A 2D layered cobalt-based metal–organic framework for photoreduction of CO2 to syngas with a controllable wide ratio range

Mei-Juan Wei a, Xian-Yan Xu *b, Jia-Qi Song a, Mei Pan *a and Cheng-Yong Su *a
aMOE Laboratory of Bioinorganic and Synthetic Chemistry, Lehn Institute of Functional Materials, School of Chemistry, Sun Yat-Sen University, Guangzhou 510006, China. E-mail: panm@mail.sysu.edu.cn; cesscy@mail.sysu.edu.cn
bCollege of Chemistry and Civil Engineering, Shaoguan University, Shaoguan 512005, China. E-mail: sofiaxxy@sgu.edu.cn

Received 16th October 2022 , Accepted 29th November 2022

First published on 29th November 2022


Abstract

Photocatalytic CO2 reduction to syngas (a mixture of CO and H2) with adjustable composition is a prospective way to mitigate the energy shortage and the greenhouse effect. Herein, we synthesize a thermally stable two-dimensional cobalt-based metal–organic framework (denoted as “Co-TBAPy”) consisting of cobalt metal centers and H4TBAPy (1,3,6,8-tetrakis(p-benzoic acid) pyrene) as an organic linker, which exhibits superior CO2 adsorption capability and a considerable specific surface area. In accordance with the energy levels of [Ru(bpy)3]Cl2·6H2O as a sensitizer, Co-TBAPy is active for photocatalytic reduction of CO2 to syngas with a wide controllable ratio range between 0.14 and 1.65 under visible-light irradiation. Moreover, the proportion of CO[thin space (1/6-em)]:[thin space (1/6-em)]H2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 and 1[thin space (1/6-em)]:[thin space (1/6-em)]3 favorable for the synthesis of methanol and methane, respectively, can be precisely regulated, which few MOF-based photocatalysts have achieved so far. Furthermore, combined with DFT calculation, we reveal the influence of the water content of the reaction system in the process of photocatalytic CO2 reduction to produce a controllable proportion of syngas, which is often a profound while elusive factor in photoreduction reactions. This work provides a feasible concept for the design and application of Co-MOF materials for the photoreduction of CO2 to syngas in practical industrial fields.


Introduction

Nowadays, the storage and transformation of carbon dioxide (CO2) are two promising approaches to reduce the excessive emission of CO2 and ease immoderate consumption of fossil resources.1 Converting CO2 into value-added chemicals or fuels by artificial photosynthesis is recognized as a reliable way to utilize CO2 and store solar energy simultaneously. Since CO2 reduction involves multi-electron processes and competition of hydrogen evolution, the photoreduction products of CO2 are diverse, including CH4, CH3OH, HCOOH, CO, H2, etc.2,3 To our knowledge, syngas, a mixture of CO and H2, is a significant feedstock for the synthesis of high-value added fuels and chemicals by Fischer–Tropsch processes. For actual industrial applications, definite ratios of syngas correspond to certain products. Typically, syngas with the ratio of CO[thin space (1/6-em)]:[thin space (1/6-em)]H2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]2 and 1[thin space (1/6-em)]:[thin space (1/6-em)]3 can be utilized to synthesize alcohol, methanol and methane, respectively.4 However, the traditional preparation of syngas through a water gas shift reaction requires harsh conditions and might result in syngas with unsuitable ratios, which restrict its practical industrial application.5–9

Therefore, the exploration of appropriate catalysts to prepare a tunable composition of syngas under mild conditions is highly attractive. To date, various photocatalysts have been verified to be suitable catalysts for photocatalytic reduction of CO2 to syngas. For instance, Co-ZIF-9 or Co3O4 can serve as a photocatalyst to photoreduce CO2 to syngas under visible light with unadjusted ratios.10 TiO2 mesoporous hollow spheres decorated with CuPt alloys and MnOx and a heterostructured CoAl-layered double hydroxide/MoS2 nanocomposite photocatalyst (CoAl-LDH/MoS2) exhibit excellent performance in CO2 photoreduction to CO and H2 with a tunable ratio.8,11 However, in order to achieve the desired CO[thin space (1/6-em)]:[thin space (1/6-em)]H2 ratio of syngas, the structure of the above composite photocatalysts must be well modified accordingly, which is complicated and the mechanism behind is difficult to reveal. Consequently, the preparation of visible light driven catalysts that can produce an adjustable ratio of syngas efficiently and the clear clarification of the structure–performance mechanism are still daunting challenges.

Metal–organic frameworks (MOFs), a type of crystalline material composed of metal nodes and organic ligands as linkers, have been widely researched in various fields,12 such as gas adsorption,13 luminescence,14 catalysis,15–17etc. Thanks to their adjustable cavities, large specific surface area, easy modification of the structure and excellent adsorption performance of CO2, some MOFs, like boron imidazolate frameworks (BIF-101) (ref. 18) and (Co/Ru)n-UiO67(bpydc),19 have been reported as excellent heterogeneous photocatalysts for CO2 photoreduction to syngas. 2D layer-like MOFs due to their readily accessible active sites and facility for mass diffusion exhibit dramatically boosted photocatalysis.20–24 However, most of the MOF-based photocatalysts are unstable and prone to structural collapse leading to deactivation, and the research on MOF-based catalysts towards CO2 photoreduction to syngas with tunable ratios is still in infancy.

Transition metals, especially Cu, Co, Zn and Ni with multiple redox states, can provide favorable CO2 adsorption sites and advantageous catalytic sites.16,25 Among them, cobalt (Co)-based complexes are not only low-cost,26 but also applicable as co-catalysts to enhance CO2 photoreduction by promoting the photogenerated electron separation.27,28 Moreover, compared with Cu and Ni, Co possesses good binding capacity for both CO2 and protons, which can serve as promising catalytic sites for CO2 reduction to syngas. In an attempt to build an efficient CO2 photoreduction system, an effective electron-transport chain and strong light-harvesting capability are crucial. Generally, pyrene and its derivatives possess rigid extended π-systems, which might display π-conjugation or intermolecular π–π interaction and provide potential for visible-light absorption. Therefore, pyrene-based metal–organic frameworks have attracted much attention as photocatalysts.29–32

Herein, a thermally stable two-dimensional cobalt-based metal–organic framework, {[Co2(TBAPy)(H2O)2]·(DMF)(H2O)8}n (Co-TBAPy), was assembled successfully with the pyrene-based ligand H4-TBAPy (1,3,6,8-tetrakis(p-benzoic acid)pyrene) serving as a linker and cobalt as a node through a facile solvothermal method. Cooperating with [Ru(bpy)3]Cl2·6H2O as a photosensitizer and triethanolamine (TEOA) as a sacrificial agent, Co-TBAPy exhibits excellent photocatalytic performance towards CO2 reduction under visible light, and can achieve a CO evolution rate of 1856.5 μmol g−1 h−1. Especially, the ratio of syngas (CO/H2) can be regulated in a wide range between 0.14 and 1.65 simply by adjusting the water content in the photocatalytic system, and a precise proportion of CO[thin space (1/6-em)]:[thin space (1/6-em)]H2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 and 1[thin space (1/6-em)]:[thin space (1/6-em)]3 favourable for fuel synthesis can be attained. Theoretical calculations were performed to clearly disclose the mechanisms.

Results and discussion

Single-crystal X-ray diffraction verifies that Co-TBAPy crystallizes in a triclinic space group P[1 with combining macron], which is the same as that reported before (Table S1).33 Each asymmetric unit possesses a Co2+ atom, a TBAPy4− ligand and a coordinated water molecule. In detail, Co-TBAPy is constructed from dinuclear Co2(CO2)4 paddle-wheel secondary building units (SBUs) capped with two water molecules at the axial positions, which are further connected by four TBAPy4− ligands (Fig. 1a). It is due to the dinuclear paddle-wheel structure that gives rise to a two-dimensional layered structure in Co-TBAPy equipped with two one-dimensional rhombus channels (size 6.34 × 7.18 Å and 7.21 × 9.08 Å, respectively) perpendicular to each layer (Fig. 1b). Confirmed by TOPOS software, Co-TBAPy retains a (4,4)-net topology, which can be simplified into an sql net (Fig. 1c). The crystal structure and refinement data of Co-TBAPy are supplied in Table S2. As shown in the SEM images (Fig. S1), Co-TBAPy presents a regular block shape with a flat surface, showing a thickness of ∼7 μm and a length of 20–50 μm. To investigate the purity, the powder X-ray diffraction patterns (PXRD) of Co-TBAPy synthesized were obtained, which exhibited excellent consistency with the simulated one (Fig. 1d).
image file: d2ta08092c-f1.tif
Fig. 1 Crystal structure of Co-TBAPy. (a) Coordination environment of Co2+, (b) 2D layer structure of Co-TBAPy along the a-axis, (c) simplified diagram for the packing of 2D layers with (4,4)-net topology, and (d) PXRD patterns of Co-TBAPy as-synthesized and simulated. Color representation: gray, C; red, O; blue, Co. H atoms are removed for clarity.

The powder of as-synthesized Co-TBAPy was immersed in various polar solvents, like acetone, methanol, ethanol, acetonitrile (MeCN), dimethylacetamide (DMA) and dichloromethane (DCM), in which Co-TBAPy can retain the pristine structure for 6 days at room temperature. Moreover, it can exist stably in water for 3 days. All of the PXRD diagrams without obvious changes have proved the solvent stability and universality of Co-TBAPy (Fig. S2). Compared with the FT-IR spectrum of H4TBAPy, the disappearance of the hydroxyl vibration peak of –COOH in the range of 3400–3000 cm−1 in the spectrum of Co-TBAPy illustrates the formation of Co–O bonds (Fig. S3). Simultaneously, the peak at 1750–1620 cm−1 originating from the vibration of the carbonyl group (–COOH) shifts to a lower wavenumber. Thermogravimetric (TG) analysis exhibits a weight loss of ∼12.2% in the range of 98–180 °C, which is attributed to the loss of H2O. And the core frameworks of Co-TBAPy can remain intact even at 405 °C (Fig. S4). Calculated by using PLATON software, the solvent accessible volume of Co-TBAPy is estimated to be 33.3% and the total potential solvent area volume is 376.6 Å3.

A type I isotherm was observed from the nitrogen adsorption/desorption curves at 77 K (Fig. 2a), indicating the presence of micropores in Co-TBAPy. The Brunauer–Emmett–Teller (BET) surface area was calculated to be 851 m2 g−1 and the average pore width is evaluated to be ∼0.67 nm, demonstrating abundant potentially active sites for photocatalysis (Fig. S5). Furthermore, Co-TBAPy displays a high affinity to CO2, whose maximum adsorption capacity is measured to be 56.5 cm3 g−1 at 273 K (Fig. 2b). The adsorption enthalpy for carbon dioxide is calculated to be 23.5–26.6 kJ mol−1, revealing that Co-TBAPy is a kind of potential material for photocatalytic reduction of CO2 (Fig. S6).


image file: d2ta08092c-f2.tif
Fig. 2 (a) N2 adsorption–desorption isotherms at 77 K. (b) CO2 isotherms at 273 K and 298 K of Co-TBAPy. (c) Tauc plot of Co-TBAPy. (d) Mott–Schottky plot of Co-TBAPy at 500 Hz and 1000 Hz, respectively.

The ultraviolet-visible absorption (UV-vis) spectrum shows that Co-TBAPy has excellent absorption capacity for visible light, in which the maximum adsorption value reaches 680 nm (Fig. S7). As calculated by the Kubelka–Munk (K–M) method34 (Fig. 2c), the band gap of Co-TBAPy is estimated to be 1.84 eV, unveiling its potential to be a kind of semiconducting photocatalyst. Mott–Schottky measurements on Co-TBAPy were conducted at 500 Hz and 1000 Hz, respectively (Fig. 2d). The positive slope of the curves elucidates that Co-TBAPy is a typical n-type semiconductor and its flat band position is determined to be ∼−0.88 V vs. Ag/AgCl (i.e., −0.68 V vs. NHE). Since it is generally accepted that the bottom of the conduction band (CB) of n-type semiconductors is more negative by about 0.1 V than the flat band position,35 the CB of Co-TBAPy is forecasted to be −0.78 V vs. NHE, which is more negative than the redox potential of E(CO2/CO) = −0.53 V (vs. NHE).36 In light of the band gap of 1.84 eV determined above, the value of the valence band (VB) is estimated to be 1.06 V vs. NHE.

Considering the prominent thermal and solvent stability, specific surface area, and the suitable band positions for CO2 redox potential, the CO2 photoreduction performance of Co-TBAPy under visible light was determined on a CO2-saturated mixture of acetonitrile and water with [Ru(bpy)3]Cl2 serving as a photosensitizer and TEOA acting as a sacrificial agent. CO was detected as the only carbonous product and H2 was discovered as the byproduct using a gas chromatograph. As the illumination time increases, the evolution of CO increased rapidly, while the yield of H2 did not decrease significantly. As displayed in Fig. 3a and S8, the amounts of CO and H2 reach 27 μmol and 156 μmol after 24 h illumination, in which the highest production yield of CO can reach 990.8 μmol g−1 h−1. The turnover number (TON) is calculated to be 5.4 for CO and 31.2 for H2 according to the content of cobalt of Co-TBAPy used in photocatalysis.


image file: d2ta08092c-f3.tif
Fig. 3 (a) Time-dependent gas evolution of CO, and the inset image shows the H2/CO ratio in CO2-saturated MeCN solution (5 mL) containing Co-TBAPy (5 μmol), [Ru(bpy)3]Cl2 (0.01 mmol), TEOA (1 mL) and H2O (0.25 mL) at room temperature and irradiated with λ > 420 nm. (b) GC-MS results of the isotopic experiment under a 13CO2 atmosphere (right) and for comparison, the air sample was detected. (c) The amounts of CO during three-run recycling experiments. (d) Plots of H2/CO generation with the tuning of the ratio of H2O under similar conditions for an 18 h reaction and the curve of the ratio of syngas (CO/H2) at different H2O contents.

To probe the key factors that influence the action of photocatalytic reduction of CO2, a series of control experiments were conducted (Table S2). In the absence of light, [Ru(bpy)3]Cl2·6H2O or TEOA, there are no gaseous products detected, unveiling that light, a photosensitizer and a sacrificed agent are essential for Co-TBAPy in the catalytic CO2 reduction process. Without Co-TBAPy, less CO and H2 are produced, which reveals that Co-TBAPy does play a catalytic role in the photoreduction process. While replacing CO2 with N2, no CO was evolved. Isotopic 13CO2 labeling experiments verified CO2 as the only carbon source (Fig. 3b). Compared with the air sample, the peak at 28 m/z shown in Fig. 3b (left) originates from the N2 of air leaked in the injection needle. Furthermore, recycling experiments were conducted to examine the stability of Co-TBAPy. After three rounds of recycling tests, the photocatalytic activity is well maintained (Fig. 3c and S9). Deduced from the PXRD pattern and FT-IR spectrum of Co-TBAPy after photocatalysis, Co-TBAPy retained integrated frameworks (Fig. S10 and S11). Obviously, Co-TBAPy is manifested to be a stable heterogeneous catalyst for photocatalytic reduction of CO2.

It is acknowledged that CO and H2 are the main components of syngas for methanol synthesis and the Fischer–Tropsch reaction.37 Since the carbon product selectivity is wonderful without an excess carbonous product during the process of photocatalytic reduction of CO2, it is theoretically feasible to use the competitive reaction of hydrogen evolution to adjust the output of CO and H2, thereby realizing an adjustment of syngas in various ratios. Photocatalytic CO2 reduction was performed under different contents of water added into MeCN solution. As shown in Fig. 3d, by using pure MeCN solvent without water, the evolution of CO (2924.2 μmol g−1) is less than that of hydrogen (20[thin space (1/6-em)]893.7 μmol g−1). As the water content increases, the production of CO gradually increases, and the output of H2 first increases accordingly but then decreases when the water content is higher than 8%. When the water content reaches 10%, the CO evolution can reach 37.13 μmol after 10 h irradiation, during which the highest CO production rate reaches 1856.5 μmol g−1 h−1 (Table S2, entry 1). Evidently, by simply adjusting the water content of the system, we can achieve a wide range of controllable synthesis of the syngas ratio CO/H2 from 0.14 to 1.65. As compared in Table S3, in the current research on the photocatalytic reduction of CO2 to produce controllable proportions of syngas, Co-TBAPy outperforms most MOF-based catalysts, homogeneous cobalt complexes and some inorganic semiconductors. It is worth mentioning that when using a H2O content of 6%, the evolution of CO and H2 keeps a ratio nearly at 1[thin space (1/6-em)]:[thin space (1/6-em)]3, which can serve as raw materials for the synthesis of methane. Moreover, when the content of H2O goes up to 8%, the yield of CO and H2 can achieve a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2, which is a profitable composition of syngas for methanol synthesis and industrial application in Fischer–Tropsch hydrocarbon formation.8,38,39

To obtain an in-depth understanding of the electron transfer process on photocatalytic CO2 reduction, a series of characterization tests on Co-TBAPy were conducted. The photocurrent response curves of Co-TBAPy were investigated (Fig. S12). A fast and regular photocurrent response was recorded for each light-on/off event under visible-light (λ > 420 nm) irradiation, indicating that the photogenerated electrons transferred quickly on the surface of Co-TBAPy without recombination. The steady state photoluminescence (PL) spectra and time-resolved transient PL decay spectra were investigated for the as-synthesized Co-TBAPy and ligand H4TBAPy. An obvious fluorescence quenching phenomenon and shorter lifetime appear on Co-TBAPy, demonstrating that photo-excited electron–hole pairs separated efficiently on the surfaces (Fig. 4a and S13).


image file: d2ta08092c-f4.tif
Fig. 4 (a) Photoluminescence (PL) spectra of as-synthesized Co-TBAPy and the ligand H4TBAPy (λex = 390 nm). (b) Schematic energy level diagram for Co-TBAPy and [Ru(bpy)3]Cl2. (c) Steady state fluorescence spectra of [Ru(bpy)3]Cl2 upon the addition of increasing amounts of Co-TBAPy (λex = 395 nm). (d) The high resolution X-ray photoelectron spectroscopy (XPS) of the Co 2p orbit of Co-TBAPy before and after photocatalysis.

Comparing different types of photosensitizers, [Ru(bpy)3]Cl2·6H2O exhibited the best light absorption performance during the photocatalytic CO2 reduction (Fig. S14). The band energy-level accordance between Co-TBAPy and [Ru(bpy)3]Cl2·6H2O is a significant factor. The highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) energy levels of [Ru(bpy)3]Cl2·6H2O are described to be 1.22 V and −1.27 V vs. NHE, respectively. As presented in Fig. 4b, the LUMO energy level is more negative than the CB position of Co-TBAPy, which is theoretically possible because photoexcited electrons can be transferred from the LUMO level of [Ru(bpy)3]Cl2·6H2O to the CB of Co-TBAPy to reduce the CO2 molecules bound on cobalt sites. PL quenching measurements were conducted on the photocatalytic system in acetonitrile. With the increasing amounts of Co-TBAPy, the PL intensity of excited [Ru(bpy)3]Cl2·6H2O gradually diminished, unveiling an oxidative quenching mechanism for the excited state [Ru(bpy)3]2+* where the excited electrons were transferred to Co-TBAPy (Fig. 4c).40 In contrast, an evident enhanced PL intensity has arisen when different amounts of TEOA were added into [Ru(bpy)3]Cl2·6H2O (Fig. S15).

Moreover, high resolution X-ray photoelectron spectroscopy (XPS) was carried out to analyze the composition and chemical states of elements in Co-TBAPy (Fig. 4d and S16). Before photocatalysis, the binding energies at 781.6 eV and 797.7 eV were ascribed to Co 2p3/2 and Co 2p1/2, respectively, accompanied by two satellite peaks at 786.1 eV and 802.6 eV, attributed to the shakeup excitation of high-spin Co2+ ions.41 After photocatalysis, a slight shift to a lower Co binding energy was observed, disclosing that some of the cobalt atoms on the surface received electrons during photocatalysis, together with a coordination environment change of cobalt.42 Analogously, the XPS spectrum of Co 3s displays a strong splitting attributed to the 3d7 high-spin configuration of Co2+ (S = 3/2) before photocatalysis, while after photocatalysis, only a single peak at 102.5 eV is observed (Fig. S17).43 These results further confirmed that Co sites served as an electron transfer intermediate station during the photocatalytic process, which can receive electrons from the photosensitizer and then transfer to the bound CO2.

On the basis of the above photocatalytic results and physical characterization, a possible mechanism of photocatalytic reduction of CO2 in the Co-TBAPy participated system is proposed (Scheme 1). Under visible-light irradiation, the light absorber [Ru(bpy)3]Cl2·6H2O is excited to generate photo-excited electron–hole pairs, in which photogenerated electrons migrate to the Co sites by oxidative quenching and excited state [Ru(bpy)3]2+* is reduced by the electron donor TEOA. Then the CO2 molecules absorbed on Co-TBAPy accept electrons from cobalt sites to form CO, accompanied by hydrogen evolution. Among these processes, adjusting the water content in the photocatalytic system can provide alterable protons for the formation of CO. Meanwhile, H2O molecules can stabilize the CO2 reaction intermediate through the hydrogen bonding effect. It might result in a more favorable CO2RR than the HER when a higher water content is involved, which effectively realizes the controlled synthesis of syngas in different proportions.


image file: d2ta08092c-s1.tif
Scheme 1 Proposed catalytic mechanism for the photocatalytic reduction of CO2 catalyzed by Co-TBAPy.

Density functional theory (DFT) calculations were carried out to reveal the effect of the water content on the performance of Co-TBAPy photocatalytic reduction of CO2 to syngas with a wide range of adjustable ratios (Fig. S18 manifesting the calculated configurations).44–46 The band gap of a photocatalyst is crucial to determine its catalytic performance; therefore, the band structure of Co-TBAPy is first calculated as shown in Fig. S19, giving the band gap at 1.79 eV, in reasonable agreement with the experimental results. In addition, the density of states (DOS) of Co-TBAPy also unveils that the bottom of the conduction band is mainly contributed by the Co and C atoms while the top of the valence band is mainly contributed by the C atoms. Next, we considered the reaction mechanism of photocatalytic CO2 reduction over Co-TBAPy.47,48 As shown in Fig. 3d, the evolution of CO gradually increases as the H2O content increases to a maximum of 8% and then slightly drops at 10%. According to the literature, with the presence of water in the MeCN solvent, the CO formation reaction may proceed via a proton-coupled electron transfer between CO2 and H2O (eqn (1)).49

 
CO2(g) + H2O(l) + 2e → CO(g) + 2OH(1)

in which the presence of H2O as a proton source may facilitate the CO formation. Furthermore, in the cases of some cobalt complexes, H2O molecules stabilize the CO2 molecules bound to the catalyst by forming hydrogen bonds, which also promotes the CO formation. Based on the above two facts, we propose that the interaction and charge transfer between CO2 and H2O are crucial for the mechanism of the CO formation reaction catalyzed by Co-TBAPy. However, the origin of the dependence of CO evolution on the H2O content remains unclear. To unravel the reaction mechanism, we investigated both the H2O activation and hydrogen bond interaction which are proposed to be important for the HER and CO2RR.50,51 We found that, for single molecule adsorption, the adsorption of H2O over a Co site is much stronger (adsorption energy Eads = −1.23 eV) than that of CO2 (Eads = −0.25 eV), indicating a preference towards H2 evolution rather than the CO2RR under a low H2O content (Fig. S20).52 With an extra electron addition to Co-TBAPy, CO2 only physisorbs on Co with up to 4H2O molecules surrounded (Fig. 5a i–iii). Interestingly, CO2 chemisorbs on Co with more than five H2O molecules presented (Fig. 5a iv and v). Upon chemisorption, CO2 attaches to Co via a side-on configuration with one of its C[double bond, length as m-dash]O double bonds.53,54 Bader charge analysis shows that barely no electrons are accumulated on the CO2 molecule upon physisorption (−0.03e for (CO2 + 2H2O) and (CO2 + 3H2O) systems and −0.04e for (CO2 + 4H2O)), while considerable electrons transfer from the catalyst to CO2 in chemisorbed (CO2 + 5H2O) and (CO2 + 6H2O) systems (−0.47e and −0.44e, respectively).


image file: d2ta08092c-f5.tif
Fig. 5 (a) The configuration of CO2 adsorption on Co-TBAPy and corresponding adsorption free energy at 298 K, which involves different numbers of H2O molecules: (i) two H2O, (ii) three H2O, (iii) four H2O, (iv) five H2O, and (v) six H2O. The brown dashed line labels the hydrogen bonding, while the blue dashed line represents the distance between atoms. (b and c) Calculated free energy diagrams for the photocatalytic HER (b), and CO2RR to CO (c) involving a multiple number of H2O molecules (pH = 8 and at 0 V vs. RHE). * represents an active site.

As a further stage in understanding the activity difference of Co-TBAPy in the HER and CO2RR and explaining the experimental observation, we investigated the reaction pathways for both the HER and the CO2RR involving different numbers of water molecules using the computational hydrogen electrode (CHE) method (Fig. 5b and c and S20).55,56 According to a previous study, HER pathways can be summarized as the following four elementary steps under alkaline and neutral conditions:57

 
H2O(l) + * → H2O*(2)
 
H2O* → OH* + H*(3)
 
image file: d2ta08092c-t1.tif(4)
 
OH* + e → OH + *(5)

For the CO2RR under a low water content (corresponding to the scenarios in which less than five H2O molecules are involved), we proposed that H2O adsorption and dissociation first occur due to the preferred H2O adsorption on Co. The adsorption species OH* and H* can be obtained after H2O dissociation. Then OH* accepts electrons and desorbs from the Co site, which facilitates an interaction between CO2 and remaining H* to form an intermediate COOH*. It is followed by the dissociation of COOH* into a CO* intermediate and OH. Finally, CO is obtained after desorption. All the processes are summarized as follows:

 
CO2(g) + H* → COOH*(6)
 
COOH* + e → CO* + OH(7)
 
CO* → CO(g) + *(8)

On the other hand, for the CO2RR under a higher water content (corresponds to the scenarios in which five and six H2O molecules are involved), CO2 chemisorption is most likely achieved and its reduction to CO is possible as proposed below:58

 
CO2(g) + * + e → *CO2(9)
 
*CO2 + H2O(l) → COOH* + OH(10)

We can find that under a low water content (corresponding to the scenarios with two to four H2O molecules), the reaction mechanisms of both the HER and the CO2RR go through via the process of H2O adsorption and dissociation. Considering the fact that the H2O dissociation (H2O* → H* + OH*) is a common step for the HER and CO2RR, although it is also the determining step for systems with 3H2O or 4H2O, the competition of the HER and CO2RR is determined by the elementary steps with maximum reaction energy subsequent to (H2O* → H* + OH*) for each reaction process, i.e.image file: d2ta08092c-t2.tif (eqn (4)) for the HER and OH* + e → OH + * (eqn (5)) for the CO2RR, which is 1.05 eV vs. 1.37 eV for the 2H2O system, 0.57 eV vs. 0.76 eV for the 3H2O system, and 0 vs. 0.42 eV for the 4H2O system. Therefore, the HER is more preferred than the CO2RR due to the lower reaction free energy of image file: d2ta08092c-t3.tif along the HER compared to that of OH* + e → OH + * along the CO2RR under a low water content.

Meanwhile under a higher water content (corresponding to the scenarios with five and six H2O molecules involved in the reaction), the HER is now determined by using image file: d2ta08092c-t4.tif (eqn (4)). On the other hand, the CO2RR will not go through H2O dissociation but direct protonation to *COOH due to the presence of CO2 chemisorption, and its determining step becomes *CO2 + H2O(l) → COOH* + OH (eqn (10)). In these cases, the competition of the HER and CO2RR is controlled by the reaction free energies of the corresponding determining steps (i.e.eqn (4) and (10), respectively). The calculated reaction free energies are 1.19 and 1.03 eV for image file: d2ta08092c-t5.tif (eqn (4)) along the HER pathway in 5H2O and 6H2O systems, respectively, while the calculated reaction free energies are 0.66 and 0.68 eV for *CO2 + H2O(l) → COOH* + OH (eqn (10)) along the CO2RR pathway in 5H2O and 6H2O systems, respectively. Therefore, the CO2RR shows lower reaction free energy than the HER and thus the CO2RR suppresses the HER under a high water content. These computational results are consistent well with the experiments, showing promise to control the ratio of CO[thin space (1/6-em)]:[thin space (1/6-em)]H2 in syngas photosynthesis from CO2 reduction by using Co-based MOFs as a photocatalyst and water content as a modulator.

Experimental section

Chemicals and characterization

All reagents of analytical grade were acquired through commercial channels and used without further purification. The PXRD spectra were recorded on a Rigaku SmartLab X-ray diffractometer using Cu-Kα radiation (λ = 1.54056 Å) at room temperature. The SC-XRD data were recorded at the SSRF (Shanghai Synchrotron Radiation Facility) 14B beamline. The X-ray photoelectron spectroscopy (XPS) data were collected on an ESCALab 250 using a monochromatic Al Kα radiation source. The FT-IR spectra were collected on a Nicolet/Nexus-670 spectrometer using a KBr disk. The thermogravimetric (TG) analyses were performed on a NETZSCH TG209 system in nitrogen with a heating rate of 5 °C min−1. The Brunauer–Emmett–Teller (BET) specific surface area and pore volume were analysed by using a Quantachrome Autosorb-iQ2-MP gas adsorption analyzer. The gases (N2 and CO2) used in the gas adsorption measurements hold ultra-high purity. The morphology of the samples was determined by using an ultra-high resolution field emission scanning electron microscope (SEM, SU8010). The UV-vis absorption data were obtained on a SHIMADZU UV-3600 equipped with an integrating sphere using BaSO4 as a reference. Photoluminescence (PL) and time-resolved fluorescence decay spectroscopy were performed on an EDINBURGH FLS980 fluorescence spectrophotometer.

Synthesis of Co-TBAPy {[Co2(TBAPy)(H2O)2]·(DMF)(H2O)8}n

The crystals of Co-TBAPy were obtained by a facile solvothermal method between H4TBAPy (1,3,6,8-tetrakis(p-benzoic acid)pyrene) (0.015 mmol) and Co(NO3)2·6H2O (0.1 mmol) in a mixed solution of DMF, H2O, dioxane and concentrated hydrochloric acid (v/v/v/v, 2/1/1/0.01) at 120 °C for 3 days. Green transparent flake crystals were collected and washed with DMF and acetone, respectively (yield at ∼76.5%, based on ligands). The Co-TBAPy synthesized was soaked in acetone for 3 days, during which the acetone was changed every 12 h. After that, Co-TBAPy was filtered and dried in an oven at 100 °C for 12 h and activated Co-TBAPy was obtained. Elemental analysis: found for C47H49O19NCo2: C 53.49, H 4.14, N 0.95%; calcd: C 53.77, H 4.67, N 1.33%.

Photocatalytic test

Photocatalytic carbon dioxide reduction tests were carried out in a 54 mL closed Pyrex reactor with a LED lamp (light intensity: 100 mW cm−2, PCX50B Discover multi-channel photochemical reaction system, Beijing Perfectlight Technology Co., Ltd) serving as a visible light source. Typically, the catalyst Co-TBAPy (0.005 mmol, activated) and [Ru(bpy)3]Cl2·6H2O (0.01 mmol) were dispersed in an acetonitrile aqueous solution of different ratios, in which TEOA (1 mL) acted as a sacrificial agent to suppress the oxidation half reaction. Then the reaction flask was evacuated and filled with CO2, repeated three times before the irradiation. The gas products were detected by using a gas chromatograph (GC9790, Fuli Analytical Instrument Co., Ltd) with a thermal conductivity detector (TCD) for H2 detection and a flame ionization detector for CO detection.

Photoelectron measurement

All of the photoelectric tests adopted a standard three-electrode system, consisting of a platinum plate electrode (1 cm × 1 cm) as the counter electrode, Ag/AgCl as the reference electrode and MOF catalyst coated fluorine doped tin oxide (FTO) glass as the working electrode with a working area of 0.5 × 0.5 cm2. In detail, 5 mg of Co-TBAPy samples and 20 μL Nafion D-520 dispersion (5% w/w in water and 1-propanol) were added into 0.5 mL of ethanol and then ultrasonicated for 5 min to obtain a suspension. 0.2 mL of the suspension liquid was added dropwise onto the FTO glass and the coating area was restricted to 0.25 cm2 by using insulating tape. The well coated FTO glass was dried at 85 °C for a day. The electrolyte was 0.5 M sodium sulfate aqueous solution. Both Mott–Schottky measurements and photocurrent response tests were performed on a CHI 660E electrochemical analyzer (CH Instruments, Chenhua Co., Shanghai, China). The Mott–Schottky measurements were performed at 500 Hz and 1000 Hz, respectively, with an amplitude at 0.01 V. During the photocurrent response tests, a Xe Lamp (300 W, λ > 420 nm, CEL-HXF300, Beijing Aulight Co.,) served as the visible light source. The on–off cycle of illumination was manually controlled by the operator at a time interval of 5 s.

Computational details

All calculations were performed in a spin-polarized fashion and by using the VASP code based on plane-wave basis sets. The generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) functional was for the exchange–correlation functional. Projector augmented wave (PAW) potentials were used for electron–ion interactions with an energy cutoff of 500 eV. The energy and force convergence settings were 10−5 eV and 0.02 eV Å−1, respectively. Hubbard U correction was also introduced for the Co atom with Ueff = 3.3. The k-point mesh was set to be 4 × 2 × 2 for geometry optimization and 8 × 5 × 3 for electronic structure calculations. The cell parameter of Co-TBAPy after the optimization is 6.72 Å × 10.85 Å × 16.06 Å, α = 90.00, β = 90.00, and γ = 102.01, in agreement with the experimental results.

As it is known that H2 evolution and the CO2RR go through a process containing proton–electron pair transfer, the computational hydrogen electrode (CHE) method thus was used to calculate the free energy of each intermediate state, of which the free energy of proton–electron pairs (H+ + e) equals image file: d2ta08092c-t6.tif. The Gibbs free energy change (ΔG) for each elemental step is calculated by

ΔG = ΔE + ΔEZPETΔS + ΔGpH
where ΔE is the electronic energy difference between free standing and adsorption states of reaction intermediates; ΔEZPE and ΔS are the changes of zero point energy and entropy, respectively, which are obtained based on vibrational calculations. T is the temperature and set to be 298 K in this work and pH is set to be 8 because acetonitrile is the solvent. ΔGpH = 2.303kBT × pH, where kB is the Boltzmann constant. The free energy of CO was calculated from the known free energy change (0.21 eV) of the reaction CO2(g) + 2H+ + 2e = CO(g) +H2O(l) under the standard conditions. The CO2RR involves two electron transfers, while the HER involves one electron transfers, and after the correction of ΔGpH, the final ΔG = 0.21 + 0.47 × 2 = 1.15 eV for the CO2RR and ΔG = 0.47 eV for the HER. The adsorption energy (Ead) is computed by using the formula:
 
Ead = Esub+AEsubEA(11)
where Esub+A, Esub and EA represent the total energies of the substrate with adsorbed species and the pure substrate without the adsorbate and adsorbate, respectively. And Gad is the corresponding adsorption free energy at 298 K.

Conclusions

In summary, a 2D cobalt-based metal–organic framework consisting of dinuclear Co2(COO)4 paddle-wheel SBUs and TBAPy4− ligands with extended π-systems was synthesized by a facile solvothermal method and used as a stable photocatalyst to photoreduce CO2 to syngas under visible light. Notably, the ratio of syngas as-prepared (CO/H2) can be adjusted in a wide range between 0.14 and 1.65 simply by adjusting the water content in the photocatalytic system. The proportion of syngas with CO[thin space (1/6-em)]:[thin space (1/6-em)]H2 = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 and 1[thin space (1/6-em)]:[thin space (1/6-em)]3 was achieved respectively, which is desirable for the practical applications of methanol and methane synthesis. Photocatalytic experiments, characterization and DFT calculation elucidate the probable electron transfer processes, and the thermodynamic and kinetic feasibility of Co-TBAPy serving as a photocatalyst. This work might present a new low-cost metal–organic framework semiconductor for the artificial photosynthesis of syngas and a more accessible preparation route of syngas at various ratios.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the NKRD Program of China (2021YFA1500401), the NSFC Projects (22171291, 21720102007, 21821003, and 21890380), the Local Innovative and Research Teams Project of Guangdong Pearl River Talents Program (2017BT01C161), and the FRF for the Central Universities. We thank Prof. Pengbo Lv of Xiangtan University for help in theoretical discussions.

References

  1. S. C. Peter, ACS Energy Lett., 2018, 3, 1557–1561 CrossRef CAS.
  2. R. P. Ye, J. Ding, W. Gong, M. D. Argyle, Q. Zhong, Y. Wang, C. K. Russell, Z. Xu, A. G. Russell, Q. Li, M. Fan and Y. G. Yao, Nat. Commun., 2019, 10, 5698 CrossRef CAS.
  3. S. Bai, H. Qiu, M. Song, G. He, F. Wang, Y. Liu and L. Guo, eScience, 2022, 2, 428–437 CrossRef.
  4. J. Li, Y. He, L. Tan, P. Zhang, X. Peng, A. Oruganti, G. Yang, H. Abe, Y. Wang and N. Tsubaki, Nat. Catal., 2018, 1, 787–793 CrossRef CAS.
  5. V. N. Nguyen and L. Blum, Chem. Ing. Tech., 2015, 87, 354–375 CrossRef CAS.
  6. J. S. Lee, D. I. Won, W. J. Jung, H. J. Son, C. Pac and S. O. Kang, Angew. Chem., Int. Ed., 2017, 56, 976–980 CrossRef CAS.
  7. M. Sun, C. Wang, C. Y. Sun, M. Zhang, X. L. Wang and Z. M. Su, J. Catal., 2020, 385, 70–75 CrossRef CAS.
  8. X. Wang, Z. L. Wang, Y. Bai, L. Tan, Y. Q. Xu, X. J. Hao, J. K. Wang, A. H. Mahadi, Y. F. Zhao, L. R. Zheng and Y. F. Song, J. Energy Chem., 2020, 46, 1–7 CrossRef.
  9. H. W. Zhang, J. T. Ming, J. W. Zhao, Q. Gu, C. Xu, Z. X. Ding, R. S. Yuan, Z. Z. Zhang, H. X. Lin, X. X. Wang and J. L. Long, Angew. Chem., Int. Ed., 2019, 58, 7718–7722 CrossRef CAS PubMed.
  10. C. Qiu, X. Hao, L. Tan, X. Wang, W. Cao, J. Liu, Y. Zhao and Y. F. Song, Chem. Commun., 2020, 56, 5354–5357 RSC.
  11. A. Li, T. Wang, X. X. Chang, Z. J. Zhao, C. C. Li, Z. Q. Huang, P. P. Yang, G. Y. Zhou and J. L. Gong, Chem. Sci., 2018, 9, 5334–5340 RSC.
  12. H. Furukawa, K. E. Cordova, M. O'Keeffe and O. M. Yaghi, Science, 2013, 341, 1230444 CrossRef.
  13. T. Ghanbari, F. Abnisa and W. M. A. Wan Daud, Sci. Total Environ., 2020, 707, 135090 CrossRef CAS.
  14. M. Pan, W.-M. Liao, S.-Y. Yin, S.-S. Sun and C.-Y. Su, Chem. Rev., 2018, 118, 8889–8935 CrossRef CAS PubMed.
  15. Y. Shi, A. F. Yang, C. S. Cao and B. Zhao, Coord. Chem. Rev., 2019, 390, 50–75 CrossRef CAS.
  16. I. I. Alkhatib, C. Garlisi, M. Pagliaro, K. Al-Ali and G. Palmisano, Catal. Today, 2020, 340, 209–224 CrossRef CAS.
  17. X. Feng, Y. Pi, Y. Song, C. Brzezinski, Z. Xu, Z. Li and W. Lin, J. Am. Chem. Soc., 2020, 142, 690–695 CrossRef CAS PubMed.
  18. Q.-L. Hong, H.-X. Zhang and J. Zhang, J. Mater. Chem. A, 2019, 7, 17272–17276 RSC.
  19. M. Liu, Y.-F. Mu, S. Yao, S. Guo, X.-W. Guo, Z.-M. Zhang and T.-B. Lu, Appl. Catal., B, 2019, 245, 496–501 CrossRef CAS.
  20. S. Fu, S. Yao, S. Guo, G.-C. Guo, W. Yuan, T.-B. Lu and Z.-M. Zhang, J. Am. Chem. Soc., 2021, 143, 20792–20801 CrossRef CAS PubMed.
  21. S. Guo, L.-H. Kong, P. Wang, S. Yao, T.-B. Lu and Z.-M. Zhang, Angew. Chem., Int. Ed., 2022, 61, e202206193 CAS.
  22. Y. Qin, Y. Wan, J. Guo and M. Zhao, Chin. Chem. Lett., 2022, 33, 693–702 CrossRef CAS.
  23. J.-W. Wang, L.-Z. Qiao, H.-D. Nie, H.-H. Huang, Y. Li, S. Yao, M. Liu, Z.-M. Zhang, Z.-H. Kang and T.-B. Lu, Nat. Commun., 2021, 12, 813 CrossRef CAS.
  24. T.-C. Zhuo, Y. Song, G.-L. Zhuang, L.-P. Chang, S. Yao, W. Zhang, Y. Wang, P. Wang, W. Lin, T.-B. Lu and Z.-M. Zhang, J. Am. Chem. Soc., 2021, 143, 6114–6122 CrossRef CAS PubMed.
  25. X.-K. Wang, J. Liu, L. Zhang, L.-Z. Dong, S.-L. Li, Y.-H. Kan, D.-S. Li and Y.-Q. Lan, ACS Catal., 2019, 9, 1726–1732 CrossRef CAS.
  26. Z. Wang, C.-Y. Zhu, H.-S. Zhao, S.-Y. Yin, S.-J. Wang, J.-H. Zhang, J.-J. Jiang, M. Pan and C.-Y. Su, J. Mater. Chem. A, 2019, 7, 4751–4758 RSC.
  27. H. Yang, D. Yang and X. Wang, Angew. Chem., Int. Ed., 2020, 59, 15527–15531 CrossRef CAS.
  28. J. Zhao, Q. Wang, C. Sun, T. Zheng, L. Yan, M. Li, K. Shao, X. Wang and Z. Su, J. Mater. Chem. A, 2017, 5, 12498–12505 RSC.
  29. Y. Xiao, Y. Qi, X. Wang, X. Wang, F. Zhang and C. Li, Adv. Mater., 2018, 30, e1803401 CrossRef.
  30. T. C. Wang, W. Bury, D. A. Gomez-Gualdron, N. A. Vermeulen, J. E. Mondloch, P. Deria, K. Zhang, P. Z. Moghadam, A. A. Sarjeant, R. Q. Snurr, J. F. Stoddart, J. T. Hupp and O. K. Farha, J. Am. Chem. Soc., 2015, 137, 3585–3591 CrossRef CAS.
  31. T. Islamoglu, S. Goswami, Z. Li, A. J. Howarth, O. K. Farha and J. T. Hupp, Acc. Chem. Res., 2017, 50, 805–813 CrossRef CAS.
  32. A. Gladysiak, T. N. Nguyen, R. Bounds, A. Zacharia, G. Itskos, J. A. Reimer and K. C. Stylianou, Chem. Sci., 2019, 10, 6140–6148 RSC.
  33. Y.-D. Huang, J.-H. Qin, X.-G. Yang, H.-R. Wang, F.-F. Li and L.-F. Ma, J. Solid State Chem., 2020, 285, 121252 CrossRef CAS.
  34. N. Li, J. Liu, J. J. Liu, L. Z. Dong, Z. F. Xin, Y. L. Teng and Y. Q. Lan, Angew. Chem., Int. Ed., 2019, 58, 5226–5231 CrossRef CAS.
  35. J. He, J. Wang, Y. Chen, J. Zhang, D. Duan, Y. Wang and Z. Yan, Chem. Commun., 2014, 50, 7063–7066 RSC.
  36. J. Schneider, H. Jia, J. T. Muckerman and E. Fujita, Chem. Soc. Rev., 2012, 41, 2036–2051 RSC.
  37. H. Xu, S. You, Z. Lang, Y. Sun, C. Sun, J. Zhou, X. Wang, Z. Kang and Z. Su, Chem.–Eur. J., 2020, 26, 2735–2740 CrossRef CAS.
  38. J. R. Rostrup-Nielsen, Catal. Today, 2000, 63, 159–164 CrossRef CAS.
  39. X. Gu, L. Qian and G. Zheng, Mol. Catal., 2020, 492, 110953 CrossRef CAS.
  40. A. N. Radhakrishnan, P. P. Rao, K. S. M. Linsa, M. Deepa and P. Koshy, Dalton Trans., 2011, 40, 3839–3848 RSC.
  41. B. Zhou, X. Zhao, H. J. Liu, J. H. Qu and C. P. Huang, Appl. Catal., B, 2010, 99, 214–221 CrossRef CAS.
  42. M.-J. Wei, J.-H. Zhang, W.-M. Liao, Z.-W. Wei, M. Pan and C.-Y. Su, J. Photochem. Photobiol., A, 2020, 387, 112137 CrossRef CAS.
  43. L. Dahéron, R. Dedryvère, H. Martinez, M. Ménétrier, C. Denage, C. Delmas and D. Gonbeau, Chem. Mater., 2008, 20, 583–590 CrossRef.
  44. G. Kresse and J. Hafner, Phys. Rev. B, 1993, 47, 558–561 CrossRef CAS PubMed.
  45. K. B. John, P. Perdew and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef PubMed.
  46. P. E. Blochl, Phys. Rev. B, 1994, 50, 17953–17979 CrossRef PubMed.
  47. Y. Kuramochi, O. Ishitani and H. Ishida, Coord. Chem. Rev., 2018, 373, 333–356 CrossRef CAS.
  48. Y. Yamazaki, H. Takeda and O. Ishitani, J. Photochem. Photobiol., C, 2015, 25, 106–137 CrossRef CAS.
  49. M. C. Figueiredo, I. Ledezma-Yanez and M. T. M. Koper, ACS Catal., 2016, 6, 2382–2392 CrossRef CAS.
  50. S. Chen, X. Li, C.-W. Kao, T. Luo, K. Chen, J. Fu, C. Ma, H. Li, M. Li, T.-S. Chan and M. Liu, Angew. Chem., Int. Ed., 2022, 61, e202206233 CAS.
  51. J. S. Derrick, M. Loipersberger, S. K. Nistanaki, A. V. Rothweiler, M. Head-Gordon, E. M. Nichols and C. J. Chang, J. Am. Chem. Soc., 2022, 144, 11656–11663 CrossRef CAS PubMed.
  52. A. Behera, A. K. Kar and R. Srivastava, Mater. Horiz., 2022, 9, 607–639 RSC.
  53. J.-C. Jiang, J.-C. Chen, M.-d. Zhao, Q. Yu, Y.-G. Wang and J. Li, Nano Res., 2022, 15, 7116–7123 CrossRef CAS.
  54. W. Gao, Z. Li, Q. Han, Y. Shen, C. Jiang, Y. Zhang, Y. Xiong, J. Ye, Z. Zou and Y. Zhou, Chem. Commun., 2022, 58, 9594–9613 RSC.
  55. A. A. Peterson, F. Abild-Pedersen, F. Studt, J. Rossmeisl and J. K. Nørskov, Energy Environ. Sci., 2010, 3, 1311–1315 RSC.
  56. Y. Lan, Y. Xie, J. Chen, Z. Hu and D. Cui, Chem. Commun., 2019, 55, 8068–8071 RSC.
  57. X. Zheng, Y. Yao, W. Ye, P. Gao and Y. Liu, Chem. Eng. J., 2021, 413, 128027 CrossRef CAS.
  58. Y.-R. Wang, Q. Huang, C.-T. He, Y. Chen, J. Liu, F.-C. Shen and Y.-Q. Lan, Nat. Commun., 2018, 9, 4466 CrossRef.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ta08092c
These authors contributed equally to this work and should be regarded as co-first authors.

This journal is © The Royal Society of Chemistry 2023