Songmei
Sun
a,
Qi
An
b,
Wenzhong
Wang
*a,
Ling
Zhang
a,
Jianjun
Liu
*a and
William A.
Goddard III
b
aState Key Laboratory of High Performance Ceramics and Superfine Microstructure, Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 200050, P. R. China. E-mail: wzwang@mail.sic.ac.cn; jliu@mail.sic.ac.cn
bDivision of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, California 91125, USA
First published on 15th November 2016
N2 reduction to ammonia by solar light represents a green and sustainable ammonia synthesis approach which helps to suppress the global warming and energy crisis. However, conventional semiconductors usually suffer from low activity or poor stability, largely suppressing the application of this technology. Here, we report that bismuth monoxide (BiO) quantum dots with an average size of 2–5 nm exhibited efficient photocatalytic activity for ammonia synthesis under simulated solar light. A highly efficient ammonia synthesis rate of 1226 μmol g−1 h−1 is achieved without the assistance of any sacrificial agent or co-catalyst, which is about 1000 times higher than that using the traditional Fe-TiO2 photocatalyst. Kinetic analysis reveals that the synergy of three low valence surface Bi(II) species markedly enhances N2 activation by electron donation, which finally resulted in the highly efficient N2 photoreduction performance. This work will shed light on designing efficient and robust N2 reduction photocatalysts.
Generally, the rate-determining step for N2 reduction is cleavage of the NN bond.5,6 N2 activation by electron donation from some reducing active sites is indispensable for the promotion of N
N bond cleavage,25 but such reducing active sites rarely exist on most of the semiconductor photocatalysts. Until now, the majority of the reported N2 photoreduction catalysts have only defect type activation centers,17–24 which result in unsatisfactory photocatalytic performance because of their limited defect concentrations. Besides, defect states often suffer from thermal instability and increased charge carrier recombination, which further prevent their practical application for artificial ammonia synthesis. Consequently, developing an ideal photocatalytic material which has a high concentration of non-defect type activation centers with robust stability holds the key to achieving a breakthrough in N2 photoreduction technology.
Using low valence metal species in low-valent metal oxide semiconductors for N2 activation may be a viable strategy to achieve this important goal. Low-valent metals in semiconductor solids have fewer coordination atoms than those in their high-valent state and are hence chemically unsaturated to facilitate chemisorption of molecules.26 Owing to the strong electron donating power, low-valent metal species such as Fe, Mo, Zr, and Ti have been widely studied for N2 activation in organometallic complexes,9,16,27–30 although a stoichiometric excess of strong reducing agents (such as Na, K) and extra proton sources are generally required to further generate NH3 in these reaction systems. Different from organometallic complexes, low-valent metal species in semiconductor materials may act as both N2 activation and hydrogenation centers without extra reducing agents, which allows quick reaction kinetics for N2 reduction, because the conduction band of a metal oxide semiconductor is usually hybridized by the metal p, d orbitals,31 and photogenerated electrons are mainly located on the metal sites. Moreover, different from defect-type active sites, lattice low-valent metal species can refrain from becoming an electron–hole recombination center and all of the surface low-valent metal sites could contribute to the catalysis reactions theoretically. In this regard, low-valent metal oxide semiconductors are expected to exhibit high photocatalytic activity for N2 photoreduction. To confirm this opinion, herein low valent bismuth monoxide (BiO) was selected as an ideal model material to study the N2 photoreduction performance. We focus on low valent bismuth(II) because it has high electron donating power and empty 6d orbitals for N2 adsorption and activation. More importantly, bismuth based materials have been widely studied with various catalytic applications.32–34
Ep = E0 + (RT/anF)ln(RTk0/anF) + (RT/anF)ln![]() |
To study the photocatalytic N2 reduction property of the BiO catalyst, 0.05 g of the BiO particles was dispersed in 200 mL of deionized water and then subsequently irradiated under simulated solar light irradiation. Control experiments showed that NH4+ cannot be generated in the absence of the BiO catalyst or solar light irradiation. In contrast, simulated solar light irradiation resulted in continuous ammonia generation by the BiO sample (Fig. 3a). After 24 h of irradiation, the BiO catalyst generated 7.4 mg L−1 of NH4+ from pure water in air. In an acidic reaction solution (pH 3.8) which could provide excess protons to decrease the kinetic barrier for N2 reduction, the generated ammonia concentration by the BiO catalyst was increased to 20.6 mg L−1 within 24 h, with a much improved ammonia synthesis rate of about 1226 μmol g−1 (catalyst) h−1, 1000 times higher than that of the traditional Fe-TiO2 photocatalyst (Fig. 3a) under the same conditions. When the reaction was performed in 15% aqueous methanol, a known sacrificial electron donor, the ammonia synthesis rate was further significantly improved, producing a total NH4+ amount of 0.7 mmol after 24 h (Fig. 3b), equivalent to a turnover number of 3.18. This suggests that ammonia evolution under these conditions is catalytic. The increase in the ammonia synthesis rate in aqueous methanol indicates that the BiO quantum dots are able to photooxidize methanol. Under the same conditions, no ammonia generation was detected with the BiO catalyst both in pure water and aqueous methanol in argon saturated solution, indicating that the ammonia in the reaction system originated from N2 reduction. The stability of the BiO catalyst under acidic conditions was also investigated. After 5 consecutive runs (120 h) for N2 reduction, the photocatalytic performance of the BiO particles was well-maintained (Fig. 3c). The TEM image of the BiO catalyst after the cycle of reaction also indicates the stability of the BiO catalyst (Fig. S3†).
To further confirm that the detected NH4+ in our experiment indeed originated from N2 reduction, we performed an isotopic labelling study using 50 vol% 15N2 as the purge gas. Infrared spectroscopy was used to characterize the NH4+ product. The IR spectra (Fig. 3d) indicate that both 15NH4+ and 14NH4+ exist in the reaction solution when using 50% 15N2 labelled N2, while only 14NH4+ is detected when using pure 14N2. The infrared absorption peak of 14NH4+ was located around 1405 cm−1, while the peak position for 15NH4+ is at 1358 cm−1 which is in good agreement with the value estimated according to the isotope effect (that is, 1405 cm−1 × (14/15)1/2 = 1357 cm−1).25 The isotope labelling study further proved that the NH4+ originates from the photocatalytic N2 reduction process. 1H NMR spectrum was also used for detecting the 15NH4+ species in the obtained ammonia product. As shown in Fig. S4,† both 15NH4+ and 14NH4+ were observed in the NMR spectrum. The ratio of the 15NH4+ content to 14NH4+ was about 2.7:
3, which is close to the initial 15N2 ratio.
It is of significant importance to understand the underlying mechanism of the high photocatalytic activity of BiO in reducing N2 in order to develop other efficient photocatalysts. Quantum chemical calculations based on density functional theory (DFT) were conducted to investigate the possible N2 activation and hydrogenation process on the BiO catalyst. According to the experimental observation from the HRTEM image, a representative BiO {010} surface was selected to investigate the N2 activation and hydrogenation performance. First, the partial charge density of the nearby Fermi level was calculated to analyze the possible interaction positions of BiO surfaces with N2. As shown in Fig. 4a and b, the valence bands of BiO near Fermi levels are mainly generated by p-electrons of Bi atoms. The charge distribution on BiO {010} surface is significantly localized and accumulated around Bi atoms, which indicates that N2 activation by electron donation may take place upon the Bi activation centers. The extent of N2 activation and hydrogenation on the BiO surface is reflected by the change of the N–N bond length. Fig. 4c presents N–N bond distance evolutions with addition of hydrogen atoms on the BiO {010} surfaces. For comparison, the corresponding evolutions without the catalyst are also plotted. As shown in Fig. 4c, the BiO {010} surface displays high catalytic activity on activation of the NN bond in N2Hn (n = 0–3). The N–N bond length could be elongated from 1.09 Å to 1.12 Å when it is near the BiO {010} surface. The N–N bond length could be further increased along with hydrogenation until three H atoms were hydrogenated, indicating that the N
N bond may be broken after three H atoms were added by the BiO catalyst. Considering that the N2 photoreduction experiment was performed in water in this study, the influence of water molecules on the N2 activation and hydrogenation was also investigated. The calculated result indicates that the catalytic activity of BiO was not changed with addition of H2O on the surface. In addition, H2O molecules covering the surface can quickly dissociate into H+ and OH−, providing sufficient hydrogen for ammonia generation. Besides, the binding energy calculations reveal a weaker interaction between N2Hn and the water-covered BiO surface than that on a dry surface (Fig. 4d). This indicates that water on the BiO {010} surface is advantageous to improve N2 photoreduction performance since hydrogenated N2Hn can easily be desorbed from the catalyst. Temperature programmed desorption (TPD) investigation was also conducted to visualize the adsorption and activation of N2 on the surface of the BiO catalyst. As shown in Fig. S5,† two peaks at 295 °C and 330 °C were clearly observed for N2 desorption, which indicated the excellent N2 chemisorption on the BiO catalyst.
Photoelectrochemical measurements further revealed the possible reasons for the high performance of the BiO catalyst. As shown in Fig. 5a, two well-defined cathodic peaks around −0.7 V and −1.2 V appear in the cyclic voltammetry experiments in N2 saturated 0.5 M Na2SO4 solution for the BiO electrode in an acidic solution (pH = 3.8), the pH value of which was adjusted using dilute H2SO4. The lower reduction peak around −0.7 V arises only after the electro-reduction process proceeds for a certain length of time (Fig. 5a and S6†), indicating that this reduction peak may have originated from further reduction of intermediates during the photoelectrochemical process. Therefore the N2 reduction peak around −1.2 V (Ep) was adopted for the kinetic study under acidic conditions. As shown in Fig. 5a, the peak current densities increased with increase in the scan rates (ν). The electron transfer number (n) involved in the N2 reduction can be calculated from the slope of Ep − lnν. The value of n was calculated to be 2.93 at the potential of 20–100 mV, indicating that a possible one-step three electron transfer N2 reduction process may occur on the BiO surface. As is well known, nitrogen fixation is a proton-coupled reaction.36 Although the BiO surface could provide sufficient low valence Bi(II) activation centers, excess H+ is also important to realize a proton coupled multi-electron reduction process. In a neutral reaction solution, the electron transfer number was decreased to 1.34 based on the analysis of the reduction peak around −0.9 V (Fig. 5b), indicating that N2 reduction on the BiO catalyst in neutral solution may be realized through six consecutive one-electron redox steps, resulting in a much lower ammonia synthesis rate. This is in line with the observed lower photocatalytic N2 reduction activity at pH = 7. Besides, a decreased cathodic current density (Fig. 5c) in a N2-saturated reaction system was observed when compared with that in air saturated solution, which proved that a larger number of electrons were transferred from the BiO catalyst to the N2 molecule in the N2-saturated reaction system under the same conditions. Fig. 5d shows the Mott–Schottky spectrum of the BiO sample, which is usually used for the analysis of the flat band potential (Efb) of semiconductor electrodes.37 The positive slope of the plot indicates that the BiO catalyst is an n-type semiconductor with electrons as the majority charge carriers. The Efb value which was calculated from the intercept of the axis with potential values was at −0.69 V vs. SCE. For many n-type semiconductors, Efb is considered to be about 0.1 V below the conduction band (Ecb).38 Based on this, the estimated Ecb value of the BiO sample was −0.55 V vs. NHE. According to the thermodynamic data of N2H3 reported by Bauer,39 the N2 reduction potential via N2 + 3H+ + 3e− → N2H3 is not larger than −0.485 V vs. NHE. This indicates that the photogenerated electrons in the BiO catalyst could energetically reduce the N2 molecule via a proton-coupled three-electron reduction process. Besides, the valence band potential of the BiO catalyst, which was estimated from the band gap and Ecb, was also energetically large enough (2.39 V vs. NHE) for water oxidation (1.23 V vs. NHE) under acidic conditions, indicating that water could efficiently act as a sacrificial electron donor for ammonia generation. All of these studies proved the possibility of a highly efficient one-step three-electron N2 reduction on the BiO catalyst, which may be realized by the synergy of three adjacent low valence Bi(II) activation centers. This process does not occur easily in a traditional N2 reduction photocatalyst, which only has isolated defect-type activation centers.
The crystal structure and coordination environment of the BiO compound may play an important role in its highly efficient photocatalytic N2 reduction performance. As shown in Fig. 2d, the BiO compound exhibits a hexagonal tunnel structure along the [010] direction, which is built up from three alternately arranged Bi groups and O groups. The entrance diameter of this tunnel structure is about 4.4–4.7 Å, large enough for structurally filling one N2 molecule (diameter 3 Å). A careful observation of the local structure of BiO indicates that BiO4 presents a pyramidal unit with two ∠OBiO angles of 114.2° and 104.2°, far from a tetrahedral structure (Fig. S7†). This indicates that bismuth atoms may undergo an orbital hybridization of sp3d resulting in the orbital distribution of a trigonal bipyramid, leaving a lone pair of electrons on the Bi atom. The N2 molecule may be stretched and activated by three alternatively arranged Bi atoms through donating electrons to the empty 6d orbitals of Bi atom and accepting electrons from the lone pairs of Bi atom to its three unoccupied anti-bonding orbitals (π*2py, π*2pz and σ*2px), resulting in a 1N2–3Bi side-on bond structure (Fig. 6a). The electron donations from the three adjacent low valence bismuth ions will strongly weaken the NN bond, which facilitates a one-step three electron N2 reduction process in a successive photocatalytic process by proton and electron transfer (Fig. 6b).
Based on the above analysis, the possible N2 reduction reactions on BiO surface under acidic conditions are proposed as follows.
BiO + hv → BiO (h+ + e−) | (1) |
N2 + 3BiO (e−) + 3H+ → N2H3 | (2) |
N2H3 + 3BiO (e−) + 5H+ → 2NH+4 | (3) |
N2H3 + BiO (e−) + H+ → N2H4 | (4) |
2N2H3 → 2N2 + 3H2 | (5) |
3H2O + 6BiO (h+) → 3/2O2 + 6H+ | (6) |
First, simulated solar light excited electrons and holes are generated on the BiO catalyst (eqn (1)). Then the photogenerated electrons may transfer to the surface and participate in a one-step three electron N2 reduction process to obtain an N2H3 intermediate (eqn (2)) due to the possible 1N2–3Bi side-on bond structure. The N2H3 intermediate will be further reduced to NH4+ (eqn (3)) or N2H4 (eqn (4)) by a subsequent one step three-electron or one-electron photocatalytic reduction process, respectively. Otherwise, the N2H3 intermediate will be decomposed, because it is a short-lived, endothermic compound.39,40 If the N2H3 intermediate is not further reduced, it will decompose into N2 and H2 spontaneously (eqn (5)). In our experiment, the reaction rates of eqn (2) and (3) may be much higher than that of eqn (5), resulting in continuous production of NH4+ under the acidic conditions. The amount of N2H4 in the final product was only about 1.6%, possibly because of the 1N2–3Bi side-on coordinating mode of N2 which facilitates the three-electron reduction process (eqn (2) and (3)). During the photocatalytic N2 reduction process, O2 was produced from water oxidation by photogenerated holes (eqn (6)). The theoretical molar ratio of the generated NH4+ to O2 from the above equations is 1.33. In a closed reaction system filled with pure N2 and water, we observed continuous O2 evolution (Fig. S8†) along with prolonged irradiation time and a total amount of 117 μmol O2 was obtained when generating 2.32 mg NH4+. The molar ratio of the generated NH4+ to O2 is 1.41, which is close to the theoretical value. Besides, a tiny amount of H2 (2.4 μmol) was also detected within 24 h in the closed reaction system, which may have originated from the decomposition of a trace amount of the unconverted N2H3 intermediate.
Footnote |
† Electronic supplementary information (ESI) available: Fig. S1–S8. See DOI: 10.1039/c6ta09275f |
This journal is © The Royal Society of Chemistry 2017 |