PreprintPDF Available

Synergy of Single Atoms Pd and Oxygen Vacancies on In2O3 for Highly Selective C1 Oxygenates Production from Methane under Visible Light

Authors:

Abstract and Figures

Methane (CH 4 ) oxidation to high value chemicals under mild conditions through photocatalysis is a sustainable and appealing pathway, nevertheless confronting the critical issues on both conversion and selectivity. Herein, under visible irradiation (420 nm), the synergy of palladium (Pd) atom cocatalyst and oxygen vacancies (OVs) on In 2 O 3 nanorods enabled superior photocatalytic CH 4 activation by O 2 . The optimised catalyst reached ca. 100 µmol·h − 1 of C1 oxygenates, with a selectivity of primary products (CH 3 OH and CH 3 OOH) up to 82.5 %. Mechanism investigation elucidated that such superior photocatalysis was induced by the dedicated function of Pd single atoms and oxygen vacancies on boosting hole and electron transfer pathway, respectively. O 2 was proven to be the only oxygen source for CH 3 OH production, while H 2 O acted as the promoter for efficient CH 4 activation through ·OH production and facilitated product desorption as indicated by DFT modelling. This work thus provides new understandings on simultaneous regulation of activity and selectivity by the significant synergy of single atom cocatalysts and oxygen vacancies.
Content may be subject to copyright.
Page 1/22
Synergy of Single Atoms Pd and Oxygen Vacancies
on In2O3 for Highly Selective C1 Oxygenates
Production from Methane under Visible Light
Lei Luo
Northwest University
Lei Fu
Northwest University
Huifen Liu
Northwest University
Youxun Xu
University College London
Jialiang Xing
Northwest University
Junwang Tang ( junwang.tang@ucl.ac.uk )
University College London https://orcid.org/0000-0002-2323-5510
Article
Keywords: methane conversion, single atom catalyst, methanol, defect engineering, photocatalysis
Posted Date: September 30th, 2021
DOI: https://doi.org/10.21203/rs.3.rs-942037/v1
License: This work is licensed under a Creative Commons Attribution 4.0 International License.
Read Full License
Page 2/22
Abstract
Methane (CH4) oxidation to high value chemicals under mild conditions through photocatalysis is a
sustainable and appealing pathway, nevertheless confronting the critical issues on both conversion and
selectivity. Herein, under visible irradiation (420 nm), the synergy of palladium (Pd) atom cocatalyst and
oxygen vacancies (OVs) on In2O3 nanorods enabled superior photocatalytic CH4 activation by O2. The
optimised catalyst reached ca. 100 µmol·h− 1 of C1 oxygenates, with a selectivity of primary products
(CH3OH and CH3OOH) up to 82.5 %. Mechanism investigation elucidated that such superior
photocatalysis was induced by the dedicated function of Pd single atoms and oxygen vacancies on
boosting hole and electron transfer pathway, respectively. O2 was proven to be the only oxygen source for
CH3OH production, while H2O acted as the promoter for ecient CH4 activation through ·OH production
and facilitated product desorption as indicated by DFT modelling. This work thus provides new
understandings on simultaneous regulation of activity and selectivity by the signicant synergy of single
atom cocatalysts and oxygen vacancies.
Introduction
As the predominant constituent of natural gas, methane hydrate and shale gas resources, selective
methane (CH4) oxidation to value-added chemicals holds considerable nancial and environmental
prospective [1–5]. However, the inert symmetrical tetrahedral structure of CH4 makes it rather dicult for
the dissociation of the rst C-H bond, which is the most important step for activation of methane. [6–8]
Industrial multistep route via steam reforming and subsequent Fischer-Tropsch synthesis could eciently
activate CH4, while it requires harsh experimental conditions (eg. > 700 oC temperature and/or high
pressure), causing huge energy-consumption and safety issues [9–13]. In parallel, it is relatively dicult
to achieve high selectivity due to the more reactive characteristics of the desired oxygenates against both
the reactant CH4 and stable product CO2. [14–17] Therefore, selective CH4 conversion to value-added
chemicals under mild conditions other than CO2 is highly attractive, while confronting considerable
challenges.
Photocatalysis offers an appealing alternative to drive many tough redox reactions under mild conditions
including CO2 conversion [18, 19], N2 reduction[20] and selective CH4 oxidation [8]. Recently, various
value-added chemicals such as methanol [1, 21–23], formaldehyde [24, 25], ethanol [26, 27], ethane and
ethylene [28–33] were produced by photocatalysis. For example, we found that up to 90 % selectivity with
a yield of 3.5 µmol·h− 1 methanol could be achieved over the optimized FeOx/TiO2 photocatalyst under
ambient condition using H2O2 as an oxidant [22]. Recently a high yield of liquid oxygenates including
CH3OH, CH3OOH and HCHO were produced under full arc irradiation over Au supported ZnO, together with
the good selectivity of primary products (CH3OH and CH3OOH) (< 70 %) [1]. Very recent, the yields of 18.7
µmol·h− 1 HCHO and 3.7 µmol·h− 1 CH3OH were reported on quantum BiVO4 with an excellent selectivity
toward HCHO (87 %) and CH3OH (99 %) under 300–400 nm or 400-780nm irradiation[25]. Given these
Page 3/22
signicant advances in photocatalytic methane conversion, the yield and/or selectivity to high value
chemicals are still quite moderate, in particular it is very challenging to achieve methane activation under
visible light irradiation instead of a full arc spectrum due to a narrowed bandgap with mitigated reduction
or oxidation potentials
To realize visible driven methane oxidation by O2 gas on narrow bandgap photocatalysts, cocatalyst is
crucial that does not only promote charge separation, more importantly manipulates the activation
energy of the methane conversion and the selectivity [34–39]. Furthermore rationally regulating the
production of reactive oxygen species (ROS) through cocatalyst modication is necessary as ·OH
radicals have been widely regarded as the main species that induced CH4 activation and over-oxidation
[40, 41]. When CH3OH served as the desired products, over-oxidation to HCHO or CO2 would be
suppressed by lowering the oxidative potential of photogenerated hole through cocatalyst modication,
thus improving the selectivity. Stimulated by molecular catalysis, single atom cocatalysts promise an
extremely high eciency, where atomic dispersed species with unsaturated coordination environment
could improve the catalytic performances based on the unique electronic structure [42–44]. Meanwhile,
high atom utilization eciency could be achieved [45, 46]. On the other hand, since CH4 exhibited low
electron and proton anity, moderate decoration of defective sites could enhance the chemical-
adsorption of non-polar molecular, then promoting the activation of CH4 [47]. Therefore, the integration of
both defects and single atom cocatalyst decoration could boost charge separation, weaken oxidative
potential and enhance CH4 activation on a photocatalyst.
Herein, atomically dispersed palladium (Pd) supported on defective In2O3 was prepared and served as the
visible-light responsive photocatalyst for CH4 conversion to high value chemicals. Under 420 nm
irradiation, the optimized production of oxygenates reached up to ca. 300 µmol in 3 h, with a very high
selectivity of 82.5 % of the primary products. In-situ XPS and EPR spectra were conducted to investigate
the charge transfer dynamics. The results indicated the dedicated roles of Pd atoms and oxygen
vacancies (OVs) in promoting the transfer of photo-induced holes and electrons, respectively. DFT
calculation results indicated H2O could also promoted the desorption of the oxygenate products, thus
suppressing over-oxidation and facilitate high selectivity of primary products. The introduction of atomic
Pd and oxygen vacancies further enhanced this effect on suppressing over-oxidation. Isotopic labelled
experiments further proved the methane conversion pathway.
Results And Discussion
Visible-light photocatalytic CH4 oxidation by O2
Atomic Pd cocatalyst was prepared by the in-situ photo-deposition method withK2PdCl4 and (NH4)2PdCl4
as the precursors on the visible driven In2O3 nanorod photocatalyst. Two types of photocatalysts were
synthesized as the defect-rich and defect-lean materials, denoted as Pd-def-In2O3 and Pd-In2O3,
respectively. For comparison, other noble metals including Pt and Au modied photocatalysts were also
Page 4/22
prepared with the same dosage, denoted as M-(def)-In2O3 (M = Pt and Au). With different K2PdCl4
dosage, the as-prepared samples were denoted as Pdx-def-In2O3, where x % represented the weight
percentage of Pd to In2O3. In the following discussion, the best sample Pd-def-In2O3 and the reference Pd-
In2O3 referred to Pd0.1-def-In2O3 and Pd0.1-In2O3 unless otherwise specied.
Typical noble metal cocatalysts (Pt, Pd, Au) loaded onIn2O3 nanorods were rst tested via photocatalytic
CH4 conversion with O2 as the oxidant (Figure 1a). Under 420 nm irradiation, the products including
CH3OH, CH3OOH and HCHO over Pd-In2O3 reaches 13.4, 32.3 and 27.5 μmol in 3 h reaction, respectively.
The selectivity of the primary products (CH3OH and CH3OOH) was 62.1 % and the selectivity to the
overoxidation products (HCHO and CO2) was 37.9 %. In comparison, Au-In2O3 and Pt-In2O3 performed
almost 100 % over-oxidized products (HCHO), with the trace yields of 1.4 and 0.9 μmol HCHO,
respectively. Such differences suggested Pd cocatalyst was more suitable than Pt and Au for CH4
activation to produce these primary products. The yield of oxygenates for Pd- In2O3 was improved further
to 179.7 μmol by the introduction of defective sites to form Pd-def-In2O3, 2.5 times higher than that of Pd-
In2O3 (73.2 μmol). Meanwhile, the selectivity of the primary products was improved from 62.1 % to 80.4
%, suggesting that deep-oxidation to HCHO and CO2 was greatly suppressed under the synergy of Pd
single atoms and oxygen vacancies (OVs). In the case of Au-def-In2O3 and Pt-def-In2O3, defect
modication exhibited the similar phenomenon on promoting CH4 conversion although the yield was
much lower than that achieved on the Pd modied photocatalyst.
The effect of Pd single atoms was explored over Pdx-def-In2O3. As shown inFigure 1b, Pdx-def-In2O3
photocatalysts exhibited much higher oxygenates production than that of the pristine In2O3. With the
raising of K2PdCl4 dosage, the production of the liquid oxygenates exhibited the volcanic trend,
increasing from 48.7 μmol on Pd0.01-def-In2O3 to 179.7 μmol on Pd0.1-def-In2O3. Furthermore, the
increasing K2PdCl4 dosage resulted in the decreased photocatalytic performance. Notably, the selectivity
exhibited slight improvement from 76.5 % to 82.8 %. With a close look at the above results, OVs and Pd
cocatalyst played the synergistic role in optimizing the activity and selectivity for photocatalytic CH4
conversion.
Molar ratio of CH4 to O2 was tuned over Pd-def-In2O3 (Figure 1c). The production of oxygenates
demonstrated the volcanic trend again, with the highest oxygenate production and selectivity achieved at
1 bar O2 pressure. Reducing molar ratio of CH4 to O2 caused the decreased production to 116.0 μmol of
primary products when CH4/O2 = 10/10, mainly ascribed to the decrease of CH4 concentration. In parallel,
the increased concentration of O2 induced over-oxidation and decreased selectivity of the primary
products from 80.4 % to 67.6 %. With the increase of H2O dosage (Figure 1d), the production of
oxygenates gradually increased, reaching the highest value with 100 mL H2O dosage. The highest
oxygenates achieved 299.0 μmol, 2.3 times improvement than that of 25 mL dosage (128.0 μmol) over
Pd-def-In2O3. Moreover, the selectivity of the primary products improved from 74.2 to 82.5 % with H2O
Page 5/22
dosage increasing from 25 to 100 mL, which could be attributed to the enhanced desorption of the
products from the surface of the photocatalyst when more water was used as discussed later. Notably, in
the absence of H2O dosage, CO2 (8.5 μmol) was produced as the only product, suggesting the critical role
of H2O in promoting CH4 activation as well suppressing over-oxidation, probably ascribed to the
production of ·OH radical and promotion desorption of oxygenates by H2O[48]. While increasing the total
pressure of the gaseous reactants, CH4 dissolved increased and the oxygenate production gradually
increased (Figure 1e), e.g. only trace amount of HCHO (4.1 μmol) produced at 1 bar and reaching the
highest yield of 179.7 μmol when the pressure was 20 bar. To investigate the stability of the optimized
photocatalyst, we carried out thecycling test experiment over Pd-def-In2O3 photocatalyst. No obvious
decrease of oxygenates was observed under 15 hours reaction (Figure 1f), demonstrating the good
stability of Pd-def-In2O3.
Structural identication
X-ray diffraction (XRD) patterns were recorded to probe the crystalline structure of the representative
photocatalysts (In2O3, Pd-In2O3 and Pd-def-In2O3) (Figure S1). The diffraction peaks on all three samples
at 30.7o, 35.5o, 51.0o and 60.7o were well matched with the standard phase of In2O3 (PDF#71-2194).
While no Pd and PdOx diffraction peaks were observed onPd-In2O3 and Pd-def-In2O3, indicating the high
dispersion of Pd species. The slightly weakened relative intensity from 100 % of In2O3 to 97 % and 93 %
of Pd-In2O3 and Pd-def-In2O3 could be probably ascribed to the introduction of defects. Raman spectra
(Figure 2a) further supported the well-established In2O3 phase. The typical Raman peaks for In2O3 were
clearly observed at 130.6, 305.1 and 494.8 cm-1 [49]. For Pd-In2O3 and Pd-def-In2O3, the dominant peak
exhibited a slight left-shift from 130.6 to 129.9 cm-1, attributed to the surface stain effect induced by the
Pd cocatalyst deposition[50].
Electron paramagnetic resonance (EPR) spectra were conducted to evaluate the spin-electrons including
oxygen vacancies (Figure 2b). For the pristine In2O3 and Pd-In2O3, a single Lorentz peak at g = 1.882 was
observed, ascribed to the electrons on the conduction band (CB)[51, 52]. In the case of Pd-def-In2O3, the
signal of this peak exhibited much stronger intensity than the others, suggesting the higher electron
density on CB. Meanwhile, an additional Lorentz peak was observed at g = 2.001, which could be
attributed to the free-electrons trapped by the oxygen vacancies[52], thus suggesting the existence of
oxygen vacancies in Pd-def-In2O3. The introduction of oxygen vacancies might contribute to the stronger
EPR peak at g = 1.882.
High-resolution transmission electron microscope (HRTEM) images further proved the defective structure
of Pd-def-In2O3 (Figure 2c-d). Pd-def-In2O3 reserved the nanorod morphology with dimension of 203 nm
in diameter and 1450 nm in length (insert ofFigure 2c) as that of the In2O3 substrate and Pd-In2O3
(Figure S2). In addition, a thin amorphous/defective layer of ca. 4 nm was observed on the edge (Figure
Page 6/22
2d). The bulk crystal plane distance of 0.415 nm was indexed to the (211) facet of In2O3. On the contrary,
there is not such defective layer as indicated in Figure S2b. Elemental distribution of the corresponding
area was analysed by the EDS-mapping images (Figure 2e). Obviously, Pd-def-In2O3 exhibited uniform
palladium distribution with indium and oxygen elements. The aberration corrected HAADF-STEM image
inFigure 2f clearly indicated the atomic distribution of Pd, where the weak intensity spots cycled by the
yellow corresponded to Pd atoms.The x-y line scan along the yellow rectangle ofFigure 2f clearly
presents the atomic dispersion of Pd as shown inFigure 2g. Therefore the best sample was composed of
single atom Pd and oxygen vacancies on In2O3 nanorods.
Mechanism investigation
UV-vis diffuse reection spectra (UV-DRS) were conducted to evaluate the photoabsorption ability of the
representative photocatalysts (In2O3, Pd-In2O3 and Pd-def-In2O3) (Figure 3a). All the three photocatalysts
exhibited the similar photoabsorption onset at ca. 450 nm, indicating that the modication of single atom
Pd and OVs has little inuence on bandgap energy and bandgap energy was not the decisive reason that
induced improvement of photocatalysis.
In-situ high-resolution Pd3d X-ray photoelectric spectra (XPS) in dark and under light were conducted to
study the charge transfer direction of Pd-def-In2O3 (Figure 3b and Figure S3). In dark, the Pd3d5/2 XPS
peak could be deconvoluted into two binding peaks at 336.55 and 335.38 eV, which were assigned to the
Pd2+ and Pd0 species, respectively [53]. Under light irradiation, the peak exhibited a left-shift to higher
binding energy (Figure S3). Further deconvoluted results (Figure 3b) suggested that Pd2+ content
increased to 26.3 %, much higher than 6.9 % in dark. Such increased Pd2+ content suggested Pd served
as the hole acceptors upon excitation. In-situ EPR spectra under light were conducted to evaluate the role
of OVs. As shown in Figure 3c, the signal at g = 2.0009 was attributed to the electrons trapped by the
oxygen vacancies, which performed gradually increasing intensity from 100 % to 226 % with the
prolonged irradiation to 360 seconds. This stronger EPR intensity suggested a higher concentration of
spin-electrons and thus demonstrated OVs served as the electron acceptor [54]. Therefore, single atom Pd
and OVs separately acted as the hole and electron acceptors under light irradiation, which would greatly
contribute to the enhanced charge separation.
Photocurrent responses (Figure S4) were tested to evaluate the charge separation eciency. Pristine
In2O3 exhibited a relatively low photocurrent density of 61.8 μA·cm-2. After photo-depositing high
dispersed Pd cocatalyst, Pd-In2O3 nearly doubled photocurrent density to 120.6 μA·cm-2. The
photocurrent density was further improved to 168.1 μA·cm-2 on Pd-def-In2O3, almost 2.7 and 1.4 times
enhancement than that of In2O3 and Pd-In2O3, respectively. Such highest photocurrent density on Pd-def-
In2O3 attributed to the most ecient charge separation, indicating the defects and single atom Pd could
greatly enhance charge transfer, which is consistent with the analysis mentioned above. Steady-state
Page 7/22
uorescence (PL) spectra further evidenced the enhanced charge separation eciency. As shown
inFigure 3d, a relatively strong PL emission peak was observed for the pristine In2O3, attributed to the
severe charge recombination. In comparison, the PL intensity for Pd-In2O3 was greatly weakened,
indicating the suppressed charge recombination. For Pd-def-In2O3 photocatalyst, the most weakened PL
peak were observed, ascribed to the most enhanced charge separation eciency, which was
corresponding with the photocurrent analysis. Time-decay PL spectra were conducted to evaluate the PL
lifetime. As shown in Figure S5, Pd-def-In2O3 photocatalyst exhibited the slowest PL decay kinetics. The
tting results (Table S1) showed that Pd-def-In2O3 exhibited the average PL lifetime at 4.99 ns, longer
than that of In2O3 (3.60 ns) and Pd-In2O3 (4.28 ns), which would be benecial to the ecient utilization of
separated charge carriers.
Reactive oxygen species including ·OOH and ·OH radicals were widely regarded as the main active
species for CH4 activation[55]and monitored by in-situ EPR spectra with 5, 5-dimethyl-1-pyrroline N-oxide
(DMPO) as the spin-electron trapping agents. As shown inFigure 4a, the DMPO-OOH adduct was
detected under light over different photocatalysts and ascribed to the presence of ·OOH, which came from
the reduction of O2 molecule with photo-induced electrons and H+.Astronger intensity of DMPO-OOH
was observed forPd-def-In2O3,suggesting the production of ·OOH radical was enhanced by the
integration of single atom Pd and OVs.On the other hand, in-situ EPR spectra under light was used to
monitor the generation of ·OH radical with DMPO as the trapping agent in H2O. The 1:2:2:1 quartet
signals wereobserved and assigned to the DMPO-OH adduct, suggesting the generation of ·OH radical
(Figure 4b). It was obvious thatPd-def-In2O3 produced much more ·OH under identical conditions than
Pd-In2O3 and In2O3was the worse. It is believed that ·OHinitially activates CH4 to methyl radical (·CH3),
thus Pd-def-In2O3 performed CH4 activation best followed byPd-In2O3, which is consistent with the step
by step enhanced photocatalytic performances by Pd and then both Pd and oxygen vacancies, indicating
that oxygen vacancies could promote charge separation and also facilitate water oxidation reaction on
Pd. Coumarin was used as the probe for ·OH radical detection due to the easy reaction between coumarin
and ·OH to produce 7-hydroxycoumain (7-HC) that could be detected by UV-vis spectra at 412 nm (Figure
4c). The results further supported that Pd-def-In2O3 held the strongest ability for ·OH production, which
facilitated CH4 activation. Therefore, single atom Pd worked as the hole acceptor, which then catalysed
·OH radical production from water oxidation. Simultaneously, OVs acted as the electron acceptor, which
then catalysed O2 reduction to generate ·OOH radical.
The reaction pathway was investigated by isotopic labelled experiments, including using H218O and 18O2.
In the presence of 3 mL H218O, 1 bar 16O2 and 19 bar CH4, no isotopic labelled CH318OH (m/z = 33 and
34) was detected by GCMS (Figure 4d), suggesting H2O was not the oxygen source that directly
participated the formation of oxygenates. In parallel, when using 3 mL H216O, 1 bar 18O2 and 19 bar CH4,
the signals at m/z = 34 and 33 were attributed to the isotopic labelled CH318OH and its fragment (Figure
4d), suggesting O2 was the only oxygen source for CH3OH formation. Carbon source for methanol
Page 8/22
production were also studied in the presence of 5 bar isotopic labelled 13CH4 (Figure 4e), where the signal
of mass spectra (MS) at m/z = 33 was ascribed to 13CH3OH, demonstrating that CH4 was the carbon
source for oxygenates production.
DFT calculations (Figure 5) were conducted to explain the improved selectivity of primary products. It
should be noted that timely desorption of the primary products on the active sites could eciently avoid
its deep-oxidation to HCHO and CO2. As ·OH radical was regarded as the main species that induced
oxidation on single atom Pd cocatalyst, it was accordingly considered that the ecient desorption of
primary products like CH3OH on Pd is critical to suppress further oxidation. Thus, the adsorption energies
of H2O and CH3OH were calculated since the stronger adsorption of H2O might promote the desorption of
CH3OH. H2O and CH3OH adsorption on In2O3 and on Pd/ In2O3 were modelled and optimized by the
density functional theory (DFT) calculations. As shown in Figure 5a and b, the adsorption energies of H2O
on In2O3, Pd-In2O3 and Pd-def-In2O3 were -0.76, -1.57 and -2.14 eV, respectively, much larger than the
CH3OH adsorption energy of -0.47, -1.38 and -1.50 eV on the specied model. Such larger adsorption
energies indicate CH3OH could be easily replaced by H2O on In2O3 or Pd atoms, promoting the desorption
of CH3OH. Moreover, adsorption energies further increased with the introduction of both Pd atom and
oxygen vacancy, demonstrating such co-modication of Pd atoms and OVs could promote the adsorption
of H2O most eciently, which is consistent with the increased production of ·OH radicals as analyzed by
the in-situ EPR and coumarin experiments. Though the adsorption of CH3OH was also enhanced due to
the introduction of Pd atoms and OVs, H2O adsorption energy was enhanced much more, thus water
could facilitate the desorption of primary products and avoid overoxidation as indicated in Figure 1d.
Based on the above results, aerobic photocatalyticCH4 conversion mechanism over Pd-def-In2O3 was
proposed (Scheme 1). Generally, under 420 nm irradiation, electrons (e-) were excited to the conduction
band of In2O3 nanorod photocatalyst, while leaving holes (h+) on the valence band. Then the photo-
induced electrons were trapped by the oxygen vacancies, activating O2 with H+ to produced ·OOH radicals
as detected by the in-situ EPR spectra. In parallel, Pd atoms served as the hole acceptors (Pd + h+
Pdδ+), and then reacted with the adsorbed H2O to produced ·OH (Pdδ+ + H2O Pd0 + ·OH + H+). CH4
molecules were next activated by the as-produced ·OH to ·CH3. The coupling reaction between ·CH3 and
·OOH then generated the primary products (CH3OOH), and subsequently transferred to CH3OH as
indicated in Scheme 1b. Compared with the pristine In2O3, the incorporation of Pd single atoms
signicantly promoted charge separation and facilitated the generation of reactive species, thus
promoting CH4 conversion to oxygenates. Pd atoms loading also moderated the oxidation ability of
photoinduced holes on In2O3 as indicated in Scheme 1a, reducing the overoxidation of the primary
products. Further decoration of oxygen vacancies could strengthen the promoted charge separation
eciency, which eventually resulted in the superior photocatalytic CH4 conversion activity and selectivity.
In order to suppress over-oxidation, it was also critical to enhance the desorption of primary oxygenate
products by H2O, as supported by the DFT calculation.
Page 9/22
In summary, visible-light-driven CH4 conversion at ambient temperature was reported over the In2O3
nanorod photocatalyst with loading of Pd single atoms cocatalysts and oxygen vacancies. Under 420 nm
irradiation, superior yield (99.7 μmol·h-1) and selectivity (82.5 %) of the primary products were achieved
on Pd-def-In2O3 photocatalyst under optimized reaction conditions. In-situ XPS and EPR spectra under
visible light irradiation indicated that Pd and oxygen vacancies acted as the hole and electron acceptors,
respectively, thus synergistically boosted charge separation and transfer. Isotopic labelled experiment
proved that O2 was the only oxygen source for oxygenates production, while H2O was the promoter of
CH4 activation through the production of ·OH radical as monitored by the in-situ EPR spectra with DMPO
as the spin-trapping agent. DFT calculation results suggested that H2O performed much larger adsorption
energies than CH3OH on either In2O3,def-In2O3 orPd-def-In2O3, suggesting the stronger adsorption of
H2O than CH3OH, which was benecial to the timely desorption of the produced CH3OH, thus avoiding
further over-oxidation. The introduction of Pd and oxygen vacancies could further improve the selectivity
of primary oxygenates mainly through the greatly enhanced adsorption of H2O and the reduced oxidation
potential of photoinduced holes. This work provided an useful avenue on co-modication by oxidative
single atom cocatalyst and oxygen vacancies for simultaneous regulation of both activity and selectivity
through enhancing charge separation, moderated photohole oxidation ability and timely promoted
desorption of primary products by a solvent.
Methods
Synthesis of In2O3 nanorods
In2O3 nanorods were prepared according to the previous study[56]. Typically, 12.0 g urea and 1.5 g InCl3
were dissolved in 135 g H2O, followed by stirring at 80 oC for 14 h. After naturally cooling down, the
reactant was centrifuged and washed with H2O for several times. The white powder was then dried in
vacuum at 60 oC overnight. After thermal-treating in air at 700 oC for 5 h at a ramping rate of 5 oC·min-1,
yellow powder was obtained and denoted as In2O3.
Synthesis of Pdx-In2O3 and Pdx-def-In2O3 nanorods
Pdx-In2O3 and Pdx-def-In2O3 nanorods were prepared through photo-deposition with ammonium
tetrachloropalladate(II) ((NH4)2PdCl4) and potassium tetrachloropalladate(II) (K2PdCl4) as the precursors,
respectively. The synthesis was conducted in the multichannel reactor (
Beijing Perfectlight Technology
Co., Ltd
). For Pdx-def-In2O3 preparation, 200 mg In2O3 was rst dispersed through sonication with the
aqueous solution containing 10 vol.% methanol. Then certain amount of K2PdCl4 solution was added.
After sealing and purging with ultrapure argon (99.999 vol.%) for 30 min, the reactor was bottom-
irradiated for 3 h to facilitate Pd photo-deposition. The suspension was then centrifuged, washed with
H2O for several times and dried under vacuum at 60 oC overnight. The as-prepared samples were denoted
as Pdx-def-In2O3, where x % represented the mass percentage of palladium to In2O3 substrates. Pdx-In2O3
Page 10/22
with no oxygen vacancies was prepared under identical conditions except the usage of (NH4)2PdCl4
instead of K2PdCl4.
Characterizations
XRD were measured to obtain the crystalline structure on the
D8 ADVANCE
diffractometer (
Bruker Co.,
Ltd
).Palladium and potassium contents were measured through inductively coupled plasma atomic
emission spectrometry (ICP-AES) on the
Agilent 7900 ICP-MS
instrument. Raman spectra were collected
on the
DXR 2DXR2
instrument (
Thermo Fisher Scientic, Co., Ltd
). HRTEM images were captured on
the
Talos F200X
instrument (
FEI Co., Ltd
). UV-DRS spectra were measured on the UV-3600 plus
spectrophotometer (
ShimadzuCo., Ltd
). Photocurrent test was conducted on the electronic workstation
(
CHI660E
) on thethree-electrode system.Ag/AgCl electrode, platinum sheet electrode and Na2SO4
solution (0.1 M) were used as the reference electrode, counter electrode and electrolyte, respectively. The
mixtures of photocatalyst, ethanol and Nion solution (
Shanghai Adamas Reagent Co., Ltd
) were
suspended and sonicated to prepare the working electrode.In-situ XPS in dark and under light were
measured on the
Thermo ESCALAB 250Xi
instrument with an Al Kα radiation source. In-situ solid-state
EPR spectra in dark and under light were measured with 20 mg photocatalyst on the
ELEXSYS II
EPR
instrument.
PhotocatalyticCH4conversion
Photocatalytic CH4 conversion was conducted in a top-irradiated high-pressure reactor with 200 mL
volume. LED lamp (420 nm,
PLS-LED100C
,
Beijing Perfectlight Technology Co., Ltd
) was used as the light
source. Typically, 20 mg photocatalyst was dispersed in 50 mL distilled water. After sealing and purging
with ultrapure O2 (99.999 vol.%) for 20 min, 1 bar O2 and 19 bar CH4 (99.999 vol.%) were owed into the
reactor. The temperature of the reactor was maintained at 25 oC by the cold-water bath. After reacting for
3 h, the gaseous and liquid products like methanol were measured by the gas chromatography
(
GC2014
,
ShimadzuCo., Ltd
) equipped with thermal conductivity detector (TCD) and ame ionization
detector (FID). CH3OOH and CH3OH were measured through 1H nuclear magnetic resonance
(NMR)(
AVANCE III
,
JEOL Ltd
). As CH3OOH and CH3OH have the same number of methyl, the area ratio of
CH3OOH to CH3OH in 1H NMR should be the molar ratio of CH3OOH to CH3OH. Thus, CH3OOH could be
quantied. HCHO was measuredthrough the colorimetric method[57]on the
UV-3600 Plus
spectrometer
(
Shimadzu Co., Ltd
).
Isotope labelling experiments
For carbon source investigation: 20 mg Pd-def-In2O3 photocatalyst was dispersed in 3 mL H2O. After the
reactor being degassed for 30 min, 1 bar O2 and 5 bar 13CH4 were injected into the reactor. After reacting
for 6 h, the suspension was ltered and then the solvent was analysed by GC-MS (QP2010,
ShimadzuCo.,
Ltd
) equipped withthe Cap WAX column.
Page 11/22
For oxygen source investigation: 20 mg Pd-def-In2O3 photocatalyst was dispersed in 3 mL H216O or
H218O. After the reactor being degassed for 30 min, 1 bar 18O2 or 16O2 and 5 bar CH4 were injected into
the reactor. After reacting for 6 h, the suspension was ltered and then the solvent was analysed by GC-
MS (QP2010,
ShimadzuCo., Ltd
).
Monitor of the reactive species
DMPO was used as the spin-trapping agent for monitor of the reactive species including ·OOH and ·OH
radicals. For ·OOH radical detection, 10 mg Pd-def-In2O3 photocatalyst was dispersed into 5 mL
DMPO/methanol solution. After purging with ultrapure O2 (99.999 vol.%) for 20 min, in-situ EPR spectra in
dark and under light irradiation were collected. For ·OH radical detection, 10 mg Pd-def-In2O3
photocatalyst was dispersed in 5 mL aqueous DMPO solution. After purging with ultrapure O2 (99.999
vol.%) for 20 min, in-situ EPR spectra in dark and under light were collected.
Analysis of hydroxyl radical (·OH)
Coumarin was used as the probe for the quantication of ·OH via the production of 7-HC[40]. Typically,
20 mg photocatalyst was dispersed in 100 mL aqueous coumarin solution (5×10-4M). After stirring for 30
min in dark, the suspension was irradiated with the LED light source (420 nm,
PLS-LED100C
,
Beijing
Perfectlight Technology Co., Ltd
). Certain amount of suspension was sampled and ltered in the 10 min
intervals. PL intensity of the produced 7-HC in the solution was then measured on the
F4500
spectrouorometer.
DFT calculation of adsorption energies
The rst-principles were employed to perform all the density functional theory (DFT) calculations within
the generalized gradient approximation (GGA) using the PBE formulation. The projected augmented wave
(PAW) potentials have been chosen to describe the ionic cores and take valence electrons into account
using a plane wave basis set with a kinetic energy cutoff of 400 eV. Partial occupancies of the
Kohn−Sham orbitals were allowed using the Gaussian smearing method and a width of 0.05 eV. The
electronic energy was considered self-consistent when the energy change was smaller than 10−4 eV. A
geometry optimization was considered convergent when the force change was smaller than 0.05 eV/Å.
Grimme’s DFT-D3 methodology was applied to describe the dispersion interactions. Three models
including In2O3 with (111) facet, def-In2O3 with one oxygen vacancy and Pd-def-In2O3 with both one
oxygen vacancy and single atom Pd modication were conducted. During structural optimizations, the
2×2×1 Monkhorst-Pack k-point grid for Brillouin zone was used for k-point sampling for structures.
Finally, the adsorption energies (
Eads
) were calculated as
Eads
=
Ead/sub
-
Ead
-
Esub
, where
Ead/sub
,
Ead
, and
Esub
were the total energies of the optimized adsorbate/substrate system, the adsorbate in the structure,
and the clean substrate, respectively.
Page 12/22
Data availability
The data that support the ndings of this study are available from the corresponding author upon
reasonable request.
References
[1] Song, H., X. Meng, S. Wang, W. Zhou, X. Wang, T. Kako, and J. Ye
.
Direct and Selective Photocatalytic
Oxidation of CH4 to Oxygenates with O2 on Cocatalysts/ZnO at Room Temperature in Water.
Journal of
the American Chemical Society
, 2019,
141
, 20507.
[2] Sushkevich, V.L., D. Palagin, M. Ranocchiari, and J.A.v. Bokhoven
.
Selective Anaerobic Oxidation of
Methane Enables Direct Synthesis of Methanol.
Science
, 2017,
356
, 523.
[3] Sher Shah, M.S.A., C. Oh, H. Park, Y.J. Hwang, M. Ma, and J.H. Park
.
Catalytic Oxidation of Methane to
Oxygenated Products: Recent Advancements and Prospects for Electrocatalytic and Photocatalytic
Conversion at Low Temperatures.
Advanced Science
, 2020,
7
, 2001946.
[4] Hu, D., V.V. Ordomsky, and A.Y. Khodakov
.
Major Routes in the Photocatalytic Methane Conversion into
Chemicals and Fuels under Mild Conditions.
Applied Catalysis B: Environmental
, 2021,
286
, 119913.
[5] Meyet, J., K. Searles, M.A. Newton, M. Wörle, A.P. van Bavel, A.D. Horton, J.A. van Bokhoven, and C.
Copéret
.
Monomeric Copper(Ii) Sites Supported on Alumina Selectively Convert Methane to Methanol.
Angewandte Chemie International Edition
, 2019,
58
, 9841.
[6] Ravi, M., M. Ranocchiari, and J.A. van Bokhoven
.
The Direct Catalytic Oxidation of Methane to
Methanol-a Critical Assessment.
Angewandte Chemie International Edition
, 2017,
56
, 16464.
[7] Zakaria, Z. and S.K. Kamarudin
.
Direct Conversion Technologies of Methane to Methanol: An
Overview.
Renewable and Sustainable Energy Reviews
, 2016,
65
, 250.
[8] Song, H., X. Meng, Z.-j. Wang, H. Liu, and J. Ye
.
Solar-Energy-Mediated Methane Conversion.
Joule
,
2019,
3
, 1606.
[9] Agarwal, N., S.J. Freakley, R.U. McVicker, S.M. Althahban, N. Dimitratos, Q. He, D.J. Morgan, R.L.
Jenkins, D.J. Willock, S.H. Taylor, C.J. Kiely, and G.J. Hutchings
.
Aqueous Au-Pd Colloids Catalyze
Selective CH4 Oxidation to CH3OH with O2 under Mild Conditions.
Science
, 2017,
358
, 223.
[10] Hammond, C., M.M. Forde, M.H. Ab Rahim, A. Thetford, Q. He, R.L. Jenkins, N. Dimitratos, J.A. Lopez-
Sanchez, N.F. Dummer, D.M. Murphy, A.F. Carley, S.H. Taylor, D.J. Willock, E.E. Stangland, J. Kang, H.
Hagen, C.J. Kiely, and G.J. Hutchings
.
Direct Catalytic Conversion of Methane to Methanol in an Aqueous
Medium by Using Copper-Promoted Fe-ZSM-5.
Angewandte Chemie International Edition
, 2012,
51
, 5129.
Page 13/22
[11] Okolie, C., Y.F. Belhseine, Y. Lyu, M.M. Yung, M.H. Engelhard, L. Kovarik, E. Stavitski, and C.
Sievers
.
Conversion of Methane into Methanol and Ethanol over Nickel Oxide on Ceria-Zirconia Catalysts
in a Single Reactor.
Angewandte Chemie International Edition
, 2017,
56
, 13876.
[12] Ma, J., K. Mao, J. Low, Z. Wang, D. Xi, W. Zhang, H. Ju, Z. Qi, R. Long, X. Wu, L. Song, and Y.
Xiong
.
Ecient Photoelectrochemical Conversion of Methane into Ethylene Glycol by WO3 Nanobar
Arrays.
Angewandte Chemie International Edition
, 2021,
60
, 9357.
[13] Zhou, Y., H. Wang, X. Liu, S. Qiao, D. Shao, J. Zhou, L. Zhang, and W. Wang
.
Direct Piezocatalytic
Conversion of Methane into Alcohols over Hydroxyapatite.
Nano Energy
, 2021,
79
, 105449.
[14] Schwach, P., X. Pan, and X. Bao
.
Direct Conversion of Methane to Value-Added Chemicals over
Heterogeneous Catalysts: Challenges and Prospects.
Chemical Reviews
, 2017,
117
, 8497.
[15] Tian, J., J. Tan, Z. Zhang, P. Han, M. Yin, S. Wan, J. Lin, S. Wang, and Y. Wang
.
Direct Conversion of
Methane to Formaldehyde and CO on B2O3 Catalysts.
Nature Communications
, 2020,
11
, 5693.
[16] Yu, T., Z. Li, L. Lin, S. Chu, Y. Su, W. Song, A. Wang, B.M. Weckhuysen, and W. Luo
.
Highly Selective
Oxidation of Methane into Methanol over Cu-Promoted Monomeric Fe/ZSM-5.
ACS Catalysis
, 2021,
11
,
6684.
[17] Jin, Z., L. Wang, E. Zuidema, K. Mondal, M. Zhang, J. Zhang, C. Wang, X. Meng, H. Yang, C. Mesters,
and F.-S. Xiao
.
Hydrophobic Zeolite Modication for in Situ Peroxide Formation in Methane Oxidation to
Methanol.
Science
, 2020,
367
, 193.
[18] Wang, Y., X. Liu, X. Han, R. Godin, J. Chen, W. Zhou, C. Jiang, J.F. Thompson, K.B. Mustafa, S.A.
Shevlin, J.R. Durrant, Z. Guo, and J. Tang
.
Unique Hole-Accepting Carbon-Dots Promoting Selective
Carbon Dioxide Reduction Nearly 100% to Methanol by Pure Water.
Nature Communications
, 2020,
11
,
2531.
[19] Wang, Y., R. Godin, J.R. Durrant, and J. Tang
.
Ecient Hole Trapping in Carbon Dot/Oxygen-Modied
Carbon Nitride Heterojunction Photocatalysts for Enhanced Methanol Production from CO2 under Neutral
Conditions.
Angewandte Chemie International Edition
, 2021,
60
, 20811.
[20] Han, Q., C. Wu, H. Jiao, R. Xu, Y. Wang, J. Xie, Q. Guo, and J. Tang
.
Rational Design of High-
Concentration Ti3+ in Porous Carbon-Doped TiO2 Nanosheets for Ecient Photocatalytic Ammonia
Synthesis.
Adv Mater
, 2021,
33
, e2008180.
[21] Zhou, W., X. Qiu, Y. Jiang, Y. Fan, S. Wei, D. Han, L. Niu, and Z. Tang
.
Highly Selective Aerobic
Oxidation of Methane to Methanol over Gold Decorated Zinc Oxide Via Photocatalysis.
Journal of
Materials Chemistry A
, 2020,
8
, 13277.
Page 14/22
[22] Xie, J., R. Jin, A. Li, Y. Bi, Q. Ruan, Y. Deng, Y. Zhang, S. Yao, G. Sankar, D. Ma, and J. Tang
.
Highly
Selective Oxidation of Methane to Methanol at Ambient Conditions by Titanium Dioxide-Supported Iron
Species.
Nature Catalysis
, 2018,
1
, 889.
[23] Feng, N., H. Lin, H. Song, L. Yang, D. Tang, F. Deng, and J. Ye
.
Ecient and Selective Photocatalytic
CH4 Conversion to CH3OH with O2 by Controlling Overoxidation on TiO2.
Nature Communications
, 2021,
12
, 10.1038/s41467-021-24912-0.
[24] Wei, S., X. Zhu, P. Zhang, Y. Fan, Z. Sun, X. Zhao, D. Han, and L. Niu
.
Aerobic Oxidation of Methane to
Formaldehyde Mediated by Crystal-O over Gold Modied Tungsten Trioxide Via Photocatalysis.
Applied
Catalysis B: Environmental
, 2021,
283
, 119661.
[25] Fan, Y., W. Zhou, X. Qiu, H. Li, Y. Jiang, Z. Sun, D. Han, L. Niu, and Z. Tang
.
Selective Photocatalytic
Oxidation of Methane by Quantum-Sized Bismuth Vanadate.
Nature Sustainability
, 2021,
4
, 509.
[26] Zhou, Y., L. Zhang, and W. Wang
.
Direct Functionalization of Methane into Ethanol over Copper
Modied Polymeric Carbon Nitride Via Photocatalysis.
Nature Communications
, 2019,
10
, 506.
[27] Yang, Z., Q. Zhang, L. Ren, X. Chen, D. Wang, L. Liu, and J. Ye
.
Ecient Photocatalytic Conversion of
CH4 into Ethanol with O2 over Nitrogen Vacancy-Rich Carbon Nitride at Room Temperature.
Chemical
Communications
, 2021,
57
, 871.
[28] Yu, X., V.L. Zholobenko, S. Moldovan, D. Hu, D. Wu, V.V. Ordomsky, and A.Y. Khodakov
.
Stoichiometric
Methane Conversion to Ethane Using Photochemical Looping at Ambient Temperature.
Nature Energy
,
2020,
5
, 511.
[29] Li, X., J. Xie, H. Rao, C. Wang, and J. Tang
.
Pt and Cuox Decorated TiO2 Photocatalyst for Oxidative
Coupling of Methane to C2 Hydrocarbons in a Flow Reactor.
Angewandte Chemie International Edition
,
2020,
59
, 19702.
[30] Ishimaru, M., F. Amano, C. Akamoto, and S. Yamazoe
.
Methane Coupling and Hydrogen Evolution
Induced by Palladium-Loaded Gallium Oxide Photocatalysts in the Presence of Water Vapor.
Journal of
Catalysis
, 2021,
397
, 192.
[31] Yu, L., Y. Shao, and D. Li
.
Direct Combination of Hydrogen Evolution from Water and Methane
Conversion in a Photocatalytic System over Pt/TiO2.
Applied Catalysis B: Environmental
, 2017,
204
, 216.
[32] Chen, Z., S. Wu, J. Ma, S. Mine, T. Toyao, M. Matsuoka, L. Wang, and J. Zhang
.
NonOxidative
Coupling of Methane: NType Doping of Niobium Single Atoms in TiO2-SiO2 Induces Electron
Localization.
Angewandte Chemie International Edition
, 2021,
60
, 11901.
[33] Li, N., R. Jiang, Y. Li, J. Zhou, Q. Ma, S. Shen, and M. Liu
.
Plasma-Assisted Photocatalysis of CH4 and
CO2 into Ethylene.
ACS Sustainable Chemistry & Engineering
, 2019,
7
, 11455.
Page 15/22
[34] Cho, Y., B. Park, D.K. Padhi, I.A.M. Ibrahim, S. Kim, K.H. Kim, K.-S. Lee, C.-L. Lee, J.W. Han, S.H. Oh, and
J.H. Park
.
Disordered-Layer-Mediated Reverse Metal–Oxide Interactions for Enhanced Photocatalytic
Water Splitting.
Nano Letters
, 2021, 10.1021/acs.nanolett.1c01368.
[35] Ma, J., X. Tan, Q. Zhang, Y. Wang, J. Zhang, and L. Wang
.
Exploring the Size Effect of Pt
Nanoparticles on the Photocatalytic Nonoxidative Coupling of Methane.
ACS Catalysis
, 2021,
11
, 3352.
[36] Song, H., X. Meng, S. Wang, W. Zhou, S. Song, T. Kako, and J. Ye
.
Selective Photo-Oxidation of
Methane to Methanol with Oxygen over Dual-Cocatalyst-Modied Titanium Dioxide.
ACS Catalysis
, 2020,
10
, 14318.
[37] Chen, F., T. Ma, T. Zhang, Y. Zhang, and H. Huang
.
Atomic-Level Charge Separation Strategies in
Semiconductor-Based Photocatalysts.
Advanced Materials
, 2021,
33
, 2005256.
[38] Ling, T., M. Jaroniec, and S.-Z. Qiao
.
Recent Progress in Engineering the Atomic and Electronic
Structure of Electrocatalysts Via Cation Exchange Reactions.
Advanced Materials
, 2020,
32
, 2001866.
[39] Lee, B.-H., S. Park, M. Kim, A.K. Sinha, S.C. Lee, E. Jung, W.J. Chang, K.-S. Lee, J.H. Kim, S.-P. Cho, H.
Kim, K.T. Nam, and T. Hyeon
.
Reversible and Cooperative Photoactivation of Single-Atom Cu/TiO2
Photocatalysts.
Nature Materials
, 2019,
18
, 620.
[40] Nosaka, Y. and A.Y. Nosaka
.
Generation and Detection of Reactive Oxygen Species in Photocatalysis.
Chemical Reviews
, 2017,
117
, 11302.
[41] Xiong, L. and J. Tang
.
Strategies and Challenges on Selectivity of Photocatalytic Oxidation of
Organic Substances.
Advanced Energy Materials
, 2021,
11
, 2003216.
[42] Lin, L., W. Zhou, R. Gao, S. Yao, X. Zhang, W. Xu, S. Zheng, Z. Jiang, Q. Yu, Y.-W. Li, C. Shi, X.-D. Wen,
and D. Ma
.
Low-Temperature Hydrogen Production from Water and Methanol Using Pt/α-MoC Catalysts.
Nature
, 2017,
544
, 80.
[43] Qiao, B., A. Wang, X. Yang, L.F. Allard, Z. Jiang, Y. Cui, J. Liu, J. Li, and T. Zhang
.
Single-Atom
Catalysis of CO Oxidation Using Pt1/FeOx.
Nature Chemistry
, 2011,
3
, 634.
[44] Wang, C., K. Wang, Y. Feng, C. Li, X. Zhou, L. Gan, Y. Feng, H. Zhou, B. Zhang, X. Qu, H. Li, J. Li, A. Li, Y.
Sun, S. Zhang, G. Yang, Y. Guo, S. Yang, T. Zhou, F. Dong, K. Zheng, L. Wang, J. Huang, Z. Zhang, and X.
Han
.
Co and Pt DualSingleAtoms with OxygenCoordinated Co-O-Pt Dimer Sites for Ultrahigh
Photocatalytic Hydrogen Evolution Eciency.
Advanced Materials
, 2021,
33
, 2003327.
[45] Guo, W., Z. Wang, X. Wang, and Y. Wu
.
General Design Concept for Single-Atom Catalysts toward
Heterogeneous Catalysis.
Advanced Materials
, 2021,
33
, e2004287.
[46] Zhang, W., Y. Chao, W. Zhang, J. Zhou, F. Lv, K. Wang, F. Lin, H. Luo, J. Li, M. Tong, E. Wang, and S.
Guo
.
Emerging DualAtomicSite Catalysts for Ecient Energy Catalysis.
Advanced Materials
, 2021,
33
,
Page 16/22
2102576.
[47] Li, Q., Y. Ouyang, H. Li, L. Wang, and J. Zeng
.
Photocatalytic Conversion of Methane: Recent
Advancements and Prospects.
Angewandte Chemie International Edition
, 2021,
10.1002/anie.202108069.
[48] Latimer, A.A., A. Kakekhani, A.R. Kulkarni, and J.K. Nørskov
.
Direct Methane to Methanol: The
Selectivity–Conversion Limit and Design Strategies.
ACS Catalysis
, 2018,
8
, 6894.
[49] Shanmuganathan, V., J.S. Kumar, R. Pachaiappan, and P. Thangadurai
.
Transition Metal Ion-Doped
In2O3 Nanocubes: Investigation of Their Photocatalytic Degradation Activity under Sunlight.
Nanoscale
Advances
, 2021,
3
, 471.
[50] Xu, C.Y., P.X. Zhang, and L. Yan
.
Blue Shift of Raman Peak from Coated TiO2 Nanoparticles.
Journal
of Raman Spectroscopy
, 2001,
32
, 862.
[51] Siedl, N., P. Gügel, and O. Diwald
.
First Combined Electron Paramagnetic Resonance and Ft-Ir
Spectroscopic Evidence for Reversible O2 Adsorption on In2O3–X Nanoparticles.
The Journal of Physical
Chemistry C
, 2013,
117
, 20722.
[52] Reyes-Gil, K.R., Y. Sun, E. Reyes-Garcı´, and D. Raftery
.
Characterization of Photoactive Centers in N-
Doped In2O3 Visible Photocatalysts for Water Oxidation.
The Journal of Physical Chemistry C
, 2009,
113
,
12558.
[53] García-Trenco, A., A. Regoutz, E.R. White, D.J. Payne, M.S.P. Shaffer, and C.K. Williams
.
Pdin
Intermetallic Nanoparticles for the Hydrogenation of CO2 to Methanol.
Applied Catalysis B:
Environmental
, 2018,
220
, 9.
[54] Wang, S., T. He, P. Chen, A. Du, K.K. Ostrikov, W. Huang, and L. Wang
.
In Situ Formation of Oxygen
Vacancies Achieving near-Complete Charge Separation in Planar BiVO4 Photoanodes.
Adv Mater
, 2020,
32
, e2001385.
[55] Zhu, W., M. Shen, G. Fan, A. Yang, J.R. Meyer, Y. Ou, B. Yin, J. Fortner, M. Foston, Z. Li, Z. Zou, and B.
Sadtler
.
Facet-Dependent Enhancement in the Activity of Bismuth Vanadate Microcrystals for the
Photocatalytic Conversion of Methane to Methanol.
ACS Applied Nano Materials
, 2018,
1
, 6683.
[56] Wang, L., M. Ghoussoub, H. Wang, Y. Shao, W. Sun, A.A. Tountas, T.E. Wood, H. Li, J.Y.Y. Loh, Y. Dong,
M. Xia, Y. Li, S. Wang, J. Jia, C. Qiu, C. Qian, N.P. Kherani, L. He, X. Zhang, and G.A. Ozin
.
Photocatalytic
Hydrogenation of Carbon Dioxide with High Selectivity to Methanol at Atmospheric Pressure.
Joule
, 2018,
2
, 1369.
[57] Chen, J., S. Stepanovic, A. Draksharapu, M. Gruden, and W.R. Browne
.
A Non-Heme Iron Photocatalyst
for Light-Driven Aerobic Oxidation of Methanol.
Angewandte Chemie International Edition
, 2018,
57
, 3207.
Page 17/22
Declarations
Acknowledgements
L. L., L. F., H. L. and J. X. are grateful for the China Postdoctoral Science Foundation (Grant No.
2019M663802) and the Shannxi Key Research Grant (China, 2020GY-244). Y. X. and J.T. are thankful for
nancial support from the UK EPSRC (EP/S018204/2), Leverhulme Trust (RPG-2017-122), Royal Society
Newton Advanced Fellowship grant (NAF\R1\191163 and NA170422) and Royal Society Leverhulme
Trust Senior Research Fellowship (SRF\R1\21000153).
Author contributions
J.T. conceived and supervised the entire project.L. L., L. F. and H. L. conducted the material synthesis,
characterizations and photocatalytic methane conversion tests.L. L. drafted the manuscript under the
guidance of J. T.. Y. X. helped to discuss the catalytic results and improve the manuscript. All authors
discussed and commented on the manuscript.
Competing intrests
The authors declare no competing interests.
Scheme
Scheme 1 is in the supplementary les section.
Figures
Page 18/22
Figure 1
Photocatalytic CH4 conversion performance under 420 nm irradiation. Investigations on (a) diverse noble
metal species, (b) K2PdCl4 dosage during synthesis, (c) molar ratio of CH4/O2, (d) H2O dosage, (e) total
pressure and (f) cycling tests over the best sample Pd-def-In2O3. Standard reaction conditions: 20 mg
photocatalyst, 50 mL distilled H2O, 19 bar CH4, 1 bar O2, 3 h. For reaction condition investigation, only
the specied parameter was changed.
Page 19/22
Figure 2
(a) Raman and (b) EPR spectra of In2O3, Pd-In2O3 and Pd-def-In2O3. (c, d) HRTEM and (e) HAADF and
EDS-mapping images of Pd-def-In2O3. Blue, red and green colors represent indium, oxygen and
palladium elements, respectively. (f) Aberration corrected HAADF-STEM image of Pd-def-In2O3, where Pd
single atoms with a weak intensity are indicated by yellow circles. (g) Line scan measured along the x-y
rectangle region marked in f.
Page 20/22
Figure 3
(a) UV-DRS spectra of In2O3, Pd-In2O3 and Pd-def-In2O3. (b) In-situ Pd3d XPS spectra and (c) in-situ EPR
spectra of Pd-def-In2O3 in dark and under light. (d) Steady-state PL spectra of In2O3, Pd-In2O3 and Pd-
def-In2O3.
Page 21/22
Figure 4
In-situ EPR spectra of (a) DMPO-OOH and (b) DMPO-OH for the monitor of reactive ·OOH and ·OH radicals
over different photocatalysts. (c) PL intensity of 7-HC versus time over different photocatalysts for the
quantication of ·OH. Isotopic labelled experiments (d) oxygen source and (e) carbon source for
methanol production in the presence of isotopic labelled H218O, 18O2 or 13CH4.
Page 22/22
Figure 5
DFT calculation of optimized geometries and adsorption energies of (a) H2O and (b) CH3OH on In2O3,
single atom Pd-In2O3 and Pd-def-In2O3.
Supplementary Files
This is a list of supplementary les associated with this preprint. Click to download.
Scheme1.png
supportinginformation.docx
Article
Direct conversion of methane (CH4) to C1-2 liquid oxygenates is a captivating approach to lock carbons in transportable value-added chemicals, while reducing global warming. Existing approaches utilizing the transformation of CH4 to liquid fuel via tandemized steam methane reforming and the Fischer-Tropsch synthesis are energy and capital intensive. Chemocatalytic partial oxidation of methane remains challenging due to the negligible electron affinity, poor C-H bond polarizability, and high activation energy barrier. Transition-metal and stoichiometric catalysts utilizing harsh oxidants and reaction conditions perform poorly with randomized product distribution. Paradoxically, the catalysts which are active enough to break C-H also promote overoxidation, resulting in CO2 generation and reduced carbon balance. Developing catalysts which can break C-H bonds of methane to selectively make useful chemicals at mild conditions is vital to commercialization. Single atom catalysts (SACs) with specifically coordinated metal centers on active support have displayed intrigued reactivity and selectivity for methane oxidation. SACs can significantly reduce the activation energy due to induced electrostatic polarization of the C-H bond to facilitate the accelerated reaction rate at the low reaction temperature. The distinct metal-support interaction can stabilize the intermediate and prevent the overoxidation of the reaction products. The present review accounts for recent progress in the field of SACs for the selective oxidation of CH4 to C1-2 oxygenates. The chemical nature of catalytic sites, effects of metal-support interaction, and stabilization of intermediate species on catalysts to minimize overoxidation are thoroughly discussed with a forward-looking perspective to improve the catalytic performance.
Article
Full-text available
The conversion of photocatalytic methane into methanol in high yield with selectivity remains a huge challenge due to unavoidable overoxidation. Here, the photocatalytic oxidation of CH 4 into CH 3 OH by O 2 is carried out on Ag-decorated facet-dominated TiO 2 . The {001}-dominated TiO 2 shows a durable CH 3 OH yield of 4.8 mmol g ⁻¹ h ⁻¹ and a selectivity of approximately 80%, which represent much higher values than those reported in recent studies and are better than those obtained for {101}-dominated TiO 2 . Operando Fourier transform infrared spectroscopy, electron spin resonance, and nuclear magnetic resonance techniques are used to comprehensively clarify the underlying mechanism. The straightforward generation of oxygen vacancies on {001} by photoinduced holes plays a key role in avoiding the formation of •CH 3 and •OH, which are the main factors leading to overoxidation and are generally formed on the {101} facet. The generation of oxygen vacancies on {001} results in distinct intermediates and reaction pathways (oxygen vacancy → Ti–O 2 • → Ti–OO–Ti and Ti–(OO) → Ti–O • pairs), thus achieving high selectivity and yield for CH 4 photooxidation into CH 3 OH.
Article
Full-text available
Abundant and affordable methane is not only a high‐quality fossil fuel, it is also a raw material for the synthesis of value‐added chemicals. Solar‐energy‐driven conversion of methane offers a promising approach to directly transform methane to valuable energy sources under mild conditions, but remains a great challenge at present. In this Review, recent advances in the photocatalytic conversion of methane are systematically summarized. Insights into the construction of effective semiconductor‐based photocatalysts from the perspective of light‐absorption units and active centers are highlighted and discussed in detail. The performance of various photocatalysts in the conversion of methane is presented, with the photooxidation classified according to the oxidant systems. Lastly, challenges and future perspectives in the photocatalytic oxidation of methane are described.
Article
Full-text available
Atomically dispersed metal catalysts with well‐defined structures have been the research hotspot in heterogeneous catalysis because of their high atomic utilization efficiency, outstanding activity, and selectivity. Dual‐atomic‐site catalysts (DASCs), as an extension of single‐atom catalysts (SACs), have recently drawn surging attention. The DASCs possess higher metal loading, more sophisticated and flexible active sites, offering more chance for achieving better catalytic performance, compared with SACs. In this review, recent advances on how to design new DASCs for enhancing energy catalysis will be highlighted. It will start with the classification of marriage of two kinds of single‐atom active sites, homonuclear DASCs and heteronuclear DASCs according to the configuration of active sites. Then, the state‐of‐the‐art characterization techniques for DASCs will be discussed. Different synthetic methods and catalytic applications of the DASCs in various reactions, including oxygen reduction reaction, carbon dioxide reduction reaction, carbon monoxide oxidation reaction, and others will be followed. Finally, the major challenges and perspectives of DASCs will be provided. As an extension of single‐atom catalysts, dual‐atomic‐site catalysts (DASCs) with high metal loading, sophisticated and flexible active sites have aroused great interest. This review summarizes the recent development of DASCs in characterization, synthetic methods, and catalytic applications.
Article
Full-text available
As a new and popular material, single‐atom catalysts (SACs) exhibit excellent activity, selectivity, and stability for numerous important reactions, and show great potential in heterogeneous catalysis due to their high atom utilization efficiency and the controllable characteristics of the active sites. The composition and coordination would determine the geometric and electronic structures of SACs, and thus greatly influence the catalytic performance. Based on atom economy, rational design and controllable synthesis of SACs have become central tasks in the fields of low‐cost and green catalysis. Herein, an introduction to the recent progress in the precise synthesis of SACs including the regulation of the coordination structure and the choice of different systems is presented. Thereafter, the potentials of SACs in different applications are comprehensively summarized and discussed. Furthermore, a detailed discussion of the recent developments regarding the large‐scale preparation of SACs is provided, including the major issues and prospects for industrialization. Finally, the main challenges and opportunities of rapid large‐scale industrialization of SACs are briefly discussed. The development and exploration of highly efficient single‐atom catalysts (SACs) is of enormous significance for achieving industrialization. The advanced progress of precise synthesis and practical application of SACs is summarized. Thereafter, the main challenges and future opportunities of large‐scale synthesis of SACs are discussed.
Article
Full-text available
The direct oxidation of methane to more desirable, one-carbon oxygenated molecules such as methanol and formaldehyde offers a pathway towards a more sustainable chemical industry as the current commercial reforming process involving two steps features a high carbon footprint and energy consumption. Here, we report the selective photocatalytic oxidation of methane at room temperature using quantum-sized bismuth vanadate nanoparticles as the catalyst and oxygen as a mild oxidant. The reaction offers a high selectivity, of 96.6% for methanol or 86.7% for formaldehyde, under optimum wavelength and intensity of light, reaction time and amount of water solvent. Comprehensive characterizations disclose a multistep reaction mechanism in which the activation of methane by the hydroxyl radical determines the reaction rate. This work broadens the avenue towards the selective conversion of the greenhouse gas methane into desirable chemical products in a sustainable way.
Article
Full-text available
Photocatalytic ammonia synthesis is exciting but quite challenging with a very moderate yield at present. One of the greatest challenges is to develop highly active centers in a photocatalyst for N2 reduction under ambient conditions. Herein, porous carbon‐doped anatase TiOx (C‐TiOx) nanosheets with high‐concentration active sites of Ti3+ are presented, which are produced by layered Ti3SiC2 through a reproducible bottom‐up approach. It is shown that the high‐concentration Ti3+ sites are the major species for the significant increase in N2 photoreduction activity by the C‐TiOx. Such bottom‐up substitutional doping of C into TiO2 is responsible for both visible absorption and generation of Ti3+ concentration. Together with the porous nanosheets morphology and the loading of a Ru/RuO2 nanosized cocatalyst for enhanced charge separation and transfer, the optimal C‐TiOx with a Ti3+/Ti4+ ratio of 72.1% shows a high NH3 production rate of 109.3 µmol g−1 h−1 under visible‐light irradiation and a remarkable apparent quantum efficiency of 1.1% at 400 nm, which is the highest compared to all TiO2‐based photocatalysts at present. A porous carbon‐doped anatase TiOx (C‐TiOx) nanosheet with controllable Ti3+ concentration is prepared by a reproducible bottom‐up strategy. The optimal C‐TiOx with a high‐concentration Ti3+ exhibits a NH3 production rate of 109.3 µmol g–1 h–1 under visible‐light irradiation and a remarkable apparent quantum efficiency of 1.1% at 400 nm, the highest compared with all TiO2‐based photocatalysts.
Article
Non-oxidative coupling of methane (NOCM, 2CH4 → C2H6 + H2) is a reaction that can directly produce ethane and hydrogen at the same time, and gallium oxide (Ga2O3) powder has been reported as an effective photocatalyst for NOCM at room temperature. In this study, we investigated the reaction conditions for Pd-loaded Ga2O3 photocatalysts to improve the production rate of C2H6 and H2. We found that the 0.1 wt% Pd/Ga2O3 exhibited high selectivity of C2H6 (75.8%, carbon-based) under the conditions of steam reforming of methane. Photocatalytic NOCM seems to proceed in the presence of small amount of water. An increase in water vapor pressure (PH2O) was essential for the steady production of C2H6 and H2. The C2H6 production rate was 0.79 μmol min⁻¹ for 50 mg of Pd/Ga2O3 powder at PH2O = 3.6 kPa. The apparent quantum efficiency (AQE) for C2H6 production was 5.1%, which is much higher than that of conventional photocatalytic NOCM in the absence of water vapor. The importance of water adsorbates on the photocatalyst surface was suggested by water vapor adsorption isotherm and Fourier transform infrared (FT-IR) spectroscopy. It is revealed that multilayered water molecules adsorbed on the photocatalyst surface play a role as a reaction field that promotes the dehydrogenative coupling of CH4.
Article
The high inertness of the C-H bond makes the photocatalytic methane conversion a significant challenge. The platinum nanoparticle is a promising cocatalyst for CH4 activation, while the study of its structure characteristics and functionality remains in its infancy. Herein, the size effect of Pt on the photocatalytic nonoxidative methane conversion efficiency was systematically investigated over x-Pt/Ga2O3 with the particle size (x) ranging from 1.5 to 2.7 nm, where a volcano-shaped relation was observed. The smaller size is beneficial to the formation of Ptδ+ species, which is mainly distributed on the terrace sites according to the DFT calculation. The corner Pt atom is the geometric active site for the CH4 polarization, and the terrace Ptδ+ helps promote C-H activation since the activity is decreased on reduced x-Pt/Ga2O3 with a lower Ptδ+ content. Meanwhile, Ptδ+ species favors the oxidation of adsorbed -CH3 group to ·CH3. The volcano-shaped size effect on the NOCM activity was finally rationalized by the balance between C-H activation and C2H6 desorption from the corner sites on different sized Pt.