Você está na página 1de 5

ARTICLES

PUBLISHED ONLINE: 2 MARCH 2014 | DOI: 10.1038/NCHEM.1873

Discovery of a Ni-Ga catalyst for carbon dioxide


reduction to methanol
Felix Studt1, Irek Sharafutdinov2, Frank Abild-Pedersen1, Christian F. Elkjær2, Jens S. Hummelshøj1,
Søren Dahl2, Ib Chorkendorff2 and Jens K. Nørskov1,3 *

The use of methanol as a fuel and chemical feedstock could become very important in the development of a more
sustainable society if methanol could be efficiently obtained from the direct reduction of CO2 using solar-generated
hydrogen. If hydrogen production is to be decentralized, small-scale CO2 reduction devices are required that operate at low
pressures. Here, we report the discovery of a Ni-Ga catalyst that reduces CO2 to methanol at ambient pressure. The
catalyst was identified through a descriptor-based analysis of the process and the use of computational methods to
identify Ni-Ga intermetallic compounds as stable candidates with good activity. We synthesized and tested a series of
catalysts and found that Ni5Ga3 is particularly active and selective. Comparison with conventional Cu/ZnO/Al2O3 catalysts
revealed the same or better methanol synthesis activity, as well as considerably lower production of CO. We suggest that
this is a first step towards the development of small-scale low-pressure devices for CO2 reduction to methanol.

N
ature reduces CO2 photochemically to store energy, and show experimentally that it has the unique property that it
devising an artificial process to replicate this remains one of reduces CO2 to methanol without producing large amounts of CO
the grand challenges in modern chemistry1–4. One possibility, via the rWGS reaction.
which is currently the subject of very active research, is a photo-elec-
trochemical process, but finding an electrocatalyst that is selective Results
and has a low overpotential is challenging5–11. An alternative A large literature exists about the methanol synthesis reaction over
approach would be to first generate molecular hydrogen via a supported copper catalysts18–28. Here, we consider the direct CO2
photo-electrochemical process or an electrochemical process using reduction to methanol. Grabow and Mavrikakis have considered
electrical power from photovoltaic cells or wind turbines12,13. If many different reaction paths and suggested the following to be
the hydrogen were then used in a heterogeneously catalysed most likely29,30:
process to reduce CO2 to methanol, a sustainable source of liquid
fuel would be established. H2 (g) + 2* ↔ 2H* (1)
Today, methanol is produced in large facilities from CO, CO2 and
H2 (derived from fossil resources) in a high-pressure (50–100 bar) CO2 (g) + H* ↔ HCOO* (2)
process using a Cu/ZnO/Al2O3 catalyst14. If hydrogen production is
to be distributed and produced in small-scale devices, it would be HCOO* + H* ↔ HCOOH* + * (3)
attractive if the subsequent conversion of H2 into a liquid fuel
could also be performed in simpler, low-pressure decentralized HCOOH* + H* ↔ H2 COOH* + * (4)
units. This is not, however, simply a case of reengineering the technol-
ogy currently optimized for high-pressure conversion of syngas into H2 COOH* + * ↔ H2 CO* + OH* (5)
methanol, because a low-pressure CO2 reduction process may
require a different catalyst. Another challenge arises with the use of H2 CO* + H* ↔ H3 CO* + * (6)
CO-free CO2 , which will lead to CO as a by-product of methanol
via the reverse water–gas shift (rWGS) reaction. The production of H3 CO* + H* ↔ CH3 OH(g) + 2* (7)
CO not only reduces the yield of methanol—it also has a negative
effect when methanol is used in fuel cells because CO poisons the OH* + H* ↔ H2 O(g) + 2* (8)
Pt catalyst used. Using the industrial Cu/ZnO/Al2O3 catalyst
(which is optimized for different reaction conditions including a The symbol * represents a surface site or an adsorbed species.
CO-rich feed) in low-pressure methanol synthesis leads to significant A simple mean-field kinetic model is used to elucidate trends in
CO production, so new catalysts are needed to advance this field. reactivity. The model treats all reaction steps as being potentially
In the present Article, we report the discovery of a new, non- rate-determining and solves the rate of methanol production
precious metal catalyst working at low pressure with similar or under steady-state conditions, similar to those described for other
higher methanol yield than the current Cu/ZnO/Al2O3 methanol reactions31,32. There are a total of eight activation energies for the
synthesis catalyst15–17. We use a computational descriptor-based forward elementary steps. Together with the eight elementary reac-
approach to guide us towards a new class of Ni-Ga catalysts and tion energies, these define the complete energy-space of the

1
SUNCAT Center for Interface Science and Catalysis, SLAC National Accelerator Laboratory, 2575 Sand Hill Road, Menlo Park, California 94025, USA,
2
Centre for Individual Nanoparticle Functionality (CINF), Department of Physics, Building 307 Technical University of Denmark, DK-2800 Lyngby, Denmark,
3
SUNCAT Center for Interface Science and Catalysis, Department of Chemical Engineering, Stanford University, Stanford, California 94305, USA.
* e-mail: norskov@stanford.edu

NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry 1

© 2014 Macmillan Publishers Limited. All rights reserved.


ARTICLES NATURE CHEMISTRY DOI: 10.1038/NCHEM.1873

1 a
Ni5Ga3 NiGa

Cu + Zn
Ni3Ga
Ni5Ga3
0 Ni5Ga3 Cu
log (TOF/TOFCu)

NiGa

–1 NiGa Pd 10 nm 10 nm

Ni b
In situ XRD
after reduction/alloying
Ni3Ga
–2
–2 –1 0 1 2
Cu
ΔEO – ΔEO (eV) α-Ni3Ga

Intensity (a.u.)
Figure 1 | Theoretical activity volcano for CO2 hydrogenation to methanol.
Turnover frequency (TOF) is plotted as a function of DEO , relative to
Cu(211). DEO for the stepped 211 surfaces of copper, nickel and palladium is δ-Ni5Ga3
depicted as open black circles, and Cu þ Zn is depicted in orange. DEO for
Ni-Ga intermetallic compounds is depicted in red. Closed circles indicate
nickel-rich sites, open circles gallium-rich sites and half-open circles mixed β-NiGa
sites. Reaction conditions are 500 K, 1 bar, and a CO2:H2 ratio of 1:3.

reaction. These energies have been calculated with density func- Cu/ZnO/Al2O3
tional theory (DFT) using the RPBE exchange-correlation energy
functional33 for a selected set of metals (see Supplementary Tables 40 50 60 70 80 90
1 and 2 for a table of all the energies). In each case we chose a 2θ (deg)
stepped face-centred cubic fcc(211) surface to represent the active
site30,34. We have shown that van der Waals (vdW) interactions Figure 2 | Characterization of the catalysts studied. a, TEM images of
can be important for the energetics of CO2 reduction for Cu(211) Ni5Ga3 and NiGa. b, In situ XRD patterns of Ni3Ga, Ni5Ga3 and NiGa
using the BEEF–vdW functional35,36. In the following we include intermetallic compounds as well as Cu/ZnO/Al2O3.
such effects by assuming that the extra effect of van der Waals inter-
actions is the same as on copper for all the other metals considered. respectively. In the figure we have also included zinc doping in a
Given the non-specific nature of the dispersion interactions, and the copper step to model the active site of the ZnO promoted commer-
fact that the catalytically interesting metals are close in bonding to cial catalyst. This has been shown theoretically and experimentally
copper, this is a very reasonable approximation. to be a good description30, and the model captures the near-
We will now describe the approach we have taken to reduce the optimal activity of such a site. Our one-descriptor model therefore
number of energy parameters in the methanol synthesis from 16 to provides a good starting point for discovering other potential cata-
only 1. In doing so we lose some accuracy, but it is important lysts. We note that even though the zinc promoted copper steps have
to build such a model for at least two reasons. First, it allows us to close to optimal activity, the density of such sites is small in a doped
understand the trends in catalytic activity among the metals. system30, and a more homogeneous catalyst with the same activity
Second, it is a very efficient way of identifying new catalyst per site but more active sites would be advantageous.
leads37,38. We find that scaling relations exist between the oxygen In Fig. 1 we have included predictions of the simple models for
adsorption energy, DEO , and the adsorption energies and tran- the mixed-metal system Ni-Ga. We chose Ni-Ga because this com-
sition-state energies of all the hydrogenated forms of CO2 when prises a series of stable intermetallic compounds with large ordering
we compare different metal surfaces (see Supplementary Figs. 1 energies (for example, Ni-Ga is calculated to have a heat of for-
and 2 for the complete set of scaling relations). The result is a com- mation of 20.64 eV/formula unit (two atoms)). This increases the
plete mapping of all the relevant energies in the methanol kinetics chance that the surfaces exhibit a truncated bulk structure,
onto only one parameter, DEO. To a first approximation this par- making modelling simpler.
ameter characterizes the catalytic properties uniquely. Several of the Ni-Ga intermetallic compounds show active
Solving the steady-state microkinetic model with the input of sites with oxygen adsorption energies close to the optimum. We
these scaling relations yields the calculated rate of CO2 hydrogen- synthesized a series of Ni-Ga catalysts with different Ni:Ga ratios sup-
ation as a function of DEO at ambient pressure and 500 K, as ported on silica using incipient wetness co-impregnation followed by
shown in Fig. 1. Values of DEO for the elemental metals copper, pal- high-temperature reduction in H2. The Ni-Ga catalysts were charac-
ladium and nickel are included in this volcano plot. The optimum in terized using X-ray diffraction (XRD) together with transmission
reaction rate is a result of competition between having a too weak electron microscopy (TEM) (Fig. 2). As can be seen from the XRD
interaction with oxygen (resulting in too unstable intermediates diffraction patterns, all three different Ni-Ga intermetallic com-
and high reaction barriers) and a too strong coupling to oxygen pounds, Ni3Ga, NiGa and Ni5Ga3 , could be prepared rather phase-
(giving rise to surface poisoning by formate, and possibly other pure, which can be attributed to the high formation energy of the
species bound through oxygen). different phases and the very sharp lines in the Ni-Ga phase
At atmospheric pressure, elemental copper is closest to the top, diagram39. The TEM images presented in Fig. 2a reveal an average
while nickel and palladium bind oxygen too strongly and weakly, particle size of 5.1 nm for the Ni5Ga3 and 6.2 nm for the NiGa.

2 NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry

© 2014 Macmillan Publishers Limited. All rights reserved.


NATURE CHEMISTRY DOI: 10.1038/NCHEM.1873 ARTICLES
a important consideration for the present group of catalysts. A CO2
MeOH
0.25 reduction process that mainly produces methanol and water is
mol [active metal]*h

highly desirable, because the CO will need to be recycled or


mol[MeOH]/

0.20
CuZnO burned. The conventional Cu-Zn catalyst has a high rate of
0.15 Ni5Ga3 rWGS. The data in Fig. 3 show that this is not the case for the
0.10 Ni-Ga catalysts.
NiGa
0.05
Ni3Ga Discussion
The experimental data raise two interesting questions: (1) what is
160 180 200 220 240 260
the relationship between the activity volcano in Fig. 1 and the
Temperature (°C)
ranking of the activity data in Fig. 3; (2) why is the rWGS activity
b of the Ni-Ga catalysts lower than for the copper-based catalyst.
100
CuZnO NiGa
Part of the answer to the first question is related to the obser-
Selectivity Ni5Ga3
vation made earlier that there are two factors affecting the rate,
MeOH + DME

80
selectivity (%)

the activity and number of active sites. The copper particles will
60
be most active at the relatively few places where they are promoted
40 Ni3Ga by zinc, whereas the active sites on the intermetallic compounds do
20
not need the presence of a promoter. The difference in activity
between Cu-Zn and the intermetallic compounds is therefore prob-
0 ably masked by differences in the number of active sites. When it
160 180 200 220 240 260
comes to the ranking of the different Ni-Ga catalysts, the main dis-
Temperature (°C)
crepancy between the volcano in Fig. 1 and the experimental data in
c Fig. 3 is the Ni3Ga mixed sites catalyst, which is predicted to be very
CO/MeOH active, but found not to be so. The reason for this, we suggest, is that
100
CO/MeOH ratio

the nickel sites become poisoned by adsorbed CO and eventually


CuZnO
(through dissociation) by carbon (see the following and the
Supplementary Fig. 3) and subsequent phase separation. In fact, if
10
Ni5Ga3
we distinguish between nickel-rich sites and gallium-rich sites (as
done in Fig. 1 by using different symbols), it can be seen that the
gallium-rich sites follow the ordering observed experimentally
1 very well.
160 180 200 220 240 260 The difference in rWGS and methanol synthesis activity can be
Temperature (°C) understood with the same picture. The gallium-rich sites facilitate
methanol synthesis and the nickel-rich sites do rWGS (and metha-
Figure 3 | The measured activity and selectivity towards methanol nation) until they become self-poisoned by CO and carbon. It is
synthesis as a function of temperature for the studied catalysts. a, Yield of different for Cu-Zn, where both rWGS and methanol synthesis
methanol of a series of NixGay catalysts compared with Cu/ZnO/Al2O3 as a proceed at the same surface site. Because CO does not bind
function of temperature at atmospheric pressure. Gas composition: 75% H2 strongly enough to copper, no poisoning effect will be observed,
and 25% CO2. Gas hourly space velocity ¼ 6,000 h21. b, CO-free selectivity which translates into a higher rWGS activity and hence lower
towards methanol and dimethyl ether in per cent. c, Comparison of the CO methanol selectivity. A more detailed analysis supporting this
to MeOH ratio of Cu/ZnO/Al2O3 with Ni5Ga3. argument can be found in the Supplementary Section ‘Reverse
water-gas-shift volcano’. We note that the high activity of the
We tested the Ni-Ga catalysts with CO2 hydrogenation at a
pressure of 1 bar in a tubular fixed-bed reactor. For comparison, a 0.20
conventional Cu/ZnO/Al2O3 catalyst was synthesized following After regeneration
the procedure described in ref. 40. This procedure has been in H2 at 350 °C
shown to produce catalysts that are at least as good as the commer- 0.15
Product (g[product]/g[cat]*h)

cial catalysts for short-term testing17. The XRD pattern of the CO


Cu/ZnO/Al2O3 catalyst is also shown in Fig. 2b. The measured *
Brunauer-Emmet-Teller surface area of this catalyst (92 m2 g21) is 0.10
*
comparable to that reported in ref. 40. MeOH
Figure 3 shows the measured activity and selectivity towards Ni5Ga3
methanol synthesis as a function of temperature for the different 200 °C
SiO2 supported Ni-Ga catalysts. The amount of active metal (Ni þ 1 atm
Ga or Cu), in moles, was the same for all catalysts under investi- CO2:H2 1:3
gation. The corresponding values of the active surface area, esti- DME
5 × 10–4
mated from XRD and TEM analysis, can be found in
CH4
Supplementary Table 3. Ni5Ga3/SiO2 stands out as being particu- 0
larly active towards methanol synthesis. In fact, at temperatures 0 20 40 60 80
above 220 8C, the yield of methanol is considerably higher than Time on stream (h)
with Cu/ZnO/Al2O3. The selectivity, including all products except
CO, is very high for Cu-Zn, Ni5Ga3 and NiGa. Only Ni3Ga produces Figure 4 | Deactivation of Ni5Ga3 with time on stream. The reaction was
significant amounts of methane. Notably, the CO-to-methanol ratio carried out at 200 8C, atmospheric pressure and with a CO2 to H2 ratio of
of Ni5Ga3 is significantly lower than that of Cu/ZnO/Al2O3 (Fig. 3c). 1:3. Regeneration of the catalyst in H2 at 350 8C is shown. Asterisks mark
The selectivity towards CO compared to methanol, that is, temperature crashes for several hours that have not been accounted for in
the rate of the rWGS versus the rate of methanol synthesis is an the total time on stream.

NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry 3

© 2014 Macmillan Publishers Limited. All rights reserved.


ARTICLES NATURE CHEMISTRY DOI: 10.1038/NCHEM.1873

Ni-Ga catalysts at high temperatures is related to the low rWGS References


activity. The amount of water in the gas is smaller, shifting the equi- 1. Schlögl, R. Chemistry’s role in regenerative energy. Angew. Chem. Int. Ed.
librium concentration of methanol up, hence substantially reducing 50, 6424–6426 (2011).
2. Olah, G. A. Towards oil independence through renewable methanol chemistry.
the backwards reaction. Angew. Chem. Int. Ed. 52, 104–107 (2013).
We performed stability tests for the best catalyst, Ni5Ga3/SiO2 , 3. Benson, E. E., Kubiak, C. P., Sathrum, A. J. & Smieja, J. M. Electrocatalytic and
under reaction conditions. Figure 4 shows the production of metha- homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev.
nol, CO, dimethyl ether and methane of Ni5Ga3/SiO2 as a function 38, 89–99 (2009).
of time on stream, at 200 8C and atmospheric pressure. After initial 4. Arakawa, H. et al. Catalysis research of relevance to carbon management:
progress, challenges, and opportunities. Chem. Rev. 101, 953–996 (2001).
deactivation of the catalyst, the activity with respect to all products
5. Hori, Y., Kikuchi, K. & Suzuki, S. Production of CO and CH4 in electrochemical
remains quite constant, with CO and methane activity dropping reduction of CO2 at metal electrodes in aqueous hydrocarbonate solution.
most, supporting the notion developed above that there are two Chem. Lett. 14, 1695–1698 (1985).
different sites, one for rWGS and methanation (nickel-rich site) 6. Kuhl, K. P., Cave, E. R., Abram, D. N. & Jaramillo, T. F. New insights into
and one for methanol formation (gallium-rich site). Ni5Ga3 was the electrochemical reduction of carbon dioxide on metallic copper surfaces.
tested for a period of over 60 h, after which we tried to regenerate Energy Environ. Sci. 5, 7050–7059 (2012).
7. Rosen, B. A. et al. Ionic liquid-mediated selective conversion of CO2 to CO at
the catalyst through reduction with hydrogen at 350 8C for 2 h. low overpotentials. Science 334, 643–644 (2011).
As can be seen in Fig. 4, the catalyst could be successfully regener- 8. Cole, E. B. et al. Using a one-electron shuttle for the multielectron reduction
ated to its original activity. Reduction with hydrogen yields an of CO2 to methanol: kinetic, mechanistic, and structural insights. J. Am. Chem.
amount of methane equivalent to poisoning of 10% of the catalyst Soc. 132, 11539–11551 (2010).
surface area (see Supplementary Table 3 for details), again 9. Schouten, K. J. P., Kwon, Y., van der Ham, C. J. M., Qin, Z. & Koper, M. T. M.
A new mechanism for the selectivity to C(1) and C(2) species in the
confirming our analysis above.
electrochemical reduction of carbon dioxide on copper electrodes. Chem. Sci.
2, 1902–1909 (2011).
Concluding remarks 10. Chen, Y., Li, C. W. & Kanan, M. W. Aqueous CO2 reduction at very low
overpotential on oxide-derived Au nanoparticles. J. Am. Chem. Soc. 134,
The Ni-Ga catalysts are not optimized but already show interesting
19969–19972 (2012).
activity, selectivity and stability for ambient-pressure CO2 reduction. 11. Peterson, A. A. & Nørskov, J. K. Activity descriptors for CO2 electroreduction
Importantly, they are superior to the existing Cu/ZnO/Al2O3 catalyst to methane on transition-metal catalysts. J. Phys. Chem. Lett. 3, 251–258 (2012).
with respect to their ability to reduce the rWGS activity in favour of 12. Lewis, N. S. & Nocera, D. G. Powering the planet: chemical challenges in
methanol production. A process producing mainly methanol and solar energy utilization. Proc. Natl Acad. Sci. USA 103, 15729–15735 (2006).
water would provide an excellent fuel for a fuel cell41,42 and could 13. Crabtree, G. & Sarrao, J. The road to sustainability. Physics World 22,
24–30 (2009).
be interesting in connection with a decentralized use of solar- or 14. Hansen, J. B. & Nielsen, P. E. H. in Handbook of Heterogeneous Catalysis
wind-generated hydrogen. There are many challenges to be overcome (eds Ertl, G., Knözinger, H. & Schüth, F.) 2920 (Wiley, 2008).
to make such a process viable. A process to efficiently separate CO2 15. Kasatkin, I., Kurr, P., Kniep, B., Trunschke, A. & Schlögl, R. Role of lattice
from air may be the largest43–46, followed by process design and, of strain and defects in copper particles on the activity of Cu/ZnO/Al2O3 catalysts
course, optimization and test, including stability and resistance to poi- for methanol synthesis. Angew. Chem. Int. Ed. 46, 7324–7327 (2007).
16. Behrens, M. Meso- and nano-structuring of industrial Cu/ZnO/(Al2O3)
soning of the catalysts. The Ni-Ga catalysts provide a good starting
catalysts. J. Catal. 267, 24–29 (2009).
point for a new catalyst system based on non-precious metals 17. Kurtz, M., Wilmer, H., Genger, T., Hinrichsen, O. & Muhler, M. Deactivation
showing new and interesting effects of suppressing the rWGS. of supported copper catalysts for methanol synthesis. Catal. Lett. 86,
77–80 (2003).
18. Campbell, C. T., Daube, K. A. & White, J. M. Cu/ZnO(0001) and ZnOx/Cu(111):
Methods model catalysts for methanol synthesis. Surf. Sci. 182, 458–476 (1987).
Density functional theory (DFT) calculations for the intermediates and transition 19. Bowker, M., Hadden, R. A., Houghton, H., Hyland, J. N. K. & Waugh, K. C.
states were carried out on the (211) surfaces of copper, silver, palladium, platinum The mechanism of methanol synthesis on copper/zinc oxide/alumina catalysts.
and rhodium using the Dacapo code (http://wiki.fysik.dtu.dk/dacapo). The
J. Catal. 109, 263–273 (1988).
computational set-up and model surfaces used are identical to those described in
20. Askgaard, T. S., Nørskov, J. K., Ovesen, C. V. & Stoltze, P. A kintic model
ref. 47. Determination of NiGa and Ni3Ga intermetallic compound stability was
of methanol synthesis. J. Catal. 156, 229–242 (1995).
performed as described in ref. 47. DEC and DEO were retrieved from ref. 47, as
21. Fisher, I. A. & Bell, A. T. In-situ infrared study of methanol synthesis from
found in CatApp48. Further information about calculations regarding Ni5Ga3 as
H2/CO2 over Cu/SiO2 and Cu/ZrO2/SiO2. J. Catal. 172, 222–237 (1997).
well as CO adsorption can be found in the Supplementary Fig. 4. Gas-phase values
22. Meitzner, G. & Iglesia, E. New insights into methanol synthesis catalysts
obtained for CO2 and HCOOH were corrected as described in refs 35 and 49.
from X-ray absorption spectroscopy. Catal. Today 53, 433–441 (1999).
Contributions from van der Waals interactions were included as estimated by
comparison to calculations performed with the BEEF–vdW functional described 23. Fujitani, T. & Nakamura, J. The chemical modification seen in the Cu/ZnO
elsewhere.35 Steady-state solutions to the microkinetic model were found as methanol synthesis catalysts. Appl. Catal. A 191, 111–129 (2000).
described in ref. 32. 24. Grunwaldt, J-D., Molenbroek, A. M., Topsøe, N-Y., Topsøe, H. & Clausen, B. S.
Ni-Ga catalysts were prepared using incipient wetness impregnation of a mixed In situ investigations of structural changes in Cu/ZnO catalysts. J. Catal.
aqueous solution of nickel and gallium nitrates (Sigma Aldrich) on silica (Saint- 194, 452–460 (2000).
Gobain NorPro). The samples were directly reduced in H2 for 2 h at 700 8C. The 25. Hansen, P. L. et al. Atom-resolved imaging of dynamic shape changes in
conventional Cu/ZnO/Al2O3 catalyst was prepared following the procedure supported copper nanocrystals. Science 295, 2053–2055 (2002).
described in ref. 40. 26. Kurtz, M. et al. New synthetic routes to more active Cu/ZnO catalysts used
Activity measurements were carried out at a total flow rate of 100 Nml min21 in for methanol synthesis. Catal. Lett. 92, 49–52 (2004).
a tubular fixed-bed reactor with a CO2 to H2 ratio of 3:1 at atmospheric pressures. 27. Waugh, K. C. Methanol synthesis. Catal. Lett. 142, 1153–1166 (2012).
Catalyst loading was 0.472 g for Ni3Ga, 0.476 g for NiGa, 0.474 g for Ni5Ga3 and 28. Yang, Y., Mims, C. A., Mei, D. H., Peden, C. H. F. & Campbell, C. T. Mechanistic
0.167 g for Cu/ZnO/Al2O3 , ensuring that the total amount of nickel and studies of methanol synthesis over Cu from CO/CO2/H2/H2O mixtures: the
gallium in moles matched the amount of copper in Cu/ZnO/Al2O3. The metal source of C in methanol and the role of water. J. Catal. 298, 10–17 (2013).
loading of the Ni-Ga and Cu/ZnO/Al2O3 catalysts was 17 wt% and 48 wt%, 29. Grabow, L. C. & Mavrikakis, M. Mechanism of methanol synthesis on Cu
respectively. The outlet stream was sampled every 15 min using a gas chromatograph through CO2 and CO hydrogenation. ACS Catal. 1, 365–384 (2011).
(Agilent 7890A). 30. Behrens, M. et al. The active site of methanol synthesis over Cu/ZnO/Al2O3
TEM measurements were performed using a FEI Technai TEM operating at industrial catalysts. Science 336, 893–897 (2012).
200 kV. XRD patterns were recorded with a PANalytical X’Pert PRO diffractometer 31. Grabow, L. C. et al. Descriptor-based analysis applied to HCN synthesis
equipped with an Anton Paar XRK in situ cell and a gas flow control system. from NH3 and CH4. Angew. Chem. Int. Ed. 50, 4601–4605 (2011).
32. Lausche, A. C., Hummelshøj, J. S., Abild-Pedersen, F., Studt, F. & Nørskov, J. K.
Application of a new informatics tool in heterogeneous catalysis: analysis of
Received 12 November 2013; accepted 14 January 2014; methanol dehydrogenation on transition metal catalysts for the production of
published online 2 March 2014 anhydrous formaldehyde. J. Catal. 291, 133–137 (2012).

4 NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry

© 2014 Macmillan Publishers Limited. All rights reserved.


NATURE CHEMISTRY DOI: 10.1038/NCHEM.1873 ARTICLES
33. Hammer, B., Hansen, L. B. & Nørskov, J. K. Improved adsorption energetics 46. House, K. Z. et al. Economic and energetic analysis of capturing CO2
within density-functional theory using revised Perdew–Burke–Ernzerhof from ambient air. Proc. Natl Acad. Sci. USA 108, 20428–20433 (2011).
functionals. Phys. Rev. B 59, 7413–7421 (1999). 47. Studt, F. et al. CO hydrogenation to methanol on Cu-Ni catalysts: theory
34. Nørskov, J. K. et al. The nature of the active site in heterogeneous metal and experiment. J. Catal. 293, 51–60 (2012).
catalysis. Chem. Soc. Rev. 37, 2163–2171 (2008). 48. Hummelshøj, J. S., Abild-Pedersen, F., Studt, F., Bligaard, T. & Nørskov, J. K.
35. Studt, F., Abild-Pedersen, F., Varley, J. B. & Nørskov, J. K. CO and CO2 CatApp: a web application for surface chemistry and heterogeneous catalysis.
hydrogenation to methanol calculated using the BEEF–vdW functional. Angew. Chem. Int. Ed. 51, 272–274 (2012).
Catal. Lett. 143, 71–73 (2013). 49. Peterson, A. A., Abild-Pedersen, F., Studt, F., Rossmeisl, J. & Nørskov, J. K.
36. Wellendorff, J. et al. Density functionals for surface science: exchange- How copper catalyzes the electroreduction of carbon dioxide into
correlation model development with Bayesian error estimation. Phys. Rev. B hydrocarbon fuels. Energy Environ. Sci. 3, 1311–1315 (2010).
85, 235149 (2012).
37. Nørskov, J. K., Abild-Pedersen, F., Studt, F. & Bligaard, T. Density functional Acknowledgements
theory in surface chemistry and catalysis. Proc. Natl Acad. Sci. USA 108, F.S., F.A-P., J.S.H. and J.K.N. acknowledge support from the US Department of Energy.
937–943 (2011). This work was partly supported by The Danish National Research Foundation’s Centre for
38. Studt, F. et al. Identification of non-precious metal alloy catalysts for Individual Nanoparticle Functionality (DNRF54) and partly by the Catalysis for
selective hydrogenation of acetylene. Science 320, 1320–1322 (2008). Sustainable Energy initiative, which is funded by the Danish Ministry of Science,
39. Okamoto, H. Ga-Ni (gallium-nickel). J. Phase Equilib. Diffus. 31, Technology, and Innovation. The authors also thank J. R. Rostrup-Nielsen for
575–576 (2010). helpful discussions.
40. Baltes, C., Vukojević, S. & Schüth, F. Correlations between synthesis,
precursor, and catalyst structure and activity of a large set of CuO/ZnO/Al2O3
catalysts for methanol synthesis. J. Catal. 258, 334–344 (2008).
Author contributions
F.S., F.A-P., J.S.H. and J.K.N. contributed to the computational work in this article. I.S.,
41. Arico, A. S., Srinivasan, S. & Antonucci, V. DMFCs: from fundamental
C.F.E., S.D. and I.C. contributed to the experimental work.
aspects to technology development. Fuel Cells 1, 133–161 (2001).
42. Kamarudin, S. K., Daud, W. R. W., Ho, S. L. & Hasran, U. A. Overview on
the challenges and developments of micro-direct methanol fuel cells (DMFC). Additional information
J. Power Sources 163, 743–754 (2007). Supplementary information is available in the online version of the paper. Reprints and
43. Lackner, K. S. A guide to CO2 sequestration. Science 300, 1677–1678 (2003). permissions information is available online at www.nature.com/reprints. Correspondence and
44. Keith, D. W. Why capture CO2 from the atmosphere? Science 325, requests for materials should be addressed to J.K.N.
1654–1655 (2009).
45. Jones, C. W. CO2 capture from dilute gases as a component of modern Competing financial interests
global carbon management. Annu. Rev. Chem. Biomol. Eng. 2, 31–52 (2011). The authors declare no competing financial interests.

NATURE CHEMISTRY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturechemistry 5

© 2014 Macmillan Publishers Limited. All rights reserved.

Você também pode gostar