Significance
Deployment of electric vehicles with long driving ranges demands new electrochemistry beyond the current Li-ion batteries. Aprotic lithium–oxygen batteries have gained considerable attention and promise to provide energy density of approximately three to five times that of state-of-the-art Li-ion batteries, but they are plagued by sluggish reaction kinetics at cathodes and induced large-voltage hysteresis. A photo-mediated strategy has been attempted to reduce the discharge/charge overvoltage, but it is obstructed by the limited ultraviolet light absorption and the inevitably high carrier recombination on semiconductor photocathodes. Here, a plasmonic heterojunction of gold nanoparticle-decorated carbon nitride with nitrogen vacancies is used as a bifunctional catalyst to accelerate the oxygen cathode reactions of the lithium–oxygen battery under visible light.
Keywords: oxygen reduction, oxygen evolution, surface plasmon, vacancy, visible light
Abstract
Aprotic lithium-oxygen (Li-O2) batteries have gained extensive interest in the past decade, but are plagued by slow reaction kinetics and induced large-voltage hysteresis. Herein, we use a plasmonic heterojunction of Au nanoparticle (NP)–decorated C3N4 with nitrogen vacancies (Au/NV-C3N4) as a bifunctional catalyst to promote oxygen cathode reactions of the visible light–responsive Li-O2 battery. The nitrogen vacancies on NV-C3N4 can adsorb and activate O2 molecules, which are subsequently converted to Li2O2 as the discharge product by photogenerated hot electrons from plasmonic Au NPs. While charging, the holes on Au NPs drive the reverse decomposition of Li2O2 with a reduced applied voltage. The discharge voltage of the Li-O2 battery with Au/NV-C3N4 is significantly raised to 3.16 V under illumination, exceeding its equilibrium voltage, and the decreased charge voltage of 3.26 V has good rate capability and cycle stability. This is ascribed to the plasmonic hot electrons on Au NPs pumped from the conduction bands of NV-C3N4 and the prolonged carrier life span of Au/NV-C3N4. This work highlights the vital role of plasmonic enhancement and sheds light on the design of semiconductors for visible light–mediated Li-O2 batteries and beyond.
The aprotic lithium-oxygen (Li-O2) battery promises ultrahigh theoretical energy density (∼3,600 Wh·kg−1) and is operated with oxygen reduction to generate the product of Li2O2 and its reverse oxidation (2Li+ + O2 + 2e− ↔ Li2O2, E0 = 2.96 V) (1–5). The sluggish oxygen cathode reactions, including the oxygen evolution reaction (OER) and the oxygen reduction reaction (ORR), lead to a high discharge/charge overvoltage (∼1.0 V) during cycles and low round-trip efficiency (6–9). Since the pioneering work on the photoinvolved Li-O2 battery using TiO2 (10) or C3N4 (11) under ultraviolet (UV)-light irradiation, reduction of the charge/discharge overvoltage via a photomediated strategy has been extensively studied and is anticipated to solve the kinetic issues of the Li-O2 battery (12–18). However, the light absorption of most semiconductors used is confined in the region of UV light, accounting for only ca. 4% of the solar spectrum (14–16). Expanding the light harvesting from UV to visible light is the long-term goal and challenge of photocatalysis (17–20). Simultaneously, high carrier recombination consumes the majority of photoelectrons and holes before catalyzing the targeted reactions, resulting in a mismatch between the carrier lifetime and kinetics of ORR or OER (19–21). This necessitates a structural design of semiconducting materials for visible-light harvesting to accelerate the cathode reactions in Li-O2 batteries.
Localized surface plasmon resonance (LSPR), which refers to the collective oscillation of conduction band (CB) electrons in metal nanocrystals under resonant excitation, has recently gained much attention (22–25). The decay of excited LSPR can produce hot electrons and holes, which initiate various chemical reactions (22, 23). Intriguingly, when plasmonic metal (e.g., Au, Ag) nanoparticles (NPs) come into contact with a semiconductor such as MoS2, TiO2, etc., an interfacial Schottky barrier forms; this barrier functions as a filter to force the energetic electrons or holes to migrate across the interface while inhibiting their reverse movement, thereby leading to effective electron–hole separation and suppressed charge–carrier recombination (26–30). LSPR systems generally are composed of plasmonic metal and semiconductors and exhibit the benefits of a low electron–hole recombination rate, enhanced light harvesting, and tailored response wavelengths from the visible to the near-infrared region (22). Recently, Au/CdSe (31) and Au/Ni(OH)2 (32) heterojunctions have been attempted for a photocatalytic hydrogen evolution reaction and OER with the aid of hot electrons and holes under visible light. Coupling the plasmonic metal with suitable semiconductors for broadened light harvesting and a plasmon-enhanced effect is highly desirable for both ORR and OER in the Li-O2 battery.
Herein, we report defective C3N4 (Au/NV-C3N4) decorated with plasmonic Au NPs as a bifunctional heterojunction catalyst that promotes cathode reactions of the Li-O2 battery under visible light. The NV on NV-C3N4 is prone to adsorb and activate O2, and the plasmon-excited electrons on Au migrate to the CB of NV-C3N4 and relax to the NV-induced defect band (DB) for O2 reduction to LiO2; then it undergoes electron reduction to Li2O2. Reversely, the Li2O2 is removed by the holes on the Au NPs driven by the applied voltage. The discharge voltage is raised to 3.16 V, and the charge voltage is lowered to 3.26 V at 0.05 mA·cm−2 with a good rate capability and cycle stability. This investigation integrates a plasmonic heterojunction into the aprotic Li-O2 battery and illustrates photoenergy conversion and storage under visible light.
Results and Discussion
Characterization of Au/NV-C3N4.
The NV-C3N4 and C3N4 were synthesized by heating guanidine hydrochloride at 600 °C and 550 °C, respectively, in an argon atmosphere (33). They were verified by X-ray diffraction (XRD) patterns (Fig. 1A) and Fourier-transform infrared (FTIR) spectra (SI Appendix, Fig. S1), in accordance with the molecular structure of C3N4 as shown in SI Appendix, Fig. S2. The presence of NV on NV-C3N4 is revealed by electron paramagnetic resonance (EPR) analysis. Compared with C3N4, which has rare defects, an enhanced EPR signal at g = 2.003 is visible on NV-C3N4, as shown in Fig. 1B, assigned to the unpaired electrons on the carbon atoms. This indicates rich defects on NV-C3N4. The NV position on NV-C3N4 is further ascertained by X-ray photoelectron spectroscopy (XPS). As shown in SI Appendix, Fig. S3, the ratio between N2c (two-coordinated N) and N3c (three-coordinated N) in the N 1s spectrum drops drastically to 1.2 from 3.7 in C3N4, indicating that the NV is preferentially located at N2c sites, as depicted in Fig. 1B, Inset. An additional C-NV peak at 289.3 eV (33), as shown in SI Appendix, Fig. S3, also confirms the presence of vacant N2c atoms for the formation of two-coordinated carbon.
Fig. 1.
(A) XRD patterns of Au/NV-C3N4 and NV-C3N4. (B) EPR spectra of C3N4, NV-C3N4, and Au/NV-C3N4. Inset shows the schematic of NV-engineered C3N4. (C) Au 4f and N 1s XPS spectra of Au/NV-C3N4. (D) C 1s XPS spectra of NV-C3N4 and Au/NV-C3N4. (E and F) Scanning transmission electron microscopy and high-resolution transmission electron microscopy images of Au/NV-C3N4.
Au NPs were prepared by a sodium citrate reduction method and deposited on NV-C3N4 to form Au/NV-C3N4. The characteristic peaks at 38.2° and 44.5° are assigned to the (111) and (200) planes of Au shown in Fig. 1A, indicating the loading of Au NPs on NV-C3N4. Fig. 1C shows the Au 4f and N 1s XPS spectra of Au/NV-C3N4, in which the Au 4f5/2 peak at 87.1 eV and the 4f7/2 peak at 83.5 eV further reveal the presence of Au NPs. The N 1s spectrum is deconvoluted to N-(C)3 (N3c, 399.8 eV), C-N = C (N2c, 398.6 eV), and C-N-H (400.9 eV), of which the ratio of N2c to N3c equals 1.1 and is similar to that of NV-C3N4, as shown in SI Appendix, Fig. S3. These findings suggest the successful preparation of Au/NV-C3N4. In addition, the C 1s XPS spectra in Fig. 1D show that the binding energies of N = C-N2 and C-NV on Au/NV-C3N4 are negatively shifted about 0.4 and 0.2 eV, respectively, as compared with those on NV-C3N4. The binding energy of Au 4f in Au/NV-C3N4 is positively shifted by ca. 0.3 eV, compared with that of pure Au NPs as shown in SI Appendix, Fig. S4. These findings confirm that the electrons transfer from Au to NV-C3N4 (30). Fig. 1E displays the high-angle annular dark-field scanning transmission electron microscopy image of Au/NV-C3N4, indicating that the Au NPs with an average size of 30 nm are well dispersed on NV-C3N4 without aggregation, in accordance with the high-resolution transmission electron microscopy images of Au/NV-C3N4 in SI Appendix, Fig. S5. The lattice spacing of 0.23 nm in Fig. 1F is well matched with the (111) plane of Au NPs, and their intimate contact with NV-C3N4 ensures facile interfacial electron transfer. Thermogravimetric analysis of Au/NV-C3N4, as shown in SI Appendix, Fig. S6, indicates that the loading of Au NPs on Au/NV-C3N4 is 1.5 wt %.
Fig. 2A displays the ultraviolet-visible (UV-Vis) absorption spectra of Au/NV-C3N4 and NV-C3N4. The NV-C3N4 shows an absorption edge of 486 nm. The converted Tauc plot shows a band gap (Eg) of 2.67 eV for NV-C3N4, as shown in SI Appendix, Fig. S7A. Compared with C3N4 in SI Appendix, Figs. S8 and S9, a decaying tail over the visible region originates from the NV-induced defect states within the band gap (34), which is consistent with the calculated projected density of state (PDOS) of NV-C3N4 as shown in Fig. 2B. The NV-induced DB value of NV-C3N4 is estimated to be 2.35 eV according to the converted Urbach plot in SI Appendix, Fig. S7B. The UV photoelectron spectrum of NV-C3N4 in SI Appendix, Fig. S10 indicates that the maximum valence band (VB) value is 4.40 V vs. Li+/Li in NV-C3N4 (35). Correspondingly, based on the formula ECB = EVB − Eg, the CB value is calculated to be 1.73 V vs. Li+/Li. The standard potential of O2/Li2O2 is located between the DB and VB of NV-C3N4 (SI Appendix, Fig. S10), which fulfills the thermodynamic prerequisites of photocatalytic ORR and OER. In contrast to NV-C3N4, Au/NV-C3N4 presents another strong UV/Vis absorption peak at 540 nm as shown in Fig. 2A, which is caused by the plasmonic excitation of Au NPs. The LSPR peak of colloidal Au NPs is at 533 nm, as shown in SI Appendix, Fig. S11. The clear red shift also provides evidence of electron interaction between Au NPs and NV-C3N4 (26), consistent with the XPS spectra in Fig. 1F.
Fig. 2.
(A) UV/Vis absorption spectra of Au/NV-C3N4 and NV-C3N4. (B) Calculated PDOS of NV-C3N4. (C) FDTD simulations of Au/NV-C3N4 and NV-C3N4. E0 and E represent the incident and localized electric field, respectively. (D) Time-resolved PL spectra of Au/NV-C3N4 and NV-C3N4.
The electric-field intensities around Au/NV-C3N4 and NV-C3N4 are simulated under illumination wavelengths of 350 to 800 nm by the finite difference time domain (FDTD) method, as shown in Fig. 2C. Compared with NV-C3N4, the electric-field intensity of Au/NV-C3N4 is increased by 10 times near the Au NPs with spatially confined “hot spots,” suggesting that the plasmon-induced hot electrons and holes are generated at the interface of NV-C3N4 and Au NPs. Steady-state and time-resolved photoluminescence (PL) spectroscopies were conducted to analyze the charge separation on NV-C3N4 and Au/NV-C3N4. The PL emission of Au/NV-C3N4 is much weaker than that of NV-C3N4, as shown in SI Appendix, Fig. S12, indicating an inhibited charge recombination on Au/NV-C3N4. The PL lifetimes of Au/NV-C3N4 and NV-C3N4 are calculated to be 3.23 and 1.76 ns, respectively, as shown in Fig. 2D. The prolonged PL lifetime also reflects the inhibitive charge recombination after loading Au onto NV-C3N4. Surface photovoltage spectroscopy (SPV) is used to analyze the charge separation degree at nanoscale and to record the surface voltage change under light irradiation. The stronger SPV signal on Au/NV-C3N4 in SI Appendix, Fig. S13 reveals the efficient charge separation and higher carrier concentration of Au/NV-C3N4.
Electrochemical Performances.
The cathodes were prepared by spraying the suspension of Au/NV-C3N4, NV-C3N4, Au/C3N4, or C3N4 onto carbon paper, as shown in SI Appendix, Fig. S14. The photoinvolved Li-O2 battery is composed of an Au/NV-C3N4 cathode, an electrolyte of 1.0 M lithium bis(trifluoromethane) sulfonimide (LiTFSI) in tetraglyme (G4), and a Li metal anode, which are assembled in a quartz glass cell (SI Appendix, Fig. S15). Fig. 3A displays the galvanostatic discharge/charge profiles of the Li-O2 battery with Au/NV-C3N4 and NV-C3N4 with or without illumination at 0.05 mA·cm−2. The Li-O2 battery with Au/NV-C3N4 under illumination (Au/NV-C3N4-L) shows an increased discharge voltage of 3.16 V and a lowered charge voltage of 3.26 V, corresponding to an ultrahigh round-trip efficiency of 97.0%. Without illumination (Au/NV-C3N4-D), it exhibits discharge and charge voltages of 2.68 and 4.22 V, respectively, with a round-trip efficiency of 63.5%. The increased round-trip efficiency under illumination is ascribed to solar energy conversion and storage during discharge and charge, respectively. In contrast, the Li-O2 battery with NV-C3N4 under illumination (NV-C3N4-L) presents charge and discharge voltages of 3.43 and 3.05 V, respectively, with an energy efficiency of 88.9%, as shown in SI Appendix, Fig. S16A. The high efficiency of Au/NV-C3N4 is attributed to the migration of hot electrons from plasmonic Au to NV-C3N4. In addition, the Li-O2 battery with Au/C3N4 under illumination (Au/C3N4-L) exhibits discharge and charge voltages of 3.02 and 3.39 V, respectively, possessing a round-trip efficiency of 89.1%, as shown in SI Appendix, Fig. S16B. The higher efficiency of Au/NV-C3N4 compared with Au/C3N4 indicates that the existence of NV promotes the oxygen cathode reaction kinetics upon discharging and charging with extended visible-light absorption, as well as C3N4, as shown in SI Appendix, Fig. S16C. In addition, the discharge/charge profile of Au/NV-C3N4 with illumination in an Ar atmosphere (SI Appendix, Fig. S17) excludes the intrinsic redox of Au/NV-C3N4 and the good stability of battery materials.
Fig. 3.
(A) Discharge/charge profiles of the Li-O2 battery with Au/NV-C3N4 and NV-C3N4 with or without illumination at 0.05 mA·cm−2. (B) Discharge and charge profiles of Au/NV-C3N4 and NV-C3N4 at various current densities under illumination. (C) Cycling performance with Au/NV-C3N4 and NV-C3N4 cathodes under illumination. (D) IPCE of Au/NV-C3N4 and NV-C3N4 at 420, 520, 550, 600, and 700 nm and absorption spectrum of Au NPs.
When the current density is raised to 0.10, 0.15, and 0.25 mA·cm−2, the charge/discharge overvoltages of the Li-O2 battery with Au/NV-C3N4-L are 0.14, 0.20, and 0.27 V, respectively, as shown in Fig. 3B. In the dark, the charge/discharge overpotentials rise quickly under the same current densities as those in SI Appendix, Fig. S18. The reduced polarization under illumination benefits from the photogenerated hot electrons and holes on Au/NV-C3N4. Accordingly, the charge/discharge overvoltages of NV-C3N4-L are 0.45, 0.54, and 0.68 V at 0.10, 0.15, and 0.25 mA·cm−2, respectively. As the current densities are raised from 0.05 to 0.25 mA·cm−2, the difference in charge/discharge overvoltages of the Li-O2 battery between Au/NV-C3N4 and NV-C3N4 becomes more obvious. This suggests that the life span of charge carriers, rather than oxygen cathode reaction kinetics, is the rate-determining step, in good agreement with the time-resolved PL spectra shown in Fig. 2D. Cycling performances of the Li-O2 battery with Au/NV-C3N4 and NV-C3N4 at 0.15 mA·cm−2 are displayed in Fig. 3C. With light irradiation, the Li-O2 battery with Au/NV-C3N4-L displays low overvoltages upon discharging and charging for 50 cycles, in sharp contrast to the greater charge/discharge overvoltages of NV-C3N4-L. This suggests that the plasmonic excitation on illuminated Au/NV-C3N4 is favorable for ORR and OER kinetics and cycling stability.
The increased voltage polarization after 50 cycles may be ascribed to the minimal side reactions at the cathode (1, 4). Iodometric titration is used to quantify the generated Li2O2 on the discharged cathode and to further analyze the proportion of side reaction. The discharge product of Li2O2 reacts with H2O via Li2O2 + 2H2O → H2O2 + 2LiOH; next, H2O2 oxidizes iodide to iodine, which is titrated by Na2S2O3 (SI Appendix, Fig. S19 and Supplementary Method). It reveals the highly reversible 1.99 e−/O2 reaction, compared with the theoretical 2.00 e−/O2. The minimal side reaction products, such as carbonates, may induce increased voltage polarization during cycles. On the other hand, upon charging, the remaining little Li2O2 particles away from the cathode/catalyst surface may lose electric contact and result in an increased charge overvoltage. The round-trip efficiency of the Li-O2 battery with Au/NV-C3N4 still reaches 92.5% after 50 cycles. In contrast, in the reported literature, the discharge/charge voltage polarization after cycles obviously increased, with low round-trip efficiencies (below 85%) due to the high charge carrier recombination and side reactions at high voltages (14–16). In addition, SI Appendix, Fig. S20A shows the discharge/charge profiles of the Li-O2 battery with Au/NV-C3N4, where illumination is only applicable for charge. It exhibits a lowered charge voltage of 3.30 V with a round-trip efficiency of 82.0%, which outperforms the other competitive cathodes for the Li-O2 battery (4, 5). A high capacity of 2.50 mAh·cm−2 for the photoinvolved Li-O2 battery has been produced, as depicted in SI Appendix, Fig. S20B, and this suggests its practicality.
In addition, the plasmonic effect strongly depends on the wavelength of incident light (36). To verify that the enhanced battery performance of Au/NV-C3N4 benefits from the plasmonic effect of Au NPs, the incident photon to current efficiency (IPCE) at different wavelengths on Au/NV-C3N4 and NV-C3N4 was evaluated (28, 29), as depicted in Fig. 3D and SI Appendix, Fig. S21 and Table S1. The IPCE of NV-C3N4 is much lower than that of Au/NV-C3N4, which reveals efficient plasmonic excitation on Au/NV-C3N4. The highest IPCE of Au/NV-C3N4 is at ∼550 nm, near the LSPR absorption peak of Au NPs, as shown in Fig. 3D. The IPCE of Au/NV-C3N4 at varied wavelengths follows the trend of extinction spectrum of Au NPs, confirming the wavelength-dependent performance of the Li-O2 battery on the plasmonic effect of Au NPs. A tabular form summarizing the cathode catalysts for the photoinvolved Li-O2 battery has been provided in SI Appendix, Table S2. The Li-O2 battery with an Au/NV-C3N4 cathode displays extended visible-light absorption, prolonged carrier lifetime, and better electrochemical performance.
Reaction Mechanism.
The discharge product of Li2O2 is identified by XRD in Fig. 4A in the Li-O2 battery with Au/NV-C3N4-L, in line with the Raman spectra in SI Appendix, Fig. S22A with a characteristic O-O stretching vibration of Li2O2 at 780 cm−1 (37). XPS spectra also confirm the formation of Li2O2 in the discharged Au/NV-C3N4 electrode, according to the Li 1s signal at 55.0 eV in Fig. 4B and the O 1s signal at 531.2 eV in Fig. 4C (11). The C 1s spectrum in SI Appendix, Fig. S22B indicates no visible parasitic products upon discharging and charging, which is consistent with the iodometric titration results in SI Appendix, Supplementary Method. After charging, the product of Li2O2 is reversibly decomposed, as verified by the elimination of its XPS signals, Raman bands, and diffraction peaks, as shown in Fig. 4 and SI Appendix, Fig. S22. As shown in Fig. 4D and SI Appendix, Fig. S23, numerous toroidal Li2O2 particles are uniformly covered on cathodes after discharge and are reversibly decomposed upon charging. The reversible behavior is recorded by scanning electron microscopy images for 50 cycles, as shown in SI Appendix, Fig. S23. These indicate the good reversibility of formation and oxidation of Li2O2 in the Li-O2 battery with Au/NV-C3N4-L. The 1H NMR spectra in SI Appendix, Fig. S24 imply inconspicuous electrolyte decomposition after discharge and charge with illumination.
Fig. 4.
(A) XRD patterns. (B) Li 1s and (C) O 1s XPS spectra. (D) Scanning electron microscopy images of discharged and charged Au/NV-C3N4 cathodes under illumination.
Density functional theory calculations are further performed to simulate the oxygen redox chemistry on Au/NV-C3N4 under illumination. Fig. 5 A and B shows the differential charge densities of O2-adsorbed NV-C3N4 and C3N4. O2 is preferentially adsorbed on NV of NV-C3N4 with a larger adsorption energy (−3.33 eV) than that of C3N4 (−1.65 eV), which is also confirmed by the large O2 desorption capacity in the temperature-programmed desorption (TPD) profile of NV-C3N4 relative to C3N4 (SI Appendix, Fig. S25). As external electrons are injected into O2 molecules, the O-O bond can be weakened to present an elongated bond length of 1.38 Å, compared with 1.23 Å in free O2, which indicates the activation of O2. Free energy diagrams are plotted to describe the initial formation or dissociation of Li4O4 at different applied potentials, as shown in Fig. 5C. The adsorption configuration of NV-C3N4-O2 clearly shows that the C atoms around N2c defect sites strongly bond with O2 and then combine with Li+ for NV-C3N4-LiO2, as shown in SI Appendix, Fig. S26. The LiO2 intermediate is further confirmed by the characteristic peak at 1,088 cm−1 in the Raman spectrum of discharged cathode with a cutoff capacity of 0.20 mAh·cm−2, as shown in SI Appendix, Fig. S27. The entire steps of Li4O4 formation are exothermal at an open circuit potential of 0 V (U = 0 V), suggesting a spontaneous process on NV-C3N4. Such an energy path can stay downhill until the applied voltage reaches 2.59 V (UDC = 2.59 V), while the reverse dissociation of Li4O4 is thermodynamically favorable with the applied voltage exceeding 3.79 V (UC = 3.79 V).
Fig. 5.
(A and B) Differential charge densities of O2-adsorbed NV-C3N4 and C3N4. The yellow and blue areas represent charge accumulation and depletion, respectively. The value of isosurface is set to 0.003 e·bohr−3. (C) Free energy diagrams of battery reactions. UDC is the theoretically maximum discharge voltage, and UC is the theoretically minimum charge voltage. (D) PDOS of NV-C3N4-O2. Inset shows the partial charge density of the defect energy level of NV-C3N4. (E) PDOS of NV-C3N4 after loading with Au4 cluster. (F) Schematic of plasmonic hot electron injection and hole transfer for the respective discharge and charge processes in a photoinvolved Li-O2 battery with Au/NV-C3N4. EF, Fermi level; SPR, surface plasmon resonance.
Moreover, the PDOS of NV-C3N4-O2 is calculated to explain the interaction of O2 and the NV of NV-C3N4 under illumination, as depicted in Fig. 5D. It shows that midgap states appear in the vicinity of the Fermi level (EF) of NV-C3N4. As O2 is adsorbed onto NV-C3N4, there is an apparent peak overlap between the p states of O and the defect states of NV-C3N4, suggesting a strong orbital hybridization, which is consistent with the differential charge density of NV-C3N4-O2 shown in Fig. 5A. The generation and relaxation of photoelectrons on NV-C3N4 with illumination are depicted in SI Appendix, Fig. S28. The defect energy levels, located between the VB and CB of NV-C3N4, can capture the photogenerated electrons and further facilitate the O2 reduction. The Au4 cluster is adopted as a simplified model to analyze the hybridization between the d states of Au clusters and the defect level of NV-C3N4. After loading the Au4 cluster onto NV-C3N4, a notable hybridization between the d bands of the Au4 cluster and the defect states of NV-C3N4 is observed, as shown in Fig. 5E. This means that the Au4 cluster can provide extra d electrons to participate in ORR and enhance the catalytic activity of NV-C3N4.
The reaction mechanism on the Au/NV-C3N4 cathode of the Li-O2 battery is described in Fig. 5F and SI Appendix, Fig. S29. On discharge, O2 is first adsorbed and activated at the NV sites of NV-C3N4. The maximum absorption edge of Au/NV-C3N4 is about 620 nm, corresponding to the photon energy of 2.00 eV. The Au/NV-C3N4 absorbs visible-light photons beyond 2.00 eV to generate hot electrons and holes, which are separated by the space charge region at the interface of Au and NV-C3N4 (38, 39). The hot electrons are transferred to the CB of NV-C3N4, relax to the NV-induced DB, and then are injected into the molecular orbitals of O2 for reduction to LiO2, which further undergoes electron reduction to Li2O2. The theoretical discharge voltage equals the potential difference between the EF and Li+/Li (SI Appendix, Fig. S29A). According to the work function (W) of Au (∼5.0 eV) and the formula EF = W − 4.5 (22, 40), the EF is calculated to be ∼0.5 V vs. the normal hydrogen electrode, namely ∼3.54 V vs. Li+/Li. Therefore, the theoretical discharge potential considering the contribution of light energy is 3.54 V, higher than the equilibrium voltage of 2.96 V, meaning light energy conversion to electricity. In a reverse charge, the holes on Au NPs are driven to oxidize the Li2O2 by an applied voltage. The theoretical charge voltage is determined by the voltage gap between the DB and Li+/Li to be 2.35 V, lower than the equilibrium voltage of 2.96 V (SI Appendix, Fig. S29B), and corresponding to light energy conversion. The achievable discharge and charge voltages of the Li-O2 battery with Au/NV-C3N4 are 3.16 and 3.26 V, respectively, at 0.05 mA·cm−2, and they deviate from the theoretical values (3.54 and 2.35 V, respectively) owing to the low utilization efficiency of light energy and the transport barrier between the catalyst and the Li2O2. Controlling the realistic loading of plasmonic metal NPs within a rationally designed three-dimensional porous cathode may improve the utilization efficiency of light and further promote the cathode reaction kinetics in the photoinvolved Li-O2 battery.
Conclusions
A plasmonic heterojunction, namely Au/NV-C3N4, is used as the bifunctional catalyst for a visible light–responsive Li-O2 battery to facilitate the formation and oxidation of Li2O2. The hot electrons generated on Au NPs transfer to NV sites for O2 reduction to LiO2, which subsequently undergoes reduction to Li2O2. Reversely, the holes on Au NPs moved by an applied voltage drive the Li2O2 oxidation to O2 and Li+. The achievable discharge voltage of the Au/NV-C3N4 cathode is up to 3.16 V, surpassing the equilibrium voltage by about 200 mV; the charge voltage is decreased to 3.26 V at 0.05 mA·cm−2 from 4.22 V in the dark, corresponding to an ultrahigh round-trip efficiency of 97.0%. Even at 0.25 mA·cm−2, the discharge and charge voltages of the Au/NV-C3N4 cathode reach 3.08 and 3.35 V, respectively. The stable discharge and charge of the Au/NV-C3N4 cathode are sustained for 50 cycles. The superior rate capability and cycle stability result in the plasmon-induced hot carrier transfer from Au NPs to NV-C3N4 and the good matching of the carrier life span with oxygen redox kinetics on Au/NV-C3N4. This investigation will promote the rational design of semiconductor electrodes for photoinvolved Li-O2 batteries and beyond.
Materials and Methods
Preparation of the NV-C3N4 and C3N4.
The NV-C3N4 and C3N4 were synthesized by heating guanidine hydrochloride (Aldrich; 99%) in an open crucible at 600 °C and 550 °C, respectively, for 4 h, with an Ar flow of 100 mL·min−1 at a ramp rate of 4 °C·min−1. The resultant yellow agglomerates were milled into powder in a mortar. To increase the hydrophilicity of NV-C3N4, 1 g of NV-C3N4 was vigorously stirred in 30 mL of concentrated hydrochloric acid for 1 h at room temperature. The mixture was filtrated and washed with deionized water repeatedly to remove the superfluous acid.
Preparation of Au/NV-C3N4 and Au/C3N4.
One milliliter of 1.0 wt % HAuCl4 solution was added to 96 mL deionized water and then heated to boil under vigorous stirring; 1.5 mL of 1.0 wt % sodium citrate solution was injected into the above solution until the color changed to wine red, indicating the formation of Au NPs. The acid-processed NV-C3N4 powder was dispersed in 20 mL of deionized water under vigorous stirring, into which the as-prepared Au colloidal solution was added dropwise. After stirring for 2 h, the resulting sample (denoted as Au/NV-C3N4) was collected, washed with deionized water several times, and dried in a vacuum at 100 °C for 10 h. The preparation of Au/C3N4 was similar to that of Au/NV-C3N4, except for the replacement of NV-C3N4 by C3N4.
Battery Assembly and Electrochemical Measurement.
The cathode suspension was prepared by mixing Au/NV-C3N4, NV-C3N4, Au/C3N4, or C3N4 powder and sodium carboxymethylcellulose with a weight ratio of 9:1. Next, the suspension was sprayed onto carbon paper and dried at 110 °C for 12 h to obtain the Au/NV-C3N4 electrode. The average mass loading of active material in each electrode was ∼0.4 mg·cm−2. The battery assembly was performed in an Ar-filled glove box with both H2O and O2 contents below 0.1 ppm. The Li-O2 battery was assembled in a sealed quartz glass cell, ensuring sufficient illumination of light on the electrodes with ignored light absorption by the glass. After purging with O2 into the cell for 30 min, galvanostatic discharge/charge cycles were conducted. The electrolyte was a solution of 1.0 M LiTFSI/G4, and Li metal foil was used as a negative electrode. The effective area of each electrode for illumination was 0.4 cm2. A 300-W Xe lamp was used as the light source with a wavelength range of 350 to 800 nm.
Characterization.
XRD (Rigaku MiniFlex600 X-ray generator; λ = 1.5406 Å, Cu Kα radiation) was used to conduct materials analyses. XPS measurements were performed on a Perkin-Elmer PHI 1600 ESCA system to characterize the cathode surface composition. Transmission electron microscopy (FEI Talos F200X G2; AEMC) was conducted to analyze the morphology and microstructure of the Au/NV-C3N4. Scanning electron microscopy (JEOL-JSM7500F) was conducted to observe the morphologies of the discharged and charged cathodes. Before the tests, dehydrated dimethoxyethane was used to wash the discharged/charged cathodes to remove residual electrolytes. The UV-vis absorption measurements were performed on a Thermo Scientific Evolution 220 UV-vis spectrophotometer. FTIR spectra were collected on a Bruker Tensor II Sample Compartment RT-DLaTGS. Raman spectra were recorded with a customized LabRAM HR800 confocal Raman microscope (Horiba Jobin Yvon). O2 TPD profiles were recorded by a Quantachrome CHEMBET3000 analyzer. 1H NMR spectra were obtained on a Bruker spectrometer (400 MHz).
Simulations of Plasmonic Near-Field Maps.
Plasmonic near-field maps were simulated at the LSPR wavelengths using the FDTD solution (Lumerical). The diameter of Au NPs was fixed at 30 nm. The scale of NV-C3N4 was set to be 15 × 60 × 150 nm3. The refractive index of NV-C3N4 was assumed to be 1.0, and the permittivity of Au was obtained by fitting experimental data from the literature.
Computational Details.
All the calculations were carried out using the Vienna Ab initio Simulation Package code (41–43). The generalized gradient approximation within the Perdew–Burke–Ernzerhof functional was adopted to describe the electronic exchange and correlation effects (43). The cutoff energy was set to 450 eV. The convergence criteria of the total energy and force were set to 10−5 eV and 0.02 eV·Å−1, respectively. The model of defective C3N4 adopted a 2 × 2 C3N4 supercell by removing a pyridine N atom. A vacuum layer of about 20 Å was used to prevent interaction between the periodic slabs. More calculation results are provided in SI Appendix, Note S1 and Table S3.
Supplementary Material
Acknowledgments
This work was supported by Nano Key Program of the Ministry of Science and Technology of China Grant 2017YFA0206700, National Natural Science Foundation of China Grants 21822506 and 51761165025, Tianjin Natural Science Foundation Grant 19JCJQJC62400, and the 111 Project of B12015.
Footnotes
The authors declare no competing interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.2024619118/-/DCSupplemental.
Data Availability
All study data are included in the article and/or SI Appendix.
References
- 1.Aurbach D., McCloskey B. D., Nazar L. F., Bruce P. G., Advances in understanding mechanisms underpinning lithium-air batteries. Nat. Energy 1, 16128 (2016). [Google Scholar]
- 2.Li F., Chen J., Mechanistic evolution of aprotic lithium-oxygen batteries. Adv. Energy Mater. 7, 1602934 (2017). [Google Scholar]
- 3.Zhu Z., et al., A hybrid Na//K+-containing electrolyte//O2 battery with high rechargeability and cycle stability. Research (Wash DC) 2019, 6180615 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Zhang P., Zhao Y., Zhang X., Functional and stability orientation synthesis of materials and structures in aprotic Li-O2 batteries. Chem. Soc. Rev. 47, 2921–3004 (2018). [DOI] [PubMed] [Google Scholar]
- 5.Zhang W., Shen Y., Sun D., Huang Z., Huang Y., Objectively evaluating the cathode performance of lithium-oxygen batteries. Adv. Energy Mater. 7, 1602938 (2017). [Google Scholar]
- 6.Chen Y., Freunberger S. A., Peng Z., Fontaine O., Bruce P. G., Charging a Li-O2 battery using a redox mediator. Nat. Chem. 5, 489–494 (2013). [DOI] [PubMed] [Google Scholar]
- 7.Lim H., et al., Rational design of redox mediators for advanced Li-O2 batteries. Nat. Energy 1, 16066 (2016). [Google Scholar]
- 8.Zhang W., et al., Promoting Li2O2 oxidation via solvent-assisted redox shuttle process for low overpotential Li-O2 battery. Nano Energy 30, 43–51 (2016). [Google Scholar]
- 9.Xia C., Kwok C. Y., Nazar L. F., A high-energy-density lithium-oxygen battery based on a reversible four-electron conversion to lithium oxide. Science 361, 777–781 (2018). [DOI] [PubMed] [Google Scholar]
- 10.Yu M., Ren X., Ma L., Wu Y., Integrating a redox-coupled dye-sensitized photoelectrode into a lithium-oxygen battery for photoassisted charging. Nat. Commun. 5, 5111 (2014). [DOI] [PubMed] [Google Scholar]
- 11.Zhu Z., Shi X., Fan G., Li F., Chen J., Photo-energy conversion and storage in an aprotic Li-O2 battery. Angew. Chem. Int. Ed. Engl. 58, 19021–19026 (2019). [DOI] [PubMed] [Google Scholar]
- 12.Schmidt D., Hager M. D., Schubert U. S., Photo-rechargeable electric energy storage systems. Adv. Energy Mater. 6, 1500369 (2016). [Google Scholar]
- 13.Gurung A., Qiao Q., Solar charging batteries: Advances, challenges, and opportunities. Joule 2, 1217–1230 (2018). [Google Scholar]
- 14.Yang X. Y., et al., An illumination-assisted flexible self-powered energy system based on a Li-O2 battery. Angew. Chem. Int. Ed. Engl. 58, 16411–16415 (2019). [DOI] [PubMed] [Google Scholar]
- 15.Gong H., et al., Photo-enhanced lithium oxygen batteries with defective titanium oxide as both photo-anode and air electrode. Energy Storage Mater. 13, 49–56 (2018). [Google Scholar]
- 16.Qiao Y., et al., Boosting the cycle life of aprotic Li-O2 batteries via a photo-assisted hybrid Li2O2-scavenging strategy. Small Methods 2, 1700284 (2018). [Google Scholar]
- 17.Lv Q., et al., Semiconducting metal-organic polymer nanosheets for a photoinvolved Li-O2 battery under visible light. J. Am. Chem. Soc. 143, 1941–1947 (2021). [DOI] [PubMed] [Google Scholar]
- 18.Li M., et al., A bifunctional photo-assisted Li-O2 battery based on a hierarchical heterostructured cathode. Adv. Mater. 32, e1907098 (2020). [DOI] [PubMed] [Google Scholar]
- 19.Liu X., et al., Utilizing solar energy to improve the oxygen evolution reaction kinetics in zinc-air battery. Nat. Commun. 10, 4767 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Fang Z., et al., Tactile UV- and solar-light multi-sensing rechargeable batteries with smart self-conditioned charge and discharge. Angew. Chem. Int. Ed. Engl. 58, 9248–9253 (2019). [DOI] [PubMed] [Google Scholar]
- 21.Du D., Zhao S., Zhu Z., Li F., Chen J., Photo-excited oxygen reduction and oxygen evolution reactions enable a high-performance Zn-air battery. Angew. Chem. Int. Ed. Engl. 59, 18140–18144 (2020). [DOI] [PubMed] [Google Scholar]
- 22.Linic S., Christopher P., Ingram D. B., Plasmonic-metal nanostructures for efficient conversion of solar to chemical energy. Nat. Mater. 10, 911–921 (2011). [DOI] [PubMed] [Google Scholar]
- 23.Ingram D. B., Linic S., Water splitting on composite plasmonic-metal/semiconductor photoelectrodes: Evidence for selective plasmon-induced formation of charge carriers near the semiconductor surface. J. Am. Chem. Soc. 133, 5202–5205 (2011). [DOI] [PubMed] [Google Scholar]
- 24.Wu K., Chen J., McBride J. R., Lian T., Efficient hot-electron transfer by a plasmon-induced interfacial charge-transfer transition. Science 349, 632–635 (2015). [DOI] [PubMed] [Google Scholar]
- 25.Xie W., Schlücker S., Hot electron-induced reduction of small molecules on photorecycling metal surfaces. Nat. Commun. 6, 7570 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Shi Y., et al., Hot electron of Au nanorods activates the electrocatalysis of hydrogen evolution on MoS2 nanosheets. J. Am. Chem. Soc. 137, 7365–7370 (2015). [DOI] [PubMed] [Google Scholar]
- 27.Li C., et al., Promoted fixation of molecular nitrogen with surface oxygen vacancies on plasmon-enhanced TiO2 photoelectrodes. Angew. Chem. Int. Ed. Engl. 57, 5278–5282 (2018). [DOI] [PubMed] [Google Scholar]
- 28.Yang J., et al., High-efficiency “working-in-tandem” nitrogen photofixation achieved by assembling plasmonic gold nanocrystals on ultrathin titania nanosheets. J. Am. Chem. Soc. 140, 8497–8508 (2018). [DOI] [PubMed] [Google Scholar]
- 29.Dong Y., et al., Plasmon-enhanced deuteration under visible-light irradiation. ACS Nano 13, 10754–10760 (2019). [DOI] [PubMed] [Google Scholar]
- 30.Wang S. S., et al., Boosting electrocatalytic hydrogen evolution over metal-organic frameworks by plasmon-induced hot-electron injection. Angew. Chem. Int. Ed. Engl. 58, 10713–10717 (2019). [DOI] [PubMed] [Google Scholar]
- 31.Shi R., et al., Self-assembled Au/CdSe nanocrystal clusters for plasmon-mediated photocatalytic hydrogen evolution. Adv. Mater. 29, 1700803 (2017). [DOI] [PubMed] [Google Scholar]
- 32.Liu G., et al., Promoting active species generation by plasmon-induced hot-electron excitation for efficient electrocatalytic oxygen evolution. J. Am. Chem. Soc. 138, 9128–9136 (2016). [DOI] [PubMed] [Google Scholar]
- 33.Niu P., Liu G., Cheng H. M., Nitrogen vacancy-promoted photocatalytic activity of graphitic carbon nitride. J. Phys. Chem. C 116, 11013–11018 (2012). [Google Scholar]
- 34.Zhao D., et al., Synergy of dopants and defects in graphitic carbon nitride with exceptionally modulated band structures for efficient photocatalytic oxygen evolution. Adv. Mater. 31, 1903545 (2019). [DOI] [PubMed] [Google Scholar]
- 35.Xiao G., Wang X., Li D., Fu X., InVO4-sensitized TiO2 photocatalysts for efficient air purification with visible light. J. Photochem. Photobiol. Chem. 193, 213–221 (2008). [Google Scholar]
- 36.Manjavacas A., Liu J. G., Kulkarni V., Nordlander P., Plasmon-induced hot carriers in metallic nanoparticles. ACS Nano 8, 7630–7638 (2014). [DOI] [PubMed] [Google Scholar]
- 37.Johnson L., et al., The role of LiO2 solubility in O2 reduction in aprotic solvents and its consequences for Li-O2 batteries. Nat. Chem. 6, 1091–1099 (2014). [DOI] [PubMed] [Google Scholar]
- 38.Zhang X., Chen Y. L., Liu R. S., Tsai D. P., Plasmonic photocatalysis. Rep. Prog. Phys. 76, 046401 (2013). [DOI] [PubMed] [Google Scholar]
- 39.Mascaretti L., et al., Plasmon-enhanced photoelectrochemical water splitting for efficient renewable energy storage. Adv. Mater. 31, 1805513 (2019). [DOI] [PubMed] [Google Scholar]
- 40.Lin H. Y., Chen Y. F., Wu J. G., Wang D. I., Chen C. C., Carrier transfer induced photoluminescence change in metal-semiconductor core-shell nanostructures. Appl. Phys. Lett. 88, 161911 (2006). [Google Scholar]
- 41.Kresse G., Furthmüller J., Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B Condens. Matter 54, 11169–11186 (1996). [DOI] [PubMed] [Google Scholar]
- 42.Kresse G., Furthmuller J., Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 6, 15–50 (1996). [Google Scholar]
- 43.Perdew J. P., Burke K., Ernzerhof M., Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All study data are included in the article and/or SI Appendix.





