You are on page 1of 12

Article

Cite This: Inorg. Chem. 2019, 58, 6016−6027 pubs.acs.org/IC

Surface, Subsurface, and Bulk Oxygen Vacancies Quantified by


Decoupling and Deconvolution of the Defect Structure of Redox-
Active Nanoceria
Rashid Mehmood,* Sajjad S. Mofarah, Wen-Fan Chen, Pramod Koshy, and Charles C. Sorrell
School of Materials Science and Engineering, Faculty of Science, UNSW Sydney, Sydney, NSW 2052, Australia
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Oxygen vacancy concentrations are critical to the redox/


Downloaded via UNIV POLITECNICA DE CATALUNYA on June 21, 2023 at 19:08:11 (UTC).

photocatalytic performance of nanoceria, but their direct analysis is


problematic under controlled atmospheres but essentially impossible under
aqueous conditions. The present work provides three novel approaches to
analyze these data from XPS data for the three main morphologies of
nanoceria synthesized under aqueous conditions and tested using in vacuo
analytical conditions. First, the total oxygen vacancy concentrations are
decoupled quantitatively into surface-filled, subsurface-unfilled, and bulk
values. Second, the relative surface areas are calculated for all exposed
crystallographic planes. Third, XPS and redox performance data are
deconvoluted according to the relative surface areas of these planes.
Correlations based on two independent empirical results from volumetric
surface XPS, combined with sequential deep XPS and independent EELS
data, confirm that these approaches provide quantitative determinations of the different oxygen vacancy concentrations.
Critically, the redox/photocatalytic performance depends not on the total oxygen vacancy concentration but on the
concentration of the active sites on each plane in the form of subsurface-unfilled oxygen vacancies. This is verified by the pH-
dependent performance, which can be increased significantly by exposing these vacancies to the surroundings. These
approaches have significance to the design and engineering of semiconducting materials exposed to the environment.

1. INTRODUCTION potential of the Ce3+ ↔ Ce4+ switching10 establishes a dynamic


Nanoceria (cerium oxide nanoparticles, CeO2‑x) recently has equilibrium of transient charge compensation between the two
attracted attention owing to its redox/photocatalytic, catalytic, valences, where the VÖ provides the driving force for the
and acid−base properties for energy, environmental, and oxidation of adsorbed molecules, thereby establishing a weak
biomedical applications. The potential performance of nano- electrostatic bond; the low reduction potential for Ce3+ ↔ Ce4+
ceria is considered widely to be associated with the intrinsic switching provides the driving force for the reverse reduction. In
defects Ce3+ and charge-compensating oxygen vacancies (VÖ ) effect, there is an equilibrium constant that effectively fixes the
present at the surface or subsurface of the nanoparticles, where Ce3+ concentration ([Ce3+]), even when the VÖ is occupied
continuous reversible Ce3+ ↔ Ce4+ switching and associated transiently by adsorbed molecules.
changes in oxygen vacancy concentration ([VÖ ]) occur; changes The physicochemical, structural, and microstructural features
in the pH in aqueous media are known to be one of the means of of nanoceria play an important role in determining the locations
initiating this switching.1−9 The insuperable challenge to and concentration of these VÖ . Consequently, the morpholo-
researchers is that defect equilibria for semiconductors such as gies, exposed surface planes, surface areas, and grain sizes are
CeO2‑x typically are described for anhydrous and high- critical, and these depend principally on the synthesis
temperature conditions. Furthermore, there does not appear method.11−20 Synthesis of nanoparticles by methods involving
to be any analytical instrumentation capable of directly assessing high temperatures not only reduces the concentrations of
defect equilibria under hydrated (water vapor) or aqueous surface vacancies but also produces hard agglomerates, which
(liquid water) conditions. Consequently, the present work cause discrepancies in the measurement of surface areas.21,22 On
presents an alternative method to assess these equilibria through the other hand, precipitation and hydrothermal methods tend to
indirect approaches of characterization of nanoceria synthesized produce monodisperse particles of higher [VÖ ]20 and soft
under aqueous conditions, the defect structure of which is agglomerates, which potentially can be dispersed by agitation.23
retained during testing using in vacuo analytical conditions in air These methods generally synthesize cubes, truncated octahedra,
and under vacuum. and square rods, all of which expose one or more of the following
Hypothetically, permanent annihilation (i.e., filling) of the VÖ
would require the formation of a covalent bond and the Received: February 4, 2019
associated electron sharing. However, the low reduction Published: April 22, 2019

© 2019 American Chemical Society 6016 DOI: 10.1021/acs.inorgchem.9b00330


Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

facets: {100}, {110}, and {111},24 which play a key role in the point, NH4OH was added using the same method until the pH
generation of reactive oxygen species (ROS) and the associated increased to ∼10.0. The suspension was aged by magnetically stirring
Ce3+ ↔ Ce4+ redox chemistry through the presence of VÖ on for 36 h at 45 °C, after which it was transferred to a 50 mL plastic
these planes.25−28 DFT calculations of the oxygen vacancy centrifuge tube and subjected to procedures identical to those applied
to the nanocubes and nanorods.
formation energies of these low-index crystallographic planes10 For all three types of nanoparticles, which are photocatalytic, storage
highlight the role of the exposed facet, suggesting that each plane between testing was done in the dark in an opaque closed container in
has a specific Ce3+ ↔ Ce4+ equilibrium constant. order to avoid light-induced reactions.13,14
Since VÖ are critical to the redox/photocatalytic performance 2.3. Characterization of Nanoceria. 2.3.1. X-ray Diffraction
of nanoceria, the concept of oxygen vacancies at the surface and (XRD). The samples were prepared for mineralogical characterization by
subsurface and in the bulk have been canvassed through hand grinding in an agate mortar and pestle and loading into aluminum
theoretical and experimental approaches,1−7,29−31 such as sample holders. The XRD powder diffraction was done using a Philips
density functional theory (DFT),2 X-ray photoelectron spec- X’Pert multipurpose X-ray diffractometer (MPD, Almelo, Nether-
lands), with Cu Kα radiation (0.15405 nm) at 20°−80° 2θ, with step
troscopy (XPS),3 synchrotron XPS,4 secondary ion mass size 0.02° 2θ and scanning speed 5.5° 2θ/min. The peaks were analyzed
spectrometry (SIMS),5 and electron energy loss spectroscopy using X’Pert High Score Plus software.
(EELS).29−31 Although EELS is capable of analyzing the oxygen The crystallite sizes and lattice strains were calculated from line
concentration ([O]) and the inferred [VÖ ] at subnanometer broadening using Scherrer’s formula and the Williamson−Hall formula,
resolution, such an analysis cannot be done under aqueous respectively).33
conditions.29,30 Furthermore, although EELS data can be used 2.3.2. Laser Raman Microspectroscopy. The nanoceria also were
to obtain lateral concentration profiles for single grains,31 data characterized mineralogically by laser Raman microspectroscopy by
for the depth concentration profiles remain unknown. using a Renishaw inVia Raman microscope (Raman; Gloucestershire,
Consequently, the present work is designed to elucidate these U.K., beam diameter 1.5 μm), which was equipped with a 35 mW
helium−neon green laser (514 nm), at 200−800 cm−1. The spectra
unknown issues by (1) decoupling the total [VÖ ] into its
were fitted and calibrated using Renishaw WiRE 4.3 software. Relative
component f illed (and thus deactivated for redox/photo- lattice strains were considered by peak shift.34
catalysis) and residual unf illed (and thus available for redox/ 2.3.3. Surface Area and Particle Size. The specific surface areas of
photocatalysis) and (2) deconvoluting the total surface analytical the nanoparticles were determined by the Brunauer−Emmett−Teller
data (XPS and surface area) into their component contributions (BET) method using a Micromeritics Tristar-3000 (Norcross, GA).
from individual exposed crystallographic planes of nanoceria Transmission electron microscopy (TEM) was used to image the
synthesized under aqueous conditions. nanoparticles directly owing to agglomeration (Supporting Information
section 1, Figure S1). Fifty individual grains of each type of
nanoparticles were selected at random for determination of the average
2. EXPERIMENTAL SECTION particle size by image analysis using ImageJ (National Institutes of
2.1. Materials. Cerium(III) nitrate hexahydrate (Ce(NO3)3·6H2O, Health, Bethesda, MD). The NC particles were of highly consistent
99.999 wt % trace metal basis), ammonium hydroxide solution size. Approximately two-thirds of the NO particles were of consistent
(NH4OH, 25 wt % in H2O), and methylene blue (dye content >82 size, with only small variation. The NR particles exhibited significantly
wt %) were obtained from Sigma-Aldrich, Australia. Hydrogen peroxide varying aspect ratios, so the length and width were determined as
(H2O2, 30 vol % in H2O, Ajex Finechem) was obtained from Thermo averages, albeit with larger variations than those of the NC and NO
Fischer Scientific, Australia. Sodium hydroxide (NaOH, 98 wt %) was particles.
obtained from Chem-Supply, Australia. 2.3.4. Transmission Electron Microscopy (TEM). Morphological
2.2. Synthesis of Nanoparticles. Nanoparticles were synthesized and structural characterization of the nanoceria particles was done by
by combination of wet-chemicals methods including precipitation and TEM using a Philips CM200 (Eindhoven, Netherlands) operating at
hydrothermal reaction.15,16 200 kV. The samples were prepared by dispersion in ethanol and
2.2.1. Nanocubes (NC) and Nanorods (NR). Nanocubes were dropping on a copper grid. TEM imaging was done using Gatan Digital
synthesized by a precipitation−hydrothermal method, where a solution Micrograph 3.9 software. Selected area electron diffraction patterns
of 30 mL of 0.9 M Ce(NO3)3·6H2O was magnetically stirred for 10 min (SAED) of the nanoparticles were done at varying magnifications in
in a 250 mL Pyrex beaker, followed by the dropwise addition of 30 mL order to assess the lattice characteristics of the nanoceria particles. The
of a 25 M NaOH solution using a 50 mL glass buret. The mixture was crystallographic planes were identified by fast Fourier transform
magnetically stirred for 50 min at room temperature. The mixture was analysis using Gatan Microscopy Suite 3 (GMS3) software. The
centrifuged at 3000 rpm for 5 min, the supernatant decanted, and the corresponding interplanar spacings were confirmed using the reported
residual solid thick slurry was redispersed magnetically in a new rinsed X-ray diffraction data for CeO2 (JCPDS 34−0394).
30 mL of 25 M NaOH solution. The mixture then was transferred to a 2.3.5. X-ray Photoelectron Spectroscopy (XPS). Surface chemical
50 mL Teflon-lined autoclave reactor, placed in an oven, heated at 5 analysis was done by XPS using a Thermo Fisher Scientific, ESCALAB
°C/min to 200 °C, soaked for 24 h, and cooled naturally. The resultant 250Xi spectrometer (Loughborough, Leicestershire, U.K.; 20 °C, 10−7
suspension was transferred to a 50 mL plastic centrifuge tube, and any Pa, 13.8 kV, 8.7 mA, monochromated Al Kα X-rays at 1487 eV, beam
residual solid was transferred following rinsing with deionized (DI) diameter 500 μm, beam exposure time 30 min). Variations in [Ce3+]
water. This suspension was centrifuged at 2500 rpm for 10 min, the and [Ce4+] were ±0.5 at. %. The data were analyzed using Thermo
supernatant decanted, the residual solid rinsed, and the suspension pH Scientific Avantage software and the peak fitting was done by using the
measured. These four steps were undertaken a total of 10 times, which standard Gaussian−Lorentzian shape function. Iterations were
was sufficient to stabilize the pH at ∼7. The final product was performed using the Marquardt method. The standard deviations
transferred to a 250 mL Pyrex beaker, the residual solid was rinsed, and between peak-fitting iterations always were <1.5%. All the peak
the suspension was dried at 70 °C for 36 h. For nanorods, the procedure positions were matched with the standard National Institute of
was identical to the above except that the autoclaving was done at 130 Standards and Technology (NIST) XPS database. In order to analyze
°C. the depth profiles of the nanoparticles, suspensions of nanoparticles
2.2.2. Nanooctahedra (NO). In order to synthesize truncated were prepared in ethanol and deposited on fused silica substrates for
nanooctahedra (NO), solid Ce(NO3)3·6H2O was dissolved in DI water analysis by deep XPS using the same instrument as for the surface XPS
in a 250 mL Pyrex beaker and magnetically stirred for 30 min to yield a analysis but with different operating parameters (13.4 kV, 10.8 mA,
concentration of 0.9 M. H2O2 was added dropwise using a glass dropper monochromated AlKα X-rays at 1487 eV, beam diameter 500 μm). An
while monitoring the pH until it decreased to a value of ∼3.5. At this argon ion beam at 1 keV was used to etch an area of 2.5 mm × 2.5 mm at

6017 DOI: 10.1021/acs.inorgchem.9b00330


Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

Figure 1. TEM images of nanoceria: rows 1−3 TEM, high resolution TEM, and SAED patterns of (a−c) nanocubes (NC), (d−f) nanorods (NR), (g−
i) and nanooctahedra (NO).

a rate of 0.14 nm/s and sampling interval of 5 s. Tantalum pentoxide that the residual lattice strain (tensile) was in the converse order
(Ta2O5) was used as a calibration reference for the depth of etching. NO > NR > NC, which confirms the degrees of crystallinity.
The expected partial of the surface by argon ion beam is explained in As the TEM data, discussed subsequently, reveal that the
results and discussion.
2.4. Photocatalytic Activity. The photocatalytic performances of nanoceria grains are single crystal, then the crystallite sizes
all three morphologies of nanoceria were assessed in terms of (Supporting Information Section 2, Table S1) and grain sizes
photobleaching of methylene blue (MB) under UV irradiation (Supporting Information Section 2, Table S2) should be
separately at pH 6.0 and pH 9.0. Methylene blue dye solutions were approximately the same. This is the case for the NR, where
prepared by dissolving MB (10−5 M) in deionized water. Nanoceria the former (based on a spherical grain) is 11 nm and the latter
particles then were dispersed in the dye solution at a concentration of 2 (based on the rod diameter) is 17 ± 9 nm. It also is the case for
mg/mL in a beaker, which then was placed in a light-obstructing
container. Irradiation was done using a UV lamp (UVP, 3UV-38, 8 W, the NO, where the former is 5 nm and the latter (based on the
350 V, ∼50 Hz, 0.16 A, light intensity 0.39 mW/cm2 with the lamp cross section) is 7 ± 2 nm. However, there is a significant
placed 9 cm above the solution) at illuminating intensity 1.6 × 10−17 J at difference between the two values for the NC, which are 20 nm
365 nm for 1 h and 7.9 × 10−17 J at 365 nm for 3 h. The pH values were for the former and 58 ± 5 nm for the latter (based on the width).
modified by addition of 1 M NaOH and 1 M HCl. Tests were done in If the size differences between the spherical and cubic
triplicate for both time points. The MB degradation from UV morphologies are calculated on the basis of the TEM data, the
irradiation was analyzed in terms of the change in the absorbance
peak height of MB (665 nm) using a PerkinElmer UV−visible crystallite sizes should be in the range 39−46 nm, which
spectrometer. represents a range of 26−27% disagreement, which is within the
ranges of variation for the other two morphologies.
3. RESULTS AND DISCUSSION However, there also is a crystallite size effect on the Raman
peak shift, where increasing crystallite size (or grain size for these
Cerium oxide nanocubes (NC), nanorods (NR), and nano-
single-crystal nanoparticles) is inversely proportional to the red
ctahedra (NO) were produced by precipitation and hydro-
shift.36 Since the observed red shift is in the order NO > NR >
thermal synthesis in the absence of surfactant or capping agent.
All three morphologies occurred as essentially single phases and NC and the grain size (using the smaller, more easily vibrated,
not as mixtures. The X-ray diffraction (XRD) data for the three width of the NR) also shows the converse relationship, then the
morphologies (Supporting Information section 1, Figure S2a− red shift also may result from a particle size effect, which
c) show that the degree of crystallinity is in the order NC > NR > generally is considered to be a result of phonon confinement7
NO. A decrease in the degree of crystallinity is reflected in the and which results in the observed trend in asymmetric line
extent of residual lattice strain, which can be assessed by both X- broadening. The longitudinal optical vibration at ∼600 cm−1 is
ray line broadening34 and laser Raman microspectroscopy peak well established as being indicative of the presence of oxygen
shift35 (Supporting Information section 1, Figure S2d). Both vacancies (VÖ ), albeit invariably at low intensities.37,8
sets of data (Supporting Information Section 2, Table S1) show Supporting Information section 1, Figure S2d shows that the
6018 DOI: 10.1021/acs.inorgchem.9b00330
Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

Figure 2. XPS spectra and relevant correlations of nanoceria: (a−c) morphology−valence correlation; (d−f) crystallographic deconvolution of the
relevant O 1s peaks.

Table 1. Correlation of [VÖ ] of Nanoceria with Redox/Photocatalytic Effectsa


concentrations (at. %)
O bon-
ded to Ce (XPS)
total total O filled total surface-adsorbed H2O residual unfilled adsorbed H2O on surface
[VÖ ]A Ce3+ Ce4+ sitesB VÖ C (XPS) VÖ D Ce3+ E
RedoxF
morphology [VÖ ]Tot [Ce3+]O [Ce4+]O [OSites] [VÖ ]Fill [H2O]Ads/Tot [VÖ ]Unfill [H2O]Ads/Ce3+ (MB)
NC 11.3 27.0 54.9 93.2 6.8 18.1 4.5 11.3 22 ± 7
NR 13.8 15.0 70.4 99.2 0.8 14.6 13.0 13.8 68 ± 4
NO 13.8 30.0 54.4 98.2 1.8 15.6 12.0 13.8 59 ± 2
a
Methods of determination of values (superscripts A−F) in table are given in the Supporting Information section 3, note 2

peak for NC is lost in the spectrum noise and the peak for NR is the {100} and {110} planes is another {110} plane, this confirms
obscured by the noise, but the peak for NO is unambiguous. that the terminating pinacoid is a {110} plane, also as indicated
Transmission electron microscopy (TEM) and selected area by others.27
electron diffraction (SAED) data are presented in Figure 1. In The low-resolution TEM images demonstrate that all three
row 1, the particle morphologies, sizes, and the natures of morphologies give the appearance of being agglomerated
agglomeration can be seen. It may be noted that the commonly (Supporting Information section 1, Figure S1). However, the
observed corner-truncated and corner- and edge-truncated13 NC show corner-to-face contact; the TEM images demonstrate
cubic morphologies were not observed. Although conversion of the occasional presence of isolated NC, NR, or NO grains; and
the (100) facets of a perfect cube are known to reconstruct to the there is no evidence of nanoparticle intergrowths; all of which
(111) corner and (110) edge truncations under strongly suggest that the nanoparticles did not form hard agglomerates.22
oxidizing conditions,12 other work16,17 shows that the Hence, the face-to-face contact of the NR and NO indicates that
morphology is strongly dependent on the temperature and
soft agglomerates22,23 formed as transient artifacts of the
NaOH concentration (i.e., pH). In row 2, the crystallographic
synthesis processes as well as the dispersion process (viz.,
orientations of the morphologies are confirmed by correlations
between measured and reported interplanar spacings.38 In row suspension in ethanol) used for TEM analysis. ImageJ39 was
3, the diffraction spots of the SAED pattern of the NC confirm used to process the TEM images in order to calculate the
the high degree of crystallinity. The XRD crystallinity data for dimensions of the nanoparticles, where 50 grains each of the NC
the NR and NO are confirmed by the nature of the rings, which and NO revealed uniform sizes while the NR were of variable
are sharper for the NR and more diffuse for the NO. The aspect ratios, so average values for the dimensions of 50 NR were
crystallographic relations of the NR confirm (Supporting determined (Supporting Information section 2, Table S2).
Information section 1, Figure S3) that it exhibits {100} and Finally, relative surface areas (RSA) for each surface-exposed
{110} pinacoids that form what are effectively the prism faces facet for all three morphologies are contrasted with the total
and these are capped by {111} and {100} domes and a {110} specific surface areas (SSA) (Supporting Information section 2,
pinacoid. Since the only plane that is mutually perpendicular to Table S3). The RSA are calculated on the basis of perfect planar
6019 DOI: 10.1021/acs.inorgchem.9b00330
Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

Figure 3. Structure−activity correlations and deep XPS of nanoceria: (a) redox/photocatalytic performance correlation with time at pH 6.0; (b)
redox/photocatalytic performance correlation with time at pH 9.0; (c) redox/photocatalytic performance correlations with [VÖ ]Tot and [VÖ ]Unfill; (d)
Microstructural and structural correlations with [VÖ ]Unfill; (e) depth profiles for [O] from deep XPS O 1s peak (depths from calibrated data based on
Ta2O5 depth profiles31) and EELS (positive locations [purple] inverted relative to negative locations [black] (Ce vacancies at greater depths not
shown); (f) depth profiles for [Ce3+] from deep XPS Ce 3d (depths from calibrated data based on Ta2O5 depth profiles31).

facets, but some degree of surface irregularity is inevitable in all of the Ce3+−O peaks in the O 1s spectra reflect the
systems. concentrations of the oxygen vacancies.
The surface XPS spectra for the Ce 3d region, which are Since, in principle, there is one oxygen vacancy per two Ce3+
represented in Figure 2a−c, demonstrate the presence of both ions, then the [Ce 3+ ] and the total oxygen vacancy
valences of cerium.40 The identification of Ce3+ confirms the concentrations ([VÖ ]Tot) for all three morphologies can be
presence of charge-compensating oxygen vacancies at the
determined (Supporting Information section 3, note 1). These
surfaces and subsurfaces of the nanoparticles.32,40−42 Further-
more, the O 1s spectra are represented in Figure 2d−f, where the calculations show that the order of [VÖ ]Tot (and hence
1
binding energies of Ce 4+ −O, Ce 3+ −O, and Ce−H 2 O /2[Ce3+]) is NO = NR > NC, which suggests the expected
(adsorbed) are given.11,41 The intensities and areas under the order of redox/photocatalytic performance.
peaks of the Ce3+-O and Ce−H2O regions have been The quantitative XPS data have been decoupled in order to
deconvoluted, which is discussed subsequently. The intensities differentiate the total oxygen vacancy concentration ([VÖ ]Tot)
6020 DOI: 10.1021/acs.inorgchem.9b00330
Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

into filled ([VÖ ]Fill) and unfilled ([VÖ ]Unfill) oxygen vacancy ance. More importantly, these data must be considered in light
concentrations, which are given in Table 1. of the effects of the morphology and the surrounding
Since it is required that the total number of O sites must environment.
equate to [VÖ ]Tot + [Ce3+]O + [Ce4+]O = 100%, then the fact that A key striking observation for Table 1 (see Supporting
[OSites] are less than theoretical (100%) indicates that the Information section 3, note 2, footnote D) is that the two
difference arises from the transient filling of the VÖ with OH− of independently obtained sets of experimental data, which are the
the H2O from the environment43 (the data have been [Ce3+] and the [H2O]Ads/Tot, yield equal values for [VÖ ]Tot and
normalized to eliminate adventitious carbon and adventitious [H2O]Ads/Ce3+ for all three morphologies with no variation
oxygen associated with carbon; these data are given in between the numerical values, viz., 100% precision. This triple
Supporting Information section 3, note 1). Consequently, set of identities shows that the [H2O]Ads/Ce3+ is associated solely
these inequalities allow the determination of the [VÖ ]Fill, which with the Ce3+ ions and not Ce4+ ions. This association of the
have been transiently filled by the adsorbed OH−, and the adsorbed H2O and Ce3+ ions is expected since the Pourbaix
[VÖ ]Unfill, which are determined by the Ce3+ ↔ Ce4+ equilibrium diagram47 shows that Ce4+ is not compatible with H2O, resulting
constants. This decoupling of the two different types of VÖ is in reductive solubility. The equality between these two variables
critical because the former are deactivated for redox/photo- actually is as expected because each oxygen vacancy in CeO2‑x is
catalysis while the latter remain active, the concentrations of integrated with an unpaired electron associated with each of the
which correlate with average decrease in methylene blue (MB) two Ce3+ ions. These two electrons thus are available to interact
dye degradation. The rationale for this decoupling is supported with the valence electrons of the OH− ions in a single adsorbed
by recent DFT calculations,2 which indicate that the formation H2O molecule. This is why there is a 1:1 ratio between the total
of the subsurface oxygen vacancies is energetically more favorable number of oxygen vacancies and the number of H2O molecules
than that at the surface. adsorbed on the free surface (viz., Ce3+ ions) of the solid.
The relevant experimental and redox/photocatalytic perform- Furthermore, for redox/photocatalytic performance, it has
ance data are shown in Figure 3. Parts a−d of Figure 3 clearly been established both theoretically and experimentally that
link the morphology to the [VÖ ]Unfill and corresponding redox/ molecular oxygen dissolved in aqueous solutions can be
photocatalytic performance. This is highlighted by the matched adsorbed at oxygen vacancy sites on nanoceria surfaces, thus
data for the latter two figures (Figure 3a,b), which reveal NC ≪ potentially generating a superoxide radical (••O2−) by accepting
NO < NR. Furthermore, the similarity in the data sets (1 h vs 5 one of the two unpaired electrons during continuous reversible
h) shown in the former two figures indicates the rapidity of the redox Ce3+ ↔ Ce4+ switching18 These superoxide radicals are
redox/photocatalytic reactions. Also, while the effect of pH also effectively ROS and so can enhance the redox/photocatalytic
is significant, the published data are for irregularly shaped performance of nanoceria.
particles, which probably explains their contradictory The concept of f illed and unfilled oxygen vacancies suggests
trends.42−46 The present work clarifies the relevant features. that the former are accessed readily by environmental H2O
At pH 6.0, the reducing conditions favor the presence of Ce3+, because they are located in the outermost surface monolayer
which should enhance the performance. However, it is pH 9.0 while the latter are blocked partially from access by H2O because
and the oxidizing conditions favoring Ce4+ that are superior. they are located at the adjacent subsurface monolayers. This
This surprising result is explained clearly by the speciation and model has been explored in the present work by depth profiling
Pourbaix diagrams.47 That is, at pH 6.0, the surface of the CeO2‑x using deep XPS. However, this technique is contentious owing
experiences solubility of both Ce4+ (reductive) and Ce3+, so the to the established partial reduction of CeO2‑x by both the X-ray
active sites are degraded continuously by partial removal of Ce3+ beam49−51 and the Ar ion beam52,53 used in XPS. Examination of
and associated surface and subsurface oxygen vacancies. In the relevant literature for pure ceria (a summary of the key
contrast, at pH 9.0, a Ce(OH)4 passivating layer is established studies is given in the Supporting Information, Table S5) reveals
on the surface of the CeO2‑x. When the UV radiation that this reduction is a function of the beam power,50−53
decomposes it,48 this results in the conversion to the soluble exposure time,50,52 and particle size.4,8,31 While differentiation
species Ce(OH)2+ (at all pH) and Ce(OH)3 (up to pH ∼ 9.7).47 between the surface altered by the Ar and X-ray beams and the
This results in removal of the surface-filled oxygen vacancies unaltered bulk has been demonstrated,50 the depth of the surface
(incorporated in the gel layer) and exposure of the subsurface alteration has been determined to be 1.5−2.052 or 2.0−3.0 nm,3
oxygen vacancies directly to the solution. This explains the depending on X-ray beam power and time.
significant increase in the redox/photocatalytic performance, as It appears that the only data reporting a cross sectional
shown in Figure 3c,d. As will be discussed subsequently, the compositional profile for pure ceria are those of Stroppa et al.,31
origin of the morphological correlation lies in the low-index who used EELS to determine the [O] across single
exposed crystallographic planes. coprecipitated octahedral (described as spherical) grains of
Furthermore, it is well-known that there are converse ceria (Figure 3e). These data reveal that the entire surface is
correlations between the particle size and other structural hyperstoichiometric in oxygen, which is surprising in light of
parameters,42 as shown in Figure 3d (also see Supporting both the expectation of the presence of oxygen vacancies and the
Information section 2, Table S4). In light of recent reports of the partial reduction from the X-ray beam. More importantly, these
role of the [VÖ ]Tot (and hence 1/2[Ce3+]),10,34 the redox/ data indicate that there is a steep downward oxygen gradient
photocatalytic performance should exhibit a converse correla- over the first monolayer (surface) and the adjacent one or two
tion with the particle size. In contrast, it is significant that parts c monolayers (subsurface), below which the data trend toward a
and d of Figure 3 show that the present work does not support constant oxygen concentration (bulk). The scatter in the bulk
the conventionally assumed direct relation between [Ce3+] and data is likely to have resulted from variations in beam
the redox/photocatalytic performance. The data supporting this penetration volume across each rounded grain, grain roughness,
apparent anomaly are given in Table 1, where it can be seen that the presence of Ce3+/Ce4+ clusters3, and/or the presence of
the [VÖ ]Unfill correlates with the redox/photocatalytic perform- oxygen vacancy clusters.54 However, the spike to higher [O] at a
6021 DOI: 10.1021/acs.inorgchem.9b00330
Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

Table 2. Exposed Surface Planes, Relative Defect Concentrations, And Redox/Photocatalytic Performance As a Function of
Morphologya
relative deconvoluted surface data for exposed facets and their redox performances
morphology SSA (m2/g) [VÖ ]Tot exposed facets RSAA (%) [VÖ ]RelB [VÖ ]FillC [VÖ ]UnfillE RedoxRelFb
NC 14 11.3 {100} 100 11.3 6.8 4.5 22
NR 105 13.8 {100} 48 5.4 0.8 4.6 10.6
{110} 49 8.1 0D 8.1 55.5
{111} 3 0.3 0D 0.3 1.5
NO 199 13.8 {100} 43 4.8 1.8 3 9.5
{111} 57 9 0D 9 49.5
a
Methods of determination of values (superscripts A−F) in the table are given in Supporting Information section 3, note 3. bCalculated by rational
analysis using the RSA.

depth of 2.1 nm and the reverse trend in the subsurface can be significantly; this region contains VÖ transiently but
explained by the presence of metal vacancies, which is discussed completely filled by OH−.
subsequently.
• Across the adjacent four (EELS) or five (deep XPS)
The nature of these EELS data31 was confirmed using deep lattice spacings, the [Ce3+] and [O] change more
XPS depth profiling for powder films of all three morphologies, gradually; this region contains VÖ transiently but only
as shown in Figure 3e. This correlation is significant partly partially filled by OH−.
because the EELS data report point lateral analyses across single
equiaxed grains while the deep XPS data report planar depth • At further depth in the grain, the [Ce3+] and [O] are
analyses into multiple grains of three different morphologies. constant (deep XPS and EELS); this region contains
The largest change in oxygen concentration is over the first unfilled VÖ .
monolayer (surface), followed by more gradual decreases over The data in Figure 3f are described in terms of the {100}
the adjacent five monolayers (subsurface); the constancy of the planar spacing (0.542 nm) since that is common to all three
oxygen vacancy concentration at greater depths is clear (bulk). morphologies; the planar spacings of the other two planes are
In light of the expected partial reduction of the surface 1.5− significantly different ({110} = 0.383 nm; {111} = 0.313 nm31).
3.0 nm by an argon ion beam, the reliability of these data must be The diameter of the beam for both surface and deep XPS is very
considered. First, the bulk composition is not affected by the large (500 μm), so the profiles of a very large number of
beam,3,8,50,52 so the data at depths greater than ∼3.0 nm are randomly oriented grains were analyzed. In contrast, the EELS
reliable. Second, since surface reduction takes place, this forces data were for individual surface profiles across single grains (6.8
the trend from the surface to the bulk to increasing [Ce3+], ± 1.8 nm) at much greater resolution (interaction diameter 0.4
which means that the overall trends of the curves are correct. nm). The difference in the levels and trends of the EELS and
Third, the acquisition time of 5 s is substantially less than those deep XPS data are likely to be a result of the interaction volume
used in similar deep XPS studies (Supporting Information, as a function of beam location, where that for deep XPS remains
Table S4, 900 s at 1487 eV and 5400 s at 50 eV), so considerably approximately constant but EELS is variable. A second point of
less reduction from the X-ray beam is expected in the present differentiation is that the grains analyzed by XPS are nominally
work. faceted, so the assumptions of the [Ce]/[O] and [Ce3+]/
Since, as shown in Figure 3f, the bulk [Ce3+] from deep XPS of [VÖ ]Tot stoichiometry ratios hold. In contrast, the EELS data are
∼7% is a maximum but the surface XPS data in Tables 1 and 2 for rounded grains, where such assumptions may not be entirely
are substantially higher, this apparent anomaly must be accurate.
explained. The reason for this lies in the two different A key observation is confirmation that all [O] are oxygen
techniques. With deep XPS, each layer is etched and analyzed hyperstoichiometric, even in the bulk, which is in contradiction
sequentially over a period of 5 s at 1487 eV, so these data provide with the Ce−O phase diagram.47 While charge compensation in
local compositional information. In contrast, surface XPS, which hypostoichiometric ceria (CeO2‑x) is achieved by Ce3+ ↔ Ce4+
is analyzed over a period of 30 min at 1487 eV, provides redox, the achievement of oxygen hyperstoichiometry in CeO2+x
volumetric data over a larger region that corresponds can be attained only through oxygen injection or cerium vacancy
approximately to the depth of the surface + subsurface. Although formation. Since it is unlikely that interstitial oxygen would exist
the surface XPS probe depth depends principally on the X-ray simultaneously with oxygen vacancies, then this indicates that
beam power and time and the X-ray absorption coefficient of the the actual formula of ceria is better expressed as
material, typical beam penetration depths are reported to be two
to three monolayers,55 1−3,56 2−5,57 and <10 nm.58 Thus, it is
expected that the surface XPS data would yield greater [Ce3+] where the bold type represents vacancies. The presence of
values than those from deep XPS, which is the case. In light of cerium vacancies is likely to enhance the redox/photocatalytic
the consistency of the data in Tables 1 and 2, it is concluded that effects owing to (1) a second such mechanism (in addition to
the surface XPS of the present work analyzed the monolayers oxygen vacancies) for these processes, (2) the presence of the
corresponding principally to the surface + subsurface. associated unpaired electrons from Ce3+, and (3) the same effect
These data confirm the perspective that ceria can be from increased surface roughness and hence surface area.59,60 It
differentiated into three regions: is clear that the simultaneous presence of vacancies and
electrons suggests a mixed ionic-electronic conductor
• Across the first lattice spacing (surface and deep XPS) or (MIEC). While this has been established for doped ceria,61,62
two lattice spacings (EELS), the [Ce3+] and [O] change it does not appear to have been suggested for the semi-
6022 DOI: 10.1021/acs.inorgchem.9b00330
Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

Figure 4. Mechanisms of action of nanoceria: (a) mechanism of H2O adsorption; (b) redox switching of Ce3+ ↔ Ce4+ (the mechanism for cerium
vacancies is effectively identical).

conductivity involving the intrinsic defects in ceria. It also is clear morphology with the highest [VÖ ]Unfill is the NR. The
that, if the charge deriving from the oxygen vacancies (VÖ ) does engineering of nanoceria surfaces to consist of surfaces with
not balance with the charge deriving from the cerium vacancies high concentrations of subsurface-unf illed oxygen vacancies will
(VÖ , VÖ ), then electronic charge compensation (e−, h*) must allow increased conversion of environmental H2O and O2
occur. molecules into free •OH and ••O2− radicals, respectively, and
While both cerium and oxygen vacancies have the potential to ROS,18,19,24 as illustrated in Figure 4. Further, since the oxygen
contribute to the redox/photocatalytic performance of nano- hyperstoichiometry decreases with depth into the grain, then it is
ceria by introduction of shallow midgap energy states,63−65 it is likely that both types of vacancies play the same role in redox/
the oxygen vacancies that are critical. Specifically, the outermost photocatalysis.
surface monolayer, which is partially f illed by OH− ions, and the For nanoceria, EELS generally has been applied to infer the
gradient and scale of the adjacent monolayers are important [VÖ ] at grain boundaries,29,30 although single grains have been
because they reflect the ease of access of both cerium and oxygen examined by Stroppa et al.31 However, the potential for VÖ or
vacancies to the adsorbed species that form free radicals and VÖ formation has not been considered. The work of Stroppa et
reactive oxygen species (ROS). These concepts are elucidated al.31 is significant because the cross sectional analysis of
generally in the mechanisms of H2O adsorption and redox/ coprecipitated CeO2 was done on octahedral grains (albeit
photocatalysis in Figure 4. The latter mechanism is validated by modeled as spheres), for which the exposed facets were {200}
the results for the methylene blue (MB) degradation by free and {111}. While these planes were clear in high-resolution
radicals and ROS shown in parts a−c of Figure 3, which TEM (HRTEM), this sensitivity was not possible with the high-
emphasize the dominant role of the subsurface-unfilled oxygen resolution high-angle annular dark field (HAADF) images that
vacancies. Hence, although the NO particles have the highest were used for the cross sectional analyses. Hence, it was not
measured and calculated surface areas (Supporting Information possible to determine which exposed facets were analyzed.
Section 2, Table S3) and, more importantly, the highest [VÖ ]Tot, The present work elucidates these data using the RSA for each
it is the NR that exhibit the best redox/photocatalytic plane. That is, the [VÖ ]Tot can be proportioned according to the
performance, which is in agreement with other work.66,67 This relative surface areas since it has been shown that both the [VÖ ]
is explained by the data in Table 1, which show that the and their formation energies for the individual planes are
6023 DOI: 10.1021/acs.inorgchem.9b00330
Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

constant, regardless of the morphology.67 As shown in Table 2, (1) All of the data in the present work are for ceria in an
the ([VÖ ]Tot is deconvoluted into [VÖ ]Rel, thereby representing aqueous environment, for which there are no analytical
the associated relative oxygen vacancy concentration for each methods for the determination of structural defects.
plane. The redox performances also are deconvoluted into (2) The deep XPS data illustrate the compositional trends
RedoxRel according to the exposed facets of each nanoceria between surface + subsurface (low [Ce3+]) and bulk (high
morphology. Consequently, it is clear that there is a Ce3+ ↔ Ce4+ [Ce3+]), the latter of which is considered to be an accurate
equilibrium constant for each plane. Furthermore, these value and the former of which must be at a lower [Ce3+].
equilibrium constants are moderated by the nature of the (3) The deep XPS and surface XPS data do not agree because
OH− electrostatic bond, the reduction potential, the oxygen the former provides local and sequential compositional
vacancy formation energy, and the planar surface area. data as a function of depth while the latter provides
In Table 2, the [VÖ ]Fill from Table 1 are weighted entirely to volumetric data for a larger volume of the solid.
{100} for two reasons: (4) The EELS and deep XPS data demonstrate the same
• The H2O adsorption energies are in the order:55 trends and support the conclusion that ceria is oxygen-
rich, the ionic charge compensation and associated
{100} = −158.37 kJ/mol > >{110} = −91.90 stoichiometry of which require the presence of cerium
vacancies.
kJ/mol > {111} = −63.10 kJ/mol (5) The surface XPS data for the total [Ce3+] can be used to
determine the total oxygen vacancy concentration. These
• Table 2 stresses that H2O adsorption on {100} dominates data also can be used to decouple total oxygen vacancy
and effectively excludes those on {110} and {111} concentration into f illed and unfilled oxygen vacancy
because [VÖ ]Fill is in the order: concentrations.
(6) The surface XPS data for the concentration of total
NC{100} = 6.9% > > NO{100} + {111} = 1.8% surface-adsorbed H2O can be used to determine the
> NR{100} + {111} + {110} = 0.8% concentration of transiently surface-f illed oxygen vacan-
cies, which then can be used to determine the
Consequently, the concentration of unfilled oxygen vacancies concentration of subsurface-unf illed oxygen vacancies.
([VÖ ]Unfill) for each plane can be calculated by difference from (7) The preceding data are verified by the differences between
[VÖ ]Rel. the surface XPS data for the concentration of total surface-
The deconvoluted data for the redox/photocatalytic perform- adsorbed H2O and the total oxygen vacancy concen-
ance in Table 2 show that the ranking of the effectiveness of the trations (i.e., 1/2[Ce3+]), which give identical values for
planes in the NR and the corresponding RSA are in the the total oxygen vacancy concentrations.
respective orders: (8) The [Ce3+] determined by surface XPS and decoupled
{110} > {100} > {111} into f illed and unf illed oxygen vacancies depends on the
Ce3+ ↔ Ce4+ equilibrium constant.
49% > 48% > >3% (9) While EELS appear to be one of the few methods with the
precision and sensitivity for chemical analyses at the
This relation is identical to evaluations by density functional
subnanometer range, it can be used only to infer oxygen
theory (DFT) calculations of oxygen vacancy formation
vacancy concentrations; the present method demon-
energies for these exposed crystallographic planes.26,68 Since
strates that surface XPS data can be used not only to
the {110} planes have the highest [VÖ ]Unfill, they have the
calculate the total oxygen vacancy concentrations but also
highest RSA, and they are unique to NR, then these confirm the
to decouple these into f illed and unf illed oxygen vacancy
superior performance of this morphology. Hence, it is the
concentrations.
contribution of the {110} planes that dominates the redox/
photocatalysis results. In contrast, the ranking of the (10) EELS using point lateral depth probing of single grains and
effectiveness of the planes in NO and the corresponding RSA deep XPS using planar depth probing of multiple grains
are not in the same respective relative orders for the two other represent complementary chemical analytical techniques
planes: at subnanometer and nanometer resolutions, respectively.
(11) The general assumption that there is one oxygen vacancy
{111} > {100} for every two Ce3+ ions is confirmed, although this
actually consists of the sum of the f illed and unf illed
57% > 43% oxygen vacancy concentrations.
Again, the reason for this is shown in Table 2, where the values (12) The surface-adsorbed water is associated solely with Ce3+
for [VÖ ]Unfill and the RSA correspond with this order. ions and not with Ce4+ ions on the free surface. The
These DFT data as well as those differentiating between number of Ce3+ ions on the free surface is the sum of the
surface and subsurface oxygen vacancies2,68 emphasize the two equal numbers of the oxygen vacancies and the
importance of the deconvolution because both morphologies surface-adsorbed water.
exhibit the same [VÖ ]Tot but this is not the parameter that (13) The OH− ions of environmental H2O transiently and
determines the redox/photocatalytic performance; it is the partially fill the oxygen vacancies. More specifically, these
[VÖ ]Unfill. Thus, the present work not only supports the view that OH− ions and H2O transiently and partially fill the
subsurface oxygen vacancies exist on the three principal low outermost surface monolayer oxygen vacancies, which are
index crystallographic planes, but it also identifies these defects deactivated for redox/photocatalysis. The access to the
as the active sites for redox/photocatalysis. oxygen vacancies at the subsurface adjacent monolayers by
More specifically, the XPS data and their analyses highlight the OH− ions of H2O is attenuated with increasing depth
the following points: into the grain and so these vacancies are partially filled and
6024 DOI: 10.1021/acs.inorgchem.9b00330
Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

unfilled. The unfilled subsurface oxygen vacancies remain the XPS and redox/photocatalytic performance data accord-
active sites for redox/photocatalysis. ingly. The key outcome of these procedures is the quantitative
(14) The consistency of the surface XPS, deep XPS, and EELS determination that the redox/photocatalytic performance
data indicate that the defect structures established during depends not on the total oxygen vacancy concentration for
the aqueous synthesis conditions were retained during the each plane assumed from the [Ce3+] from surface XPS but on
vacuum testing conditions. the concentration of the residual unfilled oxygen vacancies as
(15) The effect of pH is critical in that the expected benefit of determined by the present analysis. Furthermore, the perform-
acidic (reduction to form Ce3+) conditions is contradicted ance can be increased significantly through the imposition of
by the superior redox/photocatalytic performance of basic pH conditions. These result in the formation of a Ce(OH)4
basic conditions. passivating layer, which is decomposed by UV radiation, thereby
(16) The preceding phenomenon results from the dissolution exposing the unfilled oxygen vacancies directly to the
of Ce3+ and loss of associated oxygen vacancies from the surrounding environment. The data analysis also introduces
surface of CeO2‑x under acidic conditions while basic the importance of the acquisition of sequential deep XPS data,
conditions result in the formation of a protective which provide the bulk stoichiometry, identify the relative
passivating layer of Ce(OH)4, which, upon decomposi- stoichiometry at the surface, and indicate the trend between
tion from exposure to UV radiation, exposes the unfilled these two across the subsurface. Furthermore, deep XPS and
subsurface oxygen vacancies. corresponding EELS data demonstrate that oxygen hyper-
(17) Ceria contains not only oxygen vacancies but cerium stoichiometry requires the formation of cerium vacancies, which
vacancies, the latter of which are imposed by the also can be expected to contribute to redox/photocatalysis. The
confirmation of oxygen hyperstoichiometry on the surface approaches of the present work provide guidance to the
and in the bulk. These cerium vacancies also may engineering of different morphologies, sizes, and aspect ratios
contribute to the redox/photocatalytic effects owing to of the nanoceria grain morphologies that can be synthesized by
the associated unpaired electrons. precipitation and hydrothermal processes. They also have
significant practical ramifications to semiconductors as these
(18) The bulk of the grain contains a constant level of unfilled
materials inevitably are exposed to water vapor in the
oxygen vacancies, of which the contribution, if any, to the
environment.


redox/photocatalytic activity remains unknown.
(19) Since the oxygen vacancy concentration never has been ASSOCIATED CONTENT
deconvoluted according to the exposed facets, it remains
unrecognized that the effect of morphology on the redox/ *
S Supporting Information

photocatalytic performance of ceria is a function of the The Supporting Information is available free of charge on the
equivalent exposed crystallographic planes. Each of these ACS Publications website at DOI: 10.1021/acs.inorg-
exposed facets exhibits a Ce3+ ↔ Ce4+ equilibrium chem.9b00330.
constant, which is moderated by the nature of the OH− Transmission electron microscopy, X-ray diffraction, X-
electrostatic bond, the reduction potential, the oxygen ray photoelectron spectroscopy, laser Raman micro-
vacancy formation energy, and the planar surface area. spectroscopy, surface areas, and photocatalytic redox
(20) The match (NR) and nonmatch (NO) between the DFT performance data (PDF)


calculations for the oxygen vacancy formation energies for
the three low-index crystallographic planes and the redox/
photocatalytic performance are a reflection of the roles of AUTHOR INFORMATION
the oxygen vacancy formation energy and the planar Corresponding Author
surface area in the Ce3+ ↔ Ce4+ equilibrium constants for *(R.M.) E-mail: r.mehmood@unsw.edu.au.
the different exposed crystallographic planes. ORCID
Rashid Mehmood: 0000-0003-4073-7835
4. CONCLUSIONS Wen-Fan Chen: 0000-0002-2917-3875
The present work reports data for the structural parameters and Author Contributions
redox/photocatalytic performance of three morphologies of R.M. designed the project; undertook the syntheses, character-
nanoceria. A critical observation is that the defect structures ization, TEM imaging, and data analysis; wrote the first draft of
established during synthesis in aqueous environments are the manuscript; and worked on all subsequent drafts. S.S.M.
retained during testing under vacuum conditions, which is assisted in TEM imaging and data analysis. W.-F.C. assisted in
confirmed by the consistency of three different sets of Raman analysis. P.K. assisted with the data analysis. C.C.S.
experimental data. The overall data analysis introduces three provided the concepts for the data analysis, worked on all
novel procedures that enable volumetric surface XPS data to be subsequent drafts of the manuscript, and supervised the overall
used to determine quantitative data for surface, subsurface, and project.
bulk oxygen vacancy concentrations. First, the total oxygen
Notes
vacancy concentration is decoupled in order to differentiate
The authors declare no competing financial interest.


between (a) deactivated transiently f illed oxygen vacancies in
the outermost surface monolayer and (b) active sites in the form
of unf illed oxygen vacancies at the adjacent subsurface ACKNOWLEDGMENTS
monolayers. Second, the relative surface areas for each low- The authors wish to acknowledge grant support from the
index exposed crystallographic plane for all three morphologies, Australian Research Council (DP170104130), scholarship
each of which has a Ce3+ ↔ Ce4+ equilibrium constant, are support (R.M.) through an Australian Government Research
calculated. Third, the combined data allow the deconvolution of Training Program (RTP) Scholarship, and the characterization
6025 DOI: 10.1021/acs.inorgchem.9b00330
Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

facilities of the Mark Wainwright Analytical Centre at UNSW (17) Sakthivel, T.; Das, S.; Kumar, A.; Reid, D. L.; Gupta, A.; Sayle, D.
Sydney. C.; Seal, S. Morphological phase diagram of biocatalytically active ceria


nanostructures as a function of processing variables and their
REFERENCES properties. ChemPlusChem 2013, 78, 1446−1455.
(18) Gaynor, J. D.; Karakoti, A. S.; Inerbaev, T.; Sanghavi, S.;
(1) Esch, F.; Fabris, S.; Zhou, L.; Montini, T.; Africh, C.; Fornasiero,
Nachimuthu, P.; Shutthanandan, V.; Seal, S.; Thevuthasan, S. Enzyme-
P.; Comelli, G.; Rosei, R. Electron localization determines defect
formation on ceria substrates. Science 2005, 309, 752−755. free detection of hydrogen peroxide from cerium oxide nanoparticles
(2) Murgida, G. E.; Ganduglia-Pirovano, M. V. Evidence for immobilized on poly(4-vinylpyridine) self-assembled monolayers. J.
subsurface ordering of oxygen vacancies on the reduced CeO2(111) Mater. Chem. B 2013, 1, 3443−3450.
surface using density-functional and statistical calculations. Phys. Rev. (19) Yao, S. Y.; Xu, W. Q.; Johnston-Peck, A. C.; Zhao, F. Z.; Liu, Z.
Lett. 2013, 110, 246101. Y.; Luo, S.; Senanayake, S. D.; Martinez-Arias, A.; Liu, W. J.; Rodriguez,
(3) Holgado, J. P.; Alvarez, R.; Munuera, G. Study of CeO2 XPS J. A. Morphological effects of the nanostructured ceria support on the
spectra by factor analysis: reduction of CeO2. Appl. Surf. Sci. 2000, 161, activity and stability of CuO/CeO2 catalysts for the water-gas shift
301−315. reaction. Phys. Chem. Chem. Phys. 2014, 16, 17183−17195.
(4) Kato, S.; Ammann, M.; Huthwelker, T.; Paun, C.; Lampimaki, M.; (20) Garvie, L. A. J.; Buseck, P. R. Determination of Ce4+/Ce3+ in
Lee, M.-T.; Rothensteiner, M.; Van-Bokhoven, J. A. Quantitative depth electron-beam-damaged CeO2 by electron energy-loss spectroscopy. J.
profiling of Ce3+ in Pt/CeO2 by in situ high-energy XPS in a hydrogen Phys. Chem. Solids 1999, 60, 1943−1947.
atmosphere. Phys. Chem. Chem. Phys. 2015, 17, 5078−5083. (21) Koirala, R.; Pratsinis, S. E.; Baiker, A. Synthesis of catalytic
(5) Lane, J. A.; Kilner, J. A. Oxygen surface exchange on gadolinia materials in flames:opportunities and challenges. Chem. Soc. Rev. 2016,
doped ceria. Solid State Ionics 2000, 136-137, 927−932. 45, 3053−3068.
(6) Paidi, V. K.; Savereide, L. S.; Childers, D. J.; Notestein, J. M.; (22) Tsantilis, S.; Pratsinis, S. E. Soft- and hard-agglomerate aerosols
Roberts, C. A.; van Lierop, J. Predicting NOx catalysis by quantifying made at hightemperatures. Langmuir 2004, 20, 5933−5939.
Ce3+ from surface and lattice oxygen. ACS Appl. Mater. Interfaces 2017, (23) Djuricic, B.; Pickering, S. Nanostructured cerium oxide:
9, 30670−30678. preparation and, properties of weakly-agglomerated powders. J. Eur.
(7) Yao, S. Y.; Xu, W. Q.; Johnston-Peck, A. C.; Zhao, F. Z.; Liu, Z. Y.; Ceram. Soc. 1999, 19, 1925−1934.
Luo, S.; Senanayake, S. D.; Martinez-Arias, A.; Liu, W. J.; Rodriguez, J. (24) Ren, H.; Koshy, P.; Chen, W. F.; Qi, S.; Sorrell, C. C.
A. Morphological effects of the nanostructured ceria support on the Photocatalytic materials and technologies for air purification. J. Hazard.
activity and stability of CuO/CeO2 catalysts for the water-gas shift Mater. 2017, 325, 340−366.
reaction. Phys. Chem. Chem. Phys. 2014, 16, 17183−17195. (25) Yuan, Q.; Duan, H. H.; Li, L. L.; Sun, L. D.; Zhang, Y. W.; Yan, C.
(8) Wu, Z.; Li, M.; Howe, J.; Meyer III, H. M.; Overbury, S. H. H. Controlled synthesis and assembly of ceria-based nanomaterials. J.
Probing defect sites on CeO2 nanocrystals with well-defined surface Colloid Interface Sci. 2009, 335, 151−167.
planes by Raman spectroscopy and O2 adsorption. Langmuir 2010, 26, (26) Jiang, Y.; Adams, J. B.; Van-Schilfgaarde, M. V. Density-
16595−16606. functional calculation of CeO2 surfaces and prediction of effects of
(9) Paun, C.; Safonova, O. V.; Szlachetko, J.; Abdala, P. M.; oxygenpartial pressure and temperature on stabilities. J. Chem. Phys.
Nachtegaal, M.; Sa, J.; Kleymenov, E.; Cervellino, A.; Krumeich, F.; 2005, 123, 064701.
Van-Bokhoven, J. A. Polyhedral CeO2 nanoparticles: Size-dependent (27) Yang, Y.; Mao, Z.; Huang, W.; Liu, L.; Li, J.; Li, J.; Wu, Q. Redox
geometrical and electronic structure. J. Phys. Chem. C 2012, 116, 7312− enzyme-mimicking activities of CeO2 nanostructures: Intrinsic
7317. influence of exposed facets. Sci. Rep. 2016, 6, 35344.
(10) Yu, P.; Hayes, S. A.; O’Keefe, T. J.; O’Keefe, M. J.; Stoffer, J. O. (28) Rodriguez, J. A.; Ma, S.; Liu, P.; Hrbek, J.; Evans, J.; Perez, M.
The phase stability of cerium species in aqueous systems II. The Activity of CeOx and TiOx nanoparticles grown on Au (111) in the
Ce(III/IV)-H2O-H2O2/O2 system. Equilibrium considerations and water-gas shift reaction. Science 2007, 318, 1757−1760.
Pourbaix diagram calculations. J. Electrochem. Soc. 2006, 153, C74− (29) Feng, B.; Sugiyama, S.; Hojo, H.; Ohta, H.; Shibata, N.; Ikuhara,
C79. Y. Atomic structures and oxygen dynamics of CeO2 grain boundaries.
(11) Hsieh, C.-C.; Roy, A.; Rai, A.; Chang, Y.-F.; Banerjee, S. K. Sci. Rep. 2016, 6, 20288.
Characteristics and mechanism study of cerium oxide based random (30) Hojo, H.; Mizoguchi, T.; Ohta, H.; Findlay, S. D.; Shibata, N.;
access memories. Appl. Phys. Lett. 2015, 106, 173108.
Yamamoto, T.; Ikuhara, Y. Atomic structure of a CeO2 grain boundary:
(12) Tinoco, M.; Fernandez-Garcia, S.; Lopez-Haro, M.; Hungria, A.
the role of oxygen vacancies. Nano Lett. 2010, 10, 4668−4672.
B.; Chen, X.; Blanco, G.; Perez-Omil, J. A.; Collins, S. E.; Okuno, H.;
(31) Stroppa, D. G.; Dalmaschio, C. J.; Houben, L.; Barthel, J.;
Calvino, J. J. Critical influence of nanofaceting on the preparation and
Montoro, L. A.; Leite, E. R.; Ramirez, A. J. Analysis of dopant atom
performance of supported gold catalysts. ACS Catal. 2015, 5, 3504−
3513. distribution and quantification of oxygen vacancies on individual Gd-
(13) Kuchibhatla, S. V.; Karakoti, A. S.; Baer, D. R.; Samudrala, S.; doped CeO2 nanocrystals. Chem. - Eur. J. 2014, 20, 6288−6293.
Engelhard, M. H.; Amonette, J. E.; Thevuthasan, S.; Seal, S. Influence of (32) Paier, J.; Penschke, C.; Sauer, J. Oxygen defects and surface
aging and environment on nanoparticle chemistry - implication to chemistry of ceria: quantum chemical studies compared to experiment.
confinement effects in nanoceria. J. Phys. Chem. C 2012, 116, 14108− Chem. Rev. 2013, 113, 3949−3985.
14114. (33) Jamshidijam, M.; Mangalaraja, R. V.; Akbari-Fakhrabadi, A.;
(14) (a) Trovarelli, A. Catalytic properties of ceria and CeO2- Ananthakumar, S.; Chan, S. H. Effect of rare earth dopants on structural
containing materials. Catal. Rev.: Sci. Eng. 1996, 38, 439−520. characteristics of nanoceria d by combustion method. Powder Technol.
(b) Kumar, A.; Das, S.; Munusamy, P.; Self, W.; Baer, D. R.; Sayle, 2014, 253, 304−310.
D. C.; Seal, S. Behavior of nanoceria in biologically-relevant (34) Muhammed Shafi, P.; Chandra Bose, A. Impact of crystalline
environments. Environ. Sci.: Nano 2014, 1, 516−532. defects and size on X-ray line broadening: a phenomenological
(15) Mehmood, R.; Ariotti, N.; Koshy, P.; Yang, J.-L.; Sorrell, C. C. approach for tetragonal SnO2 nanocrystals. AIP Adv. 2015, 5, 057137.
pH-Responsive Morphology-controlled redox behaviour and cellular (35) Filtschew, A.; Hofmann, K.; Hess, C. Ceria and its defect
uptake of nanoceria in fibrosarcoma. ACS Biomater. Sci. Eng. 2018, 4, structure: new insights from a combined spectroscopic approach. J.
1064−1072. Phys. Chem. C 2016, 120, 6694−6703.
(16) Mehmood, R.; Wang, X.; Koshy, P.; Yang, J.-L.; Sorrell, C. C. (36) Spanier, J. E.; Robinson, R. D.; Zhang, F.; Chan, S.-W.; Herman,
Engineering oxygen vacancies through construction of morphology I. P. Size-dependent properties of CeO2‑y nanoparticles as studied by
maps for bio-responsive nanoceria for osteosarcoma therapy. Raman scattering. Phys. Rev. B: Condens. Matter Mater. Phys. 2001, 64,
CrystEngComm 2018, 20, 1536−1545. 245407.

6026 DOI: 10.1021/acs.inorgchem.9b00330


Inorg. Chem. 2019, 58, 6016−6027
Inorganic Chemistry Article

(37) McBride, J. R.; Hass, K. C.; Poindexter, B. D.; Weber, W. H. (56) Chen, W.-F.; Koshy, P.; Sorrell, C. C. Effects of film topology and
Raman and x-ray studies of Ce1‑x RExO2‑y, where RE = La, Pr, Nd, Eu, contamination as a function of thickness on the photo-induced
Gd, and Tb. J. Appl. Phys. 1994, 76, 2435−2441. hydrophilicity of transparent TiO2 thin films deposited on glass
(38) Lu, M.; Zhang, Y.; Wang, Y.; Jiang, M.; Yao, X. Insight into substrates by spin coating. J. Mater. Sci. 2016, 51, 2465−2480.
several factors that affect the conversion between antioxidant and (57) Gorzalski, A. S.; Donley, C.; Coronell, O. Elemental composition
oxidant activities of nanoceria. ACS Appl. Mater. Interfaces 2016, 8, of membrane foulant layers using EDS, XPS, and RBS. J. Membr. Sci.
23580−23590. 2017, 522, 31−44.
(39) Schindelin, J.; Arganda-Carreras, I.; Frise, E.; Kaynig, V.; Longair, (58) Lenser, C.; Gunkel, F.; Sohn, Y. J.; Menzler, N. H. Impact of
M.; Pietzsch, T.; Preibisch, S.; Rueden, C.; Saalfeld, S.; Schmid, B.; defect chemistry on cathode performance: A case study of Pr doped
Tinevez, J. Y.; White, D. J.; Hartenstein, V.; Eliceiri, K.; Tomancak, P.; Ceria. Solid State Ionics 2018, 314, 204−211.
Cardona, A. Fiji: an open-source platform for biological-image analysis. (59) Jaiswar, S.; Mandal, K. D. Evidence of enhanced oxygen vacancy
Nat. Methods 2012, 9, 676−682. defects inducing ferromagnetism in multiferroic CaMn7O12 Manganite
(40) Dutta, P.; Pal, S.; Seehra, M. S.; et al. Concentration of Ce3+ and with sintering time. J. Phys. Chem. C 2017, 121, 19586−19601.
oxygen vacancies in cerium oxide nanoparticles. Chem. Mater. 2006, 18, (60) Wang, D.; Jin, J.; Xia, D.; Ye, Q.; Long, J. The effect of oxygen
5144−5146. vacancies concentration to the gas-sensing properties of tin dioxide-
(41) Sudarsanam, P.; Mallesham, B.; Durgasri, D. N.; Reddy, B. M. doped Sm. Sens. Actuators, B 2000, 66, 260−262.
Physicochemical characterization and catalytic CO oxidation perform- (61) Cheng, S.; Chatzichristodoulou, C.; Søgaard, M.; Kaiser, A.;
ance of nanocrystalline Ce-Fe mixed oxides. RSC Adv. 2014, 4, 11322− Hendriksen, P. V. Ionic/electronic conductivity, thermal/chemical
expansion and oxygen permeation in Pr and Gd co-doped ceria
11330.
PrxGd0.1Ce0.9‑xO1.95‑δ. J. Electrochem. Soc. 2017, 164, F1354−F1367.
(42) Deshpande, S.; Patil, S.; Kuchibhatla, S. V. N.T.; Seal, S. Size
(62) Maheshwari, A.; Wiemhofer, H.-D. Optimized mixed ionic-
dependency variation in lattice parameter and valency states in
electronic conductivity in two-phase ceria-zirconia composite with
nanocrystalline cerium oxide. Appl. Phys. Lett. 2005, 87, 133113.
cobalt oxide and Na2CO3 as suitable additives. J. Mater. Chem. A 2016,
(43) Mansingh, S.; Padhi, D. K.; Parida, K. M. Enhanced visible light
4, 4402−4412.
harnessing and oxygen vacancy promoted N, S co-doped CeO2 (63) Cui, H.; Liu, H.; Shi, J.; Wang, C. Function of TiO2 lattice defects
nanoparticle: a challenging photocatalyst for Cr(VI) reduction. Catal. toward photocatalytic processes: View of electronic driven force. Int. J.
Sci. Technol. 2017, 7, 2772−2781. Photoenergy 2013, 2013, 364802.
(44) Mandal, B.; Mondal, A. Solar light sensitive samarium-doped (64) Ni, M.; Leung, M. K. H.; Leung, D. Y. C.; Sumathy, K. A review
ceria photocatalysts: microwave synthesis, characterization and photo- and recent developments in photocatalytic water-splitting using TiO2
degradation of acid orange 7 at atmospheric conditions and in the for hydrogen production. Renewable Sustainable Energy Rev. 2007, 11,
absence of any oxidizing agents. RSC Adv. 2015, 5, 43081−43091. 401−425.
(45) Song, S.; Xu, L.; He, Z.; Chen, J.; et al. Mechanism of the (65) Queisser, H. J.; Haller, E. E. Defects in semiconductors some
photocatalytic degradation of C.I. reactive black 5 at pH 12.0 using fatal, some vital. Science 1998, 281, 945−950.
SrTiO3/CeO2 as the catalyst. Environ. Sci. Technol. 2007, 41, 5846− (66) Liu, Z.; Li, X.; Koshy, P.; Hart, J. N.; Sorrell, C. C.; et al. Planar-
5853. dependent oxygen vacancy concentrations in photocatalysis by CeO2
(46) Ravishankar, T. N.; Ramakrishnappa, T.; Nagaraju, G.; nanoparticles. CrystEngComm 2018, 20, 204−212.
Rajanaika, H. Synthesis and characterization of CeO2 nanoparticles (67) Nolan, M.; Parker, S. C.; Watson, G. W. The electronic structure
via solution combustion method for photocatalytic and antibacterial of oxygen vacancy defects at the low index surfaces of ceria. Surf. Sci.
activity studies. ChemistryOpen 2015, 4, 146−154. 2005, 595, 223−232.
(47) Channei, D.; Phanichphant, S.; Nakaruk, A.; Mofarah, S. S.; (68) Ganduglia-Pirovano, M. V.; Da Silva, J. L. F.; Sauer, J. Density-
Koshy, P.; Sorrell, C. C. Aqueous and surface chemistries of functional calculations of the structure of near-surface oxygen vacancies
photocatalytic Fe-doped CeO2 nanoparticles. Catalysts 2017, 7, 45. and electron localization on CeO2 (111). Phys. Rev. Lett. 2009, 102,
(48) Fang, Z.; Thanthiriwatte, K. S.; Dixon, D. A.; Andrews, L.; Wang, 026101.
X. Properties of cerium hydroxides from matrix Infrared spectra and
electronic structure calculations. Inorg. Chem. 2016, 55, 1702−1714.
(49) El-Fallah, J.; Hilaire, L.; Roméo, M.; Le Normand, F. Effect of
surface treatments, photon and electron impacts on the ceria 3d core
level. J. Electron Spectrosc. Relat. Phenom. 1995, 73, 89−103.
(50) Roméo, M.; Bak, K.; El-Fallah, J.; Le Normand, F.; Hilaire, L.
XPS study of the reduction of cerium dioxide. Surf. Interface Anal. 1993,
20, 508−512.
(51) Paparazzo, E.; Ingo, G. M.; Zacchetti, N. X-ray induced reduction
effects at CeO2 surfaces: An x-ray photoelectronspectroscopy study. J.
Vac. Sci. Technol., A 1991, 9, 1416.
(52) Zhang, J.; Wong, H.; Yu, D.; Kakushima, K.; Iwai, H. X-ray
photoelectron spectroscopy study of high-k CeO2/La2O3 stacked
dielectrics. AIP Adv. 2014, 4, 117117.
(53) Liu, X.; Zhou, K.; Wang, L.; Wang, B.; Li, Y. Oxygen Vacancy
clusters promoting reducibility and activity of ceria nanorods. J. Am.
Chem. Soc. 2009, 131, 3140−3141.
(54) Grzybek, G.; Stelmachowski, P.; Gudyka, S.; Indyka, P.; Sojka, Z.;
Guillén-Hurtado, N.; Rico-Pérez, V.; Bueno-López, A.; Kotarba, A.
Strong dispersion effect of cobalt spinel active phase spread over ceria
for catalytic N2O decomposition: the role of the interface periphery.
Appl. Catal., B 2016, 180, 622−629.
(55) El-Atwani, O.; Allain, J. P.; Ortoleva, S. In-situ probing of near
and below sputter-threshold ion-induced nanopatterning on
GaSb(100). Nucl. Instrum. Methods Phys. Res., Sect. B 2012, 272,
210−213.

6027 DOI: 10.1021/acs.inorgchem.9b00330


Inorg. Chem. 2019, 58, 6016−6027

You might also like