DOI:
10.1039/C4RA01413H
(Paper)
RSC Adv., 2014,
4, 18853-18861
Efficient photocatalytic hydrogen evolution with end-group-functionalized cobaloxime catalysts in combination with graphite-like C3N4†
Received
19th February 2014
, Accepted 9th April 2014
First published on 11th April 2014
Abstract
Three comparable hybrid photocatalytic systems, comprising semiconductor g-C3N4, end-group-functionalized cobaloxime complexes (carboxy-functionalized cobaloxime, C1; pyrene-functionalized cobaloxime, C2; and non-functionalized cobaloxime, C3), and triethanolamine (TEOA), are active for visible-light-driven hydrogen production in CH3CN–H2O (9/1, v/v) solution. Upon irradiation for 12 h, the turnover numbers of hydrogen evolution are 234, 281 and 195 for the hybrid systems C1/g-C3N4, C2/g-C3N4 and C3/g-C3N4, respectively. The highest hydrogen evolution efficiency of the C2/g-C3N4 system can be attributed to the strongest π–π interactions between the pyrene moiety and g-C3N4. Based on electrochemical properties, steady-state photoluminescence spectra and theoretical analyses, the visible light absorption of g-C3N4, the catalytic H2-evolving ability of cobaloxime as well as the efficient charge separation of the excited g-C3N4 in the presence of both TEOA and cobaloxime, are responsible for the high activity of these hybrid systems.
Introduction
The visible-light-driven hydrogen evolution from water is one of the promising approaches to provide renewable and clean energy for the future.1 Therefore, a number of hydrogen-evolving systems have been actively developed including homogeneous and heterogeneous photocatalytic systems in recent years.1d,2 Studies on the homogeneous catalysts for hydrogen production are mainly focused on the first-row transition metal complexes, such as [FeFe]-hydrogenase mimics,3 cobalt and nickel complexes.4 Among various catalysts, the cobaloxime has been widely investigated for photocatalytic hydrogen evolution due to their convenient preparation and good catalytic activities.5 Many cobaloxime-based photocatalytic systems for hydrogen production have been constructed utilizing the noble-metal-based photosensitizers, such as ruthenium,6 rhenium,7 platinum,8 and iridium-based organometallic complexes.9 However, because of the high cost of noble-metal-based photosensitizers, chemists attempt to replace them with other chromophores made of earth-abundant elements, such as organic xanthene dyes,10 metalloporphyrins,11 and photoactive nanomaterials.2b,12
To our knowledge, Eisenberg and co-workers have employed the xanthene dye derivatives (S or Se in place of O in the xanthene ring) in combination with the cobaloxime for the light-driven generation of hydrogen, and reported the highest turnover of 9000 (vs. organic dye).10b Besides the organic dyes, the metalloporphyrins were also introduced as the photosensitizers to couple with cobaloxime catalysts for efficient photoinduced hydrogen production.11a,c Nevertheless, one of the inherent weaknesses of the utilized organic and organometallic photosensitizers is their instability upon long-term irradiation.13 Recently, an increasing number of noble-metal-free nanomaterials have been used as the light absorber, such as TiO2,14 CdS,13a core/shell CdSe/ZnS QDs,15 and graphite-like C3N4 (referred to hereafter as g-C3N4), which can link cobaloximes directly onto the surface and overcome the drawback of the poor stability of the organic and organometallic photosensitizers.16
Among various nanomaterials, the environmentally benign g-C3N4 is rightfully attracting increased attention owing to its relatively high stability and suitable electronic structure (Eg = 2.7 eV, conduction band at −1.42 V and valence band at 1.28 V vs. Ag/AgCl) covering the water-splitting potentials.17 However, g-C3N4 alone shows very poor photocatalytic activities for water reduction and relies on surface co-catalysts, which can accelerate the separation of photogenerated electron/hole and increase the photocatalytic performance of g-C3N4. The precedents for catalytic systems which combined g-C3N4 with a co-catalyst, such as platinum group metals and MoS2,18 do exist but it is only recently that Co- and Ni-based molecular catalysts have been used as co-catalysts.13b,16 Herein, three comparable cobaloxime complexes (C1, C2 and C3) (Fig. 1) are synthesized and studied as molecular co-catalysts for hydrogen generation from hybrid systems containing the photoactive g-C3N4 and triethanolamine (TEOA) as electron donor in acetonitrile aqueous solution. The results of photocatalysis experiments show that the TONs of hydrogen evolution are 234 and 281 (vs. complex) for the hybrid systems C1/g-C3N4 and C2/g-C3N4, respectively. In comparison, the photocatalytic activity of the non-functionalized complex C3 (TON = 195) is lower than that of the carboxy-functionalized complex C1 and pyrene-functionalized complex C2 under the same conditions. To better understand the difference of photocatalytic activities, the adsorption of the complexes to g-C3N4, the electrochemical properties and the steady-state photoluminescence properties are investigated and discussed in detail.
 |
| Fig. 1 Structures of g-C3N4 and complexes C1 to C3 used in the present work. | |
Result and discussion
Synthesis and structure of catalysts and g-C3N4
The g-C3N4 was prepared by the direct polymerization of urea at ambient conditions. The XRD pattern of the as-prepared g-C3N4 shows two characteristic peaks at 13.1 and 27.4 degree in accord with the previously reported results (Fig. S1†).19 The TEM image clearly shows the g-C3N4 possesses the layered structure (Fig. S2†). In order to investigate the effect of end-group-functionalized cobaloximes on the photocatalytic performance of as-prepared g-C3N4, three cobaloxime complexes are designed as co-catalysts and expected to assemble on the surface of g-C3N4 for photocatalytic H2 evolution. The first one (C1) is based on the relatively good adsorption performance of carboxy group, which can provide a linkage to the nanomaterial surface.20 Furthermore, in view of the conjugative π structure of g-C3N4,21 we choose another cobaloxime derivative C2, in which the pyrene moiety of π electron conjugation is expected to strongly interact with g-C3N4 via π–π interactions. As a comparison with the end-group-functionalized cobaloxime complexes C1 and C2, the complex C3 without functional-group modification is also employed as co-catalyst for g-C3N4 in this paper.
The C1 was conveniently prepared by stirring [Co(dmgH)(dmgH2)Cl2] with isonicotinic acid in CH3CN and then obtained as dark-red crystals. The C2 was prepared in an analogous manner to C1 using [Co(dmgH)(dmgH2)Cl2] and N-pyren-1-ylmethyl-isonicotinamide, which was synthesized by the amidation reaction between isonicotinoyl chloride and (pyren-1-ylmethyl)amine. All new compounds were confirmed by 1H, 13C NMR and elemental analysis. The solid-state structures of C1 and C2 were established by means of single-crystal X-ray diffraction. The crystal structures of complexes C1 and C2 are given in Fig. 2 with selected bond lengths and angles listed in Table 1. The coordination geometry around the Co-centre in C1 is strongly similar to those in other reported cobaloxime complexes.8b Because of the existence of the carboxyl, the axial Co–N(5)pyridine distance is 1.967 Å, slightly longer than that in C3 (1.959 Å). The average Co–Nimine bond distance of the glyoximate ligands in C1 is 1.896 Å, which is consistent to the value of 1.895 Å in C3.22 The asymmetric unit of C2 consists of two same Co-centred complexes and some uncoordinated solvent molecules. For clarity, we only present one Co-centre in Fig. 2, which is hexacoordinated with a slightly distorted octahedral geometry (Table 1). The pendent pyrene moiety is roughly parallel to the equatorial plane defined by the four N atoms of glyoximate ligands, with a dihedral angle of 5.6 degree.
 |
| Fig. 2 Crystal structures of C1 and C2 (the hydrogen atoms are omitted for clarity). | |
Table 1 Selected bond lengths (Å) and angles (degree) for C1 and C2
C1 |
Co(1)–N(1) |
1.910(2) |
N(3)–Co(1)–N(4) |
98.64(10) |
Co(1)–N(2) |
1.887(2) |
N(2)–Co(1)–N(4) |
178.48(11) |
Co(1)–N(3) |
1.882(2) |
N(3)–Co(1)–N(1) |
177.92(11) |
Co(1)–N(4) |
1.906(2) |
N(2)–Co(1)–N(1) |
99.33(10) |
Co(1)–N(5) |
1.967(3) |
N(4)–Co(1)–N(1) |
80.26(11) |
Co(1)–Cl(1) |
2.2301(8) |
N(3)–Co(1)–N(5) |
89.63(10) |
O(1)–N(1) |
1.364(3) |
N(2)–Co(1)–N(5) |
89.41(10) |
O(2)–N(2) |
1.340(3) |
N(4)–Co(1)–N(5) |
92.07(10) |
O(3)–N(3) |
1.336(3) |
N(1)–Co(1)–N(5) |
92.17(10) |
O(4)–N(4) |
1.352(3) |
N(3)–Co(1)–N(2) |
81.72(10) |
|
C2 |
Co(1)–N(1) |
1.895(4) |
N(3)–Co(1)–N(1) |
177.37(18) |
Co(1)–N(2) |
1.904(4) |
N(3)–Co(1)–N(4) |
82.11(17) |
Co(1)–N(3) |
1.886(4) |
N(1)–Co(1)–N(4) |
98.27(17) |
Co(1)–N(4) |
1.898(4) |
N(3)–Co(1)–N(2) |
98.08(17) |
Co(1)–N(9) |
1.974(4) |
N(1)–Co(1)–N(2) |
81.48(17) |
Co(1)–Cl(1) |
2.2241(13) |
N(4)–Co(1)–N(2) |
178.84(17) |
O(1)–N(1) |
1.362(5) |
N(3)–Co(1)–N(9) |
90.92(17) |
O(2)–N(2) |
1.337(5) |
N(1)–Co(1)–N(9) |
91.68(17) |
O(3)–N(3) |
1.330(5) |
N(4)–Co(1)–N(9) |
90.07(15) |
O(4)–N(4) |
1.343(5) |
N(2)–Co(1)–N(9) |
91.07(16) |
Table 2 Influence of the concentration of C1 and the amount of g-C3N4 on photocatalytic H2 productiona
Run |
Concentration of C1 (mM) |
Amount of g-C3N4 (mg) |
TON |
Condition: 5 vol% TEOA in CH3CN–H2O (9/1, v/v) at pH 10; irradiation time 4 h. |
1 |
0.05 |
4 |
56 |
2 |
0.1 |
4 |
84 |
3 |
0.2 |
4 |
40 |
4 |
0.1 |
2 |
57 |
5 |
0.1 |
8 |
92 |
Adsorption of the cobaloximes C1–C3 to g-C3N4
The UV-vis diffuses reflectance absorption spectrum of g-C3N4 exhibits the broad absorption from UV to visible light region (Fig. S4†). The absorption edge is at about 455 nm, corresponding to a small band-gap (2.73 eV). The UV-vis absorption spectra of complexes C1–C3 measured in CH3CN are shown in Fig. 3a. The cobaloximes C1 and C3 exhibit a high-energy absorption between 230 nm and 300 nm, which is attributed to the intra-ligand (π–π*) transitions.8b Unlike C1 and C3, the C2 displays a very different absorption spectrum with characteristic absorption bands of the pyrene group (at 342, 326, 312, 276, 265, 255, 242, and 235 nm).23 According to the report by Li et al., the cobaloximes bearing appropriate pendent ligands can be effectively adsorbed by the photoactive nanomaterials.13a To investigate the adsorption amount of the cobaloximes C1–C3 on the g-C3N4 surface, the acetonitrile solutions of C1–C3 (2 × 10−5 M, 5 mL) were measured by UV-vis spectrophotometry before and after exposure to g-C3N4 (Fig. 3). In view of the absorbance difference at the maximum adsorption wavelength, we estimate the adsorption amount of C1, C2, and C3 on g-C3N4 to be approximately 0.05, 0.07, and 0.02 μmol, respectively. Increasing the original concentration of cobaloximes from 2 × 10−5 M to 3 × 10−5 M did not lead to the increase in adsorption amount. As noted, however, no significant difference in the adsorption amount was seen between the above-mentioned cobaloximes, with a median adsorption amount for each of about 0.04 μmol.
 |
| Fig. 3 (a) UV-vis absorption spectra of C1 (2 × 10−5 M), C2 (2 × 10−5 M) and C3 (2 × 10−5 M) in CH3CN (5 mL); UV-vis absorption spectra of C2 (b), C1 (c), and C3 (d) before and after stirred with g-C3N4 (4 mg) in CH3CN for 2 h following centrifugation and filtration. | |
Photocatalytic activities of the hybrid systems
We first studied the photocatalytic H2 production of the hybrid system C1/g-C3N4 in CH3CN–H2O (9
:
1, v/v, 5 mL) solution containing 5 vol% TEOA at pH 10. Experiments under irradiation (λ > 400 nm) with different concentration of C1 (0.05–0.2 mM) showed the system with 0.1 mM C1 released the largest amount of H2 (42 μmol, TON = 84) over 4 h (Table 2). With an increase or a decrease of the concentration of C1 to 0.2 mM and 0.05 mM, the TON of H2 evolution decreased to 40 and 56, respectively. It is evident that the optimum amount (0.5 μmol) of C1 for the system is much higher than the adsorption amount (0.05 μmol) of C1 on the surface of g-C3N4, which indicates that the complex C1 predominantly exists as free molecules in reaction solution. To further clarify the role of C1, the suspension of the C1/g-C3N4 system was stirred for 2 h in the dark and subsequently filtered, resulting in the g-C3N4 residue and the clear filtrate. With readdition of reaction solution, the g-C3N4 residue adsorbing a small amount of C1 (0.05 μmol) gave a quarter of the TON of the original C1/g-C3N4 system, while with readdition of fresh g-C3N4, the clear filtrate containing the free molecules displayed a H2 evolution rate similar to that of the original system C1/g-C3N4 (Fig. S6†). These results indicate that both of the C1 molecules, adsorbed on the surface of g-C3N4 and dissolved in reaction solution, can improve the photocatalytic activities of g-C3N4. The light-induced H2 production catalyzed by C1 also depends on the amount of g-C3N4. When the concentration of C1 was 0.1 mM, the amount of H2 evolved was improved apparently with an increase of the amount of g-C3N4 from 2 mg to 4 mg. However, further increasing the amount of g-C3N4 to 8 mg did not lead to an obvious enhancement in the photocatalytic activity. Control experiments without C1, g-C3N4, or TEOA showed no H2 was released from the system C1/g-C3N4, suggesting that all three components are required for the photocatalytic H2 evolution.
Besides CH3CN, CH3CH2OH and DMF were also used as reaction solvent for the photocatalytic experiment in the C1/g-C3N4 system. Smaller quantities of hydrogen were released in systems containing either CH3CH2OH or DMF, albeit with the same 9/1 v/v ratio with water (Fig. 4). Consequently, the solvent CH3CN is superior to CH3CH2OH and DMF for H2 evolution in this system. Further studies on solvent effects were done to investigate the photocatalytic activity of the system C1/g-C3N4 by changing the ratio of CH3CN and H2O. The results showed that the C1/g-C3N4 system displayed a higher activity when the CH3CN–H2O ratio was increased. With an increase of CH3CN–H2O ratio to 9/1 from 1/1, the TON of H2 evolution was apparently increased to 88 from 10. In addition to the system C1/g-C3N4, the effects of the two factors (i.e., organic solvents and CH3CN–H2O ratios) on the photocatalytic performances in the analogous systems C2/g-C3N4 and C3/g-C3N4 were also studied in Fig. 4. It is found that the trends of the medium dependence on hydrogen evolution over the C2/g-C3N4 and C3/g-C3N4 systems agree closely with those over the C1/g-C3N4 system (Fig. 4).
 |
| Fig. 4 Effects of organic solvents and CH3CN–H2O ratios on the hydrogen evolution over each system containing cobaloxime (1 × 10−4 M), g-C3N4 (4 mg) and TEOA (5 vol%) at pH 10; irradiation time 6 h. | |
Considering the aforementioned catalytic condition, the pH value of the C1/g-C3N4 system is approximately 10 without any adjustment by acid or base. To examine the pH effects on H2 evolution from the system C1/g-C3N4, light-induced H2 evolution was performed in the pH range of 8.5–10 under the conditions described in Fig. 5. The results show that the photocatalytic H2 production is dependent on the pH value of the system, in which the optimum pH value for H2 evolution is 9, while lower amounts of H2 are obtained at either lower or higher pH values. The mechanism on the similar pH-dependent phenomenon has been elucidated in more detail by Zhang et al.10c In addition, the pH dependence on hydrogen production over the systems C2/g-C3N4 and C3/g-C3N4 also shows that the maximum hydrogen generation efficiency was achieved at pH 9. This pH-dependent effect is related to the oxidation of TEOA, which is an essential step in the catalytic cycle.8a As we know, TEOA is extensively used as electron donor in many Pt/g-C3N4-based hydrogen evolution systems.19a,21a,24 In the C1/g-C3N4 system, keeping the concentration of C1 (1 × 10−4 M) and the amount of g-C3N4 (4 mg) unchanged, increasing the concentration of TEOA from 2.5 to 5 vol% apparently improved the efficiency of H2 production. Further increasing the concentration of TEOA to 10 vol%, however, led to no obvious increase in the amount of H2 production (Fig. S9†). As a result, an optimized H2 evolution system containing 1 × 10−4 M C1, 4 mg g-C3N4, and 5 vol% TEOA in 5 mL CH3CN–H2O (9/1, v/v) at pH 9.0 was able to photocatalyze H2 evolution with a TON of 234.
 |
| Fig. 5 pH dependence on the hydrogen production over each system under the following conditions: TEOA (5 vol%), cobaloxime (1 × 10−4 M) and g-C3N4 (4 mg) in CH3CN–H2O (9/1, v/v) solution; irradiation time 12 h. | |
To examine the effect of carboxyl group on the photocatalytic activity, the photocatalytic H2 production experiment by the C3/g-C3N4 system was studied under the same conditions. A total amount of evolved H2 reached 97 μmol, corresponding to a TON of 195. The activity of the C1/g-C3N4 system is higher than that of the C3/g-C3N4 system, indicating that the carboxy group functions satisfactorily as a linkage for H2 evolution. Furthermore, it was found that the C2/g-C3N4 system displayed the highest activity among these three complexes, with TONs up to 281(Fig. 6). As expected, both of the C2 molecules, adsorbed on g-C3N4 and dissolved in solution, also displayed the highest H2 evolution efficiency among three (Fig. S7†). These results indicate that the C2 is the best co-catalyst for H2 production in the current photocatalytic system, presumably due to π–π interactions between C2 and g-C3N4.
 |
| Fig. 6 Time dependence of hydrogen production using different cobaloxime complexes C1 to C3 (1 × 10−4 M) under the following conditions: g-C3N4 (4 mg) and TEOA (5 vol%) in CH3CN–H2O (9/1, v/v) solution (5 mL) at pH 9, error bars represent standard errors of the means of three independent experiments. | |
The photocatalytic H2 production of the C1/g-C3N4 system leveled out after 9 h of irradiation (Fig. 6). To investigate the reason(s) for the cease of H2 generation, after the photocatalysis reaction, the suspension was filtered and the separated solid was washed with water and dried in vacuum. The UV-vis absorption spectrum of the resulting clear solution is different from that of the original C1 (Fig. S5†). Thus we speculate that the cessation of H2 production may be ascribed to the degradation of C1. Besides the degradation of C1 after photocatalysis reaction, the pH value of system changed to 9.6 from 9 after photocatalytic process. After 12 h of irradiation, to recover H2 evolution of the C1/g-C3N4 system, the photocatalytic experiment with readdition of C1 (0.5 μmol) to the system showed a recovered but lowered activity for hydrogen evolution under irradiation (Fig. S10†). For comparison, CoCl2 was also tested for the photocatalytic H2 evolution with g-C3N4 under the same conditions. The results show that the photocatalytic activity of the CoCl2/g-C3N4 system is fairly lower than that of the C1/g-C3N4 system (Fig. S11†), suggesting that the active species in the current system is the cobaloxime C1 rather than the decomposed products. The TEM image of the separated solid is as shown in Fig. S3,† which is similar to the TEM image before irradiation, indicating that g-C3N4 is stable in the photocatalytic process. No zero-valent cobalt colloids were observed in the TEM image, which further proved the active species is also not the zero-valent cobalt in this system. The XRD pattern of the isolated solid is very similar to that of the freshly prepared g-C3N4, except for a decrease in the intensity of the peak at 27.4 degree (Fig. S1†).
Electrochemical properties
The redox potential of the complexes is crucial to understand the electron transfer in photocatalytic process. Cyclic voltammetry studies were performed in CH3CN with Bu4NPF6 as supporting electrolyte under N2. The electrochemical reduction processes of C1–C3 are shown in Fig. S12.† The reduction potentials are versus Fc+/Fc and are summarized in Table 3. All CoIII complexes have two reductions, which can be assigned to CoIII/CoII and CoII/CoI, respectively.25 The first irreversible reduction potentials of C1, C2 and C3 are confirmed to be at about −1.06, −1.13, and −1.22 V, respectively. The first potentials of C1 and C2 are anodic shift relative to C3, which may be attributable to the axial pyridine ligand modified by electron-withdrawing group (–COOH or –CONH). But there is no above trend in the second potentials for CoII/CoI couple. The potentials at the secondary reduction are nearly equal for the three complexes. The behaviour of electrocatalytic proton reduction by C1 or C2 was studied by cyclic voltammograms in the presence of CH3COOH (Fig. S13 and S14†). Overall, complexes C1 and C2 behaved analogously toward electrocatalytic proton reduction.
Table 3 Electrochemical potentialsa (vs. Fc+/Fc) for all cobaloximes
Cobaloximes |
CoIII/CoII E1/2b |
CoII/CoI E1/2 |
ΔGc |
Electrochemical potentials were obtained by cyclic voltammetry studies under a N2 atmosphere with 0.1 M Bu4NPF6 as the supporting electrolyte in CH3CN. Irreversible reduction wave. ΔG = the potential of CB(g-C3N4) − E1/2(CoII/CoI), CB(g-C3N4) = −1.87 V. |
C1 |
−1.06 |
−1.56 |
−0.31 |
C2 |
−1.13 |
−1.48 |
−0.39 |
C3 |
−1.22 |
−1.50 |
−0.37 |
The probable mechanism on the photocatalytic H2 production
The estimated reduction potential of g-C3N4 is −1.42 V relative to Ag/AgCl,18a which can be adjusted to −1.87 V versus Fc+/Fc (E0 of Fc+/Fc = 0.45 V vs. Ag/AgCl).26 The reduction potentials of the catalysts C1–C3 are given in Table 3. According to Weller-equation, the values of free-energy changes (ΔG) for formation of CoI species are determined to be −0.31, −0.39, and −0.37 V for each system, respectively (Table 3). The negative ΔG suggests that it is thermodynamically feasible for photoinduced electron transfer from the conduction band of g-C3N4 to these catalysts. However, there is no significant difference in the ΔG value between the three hybrid systems.
To explore the charge transfer pathways, the steady-state photoluminescence properties of the C2/g-C3N4 system were investigated in acetonitrile aqueous solution. The g-C3N4 suspension excited at 375 nm results in a maximal fluorescence emission peak at 459 nm (Fig. 7a). No significant decreases in fluorescence intensity are observed upon addition of either C2 or TEOA. The reason for this phenomenon may plausibly be attributed to the fast recombination of photogenerated electron/hole pairs. However, a larger fluorescence quenching of g-C3N4 by 25% upon addition of both TEOA and C2 is observed, suggesting that both C2 and TEOA are required for the separation of the photogenerated electron/hole. Based on hydrogen production experiment and the above spectroscopic study, a plausible pathway of the photocatalytic H2 evolution reaction is proposed in Fig. 8. Under visible light irradiation, the excited state electrons of the valence band of g-C3N4 would transport to the conduction band. The conduction-band electron of g-C3N4 can give two electrons to C2 to form to CoI species. The photogenerated CoI species is protonated to obtain a Co–H intermediate, which finally reacts with another proton to generate H2.5a Meanwhile the electron donor TEOA is oxidized by the photogenerated holes in the valence band.
 |
| Fig. 7 (a) The emission spectra of g-C3N4 (0.8 g L−1, 5 mL) suspension in CH3CN–H2O (9/1, v/v) solution at pH 9 in the presence of C2 (1 × 10−4 M), TEOA (5 vol%), TEOA (5 vol%) and C2 (1 × 10−4 M); (b) the fluorescence quenching of g-C3N4 in the presence of C1 (1 × 10−4 M), C2 (1 × 10−4 M), and C3 (1 × 10−4 M) in 5 vol% TEOA acetonitrile aqueous solution. | |
 |
| Fig. 8 A proposed pathway for the photocatalytic H2 evolution by cobaloxime–g-C3N4 systems. | |
In contrast, the fluorescence intensity of g-C3N4 decreased by 20% and 14% upon addition of C1 + TEOA and C3 + TEOA (Fig. 7b), respectively. The sequence of the effectiveness quenching fluorescence of g-C3N4 is C2 > C1 > C3, which corresponds to the order of their photocatalytic H2 evolution activity. According to previous reports on hybrid photocatalytic systems containing cobaloxime, the photocatalytic activity could be influenced by the adsorption amount of cobaloximes on nanomaterials, the driving force (ΔG) of electron transfer and the attachment mode between cobaloxime and nanomaterials.13a,14a,15,16 However, no apparent difference in terms of the adsorption amount and the driving force was found between the present three systems. Therefore, the reason for the activity difference may be attributed to the attachment mode between g-C3N4 and cobaloximes. Compared with the free-collision mode between the non-functionalized C3 and g-C3N4, the direct attachment (i.e., carboxy linkage or π–π interactions) between end-group functionalized cobaloxime and g-C3N4 could promote the charge transfer from g-C3N4 to the co-catalytic centres.
Conclusions
Three end-group-functionalized cobaloxime complexes as co-catalyst were designed and assembled on the surface of g-C3N4 for photocatalytic H2 evolution in CH3CN–H2O (9/1, v/v) solution at pH 9. The photocatalytic experiments gave the highest TON of 234, 281 and 195 for C1/g-C3N4, C2/g-C3N4 and C3/g-C3N4 hybrid systems, respectively. The photocatalytic activity of these systems was attributed to the efficient charge separation of the excited g-C3N4, which was proved by steady-state photoluminescence spectra. Among the cobaloximes, the C2 was superior to C1 and C3 as co-catalyst for H2 production, which may result from the π–π interactions between C2 and g-C3N4. Due to the degradation of cobaloxime, the catalytic lifetime of the hybrid system is short, which motivates us to seek more stable and effective molecule co-catalysts for g-C3N4.
Experimental section
Chemicals
All chemicals in this work were purchased from commercial sources and used without further purification. (Pyren-1-ylmethyl)amine,27 complexes [Co(dmgH) (dmgH2)Cl2],28 and [Co(dmgH)2pyCl] (C3)28 were synthesized according to previously described procedures.
Synthesis
The preparation of g-C3N4 was following a previously reported method.19 Briefly, 10 g dry urea was added to an alumina crucible with a cover under ambient pressure in air. The precursor was put in a muffle furnace and heated to 550 °C in 2 h and maintained at this temperature for 3 h. The obtained powders were rinsed several times with distilled water and ethanol and dried in vacuum at 80 °C for 8 h.
N-Pyren-1-ylmethyl-isonicotinamide. A mixture of isonicotinic acid (0.30 g, 2.44 mmol) and thionyl chloride (20 mL) in a round-bottomed flask was refluxed for 5 h with stirring. Excess thionyl chloride was removed under reduced pressure until yellow powders of isonicotinoyl chloride were obtained. A solution of the prepared isonicotinoyl chloride in 20 mL of CH2Cl2 was treated with a mixed solution of (pyren-1-ylmethyl)amine (0.46 g, 2.00 mmol) and triethylamine (3 mL) in 20 mL of CH2Cl2 with vigorous stirring at 0 °C. After addition, the resulting mixture was allowed to react for another 8 h at room temperature. Finally, the solvent was evaporated, and the residue was purified by chromatography on a silica gel column. Elution with CH3OH–CH2Cl2 (1
:
5, v/v) furnished pure N-pyren-1-ylmethyl-isonicotinamide (0.61 g, 91%). Anal. calcd for C23H16N2O: C, 82.12%; H, 4.79%; N, 8.33%. Found: C, 82.09%; H, 4.81%; N, 8.29%. 1H NMR (400 MHz, CDCl3) δ 8.64 (2H, d, J = 5.4 Hz), 8.27 (1H, d, J = 9.2 Hz), 8.24–8.19 (2H, m), 8.16 (2H, d, J = 8.6 Hz), 8.05 (4H, dt, J = 16.9 Hz, 8.7 Hz), 7.58 (2H, d, J= 5.6 Hz), 6.63 (1H, s), 5.34 (2H, d, J = 5.1 Hz). 13C NMR (101 MHz, CDCl3) δ 165.06, 150.28, 141.50, 131.55, 131.23, 130.68, 130.06, 129.19, 128.64, 127.85, 127.49, 127.31, 126.28, 125.67, 125.56, 125.11, 124.86, 124.66, 122.51, 120.97, 42.80.
[Co(dmgH)2(4-COOH-py)Cl] (C1). A 0.36 g (1.00 mmol) sample of [Co(dmgH)(dmgH2)Cl2] was suspended in 40 mL of CH3CN. One equivalent of isonicotinic acid (0.12 g, 1.00 mmol) was then added to the flask and heated at 70 °C with stirring for 1 h. The resulting dark-red solution was filtered and left overnight. Dark-red crystals (0.39 g, 87%) of suitable for X-ray analysis were obtained. Anal. calcd for CoClC14N5O6H19·H2O: C, 36.10%; H, 4.54%; N, 15.04%. Found: C, 36.47%; H, 4.44%; N, 15. 16%. 1H NMR (400 MHz, CD3CN) δ 8.30 (2H, d, J = 6.8 Hz), 7.72 (2H, d, J = 6.7 Hz), 2.31 (12H, s).
[Co(dmgH)2(N-pyren-1-ylmethyl-isonicotinamide)Cl] (C2). Triethylamine (0.10 g, 1.00 mmol) was added to a stirred, green suspension of [CoCl2(dmgH)(dmgH2)] (0.36 g, 1.00 mmol) in CH3OH (25 mL), resulting in a brown solution. A solution of N-pyren-1-ylmethyl-isonicotinamide (0.33 g, 1.00 mmol) in 15 mL of CH2Cl2 was added after five minutes and the reaction mixture was heated at 40 °C for 1 h. The resulting brown solution was filtered and left overnight. Brown crystals (0.50 g, 76%) of suitable for X-ray analysis were obtained. Anal. calcd for CoClC31N6O5H30: C, 56.33%; H, 4.57%; N, 12.71%. Found: C, 56.47%; H, 4.62%; N, 12.60%. 1H NMR (400 MHz, CDCl3) δ 8.32 (2H, d, J = 4.8 Hz), 8.22 (3H, d, J = 7.8 Hz), 8.14 (2H, d, J = 9.2 Hz), 8.11–8.00 (3H, m), 7.98 (1H, d, J = 7.4 Hz), 7.55 (2H, d, J = 4.8 Hz), 6.71 (1H, s), 5.30 (2H, s), 2.31 (12H, s).Adsorption experiments: a solution of C1, C2 or C3 (2 × 10−5 M, 5 mL) in CH3CN solution was added to g-C3N4 (4 mg) in a Schlenk tube. The mixture was stirred for 2 h in the dark and subsequently centrifuged. The filtered, clear solution was then measured by UV-visible absorption spectra. The amount of C1, C2 or C3 adsorbed on g-C3N4 was estimated by the absorbance difference at the absorption peak before and after adsorption on to g-C3N4.
Physical measurements
Elemental analyses were carried out on a Vario MICRO Elemental Analyser. 1H, 13C NMR spectra were performed on a Bruker Avance III (400 MHz) spectrometer. TEM analyses were conducted on a JEM-2010 electron microscope at an acceleration voltage of 200 kV. UV-vis absorption spectra were measured on a Perkin-Elmer Lambda 35 UV-vis spectrophotometer.
X-ray diffraction
X-ray single-crystal data of C1 and C2 were collected on a SuperNova, Dual, Cu at zero, Atlas diffractometer. Crystal data collection, refinement and reduction were accomplished with the CrysAlisPro, Agilent Technologies, Version 1.171.36.28. The crystal structures were solved by direct methods with SHELXS-97 and refined by using the SHELXL-97 crystallographic software package. All non-hydrogen atoms were refined anisotropically. The hydrogen atoms were added in a riding model. Details of crystal data are summarized in Table 4. X-ray diffraction (XRD) was used to identify the structure of the g-C3N4 before and after photocatalytic reactions. Diffraction date were collected on a MiniFlex II diffractometer with Cu Kα radiation (30 kV × 15 mA). The 2θ scanning range was from 10 degree to 50 degree with a scanning speed of 4 degree per min.
Table 4 Crystal data and structure refinement details for complex C1 and C2
Complex |
C1·(H2O)(CH3CN) |
C2·(CH3OH)(H2O)2(CH2Cl2) |
Molecule formula |
C16H24ClCoN6O7 |
C64H70Cl4Co2N12O13 |
Formula weight |
506.79 |
1474.98 |
T (K) |
100.00(13) |
100.0(2) |
Crystal system |
Triclinic |
Triclinic |
Space group |
P![[1 with combining macron]](https://www.rsc.org/images/entities/char_0031_0304.gif) |
P![[1 with combining macron]](https://www.rsc.org/images/entities/char_0031_0304.gif) |
a (Å) |
7.9393(3) |
9.5141(4) |
b (Å) |
11.3829(4) |
17.4267(9) |
c (Å) |
13.7720(6) |
20.3947(10) |
α (°) |
111.729(4) |
75.820(5) |
β (°) |
101.772(3) |
87.861(4) |
γ (°) |
99.800(3) |
74.763(4) |
V (Å3) |
1089.74(7) |
3161.8(3) |
Z |
2 |
2 |
F (000) |
524 |
1528 |
Dcalc (g cm−3) |
1.544 |
1.549 |
R1/wR2 (I > 2σ(I)) |
0.0394/0.0860 |
0.0715/0.1822 |
R1/wR2 (all data) |
0.0447/0.0883 |
0.0994/0.2066 |
Goodness of fit |
1.084 |
1.032 |
Electrochemistry
Electrochemical measurements were made using a CH instrument Model 630A Electrochemical Workstation. The cyclic voltammetry experiments were conducted in acetonitrile solution containing 0.1 M nBu4NPF6 as the supporting electrolyte under N2. Glassy carbon and platinum wire were used as the working and counter electrodes, respectively, and the potential was measured against the Ag/AgCl reference electrode and reported relative to the internal reference of Fc+/Fc = 0.00 V.
Photocatalysis
In a typical procedure, g-C3N4 (4 mg), C1 (500 μL, 1 mM) and 5 vol% triethanolamine (TEOA) acetonitrile aqueous (9/1, v/v) solution were added to a Schlenk bottle. The mixture was magnetically stirred under N2 atmosphere for 15 min. The system was freeze–pump–thaw degassed for three times and then warmed to room temperature prior to irradiation. The reaction solution was irradiated at 25 °C using a Xe lamp (300 W) with a cutoff filter (λ > 400 nm). The gas phase of the reaction system was analyzed on a GC 7900 instrument with a 5 Å molecular sieve column, a thermal conductivity detector, and using N2 as carrying gas. The amount of hydrogen generated was determined by the external standard method. Hydrogen dissolved in the solution was not measured and the slight effect of the hydrogen generated on the pressure of the Schlenk bottle was neglected for calculation of the volume of hydrogen gas.
Acknowledgements
The authors thank the National Natural Science Foundation China (grant numbers 21071145, 21101153, 21231003 and 21203195) for financial support.
Notes and references
-
(a) N. S. Lewis and D. G. Nocera, Proc. Natl. Acad. Sci. U. S. A., 2006, 103, 15729–15735 CrossRef CAS PubMed;
(b) R. M. Navarro, M. C. Alvarez-Galvan, J. A. Villoria de la Mano, S. M. Al-Zahrani and J. L. G. Fierro, Energy Environ. Sci., 2010, 3, 1865–1882 RSC;
(c) E. S. Andreiadis, M. Chavarot-Kerlidou, M. Fontecave and V. Artero, Photochem. Photobiol., 2011, 87, 946–964 CrossRef CAS PubMed;
(d) W. T. Eckenhoff and R. Eisenberg, Dalton Trans., 2012, 41, 13004–13021 RSC.
-
(a) A. Kudo, Int. J. Hydrogen Energy, 2007, 32, 2673–2678 CrossRef CAS PubMed;
(b) P. D. Tran, L. H. Wong, J. Barber and J. S. C. Loo, Energy Environ. Sci., 2012, 5, 5902–5918 RSC;
(c) V. Artero and M. Fontecave, Chem. Soc. Rev., 2013, 42, 2338–2356 RSC.
-
(a) M. Wang, L. Chen, X. Li and L. Sun, Dalton Trans., 2011, 40, 12793–12800 RSC;
(b) F. Wang, W. G. Wang, H. Y. Wang, G. Si, C. H. Tung and L. Z. Wu, ACS Catal., 2012, 2, 407–416 CrossRef CAS.
-
(a) S. Losse, J. G. Vos and S. Rau, Coord. Chem. Rev., 2010, 254, 2492–2504 CrossRef CAS PubMed;
(b) W. Zhang, J. Hong, J. Zheng, Z. Huang, J. Zhou and R. Xu, J. Am. Chem. Soc., 2011, 133, 20680–20683 CrossRef CAS PubMed;
(c) Z. Han, W. R. McNamara, M. S. Eum, P. L. Holland and R. Eisenberg, Angew. Chem., Int. Ed., 2012, 51, 1667–1670 CrossRef CAS PubMed;
(d) H. H. Cui, J. Y. Wang, M. Q. Hu, C. B. Ma, H. M. Wen, X. W. Song and C. N. Chen, Dalton Trans., 2013, 42, 8684–8691 RSC;
(e) H. N. Kagalwala, E. Gottlieb, G. Li, T. Li, R. Jin and S. Bernhard, Inorg. Chem., 2013, 52, 9094–9101 CrossRef CAS PubMed.
-
(a) V. Artero, M. Chavarot-Kerlidou and M. Fontecave, Angew. Chem., Int. Ed., 2011, 50, 7238–7266 CrossRef CAS PubMed;
(b) M. Wang, L. Chen and L. Sun, Energy Environ. Sci., 2012, 5, 6763–6778 RSC.
- A. Fihri, V. Artero, M. Razavet, C. Baffert, W. Leibl and M. Fontecave, Angew. Chem., Int. Ed., 2008, 120, 574–577 CrossRef.
-
(a) B. Probst, M. Guttentag, A. Rodenberg, P. Hamm and R. Alberto, Inorg. Chem., 2011, 50, 3404–3412 CrossRef CAS PubMed;
(b) M. Guttentag, A. Rodenberg, R. Kopelent, B. Probst, C. Buchwalder, M. Brandstätter, P. Hamm and R. Alberto, Eur. J. Inorg. Chem., 2012, 2012, 59–64 CrossRef CAS.
-
(a) P. Du, K. Knowles and R. Eisenberg, J. Am. Chem. Soc., 2008, 130, 12576–12577 CrossRef CAS PubMed;
(b) P. Du, J. Schneider, G. Luo, W. W. Brennessel and R. Eisenberg, Inorg. Chem., 2009, 48, 4952–4962 CrossRef CAS PubMed.
- A. Fihri, V. Artero, A. Pereira and M. Fontecave, Dalton Trans., 2008, 5567–5569 RSC.
-
(a) T. Lazarides, T. McCormick, P. Du, G. Luo, B. Lindley and R. Eisenberg, J. Am. Chem. Soc., 2009, 131, 9192–9194 CrossRef CAS PubMed;
(b) T. M. McCormick, B. D. Calitree, A. Orchard, N. D. Kraut, F. V. Bright, M. R. Detty and R. Eisenberg, J. Am. Chem. Soc., 2010, 132, 15480–15483 CrossRef CAS PubMed;
(c) P. Zhang, M. Wang, J. Dong, X. Li, F. Wang, L. Wu and L. Sun, J. Phys. Chem. C, 2010, 114, 15868–15874 CrossRef CAS;
(d) J. Dong, M. Wang, P. Zhang, S. Yang, J. Liu, X. Li and L. Sun, J. Phys. Chem. C, 2011, 115, 15089–15096 CrossRef CAS;
(e) L. Gong, J. Wang, H. Li, L. Wang, J. Zhao and Z. Zhu, Catal. Commun., 2011, 12, 1099–1103 CrossRef CAS PubMed;
(f) T. M. McCormick, Z. Han, D. J. Weinberg, W. W. Brennessel, P. L. Holland and R. Eisenberg, Inorg. Chem., 2011, 50, 10660–10666 CrossRef CAS PubMed.
-
(a) P. Zhang, M. Wang, C. Li, X. Li, J. Dong and L. Sun, Chem. Commun., 2010, 46, 8806–8808 RSC;
(b) K. Peuntinger, T. Lazarides, D. Dafnomili, G. Charalambidis, G. Landrou, A. Kahnt, R. P. Sabatini, D. W. McCamant, D. T. Gryko, A. G. Coutsolelos and D. M. Guldi, J. Phys. Chem. C, 2012, 117, 1647–1655 CrossRef;
(c) M. Natali, R. Argazzi, C. Chiorboli, E. Iengo and F. Scandola, Chem.–Eur. J., 2013, 19, 9261–9271 CrossRef CAS PubMed.
- F. Wen and C. Li, Acc. Chem. Res., 2013, 46, 2355–2364 CrossRef CAS PubMed.
-
(a) F. Wen, J. Yang, X. Zong, B. Ma, D. Wang and C. Li, J. Catal., 2011, 281, 318–324 CrossRef CAS PubMed;
(b) J. Dong, M. Wang, X. Li, L. Chen, Y. He and L. Sun, ChemSusChem, 2012, 5, 2133–2138 CrossRef CAS PubMed.
-
(a) F. Lakadamyali and E. Reisner, Chem. Commun., 2011, 47, 1695–1697 RSC;
(b) F. Lakadamyali, M. Kato, N. M. Muresan and E. Reisner, Angew. Chem., Int. Ed., 2012, 51, 9381–9384 CrossRef CAS PubMed;
(c) F. Lakadamyali, A. Reynal, M. Kato, J. R. Durrant and E. Reisner, Chem.–Eur. J., 2012, 18, 15464–15475 CrossRef CAS PubMed.
- J. Huang, K. L. Mulfort, P. Du and L. X. Chen, J. Am. Chem. Soc., 2012, 134, 16472–16475 CrossRef CAS PubMed.
- S. W. Cao, X. F. Liu, Y. P. Yuan, Z. Y. Zhang, J. Fang, S. C. J. Loo, J. Barber, T. C. Sum and C. Xue, Phys. Chem. Chem. Phys., 2013, 15, 18363–18366 RSC.
- X. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domen and M. Antonietti, Nat. Mater., 2009, 8, 76–80 CrossRef CAS PubMed.
-
(a) J. Zhang, X. Chen, K. Takanabe, K. Maeda, K. Domen, J. D. Epping, X. Fu, M. Antonietti and X. Wang, Angew. Chem., Int. Ed., 2010, 49, 441–444 CrossRef CAS PubMed;
(b) X. H. Li, X. Wang and M. Antonietti, Chem. Sci., 2012, 3, 2170–2174 RSC;
(c) Y. Hou, A. B. Laursen, J. Zhang, G. Zhang, Y. Zhu, X. Wang, S. Dahl and I. Chorkendorff, Angew. Chem., Int. Ed., 2013, 52, 3621–3625 CrossRef CAS PubMed.
-
(a) S. Min and G. Lu, J. Phys. Chem. C, 2012, 116, 19644–19652 CrossRef CAS;
(b) Y. Zhang, J. Liu, G. Wu and W. Chen, Nanoscale, 2012, 4, 5300–5303 RSC.
-
(a) M. Sykora, M. A. Petruska, J. Alstrum-Acevedo, I. Bezel, T. J. Meyer and V. I. Klimov, J. Am. Chem. Soc., 2006, 128, 9984–9985 CrossRef CAS PubMed;
(b) J. Huang, D. Stockwell, Z. Huang, D. L. Mohler and T. Lian, J. Am. Chem. Soc., 2008, 130, 5632–5633 CrossRef CAS PubMed;
(c) A. J. Morris-Cohen, M. T. Frederick, L. C. Cass and E. A. Weiss, J. Am. Chem. Soc., 2011, 133, 10146–10154 CrossRef CAS PubMed.
-
(a) K. Takanabe, K. Kamata, X. Wang, M. Antonietti, J. Kubota and K. Domen, Phys. Chem. Chem. Phys., 2010, 12, 13020–13025 RSC;
(b) Y. Wang, Z. Wang, S. Muhammad and J. He, CrystEngComm, 2012, 14, 5065–5070 RSC.
- S. Geremia, R. Dreos, L. Randaccio, G. Tauzher and L. Antolini, Inorg. Chim. Acta, 1994, 216, 125–129 CrossRef CAS.
- V. Georgakilas, K. Kordatos, M. Prato, D. M. Guldi, M. Holzinger and A. Hirsch, J. Am. Chem. Soc., 2002, 124, 760–761 CrossRef CAS PubMed.
-
(a) X. Wang, K. Maeda, X. Chen, K. Takanabe, K. Domen, Y. Hou, X. Fu and M. Antonietti, J. Am. Chem. Soc., 2009, 131, 1680–1681 CrossRef CAS PubMed;
(b) Y. Zheng, J. Liu, J. Liang, M. Jaroniec and S. Z. Qiao, Energy Environ. Sci., 2012, 5, 6717–6731 RSC.
- M. Razavet, V. Artero and M. Fontecave, Inorg. Chem., 2005, 44, 4786–4795 CrossRef CAS PubMed.
- G. A. N. Felton, C. A. Mebi, B. J. Petro, A. K. Vannucci, D. H. Evans, R. S. Glass and D. L. Lichtenberger, J. Organomet. Chem., 2009, 694, 2681–2699 CrossRef CAS PubMed.
- H. Lee, E. Luna, M. Hinz, J. J. Stezowski, A. S. Kiselyo and R. G. Harvey, J. Org. Chem., 1995, 60, 5604–5613 CrossRef CAS.
- W. C. Trogler, R. C. Stewart, L. A. Epps and L. G. Marzilli, Inorg. Chem., 1974, 13, 1564–1570 CrossRef CAS.
Footnote |
† Electronic supplementary information (ESI) available: Fig. S1–S14. CCDC 955344 and 982938. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c4ra01413h |
|
This journal is © The Royal Society of Chemistry 2014 |
Click here to see how this site uses Cookies. View our privacy policy here.