Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2022 May 23;119(22):e2202913119. doi: 10.1073/pnas.2202913119

Spontaneous exciton dissociation in organic photocatalyst under ambient conditions for highly efficient synthesis of hydrogen peroxide

Huijie Yan a, Minhui Shen a, Yong Shen a, Xu-Dong Wang a, Wei Lin b, Jinhui Pan a, Jian He c, Yu-Xin Ye a,1, Xin Yang(杨欣) d, Fang Zhu a, Jianqiao Xu a, Jianguo He c, Gangfeng Ouyang a,e,f,1
PMCID: PMC9295752  PMID: 35605116

Significance

Hydrogen peroxide is a highly competitive ready-to-use product for solar energy transformation. Nevertheless, the contemporary photosynthetic systems are not efficient enough, due to severe charge recombination caused by high activation energy and binding energy of the exciton. Herein, we achieve spontaneous exciton dissociation at room temperature. Moreover, the photosynthesis of H2O2 reaches between 9,366 and 12,324 µmol·g−1 from 9 AM to 4 PM in ambient conditions, that is, sunlight irradiation, real water including fresh water and seawater, room temperature, and open air. The ultrahigh photocatalytic efficiency in ambient conditions allows the solar-to-chemical conversion in a real cost-effective and sustainable way, which represents an important step toward real applications.

Keywords: hydrogen peroxide, photocatalysis, exciton dissociation

Abstract

Hydrogen peroxide (H2O2) is a highly competitive ready-to-use product for solar energy transformation. However, charge recombination caused by the inefficient dissociation of exciton into free charges severely constrains the photocatalytic efficiencies, especially in ambient conditions. Herein, the photosynthesis of H2O2 is achieved in ambient conditions, that is, real water, open air, and sunlight irradiation, by a donor–bridge–acceptor conjugated polymeric photocatalyst with the remarkable productivity reaching between 9,366 and 12,324 µmol·g−1 from 9 AM to 4 PM. The photosynthesis efficiency of H2O2 in ambient conditions is even higher than all of the reported systems conducted in pure water and O2 atmosphere. The remarkably high efficiency is attributed to the spontaneously dissociated exciton at room temperature and the substantially suppressed back electron transfer through storing the photoinduced electron in redox electron acceptors. This efficient photosynthesis in ambient conditions allows the solar-to-chemical conversion in a real cost-effective and sustainable way, which represents an important step toward real applications.


Transformation of solar energy to ready-to-use resources is critical for transitioning toward a sustainable future (1, 2). In the most studied solar-to-chemical conversion (SCC) reactions, that is, CO2 reduction into C1 and C2 produces, dinitrogen reduction into ammonia, and water reduction into hydrogen, only half-reactions are in demand. By contrast, hydrogen peroxide (H2O2) production can fully utilize both the water oxidation and O2 reduction half-reactions, which avoids the consumption of sacrificial reagents, and thus achieves high SCC efficiency and is more cost effective (3, 4). Moreover, the Gibbs free energy change of producing H2O2 from O2 and H2O is much lower than the other conversion reactions. Thus, the photons in the full region of visible light (380 nm to 780 nm) are sufficient to actuate this reaction according to the energies of the photons. In addition, the reduction of dissolved O2 in water to H2O2 occurs prior to the reduction of protons, N2, and CO2 considering the standard reduction potentials, which indicates that it is thermodynamically possible to exclude the competitive reactions (5). Nowadays, the application of H2O2 is still expanding; besides its contemporary wide applications, it is a potential value-added fuel (6).

Organic photocatalysts that are constituted of earth-abundant elements and can be precisely tuned at molecular levels are used in most studies for the synthesis of H2O2 (2, 7, 8). However, the photosynthesis of H2O2 is not efficient enough, due to severe charge recombination. The generation of an exciton is a distinct characteristic of organic photocatalysts after absorbing photons (Fig. 1A, pathway 1), which is one pair of an electron and a hole bound by Coulombic force, owing to the low dielectric constant for organic materials (9, 10). To acquire completely dissociated charges, the Coulombic force must be overcome within the exciton lifetime (Fig. 1A, pathway 2); otherwise, the charges will undergo geminate recombination (Fig. 1A, pathway 3). The exciton binding energy (Eb) that is used to characterize the strength of the Coulombic force is typically on the order of 0.1 eV to 0.5 eV, which far outweighs the thermal energy (25 meV) at room temperature. Thus, an extra exciton activation energy (Ea) even higher than Eb is required to dissociate the exciton. Generally, the high Ea results in inefficient exciton dissociation, and the positive Eb even indicates that charge recombination is thermodynamically preferred (2, 11).

Fig. 1.

Fig. 1.

The mechanism behind the excellent performance in photosynthesis of H2O2 by TPT-alkynyl-AQ. (A and B) The formation pathways for free charges generation in previous works and this work. (C) Illustration of the pathway for electron transfer and storage.

To overcome this limitation, the strategies of promoting exciton dissociation (Fig. 1A, pathway 2) are reported (12, 13). For example, an external electric field or magnetic field is often required to provide a driving force opposite to the Coulombic force for promoting exciton dissociation (13). Also, to decrease Ea by tailoring the structures of the photocatalysts, the construction of strong charge-delocalized environments by fabricating electron donor–acceptor structures is validated as an effective way (14, 15). Through this method, the Ea can be decreased from larger than 100 meV to dozens of megaelectron volts. Still, the Ea is higher than the thermal energy at room temperature (25 meV). In addition, strong charge delocalization is a double-edged sword: Back electron transfer would be promoted simultaneously (Fig. 1A, pathway 3), which hinders the efficient generation of free charge carriers. Remarkably, the construction of built-in electric fields by fabricating heterojunctions and doping heteroatoms on traditional organic semiconductors is beneficial for decreasing both Ea and Eb, by providing driving forces and stabilizing charge carriers (1618). Nevertheless, the driving forces originating from the built-in electric fields only exist in limited regions within the photocatalysts, that is, the interfaces of the heterojunctions and the regions around the doped atoms. Moreover, heterojunctions require additional energies for driving electrons or holes across the interfaces (9).

Owing to the lack of efficient strategies for promoting exciton dissociation, most organic photocatalytic systems need a continuous supply of pure O2 and even organic hole scavengers to accelerate the reactions between charge carriers and reactants (4, 19). Thus, the apparent charge separation efficiency could be elevated. Photocatalytic synthesis of H2O2 in ambient conditions, that is, open air, real water, and solar light irradiation, remains insufficiently exploited.

Herein, we rationally design a donor–bridge–acceptor (D-B-A) conjugated polymeric photocatalyst (triphenyltriazine-alkynyl-anthraquinone [TPT-alkynyl-AQ]) exhibiting spontaneous exciton dissociation at room temperature (Fig. 1C). The spontaneous exciton dissociation is achieved by the cooperative action of the rigid alkynyl electron bridges and the redox electron acceptors (AQ) containing electron storage properties. The D-B-A structure forms a strong charge delocalization environment, and the AQ moieties provide a long-lived charge separation state (Fig. 1B, pathway 2 promoted, pathways 3 and 4 suppressed). Moreover, the alkynyl and AQ moieties present abundant reaction sites for efficient oxygen reduction reaction (ORR) and water oxidation reaction (WOR). Notably, the photosynthetic efficiency of H2O2 reaches 2,368 µmol·g−1·h−1 in open air and pure water. This efficiency is higher than all of the reported photocatalysts, even those using pure O2 for the production of H2O2. Moreover, the photosynthesis of H2O2 is achieved in ambient conditions, that is, sunlight irradiation, real water including fresh water and seawater, room temperature, and open air.

Results

Structure and Activity of the Photocatalysts.

The D-B-A conjugated polymer (CP) was synthesized with TPT moieties as the electron donors, and with built-in redox AQ moieties as the electron acceptors with alkynyl as connectors, which was named TPT-alkynyl-AQ (Fig. 2A). For comparison, D-A conjugated polymer (TPT-AQ) and another D-B-A conjugated polymer (TPT-imine-AQ; imine groups were the bridges between TPT and AQ moieties) were also synthesized (SI Appendix, Figs. S1–S9). AQ moieties exhibit excellent electron accepting ability as the lowest unoccupied molecular orbit of these CPs mainly located in them (SI Appendix, Fig. S1).

Fig. 2.

Fig. 2.

The structure and photocatalytic performance of CPs for H2O2 production. (A) Molecular structures of CPs used in this study. (B) The electronic structure of CPs. (C) The UV-visible diffuse reflectance spectra of CPs and the AQY of TPT-alkynyl-AQ at the specified wavelengths. (D) The photocatalytic performance of CPs for H2O2 production in open air and pure water. (E) The photosynthetic amount of H2O2 from TPT-alkynyl-AQ in pure water, river water collected from the Pearl River, or seawater from the South China Sea under ambient conditions, that is, open air and sunlight irradiation (from 9 AM to 4 PM).

The powder X-ray diffraction profiles revealed that all CPs exhibited the features of amorphous carbon (SI Appendix, Fig. S2) (20). The photos from scanning electron microscopy and transmission electron microscopy indicated the CPs were exfoliated structures (SI Appendix, Figs. S3 and S4). The formation of triazine ring; the retaining of AQ function groups; the existing of linkages, that is, alkynyl and imine between TPT and AQ in TPT-alkynyl-AQ and TPT-imine-AQ; and the porosity of these three polymers were confirmed carefully and in detail by various characterization methods (SI Appendix, Figs. S5–S9) (7, 21, 22).

The electronic structures of the CPs were also determined (Fig. 2B) (23, 24). As shown in Fig. 2C, CPs exhibited high absorption throughout the visible light range, which was consistent with the ash black colors of the CPs. The absorption band edges indicated that the bandgaps were 2.03, 2.36, and 2.23 eV for TPT-alkynyl-AQ, TPT-imine-AQ, and TPT-AQ, respectively (Fig. 2B). Moreover, the conduction band (CB) minima of TPT-alkynyl-AQ, TPT-imine-AQ, and TPT-AQ were determined to be −0.14, −0.3, and −0.4 eV versus the reversible hydrogen electrode via Mott–Schottky tests (SI Appendix, Fig. S10). Thereafter, the valence band maxima were calculated from the bandgaps and CBs, which were further confirmed by cyclic voltammetry measurement (SI Appendix, Fig. S11). All CPs had the capacities for 2e ORR, and 2e or 4e WOR (25).

The photosynthetic activities of the conjugated polymers were evaluated under xenon-lamp light with the UV light cut off (>400 nm, 100 mW·cm−2). No sacrificial agents or continuous O2 bubbling was adopted. TPT-alkynyl-AQ showed the highest H2O2 production rate, that is, 2,368 µmol·g−1·h−1, among the CPs, which was 2 and 5 times that of TPT-AQ and TPT-imine-AQ (Fig. 2D). And the apparent quantum yield (AQY) of TPT-alkynyl-AQ reaches as high as 25%, 18%, and 14% at 365, 425, and 450 nm, respectively, compared to only 10% and 7% at 450 nm for TPT-AQ and TPT-imine-AQ (Fig. 2C). Besides, in pure O2 atmosphere, the efficiency of TPT-alkynyl-AQ was further elevated to as high as 3,214 µmol g−1·h−1. To the best of our knowledge, the efficiency of TPT-alkynyl-AQ in open air was even higher than the reported photocatalysts used in pure O2 atmosphere (SI Appendix, Table S1) (4, 5, 17, 2631). More importantly, the efficiency of SCC is as high as 0.35% in pure water and open air, which was 3 times higher than the average solar-to-biomass conversion efficiency of typical plants (∼0.1%) (SI Appendix, Table S2) (3236). In addition, TPT-alkynyl-AQ maintained its high efficiency for five cycles without any decrease. Also, the morphology and the component of TPT-alkynyl-AQ were maintained (SI Appendix, Figs. S12 and S13). Meanwhile, the temperature-dependent photocatalytic generation of hydrogen peroxide was investigated. The photosynthetic efficiency of H2O2 reaches 564 µmol·g−1·h−1 at 277 K, which was much less than its performance at 298 and 308 K, attaining 2,030 and 2,067 µmol·g−1·h−1. At 333 K, the photosynthetic efficiency of H2O2 was only 1,503 µmol·g−1·h−1, because hydrogen peroxide was decomposed with the increase in temperature (SI Appendix, Fig. S14).

Furthermore, in ambient conditions, that is, open air and sunlight irradiation (from 9 AM to 4 PM), 9,295 µmol·g−1 of H2O2 was produced in pure water (Fig. 3 and SI Appendix, Figs. S15 and S1). For comparison, 12,042 and 9,366 µmol·g−1 of H2O2 were produced in river water; also, 10,634 and 12,324 µmol·g−1 of H2O2 were produced in seawater (Figs. 2E and 3 and SI Appendix, Figs. S15 and S16). TPT-alkynyl-AQ maintained its high efficiency for five cycles in real water without any decrease, which indicated that the complex matrices did not influence the photocatalytic efficiencies (SI Appendix, Fig. S17). Photosynthesis of H2O2 has never been realized under sunlight irradiation and in real water before.

Fig. 3.

Fig. 3.

Photosynthesis of H2O2 in ambient conditions. (A) The conditions for photosynthesis of H2O2 in ambient conditions, that is, sunlight, collected location of water samples, weather conditions, and temperature at the corresponding time. (B) Generated rates of H2O2 in various water samples under ambient conditions.

Charge Separation and Transfer Performance.

The reason behind the remarkable photosynthetic efficiency in ambient conditions was investigated. The overall recombination efficiencies of excitons (Fig. 1B, pathways 3 and 4) were monitored by steady-state photoluminescence (PL) emission spectroscopy (SI Appendix, Fig. S18) (17). The radiative recombination of excitons was negligible in TPT-alkynyl-AQ, while it was still significant in TPT-AQ and TPT-imine-AQ, indicating that the geminate charge recombination was greatly inhibited in TPT-alkynyl-AQ. The attribution of fluorescence peaks was also determined by fluorescence lifetime (SI Appendix, Fig. S19). The lifetimes of TPT-alkynyl-AQ, TPT-AQ, and TPT-imine-AQ are 0.74, 0.69, and 0.8 ns, respectively, by fitting PL decay curves, which were more consistent with the lifetime of exciton recombination than free carrier recombination. To further evaluate the influences of electron bridges and electron acceptors in CPs for inhibiting the exciton geminate recombination, exciton activation energy (Ea) was characterized through temperature-dependent PL spectra (Fig. 4 A and B and SI Appendix, Fig. S20) (8, 9). Surprisingly, the PL intensity increased with temperature for TPT-alkynyl-AQ, which was in sharp contrast to the other two organic semiconductors. This indicated that the energy barrier for exciton dissociation into free charge was lower than the thermal energy at room temperature, and the energy level of the charge-separated state was even lower than that of the exciton state (Fig. 4 C and D) (37, 38). In other words, Ea was smaller than 25 meV, and Eb was negative. This was rarely observed in organic photoactive materials, and demonstrated that the separation of excitons was spontaneous in TPT-alkynyl-AQ at room temperature. Thus, the geminate recombination (Fig. 1B, pathway 3) was highly depressed through promoting pathway 2. Based on the temperature-dependent PL, the Ea was estimated to be 42 meV for TPT-imine-AQ. In comparison, the unusual increase in PL with temperature indicated that Ea was smaller than the reverse potential barrier (Ear), as shown in Fig. 4D. Only the activation energy of charge recombination (Ear) could be estimated for TPT-alkynyl-AQ, due to the reversed temperature-dependent PL intensity, which was 22 meV. Thus, Ea for TPT-alkynyl-AQ should be even smaller than 22 meV, which verified the assumption above. Neither Ea nor Ear could be estimated for TPT-AQ, as the PL intensity varied irregularly with the temperature. Furthermore, Ea for TPT-alkynyl-TPT (a new photocatalyst without AQ moieties; SI Appendix, Figs. S20 and S21) was estimated to be 70 meV. These results demonstrated that both the alkynyl bridges and the AQ moieties were crucial for the decreasing of Ea.

Fig. 4.

Fig. 4.

Characterization of intrinsic properties of CPs. (A) Temperature-dependent photoluminescence (PL) spectra for TPT-alkynyl-AQ. (B) Evolution of PL intensity as a function of temperature from 100 K to 300 K for TPT-alkynyl-AQ. (C) Illustration of mutual transitions between the charge separated state (CS) and the lowest singlet excited state (S1) in the common materials in which the value of exciton binding energy (Eb) is positive and Ea is the activation energy from S1 to CS. (D) Illustration of mutual transitions between CS and S1 in TPT-alkynyl-AQ, in which the value of Eb is negative and Ear is the activation energy from CS to S1. (E) The EPR spectra of CPs under ambient conditions without additional illumination. (F) The odd-electron density in a single TPT-alkynyl-AQ unit, through DFT calculation.

To further investigate the embedded mechanism behind the low Ea, density functional dispersion (DFT) calculation was adopted. The introduction of the alkynyl electron bridges largely promoted the coplanarity of the TPT and AQ moieties. The dihedral angle between TPT and AQ was only 0.8° in TPT-alkynyl-AQ, while it was as large as 36° and 46° in TPT-AQ and TPT-imine-AQ, respectively (SI Appendix, Fig. S22). Thus, it was easier for AQ to withdraw the electron cloud from TPT after introducing the alkynyl bridge, to achieve a higher charge delocalization state. The calculated dihedral angle between two TPTs beside the alkynyl bridge was 0.036° in TPT-alkynyl-TPT.

It was highly remarkable that a stable charge separation state was observed in TPT-alkynyl-AQ at room temperature, as evidenced by the radical signal probed by electron paramagnetic resonance (EPR) analysis (Fig. 4E). The EPR signal was at g = 2.006, demonstrating that the radicals were mainly formed on the O atoms in the quinone groups. DFT calculations also confirmed the photoinduced electrons were mainly stored in the AQ moieties (Fig. 4F). These results indicated that AQ moieties could store photoinduced electrons under ambient conditions to avoid back electron transfer (Fig. 1B, pathway 3). Moreover, the strongest radical signal of TPT-alkynyl-AQ verified that the electron storage capacity of the AQ moieties was improved by the introduction of the alkynyl electron bridges, which should also be attributed to the better coplanarity for TPT-alkynyl-AQ.

In addition, the intramolecular charge transfer was characterized by transient absorption spectroscopy (TAS). Very broad TAS signals, between 900 and 1,200 nm, were observed, due to a range of energies for trap states (SI Appendix, Fig. S23), which were further assigned to electrons based on the amplitude decreased in the presence of AgNO3 as electron scavengers (39). The lifetimes of photoinduced electrons were at picosecond levels in TPT-alkynyl-AQ and TPT-AQ, and at nanosecond level in TPT-alkynyl-TPT (SI Appendix, Fig. S24). The ultrashort carrier lifetime was attributing to the intramolecular electron transfer from electron donor to acceptor. The photoinduced electron did not directly react with the free oxygen, but was stored in AQ moieties or used to reduce the oxygen absorbed on alkynyl moieties, which is discussed in more detail in the following section. As a result, although the carrier lifetime was very short, the ORR reaction was still very efficient. The extremely short lifetimes of electrons in TPT-alkynyl-AQ and TPT-AQ were consistent with the reported electron lifetimes for intramolecular transfer (9). The intramolecular transfer in TPT-alkynyl-AQ was also supported by DFT calculations, as no interlayer interaction between electron donors was observed (SI Appendix, Fig. S25). Thus, the recombination of free charges that occurred during intermolecular charge transfer in TPT-alkynyl-AQ was diminished (Fig. 1B, pathway 4).

Active Sites and Pathways for H2O2 Generation.

As the 4e ORR that produces H2O instead of H2O2 is a competitive reaction of the 2e ORR, the selectivity of ORR when using these three photocatalysts, that is, TPT-alkynyl-AQ, TPT-AQ, and TPT-imine-AQ, was investigated by rotating disk electrode voltammetry (SI Appendix, Fig. S26) (40). All three of these photocatalysts showed high ORR selectivity. The average electron transfer number involved in ORR was calculated to be 2.124, 2.097, and 2.004 for TPT-alkynyl-AQ, TPT- AQ, and TPT- imine-AQ, respectively, which were close to the theoretical value for directly and selectively reducing O2 into H2O2.

To elucidate the reaction routes, the active sites for ORR were studied first. DFT calculations results indicated that the spontaneous chemisorption of O2 dominantly occurred on the alkynyl electron bridges to form the endoperoxide species rather than on triazine ring or AQ moieties (SI Appendix, Fig. S27) (7). Furthermore, to investigate the subsequent ORR for the O2 absorbed on alkynyl, photosynthesis of H2O2 by TPT-alkynyl-AQ and TPT-AQ was conducted under illumination in Ar atmosphere with a hole scavenger to eliminate the O2 from the surrounding environment and water (SI Appendix, Fig. S28). The generation efficiencies of H2O2 by TPT-alkynyl-AQ (124 µmol⋅g−1⋅h−1) was 9 times higher than that in the TPT-AQ system (14 µmol⋅g−1⋅h−1) (SI Appendix, Fig. S28C). These strongly evidenced that the preadsorbed O2 on the alkynyl electron bridges could be subsequently reduced to H2O2.

Moreover, it was observed that the electrons stored in AQ moieties could reduce O2 into H2O2. The photosynthesis of H2O2 was conducted in Ar atmosphere, and then pure O2 gas was injected into the photocatalytic system in the dark. The generation of H2O2 lasted for 5 min in the TPT-alkynyl-AQ photocatalytic system after stopping the illumination, while no H2O2 was generated in TPT-alkynyl-TPT photocatalytic system (SI Appendix, Fig. S29). These phenomena indicated that the AQ moieties were another active site for ORR in TPT-alkynyl-AQ. And the reaction rate between reduced AQ and O2 was relatively slow, as it took about 5 min to consume the photoinduced electrons stored in AQ under high concentrations of O2.

Moreover, the rate-determining step of ORR in TPT-alkynyl-AQ was verified. Increasing the concentration of O2, both the photosynthesis efficiencies of H2O2 and photocurrent densities were enhanced in TPT-alkynyl-AQ but not in TPT-AQ (SI Appendix, Fig. S30). The photosynthesis efficiencies of H2O2 in open air and in pure O2 atmosphere achieved by TPT-alkynyl-AQ were 2,368 and 3,241 µmol⋅g−1⋅h−1, respectively (SI Appendix, Fig. S30A). And the photocurrent density of TPT-alkynyl-AQ in O2 was nearly twice that in Ar atmosphere (SI Appendix, Fig. S30C). All of these indicated that the supply rate of O2 determined the ORR rate in TPT-alkynyl-AQ, while the reaction rate between reduced AQ and O2 determined the ORR rate in TPT-AQ (41).

On the other hand, the products of WOR were determined by the rotating ring-disk electrode (RRDE) tests (SI Appendix, Fig. S31) (7). TPT-alkynyl-AQ TPT-AQ and TPT-imine-AQ exhibited significant reduction currents when the applied potentials at the Pt ring electrode were −0.23 V and +0.6V, which were attributed to the reduction of O2 and H2O2, respectively. Both of the reduction currents for O2 and H2O2 were improved in TPT-alkynyl-AQ. These results indicated that the efficiency of water oxidation was increased through introducing alkynyl moieties. Meanwhile, H2O2 was photosynthesized in Ar atmosphere by TPT-alkynyl-AQ (296 µmol⋅g−1⋅h−1). By contrast, only a small amount of H2O2 was generated in TPT-AQ (14 µmol⋅g−1⋅h−1), which was consistent with the results in RRDE tests that the alkynyl moieties were crucial for WOR in TPT-alkynyl-AQ.

Furthermore, time-dependent DFT calculation revealed that the photoinduced hole of excited TPT-alkynyl-AQ was mainly located on the alkynyl and AQ moieties, which demonstrated that the alkynyl and AQ moieties were the WOR active sites (Fig. 5A). Moreover, other calculation results revealed that the alkynyl moieties were the dominant sites for the generation of H2O2 by adsorbing OH (OH*), which was a crucial intermediate step in 2e WOR (SI Appendix, Fig. S32). On the other hand, DFT calculation demonstrated that free hydroxyl radical (·OH) was another potential product at the AQ or triazine active sites for WOR, which was detected by 5,5-dimethyl-1-pyrroline-N-oxide capturing experiments (SI Appendix, Fig. S33).

Fig. 5.

Fig. 5.

The proposed mechanism of photocatalytic H2O2 production. (A) The analysis for the distribution of the hole (blue) and electron (green) for TPT-alkynyl-AQ through TD-DFT. (B) The proposed pathways of photosynthetic H2O2 production.

According to the above characterizations and analyses, the photosynthetic pathways of H2O2 by TPT-alkynyl-AQ were proposed in Fig. 5B. When the O2 supply was adequate, the O2 was first adsorbed on the alkynyl electron bridges. Then the photoinduced electrons quickly reduced the O2 to generate H2O2, and the photoinduced holes were located in the alkynyl moieties to further oxidize the water to generate H2O2 (Fig. 5B, pathway 1). On the other hand, when the O2 supply was inadequate, the photoinduced electrons were first stored in the AQ moieties, and the holes were located in alkynyl moieties. Then, the holes oxidized the water to generate H2O2, and the electrons stored in AQ moieties reacted with O2 to produce H2O2 (Fig. 5B, pathway 2).

Wastewater Treatment.

In order to explore the potential of TPT-alkynyl-AQ for in situ river restoration, TPT-alkynyl-AQ was used for organic pollutant degradation under ambient conditions. The potential biological risk was investigated (SI Appendix, Fig. S34) (42). The model wastewater contained triclosan, 17α-ethinyl estradiol, and tetracycline. All of them were 10 mg⋅L−1 in the model wastewater. The wastewater was treated by dispersing 10 mg of TPT-alkynyl-AQ in 500 mL of the wastewater for 10 h under ambient conditions (xenon lamp, λ > 400 nm, 100 mW⋅cm−1, open air). Then, the treated water was mixed with the E3 nutrient solution. The normal hatching rate of zebrafish in the treated wastewater was 97%, while the normal hatching rate in nonpolluted E3 nutrient solution was 88%, and no zebrafish embryos were hatched in the pretreated wastewater after 3 d. The much higher hatching rate was attributed to the pollutant degradation and antibacterial properties of the in situ generated H2O2.

Discussion

In summary, we have reported a highly active and strong antiinterference D-B-A organic photocatalyst for H2O2 production in ambient conditions. The photosynthetic rates of H2O2 reach 9,366 and 12,324 µmol⋅g−1 during 9 AM to 4 PM in river water and seawater, respectively. To the best of our knowledge, the photosynthetic efficiency of H2O2 is the highest under ambient conditions to date, even higher than all the systems consuming pure O2. The outstanding efficiency of H2O2 production is owing to the spontaneous exciton dissociation at room temperature. Moreover, the rigid electron bridges and redox electron acceptors provide abundant active sites for both ORR and WOR to accelerate the overall rate of H2O2 production. The ultrahigh photocatalytic efficiency in ambient conditions allows SCC in a real cost-effective and sustainable way.

Materials and Methods

Synthesis of TPT-alkynyl-AQ.

Three milliliters of trifluoromethanesulfonic acid was charged into a predried flask under Ar atmosphere at 0 °C; 200 mg of 4-(2-{6-[2-(4-cyanophenyl)ethynyl]-9,10-dioxo-9,10dihydroanthracen-2-yl}ethynyl)benzonitrile in 20 mL of CHCl3 was added into the flask dropwise over 30 min and stirred for 2 h. Then the suspension was warmed up to 25 °C and stirred for another 2 h before being left overnight at 100 °C. The obtained solid was quenched in cold water and washed with excess diluted ammonia and water. After being vacuum dried at 100 °C, the finally obtained solid was ground into powder and subjected to ultrasonication for 36 h in water.

TPT-AQ was synthesized via the same procedure and the same conditions as TPT-alkynyl-AQ.

Synthesis of TPT-alkynyl-TPT

Two milliliters of trifluoromethanesulfonic acid was charged into a predried flask under Ar atmosphere at 0 °C; 200 mg of 4,4′-(ethyne-1,2-diyl)dibenzonitrile in 10 mL of CHCl3 was added into the flask dropwise over 30 min and stirred for 2 h. Then the suspension was warmed up to 25 °C and stirred for another 2 h before being left for 2 h at 100 °C. The obtained solid was quenched in cold water and washed with excess diluted ammonia, CHCl3, and water. After being vacuum dried at 100 °C, the finally obtained solid was ground into powder and subjected to ultrasonication for 36 h in water.

Synthesis of TPT-imine-AQ.

One hundred seventy-five milligrams (0.75 mmol) of 2,6-diamino-9,10-dihydroanthracene-9,10-dione, 197 mg (0.5 mmol) of 4-[bis(4-formylphenyl)-1,3,5-triazin-2-yl]benzaldehyde, N,N-dimethylformamide (DMF)/1,4-dioxane (6 mL, 2/1 by volume), and acetic acid solution (1 mL, 6 M) were charged into a predried flask. After being quickly frozen, the flask was sealed and degassed by three freeze–pump–thaw cycles. Subsequently, the reaction flask was slightly warmed to room temperature and then kept at 160 °C for 72 h. After being cooled to room temperature naturally, the brownish red precipitate was collected by filtration and washed by DMF, dichloromethane, and ethanol, sequentially. After being vacuum dried at 100 °C, the finally obtained solid was ground into powder and subjected to ultrasonication for 36 h in water.

Photocatalytic Experiments.

One milligram of photocatalyst was added to 50 mL of deionized water in a 100-mL beaker. The catalyst was dispersed by ultrasonication for 30 min and irradiated by using a Xe lamp (100 mW·cm2, λ > 400 nm) under magnetic stirring.

SCC Efficiency Measurements.

The SCC efficiency was determined by the photocatalytic experiments using an AM 1.5 G solar simulator as the light source (100 mW·cm−2). The photoreaction was performed in pure deionized water (100 mL) with photocatalyst (100 mg) in a glass bottle. The SCC efficiency (η) was calculated by following equation:

η(%)=ΔGH2O2×nH2O2tir×Sir×IAM×100%,

where ΔGH2O2 is the free energy for H2O2 generation (117 kJ·mol−1), nH2O2 is the amount of H2O2 generated, and tir is the irradiation time, 3,600 s. The overall irradiation intensity (IAM) of the AM1.5 global spectrum (300 nm to 2,500 nm) is 100 mW·cm−2, and the irradiation area (Sir) is 3.14 × 10−4 m2.

Apparent Quantum Efficiency Analysis.

The photocatalytic reaction was carried out in pure deionized water (50 mL) with photocatalyst (50 mg) in a quartz tube. The tube was irradiated by an Xe lamp for 1 h with magnetic stirring. The number of incident photons (M) is calculated by the following equation:

M=hc.

In the equation, E, λ, h, and c are the average intensity of irradiation, the wavelength of the irradiation, the Planck constant, and the speed of light, respectively. The quantum efficiency was calculated from the following equation:

AQY=2×number of evolved H2O2 moleculesnumber of incident photons ×100%.

Determination of H2O2 Concentration.

The concentration of H2O2 was quantified by a TMB-H2O2-HRP enzymatic assay, and the horseradish peroxidase (HRP) could instantaneously catalyze the reaction between H2O2 and TMB,

H2O2+TMBHRPH2O+oxTMB.

Preparation of 3,3′,5,5′-tetramethylbenzidine (TMB) solution was as follows: 0.015 g TMB was dissolved in 0.3 mL of DMSO, followed by adding 5 mL of glycerol and 45 mL of deionized water containing 0.02 g of ethylenediaminetetraacetic acid and 0.095 g of citric acid. Then the solution was filled to 500 mL with deionized water.

Preparation of HRP solution was as follows: 0.002 g of peroxidase (from horseradish) was dissolved in 10 mL of deionized water.

Determination of the calibration curve was as follows: TMB and HRP were added into the H2O2 solution with a known concentration. Note that HRP concentration should be kept at 12.5 µg·mL−1. After 3 min, 10 µL of HCl was added and measured by UV-visible spectroscopy at 450 nm. Based on the liner relationship between signal intensity and H2O2 concentration, the H2O2 concentration of the samples could be obtained.

Verification of the ORR Sites at the Alkynyl Moieties.

Photocatalyst (1 mg) and disodium ethylenediaminetetraacetate (2 mM) were added to 50 mL of deionized water in a 100-mL beaker. The catalyst was dispersed by ultrasonication for 30 min. The suspension was bubbled with O2 for 30 min to make alkynyl moieties absorb oxygen and was irradiated under Ar atmosphere for 1 h.

Verification of the ORR Sites at the AQ Moieties.

Ten milligrams of photocatalyst was added to 20 mL of deionized water in a beaker. The catalyst was dispersed by ultrasonication for 30 min. The suspension was first irradiated under Ar atmosphere for 1 h and then injected with O2 in the dark for 30 min.

TAS Measurements.

TA spectra of TPT-alkynyl-AQ, TPT-AQ, and TPT-alkynyl-TPT were measured on a Helios femtosecond transient absorption spectrometer (Ultrafast Systems, LLC). A 375-nm pump pulse was generated via an optical parametric amplifier (OPerA Solo, Coherent). The 375-nm laser intensity was 70 µW. The sample preparation was as follows: 20 mg of photocatalyst was dispersed in 3 mL of deionized water. Then the suspension was subjected to ultrasonication for 36 h. The initial suspension was centrifuged at 3,000 rpm for 10 min to remove large aggregates. The suspension was bubbled with oxygen for 20 min before atmosphere measurements.

Computational Methods.

The DFT calculation was carried out as implemented in the Gaussian 09 D.01 program package (43), using Grimme-D3 dispersion correction (44). GaussView6 was applied to visualization (45). Geometry optimization and frequency analysis were calculated at the PBE0/6-31G(g,d) level of theory (46). The optimized structures were used for single-point energy calculations with PBE0/ma-TZVP (4648). Time-dependent density functional theory (TD-DFT) was calculated at CAM-B3LYP/6-31G(d,p) level of theory and was used to investigate the transfer direction of electrons (4951). Multiwfn was used for hole–electron analysis (52, 53). The Gibbs free energy change (ΔG) during water oxidation on the surface conjugated polymers was calculated based on a computational hydrogen model (54).

Zebrafish Embryos Experiments.

Two fish culture media were prepared by adding the synthetic polluted water and the treated effluent to the standard E3 medium at a volume ratio of 1:2. The pure standard E3 medium was used as a blank control, and every group had 96 zebrafish zygotes for experiments. E3 medium includes 5 mM NaCl, 0.17 mM KCl, 0.33 mM CaCl2, and 0.33 mM MgSO4. We used ethinylestradiol, triclosan, and tetracycline as the model water pollutants. The synthetic polluted water was obtained by adding 5 mg of ethinylestradiol, 5 mg of triclosan, and 5 mg of tetracycline to 100 mL of deionized water, which was sonicated for 6 h, filtered by water filtration membrane, and filled to 500 mL with deionized water. Ten milligrams of TPT-alkynyl-AQ catalyst was added into 100 mL of solution containing 5 mg of ethinylestradiol, 5 mg of triclosan, and 5 mg of tetracycline for sonication for 6 h; then the above solution was irradiated for 10 h, filtered by water filtration membrane, and filled to 500 mL with deionized water, and named the treated effluent. Zebrafish zygotes were obtained from wild-type adults and cultured in standard E3 medium for 7 d. Dead and unfertilized eggs were discarded at 4 h postfertilization. The fertilized embryos were transferred into 48-well plates and filled with corresponding medium. Embryo development was monitored, and representative images were captured.

Supplementary Material

Supplementary File

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grants 22036003, 22076222, and 21737006), Guangdong Provincial Key R&D Programme (Grant 2020B1111350002), the National Science Foundation of Guangdong Province (Grant 2020A1515011442), Guangdong-Hongkong Joint Laboratory for Water Security (Grant 2020B1212030005), and the China Postdoctoral Science Foundation (Grant 2021M703677).

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.2202913119/-/DCSupplemental.

Data Availability

All study data are included in the article and/or SI Appendix.

References

  • 1.Fang X., Kalathil S., Reisner E., Semi-biological approaches to solar-to-chemical conversion. Chem. Soc. Rev. 49, 4926–4952 (2020). [DOI] [PubMed] [Google Scholar]
  • 2.Banerjee T., Podjaski F., Kröger J., Biswal B. P., Lotsch B. V., Polymer photocatalysts for solar-to-chemical energy conversion. Nat. Rev. Mater. 6, 168–190 (2021). [Google Scholar]
  • 3.Volokh M., Shalom M., Light on peroxide. Nat. Catal. 4, 350–351 (2021). [Google Scholar]
  • 4.Teng Z., et al. , Atomically dispersed antimony on carbon nitride for the artificial photosynthesis of hydrogen peroxide. Nat. Catal. 4, 374–384 (2021). [Google Scholar]
  • 5.Ye Y.-X., et al. , A solar-to-chemical conversion efficiency up to 0.26% achieved in ambient conditions. Proc. Natl. Acad. Sci. U.S.A. 118, e2115666118 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Xia C., Xia Y., Zhu P., Fan L., Wang H., Direct electrosynthesis of pure aqueous H2O2 solutions up to 20% by weight using a solid electrolyte. Science 366, 226–231 (2019). [DOI] [PubMed] [Google Scholar]
  • 7.Chen L., et al. , Acetylene and diacetylene functionalized covalent triazine frameworks as metal-free photocatalysts for hydrogen peroxide production: A new two-electron water oxidation pathway. Adv. Mater. 32, e1904433 (2020). [DOI] [PubMed] [Google Scholar]
  • 8.Proppe A. H., et al. , Bioinspiration in light harvesting and catalysis. Nat. Rev. Mater. 5, 828–846 (2020). [Google Scholar]
  • 9.Madhu M., Ramakrishnan R., Vijay V., Hariharan M., Free charge carriers in homo-sorted π-stacks of donor-acceptor conjugates. Chem. Rev. 121, 8234–8284 (2021). [DOI] [PubMed] [Google Scholar]
  • 10.Clarke T. M., Durrant J. R., Charge photogeneration in organic solar cells. Chem. Rev. 110, 6736–6767 (2010). [DOI] [PubMed] [Google Scholar]
  • 11.Yang L., et al. , Beyond C3N4 π-conjugated metal-free polymeric semiconductors for photocatalytic chemical transformations. Chem. Soc. Rev. 50, 2147–2172 (2021). [DOI] [PubMed] [Google Scholar]
  • 12.Gillett A. J., et al. , The role of charge recombination to triplet excitons in organic solar cells. Nature 597, 666–671 (2021). [DOI] [PubMed] [Google Scholar]
  • 13.Qian D., et al. , Design rules for minimizing voltage losses in high-efficiency organic solar cells. Nat. Mater. 17, 703–709 (2018). [DOI] [PubMed] [Google Scholar]
  • 14.Lan Z.-A., et al. , A fully coplanar donor-acceptor polymeric semiconductor with promoted charge separation kinetics for photochemistry. Angew. Chem. Int. Ed. Engl. 60, 16355–16359 (2021). [DOI] [PubMed] [Google Scholar]
  • 15.Lan Z.-A., et al. , Reducing the exciton binding energy of donor-acceptor-based conjugated polymers to promote charge-induced reactions. Angew. Chem. Int. Ed. Engl. 58, 10236–10240 (2019). [DOI] [PubMed] [Google Scholar]
  • 16.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, e1903545 (2019). [DOI] [PubMed] [Google Scholar]
  • 17.Kosco J., et al. , Enhanced photocatalytic hydrogen evolution from organic semiconductor heterojunction nanoparticles. Nat. Mater. 19, 559–565 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Kiefer D., et al. , Double doping of conjugated polymers with monomer molecular dopants. Nat. Mater. 18, 149–155 (2019). [DOI] [PubMed] [Google Scholar]
  • 19.Shiraishi Y., et al. , Resorcinol-formaldehyde resins as metal-free semiconductor photocatalysts for solar-to-hydrogen peroxide energy conversion. Nat. Mater. 18, 985–993 (2019). [DOI] [PubMed] [Google Scholar]
  • 20.Wang S., et al. , Intermolecular cascaded π-conjugation channels for electron delivery powering CO2 photoreduction. Nat. Commun. 11, 1149 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Yang Z., et al. , Transformation strategy for highly crystalline covalent triazine frameworks: From staggered AB to eclipsed AA stacking. J. Am. Chem. Soc. 142, 6856–6860 (2020). [DOI] [PubMed] [Google Scholar]
  • 22.Liu J., et al. , Solution synthesis of semiconducting two-dimensional polymer via trimerization of carbonitrile. J. Am. Chem. Soc. 139, 11666–11669 (2017). [DOI] [PubMed] [Google Scholar]
  • 23.Fang Z.-B., et al. , Boosting interfacial charge-transfer kinetics for efficient overall CO2 photoreduction via rational design of coordination spheres on metal-organic frameworks. J. Am. Chem. Soc. 142, 12515–12523 (2020). [DOI] [PubMed] [Google Scholar]
  • 24.Xia P., et al. , Designing a 0D/2D S-scheme heterojunction over polymeric carbon nitride for visible-light photocatalytic inactivation of bacteria. Angew. Chem. Int. Ed. Engl. 59, 5218–5225 (2020). [DOI] [PubMed] [Google Scholar]
  • 25.Hou H., Zeng X., Zhang X., Production of hydrogen peroxide through photocatalytic processes: A critical review of recent advances. Angew. Chem. Int. Ed. 59, 17356–17376 (2020). [DOI] [PubMed] [Google Scholar]
  • 26.Shiraishi Y., Matsumoto M., Ichikawa S., Tanaka S., Hirai T., Polythiophene-doped resorcinol-formaldehyde resin photocatalysts for solar-to-hydrogen peroxide energy conversion. J. Am. Chem. Soc. 143, 12590–12599 (2021). [DOI] [PubMed] [Google Scholar]
  • 27.Shiraishi Y., et al. , Sunlight-driven hydrogen peroxide production from water and molecular oxygen by metal-free photocatalysts. Angew. Chem. Int. Ed. Engl. 53, 13454–13459 (2014). [DOI] [PubMed] [Google Scholar]
  • 28.Cai J., et al. , Crafting mussel-inspired metal nanoparticle-decorated ultrathin graphitic carbon nitride for the degradation of chemical pollutants and production of chemical resources. Adv. Mater. 31, e1806314 (2019). [DOI] [PubMed] [Google Scholar]
  • 29.Chu C., et al. , Spatially separating redox centers on 2D carbon nitride with cobalt single atom for photocatalytic H2O2 production. Proc. Natl. Acad. Sci. U.S.A. 117, 6376–6382 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Zeng X., et al. , Simultaneously tuning charge separation and oxygen reduction pathway on graphitic carbon nitride by polyethylenimine for boosted photocatalytic hydrogen peroxide production. ACS Catal. 10, 3697–3706 (2020). [Google Scholar]
  • 31.Kofuji Y., et al. , Carbon nitride-aromatic diimide-graphene nanohybrids: Metal-free photocatalysts for solar-to-hydrogen peroxide energy conversion with 0.2% efficiency. J. Am. Chem. Soc. 138, 10019–10025 (2016). [DOI] [PubMed] [Google Scholar]
  • 32.Hirakawa H., Hashimoto M., Shiraishi Y., Hirai T., Photocatalytic conversion of nitrogen to ammonia with water on surface oxygen vacancies of titanium dioxide. J. Am. Chem. Soc. 139, 10929–10936 (2017). [DOI] [PubMed] [Google Scholar]
  • 33.Xiao C., Hu H., Zhang X., MacFarlane D. R., Nanostructured gold/bismutite hybrid heterocatalysts for plasmon-enhanced photosynthesis of ammonia. ACS Sustain. Chem.& Eng. 5, 10858–10863 (2017). [Google Scholar]
  • 34.Shiraishi Y., et al. , Nitrogen fixation with water on carbon-nitride-based metal-free photocatalysts with 0.1% solar-to-ammonia energy conversion efficiency. ACS Appl. Energy Mater. 1, 4169–4177 (2018). [Google Scholar]
  • 35.Zhang N., et al. , Refining defect states in W18O49 by Mo doping: A strategy for tuning N2 activation towards solar-driven nitrogen fixation. J. Am. Chem. Soc. 140, 9434–9443 (2018). [DOI] [PubMed] [Google Scholar]
  • 36.Li X.-H., et al. , Reduced state of the graphene oxide@polyoxometalate nanocatalyst achieving high-efficiency nitrogen fixation under light driving conditions. ACS Appl. Mater. Interfaces 11, 37927–37938 (2019). [DOI] [PubMed] [Google Scholar]
  • 37.Zhu L., et al. , Small exciton binding energies enabling direct charge photogeneration towards low-driving-force organic solar cells. Angew. Chem. Int. Ed. Engl. 60, 15348–15353 (2021). [DOI] [PubMed] [Google Scholar]
  • 38.Li X., et al. , CsPbX3 quantum dots for lighting and displays: Room-temperature synthesis, photoluminescence superiorities, underlying origins and white light-emitting diodes. Adv. Funct. Mater. 26, 2435–2445 (2016). [Google Scholar]
  • 39.Wang Y., Wang X., Antonietti M., Polymeric graphitic carbon nitride as a heterogeneous organocatalyst: From photochemistry to multipurpose catalysis to sustainable chemistry. Angew. Chem. Int. Ed. Engl. 51, 68–89 (2012). [DOI] [PubMed] [Google Scholar]
  • 40.Chen D., et al. , Efficient visible-light-driven hydrogen evolution and Cr(VI) reduction over porous P and Mo co-doped g-C3N4 with feeble N vacancies photocatalyst. J. Hazard. Mater. 361, 294–304 (2019). [DOI] [PubMed] [Google Scholar]
  • 41.Gao J., et al. , Enabling direct H2O2 production in acidic media through rational design of transition metal single atom catalyst. Chem 6, 658–674 (2020). [Google Scholar]
  • 42.Xu J., et al. , Organic wastewater treatment by a single-atom catalyst and electrolytically produced H2O2. Nat. Sustain. 4, 233–241 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Frisch M. J., et al. , Gaissoan 09 (Gaussian, Inc, Wallingford, CT, 2009). [Google Scholar]
  • 44.Grimme S., Antony J., Ehrlich S., Krieg H., A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 132, 154104 (2010). [DOI] [PubMed] [Google Scholar]
  • 45.Dennington R., et al. , GaussView (Version 6.1.1, Shawnee, Mission, KS, 2019).
  • 46.Adamo C., Barone V., Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 110, 6158–6170 (1999). [Google Scholar]
  • 47.Papajak E., Zheng J., Xu X., Leverentz H. R., Truhlar D. G., Perspectives on basis sets beautiful: Seasonal plantings of diffuse basis functions. J. Chem. Theory Comput. 7, 3027–3034 (2011). [DOI] [PubMed] [Google Scholar]
  • 48.Zheng J., Xu X., Truhlar D. G., Minimally augmented Karlsruhe basis sets. Theor. Chem. Acc. 128, 295–305 (2011). [Google Scholar]
  • 49.Yanai T., Tew D. P., Handy N. C., A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 393, 51–57 (2004). [Google Scholar]
  • 50.Francl M. M., et al. , Self‐consistent molecular orbital methods. XXIII. A polarization‐type basis set for second‐row elements. J. Chem. Phys. 77, 3654–3665 (1982). [Google Scholar]
  • 51.Ditchfield R., Hehre W. J., Pople J. A., Self‐consistent molecular‐orbital methods. IX. An extended Gaussian‐type basis for molecular‐orbital studies of organic molecules. J. Chem. Phys. 54, 724–728 (1971). [Google Scholar]
  • 52.Lu T., Chen F., Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 33, 580–592 (2012). [DOI] [PubMed] [Google Scholar]
  • 53.Tian L. U., Chen F.-W., Comparison of computational methods for atomic charges. Acta Physico Chim. Sin. 28, 1–18 (2012). [Google Scholar]
  • 54.Nørskov J. K., et al. , Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J. Phys. Chem. B 108, 17886–17892 (2004). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File

Data Availability Statement

All study data are included in the article and/or SI Appendix.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES