Co Matel Paper
Co Matel Paper
Co Matel Paper
A R T I C LE I N FO A B S T R A C T
Keywords: The photocatalytic conversion of CO2 into chemical fuels represents a promising approach for solving the future
CO2 photoreduction energy crisis. However, the construction of a photocatalyst simultaneously integrating a photosensitizer and
COFs molecular cocatalyst with intramolecular electron delivery is challenging. Herein, we designed covalent organic
Electron delocalization frameworks (COFs) with excellent extended conjugation and potential embedded redox active sites. The full
Olefin-based
-C = C- bridging in sp2c-COFdpy creates and dredges the donor-acceptor channel for intramolecular electron
Single atom
delocalization and a cascade effect. Interestingly, CO2 photoreduction can be carried out in water, and the
optimized sp2c-COFdpy-Co exhibits the highest activity and stability among COFs without noble metal involve-
ment, achieving up to 17.93 mmol g−1 CO with 81.4 % selectivity in a long-range reaction. Theoretical calcu-
lations and experimental data suggest that the structural advantages enable excitons to facilely reach single Co
sites via the electron cascade, which provides a new concept in the nanoarchitecture of COFs for efficient CO2
photoreduction.
⁎
Corresponding author at: International Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS), 1-1 Namiki,
Tsukuba, Ibaraki 305-0044, Japan.
⁎⁎
Corresponding authors at: College of Science, Huazhong Agricultural University, Wuhan 430070, PR China.
E-mail addresses: [email protected] (S. Wang), [email protected] (H. Chen), [email protected] (J. Ye).
https://doi.org/10.1016/j.apcatb.2020.119096
Received 29 January 2020; Received in revised form 26 April 2020; Accepted 3 May 2020
Available online 07 May 2020
0926-3373/ © 2020 Elsevier B.V. All rights reserved.
Y. Xiang, et al. Applied Catalysis B: Environmental 274 (2020) 119096
long-range in-plane π-conjugation of COFs have been extensively stu- for efficient CO2 photoreduction.
died to maximize long-range order to achieve high charge carrier mo-
bilities [31–33]. Currently, the most frequently used strategy for as- 2. Experimental methods
sembling COF architectures is imine linkage via the Schiff base
condensation reaction [34,35]. According to the most recent reports by 2.1. Chemicals and reagents
Huang et al. and Zou et al. [36,37], two kinds of imine-connected COFs
involving 2,2′-dipyridine as a building block to provide active sites have All the chemicals and reagents were purchased in analytical purity
yet to reach significant advances in CO2 photoreduction. The co- from commercial sources and used directly without further purification.
ordination of photosensitizers and active sites in COFs combines the Synthesis of building blocks could be found in the supporting in-
advantages of both heterogeneous and homogeneous catalysis. How- formation.
ever, these delicately designed eN]Ce bonded COFs display relatively
poor stability and weak electron delocalization due to the partial π- 2.2. Synthesis of sp2c-COFdpy
conjugation, which strictly blocks the delocalization of photoexcited
electrons and inhibits activity. As a consequence, additional molecular A 10 ml pyrex tube was charged with TFPPY (30 mg, 0.048 mmol)
noble metal complexes are inevitably introduced as photosensitizers or BPYDAN (22.7 mg, 0.096 mmol), 1,4-dioxane (1 mL) and aqueous KOH
active sites to facilitate the activity of CO2 photoreduction. As an al- solution (0.1 mL, 4 M), and then the suspension was sonicated for
ternative, it is anticipated that sp2 carbon olefin linkages will propagate 5 min, degassed thoughy with three freeze-pump-thaw cycles (liquid
the π-conjugation across the skeleton, yielding an ideal choice. Never- nitrogen bath), sealed under vacuum and heated at 110 °C for three
theless, it is extremely difficult to construct COFs with full eC]Ce days. After cooling down, the solid was collected by centrifugation,
bridging due to the low reversibility of linkage formation [31,38–40]. washed with water, THF successively until the filtrate was colorless.
Thanks to the pioneering works of Jiang et al. and Feng et al. [41,42], The resulting red power was dried at 60 °C under vacuum overnight to
sp2-carbon-linked skeleton COFs (olefin-based COFs) were successfully afford the crystalline sp2c-COFdpy (46 mg, 93 %). Elemental analysis:
designed by using Knoevenagel reactions. Such kinds of extended π- Anal. Calcd for (C72H38N8)n: C, 85.19; H, 3.77; N, 11.04. Found: C,
conjugation sp2c-COFs could offer enhanced electron delocalization, 73.22; H, 5.02; N, 8.49.
most likely affecting their light absorption and photocatalytic activities,
yet little is currently known about this behavior [43–48]. Bearing this 2.3. General procedure for synthesis of sp2c-COFdpy-M
in mind, our aim is to simultaneously achieve the desired conjugation
and refined crystallinity, as well as oriented electron delocalization, The suspension of sp2c-COFdpy (101 mg, 0.2 mmol theoretical
within well-designed olefin-based COFs for CO2 photoreduction. bounding site) in CH3OH containing MCl2 (M = Fe, Co, Ni, Cu)
Hereby, we demonstrate the clever design and novel features of (0.2 mmol) was refluxed overnight under N2 protection, and the cooled
sp2c-COFs (sp2c-COFdpy) with excellent extended aromatic conjugation precipitate was collected by filtration, washed with water and CH3OH
and potential embedded redox active sites via simple base-catalyzed thoroughly for removal of unbound metal salts, followed by drying at
Knoevenagel polycondensation (Scheme 1). With 3,6,8-tetrakis(4-for- 60 °C under vacuum overnight. As a result, sp2c-COFdpy-Fe, sp2c-
mylphenyl)pyrene (TFPPy) and 2,2′-([2,2′-bipyridine]-5,5′-diyl)diace- COFdpy-Co, sp2c-COFdpy-Ni, sp2c-COFdpy-Cu were afforded, and the
tonitrile (BPyDAN) acting as the donor and acceptor, respectively, this content of Fe, Co, Ni, Cu was determined to be 2.00 %, 3.25 %, 2.87 %,
fully π-conjugated system achieves intramolecular electron delivery via 3.48 % by ICP-MS, respectively. In addition, the content of sp2c-COFbp-
the cascade effect by using push-pull modes. Taking advantage of the Co was determined to be 0.05 %.
embedded bipyridine, a series of non-noble metals (Fe, Co, Ni, and Cu)
could be anchored by sp2c-COFdpy to provide catalytic active sites for 2.4. Photocatalytic CO2 reduction
CO2 photoreduction. The optimized sp2c-COFdpy-Co featuring single
atomic Co sites shows extremely stable performance in CO2 conversion The 20 mg COF sample, 5 ml triethanolamine (TEOA) and 45 ml
to CO, higher than that of other COFs studied to date; the conversion water (H2O) solution were dispersed in a Pyrex glass cell which was
can reach 0.99 mmol g−1 h−1 without adding any adventitious photo- connected to a gas-closed circulation system. Prior to irradiation, the
sensitizer or cocatalyst in water under visible light illumination. The suspension was evacuated and then backfilled with CO2 several times
results also provide a new point of view in the nanoarchitecture of COFs until air was removed completely. After the system with a magnetic bar
Scheme 1. (a) Synthesis of sp2c-COFdpy and (b) sp2c-COFdpy-Co. (c) Schematic CO2 photoreduction on the as-prepared sp2c-COFdpy-Co.
2
Y. Xiang, et al. Applied Catalysis B: Environmental 274 (2020) 119096
stirring was purged with high-purity CO2 (99.999 %), the 300 W Xe at 3.16, 4.73, 6.32, 9.47, 14.23 17.79 and 24.84°, corresponding to the
lamp (PLS-SXE300D, Beijing Perfectlight and PEC-L01 Peccell reflections from the (110), (020) (220), (040), (060), (370) and (001)
Technologies, Inc) equipped with a 420 nm cut-off filter was turned on facets, respectively, which is similar to the as-prepared sp2c-COFdpy
for the light source. During the reaction, the Pyrex cell was kept at the [43]. Via theoretical simulations using Materials Studio, the structure of
room temperature by circulating water through a metal jacket, and the sp2c-COFdpy was built, and Pawley refinements of the PXRD patterns
amount of evolved H2 was analyzed by an online gas chromatograph were carried out; the simulated results are in agreement with the
(GC-8A, Shimadzu Corp., Japan) with a TCD detector; the amount of abovementioned experimental results, as revealed by a difference plot,
evolved CO was analyzed by an offline gas chromatograph (GC-2014, with Rp = 1.81 % and Rwp = 2.27 % for sp2c-COFdpy-Co and Rp = 3.08
Shimadzu Corp., Japan) with a nickel conversion furnace and FID de- % and Rwp = 4.11 %, for sp2c-COFdpy respectively, suggesting the va-
tector. For the recycling measurements, the COFs after the reaction was lidity of the computational model of eclipsed AA stacking [43]. In ad-
recovered by centrifugation and washing with water, and the sample dition, the PXRD patterns obtained reference sample of sp2c-COFbp and
after drying in vacuum at 60 °C was used further for the cycling ex- the other sp2c-COFs loaded with Fe, Ni and Cu ions (Fig. S1) retained a
periments. The average intensity of irradiation was calibrated by a crystalline structure similar to that of sp2c-COFdpy [36]. The high
spectroradiometer (AvaSolar‐1, Avantes, America). The AQY measure- crystallinity and structural retention during the process of metal ion
ment at specific monochromatic wavelength was calculated based on loading can also be revealed by electron microscopy. The scanning
the following equation: AQY for CO evolution = N(CO) × 2 / N(pho- electron microscopy (SEM) and transmission electron microscopy
tons) × 100 %. In situ DRIFTS measurement were carried out by FT-IR (TEM) images (Fig. 1b and Figs. S2 and S3) of sp2c-COFdpy exhibit
spectrometer (Nicolet iS50 Thermo Scientific, USA) with a designed uniform morphologies with a large area of nanofibers whether with
reaction cell. (sp2c-COFdpy-Co, sp2c-COFdpy-Fe, sp2c-COFdpy-Ni and sp2c-COFdpy-Cu)
or without (sp2c-COFdpy) metal ions, while sp2c-COFbp is different, at-
3. Results and discussion tributed to the change in growth direction caused by the embedded
bipyridine in sp2c-COFdpy. The diameter of the nanofibers was de-
2D tetragonal sp2c-COFdpy was first synthesized by the base-cata- termined by TEM to be ca. 150 nm, and high-resolution TEM (HRTEM)
lyzed Knoevenagel condensation reaction by choosing TFPPy and images showed an almost identical lattice in both sp2c-COFdpy and sp2c-
BPyDAN as the precursors. As the low reversibility of C]C bond for- COFdpy-Co, revealing a pore dimeter of ca. 2.0 nm in these COFs (Fig. 1b
mation poses a serious challenge to high crystallinity, the solvothermal and Fig. S4). Moreover, the FFT also verified the lattice structures with
conditions were optimized and screened carefully. Finally, crystalline a pore size of 2 nm [49]. Moreover, aberration-corrected high-angle
sp2c-COFdpy was successfully prepared by using KOH catalyst at 110 °C annular dark-field (AC-HAADF) imaging demonstrated that the bipyr-
for 3 days, and only 1,4-dioxane was used as the solvent. The crystalline idine anchored the single atomic Co sites in sp2c-COFdpy-Co, as clearly
nature of COFs is a unique feature compared to other conjugated shown in Fig. 1b. Isolated Co sites, appearing as individual bright spots
polymers. Powder X-ray diffraction (PXRD) patterns were first em- with an average size of ca. 0.2 nm, are well dispersed, and the energy-
ployed to elucidate the high crystallinity of as-prepared sp2c-COFs. As dispersive X-ray (EDX) mapping images reveal that Co, N and C are
shown in Fig. 1a, sp2c-COFdpy-Co exhibited prominent diffraction peaks distributed homogeneously across the nanofibers [50]. To further
Fig. 1. (a) Experimentally observed PXRD pattern, refined pattern, difference between the experimental of sp2c-COFdpy-Co and sp2c-COFdpy and calculated patterns
for AA stacking. (b) The SEM image, TEM image, HR-TEM image with lattice, aberration-corrected HAADF-STEM and EDX elemental mapping of sp2c-COFdpy-Co. (c)
Fourier transformed k3-weighted χ(k)-function of the EXAFS spectra for Co K-edge and corresponding EXAFS fitting curve for sp2c-COFdpy-Co. (d) Solid-state 13C CP/
MAS NMR spectroscopy and (e) FT-IR spectroscopy of sp2c-COFbp, sp2c-COFdpy and sp2c-COFdpy-Co. (f) Solid-state UV/Vis DRS spectra of sp2c-COFdpy and sp2c-
COFdpy-Co with image of corresponding COFs dispersed in water. Inset: Solid-state UV/Vis DRS spectra of sp2c-COFbp and sp2c-COFbp-Co.
3
Y. Xiang, et al. Applied Catalysis B: Environmental 274 (2020) 119096
Fig. 2. (a) Photocatalytic CO evolution performance as a function of time (b) and selective of photocatalytic CO2 reduction over sp2c-COFdpy-Co, sp2c-COFdpy, sp2c-
COFbp-Co and sp2c-COFbp under visible light (420 nm cut-off filter) in water using triethanolamine (TEOA) as sacrificial agent. (c) Recyclability tests for sp2c-COFdpy-
Co in selective photoreduction of CO2 over 18 h each time. (d) Gas chromatography-mass spectra (GC–MS) of 13CO generated in the photocatalytic reduction of 13CO2
over sp2c-COFdpy-Co. The left inset shows the total ion chromatography contains O2, N2 and CO, and right inset show the standard spectra of CO in mass spectra. (e) A
long-term photocatalytic CO evolution performance as a function of time over sp2c-COFdpy-Co, sp2c-COFdpy-Ni, sp2c-COFdpy-Fe and sp2c-COFdpy-Cu.
examine the coordination configuration of single Co sites, X-ray ab- at 1475 cm−1 in the bipyridine skeleton. Moreover, no apparent peaks
sorption fine structure (XAFS) measurements were performed (Fig. S5), at 2815 cm−1 and 2720 cm−1, attributed to aldehyde groups in pre-
and Fig. 1c presents the extended X-ray absorption fine structure cursors, were detected after the polycondensation reaction, verifying a
(EXAFS) analysis of Co foil, Co(bpy)32+ and sp2c-COFdpy-Co. The high degree of polymerization [41]. No obvious difference was found
dominant peak in Co(bpy)32+ is attributed to the Co-N bond located at before and after metal coordination, which also explains the retention
1.59 Å, which is shorter than the Co-Co signal at 2.16 Å in Co foil. Si- of the main structure. Because light absorption is used to determine
milar to Co(bpy)32+, a peak located at 1.61 Å assigned to the Co-N/O whether a material can be used as a photocatalyst, UV–vis diffuse re-
bond can be observed, while no Co-Co peak can be found, corroborating flectance spectroscopy (DRS) was used to evaluate the band gap and
the atomic dispersion of single Co sites in sp2c-COFdpy-Co [51]. How- absorption edge of the as-prepared COFs (Fig. 1f and Fig. S7). sp2c-
ever, the fitting of the EXAFS spectra (Table S1) implied that the co- COFbp exhibits an absorbance edge at approximately 597 nm (inset of
ordination numbers of the Co are six, much higher than estimated two. Fig. 1f), whereas the absorbance edge of sp2c-COFdpy is redshifted to
As CeO and CeN are not distinguishable by EXAFS analysis, and it was 640 nm (Fig. 1f). According to the Kubelka-Munk equation (Fig. S8), the
inferred that each Co2+ is coordinated with two N atoms from 2,2′- absorption band edge of sp2c-COFdpy corresponds to a bandgap of
bipyridine ligands and four O atoms from coordinated water molecules, 2.03 eV, which is smaller than that of sp2c-COFbp (2.17 eV). This phe-
thus giving rise to a stabilized octahedral configuration [52]. According nomenon can be attributed to the electron delivery achieved by in-
to the fitting results, providing potential space for photocatalytic CO2 tramolecular push-pull interactions using bipyridine instead of the bi-
reduction (inset of Fig. 1c and Table S1). phenyl skeleton as an electron acceptor. Interestingly, the absorbance of
Subsequently, the exact structure of the as-prepared COFs was fur- sp2c-COFdpy-Co showed a noteworthy extension to 733 nm with Co
ther confirmed by spectral analysis. As seen in the solid-state 13C-(CP- loading (with a bandgap of 1.83 eV), which could be observed directly
MAS) NMR spectra (Fig. 1d), the chemical shifts of approximately by a change in color, due to the induced Co ions anchored by the em-
121 ppm in sp2c-COFdpy and sp2c-COFdpy-Co could be assigned to the bedded bipyridine both enhancing electron delocalization in the donor
embedded bipyridine rings according to the estimated chemical shifts of parts and promoting aggregation in the acceptor parts [53]. Despite the
different carbons (Fig. S6). The characteristic peaks at 127 ppm and significant changes in the absorbance and bandgap of sp2c-COFdpy after
140 ppm only appeared in sp2c-COFbp and were attributed to the carbon metal coordination, sp2c-COFbp retained the original absorbance and
skeleton of biphenyl. After metal ions were incorporated into sp2c- bandgap due to the lack of interaction between the involved metals and
COFdpy, the signals at 153 ppm almost disappeared, showing high ac- the biphenyl skeleton in sp2c-COFbp-Co (inset of Fig. 1f). Cyclic vol-
cordance with a previous report, which may be the result of the dif- tammetry (CV) measurements were also conducted to determine the
ferent chemical environments after metal coordination [49]. Moreover, band positions; the different positions of the oxidation peaks due to the
the success of C]C bond formation was confirmed by Fourier transform irreversible oxidation process of the COFs at the applied voltage (Fig.
infrared spectroscopy (FTIR) (Fig. 1c). A characteristic vibration band S9) revealed different energy levels (Table S2) [54]. As a result, all of
of C^N at 2210 cm−1 (vinylene CN) appeared for both sp2c-COFdpy these as-prepared COFs had enough of a negative potential to carry out
and sp2c-COFbp, which showed a slight shift in comparison to the pre- the reduction of CO2 to CO [55].
cursors, and sp2c-COFdpy showed an additional stretching mode of C]N
4
Y. Xiang, et al. Applied Catalysis B: Environmental 274 (2020) 119096
Based on the confirmed structures of the as-prepared COFs, we 32 (Fig. S15), respectively. The other nonnegligible peak in the total ion
speculate that the intramolecular electron delivery via electron cascade chromatograph (left inset of Fig. 2d) is located at 9.5 min, and the
achieved in COFs could promote the CO2 photoreduction properties of corresponding mass spectrum shows a main peak at m/z = 29 with a
sp2c-COFbp-Co. As a proof of concept, experiments of photocatalytic high abundance of 96.3 % (base peak/total ion current). Moreover, the
CO2 reduction were carried out in a closed gas circulation system in fragments produced from this peak (13C at m/z = 13 and O at m/
which only the COFs were added as photocatalysts, and triethanola- z = 16) can also be distinguished in the mass spectrum, which shows
mine (TEoA) proved to be the most effective sacrificial agent to enhance high accordance with the standard mass spectrum of CO (right inset of
the CO2 photoreduction activities over triethylamine (TEA) and as- Fig. 2d), verifying that this evident peak could be attributed to the
corbic acid (Fig. S10) [56]. Interestingly, compared to most reported generated CO. This result also provides solid evidence confirming that
polymeric photocatalysts, these olefin-based COF powder samples can the produced 13CO indeed originated from 13CO2 in the process of CO2
be readily dispersed in water by ultrasound sonication, presumably due photoreduction over sp2c-COFdpy-Co [56]. Subsequently, the long-term
to their inherent hydrophilicity caused by the Knoevenagel condensa- photocatalytic CO2 reduction properties of sp2c-COFdpy chelated with
tion reaction generating cyanide. This hydrophilicity is beneficial to different metal ions (Co, Ni, Fe and Cu) were determined under the
water wetting and permeation, thus promoting photocatalytic CO2 re- same conditions. As shown in Fig. 2e, sp2c-COFdpy-Fe and sp2c-COFdpy-
duction over the COFs when using pure water as a solvent. As displayed Cu did not show good photocatalytic activity for CO2 reduction, par-
in Fig. 2a, after 6 h of visible light (> 420 nm) irradiation, similar CO ticularly over sp2c-COFdpy-Cu, whose activity was hardly improved
generation amounts were detected when sp2c-COFbp and sp2c-COFbp-Co with respect to that of sp2c-COFdpy. In contrast, sp2c-COFdpy-Co and
were used as the photocatalyst: 0.06 mmol g−1 (0.01 mmol g−1 h−1) sp2c-COFdpy-Ni showed significant enhancements, achieving CO
and 0.07 mmol g−1 (0.01 mmol g−1 h−1), respectively. sp2c-COFdpy amounts of 17.93 mmol g−1 and 8.15 mmol g−1 after long-term reac-
generated 0.17 mmol g−1 CO with an average generated rate of tion, respectively, and sp2c-COFdpy-Co exhibited the best activity in CO2
0.03 mmol g−1 h−1. When sp2c-COFdpy-Co was applied as the photo- photoreduction among the abovementioned olefin-based COFs.
catalyst, a maximum of 6.01 mmol g−1 CO was recorded by gas chro- After scrutinizing the CO2 photoreduction activities over olefin-
matography, achieving an average generated rate of 1.00 mmol based COFs, it is of great significance and importance to unveil the
g−1 h−1, and showing approximately 36- and 100-fold enhancement underlying reason and the hidden mechanism triggering the enhance-
compared to sp2c-COFdpy and sp2c-COFbp-Co, respectively; this value is ment in CO2 photoreduction. The adsorption of CO2 is a prerequisite for
the highest reported for photocatalytically active COFs without invol- CO2 photoreduction. Although sp2c-COFdpy exhibited a larger
ving any unstable noble metal dye (Table S3). Therefore, we can rea- Brunauer-Emmett-Teller (BET) surface area of 571.6 m2 g−1 than sp2c-
sonably deduce that this high-efficiency and remarkably enhanced CO2 COFbp (156.7 m2 g−1) (Fig. S16), sp2c-COFbp showed an appreciable
photoreduction in water over sp2c-COFdpy-Co is mainly influenced by amount of CO2 adsorption, higher than that of sp2c-COFdpy (Fig. 3a). In
the intrinsic structural characteristics and anchored Co sites facilitating addition, no significant change in CO2 capture amount was observed
intramolecular electron delivery. In addition, the selectivity of olefin- between sp2c-COFbp and Co-loaded sp2c-COFbp-Co. Although the in-
based COFs for CO2 photoreduction was examined, and no significant terior cavities are partially filled by the introduced Co, sp2c-COFdpy-Co,
selectivity or product evolution rate were observed over sp2c-COFbp with a smaller surface area of 192.8 m2 g-1, shows better ability to ad-
regardless of whether Co was loaded. sp2c-COFdpy exhibited the lowest sorb CO2 than sp2c-COFdpy, which can be attributed to enhanced che-
selectivity in CO evolution (1.5 %), whereas a large amount of H2 misorption of Lewis acid-base interactions between the loaded Co ions
(0.67 mmol g−1 h−1) was generated due to the embedded bipyridine. and absorbed CO2 molecules (Fig. S17). According to previous reports,
This constructed nitrogen heterocyclic ring provides an active site for olefin-based COFs exhibit a combination of desired conjugation and
the completion of the hydrogen evolution reaction. When Co ions are simultaneously refined crystallinity, facilitating the long-range deloca-
added during the synthetic process, the Co sites in sp2c-COFdpy-Co lization of electrons [45]. Moreover, with this latent electron delocali-
promote electron transfer from the nitrogen atoms to the chelated metal zation, the electron donor and acceptor behavior of olefin-based COFs
and consequently act as a bridge to inject the aggregated electrons into could provide a driving force for intramolecular electron delivery via
CO2 molecules, thus yielding the highest selectivity of 81.4 % in CO electron cascade to adapt to multielectron CO2 photoreduction. To
generation (Fig. 2b) An apparent quantum yield (AQY) of 1.2 % was verify the established intramolecular electron delivery mechanism via
determined for sp2c-COFdpy-Co at 420 nm single-wavelength irradiation electron cascade to active sites, femtosecond transient absorption (TA)
(Fig. S11), and cyclic testing indicated that after 5 cycles, the CO was employed to study the ultrafast charge transfer in COFs. A photo-
evolution was still high (99.1 % with respect to the initial cycle) induced absorption peak is observed at 5000 nm, which is attributed to
compared to the initial values (Fig. 2c), attributed to the extended the trapped electrons in the COF. Trapped electrons generated via the
conjugation and adequate stability obtained by using bipyridine instead ultrafast intramolecular charge diffusion process are regarded as the
of biphenyl in the skeleton of sp2c-COFdpy, which could be further main excitons that drive CO2 photoreduction. Therefore, we compared
confirmed by the TGA measurement (Fig. S12) [47]. The morphology the dynamics of photoinduced absorption at 5000 nm and the fitted
and chemical crystallinity of sp2c-COFdpy-Co before and after irradia- biexponential decay over sp2c-COFdpy and sp2c-COFdpy-Co. As shown in
tion was further compared by SEM and XRD (Figs. S13 and S14), and no Fig. 3b, sp2c-COFdpy-Co exhibits increased τ1 (2.15 ps) and τ2 (76.89 ps)
obvious difference was observed, verifying excellent recyclability and values compared to the τ1 (0.77 ps) and τ2 (29.12 ps) values in sp2c-
photocorrosion-resistance of as prepared C = C bridged COFs. More- COFdpy, respectively. A much longer average lifetime was observed in
over, the Co turnover number (TONCo) of sp2c-COFdpy-Co was 54 in sp2c-COFdpy-Co (40.39 ps) than in sp2c-COFdpy (8.89 ps) during the
18 h, according to the ICP-OES Co content of 0.33 mmol g−1, thus decay process, suggesting that the incorporation of Co not only inhibits
confirming the single-site sp2c-COF photocatalytic CO2 reduction charge recombination but also promotes intramolecular electron de-
system in water as a solvent without the assistance of noble metal-based livery (Table S4) [37]. To gain more insight into this intramolecular
cocatalysts or dyes. electron delivery, we examined the aggregation of electron density by
To validate the outstanding generation of CO originating from the theoretical calculations. The results were exactly as predicted: the
catalytic splitting of CO2, we employed high-purity isotope-labeled electrons are partially concentrated on the N-substituted aromatic ring,
carbon dioxide (13CO2) as a substitute source gas to further evaluate and the introduction of Co can obviously strengthen the electron den-
CO2 photoreduction over sp2c-COFdpy-Co. As shown in Fig. 2d, the fully sity distribution, explaining that enhanced intramolecular electron de-
distinct total ion chromatographic peaks at approximately 2.2 and 3.2 livery indeed occurs from the COF skeleton to metal sites (Fig. S18).
(left inset of Fig. 2d) can be assigned to O2 and N2 because the mass After the aggregation of long-range delocalized electrons, the
spectra recorded at 2.2 and 3.2 exhibit primary peaks of m/z = 28 and smooth injection of electrons into absorbed CO2 determines subsequent
5
Y. Xiang, et al. Applied Catalysis B: Environmental 274 (2020) 119096
Fig. 3. (a) CO2 adsorption capacities of the sp2c-COFdpy-Co, sp2c-COFdpy, sp2c-COFbp-Co and sp2c-COFbp at 273.15 K. (b) Kinetics of electrons in transient absorption
over sp2c-COFdpy with or without Co under a probe wavelength of 5000 nm. (c) Time-resolved photoluminescence decay of sp2c-COFdpy with or without Co (inset)
under a CO2 or argon atmosphere. (d) In-situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) for the photocatalytic reduction of CO2 over sp2c-
COFdpy-Co. (e) DFT-calculated relative Gibbs free energy profiles for the conversion of CO2 into CO. (f) The proposed process of mechanism over sp2c-COFdpy-Co for
the CO2 photoreduction reaction.
CO2 photoreduction. Consequently, time-resolved photoluminescence activate H2O molecules generating H2. Additionally, the stability of
(TR-PL) spectroscopy was carried out under two atmospheres (argon as *COOH intermediate was closest to 0 eV on Co site, and cleavage to
an inert atmosphere and CO2 as a reactive atmosphere) to distinguish generate CO* downhills. After Co loading, the free energy barrier of
the difference in the process of electron transfer from the COFs to CO2. sp2c-COFdpy-Co limited by CO desorption is 0.60 eV, which showed a
For sp2c-COFdpy-Co (Fig. 3c), the average lifetime of electrons in the significant decrease in the energy barrier of CO2 reduction compared to
CO2 atmosphere was significantly shortened compared to that in the the sp2c-COFdpy-Co. In absence of metal, the barrier of CO2 reduction
argon atmosphere, which can be attributed to the fact that electrons over sp2c-COFdpy is 1.36 eV which is difficult for the activation of CO2.
with a relatively long lifetime (τ2) can be smoothly delivered to CO2 Moreover, the CO desorption step was also the rate-determining step
(Table S5). However, for sp2c-COFdpy without Co and sp2c-COFdp with over Fe and Ni site, with the barrier of 1.22 eV and 0.89 eV which is
or without Co, all of the abovementioned COFs showed no difference in higher than that over Co site. In the case of loading Cu site, the carbon-
TR-PL measurements between CO2 and argon atmospheres due to the based compound was weakened adsorbed, and the first hydrogenation
weak delivery of electrons to CO2 (inset of Fig. 3c and Fig. S19), which step was the rate limiting step, thus determining the barrier of 1.15 eV.
results in the disparate activation of CO2 over these olefin-based COFs. Therefore, the relative Gibbs free energy over olefin-based COFs loaded
Moreover, we also used in situ diffuse reflectance infrared Fourier with different metals (Fe, Co, Ni, or Cu) for CO2 reduction showed the
transform spectroscopy (DRIFTS) to unveil the process of CO2 photo- expected trend of sp2c-COFdpy-Co > sp2c-COFdpy-Ni > sp2c-COFdpy-
reduction occurring on sp2c-COFdpy-Co. Before light irradiation, the Fe > sp2c-COFdpy-Cu, which is consistent with the measurements of
peak located at 1631 cm−1 for sp2c-COFdpy-Co is assigned to the CO2 photoreduction activity (Fig. S20). Together, these results ration-
asymmetric stretching vibrations of surface-bound carboxylate gener- ally point to a mechanistic pathway of CO2 photoreduction over sp2c-
ated from chemically absorbed CO2. In the spectra collected after light COFdpy-Co, which is promoted by the intramolecular cascaded electron
irradiation, the main intermediate peaks of this process located at delivery. As shown in Fig. 3f, upon visible light illumination, photo-
1732 cm−1 and 1580 cm−1 significantly increased over the irradiation electrons were generated and quickly delocalized to other parts of the
time and were assigned to the asymmetric stretching vibrations of COO- sp2c-COFdpy-Co with the assistance of full -C = C- bridging. Since bi-
and COOH species (Fig. 3d), respectively, which is solidly in accordance pyridine was employed instead of embedding biphenyl as the COF
with the mechanism of CO2 conversion to CO in the previous report skeleton and subsequently chelated with Co, the electron donor and
[57]. To gain insights in the details of the difference reactivity of metal acceptor behavior of sp2c-COFdpy-Co was enhanced and facilitated in-
atoms in sp2c-COFdpy, periodic density functional theory (DFT) calcu- tramolecular electron delivery via electron cascade to the active sites.
lations were carried out. The structure of sp2c-COFdpy-Co, sp2c-COFdpy- Under strong electron flow, the initial Co(II) species sustained the cir-
Ni, sp2c-COFdpy-Fe, and sp2c-COFdpy-Cu was modeled under light ex- culation of Co(II) and Co(I), supporting the desired reduction of CO2 to
citation. The CO2 reduction reaction was occurred through absorbed CO via a proton-electron coupling pathway [58,59]. The oriented
CO2 (CO2*) and COOH (COOH*) intermediate to give CO, which was electron delivery resulting from the electron cascade through smooth
captured by the DRIFTS. The reaction free energy diagrams of the CO2 intramolecular channels thus achieves an efficient COF-based system
reduction and H2 evolution were plotted in Fig. 3e. For H2 evolution, for CO2 photoreduction.
compare to the sp2c-COFdpy-Co, it is much easier for sp2c-COFdpy to
6
Y. Xiang, et al. Applied Catalysis B: Environmental 274 (2020) 119096
7
Y. Xiang, et al. Applied Catalysis B: Environmental 274 (2020) 119096
M.A. Zwijnenburg, R.S. Sprick, A.I. Cooper, Sulfone-containing covalent organic heterogeneous photocatalysis, Angew. Chem. Int. Ed. 58 (2019) 6430–6434.
frameworks for photocatalytic hydrogen evolution from water, Nat. Chem. 10 [47] S.Q. Xu, G. Wang, B.P. Biswal, M. Addicoat, S. Paasch, W.B. Sheng, X.D. Zhuang,
(2018) 1180–1189. E. Brunner, T. Heine, R. Berger, X.L. Feng, A nitrogen-rich 2D sp2-carbon-linked
[33] P. Pachfule, A. Acharjya, J. Roeser, T. Langenhahn, M. Schwarze, R. Schomaecker, conjugated polymer framework as a high-performance cathode for lithium-ion
A. Thomas, J. Schmidt, Diacetylene functionalized covalent organic framework batteries, Angew. Chem. Int. Ed. 58 (2019) 849–853.
(COF) for photocatalytic hydrogen generation, J. Am. Chem. Soc. 140 (2018) [48] Y. Zhao, H. Liu, C. Wu, Z. Zhang, Q. Pan, F. Hu, R. Wang, P. Li, X. Huang, Z. Li, Fully
1423–1427. sp2-carbon conjugated two-dimensional covalent organic frameworks as artificial
[34] V.S. Vyas, F. Haase, L. Stegbauer, G. Savasci, F. Podjaski, C. Ochsenfeld, B.V. Lotsch, photosystem I with unprecedented efficiency, Angew. Chem. Int. Ed. 58 (2019)
A tunable azine covalent organic framework platform for visible light-induced 5376–5381.
hydrogen generation, Nat. Commun. 6 (2015) 8508. [49] M. Bhadra, H.S. Sasmal, A. Basu, S.P. Midya, S. Kandambeth, P. Pachfule,
[35] X. Zhao, P. Pachfule, S. Li, T. Langenhahn, M. Ye, G. Tian, J. Schmidt, A. Thomas, E. Balaraman, R. Banerjee, Predesigned metal-anchored building block for in situ
Silica-templated covalent organic framework-derived Fe–N-doped mesoporous generation of Pd nanoparticles in porous covalent organic framework: application
carbon as oxygen reduction electrocatalyst, Chem. Mater. 31 (2019) 3274–3280. in heterogeneous tandem catalysis, ACS Appl. Mater. Interfaces 9 (2017)
[36] W. Zhong, R. Sa, L. Li, Y. He, L. Li, J. Bi, Z. Zhuang, Y. Yu, Z. Zou, A covalent organic 13785–13792.
framework bearing single Ni sites as a synergistic photocatalyst for selective pho- [50] D.K. Wu, Q. Xu, J. Qian, X.P. Li, Y.H. Sun, Bimetallic covalent organic frameworks
toreduction of CO2 to CO, J. Am. Chem. Soc. 141 (2019) 7615–7621. for constructing multifunctional electrocatalyst, Chem. Eur. J. 25 (2019)
[37] S.Z. Yang, W.H. Hu, X. Zhang, P.L. He, B. Pattengale, C.M. Liu, M. Cendejas, 3105–3111.
I. Hermans, X.Y. Zhang, J. Zhang, J.E. Huang, 2D covalent organic frameworks as [51] T. Zhang, K. Manna, W.B. Lin, Metal-organic frameworks stabilize solution-in-
intrinsic photocatalysts for visible light-driven CO2 reduction, J. Am. Chem. Soc. accessible cobalt catalysts for highly efficient broad-scope organic transformations,
140 (2018) 14614–14618. J. Am. Chem. Soc. 138 (2016) 3241–3249.
[38] J. Xu, Y. He, S. Bi, M. Wang, P. Yang, D. Wu, J. Wang, F. Zhang, An olefin-linked [52] S.S. Acharya, B. Winther-Jensen, L. Spiccia, C.A. Ohlin, Rates of water exchange in
covalent organic framework as a flexible thin-film electrode for a high-performance 2,2-bipyridine and 1,10-phenanthroline complexes of Co-II and Mn-II, Aust. J.
micro-supercapacitor, Angew. Chem. Int. Ed. 58 (2019) 12065–12069. Chem. 70 (2017) 751–754.
[39] S. Wei, F. Zhang, W. Zhang, P. Qiang, K. Yu, X. Fu, D. Wu, S. Bi, F. Zhang, [53] H.B. Aiyappa, J. Thote, D.B. Shinde, R. Banerjee, S. Kurungot, Cobalt-modified
Semiconducting 2D triazine-cored covalent organic frameworks with unsubstituted covalent organic framework as a robust water oxidation electrocatalyst, Chem.
olefin linkages, J. Am. Chem. Soc. 141 (2019) 14272–14279. Mater. 28 (2016) 4375–4379.
[40] P.L. Wang, S.Y. Ding, Z.C. Zhang, Z.P. Wang, W. Wang, Constructing robust cova- [54] C.S. Diercks, S. Lin, N. Komienko, E.A. Kapustin, E.M. Nichols, C.H. Zhu, Y.B. Zhao,
lent organic frameworks via multicomponent reactions, J. Am. Chem. Soc. 141 C.J. Chang, O.M. Yaghi, Reticular electronic tuning of porphyrin active sites in
(2019) 18004–18008. covalent organic frameworks for electrocatalytic carbon dioxide reduction, J. Am.
[41] X.D. Zhuang, W.X. Zhao, F. Zhang, Y. Cao, F. Liu, S. Bia, X.L. Feng, A two-dimen- Chem. Soc. 140 (2018) 1116–1122.
sional conjugated polymer framework with fully sp2-bonded carbon skeleton, [55] J.J. Leung, J.A. Vigil, J. Warnan, E. Edwardes Moore, E. Reisner, Rational design of
Polym. Chem. 7 (2016) 4176–4181. polymers for selective CO2 reduction catalysis, Angew. Chem. Int. Ed. 58 (2019)
[42] E.Q. Jin, M. Asada, Q. Xu, S. Dalapati, M.A. Addicoat, M.A. Brady, H. Xu, 7697–7701.
T. Nakamura, T. Heine, Q.H. Chen, D.L. Jiang, Two-dimensional sp2 carbon-con- [56] G.X. Zhao, W. Zhou, Y.B. Sun, X.K. Wang, H.M. Liu, X.G. Meng, K. Chang, J.H. Ye,
jugated covalent organic frameworks, Science 357 (2017) 673–676. Efficient photocatalytic CO2 reduction over Co(II) species modified CdS in aqueous
[43] E. Jin, J. Li, K. Geng, Q. Jiang, H. Xu, Q. Xu, D. Jiang, Designed synthesis of stable solution, Appl. Catal., B: Environ. 226 (2018) 252–257.
light-emitting two-dimensional sp2 carbon-conjugated covalent organic frame- [57] S.W. Cao, Y. Li, B.C. Zhu, M. Jaroniec, J.G. Yu, Facet effect of Pd cocatalyst on
works, Nat. Commun. 9 (2018) 4143. photocatalytic CO2 reduction over g-C3N4, J. Catal. 349 (2017) 208–217.
[44] Q. Zhang, M. Dai, H. Shao, Z. Tian, Y. Lin, L. Chen, X.C. Zeng, Insights into high [58] A. Call, M. Cibian, K. Yamamoto, T. Nakazono, K. Yamauchi, K. Sakai, Highly ef-
conductivity of the two-dimensional iodine-oxidized sp2-c-COF, ACS Appl. Mater. ficient and selective photocatalytic CO2 reduction to CO in water by a cobalt por-
Interfaces 10 (2018) 43595–43602. phyrin molecular catalyst, ACS Catal. 9 (2019) 4867–4874.
[45] E. Jin, Z. Lan, Q. Jiang, K. Geng, G. Li, X. Wang, D. Jiang, 2D sp2 carbon-conjugated [59] S. Wang, X. Hai, X. Ding, S. Jin, Y. Xiang, P. Wang, B. Jiang, F. Ichihara, M. Oshikiri,
covalent organic frameworks for photocatalytic hydrogen production from water, X. Meng, Y. Li, W. Matsuda, J. Ma, S. Seki, X. Wang, H. Huang, Y. Wada, H. Chem,
Chemistry 5 (2019) 1632–1647. J. Ye, Intermolecular cascaded π-conjugation channels for electron delivery pow-
[46] C. Wang, R. Chen, J.-L. Shi, Y. Ma, G. Lin, X. Lang, Designed synthesis of a 2D ering CO2 photoreduction, Nat. Commun. 11 (2020) 1149.
porphyrin-based sp2 carbon-conjugated covalent organic framework for