Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Topics in Catalysis (2023) 66:205–222

https://doi.org/10.1007/s11244-022-01753-9

ORIGINAL PAPER

Structure‑Activity Relationships of Pt‑WOx/Al2O3 Prepared


with Different W Contents and Pretreatment Conditions for Glycerol
Conversion to 1,3‑Propanediol
Napaphut Dolsiririttigul1,2 · Thanapha Numpilai3 · Chularat Wattanakit4 · Anusorn Seubsai1,2 ·
Kajornsak Faungnawakij5 · Chin Kui Cheng6 · Dai‑Viet N. Vo7 · Supinya Nijpanich8 · Narong Chanlek8 ·
Thongthai Witoon1,2 

Accepted: 17 November 2022 / Published online: 30 November 2022


© The Author(s), under exclusive licence to Springer Science+Business Media, LLC, part of Springer Nature 2022

Abstract
Turning waste glycerol into a fundamental chemical used in the petrochemical industry offers an opportunity to reduce not
only waste from biodiesel production but also the consumption of petroleum-based chemicals. In the present work, glyc-
erol was converted to 1,3-propanediol over Pt-WOx/Al2O3 catalysts, while focusing on the influence of W loading contents
(10–20 wt.%) and catalyst pretreatment conditions, specifically the calcination temperatures of ­WOx/Al2O3 (700–900 °C),
calcination temperatures of Pt-WOx/Al2O3 (350–450 °C) and reduction temperatures of Pt-WOx/Al2O3 (300–400 °C). The
characteristics of the catalysts were identified by applying ­N2-sorption, XRD, Raman, CO pulse chemisorption, ­NH3-TPD,
­H2-TPR, ­H2-TPD, SEM and TEM-EDS and XPS. The W loading contents and pretreatment conditions strongly affected
the activity and products selectivity. The W content and calcination temperature of W ­ Ox/Al2O3 at 15 wt.% and 800 °C,
respectively, were evaluated as the most suitable in providing optimum W-O-W clusters to generate ­Hδ+ which was a key
parameter for the generation of 1,3-propanediol. The calcination and reduction temperatures of Pt-WO x/Al2O3 at 400 °C
and 350 °C, respectively, were mandatory to sufficiently convert P ­ tCl2 to P
­ t3O4 and P
­ t3O4 to metallic Pt, respectively. The
optimized Pt-WOx/Al2O3 catalyst achieved a high glycerol conversion (44.7%) with high selectivity toward 1,3-propanediol
(45.1%) under a reaction temperature and pressure of 220 °C and 60 bar, respectively.

Keywords  Glycerol conversion · 1,3-Propanediol · W loading · Pretreatment conditions · Pt-WOx/Al2O3 catalyst

4
* Thanapha Numpilai Department of Chemical and Biomolecular Engineering,
thanapha@tu.ac.th Vidyasirimedhi Institute of Science and Technology,
Rayong 21210, Thailand
* Thongthai Witoon
5
fengttwi@ku.ac.th National Nanotechnology Center (NANOTEC), National
Science and Technology Development Agency (NSTDA),
1
Department of Chemical Engineering, Faculty Pathum Thani 12120, Thailand
of Engineering, Center of Excellence on Petrochemical 6
Department of Chemical Engineering, Center for Catalysis
and Materials Technology, Kasetsart University,
and Separation (CeCaS), College of Engineering, Khalifa
Bangkok 10900, Thailand
University, Abu Dhabi, United Arab Emirates
2
Center for Advanced Studies in Nanotechnology 7
Centre of Excellence for Green Energy and Environmental
for Chemical, Food and Agricultural Industries, Kasetsart
Nanomaterials (CE@GrEEN), Nguyen Tat Thanh University,
University, Bangkok 10900, Thailand
Ho Chi Minh City 755414, Vietnam
3
Department of Environmental Science, Faculty 8
Synchrotron Light Research Institute,
of Science and Technology, Thammasat University,
Nakhon Ratchasima 30000, Thailand
Pathum Thani 12120, Thailand

13
Vol.:(0123456789)

206 Topics in Catalysis (2023) 66:205–222

1 Introduction ­WOx/Al2O3 and the W loading were 20 wt.% and 900 °C,


respectively. Unfortunately, the catalytic performance
1,3-Propanediol is a building block of polytrimethylene of those catalysts for the hydrogenolysis of glycerol to
terephthalate (PTT) applied in fibers for the manufactur- 1,3-propanediol was not examined.
ing of textiles, carpets and advanced thermoplastics [1]. As mentioned above, the effect of W loading and calcina-
1,3-Propanediol is produced industrially by two different tion temperature of ­WOx/Al2O3 on the structure and catalytic
routes, consisting of acrolein hydration [2] and ethylene performance of Pt-WOx/Al2O3 catalysts is still ambiguous.
oxide hydroformylation [2], using fossil resources as the Therefore, the motivation for this research was to provide
raw material. Due to the dwindling fossil resources and a better understanding on the influence of the W loading
negative environmental impacts of their usage, it is nec- (10–20 wt.%) and calcination temperatures of ­WOx/Al2O3
essary to find an alternative more sustainable raw mate- (700–900  °C) and Pt-WOx/Al2O3 (350–450  °C) and the
rial with reduced environmental impacts [3–8]. Instead of reduction temperature of Pt-WOx/Al2O3 (300–400 °C) on the
fossil resources, bio-based raw materials are considered transformation of glycerol to 1,3-propanediol. The resulting
as a carbon-neutral resource for the sustainable produc- Pt-WOx/Al2O3 catalysts were examined using N ­ 2-sorption,
tion of chemicals. The commercially available process XRD, Raman, CO pulse chemisorption, N ­ H3-TPD, ­H2-TPR,
for the transformation of bio-based raw material to fuel is ­H2-TPD, SEM and TEM-EDS and XPS to reveal struc-
the production of biodiesel derived from the reaction of ture–activity-selectivity relationships.
triglycerides with a small molecule of alcohol [9, 10]. In
addition to biodiesel, glycerol is inevitably produced as a
by-product which accounts for approximately 10 wt.% of 2 Experimental
the total product [11–13]. Therefore, the utilization of the
by-product (glycerol) as the raw material for the synthesis 2.1 Catalyst Preparation
of 1,3-propanediol offers advantages of being economi-
cally feasible and environmentally benign. 2.1.1 Preparation of ­WOx/Al2O3 Supports
Glycerol can be dehydrated and hydrogenated over
bifunctional catalysts containing acid and metal func- WOX/Al2O3 supports containing different amounts of W
tions, respectively, to form various kinds of products, (10, 15 and 20 wt.%) were prepared using the impregna-
such as 1,3-propanediol, 1,2-propanediol and propanol [2]. tion method. In a typical synthesis, the desired amount of
1,3-Propanediol is favorable from an economic viability ammonium paratungstate hydrate (Sigma-Aldrich, ≥ 99.99%,
point of view to efficiently consume H ­ 2. Two groups of CAS:11120-25-5) was added in 6 mL of deionized water
supported catalysts (Pt-WOx-supported catalysts [14–16] with stirring at 150 rpm and 60 °C. Subsequent to the com-
and Ir-ReO x -supported catalysts [17, 18]) have been plete dissolution, the nanosized alumina support (Sigma-
found to be selective for the formation of 1,3-propanediol. Aldrich, CAS: 1344-28-1) was put into the solution and
García-Fernández et al. [19] pointed out that Brønsted acid stirred at 150 rpm and 60 °C for 4 h. The obtained product
sites and close contact of Pt and ­WOx were crucial factors was dried at 100 °C for 24 h and then calcined at different
for the selective conversion of glycerol to 1,3-propanediol. temperatures (700, 800 and 900 °C) for 2 h. The heating rate
Zhu et al. [20] varied W loading contents (5, 10, 15 and for all calcination temperatures was 5 °C ­min−1.
20 wt.% of Pt-WOx/Al 2O 3 catalysts) while the calcina-
tion temperatures of W ­ Ox/Al2O3 and Pt-WOx/Al2O3 were 2.1.2 Preparation of Pt‑WOx/Al2O3
defined at 600 °C and 400 °C, respectively. The maximum
1,3-propanediol yield of 42.4% was gained over Pt-WOx/ Pt-WOX/Al2O3 catalysts (5 wt.% Pt) were prepared using
Al2O3 with 10 wt.% W loading. Lei et al. [21] varied the the impregnation method. Briefly, the obtained ­WOx/Al2O3
composition of Pt and W ­ O x in the range 1–8 wt.% and sample from Sect. 2.1.1 was added in 6 mL of an aqueous
1–40 wt.%, respectively. All the ­WOx/Al2O3 and Pt-WOx/ solution consisting of H ­ 2PtCl6·6H2O (CAS:18497-13-7).
Al2O3 were calcined at 700 °C and 300 °C, respectively. The mixture was stirred at 150 rpm and 60 °C for 4 h and
The highest 1,3-propanediol yield was attained over Pt- dried at 100 °C for 24 h. The resulting product was calcined
WOx/Al2O3 with 6 wt.% Pt and 12.9 wt.% W loading con- at different temperatures (350 °C, 400 °C and 450 °C) for
tents. Kitano et al. [22] studied the effect of the calcination 4 h with a heating rate of 2 °C ­min−1. The obtained catalysts
temperature of W ­ Ox/Al2O3 from 500 to 1150 °C and W were designated as Pt-XW/Al2O3-Y–Z where X, Y, Z are the
loading ranging from 5 to 50 wt.% on the formation of W loading content, calcination temperature of W ­ Ox/Al2O3
Brønsted acid sites. They found that the Brønsted acid and calcination temperature of Pt-WOx/Al2O3, respectively.
sites were maximized when the calcination temperature of The Pt-WOx/Al2O3 catalysts prepared with different condi-
tions with their textural properties are listed in Table 1.

13
Topics in Catalysis (2023) 66:205–222 207

Table 1  Preparation conditions of calcined Pt-WOx/Al2O3 catalysts and their physical properties analyzed using N
­ 2-sorption technique
Catalysts Preparation conditions BET surface Pore volume W surface
area ­(m2 ­g−1) ­(cm3 ­g−1) density (W
W loading Calcination temp. of Calcination temp. of ­nm−2)
(wt.%) ­WOx/Al2O3 (°C) Pt-WOx/Al2O3 (°C)

Pt-10W/Al2O3-800-400 10 800 400 142 0.68 2.31


Pt-15W/Al2O3-800-400 15 800 400 130 0.65 3.78
Pt-20W/Al2O3-800-400 20 800 400 113 0.53 5.80
Pt-15W/Al2O3-700-400 15 700 400 136 0.61 3.61
Pt-15W/Al2O3-900-400 15 900 400 96 0.56 5.10
Pt-15W/Al2O3-800-350 15 800 350 130 0.66 3.78
Pt-15W/Al2O3-800-450 15 800 450 129 0.65 3.81

2.2 Catalyst Characterization 3 Results and Discussion

The textural properties, phase and crystallinity, reducibility, 3.1 Catalyst Characterization


interaction of catalyst surface with H
­ 2, metal compositions,
structure of W­ Ox species, amount and strength of acidic Figure 1 displays the effect of W loading contents (10, 15
sites, Pt dispersion and Pt surface area, and surface chem- and 20 wt%), calcination temperatures of W ­ Ox/Al2O3 (700,
istry of Pt species of different catalysts were characterized 800 and 900 °C) and calcination temperatures of Pt-WOx/
using ­N2-physisorption, X-ray diffraction (XRD), tempera- Al2O3 (350, 400 and 450 °C) on the N ­ 2-sorption isotherms
ture-programmed reduction of ­H2 ­(H2-TPR), temperature- and pore size distribution measured using the N ­ 2-sorption
programmed desorption of ­H2 ­(H2-TPD), transmission elec- technique. All catalysts exhibited a typical type-IV isotherm
tron microscope and scanning electron microscope equipped with an H2 hysteresis loop (Fig. 1a), indicating mesoporous
with energy dispersive X-ray spectrometer (TEM-EDS and materials with ink-bottle shaped pores. The hysteresis loop
SEM-EDS), Raman spectroscopy, temperature-programmed was quite narrow, suggesting that the sizes of the neck-like
desorption of ­NH3 ­(NH3-TPD), CO chemisorption and X-ray and wide body pores were similar. In addition, the hysteresis
photoelectron spectroscopy (XPS), respectively. The details loop occurred at a high relative pressure (P/P0~0.95) cor-
of catalyst preparation for each technique were given in sup- responding to the presence of large mesopores (29.2 nm;
plementary material. Fig. 1b). Both are useful for the fast transport of reactants
and products inside the mesopores. Loading W at 10–15
wt% (Pt-10W/Al2O3-800-400 and Pt-15W/Al2O3-800-400)
2.3 Catalytic Activity Test did not affect the pore structure but further increasing the
W loading to 20 wt% (Pt-20W/Al2O3-800-400) reduced the
The hydrogenolysis of glycerol of all Pt-WOx/Al2O3 cata- peak intensity significantly, suggesting a partial pore block-
lysts was tested in the liquid phase reaction using a 50 mL age caused by the growth of ­WOx particles. Increasing the
autoclave with an agitator system. The catalyst (0.6 g) was calcination temperature of ­WOx/Al2O3 from 800 (Pt-15W/
added into the autoclave and reduced in a pure H ­ 2 flow Al2O3-800-400) to 900 °C (Pt-15W/Al2O3-900-400) led to a
(100 mL  ­min−1) at 350 °C for 2 h. After complete reduc- reduction of pores in the range 4–20 nm, suggesting the coa-
tion, the reactor temperature was cooled to 50 °C and 30 g lescence of tiny particles to form larger particles. The results
of 10 wt.% glycerol aqueous solution was placed into the corresponded to the substantial reduction in the BET surface
autoclave. The reactor was flushed with a continuous flow area of the Pt-/20W/Al2O3-800-400 and Pt-15W/Al2O3-900-
of pure ­N2 for 5 min followed by pure ­H2 for 1 min. Subse- 400 catalysts (Table 1). Increasing the calcination tempera-
quently, the reactor was pressurized to 60 bar with pure ­H2 ture (350, 400 and 450 °C) of the Pt-WOx/Al2O3 catalysts
and the reactor was heated to 220 °C with stirring at 600 rpm slightly altered pore structures with an almost constant BET
and kept at that temperature for 5 h. After the reaction, the surface area (Table 1). The W surface density, defined as the
suspension was centrifuged at 8000 rpm for 15 min to sepa- number of W atoms per surface area (atom W ­nm−2), was
rate the liquid phase and the solid catalyst. The products calculated; the results are shown in Table 1. The W surface
were analyzed according to the procedures described in our density increased from 2.31 to 5.80 W ­nm−2 with increasing
previous work [14] and turn over frequency (TOF) was cal- the W loading content from 10 to 20 wt% and from 3.61 to
culated based on the Equation reported by Jarauta-Córdoba 5.10 W ­nm−2 with increasing the calcination temperature of
et al. [23]. ­WOx/Al2O3 from 700 to 900 °C. The W surface density of

13

208 Topics in Catalysis (2023) 66:205–222

Fig. 1  N2-sorption isotherms of Pt-WOx/Al2O3 prepared with different conditions (a) and their corresponding pore size distribution (b)

the Pt-WOx/Al2O3 catalysts calcined at different tempera- strong metal-support interaction. However, the significant
tures (Pt-15W/Al2O3-800-350, Pt-15W/Al2O3-800-400 and reduction of BET surface area of the Pt-20W/Al2O3-800-
Pt-15W/Al2O3-800-450) remained unchanged because the 400 sample compared to the Pt-10 W/Al2O3-800-400 sample
calcination temperatures of the catalysts were much lower (Table 1) indicated that there was a strong Pt-W interac-
than the calcination temperature of W­ Ox/Al2O3. Notably, the tion and thus the diffraction line of P­ t3O4 was not clearly
W surface density (5.8 atom W n­ m−2) at 20 wt% W loading visible. The diffraction lines of the ­WO3 crystalline phase
(Pt-/20W/Al2O3-800-400) was higher than for the monolayer (PDF 01-075-2187) were visible for the Pt-20W/Al2O3-800-
of W surface density (5.1 atom W n­ m−2) [24] which could 400 sample, indicating the formation of bulk W ­ O3 crystals,
lead to the generation of bulk ­WO3 crystals. which was in line with its W surface density value (Table 1).
The XRD patterns of the calcined different Pt-WOx/Al2O3 The XRD pattern of Pt-15W/Al2O3-700-400 was similar
catalysts under air atmosphere are shown in Fig. 2. The XRD to that of Pt-15 W/Al2O3-800-400, suggesting that crystal
patterns of all catalysts displayed diffraction lines at 2θ val- structure of metal oxides did not change with an increase
ues of 32.80°, 37.64°, 46.60°, 67.25° and 45.85°, 66.85° in the calcination temperature of W ­ Ox/Al2O3 from 700 °C
which were matched to δ-Al2O3 (PDF 00-046-1131) and (Pt-15W/Al2O3-700-400) to 800 °C (Pt-15W/Al2O3-800-
γ-Al2O3 (PDF 00-056-0457), respectively. Small diffrac- 400). The XRD pattern of Pt-15W/Al2O3-900-400 remained
tion lines at 35.92° and 39.49° indexed to the P ­ t3O4 phase unaltered compared to that of Pt-15W/Al2O3-800-400 and
(PDF 00-021-1284) were visible for all catalysts. However, Pt-15W/Al2O3-700-400, even though the calcination tem-
the peak intensities were quite different depending on the perature increased to 900 °C and its BET surface area was
synthesis conditions. At a low W loading content (10 wt.%, 26.1% lower than that of the 15W/Al2O3-800-400 sample.
Pt-10W/Al2O3-800-400), the diffraction line of P ­ t3O4 at This affirmed that the presence of a strong Pt-W interaction
39.49° was clearly observed. Increasing the W loading to promoted the dispersion of Pt oxides species. The XRD pat-
15 (Pt-15W/Al2O3-800-400) and 20 (Pt-20W/Al2O3-800- tern of Pt-15W/Al2O3-800-350 exhibited a diffraction line
400) wt.% led to a significant decrease in the ­Pt3O4 peak of ­PtCl2, suggesting that a calcination temperature of Pt-
intensity, implying a reduction of P­ t3O4 crystallite size or a WOx/Al2O3 at 350 °C was insufficient to decompose ­PtCl2

13
Topics in Catalysis (2023) 66:205–222 209

TEM-EDS analysis. The higher Cl/Pt mass ratios of Pt-15W/


Al2O3-800-350 was likely associated with the remaining
small ­PtCl2 crystals as observed by XRD analysis (Fig. 2).
Figure 5a shows the Raman spectra of A ­ l2O3 calcined
at 800 °C ­(Al2O3-800), Pt-Al2O3 calcined at 400 °C (Pt-
Al2O3-800-400), W/Al2O3 with 15 wt% (15W/Al2O3-800)
and 20 wt% (20W/Al2O3-800) tungsten loading content. No
significant band was observed for the A ­ l2O3-800. The Pt-
Al2O3-800-400 exhibited a strong band at 590 ­cm−1 and a
small band at 335 ­cm−1 which were reported to be a charac-
teristic of highly dispersed P ­ tOx on A­ l2O3 [25]. Loading 15
wt.% tungsten on A ­ l2O3 (15W/Al2O3-800) showed a strong
band at 989 ­cm−1 attributable to ʋ(W = O) vibrations of pol-
ytungstate species [23, 26]. Additional three small bands at
271, 690 and 809 ­cm−1 assigned to crystal structure of W ­ O3
[23, 26] were visible when the tungsten loading content was
raised to 20 wt% (20W/Al2O3-800), suggesting an initial
stage in the crystalline ­WO3 formation.
After impregnation of Pt onto ­WOx/Al2O3 (Fig. 5b–d),
the Raman spectra showed a combination of bands of both
polytungstate species and amorphous P ­ tOx but the bands
intensity were different depending on the synthesis condi-
tions. As shown in Fig. 5b, the band at 974 ­cm−1 increased
with increasing W loading from 10 to 20 wt%. However,
the bands representative for crystalline W ­ O 3 were not
clearly observed as those of 20W/Al2O3-800, possibly due
to a strong interaction of Pt with W species [27]. Similarly,
the band intensity of polytungstate species increased with
increasing calcination temperature of ­WOx/Al2O3 (Fig. 5c)
from 700 to 900 °C, suggesting an increase of surface den-
sity of polytungstate. The different calcination temperatures
(350 °C, 400 °C, and 450 °C) of Pt-WOx/Al2O3 catalysts also
Fig. 2  XRD patterns of calcined Pt-WOx/Al2O3 catalysts prepared affected the surface structure of catalysts. Only small band
with different conditions at 578 ­cm−1 was visible for the catalyst calcined at 350 °C
(Pt-15W/Al2O3-800-350), which might be attributed to the
incomplete decomposition of H ­ 2PtCl6. At 450 °C (Pt-15W/
crystal. The diffraction line of ­PtCl2 was not observed for Al2O3-800-450), the band of ­PtOx became more predomi-
15W/Al2O3-800-400 sample. Further increasing calcination nant than the band of polytungstate, suggesting that amor-
temperature to 450 °C (Pt-15W/Al2O3-800-450) tended to phous ­PtOx aggregated to form a larger size covering the
enlarge the ­Pt3O4 crystallite sizes. TEM-EDS analysis was surface of polytungstate.
used to examine the dispersion of Pt and the presence of The reduction behavior of the different Pt-WOx/Al2O3
Cl in the catalysts calcined at 350 °C (Pt-15W/Al2O3-800- catalysts as well as Pt/Al2O3 is presented in Fig. 6. The
350) and 400 °C (Pt-15W/Al2O3-800-400); the results are ­H2-TPR profile of Pt/Al2O3 (in the absence of ­WOx) exhib-
shown in Figs. 3 and 4, respectively. It was found that Pt ited a main H ­ 2 consumption peak at 150 °C and a broad peak
was highly dispersed throughout the surface of both cata- at 595 °C that could be attributed to a reduction of ­Pt3O4
lysts and the Cl/Pt mass ratio of Pt-15W/Al2O3-800-350 was to metallic Pt and a reduction of P ­ tOx strongly interacted
found to be 0.161 which was 1.45 times higher than that with ­Al2O3, respectively. The reduction behavior of ­Pt3O4
of Pt-15W/Al2O3-800-400. The metal compositions of all was noticeably changed when W ­ Ox was present in the Pt-
catalysts were further investigated with SEM-EDS analysis. WOx/Al2O3 system. All Pt-WOx/Al2O3 catalysts displayed
As shown in Table 2, the metal compositions of all catalysts two main steps of H ­ 2 consumption (at 75 °C and 150 °C)
were closed to their expected ratios. In addition, the Cl/Pt that could be assigned to a reduction of ­Pt3O4 dispersed on
mass ratio of Pt-WOx/Al2O3 at 350 °C was also higher than ­WOx and a reduction of P ­ t3O4 on A­ l2O3, respectively. The
that of the other catalysts which was in good agreement with ­H2 consumption at temperatures > 200 °C could be attributed

13

210 Topics in Catalysis (2023) 66:205–222

Fig. 3  TEM-EDS mapping of Pt-15 W/Al2O3-800-350

Fig. 4  TEM-EDS mapping of Pt-15 W/Al2O3-800-400

13
Topics in Catalysis (2023) 66:205–222 211

Table 2  Metal compositions Samples Metal compositions (wt%) Cl/Pt ratio


measured by SEM–EDS and
TEM-EDS analysis of all Pt W Al O Cl Catalystsa H2PtCl6·6H2O
calcined catalysts
Pt-10W/Al2O3-800-400 4.20 8.46 34.58 52.35 0.41 0.10 1.09
Pt-15W/Al2O3-800-400 4.54 13.48 36.15 45.33 0.50 0.11
(0.11)
Pt-20W/Al2O3-800-400 5.45 16.78 33.36 43.79 0.62 0.11
Pt-15W/Al2O3-700-400 4.74 13.15 38.04 43.53 0.54 0.11
Pt-15W/Al2O3-900-400 4.62 13.65 36.14 45.08 0.51 0.11
Pt-15W/Al2O3-800-350 4.06 13.51 32.77 48.79 0.87 0.21
(0.16)
Pt-15W/Al2O3-800-450 6.16 12.91 31.53 48.78 0.62 0.10
a
 The values in parenthesis were obtained with TEM-EDS analysis

to a reduction of ­WOx species and that is the reason why the the Pt-20W/Al2O3-800-400 sample but their peaks were not
­H2-TPR profiles of the Pt-WOx/Al2O3 catalysts do not return clearly seen due to a partial reduction of bulk ­WO3 promoted
to the baseline similar to the Pt/Al2O3 catalyst. This phenom- by metallic Pt. Similar to the increase in the W loading con-
enon was also observed from the other groups using Pt/WOx tent, the metallic Pt peak intensity notably decreased when
based catalysts [27, 28]. The reduction of W ­ Ox species can the calcination temperature increased from 700 to 900 °C,
be easier when the Pt was present because the H ­ 2 could dis- suggesting a strong Pt-W interaction when the ­WOx/Al2O3
sociate over the Pt surface and then spillover to reduce W ­ Ox sample was calcined at > 700 °C. Some P ­ tCl2 species still
species. Compared to the Pt-10W/Al2O3-800-400 sample existed on the surface of the Pt-WOx/Al2O3 catalyst pre-
(Fig. 6a), the Pt-15 W/Al2O3-800-400 and Pt-/20W/Al2O3- calcined at 350 °C followed by reduction at 350 °C (Pt-15W/
800-400 samples had a smaller ­H2 consumption at 150 °C Al2O3-800-350). The XRD technique could not distinguish
but a larger ­H2 consumption at > 200 °C, indicating that a the crystal structure when the calcination temperature of Pt-
larger fraction of ­PtOx of Pt-15W/Al2O3-800-400 and Pt- WOx/Al2O3 increased from 400 °C (Pt-15W/Al2O3-800-400)
/20W/Al2O3-800-400 was in close proximity to the surface to 450 °C (Pt-15W/Al2O3-800-450). In addition, the effect
of ­WOx compared to Pt-10W/Al2O3-800-400 sample. The of reduced temperature was studied by reducing the Pt-15W/
­H2 consumption over W ­ Ox/Al2O3 calcined at 700 °C (Pt- Al2O3-800-400 sample at 300 °C (Pt-15W/Al2O3-800-400
15W/Al2O3-700-400) was lower than that of ­WOx/Al2O3 reduced at 300 °C) and 400 °C (Pt-15W/Al 2O3-800-400
calcined at 800 °C (Pt-15W/Al2O3-800-400) and 900 °C reduced at 400 °C) for 2 h. Clearly, increasing the reduction
(Pt-15W/Al2O3-900-400), indicating that increasing the cal- temperature to 400 °C caused mobility of Pt particles to
cination temperature allowed the W ­ Ox surface to be more form a larger Pt particle size.
reducible (Fig. 6b). When the synthesized Pt-WOx/Al2O3 The temperature programmed desorption (TPD) of ­H2
was pre-calcined at 350 °C (Pt-15W/Al2O3-800-350), the was conducted to gain further insight into the structure of
reduction behavior at 0–150 °C did not resemble that of the Pt-WOx/Al2O3 catalysts prepared at different conditions. The
synthesized catalysts pre-calcined at 400 °C (Pt-15W/Al2O3- ­H2-TPD profiles of all Pt-WOx/Al2O3 catalysts exhibited two
800-400) which could be attributed to the reduction of P ­ tCl2. main desorption peaks. The low-temperature desorption fea-
A slight decrease in the ­H2 consumption was observed when ture (50–200 °C) was attributed to the desorption of atomic
the calcination temperature of Pt-WOx/Al2O3 increased to H over the metallic Pt surface and oxygen vacancies of at
450 °C (Pt-15W/Al2O3-800-450) which was associated with the Pt-supports ­(WOx and ­Al2O3) interface [29, 30], while
a sintering of Pt particles (Fig. 6c). the high-temperature desorption feature (250–600 °C) could
Figure 7 shows the XRD patterns of the Pt-WOx/Al2O3 be attributed to the spillover of H atom to the surface of
catalysts reduced with H ­ 2 at 350 °C for 2 h. The XRD pat- ­Al2O3 and partially reduced ­WOx [30, 31]. The area under
tern of Pt-10W/Al2O3-800-400 displayed a prominent dif- the low-temperature peak of the Pt-10W/Al2O3-800-400
fraction line of metallic Pt, suggesting the presence of iso- sample (10 wt.% W loading) was the lowest in the series of
lated Pt particles with a low interfacial contact with W ­ Ox effects of the W loading content (Fig. 8a), possibly due to
species. The peak intensity of metallic Pt became smaller as the larger Pt crystallite size and lower interfacial contact of
the W loading content increased from 10 wt% to 15 wt% (Pt- Pt and supports. The area under the low-temperature peak of
15W/Al2O3-800-400) and 20 wt% (Pt-20W/Al2O3-800-400), Pt-20W/Al2O3-800-400 sample was slightly lower than that
confirming a strong affinity of the Pt and W species. The of the Pt-15W/Al2O3-800-400 sample due to the formation
diffraction lines of ­WO3 still remained in the XRD pattern of of crystalline ­WO3, resulting in the significant drop in the

13

212 Topics in Catalysis (2023) 66:205–222

Fig. 5  Raman spectra of ­ Al2O3 calcined at 8­00oC ­(Al2O3-800), Al2O3-800-400 (B), Pt-20W/Al2O3-800-400 (C), Pt-15W/Al2O3-700-
Pt-Al2O3 calcined at 400 oC (Pt-Al2O3-800-400),  ­WOx loaded on 400 (D), Pt-15W/Al2O3-900-400 (E), Pt-15W/Al2O3-800-350 (F) and
­Al2O3 with 15 wt% W loading (15W/Al2O3-800) and 20 wt% W load- Pt-15W/Al2O3-800-450 (G).  The meaning of the catalysts' relative
ing (20W/Al2O3-800)  (a) and Pt-WOx/Al2O3  catalysts prepared with notation is given in Table 1
different conditions including Pt-10W/Al2O3-800-400 (A), Pt-15W/

BET surface area which could lead to a lower dispersion of of ­WOx/Al2O3. At 700 °C (Pt-15W/Al2O3-700-400), the
Pt compared to that of the Pt-15W/Al2O3-800-400 sample. areas under both the low- and high-temperature features
Figure 8b shows the H­ 2-TPD profiles of the Pt-WOx/Al2O3 were substantially lower than those of Pt-15W/Al2O3-800-
catalysts prepared with different calcination temperatures 400, suggesting the low Pt-support interfaces as well as the

13
Topics in Catalysis (2023) 66:205–222 213

Fig. 6  H2-TPR profiles of different Pt-WOx/Al2O3 catalysts and Pt/Al2O3: Pt-WOx/Al2O3 catalysts prepared with different W loading contents
(a), with different calcination temperatures of W
­ Ox/Al2O3 (b) and with different calcination temperatures of Pt-WOx/Al2O3 (c)

less reducible W­ Ox species (Fig. 8b). At 900 °C (Pt-15W/


Al2O3-900-400), the area under high-temperature increased,
suggesting an increase in the reducible ­WOx species.
The calcination temperature of the Pt-WOx/Al2O3 cata-
lysts also affected the amount of H ­ 2 absorbed. As shown
in Fig. 8c, at 350 °C (Pt-15W/Al2O3-800-350), both des-
orption peaks were lower than for the Pt-15W/Al2O3-800-
400 catalyst possibly due to the loss of Pt active surface. At
450 °C (Pt-15W/Al2O3-800-450), the low-temperature peak
substantially dropped compared to that of Pt-15W/Al2O3-
800-400, suggesting sintering of the Pt particles. The effect
of the reduction temperature of the Pt-15W/Al2O3-800-400
catalyst is presented in Fig. 8d. At 300 °C, the area under the
low-temperature feature was the smallest in the series which
could be attributed to the remaining ­PtOx species. Increasing
the reduction temperature to 400 °C caused a slight decrease
in the area under the low-temperature feature, probably due
to mobility of the Pt particles on the catalyst surface to form
a larger particle at such a reduction temperature. Clearly, the
high-temperature feature peak shifted toward a higher des-
orption temperature as the reduction temperature increased,
suggesting a strengthened interaction of H with the W ­ Ox
surface.
The CO pulse chemisorption was used to determine
the Pt surface area and Pt dispersion of different Pt-WO x/
Al2O3 catalysts. As can be seen in Table 3, the Pt surface
area and dispersion of all catalysts were greatly differ-
ent despite they have the same amount of Pt loading (5
wt%). At W loading of 10 wt% (Pt-10W/Al2O3-800-400),
Fig. 7  XRD patterns of reduced Pt-WOx/Al2O3 catalysts prepared the Pt dispersion had the highest value of 52.6%. The Pt
with different conditions dispersion was found to be 31.9% and 16.2% when the W

13

214 Topics in Catalysis (2023) 66:205–222

Fig. 8  H2-TPD profiles of dif-


ferent Pt-WOx/Al2O3 catalysts:
effect of W loading contents (a)
effect of calcination tempera-
tures of ­WOx/Al2O3 (b), effect
of calcination temperatures of
Pt-WOx/Al2O3 (c) and effect of
reduction temperatures of Pt-
WOx/Al2O3 (d)

Table 3  Acidity of pre-reduced Pt-Al2O3 and Pt-WOx/Al2O3 catalysts determined by N


­ H3-TPD measurement and Pt surface area and its disper-
sion determined by CO pulse chemisorption
Samples Amount of acid sites (mmol ­g−1) Pt surface area Pt dis-
­(m2 ­g−1) persion
Weak I (τ) Weak II (β) Medium (α) Strong (γ) Total acid sites (%)

Pt-Al2O3 – 0.157 0.031 – 0.188 – –


Pt-10W/Al2O3-800-400 0.024 0.180 0.136 0.090 0.430 6.49 52.6
Pt-15W/Al2O3-800-400 0.018 0.179 0.182 0.132 0.511 3.94 31.9
Pt-20W/Al2O3-800-400 0.028 0.143 0.189 0.110 0.470 2.01 16.2
Pt-15W/Al2O3-700-400 0.021 0.163 0.173 0.092 0.449 4.29 34.7
Pt-15W/Al2O3-900-400 0.027 0.185 0.218 0.150 0.580 2.86 23.1
Pt-15W/Al2O3-800-350 0.022 0.141 0.158 0.102 0.423 1.89 15.3
Pt-15W/Al2O3-800-450 0.023 0.132 0.168 0.088 0.411 3.02 24.5

13
Topics in Catalysis (2023) 66:205–222 215

loading content increased to 15 (Pt-15W/Al2O3-800-400) increased with increasing tungsten loading content and
and 20 (Pt-20W/Al 2O3-800-400) wt%, respectively. The calcination temperature of ­WOx/Al2O3.
Pt dispersion decreased from 34.7 to 23.1% with rais-
ing calcination temperature of W ­ O x/Al 2O 3 from 700 to 3.2 Catalyst Activity Test
900 °C. Based on the results of XRD of reduced catalysts
(Fig. 7), and CO pulse chemisorption, the CO molecules The catalytic performance of the Pt-WOx/Al2O3 catalysts
selectively adsorbed on the isolated Pt particles, while the with different conditions for hydrogenolysis of glycerol is
interface of Pt-supports were inactive on CO adsorption. displayed in Figs. 10–13. As shown in Fig. 10a, Pt/Al2O3
The Pt dispersion over the Pt-WO x/Al 2O 3 catalyst cal- catalyst in the absence of W gave a low glycerol conversion
cined at 350 °C was found to be the lowest value (15.3%). of 14.0% and the glycerol was mostly converted to 1,2-pro-
The Pt dispersion increased to 31.9% when the calcination panediol (Fig. 10b). The Pt-WOx/Al2O3 catalyst with 10
temperature of the Pt-WO x/Al 2O 3 catalyst increased to wt.% W loading (Pt-10W/Al2O3-800-400) provided a higher
400 °C (Pt-15W/Al2O3-800-400). Further increase of cal- glycerol conversion of 36.5% with a significant improvement
cination temperature to 450 °C (Pt-15W/Al2O3-800-450) of 1,3-propanediol selectivity from 3.5 to 36.8%, suggest-
caused a significant drop in the Pt dispersion to 24.5%. ing that the presence of medium acid sites (Fig. 9) after the
The acid strength and number of acidity of the Pt- tungsten addition altered the product selectivity. Increasing
Al 2O 3 and different Pt-WO x/Al 2O 3 catalysts were eval- the W loading content to 15 wt.% (Pt-15W/Al2O3-800-400)
uated by N ­ H 3-TPD technique. As shown in Fig. 9, the notably improved the glycerol conversion to 44.7% as well as
­N H 3-TPD profile of Pt-Al 2O 3 catalyst exhibited a main 1,3-propanediol selectivity (45.1%). The propanol selectivity
desorption peak at 236 °C with a shoulder at 333 °C, sug- also increased to 26.8% with increasing the W loading con-
gesting that the Pt-Al2O3 catalyst mainly contained weak tent to 15 wt.%. A further increase of the W loading content
acidity corresponding to the presence of A ­ l + as Lewis to 20 wt.% (Pt-20W/Al2O3-800-400) led to a slight reduction
acid sites. The presence of tungsten on Pt-Al2O3 catalysts in both glycerol conversion (Fig. 10a) and 1,3-propanediol
remarkably enhanced the medium and strong acid sites. selectivity (Fig. 10b).
The number of acidities for each acid site was summa- In addition to the medium acid sites, previous works
rized in Table 3. It was found that the medium acid sites have shown that the type of acid site (Lewis and Brønsted)
determined the product’s selectivity toward 1,2-propanediol

Fig. 9  NH3-TPD profiles of
Pt-Al2O3 and different Pt-WOx/
Al2O3 catalysts

13

216 Topics in Catalysis (2023) 66:205–222

Fig. 10  Effect of W loading contents on conversion and yield (a), product selectivity (b) and proposed structures of Pt-WOx/Al2O3 catalysts (c)

and 1,3-propanediol, respectively [2]. Aihara et al. [32] The effect of the calcination temperature of ­WOx/Al2O3
found that no 1,3-propanediol was formed over a Pt/Al2O3 on the catalytic performance for the hydrogenolysis of glyc-
catalyst as it contained mainly Lewis acid sites. Increasing erol is presented in Fig. 11. The Pt-15W/Al2O3-700-400
the Brønsted-to-Lewis acid ratio shifted the selectivity of sample gave the lowest glycerol conversion of 37.2% and the
products toward 1,3-propanediol [21]. Based on the catalyst highest 1,2-propanediol selectivity of 32.5% in the series.
characterization and previous reports, the low 1,3-propan- Increasing the calcination temperature to 800 °C (Pt-15W/
ediol selectivity over the Pt-10W/Al2O3-800-400 catalyst Al2O3-800-400) increased the glycerol conversion to 44.7%
(10 wt.% W loading) in the current study was attributed as well as producing a substantial reduction in the 1,2-pro-
to the low density of W­ Ox groups forming slight bridging panediol selectivity to 27.1%, accompanied by an increase
W-O-W bonds, yielding a small fraction of Brønsted acid in 1,3-propanediol selectivity to 45.1%. At 900 °C (Pt-15W/
sites. Increasing the W loading content to 15 wt.% (Pt-15W/ Al2O3-900-400), both the conversion of glycerol (Fig. 11a)
Al2O3-800-400) created more bridging W–O-W bonds that and 1,3 propanediol selectivity (Fig. 11b) dropped slightly.
could delocalize electrons entirely partially reduced ­WOx The XRD and ­H2-TPD analyses (Figs. 7 and 8b) of the
domains [33, 34]. The delocalized electrons over the par- reduced Pt-15W/Al2O3-700-400 sample indicated the weak-
tially reduced ­WOx domains were essential to stabilize ­Hδ+ ened interaction of the Pt and ­WOx species and low hydro-
acting as Brønsted acid sites (Fig. 10c). A further increase gen spillover ability which resulted in more difficulty reduc-
in the W loading content to 20 wt.% (Pt-20W/Al2O3-800- ibility of the W ­ δ+. In addition, the
­ Ox surface to generate H
400) led to the formation of an inactive W
­ O3 phase, resulting high Pt dispersion (34.7%) but low interfacial contact of
in a slight drop in 1,3-propanediol and propanol selectivi- ­WOx species with Pt implied that part of Pt was dispersed
ties. Note that further catalysts analysis with in situ FTIR on ­Al2O3 support which provided a low glycerol conversion
technique with pyridine should be carried out to affirm the as well as high 1,2-Propanediol selectivity. At 900 °C (Pt-
Brønsted-to-Lewis acid ratio [23]. 15W/Al2O3-900-400), more W-O-W bonds were linked to
each other to form a large ­WOx domain (Fig. 11c), favoring

13
Topics in Catalysis (2023) 66:205–222 217

Fig. 11  Effect of calcination temperatures of ­WOx/Al2O3 on conversion and yield (a), product selectivity (b) and proposed structures of Pt-WOx/
Al2O3 catalysts (c)

the formation of Brønsted acid sites and promoting the Therefore, the low activity and selectivity of 1,3-propan-
production of 1,3-propanediol. However, the selectivity of ediol were mainly attributed to the phase of the Pt species
1,3-propanediol of the Pt-15W/Al2O3-900-400 catalyst was (Fig. 12c). Based on the XRD analysis of the reduced Pt-
slightly lower than for the Pt-15W/Al2O3-800-400 catalyst 15W/Al2O3-800-350 sample (Fig. 7), ­PtCl2 species still
but enhancement of propanol selectivity was observed that remained in the structure of the catalysts. Although P ­ tCl2
was attributed to over-hydrogenation of 1,3-propanediol to has been reported to be active for methane oxidation to
propanol due to more available H over the ­WOx surface, as methyl bisulfate [35, 36], this phase could have been inac-
confirmed by the ­H2-TPD analysis (Fig. 8b). tive for the dissociation of ­H2, retarding the formation of
Figure 12 shows the effect of the calcination tempera- ­Hδ+ and the hydrogenation ability of the catalyst. Increasing
ture of Pt-WOx/Al2O3 while maintaining the calcination the calcination temperature of the Pt-WOx/Al2O3 catalyst to
temperature of W­ Ox/Al2O3 at 800 °C. Clearly, the conver- 450 °C (Pt-15W/Al2O3-800-450) led to a slight decrease in
sion of glycerol over the Pt-WOx/Al2O3 calcined at 350 °C glycerol conversion and 1,3-propanediol and propanol selec-
(Pt-15W/Al2O3-800-350) was only 30.9%, which was much tivities to 43.7%, 41.2% and 26.1%, respectively, compared
lower than for the Pt-15W/Al2O3-800-400 sample (44.7%). to 44.7%, 45.1% and 26.8%, respectively, of the Pt-15W/
In addition, the product selectivity levels of both catalysts Al2O3-800-400 sample which were likely associated with a
were quite different, as the Pt-15W/Al2O3-800-350 sample partial sintering of Pt particles (Fig. 12c).
predominantly gave 44.3% 1,2-propanediol selectivity, while Figure 13 shows the effect of reduction temperatures
the highest product selectivity over the Pt-15W/Al2O3-800- of Pt-WOx/Al2O3 catalysts calcined at 400 °C. The low-
400 sample was altered to 1,3-propanediol (45.1%). In fact, est glycerol conversion and 1,3-propanediol and propanol
the calcination temperature of ­WOx/Al2O3 of both catalysts selectivities of 29.1%, 31.2% and 16.9%, respectively, were
was identical, providing similar ­WOx coverage on ­Al2O3. achieved over the Pt-15W/Al2O3-800-400 catalyst reduced

13

218 Topics in Catalysis (2023) 66:205–222

Fig. 12  Effect of calcination temperatures of Pt-WOx/Al2O3 on conversion and yield (a), product selectivity (b) and proposed structures of Pt-
WOx/Al2O3 catalysts (c)

at 300 °C. These values were notably lower than the Pt-15W/ 333.7 eV, belonging to Pt ­4d5/2 and Pt ­4d3/2, respectively,
Al2O3-800-400 catalyst reduced at 350 °C even though the that were then shifted toward lower binding energies at 315.5
calcination temperatures of both samples during the prepa- and 332.1 eV, respectively, and 315.3 and 331.9 eV, respec-
ration of W­ Ox/Al2O3 and Pt-WOx/Al2O3 were identical, tively, when the Pt-15W/Al2O3-800-400 catalyst reduced
suggesting similar W surface densities. Therefore, the low at 300 and 350 °C, respectively. The position of the Pt 4d
glycerol conversion and 1,3-propanediol and propanol selec- XPS spectra changed slightly with a further increase in the
tivities over the Pt-15W/Al2O3-800-400 catalyst reduced at reduction temperature to 400 °C. The results suggested that
300 °C were likely associated with an incomplete reduction the reduction of ­Pt3O4 was incomplete at 300 °C, which was
of ­PtOx to metallic Pt (Fig. 13c), lowering the capability for attributed to the low activity and selectivity of the catalyst
­H2 dissociation and the number of ­Hδ+ on the ­WOx domain, (Fig. 13).
resulting in a decrease in active sites for the generation of Figure 15a shows the T­ OF1,3-PDO-W as a function of Pt dis-
1,3-propanediol as well as propanol. Further increasing persion and W surface density. A linear correlation between
the reduction temperature to 400 °C resulted in a slightly ­TOF1,3-PDO-W and Pt dispersion was observed except that of
lower glycerol conversion and 1,3-propanediol selectivity Pt-10W/Al2O3-800-400 sample and Pt-15W/Al2O3-700–400
compared to those of the Pt-15W/Al2O3-800-400 sample, sample due to the lower amount of tungsten and different
probably due to a slight growth of Pt particles and the too tungsten species, respectively. The latter could be related
strong interaction of H with the W
­ Ox surface, as indicated by to the significant reduction of H atom adsorbed on partially
the ­H2-TPD analysis (Fig. 8d). XPS analysis was performed reduced ­WOx as observed by H ­ 2-TPD analysis (Fig. 8b).
to investigate the surface Pt species after reduction with H­ 2 These findings were analogous with the recent study
at different temperatures (Fig. 14). The calcined Pt-15W/ reported by Jarauta-Córdoba et al. [23]. The relationship
Al2O3-800-400 catalyst exhibited two peaks at 316.2 and between ­TOF1,3-PDO-W and W surface density did not clearly

13
Topics in Catalysis (2023) 66:205–222 219

Fig. 13  Effect of reduction temperatures of Pt-WOx/Al2O3 on conversion and yield (a), product selectivity (b) and proposed structures of Pt-
WOx/Al2O3 catalysts (c)

emanate in this study (Fig. 15b). For example, the Pt-WOx/ 4 Conclusions


Al2O3 catalysts calcined at 350 °C to 450 °C had almost the
same W surface density but the ­TOF1,3-PDO-W differed which The present study revealed the structure-activity relation-
was likely associated with Pt species and Pt dispersion. Nev- ships of Pt-WOx/Al2O3 catalysts prepared with different W
ertheless, it could be seen that the maximum ­TOF1,3-PDO-W loading contents and pretreatment conditions (calcination
was found at W surface density of 3.78 W ­nm−2 which was temperatures of ­WOx/Al2O3, calcination temperature of Pt-
lower than a theoretical monolayer coverage of ­WOx spe- WOx/Al2O3 and reduction temperature of Pt-WOx/Al2O3) for
cies (5.1 W ­nm−2). The similar result was also observed by the hydrogenolysis of glycerol to 1,3-propanediol. There was
Jarauta-Córdoba et al. [23] and it was suggested that glyc- a synergistic effect of Pt and W
­ Ox as the Pt particles were
erol was adsorbed and activated on W ­ δ+–O–Al interface. highly dispersed with increasing W loading content. The
However, we also found that the decline in ­TOF1,3-PDO-W optimum W loading content was 15%, which was sufficient
after the W surface density of 3.78 W ­nm−2 was also related to create W-O-W clusters providing the delocalization of
to the reduction of Pt dispersion (Fig. 15a). Further in situ electrons to stabilize H atom spillover from the Pt surface.
characterization of Pt-WOx/Al2O3 catalysts is essential to The calcination temperature of ­WOx/Al2O3 engaged the dis-
confirm the active sites and mechanism for hydrogenolysis persion of W­ Ox species over the A ­ l2O3 surface, with the
of glycerol to 1,3-propanediol. ­WOx/Al2O3 calcined 700 °C offering high dispersion of W
species, due to the negative effect on the formation of ­Hδ+
over the ­WOx domain as the surface of highly dispersed
­WOx (isolated ­WOx) was hardly reduced. The calcination
temperature of ­WOx/Al2O3 at 800 °C allowed the connection
of each isolated W­ Ox group to form W-O-W clusters that

13

220 Topics in Catalysis (2023) 66:205–222

was related to the maximum glycerol conversion (44.7%)


and 1,3-propanediol selectivity (45.1%). After impregna-
tion of ­WOx/Al2O3 with a Pt precursor (­ H2PtCl6 ­(H2O)6),
an ample calcination temperature at 400 °C was essential
to decompose the Pt precursor. The reduction temperature
of the Pt-WOx/Al2O3 catalysts at 350 °C was adequate to
achieve the active and selective Pt-WOx/Al2O3 catalysts for
glycerol conversion to 1,3-propanediol.

Supplementary Information  The online version contains supplemen-


tary material available at https://d​ oi.o​ rg/1​ 0.1​ 007/s​ 11244-0​ 22-0​ 1753-9.

Acknowledgements  This work was financially supported by the


Kasetsart University Research and Development Institute (KURDI),
(FF(KU)21.65).

Author Contributions  ND: investigation, writing-reviewing and edit-


ing. TN: investigation, validation, writing-original draft preparation,
writing-reviewing and editing. CW: visualization, writing-reviewing
and editing. AS: visualization, writing-reviewing and editing. KF:
visualization, writing-reviewing and editing. CKC: visualization,
writing-reviewing and editing. D-VNV: visualization, writing-review-
ing and editing. SN: visualization, writing-reviewing and editing. NC:
visualization, writing-reviewing and editing. TW: conceptualization,
methodology, writing-original draft preparation, writing-reviewing and
editing.

Fig. 14  Pt 4d XPS spectra of calcined Pt-15  W/Al2O3-800-400 and Data availability  Authors can confirm that all relevant data are included
Pt-15 W/Al2O3-800-400 reduced at 300, 350 and 400 °C in the article and/or its supplementary information files.

Fig. 15  TOF1,3-PDO-W as a function of Pt dispersion (a) and W surface density (b). Operating conditions: 220 °C, 60 bar, reaction time of 5 h and
10 wt.% glycerol aqueous solution

13
Topics in Catalysis (2023) 66:205–222 221

Declarations  14. Numpilai T, Cheng CK, Seubsai A, Faungnawakij K, Limtrakul


J, Witoon T (2021) Sustainable utilization of waste glycerol for
Conflict of interest  The authors declare that they have no conflicts of 1,3-propanediol production over Pt/WOx/Al2O3 catalysts: effects
interest. of catalyst pore sizes and optimization of synthesis conditions.
Environ Pollut 272:116029. https://d​ oi.o​ rg/1​ 0.1​ 016/j.e​ nvpol.2​ 020.​
116029
15. Zhao B, Liang Y, Liu L, He Q, Dong JX (2021) Facilitating Pt-
References WOx species interaction for efficient glycerol hydrogenolysis to
1,3-propanediol. ChemCatChem 13:3695–3705. https://​doi.​org/​
1. Kurian JV (2005) A new polymer platform for the future-Sorona® 10.​1002/​cctc.​20210​0773
from corn derived 1,3-propanediol. J Polym Environ 13:159–167. 16. Yang M, Wu K, Sun S, Ren Y (2022) Regulating oxygen
https://​doi.​org/​10.​1007/​s10924-​005-​2947-7 defects via atomically dispersed alumina on Pt/WOx catalyst for
2. Bhowmik S, Darbha S (2021) Advances in solid catalysts for enhanced hydrogenolysis of glycerol to 1,3-propanediol. Appl
selective hydrogenolysis of glycerol to 1,3-propanediol. Catal Catal B Environ 307:121207. https://​doi.​org/​10.​1016/j.​apcatb.​
Rev 63:1–65. https://​doi.​org/​10.​1080/​01614​940.​2020.​17947​37 2022.​121207
3. DeConto RM, Pollard D, Alley RB, Velicogna I, Gasson E, Gon- 17. Liu L, Kawakami S, Nakagawa Y, Tamura M, Tomishige K
mez N, Sadai S, Condron A, Gilford DM, Ashe EL, Kopp RE, (2022) Highly active iridium-rhenium catalyst condensed on
Li D, Dutton A (2021) The Paris climate agreement and future silica support for hydrogenolysis of glycerol to 1,3-propanediol.
sea-level rise from Antarctica. Nature 593:83–89. https://​doi.​org/​ Appl Catal B Environ 256:117775. https://​doi.​org/​10.​1016/j.​
10.​1038/​s41586-​021-​03427-0 apcatb.​2019.​117775
4. Cheong S-H, Kim D, Dang HT, Kim D, Seo B, Cheong M, Hong 18. Liu L, Asano T, Nakagawa Y, Tamura M, Okumura K, Tom-
SH, Lee H (2022) Methane oxidation to methyl trifluoroacetate ishige K (2019) Selective hydrogenolysis of glycerol to 1,3-pro-
by simple anionic palladium catalyst: comprehensive understand- panediol over rhenium-oxide-modified iridium nanoparticles
ing of K ­ 2S2O8-based methane oxidation in C ­ F3CO2H. J Catal coating rutile titania support. ACS Catal 9:10913–10930.
413:803–811. https://​doi.​org/​10.​1016/j.​jcat.​2022.​07.​031 https://​doi.​org/​10.​1021/​acsca​tal.​9b038​24
5. Dahnum D, Dang HT, Tran NT, Ha J-M, Lee H (2021) One-pot 19. García-Fernández S, Gandarias I, Requies J, Güemez MB, Ben-
synthesis of 3D-ZIF-7 supported on 2D-Zn-benzimidazole-acetate nici S, Auroux A, Arias PL (2015) New approaches to the Pt/
and its catalytic activity in the methoxycarbonylation of aniline WOx/Al2O3 catalytic system behavior for the selective glycerol
with dimethyl carbonate. J Ind Eng Chem 99:380–387. https://d​ oi.​ hydrogenolysis to 1,3-propanediol. J Catal 323:65–75. https://​
org/​10.​1016/j.​jiec.​2021.​04.​049 doi.​org/​10.​1016/j.​jcat.​2014.​12.​028
6. Nguyen DLT, Nguyen TM, Lee SY, Kim J, Kim SY, Le QV, 20. Zhu S, Gao X, Zhu Y, Li Y (2015) Promoting effect of ­WOx
Varma RS, Hwang YJ (2022) Electrochemical conversion of on selective hydrogenolysis of glycerol to 1,3-propanediol over
­CO2 to value-added chemicals over bimetallic Pd-based nano- bifunctional Pt-WO x/Al 2O 3 catalysts. J Mol Catal A Chem
structures: recent progress and emerging trends. Environ Res 398:391–398. https://​doi.​org/​10.​1016/j.​molca​ta.​2014.​12.​021
211:113116. https://​doi.​org/​10.​1016/j.​envres.​2022.​113116 21. Lei N, Zhao X, Hou B, Yang M, Zhou M, Liu F, Wang A, Zhang
7. Bok J, Lee SY, Lee B-H, Kim C, Nguyen DLT, Kim JW, Jung T (2019) Effective hydrogenolysis of glycerol to 1,3-propan-
E, Lee CW, Jung Y, Lee HS, Kim J, Lee K, Ko W, Kim YS, Cho ediol over metal-acid concerted Pt/WOx/Al2O3 catalysts. Chem-
S-P, Yoo JS, Hyeon T, Hwang YJ (2021) Designing atomically CatChem 11:3903–3912. https://​d oi.​o rg/​1 0.​1 002/​c ctc.​2 0190​
dispersed Au on tensile-strained Pd for efficient C ­ O2 electroreduc- 0689
tion to formate. J Am Chem Soc 143:5386–5395. https://​doi.​org/​ 22. Kitano T, Hayashi T, Uesaka T, Shishido T, Teramura K, Tanaka T
10.​1021/​jacs.​0c126​96 (2014) Effect of high-temperature calcination on the generation of
8. Kwon HC, Kim M, Grote J-P, Cho SJ, Chung MW, Kim H, Won Brønsted acid sites on W ­ O3/Al2O3. ChemCatChem 6:2011–2020.
DH, Zeradjanin AR, Mayrhofer KJJ, Choi M, Kim H, Choi CH https://​doi.​org/​10.​1002/​cctc.​20140​0053
(2018) Carbon monoxide as a promoter of atomically dispersed 23. Jarauta-Córdoba C, Bengoechea MO, Agirrezabal-Telleria I,
platinum catalyst in electrochemical hydrogen evolution reaction. Arias P-L, Gandarias I (2021) Insights into the nature of the active
J Am Chem Soc 140:16198–16205. https://​doi.​org/​10.​1021/​jacs.​ sites of Pt-WOx/Al2O3 catalysts for glycerol hydrogenolysis into
8b092​11 1,3-propanediol. Catalysts 11:1171. https://​doi.​org/​10.​3390/​catal​
9. Yusuff A, Kumar M, Obe BO, Mudashiru O (2021) Calcium oxide 11101​171
supported on coal fly ash (CaO/CFA) as an efficient catalyst for 24. Suwannapichat Y, Numpilai T, Chanlek N, Faungnawakij K,
biodiesel production from Jatropha curcas oil. Top Catal. https://​ Chareonpanich M, Limtrakul J, Witoon T (2018) Direct synthe-
doi.​org/​10.​1007/​s11244-​021-​01478-1 sis of dimethyl ether from C ­ O2 hydrogenation over novel hybrid
10. Jeong S-H, Lee H-S, Kim D-K, Lee J-P, Park J-Y, Hwang K-R, catalysts containing a Cu-ZnO-ZrO2 catalyst admixed with W ­ O x/
Lee J-S (2017) Biodiesel production from highly free fatty acid Al2O3 catalysts: effects of pore size of ­Al2O3 support and W load-
oils using a bifunctional solid catalyst. Top Catal 60:651–657. ing content. Energy Convers Manag 159:20–29. https://​doi.​org/​
https://​doi.​org/​10.​1007/​s11244-​017-​0772-6 10.​1016/j.​encon​man.​2018.​01.​016
11. Keogh J, Deshmukh G, Manyar H (2022) Green synthesis of 25. Graham GW, Weber WH, McBride JR, Peters CR (1991) Raman
glycerol carbonate via transesterification of glycerol using investigation of simple and complex oxides of platinum. J Raman
mechanochemically prepared sodium aluminate catalysts. Fuel Spectrosc 22:1–9. https://​doi.​org/​10.​1002/​jrs.​12502​20102
310:122484. https://​doi.​org/​10.​1016/j.​fuel.​2021.​122484 26. Wu X, Zhang L, Weng D, Liu S, Si Z, Fan J (2012) Total oxidation
12. Durán-Martín D, Lόpez Granados M, Fierro JLG, Pinel C, Mari- of propane on Pt/WOx/Al2O3 catalysts by formation of metasta-
scal R (2017) Deactivation of CuZn catalysts used in glycerol ble ­Ptδ+ species interacted with ­WOx clusters. J Hazard Mater
hydrogenolysis to obtain 1,2-propanediol. Top Catal 60:1062– 225–226:146–154. https://d​ oi.o​ rg/1​ 0.1​ 016/j.j​ hazma​ t.2​ 012.0​ 5.0​ 11
1071. https://​doi.​org/​10.​1007/​s11244-​017-​0807-z 27. Shi G, Cao Z, Xu J, Jin K, Bao Y, Xu S (2018) Effect of ­WOx dop-
13. Jiang Y, Li X, Zhao H, Hou Z (2019) Esterification of glycerol ing into Pt/SiO2 catalysts for glycerol hydrogenolysis to 1,3-pro-
with acetic acid over S ­ O3H-functionalized phenolic resin. Fuel panediol in liquid phase. Catal Lett 148:2304–2314. https://​doi.​
255:115842. https://​doi.​org/​10.​1016/j.​fuel.​2019.​115842 org/​10.​1007/​s10562-​018-​2464-7

13

222 Topics in Catalysis (2023) 66:205–222

28. Liang Y, Shi G, Jin K (2020) Promotion effect of ­Al2O3 on Pt- 34. Barton DG, Soled SL, Iglesia E (1998) Solid acid catalysts based
WOx/SiO2 catalysts for selective hydrogenolysis of bioglycerol on supported tungsten oxides. Top Catal 6:87–99. https://​doi.​org/​
to 1,3-propanediol in liquid phase. Catal Lett 150:2365–2376. 10.​1023/A:​10191​26708​945
https://​doi.​org/​10.​1007/​s10562-​020-​03140-z 35. Dang HT, Lee HW, Lee J, Choo H, Hong SH, Cheong M, Lee H
29. Li W, Chi K, Liu H, Ma H, Qu W, Wang C, Lv G, Tian Z (2017) (2018) Enhanced catalytic activity of (DMSO)2PtCl2 for the meth-
Skeletal isomerization of n-pentane: a comparative study on cata- ane oxidation in the ­SO3–H2SO4 system. ACS Catal 8:11854–
lytic properties of Pt/WOx-ZrO2 and Pt/ZSM-22. Appl Catal A 11862. https://​doi.​org/​10.​1021/​acsca​tal.​8b041​01
Gen 537:59–65. https://​doi.​org/​10.​1016/j.​apcata.​2017.​03.​005 36. Lee HW, Dang HT, Kim H, Lee U, Ha J-M, Jae J, Cheong M,
30. Panagiotopoulou P, Kondarides DI (2008) Effects of alkali addi- Lee H (2019) Pt black catalyzed methane oxidation to methyl
tives on the physicochemical characteristics and chemisorptive bisulfate in H­ 2SO4-SO3. J Catal 374:230–236. https://​doi.​org/​10.​
properties of Pt/TiO2 catalysts. J Catal 260:141–149. https://​doi.​ 1016/j.​jcat.​2019.​04.​042
org/​10.​1016/j.​jcat.​2008.​09.​014
31. Lisitsyn AS, Yakovina OA (2018) On the origin of high-temper- Publisher's Note Springer Nature remains neutral with regard to
ature phenomena in Pt/Al2O3. Phys Chem Chem Phys 20:2339– jurisdictional claims in published maps and institutional affiliations.
2350. https://​doi.​org/​10.​1039/​C7CP0​6925A
32. Aihara T, Miura H, Shishido T (2020) Investigation of the mecha- Springer Nature or its licensor (e.g. a society or other partner) holds
nism of the selective hydrogenolysis of C-O bonds over a Pt/WO3/ exclusive rights to this article under a publishing agreement with the
Al2O3 catalyst. Catal Today 352:73–79. https://​doi.​org/​10.​1016/j.​ author(s) or other rightsholder(s); author self-archiving of the accepted
cattod.​2019.​10.​008 manuscript version of this article is solely governed by the terms of
33. Barton DG, Shtein M, Wilson RD, Soled SL, Iglesia E (1999) such publishing agreement and applicable law.
Structure and electronic properties of solid acids based on tung-
sten oxide. J Phys Chem B 103:630–640. https://​doi.​org/​10.​1021/​
jp983​555d

13

You might also like