Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

RESEARCH ARTICLE | ENVIRONMENTAL SCIENCES

Dynamic defects boost in-­situ H2O2 piezocatalysis for water


cleanup
Maoxi Rana,b,c, Bibai Dub, Wenyuan Liua , Zhiyan Lianga, Lihong Lianga, Yayun Zhangd , Lixi Zengb,1 , and Mingyang Xinga,c,1

Edited by Alexis Bell, University of California, Berkeley, CA; received October 8, 2023; accepted January 3, 2024

Creating efficient catalysts for simultaneous H2O2 generation and pollutant degradation
is vital. Piezocatalytic H2O2 synthesis offers a promising alternative to traditional meth- Significance
ods but faces challenges like sacrificial reagents, harsh conditions, and low activity. In this
study, we introduce a cobalt-­loaded ZnO (CZO) piezocatalyst that efficiently generates Efficient in-­situ H2O2 yield was
H2O2 from H2O and O2 under ultrasonic (US) treatment in ambient aqueous condi- attributed to the generation of
tions. The catalyst demonstrates exceptional performance with ~50.9% TOC removal dynamic defects promoting the
of phenol and in situ generation of 1.3 mM H2O2, significantly outperforming pure adsorption and activation of
ZnO. Notably, the CZO piezocatalyst maintains its H2O2 generation capability even oxygen, which were caused by
after multiple cycles, showing continuous improvement (from 1.3 mM to 1.8 mM). This cobalt deposition, enhancing
is attributed to the piezoelectric electrons promoting the generation of dynamic defects piezoelectric activity during the
under US conditions, which in turn promotes the adsorption and activation of oxygen,
US process and activating lattice
thereby facilitating efficient H2O2 production, as confirmed by EPR spectrometry, XPS
oxygen with more piezoelectric
analysis, and DFT calculations. Moreover, the CZO piezocatalysts maintain outstanding
performance in pollutant degradation and H2O2 production even after long periods of electrons. More importantly,
inactivity, and the deactivated catalyst due to metal ion dissolution could be rejuvenated the inevitable issue of catalyst
by pH adjustment, offering a sustainable solution for wastewater purification. deactivation due to metal ion
dissolution not only got resolved
H2O2 in-­situ generation | dynamic defects | O2 activation | piezocatalytic performance | by pH adjustment, but even
water remediation
showed significant in situ
H2O2 yield. Additionally, CZO
Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

Refractory pollutants pose a threat to human health and environmental ecosystems once piezocatalyst containing defects
discharged into natural water bodies (1–4). Conventional wastewater treatment methods,
could be repeatedly utilized,
such as adsorption, can only transfer pollutants without complete degradation, while
membrane processes suffer from high costs and membrane fouling. Currently, hydrogen maintaining high H2O2
peroxide (H2O2), an environmentally friendly green oxidant, releases highly reactive free production, and simultaneously
radicals to remove pollutants (5). The anthraquinone oxidation (AO) method, which achieving advanced degradation
involved the Pd-­catalyzed cyclic hydrogenation and oxidation of alkyl-­anthraquinones in of pollutants due to their robust
organic solvents, was commercially employed for H2O2 production (6, 7). However, this self-­repair capabilities.
process was associated with significant energy consumption and environmental impacts
and safety concerns regarding H2O2 transportation and storage. Therefore, developing a
cost-­effective and environmentally friendly in situ H2O2 production and activation cascade
reaction is a sustainable strategy in the field of pollutant removal.
Piezoelectric catalysis has attracted widespread research interest in wastewater treatment
due to its simplicity, cleanliness, and high efficiency (8, 9). It is considered a potential
sustainable alternative to AO method for the production of H2O2. Similar to photocat­
alytic processes, under ultrasonic (US) activation, polarization, and built-­in electric fields
are generated in piezoelectric materials, allowing continuous separation of electrons and
holes and attracting them to opposite surfaces (10). In aqueous solutions, surface charge
carriers can react with water or dissolved oxygen to generate reactive oxygen species (ROS)
for wastewater treatment, achieving the conversion of mechanical energy into chemical
Author contributions: M.X. and L.Z. designed research;
energy (O2 + e− → •O2–, •O2– + e− + 2H+ → H2O2; 2H2O + 2h+ → H2O2 + 2H+). M.R. performed research; Y.Z. contributed analytic tools;
Currently, varieties of piezoelectric semiconductors, e.g., BaTiO3 (11), g-­C3N4 (12), M.R., B.D., W.L., Z.L., L.L., L.Z., and M.X. analyzed data;
and M.R., L.Z., and M.X. wrote the paper.
BiVO4 (13), MoS2 (8), ZnO (14), BiO2-­x (15), and CdS (16) have been widely developed.
For example, Wei et al. developed an efficient piezocatalytic system of in situ The authors declare no competing interest.
piezo-­generation and utilization of H2O2 over Fe/BiVO4, significantly promoting the This article is a PNAS Direct Submission.

production of powerful oxidizing hydroxyl radicals (•OH) (17). Lin et al. revealed the Copyright © 2024 the Author(s). Published by PNAS.
This article is distributed under Creative Commons
piezoelectric effect and piezocatalytic mechanism of oxygen-­vacancy-­rich BiO2−x NSs Attribution-­NonCommercial-­NoDerivatives License 4.0
triggered by low frequency ultrasound (15). Particularly, ZnO, as a commonly used pie­ (CC BY-­NC-­ND).
zoelectric material, has found extensive applications in fields such as water splitting and 1
To whom correspondence may be addressed. Email:
lxzeng@jnu.edu.cn or mingyangxing@ecust.edu.cn.
water remediation, due to its excellent piezoelectric properties, availability, high chemical
This article contains supporting information online at
stability, and environmental friendliness (18). However, ensuring efficient piezoelectric https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.​
catalytic performance (such as in situ piezo-­generation and utilization of H2O2 and pol­ 2317435121/-­/DCSupplemental.
lutant removal) in practical applications remains a challenge, primarily due to the rapid Published February 20, 2024.

PNAS 2024 Vol. 121 No. 9 e2317435121 https://doi.org/10.1073/pnas.2317435121 1 of 11


recombination of piezoelectric electron–hole pairs. In order to O (Fig. 1 E and F). Additionally, many small black spots were also
optimize the piezocatalytic performance of ZnO, researchers have observed on other microrods of CZO-­2 (Fig. 1G). The average
explored various strategies, including metal or non-­metal doping, diameter of these black dots was measured to be only ~1.24 nm
defect engineering, and heterostructure construction. For instance, (Fig. 1H), demonstrating that the well-­confined Co species mainly
Wang and Zheng groups have both chosen the approach of doping existed in the form of ultrafine clusters. The actual powder state
metal ions to enhance the piezoelectric effect (19, 20). Wang et al. of these catalysts was depicted in SI Appendix, Fig. S3, with the
demonstrated that V-­doped NaNbO3 single crystals improved the color gradually changing from white to yellow. Furthermore, the
catalytic efficiency via enriched active sites and enhanced polari­ results obtained from inductively coupled plasma atomic emission
zation field under US vibration (21). Additionally, metal deposi­ spectrometry (ICP-­AES) analysis, as shown in SI Appendix,
tion has been widely applied as an effective method to enhance Table S1, revealed an increase in the loaded content of Co from
carrier separation efficiency. For example, Huang et al. developed 0.51 to 0.89% with increasing dosage of cobalt nitrate. This finding
Ag single atoms (SAs) and clusters co-­anchored on carbon nitride provided additional confirmation of the successful loading of Co.
(AgSA+C-­CN) to serve as the multifunctional sites for efficient The crystalline phases of ZO and CZO-­X composite materials were
two-­electron water splitting (22). In addition to enhancing carrier investigated using X-­ray diffraction (XRD) pattern. In Fig. 1J,
separation efficiency, increasing oxygen adsorption to accelerate ZO exhibited three strong diffraction peaks at 31.8°, 34.5°, and
the oxygen reduction reaction (ORR) process is also a key factor 36.3°, which correspond to the (100), (002), and (101) crystal
in improving H2O2 yield. Therefore, introducing defect engineer­ planes of the wurtzite hexagonal ZnO (PDF#75-­0576), while
ing based on the aforementioned regulation strategies, with a focus the XRD pattern of CZO-­X was nearly identical to that of ZO,
on adjusting the electronic structure and enriching active/adsorp­ indicating that the deposition of Co clusters did not alter the
tion sites, becomes crucial. So far, there have been relatively few crystal structure of ZO. It is worth noting that no crystalline phase
studies on ZnO with dynamic oxygen defects in piezocatalysis. of cobalt was detected, likely due to the small quantity and size
Consequently, the combined use of metal deposition and defect of the Co clusters (23). Additionally, the XRD pattern displayed
engineering can significantly promote H2O2 production, while weak diffraction peaks at 33° and 60°, which were ascribed to
activating H2O2 to generate ROS for enhanced pollutant removal the (213) and (503) crystal planes of Zn(OH)2 (PDF#38-­0356).
performance. This approach holds great research significance in This attribution stemmed from the fact that zinc salts inevitably
emerging fields. produced a small amount of Zn(OH)2 during the hydrothermal
Here, we successfully prepared a Co loaded-­ZnO (CZO) pie­ synthesis of ZnO. Furthermore, it is noteworthy that varying
zocatalyst by combining a straightforward hydrothermal process material proportions exhibited similar hydroxide profiles in the
with a photo-­deposition approach. This catalyst exhibited excel­ XRD pattern. Consequently, the predominant composition of
lent capability for refractory pollutants removal and notable the catalyst was oxide crystals, allowing for the disregarding of
non-­sacrificial H2O2 production (pure water). After 3 h of US the impact of this minimal quantity of hydroxides. Raman and
Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

treatment, the removal rate of phenol reached 99% and H2O2 Fourier transform infrared (FT-­IR) spectroscopies were performed
production yield reached 1.3 mM. Experiments and characteri­ to further characterize the structure of ZO and CZO-­2, as shown
zation technique were both employed to explore the mechanism. in SI Appendix, Fig. S4. To investigate the chemical state, we
It was found that CZO piezocatalyst could be used as an excellent examined the X-­ray photoelectron spectra (XPS), where Fig. 1K
heterogeneous catalyst for H2O2 generation and activation attrib­ exhibited the Co 2p XPS spectrum of the CZO-­2 sample. The
uted to the dynamic defects in the process of US. Moreover, the peaks centered at 780.3 and 795.3 eV were assigned to Co2+ of
generation of dynamic defects contributed to the stability and Co 2p3/2 and 2p1/2 (24), respectively, while faint peaks observed
long-­term degradation performance of the catalyst. Thus, a at binding energies of 782.1 and 797.4 eV belonged to Co3+ of
piezocatalysis-­self-­Fenton system has been successfully constructed Co 2p3/2 and 2p1/2 of Co-­O, respectively. These results collectively
over CZO piezocatalyst. Compared to pristine ZnO, the CZO demonstrated the successful introduction of Co and diversity of
piezocatalyst significantly enhanced the H2O2 production and the Co valence.
purification efficiency for refractory pollutants. This study pro­
vides insights into the design of piezocatalysts and their applica­ Piezoelectricity and Electrochemical Analysis of the Catalysts.
tions in energy conversion and environmental remediation. Piezo-­response force microscopy (PFM) was employed to
characterize the piezoelectricity of ZO and CZO-­2 directly. The
Results application of an AC voltage spanning from −10 to 10 V elicited
a localized hysteresis loop (Fig. 2 A and B), where the applied
Characterizations of the Fabricated Catalysts. A straightforward external electric field caused continuous deformation of its
one-­step hydrothermal method was used to prepare ZnO microrods surface, accompanied by a significant phase angle shift of 180°.
(ZO), followed by photo-­deposition to synthesize Co-­ZnO-­X This unequivocally confirmed the inherent piezoelectricity of ZO
composites (CZO-­X, X = 0, 0.5, 2, 4). The scanning electron and CZO-­2 material (25). Furthermore, the maximum effective
microscopy (SEM) and transmission electron microscopy (TEM) piezoelectric coefficient (d33) value based on the amplitude loops
images of ZO and CZO-­X presented similar microrod structure for ZO and CZO-­2 were 41.2 and 58.4 pm/V, respectively,
with an average length exceeding 2 μm, as shown in SI Appendix, which clarified the piezoelectric response of ZO microrods was
Figs. S1 and S2 and Fig. 1 A–I. The magnified TEM image of the enhanced by Co-­deposition (26). Detailed examinations of the
circle highlighted in Fig. 1A and the corresponding high-­resolution PFM morphology, 3D morphology, piezo-­response amplitude, and
TEM (HRTEM) image of ZO (Fig. 1 B and C) displayed ordered phase images were presented in SI Appendix, Figs. S5 A–D and S6
lattice patterns with a spacing of 0.26 nm (002 lattice plane), as A–D. It can be observed that there was a clear contrast in both the
determined by line profile measurements. It could be observed amplitude and phase images of ZO and CZO-­2, demonstrating
from the TEM image in Fig. 1D that numerous black dots were the piezoelectric properties of the materials. Theoretically, the cyclic
embedded in different domains of ZO. The corresponding energy alteration of the electric field direction in piezoelectric materials
dispersive spectroscopy (EDS) elemental mapping and EDS resulting from mechanical vibrations led to a rapid recombination
spectra of CZO-­2 exhibited a uniform distribution of Co, Zn, and of piezoelectric charges, exerting a detrimental impact on the

2 of 11 https://doi.org/10.1073/pnas.2317435121 pnas.org
Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

Fig. 1. (A) TEM image and (B) the corresponding locally enlarged image of ZO; (C) HRTEM image of ZO and the d-­spacing value in the image was measured from
line profile on the Top Right; (D) TEM image, (E) EDS elemental mapping, and (F) EDS spectra of CZO-­2; (G) TEM image and (H) the corresponding locally enlarged
image of CZO-­2; (I) HRTEM image of CZO-­2 and the d-­spacing value in image was measured from line profile on the Top Right; (J) XRD spectra of ZO and CZO-­X;
(K) Co 2p XPS spectrum of CZO-­2.

efficiency of piezocatalysis. Consequently, it became imperative to As depicted in Fig. 2D, the Mott–Schottky plots of ZO exhibited
enhance both the extended lifetime of piezoelectric charges and the a steeper slope compared to CZO-­2 at a frequency of 1 kHz,
prompt charge separation in order to optimize the performance of indicating that the corresponding Nd of ZO (0.10 × 1022 cm−3)
piezoelectric-­driven redox reactions. We observed from SI Appendix, was smaller than that of CZO-­2 (0.17 × 1022 cm−3). This finding
Fig. S7 and Table S2 that moderate Co loading significantly suggested a faster charge transformation process in CZO-­2, align­
extended the charge carrier lifetime as determined through time-­ ing with the above-­mentioned results. Furthermore, a compre­
resolved photoluminescence (PL) attenuation measurements, hensive analysis of the plots enabled the determination of the
with detailed information provided in SI Appendix, Fig. S7. conduction band (CB) positions of the materials under investiga­
Subsequently, the piezo-­generated electrons and holes separation tion. The flat-­band potentials of ZO and CZO-­2 were derived as
efficiency of the samples were investigated using electrochemical approximately −0.93 V and −0.61 V, respectively (vs. saturated
impedance spectroscopy (EIS), as shown in Fig. 2C. The smaller arc calomel electrode, SCE). Taking into account that the flat-­band
radius in the Nyquist plots of EIS for CZO-­X, compared to ZO, potential typically exceeded the conduction band (CB) potential
reflected lower resistance, indicating enhanced charge separation by 0.1 to 0.3 V in n-­type semiconductors (29), the estimated ECB
efficiency and improved charge mobility in the Co and ZO phase. values for ZO and CZO-­2 were −0.79 V and −0.47 V, respectively
Consequently, this enhanced charge separation efficiency was found (vs. normal hydrogen electrode, NHE).
to be highly beneficial for catalytic reactions, similar to the findings
from PL attenuation measurements. Additionally, the charge Piezocatalytic Performance Measurement. Upon ultrasonic
densities (Nd) were calculated by Mott–Schottky measurements (US) activation, CZO-­2 demonstrated exceptional performance
from the following formula: without additional oxidants, achieving an impressive phenol
oxidation rate of 98.7%, surpassing the capabilities of ZO, which
]−1
[1] only achieved a degradation rate of 9.4% (Fig. 3A). Controlled
Nd = 2∕ e0 𝜀0 𝜀 d 1∕C2 ∕dV ,
( )[ ( )
experiments further highlighted the significance of both US
vibration and the catalyst. Without US, CZO-­2 showed a mere
where e0 represents the electron charge and ε0 and ε are the vacuum adsorption rate of 3.3% for phenol, while the degradation rate of
permittivity and dielectric constant of ZnO, respectively (27, 28). phenol without a catalyst under US treatment was 9.7%. These

PNAS 2024 Vol. 121 No. 9 e2317435121 https://doi.org/10.1073/pnas.2317435121 3 of 11


Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

Fig. 2. (A) The butterfly amplitude loop and phase curve of ZO and (B) CZO-­2; (C) Electrochemical impedance spectra (EIS) of the samples; (D) Mott–Schottky
plots of the samples at a frequency of 1 kHz (the Inset shows the electron densities of the samples).

findings underscored the crucial role played by the catalyst and efficient degradation of phenol but also effective degradation of
US activation in facilitating the highly efficient degradation RhB, AO7, MB, NB, AB, 4-­CP, SD, and quinoline (Fig. 3F and
of phenol. As shown in Fig. 3B, the kinetic constants (kobs) of SI Appendix, Table S3), highlighting its wide-­ranging applicability
phenol oxidation by the CZO-­2 were maximized (0.024 min−1) in wastewater treatment.
compared to other materials. Therefore, CZO-­2 was selected as
the optimal catalyst for subsequent trials. Subsequent research Activation of O2 and In-­Situ H2O2 Generation by CZO-­2
assessed the influence of US power and catalyst dosage on phenol Piezocatalyst. As shown in Fig. 4 A–D and SI Appendix, Fig. S9,
removal and investigated the removal efficiency at different phenol the radical quenching experiments involving different scavengers
concentrations, with detailed information provided in SI Appendix, and saturated atmosphere (Ar) were performed to explore the
Fig. S8. As widely recognized, solution pH was also a major factor activation capacity of O2 and the mechanism of H2O2 production
affecting the degradation of pollutants. Within the initial pH in the CZO-­2/US system. When IPA (as a scavenger of •OH,
range of 3 to 9, the degradation rate of phenol was approximately k•OH, IPA = 1.9 × 109 M−1 s−1) (30) was added to the CZO-­2 system
98.0%, with little variation in the kobs value, initially increasing (Fig. 4A), the degradation efficiency of phenol declined from
and then decreasing (Fig. 3C). These results indicated that the 98.7% to 45.1%, which suggested that •OH played a crucial
CZO-­2 piezoelectric system effectively degrades phenol over a role in the oxidation process. With the addition of AgNO3 (as
wide pH range. Besides, tap water, Yangtze River, Huangpu River, a scavenger of e−) and (NH4)2C2O4 (as a scavenger of h+), there
and Pearl River had no significant effect on phenol degradation was a slight inhibition effect on the oxidation of phenol, whereas
(Fig. 3D). More importantly, the CZO-­2 material maintained a significant inhibition of phenol degradation was observed in the
remarkable 95.9% removal efficiency of phenol after four cycles presence of p-­BQ (as a scavenger of •O2−, k •O2–, p-­BQ = 0.9 ×
(Fig. 3E), highlighting its sustained high catalytic activity even 109 M−1 s−1) (31, 32) or FFA (as a scavenger of 1O2, k1O2, FFA =
after repeated reactions. To demonstrate the versatility of the 1.2 × 108 M−1 s−1) (33). Furthermore, phenol oxidation reaction
CZO-­2/US system, we also selected refractory organic pollutants was further inhibited by increasing the concentrations of IPA
other than phenol as target pollutants, including rhodamine B and FFA (SI Appendix, Fig. S9A). As shown in SI Appendix,
(RhB), acid orange 7 (AO7), methylene blue (MB), nitrobenzene Fig. S9B, a substantial inhibition was observed under Ar-­
(NB), aniline (AB), tetrachlorophenol (4-­CP), sulfadiazine (SD), saturated conditions. These results verified the significance of O2
and quinoline. It was observed that CZO-­2 exhibited not only activation and the dominant role played by •OH, •O2− and 1O2

4 of 11 https://doi.org/10.1073/pnas.2317435121 pnas.org
Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

Fig. 3. (A) Phenol removal efficiency in different systems, reaction conditions: [catalyst] = 0.6 g L–1, [phenol]0 = 10 mg/L, and 3 h of US vibration; (B) the
corresponding degradation rate constant of phenol (kobs); (C) effect of initial pH on phenol removal in the CZO-­2 system; (D) the degradation of phenol in real
waters; (E) cycling runs for degradation of phenol in the CZO-­2/US system; (F) degradation of different refractory pollutants by CZO-­2 within 3 h of US.

in phenol oxidation. To clarify the significant impact of 1O2 on in phenol degradation efficiency from 98.7 to 11.4%. Moreover,
phenol degradation, we referred to our prior research showing the liquid chromatography spectrum revealed no significant
the synergistic effect of 1O2 and •OH in facilitating the ring-­ hydroxylation products. Conversely, in the CZO-­2/US system,
opening reaction of phenol (34). In our further investigation, the accumulation of intermediate products occurred rapidly, and
it was observed whether hydroxylation products were present as the US treatment progressed, phenol gradually underwent
during the phenol degradation process in the presence of 1O2 ring opening (SI Appendix, Fig. S10C). This was confirmed by
scavenger. To avoid interference from the position of the peaks in the detection of small molecules such as fumaric acid through
the high-­performance liquid chromatography (HPLC), FFA was HPLC (SI Appendix, Fig. S10D) and the high rate of total organic
replaced with TEMP as 1O2 scavenger. As shown in SI Appendix, carbon (TOC) removal, reaching 50.9% (SI Appendix, Fig. S11).
Fig. S10 A and B, the addition of TEMP resulted in a decrease Furthermore, the HPLC–MS results revealed the presence of

PNAS 2024 Vol. 121 No. 9 e2317435121 https://doi.org/10.1073/pnas.2317435121 5 of 11


Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

Fig. 4. (A) Radical quenching experiments of the oxidation of phenol and (B) concentration changes of in situ piezo-­generated H2O2 with the additions of
different scavengers and different atmospheres (Ar) in the CZO-­2/US system; (C) energy band diagram of ZO and CZO-­2; (D) concentration changes of •OH
with the additions of different scavengers and different atmospheres (Ar) in the CZO-­2/US system; (E) under air and Ar atmosphere, EPR signals for DMPO-­•OH
(water), DMPO -­•O2− (MeOH), TEMP-­1O2 over CZO-­2 powders under US irradiation; (F) in situ H2O2 production comparison of CZO-­2 with some typical reported
piezocatalysts and photocatalysts.

different intermediates and completely decomposed into CO2 and was primarily attributed to the faster charge transfer rate and longer
H2O in the end (detailed in SI Appendix, Fig. S12). Therefore, the charge lifetime resulting from Co-­deposition. Meanwhile, the con­
presence of 1O2 in this system served to promote the hydroxylation centration of H2O2 of the CZO-­2/US system in seawater was also
reaction of phenol, which was further attacked by •OH, leading to detected. As shown in SI Appendix, Fig. S16, the results revealed
ring-­opening. Meanwhile, there was a slight decrease in the peak that the H2O2 yield in seawater was slightly lower compared to that
intensity of phenol, accompanied by the accumulation of a small in pure water, but still reached 0.98 mM. The cumulative •O2− and
amount of intermediate products in the ZO/US system, which •OH production were also quantitatively monitored. Initially, nitro­
was consistent with the relatively low TOC removal rate of only tetrazolium blue chloride (NBT) was employed to measure the
3.7% (SI Appendix, Figs. S11 and S13). concentration of •O2− generated by ZO and CZO-­2 under US
Subsequently, the H2O2 concentration in the CZO-­2/US system irradiation. As depicted in SI Appendix, Fig. S17, the NBT concen­
was also detected in pure water by the iodide method using ammo­ tration gradually decreased in both ZO and CZO-­2 samples, indi­
nium molybdate. As shown in SI Appendix, Fig. S14, the concen­ cating the activation of O2 to produce •O2−. By utilizing the
tration of H2O2 associated with CZO-­2 reached 1.3 mM under stoichiometric parameter that 1 mole of NBT consumption corre­
US irradiation, nearly 6.5-­fold higher than that of ZO. To eliminate sponded to 4 moles of •O2−, the cumulative concentrations of •O2−
the influence of cobalt ions, cobalt ions were introduced into the in ZO and CZO-­2 were determined to be 0.019 and 0.037 mM,
ZO system. Interestingly, it was observed from SI Appendix, Fig. S15 respectively. Additionally, the terephthalic acid photoluminescence
that the concentration of H2O2 did not increase significantly, indi­ (TA-­PL) method was used to record the •OH concentration
cating that the enhanced H2O2 concentration in the CZO-­2 system (SI Appendix, Fig. S18). Similarly, the piezocatalytic production of

6 of 11 https://doi.org/10.1073/pnas.2317435121 pnas.org
•OH over CZO-­2 was largely enhanced compared with ZO, from suggesting both the evolution of 1O2 from •O2− and its direct
0.016 to 0.034 mM. interaction with phenol molecules. Based on these results, we
According to the above analysis, the pathway of CZO-­2 gener­ proposed a mechanism for ROS generation via ultrasound-­driven
ating H2O2 was also investigated. As shown in Fig. 4B, the in situ piezocatalysis (Eqs. 2–6). Additionally, among various piezocata­
generated H2O2 of CZO-­2 with different scavengers and Ar atmos­ lysts and photocatalysts studied, the CZO-­2/US system exhibited
phere was detected in pure water under US irradiation. The con­ the highest production of H2O2, as illustrated in Fig. 4F and
centrations of H2O2 considerably increased from 1.3 mM (control) SI Appendix, Table S5.
to 1.7 mM (IPA), 1.7 mM (FFA), and 1.6 mM [(NH4)2C2O4].
However, the concentrations of H2O2 considerably decreased to Vibration
� e− + h+ ,
CZO − 2 ���������������→ [2]
0.05 mM (AgNO3), 0.02 mM (p-­BQ), and 0.05 mM (Ar-­saturated
condition). These results indicated that H2O2 formation involved
a two-­electron process, and •O2− was an important intermediate O2 + e− → ∙ O2 − , [3]
substance for producing H2O2 (O2→•O2−→H2O2). Additionally,
by utilizing UV-­Vis diffuse reflectance spectroscopy (DRS) esti­ 2 ∙ O2 − + 2H2 O → 2H2 O2 + 2OH− +1 O2 [4]
mation in SI Appendix, Fig. S19, the optical bandgaps for ZO and
CZO-­2 were determined to be 3.09 eV and 2.93 eV, respectively, [5]
and when combined with the analysis of Mott–Schottky measure­ H2 O2 + Co2+ → Co3+ + ∙ OH + OH− ,
ment, the corresponding energy band diagrams were illustrated in
Fig. 4C. The suitable CB and valence band (VB) position were H2 O + h+ → ∙ OH + H+ . [6]
favorable for the oxygen reduction reaction (ORR), thereby initi­
ating the cascade redox reaction. Specifically, piezoelectric electrons
reacted with absorbed O2 to produce •O2−, and the intermediate Investigation on the Mechanism of Enhanced H2O2 Concen­
•O2− reacted with piezoelectric electrons to generate H2O2, which tration. It was noteworthy that the concentration of H2O2
could be further decomposed to •OH. The corresponding changes generated during the US exposure was re-­evaluated using the
in •OH radical were in accordance with the abovementioned results CZO-­2 sample after four cycles. Interestingly, there was a
(Fig. 4D). Furthermore, it was observed from SI Appendix, Fig. S20 noticeable increase in the H2O2 concentration, as depicted in
that while the CZO-­2 catalyst could decompose H2O2 under stir­ Fig. 5A, with concentration rising from 1.3 mM to 1.8 mM. To
ring conditions, it was unable to generate •OH. Moreover, in the investigate this phenomenon, we initially examined the surface
absence of a catalyst, neither US nor stirring could decompose vacancies of different catalysts using EPR characterization. As
H2O2, indicating that both US and catalyst were indispensable for shown in Fig. 5B, ZO and CZO-­2 exhibited minimal vacancies
the generation of •OH in this work. signal prior to the US reaction. However, following the US
Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

Additionally, to assess whether Co2+ was involved in the acti­ reaction, ZO displayed a weak oxygen defects signal peak at g =
vation process of H2O2, we introduced Fe2+ into the CZO-­2/US 2.004. In contrast, CZO-­2 exhibited a significantly enhanced
system, competing with Co2+ in decomposing H2O2. Due to the oxygen defects signal. Apart from the EPR technique, we employed
faster rate of Fe2+ in decomposing H2O2 compared to Co2+, the X-­ray photoelectron spectroscopy (XPS) to examine the alteration
introduction of Fe2+ served as a means to explore whether Co2+ in the surface defect ratio of the material before and after the
was involved in the activation process of H2O2 during the US reaction. Fig. 5C and SI Appendix, Table S6 revealed the O 1 s
process. If the change in the contents of cobalt ion was more XPS spectrum, which exhibited a splitting into two or three peaks:
obvious in the absence of Fe2+ compared to when Fe2+ was added, lattice oxygen was bonded to metal (Olatt, 529.5 to 530.6 eV),
it confirmed the participation of Co2+ in the activation process of oxygen deficiency (Odef, 531.0 to 531.8 eV), and the surface
H2O2 in the CZO-­2/US system. Thus, from the Co 2p XPS data, adsorbed molecule H2O (Osurf, 532. 8 eV) (35). The relative ratio
it was observed that, compared to the pure CZO-­2 catalyst, the of oxygen deficiency in ZO exhibited no significant change before
post-­reaction Co2+ content decreased from 50.2 to 39.2%, while [Odef/(Odef + Olatt) = 4.4%] and after reaction [Odef/(Odef + Olatt)
the Co3+ content increased from 49.8 to 60.8% (SI Appendix, = 3.9%], which was consistent with the previous results.
Fig. S21A and Table S4). However, when we added Fe2+ into the Conversely, the relative ratio of Odef of CZO-­2 remarkably
CZO-­2/US system (SI Appendix, Fig. S21B and Table S4), it was increased from 6.4 to 29.1% after reaction, and the surface oxygen
found that, in comparison with the pure CZO-­2 catalyst, the content dwindled from 62.4 to 51.0%, providing strong evidence
alterations in the contents of Co2+(49.9%) and Co3+(50.1%) after for the occurrence of dynamic oxygen vacancies during the CZO-­2
the reaction were not more obvious than that without adding Fe2+, US stress. Thus, we speculated that the formation of dynamic
demonstrating the involvement of Co2+ in the activation of H2O2. oxygen defects of CZO-­2 material during the US process might
These results suggested that •OH derived from H2O2 activation have been either due to the deposition of cobalt, which enhanced
(primary) and H2O oxidation (secondary) through the following piezoelectric activity, activating lattice oxygen by more piezoelectric
route: O2 → •O2− → H2O2 → •OH and H2O → •OH. electrons during migration, ultimately leading to the formation
In addition to chemical probes, the main active species that of oxygen defects, or due to the catalyst generating additional heat
contributed to the excellent performance of the phenol oxidation on the surface under US conditions, promoting oxygen overflow.
was also confirmed by the EPR characterization, as shown in Therefore, we conducted an experiment in which we subjected
Fig. 4E. Under air atmosphere, CZO-­2 piezocatalysis yielded EPR the CZO-­2 catalyst to a 3-­h heat treatment at a temperature
quadruplet DMPO-­•OH peaks, sextuplet DMPO-­•O2− peaks, equivalent to that generated under ultrasound. The results from
and triplet TEMP-­1O2 peaks with US 10 min, whereas no clear Fig. 5D showed that no defect signal was observed in the CZO-­2
signal of ROS was observed without US irradiation. Under Ar after the heat treatment. Following this, we introduced AgNO3
atmosphere, weak •OH signal was observed, consistent with the into the reaction process and observed a substantial decrease in
previous findings, and the signal of •O2− and 1O2 were negligible. the defect signal by examining the EPR signal of the catalyst after
Additionally, upon the addition of p-­BQ or phenol in the presence reaction. This finding served as strong evidence for the pivotal role
of US irradiation, the peak intensity of 1O2 was weakened, of piezoelectric electrons in the formation of dynamic defects.

PNAS 2024 Vol. 121 No. 9 e2317435121 https://doi.org/10.1073/pnas.2317435121 7 of 11


Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

Fig. 5. (A) Time dependence for yield of CZO-­2 and fourth-­CZO-­2 under US irradiation; (B) EPR spectra of the samples; (C) O 1 s XPS spectrum of ZO and CZO-­2
fresh and after the reaction; (D) EPR spectra of the samples; (E) formation energy of oxygen vacancy over ZO-­OV, CZO-­OV, and CZO-­OV-­adding electrons; (F) the
charge density difference of the O2 molecule on CZO with different oxygen vacancy concentrations (the yellow electronic cloud represents the loss of electrons,
and the blue electronic cloud represents the gain of electrons; the isosurface is set to 0.0005 eV).

Furthermore, the formation energy of a single oxygen defect over conditions of continuous defect generation, the post-­reaction
CZO was studied by DFT calculations, as shown in Fig. 5E. catalyst was stored in ambient air for a period of time. It was found
Results indicated that oxygen vacancy on surface of CZO (EOV = from Fig. 5D that the defect signal of first-­CZO-­2-­used-­storage
3.49 eV) was obviously lower than that of ZO (EOV = 3.94 eV), noticeably weakened. This indicated that the CZO-­2 catalyst
suggesting that the oxygen vacancies more easily formed on the possessed self-­repair capabilities, allowing it to maintain out­
surface of CZO due to cobalt deposition. Additionally, by standing stability during multiple cyclic tests. Additionally, since
introducing additional electrons into the CZO-­OV model to the production of H2O2 was derived from the ability of the catalyst
bolster the reducing atmosphere, a noteworthy reduction in defect to adsorb and activate dissolved oxygen, we initiated the process
formation energy was observed, decreasing from 3.49 eV to 2.25 by increasing aeration, thereby maximizing the concentration of
eV. This compelling finding unequivocally established the intricate dissolved oxygen. The results revealed that, after 1 h of aeration,
connection between defect formation and electron density. Thus, the H2O2 concentration increased from 1.3 mM to 1.6 mM
these results indicated that the formation of dynamic defects was (SI Appendix, Fig. S22). Subsequently, to gain a deeper under­
not caused by the thermal effects during the US process but rather standing of the interaction between O2 molecules and piezocatalysts
by the activation of lattice oxygen through piezoelectric electrons. with different concentrations of oxygen defects at a molecular
Additionally, we could further observe from Fig. 5B that after level, DFT simulation has been employed (36, 37). Based on
undergoing four cycles, the catalyst exhibited a stronger defect experimental characterization, we theoretically modeled the
signal. Simultaneously, the in situ generated H2O2 concentration optimal geometric structure of CZO (002) and confirmed the
also increased further, indicating a positive correlation between optimal adsorption site of the O2 molecules (SI Appendix,
the amount of H2O2 and the defect concentration. To investigate Fig. S23). As shown in Fig. 5F, CZO samples with varying oxygen
how the catalyst maintained excellent piezoelectric activity under vacancies (from OV0 to OV3) exhibited strong O2 molecule

8 of 11 https://doi.org/10.1073/pnas.2317435121 pnas.org
adsorption with a negative adsorption energy ranging from remarkable self-­healing capabilities. Next, the phenol degradation
approximately −3.05 eV to −3.24 eV, accompanied by a charge performance and H2O2 generation ability of the catalyst were
transfer ranging from 0.645 e− to 0.657 e−, indicating that the tested after it had been placed for 2 mo and 5 mo. As depicted in
more defect sites exposed on the catalyst surface, more electron Fig. 6B, it was observed that even after 3 h of US irradiation,
transfer occurs from ZO-­OV to Co clusters, leading to electron-­ phenol degradation was nearly complete. Moreover, the concen­
enriched Co clusters and thereby promoting the adsorption and tration of H2O2 remained at 1.2 mM and 1.0 mM in pure water,
activation of the O2 molecule. Meanwhile, SI Appendix, Figs. S21 respectively, indicating that the capability of the catalyst to degrade
and S24 revealed the electron cloud density of Co increased, while phenol and produce H2O2 did not significantly decline over the
that of Zn decreased in CZO-­2 after the reaction (detailed extended period. Additionally, based on the ion dissolution results
information was shown in SI Appendix, Fig. S24 and Tables S4 obtained from ICP-­AES analysis (Fig. 6C), it could be observed
and S7), indicating that the large number of defects as electron that, with prolonged US time in the CZO-­2/US system, the con­
traps produced by CZO-­2 in the piezoelectric reaction process, centrations of cobalt ions and zinc ions increased from 0.58 mg/L
effectively impeding the recombination of both piezoelectric holes to 0.75 mg/L and from 4.2 mg/L to 7.4 mg/L, respectively.
and electrons and promoting the charge transfer of ZO to Co Subsequently, after the reaction solution underwent treatment
clusters by the Zn-­O-­Co electron transfer chain, thus obtaining with a cation exchange resin, the concentrations of cobalt ions
electron rich structure. Thus, we speculated that Co clusters acted and zinc ions significantly decreased to 0.085 mg/L and 0.06
as active sites for electron enrichment, facilitating the activation mg/L, respectively, which was far lower than the emission standard
of O2 to produce H2O2 and then be further activated to •OH. and effectively prevented secondary pollution. Simultaneously,
Besides, we introduced iron into ZnO (Fe-­ZnO) using the same the pH of the reaction solution after 3 h of ultrasound was adjusted
photo-­deposition method, and its XRD spectrum closely to 10 with NaOH. Following a subsequent aging period, the dis­
resembled that of CZO-­2 (SI Appendix, Fig. S25A), displaying solution concentrations of cobalt and zinc ions decreased signifi­
prominent characteristic peaks of ZnO. Moreover, Fe-­ZnO cantly, from 0.75 mg/L to less than 0.01 mg/L and 7.4 mg/L to
exhibited excellent piezoelectric catalytic performance, with a 0.019 mg/L, respectively. The dissolved metal ions formed hydrox­
phenol degradation rate of 87% and a H2O2 production rate of ide precipitates, allowing for their effective reuse. Consequently,
1.46 mM after US treatment for 3 h (SI Appendix, Fig. S25B). To it was observed that H2O2 yield increased with prolonged US
exclude the potential interference of Fe3+ produced during the US exposure in pure water within the CZO-­2/US system (Fig. 6D).
process to the detection of H2O2 by iodometry, we mixed the However, after further extension of the time, the H2O2 production
reaction solution obtained after 3 h of ultrasound with an equal rate eventually fell below the H2O2 decomposition rate, resulting
amount of potassium thiocyanate (10 mg mL−1). The UV-­visible in a gradual decrease of H2O2 yield. Nevertheless, adjusting the
spectrum analysis of Fe3+ content in the solution, as shown in pH of the above solution to 10 and allowing for an aging period
SI Appendix, Fig. S25C, revealed the absence of Fe3+. Simult­ led to an increase in H2O2 yield again, which demonstrated that
Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

aneously, the leaching of 0.068 mg L−1 Fe from the Fe-­ZnO the deactivated catalyst due to metal ion dissolution could be
solution after 3 h of US was analyzed using ICP-­AES. Assuming rejuvenated by pH adjustment, ensuring the long-­term stability
this was equivalent to the concentration of Fe3+, it was observed for in situ H2O2 production of the CZO-­2/US system.
that the reaction of Fe3+ with potassium iodide at this concentration Furthermore, long-­term continuous degradation of phenol was
did not exhibit a peak at 351 nm (SI Appendix, Fig. S25D). examined under US exposure (Fig. 6E). The performance was
Consequently, in the Fe-­ZnO/US system, the iodometric detection evaluated after 60 h of US treatment with eight rounds of phenol
of H2O2 concentration was deemed reasonable. As a result, similar supplementation. Although the degradation rate gradually slowed
results could be achieved by substituting cobalt with alternative down, CZO-­2 maintained a high phenol degradation efficiency
metals through the application of the same photo-­deposition of 92%, showcasing its remarkable stability.
method, thereby overcoming the inherent constraints associated
with this approach. Conclusions
In order to study the effect of background material and poten­
tial application of CZO-­2 material, the impact of typical anions Piezocatalysis proved to be an exceptionally attractive and effective
(Cl−, NO3−, HCO3−, and SO42−) and dissolved organic matter technology in the field of wastewater treatment. However, the
(DOM) on phenol degradation was investigated separately. As major obstacles encountered in the development of piezocatalytic
depicted in Fig. 6A, in the presence of Cl−, NO3−, HCO3−, SO42−, reaction primarily revolved around the dual aspects of H2O2 gen­
and DOM, the oxidation efficiencies of phenol were 88.6%, eration and activation without introducing additional oxidants,
97.5%, 97.0%, 91.5%, and 91.4%, respectively (in 180 min). which were indispensable for producing sustainable ROS capable
Although the degradation efficiency was not significantly affected, of achieving thorough purification of wastewater. In this study,
the degradation rate exhibited sensitivity, particularly in the pres­ the CZO piezocatalysts prepared by the hydrothermal method
ence of Cl− ions. This was because Cl− could induce a competing combined with the photo-­deposition method demonstrate excep­
reaction with phenol oxidation by utilizing •OH. Material stabil­ tional efficacy in wastewater treatment and in situ generated
ity was also an important indicator for evaluating potential appli­ H2O2, attributed to fast charge separation efficiency and extended
cations of catalysts, we performed comprehensive investigations, charge lifetime brought by Co-­deposition and the existence of
including characterizing the catalyst after reaction, assessing the dynamic defects, as confirmed by EPR and XPS analysis. More
degradation efficiency of catalysts stored for varying time dura­ importantly, the catalysts placed for longer periods (over 2 mo)
tions, evaluating their capacity for in situ H2O2 generation, and continue to perform superior H2O2 production (~1.2 mM) and
examining their long-­term pollutant degradation capabilities. contaminant removal capabilities. This observation not only
First, the SEM images (SI Appendix, Fig. S26 A and B), XRD underscores the substantial potential of CZO piezocatalysts in
pattern (SI Appendix, Fig. S26C), and FT-­IR spectra (SI Appendix, practical wastewater treatment applications but also serves as a
Fig. S26D) of CZO-­2 after reaction were nearly identical to those reference for future studies on improving H2O2 production by
of the initial sample, indicating that CZO-­2 demonstrated robust facilitating the generation of dynamic vacancies through metal
structural stability following the reaction, attributable to its deposition.

PNAS 2024 Vol. 121 No. 9 e2317435121 https://doi.org/10.1073/pnas.2317435121 9 of 11


Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

Fig. 6. (A) Removal efficiencies and the corresponding kobs value of phenol oxidation (US for 3 h) in the presence of different water quality constituents [(anion) =
10 mg L−1, (DOM) = 10 mg L−1]; (B) under US irradiation, time dependence for phenol removal efficiency and H2O2 yield of CZO-­2 after 2 and 5 mo of placement; (C)
variation of ion concentration in the CZO-­2/US (0.6 g L–1) system (treatment-­1: cation-­exchange resin was used for adsorption of dissolved metal ions; treatment-­2:
adjusting pH = 10 with NaOH); (D) time dependence for H2O2 yield of CZO-­2. (E) Time dependence for phenol removal efficiency of CZO-­2.

Materials and Methods To prepare Co-­ZnO-­X composites (where X represents the concentra-
tion of Co(NO3)2·6H2O, X = 0, 0.5, 2, 4 mM), a light deposition method
Preparation of Catalysts. The ZnO microrods were synthesized using a hydro- was employed. A mixture containing 0.2 g ZO, varying amounts of
thermal method. Specifically, a mixture of 10 mmol Zn(NO3)2·6H2O, 10 mmol Co(NO3)2·6H2O, and 100 mL of ultrapure water was stirred under xenon
urotropin, and 48 mL of ultrapure water was stirred until homogenous. The result- lamp irradiation for 20 h. Subsequently, the solution was filtered, washed,
ing solution was transferred to a 100-­mL hydrothermal reactor and maintained and dried at 60 °C in a vacuum oven, resulting in the formation of Co-­ZnO-­0
at 95 °C for 24 h. After completion, the solution was filtered, washed, and dried (CZO-­0), Co-­ZnO-­0.5 (CZO-­0.5), Co-­ZnO-­2 (CZO-­2), and Co-­ZnO-­4 (CZO-­4)
at 60 °C to obtain purified ZnO microrods (referred to as ZO). composites.

10 of 11 https://doi.org/10.1073/pnas.2317435121 pnas.org
Iodometric Method for Detecting H2O2. The concentration of H2O2 produced Full details for DFT computations, control experiments, and materials character-
was detected by the iodometric method according to the literature following izations were introduced in SI Appendix.
the steps below: Ammonium molybdate tetrahydrate (H24Mo7N6O24·4H2O) was
dissolved in 5 mL of deionized water to prepare a 0.01 M solution, while potas- Data, Materials, and Software Availability. All study data are included in the
sium iodide (KI) was dissolved in 10 mL of deionized water to prepare a 0.1 article and/or SI Appendix.
M KI solution. During different intervals of ultrasonic irradiation, 0.5 mL of the
reaction solution was collected and mixed with 2 mL of the KI solution, along ACKNOWLEDGMENTS. This work was supported by the National Natural Science
with 50 μL of ammonium molybdate tetrahydrate solution. The resulting mixture Foundation of China (No. 22325602, 22276071, 52000101, 22176060, and
was vigorously shaken, allowed to stand for 10 min, and subsequently analyzed 22076064) and sponsored by Program of Shanghai Academic/Technology Research
using a UV-­Vis spectrophotometer at a maximum absorption wavelength of 351 Leader (23XD1421000). Project supported by Shanghai Municipal Science and
nm. The concentration of H2O2 was determined using a standard curve method. Technology Major Project (Grant No. 2018SHZDZX03), the Program of Introducing
Talents of Discipline to Universities (B16017), and the Science and Technology
Commission of Shanghai Municipality (20DZ2250400). Authors thank Research
Degradation of Organic Pollutants. The degradation efficiency of different pollut-
Center of Analysis and Test of East China University of Science and Technology for
ants [including phenol, nitrobenzene (NB), aniline (AB), sulfadiazine (SD), p-­chloro-
the help on the characterization.
phenol (4-­CP), quinoline, rhodamine B (RhB), methylene blue (MB), and acid orange
7 (AO7)] was assessed by determining their concentrations using HPLC and UV-­Vis
spectrophotometer. At specified time intervals, a 1 mL aliquot of the sample was taken
Author affiliations: aKey Laboratory for Advanced Materials and Joint International
and passed through a 0.22-­μm membrane to remove solid catalyst. The injection vol- Research Laboratory of Precision Chemistry and Molecular Engineering, Feringa Nobel
ume for HPLC analysis was 20 μL. The flow rate was set at 1 mL min−1, and the column Prize Scientist Joint Research Center, Frontiers Science Center for Materiobiology and
Dynamic Chemistry, School of Chemistry and Molecular Engineering, East China University
temperature was maintained at 30 °C. The specific HPLC methods used for analyzing of Science and Technology, Shanghai 200237, People’s Republic of China; bGuangdong
different pollutants could be found in SI Appendix, Table S3. Additionally, the solution Key Laboratory of Environmental Pollution and Health, School of Environment, Jinan
samples were analyzed for total organic carbon (TOC) content using a TOC analyzer. University, Guangzhou 511443, People’s Republic of China; cKey Laboratory for Advanced
Materials and Institute of Fine Chemicals, School of Resources and Environmental
Engineering, East China University of Science and Technology, Shanghai 200237, People’s
Statistical Analysis. All data were gained directly from the source experiment Republic of China; and dState Key Laboratory of Chemical Engineering, East China
and processed using Origin. All experiments were carried out in duplicate. University of Science and Technology, Shanghai 200237, People’s Republic of China

1. S. Li et al., Incorporating oxygen atoms in a SnS2 atomic layer to simultaneously stabilize atomic 20. B. Li et al., Ultrasound-­remote selected activation mitophagy for precise treatment of rheumatoid
hydrogen and accelerate the generation of hydroxyl radicals for water decontamination. Environ. Sci. arthritis by two-­dimensional piezoelectric nanosheets. ACS Nano 17, 621–635 (2022).
Technol. 56, 4980–4987 (2022). 21. Y. Li et al., Robust route to H2O2 and H2 via intermediate water splitting enabled by capitalizing on
2. Y. Zhang et al., Highly efficient removal of U (VI) by the photoreduction of SnO2/CdCO3/CdS minimum vanadium-­doped piezocatalysts. Nano Res. 15, 7986–7993 (2022).
nanocomposite under visible light irradiation. Appl. Catal. B 279, 119390 (2020). 22. C. Hu et al., Orthogonal charge transfer by precise positioning of silver single atoms and clusters
3. S. Zhou et al., Self-­regulating solar steam generators enable volatile organic compound removal on carbon nitride for efficient piezocatalytic pure water splitting. Angew. Chem. Int. Ed. 61,
Downloaded from https://www.pnas.org by 171.252.154.204 on March 29, 2024 from IP address 171.252.154.204.

through in situ H2O2 generation. Environ. Sci. Technol. 56, 10474–10482 (2022). e202212397 (2022).
4. Y. Chen et al., Selective recovery of precious metals through photocatalysis. Nat. Sustain. 4, 618–626 23. J. Chen et al., Nickel clusters accelerating hierarchical zinc indium sulfide nanoflowers for
(2021). unprecedented visible-­light hydrogen production. J. Colloid Interface Sci. 608, 504–512 (2022).
5. J. Ji et al., Defects on CoS2−x: Tuning redox reactions for sustainable degradation of organic 24. Q. Yan et al., Constructing an acidic microenvironment by MoS2 in heterogeneous Fenton reaction
pollutants. Angew. Chem. Int. Ed. 133, 2939–2944 (2021). for pollutant control. Angew. Chem. Int. Ed. 60, 17155–17163 (2021).
6. Q. Chen, Development of an anthraquinone process for the production of hydrogen peroxide 25. H. You et al., Harvesting the vibration energy of BiFeO3 nanosheets for hydrogen evolution. Angew.
in a trickle bed reactor-­From bench scale to industrial scale. Chem. Eng. Process. 47, 787–792 Chem. Int. Ed. 131, 11905–11910 (2019).
(2008). 26. C. Zhang et al., Piezo-­photocatalysis over metal-­organic frameworks: Promoting photocatalytic
7. Y. Zhang et al., H2O2 generation from O2 and H2O on a near-­infrared absorbing porphyrin activity by piezoelectric effect. Adv. Mater. 33, 2106308 (2021).
supramolecular photocatalyst. Nat. Energy 8, 361–371 (2023). 27. S.-­H. Chiou et al., Plasmonic gold nanoplates-­decorated ZnO branched nanorods@ TiO2 nanorods
8. W. Liu et al., Efficient hydrogen production from wastewater remediation by piezoelectricity heterostructure photoanode for efficient photoelectrochemical water splitting. J. Photochem.
coupling advanced oxidation processes. Proc. Natl. Acad. Sci. U.S.A. 120, e2218813120 (2023). Photobiol. A 443, 114816 (2023).
9. M. Ran et al., Selective production of CO from organic pollutants by coupling piezocatalysis and 28. C. Tiwari et al., Precursor mediated and defect engineered ZnO nanostructures using thermal
advanced oxidation processes. Angew. Chem. Int. Ed. 135, e202303728 (2023). chemical vapor deposition for green light emission. Thin Solid Films 762, 139539 (2022).
10. H. Huang et al., Macroscopic polarization enhancement promoting photo-­and piezoelectric-­ 29. J. Zheng et al., Incorporation of CoO nanoparticles in 3D marigold flower-­like hierarchical
induced charge separation and molecular oxygen activation. Angew. Chem. Int. Ed. 56, architecture MnCo2O4 for highly boosting solar light photo-­oxidation and reduction ability.
11860–11864 (2017). Appl. Catal. B 237, 1–8 (2018).
11. P. Zhu et al., Piezocatalytic tumor therapy by ultrasound-­triggered and BaTiO3-­mediated 30. E. J. Rosenfeldt et al., Degradation of endocrine disrupting chemicals bisphenol A, ethinyl estradiol,
piezoelectricity. Adv. Mater. 32, 2001976 (2020). and estradiol during UV photolysis and advanced oxidation processes. Environ. Sci. Technol. 38,
12. C. Hu et al., Exceptional cocatalyst-­free photo-­enhanced piezocatalytic hydrogen evolution of 5476–5483 (2004).
carbon nitride nanosheets from strong in-­plane polarization. Adv. Mater. 33, 2101751 (2021). 31. Y. Bu et al., Peroxydisulfate activation and singlet oxygen generation by oxygen vacancy for
13. F. Wang et al., Unveiling the effect of crystal facets on piezo-­photocatalytic activity of BiVO4. degradation of contaminants. Environ. Sci. Technol. 55, 2110–2120 (2021).
Nano Energy 101, 107573 (2022). 32. X. Chen et al., Photocatalytic free radical-­controlled synthesis of high-­performance single-­atom
14. Q. T. Hoang et al., Piezoelectric Au-­decorated ZnO nanorods: Ultrasound-­triggered generation of catalysts. Angew. Chem. Int. Ed. 62, e202312734 (2023).
ROS for piezocatalytic cancer therapy. Chem. Eng. J. 435, 135039 (2022). 33. C. Qi et al., Activation of peroxymonosulfate by base: Implications for the degradation of organic
15. L. Yang et al., Oxygen-­vacancy-­rich piezoelectric BiO2−x nanosheets for augmented piezocatalytic, pollutants. Chemosphere 151, 280–288 (2016).
sonothermal and enzymatic therapies. Adv. Mater. 35, 2300648 (2023). 34. H. Yu et al., Singlet oxygen synergistic surface-­adsorbed hydroxyl radicals for phenol degradation in
16. Q. Zhao et al., The collaborative mechanism of surface S-­vacancies and piezoelectric polarization for CoP catalytic photo-­Fenton. Chinese J. Catal. 43, 2678–2689 (2022).
boosting CdS photoelectrochemical performance. Chem. Eng. J. 433, 133226 (2022). 35. W. Wang et al., The confined interlayer growth of ultrathin two-­dimensional Fe3O4 nanosheets with
17. Y. Wei et al., In-­situ utilization of piezo-­generated hydrogen peroxide for efficient p-­chlorophenol enriched oxygen vacancies for peroxymonosulfate activation. ACS Catal. 11, 11256–11265 (2021).
degradation by Fe loading bismuth vanadate. Appl. Surf. Sci. 543, 148791 (2021). 36. S. Wang et al., Targeted NO oxidation and synchronous NO2 inhibition via oriented 1O2 formation
18. J. Ma et al., High efficiency bi-­harvesting light/vibration energy using piezoelectric zinc oxide based on lewis acid site adjustment. Environ. Sci. Technol. 57, 12890–12900 (2023).
nanorods for dye decomposition. Nano Energy 62, 376–383 (2019). 37. J. Cao et al., Tailoring the asymmetric structure of NH2-­UiO-­66 metal-­organic frameworks for
19. S. Cheng et al., The highly effective therapy of ovarian cancer by Bismuth-­doped oxygen-­deficient light-­promoted selective and efficient gold extraction and separation. Angew. Chem. Int. Ed. 62,
BaTiO3 with enhanced sono-­piezocatalytic effects. Chem. Eng. J. 442, 136380 (2022). e202302202 (2023).

PNAS 2024 Vol. 121 No. 9 e2317435121 https://doi.org/10.1073/pnas.2317435121 11 of 11

You might also like