Sustainable Energy & Fuels: Paper

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Sustainable

Energy & Fuels


View Article Online
PAPER View Journal

Surface structure-dependent electrocatalytic


reduction of CO2 to C1 products on SnO2 catalysts†
Published on 02 December 2019. Downloaded by University of Toronto on 1/3/2020 6:05:16 AM.

Cite this: DOI: 10.1039/c9se00678h

Minling Fang,a Zhiping Zheng,a Jiayu Chen,a Qian Chen,a Deyu Liu,c Binbin Xu,a
Jianyang Wu,a Qin Kuang *a and Zhaoxiong Xie *ab

An alternative way to mitigate the energy and environmental crisis is electrochemical CO2 reduction (ECR)
into high-value products using renewable energy. Recently, Sn-based catalysts have attracted attention
because of their high ECR selectivity to C1 products (HCOOH and CO). However, high overpotential, low
current density, and poor stability are some issues that need to be addressed for these Sn based ECR
catalysts. Resolving these problems largely depends on a comprehensive insight into the relationship
between the structure and the performance of the catalysts in the ECR. In this work, we specifically
compared the ECR activities and selectivities of three kinds of SnO2 nanocrystals (NCs) with different
dominantly exposed facets ({111}, {221} and {110}) in an effort to elucidate their surface structure–
performance relationship. We found that {111} faceted octahedral SnO2 NCs (i.e., Oct-{111} NPs) showed
not only high selectivity for C1 products (>90% FE) in a wide potential range of 0.7 V to 1.0 V but also
significant stability. The performance of Oct-{111} NPs was superior to that of the other two SnO2
morphologies, i.e. {221} faceted SnO2 octahedral NCs and {110}-dominantly faceted SnO2 rod-like NCs.
Detailed structure and composition analysis revealed that such different ECR performances of the three
Received 19th August 2019
Accepted 2nd December 2019
SnO2 facets were mainly related to the formation of different catalytic layers on their surfaces. The
findings presented here will deepen our understanding of the surface-dependent performances of Sn-
DOI: 10.1039/c9se00678h
based electrocatalysts in the ECR. This work provides a new strategy to construct high performance
rsc.li/sustainable-energy electrocatalysts through rational regulation of the metal/oxide interface.

in enhancing catalyst activity, selectivity, and stability because


Introduction of the high sensitivity of ECR to the surface structure. For
With the rapid development of industry and society, global instance, Woo et al. found that the Zn (101) facet is favourable
warming caused by the vast amount of fossil energy consump- for CO formation while the Zn (002) facet favours H2 evolution
tion and the consequent energy crisis are becoming increas- during ECR.17 Similarly, Yin and co-workers proposed that the
ingly serious.1 Converting CO2 into useful chemicals and fuels selectivity toward CH4, C2H4, C2H6, and C3H8 was higher on Cu
using renewable energy resources seems to be a promising (110) facets than on Cu (100) facets.18
solution.2 Over the past few decades, a number of catalysts Recently, low-cost and environmentally friendly Sn-based
including metals,3–5 transition-metal oxides,6–8 transition-metal catalysts have attracted attention due to their decent ECR
chalcogenides,9,10 etc. have been used for electrochemical CO2 performance in obtaining HCOOH, which is usable as a fuel in
reduction (ECR).11,12 To achieve high performances, various formic acid fuel cells.7,19,20 Various strategies to further enhance
construction strategies have been proposed. Among them, the ECR performances of Sn-based catalysts, including grain
surface engineering techniques (e.g., crystal surface effect4,13,14 boundary engineering,21 size and morphology regulation,22,23
and surface modication14–16) have proven to be highly effective and so on, have been proposed. Notably, based on theoretical
calculations, and electro-kinetic and in situ Raman spectral
analyses, researchers have found that the presence of SnOx in
a
State Key Laboratory of Physical Chemistry of Solid Surfaces, Department of Sn-based electrocatalysts plays a critical role in ECR, activating
Chemistry, College of Chemistry and Chemical Engineering, Xiamen University, CO2 to CO$ 2
2 and CO3 , which are important intermediates for
Xiamen 361005, P. R. China. E-mail: qkuang@xmu.edu.cn; zxxie@xmu.edu.cn; Fax:  24–27
generating HCOO . In spite of this progress, systematic
+86-592-2183047
b
Pen-Tung Sah Institute of Micro-Nano Science and Technology, Xiamen University,
studies are still urgently needed to reveal the surface structure–
Xiamen 361005, China performance relationships of Sn-based ECR catalysts. Espe-
c
Ningbo Institute of Materials Technology & Engineering, Chinese Academy of Sciences, cially, the inuence of exposed facets of catalysts on the
1219 Zhongguan West Road, Ningbo 315201, China formation of the Sn/SnO2 catalytic layer in the ECR process
† Electronic supplementary information (ESI) available. See DOI: cannot be ignored.
10.1039/c9se00678h

This journal is © The Royal Society of Chemistry 2019 Sustainable Energy Fuels
View Article Online

Sustainable Energy & Fuels Paper

Herein, SnO2 catalysts with well-dened surface structures and (110) surfaces. Sn atoms in the bulk of rutile SnO2 are
were used to explore the internal relationship between their sixfold-coordinated by O atoms, while Sn atoms at the surface
surface structures and ECR performances. Three kinds of SnO2 are usually coordinatively unsaturated. On the (111) and (221)
nanocrystals (NCs) with different exposed crystal surfaces were surfaces, Sn atoms are located at vefold coordinated sites
specically synthesized for this purpose. Then, the changes in (blue) with one dangling bond and fourfold-coordinated sites
the ECR performances of SnO2 catalysts with different crystal (purple) with two dangling bonds. As the most stable facets, the
surfaces were investigated. Studies revealed that Sn nano- rutile (110) surface contains rows of sixfold-coordinated Sn
particles were reduced in situ on the SnO2 NCs in the ECR atoms (grey) and vefold-coordinated Sn atoms (blue) with one
process and the status of the newly formed Sn/SnO2 interfaces dangling bond perpendicular to the surface.30 It should be
Published on 02 December 2019. Downloaded by University of Toronto on 1/3/2020 6:05:16 AM.

was the key factor affecting the ECR performance of SnO2 pointed out that {221} high-index facets can be reconsidered as
catalysts. Among the three catalysts, the octahedral SnO2 NCs a combination of (111) terraces and (110) steps. Clearly, the
(Oct-{111}) with exposed {111} facets can form more Sn/SnO2 difference in the surface structure would inuence the ECR
catalytic interfaces in the ECR process, and thus show better performances of the three morphologically distinct SnO2 NCs.
catalytic performance. Strikingly, it was observed that the ECR tests were conducted in an H-type three-electrode
Faraday efficiency (FE) for C1 products (CO and HCOOH) sur- system (Fig. S1†). Carbon paper loaded with the three
passed 90% in a wide range of potentials from 0.7 V to 1.0 V prepared morphologically distinct SnO2 NCs (1 mg cm2) acted
vs. the reversible hydrogen electrode (RHE). as a working electrode. Ag/AgCl and Pt (1  1 cm2) electrodes
were used as the reference electrode and counter electrode,
respectively. The ECR activities of the three morphologically
Results and discussion distinct SnO2 NCs were rst evaluated by linear scan voltam-
metry (LSV) in 0.5 M NaHCO3 aqueous solutions saturated with
Three kinds of SnO2 NCs with exposed {111}, {221}, and {110}
N2 and CO2, respectively. As shown in Fig. 2a, all the SnO2
facets, respectively (denoted as Oct-{111}, Oct-{221}, and Rod-
samples exhibited higher current intensities in the CO2 satu-
{110}) were synthesized based on previously reported
rated electrolyte than in the N2 saturated electrolyte, indicating
methods.28,29 Powder X-ray diffraction (XRD) analysis revealed
the catalytic activity of SnO2 NCs towards CO2 reduction. Note
that all the SnO2 samples exhibited a pure rutile phase (JCPDS
that there was a cathodic peak in the potential range of 0 V to
no. 41-1445) without the presence of any impurities (Fig. 1a).
0.4 V, which corresponds to the reduction of SnO2 (Fig. S2†).31
The morphology of these samples was studied by scanning
In the CO2-saturated electrolyte, the current densities of the
electron microscopy (SEM). Both Oct-{111} and Oct-{221} dis-
three samples at 1.0 V vs. RHE followed the order Rod-{110}
played an octahedral shape, and their particle sizes were mainly
distributed within 100–250 nm and 150–200 nm, respectively
(Fig. 1b and c). However, the tip angles of the Oct-{111} and Oct-
{221} SnO2 NCs were different, as they were actually enclosed by
different facets.19 The Rod-{110} NCs displayed a tetragonal rod-
like shape with two pyramidal tips, and the average ratio of
length to diameter was approximately 4 (Fig. 1d). According to
a previous study,28 the side surfaces of Rod-{110} NCs were
enclosed by stable {110} facets and the two pyramidal tips were
surrounded by {221} facets.
Surface scientists have demonstrated that the average
surface energy of rutile SnO2 is 2.280 J m2 for {221}, 2.209 J m2
for {111}, and 1.401 J m2 for {110}.30 Fig. 1e–g show the sche-
matic models of atomic arrangement on rutile SnO2 (111), (221),

Fig. 2 (a) LSV curves of Oct-{111}, Oct-{221} and Rod-{110} catalysts in


N2- or CO2-saturated 0.5 M NaHCO3 aqueous solution at a scan rate
Fig. 1 (a) XRD patterns and (b–d) SEM images of Oct-{111}, Oct-{221}, of 50 mV s1. (b–d) Faradaic efficiencies of H2 (black), CO (red) and
and Rod-{110}; inset: schematic model of an ideal SnO2. (e–g) Sche- HCOO (blue) over the Oct-{111}, Oct-{221} and Rod-{110} catalysts at
matic models of atomic arrangement on the (111), (221), and (110) various applied potentials, respectively. All the tests were conducted in
surfaces, respectively. CO2-saturated 0.5 M NaHCO3 aqueous solution.

Sustainable Energy Fuels This journal is © The Royal Society of Chemistry 2019
View Article Online

Paper Sustainable Energy & Fuels

(13 mA cm2) > Oct-{111} (10.8 mA cm2) > Oct-{221} (9.38 mA reaching a steady state was shorter when the applied voltage
cm2). This result indicated that the ECR activities of the three was more negative. In particular, at the reduction potential of
SnO2 samples were closely related to their exposed facets. 1.0 V vs. RHE, the ECR activity of all three samples fell
ECR tests were further conducted at selected potentials dramatically with time aer the initial surge (2 h). Furthermore,
between 0.6 and 1.0 V for 2 h in the CO2-saturated electro- we tracked the FEs of the gas products at different times by GC
lyte, to determine the selectivity of the nal products. Both to explore the intrinsic reason behind this facet dependent ECR
gaseous and liquid products were analysed using gas chroma- activity. The FE of H2 drastically decreased in the rst 1 hour for
tography (GC) and 1H nuclear magnetic resonance (1H-NMR) Oct-{111}, while the FE of CO increased and then reached
spectroscopy, respectively. The analysis results indicated that a plateau (Fig. 3d–f). In contrast to Oct-{111}, the FE of H2 over
Published on 02 December 2019. Downloaded by University of Toronto on 1/3/2020 6:05:16 AM.

H2, CO, and formate were the major products. Fig. 2b– the Oct-{221} and Rod-{110} at low potentials (0.6 V vs. RHE)
d summarize the Faraday efficiency (FE) of the products increased as the reaction progressed (Fig. S3 and S4†). Similarly,
generated over the three SnO2 samples at the selected poten- at other applied potentials (0.8 and 1.0 V vs. RHE), the time-
tials. Among these samples, the Oct-{111} showed the best dependent FE of gas products over the Oct-{221} and Rod-{110}
performance as the FE of C1 products (CO and HCOOH) was up showed different trends from that over the Oct-{111}. The above
to 90% in the potential range of 0.7 V to 1.0 V. Over the Oct- results indicated that the surface structure of the three
{111}, the hydrogen evolution reaction (HER) was remarkably morphologically distinct SnO2 samples changed signicantly
inhibited (the FE of H2 at the corresponding potentials was less during the ECR process.
than 12%). Also, FEs of H2 and CO gradually decreased, while The initial increase in currents might have been caused by
that of the formate increased with an increase in reduction the partial reduction of the surface of SnO2 catalysts under
potential. Thus it can be seen that the formation of formate was negative potentials. To conrm this the composition and
more favorable over the Oct-{111} at relatively higher potentials. morphology of all three catalysts were further characterized
Over the Oct-{221}, with an increase in applied potential, the FE aer chronoamperometry tests. As analyzed above, some
of H2 constantly decreased and the FEs of C1 products diffraction peaks assigned to metallic Sn of a tetragonal phase
remained between 60% and 70% at potentials from 0.7 V to (JCPDS no. 65-2631) were detected in the XRD patterns of the
1.0 V. In contrast to the above results, the FE of H2 over the three catalysts used (Fig. S5†). Additionally, some particles of
Rod-{110} decreased initially and then started increasing with the order of tens of nanometers were observed on the SnO2
an increase in the applied potential. catalyst surface in the SEM images (Fig. S6a and c†). Surpris-
The chronoamperometric current curves of the three SnO2 ingly, the rod-like morphology of Rod-{110} NCs was almost
samples at different potentials in CO2-saturated 0.5 M NaHCO3 completely destroyed. In contrast, the Oct-{111} and Oct-{221}
aqueous solution were analysed to evaluate the stability of the samples retained their octahedral morphology. This result was
catalysts. As shown in Fig. 3a–c, the ECR current density for the consistent with the more dramatic uctuations of the time-
SnO2 catalysts had a strong positive correlation with the applied dependent FE of the gas products in Rod-{110} as compared
potentials, consistent with the results of LSV. The current to Oct-{111} and Oct-{221}. Thus, it can be deduced that the
density of all samples increased initially with reaction time structural stability of Rod-{110} NCs in ECR was far inferior to
before attaining stability. It should be noted that the time for that of Oct-{111} and Oct-{221}.

Fig. 3 Chronoamperometric current (i–t) curves of (a) Oct-{111}, (b) Oct-{221} and (c) Rod-{110} at different potentials in CO2-saturated 0.5 M
NaHCO3 aqueous solution. Faradaic efficiency of H2 and CO for Oct-{111} at (d) 0.6, (e) 0.8, and (f) 1.0 V in CO2-saturated 0.5 M NaHCO3
aqueous solution.

This journal is © The Royal Society of Chemistry 2019 Sustainable Energy Fuels
View Article Online

Sustainable Energy & Fuels Paper

High-angle annular dark eld-scanning transmission elec-


tron microscopy (HAADF-STEM) images and energy-dispersive
X-ray (EDX) mapping images were recorded to obtain the
interface information of Sn nanoparticles formed on the SnO2
catalysts. It can be seen that the distribution of Sn was wider
than that of O in the three catalysts aer the ECR test (Fig. 4a–c).
Additionally, clear lattice fringes with 0.334 nm and 0.279 nm
spacings corresponding to SnO2 (110) and Sn (101), respectively,
were observed in the high-resolution TEM (HRTEM) images
Published on 02 December 2019. Downloaded by University of Toronto on 1/3/2020 6:05:16 AM.

taken from the surface regions of the three catalysts aer the
ECR test (Fig. 4d–f). Metallic Sn nanoparticles gradually formed
on the SnO2 NCs during the ECR.
We tracked the change in catalyst structure under different
potentials with reaction time using transmission electron
microscopy (TEM) to elucidate the relationship between the
formed Sn/SnO2 interfaces on the catalysts and their ECR
performances (Fig. 5). We focused on the Oct-{111} and Oct-
{221}, because even though their apparent morphologies are
similar, there were signicant differences in their ECR perfor-
mances (especially at low reduction potentials). The results
indicated that the construction of an effective Sn/SnO2 interface
was the key to ECR activity. At low potentials (0.6 V vs. RHE),
no signicant changes were observed in the morphology of Oct-
{111} and Oct-{221} aer 1 h (Fig. 5a). However, some Sn
nanoparticles began to form on the surface of SnO2 NCs aer Fig. 5 TEM images of Oct-{111} and Oct-{221} after ECR for different
2 h (Fig. 5b). In the subsequent reaction process, the octahedral times at potentials of (a and b) 0.6 V vs. RHE, (c and d) 0.8 V vs. RHE,
morphology of SnO2 NCs did not change signicantly, but the and (e and f) 1.0 V vs. RHE. The inset is the corresponding image at
formation of Sn nanoparticles showed different trends in Oct- a higher magnification (scale bar: 20 nm).
{111} and Oct-{221}. In the case of the Oct-{111}, the number of
Sn nanoparticles formed on the surface of the particles gradu-
ally increased, whilst the individual particles did not grow in Oct-{221} indicate that the Sn/SnO2 interface is the actual active
size. In other words, more and more Sn/SnO2 interfaces were site for ECR rather than Sn or SnO2, which may account for the
generated in the Oct-{221} as the ECR process went on. On the different FE–time variations of the products of the two catalysts.
other hand, in Oct-{221} the initially formed Sn nanoparticles This result is similar to previous reports.20,32 In fact, under the
grew larger but their number didn't increase signicantly as the potential of 0.8 V vs. RHE, the Sn/SnO2 interface could be
reaction progressed. This differences in the number and size formed more rapidly on the surfaces of Oct-{111} and Oct-{221},
evolution of the formed Sn nanoparticles between Oct-{111} and and the FEs of ECR products tend to achieve stability more
rapidly (Fig. 5c and d). When compared to Oct-{221}, Sn nano-
particles formed on the Oct-{111} had smaller size and were
more in number. This resulted in Oct-{111} exhibiting better
ECR activity. A similar relationship between the Sn/SnO2
interfaces and the FE of ECR products with time was also
observed in the ECR test under high potentials (1.0 V vs. RHE)
(Fig. 5e and f). Thus, owing to the different stabilities in the
ECR, different types of Sn/SnO2 interfaces were established on
the {111} and {221} surfaces of the SnO2 catalysts, ultimately
leading to distinct ECR activities of the SnO2 catalysts and
selectivity of the products.
In order to further conrm the role of the Sn/SnO2 interface,
formed in situ in the ECR process, in the catalyst performance,
we compared the changes in the electrochemical surface area
(ECSA) and the interfacial charge transfer resistance (Rct) of the
three catalysts before and aer the reaction. The ECSAs of the
Fig. 4 HADDF-STEM images and the corresponding elemental three catalysts were measured via double-layer capacitance tests
mapping images of (a) Oct-{111}, (b) Oct-{221} and (c) Rod-{110} after in CO2-saturated 0.5 M NaHCO3 solution. As shown in Fig. 6a,
CO2RR tests; the corresponding HRTEM images of (d) Oct-{111}, (e)
the initial ECSAs of the three morphologically distinct SnO2 NCs
Oct-{221}, and (f) Rod-{110} to show the interplanar lattice fringe
spacings. were similar. Aer a 6 hour reaction, the ECSAs of the three

Sustainable Energy Fuels This journal is © The Royal Society of Chemistry 2019
View Article Online

Paper Sustainable Energy & Fuels

that the initial O content on the surface of the Sn electrodes


correlated with the formate selectivity in terms of the faradaic
efficiency.23 The heat-treated Sn dendrite electrode (i.e., 71.6%
for formate and 11.8% for CO) shows values 24.9 and 21.5
percentage points higher than the values for the Sn foil and Sn
dendrite electrodes, respectively. From the above Sn/SnOx
studies, it can be concluded that the absence of SnOx in Sn-
based electrocatalysts would lead to a worse catalytic activity.
Fig. 6 (a) Charging current density differences plotted against scan In fact, the above studies also verify our opinion that the in situ
Published on 02 December 2019. Downloaded by University of Toronto on 1/3/2020 6:05:16 AM.

rates. (b) Nyquist plots for Oct-{111}, Oct-{221} and Rod-{110}. Z0 and formation of a Sn/SnO2 interface on the surface of SnO2 cata-
-Z00 are the real impedance and imaginary impedance, respectively. EIS lysts is the key factor for efficient ECR.
measurements for the three samples were conducted under
a potential of 0.7 V versus RHE in CO2-saturated 0.5 M NaHCO3
aqueous solution. Conclusions
In summary, by using SnO2 NCs enclosed with well-dened
facets as the research reference, we probed the intrinsic rela-
samples increased signicantly. Typically, the ECSA of the Oct-
tionship between the surface structure of SnO2 catalysts with
{111} aer the ECR test was approximately threefold larger than
different exposed facets and their catalytic performances in the
that of the initial samples. Clearly, in situ formation of the Sn
ECR. Our experimental results showed that the in situ formation
nanoparticles during the ECR process roughens the surface of
of a Sn/SnO2 interface on the surface of SnO2 catalysts is the key
the catalyst at different levels, thereby leading to an increase in
factor for efficient ECR. The ECR activity of the three SnO2
the catalytically active sites and improvement in catalytic
crystal planes followed the order {111} > {221}> {110}. Compared
activity for CO2 reduction. The order of ECSAs of the three SnO2
to the other two samples, more Sn/SnO2 catalytic interfaces were
samples aer ECR tests were Oct-{111} (slope: 1.87 mF cm2,
in situ constructed on the {111} surface in the ECR, thus
area: 31.2 cm2) > Rod-{110} (slope: 1.61 mF cm2, area: 26.8
exhibiting better ECR activity and higher selectivity over {111}-
cm2) > Oct-{221} (slope 1.47 mF cm2, area 24.5 cm2), which was
Oct for C1 products. This work reminds us that more attention
consistent with the order of the catalytic activity for CO2
should be paid to the real surface states of the catalysts in the
reduction.
catalysis process when we study the properties of catalysts in
The difference between the Sn/SnOx interfaces, formed in
ECR. This concept can be further extended to the design of
situ on the three SnO2 samples, was also reected in the
other oxide catalyst structures.
changes in the interfacial charge transfer impedances before
and aer the ECR. Fig. 6b shows the electrochemical impedance
spectroscopy responses (Nyquist plots), in which the radius of Experimental section
the semicircles represents the resistance (Rct) of interfacial Chemicals
charge transfer. Aer the ECR tests, the Rct decreased, which
Hydrated stannic chloride (SnCl4$5H2O), hydrochloric acid
indicated that the formation of the Sn/SnOx interface was
(HCl, 37.5%), poly(vinyl pyrrolidone) (PVP, K-30) and tetrame-
conducive to the electron transfer. Among the three SnO2
thylammonium hydroxide (TMAH) were purchased from Sino-
samples, Rct of the Oct-{111} decreased the most aer the
pharm Chemical Reagent Co. Ltd (Shanghai, China). All
reaction, indicating that the surface states of Oct-{111} catalysts
reagents were used directly without any further purication.
changed the most during the ECR process. This EIS result of the
Oct-{111} is consistent with the TEM images (Fig. 5). More and
smaller Sn domains are formed on the surface of the catalyst, Synthesis of octahedral SnO2 NCs enclosed by {221} facets and
which means more creation of Sn/SnOx interfaces. Therefore, rod-like SnO2 NCs dominated by {110} facets
the Oct-{111} aer the CO2RR has a smaller Rct than Rod-{110} In a typical synthesis of Oct-{221} SnO2 NCs, SnCl4$5H2O
and Oct-{221}.33 Thus it was reasonable to assume that Oct-{111} (0.350 g, 1 mmol), HCl (0.60 mL), and PVP (0.315 g, 0.006 mmol)
exhibited superior ECR performance as compared to Oct-{221} were added in order to a mixed solution of ethanol and distilled
and Rod-{110}. water (6.00 mL, 1/1 v/v) under intense ultrasonic treatment.
On the basis of the results above, we propose that the Sn/ Then the resulting solution was transferred to a Teon-lined
SnO2 interface is crucial for the ECR, and individual Sn or SnO2 stainless steel autoclave and maintained at 200  C for 12 h.
may not be the real active site in ECR. Many previous studies The products were collected by centrifugation at 4000 rpm, and
have revealed that the presence of SnOx in Sn-based electro- washed several times with deionized water and ethanol. The
catalysts can activate CO2 to CO$ 2 and CO32, which are rod-like SnO2 NCs (i.e., Rod-{110}) were synthesized in a similar
important intermediates for generating HCOO.24–27 In the way except that 0.20 mL of HCl was used instead of 0.60 mL HCl.
absence of SnOx, Sn0 catalyzes only H2 evolution because the
electron transfer to CO2 is prohibitively slow.20 Removing the Synthesis of octahedral SnO2 enclosed by {111} facets
surface oxide of the Sn-based catalysts will cause degradation of
In a typical synthesis, SnCl4$5H2O (0.350 g, 1.00 mmol) and
the ECR performance.34 Furthermore, Seong Ihl Woo has found
aqueous TMAH (17.00 mL) were successively added to ethanol

This journal is © The Royal Society of Chemistry 2019 Sustainable Energy Fuels
View Article Online

Sustainable Energy & Fuels Paper

(3.00 mL) under intense ultrasonic treatment. Then the result- versus Ag/AgCl (0.5 M NaHCO3 solution). The Nyquist plots,
ing solution was transferred to a Teon lined stainless steel showing electrochemical impedance spectroscopy (EIS)
autoclave (25 mL) and maintained at 200  C for 12 h. The responses, were measured with frequencies ranging from 100
products were collected by centrifugation at 4000 rpm, and kHz to 0.01 Hz, and the amplitude of the applied voltage was
washed several times with deionized water and ethanol. 5 mV.

Characterization of samples
Conflicts of interest
The morphology and structure of the products were character-
ized by transmission electron microscopy (TEM, JEOL JEM- There are no conicts to declare.
Published on 02 December 2019. Downloaded by University of Toronto on 1/3/2020 6:05:16 AM.

2100) at 200 kV accelerating voltage. High-resolution TEM


(HRTEM), high-angle annular dark-eld scanning TEM
(HAADF-STEM) and energy-dispersive X-ray (EDX) studies were
Acknowledgements
conducted on a TECNAI F-30 high-resolution transmission This work was supported by the National Key Research and
electron microscope at 300 kV, and scanning electron micros- Development Program of China (2017YFA0206500 and
copy (SEM, Hitachi S-4800) was performed at 10 kV accelerating 2017YFA0206801), the National Basic Research Program of
voltage. The compositions of the products were determined by China (2015CB932301), and the National Natural Science
powder X-ray diffraction (XRD, PANalytical X'Pert). Foundation of China (21671163, 21721001, 21773190, and
21931009). Jianyang Wu is grateful for the support of NFFTBS
Electrochemical measurements (No. J1310024).
All electrochemical properties were tested using a Princeton
4000+ with a three-electrode system at room temperature. The References
as-loaded catalyst carbon paper served as the working electrode,
a platinum plate (1  1 cm2) served as the counter electrode, 1 H. Yang, Z. Xu, M. Fan, R. Gupta, R. B. Slimane, A. E. Bland
and a Ag/AgCl (saturated KCl) electrode served as the reference and I. Wright, J. Environ. Sci., 2008, 20, 14–27.
electrode in 0.5 M NaHCO3 aqueous electrolyte (pH 7.2). For 2 A. Goeppert, M. Czaun, J. P. Jones, G. K. Surya Prakash and
preparation of the working electrode, 1 mg catalyst was added G. A. Olah, Chem. Soc. Rev., 2014, 43, 7995–8048.
into 200 mL ethanol and then sonicated for 30 min. 200 mL of 3 W. Zhu, R. Michalsky, O. Metin, H. Lv, S. Guo, C. J. Wright,
catalyst suspension was dropped on carbon paper with X. Sun, A. A. Peterson and S. Sun, J. Am. Chem. Soc., 2013,
a geometric area of 1  1 cm2. Linear sweep voltammetry (LSV) 135, 16833–16836.
and chronoamperometry were both performed in an H-type 4 K. Manthiram, B. J. Beberwyck and A. P. Alivisatos, J. Am.
electrolytic cell, where the cathode and anode compartments Chem. Soc., 2014, 136, 13319–13325.
were separated by Naon-115 membrane. Each half-cell 5 L. Zhang, Z. J. Zhao and J. Gong, Angew. Chem., Int. Ed., 2017,
contains 25 mL electrolyte with a gas headspace of 5 mL. 56, 11326–11353.
Before testing, the electrolyte was purged with CO2 or N2 for 6 M. Le, M. Ren, Z. Zhang, P. T. Sprunger, R. L. Kurtz and
least 30 min with a ow rate of 10 mL min1. During the J. C. Flake, J. Electrochem. Soc., 2011, 158, E45.
chronoamperometry measurements, the gas products were 7 F. Lei, W. Liu, Y. Sun, J. Xu, K. Liu, L. Liang, T. Yao, B. Pan,
analyzed using a TCD (thermal conductivity detector) for H2 and S. Wei and Y. Xie, Nat. Commun., 2016, 7, 12697.
an FID (ame ionization detector) for CO and hydrocarbons. 8 S. Gao, Z. Sun, W. Liu, X. Jiao, X. Zu, Q. Hu, Y. Sun, T. Yao,
The gas products were tested every 10 min. A gas chromato- W. Zhang, S. Wei and Y. Xie, Nat. Commun., 2017, 8, 14503.
graph (GC-2014C, Shimadzu) equipped with PLOT Mol Sieve 5A 9 F. Li, L. Chen, M. Xue, T. Williams, Y. Zhang,
and Q-bond PLOT columns was used for quantications. Ar D. R. MacFarlane and J. Zhang, Nano Energy, 2017, 31, 270–
(99.999%) was used as the carrier gas. Liquid products were 277.
analyzed using 1H NMR spectra, which were recorded on an 10 J. Deep, V. K. Sangwan, L. J. Lauhon, T. J. Marks and
Advance III 500 MHz Unity Plus spectrometer (Bruker), in which M. C. Hersam, ACS Nano, 2014, 8, 1102–1120.
0.4 mL of the electrolyte was mixed with 0.1 mL of D2O 11 J. Qiao, Y. Liu, F. Hong and J. Zhang, Chem. Soc. Rev., 2014,
(deuterated water), and 0.1 mL of dimethyl sulfoxide (DMSO) 43, 631–675.
(99.99%, Sigma) was added as an internal standard. The 12 E. E. Benson, C. P. Kubiak, A. J. Sathrum and J. M. Smieja,
Nyquist plots were measured with frequencies ranging from 100 Chem. Soc. Rev., 2009, 38, 89–99.
kHz to 0.1 Hz, and the amplitude of the applied voltage was 13 S. Liu, H. Tao, L. Zeng, Q. Liu, Z. Xu, Q. Liu and J. L. Luo, J.
1 mV. The impedance data were tted to a simplied Randles Am. Chem. Soc., 2017, 139, 2160–2163.
circuit to extract the charge-transfer resistances. The electro- 14 Y. Fang and J. C. Flake, J. Am. Chem. Soc., 2017, 139, 3399–
chemically active surface area (ECSA) was determined by 3405.
measuring the capacitive current associated with double-layer 15 T. Shinagawa, G. O. Larrazábal, A. J. Martı́n, F. Krumeich and
(Cdl) charging from the scan-rate dependence of CVs (the J. Pérez-Ramı́rez, ACS Catal., 2018, 8, 837–844.
potential window was 0.25 V to 0.35 V versus Ag/AgCl). Cdl 16 P. Wang, M. Qiao, Q. Shao, Y. Pi, X. Zhu, Y. Li and X. Huang,
was estimated by plotting the current density (Dj) at 0.3 V Nat. Commun., 2018, 9, 4933.

Sustainable Energy Fuels This journal is © The Royal Society of Chemistry 2019
View Article Online

Paper Sustainable Energy & Fuels

17 H. Won da, H. Shin, J. Koh, J. Chung, H. S. Lee, H. Kim and 26 Y. Chen and M. W. Kanan, J. Am. Chem. Soc., 2012, 134,
S. I. Woo, Angew. Chem., Int. Ed., 2016, 55, 9297–9300. 1986–1989.
18 Z. Wang, G. Yang, Z. Zhang, M. Jin and Y. Yin, ACS Nano, 27 A. Dutta, A. Kuzume, M. Rahaman, S. Vesztergom and
2016, 10, 4559–4564. P. Broekmann, ACS Catal., 2015, 5, 7498–7502.
19 Y. Zhao, J. Liang, C. Wang, J. Ma and G. G. Wallace, Adv. 28 X. Han, M. Jin, S. Xie, Q. Kuang, Z. Jiang, Y. Jiang, Z. Xie and
Energy Mater., 2018, 8, 1702524. L. Zheng, Angew. Chem., Int. Ed., 2009, 48, 9180–9183.
20 Y. Chen and M. W. Kanan, J. Am. Chem. Soc., 2012, 134, 29 X. Wang, X. Han, S. Xie, Q. Kuang, Y. Jiang, S. Zhang, X. Mu,
1986–1989. G. Chen, Z. Xie and L. Zheng, Chem.–Eur. J., 2012, 18, 2283–
21 B. Kumar, V. Atla, J. P. Brian, S. Kumari, T. Q. Nguyen, 2289.
Published on 02 December 2019. Downloaded by University of Toronto on 1/3/2020 6:05:16 AM.

M. Sunkara and J. M. Spurgeon, Angew. Chem., Int. Ed., 30 B. Slater, C. R. A. Catlow, D. H. Gay, D. E. Williams and
2017, 56, 3645–3649. V. Dusastre, J. Phys. Chem. B, 1999, 103, 10644–10650.
22 S. Zhang, P. Kang and T. J. Meyer, J. Am. Chem. Soc., 2014, 31 Q. Li, J. Fu, W. Zhu, Z. Chen, B. Shen, L. Wu, Z. Xi, T. Wang,
136, 1734–1737. G. Lu, J. J. Zhu and S. Sun, J. Am. Chem. Soc., 2017, 139, 4290–
23 H. Won da, C. H. Choi, J. Chung, M. W. Chung, E. H. Kim 4293.
and S. I. Woo, ChemSusChem, 2015, 8, 3092–3098. 32 X. Chang, T. Wang, Z. J. Zhao, P. Yang, J. Greeley, R. Mu,
24 W. Luc, C. Collins, S. Wang, H. Xin, K. He, Y. Kang and G. Zhang, Z. Gong, Z. Luo, J. Chen, Y. Cui, G. A. Ozin and
F. Jiao, J. Am. Chem. Soc., 2017, 139, 1885–1893. J. Gong, Angew. Chem., Int. Ed., 2018, 57, 15415–15419.
25 E. R. Cave, C. Shi, K. P. Kuhl, T. Hatsukade, D. N. Abram, 33 F. Lei, W. Liu, Y. Sun, J. Xu, K. Liu, L. Liang, T. Yao, B. Pan,
C. Hahn, K. Chan and T. F. Jaramillo, ACS Catal., 2018, 8, S. Wei and Y. Xie, Nat. Commun., 2016, 7, 12697.
3035–3040. 34 M. F. Baruch, J. E. Pander, J. L. White and A. B. Bocarsly, ACS
Catal., 2015, 5, 3148–3156.

This journal is © The Royal Society of Chemistry 2019 Sustainable Energy Fuels

You might also like