You are on page 1of 10

Articles

https://doi.org/10.1038/s41929-022-00745-y

Modulating the dynamics of Brønsted acid sites


on PtWOx inverse catalyst
Jiayi Fu1,2,11, Shizhong Liu2,11, Weiqing Zheng2,11, Renjing Huang   2,3, Cong Wang2, Ajibola Lawal   2,4,
Konstantinos Alexopoulos2, Sibao Liu   5, Yunzhu Wang2, Kewei Yu1,2, J. Anibal Boscoboinik   2,6,
Yuefeng Liu   7, Xi Liu   8, Anatoly I. Frenkel   2,9,10, Omar A. Abdelrahman   2,4, Raymond J. Gorte2,3,
Stavros Caratzoulas2 and Dionisios G. Vlachos   1,2 ✉

Brønsted acid sites on the oxide overlayers of metal–metal oxide inverse catalysts are often hypothesized to drive selective
C–O bond activation. However, the Brønsted acid site nature and dynamics under working conditions remain poorly under-
stood due to the functionalities of the constituent materials. Here we investigate the formation and the dynamics of Brønsted
acid and redox sites on PtWOx/C under working conditions. Density functional theory-based thermodynamic calculations and
microkinetic modelling reveal a complex interplay between Brønsted acid and redox sites and potentially fast catalyst dynamics
at comparable timescales to the Brønsted acid catalysed dehydration chemistry. Combining in situ characterization and probe
chemistry, we demonstrate that the density of Brønsted acid sites on the PtWOx/C inverse catalyst could be modulated by up
to two orders of magnitude by altering the reaction parameters and by the chemistry itself. We elicit an order of magnitude
increase in the acid-catalysed dehydration average reaction rate by periodic hydrogen pulsing.

M
etal–metal oxide (M–MO) catalysts are central to activity to Brønsted acid sites on partially reduced WOx (ref.14).
petrochemical, fine chemical, pharmaceutical and Therefore, it is crucial to understand the dynamics and coupling
biomass upgrade reactions1–10 owing to cooperative of acid and redox sites under various reaction environments. Aside
catalysis among multiple active sites. For example, synthesizing from the ambiguity on the active sites of these materials, the active
selectively high-value C3–C6 terminal diols and aromatics over sites are taken to be stationary.
noble metal catalysts from biomass-derived polyols and furanics In this Article, we investigate Brønsted acid site dynamics on the
by hydrodeoxygenation is challenging. The former molecules core–shell metal/metal oxide PtWOx/C inverse catalyst as a case
require cleaving secondary C–O bonds instead of accessible pri- study. The catalyst is synthesized by atomic layer deposition (ALD).
mary hydroxyls2,3,6,10; the latter entail activating the primary C–O Using density functional theory (DFT) calculations, microkinetic
bonds without reducing the aromatic groups5,7,11,12 that interact modelling (MKM) and experimental kinetics with probe molecules,
strongly with extended noble metal surfaces. Surprisingly, deco- we elucidate the structure of the WOx overlayer, the catalyst phase
rating the surface of supported noble metals (for example, Pt, Ir, behaviour and the dynamics of the acid sites under various work-
Rh) with reducible oxides (WOx, MoOx, ReOx) enables the selec- ing conditions. Unlike the metal and the oxide that are not acidic
tive production of many chemicals, such as terminal diols, and also under our conditions, we demonstrate Brønsted acid catalysis by the
C–O bond scission in m-cresol while preserving aromaticity2–9 and synergy of the elements and identify water dissociation, hydrogen
the ring-opening of tetrahydrofurfuryl alcohol5,6. These multifunc- spillover and oxide reduction as crucial contributors to the dynam-
tional catalysts consist of metal sites, redox sites and Brønsted and ics and density of Brønsted acid sites, leading to profound changes
Lewis acid sites. in the dehydration rate of probe reactions. While the acidity of these
It has been suggested that the Brønsted acid sites activate the sec- materials can be strong compared to microporous aluminosilicate
ondary C–O bond and stabilize the cationic intermediates towards zeolites, their site density is modulated by the chemistry, the oper-
the formation of diols from biomass-derived substrates such as ating conditions, the catalyst pretreatment and the longitudinal
glycerol and tetrahydrofurfuryl alcohol2,3,6,8,13. Establishing struc- position in a reactor. The fundamental insights provide methods
ture–activity relationships is complicated by the intimate coupling for modulating the sites in the selective C–O bond activation pav-
of Brønsted acidity and catalyst reducibility. Specifically, in situ ing numerous routes to catalyst design. Using the dehydration of
Brønsted acid formation is often accompanied by partial reduction tert-butanol as a case study, we demonstrate that periodic pulsing
of the oxide, leading to redox centres (vacancy formation)14. Such of hydrogen allows the PtWOx/C catalyst to achieve alternate states
redox centres have been proposed to activate the primary hydroxyl with about one order of magnitude higher activity than under con-
groups in aromatic compounds5,15,16; others have attributed the stant hydrogen feed.

Department of Chemical and Biomolecular Engineering, University of Delaware, Newark, DE, USA. 2Catalysis Center for Energy Innovation, University
1

of Delaware, Newark, DE, USA. 3Department of Chemical and Biomolecular Engineering, University of Pennsylvania, Philadelphia, PA, USA. 4Department
of Chemical Engineering, University of Massachusetts Amherst, Amherst, MA, USA. 5Key Laboratory for Green Chemical Technology of Ministry of
Education, School of Chemical Engineering and Technology, Tianjin University, Tianjin, China. 6Center for Functional Nanomaterials, Brookhaven National
Laboratory, Upton, NY, USA. 7Dalian National Laboratory for Clean Energy, Dalian Institute of Chemical Physics, Chinese Academy of Science, Dalian,
China. 8In-situ Center for Physical Science, School of Chemistry and Chemical Engineering, Shanghai Jiao Tong University, Shanghai, China. 9Department
of Material Science and Chemical Engineering, Stony Brook University, Stony Brook, NY, USA. 10Division of Chemistry, Brookhaven National Laboratory,
Upton, NY, USA. 11These authors contributed equally: Jiayi Fu, Shizhong Liu, Weiqing Zheng. ✉e-mail: vlachos@udel.edu

144 Nature Catalysis | VOL 5 | February 2022 | 144–153 | www.nature.com/natcatal


Nature Catalysis Articles
a b c

d e
WO3 monolayer W L1 WO3 W L1
WO3 trimer WOx /C-1R (fresh)
1.5 1.5
WO2 trimer PtWOx /C-1R (fresh)
PtWOx /C-1R (reduced)
Normalized µ(E )

Normalized µ(E )
1.0 1.0

0.5 0.5

Simulation Experiment
0 0
12,080 12,120 12,160 12,080 12,120 12,160
Energy (eV) Energy (eV)

Fig. 1 | Model structures and XANES spectra. a, WO3 monolayer on Pt(111). b, W3O9 cluster on Pt(111). c, W3O6 cluster on Pt(111). Colour code: blue, Pt;
grey, W; red, O. d, Simulated W L1 XANES spectra of model structures. μ(E), absorption coefficient; E, photon energy of the incident X-ray. e, Experimental
W L1 XANES spectra of WO3 standard, WOx/C-1R and PtWOx/C-1R. Reduction at 400 °C. The scaled raw data without background removal are shown in
Supplementary Fig. 1. 1R indicates 1 ALD cycle.

Results dissociation on WOx and direct H2O dissociation on WOx.


Morphology, coordination and oxidation state of WOx/Pt. The dissociative adsorption of H2 on Pt has an enthalpy
Establishing a representative structural model is important. change of −0.95 eV and is non-activated. Upon dissociation, H
Although tetrahedral WOx clusters on various oxide surfaces have diffusion from Pt onto the W3O9 cluster and formation of an acidic
been reported5, the structure of WOx on noble metals under reduc- terminal hydroxyl group (Supplementary Fig. 2a) has an activation
tive conditions has not been explored. We investigated three struc- barrier of 0.15 eV and releases 0.08 eV of enthalpy (driving force
tural models: a monolayer on Pt(111) with W atoms coordinating for spillover). Various terminal sites for the OH group have been
to five oxygen atoms (Fig. 1a), as previously reported17; a tetrahedral investigated; two examples are shown in Supplementary Fig. 3a and
W3O9 cluster with two terminal W centres and a bridge W centre5, Supplementary Fig. 3b. Bridging W–OH–W sites (Supplementary
representative of oligomeric clusters on Pt(111) (Fig. 1b); and a Fig. 3c) are energetically less favoured than terminal ones by 1.10 eV.
W3O6 trimer representing a reduced structure (Fig. 1c). Figure 1d The direct H2 dissociation on the WO3 monolayer has a prohibitively
shows simulated X-ray absorption near edge structure (XANES) high barrier of 1.97 eV (transition state shown in Supplementary
spectra (W L1 edge). The W3O9 cluster model shows an intense Fig. 2b), consistent with experimental reports that hydrogen
pre-edge feature, characteristic of tetrahedral W(VI) centres—this activation over WOx–C and WOx–ZrO2 catalysts requires tempera-
feature is absent from the monolayer model18,19. Our experimen- tures over 200 °C (refs. 6,20). The transition state for proton forma-
tal W L1 XANES spectra display a similar intense pre-edge fea- tion by dissociative water adsorption is shown in Supplementary
ture at 12,095 eV for PtWOx/C-1R and a smaller pre-edge peak for Fig. 2c. The reaction is exothermic by −0.10 eV and requires only
WOx/C-1R and WO3 (Fig. 1e), indicating more tetrahedrally coor- 0.27 eV of activation. H2O dissociation creates a proton on a ter-
dinated W centres on Pt than on carbon or a bulk oxide. The smaller minal O site and a hydroxyl on a W site (the bridging OH is not
pre-edge peak is attributed to the distorted octahedral W centres. energetically favoured).
Reducing the W(VI) trimer to a W(IV) trimer decreases the calcu-
lated pre-edge feature, also seen in the experimental spectrum of the Pt-assisted self-inhibited catalyst dynamics. Figure 2a illustrates
reduced PtWOx/C-1R (Fig. 1e). The scanning transmission electron pathways encountered in the inverse-catalyst dynamics under vari-
microscopy (STEM) characterization6 and the XANES spectra sug- ous working conditions. As we show below, the extent of these sur-
gest a sub-monolayer of WOx. Thereby, we conclude that the WOx face reactions governs the active site densities. The facile Pt-assisted
overlayer in PtWOx/C-1R is distinct from WO6 units in bulk WO3 reduction of WO3 on several oxides below 400 °C was demonstrated
and WO2 and is faithfully represented by a W3O9 trimer on a Pt(111) using temperature-programmed reduction (TPR) and X-ray photo-
slab (Fig. 1b). electron spectroscopy (XPS)21,22. By contrast, bulk WO3 (without Pt)
remains fully oxidized at the same conditions23. Figure 2b depicts
Formation mechanisms of Brønsted acid sites. We have exam- the reduction/oxidation cycle on the W3O9 trimer, including H2
ined three formation mechanisms of Brønsted acid sites on the activation, H spillover, WOx reduction, the reverse WOx hydration
PtWOx catalyst, including H2 dissociation on Pt followed by H dif- and the first-principles predicted energy profile of the cycle. Pt acti-
fusion and attachment to a terminal or bridge O of the WOx cluster vates H2 and protonates W3Ox (Supplementary Fig. 4). Additional
(hereinafter referred to as the spillover mechanism), direct H2 H spillover can reduce W3Ox by H2O elimination (Supplementary

Nature Catalysis | VOL 5 | February 2022 | 144–153 | www.nature.com/natcatal 145


Articles Nature Catalysis

a c
Protonation

Brønsted acid site +1


Oxidation state –1 W3O7H3 W3O7H1
0.10
Hydrogen pressure

Reduction-I
0.08
Brønsted acid site +0 W3O5
0.06

PH (bar)
Oxidation state –2
0.04

2
Reduction-II
0.02
Brønsted acid site –1
Oxidation state –1 0 W3O9
100 150 200 250 300
Temperature (°C)
Oxidation-I d
0.10 0.86
Brønsted acid site +0

Brønsted acid density


Oxidation state +2 0.08

Oxidation-lI 0.06

PH (bar)
Water pressure

0.04

2
Brønsted acid site +1
Oxidation state +1
0.02
Hydration
0
Brønsted acid site +2 100 150 200 250 300 0
Oxidation state +0 Temperature (°C)
e
0.66
0.10

Dehydration

Redox centre density


0.08

0.06
PH (bar)

Brønsted acid site –2


Temperature

Oxidation state +0 0.04


2

Deprotonation
0.02
Brønsted acid site –1 0
Oxidation state +1 100 150 200 250 300 0
Temperature (°C)

b 1

0
TS2
Potential energy (eV)

W3O6 + H2 TS1

–1
W3O9 + H2

–2

–3
Reaction coordinate

Fig. 2 | Pathways, energy profiles of W3Ox reduction and phase behaviour. a, Possible reactions on W3Ox. Colour code: blue, Pt; grey, W; red, O; white, H.
b, Energy profiles of Brønsted acid and redox site formation by H spillover to W3O9 and W3O6 and oxide reduction. TS, transition state. c, Ab initio phase
diagram for the W3Ox/Pt(111) model. Each square represents the most abundant state under the corresponding conditions. Colour code: black squares,
W3O9; grey squares, W3O5; orange squares, W3O7H1; red squares, W3O7H3. d, Microkinetic heatmap of Brønsted acid site density (number of OHs per
W) at equilibrium. e, Microkinetic heatmap of redox centre density (number of Ws with oxidation state ≤ +4 per W) at equilibrium. Larger clusters and
heterogeneous distribution of cluster sizes provide a continuum spectrum of catalyst states. Water partial pressure is 0.1 bar. PH2, hydrogen partial pressure.

Fig. 5), forming an oxygen vacancy. The reduction of W3O9 to adsorption followed by the notably activated H2O dissociation
W3O8 is thermodynamically downhill and requires minor activa- (0.93 eV). In summary, catalyst re-oxidation is energetically harder
tion (0.3 eV). The reverse process of W3O8 oxidation entails H2O than reduction.

146 Nature Catalysis | VOL 5 | February 2022 | 144–153 | www.nature.com/natcatal


Nature Catalysis Articles
a
t=0s t = 10–2 s t=1s t = 102 s

0.10

0.08

0.06
PH (bar)
2

0.04

0.02

0
100 150 200 250 300 100 150 200 250 300 100 150 200 250 300 100 150 200 250 300
Temperature (°C)

b
t=0s t=1s t = 103 s t = 106 s

0.10

0.08

0.06
PH (bar)
2

0.04

0.02

0
100 150 200 250 300 100 150 200 250 300 100 150 200 250 300 100 150 200 250 300
Temperature (°C)

Fig. 3 | Predicted dynamics of the catalyst state. a, Catalyst state versus time (W3O9 at t = 0). b, Phase behaviour for W3O5 at t = 0. Water partial pressure
is 0.1 bar, surface-to-volume ratio = 10−20 cm−1. Black, W3O9; grey, W3O5; light grey, W3O8; orange, W3O7H1; light orange, W3O8H1; red, W3O7H3.

Notably, the kinetics of WOx reduction depends strongly on diagram (Fig. 2c). In the absence of hydrogen, W3O9 remains
the oxidation state of WOx. Figure 2b compares the electronic oxidized (the black region in Fig. 2c) and does not possess either
energies of reduction of W3O9 and W3O6, chosen as representative Brønsted acid or redox sites. With increasing hydrogen pressure,
examples. Reduction of W3O6 is not as thermodynamically driven the catalyst is first partially reduced to W3O7H1 over a broad tem-
as that of W3O9. Desorption of H2O (the last step of the cycle) is perature range, then protonated to W3O7H3 at low temperatures or
energetically costly. reduced to W3O5 at high temperatures with a concomitant decrease
Figure 2b depicts the energy profiles for the first protonation. in Brønsted acid density (Fig. 2d) and an increase in redox centre
We conducted comprehensive DFT calculations for subsequent density (Fig. 2e). The degree of protonation and reduction is consis-
protonations on WOx from W3O4 to W3O9 (Supplementary Table 1 tent with the DFT energy profiles (Fig. 2b).
and Supplementary Fig. 6 present all reactions and models, respec- To unravel the catalyst dynamics, we built a discrete state-based
tively). The enthalpy change for the protonation of W3O9 to W3O9H1 microkinetic model24 to simulate the catalyst state when it is
is −0.08 eV; for the protonation of W3O9H2 to W3O9H3 it is 0.43 eV exposed to different conditions. When the catalyst is not trapped
(reactions 64 and 67, respectively, Supplementary Table 1). This trend in a metastable state, the long-time kinetic behaviour (Fig. 3a) coin-
holds for reduced clusters as well. For example, protonation of W3O5 cides with the thermodynamic behaviour, as happens here (Fig. 2c).
by H spillover becomes endothermic (reaction 5, Supplementary High H2O partial pressure leads to more Brønsted acid sites at high
Table 1); H prefers binding to Pt rather than W3O5. As a result, and temperatures because equilibrium favours the dissociative adsorp-
depending on conditions, not all terminal oxygens are protonated. tion of H2O (Supplementary Fig. 7).
The protonation slows down as it proceeds, that is, it is self-inhibited. Figure 3a shows the predicted predominant catalyst state
In summary, H spillover first protonates the terminal oxygens as a function of time when the catalyst is exposed to different
of the catalyst and eventually eliminates them in the form of water, environments starting from the fully oxidized state, W3O9.
creating redox centres, whereas water hydroxylates and re-oxidizes At lower temperatures, Brønsted acid sites form quickly by
the catalyst. Interestingly, the protonation and reduction are hydrogen spillover and mild reduction occurs below 140 °C.
self-inhibited; WOx does not become either fully reduced or fully Further catalyst reduction occurs slowly at 150–200 °C and accel-
protonated. Instead, it dynamically responds to the reaction condi- erates above 200 °C, due to the higher energy barriers for the
tions, controlling site identity and density, as discussed below. The reduction of W3O7H1 to W3O5. This analysis has several impli-
protonation, reduction and oxidation kinetics depend on the degree cations. First, the in situ formation of Brønsted acid sites is
of catalyst protonation and oxidation. rapid. Second, a gradual increase in redox site density at the
expense of Brønsted acidity is anticipated at 140–250 °C. Third,
Phase diagram and dynamic behaviour. To examine the active above 250 °C, a swift reduction of WO3 leads to a significant loss of
sites under working conditions, we calculated the ab initio phase Brønsted acidity.

Nature Catalysis | VOL 5 | February 2022 | 144–153 | www.nature.com/natcatal 147


Articles Nature Catalysis

a 1.0 b 6 XANES at the Pt L2 edge indicates a readily reduced Pt/C to metal


0.1 bar H2 No H2 0.1 bar H2 No H2 at 250 °C and a slightly oxidized Pt on PtWOx/C-1R even at 400 °C,
as indicated by the somewhat higher white line peak at 13,276.5 eV
0.8
(Supplementary Fig. 12). The results suggest charge transfer from Pt
Brønsted acid density

125 °C 5 to the WOx overlayer with a stable, partially reduced WOx overlayer.

Oxidation state
0.6 Experimental results are qualitatively consistent with model predic-
tions (Fig. 2c,e).
0.4 In situ XPS was also conducted over WOx/C-1R, PtWOx/C-1R
150 °C 4 125 °C and 10%PtWOx/C-15R. Due to the high WOx loading of
0.2 150 °C 10%PtWOx/C-15R, its W 4f signal was much stronger than that of
175 °C 175 °C PtWOx/C-1R, allowing a reliable fit of the spectra. Before reduction,
all catalysts were treated under reduced pressure at 250 °C. At 250 °C,
0 3
0 5 10 15 20 0 5 10 15 20
WOx/C-1R undergoes minimal changes (Fig. 5b), whereas W(V) of
Time (s) Time (s) 10%PtWOx/C-15R increases slightly (Fig. 5c). At 400 °C, W(VI)
doublets of the WOx/C-1R still dominate; by contrast, the W(V) dou-
Fig. 4 | Predicted kinetic effect of H2 starting from a fully oxidized W3O9 blet contribution of the 10%PtWOx/C-15R increases significantly.
state. a, Brønsted acid site density (number of OHs per W atom). b, The spectrum of PtWOx/C-1R resembles that of 10%PtWOx/C-15R
Oxidation state. Initially, the catalyst is exposed to 0.1 bar H2 and 0.1 bar (Supplementary Fig. 13). Noble-metal-promoted reduction of WOx
H2O. H2 is turned off at 10 s. The oxidation state of W is calculated from the to partially reduced 5+ and 4+ states above 400 °C was reported
number of W=O, W–O–W and W–OH bonds. previously25,26 and is supported by TPR results (Supplementary
Fig. 14). XPS results are qualitatively consistent with the X-ray
absorption spectroscopy (XAS) data and demonstrate the impor-
Expectedly, the kinetic behaviour varies with the initial catalyst tance of Pt in modulating the oxidation state of the WOx overlayer
state. Figure 3b depicts the dynamics starting from a reduced state, by facilitating partial reduction only.
W3O5. Unlike W3O9, which suffers rapid reduction under high H2
partial pressure, W3O5 undergoes reduction at a slower rate. First, Exposing the dynamics of acidity using alcohol dehydration. We
the W3O5 domain is replaced with W3O7H1 at 120–250 °C; below have used the dehydration of 2-propanol to correlate the catalyst state
120 °C the catalyst is oxidized to W3O7H3 after 1,000 s. Up to this with activity. All samples were pre-reduced at 250 °C for 1 h to remove
point, the kinetic catalyst state agrees with the thermodynamic one any PtOx. Over the metallic sites of Pt/C, 2-propanol undergoes
except when no H2 is present. It takes 106 s to overcome the kinetic almost exclusively dehydrogenation to produce acetone (Fig. 6a,b),
barriers and fully oxidize W3O8Hx to W3O9. A slightly reduced cata- consistent with previous work10. Pt and C show negligible dehydra-
lyst would not become fully oxidized at relevant timescales. tion activity. On the other hand, WOx dehydrates 2-propanol to
To better understand the catalyst dynamics upon exposure to propene (Fig. 6b). PtWOx/C-1R exposes multifunctional behaviour:
H2 and H2O, we plot the time-dependent Brønsted acid site density WOx Brønsted acid sites dehydrate 2-propanol; Pt dehydrogenates
and oxidation state at three temperatures (Fig. 4). Upon exposure to 2-propanol to acetone and H2 and subsequently hydrogenates pro-
H2, the Brønsted acid density increases rapidly at short times owing pene—with the produced H2—to propane. The substantially lower
to fast H spillover from Pt. Consistent with the phase diagram, the dehydrogenation activity compared to Pt reveals a large fraction
maximum Brønsted acid density is anticorrelated with temperature. of Pt sites covered with WOx, in agreement with energy disper-
Interestingly, after reaching a maximum, the Brønsted acid site den- sive X-ray spectroscopy (EDX) results (Supplementary Fig. 15)6.
sity decreases due to the partial reduction of WOx (Fig. 4b). When Increasing the number of ALD cycles of W on Pt/C increases the
H2 is turned off, the Brønsted acid site density decreases sharply at dehydration activity at the expense of metal-catalysed 2-propanol
first and then gradually at lower temperatures (<125 °C). By con- dehydrogenation and propene hydrogenation.
trast, at 175 °C, the Brønsted acid site density first increases, owing To eliminate the complication from the in situ-formed hydro-
to the re-oxidation of the redox centres by H2O and eventually gen (from 2-propanol dehydrogenation) causing online catalyst
decreases via backward H spillover to Pt and H2 desorption. Overall, reduction, we investigated tert-butanol dehydration (Fig. 6a). We
the complex dynamic catalyst behaviour is found to be controlled compared the PtWOx/C-1R catalysts, reduced at 250 and 400 °C,
by rapidly forming Brønsted acid sites, slower catalyst reduction by with the untreated (fresh) catalyst. Figure 6c depicts a monotonic
H2 and even slower re-oxidation by H2O. As the amounts of H2 and decrease in dehydration rate with increasing reduction temperature
H2O change along the reactor, as well as with time and temperature, by nearly eight times due to the reduction of W(VI) to W(V) and
the catalyst protonation and oxidation state will change dynamically. W(IV) (XANES data; Fig. 5a) and the lowering of the Brønsted acid
Clearly, a static picture of acidity does not hold for these materials. site density, consistent with the predicted phase behaviour (Fig. 2c).
To disentangle the interplay of redox and Brønsted acid sites,
In situ catalyst characterization under reducing conditions. To we annealed the catalyst at 400 °C in a mixture of N2 and He. The
assess the computational predictions, we investigated the stabil- annealed catalyst is less active than the fresh one (Fig. 6d), due to the
ity and the electronic state of PtWOx/C-1R at various reduction loss of OHs by dehydration, but more active than the reduced cata-
temperatures and hydrogen partial pressures. CO chemisorption lyst. The surface OH groups are partially regenerated by steaming the
reveals that roughly 32% of the Pt surface is covered by WOx clus- annealed catalyst in 0.1 bar steam at 400 °C (grey square in Fig. 6d).
ters (Supplementary Note 1) and most of those clusters remain on The initial dehydration activities of the pretreated catalysts (Fig. 6d)
Pt after reduction and annealing treatments at 400 °C. The W L3 are in satisfactory agreement with the Brønsted acid site densities
edge energy of PtWOx/C-1R is 10,209.8 eV (Fig. 5a), close to that of measured by reactive gas chromatography with tert-butylamine
WO3 (10,209.4 eV). Mild reduction at 250 °C with 0.1 bar of H2 leads (Supplementary Fig. 16). For PtWOx/C-1R, this acid quantification
to a white line intensity decrease and a slight edge energy reduction technique is preferred over Fourier transform infrared spectroscopy
(0.2 eV), suggesting a W(VI) oxidation state (Fig. 5a). Reduction of adsorbed pyridine because of the low Brønsted acid site density
at 400 °C further reduces WOx to a state between 4+ and 5+, esti- of PtWOx/C-1R and the strong absorption of infrared by the carbon
mated based on the W L3 edge energy (Supplementary Fig. 10). support. The adsorbed pyridine cannot be observed with Fourier
By contrast, WOx/C-1R is not reduced (Supplementary Fig. 11). transform infrared spectroscopy even after a tenfold dilution with

148 Nature Catalysis | VOL 5 | February 2022 | 144–153 | www.nature.com/natcatal


Nature Catalysis Articles
a b c
6
PtWOx/C-1R (fresh) WOx/C-1R W 4f 10%PtWOx/C-15R W 4f
PtWOx/C-1R (Tr = 250 °C) Vacuum, 250 °C Vacuum, 250 °C

PtWOx/C-1R (Tr = 400 °C)


WO3
WO2

4
0.5 mbar H2, 250 °C 0.5 mbar H2, 250 °C
Normalized µ(E)

Intensity (a.u.)

Intensity (a.u.)
2
0.5 mbar H2, 400 °C 0.5 mbar H2, 400 °C

PtWOx/C-1R Fresh 250 °C 400 °C

W ox. state 6+ <6+ 4 to 5+


0
10,205 10,210 10,215 10,220 40 38 36 34 32 40 38 36 34 32
Energy (eV) Binding energy (eV) Binding energy (eV)

Fig. 5 | In situ characterization of the W oxidation state. a, W L1 XANES spectra of PtWOx/C-1R. Reduction in 0.1 bar of flowing hydrogen for 1 h. Pink
(tiny) area, standard deviations of triplicates; Tr, reduction temperature. b, W 4f XPS spectra of WOx/C-1R. c, W 4f XPS spectra of 10%PtWOx/C-15R.
Sample heated under reduced presure at 250 °C for 1 h, cooled down to room temperature overnight and reduced for 1 h in 0.5 mbar of H2 at temperatures
indicated. Black line, experimental data; grey line, background; red line, XPS fitting curve; violet line, W(VI); blue line, W(V).

a b
70
Propene
Propane
60 Acetone

50
Initial yield (%)

40

30

20

10

0
Pt/C WOx /C-1R PtWOx /C-1R PtWOx /C-2R

c d e f
Initial dehydration rate (×10–4 mol s−1 gcatalyst−1)

Initial dehydration rate (×10–4 mol s−1 gcatalyst−1)


Initial dehydration rate (×10–4 mol s−1 gcatalyst−1)

gcatalyst )

5 5
−1

No pretreatment 5 No pretreatment Annealed at 400 °C (0.1 bar H2 cofeed)

Annealed at 400 °C; 2.5 Annealed at 400 °C (no H2 cofeed)


steamed at 400 °C Reduced at 400 °C (0.1 bar H2 cofeed)
4 4 Reduced at 400 °C (no H2 cofeed)
Annealed at 400 °C 4
−1

2.0
Dehydration rate (×10 mol s

3 3 3
1.5
–4

2 2 2 1.0

1 1 1 0.5

0 0 0 0
No 250 400 0 0.05 0.15 0 0.02 0.10 0 2 4 6 8
reduction Partial pressure of water Partial pressure of hydrogen Time on stream (h)
Reduction temperature (°C) cofeed (bar) cofeed (bar)

Fig. 6 | Probing active sites with alcohols. a, Reaction schemes. b, Reaction of 2-propanol. Catalysts reduced in 0.1 bar H2 at 250 °C for 1 h and purged
with pure He before reaction. Reaction conditions: 4 mol% 2-propanol; 100 ml min−1 N2/2-propanol mixture; 10 mg catalyst; reaction temperature, 250 °C.
c, Reaction of tert-butanol. Catalysts reduced in 0.1 bar H2 for 1 h. Carrier gas, 100 ml min−1 N2/He (1:1) mixture. d, Modulating surface OHs with annealing
and water. Annealing time, 1 h; steaming time, 1 h; water partial pressure, 0.1 bar; carrier gas, 100 ml min−1 N2/He mixture or water/N2/He mixture. e, Effect
of hydrogen partial pressure on the dehydration activity of fresh PtWOx/C-1R. Carrier gas, 100 ml min−1 N2/He mixture or H2/N2/He mixture. f, Effect of
oxidation state on the in situ formation of Brønsted acid sites by hydrogen. Pretreatment time, 1 h; reduction in 0.1 bar H2; carrier gas, 100 ml min−1 N2/He
mixture or H2/N2/He mixture. Reaction conditions for c–f: 140 °C, 4 mol% tert-butanol, 10 mg catalyst. The dehydration rate accounts for the sum
of isobutene and isobutane. The initial dehydration rates are estimated by linearly extrapolating within 1 h on stream to zero reaction time. Error bars
represent the standard deviation from triplicates.

Nature Catalysis | VOL 5 | February 2022 | 144–153 | www.nature.com/natcatal 149


Articles Nature Catalysis

a
Hydrogen ON
Periodic hydrogen pulse

High Brønsted acid density

Hydrogen OFF

b c

3.0 1.5
No H2 cofeed No H2 cofeed

Average dehydration rate (×10–4 mol s−1 gcatalyst−1)


0.1 bar H2 cofeed 0.1 bar H2 cofeed
Dehydration rate (×10–4 mol s−1 gcatalyst−1)

2.5 Oscillatory H2 cofeed Oscillatory H2 cofeed


0.1 bar H2 0.1 bar H2 0.1 bar H2 0.1 bar H2

2.0 1.0

1.5

×9
1.0
0.5

0.5

0
0 4 8 12 0
Time on stream (h) 10–14 h

Fig. 7 | Impact of hydrogen pulsing on reaction rate. a, Schematic of periodic hydrogen pulse for sustaining high Brønsted acid site density.
b, Time-on-stream data of the hydrogen pulse co-feed experiment. PtWOx/C-1R is pre-reduced at 400 °C in 0.1 bar H2 for 1 h prior to reaction. Reaction
conditions, 140 °C; 4 mol% tert-butanol; 10 mg catalyst; carrier gas, 100 ml min−1 H2/N2/He mixture (0.1 bar H2) or N2/He mixture. Each oscillation cycle
lasts for approximately 4 h. The dehydration rate is the sum of formation rates of isobutene and isobutane. c, Average dehydration rates from 10 h to 14 h.

KBr (Supplementary Fig. 17). On the other hand, reactive gas chro- Tuning activity by periodic pulsing of H2. The fundamental
matography is not affected by the opacity of the support, is Brønsted insights from pretreatment and co-feed studies are consistent with
acid site specific and is capable of measuring Brønsted acid site den- the dynamic behaviour of the WOx overlayer depicted in Fig. 2b
sity below 1 µmol H+ gcatalyst−1 (ref. 27). and Fig. 3. Water hydroxylates the oxide and generates Brønsted
Water co-feed experiments (Fig. 6d) over the annealed catalyst at acid sites, slowly oxidizes the reduced WOx overlayer at 140 °C and
140 °C indicate, counterintuitively, an increased initial dehydration retards the dehydration reaction by promoting its reverse. Hydrogen
rate, despite water product enhancing the reverse, equilibrium-limited generates Brønsted acid sites by spillover at shorter times but slowly
dehydration reaction. The lack of an induction period suggests reduces the catalyst at longer reaction times. This dynamic behav-
that water rapidly hydroxylates the WOx and regenerates OHs iour of WOx can be leveraged to increase the reaction rate by pulsing
(Supplementary Fig. 18). Increasing the steam partial pressure from hydrogen (Fig. 7a).
0.05 to 0.15 bar slightly decreases the initial dehydration activity due Figure 7b shows time-on-stream data for the reduced
to either the competitive adsorption of water on the active sites or the PtWOx/C-1R. After an initial increase in the rate when exposed to
reverse reaction. The inhibiting effect of steam is also illustrated by its hydrogen, the catalyst is deactivated in the first 2 h. When the hydro-
removal after 8 h on stream (Supplementary Fig. 18). gen is switched off after 2 h, the activity of the catalyst decreases
The in situ formation of Brønsted acid sites by co-feeding hydro- immediately but gradually increases due to re-oxidization by the
gen is illustrated in Fig. 6e. While catalyst reduction adversely affects in situ formed water. Since the oxidized WOx overlayer can sustain a
the activity, on the fresh catalyst, hydrogen generates Brønsted acid higher Brønsted acid density from hydrogen spillover, the dehydra-
sites; its partial pressure correlates positively with the dehydration tion activity substantially increases when H2 is re-introduced after
activity (Fig. 6e). On the annealed catalyst, hydrogen increases 4 h. This trend repeats itself. Periodic hydrogen pulsing allows the
the initial rate threefold (Fig. 6f). By contrast, although annealed catalyst to reach alternate states with a much higher acid site den-
WOx/C-1R shows comparable dehydration activity (Supplementary sity that cannot be sustained at steady state. The average dehydra-
Fig. 19), hydrogen exposure has no promotional effect, confirming tion rate over the last 4 h is more than 8 times higher than that with
the pivotal role of Pt in forming Brønsted acid sites by hydrogen constant co-fed hydrogen (Fig. 7c). One may expect more than one
spillover. The reduced catalyst (Fig. 6f) shows distinct dynamic order of magnitude increase in catalyst activity upon optimizing the
behaviour under different environments. In hydrogen, Brønsted frequency of periodic pulsing.
acid sites form at short times, but the catalyst is gradually reduced
at longer times and becomes inactive when the hydrogen is turned Discussion
off after 8 h (Supplementary Fig. 20). Without hydrogen, the catalyst Inverse catalysts have been demonstrated in many essential appli-
is initially less active but becomes more active with time on stream cations1–10, but their active sites and dynamics have remained elu-
due to the in situ generated water via the reverse pathway shown in sive. While catalyst reduction by H2 is typical, in many of these
Fig. 2b. These dynamic changes occur at longer timescales of hours. reactions, hydrogen and water are reactants and/or products and

150 Nature Catalysis | VOL 5 | February 2022 | 144–153 | www.nature.com/natcatal


Nature Catalysis Articles
alter the mode of action of the catalyst. Computational modelling, reactions producing high-value terminal diols and aromatics, where
in situ catalyst characterization and dehydration probe reactions acid sites have been proposed to play a crucial role. The fundamen-
reveal a complex interplay of water dissociation, hydrogen spillover tal insights from this study could enable the development of more
and catalyst reduction in controlling the Brønsted acid and redox efficient inverse catalysts. While these catalysts are far from static
site density of PtWOx/C-1R, a prototypical inverse catalyst. The and are inherently complex, simple ab initio thermodynamic and
time-scales of various elementary steps are quite different, giving microkinetic models in conjunction with probe reactions and char-
rise to complex dynamic behaviour with time-on-stream or location acterization form a blueprint for understanding and manipulating
in a chemical reactor. Brønsted acid sites on PtWOx/C-1R are more the Brønsted versus redox properties of the catalyst and the under-
active than the acid sites in microporous materials such as zeolites lying chemistry.
(Supplementary Fig. 21). Different from typical Brønsted acid cata-
lysts with localized protons near the oxygens adjacent to the metal Methods
atoms with charge imbalance, whose site density is constant, the DFT calculations. Spin-polarized DFT calculations were performed using VASP
site density here changes with reaction conditions and location in a (ref. 32). Exchange and correlation effects were approximated with the Perdew–
Burke–Ernzerhof (PBE) exchange–correlation energy functional33 and the core
reactor. Excitingly, the chemistry itself can regulate the active sites electrons were calculated using the projector augmented wavefunction. For the
and change the rate and selectivity for various pathways. WxOy cluster/Pt(111), we put a W3Ox cluster on Pt(111) slabs of 4 layers. In order
Unlike hydrogen or other reducing agents, in situ generated to eliminate periodic image interactions, we used a big Pt(111) slab (5 × 4) to
water or steaming creates Brønsted acid sites on the WOx overlayer support one trimer. For the (WO3)n monolayer/Pt (111), we used a (4 × 2) Pt(111)
slab to support the (WO3)4 2D monolayer structure (Fig. 1a). For each calculation,
by direct, modestly activated dissociation without lowering the oxi-
the bottom two layers of the Pt slab were fixed at the bulk position while the top
dation state of W, as revealed by XPS (Supplementary Fig. 22). Water two layers with the supported oxide cluster and adsorbates were fully relaxed.
can also slowly re-oxidize the catalyst, forming W–OH centres and We applied a Monkhorst–Pack mesh34. The k-points were (3 × 2 × 1) and (2 ×
thus increasing the acid density. Finally, it inhibits dehydration by 4 × 1) for the 5 × 4 slabs and the monolayer supercells, respectively. Transition
promoting its backward reaction. Although direct hydrogen disso- state searches were carried out using the climbing-image VTST NEB method35.
For geometry optimization calculations, the cut-off energy was set to 520 eV and
ciative adsorption on WOx is unfavourable below 400 °C, Pt readily the energy and force convergence criteria were set to 10−5 eV and 0.01 eV Å−1,
dissociates H2 and facilitates the rapid formation of Brønsted acid respectively. In the NEB calculations, the force criterion was set to 0.025 eV Å−1.
sites by spillover. It can also dehydrogenate a substrate and form We used pMuTT (ref. 36), together with the DFT-calculated reaction energies and
hydrogen in situ. Pt stabilizes the WOx overlayer inverse structure, vibrational frequencies, to simulate a phase diagram.
by allowing partial but not complete reduction due to self-inhibition
Simulated phase diagrams. Thermodynamic phase diagrams were generated
and enables charge transfer to WOx even at 400 °C, as shown by the using the PhaseDiagram function in pMuTT (ref. 36). DFT reaction energies and
EDX (ref. 6) and Pt L2 XANES spectra (Supplementary Fig. 12b). frequencies were used to calculate Gibbs free energies. The cluster model we used
Finally, hydrogen enables a slower reduction of WOx that increases has a formula of W3OxHy·nH2O (x = 4 to 10, y = 0 to 5, n = 0 or 1) (Supplementary
the density of redox centres and decreases the Brønsted acid Table 1). The PhaseDiagram function in pMuTT generates a graph of grids, where
site density. each grid represents the most stable state under a certain reaction condition.
State-based MKM24 was performed using the CHEMKIN package. A batch
The temperature has a dominant effect on the distribution of reactor was chosen to study the reduction kinetics of WOx/Pt using H2/H2O
active sites. The in situ XANES (Fig. 5a) and reduction studies mixed gas at different temperatures. Two sites were considered in this model: the
(Fig. 6c) show that reducing the catalyst at high temperatures low- Pt(111) site and the WOx site. The Pt site is responsible for H2 activation, while the
ers the oxidation state to W(V) and W(IV), preventing the genera- tungsten oxides can be hydrogenated and hydrated to generate hydroxyl groups.
tion of Brønsted acid sites by hydrogen spillover. High-temperature H-spillover from Pt to WOx sites can also reduce the oxide by forming H2O. Due to
the huge set of structures of all possible hydrogenated/hydrated tungsten oxides, we
annealing in an inert atmosphere facilitates surface dehydroxyl- did not calculate all activation energies in the reaction network using DFT. Instead,
ation, lowering the Brønsted acid density again without significantly we estimated them using Brønsted–Evans–Polanyi relationships developed herein
reducing the oxide. At low temperatures, hydrogen spillover and from a smaller subset of DFT calculations (Supplementary Fig. 25). To keep the
hydroxylation by water (Fig. 3d) significantly increase the number chemical potential of H2 and H2O constant during the reaction, we chose a small
of Brønsted acid sites due to low energy barriers without signifi- surface area to volume ratio of 10−20 cm−1 for the batch reactor. For the calculations
starting with fully oxidized W3O9 clusters, the reaction time for every condition in
cantly changing the oxidation state. MKM was set to 36,000 s. For the re-oxidation of W3O5 clusters, the MKM reaction
At intermediate temperatures, slow reduction of PtWOx not only times were set to 106 s. The Brønsted acid density heatmaps were generated by
causes loss of protons but also lowers the activity of WOx by cleaving calculating the OH group coverages at each reaction condition at long times. The
the W=O bond, which is essential for stabilizing and delocalizing heatmaps were smoothed by interpolation.
the charge-compensating electron28. This slow reduction is observed
XANES simulation. XANES simulation of W atoms was carried out using
using a time-dependent XANES study (Supplementary Note 2). The FDMNES (refs. 37,38), where the convoluted L1 edge spectra of all W atoms in
state of MOx can easily be tuned by changing the ratio of metal to each supercell were calculated using the muffin-tin approximation with a radius
the metal oxide21,29,30, the dispersion of M, the surface density of of 4.0 Å. To account for broadening effects (for example, core–hole lifetime), a
MOx (refs. 22,31), the distance between M and MO (Supplementary Lorentzian energy-dependent convolution was applied to the calculated spectrum.
The energy range was set to −20 to 80 eV relative to the Fermi level. The simulated
Note 2), the support reducibility, the reaction temperature and the
spectra were aligned with the experimental spectra of W standards (WO3 and
hydrogen pressure. In situ XPS characterization (Supplementary WO2) using the first inflection point.
Fig. 24) shows that reducible metal oxides, such as WOx, can easily
be re-oxidized in air to restore the Brønsted acidity. Catalyst synthesis. The 4 wt% Pt/C catalyst was prepared by the conventional
The preferred Brønsted acid site formation pathway depends incipient wetness impregnation method, as described previously5,12. First, 0.08 g Pt
on operating conditions, such as the hydrogen and water partial precursor (Pt(NH3)4(NO3)2, 99.99%, Alfa Aesar) was dissolved in an 8 g water/
pressures and the WOx coverage on Pt. As the coverage of WOx ethanol (mass ratio = 3:1) solution. Then, 0.96 g carbon black (225 m2 g−1, Vulcan
XC-72R) was added to the solution and it was kept at room temperature for 12 h.
approaches one monolayer, Brønsted acid sites could form by water After drying the catalyst in an oven at 60 °C for 12 h, the resulting powder was
dissociation only, as the hydrogen dissociation will be hindered by reduced in a 7-mm-inner diameter quartz tube with 60 ml min−1 He and 5 ml min−1
unavailable Pt sites and its high activation energy barrier on the H2 at 500 °C for 2 h using a heating ramp of 3 °C min−1.
WOx overlayer. Our study demonstrates that the Brønsted acid site The WOx overlayer was prepared by ALD using a home-built, static system, as
density can significantly be modulated by changing the pretreat- described previously5,6,39. In a typical experiment, the tungsten precursor (W(CO)6,
99%, Strem) was first evacuated and then heated to 130 °C. The W(CO)6 vapour
ment conditions (temperature, hydrogen pressure, steam pressure) was then introduced to approximately 0.3 g catalyst, which was evacuated and
and by periodic hydrogen pulsing, leading to a significant activity held at 200 °C before the deposition, for 1.5 min. Next, the catalyst chamber was
increase. The idea could in principle be extended to other biomass evacuated for 5 min to remove excess precursors. The catalyst was then exposed to

Nature Catalysis | VOL 5 | February 2022 | 144–153 | www.nature.com/natcatal 151


Articles Nature Catalysis
air for 6 min to oxidize the adsorbed W(CO)6 precursor. After completing all ALD column (Agilent, 19091P-Q04) and a flame ionization detector. Unless otherwise
cycles, the catalyst was treated twice by rapid thermal annealing (400 °C, 1 min) to noted, the catalyst was pre-treated in situ at 400 °C under a flow of H2, pure He
ensure the complete removal of ligands. Samples prepared using n ALD cycles are or 10% H2O/He mixture to reduce, anneal or steam the catalyst, respectively. The
referred to here as PtWOx/C-nR. pre-treated catalyst was then saturated at 140 °C with 10 consecutive 1 µl pulses of
10%PtWOx/C-15R was prepared by the same procedure as PtWOx/C-nR using tert-butylamine, delivered at a rate of 1 µl min−1 using an automated liquid sampler.
0.2 g Pt precursor and 0.9 g carbon black. WOx/C-1R was prepared by the same The saturated catalyst was then purged in 1,000 sccm of He to remove any weakly
procedure as PtWOx/C-nR except that the carbon black was impregnated with bound tert-butylamine. The catalyst temperature was then ramped up to 400 °C at
water/ethanol solution without the addition of Pt precursor. The carbon black 10 °C min−1 and held for 30 min, with the chromatography column held at 30 °C to
samples, used as the reference for reaction and characterization, were prepared trap desorbing butenes resulting from the Hofmann elimination of tert-butylamine
similarly by impregnation without salts, reduction and finally rapid thermal adsorbed on Brønsted acid sites. The temperature of the chromatography column
annealing. The actual Pt and W loadings were determined using inductively was then ramped to 270 °C eluting any trapped isobutene which was then
coupled plasma–atomic emission spectrometry (PerkinElmer, Optima 5300V) by quantified and taken as a direct measure of the Brønsted acid site density.
Galbraith Laboratories.
Data availability
Kinetics experiments. Kinetics experiments were performed using a fixed-bed The data to support the findings of this study are provided in the Supplementary
reactor at near ambient pressure. The reactor was a quartz tube placed in the Information and paper or from the corresponding author upon request. The atomic
middle of a resistance furnace. For a typical experiment, the catalyst bed was coordinates of the DFT-optimized WxOy/Pt(111) models and data that support the
fixed between two inert quartz beds. The catalyst bed was placed in the constant plots in this paper are available on Mendeley Data41.
temperature zone of the furnace to minimize the axial temperature gradient. The
temperature of the catalyst bed was measured using a thermocouple covered by a
quartz sleeve. Three mass flow controllers were used to feed a mixture of nitrogen, Code availability
hydrogen and helium at specified ratios. The typical flow rate of the gas mixture The code to convert ab initio data to microkinetic model inputs is available on
was 100 ml min−1 at room temperature. Mendeley Data41. No specialized, home written software was used for this work.
The alcohol flow rate was adjusted by controlling the temperature of the
bubbler and the flow rates of the carrier gas (nitrogen, 99.999% purity) and the Received: 23 March 2021; Accepted: 13 January 2022;
balance gas (helium, 99.999% purity). All the pipelines containing alcohol and Published online: 21 February 2022
water were heated to above 150 °C to avoid condensation. The reactor outlet was
connected to a gas chromatograph (Agilent, 7890A) equipped with a Plot-Q
column and a flame ionization detector. For a typical experiment, the catalyst was References
pre-reduced with a 100 ml min−1 H2–He mixture (10% H2) at 250 °C for 1 h with a 1. Védrine, J. C. Heterogeneous catalysis on metal oxides. Catalysts 7, 341 (2017).
ramping rate of 10 °C min−1. The reactor was then purged with He for 1 h before 2. Tomishige, K., Nakagawa, Y. & Tamura, M. Selective hydrogenolysis and
the reaction. hydrogenation using metal catalysts directly modified with metal oxide
species. Green Chem. 19, 2876–2924 (2017).
In situ XAS. XAS measurements were performed on beamline 5BM-D at the 3. Chia, M. et al. Selective hydrogenolysis of polyols and cyclic ethers over
Advanced Photon Source, Argonne National Laboratory. Spectra collected in bifunctional surface sites on rhodium–rhenium catalysts. J. Am. Chem. Soc.
transmission mode were recorded using ionization chambers. Samples were 133, 12675–12689 (2011).
pressed into pellets and loaded into a three-well sample holder. The sample holder 4. Daniel, O. M. et al. X-ray absorption spectroscopy of bimetallic Pt–Re
was then placed in a tubular reactor with two Kapton windows at the ends for catalysts for hydrogenolysis of glycerol to propanediols. ChemCatChem 2,
in situ reduction studies. After purging the reactor with a flowing hydrogen– 1107–1114 (2010).
helium mixture (10% hydrogen) and acquiring the spectra of the fresh samples, 5. Wang, C. et al. Mechanistic study of the direct hydrodeoxygenation of
the reactor was heated to the desired reduction temperature. The XAS spectra of m-cresol over WOx-decorated Pt/C catalysts. ACS Catal. 8, 7749–7759 (2018).
the reduced samples were collected in situ at the desired reduction temperatures 6. Wang, C. et al. A study of tetrahydrofurfuryl alcohol to 1,5-pentanediol over
under a flowing hydrogen–helium mixture after 1 h reduction. X-ray absorption Pt–WOx/C. Catal. Letters 148, 1047–1054 (2018).
near edge structure (XANES) spectra were recorded at the W L1 edge, W L3 edge 7. Yang, F. et al. Enhancement of m-cresol hydrodeoxygenation selectivity
and Pt L2 edge and extended X-ray absorption fine structure (EXAFS) spectra were on Ni catalysts by surface decoration of MoOx species. ACS Catal. 9,
recorded at the W L3 edge. The XAS spectra of PtWOx/C-1R were also collected 7791–7800 (2019).
before and after reduction at 250 °C in fluorescence mode. The measurements 8. Zhou, W. et al. Insight into the nature of Brönsted acidity of Pt-(WOx)n-H
were performed using a Linkam cell and a Vortex-ME4 silicon drift detector. The model catalysts in glycerol hydrogenolysis. J. Catal. 388, 154–163 (2020).
time-dependent XANES spectra for PtWOx/C-1R at W L1 and W L3 edges were 9. Stephens, K. J. et al. A mechanistic study of polyol hydrodeoxygenation over
recorded on the Quick X-ray Absorption and Scattering (QAS) beamline at the a bifunctional Pt-WOx/TiO2 catalyst. ACS Catal. 10, 12996–13007 (2020).
National Synchrotron Light Source (NSLS) II, Brookhaven National Laboratory 10. Lei, N. et al. Effective hydrogenolysis of glycerol to 1,3-propanediol over
(BNL). Fluorescence spectra were measured using a passivated implanted planar metal-acid concerted Pt/WOx/Al2O3 catalysts. ChemCatChem 11,
silicon (PIPS) detector. The spectra of reference foils were measured using 3903–3912 (2019).
ionization chambers. XAS data processing was performed using the Athena 11. Fu, J. et al. C–O bond activation using ultra-low loading noble metal catalysts
(0.9.26) program following standard procedures40. The first maximum of the first on moderately reducible oxides. Nat. Catal. 3, 446–453 (2020).
derivative of the raw XAS data was selected to be the edge energy. 12. Luo, J. et al. Mechanisms for high selectivity in the hydrodeoxygenation
of 5-hydroxymethylfurfural over PtCo nanocrystals. ACS Catal. 6,
4095–4104 (2016).
In situ XPS. The near-ambient pressure X-ray photoelectron spectroscopy 13. Falcone, D. D. et al. Evidence for the bifunctional nature of Pt–Re catalysts
(NAP-XPS) experiments were performed at the Center for Functional for selective glycerol hydrogenolysis. ACS Catal. 5, 5679–5695 (2015).
Nanomaterials (CFN), Brookhaven National Laboratory (BNL). Catalyst samples 14. Rasmussen, M. J. & Medlin, J. W. Role of tungsten modifiers in bimetallic
were scanned with a laboratory-based spectrometer (Phoibos NAP 150, SPECS catalysts for enhanced hydrodeoxygenation activity and selectivity. Catal. Sci.
Surface Nano Analysis). The instrument was equipped with a monochromated Technol. 10, 414–423 (2020).
Al Kα (1,486.6 eV) photon source and a differentially pumped hemispherical 15. Murugappan, K. et al. Operando NAP-XPS unveils differences in MoO3 and
analyser (SPECS Phoibos). Each powder sample was pressed to a cleaned thin Cu Mo2C during hydrodeoxygenation. Nat. Catal. 1, 960–967 (2018).
foil and transferred into the main chamber. Pt 4f, W 4f, O 1s and C 1s spectra were 16. Deo, S., Medlin, W., Nikolla, E. & Janik, M. J. Reaction paths for
collected during all procedures. Each sample was heated to 250 °C under reduced hydrodeoxygenation of furfuryl alcohol at TiO2/Pd interfaces. J. Catal. 377,
pressure for 1 h to remove adsorbed surface species. During in situ reduction and 28–40 (2019).
the following oxidation studies, hydrogen or oxygen were introduced into the main 17. Li, Z. et al. Growth of ordered ultrathin tungsten oxide films on Pt(111).
chamber through a variable leak valve. The pressure of the main chamber was J. Phys. Chem. C. 115, 5773–5783 (2011).
maintained at 0.5 mbar. XPS data analysis was performed using CasaXPS version 18. Horsley, J. A., Wachs, I. E., Brown, J. M., Via, G. H. & Hardcastle, F. D.
2.3.23. The peak model parameters are shown in Supplementary Table 4. Structure of surface tungsten oxide species in the WO3/Al2O3 supported oxide
system from X-ray absorption near-edge spectroscopy and Raman
Brønsted acid site quantification. The Brønsted acid site density of PtWOx/C-1R spectroscopy. J. Phys. Chem. 91, 4014–4020 (1987).
was measured under different gaseous environments using reactive gas 19. Yamazoe, S., Hitomi, Y., Shishido, T. & Tanaka, T. XAFS study of tungsten
chromatography27, equipped with H2 or He to simulate reducing or annealing L1- and L3-edges: structural analysis of WO3 species loaded on TiO2 as a
treatment environments, respectively. Initially, 20 mg of catalyst was sandwiched catalyst for photo-oxidation of NH3. J. Phys. Chem. C. 112, 6869–6879 (2008).
between two plugs of deactivated glass wool (Restek, 24324) and placed within a 20. Baertsch, C. D., Soled, S. L. & Iglesia, E. Isotopic and chemical titration of
quartz inlet liner (Agilent, 5062-3587). The catalyst-containing liner was placed acid sites in tungsten oxide domains supported on zirconia. J. Phys. Chem. B
in the inlet of a gas chromatograph (Agilent, 7890B) equipped with an HP-PlotQ 105, 1320–1330 (2001).

152 Nature Catalysis | VOL 5 | February 2022 | 144–153 | www.nature.com/natcatal


Nature Catalysis Articles
21. He, J. et al. Synthesis of 1,6-hexanediol from cellulose derived Acknowledgements
tetrahydrofuran-dimethanol with Pt-WOx/TiO2 catalysts. ACS Catal. 8, This work was financially supported primarily by the Catalysis Center for Energy
1427–1439 (2018). Innovation, an Energy Frontier Research Center funded by the US Department of
22. Iglesia, E. et al. Selective isomerization of alkanes on supported tungsten Energy, Office of Science, Office of Basic Energy Sciences under award number
oxide acids. Stud. Surf. Sci. Catal. 101, 533–543 (1996). DESC0001004. Portions of this work were performed at the DuPont–Northwestern–Dow
23. Fouad, N. E., Attyia, K. M. E. & Zaki, M. I. Thermogravimetry of WO3 Collaborative Access Team (DND-CAT) located at Sector 5 of the Advanced Photon
reduction in hydrogen: kinetic characterization of autocatalytic effects. Source. DND-CAT is supported by Northwestern University, E.I. DuPont de Nemours
Powder Technol. 74, 31–37 (1993). and Co. and The Dow Chemical Company. This research used resources of the Advanced
24. Alexopoulos, K., Wang, Y. & Vlachos, D. G. First-principles kinetic Photon Source, a US Department of Energy, Office of Science user facility operated
and spectroscopic insights into single-atom catalysis. ACS Catal. 9, for the Department of Energy, Office of Science by Argonne National Laboratory
5002–5010 (2019). under contract no. DE-AC02-06CH11357. The research was carried out in part at the
25. Regalbuto, J. R., Fleisch, T. H. & Wolf, E. E. An integrated study of Pt/WO3/ 7-BM (QAS) beamline of the National Synchrotron Light Source II at Brookhaven
SiO2 catalysts for the NO–CO reaction: i. catalyst characterization by XRD, National Laboratory, which is supported by the Synchrotron Catalysis Consortium,
chemisorption, and XPS. J. Catal. 107, 114–128 (1987). US Department of Energy under grant no. DE-SC0012335 and at the Center for
26. Kammert, J. D., Brezicki, G., Miyake, N., Stavitski, E. & Davis, R. J. Reduction Functional Nanomaterials, Brookhaven National Laboratory, which is supported by
of propanoic acid over Pd-promoted supported WOx catalysts. ChemCatChem the US Department of Energy, Office of Basic Energy Sciences under contract no.
12, 314–325 (2020). DE-SC0012704. Some of the XPS measurements were carried out using the Thermo
27. Abdelrahman, O. A. et al. Simple quantification of zeolite acid site Scientific K-Alpha+ XPS System at the University of Delaware surface analysis facility
density by reactive gas chromatography. Catal. Sci. Technol. 7, supported by NSF (no. 1428149). Y.L. acknowledges financial support from the Chinese
3831–3841 (2017). Academy of Science Youth Innovation Promotion Association (no. 2018220) and
28. Bernholc, J., Horsley, J. A., Murrell, L. L., Sherman, L. G. & Soled, S. Brønsted LiaoNing Revitalization Talents Program (no. XLYC1907053), and X.L. acknowledges
acid sites in transition metal oxide catalysts: modeling of structure, acid financial support from the National Natural Science Foundation of China (nos.
strengths, and support effects. J. Phys. Chem. 91, 1526–1530 (1987). 21872163, 22072090, 21991153 and 21991150) for the annular dark-field (ADF)-STEM
29. Liu, L. et al. Selective hydrogenolysis of glycerol to 1,3-propanediol over analysis. The authors acknowledge H. Matsumoto for assistance with STEM data
rhenium-oxide-modified iridium nanoparticles coating rutile titania support. acquisition from a Hitachi HF5000 microscope. We acknowledge W. Wu for assisting
ACS Catal. 9, 10913–10930 (2019). with the XPS measurements, Q. Ma, S. Ehrlich, N. S. Marinkovic and L. Ma for helping
30. Liu, S., Amada, Y., Tamura, M., Nakagawa, Y. & Tomishige, K. One-pot with the XAS measurements and S. Prodinger for valuable discussions.
selective conversion of furfural into 1,5-pentanediol over a Pd-added
Ir-ReOx/SiO2 bifunctional catalyst. Green Chem. 16, 617–626 (2014).
31. Shi, D. et al. WOx supported on ɣ-Al2O3 with different morphologies as Author contributions
model catalysts for alkanol dehydration. J. Catal. 363, 1–8 (2018). J.F. performed experimental kinetics, XAS and XPS analyses. S.L. carried out the DFT
32. Kresse, G. & Furthmüller, J. Efficient iterative schemes for ab initio calculations, the XANES simulation and the microkinetic modelling. W.Z. performed
total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, XPS and XAS studies. R.H. and C.W. synthesized the catalysts. A.L. conducted reactive
11169–11186 (1996). gas chromatography. K.A. provided guidance for XANES simulation. K.A. and S.L.
33. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation assisted with XAS measurements. S.L., Y.W. and J.A.B. conducted the XPS measurements.
made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). K.Y. performed the TPR measurements. Y.L. and X.L. carried out the ADF-STEM
34. Pack, J. D. & Monkhorst, H. J. “Special points for Brillouin-zone analysis. A.I.F. provided guidance for the XAS data analysis. O.A.A. provided guidance
integrations”—a reply. Phys. Rev. B 16, 1748–1749 (1977). for the acid site density measurements. R.J.G. provided insights into the experimental
35. Henkelman, G., Uberuaga, B. P. & Jónsson, H. A climbing image nudged work. S.C. guided the DFT calculations. D.G.V. directed the project and provided
elastic band method for finding saddle points and minimum energy paths. guidance for the experimental and theoretical work. The manuscript was written by J.F.,
J. Chem. Phys. 113, 9901–9904 (2000). S.L. and D.G.V. with input from all authors.
36. Lym, J., Wittreich, G. R. & Vlachos, D. G. A python multiscale
thermochemistry toolbox (pMuTT) for thermochemical and kinetic Competing interests
parameter estimation. Comput. Phys. Commun. 247, 106864 (2020). The authors declare no competing interests.
37. Joly, Y. X-ray absorption near-edge structure calculations beyond the
muffin-tin approximation. Phys. Rev. B 63, 125120 (2001). Additional information
38. Bunǎu, O. & Joly, Y. Self-consistent aspects of X-ray absorption calculations. Supplementary information The online version contains supplementary material
J. Phys. Condens. Matter 21, 345501 (2009). available at https://doi.org/10.1038/s41929-022-00745-y.
39. Onn, T. M. et al. Smart Pd catalyst with improved thermal stability supported
on high-surface-area LaFeO3 prepared by atomic layer deposition. J. Am. Correspondence and requests for materials should be addressed to Dionisios G. Vlachos.
Chem. Soc. 140, 4841–4848 (2018). Peer review information Nature Catalysis thanks Gianfranco Pacchioni and the other,
40. Ravel, B. & Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: data analysis anonymous, reviewer(s) for their contribution to the peer review of this work.
for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 12, Reprints and permissions information is available at www.nature.com/reprints.
537–541 (2005).
41. Fu, J. et al. Dataset for modulating the dynamics of Brønsted acid sites on Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
PtWOx inverse catalyst. Mendeley Data https://doi.org/10.17632/hp45n9ykd6.1 published maps and institutional affiliations.
(2021). © The Author(s), under exclusive licence to Springer Nature Limited 2022

Nature Catalysis | VOL 5 | February 2022 | 144–153 | www.nature.com/natcatal 153

You might also like