Next Article in Journal
Antimicrobial Peptides in the Battle against Orthopedic Implant-Related Infections: A Review
Next Article in Special Issue
Polylactide Perspectives in Biomedicine: From Novel Synthesis to the Application Performance
Previous Article in Journal
Cyclodextrin as Functional Carrier in Development of Mucoadhesive Tablets Containing Polygoni cuspidati Extract with Potential for Dental Applications
Previous Article in Special Issue
Chitosan-Based Biocompatible Copolymers for Thermoresponsive Drug Delivery Systems: On the Development of a Standardization System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Stimuli Responsive Nitric Oxide-Based Nanomedicine for Synergistic Therapy

1
Zhuhai Institute of Translational Medicine, Zhuhai Precision Medical Center, Zhuhai People’s Hospital (Zhuhai Hospital Affiliated with Jinan University), Zhuhai 519000, China
2
The Department of Medical Imaging, Zhuhai People’s Hospital (Zhuhai Hospital Affiliated with Jinan University), Zhuhai 519000, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Pharmaceutics 2021, 13(11), 1917; https://doi.org/10.3390/pharmaceutics13111917
Submission received: 28 August 2021 / Revised: 14 October 2021 / Accepted: 22 October 2021 / Published: 12 November 2021
(This article belongs to the Special Issue Polymer Therapeutics: From Synthesis to Biomedical Applications)

Abstract

:
Gas therapy has received widespread attention from the medical community as an emerging and promising therapeutic approach to cancer treatment. Among all gas molecules, nitric oxide (NO) was the first one to be applied in the biomedical field for its intriguing properties and unique anti-tumor mechanisms which have become a research hotspot in recent years. Despite the great progress of NO in cancer therapy, the non-specific distribution of NO in vivo and its side effects on normal tissue at high concentrations have impaired its clinical application. Therefore, it is important to develop facile NO-based nanomedicines to achieve the on-demand release of NO in tumor tissue while avoiding the leakage of NO in normal tissue, which could enhance therapeutic efficacy and reduce side effects at the same time. In recent years, numerous studies have reported the design and development of NO-based nanomedicines which were triggered by exogenous stimulus (light, ultrasound, X-ray) or tumor endogenous signals (glutathione, weak acid, glucose). In this review, we summarized the design principles and release behaviors of NO-based nanomedicines upon various stimuli and their applications in synergistic cancer therapy. We also discuss the anti-tumor mechanisms of NO-based nanomedicines in vivo for enhanced cancer therapy. Moreover, we discuss the existing challenges and further perspectives in this field in the aim of furthering its development.

1. Introduction

Globally, cancer is one of the most serious diseases threatening human health. Despite the rapid advances made in diagnosis and treatment technologies, it remains a great challenge due to the heterogeneity and complexity of tumors themselves [1,2,3,4]. Therefore, basic and clinical studies for effective cancer treatment are still in desperate need in order to improve the health of human beings [5,6,7,8]. Presently, common treatments for tumors are chemotherapy, radiotherapy, surgery and biological treatment [9,10,11]. However, clinical outcomes have proven that these treatments exhibit certain limitations and side effects such as an off-targeting phenomenon and damage to normal cells, which hinder the significant improvement of the survival rate of patients [12,13,14,15]. On this accession, it is significant to explore new types of tumor treatment methods for superior therapeutic efficiency and reduced side effects. Compared with traditional anti-tumor methods, gas therapy has received widespread attention as an emerging and promising therapeutic approach to cancer treatment [5]. Many gas molecules such as nitric oxide (NO), carbon monoxide (CO), hydrogen sulfide (H2S) and hydrogen selenide (H2Se) have been proven to participate in the physiological and pathological processes of living subjects as signal transduction molecules [16]. Therefore, regulating the gas concentrations in the tumor microenvironment (TME) could profoundly affect the occurrence and development of tumors [17]. To begin with, previous studies have demonstrated that therapeutic gas molecules could effectively pass through the biological barriers owing to their diffusivity [18]. Secondly, some therapeutic gas molecules could reverse the Warburg effect in tumor tissue, thereby inhibiting the activity of tumor cells without harming the normal cells at certain concentrations [19,20]. In addition, the anti-tumor effects of gas therapy are multiple. For example, NO has been demonstrated to strengthen the enhanced permeability and retention (EPR) effect by vasodilation, which could enhance the accumulation efficiency of nanomedicines [21]. Moreover, NO, CO and sulfur dioxide (SO2) could enhance the effect of chemotherapy by overcoming the multidrug resistance (MDR) of tumor cells [22,23,24]. Furthermore, hydrogen (H2) has been found to kill tumor cells and protect normal cells at the same time at proper concentrations [25]. In addition, our research group also achieved an anti-tumor effect by using responsively released H2Se which effectively inhibited tumor metastasis by down-regulating the expression of metastasis-related proteins [26]. Finally, some gas molecules could enhance the effect of radiotherapy, photodynamic therapy (PDT) or immunotherapy by increasing the oxygenation ability of tumor cells and so on [27,28,29,30,31]. Therefore, as a new and promising anti-tumor method, gas therapy has increasingly received attention in anti-tumor therapy and presented itself as a new opportunity for the treatment of cancer [18,22,32].
To date, commonly used gas molecules in cancer therapy have been NO, CO, H2, H2S, SO2 and H2Se, among which NO was the first one to be developed in the biomedicine field [33,34,35]. In 1998, Dr. Louis J. Ignarro won the Nobel Prize for discovering the role of NO in the cardiovascular system. Since then, numerous studies have focused on the biomedical applications of NO [36,37]. It has been found that NO plays an important role in the cardiovascular system, respiratory system, nervous system, immune system, antibacterial treatment, wound healing and other biomedical fields [38,39,40]. It is especially in the field of cancer treatment that an increasing number of studies have demonstrated that NO presents direct or indirect anti-tumor effects at different concentrations [41,42,43,44]. The main anti-tumor mechanisms of NO are listed as follows: (1) NO firmly binds to the blood iron center of many proteins and thus inhibits the mitochondrial respiratory metabolism of cancer cells [5,33,45]; (2) NO combines with intracellular superoxide anions to produce free radicals that damage the DNA of tumor cells [33,39]; (3) NO could effectively dilate the blood vessels at the tumor site leading to an enhanced EPR effect [27,28,46]. The main mechanism is the combination of NO to the heme group of the soluble guanylyl cyclase (sGC) which enhances the activity of sGC, thereby enhancing the activity of protein kinase G and reducing the intracellular Ca2+ level, leading to a vasodilation effect [47]. Meanwhile, NO can also reduce platelet aggregation by decreasing Ca2+ level in platelets, thus inhibiting the integrity of tumor blood vessels and ultimately enhancing the EPR effect [48]; (4) P-glycoprotein (P-gp), one of the adenosine triphosphate (ATP) binding cassette transporters, is over-expressed in tumor cell membranes [30,49]. NO can nitrate tyrosine residues in P-gp and its transport function, thus inhibiting the drug outflow from tumor cells and overcoming the MDR of tumor cells to chemotherapy [48]; (5) NO could also be applied as a radiosensitizer to improve the efficacy of radiotherapy for hypoxic tumors [50,51]. Therefore, NO-based gas therapy has been widely applied in tumor therapy with multiple functions and applications.
Despite the many advantages of the therapeutic effect of NO gas in tumor therapy, there are still limitations to its direct application in vivo. First of all, NO is a double-edged sword gas which means that NO could promote the growth of cancer cells at low concentrations (<nM level) mediating antioxidants, signaling and positive bioenergetic mechanisms, while effectively killing high concentrations (≥nM level) of cancer cells by inhibiting mitochondrial respiratory metabolism [52,53]. It should be noted that a previous study found that NO causes blood poisoning above the concentration of 25 ppm in gas inhalation concentration [5]. Therefore, only in a certain range of concentration could NO exert an effective anti-tumor effect [54,55,56]. Secondly, the half-life of NO in vivo is very short, and it is difficult for the concentration of free NO to reach desired levels in tumor tissue [57]. In addition, NO exhibited nonspecific distribution in vivo, leading to its limited therapeutic effect and severe side effects [56]. Therefore, achieving the on-demand release of NO in tumor tissue is a key point to achieve the safe and effective treatment of tumors [56,58,59].
In recent years, numerous studies have focused on the controlled release of NO by designing NO-based nanomedicines [60]. The release of NO could be controlled by an exogenous stimulus such as light, ultrasound and X-rays, which shows the characteristics of non-invasive and precise control with space–time selection. In addition, the TME has been found to be a critical foundation for tumor initiation and progression [61,62]. It exhibits unique pathological features such as a low oxygen level, low pH level, high-level glutathione (GSH), high-level glucose, high reducibility, high-level hydrogen peroxide (H2O2) and over-expressed enzymes [62,63,64,65]. Taking advantage of the TME, NO-based nanomedicines which could respond to low pH, high-level GSH and high-level glucose were developed [62,63,64]. The developed nanomedicines could remain stable in vitro or in normal tissues, which guaranteed the biosafety of the nanomedicines. Upon reaching the tumor(s), an exogenous or endogenous stimulus could trigger the on-demand release of NO which achieves the effective treatment of tumor. The most commonly used NO donors are N-Diazeniumdiolates (NONOates), S-Nitrosothiols (RSNOs), metal–nitrosyl complexes (such as Roussin’s black salt, red salt dianinion and ruthenium-nitrosyl) and nitrobenzene derivatives (such as N,N′-di-sec-butyl-N,N′-dinitroso-1,4-phenylenediamine, BNN6) [66]. Diverse NO donors usually exhibit different characteristics and properties. For example, NONOates can achieve the rapid release of NO gas under acidic conditions, but they are also not very stable under physiological conditions, leading to non-negligible premature leakage; RSNOs are relatively stable under physiological conditions while chemical bonds’ breakage and NO release can be achieved under hyperthermia [67]. Based on the advantages and disadvantages of different NO donors, many new responsive NO-releasing nanomedicines have been developed, as shown in Table 1. This review summarizes the design principles and release mechanisms of NO-based nanomedicines upon various stimuli and their applications in synergistic cancer therapy—as shown in Scheme 1.

2. Exogenous Stimuli-Responsive NO Nanomedicines

Since the exogenous stimuli can be conveniently controlled in vitro, it is advantageous for the on-demand release of NO-based nanomedicines in a spatio-temporal manner [73,106]. Therefore, numerous studies have reported NO-based nanomedicines controlled by exogenous stimulus in recent years [107,108]. External stimuli including light, ultrasound and X-rays provide a promising means of implementing the controlled release of NO which has achieved effective anti-tumor effects.

2.1. Light-Triggered NO Nanomedicines

2.1.1. Ultraviolet–Visible Triggered NO Nanomedicines

Due to their non-invasive, practical, inexpensive and easily controllable nature, light has received significant attention as an exogenous stimulus in biomedical applications [109,110,111,112]. It has been reported that a large part of the light-sensitive NO donors could release NO in a controlled manner with the irradiation of ultraviolet–visible (UV–vis) light [113,114,115]. For example, a type of NO donor, named N,N′-Di-sec-butyl-N,N′-dinitroso-1,4-phenylenediamine (BNN6), can achieve the controlled release of NO under UV/vis irradiation. However, the hydrophobicity and toxicity of BNN6 hinder its direct biological application [82]. In view of this, Fan et al. encapsulated BNN6 with block polymer mPEG-PLGA, which exhibited superior aqueous dispersibility and stability. The obtained nanosystem could achieve the on-demand release of NO under UV–vis irradiation. The scanning electron microscopy (SEM) image indicated that the nanomedicine showed a uniform sperm shape before UV light irradiation and a collapsed state after UV light irradiation for 2 min. Meanwhile, NO was simultaneously produced, which effectively reverses the MDR of tumor cells by inhibiting the expression of P-gp on tumor cells (Figure 1A) [68]. Given that mitochondria are an important target for anti-tumor therapy, Wu et al. designed a mitochondrial-targeting photo-responsive NO-releasing nanomedicine using carbon dots (C-dots) as the carrier loaded with NO donor S-nitrosothiol (R-SNO) derivatives. It was found that the obtained nanomedicine was able to target mitochondria and release NO in response to the irradiation by UV–vis laser, which specifically damaged the mitochondria in tumor cell (Figure 1B) [69]. Furthermore, Liu et al. designed a folate-modified NO-based nanomedicine using metal nitrosyls Ru-NO as the NO donor. It was found that the nanomedicine could target human cervical HeLa cells that over-expressed folate receptors and released NO as well as singlet oxygen in response to UV–vis stimulation, resulting in a synergistic anti-tumor effect (Figure 1C) [70].
In addition to its therapeutic effect on tumors, NO has also shown unpredictable potential in the treatment of other diseases [66]. For example, as a gaseous signaling molecule, NO has been demonstrated to facilitate the processes of wound healing [116]. NO is involved in vascular dilation and anti-platelet activity during inflammation, which can promote angiogenesis at the stage of proliferation and collagen deposition at the stage of remodeling. Therefore, Duan et al. designed a nanosystem by directly polymerizing an NO donor: n-nitrosoamine-based monomers and wrapping them in a vesicle bilayer structure. The nanosystem could accurately release NO when stimulated by 365 nm UV, which could promote wound healing in a corneal wound model [117].

2.1.2. First Near-Infrared (NIR-I) Photothermal Triggered NO Nanomedicines

Although many advances have been achieved in the UV–vis light-triggered NO releasing system, the low tissue penetration and high phototoxicity of UV–vis light limit its application in the field of anti-tumor therapy [118]. Compared with UV–vis light, NIR light has become an emerging hotspot because of its high tissue penetration and low photoxicity [53,119,120,121]. NIR can not only stimulate the release of NO from photo-sensitive NO donors but also mediate the role of photosensitizers in phototherapy, including photothermal therapy (PTT) and photodynamic therapy (PDT) [81,122,123]. PTT refers to the use of photosensitive agents to absorb NIR and convert it into heat to kill tumor cells [124]. Therefore, the use of photosensitive NO donors can effectively combine PTT and NO to play a synergistic anti-tumor role [125,126,127,128]. Depending on the wavelength, NIR could be divided into two categories: first NIR window (650–900 nm) and second NIR window (900–1200 nm). NIR-I light has been the most used NIR light in the treatment of tumors in previous studies. In this section, NIR-I light triggered NO nanomedicines are introduced and analyzed.
It has been reported that MDR is an important reason for chemotherapy failure [129,130]. Thus, Yang et al. used Fe3O4@polydopamine@mSiO2 as a photothermal carrier to prepare NIR light-triggered nanomedicine. By receiving NIR stimulation and converting NIR light into heat, the nanomedicine could break the chemical bond of R-SNO and achieve the controlled release of NO, which could reverse the MDR and realize the combined treatment of tumor with NO therapy, chemotherapy and PTT (Figure 2A) [71]. To further enhance the light absorption rate and photothermal conversion efficiency, Huang et al. used intelligent protoporphyrin with high light absorptivity as a photosensitizer to design a novel polymer nanomedicine. The nanomedicine could efficiently release NO under the irradiation of NIR, which could effectively kill tumor cells with low laser power density (Figure 2B) [72]. Furthermore, Liu et al. designed a novel multifunctional NIR-responsive magnetic nanoplatform, which can be guided by a magnetic field to achieve high nanoparticle accumulation in tumor tissue. Under the irradiation of an 808 nm laser, the photothermal effects led to the on-demand release of NO, which achieved remarkable anti-tumor effects [73].
In the process of PTT, high temperature is needed to effectively kill tumor cells, but it will inevitably cause damage to the surrounding normal tissues. In order to prevent this, researchers have tried to use a mild photothermal therapy (mPTT) strategy to solve this dilemma, which, however, can easily lead to the self-repair of tumor cells after insufficient heat damage [131]. In this context, Zhao et al. designed a nanocomposite using bismuth sulfide (Bi2S3) nanoparticles as a photothermal carrier loaded with an NO donor. The obtained nanocomposite could achieve the mPTT effect under the irradiation of NIR laser, which led to the controlled release of NO, thereby improving the anti-tumor effect of mPTT (Figure 2C) [74]. Furthermore, Dong et al. designed an NIR responsive nanocomposite platform to combine the mPTT with gas and chemotherapy. Under the stimulation of the NIR laser, the thermosensitive NO donor achieved the on-demand release of NO, which could overcome the MDR of tumor cells and enhanced the chemotherapy effect of doxorubicin (DOX). At the same time, NO and chemotherapy could effectively compensate for the therapeutic deficiency of mPTT, achieving a combined anti-tumor effect [75]. With the further exploration of mPTT in tumor therapy, studies have found that cunning tumor cells produced a series of heat shock proteins, which protected tumor cells from damage at low heat levels [132]. Therefore, Wang et al. encapsulated heat shock protein 70 (HSP70) inhibitor 2-phenylethyne sulfonamide (PES) into the NO-based nanomedicine. The mild heat generated by the nanomedicine under the irradiation of NIR could not only induce cell apoptosis with the assistance of PES at a relatively low temperature, but also released NO, thus achieving the synergistic effect of mPTT and NO therapy (Figure 2D) [76].

2.1.3. Second NIR (NIR-II) Photothermal Triggered NO Nanomedicines

Compared with NIR-I light, NIR-II exhibited deeper tissue penetration depth and higher maximum skin exposure density (1.0 w cm−2), which has emerged as a research hotspot in recent years. For example, Fan et al. designed semiconductor polymer nanoparticles (PFTDPP) with spectral adjustability and electro-optical property. Under the irradiation of 808 nm laser, the nanomedicine could absorb NIR and emit NIR-II (1000–1700 nm) to achieve the accurate tumor location by NIR-II/photoacoustic (PA) imaging. At the same time, the photothermal effect could also stimulate the release of NO from the NO donor (S-nitroso-N-acetylpenicillamine, SNAP), which represented a combination of dual-mode imaging and NO therapy to achieve the better anti-tumor effect (Figure 3A) [77]. In addition to the first NIR biological window (NIR-I) (700–950 nm), the second NIR biological window (NIR-II) (1000–1700 nm) can also be used as the stimulus to trigger the release of NO [133,134,135,136]. Compared with NIR-I, NIR-II has a deeper penetration depth of mammalian tissue and higher maximum allowable skin exposure (MPE) [137,138]. For example, owing to the excellent adsorption ability of MXenes in the NIR-II window, Xu et al. designed a type of MXenes-based nanomedicine modified with the NO donor which could achieve the synergistic/sequential anti-tumor therapy [139,140,141]. MXenes could effectively convert NIR-II laser into heat to break SNO and released NO at high temperature, thus realizing the effective inhibition effect on the tumor(s) (Figure 3B) [33]. It has been found that alkyl radical, a type of free radical independent from oxygen, can damage lipid and DNA by increasing the oxidative stress of tumor cells. Thus, Xue et al. prepared a novel nanomedicine by encapsulating the NIR-II molecule (IR 1061), the NO donor BNN6 and alkyl radical initiator 2,2-azobis[2-(2-imidazolin-2-yl) propane] dihydrochloride (AIPH) in an amphiphilic lecithin stabilized phase change material with a eutectic point of 39 °C. The nanomedicine produced photothermal effects under NIR-II irradiation to release BNN6 and AIPH, which then decomposed to form highly active NO and alkyl radical. The synergistic action of NO and alkyl radicals induced the high oxidative stress in mitochondria, which could not only accelerate the release of cytochrome C to cause mitochondrial membrane damage but also significantly down-regulate the Bcl-2 protein to trigger the apoptosis of tumor cells. Animal experiments have shown that mice treated with this strategy can survive for more than 40 days and have a tumor suppression rate of more than 70% (Figure 3C) [78]. Furthermore, Tian et al. designed a type of microsphere for the treatment of cold tumors using the photothermal agent IR1061, NO donor SNO and 1-methyl-tryptophan(1-MT). The IR1061 could achieve PTT by converting the light energy of NIR-II laser into heat energy. PTT not only induced immunogenic cell death (ICD) to recruit CD8+ cytotoxic T lymphocytes (CTL), but also act as an external stimulus which accurately triggered NO release by breaking the S-NO bond with the action of heat. The released NO normalized the tumor blood vessels by inducing endogenous angiogenic factors, which alleviated the hypoxia of TME and reprogrammed the immunosuppressive TME by, e.g., decreasing the level of immunosuppressive cells (Tregs, M2-like macrophages) and the expression of immune checkpoints (PD-L1) [142]. Therefore, a therapeutic strategy that achieved the PTT-augmented NO therapy in response to NIR-II stimulation could effectively reverse the “cold tumor” into highly immunity-invasive “hot tumor” and reprogram the immunosuppressive TME into an immunostimulating phenotype. The in vivo experiments showed that, on day 21, four of the six tumor-bearing mice survived without significant weight loss. The survival rate of tumor-bearing mice reached 66.7% on day 48 after the treatment, with efficacy in both primary breast cancer and metastatic tumor (Figure 3D) [79].

2.1.4. NIR Photodynamic Triggered NO Nanomedicines

In addition to PTT, PDT is also an effective approach to trigger the release of NO [143]. PDT involves NIR stimulating the photosensitive material, which transfers energy to the surrounding tissues to produce a large number of reactive ROS [144,145]. For example, Xiang et al. designed a dual-targeting nanoplatform to target the lysosomal organelles in the folate receptor over-expressed cancer cells. NO and ROS were released under the NIR irradiation, exerting the synergistic anti-tumor effect of gas therapy and PDT [87]. Among the conventional targeted nanoplatforms that have been reported, there are often challenges such as difficulty in synthesis and the potential immune response of the modified ligands [146,147,148]. Therefore, based on the abundance antigens on the surface of cell membrane, Zhang et al. prepared a porous coordination network that was coated with the cell membrane of the murine mammary carcinoma 4T1 cells. This nanosystem could achieve the effective tumor accumulation and negligible immune response. Under the stimulation of the NIR laser, a large amount of ROS could be directly generated to realize the PDT in tumor tissue. ROS generated by PDT could further react with NO to produce more toxic reactive nitrogen (RNS), which represents one of the anti-tumor mechanisms of NO (Figure 4A) [83]. Wei et al. synthesized a novel multifunctional nanosystem by integrating the NO donor into a photosensitizer. The nanosystem releases NO under NIR laser irradiation, which could promote vascular dilatation and overcome the MDR of tumor cells. In addition, the released ROS by PDT could react with NO, which ensured the generation efficiency of RNS with high toxicity. In addition, NO could also destroy the antioxidant system of tumor cells by consuming high concentrations of GSH in tumor cells, which further improved the efficiency of oxidative damage of RNS. This multifunctional synergistic effect could effectively guarantee the anti-tumor therapeutic effect (Figure 4B) [84]. In order to further improve the anti-tumor effect, Wang et al. designed a nanosystem to sequentially release the loaded drugs. Under the stimulation of the NIR laser, the nanosystem absorbed the laser to produce ROS, which could oxidize the NO donor L-Arginine (L-Arg) to produce NO. When NO gradually increased, the liposome layer of the nanosystem was gradually destroyed and followed by the release of DOX. The sequential release of NO and DOX could firstly enable NO to reverse MDR and provided a favorable time window and cellular environment for the anti-tumor effect of DOX (Figure 4C) [85]. In addition, hydrogels, as carriers with low toxicity, good biocompatibility, local drug delivery and slow release, are increasingly being applied in anti-tumor gas therapy [149]. For example, Sun et al. designed a responsive hydrogel using indocyanine green (ICG) as the photosensitizer and L-Arg as the NO donor. In 4T1 tumor-bearing mice, the obtained hydrogel was locally injected into tumors tissues, where ROS and NO could be produced in situ under NIR irradiation to exert synergistic anti-tumor effects. Meanwhile, studies have shown that components such as collagen or hyaluronic acid, which are highly expressed in the extracellular matrix of tumor, could block the entry of nanoparticles and immune cells into solid tumors and weaken the anti-tumor effect. The matrix metalloproteinases (MMPs) could effectively degrade tumor extracellular matrix and improve the penetration of nanoparticles and immune cells [150,151,152]. Therefore, the RNS generated by ROS and NO could activate MMPs in tumor cells which could further degrade the collagen in the extracellular matrix, which modulated the TME, leading to the improved drug sensitivity and immunotherapy. In animal experiments, the bioluminescent imaging signal intensity in mice treated with this method showed a significant increase in the efficiency of tumor growth inhibition of up to 60% (Figure 4D) [86].
Additionally, osteoporosis, a bone metabolic disease with a high risk of fracture, has been found to be reversed by NO which could promote the proliferation and differentiation of osteoblasts. Therefore, Ye et al. developed a bone-targeted upconversion nanoparticles (UCNP) by loading modified bisphosphonate—alendronate (Ald)—onto a mesoporous silica surface followed by an NO donor—BNN6. Under the irradiation of 808 nm NIR, the nanosystem could achieve the precise release of NO in the bone, thereby promoting osteoblast formation and ultimately reversing osteoporosis [153].

2.2. Ultrasound Triggered NO Nanomedicines

Compared to other stimuli, ultrasound has the advantage of deeper tissue penetration with a maximum depth reaching 20 cm at the power density of 1 MHz [154]. In addition, ultrasound exhibited the characteristics of non-invasiveness, easy to control and easy to focus on a particular area. Therefore, ultrasound responsive nanomedicines are good candidates for the controlled release of NO [155,156,157]. In the drug delivery process, it was found that the intercellular matrix hindered the efficient transport of nanomedicines to tumor tissues [158]. Therefore, Chen et al. designed ultrasonic-responsive nanosystems modified with cyclic decapeptides which can target fibrinogen and achieve the ‘relay’ transmission of the nanomedicines from the blood vessel to the stroma and then to the tumor cells, which increased the retention rate of the nanomedicines in tumor cells by almost seven times. Under the local ultrasound stimulation, the responsive release of NO from the nanomedicines could be achieved, which played a significant anti-tumor effect [93]. In addition, in order to accurately target the tumor site, He et al. designed an ultrasonic-triggered nanomedicine using paramagnetic iron oxide nanoparticles with MRI imaging guidance functions. The developed nanomedicine could be clearly localized under the MRI guidance and released NO under ultrasonic stimulation to achieve efficient and safe tumor therapy (Figure 5A) [88]. In addition, ultrasonic responsive nanomedicines can also be used in combination with sonodynamic therapy (SDT) to treat tumors. For example, Feng et al. designed a novel ultrasonic-responsive sequential nanomedicine using hollow mesoporous titanium dioxide nanoparticles loaded with anoxic active prodrug tirapazamine (TPZ) and SNO. Under the stimulation of ultrasound, the nanosystem generated ROS through SDT, which caused the breakage of SNO and released NO in a controlled manner. With the continuous function of SDT, the hypoxic environment formed by the consumption of oxygen could further effectively activate TPZ to lyse the DNA in tumor cells. The ultrasonic-responsive sequential SDT/NO therapy could effectively enhance the anti-tumor effect (Figure 5B) [89]. High-intensity focused ultrasound (HIFU) therapy can kill tumor cells by forming high temperature in the target area through the thermal effect at the focus. Kang et al. combined the thermal effect of HIFU with ultrasound responsive nanosystem to strengthen the anti-tumor effect. Under the action of HIFU, ultrasound converged to the tumor site to produce high temperature, resulting in tumor tissue degeneration and necrosis. Meanwhile, NO was effectively released at high temperature to dilate the blood vessels at the tumor site, thus promoting the effective accumulation of the injected nanomedicines (Figure 5C) [90]. In order to enable the nanosystem to target tumor cells more effectively, Zhao et al. took advantage of the homologous targeting of the tumor cell membrane by encasing NO donor GSNO and dihydroporphyrin e6 (Ce6) in the tumor cell membrane, and finally designed a bionic nanosystem with both pH and ultrasonic response. Under the action of the tumor cell membrane, the nanosystem could effectively target to the tumor tissue. In the tumor acidic environment, GSNO and Ce6 were released with the stimulation of exogenous ultrasound. In addition, ROS produced by Ce6 under ultrasound could interact with NO to produce more toxic RNS to kill tumor cells. This work is the first example to realize gas-sonodynamic combined therapy for anti-tumor therapy, thus opening a new avenue for ultrasound-triggered gas therapy (Figure 5D) [91]. In addition, platelets are known to maintain vascular integrity in tumors, whose function could be inhibited by NO. Xu et al. loaded the chemotherapy drug paclitaxel (PTX) into an albumin shell modified with the NO donor SNO, which designed an NO-based nanosystem, SNO–HSA–PTX. The nanosystem was able to respond to ultrasound stimulation which triggered the release of NO. Therefore, the function of tumor-associated platelets was inhibited, which induces the opening of the tumor vascular barrier and promotes the accumulation of PTX and immune cell invasion. Through the combination of chemotherapy, immunotherapy and NO gas therapy, remarkable tumor suppression was finally achieved with negligible side effects [92].

2.3. X-ray-Triggered NO Nanomedicines

X-rays, as an exogenous excitation source, have been widely used in imaging and cancer therapy in clinics. In recent years, X-rays have also been used as exogenous stimulation to achieve the controlled release of NO [159]. For example, Shi et al. successfully constructed intelligent X-ray-responsive upconversion nanosystems which could track and monitor the drug delivery and release process under the radiation of X-rays. Meanwhile, the water in the tumor could be decomposed into ROS under the X-rays irradiation, which reacted with SNO to trigger the release of NO, thereby achieving the combined tumor treatment (Figure 6A) [94]. Additionally, bismuth nanoparticles with high X-ray absorption coefficients can also be used to track the delivery process of the NO-based nanomedicines in real time [160]. For example, Zhang et al. used bismuth-based nanoparticles to design a multifunctional X-ray responsive nanosystem, which could be used for computed tomography (CT) imaging under X-ray radiation. Besides, the chemical bond of SNO was broken, leading to the controlled release of NO with X-ray irradiation. In addition to directly killing tumor cells, a high concentration of NO could also enhance the radiotherapy effect through Compton scattering and Auger effect, which synergistically enhanced the anti-tumor effect (Figure 6B) [95]. Although numerous progress has been made in X-ray responsive nanosystems, most of the previously reported X-ray responsive NO releasing nanosystems requires high X-ray doses to break the chemical bond of the NO donor. In this context, Hao et al. designed a novel soft X-ray responsive continuous luminescent release nanosystem by using persistent luminescent nanoparticles (PLNPs) as the carrier. When irradiated by soft X-rays (doses as low as 0.9 mGy), the nanosystem stored energy through PLNPs which continuously released NO in deep tumor tissues (up to 20 mm deep). The nanosystem continued to release NO even 40 min after the cessation of X-ray irradiation (Figure 6C) [96]. This complementary therapy strategy of the X-ray and NO brings a broader idea for tumor therapy and becomes one of the key points of anti-tumor research on stimuli-responsive NO nanomedicines in the future [161,162,163,164,165].

3. Endogenous Stimuli-Responsive NO Nanomedicines

As is already known, TME is characterized by low pH, high-level glucose, high-level GSH, high-level H2O2 and over-expressed enzymes, etc. [166,167,168,169]. Therefore, many researchers have designed endogenous responsive NO release nanomedicines taking advantage of these characteristics [106]. Compared to exogenous stimuli, endogenous stimuli are not limited by tissue depth and do not require external equipment, so they can reduce treatment costs and significantly improve the compliance of the patients. In addition, nanomedicines can be spontaneously activated at the tumor site under the action of endogenous stimulation to achieve precisely controlled NO release [170,171]. Therefore, endogenous stimulations, including low pH, high-level glucose and high-level GSH, also provide a promising pathway for the controlled release of NO in vivo.

3.1. Glutathione-Triggered NO Nanomedicines

As is already well-known, the TME is characterized by high GSH with concentration 7–10 times higher than that of normal tissues. Therefore, GSH could act as an endogenous stimulus to trigger the release of NO [172,173]. Song et al. previously synthesized a new type of NO donor, named a nitrate-functionalized D-α-tocopheryl polyethylene glycol succinate (TNO3). It was found that TNO3 could chemically react with a high concentration of GSH in TME to achieve the controlled release of NO. The accumulation of NO in tumor tissues could enhance the EPR effect and increase the accumulation of DOX in tumor sites, thus improving the efficacy of chemotherapy [101]. Zhang et al. designed a novel nanosystem which encapsulated the NO donor O2-(2,4-dinitrophenyl) 1-[4(propargyloxycarbonyl) piperazin-1-yl] diazen-1-ium-1,2-diolate (alkynyl-JSK) in an amphiphilic block copolymer via a click reaction. After entry into tumor cells, alkynyl-JSK locally generated NO through a nucleophilic reaction with GSH. Especially, when this nanosystem is co-treated in combination with DOX, the anti-tumor efficacy can be further enhanced through a synergistic effect (Figure 7A) [97]. Moreover, Park et al. used a hydrophobic nitrifying glucan (NO-Dex) as the NO donor, and encapsulated it in hydrophilic PEG to obtain a new type of nanomedicine. After entering the tumor site, NO is produced by nitrate reduction reaction between GSH and NO-Dex. The release of NO could dilate blood vessels, increase vascular permeability, and promote the positive feedback effect of EPR (Figure 7B) [98]. Furthermore, Yuan et al. used the chemotherapeutic drug hydroxycamptothecin to prepare 10-hydroxycamptothecin-loaded chitosan nanoparticles (HCPT) by the ionic gel and chemical crosslinking method. The NO donor in the nanosystem could effectively release NO with the stimulation of GSH. It was found that NO could nitrate some tyrosine residues of P-gp on the surface of tumor cell membrane, thus inhibiting the pump function and enhancing the effective accumulation of chemotherapeutic drugs in tumor cells (Figure 7C) [99]. Moreover, Deng et al. found that tumor-derived microvesicles (TMVs), which are rich in proteins, lipids and RNA, promoted tumor progression and metastasis by regulating TME. The energy required for the production and release of TMV from tumor cells was closely related to mitochondria. Therefore, they designed a nanosystem capable of delivering SNO to mitochondria for targeting the release of NO. The amide bond in the nanosystem was hydrolyzed in the acidic TME and selectively targeted the mitochondria, which enabled the NO donor to accumulate in the mitochondria. Studies have shown that GSH can react with the NO donor to generate NO and directly cut off the mitochondrial energy supply, thereby effectively inhibiting TMV and thus preventing cancer metastasis (Figure 7D) [100].

3.2. pH-Triggered NO Nanomedicines

Tumor cells produce a large amount of lactic acid because of their rapid growth and enhanced glycolysis metabolism, leading to the weakly acidic TME (pH 6.5~6.8) [174,175]. Therefore, an increasingly number of studies have taken advantage of the acidic TME and designed different pH-responsive NO-releasing nanomedicines based on it in recent years [173,176,177,178]. For example, Sung et al. designed a pH-sensitive release nanosystem loaded with chemotherapeutic drugs and the NO donor diethylenetriamine diazeniumdiolate (DETA NONOate). In the acidic environment of the tumor, H+ chemically reacts with DETA NONOate to release NO. The responsive release of NO could overcome the MDR and increase the accumulation of chemotherapeutic drugs, thereby reducing the survival rate of tumor cells (Figure 8A) [49]. In addition, based on the unique solution behavior of the biominerals’ calcium carbonate (CaCO3), Lee et al. designed a pH-responsive release nanosystem using CaCO3 as a carrier. When the nanosystem was internalized into tumor cells, acidic organelles (pH~5.0) dissolved CaCO3 and accelerated the release of NO donor GSNO. GSNO could react with reductive ascorbic acid in tumor cells to produce NO. Furthermore, the research group showed that the release of NO was much higher in the acidic pH environment than that in the neutral pH environment. It was found that a low pH is indeed a key factor to trigger the release of NO, which is expected to achieve the intracellular delivery of NO (Figure 8B) [102]. In addition, Wang et al. used the acid-responsive amphiphilic diblock copolymers to encapsulate the NO donor DETA NONOate. The obtained nanocapsules were found to be degraded in the acidic environment of tumor cell lysosomes which triggered the release of NO from DETA NONOate. More importantly, photoacoustic cavitation could stimulate ROS production and the formation water molecules in tumor cells under a pulsed laser. ROS could react with NO in situ to produce more toxic RNS, which could effectively inhibit the growth of tumor cells (Figure 8C) [103].

3.3. Glucose-Triggered NO Nanomedicines

It is well known that tumors require more nutrients and energy than normal tissues to maintain growth, leading to the higher concentration of glucose in the TME than in normal tissue [177,179]. Taking advantage of this, Fan et al. designed a novel glucose-responsive NO release nanosystem by loading glucose oxidase (GOx) and L-Arg in hollow silicone nanoparticles (HMON). The GOx in the nanosystem could consume glucose in the TME and inhibit tumor cell growth by glucose consumption (starvation therapy). On the other hand, H2O2 produced by glucose consumption could oxidize NO donor L-Arg into NO, which subsequently reacted with NO to generate more toxic peroxynitrite molecules to kill tumor cells. Using U87MG cells as target cells, the researchers found that with the increase in glucose concentration in the cell medium, the release of NO from the cells increased as well. In addition, the activity of tumor cells was significantly reduced when the nanosystem was used in a high concentration glucose medium. The results of the in vitro experiment proved that endogenous glucose in tumor cells could effectively stimulate the release of NO. Meanwhile, GOx-triggered starvation therapy could further synergistically kill tumor cells. In the U87 tumor-bearing mice, it was observed that the mice treated with the nanosystem had significantly smaller tumors than the mice treated with the other groups. This further confirmed the remarkable therapeutic effect of starvation therapy combined with the release of NO under the stimulation of glucose in tumors (Figure 9A) [104]. Recently, Su et al. innovatively designed a multimodal nanosystem for collaborative tumor treatment, which also used glucose-consuming starvation therapy to fight tumors. They used black phosphorus nanosheets (BPNs) with high photothermal conversion efficiency as carriers which loaded with NO donor L-Arg and GOx to obtain BPNs–Arg–GOx. BPNs–Arg–GOx was further modified with manganese dioxide nanosheets (MN) to obtain a nanosystem (BAGN). When the nanosystem entered the solid tumor site, it reacted with glucose to generate H2O2 which could promote the oxidation response of NO donor L-Arg to release NO. The NO produced by the nanosystem degrades the extracellular matrix in solid tumors by up-regulating the expression of MMP, which facilitated the accumulation of the nanosystem in the tumor site. In addition, the nanosystem could decompose H2O2 to produce oxygen to alleviate the hypoxia of solid tumors. Manganese ion and oxygen also provided good T1-weighted magnetic resonance and ultrasound imaging to monitor the therapeutic process. Therefore, the nanosystem demonstrated a powerful multimodal combinational tumor therapy based on NO nanomedicine (Figure 9B) [105].
The ability of starvation therapy to induce hypoglycemia has also attracted the attention of researchers. Several studies have found that a low rate of NO in diabetic patients may be a contributing factors leading to systemic vascular lesions in diabetic nephropathy. Therefore, NO shows important application prospects as a therapeutic gas in the treatment of diabetic nephropathy. For example, Yang et al. designed an NO nanoprodrug by wrapping the NO donor L-Arg with magnetic nanoparticles as a shell and loading GOx on the surface. Under high glucose levels in blood, GOx on the surface of the nanosystem could react with glucose to reduce the level of glucose and generate H2O2. Under the action of the magnetic field, the NO donor could be oxidized by H2O2 to release NO gas, and then realize the therapeutic effect of NO gas. Finally, this strategy may be a potential therapeutic method for lowering blood glucose and alleviating the progression of diabetic nephropathy [180].

4. Conclusions and Outlook

As a burgeoning anti-tumor treatment method, gas therapy has attracted increasing attention—especially that of NO gas, owing to its unique disease treatment application in anti-tumor applications, which has been widely studied by researchers. To address the side effect of current common anti-tumor therapy, as well as the short half-life and poor stability of NO gas, it is of great significance to construct stimuli-responsive NO releasing nanomedicines. Based on the characteristics of NO donors, this paper summarizes the current progress of research into stimuli-responsive nanomedicines for the controlled release of NO, including via exogenous stimulation such as by light, ultrasound, X-ray and endogenous stimulation such as by pH, glucose and GSH. This paper also discussed the application of NO in the treatment of various diseases, especially in tumor treatment, and discussed the anti-tumor mechanism of NO-based nanodrugs in anti-tumor therapy and their combined therapy with chemotherapy, PDT, PTT, etc. However, the NO-based nanodrugs still have a long way to go before they can be clinically implemented.
Firstly, the continuous investigation of NO-based nanodrugs is a prerequisite for advancing the development of this field. Only by exploring stable and intelligent NO-based prodrugs can NO be released when stimulated, which ensured the on-demand release of NO at specifical locations and in specific required doses, avoiding the damage to normal tissues. However, it is difficult for NO donors to achieve complete stability without any leakage in the physiological environment while achieving complete release with the stimulation. Therefore, how to control the leakage dose of NO in a safe range becomes particularly important. This requires researchers to consider the tolerance of normal human cells to NO gas. Only in this way, the early release of a small amount of NO gas from NO-based nanomaterials may be acceptable as long as the upper limit is not exceeded. Secondly, it remains to ensure that the stimulus-responsive NO nanosystems can effectively target the desired sites, allowing NO to be released at a specific location without affecting normal tissues and cells. The biological barriers include rapid clearance in blood circulation, penetration into tumor sites, high interstitial pressure in tumor sites, tumor membrane barrier and lysosome/endosome barrier. These biological barriers impair the delivery efficiency of most nanomedicines, which limits their clinical transitions. Therefore, it is essential to design more effective delivery systems to improve the therapeutic efficiency. Although there have been various studies on the targeting of NO-based nanosystems, the versatility of these studies remains to be explored in the hope of finding a universal NO releasing strategy with a targeting effect on the vast majority of tumors. Third, actuating the evaluation of the biocompatibility and safety of the NO release system is the first prerequisite for the clinical application. For both endogenous and exogenous stimuli, we should ensure their effective controllability and safety to ensure the accurate release of NO. Therefore, there is also a need to develop multimodal imaging functions to achieve the accurate real-time monitoring of the position of the NO nanosystem and the accurate evaluation of the NO release. Given that immune therapy has shown its excellent anti-tumor effect and NO could increase the tumor immune response, investigators are reminded to further explore the therapeutic effect of NO therapy combined with immunotherapy. The combination of NO therapy with immunotherapy may prevent tumor recurrence and metastasis, which is also a very important research point. To sum up, a new tumor treatment strategy, NO gas therapy, needs us to constantly explore, which will bring hope for more patients, especially cancer patients, and a brighter prospects in the medical field.

Author Contributions

Writing—original draft preparation, Y.Z. and X.O.; writing—review and editing, Y.P. and S.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by Ligong Lu with the National Key Research and Development Program of China (2017YFA0205200) and Shaojun Peng with the Natural Science Foundation of China (81903165 and 82172073).

Acknowledgments

We acknowledged Ligong Lu for the funding support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gu, Z.; Dong, Y.; Xu, S.; Wang, L.; Liu, Z. Molecularly Imprinted Polymer-Based Smart Prodrug Delivery System for Specific Targeting, Prolonged Retention, and Tumor Microenvironment-Triggered Release. Angew. Chem. Int. Ed. Engl. 2020, 60, 2663–2667. [Google Scholar] [CrossRef]
  2. Zhang, Y.; Dong, Y.; Fu, H.; Huang, H.; Wu, Z.; Zhao, M.; Yang, X.; Guo, Q.; Duan, Y.; Sun, Y. Multifunctional tumor-targeted PLGA nanoparticles delivering Pt(IV)/siBIRC5 for US/MRI imaging and overcoming ovarian cancer resistance. Biomaterials 2020, 269, 120478. [Google Scholar] [CrossRef]
  3. Sun, D.; Zhou, S.; Gao, W. What Went Wrong with Anticancer Nanomedicine Design and How to Make It Right. ACS Nano 2020, 14, 12281–12290. [Google Scholar] [CrossRef]
  4. Zhang, J.; Zhao, B.; Chen, S.; Wang, Y.; Zhang, Y.; Wang, Y.; Wei, D.; Zhang, L.; Rong, G.; Weng, Y.; et al. Near-Infrared Light Irradiation Induced Mild Hyperthermia Enhances Glutathione Depletion and DNA Interstrand Cross-Link Formation for Efficient Chemotherapy. ACS Nano 2020, 14, 14831–14845. [Google Scholar] [CrossRef] [PubMed]
  5. Wang, Y.; Yang, T.; He, Q. Strategies for engineering advanced nanomedicines for gas therapy of cancer. Nat. Sci. Rev. 2020, 7, 1485–1512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Hu, R.; Chen, Z.; Dai, C.; Guo, X.; Feng, W.; Liu, Z.; Lin, H.; Chen, Y.; Wu, R. Engineering two-dimensional silicene composite nanosheets for dual-sensitized and photonic hyperthermia-augmented cancer radiotherapy. Biomaterials 2020, 269, 120455. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, J.; Gao, Y.; Liu, P.; Xu, S.; Luo, X. Core-Shell Multifunctional Nanomaterial-Based All-in-One Nanoplatform for Simultaneous Multilayer Imaging of Dual Types of Tumor Biomarkers and Photothermal Therapy. Anal. Chem. 2020, 92, 15169–15178. [Google Scholar] [CrossRef] [PubMed]
  8. Chen, X.; Zhang, X.; Zhang, L.; Zhao, G.; Xu, S.; Li, L.; Su, Z.; Liu, R.; Wang, C. An EPR-independent therapeutic strategy: Cancer cell-mediated dual-drug delivery depot for diagnostics and prevention of hepatocellular carcinoma metastasis. Biomaterials 2020, 268, 120541. [Google Scholar] [CrossRef]
  9. Lin, H.; Chen, Y.; Shi, J. Nanoparticle-triggered in situ catalytic chemical reactions for tumour-specific therapy. Chem. Soc. Rev. 2018, 47, 1938–1958. [Google Scholar] [CrossRef]
  10. Fu, J.; Li, T.; Yang, Y.; Jiang, L.; Wang, W.; Fu, L.; Zhu, Y.; Hao, Y. Activatable nanomedicine for overcoming hypoxia-induced resistance to chemotherapy and inhibiting tumor growth by inducing collaborative apoptosis and ferroptosis in solid tumors. Biomaterials 2020, 268, 120537. [Google Scholar] [CrossRef]
  11. Guo, X.; Liu, F.; Deng, J.; Dai, P.; Qin, Y.; Li, Z.; Wang, B.; Fan, A.; Wang, Z.; Zhao, Y. Electron-Accepting Micelles Deplete Reduced Nicotinamide Adenine Dinucleotide Phosphate and Impair Two Antioxidant Cascades for Ferroptosis-Induced Tumor Eradication. ACS Nano 2020, 14, 14715–14730. [Google Scholar] [CrossRef]
  12. Zhou, S.; Zhen, Z.; Paschall, A.V.; Xue, L.; Yang, X.; Bebin Blackwell, A.-G.; Cao, Z.; Zhang, W.; Wang, M.; Teng, Y.; et al. FAP-Targeted Photodynamic Therapy Mediated by Ferritin Nanoparticles Elicits an Immune Response against Cancer Cells and Cancer Associated Fibroblasts. Adv. Funct. Mater. 2020, 31, 2007017. [Google Scholar] [CrossRef]
  13. Jiang, M.; Mu, J.; Jacobson, O.; Wang, Z.; He, L.; Zhang, F.; Yang, W.; Lin, Q.; Zhou, Z.; Ma, Y.; et al. Reactive Oxygen Species Activatable Heterodimeric Prodrug as Tumor-Selective Nanotheranostics. ACS Nano 2020, 14, 16875–16886. [Google Scholar] [CrossRef]
  14. Li, Y.; Xie, J.; Um, W.; You, D.G.; Kwon, S.; Zhang, L.; Zhu, J.; Park, J.H. Sono/Photodynamic Nanomedicine-Elicited Cancer Immunotherapy. Adv. Funct. Mater. 2020, 31, 2008061. [Google Scholar] [CrossRef]
  15. Zhuang, Q.; Xu, J.; Deng, D.; Chao, T.; Li, J.; Zhang, R.; Peng, R.; Liu, Z. Bacteria-derived membrane vesicles to advance targeted photothermal tumor ablation. Biomaterials 2020, 268, 120550. [Google Scholar] [CrossRef]
  16. Głowacka, U.; Brzozowski, T.; Magierowski, M. Synergisms, Discrepancies and Interactions between Hydrogen Sulfide and Carbon Monoxide in the Gastrointestinal and Digestive System Physiology, Pathophysiology and Pharmacology. Biomolecules 2020, 10, 445. [Google Scholar] [CrossRef] [Green Version]
  17. Wu, D.; Duan, X.H.; Guan, Q.Q.; Liu, J.; Yang, X.; Zhang, F.; Huang, P.; Shen, J.; Shuai, X.T.; Cao, Z. Mesoporous Polydopamine Carrying Manganese Carbonyl Responds to Tumor Microenvironment for Multimodal Imaging-Guided Cancer Therapy. Adv. Funct. Mater. 2019, 29, 11. [Google Scholar] [CrossRef]
  18. Chen, L.; Zhou, S.F.; Su, L.; Song, J. Gas-Mediated Cancer Bioimaging and Therapy. ACS Nano 2019, 13, 10887–10917. [Google Scholar] [CrossRef]
  19. Szabo, C. Gasotransmitters in cancer: From pathophysiology to experimental therapy. Nat. Rev. Drug Discov. 2016, 15, 185–203. [Google Scholar] [CrossRef] [Green Version]
  20. He, Q. Precision gas therapy using intelligent nanomedicine. Biomater. Sci. 2017, 5, 2226–2230. [Google Scholar] [CrossRef]
  21. Yoshikawa, T.; Mori, Y.; Feng, H.; Phan, K.Q.; Kishimura, A.; Kang, J.H.; Mori, T.; Katayama, Y. Rapid and continuous accumulation of nitric oxide-releasing liposomes in tumors to augment the enhanced permeability and retention (EPR) effect. Int. J. Pharm. 2019, 565, 481–487. [Google Scholar] [CrossRef]
  22. Yao, X.; Ma, S.; Peng, S.; Zhou, G.; Xie, R.; Jiang, Q.; Guo, S.; He, Q.; Yang, W. Zwitterionic Polymer Coating of Sulfur Dioxide-Releasing Nanosystem Augments Tumor Accumulation and Treatment Efficacy. Adv. Healthc. Mater. 2020, 9, e1901582. [Google Scholar] [CrossRef]
  23. Li, S.; Song, X.; Zhu, W.; Chen, Y.; Zhu, R.; Wang, L.; Chen, X.; Song, J.; Yang, H. Light-Switchable Yolk-Mesoporous Shell UCNPs@MgSiO3 for Nitric Oxide-Evoked Multidrug Resistance Reversal in Cancer Therapy. ACS Appl. Mater. Interfaces 2020, 12, 30066–30076. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, J.; Deng, M.; Shi, X.; Zhang, C.; Qu, X.; Hu, X.; Wang, W.; Kong, D.; Huang, P. Cascaded amplification of intracellular oxidative stress and reversion of multidrug resistance by nitric oxide prodrug based-supramolecular hydrogel for synergistic cancer chemotherapy. Bioact. Mater. 2021, 6, 3300–3313. [Google Scholar] [CrossRef]
  25. Zhao, P.; Jin, Z.; Chen, Q.; Yang, T.; Chen, D.; Meng, J.; Lu, X.; Gu, Z.; He, Q. Local generation of hydrogen for enhanced photothermal therapy. Nat. Commun. 2018, 9, 4241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Peng, S.; Wang, H.; Xin, Y.; Zhao, W.; Zhan, M.; Li, J.; Cai, R.; Lu, L. Second near-infrared photoactivatable hydrogen selenide nanogenerators for metastasis-inhibited cancer therapy. Nano Today 2021, 40, 101240. [Google Scholar] [CrossRef]
  27. Maeda, H.; Wu, J.; Sawa, T.; Matsumura, Y.; Hori, K. Tumor vascular permeability and the EPR effect in macromolecular therapeutics: A review. J. Control. Release 2000, 65, 271–284. [Google Scholar] [CrossRef]
  28. Jain, R.K.; Stylianopoulos, T. Delivering nanomedicine to solid tumors. Nat. Rev. Clin. Oncol. 2010, 7, 653–664. [Google Scholar] [CrossRef] [Green Version]
  29. Torchilin, V. Tumor delivery of macromolecular drugs based on the EPR effect. Adv. Drug Deliv. Rev. 2011, 63, 131–135. [Google Scholar] [CrossRef] [PubMed]
  30. Kim, J.; Yung, B.C.; Kim, W.J.; Chen, X. Combination of nitric oxide and drug delivery systems: Tools for overcoming drug resistance in chemotherapy. J. Control. Release 2017, 263, 223–230. [Google Scholar] [CrossRef]
  31. Folkes, L.K.; O’Neill, P. Modification of DNA damage mechanisms by nitric oxide during ionizing radiation. Free Radic. Biol. Med. 2013, 58, 14–25. [Google Scholar] [CrossRef] [PubMed]
  32. Qian, Y.; Matson, J.B. Gasotransmitter delivery via self-assembling peptides: Treating diseases with natural signaling gases. Adv. Drug Deliv. Rev. 2017, 110–111, 137–156. [Google Scholar] [CrossRef]
  33. Yin, H.; Guan, X.; Lin, H.; Pu, Y.; Fang, Y.; Yue, W.; Zhou, B.; Wang, Q.; Chen, Y.; Xu, H. Nanomedicine-Enabled Photonic Thermogaseous Cancer Therapy. Adv. Sci. 2020, 7, 1901954. [Google Scholar] [CrossRef]
  34. Yang, T.; Zelikin, A.N.; Chandrawati, R. Progress and Promise of Nitric Oxide-Releasing Platforms. Adv. Sci. 2018, 5, 1701043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Quinn, J.F.; Whittaker, M.R.; Davis, T.P. Delivering nitric oxide with nanoparticles. J. Control. Release 2015, 205, 190–205. [Google Scholar] [CrossRef]
  36. SoRelle, R. Nobel prize awarded to scientists for nitric oxide discoveries. Circulation 1998, 98, 2365–2366. [Google Scholar] [CrossRef] [Green Version]
  37. Schulz, R.; Rassaf, T.; Massion, P.B.; Kelm, M.; Balligand, J.L. Recent advances in the understanding of the role of nitric oxide in cardiovascular homeostasis. Pharm. Ther. 2005, 108, 225–256. [Google Scholar] [CrossRef] [PubMed]
  38. Feng, T.; Wan, J.; Li, P.; Ran, H.; Chen, H.; Wang, Z.; Zhang, L. A novel NIR-controlled NO release of sodium nitroprusside-doped Prussian blue nanoparticle for synergistic tumor treatment. Biomaterials 2019, 214, 119213. [Google Scholar] [CrossRef]
  39. Carpenter, A.W.; Schoenfisch, M.H. Nitric oxide release: Part II. Therapeutic applications. Chem. Soc. Rev. 2012, 41, 3742–3752. [Google Scholar] [CrossRef]
  40. Wang, P.G.; Xian, M.; Tang, X.; Wu, X.; Wen, Z.; Cai, T.; Janczuk, A.J. Nitric oxide donors: Chemical activities and biological applications. Chem. Rev. 2002, 102, 1091–1134. [Google Scholar] [CrossRef]
  41. Somasundaram, V.; Basudhar, D.; Bharadwaj, G.; No, J.H.; Ridnour, L.A.; Cheng, R.Y.S.; Fujita, M.; Thomas, D.D.; Anderson, S.K.; McVicar, D.W.; et al. Molecular Mechanisms of Nitric Oxide in Cancer Progression, Signal Transduction, and Metabolism. Antioxid. Redox Signal. 2019, 30, 1124–1143. [Google Scholar] [CrossRef]
  42. Holotiuk, V.V.; Kryzhanivska, A.Y.; Churpiy, I.K.; Tataryn, B.B.; Ivasiutyn, D.Y. Role of nitric oxide in pathogenesis of tumor growth and its possible application in cancer treatment. Exp. Oncol. 2019, 41, 210–215. [Google Scholar] [CrossRef] [PubMed]
  43. Li, B.; Ji, P.; Peng, S.Y.; Pan, P.; Zheng, D.W.; Li, C.X.; Sun, Y.X.; Zhang, X.Z. Nitric Oxide Release Device for Remote-Controlled Cancer Therapy by Wireless Charging. Adv. Mater. 2020, 32, e2000376. [Google Scholar] [CrossRef]
  44. Singh, N.; Patel, K.; Sahoo, S.K.; Kumar, R. Human nitric oxide biomarker as potential NO donor in conjunction with superparamagnetic iron oxide @ gold core shell nanoparticles for cancer therapeutics. Colloids Surf. B Biointerfaces 2018, 163, 246–256. [Google Scholar] [CrossRef] [PubMed]
  45. Mocellin, S.; Bronte, V.; Nitti, D. Nitric oxide, a double edged sword in cancer biology: Searching for therapeutic opportunities. Med. Res. Rev. 2007, 27, 317–352. [Google Scholar] [CrossRef] [PubMed]
  46. Seki, T.; Fang, J.; Maeda, H. Enhanced delivery of macromolecular antitumor drugs to tumors by nitroglycerin application. Cancer Sci. 2009, 100, 2426–2430. [Google Scholar] [CrossRef]
  47. Farah, C.; Michel, L.Y.M.; Balligand, J.L. Nitric oxide signalling in cardiovascular health and disease. Nat. Rev. Cardiol. 2018, 15, 292–316. [Google Scholar] [CrossRef] [PubMed]
  48. Jin, G.; Gao, Z.; Liu, Y.; Zhao, J.; Ou, H.; Xu, F.; Ding, D. Polymeric Nitric Oxide Delivery Nanoplatforms for Treating Cancer, Cardiovascular Diseases, and Infection. Adv. Healthc. Mater. 2021, 10, e2001550. [Google Scholar] [CrossRef]
  49. Chung, M.F.; Liu, H.Y.; Lin, K.J.; Chia, W.T.; Sung, H.W. A pH-Responsive Carrier System that Generates NO Bubbles to Trigger Drug Release and Reverse P-Glycoprotein-Mediated Multidrug Resistance. Angew. Chem. Int. Ed. Engl. 2015, 54, 9890–9893. [Google Scholar] [CrossRef]
  50. Scicinski, J.; Oronsky, B.; Ning, S.; Knox, S.; Peehl, D.; Kim, M.M.; Langecker, P.; Fanger, G. NO to cancer: The complex and multifaceted role of nitric oxide and the epigenetic nitric oxide donor, RRx-001. Redox Biol. 2015, 6, 1–8. [Google Scholar] [CrossRef] [Green Version]
  51. De Ridder, M.; Verellen, D.; Verovski, V.; Storme, G. Hypoxic tumor cell radiosensitization through nitric oxide. Nitric Oxide 2008, 19, 164–169. [Google Scholar] [CrossRef]
  52. Sun, F.; Wang, Y.; Luo, X.; Ma, Z.; Xu, Y.; Zhang, X.; Lv, T.; Zhang, Y.; Wang, M.; Huang, Z.; et al. Anti-CD24 Antibody-Nitric Oxide Conjugate Selectively and Potently Suppresses Hepatic Carcinoma. Cancer Res. 2019, 79, 3395–3405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Wang, H.; Liu, Y.; Wang, Z.; Yang, M.; Gu, Y. 808 nm-light-excited upconversion nanoprobe based on LRET for the ratiometric detection of nitric oxide in living cancer cells. Nanoscale 2018, 10, 10641–10649. [Google Scholar] [CrossRef]
  54. Xu, W.; Liu, L.Z.; Loizidou, M.; Ahmed, M.; Charles, I.G. The role of nitric oxide in cancer. Cell Res. 2002, 12, 311–320. [Google Scholar] [CrossRef] [Green Version]
  55. Ridnour, L.A.; Thomas, D.D.; Donzelli, S.; Espey, M.G.; Roberts, D.D.; Wink, D.A.; Isenberg, J.S. The biphasic nature of nitric oxide responses in tumor biology. Antioxid. Redox Signal. 2006, 8, 1329–1337. [Google Scholar] [CrossRef] [PubMed]
  56. Fan, W.; Yung, B.C.; Chen, X. Stimuli-Responsive NO Release for On-Demand Gas-Sensitized Synergistic Cancer Therapy. Angew. Chem. Int. Ed. Engl. 2018, 57, 8383–8394. [Google Scholar] [CrossRef]
  57. Xu, J.; Zeng, F.; Wu, H.; Wu, S. A mitochondrial-targeting and NO-based anticancer nanosystem with enhanced photo-controllability and low dark-toxicity. J. Mater. Chem. B 2015, 3, 4904–4912. [Google Scholar] [CrossRef] [PubMed]
  58. Keshet, R.; Erez, A. Arginine and the metabolic regulation of nitric oxide synthesis in cancer. Dis. Model. Mech. 2018, 11, dmm033332. [Google Scholar] [CrossRef] [Green Version]
  59. Xu, Y.; Ren, H.; Liu, J.; Wang, Y.; MEng, Z.; He, Z.; Miao, W.; Chen, G.; Li, X. A switchable NO-releasing nanomedicine for enhanced cancer therapy and inhibition of metastasis. Nanoscale 2019, 11, 5474–5488. [Google Scholar] [CrossRef]
  60. Raju, G.S.R.; Dariya, B.; Mungamuri, S.K.; Chalikonda, G.; Kang, S.M.; Khan, I.N.; Sushma, P.S.; Nagaraju, G.P.; Pavitra, E.; Han, Y.K. Nanomaterials multifunctional behavior for enlightened cancer therapeutics. Semin. Cancer Biol. 2021, 69, 178–189. [Google Scholar] [CrossRef]
  61. Wang, Y.; Huang, X.; Tang, Y.; Zou, J.; Wang, P.; Zhang, Y.; Si, W.; Huang, W.; Dong, X. A light-induced nitric oxide controllable release nano-platform based on diketopyrrolopyrrole derivatives for pH-responsive photodynamic/photothermal synergistic cancer therapy. Chem. Sci. 2018, 9, 8103–8109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. FEng, L.; Liu, B.; Xie, R.; Wang, D.; Qian, C.; Zhou, W.; Liu, J.; Jana, D.; Yang, P.; Zhao, Y. An Ultrasmall SnFe2O4 Nanozyme with Endogenous Oxygen Generation and Glutathione Depletion for Synergistic Cancer Therapy. Adv. Funct. Mater. 2020, 31, 2006216. [Google Scholar] [CrossRef]
  63. Quader, S.; Liu, X.; Toh, K.; Su, Y.L.; Maity, A.R.; Tao, A.; Paraiso, W.K.D.; Mochida, Y.; Kinoh, H.; Cabral, H.; et al. Supramolecularly enabled pH- triggered drug action at tumor microenvironment potentiates nanomedicine efficacy against glioblastoma. Biomaterials 2021, 267, 120463. [Google Scholar] [CrossRef]
  64. Yu, W.; Liu, T.; Zhang, M.; Wang, Z.; Ye, J.; Li, C.X.; Liu, W.; Li, R.; FEng, J.; Zhang, X.Z. O2 Economizer for Inhibiting Cell Respiration to Combat the Hypoxia Obstacle in Tumor Treatments. ACS Nano 2019, 13, 1784–1794. [Google Scholar] [CrossRef]
  65. Sun, S.; Chen, Q.; Tang, Z.; Liu, C.; Li, Z.; Wu, A.; Lin, H. Tumor Microenvironment Stimuli-Responsive Fluorescence Imaging and Synergistic Cancer Therapy by Carbon-Dot-Cu2+ Nanoassemblies. Angew. Chem. Int. Ed. Engl. 2020, 59, 21041–21048. [Google Scholar] [CrossRef]
  66. Cheng, J.; He, K.; Shen, Z.; Zhang, G.; Yu, Y.; Hu, J. Nitric Oxide (NO)-Releasing Macromolecules: Rational Design and Biomedical Applications. Front. Chem. 2019, 7, 530. [Google Scholar] [CrossRef] [Green Version]
  67. Miller, M.R.; Megson, I.L. Recent developments in nitric oxide donor drugs. Br. J. Pharm. 2007, 151, 305–321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Fan, J.; He, Q.; Liu, Y.; Zhang, F.; Yang, X.; Wang, Z.; Lu, N.; Fan, W.; Lin, L.; Niu, G.; et al. Light-Responsive Biodegradable Nanomedicine Overcomes Multidrug Resistance via NO-Enhanced Chemosensitization. ACS Appl. Mater. Interfaces 2016, 8, 13804–13811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Xu, J.; Zeng, F.; Wu, H.; Hu, C.; Yu, C.; Wu, S. Preparation of a mitochondria-targeted and NO-releasing nanoplatform and its enhanced pro-apoptotic effect on cancer cells. Small 2014, 10, 3750–3760. [Google Scholar] [CrossRef] [PubMed]
  70. Xiang, H.J.; An, L.; Tang, W.W.; Yang, S.P.; Liu, J.G. Photo-controlled targeted intracellular delivery of both nitric oxide and singlet oxygen using a fluorescence-trackable ruthenium nitrosyl functional nanoplatform. Chem. Commun. 2015, 51, 2555–2558. [Google Scholar] [CrossRef]
  71. Guo, R.R.; Tian, Y.; Wang, Y.J.; Yang, W.L. Near-Infrared Laser-Triggered Nitric Oxide Nanogenerators for the Reversal of Multidrug Resistance in Cancer. Adv. Funct. Mater. 2017, 27, 1606398. [Google Scholar] [CrossRef]
  72. Huang, X.; Xu, F.; Hou, H.; Hou, J.; Wang, Y.; Zhou, S. Stimuli-responsive nitric oxide generator for light-triggered synergistic cancer photothermal/gas therapy. Nano Res. 2019, 12, 1361–1370. [Google Scholar] [CrossRef]
  73. Yu, Y.T.; Shi, S.W.; Wang, Y.; Zhang, Q.L.; Gao, S.H.; Yang, S.P.; Liu, J.G. A Ruthenium Nitrosyl-Functionalized Magnetic Nanoplatform with Near-Infrared Light-Controlled Nitric Oxide Delivery and Photothermal Effect for Enhanced Antitumor and Antibacterial Therapy. ACS Appl. Mater. Interfaces 2020, 12, 312–321. [Google Scholar] [CrossRef]
  74. Zhang, X.; Du, J.; Guo, Z.; Yu, J.; Gao, Q.; Yin, W.; Zhu, S.; Gu, Z.; Zhao, Y. Efficient Near Infrared Light Triggered Nitric Oxide Release Nanocomposites for Sensitizing Mild Photothermal Therapy. Adv. Sci. 2019, 6, 1801122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Li, J.; Jiang, R.; Wang, Q.; Li, X.; Hu, X.; Yuan, Y.; Lu, X.; Wang, W.; Huang, W.; Fan, Q. Semiconducting polymer nanotheranostics for NIR-II/Photoacoustic imaging-guided photothermal initiated nitric oxide/photothermal therapy. Biomaterials 2019, 217, 119304. [Google Scholar] [CrossRef]
  76. You, C.; Li, Y.; Dong, Y.; Ning, L.; Zhang, Y.; Yao, L.; Wang, F. Low-Temperature Trigger Nitric Oxide Nanogenerators for Enhanced Mild Photothermal Therapy. ACS Biomater. Sci. Eng. 2020, 6, 1535–1542. [Google Scholar] [CrossRef] [PubMed]
  77. Ding, Y.; Du, C.; Qian, J.; Dong, C.M. NIR-Responsive Polypeptide Nanocomposite Generates NO Gas, Mild Photothermia, and Chemotherapy to Reverse Multidrug-Resistant Cancer. Nano Lett. 2019, 19, 4362–4370. [Google Scholar] [CrossRef] [PubMed]
  78. Wu, W.; Yang, Y.; Liang, Z.; Song, X.; Huang, Y.; Qiu, L.; Qiu, X.; Yu, S.; Xue, W. Near infrared II laser controlled free radical releasing nanogenerator for synergistic nitric oxide and alkyl radical therapy of breast cancer. Nanoscale 2021, 13, 11169–11187. [Google Scholar] [CrossRef] [PubMed]
  79. Yang, Z.; Gao, D.; Guo, X.; Jin, L.; ZhEng, J.; Wang, Y.; Chen, S.; ZhEng, X.; ZEng, L.; Guo, M.; et al. Fighting Immune Cold and Reprogramming Immunosuppressive Tumor Microenvironment with Red Blood Cell Membrane-Camouflaged Nanobullets. ACS Nano 2020, 14, 17442–17457. [Google Scholar] [CrossRef]
  80. Ding, Y.; Ma, Y.X.; Du, C.; Wang, C.W.; Chen, T.T.; Wang, Y.; Wang, J.; Yao, Y.; Dong, C.M. NO-releasing polypeptide nanocomposites reverse cancer multidrug resistance via triple therapies. Acta Biomater. 2021, 123, 335–345. [Google Scholar] [CrossRef] [PubMed]
  81. Chen, L.; He, Q.; Lei, M.; Xiong, L.; Shi, K.; Tan, L.; Jin, Z.; Wang, T.; Qian, Z. Facile Coordination-Precipitation Route to Insoluble Metal Roussin’s Black Salts for NIR-Responsive Release of NO for Anti-Metastasis. ACS Appl. Mater. Interfaces 2017, 9, 36473–36477. [Google Scholar] [CrossRef]
  82. Fan, J.; He, N.; He, Q.; Liu, Y.; Ma, Y.; Fu, X.; Liu, Y.; Huang, P.; Chen, X. A novel self-assembled sandwich nanomedicine for NIR-responsive release of NO. Nanoscale 2015, 7, 20055–20062. [Google Scholar] [CrossRef]
  83. Wan, S.S.; ZEng, J.Y.; Cheng, H.; Zhang, X.Z. ROS-induced NO generation for gas therapy and sensitizing photodynamic therapy of tumor. Biomaterials 2018, 185, 51–62. [Google Scholar] [CrossRef] [PubMed]
  84. Shi, X.; Zhan, Q.; Li, Y.; Zhou, L.; Wei, S. Multiple Functions Integrated inside a Single Molecule for Amplification of Photodynamic Therapy Activity. Mol. Pharm. 2020, 17, 190–201. [Google Scholar] [CrossRef] [PubMed]
  85. Wang, L.; Chang, Y.; FEng, Y.; Li, X.; ChEng, Y.; Jian, H.; Ma, X.; ZhEng, R.; Wu, X.; Xu, K.; et al. Nitric Oxide Stimulated Programmable Drug Release of Nanosystem for Multidrug Resistance Cancer Therapy. Nano Lett. 2019, 19, 6800–6811. [Google Scholar] [CrossRef] [PubMed]
  86. Sun, Z.; Wang, X.; Liu, J.; Wang, Z.; Wang, W.; Kong, D.; Leng, X. ICG/l-Arginine Encapsulated PLGA Nanoparticle-Thermosensitive Hydrogel Hybrid Delivery System for Cascade Cancer Photodynamic-NO Therapy with Promoted Collagen Depletion in Tumor Tissues. Mol. Pharm. 2021, 18, 928–939. [Google Scholar] [CrossRef]
  87. Xiang, H.J.; DEng, Q.; An, L.; Guo, M.; Yang, S.P.; Liu, J.G. Tumor cell specific and lysosome-targeted delivery of nitric oxide for enhanced photodynamic therapy triggered by 808 nm near-infrared light. Chem. Commun. 2016, 52, 148–151. [Google Scholar] [CrossRef]
  88. Jin, Z.; Wen, Y.; Hu, Y.; Chen, W.; ZhEng, X.; Guo, W.; Wang, T.; Qian, Z.; Su, B.L.; He, Q. MRI-guided and ultrasound-triggered release of NO by advanced nanomedicine. Nanoscale 2017, 9, 3637–3645. [Google Scholar] [CrossRef]
  89. FEng, Q.; Li, Y.; Yang, X.; Zhang, W.; Hao, Y.; Zhang, H.; Hou, L.; Zhang, Z. Hypoxia-specific therapeutic agents delivery nanotheranostics: A sequential strategy for ultrasound mediated on-demand tritherapies and imaging of cancer. J. Control. Release 2018, 275, 192–200. [Google Scholar] [CrossRef]
  90. Kang, Y.; Kim, J.; Park, J.; Lee, Y.M.; Saravanakumar, G.; Park, K.M.; Choi, W.; Kim, K.; Lee, E.; Kim, C.; et al. Tumor vasodilation by N-Heterocyclic carbene-based nitric oxide delivery triggered by high-intensity focused ultrasound and enhanced drug homing to tumor sites for anti-cancer therapy. Biomaterials 2019, 217, 119297. [Google Scholar] [CrossRef]
  91. An, J.; Hu, Y.G.; Li, C.; Hou, X.L.; ChEng, K.; Zhang, B.; Zhang, R.Y.; Li, D.Y.; Liu, S.J.; Liu, B.; et al. A pH/Ultrasound dual-response biomimetic nanoplatform for nitric oxide gas-sonodynamic combined therapy and repeated ultrasound for relieving hypoxia. Biomaterials 2020, 230, 119636. [Google Scholar] [CrossRef]
  92. Xu, Y.; Liu, J.; Liu, Z.; Chen, G.; Li, X.; Ren, H. Damaging Tumor Vessels with an Ultrasound-Triggered NO Release Nanosystem to Enhance Drug Accumulation and T Cells Infiltration. Int. J. Nanomed. 2021, 16, 2597–2613. [Google Scholar] [CrossRef]
  93. Zhang, K.; Xu, H.; Jia, X.; Chen, Y.; Ma, M.; Sun, L.; Chen, H. Ultrasound-Triggered Nitric Oxide Release Platform Based on Energy Transformation for Targeted Inhibition of Pancreatic Tumor. ACS Nano 2016, 10, 10816–10828. [Google Scholar] [CrossRef]
  94. Fan, W.; Bu, W.; Zhang, Z.; Shen, B.; Zhang, H.; He, Q.; Ni, D.; Cui, Z.; Zhao, K.; Bu, J.; et al. X-ray Radiation-Controlled NO-Release for On-Demand Depth-Independent Hypoxic Radiosensitization. Angew. Chem. Int. Ed. Engl. 2015, 54, 14026–14030. [Google Scholar] [CrossRef] [PubMed]
  95. Zhang, F.; Liu, S.; Zhang, N.; Kuang, Y.; Li, W.; Gai, S.; He, F.; Gulzar, A.; Yang, P. X-ray-triggered NO-released Bi-SNO nanoparticles: All-in-one nano-radiosensitizer with photothermal/gas therapy for enhanced radiotherapy. Nanoscale 2020, 12, 19293–19307. [Google Scholar] [CrossRef]
  96. Xue, Z.; Jiang, M.; Liu, H.; Zeng, S.; Hao, J. Low dose soft X-ray-controlled deep-tissue long-lasting NO release of persistent luminescence nanoplatform for gas-sensitized anticancer therapy. Biomaterials 2020, 263, 120384. [Google Scholar] [CrossRef] [PubMed]
  97. Zhang, J.; Song, H.; Ji, S.; Wang, X.; Huang, P.; Zhang, C.; Wang, W.; Kong, D. NO prodrug-conjugated, self-assembled, pH-responsive and galactose receptor targeted nanoparticles for co-delivery of nitric oxide and doxorubicin. Nanoscale 2018, 10, 4179–4188. [Google Scholar] [CrossRef] [PubMed]
  98. Deepagan, V.G.; Ko, H.; Kwon, S.; Rao, N.V.; Kim, S.K.; Um, W.; Lee, S.; Min, J.; Lee, J.; Choi, K.Y.; et al. Intracellularly Activatable Nanovasodilators to Enhance Passive Cancer Targeting Regime. Nano Lett. 2018, 18, 2637–2644. [Google Scholar] [CrossRef] [PubMed]
  99. Niu, X.; Cao, J.; Zhang, Y.; Gao, X.; ChEng, M.; Liu, Y.; Wang, W.; Yuan, Z. A glutathione responsive nitric oxide release system based on charge-reversal chitosan nanoparticles for enhancing synergistic effect against multidrug resistance tumor. Nanomedicine 2019, 20, 102015. [Google Scholar] [CrossRef]
  100. Deng, Y.; Jia, F.; Chen, X.; Jin, Q.; Ji, J. ATP Suppression by pH-Activated Mitochondria-Targeted Delivery of Nitric Oxide Nanoplatform for Drug Resistance Reversal and Metastasis Inhibition. Small 2020, 16, e2001747. [Google Scholar] [CrossRef]
  101. Song, Q.; Tan, S.; Zhuang, X.; Guo, Y.; Zhao, Y.; Wu, T.; Ye, Q.; Si, L.; Zhang, Z. Nitric oxide releasing d-α-tocopheryl polyethylene glycol succinate for enhancing antitumor activity of doxorubicin. Mol. Pharm. 2014, 11, 4118–4129. [Google Scholar] [CrossRef]
  102. Lee, H.J.; Park, D.J.; Choi, G.H.; Yang, D.N.; Heo, J.S.; Lee, S.C. pH-Responsive mineralized nanoparticles as stable nanocarriers for intracellular nitric oxide delivery. Colloids Surf. B Biointerfaces 2016, 146, 1–8. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, Z.; Zhan, M.; Li, W.; Chu, C.; Xing, D.; Lu, S.; Hu, X. Photoacoustic Cavitation-Ignited Reactive Oxygen Species to Amplify Peroxynitrite Burst by Photosensitization-Free Polymeric Nanocapsules. Angew. Chem. Int. Ed. Engl. 2021, 60, 4720–4731. [Google Scholar] [CrossRef]
  104. Fan, W.; Lu, N.; Huang, P.; Liu, Y.; Yang, Z.; Wang, S.; Yu, G.; Liu, Y.; Hu, J.; He, Q.; et al. Glucose-Responsive Sequential Generation of Hydrogen Peroxide and Nitric Oxide for Synergistic Cancer Starving-Like/Gas Therapy. Angew. Chem. Int. Ed. Engl. 2017, 56, 1229–1233. [Google Scholar] [CrossRef] [PubMed]
  105. Su, Y.; Zhang, X.; Lei, L.; Liu, B.; Wu, S.; Shen, J. Tumor Microenvironment-Activatable Cyclic Cascade Reaction to Reinforce Multimodal Combination Therapy by Destroying the Extracellular Matrix. ACS Appl. Mater. Interfaces 2021, 13, 12960–12971. [Google Scholar] [CrossRef]
  106. Hou, J.; Pan, Y.; Zhu, D.; Fan, Y.; FEng, G.; Wei, Y.; Wang, H.; Qin, K.; Zhao, T.; Yang, Q.; et al. Targeted delivery of nitric oxide via a ‘bump-and-hole’-based enzyme–prodrug pair. Nat. Chem. Biol. 2018, 15, 151–160. [Google Scholar] [CrossRef] [PubMed]
  107. Guo, M.; Xiang, H.J.; Wang, Y.; Zhang, Q.L.; An, L.; Yang, S.P.; Ma, Y.; Wang, Y.; Liu, J.G. Ruthenium nitrosyl functionalized graphene quantum dots as an efficient nanoplatform for NIR-light-controlled and mitochondria-targeted delivery of nitric oxide combined with photothermal therapy. Chem. Commun. 2017, 53, 3253–3256. [Google Scholar] [CrossRef]
  108. Su, C.H.; Li, W.P.; Tsao, L.C.; Wang, L.C.; Hsu, Y.P.; Wang, W.J.; Liao, M.C.; Lee, C.L.; Yeh, C.S. Enhancing Microcirculation on Multitriggering Manner Facilitates Angiogenesis and Collagen Deposition on Wound Healing by Photoreleased NO from Hemin-Derivatized Colloids. ACS Nano 2019, 13, 4290–4301. [Google Scholar] [CrossRef]
  109. Tessaro, A.L.; Fraix, A.; Pedrozo da Silva, A.C.; Gazzano, E.; Riganti, C.; Sortino, S. “Three-Bullets” Loaded Mesoporous Silica Nanoparticles for Combined Photo/Chemotherapy. Nanomaterials 2019, 9, 823. [Google Scholar] [CrossRef] [Green Version]
  110. Yang, S.; Chen, C.; Qiu, Y.; Xu, C.; Yao, J. Paying attention to tumor blood vessels: Cancer phototherapy assisted with nano delivery strategies. Biomaterials 2020, 268, 120562. [Google Scholar] [CrossRef] [PubMed]
  111. Kumar, G.S.; Lin, Q. Light-Triggered Click Chemistry. Chem. Rev. 2021, 121, 6991–7031. [Google Scholar] [CrossRef] [PubMed]
  112. Cho, H.J.; Park, S.J.; Jung, W.H.; Cho, Y.; Ahn, D.J.; Lee, Y.S.; Kim, S. Injectable Single-Component Peptide Depot: Autonomously Rechargeable Tumor Photosensitization for Repeated Photodynamic Therapy. ACS Nano 2020, 14, 15793–15805. [Google Scholar] [CrossRef]
  113. Karaki, F.; Kabasawa, Y.; Yanagimoto, T.; Umeda, N.; Firman Urano, Y.; Nagano, T.; Otani, Y.; Ohwada, T. Visible-light-triggered release of nitric oxide from N-pyramidal nitrosamines. Chemistry 2012, 18, 1127–1141. [Google Scholar] [CrossRef]
  114. Fraix, A.; Marino, N.; Sortino, S. Phototherapeutic Release of Nitric Oxide with Engineered Nanoconstructs. Top. Curr. Chem. 2016, 370, 225–257. [Google Scholar] [CrossRef] [PubMed]
  115. Chegaev, K.; Fraix, A.; Gazzano, E.; Abd-Ellatef, G.E.; Blangetti, M.; Rolando, B.; Conoci, S.; Riganti, C.; Fruttero, R.; Gasco, A.; et al. Light-Regulated NO Release as a Novel Strategy To Overcome Doxorubicin Multidrug Resistance. ACS Med. Chem. Lett. 2017, 8, 361–365. [Google Scholar] [CrossRef]
  116. Zhang, Y.; Tang, K.; Chen, B.; Zhou, S.; Li, N.; Liu, C.; Yang, J.; Lin, R.; Zhang, T.; He, W. A polyethylenimine-based diazeniumdiolate nitric oxide donor accelerates wound healing. Biomater. Sci. 2019, 7, 1607–1616. [Google Scholar] [CrossRef] [Green Version]
  117. Duan, Y.; Wang, Y.; Li, X.; Zhang, G.; Zhang, G.; Hu, J. Light-triggered nitric oxide (NO) release from photoresponsive polymersomes for corneal wound healing. Chem. Sci. 2020, 11, 186–194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Zhang, X.; Tian, G.; Yin, W.Y.; Wang, L.M.; ZhEng, X.P.; Yan, L.; Li, J.X.; Su, H.R.; Chen, C.Y.; Gu, Z.J.; et al. Controllable Generation of Nitric Oxide by Near-Infrared-Sensitized Upconversion Nanoparticles for Tumor Therapy. Adv. Funct. Mater. 2015, 25, 3049–3056. [Google Scholar] [CrossRef]
  119. Pu, K.; Huang, J.; Jiang, Y.; Li, J.; Huang, J. Molecular Chemiluminescent Probes with a Record Long Near-infrared Turn-on Wavelength for In vivo Imaging. Angew. Chem. Int. Ed. Engl. 2020, 60, 3999–4003. [Google Scholar] [CrossRef]
  120. Lee, H.; Kim, J.; Lee, J.; Park, H.; Park, Y.; Jung, S.; Lim, J.; Choi, H.C.; Kim, W.J. In vivo self-degradable graphene nanomedicine operated by DNAzyme and photo-switch for controlled anticancer therapy. Biomaterials 2020, 263, 120402. [Google Scholar] [CrossRef]
  121. Liu, J.S.; PEng, S.J.; Li, G.F.; Zhao, Y.X.; MEng, X.Y.; Yu, X.R.; Li, Z.H.; Chen, J.M. Polydopamine Nanoparticles for Deep Brain Ablation via Near-Infrared Irradiation. ACS Biomater. Sci. Eng. 2020, 6, 664–672. [Google Scholar] [CrossRef] [PubMed]
  122. Zhang, C.; Li, Q.; Zhao, Y.; Liu, H.; Song, S.; Zhao, Y.; Lin, Q.; Chang, Y. Near-infrared light-mediated and nitric oxide-supplied nanospheres for enhanced synergistic thermo-chemotherapy. J. Mater. Chem. B 2019, 7, 548–555. [Google Scholar] [CrossRef]
  123. Li, X.M.; Zhang, Y.H.; Ma, Z.Q.; He, C.J.; Wu, Y.L.; An, Q. Designing cancer nanodrugs that are highly loaded, pH-responsive, photothermal, and possess a favored morphology: A hierarchical assembly of DOX and layer-by-layer modified rGO. Chin. Chem. Lett. 2019, 30, 489–493. [Google Scholar] [CrossRef]
  124. Zhang, A.M.; Hai, L.; Wang, T.Z.; ChEng, H.; Li, M.; He, X.X.; Wang, K.M. NIR-triggered drug delivery system based on phospholipid coated ordered mesoporous carbon for synergistic chemo-photothermal therapy of cancer cells. Chin. Chem. Lett. 2020, 31, 3158–3162. [Google Scholar] [CrossRef]
  125. Choi, H.W.; Kim, J.; Kim, J.; Kim, Y.; Song, H.B.; Kim, J.H.; Kim, K.; Kim, W.J. Light-Induced Acid Generation on a Gatekeeper for Smart Nitric Oxide Delivery. ACS Nano 2016, 10, 4199–4208. [Google Scholar] [CrossRef]
  126. Sung, Y.C.; Jin, P.R.; Chu, L.A.; Hsu, F.F.; Wang, M.R.; Chang, C.C.; Chiou, S.J.; Qiu, J.T.; Gao, D.Y.; Lin, C.C.; et al. Delivery of nitric oxide with a nanocarrier promotes tumour vessel normalization and potentiates anti-cancer therapies. Nat. Nanotechnol. 2019, 14, 1160–1169. [Google Scholar] [CrossRef]
  127. Luo, D.; Wang, X.; Walker, E.; Wang, J.; Springer, S.; Lou, J.; Ramamurthy, G.; Burda, C.; Basilion, J.P. Nanoparticles Yield Increased Drug Uptake and Therapeutic Efficacy upon Sequential Near-Infrared Irradiation. ACS Nano 2020, 14, 15193–15203. [Google Scholar] [CrossRef]
  128. Han, R.; Xiao, Y.; Yang, Q.; Pan, M.; Hao, Y.; He, X.; PEng, J.; Qian, Z. Ag2S nanoparticle-mediated multiple ablations reinvigorates the immune response for enhanced cancer photo-immunotherapy. Biomaterials 2021, 264, 120451. [Google Scholar] [CrossRef] [PubMed]
  129. Wei, G.; Yang, G.; Wei, B.; Wang, Y.; Zhou, S. Near-infrared light switching nitric oxide nanoemitter for triple-combination therapy of multidrug resistant cancer. Acta Biomater. 2019, 100, 365–377. [Google Scholar] [CrossRef]
  130. Zhang, H.; Tian, X.T.; Shang, Y.; Li, Y.H.; Yin, X.B. Theranostic Mn-Porphyrin Metal-Organic Frameworks for Magnetic Resonance Imaging-Guided Nitric Oxide and Photothermal Synergistic Therapy. ACS Appl. Mater. Interfaces 2018, 10, 28390–28398. [Google Scholar] [CrossRef]
  131. Yang, Y.; Zhu, W.; Dong, Z.; Chao, Y.; Xu, L.; Chen, M.; Liu, Z. 1D Coordination Polymer Nanofibers for Low-Temperature Photothermal Therapy. Adv. Mater. 2017, 29, 3588. [Google Scholar] [CrossRef]
  132. Zhang, K.; MEng, X.D.; Cao, Y.; Yang, Z.; Dong, H.F.; Zhang, Y.D.; Lu, H.T.; Shi, Z.J.; Zhang, X.J. Metal-Organic Framework Nanoshuttle for Synergistic Photodynamic and Low-Temperature Photothermal Therapy. Adv. Funct. Mater. 2018, 28, 10. [Google Scholar] [CrossRef]
  133. Zhao, J.; Zhong, D.; Zhou, S. NIR-I-to-NIR-II fluorescent nanomaterials for biomedical imaging and cancer therapy. J. Mater. Chem. B 2018, 6, 349–365. [Google Scholar] [CrossRef]
  134. Wang, J.; Wu, C.C.; Qin, X.R.; Huang, Y.Y.; Zhang, J.N.; Chen, T.T.; Wang, Y.; Ding, Y.; Yao, Y. NIR-II light triggered nitric oxide release nanoplatform combined chemo-photothermal therapy for overcoming multidrug resistant cancer. J. Mater. Chem. B 2021, 9, 1698–1706. [Google Scholar] [CrossRef]
  135. Sun, H.T.; Zhang, Q.; Li, J.C.; PEng, S.J.; Wang, X.L.; Cai, R. Near-infrared photoactivated nanomedicines for photothermal synergistic cancer therapy. Nano Today 2021, 37, 29. [Google Scholar] [CrossRef]
  136. PEng, S.J.; Ouyang, B.S.; Men, Y.Z.; Du, Y.; Cao, Y.B.; Xie, R.H.; Pang, Z.Q.; Shen, S.; Yang, W.L. Biodegradable zwitterionic polymer membrane coating endowing nanoparticles with ultra-long circulation and enhanced tumor photothermal therapy. Biomaterials 2020, 231, 13. [Google Scholar] [CrossRef]
  137. Wang, C.; Dai, C.; Hu, Z.; Li, H.; Yu, L.; Lin, H.; Bai, J.; Chen, Y. Photonic cancer nanomedicine using the near infrared-II biowindow enabled by biocompatible titanium nitride nanoplatforms. Nanoscale Horiz 2019, 4, 415–425. [Google Scholar] [CrossRef] [PubMed]
  138. She, D.J.; PEng, S.J.; Liu, L.; Huang, H.H.; ZhEng, Y.Y.; Lu, Y.P.; GEng, D.Y.; Yin, B. Biomimic FeS2 nanodrug with hypothermal photothermal effect by clinical approved NIR-II light for augmented chemodynamic therapy. Chem. Eng. J. 2020, 400, 13. [Google Scholar] [CrossRef]
  139. Lin, S.; Lin, H.; Yang, M.; Ge, M.; Chen, Y.; Zhu, Y. A two-dimensional MXene potentiates a therapeutic microneedle patch for photonic implantable medicine in the second NIR biowindow. Nanoscale 2020, 12, 10265–10276. [Google Scholar] [CrossRef]
  140. Han, X.; Jing, X.; Yang, D.; Lin, H.; Wang, Z.; Ran, H.; Li, P.; Chen, Y. Therapeutic mesopore construction on 2D Nb2C MXenes for targeted and enhanced chemo-photothermal cancer therapy in NIR-II biowindow. Theranostics 2018, 8, 4491–4508. [Google Scholar] [CrossRef] [PubMed]
  141. Lin, H.; Gao, S.; Dai, C.; Chen, Y.; Shi, J. A Two-Dimensional Biodegradable Niobium Carbide (MXene) for Photothermal Tumor Eradication in NIR-I and NIR-II Biowindows. J. Am. Chem. Soc. 2017, 139, 16235–16247. [Google Scholar] [CrossRef]
  142. Wen, S.F.; Zhang, K.; Li, Y.; Fan, J.Q.; Chen, Z.M.; Zhang, J.P.; Wang, H.; Wang, L. A self-assembling peptide targeting VEGF receptors to inhibit angiogenesis. Chin. Chem. Lett. 2020, 31, 3153–3157. [Google Scholar] [CrossRef]
  143. Chen, J.; Lin, S.; Zhao, D.; Guan, L.; Hu, Y.; Wang, Y.; Lin, K.; Zhu, Y. Palladium Nanocrystals-Engineered Metal-Organic Frameworks for Enhanced Tumor Inhibition by Synergistic Hydrogen/Photodynamic Therapy. Adv. Funct. Mater. 2020, 31, 2006853. [Google Scholar] [CrossRef]
  144. Tan, L.; Huang, R.; Li, X.; Liu, S.; Shen, Y.M. Controllable release of nitric oxide and doxorubicin from engineered nanospheres for synergistic tumor therapy. Acta Biomater. 2017, 57, 498–510. [Google Scholar] [CrossRef]
  145. Shen, L.J.; Zhou, T.J.; Fan, Y.T.; Chang, X.; Wang, Y.; Sun, J.G.; Xing, L.; Jiang, H.L. Recent progress in tumor photodynamic immunotherapy. Chin. Chem. Lett. 2020, 31, 1709–1716. [Google Scholar] [CrossRef]
  146. Li, S.Y.; ChEng, H.; Qiu, W.X.; Zhang, L.; Wan, S.S.; ZEng, J.Y.; Zhang, X.Z. Cancer cell membrane-coated biomimetic platform for tumor targeted photodynamic therapy and hypoxia-amplified bioreductive therapy. Biomaterials 2017, 142, 149–161. [Google Scholar] [CrossRef]
  147. Fang, R.H.; Hu, C.M.; Luk, B.T.; Gao, W.; Copp, J.A.; Tai, Y.; O’Connor, D.E.; Zhang, L. Cancer cell membrane-coated nanoparticles for anticancer vaccination and drug delivery. Nano Lett. 2014, 14, 2181–2188. [Google Scholar] [CrossRef]
  148. Li, S.Y.; ChEng, H.; Xie, B.R.; Qiu, W.X.; ZEng, J.Y.; Li, C.X.; Wan, S.S.; Zhang, L.; Liu, W.L.; Zhang, X.Z. Cancer Cell Membrane Camouflaged Cascade Bioreactor for Cancer Targeted Starvation and Photodynamic Therapy. ACS Nano 2017, 11, 7006–7018. [Google Scholar] [CrossRef] [PubMed]
  149. Jiang, L.; Ding, Y.; Xue, X.L.; Zhou, S.S.; Li, C.; Zhang, X.K.; Jiang, X.Q. Entrapping multifunctional dendritic nanoparticles into a hydrogel for local therapeutic delivery and synergetic immunochemotherapy. Nano Res. 2018, 11, 6062–6073. [Google Scholar] [CrossRef]
  150. Wang, T.; Wang, D.; Liu, J.; FEng, B.; Zhou, F.; Zhang, H.; Zhou, L.; Yin, Q.; Zhang, Z.; Cao, Z.; et al. Acidity-Triggered Ligand-Presenting Nanoparticles To Overcome Sequential Drug Delivery Barriers to Tumors. Nano Lett. 2017, 17, 5429–5436. [Google Scholar] [CrossRef] [PubMed]
  151. Chen, Y.; Liu, X.; Yuan, H.; Yang, Z.; von Roemeling, C.A.; Qie, Y.; Zhao, H.; Wang, Y.; Jiang, W.; Kim, B.Y.S. Therapeutic Remodeling of the Tumor Microenvironment Enhances Nanoparticle Delivery. Adv. Sci. 2019, 6, 1802070. [Google Scholar] [CrossRef] [Green Version]
  152. Zinger, A.; Koren, L.; Adir, O.; Poley, M.; Alyan, M.; Yaari, Z.; Noor, N.; Krinsky, N.; Simon, A.; Gibori, H.; et al. Collagenase Nanoparticles Enhance the Penetration of Drugs into Pancreatic Tumors. ACS Nano 2019, 13, 11008–11021. [Google Scholar] [CrossRef] [PubMed]
  153. Ye, J.; Jiang, J.; Zhou, Z.; WEng, Z.; Xu, Y.; Liu, L.; Zhang, W.; Yang, Y.; Luo, J.; Wang, X. Near-Infrared Light and Upconversion Nanoparticle Defined Nitric Oxide-Based Osteoporosis Targeting Therapy. ACS Nano 2021, 15, 13692–13702. [Google Scholar] [CrossRef]
  154. Zaccagna, F.; Anzidei, M.; Sandolo, F.; Marincola, B.C.; Palla, C.; Leonardi, A.; Caliolo, G.; Andreani, F.; De Soccio, V.; Catalano, C.; et al. MRgFUS for liver and pancreas cancer treatments: The Umberto I hospital experience. Transl. Cancer Res. 2014, 3, 430–441. [Google Scholar] [CrossRef]
  155. Min, H.S.; You, D.G.; Son, S.; Jeon, S.; Park, J.H.; Lee, S.; Kwon, I.C.; Kim, K. Echogenic Glycol Chitosan Nanoparticles for Ultrasound-Triggered Cancer Theranostics. Theranostics 2015, 5, 1402–1418. [Google Scholar] [CrossRef]
  156. Min, H.S.; Son, S.; You, D.G.; Lee, T.W.; Lee, J.; Lee, S.; Yhee, J.Y.; Lee, J.; Han, M.H.; Park, J.H.; et al. Chemical gas-generating nanoparticles for tumor-targeted ultrasound imaging and ultrasound-triggered drug delivery. Biomaterials 2016, 108, 57–70. [Google Scholar] [CrossRef]
  157. Liang, S.; DEng, X.; Ma, P.; Cheng, Z.; Lin, J. Recent Advances in Nanomaterial-Assisted Combinational Sonodynamic Cancer Therapy. Adv. Mater. 2020, 32, e2003214. [Google Scholar] [CrossRef] [PubMed]
  158. Xiang, H.J.; Guo, M.; An, L.; Yang, S.P.; Zhang, Q.L.; Liu, J.G. A multifunctional nanoplatform for lysosome targeted delivery of nitric oxide and photothermal therapy under 808 nm near-infrared light. J. Mater. Chem. B 2016, 4, 4667–4674. [Google Scholar] [CrossRef] [PubMed]
  159. Du, Z.; Zhang, X.; Guo, Z.; Xie, J.; Dong, X.; Zhu, S.; Du, J.; Gu, Z.; Zhao, Y. X-Ray-Controlled Generation of Peroxynitrite Based on Nanosized LiLuF4: Ce3+ Scintillators and their Applications for Radiosensitization. Adv. Mater. 2018, 30, e1804046. [Google Scholar] [CrossRef]
  160. Nosrati, H.; Charmi, J.; Salehiabar, M.; Abhari, F.; Danafar, H. Tumor Targeted Albumin Coated Bismuth Sulfide Nanoparticles (Bi2S3) as Radiosensitizers and Carriers of Curcumin for Enhanced Chemoradiation Therapy. ACS Biomater. Sci. Eng. 2019, 5, 4416–4424. [Google Scholar] [CrossRef] [PubMed]
  161. Ning, S.; Bednarski, M.; Oronsky, B.; Scicinski, J.; Knox, S.J. Novel nitric oxide generating compound glycidyl nitrate enhances the therapeutic efficacy of chemotherapy and radiotherapy. Biochem. Biophys. Res. Commun. 2014, 447, 537–542. [Google Scholar] [CrossRef]
  162. Gao, S.; Zhang, W.; Wang, R.; Hopkins, S.P.; Spagnoli, J.C.; Racin, M.; Bai, L.; Li, L.; Jiang, W.; Yang, X.; et al. Nanoparticles Encapsulating Nitrosylated Maytansine to Enhance Radiation Therapy. ACS Nano 2020, 14, 1468–1481. [Google Scholar] [CrossRef] [PubMed]
  163. Dou, Y.; Liu, Y.; Zhao, F.; Guo, Y.; Li, X.; Wu, M.; Chang, J.; Yu, C. Radiation-responsive scintillating nanotheranostics for reduced hypoxic radioresistance under ROS/NO-mediated tumor microenvironment regulation. Theranostics 2018, 8, 5870–5889. [Google Scholar] [CrossRef] [PubMed]
  164. Tu, J.; Tu, K.; Xu, H.; Wang, L.; Yuan, X.; Qin, X.; Kong, L.; Chu, Q.; Zhang, Z. Improving tumor hypoxia and radiotherapy resistance via in situ nitric oxide release strategy. Eur. J. Pharm. BioPharm. 2020, 150, 96–107. [Google Scholar] [CrossRef]
  165. Wilson, A.; Menon, V.; Khan, Z.; Alam, A.; Litovchick, L.; Yakovlev, V. Nitric oxide-donor/PARP-inhibitor combination: A new approach for sensitization to ionizing radiation. Redox Biol. 2019, 24, 101169. [Google Scholar] [CrossRef]
  166. Fu, J.; Liu, L.; Huang, Z.; Lai, Y.; Ji, H.; PEng, S.; Tian, J.; Zhang, Y. Hybrid molecule from O2-(2,4-dinitrophenyl)diazeniumdiolate and oleanolic acid: A glutathione S-transferase π-activated nitric oxide prodrug with selective anti-human hepatocellular carcinoma activity and improved stability. J. Med. Chem. 2013, 56, 4641–4655. [Google Scholar] [CrossRef]
  167. Sun, Z.; Yi, Z.; Cui, X.; Chen, X.; Su, W.; Ren, X.; Li, X. Tumor-targeted and nitric oxide-generated nanogels of keratin and hyaluronan for enhanced cancer therapy. Nanoscale 2018, 10, 12109–12122. [Google Scholar] [CrossRef] [PubMed]
  168. Jia, X.; Zhang, Y.; Zou, Y.; Wang, Y.; Niu, D.; He, Q.; Huang, Z.; Zhu, W.; Tian, H.; Shi, J.; et al. Dual Intratumoral Redox/Enzyme-Responsive NO-Releasing Nanomedicine for the Specific, High-Efficacy, and Low-Toxic Cancer Therapy. Adv. Mater. 2018, 30, e1704490. [Google Scholar] [CrossRef]
  169. Hou, L.; Zhang, Y.; Yang, X.; Tian, C.; Yan, Y.; Zhang, H.; Shi, J.; Zhang, H.; Zhang, Z. Intracellular NO-Generator Based on Enzyme Trigger for Localized Tumor-Cytoplasm Rapid Drug Release and Synergetic Cancer Therapy. ACS Appl. Mater. Interfaces 2019, 11, 255–268. [Google Scholar] [CrossRef]
  170. Sun, B.; DEng, C.; MEng, F.; Zhang, J.; Zhong, Z. Robust, active tumor-targeting and fast bioresponsive anticancer nanotherapeutics based on natural endogenous materials. Acta Biomater. 2016, 45, 223–233. [Google Scholar] [CrossRef]
  171. Feng, X.; Xu, W.; Li, Z.; Song, W.; Ding, J.; Chen, X. Immunomodulatory Nanosystems. Adv. Sci. 2019, 6, 1900101. [Google Scholar] [CrossRef]
  172. Hu, Y.; Lv, T.; Ma, Y.; Xu, J.; Zhang, Y.; Hou, Y.; Huang, Z.; Ding, Y. Nanoscale Coordination Polymers for Synergistic NO and Chemodynamic Therapy of Liver Cancer. Nano Lett. 2019, 19, 2731–2738. [Google Scholar] [CrossRef]
  173. Li, K.; Lin, C.; He, Y.; Lu, L.; Xu, K.; Tao, B.; Xia, Z.; ZEng, R.; Mao, Y.; Luo, Z.; et al. Engineering of Cascade-Responsive Nanoplatform to Inhibit Lactate Efflux for Enhanced Tumor Chemo-Immunotherapy. ACS Nano 2020, 14, 14164–14180. [Google Scholar] [CrossRef]
  174. Helmlinger, G.; Yuan, F.; Dellian, M.; Jain, R.K. Interstitial pH and pO2 gradients in solid tumors in vivo: High-resolution measurements reveal a lack of correlation. Nat. Med. 1997, 3, 177–182. [Google Scholar] [CrossRef]
  175. Cao, Y.; Liu, M.; ChEng, J.; Yin, J.; Huang, C.; Cui, H.; Zhang, X.; Zhao, G. Acidity-Triggered Tumor-Targeted Nanosystem for Synergistic Therapy via a Cascade of ROS Generation and NO Release. ACS Appl. Mater. Interfaces 2020, 12, 28975–28984. [Google Scholar] [CrossRef]
  176. Li, Y.; Lin, J.; Zhi, X.; Li, P.; Jiang, X.; Yuan, J. Triple stimuli-responsive keratin nanoparticles as carriers for drug and potential nitric oxide release. Mater. Sci. Eng. C Mater. Biol. Appl. 2018, 91, 606–614. [Google Scholar] [CrossRef] [PubMed]
  177. Wang, J.; Bi, Y.; Ruan, H.; Sun, G.; Cui, X.; Yang, X.; Qin, C. Hollow S-nitrosothiols nanoparticle with polymeric brushes for nitric oxide (NO)-releasing as tumor targeted chemotherapy. J. Biomater. Sci. Polym Ed. 2019, 30, 122–136. [Google Scholar] [CrossRef] [PubMed]
  178. Zhang, P.; Zhang, Y.; Ding, X.; Shen, W.; Li, M.; Wagner, E.; Xiao, C.; Chen, X. A Multistage Cooperative Nanoplatform Enables Intracellular Co-Delivery of Proteins and Chemotherapeutics for Cancer Therapy. Adv. Mater. 2020, 32, e2000013. [Google Scholar] [CrossRef] [PubMed]
  179. Ling, P.; Qian, C.; Gao, F.; Lei, J. Enzyme-immobilized metal-organic framework nanosheets as tandem catalysts for the generation of nitric oxide. Chem. Commun. 2018, 54, 11176–11179. [Google Scholar] [CrossRef]
  180. Yang, F.; Li, M.; Liu, Y.; Wang, T.; Feng, Z.; Cui, H.; Gu, N. Glucose and magnetic-responsive approach toward in situ nitric oxide bubbles controlled generation for hyperglycemia theranostics. J. Control. Release 2016, 228, 87–95. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. A review of the different endogenous and exogenous stimulus-responsive NO nanomedicines for the treatment of cancer. (Abbreviations—US: ultrasound; SDT: sonodynamic therapy; NIR: near infrared; PDT: photodynamic therapy; PTT: photothermal therapy; GSH: glutathione.)
Scheme 1. A review of the different endogenous and exogenous stimulus-responsive NO nanomedicines for the treatment of cancer. (Abbreviations—US: ultrasound; SDT: sonodynamic therapy; NIR: near infrared; PDT: photodynamic therapy; PTT: photothermal therapy; GSH: glutathione.)
Pharmaceutics 13 01917 sch001
Figure 1. (A) Schematic diagram of NO release in cells by the mPEG–PLGA–BNN6–DOX nanosystem and the SEM images of the nanoparticles. Adapted with permission from [68], ACS Applied Materials & Interfaces, 2016; (B) schematic diagram of Cdot–TPP–SNO nanosystem releasing NO in cancer cells for anti-tumor therapy. Reproduced with permission from [69], Small, 2014; (C) schematic diagram of the NO release in tumor cells by the Ru–NO@TiO2NPs nanosystem. Reproduced with permission from [70], Chemical Communications, 2015.
Figure 1. (A) Schematic diagram of NO release in cells by the mPEG–PLGA–BNN6–DOX nanosystem and the SEM images of the nanoparticles. Adapted with permission from [68], ACS Applied Materials & Interfaces, 2016; (B) schematic diagram of Cdot–TPP–SNO nanosystem releasing NO in cancer cells for anti-tumor therapy. Reproduced with permission from [69], Small, 2014; (C) schematic diagram of the NO release in tumor cells by the Ru–NO@TiO2NPs nanosystem. Reproduced with permission from [70], Chemical Communications, 2015.
Pharmaceutics 13 01917 g001aPharmaceutics 13 01917 g001b
Figure 2. (A) Schematic diagram of the preparation and release of NO from photo-triggered NO nanogenerators (PTNG). Reproduced with permission from [71], Advanced Functional Materials, 2017; (B) schematic diagram representation of the therapeutic function of PpRE@PEG–PpIX NPs. Reproduced with permission from [72], Nano Research, 2019; (C) schematic diagram of the synthesis and NIR light-triggered NO release from the BNN–Bi2S3. Reproduced with permission from [74], Advanced Science, 2019; (D) schematic diagram of the preparation of PEG–PAu@SiO2–SNO and the process of mild heat-enhanced NO therapy under NIR irradiation in cancer cells. Reproduced with permission from [76], ACS Biomaterials Science & Engineering, 2020.
Figure 2. (A) Schematic diagram of the preparation and release of NO from photo-triggered NO nanogenerators (PTNG). Reproduced with permission from [71], Advanced Functional Materials, 2017; (B) schematic diagram representation of the therapeutic function of PpRE@PEG–PpIX NPs. Reproduced with permission from [72], Nano Research, 2019; (C) schematic diagram of the synthesis and NIR light-triggered NO release from the BNN–Bi2S3. Reproduced with permission from [74], Advanced Science, 2019; (D) schematic diagram of the preparation of PEG–PAu@SiO2–SNO and the process of mild heat-enhanced NO therapy under NIR irradiation in cancer cells. Reproduced with permission from [76], ACS Biomaterials Science & Engineering, 2020.
Pharmaceutics 13 01917 g002
Figure 3. (A) Schematic diagram of the PFTDPP–SNAP which achieved PTT and NO therapy with NIR light irradiation. Reproduced with permission from [77], Nano Letters, 2019; (B) schematic diagram representation of the therapeutic function of Nb2C–MSNS–SNO. Reproduced with permission from [33], Advanced Science, 2020; (C) schematic diagram of preparation of P(IR/BNN6/AIPH)@Lip–RGD and the process of PTT and NO therapy with NIR-II irradiation in cancer cells. Reproduced with permission from [78], Nanoscale, 2021; (D) schematic diagram of the preparation and drugs releasing process of nanomedicine and the process of reprogramming tumor immunosuppressive microenvironment and fighting immune cold tumor. Reproduced with permission from [79], ACS Nano, 2020.
Figure 3. (A) Schematic diagram of the PFTDPP–SNAP which achieved PTT and NO therapy with NIR light irradiation. Reproduced with permission from [77], Nano Letters, 2019; (B) schematic diagram representation of the therapeutic function of Nb2C–MSNS–SNO. Reproduced with permission from [33], Advanced Science, 2020; (C) schematic diagram of preparation of P(IR/BNN6/AIPH)@Lip–RGD and the process of PTT and NO therapy with NIR-II irradiation in cancer cells. Reproduced with permission from [78], Nanoscale, 2021; (D) schematic diagram of the preparation and drugs releasing process of nanomedicine and the process of reprogramming tumor immunosuppressive microenvironment and fighting immune cold tumor. Reproduced with permission from [79], ACS Nano, 2020.
Pharmaceutics 13 01917 g003
Figure 4. (A) Schematic diagram of the preparation of the L-Arg@PCN@Mem and the tumor-killing mechanism. Reproduced with permission from [83], Biomaterials, 2018; (B) mechanism diagram of cancer cells apoptosis facilitated by the integration of various functions within L-arginine ethyl ester-modified zinc phthalocyanine (Arg–ZnPc). Reproduced with permission from [84], Molecular Pharmaceutics, 2020; (C) schematic diagram of the preparation of L-arginine/Dox-loaded gold@copper sulfide yolk−shell nanoparticles (ADLAu@CuS YSNPs), the mechanism diagram of the process of NO and DOX being sequentially released plays an anti-tumor effect under NIR irradiation in cancer cells. Reproduced with permission from [85], Nano Letters, 2019; (D) schematic diagram of the preparation of PLGA@ICG@L-Arg nanoparticles and the process of anti-tumor effects under NIR irradiation in cancer cells. Reproduced with permission from [86], Molecular Pharmaceutics, 2021.
Figure 4. (A) Schematic diagram of the preparation of the L-Arg@PCN@Mem and the tumor-killing mechanism. Reproduced with permission from [83], Biomaterials, 2018; (B) mechanism diagram of cancer cells apoptosis facilitated by the integration of various functions within L-arginine ethyl ester-modified zinc phthalocyanine (Arg–ZnPc). Reproduced with permission from [84], Molecular Pharmaceutics, 2020; (C) schematic diagram of the preparation of L-arginine/Dox-loaded gold@copper sulfide yolk−shell nanoparticles (ADLAu@CuS YSNPs), the mechanism diagram of the process of NO and DOX being sequentially released plays an anti-tumor effect under NIR irradiation in cancer cells. Reproduced with permission from [85], Nano Letters, 2019; (D) schematic diagram of the preparation of PLGA@ICG@L-Arg nanoparticles and the process of anti-tumor effects under NIR irradiation in cancer cells. Reproduced with permission from [86], Molecular Pharmaceutics, 2021.
Pharmaceutics 13 01917 g004
Figure 5. (A) Schematic diagram of the preparation of BNN6–SPION@hMSN which could achieve the controlled release of NO with ultrasound stimulation. Reproduced with permission from [88], Nanoscale, 2017; (B) schematic diagram of the therapeutic modes of TPZ/HMTNPs–SNO in cancer cells. Reproduced with permission from [89], Journal of Controlled Release, 2018; (C) Schematic diagram of the anti-tumor effect of nanomicelles IMesNO/DOX@MCs under HIFU irradiation. Reproduced with permission from [90], Biomaterials, 2019; (D) schematic diagram of the preparation of the GSNO/Ce6@ZIF-8@Cytomembrane (GCZ@M) and the process of anti-tumor effects in cells. Reproduced with permission from [91], Biomaterials, 2019.
Figure 5. (A) Schematic diagram of the preparation of BNN6–SPION@hMSN which could achieve the controlled release of NO with ultrasound stimulation. Reproduced with permission from [88], Nanoscale, 2017; (B) schematic diagram of the therapeutic modes of TPZ/HMTNPs–SNO in cancer cells. Reproduced with permission from [89], Journal of Controlled Release, 2018; (C) Schematic diagram of the anti-tumor effect of nanomicelles IMesNO/DOX@MCs under HIFU irradiation. Reproduced with permission from [90], Biomaterials, 2019; (D) schematic diagram of the preparation of the GSNO/Ce6@ZIF-8@Cytomembrane (GCZ@M) and the process of anti-tumor effects in cells. Reproduced with permission from [91], Biomaterials, 2019.
Pharmaceutics 13 01917 g005
Figure 6. (A) (a) Schematic diagram of PEG–USMSs–SNO construction; (b) transmission electron microscope images of USMSs; (c) SEM images of USMSs; (d) elemental mapping of the merging of Si and Y of USMSs; (e) elemental mapping of merging of Si of USMSs; (f) elemental mapping of merging of Y of USMSs (USMSs: a dense silica shell was coated on upconversion nanoparticles). Reproduced with permission from [94], Angewandte Chemie International Edition, 2015. (B) schematic diagram of the preparation of Bi–SNO-NPs and the process of enhanced NO therapy with X-ray irradiation in cancer cells. Reproduced with permission from [95], Nanoscale, 2020. (C) schematic diagram of the preparation of the ZGO:Mn-RBS nanosystem and the process of anti-tumor effects. Reproduced with permission from [96], Biomaterials, 2020.
Figure 6. (A) (a) Schematic diagram of PEG–USMSs–SNO construction; (b) transmission electron microscope images of USMSs; (c) SEM images of USMSs; (d) elemental mapping of the merging of Si and Y of USMSs; (e) elemental mapping of merging of Si of USMSs; (f) elemental mapping of merging of Y of USMSs (USMSs: a dense silica shell was coated on upconversion nanoparticles). Reproduced with permission from [94], Angewandte Chemie International Edition, 2015. (B) schematic diagram of the preparation of Bi–SNO-NPs and the process of enhanced NO therapy with X-ray irradiation in cancer cells. Reproduced with permission from [95], Nanoscale, 2020. (C) schematic diagram of the preparation of the ZGO:Mn-RBS nanosystem and the process of anti-tumor effects. Reproduced with permission from [96], Biomaterials, 2020.
Pharmaceutics 13 01917 g006
Figure 7. (A) Schematic diagram of the structure and preparation of the p(Gd-Az-JSK) NPs and their delivery process. Reproduced with permission from [97], Nanoscale, 2018; (B) schematic diagram of PEG-b-NO-Dex nanoparticles releasing NO and boosting EPR in the presence of GSH. Reproduced with permission from [98], Nano Letters, 2018; (C) schematic diagram of the structure and preparation of the HCPT/CTS–NO–DMMA NPs and its delivery process. Reproduced with permission from [99], Nanomedicine, 2019; and (D) schematic diagram of preparation of the α-CD–DOXNO–DA NPs and the process of exerting anti-tumor effects in cancer cells. Reproduced with permission from [100], Small, 2020.
Figure 7. (A) Schematic diagram of the structure and preparation of the p(Gd-Az-JSK) NPs and their delivery process. Reproduced with permission from [97], Nanoscale, 2018; (B) schematic diagram of PEG-b-NO-Dex nanoparticles releasing NO and boosting EPR in the presence of GSH. Reproduced with permission from [98], Nano Letters, 2018; (C) schematic diagram of the structure and preparation of the HCPT/CTS–NO–DMMA NPs and its delivery process. Reproduced with permission from [99], Nanomedicine, 2019; and (D) schematic diagram of preparation of the α-CD–DOXNO–DA NPs and the process of exerting anti-tumor effects in cancer cells. Reproduced with permission from [100], Small, 2020.
Pharmaceutics 13 01917 g007
Figure 8. (A) Schematic structure/composition of hollow microspheres (HMs) and the mechanism in the treatment of tumors. Reproduced with permission from [49], Angewandte Chemie International Edition, 2015; (B) schematic diagram of the preparation of the GSNO–MNPs and the intracellular NO release from the GSNO-loaded CaCO3-mineralized nanoparticles (GSNO-MNPs). Reproduced with permission from [102], Colloids and Surfaces B: Biointerfaces, 2016; and (C) schematic diagram of the preparation of the NO-based nanosystem. Reproduced with permission from [103], Angewandte Chemie International Edition, 2021.
Figure 8. (A) Schematic structure/composition of hollow microspheres (HMs) and the mechanism in the treatment of tumors. Reproduced with permission from [49], Angewandte Chemie International Edition, 2015; (B) schematic diagram of the preparation of the GSNO–MNPs and the intracellular NO release from the GSNO-loaded CaCO3-mineralized nanoparticles (GSNO-MNPs). Reproduced with permission from [102], Colloids and Surfaces B: Biointerfaces, 2016; and (C) schematic diagram of the preparation of the NO-based nanosystem. Reproduced with permission from [103], Angewandte Chemie International Edition, 2021.
Pharmaceutics 13 01917 g008
Figure 9. (A) Schematic diagram of the construction of L-Arg–HMON–GOx for synergistic cancer therapy. Reproduced with permission from [104], Angewandte Chemie International Edition, 2017; and (B) schematic diagram of the preparation of the BPNs–Arg–GOx@MnO2 (BAGN) nanosystem and the intracellular NO release. Reproduced with permission from [105], ACS Applied Materials & Interfaces, 2021.
Figure 9. (A) Schematic diagram of the construction of L-Arg–HMON–GOx for synergistic cancer therapy. Reproduced with permission from [104], Angewandte Chemie International Edition, 2017; and (B) schematic diagram of the preparation of the BPNs–Arg–GOx@MnO2 (BAGN) nanosystem and the intracellular NO release. Reproduced with permission from [105], ACS Applied Materials & Interfaces, 2021.
Pharmaceutics 13 01917 g009
Table 1. Summary of the release mechanism and theranostic method of stimulus-responsive NO-based nanomedicine.
Table 1. Summary of the release mechanism and theranostic method of stimulus-responsive NO-based nanomedicine.
Trigger for NO
Release
Nanomedicine FormulationNO DonorNO Release MechanismTheranostic MethodReference
UV–VismPEG–PLGA–BNN6–DOXBNN6The breaking of bondsNO therapy–chemotherapy[68]
Cdot–TPP–SNOR-SNOThe breaking of bondsNO therapy[69]
Ru-NO@TiO2Ru-NOThe breaking of bondsNO therapy–PDT[70]
NIRPTNGR-SNOThe breaking of bondsNO therapy–PTT–chemotherapy[71]
PpRE@PEG–PpIXPpREThe breaking of bondsNO therapy–PTT[72]
Fe3O4@PDA@Ru-NO@FARu-NOThe breaking of bondsNO therapy–PTT–MRI[73]
BNN–Bi2S3BNN6The breaking of bondsNO therapy–mild PTT[74]
PNOC–PDAR-SNOThe breaking of bondsNO therapy–mild PTT–chemotherapy[75]
PEG–PAu@SiO2–SNOR-SNOThe breaking of bondsNO therapy–mild PTT[76]
PFTDPP–SNAPSNAPThe breaking of bondsNO therapy–PTT– NIR II/PA imaging[77]
Nb2C–MSNs–SNOR-SNOThe breaking of bondsNO therapy–PTT– PA imaging[33]
P(IR/BNN6/AIPH)@Lip–RGDBNN6The breaking of bondsNO therapy–alkyl radicals–PTT[78]
RBCm/PAAV–SNOR-SNOThe breaking of bondsNO therapy–PTT–reprogramming tumor immunosuppressive microenvironment[79]
PDA–PLC/DOXR-SNOThe breaking of bondsNO therapy–PTT–chemotherapy[80]
Me–RBSsRBSThe breaking of bondsNO therapy[81]
GO–BNN6BNN6The breaking of bondsNO therapy[82]
L-Arg@PCN@MemL-ArgOxidation–reduction reactionNO therapy–PDT[83]
Arg–ZnPcL-ArgOxidation–reduction reactionNO therapy–PDT[84]
ADLAu@CuS YSNPsL-ArgOxidation–reduction reactionNO therapy–PDT–chemotherapy[85]
PLGA@ICG@L-ArgL-ArgOxidation–reduction reactionNO therapy–PDT[86]
Lyso–Ru-NO@FA@C–TiO2Ru-NOThe breaking of bondsNO therapy–PDT[87]
USBNN6–SPION@hMSNBNN6The breaking of bondsNO therapy–MRI[88]
TPZ/HMTNPs–SNOR-SNOThe breaking of bondsNO therapy–SDT–USI[89]
IMesNO/DOX@MCsIMesNOThe breaking of bondsNO therapy–chemotherapy[90]
GSNO/Ce6@ZIF–8@Cytomembrane (GCZ@M) GSNOOxidation–reduction reactionNO therapy–SDT[91]
SNO–HSA–PTXR-SNOThe breaking of bondsNO therapy–chemotherapy–immunotherapy[92]
peptide−HMSN−LAL-ArgOxidation–reduction reactionNO therapy–SDT[93]
X-rayPEG–USMSs–SNOR-SNOThe breaking of bondsNO therapy–radiotherapy [94]
Bi–SNOR-SNOThe breaking of bondsNO therapy–radiotherapy–CT imaging–PTT[95]
ZGO:Mn–RBSRBSThe breaking of bondsNO therapy–radiotherapy [96]
GSHp(Gd–Az–JSK)alkynyl-JSKThe nucleophilic attackingNO therapy–chemotherapy[97]
PEG–b–NO-Dex–DOXNO-DexOxidation–reduction reactionNO therapy–chemotherapy[98]
HCPT/CTS–NO–DMMAPSFOxidation–reduction reactionNO therapy[99]
α–CD–DOX–NO–DAR-SNO/NO therapy–chemotherapy[100]
TNO3–DOXTNO3Oxidation–reduction reactionNO therapy–chemotherapy[101]
pHhollow microsphere (HM)DETA NONOateHydrolysis reactionNO therapy–chemotherapy[49]
GSNO-MNPsGSNOHydrolysis reaction–oxidation–reduction reactionNO therapy–chemotherapy[102]
NO–NCPsDETA NONOateHydrolysis reactionNO therapy–PDT–PA imaging[103]
GlucoseL-Arg–HMON–GOxL-ArgOxidation–reduction reactionNO therapy–starving therapy–USI[104]
BPNs–Arg–GOx@MnO2(BAGM)L-ArgOxidation–reduction reactionNO therapy–starving therapy–PTT[105]
Abbreviations—BNN6: N,N′-di-sec-butyl-N,N′-dinitroso-1,4-phenylenediamine; R-SNO: S-nitrosothiol; Ru-NO: ruthenium nitrosyls; PpRE: porphyrin-modified Roussin’s red-salt ester; SNAP: S-nitroso-N-acetylpenicillamine; RBS: Roussin’s Black Salt([NH4][Fe4S3(NO)7]); L-Arg: L-arginine; IMesNO: 1,3-bis-(2,4,6-trimethylphenyl)imidazolylidene nitric oxide; GSNO: nitrosoglutathione; alkynyl-JSK: O2-(2,4-dinitrophenyl) 1-[4(propargyloxycarbonyl) piperazin-1-yl] diazen-1-ium-1,2-diolate; NO-Dex: hydrophobic nitrifying glucan; PSF: phenylsulfonyl furoxan; TNO3: nitrate functionalized D-α-tocopheryl polyethylene glycol succinate; DETA NONOate: diethylenetriamine diazeniumdiolate; PDT: photodynamic therapy; PTT: photothermal therapy; MRI: magnetic resonance imaging; NIR II/PA imaging: second NIR window and photoacoustic imaging; PA: photoacoustic; SDT: sonodynamic therapy; USI: ultrasound imaging; CT: computed tomography.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, Y.; Ouyang, X.; Peng, Y.; Peng, S. Stimuli Responsive Nitric Oxide-Based Nanomedicine for Synergistic Therapy. Pharmaceutics 2021, 13, 1917. https://doi.org/10.3390/pharmaceutics13111917

AMA Style

Zhao Y, Ouyang X, Peng Y, Peng S. Stimuli Responsive Nitric Oxide-Based Nanomedicine for Synergistic Therapy. Pharmaceutics. 2021; 13(11):1917. https://doi.org/10.3390/pharmaceutics13111917

Chicago/Turabian Style

Zhao, Yijun, Xumei Ouyang, Yongjun Peng, and Shaojun Peng. 2021. "Stimuli Responsive Nitric Oxide-Based Nanomedicine for Synergistic Therapy" Pharmaceutics 13, no. 11: 1917. https://doi.org/10.3390/pharmaceutics13111917

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop