Вы находитесь на странице: 1из 12

Nano Research

https://doi.org/10.1007/s12274-017-1943-2

Ni-doped ZnCo2O4 atomic layers to boost the selectivity


in solar-driven reduction of CO2
Katong Liu, Xiaodong Li, Liang Liang, Ju Wu, Xingchen Jiao, Jiaqi Xu, Yongfu Sun (), and Yi Xie ()

Hefei National Laboratory for Physical Sciences at Microscale, CAS Center for Excellence in Nanoscience, University of Science and
Technology of China, Hefei 230026, China

Received: 31 October 2017 ABSTRACT


Revised: 20 November 2017 Regulating the selectivity of CO2 photoreduction is particularly challenging.
Accepted: 29 November 2017 Herein, we propose ideal models of atomic layers with/without element doping
to investigate the effect of doping engineering to tune the selectivity of CO2
© Tsinghua University Press photoreduction. Prototypical ZnCo2O4 atomic layers with/without Ni-doping
and Springer-Verlag GmbH were first synthesized. Density functional theory calculations reveal that introducing
Germany, part of Springer Ni atoms creates several new energy levels and increases the density-of-states at
Nature 2017 the conduction band minimum. Synchrotron radiation photoemission spectroscopy
demonstrates that the band structures are suitable for CO2 photoreduction,
KEYWORDS while the surface photovoltage spectra demonstrate that Ni doping increases the
atomic layers, carrier separation efficiency. In situ diffuse reflectance Fourier transform
Ni-doped, infrared spectra disclose that the CO2− radical is the main intermediate, while
zinc cobaltite, temperature-programed desorption curves reveal that the ZnCo2O4 atomic
selectivity of CO2 reduction layers with/without Ni doping favor the respective CO and CH4 desorption.
The Ni-doped ZnCo2O4 atomic layers exhibit a 3.5-time higher CO selectivity
than the ZnCo2O4 atomic layers. This work establishes a clear correlation between
elemental doping and selectivity regulation for CO2 photoreduction, opening
new possibilities for tailoring solar-driven photocatalytic behaviors.

1 Introduction the concentration of CO2 in the atmosphere while


simultaneously solving the energy and environmental
Carbon dioxide emissions have substantially con- problems. In this regard, solar-driven reduction of
tributed to climate change and global warming, which CO2 is a valuable method that has used diverse
have destroyed marine ecosystems and negatively semiconductors, including ZnO, TiO2, BiWO4, and
impacted human health [1–4]. Thus, it is urgent to ZnIn2S4, as photocatalysts [3, 5–10]. To date, no effective
convert CO2 into an energy-bearing carbon form via a methods have been developed to tune the product
sustainable approach i.e., one that does not increase selectivity of CO2 photoreduction; the various carbon-

Address correspondence to Yongfu Sun, yfsun@ustc.edu.cn; Yi Xie, yxie@ustc.edu.cn


2 Nano Res.

based products are difficult to separate and hence photovoltage spectra (SPV) showed that the carrier
impede practical applications. Doping engineering has separation efficiency of the Ni-doped ZnCo2O4 atomic
been widely utilized to tailor the electronic structure layers was higher than that of the ZnCo2O4 atomic
of semiconductors [11–14] and hence change the layers. Hence, in situ diffuse reflectance infrared Fourier
behavior of CO2 reduction. Significant progress has transform spectroscopy (DRIFTS) was performed to
been made in the field of water splitting using element detect the intermediates during the reduction process,
doping engineering; CO2 is catalytically reduced in where the CO2− radical was the main intermediate
H2O, and the element doping engineering is applied during the CO2 photoreduction for both samples. In
to suppress the generation of H2. For instance, Sato addition, temperature-programed desorption (TPD)
and colleagues reported a nitrogen-doped Ta2O5 measurements were used to investigate the product
semiconductor as an effective photocatalyst for CO2 desorption processes; CO preferentially desorbed from
photoreduction; it had a higher selectivity toward the Ni-doped ZnCo2O4 atomic layers, and CH4 was
HCOOH evolution during the photocatalytic conversion prone to desorb from the ZnCo2O4 atomic layers. The
with a selectivity for HCOOH > 75% [15]. However, real-time product monitoring system demonstrated
atomic-level insights into the role of doping engineering that the Ni-doped ZnCo2O4 atomic layers contained
for tailoring product selectivity of CO2 photoreduction higher proportions of CO in the products, suggesting
are still not clear because of abundant micro- and that the Ni-doped ZnCo2O4 atomic layers boosted
nanostructures on the surface of the reported the selectivity for solar-driven reduction of CO2. Based
photocatalysts [16]. on the characterizations and photoreduction perfor-
Herein, we begin by building an ideal model of mances, the as-synthesized ZnCo2O4 atomic layers
atomically thin, two-dimensional layers; the atomic with Ni-doping successfully changed the selectivity
thickness enables an ultrahigh fraction of low- toward CO and CH4 evolution. This work provides
coordinated surface atoms to serve as highly catalytically valuable insights into the role of element doping
active sites. We deliberately doped transition metal engineering for selectively regulating CO2 photore-
atoms into the atomic layers to tailor their electronic duction, showing great promise for designing catalysts
structure and regulate the product selectivity during with highly selective performances.
CO2 photoreduction to hydrocarbon fuels. Taking the
nontoxic semiconductor of cubic-ZnCo2O4 (its crystal
2 Experimental section
structure is shown in Fig. S1(a) in the Electronic
Supplementary Material (ESM)) as an example, 2.1 Synthesis of ZnCo2O4 atomic layers
2.5-nm-thick ZnCo2O4 atomic layers with/without
Ni-doping were successfully fabricated. Density In a typical procedure [17, 18], 300 mg of Co(NO3)2·6H2O,
functional theory (DFT) calculations disclosed that Ni 150 mg of Zn(NO3)2·6H2O, and 100 mg of KNO3 were
doping confined in the ZnCo2O4 atomic layers increased added (in that order) to a 5-mL transparent solution
the density of states (DOS) at the conduction band containing 100 mg of hexadecyl trimethyl ammonium
minimum (CBM) with respect to the perfect ZnCo2O4 bromide (CTAB) dissolved in deionized water under
atomic layers, thus favoring the transport and constant magnetic stirring. The as-prepared solution
separation process of the photoexcited charge carrier. was transferred to a 40-mL autoclave after 10 min of
Moreover, the band structures for the Ni-doped stirring, and 30 mL of ethylene glycol was added to
ZnCo2O4 atomic layers and the ZnCo2O4 atomic the autoclave followed by another 10 min of stirring.
layers were measured using synchrotron radiation The system was maintained at 140 °C for 36 h. After
photoemission spectroscopy (SRPES) and ultraviolet naturally cooling to room temperature, the obtained
visible absorption spectroscopy (UV–vis). The results precipitate was washed several times with water and
implied that both the ZnCo2O4 atomic layers with/ ethanol to obtain the amaranthine product, and then
without Ni-doping possessed appropriate band dried using vacuum freeze-drying equipment for
structures for CO2 reduction to CO and CH4. Surface further characterization.

| www.editorialmanager.com/nare/default.asp
Nano Res. 3

2.2 Synthesis of Ni-doped ZnCo2O4 atomic layers A surface area and porosity analyzer (Micromeritics
ASAP 2020 M PLUS HD88) was used to measure the
In a typical procedure, 300 mg of Co(NO3)2·6H2O, Brunauer–Emmett–Teller (BET) isothermal adsorption,
150 mg of Zn(NO3)2·6H2O, and 30 mg of Ni(NO3)2·6H2O Barrett–Joyner–Halenda pore-size distribution, and
were added (in that order) to a 5-mL transparent CO2 isothermal adsorption. A chemisorption analyzer
solution containing 100 mg of CTAB dissolved in (Micromeritics Autochem 2920) was used to conduct
deionized water under constant magnetic stirring. CO and CH4 TPD measurements. The synchrotron
The as-prepared solution was transferred to a 40-mL radiation photoemission spectroscopy spectra were
autoclave after 10 min of stirring, and 30 mL of measured using a photon energy of 40 eV at the
ethylene glycol was added to the autoclave followed Catalysis and Surface Science Endstation (BL11U) at
by another 10 min of stirring. The system was the University of Science and Technology of China. The
maintained at 140 °C for 36 h. After naturally cooling work function (Φ) was calculated from Φ = hν − Ecutoff
to room temperature, the obtained precipitate was (hν is the incident photon energy of 40 eV), while a
washed several times with water and ethanol to obtain bias of −5 V was used to gain the secondary electron
the violet product, which was then dried using vacuum cutoff (Ecutoff).
freeze-drying before further characterization.
2.4 Surface photovoltage measurements
2.3 Characterization
SPV spectra were obtained using a lock-in amplifier.
Field-emission scanning electron microscopy images The measurement systems include a lock-in amplifier
were obtained using an FEI Sirion-200 scanning electron (SR830, Stanford Research Systems, Inc.), mono-
microscope. Transmission electron microscopy (TEM) chromatic light, light chopper (SR540, Stanford Research
images and high-resolution TEM images were obtained Systems, Inc.), and electroconductive chamber. Mono-
using a JEOL-2010 transmission electron microscope chromatic light was generated from a 500-W xenon
with an acceleration voltage of 200 kV. X-ray diffraction lamp (CHF-XM-500 W, global xenon lamp power)
(XRD) patterns were recorded using a Philips X’Pert using a monochromator (Omni-3007, No.16047, Zolix
Pro Super diffractometer with Cu Kα radiation (λ = Instruments Co.).
1.54178 Å). Atomic force microscopy (AFM) was
performed using a Veeco DI Nano-scope MultiMode V 2.5 Photocatalytic CO2 reduction measurements
system. X-ray photoelectron spectra (XPS) were Photocatalytic CO2 reduction measurements were
acquired on an ESCALAB MKII using Mg Kα (hν = conducted using a Lab SolarIII AG system (Perfectlight
1,253.6 eV) as the excitation source. The binding Limited, Beijing). The temperature of the reaction cell
energies obtained in the XPS spectral analysis were was maintained at 25 ± 0.2 °C by recirculating cooling
corrected for specimen charging by referencing C 1s water during irradiation. In the typical CO2 photo-
to 284.8 eV. Inductively coupled plasma atomic catalytic reduction process, 20 mg of the as-synthesized
emission spectrometry (ICP-AES) was performed photocatalyst (ZnCo2O4 atomic layers with/without
using an Atomscan Advantage (Thermo Jarrell Ash Ni doping) was homogeneously dispersed onto a
Corporation (USA)). Room-temperature UV–vis diffuse quartz plate in the reaction cell containing 2 mL of
reflectance spectroscopy (DRS) and UV–vis absorption deionized water. A PLS-SXE300/300UV xenon lamp
spectra were recorded on a Perkin Elmer Lambda (Trusttech Co., Ltd. Beijing) with a standard AM 1.5
950 UV-vis-near infrared spectrophotometer. Fourier filter provided light irradiation with an output light
transform infrared (FT-IR) spectra were acquired on a density of about 100 mW·cm−2. The instrument was
Nicolet FT-IR spectrometer (Thermo Fisher Scientific) initially vacuum-treated three times and then pumped
using KBr tablets, scanning from 4,000 to 400 cm−1 at using high-purity CO2 (99.99%) to reach atmospheric
room temperature. In situ DRIFTS was performed pressure. During light irradiation, the gas products
using the Nicolet FT-IR spectrometer with a gas were qualitatively analyzed using a Techcomp GC7900
path system and a reaction chamber (HVC-DRM-5). gas chromatograph (flame ionization detector, TDX-01

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


4 Nano Res.

column) by identifying the chromatographic peaks. A 3 Results and discussion


sampling interval of 1 h was set using an automatic
sampler with a single cycle time of 21 h. 3.1 Characterizations of Ni-doped ZnCo2O4 atomic
layers
2.6 Density functional theory calculations
ZnCo2O4 atomic layers and Ni-doped ZnCo2O4 atomic
The first-principles calculations were performed using layers were first successfully synthesized using a
the Vienna ab initio simulation package [19]. The low-temperature and straightforward solvothermal
interaction between the ions and valence electrons method. For example, the full characterizations of
was described using projector augmented wave the Ni-doped ZnCo2O4 atomic layers were shown as
potentials, and the exchange-correlation between follows. The XRD patterns were readily assigned
electrons was treated using the generalized gradient to cubic ZnCo2O4 (JCPDS 23-1390; Figs. S1–S3 in the
approximation in the Perdew–Burke–Ernzerhof form ESM). Additionally, the XPS spectra in Fig. S4 in the
[20]. The screened hybrid functional proposed by ESM indicate the presence of Zn2+, Co3+, and O2−;
Heyd, Scuseria, and Ernzerhof [21] was adopted to meanwhile, the Ni 2p3/2 peak at 856.0 eV indicate that
obtain accurate densities for the electronic states; the Ni was successfully doped in the lattice; Ni2+ was
plane wave cutoff energy was 480 eV, and a 3 × 3 × 1 the major species in the products (Fig. 1(d)). Element
forsheet k-point mesh was used. Ionic relaxations mapping was further conducted to determine the
were carried out under the conventional energy distribution of component elements; elemental Ni
(10−4 eV) and force (0.01 eV·Å−1) convergence criteria. was uniformly distributed in the synthetic products,
The single-layer along the [100] projection was used which is consistent with the XPS results. In addition,
to mimic the as-prepared 2.5-nm-thick atomic layers, the measured atomic percentage of doped Ni atoms
in which a 1.5-nm-thick vacuum layer was added to was 4.84% using ICP-AES. The FT-IR spectrum in
avoid interactions between adjacent layers. Fig. S5 in the ESM indicates the absence of organic

Figure 1 Characterizations of the Ni-doped ZnCo2O4 atomic layers: (a) TEM image, (b) high-resolution TEM image, and (c) the
corresponding interplanar spacing profiles along the dashed green lines; (d) X-ray photoelectron spectroscopy of Ni 2p, (e) atomic force
microscopy image, and (f) the corresponding height profiles; (g) high-angle annular dark-field scanning transmission electron
microscopy image and the corresponding energy-dispersive X-ray spectroscopy mapping images of Ni-doped ZnCo2O4 atomic layers.

| www.editorialmanager.com/nare/default.asp
Nano Res. 5

matter on the surfaces and the formation of clean 3.2 Suitable band structures for CO2 photoreduction
Ni-doped ZnCo2O4. Moreover, the TEM image in
Generally, when the semiconductor photocatalyst
Fig. 1(a) reveals that the as-obtained products possess
absorbs photons with energies higher than its band
a sheet-like morphology. As revealed by the high-
gap, the electrons would be excited into the conduction
resolution TEM image in Figs. 1(b) and 1(c), the 2.08 Å
band, leaving holes in the valence band [22, 23]. The
interplanar spacing matches well with the d400 spacings,
photogenerated electrons would be captured by CO2
and the corresponding dihedral angle of 90° agrees
and hence reduce it to carbon-based fuels, while the
with the calculated angle between the (040) and (004)
photoexcited holes would simultaneously oxidize H2O
planes of cubic ZnCo2O4, suggesting the [100] orientation
to O2. Thus, a suitable band structure is crucial for
of the synthetic two-dimensional Ni-doped ZnCo2O4
photocatalytic reduction of CO2. Two prerequisites
atomic layers. The AFM image and corresponding
exist for the band structure to realize solar-driven
height profiles in Figs. 1(e) and 1(f) show average CO2 reduction: (1) The conduction band edge should
thicknesses of approximately 2.5 nm, which correspond be more negative than the potential of CO2 reduction,
to the three-unit cell thickness along the [100] projection. and (2) the valence band edge should be more positive
Consequently, it was verified that clean Ni-doped than the potential of water oxidation. Hence, UV–vis
ZnCo2O4 atomic layers were first successfully fabricated. DRS and SRPES were performed to investigate their
For comparison, non-doped ZnCo2O4 atomic layers of band structures. As shown in Fig. 2(a), the Ni-doping
the same thickness were also synthesized under the endowed the ZnCo2O4 atomic layers with significantly
same conditions (except for the Ni source; Figs. S1–S6 increased photoabsorption from 300 to 2,500 nm. In
in the ESM). Therefore, two ideal models of atomic addition, the inset graph showed calculated bandgaps
layers with/without element doping were subsequently of 2.51 and 2.67 eV for the Ni-doped ZnCo2O4 atomic
built to study the role of doping engineering for layers and the ZnCo2O4 atomic layers, respectively.
selectivity regulating solar-driven CO2 reduction. Specifically, the extra adsorption band at 387 nm for

Figure 2 (a) Surface photovoltage response and temperature-programmed desorption curves of (b) CO2, (c) CH4, and (d) CO for the
Ni-doped ZnCo2O4 atomic layers and ZnCo2O4 atomic layers.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


6 Nano Res.

the Ni-doped ZnCo2O4 atomic layers belonged to d–d maxima that are 2.72 and 2.78 eV below the Fermi
crystal-field transitions (3T1(3P)→3A2(3F) transitions) level, respectively (Fig. 3(d)).
for high-spin Ni2+ (3d8). Three absorption bands were Therefore, the valence band maxima relative to
observed at 524, 580, and 628 nm, which could be the vacuum level for the ZnCo2O4 atomic layers
assigned to the d–d crystal-field transitions (1T2(3G), with/without Ni-doping are 1.26 and 1.34 eV, respec-
1
T1(3D), 3T2(3P)→ground state 1A1(5D) transitions) for tively. Along with the optical band gaps, the electronic
the low-spin Co3+ (3d6) in tetrahedral coordination [24]. band energies were calculated for both samples relative
The absorption peak around 1,250 nm corresponded to a normal hydrogen electrode, as shown in Fig. 3(b);
to the transition from 3T1(3P)→ground state 1A1(5D), their valence bands could produce oxygen and their
3
T2(3F)→3A2(3F), and 3T2(3F)→3A2(3F) for low-spin Co3+ conduction bands could satisfy both the carbon
(3d6), according to Tanabe–Sugano theory [25], in monoxide and methane production potentials, which
which the electrons in fully occupied electron levels further verify their photocatalytic performances to
can be easily excited to the unoccupied levels under produce carbon forms and oxygen.
irradiation, thus providing a superior light conversion
3.3 Enhanced electron–hole separation efficiency
efficiency. Moreover, SRPES, which was a powerful
and chemisorption properties
technique for resolving the energy band structures,
was also performed to determine the band structure The separation process of photoexcited electron–hole
positions. The calculated work functions of the ZnCo2O4 pairs is a major concern for promoting solar-driven
atomic layers with/without Ni-doping are 3.04 and CO2 reduction. Thus, an in-depth understanding of
3.06 eV, respectively, using the equation provided in the charge separation is necessary to achieve efficient
section 2.3 (Fig. 3(c)). Meanwhile, the ZnCo2O4 atomic CO2 photoreduction [22]. In this case, SPV spectra
layers with/without Ni-doping showed valence band were measured to investigate the charge carrier

Figure 3 Electronic band structures for the Ni-doped ZnCo2O4 atomic layers and the ZnCo2O4 atomic layers. (a) UV−vis diffuse
reflectance spectra, (b) band structures and corresponding reduction potentials, (c) secondary electron cutoff (Ecutoff) measured by SRPES,
and (d) SRPES valence band spectra.

| www.editorialmanager.com/nare/default.asp
Nano Res. 7

separation process of both samples [7, 8]. As illustrated of the ZnCo2O4 atomic layers because of the advantages
in Fig. 2(a), the Ni-doped ZnCo2O4 atomic layers exhi- discussed above. Their photoreduction measurements
bited higher SPV responses than the ZnCo2O4 atomic were performed for a gas-solid system, of which
layers, suggesting that the doped Ni atoms promoted the gases were a mixture of CO2 and water vapor.
the separation efficiency of the photoexcited carriers. As shown in Fig. 4(a), the Ni-doped ZnCo2O4 atomic
Despite the improvement, it was still not clear if the layers provided CO and CH4 yields of 31.4 and
selectivity change was caused by Ni-doping. In this 20.2 μmol·g−1·h−1, respectively, while the ZnCo2O4
case, specific surface area measurements (BET) and CO2 atomic layers exhibited CO and CH4 yields of 12.4
isothermal adsorption were conducted. Intriguingly, and 28.1 μmol·g−1·h−1, respectively. Conspicuously,
because of the similar morphologies, thicknesses, and the selectivity of CO/CH4 for the Ni-doped ZnCo2O4
crystal orientations, both the Ni-doped ZnCo2O4 atomic atomic layers was 352% greater than that of the
layers and the ZnCo2O4 atomic layers exhibited nearly ZnCo2O4 atomic layers, which agreed with the TPD
the same specific surface area (43.77 vs. 45.18 m2·g−1, results shown in Figs. 2(c) and 2(d). The ZnCo2O4 atomic
respectively) (Fig. S7(a) in the ESM). In addition, the layers with/without Ni-doping could simultaneously
similar adsorption amount of CO2 at ambient tem- realize H2O oxidation to O2 with average O2 evolution
perature (Fig. S7(b) in the ESM) further demonstrated rates of approximately 51.6 and 59.7 μmol·g−1·h−1,
the consistent physical properties of the ZnCo2O4 respectively, during the 20-h photocatalysis tests
atomic layers with/without Ni-doping; thus, the (Fig. S8 in the ESM). The yields of O2 were close to
different selectivity might be caused by chemical the theoretical yields of 56.1 and 62.4 μmol·g−1·h−1
doping by Ni atoms. Moreover, TPD analyses were derived from the CO and CH4 yields, respectively.
conducted to study the chemisorption process of the These results verified the crucial role of Ni-doping in
involved species in the reduction process. Note that improving the photocatalytic selectivity of CO/CH4
the chemidesorption temperatures of CO2 from the production in CO2 photoreduction. Both samples
ZnCo2O4 atomic layers with/without Ni-doping were survived three 21-h cycles (Fig. 4(b)) and maintained
69 and 47 °C (Fig. 2(b)), respectively, which suggests nearly the same CO2 photoreduction activities and
the strong adsorption of CO2 with Ni-doped ZnCo2O4 selectivities, indicating their superior stabilities. Based
atomic layers, thus favoring the photoreduction on the above analysis, Ni-doping could account for the
process. More importantly, the chemidesorption tem- 3.5 times greater selectivity of CO/CH4 production
peratures of CH4 for the Ni-doped ZnCo2O4 atomic during CO2 photoreduction.
layers and the ZnCo2O4 atomic layers were 61 and
3.5 In situ DRIFTS and possible mechanism of
18 °C, respectively, while their CO chemidesorption
CO2 photoreduction
temperatures were 10 and 74 °C, respectively (Figs. 2(c)
and 2(d)). Thus, CO desorption was favored over CH4 To disclose the insight of the CO2 photoreduction
desorption on the Ni-doped ZnCo2O4 atomic layers, process, in situ DRIFTS of both samples was
whereas the reverse occurred for the ZnCo2O4 atomic conducted to detect their intermediates during CO2
layers. Ni-doped ZnCo2O4 atomic layers might possess photoreduction; in situ DRIFTS is a real-time technique
a higher CO production rate, while the ZnCo2O4 for observing the surface species that exist in the
atomic layers exhibit a higher CH4 production rate. transition states. As an example, the synthetic Ni-doped
In short, the doping engineering confined in atomic ZnCo2O4 atomic layers generated new species with
layers would likely regulate product selectivity for increasing illumination time (Fig. 5(a)).
CO2 photoreduction. The concentrations of the involved species in the
photoreduction process correspond to the response
3.4 Boosted performances of solar-driven CO2
intensity depicted by the color in the three-dimensional
reduction
(3D) color-fill surface and heat maps of Fig. 5(a). Broad
The Ni-doped ZnCo2O4 atomic layers exhibit an peaks were visible from 1,600 to 1,700 cm−1 (Fig. 5(b));
enhanced CO2 photoreduction activity relative to that the peaks at 1,679 and 1,123 cm−1 could be attributed

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


8 Nano Res.

Figure 4 3D histogram for the photoreduction of CO2 under irradiation from a 300-W xenon lamp for the Ni-doped ZnCo2O4 atomic
layers (cylinders cubes) and ZnCo2O4 atomic layers (cylinders). (a) Reduction rates during one 21-h cycle and (b) stability of three
cycles.

Figure 5 In situ diffuse reflectance infrared Fourier transform spectroscopy of CO2 photoreduction process for the Ni-doped ZnCo2O4
atomic layers. (a) 3D color-fill surface of photocatalysis for up to 40 min and (b) the corresponding heat maps.

to the absorption band of bicarbonate (asymmetric atomic layers, indicating that the Ni-doped ZnCo2O4
stretching of b-HCO3−), while the peak at 1,565 cm−1 atomic layers provided a higher CO yield. Overall,
was assigned to the symmetric OCO stretching of these spectroscopic data confirmed that the surface
CO2− radicals [26, 27], implying that an electron bound CO2− species were the most likely reaction
could be spontaneously transferred to CO2 under intermediates from CO2 photoreduction using both
illumination, ascribed to the transition process of CO2 Ni-doped ZnCo2O4 atomic layers and ZnCo2O4 atomic
+ e− → CO2−. Peaks at 2,109 and 1,320 cm−1 gradually layers. To further confirm the crucial role of Ni-doping
increased as the photolysis time increased; these engineering in tailoring the electronic structure at
peaks could be assigned to physically adsorbed CO an atomic level, we performed DFT calculations to
(COad) [28] and the H–C–H bending vibration in CH4 investigate the variation in the DOS. Evidently, as
[29], indicating an accumulation of photoreduction revealed by the calculated DOS shown in Fig. 6, the
products [30]. These as-assigned peaks also existed in presence of the Ni dopant endowed the ZnCo2O4
the photocatalytic process of the non-doped ZnCo2O4 atomic layers with an increased DOS at the CBM
atomic layers, as shown in Fig. S9 in the ESM, in with respect to the perfect ZnCo2O4 atomic layers in
which the vibration peaks of COad and intermediate the slab. In addition, the spatial distribution of the
CO2− were weaker than that of the Ni-doped ZnCo2O4 orbital wave functions at the CBM revealed that the

| www.editorialmanager.com/nare/default.asp
Nano Res. 9

Figure 6 Density functional theory calculations. Calculated density of states of (a) Ni-doped ZnCo2O4 atomic-layer slab and (b) ZnCo2O4
atomic-layer slab. Charge density distribution calculation of (c) Ni-doped ZnCo2O4 atomic-layer slab and (d) ZnCo2O4 atomic-layer slab.

increased charge density originated from the Co and


Zn atoms with the introduction of Ni atoms (Figs. 6(b)
and 6(d)). More importantly, the introduction of Ni
atoms produced several new energy levels in the
Ni-doped ZnCo2O4 atomic layers in the slab (Fig. 6),
which could be ascribed to splitting of the Ni 3d
states [18]. In this case, the photogenerated electrons
could be easily excited via the d→d internal transition
of tetrahedrally coordinated Ni ions under solar light Scheme 1 Schematic illustration for the possible mechanism
irradiation, suggesting that the Ni dopants confined of CO2 reduction with the as-synthesized ZnCo2O4 atomic layers
in the atomic layers could be directly involved in the with/without Ni-doping under illumination.
photocatalytic CO2 reduction and hence significantly
improve the photocatalytic activity and selectivity. following reduction reaction, and (5) desorbing the
Hence, a possible solar-driven reduction path was reduction products. Interestingly, as verified by the
proposed (Scheme 1) according to the intermediates SPV spectra, Ni-doping assisted the electron–hole
revealed by in situ DRIFTS [10, 31–33]. pair process in the ZnCo2O4 atomic layers; thus,
Generally, the CO2 photoreduction includes the more photogenerated electrons are anticipated in the
following crucial processes: (1) generating electron– following reduction of CO2. UV–vis spectroscopy
hole pairs with illumination on the semiconductor, and SRPES indicated that the band structures of both
(2) separating photoexcited electron–hole pairs, the Ni-doped ZnCo2O4 atomic layers and the ZnCo2O4
(3) adsorbing CO2, (4) transferring photogenerated atomic layers were suitable for CO2 reduction to CO
electrons to CO2, forming CO2− and allowing the and CH4. Moreover, the reaction intermediates were

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


10 Nano Res.

detected as CO2− radicals using in situ DRIFTS, calculations illustrated that the Ni doping led to a
suggesting that the photogenerated electrons would higher DOS near the CBM and new doping energy
initially reduce CO2 to CO2− radicals. The as-produced levels, beneficial for a higher separation efficiency of
intermediates would further react with photoexcited photogenerated electron–hole pairs. As a result, the
electrons and H+ to transform into CO and CH4 via Ni-doped ZnCo2O4 atomic layers possessed CO and
different paths, as shown in Scheme 1. The TPD CH4 formation rates of 31.4 and 20.2 μmol·g−1·h−1,
results revealed that the CO desorption process was respectively, while those of the ZnCo2O4 atomic
easier using the Ni-doped ZnCo2O4 atomic layers, layers were 12.4 and 28.1 μmol·g−1·h−1, respectively; the
while the CH4 desorption process was easier using former shows a 3.5 times higher selectivity toward
the ZnCo2O4 atomic layers; this implies that the CO evolution than CH4 production. Briefly, this work
Ni-doping engineering benefitted the CO a desorption developed an element-doping strategy for tuning the
process. The Ni atoms doped in the lattice of the product selectivity of solar-driven CO2 reduction with
ZnCo2O4 atomic layers changed the adsorption energy atomically thin semiconductors, providing important
of different species for the extra adsorption structure clues for controlling and tailoring the selectivity of
of Ni–O; Co–O and Zn–O (O in CO2 and CO) the desired products during CO2 photoreduction.
(Scheme S1 in the ESM) could account for the
different chemisorption ability of the products for Acknowledgements
Ni-doped ZnCo2O4 atomic layers and ZnCo2O4 atomic
layers. This work was financially supported by the National
Key Research and Development Program of China
(Nos. 2017YFA0303500 and 2017YFA0207301), the
4 Conclusions National Natural Science Foundation of China (Nos.
Element doping confined to atomically thin two- 21422107, U1632147, 21331005, U1532265, and 11621063),
dimensional layers was first applied as a selective Youth Innovation Promotion Association of CAS (No.
tuning approach for photocatalytic CO2 reduction. As CX2340000100), Key Research Program of Frontier
a proof-of-concept prototype, ZnCo2O4 atomic layers Sciences of CAS (No. QYZDY-SSW-SLH011), the
with/without Ni doping were first synthesized using Fundamental Research Funds for the Central
a low temperature (i.e., 140 °C) solvothermal strategy. Universities (Nos. WK2340000063 and WK2340000073)
and Scientific Research Grant of Hefei Science Center
Energy-dispersive X-ray spectroscopy mapping
of CAS (No. 2016HSC-IU002).
clarified the uniform distribution of doped Ni atoms
in the ZnCo2O4 atomic layers. The band structures
Electronic Supplementary Material: Supplementary
obtained by UV–vis spectroscopy and SRPES indicated
material (Figs. S1–S9) is available in the online version
that both the Ni-doped ZnCo2O4 atomic layers and
of this article at https://doi.org/10.1007/s12274-017-
the ZnCo2O4 atomic layers could photoreduce CO2 to
1943-2.
CO and CH4. The higher separation efficiency of the
photo-excited electron–hole pairs for the Ni-doped
ZnCo2O4 atomic layers was verified by the SPV spectra. References
In situ DRIFTS spectra provided evidence for the
[1] Lei, F. C.; Liu, W.; Sun, Y. F.; Xu, J. Q.; Liu, K. T.;
existence of CO2− radical as the main intermediate in
Liang, L.; Yao, T.; Pan, B. C.; Wei, S. Q.; Xie, Y. Metallic
the photoreduction of CO2 for both samples. TPD
tin quantum sheets confined in graphene toward high-
curves revealed that the Ni-doped ZnCo2O4 atomic efficiency carbon dioxide electroreduction. Nat. Commun.
layers showed a lower CO desorption temperature 2016, 7, 12697.
than the ZnCo2O4 atomic layers, indicating that the [2] Gao, S.; Lin, Y.; Jiao, X. C.; Sun, Y. F.; Luo, Q. Q.; Zhang,
chemisorption ability with products would account W. H.; Li, D. Q.; Yang, J. L.; Xie, Y. Partially oxidized
for the preferential selectivity of CO evolution for atomic cobalt layers for carbon dioxide electroreduction to
Ni-doped ZnCo2O4 atomic layers. Furthermore, DFT liquid fuel. Nature 2016, 529, 68–71.

| www.editorialmanager.com/nare/default.asp
Nano Res. 11

[3] Liang, L.; Lei, F. C.; Gao, S.; Sun, Y. F.; Jiao, X. C.; Wu, J.; evolution. Nano Energy 2016, 30, 810–817.
Qamar, S.; Xie, Y. Single unit cell bismuth tungstate layers [15] Sato, S.; Morikawa, T.; Saeki, S.; Kajino, T.; Motohiro, T.
realizing robust solar CO2 reduction to methanol. Angew. Visible-light-induced selective CO2 reduction utilizing a
Chem., Int. Ed. 2015, 54, 13971–13974. ruthenium complex electrocatalyst linked to a p-type
[4] Xu, J. Q.; Li, X. D.; Liu, W.; Sun, Y. F.; Ju, Z. Y.; Yao, T.; nitrogen-doped Ta2O5 semiconductor. Angew. Chem., Int.
Wang, C. M.; Ju, H. X.; Zhu, J. F.; Wei, S. Q. et al. Carbon Ed. 2010, 49, 5101–5105.
dioxide electroreduction into syngas boosted by a partially [16] Teramura, K.; Wang, Z.; Hosokawa, S.; Sakata, Y.; Tanaka,
delocalized charge in molybdenum sulfide selenide alloy T. A doping technique that suppresses undesirable H2
monolayers. Angew. Chem., Int. Ed. 2017, 56, 9121–9125. evolution derived from overall water splitting in the highly
[5] Qiu, Q. Q.; Li, S.; Jiang, J. J.; Wang, D. J.; Lin, Y. H.; Xie, selective photocatalytic conversion of CO2 in and by water.
T. F. Improved electron transfer between TiO2 and FTO Chemistry 2014, 20, 9906–9909.
interface by N-doped anatase TiO2 nanowires and its [17] Wang, S.; Ding, Z. X.; Wang, X. C. A stable ZnCo2O4
applications in quantum dot-sensitized solar cells. J. Phys. cocatalyst for photocatalytic CO2 reduction. Chem. Commun.
Chem. C 2017, 121, 21560–21570. 2015, 51, 1517–1519.
[6] Qamar, S.; Lei, F. C.; Liang, L.; Gao, S.; Liu, K. T.; Sun, Y. [18] Zhu, Y. Q.; Cao, C. B.; Zhang, J. T.; Xu, X. Y. Two-
F.; Ni, W. X.; Xie, Y. Ultrathin TiO2 flakes optimizing dimensional ultrathin ZnCo2O4 nanosheets: General formation
solar light driven CO2 reduction. Nano Energy 2016, 26, and lithium storage application. J. Mater. Chem. A 2015, 3,
692–698. 9556–9564.
[7] Li, H. Y.; Wang, D. J.; Fan, H. M.; Jiang, T. F.; Li, X. L.;
[19] Surendranath Y.; Kanan M. W.; Nocera D. G. Mechanistic
Xie, T. F. Synthesis of ordered multivalent Mn-TiO2
studies of the oxygen evolution reaction by a cobalt-phosphate
nanospheres with tunable size: A high performance visible-
catalyst at neutral pH. J. Am. Chem. Soc. 2010, 132, 16501–
light photocatalyst. Nano Res. 2011, 4, 460–469.
16509.
[8] Jiang, T. F.; Xie, T. F.; Zhang, Y.; Chen, L. P.; Peng, L. L.;
[20] Perdew J. P.; Burke K.; Ernzerhof M. Generalized gradient
Li, H. Y.; Wang, D. J. Photoinduced charge transfer
approximation made simple. Phys. Rev. Lett. 1996, 77,
in ZnO/Cu2O heterostructure films studied by surface
3865–3868.
photovoltage technique. Phys. Chem. Chem. Phys. 2010,
[21] Shek, C. H.; Lai, J. K. L.; Lin, G. M. Investigation of interface
12, 15476–15481.
defects in nanocrystalline SnO2 by positron annihilation.
[9] Fletcher, C.; Jiang, Y. J.; Sun, C. H.; Amal, R. Morphological
J. Phys. Chem. Solids 1999, 60, 189–193.
evolution and electronic alteration of ZnO nanomaterials
[22] Jiao, X. C.; Chen, Z. W.; Li, X. D.; Sun, Y. F.; Gao, S.;
induced by Ni/Fe co-doping. Nanoscale 2014, 6, 7312–7318.
Yan, W. S.; Wang, C. M.; Zhang, Q.; Lin, Y.; Luo, Y. et al.
[10] Ong, W.-J.; Tan, L.-L.; Chai, S.-P.; Yong, S.-T.; Mohamed,
Defect-mediated electron–hole separation in one-unit-cell
A. R. Self-assembly of nitrogen-doped TiO2 with exposed
ZnIn2S4 layers for boosted solar-driven CO2 reduction. J.
{001} facets on a graphene scaffold as photo-active hybrid
Am. Chem. Soc. 2017, 139, 7586–7594.
nanostructures for reduction of carbon dioxide to methane.
[23] Iizuka, K.; Wato, T.; Miseki, Y.; Saito, K.; Kudo, A.
Nano Res. 2014, 7, 1528–1547.
Photocatalytic reduction of carbon dioxide over Ag
[11] Mao, J.; Li, K.; Peng, T. Y. Recent advances in the
photocatalytic CO2 reduction over semiconductors. Catal. cocatalyst-loaded AlA4Ti4O15 (A = Ca, Sr, and Ba) using
Sci. Technol. 2013, 3, 2481–2498. water as a reducing reagent. J. Am. Chem. Soc. 2011, 133,
[12] Lei, F. C.; Sun, Y. F.; Liu, K. T.; Gao, S.; Liang, L.; Pan, B. 20863–20868.
C.; Xie, Y. Oxygen vacancies confined in ultrathin indium [24] Lei, F. C.; Zhang, L.; Sun, Y. F.; Liang, L.; Liu, K. T.;
oxide porous sheets for promoted visible-light water splitting. Xu, J. Q.; Zhang, Q.; Pan, B. C.; Luo, Y.; Xie, Y. Atomic-
J. Am. Chem. Soc. 2014, 136, 6826–6829. layer-confined doping for atomic-level insights into visible-
[13] Liu, Y. W.; Xiao, C.; Li, Z.; Xie, Y. Vacancy engineering for light water splitting. Angew. Chem., Int. Ed. 2015, 54,
tuning electron and phonon structures of two-dimensional 9266–9270.
materials. Adv. Energy Mater. 2016, 6, 1600436. [25] Balti, I.; Mezni, A.; Dakhlaoui-Omrani, A.; Léone, P.; Viana,
[14] Liu, K. T.; Zhang, W. S.; Lei, F. C.; Liang, L.; Gu, B. C.; B.; Brinza, O.; Smiri, L.-S.; Jouini, N. Comparative study
Sun, Y. F.; Ye, B. J.; Ni, W. X.; Xie, Y. Nitrogen-doping of Ni- and Co-substituted ZnO nanoparticles: Synthesis,
induced oxygen divacancies in freestanding molybdenum optical, and magnetic properties. J. Phys. Chem. C 2011, 115,
trioxide single-layers boosting electrocatalytic hydrogen 15758–15766.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


12 Nano Res.

[26] Liu, Y. M.; Chen, S.; Quan, X.; Yu, H. T. Efficient [30] Cao, Y.; Li, H. R.; Zhang, J. P.; Shi, L. Y.; Zhang, D. S.
electrochemical reduction of carbon dioxide to acetate on Promotional effects of rare earth elements (Sc, Y, Ce, and
nitrogen-doped nanodiamond. J. Am. Chem. Soc. 2015, 137, Pr) on nimgal catalysts for dry reforming of methane. RSC
11631–11636. Adv. 2016, 6, 112215–112225.
[27] Neaţu, Ş.; Maciá-Agulló, J. A.; Concepción, P.; Garcia, H. [31] Wang, W.; Gong, J. L. Methanation of carbon dioxide: An
Gold-copper nanoalloys supported on TiO2 as photocatalysts overview. Front. Chem. Sci. Eng. 2011, 5, 2–10.
for CO2 reduction by water. J. Am. Chem. Soc. 2014, 136, [32] Zhang, L.; Zhao, Z. J.; Gong, J. L. Nanostructured materials
15969–15976. for heterogeneous electrocatalytic CO2 reduction and their
[28] Grabow, L. C.; Mavrikakis, M. Mechanism of methanol related reaction mechanisms. Angew. Chem., Int. Ed. 2017,
synthesis on Cu through CO2 and CO hydrogenation. ACS 56, 11326–11353.
Catal. 2011, 1, 365–384. [33] Zhang, L.; Wang, W. Z.; Jiang, D.; Gao, E. P.; Sun, S. M.
[29] Wilcox, E. M.; Roberts, G. W.; Spivey, J. J. Direct catalytic Photoreduction of CO2 on BiOCl nanoplates with the
formation of acetic acid from CO2 and methane. Catal. assistance of photoinduced oxygen vacancies. Nano Res.
Today 2003, 88, 83–90. 2015, 8, 821–831.

| www.editorialmanager.com/nare/default.asp

Вам также может понравиться