Next Article in Journal
New Insights and Experimental Investigation of High-Temperature Gel Reinforced by Nano-SiO2
Next Article in Special Issue
Unravelling the Supramolecular Driving Forces in the Formation of CO2-Responsive Pseudopeptidic Low-Molecular-Weight Hydrogelators
Previous Article in Journal
Poloxamer-Based Scaffolds for Tissue Engineering Applications: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Studies on Hydrogels Based on H2O2-Responsive Moieties: Mechanism, Preparation and Application

1
Affiliated Hospital of Hebei University, Baoding 071000, China
2
College of Chemistry and Environmental Science, Hebei University, Baoding 071002, China
*
Author to whom correspondence should be addressed.
Submission received: 27 April 2022 / Revised: 29 May 2022 / Accepted: 30 May 2022 / Published: 8 June 2022
(This article belongs to the Special Issue Advances in Stimuli-Responsive Polymer Gels)

Abstract

:
H2O2 is essential for cellular processes and plays a vital role in the regulation of cell signaling pathways, which can be viewed as a warning signal for many kinds of disease including cancer, cardiovascular disease, reproductive abnormalities, diabetes, and renal failure. A H2O2-responsive hydrogel (H2O2-Gel) is a promising candidate for biomedical applications because of its good biocompatibility, similarity to soft biological tissues, ease of preparation, and its ability to respond to H2O2. In this study, the H2O2-responsive moieties used to fabricate H2O2-Gels were reviewed, including thioethers, disulfide bonds, selenides, diselenium bonds, diketones, boronic, and others. Next, the preparation method of H2O2-Gel was divided into two major categories according to their reaction mechanisms: either self-crosslinking or mechanisms entailing the addition of difunctional crosslinkers. Last, the applications of H2O2-Gels were emphasized, which have been viewed as desirable candidates in the fields of drug delivery, the detection of H2O2, glucose-responsive systems, ROS scavengers, tissue engineering, and cell-encapsulation.

1. Introduction

In recent years, relevant researchers have developed a variety of stimulation mechanisms by investigating the changes in the microenvironment of the human body [1,2,3]. Following this research, quite a number of stimuli-responsive materials have been developed with sensibility to acidity [4], alkali [5], temperature [6], mechanical force [7], magnetic fields [8], ultrasound [9], reactive oxygen species (ROS) [10], etc. ROS are emerging as critical signaling molecules which are abundant in the human microenvironment [11]. The term ROS encompasses a wide range of molecules, such as hydrogen peroxide (H2O2), singlet oxygen (1O2), hydroxyl radical, superoxide, etc. [12]. The stable control of reactive oxygen species level is of great significance to maintain the normal physiological activities of human bodies [13]. The overexpression of reactive oxygen species can easily destroy the structure of many biological macromolecules, such as proteins and nucleic acids, and thus result in a variety of diseases [14]. Therefore, it is of great theoretical and practical significance to design and develop smart materials with ROS responsive behaviors.
H2O2 is a representative case of ROS that is particularly important in the physiological regulation in organisms [15]. It plays an important role in cell signal transduction, differentiation, and proliferation, and also plays an irreplaceable role in the diagnosis and treatment of many diseases [16]. Especially in diseases related to oxidative stress, a change in the H2O2 level is a good indication of the location of the lesion site [17]. When oxidative stress and injury occur in the body, the concentration of H2O2 is higher than the normal level [18]. Therefore, the over expression of H2O2 in vivo is directly related to inflammation, bleeding, Alzheimer’s disease, diabetes, renal function decline or tumors, and other diseases (Scheme 1) [19]. For example, the H2O2 level of tumor tissue is significantly higher than that of normal tissue. The H2O2 concentration in normal tissue usually maintains about 10−8 M, while the H2O2 concentration in inflammatory tissue or tumors is about dozens or even thousands of times higher than that in normal tissue [16]. Therefore, H2O2 is more suitable to serve as a pathological marker than pH, temperature, ionic charge, etc. The concentration of H2O2 is also overexpressed during the proliferation and metastasis of tumor cells. Therefore, the overexpression of H2O2 can be used to monitor changes in the microenvironment. The development of H2O2-responsive materials with high specificity, sensitivity, and accuracy is a meaningful and interesting topic, which may exhibit much more promising applications for early diagnosis than other stimuli responsive materials such as those that respond to pH, temperature, ionic charge, etc.
Until now, different types of H2O2-responsive materials have been explored, including MRI contrast agents, fluorescent sensors, transistors, drug delivery systems, mitochondria-directed tools, etc. [20,21]. Gu et al. published a review paper on the topic of H2O2-responsive materials, in which the biomedical applications were emphasized [17]. It is worth noting that hydrogels are widely used in the fields of drug delivery, tissue engineering, surgical dressing, etc. [22,23,24,25]. As a result, hydrogel based on H2O2-responsive moieties is a major kind of H2O2-responsive material [26,27] which can be defined as H2O2-responsive hydrogel (H2O2-Gel). There is a great deal of significance in performing a review of the development of H2O2-Gel and in the explication of the guidelines for designing suitable formations, which is also our interest in this study. Hydrogels are three-dimensional (3D) cross-linked polymer networks, which can absorb and retain large amounts of water [28]. Due to their flexible structure, stimuli-responsive hydrogels can perform work by converting an external stimulation into mechanical motions or chemical conversions, which can be an effective framework for soft actuators. In this review, we focus on overviewing the responsive mechanism of H2O2-responsive moieties and the applications of H2O2-Gel, especially in biomedicine. The preparation methods are also summarized. We expect to provide guidance and designate the direction of further studies on H2O2-responsive materials.

2. H2O2-Responsive Moieties

2.1. Thioether (-S-)

Thioether can be regarded as a compound analogous to ether in which the oxygen is replaced by sulfur. Thioether is commonly synthesized from potassium or sodium sulfide and halohydrocarbon and is widely distributed in proteins [29]. Benefiting from the higher atomic number of sulfur compared to oxygen, thioether usually shows higher chemical activity than ether and can be transferred into sulfoxide and sulfone with the treatment of H2O2 (Table 1) [30].
Napoli et al. investigated the mechanism for the oxidative destabilization of poly(propylene sulfide) [31]. Even at concentrations as low as 0.03-vol.% H2O2, the transformation from thioether to sulfoxide or sulfone can be achieved. The NMR shifts of the methyl group in the α and β position to sulfur were examined in detail to reveal the changes. After treating the material with H2O2, a doublet at ~1.53 ppm apparated corresponding to the methyl in β position to –SOx– moieties. Moreover, a high-frequency shift of the peaks corresponding to the protons in the α position to the –S– (2.65 and 2.93 ppm) was also observed. These phenomena indicate that both the sulfoxide and sulphone moieties are included in the backbone of H2O2-treated poly(propylene sulfide). Kramer and Deming further demonstrated the H2O2-sensitive mechanism of thioether, in which the thioether moiety in a pyranose derivative can be oxidized into sulphone-containing product with a high yield [32].

2.2. Disulfide Bond (-S-S-)

Disulfide bonds are commonly included in a variety of proteins, which can be generated from the oxidation of thiol groups. An S anion from one sulfhydryl group can act as a nucleophile and attack another sulfhydryl moiety to create a disulfide bond. Particularly, H2O2 can be used to break the S-S bond and yield corresponding thiolated compounds (Table 1). Due to the high activity of sulfhydryl groups, the degraded thiolated compounds have the capacity of forming the original disulfide-bearing substance. Therefore, the disulfide-based H2O2-responsive behavior can be regarded as a reversible process [33].

2.3. Selenide (-Se-)

Selenium (Se) is a nonmetallic element in the oxygen family that mainly exists in several allotropic forms. It is widely spread in volcanic areas and sulfide ores, which can be used in photocells, solar cells, and in xerography. Selenium can be bound into proteins and serves as an indispensable microelement in humans. Its chemical properties are similar to sulfur, but with a higher reactivity. As a result, the H2O2-treated selenide derivatives are usually transferred into the selenone form (Table 1).
Zhang’s group illustrated the structural change with the oxidation of a selenide block copolymer with H2O2 at 0.1%, and the oxidized residue was analyzed by XPS. The binding energy of Se 3d5’s shift from 55.9 eV to 59.5 eV was clearly observed, suggesting the higher valency of selenium which greatly matches the formation of selenone groups. The presence of O=Se=O groups can also be demonstrated based on the characteristic absorption bands at 904 and 880 cm−1. 1H-NMR and 77Se-NMR further confirmed the transformation from -Se- to O=Se=O groups, which shows a H2O2-responsive behavior [35].

2.4. Diselenium Bond (-Se-Se-)

A diselenium bond can be regarded as a linkage containing two atoms of selenium combined with an element or radical. Due to the similar chemical properties of selenium and sulfur, a diselenium bond can also be broken with treatment from H2O2. The diselenide bond has an energy of around 176 kJ/mol, indicating its easy oxidization in the presence of H2O2 while undergoing a phase transition from hydrophobic diselenide to hydrophilic selenic acid (Table 1) [41]. An attractive advantage of diselenide and disulfide bonds is their hydrolytic stability in physiological conditions [36]. Moreover, the diselenid-based H2O2-responsive behavior can also be regarded as a reversible process. The degraded hydrophilic selenic acid can rebuild the diselenium bond under ambient atmosphere.
Diselenium derivatives can be prepared by the oxidation of aryl selenols [42]. However, as aryl selenols are most conveniently prepared from the diselenides themselves, this method is not likely to have preparative value. Thus, diselenium derivatives are usually synthesized via other routes. For example, electron-deficient aryl halides can be used as substrates [43], and their reaction with the diselenide dianion gives rise to the desired molecules. Arylamines can also serve as potential precursors via the reaction of diazonium salts with sodium hydrogenselenide. Another typical route to diaryl diselenides is based on the reaction of arylaldehydes with sodium hydrogenselenide, which, after treatment with sodium borohydride and piperidine, leads to the target compounds.

2.5. Diketone

Benzil is a yellow crystalline containing a diketone moiety made by oxidizing benzoin [44]. Sawaki et al. have reported that benzil can transform into benzoic anhydride via a Baeyer-Villiger type reaction with H2O2 in alkaline organic solvents [39]. Then the products are usually presented in benzoic form due to hydrolysis (Table 1). However, the reported conditions were far from biological. Nagano’s group developed a simple strategy to adjust the reactivity between benzil and H2O2 based on the modification of the benzene ring [40]. Following this method, 5-benzoylcarbonylfluorescein derivatives were synthesized for the detection of hydrogen peroxide at the cellular level together with the d-PeT mechanism to control fluorescence. Moreover, other peroxalate ester-containing derivatives have been reported to show similar H2O2-responsive properties [45,46,47].

2.6. Boronic

The reaction between H2O2 and benzeneboronic acid in aqueous solutions can generate phenol and boric acid in high yields, which can be viewed as a homolytic orelectrophilic substitution in the benzene ring (Table 1) [48]. Generally, the rate of the reaction follows a dose-dependent manner. Moreover, in addition to benzeneboronic acid, quite a number of aryl-substituted boronic acids were developed, which can be exploited as H2O2-responsive tools with high accuracy, specificity, and sensitivity. Chang’s group proposed a strategy for the optical detection of H2O2 relying on the selective H2O2-triggered transformation of arylboronates to phenols [37]. With the H2O2-triggered hydrolytic deprotection of the boronates, our groups have developed a series of open, colored, and fluorescent products [49,50,51,52].
Interestingly, when the arylboronic acid ester is connected to a suitable linker, e.g., a typical –Ar–CH2–O– linkage, after the formation of a phenol derivative triggered by H2O2, a sequential self-immolative process can take place via a 1,4-/1,6 elimination reaction. This character of arylborates has enabled the development of H2O2-triggered degradable materials [38].

2.7. Others

To meet the increasing demand for H2O2-responsive materials and ROS scavengers, some other types of H2O2-sensitive linkages have been employed. Tellurium (Te)-containing organic compounds have gained attention because of their lower electronegativity, which is thought to be more easily oxidized by H2O2 than selenium/sulfur–containing ones [53]. Moreover, tellurium-containing compounds were evidenced to exhibit less toxicity than selenium analogues. It has been reported that the tellurium atoms in Te-containing organic compounds can be completely oxidized with 100 μM H2O2.
Peptides and enzymes have also been used to fabricate H2O2-responsive materials. Sung et al. reported a peptide oligomer as a H2O2-responsive crosslinker for long-term tissue engineering applications. The degradation can take place by reacting with 5 mM H2O2 [54,55].
Otherwise, some metallic oxides and metal salts can be used as H2O2-responsive moieties owing to their catalytic action or redox activity. For example, the catalytic reaction between the H2O2 and MnO2 contributes to the formation of H2O and O2 [56]. Fe(II) can react with H2O2 to generate hydroxyl radicals [57].

3. Preparation

Generally, hydrogels fall into two categories [58,59,60]:
(i)
Chemical gel: Hydrogels that are covalently cross-linked networks in which the equilibrium swelling state depends on the crosslink density and interaction parameters [61,62];
(ii)
Physical gel: The networks are constructed by noncovalent linkages, e.g., hydrogen bonding, ionic bonding, hydrophobic interactions, molecular entanglements, etc. These interactions are reversible.
As for H2O2-responsisve hydrogels, the responsiveness is mainly based on the linkages summarized in the abovementioned section. Such responsiveness is quite different from a response to temperature, pH, or electric fields, and the latter of these is mainly presented in physical gels. As a result, H2O2-responsisve hydrogels are mainly presented as chemical gels in the literature, in which a chemical cross-linking approach is usually involved. The cross-linking approach can be classified into two generic groups.

3.1. Self-Crosslinking

For this case, the H2O2-Gels were usually prepared based on the self-crosslinking behavior of the polymeric chains without the addition of any other crosslinkers. The H2O2-sensitive moieties, e.g., sulfydryl, selenol, phenol, etc., are mainly incorporated in the polymeric chains as pendent or terminal groups which can be further involved in the condensation reactions to achieve the three-dimensional structures.
Wan’s group synthesized a sulfydryl-bearing oligomer (o-DHLA) by treating dihydrolipoic acid with equimolar 2,2-dimethoxypropane under the catalysis of p-toluenesulfonic acid [63]. The obtained o-DHLA was achieved with sulfydryl units as terminal groups, the latter of which can react with each other to yield –S–S– containing microgels with the addition of polyvinyl alcohol. The formation of an –S–S– linkage makes the obtained microgels desirable H2O2–responsive drug delivery systems. Yang’s group employed cysteine-containing keratin as the base material, in which the sulfydryl units can afford –S–S– linkages under the effect of oxygen [64].
Ma et al. [65] prepared a boronate-based copolymer comprising N-isopropylacrylamide, hydroxyethyl methacrylate, and (4-(hydroxymethyl)-phenylboronic acid. The presence of N-isopropylacrylamide endowed the obtained copolymer’s temperature-responsive characters. As a result, the sol-gel transformation occurs when the temperature is higher than a certain value ranging from 18 to 37 °C. The lower critical solution temperature (LCST) and upper critical solution temperature (UCST) can be controlled by the proportion of N-isopropylacrylamide. The presence of boronate moieties makes the obtained hydrogel exhibit H2O2–responsive characters. A similar strategy can also be found in another patent [66].

3.2. Addition of Difunctional Crosslinkers

Zhao et al. [67] prepared H2O2–responsive hydrogels using cystamine dihydrochloride as the difunctional crosslinkers. In their strategy, catecholamine-modified chitosan was prepared and used to fabricate the hydrogels. The amino groups in cystamine can react with catecholamine moieties via the formation of benzoquinone. As a result, the disulfide bonds were introduced into the matrix of the obtained hydrogels, which endowed the hydrogels with H2O2–responsive degradation abilities.
Our group proposed an alternative approach to synthesizing H2O2–responsive hydrogels, in which the difunctional crosslinkers and polyhydric matrix are used. For example, the arylboronic acid unit on 1,4-phenylenediboronic acid can react with the vicinal diol groups of a polysaccharide to form a chemical hydrogel with high efficiency. The B−C bond in the obtained hydrogel can be broken with the addition of H2O2 in a dose-dependent manner, which can give rise to the degradation behavior and endow the H2O2–responsive behavior to the hydrogel (Figure 1) [50]. The boronate-based difunctional crosslinkers have been also reported in other studies [68].

4. Application

4.1. Drug Delivery

Drug-loaded hydrogels have generated increasing interest in recent years because of their extraordinary advantages in versatile drug administration, ease of preparation, high biocompatibility, and improved patient compliance [69]. Moreover, the stimuli-responsiveness can be achieved by including the responsive linkages into the hydrogel networks, e.g., temperature-responsive, pH-responsive, photo-responsive, etc. H2O2 is usually involved in a variety of pathological effects in organisms, in which the aberrant expression of H2O2 is directly linked with cancer, inflammation, phyma, etc. The exploration of H2O2-responsive drug delivery systems with high selectivity and specificity has great significance for improving antitumor efficacy and preventing inflammation.
In Wan’s study, a serious of thioketal-containing microgels have been developed that show a dual responsive behavior to H2O2 and glutathione. Paclitaxel was loaded into the dual responsive thioketal microgels to achieve an antitumor preparation which is sensitive enough to the tumor-site H2O2 and glutathione levels and able to release paclitaxel accordingly. Moreover, the obtained dual responsive thioketal microgels can be degradable into biocompatible and low toxic byproducts after performing their delivering tasks. Following this method, the antitumor efficacy of chemotherapy drugs can be significantly improved by loading them into the dual responsive platform. Meanwhile, side effects can also be effectively reduced [63].
Duvall et al. [70] explored poly(propylene sulfide) nanoparticles and hydrogels as H2O2-responsive drug carriers. Generally, the conversion between sulfoxides/sulfones and hydrophobic sulfide can cause solubility changes. The hydrophobicity of poly(propylene sulfide) makes it a good candidate for loading hydrophobic drugs, similar to curcumin. After delivery in vivo, the exposure of poly(propylene sulfide) to H2O2 enables a transition to its hydrophilic form, which triggers particle swelling and H2O2-responsive drug release.
Our group developed H2O2-Gels by using difunctional crosslinkers. A high level of H2O2 can destroy the B-C bond in the obtained hydrogels and thereby break the three-dimensional networks in H2O2-Gels. When the drug was loaded into the hydrogel, a typical H2O2-responsive behavior was easily achieved. It should be noted that hydrogels, which are diffusion-controlled drug delivery systems, usually exhibit an initial burst release. Since drug-release during the initial stage depends mostly on diffusional escape, a sudden release in a short period usually takes place. It can be regarded as an “initial burst release” [71,72]. A serious initial burst release may be toxic and give rise to suppressed bioavailability. In our study, the model drug was poured into halloysite nanotubes, a natural aluminosilicate clay mineral with a hollow tubular structure, before loading it into the matrix of H2O2-Gels (Figure 2, DLHCH-1: the therapeutic agent was pre-loaded into the halloysite nanotubes; DLHCH-2: the therapeutic agent was directly loaded into the matrix of hydrogels). Following this method, the “initial burst effect” was effectively suppressed due to the halloysite nanotubes. Our strategy provides an alternative way to design H2O2-Gels which may meet the requirements for the application of topical preparations to prevent inflammation (Figure 2) [50].
Zhen et al. focused on exploiting the boronate-based polymeric network as microneedles, which can be used as skin patches for acne vulgaris treatment. Due to the presence of boronate moieties, the antibiotic loaded composition possesses a H2O2-responsive release behavior [68].
There numerous reports focused on the development of H2O2-Gel as drug delivery systems [73,74]. However, there is still a lack of commercially available products. Further studies should pay attention to this point. Some other studies are still required. The potential side effects of the degraded products should also be investigated. Additionally, the specificity in non- enzyme systems should also be improved since the body fluid is a very complicated system.

4.2. Detection

H2O2 has been demonstrated to be a metabolite of many biochemical reactions. Qualitative H2O2 detection is an important tool which can be used to discover early lesions such as those present in cancer. As typical soft materials, hydrogels show a biophysical similarity to soft biological tissues and thereby possess a desirably good affinity with them. Moreover, the chemosensor moieties can be covalently linked or physically incorporated into the matrix of hydrogels to endow the product with various diagnostic applications. In past decades, studies on utilizing H2O2-Gels as detection tools were reported [75,76].

4.2.1. Electrical

Tian’s group developed a hydrogel-based biosensor system by incorporating Cytochrome c into the matrix of hydrogel consisting of Fmoc-L-lysine, Fmoc-L-phenylalanine, and sodium carbonate, in which the inherent bioactive activity of Cytochrome c was retained. After treatment with H2O2, the biochemical reaction between Cytochrome c and H2O2 can result in a fast and sensitive change in the redox formal potential. Moreover, the cellular level H2O2 can be quantitively determined. Such H2O2-responsive behavior was demonstrated to exhibit excellent specificity over commonly used ions, ascorbic acid, and other ROS. The good stability and reproducibility, as well as the high specificity and sensitivity, may pave a new path to further understanding the role of H2O2 in pathological changes [18].
Li et al. prepared luminophore-incorporated polyaniline-polyacrylamide hydrogel that exhibited favorable biocompatibility and good conductivity. The obtained hydrogel showed a high detection sensitivity upon H2O2 based on an electrochemiluminescence method. The limit of detection of H2O2 was as low as 2.9 × 10−9 M. The detection concentration ranged from 5 × 10−5 to 1 × 10−8 M. The results indicated that the prepared H2O2-Gel can be utilized to detect cellular H2O2 [77].
The electrochemical detection of H2O2 can also be achieved by constructing Hemin-G4/Au-containing composite hydrogels. In Hou’s study [78], the excellent catalytic activity of Hemin-G4 and the high conductivity of gold nanoparticles were employed to endow the obtained hydrogels with a good electrochemical response towards H2O2. Moreover, A549 cells can adhere to the surface of the obtained hydrogels, which can be used to detect extracellular H2O2.
Electrical detection methods usually show low limit of detection (LOD) levels. The enzyme-incorporated ones exhibit good specificity, making them desirable detection tools in bioanalysis. Nevertheless, there is still a need for other detention methods that are low-cost and free of energy.

4.2.2. Fluorescent

Hydrogels with a fluorescent response to H2O2 are fundamentally important in biology and pathophysiology which have practical value in early clinical diagnosis. In Yang’s study [79], an intensive chemiluminescence hydrogel was proposed by incorporating luminol and hemin into the matrix of guanosine-derived hydrogel. The obtained composite hydrogel demonstrated enzyme-like activity towards the H2O2-mediated oxidation of luminol, in which the blue light can be visibly observed.
Our group synthesized a novel type of fluorescein-derived fluorescence probe by installing boronic acid groups at the 3′ and 6′-positions of the xanthenone moiety in fluorescein. The obtained fluorescein derivative displayed almost no emission behavior (Φf = ca.0.01). After treatment with H2O2, the fluorescein derivative can be degraded into fluorescein with strong fluorescence (Φf = 0.94) [80]. The “turn-on” response in fluorescence resulted in a H2O2-detecting ability. Moreover, the two arylboronic acid groups in fluorescein derivative can react to diol units, which makes it a desirable crosslinker to fabricate the H2O2-responsive fluorescein change in the obtained hydrogels. Following this method, hydrogels with a “turn-on” fluorescence behavior upon H2O2 administration were developed (Figure 3A,B). The obtained hydrogel shows no emission behavior under the physiological H2O2 concentration, while a fast non-fluorescent to fluorescent change can be achieved under pathological H2O2 concentration [51].
Concerning the lack of an indicator to monitor the release behavior in real time for H2O2-Gels, the obtained hydrogels were also explored as drug carriers. As shown in Figure 3C, a very small amount of agent was released from DHNTs@PVA@PA under a low level of H2O2 (0.02 μM), while the 4-fold increased initial-burst amount can be tracked for PTX@PVA@PA (DHNTs@PVA@PA: the therapeutic agent was pre-loaded into the halloysite nanotubes; PTX@PVA@PA: the therapeutic agent was directly loaded into the matrix of hydrogels). The significant increase of the release rates can be observed in 200 μM H2O2 and compared to those in 0.02 μM H2O2, suggesting a typical H2O2-responsive release behavior. Figure 3D shows a remarkable increase of the fluorescence intensity over time from 5 to 40 min. The relationship between the release rate and fluorescence intensity was investigated and displayed in Figure 3E. The good relationship is quite beneficial to invisibly monitoring the drug release behavior.

4.2.3. Colorimetric

As a typical pH indicator, phenolphthalein solution can present a red color under an alkalescence environment based on the transformation between the phenolic hydroxyl groups and the benzoquinone unit. In our study, this colorimetric behavior was employed for the visual detection of H2O2. Boronic acid pinacol ester groups were used to substitute the hydrogel groups in phenolphthalein. As a result, the transformation between the phenolic hydroxyl groups and benzoquinone unit was blocked. Then, the phenolphthalein derivative was used to crosslink polyvinyl alcohol chains in a weak basic solution to yield a H2O2-Gel based on the dynamic covalent bonds. After treatment with H2O2 under a pathological concentration, the B–C linkage in the network can be rapidly degraded into Ar-OH groups and B-OH groups result in the formation of phenolphthalein. Since the hydrogel was prepared under an alkalescence environment, the obtained hydrogel showed a distinct colorimetric response to H2O2 (Figure 4) [52].
Yang et al. [81] reported a 2D photonic crystal-based horseradish peroxidase/bovine serum albumin (HRP/BSA) composite hydrogel. BSA and HRP was crosslinked by glutaraldehyde, in which BSA and HRP acts as a scaffold and a recognition unit, respectively. The obtained hydrogel can monitor H2O2 since HRP can selectively decompose H2O2 along with heme inactivation. The conformational change of HRP can decrease the crosslinking density and enlarge the particle spacing. As a result, the Debye ring of the 2-D photonic crystals also decreased, which provided a colorimetric response to H2O2 for the obtained hydrogel.
Compared to electrical and fluorescent methods, the colorimetric method shows a superior low cost, an easy operation, and operates free of energy. Moreover, the latter method can be used without any sophisticated or expensive analytical instruments, which makes it a desirable detection tool for actual production and daily life.

4.3. Glucose-Responsive

Diabetes is regarded as a serious metabolic disorder, which affects more than one-tenth of the population in the world [82,83]. Long-term high glucose levels can result in serious cardiovascular disease, renal failure, abnormal bone metabolism, and many other complications. The glucose-controlled insulin release system is usually required in the treatment of diabetes, especially for suppressing diabetes ulcers [84]. Among these systems, glucose oxidase is an important intermediary for the transformation of glucose signals into the generation of H2O2 [85]:
Glucose + O 2 Glucose   oxidase H 2 O 2 + Gluconic   acid
Therefore, with the addition of glucose oxidase, H2O2-Gel can also be used as glucose-responsive systems. Kiritoshi et al. prepared H2O2-Gel via a typical radical copolymerization of 2-methacryloyloxyethyl phosphorylcholine and triethylene glycol dimethacrylate. The obtained polymer can be easily degraded in the presence of H2O2 (974 mmol/L). In Kiritoshi’s strategy, the glucose oxidase can be loaded into the matrix of the obtained hydrogel without a loss in catalytic activity. Following this method, the obtained H2O2-Gel is also available in glucose-responsive insulin delivery systems in a concentration-dependent manner [86].
In past decades, an increasing number of studies have focused on transforming H2O2-Gels into glucose-responsive systems with the addition of glucose oxidase [64,87]. For example, Yang et al. developed a hydrogel that formed in situ by loading glucose oxidase into keratin hydrogels, which showed promising applications for the treatment of diabetic wounds [64].

4.4. ROS Scavengers

An excessive ROS level can raise the oxidative damage to cellular macromolecules and thereby deteriorate many inflammation-related diseases. Moreover, the ROS accumulated in a wound can induce strong inflammation and serve as a strong barrier that inhibits tissue regeneration, which is a puzzling problem for diabetic ulcers [88]. Some H2O2-Gels are promising ROS scavengers in clinical applications [35,89], especially those that have the capacity to consume H2O2 in a concentration-dependent manner [90].
Liu’s group developed a type of ROS-scavenger by using a difunctional arylboronic derivative as crosslinker [91]. The obtained H2O2-Gels can act as an effective ROS-scavenger agent to promote wound-healing by reducing the ROS level around the wound, which is quite beneficial to inhibiting aseptic inflammation. Moreover, the H2O2-Gel was also explored as a drug carrier to allow the release of therapeutic agents, such as antibacterial agents and growth factors. As a result, the H2O2-Gel supported a systemic therapy strategy to promote wound-healing in complicated infections.
Gao et al. presented a novel type of thioether-containing injectable hydrogel with H2O2-responsive behaviors, which can be used as ROS-scavengers for myocardial infarction treatment [92]. The resulting animal model showed that the obtained H2O2-Gels can significantly consume excessive ROS, accelerate angiogenesis, and improve cardiac functions.

4.5. Tissue Engineering

Hydrogels have been demonstrated to be a desirable candidate for cartilage regenerative substrates owing to their attractive advantages including mild reaction conditions and tunable structures. The controllable gelatinizing process, especially for the gelation time, is the key issue when hydrogel is used for articular cartilage engineering. Rahbarghazi et al. developed a library of alginate-based hydrogels crosslinked via horseradish peroxidase. Generally, horseradish peroxidase can serve as an efficient enzyme catalysis in catalyzing the conversion from H2O2 to H2O and O2 [93]:
H 2 O 2 Horseradish   peroxidase 2 H 2 O + O 2
As a result, the gelation time can be well-controlled based on the concentration of H2O2, which is a result of the deactivation effect of horseradish peroxidase by H2O2 [94].
It is worth noting that the tissue engineering materials are usually composed of different parts to meet the demand in terms of therapeutics. Liu’s group developed an injectable, post-trauma microenvironment-responsive, H2O2 depleting, and drug-loaded hydrogel [95], which can be used to lower the H2O2 levels in damaged brain tissue. The poly (propylene sulfide)-containing matrix endowed the obtained hydrogel with H2O2-responsive abilities. The released curcumin can further eliminate the ROS, promoting the regeneration and recovery of neurons. Following this method, a systematic and multi-channel promoting effect can be achieved, which is quite useful in tissue engineering materials.

4.6. Cell-Encapsulation

Hydrogel particles have been studied in the latest decade for encapsulating and delivering cells in humans, which meets many of the demands for treating diseases and producing antibodies for vaccine use. The main advantage of hydrogel-based cell-encapsulation technology is a mild reaction condition which does not threaten the lives of the enclosed cells.
Sakai et al. reported a facile approach for preparing mammalian cell-encapsulating hydrogel particles based on the horseradish peroxidase-crosslinked hydrogelation [96]. Gelatin bearing phenolic hydroxyl moieties were employed as the polymeric substrate and dissolved in a cell-suspending solution containing horseradish peroxidase. The mixed solution was then dropped into 1 mM H2O2 to form hydrogels with a spherical shape. The α-H in phenolic hydroxyl moieties usually exhibits high activity with the presence of horseradish peroxidase and H2O2 [19]. These phenomena suggested that the H2O2-responsive behavior can be used to establish a cell-encapsulation method in hydrogel particles.
In Sakai’s proposed strategy [19], the individual cell can be encapsulated into a thin hydrogel sheath. This approach can be controlled by the addition of HRP and exhibits no significant adverse effects to the encapsulated cells. Apparently, the proposed approach based on the H2O2-responsive behavior is expected to be used to extend the applications of cell encapsulation technology.

5. Conclusions and Perspective

Recent studies of H2O2-Gels have been reviewed in this paper. The summarized H2O2-responsive linkages and preparation methods may serve as guides for researchers who are interested in H2O2-Gels. We expect to provide guidance and designate the direction of further studies on H2O2-Gels. In combination with drug delivery, the detection of H2O2, glucose-responsive systems, ROS scavengers, tissue engineering, and cell-encapsulation, H2O2-Gels could be advanced to have much more promising applications in biomedicine. Nevertheless, commercially available H2O2-Gels, as well as official standards or criteria, are also urgently needed. There are still some limitations for this issue. The level of H2O2 usually differs among individuals, and the complexity of bodily fluids should also be considered. Therefore, it is far more difficult to establish in vitro and in vivo correlation (IVIVC), which is rarely studied but indeed remains a key issue that should be focused on. On the other hand, costly proteinases seriously limit large-scale production. Non-proteinase ones usually suffer from other problems such as stability, complexity in the synthesis approach, the potential toxicity of the degraded by-products, etc. Thus, longer-duration studies are encouraged to conduct deeper investigations to make H2O2-Gels “over the counter”.

Author Contributions

Conceptualization, H.Z.; data collection: W.S., J.Y., Y.Z., Q.Y., J.J.; Original draft preparation: W.S. and H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Project of Science & Technology Bureau of Baoding City (no.1941ZF081).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lavrador, P.; Esteves, M.R.; Gaspar, V.M.; Mano, J.F. Stimuli-responsive nanocomposite hydrogels for biomedical applications. Adv. Funct. Mater. 2021, 31, 2005941. [Google Scholar] [CrossRef]
  2. Panja, S.; Adams, D.J. Stimuli responsive dynamic transformations in supramolecular gels. Chem. Soc. Rev. 2021, 50, 5165–5200. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, P.; Zhu, F.; Zhang, Z.; Cheng, Y.; Wang, Z.; Li, Y. Stimuli-responsive polydopamine-based smart materials. Chem. Soc. Rev. 2021, 50, 8319–8343. [Google Scholar] [CrossRef] [PubMed]
  4. Meng, J.; Jin, Z.; Zhao, P.; Zhao, B.; Fan, M.; He, Q. A multistage assembly/disassembly strategy for tumor-targeted co delivery. Sci. Adv. 2020, 6, eaba1362. [Google Scholar] [CrossRef] [PubMed]
  5. Li, Q.; Wang, D.; Fang, X.; Zong, B.; Liu, Y.; Li, Z.; Mao, S.; Ostrikov, K.K. Rapid synthesis of multifunctional beta-cyclodextrin nanospheres as alkali-responsive nanocarriers and selective antibiotic adsorbents. Chem. Commun. 2021, 57, 1161–1164. [Google Scholar] [CrossRef] [PubMed]
  6. Guo, Y.; Bae, J.; Fang, Z.; Li, P.; Zhao, F.; Yu, G. Hydrogels and hydrogel-derived materials for energy and water sustainability. Chem. Rev. 2020, 120, 7642–7707. [Google Scholar] [CrossRef]
  7. Gao, G.; Yang, F.; Zhou, F.; He, J.; Lu, W.; Xiao, P.; Yan, H.; Pan, C.; Chen, T.; Wang, Z.L. Bioinspired self-healing human-machine interactive touch pad with pressure-sensitive adhesiveness on targeted substrates. Adv. Mater. 2020, 32, 2004290. [Google Scholar] [CrossRef]
  8. Liu, Z.; Liu, J.; Cui, X.; Wang, X.; Zhang, L.; Tang, P. Recent advances on magnetic sensitive hydrogels in tissue engineering. Front. Chem. 2020, 8, 124. [Google Scholar] [CrossRef]
  9. Lea-Banks, H.; Meng, Y.; Wu, S.-K.; Belhadjhamida, R.; Hamani, C.; Hynynen, K. Ultrasound-sensitive nanodroplets achieve targeted neuromodulation. J. Control. Release 2021, 332, 30–39. [Google Scholar] [CrossRef]
  10. Lin, X.; Liu, S.; Zhang, X.; Zhu, R.; Chen, S.; Chen, X.; Song, J.; Yang, H. An ultrasound activated vesicle of janus au-mno nanoparticles for promoted tumor penetration and sono-chemodynamic therapy of orthotopic liver cancer. Angew. Chem. Int. Ed. 2020, 59, 1682–1688. [Google Scholar] [CrossRef]
  11. Houstis, N.; Rosen, E.D.; Lander, E.S. Reactive oxygen species have a causal role in multiple forms of insulin resistance. Nature 2006, 440, 944–948. [Google Scholar] [CrossRef] [PubMed]
  12. Fubini, B.; Hubbard, A. Reactive oxygen species (ROS) and reactive nitrogen species (RNS) generation by silica in inflammation and fibrosis. Free Radical Bio. Med. 2003, 34, 1507–1516. [Google Scholar] [CrossRef]
  13. Wang, S.; Wang, Z.; Yu, G.; Zhou, Z.; Jacobson, O.; Liu, Y.; Ma, Y.; Zhang, F.; Chen, Z.-Y.; Chen, X. Tumor-specific drug release and reactive oxygen species generation for cancer chemo/chemodynamic combination therapy. Adv. Sci. 2019, 6, 1801986. [Google Scholar] [CrossRef]
  14. Wang, Y.Q.; Li, L.L.; Zhao, W.B.; Dou, Y.; An, H.J.; Tao, H.; Xu, X.Q.; Jia, Y.; Lu, S.; Zhang, J.X.; et al. Targeted therapy of atherosclerosis by a broad-spectrum reactive oxygen species scavenging nanoparticle with intrinsic anti-inflammatory activity. ACS Nano 2018, 12, 8943–8960. [Google Scholar] [CrossRef] [PubMed]
  15. Rhee, S. H2O2, a necessary evil for cell signaling. Science 2006, 312, 1882–1883. [Google Scholar] [CrossRef]
  16. Sies, H. Role of metabolic h2o2 generation: Redox signaling and oxidative stress. J. Biol. Chem. 2014, 289, 8735–8741. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, J.; Zhang, Y.; Archibong, E.; Ligler, F.S.; Zhen, G. Leveraging H2O2 levels for biomedical applications. Adv. Biosyst. 2017, 1, 1700084. [Google Scholar] [CrossRef]
  18. Zhou, J.; Liao, C.; Zhang, L.; Wang, Q.; Tian, Y. Molecular hydrogel-stabilized enzyme with facilitated electron transfer for determination of h2o2 released from live cells. Anal. Chem. 2014, 86, 4395–4401. [Google Scholar] [CrossRef]
  19. Cheng, Y.; Dai, J.; Sun, C.L.; Liu, R.; Zhai, T.Y.; Lou, X.D.; Xia, F. An intracellular H2O2-responsive aiegen for the peroxidase-mediated selective imaging and inhibition of inflammatory cells. Angew. Chem. Int. Edit. 2018, 57, 3123–3127. [Google Scholar] [CrossRef]
  20. Saxon, E.; Peng, X. Recent advances in hydrogen peroxide responsive organoborons for biological and biomedical applications. ChemBioChem 2022, 23, e202100366. [Google Scholar] [CrossRef]
  21. Yang, N.; Xiao, W.; Song, X.; Wang, W.; Dong, X. Recent advances in tumor microenvironment hydrogen peroxide-responsive materials for cancer photodynamic therapy. Nano-Micro Lett. 2020, 12, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Bastiancich, C.; Danhier, P.; Preat, V.; Danhier, F. Anticancer drug-loaded hydrogels as drug delivery systems for the local treatment of glioblastoma. J. Control. Release 2016, 243, 29–42. [Google Scholar] [CrossRef] [PubMed]
  23. Daly, A.C.; Riley, L.; Segura, T.; Burdick, J.A. Hydrogel microparticles for biomedical applications. Nat. Rev. Mater. 2020, 5, 20–43. [Google Scholar] [CrossRef]
  24. Liu, X.; Liu, J.; Lin, S.; Zhao, X. Hydrogel machines. Mater. Today 2020, 36, 102–124. [Google Scholar] [CrossRef]
  25. Yang, J.; Bai, R.; Chen, B.; Suo, Z. Hydrogel adhesion: A supramolecular synergy of chemistry, topology, and mechanics. Adv. Funct. Mater. 2020, 30, 1901693. [Google Scholar] [CrossRef]
  26. Yoshii, T.; Onogi, S.; Shigemitsu, H.; Hamachi, I. Chemically reactive supramolecular hydrogel coupled with a signal amplification system for enhanced analyte sensitivity. J. Am. Chem. Soc. 2015, 137, 3360–3365. [Google Scholar] [CrossRef]
  27. Xu, Q.; He, C.; Ren, K.; Xiao, C.; Chen, X. Thermosensitive polypeptide hydrogels as a platform for ROS-triggered cargo release with innate cytoprotective ability under oxidative stress. Adv. Healthc. Mater. 2016, 5, 1979–1990. [Google Scholar] [CrossRef]
  28. Mateen, R.; Ali, M.M.; Hoare, T. A printable hydrogel microarray for drug screening avoids false positives associated with promiscuous aggregating inhibitors. Nat. Commun. 2018, 9, 602. [Google Scholar] [CrossRef] [Green Version]
  29. Tabor, A.B. The challenge of the lantibiotics: Synthetic approaches to thioether-bridged peptides. Org. Biomol. Chem. 2011, 9, 7606–7628. [Google Scholar] [CrossRef]
  30. Kholdeeva, O.A.; Maksimov, G.M.; Maksimovskaya, R.I.; Kovaleva, L.A.; Hill, C.L. A dimeric titanium-containing polyoxometalate. Synthesis, characterization, and catalysis of H2O2-based thioether oxidation. Inorg. Chem. 2000, 39, 3828–3837. [Google Scholar] [CrossRef]
  31. Napoli, A.; Valentini, M.; Tirelli, N.; Müller, M.; Hubbell, J.A. Oxidation-responsive polymeric vesicles. Nat. Mater. 2004, 3, 183–189. [Google Scholar] [CrossRef] [PubMed]
  32. Kramer, J.R.; Deming, T.J. Glycopolypeptides with a redox-triggered helix-to-coil transition. J. Am. Chem. Soc. 2012, 134, 4112–4115. [Google Scholar] [CrossRef]
  33. Miller, E.W.; Bian, S.X.; Chang, C.J. A fluorescent sensor for imaging reversible redox cycles in living cells. J. Am. Chem. Soc. 2007, 129, 3458–3459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Xu, Q.; Venet, M.; Wang, W.; Creagh-Flynn, J.; Wang, X.; Li, X.; Gao, Y.; Zhou, D.; Zeng, M.; Lara-Saez, I.; et al. Versatile hyperbranched poly(beta-hydrazide ester) macromers as injectable antioxidative hydrogels. ACS Appl. Mater. Interfaces 2018, 10, 39494–39504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Ma, N.; Li, Y.; Ren, H.; Xu, H.; Li, Z.; Zhang, X. Selenium-containing block copolymers and their oxidation-responsive aggregates. Polym. Chem. 2010, 1, 1609–1614. [Google Scholar] [CrossRef]
  36. Deepagan, V.G.; Kwon, S.; You, D.G.; Nguyen, V.Q.; Park, J.H. In situ diselenide-crosslinked polymeric micelles for ROS-mediated anticancer drug delivery. Biomaterials 2016, 103, 56–66. [Google Scholar] [CrossRef]
  37. Chang, M.; Pralle, A.; Isacoff, E.Y.; Chang, C.J. A selective, cell-permeable optical probe for hydrogen peroxide in living cells. J. Am. Chem. Soc. 2004, 126, 15392–15393. [Google Scholar] [CrossRef] [Green Version]
  38. Broaders, K.E.; Grandhe, S.; Fréchet, J.M. A biocompatible oxidation-triggered carrier polymer with potential in therapeutics. J. Am. Chem. Soc. 2011, 133, 756–758. [Google Scholar] [CrossRef]
  39. Sawaki, Y.; Foote, C.S. Acyclic mechanism in the cleavage of benzils with alkaline hydrogen peroxide. J. Am. Chem. Soc. 1979, 101, 6292–6296. [Google Scholar] [CrossRef]
  40. Abo, M.; Urano, Y.; Hanaoka, K.; Terai, T.; Komatsu, T.; Nagano, T. Development of a highly sensitive fluorescence probe for hydrogen peroxide. J. Am. Chem. Soc. 2011, 133, 10629. [Google Scholar] [CrossRef]
  41. Chivers, T. Main-group elements, including noble gases. In Comprehensive Inorganic Chemistry II; Elsevier: Amsterdam, The Netherlands, 2013. [Google Scholar]
  42. Etal, A. Synthesis: Carbon with one heteroatom attached by a single bond. In Comprehensive Organic Functional Group Transformations III; Taylor: Oxfordshire, UK, 1995. [Google Scholar]
  43. Wang, Y.; Lv, M.-Z.; Song, N.; Liu, Z.-J.; Wang, C.; Yang, Y.-W. Dual-stimuli-responsive fluorescent supramolecular polymer based on a diselenium-bridged pillar 5 arene dimer and an AIE-active tetraphenylethylene guest. Macromolecules 2017, 50, 5759–5766. [Google Scholar] [CrossRef]
  44. Zhang, K.-M.; Dou, W.; Li, P.-X.; Shen, R.; Ru, J.-X.; Liu, W.; Cui, Y.-M.; Chen, C.-Y.; Liu, W.-S.; Bai, D.-C. A coumarin-based two-photon probe for hydrogen peroxide. Biosens. Bioelectron. 2015, 64, 542–546. [Google Scholar] [CrossRef] [PubMed]
  45. Lee, Y.-D.; Lim, C.-K.; Singh, A.; Koh, J.; Kim, J.; Kwon, I.C.; Kim, S. Dye/peroxalate aggregated nanoparticles with enhanced and tunable chemiluminescence for biomedical imaging of hydrogen peroxide. ACS Nano 2012, 6, 6759–6766. [Google Scholar] [CrossRef] [PubMed]
  46. Seo, Y.H.; Singh, A.; Cho, H.-J.; Kim, Y.; Heo, J.; Lim, C.-K.; Park, S.Y.; Jang, W.-D.; Kim, S. Rational design for enhancing inflammation-responsive in vivo chemiluminescence via nanophotonic energy relay to near-infrared aie-active conjugated polymer. Biomaterials 2016, 84, 111–118. [Google Scholar] [CrossRef] [PubMed]
  47. Liu, C.; Zhu, X.; Duan, J.; Liang, X.; Li, X.; Yang, J. Hydrogen peroxide-responsive peroxalate ester-linked PCL-PEG micelles as drug carrier. J. Control. Release 2017, 259, e17. [Google Scholar] [CrossRef]
  48. Cui, Y.; Zhang, M.; Du, F.S.; Li, Z.C. Facile synthesis of h2o2-cleavable poly(ester-amide)s by passerini multicomponent polymerization. ACS Macro Lett. 2016, 6, 11–15. [Google Scholar] [CrossRef]
  49. Zhang, H.; Ren, T.; Ji, Y.; Han, L.; Wu, Y.; Song, H.; Bai, L.; Ba, X. Selective modification of halloysite nanotubes with 1-pyrenylboronic acid: A novel fluorescence probe with highly selective and sensitive response to hyperoxide. ACS Appl. Mater. Interfaces 2015, 7, 23805–23811. [Google Scholar] [CrossRef]
  50. Liu, F.; Bai, L.; Zhang, H.; Song, H.; Hu, L.; Wu, Y.; Ba, X. Smart H2O2-responsive drug delivery system made by halloysite nanotubes and carbohydrate polymers. ACS Appl. Mater. Interfaces 2017, 9, 31626–31633. [Google Scholar] [CrossRef]
  51. Cheng, C.; Gao, Y.; Song, W.; Zhao, Q.; Zhang, H.; Zhang, H. Halloysite nanotube-based H2O2-responsive drug delivery system with a turn on effect on fluorescence for real-time monitoring. Chem. Eng. J. 2020, 380, 122474. [Google Scholar] [CrossRef]
  52. Tang, B.; Zhang, H.; Cheng, C.; Jiang, H.; Bai, L.; Ba, X.; Wu, Y. Development of halloysite nanotube-based hydrogel with colorimetric H2O2-responsive character. Appl. Clay Sci. 2021, 212, 106230. [Google Scholar] [CrossRef]
  53. Li, F.; Li, T.; Cao, W.; Wang, L.; Xu, H. Near-infrared light stimuli-responsive synergistic therapy nanoplatforms based on the coordination of tellurium-containing block polymer and cisplatin for cancer treatment. Biomaterials 2017, 133, 208–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Yu, S.S.; Koblin, R.L.; Zachman, A.L.; Perrien, D.S.; Hofmeister, L.H.; Giorgio, T.D.; Sung, H.-J. Physiologically relevant oxidative degradation of oligo(proline) cross-linked polymeric scaffolds. Biomacromolecules 2011, 12, 4357–4366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Lee, H.S.; Boire, T.C.; Lee, J.B.; Gupta, M.K.; Zachman, A.L.; Rath, R.; Sung, H.-J. ROS-cleavable proline oligomer crosslinking of polyca pro lactone for pro-angiogenic host response. J. Mater. Chem. B 2014, 2, 7109–7113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ye, Y.; Sun, X.; Zhang, Y.; Han, X.; Sun, X. A novel cell-based electrochemical biosensor based on mno2 catalysis for antioxidant activity evaluation of anthocyanins. Biosens. Bioelectron. 2022, 202, 113990. [Google Scholar] [CrossRef]
  57. Xie, M.; Liu, X.; Wang, S. Degradation of methylene blue through fenton-like reaction catalyzed by MoS2-doped sodium alginate/fe hydrogel. Colloids Surf. B 2022, 214, 112443. [Google Scholar] [CrossRef]
  58. Sallach, R.E.; Cui, W.; Wen, J.; Martinez, A.; Conticello, V.P.; Chaikof, E.L. Elastin-mimetic protein polymers capable of physical and chemical crosslinking. Biomaterials 2009, 30, 409–422. [Google Scholar] [CrossRef] [Green Version]
  59. Leonardi, A.B.; Fasce, L.A.; Zucchi, I.A.; Hoppe, C.E.; Soule, E.R.; Perez, C.J.; Williams, R.J.J. Shape memory epoxies based on networks with chemical and physical crosslinks. Eur. Polym. J. 2011, 47, 362–369. [Google Scholar] [CrossRef]
  60. Xu, J.; Liu, X.; Ren, X.; Gao, G. The role of chemical and physical crosslinking in different deformation stages of hybrid hydrogels. Eur. Polym. J. 2018, 100, 86–95. [Google Scholar] [CrossRef]
  61. Nezhad-Mokhtari, P.; Ghorbani, M.; Roshangar, L.; Rad, J.S. Chemical gelling of hydrogels-based biological macromolecules for tissue engineering: Photo- and enzymatic-crosslinking methods. Int. J. Biol. Macromol. 2019, 139, 760–772. [Google Scholar] [CrossRef]
  62. Annaka, M.; Ogata, Y.; Nakahira, T. Swelling behavior of covalently cross-linked gellan gels. J. Phys. Chem. B 2000, 104, 6755–6760. [Google Scholar] [CrossRef]
  63. Chen, D.; Zhang, G.; Li, R.; Guan, M.; Wang, X.; Zou, T.; Zhang, Y.; Wang, C.; Shu, C.; Hong, H.; et al. Biodegradable, hydrogen peroxide, and glutathione dual responsive nanoparticles for potential programmable paclitaxel release. J. Am. Chem. Soc. 2018, 140, 7373–7376. [Google Scholar] [CrossRef] [PubMed]
  64. Chen, Y.; Li, Y.; Yang, X.; Cao, Z.; Nie, H.; Bian, Y.; Yang, G. Glucose-triggered in situ forming keratin hydrogel for the treatment of diabetic wounds. Acta Biomater. 2021, 125, 208–218. [Google Scholar] [CrossRef] [PubMed]
  65. Ma, J.; Guan, J.; Li, H.; Niu, H.; Guan, Y.; Tan, T. Cosmeceutical Composition. WO Patent 2021150383A1, 29 July 2021. [Google Scholar]
  66. Duvall, G.L.; Gupta, M.K.; Martin, J.R.; Dollinger, B.R. ROS-Degradeable Hydrogels. U.S. Patent 201603003241A1, 20 October 2016. [Google Scholar]
  67. Zhao, A.; Liu, L.; Lu, B.; Yang, P. Hydrogel with Ros Response and Preparation Method and Application. CN Patent 113101264A, 8 April 2021. [Google Scholar]
  68. Gu, Z.; Zhang, Y. ROS-Responsive Microneedle Patch for Acne Vulgaris Treatment. U.S. Patent 20210145984A, 20 May 2021. [Google Scholar]
  69. Lin, Z.Q.; Gao, W.; Hu, H.X.; Ma, K.; He, B.; Dai, W.B.; Wang, X.Q.; Wang, J.C.; Zhang, X.; Zhang, Q. Novel thermo-sensitive hydrogel system with paclitaxel nanocrystals: High drug-loading, sustained drug release and extended local retention guaranteeing better efficacy and lower toxicity. J. Control. Release 2014, 174, 161–170. [Google Scholar] [CrossRef]
  70. Duvall, C.L.; Martin, J.R.; O’Grady, K.P.; Nelson, C.E. Reactive Oxygen Species (ROS)-Responsive Compositions and Method. U.S. Patent 10695288B2, 23 June 2016. [Google Scholar]
  71. Yamaguchi, Y.; Takenaga, M.; Kitagawa, A.; Ogawa, Y.; Mizushima, Y.; Igarashi, R. Insulin-loaded biodegradable plga microcapsules: Initial burst release controlled by hydrophilic additives. J. Control. Release 2002, 81, 235–249. [Google Scholar] [CrossRef]
  72. Allison, S.D. Analysis of initial burst in PLGA microparticles. Expert Opin. Drug Del. 2008, 5, 615–628. [Google Scholar] [CrossRef] [PubMed]
  73. Williams, G.T.; Sedgwick, A.C.; Sen, S.; Gwynne, L.; Gardiner, J.E.; Brewster, J.T., II; Hiscock, J.R.; James, T.D.; Jenkins, A.T.A.; Sessler, J.L. Boronate ester cross-linked pva hydrogels for the capture and H2O2-mediated release of active fluorophores. Chem. Commun. 2020, 56, 5516–5519. [Google Scholar] [CrossRef] [Green Version]
  74. Wang, W.; Liu, X.; Ding, L.; Jin, H.J.; Li, X. Rna hydrogel combined with MnO2 nanoparticles as a nano-vaccine to treat triple negative breast cancer. Front. Chem. 2021, 9, 797094. [Google Scholar] [CrossRef]
  75. Wang, Z.-L.; Shen, Y.-H.; Sun, X.; Li, Z.-H.; Wang, X.-Y.; Zhao, Z. High biocompatible auncs-silk fibroin hydrogel system for visual detection of H2O2. Microchem. J. 2020, 157, 105036. [Google Scholar] [CrossRef]
  76. Manickam, P.; Vashist, A.; Madhu, S.; Sadasivam, M.; Sakthivel, A.; Kaushik, A.; Nair, M. Gold nanocubes embedded biocompatible hybrid hydrogels for electrochemical detection of H2O2. Bioelectrochemistry 2020, 131, 107373. [Google Scholar] [CrossRef]
  77. Xiaojin, G.; Li, Y.; Yingchun, L.; Ye, Z.; Zhang, J.; Zhu, T.; Li, F. An l012@pani-paam hydrogel composite based-electrochemiluminescence biosensor for in situ detection of H2O2 released from cardiomyocytes. Electrochim. Acta 2020, 354, 136763. [Google Scholar]
  78. Zhou, S.Y.; Wang, X.F.; Jiang, L.Y.; Sun, H.M.; Huo, D.Q.; Hou, C.J. A three-dimensional hydrogel-modified indium tin oxide electrode with enhanced performance for in situ electrochemical detection of extracellular H2O2. Analyst 2021, 146, 5403–5412. [Google Scholar] [CrossRef] [PubMed]
  79. Ye, J.; Zhu, L.P.; Yan, M.X.; Xiao, T.; Fan, L.B.; Xue, Y.; Huang, J.S.; Yang, X.R. An intensive and glow-type chemiluminescence of luminol-embedded, guanosine-derived hydrogel. Talanta 2021, 230, 122351. [Google Scholar] [CrossRef] [PubMed]
  80. Dong, J.; Zhao, Z.; Liu, R.; Zhang, H.; Wu, Y.; Ba, X. Investigation of a halloysite-based fluorescence probe with a highly selective and sensitive "turn-on" response upon hydrogen peroxide. RSC Adv. 2017, 7, 55067–55073. [Google Scholar] [CrossRef] [Green Version]
  81. Liu, R.X.; Cai, Z.Y.; Zhang, Q.S.; Yuan, H.; Zhang, G.L.; Yang, D. Colorimetric two-dimensional photonic crystal biosensors for label-free detection of hydrogen peroxide. Sens. Actuators B Chem. 2022, 354, 131236. [Google Scholar] [CrossRef]
  82. Diabetes Control and Complications Trial Research Group; Nathan, D.M.; Genuth, S.; Lachin, J.; Cleary, P.; Crofford, O.; Davis, M.; Rand, L.; Siebert, C. The effect of intensive treatment of diabetes on the development and progression of long-term complications in insulin-dependent diabetes mellitus. N. Engl. J. Med. 1993, 329, 977–986. [Google Scholar] [CrossRef]
  83. Wild, S.; Roglic, G.; Green, A.; Sicree, R.; King, H. Global prevalence of diabetes - estimates for the year 2000 and projections for 2030. Diabetes Care 2004, 27, 1047–1053. [Google Scholar] [CrossRef] [Green Version]
  84. The ADVANCE Collaborative Group; Patel, A.; MacMahon, S.; Chalmers, J.; Neal, B.; Billot, L.; Woodward, M.; Marre, M.; Cooper, M.; Glasziou, P.; et al. Intensive blood glucose control and vascular outcomes in patients with type 2 diabetes. N. Engl. J. Med. 2008, 358, 2560–2572. [Google Scholar] [CrossRef] [Green Version]
  85. Zhu, H.G.; Srivastava, R.; Brown, J.Q.; McShane, M.J. Combined physical and chemical immobilization of glucose oxidase in alginate microspheres improves stability of encapsulation and activity. Bioconjug. Chem. 2005, 16, 1451–1458. [Google Scholar] [CrossRef] [Green Version]
  86. Uchiyama, T.; Kiritoshi, Y.; Watanabe, J.; Ishihara, K. Degradation of phospholipid polymer hydrogel by hydrogen peroxide aiming at insulin release device. Biomaterials 2003, 24, 5183–5190. [Google Scholar] [CrossRef]
  87. Jeong, D.I.; Kim, S.; Lee, S.Y.; Kim, H.-J.; Lee, J.; Lee, K.; Cho, H.-J. Iron sulfate-reinforced hydrogel reactors with glucose deprivation, serial reactive oxygen species generation, ferroptosis induction, and photothermal ablation for cancer therapy. Chem. Eng. J. 2022, 438, 135584. [Google Scholar] [CrossRef]
  88. Kim, Y.E.; Kim, J. Ros-scavenging therapeutic hydrogels for modulation of the inflammatory response. ACS Appl. Mater. Interfaces 2021, 14, 23002–23021. [Google Scholar] [CrossRef] [PubMed]
  89. Phuong Le, T.; Lee, Y.; Dieu Linh, T.; Thai Thanh Hoang, T.; Kang, J.I.; Park, K.M.; Park, K.D. In situ forming and reactive oxygen species-scavenging gelatin hydrogels for enhancing wound healing efficacy. Acta Biomater. 2020, 103, 142–152. [Google Scholar]
  90. Wu, T.; Liu, W. Functional hydrogels for the treatment of myocardial infarction. NPG Asia Mater. 2022, 14, 9. [Google Scholar] [CrossRef]
  91. He, Z.A.; Jie, H.A.; Yan, L.A.; Xl, A.; Hz, A.; Hw, A.; Yx, A.; Chao, W.B.; Jian, W.A.; Zhuang, L.B. ROS-scavenging hydrogel to promote healing of bacteria infected diabetic wounds. Biomaterials 2020, 258, 120286. [Google Scholar] [CrossRef]
  92. Ding, J.; Yao, Y.; Li, J.; Duan, Y.; Nakkala, J.R.; Feng, X.; Cao, W.; Wang, Y.; Hong, L.; Shen, L.; et al. A reactive oxygen species scavenging and o2 generating injectable hydrogel for myocardial infarction treatment in vivo. Small 2020, 16, 2005038. [Google Scholar] [CrossRef]
  93. Chen, Z.; Xu, L.; Liang, Y.; Zhao, M. Ph-sensitive water-soluble nanospheric imprinted hydrogels prepared as horseradish peroxidase mimetic enzymes. Adv. Mater. 2010, 22, 1488–1492. [Google Scholar] [CrossRef]
  94. Saghati, S.; Khoshfetrat, A.B.; Tayefi Nasrabadi, H.; Roshangar, L.; Rahbarghazi, R. Fabrication of alginate-based hydrogel cross-linked via horseradish peroxidase for articular cartilage engineering. BMC Res. Notes 2021, 14, 384. [Google Scholar] [CrossRef]
  95. Qian, F.; Han, Y.; Han, Z.; Zhang, D.; Liu, H. In situ implantable, post-trauma microenvironment-responsive, ROS depletion hydrogels for the treatment of traumatic brain injury. Biomaterials 2021, 270, 120675. [Google Scholar] [CrossRef]
  96. Sakai, S.; Ashida, T.; Ogino, S.; Taya, M. Horseradish peroxidase-mediated encapsulation of mammalian cells in hydrogel particles by dropping. J. Microencapsul. 2014, 31, 100–104. [Google Scholar] [CrossRef]
Scheme 1. Relation between the over expression of H2O2 and related diseases.
Scheme 1. Relation between the over expression of H2O2 and related diseases.
Gels 08 00361 sch001
Figure 1. The H2O2-responsive release mechanism of the 1,4-phenylenediboronic acid-crosslinked hydrogels reported by our group (Reprinted/adapted with permission from Ref. [50]. Copyright year: 2022, copyright owner’s name: Hailei Zhang).
Figure 1. The H2O2-responsive release mechanism of the 1,4-phenylenediboronic acid-crosslinked hydrogels reported by our group (Reprinted/adapted with permission from Ref. [50]. Copyright year: 2022, copyright owner’s name: Hailei Zhang).
Gels 08 00361 g001
Figure 2. The H2O2-responsive behaviour of 1,4-phenylenebisdiboronic acid-crosslinked hydrogels reported by our group (A): H2O2-Gel in a transparent vial; (B): H2O2-Gel immersing in H2O2 aqueous solution for 100 min; (C): H2O2-Gel immersing in aqueous solution for 100 min; (D): the drug release profiles of DLHCH-1 and DLHCH-2) (Reprinted/adapted with permission from Ref. [50]. Copyright year: 2022, copyright owner’s name: Hailei Zhang).
Figure 2. The H2O2-responsive behaviour of 1,4-phenylenebisdiboronic acid-crosslinked hydrogels reported by our group (A): H2O2-Gel in a transparent vial; (B): H2O2-Gel immersing in H2O2 aqueous solution for 100 min; (C): H2O2-Gel immersing in aqueous solution for 100 min; (D): the drug release profiles of DLHCH-1 and DLHCH-2) (Reprinted/adapted with permission from Ref. [50]. Copyright year: 2022, copyright owner’s name: Hailei Zhang).
Gels 08 00361 g002
Figure 3. H2O2-responsive behaviors of the fluorescein-crosslinked hydrogels reported by our group: (A) The transformation mechanism of fluorescein-crosslinked hydrogels from arylboronates to phenols to afford fluorescein with high fluorescence in the presence of H2O2; (B) the changes of the hydrogel from non-fluorescent to fluorescent; (C) the drug release profiles in different concentration of H2O2; (D) the fluorescence spectra of the release medium (H2O2 = 200 μM) after addition of DHNTs@PVA@PA; (E) plots of fluorescence intensity vs. the drug release rate (Reprinted/adapted with permission from Ref. [51]. Copyright year: 2022, copyright owner’s name: Hailei Zhang).
Figure 3. H2O2-responsive behaviors of the fluorescein-crosslinked hydrogels reported by our group: (A) The transformation mechanism of fluorescein-crosslinked hydrogels from arylboronates to phenols to afford fluorescein with high fluorescence in the presence of H2O2; (B) the changes of the hydrogel from non-fluorescent to fluorescent; (C) the drug release profiles in different concentration of H2O2; (D) the fluorescence spectra of the release medium (H2O2 = 200 μM) after addition of DHNTs@PVA@PA; (E) plots of fluorescence intensity vs. the drug release rate (Reprinted/adapted with permission from Ref. [51]. Copyright year: 2022, copyright owner’s name: Hailei Zhang).
Gels 08 00361 g003
Figure 4. Colorimetric H2O2-responsive mechanism of phenolphthalein-crosslinked hydrogel (PPH@PVA) reported by our group (Reprinted/adapted with permission from Ref. [52]. Copyright year: 2022, copyright owner’s name: Hailei Zhang).
Figure 4. Colorimetric H2O2-responsive mechanism of phenolphthalein-crosslinked hydrogel (PPH@PVA) reported by our group (Reprinted/adapted with permission from Ref. [52]. Copyright year: 2022, copyright owner’s name: Hailei Zhang).
Gels 08 00361 g004
Table 1. Summary of responsive mechanism to H2O2.
Table 1. Summary of responsive mechanism to H2O2.
No.MechanismLiterature
1 Gels 08 00361 i001[31,32]
2 Gels 08 00361 i002[33,34]
3 Gels 08 00361 i003[35]
4 Gels 08 00361 i004[36]
5 Gels 08 00361 i005[37]
6 Gels 08 00361 i006[38]
7 Gels 08 00361 i007[39,40]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Song, W.; You, J.; Zhang, Y.; Yang, Q.; Jiao, J.; Zhang, H. Recent Studies on Hydrogels Based on H2O2-Responsive Moieties: Mechanism, Preparation and Application. Gels 2022, 8, 361. https://0-doi-org.brum.beds.ac.uk/10.3390/gels8060361

AMA Style

Song W, You J, Zhang Y, Yang Q, Jiao J, Zhang H. Recent Studies on Hydrogels Based on H2O2-Responsive Moieties: Mechanism, Preparation and Application. Gels. 2022; 8(6):361. https://0-doi-org.brum.beds.ac.uk/10.3390/gels8060361

Chicago/Turabian Style

Song, Weihua, Jipeng You, Yuangong Zhang, Qi Yang, Jin Jiao, and Hailei Zhang. 2022. "Recent Studies on Hydrogels Based on H2O2-Responsive Moieties: Mechanism, Preparation and Application" Gels 8, no. 6: 361. https://0-doi-org.brum.beds.ac.uk/10.3390/gels8060361

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

Article Metrics

Back to TopTop