Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Skip to content
BY 4.0 license Open Access Published by De Gruyter Open Access March 10, 2023

Application prospect of calcium peroxide nanoparticles in biomedical field

  • Xincai Wu , Xu Han , Yang Guo , Qian Liu , Ran Sun , Zhaohui Wen EMAIL logo and Changsong Dai

Abstract

In recent years, calcium peroxide (CaO2) has attracted widespread attention in the medical community due to its excellent antitumor and antibacterial properties, and has gradually become a hot research topic in the biomedical field. CaO2 reacts with water (H2O) to produce calcium ion (Ca2+), oxygen (O2), and hydrogen peroxide (H2O2), where Ca2+ is suitable for calcium death caused by calcium overload, O2 is suitable for O2-dependent anticancer therapy, and H2O2 is suitable for H2O2-dependent anticancer therapy. In addition, H2O2 can also be used in the antibacterial field to treat bacterial infections. All these make the CaO2 to become a kind of excellent antitumor and antibacterial drug. This study mainly reviews the preparation and surface modification of CaO2, probes into the latest progress about CaO2 nanoparticles in the field of tumor treatment and antimicrobial therapy. Finally, the challenges that CaO2 still faces in the future research field are clarified, and its prospects are forecasted.

Abbreviations

CaO2

calcium peroxide

CaO

calcium oxide

Ca(OH)2

calcium hydroxide

Ca2+

calcium ion

CAT

catalase

CaCl2

calcium chloride

CDT

chemodynamic therapy

CPO

chloroperoxidase

CT

computed tomography

C. tetani

Clostridium tetani

DMOS

dendritic mesoporous organosilica

DOX

doxorubicin

ECM

extracellular matrix

EDT

enzyme dynamic therapy

EDTA·2Na

ethylenediamine tetra acetic acid disodium salt

EPR

enhanced permeability and retention

E. coli

Escherichia coli

F. nucleatum

Fusobacterium nucleatum

GOx

glucose oxidase

GNPs

gold nanoparticles

GSH

glutathione

GSSH

oxidized glutathione

HA

hyaluronic acid

HCl

hydrochloric acid

HIF-1a

hypoxia inducible factor-1 alpha

HRP

horseradish peroxidase

H2S

hydrogen sulfide

H2O2

hydrogen peroxide

H2O

water

ICG

indocyanine green

IFP

interstitial fluid pressure

LA

lauric acid

l-arg

l-arginine

MnO2

manganese dioxide

MSN

manganese silicate nanoparticle

MOFs

metal–organic frameworks

MRP2

multidrug resistance protein 2

MRI

magnetic resonance imaging

MS

monostearin

NIR

near-infrared

NO

nitric oxide

NP

nanoparticle

O2

oxygen

1O2

singlet oxygen

O2 ·–

superoxide anion

OA

oleyl-amine

OD

optical density

·OH

hydroxyl radical

OH

hydroxyl ions

PAA

polyacrylic acid

PDA

polydopamine

PDT

photodynamic therapy

PEG

polyethylene glycol

P-gp

P-glycoprotein

PS

photosensitizer

PTT

photothermal therapy

PVP

polyethylene pyrrolidone

ROS

reactive oxygen species

SEM

scanning electron microscope

SC

sodium citrate

SDT

sonodynamic therapy

SH

sodium hyaluronate

ST

starvation therapy

S. aureus

Staphylococcus aureus

TA

tannic acid

TACE

transcatheter arterial chemoembolization

TEM

transmission electron microscopy

TME

tumor microenvironment

1 Introduction

So far, cancer is still a major risk factor threatening human health. The traditional methods of cancer treatment mainly include surgery, radiotherapy, and chemotherapy. Although it has a certain therapeutic effect, the unavoidable side effects and poor specificity of cancer cells, so far, cannot meet the needs of cancer treatment. With the progress and development of modern science and technology, new tumor treatment methods such as chemodynamic therapy (CDT) [1,2], photodynamic therapy (PDT) [3,4,5], photothermal therapy (PTT) [6,7], sonodynamic therapy (SDT) [8], and calcium overload [9] emerge as the times requirement, which broadens the mode of diagnosis and treatment related to cancer. It brings more hope for cancer patients. However, affected by the tumor microenvironment (TME), the emerging tumor treatment methods are also difficult to play effectively.

For bacterial infection, antibiotics are the most commonly used method to treat bacterial infection [10]. Although it reduces the morbidity and mortality of human beings, its long-term and overuse makes multi-drug-resistant bacterial infection one of the world’s public health threats [11,12]. At present, oxidative stress induced by reactive oxygen species (ROS) can effectively treat bacterial infection, which is favored by researchers because of its destructive effect on bacteria.

Under acidic conditions, CaO2 with structural peroxy bond reacts with H2O to form a large amount of H2O2(0.47 g H2O2/g CaO2) and a small amount of O2(0.2222 O2/g CaO2) [13,14], accompanied by the release of Ca2+. Therefore, the introduction of CaO2, on the one hand, reshape the TME to ensure the effective play of anticancer therapy such as O2-dependent (chemotherapy, PDT), H2O2-dependent (CDT), and calcium overload, while, on the other hand, effectively inhibit bacterial infection.

This article mainly summarized the preparation and surface modification of CaO2, probes into the latest progress about CaO2 nanoparticles in the field of tumor treatment and antimicrobial therapy. The schematic diagram of the review is shown in Figure 1.

Figure 1 
               Schematic diagram of the review: Preparation, surface modification, and application of CaO2 NPs in tumor and bacteria. Ion interference therapy: reproduced with permission from previous study [15]. © 2019 Elsevier Inc. O2 dependence therapy: reproduced with permission from previous study [16]. Copyright © 2021, American Chemical Society. H2O2 dependence therapy (the field of tumor therapy): reproduced with permission from previous study [17]. © 2021 Elsevier B.V. All rights reserved. H2O2 dependence therapy (the field of bacterial therapy): reproduced with permission from previous study [18]. © 2021 Wiley‐VCH GmbH; Surface modification: reproduced with permission from previous study [19]. © 2021 Elsevier B.V. All rights reserved. Preparation method: reproduced with permission from previous study [18]. © 2021 Wiley‐VCH GmbH. Notes: polyethylene glycol (PEG), polyethylene pyrrolidone (PVP), hyaluronic acid (HA), sodium hyaluronate (SH), tannic acid (TA), calcium chloride (CaCl2), calcium oxide (CaO), calcium hydroxide (Ca(OH)2).
Figure 1

Schematic diagram of the review: Preparation, surface modification, and application of CaO2 NPs in tumor and bacteria. Ion interference therapy: reproduced with permission from previous study [15]. © 2019 Elsevier Inc. O2 dependence therapy: reproduced with permission from previous study [16]. Copyright © 2021, American Chemical Society. H2O2 dependence therapy (the field of tumor therapy): reproduced with permission from previous study [17]. © 2021 Elsevier B.V. All rights reserved. H2O2 dependence therapy (the field of bacterial therapy): reproduced with permission from previous study [18]. © 2021 Wiley‐VCH GmbH; Surface modification: reproduced with permission from previous study [19]. © 2021 Elsevier B.V. All rights reserved. Preparation method: reproduced with permission from previous study [18]. © 2021 Wiley‐VCH GmbH. Notes: polyethylene glycol (PEG), polyethylene pyrrolidone (PVP), hyaluronic acid (HA), sodium hyaluronate (SH), tannic acid (TA), calcium chloride (CaCl2), calcium oxide (CaO), calcium hydroxide (Ca(OH)2).

2 Preparation and surface modifiers of CaO2

2.1 Preparation of CaO2

At present, there are three main methods to prepare CaO2 at home and abroad: CaCl2 method, CaO method, and Ca(OH)2 method. Ca(OH)2 method is divided into traditional method, spray drying method, and air cathode method. Among them, CaCl2 method and traditional Ca(OH)2 method belong to hydrolytic precipitation method and are the most commonly used methods for preparing CaO2 [20].

2.1.1 CaCl2 method

General preparation process: first of all, ammonia water is injected into an alkaline solution of metal chloride (such as CaCl2), and in the agitation state, H2O2 containing a stabilizer is injected to activate the reaction [21]. The reaction equation is as follows:

(1) CaC l 2 + H 2 O 2 + 2 N H 3 · H 2 O + 6 H 2 O = Ca O 2 · 8 H 2 O + 2 N H 4 Cl ,

(2) Ca O 2 · 8 H 2 O Ca O 2 + H 2 O .

Without the addition of stabilizer, low temperature reaction is needed, and the preparation process is complicated and the cost is high. At present, the method of preparing CaO2 at room temperature by introducing stabilizer is widely used to improve the utilization rate of H2O2 and the yield of CaO2, and to overcome the problems of complex preparation process and high cost under low temperature production conditions.

The injecting of ammonia provides the required alkaline environment for the synthesis of CaO2 material and neutralizes the by-product hydrochloric acid (HCl), which promotes the forward progress of the reaction [22]. Similarly, sodium hydroxide as an alkaline substance, cannot replace ammonia to provide a suitable alkaline environment for the synthesis of CaO2 NPs. The reason might be that sodium hydroxide is a strong base environment which can easily generate Ca(OH)2 precipitation and excessive hydroxyl ions to promote H2O2 decomposition.

CaCl2 method has the advantages of simple preparation process, low cost, mature technology, and suitable for small-scale production. However, due to the use of dilute solution production, the decomposition loss of H2O2 in mother liquor is large, and energy consumption is large, resulting in low product content of CaO2, generally 50–60%.

2.1.2 CaO method

The general preparation process: CaO is dissolved in H2O to generate Ca(OH)2 solution, and after the temperature of the reaction solution is stable, H2O2 is added to obtain CaO2·8H2O. After filtering and drying the filter cake, anhydrous calcium peroxide (CaO2) is obtained. The reaction equation is as follows:

(3) CaO + H 2 O Ca ( OH ) 2 ,

(4) Ca ( OH ) 2 + H 2 O 2 + 6 H 2 O Ca O 2 · 8 H 2 O ,

(5) Ca O 2 · 8 H 2 O Ca O 2 + H 2 O .

The preparation of CaO2 by CaO method has the advantages of cheap raw materials, simple preparation process, simple equipment, no need to add ammonia and other substances, and basically no problem of discharge of the three wastes.

2.1.3 Ca(OH)2 method

2.1.3.1 Traditional Ca(OH)2 method

General preparation process: under the action of stabilizer, Ca(OH)2 is slowly added to H2O2 to get CaO2. After further drying anhydrous, CaO2 material can be obtained [23]. The reaction equation is as follows:

(6) Ca ( OH ) 2 + H 2 O 2 + 6 H 2 O Ca O 2 · 8 H 2 O ,

(7) Ca O 2 · 8 H 2 O Ca O 2 + 8 H 2 O .

The purpose of introducing stabilizer is to make the reaction proceed at room temperature and improve the utilization rate of H2O2 and the stability of the product. When no stabilizer is added, low temperature reaction is required where the reaction temperature should be controlled below 5℃, which is of high cost and high energy consumption.

Because CaO2 is slightly soluble in H2O, most of the domestic Ca(OH)2 is dissolved in ammonium solution to generate ammonium complex, and then the free Ca2+ in the complex reacts with H2O2 to prepare CaO2 substance. The reaction equation is as follows:

(8) Ca ( OH ) 2 + 2 N H 4 Cl Ca ( N H 3 ) 2 C l 2 · 2 H 2 O CaC l 2 + 2 N H 3 · H 2 O ,

(9) CaC l 2 + H 2 O 2 + 2 N H 3 · H 2 O + 6 H 2 O Ca O 2 · 8 H 2 O + 2 N H 4 Cl .

The preparation of CaO2 by Ca(OH)2 method simplifies the production and preparation process, reduces energy consumption and production cost, and the technology is mature. However, because it is also produced by dilute solution, the product yield and the content of CaO2 are not high, and the content of CaO2 is generally 50–60%.

2.1.3.2 Spray drying method

Under the condition of cooling and constant stirring, the concentrated suspension of Ca(OH)2 water reacted with the concentrated H2O2, and the resulting mixture was spray dried directly after the reaction [24]. This method can obtain anhydrous CaO2 with good dispersion and uniform particle size without separation and purification. The reaction equation is as follows:

(10) Ca ( OH ) 2 + H 2 O 2 Ca O 2 + 2 H 2 O ,

(11) CaO + H 2 O 2 Ca O 2 + H 2 O .

Compared with the traditional preparation method, spray drying method has the following advantages: (a) H2O2 and Ca(OH)2 are cheap and easy to obtain. (b) It saves time and cost without separation and purification. (c) It improves the utilization rate of H2O2 and the yield of CaO2. (d) Energy saving. (e) Continuous production and intermittent production can be carried out, so that the product has good uniformity.

However, this method is also faced with problems such as heavy equipment investment, easy blockage of pipes, high energy consumption, and explosive risk. In addition, high concentration of H2O2 is lacking in China.

2.1.3.3 Air cathode method

In order to further solve the problem of high cost of H2O2, this method is a low-cost H2O2 production process after the industrialization of producing H2O2 by anthraquinone method. Ca(OH)2 reacts with diluted H2O2 to generate CaO2·8H2O. The mother liquor is removed after centrifugation, and the obtained NaOH is recycled or discharged in time. CaO2 product can be obtained by grinding the product after vacuum drying of filter cake. The reaction equation is as follows:

(12) Cathode : O 2 + H 2 O + 2 e H O 2 + O H ,

(13) Anode : 2 O H 2 e + + H 2 O + 1 2 O 2 .

Compared with traditional preparation methods, CaO2 prepared by air cathode method can reduce production cost, but it still has the following limitations: (a) The utilization rate of H2O2 and the yield of CaO2 are low. (b) High energy consumption. (c) The separation of CaO2 is difficult and the production cost is increased due to its fine particles and colloids. (d) Additional crushing device is required. (e) CaO2 products obtained are heterogeneous.

Table 1 presents a comparison of advantages and disadvantages between different preparation methods.

Table 1

Comparison of advantages and disadvantages between different preparation methods

Advantages Disadvantages
CaCl2 method
  1. Simple process

  2. Low cost

  3. Mature technology

  4. Suitable for small-scale production

  1. Low utilization of H2O2

  2. High energy consumption

  3. The output is low

CaO method
  1. Raw materials are cheap

  2. The preparation process is simple

  3. The equipment is simple

  4. There is no discharge of three wastes

Not exactly mentioned
Traditional method
  1. Process simplification

  2. Energy consumption reduction

  3. Reduction of production cost

  4. Technical maturity

The output is low
Spray drying method
  1. Raw materials are cheap

  2. No need for separation and purification

  3. Saves time and cost

  4. Increased H2O2 utilization

  5. Increase in output

  6. Saves energy

  1. Large investment in equipment

  2. The pipeline is easy to be blocked

  3. High energy consumption

  4. Explosive

Air cathode method Reduction in production cost
  1. Low utilization of H2O2

  2. The output is low

  3. High energy consumption

  4. The difficulty of separation increases the cost

  5. Need crushing equipment

  6. The product is heterogeneous

2.2 Surface modifiers of CaO2

CaO2 NPs prepared by traditional methods often has some phenomena such as different particle size, different morphology, and poor stability. Therefore, surface modifiers are often used to regulate the particle size, morphology, and stability of CaO2 NPs. In addition, some surface modifiers can endow CaO2 NPs with different functions.

Common surface modifiers include PEG, PVP, CO-520, HA, SH, TA, polyacrylic acid (PAA), ethylenediamine tetra acetic acid disodium salt (EDTA·2Na), sodium citrate (SC), polydopamine (PDA) etc., among which PEG is the most common method [23,25,26,27].

Zheng [28] prepared CaO2 NPs without adding surface modifier and with adding surface modifier, and the results are as follows: 1) Pure CaO2 NPs: the particle size is different and the dispersion is poor, showing agglomerate morphology; 2) PDA-CaO2 NPs: the particle size is different, about ∼ 150 nm, and the agglomeration is serious; 3) PDA-SC-CaO2 NPs: the particle size is small, about 30 nm, with a filamentous structure; 4) SC-CaO2 NPs: uniform particle size, monodisperse spherical structure, and good dispersibility; 5) PAA-CaO2 NPs: the particle size is different, and the phenomenon of agglomeration is serious; 6) EDTA·2Na-CaO2 NPs: different particle size, showing cross-linked reticular structure.

Park et al. [29] successfully synthesized/CaO2 NPs by introducing TA in the process of CaO2 NPs synthesis. Figure 2A shows the particle size distribution and scanning electron microscope (SEM) images of three kinds of TA/CaO2 NPs. It can be seen from the diagram that the three kinds of TA/CaO2 NPs have the problems of relatively uniform particle size, relatively regular morphology, and poor dispersion.

Figure 2 
                  (A) Particle size distributions of (a) TA (10 mg)/CaO2, (b) TA (25 mg)/CaO2, and (c) TA (50 mg)/CaO2. SEM images of (d) TA (10 mg)/CaO2, (e) TA (25 mg)/CaO2, and (f) TA (50 mg)/CaO2 [29]. (B–D) The TEM of PEG/CaO2 NPs [21,30,31]. (E) The TEM of SH-CaO2 NPs [15]. (F) TEM images of PVP/CaO2 spherical aggregates synthesized in the presence of CaCl2 at different concentrations: (a) 2.1, (b) 4.2, (c) 8.4, (d) 25.2, (e) 42, and (f) 168 × 10−3 M, respectively [32]. The inset in (a) shows an image at a higher magnification [32]. (G) TEM images of PVP/CaO2 spherical aggregates synthesized in the presence of PVP at different concentrations: (a) 43.2, (b) 30.9, (c) 21.6, (d) 6.17, (e) 1.23, and (f) 0 mg·mL−1, respectively [32]. The inset in (a) shows an image at a higher magnification [32]. (A) Reproduced with permission from a previous study [29]. © 2020 by the authors. Licensee MDPI, Basel, Switzerland. (B) Reproduced with permission from a previous study [21]. Copyright © 2011 Elsevier B.V. All rights reserved. (C) Reproduced with permission from a previous study [30]. © 2021 Elsevier B.V. All rights reserved. (D) Reproduced with permission from a previous study [31]. © 2017 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim. (E) Reproduced with permission from a previous study [15]. © 2019 Elsevier Inc. (F and G) Reproduced with permission from a previous study [32]. © 2019 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 2

(A) Particle size distributions of (a) TA (10 mg)/CaO2, (b) TA (25 mg)/CaO2, and (c) TA (50 mg)/CaO2. SEM images of (d) TA (10 mg)/CaO2, (e) TA (25 mg)/CaO2, and (f) TA (50 mg)/CaO2 [29]. (B–D) The TEM of PEG/CaO2 NPs [21,30,31]. (E) The TEM of SH-CaO2 NPs [15]. (F) TEM images of PVP/CaO2 spherical aggregates synthesized in the presence of CaCl2 at different concentrations: (a) 2.1, (b) 4.2, (c) 8.4, (d) 25.2, (e) 42, and (f) 168 × 10−3 M, respectively [32]. The inset in (a) shows an image at a higher magnification [32]. (G) TEM images of PVP/CaO2 spherical aggregates synthesized in the presence of PVP at different concentrations: (a) 43.2, (b) 30.9, (c) 21.6, (d) 6.17, (e) 1.23, and (f) 0 mg·mL−1, respectively [32]. The inset in (a) shows an image at a higher magnification [32]. (A) Reproduced with permission from a previous study [29]. © 2020 by the authors. Licensee MDPI, Basel, Switzerland. (B) Reproduced with permission from a previous study [21]. Copyright © 2011 Elsevier B.V. All rights reserved. (C) Reproduced with permission from a previous study [30]. © 2021 Elsevier B.V. All rights reserved. (D) Reproduced with permission from a previous study [31]. © 2017 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim. (E) Reproduced with permission from a previous study [15]. © 2019 Elsevier Inc. (F and G) Reproduced with permission from a previous study [32]. © 2019 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim.

Khodaveisi et al. [21], Yin et al. [30], and Liu et al. [31] successfully prepared PEG-modified CaO2 NPs (Figure 2B–D). Transmission electron microscopy (TEM) images showed that although PEG as a surface modifier could alleviate the agglomeration problem of CaO2 NPs in aqueous solution [21], NPs often showed irregular size and morphology. In addition, the method requires a time-consuming washing process to stabilize pH [29].

He et al. [26] successfully prepared CaO2 NPs modified by CO-520 by reverse microemulsion method. The introduction of CO-520 gives CaO2 NPs a good dispersion, but it still faces the problems of different size and irregular morphology.

Shen et al. [32] successfully synthesized CaO2 nanocrystals and spherical aggregates with controllable size and uniform morphology by adjusting the concentration of CaCl2 and PVP on the basis of solving the CaO2 NPs agglomeration problem (Figure 2F and G). Most of all, CaO2 NPs modified by PVP should be dispersed in ethanol for preservation after synthesis, as drying might promote CaO2 NPs irreversible aggregation [32].

Zhang et al. [15] successfully prepared SH-CaO2 NPs by using SH as surface modifiers (Figure 2E). In addition, Han et al. [25] successfully prepared HA-CaO2 NPs by using HA as surface modifiers. The introduction of SH and HA not only gives CaO2 NPs the advantages of good dispersion, uniform size, and uniform morphology but also gives it the function of active targeting to improve its enrichment rate in tumor tissue.

Table 2 describes the effects of the presence or absence of surface modifiers on the morphology, particle size, and stability of CaO2 NPs, and the potential application of partially modified NPs to cancer cells.

Table 2

Comparison of the differences between CaO2 NPs after different modification

NPs Surface modification Appearance Particle size Stability Other
CaO2 None Inhomogeneity Inhomogeneity Poor None
CaO2 PDA Inhomogeneity About ∼150 nm Poor PTT
CaO2 PDA, SC Filamentous structure About ∼30 nm Relatively poor PTT
CaO2 SC Spherical type Homogeneity Good None
CaO2 PAA Inhomogeneity Inhomogeneity Poor Responsive drug release
CaO2 EDTA·2Na Cross-linked reticular structure Inhomogeneity Relatively poor None
CaO2 TA Comparative homogeneity Comparative homogeneity Poor Improves the catalytic efficiency of Fenton reaction
CaO2 PEG Irregular Inhomogeneity Good Prolonging the half-life of blood circulation of NPs
CaO2 CO-520 Irregular Inhomogeneity Good None
CaO2 PVP Regular Homogeneity Good None
CaO2 SH Regular Homogeneity Good Active targeting
CaO2 HA Regular Homogeneity Good Active targeting

3 Application of CaO2 in tumor therapy

CaO2 as a promising anticancer material, introduced into the nanodrugs delivery system might regulate the TME. The construction of a multi-functional nano-therapy platform based on CaO2 NPs can achieve a single treatment for tumors, and even achieve a combined antitumor effect. This chapter mainly focuses on several anticancer therapies based on CaO2 NPs.

3.1 O2-dependent anticancer therapy

3.1.1 Enhanced chemotherapy

Chemotherapy refers to the use of highly cytotoxic chemotherapeutic drugs to interfere with the proliferation of cancer cells and then kill cancer cells to achieve the purpose of tumor treatment, independent of external energy, and also suitable for the treatment of deep tumors that are difficult to reach by laser.

The permeability of nanodrugs in the tumor site is an important factor affecting the effectiveness of chemotherapy. The poor permeability of nanodrugs in tumor site is mainly affected by TME, including heterogeneous blood supply, interstitial fluid pressure (IFP), and extracellular matrix (ECM) [33]. First of all, because the heterogeneous blood supply of tumor tissue is mainly distributed around the tissue, coupled with the high oxygen consumption of tumor cells, hypoxia has become a prominent feature of advanced solid tumors. Compared with normal physiological blood vessels, hypoxia up-regulates hypoxia inducible factor-1alpha (HIF-1a) and multidrug resistance protein 2 (MRP2), inducing immunosuppression and immune escape of tumor cells, resulting in multidrug resistance and rendering tumor-related chemotherapy drugs ineffective [34,35,36]. Therefore, hypoxia poses a great challenge to O2-dependent chemotherapy. Second, in tumor tissue, IFP increases gradually from outside to inside, which hinders the further spread of nanodrugs to deep tumors. In addition, ECM further hinders the penetration of nanodrugs in the tumor site, due to the fact that collagen is the main component of tumor tissue and is expressed in a HIF-1a-dependent manner. In the hypoxic environment of tumor tissue, collagen deposition increases the density of ECM, which affects the penetration of nanodrugs in the tumor site. This section mainly discusses how to improve the antitumor effect of chemotherapy drugs from the point of view of hypoxia and ECM.

CaO2 has more advantages than other O2 producing materials, and its biocompatibility and biodegradability are better.

CaO2 can not only produce O2 in situ but also provide reaction substrate for other O2-producing materials (for example, provide raw material for manganese dioxide (MnO2) to produce O2 under acidic TME), as shown in equation (14) [37].

(14) Mn O 2 + H 2 O 2 + 2 H + Mn 2 + + 2 H 2 O + O 2 .

In situ O2 production of CaO2 can regulate TME, downregulate HIF-1a and MRP2, reverse multidrug resistance of tumor cells, overcome hypoxia-induced chemotherapy limitation etc., and improve the therapeutic effect of chemotherapy drugs. Zhang et al. [38] constructed an intelligent O2 nanocarrier coated with solid lipid monostearin (MS) for CaO2/MnO2 to comprehensively optimize doxorubicin (DOX) transport and enhance chemotherapy effect. Due to the high expression of lipase in tumor cells, NPs are passively accumulated in tumor tissues through enhanced permeability and retention (EPR) effect and ingested by tumor cells [39]. The MS shell of NPs is degraded by lipase, exposing core NPs (CaO2/MnO2) and releasing O2 and DOX. After cancer cells death, the incomplete reaction core NPs are released and further reacted in the ECM to continue to release O2 and DOX. O2 reduces the collagen deposition in ECM through the down regulation of HIF-1a, improves the permeability of DOX, makes O2 and DOX molecules easy to spread to the deep part of tumor, and induces further death of cancer cells.

Considering the complexity, heterogeneity, and diversity of tumor tissue, the effect of single therapy is effective after all, so He et al. [40] prepared a kind of nanocarrier (DOX-CaO2-Fe/MS) with self-supply of O2 and H2O2 to realize the combined antitumor therapy of chemotherapy and CDT. Due to the overexpression of lipase in cancer cells, MS is degraded and iron oleate and DOX-CaO2 are released when nanocarriers accumulate in tumor tissue through EPR effect and are absorbed by cancer cells. Iron oleate releases Fe3+ by hydrolysis. CaO2 reacts with water in acidic environment to form H2O2 and release loaded DOX. Subsequently, Fe3+ reacts with H2O2 to produce O2 and Fe2+, the former downregulated the expression of HIF-1a and P-glycoprotein (P-gp) and enhanced the chemotherapy effect of DOX, while the latter continued to react with H2O2 to produce hydroxyl radical (·OH) and enhance the effect of CDT. It has been proved that nanomaterials have excellent anticancer effect and good biosafety through in vivo and in vitro experiments.

Wang et al. [41] used CaO2 NPs as a synergist in cooperation with transcatheter arterial chemoembolization (TACE) to improve its antitumor effect. After intratumoral injection, CaO2 NPs react with H2O to form O2, H2O2, hydroxyl ions (OH), and Ca2+, to reshape TME and induce calcium overload. In addition, due to the regulation of TME, the therapeutic effect of chemotherapy drugs in TACE has been effectively improved. The results in vitro and in vivo showed that the synergistic antitumor effect of CaO2 NPs + TACE group was significant.

Zhang et al. [42] constructed a nanohybrid material (CaO2@FePt-DOX@PDA@CM) camouflaged by the membrane of 4T1 cancer cells, which enhances the therapeutic effect of cancer by cooperating with chemotherapy, CDT, Ca2+ overload, and PTT. 4T1 cancer cell membrane has the ability of immune escape and homologous targeting to tumor cells. After the nanohybrid material was internalized by cancer cells, the O2-producing characteristics of CaO2 were used to alleviate the hypoxia of tumor cells, downregulate the expression of HIF-1a and P-gp, and enhance the chemotherapeutic effect of DOX, the H2O2-producing characteristics of CaO2 were used to enhance the CDT effect of FePt, and the Ca2+-releasing characteristics of CaO2 were used to induce mitochondrial damage, upregulate Cytochrome C and Caspace-3, and achieve calcium overload. In addition, the existence of PDA gives nanohybrid materials excellent photothermal conversion properties, which can enhance the efficiency of Fenton reaction while giving full play to PTT. In vitro and in vivo studies have shown that nanohybrid materials can cooperate with a variety of anticancer therapies and effectively improve the therapeutic effect by optimizing individual advantages.

3.1.2 Enhanced PDT

PDT is one of the effective antitumor therapies approved by the US Food and Drug Administration (FDA), which is a promising antitumor strategy. PDT consists of three basic components: light, photosensitizer (PS), and oxygen source [43]. The antitumor mechanism of PDT is that PS reacts with molecular oxygen under laser irradiation at a specific wavelength to produce efficient singlet oxygen (1O2), which causes irreversible damage to tumor cells and blood vessels and leads to photoinduced tumor cell death.

Figure 3 illustrates the basic principles of PDT, PS change from ground state to instantaneous singlet state (1PS*) under the irradiation of light source. There are two alternative paths for instantaneous singlet states: (1) Transition from instantaneous singlet state to ground state. (2) The transition from instantaneous singlet state to long-lived excited triplet state (3PS*). The excitation triplet also includes type I and type II reaction processes: (1) Type I: The cation or anion radical is formed by electron or proton transfer between 3PS* and substrate. Subsequently, these free radicals react with molecules such as H2O or O2 to form cytotoxic ROS (superoxide anion (O2 ·–), ·OH, and H2O2) [44]. (2) Type II: Excited triplet transfers energy directly to oxygen molecules, producing highly active 1O2 [45]. Type I and Type II photochemical reactions can occur simultaneously, and the ratio between the two is determined by the type of PS and the concentration of O2 [23,46]. According to the latest research progress, compared with O2-dependent type II PDT, type I PDT is not O2-dependent, ROS is more toxic, and has a broader prospect in the treatment of hypoxic tumors [44,47]. However, for most PDT photosensitizers, type II photochemical processes dominate [48]. Therefore, this chapter mainly discusses the application of CaO2 NPs in enhancing O2-dependent type II PDT.

Figure 3 
                     PDT and CaO2 enhanced PDT basic schematic diagram.
Figure 3

PDT and CaO2 enhanced PDT basic schematic diagram.

However, whether PDT effect can play effectively is mainly limited by O2. The consumption of O2 in tumor cells by PDT will further aggravate the hypoxia of tumor cells which will lead to further weakening of PDT treatment effect or even treatment failure and tumor cells metastasis which form a vicious cycle. The effective way to improve the PDT effect of O2 consumption is to establish an O2 supply system to reduce O2 consumption and improve the utilization rate of O2. CaO2 NPs as an O2 supply system is a promising material for regulating acid TME and enhancing PDT. After introducing CaO2 NPs into tumor cells, O2 generated through disproportionation reaction with H2O compensated for the O2 consumed during PDT treatment and promoted the effective play of PDT effect (Figure 3).

Sun et al. [49] loaded CaO2 NPs (PCN-224-CaO2-HA) modified by HA on the basis of porphyrin metal–organic frameworks (MOFs) to target and enhance PDT (Figure 4). Among them, HA can target tumor cells with overexpression of CD44+ receptor, while protecting CaO2 NPs. PCN-224-CaO2-HA releases PS and O2 after targeting and uptake by tumor cells. Subsequently, PS was irradiated by 650 nm laser to achieve enhanced PDT. In vitro and in vivo experiments showed that PCN-224-CaO2-HA could target tumor cells and enhance PDT.

Figure 4 
                     Schematic representation of PCN-224 combined with HA-modified CaO2 NPs for enhanced PDT [49]. Reproduced with permission from a previous study [49]. © The Royal Society of Chemistry.
Figure 4

Schematic representation of PCN-224 combined with HA-modified CaO2 NPs for enhanced PDT [49]. Reproduced with permission from a previous study [49]. © The Royal Society of Chemistry.

In order to accelerate the decomposition of H2O2 into O2, Ren et al. [50] designed and synthesized a cascade of catalytically enhanced UIO@Ca-Pt NPs. Among them, the multi-nanometer MOFs material modified by porphyrin PS can effectively avoid the accumulation of PS. Under acidic TME, CaO2 reacts with H2O to produce H2O2, which can be decomposed into H2O and O2 due to its instability, but its decomposition rate is slow. While the introduced PT-NPs can be used as an effective catalyst to accelerate the decomposition of H2O2 into O2. As a result, the NPs greatly improved the antitumor efficacy of PDT.

Although the above cases partly overcome the problem of insufficient oxygen source of PDT, the consumption of ROS by the highly expressed glutathione (GSH) in tumor cells still limits the antitumor effect of PDT. Therefore, the effect of single treatment is limited after all, and combined antitumor therapy is necessary.

Gulzar et al. [51] constructed a multi-mode image-guided nanocomposite CaO2-MnO2-UCNPs-Ce6/DOX (Ca-Mn-NUC). When the nanocomposite is endocytosed by tumor cells, CaO2 generates O2 in situ under acidic TME, providing sufficient oxygen source for PDT. After MnO2 converts GSH to oxidized glutathione (GSSH), it can reduce the consumption of ROS, further enhance the oxidative stress of tumor cells, and finally enhance PDT. In addition, the O2 generated can also improve chemotherapy, and ultimately achieve chemotherapy/PDT combined treatment, which is highly anticancer. The nanocomposite can also provide computed tomography (CT) and magnetic resonance imaging (MRI) dual-mode imaging for real-time monitoring of anticancer effects.

Chen et al. [52] developed a TME-responsive CaO2-based nanosystem (CF@CO@HC) using a bottom-up approach, which has the ability of GSH consumption and ROS self-amplification, and can cooperate with PDT and CDT for cancer treatment. The nanosystem consists of CaO2 (CaO2-FM) doped with PS as core, hybrid silicone skeleton (Cu-ONS), and local hydrophobic cage (HC) as shell. After the nanosystem was internalized by cancer cells, the hybrid organosilicon skeleton was cleaved under the reduction of high concentration of GSH to release Cu+. CaO2 reacts with H2O in acidic environment to form H2O2 and O2, and releases the doped PS derivative 4-FM. Subsequently, the generation of H2O2 enhances Cu+-mediated CDT. The generation of O2 enhances 4-FM-mediated PDT. In addition, GSH consumption further increases the production of ROS. Therefore, the nanosystem provides a strategy to increase revenue and reduce expenditure for cancer treatment, and effectively enhances cancer treatment.

In addition, Yan et al. [53] designed and synthesized a multistimulus TME response nanoplatform composed of MCMnH NPs and CaO2 NPs to improve the antitumor effect of the guidance of MRI and photothermal imaging. HA can target tumor cells with overexpression of CD44+ receptor and induce tumor cells to uptake MCMn NPs. Subsequently, MNP realizes PTT under the irradiation of 808 nm laser, and PDT is realized through energy transfer between PTT and Ce6. MnO2 produces Mn2+ and O2 through a series of reactions in tumor cells. The former reacts with H2O2 to form ·OH through Fenton-like reaction to achieve CDT. CaO2 NPs produces H2O2 and O2 through a series of reactions in tumor cells, and the former enhances CDT. O2 generated by MnO2 and CaO2 NPs downregulates HIF-1a and enhances PDT. In vivo and in vitro experiments show that the nanoplatform has significant antitumor effect through synergistic promotion.

3.2 H2O2-dependent anticancer therapy

3.2.1 Enhanced CDT

CDT is a novel and efficient antitumor technology based on ROS and its basic principle is delivering Fenton reaction or type of Fenton-like reaction catalyst using nanotechnology to tumor tissue, under the acidic TME, making H2O2 to decompose in the tumor tissue, producing ROS with strong oxidation capacity (·OH), inducing tumor cell apoptosis and killing cancer cells [54,55]. With the gradual deepening of CDT research, various Fenton or Fenton-like reaction catalysts emerged, including Fe2+, Fe3+, Cu2+, Mn2+, Mo4+, Cr4+, V2+, Ti3+, Graphene oxide, etc. [56,57,58]. Among them, Fe2+ and Fe3+ are typical Fenton and Fenton-like catalysts. The reaction equation is as follows:

(15) Ca O 2 + H 2 O Ca ( OH ) 2 + 1 / 2 O 2 ,

(16) Ca O 2 + 2 H 2 O Ca ( OH ) 2 + H 2 O 2 ,

(17) Fe 2 + + H 2 O 2 Fe 3 + + OH + ·OH ,

(18) Fe 3 + + H 2 O 2 Fe 2 + + H + + HO 2 · .

CDT works independent of O2 and external energy compared to other cancer treatments [59,60]. Therefore, CDT is more suitable for deep tumors with hypoxia or difficult to reach by laser.

At present, antitumor research on CDT is still in the initial stage. However, there are many factors affecting the efficiency of CDT. In order to improve the therapeutic effect of CDT, researchers mainly proposed the following methods: Change TME (increase H2O2 concentration, decrease GSH level, decrease pH value) and select suitable catalyst and combine with other anticancer therapy effectively. In addition, in order to improve the efficacy of CDT, current studies mostly focus on the catalytic activity of nanomaterials, but ignore their biocompatibility and biodegradation.

CaO2 NPs have good biocompatibility and biodegradability. CaO2 NPs and H2O generate H2O2 and O2 through disproportionation reaction (equations (15) and (16)), in which the former provides sufficient substrates for subsequent CDT, thus ensuring the effective play of CDT effect and enhancing the antitumor effect of CDT. The latter regulates TME and reverses hypoxia in tumor cells [61,62,63]. Therefore, considering the advantages of CaO2 NPs, it is particularly important to construct a nano-therapy platform based on CaO2 NPs.

Due to the limited H2O2 concentration (∼100 µM) in tumor cells, ROS production is difficult to maintain. Based on this, Han et al. [25] prepared a H2O2 self-supporting nanotubule platform CaO2-Fe3O4@HA-Cy7 NPs for guiding targeted and imaging CDT. The HA can target the tumor cell receptor over-expressed by CD44+ and reach the cell through endocytosis, which is hydrolyzed by hyaluronidase in the cytoplasm and releases near-infrared (NIR) fluorescent group markers, CaO2, and Fe3O4. Through further reaction, the H2O2 consumed by CDT can be compensated and the therapeutic effect of CDT can be enhanced. NIR fluorescence and magnetic resonance bimodal imaging provide timely CDT therapeutic effect of NPs in vivo. In vivo antitumor studies have shown that CaO2-Fe3O4@HA-Cy7 NPs have excellent tumor specificity, enhanced CDT efficacy, and good biocompatibility.

Mamat et al. [64] constructed CaO2/Fe3O4 nanocomposites coated with oleyl-amine (OA). The NPs inhibited the premature reaction of CaO2, overcame the deficiency of H2O2 in tumor tissues, and realized the non-oxidative generation of ROS and efficient CDT. The results show that the nanocomposites material has significant tumor growth inhibition ability and good biocompatibility, and has great clinical transformation value in tumor therapy.

GSH is another important factor that limits the efficacy of CDT therapy. Therefore, reducing GSH expression level in cancer cells to reduce ROS consumption is an effective way to improve the efficiency of CDT treatment. In addition, the commonly used Fenton and Fenton-like reaction catalysts (Fe2+ and Fe3+) have high catalytic efficiency only under strong acidic conditions (pH 2–4), maximizing the therapeutic effect of CDT, on the contrary, the catalytic efficiency is relatively low under neutral and weak acidic conditions. Therefore, Kong et al. [65] explored a new combination of antitumor therapy by designing and synthesizing a Cu-ferrocene modified CaO2 NPs (CaO2/Cu-ferrocene). In terms of CDT enhancement, this scheme not only solves the problem of insufficient H2O2 in tumor cells, but also overcomes the problems of high expression of GSH and low efficiency of Fenton and Fenton-like reaction catalysts in tumor cells. The NPs remain stable under neutral conditions and rapidly release H202 under acidic TME. Among them, H2O2 and ferrocene generate ·OH through Fenton reaction, which achieves CDT, while Cu2+ reacts with GSH to generate GSSH and Cu+, which reduces the consumption of ROS by GSH and improves the catalytic efficiency of Fenton and Fenton-like reaction catalysts, thus greatly improving the therapeutic effect of CDT. In vitro and in vivo experimental results showed that the consumption of GSH and the production of ·OH significantly enhanced the therapeutic effect of CDT.

In addition, in order to improve the killing effect of NPs on tumor cells, the combined antitumor application based on CDT has also been reported one after another [66,67,68]. Compared with single anticancer therapy, two or more CDT-based treatment modes can achieve better anticancer effects.

The combination of chemotherapeutic drugs and ROS for tumor therapy is also an effective treatment strategy to improve the efficiency of CDT [69,70,71,72]. Gao et al. [73] designed and synthesized a self-supplied O2/H2O2 nanocatalysis drug CaO2@DOX@ZIF-67, and constructed a pH responsive CDT/chemotherapy combined antitumor treatment platform (Figure 5a). The NPs are taken up by acid lysosomes after entering the tumor cells. Subsequently, the MOFs decomposes and releases Co2+, CaO2 and DOX. The Co2+ reacts with H2O2 to achieve CDT, and DOX to achieve chemotherapy. In addition, CaO2 further reacts to generate H2O2 and O2 to improve the TME, reverse hypoxia of tumor cells, avoid multidrug resistance of cancer cells, enhance CDT and chemotherapy, and enhance the combination therapy of the two. The experimental results show that CaO2@DOX@ZIF-67 had good antitumor activity. Therefore, the nanocomposite is expected to be a candidate for pH-responsive chemotherapy/CDT combined antitumor therapy, and has great clinical transformation value.

Figure 5 
                     CDT-based anticancer therapies: (a) Combined treatment of CDT and chemotherapy [73]. (b) Combined treatment of CDT and ST [76]. (c) Combined treatment of CDT and PDT [77]. (a) Reproduced with permission from a previous study [73]. © 2019 The Authors. Published by WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim. (b) Reproduced with permission from a previous study [76]. © 2021 Elsevier Ltd. All rights reserved. (c) Reproduced with permission from a previous study [77].
Figure 5

CDT-based anticancer therapies: (a) Combined treatment of CDT and chemotherapy [73]. (b) Combined treatment of CDT and ST [76]. (c) Combined treatment of CDT and PDT [77]. (a) Reproduced with permission from a previous study [73]. © 2019 The Authors. Published by WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim. (b) Reproduced with permission from a previous study [76]. © 2021 Elsevier Ltd. All rights reserved. (c) Reproduced with permission from a previous study [77].

Starvation therapy (ST) is to starve the tumor cells by cutting off the blood supply to the tumor tissue [74] or cutting off the nutrition source of the tumor cells [75]. Glucose is the energy source for the growth and proliferation of tumor cells. Reducing the concentration of glucose in tumor cells can effectively cut off the energy supply of cancer cells and inhibit or kill cancer cells. Glucose oxidase (GOx) is an endogenous redox enzyme that specifically oxidizes glucose to gluconic acid and H2O2. In recent years, GOx has been widely used in the treatment of tumor starvation, but its anticancer effect is limited by the hypoxic condition of tumor. Based on this, Zhang et al. [76] coated CaO2@Fe(OH)3 nanometer composites and GOx in bio-compatible liposomes to prepare nanometer composite drugs, and their synthesis steps and antitumor mechanism are shown in Figure 5b. When the NPs entered the tumor cells, the degradation of CaO2@Fe(OH)3 nanocomposites were triggered in acidic TME to generate H2O2 and Fe3+. Second, H2O2 and Fe3+ react to generate O2 and Fe2+. The former can reverse the problem of tumor hypoxia, make up for the O2 consumed by GOx in the process of oxidizing glucose, cut off the energy material (glucose) of tumor tissue, realize the purpose of starving tumor cells, and strengthen ST, the latter triggers Fenton reaction to produce ·OH with strong oxidation capacity, which realizes CDT. In addition, the products gluconic acid and H2O2 obtained by GOx in the process of glucose oxidation reduce the pH value of tumor tissue, provide the best reaction conditions for CDT, and further supplement its required substrate (H2O2) and improve the effect of CDT. The experimental results confirmed that the anticancer effect of CDT/ST combination was significant.

PDT is an anticancer therapy based on ROS. Compared with CDT, both of them have the common goal of achieving tumor therapy by means of ROS. Liu et al. [77] designed a thermosensitive nano-system of H2O2/O2 MSNs@CaO2-ICG@LA for CDT/PDT combination therapy of tumor cells (Figure 5c). The NPs were prepared by loading CaO2 NPs and indocyanine green (ICG) with manganese silicate nanoparticles (MSNs) and further modifying lauric acid (LA) on their surface. When the NPs reached the tumor cells through the endocytic pathway, ICG generated heat and 1O2 under the irradiation of 808 nm laser, the former caused the phase transformation of LA (from solid phase to liquid phase), and the latter realized PDT. In addition, exposed CaO2 NPs further react with H2O to generate H2O2 and O2. The production of H2O2 provides sufficient substrates for subsequent CDT and enhances CDT. The production of O2 alleviates tumor hypoxia and enhances PDT. MSNs are gradually exposed with further consumption of CaO2 NPs, and react with GSH in cancer cells to release a large amount of Fenton-like reaction catalyst Mn2+, which is used to self-supply CDT of H2O2. The consumption of GSH by MSNs further promotes the continuous production of ROS in cancer cells. In conclusion, the team utilized PDT/CDT combined anticancer therapy to increase ROS sources in tumor cells while maintaining ROS production in tumor cells through GSH consumption. Experimental results in vivo and in vitro show that the strategy has remarkable anticancer effect and can effectively inhibit the growth of cancer cells.

3.3 Calcium overload

Calcium overload belongs to ion interference therapy, which is a new and efficient antitumor technology and has attracted more and more attention in recent years [17,78]. Calcium overload refers to the dysfunction of calcium balance system and disorder of calcium distribution under the action of some factors, resulting in abnormal increase in intracellular calcium concentration. Calcium overload has been linked to tumor death, but the root cause of cancer cell death remains unclear. Therefore, cancer cell death caused by calcium overload may be related to the following factors or the result of a combination of factors:

  1. Induction of tumor cell death by activating mitochondrial apoptosis pathway [15,79,80,81,82,83].

  2. It is related to macroscopic calcification induced by microscopic calcium overload, leading to the death of cancer cells [15] (it is worth noting that tumor cell calcification can assist CT imaging and monitor the treatment process [84]).

  3. It is related to interfering the process of mitochondrial oxidative phosphorylation, damaging mitochondria and inactivating phospholipase in cytoplasm, thus causing irreversible damage to tumor cells [85].

Ca2+ are mainly stored in organelles such as cytoplasm, mitochondria, and endoplasmic reticulum, and are endogenous substances necessary for the maintenance of organisms in vivo. As one of the second messengers, Ca2+ play an important regulatory role in the proliferation and migration of tumor cells [86,87]. As tumor cells are characterized by abnormal differentiation and proliferation, reduced apoptosis, and high metastasis, low concentration of Ca2+ in the cytoplasm is a necessary condition to avoid apoptosis of tumor cells when cancer occurs [88,89]. As a result, improving the Ca2+ concentration in the cytoplasm of tumor is an effective treatment strategy for antitumor. Unfortunately, due to the long neglect of Ca2+-induced cell damage, there have been few reports on the antitumor application of calcium overload when designing anticancer drugs based on calcium nanomaterials. This section introduces the strategy of calcium overload in the treatment of tumors [30].

Zhang et al. [15] designed and synthesized a kind of ultra-small SH-CaO2 NPs by taking advantage of the special cellular biological effect of Ca2+ and its enhancement effect on H2O2 oxidative stress, and systematically studied the anticancer effect induced by Ca2+. Due to the pH responsiveness of CaO2 NPs, under acidic TME, CaO2 slowly decomposes into free Ca2+ and H2O2, leading to calcium overload and oxidative stress, respectively. In addition, the low expression of catalase (CAT) in tumor cells [90] and continuous oxidative stress can lead to the functional disorder of protein, trigger desensitization of calcium-related channels, and lead to uncontrolled accumulation of Ca2+ in tumor cells [91]. The anticancer mechanism of the nanocomposite is shown in Figure 6a. The experimental results showed that the survival rate of tumor cells decreased with the increase in SH-CaO2 NPs concentration, (Figure 6b). The significant killing effect of SH-CaO2 NPs was not limited by tumor type, and it could effectively kill multiple types of cancer cells (Figure 6c). In addition, Alizarin Bordeaux staining and von Kossa staining could clearly identify the calcification area caused by calcium overload in Figure 6d. According to the above results, SH-CaO2 NPs have significant anticancer effect.

Figure 6 
                  (a) Schematic representation of the functional pattern of SH-CaO2 NPs. (b) Cell viability assay in 4T1 cells incubated with CaCl2 or SH-CaO2 NPs. (c) Cell viability assay in human tumor cells (HeLa, A549, and MB231) or normal cells (LO2) incubated with SH-CaO2 NPs. Normal cells were more tolerant to SH-CaO2 NPs. (d) Identification of the products of exocytosis in vitro with Alizarin Bordeaux staining and von Kossa staining, which show the calcified areas in red and black, respectively [15]. Reproduced with permission from a previous study [15]. © 2019 Elsevier Inc.
Figure 6

(a) Schematic representation of the functional pattern of SH-CaO2 NPs. (b) Cell viability assay in 4T1 cells incubated with CaCl2 or SH-CaO2 NPs. (c) Cell viability assay in human tumor cells (HeLa, A549, and MB231) or normal cells (LO2) incubated with SH-CaO2 NPs. Normal cells were more tolerant to SH-CaO2 NPs. (d) Identification of the products of exocytosis in vitro with Alizarin Bordeaux staining and von Kossa staining, which show the calcified areas in red and black, respectively [15]. Reproduced with permission from a previous study [15]. © 2019 Elsevier Inc.

However, compared with most solid tumors, the antitumor effects of monotherapy are limited. Therefore, researchers have successively developed combined therapies based on calcium overload to improve the therapeutic effect of tumors.

Liu et al. [84] used dendritic mesoporous organosilica (DMOS) bound by tetra-sulfide bonds as nanometer carriers, loaded with chloroperoxidase (CPO) and sodium-hyaluronate-modified CaO2 NPs (CaO2-HA NPs), and improved the stability of CPO and the limited H2O2 level in tumor cells. A novel combined treatment strategy of hydrogen sulfide (H2S) gas, enzyme dynamic therapy (EDT), and Ca2+ interference was developed, as shown in Figure 7A. The NPs are enriched in tumor cells through enhanced EPR effect and specific targeting effect of HA on tumor cells. Subsequently, HA is degraded by hyaluronidase and CaO2 further reacts with H2O to generate H2O2 and Ca2+. In addition, DMOS have GSH responsiveness, releasing CPO, consuming GSH, and generating H2S gas under high GSH environment in tumor cells, and finally realizing multi-modal antitumor therapy under CT imaging. In vitro experiment results confirmed that when DCC-HA NPs were co-cultured with mouse fibroblasts (L929 cells) for a certain period of time, the survival rate of L929 cells remained high even at a high concentration of NPs, indicating that DCC-HA NPs had good biocompatibility (Figure 7B). Subsequently, DCC-HA NPs were co-cultured with mouse 4T1 cells for a period of time. It was found that the survival rate of 4T1 cells not only decreased with the increase in NPs concentration, but also decreased much more than the survival rate of tumor cells under other conditions (Figure 7C). In vivo results confirmed that 4T1 tumor cells were transplanted into mice and significantly inhibited after 14 days of NPs treatment (Figure 7D). All the above results indicate that the TME responsive nanocomposite has significant anticancer effect.

Figure 7 
                  (A) ((a) Schematic illustration of the synthesis and (b) antitumor performance of the DCC-HA NPs). (B) The cell viability of L929 cells after incubation with different concentrations of DCC-HA NPs for 24 h (n = 3). (C) Cell cytotoxicity of 4T1 cells under different conditions (n = 3). ***p < 0.001. (D) Digital photographs of tumor-bearing mice after different treatments for 14 days [84]. Reproduced with permission from a previous study [84]. © 2021 Wiley‐VCH GmbH.
Figure 7

(A) ((a) Schematic illustration of the synthesis and (b) antitumor performance of the DCC-HA NPs). (B) The cell viability of L929 cells after incubation with different concentrations of DCC-HA NPs for 24 h (n = 3). (C) Cell cytotoxicity of 4T1 cells under different conditions (n = 3). ***p < 0.001. (D) Digital photographs of tumor-bearing mice after different treatments for 14 days [84]. Reproduced with permission from a previous study [84]. © 2021 Wiley‐VCH GmbH.

Jiang et al. [92] constructed a nanoplatform containing UCNPs-Ce6@RuR@mSiO2@PL-HA NPs (UCRSPH) and SA-CaO2 NPs. After entering tumor cells, the composite nanometer platform was irradiated by NIR (980 nm) to alter TME, reverse tumor hypoxia, and achieve Ca2+ interference therapy and PDT self-enhancement. In addition, the nanoplatform provides fluorescence imaging and CT imaging to further guide in vivo antitumor therapy. In conclusion, the nanometer platform shows a potent combined anticancer potential with minimal toxic side effects on normal cells.

Chen et al. [93] designed and synthesized a new CaO2@TA-FEIII nanodrug delivery system for enhancing CDT. Its anticancer mechanism is: TA and FeIII form TA–Fe nanocoating on the surface of spherical CaO2 nanoaggregate. When the nanocomposite enters tumor cells, H2O2 is generated through degradation, which solves the problem of insufficient H2O2 in CDT. In addition, TA of nanomaterials reduced FeIII to FeII, improved the catalytic efficiency of Fenton reaction, generated ·OH, induced irreversible oxidative damage to tumor cells, promoted calcium overload, and accelerated the death of tumor cells. The experimental results in vitro and in vivo indicate that the nanocomposite is a promising novel and highly effective antitumor nanoplatform with good tumor treatment effect.

In addition to the anticancer therapies mentioned above, other anticancer therapies based on CaO2 NPs are shown in Table 3.

Table 3

Application of hybrid nanomaterials based on CaO2 NPs in tumor field

Synthesized NPs Applications Year Ref.
PLLA/CaO2@ZIF-67 CDT 2023 [94]
CaO2-MNPs Chemotherapy 2021 [95]
CaO2@Mn-PDA NPs PTT/CDT 2021 [96]
CFM Radiotherapy/CDT 2020 [97]
CaO2@ZIF-8@MPN CDT/Ca2+ overload 2021 [17]
CaO2/DOX@Cu/ZIF-8@HA CDT/chemotherapy 2021 [98]
DOX-CaO2-Fe/MS NPs CDT/chemotherapy 2021 [40]
CaO2@ZIF-8@THPP PDT/ion-interference 2023 [99]
CaO2@Co-ferrocene Calcium-overload/CDT 2021 [100]
CaO2@ZIF-Fe/Ce6@PEG CDT/PDT/Ca2+ overload 2021 [101]
CaO2@Au nano-shells CDT/PTT/chemotherapy 2022 [102]
Fe-GA/CaO2@PCM Calcium-overload/CDT/PTT 2020 [103]
RB/CaO2 (RBC NPs) Immune-mediated therapeutic/SDT 2021 [104]
CaO2/Cu-ferrocene Ca2+/ROS synergistic tumor therapy 2021 [9]
CaO2@CuS-MnO2@HA PTT/PDT/Ca2+ overload/immunotherapy 2022 [105]
CaO2/TF/CUR Calcium-overload/apoptosis of iron/chemotherapy 2021 [30]
CaO2@HMSNs-PAA Inducing apoptosis of cancer cells by oxidative stress 2020 [106]

4 Antimicrobial therapy

Bacterial infection has become one of the main problems threatening human health [107,108,109]. CaO2 NPs have attracted much attention due to their small size, increased surface contact area with bacteria and damage to bacteria [110]. The antibacterial mechanism of CaO2 NPs is related to the production of ROS (1O2, O2 ·–, H2O2 and ·OH). Relevant literature has shown that ROS induced oxidative stress is one of the important antibacterial mechanisms for the construction of ROS, which can effectively treat bacterial infection [111].

ROS treats bacterial infections by destroying cell membranes [112], and lipid peroxidation caused by free radicals is one of the ways to change cell membranes [113]. However, due to the thicker cell wall structure of Gram-positive bacteria, the lipids seen in Staphylococcus aureus (S. aureus) are not affected. In addition, negatively charged ROS cannot easily penetrate negatively charged cell membranes, but H2O2, as a type of ROS, can easily penetrate cell membranes for bacteriostatic purposes [114].

ROS on cell walls are produced by negatively charged cell membranes interacting with positively charged NPs [115]. ROS can damage the cell membrane by interacting with the cell wall of bacteria, inhibit further growth of cells, leak the internal cell components, and finally lead to the death of bacteria. Therefore, it is necessary to develop a new kind of biological material with high antibacterial ability. Shen et al. [32] synthesized CaO2 nanocrystals and spherical aggregates with controllable size and uniform morphology using CaCl2 as precursor by wet chemical process. By changing the concentration of CaCl2 and PVP, the particle size of CaO2 nanocrystals and their spherical aggregates can be adjusted in the range of 15–100 nm. The results of bacteriostatic experiment showed that the spherical aggregates showed obvious size dependent bacteriostatic effect. In addition, CaO2 NPs synthesized with Ca(NO3)2 as the precursor showed similar antibacterial activity (Table 4). It can be seen that the small-sized CaO2 nanocrystals and their spherical aggregates have great potential in antibacterial applications.

Table 4

Minimum inhibitory concentrations (MIC) of CaO2 NPs prepared using Ca(NO3)2 as precursor [32]

Bacteria MIC at pH 7.4 (µg·mL−1) MIC at pH 5.8 (µg·mL−1)
CaCl2 as precursor Ca(NO3)2 as precursor CaCl2 as precursor Ca(NO3)2 as precursor
C. tetani 60 60 60 60
F. nucleatum 60 60 60 60
E. coli 120 120 120 100

Reproduced with permission from a previous study [32]. © 2019 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim.

Note: C. tetani: Clostridium tetani, F. nucleatum: Fusobacterium nucleatum, E. coli: Escherichia coli.

Moreover, Thi et al. [116] developed a new hydrogel preparation method for antimicrobial therapy by promoting in situ cross-linking hydrogel formation mediated by horseradish peroxidase (HRP) using CaO2. General preparation process: Gelatin-hydroxyphenyl propionic acid (GH) polymer, HRP, and CaO2 were used as raw materials, and the GH/C hydrogel was prepared by mixing the polymer solution with different concentrations of HRP and CaO2 solution (volume ratios of GH:HRP:CaO2 = 8:1:1). The introduction of CaO2 can give hydrogel system the following two advantages:

  1. On the premise of not affecting biocompatibility, H2O2 is gradually generated to form hydrogel system.

  2. Dynamic hydrogel matrix was constructed to stimulate surrounding tissues by continuously releasing Ca2+, H2O2, and O2.

Then, the antibacterial activity of GH/C hydrogel was evaluated by measuring the optical density (OD) value of the bacterial suspension at 600 nm. The antibacterial activity of GH/C hydrogel is related to the content of CaO2, and its antibacterial activity increases with the increase in the dosage of CaO2, among which GH/C 0.5 shows the highest antibacterial effect (Figure 8). Therefore, although GH/C hydrogel could not kill bacteria, it could also effectively inhibit the growth rate of E. coli and S. aureus. We believe that the formation of in situ crosslinked hydrogels mediated by HRP/CaO2 has potential application value in the treatment of bacterial infection.

Figure 8 
               
                  In vitro antibacterial activities of GH/C hydrogels depending on CaO2 concentration against (a) E. coli and (b) S. aureus [116].
Figure 8

In vitro antibacterial activities of GH/C hydrogels depending on CaO2 concentration against (a) E. coli and (b) S. aureus [116].

Traditional titanium implants can effectively treat bacterial infection after modification, but the resulting anti-inflammatory response is also ignored. Therefore, He et al. [18] constructed a dual drug loading system (TNT@IL-4/GelMA@CaO2) to treat bacterial infections while inhibiting the occurrence of pro-inflammatory reactions (Figure 9). In vitro and in vivo experiments show that the drug loading system can effectively treat bacterial infection and regulate pro-inflammatory response. In short, the drug loading system provides a good example for the development of advanced titanium implants with anti-inflammatory and antibacterial properties.

Figure 9 
               Schematic illustration of the underlying antibacterial and anti-inflammatory mechanisms of the prepared samples [18]. Reproduced with permission from a previous study [18]. © 2021 Wiley‐VCH GmbH.
Figure 9

Schematic illustration of the underlying antibacterial and anti-inflammatory mechanisms of the prepared samples [18]. Reproduced with permission from a previous study [18]. © 2021 Wiley‐VCH GmbH.

Zhang et al. [117] prepared a hydrogel library (OMCN) with self-activated nitric oxide (NO) release to cooperate with NO and PTT in the fight against bacterial infection. First of all, H2O2 formed by the reaction of CaO2 with H2O can oxidize l-arginine (l-arg) to form NO, which is used to treat bacterial infection. In addition, thanks to the excellent photothermal effect of OMCN, PTT can increase the production of NO gas and cooperate with sterilization at the same time. In vivo and in vitro studies showed that the material had good synergistic antibacterial activity under the laser irradiation of 808 nm, and there was no obvious toxicity. Therefore, the prepared hydrogel library provides a promising strategy for antibacterial in the future.

In addition, Table 5 describes other antibacterial applications based on CaO2 NPs.

Table 5

Application of CaO2 NPs-based hybrid nanomaterials in the field of bacteria

Synthesized NPs Application Year Ref.
Ca@PDAFe-CNO Bacteriostasis 2022 [118]
CaO2@SiO2/CS Bacteriostasis 2022 [119]
CaO2/GQDs@ZIF-67 Bacteriostasis 2021 [16]
BNP@TI Bacteriostasis 2022 [120]
CaO2@PDA Bacteriostasis 2022 [121]
CaO2-TiOx@Ti3C2 Bacteriostasis 2023 [122]
CPA Bacteriostasis 2021 [123]

5 Biological effects and biosafety of CaO2 NPs

Whether CaO2 NPs can achieve further clinical progress is closely related to its biological effects and biosafety.

The biological effects and biosafety of CaO2 NPs have been preliminarily explored in the relevant data [124,125]. The relevant data are encouraging, but there is still a certain distance from further clinical conversion. For example, when CaO2 NPs circulate in the vessel, it will react with H2O in advance, and its product H2O2 is toxic and may affect normal tissues and cells. In the following studies, a more systematic in vitro and in vivo biosafety assessment should be conducted to further reveal the biological distribution and excretion of CaO2 NPs in vivo, providing more powerful evidence for biological effects and biosafety.

6 Conclusion and future perspective

The preparation technology and surface modification of CaO2 NPs have been relatively mature. Although CaO2 NPs have achieved good results in tumor treatment and bacterial infection, there are still some challenges to be overcome in future research. If the following problems are solved, CaO2 NPs are expected to continue to move forward in the field of biomedicine.

  1. There are many factors affecting the effect of chemotherapy, including hypoxia, low pH value and overexpression of GSH etc. [126,127,128,129]. At present, most research are from the perspective of hypoxia, that is, through exogenous supply of O2 and endogenous production of O2 to alleviate the hypoxic environment in tumor cells, reduce MDR and enhance the effect of chemotherapy. However, there are few reports on how to overcome low pH value and overexpression of GSH to enhance the effect of chemotherapy. Therefore, in the future, it is necessary to study how to enhance the effect of chemotherapy from the aspects of low pH value and GSH overexpression.

  2. The effect of PDT is affected by TME, the selection of effective PS and the management of light dose [130,131]. Up to now, great progress and achievements have been made in the study of PS selection and optical quantum management [132,133,134,135]. However, there are still many challenges in cancer treatment that regulates TME to enhance type II PDT.

  3. As we all know, the best pH range to maximize the effect of CDT is 2–4 [136], while the pH value of tumor tissue is about 6.5 [54]. Although GOx and gold nanoparticles (GNPs) are often used to improve pH in tumor tissues, it is necessary to develop pH-independent CDT nanocatalysts. In addition, most of the previously reported CDT nanocatalysts are transition metal-based nanoparticles, which can cause acute inflammation and toxicity in vivo. Therefore, the development of CDT nanocatalyst with high biosafety and good biodegradability is a more feasible scheme for the development of CDT in future.

  4. In a certain range, the smaller the size of NPs, the stronger the penetration of tumor tissue, and the better the therapeutic effect on bacterial infection. In addition, most studies have shown that spherical NPs are slightly less permeable in tumor tissue than other shapes [33]. Therefore, it is necessary to adjust the shape and size of NPs to improve the permeability and antibacterial activity of tumor tissues.

  5. Oxidative stress induced by ROS is one of the effective bacteriostatic mechanisms. CaO2 NPs can self-amplify ROS and effectively treat bacterial infection. Although the bacteriostatic effect is remarkable, the treatment mode is relatively simple. Therefore, it is necessary to use CaO2 NPs to design a nanohybrid material based on CaO2 and to develop a novel bacteriostatic mode with better therapeutic effect by using the products of CaO2 NPs.

In a word, the reports of antitumor and antibacterial infection related to CaO2 NPs have increased in recent years, but their application in clinical transformation still has a long way to go. Solving the above problems is helpful to speed up their application in clinical transformation. I believe that the hybrid nanomaterials related to CaO2 NPs will provide a better way for tumor treatment and bacterial infection in the future through continuous optimization and improvement.


tel: +86 451 85555848, fax: +86 451 53670428

  1. Funding information: This work was supported by the National Natural Science Foundation of China (no. 51873052).

  2. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: The authors state no conflict of interest.

References

[1] Lin, L., S. Wang, H. Deng, W. Yang, L. Rao, R. Tian, et al. Endogenous labile iron pool-mediated free radical generation for cancer chemodynamic therapy. Journal of the American Chemical Society, Vol. 142, No. 36, 2020, pp. 15320–15330.10.1021/jacs.0c05604Search in Google Scholar PubMed

[2] Gao, F., M. Sun, J. Zhang, Y. Chang, W. Gao, G. Ma, et al. Fenton-like reaction and glutathione depletion by chiral manganese dioxide nanoparticles for enhanced chemodynamic therapy and chemotherapy. Journal of Colloid and Interface Science, Vol. 616, 2022, pp. 369–378.10.1016/j.jcis.2022.02.060Search in Google Scholar PubMed

[3] Wan, Y., L. H. Fu, C. Li, J. Lin, and P. Huang. Conquering the hypoxia limitation for photodynamic therapy. Advanced Materials, Vol. 33, No. 48, 2021, id. 2103978.10.1002/adma.202103978Search in Google Scholar PubMed

[4] Zhao, X., J. Liu, J. Fan, H. Chao, and X. Peng. Recent progress in photosensitizers for overcoming the challenges of photodynamic therapy: from molecular design to application. Chemical Society Reviews, Vol. 50, No. 6, 2021, pp. 4185–4219.10.1039/D0CS00173BSearch in Google Scholar PubMed

[5] Duosiken, D., R. Yang, Y. Dai, Z. Marfavi, Q. Lv, H. Li, et al. Near-infrared light-excited reactive oxygen species generation by thulium oxide nanoparticles. Journal of the American Chemical Society, Vol. 144, No. 6, 2022, pp. 2455–2459.10.1021/jacs.1c11704Search in Google Scholar PubMed

[6] Ding, L., Y. Chang, P. Yang, W. Gao, M. Sun, Y. Bie, et al. Facile synthesis of biocompatible l-cysteine-modified MoS2 nanospheres with high photothermal conversion efficiency for photothermal therapy of tumor. Materials Science and Engineering: C, Vol. 117, 2020, id. 111371.10.1016/j.msec.2020.111371Search in Google Scholar PubMed

[7] Xie, L., M. Yan, T. Liu, K. Gong, X. Luo, B. Qiu, et al. Kinetics-controlled super-assembly of asymmetric porous and hollow carbon nanoparticles as light-sensitive smart nanovehicles. Journal of the American Chemical Society, Vol. 144, No. 4, 2022, pp. 1634–1646.10.1021/jacs.1c10391Search in Google Scholar PubMed

[8] Li, G., S. Wang, D. Deng, Z. Xiao, Z. Dong, Z. Wang, et al. Fluorinated chitosan to enhance transmucosal delivery of sonosensitizer-conjugated catalase for sonodynamic bladder cancer treatment post-intravesical instillation. ACS Nano, Vol. 14, No. 2, 2020, pp. 1586–1599.10.1021/acsnano.9b06689Search in Google Scholar PubMed

[9] Kong, H., Q. Chu, C. Fang, G. Cao, G. Han, and X. Li. Cu–ferrocene‐functionalized CaO2 nanoparticles to enable tumor‐specific synergistic therapy with GSH depletion and calcium overload. Advanced Science, Vol. 8, No. 14, 2021, id. 2100241.10.1002/advs.202100241Search in Google Scholar PubMed PubMed Central

[10] Qiao, Y., Y. Xu, X. Liu, Y. Zheng, B. Li, Y. Han, et al. Microwave assisted antibacterial action of Garcinia nanoparticles on Gram-negative bacteria. Nature Communications, Vol. 13, No. 1, 2022, pp. 1–13.10.1038/s41467-022-30125-wSearch in Google Scholar PubMed PubMed Central

[11] Willyard, C. Drug-resistant bacteria ranked. Nature, Vol. 543, No. 7643, 2017, id. 15.10.1038/nature.2017.21550Search in Google Scholar

[12] Li, X., S. M. Robinson, A. Gupta, K. Saha, Z. Jiang, D. F. Moyano, et al. Functional gold nanoparticles as potent antimicrobial agents against multi-drug-resistant bacteria. ACS Nano, Vol. 8, No. 10, 2014, pp. 10682–10686.10.1021/nn5042625Search in Google Scholar PubMed PubMed Central

[13] Northup, A. and D. Cassidy. Calcium peroxide (CaO2) for use in modified Fenton chemistry. Journal of Hazardous Materials, Vol. 152, No. 3, 2008, pp. 1164–1170.10.1016/j.jhazmat.2007.07.096Search in Google Scholar PubMed

[14] Northup, A. M. Modified fenton oxidation of hydrocarbon contaminated soils with CaO2: Microbial survival and surfactant production, 2007. Dissertations. 3507.Search in Google Scholar

[15] Zhang, M., R. Song, Y. Liu, Z. Yi, X. Meng, J. Zhang, et al. Calcium-overload-mediated tumor therapy by calcium peroxide nanoparticles. Chem, Vol. 5, No. 8, 2019, pp. 2171–2182.10.1016/j.chempr.2019.06.003Search in Google Scholar

[16] Ma, Y., H. Xu, B. Sun, S. Du, S. Cui, L. Zhang, et al. pH-responsive oxygen and hydrogen peroxide self-supplying nanosystem for photodynamic and chemodynamic therapy of wound infection. ACS Applied Materials & Interfaces, Vol. 13, No. 50, 2021, pp. 59720–59730.10.1021/acsami.1c19681Search in Google Scholar PubMed

[17] Liu, J., Y. Jin, Z. Song, L. Xu, Y. Yang, X. Zhao, et al. Boosting tumor treatment by dredging the hurdles of chemodynamic therapy synergistic ion therapy. Chemical Engineering Journal, Vol. 411, 2021, id. 128440.10.1016/j.cej.2021.128440Search in Google Scholar

[18] He, Y., K. Li, X. Yang, J. Leng, K. Xu, Z. Yuan, et al. Calcium peroxide nanoparticles‐embedded coatings on anti‐inflammatory TiO2 nanotubes for bacteria elimination and inflammatory environment amelioration. Small, Vol. 17, No. 47, 2021, id. 2102907.10.1002/smll.202102907Search in Google Scholar PubMed

[19] Wang, X., C. Li, H. Jin, X. Wang, C. Ding, D. Cao, et al. Mutual promotion of oxidative stress amplification and calcium overload by degradable spatially selective self-cascade catalyst for synergistic tumor therapy. Chemical Engineering Journal, Vol. 432, 2022, id. 134438.10.1016/j.cej.2021.134438Search in Google Scholar

[20] He, J., L.-H. Fu, C. Qi, J. Lin, and P. Huang. Metal peroxides for cancer treatment. Bioactive Materials, Vol. 6, No. 9, 2021, pp. 2698–2710.10.1016/j.bioactmat.2021.01.026Search in Google Scholar PubMed PubMed Central

[21] Khodaveisi, J., H. Banejad, A. Afkhami, E. Olyaie, S. Lashgari, and R. Dashti. Synthesis of calcium peroxide nanoparticles as an innovative reagent for in situ chemical oxidation. Journal of Hazardous Materials, Vol. 192, No. 3, 2011, pp. 1437–1440.10.1016/j.jhazmat.2011.06.060Search in Google Scholar PubMed

[22] Hu, Y., X. Wang, P. Zhao, H. Wang, W. Gu, and L. Ye. Nanozyme-catalyzed oxygen release from calcium peroxide nanoparticles for accelerated hypoxia relief and image-guided super-efficient photodynamic therapy. Biomaterials Science, Vol. 8, No. 10, 2020, pp. 2931–2938.10.1039/D0BM00187BSearch in Google Scholar PubMed

[23] Ji, C., Z. Lu, Y. Xu, B. Shen, S. Yu, and D. Shi. Self‐production of oxygen system CaO2/MnO2@ PDA‐MB for the photodynamic therapy research and switch‐control tumor cell imaging. Journal of Biomedical Materials Research, Part B: Applied Biomaterials, Vol. 106, No. 7, 2018, pp. 2544–2552.10.1002/jbm.b.34071Search in Google Scholar PubMed

[24] Zhang, M., T. Kiratiwongwan, and W. Shen. Oxygen‐releasing polycaprolactone/calcium peroxide composite microspheres. Journal of Biomedical Materials Research, Part B: Applied Biomaterials, Vol. 108, No. 3, 2020, pp. 1097–1106.10.1002/jbm.b.34461Search in Google Scholar PubMed

[25] Han, Y., J. Ouyang, Y. Li, F. Wang, and J. H. Jiang. Engineering H2O2 self-supplying nanotheranostic platform for targeted and imaging-guided chemodynamic therapy. ACS Applied Materials & Interfaces, Vol. 12, No. 1, 2020, pp. 288–297.10.1021/acsami.9b18676Search in Google Scholar PubMed

[26] He, C., X. Zhang, R. Yan, P. Zhao, Y. Chen, M. Li, et al. Enhancement of cisplatin efficacy by lipid–CaO2 nanocarrier-mediated comprehensive modulation of the tumor microenvironment. Biomaterials Science, Vol. 7, No. 10, 2019, pp. 4260–4272.10.1039/C9BM00797KSearch in Google Scholar

[27] Yu, Q., T. Huang, C. Liu, M. Zhao, M. Xie, G. Li, et al. Oxygen self-sufficient NIR-activatable liposomes for tumor hypoxia regulation and photodynamic therapy. Chemical Science, Vol. 10, No. 39, 2019, pp. 9091–9098.10.1039/C9SC03161HSearch in Google Scholar

[28] Zheng, A. Studies on controlled synthesis of calcium carbonate and calcium peroxide mediated by dopamine and sodium citrate and their applications. Hangzhou Normal University, Hangzhou, 2022.Search in Google Scholar

[29] Park, J. S., Y. J. Song, Y. G. Lim, and K. Park. Facile fabrication of oxygen-releasing tannylated calcium peroxide nanoparticles. Materials, Vol. 13, No. 17, 2020, id. 3864.10.3390/ma13173864Search in Google Scholar PubMed PubMed Central

[30] Yin, Y., T. Jiang, Y. Hao, J. Zhang, W. Li, Y. Hao, et al. Cascade catalytic nanoplatform based on ions interference strategy for calcium overload therapy and ferroptosis. International Journal of Pharmaceutics, Vol. 606, 2021, id. 120937.10.1016/j.ijpharm.2021.120937Search in Google Scholar PubMed

[31] Liu, L. H., Y. H. Zhang, W. X. Qiu, L. Zhang, F. Gao, B. Li, et al. Dual‐stage light amplified photodynamic therapy against hypoxic tumor based on an O2 self‐sufficient nanoplatform. Small, Vol. 13, No. 37, 2017, id. 1701621.10.1002/smll.201701621Search in Google Scholar PubMed

[32] Shen, S., M. Mamat, S. Zhang, J. Cao, Z. D. Hood, L. Figueroa-Cosme, et al. Synthesis of CaO2 nanocrystals and their spherical aggregates with uniform sizes for use as a biodegradable bacteriostatic agent. Small, Vol. 15, No. 36, 2019, id. e1902118.10.1002/smll.201902118Search in Google Scholar PubMed

[33] Ding, J., J. Chen, L. Gao, Z. Jiang, Y. Zhang, M. Li, et al. Engineered nanomedicines with enhanced tumor penetration. Nano Today, Vol. 29, 2019, id. 100800.10.1016/j.nantod.2019.100800Search in Google Scholar

[34] Comerford, K. M., T. J. Wallace, J. Karhausen, N. A. Louis, M. C. Montalto, and S. P. Colgan. Hypoxia-inducible factor-1-dependent regulation of the multidrug resistance (MDR1) gene. Cancer Research, Vol. 62, No. 12, 2002, pp. 3387–3394.Search in Google Scholar

[35] Comerford, K. M., E. P. Cummins, and C. T. Taylor. c-Jun NH2-terminal kinase activation contributes to hypoxia-inducible factor 1α–dependent P-glycoprotein expression in hypoxia. Cancer Research, Vol. 64, No. 24, 2004, pp. 9057–9061.10.1158/0008-5472.CAN-04-1919Search in Google Scholar PubMed

[36] Chen, J., Z. Jiang, W. Xu, T. Sun, X. Zhuang, J. Ding, et al. Spatiotemporally targeted nanomedicine overcomes hypoxia-induced drug resistance of tumor cells after disrupting neovasculature. Nano Letters, Vol. 20, No. 8, 2020, pp. 6191–6198.10.1021/acs.nanolett.0c02515Search in Google Scholar PubMed

[37] Yuan, J., Y. Cen, X.-J. Kong, S. Wu, C. L. Liu, R. Q. Yu, et al. MnO2-nanosheet-modified upconversion nanosystem for sensitive turn-on fluorescence detection of H2O2 and glucose in blood. ACS Applied Materials & Interfaces, Vol. 7, No. 19, 2015, pp. 10548–10555.10.1021/acsami.5b02188Search in Google Scholar PubMed

[38] Zhang, X., C. He, Y. Sun, X. Liu, Y. Chen, C. Chen, et al. A smart O2-generating nanocarrier optimizes drug transportation comprehensively for chemotherapy improving. Acta Pharmaceutica Sinica B, Vol. 11, No. 11, 2021, pp. 3608–3621.10.1016/j.apsb.2021.04.021Search in Google Scholar PubMed PubMed Central

[39] Zhang, X., C. He, R. Yan, Y. Chen, P. Zhao, M. Li, et al. HIF-1 dependent reversal of cisplatin resistance via anti-oxidative nano selenium for effective cancer therapy. Chemical Engineering Journal, Vol. 380, 2020, id. 122540.10.1016/j.cej.2019.122540Search in Google Scholar

[40] He, C., X. Zhang, C. Chen, X. Liu, Y. Chen, R. Yan, et al. A solid lipid coated calcium peroxide nanocarrier enables combined cancer chemo/chemodynamic therapy with O2/H2O2 self-sufficiency. Acta Biomaterialia, Vol. 122, 2021, pp. 354–364.10.1016/j.actbio.2020.12.036Search in Google Scholar PubMed

[41] Wang, Y., X. He, C. Zhou, Y. Bai, T. Li, J. Liu, et al. Nanoscale CaO2 materials for synergistic transarterial chemoembolization in a VX2 orthotopic rabbit liver cancer model. Acta Biomaterialia, Vol. 154, 2022, pp. 536–548.10.1016/j.actbio.2022.09.074Search in Google Scholar PubMed

[42] Zhang, F., C. Xin, Z. Dai, H. Hu, Q. An, F. Wang, et al. Oncocyte membrane-camouflaged multi-stimuli-responsive nanohybrids for synergistic amplification of tumor oxidative stresses and photothermal enhanced cancer therapy. ACS Applied Materials & Interfaces, Vol. 14, No. 36, 2022, pp. 40633–40644.10.1021/acsami.2c11200Search in Google Scholar PubMed

[43] Sun, Y., D. Zhao, G. Wang, Y. Wang, L. Cao, J. Sun, et al. Recent progress of hypoxia-modulated multifunctional nanomedicines to enhance photodynamic therapy: opportunities, challenges, and future development. Acta Pharmaceutica Sinica B, Vol. 10, No. 8, 2020, pp. 1382–1396.10.1016/j.apsb.2020.01.004Search in Google Scholar PubMed PubMed Central

[44] Wang, Y.-Y., Y.-C. Liu, H. Sun, and D. S. Guo. Type I photodynamic therapy by organic–inorganic hybrid materials: From strategies to applications. Coordination Chemistry Reviews, Vol. 395, 2019, pp. 46–62.10.1016/j.ccr.2019.05.016Search in Google Scholar

[45] Feng, G., G.-Q. Zhang, and D. Ding. Design of superior phototheranostic agents guided by Jablonski diagrams. Chemical Society Reviews, Vol. 49, No. 22, 2020, pp. 8179–8234.10.1039/D0CS00671HSearch in Google Scholar

[46] Ochsner, M. Photophysical and photobiological processes in the photodynamic therapy of tumours. Journal of Photochemistry and Photobiology, B: Biology, Vol. 39, No. 1, 1997, pp. 1–18.10.1016/S1011-1344(96)07428-3Search in Google Scholar

[47] Chen, D., Q. Xu, W. Wang, J. Shao, W. Huang, and X. Dong. Type I photosensitizers revitalizing photodynamic oncotherapy. Small, Vol. 17, No. 31, 2021, id. 2006742.10.1002/smll.202006742Search in Google Scholar PubMed

[48] Lo, P.-C., M. S. Rodríguez-Morgade, R. K. Pandey, D. Ng, T. Torres, and F. Dumoulin. The unique features and promises of phthalocyanines as advanced photosensitisers for photodynamic therapy of cancer. Chemical Society Reviews, Vol. 49, No. 4, 2020, pp. 1041–1056.10.1039/C9CS00129HSearch in Google Scholar PubMed

[49] Sun, X., K. Chen, Y. Liu, G. Zhang, M. Shi, P. Shi, et al. Metal–organic framework combined with CaO2 nanoparticles for enhanced and targeted photodynamic therapy. Nanoscale Advances, Vol. 3, No. 23, 2021, pp. 6669–6677.10.1039/D1NA00610JSearch in Google Scholar

[50] Ren, S.-Z., X.-H. Zhu, B. Wang, M. Liu, S. K. Li, Y. S. Yang, et al. A versatile nanoplatform based on multivariate porphyrinic metal–organic frameworks for catalytic cascade-enhanced photodynamic therapy. Journal of Materials Chemistry B, Vol. 9, No. 23, 2021, pp. 4678–4689.10.1039/D0TB02652BSearch in Google Scholar PubMed

[51] Gulzar, A., F. He, A. Gulzar, Y. Kuang, F. Zhang, S. Gai, et al. In situ oxygenating and 808nm light-sensitized nanocomposite for multimodal imaging and mitochondria-assisted cancer therapy. Journal of Materials Chemistry B, Vol. 9, No. 1, 2021, pp. 131–146.10.1039/D0TB01967DSearch in Google Scholar PubMed

[52] Chen, M., S. Zhao, J. Zhu, E. Feng, F. Lv, W. Chen, et al. Open-source and reduced-expenditure nanosystem with ROS self-amplification and glutathione depletion for simultaneous augmented chemodynamic/photodynamic therapy. ACS Applied Materials & Interfaces, Vol. 14, No. 18, 2022, pp. 20682–20692.10.1021/acsami.2c01782Search in Google Scholar PubMed

[53] Yan, J.-H., W. Meng, H. Shan, X. P. Zhang, L. M. Zou, L. L. Wang, et al. Melanin nanoparticles combined with CaO2 nanoparticles for image-guided tumor microenvironment-responsive multimodal therapy. ACS Applied Nano Materials, Vol. 4, No. 2, 2021, pp. 1351–1363.10.1021/acsanm.0c02916Search in Google Scholar

[54] Zhang, L., C. X. Li, S. S. Wan, and X. Z. Zhang. Nanocatalyst‐mediated chemodynamic tumor therapy. Advanced Healthcare Materials, Vol. 11, No. 2, 2022, id. 2101971.10.1002/adhm.202101971Search in Google Scholar PubMed

[55] Wang, X., X. Zhong, Z. Liu, and L. Cheng. Recent progress of chemodynamic therapy-induced combination cancer therapy. Nano Today, Vol. 35, 2020, id. 100946.10.1016/j.nantod.2020.100946Search in Google Scholar

[56] Zhang, D., Y. Meng, Y. Song, P. Cui, Z. Hu, and X. Zheng. Precision therapy through breaking the intracellular redox balance with an MOF-based hydrogel intelligent nanobot for enhancing ferroptosis and activating immunotherapy. Nanoscale, Vol. 14, No. 23, 2022, pp. 8441–8453.10.1039/D2NR00950ASearch in Google Scholar PubMed

[57] Zhou, Y., S. Fan, L. Feng, X. Huang, and X. Chen. Manipulating intratumoral fenton chemistry for enhanced chemodynamic and chemodynamic-synergized multimodal therapy. Advanced Materials, Vol. 33, No. 48, 2021, id. e2104223.10.1002/adma.202104223Search in Google Scholar PubMed

[58] Ruan, J., H. Liu, B. Chen, F. Wang, W. Wang, Z. Zha, et al. Interfacially engineered Zn(x)Mn(1-x)S@Polydopamine hollow nanospheres for glutathione depleting photothermally enhanced chemodynamic therapy. ACS Nano, Vol. 15, No. 7, 2021, pp. 11428–11440.10.1021/acsnano.1c01077Search in Google Scholar PubMed

[59] Wang, J., J. Sun, W. Hu, Y. Wang, T. Chou, B. Zhang, et al. A porous Au@Rh bimetallic core-shell nanostructure as an H2O2 -driven oxygenerator to alleviate tumor hypoxia for simultaneous bimodal imaging and enhanced photodynamic therapy. Advanced Materials, Vol. 32, No. 22, 2020, id. e2001862.10.1002/adma.202001862Search in Google Scholar PubMed PubMed Central

[60] Shi, L., F. Hu, Y. Duan, W. Wu, J. Dong, X. Meng, et al. Hybrid nanospheres to overcome hypoxia and intrinsic oxidative resistance for enhanced photodynamic therapy. ACS Nano, Vol. 14, No. 2, 2020, pp. 2183–2190.10.1021/acsnano.9b09032Search in Google Scholar PubMed

[61] Fan, J. X., M. Y. Peng, H. Wang, H. R. Zheng, Z. L. Liu, C. X. Li, et al. Engineered bacterial bioreactor for tumor therapy via Fenton-like reaction with localized H2O2 generation. Advanced Materials, Vol. 31, No. 16, 2019, id. e1808278.10.1002/adma.201808278Search in Google Scholar PubMed

[62] Feng, L., R. Xie, C. Wang, S. Gai, F. He, D. Yang, et al. Magnetic targeting, tumor microenvironment-responsive intelligent nanocatalysts for enhanced tumor ablation. ACS Nano, Vol. 12, No. 11, 2018, pp. 11000–11012.10.1021/acsnano.8b05042Search in Google Scholar PubMed

[63] Wu, Y., T. Guo, Y. Qiu, Y. Lin, Y. Yao, W. Lian, et al. An inorganic prodrug, tellurium nanowires with enhanced ROS generation and GSH depletion for selective cancer therapy. Chemical Science, Vol. 10, No. 29, 2019, pp. 7068–7075.10.1039/C9SC01070JSearch in Google Scholar PubMed PubMed Central

[64] Mamat, M., X. Wang, L. Wu, R. Zhao, J. Cao, X. Qi, et al. CaO2/Fe3O4 nanocomposites for oxygen-independent generation of radicals and cancer therapy. Colloids and Surfaces B, Biointerfaces, Vol. 204, 2021, id. 111803.10.1016/j.colsurfb.2021.111803Search in Google Scholar PubMed

[65] Kong, H., Q. Chu, C. Fang, G. Cao, G. Han, and X. Li. Cu-ferrocene-functionalized CaO2 nanoparticles to enable tumor-specific synergistic therapy with GSH depletion and calcium overload. Adv Sci (Weinh), Vol. 8, No. 14, 2021, id. e2100241.10.1002/advs.202100241Search in Google Scholar PubMed PubMed Central

[66] Zhang, L., S.-S. Wan, C.-X. Li, L. Xu, H. Cheng, and X. Z. Zhang. An adenosine triphosphate-responsive autocatalytic fenton nanoparticle for tumor ablation with self-supplied H2O2 and acceleration of Fe (III)/Fe (II) conversion. Nano Letters, Vol. 18, No. 12, 2018, pp. 7609–7618.10.1021/acs.nanolett.8b03178Search in Google Scholar PubMed

[67] Fu, L. H., Y. R. Hu, C. Qi, T. He, S. Jiang, C. Jiang, et al. Biodegradable manganese-doped calcium phosphate nanotheranostics for traceable cascade reaction-enhanced anti-tumor therapy. ACS Nano, Vol. 13, No. 12, 2019, pp. 13985–13994.10.1021/acsnano.9b05836Search in Google Scholar PubMed

[68] Qi, C., J. He, L. H. Fu, T. He, N. T. Blum, X. Yao, et al. Tumor-specific activatable nanocarriers with gas-generation and signal amplification capabilities for tumor theranostics. ACS Nano, Vol. 15, No. 1, 2021, pp. 1627–1639.10.1021/acsnano.0c09223Search in Google Scholar PubMed

[69] Dai, Y., Z. Yang, S. Cheng, Z. Wang, R. Zhang, G. Zhu, et al. Toxic reactive oxygen species enhanced synergistic combination therapy by self‐assembled metal‐phenolic network nanoparticles. Advanced Materials, Vol. 30, No. 8, 2018, id. 1704877.10.1002/adma.201704877Search in Google Scholar PubMed

[70] Wang, Z., B. Liu, Q. Sun, S. Dong, Y. Kuang, Y. Dong, et al. Fusiform-like copper(II)-based metal–organic framework through relief hypoxia and GSH-depletion co-enhanced starvation and chemodynamic synergetic cancer therapy. ACS Applied Materials & Interfaces, Vol. 12, No. 15, 2020, pp. 17254–17267.10.1021/acsami.0c01539Search in Google Scholar PubMed

[71] Ren, Z., S. Sun, R. Sun, G. Cui, L. Hong, B. Rao, et al. A metal-polyphenol-coordinated nanomedicine for synergistic cascade cancer chemotherapy and chemodynamic therapy. Advanced Materials, Vol. 32, No. 6, 2020, id. e1906024.10.1002/adma.201906024Search in Google Scholar PubMed

[72] Sun, P., Q. Deng, L. Kang, Y. Sun, J. Ren, and X. Qu. A smart nanoparticle-laden and remote-controlled self-destructive macrophage for enhanced chemo/chemodynamic synergistic therapy. ACS Nano, Vol. 14, No. 10, 2020, pp. 13894–13904.10.1021/acsnano.0c06290Search in Google Scholar PubMed

[73] Gao, S., Y. Jin, K. Ge, Z. Li, H. Liu, X. Dai, et al. Self‐supply of O2 and H2O2 by a nanocatalytic medicine to enhance combined chemo/chemodynamic therapy. Advanced Science, Vol. 6, No. 24, 2019, id. 1902137.10.1002/advs.201902137Search in Google Scholar PubMed PubMed Central

[74] Li, S., Q. Jiang, S. Liu, Y. Zhang, Y. Tian, C. Song, et al. A DNA nanorobot functions as a cancer therapeutic in response to a molecular trigger in vivo. Nature Biotechnology, Vol. 36, No. 3, 2018, pp. 258–264.10.1038/nbt.4071Search in Google Scholar PubMed

[75] Chen, C., W. Su, Y. Liu, J. Zhang, C. Zuo, Z. Yao, et al. Artificial anaerobic cell dormancy for tumor gaseous microenvironment regulation therapy. Biomaterials, Vol. 200, 2019, pp. 48–55.10.1016/j.biomaterials.2019.02.007Search in Google Scholar PubMed

[76] Zhang, X., C. He, Y. Chen, C. Chen, R. Yan, T. Fan, et al. Cyclic reactions-mediated self-supply of H2O2 and O2 for cooperative chemodynamic/starvation cancer therapy. Biomaterials, Vol. 275, 2021, id. 120987.10.1016/j.biomaterials.2021.120987Search in Google Scholar PubMed

[77] Liu, C., Y. Cao, Y. Cheng, D. Wang, T. Xu, L. Su, et al. An open source and reduce expenditure ROS generation strategy for chemodynamic/photodynamic synergistic therapy. Nature Communications, Vol. 11, No. 1, 2020, id. 1735.10.1038/s41467-020-15591-4Search in Google Scholar PubMed PubMed Central

[78] Xiao, Y., Z. Li, A. Bianco, and B. Ma. Recent advances in calcium‐based anticancer nanomaterials exploiting calcium overload to trigger cell apoptosis. Advanced Functional Materials, 2022, id. 2209291.10.1002/adfm.202209291Search in Google Scholar

[79] Giorgi, C., S. Marchi, and P. Pinton. The machineries, regulation and cellular functions of mitochondrial calcium. Nature Reviews Molecular Cell Biology, Vol. 19, No. 11, 2018, pp. 713–730.10.1038/s41580-018-0052-8Search in Google Scholar PubMed

[80] Zheng, P., B. Ding, Z. Jiang, W. Xu, G. Li, J. Ding, et al. Ultrasound-augmented mitochondrial calcium ion overload by calcium nanomodulator to induce immunogenic cell death. Nano Letters, Vol. 21, No. 5, 2021, pp. 2088–2093.10.1021/acs.nanolett.0c04778Search in Google Scholar PubMed

[81] Zheng, P. and J. Ding. Calcium ion nanomodulators for mitochondria-targeted multimodal cancer therapy. Asian Journal of Pharmaceutical Sciences, Vol. 17, No. 1, 2022, pp. 1–3.10.1016/j.ajps.2021.10.004Search in Google Scholar PubMed PubMed Central

[82] Zheng, P., B. Ding, R. Shi, Z. Jiang, W. Xu, G. Li, et al. A multichannel Ca2+ nanomodulator for multilevel mitochondrial destruction‐mediated cancer therapy. Advanced Materials, Vol. 33, No. 15, 2021, id. 2007426.10.1002/adma.202007426Search in Google Scholar PubMed

[83] Yang, G., Y. Liu, J. Chen, J. Ding, and X. Chen. Self-adaptive nanomaterials for rational drug delivery in cancer therapy. Accounts of Materials Research, Vol. 3, No. 12, 2022, pp. 1232–1247.10.1021/accountsmr.2c00147Search in Google Scholar

[84] Liu, B., S. Liang, Z. Wang, Q. Sun, F. He, S. Gai, et al. A tumor-microenvironment-responsive nanocomposite for hydrogen sulfide gas and trimodal-enhanced enzyme dynamic therapy. Advanced Materials, Vol. 33, No. 30, 2021, id. e2101223.10.1002/adma.202101223Search in Google Scholar PubMed

[85] Sitkovsky, M. V., S. Hatfield, R. Abbott, B. Belikoff, D. Lukashev, and A. Ohta. Hostile, hypoxia-A2-adenosinergic tumor biology as the next barrier to overcome for tumor immunologists. Cancer Immunology Research, Vol. 2, No. 7, 2014, pp. 598–605.10.1158/2326-6066.CIR-14-0075Search in Google Scholar PubMed PubMed Central

[86] An, J., K. Zhang, B. Wang, S. Wu, Y. Wang, H. Zhang, et al. Nanoenabled disruption of multiple barriers in antigen cross-presentation of dendritic cells via calcium interference for enhanced chemo-immunotherapy. ACS Nano, Vol. 14, No. 6, 2020, pp. 7639–7650.10.1021/acsnano.0c03881Search in Google Scholar PubMed

[87] Bong, A. H. L. and G. R. Monteith. Calcium signaling and the therapeutic targeting of cancer cells. Biochimica et Biophysica Acta, Molecular Cell Research, Vol. 1865, No. 11 Pt B, 2018, pp. 1786–1794.10.1016/j.bbamcr.2018.05.015Search in Google Scholar PubMed

[88] Peters, A. A., M. J. Milevskiy, W. C. Lee, M. C. Curry, C. E. Smart, J. M. Saunus, et al. The calcium pump plasma membrane Ca2+-ATPase 2 (PMCA2) regulates breast cancer cell proliferation and sensitivity to doxorubicin. Scientific Reports, Vol. 6, 2016, id. 25505.10.1038/srep25505Search in Google Scholar PubMed PubMed Central

[89] Seo, J. A., B. Kim, D. N. Dhanasekaran, B. K. Tsang, and Y. S. Song. Curcumin induces apoptosis by inhibiting sarco/endoplasmic reticulum Ca2+ ATPase activity in ovarian cancer cells. Cancer Letters, Vol. 371, No. 1, 2016, pp. 30–37.10.1016/j.canlet.2015.11.021Search in Google Scholar PubMed

[90] Doskey, C. M., V. Buranasudja, B. A. Wagner, J. G. Wilkes, J. Du, J. J. Cullen, et al. Tumor cells have decreased ability to metabolize H2O2: Implications for pharmacological ascorbate in cancer therapy. Redox Biology, Vol. 10, 2016, pp. 274–284.10.1016/j.redox.2016.10.010Search in Google Scholar PubMed PubMed Central

[91] Ermak, G. and K. J. Davies. Calcium and oxidative stress: from cell signaling to cell death. Molecular Immunology, Vol. 38, No. 10, 2002, pp. 713–721.10.1016/S0161-5890(01)00108-0Search in Google Scholar PubMed

[92] Jiang, Y., W. Meng, L. Wu, K. Shao, L. Wang, M. Ding, et al. Image-guided TME-improving nano-platform for Ca2+ signal disturbance and enhanced tumor PDT. Advanced Healthcare Materials, Vol. 10, No. 19, 2021, id. e2100789.10.1002/adhm.202100789Search in Google Scholar PubMed

[93] Chen, F., B. Yang, L. Xu, J. Yang, and J. Li. A CaO2@Tannic Acid‐FeIII nanoconjugate for enhanced chemodynamic tumor therapy. ChemMedChem, Vol. 16, No. 14, 2021, pp. 2278–2286.10.1002/cmdc.202100108Search in Google Scholar PubMed

[94] Qian, G., J. Wang, L. Yang, Z. Zeng, Z. Zhao, S. Peng, et al. A pH-responsive CaO2@ZIF-67 system endows a scaffold with chemodynamic therapy properties. Journal of Materials Science, Vol. 58, 2023, pp. 1–5.10.1007/s10853-022-08103-wSearch in Google Scholar

[95] Cheng, F. Y., C. H. Chan, B. J. Wang, Y. L. Yeh, Y. J. Wang, and H. W. Chiu. The oxygen-generating calcium peroxide-modified magnetic nanoparticles attenuate hypoxia-induced chemoresistance in triple-negative breast cancer. Cancers (Basel), Vol. 13, No. 4, 2021, id. 606.10.3390/cancers13040606Search in Google Scholar PubMed PubMed Central

[96] Ruan, L., G. Song, X. Zhang, T. Liu, Y. Sun, J. Zhu, et al. Transdermal delivery of multifunctional CaO2@Mn-PDA nanoformulations by microneedles for NIR-induced synergistic therapy against skin melanoma. Biomaterials Science, Vol. 9, No. 20, 2021, pp. 6830–6841.10.1039/D1BM01117KSearch in Google Scholar PubMed

[97] Suo, M., Z. Liu, W. Tang, J. Guo, W. Jiang, Y. Liu, et al. Development of a novel oxidative stress-amplifying nanocomposite capable of supplying intratumoral H2O2 and O(2) for enhanced chemodynamic therapy and radiotherapy in patient-derived xenograft (PDX) models. Nanoscale, Vol. 12, No. 45, 2020, pp. 23259–23265.10.1039/D0NR06594CSearch in Google Scholar PubMed

[98] Cui, R., J. Shi, and Z. Liu. Metal-organic framework-encapsulated nanoparticles for synergetic chemo/chemodynamic therapy with targeted H2O2 self-supply. Dalton Transactions, Vol. 50, No. 43, 2021, pp. 15870–15877.10.1039/D1DT03110DSearch in Google Scholar

[99] Yang, P., Y. Chang, J. Zhang, F. Gao, X. Liu, Q. Wei, et al. The combination of in situ photodynamic promotion and ion-interference to improve the efficacy of cancer therapy. Journal of Colloid and Interface Science, Vol. 629, 2023, pp. 522–533.10.1016/j.jcis.2022.08.125Search in Google Scholar PubMed

[100] Kong, H., C. Fang, Q. Chu, Z. Hu, Y. Fu, G. Han, et al. Catalytic core-shell nanoparticles with self-supplied calcium and H2O2) to enable combinational tumor inhibition. Journal of Nanobiotechnology, Vol. 19, No. 1, 2021, id. 313.10.1186/s12951-021-01055-4Search in Google Scholar PubMed PubMed Central

[101] Shen, J., H. Yu, Y. Shu, M. Ma, and H. Chen. A robust ROS generation strategy for enhanced chemodynamic/photodynamic therapy via H2O2/O2 self‐supply and Ca2+ overloading. Advanced Functional Materials, Vol. 31, No. 50, 2021, id. 2106106.10.1002/adfm.202106106Search in Google Scholar

[102] Wang, M., L. Zhang, H. Hao, X. Hu, Z. Xin, Y. Zhu, et al. Synergistic H2O2 self-supplying and NIR-responsive drug delivery nanoplatform for chemodynamic-photothermal-chemotherapy. Colloids and Surfaces B, Biointerfaces, Vol. 213, 2022, id. 112412.10.1016/j.colsurfb.2022.112412Search in Google Scholar PubMed

[103] Zhang, S., C. Cao, X. Lv, H. Dai, Z. Zhong, C. Liang, et al. A H2O2 self-sufficient nanoplatform with domino effects for thermal-responsive enhanced chemodynamic therapy. Chemical Science, Vol. 11, No. 7, 2020, pp. 1926–1934.10.1039/C9SC05506ASearch in Google Scholar PubMed PubMed Central

[104] Nicholas, D., H. Nesbitt, S. Farrell, K. Logan, E. McMullin, T. Gillan, et al. Exploiting a Rose Bengal-bearing, oxygen-producing nanoparticle for SDT and associated immune-mediated therapeutic effects in the treatment of pancreatic cancer. European journal of pharmaceutics and biopharmaceutics : official journal of Arbeitsgemeinschaft für Pharmazeutische Verfahrenstechnik e.V, Vol. 163, 2021, pp. 49–59.10.1016/j.ejpb.2021.03.005Search in Google Scholar PubMed

[105] Huang, C., B. Lin, C. Chen, H. Wang, X. Lin, J. Liu, et al. Synergistic reinforcing of immunogenic cell death and transforming tumor‐associated macrophages via a multifunctional cascade bioreactor for optimizing cancer immunotherapy. Advanced Materials, Vol. 34, 2022, id. 2207593.10.1002/adma.202207593Search in Google Scholar PubMed

[106] Wu, D., Z. Q. Zhu, H. X. Tang, Z. E. Shi, J. Kang, Q. Liu, et al. Efficacy-shaping nanomedicine by loading calcium peroxide into tumor microenvironment-responsive nanoparticles for the antitumor therapy of prostate cancer. Theranostics, Vol. 10, No. 21, 2020, pp. 9808–9829.10.7150/thno.43631Search in Google Scholar PubMed PubMed Central

[107] Wang, Y., Y. Yang, Y. Shi, H. Song, and C. Yu. Antibiotic‐free antibacterial strategies enabled by nanomaterials: progress and perspectives. Advanced Materials, Vol. 32, No. 18, 2020, id. 1904106.10.1002/adma.201904106Search in Google Scholar PubMed

[108] Liu, P., Y. Wu, B. Mehrjou, K. Tang, G. Wang, and P. K. Chu. Versatile phenol‐incorporated nanoframes for in situ antibacterial activity based on oxidative and physical damages. Advanced Functional Materials, Vol. 32, No. 17, 2022, id. 2110635.10.1002/adfm.202110635Search in Google Scholar

[109] Liu, C., S. Feng, L. Ma, M. Sun, Z. Wei, J. Wang, et al. An amphiphilic carbonaceous/nanosilver composite-incorporated urinary catheter for long-term combating bacteria and biofilms. ACS Applied Materials & Interfaces, Vol. 13, No. 32, 2021, pp. 38029–38039.10.1021/acsami.1c07399Search in Google Scholar PubMed

[110] Pachaiappan, R., S. Rajendran, P. L. Show, K. Manavalan, and M. Naushad. Metal/metal oxide nanocomposites for bactericidal effect: A review. Chemosphere, Vol. 272, 2021, id. 128607.10.1016/j.chemosphere.2020.128607Search in Google Scholar PubMed

[111] Wang, M., X. Zhou, Y. Li, Y. Dong, J. Meng, S. Zhang, et al. Triple-synergistic MOF-nanozyme for efficient antibacterial treatment. Bioactive Materials, Vol. 17, 2022, pp. 289–299.10.1016/j.bioactmat.2022.01.036Search in Google Scholar PubMed PubMed Central

[112] Wang, L., H. He, Y. Yu, L. Sun, S. Liu, C. Zhang, et al. Morphology-dependent bactericidal activities of Ag/CeO2 catalysts against Escherichia coli. Journal of Inorganic Biochemistry, Vol. 135, 2014, pp. 45–53.10.1016/j.jinorgbio.2014.02.016Search in Google Scholar PubMed

[113] Kim, J. S., E. Kuk, K. N. Yu, J. H. Kim, S. J. Park, H. J. Lee, et al. Antimicrobial effects of silver nanoparticles. Nanomedicine, Vol. 3, No. 1, 2007, pp. 95–101.10.1016/j.nano.2006.12.001Search in Google Scholar PubMed

[114] Padmavathy, N. and R. Vijayaraghavan. Enhanced bioactivity of ZnO nanoparticles-an antimicrobial study. Science and Technology of Advanced Materials, Vol. 9, No. 3, 2008, id. 035004.10.1088/1468-6996/9/3/035004Search in Google Scholar PubMed PubMed Central

[115] Ivask, A., A. Elbadawy, C. Kaweeteerawat, D. Boren, H. Fischer, Z. Ji, et al. Toxicity mechanisms in Escherichia coli vary for silver nanoparticles and differ from ionic silver. ACS Nano, Vol. 8, No. 1, 2014, pp. 374–386.10.1021/nn4044047Search in Google Scholar PubMed

[116] Thi, P. L., Y. Lee, D. L. Tran, T. T. Hoang Thi, K. M. Park, and K. D. Park. Calcium peroxide-mediated in situ formation of multifunctional hydrogels with enhanced mesenchymal stem cell behaviors and antibacterial properties. Journal of Materials Chemistry B: Materials for Biology and Medicine, Vol. 8, No. 48, 2020, pp. 11033–11043.10.1039/D0TB02119ASearch in Google Scholar PubMed

[117] Zhang, S., K. Guan, Y. Zhang, J. Zhang, H. Fu, T. Wu, et al. A self-activated NO-releasing hydrogel depot for photothermal enhanced sterilization. Nano Research, Vol. 15, No. 12, 2022, pp. 1–11.10.1007/s12274-022-5239-9Search in Google Scholar

[118] Xie, S., K. Huang, J. Peng, Y. Liu, W. Cao, D. Zhang, et al. Self‐propelling nanomotors integrated with biofilm microenvironment‐activated NO release to accelerate healing of bacteria‐infected diabetic wounds. Advanced Healthcare Materials, Vol. 11, No. 19, 2022, id. 2201323.10.1002/adhm.202201323Search in Google Scholar PubMed

[119] Lei, J., S. Li, S. Liu, Q. Wu, B. Xu, Z. Huang, et al. A bioactive nanocomposite sponge for simultaneous hemostasis and antimicrobial therapy. Nano Research, Vol. 15, No. 12, 2022, pp. 1–9.10.1007/s12274-022-5226-1Search in Google Scholar

[120] Zhang, S., Q. Chai, Z. Man, C. Tang, Z. Li, J. Zhang, et al. Bioinspired nano-painting on orthopedic implants orchestrates periprosthetic anti-infection and osseointegration in a rat model of arthroplasty. Chemical Engineering Journal, Vol. 435, 2022, id. 134848.10.1016/j.cej.2022.134848Search in Google Scholar

[121] Zeng, Z., G. Jiang, Y. Sun, U. E. Aharodnikau, K. E. Yunusov, X. Gao, et al. Rational design of flexible microneedles coupled with CaO2@ PDA-loaded nanofiber films for skin wound healing on diabetic rats. Biomaterials. Science, Vol. 10, No. 18, 2022, pp. 5326–5339.10.1039/D2BM00861KSearch in Google Scholar PubMed

[122] Yu, Y., Q. Lu, J. Sun, P. Zhang, L. Zeng, K. Vasilev, et al. Spontaneous Formation of MXene-Oxidized Sono/Chemo-Dynamic Sonosensitizer/Nanocatalyst for Antibacteria and Bone-Tissue Regeneration, 2023.10.21203/rs.3.rs-2412598/v1Search in Google Scholar

[123] Huang, L., S. Jiang, B. Cai, G. Wang, Z. Wang, and L. Wang. pH-triggered nanoreactors as oxidative stress amplifiers for combating multidrug-resistant biofilms. Chemical Communications, Vol. 57, No. 38, 2021, pp. 4662–4665.10.1039/D1CC00247CSearch in Google Scholar

[124] Gao, S., X. Lu, P. Zhu, H. Lin, L. Yu, H. Yao, et al. Self-evolved hydrogen peroxide boosts photothermal-promoted tumor-specific nanocatalytic therapy. Journal of Materials Chemistry B, Vol. 7, No. 22, 2019, pp. 3599–3609.10.1039/C9TB00525KSearch in Google Scholar

[125] Dong, S., Y. Chen, L. Yu, K. Lin, and X. Wang. Magnetic hyperthermia–synergistic H2O2 self‐sufficient catalytic suppression of osteosarcoma with enhanced bone‐regeneration bioactivity by 3D‐printing composite scaffolds. Advanced Functional Materials, Vol. 30, No. 4, 2020, id. 1907071.10.1002/adfm.201907071Search in Google Scholar

[126] Kato, Y., S. Ozawa, C. Miyamoto, Y. Maehata, A. Suzuki, T. Maeda, et al. Acidic extracellular microenvironment and cancer. Cancer Cell International, Vol. 13, No. 1, 2013, pp. 1–8.10.1186/1475-2867-13-89Search in Google Scholar PubMed PubMed Central

[127] Lin, L. S., J. Song, L. Song, K. Ke, Y. Liu, Z. Zhou, et al. Simultaneous Fenton‐like ion delivery and glutathione depletion by MnO2‐based nanoagent to enhance chemodynamic therapy. Angewandte Chemie, Vol. 130, No. 18, 2018, pp. 4996–5000.10.1002/ange.201712027Search in Google Scholar

[128] Huang, C.-C., W.-T. Chia, M.-F. Chung, K. J. Lin, C. W. Hsiao, C. Jin, et al. An implantable depot that can generate oxygen in situ for overcoming hypoxia-induced resistance to anticancer drugs in chemotherapy. Journal of the American Chemical Society, Vol. 138, No. 16, 2016, pp. 5222–5225.10.1021/jacs.6b01784Search in Google Scholar PubMed

[129] Trédan, O., C. M. Galmarini, K. Patel, and I. F. Tannock. Drug resistance and the solid tumor microenvironment. Journal of the National Cancer Institute, Vol. 99, No. 19, 2007, pp. 1441–1454.10.1093/jnci/djm135Search in Google Scholar PubMed

[130] Lucky, S. S., K. C. Soo, and Y. Zhang. Nanoparticles in photodynamic therapy. Chemical Reviews, Vol. 115, No. 4, 2015, pp. 1990–2042.10.1021/cr5004198Search in Google Scholar PubMed

[131] Masoud, G. N. and W. Li. HIF-1α pathway: role, regulation and intervention for cancer therapy. Acta Pharmaceutica Sinica B, Vol. 5, No. 5, 2015, pp. 378–389.10.1016/j.apsb.2015.05.007Search in Google Scholar PubMed PubMed Central

[132] Cheruku, R. R., J. Cacaccio, F. A. Durrani, W. A. Tabaczynski, R. Watson, A. Marko, et al. Epidermal growth factor receptor-targeted multifunctional photosensitizers for bladder cancer imaging and photodynamic therapy. Journal of Medicinal Chemistry, Vol. 62, No. 5, 2019, pp. 2598–2617.10.1021/acs.jmedchem.8b01927Search in Google Scholar PubMed

[133] Wu, W., D. Mao, S. Xu, M. Panahandeh‐Fard, Y. Duan, F. Hu, et al. Precise molecular engineering of photosensitizers with aggregation‐induced emission over 800 nm for photodynamic therapy. Advanced Functional Materials, Vol. 29, No. 42, 2019, id. 1901791.10.1002/adfm.201901791Search in Google Scholar

[134] Chien, Y. H., K. K. Chan, T. Anderson, K. V. Kong, B. K. Ng, and K. T. Yong. Advanced near‐infrared light‐responsive nanomaterials as therapeutic platforms for cancer therapy. Advanced Therapeutics, Vol. 2, No. 3, 2019, id. 1800090.10.1002/adtp.201800090Search in Google Scholar

[135] Hu, Y., J. F. Honek, B. C. Wilson, and Q. B. Lu. Design, synthesis and photocytotoxicity of upconversion nanoparticles: Potential applications for near‐infrared photodynamic and photothermal therapy. Journal of Biophotonics, Vol. 12, No. 11, 2019, id. e201900129.10.1002/jbio.201900129Search in Google Scholar PubMed

[136] Li, T., F. He, B. Liu, T. Jia, B. Shao, R. Zhao, et al. In situ synthesis of FeOCl in hollow dendritic mesoporous organosilicon for ascorbic acid-enhanced and MR imaging-guided chemodynamic therapy in neutral pH conditions. ACS Applied Materials & Interfaces, Vol. 12, No. 51, 2020, pp. 56886–56897.10.1021/acsami.0c19330Search in Google Scholar PubMed

Received: 2022-10-09
Revised: 2023-01-27
Accepted: 2023-02-20
Published Online: 2023-03-10

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 17.8.2024 from https://www.degruyter.com/document/doi/10.1515/rams-2022-0308/html
Scroll to top button