You are on page 1of 25

Journal Pre-proof

A leaf-branch TiO2 /Carbon@MOF composite for selective CO2


photoreduction

Awu Zhou, Yibo Dou, Chen Zhao, Jian Zhou, Xue-Qian Wu,
Jian-Rong Li

PII: S0926-3373(19)31265-2
DOI: https://doi.org/10.1016/j.apcatb.2019.118519
Reference: APCATB 118519

To appear in: Applied Catalysis B: Environmental

Received Date: 4 September 2019


Revised Date: 10 November 2019
Accepted Date: 9 December 2019

Please cite this article as: Zhou A, Dou Y, Zhao C, Zhou J, Wu X-Qian, Li J-Rong, A
leaf-branch TiO2 /Carbon@MOF composite for selective CO2 photoreduction, Applied
Catalysis B: Environmental (2019), doi: https://doi.org/10.1016/j.apcatb.2019.118519

This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2019 Published by Elsevier.


A leaf-branch TiO2/Carbon@MOF composite for selective CO2

photoreduction

Awu Zhou, Yibo Dou, Chen Zhao, Jian Zhou, Xue-Qian Wu, Jian-Rong Li*

Beijing Key Laboratory for Green Catalysis and Separation and Department of Chemistry and

Chemical Engineering, College of Environmental and Energy Engineering, Beijing University of

Technology, Beijing 100124, P. R. China

E-mail: jrli@bjut.edu.cn

of
ro
Graphical Abstract
-p
re
A leaf-branch TiO2/Carbon@MOF composite for selective CO2
lP

photoreduction
na
ur
Jo

1
A leaf-branch TiO2/C@ZIF-L is fabricated by the in-situ growth of leaf-like zeolitic imidazolate
frameworks (ZIF-L) on branch-like TiO2/C nanofiber. The resulting composite photocatalyst exhibits
improved catalytic efficiency and good selectivity for the CO2 photoreduction on synergistic effect.

Research highlights

 A leaf-branch TiO2/C@MOF composite photocatalyst is well designed and

fabricated.

of
ro
 The catalyst shows outstanding activity and selectivity for CO2 photoreduction.
-p
re
 The synergistic effect on enhanced photocatalytic performance is investigated.
lP
na

Abstract: Photocatalytic CO2 reduction provides a promising way for solving the energy crisis and

environmental issues. Herein, a well-integrated photocatalyst is fabricated through growing leaf-like


ur

zeolitic imidazolate frameworks (ZIF-L) on branch-like titanium dioxide/carbon (TiO2/C) nanofiber


Jo

for selective CO2 reduction. This platform enables the study of the synergistic effect of the

photocatalytic performance, where the strong CO2 adsorption/activation, highly efficient

light-harvesting, promoted electron-hole pairs separation, and specific metal Lewis sites are met by

integrating suitable band matched ZIF and TiO2, as well as interfaced graphitic carbon. As a result, a

2
largely increased CO generation rate of 28.6 μmol h−1 g−1 with remarkable selectivity of 99% is

achieved in the TiO2/C@ZnCo-ZIF-L without using sacrificial reagent. This work thus provides an

insight into the rational construction of well-integrated composites to enhance photocatalytic

performance via a well-balanced synergistic effect.

Keywords: Leaf-branch structure, Metal-organic framework (MOF), Photocatalysis, TiO2 fibers,

Z-scheme

of
ro
-p
re
lP
na
ur
Jo

3
1. Introduction

Artificial photosynthesis via solar-driven carbon dioxide (CO2) conversion to value-added

chemicals is a promising way for simultaneously alleviating the ever-growing energy crisis and

environmental pressure [1-6]. Great efforts have been devoted to developing new photocatalysts for

improving the conversion and selectivity of CO2 reduction. Up to now, various photocatalysts,

especially semiconductors-based ones have been widely explored [7-11]. Nevertheless, the

of
photocatalytic efficiency achieved so far is still not at an acceptable level yet, mainly because of the

high recombination of photogenerated charge carriers, low utilization of visible light, and weak

ro
adsorption/activation capability of CO2 on the catalysts active sites [12-15]. In addition, the
-p
photoreduction of CO2 involves complicated multi-electron transfer processes, leading to a variety of
re
products and thus poor selectivity [16-20].

Among various developed strategies to improve the photocatalytic performance, the


lP

construction of composite catalysts by optimizing the structure and composition has attracted great

interest [21-27]. For example, Tahir et al. presented that the synergistic effect in Au-MM/TiO2
na

catalyst remarkably enhances CO2 photoreduction to CO [28]. Liu et al. reported ZIF-8 modified
ur

tubular g-C3N4 photocatalysts with a great enhancement for photocatalytic CO2 reduction into

CH3OH [29]. Chew et al. constructed hierarchical Zn2In2S5/Ti3C2(O, OH)x hybrids toward
Jo

photocatalytic hydrogen evolution under visible light [30]. Despite of great endeavors, enhancing the

conversion efficiency and the selectivity of composite photocatalysts is still inhibited by the

following main challenges: (1) the developed strategies could not simultaneously guarantee the

composite photocatalysts possessing the high solar light utilization, efficient charge separation, as

4
well as highly exposed active sites of activating CO2, which lead to unremarkable synergistic

photocatalytic effect [29,31-33]; (2) the illustration of complicated synergistic effect of each

component on the photocatalytic performance is still rare, but critical for the design and construction

of novel efficient photocatalysts [34,35]; (3) the weak interfacial interactions in some explored

composites cannot ensure the efficient electron transfer among components and the structural

integrity, resulting in unfavorable electron-hole pair recombination and poor stability [13,16,36,37];

and (4) the catalytic active sites in some composites are normally uncontrollable, which distribute

of
on different components randomly, thus leading to poor selectivity [30]. Therefore, exploring

ro
well-structured composite photocatalysts to control, and then optimize the interplay among

efficiency, selectivity, and stability remains a challenge [4,39].


-p
Herein, we propose the well-integration of porous metal-organic frameworks (MOFs) with
re
carbon/semiconductors to fabricate sophisticated photocatalysts for high-performance CO2 reduction
lP

(Scheme 1). As a proof of concept, leaf-branch TiO2/C@ZIF-L fabricated by the in-situ growth of

leaf-like zeolitic imidazolate frameworks (ZIF-L) nanoplates on the branch-like TiO2/C nanofiber is
na

firstly presented as an ideal model for disclosing insights into the roles of synergistic photocatalytic

effect. The fabricated ternary photocatalysts could afford the advantages of each component
ur

including favorable CO2 adsorption capability, efficient solar light utilization, rich exposed catalytic
Jo

active sites. Moreover, the synergistic effect of components that including suitable band matching

between ZIF-L and TiO2, and interfacial superior electron mobility from graphitic carbon could

facilitate the effective electron-hole separation. In addition, the metal Lewis sites in ZIF-L are

favorable for driving CO2 into CO. As a result, the TiO2/C@ZnCo-ZIF-L photocatalysts display

remarkable CO conversion (28.6 μmol h−1 g−1) and selectivity (99%) without using photosensitizer
5
and sacrificial reagent. The CO production of the composite is obviously improved compared with its

components (> seven-fold increase) and the commercial P25 benchmark photocatalyst (>

seventy-fold increase). The investigation of the relationship between synergistic effect and

photocatalytic performance may provide general insight into the rational design and fabrication of

well-integrated composite photocatalysts.

of
ro
-p
re
Scheme 1. Schematic illustration for the fabrication of leaf-branch TiO2/C@ZIF-L photocatalysts by
lP

three steps including electrospinning of TiO2/PAN fiber, pyrolysis of TiO2/PAN fiber, and in-situ

growth of ZIF-L on the TiO2/C@ZIF-L fiber.


na

2. Experimental section
ur

2.1 Preparation of TiO2/C mat


Jo

The TiO2/PAN fibers were firstly prepared via facile electrospinning and pyrolysis process

reported previously [40,41]. Then, the fibrous TiO2/PAN mat was firstly oxidized at 280 °C for 1 h

in air atmosphere with a heating rate of 5 °C min−1, then carbonized at 400 °C for 0.5 h in Ar

atmosphere with a heating rate of 5 °C min−1.

6
2.2 Preparation of TiO2/C@ZnCo-ZIF-L, TiO2/C@Zn-ZIF-L, and TiO2/C@Co-ZIF-L mat

In a typical synthesis, Zn(NO3)2·6H2O (1 mmol) and Co(NO3)2·6H2O (1 mmol) were dissolved

in 40 mL deionized water under vigorous stirring (donated as A). Then, 2-methylimidazole (2-MeIM,

16 mmol) was dissolved in another 40 mL deionized water under vigorous stirring (donated as B).

Subsequently, the metal nitrate solution (A) was poured into the 2-MeIM solution (B) quickly, and a

piece of TiO2/C mat (2  2 cm2) was placed in the above mixing solution instantly without stirring at

room temperature for 8 h. Finally, the resulting TiO2/C@ZnCo-ZIF-L mat was withdrawn and

of
washed with deionized water and anhydrous ethanol three times and then dried under vacuum at

ro
60 °C for 5 h. In addition, the TiO2/C@Co-ZIF-L and TiO2/C@Zn-ZIF-L mats were fabricated with

-p
similar procedures by using Co(NO3)2·6H2O and Zn(NO3)2·6H2O as the metal source, respectively.

2.3 Photoreduction of CO2


re
Photoreduction CO2 experiment was conducted in a Teflon-lined stainless reaction chamber
lP

with a quartz window at the top for light irradiation [16,42,43]. The photocatalyst (43 mg) was spread

in the reaction chamber. Prior to the photoreaction, the chamber was evacuated by a mechanical pump.
na

Then, CO2 (Beijing Beiwen Gas Factory, ≥ 99.999%) bubbled through water was added to the
ur

chamber to achieve a pressure of 0.1 MPa. The products formed under full-spectrum irradiation (300

W Xe lamp, Beijing Perfect-Light Technology Co. Ltd) were determined at regular intervals from the
Jo

chamber. The products were analyzed by an Agilent 7890B gas chromatograph (GC) using He as the

carrier gas and a thermal conductivity detector (TCD), equipped with an active-carbon-packed

column. The amount of produced H2 was determined by gas chromatography (Shimadzu GC-2014C),

using a thermal conductivity detector (TCD) with N2 as the carrier gas. The isotope-labeled

7
experiments were examined using 13CO2 instead of 12CO2, and the products were analyzed by a gas

chromatography-mass spectrometry (GCMS-QP2010 SE, equipped with Micropacked column).

3. Results and discussion

TiO2/C@ZIF-L fibers were a fabrication by the following steps including electrospinning,

pyrolysis and in-situ growth process. The scanning electron microscopy (SEM) and transmission

electron microscopy (TEM) were used to track the fabricating process from TiO 2/PAN, TiO2/C to

final TiO2/C@ZnCo-ZIF-L fibers (Fig. 1). First of all, the TiO2/PAN fibers with a diameter of ~350

of
nm were fabricated by the electrospinning (Fig. S1). The subsequent pyrolysis treatment resulted in

ro
the TiO2/C nanofibers with rough and wrinkled surface (Fig. 1A). Finally, the uniform leaf-like

-p
ZnCo-ZIF-L nanoplates (Fig. S2) were in-situ grown on the surface of the TiO2/C fibers (Fig. 1B).

The TEM image further revealed that the TiO2/C@ZnCo-ZIF-L fibers displayed leaf-branch structure
re
(Fig. 1C). Moreover, the high-resolution TEM (HR-TEM) image of the edge section of the ternary
lP

composite confirmed the well-connected interfaces among ZnCo-ZIF-L, carbon, and TiO2.

Meanwhile, the fringes with lattice spaces of ~0.35 nm can be indexed to the (101) plane of the
na

anatase TiO2 phase (Fig. 1D) [44,45]. In addition, Energy-dispersive X-ray spectroscopy (EDX)

mapping with signals for Ti, C, Co, and Zn elements confirmed the homogeneous distribution of
ur

carbon and ZIF-L on the TiO2 nanofibers (Fig. 1E and S3). Similarly, leaf-branch TiO2/C@Co-ZIF-L
Jo

and TiO2/C@Zn-ZIF-L were fabricated, respectively (Fig.S4–S7).

The composition and phase structure of the obtained TiO2/C@ZIF-L was investigated by X-ray

diffraction (XRD). As shown in Fig. 2A, the XRD pattern of bare TiO2 displays a tetragonal anatase

structure [46,47]. After the decoration of ZIF-L (Fig. S8), a superimposition of the patterns for the

8
TiO2 and ZIF-L without other new diffraction peaks was observed from TiO2/C@ZIF-L. The Fourier

of
ro
-p
Fig. 1. (A) SEM images of TiO2/C fibers. (B) SEM, (C) TEM, (D) HR-TEM, and (E) EDX

elemental mapping images for TiO2/C@ZnCo-ZIF-L fibers.


re
transform infrared spectroscopy (FT-IR) also confirmed the existence of both TiO2 and ZIF-L in
lP

TiO2/C@ZIF-L (Fig. S9). Moreover, a shoulder peak at 804 cm−1 in the spectrum of TiO2/C and

TiO2/C@ZnCo-ZIF-L assigned to the Ti–O–C vibration was observed [48,49]. The results indicate
na

that the pyrolysis treatment results in the formation of chemisorption interfaces via Ti–O–C bonding,
ur

probably facilitating the charge transfer. In addition, the Raman spectrum was used to identify the

existence of carbon (Fig. 2B). Both of TiO2/C and TiO2/C@ZIF-L showed two intense bands at 1361
Jo

(D-band) and 1575 cm−1 (G-band), corresponding to the disordered or defect carbon and graphitic sp2

carbon (Fig. 2B), respectively [50]. Moreover, the low ID/IG (the intensity ratio of D to G bands)

values of 1.01 manifest a distinct degree of graphitization in the TiO2/C@ZIF-L fibers, being

favorable for improving conductivity. Besides the peaks for the carbon, the peaks at 146, 357, 509 and

9
628 cm−1 could be assigned to Eg, B1g, A1g + B1g and Eg of TiO2 [51,52], being consistent with above

XRD results.

X-ray photoelectron spectroscopy (XPS) was performed to investigate the interaction of

components in TiO2/C@ZIF-L. The C 1s peak of TiO2/C at 284.3 eV, 285.3 and 286.3 eV attributed to

sp2 and sp3 hybridized carbon (Fig. 2C), highly agreed with the Raman result. The peak at 288.2 eV

was assigned to Ti–O–C band [53], probably generated from the strong interaction between carbon

and TiO2. Fig. 3D showed the Ti 2p XPS spectra, which were deconvoluted by the Ti 2p3/2 and Ti 2p1/2

of
peaks. For TiO2, the Ti4+ state was determined by the binding energies of 457.75 and 463.47 eV for Ti

ro
2p3/2 and Ti 2p1/2, respectively. In contrast, the two Ti 2p peaks of TiO2/C increased to 458.87 and

464.62 eV, further proved well interaction between TiO2 and C, accorded with the FT-IR results. After
-p
further decorating ZnCo-ZIF-L, the binding energies of Ti 2p3/2 and Ti 2p1/2 in the
re
TiO2/C@ZnCo-ZIF-L fibers showed a positive shift to 458.65 and 464.08 eV. Simultaneously, the
lP

observed negative shift of the Co 2p (Fig. 2E) binding energies indicated an electron transfer from

TiO2 to ZIF-L (Fig. S10–S13), being favorable for the fast charge transfer at the interface. Moreover,
na

in comparison to pristine TiO2 and TiO2/C fibers, some new small peaks of TiO2/C@ZnCo-ZIF-L

located at 462.72 eV and 457.59 eV indicate the presence of Ti3+ bonds, which implies the formation
ur

of oxygen vacancy generated by the integration with ZIF-L [44,55]. In addition, the CO2 uptake
Jo

capability was also evaluated. Compared with bare TiO2/C, TiO2/C@ZIF-L exhibits increased CO2

uptakes of 0.79 mmol g−1 (Fig. 2F), which is attributed to high CO2 absorbing property (1 mmol g−1)

of ZIF-L with unique cushion-shaped cavities [56,57].

The UV–vis diffuse reflectance spectra were conducted to investigate the optical absorption

property (Fig. S14 and S15). The TiO2 displayed a steep absorption edge located at ~382 nm assigned
10
to the TiO bond [58], while the ZnCo-ZIF-L shell showed strong absorption in the visible light range

(450~700 nm) (Fig. 3A). As expected, the TiO2/C@ZnCo-ZIF-L exhibits a significantly extended

scale consisting of a visible light absorption band (400800 nm) and a UV absorption band. In

addition, a slight redshift to higher wavelength of 390 nm for steep absorption edge was observed,

which is probably attributed to the effective interactions between TiO2, carbon, and ZIF-L [49]. And,

the band gaps of 3.01 eV (TiO2) and 1.89 eV (ZnCo-ZIF-L) were calculated by the Kubelka-Munk

(KM) method (Fig. S16 and S17) [59,60]. The electron spin resonance (ESR) was utilized to explore

of
the unpaired electrons in photocatalysts (Fig. 3B). Compared with TiO2 and ZIF-L with a negligible

ro
signal peak, the TiO2/C and TiO2/C@ZIF-L shows a very strong EPR signal at g = 2.002. The result

indicated the efficient combination leads to the unpaired electrons trapped on surface oxygen
-p
vacancies in TiO2, which is consistent with the XPS results. More importantly, after decoration of
re
various ZIF-L, the TiO2/C@ZnCo-ZIF-L gave the strongest noted Ti3+ signal (Fig. 3B). These
lP

observed results reveal that introduction of ZIF-L enables the alternation of surface states on TiO2 and

generation of defects such as Ti3+ and oxygen vacancy.


na

The amperometric I-t curves under chopped-light illumination were also tested (Fig. 3C). The

current density follows the sequence of TiO2 < TiO2/C < TiO2/C@ZIF-L. In addition, the pure
ur

ZnCo-ZIF-L sample shows a relative low photocurrent response (Fig. S18). The result indicated that
Jo

well-integrated ternary composites extend the light harvest absorption range, improve electron-hole

separation, and boost the mobility of the charge carriers. Especially, the growth on ZnCo-ZIF-L and

Co-ZIF-L affords remarkable improved current density compared with Zn-ZIF-L, indicating the

transition Co ion in the ZIF could probably act as co-catalysis sites for promoting the transient

photocurrent response. Based on the triple-exponential fitting of the emission decay curves (Fig. S20),
11
the fluorescence lifetimes followed the sequence of TiO2 < TiO2/C < TiO2/C@Zn-ZIF-L <

of
ro
-p
re
lP

Fig. 2. (A) XRD patterns of (a) TiO2, (b) TiO2/C, (c) TiO2/C@Zn-ZIF-L, (d) TiO2/C@Co-ZIF-L and

(e) TiO2/C@ZnCo-ZIF-L. (B) Raman spectra of TiO2/C, TiO2/C@Zn-ZIF-L, TiO2/C@Co-ZIF-L,


na

and TiO2/C@ZnCo-ZIF-L. (C) C 1s XPS spectra of TiO2/C. (D) Ti 2p XPS spectra of TiO2, TiO2/C,

and TiO2/C@ZnCo-ZIF-L. (E) Co 2p XPS spectra of ZnCo-ZIF-L and TiO2/C@ZnCo-ZIF-L. (F)

CO2 adsorption isotherms of TiO2/C, Zn-ZIF-L, Co-ZIF-L, ZnCo-ZIF-L, and TiO2/C@ZIF-L


ur

measured at 298 K.
Jo

TiO2/C@Co-ZIF-L < TiO2/C@ZnCo-ZIF-L, which is in coincide with the trend of PL intensities (Fig.

S19). In addition, TiO2/C@ZnCo-ZIF-L has the longest excited-state lifetime, which is beneficial for

suppressing recombination of electron-hole pair in the TiO2/C@ZnCo-ZIF-L. To investigate the

charge transfer efficiency, the surface photovoltage (SPV) spectroscopy of these photocatalysts were

12
investigated. As shown in Fig. S21, the SPV positive response intensity orders are as follows: TiO2 <

TiO2/C < TiO2/C@Zn-ZIF-L < TiO2/C@Co-ZIF-L < TiO2/C@ZnCo-ZIF-L, which is in good

agreement with the above results. Therefore, the photogenerated charge separation of TiO2 could be

greatly improved after the decoration of ZnCo-ZIF-L. Furthermore, the electrochemical impedance

spectroscopy (EIS) illustrated the TiO2/C@ZnCo-ZIF-L (Fig. 3D) exhibited the smallest semicircular

diameter compared to TiO2/C, TiO2/C@Zn-ZIF-L, and TiO2/C@Co-ZIF-L (Fig. S21 and S23). The

result further confirmed the lowest electron-transfer resistance of TiO2/C@ZnCo-ZIF-L, being

of
favorable for the electron-hole generation and separation.

ro
-p
re
lP
na
ur

Fig. 3. (A) UV-vis diffuse reflectance spectra of TiO2/C, TiO2, ZnCo-ZIF-L, and
Jo

TiO2/C@ZnCo-ZIF-L. (B) ESR spectra of TiO2, TiO2/C, ZnCo-ZIF-L, and TiO2/C@ZnCo-ZIF-L.

(C) Photocurrent responses under chopped-light illumination of TiO2, TiO2/C, TiO2/C@Zn-ZIF-L,

TiO2/C@Co-ZIF-L, and TiO2/C@ZnCo-ZIF-L. (D) Electrochemical impedance spectra (EIS) of

TiO2/C and TiO2/C@ZnCo-ZIF-L.

13
The photocatalytic CO2 reduction over different catalysts was then evaluated without the

addition of any sacrificial reagents. The conversion of main product CO with the function of time as

shown in Fig. 4A. The pristine TiO2 and ZIF-L showed relative low CO production rate of 4.1 and

0.36 μmol h−1 g−1, respectively. A slightly increased CO production rate of 7.8 μmol h −1 g−1 was

achieved for TiO2/C. Obviously, the TiO2/C@ ZnCo-ZIF-L displays largely improved CO production

rate, owing to the efficient visible light harvesting and high CO2 adsorption/activation capacity of

MOF. In contrast to TiO2@ZnCo-ZIF-L with a CO production rate of 6.6 μmol h−1 g−1, the results

of
illustrated that graphitic carbon with superior electron mobility is favorable for accepting electron

ro
transfer between TiO2 and ZnCo-ZIF-L. In addition, the mixture of TiO2/ZnCo-ZIF-L showed poor

CO2 reduction ability (1.3 μmol h−1 g−1), further illustrated that the leaf-branch structure would afford
-p
abundant exposed active sites and strong electronic coupling for electron-hole separation (Fig. S24).
re
Remarkably, the CO evolution rate of TiO2/C@ZnCo-ZIF-L reached 28.6 μmol h−1 g−1 (Fig. 4A),

higher than those of TiO2/C@Co-ZIF-L (22.7 μmol h−1 g−1) and TiO2/C@Zn-ZIF-L (18.7 μmol h−1
lP

g−1). It can be seen from the results that the bimetallic Zn/Co-ZIF exhibits higher photocatalytic
na

activity than the single metal ZIF. In addition, the properties characterization results demonstrated

that the combination of both Co2+ and Zn2+ with synergistic effect could enhance the CO2 adsorption
ur

and light-harvesting (Fig. 3A, 3C, S20, and S21). As a result, the synergistic effect leads to an
Jo

improvement in catalytic performance. As shown in Fig. S27, the H2 yield in 10 h for the

TiO2/C@ZnCo-ZIF-L is 3.6 μmol g−1, which is much lower than that of TiO2 (12.3 μmol g−1). The

result shows that the proton reduction into H2 is efficiently inhibited on the TiO2/C@ZnCo-ZIF-L

catalyst, being probably related to its higher CO2 adsorption capacity. In addition, the photogenerated

holes were quenched by H2O to generate O2 (Fig. S25).


14
To make a fair comparison, the commercial TiO2 (P25) was tested. And the CO evolution rate of

0.4 μmol g−1 h−1 for P25 was obviously lower than that of the leaf-branch photocatalysts. The

apparent quantum efficiency (AQE) of TiO2/C@ZnCo-ZIF-L composites for photocatalytic CO

production was almost 70 times more than that of P25. Most importantly, the CO selectivity could

achieve 99% for TiO2/C@ZnCo-ZIF-L

( , where Ci is the mole of carbon species

i (CO, CH4, C2H4, and C2H6) and H2 in the products mixture, and C is the mole of total carbon

compounds and H2 produced). In contrast, a mixture of H2, CH4, CO, C2H6, and C2H4 was detected

for TiO2 (Fig. S26 and S27). These results indicated that the well integration of ZIF-L and graphitic

of
carbon on TiO2 would efficiently inhibit the formation of by-product. Subsequently, the in-situ FT-IR

was performed to gain an in-depth illustration of high selectivity. As shown in Fig. 4B, a weak band at

ro
2267 cm−1 for the TiO2/C@ZnCo-ZIF-L after absorption of CO2 was observed, which can be

assigned to the asymmetric stretching vibrations of CO2 molecules adsorbed on Zn/Co ions [61-63].
-p
It should be noted that the metal sites in ZIF-L acting as Lewis acid is favorable for binding and
re
activating the polar carbon-oxygen bonds of CO2, implying the existence of different active centers

for the activation of CO2 between TiO2/C@ZnCo-ZIF-L and TiO2. Moreover, in-situ FT-IR also was
lP

used to monitor the photocatalysis process. Before photocatalytic reaction, the monodentate

carbonate (m–CO32−, 1298, 1314, 1472, and 1541 cm−1), bidentate carbonate (b–CO32−, 1341, 1651

cm−1) and bicarbonate (HCO32−, 1238, 1396, and 1420 cm−1) was observed for TiO2/C@ZnCo-ZIF-L
na

(Fig. 4C). After illumination, two new peaks at 1356 and 1525 cm−1 assigned to COO− stretching

appeared [64-66]. It should be noted that the formation of COOH* is an important intermediate for
ur

the formation of CO. The results illustrate that the Lewis acid sites in ZIF-L are favorable for

adsorption/activation of CO2 into COOH*, contributing to the remarkable CO selectively.


Jo

The 13CO2 isotope-labeling experiments and other control tests were also conducted (Fig. 4D). As

summarized in Table S1, the results illustrated the products (CO and CH4) are only detected for the

TiO2/C@ZnCo-ZIF-L and ZnCo-ZIF-L with CO2 under light irradiation, confirming the CO is

produced by CO2 over TiO2/C@ZnCo-ZIF-L and ZnCo-ZIF-L photocatalyst. In addition, the signals
15
of the mass spectrum showed that the peak at m/z = 29 can be attributed to 13CO, firmly testifying the

produced CO generated from CO2 reduction but not from impurities or any other carbon source. The

photocatalytic stability of TiO2/C@ZIF-L was examined upon cycled photoreactions. Only a slight

decrease in CO generation was observed after three cycles (Fig. S28), illustrating the TiO2/C@ZIF-L

based fibrous mat can be easily reused without losing activity. In addition, no obvious change of XRD

patterns, SEM and HRTEM images, and Raman spectra (Fig. S29−S32) after three cycles were

observed, proving its high stability. These results thus illustrate that the fabricated

of
semiconductor/carbon@MOF composites can serve as promising photocatalysts toward CO 2

ro
reduction with highly efficient conversion and remarkable selectivity, as well as high recyclability.

-p
re
lP
na
ur
Jo

Fig. 4. (A) The CO production as a function of reaction time under dark and irradiation conditions on

TiO2, TiO2/C, TiO2/C@Zn-ZIF-L, TiO2/C@Co-ZIF-L, and TiO2/C@ZnCo-ZIF-L. In-situ FT-IR

spectra of TiO2/C@ZnCo-ZIF-L (B) before (blue curve) and after (red curve) absorption of CO2, (C)

before (blue curve) and in (red curve) photocatalytic CO2 reduction. (D) Gas chromatogram and mass

spectra of 13CO (m/z = 29) produced over TiO2/C@ZnCo-ZIF-L.

16
The theoretical calculations were performed to investigate the electronic structure of

TiO2/C@ZnCo-ZIF-L (Fig. S33). As shown in Fig. 5C, the partial density of states (PDOS) illustrated

that the forbidden zones of both TiO2 and ZIF-L in TiO2/C@ZnCo-ZIF-L became narrower compared

with individual TiO2 and ZIF-L (Fig. 6A–C, and S34), which was probably attributed to their

conjugation with carbon and the existence of electronic coupling. Afterward, the energy levels of the

conduction band minimum (ECBM) and valence band maximum (EVBM) for TiO2 and ZIF-L, together

with the Fermi level of carbon were calculated (Fig. S35 and S36). The calculated band gap energies

for TiO2 and ZnCo-ZIF-L were 3.036 eV and 1.808 eV, being consistent with experimental results

(Fig. S16 and S17). Most importantly, the conduction band (CB) and the valence band (VB) of TiO2

of
were respectively higher and lower than the corresponding bands of ZIF-L, indicating the dual

excitation paths in TiO2/C@ZnCo-ZIF-L with a Z-scheme heterojunction (Fig. 5D). Thus, under light

ro
irradiation, the photogenerated electrons on the TiO2 migrate to the VB of the ZIF-L via an electron

shuttle carbon, being favorable for the electrons further exciting to the CB of ZIF-L for efficient CO2
-p
reduction to CO. It should be noted that the electron-hole pairs cannot be effectively separated for the

type I heterojunction TiO2@ZnCo-ZIF-L photocatalyst (Fig. S37) because both electrons and holes
re
will accumulate at the CB and the VB levels of ZIF-L [67,68], is different from the electron transfer

path of TiO2/C@ZnCo-ZIF-L system. The reaction pathways of CO2 reduction to CO were also
lP

shown in Fig. 5E and Equations [69-72]:

(1)
na

(2)

(3)
ur

(4)
Jo

(5)

where the asterisks denote catalytically active sites. The CO2 is initially adsorbed on the Lewis acid

centers in ZIF-L and activated to *COOH intermediate. The formation of *COOH as the critical

intermediate for generation CO is then reduced to *CO intermediate. Finally, the CO* adsorbed on

17
the Lewis acid centers is converted into CO efficiently. In addition, the calculated adsorption energy

of CO on TiO2/C@ZnCo-ZIF-L was –1.08 eV, which was clearly lower than that on TiO2 (–3.21 eV)

and ZnCo-ZIF-L (–2.23 eV) (Fig. 5F). The results thus indicated that the binding energy of CO*

adsorbed on the TiO2/C@ZIF-L is too weak and the CO could be easily desorbed from the active sites,

contributing to a high CO selectivity.

of
ro
-p
re
lP
na
ur

Fig. 5. The PDOS of (A) TiO2, (B) ZnCo-ZIF-L, (C) TiO2, carbon, and ZIF-L in
Jo

TiO2/C@ZnCo-ZIF-L. Schematic illustration of (D) the electron-hole generation and separation

process and (E) main reaction pathway for the CO2 reduction on TiO2/C@ZnCo-ZIF-L under light

irradiation. (F) The CO adsorption energy on TiO2, ZnCo-ZIF-L, and TiO2/C@ZnCo-ZIF-L,

respectively.

18
4. Conclusions

In summary, a ternary semiconductor/carbon@MOF photocatalyst was fabricated to disclose

the important role of synergistic effect in high-performance photocatalytic CO2 reduction. The

resulting TiO2/C@MOF exhibits a significantly enhanced photocatalytic efficiency, selectivity, and

stability. The excellent performance could be attributed to rich exposed active sites, improved solar

energy harvesting, and high CO2 absorption capability afforded by the well-integrated leaf-branch

of
structure. Most importantly, the strong electronic coupling and suitable band matching for three

components could synergistically suppress the recombination of electron-hole pairs. Meanwhile, the

ro
Lewis sites existing in the MOF could promote the activated CO2 to form *COOH intermediate,
-p
being favorable for improving the CO selectivity. The developed strategy, as well as the mechanistic
re
insight, conveyed in our work will open up a new avenue to rationally design and fabricate other

composited photocatalysts for advanced energy conversion systems.


lP

Author contributions
na

Jian-Rong Li, Yibo Dou and Awu Zhou conceived and designed the experiments. Awu Zhou, Chen

Zhao and Jian Zhou mainly participated in catalysts synthesis and structural characterizations. Awu
ur

Zhou, Xue-Qian Wu and Chen Zhao performed the catalytic test. Yibo Dou and Awu Zhou performed
Jo

DFT calculations. Jian-Rong Li, Yibo Dou and Awu Zhou co-wrote the manuscript. All authors

discussed the results and commented on the manuscript.

Declaration of Interest Statement

19
All authors (Awu Zhou, Yibo Dou, Chen Zhao, Jian Zhou, Xue-Qian Wu, Jian-Rong Li) declare

NO affiliations with or involvement in any organization or entity with any competing interest in this

manuscript (A leaf-branch TiO2/Carbon@MOF composite for selective CO2 photoreduction).

Acknowledgements

We acknowledge financial support from the National Natural Science Foundation of China

of
(21606006, 51621003, 21771012), the Beijing Natural Science Foundation (2174064). We also

ro
thank the National Supercomputing Center in Shenzhen for providing computational resources.

References
-p
re
[1] I. Shown, S. Samireddi, Y. C. Chang, R. Putikam, P. H. Chang, A. Sabbah, F. Y. Fu, W. F.

Chen, C. I. Wu, T. Y. Yu, P. W. Chung, M. C. Lin, L. C. Chen, K. H. Chen, Nat. Commun.


lP

9 (2018) 169.

[2] J. Yu, L.-H. Xie, J.-R. Li, Y. Ma, J. M. Seminario, P. B. Balbuena, Chem. Rev. 117 (2017)
na

9674−9754.

[3] Q. Xia, H. Wang, B. Huang, X. Yuan, J. Zhang, J. Zhang, L. Jiang, T. Xiong, G. Zeng,
ur

Small 15 (2019) 1803088.

[4] X. Chang, T. Wang, J. Gong, Energy Environ. Sci. 9 (2016) 2177−2196.


Jo

[5] S. K. Dutta, S. Chakraborty, Ind. Eng. Chem. Res. 58 (2019) 8963−8978.

[6] A. A. Khan, M. Tahir, J. CO2 Util. 29 (2019) 205−239.

[7] S. Cao, J. Low, J. Yu, M. Jaroniec, Adv. Mater. 27 (2015) 2150−2176.

[8] M. Grätzel, Nature 414 (2001) 338.

[9] J. Ran, M. Jaroniec, S.-Z. Qiao, Adv. Mater. 30 (2018) 1704649.

[10] J. L. White, M. F. Baruch, J. E. Pander, III, Y. Hu, I. C. Fortmeyer, J. E. Park, T. Zhang, K.


20
Liao, J. Gu, Y. Yan, T. W. Shaw, E. Abelev, A. B. Bocarsly, Chem. Rev. 115 (2015)

12888−12935.

[11] S. Chen, H. Wang, Z. Kang, S. Jin, X. Zhang, X. Zheng, Z. Qi, J. Zhu, B. Pan, Y. Xie, Nat.

Commun. 10 (2019) 788.

[12] L. Cao, S. Sahu, P. Anilkumar, C. E. Bunker, J. Xu, K. A. S. Fernando, P. Wang, E. A.

Guliants, K. N. Tackett, II, Y.-P. Sun, J. Am. Chem. Soc. 133 (2011) 4754−4757.

[13] S. Wang, M. Xu, T. Peng, C. Zhang, T. Li, I. Hussain, J. Wang, B. Tan, Nat. Commun. 10

(2019) 676.

[14] H.-Q. Xu, J. Hu, D. Wang, Z. Li, Q. Zhang, Y. Luo, S.-H. Yu, H.-L. Jiang, J. Am. Chem.

of
Soc. 137 (2015) 13440−13443.

[15] J. Qin, S. Wang, X. Wang, Appl. Catal. B 209 (2017) 476−482.

ro
[16] A. Crake, K. C. Christoforidis, A. Gregg, B. Moss, A. Kafizas, C. Petit, Small 15 (2019)

1805473.

[17]
-p
W. G. Tu, Y. Zhou, Z. G. Zou, Adv. Mater. 26 (2014) 4607−4626.

[18] X. Li, J. Yu, M. Jaroniec, Chem. Soc. Rev. 45 (2016) 2603−2636.


re
[19] A. K. Singh, J. H. Montoya, J. M. Gregoire, K. A. Persson, Nat. Commun. 10 (2019) 443.

[20] M. Tahir, Appl. Surf. Sci. 467−468 (2019) 1170−1180.


lP

[21] A. Zhou, R.-M. Guo, J. Zhou, Y. Dou, Y. Chen, J.-R. Li, ACS Sustainable Chem. Eng. 6

(2018) 2103−2111.
na

[22] R. Li, J. Hu, M. Deng, H. Wang, X. Wang, Y. Hu, H. L. Jiang, J. Jiang, Q. Zhang, Y. Xie, Y.

Xiong, Adv. Mater. 26 (2014) 4783−4788.


ur

[23] K. Maeda, Adv. Mater. 31 (2019) 1808205.

[24] A. Dhakshinamoorthy, A. M. Asiri, H. Garcia, Angew. Chem. Int. Ed. 55 (2016)


Jo

5414−5445.

[25] S. N. Habisreutinger, L. Schmidt-Mende, J. K. Stolarczyk, Angew. Chem. Int. Ed. 52 (2013)

7372−7408.

[26] Y. Wu, H. Wang, W. Tu, S. Wu, J. W. Chew, Appl. Catal. B 256 (2019) 117810.

[27] J. Zhang, H. Wang, X. Yuan, G. Zeng, W. Tu, S. Wang, J. Photoch. Photo. C 38 (2019) 1−26.

[28] M. Tahir, Appl. Catal. B. 219 (2017) 329−343.


21
[29] S. Liu, F. Chen, S. Li, X. Peng, Y. Xiong, Appl. Catal. B 211 (2017) 1−10.

[30] H. Wang, Y. Sun, Y. Wu, W. Tu, S. Wu, X. Yuan, G. Zeng, Z. J. Xu, S. Li, J. W. Chew, Appl.

Catal. B 245 (2019) 290−301.

[31] H. Zhao, X. S. Wang, J. F. Feng, Y. N. Chen, X. Yang, S. Y. Gao, R. Cao, Catal. Sci. Technol.

8 (2018) 1288−1295.

[32] F. Wang, X. He, L. Sun, J. Chen, X. Wang, J. Xu, X. Han, J. Mater. Chem. A 6 (2018)

2091−2099.

[33] Z. H. Yan, M. H. Du, J. X. Liu, S. Y. Jin, C. Wang, G. L. Zhuang, X. J. Kong, L. S. Long, L.

S. Zheng, Nat. Commun. 9 (2018) 3353.

of
[34] S. Xie, Y. Wang, Q. Zhang, W. Deng, Y. Wang, ACS Catal. 4 (2014) 3644−3653.

[35] Y. X. Pan, Y. You, S. Xin, Y. Li, G. Fu, Z. Cui, Y. L. Men, F. F. Cao, S. H. Yu, J. B.

ro
Goodenough, J. Am. Chem. Soc. 139 (2017) 4123−4129.

[36] Y. Wang, P. Han, X. Lv, L. Zhang, G. Zheng, Joule 2 (2018) 2551−2582.

[37]
-p
B. Di Credico, M. Redaelli, M. Bellardita, M. Calamante, C. Cepek, E. Cobani, M.

D'Arienzo, C. Evangelisti, M. Marelli, M. Moret, L. Palmisano, R. Scotti, Catalysts 8 (2018)


re
353.

[38] F. Ke, L. Wang, J. Zhu, Nano Research 8 (2015) 1834−1846.


lP

[39] C. S. Diercks, Y. Liu, K. E. Cordova, O. M. Yaghi, Nat. Mater. 17 (2018) 301−307.

[40] Y. Shi, D. Yang, Y. Li, J. Qu, Z.-Z. Yu, Appl. Surf. Sci. 426 (2017) 622−629.
na

[41] S. Kim, S. K. Lim, Appl. Catal. B 84 (2008) 16−20.

[42] L. Tan, S. M. Xu, Z. Wang, Y. Xu, X. Wang, X. Hao, S. Bai, C. Ning, Y. Wang, W. Zhang, Y.
ur

K. Jo, S. J. Hwang, X. Cao, X. Zheng, H. Yan, Y. Zhao, H. Duan, Y. F. Song, Angew. Chem.

Int. Ed. 58 (2019) 2−10.


Jo

[43] Y. Wang, Z. Zhang, L. Zhang, Z. Luo, J. Shen, H. Lin, J. Long, J. C. S. Wu, X. Fu, X. Wang,

C. Li, J. Am. Chem. Soc. 140 (2018) 14595−14598.

[44] M. Wang, D. Wang, Z. Li, Appl. Catal. B 183 (2016) 47−52.

[45] L. Liu, H. Zhao, J. M. Andino, Y. Li, ACS Catal. 2 (2012) 1817−1828.

[46] W. Wang, J. Fang, S. Shao, M. Lai, C. Lu, Appl. Catal. B 217 (2017) 57−64.

[47] Y. Zhao, Y. Zhao, R. Shi, B. Wang, G. I. N. Waterhouse, L. Z. Wu, C. H. Tung, T. Zhang,


22
Adv. Mater. 31 (2019) 1806482.

[48] W. Tu, Y. Zhou, Q. Liu, S. Yan, S. Bao, X. Wang, M. Xiao, Z. Zou, Adv. Funct. Mater. 23

(2013) 1743−1749.

[49] H. Zhang, X. Lv, Y. Li, Y. Wang, J. Li, ACS Nano 4 (2010) 380−386.

[50] J. Zhou, Y. Dou, A. Zhou, L. Shu, Y. Chen, J.-R. Li, ACS Energy Lett. 3 (2018) 1655−1661.

[51] G. Lui, J.-Y. Liao, A. Duan, Z. Zhang, M. Fowler, A. Yu, J. Mater. Chem. A 1 (2013)

12255−12262.

[52] K. Alamelu, V. Raja, L. Shiamala, B. M. Jaffar Ali, Appl. Surf. Sci. 430 (2018) 145−154.

[53] S. Pu, R. Zhu, H. Ma, D. Deng, X. Pei, F. Qi, W. Chu, Appl. Catal. B 218 (2017) 208−219.

of
[54] X. Feng, F. Pan, H. Zhao, W. Deng, P. Zhang, H.-C. Zhou, Y. Li, Appl. Catal. B 238 (2018)

274−283.

ro
[55] J. Yang, Y. Guo, R. Jiang, F. Qin, H. Zhang, W. Lu, J. Wang, J. C. Yu, J. Am. Chem. Soc.

140 (2018) 8497−8508.

[56]
-p
R. Chen, J. Yao, Q. Gu, S. Smeets, C. Baerlocher, H. Gu, D. Zhu, W. Morris, O. M. Yaghi,

H. Wang, Chem. Commun. 49 (2013) 9500−9502.


re
[57] H. Jiang, S. Xue, Y. Liu, W. Xing, R. Chen, Micropor. Mesopor. Mat. 243 (2017) 16−21.

[58] Y. Wei, X. Wu, Y. Zhao, L. Wang, Z. Zhao, X. Huang, J. Liu, J. Li, Appl. Catal. B 236 (2018)
lP

445−457.

[59] N. Li, J. Liu, J. J. Liu, L. Z. Dong, Z. F. Xin, Y. L. Teng, Y. Q. Lan, Angew. Chem. Int. Ed.
na

58 (2019) 5226−5231.

[60] J. C. Cardoso, S. Stulp, J. F. de Brito, J. B. S. Flor, R. C. G. Frem, M. V. B. Zanoni, Appl.


ur

Catal. B 225 (2018) 563−573.

[61] M. Wang, J. Liu, C. Guo, X. Gao, C. Gong, Y. Wang, B. Liu, X. Li, G. G. Gurzadyan, L. Sun,
Jo

J. Mater. Chem. A 6 (2018) 4768−4775.

[62] C. Chizallet, S. Lazare, D. Bazer-Bachi, F. Bonnier, V. Lecocq, E. Soyer, A. A. Quoineaud,

N. Bats, J. Am. Chem. Soc. 132 (2010) 12365−12377.

[63] A. Zanon, S. Chaemchuen, B. Mousavi, F. Verpoort, J. CO2 Util. 20 (2017) 282−291.

[64] Z. Jiang, H. Sun, T. Wang, B. Wang, W. Wei, H. Li, S. Yuan, T. An, H. Zhao, J. Yu, P. K.

Wong, Energy Environ. Sci. 11 (2018) 2382−2389.


23
[65] T. Di, B. Zhu, B. Cheng, J. Yu, J. Xu, J. Catal. 352 (2017) 532−541.

[66] S. Kattel, W. Yu, X. Yang, B. Yan, Y. Huang, W. Wan, P. Liu, J. G. Chen, Angew. Chem. Int.

Ed. 55 (2016) 7968−7973.

[67] P. Zhou, J. Yu, M. Jaroniec, Adv. Mater. 26 (2014) 4920−4935.

[68] J. Low, J. Yu, M. Jaroniec, S. Wageh, A. A. Al-Ghamdi, Adv. Mater. 29 (2017) 1601694.

[69] X. Li, Y. Sun, J. Xu, Y. Shao, J. Wu, X. Xu, Y. Pan, H. Ju, J. Zhu, Y. Xie, Nat. Energy 4 (2019)

690−699.

[70] M. Tahir, B. Tahir, N. A. S. Amin, Z. Y. Zakaria, J. CO2 Util. 18 (2017) 250−260.

[71] M. Tahir, B. Tahir, N. A. S. Amin, Appl. Catal. B 204 (2017) 548−560.

of
[72] B. Tahir, M. Tahir, N. A. S. Amin, Appl. Catal. B 248 (2019) 167−183.

ro
-p
re
lP
na
ur
Jo

24

You might also like