I. Tamiolakis,
I. T. Papadas,
K. C. Spyridopoulos and
G. S. Armatas*
Department of Materials Science and Technology, University of Crete, 71003 Heraklion, Crete, Greece. E-mail: garmatas@materials.uoc.gr
First published on 31st May 2016
Photocatalytic water splitting to produce hydrogen using solar energy is a particularly attractive solution to increasing energy demands. However, to be of practical use, semiconductor electrodes need to be made of inexpensive, abundant elements and have a high, yet stable, photocatalytic H2-production activity. Here we report the first demonstration of 3D mesoporous networks of Cu2O and TiO2 nanoparticles as highly efficient photocatalysts for hydrogen generation from water. These assembled structures feature a highly accessible pore surface that exposes a large fraction of anatase TiO2 and Cu2O nanoparticles to electrolytes, and has a small grain size of the constituent nanocrystals, which lead to excellent activity for H2 evolution via a UV-visible light-driven reduction of protons. Catalytic results associated with optical UV-vis/NIR absorption and photoluminescence (PL) data indicated that the large separation of photogenerated electrons and holes at the Cu2O–TiO2 p–n junctions was the main reason attributed to the improved photochemical performance. Consequently, the mesoporous Cu2O/TiO2 catalyst containing ∼1.5 wt% Cu reaches an average H2 evolution rate of ∼542 μmol h−1 (or ∼36133 μmol h−1 g−1) with an apparent quantum efficiency (QE) of 13.5% at 365 nm and an incident photon conversion efficiency of ∼4.1% under UV-visible light illumination (360–780 nm), which is one of the best HER activities among TiO2-based semiconductor systems reported to date.
Nowadays, TiO2 is the most widely study material for energy conversion and environmental applications, mainly because of its appropriate energy edge potentials, chemical stability, non-toxicity and low cost.10 Nevertheless, TiO2 photocatalyst is poorly active for producing H2 from water, as a results of the low solar light-capturing capability (it absorbs only in UV region) and low charge carrier separation yield.11 To this end, various TiO2-based heterojunction materials have been studied in an effort to sustain high photon collection and chemical STH efficiency. Metal doping with Pt, Pd, Rh and Au co-catalysts has been a prominent quest to improve hydrogen production performance.12,13 It is well known that these metal NPs retard the electron–hole recombination by acting as electron sinks, and facilitating the interparticle metal/TiO2 electron-transfer process.14 However, although noble metal NPs, like Pt, allow considerable reduction in energy cost for H2 generation, they are too scare and costly for large-scale deployment.
An alternative and cost-efficient TiO2-based photocatalyst system is based on loading 3d-transition metal oxides, such as Fe2O3,15 Cu2O,16,17 NiO18 and Co3O4,18 to promote HER catalysis. Among them, Cu2O/TiO2 p–n heterojunction composites have attracted particular attention for their potential of achieving high H2 evolution performance from water splitting.19–23 In fact, owing to the intrinsic electric field across the Cu2O–TiO2 junctions, the photogenerated carriers can be spatially separated into the composite structure, thus suppressing the recombination of electron–hole pairs.24 Thanks to the proper matching of band structure between Cu2O and TiO2, the photogenerated electrons can transfer from the conduction band (CB) of Cu2O to TiO2, while the holes at the valence band (VB) of TiO2 can move to the VB of Cu2O; both conduction and valence band levels of Cu2O (−1.4 V and 0.8 V vs. NHE, pH = 7) lie above those of TiO2 (−0.6 V and 2.6 V vs. NHE, pH = 7). Recent studies have shown that the dimensional attributes of copper and titanium oxide components in a Cu-loaded TiO2 photocatalyst are strongly related with the photocatalytic activity. For example, one-dimensional (1D) CuO/TiO2 nanotubes showed an efficient H2-production activity as a result of the movement of the electrons and holes along the tube length.25,26 Therefore, as a consequence of the different chemical reactivity of the oxide nanostructures, a beneficial improvement in charge carriers' separation and hence better photocatalytic activity can be expected.
Recently, we have reported a facile synthesis of well-controlled NP-based mesoporous titania (MTA), using a polymer-templated aggregating self-assembly of TiO2 NPs.27 Mesoporous networks of TiO2 NPs were obtained via a sol–gel polymerization of TiCl4 and titanium(IV)propoxide (TPP) compounds at the surface of polyoxoethylene cetyl ether (Brij-58) block copolymer micelles, followed by calcination in air (350 °C). Because of the large expose surface area and small grain size of anatase particles, MTA architecture offers a unique platform to develop new composite materials and investigate their electrochemical properties. One such interesting feature of small-sized TiO2 nanocrystals is the efficient electron-transport to the solid/electrolyte interface.28,29 In this work, we use a simple chemical reduction deposition method to grow Cu2O nanocrystals on the surface of mesoporous titania assemblies. The resulting materials possess a three-dimensional (3D) network consisting of connected Cu2O and TiO2 NPs and exhibit high internal surface area and narrow pore size distribution. To our knowledge, this is the first report on the use of a highly porous network of Cu2O-conjugated TiO2 NPs for photocatalytic hydrogen production. Previous reports have demonstrated the use of Cu-loaded TiO2 materials in nanotube,26 isolated nanoparticle, nanoplate18 or thin film morphologies20 as photocatalysts for photoelectrochemical water splitting. Here we show that the assembled structure of these composites is highly effective and well tolerated in producing H2 from water under UV-visible irradiation. We found that mesoporous Cu2O/TiO2 composite at 1.5 wt% Cu loading leads to a considerable H2-production enhancement, which corresponds to a ∼135-fold increase of HER activity compared to that of pristine mesoporous TiO2 (MTA). The associated mechanism of the photocatalytic H2 production under UV-visible light is also discussed.
Sample | Atomic ratioa (Ti![]() ![]() |
Cu loading (wt%) | Surface area (m2 g−1) | Pore volume (cm3 g−1) | Pore size (nm) |
---|---|---|---|---|---|
a Based on the EDS analysis. | |||||
CuMTA-1 | 99.38![]() ![]() |
0.49% | 141 | 0.27 | 7.1 |
CuMTA-2 | 98.80![]() ![]() |
0.96% | 130 | 0.24 | 7.0 |
CuMTA-3 | 98.17![]() ![]() |
1.47% | 129 | 0.23 | 7.0 |
CuMTA-4 | 97.59![]() ![]() |
1.93% | 126 | 0.24 | 6.8 |
CuMTA-3P | 98.19![]() ![]() |
1.45% | 121 | 0.23 | 6.7 |
The mesostructure and crystallinity of the obtained porous materials were carefully investigated by transmission electron microscopy (TEM), selected-area electron diffraction (SAED), X-ray diffraction (XRD), and nitrogen physisorption analysis. A typical TEM image of the CuMTA-3, which is the most active catalyst studied in this work, is shown in Fig. 1a. The image shows that this material consists of a porous network of connected TiO2 NPs with uniform size (ca. 7.3 nm). Also, it is clearly observed that some discrete Cu2O nanocrystals (appeared as dark areas) with an average diameter of 5.7 nm, as inferred by TEM analysis (Fig. 1b), are well dispersed over the TiO2 framework (bright region). The high resolution TEM (HRTEM) image, in Fig. 1c, shows the lattice fringes of both TiO2 and Cu2O components, suggesting the well-defined crystal structures. The lattice fringes with d-spacing of 3.5 Å are assigned to the (101) planes of anatase TiO2, while those with 2.4 Å lattice spacing correspond to the (111) faces of the cubic cuprite structure (Pnm, a = 4.267 Å). It is worth noting that Cu2O NPs appear to be in intimate contact with TiO2 particles. Such a close interconnection between Cu2O and TiO2 particles is particularly important for photocatalytic applications where the interfacial electron transfer from Cu2O to TiO2 is believed to enhance the charge carrier separation and, thus, the photocatalytic efficiency. In agreement with TEM observations, scanning electron microscopy (SEM) images of CuMTA-3 show that very small particles of size less than 10 nm are agglomerated into large structures, forming a 3D network (Fig. S1†). Further structural information was obtained from selected area electron diffraction analysis. The SAED pattern of CuMTA-3 sample depicts several Debye–Scherrer diffraction rings that can be indexed to the crystal planes of the TiO2 anatase structure (marker with red lines) and the cubic lattice of Cu2O (marker with yellow lines), see Fig. 1d.
The powder XRD patterns of the mesoporous CuMTA materials clearly demonstrate the well-developed nanocrystalline structure of TiO2, showing several broad reflections in the 2θ = 20–80° range (Fig. S2†). All reflections peaks match well with the Bragg diffractions of the standard anatase structure of TiO2 (JCPDS card no. 21-1272) with an average grain size, as estimated from the broadening of the (101) peak and the Scherrer equation,30 of about 9 nm, in good consistency with TEM results. However, Cu2O phase is not detected by XRD analysis in CuMTA composites owing to the low amounts present and high dispersion of Cu2O particles on the surface of TiO2.
The mesoporosity of the title materials was determined by nitrogen physisorption at 77 K. The N2 adsorption–desorption isotherms of mesoporous CuMTA are shown in Fig. 2a. All isotherms exhibit typical type-IV curves, according to the IUPAC classification, with a sharp capillary condensation step at a relative pressure (P/P0) of 0.75, indicating mesoporous structure with narrow-sized pores.31 Of note, the H3-type hysteresis loop between the adsorption and desorption branch of isotherms suggests the existence of an interconnected pore system between the nanoparticles. The CuMTA composites exhibited Brunauer–Emmett–Teller (BET) surface areas in the range of 126–141 m2 g−1 and total pore volumes in the range of 0.23–0.27 cm3 g−1, which are slightly lower than that of the pristine MTA sample (153 m2 g−1, 0.29 cm3 g−1, see ESI Fig. S3†), plausibly due to the deposition of Cu2O NPs on the titania surface. The pore width in CuMTA materials was determined by using the non-local density functional theory (NLDFT) model on the adsorption data, and was found to be in the range 6.8–7.1 nm with narrow size distribution (Fig. 2b). In comparison with the pore size of the pristine MTA (ca. 7.2 nm), the pore diameter in CuMTA is slightly lower and progressively decreased with increasing Cu content. This suggests that Cu2O particles are embedded in the porous structure of MTA. All the textural properties of the mesoporous CuMTA materials are listed in Table 1.
The presence of copper oxide NPs in CuMTA mesoporous was also verified by ultraviolet-visible/near-IR (UV-vis/NIR) light diffuse reflectance spectroscopy. The sharp optical absorption at wavelengths shorter than 390 nm is assigned to the intrinsic interband (O2p → Ti3d) electron transitions in TiO2 (Fig. 3). The band gap of the CuMTA semiconductors can be estimated from the Tauc plot [(αhν)2 versus hν, in which α, h and ν are the absorption coefficient, Planck's constant and light frequency, respectively], and thus was found to be 3.2–3.3 eV (Fig. 3, inset). Furthermore, an apparent tail in the visible light region (at wavelength over 400 nm) of the UV-vis/NIR spectra can be clearly observed with the increase of Cu2O loading. Given that the absorption feature at ∼460 nm is very close to the interfacial transition energy from the VB of TiO2 (2.6 V vs. NHE, pH = 7) to Cu2O/Cu (0.05 V vs. NHE, pH = 7),32 it is reasonable to suggest that this absorption may be due to the interface charge transfer (IFCT) from the VB of TiO2 to Cu2O.33 Such photoinduced electron transfer can cause a partial reduction of Cu2O to Cu0 in CuMTA composites, the presence of which has been identified by HRTEM, as we shall discuss below. The absorption at wavelengths longer than 500 nm is attributed to the band edge transition in Cu2O.34 Notably, we did not observe any appreciable red shift in the energy gap of TiO2 with increasing the Cu loading. This suggests that Cu elements do not incorporated into the titania lattice but rather form Cu2O deposits on the surface.
Catalyst | Evolved H2, 3 h (μmol) | Rate of H2 evolutiona (μmol h−1) |
---|---|---|
a Average H2 evolution rate over 3 h irradiation period.b Total amount and average H2 evolution rate under visible light irradiation (λ > 420 nm). | ||
MTA | 15 | ∼4 |
CuMTA-1 | 241 | 82 |
CuMTA-2 | 988 | 341 |
CuMTA-3 | 1630 (∼11)b | 542 (∼3.6)b |
CuMTA-4 | 1224 | 409 |
CuMTA-3P | 374 | 120 |
Note here that the HER activity of CuMTA-3 is higher to that obtained for Cu2O/rGO (rGO, reduced graphene oxide) (∼264 μmol g−1 h−1)36 and Cu2O–TiO2/rGO (110968 μmol g−1 h−1)37 composites, Cu2O/TiO2 nanoplates (37.2 μmol h−1, 372 μmol g−1 h−1),18 CuO/TiO2 nanotubes (499 μmol h−1, 99
823 μmol g−1 h−1),25 and Cu2O/TiO2 (P25) (∼134 μmol h−1, 668 μmol g−1 h−1),19 Cu–Cu2O/TiO2 (P25) (511 μmol h−1, 12
779 μmol g−1 h−1),38 Cu2O/TiO2 (P25) (290 μmol h−1, 2900 μmol g−1 h−1)39 and Au/Cu2O–TiO2 (1140 μmol h−1, 11
400 μmol g−1 h−1, QE ∼ 7–8%)40 NPs, and it is comparable to the activity of mesoporous Cu–CuO/TiO2 ternary structures (7150 μmol h−1, 286
000 μmol g−1 h−1, QE ∼ 10.5%).41 To our knowledge, the prepared CuMTA composite system is one of the most efficient photocatalysts reported until now for TiO2-based photocatalytic hydrogen generation.
The CuMTA-3 catalyst also demonstrated very good stability under the examined conditions. The reusability study was carried out by purging the reaction cell with argon at certain intervals (5 h), until no H2 and O2 were detected (by GC analysis). As shown in Fig. 5a, the CuMTA-3 sustains its activity after five catalytic cycles. After 25 hours of illumination, a total amount of H2 of ∼11978 μmol (∼268 mL) was obtained, which corresponds to an average HER rate of ∼479 μmol h−1. Assuming that this H2 evolution rate corresponds to a 114 J energy generated by the reduction of water and the catalyst particles absorb all incident UV and visible light (0.51 W cm−2), the photon-to-hydrogen energy conversion efficiency was determined to be ∼2% (in the 360–780 nm interval), which is among the highest values reported in the literature.42,43 Moreover, elemental X-ray microanalysis and N2 physisorption studies showed no detachment of Cu2O NPs from the TiO2 surface and dissolution of the TiO2 assembled structure after 25 h of reaction, confirming the high durability of the CuMTA-3 composite under the examined conditions. These results indicated that the regenerated sample have a Cu loading of ∼1.42 wt%, a surface area of 121 m2 g−1 and a total pore volume of 0.22 cm3 g−1 (Fig. S5 and S6†). Furthermore, the CuMTA-3 catalyst retains, to a great extent, the high distribution of Cu2O nanocrystals on the titania matrix after catalysis, as derived from the contrast different and the phase assignment from the TEM images in Fig. 5b and c. The HRTEM also revealed that some Cu nanocrystals with a diameter of about 4–5 nm were formed between the Cu2O and TiO2 particles; the image in Fig. 5c shows the lattice fringes with d-spacings of 1.8 Å and 2.1 Å corresponding to the (200) and (111) facets of Cu, along with the lattice fringes of the Cu2O (111) and anatase TiO2 (101) structures. The color change of the obtained catalyst to darker gray is also a reasonable indication of the presence of metallic Cu NPs. However, the regenerated catalyst fully regained its original color upon exposure to air for a few hours. It is noted that, for TEM analysis, significant effort was taken to minimize the air exposure time of the reused sample (it was handled under N2 atmosphere) in order to avoid spontaneous oxidation of Cu metals to Cu2O.44 In agreement with TEM results, the optical absorption spectrum of the regenerated catalyst suggested that metallic Cu is also present in the composite structure, displaying a broad feature at ∼600 nm due to the localized surface plasmon resonance (LSPR) of Cu NPs (Fig. S7†).45,46 Taken together, it is obvious that a partial transformation of Cu2O into metallic Cu occurs during the photocatalytic process. A similar structural transformation of Cu2O into Cu0 upon UV irradiation of Cu2O/TiO2 composites have been observed by other researchers.19,38,40
To better understand the effect of Cu2O NPs on H2 production activity of CuMTA, we ran visible light (λ > 420 nm) irradiated HER using the CuMTA-3 as catalyst. In this experiment, CuMTA-3 exhibited little H2 evolution activity (∼3.6 μmol h−1), which is more than an order of magnitude lower than that under UV-light illumination (Table 2). The results from this study clearly indicate that even through incorporation of Cu2O NPs on the titania surface can induces visible light absorption (Cu2O has an energy gap of ∼2.1–2.2 eV),32 as inferred by UV-vis spectroscopy, it does not seem to be a reliable explanation for the increased photoactivity. Therefore, we can infer that the greatly improved HER performance of CuMTA mesoporous mainly arising from the efficient charge carrier dissociation at the Cu2O/TiO2 interface that retards the recombination of photogenerated electrons and holes. To investigate whether electron–hole pair separation occurs in the CuMTA heterostructures, we performed photoluminescence (PL) measurements. As seen in ESI Fig. S8,† the mesoporous TiO2 (MTA) exhibits two intense light emissions at around 414 and 470 nm due to the radiative recombination of self-trapped excitons (between the sub-bandgap states of TiO2).47 The weak signal at 500–550 nm range could be related to surface oxygen vacancies and defects in TiO2 lattice.44 Comparatively, the PL emission of CuMTA-3 heterostructure is almost vanished. This suggests a limited interband recombination of photoexcited electrons and holes in CuMTA-3 due to the resonance charge transport within the Cu2O/TiO2 heterojunctions.
It should be stressed that the chemical nature and phase morphology of Cu species attached on the titania surface are important for promoting photocatalytic H2 production. As such, we also prepared a Cu-loaded TiO2 catalyst by photo-deposition of ∼1.5 wt% Cu on the surface of MTA (CuMTA-3P) and we compared its HER activity with that of CuMTA-3. These two materials exhibit very similar textural properties and chemical composition (see Table 1 and Fig. S9†), but feature copper species with different valence state on the titania surface. The HRTEM images in ESI Fig. S10† show that the surface of CuMTA-3P is dominated by metallic Cu deposits with an average size of about 4–5 nm; these NPs display lattice fringe spacings of 2.1 and 1.8 Å nm, which are characteristic of the (111) and (200) planes of cubic Cu. Consistent with TEM observation, UV-vis/NIR absorption spectrum of CuMTA-3P showed a broad feature in the 700–800 nm range due to the plasmonic excitation of metallic Cu (Fig. S11†),46 further confirming the Cu chemical nature. Catalytic results indicated that CuMTA-3P affords a ∼120 μmol h−1 H2 evolution rate under identical illumination conditions, which is remarkably low when compared to the activity of mesoporous CuMTA-3. This observation is consistent with previous studies,48 and suggests that Cu/TiO2 composite is not a suitable photocatalyst for H2 evolution.
On the basis of the above results, a schematic representation of photocatalytic H2 production from the mesoporous CuMTA is shown in Scheme 1. In short, upon UV-visible light illumination, both Cu2O and TiO2 semiconductor components are photoexcited to produce electrons (e−) and holes (h+) in the conduction and valence band, respectively. Next, because of the potential gradient at the Cu2O/TiO2 interface, the photogenerated electrons from Cu2O can transfer to the CB of TiO2 and promote the hydrogen production. In addition, during the course of irradiation a fraction of photogenerated electrons in CB of Cu2O may be also consumed by the reduction of Cu+ ions to Cu NPs. This could be due to the lower reduction potential of Cu2O to Cu0 (E0 = 0.05 V vs. NHE) than the CB edges of Cu2O and TiO2. Also, the direct IFCT from the VB of TiO2 to Cu2O, which results in the partial reduction of Cu2O to Cu, is a reasonable option. The presence of Cu metal clusters between the TiO2 and Cu2O particles in the CuMTA-3 was evidenced by HRTEM analysis and UV-vis/NIR spectroscopy. It was recognized that such Cu/Cu2O/TiO2 ternary heterojunctions can enhance tunneling of electrons across the junction, dissociating more efficiently the photoexcited electrons and holes that are available for reactions.37,49 Meanwhile, the photogenerated holes in the VB of TiO2 can be transferred to the Cu2O VB (or to the metallic Cu NPs oxidizing Cu0 to Cu+) and consumed by methanol. As noted above, the Cu2O–TiO2 junctions can facilitate the spatial separation of photoexcited excitons (electron–hole pairs) and promote HER activity. The improved separation of photogenerated electrons and holes within the structure of CuMTA was verified by PL spectroscopy.
![]() | ||
Scheme 1 Photocatalytic H2 production mechanism on the Cu/Cu2O/TiO2 interface under UV-visible light irradiation (VB: valence band, CB: conduction band, OPs: oxidation products). |
The apparent quantum efficiency (QE = 2NH2/Nip, where NH2 and Nip are the number of evolved hydrogen molecules and number of incident photons) and the UV-visible photon conversion efficiency of the CuMTA-3 sample were determined by using a StarLite power meter equipped with a FL400A-BB-50 fan cooled thermal sensor (Ophir Optronics Ltd). The average intensity of incident light was measured to be 40 mW cm−2 using a band-pass cutoff filter of λ = 365 ± 10 nm (Asahi Spectra, Japan) and the total incident power of irradiation was 0.51 W cm−2 in the 360–780 nm interval.
Footnote |
† Electronic supplementary information (ESI) available: SEM images, XRD patterns and catalytic data of CuMTA samples, N2 physisorption isotherms of MTA, EDS, N2 physisorption and UV-vis/NIR data of the regenerated CuMTA-3 catalyst, PL spectra of MTA and CuMTA-3, N2 physisorption isotherms, TEM images and UV-vis/NIR spectrum of CuMTA-3P. See DOI: 10.1039/c6ra08546f |
This journal is © The Royal Society of Chemistry 2016 |