Cancer nanotechnology is a promising area of cross-disciplinary research aiming to develop facile, effective, and noninvasive strategies to improve cancer diagnosis and treatment. Catalytic therapy based on exogenous stimulus-responsive semiconductor nanomaterials has shown its potential to address the challenges under the most global medical needs. Semiconductor nanocatalytic therapy is usually triggered by the catalytic action of hot electrons and holes during local redox reactions within the tumor, which represent the response of nontoxic semiconductor nanocatalysts to pertinent internal or external stimuli. However, careful architecture design of semiconductor nanocatalysts has been the major focus since the catalytic efficiency is often limited by facile hot electron/hole recombination. Addressing these challenges is vital for the progress of cancer catalytic therapy. In recent years, diverse strategies have been developed, with heterojunctions emerging as a prominent and extensively explored method. The efficiency of charge separation under exogenous stimulation can be heightened by manipulating the semiconducting performance of materials through heterojunction structures, thereby enhancing catalytic capabilities. This review summarizes the recent applications of exogenous stimulus-responsive semiconducting nanoheterojunctions for cancer theranostics. The first part of the review outlines the construction of different heterojunction types. The next section summarizes recent designs, properties, and catalytic mechanisms of various semiconductor heterojunctions in tumor therapy. The review concludes by discussing the challenges and providing insights into their prospects within this dynamic and continuously evolving field of research.

Catalysis is a vital process of biochemical reactions, playing a pivotal role in orchestrating numerous metabolic pathways and physiological functions within our bodies with remarkable precision and efficiency.1,2 In catalysis, a key player, a catalytic agent or catalyst, accelerates a chemical reaction by lowering the activation energy without being used itself. This efficiency is crucial for various biochemical processes, including metabolism, digestion, energy production, detoxification, DNA maintenance, and cellular signaling.3 Catalytic medicine is an exciting new field utilizing this concept for therapeutic purpose. Instead of traditional drugs, it uses nontoxic artificial catalysts, especially metal-containing or metal based nanoparticles (NPs) known as nanozymes, to treat diseases.4–7 These catalysts work by accelerating desired reactions at the disease site, producing beneficial molecules that promote healing. Scientists are interested in this new approach to fighting cancer using specially designed entities called “catalytic nanomedicines.” These entities target unique chemical features found within tumors, possibly improving treatment and reducing drug resistance. Our bodies naturally produce powerful oxidizing molecules such as hydroxyl radical (•OH) through hydrogen peroxide (H2O2) breakdown catalyzed by metal ions. Similarly, peroxidase (POD)-like nanozymes were used to create •OH radicals from endogenously overexpressed H2O2 in tumor microenvironment (TME) to kill cancer cells.8 Based on their architecture and properties, nanozymes show different catalytic activities, such as POD-, catalase (CAT)-, superoxide dismutase (SOD)-, glucose oxidase (GOx)-, and glutathione peroxidase (GPx)-like catalytic activities.9 However, this self-sustained activity faces challenges related to targeting and specificity issues, as nanozymes can cause off-target effects and toxicity to other organs, underscoring the importance of precise targeting strategies and thorough biocompatibility assessments for their successful implementation in clinical settings.

Catalytic therapy can be further enhanced by introducing external triggers, such as light, temperature, ultrasound (US), and magnetic force, a strategy called exogenous stimulation therapy. This approach offers several advantages compared to endogenous therapy. Unlike relying solely on the body's natural catalysts, exogenous stimulation therapy utilizes various catalytic substrates, increasing both efficiency and effectiveness. The success of this approach depends on the carefully chosen catalyst and its unique physical and chemical properties. Selecting the right catalyst ensures optimal performance and maximizes the therapeutic potential of exogenous stimulation therapy. The recent surge in research has focused on the synergy between specific semiconductors and external energy sources, such as light, US, temperature, and magnetic force. This powerful combination is promising for generating clean energy sources and decontaminating water. A key discovery by researchers K. Honda and A. Fujishima substantially advanced the photocatalysis field (light as the stimuli) in 1972.10 They demonstrated that titanium dioxide (TiO2) could split water molecules when exposed to light. This breakthrough laid the foundation for further exploration of photocatalysis. Meanwhile, sonocatalysis and radiocatalysis utilize sound or radio waves, respectively, to activate semiconducting materials as catalysts. When the catalysts are exposed to these external stimuli with energy higher than their band gaps (hν ≥ Eg), an electron (e) in the catalyst valence band (VB) gets promoted to the conduction band (CB). This creates a hole (h+) in the VB, forming an electron–hole pair known as an exciton. These excited electrons and holes are the key players responsible for various chemical reactions. However, a major obstacle in optimizing these catalytic processes is the tendency of excitons to recombine. This recombination, where the excited electron and hole reunite, releases energy and reduces overall efficiency. Therefore, strategies to minimize electron–hole recombination and extend exciton lifetime are crucial for maximizing the performance of these technologies. Following the discovery of interparticle electron transfer by Serpone et al. in 1984, pairing semiconductors became a popular strategy in photo-, sono-, and radiocatalysis. This approach suppresses charge recombination by facilitating electron transfer between materials. A decade later, the confirmation of interparticle hole transfer further advanced the research on integrating metal NPs or forming heterojunctions with semiconductors to improve catalytic efficiency.11 

Cancer remains a global issue, demanding innovative and well-tolerated treatment.12 However, effective, traditional methods, such as surgery, chemotherapy, and radiation therapy, often inflict significant side effects or limitations on patients.13 A new therapeutic strategy for cancer treatment based on the in situ production of therapeutic compounds or molecules within the tumor, which can effectively kill cancer cells or impair their proliferation, is under development.14 Thus, the strategy minimizes the use of invasive treatment and toxic agents and reduces their accumulation in normal cells. Unlike traditional metal catalysts, semiconductor catalysis uses the power of excitons—electron–hole pairs generated upon light or energy absorption.15 These excitons play a central role in driving various redox reactions at the surface of the catalyst. Electrons and holes readily participate in redox reactions with surrounding molecules. In cancer cells, endogenously reduced substrates, such as intracellular overexpressed glutathione (GSH), can be oxidized into GSSG (≥0.24 V for GSH to GSSG) by holes, while the generated H+ can be converted into molecular hydrogen (H2) (H+/H2 = 0 V vs normal hydrogen electrode [NHE], pH = 0) by hot electrons generated on the surface of the catalyst.16 The generated H2 gas can induce cancer cell apoptosis by diffusing into mitochondria, impairing their function, and hindering adenosine triphosphate (ATP) synthesis.17 These agents can be converted into oxidized products by generated holes, consequently modulating the TME. Simultaneously, high generated electron levels can reduce oxygen (O2) and carbon dioxide (CO2) to produce cytotoxic superoxide (•O2) and carbon monoxide (CO), respectively.18,19 However, single-component catalysts often face challenges because of the tendency of rapid exciton recombination, which significantly reduces their overall catalytic activity.20 Fortunately, researchers identified a promising solution: the formation of heterojunctions. These junctions, where two different semiconductors are joined, are at the forefront of developing highly efficient catalysts.21 Heterojunctions not only address the recombination issue but also offer an additional benefit of enhanced selectivity. The specific redox potentials of excitons can be tuned by engineering different types of heterojunctions, allowing more precise control over the desired chemical reactions. Heterojunction nanomedicine could be a promising effective cancer treatment. This approach addresses two key challenges in traditional photocatalysis: minimizing exciton recombination and enhancing selectivity. Creating heterojunctions can significantly improve the utilization of external energy, leading to more efficient generation of reactive species. Additionally, these junctions can provide multiple active sites with distinct properties, allowing better simultaneous control of targeted reactions within the TME.22 This targeted therapy control can potentially enhance the overall efficacy and safety of catalytic cancer therapy (Scheme 1). Moreover, metal particles commonly employed in semiconductor catalysts can serve as theranostic agents, particularly in photoacoustic (PA) imaging, computed tomography (CT) imaging, and photothermal imaging. Consequently, this review explores the basic working principles of semiconductor heterojunction catalysts. It clarifies the latest progress in designing and utilizing heterojunctions for tumor theranostics applications. Then, the review examines strategies to optimize existing heterojunctions and explores their potential for future therapeutic approaches beyond cancer treatment. It discusses the exciting prospects for integrating heterojunctions into different biomedical fields. Ultimately, this review summarizes and discusses the reported literature on exogenously activated semiconducting nano-heterojunctions in the field of cancer medicine, aiming to inspire further innovation and promote the development of heterojunctions as powerful tools for advancing biomedicine.

SCHEME 1.

A schematic representation of the classification, stimuli, and modes of cancer treatment using heterojunction nanocatalysts.

SCHEME 1.

A schematic representation of the classification, stimuli, and modes of cancer treatment using heterojunction nanocatalysts.

Close modal

Photocatalytic reactions occur on semiconductor surfaces, utilizing excited electrons and holes as catalysts for the reduction and oxidation, respectively, of target molecules. Generally, when a semiconductor receives light irradiation with energy equal to or higher than its bandgap, electrons are excited from the VB to the CB, leaving holes behind in the VB. These high-energy charge carriers on the catalyst surface can drive various redox reactions depending on the VB and CB potentials. However, in single-phase semiconductors, excited electrons are prone to recombine with holes in the VB, thereby deteriorating their catalytic power for photocatalytic applications. Several techniques in structural modification/rearrangement, including doping with impurities, heterojunction formation, defect engineering, and crystal facet engineering, were developed to address this challenge. Among them, heterojunction formation is a promising approach to enhancing photocatalytic activity.23 Due to band bending that occurs at the junction of two semiconductors/components, a difference in the potential develops between the two semiconductor regions. This induces a built-in electric field in the space charge zone and plays a crucial role in facilitating the spatial separation of photogenerated excitons, known as heterojunction.24–26 Heterojunction construction involves a combination of a semiconductor with another material, such as another semiconductor (S-S), metal (S-M), or carbonaceous material. The Fermi levels of heterostructure components align, causing band bending that hinders electron backflow and extends the lifetime of charge carriers (Fig. 1).

FIG. 1.

Schematic of the charge transfer mechanisms in different heterojunctions adopted in the photo/sonocatalytic process. Four types (a) type I, (b) type II, (c) Z-scheme, and (d) Schottky junction are included. CB—conduction band, VB—valence band, and Ef—Fermi level.

FIG. 1.

Schematic of the charge transfer mechanisms in different heterojunctions adopted in the photo/sonocatalytic process. Four types (a) type I, (b) type II, (c) Z-scheme, and (d) Schottky junction are included. CB—conduction band, VB—valence band, and Ef—Fermi level.

Close modal

However, the rapid recombination of these photogenerated electron–hole pairs often limits their efficiency when using a single-phase semiconductor as a photocatalyst. Thus, researchers explored various strategies, such as element doping, compositing, and crystal facet engineering, to overcome this limitation and enhance photocatalytic activity. One of the most promising approaches involves the creation of heterojunctions by using a semiconductor with another semiconductor or a metal or a carbonaceous compound to decrease the electron–hole recombination. After hybridization, a new electronic structure is formed by the combination of materials with dissimilar band gaps. Semiconductor heterojunctions are classified into two main categories: S-S (semiconductor-semiconductor) and S-M (semiconductor-metal), which are differentiated based on the inherent electrical properties of the materials involved.

S-S heterojunctions are developed at the contact interface of two semiconductor materials, each with unique electronic properties and energy band structure. The unique chemical and physical properties of such materials are developed through doping or surface functionalization to form various intrinsic n- or p-type semiconductors.27–29 Generally, the alignment of electronic band edges and work functions of two semiconductors are the key factors regulating the charge transfer mechanism at the heterojunction interface, directly impacting the catalytic performance of materials. Based on the charge transfer mechanism, S-S heterojunctions can be classified into types I, II, and III or Z-scheme heterojunctions.30 

In type I heterojunction with straddling band structure, the energy band alignment of two catalytic materials is completely nested within each other.31 As shown in Fig. 1(a), both the VB and CB of the narrow-bandgap catalyst (referred to as semiconductor II) reside within the forbidden zone of the wideband gap catalyst (referred to as semiconductor I). When exposed to sufficient energy, holes and electrons transfer from semiconductor I to semiconductor II due to a more negative CB and a more positive VB of semiconductor I.32 Thus, type I heterojunction, in theory, does not lead to enhanced charge carrier separation and does not improve photocatalytic activity. In a type II heterojunction catalytic system comprising two semiconductors, the CB position of one semiconductor (referred to as semiconductor I) is more negative than that of the second one [referred to as semiconductor II; Fig. 1(b)], while semiconductor II has a more positive VB position. When these two semiconductors are in contact, their Fermi levels rearrange, causing semiconductor I to donate electrons to semiconductor II. Therefore, the electrons and holes transfer in opposite directions, leading to improved charge carrier separation, reduced recombination probability, increased charge carrier lifetimes, and improved photocatalytic activity.33 During light irradiation, excited charge carriers are accumulated at the VB and CB of catalysts I and II, respectively. Compared to type I heterojunctions, type II heterojunctions trap excited charge carriers in two distinct spatial potential wells.34 This configuration improves catalytic activity by enhancing charge carrier longevity at the junction interface.

Z-scheme or type III heterojunctions exhibit a unique energy band alignment where the conduction band minimum (CBM) of one catalyst (referred to as semiconductor II) directly aligns with the valence band maximum (VBM) of the other catalyst (referred to as semiconductor I), resulting in a Z-shaped band structure [Fig. 1(c)]. This structure enables effective charge separation, enhanced redox reactions, and improved catalytic performance. In 1979, Bard introduced the concept of a liquid-phase Z-scheme heterojunction as a photocatalyst.35 This heterojunction catalytic system generally comprises two semiconductor components, where one acts as a catalytic center and the other serves as a redox mediator.36 However, liquid phase Z-scheme heterojunction catalysts face challenges, such as lower stability, limited application range due to decreased electron transfer efficiency, synthesis complexity, contamination risks, potential efficiency concerns, and material compatibility limitations. In 2001, Grätzel first adopted a direct Z-scheme mechanism to explain the photocatalytic activity enhancement in the WO3/TiO2 heterojunction. Direct Z-scheme heterojunctions are designed semiconductor interfaces that directly transfer photogenerated charge carriers (electrons and holes) between two distinct materials. This eliminates the need for an external electron mediator while maintaining the catalytic activity of both materials. It involves two semiconductors with staggered energy gaps, designated as semiconductor II (higher work function) and semiconductor I (lower work function).37,38 Following the contact of two semiconductors, electrons migrate from semiconductor II to semiconductor I until equilibrium in the Fermi levels is reached. Consequently, the energy bands of semiconductor II bend upward, while the bands of semiconductor I bend downward. This bending allows the excited holes of semiconductor II and the excited electrons of semiconductor I to be combined through the tunneling effect.39 Additionally, the internal electric field caused by Fermi-level rearrangements in semiconductors I and II minimizes the charge recombination and maximizes the utilization of excited charge carriers in redox reactions. Despite wide applicability, direct Z-scheme heterojunctions encounter challenges in selecting semiconductor materials with matched band alignment and establishing efficient charge transfer pathways at the heterojunction interface to facilitate direct Z-scheme formation over type II heterojunctions.40–42 Researchers address direct Z-scheme limitations by incorporating charge transfer bridges within heterostructures, enhancing efficiency, and directing interfacial charge flow.42 These charge transfer bridges are developed by incorporating noble metal NPs, such as Ag, Au, and Pd.43–45 These bridges not only modify the band alignment but also sustain the rapid charge transfer process.46 

The oldest practical semiconductor devices use S-M heterojunctions created by combining semiconductors with metals. S-M heterojunctions are categorized in Schottky junctions [Fig. 1(d)] and Ohmic contacts and depend on semiconductor type and their relative work function. Schottky junctions are formed at the interface of metals and semiconductors, with the metal having a higher work function than the semiconductor (n-type) or the semiconductor having a higher work function than the metal (p-type). Due to the difference in the work function between metals and semiconductors, Schottky junctions effectively trap electrons and promote the separation of excited holes and electrons.47 Ohmic junctions occur when semiconductors with different work function values are connected to metals. In contrast to Schottky junctions, the band edges of semiconductors bend downward, allowing excited electrons to flow back to the VB and causing recombination. Hence, Ohmic junctions are not appropriate for catalysis.

1. Photocatalysis

Photocatalysis is a process of utilizing light energy to initiate catalytic reactions by rearranging the electronic band structure of semiconductors. The excitation causing electron transition from VB to CB can be achieved by utilizing light energy equivalent to or greater than the bandgap energy of the semiconducting material. Thus, the generated hot electron and hole pairs facilitate semiconducting nanomaterials to engage in redox reactions at the catalyst surface. Semiconducting materials act as reducing agents when their CB potential is between +0.5 and −1.5 V (vs hydrogen electrode). VB holes demonstrate oxidizing properties when their oxidative potential is between +1.0 and +3.5 V (vs hydrogen electrode).48 For an effective photocatalytic reaction with minimized recombination of the charged species, the most common strategy adopted by the researchers is the inclusion of sacrificial agents. The sacrificial agents will capture the holes, enhancing the electron concentration for prolonged and efficient reduction reactions. Meanwhile, the holes oxidize the sacrificial agents and can be converted to value-added products. Researchers have employed similar strategies for treating tumors in recent years by utilizing photocatalysts and different reductive chemical factors in the tumor microenvironment. For example, Yang et al. used intracellularly accumulated lactic acid as a hole sacrificial agent in photocatalytic H2 production for NIR-induced cancer photoimmunotherapy.49 Meanwhile, Chen et al. utilized glucose as the sacrificial agent to generate H2 inside the diseased site.50 Similarly, several other intratumoral/cellular substrate molecules, such as OH (E[OH/•OH] = 1.99 eV vs NHE), water (E[H2O/•OH] = 2.73 eV vs NHE), and H2O2 (E[H2O2/•OH] = 0.32 eV NHE), have been explored to produce •OH radicals, which is highly reactive and toxic. O2 molecules present inside the tumor can act as a substrate to form •O2 by accepting the electrons from the CB of the semiconducting material (E[O2 /•O2] = −0.13 V vs NHE). The generated •O2, as the primary reactive oxygen species (ROS), can be further converted into other more potent ROS (such as H2O2 and •OH) through various approaches (•O2+ e + 2H+ →2H2O2; •O2 + H2O2 → •OH + OH + O2), causing substantial damage to cancer cells.51–54 Meanwhile, generated holes can oxidize H2O2 (E[H2O2/O2] = 0.69 V vs NHE) and water to O2 to relieve hypoxia, as well as GSH to GSSG (E[GSH/GSSG] = 0.24 V vs NHE) to facilitate the destruction of the redox TME.55 Therefore, photoactive nanomaterials with more negative CB (<−0.13 V) and more positive VB (>1.23 V) can produce ROS and O2 that are beneficial for efficient photodynamic therapy (PDT) in hypoxia [Fig. 2(a)]. Semiconducting materials with bandgap <1.53 eV can be excited with 808 nm near-infrared (NIR) laser irradiation.22,56 In response to the limitations imposed by rapid electron–hole recombination in single-component (unitary) semiconductors, researchers have been exploring heterojunction semiconductors as promising candidates in photocatalysis and, thus, a potential solution. Pan et al. put forward the initial proposal for the concept of heterojunction nanomedicine and its future.22 However, despite these advancements, several challenges in the exploration of heterojunction photocatalysts for tumor therapy remain.

FIG. 2.

(a) Possible photocatalytic redox reactions to produce/increase ROS for oxidative damage on cancer cells. (b) Schematic representation of Z‐scheme Bi2S3@Bi heterostructure and the possible mechanism underlying the generation of O2 and ROS under NIR laser irradiation at 808 nm. Reproduced with the permission from Cheng et al., Adv. Mater. 32(11), 1908109 (2020). Copyright 2020 John Wiley and Sons.57 (c) Mechanism of Bi2S3-x-Au@HA (hyaluronic acid) for photocatalytic antitumor reactions against 4T1 tumor cells under 1064 nm NIR-II laser irradiation. Reproduced with the permission from Meng et al., J. Colloid Interface Sci. 644, 437–453 (2023). Copyright 2023 Elsevier.58 (d) The charge transfer mechanism and the generation of ROS in Bi-HA (humic acid)/FK866 heterojunction under NIR laser for effectively inhibiting NAD/ERK/NF-κB signal pathway and cancer cell migration. Reproduced with the permission from Song et al., Biomaterials 254, 120140 (2020). Copyright 2020 Elsevier.59 (e) The overall mechanisms of ROS generation and GSH depletion in Z-scheme BP@CPP heterojunctions for enhanced lipid peroxidation and amplified ferroptotic cell death. Reproduced with the permission from Liu et al., Small 20, 2309206 (2023). Copyright 2023 John Wiley and Sons.60 

FIG. 2.

(a) Possible photocatalytic redox reactions to produce/increase ROS for oxidative damage on cancer cells. (b) Schematic representation of Z‐scheme Bi2S3@Bi heterostructure and the possible mechanism underlying the generation of O2 and ROS under NIR laser irradiation at 808 nm. Reproduced with the permission from Cheng et al., Adv. Mater. 32(11), 1908109 (2020). Copyright 2020 John Wiley and Sons.57 (c) Mechanism of Bi2S3-x-Au@HA (hyaluronic acid) for photocatalytic antitumor reactions against 4T1 tumor cells under 1064 nm NIR-II laser irradiation. Reproduced with the permission from Meng et al., J. Colloid Interface Sci. 644, 437–453 (2023). Copyright 2023 Elsevier.58 (d) The charge transfer mechanism and the generation of ROS in Bi-HA (humic acid)/FK866 heterojunction under NIR laser for effectively inhibiting NAD/ERK/NF-κB signal pathway and cancer cell migration. Reproduced with the permission from Song et al., Biomaterials 254, 120140 (2020). Copyright 2020 Elsevier.59 (e) The overall mechanisms of ROS generation and GSH depletion in Z-scheme BP@CPP heterojunctions for enhanced lipid peroxidation and amplified ferroptotic cell death. Reproduced with the permission from Liu et al., Small 20, 2309206 (2023). Copyright 2023 John Wiley and Sons.60 

Close modal

Plants utilize solar light (photosensitization) to generate and separate charges during photosynthesis, ultimately driving catalytic reactions producing O2 from H2O. Inspired by natural photosynthesis, researchers attempted to develop artificial photosynthetic materials that can capture light, split charges, and drive valuable catalytic reactions. These artificial systems aim to mimic not only the remarkable efficiency of natural photosynthesis for O2 production from abundant H2O but also develop different materials for possible catalytic reactions, such as H2 and O2 evolution, water purification, and pollutant degradation. Inspired by solar-enabled water-splitting for clean and sustainable energy, Zheng et al. developed a 2D metal-free C3N4 nanomaterial as a promising water-splitting system for tumor treatment for the first time with carbon dots and protoporphyrin IX to construct a heterostructure with enhanced red-light absorption. The RGD-conjugated nanosheet heterojunction was substantially accumulated in αvβ3-integrin-overexpressed 4T1 breast cancer cells. Under 630 nm light irradiation, the heterostructures accumulating within the tumor could split intratumoral H2O and generate sufficient O2 to reduce the expression of hypoxia-associated proteins, such as hypoxia-inducible factor-1α and carbonic anhydrase 9. O2 could also be partially converted to cytotoxic 1O2 with the photosensitizer, protoporphyrin IX, for effective photodynamic therapy (PDT). Overall, both PDT and hypoxia relief significantly improve the prognosis for treating hypoxic tumors.61 Due to the matter of concern posed by tissue penetration depth from light absorption and scattering in tissues, the delivery of visible and ultraviolet (UV) light is constrained. The penetration is dependent on light wavelength and is generally between 0.5 and 2.5 mm in healthy skin tissues making the dose availability for photosensitizer activation following tissue penetration limited.62 Thus, high-intensity UV and visible light irradiation cannot be employed for biomedical application because of the resulting tissue damage. Some other electromagnetic waves, such as NIR and x-rays, have good tissue penetration and are preferred for light-activated therapeutics.63 In Sec. II, we have discussed the structural design of various heterojunctions to enhance the separation of photogenerated electron–hole pairs. By carefully aligning the energy levels of two materials within a heterojunction, it becomes easier to direct electrons and holes into separate regions. This minimizes recombination and enhances photocatalytic efficiency, an essential factor for effective tumor treatment. Several synthesis methods have been developed to engineer these heterojunction catalysts (Table I). One widely explored method is surface oxidation, which involves oxidizing the surface of a material to form an oxide layer. This oxide layer can then interact with the core material to create a heterojunction. Surface oxidation improves adhesion between the two materials, leading to a more stable and uniform heterojunction, but also serves as a bonding interface that reduces defects and enhances the overall quality of heterojunctions. A key example is the development of surface-oxidized arsenene nanosheets (As/AsxOy NSs) with a type II heterojunction structure, used as imaging-guided, noninvasive nanomedicines for cancer therapy. Under 808 nm irradiation, photoexcited electrons in the CB of arsenene transfer to the CB of arsenic oxide (AsxOy), catalyzing the formation of ·O2 from intratumoral O2. At the same time, holes in the VB of AsxOy transfer to the VB of As, facilitating the oxidation of intratumoral GSH. The heterostructure was further modified with polydopamine and a cancer cell membrane coating to enhance its biocompatibility and targeting ability. The cancer cell membrane provides a biomimetic camouflage that allows for specific recognition of cancer cells.64 The natural process of photosynthesis has evolved on Earth for billions of years, and chlorophyll molecules are the most abundant photosynthetic pigments indispensable for natural photosynthesis. Ou et al. explored a novel strategy to develop an effective photocatalysis-mediated tumor therapy using a type II heterojunction engineered with black phosphorus nanosheets and photosynthetic system of chlorophyll cells (Chls), named Chl@BP-Fe. BP nanosheets modified with polyaspartic acid (PASP) were loaded onto the surface of Chls through the chelation reaction between Fe3+ and PASP. When the heterojunction was exposed to NIR laser at 658 nm, the electrons in the CB of Chls were preferably transferred to the CB of BP nanosheets. Meanwhile, the holes in the VB of BP nanosheets tended to migrate to the VB of Chls. The transfer in both electrons and holes facilitated the segregation of photoexcited charges into the CB and VB of two different photosensitizers, fully preventing the recombination of electron–hole pairs. The Fe3+ present inside the heterostructure could both deplete GSH and catalyze the Fenton reaction to generate •OH, increasing therapeutic efficacy. Chls can also serve as an immune adjuvant by triggering T cell activation and maturation of dendritic cells, thereby facilitating immunotherapy.65 This study opens new avenues for researchers to explore biological and natural materials in creating novel heterojunctions.

TABLE I.

List of different stimuli responsive semiconductor heterojunction nanocatalysts for ROS catalysis in tumor therapy.

Materials Synthesis method ROS species Band gap Stimulus used Modes of treatment Year
Photo activated heterojunctions  p-MoS2/n-rGO–MnO2–PEG  Liquid exfoliation, Hydrothermal  1O2, •OH, •O2  NA  NIR (980 nm)  PDT  2018137  
NIR-CD/MoS2  Hydrothermal  NA  1.9 eV  NIR (808 nm)  PTT  2019138  
CeVO4/Ag  Hydrothermal, Surface reduction of Ag+ by Ce3+ •O2  3.07 eV  NIR (808 nm)  PTT, PDT  2019139  
TOPY-PEG NSs  Ball grinding, Liquid exfoliation, Thermal oxidation  •O2, •OH  FeS2: 0.9 eV
Fe2O3: 2.1 eV 
NIR (808 nm)  PTT, PDT, CDT  201967  
NdVO4/Au  Hydrothermal, NaBH4 reduction  •O2, •OH  NA  NIR (808 nm)  PTT, PDT  2019140  
g-C3N4-AuNP  Calcination, Photodeposition  •O2, •OH  NA  NIR (670 nm)  PDT  2019141  
Bi2Se3/MoSe2/Bi2Se3 @PEG-Dox  Liquid exfoliation, cation-exchange  •OH  1.17 eV  NIR (808 nm)  PTT, PDT, CDT, Chemotherapy  2019142  
Bi–Bi2S3/BSA&FA  NaBH4 reduction  •OH  NA  NIR (808 nm)  PTT, CT imaging  2019143  
CeVO4/Au  Hydrothermal, NaBH4 reduction  •O2, •OH  NA  NIR (808 nm)  PTT, PDT  2020144  
FeTiO3@Fe2O3  Liquid exfoliation  •O2, •OH  NA  NIR (650 and 808 nm)  PTT, PDT  2020145  
g-C3N4/Cu3 Calcination  •O2  NA  NIR (980 nm)  PTT, PDT, CDT  2020146  
MoSe2/CoSe2@PEGs  Coprecipitation  H2O2, •O2, •OH  1.3 eV  NIR (808 nm)  PTT, CDT  2020147  
NaErF4@ZnO  Hydrothermal, high temperature annealing  1O2  NA  NIR (980 nm)  PTT, PDT, CDT  2020148  
Bi2S3@Bi  Solvothermal, hydrazine treatment  •O2, •OH  NA  NIR (808 nm)  PDT  202057  
Cu2-xSe-Au  Oxidation-reduction  •OH  NA  NIR (808 nm)  PTT, PDT  2020149  
BiNS–Fe@Fe  Solvothermal  •O2, •OH  NA  NIR (808 nm)  PTT, PDT, CDT  2021150  
Sb–THPP–PEG  Liquid exfoliation  1O2, •O2  1.75 eV  NIR (808 nm)  PTT, PDT  2020151  
As/AsxOy@PDA@M  Ball grinding, liquid exfoliation  1O2, •O2  NA  NIR (808 nm)  PTT, PDT, CDT  202164  
Bi2Se3/Au @PLGA-PEG-DOX  Hydrothermal, oxidation-reduction  1O2, •O2, •OH  NA  NIR (808 nm)  PTT, PCT, chemotherapy  2021152  
M-RP/BP@ZnFe2O4  Hydrothermal, liquid exfoliation  1O2, •O2, •OH  NA  NIR (660 nm)  PDT  2021153  
Mo2C@N-Carbon@PEG  Calcination  •OH  ∼2.10  NIR (808 nm)  PTT, PDT,  2021154  
PTh@Au NCs  Oxidation-reduction  1O2, •O2, •OH  1.80 eV  NIR (650 nm)  PDT  2021155  
Bi/BiOx  Regioselective coordination, orientational oxidation  1O2, •O2, •OH  1.83 eV  NIR (660 nm)  PTT, PDT  2021156  
Cu2−xSe/Bi2Se3@PEG  Thermal injection, cation-exchange  1O2, •O2, •OH  NA  (NIR 808 nm)  PTT. PDT, CDT  2021157  
CNMS  Hydrothermal, calcination  •O2, •OH    (NIR 670 nm)  PDT  2021158  
MoSe2/Au@PEG  Hydrothermal, photodeposition  •OH  1.33 eV  NIR (808 nm)  PDT, PDT, CDT  2022159  
g-C3N4/rGO/ZnO-Ag  Hydrothermal, calcination  •O2, •OH  2.6 eV  Visible-light  PDT  2022160  
CeO2@MXne  Liquid exfoliation, hydrothermal  1O2, •O2, •OH  NA  NIR (808 nm)  PDT  2024161  
Ultrasound activated heterojunctions  HAu-TiO2  Photoreduction  1O2, •OH  NA  US (1.5 MHz)  SDT  2016162  
Au-TiO2-A-TPP  Hydrothermal, NaBH4 reduction  1O2, •OH  2.90 eV  US (1.0 MHz and 1.5 W cm−2 SDT  2019163  
Fe-TiO2  Thermal decomposition  1O2, •OH  2.3 eV  US (40 kHz, 3 W/cm2 SDT, CDT  202080  
Pt–TiO2  Hydrothermal, vacuum metal sputter deposition  1O2, •O2, •OH  NA  1.0 MHz, 1.5 W cm-2  SDT, Chemotherapy  202096  
TiH1.924  Liquid exfoliation  •O2,1O2, •OH  2.7 eV  40 kHz, 3 W cm−2  SDT PTT  2020164  
TiN  Liquid exfoliation  1O2, •O2, •OH  NA  US (40 kHz, 3.0 W/cm2 SDT, PTT  2021165  
α-Fe2O3@Pt  Hydrothermal, NaBH4 reduction  1O2  1.83 eV  US (1.0 MHZ, 1.0 W cm−2 SDT  202185  
Cu2−xO–BaTiO3  Hydrothermal, electrostatic interaction, calcination  1O2, •OH  NA  1.0 MHz, 1.0 W/cm2  CDT, SDT  2022166  
N-CD@LiFePO4  Microwave-assisted hydrothermal, ultrafast microwave reaction  1O2, •OH  NA  US irradiation (50 kHz, 3 Wcm−2 SDT, CDT  202294  
SMISO  Hydrothermal  1O2, •OH  NA  US (1.0 MHz, 1 W/cm2 SDT  2023167  
Ru/TiO2−x@TiCN  Liquid exfoliation, hydrothermal, NaBH4 reduction  1O2, •OH  1.8 eV  US (50 kHz and 1 W/cm2 SDT  202492  
x-ray activated heterojunctions  BiOI@Bi2S3@BSA  Hydrothermal  •O2, •OH  1.63 eV  x-ray (6 Gy), and NIR (808 nm)  RT, PDT, PTT  2017114  
Bi2WO6  Hydrothermal  •OH  ∼2.62 eV  x-ray (6 Gy)  RT  2019111  
Au–Bi2S3  Precipitation, Reduction  •OH  NA  x-ray (6 Gy)  RT  2019113  
WO2.9-WSe2  Hydrothermal, Selenization  •OH  NA  x-ray (4 Gy)  RT, PTT, CBT  2020168  
WO3@Ag2WO4@CS  Hydrothermal, deposition-precipitation  •OH  NA  x-ray (6 Gy)  RT  2020110  
BiOI/Bi2S3@polydopamine  Anion-exchange  •OH  NA  x-ray (6 Gy)  RCT  2021120  
Thermoelectric heterojunctions  BST/BTS  Hydrothermal  •O2, •OH  NA  ΔT  TET  2022128  
SrTiO3/Cu2Se  Hydrothermal  •O2, •OH    ΔT  TET  2023135  
BST/CaO2  Hydrothermal, oxidation-reduction  •O2, •OH  NA  ΔT  TET  2023136  
Materials Synthesis method ROS species Band gap Stimulus used Modes of treatment Year
Photo activated heterojunctions  p-MoS2/n-rGO–MnO2–PEG  Liquid exfoliation, Hydrothermal  1O2, •OH, •O2  NA  NIR (980 nm)  PDT  2018137  
NIR-CD/MoS2  Hydrothermal  NA  1.9 eV  NIR (808 nm)  PTT  2019138  
CeVO4/Ag  Hydrothermal, Surface reduction of Ag+ by Ce3+ •O2  3.07 eV  NIR (808 nm)  PTT, PDT  2019139  
TOPY-PEG NSs  Ball grinding, Liquid exfoliation, Thermal oxidation  •O2, •OH  FeS2: 0.9 eV
Fe2O3: 2.1 eV 
NIR (808 nm)  PTT, PDT, CDT  201967  
NdVO4/Au  Hydrothermal, NaBH4 reduction  •O2, •OH  NA  NIR (808 nm)  PTT, PDT  2019140  
g-C3N4-AuNP  Calcination, Photodeposition  •O2, •OH  NA  NIR (670 nm)  PDT  2019141  
Bi2Se3/MoSe2/Bi2Se3 @PEG-Dox  Liquid exfoliation, cation-exchange  •OH  1.17 eV  NIR (808 nm)  PTT, PDT, CDT, Chemotherapy  2019142  
Bi–Bi2S3/BSA&FA  NaBH4 reduction  •OH  NA  NIR (808 nm)  PTT, CT imaging  2019143  
CeVO4/Au  Hydrothermal, NaBH4 reduction  •O2, •OH  NA  NIR (808 nm)  PTT, PDT  2020144  
FeTiO3@Fe2O3  Liquid exfoliation  •O2, •OH  NA  NIR (650 and 808 nm)  PTT, PDT  2020145  
g-C3N4/Cu3 Calcination  •O2  NA  NIR (980 nm)  PTT, PDT, CDT  2020146  
MoSe2/CoSe2@PEGs  Coprecipitation  H2O2, •O2, •OH  1.3 eV  NIR (808 nm)  PTT, CDT  2020147  
NaErF4@ZnO  Hydrothermal, high temperature annealing  1O2  NA  NIR (980 nm)  PTT, PDT, CDT  2020148  
Bi2S3@Bi  Solvothermal, hydrazine treatment  •O2, •OH  NA  NIR (808 nm)  PDT  202057  
Cu2-xSe-Au  Oxidation-reduction  •OH  NA  NIR (808 nm)  PTT, PDT  2020149  
BiNS–Fe@Fe  Solvothermal  •O2, •OH  NA  NIR (808 nm)  PTT, PDT, CDT  2021150  
Sb–THPP–PEG  Liquid exfoliation  1O2, •O2  1.75 eV  NIR (808 nm)  PTT, PDT  2020151  
As/AsxOy@PDA@M  Ball grinding, liquid exfoliation  1O2, •O2  NA  NIR (808 nm)  PTT, PDT, CDT  202164  
Bi2Se3/Au @PLGA-PEG-DOX  Hydrothermal, oxidation-reduction  1O2, •O2, •OH  NA  NIR (808 nm)  PTT, PCT, chemotherapy  2021152  
M-RP/BP@ZnFe2O4  Hydrothermal, liquid exfoliation  1O2, •O2, •OH  NA  NIR (660 nm)  PDT  2021153  
Mo2C@N-Carbon@PEG  Calcination  •OH  ∼2.10  NIR (808 nm)  PTT, PDT,  2021154  
PTh@Au NCs  Oxidation-reduction  1O2, •O2, •OH  1.80 eV  NIR (650 nm)  PDT  2021155  
Bi/BiOx  Regioselective coordination, orientational oxidation  1O2, •O2, •OH  1.83 eV  NIR (660 nm)  PTT, PDT  2021156  
Cu2−xSe/Bi2Se3@PEG  Thermal injection, cation-exchange  1O2, •O2, •OH  NA  (NIR 808 nm)  PTT. PDT, CDT  2021157  
CNMS  Hydrothermal, calcination  •O2, •OH    (NIR 670 nm)  PDT  2021158  
MoSe2/Au@PEG  Hydrothermal, photodeposition  •OH  1.33 eV  NIR (808 nm)  PDT, PDT, CDT  2022159  
g-C3N4/rGO/ZnO-Ag  Hydrothermal, calcination  •O2, •OH  2.6 eV  Visible-light  PDT  2022160  
CeO2@MXne  Liquid exfoliation, hydrothermal  1O2, •O2, •OH  NA  NIR (808 nm)  PDT  2024161  
Ultrasound activated heterojunctions  HAu-TiO2  Photoreduction  1O2, •OH  NA  US (1.5 MHz)  SDT  2016162  
Au-TiO2-A-TPP  Hydrothermal, NaBH4 reduction  1O2, •OH  2.90 eV  US (1.0 MHz and 1.5 W cm−2 SDT  2019163  
Fe-TiO2  Thermal decomposition  1O2, •OH  2.3 eV  US (40 kHz, 3 W/cm2 SDT, CDT  202080  
Pt–TiO2  Hydrothermal, vacuum metal sputter deposition  1O2, •O2, •OH  NA  1.0 MHz, 1.5 W cm-2  SDT, Chemotherapy  202096  
TiH1.924  Liquid exfoliation  •O2,1O2, •OH  2.7 eV  40 kHz, 3 W cm−2  SDT PTT  2020164  
TiN  Liquid exfoliation  1O2, •O2, •OH  NA  US (40 kHz, 3.0 W/cm2 SDT, PTT  2021165  
α-Fe2O3@Pt  Hydrothermal, NaBH4 reduction  1O2  1.83 eV  US (1.0 MHZ, 1.0 W cm−2 SDT  202185  
Cu2−xO–BaTiO3  Hydrothermal, electrostatic interaction, calcination  1O2, •OH  NA  1.0 MHz, 1.0 W/cm2  CDT, SDT  2022166  
N-CD@LiFePO4  Microwave-assisted hydrothermal, ultrafast microwave reaction  1O2, •OH  NA  US irradiation (50 kHz, 3 Wcm−2 SDT, CDT  202294  
SMISO  Hydrothermal  1O2, •OH  NA  US (1.0 MHz, 1 W/cm2 SDT  2023167  
Ru/TiO2−x@TiCN  Liquid exfoliation, hydrothermal, NaBH4 reduction  1O2, •OH  1.8 eV  US (50 kHz and 1 W/cm2 SDT  202492  
x-ray activated heterojunctions  BiOI@Bi2S3@BSA  Hydrothermal  •O2, •OH  1.63 eV  x-ray (6 Gy), and NIR (808 nm)  RT, PDT, PTT  2017114  
Bi2WO6  Hydrothermal  •OH  ∼2.62 eV  x-ray (6 Gy)  RT  2019111  
Au–Bi2S3  Precipitation, Reduction  •OH  NA  x-ray (6 Gy)  RT  2019113  
WO2.9-WSe2  Hydrothermal, Selenization  •OH  NA  x-ray (4 Gy)  RT, PTT, CBT  2020168  
WO3@Ag2WO4@CS  Hydrothermal, deposition-precipitation  •OH  NA  x-ray (6 Gy)  RT  2020110  
BiOI/Bi2S3@polydopamine  Anion-exchange  •OH  NA  x-ray (6 Gy)  RCT  2021120  
Thermoelectric heterojunctions  BST/BTS  Hydrothermal  •O2, •OH  NA  ΔT  TET  2022128  
SrTiO3/Cu2Se  Hydrothermal  •O2, •OH    ΔT  TET  2023135  
BST/CaO2  Hydrothermal, oxidation-reduction  •O2, •OH  NA  ΔT  TET  2023136  

As explained in Sec. II, in type II heterojunctions, the CB and VB of two semiconductors are staggered. This alignment in type II structure often results in reduced redox potentials because the electrons and holes are at lower energy levels after separation, which may limit the range of reactions they can effectively drive. Rather than creating type II heterojunctions, various approaches have been employed to enhance the production of ROS using semiconductor materials. In a Z-scheme heterojunction, the resulting electrons and holes possess higher redox potentials due to the electron transfer between the two semiconductors. This is crucial to driving more demanding reactions, like water splitting, where strong oxidizing and reducing agents are needed. Several approaches have been employed to construct a successful Z-scheme photocatalytic system, including those by ensuring proper band alignment, utilizing appropriate preparation methods, and achieving effective cooperation among various materials.66 For example, Pan et al. in 2019 reported the first direct Z-scheme heterojunction using thermally oxidized ultrathin pyrite nanosheets comprising FeS2 as cores and Fe2O3 shells. These nanosheets were synthesized through mechanical grinding and ultrasonic-assisted exfoliation. A thermal oxidation process employing a two-step probe sonication in an N-methyl-pyrrolidone (NMP)/water solution under an O2 atmosphere was then used to modify the surface of the FeS2 sheets. The sonication results in sufficient oxidization of the FeS2 nanosheets to form Fe2O3 shells (TOPY NSs) with a well-defined Z-scheme heterojunction. The Z-scheme electronic band structures of FeS2 and Fe2O3 in the TOPY NSs allow for efficient charge separation by means of the transfer of photoexcited electrons in the CB of Fe2O3 to the VB of FeS2, thereby preventing the recombination of electron–hole pairs in FeS2. The Fe2O3 shell and Fe3+/Fe2+ inside the heterostructure induced tumor tissue damage with the disruption of the redox homeostasis due to GSH oxidation and •OH production by the Fenton reaction. The generation of •O2 from O2 and •OH from OH on the CB of FeS2 and VB of Fe2O3, respectively, was largely enhanced upon 650 nm laser irradiation. In addition, 808 nm laser irradiation generated local hyperthermia for photothermal therapy (PTT) with a conversion efficiency of 41.3%.67 Semiconducting heterojunction NPs are not only explored for tumor treatment but also for enhanced imaging contrast. For example, PA imaging, an emerging in vivo imaging technique, relies on the ability of contrasting agent to absorb light and convert it to heat. The development and optimization of these contrasting agents is obviously a vital research area with the potential to significantly enhance the capabilities and applications of PA imaging. In an S-M NP, with the heterojunction formation, photoluminescence quenching usually occurs in the semiconductor part due to increased non-radiative recombination by transferring the exciton from the semiconductor to the metallic domain. The non-radiative recombination thus generates localized heat inside the tissues, leading to thermal expansion and generation of acoustic waves, being detected as photoacoustic signals.68 Introducing vacancies or doping impurities and narrowing the bandgap can also enhance photon absorption and thereby the non-radiative recombination and the PA signals. TOPY NSs show excellent PA imaging due to the strong absorbance in NIR wavelength and high photothermal conversion efficiency. Moreover, the fluorescent and photothermal imaging capabilities of heteronanostructures allow them to act as promising candidates for multimodal imaging-guided cancer theranostics.67 

Due to the large atomic nuclei and more electrons, high-Z elements, such as Au, W, and Bi, are more likely to interact with x-rays. When these high-Z elements are introduced to make a heterojunction, the resulting structures can be designed to absorb x-rays more efficiently. This is particularly useful in CT imaging, where contrast and resolution are critical. Bi2S3 nanomaterials have been recently used as PDT and PTT agents due to their higher absorption in the NIR region. The photothermal and photodynamic properties of Bi2S3 NPs can be further improved by forming heterostructures with other materials. Generally, PDT displays low efficiency in hypoxic conditions due to lower O2 concentration. Inspired by the VB position of Bi2S3 nanomaterials (1.48 V vs NHE), Cheng et al. fabricated Bi2S3@Bi Z-scheme heterostructured nanorods for photoinduced hole-mediated oxidation of water to produce O2 for alleviating tumor hypoxia [Fig. 2(b)].57 With the heterojunction formation, the electrons move from the higher Fermi level semi-metallic Bi to the lower Fermi level Bi2S3 until it reaches an equilibrium. Under 808 nm laser irradiation, the electric field stimulates the flow of electrons from the Bi2S3 to Bi, resulting in the recombination of electrons and holes in the CB of Bi2S3 and VB of Bi, respectively. This phenomenon results in the enhanced availability of holes in the VB of Bi2S3 and electrons in the CB of Bi. The free holes thus formed in the VB of Bi2S3 split water to generate O2. A self-supply of O2 and ROS production is showcased, when photoexcited electrons on the CB of Bi in Bi2S3@Bi Z-scheme heterostructured nanorods reacted with O2, which is deemed favorable in hypoxic tumor therapy, with an additional CT imaging capability. Generally, the conjugation of targeting ligands to the nanomaterials can enhance their accumulation in the tumor site and reduce systemic toxicity in normal tissues. PEG-SH and LyP-1 peptide with a cysteine moiety was functionalized on the surface of Bi2S3@Bi heterostructure for enhancing biocompatibility and tumor targeting. Additionally, Bi2S3@Bi Z-scheme heterostructured nanorods hold great promise for improved cancer diagnosis based on their CT imaging capability.

Various approaches have been employed to enhance the production of ROS by semiconductor nanosonosensitizers. These methods include building heterostructures, incorporating dopants, and generating surface vacancies. For example, noble metals (Au, Ag, Pd, Pt, etc.) have been deposited onto the semiconductor surface of nanosonosensitizers to create metal/semiconductor Schottky junctions. These junctions efficiently increase ROS generation by enhancing the utilization of external ultrasound mechanical energy.69 Vacancies are the point defects generally seen in semiconductor materials. Especially, anion vacancies are reported in metal oxides (oxygen vacancy) and metal sulfides (sulfur vacancy). These defects can alter the band structure and act as an electron trap, thereby preventing the recombination of charge carriers and altering the physiochemical properties. This leads to enhanced photocatalytic activity. Different methods have been used to engineer vacancy-rich semiconducting materials, such as introducing vacancies into pre-synthesized semiconducting material and in situ generation of vacancies during the synthesis procedure. Chemical reduction methods, such as material treatment with NaBH4, and high-temperature processes, such as annealing, have been reported to remove O2 atoms from the surface of photocatalysts to generate vacancies.70 In addition, hydrothermal, solvothermal, UV reduction, and chemical etching approaches can generate vacancies in semiconducting materials.71 Meng et al. prepared the sulfur-deficient Bi2S3−x nanorods by one-pot hydrothermal reaction, followed by the in situ deposition of Au nanoparticles on the surface.58 The electron paramagnetic resonance signal at g = 2.004 and the binding energy reduction of sulfur in x-ray photoelectron spectroscopy spectra help them to confirm the generation of sulfur vacancy in Bi2S3−x-Au heterojunction. Bi2S3−x-Au nanorods were then surface functionalized with hyaluronic acid to target the CD44 protein selectively overexpressed 4T1 tumor. An NIR-II light-responsive sulfur-vacant Bi2S3−x-Au heterojunction was developed as an all-in-one phototheranostic platform for increased ROS generation via enhanced electron–hole separation [Fig. 2(c)].58 This heterojunction NPs show improved photothermal efficiency at 1064 nm laser irradiation compared to the individual components. With its magnified absorption in the NIR-II region, the PA imaging contrast is significantly improved. PDT is one of the most promising strategies for advanced combination treatment using chemotherapy. Song et al.59 reported a Bi-HA/FK866 heterojunction photosensitizer for NIR-triggered ROS generation and nicotinamide phosphoribosyl transferase inhibition. Nicotinamide phosphoribosyl transferase participates in cancer migration by activating the ERK1/2 pathway and downregulating E-cadherin. Humic acid reduces Bi (III) to elemental Bi, enabling NIR (808 nm) absorption and electron/hole pair generation due to the narrow bandgap [Fig. 2(d)]. While transmission electron microscopy confirmed its particle size (81 nm), x-ray diffraction analysis verified the formation of Bi0, which is responsible for elevated photothermal heating (52.7 °C, 5 min, 0.5 W/cm2). The Bi-HA platform facilitated efficient electron/hole separation via band alignment (Bi CB: 0.544 eV, VB: −0.986 eV; HA LUMO: −0.45 eV, HOMO: −1.284 eV). The heterojunction promoted charge separation, leading to enhanced ROS production and DNA damage in tumor cells. Photogenerated electrons reduce H2O2/H2O to •O2/•OH, while holes oxidize H2O/GSH to •OH/GSSG. An increase in Bi-HA concentrations led to a linear increase in multi-spectral optoacoustic tomography and CT signals. Enhanced ROS generation and DNA damage increase the sensitivity to FK866 (inhibitor of nicotinamide phosphoribosyl transferase), downregulating the NAD+/ERK/NF-κB pathway and inhibiting tumor progression. Downregulation of NAD+ prevents cancer proliferation and GSSG reduction through reduced NADPH expression. At the same time, reduction in ERK1/2 and NF-κB expression prevents tumor migration. In vivo blood biochemistry showed no significant differences, indicating high biocompatibility of the Bi-HA/FK866 platform. Altogether, this study sheds light on the impact of combinatorial therapy utilizing heterojunction nanostructures, revealing its effectiveness in tumor treatment.

Beyond Bi-based heterojunctions, novel materials with heterojunction architectures have also been actively explored for photocatalytic tumor therapy.72 Most heterojunction nanomedicines focus on killing cancer cells by inducing apoptosis. Since anti-apoptosis is a common trait of neoplasm, the introduction of new therapeutic approaches for non-apoptotic regulated cell death is often in great demand.73 In a recent study, Liu et al.60 proposed PEGylated Cu+/Cu2+ doped black phosphorus (BP) and polypyrrole (CP) heterojunction (BP@CPP) NPs as a potential therapeutic approach for malignant breast cancer. TEM analysis revealed sheet-like BP@CPP (421.7 × 327.4 nm2) with successful Cu ion doping, confirmed by scanning transmission electron microscope and x-ray photoelectron spectroscopy analyses. BP@CPP exhibits superior light absorption across the visible light to NIR region compared to bare BP, translating to a remarkable 51.7% photothermal conversion efficiency under 808 nm laser irradiation. Its therapeutic potential lies in the engineered Z-scheme heterojunction. Upon laser irradiation, photogenerated electron and hole efficiently separate on BP and CP surfaces, respectively. This design fosters preferential recombination of photoexcited electrons in the CB of CP with h+ in the VB of BP. This spatial separation minimizes charge recombination, leading to high oxidative and reductive activities in the VB of CP and CB of BP, and promotes the generation of •OH and •O2 due to favorable band potentials. Moreover, an obvious GSH depletion was achieved due to the presence of Cu2+ in the BP@CPP. Typically, GSH depletion results in the inactivity of glutathione peroxidase 4, leading to lipid peroxide accumulation in the cell membrane, finally accelerating ferroptosis of tumor cells [Fig. 2(d)]. Work function is an important parameter in determining heterostructure NP's band alignment. It is the minimum energy required to release an electron from the surface of a material to the vacuum (a point just outside the material). In a typical heterojunction, the difference in the work function between the two components affects the alignment of their energy bands when forming the heterojunction. This band alignment will define the charge separation and migration across the junction between two materials, which is crucial for the catalytic activity. By carefully selecting materials with appropriate work functions and electronic properties, designing an efficient nanoheterojunction is possible. A Pt/hollow-TiN Schottky heterojunction NPs was developed for 4T1 cancer therapy in a recent work.74 Generally, noble metals with high work function and n-type semiconductors with low work function are suitable for constructing Schottky heterojunctions. The work function and fermi level measurement by ultraviolet photoelectron spectroscopy reveal that the fermi level of hollow-TiN NPs, Pt NPs, and Pt/hollow-TiN NPs was 4.46, 4.25, and 4.34 eV, respectively, and their work function was 4.94, 5.66, and 5.13 eV, respectively. The data shows that the work function of hollow-TiN is smaller than the metal Pt, and the fermi level is higher. When these two materials form the heterojunction, the electrons will be transferred from hollow-TiN to Pt until the fermi level at the junction becomes equilibrium. At the same time, the space charge regions formed on the hollow-TiN result in the bending of the energy band upward and form a Schottky barrier. When this heterojunction was irradiated using the US, the excited electrons from the CB of hollow-TiN were transferred to the Fermi level of metallic Pt through the Schottky barrier, preventing their backflow. The holes left behind in the VB of hollow-TiN remain free without recombination. This effective charge separation enhanced the SCT performance and, thus, the ROS production and ferroptosis. Several other heterojunction NPs, such as Bi2MoO6-MXene Schottky heterojunction,75 l-arginine CoWO4/FeWO4 S-scheme heterojunction,76 and Ti3C2@ BaTiO3 Schottky heterojunction,77 have been reported to induce ferroptosis. However, photocatalytic heterojunction semiconductors face numerous challenges in their application for tumor therapy. One critical requirement is that the heterojunction materials must possess a bandgap <1.53 eV. Additionally, a significant drawback is the limited penetration of light into the inner core of the tumor, leading to a lower catalytic efficiency. Despite its challenges, phototherapy offers several advantages over other therapies, including precise cancer targeting, minimal invasiveness, and the capability of NIR-activated heterojunction catalysts to generate heat for photothermal treatment, which will be further discussed in Sec. III B. Table I summarizes additional semiconducting heterojunction materials investigated for the generation of ROS via photocatalysis for tumor therapy.

2. Sonocatalysis

Sonodynamic therapy (SDT), a derivative of PDT, operates by utilizing US energy to activate sonosensitizers and O2 molecules for ROS production. Sonosensitizers respond to US waves by generating ROS, including •OH and 1O2. Mechanistically, ultrasound energy triggers electron transitions in sonosensitizers, causing ROS production through reactions with O2 and water. Important parameters that are critical for sonocatalytic therapy (SCT) are species and concentrations of therapeutically generated products and the active site of the sonosensitizer. It is vitally important to augment quantum yield and catalytic efficacy by tailoring sonosensitizer structures and properties under limited external stimulatory energy and substrate types and levels.

Inorganic sonosensitizers, such as zinc oxide, similarly produce ROS through electron/hole separation and redox reactions with O2 and water. This ROS-induced oxidative stress triggers tumor cell apoptosis and autophagy, contributing to tumor eradication.78 TiO2 (Eg = 3.2 eV) is one of the most representative materials studied in the biomedical application of inorganic sonosensitizers due to their excellent biocompatibility and high stability for in vivo translation.79 Many strategies, including heterostructure formation, have been employed to improve the therapeutic outcome of TiO2 NPs.79,80 Zhou et al. demonstrated an engineered TiO2@MnO2-x-PEG heterostructure that exhibited substantially improved sonocatalytic efficacy compared to pure TiO2 semiconductors due to enhanced electron/hole pair separation [Fig. 3(a)].81 Additionally, the incorporation of MnO2-x imparts the heterostructure with the ability to modulate tumor hypoxia, deplete GSH, and initiate Fenton-like reactions. High aspect ratio surface area and porosity are essential for better catalyst performance to absorb and oxidize/reduce the target species. In general, the porous structure enhances the light-harvesting capability of the material. It helps in the mass adsorption and diffusion of substrates and products, thus enhancing the active sites for the redox reaction and minimizing the hot electron and hole migration distance from bulk to surface.82 Moreover, for cancer therapy, the porous photocatalysts allow the loading of anticancer agents and deliver them to the tumor site, paving a path for combination therapy. In a study, US-activatable, metal-free TiO2@g-C3N4 heterostructure nanotherapeutics loaded with histone deacetylase inhibitor, romidepsin, for effective chemo/sonocatalytic combination therapy has been recently developed by He et al.83 Poly(vinylpyrrolidone) assisted solvothermal method was used to prepare hollow and porous TiO2, and the total surface area of the particle was found to be 67.183 m2/g, with a pore diameter calculated to be 2.896 nm based on the N2 adsorption/desorption isotherms. Later, g-C3N4 QDs were electrostatically deposited on the surface of TiO2 via precipitation. The authors found that the heterojunction formation between TiO2 and g-C3N4 quantum dots reduced the bandgap from 3.23 to 2.98 eV, determined by extrapolating the linear part of the Tauc plot from the transformed Kubelka–Munk function. The reduction in the bandgap may promote the redox reactions under low-power US irradiation. The VB potentials of TiO2 and g-C3N4 QDs were 2.31 and 1.8 eV vs normal hydrogen electrode, respectively, as estimated from XPS-VB spectra. The CB potentials of TiO2 and g-C3N4 QDs were determined by subtracting the bandgap energy from the VB potential that gave rise to −0.92 and −1.15 eV, respectively. The band alignment led the heterostructure when it was formed to a direct Z-scheme heterojunction, allowing the migration of hot electrons from TiO2 to g-C3N4 QDs, yet leaving holes in the VB. The hot electrons in the CB of g-C3N4 QDs after US stimulation could effectively reduce O2 to •O2−, while the holes remaining on VB of TiO2 facilitated catalytic oxidation from H2O to •OH for effective tumor therapy. Rather than TiO2-based heterojunctions, several metal oxide NP-based heteronanostructures have been employed for cancer therapy.84,85 Metal-organic frameworks (MOF) are a type of porous coordination material built with metal ions and organic linkers and widely used in catalytic applications due to their porous architecture and efficient catalytic activity.86 However, the organic linkers coordinating the metal ions or clusters reduce the charge separation and thus cause rapid exciton quenching. However, this issue can be addressed by introducing metal NPs on the surface of MOF. Meng et al. developed Ti-based MOF (MIL) deposited with Ag for SCT.87 Under US stimulation, the excited electrons from the MIL move to the deposited Ag NPs by metal-to-metal electron transfer, thus effectively separating electrons and holes for effective redox reactions. Several metal nanoparticles and carbon dots exhibit POD, CAT, SOD, and GPx enzyme-like activity (nanozymes), offering considerable potential for cancer treatment. Integrating enzyme-mimicking metal NPs with semiconductors can facilitate the formation of the Schottky junction to improve the therapeutic efficacy by both semiconducting catalytic and enzymatic routes. A simple defect-rich Pt-ZnO S-M heterojunction sonosensitizer was developed by Li et al., demonstrating dual nanozyme activities and excellent sonocatalytic properties, resulting in a triple ROS amplification for synergistic tumor therapy. The heterostructure formation and defects present in Pt-ZnO improved the efficiency of electron and hole separation under US, prolonging the lifespan of electron and enhancing ROS generation during SCT. Additionally, Pt-ZnO exhibited significant POD- and CAT-like activities, mitigating tumor hypoxia and increasing ROS generation [Fig. 3(b)]. Importantly, Pt-ZnO effectively depleted GSH levels, therefore minimizing ROS consumption and augmenting oxidative stress, resulting in a 614.4% increase in •OH yield and an 859.1% increase in 1O2 yield. Consequently, Pt-ZnO represented highly effective synergistic SCT, achieving a remarkable tumor inhibition rate of 98.1%.88 Several heterojunction NPs, such as carbon dots@Nb2C nanozymes (POD, CAT),89 Fe3O4/Ag/Bi2MoO6 NPs (POD, CAT, SOD),90 Co3O4@TiO2−x (POD, CAT),91 Ru/TiO2−x@TiCN (POD, GPx, CAT),92 and hollow black TiO2 nanosphere-carbon dots (CAT, GPx, POD like),93 have been reported to show the capability of acting as both nanozymes and semiconducting catalysts working together to kill tumor cells.

FIG. 3.

(a) Schematic illustration of synthesis procedure of TiO2@MnO2-x-PEG nanosheets, HAADF-STEM image, and HR-TEM image of TiO2@MnO2-x nanosheets and 4T1 tumor photographs after different treatments. Reproduced with the permission from Zhou et al., Chem. Eng. J. 431, 134017 (2022). Copyright 2022 Elsevier.81 (b) Schematic illustration of synergistic sonocatalytic therapy of Pt-ZnO with triple ROS amplification effect for effective tumor therapy. Reproduced with the permission from Li et al., Acta Biomater. 171, 543–552 (2023). Copyright 2023 Elsevier.88 (c) Schematic illustration of a type II N-CD@LiFePO4 heterojunction and the photocatalytic ROS generation mechanism and HRTEM image featuring the deposition of CD on the surface of LiFePO4. Also shown is the relative cell viability of 143B cells treated with N-CD@LiFePO4 + H2O2 under US irradiation. Reproduced with the permission from Hu et al., Chem. Eng. J. 446, 137320 (2022). Copyright 2022 Elsevier.94 (d) The proposed mechanism of enhanced sonodynamic performance, including GSH depletion, hypoxia alleviation, and photothermal effect of Bi@Bi2O3@Bi2S3-PEG NPs. Reproduced with the permission from Song et al., Adv. Healthcare Mater. 11(11), 2102503 (2022). Copyright 2022 John Wiley and Sons.95 

FIG. 3.

(a) Schematic illustration of synthesis procedure of TiO2@MnO2-x-PEG nanosheets, HAADF-STEM image, and HR-TEM image of TiO2@MnO2-x nanosheets and 4T1 tumor photographs after different treatments. Reproduced with the permission from Zhou et al., Chem. Eng. J. 431, 134017 (2022). Copyright 2022 Elsevier.81 (b) Schematic illustration of synergistic sonocatalytic therapy of Pt-ZnO with triple ROS amplification effect for effective tumor therapy. Reproduced with the permission from Li et al., Acta Biomater. 171, 543–552 (2023). Copyright 2023 Elsevier.88 (c) Schematic illustration of a type II N-CD@LiFePO4 heterojunction and the photocatalytic ROS generation mechanism and HRTEM image featuring the deposition of CD on the surface of LiFePO4. Also shown is the relative cell viability of 143B cells treated with N-CD@LiFePO4 + H2O2 under US irradiation. Reproduced with the permission from Hu et al., Chem. Eng. J. 446, 137320 (2022). Copyright 2022 Elsevier.94 (d) The proposed mechanism of enhanced sonodynamic performance, including GSH depletion, hypoxia alleviation, and photothermal effect of Bi@Bi2O3@Bi2S3-PEG NPs. Reproduced with the permission from Song et al., Adv. Healthcare Mater. 11(11), 2102503 (2022). Copyright 2022 John Wiley and Sons.95 

Close modal

Recent advancements in the development of innovative sonocatalysts have opened up avenues for the creation of diverse heterojunctions aimed at enhancing the therapeutic efficacy of tumor treatment. Hu et al. introduced a novel biocompatible narrow-bandgap N-CD@LiFePO4 sonocatalyst [Fig. 3(c)].94 This sonosensitizer effectively catalyzed the decomposition of H2O2 into •OH radicals while simultaneously depleting overexpressed GSH through Fe3+ oxidation. The (PO4)3- groups significantly improve H2O2 decomposition compared to traditional iron-based nanozymes. Furthermore, the p–n junctions within the structure enhance ROS production by efficiently separating charge carriers under US irradiation. The p–n junction boosted sonodynamic therapy, Fe2+/Fe3+ mediated TME reconstruction, and targeted tumor accumulation achieved complete eradication of human osteosarcoma in a single treatment using US irradiation. Crucially, this p–n junction sonocatalyst represents a new class of semiconductor heterojunction nanomedicines that can be gradually metabolized from the body, opening avenues for their clinical translation.94,96 Instead of binary heterostructures, a three-component heterostructure system named ternary heterojunctions can create more complex band alignments, leading to more effective separation of charge carriers. This can reduce recombination rates and enhance the efficiency of photocatalytic reactions. In an approach, Song et al. developed a ternary heterojunction, Bi@Bi2O3@Bi2S3-PEG, to enhance the effectiveness of SCT, synergized by PTT, for tumors. This nano-heterojunction was prepared using the coprecipitation and reduction method. Upon US activation, the nanosonocatalyst generated abundant ROS (•O2 and •OH), facilitated by oxygen vacancies formation, localized surface plasmon resonance effect induced by Bi, and sonoluminescence effect. These mechanisms collectively improved the electron/hole separation capability of Bi@Bi2O3@Bi2S3-PEG. Furthermore, the bismuth-based nanomaterial continuously depleted GSH, disrupting redox homeostasis and inducing sustained oxidative stress while also enabling easy degradation and metabolism. Additionally, Bi@Bi2O3@Bi2S3-PEG exhibited remarkable photothermal conversion ability upon NIR stimulation, causing thermal tumor damage and relieving tumor hypoxia while also providing endogenous O2 for SDT [Fig. 3(d)].95 Immunotherapy has advanced in the cancer treatment recently in various aspects, especially in SCT where immunogenic cell death is induced to elicit antitumor responses. Liu and his group decorated Au and carbon dots on hollow black TiO2, which showed a spherical hollow structure with a diameter of 100–200 nm. The heterojunction showed significant ROS generation under US due to the presence of oxygen vacancies, Ti3+ in the black TiO2, as well as the localized surface plasmon resonance of Au NPs. Hypoxia can be alleviated in TME and inhibit the immunosuppressive mediators expression by the triple-enzyme mimetic activity of the heteronanostructure to enhance the antitumor effect by SCT. Bi2O3, a wide energy gap semiconductor, has recently been employed and fabricated for combined photothermal/sonocatalytic therapy by surface decoration with Au nanorods to form a Schottky junction. Under US treatment, the hot electrons generated from Bi2O3 migrated toward metallic Au nanorods through the Schottky junction, becoming prohibited from backflow. Therefore, their recombination with holes was significantly reduced. The high electron density thereby rendered the Au nanorod of sufficient potential to reduce O2 to •O2−, while the holes in the VB of the wideband gap Bi2O3 possessed the catalytic ability to drive H2O oxidation to •OH. Due to the localized surface plasmon resonance (LSPR) effect of Au nanorods, the heterostructure showed good photothermal conversion under NIR irradiation at 808 nm, while Bi2O3 could deplete intratumoral GSH by forming a coordination bond. Altogether, the disruption of redox balance and photothermal heat generation by the catalytic system induces immunogenic cell death, thereby reversing tumor immunosuppressive microenvironment and promoting dendritic cell maturation and CD8+ T cell infiltration.97 Altogether, SCT using heterojunction NPs, as an emerging field in nanotechnology-assisted tumor therapy, is garnering significant attention due to its remarkable outcomes. However, further research is needed to translate the utilization of heterojunction NPs in SCT from the laboratory bench to clinical applications. Table I summarizes recent studies on heterojunction materials employed for sonocatalysis-based tumor therapy.

3. Radiocatalysis

The radiocatalytic activity of nanocatalysts depends on the process and products created when materials interact physically with gamma rays or x-rays. X ray radiation can cause physical reactions in NPs, including the Auger effect, photoelectric scattering, and Compton effect, creating secondary x-rays, free electron, and fluorescence emission.98,99 X ray irradiation of heavy metals ejects photoelectrons with characteristic kinetic energies in the keV range. The specific energy distribution of these photoelectrons is influenced by the incident x-ray energy and the elemental composition of NPs.100 The generation of photoelectrons when elements interact with x-rays is directly related to their atomic number (Z) but inversely related to x-ray energy. NPs, such as Bi (Z = 83) and Au (Z = 79), can be utilized to enhance the generation of active molecules or species with therapeutic properties, thereby enhancing radiotherapy efficacy.98,101 When photoelectrons are emitted from lower atomic orbitals, electron from higher orbitals shift to fill the void, creating a hole and releasing excessive energy. This energy can either lead to the emission of fluorescent photons or the ionization of an electron in a higher orbital, known as Auger electron. Auger electron has low energy and limited travel distance, primarily depositing near the surface of NPs.102 Compton electron is generated through Compton scattering, which involves a single photon interacting with a free electron. Scintillation NPs exploit luminescence to convert x-ray energy into light. X ray irradiation excites electron (VB to CB), leading to short-lived fluorescence or long-lived phosphorescence upon relaxation.103,104 Various high-Z NPs, such as noble metals (Ag, Au, and Pt), rare earth elements (Gd, Tm, and Ho), and other heavy elements (Bi, Hf, W, and Ta), have been utilized as sensitizers in radiation therapy.105–109 Generally, x-rays interact with the above-mentioned metals (directly) and water (indirectly) to produce ROS. In recent years, inorganic semiconductor nanomaterials incorporating high-Z elements have garnered significant interest as potential radiosensitizers due to their ability to enhance therapeutic efficacy.110 Their mechanism of action shares similarities with photocatalysis. However, ionizing radiation excitation forms charge carriers through a more intricate pathway. Upon exposure to high-energy x-rays or gamma photons, the photoelectric effect leads to the ejection of a core electron, creating hole in the core orbital.111 These core holes are short-lived, rapidly relaxing to the VB within picoseconds via Auger processes. Consequently, they are not directly involved in surface reactions. The dislodged, high-energy electron (hot electron) triggers a cascade of secondary ionization and bandgap excitation events. This ultimately results in the generation of long-lived electron/hole pairs with a yield of approximately 102/keV of absorbed energy in typical semiconductors. Additionally, the escape of these ionized electrons as lower-energy particles contributes to enhanced water radiolysis, further amplifying the therapeutic effect.112 This unique band structure enables the semiconducting nanomaterials to produce free radicals independent of O2 availability.113 

Akin to photocatalysis being limited by material properties, employing solely single-component semiconductors might not achieve optimal results in radiocatalytic therapy (RCT). This is attributed to the intrinsic radioresistance of tumors arising from hypoxia, a characteristic feature of TME, which can induce systemic acquired radioresistance, rendering cancer cells 2–3 times more resistant to radiation compared to oxygenated (normoxic) tumor cells.115 Heterojunction semiconducting nanomaterials play a vital role in overcoming radioresistance and thereby enhancing treatment efficacy. Research in radiocatalysis suggests that combining metal oxide semiconductor NPs with ionizing radiation offers a potentially faster alternative to traditional photocatalysis methods using visible or UV light.117 Due to the emphasis on heterojunctions for enhanced radiocatalytic performance, researchers are actively investigating metal oxide/sulfide based composites with this architecture. In heterojunction catalysts, apart from the material selection, the crystal structures play a significant role in their catalytic performance. High crystallinity of the materials can contribute to better lattice matching while forming the heterojunction and promote effective charge transfer across the interface, thus enhancing the overall catalytic performance. Liquid phase deposition and chemical vapor deposition are two useful techniques employed to deposit highly crystalline and pure materials on the substrate. Although highly crystalline materials are more resistant to photo-leaching and stable under x-ray irradiation, amorphous materials may exhibit special properties in heterojunction catalysis due to their electronic characteristics. Inspired by this, amorphous/crystalline heterojunctions for effective catalytic application were built and their properties were studied.118,119 Generally, the amorphous phase is incorporated into the crystalline phase of the material by two strategies. One method is the amorphization of the crystalline materials, and the other is to grow the amorphous phase on the crystalline structure. One of the common routes used for the amorphization of the crystalline phase is the ion exchange method. In a recent study, Wang et al.120 developed a Z-scheme BiOI/Bi2S3@polydopamine nanosheets for RCT [Fig. 4(a)]. The nanosheets modified by glucose oxidase can generate an environment with excessive H2O2, supporting RCT. This not only aids ROS generation but also helps cut off energy to tumor cells via starvation. The energy level diagram by Wang and coworkers suggests a traditional type III exciton separation, involving significantly larger band gaps between the individual CB and VB of BiOI and Bi2S3 (0.96 and 1.68 V) compared to the gap between their combined CB and VB (0.36 V) point toward a more favorable mechanism, a Z-scheme-mediated charge transfer process. In this scenario, photogenerated electrons from BiOI CB directly recombine with hole in Bi2S3 VB. This facilitates the participation of both electron and hole in the reduction of H2O2 and oxidation of H2O, leading to the observed generation of •OH. In vitro and in vivo studies utilizing HeLa tumor cells demonstrated that BiOI/Bi2S3 can exploit a synergistic combination of starvation and RCT. This combined approach effectively increases ROS accumulation and chromosomal damage within tumor cell. Similarly, Guo et al.114 used BSA-coated BiOI/Bi2S3 nanoheterojunction for the combination of radio/PDT/PTT in an earlier study. BiOI/Bi2S3 offers a multimodal therapeutic approach to cancer treatment. Under x-ray irradiation, generated electron/hole pairs cause ROS production, resulting in a photodynamic effect. Additionally, the superior NIR absorption of Bi2S3 contributes to PTT. Furthermore, high x-ray attenuation and NIR absorption of the NPs make them suitable for use as a contrast agent in CT and PA imaging, enabling effective tumor visualization. In vitro studies using the BEL-7402 hepatocarcinoma cell line demonstrated a significant enhancement in tumor growth inhibition, highlighting the potential of this platform for cancer therapy. Recently, polyoxometalates (POMs), frequently used in the preparation of semiconducting nanocatalysts in transitional metals, such as Mo, W, and Nb, have shown wide potential in catalytic application due to their tunable composition that helps adjust the bandgap to improve the catalytic performance from catalytic active centers that are located and exposed on the surface. POMs can be easily functionalized or loaded onto several semiconducting materials for building a heterojunction nanocomposite. In yet another recent study, Zhou115 and coworkers have explored a novel Bi-based POM material (BiP5W3) for enhanced radiotherapy efficacy. This hybrid material combines BiP5W30 nanoclusters, containing elements with high atomic numbers, and reduced graphene oxide (rGO) to exploit synergistic functionalities. BiP5W30 nanoclusters effectively enhance the radiation-induced ROS generation in tumor cells due to their unique electronic properties [Fig. 4(b)]. Additionally, they deplete GSH, a key cellular antioxidant, via redox reactions and catalyze the conversion of H2O2 into cytotoxic •OH, significantly amplifying ROS generation upon x-ray exposure. rGO further enhances the efficacy of the platform by promoting electron/hole separation within the material, improving radiocatalysis. Moreover, the superior NIR absorption of rGO contributes to photothermal effects, elevating tumor temperature and compromising cancer cell viability. Promising results from both in vitro and in vivo studies using HeLa cells suggest the potential of this hybrid material for a combined therapeutic approach in the cancer treatment, highlighting its potential as a promising candidate for further investigation. Other Bi-based heterojunctions are also explored toward RCT, including BiVO4@Bi2S3 by Wang et al.,121 PVP-functionalized Bi/GO by Zhou et al.,122 and Au-Bi2S3 by Wang et al.113 These materials offer not only enhanced radiocatalytic efficiency but also proved to be useful for multimodal imaging applications owing to their high x-ray attenuation. Cai et al.123 investigated a plasmonic gold-based NPs (AuPt@CuS) for its potential in combined radiophotothermal therapy and PA/CT imaging. They highlighted the unique properties of AuPt@CuS heterojunction for enhanced therapeutic efficacy [Fig. 4(c)]. AuPt@CuS NPs demonstrated high photothermal efficiency (41.56%) due to localized electromagnetic field enhancement at the heterojunction interface, significantly rising the temperature (up to 50.1 °C) in the NIR region. This localized heating can directly ablate cancer cells. Furthermore, the favorable energy level alignment of the material facilitates the separation of radio-generated electron/hole. Then, these separated charges can interact with surrounding molecules: electron reduces O2 to ROS, while the hole depletes GSH, a cellular antioxidant. This depletion of GSH and the concomitant rise in ROS levels synergistically enhance the cytotoxicity of radiotherapy. Additionally, AuPt@CuS NPs exhibit high x-ray attenuation due to their composition, making them suitable for PA/CT imaging. The interaction of x-rays with NPs generates a substantial electron pool. These electrons can further react with water molecules to produce hydrated electrons that disrupt cellular redox balance, ultimately leading to tumor cell death. In vitro and in vivo studies conducted on murine mammary carcinoma (4T1) cells confirmed the therapeutic potential of this novel nanoheterojunction in cancer treatment. Photocatalysts have been widely developed to reduce the toxic metal ions in water and relieve environmental toxicity in the environmental engineering field.124 The same story but a different application is used to convert less toxic metal ions to higher toxic metal ions in tumor tissues for cancer treatment. For example, Wang et al.116 have constructed a Z-scheme graphdiyne heterojunction (CuO@GDY) for photocatalytic O2 generation and radiosensitization for tumor therapy [Fig. 4(d)]. Cu+, as a high-energy species, is more active to undergo the Fenton-like reaction via reaction with H2O2 than Cu2+. When the CuO@GDY nanocatalyst is stimulated by x-ray irradiation, Cu2+ ions are mostly reduced to Cu+ and the latter acts as an active site for accelerated Fenton-like reaction for ROS production in response to endogenous H2O2, causing tumor cell death. This combination of photo- and radiocatalytic activities leads to a controllable and precise therapy. Wang et al.125 presented a novel radiosensitizing material, a lanthanide pyrosilicate scintillator coated with titanium dioxide (LnPS@TiO2). This composite material leverages the unique properties of both components to achieve enhanced radiation therapy. LnPS scintillator converts incoming radiation energy into UV light. Then, this emitted UV light excites TiO2 NPs via fluorescence resonance energy transfer, which occurs due to the spectral overlap between the LnPS emission wavelength (368 nm) and the absorption band of TiO2. This efficient energy transfer significantly increases the generation of •O2 radicals by TiO2. •O2 produced by excited TiO2 NPs target cellular iron-sulfur clusters, disrupting essential DNA metabolic processes. Additionally, these radicals can interact with various biomolecules, such as peptides, carbohydrates, nucleic acids, and lipids, inducing long-term oxidative stress within the cell. The combined effect of impaired DNA metabolism and oxidative stress ultimately leads to tumor cell death. As discussed above, many researchers have recently described the implications of high-Z semiconductor heterojunction NPs in radiotherapy due to their versatile physicochemical properties. Table I summarizes recent studies of heterojunction materials employed for radiocatalysis-based tumor therapy.

FIG. 4.

(a) Schematic representation of the generation of ROS under x-ray irradiation for radiocatalytic therapy using BiOI@Bi2S3 heterojunction NPs. Reproduced with the permission from Guo et al., Adv. Mater. 29(44), 1704136 (2017). Copyright 2017 John Wiley and Sons.114 (b) The possible charge transfer mechanism to generate cytotoxic free radicals in the tumor treatment by BiP5W30@rGO (PVP-PG) heterostructures under x-ray irradiation along with the in vitro ROS fluorescence imaging under different treatments. Reproduced with the permission from Zhou et al., Biomaterials 189, 11–22 (2019). Copyright 2019 Elsevier.115 (c) Illustration of possible charge transfer routes and the working principles of AuPt@CuS nanosheets to generate cytotoxic free radicals under x-ray irradiation for effective tumor therapy. Reproduced with the permission from Cai et al., J. Am. Chem. Soc. 143(39), 16113–16127 (2021). Copyright 2022 American Chemical Society.123 (d) Illustration of the combined radiotherapy and CuO@GDY-mediated radiocatalytic therapy and their effects on colony formation of 4T1 cells and HUVECs under various treatments. Reproduced with permission from Wang et al., ACS Nano 16(12), 21186–21198 (2022). Copyright 2022 American Chemical Society.116 

FIG. 4.

(a) Schematic representation of the generation of ROS under x-ray irradiation for radiocatalytic therapy using BiOI@Bi2S3 heterojunction NPs. Reproduced with the permission from Guo et al., Adv. Mater. 29(44), 1704136 (2017). Copyright 2017 John Wiley and Sons.114 (b) The possible charge transfer mechanism to generate cytotoxic free radicals in the tumor treatment by BiP5W30@rGO (PVP-PG) heterostructures under x-ray irradiation along with the in vitro ROS fluorescence imaging under different treatments. Reproduced with the permission from Zhou et al., Biomaterials 189, 11–22 (2019). Copyright 2019 Elsevier.115 (c) Illustration of possible charge transfer routes and the working principles of AuPt@CuS nanosheets to generate cytotoxic free radicals under x-ray irradiation for effective tumor therapy. Reproduced with the permission from Cai et al., J. Am. Chem. Soc. 143(39), 16113–16127 (2021). Copyright 2022 American Chemical Society.123 (d) Illustration of the combined radiotherapy and CuO@GDY-mediated radiocatalytic therapy and their effects on colony formation of 4T1 cells and HUVECs under various treatments. Reproduced with permission from Wang et al., ACS Nano 16(12), 21186–21198 (2022). Copyright 2022 American Chemical Society.116 

Close modal

4. Thermoelectric NPs in cancer therapy

Thermoelectric phenomena bridge the fascinating interplay between heat, electricity, and charge carrier behavior within a material. Two key effects that govern this interplay are the Seebeck and Peltier effects. The Seebeck effect describes the conversion of a temperature difference into electricity, while the Peltier effect utilizes electricity to manipulate heat flow.126 In the nanomedicine field, there is a growing interest in thermoelectric nanomaterials utilizing the catalytic potential of electron and hole formed by temperature fluctuations near body temperature.127,128 For such applications, NPs with high figures of merit (ZT) are crucial. ZT is defined as ZT=S2σT/κ, where S is the Seebeck coefficient, σ is the electrical conductivity, T is the absolute temperature, and κ is the thermal conductivity. A two-pronged approach is necessary to achieve high ZT. First, semiconductor NPs with narrow bandgaps are preferred for a high-power factor (S2σ). Second, semiconductors containing heavy elements with significant lattice anharmonicity or complex structures are desirable to achieve low thermal conductivity (κ).129 For example, canonical binary compounds, such as Bi2Te3,130 SnSe,131 and PbTe,132 are extensively used thermoelectric materials.

Recent research in thermoelectric nanocatalysts has shown promise in cancer therapy. Wang et al. investigated Bi13S18I2, demonstrating the generation of pyrocurrent by an electrochemical workstation under the heat and cool cycle of an aqueous solution. The pyroelectric electron generated from temperature fluctuation was used to convert O2 to •O2. The anticancer property was evaluated in 4T1 tumor-bearing mice, in which the synergistic effect of Bi13S18I2 nanorods mediated PTT and pyroelectric dynamic therapy (PEDT) selectively killed cancer cells.127 Building on this concept, Dong et al. reported the photothermoelectric property of ternary copper chalcogenide Cu3VS4 with the mitochondrial-targeting ability for PA imaging–guided synergistic photothermoelectric therapy/CDT (PTET/CDT).133 The development of Bi2Te3 NPs by Jiang et al. demonstrated the potential of photothermoelectric catalysis.134 Kang et al. engineered a p–n junction heterojunction combining SrTiO3 (n-type) and Cu2Se (p-type) to construct a new interfacial electric field to facilitate charge separation and prevent rapid recombination.135 Yual et al. utilized the exothermic reaction of CaO2 hydrolysis in the acidic TME to trigger the thermoelectric property of Bi0.5Sb1.5Te3 @ CaO2 NPs, release Ca2+ ions, and produce H2O2. The heat acts as a stimulus to excite electron to VB, and the built-in electric field minimizes charge recombination. Thus, excited electrons are used to reduce O2 to •O2 radical, which in combination with Ca2+ ions overload increased osmotic pressure in cancer cells, ultimately leading to cell death.136 Ji et al. synthesized Z-scheme heterojunction NPs combining n-type Bi2Te3−xSex and p-type BixSb2−xTe3 nanoplates to improve thermoelectric properties at the near-body temperature. This approach forms a new Fermi level and band bending due to the flow of electron from BST to BTS. Overall, the CB of heterojunction approached the reduction potential of O2 to •O2, resulting in ROS production at temperature fluctuations at the near-body temperature.

Photothermal refers to a process or effect related to the conversion of light (photons) into heat (thermal energy).169 In the context of PTT, it specifically refers to a therapeutic approach in which light-absorbing materials (NPs) are used to convert absorbed light into heat.170 Then, this localized generation of heat is employed to induce hyperthermia in a targeted region, which can have various therapeutic applications, including the destruction of cancer cells or treatment of certain diseases.171 Recently, semiconductor nanomaterials with a wide-range bandgap have been used as photosensitizers in cancer therapy.172 Nanomaterials with narrow bandgap or high carrier concentration are activable under NIR irradiation, while the engineering of low carrier concentration nanomaterials or wide bandgap materials makes them a new member of NIR activable therapeutic agents in the field of cancer nanomedicine.173 Heterojunction formation is a popular strategy to engineer the bandgap of NPs. The formation of nanoheterojunctions induces a reduction in the bandgap of nanomaterials by modifying the electronic structure at the interface. This reduction occurs due to the alignment of energy levels between two distinct semiconductor materials, resulting in a more favorable configuration for electron transitions across the bandgap. The reduced bandgap enables the nanomaterials to absorb a broader range of photons, especially in the NIR region. The performance of semiconducting PTT agent can be improved mainly through three mechanisms: (i) increasing the absorbance and utilization of light; (ii) enhancing the electron/hole separation efficiency; and (iii) inhibiting the rapid radiative recombination of electron/hole pairs. Recent studies have shown that matched energy levels of heterostructures could significantly improve photothermal performance. Many strategies, such as doping atoms or impurities into the lattice, introducing disorder, and creating defects in traditional wide bandgap semiconductors, have been employed to modify the band energies of photocatalysts, thereby extending their activity into the NIR region.

The increased light absorption by the nanoheterojunction translates to a higher probability of exciting electron from the VB to the CB. The nanoheterojunction interface introduces a built-in electric field due to the difference in work function and electron affinities between the two materials. This electric field acts as a driving force spatially separating electron/hole. Then, the separated charges are available for participating in photothermal processes, such as electron transfer reactions or interactions with surrounding molecules.174 The efficient charge separation and reduced recombination contribute to an overall enhancement of the photothermal conversion efficiency of the nanoheterojunction.175,176 The absorbed energy is more effectively converted into localized heat, which is crucial for inducing photothermal ablation of cancer cells. Nuo et al. synthesized SnO2 NPs with a bandgap of 3.6 eV and doped them with Sb to obtain an NIR-responsive Sb0.2-SnO2 nanocrystal [Fig. 5(a)]. SnO2 originally did not respond to light in the visible or NIR spectrum due to a large bandgap; however, after Sb doping, an all-in-one semiconductor NP was formed, which was then used for NIR-II-responsive photothermal therapy in combination with CT and PA imaging.177 Similarly, high bandgap TiO2 NPs were doped with various elements, such as boron,178 carbon,179 nitrogen,180 iron,181 and tungsten,182 to reduce the bandgap and increase the absorption in the NIR region. Gao et al.183 synthesized TiO2 NPs and doped them with 15% W, increasing the free carrier concentration and leading to excellent absorbance in the NIR region. Then, WTO NPs were used to treat 4T1 by PTT and radiation therapy.

FIG. 5.

(a) Illustration of the resonance energy transfer process in Au@CuS NPs, in vitro western blot analysis of HSP70 and HO-1, thermal infrared images and H&E staining of Au@CuS NPs treated tumor. Reproduced with the permission from Chang et al., Nano Lett. 18(2), 886–897 (2018). Copyright 2018 American Chemical Society.184 (b) The FDTD simulation showing the electric field distributions of CuS and Pt-CuS under 808 nm laser irradiation for 7 min to reveal the mechanism of the photothermal effect along with experimental photothermal elevation profiles of CuS and Pt-CuS. Reproduced with the permission from Liang et al., Nano Lett. 19(6), 4134–4145 (2019). Copyright 2019 American Chemical Society.185 (c) Schematic illustration of photo/radiocatalytic ROS generation and immunogenic cell death for cancer immunotherapy by WO2.9-WSe2-PEG NPs under x-ray irradiation. Reproduced with the permission from Dong et al., ACS Nano 14(5), 5400–5416 (2020). Copyright 2020 American Chemical Society.168 

FIG. 5.

(a) Illustration of the resonance energy transfer process in Au@CuS NPs, in vitro western blot analysis of HSP70 and HO-1, thermal infrared images and H&E staining of Au@CuS NPs treated tumor. Reproduced with the permission from Chang et al., Nano Lett. 18(2), 886–897 (2018). Copyright 2018 American Chemical Society.184 (b) The FDTD simulation showing the electric field distributions of CuS and Pt-CuS under 808 nm laser irradiation for 7 min to reveal the mechanism of the photothermal effect along with experimental photothermal elevation profiles of CuS and Pt-CuS. Reproduced with the permission from Liang et al., Nano Lett. 19(6), 4134–4145 (2019). Copyright 2019 American Chemical Society.185 (c) Schematic illustration of photo/radiocatalytic ROS generation and immunogenic cell death for cancer immunotherapy by WO2.9-WSe2-PEG NPs under x-ray irradiation. Reproduced with the permission from Dong et al., ACS Nano 14(5), 5400–5416 (2020). Copyright 2020 American Chemical Society.168 

Close modal

Fusing plasmonic metals or some nonmetallic semiconductors with wideband gap semiconductors is an effective approach to effective capturing of NIR light. Generally, NIR active plasmonic photocatalysts consist of a combination of plasmonic elements and semiconductors, by which the coherent oscillation of electrons within the plasmonic component can be transformed into separated electrons and holes. Under NIR irradiation, the plasmonic materials convert the incident photons into different forms of energy, including heat.186 Good photothermal performance of semiconductors can be generally attained by selecting materials with good light-to-heat conversion efficiency (for example, CuS NPs H2 was also approved as a food additive by the US Food and Drug Administration). Additionally, incorporating metal NPs with high LSPR property into the semiconducting materials is an alternative strategy to improve their photothermal property. First, the LSPR effect of metal NPs significantly enhances the hyperthermia effect under light irradiation. Additionally, the metal NPs can act as an electron sink to scavenge excited electrons from the semiconductor, preventing electrons from backflow and thus enhancing the catalytic properties. Chang et al. reported that resonant energy transfer in Au@CuS NPs resulted in electron/hole pair generation and improved phototherapeutic performance.184 Nuo et al.187 reported Nb-doped TiO2 with increased photothermal properties for cancer therapy. Due to the substitution of Nb5+ on the Ti4+ site, Nb-TiO2 NPs have high free carrier concentration in the CB, increasing the localized surface plasmon resonance (LSPR) activity of NPs due to enhanced electron transport. The free carrier concentration in 5% Nb-doped NPs was calculated using the LSPR frequency as 2 × 1021 cm−3, greater than the minimum free carrier concentration required for satisfactory absorption in the NIR region (1021 cm−3). Bismuth sulfide is another semiconductor NP widely used in cancer nanomedicine. However, Bi2S3 has poor LSPR due to a low carrier concentration of 1015–1016 cm−3. Janus nanostructures have received significant attention for combining multiple functionalities and showing the all-in-one theranostic capability [Fig. 5(b)]. Cheng et al. found that the crystal lattices of Bi2S3 contain S atom-deficient regions, either vacant (S vacancy) or replaced by Bi atoms (BiS antisite). The S vacancies and BiS antisite work as deep-level defects in Bi2S3 nanomaterials, promoting nonradiative recombination to emit phonons and producing photothermal properties. Following this approach, they further enhanced the photothermal conversion efficiency of Bi2S3 nanorods by depositing Au NPs on the surfaces. With the formation of the heterojunction with Au, the higher Pauling electronegativity of Au (2.54) compared to Bi (2.02) causes the lattice S atoms in Bi2S3 to bind with the Au atoms, leading to additional BiS defects and promoting photothermal effect. This work demonstrates that increasing the number of deep-level defects could increase the photothermal efficiency of semiconducting nanomaterials. Both in vitro and in vivo studies proved that the photothermal efficacy of Bi2S3 NRs could be improved by growing gold nanodots on their surface to form Bi2S3-Au NRs as a safe and practical NIR light-triggered photothermal therapeutic agent.173 In another study, a high-density S-vacant Bi2S3−x-Au@HA was constructed. The S vacancies reduced the bandgap (1.21 eV) of the composite and also served as an e trap. Bi2S3−x-Au@HA when compared to Bi2S3−x, the absorption band of the former extended from the NIR-I region to the NIR-II region with broad and intense absorption peaks. Bi2S3−x-Au@HA has excellent photothermal conversion efficiency reaching 43.0%, which is significantly higher than other reported agents like CuS-Au-MnO2 NPs (28.0%), Gd/CuS nanogel (26.7%), and CuS/MnO2 NPs (30.17%). Furthermore, it was even superior to photothermal agents active in NIR-I, such as CuS@Cu2S@Au nanohybrid (35%).58 A novel Bi2S3/titanium carbide (Ti3C2) two-dimensional nanoheterostructure was designed by Jiang et al., who found that compared to simple Bi2S3 NPs, Bi2S3/Ti3C2-TPP significantly extended absorption to the NIR region, enhanced photocatalytic activity due to higher photogenerated carrier separation and electron transfer efficiency, and effectively accumulated in tumor cells by targeting mitochondria. These heterostructures exhibited excellent capabilities in CT imaging, rendering them highly effective as theranostic agents.188 A p–n heterojunction of BiOCl-Bi2S3 that promotes the effective electron/hole separation, offering a promising option to improve PTT efficiency, was used to ablate a subcutaneous hepatoma. Pt-CuS Janus heterojunction was prepared for efficient PTT. The photothermal conversion efficiencies of Pt-CuS and CuS NPs were 34.5% and 23%, respectively.185 

Li et al.189 synthesized Pt-doped Prussian blue nanozymes for catalytic cancer therapy. To obtain high LSPR frequency, Pt (20 μg/mL in terms of PB) was doped in PB, shifting the maximum absorption peak from 720 to 810 nm and improving the photothermal property by 1.7 times. PtPB showed an excellent photothermal conversion efficiency of 58.2%. Deng et al.190 prepared a heterojunction of black phosphorus and FeSe2, in which the lone pair of electrons on the BP surface was covalently combined with FeSe2 through the P–Se bond to improve the particle photostability. Heterostructure BPs-FeSe2-PEG showed enhanced photothermal properties due to BP with prolonged separation of photoexcited electron/hole. This heterojunction also enhanced magnetic resonance imaging contrast due to the formation of hydrogen bonds while FeSe2 aggregated on BP. Xu et al.191 synthesized plasmonic heterostructure of Au and metalloporphyrin NPs to prevent rapid electron/hole recombination. The high LSPR frequency of Au caused a bathochromic shift in the absorption wavelength of the heterojunction to the NIR region. Upon US and NIR irradiation, the nanoheterostructure generating ROS and hyperthermia was combined with PA imaging for theranostic cancer treatment. Dong et al.168 combined two semiconductors to form a heterojunction of WO2.9 and WSe2 NPs to obtain the WO2.9-WSe2 heterojunction that utilized x-ray and mild PTT to cure cancer, in which mild PTT induced tumor cell death and increased blood flow to the tumor for improving radiation therapy [Fig. 5(c)]. The heterojunction structure showed good photocurrent due to prolonged time in the electron/hole recombination. The heterostructure also demonstrated good photothermal efficiency and improved photostability than WO2.9 alone, confirming the benefit of the heterojunction. Overall, doping metal NPs into semiconductor NPs with low free carrier concentration or forming heterojunction between semiconductors to prevent electron/hole recombination improves absorption in the NIR region and enhances their photothermal efficiency.

Gas therapy is a promising cancer treatment strategy due to its high therapeutic efficacy and few side effects. However, a major challenge hindering its clinical translation is the lack of control over the distribution of gas molecules. Systemic administration often leads to nonspecific delivery, resulting in significant off-target toxicity to healthy tissues. Therefore, targeted delivery and on-demand release of therapeutic gas molecules at the tumor site are required for clinically successful gas therapy. The rapid advancement in nanotechnology made gas precision treatment possible via on-target delivery, reducing the risk of undesired toxicity. Recently, semiconductor heterojunction NPs have been employed to generate various gases in TME. The highly reductive electron (≥1.23 eV for H2 evolution from H+) generated after excitations can reduce endogenous CO2, H+, and NO3 into CO, H2, and NO, respectively, offering opportunities for gas therapy.192–195 Meanwhile, the hole generated can oxidize H2O2 (E [H2O2/O2] = 0.69 V vs NHE) and water to generate O2, which can be employed for hypoxia relief.

1. Hydrogen gas

H2, often considered an unreactive gas, is FDA-approved as a food additive due to its high biosafety and antioxidative properties. H2 gas exhibits anti-inflammatory properties by selectively eliminating cytotoxic free radicals, such as •OH and peroxynitrite (ONOO), from diseased cells while keeping the functions of healthy cells unaffected. This targeted action shows its potential in treating inflammation-related diseases, including cancer, atherosclerosis, stroke, and arthritis.196 While the anti-inflammatory action of H2 eliminates •OH and peroxynitrite (ONOO) in the affected cells, the physiological functions of healthy cells are unaffected.196 However, in tumors, the complete understanding of the apoptotic ability and cellular pathway modulation of H2 is yet to be clarified. Previous studies reported that the low-concentration H2 gas could modulate the inflammatory microenvironment. Meanwhile, its high concentrations could induce mitochondrial damage and disrupt redox homeostasis, causing cancer cell dysfunction or apoptosis.52,197,198 However, the lesion site, duration, and concentration of H2 determine the therapeutic efficiency. The practiced delivery method of H2 is direct inhalation, oral consumption of H2-rich water or capsules, and H2 bathing. The drawback of these approaches is their limited delivery to tumors that are challenging to access due to the high diffusion and low solubility of the gaseous molecule. H2 therapy combined with conventional methods to relieve harmful side effects is challenging, requiring optimization of therapeutic protocols. H2 nanomedicine is a potential method that introduces nontoxic nanocatalysts into lesions, thereby producing H2 in situ by triggering water-splitting reactions. Long-term H2 release is challenging and has limited clinical translation, underscoring the need for advanced research. In the energy sector, photocatalytic H2 production technology is thriving, while in practice, materials for catalyzing H2 production are mostly in the visible light spectrum.

In theory, endogenous H+/H2O can be reduced to generate H2 as long as CB edge of the semiconductor nanocatalyst is more negative than the redox potential of H+/H2 (vs NHE, 0 V, pH 0). GSH and lactate, which are generally overexpressed in TME can serve as a natural hole-sacrificing agents. They can deplete the holes on the surface of the catalyst consequently reducing the possibility of electron–hole recombination and enhancing catalytic H2 production. Zhao et al. developed an NIR-activatable Z-scheme SnS1.68-WO2.41 nanocatalyst that aligns with the concept of electron/hole in which catalysts generate hole as a therapeutic methodology. In a one-pot synthesis method, WO2.41 nanodots were heterogeneously grown over the surfaces of SnS1.68 nanoplates.52 Characterization of semiconductor heterojunction band alignment is indeed essential, as it directly influences the catalytic properties. Several techniques like x-ray photoelectron spectroscopy for evaluating the VB offset, ultraviolet photoelectron spectroscopy to assess the work function and VB maximum, photoluminescence spectroscopy for analyzing the recombination of electron–hole pairs, diffuse reflectance spectroscopy for the bandgap determination, Mott-Schottky analysis providing valuable information like flatband potential, and type of semiconductor (n-type or p-type) are some of them. The Kubelka–Munk function was incorporated into the reflectance data to obtain a plot from which the Eg of the materials was estimated. The Eg of SnS and WO in SnS1.68-WO2.41 heterojunction was found to be 1.49 and 2.43 eV, respectively.52 By narrowing the bandgap of SnS1.68, the NIR photoelectric conversion, GSH oxidation, and H2 production rate were enhanced. The NIR-photocurrent measurement showed that the current density of SnS1.68-WO2.41 is higher than that of SnS2. This suggests that under NIR irradiation, a large amount of hot electrons from WO2.41 are injected into conjugated SnS1.68 nanoplates, thereby enhancing NIR-photocatalytic reduction for H2 generation. This was demonstrated in a 4T1 breast cancer model where the nanocatalyst oxidized intratumoral overexpressed GSH and generated H2 on irradiation in a controlled and sustained manner. With the consumption of GSH and H2, the intratumoral redox balance was disturbed, reducing cancer cell energy and subsequently inducing DNA damage. Similarly, Pt-Bi2S3 Schottky junction NPs were developed and employed to produce a large amount of H2 as a gas therapy by inducing severe mitochondrial dysfunction and cell death on US irradiation–triggered activation.16 The authors used Mott–Schottky analysis and UV–vis diffuse reflectance spectroscopy to evaluate the band structure of the heterojunction NPs. The Eg of Bi2S3 nanorods and Pt-Bi2S3 heterojunctions were found to be 1.1 and 1.2 eV while the CB potentials for Bi2S3 and Pt-Bi2S3 NPs determined from the Mott–Schottky calculations, were −0.22 and −0.01 V, respectively. This leads to 0.89 V (1.19 V) for the VB potential of Bi2S3 (Pt-Bi2S3) NPs. This experimental analysis demonstrated that the VB of the heterojunction material is sufficient to oxidize GSH, while the CB potential showed the capability of reducing H+ to H2. Meng et al.156 utilized metallic Bi and a semiconductor BiOx to form a Schottky-type junction nanostructure for the reduction of intratumoral H+ to generate H2 with 808 nm laser, exploring the photovoltaic effect of the heterojunction [Fig. 6(a)]. Taking inspiration from H2 generation via solar-powered water splitting for clean energy applications, researchers developed an engineered red polymeric carbon nitride is characterized by a relatively narrow bandgap of approximately 1.71 eV alongside strong light absorption in the NIR. These characteristics facilitated efficient photocatalytic H2 evolution from water under NIR irradiation, resulting in the highest apparent quantum efficiency observed among the studied materials.186 Another group prepared Cu2+-anchored carbon photocatalysts by doping carbon dots (CDs) into carbon nitride (CN) in the presence of urea to produce CDCN by a thermal decomposition method. Then, they anchored it with Cu2+ to generate Cu@CDCN. Under visible light irradiation, photoexcited electrons split water into H2 and H2O2 through a two-electron reduction pathway. Interestingly, upon light irradiation, Cu2+ anchored on the nanocomposites was reduced to Cu+, activating intracellular cuproptosis, ATP and GSH depletion, severe lipid peroxidation, and aggregation of lipoylation proteins [Fig. 6(b)], thereby effectively inhibiting the tumor growth in tumor-bearing mice.199 The penetration of visible light to the tumor site is minimal. However, despite this limitation, NPs demonstrated notable photocatalytic efficacy both in vivo and in vitro, possibly owing to their higher efficiency. X-rays would be preferred to enhance penetration and photocatalytic efficacy in tumor tissues. Following x-ray irradiation, residual scintillation light persists temporarily, offering an opportunity for improved photocatalytic performance. Utilizing the excessive energy from x-rays and afterglow to sustain H2 production after irradiation cessation holds significant promise for catalytic applications within the tumor. A radiocatalytic nanotherapeutic system was developed to attain the synergistic effect of H2 gas therapy with radiotherapy using dumbbell-shaped Au-TiO2 heterostructure coated with long afterglow Co-A/Cu/ZnS NPs (Au-TiO2@ZnS). Under x-ray irradiation, the electrons and holes were produced in TiO2, and the excited electrons moved to the Au nanorod due to the p-type antiblocking layer, reducing recombination. Immediately after x-ray exposure, visible light, emitted at 578 nm from the afterglow material deposited on the Au-TiO2 heterostructure over 10 min, activated the Au nanorods for further hot electron generation. The hot electrons generated smoothly flowed across the metal-semiconductor Schottky barrier to the CB of TiO2. Separate production of hot electrons by the x-ray irradiation on TiO2 and exposure of the Au nanorods to visible light prolonged the H2 production and enhanced the efficiency of electron-catalyzed reduction reaction of protons (H+) in the tumor. Meanwhile, the Au nanorods improved the radiosensitivity and acted as a PA imaging agent. Altogether, the Au-TiO2@ZnS heterojunction structure significantly improved and prolonged the generation of H2 gas upon x-ray exposure for effective tumor therapy.200 

FIG. 6.

(a) Illustration of the charge transfer mechanism for catalytic ROS and H2 generation by Bi/BiOx lateral nano-heterostructure under normoxic and hypoxic conditions. Reproduced with the permission from Qiu et al., Adv. Mater. 33(49), 2102562 (2021). Copyright 2021 John Wiley and Sons.156 (b) Photocatalytic mechanism of visible light driven combinatorial H2 therapy and cuproptosis by Cu@CDCN nanosheets. Reproduced with the permission from Ding et al., Angew. Chem., Int. Ed. 62(44), 202311549 (2023). Copyright 2023 John Wiley and Sons.199 (c) Sonocatalytic charge transfer mechanism in Z-scheme FeOCl/FeOOH nanosheets for O2 generation and H2O2 generation for effective Fenton-like reaction. Reproduced with the permission from Kang et al., Nat. Commun. 13(1), 2425 (2022). Copyright 2022 Authors, licensed under Creative Commons Attribution (CC BY) license.201 (d) Schematic representation of photocatalytic mechanism in FeS2/CoS2@PEG nanosheets for the production of O2 in combination with PTT/PDT/CDT and immunotherapy for effective tumor treatment. Reproduced with the permission from Wang et al., J. Colloid Interface Sci. 625, 145–157 (2022). Copyright 2022 Elsevier.202 (e) TEM image and HR TEM image of BiOBr@Bi2S3 showing the close contact of BiOBr and Bi2S3 nanorod for the effective formation of heterojunction. Reproduced with the permission from Yuan et al., Adv. Sci. 11, 2308546 (2024). Copyright 2024 John Wiley and Sons.174 

FIG. 6.

(a) Illustration of the charge transfer mechanism for catalytic ROS and H2 generation by Bi/BiOx lateral nano-heterostructure under normoxic and hypoxic conditions. Reproduced with the permission from Qiu et al., Adv. Mater. 33(49), 2102562 (2021). Copyright 2021 John Wiley and Sons.156 (b) Photocatalytic mechanism of visible light driven combinatorial H2 therapy and cuproptosis by Cu@CDCN nanosheets. Reproduced with the permission from Ding et al., Angew. Chem., Int. Ed. 62(44), 202311549 (2023). Copyright 2023 John Wiley and Sons.199 (c) Sonocatalytic charge transfer mechanism in Z-scheme FeOCl/FeOOH nanosheets for O2 generation and H2O2 generation for effective Fenton-like reaction. Reproduced with the permission from Kang et al., Nat. Commun. 13(1), 2425 (2022). Copyright 2022 Authors, licensed under Creative Commons Attribution (CC BY) license.201 (d) Schematic representation of photocatalytic mechanism in FeS2/CoS2@PEG nanosheets for the production of O2 in combination with PTT/PDT/CDT and immunotherapy for effective tumor treatment. Reproduced with the permission from Wang et al., J. Colloid Interface Sci. 625, 145–157 (2022). Copyright 2022 Elsevier.202 (e) TEM image and HR TEM image of BiOBr@Bi2S3 showing the close contact of BiOBr and Bi2S3 nanorod for the effective formation of heterojunction. Reproduced with the permission from Yuan et al., Adv. Sci. 11, 2308546 (2024). Copyright 2024 John Wiley and Sons.174 

Close modal

2. Oxygen gas

Hypoxia is a negative prognostic factor for solid tumors. It enhances tumor invasiveness and reduces tumor sensitivity to treatments, including PDT, radiotherapy, and chemotherapy. Supplying O2 is one of the most adopted techniques to alleviate tumor hypoxia. Several CAT-like nanozymes that decomposed intratumoral H2O2 to generate O2 have been developed and studied for tumor therapy. Recently, photo/sonocatalytic water splitting has achieved significant results in generating on-demand intratumoral O2 to alleviate hypoxia. Under exogenous stimulation, heterojunction catalysts can supply O2 through water splitting and produce ROS by reducing the generated O2, rendering them promising for hypoxic tumor therapy.203 Rather than water splitting, H2O2 has also been used as a substrate for O2 production in tumors by semiconducting heterojunction catalysis.16 

Photoactive NPs can usually generate O2 based on their unique electronic structures. Specifically, photoactive NPs with a narrow bandgap can be excited by NIR laser to generate electron and hole in CB and VB. These holes can react with water (H2O) to produce O2 (2H2O + 4H+ → O2 + 4H+, 1.23 V vs NHE). O2 self-supply and ROS-producing semiconductor heterojunction NPs have been used to treat hypoxic tumors. A 2-dimensional interplanar Z-schemed FeOCl/FeOOH nanosheet was developed and adopted for intratumoral oxidation of H2O to generate O2 under US stimulation [Fig. 6(c)]. The 2-dimensional Z-scheme heterojunction with good band alignment was prepared by hydrothermal process, liquid exfoliation, and NaOH etching. The authors evaluated the effect of alkali treatment on the FeOCl/FeOOH nanosheet by spin-polarized density functional theory calculations. The fully optimized structures revealed that the Fe-Cl bond length in FeOCl is 1.72 Å, while in FeOOH, the Fe–O and O–H bond lengths are 2.03 and 0.98 Å, respectively, with a Fe–O–H bond angle of 111.5°. Substituting −Cl with −OH elongated the Fe–Cl bond from 1.72 to 1.83 Å. The authors further evaluated the interaction of water molecules with FeOCl and FeOOH. The calculated results reveal that the adsorption characteristics of water molecules with the FeOOH surface are closer than that with FeOCl, thereby indicating significant improvement in the hydrophilic properties of the material and increasing interlayer water molecule infiltration. The FeOCl/FeOOH nanosheets receiving the US stimulation induced charge separation in both FeOCl and FeOOH. The hot electrons in the CB of FeOCl recombined with the holes on the VB of FeOOH, leaving the holes in the VB of FeOCl with high oxidation potential. Meanwhile, the hot electrons left on the CB of FeOOH without recombination otherwise showed high reduction potential. This phenomenon created a built-in electric field along with the formation of a Schottky barrier at their interface, rendering the active holes of FeOCl highly effective in catalyzing H2O oxidation and O2 production.201 Yang et al.204 developed a novel poly(vinylpyrrolidone)-modified BiFeO3/Bi2WO6 with a p–n type heterojunction for the reconstruction of the immunosuppressive TME. Irradiation with 660 nm light generates holes that trigger intratumoral H2O2 decomposition, boosting intratumoral O2 levels and alleviating tumor hypoxia. This multi-pronged effect enhances PDT and radiotherapy, while promoting a switch in tumor-associated macrophages from immunosuppressive M2 to immunostimulatory M1 phenotype. Additionally, BFO/BWO-PVP nanoparticles demonstrate excellent performance as a CT imaging contrast agent. However, the depth penetration of 660 nm into the tumor remained an obstacle to achieving maximum efficacy. In another work, An NIR-II-responsive FeS2/CoS2@PEG (FCs@PEG) Z-scheme heterostructure was developed for intracellular O2 generation to alleviate hypoxia in cancer treatment [Fig. 6(d)]. The formation of the heterostructure by FeS2 and CoS2 established a Z-scheme heterojunction through band bending, resulting in the migration of electrons from the CB of FeS2 to the VB of CoS2 under NIR-II activation. The holes left in the VB of FeS2 exhibited an oxidation potential of 1.31 eV, sufficing to oxidize H2O to O2.202 Yuan et al.16 designed Bi2S3@Pt Schottky heterostructure for sonocatalytic tumor therapy. Pt NPs decorated over Bi2S3 nanorods catalyzed the excessive H2O2 in the TME, generating O2. Furthermore, the sonoexcited holes could also oxidize H2O2 to generate O2, alleviating tumor hypoxia. In addition to transforming tumor-associated M2 macrophages into M1 phenotype, the production of O2 alleviated the hypoxic TME and downregulated the expression of HIF-1α to inhibit tumor angiogenesis, demonstrating a successful reversal of immunosuppressive TME. Overall, O2 is produced by heterojunction catalysts via H2O or H2O2 oxidation in the TME by reactive electron/hole.

3. Carbon monoxide

Carbon monoxide (CO) is an endogenous signaling molecule studied in preclinical studies using multiple experimental models. CO is a promising therapeutic agent that benefits the treatment of diseases, including cancer, by increasing its endogenous production and exogenous delivery.205 Since CO has a high affinity toward hemoglobin but no tumor selectivity, direct administration poses a threat. Therefore, sophisticated techniques should be implemented for in situ CO generation in tumor tissues. Many approaches are under investigation, among which photocatalytic CO2 reduction has emerged as a promising method. CO produced by photocatalysis sensitizes cancer cells to chemotherapy while shielding healthy cells. In a recent study, Ag3PO4-doped CD-decorated C3N4 NPs and functionalized with histidine-rich peptide (HisAgCCNs) were prepared as a Z-scheme system to convert CO2 to CO under 630 nm light irradiation. The CO production rate of HisAgCCNs reached 65 μmol/h−1/g mat−1, significantly increasing the cytotoxicity of anticancer drug DOX by 70%.206 However, the limited penetration of the light, low internal CO2 concentration, and the strong dependence on internal CO2 hinder efficient CO2 photoreduction in vivo. Thus, developing novel photocatalytic CO generation systems to overcome these problems is essential. Wang et al. developed a bicarbonate-conjugated nanocatalytic system to generate CO gas for tumor therapy. They prepared defective WO3 nanosheets (DW) decorated with lipoic acid–conjugated dopamine (LA-DPA) through W-S bonds. Then, they conjugated ferric ions as a coordination center to bridge both DPA and bicarbonate and termed it P@DW/BC. Upon 808-nm NIR laser irradiation, DW nanosheets serve as a PTT agent and a catalyst to convert CO2 to CO. The CO-mediated anti-inflammatory effect greatly improved the survival rate of mice after PTT.207 

US-assisted catalysis has recently become a mainstream treatment strategy primarily due to its operability and penetration depth. Recently, BiOCl/Bi2O3 was developed by wet chemical method by integrating interplanar heterojunction synthesis and 2D ultrathin heterojunction exfoliation in single step. The active catalytic sites located at the CB (−1.1 eV) of BiOCl could catalyze the generation of •O2 (O2 + e → •O2) and CO (CO2 + 2H+ + 2e → CO + H2O). Meanwhile, the VB (0.4 eV) of Bi2O3 could catalyze the generation of •OH (H2O + h+ → •OH + H+). Based on US excitation, BiOCl NPs and BiOCl/Bi2O3 nanosheets catalyze CO2 reduction, and BiOCl/Bi2O3 nanosheets exhibited more efficient CO generation than BiOCl NPs. Enhanced CO yields were observed in hypoxia.208 The CB values of BiOCl and Bi2O3 were −1.1 and 0.4 eV, respectively, and the E0 reduction of CO2/CO (−0.53 eV) was lower than that of the CB of BiOCl. In addition to material properties such as crystallinity, porosity, surface area, and surface energy the orientation and interfacial connection of the constituent materials within a heterojunction should also be considered as one of the critical factors.209,210 Xiong et al. fabricated three different heterojunction interfaces with n-type α-Fe2O3 and p-type Bi2O3.210 The three heterojunctions fabricated were ring-to-face (I), face-to-face (II), and rod-to-face (III), respectively. The photocatalytic ability of the heterojunctions was superior to α-Fe2O3, and the photocatalytic ability among the heterojunctions was on the order of face-to-face > rod-to-face > ring-to-face. The authors found out the reason behind the superior catalytic property as the interfacial connection of face-to-face was superior to that of rod-to-face and ring-to-face by electrochemical impedance spectroscopy. This study has highlighted that constructing an efficient heterojunction for photocatalysis requires matching the band positions and considering the interfacial connection between the materials. In the work of Yuan et al.,174 a BiOBr@Bi2S3 S-scheme heterojunction was developed for US-triggered CO therapy. Bi2S3 nanorods were grown on the edges of BiOBr nanosheets for better interfacial connection and superior catalytic ability [Fig. 6(e)]. BiOBr nanosheets were prepared by hot injection method and the length of the Bi2S3 nanorods were adjusted by altering the concentration of thioacetamide. Similar to type II heterojunctions, the S-scheme heterojunction also consists of both reduction and oxidation catalysts but entirely differs in the charge transfer route from the type II counterparts. In the S-scheme heterojunction, the reduction catalyst has higher CB and VB, as well as a smaller work function, compared to the oxidation catalyst. When these two semiconductors are in close contact, electrons in the reduction catalyst spontaneously diffuse to the oxidation catalyst, creating a significant electron gradient between the electron depletion layer and electron accumulation layer near the interface in the reduction catalyst and oxidation catalyst, respectively. This gradient accounts for the Fermi level bending in the interface region [Fig. 6(e)].174 In the BiOBr@Bi2S3 S-scheme heterojunction, the CB of Bi2S3 (−0.49 V) acted as the reduction site for electron-catalyzed conversion of intracellular CO2 to CO for CO therapy, while the VB of BiOBr (2.5 V) served as the oxidation site to convert water into cytotoxic •OH radicals. Thus, the attained CO gas effectively damaged mitochondria and reduced intracellular energy production, synergistically inducing tumor cell apoptosis in cooperation with cytotoxic •OH radicals.

Despite considerable progress in patient care, the global incidence of various cancer types continues to grow. Developing safer and more efficient anti-cancer treatment approaches are of great demand. In recent decades, nanotechnology has emerged as a promising and innovative medical approach for cancer diagnosis and treatment. In the hunt for cancer, nanomedicine has been diversified into different fields. Among them, heterojunction tumor nanomedicine is a burgeoning field that involves multiple disciplines. Herein, we summarize and explain the recent developments and achievements in exogenous-activated nano-heterojunction for tumor treatment. The first part of this review article includes the fundamental aspects and working principles of different heterojunctions. The latter consists of the summarized report on applying literature in heterojunction semiconductor nanocatalysts as cancer theranostics with specific examples and detailed mechanisms. In short, Exogeneous stimuli activated heterojunction catalytic therapy activates in situ tumor-localized catalytic reactions through nontoxic or low-toxicity nanocatalysts in response to external stimuli (light, US, or temperature), converting substrates in the TME into effective therapeutic agents to induce tumor cell death without affecting normal tissues. This emerging technology offers promising avenues for alternative therapeutic modalities, such as PTT, radiotherapy, gas therapy, and starvation therapy, to cancer treatment. However, several factors control the efficacy of heterojunctions. For the effective charge separation and an inbuilt electric field formation at the interface, a direct contact between the components is essential. For effective redox reactions inside the TME, The CB and VB positions plays a crucial role in the effective transfer of electron/hole and are closely associated with the type of heterojunction achieved. The built-in electric field and potential barriers at the heterojunction interface not only rely on the band positions but also other factors, such as semiconductivity (n- or p-type), work function and Fermi level. The advancement of research in this field is implementing heterojunction nanoparticles as a promising new candidate for cancer treatment. However, the clinical translation of heterojunction catalyst is in its infancy and requires further substantial progress since severe concerns exist.

Below is a list summarizing the factors that need to be addressed to maximize the effectiveness of heterojunction-based catalytic systems from laboratory scale to clinical application.

Improper band alignment can create barriers to electron and/or hole transfer, leading to the recombination of charge carriers at the interface and reducing the photocatalytic efficiency. Achieving the desired band alignment often requires careful selection of materials and precise control of their interface. This alignment ensures that the energy levels of the materials involved are appropriately matched, enabling efficient charge transfer. When the energy bands of different materials in the heterojunction are favorably aligned, it facilitates the separation of photogenerated electron/hole pairs, which is crucial for the catalytic activity. Second, efficient migration of these photogenerated electron/hole pairs is vital for catalytic performance. Effective transport of charge carriers within the heterojunction ensures they reach the catalytic sites promptly, minimizing recombination losses and maximizing the utilization of photogenerated charges for catalysis. The orientation and interfacial contact between the two components forming the heterojunction should be considered along with the band matching to enhance catalytic performance. The prior criteria for constructing a desirable heterojunction include careful selection of materials and preparation methods with appropriate band structure and bandgap matching. Utilization of theoretical models (for example, density functional theory calculation) and simulation of heterojunctions to predict and optimize the band alignment before experimental implementation would be helpful in the construction of an effective heterojunction.

One of the main challenges in fabricating heterojunctions is the integration of materials of different crystal structures as lattice mismatches and defects at the interfaces often occur. These defects can act as recombination centers for charge carriers, reducing the efficiency of the heterojunction. To address this issue, it requires precise control over the synthesis conditions, including temperature, pressure, and techniques such as atomic layer deposition or molecular beam epitaxy, which allow for the growth of high-quality heterojunction with minimal defects. The surface properties, porosity, and crystallinity of heterojunction materials also play crucial roles in determining their effectiveness. Surface properties, including surface charge and functional groups, influence the interactions with biological environments and the stability of the NPs. Porosity affects the diffusion of reactants and products and the availability of active sites for catalysis. Crystallinity, on the other hand, impacts the charge mobility and the ability of the heterojunction to maintain efficient charge separation. High crystallinity typically enhances charge transport, while poor crystallinity often induces recombination losses. However, recent research has shown that using amorphous materials fused with crystalline materials exhibited superior efficacy. While the design and fabrication of heterojunction nanoparticles for cancer therapy present several challenges, particularly in integrating different crystal structures, advancements in fabrication techniques and a better understanding of the material properties can lead to successful development of highly effective therapeutic agents. The interactions between catalysts and reactants predominantly occurs on the surface or at interfaces, rendering material efficiency partially reliant on specific surface area. A larger surface area provides more active sites for reacting with absorbed water and other biomolecules to execute redox reactions. A complex interplay of physical and chemical properties influences the overall efficiency of heterojunction photocatalytic systems. While it is well-established that these factors impact photocatalytic performance in various ways, comprehensive and detailed analyses that encompass all these aspects are still lacking.

Additionally, ensuring the size, shape, and uniformity of heterojunctions is paramount for consistent and reliable catalytic performance, intratumoral accumulation, and anticancer activity. Irregularities or variations in the shape and size of heterojunction components can cause differences in catalytic activity across the material, resulting in reduced overall efficiency. Meanwhile, the NP size is a critical factor in determining the blood circulation time and bioelimination, the important parameters strongly associated with the material therapeutic efficacy. NP size ranging from 20 to 200 nm is widely accepted since these nanoparticles can passively extravasate through the leaky tumor vasculature and preferentially accumulate in tumors due to the enhanced permeability and retention effects. The stability of heterojunction is crucial for the application in cancer therapy. Issues concerning aggregation or interactions with biological components could impair the heterojunction's effectiveness in targeted applications, such as tumor treatment. Surface functionalization is a viable solution to enhance the stability of heterojunctions during systemic circulation. Modifying the surface of the heterojunctions with biocompatible polymers or targeting ligands and proteins could improve its resistance to degradation, prevent aggregation, and facilitate specific interactions with target cells or tissues. The interface between the heterojunction materials should remain intact and effective during the treatment period inside the TME. Any degradation at the interface, particularly under acidic tumor conditions, can significantly compromise the overall performance of the heterojunction in tumor treatment applications, which should be considered while designing heterojunctions.

The limited ability of external stimuli, such as light, to penetrate biological tissues often necessitates a significant amount of nanocatalysts for optimal efficacy. However, this approach can introduce various biosafety concerns. Consequently, there is a growing interest in combining different therapies, such as PDT and SDT, or radiotherapy, to utilize their respective advantages. The development of multifunctional sensitizing nanomaterials for sonophototherapy holds considerable promise. By integrating PDT and SDT into a collaborative treatment approach, sonophototherapy not only addresses the issue of low tissue penetration but also achieves therapeutic efficacy with a reduced dose of nanomaterials. This combined therapy is expected to be a significant research focus in the future, offering enhanced therapeutic outcomes while mitigating safety concerns. X-rays have significant tissue penetration capability, allowing them to reach deep-seated tumors. They can penetrate the tumor to activate the catalytic nanoparticles, making radiocatalysis an effective strategy against deeply set tumors. However, high doses of x-rays can cause collateral damage to adjacent healthy tissues during the cancer treatment. Consequently, developing nanoheterojunctions with high catalytic efficiency that necessitate low dose x-ray may offer a promising solution.

Nanoheterojunction materials, often entailing metal-based semiconductors, are constrained to laboratory research alone due to toxicity and biosafety issues. One of the major requirements for the clinical translation of heterojunction catalysts is to ensure biocompatibility in particular by means of minimizing off-target effects. To reduce off-target, various surface modifications with biorecognizable ligands have been employed to impart active targeting. Despite acquiring target specificity, selectivity and sensitivity must be enhanced for preferred biodistribution. For specific control over the location and timing of activating therapeutic agents, exogenous stimulation helps in the controlled activation, thereby minimizing off-target toxicity. In addition, more studies should be carried out on low-dose heterojunction materials with high elimination rate from the body. Currently, the biosafety study centers on the short-term toxic effects in the laboratories. However, a substantial and comprehensive study of cytotoxicity, genotoxicity, and long-term effects is required to take this promising approach further. The unpredictable variation in the level of substrates undergoing catalytic redox reactions as the onset of the treatment remains a great challenge in tumor-specific and patient-specific therapy. Due to their wideband gap, certain heterojunction materials generate high-energy electrons and holes. A major concern ensues with their broad and nonspecific reactivity with proteins and biological molecules. Extensive research and new material design on the specificity of heterojunction catalysts are important to divulge the potential of heterojunction in biomedicine. Other challenges are the high cost and complexity of manufacturing heterojunctions. The fabrication process often requires specialized equipment and techniques, which can be difficult to scale up for mass production.

Therefore, multidimensional modulation of heterojunction nanocatalyst structures and properties to maximize quantum yields and catalytic efficiencies under limited external stimulative energy and substrate types and concentrations is key to elevating the clinical potential of nanocatalytic therapy. By addressing these requirements, researchers can enhance the catalytic efficacy of heterojunction-based systems, unlocking their full potential for various biomedical applications.

We acknowledge the financial support by the National Science and Technology Council, Taiwan, NSTC 111-2314-B-007-001-MY3 and NSTC 110-2221-E-007-018-MY3.

The authors have no conflicts to disclose.

Ethics approval is not required.

Arjun Sabu: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Supervision (equal); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal). Manoj Kandel: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Ritwick Ranjan Sarma: Data curation (equal); Formal analysis (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Lakshminarayan Ramesan: Data curation (equal); Formal analysis (equal); Software (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Ekta Roy: Conceptualization (equal); Methodology (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Ramalingam Sharmila: Conceptualization (equal); Supervision (equal); Visualization (equal); Writing – review & editing (equal). Hsin-Cheng Chiu: Conceptualization (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Project administration (equal); Resources (equal); Supervision (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
A.
Szöke
,
W. G.
Scott
, and
J.
Hajdu
, “
Catalysis, evolution and life
,”
FEBS Lett.
553
(
1–2
),
18
20
(
2003
).
2.
N.
Das
and
C.
Maity
, “
Switchable aqueous catalytic systems for organic transformations
,”
Commun. Chem.
5
(
1
),
115
(
2022
).
3.
T.
Sobanski
,
M.
Rose
,
A.
Suraweera
,
K.
O'Byrne
,
D. J.
Richard
, and
E.
Bolderson
, “
Cell metabolism and DNA repair pathways: Implications for cancer therapy
,”
Front. Cell Dev. Biol.
9
,
633305
(
2021
).
4.
R.
Zhang
,
X.
Yan
, and
K.
Fan
, “
nanozymes inspired by natural enzymes
,”
Acc. Mater. Res.
2
(
7
),
534
547
(
2021
).
5.
R.
Zhang
,
K.
Fan
, and
X.
Yan
, “
Nanozymes: Created by learning from nature
,”
Sci. China Life Sci.
63
(
8
),
1183
1200
(
2020
).
6.
S.
Zhang
,
Y.
Li
,
S.
Sun
,
L.
Liu
,
X.
Mu
,
S.
Liu
,
M.
Jiao
,
X.
Chen
,
K.
Chen
,
H.
Ma
,
T.
Li
,
X.
Liu
,
H.
Wang
,
J.
Zhang
,
J.
Yang
, and
X.-D.
Zhang
, “
Single-atom nanozymes catalytically surpassing naturally occurring enzymes as sustained stitching for brain trauma
,”
Nat. Commun.
13
(
1
),
4744
(
2022
).
7.
X.
Meng
,
H.
Fan
,
L.
Chen
,
J.
He
,
C.
Hong
,
J.
Xie
,
Y.
Hou
,
K.
Wang
,
X.
Gao
,
L.
Gao
,
X.
Yan
, and
K.
Fan
, “
Ultrasmall metal alloy nanozymes mimicking neutrophil enzymatic cascades for tumor catalytic therapy
,”
Nat. Commun.
15
(
1
),
1626
(
2024
).
8.
Q.
An
,
C.
Sun
,
D.
Li
,
K.
Xu
,
J.
Guo
, and
C.
Wang
, “
Peroxidase-like activity of Fe3O4@Carbon nanoparticles enhances ascorbic acid-induced oxidative stress and selective damage to PC-3 prostate cancer cells
,”
ACS Appl. Mater. Interfaces
5
(
24
),
13248
13257
(
2013
).
9.
Y.
Ai
,
Z.
Hu
,
X.
Liang
,
H.
Sun
,
H.
Xin
, and
Q.
Liang
, “
Recent advances in nanozymes: From matters to bioapplications
,”
Adv. Funct. Mater.
32
(
14
),
2110432
(
2022
).
10.
X.
Li
,
Y.
Chen
,
Y.
Tao
,
L.
Shen
,
Z.
Xu
,
Z.
Bian
, and
H.
Li
, “
Challenges of photocatalysis and their coping strategies
,”
Chem. Catal.
2
(
6
),
1315
1345
(
2022
).
11.
A.
Balapure
,
J.
Ray Dutta
, and
R.
Ganesan
, “
Recent advances in semiconductor heterojunctions: A detailed review of the fundamentals of photocatalysis, charge transfer mechanism and materials
,”
RSC Appl. Interfaces
1
(
1
),
43
69
(
2024
).
12.
R.
Bazak
,
M.
Houri
,
S.
El Achy
,
S.
Kamel
, and
T.
Refaat
, “
Cancer active targeting by nanoparticles: A comprehensive review of literature
,”
J. Cancer Res. Clin. Oncol.
141
(
5
),
769
784
(
2015
).
13.
C.
Pucci
,
C.
Martinelli
, and
G.
Ciofani
, “
Innovative approaches for cancer treatment: Current perspectives and new challenges
,”
ecancer.
13
, 961 (
2019
).
14.
Q.
Huo
,
Q.
Yang
,
R. M.
Gorczynski
, and
T.
Li
, “
Oncological minimally invasive surgery
,”
J. Oncol.
2019
,
1
2
.
15.
H.
Wang
,
W.
Liu
,
X.
He
,
P.
Zhang
,
X.
Zhang
, and
Y.
Xie
, “
An excitonic perspective on low-dimensional semiconductors for photocatalysis
,”
J. Am. Chem. Soc.
142
(
33
),
14007
14022
(
2020
).
16.
M.
Yuan
,
S.
Liang
,
L.
Yang
,
F.
Li
,
B.
Liu
,
C.
Yang
,
Z.
Yang
,
Y.
Bian
,
P.
Ma
,
Z.
Cheng
, and
J.
Lin
, “
Rational design of platinum–bismuth sulfide schottky heterostructure for sonocatalysis‐mediated hydrogen therapy
,”
Adv. Mater.
35
(
10
),
2209589
(
2023
).
17.
S.
Dong
,
Y.
Dong
,
T.
Jia
,
S.
Liu
,
J.
Liu
,
D.
Yang
,
F.
He
,
S.
Gai
,
P.
Yang
, and
J.
Lin
, “
GSH‐depleted nanozymes with hyperthermia‐enhanced dual enzyme‐mimic activities for tumor nanocatalytic therapy
,”
Adv. Mater.
32
(
42
),
2002439
(
2020
).
18.
Q.
Wang
,
Z.
He
,
R.
Zhang
,
J.
Du
,
L.
Zhu
,
X.
Li
,
H.
Yang
,
Y.
Miao
, and
Y.
Li
, “
Carbon monoxide-based immunogenic cell death amplifier remodels the hypoxic microenvironment for tumor sono-immunotherapy
,”
Chem. Eng. J.
480
,
148269
(
2024
).
19.
S.
Kumari
,
A. K.
Badana
,
M. M.
G
,
S.
G
, and
R.
Malla
, “
Reactive oxygen species: A key constituent in cancer survival
,”
Biomarker Insights
13
, 1177271918755391 (
2018
).
20.
L.
Wang
,
X.
Huang
,
M.
Han
,
L.
Lyu
,
T.
Li
,
Y.
Gao
,
Q.
Zeng
, and
C.
Hu
, “
Efficient inhibition of photogenerated electron-hole recombination through persulfate activation and dual-pathway degradation of micropollutants over iron molybdate
,”
Appl. Catal. B
257
,
117904
(
2019
).
21.
Z.
Wang
,
Z.
Lin
,
S.
Shen
,
W.
Zhong
, and
S.
Cao
, “
Advances in designing heterojunction photocatalytic materials
,”
Chin. J. Catal.
42
(
5
),
710
730
(
2021
).
22.
C.
Pan
,
Z.
Mao
,
X.
Yuan
,
H.
Zhang
,
L.
Mei
, and
X.
Ji
, “
Heterojunction nanomedicine
,”
Adv. Sci.
9
(
11
),
2105747
(
2022
).
23.
I.
Ahmad
,
Y.
Zou
,
J.
Yan
,
Y.
Liu
,
S.
Shukrullah
,
M. Y.
Naz
,
H.
Hussain
,
W. Q.
Khan
, and
N. R.
Khalid
, “
Semiconductor photocatalysts: A critical review highlighting the various strategies to boost the photocatalytic performances for diverse applications
,”
Adv. Colloid Interface Sci.
311
,
102830
(
2023
).
24.
J.
Low
,
J.
Yu
,
M.
Jaroniec
,
S.
Wageh
, and
A. A.
Al‐Ghamdi
, “
Heterojunction photocatalysts
,”
Adv. Mater.
29
(
20
),
1601694
(
2017
).
25.
H.
Wang
,
L.
Zhang
,
Z.
Chen
,
J.
Hu
,
S.
Li
,
Z.
Wang
,
J.
Liu
, and
X.
Wang
, “
Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances
,”
Chem. Soc. Rev.
43
(
15
),
5234
(
2014
).
26.
S. J. A.
Moniz
,
S. A.
Shevlin
,
D. J.
Martin
,
Z.-X.
Guo
, and
J.
Tang
, “
Visible-light driven heterojunction photocatalysts for water splitting—A critical review
,”
Energy Environ. Sci.
8
(
3
),
731
759
(
2015
).
27.
F.
Li
,
G.
Zhu
,
J.
Jiang
,
L.
Yang
,
F.
Deng
,
Arramel
, and
X.
Li
, “
A review of updated S-scheme heterojunction photocatalysts
,”
J. Mater. Sci.
177
,
142
180
(
2024
).
28.
G.
Wang
,
S.
Lv
,
Y.
Shen
,
W.
Li
,
L.
Lin
, and
Z.
Li
, “
Advancements in heterojunction, cocatalyst, defect and morphology engineering of semiconductor oxide photocatalysts
,”
J. Materiomics
10
(
2
),
315
338
(
2024
).
29.
Y.
Sun
,
Z.
Jia
,
N.
Wan
, and
W.
Feng
, “
Biomimetic guided Bi2WO6/Bi2O3 vertical heterojunction with controllable microstructure for efficient photocatalysis
,”
Molecules
28
(
7
),
3123
(
2023
).
30.
S.
Li
,
W.
Xu
,
L.
Meng
,
W.
Tian
, and
L.
Li
, “
Recent progress on semiconductor heterojunction‐based photoanodes for photoelectrochemical water splitting
,”
Small Sci.
2
(
5
),
2100112
(
2022
).
31.
L.
Schumacher
and
R.
Marschall
, “
Recent advances in semiconductor heterojunctions and Z-schemes for photocatalytic hydrogen generation
,”
Top. Curr. Chem. (Cham)
380
(
6
),
53
(
2022
).
32.
T.
Shao
,
Y.
Chang
,
Z.
Li
,
Y.
Song
,
D.
Jin
,
J.
Gao
,
L.
Sun
, and
J.
Hou
, “
A chemically bonded and plasmonic Z-scheme junction for high-performance artificial photosynthesis of hydrogen peroxide
,”
J. Mater. Chem. A
11
(
3
),
1199
1207
(
2023
).
33.
R.
Marschall
, “
Semiconductor composites: Strategies for enhancing charge carrier separation to improve photocatalytic activity
,”
Adv. Funct. Mater.
24
(
17
),
2421
2440
(
2014
).
34.
J. K.
Hyun
,
S.
Zhang
, and
L. J.
Lauhon
, “
Nanowire heterostructures
,”
Annu. Rev. Mater. Res.
43
(
1
),
451
479
(
2013
).
35.
A. J.
Bard
, “
Photoelectrochemistry and heterogeneous photo-catalysis at semiconductors
,”
J. Photochem.
10
(
1
),
59
75
(
1979
).
36.
K.
Maeda
, “
Z-scheme water splitting using two different semiconductor photocatalysts
,”
ACS Catal.
3
(
7
),
1486
1503
(
2013
).
37.
M. S.
Athar
and
M.
Muneer
, “
Construction of visible light responsive GO-stimulated In2O3/BiVO4 Z-scheme heterostructure: Photocatalytic degradation of dye & antibiotics, anticancer activity and toxicity elimination
,”
J. Alloys Compd.
980
,
173385
(
2024
).
38.
C.
Feng
,
J.
Hu
,
C.
Xiao
,
J.
Yang
,
B.
Xin
,
Z.
Jia
,
S.
Zhang
,
G.
Tian
,
D.
Zhang
,
L.
Geng
,
L.
Yan
,
L.
Wang
, and
B.
Geng
, “
Shell-core COF@Co3O4 Z-scheme heterojunctions for triple amplification of oxidative stress to enhance nanocatalytic-sonodynamic tumor therapy
,”
Chem. Eng. J.
460
,
141874
(
2023
).
39.
Y.
Yuan
,
R.
Guo
,
L.
Hong
,
X.
Ji
,
Z.
Lin
,
Z.
Li
, and
W.
Pan
, “
A review of metal oxide-based Z-scheme heterojunction photocatalysts: Actualities and developments
,”
Mater. Today Energy
21
,
100829
(
2021
).
40.
X.
Li
,
C.
Garlisi
,
Q.
Guan
,
S.
Anwer
,
K.
Al-Ali
,
G.
Palmisano
, and
L.
Zheng
, “
A review of material aspects in developing direct Z-scheme photocatalysts
,”
Mater. Today
47
,
75
107
(
2021
).
41.
C.
Yang
,
X.
Yu
,
S.
Heißler
,
A.
Nefedov
,
S.
Colussi
,
J.
Llorca
,
A.
Trovarelli
,
Y.
Wang
, and
C.
Wöll
, “
Surface faceting and reconstruction of ceria nanoparticles
,”
Angew. Chem., Int. Ed.
56
(
1
),
375
379
(
2017
).
42.
G.
Zhang
,
T.
Dai
,
Y.
Meng
,
L.
Zhang
,
C.
Yang
,
G.
Pan
,
Z.
Ni
, and
S.
Xia
, “
Z–scheme heterojunction ZnO-Au-ZnAl2O4: Bridge-type hot carrier transfer and reaction kinetics in the photodegradation of catechol
,”
Appl. Surf. Sci.
532
,
147456
(
2020
).
43.
F.
Ye
,
Y.
Su
,
X.
Quan
,
S.
Chen
,
H.
Yu
, and
H.
Li
, “
Constructing desired interfacial energy band alignment of Z-scheme TiO2-Pd-Cu2O hybrid by controlling the contact facet for improved photocatalytic performance
,”
Appl. Catal. B
244
,
347
355
(
2019
).
44.
L.
Jing
,
Y.
Xu
,
S.
Huang
,
M.
Xie
,
M.
He
,
H.
Xu
,
H.
Li
, and
Q.
Zhang
, “
Novel magnetic CoFe2O4/Ag/Ag3 VO4 composites: Highly efficient visible light photocatalytic and antibacterial activity
,”
Appl. Catal. B
199
,
11
22
(
2016
).
45.
R.
Shi
,
Y.
Cao
,
Y.
Bao
,
Y.
Zhao
,
G. I. N.
Waterhouse
,
Z.
Fang
,
L.
Wu
,
C.
Tung
,
Y.
Yin
, and
T.
Zhang
, “
Self‐assembled Au/CdSe nanocrystal clusters for plasmon‐mediated photocatalytic hydrogen evolution
,”
Adv. Mater.
29
(
27
),
1700803
(
2017
).
46.
H.
Tada
,
T.
Mitsui
,
T.
Kiyonaga
,
T.
Akita
, and
K.
Tanaka
, “
All-solid-state Z-scheme in CdS–Au–TiO2 three-component nanojunction system
,”
Nat. Mater.
5
(
10
),
782
786
(
2006
).
47.
R.
Ouyang
,
P.
Cao
,
P.
Jia
,
H.
Wang
,
T.
Zong
,
C.
Dai
,
J.
Yuan
,
Y.
Li
,
D.
Sun
,
N.
Guo
,
Y.
Miao
, and
S.
Zhou
, “
Bistratal Au@Bi2S3 nanobones for excellent NIR-triggered/multimodal imaging-guided synergistic therapy for liver cancer
,”
Bioact. Mater.
6
(
2
),
386
403
(
2021
).
48.
H.
Tong
,
S.
Ouyang
,
Y.
Bi
,
N.
Umezawa
,
M.
Oshikiri
, and
J.
Ye
, “
Nano‐photocatalytic materials: Possibilities and challenges
,”
Adv. Mater.
24
(
2
),
229
251
(
2012
).
49.
C.
Yang
,
J.
Zhang
,
M.
Chang
,
J.
Tan
,
M.
Yuan
,
Y.
Bian
,
B.
Liu
,
Z.
Liu
,
M.
Wang
,
B.
Ding
,
P.
Ma
, and
J.
Lin
, “
NIR-Activatable heterostructured nanoadjuvant CoP/NiCoP executing lactate metabolism interventions for boosted photocatalytic hydrogen therapy and photoimmunotherapy
,”
Adv. Mater.
36
(
5
),
e2308774
(
2024
).
50.
S.
Chen
,
Y.
Zhu
,
Q.
Xu
,
Q.
Jiang
,
D.
Chen
,
T.
Chen
,
X.
Xu
,
Z.
Jin
, and
Q.
He
, “
Photocatalytic glucose depletion and hydrogen generation for diabetic wound healing
,”
Nat. Commun.
13
,
5684
(
2022
).
51.
Q.
Fu
,
L.
Yu
,
M.
Zhang
,
S.
Li
, and
L.
Liu
, “
Engineering nanosystems for ROS-bridged cancer cascade therapy
,”
Chem. Eng. J.
473
,
145415
(
2023
).
52.
B.
Zhao
,
Y.
Wang
,
X.
Yao
,
D.
Chen
,
M.
Fan
,
Z.
Jin
, and
Q.
He
, “
Photocatalysis-mediated drug-free sustainable cancer therapy using nanocatalyst
,”
Nat. Commun.
12
(
1
),
1345
(
2021
).
53.
B. M.
Rajbongshi
, “
Photocatalyst: Mechanism, challenges, and strategy for organic contaminant degradation
,” in
Handbook of Smart Photocatalytic Materials
(
Elsevier
,
2020
), pp.
127
149
.
54.
X.
Wang
,
J.
Xu
,
D.
Yang
,
C.
Sun
,
Q.
Sun
,
F.
He
,
S.
Gai
,
C.
Zhong
,
C.
Li
, and
P.
Yang
, “
Fe3O4@MIL-100(Fe)-UCNPs heterojunction photosensitizer: Rational design and application in near infrared light mediated hypoxic tumor therapy
,”
J. Chem. Eng.
354
,
1141
1152
(
2018
).
55.
P.
Liu
,
L.
Hao
,
M.
Liu
, and
S.
Hu
, “
Glutathione-responsive and -exhausting metal nanomedicines for robust synergistic cancer therapy
,”
Front. Bioeng. Biotechnol.
11
,
1161472
(
2023
).
56.
C.
Yang
,
J.
Zhang
,
M.
Chang
,
J.
Tan
,
M.
Yuan
,
Y.
Bian
,
B.
Liu
,
Z.
Liu
,
M.
Wang
,
B.
Ding
,
P.
Ma
, and
J.
Lin
, “
NIR‐activatable heterostructured nanoadjuvant CoP/NiCoP executing lactate metabolism interventions for boosted photocatalytic hydrogen therapy and photoimmunotherapy
,”
Adv. Mater.
36
(
5
),
2308774
(
2024
).
57.
Y.
Cheng
,
X.
Kong
,
Y.
Chang
,
Y.
Feng
,
R.
Zheng
,
X.
Wu
,
K.
Xu
,
X.
Gao
, and
H.
Zhang
, “
Spatiotemporally synchronous oxygen self‐supply and reactive oxygen species production on Z‐scheme heterostructures for hypoxic tumor therapy
,”
Adv. Mater.
32
(
11
),
1908109
(
2020
).
58.
N.
Meng
,
P.
Xu
,
C.
Wen
,
H.
Liu
,
C.
Gao
,
X.-C.
Shen
, and
H.
Liang
, “
Near-infrared-II-activatable sulfur-deficient plasmonic Bi2S3-x-Au heterostructures for photoacoustic imaging-guided ultrasound enhanced high performance phototherapy
,”
J. Colloid Interface Sci.
644
,
437
453
(
2023
).
59.
Y.
Song
,
L.
Yang
,
M.
Xu
,
Q.
Lu
,
W.
Li
,
C.
Ren
,
P.
Liu
,
Y.
Wang
,
Y.
Zhu
,
F.
Tan
, and
N.
Li
, “
Boosted photocatalytic activity induced NAMPT-Regulating therapy based on elemental bismuth-humic acids heterojunction for inhibiting tumor proliferation/migration/inflammation
,”
Biomaterials
254
,
120140
(
2020
).
60.
Y.
Liu
,
Y.
Yi
,
S.
Sun
,
T.
Wang
,
J.
Tang
,
Z.
Peng
,
W.
Huang
,
W.
Zeng
, and
M.
Wu
, “
Biodegradable and efficient charge‐migrated Z‐scheme heterojunction amplifies cancer ferroptosis by blocking defensive redox system
,”
Small
20
,
2309206
(
2024
).
61.
D.-W.
Zheng
,
B.
Li
,
C.-X.
Li
,
J.-X.
Fan
,
Q.
Lei
,
C.
Li
,
Z.
Xu
, and
X.-Z.
Zhang
, “
Carbon dot decorated carbon nitride nanoparticles for enhanced photodynamic therapy against hypoxic tumor via water splitting
,”
ACS Nano
10
(
9
),
8715
8722
(
2016
).
62.
A.
Douplik
,
G.
Saiko
,
I.
Schelkanova
, and
V. V.
Tuchin
, “
The response of tissue to laser light
,” in
Lasers for Medical Applications
(
Elsevier
,
2013
), pp.
47
109
.
63.
Z.
Xue
,
X.
Li
,
Y.
Li
,
M.
Jiang
,
H.
Liu
,
S.
Zeng
, and
J.
Hao
, “
X-ray-activated near-infrared persistent luminescent probe for deep-tissue and renewable in vivo bioimaging
,”
ACS Appl. Mater. Interfaces
9
(
27
),
22132
22142
(
2017
).
64.
N.
Kong
,
H.
Zhang
,
C.
Feng
,
C.
Liu
,
Y.
Xiao
,
X.
Zhang
,
L.
Mei
,
J. S.
Kim
,
W.
Tao
, and
X.
Ji
, “
Arsenene-mediated multiple independently targeted reactive oxygen species burst for cancer therapy
,”
Nat. Commun.
12
(
1
),
4777
(
2021
).
65.
M.
Ou
,
C.
Lin
,
Y.
Wang
,
Y.
Lu
,
W.
Wang
,
Z.
Li
,
W.
Zeng
,
X.
Zeng
,
X.
Ji
, and
L.
Mei
, “
Heterojunction engineered bioactive chlorella for cascade promoted cancer therapy
,”
J. Controlled Release
345
,
755
769
(
2022
).
66.
X.
Li
,
H.
Sun
,
Y.
Xie
,
Y.
Liang
,
X.
Gong
,
P.
Qin
,
L.
Jiang
,
J.
Guo
,
C.
Liu
, and
Z.
Wu
, “
Principles, synthesis and applications of dual Z-scheme photocatalysts
,”
Coord. Chem. Rev.
467
,
214596
(
2022
).
67.
C.
Pan
,
M.
Ou
,
Q.
Cheng
,
Y.
Zhou
,
Y.
Yu
,
Z.
Li
,
F.
Zhang
,
D.
Xia
,
L.
Mei
, and
X.
Ji
, “
Z‐scheme heterojunction functionalized pyrite nanosheets for modulating tumor microenvironment and strengthening photo/chemodynamic therapeutic effects
,”
Adv. Funct. Mater.
30
(
3
),
1906466
(
2020
).
68.
C.
Lee
,
M.
Jeon
, and
C.
Kim
, “
Photoacoustic imaging in nanomedicine
,” in
Applications of Nanoscience in Photomedicine
(
Elsevier
,
2015
), pp.
31
47
.
69.
X.
Cao
,
Y.
Wang
,
X.
Song
,
W.
Lou
,
X.
Li
,
W.
Lu
,
K.
Chen
,
L.
Chen
,
Y.
Chen
, and
B.
Huang
, “
Defect‐engineering bismuth‐based homologous schottky heterojunction for metabolic regulation‐augmented sonodynamic tumor therapy
,”
Adv. Funct. Mater.
33
(
25
),
2300777
(
2023
).
70.
S.
Bai
,
N.
Zhang
,
C.
Gao
, and
Y.
Xiong
, “
Defect engineering in photocatalytic materials
,”
Nano Energy
53
,
296
336
(
2018
).
71.
J.
Tian
,
J.
Li
,
Y.
Guo
,
Z.
Liu
,
B.
Liu
, and
J.
Li
, “
Oxygen vacancy mediated bismuth-based photocatalysts
,”
Adv. Powder Mater.
3
(
4
),
100201
(
2024
).
72.
J.
Ping
,
J.
Du
,
R.
Ouyang
,
Y.
Miao
, and
Y.
Li
, “
Recent advances in stimuli-responsive nano-heterojunctions for tumor therapy
,”
Colloids Surf. B
226
,
113303
(
2023
).
73.
X.
Wang
,
P.
Hua
,
C.
He
, and
M.
Chen
, “
Non-apoptotic cell death-based cancer therapy: Molecular mechanism, pharmacological modulators, and nanomedicine
,”
Acta Pharm. Sin. B
12
(
9
),
3567
3593
(
2022
).
74.
J.
Liu
,
S.
Dong
,
S.
Gai
,
S.
Li
,
Y.
Dong
,
C.
Yu
,
F.
He
, and
P.
Yang
, “
Four birds with one Stone: A bandgap regulated multifunctional schottky heterojunction for robust synergistic antitumor therapy upon endo-/exogenous stimuli
,”
ACS Nano
18
,
23579
(
2024
).
75.
S.
Cheng
,
T.
Zhou
,
Y.
Luo
,
J.
Zhang
,
K.
Dong
,
Q.
Zhang
,
W.
Shu
,
T.
Zhang
,
Q.
Zhang
,
R.
Shi
,
Y.
Yao
, and
H.
Wang
, “
Ultrasound-responsive Bi2MoO6-MXene heterojunction as ferroptosis inducers for stimulating immunogenic cell death against ovarian cancer
,”
J. Nanobiotechnol.
22
(
1
),
408
(
2024
).
76.
Z.
Yang
,
C.
Yang
,
D.
Yang
,
Y.
Zhang
,
Q.
Yang
,
F.
Qu
, and
W.
Guo
, “
l-Arginine-modified CoWO4/FeWO4S-scheme heterojunction enhances ferroptosis against solid tumor
,”
Adv. Healthcare Mater.
12
(
18
),
2203092
(
2023
).
77.
Z.
Lv
,
Y.
Wu
,
J.
Lin
,
W.
Li
,
H.
Weisbecker
,
P.
Wang
,
X.
Song
,
W.
Sun
,
Z.
Sun
,
Y.
Xie
,
J.
Meng
,
J.
Dong
,
X.
An
,
J.
Chen
,
S.
Yang
,
T.
Yuan
,
H.
Jiang
,
C.
Sun
,
X.
Yang
,
H.
Qian
,
H.
Cai
,
J.
Zhao
,
W.
Bai
,
D.
Shi
, and
N.
Bao
, “
Schottky heterojunction facilitates osteosarcoma ferroptosis and enhances bone formation in a switchable mode
,”
Adv. Funct. Mater.
34
(
10
),
2312032
(
2024
).
78.
J.
Li
,
Z.
Yue
,
M.
Tang
,
W.
Wang
,
Y.
Sun
,
T.
Sun
, and
C.
Chen
, “
Strategies to reverse hypoxic tumor microenvironment for enhanced sonodynamic therapy
,”
Adv. Healthcare Mater.
13
(
1
),
2302028
(
2024
).
79.
R.
Zhou
,
M.
Chang
,
M.
Shen
,
Y.
Cong
,
Y.
Chen
, and
Y.
Wang
, “
Sonocatalytic optimization of titanium‐based therapeutic nanomedicine
,”
Adv. Sci.
10
(
25
),
2301764
(
2023
).
80.
S.
Bai
,
N.
Yang
,
X.
Wang
,
F.
Gong
,
Z.
Dong
,
Y.
Gong
,
Z.
Liu
, and
L.
Cheng
, “
Ultrasmall iron-doped titanium oxide nanodots for enhanced sonodynamic and chemodynamic cancer therapy
,”
ACS Nano
14
(
11
),
15119
15130
(
2020
).
81.
Y.
Zhou
,
L.
Yu
,
C.
Dong
,
J.
Liu
,
B.
Yang
,
Y.
Chen
, and
Z.
Hu
, “
Two-dimensional semiconductor heterojunction nanostructure for mutually synergistic sonodynamic and chemoreactive cancer nanotherapy
,”
Chem. Eng. J.
431
,
134017
(
2022
).
82.
A. K.
Ganguli
,
G. B.
Kunde
,
W.
Raza
,
S.
Kumar
, and
P.
Yadav
, “
Assessment of performance of photocatalytic nanostructured materials with varied morphology based on reaction conditions
,”
Molecules
27
(
22
),
7778
(
2022
).
83.
M.
He
,
H.
Yu
,
Y.
Zhao
,
J.
Liu
,
Q.
Dong
,
Z.
Xu
,
Y.
Kang
, and
P.
Xue
, “
Ultrasound-activatable g-C3N4-anchored titania heterojunction as an intracellular redox homeostasis perturbator for augmented oncotherapy
,”
Small
19
,
2300244
(
2023
).
84.
C.
Feng
,
L.
Wang
,
D.
Zhang
,
L.
Geng
,
L.
Zhou
,
L.
Wang
,
G.
Tian
,
Q.
Tang
,
J.
Hu
,
B.
Geng
, and
L.
Yan
, “
Tumour microenvironment-responded Fe-doped carbon dots-sensitized cubic Cu2O for Z-scheme heterojunction-enhanced sono-chemodynamic synergistic tumor therapy
,”
J. Colloid Interface Sci.
665
,
681
692
(
2024
).
85.
T.
Zhang
,
Q.
Zheng
,
Y.
Fu
,
C.
Xie
,
G.
Fan
,
Y.
Wang
,
Y.
Wu
,
X.
Cai
,
G.
Han
, and
X.
Li
, “
α-Fe2O3@Pt heterostructure particles to enable sonodynamic therapy with self-supplied O2 and imaging-guidance
,”
J. Nanobiotechnol.
19
(
1
),
358
(
2021
).
86.
Q.
Wang
,
Q.
Gao
,
A. M.
Al-Enizi
,
A.
Nafady
, and
S.
Ma
, “
Recent advances in MOF-based photocatalysis: Environmental remediation under visible light
,”
Inorg. Chem. Front.
7
(
2
),
300
339
(
2020
).
87.
X.
Meng
,
S.
Sun
,
C.
Gong
,
J.
Yang
,
Z.
Yang
,
X.
Zhang
, and
H.
Dong
, “
Ag-doped metal–organic frameworks' heterostructure for sonodynamic therapy of deep-seated cancer and bacterial infection
,”
ACS Nano
17
(
2
),
1174
1186
(
2023
).
88.
Y.
Li
,
W.
Li
,
Y.
Liu
,
J.
Liu
,
X.
Yuan
,
J.
Zhang
, and
H.
Shen
, “
Defect-rich platinum–zinc oxide heterojunction as a potent ROS amplifier for synergistic sono-catalytic therapy
,”
Acta Biomater.
171
,
543
552
(
2023
).
89.
B.
Geng
,
L.
Yan
,
Y.
Zhu
,
W.
Shi
,
H.
Wang
,
J.
Mao
,
L.
Ren
,
J.
Zhang
,
Y.
Tian
,
F.
Gao
,
X.
Zhang
,
J.
Chen
, and
J.
Zhu
, “
Carbon dot@MXene nanozymes with triple enzyme-mimic activities for mild NIR-II photothermal-amplified nanocatalytic therapy
,”
Adv. Healthcare Mater.
12
(
5
),
2202154
(
2023
).
90.
C.
Cao
,
H.
Zou
,
N.
Yang
,
H.
Li
,
Y.
Cai
,
X.
Song
,
J.
Shao
,
P.
Chen
,
X.
Mou
,
W.
Wang
, and
X.
Dong
, “
Fe3O4/Ag/Bi2MoO6 Photoactivatable nanozyme for self-replenishing and sustainable cascaded nanocatalytic cancer therapy
,”
Adv. Mater.
33
,
2106996
(
2021
).
91.
S.
Dong
,
Y.
Dong
,
T.
Jia
,
S.
Liu
,
J.
Liu
,
D.
Yang
,
F.
He
,
S.
Gai
,
P.
Yang
, and
J.
Lin
, “
GSH-depleted nanozymes with hyperthermia-enhanced dual enzyme-mimic activities for tumor nanocatalytic therapy
,”
Adv. Mater.
32
(
42
),
e2002439
(
2020
).
92.
Y.
Zhao
,
B.
Yuan
,
L.
Yan
,
Z.
Wang
,
Z.
Xu
,
B.
Geng
,
X.
Guo
, and
X.
Chen
, “
In situ synthesis of Ru/TiO2−x@TiCN ternary heterojunctions for enhanced sonodynamic and nanocatalytic cancer therapy
,”
Adv. Sci.
11
(
4
),
2307029
(
2024
).
93.
N.
Tao
,
H.
Li
,
L.
Deng
,
S.
Zhao
,
J.
Ouyang
,
M.
Wen
,
W.
Chen
,
K.
Zeng
,
C.
Wei
, and
Y.-N.
Liu
, “
A cascade nanozyme with amplified sonodynamic therapeutic effects through comodulation of hypoxia and Immunosuppression against cancer
,”
ACS Nano
16
(
1
),
485
501
(
2022
).
94.
J.
Hu
,
B.
Geng
,
J.
Glowacki
,
S.
Zhang
,
X.
Yang
,
D.
Pan
, and
L.
Shen
, “
A biodegradable p-n junction sonosensitizer for tumor microenvironment regulating sonodynamic tumor therapy
,”
Chem. Eng. J.
446
,
137320
(
2022
).
95.
K.
Song
,
J.
Du
,
X.
Wang
,
L.
Zheng
,
R.
Ouyang
,
Y.
Li
,
Y.
Miao
, and
D.
Zhang
, “
Biodegradable bismuth‐based nano‐heterojunction for enhanced sonodynamic oncotherapy through charge separation engineering
,”
Adv. Healthcare Mater.
11
(
11
),
2102503
(
2022
).
96.
S.
Liang
,
X.
Deng
,
G.
Xu
,
X.
Xiao
,
M.
Wang
,
X.
Guo
,
P.
Ma
,
Z.
Cheng
,
D.
Zhang
, and
J.
Lin
, “
A novel Pt–TiO2 heterostructure with oxygen‐deficient layer as bilaterally enhanced sonosensitizer for synergistic chemo‐sonodynamic cancer therapy
,”
Adv. Funct. Mater.
30
(
13
),
1908598
(
2020
).
97.
G.
Chen
,
J.
Du
,
L.
Gu
,
Q.
Wang
,
Q.
Qi
,
X.
Li
,
R.
Zhang
,
H.
Yang
,
Y.
Miao
, and
Y.
Li
, “
Metal-sensitized Au-Bi2O3 nanoheterojunction for immunogenic cell death-boosted sono-immuno cancer therapy
,”
Chem. Eng. J.
482
,
148953
(
2024
).
98.
Y.
Wang
,
H.
Zhang
,
Y.
Liu
,
M. H.
Younis
,
W.
Cai
, and
W.
Bu
, “
Catalytic radiosensitization: Insights from materials physicochemistry
,”
Mater. Today
57
,
262
278
(
2022
).
99.
J. D.
Carter
,
N. N.
Cheng
,
Y.
Qu
,
G. D.
Suarez
, and
T.
Guo
, “
Nanoscale energy deposition by X-ray absorbing nanostructures
,”
J. Phys. Chem. B
111
(
40
),
11622
11625
(
2007
).
100.
E.
Chelnokov
,
V.
Cuba
,
D.
Simeone
,
J.-M.
Guigner
,
U.
Schmidhammer
,
M.
Mostafavi
, and
S.
Le Caër
, “
Electron transfer at oxide/water interfaces induced by ionizing radiation
,”
J. Phys. Chem. C
118
(
15
),
7865
7873
(
2014
).
101.
W.
Liu
,
B.
Chen
,
H.
Zheng
,
Y.
Xing
,
G.
Chen
,
P.
Zhou
,
L.
Qian
, and
Y.
Min
, “
Advances of nanomedicine in radiotherapy
,”
Pharmaceutics
13
(
11
),
1757
(
2021
).
102.
C.
Lee
,
N. N.
Cheng
,
R. A.
Davidson
, and
T.
Guo
, “
Geometry enhancement of nanoscale energy deposition by X-rays
,”
J. Phys. Chem. C
116
(
20
),
11292
11297
(
2012
).
103.
C. C. L.
McCrory
,
S.
Jung
,
J. C.
Peters
, and
T. F.
Jaramillo
, “
Benchmarking heterogeneous electrocatalysts for the oxygen evolution reaction
,”
J. Am. Chem. Soc.
135
(
45
),
16977
16987
(
2013
).
104.
W.
Fan
,
W.
Bu
,
Z.
Zhang
,
B.
Shen
,
H.
Zhang
,
Q.
He
,
D.
Ni
,
Z.
Cui
,
K.
Zhao
,
J.
Bu
,
J.
Du
,
J.
Liu
, and
J.
Shi
, “
X‐ray radiation‐controlled NO‐release for on‐demand depth‐independent hypoxic radiosensitization
,”
Angew. Chem., Int. Ed.
54
(
47
),
14026
14030
(
2015
).
105.
M.
Zhang
,
Z.
Cui
,
R.
Song
,
B.
Lv
,
Z.
Tang
,
X.
Meng
,
X.
Chen
,
X.
Zheng
,
J.
Zhang
,
Z.
Yao
, and
W.
Bu
, “
SnWO4-based nanohybrids with full energy transfer for largely enhanced photodynamic therapy and radiotherapy
,”
Biomaterials
155
,
135
144
(
2018
).
106.
Y.
Yong
,
X.
Cheng
,
T.
Bao
,
M.
Zu
,
L.
Yan
,
W.
Yin
,
C.
Ge
,
D.
Wang
,
Z.
Gu
, and
Y.
Zhao
, “
Tungsten sulfide quantum dots as multifunctional nanotheranostics for in vivo dual-modal image-guided photothermal/radiotherapy synergistic therapy
,”
ACS Nano
9
(
12
),
12451
12463
(
2015
).
107.
Z.
Wang
,
X.
Ren
,
D.
Wang
,
L.
Guan
,
X.
Li
,
Y.
Zhao
,
A.
Liu
,
L.
He
,
T.
Wang
,
A. V.
Zvyagin
,
B.
Yang
, and
Q.
Lin
, “
Novel strategies for tumor radiosensitization mediated by multifunctional gold-based nanomaterials
,”
Biomater. Sci.
11
(
4
),
1116
1136
(
2023
).
108.
Q.
Liu
,
Y.
Shi
,
Y.
Chong
, and
C.
Ge
, “
Pharmacological ascorbate promotes the tumor radiosensitization of Au@Pd nanoparticles with simultaneous protection of normal tissues
,”
ACS Appl. Bio Mater.
4
(
2
),
1843
1851
(
2021
).
109.
B.
Yang
,
Y.
Chen
, and
J.
Shi
, “
Reactive oxygen species (ROS)-based nanomedicine
,”
Chem. Rev.
119
(
8
),
4881
4985
(
2019
).
110.
R.
Cheng
,
H.
Liu
,
X.
Dong
,
S.
Zhu
,
R.
Zhou
,
C.
Wang
,
Y.
Wang
,
X.
Wang
,
C.
Su
, and
Z.
Gu
, “
Semiconductor heterojunction-based radiocatalytic platforms for tumors treatment by enhancing radiation response and reducing radioresistance
,”
Chem. Eng. J.
394
,
124872
(
2020
).
111.
Y.
Zang
,
L.
Gong
,
L.
Mei
,
Z.
Gu
, and
Q.
Wang
, “
Bi2WO6 Semiconductor nanoplates for tumor radiosensitization through high-Z Effects and radiocatalysis
,”
ACS Appl. Mater. Interfaces
11
(
21
),
18942
18952
(
2019
).
112.
S. P.
Sahu
and
E. L.
Cates
, “
X-ray radiocatalytic activity and mechanisms of bismuth complex oxides
,”
J. Phys. Chem. C
121
(
19
),
10538
10545
(
2017
).
113.
X.
Wang
,
C.
Zhang
,
J.
Du
,
X.
Dong
,
S.
Jian
,
L.
Yan
,
Z.
Gu
, and
Y.
Zhao
, “
Enhanced generation of non-oxygen dependent free radicals by schottky-type heterostructures of Au–Bi2S3 nanoparticles via X-ray-induced catalytic reaction for radiosensitization
,”
ACS Nano
13
(
5
),
5947
5958
(
2019
).
114.
Z.
Guo
,
S.
Zhu
,
Y.
Yong
,
X.
Zhang
,
X.
Dong
,
J.
Du
,
J.
Xie
,
Q.
Wang
,
Z.
Gu
, and
Y.
Zhao
, “
Synthesis of BSA‐Coated BiOI@Bi2S3 semiconductor heterojunction nanoparticles and their applications for radio/photodynamic/photothermal synergistic therapy of tumor
,”
Adv. Mater.
29
(
44
),
1704136
(
2017
).
115.
R.
Zhou
,
H.
Wang
,
Y.
Yang
,
C.
Zhang
,
X.
Dong
,
J.
Du
,
L.
Yan
,
G.
Zhang
,
Z.
Gu
, and
Y.
Zhao
, “
Tumor microenvironment-manipulated radiocatalytic sensitizer based on bismuth heteropolytungstate for radiotherapy enhancement
,”
Biomaterials
189
,
11
22
(
2019
).
116.
D.
Wang
,
Y.
Liao
,
H.
Yan
,
S.
Zhu
,
Y.
Liu
,
J.
Li
,
X.
Wang
,
X.
Guo
,
Z.
Gu
, and
B.
Sun
, “
In situ formed Z-scheme graphdiyne heterojunction realizes NIR-photocatalytic oxygen evolution and selective radiosensitization for hypoxic tumors
,”
ACS Nano
16
(
12
),
21186
21198
(
2022
).
117.
J.
Kosco
,
M.
Bidwell
,
H.
Cha
,
T.
Martin
,
C. T.
Howells
,
M.
Sachs
,
D. H.
Anjum
,
S.
Gonzalez Lopez
,
L.
Zou
,
A.
Wadsworth
,
W.
Zhang
,
L.
Zhang
,
J.
Tellam
,
R.
Sougrat
,
F.
Laquai
,
D. M.
DeLongchamp
,
J. R.
Durrant
, and
I.
McCulloch
, “
Enhanced photocatalytic hydrogen evolution from organic semiconductor heterojunction nanoparticles
,”
Nat. Mater.
19
(
5
),
559
565
(
2020
).
118.
L.
Zhang
,
Y.
Wu
,
J.
Li
,
Z.
Jin
,
Y.
Li
, and
N.
Tsubaki
, “
Amorphous/crystalline heterojunction interface driving the spatial separation of charge carriers for efficient photocatalytic hydrogen evolution
,”
Mater. Today Phys.
27
,
100767
(
2022
).
119.
S.
Guo
,
S.
Han
,
H.
Mao
,
S.
Dong
,
C.
Wu
,
L.
Jia
,
B.
Chi
,
J.
Pu
, and
J.
Li
, “
Structurally controlled ZnO/TiO2 heterostructures as efficient photocatalysts for hydrogen generation from water without noble metals: The role of microporous amorphous/crystalline composite structure
,”
J. Power Sources
245
,
979
985
(
2014
).
120.
Z.
Wang
,
L.
Wang
,
S.
Liu
,
M.
Zhang
,
Y.
Li
,
L.
Rong
,
Y.
Liu
, and
H.
Zhang
, “
Z-Scheme heterostructures for glucose oxidase-sensitized radiocatalysis and starvation therapy of tumors
,”
Nanoscale
14
(
6
),
2186
2198
(
2022
).
121.
Z.
Wang
,
S.
Liu
,
L.
Wang
,
H.
Zou
,
Z.
Wang
,
X.
Tang
,
W.
Feng
,
Y.
Chong
,
Y.
Liu
,
B.
Yang
, and
H.
Zhang
, “
BiVO4@Bi2S3 heterojunction nanorods with enhanced charge separation efficiency for multimodal imaging and synergy therapy of tumor
,”
ACS Appl. Bio Mater.
3
(
8
),
5080
5092
(
2020
).
122.
R.
Zhou
,
X.
Liu
,
Y.
Wu
,
H.
Xiang
,
J.
Cao
,
Y.
Li
,
W.
Yin
,
Y.
Zu
,
J.
Li
,
R.
Liu
,
F.
Zhao
,
Z.
Liu
,
C.
Chen
,
Z.
Gu
,
L.
Yan
, and
Y.
Zhao
, “
Suppressing the radiation-induced corrosion of bismuth nanoparticles for enhanced synergistic cancer radiophototherapy
,”
ACS Nano
14
(
10
),
13016
13029
(
2020
).
123.
R.
Cai
,
H.
Xiang
,
D.
Yang
,
K.-T.
Lin
,
Y.
Wu
,
R.
Zhou
,
Z.
Gu
,
L.
Yan
,
Y.
Zhao
, and
W.
Tan
, “
Plasmonic AuPt@CuS heterostructure with enhanced synergistic efficacy for radiophotothermal therapy
,”
J. Am. Chem. Soc.
143
(
39
),
16113
16127
(
2021
).
124.
J.
Liu
,
Z.
Zhang
,
Z.
Fan
,
X.
Tang
, and
Q.
Zhong
, “
Strategies for boosting the photocatalytic reduction of toxic metal ions: Progress and prospects
,”
J. Water Process Eng.
57
,
104683
(
2024
).
125.
C.
Wang
,
H.
Wang
,
L.
Luo
,
S.
Gan
,
Y.
Yao
,
Q.
Wei
,
J.
Wu
,
A.
Yuan
,
Y.
Hu
,
C.
Wu
, and
Z.
Zou
, “
Scintillator-based radiocatalytic superoxide radical production for long-term tumor DNA damage
,”
Biomater. Sci.
10
(
13
),
3433
3440
(
2022
).
126.
L.
Yin
,
F.
Yang
,
X.
Bao
,
W.
Xue
,
Z.
Du
,
X.
Wang
,
J.
Cheng
,
H.
Ji
,
J.
Sui
,
X.
Liu
,
Y.
Wang
,
F.
Cao
,
J.
Mao
,
M.
Li
,
Z.
Ren
, and
Q.
Zhang
, “
Low-temperature sintering of Ag nanoparticles for high-performance thermoelectric module design
,”
Nat. Energy
8
(
7
),
665
674
(
2023
).
127.
Y.
Wang
,
X.
Dai
,
C.
Dong
,
W.
Guo
,
Z.
Xu
,
Y.
Chen
,
H.
Xiang
, and
R.
Zhang
, “
Engineering electronic band structure of binary thermoelectric nanocatalysts for augmented pyrocatalytic tumor nanotherapy
,”
Adv. Mater.
34
(
4
),
2106773
(
2022
).
128.
X.
Ji
,
Z.
Tang
,
H.
Liu
,
Y.
Kang
,
L.
Chen
,
J.
Dong
,
W.
Chen
,
N.
Kong
,
W.
Tao
, and
T.
Xie
, “
Nanoheterojunction‐mediated thermoelectric strategy for cancer surgical adjuvant treatment and β‐elemene combination therapy
,”
Adv. Mater.
35
(
8
),
2207391
(
2023
).
129.
J.
Mao
,
Z.
Liu
,
J.
Zhou
,
H.
Zhu
,
Q.
Zhang
,
G.
Chen
, and
Z.
Ren
, “
Advances in thermoelectrics
,”
Adv. Phys
67
(
2
),
69
147
(
2018
).
130.
S. I.
Kim
,
K. H.
Lee
,
H. A.
Mun
,
H. S.
Kim
,
S. W.
Hwang
,
J. W.
Roh
,
D. J.
Yang
,
W. H.
Shin
,
X. S.
Li
,
Y. H.
Lee
,
G. J.
Snyder
, and
S. W.
Kim
, “
Dense dislocation arrays embedded in grain boundaries for high-performance bulk thermoelectrics
,”
Science
348
(
6230
),
109
114
(
2015
).
131.
A. T.
Duong
,
V. Q.
Nguyen
,
G.
Duvjir
,
V. T.
Duong
,
S.
Kwon
,
J. Y.
Song
,
J. K.
Lee
,
J. E.
Lee
,
S.
Park
,
T.
Min
,
J.
Lee
,
J.
Kim
, and
S.
Cho
, “
Achieving ZT=2.2 with Bi-doped n-type SnSe single crystals
,”
Nat. Commun.
7
(
1
),
13713
(
2016
).
132.
Y.
Pei
,
X.
Shi
,
A.
LaLonde
,
H.
Wang
,
L.
Chen
, and
G. J.
Snyder
, “
Convergence of electronic bands for high performance bulk thermoelectrics
,”
Nature
473
(
7345
),
66
69
(
2011
).
133.
Y.
Dong
,
S.
Dong
,
C.
Yu
,
J.
Liu
,
S.
Gai
,
Y.
Xie
,
Z.
Zhao
,
X.
Qin
,
L.
Feng
,
P.
Yang
, and
Y.
Zhao
, “
Mitochondria-targeting Cu3VS4 nanostructure with high copper ionic mobility for photothermoelectric therapy
,”
Sci. Adv.
9
(
44
),
eadi9980
(
2023
).
134.
X.
Jiang
,
M.
Yang
,
Y.
Fang
,
Z.
Yang
,
X.
Dai
,
P.
Gu
,
W.
Feng
, and
Y.
Chen
, “
A photo‐activated thermoelectric catalyst for ferroptosis‐/pyroptosis‐boosted tumor nanotherapy
,”
Adv. Healthcare Mater.
12
(
24
),
2300699
(
2023
).
135.
Y.
Kang
,
N.
Kong
,
M.
Ou
,
Y.
Wang
,
Q.
Xiao
,
L.
Mei
,
B.
Liu
,
L.
Chen
,
X.
Zeng
, and
X.
Ji
, “
A novel cascaded energy conversion system inducing efficient and precise cancer therapy
,”
Bioact. Mater.
20
,
663
676
(
2023
).
136.
X.
Yuan
,
Y.
Kang
,
J.
Dong
,
R.
Li
,
J.
Ye
,
Y.
Fan
,
J.
Han
,
J.
Yu
,
G.
Ni
,
X.
Ji
, and
D.
Ming
, “
Self-triggered thermoelectric nanoheterojunction for cancer catalytic and immunotherapy
,”
Nat. Commun.
14
(
1
),
5140
(
2023
).
137.
S.
Kapri
and
S.
Bhattacharyya
, “
Molybdenum sulfide–reduced graphene oxide p–n heterojunction nanosheets with anchored oxygen generating manganese dioxide nanoparticles for enhanced photodynamic therapy
,”
Chem. Sci.
9
(
48
),
8982
8989
(
2018
).
138.
B.
Geng
,
H.
Qin
,
F.
Zheng
,
W.
Shen
,
P.
Li
,
K.
Wu
,
X.
Wang
,
X.
Li
,
D.
Pan
, and
L.
Shen
, “
Carbon dot-sensitized MoS2 nanosheet heterojunctions as highly efficient NIR photothermal agents for complete tumor ablation at an ultralow laser exposure
,”
Nanoscale
11
(
15
),
7209
7220
(
2019
).
139.
M.
Chang
,
M.
Wang
,
Y.
Chen
,
M.
Shu
,
Y.
Zhao
,
B.
Ding
,
Z.
Hou
, and
J.
Lin
, “
Self-assembled CeVO4/Ag nanohybrid as photoconversion agents with enhanced solar-driven photocatalysis and NIR-responsive photothermal/photodynamic synergistic therapy performance
,”
Nanoscale
11
(
20
),
10129
10136
(
2019
).
140.
M.
Chang
,
M.
Wang
,
M.
Shu
,
Y.
Zhao
,
B.
Ding
,
S.
Huang
,
Z.
Hou
,
G.
Han
, and
J.
Lin
, “
Enhanced photoconversion performance of NdVO4/Au nanocrystals for photothermal/photoacoustic imaging guided and near infrared light-triggered anticancer phototherapy
,”
Acta Biomater.
99
,
295
306
(
2019
).
141.
J.
Dai
,
J.
Song
,
Y.
Qiu
,
J.
Wei
,
Z.
Hong
,
L.
Li
, and
H.
Yang
, “
Gold nanoparticle-decorated g-C3N4 nanosheets for controlled generation of reactive oxygen species upon 670 nm laser illumination
,”
ACS Appl. Mater. Interfaces
11
(
11
),
10589
10596
(
2019
).
142.
Y.
Wang
,
J.
Zhao
,
Z.
Chen
,
F.
Zhang
,
Q.
Wang
,
W.
Guo
,
K.
Wang
,
H.
Lin
, and
F.
Qu
, “
Construct of MoSe2/Bi2Se3 nanoheterostructure: Multimodal CT/PT imaging-guided PTT/PDT/chemotherapy for cancer treating
,”
Biomaterials
217
,
119282
(
2019
).
143.
L.
Dong
,
P.
Zhang
,
X.
Liu
,
R.
Deng
,
K.
Du
,
J.
Feng
, and
H.
Zhang
, “
Renal clearable Bi–Bi2S3 heterostructure nanoparticles for targeting cancer theranostics
,”
ACS Appl. Mater. Interfaces
11
(
8
),
7774
7781
(
2019
).
144.
J.
Wang
,
Y.
Hu
,
J.
Chen
, and
C.
Ye
, “
Self-assembled CeVO4/Au heterojunction nanocrystals for photothermal/photoacoustic bimodal imaging-guided phototherapy
,”
RSC Adv.
10
(
5
),
2581
2588
(
2020
).
145.
M.
Ou
,
C.
Pan
,
Y.
Yu
,
X.
Wang
,
Y.
Zhou
,
H.
Zhang
,
Q.
Cheng
,
M.
Wu
,
X.
Ji
, and
L.
Mei
, “
Two-dimensional highly oxidized ilmenite nanosheets equipped with Z-scheme heterojunction for regulating tumor microenvironment and enhancing reactive oxygen species generation
,”
Chem. Eng. J.
390
,
124524
(
2020
).
146.
Q.
Wang
,
Y.
Ji
,
J.
Shi
, and
L.
Wang
, “
NIR-driven water splitting H2 production nanoplatform for H2-mediated cascade-amplifying synergetic cancer therapy
,”
ACS Appl. Mater. Interfaces
12
(
21
),
23677
23688
(
2020
).
147.
Y.
Li
,
R.
Jia
,
H.
Lin
,
X.
Sun
, and
F.
Qu
, “
Synthesis of MoSe2/CoSe2 nanosheets for NIR‐enhanced chemodynamic therapy via synergistic in‐situ H2O2 production and ctivation
,”
Adv. Funct. Mater.
31
(
8
),
2008420
(
2021
).
148.
Y.
Li
,
Y.
Li
,
Y.
Bai
,
R.
Wang
,
L.
Lin
, and
Y.
Sun
, “
Semiconductor ZnO based photosensitizer core–shell upconversion nanoparticle heterojunction for photodynamic therapy
,”
RSC Adv.
10
(
63
),
38416
38423
(
2020
).
149.
Y.
Wang
,
Z.
Li
,
Y.
Hu
,
J.
Liu
,
M.
Guo
,
H.
Wei
,
S.
Zheng
,
T.
Jiang
,
X.
Sun
,
Z.
Ma
,
Y.
Sun
,
F.
Besenbacher
,
C.
Chen
, and
M.
Yu
, “
Photothermal conversion-coordinated Fenton-like and photocatalytic reactions of Cu2−xSe-Au Janus nanoparticles for tri-combination antitumor therapy
,”
Biomaterials
255
,
120167
(
2020
).
150.
S.
Ma
,
J.
Xie
,
L.
Wang
,
Z.
Zhou
,
X.
Luo
,
J.
Yan
, and
G.
Ran
, “
Hetero-core–shell BiNS–Fe@Fe as a potential theranostic nanoplatform for multimodal imaging-guided simultaneous photothermal–photodynamic and chemodynamic treatment
,”
ACS Appl. Mater. Interfaces
13
(
9
),
10728
10740
(
2021
).
151.
Y.
Kang
,
Z.
Li
,
Y.
Yang
,
Z.
Su
,
X.
Ji
, and
S.
Zhang
, “
Antimonene nanosheets‐based Z‐Scheme heterostructure with enhanced reactive oxygen species generation and photothermal conversion efficiency for photonic therapy of cancer
,”
Adv. Healthcare Mater.
10
(
3
),
2001835
(
2021
).
152.
C.
Yang
,
M.
Chang
,
M.
Yuan
,
F.
Jiang
,
B.
Ding
,
Y.
Zhao
,
P.
Dang
,
Z.
Cheng
,
A. A. A.
Kheraif
,
P.
Ma
, and
J.
Lin
, “
NIR‐triggered multi‐mode antitumor therapy based on Bi2Se3/Au heterostructure with enhanced efficacy
,”
Small
17
(
28
),
2100961
(
2021
).
153.
Y.
Kang
,
Z.
Li
,
F.
Lu
,
Z.
Su
,
X.
Ji
, and
S.
Zhang
, “
Synthesis of red/black phosphorus-based composite nanosheets with a Z-scheme heterostructure for high-performance cancer phototherapy
,”
Nanoscale
14
(
3
),
766
779
(
2022
).
154.
H.
Hou
,
Z.
Wang
,
Y.
Ma
,
K.
Yu
,
J.
Zhao
,
H.
Lin
, and
F.
Qu
, “
NIR-driven intracellular photocatalytic oxygen-supply on metallic molybdenum carbide@N-carbon for hypoxic tumor therapy
,”
J. Colloid Interface Sci.
607
,
1
15
(
2022
).
155.
J.
Bai
,
C.
Peng
,
W.
Lv
,
J.
Liu
,
Y.
Hei
, and
X.
Bo
, “
Vacancy engineering to regulate photocatalytic activity of polymer photosensitizers for amplifying photodynamic therapy against hypoxic tumors
,”
ACS Appl. Mater. Interfaces
13
(
33
),
39055
39065
(
2021
).
156.
M.
Qiu
,
D.
Wang
,
H.
Huang
,
T.
Yin
,
W.
Bao
,
B.
Zhang
,
Z.
Xie
,
N.
Xie
,
Z.
Wu
,
C.
Ge
,
Q.
Wang
,
M.
Gu
,
H. L.
Kutscher
,
L.
Liu
,
S.
Bao
,
P. N.
Prasad
, and
H.
Zhang
, “
A regioselectively oxidized 2D Bi/BiOx lateral nano‐heterostructure for hypoxic photodynamic therapy
,”
Adv. Mater.
33
(
49
),
2102562
(
2021
).
157.
Y.
Wang
,
W.
Wang
,
D.
Sang
,
K.
Yu
,
H.
Lin
, and
F.
Qu
, “
Cu2−xSe/Bi2Se3@PEG Z-scheme heterostructure: A multimode bioimaging guided theranostic agent with enhanced photo/chemodynamic and photothermal therapy
,”
Biomater. Sci.
9
(
12
),
4473
4483
(
2021
).
158.
J.
Dai
,
J.
Chen
,
J.
Song
,
Y.
Ji
,
Y.
Qiu
,
Z.
Hong
,
H.
Song
,
L.
Yang
,
Y.
Zhu
,
L.
Li
,
H.
Yang
, and
Z.
Hu
, “
Photodynamic therapy: When van der Waals heterojunction meets tumor
,”
Chem. Eng. J.
421
,
129773
(
2021
).
159.
Y.
Li
,
L.
Wang
,
K.
Kang
,
Y.
Ma
,
K.
Yu
,
T.
Lu
,
F.
Qu
, and
H.
Lin
, “
All in one theranostic nanoagent based on MoSe2/Au@PEG hollow nanospheres for the enhanced synergetic antitumor
,”
Chem. Eng. J.
429
,
132330
(
2022
).
160.
N.
Elavarasan
,
S.
Vignesh
,
M.
Srinivasan
,
G.
Venkatesh
,
G.
Palanisamy
,
P.
Ramasamy
,
B.
Palanivel
,
A. M.
Al-Enizi
,
M.
Ubaidullah
,
V. R.
Minnam Reddy
, and
W. K.
Kim
, “
Synergistic S-Scheme mechanism insights of g-C3N4 and rGO combined ZnO-Ag heterostructure nanocomposite for efficient photocatalytic and anticancer activities
,”
J. Alloys Compd.
906
,
164255
(
2022
).
161.
B.
Zheng
,
R.
Zhang
,
F.
Kuang
,
T.
Hui
,
C.
Fu
,
L.
Zhang
,
C.
Zhou
,
M.
Qiu
, and
B.
Yue
, “
Schottky heterojunction CeO2 @MXene nanosheets with synergistic type I and type II PDT for anti-osteosarcoma
,”
J. Mater. Chem. B
12
(
7
),
1816
1825
(
2024
).
162.
V. G.
Deepagan
,
D. G.
You
,
W.
Um
,
H.
Ko
,
S.
Kwon
,
K. Y.
Choi
,
G.-R.
Yi
,
J. Y.
Lee
,
D. S.
Lee
,
K.
Kim
,
I. C.
Kwon
, and
J. H.
Park
, “
Long-circulating Au-TiO2 nanocomposite as a sonosensitizer for ROS-mediated eradication of cancer
,”
Nano Lett.
16
(
10
),
6257
6264
(
2016
).
163.
Y.
Cao
,
T.
Wu
,
W.
Dai
,
H.
Dong
, and
X.
Zhang
, “
TiO2 nanosheets with the Au nanocrystal-decorated edge for mitochondria-targeting enhanced sonodynamic therapy
,”
Chem. Mater.
31
(
21
),
9105
9114
(
2019
).
164.
F.
Gong
,
L.
Cheng
,
N.
Yang
,
Y.
Gong
,
Y.
Ni
,
S.
Bai
,
X.
Wang
,
M.
Chen
,
Q.
Chen
, and
Z.
Liu
, “
Preparation of TiH1.924 nanodots by liquid-phase exfoliation for enhanced sonodynamic cancer therapy
,”
Nat. Commun.
11
(
1
),
3712
(
2020
).
165.
X.
Wang
,
X.
Wang
,
Q.
Yue
,
H.
Xu
,
X.
Zhong
,
L.
Sun
,
G.
Li
,
Y.
Gong
,
N.
Yang
,
Z.
Wang
,
Z.
Liu
, and
L.
Cheng
, “
Liquid exfoliation of TiN nanodots as novel sonosensitizers for photothermal-enhanced sonodynamic therapy against cancer
,”
Nano Today
39
,
101170
(
2021
).
166.
Y.
Zhao
,
S.
Wang
,
Y.
Ding
,
Z.
Zhang
,
T.
Huang
,
Y.
Zhang
,
X.
Wan
,
Z. L.
Wang
, and
L.
Li
, “
Piezotronic effect-augmented Cu2–xO–BaTiO3 sonosensitizers for multifunctional cancer dynamic therapy
,”
ACS Nano
16
(
6
),
9304
9316
(
2022
).
167.
T.
Zhang
,
Q.
Zheng
,
C.
Xie
,
G.
Fan
,
Y.
Wang
,
Y.
Wu
,
Y.
Fu
,
J.
Huang
,
D. Q. M.
Craig
,
X.
Cai
, and
X.
Li
, “
Integration of silica nanorattles with manganese-doped In2S3/InOOH to enable ultrasound-mediated tumor theranostics
,”
ACS Appl. Mater. Interfaces
15
(
4
),
4883
4894
(
2023
).
168.
X.
Dong
,
R.
Cheng
,
S.
Zhu
,
H.
Liu
,
R.
Zhou
,
C.
Zhang
,
K.
Chen
,
L.
Mei
,
C.
Wang
,
C.
Su
,
X.
Liu
,
Z.
Gu
, and
Y.
Zhao
, “
A heterojunction structured WO2.9-WSe2 nanoradiosensitizer increases local tumor ablation and checkpoint blockade immunotherapy upon low radiation dose
,”
ACS Nano
14
(
5
),
5400
5416
(
2020
).
169.
Y.
Xiong
,
Y.
Rao
,
J.
Hu
,
Z.
Luo
, and
C.
Chen
, “
Nanoparticle-based photothermal therapy for breast cancer noninvasive treatment
,”
Adv. Mater.
e2305140
(published online) (
2023
).
170.
L.
Zhao
,
X.
Zhang
,
X.
Wang
,
X.
Guan
,
W.
Zhang
, and
J.
Ma
, “
Recent advances in selective photothermal therapy of tumor
,”
J. Nanobiotechnol.
19
(
1
),
335
(
2021
).
171.
X.
Wan
,
Y.
Ge
,
J.
Zhang
,
W.
Pan
,
N.
Li
, and
B.
Tang
, “
A covalent organic framework derived n-doped carbon nanozyme as the all-rounder for targeted catalytic therapy and NIR-II photothermal therapy of cancer
,”
ACS Appl. Mater. Interfaces
15
(
38
),
44763
44772
(
2023
).
172.
X.
Huang
,
W.
Zhang
,
G.
Guan
,
G.
Song
,
R.
Zou
, and
J.
Hu
, “
Design and functionalization of the NIR-responsive photothermal semiconductor nanomaterials for cancer theranostics
,”
Acc. Chem. Res.
50
(
10
),
2529
2538
(
2017
).
173.
Y.
Cheng
,
Y.
Chang
,
Y.
Feng
,
H.
Jian
,
Z.
Tang
, and
H.
Zhang
, “
Deep‐level defect enhanced photothermal performance of bismuth sulfide–gold heterojunction nanorods for photothermal therapy of cancer guided by computed tomography imaging
,”
Angew. Chem., Int. Ed.
57
(
1
),
246
251
(
2018
).
174.
M.
Yuan
,
L.
Yang
,
Z.
Yang
,
Z.
Ma
,
J.
Ma
,
Z.
Liu
,
P.
Ma
,
Z.
Cheng
,
A.
Maleki
, and
J.
Lin
, “
Fabrication of interface engineered S‐scheme heterojunction nanocatalyst for ultrasound‐triggered sustainable cancer therapy
,”
Adv. Sci.
11
,
2308546
(
2024
).
175.
Z.
Zhang
and
J. T.
Yates
, “
Band bending in semiconductors: Chemical and physical consequences at surfaces and interfaces
,”
Chem. Rev.
112
(
10
),
5520
5551
(
2012
).
176.
J.
Li
,
Z.
Li
,
X.
Liu
,
C.
Li
,
Y.
Zheng
,
K. W. K.
Yeung
,
Z.
Cui
,
Y.
Liang
,
S.
Zhu
,
W.
Hu
,
Y.
Qi
,
T.
Zhang
,
X.
Wang
, and
S.
Wu
, “
Interfacial engineering of Bi2S3/Ti3C2Tx MXene based on work function for rapid photo-excited bacteria-killing
,”
Nat. Commun.
12
(
1
),
1224
(
2021
).
177.
N.
Yu
,
C.
Peng
,
Z.
Wang
,
Z.
Liu
,
B.
Zhu
,
Z.
Yi
,
M.
Zhu
,
X.
Liu
, and
Z.
Chen
, “
Dopant-dependent crystallization and photothermal effect of Sb-doped SnO2 nanoparticles as stable theranostic nanoagents for tumor ablation
,”
Nanoscale
10
(
5
),
2542
2554
(
2018
).
178.
S.
In
,
A.
Orlov
,
R.
Berg
,
F.
García
,
S.
Pedrosa-Jimenez
,
M. S.
Tikhov
,
D. S.
Wright
, and
R. M.
Lambert
, “
Effective visible light-activated B-doped and B, N-codoped TiO2 Photocatalysts
,”
J. Am. Chem. Soc.
129
(
45
),
13790
13791
(
2007
).
179.
A. W.
Morawski
,
M.
Janus
,
B.
Tryba
,
M.
Inagaki
, and
K.
Kałucki
, “
TiO2-anatase modified by carbon as the photocatalyst under visible light
,”
C. R. Chim.
9
(
5–6
),
800
805
(
2005
).
180.
R.
Asahi
,
T.
Morikawa
,
T.
Ohwaki
,
K.
Aoki
, and
Y.
Taga
, “
Visible-light photocatalysis in nitrogen-doped titanium oxides
,”
Science
293
(
5528
),
269
271
(
2001
).
181.
W. Y.
Teoh
,
R.
Amal
,
L.
Mädler
, and
S. E.
Pratsinis
, “
Flame sprayed visible light-active Fe-TiO2 for photomineralisation of oxalic acid
,”
Catal. Today
120
(
2
),
203
213
(
2007
).
182.
J.
Gong
,
C.
Yang
,
W.
Pu
, and
J.
Zhang
, “
Liquid phase deposition of tungsten doped TiO2 films for visible light photoelectrocatalytic degradation of dodecyl-benzenesulfonate
,”
Chem. Eng. J.
167
(
1
),
190
197
(
2011
).
183.
K.
Gao
,
W.
Tu
,
X.
Yu
,
F.
Ahmad
,
X.
Zhang
,
W.
Wu
,
X.
An
,
X.
Chen
, and
W.
Li
, “
W-doped TiO2 nanoparticles with strong absorption in the NIR-II window for photoacoustic/CT dual-modal imaging and synergistic thermoradiotherapy of tumors
,”
Theranostics
9
(
18
),
5214
5226
(
2019
).
184.
Y.
Chang
,
Y.
Cheng
,
Y.
Feng
,
H.
Jian
,
L.
Wang
,
X.
Ma
,
X.
Li
, and
H.
Zhang
, “
Resonance energy transfer-promoted photothermal and photodynamic performance of gold–copper sulfide yolk–shell nanoparticles for chemophototherapy of cancer
,”
Nano Lett.
18
(
2
),
886
897
(
2018
).
185.
S.
Liang
,
X.
Deng
,
Y.
Chang
,
C.
Sun
,
S.
Shao
,
Z.
Xie
,
X.
Xiao
,
P.
Ma
,
H.
Zhang
,
Z.
Cheng
, and
J.
Lin
, “
Intelligent hollow Pt-CuS janus architecture for synergistic catalysis-enhanced sonodynamic and photothermal cancer therapy
,”
Nano Lett.
19
(
6
),
4134
4145
(
2019
).
186.
Y.
Huang
,
H.
Yang
,
X.
Lu
,
M.
Chen
, and
W.
Shi
, “
Near infrared-driven photocatalytic overall water splitting: Progress and perspective
,”
Chin. J. Catal.
58
,
105
122
(
2024
).
187.
N.
Yu
,
Y.
Hu
,
X.
Wang
,
G.
Liu
,
Z.
Wang
,
Z.
Liu
,
Q.
Tian
,
M.
Zhu
,
X.
Shi
, and
Z.
Chen
, “
Dynamically tuning near-infrared-induced photothermal performances of TiO2 nanocrystals by Nb doping for imaging-guided photothermal therapy of tumors
,”
Nanoscale
9
(
26
),
9148
9159
(
2017
).
188.
H.
Jiang
,
J.
Sun
,
F.
Liu
,
Y.
Zhao
,
X.
Chen
,
C.
Dai
, and
Z.
Wen
, “
Bi2S3/Ti3C2-TPP nano-heterostructures induced by near-infrared for photodynamic therapy combined with photothermal therapy on hypoxic tumors
,”
J. Nanobiotechnol.
22
(
1
),
123
(
2024
).
189.
Z.-H.
Li
,
Y.
Chen
,
Y.
Sun
, and
X.-Z.
Zhang
, “
Platinum-doped prussian blue nanozymes for multiwavelength bioimaging guided photothermal therapy of tumor and anti-inflammation
,”
ACS Nano
15
(
3
),
5189
5200
(
2021
).
190.
X.
Deng
,
H.
Liu
,
Y.
Xu
,
L.
Chan
,
J.
Xie
,
Z.
Xiong
,
Z.
Tang
,
F.
Yang
, and
T.
Chen
, “
Designing highly stable ferrous selenide-black phosphorus nanosheets heteronanostructure via P-Se bond for MRI-guided photothermal therapy
,”
J. Nanobiotechnol.
19
(
1
),
201
(
2021
).
191.
P.
Xu
,
C.
Wen
,
C.
Gao
,
H.
Liu
,
Y.
Li
,
X.
Guo
,
X.-C.
Shen
, and
H.
Liang
, “
Near-infrared-II-activatable self-assembled manganese porphyrin-gold heterostructures for photoacoustic imaging-guided sonodynamic-augmented photothermal/photodynamic therapy
,”
ACS Nano
18
(
1
),
713
727
(
2024
).
192.
L.
Yang
,
Y.
Zhu
,
A.
Ma
,
W.
Gao
,
Z.
Guo
,
Y.
Xin
,
R.
Ran
, and
X.
Yang
, “
Near-infrared laser/H2O2 dual-responsive multifunctional mesoporous polydopamine nanoparticles for tumor photothermal/chemodynamic/CO combination therapy
,”
J. Photochem. Photobiol. A
435
,
114341
(
2023
).
193.
D.-W.
Zheng
,
Y.
Chen
,
Z.-H.
Li
,
L.
Xu
,
C.-X.
Li
,
B.
Li
,
J.-X.
Fan
,
S.-X.
Cheng
, and
X.-Z.
Zhang
, “
Optically-controlled bacterial metabolite for cancer therapy
,”
Nat. Commun.
9
(
1
),
1680
(
2018
).
194.
R. K.
Vayalakkara
,
C.-L.
Lo
,
H.-H.
Chen
,
M.-Y.
Shen
,
Y.-C.
Yang
,
A.
Sabu
,
Y.-F.
Huang
, and
H.-C.
Chiu
, “
Photothermal/NO combination therapy from plasmonic hybrid nanotherapeutics against breast cancer
,”
J. Controlled Release
345
,
417
432
(
2022
).
195.
P. P.
Oza
and
K.
Kashfi
, “
The triple crown: NO, CO, and H2S in cancer cell biology
,”
Pharmacol. Ther.
249
,
108502
(
2023
).
196.
A.
Ahmad
,
A. A.
Baig
,
M.
Hussain
,
M. U.
Saeed
,
M.
Bilal
,
N.
Ahmed
,
H.
Chopra
,
M.
Hassan
,
M.
Rachamalla
,
S. K.
Putnala
,
M.
Khaliq
,
Z.
Tahir
, and
M. A.
Kamal
, “
Narrative on hydrogen therapy and its clinical applications: Safety and efficacy
,”
Curr. Pharm. Des.
28
(
31
),
2519
2537
(
2022
).
197.
Z.
Kou
,
P.
Zhao
,
Z.
Wang
,
Z.
Jin
,
L.
Chen
,
B.-L.
Su
, and
Q.
He
, “
Acid-responsive H2-releasing Fe nanoparticles for safe and effective cancer therapy
,”
J. Mater. Chem. B
7
(
17
),
2759
2765
(
2019
).
198.
M.
Fan
,
Y.
Wen
,
D.
Ye
,
Z.
Jin
,
P.
Zhao
,
D.
Chen
,
X.
Lu
, and
Q.
He
, “
Acid‐responsive H2‐releasing 2D MgB2 nanosheet for therapeutic synergy and side effect attenuation of gastric cancer chemotherapy
,”
Adv. Healthcare Mater.
8
(
13
),
1900157
(
2019
).
199.
H.
Ding
,
F.
Ren
,
P.
Liu
,
Y.
Feng
,
X.
Ma
,
Z.
Shen
,
Q.
Shi
,
M.
Xu
,
W.
Li
, and
H.
Chen
, “
Cu2+ anchored carbon nano‐photocatalysts for visible water splitting to boost hydrogen cuproptosis
,”
Angew. Chem., Int. Ed.
62
(
44
),
202311549
(
2023
).
200.
Y.
Wu
,
L.
Su
,
M.
Yuan
,
T.
Chen
,
J.
Ye
,
Y.
Jiang
,
J.
Song
, and
H.
Yang
, “
In vivo X‐ray triggered catalysis of H2 generation for cancer synergistic gas radiotherapy
,”
Angew. Chem., Int. Ed.
60
(
23
),
12868
12875
(
2021
).
201.
Y.
Kang
,
Z.
Mao
,
Y.
Wang
,
C.
Pan
,
M.
Ou
,
H.
Zhang
,
W.
Zeng
, and
X.
Ji
, “
Design of a two-dimensional interplanar heterojunction for catalytic cancer therapy
,”
Nat. Commun.
13
(
1
),
2425
(
2022
).
202.
L.
Wang
,
K.
Kang
,
H.
Hou
,
Y.
Ma
,
K.
Yu
,
F.
Qu
, and
H.
Lin
, “
NIR-II-driven intracellular photocatalytic oxygen-generation on Z-Scheme iron sulfide/cobalt sulfide nanosheets for hypoxic tumor therapy
,”
J. Colloid Interface Sci.
625
,
145
157
(
2022
).
203.
X.
Cui
,
Q.
Zhao
,
Z.
Huang
,
Y.
Xiao
,
Y.
Wan
,
S.
Li
, and
C.
Lee
, “
Water‐splitting based and related therapeutic effects: Evolving concepts, progress, and perspectives
,”
Small
16
(
47
),
2004551
(
2020
).
204.
Z.
Yang
,
Y.
Luo
,
H.
Yu
,
K.
Liang
,
M.
Wang
,
Q.
Wang
,
B.
Yin
, and
H.
Chen
, “
Reshaping the tumor immune microenvironment based on a light‐activated nanoplatform for efficient cancer therapy
,”
Adv. Mater.
34
(
11
),
2108908
(
2022
).
205.
R.
Motterlini
and
L. E.
Otterbein
, “
The therapeutic potential of carbon monoxide
,”
Nat. Rev. Drug Discov.
9
(
9
),
728
743
(
2010
).
206.
D.
Zheng
,
B.
Li
,
C.
Li
,
L.
Xu
,
J.
Fan
,
Q.
Lei
, and
X.
Zhang
, “
Photocatalyzing CO2 to CO for enhanced cancer therapy
,”
Adv. Mater.
29
(
44
),
1703822
(
2017
).
207.
S.-B.
Wang
,
C.
Zhang
,
J.-J.
Ye
,
M.-Z.
Zou
,
C.-J.
Liu
, and
X.-Z.
Zhang
, “
Near-infrared light responsive nanoreactor for simultaneous tumor photothermal therapy and carbon monoxide-mediated anti-inflammation
,”
ACS Cent. Sci.
6
(
4
),
555
565
(
2020
).
208.
L.
Chen
,
Z.
Mao
,
Y.
Wang
,
Y.
Kang
,
Y.
Wang
,
L.
Mei
, and
X.
Ji
, “
Edge modification facilitated heterogenization and exfoliation of two-dimensional nanomaterials for cancer catalytic therapy
,”
Sci. Adv.
8
(
39
),
eabo7372
(
2022
).
209.
L.
Pan
,
H.
Huang
,
C. K.
Lim
,
Q. Y.
Hong
,
M. S.
Tse
, and
O. K.
Tan
, “
TiO2 rutile–anatase core–shell nanorod and nanotube arrays for photocatalytic applications
,”
RSC Adv.
3
(
11
),
3566
3571
(
2013
).
210.
Z.
Xiong
,
Q.
Liu
,
Z.
Gao
,
J.
Yang
,
X.
Zhang
,
Q.
Yang
, and
C.
Hao
, “
Heterogeneous interface design to enhance the photocatalytic performance
,”
Inorg. Chem.
60
(
7
),
5063
5070
(
2021
).
Published open access through an agreement withNational Tsing Hua University