You are on page 1of 15

ll

Article
Boosting electrosynthesis of urea from N2 and
CO2 by defective Cu-Bi
Wenjie Wu, Yulu Yang, Yitong
Wang, ..., Qingchao Liu,
Zhaomin Hao, Shuyan Song

zmhao@henu.edu.cn (Z.H.)
songsy@ciac.ac.cn (S.S.)

Highlights
It is the first time that Cu-Bi alloy
was applied to the
electrosynthesis of urea

The mechanism of defective Cu-Bi


alloy in electrosynthesis of urea
was revealed

In situ Raman spectroscopy was


used to monitor the co-reduction
of N2 and CO2

This work switches on a light to the


application of defects in
electrosynthesis of urea

In this research, as exemplified by defective Cu-Bi, the authors focused on the


strategy that the application of defects could enable an electrosynthesis of urea
from N2 and CO2 molecules. To support the experimental and theoretical results,
atomic images, electrochemical signals, isotope labeling experiments, in situ
Raman spectroscopy, and density functional theory (DFT) calculations were
systematically involved.

Wu et al., Chem Catalysis 2, 1–14


November 17, 2022 ª 2022 Elsevier Inc.
https://doi.org/10.1016/j.checat.2022.09.012
Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll

Article
Boosting electrosynthesis of urea
from N2 and CO2 by defective Cu-Bi
Wenjie Wu,1 Yulu Yang,1 Yitong Wang,1 Tiantian Lu,1 Qingsong Dong,1 Junwei Zhao,1 Jingyang Niu,1
Qingchao Liu,2 Zhaomin Hao,1,4,* and Shuyan Song3,*

SUMMARY THE BIGGER PICTURE


Electrosynthesis of urea from N2 and CO2 molecules is of immense Converting N2 and CO2 into urea
significance but has so far been plagued by the lack of a universal via the C–N coupling reaction in
design principle to guide the search for new catalysts. Herein, an aqueous solution is a promising
as exemplified by the Cu-Bi catalyst, our findings revealed alternative to traditional energy-
that the defective Cu-Bi catalyst could offer a maximum urea con- intensive processes in the
centration of 0.45 G 0.06 mg L1 and the largest faradaic efficiency industry. However, the co-
of 8.7% G 1.7% at 0.4 V (versus reversible hydrogen electrode reduction of inert N2 and CO2
[RHE]), exhibiting outstanding activity in contrast to monometallic molecules has so far been
(Cu and Bi) and bimetallic (Cu-Bi mixture and intact Cu-Bi) catalysts. plagued by the lack of a universal
Moreover, the evidence from isotope tracing experiments, in situ design principle for rational
Raman spectroscopy, and density functional theory (DFT) calcula- activity enhancement in
tions suggested that it is a promising strategy for the co-reduction electrocatalysts. In this work, we
of N2 and CO2 by defective Cu-Bi catalysts, switching on a light to specifically focus our attention on
the application of defects in electrosynthesis of urea. the feasibility of producing urea
from N2 and CO2 molecules by
defective metals, synthesizing Cu-
INTRODUCTION Bi catalysts and control samples
Urea, subconsciously a neutral fertilizer with high nitrogen content, is of great signif- that included bimetallic and
icance to meet the demands of agricultural, industrial, and medical fields.1–3 None- monometallic catalysts, which
theless, almost all deployments of urea in industry are still inseparable from an en- enabled us to establish a
ergy-intensive process.4,5 Recent progress in synthesizing urea by electrochemical structure-to-activity correlation of
technologies has offered a promising route for artificial CO2 and N2 utilization under Cu-Bi catalyst. In combination
ambient conditions.6–13 Unfortunately, the production of urea via direct coupling of with electrochemical evidence of
N2 and CO2 remains below that required for economic viability due to the limited controlled samples, isotope
activity and poor selectivity of catalysts. Exploring more efficient and stable catalysts labeling experiments and in situ
is timely in light of the comparatively mature electrosynthesis of urea from N2 and Raman spectroscopy, our studies
CO2 molecules. pointed to a strategy in which the
rational utilization of defects is a
On the path to producing urea by N2 and CO2, a central challenge is how to substan- promising approach for the
tially address the intrinsic inertness of adsorbed molecules, which is essential for the electrosynthesis of urea by
rational design and synthesis of highly efficient electrocatalysts. In reality, several av- coupling N2 and CO2 in an
enues have previously been employed to correlate the catalytic activity of electroca- aqueous solution.
talysts with independent N2 and CO2. For example, Bi-based materials have so far
been one of the most efficient in electrocatalytic N2 reduction reaction (NRR).14–16
In parallel, Cu-based materials are one of the most promising and intensely studied
electrocatalysts to activate CO2 molecules.17–21 Noting the performance to activate
independent N2 and CO2, if we postulated a Cu-Bi alloy that combined the merits of
both Bi- and Cu-based materials, will it be effective in co-reduction of N2 and CO2?

To test the above hypothesis, we then embarked on a systematic study on the feasi-
bility of producing urea from N2 and CO2 molecules, synthesizing Cu-Bi catalysts

Chem Catalysis 2, 1–14, November 17, 2022 ª 2022 Elsevier Inc. 1


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

and some control samples that included bimetallic and monometallic catalysts,
which enabled us to establish a structure-to-activity correlation of Cu-Bi catalyst.
Unexpectedly, the study reported herein identifies a successful synthesis of defec-
tive Cu-Bi by employing characterizations from the double-corrected spherical
aberration electron microscope and electron paramagnetic resonance (EPR). The
subsequent experimental result indicated that the defective Cu-Bi catalyst offered
a maximum urea concentration of 0.45 G 0.06 mg L1 and the largest faradaic
efficiency of 8.7% G 1.7% at 0.4 V (versus reversible hydrogen electrode [RHE]),
in combination with electrochemical evidence of controlled samples, isotope label-
ing experiments, and in situ Raman spectroscopy, our studies pointed to a strategy
in which the rational utilization of defects is a promising approach for the electrosyn-
thesis of urea by coupling N2 and CO2 in an aqueous solution. Theoretical calcula-
tions further confirmed that the defective Cu-Bi catalyst could offer important
advantages over that intact Cu-Bi alloy on coupling C–N intermediates. Taken
together, the experimental and computational trends established on Cu-Bi catalysts
provide insights into intriguing relationships between defective structure and cata-
lytic activity, shedding light on the rational design of more defective electrocatalysts
for producing urea from N2 and CO2 molecules.

RESULTS AND DISCUSSION


In this work, our initial attempts were made to explore the experimental evidence in
identifying the Cu-Bi catalyst as a promising candidate for the production of urea
from N2 and CO2 molecules. To systematically figure out the structure-to-activity cor-
relation of the Cu-Bi catalyst, a series of side-by-side comparisons based on mono-
metallic (Cu and Bi) and bimetallic (Cu-Bi mixture and alloy) catalysts were designed
and synthesized using a one-pot wet-chemical approach (see experimental proced-
ures). The crystal structure of all catalysts was checked by the powder X-ray diffraction
(XRD) pattern. Specifically, the diffraction peaks of Cu-Bi are different from those of
monometallic Cu and Bi, and the peak shift in XRD has emerged in the Cu-Bi catalyst
(Figures 1A and S1), which alluded to the formation of an alloy phase.22,23 Together
with monometallic Cu and Bi, X-ray photoelectron spectroscopy (XPS) measurements
were used to compare the surface composition. The XPS measurements confirmed
the co-existence of Cu and Bi signals in the Cu-Bi catalyst. Accurately, the peaks
of the monometallic Cu are located at 931.93 (Cu2p3/2) and 951.80 eV (Cu2p1/2), while
the binding energies of Cu-Bi centered at 932.20 and 952.15 eV are ascribed to
Cu2p3/2 and Cu2p1/2, respectively. Meanwhile, the binding energies of Bi4f in Cu-Bi
are located at 164.10 (Bi4f5/2) and 158.80 eV (Bi4f7/2) with a valence state of 0, while
the peaks of the monometallic Bi are located at 156.25, 158.45 161.55, and 163.80
eV (Figure 1B). The morphologies of the Cu-Bi were checked by performing field-
emission scanning electron microscopy (SEM) and transmission electron microscopy
(TEM). Figures 1C–1F showed the elemental mapping images of the Cu-Bi catalyst, 1Henan Key Laboratory of Polyoxometalate
which confirmed the co-existence and uniform distribution of Cu and Bi elements. Chemistry, College of Chemistry and Chemical
Engineering, Henan University, Kaifeng 475004,
In addition, the lattice spacing from a high-resolution TEM (HRTEM) image and
China
well-defined diffraction rings in the selected-area electron diffraction (SAED) pattern 2Institute
of Green Catalysis, College of
in Cu-Bi is different from standard Cu or Bi, which may be attributed to a strong inter- Chemistry, Zhengzhou University, Zhengzhou
action of the metal lattice effect between Cu and Bi (Figures 1G and S2–S7).24 450001, China
3Changchun Institute of Applied Chemistry,
Chinese Academy of Sciences Changchun,
To illustrate a more elaborate and accurate structure of Cu-Bi, a double-corrected Changchun 130022, China
spherical aberration electron microscope was employed to study the surface and 4Lead contact
local atomic structure. Unexpectedly, Figure 1H indicates that, besides ‘‘intact’’ re- *Correspondence: zmhao@henu.edu.cn (Z.H.),
gions, the bimetallic Cu-Bi has ordered regions consisting of missing-atom defects, songsy@ciac.ac.cn (S.S.)
which provide clear evidence that the defective Cu-Bi has been successfully https://doi.org/10.1016/j.checat.2022.09.012

2 Chem Catalysis 2, 1–14, November 17, 2022


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

Figure 1. Characterization of defective Cu-Bi


(A) XRD pattern.
(B) Comparison of XPS patterns for Bi 4f (blue) and Cu 2p (red).
(C–F) Elemental mapping.
(G) SAED pattern.
(H) TEM image from a double-corrected spherical aberration electron microscope.
(I) Comparison of EPR spectra for intact and defective Cu-Bi.

prepared using our synthesis protocol. Since electrons or holes trapped in or around
the defect can form materials with a single electron, the EPR assay was performed to
check the vacancy on the surface of the defective Cu-Bi alloy.25 Compared with
intact Cu-Bi that was prepared through the method of electro-deposition, the EPR
spectra and higher signal intensity of g-factor (g = 2.107 G 0.002) reflected the
higher concentration of unsaturated Bi sites with unpaired electrons in defective
Cu-Bi alloy (Figures 1I and S8).26,27

Electrocatalytic activity of electrosynthesis of urea by coupling N2 and CO2 in an


aqueous solution was investigated at room temperature in a standard three-elec-
trode system. To eliminate the overestimation of catalytic activity due to the
commonly present contamination either by artificial or environmental factors, all

Chem Catalysis 2, 1–14, November 17, 2022 3


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

measurements were performed in a urea-free state (super-clean room, tight-junction


protective clothing, three-time-washing glass instruments, etc.). Besides, each
experimental result was repeated at least three times, and several control experi-
ments were designed to reduce accidental error (Figure S9). Beyond that, it may
lead to false-positive results because of the NOx contamination in our experiment.
In order to exclude this factor, some control experiments were also performed to
demonstrate that the urea in this study was derived only from the co-reduction reac-
tions of CO2 and N2 (Figures S10–S12).

Combined with the collected data from online gas chromatography, ion chromatog-
raphy, nuclear magnetic resonance hydrogen spectrum (1H NMR), and UV-visible
absorption spectra (Figures S13–S15), the urea concentrations and the calculated
faradaic efficiencies (FEs) were plotted (Figure 2A). Chronoamperometric curves of
defective Cu-Bi demonstrated the stability of the catalyst for at least a 7,200 s dura-
tion of the experiment at the range of 0.2 to 0.6 V (versus RHE) (Figure 2B and
S16–S18). For quantitative analysis of the concentration of urea, the diacetyl monox-
ime (DAMO) method was conducted by measuring the absorbance of the aqueous
solution at 525 nm (see Figures S19 and S20). Specifically, it is shown that the defec-
tive Cu-Bi delivered low urea concentration at high biases of 0.5 and 0.6 V (versus
RHE) mainly because of the competitive hydrogen evolution reaction (HER). When
the applied potential is 0.4 V (versus RHE), the defective Cu-Bi offered a maximum
urea concentration of 0.45 G 0.06 mg L1 and the largest FE of 8.7% G 1.7% (Fig-
ure 2C). Accordingly, the calculated rate of urea generation is 5.27 3 1011 mol s1
cm2, which is higher than most of the latest work.6–8 Both the concentration and FE of
urea decreased at 0.3 and 0.2 V (versus RHE), primarily from the increased domi-
nance of the CO2 reduction reaction (CO2RR) that prohibited effective N2 adsorp-
tion. For the co-reduction of CO2 and N2 in an aqueous solution, the CO2RR and
the HER are the main competitive reactions. The introduction of Bi could induce
the positive shift of the d-band center for Cu, and these could contribute to reduced
formation energies of *CH2O and *CH3O intermediates, which is the potential-
determining step of CO2RR and thus facilitates the electrocatalytic CO2 reduction
to CH4.28,29 In addition, CO is also a major by-product. The release of CO occupied
the adsorption sites of CO2 and N2, resulting in a substantial decrease in urea pro-
duction and FEs, and further induced the diversification of the distribution of reduc-
tion products. For the by-product of H2, the H2 content could increase significantly
with the increase of overpotential, which was consistent with the previously reported
literature.30

From side-by-side comparisons, convincing new evidence supported the fact that
the defective Cu-Bi could be explored as a potential electrocatalyst for efficient pro-
duction of urea from N2 and CO2. As a greater relative error may be ultimately af-
forded by a low concentration of urea, our attempts have been made to longer cat-
alytic time (10 h). When different gas atmospheres were used for producing urea by
the defective Cu-Bi, the urea could be generated only by coupling N2 and CO2
(Figures 2D, S21, and S22). Moreover, the defective Cu-Bi could be proved to be
highly active in producing urea from N2 and CO2 by comparing Cu, Bi, mixed Cu-
Bi, and intact Cu-Bi alloy (Video S1; Figures 2E and S23–S26). The electrocatalytic
activity of urea synthesis by coupling N2 and CO2 in an aqueous solution was also
comparable to the values reported for the maximum formation rates of urea, such
as BiBiVO4 (4.9 3 1011 mol s1 cm2) and BiFeO3/BiVO4 hybrids (1.39 3 1011
mol s1 cm2) (Figures 2F; Table S1),7,8 which indicated that the defective Cu-Bi
was as good as the most excellent catalysts for the production of urea from N2
and CO2 molecules.

4 Chem Catalysis 2, 1–14, November 17, 2022


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

Figure 2. Evaluation of the electrocatalytic performance of defective Cu-Bi


(A) Faradaic efficiency toward the major N 2 and CO 2 co-reduction products.
(B) The current densities and UV-vis spectrum at various biases.
(C) The concentrations and faradaic efficiencies of urea.
(D) Comparison of the urea concentrations with N2 +CO2 (blue), N 2 (green), CO 2 (red), and Ar
(yellow) as feeding gases, respectively.
(E) Comparison of the urea concentrations by defective Cu-Bi (blue), the mixture of Cu-Bi (green),
Cu (red), intact Cu-Bi (yellow), and Bi (gray) catalysts.
(F) Comparison of the urea formation rate for defective Cu-Bi with reported results.

To provide more convincing evidence for the electrosynthesis of urea from N2 and
CO2 molecules and further clarify the nature of the electrocatalytic activity, the
isotope tracing technique based on NMR and mass spectrometry (MS), as well
as the in situ electrochemical surface-enhanced Raman spectroscopy (SERS),
were employed to trace the evidence of the urea catalyzed by defective Cu-
Bi.31,32 The 1H-NMR spectrum based on a 700 M NMR instrument indicated that
a peak belonging to urea appeared at about 5.64 ppm for the gas atmosphere
of 12CO2+14N2, while the two peaks (5.54 and 5.67 ppm) from an additional qual-
itative control experiment using 15N2 isotope confirmed the activity of our catalysts
toward producing urea (Figure 3A).7,8 Besides, the comparison of standard
12
CO(14NH2)2 and 12CO(15NH2)2 with practical chemicals from 12CO2+14N2 and

Chem Catalysis 2, 1–14, November 17, 2022 5


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

Figure 3. Isotope measurements and in situ Raman signals to check the activity of defective Cu-Bi
(A) 1 HNMR spectrum of CO( 14 NH 2 ) 2 (produced by 12 CO 2 + 14 N 2 ) and CO( 15 NH 2 ) 2 (produced by
12
CO2 + 15 N 2 ).
(B) Comparison of mass spectrometry (MS) for standard CO(14NH 2 ) 2 ( 14 urea, green) and
CO(15NH 2 ) 2 ( 15 urea, yellow) with in situ-generated urea by 12 CO2 + 14 N 2 ( 14 urea, red) and
12
CO2 + 15 N 2 ( 15 urea, blue).
(C–F) In situ Raman signals for defective Cu-Bi during the electrocoupling of N 2 and CO 2 for various
biases (C), zoom-in Raman signals from 0.35 to 0.45 V (RHE) (D), and comparison of Raman
signals with different feeding gases (E) and catalysts (F).

12
CO2+15N2 further proved that the defective Cu-Bi could be effective in the direct
synthesis of urea from N2 and CO2 (Figure 3B). Moreover, in situ Raman signals for
defective Cu-Bi during the electrocoupling of N2 and CO2 were also used to detect
the microstructure information of electrode surface (interface) molecules. In this
work, no peaks have been observed when there is no bias on the defective Cu-
Bi electrode. During the negative potential excursion from 0.2 to 0.6 V (versus

6 Chem Catalysis 2, 1–14, November 17, 2022


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

RHE), the peaks around 1,540 cm1 were attributed to the signal of C–N interme-
diates, while the peak around 3,150 cm1 could be assigned to the N–H stretching
vibration (Figure 3C).33 The detailed SERS, as exemplified from 0.35 to 0.45 V
(versus RHE), suggested that the co-reduction of N2 and CO2 may present a pref-
erence for an appropriate range of bias (Figure 3D). To eliminate the overestima-
tion of the catalytic activity due to the commonly present operation error or inter-
ference of the environment, several control experiments using Ar, independent N2,
and CO2 instead of the gas source from urea were measured at 0.4 V (versus
RHE). Impressively, there was no observable Raman signal for C–N intermediates
in the range from 1,450 to 3,200 cm1, except for the defective Cu-Bi immersed
in both N2 and CO2 atmosphere (Figure 3E). Similarly, no evidence of co-reduction
N2 and CO2 was found for Raman signals when different samples including Cu, Bi,
mixed Cu-Bi, and intact Cu-Bi were checked for evidence of C–N intermediates
(Figure 3F), which is consistent with the conclusion of electrochemical measure-
ments, suggesting that the defective Cu-Bi have effective active sites to produce
urea by coupling N2 and CO2.

To gain molecular-level insight into the mechanism toward simultaneous coupling of


N2 and CO2 to produce urea by defective Cu-Bi catalyst, we calculated the density of
states (DOS), differential charge densities, and binding energies (Figure 4, S27, and
S28). According to the differences in the surface structure between the intact and
defective Cu-Bi, two representative atomic models are studied to illustrate the na-
ture of the high activity in N2 and CO2 co-reduction of defective Cu-Bi by using
the Vienna Ab initio Simulation Package (VASP) (Figure 4A). The DOS results indi-
cated that the qualitative handle on the nature of electron hybridization is different
between intact and defective Cu-Bi (Figure 4B). Differential charge density analysis,
as exemplified by *NN reaction coordinate, reveals apparent interfacial electron
transfer from bimetallic Cu-Bi to a metal-free nitrogen intermediate (Figure S29).
In addition, more metal sites are activated when the *NN is adsorbed to the surface
of the defective Cu-Bi (Figure S29). To obtain deeper insight into the differences be-
tween the intact and defective Cu-Bi, density functional theory (DFT) calculations
were further employed to assess the intermediate variation and energy barrier dur-
ing the C–N reaction. Competing with the independent electrochemical CO2 reduc-
tion in aqueous solutions—also known as the CO2RR—the electrosynthesis-urea
pathways are kinetically sluggish and need to overcome many more scientific hur-
dles because of the factors from inert N2 molecules. The competition between
*COOH (*N2) and *COOH could be one important factor in determining the catalyst
activity toward the co-reduction of N2 and CO2.9,10 Of particular interest, the defec-
tive Cu-Bi Gibbs free energy of 0.64 eV reduced CO2, while the free energy
decreased to 0.1 eV when both CO2 and N2 molecules emerged on the surface of
catalysts. For comparison, the intact Cu-Bi didn’t show a significant preference for
*COOH (*N2) and *COOH, which suggested that the defective Cu-Bi could offer
important advantages over that intact Cu-Bi on coupling C–N intermediates
(Figure 4C). Another important parameter that could be used to evaluate the cata-
lytic activity is the rate-determining step (RDS) barriers.34,35 The RDS for the distal
and alternating styles of intact Cu-Bi is from *NN+*CO to *NCON* step, amounting
to a barrier of 2.63 eV (Figures 4D and 4E). Relative lower RDS barriers are posed by
comparing the distal (from *NHCON to *NHHCON, 0.78 eV) and alternating (from
*NN+*CO to *NCON*, 0.76 eV) styles in defective Cu-Bi (Figures 4D and 4E). Theo-
retical simulations rationalize the possibility of coupling N2 and CO2 to produce urea
and demonstrate that defective Cu-Bi could pose a catalytic favor to the formation of
urea, providing theoretical support for our experimental results in the co-reduction
of N2 and CO2 molecules.

Chem Catalysis 2, 1–14, November 17, 2022 7


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

Figure 4. The theoretical calculations for intact and defective Cu-Bi


(A) Atomic models for DFT calculation.
(B) Calculated DOS.
(C) Comparison of free-energy diagram of the CO2 reduction with and without the assistance of N2 .
(D and E) Free-energy diagram of the electrolytic urea production via distal and alternating
mechanisms.

Conclusion
In summary, we have identified, through a combined experimental and theoretical
approach, that the defective Cu-Bi could be a promising electrocatalyst for produc-
ing urea by coupling N2 and CO2 molecules. The new understanding we develop
herein and the experimental implementation pave the way to further advances in
designing other defective electrocatalysts for N2 and CO2 co-reduction operation,
making it possible for the practical applications of urea production.

EXPERIMENTAL PROCEDURES
Resource availability
Lead contact
Further information and requests for resources should be directed to and will be ful-
filled by the lead contact, Zhaomin Hao (zmhao@henu.edu.cn).

Materials availability
All materials generated in this study are available from the lead contact upon
request.

8 Chem Catalysis 2, 1–14, November 17, 2022


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

Data and code availability


Data and code generated during this study are available from the lead contact upon
request.

Chemicals
Copper chloride (CuCl2 2H2O; 99.99%); ammonia solution (NH3 H2O; 30%); ammo-
nium bismuth citrate (C12H22BiN3O14; 99.99%); bismuth nitrate pentatydrate
(Bi(NO3)3$5H2O; 99.99%); dimethyl sulfoxide (DMSO; 99.9%); ethylene glycol
(C2H6O2; R99.5%); phosphoric acid (H3PO4; R85%); sulfuric acid (H2SO4; R85%);
sodium borohydride (NaBH4; 99.9%); copper nitrate (Cu(NO3)2 3H2O; 99.99%); citric
acid (C6H8O7; 99.9%); nitric acid (HNO3; 65%); diacetylmonoxime (C4H7NO2; AR);
thiosemicarbazide (CH5N3S; 99%); and potassium bicarbonate (KHCO3; R99.99%)
were purchased from Aladdin. All reagents are analytical rank and were used without
further purification.

Synthesis of Cu
The sample of Cu was synthesized following previous work.14 Firstly, a mixed solu-
tion (6 mL DMSO, 2 mL NH3 H2O, 4 mL H2O) was added into CuCl2 2H2O
(400 mg), and the mixture was placed in an ice-water bath. Next, NaBH4 (5 M,
2 mL) solution was injected rapidly into the mixed system until no bubbles formed.
The as-prepared precipitate was subsequently washed three times with water and
acetone to completely remove the unreacted precursors and other possible by-
products. Finally, the powder of Cu was gotten by drying under vacuum conditions.

Synthesis of Bi
The synthesis scheme of Bi is similar to the above Cu except that the CuCl2 2H2O was
replaced with C12H22BiN3O14 (400 mg).

Synthesis of mixed Cu-Bi


As a control experiment, the mixture of Cu-Bi was gotten by physically mixing the
above Cu and Bi at a 1:1 atomic ratio.

Synthesis of intact Cu-Bi alloy


The Cu-Bi alloy was synthesized following a similar work.36 Firstly, citric acid (0.01 M)
and HNO3 (0.5 M) were mixed to get a uniform aqueous solution. Then,
Bi(NO3)3$5H2O (0.025 M) and Cu(NO3)2 3H2O (0.025 M) were added into the above
solution. The Cu-Bi alloy was fabricated by the electrochemical method using a
three-electrode system, where the platinum was employed as the counter electrode
and the titanium as the working electrode. A single cycle of deposition consists of
constant-potential electrodeposition at 0.8 V versus Ag/AgCl for 150 s and a sub-
sequent resting step for 5 s.

Synthesis of defective Cu-Bi alloy


The defective Cu-Bi alloy was synthesized following the scheme of Cu except that
the CuCl2 2H2O was replaced with a mixture of C12H22BiN3O14 (320 mg) and
CuCl2 2H2O (80 mg).

Mass spectra measurement


The mass spectra measurement follows a similar work.6 20 mL 1,1,3,3-tetramethox-
ypropane aqueous solution (0.3 M) and 40 mL hydrochloric acid (250 g/L) were added
to the 600 mL electrolyte. Next, the mixture was placed on the shaking bed for 1 h at
room temperature, and then the above product was freeze dried. After that, 200 mL
anhydrous acetonitrile was added to the reactor, followed by the addition of 20 mL

Chem Catalysis 2, 1–14, November 17, 2022 9


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

N-methyl-N- (trimethylsilyl)trifluoroacetamide. The final product was analyzed on a


GCMS-QP2010 S spectrometer (Shimadzu) after the solution was heated for 1 h at
60 C.

In situ Raman experiments


In situ Raman spectroscopy was carried out by a Renishaw inVia Raman spectrometer
equipped with a 532 nm He-Ne laser, 350 objective, monochromator (600 grooves/
mm grating), and a charge-coupled device detector. The signal acquisition time is
90 s for each sample.

Characterization
XRD pattern was recorded on a BrukerD8 Focus Powder X-ray diffractometer using
Cu Ka (l = 0.15405 nm) radiation (40 kV, 40 mA). XPS was performed on a VG
ESCALAB MK (VK Company, UK) at room temperature by using an Al Ka X-ray source
at 12 kV and 20 mA. The morphology was observed on a JSM-7610 field-emission
SEM (FE-SEM) with 10 kV and 10 mA. TEM was performed using an FEI TECNAI
G2 S-Twin instrument with an FE gun operating at 200 kV. Inductively coupled
plasma (ICP) emission spectroscopy analysis was conducted on an Optima
2100DV instrument. Formate (HCOO) concentration was analyzed in duplicate by
ion chromatography (Thermo Fisher Scientific ICS-900). The gaseous products for
electrocatalytic reaction were analyzed on PANNA-A91Plus gas chromatography.
The 1HNMR spectra were measured on a Bruker 700M NMR spectrometer equipped
with an ultralow temperature probe. Moreover, the UV-visible absorption spectra
were recorded on a spectrophotometer (UH4150). The atomic image was filmed
by a double-corrected spherical aberration electron microscope (TEM, FEI Theims
Z). The signal of defects was measured by EPR, which was carried out by a Bruker
EMXplus-6/1 EPR spectrometer at the room temperature of 300 K. The specific pa-
rameters in the EPR measurement are as follows: center field, 3504.00 G; sweep
width, 200G; microwave power, 6.325 mW; power attenuation, 15.0 dB; sweep
time, 30.00 s; modulation amplitude, 3.00 G; modulation frequency, 100.00 kHz.
In situ Raman spectroscopy was carried out with a Renishaw inVia Raman spectrom-
eter. Mass spectra measurement was analyzed on a GCMS-QP2010 S spectrometer
(Shimadzu). Platinum electrodes and Hg/Hg2Cl2 were used as the counter and refer-
ence electrodes, respectively. As-prepared Defective Cu-Bi or control sample was
used as the working electrode with 0.1 M KHCO3 electrolyte.

The signal of defective was measured by EPR. It was carried out by a Bruker EMX-
plus-6/1 EPR spectrometer at the room temperature of 300 K. The detailed param-
eters followed: Center Field is 3504.00G, Sweep Width is 100G or 200G, Power is
6.325 mW, Power Atten is 15.0 dB, Frequency Mon is 9.832416 GHz, Sweep Time
is 30.00 s, Mod Amp is 3.000 G and Mod Freq is 100.00 kHz.

Electrochemical measurement
A Modulab-XM electrochemical workstation was used in all electrochemical urea ex-
periments at fixed potentials with Nafion-117 membrane in an H-cell arrangement.
Both catholyte and anolyte are 0.1M KHCO3 aqueous solutions. All electrochemical
measurements were based on a typical three-electrode test system, a nickel foil and
an SCE were used as the counter and reference electrodes, respectively. To prepare
the working electrode, the catalyst (2 mg) was dispersed in 450mL of ethanol
and 25 mL of Nafion (5 wt % aqueous solution) with sonication for 30min to form a
homogeneous ink. Then, all catalyst ink was loaded onto a piece of carbon paper
(1cm * 1cm, from FuelCell Store) and dried naturally.

10 Chem Catalysis 2, 1–14, November 17, 2022


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

In the electrocatalytic experiment, the catholyte was pre-saturation with the corre-
sponding gases before starting the test. The N2 was pre-purified through flowing
into alkaline solution (pH = 13 KOH aqueous solution) and acid solution
(pH = 1H2SO4 aqueous solution) in the electrocatalytic experiment. After that, the
gas flow rate was maintained at 40 sccm for electrochemical urea synthesis, which
contained 50 vol % N2 (20 sccm) and 50 vol % CO2 (20 sccm). All potentials vs
Hg/Hg2Cl2 reference electrode were converted to values vs RHE using the equation:
ERHE = ESCE + 0.241V + 0.0591 3 pH. For quantitative comparison of catalytic per-
formance, the volume of each electrolyte was set to 60 mL. The pH value was 6.8 for
the CO2-saturated electrolyte or N2- and CO2-saturated electrolyte, and pH = 8.3 for
the 0.1M KHCO3 aqueous solution saturated solely with N2, and pH = 8.0 for Ar-
saturated electrolyte in 0.1 M KHCO3.

Product quantification and identification


After continuous pouring the high purity CO2 and N2 to the 0.1M KHCO3 solution for
120 min, the quantification gas and liquid products were analyzed at each potential
at least three times during urea electrochemical synthesis. The gas-phase products
were detected by online gas chromatography (PANNA- A91Plus, China) with a ther-
mal conductivity detector (TCD) for detecting hydrogen and a flame ionization
detector (FID) for detecting carbon monoxide and other hydrocarbons. Formate
(HCOO) concentration was analyzed in duplicate by Ion Chromatography (Thermo
Fisher ICS-900). The identification and quantification of urea were achieved by NMR
spectroscopy and the DAMO method.37 For the identification of urea from electro-
chemical synthesis, the extracted electrolyte was acidized to reach the pH value of
3 by addition of an appropriate amount of 0.5M H2SO4, and a known amount of
dimethylsulfoxide-d6 was used as an internal standard. The 1HNMR spectrum was
collected on a Bruker 700 NMR spectrometer equipped with an ultralow tempera-
ture probe. The presented data are the accumulated result of 2,048 scans. For the
quantification of urea from the electrocatalytic coupling reactions, the urea concen-
tration was determined as follows: 1.25 g DAMO and 25 mg thiosemicarbazide (TSC)
were dissolved in ultrapure water and diluted to 250 mL (named as DAMO-TSC
solution). Next, 25 mL concentrated phosphoric acid was mixed with 75 mL sulfuric
acid and 150 mL ultrapure water (>18.25 MU cm), then 25 mg FeCl3 was dissolved in
the above solution (named as the acid solution). 1 mL electrolyte was removed from
the cathodic chamber. Afterward, 1 mL DAMO-TSC solution and 2 mL acid solution
were added into 1 mL electrolyte. Finally, the mixed solution was heated to 100 C
and maintained for 15 min. When the solution was cooled to room temperature,
the UV-visible (UV-vis) absorption spectrum was checked at a wavelength of
525 nm. To prepare the calibration curve, a series of urea standard solutions (0, 1,
2, 3, 4, and 5 ppm) were also checked by UV-vis absorption spectrum. The calibra-
tion curve for quantification of urea could be adopted when the linear relation
between the intensity of absorbance and urea concentration is similar (R2 > 0.99)
during the three-times independent measurement.

Calculation of FE
The FE is an important reference to electrocatalytic performance. Herein, the FE
was defined as the charge ratio between producing urea and the total charge
passed through the electrodes during the electrolysis. Assuming six electrons
were needed to form one urea molecule, the FE for urea could be calculated as
follows:

FEð%Þ = 6 3 F 3 C urea 3 V = ð60:06 3 QÞ 3 100; (Equation 1)

Chem Catalysis 2, 1–14, November 17, 2022 11


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

where F is the Faraday constant, Q is the electric quantity, C is the concentration of


generated urea, and V is the volume of the electrolyte. The FE for other gaseous
products (FE) in this work could be calculated as follows:

ngaseous products 3 Z 3 F
FEgaseous products = : (Equation 2)
Q
ngaseous products are moles of gas products, Z is the number of electrons transferred
(H2 is 2e, CO is 2e, and CH4 is 8e), F is the Faraday constant, and Q is the electric
quantity. The ngaseous products in this work could be calculated as follows:

P 3 Vgaseous products
ngaseous products = ; (Equation 3)
R3T

Vgaseous products = V 3 Cgaseous products ; and (Equation 4)

V = VMFC 3 t; (Equation 5)
where P is pressure, Vgaseous products is gaseous products volume, R is the molar gas
constant, T is temperature, Cgaseous products is gaseous products concentration (the
values obtained from gas chromatography), V is the volume of gas passed through
during the test, VMFC is the value of flowmeter, and t is time.

Theoretical section
DFT calculations in this study were carried out using the VASP code. The ion-electron
interaction was described with the projector augmented wave (PAW) method. Elec-
tron exchange correlation was represented by the functional of Perdew, Burke, and
Ernzerhof (PBE) of generalized gradient approximation (GGA).38 The energy cutoff
for the plane-wave basis set was 400 eV, and a (1 3 1 3 1) k-point sampling was
used following the Monkhorst-Pack scheme. In all the calculations, the convergence
criterion of the electronic structures was set to 105 eV, and the atomic positions
were allowed to relax until the forces were less than 0.01 eV Å1. A vacuum region
of about 20 Å was used to decouple the periodic replicas. For the differential charge
density, we followed a technique similar to that of ground-state energy calculations:
Dr = rAB  rA – rB, where rAB, rA, and rB are the charge densities of Cu-Bi*NN,
Cu-Bi, and *NN, respectively. To get an appropriate Dr, the NGX, NGY, NGZ,
NGXF, NGYF, and NGZF of rAB, rA, and rB must be consistent with each other
in the same catalyst.

SUPPLEMENTAL INFORMATION
Supplemental information can be found online at https://doi.org/10.1016/j.checat.
2022.09.012.

ACKNOWLEDGMENTS
This work was supported by financial aid from the National Key Research and Devel-
opment Program of China (2020YFE0204500). In addition, the authors are grateful
for financial aid from the National Natural Science Foundation of China (grants
21401203 and 21702045) and the Educational Administration of Henan Province
(22A150033). The authors would like to thank Zhangbin Zheng from Shiyanjia Lab
(https://www.shiyanjia.com) for the EPR analysis.

AUTHOR CONTRIBUTIONS
W.W. performed the experiments and wrote this manuscript; Y.Y. prepared the sam-
ples; Y.W. and T.L. performed of the electrochemical tests; Q.D. and J.Z. performed
in situ Raman tests; J.N. and Q.L. performed the theoretical calculations; Z.H. and

12 Chem Catalysis 2, 1–14, November 17, 2022


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

S.S. instructed the project and wrote this manuscript. All authors discussed the re-
sults and commented on the article.

DECLARATION OF INTERESTS
A China patent application based on the technology described in this work was
granted to W.W., Z.H., Y.Y., Y.W., and T.L. at Henan University (patent number:
CN114232023A).

Received: March 4, 2022


Revised: August 7, 2022
Accepted: September 13, 2022
Published: October 10, 2022

REFERENCES
1. Erisman, J.W., Sutton, M.A., Galloway, J., Nat. Commun. 12, 4080–4088. https://doi.org/ Chem. 10, 974–980. https://doi.org/10.1016/j.
Klimont, Z., and Winiwarter, W. (2008). How a 10.1038/s41467-021-24400-5. mattod. 2019.03.002.
century of ammonia synthesis changed the
world. Nat. Geosci. 1, 636–639. https://doi.org/ 10. Zhang, Z., and Guo, L. (2021). Electrochemical 18. Dinh, C.T., Burdyny, T., Kibria, M.G.,
10.1038/ngeo325. reduction of CO2 and N2 to synthesize urea on Seifitokaldani, A., Gabardo, C.M., Garcı́a de
metal–nitrogen-doped carbon catalysts: a Arquer, F.P., Kiani, A., Edwards, J.P., De Luna,
2. Liu, Y., Zhao, X., and Ye, L. (2016). A novel theoretical study. Dalton Trans. 50, 11158– P., Bushuyev, O.S., et al. (2018). CO2
elastic urea-melamine-formaldehyde foam: 11166. https://doi.org/10.1039/D1DT01390D. electroreduction to ethylene via hydroxide-
structure and properties. Ind. Eng. Chem. Res. mediated copper catalysis at an abrupt
55, 8743–8750. https://doi.org/10.1021/acs. 11. Mukherjee, J., Paul, S., Adalder, A., Kapse, S., interface. Science 360, 783–787. https://doi.
iecr.6b01957. Thapa, R., Mandal, S., Ghorai, B., Sarkar, S., and org/10.1038/s41557-018-0092-x.
Ghorai, U.K. (2022). Understanding the site-
3. Huang, H.M., McDouall, J.J.W., and Procter, selective electrocatalytic Co-reduction 19. Hoang, T.T.H., Verma, S., Ma, S., Fister, T.T.,
D.J. (2018). Radical anions from urea type mechanism for green urea synthesis using Timoshenko, J., Frenkel, A.I., Kenis, P.J.A., and
carbonyls: radical cyclizations and cyclization copper phthalocyanine nanotubes. Adv. Funct. Gewirth, A.A. (2018). Nanoporous copper-
cascades. Angew. Chem. Int. Ed. Engl. 57, Mater. 32, 2200882. https://doi.org/10.1002/ silver alloys by additive-controlled
4995–4999. https://doi.org/10.1002/anie. adfm.202200882. electrodeposition for the selective
201800667. electroreduction of CO2 to ethylene and
12. Zhu, C., Wang, M., Wen, C., Zhang, M., Geng,
ethanol. J. Am. Chem. Soc. 140, 5791–5797.
4. Barzagli, F., Mani, F., and Peruzzini, M. (2011). Y., Zhu, G., and Su, Z. (2022). Establishing the
https://doi.org/10.1021/jacs.8b01868.
From greenhouse gas to feedstock: formation principal descriptor for ElectrochemicalUrea
of ammonium carbamate from CO2 and NH3 in production via the dispersed dual-metals 20. Fan, Q., Zhang, X., Ge, X., Bai, L., He, D., Qu, Y.,
organic solvents and its catalytic conversion anchoredon the N-decorated graphene. Adv. Kong, C., Bi, J., Ding, D., Cao, Y., et al. (2021).
into urea under mild conditions. Green Chem. Sci. 9, 2105697. https://doi.org/10.1002/advs. Manipulating Cu nanoparticle surface
13, 1267–1274. https://doi.org/10.1039/ 202105697. oxidation states tunes catalytic selectivity
C0GC00674B. toward CH4 or C2+ products in CO2
13. Zhu, C., Wen, C., Wang, M., Zhang, M., Geng,
electroreduction. Adv. Energy Mater. 11,
Y., and Su, Z. (2022). Non-metal boron atoms
5. Pérez-Fortes, M., Bocin-Dumitriu, A., and 2101424. https://doi.org/10.1002/aenm.
on a CuB12 monolayer as efficient catalytic sites
Tzimas, E. (2014). CO2 utilization pathways: 202101424.
for urea production. Chem. Sci. 13, 1342–1354.
techno-economic assessment and market
https://doi.org/10.1039/d1sc04845g.
opportunities. Energy Proc. 63, 7968–7975. 21. Jeong, H.M., Kwon, Y., Won, J.H., Lum, Y.,
https://doi.org/10.1016/j.egypro.2014.11.834. 14. Wang, Y., Shi, M.M., Bao, D., Meng, F.L., Cheng, M., Kim, K.H., Head-Gordon, M., and
Zhang, Q.H., Zhou, Y.T., Liu, K.H., Zhang, Y., Kang, J.K. (2020). Atomic-Scale spacing
6. Chen, C., Zhu, X., Wen, X., Zhou, Y., Zhou, L., Li, Wang, J.Z., Chen, Z.W., et al. (2019). between copper facets for the electrochemical
H., Tao, L., Li, Q., Du, S., Liu, T., et al. (2020). Generating defect-rich bismuth for enhancing reduction of carbon dioxide. Adv. Energy
Coupling N2 and CO2 in H2O to synthesize the rate of nitrogen electroreduction to Mater. 10, 1903423. https://doi.org/10.1002/
urea under ambient conditions. Nat. Chem. 12, ammonia. Angew. Chem. Int. Ed. Engl. 58, aenm.201903 423.
717–724. https://doi.org/10.1038/s41557-020- 9464–9469. https://doi.org/10.1002/anie.
0481-9. 22. Li, Y.C., Wang, Z., Yuan, T., Nam, D.H., Luo, M.,
201903969.
Wicks, J., Chen, B., Li, J., Li, F., de Arquer,
7. Yuan, M., Chen, J., Bai, Y., Liu, Z., Zhang, J., 15. Wan, Y., Xu, J., and Lv, R. (2019). F.P.G., et al. (2019). Binding site diversity
Zhao, T., Wang, Q., Li, S., He, H., and Zhang, G. Heterogeneous electrocatalysts design for promotes CO2 electroreduction to ethanol.
(2021). Unveiling Electrochemical Urea nitrogen reduction reaction under ambient J. Am. Chem. Soc. 141, 8584–8591. https://doi.
synthesis by co-Activation of CO2 and N2 with conditions. Mater. Today 27, 69–90. https:// org/10.1021/jacs.9b02945.
Mott–Schottky heterostructure catalysts. doi.org/10.1021/acsmaterialslett.0c00611.
Angew. Chem. Int. Ed. Engl. 60, 10910–10918. 23. Chen, C., Li, Y., Yu, S., Louisia, S., Jin, J., Li, M.,
https://doi.org/10.1002/anie.202101275. 16. Yao, D., Tang, C., Li, L., Xia, B., Vasileff, A., Jin, Ross, M.B., and Yang, P. (2020). Cu-Ag tandem
H., Zhang, Y., and Qiao, S. (2020). In situ catalystsfor high-rate CO2 electrolysis towards
8. Yuan, M., Chen, J., Bai, Y., Liu, Z., Zhang, J., fragmented bismuth nanoparticles for multi-carbons. Joule 4, 1688–1699. https://doi.
Zhao, T., Shi, Q., Li, S., Wang, X., and Zhang, G. electrocatalytic nitrogen reduction. Adv. org/10.1016/j.joule.2020.07.009.
(2021). Electrochemical C–N coupling with Energy Mater. 10, 2001289. https://doi.org/10.
perovskite hybrids toward efficient urea 1002/aenm.202001289. 24. Jiang, X., Wang, X., Liu, Z., Wang, Q., Xiao, X.,
synthesis. Chem. Sci. 12, 6048–6058. https:// Pan, H., Li, M., Wang, J., Shao, Y., Peng, Z.,
doi.org/10.1039/D1SC01467F. 17. Zhou, Y., Che, F., Liu, M., Zou, C., Liang, Z., De et al. (2019). A highly selective tin-copper
Luna, P., Yuan, H., Li, J., Wang, Z., Xie, H., et al. bimetallic electrocatalyst for the
9. Zhu, X., Zhou, X., Jing, Y., and Li, Y. (2021). (2018). Dopant-induced electron localization electrochemical reduction of aqueous CO2 to
Electrochemical synthesis of urea on Mbenes. drives CO2 reduction to C2 hydrocarbons. Nat. formate. Appl. Catal. B Environ. 259,

Chem Catalysis 2, 1–14, November 17, 2022 13


Please cite this article in press as: Wu et al., Boosting electrosynthesis of urea from N2 and CO2 by defective Cu-Bi, Chem Catalysis (2022),
https://doi.org/10.1016/j.checat.2022.09.012

ll
Article

118040–118047. https://doi.org/10.1016/j. metal–organic frameworks for enhancing 34. Hao, Z., Yuan, T., Dong, Q., Singh, K., Dinic, F.,
apcatb.2019.118040. electroreduction of CO2 to CH4. Adv. Funct. Zou, Y., Niu, J., and Voznyy, O. (2021).
Mater. 32. https://doi.org/10.1002/adfm. Underappreciated role of lowenergy facets in
25. Barreca, D., Morazzoni, F., Andrea Rizzi, G., 202203677. nitrogen electroreduction. ACS Mater. Lett. 3,
Scotti, R., and Tondello, E. (2001). Molecular 327–330. https://doi.org/10.1021/
oxygen interaction with a spectroscopic and 30. Yuan, M., Zhang, H., Xu, Y., Liu, R., Wang, R., acsmaterialslett.0c00611.
Bi2O3: spectromagnetic investigation. Phys. Zhao, T., Zhang, J., Liu, Z., He, H., Yang, C.,
Chem. Chem. Phys. 3, 1743–1749. https://doi. et al. (2022). Artificial frustrated Lewis 35. Yang, Y., Wu, W., Wang, Y., Liu, J., Li, N., Fan,
org/10.1039/B009482J. pairs facilitating the electrochemical N2 and Y., Zhang, S., Dong, Q., Zhao, J., Niu, J., et al.
CO2 conversion to urea. Chem Catal. 2, (2022). Enhanced electrochemical O2-to-
26. Schneider, C., Mendt, M., Pöppl, A., Crocellà, 309–320. https://doi.org/10.1016/j.checat. H2O2 synthesis via Cu-Pb synergistic interplay.
V., and Fischer, R.A. (2020). Scrutinizing the 2021.11.009. Small 18, 2106534. https://doi.org/10.1002/
pore chemistry and the importance of Cu(I) smll.202106534.
defects in TCNQ-Loaded Cu3(BTC)2 by a 31. Wei, D.Y., Yue, M.F., Qin, S.N., Zhang, S., Wu,
multitechnique spectroscopic approach. ACS Y.F., Xu, G.Y., Zhang, H., Tian, Z.Q., and Li, J.F. 36. Koh, J.H., Won, D.H., Eom, T., Kim, N.K., Jung,
Appl. Mater. Interfaces 12, 1024–1035. https:// (2021). In situ Raman observation of oxygen K.D., Kim, H., Hwang, Y.J., and Min, B.K. (2017).
doi.org/10.1021/acsami.9b16663. activation and reaction at platinum–ceria CO2 electro-reduction to formate via
27. Zhou, H., Wang, M., and Wang, F. (2021). interfaces during CO Oxidation. J. Am. Chem. oxygen bidentate intermediate
Oxygen vacancy mediated catalytic Soc. 143, 15635–15643. https://doi.org/10. stabilized by high-index planes of Bi dendrite
methanation of lignocellulose at temperatures 1021/jacs.1c04590. catalyst. ACS Catal. 7, 5071–5077. https://doi.
below 200  C. Joule 5, 3031–3044. https://doi. org/10.1021/acscatal.7b00707.
org/10.1016/j.joule.2021.07.001. 32. Radjenovic, P.M., Zhou, R.Y., Dong, J.C., and Li,
J.F. (2021). Watching reactions at solidliquid 37. Rahmatullah, M., and Boyde, T.R.C. (1980).
28. Wang, Y., Cheng, L., Zhu, Y., Liu, J., Xiao, C., interfaces with in situ Raman spectroscopy. Improvements in the determination of urea
Chen, R., Zhang, L., Li, Y., and Li, C. (2022). J. Phys. Chem. C 125, 26285–26295. https://doi. using diacetyl monoxime; methods with and
Tunable selectivity on copper-bismuth org/10.1021/acs.jpcc.1c08640. without deproteinization. Clin. Chim. Acta 107,
bimetallic aerogels for electrochemical 3–9. https://doi.org/10.1016/0009-8981(80)
CO2 reduction. Appl. Catal. B Environ. 317, 33. Zhao, Y., Li, F., Li, W., Li, Y., Liu, C., Zhao, Z., 90407-6.
121650. https://doi.org/10.1016/j.apcatb.2022. Shan, Y., Ji, Y., and Sun, L. (2021). Identification
121650. of M-NH2-NH2 intermediate and rate 38. Hafner, J. (2008). Ab-InitioSimulations
determining step for nitrogen reduction with of materials using VASP: density-
29. Zhang, Y., Zhou, Q., Qiu, Z., Zhang, X., Chen, J., bioinspired sulfur-bonded FeW catalyst. functional theory and beyond. J. Comput.
Zhao, Y., Gong, F., and Sun, W. (2022). Tailoring Angew. Chem. Int. Ed. Engl. 60, 20331–20341. Chem. 29, 2044–2078. https://doi.org/10.1002/
coordination microenvironment of Cu(I) in https://doi.org/10.1002/anie.202104918. jcc.21057.

14 Chem Catalysis 2, 1–14, November 17, 2022

You might also like