
maintaining similar activities as Pd on Al2O3.5,6 Additionally, The BaGa2 crystal structure consists of planar honeycomb
computational screening methods that determine the relative networks of Ga− anions separated by Ba2+ cations (Figure 1a).
binding affinity of substrates to a catalyst surface have been
fruitful for discovering new heterogeneous catalysts.12 For Received: September 11, 2019
example, initial predictions of selective alkyne semihydrogena- Published: December 9, 2019

© 2019 American Chemical Society 19969 DOI: 10.1021/jacs.9b09856


J. Am. Chem. Soc. 2019, 141, 19969−19972
Journal of the American Chemical Society Communication

Figure 1. (a) Diagram showing conversion of BaGa2 to BaGa2H2.


Blue, red, and green spheres correspond to Ga, Ba, and H,
respectively. (b) X-ray diffraction of BaGa2H2 (red) and BaGa2
(blue). (c) Secondary electron micrograph of BaGa2 powder.

This structure is isoelectronic to MgB2 or graphite, suggesting


that the Ga networks have appreciable π-bonding. Ga π-
bonding is considerably weaker than in the first row elements.
Consequently, under moderate conditions (50 bar H2, 170
°C), BaGa2 reacts with H2 to form BaGa2H2.25 The
polyanionic gallium hydride layer retains the honeycomb
configuration but puckers, indicative of Ga sp3-hybridization.
We synthesized BaGa2 from the elements in a sealed tube
synthesis. Figure 1b shows the powder X-ray diffraction (XRD)
pattern of the BaGa2 product. Rietveld analysis confirms that
BaGa2 is >97% phase pure and crystallizes into a P6/mmm
space group with a = 4.4320(3) Å and c = 5.0731(2) Å (Figure
S1, Tables S1 and S2). Due to the peritectic nature of BaGa2,26 Figure 2. (a) BaGa2 catalyzed hydrogenation of phenylacetylene
(red) to styrene (blue) and ethylbenzene (green). (b) Time
trace quantities of BaGa4 form under a wide range of synthetic dependent conversion products using 1 bar H2, 40 °C, 2.4 mL of
conditions. Therefore, we synthesized pure BaGa4 to compare DMF, 0.87 mmol of phenylacetylene, and 0.043 mmol of BaGa2. (c)
its activity as a potential catalytic impurity (Figure S2). Also, Time dependent conversion products using 51 bar H2, 40 °C, 2.4 mL
we confirmed that the majority of the BaGa2 phase is of DMF, 1.03 mmol of phenylacetylene, and 0.060 mmol of BaGa2.
transformed into BaGa2H2 upon annealing for 7 days at 170 The lines are fit kinetic traces assuming semihydrogenation, and full
°C under 50 bar H2. Figure 1b also shows the XRD pattern of hydrogenation reactions are both first order in phenylacetylene and
our hydrogenated sample, showing a majority of the BaGa2H2 styrene, respectively.
phase. Rietveld analysis (Figure S3, Tables S3 and S4)
confirms that the majority (84.7 wt %) transforms into a air exposure during sampling. No other products other than
P3̅m1 space group with a = 4.530(8) Å and c = 4.910(9) Å, ethylbenzene or styrene were detected via 1H NMR or HPLC.
with 11.6% unreacted BaGa2, and ∼3.7 wt % is BaGa4. Additionally, BaGa4 and BaGa2H2 controls showed minimal
Secondary electron microscopy (SEM) images of the ground catalytic activity even at 51 bar H2 (<10% conversion, 20 h) for
BaGa2 powders show particles with dimensions ranging from all reactions. BaGa4 consists of layers of 2D tetrahedrally
320 nm to 13 μm (Figure 1c). coordinated Ga42− sheets. The lack of activity using BaGa4
Preliminary hydrogenation experiments were conducted implies that the unsaturated π-bonding honeycomb network is
using phenylacetylene as a substrate to evaluate the catalytic essential for activity. We hypothesize that, when the Ga atoms
semihydrogenation to styrene and full hydrogenation to are coordinatively saturated upon forming BaGa2H2, catalysis is
ethylbenzene (Figure 2a). Using 1 bar H2 and only 0.5−5 disrupted via blocking substrate adsorption upon elimination
mol % BaGa2, we found that the dried, freshly distilled, and of the π-network. This loss of activity in the hydrogenated
degassed solvents DMF (40 °C) and n-butanol (70 °C) had phase is consistent with the loss of catalytic activity in
the highest conversion of phenylacetylene to styrene and LaCu0.67Si1.33H0.3.18 Finally, once the BaGa2 catalyst is
ethylbenzene (15−40%, 24 h). Other solvents including THF removed by filtering in a glovebox, the activity is eliminated,
and MeOH had much less conversion of phenylacetylene (5− indicating the absence of a trace homogeneous catalytic
10%, 24 h). As shown in Figure 2b, 80% conversion of impurity being responsible for catalysis. Further confirming
phenylacetylene to styrene and ethylbenzene was observed this, inductively coupled plasma optical emission spectroscopy
after 48 h in DMF, 1 bar H2, and 4.7 mol % BaGa2. Longer of the n-butanol supernatant after a 24 h reaction shows <1
reaction times did not lead to 100% conversion, which we ppm of dissolved Ga. The XRD pattern of the removed BaGa2
attribute to catalyst deactivation via partial oxidation caused by shows minimal changes (Figure S6).
19970 DOI: 10.1021/jacs.9b09856
J. Am. Chem. Soc. 2019, 141, 19969−19972
Journal of the American Chemical Society Communication

The SSA was calculated by measuring the surface area per two different BaGa2 catalysts: one that has been exposed to air
gram of the BaGa2 catalyst via Kr Braunauer-Emmett-Teller and one that has only been exposed to an Ar-filled glovebox.
(BET) measurements, along with the conversion vs time After the BaGa2 catalyst is exposed to air, the 2p3/2 and 2p1/2
(Supporting Information methods). The average surface area peaks appear at 1117.90 and 1144.55 eV, indicating an
of BaGa2 powder was determined to be 0.272 m2g−1 (Figures oxidized Ga3+.28 BaGa2 that has been solely handled in a
S4 and S5). The SSA at 1 bar H2 is estimated to be 425 h−1. glovebox has a surface that is still partially oxidized. However,
This makes BaGa2 the most active Pt/Pd-free heterogeneous the presence of Ga 2p3/2 and 2p1/2 shoulders at 1115.70 and
catalyst for phenylacetylene hydrogenation and is within 1−2 1142.69 eV are indicative of a reduced anionic Ga surface
orders of magnitude of Pt/Pd catalysts under similar species, as would be expected in BaGa2. Further confirming
conditions (Table S5). that surface oxidation inhibits catalysis, the exposure of BaGa2
Additional hydrogenation experiments were performed to either undistilled solvents or air/H 2O immediately
under elevated pressure using the solvents with the best eliminates all catalytic activity. The XPS of a catalyst that has
activity, DMF and n-butanol. In DMF, phenylacetylene is lost activity over long reaction times shows only the oxidized
completely converted to styrene and ethylbenzene within 2 h Ga 2p3/2 and 2p1/2 peaks at 1117.90 and 1144.55 eV (Figure
at 40 °C using 51 bar H2 and 5.5 mol % catalyst (Figure 2c). S8). Annealing air-exposed BaGa2 at 600−800 °C restores the
Within 15 h, the styrene is completely hydrogenated to activity, and the 1115.70 eV Ga 2p3/2 shoulder re-emerges in
ethylbenzene. The maximum SSA for this catalyst is ∼8390 the XPS (Figure S8). Thus, the nonoxidized surface species is
h−1. We modeled the kinetic traces at both low and high H2 regenerated at these high temperatures. It is not unexpected
pressure to consecutive reactions that both exhibit first order that this Ga containing compound is prone to deactivation via
reaction kinetics to their respective substrates. At 1 bar H2, the oxidation, as even catalytic activity in PdGa, Pd2Ga, and
first order rate constant of phenylacetylene to styrene fits to k1 Pd3Ga7 is significantly reduced upon oxidation.29−31 Future
= 0.035(4) h−1, and the first order rate constant of styrene to explorations will determine whether the surface oxide can be
ethylbenzene fits to k2 = 0.035(7) h−1. The slower onset of removed in nonaqueous solution conditions by exploiting the
conversion from phenylacetylene to styrene is indicative of an amphoteric nature of Ga.
induction period in this BaGa2 sample. An induction period is Taken together, we have shown for the first time that BaGa2,
not always observed. Figure S7 shows a kinetic trace using a a transition metal-free intermetallic Zintl phase, has an
second BaGa2 catalyst under the same conditions as Figure 2b, extremely high catalytic activity for alkyne hydrogenation. In
exhibiting no induction period. Future studies are needed to the conversion of phenylacetylene to styrene, the SSAs are
elucidate the origin of the induction period when present. At among the highest for Pd/Pt-free catalysts. The exploration of
51 bar H2, the rate constants are still first order and much these materials for catalysis is in its infancy, and future efforts
faster, with k1 = 1.16(19) h−1 and k2 = 0.337(23) h−1. These are needed to achieve long-term oxidative stability, substrate
differences with respect to H2 pressure indicate the phenyl- selectivity, and a mechanistic understanding. BaGa2 is one
acetylene to styrene conversion reaction approaches first order member of a large family of hydrogen-absorbing layered Zintl
in H2, and the styrene to ethylbenzene conversion approaches phases, all likely to have complementary catalytic properties.
1/2 order in H2. It has been well-established, in the conversion
of ethylene to ethane, that the rate law with respect to H2
varies from 1/2 to first order depending on the catalyst and

*
ASSOCIATED CONTENT
S Supporting Information
conditions.27 In the conversion of acetylene to ethylene, most The Supporting Information is available free of charge at
metals show first order kinetics with respect to H2, comparable https://pubs.acs.org/doi/10.1021/jacs.9b09856.
to our observations with phenylacetylene. Experimental methods, Rietveld analysis of all XRD,
In n-butanol, the conversion rate was slower than DMF, crystallographic data, atomic coordinates and displace-
giving 95% phenylacetylene conversion at 100 °C after 20 h ment parameters, powder XRD patterns, Kr adsorption
(Figure 3a). Figure 3a also demonstrates the reusability of the isotherm, BET measurements and analysis, time depend-
catalyst. The same catalyst retains greater than 50% conversion ent conversion products, X-ray photoelectron spectros-
after four 20 h cycles. The decreased activity over time is copy, and comparison of the catalytic activities of
attributed to the formation of a surface oxide. Figure 3b shows phenylacetylene to styrene with different catalysts
the Ga 2p region of the X-ray photoelectron spectra (XPS) of (PDF)

■ AUTHOR INFORMATION
Corresponding Author
*goldberger.4@osu.edu
ORCID
Joshua E. Goldberger: 0000-0003-4284-604X
Notes
The authors declare no competing financial interest.

Figure 3. (a) Catalyst recycling experiments (51 bar H2, 100 °C, 20 h,
0.058 mmol of BaGa2, 1.1 mmol of phenylacetylene, 2.4 mL of n-
■ ACKNOWLEDGMENTS
K.L.H. and J.E.G. gratefully acknowledge Caitlin Bien and
butanol solvent). (b) XPS spectra of the Ga 2p3/2 and 2p1/2 peaks of Casey Wade for BET measurements and Daniel Weber for
BaGa2 before (red) and after (blue) exposure to air. (*) corresponds SEM measurements. Primary support for this work (K.L.H.)
to the nonoxidized BaGa2. was provided by the Center for Emergent Materials: an NSF
19971 DOI: 10.1021/jacs.9b09856
J. Am. Chem. Soc. 2019, 141, 19969−19972
Journal of the American Chemical Society Communication

MRSEC under Award DMR-1420451. J.E.G. acknowledges the (18) Ye, T. N.; Lu, Y. F.; Li, J.; Nakao, T.; Yang, H. S.; Tada, T.;
Camille and Henry Dreyfus Foundation for partial support. Kitano, M.; Hosono, H. Copper-Based Intermetallic Electride


Catalyst for Chemoselective Hydrogenation Reactions. J. Am. Chem.
Soc. 2017, 139, 17089−17097.
REFERENCES (19) Haussermann, U.; Kranak, V. F.; Puhakainen, K. Hydrogenous
(1) Lindlar, H.; Dubuis, R. Palladium catalyst for partial reduction of Zintl Phases: Interstitial Versus Polyanionic Hydrides. In Zintl Phases:
acetylenes. Org. Synth. 1966, 46, 89. Principles and Recent Developments; Fassler, T. F., Ed.; 2011; Vol. 139,
(2) Kandemir, T.; Schuster, M. E.; Senyshyn, A.; Behrens, M.; pp 143−161.
Schlogl, R. The Haber-Bosch Process Revisited: On the Real (20) Worsham, J. E.; Wilkinson, M. K.; Shull, C. G. Neutron-
Structure and Stability of ″Ammonia Iron″ under Working Diffraction Observations on the Palladium-Hydrogen and Palladium-
Conditions. Angew. Chem., Int. Ed. 2013, 52, 12723−12726. Deuterium Systems. J. Phys. Chem. Solids 1957, 3, 303−310.
(3) Song, Z.; Cai, T. H.; Hanson, J. C.; Rodriguez, J. A.; Hrbek, J. (21) Garcia-Mota, M.; Bridier, B.; Perez-Ramirez, J.; Lopez, N.
Structure and reactivity of Ru nanoparticles supported on modified Interplay between carbon monoxide, hydrides, and carbides in
graphite surfaces: A study of the model catalysts for ammonia selective alkyne hydrogenation on palladium. J. Catal. 2010, 273,
synthesis. J. Am. Chem. Soc. 2004, 126, 8576−8584. 92−102.
(4) Zaera, F. The Surface Chemistry of Metal-Based Hydrogenation (22) Pradier, C. M.; Mazina, M.; Berthier, Y.; Oudar, J.
Catalysis. ACS Catal. 2017, 7, 4947−4967. Hydrogenation of Acetylene on Palladium. J. Mol. Catal. 1994, 89,
(5) Armbruster, M.; Kovnir, K.; Friedrich, M.; Teschner, D.; 211−220.
Wowsnick, G.; Hahne, M.; Gille, P.; Szentmiklosi, L.; Feuerbacher, (23) Teschner, D.; Borsodi, J.; Wootsch, A.; Revay, Z.; Havecker,
M.; Heggen, M.; Girgsdies, F.; Rosenthal, D.; Schlogl, R.; Grin, Y. M.; Knop-Gericke, A.; Jackson, S. D.; Schlogl, R. The roles of
Al13Fe4 as a low-cost alternative for palladium in heterogeneous subsurface carbon and hydrogen in palladium-catalyzed alkyne
hydrogenation. Nat. Mater. 2012, 11, 690−693. hydrogenation. Science 2008, 320, 86−89.
(6) Armbruster, M.; Kovnir, K.; Behrens, M.; Teschner, D.; Grin, Y.; (24) Teschner, D.; Vass, E.; Havecker, M.; Zafeiratos, S.; Schnorch,
Schlogl, R. Pd-Ga Intermetallic Compounds as Highly Selective P.; Sauer, H.; Knop-Gericke, A.; Schloegl, R.; Chamam, M.; Wootsch,
Semihydrogenation Catalysts. J. Am. Chem. Soc. 2010, 132, 14745− A.; Canning, A. S.; Gamman, J. J.; Jackson, S. D.; McGregor, J.;
14747. Gladden, L. F. Alkyne hydrogenation over Pd catalysts: A new
(7) Furukawa, S.; Komatsu, T. Intermetallic Compounds: Promising paradigm. J. Catal. 2006, 242, 26−37.
Inorganic Materials for Well-Structured and Electronically Modified (25) Bjorling, T.; Noreus, D.; Haussermann, U. Polyanionic
Reaction Environments for Efficient Catalysis. ACS Catal. 2017, 7, hydrides from polar intermetallics AeE(2) (Ae = Ca, Sr, Ba; E =
735−765. Al, Ga, In). J. Am. Chem. Soc. 2006, 128, 817−824.
(8) Shen, H.; Li, Y. J.; Shi, Z. Q. A Novel Graphdiyne-Based Catalyst (26) Okamoto, H. Supplemental Literature Review of Binary Phase
for Effective Hydrogenation Reaction. ACS Appl. Mater. Interfaces Diagrams: Al-P, B-Ga, B-Nd, Ba-Ga, Bi-Cs, Ca-Ga, Cd-Gd, Cr-Mo,
2019, 11, 2563−2570. Gd-Ni, Ni-Pb, Ni-Sc, and Sc-Sn. J. Phase Equilib. Diffus. 2015, 36,
(9) Caban-Acevedo, M.; Stone, M. L.; Schmidt, J. R.; Thomas, J. G.; 518−530.
Ding, Q.; Chang, H. C.; Tsai, M. L.; He, J. H.; Jin, S. Efficient (27) Bond, G. Metal-Catalyzed Reactions of Hydrocarbons; Springer
hydrogen evolution catalysis using ternary pyrite-type cobalt Science: New York, NY, 2005.
phosphosulphide. Nat. Mater. 2015, 14, 1245−1251. (28) Priyantha, W.; Radhakrishnan, G.; Droopad, R.; Passlack, M.
(10) Ledendecker, M.; Calderon, S. K.; Papp, C.; Steinruck, H. P.; In-situ XPS and RHEED study of gallium oxide on GaAs deposition
Antonietti, M.; Shalom, M. The Synthesis of Nanostructured Ni5P4 by molecular beam epitaxy. J. Cryst. Growth 2011, 323, 103−106.
Films and their Use as a Non-Noble Bifunctional Electrocatalyst for (29) Kovnir, K.; Osswald, J.; Armbruster, M.; Teschner, D.;
Full Water Splitting. Angew. Chem., Int. Ed. 2015, 54, 12361−12365. Weinberg, G.; Wild, U.; Knop-Gericke, A.; Ressler, T.; Grin, Y.;
(11) Callejas, J. F.; Read, C. G.; Popczun, E. J.; McEnaney, J. M.; Schlogl, R. Etching of the intermetallic compounds PdGa and Pd3Ga7:
Schaak, R. E. Nanostructured Co2P Electrocatalyst for the Hydrogen An effective way to increase catalytic activity? J. Catal. 2009, 264, 93−
Evolution Reaction and Direct Comparison with Morphologically 103.
Equivalent CoP. Chem. Mater. 2015, 27, 3769−3774. (30) Wowsnick, G.; Teschner, D.; Kasatkin, I.; Girgsdies, F.;
(12) Studt, F.; Abild-Pedersen, F.; Bligaard, T.; Sorensen, R. Z.; Armbruster, M.; Zhang, A. P.; Grin, Y.; Schlogl, R.; Behrens, M.
Christensen, C. H.; Norskov, J. K. Identification of non-precious Surface dynamics of the intermetallic catalyst Pd2Ga, Part I -
metal alloy catalysts for selective hydrogenation of acetylene. Science Structural stability in UHV and different gas atmospheres. J. Catal.
2014, 309, 209−220.
2008, 320, 1320−1322.
(31) Wowsnick, G.; Teschner, D.; Armbruster, M.; Kasatkin, I.;
(13) Liu, Y. X.; Liu, X. W.; Feng, Q. C.; He, D. S.; Zhang, L. B.;
Girgsdies, F.; Grin, Y.; Schlogl, R.; Behrens, M. Surface dynamics of
Lian, C.; Shen, R. A.; Zhao, G. F.; Ji, Y. J.; Wang, D. S.; Zhou, G.; Li,
the intermetallic catalyst Pd2Ga, Part II - Reactivity and stability in
Y. D. Intermetallic NixMy (M = Ga and Sn) Nanocrystals: A Non-
liquid-phase hydrogenation of phenylacetylene. J. Catal. 2014, 309,
precious Metal Catalyst for Semi-Hydrogenation of Alkynes. Adv.
221−230.
Mater. 2016, 28, 4747−4754.
(14) Park, H.; Encinas, A.; Scheifers, J. P.; Zhang, Y. M.; Fokwa, B.
P. T. Boron-Dependency of Molybdenum Boride Electrocatalysts for
the Hydrogen Evolution Reaction. Angew. Chem., Int. Ed. 2017, 56,
5575−5578.
(15) Grant, J. T.; Carrero, C. A.; Goeltl, F.; Venegas, J.; Mueller, P.;
Burt, S. P.; Specht, S. E.; McDermott, W. P.; Chieregato, A.;
Hermans, I. Selective oxidative dehydrogenation of propane to
propene using boron nitride catalysts. Science 2016, 354, 1570−1573.
(16) Gong, Y. T.; Wu, J. Z.; Kitano, M.; Wang, J. J.; Ye, T. N.; Li, J.;
Kobayashi, Y.; Kishida, K.; Abe, H.; Niwa, Y.; Yang, H. S.; Tada, T.;
Hosono, H. Ternary intermetallic LaCoSi as a catalyst for N2
activation. Nat. Catal. 2018, 1, 178−185.
(17) Wu, J. Z.; Li, J.; Gong, Y. T.; Kitano, M.; Inoshita, T.; Hosono,
H. Intermetallic Electride Catalyst as a Platform for Ammonia
Synthesis. Angew. Chem., Int. Ed. 2019, 58, 825−829.

19972 DOI: 10.1021/jacs.9b09856


J. Am. Chem. Soc. 2019, 141, 19969−19972

You might also like