ArticlePDF Available

Electronic metal-support interaction modulates Cu electronic structures for CO2 electroreduction to desired products

Springer Nature
Nature Communications
Authors:

Abstract and Figures

In this work, the Cu single-atom catalysts (SACs) supported by metal-oxides (Al2O3-CuSAC, CeO2-CuSAC, and TiO2-CuSAC) are used as theoretical models to explore the correlations between electronic structures and CO2RR performances. For these catalysts, the electronic metal-support interaction (EMSI) induced by charge transfer between Cu sites and supports subtly modulates the Cu electronic structure to form different highest occupied-orbital. The highest occupied 3dyz orbital of Al2O3-CuSAC enhances the adsorption strength of CO and weakens C-O bonds through 3dyz-π* electron back-donation. This reduces the energy barrier for C-C coupling, thereby promoting multicarbon formation on Al2O3-CuSAC. The highest occupied 3dz2 orbital of TiO2-CuSAC accelerates the H2O activation, and lowers the reaction energy for forming CH4. This over activated H2O, in turn, intensifies competing hydrogen evolution reaction (HER), which hinders the high-selectivity production of CH4 on TiO2-CuSAC. CeO2-CuSAC with highest occupied 3dx2-y2 orbital promotes CO2 activation and its localized electronic state inhibits C-C coupling. The moderate water activity of CeO2-CuSAC facilitates *CO deep hydrogenation without excessively activating HER. Hence, CeO2-CuSAC exhibits the highest CH4 Faradaic efficiency of 70.3% at 400 mA cm⁻².
This content is subject to copyright. Terms and conditions apply.
Article https://doi.org/10.1038/s41467-025-57307-6
Electronic metal-support interaction
modulates Cu electronic structures for CO
2
electroreduction to desired products
Yong Zhang
1
,FeifeiChen
1
,XinyiYang
1
,YiranGuo
1
, Xinghua Zhang
2
,
Hong Dong
1
,WeihuaWang
1
,FengLu
1
,ZunmingLu
2
,HuiLiu
3
,HuiLiu
1
,
Yao Xiao
4
&YahuiCheng
1
In this work, the Cu single-atom catalysts (SACs) supported by metal-oxides
(Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
) are used as theoretical models to
explore the correlations between electronic structures and CO
2
RR perfor-
mances. For these catalysts, the electronic metal-support interaction (EMSI)
induced by charge transfer between Cu sites and supports subtly modulates
the Cu electronic structure to form different highest occupied-orbital. The
highest occupied 3d
yz
orbital of Al
2
O
3
-Cu
SAC
enhances the adsorption strength
of CO and weakens C-O bonds through 3d
yz
-π* electron back-donation. This
reduces the energy barrier for C-C coupling, thereby promoting multicarbon
formationonAl
2
O
3
-Cu
SAC
. The highest occupied 3d
z2
orbital of TiO
2
-Cu
SAC
accelerates the H
2
O activation, and lowers the reaction energy for forming
CH
4
. This over activated H
2
O, in turn, intensies competing hydrogen evolu-
tion reaction (HER), which hinders the high-selectivity production of CH
4
on
TiO
2
-Cu
SAC
.CeO
2
-Cu
SAC
with highest occupied 3d
x2-y2
orbital promotes CO
2
activation and its localized electronic state inhibits C-C coupling. The mod-
erate water activity of CeO
2
-Cu
SAC
facilitates *CO deep hydrogenation without
excessively activating HER. Hence, CeO
2
-Cu
SAC
exhibits the highest CH
4
Far-
adaic efciency of 70.3% at 400 mA cm2.
Renewable energy-driven electrochemical carbon dioxide reduction
reaction (CO
2
RR) is a promising method for achieving carbon cycle
and clean production of chemicals1,2.CO
2
molecules are deeply
reduced on the Cu-based catalyst surface into hydrocarbons and
oxygen-containing compounds such as methane (CH
4
), ethylene
(C
2
H
4
), ethanol (EtOH), etc., through multiple electron-proton coupled
steps of CO
2
RR3. Although these high value chemicals and high energy
density fuels have broad markets, industrial-scale implementation of
CO
2
RR still has a long way to go4. On the one hand, in aqueous media,
water (H
2
O) molecules serve as the proton source for electrochemical
reactions, leading to a conict between CO
2
RR and hydrogen evolu-
tion reaction (HER)58. The proton-electron coupling properties of
CO
2
RR require effective activation of H
2
O and smooth proton transfer
to avoid excessive activation of H
2
O, otherwise HER would competi-
tively overwhelm CO
2
RR. On the other hand, deep hydrogenation of
adsorbed CO (*CO) and C-C coupling often coexist and compete with
each other, resulting in lowproduct selectivity7,911. The adsorption and
coverage of key *CO intermediate on the catalyst surface are crucial in
controlling the selectivi ty of C
2
products10,11. Hence, the rational design
and controllable synthesis of catalysts based on the deep under-
standing of reaction mechanism and structure-activity relationship is
crucial for precise regulation of the competitive pathways for CO
2
RR.
Modulating the electronic structure of a catalyst and elucidating
its inuence over catalytic activity is an effective approach for studying
Received: 23 August 2024
Accepted: 18 February 2025
Check for updates
1
Department of Electronic Science and Engineering, Nankai University, Tianjin, China.
2
School of Material Science and Engineering, Hebei University of
Technology, Tianjin, China.
3
Institute of New-Energy Materials, Tianjin University, Tianjin, China.
4
College of Chemistry and Materials Engineering, Wenzhou
University, Wenzhou, China. e-mail: hui_liu@tju.edu.cn;liuhui@nankai.edu.cn;xiaoyao@wzu.edu.cn;chengyahui@nankai.edu.cn
Nature Communications | (2025) 16:1956 1
1234567890():,;
1234567890():,;
Content courtesy of Springer Nature, terms of use apply. Rights reserved
the structure-activity relationships. To modulate Cu electronic struc-
ture, research interest has concentrated on alloying12,13,doping
engineering14,15, metal-support interaction (MSI) modulation16,17,andso
on. In heterogeneous catalysis, MSI signicantly affects the catalytic
performance as it modulates the electronic and geometric structures
of metal as well as coordination environments. The electronic metal-
support interaction (EMSI) was further proposed by Campbell, which
goes beyond MSI and provides a much more detailed explanation of
the enhanced properties of supported catalysts than MSI1820.EMSIis
associated with the orbital rehybridization and charge transfer across
the metal-support interface, leading to the formation of new chemical
bonds and the realignment of molecular energy levels1921. By precisely
controlling the d-band structure of metal through EMSI, the adsorp-
tion scaling relation can be disrupted to regulate the adsorption of key
intermediates. Thisis crucial for the precise control of CO
2
RR pathway
because protonating to form key intermediates (such as *COOH and
*CHO) is difcult on weakly adsorption surface, while over-
strengthening of the adsorption energy would favor HER according
to the adsorption scaling relation10,22. However, limited by the intrinsic
bulk effects, accurately identifying the orbital information of metal
particles or clusters still poses challenges inproviding design guidance
or performance enhancement explanations for supported catalysts20.
The EMSI based on single-atom catalysts (SACs) provides a bridge
for theoretical electronic structure studies and the design of hetero-
genous catalysts because their electronic structure can be easily
characterized through experiments and theoretical calculations20,2325.
Strong EMSI not only stabilizes the single-atom metals due to the
thermodynamicallyfavorable metal-support bond formation, but also
leads to the charge redistribution induced via electron transfer,
affecting the energy level distribution of the 3 d orbitals of single
atoms2630. For example, Ma et al. reported that the coordination
environment of Fe-N
5
raises the energy level of Fe 3d
z2
orbital com-
pared to that of Fe-N
4
, thereby enhancing COOH adsorption and
promoting *CO desorption26. Therefore, the energy level of Cu 3dstate
may be manipulated by coordination structural perturbations of the
Cu atom30. Although SACs have been widely used for CO
2
RR, general
relationships between electronic structures and the catalytic behaviors
of SAC remain unclear. Hence, a comprehensive atomic-level insights
into the structure-activity relationship of Cu SAC is urgently needed
for guiding the regulation of *CO deep hydrogenation or coupling.
Herein, the Cu SACs supported by Al
2
O
3
(Al
2
O
3
-Cu
SAC
), CeO
2
(CeO
2
-Cu
SAC
), and TiO
2
(TiO
2
-Cu
SAC
) are constructed by atomic layer
deposition (ALD) technique and the correlations between the char-
acteristics of electronic structures and the CO
2
RR performance are
rationalized through detailed characterization and density functional
theory (DFT) calculations. The switching of supports subtly modulates
the electronic structure of Cu sites to form three SACs with completely
different highest occupied orbitals (3d
yz
orbital for Al
2
O
3
-Cu
SAC
,3d
x2-y2
orbital for CeO
2
-Cu
SAC
,and3d
z2
orbital for TiO
2
-Cu
SAC
). The 3d
yz
of
Al
2
O
3
-Cu
SAC
tends to interact with the π* anti-bonding orbital of CO,
which enhance CO adsorption on Al
2
O
3
-Cu
SAC
and weaken the C-O
bonds through 3d
yz
-π* electron back-donation. Meanwhile, the Cu
electron delocalization increases the CO adjacent-adsorption energies
on Al
2
O
3
-Cu
SAC
. These enhance the C-C coupling on Al
2
O
3
-Cu
SAC
,
resulting in a lowest ratio of CH
4
Faradaic efciency (FE) to C
2
FE
(FE
CH4
/FE
C2
, 1.08). The3d
z2
of TiO
2
-Cu
SAC
exhibits strong hybridization
with the σbonding and σ* anti-bonding orbitals of H
2
O, which
enhances the adsorption of H
2
O and promotes H
2
O dissociation by
weakening O-H bonds. Therefore, TiO
2
-Cu
SAC
exhibits a highest FE
CH4
/
FE
C2
of 4.14. However, this over activated H
2
O, in turn, intensies
competing HER, which hinders the high-selectivity production of CH
4
on TiO
2
-Cu
SAC
.The3d
x2-y2
orbital effectively promotes CO
2
activation
and its localized Cu electronic state inhibits C-C coupling. This makes
CeO
2
-Cu
SAC
exhibit a FE
CH4
/FE
C2
of 3.16. Meanwhile, the moderate
water activity of CeO
2
-Cu
SAC
facilitates *CO deep hydrogenation
without excessively activating HER, resulting in the highest CH
4
FE of
70.3% at 400 mA cm2.
Results
Theoretical analysis of electronic structure
To explore the structure-activity relationship of Cu single-atom elec-
tronic structure in selective *CO deep hydrogenation or coupling,
theoretical analysis based on DFT + U calculations was conducted. By
loading Cu single atoms on Al
2
O
3
,CeO
2
, and TiO
2
supports to mod-
ulate the electronic structure of Cu single-atom (Supplementary Fig. 1
and Supplementary data 1), named as Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,and
TiO
2
-Cu
SAC
. As shown in Supplementary Fig. 2, the partial density of
states (PDOS) of Cu 3 dorbitals of Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
, and TiO
2
-
Cu
SAC
surfaces are calculated, and the respective energy levels of 3 d
xy
,
3d
x2-y2
,3d
yz
,3d
xz
,and3d
z2
are plotted based on their d-band center
(Fig. 1a)30. Instead of forming a broad d-band, each 3d state is spatially
highlylocalized with a narrow energy window. On Al
2
O
3
-Cu
SAC
,the3d
yz
is the highest occupied-orbital; on CeO
2
-Cu
SAC
,the3d
x2-y2
is the
highest occupied-orbital; while on TiO
2
-Cu
SAC
,the3d
z2
is the highest
occupied-orbital. Clearly, changing the support can signicantly affect
the highest occupied-orbital of the Cu single-atom due to the orbital
rehybridization and charge transfer across the metal-support interface
(Supplementary Fig. 3)21.Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
each present Cu single-atoms with three completely different highest
occupied-orbitals. The highest occupied d-orbitals often play an
important role in regulating the binding modes of intermediates
because the electrons lled on the orbital closer to the Fermilevel (E
f
)
are more active27. From the orbital wave functions, it can be observed
that the 3d
yz
orbital (localize in Zaxis) symmetry and energy-match
with the π* orbital of CO, facilitating the formation of 3d
yz
-π*bonds
(Fig. 1b). The 3d
z2
orbital symmetry and energy-match with the σ
orbital of CO as well as σ/σ* orbital of H
2
O, thereby aiding in the
formation of d
z2
-σ/σ* bonds (Fig. 1b and Supplementary Fig. 4). The
3d
x2-y2
orbital symmetry and energy-match with the π*orbitalsofCO
2
,
promoting the formation of 3d
x2-y2
-π* bonds.
Figure 1c and Supplementary Figs. 5, 6 show the PDOS of CO and
H
2
O molecular frontier orbitals as well as the PDOS of CO and H
2
O
adsorption on Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
surfaces. The
PDOS proves that the strong hybridization between CO and Cu atom
mainly contributes from the 3d
yz
and 3d
z2
orbital, and the hybridiza-
tion between H
2
O and Cu mainly contributes from the 3d
z2
orbital. The
hybridization of the Cu 3dorbital with the CO σis primarily con-
tributed by the 3d
z2
orbital, while the contribution to the CO π*is
mainly from the 3d
yz
orbital. This aligns with the analysis of the orbital
wave functions. On Al
2
O
3
-Cu
SAC
,the3d
z2
orbitalof Cu hybridize with σ
orbitalof CO, while stronger hybridization occurs between the highest
occupied 3d
yz
orbital and π* orbital of CO than that of CeO
2
-Cu
SAC
and
TiO
2
-Cu
SAC
31. The d-band center of 3d
yz
(0.26 eV) on Al
2
O
3
-Cu
SAC
are
closer to the E
f
than that on CeO
2
-Cu
SAC
(0.50 eV) and TiO
2
-Cu
SAC
(0.64 eV), making more active electrons in 3 d
yz
easily back-donate to
the π* of CO as formation of Cu-CO bonds29,32. Moreover, due to anti-
bonding feature of the π* orbital, the 3d
yz
-π* bonding will weaken C-O
bonds in *CO, which can facilitate the subsequent reactions of *CO
(hydrogenation or coupling)29,32. In contrast, the highest occupied 3d
z2
orbital of TiO
2
-Cu
SAC
exhibits strong hybridization with the σorbital of
CO, while the hybridization between 3d
yz
and π* is lower than that of
Al
2
O
3
-Cu
SAC
(Supplementary Fig. 6a). Due to the bonding and anti-
bonding orbitals of H
2
Oisσand σ*, the hybridization between Cu 3d
and H
2
OonTiO
2
-Cu
SAC
are remarkably stronger than that on Al
2
O
3
-
Cu
SAC
and CeO
2
-Cu
SAC
(Fig. 1c and Supplementary Fig. 6b, d), leading
to an enhanced ability for TiO
2
-Cu
SAC
to adsorb H
2
O. In addition, the
interaction between the d
z2
orbital and the anti-bonding σ*orbital
weakens the O-H bonds, thereby promoting the dissociation of H
2
O.
On CeO
2
-Cu
SAC
, the highest occupied 3d
x2-y2
orbital tends to couple
with the lowest π*orbitalofCO
2
(Supplementary Fig. 7), weakening
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 2
Content courtesy of Springer Nature, terms of use apply. Rights reserved
C-O bonds and favoring hydrogenation on the low-coordinate O atom
to form the crucial *COOH intermediate, thereby promoting the acti-
vation of CO
2
to form *COOH27,33.
The CO adsorption energies of Cu single-atom on these three
supports follows the order Al
2
O
3
-Cu
SAC
(1.22 eV) > CeO
2
-Cu
SAC
(0.74 eV) > TiO
2
-Cu
SAC
(0.46 eV), consistent with the energy level
arrangement of the 3d
yz
orbital (Supplementary Fig. 8 and Fig. 1a). The
H
2
O adsorption energies are strongly related to the position of the 3d
z2
orbital energy level, showing the order TiO
2
-Cu
SAC
(0.89 eV) > Al
2
O
3
-
Cu
SAC
(0.30 eV) > CeO
2
-Cu
SAC
(0.16 eV). The crystal orbital Hamilton
population (COHP) was further used to analyze the Cu-C bond
strength in Cu-CO andthe Cu-O bond strength in Cu-H
2
O(Fig.1d, e and
Supplementary Fig. 9)34.OnAl
2
O
3
-Cu
SAC
, the Cu single-atom exhibits
the strongest Cu-C bonding with an integrated COHP (ICOHP) of
3.88 eV, followed by CeO
2
-Cu
SAC
(3.71 eV) and TiO
2
-Cu
SAC
(3.65 eV).
Below the E
f
, there are fewer lling electrons in the anti-bonding
orbitals for Cu-C compared to that of Cu-O on Al
2
O
3
-Cu
SAC
, indicating
that 3d
yz
is more favorable for bonding with CO. In contrast, TiO
2
-
Cu
SAC
exhibits the highest ICOHP of 2.45 eV, suggesting that the
3d
z2
orbital is favor for 3d
z2
-H
2
O bonds. Moreover, TiO
2
-Cu
SAC
exhibits
the lowest H
2
O dissociation energy of 0.52 eV, succeeded by Al
2
O
3
-
Cu
SAC
and CeO
2
-Cu
SAC
(Fig. 1f). This provides sufcient activated
protons for *CO deep hydrogenation to form CH
4
.However,excessive
activation of H
2
O could enhance competitive HER, and the formation
energy of H
2
on TiO
2
-Cu
SAC
is as low as 0.18 eV, which may signicantly
suppress CO
2
RR activity (Supplementary Fig. 10). On the pure CeO
2
(111) surface far away from Cu single-atom sites, the dissociation
energy of H
2
O is 1.33eV, which can also provide sufcient protons for
CO adsorbed on Cu single-atom sites hydrogenation.
After coordination between Cu and O, a large amount of charge
accumulates around the O atom, while Cu carries the positive charge
due to the higher electronegativity of O than that of Cu. Bader charge
analysis shows that the Bader charge of Cu is 10.13 |e
|inAl
2
O
3
-Cu
SAC
,
10.57 |e
|inCeO
2
-Cu
SAC
,and10.37|e
|inTiO
2
-Cu
SAC
(Supplementary
Al2O3-CuSAC TiO2-CuSAC
d
Cu Cu
A
B
C
g
Al2O3-CuSAC CeO2-CuSAC TiO2-CuSAC
Cu Cu
Cu 3dyz
CO π* CO σ
abc
ef
hi
PDOS (a.u.)
dyz-π*
dyz/dz2-σ/π
3dyz-π*
Cu 3dz2
3dz2
y
x
z
Cu 3dz2
3dz2
H2O σ
1
0
dz2-σ*
dz2
π
E-EF (eV)
Fig. 1 | Electronic structure analysis. a The Cu 3 ddiagrams of Al
2
O
3
-Cu
SAC
,CeO
2
-
Cu
SAC
, and TiO
2
-Cu
SAC
, respectively. bThe binding modes of CO and H
2
O inter-
acting withCu site. cThe Cu 3 dPDOS of Al
2
O
3
-CuSAC with adsorbed CO and H
2
O,
as well as PDOSof CO and H
2
O orbitals after adsorbed(The inset is anenlarged view
ofthedashedlineinc). The COHP curves of CuC bonds and Cu-O bonds in
dAl
2
O
3
-Cu
SAC
and eTiO
2
-Cu
SAC
.fFree energy diagram for the dissociation of *H
2
O.
gThe ELF of Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
, and TiO
2
-Cu
SAC
.hSchematic diagram of CO
adjacent-adsorption on Al
2
O
3
-Cu
SAC
surface. iThe CO adsorption energy of Al
2
O
3
-
Cu
SAC
,CeO
2
-Cu
SAC
, and TiO
2
-Cu
SAC
. Relevant source data are provided as a Source
Data le.
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 3
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Fig. 11). The electron localization function (ELF) obtained from DFT
calculations indicates that Al
2
O
3
attracts more electrons from Cu,
leading to electron delocalization around the Cu single atom (Fig.1g)35.
Further calculations are performed for the top-adsorption energy and
adjacent-adsorption energies of CO on the Cu single-atom, and the
corresponding adsorption modes are shown in Fig. 1h and Supple-
mentary Fig. 12. On Al
2
O
3
-Cu
SAC
, there is only a slight decrease in the
CO adjacent-adsorption energies (0.35 to 1.44 eV) compared to the
top-adsorption energy (Fig. 1i). This indicates that the delocalized
electronic state of Cu on Al
2
O
3
-Cu
SAC
enables more CO molecules to
adsorb around the Cu atom for C-C coupling36.OnCeO
2
-Cu
SAC
,the
electronegativity of Cu is higher than that of Ce, enabling Cu to attract
electrons from Ce, signicantly reducing the electron delocalization
around the Cu single-atom (Fig. 1g). As a result, the adjacent-
adsorption energies of CO (0.35 to 0.65 eV) on the Cu atom sig-
nicantly decreases compared to top-adsorption energy, making *CO
more likely to undergo deep hydrogenation in the form of individual
top-adsorption rather than C-C coupling.
Synthesis and characterization
To verify the results of DFT calculations, Cu atoms were deposited on
α-Al
2
O
3
, rutile TiO
2
,anduorite CeO
2
supports using ALD technique.
Scanning electron microscopy (SEM) and transmission electron
microscopy (TEM) images indicate that the morphologies of these
three supports consist of irregular nanoparticles (NPs) with particle
sizes larger than 50 nm, thus avoiding size effects of supports (Sup-
plementary Figs. 13 and 14)37. X-ray diffraction (XRD) patterns and high-
resolution TEM (HRTEM) images conrm that the supports are pure
phases of Al
2
O
3
,TiO
2
,andCeO
2
(Supplementary Figs. 14 and 15). Ben-
eting from the ALD technique relying on sequential molecular-level
self-limiting surface reactions38,39, the Cu atom can anchor on supports
in the form of a single atom (Fig. 2a). The Inductively coupled plasma
optical emission spectrometer (ICP-OES) measurements indicate that
the Cu loading amounts for Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
are 1.28 ± 0.23, 1.43 ± 0.09, and 1.76 ± 0.17 wt%, respectively (Supple-
mentary Table 1). Aberration-corrected high-resolution high-angle
annular dark-eld scanning TEM (HAADF-STEM) images show that Cu
atoms are atomically dispersed in Al
2
O
3
,TiO
2
,andCeO
2
supports
(Fig. 2bd), without the presence of any visible clusters at both low and
high magnications (Supplementary Fig. 16). XRD patterns and SEM
images of the samples after ALD further conrm that no observable Cu
NPs are deposited on supports (Supplementary Figs. 17 and 18).
HAADF-STEM and energy dispersive X-ray spectroscopy (EDS) element-
mapping images indicate that Cu is evenly distributed throughout the
Al
2
O
3
,TiO
2
,andCeO
2
supports (Fig. 2e, f). By controlling the ALD
process, Cu NPs loaded heterogenous catalysts were also prepared to
serve as an expansion of EMSI catalysts (Supplementary Fig. 19). The
catalysts loaded with Cu NPs on Al
2
O
3
and CeO
2
are named Al
2
O
3
-
CuNPs and CeO
2
-CuNPs, respectively. The Cu loading amounts for
Al
2
O
3
-CuNPs and CeO
2
-CuNPs are 2.91 ± 0.34 and 3.34 ± 0.53 wt%,
respectively (Supplementary Table 2). The statistical particle size dis-
tribution of Cu NPs based on TEM images displays that the average
diameter of Cu NPs loaded onto Al
2
O
3
and CeO
2
are 20.0 ± 0.3 nm and
17.4 ± 0.4 nm, respectively (Supplementary Figs. 20 and 21). HRTEM
images of Al
2
O
3
-CuNPs and CeO
2
-CuNPs both reveal interfaces
between Cu NPs and the supports, with the Cu NPs partly and rmly
socketed onto the support surface (Supplementary Figs. 20e and 21b).
To acquire the structural information of Cu/metal-oxides SAC,
synchrotron-based X-ray adsorption spectroscopy and X-ray photo-
electron spectroscopy (XPS) were performed. In the X-ray absorption
near-edge structure (XANES) spectra, the absorption edges for Al
2
O
3
-
Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
shift to higher energy than that of
Cu-foil, indicating that the Cu atoms are in an oxidized state (Supple-
mentary Fig. 22)40.Therst peaks of Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,and
TiO
2
-Cu
SAC
are similar and close to that of CuO (8986.4 eV) in the rst
derivative of XANES spectra (Fig. 3a), indicating that the valence of Cu
species in Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
are close to +241.
Only one apparent peak at 1.50 Å corresponding to the rst coordi-
nation shell of Cu-O scattering can be detected in the extended X-ray
absorption ne structure (EXAFS) spectra of Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,
and TiO
2
-Cu
SAC
(Fig. 3b). This conrms that Cu atoms are atomically
dispersed because no Cu-O-Cu or CuCu metallic bonds are observed.
The EXAFS tting data for Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
are shown in Fig. 3c, Supplementary Fig. 23, and Supplementary
Table 3. The average Cu-O coordination numbers of Al
2
O
3
-Cu
SAC
,
CeO
2
-Cu
SAC
, and TiO
2
-Cu
SAC
are 3.37, 3.67, and 3.41, respectively.
Combined with the results of theoretical calculations (Cu
1
-O
3
has the
lowest free energy; Supplementary Fig. 1), these data reveal that the
local structure of Cu/metal-oxides SAC comprised one isolated Cu
atom may coordinate with about three oxygen atoms (Cu
1
-O
3
units) in
the triangular pyramid structure between the Cu and O atoms
(Fig. 3d)41,42.Thetting of Cu 2pXPS spectra indicate that the pre-
dominant oxidation state on the surfaces of Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,
and TiO
2
-Cu
SAC
is Cu2+, with proportions of 86.2%, 85.5%, and 86.4%,
respectively. (Fig. 3e). Compared to CuO, the Cu 2pXPS peaks of Al
2
O
3
-
Cu
SAC
and TiO
2
-Cu
SAC
shift to higher binding energy, while Ti 2pand Al
2pXPS peaks shift to lower binding energy than that of pureAl
2
O
3
and
TiO
2
(Supplementary Fig. 24). It means that the electron transfers from
Cu to the adjacent Al
2
O
3
and TiO
2
supports, which results from the
strong EMSI between Cu and metal-oxide supports. In contrast, the Cu
2pXPS peak shifts to lower binding energy on the surface of CeO
2
-
Cu
SAC
, indicating that Cu attracts electrons from CeO
2
supports, pos-
sibly due to the higher electronegativity of Cu compared to Ce atom.
These results of electron transfer between Cu and the supports are
consistent with DFT calculations.
Using CO as a probe molecule, the atomic dispersion of Cu atoms
and the stability of Cu single-atoms during the CO
2
RR process can be
characterized by in-situ Raman spectroscopy (Fig. 3f)43.Inthe0.1M
CO-saturated KHCO
3
solution, Raman peaks attributed to linear top
adsorption (atop) of CO (C-O stretching vibration) are observed at
2087 and 2058 cm1for Cu
2
O and Cu, respectively44. Bridged adsorp-
tion of CO (i.e., CO adsorbed on two adjacent Cu atoms) is also
observed on the Cu surface. In contrast, bridged adsorption of CO is
not detected on the Cu SAC surface. Raman spectra collected on the
supports show that there is no peak in the range from 1800 cm1to
2200 cm1, suggesting that there is no CO adsorption on the supports
(Supplementary Fig. 25). Therefore, the Raman peaks located at 2110,
2105, and 2123 cm1can be attributed to CO atop on the Cu single-sites
of Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
45. DFT calculations based
on the Blyholder model reveal differences in the frequency of C-O
stretching vibrations between Cu single sites and Cu cluster sites
(Supplementary Fig. 26), providing the basis for spectroscopic mea-
surements to assess the stability of Cu single atoms. The stability of Cu
single-atoms correlates well with the afnity between the Cu sites and
supports, specically, strong interactions between Cu and O lead to
high stability43. To evaluate the stability of Cu single-atoms during the
electroreduction process, a constant potential of 1.5 V (all potentials
in this work are vs. the reversible hydrogen electrode, RHE) was
applied to Al
2
O
3
-Cu
SAC
electrode for 60 min. After electrolysis at 1.5 V
for 60 min, no corresponding Raman peaks are observed for CO
bridge-adsorption on metal Cu sites (Fig. 3g), indicating that Cu
remains anchored in supports in the form of single-atoms during the
electrolysis process. The reduce in intensity of CO atop Raman peaks
could be attributed to the accumulation of bubbles during long-term
electrolysis. Similarly, the stability of the supports during electrolysis is
maintained due to the high reduction potentials of Al
2
O
3
,CeO
2
,and
TiO
2
.Specically, after prolonged CO
2
electrolysis, CeO
2
-Cu
SAC
does
not show Raman peaks caused by oxygen defects of CeO
2
(570 cm1),
and the rst-order F
2g
Raman peak characteristic of CeO
2
(463 cm1)
remains present during the whole CO
2
RR process (Fig. 3h)46.In-situ
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 4
Content courtesy of Springer Nature, terms of use apply. Rights reserved
[0 0 1]
1 nm
Cu
Cu
2 nm
[1 1 1]
Cu
[0 1 0]
1 nm
HAADF
50 nm
Cu Al O
100 nm
100 nm
HAADF Cu Ce O
HAADF Cu Ti O
a
bcd
e
f
g
50 nm
100 nm
100 nm
50 nm
100 nm
100 nm
50 nm
100 nm
100 nm
Fig. 2 | Synthesis and structural characterization. a Schematic of the synthesis
process for Cu SAC support with metal-oxide. High-resolution HAADF-STEM ima-
ges of bAl
2
O
3
-Cu
SAC
,cCeO
2
-Cu
SAC
,anddTiO
2
-Cu
SAC
. Atomic models: Al, gray; Ce,
yellow; Ti, bule; and O, red. HAADF-STEM and EDS element-mapping images of
eAl
2
O
3
-Cu
SAC
,fCeO
2
-Cu
SAC
,andgTiO
2
-Cu
SAC
.
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 5
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Raman spectra of Al
2
O
3
-Cu
SAC
and TiO
2
-Cu
SAC
also conrm their sta-
bility during long-term electrolysis (Supplementary Fig. 27)47.
CO
2
RR performance
Electrochemical experiments on Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
, and TiO
2
-
Cu
SAC
were conducted in a ow cell with constant potential electro-
lysis. The main gas products observed during CO
2
RR are CH
4
,C
2
H
4
,
CO, and H
2
, with minor liquid products (Supplementary Fig. 28). Prior
to assessing the CO
2
RR performance of Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,and
TiO
2
-Cu
SAC
, thorough consideration was givento exclude the potential
impacts of surface hydroxide adsorption, charge transfer resistance of
the support, and Cu loading amounts on the CO
2
RR performance of
catalysts through a series of electrochemical, BET, and SEM measure-
ments (Supplementary Figs. 2931 and Supplementary Note 1). The
linear sweep voltammetry (LSV) curves indicate that the current den-
sities of Al
2
O
3
-Cu
SAC
increase the fastest with potential, followed by
CeO
2
-Cu
SAC
and TiO
2
-Cu
SAC
(Fig. 4a). Meanwhile, the current densities
of the samples measured with CO
2
gas at all applied potentials are
higher than those recorded with Ar gas, indicating the occurrence of
CO
2
RR. The smallest difference in current density values between
measurements of TiO
2
-Cu
SAC
in CO
2
and Ar atmospheres indicates that
TiO
2
-Cu
SAC
exhibits the highest HER activity and H
2
O dissociation
ability, followed by CeO
2
-Cu
SAC
and Al
2
O
3
-Cu
SAC
. The product dis-
tribution from 1.1 V to 1.5 V is displayed in Supplementary Fig. 28. At
1.1 V, the ratio of CH
4
FE to C
2
FE for Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,and
TiO
2
-Cu
SAC
is 1.08, 3.16, and 4.14 respectively, and it increases as the
potential becomes more negative (Fig. 4b). As predicted by the elec-
tronic structure, Al
2
O
3
exhibits the highest selectivity of C
2
among
these three catalysts, while CeO
2
-Cu
SAC
and TiO
2
-Cu
SAC
show selec-
tivity towards CH
4
.AnanalysisoftheCH
4
kinetic isotopic effect (KIE),
dened as the ratio of CH
4
generation rate in H
2
OandD
2
O-based
electrolytes, was performed to assess the impact of electronic struc-
tures on the proton transfer process for CO
2
RR (Fig. 4candSupple-
mentary Fig. 32). The KIE values of Al
2
O
3
-Cu
SAC
are higher than CeO
2
-
e
ab
c
fgh
d
Top view
Side view
1.96 Å 1.96 Å
1.96 Å
Fig. 3 | Characterization of Cu single-sites. a The rst derivatives of XANES
spectra. bThe corresponding k
2
-weighted Fourier transform spectra of EXAFS.
cFitting of k
2
-weighted EXAFS data of Al
2
O
3
-Cu
SAC,
CeO
2
-Cu
SAC
, and TiO
2
-Cu
SAC
in
the region of 1.02.5 Å. dThe proposed Cu
1
-O
3
conguration of Cu/metal-oxide
SAC. Atomic models: Metal,yellow, Cu, bule; and O, red. eThe Cu 2pXPS spectra.
fThe Ramanspectra of CO adsorption on Cu/metal-oxide SACs, Cu
2
O, and Cu. gIn-
situ Raman spectra of CO adsorption on Al
2
O
3
-Cu
SAC
for long-term electrolysis at
1.5 V without iR correction. hIn-situ Raman spectra of CeO
2
-Cu
SAC
for long-term
electrolysis at 1.5 V without iR correction. Relevant source data are provided as a
Source Data le.
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 6
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Cu
SAC
and TiO
2
-Cu
SAC
at all potentials and close to 2, indicating that its
rate-determining step involves H
2
O dissociation6.TiO
2
-Cu
SAC
has the
lowest KIE value, indicating that the 3d
z2
orbital signicantly accel-
erates the activation process of H
2
O, while CeO
2
-Cu
SAC
shows mod-
erate dissociation rates of H
2
O. Considering the moderate CO
adsorption energy and the localized electronic state of CeO
2
-Cu
SAC
,
CeO
2
-Cu
SAC
demonstrates the highest CH
4
FE of 70.3%, especially at
high current density of 400 mA cm2(Fig. 4d). Al
2
O
3
-Cu
SAC
and TiO
2
-
Cu
SAC
have relatively low CH
4
FE values of 46.5% and 53.7%, respec-
tively. CeO
2
-Cu
SAC
exhibits excellent performance in terms of the CH
4
FE and partial current density compared to previously reported
excellent catalysts (Supplementary Fig. 33 and Supplementary
Table 4)48.
Further testing was conducted on the CO
2
RR performance of
Al
2
O
3
-CuNPs and CeO
2
-CuNPs to validate the structure-activity rela-
tionships based on EMSI for guiding the design of supported catalysts.
Due to the promotion of C-C coupling bythe Cu electronic structure in
Al
2
O
3
-Cu
SAC
, it exhibits a higher FE
CH4
/FE
C2
. When extending Cu single-
atoms to Cu nanoparticles, more sites for CO or CHO adsorption and
coupling on the Cu surface can lead to the increased selectivity of C
2
product (Supplementary Fig. 34). The FEs of various C
2
products,
including C
2
H
4
, acetate (AcO
), and EtOH, over the Al
2
O
3
-CuNPs within
a current density range of 200 to 1000 mA cm 2, are depicted in Fig. 4e.
Specically, under the condition of constant current electrolysis, the
enhanced C
2
selectivity of Al
2
O
3
-CuNPs is observed, achieving the
highest C
2
FE of 81.3% at 900 mA cm2. In contrast, the selectivity of C
2
ab
cde
fg
CeO2-CuSAC Al2O3-CuNPs
Fig. 4 | CO
2
RR performance. a The LSV curves of Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,and
TiO
2
-Cu
SAC
with CO
2
and Ar as supplygases. bThe FE
CH4
/FE
C2
of Al
2
O
3
-Cu
SAC
,CeO
2
-
Cu
SAC
, and TiO
2
-Cu
SAC
at the potentials form 1.1 V to 1.5 V. cKIE of H/D for CO
2
RR-
to-CH
4
conversion at different potentials. dThe corresponding FEs at high current
densities. eFEs of Al
2
O
3
-CuNPs for CO
2
RR at different current densities. Long-term
stability of fCeO
2
-Cu
SAC
and gAl
2
O
3
-CuNPs at a constant current density of
400 mA cm2.agNo iR correction was applied to calculate the applied potential.
The error bars represent standard deviations from at least three independent
measurements. Relevant source data are provided as a Source Data le.
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 7
Content courtesy of Springer Nature, terms of use apply. Rights reserved
on CeO
2
-CuNPs do not signicancy increase as the sites for C-C cou-
pling increased, with the C
2
FE of 35.8% and CH
4
FE of 40.2% at
500mAcm
2(Supplementary Fig. 35). The LSV curves also reveal that
Al
2
O
3
-CuNPs displays a larger CO
2
response current density than CeO
2
-
CuNPs (Supplementary Fig. 36). The long-term durability of CeO
2
-
Cu
SAC
and Al
2
O
3
-CuNPs were evaluated by continuous CO
2
RR at
400 mA cm2, with timely replacement of the electrolyte and clean the
GDE to avoid salt deposit (Fig. 4f, g). The applied potential of CeO
2
-
Cu
SAC
maintains stability for 70 h, with the CH
4
FE consistently
remaining above 55%. Al
2
O
3
-CuNPs also shows stability for 70h with a
FE
C2
stable above 62%. Finally, SEM, XRD, XPS, and TEM measurements
were conducted on CeO
2
-Cu
SAC
and Al
2
O
3
-CuNPs catalysts after sta-
bility tests, along with ICP-MS measurement on the electrolyte (Sup-
plementary Figs. 37, 38 and Supplementary Table 5). These results
collectively conrm the stability of the catalysts during long-term
CO
2
RR tests (Supplementary Note. 2).
DFT calculations and spectroscopic insights into CO
2
RR
Afterward, the DFT calculations were further carried out to reveal the
effect of the electronic structures on the selectivity of CO
2
RR
products. First, the energy required for converting CO
2
to *CO was
calculated for these three models, followed by the formation energy of
CH
4
through deep hydrogenation of *CO (Fig. 5a and Supplementary
Tables 6, 7). Consistent with the conclusions from electronic structure
analysis, the highest occupied 3d
x2-y2
orbital reduces the reaction
energy for CO
2
activation to form *COOH on CeO
2
-Cu
SAC
(0.38 eV),
which is signicantly lower than that on Al
2
O
3
-Cu
SAC
and TiO
2
-Cu
SAC
surfaces. The *COOH intermediate is reduced to the CO* species by
reacting with a proton and releasing a H
2
O molecule. Subsequently, in
the branch of *CO hydrogenation to *CHO or *COH, the *CHO should
be the major species formed from *CO hydrogenation on these sur-
faces because the reaction energy required for *CHO formation is
lower than that for *COH (Supplementary Fig. 39). The reaction energy
for *CO hydrogenation on TiO
2
-Cu
SAC
is the lowest with 0.02 eV, fol-
lowed by CeO
2
-Cu
SAC
of 0.61 eV and Al
2
O
3
-Cu
SAC
of 0.66 eV. In the
pathway for *CHO to form CH
4
, the reaction of *CHOH with a proton
and release of H
2
O to form *CH is the rate-limiting step for CO
2
RR-to-
CH
4
.Thereactionenergiesfor*CHOHto*CHonAl
2
O
3
-Cu
SAC
and
CeO
2
-Cu
SAC
are as high as 1.31 eV and 0.90 eV, respectively. In contrast,
the reaction energy for *CH is reduced to 0.46 eV due to the active
Fig. 5 | DFT calculations and in-situ ATR-IRAS. a Free energy diagram for CO
2
RR
to form CH
4
.bFree energy diagram for *CO-*CHO coupling. In-situ ATR-IRAS spectra
of CO
2
RR over c,dAl
2
O
3
-Cu
SAC
and eCeO
2
-Cu
SAC
. In-situ ATR-IRAS of O-H stretching
vibration for fAl
2
O
3
-Cu
SAC
,gCeO
2
-Cu
SAC
,andhTiO
2
-Cu
SAC
.iFraction of free H
2
Oin
the electrode-electrolyte interface. ciNo iR correction was applied to calculate the
applied potential. Relevant source data are provided as a Source Data le.
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 8
Content courtesy of Springer Nature, terms of use apply. Rights reserved
protons on TiO
2
-Cu
SAC
. Subsequently, the energy barriers for C-C
coupling on these three surfaces were calculated, which is crucial
for the formation of C
2
products and competes with *CO deep
hydrogenation. Specically, the energy barrier for *CO-*CHO coupling
is the lowest with 0.37 eV on Al
2
O
3
-Cu
SAC
, followed by TiO
2
-Cu
SAC
of
0.61 eV and CeO
2
-Cu
SAC
of 0.80 eV (Fig. 5b and Supplementary
Fig. 40). It is noteworthy that on all three surfaces, the reaction ener-
gy for *CO dimerization is higher than that for asymmetric coupling
of *CO-*CHO (Supplementary Fig. 41), suggesting a coupling mechan-
ism on Cu SAC surfaces where *CO is hydrogenated to *CHO
and couples with surrounding CO intermediates. DFT calculations
further indicate that when the active sites on the Al
2
O
3
support switch
from Cu single-atom to Cu cluster, the consecutive Cu sites strongly
inhibit the thermodynamic generation of CH
4
while maintaining high
selectivity towards C
2
products (Supplementary Fig. 42).
The in-situ attenuated total reectance infrared absorption spec-
troscopy (ATR-IRAS) was further carried out at different potentials to
reveal the intermediate absorption manners in different reaction
pathways. The infrared (IR) bands at Al
2
O
3
-Cu
SAC
and CeO
2
-Cu
SAC
are
assigned to corresponding intermediates based on independently
reported studies (Supplementary Table 8). The IR band at 1301cm1
can be attributed to O-H deformation of *COOH, a critical intermediate
in the CO
2
to *CO conversion pathway (Fig. 5c, e)49,50. The IR bands at
1242 and 1263 cm1can be attributed to C-H stretching vibration of
*CHO and *C-H deformation in the pathway for *CO deep hydrogena-
tion to produce CH
4
5153.OnCeO
2
-Cu
SAC
, IR band (1738 cm1) attributed
to the C = O stretching vibration of *CHO can also be detected54.TheIR
bands at 1340 and 1415 cm1can be assigned to the O*CCHO inter-
mediate formed after asymmetric coupling of *CO and *CHO55.The
CHO part of O*CCHO is at 1340 cm1,andtheCOpartofO*CCHOisat
1415 cm1. The IR band at 2105 cm1on Al
2
O
3
-Cu
SAC
can be attributed to
top-adsorption of CO (Fig. 5d), while no *CO intermediate is detected
on CeO
2
-Cu
SAC
(Supplementary Fig. 43), possibly due to low *CO
concentration or rapid consumption through hydrogenation56.The
broad IRband in the range of 15501700 cm1can be attributed to H-O-
HbendofadsorbedH
2
O, with the peak on Al
2
O
3
-Cu
SAC
surface
appearing noticeably larger than that on CeO
2
-Cu
SAC
surface, con-
sistent with DFT calculations. To avoid the inuence of the broad IR
band from H
2
O and verify whether the adsorbed species on Al
2
O
3
-
Cu
SAC
surface contain hydrogen, in-situ ATR-IRAS were repeated using
D
2
O insteadof H
2
O(0.1MCO
2
-saturated KDCO
3
in D
2
O). The H
2
Oband
disappearsandisredshiftedto1188cm
1, which is consistent with the
in-plane bending vibration (D-O-D) of D
2
O (Supplementary Fig. 44)54,57.
The adsorbed carbonate band is blueshifted in the deuterated system
to 1496 cm1, which is consistent with recent reported work52.Thisis
likely due to the deuteration altering the carbonate-bicarbonate
equilibrium and the carbonate adsorption process55. The asymmetric
C = O stretching vibration of *CO
2
can be detected at 1601 cm157.
Notability, the IR bands around 1242 (*CHO), 1263 (*C-H), 1301
(*COOH), and 1340 (O*CCHO) cm1obviously shifted to 1227 (*CDO),
1257 (*C-D), 1285 (*COOD), and 1317 (O*CCDO) cm1in D
2
O, suggesting
their relation with hydrogen53.OntheAl
2
O
3
-Cu
SAC
surface, the IR bands
of *CHO and *CO both undergo a redshift with increasing potential,
which is caused by the Stark effect (Fig. 5c, d)57. This indicates that *CO
and *CHO are in-situ generated, and coupling reactions occur between
them58. Moreover, the intensity of O*CCHO is signicantly higher than
that of *C-H, indicating that the Al
2
O
3
-Cu
SAC
is more favorable for C-C
coupling reactions. In contrast, on the CeO
2
-Cu
SAC
surface, the inten-
sity of *C-H is signicancy higher than that of O*CCHO, suggesting that
*CO on the CeO
2
-Cu
SAC
surface is more conducive to hydrogenation,
leading to the generation of a large amount of CH
4
.
The state of H
2
O molecules near the electrode-electrolyte inter-
faces of Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
were then analyzed.
The broad band in 30003800 cm1represents the stretching vibra-
tion mode of the O-H bond (Fig. 5fh), and the broad band can be
divided into three peaks near 3610, 3450, and 3250 cm1through
Gaussian tting according to the distinct hydrogen bonding environ-
ments of water (Supplementary Fig. 45)6,59. They belong to water
without hydrogen-bonds (free H
2
O), with weak hydrogen-bonds
(liquid-like H
2
O), and strong hydrogen-bonds (ice-like H
2
O), respec-
tively. In general, an increased degree of hydrogen-bonds lowers the
energy of the O-H stretch, and thus the dissociation energies of H
2
O
increased in the order: free H
2
O <liquid-like H
2
O <ice-like H
2
O59,60.On
TiO
2
-Cu
SAC
, the initial proportion of free H
2
O is 98%, much higher than
CeO
2
-Cu
SAC
of 30% and Al
2
O
3
-Cu
SAC
of 20%. This suggests that the
TiO
2
-Cu
SAC
surface is enriched with free H
2
O, providing more active
protons for *CO deep hydrogenation and the HER reaction. As the
potential becomes more negative, the proportion of free H
2
Oonthe
three electrode surfaces gradually decreases, indicating the con-
sumption of free H
2
O involved in surface reactions (Fig. 5i). The con-
sumption rate of free H
2
OontheTiO
2
-Cu
SAC
surface is the fastest
(30% V1) due to the strong hydrolytic ability of TiO
2
-Cu
SAC
. However,
an excess of active hydrogen enhances competing HER, increasing the
H
2
FE of TiO
2
-Cu
SAC
. In contrast, the low abundance of free H
2
Oand
weak hydrolytic ability of Al
2
O
3
-Cu
SAC
(10.8% V1)make*COmore
prone to coupling rather than deep hydrogenation. CeO
2
-Cu
SAC
,
positioned in between Al
2
O
3
-Cu
SAC
and TiO
2
-Cu
SAC
,balances*COdeep
hydrogenation and the HER well, resulting in a CH
4
FE as high as 70.3%.
Discussion
In summary, Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
are successfully
constructed by ALD technique, and based on these catalysts, the cor-
relations between the characteristics of electronic structures and the
CO
2
RR performance are proposed through detailed characterization
and DFT calculations. The EMSI between Cu sites and Al
2
O
3
support
ranks the 3d
yz
orbital as the highest occupied orbital, which enhances
theadsorptionstrengthofCOandweakensC-Obondsthrough3d
yz
-π*
electron back-donation. This reduces the energy barrier for C-C cou-
pling, thereby promoting the formation of C
2
products on Al
2
O
3
-
Cu
SAC
. The EMSI between Cu sites and TiO
2
support ranks the 3d
z2
orbital as the highest occupied orbital, which enhances the adsorption
of H
2
OandpromotesH
2
O dissociation, thus lowering the energy
barrier for forming CH
4
. However, this over activated H
2
O, in turn,
intensies competing HER, which hinders the high-selectivity pro-
duction of CH
4
on TiO
2
-Cu
SAC
.TheEMSIbetweenCusitesandCeO
2
support ranks the 3d
x2-y2
orbital as the highest occupied orbital, which
promotes CO
2
activation and protonation. CeO
2
-Cu
SAC
effectively
balances the CO adsorption strength with the activation of H
2
O,
resulting in the highest CH
4
FE of 70.3% among the three SACs at
400 mA cm2. This structure-activity relationship based on EMSI can
provide inspiration for the design of supported catalysts. Under this
guidance, it is predicted that Cu nanoparticles loaded on Al
2
O
3
exhibit
higher C
2
selectivity than those on CeO
2
,andaC
2
FE of 81.3% at
900mAcm
2is achieved. This work not only provides an efcient
catalyst for potential industrial application, but also gives an in-depth
understanding of the CO
2
RR mechanisms that may inspire the rational
design of other catalysts for the controlled CO
2
RR.
Methods
Preparation of Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
The pure α-Al
2
O
3
,uorite CeO
2
, and rutile TiO
2
were purchased from
Shanghai Yaoyi Alloy Material Co.,Ltd. To deposit Cu single-atoms on
Al
2
O
3
,CeO
2
,andTiO
2
supports, the atomic layer deposition (ALD)
method was used. Firstly, 0.1 g of supports was evenly dispersed in a
crucible, and then the crucible was placed inside the vacuum chamber,
where the ALD took place. The vacuum chamber temperature was
maintained at 280 °C during the ALD process. Bis(hexa-uor-
oacetylacetonato) copper(II) (Cu(hfac)
2
, 99.99%, Nanjing Ai Mou Yuan
Scientic Equipment Co. Ltd.) was used as Cu precursors, and H
2
Owas
used as reducing agent. Cu(hfac)
2
was kept at 150 °C and the N
2
carrier
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 9
Content courtesy of Springer Nature, terms of use apply. Rights reserved
gas was kept at 170 °C. The ALD cycle followed a precise sequence:
introducing Cu(hfac)
2
through a solenoid valve for 50 ms and purging
the chamber with N
2
for 12 s. Atthis point, the pressure response inthe
vacuum chamber was 0.5 Pa. Thereafter, H
2
O was injected for 100 ms
and another round of N
2
cleaning was executed for 30 s. This deposi-
tion cycle was repeated 150 times. After ending the ALD cycles, the
vacuum chamber was immediately heated to 480 °C with 5 °C min1
and kept at 480 °C for 3 h to rmly anchor Cu single atoms on
supports.
Preparation of Al
2
O
3
-CuNPs and CeO
2
-CuNPs
The preparation of Al
2
O
3
-CuNPs and CeO
2
-CuNPs was generally similar
to the method used for SAC. The difference was: introducing Cu(hfac)
2
through a solenoid valve for 3 s and purging the chamber with N
2
for
12 s. At this point, the pressure response in the vacuum chamber was
10 Pa. Thereafter, H
2
O was injected for 200ms and another round of
N
2
cleaning was executed for 15 s. This deposition cycle was repeated
100 times.
Electrochemical CO
2
RR measurements
The obtained catalysts were made into electrode ink for CO
2
RR mea-
surements. 2 mg of catalystwas ultrasonically dispersed into a solution
containing 200 μL of isopropanol (99.9%, MERYER Co. Ltd.), 300 μLof
deionized water, and 20 μLofNaon (5 wt%, Alfa Aesar Co. Ltd.) to
form electrode ink. For CO
2
RR test, drop 80 μL of electrode ink onto
the gas diffusion electrode (GDE, 1.0 cm2active area) and then GDE
were heated at 60 °C for 0.5 h to remove residual isopropanol. The
activity and selectivity of the obtained catalysts were tested in a ow
cell comprising the GDE cathode (YLS-30T, Sinoro Energy Mall), anion-
exchange membrane (Fumasep FAB-PK-130, 130 μm, Sinoro Energy
Mall), Ag/AgCl electrode (3.0M KCl), and the IrO
2
@Ti mesh (0.5 mm
thickness) anode. Supply 20 standard cubic centimeter per minutes
(sccm) of CO
2
to the cathode side and 1.0M KOH (20.0mL) anolyte
stream that owed through the anode at a rate of 10.0 mL min1.A
solution of 1.0 M KOH (20.0 mL) was used as cathode electrolyte. The
CO
2
RR performance for the obtained catalysts was evaluated by
applying different currents with a current amplier in the three-
electrode system at the electrochemical workstation (CHI660E,
Chenhua) with a current amplier. The LSV curves test was conducted
at potentials from 0 V to 1.5Vvs.RHEwiththescanrateof50mVs
1.
Potentials vs. Ag/AgCl reference electrode were converted to RHE
reference scale using the following equation without iRcompensation:
ERHE =EAg=Ag Cl +0:197 + 0:0591 × pH ,ð1Þ
The gas products were quantied by gas chromatography (GC,
FULI INSTRUMENTS GC9790 Plus) equipped with thermal con-
ductivity detector(TCD) and ame ionization detector (FID) detectors.
The GC was calibrated by ve standard gases (H
2
,CO,CH
4
,C
2
H
4
and
C
2
H
6
in CO
2
) before use. The FEs of liquid products were calculated
using the total amount of the products collected at anodes and cath-
odes. The liquid product was detected by liquid 1H nuclear magnetic
resonance (NMR) spectroscopy (Bruker, AVANCE III 400 MHz Nano-
BAY), and 500 μL of electrolyzed electrolyte, 100 μLofD
2
O (99.9%,
MERYER Co. Ltd.) and 10 μLofDMSO(0.4μLmL
1, MERYER Co. Ltd.)
were mixed uniformly in the NMR tube. FE of the CO
2
RR products were
computed from:
FEð%Þ=Fne
Q×100%,ð2Þ
where F is the Faraday constant (96485 C mol1), nis the total
product (in mole), eis the number of transferred electrons for each
product, and Qis the current time integral the amount of charge
obtained.
Characterizations
The powder XRD patterns were obtained on a powder diffractometer
(Rigaku Smart Lab 3 kW) using Cu K
α
X-ray source. The SEM images of
the samples were obtained using a JSM-7800F SEM. The TEM images
were obtained with a Talos 200X G2 transmission scanning microscopy
at 200 kV. High-angle annular dark-eld-scanning transmission elec-
tron microscopy (HAADF-STEM) characterizations were carried out on
aFEITitan
3G2 60300 equipped with double aberration correctors,
which was operated at 200 kV. The X-ray absorption spectroscopy a t
the Cu K-edge was obtained at the BL14W beamlines at the Shanghai
Synchrotron Radiation Facility (SSRF) (Shanghai, China), using a Si (111)
crystal monochromator operated in transmission mode. The spectra
were processed and analyzed by the software codes Athena and Arte-
mis. XPS was performed on a Thermo Scientic ESCALAB 250Xi
instrument using Al K
α
X-ray source. The binding energy of the col-
lected spectra was calibrated by the C 1 s binding energy of 284.8 eV.
The inductively coupled plasma mass spectrometry (ICP-MS) datas
were determined using an Aglient 7850 (Ms) system. Nitrogen sorption
isotherms at 77 K were obtained on Micromeritics ASAP 2460.
In-situ Raman spectroscopy measurements
In-situ Raman spectra was recorded on laser Raman spectrometer
(LabRAM HR Evolution) with 785 nm laser (25% intensity). The in-situ
Raman spectroscopy measurements wereperformed in the Raman cell
(Tianjin Gaoss Union Technology Co. Ltd.) with a quartz optical win-
dow, an Ag/AgCl (3.0 M KCl) reference electrode and a Pt counter
electrode. Each Raman spectra was recorded for three accumulations
with an acquisition time of 20 s. The Raman cell was lled with 0.1 M
CO-saturated KHCO
3
electrolyte, and the CO ow rate was maintained
at 2 sccm during the CO adsorption measurements. The structural
evolution of the supports on Al
2
O
3
-Cu
SAC
,CeO
2
-Cu
SAC
,andTiO
2
-Cu
SAC
during the CO
2
RR process was conducted in 0.1 M CO
2
-saturated
KHCO
3
electrolyte, and the ow rate of electrolyte was sat as 2 sccm to
remove bubbles.
In-situ ATR-IRAS measurements
In situ ATR-IRAS was performed on a Nicolet iS50 spectrometer
equipped with an HgCdTe (MCT) detector and a VeeMax III (PIKE
Technologies) accessory. The measurement was conducted in an
electrochemical single-cell furnished with Pt and Ag/AgCl as counter
and reference electrodes. The cell was lled with 0.1 M CO
2
-saturated
KHCO
3
electrolyte. A xed-angle Si prism (60°) coated with catalysts
embed into the bottom of the cell served as the working electrode.
Chronoamperometry was used for CO
2
RR test and was accompanied
by the spectrum collection (32 scans, 4 cm1resolution). All spectra
were subtracted with the background.
DFT calculations
All DFT calculations were performed using Vienna ab initio simulation
package (VASP)61. The ion-electron interactions are represented by the
projector augmented wave (PAW) method and the electron exchange-
correlation by the generalized gradient approximation (GGA) with the
Perdew-Burke-Ernzerhof (PBE) exchange-correlation functional62,63.The
Kohn-Shamvalencestateswereexpandedinaplane-wavebasissetwith
a cut-off energy of 450 eV and Brillouin-zone integrations were per-
formed using a (3 × 3 × 1) Monkhorst-Pack mesh during the optimiza-
tion. The bottom three atomic layers were xed and the vacuum space
was 15 Å to avoid interactions with their periodic images in all calcula-
tions. The structure converges until all the forces on the free atoms were
less than 0.05 eV Å1. The electronic self-consistent eld cycles were set
at 105eV. The calculation was performed using generalized gradient
approximation with DFT + U and U - J = 3.0 eV for Cu 3 d;U-J=3.0eVfor
Ti 3 d;U-J=5.0eVforCe4f. The DFT + U method was employed to
calculate nal structures and energies without entropy. The chemical
potential of the proton-electron pair (H++e
) was equated with the
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 10
Content courtesy of Springer Nature, terms of use apply. Rights reserved
chemical potential of a hydrogen molecule (H
2
) at the standard hydro-
gen electrode (SHE) potential reference64. The climbing-image-nudged-
elastic-band (CI-NEB) method was applied to locate transition structures
and the prole of the potential-energy surface (PES) was constructed
accordingly65. Frequency calculations were applied to verify the adsor-
bed intermediates and the transition states (with only one imaginary
frequency). The detailed DFT calculation process and descriptions can
be found in Supplementary Figs. 46, 47 and Supplementary Note 3.
Data availability
Source data are provided with this paper and are available from the
corresponding authors upon request. Source data are provided as a
Source Data le. Source data are provided with this paper.
References
1. Nitopi, S. et al. Progress and perspectives of electrochemical CO
2
reduction on copper in aqueous electrolyte. Chem. Rev. 119,
76107672 (2019).
2. Bushuyev,O.S.etal.WhatshouldwemakewithCO
2
and how can
we make it? Joule 2,825832 (2018).
3. Wang, Z. et al. Advanced catalyst design and reactor conguration
upgrade in electrochemical carbon dioxide conversion. Adv. Mater.
35, e2303052 (2023).
4. Jouny, M., Luc, W. & Jiao, F. General techno-economic analysis of
CO
2
electrolysis systems. Ind. Eng. Chem. Res. 57,21652177 (2018).
5. Ge,W.etal.Dynamicallyformedsurfactantassemblyattheelec-
tried electrode-electrolyte interface boosting CO
2
electroreduc-
tion. J. Am. Chem. Soc. 144, 66136622 (2022).
6. Ni, W. et al. Molecular engineering of cation solvation structure for
highly selective carbon dioxide electroreduction. Angew. Chem.
Int. Ed. 62, e202303233 (2023).
7. Liu, J. et al. Switching between C
2+
products and CH
4
in CO
2
electrolysis by tuning the composition and structure of rare-earth/
copper catalysts. J. Am. Chem. Soc. 145, 2303723047 (2023).
8. Lin, Y. et al. Tunable CO
2
electroreduction to ethanol and ethylene
with controllable interfacial wettability. Nat. Commun. 14,
3575 (2023).
9. Birdja, Y. Y. et al. Advances and challenges in understanding the
electrocatalytic conversion of carbon dioxide to fuels. Nat. Energy
4,732745 (2019).
10. Zheng,M.etal.ElectrocatalyticCO
2
-to-C
2+
with ampere-level
current on heteroatom-engineered copper via tuning *CO inter-
mediate coverage. J. Am. Chem. Soc. 144,1493614944 (2022).
11. Zhang, J. et al. Accelerating electrochemical CO
2
reduction to
multi-carbon products via asymmetric intermediate binding at
conned nanointerfaces. Nat. Commun. 14, 1298 (2023).
12. Li, Y. C. et al. Binding site diversity promotes CO
2
electroreduction
to ethanol. J. Am. Chem. Soc. 141,85848591 (2019).
13. Zhong, M. et al. Accelerated discovery of CO
2
electrocatalysts
using active machine learning. Nature 581,178183 (2020).
14. Sun, Y. et al. Boosting CO
2
electroreduction to C
2
H
4
via uncon-
ventional hybridization: high-order Ce4+ 4f and O 2p interaction in
Ce-Cu
2
OforstabilizingCu
+.ACS Nano 17,1397413984 (2023).
15. Zhang, L. et al. Oxophilicity-controlled CO
2
electroreduction to C
2+
alcohols over Lewis acid metal-doped Cuδ+catalysts. J. Am. Chem.
Soc. 145,2194521954 (2023).
16. Lee, C. W. et al. Metaloxide interfaces for selective electrochemical
CC coupling reactions. ACS Energy Lett 4,22412248 (2019).
17. Belgamwar, R. et al. Defects tune the strong metal-support inter-
actions in copper supported on defected titanium dioxide catalysts
for CO
2
reduction. J. Am. Chem. Soc. 145,86348646 (2023).
18. Bruix, A. et al. A new type of strong metal-support interaction and
the production of H
2
through the transformation of water on Pt/
CeO
2
(111) and Pt/CeO
x
/TiO
2
(110) catalysts. J. Am. Chem. Soc. 134,
89688974 (2012).
19. Campbell, C. T. Catalyst-support interactions: electronic pertur-
bations. Nat. Chem. 4,597598 (2012).
20. Yang, J., Li, W., Wang, D. & Li, Y. Electronic metal-support interac-
tion of single-atom catalysts and applications in electrocatalysis.
Adv. Mater. 32, e2003300 (2020).
21. Shi,Y.etal.Electronicmetal-support interaction modulates single-
atom platinum catalysis for hydrogen evolution reaction. Nat.
Commun. 12,3021(2021).
22. Nørskov, J. K., Bligaard, T., Rossmeisl, J. & Christensen, C. H.
Towards the computational design of solid catalysts. Nat. Chem. 1,
3746 (2009).
23. Liang, X., Fu, N., Yao, S., Li, Z. & Li, Y. The progress and outlook
of metal single-atom-site catalysis. J. Am. Chem. Soc. 144,
1815518174 (2022).
24. Liu,D.,He,Q.,Ding,S.&Song,L.Structuralregulationandsupport
coupling effect of single-atom catalysts for heterogeneous cata-
lysis. Adv Energy Mater 10, 202001482 (2020).
25. Lee, B.-H. et al. Electronic interaction between transition metal
single-atoms and anatase TiO
2
boosts CO
2
photoreduction with
H
2
O. Energy Environ. Sci. 15,601609 (2022).
26. Huang, J. R. et al. Single-product faradaic efciency for electro-
catalytic of CO
2
to CO at current density larger than 1.2 A cm-2 in
neutral aqueous solution by a single-atom nanozyme. Angew.
Chem. Int. Ed. 61, e202210985 (2022).
27. Wu, Q. J. et al. Atomically precise copper nanoclusters for highly
efcient electroreduction of CO
2
towards hydrocarbons via
breaking the coordination symmetry of Cu site. Angew. Chem. Int.
Ed. 62, e202306822 (2023).
28. Wang, J. et al. Atomically dispersed metal-nitrogen-carbon cata-
lysts with d-orbital electronic conguration-dependent selectivity
for electrochemical CO
2
-to-CO reduction. ACS Catal 13,23742385
(2023).
29. Zhu, W. et al. Activating *CO by strengthening FeCO π-backbonding
to enhance two-carbon products formation toward CO
2
electro-
reductiononFeN
4
sites. Adv. Funct. Mater.34, 202402537 (2024).
30. He,C.,Lee,C.H.,Meng,L.,Chen,H.T.&Li,Z.Selectiveorbital
coupling: an adsorption mechanism in single-atom catalysis. J. Am.
Chem. Soc. 146,1239512400 (2024).
31. Wang, Q. et al. Atomically dispersed s-block magnesium sites for
electroreduction of CO
2
to CO. Angew. Chem. Int. Ed. 60,
2524125245 (2021).
32. Ding, J. et al. Atomic high-spin cobalt(II) center for highly selective
electrochemical CO reduction to CH
3
OH. Nat. Commun. 14, 6550
(2023).
33. Wang, J. et al. Direct electrochemical synthesis of acetamide from
CO
2
and N
2
on a single-atom alloy catalyst. ACS Appl. Mater.
Interfaces 15, 5343653445 (2023).
34. Ouyang, Y. et al. Selectivity of electrochemical CO
2
reduction
toward ethanol and ethylene: the key role of surface-active hydro-
gen. ACS Catal. 13,1544815456 (2023).
35. Yang, J. et al. The electronic metal-support interaction directing the
design of single atomic site catalysts: achieving high efciency
towards hydrogen evolution. Angew.Chem.Int.Ed.60,1908519091
(2021).
36. Zhou, Y. et al. Dopant-induced electron localization drives CO
2
reductiontoC
2
hydrocarbons. Nat. Chem. 10,974980 (2018).
37. Parastaev,A.etal.BoostingCO
2
hydrogenation via size-dependent
metal-support interactions in cobalt/ceria-based catalysts. Nat.
Catal. 3,526533 (2020).
38. Li, J. et al. Highly active and stable metal single-atom catalysts
achieved by strong electronic metal-support interactions. J. Am.
Chem. Soc. 141,1451514519 (2019).
39. Gao, J. et al. Solar reduction of carbon dioxide on copper-tin elec-
trocatalysts with energy conversion efciency near 20%. Nat.
Commun. 13, 5898 (2022).
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 11
Content courtesy of Springer Nature, terms of use apply. Rights reserved
40. Jiang, Y. et al. Pushing the performance limit of Cu/CeO
2
catalyst in
CO
2
electroreduction: a cluster model study for loading single
atoms. ACS Nano 17,26202628 (2023).
41. Zhao, H. et al. The role of Cu
1
-O
3
species in single-atom Cu/ZrO
2
catalyst for CO
2
hydrogenation. Nat. Catal. 5,818831 (2022).
42. Chen, S. et al. Lewis acid site-promoted single-atomic Cu catalyzes
electrochemical CO
2
methanation. Nano. Lett. 21,73257331 (2021).
43. Zhang, L. et al. Elucidating the structure-stability relationship of Cu
single-atom catalysts using operando surface-enhanced infrared
absorption spectroscopy. Nat. Commun. 14, 8311 (2023).
44. Gunathunge, C. M., Li, J., Li, X., Hong, J. J. & Waegele, M. M. Revealing
the predominant surface facets of rough Cu electrodes under elec-
trochemical conditions. ACS Catal. 10, 69086923 (2020).
45. Yang, T. et al. Catalytic structure design by AI generating with spec-
troscopic descriptors. J. Am. Chem. Soc. 145,2681726823 (2023).
46. Wang, F. et al. Active site dependent reaction mechanism over Ru/
CeO
2
catalyst toward CO
2
methanation. J. Am. Chem. Soc. 138,
62986305 (2016).
47. Ilie, A. G. et al. Principal component analysis of Raman spectra for
TiO
2
nanoparticle characterization. Appl. Surf. Sci. 417,93103 (2017).
48. Zhou, X. et al. Stabilizing Cu2+ ions by solid solutions to promote CO
2
electroreduction to methane. J. Am. Chem. Soc. 144,20792084
(2022).
49. Zhang, X. D. et al. Asymmetric low-frequency pulsed strategy
enables ultralong CO
2
reduction stability and controllable product
selectivity. J. Am. Chem. Soc. 145,21952206 (2023).
50. Firet, N. J. & Smith, W. A. Probing the reaction mechanism of CO
2
electroreduction over Ag lms via operando infrared spectroscopy.
ACS Catal. 7,606612 (2016).
51. Shao, F. et al. In situ spectroelectrochemical probing of CO redox
landscape on copper single-crystal surfaces. Proc. Natl Acad. Sci.
USA 119, e2118166119 (2022).
52. Shan, W. et al. In situ surface-enhanced Raman spectroscopic evi-
dence on the origin of selectivity in CO
2
electrocatalytic reduction.
ACS Nano 14,1136311372 (2020).
53. Yan,X.etal.SynergyofCu/C
3
N
4
interface and cu nanoparticles
dual catalytic regions in electrolysis of CO to acetic acid. Angew.
Chem. Int. Ed. 62, e202301507 (2023).
54. Kim, Y. et al. Time-resolved observation of C-C coupling inter-
mediates on Cu electrodes for selective electrochemical CO
2
reduction. Energy Environ. Sci. 13,43014311 (2020).
55. Delmo, E. P. et al. In situ infrared spectroscopic evidence of
enhanced electrochemical CO
2
reduction and C-C coupling on
oxide-derived copper. J. Am. Chem. Soc. 146,19351945 (2024).
56. Tao, Z., Pearce, A. J., Mayer, J. M. & Wang, H. Bridge sites of Au
surfaces are active for electrocatalytic CO
2
reduction. J. Am. Chem.
Soc. 144, 86418648 (2022).
57. Chernyshova, I. V., Somasundaran, P. & Ponnurangam, S. On the
origin of the elusive rst intermediate of CO
2
electroreduction.
Proc.NatlAcad.Sci.USA115, E9261E9270 (2018).
58. Liang, Y. et al. Stabilizing copper sites in coordination polymers
toward efcient electrochemical C-C coupling. Nat. Commun. 14,
474 (2023).
59. Wang, Y. et al. Strong hydrogen-bonded interfacial water inhibiting
hydrogen evolution kinetics to promote electrochemical CO
2
reductiontoC
2+
.ACS Catal. 14,34573465 (2024).
60. Wang, Y. H. et al. In situ Raman spectroscopy reveals the structure
and dissociation of interfacial water. Nature 600,8185 (2021).
61. Kressea, G. & Furthmüller, J. Efciency of ab-initio total energy
calculations for metals and semiconductors using a plane-wave
basis set. Comput.Mater.Sci.6,1550 (1996).
62. Kresse, G. & Joubert, D. From ultrasoft pseudopotentials to the pro-
jector augmented-wave method. Phys.Rev.B59,17581775 (1999).
63. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient
approximation made simple. Phys. Rev. Lett. 77, 3865 (1996).
64. Peterson, A. A., AbildPedersen,F.,Studt,F.,Rossmeisl,J.&Nørskov,
J. K. How copper catalyzes the electroreduction of carbon dioxide
into hydrocarbon fuels. Energy Environ. Sci. 3,13111315 (2010).
65. Henkelman,G.,Uberuaga,B.P.&Jónsson,H.Aclimbingimage
nudged elastic band method for nding saddle points and mini-
mum energy paths. JChem.Phys.113,99019904 (2000).
Acknowledgements
This work was supported by the following grants: National Natural Sci-
ence Foundation of China (No. 52071183, 51871122). The authors would
like to thank Meiyu Li from Shiyanjia Lab (www.shiyanjia.com)fortestsof
the XANES and EXAFS.
Author contributions
Y.C., H.L., (Nankai University), and H.L. (Tianjin University) supervised this
project. Y.Z. designed and performed most of the experiments and
analyzed the experimental data. F. C. performed the SEM measurements
and analyzed the results. Y. Z., X.Y., and Y. G. conducted the catalytic
tests. X.Z. and Z.L. performed the Raman measurements. Y. Z. and H. L.
(Tianjin University) performed the ATR-IRAS measurements. W.W., F.L.,
D.H., Y.X., and Y. Z. carried out the theoretical calculations. Y.Z., Y.C.,
and F.L. performed the TEM measurements and analyzed the results.
Y.X. and Y.Z. performed the BET and XAFS measurements. All authors
participated in the discussion of the research.
Competing interests
The authors declare no competing interests.
Additional information
Supplementary information The online version contains
supplementary material available at
https://doi.org/10.1038/s41467-025-57307-6.
Correspondence and requests for materials should be addressed to
Hui Liu, Hui Liu, Yao Xiao or Yahui Cheng.
Peer review information Nature Communications thanks the anon-
ymous reviewer(s) for their contribution to the peer review of this work. A
peer review le is available.
Reprints and permissions information is available at
http://www.nature.com/reprints
Publishers note Springer Nature remains neutral with regard to jur-
isdictional claims in published maps and institutional afliations.
Open Access This article is licensed under a Creative Commons
Attribution-NonCommercial-NoDerivatives 4.0 International License,
which permits any non-commercial use, sharing, distribution and
reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the
Creative Commons licence, and indicate if you modied the licensed
material. You do not have permission under this licence to share adapted
material derived from this article or parts of it. The images or other third
party material in this article are included in the articles Creative
Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the articles Creative Commons
licence and your intended use is not permitted by statutory regulation or
exceeds the permitted use, you will need to obtain permission directly
from the copyright holder. To view a copy of this licence, visit http://
creativecommons.org/licenses/by-nc-nd/4.0/.
© The Author(s) 2025
Article https://doi.org/10.1038/s41467-025-57307-6
Nature Communications | (2025) 16:1956 12
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
Article
Full-text available
Many non‐precious metal‐nitrogen (M–Nx)‐containing catalysts are highly efficient for electrochemical reduction of CO2 to CO and yet encounter challenges in further converting CO to more valuable two‐carbon products (C2), such as ethanol and acetic acid. The ambiguous structure‐activity relationship of the M–Nx moieties toward CO2 reduction reaction (CO2RR) results in difficulties in regulating the CO2RR product selectivity on the M–Nx‐containing catalysts. Herein, by using fluorinated iron phthalocyanines with axial‐coordinated ligands (L–FePc–F) as an M–N4‐based model electrocatalyst for CO2RR, a correlation between the electronic structure and C2 selectivity of Fe–N4 is revealed and a comprehensive descriptor based on the Fe–CO π‐backbonding is proposed for guiding the regulation of M–N4 toward higher C2 selectivity. Based on the regulation principle, the Br‐axial‐coordinated FePc–F (Br–FePc–F) remarkably increases the Faradic efficiency (FE) of C2 products from 0% (i.e., the FEC2 of FePc–F) to 34% due to the strengthened Fe–CO π‐backbonding stemming from the elevated 3dxz/dyz orbital energy and enhanced electron‐donating ability of the Fe centers of Fe–N4. This work provides a strategy and mechanism insights on the regulation of the C2 selectivity of CO2RR on the Fe–N4 moieties, which may be inspiring for precise construction and regulation of the M–Nx‐containing catalysts for specific CO2RR products.
Article
Full-text available
Understanding the structure-stability relationship of catalysts is imperative for the development of high-performance electrocatalytic devices. Herein, we utilize operando attenuated total reflectance surface-enhanced infrared absorption spectroscopy (ATR-SEIRAS) to quantitatively monitor the evolution of Cu single-atom catalysts (SACs) during the electrochemical reduction of CO2 (CO2RR). Cu SACs are converted into 2-nm Cu nanoparticles through a reconstruction process during CO2RR. The evolution rate of Cu SACs is highly dependent on the substrates of the catalysts due to the coordination difference. Density functional theory calculations demonstrate that the stability of Cu SACs is highly dependent on their formation energy, which can be manipulated by controlling the affinity between Cu sites and substrates. This work highlights the use of operando ATR-SEIRAS to achieve mechanistic understanding of structure-stability relationship for long-term applications.
Article
Full-text available
In this work, via engineering the conformation of cobalt active center in cobalt phthalocyanine molecular catalyst, the catalytic efficiency of electrochemical carbon monoxide reduction to methanol can be dramatically tuned. Based on a collection of experimental investigations and density functional theory calculations, it reveals that the electron rearrangement of the Co 3d orbitals of cobalt phthalocyanine from the low-spin state (S = 1/2) to the high-spin state (S = 3/2), induced by molecular conformation change, is responsible for the greatly enhanced CO reduction reaction performance. Operando attenuated total reflectance surface-enhanced infrared absorption spectroscopy measurements disclose accelerated hydrogenation of CORR intermediates, and kinetic isotope effect validates expedited proton-feeding rate over cobalt phthalocyanine with high-spin state. Further natural population analysis and density functional theory calculations demonstrate that the high spin Co²⁺ can enhance the electron backdonation via the dxz/dyz−2π* bond and weaken the C-O bonding in *CO, promoting hydrogenation of CORR intermediates.