Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of Catalysis 404 (2021) 96–108

Contents lists available at ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Facet-Dependent selectivity of CeO2 nanoparticles in 2-Propanol


conversion
Berlin Sudduth a, Dongmin Yun a, Junming Sun a, Yong Wang a,b,⇑
a
The Gene and Linda Voiland School of Chemical Engineering and Bioengineering, Washington State University, Pullman, WA 99164, USA
b
Insititute for Integrated Catalysis, Pacific Northwest National Laboratory, Richland, Washington 99354, USA

a r t i c l e i n f o a b s t r a c t

Article history: CeO2 nanoshapes, cubes with dominant (1 0 0) facets and octahedra with dominant (1 1 1) facets, were
Received 8 July 2021 synthesized to investigate the influence of surface structure on acid-base properties. An optimization
Revised 4 September 2021 of calcination temperatures, coupled with Raman and TEM, was employed to minimize the intrinsic
Accepted 11 September 2021
(Frenkel-type) defect sites and their potential complications on the facet studies involved in this work.
Available online 20 September 2021
The acid-base properties of these CeO2 nanoshapes were characterized with in situ pyridine and CO2
adsorption using infrared spectroscopy and quantified using pyridine, ammonia, and CO2 temperature-
Keywords:
programmed desorption (TPD). The (1 0 0) facet displayed weaker acid sites with lower site density than
Ceria
Facets
the (1 1 1) facet which had a similar density of base sites but those on the (1 0 0) facet were stronger. A
Acid-base properties strong correlation was observed between 2-propanol conversion and each facet’s acid-base properties.
2-Propanol CeO2 cubes exhibited greater base-site catalyzed dehydrogenation to acetone while the more acidic
Dehydrogenation CeO2 octahedra were more active in dehydration to propene.
Dehydration Ó 2021 Published by Elsevier Inc.

1. Introduction [2,12,13]. The type and strength of these sites are determined by
IR features (peak frequency and intensity) of the adsorbed probe
The acid-base properties of solid metal oxides such as titania molecules and changes to associated functional groups (e.g.,
(TiO2), zirconia (ZrO2), and ceria (CeO2) play key roles in catalytic hydroxyl groups) on the surface [6,14–17]. However, identifying
reactions [1–3]. Studies characterizing the acid-base properties of the structure–function relationships between metal oxide surfaces
such metal oxides commonly reported that coordinatively unsatu- and their acid-base catalytic behavior remains elusive owing to the
rated metal cations (Mn+) exposed on the surface act as Lewis acids, structural heterogeneity of active sites on traditional polycrys-
while coordinatively unsaturated oxygen anions (O2–) act as Lewis talline nanoparticles [5]. Additionally, it has been shown that the
bases [4,5]. Surface functional groups can also influence acid-base concentration of oxygen defect sites varies depending on both
properties as demonstrated with hydroxyl groups (OH) which can the exposed facet and the method of synthesis. The presence of
act as Brønsted acid or base sites depending on the metal oxide oxygen vacancies has been closely associated with the redox prop-
[6,7]. The types and strengths of acid and base sites can vary signif- erties of CeO2 [18–20]; with more defective nanocrystals express-
icantly between different metal oxides [5,8,9], but surface science ing increased reactivity in redox reactions regardless of the
and theoretical studies have shown that surface structure alone dominant facet exposed [21,22]. The influence of oxygen vacancies
may be a significant contributor. For example, in MgO [10] and on acid-base properties is relatively less explored but may further
Al2O3 [11] studies, the coordination environment of the surfaces complicate the development of structure–function relationships;
leads to variations in acid-base properties. Acid-base properties some studies suggest defect sites increase the concentration of
are often characterized by obtaining infrared (IR) spectra of the Lewis acid sites [23–25].
metal oxide after adsorption of basic probe molecules (ammonia, Metal oxide nanoparticles with specific morphologies are
pyridine, and CD3CN) or acidic probe molecules (CO2, CHCl3 and promising due to their more uniform surface terminations; analo-
benzaldehyde) to identify acid or base sites, respectively gous to the crystallographically synthesized single crystals used in
surface science studies [26–28]. A recent study demonstrated that
⇑ Corresponding author at: The Gene and Linda Voiland School of Chemical both acid site density, presence of Brønsted acids, and the 2-
Engineering and Bioengineering, Washington State University, Pullman, WA 99164, propanol adsorption species are influenced by the dominantly
USA. exposed facet on TiO2 model nanoparticles [29]. These structural
E-mail address: yong.wang@pnnl.gov (Y. Wang).

https://doi.org/10.1016/j.jcat.2021.09.009
0021-9517/Ó 2021 Published by Elsevier Inc.
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

differences resulted in increased acid site catalyzed dehydration of the calcination temperatures of the synthesized nanoshapes. Com-
2-propanol observed on TiO2(1 0 1) due to a lower activation bar- mercial CeO2 was also studied due to its unique defect density and
rier associated with the molecularly adsorbed reaction mechanism polycrystalline structure. The catalytic behavior of these different
compared to that of TiO2 (0 0 1) which was supported by DFT calcu- CeO2 nanocrystals was examined using 2-propanol dehydrogena-
lations. CeO2 with different nanoshapes have been hydrothermally tion and dehydration as a probe reaction. 2-propanol has been
synthesized which possess well-defined surface structures [30– widely used to study acid-base catalysis of metal oxides
32]. CeO2 nanoshapes in the form of octahedra dominantly expos- [9,29,42–45] because, while ethanol and 1-propanol follow similar
ing the (1 1 1), cubes exposing the (1 0 0), and rods with a mixture pathways, 2-propanol tends to be more sensitive to differences in
of (1 1 0) and (1 0 0) facets provide an opportunity to probe these acid-base properties [27,46]. The strength and amount of acid
specific surface structures while providing meaningful reactivity and base sites of ceria nanocrystals are reflected in 2-propanol con-
measurements on high surface area nanocrystals. The surface- version and selectivity toward acetone and propene.
dependent acid-base properties of well-defined ceria nanocrystals
were investigated by Wu et al., by means of in situ IR spectroscopy
2. Experimental methods
and temperature programmed desorption (TPD) [33]. It was sug-
gested that the strengths of Lewis acid sites (Ce cations) are rela-
2.1. Catalyst preparation
tively similar regardless of the shape of the ceria nanocrystals
while the amount and strength of base sites (O anions and hydro-
CeO2 nanoshapes (cubes and octahedra) were synthesized using
xyl groups) are shape and defect dependent. However, the relative
hydrothermal methods with a Ce(NO3)36H2O (Sigma-Aldrich
quantity of acid sites on these CeO2 nanoshapes, the contribution
99.99%) cerium source and a NaOH precipitator, as reported else-
of defect density and the impact of these acid-base properties on
where [30,32]. Briefly, the morphology of the ceria was determined
a catalytic reaction has yet to be investigated.
by varying the concentrations of the precursors and the hydrother-
In this study, we characterized the acid-base properties of CeO2
mal treatment conditions. CeO2 cubes (denoted as CeO2-C) were
nanocrystals with well-defined (1 1 1) facets in the form of octahe-
synthesized using an aqueous solution of 36 g Ce(NO3)36H2O in
dra and (1 0 0) facets in the form of cubes using in situ IR spec-
350 mL of DI water and an aqueous solution of 2.17 g NaOH in
troscopy and temperature programmed desorption. CeO2(1 1 1),
50 mL DI water while ceria octahedra (denoted as CeO2-O) used
Fig. 1a, is the most stable surface with the exposed O and Ce each
15 g and 0.025 g of each respective precursor. The two precursor
having a single coordination vacancy [34]. Scanning tunneling
solutions were mixed at room temperature before being placed
microscopy (STM) [35], atomic force microscopy (AFM) [36] and
in a four 125 mL Teflon lined steel autoclave and heated to
ion scattering [36] studies have shown that CeO2(1 1 1) is termi-
180 °C for 24 h (cubes) or 12 h (octahedra). The resulting CeO2 pre-
nated by a layer of O. A simple truncation along CeO2(1 0 0) results
cipitate was separated via centrifuge and washed with deionized
in an unstable surface due to the dipole caused by the presence of
water and ethanol three times each in an ultrasonic bath. Impuri-
twice as many O as Ce in adjacent layers. This dipole is eliminated
ties, such as the sodium used during synthesis, has been shown to
when half of the surface O are moved from the top face to the bot-
influence the acid-base properties of CeO2 but additional dilute
tom face [37]. This rearrangement has been supported by STM [38]
acid and base washes have shown to minimize the effect of impu-
and high-resolution transmission election microscopy (HRTEM)
rities [33]. Removal of additional sodium impurities was accom-
[39,40] experiments which suggest the surface is not terminated
plished by further washing the products with a solution of 0.1 M
by O or Ce exclusively, but a mixture of both ions. The O and Ce
NH4OH before three additional water rinses followed by washing
on the CeO2(1 0 0) structure in Fig. 1b is the reconstructed configu-
with a 0.1 M HNO3 and three final water rinses with each rinse
ration [41] in which both ions have two coordination vacancies.
being done in an ultrasonic bath [47]. The washed precipitate
In addition to studying the two different surface facets, the
was dried at 50 °C for 12 h and calcined for 4 h in air. Calcination
effect of defects on acid-base properties was studied by varying
temperatures were varied from 450 to 750 °C as a means of con-
trolling ceria defect density [23]. Commercial ceria (denoted as
CeO2-Com) was obtained from Sigma-Aldrich (99.995%) and used
without any further modification.

2.2. Catalyst characterization

Samples were prepared for TEM characterization by dispersing


approximately 1 mg of the powder CeO2 in 5 mL of ethanol. These
solutions were sonicated for 20 min, added dropwise over a nickel
200 mesh TEM grid with Formvar/carbon coating, and dried over-
night under a heat lamp in air. TEM was performed using an FEI
Technai G2 20 Twin operated at 200 kV and equipped with a 4 k
Eagle CCD camera. Brunauer-Emmett-Teller (BET) analysis was
used to determine the specific surface areas (SSA) of the CeO2 sam-
ples. Approximately 0.5 g of samples were placed in quartz tubes
and degassed under vacuum at 300 °C for 3 h. N2 adsorption and
desorption isotherms were obtained at 77 K using surface area
analyzer (Tristar II, Micromeritics). In situ Raman spectroscopy
was used to measure the defect density of the ceria catalysts.
About 50 mg of samples were loaded into an in situ reactor cell
Fig. 1. Top and side views of (a) CeO2(1 1 1) and (b) reconstructed CeO2(1 0 0) (Linkam CCR1000) and heated to 400 °C for 1 h under 10% O2/He
surfaces. Ce is shown in blue and O in red. The coordinated vacancies (CV) of the
first layer O and second layer Ce are indicated. (SINGLE COLUMN). (For interpre-
(20 mL/min). The pretreated samples were cooled to 100 °C before
tation of the references to colour in this figure legend, the reader is referred to the measurements from 0 to 1400 cm1 using a Raman spectroscope
web version of this article.) (Horiba LabRAM HR 800) equipped with a 532 nm (Ventus LP
97
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

532) laser source and Synapse CCD (Charge Coupled Device) detec- generated in a flask of Py held at room temperature with 10 mL/
tor. XRD patterns were obtained using a Rigaku MiniFlex X-ray min He flow until saturation. The sample was purged under 5%
diffractometer using Cu-Ka (k = 1.54178 A).€ Normal operating Ar/He flow (50 mL/min) for 2 h to remove physically adsorbed
power was 40 kV, 30 mA and measurements were obtained from pyridine. The temperature was then increased to 600 °C at various
2h of 3° to 90° at scanning speeds of 0.5°/min. Diffuse reflectance ramping rates (2, 5, 15 °C/min) under 5% Ar/He flow (50 mL/min).
infrared Fourier transform spectroscopy (DRIFTS) was conducted The QMS was used to measure the argon tracer (m/z = 40) and des-
using a Bruker Tensor 27 FTIR spectrometer equipped with a Har- orbed pyridine (m/z = 79) during TPD.
rick Praying Mantis attachment by averaging 256 scans with a
spectral resolution of 4 cm1. For analysis of the surface hydroxyl 2.4. Catalyst activity tests
groups, samples were pretreated at 400 °C for 1 h under 10% O2/
He flow (20 mL/min) and cooled to 150 °C before spectral acquisi- The dehydrogenation and dehydration of 2-propanol was used
tion under He flow (20 mL/min). The hydroxyl group vibrations are as a probe reaction to test the catalytic activity of the CeO2
difference spectra using potassium bromide (KBr) at 150 °C as the nanoshapes. Isothermal rate measurements were carried out in a
reference background. fixed-bed flow reactor with catalyst packed into a quartz tube.
Approximately 50 mg of the CeO2 samples were diluted with inert
SiC at a 1:10 CeO2:SiC mass ratio to prevent temperature gradients.
2.3. Acid and base site characterization
A syringe pump was used to inject the 2-propanol upstream of the
catalyst and heated to 180 °C to ensure complete evaporation and
Pyridine (Py = C5H5N) adsorption in the previous mentioned
mixing with a He carrier gas before reaching the reactor. The outlet
DRIFTS equipment was used to probe the surface acid sites of each
of the reactor was heated to 150 °C to avoid product condensation.
ceria catalyst. Samples were pretreated with 10% O2/He at 400 °C
An online Micro-GC (Agilent 490) equipped with four columns
for 1 h and cooled to ambient temperature under He (20 mL/
(Molsieve, CP-PoraPLOT U, aluminum oxide column and CP-Sil 5
min). He (10 mL/min) was used as a carrier gas and flowed through
CB) and TCD detectors was used to analyze the reaction products.
a bubbler with pyridine at room temperature to continuously fill a
Catalysts were pretreated with dry air (50 mL/min) at 400 °C
0.5 mL loop attached to a six-way valve. Pulses of pyridine were
before lowering the reactor to the desired reaction temperature.
injected from the loop to the catalysts using He (20 mL/min) until
Selectivity and activity were monitored with an inlet 2-propanol
saturation, as monitored by the pyridine signal (m/z = 79) from an
partial pressure of 6 kPa (weight hourly space velocity of
online quadrupole mass spectrometer (QMS, Quadra 220, Pfeiffer)
7.5 g2-propanolh1g-1
CeO2) at 10 °C increments between the tempera-
attached to the effluent stream. Samples were purged under He
tures of 250 and 290 °C. Fresh, pretreated catalysts were used for
flow (20 mL/min) for 30 min before acquiring spectra. Spectra col-
measurements at each temperature. Reaction rates were measured
lected at the various desorption temperatures used the pretreated
under the differential kinetic regime (conversions below 5%) and
ceria at the same temperature, before pyridine adsorption, as the
normalized to turnover frequencies (TOF) based on the surface area
background. A similar procedure was followed for DRIFTS analysis
of the CeO2 catalysts as obtained by BET.
of CO2 adsorption and desorption. After pretreatment, 10% CO2/He
(20 mL/min) is introduced at room temperature until saturation.
Samples are then purged under He flow (20 mL/min) for 30 min 3. Results and discussion
before acquiring room temperature spectra and increasing the
temperature to 400 °C at a ramp rate of 10 °C/min. The pretreated 3.1. Physical and structural properties
CeO2 at the various temperatures, before CO2 adsorption, were
used as the background. For a meaningful comparison of acid and base properties of
Temperature-programmed desorption of various probe mole- CeO2, the concentration of intrinsic (Frenkel-type) defects may be
cules was performed using a chemisorption unit (AutoChem II minimized to reduce the potential impact of these defects on the
2920 Chemisorption Analyzer, Micromeritics) coupled with an acid-base moieties [25,33,48]. It should be highlighted that the
online QMS (Quadra 220, Pfeiffer). Samples were measured to intrinsic defects are different from oxygen vacancies. The oxygen
matching surface areas and loaded into a quartz tube supported vacancies are dynamically formed upon reduction while the intrin-
by quartz wool. Each sample was first pretreated at 400 °C under sic defects are present in the fully oxidized CeO2 structure, com-
10% O2/He flow (50 mL/min) for 1 h before cooling to ambient tem- monly caused by the relocation of oxygen anion (O2–) from
perature and exposing the sample to the probe molecule. tetrahedral to octahedral coordination [21,49]. For this purpose,
Ammonia-TPD (NH3-TPD) was conducted by then exposing the Raman spectroscopy and TEM imaging were carried out to find
sample to 10% NH3/He (50 mL/min) until saturation. The sample the optimum calcination temperature which produces the desired
was purged under 5% Ar/He flow (50 mL/min) for 2 h to remove ceria nanoshape with the minimum intrinsic defect concentration.
physically adsorbed ammonia. The temperature was then Fig. 2 shows the Raman spectra of CeO2-C and CeO2-O calcined at
increased to 600 °C at a ramping rate of 10 °C/min under 5% Ar/ different temperatures. Regardless of the calcination temperature
He flow (50 mL/min). The QMS was used to measure the argon tra- and shape of ceria nanocrystals, Raman spectra of ceria show an
cer (m/z = 40) and desorbed NH3 (m/z = 16) during TPD. The QMS intense peak at 465 cm1, assigned to the Ce-O vibration mode
response was calibrated by flowing various concentrations of (F2g) whose intensity was used to normalize the entire spectra
NH3 with the Ar/He carrier. Carbon dioxide-TPD (CO2-TPD) was for comparisons [21]. Two weak Raman bands at 241 and
conducted by exposing the sample to 10% CO2/He (50 mL/min) 598 cm1 were also observed which represent the second-order
until saturation. The sample was purged under 5 %Ar/He flow transverse acoustic (2TA) and defect-induced (D) vibrational
(50 mL/min) for 2 h to remove physically adsorbed carbon dioxide. modes of CeO2, respectively [50,51]. The relative intensity (ID/
The temperature was then increased to 600 °C at a ramping rate of IF2g) between the defect-induced vibrational mode at 598 cm1
10 °C/min under 5 %Ar/He flow (50 mL/min). The QMS was used to (ID) and F2g bands at 465 cm1 (IF2g) at different temperatures, as
measure the argon tracer (m/z = 40) and desorbed CO2 (m/z = 44) shown in Fig. 2c, was used to determine the change of defect den-
during TPD. The QMS response was calibrated by flowing various sity of CeO2 [21]. For both nanoshapes, defect density decreases as
concentrations of CO2 with the Ar/He carrier. Pyridine-TPD (Py- calcination temperatures increase, a result of the collapse of tetra-
TPD) was conducted by exposing the sample to pulses of Py vapor hedral sites where the intrinsic defects sites were present [23].
98
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

Fig. 2. Normalized Raman spectra of CeO2-C (a) and CeO2-O (b) calcined at 450, 550, 650, and 750 °C. All spectra were normalized to the height of the dominant band at
465 cm1. (c) The defect density (Raman band peak area at 593 cm1 /peak area at 463 cm1) as a function of calcination temperature (°C). (d) Normalized Raman spectra of
CeO2-C, CeO2-O, and CeO2-Com calcined under the conditions used for the remainder of the study. (2-COLUMN).

The morphological stability of ceria nanocrystals was deter- TEM images of the synthesized CeO2-C calcined at 450 °C, CeO2-
mined using TEM. Fig. 3 shows TEM images obtained for CeO2 O calcined at 700 °C and CeO2-Com are shown in Fig. 4. Analysis of
nanoshapes (CeO2-C and CeO2-O) calcined at different tempera- the TEM images indicates that the average particle sizes of CeO2-C
tures (450, 550, 650, and 750 °C). The morphology of CeO2-C (Fig. 4a), CeO2-O (Fig. 4b) and CeO2-Com (Fig. 4c) are approxi-
remained stable up to 450 °C and began to show changes above mately be 47, 28, and 43 nm, respectively. The CeO2-C contain
550 °C, first with the rounding of the edges and corners and even- mostly the (1 0 0) plane and a small portion of the (1 1 0) plane,
tually with complete loss of their initial cubic shape at 700 °C. In while the surface of CeO2-O dominantly expose the (1 1 1) plane,
contrast, the ceria octahedra are thermally stable up to 700 °C, as indicated in our previous contribution [52]. By contrast, the
indicating that CeO2 octahedra are robust and largely stable at morphology of the CeO2-Com, however, is not well defined. The
higher temperatures. These observations are consistent with DFT main shape of commercial CeO2 is truncated octahedra containing
calculations which predict that CeO2(1 1 1) has the lowest surface a majority of (1 1 1) with truncated ends exposing the (1 0 0) planes.
energy of ceria’s low index surfaces [34]. Based on these results, The XRD patterns of the CeO2-Com, CeO2-C, and CeO2-O are
calcination temperatures of 450 and 700 °C were used for CeO2-C shown in Fig. 5. Each CeO2 nanocrystals exhibits peaks at 28.5°,
and CeO2-O, respectively. These conditions assure the minimum 33.1°, 47.5°, 56.4°, 59.1°, 69.4°, 76.7°, 79.1°, and 88.5°, which are
concentration of intrinsic defect without morphological change. A associated with the (1 1 1), (2 0 0), (2 2 0), (3 1 1), (2 2 2), (4 0 0),
comparison of the Raman spectra for the synthesized CeO2 (3 3 1), (4 2 0), (4 2 2)} facets of pure cubic fluorite structure (space
nanoshapes and the as-obtained commercial CeO2 is shown in group Fm3m 225), as identified in JCPDS 34–0394. None of the
Fig. 2d indicating that the defect density of the samples are nanocrystals exhibit any peaks from Ce(OH)3 or Ce(OH)CO3 impu-
CeO2-C > CeO2-O > CeO2-Com. rities which indicates sufficiently high calcination temperatures

99
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

Fig. 3. (Top) TEM images of CeO2-C calcined at (a) 450 °C, (b) 550 °C, (c) 650 °C and (d) 750 °C. (Bottom) TEM images of CeO2-O calcined at (e) 450 °C, (f) 550 °C, (g) 650 °C and
(h) 750 °C. (2-COLUMN).

were used [30]. The average crystallite size of CeO2-Com, CeO2-C, shown to influence the vibrational frequency of pyridine’s v8a
and CeO2-O estimated using Scherrer equation are 40, 44, and mode, with stronger acid sites causing a blue shift of this mode
29 nm, respectively, comparable with estimates from TEM imag- [54,55]. The v8a mode of each CeO2 nanoparticle appears at the
ing. The BET surface areas of CeO2-C, CeO2-O and CeO2-Com are same 1600 cm1 frequency suggesting that pyridine is unable to
10.4, 25.3 and 17.6 m2/g, respectively. The physical properties of distinguish any differences in the strengths of each surface’s Ce
each CeO2 sample are summarized in Table 1. cations. This is consistent with previous studies comparing pyri-
dine adsorption on CeO2 nanoshapes or different metal oxides
3.2. Acid sites characterization [12,33]. Pyridine is too strong a basic molecule to produce measur-
able shifts in the v8a mode if the acid site strengths are relatively
In situ IR spectroscopy has been extensively used to analyze the similar and the vibrational modes of a weaker base, like deuterated
acid properties (e.g., type and strength) of metal oxides [12]. Pyri- acetonitrile, are more sensitive to small changes in acid site
dine is commonly used as a probe molecule because its vibrational strength. The surface OH groups of the CeO2 nanoparticles exhibit
modes are sensitive to adsorption at Lewis and Brønsted acid sites. no Brønsted acidity as the v19b mode of protonated pyridine (i.e.,
DRIFTS spectra of pyridine adsorbed on CeO2-Com, CeO2-C, and ring vibration) is not observed at 1540 cm1. DRIFTS spectra of the
CeO2-O are shown in Fig. 6a. The spectra were acquired after hydroxyl region after Py adsorption (Figure S1) shows that, while
adsorption at room temperature, purging under He flow for no Brønsted acid associated pyridine vibrations could be seen at
30 min, and desorption at 150 °C for 30 min to remove 150 °C, pyridine is hydrogen bonded to all three sample as evi-
hydrogen-bonded pyridine [1]. The bands around 1600 and denced by the negative peaks most prominent in the bridged Ce-
1443 cm1 are associated with the v8a and v19b modes of OH region (3686 – 3633 cm1).
adsorbed pyridine on Lewis acid sites, respectively [5,53]. All three While Py-IR was able to identify the types of acid sites on the
CeO2 nanoparticles exhibit pyridine adsorbed on Lewis acid sites. CeO2 nanoparticles, it was unable to determine the quantity of
The Lewis acid strength of a metal oxide’s cations sites has been these sites NH3-TPD is often employed to investigate these charac-

Fig. 4. TEM images of (a) CeO2-C calcined at 450 °C, (b) as-obtained CeO2-Com, (c) CeO2-O calcined at 700 °C. (2-COLUMN).

100
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

sorption on the observed trends. The linear equation obtained by


altering b and tracking its influence on TM provides a slope Ed/R
[59]. The energy of desorption was determined highest on CeO2-
Com (141 kJ/mol), slightly lower on CeO2-O (132 kJ/mol), and low-
est on CeO2-C (93 kJ/mol). These values are consistent with the
NH3-TPD results which suggested greater acid site strengths on
CeO2-Com and CeO2-O with noticeably weaker acid sites on
CeO2-C.
There are clear differences in the acid site properties among
three ceria samples: (1 1 1) dominant CeO2-Com, (1 1 1) dominant
CeO2-O, and (1 0 0) dominant CeO2-C. While calcination tempera-
tures were used to minimize the defect density of these samples,
some have suggested that defects increase the density of acid sites
on CeO2 [23,25,47]. Thus, it was important to explore the effect of
defect density on the CeO2 nanoshapes’ acid site densities. Fig. 8a
details the changes in acid site density with increasing calcination
temperature for both CeO2-O and CeO2-C. There is no clear trend
between calcination temperature and acid site density. There is a
small decrease from 1.14 to 0.85 mmol NH3 / nm2 when octahedra
are calcined at 450 °C and 750 °C, respectively, while cubes expe-
rience a slight increase from 0.6 to 0.72 mmol NH3 / nm2 with the
Fig. 5. XRD patterns of CeO2-Com, CeO2-C, and CeO2-O. (SINGLE COLUMN). same calcination temperature increase. In a previous study mea-
suring the acid site density of polycrystalline CeO2 with increasing
temperature [23], the density was normalized to the mass of the
teristics of metal oxides as it binds to both Lewis and Brønsted acid CeO2, which does not account for the decrease in specific surface
[56–58]. A comparison of the NH3-TPD profiles of the CeO2 cata- area after increased calcination temperatures, and may explain this
lysts is shown in Fig. 6b. All three nanoshapes exhibit slight NH3 discrepancy. When the acid site density is directly compared to the
desorption when the temperature is elevated above ambient tem- defect densities, as shown in Fig. 8b, there is a clear difference
perature with an increase in desorption above 70 °C. The NH3 des- between the two CeO2 nanoshapes regardless of their defect den-
orption behavior of CeO2-Com and CeO2-O are similar and have the sity. After cubes are calcined at 750 °C, under which conditions it
same desorption maximum at 119 °C with the remaining NH3 des- experiences significant morphological changes to the more stable
orbed as temperature reaches 250 °C. On the other hand, CeO2-C (1 1 1) facet, the acid site density begins converging with those
rapidly reach peak desorption at 100 °C suggesting CeO2-C have measured for octahedra. This observation provides further evi-
slightly weaker acid sites than the (1 1 1) dominant CeO2-Com dence that the acid properties of CeO2 are dependent on their
and CeO2-O. Integration of the NH3 desorption peak areas allows exposed facet and that defect density, while it may still play a role,
for the quantification of the total amount of acid sites and, after does not fully explain these differences.
normalizing to the surface area, the density of the acid sites on
the CeO2. CeO2-Com possess the highest density of acid sites
(1.13 NH3/nm2), closely followed by CeO2-O (0.94 NH3/nm2), with 3.3. Base site characterization
CeO2-C having the lowest (0.60 NH3/nm2).
Pyridine-TPD was used to further distinguish differences in CO2 is a simple acidic probe molecule commonly used to probe
Lewis acid site strength between the three nanoshapes. Three sep- the two types of basic sites on metal oxides, hydroxyl groups and
arate TPD experiments were performed using different tempera- surface oxygen anions [12,60,61]. CO2 forms bicarbonate species
ture ramp rates (b = 2, 5, 15 °C) and the desorption peak from interactions with basic OH groups and a range of carbonate
temperature (TM) from each desorption profile was used to deter- species from oxygen anions which can be identified by their char-
mine the energy of desorption for each CeO2 nanoparticle using the acteristic IR vibrational modes whose assignments are summarized
Arrhenius equation and assuming first order desorption kinetics, in Table 2 and based on previous CO2 adsorption studies [13,62,63].
Fig. 9 shows in situ DRIFTS spectra collected after ambient temper-
kd ¼ Ad expðEd =RTÞ
ature adsorption of CO2 on CeO2-C and CeO2-O, respectively, and
purging at increased temperatures. IR spectra after adsorption of
2lnðTM Þ  lnðbÞ ¼ Ed =RTm þ lnðEd =RAd Þ
CO2 on CeO2-Com are similar to those from CeO2-O and can be
where Ed is the energy of desorption and Ad is the exponential found in Figure S3.
prefactor. A summary of the Py-TPD results is shown in Fig. 7 Bicarbonate vibrations [HCO3–: mas(CO3–) = 1635 cm1, ms(CO3–
where 2lnðTM Þ  lnðbÞ is plotted versus 1/TM. The individual des- ) = 1405 cm1, and d(OH) = 1208 cm1] can be seen on CeO2-C indi-
orption profiles can be found in Figure S2. Pyridine desorbed with- cating that the surface OH on the (1 0 0) surface are basic in nature
out significant oxidation or decomposition as no products of these and persist up to 250 °C. Analysis of the hydroxyl group region dur-
reactions were be observed via QMS. Particle size and bed height ing CO2 desorption, shown in Figure S4, shows the CO2 interacts
were similar for each sample to minimize the influence of read- with the OH species on CeO2-C as indicated by the negative peak

Table 1
Summary of the physical properties of the CeO2 nanoshapes.

Sample Dominant Facet BET SSA (m2/g) Particle Size (nm) Defect Density (ID/IF2g)
CeO2-Com (1 1 1) 17.6 43 0.009
CeO2-C (1 0 0) 10.4 47 0.016
CeO2-O (1 1 1) 25.3 28 0.012

101
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

Fig. 6. Acid site characterization of the CeO2-Com, CeO2-C and CeO2-O. (a) DRIFTS spectra of pyridine adsorbed on the ceria nanoshapes after desorption at 150 °C. The
characteristic locations of modes associated with Lewis and Brønsted acid sites are indicated with dashed lines. (b) QMS profiles of desorbed NH3 during NH3-TPD normalized
by the surface area of the loaded CeO2 sample. Quantity of desorbed NH3 calculated using the area under the desorption profile. (2-COLUMN).

at 3711 cm1 which is the characteristic vibration of terminal Ce- 1246, 1053 cm1) and monodentate (1584, 1405, 1053 cm1) car-
OH and the stretching mode vibration of the bicarbonate at bonates contribute to most of the remaining features found in
3629 cm1. CeO2-O exhibit a smaller shoulder centered around CeO2-C’s spectra. Bridged carbonates appear to desorb at a similar
1627 cm1 which may be attributed to bicarbonate, but the hydro- temperature as the bicarbonate species with their features largely
xyl group region is largely unchanged during CO2 adsorption sug- disappearing above 200 °C while monodentate species remain up
gesting that CeO2-O’s hydroxyls are less basic than those on to 300 °C. Bidentate carbonate (1584 and 1293 cm1) formation
CeO2-C. occurs upon CO2 reacting with a Lewis acid-base pair such as a
There are numerous peaks assigned to bidentate, monodentate coordinatively unsaturated surface Ce cation and O anion. These
and bridged carbonates between 1000 and 1700 cm1 in the spec- features are relatively weak on the cubes but stable up to 300 °C.
tra of each nanoshape. This confirms the presence of surface oxy- In contrast with CeO2-C, CeO2-O and CeO2-Com spectra are
gen with Lewis base behavior on all three nanoshapes though dominated by bidentate carbonate peaks (1575, 1295,
there are clear surface dependent differences in the relative distri- 1023 cm1) with no clear features associated with neither bridged
bution of these carbonates between the (1 0 0) dominant CeO2-C nor monodentate carbonates. Although the bidentate carbonate
and (1 1 1) dominant CeO2-O and CeO2-Com. Bridged (1695, 1465, features are much stronger at low temperatures on the (1 1 1) dom-
inant nanoshapes, they are completely desorbed by 200 °C which
suggests that the Lewis acid-base pairs on this facet are weaker
than those on CeO2-C’s (1 0 0) facet. This is consistent with the sur-
face Ce cations and O anions being more coordinatively unsatu-
rated on the (1 0 0) facet.
Oxygen anions act as Lewis base sites while the cerium cations
are Lewis acid sites, but the surfaces of metal oxides nanoparticles
are hydroxylated under normal conditions, creating a population of
amphoteric hydroxyl groups on the surface. The acidity or basicity
of these hydroxyl groups depends on the metal oxide and the coor-
dination of the OH groups [7,42,64]. Highly coordinated OH groups
are more strongly polarized by the cations which loosens the bond
to hydrogen, making it easier to deprotonate, thus having acidic
characteristics. Single coordinated OH groups have a stronger ten-
dency to dissociate as OH– ions, making them easier to exchange
for other anions, thus having basic characteristics. Fig. 10a displays
the IR spectra of the OH stretching region for the CeO2 nanoshapes.
Bands around 3709 cm1 are assigned to terminal OH and bridged
OH appear around 3637 cm1. CeO2-Com and CeO2-O are domi-
nated by bridged OH which is consistent with theoretical studies
suggesting bridged OH are most stable on the CeO2(1 1 1) surface
[65]. There is a minor band associated with terminal OH groups
on CeO2-Com which may be a result of the more heterogenous nat-
Fig. 7. Pyridine desorption peak temperatures (Tm) during Py-TPD under varying ure of these nanoparticles. CeO2-O have the weakest band associ-
temperature ramp rates (b). (SINGLE COLUMN). ated with terminal OH whose limited presence may be due to
102
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

Fig. 8. Acid site density as measured using NH3-TPD of CeO2-O and CeO2-C calcined at different temperatures as a function of (a) calcination temperature and (b) defect
density as measured in Raman. (2-COLUMN).

Table 2
General assignment of IR bands from CO2 adsorption and TPD on CeO2-C, CeO2-O, and CeO2-Com.

Species v(CO3) d(OH) v(OH)


CeO2-C bicarbonate 1635, 1405 1208 3629
bridged carbonate 1695, 1465, 1246, 1053
monodentate carbonate 1584, 1405, 1053
bidentate carbonate 1584, 1293, 1023
CeO2-O CeO2-Com bicarbonate 1627
bidentate carbonate 1575, 1295, 1023

slight variations in the surface structure at the edges and corners of mation upon CO2 reacting with basic OH primarily occurring on
the nanoparticles. CeO2-C are predominately populated with the CeO2-C. The broad band at 3473 cm1 on CeO2-C and CeO2-O is
more basic terminal OH with only a minor presence of bridged likely due to cerium oxyhydroxide impurities [66] which appear
OH groups. These results are consistent with the bicarbonate for- to be largely inactive or inaccessible to both acidic (e.g., CO2) and

Fig. 9. DRIFTS spectra on (a) CeO2-C and (b) CeO2-O after CO2 adsorption at room temperature and purged under He flow at increased desorption temperatures. (2-COLUMN).

103
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

Fig. 10. (a) IR spectra of hydroxyl groups on CeO2-Com, CeO2-C, and CeO2-O. (b) QMS profiles of desorbed CO2 (m/z = 44) from CeO2-Com, CeO2-C, and CeO2-O normalized by
surface area of the loaded CeO2 sample. (2-COLUMN).

basic (e.g., pyridine) probe molecules as suggested by the minimal a referenced for their impact on the acid-base sensitive 2-propanol
changes to the spectra in this region after adsorption (Figure S1 probe reaction detailed in the following section.
and Figure S4).
Having identified the basic site types using CO2 DRIFTS, CO2- 3.4. 2-Propanol conversion
TPD, shown in Fig. 10b, was used to characterize the overall
strength and quantity of the basic sites. CO2 was adsorbed at ambi- The 2-propanol transformations were carried out under He flow
ent temperature and purged under He flow for 2 h to remove phy- and constant 2-propanol partial pressure. Pure He was chosen
sisorbed CO2. Each CeO2 nanoshape immediately begins to desorb because, although 2-propanol conversion is more active and stable
CO2 upon increasing the temperature above ambient temperature. under oxidizing conditions, only acetone is produced at low tem-
The (1 1 1) dominant CeO2-Com and CeO2-O have a single, low tem- peratures and complete oxidation to CO2 occurs at elevated tem-
perature desorption peak at 98 °C. CeO2-C demonstrate a broader peratures [9]. Example time-on-stream plots for 2-propanol
desorption profile from 50 °C to 250 °C with a low temperature conversion under He over each sample can be seen in the support-
desorption peak around 100 °C and a high temperature desorption ing information (Fig. S5-7). Each CeO2 sample displayed 100%
peak at approximately 170 °C. The higher CO2 desorption temper- combined selectivity to the dehydrogenation product, acetone, or
atures seen from CeO2-C may be attributed to monodentate and the dehydration product, propene. Only trace amount of decompo-
bidentate carbonates species whose features are visible in the IR sition products (e.g., CO, CO2, CH4) were observed. Overall conver-
spectra up to 300 °C. The increased decomposition temperatures sion of 2-propanol under these conditions is kept below 5% to
of monodentate and bidentate carbonates over bicarbonate and ensure the reaction remains within the kinetic region.
bridged carbonate species are consistent with previous investiga- During the first 30 min of 2-propanol conversion over each
tions of CO2 desorption from CeO2 materials [33,67]. Integration sample there is a decrease in acetone rates and a slight increase
of the CO2 desorption peak areas allows for the quantification of in propene rates. 2-propanol has been shown to act as a reductant
the total number of base sites which is summarized in Table 3. under UHP conditions, this is particularly evident in 2-propanol
CeO2-C possess the highest density of base sites (0.57 CO2/nm2), TPD on CeO2(1 0 0) films on which the acetone production depletes
followed by CeO2-Com (0.54 CO2/nm2), with CeO2-O having the available surface O [68]. On both CeO2(1 1 1) and CeO2(1 0 0), sur-
lowest (0.45 CO2/nm2). face reduction decreases the amount of acetone formed in favor
To summarize the acid-base properties measured: (1 1 1) domi- of propene production during 2-propanol TPD [69]. In the current
nant CeO2-O and CeO2-Com have a higher density of stronger acid studies, the initial decrease in acetone formation and the corre-
sites while the (1 0 0) dominant CeO2-C have considerably stronger sponding increase in propene may be explained by a partial reduc-
base sites with a comparable site density. These properties will be tion of surface CeO2 catalysts by 2-propanol. The effects of
reduction are most clearly seen on the CeO2-C samples which typ-
ically displays greater than 60% decrease in acetone formation
rates during the first 30 min. In contrast, acetone formation rates
on CeO2-O and CeO2-Com only decrease by approximately 30%.
Table 3
Summary of the acid-base properties of CeO2-Com, CeO2-O, and CeO2-C.
This is consistent with UHP studies demonstrating that 2-
propanol acts as a reductant during 2-propanol TPD on
Sample Acid Site Density Py Desorption Base Site Density CeO2(1 0 0) but is a poor reductant of CeO2(1 1 1) [69]. The presence
(NH3/nm2) Energy (KJ/mol) (CO2/nm2)
of surface defects on CeO2-O increases its reducibility in compar-
CeO2-Com 1.13 141 0.54 ison to pristine CeO2(1 1 1) single crystal films [22]. To minimize
CeO2-C 0.60 93 0.57
CeO2-O 0.94 132 0.45
potential complications from deactivation and differences in the
extents of reduction dependent on the nanoshape and reaction
104
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

Fig. 11. (a) Acetone and propene production rate normalized by sample surface area during 2-propanol conversion over CeO2-Com, CeO2-O, and CeO2-C (a) normalized by
sample surface area and (b) normalized by number of base sites and acid sites as determined by CO2 and NH3 TPD, respectively. 6 kPa 2-propanol feed at 270 °C,
WHSV = 7.5 g2-propanol / hr / gCeO2. (2-COLUMN).

temperature, this study evaluates the initial reaction rates by the basic product at higher rates which suggests that the observed
extrapolating the experimental results within the first 20 min of differences in the strengths of these sites play a key role. The rates
time-on-stream. Conducting the reaction under He flow causes for dehydration of 2-propanol suggest the opposite trend, in that
the samples to reduce and deactivate over time but highlights CeO2-O and CeO2-Com have rates of 0.0062 and 0.0049 lmol-
surface-dependent dehydration and dehydrogenation rates. Note propene/m2/s, respectively, while CeO2-C produce much less pro-
that TEM analysis of the CeO2 samples after 2-propanol conversion pene at 0.0003 lmol-propene/m2/s. The observed difference in
up to 450 °C shows no observable morphological changes propene production between the (1 1 1) dominant CeO2-O and
(Figure S8). CeO2-Com and the (1 0 0) dominant CeO2-C is consistent with the
The initial production rates for acetone and propene during higher density and strength of acid sites, as measured by NH3-
6 kPa 2-propanol feed at 270 °C are shown in Fig. 11a. CeO2-O TPD, for the (1 1 1) dominant nanocrystals.
and CeO2-Com have dehydrogenation rates of 0.0093 and To clarify whether the strength of the acid and base sites plays a
0.0095 lmol-acetone/m2/s, respectively, while CeO2-C exhibit a role in 2-propanol dehydrogenation and dehydration rates, or just
greater rate of acetone production at 0.0141 lmol-acetone/m2/s. the differences in site density, the initial acetone and propene pro-
So, despite the observation that the number of basic sites is similar duction rates normalized by the number of base and acid sites,
between the three CeO2 nanocrystals, CeO2-C are able to produce respectively, are plotted in Fig. 11b. Dehydrogenation to acetone

Fig. 12. Arrhenius plots of 2-propanol dehydrogenation and dehydration on CeO2-C ( ), CeO2-Com ( ), and CeO2-O (▲) CeO2 normalized by surface area. 6 kPa 2-propanol,
250 – 290 °C, WHSV = 7.5 g2-propanolh1g-1
CeO2 (2-COLUMN).

105
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

Table 4
Summary of initial 2-propanol dehydrogenation and dehydration production rates. Apparent activation energies (Ea) and preexponential factors (A) for 2-propanol
dehydrogenation and dehydration.

Sample Dehydrogenation Rate Dehydration Rate Dehydrog. Ea Dehydrat. Ea Dehydrog. A (s1) Dehydrat. A (s1)
(lmol-acetone/m2/s) (lmol-propene/m2/s) (kJ/mol) (kJ/mol)
Commercial 0.0095 0.0049 69 164 4.3x104 5.7x1012
Octahedra 0.0093 0.0062 74 170 1.1x105 1.3x1014
Cubes 0.0141 0.0003 47 76 4.7x102 2.2x104

over the CeO2-C remains the highest at 0.0025 acetone/base-site/s Acknowledgements


while CeO2-O and CeO2-Com produce 0.0021 and 0.0020 acetone/
base-site/s, respectively. This suggests that the stronger basic sites This work was supported by the U.S. Department of Energy,
of (1 0 0) dominant CeO2-C compared to those of the (1 1 1) domi- Office of Science, Office of Basic Energy Sciences, Division of Chem-
nant CeO2-O and CeO2-Com are responsible for increased rate of ical Sciences, Geosciences, and Biosciences’ Catalysis Program
dehydrogenation. Similarly, dehydration to propene remains great- (FWP-47319).
est over the (1 1 1) dominant CeO2-O and CeO2-Com, producing
0.00066 and 0.00047 propene/acid-site/s, respectively, while
CeO2-C only produce 0.00019 propene/acid-site/s. This is due to Appendix A. Supplementary data
increased strength of the acid sites on CeO2-O and CeO2-Com as
observed in their higher temperature desorption profiles during Supplementary data to this article can be found online at
NH3-TPD and the larger calculated pyridine desorption energy. https://doi.org/10.1016/j.jcat.2021.09.009.
Arrhenius plots for each sample’s initial dehydrogenation and
dehydration rates are shown in Fig. 12. The apparent activation
energies for dehydrogenation to acetone were calculated to be References
47, 74, and 69 kJ/mol for CeO2-C, CeO2-O, and CeO2-Com, respec-
[1] K.-I. Tanaka, A. Ozaki, Acid-Base Properties and Catalytic Activity of Solid
tively. The apparent activation energies for dehydration to propene Surfaces, J. Catal. 8 (1967) 1–7, https://doi.org/10.1016/0021-9517(67)90274-
were calculated to be 76, 170, and 164 kJ/mol for CeO2-C, CeO2-O, 6.
[2] K. Tanabe, M. Misono, H. Hattori, New Solid Acids and Bases, Their Catalytic
and CeO2-Com, respectively. The apparent activation energy for
Properties (1990). https://www.sciencedirect.com/bookseries/studies-in-
dehydrogenation to acetone is lower for all three CeO2 nanocrys- surface-science-and-catalysis/vol/51.
tals than that for dehydration to propene. These activation energy [3] K. Tanabe, H. Hattori, in: G. Ertl, H. Knoezinger, J. Weitkamp (Eds.) Handbook of
differences and the overall increased production rate for acetone Heterogeneous Catalysis, Wiley-VCH, Weinheim, 1997, pp. 404-412.
https://doi.org/10.1002/9783527619474.
over propene are in agreement with the consensus from 2- [4] J.-M. Basset, B.C. Gates, J.-P. Candy, A. Choplin, M. Leconte, F. Quignard, C.
propanol studies on metal oxides with varying acid-base proper- Santini (Eds.), Surface Organometallic Chemistry: Molecular Approaches to
ties which demonstrate that CeO2 is a more basic metal oxide Surface Catalysis, Springer Netherlands, Dordrecht, 1988.
[5] M.I Zaki, M.A Hasan, F.A Al-Sagheer, L. Pasupulety, In situ FTIR spectra of
[9,46]. Interestingly, while CeO2-C produce much less propene than pyridine adsorbed on SiO2–Al2O3, TiO2, ZrO2 and CeO2: general
CeO2-O or CeO2-Com, the activation energy for dehydration over considerations for the identification of acid sites on surfaces of finely divided
CeO2-C is significantly less than its (1 1 1) dominant counterparts. metal oxides, Colloids Surf. Physicochem. Eng. Aspects 190 (3) (2001) 261–
274, https://doi.org/10.1016/S0927-7757(01)00690-2.
This suggests that the decreased density of acid sites on CeO2-C [6] H.P. Boehm, H. Knözinger, in: J.A. Anderson, M. Boudart (Eds.) Catal. Sci.
plays the largest factor in its rate of dehydration, which is demon- Technol., Springer, Berlin, 1983, pp. 40-189. https://doi.org/10.1007/978-3-
strated by a calculated pre-exponential factor ten orders of magni- 642-93229-8.
[7] H.P. Boehm, Acidic and basic properties of hydroxylated metal oxide surfaces,
tude lower, as summarized in Table 4.
Discuss. Faraday Soc. 52 (1971) 264–275, https://doi.org/10.1039/
DF9715200264.
[8] M.I. Zaki, M.A. Hasan, L. Pasupulety, Surface Reactions of Acetone on Al2O3,
4. Conclusions TiO2, ZrO2, and CeO2: IR Spectroscopic Assessment of Impacts of the Surface
AcidBase Properties, Langmuir 17 (2001) 768–774, https://doi.org/10.1021/
la000976p.
CeO2-C with well-defined (1 0 0) crystal facets and CeO2-O [9] D. Haffad, A. Chambellan, J.C. Lavalley, Propan-2-ol transformation on simple
exposing the (1 1 1) facet were synthesized and their acid and base metal oxides TiO2, ZrO2 and CeO2, J. Mol. Catal. A: Chem. 168 (1-2) (2001)
sites were characterized using in situ DRIFTS spectroscopy and TPD 153–164, https://doi.org/10.1016/S1381-1169(00)00516-1.
[10] C. Chizallet, M.L. Bailly, G. Costentin, H. Lauron-Pernot, J.M. Krafft, P. Bazin, J.
of basic and acidic probe molecules. CeO2-C exhibit stronger base Saussey, M. Che, Thermodynamic bronsted basicity of clean MgO surfaces
sites while CeO2-O and CeO2-Com have a higher density of stronger determined by their deprotonation ability: Role of Mg(2+)-O(2-) pairs, Catal.
acid sites. These acid-base characteristics are reflected in 2- Today 116 (2006) 196–205, https://doi.org/10.1016/j.cattod.2006.01.030.
[11] D.T. Lundie, A.R. McInroy, R. Marshall, J.M. Winfield, P. Jones, C.C. Dudman, S.F.
propanol dehydrogenation and dehydration. CeO2 samples with Parker, C. Mitchell, D. Lennon, Improved description of the surface acidity of
dominant (1 1 1) facet such as CeO2-O demonstrate greater dehy- eta-alumina, J. Phys. Chem. B 109 (2005) 11592–11601, https://doi.org/
dration activity to propene facilitated by the higher strength and 10.1021/jp0405963.
[12] M. Tamura, K.-i. Shimizu, A. Satsuma, Comprehensive IR study on acid/base
number of acid sites. The higher coordination vacancy of O anions
properties of metal oxides, Appl. Catal., A, 433–434 (2012) 135-145.
on the CeO2(1 0 0) surface increases the rate of dehydrogenation to https://doi.org/10.1016/j.apcata.2012.05.008.
acetone while the lower density of weaker acid sites and more [13] C. Binet, M. Daturi, J.-C. Lavalley, IR study of polycrystalline ceria properties in
oxidised and reduced states, Catal. Today 50 (2) (1999) 207–225, https://doi.
basic surface hydroxyls limit its dehydration activity.
org/10.1016/S0920-5861(98)00504-5.
[14] N. Sheppard, in: R.F. Willis (Ed.) Vibrational Spectroscopy of Adsorbates,
Springer, Berlin, 1980, pp. 165-176. https://doi.org/10.1007/978-3-642-
Declaration of Competing Interest 88644-7.
[15] J.B. Peri, in: J.R. Anderson, M. Boudart (Eds.) Catal. Sci. Technol., Springer,
The authors declare that they have no known competing finan- Berlin, 1984. https://doi.org/10.1007/978-3-642-93247-2
[16] A.A. Davydov, in: C.H. Rochester (Ed.) Infrared Spectroscopy of Adsorbed
cial interests or personal relationships that could have appeared Species on the Surface of Transition Metal Oxides, Wiley, Chichester, 1990.
to influence the work reported in this paper. https://doi.org/10.1002/sia.740170510.

106
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

[17] A.T. Bell, Methods of Surface Characterization, in: J. Yates, T.E. Madey (Eds.), [41] M. Nolan, Hybrid density functional theory description of oxygen vacancies in
Vibrational Spectroscopy of Molecultes on Surfaces, Plenum Press, New York, the CeO2 (110) and (100) surfaces, Chem. Phys. Lett. 499 (1-3) (2010) 126–
1987, pp. 105–133, https://doi.org/10.1007/978-1-4684-8759-6. 130, https://doi.org/10.1016/j.cplett.2010.09.016.
[18] E. Mamontov, T. Egami, R. Brezny, M. Koranne, S. Tyagi, Lattice Defects and [42] M.I. Zaki, N. Sheppard, An Infrared Spectroscopic Study of the Adsorption and
Oxygen Storage Capacity of Nanocrystalline Ceria and Ceria-Zirconia, J. Phys. Mechanism of Surface-Reactions of 2-propanol on Ceria, J. Catal. 80 (1983)
Chem. B 104 (47) (2000) 11110–11116, https://doi.org/10.1021/jp0023011. 114–122, https://doi.org/10.1016/0021-9517(83)90235-X.
[19] P. Dutta, S. Pal, M.S. Seehra, Y. Shi, E.M. Eyring, R.D. Ernst, Concentration of Ce3 [43] G.A.M. Hussein, N. Sheppard, M.I. Zaki, R.B. Fahim, Infrared spectroscopic
+ and Oxygen Vacancies in Cerium Oxide Nanoparticles, Chem. Mater. 18 (21) studies of the reactions of alcohols over group IVB metal oxide catalysts. Part
(2006) 5144–5146, https://doi.org/10.1021/cm061580n. 1.-Propan-2-ol over TiO2, ZrO2 and HfO2, J. Chem. Soc., Faraday Trans. 1, 85
[20] Y. Madier, C. Descorme, A.M. Le Govic, D. Duprez, Oxygen Mobility in CeO2 and (1989) 1723-1741. http://dx.doi.org/10.1039/F19898501723.
CexZr(1–x)O2 Compounds: Study by CO Transient Oxidation and 18O/16O [44] W. Turek, A. Krowiak, Evaluation of oxide catalysts’ properties based on
Isotopic Exchange, J. Phys. Chem. B 103 (1999) 10999–11006, https://doi.org/ isopropyl alcohol conversion, Appl. Catal., A, 417–418 (2012) 102-110.
10.1021/jp991270a. https://doi.org/10.1016/j.apcata.2011.12.030.
[21] Z. Wu, M. Li, J. Howe, H.M. Meyer, S.H. Overbury, Probing Defect Sites on CeO2 [45] S.A. Fuente, C.A. Ferretti, N.F. Domancich, V.K. Díez, C.R. Apesteguía, J.I. Di
Nanocrystals with Well-Defined Surface Planes by Raman Spectroscopy and O- Cosimo, R.M. Ferullo, N.J. Castellani, Adsorption of 2-propanol on MgO surface:
2 Adsorption, Langmuir 26 (21) (2010) 16595–16606, https://doi.org/10.1021/ A combined experimental and theoretical study, Appl. Surf. Sci. 327 (2015)
la101723w. 268–276, https://doi.org/10.1016/j.apsusc.2014.11.159.
[22] Y. Li, Z. Wei, F. Gao, L. Kovarik, R.A.L. Baylon, C.H.F. Peden, Y. Wang, Effect of [46] D. Kulkarni, I.E. Wachs, Isopropanol oxidation by pure metal oxide catalysts:
Oxygen Defects on the Catalytic Performance of VOx/CeO2 Catalysts for number of active surface sites and turnover frequencies, Appl. Catal., A 237 (1-
Oxidative Dehydrogenation of Methanol, ACS Catalysis 5 (5) (2015) 3006– 2) (2002) 121–137, https://doi.org/10.1016/S0926-860X(02)00325-3.
3012, https://doi.org/10.1021/cs502084g. [47] A.K.P. Mann, Z. Wu, F.C. Calaza, S.H. Overbury, Adsorption and Reaction of
[23] Y. Wang, F. Wang, Q. Song, Q. Xin, S. Xu, J. Xu, Heterogeneous Ceria Catalyst Acetaldehyde on Shape-Controlled CeO2 Nanocrystals: Elucidation of
with Water-Tolerant Lewis Acidic Sites for One-Pot Synthesis of 1,3-Diols via Structure-Function Relationships, ACS Catalysis 4 (8) (2014) 2437–2448,
Prins Condensation and Hydrolysis Reactions, J. Am. Chem. Soc. 135 (4) (2013) https://doi.org/10.1021/cs500611g.
1506–1515, https://doi.org/10.1021/ja310498c. [48] S. Agarwal, L. Lefferts, B.L. Mojet, Ceria Nanocatalysts: Shape Dependent
[24] H. Metiu, S. Chrétien, Z. Hu, B. Li, X. Sun, Chemistry of Lewis Acid-Base Pairs on Reactivity and Formation of OH, Chemcatchem 5 (2) (2013) 479–489, https://
Oxide Surfaces, J. Phys. Chem. C 116 (19) (2012) 10439–10450, https://doi.org/ doi.org/10.1002/cctc.201200491.
10.1021/jp301341t. [49] S. Agarwal, X. Zhu, E.J.M. Hensen, B.L. Mojet, L. Lefferts, Surface-Dependence of
[25] S. Zhang, Z.Q. Huang, Y.Y. Ma, W. Gao, J. Li, F.X. Cao, L. Li, C.R. Chang, Y.Q. Qu, Defect Chemistry of Nanostructured Ceria, J. Phys. Chem. C 119 (22) (2015)
Solid frustrated-Lewis-pair catalysts constructed by regulations on surface 12423–12433, https://doi.org/10.1021/acs.jpcc.5b02389.
defects of porous nanorods of CeO2, Nat. Commun. 8 (2017) 11, https://doi. [50] T. Taniguchi, T. Watanabe, N. Sugiyama, A.K. Subramani, H. Wagata, N.
org/10.1038/ncomms15266. Matsushita, M. Yoshimura, Identifying Defects in Ceria-Based Nanocrystals by
[26] D.R. Mullins, P.M. Albrecht, T.-L. Chen, F.C. Calaza, M.D. Biegalski, H.M. UV Resonance Raman Spectroscopy, J. Phys. Chem. C 113 (46) (2009) 19789–
Christen, S.H. Overbury, Water Dissociation on CeO2(100) and CeO2(111) Thin 19793, https://doi.org/10.1021/jp9049457.
Films, J. Phys. Chem. C 116 (36) (2012) 19419–19428, https://doi.org/10.1021/ [51] J.E. Spanier, R.D. Robinson, F. Zheng, S.W. Chan, I.P. Herman, Size-dependent
jp306444h. properties of CeO2-y nanoparticles as studied by Raman scattering, Phys. Rev.
[27] D.R. Mullins, S.D. Senanayake, T.-L. Chen, Adsorption and Reaction of C1-C-3 B 64 (2001) 8, https://doi.org/10.1103/PhysRevB.64.245407.
Alcohols over CeOx(111) Thin Films, J. Phys. Chem. C 114 (40) (2010) 17112– [52] Y. Li, Z.H. Wei, F. Gao, L. Kovarik, C.H.F. Peden, Y. Wang, Effects of CeO2 support
17119, https://doi.org/10.1021/jp103905e. facets on VOx/CeO2 catalysts in oxidative dehydrogenation of methanol, J.
[28] D.R. Mullins, M.D. Robbins, J. Zhou, Adsorption and reaction of methanol on Catal. 315 (2014) 15–24, https://doi.org/10.1016/j.jcat.2014.04.013.
thin-film cerium oxide, Surf. Sci. 600 (7) (2006) 1547–1558, https://doi.org/ [53] F. Bonino, A. Damin, S. Bordiga, C. Lamberti, A. Zecchina, Interaction of CD3CN
10.1016/j.susc.2006.02.011. and Pyridine with the Ti(IV) Centers of TS-1 Catalysts: a Spectroscopic and
[29] F. Lin, Y. Chen, L. Zhang, D. Mei, L. Kovarik, B. Sudduth, H. Wang, F. Gao, Y. Computational Study, Langmuir 19 (2003) 2155–2161, https://doi.org/
Wang, Single-Facet Dominant Anatase TiO2 (101) and (001) Model Catalysts to 10.1021/la0262194.
Elucidate the Active Sites for Alkanol Dehydration, ACS Catalysis 10 (7) (2020) [54] D. Haffad, A. Chambellan, J.C. Lavalley, Characterisation of acid-treated
4268–4279, https://doi.org/10.1021/acscatal.9b0465410.1021/ bentonite. Reactivity, FTIR study and Al-27 MAS NMR, Catal. Lett. 54 (1998)
acscatal.9b04654.s001. 227–233, https://doi.org/10.1023/A:1019064930150.
[30] H.-X. Mai, L.-D. Sun, Y.-W. Zhang, Rui Si, W. Feng, H.-P. Zhang, H.-C. Liu, C.-H. [55] A. Travert, A. Vimont, J.C. Lavalley, V. Montouillout, M.R. Delgado, J.J.C. Pascual,
Yan, Shape-selective synthesis and oxygen storage behavior C.O. Arean, Evidence for discrepancy between the surface Lewis
of ceria nanopolyhedra, nanorods, and nanocubes, J. Phys. Chem. B 109 (51) acid site strength and infrared spectra of adsorbed molecules: The case of
(2005) 24380–24385, https://doi.org/10.1021/jp055584b10.1021/jp055584b. boria-silica, J. Phys. Chem. B 108 (2004) 16499–16507, https://doi.org/
s001. 10.1021/jp0479365.
[31] K ZHOU, X WANG, X SUN, Q PENG, Y LI, Enhanced catalytic activity of ceria [56] S. Wang, L. Zhao, W. Wang, Y. Zhao, G. Zhang, X. Ma, J. Gong, Morphology
nanorods from well-defined reactive crystal planes, J. Catal. 229 (1) (2005) control of ceria nanocrystals for catalytic conversion of CO2 with methanol,
206–212, https://doi.org/10.1016/j.jcat.2004.11.004. Nanoscale 5 (2013) 5582–5588, https://doi.org/10.1039/C3NR00831B.
[32] H. Tan, J. Wang, S. Yu, K. Zhou, Support Morphology-Dependent Catalytic [57] C. Lahousse, F. Maugé, J. Bachelier, J.-C. Lavalley, Acidic and basic properties of
Activity of Pd/CeO2 for Formaldehyde Oxidation, Environ. Sci. Technol. 49 (14) titania-alumina mixed oxides; active sites for propan-2-ol dehydration, J.
(2015) 8675–8682, https://doi.org/10.1021/acs.est.5b01264. Chem. Soc., Faraday Trans. 91 (17) (1995) 2907–2912, https://doi.org/10.1039/
[33] Z. Wu, A.K.P. Mann, M. Li, S.H. Overbury, Spectroscopic Investigation of FT9959102907.
Surface-Dependent Acid Base Property of Ceria Nanoshapes, J. Phys. Chem. C [58] C.D. Baertsch, K.T. Komala, Y.-H. Chua, E. Iglesia, Genesis of Brønsted Acid Sites
119 (13) (2015) 7340–7350, https://doi.org/10.1021/acs.jpcc.5b00859. during Dehydration of 2-Butanol on Tungsten Oxide Catalysts, J. Catal. 205 (1)
[34] M. Nolan, S. Grigoleit, D.C. Sayle, S.C. Parker, G.W. Watson, Density functional (2002) 44–57, https://doi.org/10.1006/jcat.2001.3426.
theory studies of the structure and electronic structure of pure and defective [59] R. Vanselow, R. Howe, Chemistry and Physics of Solid Surfaces V (1984),
low index surfaces of ceria, Surf. Sci. 576 (1-3) (2005) 217–229, https://doi. https://doi.org/10.1007/978-3-642-82253-7.
org/10.1016/j.susc.2004.12.016. [60] S. Chen, T. Cao, Y. Gao, D. Li, F. Xiong, W. Huang, Probing Surface Structures of
[35] H. Nörenberg, G.A.D. Briggs, Surface structure of CeO2(111) studied by low CeO2, TiO2, and Cu2O Nanocrystals with CO and CO2 Chemisorption, J. Phys.
current STM and electron diffraction, Surf. Sci. 402–404 (1998) 734–737, Chem. C 120 (38) (2016) 21472–21485, https://doi.org/10.1021/acs.
https://doi.org/10.1016/S0039-6028(97)00999-0. jpcc.6b0615810.1021/acs.jpcc.6b06158.s001.
[36] D.R. Mullins, P.V. Radulovic, S.H. Overbury, Ordered cerium oxide thin films [61] Z. Cui, J. Gan, J. Fan, Y. Xue, R. Zhang, Size-Dependent Surface Basicity
grown on Ru(0001) and Ni(111), Surf. Sci. 429 (1-3) (1999) 186–198, https:// of Nano-CeO2 and Desorption Kinetics of CO2 on Its Surface, Ind. Eng.
doi.org/10.1016/S0039-6028(99)00369-6. Chem. Res. 57 (32) (2018) 10977–10984, https://doi.org/10.1021/acs.
[37] J.C. Conesa, Computer Modeling of Surfaces and Defects on Cerium Dioxide, iecr.8b01247.
Surf. Sci. 339 (3) (1995) 337–352, https://doi.org/10.1016/0039-6028(95) [62] K.R. Hahn, M. Iannuzzi, A.P. Seitsonen, J. Hutter, Coverage Effect of the CO2
00595-1. Adsorption Mechanisms on CeO2(111) by First Principles Analysis, J. Phys.
[38] H. Nörenberg, J.H. Harding, The surface structure of CeO2(001) single crystals Chem. C 117 (4) (2013) 1701–1711, https://doi.org/10.1021/jp309565u.
studied by elevated temperature STM, Surf. Sci. 477 (1) (2001) 17–24, https:// [63] S. Agarwal, X. Zhu, E.J.M. Hensen, L. Lefferts, B.L. Mojet, Defect Chemistry of
doi.org/10.1016/S0039-6028(01)00700-2. Ceria Nanorods, J. Phys. Chem. C 118 (8) (2014) 4131–4142, https://doi.org/
[39] Y. Lin, Z. Wu, J. Wen, K.R. Poeppelmeier, L.D. Marks, Imaging the Atomic 10.1021/jp409989y.
Surface Structures of CeO2 Nanoparticles, Nano Lett. 14 (1) (2014) 191–196, [64] M.I. Zaki, H. Knözinger, Carbon monoxide — A low temperature infrared probe
https://doi.org/10.1021/nl403713b. for the characterization of hydroxyl group properties on metal oxide surfaces,
[40] M. Bugnet, S.H. Overbury, Z.L. Wu, T. Epicier, Direct Visualization and Control Mater. Chem. Phys. 17 (1-2) (1987) 201–215, https://doi.org/10.1016/0254-
of Atomic Mobility at 100 Surfaces of Ceria in the Environmental Transmission 0584(87)90056-3.
Electron Microscope, Nano Lett. 17 (12) (2017) 7652–7658, https://doi.org/ [65] G. Vicario, G. Balducci, S. Fabris, S. de Gironcoli, S. Baroni, Interaction of
10.1021/acs.nanolett.7b0368010.1021/acs.nanolett.7b03680.s00110.1021/ Hydrogen with Cerium Oxide Surfaces: a Quantum Mechanical Computational
acs.nanolett.7b03680.s00210.1021/acs.nanolett.7b03680.s00310.1021/acs. Study, J. Phys. Chem. B 110 (2006) 19380–19385, https://doi.org/10.1021/
nanolett.7b03680.s004. jp061375v.

107
B. Sudduth, D. Yun, J. Sun et al. Journal of Catalysis 404 (2021) 96–108

[66] A. Badri, C. Binet, J.-C. Lavalley, An FTIR study of surface ceria hydroxy groups [68] D.R. Mullins, The surface chemistry of cerium oxide, Surf. Sci. Rep. 70 (1)
during a redox process with H2, J. Chem. Soc., Faraday Trans. 92 (1996) 4669– (2015) 42–85, https://doi.org/10.1016/j.surfrep.2014.12.001.
4673, https://doi.org/10.1039/FT9969204669. [69] D.R. Mullins, P.M. Albrecht, F. Calaza, Variations in Reactivity on Different
[67] G. Finos, S. Collins, G. Blanco, E. del Rio, J.M. Cíes, S. Bernal, A. Bonivardi, Crystallographic Orientations of Cerium Oxide, Top. Catal. 56 (15-17) (2013)
Infrared spectroscopic study of carbon dioxide adsorption on the surface of 1345–1362, https://doi.org/10.1007/s11244-013-0146-7.
cerium–gallium mixed oxides, Catal. Today 180 (1) (2012) 9–18, https://doi.
org/10.1016/j.cattod.2011.04.054.

108

You might also like