Sei sulla pagina 1di 5

DOI: 10.1002/cctc.

201600245 Full Papers

A New Approach to the Mechanism of FischerTropsch


Syntheses Arising from Gas Phase NMR and Mass
Spectrometry
Alexis Bordet, Lise-Marie Lacroix, Katerina Soulantica,* and Bruno Chaudret*[a]

We used 13CO labeling to show that gas-phase NMR spectros- yield only 12C-labeled FT products, which evidences the ab-
copy and mass spectrometry are simple tools for mechanistic sence of the incorporation of FeCx carbon atoms in the prod-
investigations of the FischerTropsch (FT) reaction. Thus, mon- ucts. In addition, this approach shows for the first time that
odisperse Fe nanoparticles (NPs) react with syngas to form the formation of 13CH4 at 250 8C does not result from direct
monodisperse iron carbide (FeCx) NPs. As expected, the heat- carbide hydrogenation but from an intermediate step that in-
ing of 13C-labeled monodisperse FeCx NPs under H2 results in volves a reaction between Fe13Cx and H2O to give 13CO2, which
the desorption of the carbide carbons as 13CH4 and, interest- is subsequently hydrogenated. These results rule out the in-
ingly, restores the initial Fe NPs in terms of size and dispersity. volvement of FeCx carbon atoms in the chain growth process
The Fe13Cx NPs catalyze the hydrogenation of 12CO at 210 8C to under our conditions.

Introduction

The mechanism of Fe-catalyzed FischerTropsch Syntheses reaction products (initial fraction of 13C in C2C4 : 1016 %), but
(FTS) has been studied extensively for several decades but still the isotope distribution in the C2C4 fraction led them to con-
remains controversial. One of the difficulties of the study of clude that the carbide may play the role of a chain growth ini-
this mechanism lies in the ability of metals to transform into tiator. The enol mechanism[8] deals with the nondissociative
carbides under catalysis conditions.[1] The exact role of the iron chemisorption of CO that reacts with adsorbed hydrogen to
carbide phase is not totally clear and represents an important form enolic units. A combination of surface polymerization,
source of discussion. It is quite well accepted that iron carbide condensation, and water elimination steps by adjacent groups
is the most active iron phase in FTS catalysis,[2] but the main leads to the formation of hydrocarbons. In an early study,
question is its exact role in the reaction mechanism. Among Emmett and co-workers showed that the incorporation of the
the different FTS mechanisms proposed, three are accepted carbide in the hydrocarbon products (carbide mechanism) is
widely today that involve, respectively, surface carbides, enol only marginal (maximum 30 %).[9] Further investigations, espe-
formation, and CO insertion. cially with 14C-labeled alcohols, led several groups to support
In the surface carbide mechanism,[3] CO undergoes a dissocia- the enol mechanism as the predominant one.[2d, 10, 11] The CO in-
tive adsorption at the surface of Fe to give carbides that are sertion mechanism, also known as the PichlerSchulz mecha-
then hydrogenated to form hydrocarbons. This mechanism nism, was introduced later and consists of the insertion of CO
became accepted widely thanks to the development of surface into a metalalkyl bond that is then hydrogenated to produce
science techniques, which evidenced the absence of oxygen alcohols, alkenes, and alkanes.[12] Variants of this mechanism in-
atoms at the catalyst surface.[4] The surface carbide mechanism clude the insertion of CO into the OH bond of an adsorbed
has received support from many groups,[5] and more recently OH group[13] and even the hydrogenation of CO to lead to
from Niemantsverdriet and co-workers[6] who used DFT calcula- oxygen atoms bonded directly to the metal surface.[14]
tions to model CO hydrogenation over Fe5C2. Very recently, We have shown recently that it is possible to transform
Khodakov et al.[7] investigated the role of the carbon atoms of monodisperse Fe0 nanoparticles (NPs) into monodisperse iron
supported 13C-labeled iron carbide in FTS at 300 8C. They evi- carbide (FeCx) NPs by a two-step procedure. The first step in-
denced the incorporation of the carbide carbon atoms into the volves the synthesis of Fe0 NPs by the reduction of an organo-
metallic complex [Fe{N(SiMe3)2}2]2 under H2 in the presence of
a mixture of hexadecylamine (HDA) and HDAHCl as li-
[a] A. Bordet, Dr. L.-M. Lacroix, Dr. K. Soulantica, Dr. B. Chaudret
INSA; Laboratoire de Physique et Chimie des Nano-Objets (LPCNO) gands.[1516] In the second step, the resulting NPs were reacted
CNRS; UMR 5215 with Fe(CO)5 and H2 to produce iron carbide nanocrystals.[17]
Universit de Toulouse Here we show that monodisperse iron carbide NPs can be
135 Avenue de Rangueil, 31077 Toulouse (France)
formed directly and more easily through the reaction of a mix-
E-mail: bruno.chaudret@insa-toulouse.fr
ture of CO and H2 with Fe0 NPs obtained by the reduction of
Supporting information and the ORCID identification number(s) for the
author(s) of this article can be found under http://dx.doi.org/10.1002/ [Fe{N(SiMe3)2}2]2 under H2 in the presence of palmitic acid (PA)
cctc.201600245. and HDA[15] in a FTS-like process. The carburation is reversible

ChemCatChem 2016, 8, 1727 1731 1727 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
under H2 and leads back to monodisperse Fe0 NPs. In addition, XRD (Figure 1 ad). Thus, after 24 h under syngas, TEM eviden-
following previous work on the use of gas-phase NMR spec- ces the presence of spherical NPs with a mean size of 9.7 nm.
troscopy for reactivity studies,[18] we show here that simple The XRD pattern of these NPs displays the characteristic peaks
NMR experiments can shed light on the role of the carbide of Fe2.2C and no sign of the presence of body-centered cubic
carbon atoms in the FTS and rule out the involvement of car- (bcc) Fe0 peaks. The carbide was further characterized by using
bide carbon atoms in the chain growth process under these Mssbauer spectroscopy at 4 K (Figure 1 e). The analysis shows
conditions. that the NPs contain a mixture of two different iron carbide
phases Fe2.2C (67 %) and Fe5C2 (21 %) together with a small
amount of Fe0 (12 %). Interestingly, the Fe5C2 phase is not visi-
Results and Discussion ble in the XRD pattern, probably because of its lack of crystal-
linity, associated with a low crystallographic symmetry.
The reduction of the precursor [Fe{N(SiMe3)2}2]2 under H2 in the
As the NPs are carburated easily with CO in the presence of
presence of HDA and PA gives homogeneous Fe0 nanocrystals
H2, it was of interest to synthesize 13C-labeled NPs using 13CO
with a tunable average size between 2 and 30 nm. As reported
instead of 12CO to study the low-temperature, low-pressure
previously, the size and the shape of the NPs can be tuned by
FischerTropsch (FT) mechanism and to avoid the presence of
varying the experimental conditions (ligand concentration,
artifacts that result from ligand decomposition, for example.
temperature, reaction time, etc).[12] In this study 9.0 nm Fe0 NPs
The aim of the first experiment was to determine if the in-
were prepared as described in the Experimental Section. The
corporation of carbon into the Fe NPs is reversible and if CC
formation of carbides, inspired by the FTS process, was per-
bonds could be formed during the desorption step. Thus, the
formed by heating the Fe0 NPs at 150 8C in solution in mesity-
isolated 13C-labeled iron carbide NPs were heated (at 210 and
lene in the presence of CO and H2 (syngas, 2 bar/2 bar). Both
250 8C) in a batch reactor under 3 bar H2 for 48 h. At the end
Fe0 and iron carbide NPs were characterized by using TEM and
of the reaction, the NPs were analyzed by using XRD and TEM
(Figure 1 f and g) and the gas-phase was analyzed by using 1H
and 13C NMR spectroscopy (Figure 2). TEM observations evi-
dence the conservation of the morphological characteristics of
the NPs throughout the process. After heating at both temper-
atures, XRD analysis evidences the presence of pure Fe0 NPs,
which indicates that the carbide desorption is complete. The

Figure 1. Analysis of Fe NPs before and after the incorporation of carbon


and the after desorption of carbon: a) TEM image and b) XRD pattern of Fe0
NPs; c) TEM image, d) XRD pattern, and e) Mssbauer spectrum of the iron Figure 2. Gas-phase NMR spectra after carbide desorption under H2 at
carbide NPs; f) TEM image and g) XRD pattern of the NPs after carbon de- 210 8C. a) Gas-phase 1H NMR spectrum and b) gas-phase 13C{1H} NMR spec-
sorption at 210 8C. TEM scale bar: 100 nm. trum.

ChemCatChem 2016, 8, 1727 1731 www.chemcatchem.org 1728 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
NMR spectra show the formation of 13CH4 and H2O as well as the presence of significant amounts of 13CO2 clearly (Fig-
that of very small amounts of 12CH4, which presumably results ure 3 d). The first possibility to form CO2 under these condi-
from ligand decomposition (see Figure S1 for desorption at tions is the water gas shift reaction (WGSR) with water pro-
250 8C). Carbide formation is thus an easily reversible process duced from CO hydrogenation. As we used nonlabeled CO,
that does not modify the NP morphology and that, under H2, a direct WGSR cannot be the origin of 13CO2. An alternative
produces methane. Interestingly, but not surprisingly according possibility to account for the presence of 13CO2 could be direct
to previous studies,[6, 19] we could not find any evidence for the exchange between 12CO and Fe13C to give 13CO and Fe12C after
formation of hydrocarbons other than methane, which means CO dissociation. To test this hypothesis, labeled iron carbide
that no chain growth occurs during the carbide desorption in NPs were heated at 250 8C under 12CO. MS of the gas phase
the presence of H2 alone under these conditions. after 48 h did not show any enrichment of 13CO (Figure S4).
In a further experiment, 13C-labeled iron carbide NPs were This result is in agreement with previous observations of Ben-
reacted with a 1:1 H2/12CO mixture at 210 8C for 48 h. NMR nett and co-workers for a Fe/Al2O3 catalyst at 285 8C.[20] Other
spectroscopy of the gas phase showed the formation of sever- pathways for CO2 formation under FT conditions include the
al hydrocarbons and olefins, as expected for a FTS reaction, reaction between adsorbed oxygen atoms, which likely arises
but none of these products showed any evidence of 13C en- either from CO dissociation or from water, namely, one of the
richment (Figure 3 a and b). This demonstrates the absence of two hypotheses tested here, and adsorbed CO,[21] or decarbox-
the incorporation of the carbide carbon atom into the reaction ylation of organic acids formed through the oxidation of FTS
products under these conditions. Furthermore, XRD analysis of alcohols.[22] Although the decarboxylation of organic acids
the NPs after the reaction confirmed that they still consisted would not explain the formation of CO2 that contains carbide
entirely of iron carbide (Figure SI2). Therefore, the iron carbide carbon atoms, a reaction between carbide carbon atoms and
NPs remain unchanged during the FTS reaction, and the adsorbed oxygen atoms is possible. However, such a reaction
carbon contained in the hydrocarbons formed comes exclu- would also lead to the formation of 13CO, which does not
sively from the gas phase. occur here.
A similar experiment was performed at 250 8C. In this case, To explain the formation of 13C-labeled CO2 here, we consid-
the 1H NMR spectrum evidences again the formation of nonla- ered a direct reaction that occurs between the carbide and
beled hydrocarbons together with a very small amount of water. Indeed, a reaction between carbon and water to pro-
13
CH4 (Figure 3 c). XRD analysis of the NPs after reaction con- duce either CO and H2 or CO2 and H2 is possible but at a very
firmed that they still consisted entirely of iron carbide (Fig- high temperature (> 550 8C).[23] In this case, to investigate this
ure S3). In addition, gas-phase 13C NMR spectroscopy evidences hypothesis, 13C-labeled iron carbide NPs were reacted with H2O

Figure 3. NMR spectra after FT reaction (500 MHz). a) 1H NMR spectrum after reaction at 210 8C. b) 13C NMR spectrum after reaction at 210 8C. c) 1H NMR spec-
trum after reaction at 250 8C. d) 13C NMR spectrum after reaction at 250 8C.

ChemCatChem 2016, 8, 1727 1731 www.chemcatchem.org 1729 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
under Ar at 250 8C. By using both 13C{1H} NMR spectroscopy the exclusive formation of methane without inducing any coa-
and MS, we evidenced the presence of excess 13CO2 in the gas lescence, restructuration, or morphology changes in the NPs. A
phase at the end of the reaction (Figure S5). Therefore, we can similar reaction performed under FischerTropsch synthesis
assume that at 250 8C, the labeled carbide reacts with water (FTS) conditions, that is, in the presence of H2 and CO at
produced during the FTS process to give 13CO2. The reaction of 210 8C, leads to a mixture of unlabeled hydrocarbons and
the as-obtained 13CO2 with H2 can then lead to the formation hence evidences the absence of direct exchange between the
of methane and other hydrocarbons at higher temperature carbide carbon atoms and the carbon oxides present in the
and hence account for the presence of 13CO2 and 13CH4 in the gas phase. However, at 250 8C, the presence of some 13CH4 in
gas phase. At 250 8C, the WGSR (forward and reverse) is slow the products was demonstrated to result from a reaction be-
and the reaction of iron carbide with water is not very efficient, tween iron carbides (Fe13Cx) and water that arises from CO
which explains the amount of 13CO2 and 13CH4 observed by hydrogenation. Therefore, under our reaction conditions, this
using NMR spectroscopy. In addition, the carbidization degree experiment rules out the direct participation of carbide carbon
of the NPs at the end of the reaction remains constant accord- atoms in the FTS process through the surface carbide mecha-
ing to the XRD pattern. This means that some CO present in nism. It could, however, explain some 13C enrichment in hydro-
the syngas is incorporated into Fe NPs to restore the carbide carbons observed previously in FischerTropsch reactions
phase after the exclusion of some 13C converted into 13CO2. through a complex series of reactions that involve water pro-
This hypothesis was confirmed experimentally. Fe12C NPs were duced during the FTS process.
reacted with 13CO and H2 at 250 8C. The resulting NPs were We show in this study how the use of gas-phase NMR spec-
washed and treated again with H2 at 250 8C. NMR spectroscopy troscopy can help in the study of complex mechanisms such
of the gas phase shows the presence of 13CH4 clearly among as that that involves iron carbide in FTS.
the desorption products and thus confirms the incorporation
of carbon that arises from CO(g) into the NPs under FT condi-
tions at 250 8C (Figure S6). This reaction pathway is likely to ex- Experimental Section
plain the incorporation of 13C into the FT products if both All manipulations were performed in anhydrous solvents under an
Fe13Cx and 12CO-containing syngas are used. This can also ac- Ar atmosphere.
count for previous observations as at 300 8C the WGSR is fast
and could lead to a higher amount of labeled hydrocarbons,
as reported by Emmett et al.[9] and Khodakov et al.[7] Preparation of iron carbide NPs
Iron carbide NPs were obtained through the carbidization of pre-
formed Fe0 NPs. Typically, 9.0 nm Fe0 NPs (50 mg, 0.45 mmol of Fe)
were dispersed in mesitylene (9 mL), and the mixture was pressur-
Conclusions ized with CO/H2 (2 bar/2 bar; 6.5 mmol for each gas) at 150 8C for
We report the easy and reversible low-temperature incorpora- 24 h. At the end of the reaction, the 9.7 nm NPs were recovered by
tion of carbon into Fe0 nanoparticles (NPs) to lead to the syn- decantation assisted by a magnet and were washed three times
with toluene (3 5 mL). The NPs were then dried under vacuum.
thesis of monodisperse iron carbide NPs. We show that this
According to elemental analysis, the black powder obtained con-
process preserves the size and shape of the NPs. Furthermore, tained ~ 65 % of Fe. TEM grids were prepared by drop casting a col-
the iron carbide NPs could be labeled easily with 13C to investi- loidal solution that contained NPs on a copper grid covered with
gate the role of iron carbide in Fe-catalyzed CO hydrogenation amorphous carbon. The grids were observed by using a MET JEOL
using simple analytical techniques, namely, gas-phase NMR JEM 1011. XRD samples were prepared in the glovebox and were
spectroscopy and MS (Figure 4). The heating of iron carbide analyzed by using a PANalytical Empyrean diffractometer using
NPs at 210 or 250 8C under H2 leads to the complete desorp- CoKa radiation at 45 kV and 40 mA.13C-labeled iron carbide NPs
tion of the carbon and, in agreement with previous studies, to were prepared by using 13CO in the syngas mixture.

Carbide carbon desorption


Dried 13C-labeled iron carbide NPs (10 mg) were reacted in a batch
reactor under H2 (3 bar) at 210/250 8C for 48 h in the absence of
solvent. At the end of the reaction, the gas phase was analyzed by
1
H and 13C NMR spectroscopy. The 8.8 nm NPs were collected in
a glovebox and characterized by using XRD and TEM (Figure 1 and
Figure S1).

FT reaction
Dried 13C-labeled iron carbide NPs (10 mg) were reacted in a batch
reactor under H2/CO (2 bar/2 bar; 6.5 mmol for each gas) at 210/
Figure 4. Summary of the different reactions investigated using iron carbide 250 8C overnight in the absence of solvent. At the end of the reac-
NPs. tion, the gas phase was analyzed by gas-phase 1H and 13C NMR

ChemCatChem 2016, 8, 1727 1731 www.chemcatchem.org 1730 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
spectroscopy (Figure 3). The NPs were collected in a glovebox and [8] a) R. B. Anderson, R. A. Friedel, H. H. Storch, J. Chem. Phys. 1951, 19, 313.
characterized by using XRD and TEM (Figures S2 and S3). [9] a) J. T. Kummer, L. C. Browning, P. H. Emmett, J. Chem. Phys. 1948, 16,
739; b) J. T. Kummer, T. W. DeWitt, P. H. Emmett, J. Am. Chem. Soc. 1948,
70, 3632 3643; c) H. H. Podgurski, J. T. Kummer, T. W. DeWitt, P. H.
Acknowledgements Emmett, J. Am. Chem. Soc. 1950, 72, 5382.
[10] a) H. H. Storch, N. Golumbic, R. B. Anderson, The Fischer Tropsch and
Related Synthesis, Wiley, New York 1951; b) R. J. Kokes, W. K. Hall, P. H.
The authors thank ERC Advanced Grant (NANOSONWINGS 2009- Emmett, J. Am. Chem. Soc. 1957, 79, 2989 2996.
246763) for financial support, C. Bijani and Y. Coppel for NMR [11] a) B. H. Davis, Fuel Chem. Div. 2002, 237 239.
measurements, and F. Dumestre for discussions concerning the [12] a) H. Pichler, H. Schulz, Chem. Eng. Technol. 1970, 42, 1162; b) G. Henrici-
Oliv, S. Oliv, J. Mol. Catal. 1982, 16, 111 115; G. Henrici-Oliv, S.
synthesis of Fe0 nanoparticles.
Oliv, Angew. Chem. Int. Ed. Engl. 1976, 15, 136 141; Angew. Chem.
1976, 88, 144 150.
[13] a) A. Deluzarche, R. Kieffer, A. Muth, Tetrahedron Lett. 1977, 18, 3357.
Keywords: iron isotope labeling nanoparticles NMR
[14] a) R. S. Sapienza, M. J. Sansone, L. D. Spaulding, J. F. Lynch in Fundamen-
spectroscopy reaction mechanisms tal Research in Homogeneous Catalysis, Vol. 3 (Ed.: M. Tsutsui), Plenum
Press 1979, p. 179.
[1] a) C. S. Huang, B. Ganguly, G. P. Huffman, F. E. Huggins, B. H. Davis, Fuel [15] a) F. Dumestre, B. Chaudret, C. Amiens, P. Renaud, P. Fejes, Science 2004,
Sci. Technol. Int. 1993, 11, 1289 1312; b) M. D. Shroff, D. S. Kalakkad, 303, 821 823; b) L.-M. Lacroix, S. Lachaize, A. Falqui, M. Respaud, B.
K. E. Coulter, S. D. Koehler, M. S. Harrington, N. B. Jackson, A. G. Sault, Chaudret, J. Am. Chem. Soc. 2009, 131, 549 557.
A. K. Datye, J. Catal. 1995, 156, 185 207; c) R. Zhao, J. G. Goodwin, K. [16] a) A. Meffre, S. Lachaize, C. Gatel, M. Respaud, B. Chaudret, J. Mater.
Jothimurugesan, S. K. Gangwal, J. J. Spivey, Ind. Eng. Chem. Res. 2001, Chem. 2011, 21, 13464 13469.
40, 1320 1328; d) N. Sirimanothan, H. H. Hamdeh, Y. Zhang, B. H. Davis, [17] a) A. Meffre, B. Mehdaoui, V. Kelsen, P. F. Fazzini, J. Carrey, S. Lachaize, M.
Catal. Lett. 2002, 82, 181 191. Respaud, B. Chaudret, Nanoletters 2012, 12, 4722 4728.
[2] a) R. B. Anderson, Academic Press, Orlando 302p, 1984; b) M. E. Dry, [18] a) J. Matthes, T. Pery, S. Grndemann, G. Buntkowsky, S. Sabo-Etienne, B.
Appl. Catal. A 1996, 138, 319 344; c) T. R. Motjope, H. T. Dlamini, G. R. Chaudret, H.-H. Limbach, J. Am. Chem. Soc. 2004, 126, 8366; b) F. Novio,
Hearne, N. J. Coville, Catal. Today 2002, 71, 335- 341; d) B. H. Davis, D. Monahan, Y. Coppel, G. Antorrena, P. Lecante, K. Philippot, B. Chau-
Catal. Today 2009, 141, 25 33; e) S. Li, W. Ding, G. D. Meitzner, E. Iglesia, dret, Chem. Eur. J. 2014, 20, 1287 1297; c) L. M. Martnez-Prieto, S. Care-
J. Phys. Chem. B 2002, 106, 85 91. nco, C. H. Wu, E. Bonnefille, S. Axnanda, Z. Liu, P. F. Fazzini, K. Philippot,
[3] a) F. Fischer, H. Tropsch, Brennst.-Chem. 1923, 4, 276; b) F. Fischer, H. M. Salmeron, B. Chaudret, ACS Catal. 2014, 4, 3160 3168.
Tropsch, Brennst.-Chem. 1926, 7, 79 116. [19] a) G. B. Raupp, W. N. Delgass, J. Catal. 1979, 58, 361 369; b) P. Biloen,
[4] a) P. Winslow, A. Bell, J. Catal. 1984, 86, 186; b) J. P. Hindermann, G. J. J. N. Helle, W. M. H. Sachtler, J. Catal. 1979, 58, 95 107.
Hutchings, A. Kiennemann, Catal. Rev. Sci. Eng. 1993, 35, 1. [20] a) D. M. Stockwell, D. Bianchi, C. O. Bennett, J. Catal. 1988, 113, 13; b) H.
[5] a) R. C. Brady III, R. Pettit, J. Am. Chem. Soc. 1981, 103, 1287; b) M. Rit- Matsumoto, C. O. Bennett, J. Catal. 1978, 53, 331.
schel, W. Vielstich, Chem. Ing. Tech. 1980, 52, 327; c) M. Ritschel, H.-W. [21] a) S. Krishnamoorthy, A. Li, E. Iglesia, Catal. Lett. 2002, 80, 77 86.
Buschmann, W. Vielstich, Grundlagenforschung zur Fischer Tropsch-Syn- [22] a) W. K. Hall, R. J. Kokes, P. H. Emmett, J. Am. Chem. Soc. 1960, 82, 1027;
these, BMFT-FB-T 82-020, January 1982; d) A. T. Bell, Catal. Rev. Sci. Eng. b) L. Xu, S. Bao, L.-M Tau, B. Chawla, H. Dabbagh, B. H. Davis, 11th Annual
1981, 23, 203; e) S. C. Chuang, Y. Tian, J. G. Goodwin Jr., I. Wender, J. International Pittsburg Coal Conference Proceedings 1994, 88; c) W. K.
Catal. 1985, 96, 396; f) C. J. Wang, J. G. Ekerdt, J. Catal. 1984, 86, 239; Hall, R. J. Kokes, P. J. Emmett, J. Am. Chem. Soc. 1957, 79, 2983.
g) F. A. P. Cavalcanti, D. Blackmond, I. Wender, R. Oukaci, J. Catal. 1990, [23] a) G. Hermann, K. J. Httinger, Carbon 1986, 24, 429 435.
123, 270.
[6] a) J. M. Gracia, F. F. Prinsloo, J. W. Niemantsverdriet, Catal. Lett. 2009,
133, 257.
[7] a) V. V. Ordomsky, B. Legras, K. Cheng, S. Pul, A. Y. Khodakov, Catal. Sci. Received: March 1, 2016
Technol. 2015, 5, 1433 1437. Published online on April 26, 2016

ChemCatChem 2016, 8, 1727 1731 www.chemcatchem.org 1731 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Potrebbero piacerti anche