You are on page 1of 15

Top Catal

DOI 10.1007/s11244-013-0115-1

ORIGINAL PAPER

Aqueous Phase Glycerol Reforming with Pt and PtMo Bimetallic


Nanoparticle Catalysts: The Role of the Mo Promoter
Paul J. Dietrich • Tianpin Wu • Aslihan Sumer •
James A. Dumesic • Julius Jellinek • W. Nicholas Delgass •

Fabio H. Ribeiro • Jeffrey T. Miller

Ó Springer Science+Business Media New York 2013

Abstract The turnover rate (TOR, normalized to sites Keywords Glycerol reforming  Aqueous reforming 
measured by CO chemisorption before reaction) and PtMo bimetallic catalyst  Delta-XANES (DXANES) 
selectivity for the aqueous phase reforming of glycerol Operando X-ray absorption spectroscopy  Mo promoter
have been determined for Pt/C and PtMo/C catalysts. of Pt  Hydrogen production from biomass
While the TOR of PtMo/C is higher than that of Pt/C by
about 4 times at comparable conversion, the selectivity to
C–O bond cleavage is higher, thus reducing the H2 yield at 1 Introduction
high conversion. Under reaction conditions on Pt/C, CO is
observed as the most abundant Pt surface species with a Biomass has been proposed as an alternative feedstock for
fractional coverage of about 0.6 using operando X-ray fuels and chemicals in response to calls to lower depen-
absorption spectroscopy. Since there is little CO in the dence on petroleum. However, petroleum contains low
effluent (CO2:CO ratios [ 100:1, when CO is detected), it oxygen levels, in contrast to the high oxygen content of
is thought that surface CO is converted to H2 and CO2 by biomass. Even conversion by acid hydrolysis or pyrolysis
the water gas shift reaction. DFT calculations suggest that generates oxygen rich bio-oils, which then must be treated
the role of metallic Mo is to alter the electronic properties in extensive downstream processing [1, 2]. To turn bio-oils
of Pt lowering the binding energy of CO and reducing the or biomass derived carbohydrates into usable products,
activation energies of dehydrogenation and C–O bond e.g., energy dense fuels with similar properties to petro-
cleavage. Because the activation energy for C–O cleavage leum derived fuels that can be used as drop-in replace-
is lowered more than for dehydrogenation, the selectivity ments, a large amount of hydrogen must be added to
for C–O bond cleavage is increased, ultimately lowering remove the oxygen. To maintain biomass conversion pro-
the H2 yield compared to Pt/C. cesses as carbon neutral, hydrogen generation must utilize
renewable feed stocks.
Catalytic conversion of biomass via aqueous phase
reforming has been demonstrated as a technically feasible
process for generating hydrogen [3]. These liquid phase
P. J. Dietrich  W. N. Delgass  F. H. Ribeiro (&)
School of Chemical Engineering, Purdue University, reactions are thermodynamically favorable [4] and operate
W. Lafayette, IN 47907, USA at mild temperatures and pressures compared to traditional
e-mail: fabio@purdue.edu hydrocarbon steam reforming. Product distributions can be
tuned by catalyst formulation or conditions to make either
T. Wu  A. Sumer  J. Jellinek  J. T. Miller (&)
Argonne National Laboratory, Chemical Science and gas phase products (H2, syngas) [3, 5, 6] or liquid phase
Engineering, Argonne, IL 60439, USA oxygenates that can be further upgraded into transportation
e-mail: millerjt@anl.gov fuels [7]. Carbon or metal oxide supported Pt catalysts have
been studied in the past, often with a second metal to act as
J. A. Dumesic
Department of Chemical and Biological Engineering, University a rate promoter [8]. Carbon supports have been demon-
of Wisconsin-Madison, Madison, WI 53706, USA strated to be more stable and have high activity under harsh

123
Top Catal

aqueous phase conditions compared to metal oxide sup- and X-ray absorption spectroscopy [11]. Quartz tubes
ports [9]. (6.35 mm/3.124 mm OD/ID, ChemGlass) or Sigradur
Previously, a PtMo/C bimetallic reforming catalyst was G-tube glassy carbon tubes (10 mm/4 mm OD/ID, Hoch-
characterized using X-ray absorption spectroscopy, trans- temperatur-Werkstoffe GmbH) connected to standard
mission electron microscopy, X-ray photoelectron spec- Swagelok fittings with graphite ferrules (Chromalytic
troscopy, density functional theory, and reaction kinetics to Technology, Pty Ltd.) were used as the reactors. The glassy
understand the active phase of the catalyst and how Mo carbon tubes are a unique material that allow for excellent
changes particle morphology [10]. In this work, we com- X-ray transparency and operation at high temperature and
pare this previously characterized catalyst to its monome- pressure [12]. Catalyst loadings varied between 20 and
tallic analog, a Pt/Norit C catalyst, to understand the 250 mg, and the catalyst bed was held in place by stainless
effect of Mo on the catalytic activity. In particular, we steel mesh frits with plugs of quartz wool between the frits
investigate the effect of the secondary metal on rates and and catalyst bed. A type K thermocouple (Omega Engi-
product selectivity, and combine with operando spectros- neering) was inserted into the top of the catalyst bed and
copy and DFT calculations to understand how the elec- used to continuously monitor and control reactor temper-
tronic and morphological properties of the catalyst ature via a PID controller (Omega Engineering) and an
contribute to the differences in activity and selectivity. aluminum heating block. A syringe pump (Teledyne ISCO)
Reaction kinetics and operando experiments were con- was used to control the feed rate of a 30 wt% glycerol in
ducted in a fixed-bed, plug flow system, using aqueous water solution between 0.005 and 0.5 mL min-1, corre-
glycerol as a model compound feed with a 1:1 C:O stoi- sponding to a weight hourly space velocity (WHSV, g feed
chiometry typical of biomass derived carbohydrates. The g cat-1 h-1) range of 0.5 to 100 h-1. System pressure was
results indicate that the addition of Mo increases the rate of controlled by a back pressure regulator (GO Regulators),
reaction, but has an undesired side effect of shifting and 27 mL min-1 Ar was used as a sweep gas through the
product selectivity towards products displaying significant gas/liquid phase separator (Jergusen) to remove gas phase
deoxygenation. Spectroscopy and DFT results indicate that products, which were analyzed online by 6890 and 7890
surface Mo plays a key role in the undesired side reaction, gas chromatographs (Agilent Technologies) equipped with
while still having a promotional effect on the overall rate. packed columns (Carboxen 1000) and thermal conductivity
detectors. Liquid effluent was periodically collected from
the phase separator and analyzed offline in a 7890 GC
2 Experimental equipped with a flame ionization detector and capillary
column (DB-WAX). System mass balances generally
2.1 Catalyst Preparation closed to 100 ± 2 %.

The carbon supported Pt catalyst was prepared by incipient 2.3 Transmission Electron Microscopy
wetness impregnation of a commercially available carbon
black (Norit-SX1G, 1,000 m2 g-1) with an aqueous solu- A suspension of catalyst powder (1–2 mg) in pure ethanol
tion of H2PtCl66H2O (Strem Chemicals) to yield a catalyst (1–3 cc) was dispersed using an ultrasonic bath for about
with 5 wt% Pt loading. For the bimetallic PtMo catalyst, 5 min. Several drops of the resulting solution were spread
Mo was added by incipient wetness impregnation of a over a 200 mesh copper TEM grid coated with a lacey
Mo(CO)6 in THF solution on the dried and unreduced carbon film. HRTEM images of the samples were taken on
5 wt% Pt on carbon catalyst (atomic Pt:Mo ratio of 1:1). an FEI Titan apparatus operating at 300 kV equipped with
The catalysts were dried in air for 12 h at 25 °C. After a Gatan Imaging Filter (GIF). The images were captured
drying, each catalyst was reduced for 2 h in flowing H2 at digitally on a CCD camera (1024 9 1024 pixel) and
200 sccm by heating to 450 °C at 0.5 °C min-1. The cat- recorded using the Gatan DigitalMicrograph software.
alyst was cooled in He and passivated in flowing 2 % O2 in Metal particle sizes and distributions were determined for
He at 25 °C. Prior to reaction, the passivated catalyst was fresh and used catalysts using the ImageJ software pack-
re-reduced for 2 h under 50 sccm of 4 % H2 in Ar at age. The total number of particles counted was 79 and 71,
300 °C. respectively.

2.2 Glycerol Reforming Experiments 2.4 X-Ray Absorption Spectroscopy (XAS)

Glycerol reforming experiments were performed in a spe- Pt L3-edge (11.564 keV) X-ray absorption measurements
cially designed operando plug flow reactor system that were conducted on the insertion device beamline of the
allows for the simultaneous measurement of both kinetics Materials Research Collaborative Access Team (MRCAT,

123
Top Catal

10-ID) at the Advanced Photon Source (APS) at Argonne by standard methods for XANES on the reduced catalyst and
National Laboratory. Ionization chambers were optimized catalyst after gas adsorption. The edge energy was determined
for the maximum current with linear response (ca. 1010 from the position of the maximum in the first peak of the 1st
photons detected s-1). A third detector in the series derivative of the XANES. After adsorption of different gases,
simultaneously collected a Pt foil reference spectrum with the Pt L3 adsorption edge shifts and the DXANES spectra were
each measurement for energy calibration. The X-ray beam obtained by subtracting the XANES spectrum of the reduced
was 0.7 mm 9 1.0 mm, and data was collected in trans- catalyst scanned in He from the spectrum after saturation by
mission mode with 3 min quick scans. adsorbing gas. For extended X-ray absorption fine structure
For ex situ measurements, Pt/C catalysts were reduced in (EXAFS) analysis, experimental phase shift and backscat-
a continuous-flow reactor, which consisted of a quartz tube tering amplitudes were obtained from the Pt foil for Pt–Pt
(100 OD, 1000 length) with Kapton windows sealed by two (coordination number 12 at bond distance 2.77 Å) scattering.
Ultra-Torr fittings. Two ball valves welded to the Ultra-Torr Standard procedures using WINXAS 3.1 software were
fittings served as gas inlet and outlet. The catalyst was gently employed to fit the XAS data. The EXAFS parameters were
pressed into a cylindrical sample holder consisting of 6 wells, obtained by a least square fit in R-space of the k2-weighted
forming a self-supporting wafer. The catalyst amount was Fourier transform (FT) data from 2.7 to 12.2 Å-1 and fit from
calculated to give an absorbance (lx) of approximately 2.0. 1.6 to 3.2 Å. The fit parameters were determined by fitting
The catalysts were reduced in 3.5 % H2/He (100 mL min-1) both the real and imaginary parts of the Fourier transform in
at 300 °C (measured by an internal thermocouple at the R-space of the k2-weighted EXAFS spectra.
position of the samples), purged with He for 10 min, and then
cooled to RT. Traces of oxygen and moisture were removed 2.5 Computational Details
by means of a gas purifier (Matheson PUR-Gas Triple Purifier
Cartridge). XAS spectra were collected on the reduced sam- The computations were performed within the DFT framework
ples in helium at RT. Following spectrum collection on the with the PW91 exchange–correlation functional [13, 14] and
reduced catalyst, the catalyst was exposed to 3.5 % H2/He or the Stuttgart pseudo-potentials/basis sets for Pt and Mo [15] as
1 % CO/He at room temperature. The catalysts were then implemented in the NWChem package [16]. This selection is
purged with He at RT before XAS spectrum collection. To based on extensive tests of various alternative choices, which
remove adsorbed gases, the samples were heated to 300 °C were performed on Pt and Mo atoms and dimers [the details
under He for 30 min, then cooled to RT for subsequent will be given elsewhere (J. Jellinek, A. Sumer, in prepara-
adsorption. Spectra with adsorbed water were captured in the tion)]. For the C, O, and H atoms, the corresponding all-
operando reactor after reduction by filling the reactor with DI electron 6-311G(2df,2pd) basis sets were utilized [15]. The
H2O purged with N2, and scanned at ambient conditions, structural optimizations were carried out using gradient-based
20 °C and 26 bar, and 230 °C and 26 bar (all liquid phase techniques. Spin-unrestricted formalism was used to allow for
water) prior to introducing the glycerol feed. optimization of not only the interatomic distances, but also the
Operando XAS measurements were performed under spin-defined electronic states.
glycerol reforming conditions identical to those of the reaction For the CO adsorption studies, we consider the case of
studies using the reactor described above. Reactions at WHSV 55-atom clusters. Optimized icosahedral (ico, Ih), cubocta-
of 16 and 0.5 h-1, characteristic of low and high conversion, hedral (cubo, Ih), and Cs-symmetry conformations, which are
were conducted. Multiple XAS spectra were collected and about 1 nm in diameter, were chosen as paradigmatic repre-
averaged to obtain adequate signal-to-noise ratio. The Debye– sentatives of Pt55 and its Mo-doped derivatives. The Cs con-
Waller factor (DWF, Dr2) is temperature dependent, and the formation is generated by reflecting half of the ico structure
value at reaction temperature was determined by obtaining the with respect to a plane that contains its central atom and is
XAS spectrum first at the reaction temperature and again after perpendicular to its fivefold axis. The cubo conformation is
cooling to room temperature and assuming the same coordi- consistent with the FCC lattice structure of bulk Pt, but neither
nation number at both temperatures. The gas reaction products the ico, the cubo, nor the Cs conformation is necessarily the
were continuously monitored to determine when the system energetically most preferred form of Pt55 (the imposed sym-
had reached steady state. Liquid products were collected after metries result in a significant speed-up of the computations).
the catalyst had been at steady state for several hours and were The average energy of CO adsorption on the clusters
analyzed off-line. was computed as
EAds ¼ ½ECluster þ pECO  ETotal =p
2.4.1 XAS Data Analysis
where ECluster, ECO and ETotal are the equilibrium energies
The normalized energy calibrated Pt L3 edge X-ray absorp- of the bare cluster, the CO molecule, and the CO-covered
tion near edge spectroscopy (XANES) spectra were obtained cluster, and p is the number of the adsorbed CO molecules.

123
Top Catal

To analyze the energetics of the catalytic reactions of Table 1 Turnover rate (normalized to Pt surface area determined by
interest (removal of a hydrogen atom or a hydroxyl group) CO chemisorption) comparisons for Pt/C and PtMo/C catalysts
including glycerol consumption and gas phase production rates
from glycerol or its fragments, we considered Pt7, Pt5Mo2
and Pt3Mo4, each in its corresponding lowest energy con- Conversion (%) TOR/10-3 s-1 CO2/CH4
figuration (J. Jellinek, A. Sumer, in preparation). All pos- Glycerol H2 CO2 CH4
sible adsorption conformations were obtained on each of
these nano-catalysts, and the complete minimum energy Pt/C 0.4 9.3 6.7 3.7 0.5 7.4
paths for various possible dehydrogenation and dehydr- 6 5.8 17 5.9 – –
oxylation reactions were mapped out for the three lowest 12 3.7 9.9 5.0 1.6 3.0
energy catalyst–reactant complexes. This was done by 14 2.3 5.2 5.6 1.5 3.7
performing constrained energy minimizations on grids of PtMo/C 0.3 17 6.1 6.6 0.7 9.7
C–H, O–H, and C–OH distances. Further refinement of the 5 37 47 41 2.3 –
transition states was obtained by using the structures of the 8 41 64 31 4.4 8.7
highest energy configuration along the grid-based mini- 21 17 25 26 4.6 5.6
mum energy paths as inputs for the so-called mode-fol- 26 5.1 4.6 8.3 1.3 6.6
lowing procedure as implemented in NWChem [16]. The Data collected at 230 °C and 30 bar with varying WHSV
transition states were finally verified by normal mode
analysis and through their connectivity to the reactants and
products of interest. Although the size of the nano-catalysts times higher rate than the Pt/C at a conversion of 5 %.
considered in computations was smaller than those Below 1 % conversion, the rates of hydrogen generation
explored experimentally, it allowed for elucidation of the are similar within the range of the error of measurement. At
global and generic qualitative effects of the nano-catalyst conversions above 1 %, PtMo/C had higher rates than Pt/C.
composition (pure Pt vs. mixed Pt–Mo). PtMo/C also displayed an overall increase in glycerol
conversion rate at all studied conversions. At conversions
up to 40 %, CO2 was preferentially produced over CH4
3 Results (Fig. 1), with CO2/CH4 ratios (a measure of pathway
selectivity) between 3 and 10 (Table 1), indicating a gen-
3.1 Kinetic Characterization eral selectivity to the hydrogen formation pathway. At
conversions below 5 %, selectivity to CO2 and CH4 were
Conversion of glycerol in the aqueous phase was studied nearly equal between Pt/C and PtMo/C. At conversions
over both PtMo/C and Pt/C catalysts. Weight hourly space higher than 5 %, Pt/C showed higher selectivity to gas
velocity (WHSV) was varied to investigate the product phase molecules compared to PtMo/C (5 % higher CO2
distributions as a function of conversion and to probe major
reaction pathways. As in a previous study of the PtMo/C
catalyst [10], the products observed were consistent with
both C–OH and C–C bond cleavage reactions. Carbon–
carbon cleavage reactions of oxygenated hydrocarbons
result in the formation of CO2 and H2, the desired products
for this reaction, whereas carbon–oxygen bond cleavage
results in the formation of liquid phase species with satu-
rated hydrocarbon fragments (i.e. R-CH3 type compounds)
which require hydrogen to form. Thus, the formation of
products resulting from C–C cleavage reactions leads to the
generation of H2, whereas products resulting from C–OH
cleavage reduce hydrogen yields and represent an unde-
sirable side reaction ultimately leading to the formation of
light hydrocarbons.
The major gas phase products identified were H2, CO2,
and CH4. CO was not a major product and was generally
only observed in trace quantities; when it was present,
Fig. 1 Gas phase carbon selectivity to carbon dioxide (diamond) and
CO2:CO ratios exceeded 100:1. A comparison of turnover
methane (square) as a function of conversion comparing the Pt/C
rates for hydrogen at similar conversions (Table 1) indi- catalyst (blue, hollow markers, solid line) to the PtMo/C catalyst (red,
cates that the PtMo/C catalyst produces hydrogen at a 4 filled markers, dashed line)

123
Top Catal

Table 2 Carbon distribution of [C–C] and [C–O] products in the


liquid phase effluent as a function of conversion
Conversion (%) [C–C] (%) [C–O] (%)

Pt/C 0.4 34 66
6 21 79
12 18 82
14 19 81
40 25 75
PtMo/C 0.3 9 91
5 18 82
8 18 82
21 8 92
26 8 92
Data collected at 230 °C and 30 bar with varying WHSV

Fig. 2 Overall carbon selectivity to liquid phase products comparing


selectivity to [C-C] products (squares) and [C-O] products (dia-
selectivity and 2 % higher CH4 selectivity). Despite the monds) as a function of conversion for Pt/C (blue, hollow markers,
differences in selectivity, both Pt/C and PtMo/C had sim- solid line) and PtMo/C (red, filled markers, dashed line) catalysts
ilar CO2/CH4 ratios at the reported conversions (Table 1).
A significant portion of reacted carbon was contained in of the fresh catalyst was 1.8 ± 0.6 nm, with 11 % of the
the liquid phase effluent, from 90 % at low conversions to particles larger than 3 nm. After 6 days on stream under
50 % at intermediate conversions (the remaining carbon glycerol APR conditions, there is an increase in size of the
was converted to CO2 and CH4). Products identified as nanoparticles (Fig. 3c, d and in the histogram Fig. 4). The
liquid phase intermediates include propylene glycol, eth- average size and distribution of the used catalyst is much
ylene glycol, hydroxyacetone, ethanol, 1- and 2-propanol, larger than in the fresh sample (5.2 ± 4.2 nm), indicating
methanol, propionaldehyde, and acetaldehyde. Liquid that the nanoparticles sinter under the reaction conditions.
phase products were grouped by the type of bond cleavage. These particles are of similar average size to those that
Those that undergo solely carbon–carbon bond cleavage were observed in the characterization of the PtMo/C cata-
(ethylene glycol, methanol) were classified as C–C scission lyst (2.0 ± 0.5 nm fresh, 5.1 ± 3.1 nm used [10]).
intermediates ([C–C]). Products resulting from a mixed
pathway of C–OH and C–C bond cleavage, or products 3.2.2 XAS of Reduced Catalyst
that have been deoxygenated (propylene glycol, hydrox-
yacetone, 1- and 2-propanol, ethanol, acetaldehyde, pro- The Pt L3 XANES of Pt foil was compared to that of the Pt/C
pionaldehyde), were designated as C–OH scission catalyst. Prior to reduction, the catalyst had small amounts of
intermediates ([C–O]). The products in the liquid phase oxidized Pt, as expected from a passivated Pt particle. Pt
were generally dominated by [C–O] products for both Pt reduces completely to the metallic state (Pt0) upon H2
(60–80 %) and PtMo (80–90 %) catalysts (Table 2). There reduction at 300 °C. The edge energy, which is determined
are also differences in the liquid selectivity profiles for the from the maximum in the first peak of the 1st derivative of the
Pt vs. PtMo catalyst. As Fig. 2 demonstrates, the Pt catalyst XANES, suggests no significant edge position shift between
had lower selectivity to [C–O] scission products and gen- Pt/C catalyst and Pt foil. The white line intensity above the
erally higher selectivity to [C–C] scission products when edge of XANES on Pt/C is lower than that of Pt foil (Fig. 5),
compared with the PtMo catalyst at all tested conversions. indicating complete reduction of the Pt/C catalyst and nano-
This indicates that the addition of Mo shifts selectivity particles smaller than about 3 nm [17, 18].
towards deoxygenation. Additional information on the Pt particle sizes were
determined by EXAFS spectroscopy at room temperature
3.2 Catalyst Characterization after the catalysts were reduced in 3.5 % H2/He at 300 °C for
30 min. The fit of the Pt–Pt first-shell EXAFS spectra were
3.2.1 TEM obtained by a Fourier transform of the k2-weighted data and
are summarized in Table 3. Figure 6 shows the magnitude of
Transmission electron micrographs for the Pt/C catalyst the Fourier transform (uncorrected phase and amplitude) for
were obtained before and after 6 days under reaction representative data of the Pt/C catalyst and Pt foil. The Pt/C
(Fig. 3). Analysis of the images indicates the average size catalyst has a lower peak intensity in the first shell Pt–Pt

123
Top Catal

Fig. 3 Transmission electron


micrographs of the Pt/C catalyst
for samples of fresh (a, b) and
used for 6 days on stream (c,
d) catalysts, and representative
images of the PtMo/C catalyst
fresh (e) and used (f)

distance (between 1.7 and 3.1 Å) compared to the foil. The show that for the reduced catalyst there is a contraction in the
FT peak size is proportional to the coordination number [17] Pt bond distances (2.71 Å) compared to bulk Pt (2.77 Å),
and can be used as a measure of particle size, thus allowing with a Pt–Pt coordination number of 6.1. Assuming spherical
for the observation of relative particle size during reaction. In nanoparticles [19, 20], the average particle size was esti-
addition, there is a shift to lower R indicating a decrease in mated as 1.7 nm for the Pt/C catalyst, in reasonable agree-
the Pt–Pt bond distance for the Pt/C catalyst, a characteristic ment with TEM results. These estimates were based on
of nanoparticles less than about 3 nm [18]. The fit results previous correlations of the NPt–Pt with dispersion [17].

123
Top Catal

the intensity of these changes are linearly dependent on the


amount of the adsorbed species [21]. The advantage of
XANES spectroscopy is that it is specific only to the
amount of gas adsorbed on Pt and is independent of any
adsorbates on the substrate or other components in the
catalyst. Thus, the relative coverage of gases (i.e. CO and
H2) on the Pt surface under reaction conditions can be
determined from the XANES spectra.
The Pt L3 edge XANES spectra of Pt/C with and without
CO and H2 adsorption (ambient conditions) and water
adsorption (230 °C and 26 bar) are shown in Fig. 7. Using
CO adsorption as a representative example, the white line
intensity of the XANES increased after adsorption, and the
edge shifts to higher energy (*0.5 eV shift). If one sub-
tracts the XANES in He from that with adsorbed CO, the
difference (or DXANES) isolates changes in shape and
Fig. 4 Histogram showing changes in particle size distribution for Pt/ intensity of the edge due to CO adsorption. Figure 8 shows
C catalysts before and after reaction; fresh (1.8 ± 0.6 nm, red) and the DXANES spectra for each of the adsorbate standards
used (5.2 ± 4.2 nm, black, hatched) on Pt/C. Each of the adsorbed species has a distinct shift in
the edge position and XANES shape compared to the
spectra in He.

3.2.4 XAS of Catalysts Under Reaction Condition


(Operando XAS)

Operando XAS studies were carried out to study the


structure of the catalyst during aqueous phase glycerol
reforming at 26 bar and 230 °C at different space velocities
in order to achieve low and high glycerol conversions.
Figure 6 compares the magnitude of the Fourier transform
of the Pt L3 edge EXAFS spectra for Pt/C catalyst during
glycerol reforming at high and low conversion. The mag-
nitudes of the FT are smaller than a Pt foil and similar in
size to that of the reduced catalyst. However, the former
were obtained at reaction temperature which lowers the
intensity of the magnitude of the Fourier transform. The fit
Fig. 5 Pt L3 absorption near edge spectra from 11.54 to 11.59 keV parameters are summarized in Table 3. The fits suggest Pt–
comparing the Pt nanoparticles on the reduced Pt/C catalyst (red,
dashed line) to a Pt foil (black) Pt contributions with coordination numbers around 8.2 and
an interatomic distance of 2.72 Å for both high and low
glycerol conversion. The particle size is estimated at about
3.2.3 XANES of Catalysts After Adsorption of CO, H2 3 nm [17], indicating a slight increase in particle size
and H2O during the first few hours under reaction.
The Pt L3 XANES spectra of the Pt/C catalyst reduced
Reactant and product chemisorption cause changes in the and under glycerol reforming reaction at high and low
Pt L3 XANES spectra and can be used to determine conversion are shown in Fig. 9. The XANES under reac-
adsorbed surface species under reaction conditions [21]. tion shares very similar spectra shape and edge position for
Because the L3 XANES is a 2p to 5d (valence level) both high and low conversions, and both show a positive
transition, the edge energy and shape are sensitive to *0.2 eV edge position shift relative to the spectra taken in
changes in the Pt d-orbital electron density due to bond helium after reduction. The XANES are characteristic of
formation. More specifically, chemisorption of gases, e.g. metallic Pt under reaction conditions. Figure 10 shows the
CO, H2 and H2O, result in different changes to the shape of DXANES of Pt/C catalyst under glycerol reforming con-
the Pt L3 XANES [22]. Thus, changes induced by adsorbed ditions at high and low conversion along with that for CO
species can be used to determine their identity. In addition, saturation at RT (shown in Fig. 8). The DXANES spectra

123
Top Catal

Table 3 EXAFS fit parameters for Pt/C catalysts measured ex situ and in operando
Sample TreatmentScan conditions Edge energy (keV) Pt–Pt N Pt–Pt R (Å) Dr2 (Å2) E0 (eV)

Pt foil As is 11.5640 12 2.77 0.000 -1.1


Pt/C Reduced in H2, RT in He 11.5640 6.1 2.71 0.002 -3.4
Pt/C Reduced in H2, glycerol reforming at 230 °C
High conversion 11.5642 8.2 2.72 0.005 -2.5
Low conversion 8.2 2.73 0.005 -2.9

Fig. 6 Pt L3 Magnitude of the k2-weighted, uncorrected phase and Fig. 8 DXANES from 11.54 to 11.59 keV showing the isolated
amplitude Fourier transform (Dk = 2.7–12.2 Å-1) of the EXAFS for changes to the XANES caused by CO (red, dashed line), H2 (blue,
the Pt/C catalyst under different conditions. Pt foil (black, solid line), dotted line) and H2O (green, dash-dot line, at 230 °C, 26 bar)
Pt/C (red, dashed line), Pt/C with high glycerol conversion (green, adsorption; all spectra recorded at ambient conditions, except where
dotted line), Pt/C with low glycerol conversion (blue, dash-dot line) noted

at both glycerol conversions are very similar in position


and shape to that for Pt/C saturated with CO at RT for Pt/C
(Fig. 8), and distinct from that of either H2 or H2O. Thus,
although CO is observed in trace amounts in the gas phase
effluent, the XANES spectra indicate it is the most abun-
dant surface species under reforming conditions.

3.2.5 DFT Computational Results

Density functional theory calculations were performed on


the Pt/C and PtMo/C catalyst. The latter was described in
previous work [10] and demonstrated the effects of alloy-
ing on the electronic properties of the nanoparticles. For a
monometallic Pt system, three conformations of Pt55 are
shown in Fig. 11. Their computed binding energies are
245.66, 245.35, and 244.40 eV with preferred spin states of
6, 6, and 5 for the Cs, Ih and Oh conformations. Figure 12
Fig. 7 Pt L3 absorption near edge spectra from 11.54 to 11.59 keV shows three possible topologically different arrangements
showing the effects of gas adsorption on the XANES for Pt/C reduced for adsorption of two CO molecules on the energetically
at 300 °C. Pt/C in He (black, solid line) and with adsorbed CO (red,
most stable (of the three considered) Cs Pt55; the two CO
dashed line), H2 (blue, dotted line), and H2O (green, dot-dash line, at
230 °C, 26 bar); all spectra recorded ambient conditions, except molecules are added in a manner that preserves the Cs
where noted symmetry. The computations indicate that atop adsorption

123
Top Catal

Figure 13 displays the results for the cases of the ico and
cubo conformations of the pure Pt55 and mixed Pt43Mo12.
The stoichiometry of the mixed clusters was chosen to
allow for preserving the Ih and Oh symmetries of the cor-
responding parent Pt55. Both symmetries allow for two
types of replacements of 12 Pt atoms by Mo atoms, which
result in different so-called homotopic forms; the latter
differ by the occupation of the sites of a given geometric
structure by different types of atoms at a fixed elemental and
percentile composition [10, 25–27]. The first places the Mo
atoms in the interior of the clusters (Mo forms a 12-atom
shell around the central Pt atom), and the second places the
12 Mo atoms in symmetry-consistent sites on the surface of
the cluster. The homotops with Mo in the interior are more
stable than their counterparts with Mo in the surface [10] (J.
Jellinek, A. Sumer, in preparation). This apparent incon-
Fig. 9 Pt L3 XANES spectra from 11.54 to 11.59 keV of reduced Pt/ sistency with our experimental results was discussed in
C catalyst in He (black, solid line) under low (green, dotted line) and detail in our previous work [10].
high (blue, dash-dot line) conversions compared to the Pt/C catalyst
In each case, 30 CO molecules have been added to the
same 30 surface sites of the ico and cubo conformations.
Inspection of the CO adsorption energies (Fig. 13) indi-
cates that the presence of Mo in the 55-atom Pt-based
nanoparticles leads to a significant reduction in the strength
of adsorption of CO to the Pt sites, regardless of the
location of the Mo. However, the degree of reduction is
larger in the case of homotops with Mo in the interior. The
same qualitative conclusions follow from the results
obtained for Pt and PtMo clusters of other sizes, structures,
and percentile compositions (J. Jellinek, A. Sumer, in
preparation). Our analysis reveals that the fundamental
reason for the general reduction in the CO binding energy
is lowering of the density of electronic states in the vicinity
of the ‘‘Fermi level’’ (i.e. the HOMO) caused by admixing
Mo to Pt [the details will be discussed elsewhere (J. Jel-
linek, A. Sumer, in preparation)].
The computations also clarify the effects of Mo on the
Fig. 10 Pt L3 DXANES spectra from 11.55 to 11.59 keV for the Pt/C
catalysts under operando conditions at low (green, dotted line) and reaction energetics of dehydrogenation versus dehydroxy-
high (blue, dash-dot line) conversions; CO saturated at RT (red, lation, with dehydrogenation being a representative reaction
dashed line) for hydrogen generation, as it has been shown that signifi-
cant dehydrogenation is required for the desired C–C bond
is the energetically most favored configuration (Ead- cleavage to become favorable [28]. Figure 14 displays the
s = 2.62 eV), followed by over an edge (Eads = 2.20 eV), relative energies of the transition states (activation barriers)
and then over a face (Eads = 1.92 eV). The same trend was and the products for the dehydrogenation and dehydroxy-
observed for Pt clusters of other sizes and structures (J. lation reactions of C3O3H7 on Pt7, Pt5Mo2 and Pt3Mo4
Jellinek, A. Sumer, in preparation). The atop arrangement nanocatalysts (C3O3H7 was chosen as a prototypical inter-
has been found experimentally to be the strongest for CO mediate in glycerol reforming); in each case, the energy of
adsorption on small Pt clusters [23] and the (111) face of the reactant C3O3H7-nanocatalyst complex is used as the
the bulk Pt [24]. reference zero level. As the Mo content of the catalyst
The strength of CO adsorption was also used as a increases, the barriers for both dehydrogenation and dehy-
descriptor to analyze the effects of admixing Mo to the Pt in droxylation decrease. The degree of the barrier reduction
metallic nanoparticles. The studies were performed for for dehydroxylation is larger than that for dehydrogenation.
various sizes, structures and composition of pure Pt and These results help to explain the mechanisms through which
mixed Pt/Mo clusters (J. Jellinek, A. Sumer, in preparation). Mo changes the activity of Pt nanoparticles by both

123
Top Catal

Fig. 11 Three symmetric Cs Ih Oh


conformations and binding
energies (BE) of Pt55 (see the
text for details)

BE (eV/atom): 4.47 4.46 4.44

Fig. 12 Configurations and Atop Over Edge Over Face


energies of atop, over bridge,
and over face CO adsorption on
the Cs conformation of Pt55
(orange circle Pt, green circle C
and red circle O)

EAds (eV/CO): 2.62 2.20 1.92

mitigation of the CO poisoning and decrease in the acti- During the first few hours under reforming conditions,
vation energy for dehydrogenation and dehydroxylation. the EXAFS indicates the Pt nanoparticles remains metal-
They also indicate that one may expect lower H2 selectivity lic, but increases in size from 1.7 nm (ca. 0.6 dispersion)
on the Mo containing catalysts, which is a consequence of to about 3 nm (0.3 dispersion) Pt particles. To determine
the larger degree of reduction in the barrier for dehydr- the likely adsorbed species on the catalyst metal particles
oxylation than for dehydrogenation. under reaction conditions, the DXANES spectra of the
As a final remark, the computational results presented sample was compared with the DXANES spectra of the
above are obtained for gas phase nanoparticles, and same sample with adsorbed CO, H2, or H2O. Comparing
therefore they do not incorporate support or solvent effects. the DXANES of the Pt/C catalyst under glycerol reform-
The analysis of the consequences of Mo-doping is, how- ing reaction with that of CO adsorption at RT (Fig. 9), it
ever, general and robust, and it helps explain the experi- is observed that the shapes of the DXANES are very
mentally observed improvement in catalytic activity of similar, indicating that the major surface species during
Mo-doped metallic Pt nanoparticles compared to their reaction is CO, likely a result of the hydrogen generation
mono-metallic Pt. pathway described in our previous work [10]. The
DXANES under reaction is about 0.3 that of CO adsorp-
tion at RT, and nearly identical for both conversions.
4 Discussion Including the DXANES of H2 and H2O in the fits did not
improve the quality of the fit, indicating the coverage of
Spectra of the Pt/C sample before and after adsorption of these species are too low to be reliably estimated. Since
CO (Fig. 7) shows that the presence of CO influences the the Pt under reaction is a different size than that of the
Pt L3 XANES features. Since Pt L3 XANES intensity and fresh catalyst, in order to determine the true CO surface
position are dependent on the d orbital occupancy, these coverage, one must account for the difference in the
changes indicate that CO is chemically bonded. The white number of surface atoms. With the increase in size alone,
line intensity increase can be explained as a change in the DXANES would be lower than that in the fresh cat-
partial density of the d state and the d-band due to a alyst. Taking the ratio of the dispersions (0.6/0.3), the
charge transfer event, i.e. the partial removal of d electron number of surface atoms in the reduced catalyst is twice
density from the metal to the p anti-bonding orbital of that under glycerol reforming. Thus, the true CO surface
CO [29]. coverage is about 0.6.

123
Top Catal

Fig. 13 Configurations and Ico Cubo


energies of CO adsorption on 30
identical surface Pt sites for,
respectively, ico and cubo
conformations of Pt55,
Pt43Mo12(core), and
Pt43Mo12(surface) (orange circle Pt 55
Pt, blue circle Mo, green circle
C and red circle O)

EAds (eV/CO): 2.14 2.11

Pt 43Mo12(core)

EAds (eV/CO): 1.52 1.72

Pt 43 Mo12(surface)

EAds (eV/CO): 1.72 1.80

Fig. 14 Energetics of the most


facile H and OH removal from
C3O3H7 on Pt7, Pt5Mo2 and
Pt3Mo4. R, TS and P stand for
reactant, transition state and
product, respectively; the
energies of the reactant
C3O3H7-nanocatalyst
complexes are chosen as the
reference zero level of energy

While the DXANES spectrum of Pt/C indicates that CO that any CO present was below the detection limit. This
is the predominant surface species, similar analysis on a may be explained by the structure of the PtMo/C catalyst in
PtMo/C catalyst during reforming [10] did not show a which the fractional surface coverage of Pt is about 0.25
measurable contribution from CO adsorption, suggesting versus a similar sized monometallic Pt particle. In addition,

123
Top Catal

the 2 nm bimetallic PtMo nanoparticles rapidly sinter to the initial reaction products and study the relative selec-
about 5 nm under reforming conditions, further reducing tivity of C–C versus C–OH bond cleavage (Fig. 15). At this
the Pt dispersion. Density functional theory also indicates conversion level, significant differences between the Pt and
that CO is more weakly bound to Pt in the bimetallic PtMo/ PtMo catalysts are observed. Considering only the 5 most
C catalyst. Thus, because of the lower dispersion, low Pt abundant species (which at this conversion make up 92 and
surface composition, and weaker CO binding (leading to 96 % of the reacted carbon), the PtMo has a significantly
higher rates of water gas shift, see below), changes to the higher selectivity to [C–O] cleavage compared to Pt/C (79
XANES spectrum by CO adsorption on PtMo catalysts vs. 57 %), whereas the Pt/C catalyst has a much higher
were not observed. selectivity to the [C–C] cleavage compounds (29 vs. 8 %).
The addition of Mo to a Pt catalyst causes changes in The gas phase products observed below 1 % conversion are
both rate and selectivity of the glycerol reforming reaction. predominantly CO2 (and H2). Under these conditions, the
In the PtMo particles, the Mo alloys with the Pt, changing selectivity to CO2 and CH4 are equal (CO2: 12 vs. 12 %;
both the structure and electronic properties of the nano- CH4: 2 vs. 1 %; for PtMo compared to Pt), resulting in
particles. In particular, DFT calculations indicate that the comparable CO2/CH4 ratios. At this conversion level, the
addition of Mo to a Pt system decreases the binding energy differences between the two catalysts are not apparent by
for CO on the surface Pt of the catalyst. These differences only measuring the gas products. However, from an anal-
in CO binding energy help explain the observed activity ysis of the liquid products, the differences in catalyst
differences between the two catalysts. First, a lower CO selectivity are clearly distinguishable. As the conversion
binding energy suggests that it is more difficult for the CO increases, the selectivity to gas products on Pt/C becomes
to stick on the PtMo surface, freeing up reaction sites. higher than for PtMo/C, while maintaining CO2/CH4 ratios
However, perhaps more importantly, the gas phase effluent between 3 and 6, which suggests that the Pt/C catalyst
from both catalysts contains negligible amounts of CO better catalyzes reactions that produce H2 (i.e. C–C bond
(CO2/CO [ 100, when observable). Thus, it is unlikely that cleavage) versus those that yield light alkanes (i.e. C–O
the decrease in binding energy is simply causing additional bond cleavage).
CO to desorb from the PtMo surface and exit the reactor in The underlying cause for this increase in [C–O] cleav-
the gas phase, leading to the negligible coverage observed. age rates is the result of adding Mo to the Pt nanoparticles.
Rather, we note that CO binding energy is often used as a There are two proposed mechanisms for this cleavage
rate descriptor for the water gas shift (WGS) reaction [30]. reaction. The first is via an acid catalyzed dehydration of a
Thus, a decrease in CO binding energy leads to an increase C–OH group by MoOx clusters on the carbon support (as
in the rate of the water gas shift reaction that converts the proposed in [10], followed by subsequent hydrogenation.
surface CO to CO2, removing rate inhibiting CO and Supported MoO3 has been demonstrated to have weakly
freeing up surface sites for reaction. In-house water gas acidic functionality [32, 33], but experiments on MoOx/C
shift experiments on supported Pt and PtMo catalysts have shown no activity for loss of C–O bonds of biomass
confirm that the PtMo has a higher gas phase WGS turn- compounds [34, 35]. Koso et al. [35] have reported activity
over rate than Pt [31]. for C–O bond cleavage on Rh–Mo, and proposed that Mo
While the PtMo/C displayed increased glycerol and H2 sits on the surface of a Rh particle. It was also suggested
TORs compared to Pt/C, the selectivity to intermediate that the Mo and Rh have a synergistic effect in which the
liquid products must also be considered. While the selec- oxygenated group adsorbs to the Mo, and the hydrogen-
tivity ratios in the gas phase, particularly CO2/CH4 ratios, olysis occurs with a nearby Rh. The selectivity for the
were similar between Pt and PtMo at the reported con- PtMo catalyst described here are similar to those on RhMo
versions, the liquid phase carbon distribution was different. catalysts with similar surface configurations, suggesting
The Pt/C catalyst had a higher selectivity to [C–C] type that the reaction mechanisms for deoxygenation may be
cleavage products, whereas the PtMo/C catalyst had a similar. It has also been reported that oxygen binding
higher selectivity to [C–O] cleavage products. As these energy on the PtMo species is significantly higher than on
products are completely converted into gas phase products the pure Pt species [34], which is consistent with the
as conversion approaches 100 %, the PtMo/C catalyst will mechanism proposed by Koso. Following C–O cleavage,
have a higher selectivity to light alkane products, which the deoxygenated species could then be either desorbed
will lower the overall hydrogen yield. In addition, the (likely as the aldehyde or ketone, as keto-enol tautomer-
propensity of the PtMo/C catalyst to catalyze deoxygen- ization generally favors the keto form, i.e. hydroxyacetone)
ation reactions suggests that at higher conversions the or be hydrogenated by the Pt to the corresponding alcohol
differences in liquid phase selectivity will increase. (i.e. propylene glycol). Selectivity results indicate that as
The two catalysts were compared at glycerol conversion hydrogen partial pressures in the reactor increase, partic-
below 1 % (0.3 % for PtMo and 0.4 % for Pt) to determine ularly at higher conversions, the favored products are the

123
Top Catal

it preferentially occupies surface sites, helps to drive down


the Pt–CO binding energy increasing the catalytic activity.
DFT calculations also show that for a PtMo morphology
where the Mo segregates to the center of the particle giving
a bimetallic particle with surface Pt only, the Pt–CO
binding energy decreases even further. With the increase in
the fraction of surface Pt and the much lower Pt–CO
binding energy, this would be expected to significantly
increase the rate over these catalysts. The effect of this
alternate configuration on the selectivity is not known. If
the Mo acts as an oxophilic center as proposed by Koso,
then non-surface Mo should promote both the rate and
display selectivities similar to that of monometallic Pt.
However, if the Mo is an electronic Pt promoter, then
subsurface Mo might still display high C–O bond cleavage
Fig. 15 Comparison of carbon selectivity for various classes of selectivity and low H2 yield.
compounds for Pt/C (blue, solid bars) and PtMo/C (red, hatched bars)
catalysts at 230 °C, 30 bar with 30 wt% glycerol at conversions of
conversions of 0.4 ± 0.1 % (Pt/C) and 0.3 ± 0.02 % (PtMo/C) 5 Conclusions

hydrogenated species. The keto-species made up a signif- The TORs and gas and liquid product selectivity for
icant portion of the liquid products at conversion levels aqueous phase reforming of glycerol on Pt/C and PtMo/C
below 1 % (24 and 37 % selectivity for Pt and PtMo), but have been determined. The TORs of PtMo/C are approxi-
decreased significantly as conversion increased (2 % mately 4 times higher than that of Pt/C at comparable
selectivity at conversion above 20 % for both Pt and conversions. While at low conversions the selectivity for
PtMo). H2/CO2 and CH4 are similar for both Pt and PtMo catalysts,
An alternative explanation for the increased C–O the partially converted liquid products indicate a higher
cleavage selectivity is the Mo modifies the electronic selectivity to C–O bond cleavage products, i.e. those with
properties of the Pt active site. While gas and liquid saturated hydrocarbon fragments, on PtMo/C catalysts. At
product distribution of PtMo/C differs slightly from that of higher conversions these fragments form methane, sug-
Pt/C, the catalysts operate at similar reaction temperature, gesting that PtMo/C will have lower H2 selectivity as
and have comparable TORs and selectivity. While the TOR conversion approaches 100 %.
and selectivity of Mo is not known, it would be expected to Operando EXAFS spectroscopy indicates that the 2 nm
be very different from that of Pt. Thus, in the bimetallic Pt/C nanoparticles are metallic under reaction conditions,
PtMo where Mo occupies 0.75 of the surface, if the Mo and changes in the XANES indicate that CO is the most
TOR were near that of Pt, the gas and liquid products abundant surface species despite CO2:CO ratios above 100
would be expected to be significantly different from that of in the reaction effluent. The CO surface coverage for Pt/C
Pt. The experimental results indicate that the CO2(H2):CH4 is estimated to be about 0.6.
ratio, liquid products and TOR are more similar than fun- DFT calculations indicate that metallic Mo lowers the
damentally different. The changes in rate and selectivity Pt–CO binding energy and also lowers the activation
suggest that Pt is the active site and the Mo induces energies for dehydrogenation and C–O bond cleavage. The
changes in the Pt in the bimetallic PtMo/C catalyst. It is latter is more strongly affected leading to an increase in the
suggested that the role of the Mo, therefore, is to modify C–O bond cleavage selectivity and ultimately lower H2
the energy of the Pt-adsorbate bond energy and activation yields, consistent with experimental results.
energies, altering its chemistry rather than to directly par- Finally, the product selectivity and DFT calculations
ticipate in the reaction. Specifically, Mo appears to pref- suggest that the surface Pt atoms are the active site and the
erentially lower the C–O bond cleavage pathway leading to role of the Mo is to modify the electronic properties of the
lower H2 yield (and higher CH4 yield). Pt-adsorbate bond energies, which lead to the observed
The addition of Mo has both positive and negative changes in TOR and selectivity.
effects for hydrogen generation catalysts. The DFT results
Acknowledgments This material is based upon work supported as
presented here and in previous work [10] provide a blue-
part of the Institute for Atom-efficient Chemical Transformations
print for improving future catalysts. The DFT results (IACT), an Energy Frontier Research Center funded by the U.S.
indicate that the addition of Mo to the particle, even when Department of Energy, Office of Science, Office of Basic Energy

123
Top Catal

Sciences. Use of the Advanced Photon Source is supported by the 15. Schuchardt KL, Didier BT, Elsethagen T, Sun L, Gurumoorthi V,
U. S. Department of Energy, Office of Science, and Office of Basic Chase J, Li J, Windus TL (2007) Basis set exchange: a com-
Energy Sciences, under Contract DE-AC02-06CH11357. MRCAT munity database for computational sciences. J Chem Inf Model
operations are supported by the Department of Energy and the 47:1045–1052
MRCAT member institutions. J.J. was also supported by the Office of 16. Valiev M, Bylaska EJ, Govind N, Kowalski K, Straatsma TP,
Basic Energy Sciences, Division of Chemical Sciences, Geosciences Van Dam HJJ, Wang D, Nieplocha J, Apra E, Windus TL, de
and Biosciences, U.S. Department of Energy under Contract DE- Jong WA (2010) NWChem: a comprehensive and scalable open-
AC02-06CH11357. This research used the resources of the National source solution for large scale molecular simulations. Comput
Energy Research Scientific Computing Center (NERSC), which is Phys Commun 181:1477–1489
supported by the Office of Science of the U.S. Department of Energy 17. Miller JT, Kropf AJ, Zha Y, Regalbuto JR, Delannoy L, Louis C,
under Contract DE-AC02-05CH11231 and of the Laboratory Com- Bus E, van Bokhoven JA (2006) The effect of gold particle size
puting Resource Center (Fusion/LCRC) at Argonne National Labo- on AuAu bond length and reactivity toward oxygen in supported
ratory. The authors would like to thank M. Cem Akatay for his catalysts. J Catal 240:222–234
assistance with the TEM images. 18. Lei Y, Jelic J, Nitsche L, Meyer R, Miller J (2011) The effect of
particle size and adsorbates on L3, L2 and L1 X-ray absorption
near-edge spectra of supported Pt nanoparticles. Top Catal 54:
334–348
References 19. de Graaf J, van Dillen AJ, de Jong KP, Koningsberger DC (2001)
Preparation of highly dispersed Pt particles in zeolite Y with a
1. Huber GW, Iborra S, Corma A (2006) Synthesis of transportation narrow particle size distribution: characterization by hydrogen
fuels from biomass: chemistry. Catal Eng Chem Rev 106:4044–4098 chemisorption, TEM, EXAFS spectroscopy, and particle model-
2. Bridgwater AV, Peacocke GVC (2000) Fast pyrolysis processes ing. J Catal 203:307–321
for biomass. Renew Sustain Energy Rev 4:1–73 20. Frenkel AI, Hills CW, Nuzzo RG (2001) A View from the Inside:
3. Cortright RD, Davda RR, Dumesic JA (2002) Hydrogen from complexity in the atomic scale ordering of supported metal
catalytic reforming of biomass-derived hydrocarbons in liquid nanoparticles. J Phys Chem B 105:12689–12703
water. Nature 418:964–967 21. Guo N, Fingland BR, Williams WD, Kispersky VF, Jelic J,
4. Luo N, Zhao X, Cao F, Xiao T, Fang D (2007) Thermodynamic Delgass WN, Ribeiro FH, Meyer RJ, Miller JT (2010) Determi-
study on hydrogen generation from different glycerol reforming nation of CO, H2O and H2 coverage by XANES and EXAFS on
processes. Energy Fuels 21:3505–3512 Pt and Au during water gas shift reaction. Phys Chem Chem Phys
5. Kunkes EL, Simonetti DA, Dumesic JA, Pyrz WD, Murillo LE, 12:5678–5693
Chen JG, Buttrey DJ (2008) The role of rhenium in the conver- 22. Stakheev AY, Zhang Y, Ivanov AV, Baeva GN, Ramaker DE,
sion of glycerol to synthesis gas over carbon supported platinum- Koningsberger DC (2007) Separation of geometric and electronic
rhenium catalysts. J Catal 260:164–177 effects of the support on the CO and H2 chemisorption properties
6. Kunkes EL, Soares RR, Simonetti DA, Dumesic JA (2009) An of supported Pt particles: the effect of ionicity in modified alu-
integrated catalytic approach for the production of hydrogen by mina supports. J Phys Chem C 111:3938–3948
glycerol reforming coupled with water-gas shift. Appl Catal B 23. Gruene P, Fielicke A, Meijer G, Rayner DM (2008) The
90:693–698 adsorption of CO on group 10 (Ni, Pd, Pt) transition-metal
7. Kunkes EL, Simonetti DA, West RM, Serrano-Ruiz JC, Gärtner clusters. Phys Chem Chem Phys 10:6144–6149
CA, Dumesic JA (2008) Catalytic conversion of biomass to 24. Rupprechter G, Dellwig T, Unterhalt H, Freund HJ (2001) High-
monofunctional hydrocarbons and targeted liquid-fuel classes. pressure carbon monoxide adsorption on Pt(111) revisited: a sum
Science 322:417–421 frequency generation study. J Phys Chem B 105:3797–3802
8. Huber GW, Shabaker JW, Evans ST, Dumesic JA (2006) 25. Jellinek J, Krissinel EB (1996) NinAlm alloy clusters: analysis of
Aqueous-phase reforming of ethylene glycol over supported Pt structural forms and their energy ordering. Chem Phys Lett 258:
and Pd bimetallic catalysts. Appl Catal B 62:226–235 283–292
9. Soares RR, Simonetti DA, Dumesic JA (2006) Glycerol as a 26. Ferrando R, Jellinek J, Johnston RL (2008) Nanoalloys: from
source for fuels and chemicals by low-temperature catalytic theory to applications of alloy clusters and nanoparticles. Chem
processing. Angew Chem Int Ed 45:3982–3985 Rev 108:845–910
10. Dietrich P, Lobo-Lapidus R, Wu T, Sumer A, Akatay M, Fin- 27. Jellinek J (2008) Nanoalloys: tuning properties and characteris-
gland B, Guo N, Dumesic J, Marshall C, Stach E, Jellinek J, tics through size and composition. Faraday Discuss 138:11–35
Delgass W, Ribeiro F, Miller J (2012) Aqueous phase glycerol 28. Liu B, Greeley J (2011) Decomposition pathways of glycerol via
reforming by PtMo bimetallic nano-particle catalyst: product C–H, O–H, and C–C bond scission on Pt(111): a density func-
selectivity and structural characterization. Top Catal 55:53–69 tional theory study. J Phys Chem C 115:19702–19709
11. Fingland B, Ribeiro F, Miller J (2009) Simultaneous measure- 29. Shriver D, Atkins P, Overton T, Rourke J (2009) Inorganic
ment of X-ray absorption spectra and kinetics: a fixed-bed, plug- chemistry. W. H. Freeman, New York
flow operando reactor. Catal Lett 131:1–6 30. Schumacher N, Boisen A, Dahl S, Gokhale AA, Kandoi S,
12. Kispersky VF, Kropf AJ, Ribeiro FH, Miller JT (2012) Low Grabow LC, Dumesic JA, Mavrikakis M, Chorkendorff I (2005)
absorption vitreous carbon reactors for operando XAS: a case Trends in low-temperature water–gas shift reactivity on transition
study on Cu/Zeolites for selective catalytic reduction of NOx by metals. J Catal 229:265–275
NH3. Phys Chem Chem Phys 14:2229–2238 31. Williams WD, Bollmann L, Miller JT, Delgass WN, Ribeiro FH
13. Perdew JP, Wang Y (1992) Accurate and simple analytic repre- (2012) Effect of molybdenum addition on supported platinum cat-
sentation of the electron–gas correlation energy. Phys Rev B alysts for the water–gas shift reaction. Appl Catal B 125:206–214
45:13244 32. Nair H, Baertsch CD (2008) Method for quantifying redox site
14. Perdew JP, Chevary JA, Vosko SH, Jackson KA, Pederson MR, densities in metal oxide catalysts: application to the comparison
Singh DJ, Fiolhais C (1992) Atoms, molecules, solids, and sur- of turnover frequencies for ethanol oxidative dehydrogenation
faces: applications of the generalized gradient approximation for over alumina-supported VOx, MoOx, and WOx catalysts. J Catal
exchange and correlation. Phys Rev B 46:6671 258:1–4

123
Top Catal

33. Nair H, Gatt JE, Miller JT, Baertsch CD (2011) Mechanistic surface sites on rhodium–rhenium catalysts. J Am Chem Soc
insights into the formation of acetaldehyde and diethyl ether from 133:12675–12689
ethanol over supported VOx, MoOx, and WOx catalysts. J Catal 35. Koso S, Ueda N, Shinmi Y, Okumura K, Kizuka T, Tomishige K
279:144–154 (2009) Promoting effect of Mo on the hydrogenolysis of tetra-
34. Chia M, Pagán-Torres YJ, Hibbitts D, Tan Q, Pham HN, Datye hydrofurfuryl alcohol to 1,5-pentanediol over Rh/SiO2. J Catal
AK, Neurock M, Davis RJ, Dumesic JA (2011) Selective 267:89–92
hydrogenolysis of polyols and cyclic ethers over bifunctional

123

You might also like