Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Volume 15

Number 2

Energy &
February 2022
Pages 401–868

Environmental
Science
rsc.li/ees

ISSN 1754-5706

PAPER
Hyungjun Kim, Su-Il In, Taeghwan Hyeon et al.
Electronic interaction between transition metal single-atoms
and anatase TiO2 boosts CO2 photoreduction with H 2 O
Energy &
Environmental
Science
View Article Online
PAPER
Published on 20 September 2021. Downloaded by CHONNAM NATIONAL UNIVERSITY on 2/10/2023 5:32:35 AM.

View Journal | View Issue

Electronic interaction between transition metal


single-atoms and anatase TiO2 boosts CO2
Cite this: Energy Environ. Sci.,
2022, 15, 601 photoreduction with H2O†
Byoung-Hoon Lee, ‡ab Eunhee Gong, ‡c Minho Kim, ‡de Sunghak Park, f

Hye Rim Kim,c Junho Lee,c Euiyeon Jung, ab Chan Woo Lee,ab Jinsol Bok,ab
Yoon Jung,ab Young Seong Kim,a Kug-Seung Lee, g Sung-Pyo Cho,h
Jin-Woo Jung,i Chang-Hee Cho, i Sébastien Lebègue, d Ki Tae Nam, f
Hyungjun Kim,*j Su-Il In *c and Taeghwan Hyeon *ab

Single-atom catalysts are playing a pivotal-role in understanding atomic-level photocatalytic processes.


However, single-atoms are typically non-uniformly distributed on photocatalyst surfaces, hindering the
systematic investigation of structure–property correlation at atomic precision. Herein, by combining
material design, spectroscopic analyses, and theoretical studies, we investigate the atomic-level CO2
photoreduction process on TiO2 photocatalysts with uniformly stabilized transition metal single-atoms.
First, the electronic interaction between single Cu atoms and the surrounding TiO2 affects the
reducibility of the TiO2 surface, leading to spontaneous O vacancy formation near Cu atoms. The
Received 24th May 2021, coexistence of Cu atoms and O vacancies cooperatively stabilizes CO2 intermediates on the TiO2
Accepted 20th September 2021 surface. Second, our approach allows us to control the spatial distribution of uniform single Cu atoms
DOI: 10.1039/d1ee01574e on TiO2, and demonstrate that neighboring Cu atoms simultaneously engage in the interaction with CO2
intermediates by controlling the charge localization. Optimized Cu1/TiO2 photocatalysts exhibit 66-fold
rsc.li/ees enhancement in CO2 photoreduction performance compared to the pristine TiO2.

Broader context
Heterogeneous photocatalysis represents an attractive platform for the storage of renewable energy and production of solar-chemicals. TiO2 is the most
important photocatalyst that occupies 490% of the entire photocatalyst industry due to its environmental benignity, high photocatalytic activity and cost-
effectiveness. Although various important advances have been effective in tailoring macroscopic photocatalytic properties of TiO2, atomic-level strategies to
enhance the photocatalytic performance of TiO2 have been scarce because the position and valence of cocatalysts are difficult to control on an atomic scale.
Reversible electron transfer from metal cofactors to surrounding redox-mediating ligands controls catalytic processes in homogeneous catalysts and
metalloenzymes. Similarly, redox-active reducible oxides (e.g. TiO2 and CeO2) stabilize a single atom in an extended system, and function as an electron
reservoir, which can influence the degree of charge transfer at the interface. To this end, we investigate how electronic interaction of a single atom and
surrounding TiO2 regulates photocatalytic CO2 reduction. In the current work, we demonstrate that single Cu atoms, which are site-specifically stabilized in the
Ti vacancies of TiO2, interact with surrounding TiO2 and control the overall electronic properties (e.g. reducibility, defect formation) of TiO2. We also found that
the electronic interaction is significantly dependent upon the presence of nearby single atoms.

a
Center for Nanoparticle Research, Institute for Basic Science (IBS), Seoul 08826, Republic of Korea
b
School of Chemical and Biological Engineering, and Institute of Chemical Processes, Seoul National University, Seoul 08826, Republic of Korea. E-mail: thyeon@snu.ac.kr
c
Department of Energy Science & Engineering, Daegu Gyeongbuk Institute of Science and Technology (DGIST), Daegu, 42988, Republic of Korea. E-mail: insuil@dgist.ac.kr
d
Université de Lorraine and CNRS, LPCT, UMR 7019, Vandœuvre-lès-Nancy 54506, France
e
Department of Applied Chemistry, Kyung Hee University, Yongin, Gyeonggi 17104, Republic of Korea
f
Department of Materials Science and Engineering, Seoul National University, Seoul 08826, Republic of Korea
g
Pohang Accelerator Laboratory (PAL), Pohang University of Science and Technology (POSTECH), Pohang 37673, Republic of Korea
h
National Center for Inter-University Research Facilities, Seoul National University, Seoul 08826, Republic of Korea
i
Department of Emerging Materials Science, DGIST, Daegu 42988, South Korea
j
Department of Chemistry, Korea Advanced Institute of Science and Technology (KAIST), Daejeon 34141, Republic of Korea. E-mail: linus16@kaist.ac.kr
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1ee01574e
‡ These authors contributed equally to this work.

This journal is © The Royal Society of Chemistry 2022 Energy Environ. Sci., 2022, 15, 601–609 | 601
View Article Online

Paper Energy & Environmental Science

Introduction Herein, combining material design, photocatalytic experi-


ments and theoretical analyses, we investigate how electronic
Among numerous photocatalytic materials,1–5 TiO2 is the most interaction between redox-active TiO2 and single Cu atoms
intensively studied photocatalyst due to its environmental affects atomic-level photocatalytic CO2 reduction. The site-
benignity and cost-effectiveness along with intrinsically high specific stabilization of a metal atom in a Ti vacancy eliminates
Published on 20 September 2021. Downloaded by CHONNAM NATIONAL UNIVERSITY on 2/10/2023 5:32:35 AM.

photocatalytic activity.6–8 Various strategies including facet the possible coexistence of other atomic configuration or sub-
engineering,9 vacancy formation,10 structural disorder,11,12 ele- nanometric clusters, enabling investigation of the photocataly-
mental doping,13,14 band engineering15 and cocatalyst tic CO2 reduction on a TiO2 surface with atomic precision. We
design16–19 have been applied to improve the photocatalytic first identify the electronic properties (e.g. reducibility, defect
properties of TiO2 including light absorption, charge separa- formation) of TiO2 through interacting with stabilized single
tion and surface catalytic property. However, atomic-level stra- atoms, which arise from electronic metal–support interaction.
tegies to enhance the photocatalytic performance of TiO2 have Interestingly, spontaneous O vacancy formation is favoured
been scarce and deposition of a noble metal cocatalyst (Pt, Pd, near Cu atoms in Ti vacancies, and the coexistence of Cu atoms
Rh and Ir) is generally required to improve the intrinsically low and O vacancies synergistically boosts photocatalytic CO2
surface catalytic activity of TiO2, which has limited broader reduction by eliminating the CO2 activation barrier, enabling
photocatalytic applications of TiO2.1–3,16–18 66-fold enhancement in CO2 reduction activity compared to
Earth-abundant first-row transition metal atoms (Fe, Co, Ni, that of pristine TiO2. Our synthetic approach further enables
Mn and Cu), which are highly effective as active catalytic the control of the spatial distribution of uniform single
species for the conversion of small molecules such as H2O, Cu atoms on TiO2 and demonstrates that the presence of
O2, CO2 and N2, have attracted significant research interest in multiple adjacent single Cu atoms leads to different structural
designing high-performance single-atom catalysts (SACs) com- relaxation processes through cooperative molecular CO2 inter-
prising earth-abundant elements.20–33 Likewise, a number of action by controlling the charge localization.
atomically dispersed metals on a photocatalyst surface have
emerged as effective platforms for various photochemical reac-
tions. These include H2 evolution,24–30 CH4 conversion,19 N2 Results and discussion
reduction,31 and CO2 reduction,16,34–37 demonstrating the ver-
satility of a single atom as an effective cocatalyst system. Electronic interaction between a single atom and support
However, the single-atom photocatalyst is still in its infancy The redox-mediating property of a support can directly affect
and the determining factor that influences the photocatalytic the electronic interactions at a metal–support interface. To
performance is not well understood. Considering the impor- investigate, we first considered the influence of electronic
tance of TiO2 in photocatalytic applications, it is significant to interaction on both a single Cu atom and the surrounding
establish the structure–activity correlation that regulates the TiO2. The d states of an active metal atom in ML6 coordination
photocatalytic performance of single-atom cocatalysts on TiO2. (that is, an MO6 octahedron in TiO2) experience octahedral
Very recently, it was identified that the catalytic performance splitting to eg and t2g states, which individually interact with
of SACs can be controlled by tuning the local atomic symmetrically matched p states of O valence states (ps and pp),
configuration.21,25,38–44 Especially, significant differences have leaving the rest of the p states nonbonded (pNB) (Fig. 1a).
been observed in the catalytic performance of single atom/ Accordingly, charge transfer at the metal/support interface
reducible oxide systems (e.g. TiO2 and CeO2), depending on the determines the relative alignments of metal/support electronic
local environment of single atoms.25,41–44 Redox-active reduci- states and thus can affect the electronic properties of the
ble oxides stabilize a single atom in an extended system and support material as well as the reactivity of the active metal
function as an electron reservoir, which can influence the center. Interestingly, for a single Cu atom on anatase TiO2
degree of charge transfer at the interface.45,46 Considering that (Fig. 1b), the coordination environment of a single metal atom
the charge transfer between a metal cofactor and redox- influences the valence states and lability of a metal and support
mediating ligands regulates the entire catalytic process in material. A Cu atom in a Ti vacancy induces an oxidizing
biological enzymes, the surrounding support in a heteroge- (O-rich) condition that partially oxidizes an oxide anion (O2 )
neous system is equally important to that of supported metal to stabilize the metal–support interface. As a result, the mid-
atoms. Only recently, dynamic processes such as structural gap O states of Cu1/TiO2 are exposed above the Fermi level and
distortion and electronic structure relaxation on a single atom increase surface reducibility. This contrasts to pristine TiO2
catalyst have been correlated with the coordinating support, where the O states are below the Fermi level in the absence of
which significantly affected entire catalytic characteristics mid-gap states. Due to the upshift in the O state, TiO2 becomes
including molecular interaction and charge transfer vulnerable to hydrogenation and consequently, favours O
property.47,48 Therefore, understanding the influence of elec- vacancy formation near the Cu atom in ambient conditions.
tronic interaction between an atomic catalyst and surrounding This is supported by the thermodynamic preference to
support toward catalytic reactivity is very important to accu- O vacancy formation as shown in I - IV of Fig. 1c and d,
rately establish structure–activity correlation in a single atom where Cu1/TiO2 results in 0.75 eV downhill in contrast with
catalyst system. pristine TiO2 of +4.41 eV uphill. Critically, our theoretical result

602 | Energy Environ. Sci., 2022, 15, 601–609 This journal is © The Royal Society of Chemistry 2022
View Article Online

Energy & Environmental Science Paper


Published on 20 September 2021. Downloaded by CHONNAM NATIONAL UNIVERSITY on 2/10/2023 5:32:35 AM.

Fig. 1 Enhanced surface reducibility of a single-atom catalyst by electronic metal–support interfacial hybridization. (a) Schematic alignments of the
d orbital splitting of a metal atom in a coordinate complex (green on the left), the energy levels of the support material (middle) and their metal–support
hybrid states based on ligand field theory (black in the middle). The resulting states of a single-atom catalyst are shown on the right. (b) Structure and
schematic illustration of the density of states of Cu1/TiO2 (left) and pristine TiO2. For Cu1/TiO2, the spin density of the Cu(dz2)–O state near the Fermi level
is shown in yellow. (c) DFT energetics of oxygen vacancy formation (Cu1/TiO2 in red and pristine TiO2 in black), where the chemical potential of oxygen
was considered as m(O) = 1/2 E(O2) and the corresponding structural change of Cu–O4 (in green) in Cu1/TiO2 where Cu is in blue, O is in red, and
Ti is in cyan.

indicates that the intimate electronic interaction between a state) (Fig. S1, ESI†). Based on DFT energetics, all the elemen-
metal atom catalyst and the surrounding support material tary steps of CO2 reduction in Cu1/TiO2 x become energetically
cooperatively changes their electronic properties (e.g., the elec- favourable (Fig. 2b). CO2 can be first chemically adsorbed
tronic states of a single atom and the reducibility and defect favourably at the O vacancy site near the Cu atom (1 - 2 in
formation of the support material). Fig. 2a, 0.30 eV) where the square planar Cu–O4 (1) is
distorted into seesaw Cu–O4 (2) to form a Cu–C bond. The
CO2 photoreduction mechanism on Cu1/TiO2 with an O following proton–electron transfer reduces terminal O into
vacancy water and forms a favourable metal–CO bond (4 in Fig. 2a) that
Considering that the presence of an O vacancy on a photo- efficiently removes the potential barrier for the deoxygenation
catalyst surface has been shown to influence photocatalytic of CO2 to CO (3 - 4 in Fig. 2a, 0.59 eV, red in Fig. 2b), which
performance substantially,10,45,49 the identification of sponta- is the energetically most unfavourable step in perfect TiO2
neous O vacancy formation near the Cu atom is significant. We (+1.35 eV, blue in Fig. S2, ESI†) and one of the unfavourable
can understand the photocatalytic CO2 reduction mechanism steps in defective TiO2 x (+0.37 eV, black in Fig. 2b). The
on Cu1/TiO2 with an O vacancy (denoted as Cu1/TiO2 x) based following reduction step selectively hydrogenates the carbon
on the fast-hydrogenation pathway (Fig. 2).10 A photoexcited (4 - 5 in Fig. 2a, 0.02 eV) compared with either copper
state of Cu1/TiO2 x can be described with a proton that (+0.72 eV) or terminal oxygen (+2.02 eV) (Fig. S3, ESI†), since it
balances the surface charge with a photoexcited electron loca- can simultaneously form favourable C–H and Ti–O bonds,
lized to the dx2 y2 state of the Cu atom resulting in a Cu+ state leading to a successive reduction of carbon to methane
(Bader spin population, ms = 0.00 for Cu+ in the photoexcited (5 - 8 in Fig. 2a). The oxygen vacancy can be readily replenished
Cu1/TiO2 x state vs. 0.72 for Cu2+ in the resting Cu1/TiO2 x by favourable vacancy formation (9 - 1 in Fig. 2a, 1.56 eV) as

This journal is © The Royal Society of Chemistry 2022 Energy Environ. Sci., 2022, 15, 601–609 | 603
View Article Online

Paper Energy & Environmental Science


Published on 20 September 2021. Downloaded by CHONNAM NATIONAL UNIVERSITY on 2/10/2023 5:32:35 AM.

Fig. 2 Photocatalytic CO2 reduction at Cu1/TiO2 with an oxygen vacancy. (a) CO2 reduction mechanism at Cu1/TiO2 x and (b) DFT energetics of CO2
reduction at Cu1/TiO2 x (red) and TiO2 x (black). During the reduction process, the photoexcited electron is assumed to be transferred from the
conduction band minimum of TiO2 (–0.25 VRHE).

shown in Fig. 1c, where the defective TiO2 x suffers from a spontaneous O vacancy formation adjacent to Cu atoms. There-
thermodynamic barrier (+0.14 eV). Considering that the regenera- fore, Cu1/TiO2 would possess both Ti3+ (near Cu atoms) and
tion of the oxygen vacancy is the bottleneck for the CO2 reduction Ti4+ (far from Cu atoms) on the surface. To investigate the
reaction at TiO2,10 the promoted O vacancy formation near the electronic structure change of TiO2 upon stabilizing Cu atoms,
single Cu atom effectively enhances catalytic efficiency. electron energy-loss spectroscopy (EELS) was performed (Fig. 3e
and f). The L2,3 edge of 3d transition metals provides informa-
Material synthesis and characterization tion of the coordination structure, symmetry and oxidation
Various site-uniform single metal atom TiO2 photocatalysts state of central atoms. The Ti L2 and L3 edges are further split
have been synthesized through template confined thermody- into two characteristic peaks (eg and t2g) that develop through
namic redistribution of single atoms (Fig. S4, ESI†).25 The site- octahedral crystal field splitting. Therefore, crystalline anatase
uniform single atoms in the Ti vacancies of TiO2 eliminate the TiO2 possesses four characteristic peaks, as shown in Fig. 3f.
coexistence of multiple atomic sites and sub-nanometric clus- When Ti4+ is reduced to Ti3+, the octahedral crystal field
ters, which enables us to correlate the photocatalytic CO2 splitting is disturbed and eg and t2g are not well distinguished,
reduction performance directly to the single atoms in the exhibiting two peaks at slightly lower energy loss. Interestingly,
Ti vacancies. Structural characterization was performed on the EELS spectra of our Cu1/TiO2 (0.21) possess features of both
Cu1/TiO2 samples (Fig. 3 and Fig. S5–S14, ESI†). Inductively four peaks (site 2) and two peaks (site 1) when taken upon
coupled plasma-atomic emission spectroscopy (ICP-AES) was different sites, indicating that both Ti3+ and Ti4+ coexist on the
performed to determine the loading amount of the Cu atoms surface. Note that the peak occurs at slightly lower energy loss
(Table S1, ESI†). Hereafter, we denote Cu1/TiO2 (x), where x is even in the case of the Ti4+ site due to the presence of abundant
the Cu atom concentration in at%. The scanning transmission Ti3+ nearby. The electronic structure of TiO2 upon stabilizing
electron microscopy (STEM) and the transmission electron Cu atoms was further studied by X-ray photoelectron spectro-
microscopy (TEM) images reveal the absence of observable Cu scopy (XPS). The Ti 2p peak of pristine TiO2 is located at
nanoparticles in the prepared Cu1/TiO2 (Fig. 3a, Fig. S5 and S9, 458.4 eV, which is a typical Ti4+ peak of anatase TiO2.51
ESI†). Energy-dispersive X-ray spectroscopy (EDS) analysis with Interestingly, the Ti 2p peak of Cu1/TiO2 shows a downshift
STEM shows that Cu atoms are uniformly dispersed on the TiO2 along with increasing Cu concentration (0.2 eV at 1.42 at%),
surface even at a high Cu atom concentration (0.70 at%) consistent with the EELS spectra, indicating that TiO2 is
(Fig. 3b). The results are consistent with the X-ray diffraction partially reduced upon stabilizing single Cu atoms (Fig. S8,
(XRD) patterns, which show no additional peaks other than ESI†). Then, the atomic structure of Cu1/TiO2 with different Cu
anatase TiO2 in all the Cu1/TiO2 with various Cu at% (Fig. S7, concentrations was analysed using X-ray absorption spectro-
ESI†). From our theoretical results, stabilizing Cu atoms causes scopy (XAS) (Fig. 3g and h and Fig. S11–S14, ESI†). From the

604 | Energy Environ. Sci., 2022, 15, 601–609 This journal is © The Royal Society of Chemistry 2022
View Article Online

Energy & Environmental Science Paper


Published on 20 September 2021. Downloaded by CHONNAM NATIONAL UNIVERSITY on 2/10/2023 5:32:35 AM.

Fig. 3 Material characterization of Cu1/TiO2 samples. (a) Low-resolution STEM and (b) STEM-EDS elemental mapping image of Cu1/TiO2. High-
resolution STEM Z-contrast image of (c) Cu1/TiO2 (0.70) and (d) Cu1/TiO2 (0.21). (e) STEM image and (f) corresponding EELS spectra of Cu1/TiO2 (0.21) at
different sites. Cu K edge (g) XANES, and (h) FT-EXAFS spectra of Cu1/TiO2 samples.

Cu K edge X-ray absorption near-edge structure (XANES) directly imaged the local structure of Cu atoms on TiO2 using Cs-
spectroscopy, we found that the oxidation state of the Cu atoms corrected STEM analysis (Fig. 3c and d and Fig. S9 and S10, ESI†).
is the Cu2+ state in all the Cu1/TiO2 samples exhibiting the From the low-resolution STEM and TEM images, agglomeration of
absorption edge at 8996.3 eV and peak at 8997.3 eV, regardless Cu was not observed, which is consistent with the XRD and XAS
of the Cu loading amount (Fig. 3g). The absence of the peaks at results. Because the atomic mass of Cu is greater than that of Ti, we
8981.4 eV (Cu1+) and 8980.8 eV (Cu0) indicates that oxidation could observe that bright dots (indicating Cu atoms) are uniformly
states other than Cu2+ are absent. XANES spectra of CuO, Cu2O, distributed on the TiO2 surface. The bright dots are exclusively
and Cu foil were measured as the references, confirming the located on the Ti position, suggesting that the thermodynamic
Cu2+ oxidation state in all the Cu1/TiO2 samples. The Cu K edge redistribution process leads Cu atoms to be stabilized in the
extended X-ray absorption-fine-structure (EXAFS) analysis Ti vacancies. EELS experiment was performed to verify that the
reveals that the absence of the Cu-based nanoparticle for- bright dots are single Cu atoms (Fig. S10, ESI†). In single Pt atom
mation is consistent with the XRD, EDS, and TEM results, catalysts on metal oxide supports that were synthesized via a typical
and the identical local coordination structure of Cu atoms in wet-impregnation method, inevitably sub-nanometer-sized Pt oxide
all the Cu1/TiO2 samples. The two characteristic peaks of Cu–O clusters are generated above 0.04–0.08 Pt at% (or 0.1–0.2 wt%),
and Cu–O–Ti are observed in all the samples, indicating that complicating the precise interpretation of catalyst properties at high
the Cu atoms are present in the Ti vacancies of anatase TiO2. Pt concentrations.43 In contrast, our thermodynamic redistribution
Profile fitting on the EXAFS spectra was performed, confirming of single atoms in a nano-confined space at a high temperature
that the Cu atoms are stabilized in the Ti vacancies. The (1223 K) ensured stabilization of single atoms exclusively in the
coordination number (CN) of the Cu–O bonds in all Cu1/TiO2 Ti vacancies even at a high concentration. Interestingly, upon
samples was similar (4.6–4.8), showing that negligible differ- imaging Cu1/TiO2 with high Cu concentration (0.70 at%), we found
ences exist in the local coordination structure of Cu atoms at that the spatial distance between multiple Cu atoms becomes very
different concentrations (Fig. 3h, Fig. S14 and Table S2, ESI†). close (Fig. 3c). From these material characterization results, we
Note that the CN of Cu–O bonding is smaller than that of Ti–O could find two distinctive characteristics regarding Cu1/TiO2. Firstly,
bonding in typical anatase TiO2, indicating that Cu atoms are Cu atom stabilization in the Ti vacancy influences the reducibility of
undercoordinated. Considering the downshift of Ti 2p spectra the TiO2 surface. Secondly, upon increasing the Cu content, the
in the XPS analysis of Cu1/TiO2, undercoordinated Cu atoms spatial distance between adjacent Cu atoms becomes very close in
indicate that partial reduction of the TiO2 surface is associated that multiple Cu atoms can simultaneously interact with one CO2
with favorable O vacancy formation near Cu atoms, which is molecule.
consistent with our theoretical prediction.
Considering the negligible differences observed in the STEM, Photocatalytic CO2 reduction experiment
XRD, XANES and EXAFS analyses, we could exclude the influence of The photocatalytic CO2 reduction experiments were performed
the difference in the local coordination structure of single Cu atoms in a gas–solid reaction setup (1000 ppm CO2 feed, He balance)
upon increasing Cu concentration. To gain further insights, we under simulated solar irradiation without any sacrificial

This journal is © The Royal Society of Chemistry 2022 Energy Environ. Sci., 2022, 15, 601–609 | 605
View Article Online

Paper Energy & Environmental Science

reagent or photosensitizer (Fig. S15, ESI†). All the prepared Isolated Cu atoms in an MO6 octahedron in TiO2 possess
single-atom TiO2 photocatalysts incorporated with Cu, Co, Ni, d orbitals inside the bandgap of TiO2 and photogenerated
and Rh atoms (metal loading of 0.21 at%) exhibited enhanced electrons transfer from the conduction band of TiO2 to the
photocatalytic CO2 reduction compared to pristine TiO2 d orbitals of the isolated copper atoms.25 The TRPL analysis
(Fig. 4a). Whereas the pristine TiO2 exhibited a CH4 production confirms that isolated Cu atoms facilitate charge transfer,
Published on 20 September 2021. Downloaded by CHONNAM NATIONAL UNIVERSITY on 2/10/2023 5:32:35 AM.

rate of 21.6 ppm g 1 h 1, the CH4 production rates of various promoting the CO2 reduction reaction.
single-atom incorporated TiO2 were at least doubled, exhibiting Previously, copper nanoparticle-based cocatalysts (e.g. CuO,
44.6 (Rh1/TiO2), 97.3 (Ni1/TiO2), 123.5 (Co1/TiO2), and 1416.9 Cu2O, Cu(OH)2 and Cu) on TiO2 were shown to be effective for
(Cu1/TiO2) ppm g 1 h 1, indicating that the incorporation of photocatalytic CO2 reduction with H2O.17,56,57 It is noteworthy
single atoms in TiO2 can improve the surface catalytic activity that the CH4 production rate of Cu1/TiO2 is also 30 times higher
of TiO2 to favour photocatalytic CO2 reduction. Specifically, the than the reported copper oxide nanoparticle-cocatalyst system
CH4 production rate of Cu1/TiO2 was significantly increased, on TiO2, demonstrating that the atomically dispersed
with a 66-fold enhancement, even with the low loading amount Cu exhibits a superior photocatalytic performance compared
of Cu atom (0.21 at% or 0.17 wt%). Time-resolved photolumi- to copper oxide nanoparticles.56 We modified our synthetic
nescence (TRPL) spectroscopy was performed to investigate the methods and synthesized a Cu nanoparticle (CuNP)/TiO2 photo-
influence of Cu atoms on the charge carrier dynamics (Fig. S16 catalyst with various concentrations and compared the
and Table S3, ESI†). TRPL decay was fitted using a bi- photocatalytic activity with Cu1/TiO2. All the CuNP/TiO2 photo-
exponential fitting model (eqn (1)) where t1 and t2 are the catalysts with different Cu concentrations exhibited a much
decay times, and A1 and A2 are the corresponding magnitudes. lower CO2 reduction performance compared to optimized
Cu1/TiO2 (0.21) (Fig. S17, ESI†). Specifically, the best
I(t) = A1exp( t/t1) + A2exp( t/t2) (1)
CuNP/TiO2 exhibited 240.0 ppm g 1 h 1 for CO2 to CH4 conver-
sion while Cu1/TiO2 (0.21) exhibited 1416.9 ppm g 1 h 1. The
The signal decays when the material undergoes electron– photocatalytic performance remained stable upon 5 cycles
hole recombination (longer lifetime, t1) and charge transfer/ without noticeable performance decrease (Fig. 4b). We per-
trapping (shorter lifetime, t2).52,53 Cu1/TiO2 (0.21) has a smaller formed various material characterizations including EELS,
average lifetime (0.477 ns) than pure TiO2 (0.711 ns), suggesting STEM-EDS and Cs-STEM analysis on Cu1/TiO2 (0.21) to inves-
that an electron is transferred from TiO2 to Cu (Fig. S16, ESI†). tigate the recyclability and material stability of Cu1/TiO2 after

Fig. 4 Comparison of the photocatalytic CO2 reduction activity. (a) Photocatalytic CO2 reduction performance of various M1/TiO2 (M = Cu, Co, Ni, Rh,
and pristine TiO2). (b) Cyclic stability of CO2 reduction on Cu1/TiO2. (c) Cu concentration-dependent CO2 reduction performance on Cu1/TiO2. The DFT
optimized structure of (d) Cu1/TiO2 x and (e) Cu2/TiO2 x after photoexcitation and CO2 adsorption where the binding pocket at the single Cu atom is
highlighted by an orange dotted circle. The charge density difference between CO2 adsorption [Dr = r(CO2–Cun/TiO2 x) r(Cun/TiO2 x) r(CO2)] is
shown in yellow and cyan for charge accumulation and depletion, respectively, with the isosurface level being 5  10 3 e Å 3.

606 | Energy Environ. Sci., 2022, 15, 601–609 This journal is © The Royal Society of Chemistry 2022
View Article Online

Energy & Environmental Science Paper

5 cycles of CO2 photoreduction. As can be seen from the STEM relaxation processes. Although cooperative catalytic interaction
images (Fig. S20a and b, ESI†), Cu1/TiO2 (0.21) maintains its of multiple adjacent active sites is not often considered in
initial morphology after a sequential CO2 photoreduction. heterogeneous photocatalysts, it has been demonstrated that
STEM-EDS analysis shows that the Cu atoms maintain their cooperation of adjacent active sites in homogeneous catalysts
homogeneous dispersion in the spent Cu1/TiO2 (0.21) (Fig. S21, or metal–organic-frameworks leads to different molecular inter-
Published on 20 September 2021. Downloaded by CHONNAM NATIONAL UNIVERSITY on 2/10/2023 5:32:35 AM.

ESI†). Atomic level imaging using Cs-STEM directly visualizes actions with CO2.54,55 The electronic effect of adjacent
single Cu atoms on the surface of Cu1/TiO2 (0.21) (Fig. S20c, Cu atoms on the active Cu site was investigated through
ESI†). These results are mutually consistent, demonstrating theoretical calculation by locating an additional copper atom
that the Cu atoms stably maintain their initial atomic disper- in Cu1/TiO2 x, denoted as Cu2/TiO2 x henceforth. While the Cu
sion after the CO2 photoreduction. EELS analysis was per- atom in Cu1/TiO2 x preferably binds CO2 with d(Cu–C) = 2.02 Å
formed on the spent Cu1/TiO2, featuring both four peaks (Fig. 4d) by the aforementioned Cu–O4 distortion, evidenced by
(Ti4+, site 2) and two peaks (Ti3+, site 1) (Fig. S22, ESI†), the axial Cu–O bond elongation from 1.86 to 2.17 Å and vertical
indicating that Cu atoms favourably regenerate the O vacancy, ascent of Cu atom by Dz(Cu) = +0.68 Å, Cu2/TiO2 x does not
as predicted by the theoretical calculation. Isotopic 13CO2 allow chemisorption of CO2 at the O vacancy site with
experiments and control experiments confirm CO2 as the d(Cu–C) = 3.05 Å and vertical descent of Cu atom by
hydrocarbon source (Fig. S18 and S19, ESI†). Cu1/TiO2 also Dz(Cu) = 0.20 Å (Fig. 4e). As shown in the molecular orbital
exhibited notable C2H6 production (64.2 ppm g 1 h 1). Recent density (Fig. S23, ESI†), the Cu–O4 active centre has well
studies reported that oxidative CH4 coupling to C2H6 can occur localized electrons in its low symmetry of the seesaw (C2v)
on the TiO2 surface upon light irradiation.18,58 Considering that structure, leading to strong Lewis basicity of the Cu atom.
C2H6 production is only observed on Cu1/TiO2, a high CH4 Therefore, the Cu atom can efficiently transfer electrons to
production rate is essential for C2H6 formation through further the oxygen p state (1a2) of CO2 and break the symmetry of CO2
coupling of CH4. We then investigated the effect of the loading into two p states of each terminal oxygen atom (Fig. S23a and b,
amount of Cu atoms on the photocatalytic CO2 reduction ESI†) and bend the C–O–C bond angle (1801 - 1471). Then, the
performance (Fig. 4c). The photocatalytic performance was single Cu site and the nearby O vacancy offer an active binding
heavily dependent on the concentration of Cu atoms. Interest- pocket (highlighted by an orange dotted circle in Fig. 4d), at
ingly, the photocatalytic CO2 reduction performance was max- which the bent CO2 can be favourably accommodated through
imized with a very low Cu atom amount (0.21 at%) and a bidentate interaction,50 while the terminal O p states are
gradually decreased with increasing Cu atom concentration. respectively interacting with the Cu–C bond and Ti–O bond as
Specifically, a significant decrease in CH4 production rate was shown in the structure and charge density difference in Fig. 4d.
observed for the catalyst with a Cu atom concentration of On the other hand, Cu2/TiO2 x can stabilize d electrons by a
1.42 at%, which is approximately 16 times lower than that with strong charge delocalization with the nearby Cu atom through
0.21 at%. In general, for nanoparticle-based cocatalysts, symmetrically matched d orbitals, which results in the distor-
enhancement in photocatalytic performance is observed upon tion of the Cu–O4 complex into a more symmetrical square
increasing the cocatalyst amount (typically up to 1–3 wt%), planar (D4h) structure, as shown in Fig. S23c and S24 (ESI†). As
which is known to originate from the increase in the number of a result, the structure relaxation of Cu–O4 induced by charge
active sites.1,2 In contrast, Cu1/TiO2 exhibited the highest CO2 delocalization prevents (1) the charge localization to the
reduction performance at very low concentrations (0.21 at% or Cu atom for strong basicity of a Cu atom and (2) the conserva-
0.17 wt%). Considering that all the Cu1/TiO2 samples with tion of a binding pocket structure, both of which are electro-
different concentrations are composed of single Cu atoms as nically and sterically required for chemical bond formation
cocatalysts, the observed phenomenon is apparently different between CO2 and Cu (Fig. 4d).
from the previous understanding. This indicates that factors
other than the number of active sites should be taken into
account to determine what significantly affects the photocata- Conclusions
lytic performance of single-atom photocatalysts.
In summary, we have investigated the atomic-level photocata-
Electronic effect of adjacent Cu atoms on CO2 photoreduction lytic process on the TiO2 surface using uniformly stabilized
Very recently, charge transfer between the support and single single Cu atoms on TiO2. Most importantly, our experimental
atom has been shown to modify the d state of stabilized metal and theoretical results are mutually consistent, highlighting
atoms through structural relaxation. This result indicates that that the redox-mediating property of the support can directly
the catalytic behaviour of SACs can be predicted by analogy to affect electronic interaction at the metal atom–support inter-
coordination chemistry.48 Structural relaxation heavily depends face. In the case of Cu1/TiO2, the presence of a single Cu atom
on neighbouring metal cofactors in coordination complexes partially oxidizes an adjacent surface O atom to stabilize the
where single or multiple metal atoms can lead to different metal–support interface, and thereby spontaneously generates
relaxation processes. We hypothesized that if catalytic inter- an O vacancy near the single Cu atom. CO2 is chemically
action on SACs resembles that of coordination complexes, the adsorbed on the O vacancy site by strong charge transfer of
presence of an adjacent single atom can also lead to different the localized electron in the single Cu atom. Consequently, an

This journal is © The Royal Society of Chemistry 2022 Energy Environ. Sci., 2022, 15, 601–609 | 607
View Article Online

Paper Energy & Environmental Science

isolated Cu atom on the TiO2 surface plays two important roles 5 X. Li, Y. Sun, J. Xu, Y. Shao, J. Wu, X. Xu, Y. Pan, H. Ju, J. Zhu
in the photocatalytic CO2 reduction. Firstly, it promotes the and Y. Xie, Nat. Energy, 2019, 4, 690.
O vacancy formation and replenishment by enhancing the 6 M. R. Hoffmann, S. T. Martin, W. Choi and D. W.
surface reducibility of the surrounding TiO2. Secondly, it Bahnemann, Chem. Rev., 1995, 95, 69.
induces electron localization to the active Cu centre that can 7 A. Holm, E. D. Goodman, J. H. Stenlid, A. Aitbekova,
Published on 20 September 2021. Downloaded by CHONNAM NATIONAL UNIVERSITY on 2/10/2023 5:32:35 AM.

offer a binding pocket for favourable chemisorption of a CO2 R. Zelaya, B. T. Diroll, A. C. Johnston-Peck, K. C. Kao,
molecule. The charge localization to the Cu atoms is disturbed C. W. Frank, L. G. M. Pettersson and M. Cargnello, J. Am.
when additional Cu atoms are closely located, which directly Chem. Soc., 2020, 142, 14481.
affects the CO2 photoreduction performance. Optimized 8 M. Cargnello, T. Montini, S. Y. Smolin, J. B. Priebe,
Cu1/TiO2 photocatalysts exhibit 66-fold enhancement in CO2 J. J. D. Jaén, V. V. T. Doan-Nguyen, I. S. McKay,
photoreduction performance compared to the pristine TiO2. J. A. Schwalbe, M. M. Pohl, T. R. Gordon, Y. Lu,
J. B. Baxter, A. Brückner, P. Fornasiero and C. B. Murray,
Proc. Natl. Acad. Sci. U. S. A., 2016, 113, 3966.
Author contributions 9 J. Yu, J. Low, W. Xiao, P. Zhou and M. Jaroniec, J. Am. Chem.
B. H. L., E. G., M. K., S. P., K. T. N., H. K., S. I. I. and T. H. Soc., 2014, 136, 8839.
conceived the research. B. H. L., E. G. and M. K. designed the 10 Y. Ji and Y. Luo, J. Am. Chem. Soc., 2016, 138, 15896.
experiments. B. H. L., E. J., C. W. L., Y. S. K. and Y. J. performed 11 S. Selcuk, X. Zhao and A. Selloni, Nat. Mater., 2018, 17, 923.
and analysed the material synthesis. E. G., J. L. and H. R. K. 12 X. Chen, L. Liu, P. Y. Yu and S. S. Mao, Science, 2011, 331, 746.
performed photochemical measurements. M. K., S. L. and H. K. 13 R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki and Y. Taga,
performed computational analysis. S. P. C. conducted the Science, 2011, 293, 269.
HAADF-STEM analysis. K. S. L. conducted the XAFS analysis. 14 S. U. M. Khan, M. Al-Shahry and W. B. Ingler Jr., Science,
J. W. J. and C. H. C. performed TRPL analysis. B. H. L., E. G., 2002, 297, 2243.
M. K., Y. S. K., S. P., K. T. N., H. K., S. I. I. and T. H. wrote the 15 D. O. Scanlon, C. W. Dunnill, J. Buckeridge, S. A. Shevlin,
manuscript. H. K., S. I. I. and T. H. supervised the project. All A. J. Logsdail, S. M. Woodley, C. R. A. Catlow, M. J. Powell,
the authors commented on the manuscript. R. G. Palgrave, I. P. Parkin, G. W. Watson, T. W. Keal,
P. Sherwood, A. Walsh and A. A. Sokol, Nat. Mater., 2013, 12, 798.
16 R. Long, Y. Li, Y. Liu, S. Chen, X. Zheng, C. Gao, C. He,
Conflicts of interest N. Chen, Z. Qi, L. Song, J. Jiang, J. Zhu and Y. Xiong, J. Am.
Chem. Soc., 2017, 139, 4486.
There are no conflicts to declare.
17 S. Neatu, J. A. Macia-Agullo, P. Concepcion and H. Garcia,
J. Am. Chem. Soc., 2014, 136, 15969.
Acknowledgements 18 S. Sorcar, J. Thompson, Y. Hwang, Y. H. Park, T. Majima,
C. A. Grimes, J. R. Durrant and S. I. In, Energy Environ. Sci.,
Synthesis and physicochemical property analysis of the nano- 2018, 11, 3183.
material samples was supported by the Research Center Pro- 19 J. Xie, R. Jin, A. Li, Y. Bi, Q. Ruan, Y. Deng, Y. Zhang, S. Yao,
gram of the IBS (IBS-R006-D1) in Korea (T.H.). X-ray absorption G. Sankar, D. Ma and J. Tang, Nat. Catal., 2018, 1, 889.
spectra characterization at Pohang Accelerator Laboratory (PAL) 20 J. Gu, C. S. Hsu, L. Bai, H. M. Chen and X. Hu, Science, 2019,
8C beamline was supported by a National Research Foundation 364, 1091.
of Korea (NRF) grant funded by the Korean government (No. 21 E. Jung, H. Shin, B. H. Lee, V. Efremov, S. Lee, H. S. Lee,
2019M3D1A1079309) (K.S.L.). Photochemical property analysis J. Kim, W. H. Antink, S. Park, K. S. Lee, S. P. Cho, J. S. Yoo,
of the materials was supported by the Ministry of Science and Y. E. Sung and T. Hyeon, Nat. Mater., 2020, 19, 436.
ICT in Korea (2021R1A2C2009459) (S.I.I.), and theoretical ana- 22 A. Wang, J. Li and T. Zhang, Nat. Rev. Chem., 2018, 2, 65.
lysis was supported by the Creative Materials Discovery Pro- 23 Y. Sun, L. Silvioli, N. R. Sahraie, W. Ju, J. Li, A. Zitolo, S. Li,
gram (Grant 2017M3D1A1039378) (M. K. S. L. and H. K.). A. Bagger, L. Arnarson, X. Wang, T. Moeller, D. Bernsmeier,
J. Rossmeisl, F. Jaouen and P. Strasser, J. Am. Chem. Soc.,
References 2019, 141, 12372.
24 D. W. Su, J. Ran, Z. W. Zhuang, C. Chen, S. Z. Qiao, Y. D. Li
1 J. Yang, D. Wang, H. Han and C. Li, Acc. Chem. Res., 2013, and G. X. Wang, Sci. Adv., 2020, 6, eaaz8447.
46, 1900. 25 B. H. Lee, S. Park, M. Kim, A. K. Sinha, S. C. Lee, E. Jung,
2 Y. Zhang, B. Xia, J. Ran, K. Davey and S. Z. Qiao, Adv. Energy W. J. Chang, K. S. Lee, J. H. Kim, S. P. Cho, H. Kim,
Mater., 2020, 10, 1903879. K. T. Nam and T. Hyeon, Nat. Mater., 2019, 18, 620.
3 X. Li, J. Yu, M. Jaroniec and X. Chen, Chem. Rev., 2019, 26 D. Yi, F. Lu, F. Zhang, S. Liu, B. Zhou, D. Gao, X. Wang and
119, 3962. J. Yao, Angew. Chem., Int. Ed., 2020, 132, 15989.
4 Y. Wang, A. Vogel, M. Sachs, R. S. Sprick, L. Wilbraham, 27 M. Xiao, L. Zhang, B. Luo, M. Lyu, Z. Wang, H. Huang,
S. J. A. Moniz, R. Godin, M. A. Zwijnenburg, J. R. Durrant, S. Wang, A. Du and L. Wang, Angew. Chem., Int. Ed., 2020,
A. I. Cooper and J. Tang, Nat. Energy, 2019, 4, 746. 132, 7297.

608 | Energy Environ. Sci., 2022, 15, 601–609 This journal is © The Royal Society of Chemistry 2022
View Article Online

Energy & Environmental Science Paper

28 S. Hejazi, S. Mohajernia, B. Osuagwu, G. Zoppellaro, 43 J. Resasco, L. DeRita, S. Dai, J. P. Chada, M. Xu, X. Yan,
P. Andryskova, O. Tomanec, S. Kment, R. Zbořil and J. Finzel, S. Hanukovich, A. S. Hoffman, G. W. Graham,
P. Schmuki, Adv. Mater., 2020, 32, 1908505. S. R. Bare, X. Pan and P. Christopher, J. Am. Chem. Soc.,
29 X. Li, W. Bi, L. Zhang, S. Tao, W. Chu, Q. Zhang, Y. Luo, 2020, 142, 169.
C. Wu and Y. Xie, Adv. Mater., 2016, 28, 2427. 44 J. Li, Q. Guan, H. Wu, W. Liu, Y. Lin, Z. Sun, X. Ye, X. Zheng,
Published on 20 September 2021. Downloaded by CHONNAM NATIONAL UNIVERSITY on 2/10/2023 5:32:35 AM.

30 X. Jin, R. Wang, L. Zhang, R. Si, M. Shen, M. Wang, J. Tian H. Pan, J. Zhu, S. Chen, W. Zhang, S. Wei and J. Lu, J. Am.
and J. Shi, Angew. Chem., Int. Ed., 2020, 132, 6894. Chem. Soc., 2019, 141, 14515.
31 Y. Zhao, Y. Zhao, R. Shi, B. Wang, G. I. N. Waterhouse, 45 R. Rousseau, V. A. Glezakou and A. Selloni, Nat. Rev. Mater.,
L. Z. Wu, C. H. Tung and T. Zhang, Adv. Mater., 2019, 2020, 5, 460.
31, 1806482. 46 R. Lang, X. Du, Y. Huang, X. Jiang, Q. Zhang, Y. Guo, K. Liu,
32 C. Liu, J. Qian, Y. Ye, H. Zhou, C. J. Sun, C. Sheehan, B. Qiao, A. Wang and T. Zhang, Chem. Rev., 2020, 120, 11986.
Z. Zhang, G. Wan, Y. S. Liu, J. Guo, S. Li, H. Shin, 47 B. Moss, Q. Wang, K. T. Butler, R. Grau-Crespo, S. Selim,
S. Hwang, T. B. Gunnoe, W. A. Goddard III and S. Zhang, A. Regoutz, T. Hisatomi, R. Godin, D. J. Payne, A. Kafizas,
Nat. Catal., 2020, 4, 36. K. Domen, L. Steier and J. R. Durrant, Nat. Mater., 2021,
33 M. D. Hossain, Y. Huang, T. H. Yu, W. A. Goddard III and 20, 511.
Z. Luo, Nat. Commun., 2020, 11, 2256. 48 J. Hulva, M. Meier, R. Bliem, Z. Jakub, F. Kraushofer,
34 J. Di, C. Chen, S. Z. Yang, S. Chen, M. Duan, J. Xiong, C. Zhu, M. Schmid, U. Diebold, C. Franchini and G. S. Parkinson,
R. Long, W. Hao, Z. Chi, H. Chen, Y. X. Weng, J. Xia, L. Song, Science, 2021, 371, 375.
S. Li, H. Li and Z. Liu, Nat. Commun., 2019, 10, 2840. 49 J. Y. Liu, X. Q. Gong, R. Li, H. Shi, S. B. Cronin and
35 J. Wang, T. Heil, B. Zhu, C. W. Tung, J. Yu, H. M. Chen, A. N. Alexandrova, ACS Catal., 2020, 10, 4048.
M. Antonietti and S. Cao, ACS Nano, 2020, 14, 8584. 50 J. Jia, C. Qian, Y. Dong, Y. F. Li, H. Wang, M. Ghoussoub,
36 G. Wang, C. T. He, R. Huang, J. Mao, D. Wang and Y. Li, K. T. Butler, A. Walsh and G. A. Ozin, Chem. Soc. Rev., 2017,
J. Am. Chem. Soc., 2020, 142, 19339. 46, 4631.
37 G. Gao, Y. Jiao, E. R. Waclawik and A. Du, J. Am. Chem. Soc., 51 H. G. Yang and H. C. Zeng, J. Phys. Chem. B, 2004, 108, 3492.
2016, 138, 6292. 52 T. Du, J. Kim, J. Ngiam, S. Xu, P. R. F. Barnes, J. R. Durrant
38 H. Fei, J. Dong, Y. Feng, C. S. Allen, C. Wan, B. Volosskiy, and M. A. McLachlan, Adv. Funct. Mater., 2018, 28, 1801808.
M. Li, Z. Zhao, Y. Wang, H. Sun, P. An, W. Chen, Z. Guo, 53 E. V. Péan, S. Dimitrov, C. S. D. Castroc and M. L. Davies,
C. Lee, D. Chen, I. Shakir, M. Liu, T. Hu, Y. Li, A. I. Kirkland, Phys. Chem. Chem. Phys., 2020, 22, 28345.
X. Duan and Y. Huang, Nat. Catal., 2018, 1, 63. 54 B. An, Z. Li, Y. Song, J. Zhang, L. Zeng, C. Wang and W. Lin,
39 S. Ji, Y. Chen, X. Wang, Z. Zhang, D. Wang and Y. Li, Chem. Nat. Catal., 2019, 2, 709.
Rev., 2020, 120, 11900. 55 Z. Guo, G. Chen, C. Cometto, B. Ma, H. Zhao, T. Groizard,
40 P. Zhou, Q. Zhang, Z. Xu, Q. Shang, L. Wang, Y. Chao, Y. Li, L. Chen, H. Fan, W. L. Man, S. M. Yiu, K. C. Lau, T. C. Lau
H. Chen, F. Lv, Q. Zhang, L. Gu and S. Guo, Adv. Mater., and M. Robert, Nat. Catal., 2019, 2, 801.
2020, 32, 1904249. 56 S. I. In, D. D. Vaughn II and R. E. Schaak, Angew. Chem., Int.
41 N. Daelman, M. Capdevila-Cortada and N. Lopez, Nat. Ed., 2012, 51, 3915.
Mater., 2019, 18, 1215. 57 Y. Lan, Y. Xie, J. Chen, Z. Hu and D. Cui, Chem. Commun.,
42 L. DeRita, J. Resasco, S. Dai, A. Boubnov, H. V. Thang, 2019, 55, 8068.
A. S. Hoffman, I. Ro, G. W. Graham, S. R. Bare, G. Pacchioni, 58 X. Li, J. Xie, H. Rao, C. Wang and J. Tang, Angew. Chem., Int.
X. Pan and P. Christopher, Nat. Mater., 2019, 18, 746. Ed., 2020, 59, 19702.

This journal is © The Royal Society of Chemistry 2022 Energy Environ. Sci., 2022, 15, 601–609 | 609

You might also like