Abstract
Photocatalytic water disinfection technology is highly promising in off-grid areas due to abundant year-round solar irradiance. However, the practical use of powdered photocatalysts is impeded by limited recovery and inefficient inactivation of stress-resistant bacteria in oligotrophic surface water. Here we prepare a floatable monolithic photocatalyst with ZIF-8-NH2 loaded Ag single atoms and nanoparticles (AgSA+NP/ZIF). Atomically dispersed Ag sites form an Ag−N charge bridge, extending the lifetime of charge carriers and thereby promoting reactive oxygen species (ROS) generation. The photothermal effect of the plasmonic Ag nanoparticles reduces the bacterial resistance to ROS and impairs DNA repair capabilities. Under sunlight irradiation, the synergistic effect of Ag single atoms and nanoparticles enables 4.0 cm2 AgSA+NP/ZIF to achieve over 6.0 log inactivation (99.9999%) for the stress-resistant Escherichia coli (E. coli) in oligotrophic surface water within 30 min. Furthermore, 36 cm2 AgSA+NP/ZIF is capable of disinfecting at least 10.0 L of surface water, which meets the World Health Organization (WHO) recommended daily per capita drinking water allocation (8.0 L). This study presents a decentralized and sustainable approach for water disinfection in off-grid areas.
Similar content being viewed by others
Introduction
Water pollution caused by pathogenic microorganisms poses a serious threat to public health. Millions of individuals in developing nations succumb to mortality each year due to the consumption of water contaminated with pathogenic microorganisms1,2. Every 2 min a child under the age of five perishes from diarrheal diseases caused by poor sanitation and substandard water quality3,4. Since the 20th century, the implementation of urban water supply network systems and centralized water treatment facilities has significantly mitigated infections caused by waterborne pathogens. However, effective access to safe drinking water remains scarce in off-grid rural areas, such as in sub-Saharan Africa, North Africa, and the Middle East5, often due to lack of electricity and other vital water treatment infrastructures to ensure reliable and safe drinking water access. With the intensification of climate change, it is estimated that 3.6 billion people worldwide currently live in regions that may face potential water scarcity, and this number could rise to 5.7 billion by 20506. Therefore, improving water management and sanitation is critical for communities that have no or limited access to traditional water disinfection methods.
Point-of-use (POU) disinfection at household or community scale is one of the most promising solutions in addressing this issue5,7. POU disinfection technologies eliminate the need for expensive large-scale water treatment equipment and avoid the complexities associated with transportation and storage procedures. Despite decades of technological advances, existing POU disinfection technologies still face issues with high energy consumption (e.g., boiling water), generation of carcinogenic by-products (e.g., chlorination), and high costs (e.g., reverse osmosis)8,9,10. Fortunately, among the 64 countries with <80% rural access to basic drinking water services, solar irradiance (5.59 kWh m−2 per day) significantly exceeds the global average (4.70 kWh m−2 per day)5, offering valuable opportunities for solar energy-driven water treatment. Therefore, the development of advanced materials and processes that can fully harness solar energy to inactivate microorganisms will contribute to the attainment of environmentally friendly and sustainable water disinfection in off-grid communities.
Single-atom photocatalysts (SAPCs) have been reported for disinfection due to their maximum metal atom utilization and efficient charge separation11,12,13. SAPCs are known to generate ROS that attack cell membranes and intracellular components like DNA14. However, microorganisms possess innate damage repair system, such as DNA repair process (SOS response), which confer resistance to ROS stresses15. Such resistance is more pronounced in microorganisms at late stationary phase in natural waters16,17,18. In oligotrophic surface water, bacteria at this growth phase are usually starved, and their adaptation to starvation enhances their tolerance to ROS and other environmental stresses19. This is why these bacteria cultured under above specific conditions are usually called stress-resistant bacteria. Consequently, numerous photocatalysts exhibiting exceptional antimicrobial performance in laboratory show limited bactericidal effectiveness in real water environments, posing challenges for their large-scale applications. Nanoparticles with plasmonic effects, such as Ag and Au, exhibit excellent photothermal properties through collective resonance of free charge carriers and non-radiative decay processes20,21. Previous studies have shown that photothermal ablation can target tumor cell proteins and promote protein aggregation, consequently interfering with DNA damage repair22. Given that the bacterial inactivation by SAPCs relies on ROS-mediated intracellular DNA damage, any potential reduction in bacterial DNA repair capacity due to the photothermal effect would significantly enhance the bactericidal efficiency. Hence, we speculate that SAPCs integrated with plasmonic nanoparticles could effectively inactivate stress-resistant microorganisms in water.
In the synthesis of SAPCs, the prevalent use of powdered supports such as carbon nitride and metal oxide materials to anchor single atoms may interfere their recovery in aqueous reactions23,24,25, thereby increasing the contamination risk throughout the disinfection process16. In addition, the widespread presence of suspended sediment and natural organic matter in the natural water significantly interferes with the light absorbance of SAPCs5. To improve catalyst recovery and light absorption, previous studies have employed polymer binders to immobilize catalysts on the platforms such as nickel foam and ceramic plate26,27, resulting in compromised accessibility of the potential active reaction sites and hindered electron transfer28,29,30. Consequently, there is an urgent need to avoid using polymer binders while strengthening the stability of the catalytic layer on the support. This requirement can be effectively fulfilled by the fabrication of binder-free floatable monolithic photocatalysts via loading metal cocatalysts into the floatable monolithic supports.
Here, we prepared a silver single-atom and nanoparticle-loaded floatable monolithic photocatalyst (AgSA+NP/ZIF) via a two-step synthesis method as shown in Fig. 1a. Firstly, a monolithic metal-organic frameworks ZIF-8-NH2 layer was synthesized in situ by a ligand evaporation deposition strategy, employing 3D melamine sponges with zinc ions as a substrate, and utilizing 2-methylimidazole and 2-aminobenzimidazole as organic ligands. ZIF-8-NH2 material possesses semiconducting properties and is frequently employed as support for loading metal cocatalysts31,32. Abundant N atoms with lone pair of electrons in ZIF-8-NH2 can form strong coordination with Ag metal atoms, effectively preventing their migration and aggregation23,33. The low density of sponges ensures that the monolithic photocatalyst can float on water. The monolithic photocatalyst was selectively loaded with AgSA and AgNP using the ice-photoreduction method, wherein the photocatalytic reaction primarily occurs on the submerged surface with sunlight exposure (Fig. 1b). AgSA+NP/ZIF can be utilized in a tablet-like form for convenient recyclable water disinfection. Under natural light irradiation, the photothermal effect of AgSA+NP/ZIF synergizes with photogenerated ROS to inactivate stress-resistant bacteria in oligotrophic surface water. This monolithic photocatalyst exhibits practical advantages for POU water disinfection, as evidenced by its performance, recyclability, durability, and cost-effectiveness validated in this study.
Results
Material characterization
The photocatalyst support consists of a 3D melamine sponge and ZIF-8-NH2 layer. As shown in Fig. 2a, a dense and continuous ZIF-8-NH2 layer was deposited on the surface of the melamine sponge. The thickness of ZIF-8-NH2 coating is 1.0-1.4 μm. Cross-sectional scanning electron microscope (SEM) image (Supplementary Fig. 1) shows that some ZIF-8-NH2 crystals were infused into the inner sponge layer, which potentially enhances the adhesion between the surface ZIF-8-NH2 layer and substrate. X-ray diffraction (XRD) patterns verified that the surface layer exhibits intact ZIF-8-NH2 crystal structure (Supplementary Fig. 2)34. Transmission electron microscopy (TEM) images show successful loading of Ag nanoparticles with an average size of 13.1 nm (Fig. 2b and Supplementary Fig. 3).
a Top (i) and (ii) cross-sectional view SEM images of monolithic ZIF-8-NH2 support. b TEM image of AgSA+NP/ZIF. Inset is the statistical size distribution of Ag nanoparticles. c AC-HAAD-STEM image of AgSA+NP/ZIF (the bright dots in the red circles are Ag single atoms). Inset is the HAAD-STEM image of AgSA+NP/ZIF. d HAADF-STEM with elemental mapping images of AgSA+NP/ZIF. e Fourier-transformed EXAFS results of AgSA/ZIF, AgNP/ZIF, AgSA+NP/ZIF, Ag foil and Ag2O at the Ag K-edge. f EXAFS fitting curve of AgSA+NP/ZIF in R space.
Aberration-corrected high-angle-annular-dark-field scanning transmission electron microscopy (AC-HAADF-STEM) images (Fig. 2c and Supplementary Fig. 4) show atomically dispersed Ag (circled bright spots). The elemental mapping images also confirm that Ag atoms and nanoparticles dispersed on ZIF-8-NH2 (Fig. 2d). Thus, Ag single atoms and nanoparticles were both present on this monolithic photocatalyst. The total Ag loading content was 1.28 wt% as detected by an inductively coupled plasma optical emission spectrometer (ICP-OES), where the proportion of Ag single atoms and Ag nanoparticles was 46.9% and 53.1%, respectively. In addition, the 1.21 wt% Ag single-atom-loaded ZIF-8-NH2 (AgSA/ZIF) and 1.16 wt% Ag nanoparticle-loaded ZIF-8-NH2 (AgNP/ZIF) were also prepared for comparative studies and their AC-HAADF-STEM and TEM images are shown in Supplementary Fig. 5.
A measurement of X-ray absorption fine structure (XAFS) was carried out to further investigate the dispersion and coordination environment of Ag species35. Fourier-transformed (FT) extended Ag K-edge X-ray absorption fine structure (EXAFS) spectra show that AgSA/ZIF and AgSA+NP/ZIF share a common peak at 1.56 Å, which may be attributed to the coordination of Ag single atoms with N or C (Fig. 2e). The best-fitting result for EXAFS spectrum of AgSA+NP/ZIF shows that the dominant peak at 1.56 Å can be attributed to the Ag−N coordination (Fig. 2f and Supplementary Table 1). The Ag−N and Ag−Ag coordination numbers are 1.5 and 4.6 at bond distances of 2.02 Å and 2.83 Å, respectively. In addition, a peak at 2.7 Å was observed in both AgSA+NP/ZIF and AgNP/ZIF, resembling the characteristic Ag−Ag peak of metallic Ag, thereby confirming the presence of Ag nanoparticles in these samples.
Optical and photothermal properties
Light capture and photogenerated carrier separation play key roles in photocatalytic disinfection reactions16,36. As shown in Fig. 3a, compared to ZIF-8-NH2, AgSA+NP/ZIF exhibits the strongest light absorption in the visible region, which may result from the introduction of Ag single atoms and nanoparticles37,38. The light absorption intensity of AgNP/ZIF is higher than that of AgSA/ZIF, indicating a more significant contribution from the plasmonic Ag nanoparticles to light adsorption capacity. The noise at 300-380 nm originates from the sponge (Supplementary Fig. 6a). The plasmonic peak of Ag nanoparticles is absent at 450 nm, probably due to the low Ag loading (1.16 wt% for AgNP/ZIF) and the overlaps of the significant optical absorption of ZIF-8-NH2 with the weak plasmonic peak39.
a UV-Vis diffuse reflectance spectra of photocatalysts. b Electromagnetic simulations: enhancement factor of electric field on the Ag nanoparticle-ZIF-8-NH2 surface under visible light (450 nm) irradiation. c, d Steady-state fluorescence spectra (c) and TRPL decay curves (d) of photocatalysts, the excitation wavelength is 375 nm. e Electrochemical impedance spectra of photocatalysts, inset shows the corresponding equivalent circuit diagram. Rs: Solution resistance; CPE: Constant phase element. f Infrared thermal images of photocatalysts under simulated visible light (400-760 nm) irradiation at an intensity of 100 mW cm−2 for 15 min. g Temperature elevation curves of photocatalysts under simulated visible light (400-760 nm) irradiation at an intensity of 100 mW cm−2. h Temperature elevation curves of water after adding the monolithic photocatalyst under natural light irradiation. Reaction condition: 2 × 2 cm2 catalyst; 0.1 L Yellow River water from upper reaches (Longyang Gorge); 33.0-35.0 mW cm−2 sunlight intensity.
The increased light absorption by Ag nanoparticles can be attributed to the local electromagnetic enhancement resulting from the localized surface plasmon resonance40. Noble metal (e.g., gold and silver) nanoparticles contain large numbers of freely mobile electrons that can strongly interact with light by either absorbing or scattering photons20,21. The oscillating electric field of the incoming radiation induces coherent collective oscillation of the free electrons on the metal surface20,21. Supplementary Fig. 6b shows the simulation of the electromagnetic field at the interface between one single Ag nanoparticle (e.g., 13 nm in diameter) and ZIF-8-NH2 under 450 nm irradiation using COMSOL Multiphysics. The average enhancement factor, defined as P = | E | /E0, quantifies the amplification of the localized electric field (E) after light irradiation relative to the electric field (E0) before light irradiation in the near-field region of the Ag nanoparticle. Clearly, along the distance from the core of the Ag nanoparticle, the local electromagnetic field yielded a P-value of >10-fold due to the dynamic charge accumulation at the Ag nanoparticle surface under a 450 nm planewave excitation. In addition, the effects of ZIF-8-NH2 thickness and Ag nanoparticle size on the electromagnetic field enhancement are evident. The results presented in Fig. 3b demonstrate a proportional increase in the enhancement factor as the radius of Ag nanoparticles ranges from 1.0 to 8.0 nm. It is observed that the thickness of ZIF-8-NH2 does not significantly impact the enhancement factor when exceeding 4.0 nm. Furthermore, there is no enhancement of the electric field intensity within the near field of pure ZIF-8-NH2 (Supplementary Fig. 6c), suggesting that the interface between ZIF-8-NH2 substrate and Ag nanoparticle is essential for enhancing the light absorbance of the photocatalyst22.
Furthermore, the band gap of AgSA+NP/ZIF was determined to be 2.78 eV, which is smaller than that of pure ZIF-8-NH2 (3.47 eV) as compared by the Tauc plots in Supplementary Fig. 7a, b. The strong local electric field generated by Ag nanoparticles or the formation of Ag−N coordination might influence the electronic density distribution in ZIF-8-NH2, consequently modulating its band gap41,42. Mott-Schottky plots of AgSA+NP/ZIF were employed to determine the conduction band minimum (CBM) (Supplementary Fig. 7c). Higher CBM (−0.43 V vs. NHE) of AgSA+NP/ZIF than reduction potential of O2 to O2•−(−0.33 V vs. NHE) implies the feasibility of AgSA+NP/ZIF for ROS generation43.
The concentration of ROS production depends on the separation of photogenerated carriers, which was investigated by steady-state fluorescence spectroscopy and time-resolved photoluminescence (TRPL) spectroscopy. As shown in Fig. 3c, both AgSA+NP/ZIF and AgSA/ZIF exhibited lower fluorescence intensities compared to AgNP/ZIF, indicating that the recombination of electron-hole pairs was inhibited. This is because the atomically dispersed Ag provides more reactive sites to capture the photogenerated electrons44. As shown in Fig. 3d and Supplementary Table 2, the average lifetime of photogenerated charge carriers in AgSA/ZIF (1.67 ns) exceeds those in AgSA+NP/ZIF (1.53 ns) and AgNP/ZIF (1.10 ns). The Ag content of AgSA/ZIF is 1.21 wt%, which is lower than that of AgSA+NP/ZIF (1.28 wt%). Nonetheless, AgSA/ZIF exhibits a longer charge carrier lifetime, indicating that AgSA is more conducive to the capture of photogenerated electrons. Moreover, AgSA+NP/ZIF and AgSA/ZIF show smaller impedance arc radii in electrochemical impedance spectra compared to AgNP/ZIF (Fig. 3e). The charge transfer resistance (Rct) of AgSA/ZIF is 25.5 KΩ, which is lower than that of AgSA+NP/ZIF (39.0 KΩ). The atomically dispersed Ag atoms bonded with N atoms facilitates the formation of Ag−N charge bridge. This configuration significantly enhances the metal-support interaction, thereby boosting the charge transfer39. Hence, we conclude that the incorporation of Ag single atoms endows monolithic ZIF-8-NH2 with intensive photochemical properties.
The photothermal properties of the dry monolithic photocatalysts were investigated by infrared thermal imaging. After exposure to simulated visible light (400-760 nm) at an intensity of 100 mW cm−2, the dry surface temperatures of AgSA+NP/ZIF and AgNP/ZIF rapidly increase to 67.1 and 72 °C within 15 min, respectively, followed by a plateau (Fig. 3f, g). In comparison, monolithic ZIF-8-NH2 and AgSA/ZIF demonstrated weak photothermal effects (<36 °C) after 60 min light irradiation. Both AgSA+NP/ZIF and AgNP/ZIF exhibit greater photothermal effects due to surface plasmon resonance of Ag nanoparticles22. We further evaluated the water temperature elevation induced by monolithic photocatalysts under natural light irradiation of 33.0-35.0 mW cm−2, which is approximately equivalent to the average daily irradiance (5.59 kWh m−2 per day) for most off-grid areas5. As shown in Fig. 3h, the utilization of AgSA+NP/ZIF resulted in a significant elevation of water temperature to 41.8 °C within 15 min light irradiation, whereas the ZIF-8-NH2 and AgSA/ZIF-treated group exhibited only a minor increase in water temperature (<36.7 °C). These results indicate that AgSA+NP/ZIF possesses pronounced photothermal heating properties.
Natural water disinfection performance
We investigated the water disinfection performance of AgSA+NP/ZIF toward stress-resistant bacteria E. coli at late stationary phase cultivated in the river water from the upper reaches of Yellow River. The chemical parameters of the water are shown in Supplementary Table 3, wherein the total organic carbon (TOC) is 5.49 mg L−1 and primarily originates from natural organic matter that exhibits light absorption properties. However, unlike water-dispersed particulate catalysts, AgSA+NP/ZIF floating at the water surface (Fig. 4a) remains unaffected by impurities in the water, thereby ensuring efficient light delivery.
a Disinfection operation schematic of AgSA+NP/ZIF. b Bactericidal efficiencies of 1, 2 and 4 cm2 AgSA+NP/ZIF after 30 min solar irradiation in 0.1 L Yellow River water from upper reaches (Longyang Gorge). c Photocatalytic bactericidal performance of 1 cm2 AgSA+NP/ZIF against E. coli in 0.1 L Yellow River water from upper reaches (Longyang Gorge) and laboratory-derived E. coli under solar irradiation. The lines represent time curves. d Time profiles of E. coli inactivation by 39.8 μg L−1 Ag+ and 207.2 μg L−1 Zn2+. e Time profiles of E. coli inactivation in 0.1 L Yellow River water from upper reaches (Longyang Gorge) by 4 cm2 AgSA+NP/ZIF, AgSA/ZIF, AgNP/ZIF and ZIF-8-NH2 photocatalysts. f Time profiles of S. aureus and virus MS2 inactivation in 0.1 L Yellow River water from upper reaches (Longyang Gorge) by 4 cm2 AgSA+NP/ZIF photocatalyst. g Time profiles of E. coli inactivation in 0.1 L Yellow River water from middle reaches (Toudaoguai Station) and lake water (Shichahai) by 4 cm2 AgSA+NP/ZIF photocatalysts. h E. coli inactivation efficiency in 1.0 L Yellow River water from upper reaches (Longyang Gorge) by four 3×3 cm2 AgSA+NP/ZIF, corresponding to a total area of 36 cm2. The line represents the time curve. i Disinfection efficiency of AgSA+NP/ZIF during recycling process. Experiments in b-i were conducted in triplicate, and the error bars represent the arithmetic mean ± standard deviation.
Firstly, we assessed the disinfection activity of different areas of AgSA+NP/ZIF against the stress-resistant E. coli in 0.1 L river water under 30 min of 33.0-35.0 mW cm−2 solar irradiation (Fig. 4b). Four cm2 AgSA+NP/ZIF achieved a 6.50 log10 reduction of E. coli, implying almost complete inactivation of the bacteria. Under a dark condition, a mere 0.44 log10 reduction was detected, indicating that the inactivation of bacteria was light-driven, thus ruling out adsorptive removal of bacteria by AgSA+NP/ZIF. The bactericidal activity of the 1 cm2 and 2 cm2 AgSA+NP/ZIF exhibited a reduction of at least 4 log10 compared to the 4 cm2 AgSA+NP/ZIF. This decrease can be attributed not only to the lower photocatalyst dosage but also to the inability of the smaller samples to raise the water temperature above 37 °C (Supplementary Fig. 8). In contrast, 4 cm2 AgSA+NP/ZIF effectively increased the water temperature to over 41 °C (Fig. 3h), surpassing the optimal growth temperature for E. coli at 37 °C.
In addition, we found that the bactericidal efficiency of 1 cm2 AgSA+NP/ZIF was enhanced by ~4.0 log10 after incubating E. coli to the logarithmic phase with saline (Fig. 4c), specifically referring to E. coli cultured under laboratory conditions without any stress resistance. The saline solution was prepared using the same river water, thus the different bactericidal activity toward stress-resistant bacteria and laboratory-cultured bacteria was not caused by other factors in the river water. This result confirms the stress resistance of E. coli at late stationary phase and highlights the necessity and precision of employing resistant bacteria to study the bactericidal activity of catalysts in actual aqueous environments. After 30 min of light exposure, 4 cm2 AgSA+NP/ZIF dissolved Ag+ and Zn2+ in 0.1 L river water, which were detected by an inductively coupled plasma mass spectrometry (ICP-MS) as 39.8 μg L−1 and 207.2 μg L−1, respectively. These released metal ions demonstrated limited inhibitory efficiency against E. coli (Fig. 4d).
The time profiles of E. coli inactivation revealed that 4 cm2 AgSA+NP/ZIF exhibited superior bactericidal efficiency compared to AgSA/ZIF and AgNP/ZIF (Fig. 4e). The converted atomic inactivation efficiency of AgSA+NP/ZIF within 30 min was 5.69 log10 per mg Ag (Supplementary Table 4), which was over 195 and 2630 times higher than those of AgSA/ZIF (3.40 log10 per mg Ag) and AgNP/ZIF (2.27 log10 per mg Ag), respectively. The bactericidal efficiency of AgSA+NP/ZIF correlates with the temperature rise curve (Supplementary Fig. 9). During the initial 15 min, as the water temperature gradually rises, the bactericidal efficiency increases slowly. When the temperature reaches 41.8 °C after 15 min, there is a marked increase in bactericidal efficiency. This result suggests that the photothermal effect of Ag nanoparticles significantly enhances the bactericidal activity. In addition, the disinfection efficiencies of 4 cm2 AgSA+NP/ZIF against Gram-positive bacteria Staphylococcus aureus (S. aureus) and virus MS2 were 2.10 log10 and 5.77 log10 (Fig. 4f), respectively, thereby confirming the high efficiency of AgSA+NP/ZIF to disinfect a broader range of microorganisms.
To verify the high bactericidal activity of the catalyst under varying water quality conditions, we studied its effectiveness against stress-resistant E. coli cultured in water from the middle reaches of Yellow River (Toudaoguai Station) and lake water (Shichahai). Despite differences in water types (Supplementary Table 3), the 4 cm2 AgSA+NP/ZIF achieved significant bacterial reductions of 4.35 log10 in the middle reaches of Yellow River and 6.27 log10 in lake water. The bactericidal activity in the middle reaches of the Yellow River was ~2 log10 lower than that observed in the upper reaches (TOC: 5.49 mg L−1), potentially attributed to higher concentrations of natural organic matter (TOC: 8.31 mg L−1) in the middle reaches that effectively quench ROS. Additionally, the 4 cm2 AgSA+NP/ZIF catalyst inactivated nearly 100% of the microorganisms in untreated real river water (Supplementary Fig. 10), demonstrating its potential for practical disinfection applications.
The application of AgSA+NP/ZIF was further extended to a large-scale water disinfection system in 1.0 L river water. As illustrated in Supplementary Fig. 11, four 3 × 3 cm2 AgSA+NP/ZIF were employed to sterilize water with shaking. To consistently raise the water temperature above 41 °C, the system required a minimum catalyst area of 36 cm2 and a velocity gradient of 70.7 s−1 (Supplementary Fig. 12a). This velocity corresponds to an extremely low rotational speed of ~100 rpm, which is insufficient to cause catalyst turnover and thereby fully expose the upper catalyst layer to light irradiation. The water temperature rose to 41.5 °C (Supplementary Fig. 12b) and 6.60 log10 E. coli inactivation was achieved within 1.0 h (Fig. 4h) under solar irradiation of 33.0-35.0 mW cm−2.
Durability experiments showed that the bactericidal activity of AgSA+NP/ZIF barely decreased after disinfecting a cumulative volume of 10.0 L water (Fig. 4i). This demonstrates its potential to meet the WHO recommended daily drinking water allocation of 8.0 L per person. The total loss of Ag compared to the total Ag present in the entire photocatalyst is 5.57% detected by ICP-MS. AC-HAADF-STEM image shows that Ag single atoms remain atomically dispersed and Ag nanoparticles are not depleted after the durability experiments (Supplementary Fig. 13), confirming the stability of AgSA+NP/ZIF. In addition, the average concentration of Ag+ and Zn2+ dissolved by AgSA+NP/ZIF for a 1-h cycle was 40.2 μg L−1 and 331.4 μg L−1 (Supplementary Table 5), respectively. The Chinese standard for drinking water quality (GB 5749-2022) stipulates that the concentration of Ag and Zn in drinking water should not exceed 50.0 μg L−1 and 1000.0 μg L−1, respectively, which indicated that our detected levels are complying.
The price of AgSA+NP/ZIF-based water disinfection was also evaluated as shown in Supplementary Table 6. Four 3 × 3 cm2 AgSA+NP/ZIF can disinfect at least 10.0 L water at a cost of $0.27, which is relatively higher compared to conventional treatment technologies such as chlorine dioxide disinfection, which costs $0.11. This higher cost arises from the synthesis process, which is conducted on a laboratory scale utilizing small packages of high-purity reagents (analytical reagent grade). However, the cost could be substantially reduced by scaling production to an industrial level, utilizing more cost-effective, technical-grade reagents. This adjustment could satisfy the economic and practical demands for disinfection applications.
Photothermal effect synergizes water disinfection
Previous studies have demonstrated that the photothermal action of plasmonic nanoparticle promotes protein aggregation in tumor cells and interferes their DNA damage repair capabilities22. However, the impacts of photothermal action on bacterial cells remain unclear. We hypothesize that the bacterial stress-resistant systems may potentially be compromised primarily due to the photothermal heating of Ag nanoparticles, resulting in increased bacterial susceptibility to ROS.
The transcriptomic analysis was employed to elucidate the impact of photothermal effect on the stress-resistant system of E. coli at the gene expression level. Compared to AgSA/ZIF treated E. coli, a total of 287 genes exhibited differential expression in E. coli treated with AgSA+NP/ZIF for 15 min ( | log2 fold-change | ≥ 1, p < 0.05), with 121 genes being downregulated and 166 genes being up-regulated (Fig. 5a). Gene Ontology (GO) functional enrichment analysis results shown in Fig. 5b and Supplementary Table 7 indicate that the downregulated genes are enriched in response to hydrogen peroxide, response to reactive oxygen species, and DNA repair items (p-value < 0.05), with the rich factors of 0.16, 0.11 and 0.07, respectively. The expression of downregulated genes in AgSA/ZIF-treated E. coli was 2286 (Supplementary Fig. 14). However, no significant enrichment was found in the aforementioned three GO categories (Supplementary Table 8), which suggests that the photothermal effect of AgSA+NP/ZIF at 41.8 °C can decrease the bacterial resistance to ROS and DNA repair ability45, with a higher proportion of downregulated genes responding to ROS (particularly hydrogen peroxide) compared to those involved in DNA damage repair. This result was further validated through disinfection experiments (Supplementary Fig. 15), wherein the 30-min inactivation of E. coli exhibited a >2.0 log increase when the temperature of the AgSA/ZIF system was elevated from 36 °C to 41 °C, approximating the temperature of the AgSA+NP/ZIF system. It is worth noting that the sole increase in temperature to 41 °C did not result in a significant impact on bacterial survival.
a Volcano plot of RNA sequencing data from E. coli culture exposed to AgSA+NP/ZIF for 15 min. b GO functional enrichment of the downregulated genes subset in E. coli exposure to AgSA+NP/ZIF for 15 min. Rich factor refers to the ratio of the number of differentially expressed genes to the total number of all annotated genes located in the same term. The dot color represents the p-value, and the dot size represents the number of differentially expressed genes in each term. c Intracellular ROS levels of E. coli exposed to monolithic photocatalysts for 15 min. d O2•− generation kinetics indicated by the absorbance of formazan dye that generated by the reaction of NBT with O2•−and H2O2 generation kinetics of AgSA+NP/ZIF. e •OH generation kinetics of AgSA+NP/ZIF and time profile of ABMDMA degradation by AgSA+NP/ZIF-generated 1O2. Reaction conditions: 4 cm2 monolithic photocatalysts; 0.1 L Yellow River water from upper reaches (Longyang Gorge); 33.0-35.0 mW cm−2 sunlight intensity. Experiments in d and e were conducted in triplicate, and the error bars represent the arithmetic mean ± standard deviation.
Given that the bacterial anti-stress system involves the removal of intracellular ROS, the fluorescent probe 2,7-dichlorodihydrofluorescein diacetate (DCFH-DA) was employed in conjunction with flow cytometry for semi-quantitative analysis of the total intracellular ROS accumulation. In the early phase of the reaction, the mean fluorescence intensity across the groups exhibited no significant variation, suggesting a similar accumulation of ROS (Supplementary Fig. 16). After 15 min (Fig. 5c), the mean fluorescence intensity in the AgSA+NP/ZIF-treated cells increased 33.6-fold when compared to that in the AgSA/ZIF-treated cells. Upon comparing the mean fluorescence intensity between groups following 30 min of reaction (Supplementary Fig. 16), the observed trend remained consistent. This phenomenon can be attributed to the photothermal effect of AgSA+NP/ZIF, which effectively down-regulates the cellular response to ROS and consequently hinders their removal45.
The mean fluorescence intensity of AgSA/ZIF-treated cells was 1.7 times higher compared to that of AgNP/ZIF-treated cells (Fig. 5c). This suggests that a sole photothermal effect from Ag nanoparticles is not significant to elevate intracellular ROS levels in E. coli. The reaction system was subsequently treated with 20 μM melatonin to eliminate exogenous ROS, while low levels of endogenous ROS were detected in E. coli as indicated by low 2,7-dichlorofluorescein (DCF) intensity (Supplementary Fig. 17). Therefore, the accumulated intracellular ROS mainly originated from exogenous ROS generated by the photocatalyst. The absence of Ag single-atom sites in AgNP/ZIF leads to insufficient generation of exogenous ROS, whose transmembrane import is the main cause of the rise in intracellular ROS levels. Thus, AgSA+NP/ZIF, which possesses atomically dispersed Ag sites and plasmonic Ag nanoparticles, could respectively facilitate intracellular ROS generation and accumulation in cells, thereby synergistically inactivating stress-resistant E. coli cells.
To comprehend the disinfection mechanisms associated with extracellular ROS, nitro blue tetrazolium chloride (NBT, O2•−-specific indicator), coumarin (Cou, •OH-specific indicator), 9,10-anthracenediyl-bis(methylene)dimalonic acid (ABMDMA, 1O2-specific indicator), p-hydroxyphenylacetic acid (HPA, H2O2-specific indicator) were applied to monitor specific ROS generated by AgSA+NP/ZIF in deionize water. Under solar irradiation, the absorbance of formazan dye generated by the reaction of NBT with O2•− increases in a time-dependent manner (Fig. 5d). Similarly, the fluorescent products generated by the reaction of H2O2 with HPA were detected, and the concentration of H2O2 after 30 min reaction was 5.22 µmol L−1 (Fig. 5d). In the disinfection system, H2O2 can be generated through both the oxygen reduction pathway (O2 + 2H+ + 2e− → H2O2) and the disproportionation reaction of O2•− (2O2•− + H+ + H2O → H2O2 + O2 + OH−)43. When investigating the production of •OH and 1O2, we found no significant increase in fluorescence intensity of Cou and decrease in absorbance of ABMDM (Fig. 5e), indicating the absence of detectable levels of •OH and 1O2. •OH generation is thermodynamically unfavorable as the valence band potential of AgSA+NP/ZIF (2.35 V vs. NHE) may not be sufficient to oxidize water (2.68 V vs. NHE)43.
The generation of extracellular ROS was further validated using two independent tests: electron paramagnetic resonance (EPR) spectroscopy and scavenger quenching analysis. A characteristic 5-tert-butoxycarbonyl-5-methyl-1-pyrroline N-oxide (BMPO)-O2•− peak indicated the production of O2•− (Supplementary Fig. 18a), while no EPR signals were detected for •OH and 1O2 (Supplementary Fig. 18b, c). Scavenging H2O2 and O2•− significantly reduced E. coli inactivation efficiency (Supplementary Fig. 19). Therefore, O2•− and H2O2 are the primary ROS responsible for AgSA+NP/ZIF disinfection, with h+ playing a secondary role in E. coli inactivation. It is worth mentioning that O2•− and H2O2 both exhibit longer lifespans compared to •OH and elicit longer disinfection power46,47. The direct contact between bacterial cells and photocatalyst is less important for our disinfection system, as H2O2 and O2•− can effectively diffuse into the bulk solution and inactivate bacterial cells that are not in contact with the photocatalyst48.
TEM analysis showed no significant damage to the cell membrane structure of AgSA+NP/ZIF-treated bacteria compared to untreated bacteria (Supplementary Fig. 20a, b). Bacteria with intact membrane selectivity remain impermeable to propidium iodide (PI) staining but can be stained by SYTO9, while the bacteria with disrupted membrane selectivity are PI-permeable. Approximately 71.2% of bacteria exhibited disrupted membrane selectivity after 30-min treatment with AgSA+NP/ZIF (Supplementary Fig. 20c-e). The disruption interfered with normal cellular transport, as evidenced by the down-regulation of genes related to electron transport coupled proton transport, amino acid transmembrane transport, and anion transmembrane transport (Fig. 5b). Quantitative polymerase chain reaction (qPCR) analysis revealed a 1.27 log reduction in the 16s gene copy number in the AgSA+NP/ZIF-treated group after 30 min (Supplementary Fig. 21), indicating significant DNA damage. This result is consistent with transcriptome analysis (Fig. 5b) showing down-regulation of DNA repair genes. While direct structural damage to the cell membrane was minimal due to limited contact, AgSA+NP/ZIF effectively compromised membrane selectivity and induced intracellular DNA degradation through ROS generation.
Discussion
In this study, we successfully developed a floatable monolithic AgSA+NP/ZIF photocatalyst for water disinfection. Atomically dispersed Ag promotes charge transfer to facilitate the generation of H2O2 and O2•−. Transcriptomics analysis suggests that the photothermal effect of plasmonic Ag nanoparticles disrupted the anti-stress system of E. coli, thereby reducing their resistance to ROS. Under solar irradiation, the synergistic effect of Ag single atoms and nanoparticles enables 36 cm2 AgSA+NP/ZIF to disinfect at least 10.0 L of surface water, which meets the WHO’s recommended daily per capita drinking water allocation. The present study demonstrates a potential approach to addressing the challenges of low photocatalytic disinfection efficiency in real-world water and the limited recyclability of traditional granular photocatalysts. It provides a foundation for sustainable and environmentally friendly POU water disinfection. However, the findings are based on the evaluation of a single photocatalyst under controlled laboratory conditions. To apply this approach in real-world settings, particularly in off-grid regions, requires overcoming several critical challenges. These include reducing material costs, improving scalability, and ensuring sustained photocatalytic performance under varying environmental conditions. Furthermore, overcoming practical obstacles such as efficient system integration and optimizing energy consumption is essential for the successful deployment and long-term viability of this technology in large-scale water treatment systems.
Methods
Monolithic ZIF-8-NH2 support preparation
Commercially available sponges are mainly composed of melamine and have a porous structure with a density of 45.0 mg cm−3. The 4×2×1 cm3 three-dimensional melamine sponge was utilized as a substrate to absorb 10 mL 2.0 wt% aqueous sodium alginate solution (99%; Aladdin), followed by immersion in 30 mL 2.0 wt% zinc nitrate hexahydrate solution (99%; Aladdin) for 12 h to form gel-coated sponge. The gel-coated sponge was then placed in an oven and dried at 150 °C for 2 h to facilitate water evaporation and obtain a desiccated sponge. Then 0.279 g 2-methylimidazole (99%; Sigma Aldrich) and 0.08 g 2-aminobenzimidazole (97%; Aladdin) were placed at the bottom of the customized Teflon-lined stainless-steel autoclave, and a 1×1×0.2 cm3 dried sponge situated above the organic ligands49. The dried sponge and organic ligand powder were kept separate. The autoclave was heated at 150 °C for ligand evaporation and deposition. After the reaction for 2 h, the obtained monolithic ZIF-8-NH2 support was washed three times with deionized water and dried in a vacuum oven at 60 °C for 12 h. Larger size monolithic ZIF-8-NH2 was prepared through proportionally increasing the amount of reaction reagents.
Monolithic photocatalyst preparation
The 1 × 1 × 0.2 cm3 monolithic ZIF-8-NH2 sponge and 120 µL 67.9 mg mL−1 silver nitrate solution (99.9%; Sigma Aldrich) were added to 10 mL of deionized water and left for 12 h to sufficiently adsorb Ag ions. The mixture was then snap-frozen in liquid nitrogen and the upper surface (1×1 cm2) of monolithic ZIF-8-NH2 sponge was irradiated under UV lamp (Beijing Perfectlight, PCX-50C Discover) with a peak intensity at 365 nm for 10 min. The light intensity on the reaction solution surface was 52.8 mW cm−2 as measured using the UV power meter (UV-A, Beijing Normal University Photoelectric Instrument Factory). The ambient temperature is maintained at 25 °C to allow the ice to melt during irradiation. The ice lattice inhibited nucleation50, thus reducing Ag ions to single atoms. With the melting of ice, a portion of Ag was transformed into nanoparticles, leading to the formation of photocatalysts that incorporate both Ag single atoms and Ag nanoparticles. The obtained monolithic AgSA+NP/ZIF was washed three times with deionized water and dried in a vacuum oven at 60 °C. In addition, the synthesis of AgSA/ZIF was carried out under the same conditions except that the ambient temperature was kept at 0 °C. AgNP/ZIF was synthesized by the same process without liquid nitrogen freezing.
Material characterization
The microstructure and morphology of photocatalysts were obtained by SEM (FEI Inspect F50) and high-resolution TEM (FEI Talos F200S). Ag single-atom was observed by AC-HAADF-STEM (Titan Themis 60-300) operated at 200 kV, coupled with a probe spherical aberration corrector. Atomic dispersion and coordination environment of Ag species were confirmed by XAFS measurements, which were carried out on the sample at 11S2 X-ray absorption beamline of Aichi Synchrotron Radiation Center. The radiation was monochromatized by a Si (111) double-crystal monochromator. The measurements were carried out with a passivated implanted planar silicon fluorescence detector using Ag foil to calibrate the energy. XAFS data was analyzed by the Athena (version 0.9.26) and Artemis software (version 0.9.26) as shown in Supplementary Table 1 and Supplementary Note. The Ag loading content and dissolved Ag+ concentration of photocatalysts were measured by ICP-OES (Agilent 725-ES) and ICP-MS (Agilent 7700), respectively. The UV-Vis diffuse reflectance spectra of the photocatalysts were recorded on a PE Lambda750 spectrophotometer. The steady-state and time-resolved photoluminescence spectra were carried out for the characterization of electron-hole pairs separation using the Hitachi F-7000 spectrophotometer and Edinburgh FLS1000 spectrofluorometer with an excitation wavelength of 375 nm, respectively. The average lifetime was obtained by fitting the luminescence decay curves with a biexponential function. Electrochemical impedance spectroscopy and Mott–Schottky measurements were performed in 0.5 M Na2SO4 (99%; Aladdin) solution by using a CHI-660E electrochemical workstation (Shanghai, China). Photothermal properties of photocatalysts were characterized by a digital infrared thermal camera (FLIR E50).
Since Ag single atoms and Ag nanoparticles account for 100% of the silver content, the proportion of Ag nanoparticles can be deduced from the Ag single atoms proportion. The proportion of Ag single atoms can be estimated using the average Ag−N coordination number51. The Ag−N coordination number of atomically dispersed Ag sites in AgSA/ZIF catalyst is 3.2. Therefore, the average Ag−N coordination number in AgSA+NP/ZIF catalyst can be calculated by the following equation:
where n represents the coordination number of Ag−N in AgSA+NP/ZIF catalyst, NAg−N represents the number of total Ag−N bond, NAg represents the number of total Ag atoms, and X1 represents the proportion of Ag single atoms in AgSA+NP/ZIF catalyst. Therefore, Using the observed coordination number (Supplementary Table 1), the proportion of Ag single atoms in the AgSA+NP/ZIF catalyst can be approximately calculated as follow:
Evaluation of the photocatalytic disinfection performance
River water and lake water with various water qualities, including the upper reaches of Yellow River (Longyang Gorge, Qinghai Province, China), the middle reaches of Yellow River (Toudaoguai Station, Inner Mongolia, China), and lake water (Shichahai, Beijing, China), were collected to test disinfection activity. The chemical parameters of the water samples were detected by the water quality analyzer (ProQuatro, YSI) as shown in Supplementary Table 3. Briefly, river water was filtered through cellulose membranes with the pore size of 0.45 μm to remove suspended sediment. The filtered river water was supplemented with 100 mg L−1 of yeast extract and autoclaved to remove background microorganisms. This sterilized water samples were used to grow bacteria for disinfection experiments and RNA extraction.
E. coli BW25113 purchased from the Coli Genetic Stock Center (CGSC, Yale University, USA) was used as the model microorganism. 5 μL of E. coli strain were inoculated in 5 mL of Luria-Bertani (LB) medium at 37 °C for 12 h, then 5 mL of culture was transferred to 50 mL of LB medium and cultured for another 2 h. E. coli culture was collected and washed by centrifugation, and re-suspended in 50 mL sterilized river water to achieve an optical density of 0.06 at 670 nm ( ≈ 1×108 bacteria mL−1). E. coli cells was allowed to stand in the sterilized river water at 25 °C for 24 h to acclimatize to the new growth conditions and collected by centrifugation. Collected E. coli cells were then transferred to 50 mL fresh sterilized river water for 48 h incubation with shaking at 160 rpm at 25 °C to reach the late stationary phase (Supplementary Fig. 22). This growth phase better resembles bacterial cell conditions observed in oligotrophic surface water, where cellular responses are typically activated to counter starvation and other environmental stressors17. After 48 h of incubation, E. coli cultures were diluted 100-fold by fresh river water for disinfection testing. The monolithic photocatalyst with surface area of 1-4 cm2 was plunged into 0.1 L E. coli suspension and irradiated without any agitation under sunlight (14:00-15:00 pm in September in Beijing, China) with a light intensity of 33.0-35.0 mW cm−2. For large-scale water disinfection, 1.0 L E. coli suspension was placed in a 2.0 L beaker and shaken on a shaker at a constant velocity. The Camp-Stein velocity gradient (G) for a shaker was calculated by the following Eq. (3):
where P represents the power input into the control volume (0-14 W), p represents the density of the fluid (997 kg m−3), v represents the volume of the fluid (0.001 m3), and ∀ represents volume of the control volume (0.002 m3).
Five mL of the sample was carefully pipetted out at the sampling time and diluted 10, 100, and 1000 times with sterile normal saline, respectively. 100 μL of the diluted E. coli samples were spread evenly on LB agar plates and incubated at 37 °C overnight for colony counting. The concentration of E. coli was calculated by the formula: C = n × dilution times × 10, where C represented the concentration of E. coli in the sample (CFU mL−1) and n represented the number of colonies. The concentration of E. coli without photocatalytic treatment was represented by C0, thus the bactericidal efficiency was expressed as Log10(C0/C). Each experiment was performed three times independently.
Transcriptome analysis
To analyze the impact of photothermal stress on biological functions in E. coli, transcriptomics analysis was conducted by extracting RNA from E. coli cells exposed to AgSA/ZIF and AgSA+NP/ZIF for 15 min, as well as a control group. Total RNA extraction was performed using the Total RNA Extractor kit (Sangon Biotech, Shanghai) for subsequent RNA-seq analysis. The quality and quantity of total isolated RNA in the samples were analyzed by a Qubit 2.0 fluorometer using Qubit RNA Quantification Kit (Invitrogen). The rRNA depletion was carried out using Ribo-off rRNA Depletion Kit (Vazyme, Nanjing) according to manufacturer’s specifications. The rRNA-deleted samples were further purified to obtain purified mRNA, and the RNA libraries of purified mRNA were constructed with VAHTS Stranded mRNA-seq V3 Library Prep Kit for Illumina (Vazyme, Nanjing). The quality of the RNA libraries was checked by 8% polyacrylamide gel electrophoresis. After that, all libraries were sequenced using an Illumina HiSeq 3000 system with paired-end 150 base-pair reads. Significantly differentially expressed genes were analyzed by DESeq2 R-package (version 1. 12. 4) with a screening condition of p-value ≤ 0.05 and |log2 fold-change | ≥1. GO enrichment analysis was performed by topGO R-package (version 2. 24. 0) to evaluate the biological function enrichment corresponding to the differentially expressed genes sets. In general, GO terms with an adjusted p-value < 0.05 were considered significantly enriched.
ROS measurements
When bacterial cells are exposure to monolithic photocatalyst, ROS may penetrate the cell membrane causing intracellular ROS accumulation, which was detected by using the fluorescent probe DCFH-DA (99.81%; MCE) in conjunction with flow cytometer (ACEA NoveCyteTM). DCFH-DA can freely penetrate cellular membranes and undergo intracellular hydrolysis by esterase, resulting in the formation of 2,7-dichlorodihydrofluorescein (DCFH) that becomes sequestered within the cytomembrane. Subsequently, intracellular ROS oxidizes DCFH to generate fluorescent DCF, exhibiting maximum excitation and emission wavelengths at 488 nm and 525 nm, respectively. Briefly, E. coli treated for 0-30 min in the photocatalytic disinfection system were incubated with DCFH-DA at a concentration of 25 μM for 30 min at 37 °C. The bacterial cells after incubation were collected and washed with saline (85% NaCl solution) to remove excess fluorescent probe. The flow cytometer laser was selected to be 488 nm, and the number of the collected cells was 100,000. The sample collection speed was adjusted to the low mode (14 μL min−1), and the threshold of FSC-H was set to 1000. The mean intracellular fluorescence intensity of bacteria was analyzed by FlowJo V10 software.
For H2O2 concentration detection, a working solution was prepared by adding 2.7 mg of p-hydroxyphenylacetic acid (98%; Aladdin) and 1 mg of horseradish peroxidase (200 U mg−1; Aladdin) to 10 mL of potassium hydrogen phthalate buffer solution (8.2 g L−1). Subsequently, 50 μL of working solution was added to 1 mL sample, diluted with 1 mL deionized water, and the reaction was quenched by adding 1 mL 0.1 M NaOH (99%; Aladdin) after 10 min. The fluorescence spectra of samples were measured with an excitation wavelength of 315 nm and detected at an emission wavelength of 409 nm. For the detection of O2•−, 30 µL of 1 mM NBT (98%; Aladdin) dissolved in dimethyl sulfoxide (99.8%; Aladdin) and 3 mL of sample were mixed and incubated for 10 min. The reaction generated formazan dye with maximum absorbance at 680 nm was detected by UV-Vis spectrophotometer. The concentration of •OH generation was determined by the fluorescence method using coumarin as a probe43. One mM coumarin (99%; Aladdin) and AgSA+NP/ZIF were mixed in 0.1 L water. Coumarin reacts with •OH to produce 7-hydroxycoumarin, which can emit fluorescence at 451 nm with excitation wavelength of 332 nm. The presence of 1O2 was detected by using ABMDMA (90%; MCE) as a sensitive indicator. The bleaching of ABMDMA by singlet oxygen resulted in a decrease in its absorbance at 400 nm. 0.5 mL ABMDMA (2 mM) and AgSA+NP/ZIF were mixed in 0.1 L water. The consumption of ABMDMA was monitored by recording the absorbance at 400 nm with a UV-Vis spectrophotometer. The sunlight intensity was 33.0-35.0 mW cm−2 and AgSA+NP/ZIF size is 4 cm2.
ROS were detected by EPR spectroscopy on a Bruker A300 spectrometer operating at an X-band frequency of 9.85 GHz. The monolithic catalyst was crushed and prepared as a suspension sample (100 mg L−1). Ten µL of 50 mg mL−1 BMPO (99%; Aladdin) and 10 µL of water were mixed with 30 µL of catalyst sample. The resulting liquid was aspirated using a capillary tube, followed by light exposure for 10 min to detect both •OH and O2•−. For the detection of •OH alone, the procedure was modified by substituting 10 µL of water with 10 µL of superoxide dismutase (SOD, 5 U mL−1; Aladdin) to quench O2•−. To detect 1O2, 25 µL of 10 mM 2,2,6,6-tetramethylpiperidine (TEMP, 99%; Aladdin) was mixed with 25 µL of catalyst sample, and measurements were conducted under both 10-min light exposure and in dark conditions.
The scavengers used for •OH elimination, H2O2 neutralization, O2•− scavenging, hole scavenger, electron capture, and 1O2 quenching were 2.5 mM isopropanol (99.5%; Sigma), catalase (CAT, 300 U mL−1; Sigma), SOD (400 U mL−1; Sigma), 2.5 mM sodium oxalate (99.5%; Sigma), 2.5 mM sodium chromate (Cr(VI)), (99.5%; Sigma), and 2.5 mM furfuryl alcohol (99%; Sigma), respectively. The scavengers applied in this study showed no significant impact on bacterial activity (Supplementary Fig. 23). These scavengers were added to the bacterial suspension prior to illumination. Bacterial concentrations in solution were determined at different time intervals using standard spread plating method. Each sample was serially diluted and plated in triplicate onto nutrient agar before incubation at 37 °C for 24 h.
Membrane damage and DNA degradation
Cell membrane structure damage was analyzed by TEM on sliced bacterial specimen. Briefly, the harvested bacterial cells were first fixed in a solution of 4% paraformaldehyde (95%; Aladdin) and 2.5% glutaraldehyde (50%; Aladdin) for 12 h at 4 °C. The samples were then infiltrated and embedded in Spurr’s resin with propylene oxide (99%; Aladdin) and cured overnight at 70 °C. Subsequently, ultrathin sections were prepared using an ultramicrotome (MT-X, RMC) and examined with a Hitachi H7700 TEM operated at 75 kV. The collected samples were stained with LIVE/DEAD BacLight Viability Kit (Thermo Fisher scientific) and analyzed with flow cytometry (ACEA NoveCyteTM). This kit contains two fluorescent nucleic acids dyes, namely SYTO 9 and PI. E. coli cells with intact membrane can only be stained green by SYTO 9 (Ex. 488 nm and Em. 530 nm), and E. coli cells with disrupted membrane selectivity can be stained red by PI (Ex. 488 nm and Em. 630 nm). The defined gates include: complete membrane destruction (Q1), disrupted membrane selectivity (Q2), impurities (Q3), and intact membrane (Q4). The proportion of bacteria (P) with disrupted membrane selectivity was determined using Eq. (4):
The qPCR amplicon targeting the 16S gene was employed for highly sensitive detection of DNA damage52, and the primer sequence can be found in Supplementary Table 9. The qPCR reaction parameters were as follows: pre-incubation at 95 °C for 3 min; amplification of 35 cycles with each cycle being 95 °C for 30 s, 60 °C for 30 s, and 72 °C for 45 s; melt curve generation at 90 °C for 10 s, and 60 °C for 60 s; and a gradual heating to 90 °C at a rate of 0.03 °C s−1.
Electromagnetic simulation
To explain the enhancement local electromagnetic field near the surface between the Ag nanoparticles and the ZIF-8-NH2, we used COMSOL Multiphysics (version 6.2), employing the Electromagnetic Waves, Frequency Domain physics interface and Wavelength Domain in the optics module. Briefly, the refractive indices of Ag nanoparticles and air were obtained from the COMSOL built-in material library, while the refractive index of ZIF-8-NH2 substrate was set to 1.38. A series of parallel simulations is designed to show the effect of varying ZIF-8-NH2 thickness and Ag nanoparticle size on the electromagnetic field at the interface between Ag nanoparticle and ZIF-8-NH2 under 450 nm irradiation. The variations include ZIF-8-NH2 thicknesses ranging from 1 to 15 nm and Ag nanoparticle radius of 1, 3.5, 5, 6.5, 8 nm. The geometric model consisted of one Ag spherical particle on the center of a ZIF-8-NH2 substrate as shown in the Supplementary Fig. 24. Incident light with a wavelength of 450 nm was directed vertically downward pointing the x-y plane. The air domain is confined by the inclusion of a Perfect Matching Layer (PML), which restricts the model to a specific area of interest to conserve computing resources while maintaining accuracy consistent with results from a large, unrestricted solution domain53. The solution within this domain remains unaffected by the introduced PML. The Perfect Electrical Conductor (PEC) boundary conditions were used on the x-y symmetry plane. A Frequency Domain was defined as the initial incident electric field E0, serving as the background for the scattered field. The distribution of |E | / | E0| was computed to analyze the electromagnetic field enhancement at the interface of Ag nanoparticles loaded on ZIF-8-NH2. The COMSOL simulation parameters are provided in Supplementary Method and Supplementary Table 10.
Reporting summary
Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.
Data availability
The RNA sequencing data generated in this study have been deposited in the NCBI Sequence Read Archive (SRA) database under accession code PRJNA1199748. Data supporting the findings of this work are available within the paper and its Supplementary Information files. Source data are provided with this paper. The full image data are available from the corresponding author upon request. Source data are provided with this paper.
References
World Health Organization & United Nations Children’s Fund (UNICEF). Progress on Sanitation and Drinking Water—2015 Update and MDG Assessment (World Health Organization, Geneva, 2015).
Alvarez, P. J. J., Chan, C. K., Elimelech, M., Halas, N. J. & Villagrán, D. Emerging opportunities for nanotechnology to enhance water security. Nat. Nanotechnol. 13, 634–641 (2018).
Liu, L. et al. Global, regional, and national causes of child mortality: an updated systematic analysis for 2010 with time trends since 2000. The Lancet 379, 2151–2161 (2012).
Prüss‐Ustün, A. et al. Burden of disease from inadequate water, sanitation and hygiene in low‐ and middle‐income settings: a retrospective analysis of data from 145 countries. Trop. Med. Int. Health 19, 894–905 (2014).
Chu, C., Ryberg, E. C., Loeb, S. K., Suh, M.-J. & Kim, J.-H. Water Disinfection in Rural Areas Demands Unconventional Solar Technologies. Acc. Chem. Res. 52, 1187–1195 (2019).
Richards, T. et al. A residue-free approach to water disinfection using catalytic in situ generation of reactive oxygen species. Nat. Catal. 4, 575–585 (2021).
Montgomery, M. A. & Elimelech, M. Water And Sanitation in Developing Countries: Including Health in the Equation. Environ. Sci. Technol. 41, 17–24 (2007).
Smieja, J. A. Household Water Treatments in Developing Countries. J. Chem. Educ. 88, 549–553 (2011).
Palika, A. et al. An antiviral trap made of protein nanofibrils and iron oxyhydroxide nanoparticles. Nat. Nanotechnol. 16, 918–925 (2021).
Zhou, X. et al. Bacteria inactivation by sulfate radical: progress and non-negligible disinfection by-products. Front. Environ. Sci. Eng. 17, 29 (2023).
Chen, R., Chen, S., Wang, L. & Wang, D. Nanoscale Metal Particle Modified Single‐Atom Catalyst: Synthesis, Characterization, and Application. Adv. Mater. 36, 2304713 (2024).
Gloag, L., Somerville, S. V., Gooding, J. J. & Tilley, R. D. Co-catalytic metal–support interactions in single-atom electrocatalysts. Nat. Rev. Mater. 9, 173–189 (2024).
Liu, Y., Zhao, H. & Zhao, Y. Designing Efficient Single Metal Atom Biocatalysts at the Atomic Structure Level. Angew. Chem. Int. Ed. 63, e202315933 (2024).
Li, Y. et al. Visible-light-driven photocatalytic disinfection mechanism of Pb-BiFeO3/rGO photocatalyst. Water Res. 161, 251–261 (2019).
Zhou, Z. et al. Structural insights into the inhibition of bacterial RecA by naphthalene polysulfonated compounds. iScience 24, 101952 (2021).
Wu, T. et al. Solar-driven efficient heterogeneous subminute water disinfection nanosystem assembled with fingerprint MoS2. Nat. Water 1, 462–470 (2023).
Daer, S., Goodwill, J. E. & Ikuma, K. Effect of ferrate and monochloramine disinfection on the physiological and transcriptomic response of Escherichia coli at late stationary phase. Water Res. 189, 116580 (2021).
Zhu, Z. et al. Entropy of a bacterial stress response is a generalizable predictor for fitness and antibiotic sensitivity. Nat. Commun. 11, 4365 (2020).
Lisle, J. T. et al. Effects of Starvation on Physiological Activity and Chlorine Disinfection Resistance in Escherichia coli O157:H7. Appl. Environ. Microbiol. 64, 4658–4662 (1998).
Jiang, H. et al. Light-driven CO2 methanation over Au-grafted Ce0.95Ru0.05O2 solid-solution catalysts with activities approaching the thermodynamic limit. Nat. Catal. 6, 519–530 (2023).
Jiang, J., Wang, X. & Guo, H. Enhanced Interfacial Charge Transfer/Separation By LSPR‐Induced Defective Semiconductor Toward High Co2 RR Performance. Small 19, 2301280 (2023).
Cai, R. et al. Plasmonic AuPt@CuS Heterostructure with Enhanced Synergistic Efficacy for Radiophotothermal Therapy. J. Am. Chem. Soc. 143, 16113–16127 (2021).
Xia, C. et al. General synthesis of single-atom catalysts with high metal loading using graphene quantum dots. Nat. Chem. 13, 887–894 (2021).
Hai, X. et al. Scalable two-step annealing method for preparing ultra-high-density single-atom catalyst libraries. Nat. Nanotechnol. 17, 174–181 (2022).
Sun, T. et al. Single-atomic cobalt sites embedded in hierarchically ordered porous nitrogen-doped carbon as a superior bifunctional electrocatalyst. Proc. Natl. Acad. Sci. 115, 12692–12697 (2018).
Peugeot, A. et al. Benchmarking of oxygen evolution catalysts on porous nickel supports. Joule 5, 1281–1300 (2021).
Cai, C., Sun, W., He, S., Zhang, Y. & Wang, X. Ceramic membrane fouling mechanisms and control for water treatment. Front. Environ. Sci. Eng. 17, 126 (2023).
Zhao, C. et al. Solid-Diffusion Synthesis of Single-Atom Catalysts Directly from Bulk Metal for Efficient CO2 Reduction. Joule 3, 584–594 (2019).
Liang, Q. et al. Macroscopic 3D Porous Graphitic Carbon Nitride Monolith for Enhanced Photocatalytic Hydrogen Evolution. Adv. Mater. 27, 4634–4639 (2015).
Zhu, L. et al. Designing 3D‐MoS2 Sponge as Excellent Cocatalysts in Advanced Oxidation Processes for Pollutant Control. Angew. Chem. Int. Ed. 59, 13968–13976 (2020).
Jiang, Z. et al. Filling metal–organic framework mesopores with TiO2 for CO2 photoreduction. Nature 586, 549–554 (2020).
Li, R., Chen, T. & Pan, X. Metal–Organic-Framework-Based Materials for Antimicrobial Applications. ACS Nano 15, 3808–3848 (2021).
Zhang, X. et al. A General Method for Transition Metal Single Atoms Anchored on Honeycomb‐Like Nitrogen‐Doped Carbon Nanosheets. Adv. Mater. 32, 1906905 (2020).
Yu, S. et al. Covalently bonded zeolitic imidazolate frameworks and polymers with enhanced compatibility in thin film nanocomposite membranes for gas separation. J. Membr. Sci. 540, 155–164 (2017).
Ji, S. et al. Chemical Synthesis of Single Atomic Site Catalysts. Chem. Rev. 120, 11900–11955 (2020).
You, J., Guo, Y., Guo, R. & Liu, X. A review of visible light-active photocatalysts for water disinfection: Features and prospects. Chem. Eng. J. 373, 624–641 (2019).
Cao, S. et al. Single-Atom Engineering of Directional Charge Transfer Channels and Active Sites for Photocatalytic Hydrogen Evolution. Adv. Funct. Mater. 28, 1802169 (2018).
Zhu, Z. et al. Construction of high-dispersed Ag/Fe3O4/g-C3N4 photocatalyst by selective photo-deposition and improved photocatalytic activity. Appl. Catal. B Environ. 182, 115–122 (2016).
Wu, Z. et al. Dual‐Site Activation Coupling with a Schottky Junction Boosts the Electrochemiluminescence of Carbon Nitride. Angew. Chem. Int. Ed. 62, e202308257 (2023).
Zhang, W., Li, Y., Niu, J. & Chen, Y. Photogeneration of Reactive Oxygen Species on Uncoated Silver, Gold, Nickel, and Silicon Nanoparticles and Their Antibacterial Effects. Langmuir 29, 4647–4651 (2013).
Abdi, J. Synthesis of Ag-doped ZIF-8 photocatalyst with excellent performance for dye degradation and antibacterial activity. Colloids Surf. Physicochem. Eng. Asp. 604, 125330 (2020).
Guo, X., He, S., Meng, Z., Wang, Y. & Peng, Y. Ag@ZIF-8/g-C3N4 Z-scheme photocatalyst for the enhanced removal of multiple classes of antibiotics by integrated adsorption and photocatalytic degradation under visible light irradiation. RSC Adv. 12, 17919–17931 (2022).
Li, P. et al. Metal-organic frameworks with photocatalytic bactericidal activity for integrated air cleaning. Nat. Commun. 10, 2177 (2019).
Zhou, P., Luo, M. & Guo, S. Optimizing the semiconductor–metal-single-atom interaction for photocatalytic reactivity. Nat. Rev. Chem. 6, 823–838 (2022).
Yang, L. et al. Cellular responses to reactive oxygen species are predicted from molecular mechanisms. Proc. Natl. Acad. Sci. 116, 14368–14373 (2019).
Xia, D. et al. Single Ag atom engineered 3D-MnO2 porous hollow microspheres for rapid photothermocatalytic inactivation of E. coli under solar light. Appl. Catal. B Environ. 245, 177–189 (2019).
Nosaka, Y. & Nosaka, A. Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 117, 11302–11336 (2017).
Wang, W., Huang, G., Yu, J. C. & Wong, P. K. Advances in photocatalytic disinfection of bacteria: Development of photocatalysts and mechanisms. J. Environ. Sci. 34, 232–247 (2015).
Li, W. et al. Ultrathin metal–organic framework membrane production by gel–vapour deposition. Nat. Commun. 8, 406 (2017).
Wei, H. et al. Iced photochemical reduction to synthesize atomically dispersed metals by suppressing nanocrystal growth. Nat. Commun. 8, 1490 (2017).
Fu, N. et al. Controllable Conversion of Platinum Nanoparticles to Single Atoms in Pt/CeO2 by Laser Ablation for Efficient CO Oxidation. J. Am. Chem. Soc. 145, 9540–9547 (2023).
Yang, C., Sun, W. & Ao, X. Bacterial inactivation, DNA damage, and faster ATP degradation induced by ultraviolet disinfection. Front. Environ. Sci. Eng. 14, 13 (2020).
Yan, H., Song, X., Wang, X. & Wang, Y. Electromagnetic wave absorption and scattering analysis for Fe3O4 with different scales particles. Chem. Phys. Lett. 723, 51–56 (2019).
Acknowledgements
This study was financially supported by the Beijing Natural Science Foundation (JQ24054, Y.L.), and the National Natural Science Foundation of China (grant numbers T2421005, X.H.X. and 52170024, Y.L.).
Author information
Authors and Affiliations
Contributions
Y.L., W.Z. and J.W. conceived the idea and designed the research. J.W. conducted the experiments and performed the data analysis. J.Z. performed the electromagnetic simulation. Y.L., W.Z., J.-H.K., H.Y. and J.W. drafted the manuscript. J.W., Y.L. and X.X. reviewed and edited the manuscript.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks the anonymous, reviewers for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Wang, J., Zhang, J., Li, Y. et al. Silver single atoms and nanoparticles on floatable monolithic photocatalysts for synergistic solar water disinfection. Nat Commun 16, 981 (2025). https://doi.org/10.1038/s41467-025-56339-2
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-025-56339-2
This article is cited by
-
Reusable photocatalytic film for efficient water disinfection under low light intensity
Nature Water (2025)
-
Photodynamic antimicrobial chemotherapy outperforms solar disinfection in the treatment of water to remove bacterial pathogens
Sustainable Water Resources Management (2025)
-
Redox-mediated stabilization of ultrasmall Au25 nanoclusters in amine-functionalized MOF for robust photocatalytic antibacterial applications
Science China Materials (2025)