Next Article in Journal
Effect of Sol–Gel Silica Matrices on the Chemical Properties of Adsorbed/Entrapped Compounds
Next Article in Special Issue
Dual-Temperature/pH-Sensitive Hydrogels with Excellent Strength and Toughness Crosslinked Using Three Crosslinking Methods
Previous Article in Journal
Streamlining Skin Regeneration: A Ready-To-Use Silk Bilayer Wound Dressing
Previous Article in Special Issue
Development of a Self-Healing Gel with Self-Healing Kinetics That Can Be Controlled by Heat
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Stimulus-Responsive Hydrogels for Targeted Cancer Therapy

Department of Biological Sciences and Engineering, Indian Institute of Technology Gandhinagar, Palaj 382355, Gujarat, India
*
Authors to whom correspondence should be addressed.
Gels 2024, 10(7), 440; https://doi.org/10.3390/gels10070440
Submission received: 10 June 2024 / Revised: 27 June 2024 / Accepted: 29 June 2024 / Published: 1 July 2024
(This article belongs to the Special Issue Stimuli-Responsive Composite Gels)

Abstract

:
Cancer is a highly heterogeneous disease and remains a global health challenge affecting millions of human lives worldwide. Despite advancements in conventional treatments like surgery, chemotherapy, and immunotherapy, the rise of multidrug resistance, tumor recurrence, and their severe side effects and the complex nature of the tumor microenvironment (TME) necessitates innovative therapeutic approaches. Recently, stimulus-responsive nanomedicines designed to target TME characteristics (e.g., pH alterations, redox conditions, enzyme secretion) have gained attention for their potential to enhance anticancer efficacy while minimizing the adverse effects of chemotherapeutics/bioactive compounds. Among the various nanocarriers, hydrogels are intriguing due to their high-water content, adjustable mechanical characteristics, and responsiveness to external and internal stimuli, making them promising candidates for cancer therapy. These properties make hydrogels an ideal nanocarrier for controlled drug release within the TME. This review comprehensively surveys the latest advancements in the area of stimulus-responsive hydrogels for cancer therapy, exploring various stimuli-responsive mechanisms, including biological (e.g., pH, redox), chemical (e.g., enzymes, glucose), and physical (e.g., temperature, light), as well as dual- or multi-stimuli responsiveness. Furthermore, this review addresses the current developments and challenges in hydrogels in cancer treatment. Our aim is to provide readers with a comprehensive understanding of stimulus-responsive hydrogels for cancer treatment, offering novel perspectives on their development for cancer therapy and other medical applications.

1. Introduction

Cancer is one of the most detrimental diseases, affecting millions of human lives globally and imposing a significant economic burden. As per the reports from the International Agency for Research on Cancer (IARC), there were about 20 million new cancer cases diagnosed, as well as 9.7 million cancer-related deaths in 2022 [1]. In 2040, about 28.4 million new cancer cases are expected [1,2]. Cancer is seriously affecting patients and their families and leads to high economic expenses including medical expenses and the loss of human capital due to early deaths [2]. Conventional cancer therapies, such as chemotherapy and radiotherapy, often suffer from limited efficacy, multidrug resistance, tumor recurrence, and severe side effects such as cardiotoxicity and nephrotoxicity due to their non-specific targeting of healthy tissues [3]. Recently, the emergence of cancer biomarkers in oncology has transformed cancer treatment, leading to significant advances in cancer therapies and improved patient prognoses [4]. Modern advances in interdisciplinary fields; the evolution of molecular imaging tools such as magnetic resonance imaging (MRI), computed tomography (CT), positron emission tomography (PET), single photon emission computed tomography (SPECT), and photoacoustic tomography (PAT) for cancer diagnostics [5,6]; developments in artificial intelligence (AI) in oncology [7]; significant advancements in single-cell multi-omics technology such as FIPRESCI [8]; and cancer vaccines [9] have revolutionized cancer therapy. However, these advances in cancer therapy are still insufficient to combat cancer completely. Hence, there is a pressing need to explore innovative approaches that can enable targeted drug delivery to cancer cells while minimizing systemic toxicity.
Nanotechnology represents a growing and promising field with diverse applications in drug delivery, bioimaging, and therapeutics [10]. Nano-drug delivery systems offer beneficial properties such as enhanced bioavailability, sustained drug release, stability, and improved aqueous solubility, thereby enhancing the anticancer activity of drugs [11]. Various nanocarriers, including hydrogels; polymeric, metallic, and lipid-based nanocarriers; micelles; dendrimers; and so on, have been utilized for the delivery of chemotherapeutic drugs/bioactive compounds [12].
Stimulus-responsive nanomedicine offers significant potential advantages in the realm of targeted therapy and personalized medicine [13]. By leveraging nanomaterials that respond to specific physiological stimuli—such as pH changes, temperature variations, or enzyme presence—these advanced therapeutic systems can deliver drugs with high precision directly to diseased tissues; minimizing off-target effects and reducing systemic toxicity. This targeted approach not only enhances the efficacy of treatments, particularly in cancer and inflammatory diseases, but also allows for controlled drug release, improving patient compliance and outcomes. Additionally, the adaptability of these nanomedicines can lead to real-time monitoring and responsive treatment adjustments, paving the way for more dynamic and effective healthcare solutions.
Hydrogels are three-dimensional (3D) structures of cross-linked polymers and holds great potential in drug delivery [14]. Their water-holding characteristics and structural similarity to biological tissues make them suitable carriers for drug delivery and tissue bioengineering applications. Hydrogels have modifiable durability, biocompatibility, and biodegradability as well as versatility in incorporating a wide range of therapeutic agents. Their adaptability for modification makes hydrogels a promising tool in improving the efficacy and safety of cancer treatments. Other advantages such as swelling behavior, porosity, mechanical properties, and responsiveness to various stimuli also make them ideal platforms for developing smart drug delivery systems. The TME has specific characteristics such as hypoxia, acidity, increased enzyme levels, and aberrant temperatures, all of which can be used to selectively induce drug release from hydrogels. In response to specific stimuli present in the TME, hydrogel-based drug delivery systems could be a promising approach to combat cancer.
Targeted drug delivery through hydrogels is achieved via various stimuli such as pH, temperature, light, and so on [15]. Hydrogels can be synthesized using different biomaterials, including carbohydrates, proteins, lipids, or DNA/RNA-based nanomaterials, depending on the targeted cancer sites [16,17]. Hydrogels can be administered through various methods including injection, topical application, or implantation, depending on the therapeutic need. Hydrogels are widely used in biomedical applications such as drug delivery systems, wound dressings, and tissue engineering due to their biocompatibility and ability to mimic the TME (Figure 1).
We present here an in-depth overview of the current status of the research on stimulus-responsive hydrogels as a cancer therapy. We will explore the various types of biomaterials used in hydrogel fabrication, their responsiveness to different stimuli, and their applications in cancer targeting (Figure 1). Biological, chemical, and physical stimuli are thoroughly discussed with recent studies on hydrogels. By defining the fundamental concepts and focusing on current advances in the field, we are interested in shedding light on the potential of stimulus-responsive hydrogels to transform cancer treatment strategies.

2. Tumor Microenvironment (TME)

The scientific community has been unable to eliminate cancer despite many years of effort. One reason behind this is the complex nature of the TME. A tumor is a heterogeneous collection of invading cells rather than just a collection of cancer cells. This comprises the extracellular matrix (ECM), substances that are secreted, and resident host cells. In order to facilitate the growth and spread of tumors, cancer cells fundamentally alter the host tissues’ cellular, molecular, and physical makeup [18]. A developing TME is a dynamic and evolving entity. While the TME’s makeup varies depending on the kind of tumor, the immune cells, stromal cells, blood vessels, and extracellular matrix are its common components. The theory states that the “tumor microenvironment actively promotes the progression of cancer rather than merely acting as a silent bystander” [19]. The development of the TME is primarily linked to the uncontrolled growth of cancer cells and the expansion of defective blood vessels.
Recent research has revealed that key cellular metabolism pathways, which are controlled by internal genetic changes and TME-induced cell-extrinsic responses, are different in cancer cells from those in most normal tissue cells. These changes involve, but are not restricted to, altered signal transduction (e.g., hypoxia-inducible factor 1 (HIF-1), aerobic glycolysis, reduced oxidative phosphorylation) and biomolecule metabolism (e.g., glucose, lactate, glutathione). These factors lead to increased tumor proliferation, invasiveness, metastasis, and strong treatment resistance, enabling tumor cells to elude tumor-targeted therapies. They also promote vascular regeneration, nutrient uptake, adenosine triphosphate (ATP) generation, increased macromolecule biosynthesis, and elevated redox levels [20].
Acidic pHs, hypoxia, increased levels of GSH, higher reactive oxygen species (ROS) generation, and overexpressed enzymes are key characteristics of the TME, which promote tumor angiogenesis and metastasis while also being responsible for therapeutic resistance and treatment failure (Figure 2) [21]. The design and development of TME-responsive intelligent drug delivery systems has gained interest in improving drug therapeutic effects owing to their potential to address some significant therapeutic issues, such as low therapeutic efficacy and serious side effects like cardiotoxicity, nephrotoxicity, and so on.

3. Hydrogels: An overview

In 1960s, Wichterle and Lim invented the word “hydrogel” to describe a 3D structure composed of hydrophilic polymers in biological contexts [22]. They named it poly (2-hydroxyethyl methacrylate) (pHEMA), and it was utilized in the contact lens industry [23]. pHEMA demonstrated how a hydrogel could absorb moisture and maintain its network structure. Hydrogels have emerged as fascinating drug delivery materials because of their distinct characteristics. One of their primary advantages is their high-water content, which replicates the natural environment of biological tissues, enhancing compatibility and decreasing discomfort during administration [24]. Furthermore, hydrogels have variable porosity and swelling behaviors, which allows for controlled drug release kinetics and maintains therapeutic concentrations over time [25]. Their soft and flexible structure allows for high entrapment of anticancer drugs including hydrophilic and hydrophobic drugs, while maintaining stability and bioactivity [26]. Furthermore, hydrogels have the benefit of being easily adaptable to achieve sustained release profiles for various pharmaceutical and medicinal applications. These promising characteristics make hydrogels an appealing candidate for precise and effective drug delivery systems. Hydrogels can be classified as homo or copolymeric based on their chemical composition; macro gels, microgels, or nanogels based on their network size; anionic, cationic, zwitterionic, or non-ionic based on their ion charge; and physical or chemical based on their cross-linking (Figure 3) [27,28].
Hydrogels can be synthesized using polymers through covalent or non-covalent bonding and classified into physically cross-linked, chemically cross-lined or hybrid cross-linking (Figure 3) [27]. Physical hydrogels are formed through interactions such as hydrogen bonding, Van der Waals interactions, ionic forces, polyelectrolyte complexation and hydrophobic forces which cause cross-linking among the polymer chains [29]. Due to the weak secondary connections between the polymer chains, physical hydrogels respond reversibly to environmental changes and are mechanically weak, disorganized, and fragile when subjected to external stimuli. In contrast, chemically cross-linked hydrogels are formed by covalent bonding between the polymer chains. Physically cross-lined hydrogels typically disintegrate in organic solvents and water (when heated) due to weak interactions, whereas chemical hydrogels do not dissolve in the surrounding medium and do not exhibit a reversible sol-gel transition. Chemical cross-linking agents such as formaldehyde, glutaraldehyde (GA), genipin, diglycidyl ether, and N, N′-methylenebisacrylamide stabilize the network by forming covalent bonds between small molecules and polymers through condensation or free radical mechanisms [30]. This process imparts excellent thermal, mechanical, chemical, and surface properties to the hydrogels, maintaining their network structure even when fully swollen. Additionally, advanced synthesis methods such as electrospinning and 3D printing (Figure 3) have been employed to create hydrogels with tailored structures and properties, enabling precise control over their architecture and functionality for various biomedical applications [31,32].

4. Stimulus-Responsive Hydrogels for Cancer Therapy

Stimulus-responsive hydrogels for cancer therapy represent a promising approach in drug delivery and treatment strategies. These hydrogels can be tailored to release therapeutic agents in response to specific stimuli such as pH, temperature, light, or enzymatic activity found in tumors, allowing for targeted and controlled drug delivery. Their ability to integrate with biological systems and respond dynamically to the TME makes them valuable tools for enhancing the efficacy and reducing the side effects of cancer treatments [33]. Stimulus-responsive hydrogels, a novel class of smart biomaterials, are engineered for drug delivery applications that utilize physical stimuli (such as temperature and light), chemical stimuli (such as pH and redox reactions), biochemical signals (such as glucose and enzymes), and can respond to dual or multiple stimuli.
These hydrogels exhibit responsiveness to stimuli, enabling controlled modulation of therapeutic agent release. This capability facilitates precise and controlled drug delivery strategies, thereby enhancing treatment efficacy and minimizing adverse effects in medical applications. Several hydrogel-based formulations have been developed over time for cancer treatment, prevention, and diagnosis with some undergoing clinical trials. More than 30 hydrogels have received FDA approval and are now used in clinical applications. Recently, UGN-101® (Jelmyto™, formerly MitoGel, UroGen Pharma, Israel) [34] and Vantas® (Endo Pharmaceuticals) [35] have been approved by the FDA for the treatment of cancers. UGN-101 is a temperature-sensitive, water-soluble, and novel polymer-based formulation of mitomycin C for the treatment of upper urothelial carcinomas [36]. The hydrogel composition is liquid at room temperature and gels at body temperature, conforming to each patient’s unique pelvicalyceal and ureteric anatomy. Vantas is a hydrogel depot made of polymers (hydroxyethyl methacrylate and hydroxypropyl) cross-linked with trimethylopropane trimethacrylate [37,38]. It releases histrelin acetate into the upper arm over 12 months following subcutaneous implementation, providing palliative treatment for advanced prostate cancer. Another FDA-approved gel-based product, ELIGARD® (Atrix Laboratories, currently known as Tolmar) is an in situ polymeric gel used for the delivery of leuprolide acetate and has been approved for palliative therapy in advanced prostate cancer [38,39]. Utilizing a novel in situ polymeric gel extended delivery system, this formulation forms a solid implant upon injection. This allows for the controlled and sustained release of leuprolide acetate as the biodegradable gel. These are just some examples of hydrogels that are currently used in clinical practice (represented in Figure 4), but more research is ongoing, with many being tested in preclinical and clinical trials. In this review, we will discuss various stimulus-responsive hydrogels, providing recent examples from research studies.
In addition to significantly increasing the therapeutic efficacy of the loaded medications, responsive hydrogels can recognize tumor tissue and release drugs on demand based on specific conditions within the tumor microenvironment, thereby reducing harm to healthy tissues. These smart materials respond to multiple stimuli, such as changes in pH, temperature, and the presence of certain molecules. For instance, redox–pH hydrogels respond to altered glutathione in the cytoplasm of tumor cells, with the acidic TME triggering drug release, while thermo-pH-sensitive hydrogels respond to the elevated temperature and acidic environment of tumor cells. This targeted approach ensures effective drug delivery and controlled in situ release, enhancing treatment outcomes and minimizing side effects on normal tissues. The detailed mechanism of hydrogels in response to different stimuli are discussed in depth in their respective sections.

4.1. Chemical Stimuli-Responsive Hydrogels

Chemical response-mediated hydrogels dynamically change their structure or properties in response to specific chemical signals, offering tailored solutions for drug delivery and tissue engineering [40]. pH-responsive hydrogels undergo swelling or contraction in response to changes in environmental acidity or alkalinity, making them ideal for targeted drug release in acidic tumor microenvironments [41]. Similar to this, redox-responsive hydrogels can be made to change reversibly in response to changes in the redox state of their environment. This allows for controlled release in environments that are either reducing or oxidative [42]. We discuss pH- and redox-responsive hydrogels below, along with several research studies that have utilized these stimuli for drug delivery.

4.1.1. pH-Responsive Hydrogels

Hydrogels that respond to pH fluctuations are an exceptional category of biomaterials designed to exhibit reversible changes [43]. These hydrogels contain pH-sensitive functional groups, such as acidic or basic residues, within their polymer networks, enabling them to swell or contract in acidic or alkaline environments. This pH-triggered swelling behavior can be harnessed for targeted drug delivery, where the hydrogel acts as a carrier for therapeutic agents, releasing them selectively in response to specific pH conditions characteristic of diseased tissues or cellular compartments [44]. Their tunable responsiveness to pH gradients renders them invaluable tools in biomedicine, offering tailored solutions for controlled release, tissue engineering, and diagnostic applications, with the potential to revolutionize drug delivery strategies and improve healthcare outcomes [45]. pH-sensitive hydrogels can deliver oral drugs to sites in the gastrointestinal (GI) tract due to pH variations [43].
Hydrogels act as cargo for therapeutic drugs, gradually releasing them through interactions with charged ions that modulate the release rate [46]. Polymers like PEG, PAA, pluronics, and chitosan are preferred for targeted drug delivery due to their pH-responsive nature, facilitated by protonable cationic (NH2) or deprotonable anionic (-COOH) groups in the hydrogel matrix. Adjusting pH levels effectively controls the swelling capacity of pH-responsive hydrogels. The ionization of hydrogels is altered with pH changes, influencing the swelling behavior of cross-linked polyelectrolytes [47]. Variations in pH, ionic strength, and electrostatic repulsion significantly affect these characteristics, making pH-sensitive hydrogels suitable for targeted oral drug-release formulations, leveraging pH differences in body regions like the colon (pH > 7) and stomach (pH < 3) [48]. Cationic hydrogels, such as those composed of polycationic polymers, exhibit maximal drug release at low pH levels, highlighting their utility in stomach-targeted drug delivery systems [49].
Ortiz, J.A. et al. synthesized novel pH-responsive hydrogels by combining carboxymethylagarose (CMA) and chitosan (CS) at different weight ratios [50]. Diclofenac sodium (DS) was used as a model drug and was successfully incorporated into CMA/CS PECs. The viability of HaCaT cells was nearly 100% in the presence of the hydrogels and DS. The study demonstrated that the prepared hydrogels could be used as a pH-responsive nano system for transdermal drug delivery. Another research team synthesized pH-responsive injectable and covalently cross-linked hydrogels by mixing a dibenzaldehyde-terminated PEG (DF-PEG) solution and polyaspartylhydrazide (PAHy) solution (Figure 5A), containing doxorubicin (DOX) into the prepared hydrogel [51].
The results indicated that the sol–gel transitions of the produced hydrogel are reversible in response to pH fluctuations (Figure 5B). In an in vivo mouse model, the release rate of the DOX encapsulated within the hydrogel and its accumulation in the tumor were significantly slower compared to free DOX. The drug-loaded hydrogel exhibited enhanced efficacy, achieving approximately 80% tumor inhibition by day 20 (Figure 5C), suggesting its potential as a highly effective treatment for human fibrosarcoma with reduced side effects.
Various polymers, such as chitosan [52,53,54,55,56], poly(ethylene oxide) [57,58,59], polyethylene glycol [60,61,62], methacrylic acid [63,64,65], gelatin [66,67,68], laponite RD [69], polyethyleneimine [70], β-cyclodextrin [71,72,73], polyacrylamide, polyurethanes [74], carboxymethylagarose [50], and dibenzaldehyde-terminated PEG [51], have been used in the synthesis of pH-responsive hydrogels. Table 1 summarizes the diverse polymers that have been conjugated or encapsulated with drugs/ligands to synthesize chemical-responsive hydrogels, along with a summary of the findings.
Parisa Alipournazari et al. synthesized a pH-sensitive starch/PVA/g-C3N4 hydrogel for the delivery of DOX to breast cancer cells [80]. DOX was released faster in the tumor acidic environment compared to physiological pH conditions. Additionally, biological studies suggested that the prepared hydrogels enhanced anticancer activity and apoptosis and were non-toxic compared to free DOX. In another study, Bing Ma et al. prepared pH-responsive benzimidazole–chitosan quaternary ammonium salt (BIMIXHAC) nanogels for the delivery of DOX [81]. BIMIXHAC was cross-linked with chitosan, HA, and sodium alginates via the ion cross-linking method. Cytotoxicity studies showed that these nanogels were biocompatible and exhibited significant anticancer effects against MCF-7 and A549 cells. Zijian Zhang et al. synthesized pH- and thermo-sensitive composite hydrogels using oxidized cellulose nanofiber, PVA, and PDA via physical cross-linking [82]. In vitro studies showed that the prepared composite hydrogel exhibited a high NIR photothermal conversion efficiency, low cytotoxicity to L929 cells, pH-dependent release, and potent inhibitory effects on MCF-7 cells. Olmesartan is a hydrophobic drug that has solubility issues. To enhance its solubility and sustained release, Tahir Mahmood et al. synthesized polymeric pH-responsive nanogels developed from PEG-g-poly(MAA) using a free radical polymerization technique [63]. The resultant nanogel was able to increase the solubility by 12.3- and 13.29-fold at pH 1.2 and pH 6.8, respectively.

4.1.2. Redox-Responsive Hydrogels

Redox is advantageous over other stimuli because it can change both the charge and spin states of a molecule and its assemblies at the same time. In the 18th century, Lavoisier coined the terms “oxidation” and “reduction” to describe oxygen uptake and loss, respectively [83]. Redox-active compounds offer promise in biomedical applications including drug delivery. Due to their response to redox stimuli, disulfide-cross-linked hydrogels are frequently used as a drug delivery carrier [78]. Thiols can be present in the cysteine residues of synthetic polymers as well as protein-based compounds. Therefore, thiol-based cross-linking can be used to quickly make hydrogels [84]. For the synthesis of thiol–ene and thiol–Michael cross-linked hydrogels with good biocompatibility, maleimide and divinyl sulfone (DVS) are efficient chemicals [85]. Through oxidation, thiols can also form disulfide linkages [86] and oxidants like periodate [87], hydrogen peroxide [88], and ferricyanide [89] can be used to enhance the process.
Since disulfide bonds are converted to sulfhydryl (-SH) groups in tumor tissue by high levels of reduced glutathione, most redox-responsive hydrogels are synthesized from disulfide bond-bearing compounds [42]. In normal tissue with low reducing activity, the disulfide bond is reasonably stable; nevertheless, in a tumor environment, its reduction causes structural disruptions of the drug delivery system, leading to the abrupt release of the drug [90]. Direct and indirect introduction are the two accepted methods for creating disulfide links in polymers [91]. The disulfide bond is created by oxidizing sulfhydryl groups in indirect introduction; a disulfide bond-containing moiety is inserted into the nanocarrier structure in direct introduction. The indirect method is the simpler and more popular approach among these strategies.
Kilic Boz, R. et al. used a highly effective thiol–disulfide exchange technique to develop redox-responsive hydrogels [33]. Gelation was achieved by combining linear telechelic PEG-based polymers with pyridyl disulfide units at the chain ends with thiol-terminated tetra-arm PEG polymers. Fast gelation provides macroporous hydrogels with significant water absorption through excellent conversions (>85%). Furthermore, the resultant hydrogels have the ability to self-heal because of the disulfide linkages. The hydrogels totally break down in an environment high in thiol-containing substances such as L-glutathione (GSH) and dithiothreitol (DTT). Furthermore, the molecular weight of the polymeric precursors can be changed to adjust the release profile of the encapsulated protein, in this case, bovine serum albumin. A live/dead cell viability assay confirmed the materials’ cytocompatibility. Their study suggested that redox responsive hydrogels may be appealing for a variety of biological applications due to their ease of manufacturing, and capacity to disintegrate on demand and release their payload like BSA. Other examples include DTSSP-cross-linked RZ10-RGD [79], ferric ethylenediaminetetraacetic acid (Fe-EDTA) [78], PEG-b-PBSe block copolymers [92], poly (2-methacryloyloxyethyl phosphorylcholine) (PMPC) [93], and carboxymethyl chitosan [94] polymer-based nano systems synthesized for redox-responsive drug delivery (Table 1).
Di Huang et al. prepared pH- and redox-sensitive microgels using an emulsion polymerization method [95]. These microgels were synthesized from bis(2-methacryloyloxyethyl) disulfide (DSDMA), which serves as a redox-responsive cross-linker, PEG, as the hydrophilic shell, and 2-(diisopropylamino) ethyl methacrylate (DPA) as the monomers. In vitro and in vivo studies showed the inhibition of proliferation and suppression of tumor growth. In another study, Weiyong Tao et al. synthesized a selenite-containing hydrogel composed of calcium selenite/L-arginine (L-Arg) nanospheres (SL NPs), sodium alginate, glucose oxidase, and 6-aminonicotinamide [96]. This injectable hydrogel could form a gel in a physiological environment. The prepared hydrogels efficiently induced the immunogenic cell death of tumor cells via enhanced ROS/reactive nitrogen species (RNS)-mediated combination therapy by disrupting NAPDH homeostasis and enhancing antitumor immunity. Furthermore, a TME-selective PAA-MnO2 hydrogel was synthesized by Akhmad Irhas Robby et al. [97]. The in vivo experiments on the cancer cell-triggered sol–gel transformation of the mineralized hydrogel showed that the PAA-MnO2 hydrogel can be formed in tumor-bearing mice due to its excellent ROS-scavenging activity at the tumor site, as confirmed by a gene expression analysis. Based on these reported studies, redox-responsive hydrogels could be promising in the treatment of cancer.

4.2. Biological Stimuli-Responsive Hydrogels

Biological stimuli-responsive hydrogels exhibit dynamic changes in their structure or properties in response to specific biological cues, enabling tailored drug delivery and tissue engineering applications [40,98]. Enzyme-responsive hydrogels, such as those triggered by proteases, enable targeted drug release at sites of inflammation or disease where elevated enzyme levels are present [99]. Glucose-responsive hydrogels, another prominent example, undergo swelling or degradation in response to variations in glucose concentration, offering precise insulin delivery for diabetes management [100]. These biological stimuli-responsive hydrogels intelligently respond to biological signals, enhancing therapeutic efficacy, and hold great promise for cancer therapy and the treatment of other medical conditions.

4.2.1. Enzyme-Responsive Hydrogels

Among the several stimulus-responsive nanocarriers, the enzyme-responsive approach has gained increasing interest in the development of functional biomaterials, particularly for designing drug delivery systems, for the following reasons: (1) enzymes play important functions in most biochemical processes, and the enzyme-based method is extremely biocompatible, and (2) therapeutics can be delivered to enzyme-overexpressed tumor sites with great selectivity and efficiency using enzyme-based recognition strategies [101]. Enzyme-responsive hydrogels, leveraging specific enzyme-cleavable motifs within their polymer network, offer a dynamic platform for targeted drug delivery, tissue engineering, biosensing, and regenerative medicine [102]. These hydrogels enable spatiotemporal control over drug release, mimicking the dynamic nature of biological tissues, and providing a biomimetic matrix for promoting cell proliferation and tissue regeneration [103]. Sakai S. and Kawakami K. synthesized a novel alginate-based hydrogel incorporating phenol groups via carbodiimide coupling, aiming to develop an enzymatically cross-linkable hydrogel with gelation properties [104]. By controlling phenol incorporation, they achieved aqueous polymer solutions capable of gelation through both ionic and enzymatic cross-linking, overcoming destabilization challenges in biomedical applications. Their findings suggest the potential of phenol-modified alginate as an alternative to conventional forms for various biomedical uses.
MMPs are enzymes crucial for regulating the ECM in various physiological processes, including tissue remodeling, wound healing, and immune response modulation [105]. In cancer, MMPs play a significant role in tumor invasion and metastasis [106]. Exploiting this enzymatic activity, MMP-responsive nanocarriers, including nanoparticles and hydrogels, have emerged as promising platforms for targeted drug delivery in cancer therapy [107,108]. MMP-responsive hydrogels enable controlled drug release within tumor microenvironments, minimizing systemic toxicity. They hold promise for precision cancer therapy, enhancing efficacy while reducing adverse effects. Various MMP-based targeted hydrogels have been synthesized and explored as nanocarriers for cancer treatment [109,110,111].
Su T. et al. designed an enzyme-responsive hydrogel composed of glucose oxidase (GOx), an N-hydroxyimide–heparin conjugate, and β-D-glucose [112]. The developed enzyme-responsive hydrogel enables controlled drug release triggered by heparin-specific cleavage by heparanase, targeting cancer cells with heparanase overexpression while minimizing the negative effects of premature drug release on normal cells. Another group of researchers synthesized a matrix metalloproteinase (MMP)-responsive hydrogel to target the overexpression of O6-methylguanine-DNA methyltransferase (MGMT) by glioma cells [109]. They loaded O6-benzylguanine (an inhibitor for MGMT) and temozolomide (TMZ, a first-line agent for treatment) for a therapy targeting TMZ-resistant gliomas. The drug-loaded hydrogels reduced MGMT expression in vivo, rendering TMZ-resistant glioma cells more responsive to TMZ treatment. Additionally, post-surgery, these hydrogels significantly enhanced TMZ efficacy in glioma growth inhibition and reduced the recurrence of TMZ-resistant gliomas, suggesting their potential as localized medication delivery for preventing glioma recurrence. The enzymes such as proteases, hydrolases, oxidoreductases such as glucose oxidase, kinases, phosphatases, and glycosidases that are commonly utilized in enzyme-responsive hydrogels are listed in Table 2.
Other than MMPs, Hovgaard L. and Brøndsted H. synthesized dextranase-targeted enzyme-responsive hydrogels based on dextran [114]. Increasing the molecular weight of dextran, adding cross-linking agents, or lowering the amount of DMSO in the reaction mixture resulted in hydrogels with decreased swelling and higher mechanical strength. Dextran hydrogels were degraded in vitro by dextranase, as well as in vivo in rats and a human colonic fermentation model. A detailed list of enzyme-responsive hydrogels is given in Table 2. All the evidence underscores the potential of enzyme-responsive hydrogels for effective cancer treatment, offering targeted drug delivery and minimizing systemic side effects.

4.2.2. Glucose-Responsive Hydrogels

Glucose-responsive systems dynamically respond to fluctuations in the surrounding glucose concentration, offering significant promise for diabetes and cancer treatments [117]. Research on glucose-responsive systems can be categorized along two dimensions: the mechanism of glucose responsiveness and its application [118]. Mechanistically, these systems fall into three main groups: lectin-based, glucose oxidase (GOx)-based, and phenylboronic acid (PBA)-modified systems. Application-wise, they are typically classified into two groups: glucose concentration diagnostics and insulin release facilitation. In major studies, glucose-responsive components such as glucose-binding proteins [119], phenylboronic acid (PBA) [115,120,121] and enzymes such as glucose oxidase/catalase (GOx/CAT)) [122,123] are used for the delivery of insulin (Table 2).
Li X. et al. designed novel, biocompatible, glucose-responsive hydrogels containing GOx, catalase, and insulin in peptide IA-0 for insulin delivery via the self-assembly approach in physiological conditions [116]. The in vitro and in vivo studies demonstrated that the developed hydrogels regulated the blood glucose levels, even in mouse models with STZ-induced diabetes. Glucose-responsive hydrogels find more extensive application in diabetic disease management compared to cancer therapy, primarily due to the higher prevalence of diabetes and the urgent need for precise insulin delivery in diabetic patients. These studies demonstrated that biological stimuli-responsive hydrogels, including enzyme- and glucose-responsive ones, are increasingly popular in the treatment of various diseases, including cancer, due to their ability to deliver drugs precisely and effectively in response to specific biological triggers.

4.3. Physical Stimuli-Responsive Hydrogels

To ensure the effective transport of nanocarriers within cells, it is imperative to thoroughly evaluate the appropriate designs of the drug carriers that use either active or passive targeting strategies. Common mechanisms involve receptor-mediated endocytosis, passive diffusion, or particle phagocytosis [124]. For example, the enhanced permeability and retention (EPR) effect has frequently underpinned passive targeting approaches in cancer therapy [125]. Conversely, active targeting can be achieved by incorporating specialized ligands, such as antibodies, oligonucleotides, and peptides, that bind to specific cell surface biomarkers or overexpressed receptors. Ensuring the efficient release of drugs that are conjugated, complexed, or encapsulated into the nanocarriers is a critical consideration [126]. The stability of the drug delivery system, the drug’s bioavailability, and the requisite mechanism for controlled release can all be influenced by the chemical nature of the bond and linker [127]. Recently, controlled drug release has also been achieved through the application of physical stimuli such as light, temperature, pressure, magnetic fields, and ultrasound to overcome hurdles observed in chemical stimuli-responsive hydrogels [128]. Here, we discuss thermo- and light-responsive hydrogels used for drug delivery (Table 3).

4.3.1. Thermo-Responsive Hydrogels

This form of drug delivery has significantly enhanced patient convenience and maximized the efficiency of the drugs used. Nevertheless, the hypodermic route of administration has not fully addressed certain challenges, notably cardiotoxicity, that are associated with the intravenous delivery of chemotherapeutic agents such as Herceptin [142]. An alternative therapeutic approach, aiding localized continuous medication administration, has emerged as a promising strategy for enhancing therapeutic efficacy while mitigating systemic adverse effects. Thermo-responsive hydrogels undergo a reversible sol–gel transition in response to temperature changes, making them valuable for controlled drug delivery and tissue engineering [143]. For instance, injectable and thermosensitive hydrogels have been widely used for targeted drug delivery for the treatment of various diseases [144]. Below their lower critical solution temperature (LCST), they exist as a sol, becoming a gel above this threshold. This transition allows the encapsulation and targeted release of drugs, minimizing systemic side effects and improving therapeutic efficacy [145]. Common polymers used include poloxamers, poly(N-isopropylacrylamide) (PNIPAAm), CS/β-glycerophosphate, hyaluronic acid (HA), poly (ethylene glycol) (PEG), poly(phosphazene), and poly (lactic-co-glycolic acid) (PLGA), offering versatility and tunability for various biomedical applications [146,147].
In general, these materials are low-viscosity sols at low or room temperature, making it easy to entrapped anticancer therapeutic agents in polymer-, polypeptide-, and protein-based hydrogels [148]. Following injection into the body, they undergo a temperature-triggered sol–gel transition, transforming into semisolid gels. This enables the controlled release of the loaded anticancer drugs from the hydrogel in a predetermined and regulated manner [149]. Among them, thermosensitive hydrogels composed of PLGA and PEG triblock copolymers have gained popularity due to their convenient one-pot synthesis method and a good safety profile. Furthermore, both are FDA approved (PLGA and PEG) and have been used clinically for many years [150]. Chen X. et al. developed thermo-responsive injectable hydrogels composed of PLGA-PEG-PLGA triblock copolymers, with different ratios of PLGA and PEG using bulk ring-opening copolymerization [142]. The mixture of PLGA and PEG was optimized in such a way that the hydrogel exhibited sol–gel transitions in water as the temperature increased. Then, Herceptin was loaded into this designed hydrogel and then its in vitro and in vivo antitumor efficacy was assessed using a HER2+ breast cancer tumor model (Figure 6).
A single injection of the Herceptin-loaded hydrogel, combined with weekly injections of a Herceptin solution, effectively prevented tumor recurrence in a locally recurrent HER2+ breast tumor nude mouse model. Furthermore, the weekly pulsed injection of the Herceptin solution led to a significant decrease in the left ventricular ejection fraction (LVEF) in both the anticancer and anti-relapse studies. Conversely, the slow and sustained release of Herceptin from the hydrogel depot effectively mitigated cardiotoxicity. These findings suggest that the hydrogel system can enhance therapeutic efficacy, reduce systemic adverse effects, and decrease the frequency of administration in the treatment of HER2+ breast tumors. In addition to PLGA and PEG polymers, various other polymers have been utilized, such as chitosan [129], MXene nanosheets [133], L-lysine- and L-alanine-based diblock copolypeptides [131], and poly(N-isopropylacrylamide) [134,136], to synthesize thermo-responsive hydrogels. Anticancer drugs such as DOX, PTX, and CTX have been encapsulated within polymer-grafted hydrogels, which release them at predetermined times at the tumor target site. This ultimately enhances the anticancer efficacy of the agents.

4.3.2. Light-Responsive Hydrogels

Light-responsive hydrogels represent a cutting-edge class of biomaterials designed to undergo controlled changes in response to light [151]. These hydrogels typically incorporate light-sensitive moieties, such as photoactive molecules or photo-responsive polymers, enabling precise spatiotemporal control over their properties and functions [152]. Through techniques like photopolymerization or photoisomerization, light can trigger reversible alterations in the hydrogel’s structure, leading to phenomena such as swelling, contraction, or release of encapsulated drugs for cancer therapy [153]. This unique responsiveness to light makes these hydrogels invaluable for a wide range of applications, including drug delivery, tissue engineering, biosensing, and actuators in soft robotics, offering unprecedented levels of precision and versatility in the biomedical and technological fields [154]. Drug release mechanisms including photo-induced transformation [155], bond cleavage [156], photoisomerization [157], or photo-cross-linking [158] can be activated by UV light. Consequently, certain linkers containing aromatic rings are susceptible to UV light-induced drug release via one or more photon excitation mechanisms operating at specific wavelengths. Hou M. et al. developed a light-induced hydrogel composed of the natural product humic acid/agarose [140]. This hydrogel was utilized to encapsulate sodium humate (SH) and DOX, resulting in combined chemo- and photothermal therapeutic effects (Figure 7A). Since SH is a strong absorber of near-infrared (NIR) light, it has the ability to effectively convert light energy into thermal energy, which causes localized hyperthermia. This, in turn, causes sustained drug release from the hydrogel complex via a standard gel–sol transition, which improves the uptake of therapeutic drugs by cells. Furthermore, when solid tumors were exposed to NIR laser radiation, intratumoral injection of the drug-loaded hydrogel produced a simultaneous chemo-photothermal therapeutic response that may prevent overall tumor recurrence. In vivo studies suggested an improved antineoplastic efficacy of this prepared hydrogel compared to local delivery of free drugs to tumoral tissues (Figure 7B).
This research offers a promising start towards creating an affordable, light-responsive hydrogel for targeted tumor treatment. Chondroitin sulfate methacryloyl (CSMA)- [138] and methylcellulose [141]-like polymers are being investigated for crafting light-induced hydrogels, which could be pivotal in cancer treatment. These innovative hydrogels that are responsive to light stimuli, can adeptly encapsulate potent therapeutic agents like DOX, facilitating precise chemo- and photothermal therapy, thereby heightening efficacy in combating cancer cells [137,140,159].

4.4. Multi-Responsive Hydrogels

Multi-responsive hydrogels exhibit dynamic behavior in response to multiple external stimuli, such as pH, temperature, light, ions, or enzymes [160]. These versatile materials possess the ability to undergo reversible changes in their structure, properties, and functionalities upon exposure to specific physical and chemical environmental cues, allowing for the fine tuning of their behavior [98]. By incorporating multiple responsive elements into their polymer networks, these hydrogels offer enhanced adaptability and functionality, making them highly attractive for a wide range of applications in drug delivery, tissue engineering, sensing, and soft robotics [161]. Their ability to integrate and respond to various stimuli provides researchers with a powerful platform for designing sophisticated biomaterials capable of addressing complex biological and technological challenges, with potential implications for personalized medicine, regenerative therapies, and advancements in material science [162].
Multi-responsive hydrogels are comprised of three-dimensional polymer networks constructed from either natural or synthetic polymers [163]. Commonly used natural polymers include alginate, chitosan, collagen, hyaluronic acid, and gelatin. Additionally, synthetic polymers like polyethylene glycol (PEG), polyacrylamide (PAAm), and poly(N-isopropylacrylamide) (PNIPAAm) have found wide application in hydrogel synthesis (Table 4). The selection of the polymer is contingent upon the desired properties and intended applications of the hydrogel.
Liang, Y. et al. designed a pH and glucose dual-responsive multifunctional hydrogel for the specific release of metformin [121]. The hydrogel was formed using phenylboronic acid and benzaldehyde bifunctional polyethylene glycol-co-poly(glycerol sebacic acid)/dihydrocaffeic acid and l-arginine-cografted chitosan (PEGS-PBA-BA/CS-DA-LAG) (PC hydrogels) via dynamic bonds between Schiff bases and phenylboronate esters. The conductivity and hemostasis were enhanced with rGO@PDA. They evaluated tissue adhesion, blood coagulation, in vivo hemostasis, biocompatibility, antibacterial and antioxidant effects, the self-healing ability, rheological and mechanical properties, as well as in vitro metformin release. They also assessed the healing of type II diabetes foot wounds in a rat model. Their research revealed that the developed PC hydrogels were effective in accelerating the healing of chronic athletic diabetic foot wounds. These hydrogels displayed increased adhesion, stimuli-responsive metformin release, and self-healing properties.
Another study conducted by Wu Y. et al. developed a pH/ROS dual-sensitive injectable glycopeptide hydrogel by combining PBA-grafted oxidized dextran (POD) and caffeic acid-grafted ε-polylysine (CE) [167]. The hydrogel exhibited inherent antibacterial and antioxidant properties. pH-responsive micelles were used to encapsulate the angiogenesis-promoting compound mangiferin (MF). Following this, drug-loaded micelles and diclofenac sodium (DS), which is well known for its anti-inflammatory qualities, were added to the hydrogel. The results of the in vitro and in vivo experiments demonstrated the biocompatibility of the hydrogel in addition to its first-rate anti-infection, anti-oxidation, and anti-inflammatory qualities, which were subsequently followed by enhanced angiogenesis and expedited wound healing [167]. All things considered, this innovative glycopeptide hydrogel offers a straightforward and effective method for curing long-term diabetic wounds.
Recent advancements in this field include the development of smart hydrogels capable of responding to dynamic changes within the TME, offering personalized and responsive treatment strategies. The available literature highlights the potential of multi-responsive hydrogels in revolutionizing cancer therapy through enhanced drug delivery and therapeutic efficacy [171,172]. Multi-stimuli-responsive hydrogels, synthesized from hyaluronic acid and a diselenide-based cross-linker, can facilitate the controlled release of DOX [173]. In vitro studies confirm the hydrogels’ cytocompatibility and demonstrated their comparable antitumor efficacy to free DOX. Moreover, DOX + ICG-loaded hydrogels display enhanced antitumor effects post-NIR irradiation [173]. Gou S. et al. also synthesized silk fibroin-based hydrogels (HSFs) for localized treatment [174]. The resulting HSF hydrogel demonstrated a clear thixotropic behavior, viscoelasticity, and self-healing capabilities. Interestingly, upon the usage of different stimuli like ROS, acidity, glutathione, heat, and NIR irradiation, this hydrogel also demonstrated good stimuli-responsive drug release characteristics (Figure 8).
This implies that in reaction to the NIR irradiation and the TME, the hydrogel could enable on-demand drug release at the desire location and time. The combination of an intratumoral injection of the prepared DOX-loaded hydrogel and NIR irradiation produced the most potent anticancer effect of all the treatment groups. This demonstrates the strong synergistic benefits of photothermal, photodynamic, and chemotherapy treatments. Interestingly, the hydrogel was able to almost completely remove the tumor masses, significantly prolonging the tumor-bearing mice’s survival time for almost 60 days without causing any negative effects (Figure 8). This underscores the potential of developing a multi-stimuli-responsive injectable HSF hydrogel as a targeted and cooperative cancer treatment platform.
Nanogels are nanoscale hydrogels, and can offer similar properties to macroscopic hydrogels but with the advantage of intravenous injection, enabling access to hard-to-reach body parts through enhanced permeability and retention (EPR). Their cellular uptake makes them ideal for safely delivering therapeutic substances, such as chemotherapeutics and nucleic acid-based medications, directly to the target cell’s cytoplasm [175]. As an example, for the delivery of DOX, Xuejiao Zhang et al. developed a pH and redox dual-responsive prodrug nanogel using an inverse nanoprecipitation technique (Figure 9A) [176]. Microscopy and flow cytometry studies on cell cultures demonstrated that the tumor cells effectively internalized the prodrug nanogels (Figure 9C).
The various nanogels and microgels that were previously discussed have shown significant potential for the encapsulation, intracellular delivery, and facilitating the enhanced anticancer efficiency of therapeutic drugs. The research findings indicate that single-, dual-, and multi-stimuli-responsive hydrogels represent a promising platform for targeted cancer treatment.

5. Limitations of Hydrogels

Stimulus-responsive hydrogels offer versatile platforms for drug delivery and tissue engineering, yet they are not without limitations. Hydrogels, while versatile in drug delivery, present drawbacks including a limited loading capacity, potential burst release leading to uneven drug concentrations, stability issues under varying environmental conditions, and sometimes complex fabrication processes [177]. Another significant challenge is achieving precise control over their responsiveness to external cues such as pH, temperature, or biomolecular signals, which can vary widely within biological systems [178]. This lack of fine tuning may lead to inadequate therapeutic responses or unintended side effects. Moreover, the mechanical properties of stimuli-responsive hydrogels, including their stiffness and elasticity, may not always match the requirements of the target tissue, potentially limiting their applicability in certain clinical scenarios [161]. Hydrogels like calcium alginate have low mechanical strength whereas agar/alginate hydrogel beads provide improved mechanical strength and controlled drug release [179,180].
Pre-formed hydrogels and scaffolds in tissue engineering face challenges such as infection risk and difficulties with surgical implantation. Injectable hydrogels systems offer a solution to some of these problems [181,182]. Hydrogels are often insufficient carriers for small-molecular-weight and hydrophobic drugs [183,184]. Copolymers of methyl methacrylate and acrylic acid can incorporate drug-loaded colloidal carriers to achieve homogeneous dispersion of hydrophobic drugs [185]. Introducing hydrophobic domains and using stimulated release triggers can improve controlled drug release [186]. Incorporating drug-loaded colloidal carriers into hydrogels is challenging, prompting the development of mixed delivery systems such as liposome-in-hydrogel formulations. pH-dependent hydrogels often exhibit limited solubility, and a slow sol–gel transition phase. Chemical modifications can enhance their solubility profile and mucoadhesive property [187]. Light-responsive hydrogels tend to have slow and variable responses to light stimuli. Solutions include fabricating cross-linked polymers of 2-hydroxyethyl methacrylate functionalized with azobenzene groups and forming an interpenetrating polymer network between polyacrylamide and polyacrylic acid [188,189].
Additionally, concerns regarding biocompatibility, immunogenicity, toxicity, and the potential for the long-term accumulation of degradation by-products in vivo necessitate rigorous evaluations for their safe and effective use in clinical practices [190]. The challenges found with hydrogels reveal an opportunity to enhance them by improving their responsiveness to various signals, such as temperature and light changes. This could result in the development of smarter and more sensitive hydrogels that respond better to stimuli like temperature, significantly enhancing the precision of treatment delivery and facilitating their use in medical applications.

6. Conclusions and Future Directions

Stimulus-responsive hydrogels represent a rapidly advancing field that is setting new benchmarks for the advancement of drug delivery platforms, offering numerous benefits over conventional therapeutics and delivery methods. As outlined in this review, the versatility of stimulus-responsive hydrogels allows for the targeted engineering for diverse biomedical applications, demonstrating a gain-of-function effect. This review presents the promising potential of stimulus-responsive hydrogels that are engineered considering biological, chemical, and physical stimuli, particularly for drug delivery in cancer therapy. While this review primarily addressed drug delivery applications of stimulus-responsive hydrogels, the ongoing research extends across multiple domains including controlled drug delivery, biocatalysis, optical sensing, imaging, cell encapsulation, and tissue engineering. In the future, synergistic therapeutic approaches (i.e., local and systemic) to regulate prolonged drug release via specialized stimulus responsive hydrogels could signify a new paradigm in cancer treatment. Future studies should also aim to further enhance the application and effectiveness of stimulus-responsive hydrogels in these diverse areas.

Author Contributions

R.S.: Conceptualization, Writing—Original Draft, Data Curation, Software. D.B.: Conceptualization, Writing—Review and Editing, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

R.S. acknowledges the Science and Engineering Research Board (SERB), Govt. of India, for a National Post Doctoral Fellowship (NPDF). R.S. and D.B. acknowledge the Indian Institute of Technology (IIT), Gandhinagar, for their financial support and facilities.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are openly available in article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bray, F.; Laversanne, M.; Sung, H.; Ferlay, J.; Siegel, R.L.; Soerjomataram, I.; Jemal, A. Global Cancer Statistics 2022 GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2024, 74, 229–263. [Google Scholar] [CrossRef] [PubMed]
  2. Kumar, R.R.; Kumar, A.; Chuang, C.-H.; Shaikh, M.O. Recent Advances and Emerging Trends in Cancer Biomarker Detection Technologies. Ind. Eng. Chem. Res. 2023, 62, 5691–5713. [Google Scholar] [CrossRef]
  3. Zhu, Y.; Jia, H.; Duan, Q.; Wu, F. Nanomedicines for Combating Multidrug Resistance of Cancer. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2021, 13, e1715. [Google Scholar] [CrossRef] [PubMed]
  4. Passaro, A.; Al Bakir, M.; Hamilton, E.G.; Diehn, M.; André, F.; Roy-Chowdhuri, S.; Mountzios, G.; Wistuba, I.I.; Swanton, C.; Peters, S. Cancer Biomarkers: Emerging Trends and Clinical Implications for Personalized Treatment. Cell 2024, 187, 1617–1635. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, X.; Jiang, H.; Wang, X. Advances in Cancer Research: Current and Future Diagnostic and Therapeutic Strategies. Biosensors 2024, 14, 100. [Google Scholar] [CrossRef] [PubMed]
  6. Gupta, D.; Roy, P.; Sharma, R.; Kasana, R.; Rathore, P.; Gupta, T.K. Recent Nanotheranostic Approaches in Cancer Research. Clin. Exp. Med. 2024, 24, 8. [Google Scholar] [CrossRef] [PubMed]
  7. Sufyan, M.; Shokat, Z.; Ashfaq, U.A. Artificial intelligence in cancer diagnosis and therapy: Current status and future perspective. Comput. Biol. Med. 2023, 165, 107356. [Google Scholar] [CrossRef] [PubMed]
  8. Liang, A.; Kong, Y.; Chen, Z.; Qiu, Y.; Wu, Y.; Zhu, X.; Li, Z. Advancements and Applications of Single-Cell Multi-Omics Techniques in Cancer Research: Unveiling Heterogeneity and Paving the Way for Precision Therapeutics. Biochem. Biophys. Rep. 2024, 37, 101589. [Google Scholar] [CrossRef]
  9. Ogasawara, M. Wilms’ Tumor 1-Targeting Cancer Vaccine: Recent Advancements and Future Perspectives. Hum. Vaccin. Immunother. 2024, 20, 2296735. [Google Scholar] [CrossRef]
  10. Mir, M.; Ishtiaq, S.; Rabia, S.; Khatoon, M.; Zeb, A.; Khan, G.M.; ur Rehman, A.; ud Din, F. Nanotechnology: From in Vivo Imaging System to Controlled Drug Delivery. Nanoscale Res. Lett. 2017, 12, 500. [Google Scholar] [CrossRef]
  11. Solanki, R.; Shankar, A.; Modi, U.; Patel, S. New Insights from Nanotechnology in SARS-CoV-2 Detection, Treatment Strategy and Prevention. Mater. Today Chem. 2023, 29, 101478. [Google Scholar] [CrossRef] [PubMed]
  12. Solanki, R.; Jodha, B.; Prabina, K.E.; Aggarwal, N.; Patel, S. Recent Advances in Phytochemical Based Nano-Drug Delivery Systems to Combat Breast Cancer: A Review. J. Drug Deliv. Sci. Technol. 2022, 77, 103832. [Google Scholar] [CrossRef]
  13. Qiao, Y.; Wan, J.; Zhou, L.; Ma, W.; Yang, Y.; Luo, W.; Yu, Z.; Wang, H. Stimuli-responsive Nanotherapeutics for Precision Drug Delivery and Cancer Therapy. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol 2019, 11, e1527. [Google Scholar] [CrossRef] [PubMed]
  14. Geckil, H.; Xu, F.; Zhang, X.; Moon, S.; Demirci, U. Engineering Hydrogels as Extracellular Matrix Mimics. Nanomedicine 2010, 5, 469–484. [Google Scholar] [CrossRef] [PubMed]
  15. Thakuria, A.; Kataria, B.; Gupta, D. Nanoparticle-Based Methodologies for Targeted Drug Delivery—An Insight. J. Nanopart. Res. 2021, 23, 87. [Google Scholar] [CrossRef]
  16. Niazi, M.; Alizadeh, E.; Zarebkohan, A.; Seidi, K.; Ayoubi-Joshaghani, M.H.; Azizi, M.; Dadashi, H.; Mahmudi, H.; Javaheri, T.; Jaymand, M. Advanced Bioresponsive Multitasking Hydrogels in the New Era of Biomedicine. Adv. Funct. Mater. 2021, 31, 2104123. [Google Scholar] [CrossRef]
  17. Solanki, R.; Patel, S. Protein Nanocarriers for the Delivery of Phytoconstituents. In Nanotechnology Based Delivery of Phytoconstituents and Cosmeceuticals; Pooja, D., Kulhari, H., Eds.; Springer Nature: Singapore, 2024; pp. 229–264. ISBN 978-981-99-5314-1. [Google Scholar]
  18. Anderson, N.M.; Simon, M.C. The Tumor Microenvironment. Curr. Biol. 2020, 30, R921–R925. [Google Scholar] [CrossRef] [PubMed]
  19. Truffi, M.; Sorrentino, L.; Corsi, F. Fibroblasts in the Tumor Microenvironment; Springer: Berlin/Heidelberg, Germany, 2020; pp. 15–29. [Google Scholar]
  20. Wang, L.; Huo, M.; Chen, Y.; Shi, J. Tumor Microenvironment-enabled Nanotherapy. Adv. Heal. Mater. 2018, 7, 1701156. [Google Scholar] [CrossRef] [PubMed]
  21. Gong, F.; Yang, N.; Wang, X.; Zhao, Q.; Chen, Q.; Liu, Z.; Cheng, L. Tumor Microenvironment-Responsive Intelligent Nanoplatforms for Cancer Theranostics. Nano Today 2020, 32, 100851. [Google Scholar] [CrossRef]
  22. Wichterle, O.; Lim, D. Hydrophilic Gels for Biological Use. Nature 1960, 185, 117–118. [Google Scholar] [CrossRef]
  23. Nogueira, N.; Conde, O.; Minones, M.; Trillo, J.M.; Miñones, J., Jr. Characterization of Poly (2-Hydroxyethyl Methacrylate)(PHEMA) Contact Lens Using the Langmuir Monolayer Technique. J. Colloid. Interface Sci. 2012, 385, 202–210. [Google Scholar] [CrossRef] [PubMed]
  24. Malda, J.; Visser, J.; Melchels, F.P.; Jüngst, T.; Hennink, W.E.; Dhert, W.J.A.; Groll, J.; Hutmacher, D.W. 25th Anniversary Article: Engineering Hydrogels for Biofabrication. Adv. Mater. 2013, 25, 5011–5028. [Google Scholar] [CrossRef] [PubMed]
  25. Lin, C.-C.; Metters, A.T. Hydrogels in Controlled Release Formulations: Network Design and Mathematical Modeling. Adv. Drug Deliv. Rev. 2006, 58, 1379–1408. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, H.; Zhai, Y.; Wang, J.; Zhai, G. New Progress and Prospects: The Application of Nanogel in Drug Delivery. Mater. Sci. Eng. C 2016, 60, 560–568. [Google Scholar] [CrossRef]
  27. Bustamante-Torres, M.; Romero-Fierro, D.; Arcentales-Vera, B.; Palomino, K.; Magaña, H.; Bucio, E. Hydrogels Classification According to the Physical or Chemical Interactions and as Stimuli-Sensitive Materials. Gels 2021, 7, 182. [Google Scholar] [CrossRef] [PubMed]
  28. Mohapatra, S.; Mirza, M.A.; Hilles, A.R.; Zakir, F.; Gomes, A.C.; Ansari, M.J.; Iqbal, Z.; Mahmood, S. Biomedical Application, Patent Repository, Clinical Trial and Regulatory Updates on Hydrogel: An Extensive Review. Gels 2021, 7, 207. [Google Scholar] [CrossRef] [PubMed]
  29. Yang, J.; Chen, Y.; Zhao, L.; Zhang, J.; Luo, H. Constructions and Properties of Physically Cross-Linked Hydrogels Based on Natural Polymers. Polym. Rev. 2023, 63, 574–612. [Google Scholar] [CrossRef]
  30. Bashir, S.; Hina, M.; Iqbal, J.; Rajpar, A.H.; Mujtaba, M.A.; Alghamdi, N.A.; Wageh, S.; Ramesh, K.; Ramesh, S. Fundamental Concepts of Hydrogels: Synthesis, Properties, and Their Applications. Polymers 2020, 12, 2702. [Google Scholar] [CrossRef] [PubMed]
  31. Chen, T.; Bakhshi, H.; Liu, L.; Ji, J.; Agarwal, S. Combining 3D Printing with Electrospinning for Rapid Response and Enhanced Designability of Hydrogel Actuators. Adv. Funct. Mater. 2018, 28, 1800514. [Google Scholar] [CrossRef]
  32. Zaszczyńska, A.; Niemczyk-Soczynska, B.; Sajkiewicz, P. A Comprehensive Review of Electrospun Fibers, 3D-Printed Scaffolds, and Hydrogels for Cancer Therapies. Polymers 2022, 14, 5278. [Google Scholar] [CrossRef]
  33. Kilic Boz, R.; Aydin, D.; Kocak, S.; Golba, B.; Sanyal, R.; Sanyal, A. Redox-Responsive Hydrogels for Tunable and “On-Demand” Release of Biomacromolecules. Bioconjug Chem. 2022, 33, 839–847. [Google Scholar] [CrossRef] [PubMed]
  34. Jacob, J.M.; Woldu, S.L.; Linehan, J.; Labbate, C.; Rose, K.M.; Sexton, W.J.; Tachibana, I.; Kaimakliotis, H.; Nieder, A.; Bjurlin, M.A. First Analysis of the Safety and Efficacy of UGN-101 in the Treatment of Ureteral Tumors. Urol. Oncol. Semin. Orig. Investig. 2024, 42, 20.e17–20.e23. [Google Scholar] [CrossRef] [PubMed]
  35. Woolen, S.; Holzmeyer, C.; Nesbitt, E.; Siami, P.F. Long-Term Efficacy and Tolerability of Abdominal Once-Yearly Histrelin Acetate Subcutaneous Implants in Patients with Advanced Prostate Cancer. Prostate Cancer 2014, 2014, 490315. [Google Scholar] [CrossRef] [PubMed]
  36. Porta, C.; Giannatempo, P.; Rizzo, M.; Lucarelli, G.; Ditonno, P.; Battaglia, M. An Evaluation of UGN-101, a Sustained-Release Hydrogel Polymer-Based Formulation Containing Mitomycin-C, for the Treatment of Upper Urothelial Carcinomas. Expert. Opin. Pharmacother. 2020, 21, 2199–2204. [Google Scholar] [CrossRef]
  37. Rafael, D.; Melendres, M.M.R.; Andrade, F.; Montero, S.; Martinez-Trucharte, F.; Vilar-Hernandez, M.; Durán-Lara, E.F.; Schwartz, S., Jr.; Abasolo, I. Thermo-Responsive Hydrogels for Cancer Local Therapy: Challenges and State-of-Art. Int. J. Pharm. 2021, 606, 120954. [Google Scholar] [CrossRef] [PubMed]
  38. Andrade, F.; Durán-Lara, E.; Rafael, D. Multicomponent Hydrogels for Cancer Diagnosis and Therapy; The Royal Society of Chemistry: London, UK, 2023; pp. 542–577. [Google Scholar] [CrossRef]
  39. Abulateefeh, S.R. Long-Acting Injectable PLGA/PLA Depots for Leuprolide Acetate: Successful Translation from Bench to Clinic. Drug Deliv. Transl. Res. 2023, 13, 520–530. [Google Scholar] [CrossRef] [PubMed]
  40. Municoy, S.; Alvarez Echazu, M.I.; Antezana, P.E.; Galdopórpora, J.M.; Olivetti, C.; Mebert, A.M.; Foglia, M.L.; Tuttolomondo, M.V.; Alvarez, G.S.; Hardy, J.G. Stimuli-Responsive Materials for Tissue Engineering and Drug Delivery. Int. J. Mol. Sci. 2020, 21, 4724. [Google Scholar] [CrossRef] [PubMed]
  41. Qiu, Y.; Park, K. Environment-Sensitive Hydrogels for Drug Delivery. Adv. Drug Deliv. Rev. 2001, 53, 321–339. [Google Scholar] [CrossRef] [PubMed]
  42. Raza, A.; Hayat, U.; Rasheed, T.; Bilal, M.; Iqbal, H.M.N. Redox-Responsive Nano-Carriers as Tumor-Targeted Drug Delivery Systems. Eur. J. Med. Chem. 2018, 157, 705–715. [Google Scholar] [CrossRef]
  43. Rizwan, M.; Yahya, R.; Hassan, A.; Yar, M.; Azzahari, A.D.; Selvanathan, V.; Sonsudin, F.; Abouloula, C.N. PH Sensitive Hydrogels in Drug Delivery: Brief History, Properties, Swelling, and Release Mechanism, Material Selection and Applications. Polymers 2017, 9, 137. [Google Scholar] [CrossRef]
  44. Ding, H.; Tan, P.; Fu, S.; Tian, X.; Zhang, H.; Ma, X.; Gu, Z.; Luo, K. Preparation and Application of PH-Responsive Drug Delivery Systems. J. Control. Release 2022, 348, 206–238. [Google Scholar] [CrossRef] [PubMed]
  45. Ramburrun, P.; Khan, R.A.; Choonara, Y.E. Design, Preparation, and Functionalization of Nanobiomaterials for Enhanced Efficacy in Current and Future Biomedical Applications. Nanotechnol. Rev. 2022, 11, 1802–1826. [Google Scholar] [CrossRef]
  46. Yang, Z.; McClements, D.J.; Li, C.; Sang, S.; Chen, L.; Long, J.; Qiu, C.; Jin, Z. Targeted Delivery of Hydrogels in Human Gastrointestinal Tract: A Review. Food Hydrocoll. 2023, 134, 108013. [Google Scholar] [CrossRef]
  47. Alvarez-Lorenzo, C.; Blanco-Fernandez, B.; Puga, A.M.; Concheiro, A. Crosslinked Ionic Polysaccharides for Stimuli-Sensitive Drug Delivery. Adv. Drug Deliv. Rev. 2013, 65, 1148–1171. [Google Scholar] [CrossRef] [PubMed]
  48. Chen, S.; Liu, M.; Jin, S.; Chen, Y. Synthesis and Swelling Properties of PH-sensitive Hydrogels Based on Chitosan and Poly (Methacrylic Acid) Semi-interpenetrating Polymer Network. J. Appl. Polym. Sci. 2005, 98, 1720–1726. [Google Scholar] [CrossRef]
  49. Gomte, S.S.; Agnihotri, T.G.; Khopade, S.; Jain, A. Exploring the potential of pH-sensitive polymers in targeted drug delivery. J. Biomater. Sci. Polym. Ed. 2024, 35, 228–268. [Google Scholar] [CrossRef]
  50. Ortiz, J.A.; Sepúlveda, F.A.; Panadero-Medianero, C.; Murgas, P.; Ahumada, M.; Palza, H.; Matsuhiro, B.; Zapata, P.A. Cytocompatible Drug Delivery Hydrogels Based on Carboxymethylagarose/Chitosan PH-Responsive Polyelectrolyte Complexes. Int. J. Biol. Macromol. 2022, 199, 96–107. [Google Scholar] [CrossRef] [PubMed]
  51. Li, L.; Gu, J.; Zhang, J.; Xie, Z.; Lu, Y.; Shen, L.; Dong, Q.; Wang, Y. Injectable and Biodegradable PH-Responsive Hydrogels for Localized and Sustained Treatment of Human Fibrosarcoma. ACS Appl. Mater. Interfaces 2015, 7, 8033–8040. [Google Scholar] [CrossRef] [PubMed]
  52. Patel, V.R.; Amiji, M.M. Preparation and Characterization of Freeze-Dried Chitosan-Poly(Ethylene Oxide) Hydrogels for Site-Specific Antibiotic Delivery in the Stomach. Pharm. Res. 1996, 13, 588–593. [Google Scholar] [CrossRef]
  53. Qu, X.; Wirsen, A.; Albertsson, A.-C. Novel PH-Sensitive Chitosan Hydrogels: Swelling Behavior and States of Water. Polymer 2000, 41, 4589–4598. [Google Scholar] [CrossRef]
  54. Qu, J.; Zhao, X.; Ma, P.X.; Guo, B. PH-Responsive Self-Healing Injectable Hydrogel Based on N-Carboxyethyl Chitosan for Hepatocellular Carcinoma Therapy. Acta Biomater. 2017, 58, 168–180. [Google Scholar] [CrossRef] [PubMed]
  55. Abdelaal, M.Y.; Abdel-Razik, E.A.; Abdel-Bary, E.M.; El-Sherbiny, I.M. Chitosan-based Interpolymeric PH-responsive Hydrogels for in Vitro Drug Release. J. Appl. Polym. Sci. 2007, 103, 2864–2874. [Google Scholar] [CrossRef]
  56. Goycoolea, F.M.; Heras, A.; Aranaz, I.; Galed, G.; Fernández-Valle, M.E.; Argüelles-Monal, W. Effect of Chemical Crosslinking on the Swelling and Shrinking Properties of Thermal and PH-responsive Chitosan Hydrogels. Macromol. Biosci. 2003, 3, 612–619. [Google Scholar] [CrossRef]
  57. Bilia, A.; Carelli, V.; Di Colo, G.; Nannipieri, E. In Vitro Evaluation of a PH-Sensitive Hydrogel for Control of GI Drug Delivery from Silicone-Based Matrices. Int. J. Pharm. 1996, 130, 83–92. [Google Scholar] [CrossRef]
  58. Monir, T.S.B.; Afroz, S.; Khan, R.A.; Miah, M.Y.; Takafuji, M.; Alam, M.A. PH-Sensitive Hydrogel from Polyethylene Oxide and Acrylic Acid by Gamma Radiation. J. Compos. Sci. 2019, 3, 58. [Google Scholar] [CrossRef]
  59. Nho, Y.C.; Lim, Y.M.; Lee, Y.M. Preparation, Properties and Biological Application of PH-Sensitive Poly(Ethylene Oxide)(PEO) Hydrogels Grafted with Acrylic Acid (AAc) Using Gamma-Ray Irradiation. Radiat. Phys. Chem. 2004, 71, 239–242. [Google Scholar] [CrossRef]
  60. Amiji, M.; Tailor, R.; Ly, M.-K.; Goreham, J. Gelatin-Poly(Ethylene Oxide) Semi-Interpenetrating Polymer Network with PH-Sensitive Swelling and Enzyme-Degradable Properties for Oral Drug Delivery. Drug Dev. Ind. Pharm. 1997, 23, 575–582. [Google Scholar] [CrossRef]
  61. Tenório-Neto, E.T.; Guilherme, M.R.; Lima-Tenório, M.K.; Scariot, D.B.; Nakamura, C.V.; Rubira, A.F.; Kunita, M.H. Synthesis and Characterization of a PH-Responsive Poly(Ethylene Glycol)-Based Hydrogel: Acid Degradation, Equilibrium Swelling, and Absorption Kinetic Characteristics. Colloid. Polym. Sci. 2015, 293, 3611–3622. [Google Scholar] [CrossRef]
  62. Wang, S.; Attah, R.; Li, J.; Chen, Y.; Chen, R. A PH-Responsive Amphiphilic Hydrogel Based on Pseudopeptides and Poly(Ethylene Glycol) for Oral Delivery of Hydrophobic Drugs. ACS Biomater. Sci. Eng. 2018, 4, 4236–4243. [Google Scholar] [CrossRef]
  63. Mahmood, T.; Sarfraz, R.M.; Mahmood, A.; Salem-Bekhit, M.M.; Ijaz, H.; Zaman, M.; Akram, M.R.; Taha, E.I.; Sahu, R.K.; Benguerba, Y. Preparation, In Vitro Characterization, and Evaluation of Polymeric PH-Responsive Hydrogels for Controlled Drug Release. ACS Omega 2024, 9, 10498–10516. [Google Scholar] [CrossRef]
  64. Halacheva, S.S.; Freemont, T.J.; Saunders, B.R. PH-Responsive Physical Gels from Poly(Meth)Acrylic Acid-Containing Crosslinked Particles: The Relationship between Structure and Mechanical Properties. J. Mater. Chem. B 2013, 1, 4065–4078. [Google Scholar] [CrossRef] [PubMed]
  65. Milosavljević, N.B.; Milašinović, N.Z.; Popović, I.G.; Filipović, J.M.; Kalagasidis Krušić, M.T. Preparation and Characterization of PH-sensitive Hydrogels Based on Chitosan, Itaconic Acid and Methacrylic Acid. Polym. Int. 2011, 60, 443–452. [Google Scholar] [CrossRef]
  66. Frutos, G.; Prior-Cabanillas, A.; París, R.; Quijada-Garrido, I. A Novel Controlled Drug Delivery System Based on PH-Responsive Hydrogels Included in Soft Gelatin Capsules. Acta Biomater. 2010, 6, 4650–4656. [Google Scholar] [CrossRef]
  67. Oh, J.; Kim, B. Mucoadhesive and PH-Responsive Behavior of Gelatin Containing Hydrogels for Protein Drug Delivery Applications. Korea-Aust. Rheol. J. 2020, 32, 41–46. [Google Scholar] [CrossRef]
  68. Hussain, K.; Aslam, Z.; Ullah, S.; Shah, M.R. Synthesis of PH Responsive, Photocrosslinked Gelatin-Based Hydrogel System for Control Release of Ceftriaxone. Chem. Phys. Lipids 2021, 238, 105101. [Google Scholar] [CrossRef] [PubMed]
  69. Jafari, H.; Namazi, H. PH-Sensitive Biosystem Based on Laponite RD/Chitosan/Polyvinyl Alcohol Hydrogels for Controlled Delivery of Curcumin to Breast Cancer Cells. Colloids Surf. B Biointerfaces 2023, 231, 113585. [Google Scholar] [CrossRef]
  70. Asad, S.; Khan, S.A.; Ullah, K.; Mannan, A.; Murtaza, G. Synthesis, Characterization, and in-Vitro Evaluation of PH-Responsive PEI-MAA Polymeric Matrices Decorated with Mesalazine for Colonic Delivery. J. Drug Deliv. Sci. Technol. 2023, 88, 104926. [Google Scholar] [CrossRef]
  71. Noor, F.; Mahmood, A.; Zafar, N.; Sarfraz, R.M.; Rehman, U.; Ijaz, H.; Hussain, Z.; Ahmed, I.A.; Imam, M.T.; Al Abdulmonem, W. Fabrication of PH-Responsive Hydrogels of Perindopril Erbumine Using Black Seed Extract and β-Cyclodextrin Co-Polymerized with Methacrylic Acid and Methylene Bisacrylamide. J. Drug Deliv. Sci. Technol. 2023, 88, 104924. [Google Scholar] [CrossRef]
  72. Yuan, L.; Li, Z.; Li, X.; Qiu, S.; Lei, J.; Li, D.; Mu, C.; Ge, L. Functionalization of an Injectable Self-Healing PH-Responsive Hydrogel by Incorporating a Curcumin/Polymerized β-Cyclodextrin Inclusion Complex for Selective Toxicity to Osteosarcoma. ACS Appl. Polym. Mater. 2022, 4, 1243–1254. [Google Scholar] [CrossRef]
  73. Yang, X.; Kim, J. Β-Cyclodextrin Hydrogels Containing Naphthaleneacetic Acid for PH-sensitive Release. Biotechnol. Bioeng. 2010, 106, 295–302. [Google Scholar] [CrossRef]
  74. Li, Z.; Crago, M.; Schofield, T.; Zeng, H.; Vyas, H.K.N.; Müllner, M.; Mai-Prochnow, A.; Farajikhah, S.; Naficy, S.; Dehghani, F. Synthesis and Evaluation of Functionalized Polyurethanes for PH-Responsive Delivery of Compounds in Chronic Wounds. Gels 2023, 9, 611. [Google Scholar] [CrossRef] [PubMed]
  75. Xu, L.; Qiu, L.; Sheng, Y.; Sun, Y.; Deng, L.; Li, X.; Bradley, M.; Zhang, R. Biodegradable PH-Responsive Hydrogels for Controlled Dual-Drug Release. J. Mater. Chem. B 2018, 6, 510–517. [Google Scholar] [CrossRef]
  76. Zhang, T.; Cheng, X.; Xiu, J.; Liu, M.; Liu, S.; Zhang, B.; Miao, Q.; Cun, D.; Yang, C.; Li, K. PH-Responsive Injectable Multifunctional Pluronic F127/Gelatin-Based Hydrogels with Hydrogen Production for Treating Diabetic Wounds. ACS Appl. Mater. Interfaces 2023, 15, 55392–55408. [Google Scholar] [CrossRef] [PubMed]
  77. Meena, P.; Singh, P.; Warkar, S.G. Development and Assessment of Carboxymethyl Tamarind Kernel Gum-Based PH-Responsive Hydrogel for Release of Diclofenac Sodium. Eur. Polym. J. 2023, 197, 112340. [Google Scholar] [CrossRef]
  78. Lin, C.-Y.; Battistoni, C.M.; Liu, J.C. Redox-Responsive Hydrogels with Decoupled Initial Stiffness and Degradation. Biomacromolecules 2021, 22, 5270–5280. [Google Scholar] [CrossRef] [PubMed]
  79. Su, R.S.; Galas, R.J., Jr.; Lin, C.; Liu, J.C. Redox-responsive Resilin-like Hydrogels for Tissue Engineering and Drug Delivery Applications. Macromol. Biosci. 2019, 19, 1900122. [Google Scholar] [CrossRef] [PubMed]
  80. Alipournazari, P.; Pourmadadi, M.; Abdouss, M.; Rahdar, A.; Pandey, S. Enhanced Delivery of Doxorubicin for Breast Cancer Treatment Using PH-Sensitive Starch/PVA/g-C3N4 Hydrogel. Int. J. Biol. Macromol. 2024, 265, 130901. [Google Scholar] [CrossRef] [PubMed]
  81. Ma, B.; Li, Q.; Mi, Y.; Zhang, J.; Tan, W.; Guo, Z. PH-Responsive Nanogels with Enhanced Antioxidant and Antitumor Activities on Drug Delivery and Smart Drug Release. Int. J. Biol. Macromol. 2024, 257, 128590. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, Z.; Cui, H.; Wang, X.; Liu, J.; Liu, G.; Meng, X.; Lin, S. Oxidized Cellulose-Filled Double Thermo/PH-Sensitive Hydrogel for Local Chemo-Photothermal Therapy in Breast Cancer. Carbohydr. Polym. 2024, 332, 121931. [Google Scholar] [CrossRef]
  83. Fukino, T.; Yamagishi, H.; Aida, T. Redox-responsive Molecular Systems and Materials. Adv. Mater. 2017, 29, 1603888. [Google Scholar] [CrossRef]
  84. Nair, D.P.; Podgorski, M.; Chatani, S.; Gong, T.; Xi, W.; Fenoli, C.R.; Bowman, C.N. The Thiol-Michael Addition Click Reaction: A Powerful and Widely Used Tool in Materials Chemistry. Chem. Mater. 2014, 26, 724–744. [Google Scholar] [CrossRef]
  85. Jiang, Y.; Chen, J.; Deng, C.; Suuronen, E.J.; Zhong, Z. Click Hydrogels, Microgels and Nanogels: Emerging Platforms for Drug Delivery and Tissue Engineering. Biomaterials 2014, 35, 4969–4985. [Google Scholar] [CrossRef]
  86. Poole, L.B. The Basics of Thiols and Cysteines in Redox Biology and Chemistry. Free Radic. Biol. Med. 2015, 80, 148–157. [Google Scholar] [CrossRef]
  87. Asai, D.; Xu, D.; Liu, W.; Quiroz, F.G.; Callahan, D.J.; Zalutsky, M.R.; Craig, S.L.; Chilkoti, A. Protein Polymer Hydrogels by in Situ, Rapid and Reversible Self-Gelation. Biomaterials 2012, 33, 5451–5458. [Google Scholar] [CrossRef]
  88. Yang, F.; Wang, J.; Cao, L.; Chen, R.; Tang, L.; Liu, C. Injectable and Redox-Responsive Hydrogel with Adaptive Degradation Rate for Bone Regeneration. J. Mater. Chem. B 2014, 2, 295–304. [Google Scholar] [CrossRef] [PubMed]
  89. Chen, L.; Annis, I.; Barany, G. Disulfide Bond Formation in Peptides. Curr. Protoc. Protein Sci. 2001, 23, 16–18. [Google Scholar] [CrossRef]
  90. Lee, M.H.; Sessler, J.L.; Kim, J.S. Disulfide-Based Multifunctional Conjugates for Targeted Theranostic Drug Delivery. Acc. Chem. Res. 2015, 48, 2935–2946. [Google Scholar] [CrossRef]
  91. Iqbal, H.M.N.; Keshavarz, T. Bioinspired Polymeric Carriers for Drug Delivery Applications. In Stimuli Responsive Polymeric Nanocarriers for Drug Delivery Applications, Volume 1; Elsevier: Amsterdam, The Netherlands, 2018; pp. 377–404. [Google Scholar]
  92. Zhai, S.; Hu, X.; Hu, Y.; Wu, B.; Xing, D. Visible Light-Induced Crosslinking and Physiological Stabilization of Diselenide-Rich Nanoparticles for Redox-Responsive Drug Release and Combination Chemotherapy. Biomaterials 2017, 121, 41–54. [Google Scholar] [CrossRef] [PubMed]
  93. Tian, Y.; Lei, M.; Yan, L.; An, F. Diselenide-Crosslinked Zwitterionic Nanogels with Dual Redox-Labile Properties for Controlled Drug Release. Polym. Chem. 2020, 11, 2360–2369. [Google Scholar] [CrossRef]
  94. Wang, J.; Liu, J.; Lu, D.-Q.; Chen, L.; Yang, R.; Liu, D.; Zhang, B. Diselenide-Crosslinked Carboxymethyl Chitosan Nanoparticles for Doxorubicin Delivery: Preparation and in Vivo Evaluation. Carbohydr. Polym. 2022, 292, 119699. [Google Scholar] [CrossRef]
  95. Huang, D.; Zhu, J.; Xu, M.-F.; Chen, J.; Gao, X.; Zhao, L.; Ding, F.; Wu, C.-Z. One-Pot Preparation of PH-and Redox-Responsive Polymeric Microgel as an Efficient Carrier for Improved Breast Cancer Therapy. Colloids Surf. A Physicochem. Eng. Asp. 2024, 685, 133320. [Google Scholar] [CrossRef]
  96. Tao, W.; Wu, X.; Li, J.; Wu, F.; Chen, C.; Jiang, T.; Xu, C.; Jiang, S.; Wang, J.; Xiao, B. An Injectable Selenite-Containing Hydrogel for Synergistic Tumor Therapy by Triggering ROS/RNS Generation and Disrupting NADPH Homeostasis. Chem. Eng. J. 2024, 479, 147437. [Google Scholar] [CrossRef]
  97. Robby, A.I.; Yang, J.; Jin, E.; Park, S.Y. Tumor Microenvironment-Selective Sol–Gel Mineralization of ROS-Responsive Stretchable and Conductive Hydrogel. Adv. Funct. Mater. 2024; early view. [Google Scholar] [CrossRef]
  98. Mohamed, M.A.; Fallahi, A.; El-Sokkary, A.M.A.; Salehi, S.; Akl, M.A.; Jafari, A.; Tamayol, A.; Fenniri, H.; Khademhosseini, A.; Andreadis, S.T. Stimuli-Responsive Hydrogels for Manipulation of Cell Microenvironment: From Chemistry to Biofabrication Technology. Prog. Polym. Sci. 2019, 98, 101147. [Google Scholar] [CrossRef]
  99. Shahriari, M.; Zahiri, M.; Abnous, K.; Taghdisi, S.M.; Ramezani, M.; Alibolandi, M. Enzyme Responsive Drug Delivery Systems in Cancer Treatment. J. Control. Release 2019, 308, 172–189. [Google Scholar] [CrossRef]
  100. Webber, M.J.; Anderson, D.G. Smart Approaches to Glucose-Responsive Drug Delivery. J. Drug Target. 2015, 23, 651–655. [Google Scholar] [CrossRef] [PubMed]
  101. Hou, X.-F.; Chen, Y.; Liu, Y. Enzyme-Responsive Protein/Polysaccharide Supramolecular Nanoparticles. Soft Matter 2015, 11, 2488–2493. [Google Scholar] [CrossRef]
  102. Sobczak, M. Enzyme-Responsive Hydrogels as Potential Drug Delivery Systems—State of Knowledge and Future Prospects. Int. J. Mol. Sci. 2022, 23, 4421. [Google Scholar] [CrossRef]
  103. He, W.; Reaume, M.; Hennenfent, M.; Lee, B.P.; Rajachar, R. Biomimetic Hydrogels with Spatial-and Temporal-Controlled Chemical Cues for Tissue Engineering. Biomater. Sci. 2020, 8, 3248–3269. [Google Scholar] [CrossRef] [PubMed]
  104. Sakai, S.; Kawakami, K. Synthesis and Characterization of Both Ionically and Enzymatically Cross-Linkable Alginate. Acta Biomater. 2007, 3, 495–501. [Google Scholar] [CrossRef]
  105. Sekhon, B.S. Matrix Metalloproteinases—An Overview. Res. Rep. Biol. 2010, 1, 1–20. [Google Scholar]
  106. Conlon, G.A.; Murray, G.I. Recent Advances in Understanding the Roles of Matrix Metalloproteinases in Tumour Invasion and Metastasis. J. Pathol. 2019, 247, 629–640. [Google Scholar] [CrossRef] [PubMed]
  107. Yao, Q.; Kou, L.; Tu, Y.; Zhu, L. MMP-Responsive ‘Smart’ Drug Delivery and Tumor Targeting. Trends Pharmacol. Sci. 2018, 39, 766–781. [Google Scholar] [CrossRef] [PubMed]
  108. Chauhan, S.; Solanki, R.; Jangid, A.K.; Jain, P.; Pranjali, P.; Patel, S.; Guleria, A.; Pooja, D.; Kulhari, H. Manganese Nanocarrier for Matrix Metalloproteinase 9 Responsive Delivery of Irinotecan for Colon Cancer Treatment. J. Ind. Eng. Chem. 2023, 128, 258–267. [Google Scholar] [CrossRef]
  109. Zhao, Z.; Shen, J.; Zhang, L.; Wang, L.; Xu, H.; Han, Y.; Jia, J.; Lu, Y.; Yu, R.; Liu, H. Injectable Postoperative Enzyme-Responsive Hydrogels for Reversing Temozolomide Resistance and Reducing Local Recurrence after Glioma Operation. Biomater. Sci. 2020, 8, 5306–5316. [Google Scholar] [CrossRef] [PubMed]
  110. Nazli, C.; Demirer, G.S.; Yar, Y.; Acar, H.Y.; Kizilel, S. Targeted Delivery of Doxorubicin into Tumor Cells via MMP-Sensitive PEG Hydrogel-Coated Magnetic Iron Oxide Nanoparticles (MIONPs). Colloids Surf. B Biointerfaces 2014, 122, 674–683. [Google Scholar] [CrossRef] [PubMed]
  111. Nagel, G.; Sousa-Herves, A.; Wedepohl, S.; Calderón, M. Matrix Metalloproteinase-Sensitive Multistage Nanogels Promote Drug Transport in 3D Tumor Model. Theranostics 2020, 10, 91. [Google Scholar] [CrossRef] [PubMed]
  112. Su, T.; Tang, Z.; He, H.; Li, W.; Wang, X.; Liao, C.; Sun, Y.; Wang, Q. Glucose Oxidase Triggers Gelation of N-Hydroxyimide–Heparin Conjugates to Form Enzyme-Responsive Hydrogels for Cell-Specific Drug Delivery. Chem. Sci. 2014, 5, 4204–4209. [Google Scholar] [CrossRef]
  113. Ono, K.; Hashimoto, H.; Katayama, T.; Ueda, N.; Nagahama, K. Injectable Biocatalytic Nanocomposite Hydrogel Factories for Focal Enzyme-Prodrug Cancer Therapy. Biomacromolecules 2021, 22, 4217–4227. [Google Scholar] [CrossRef] [PubMed]
  114. Hovgaard, L.; Brøndsted, H. Dextran Hydrogels for Colon-Specific Drug Delivery. J. Control. Release 1995, 36, 159–166. [Google Scholar] [CrossRef]
  115. Dong, Y.; Wang, W.; Veiseh, O.; Appel, E.A.; Xue, K.; Webber, M.J.; Tang, B.C.; Yang, X.-W.; Weir, G.C.; Langer, R. Injectable and Glucose-Responsive Hydrogels Based on Boronic Acid–Glucose Complexation. Langmuir 2016, 32, 8743–8747. [Google Scholar] [CrossRef]
  116. Li, X.; Fu, M.; Wu, J.; Zhang, C.; Deng, X.; Dhinakar, A.; Huang, W.; Qian, H.; Ge, L. PH-Sensitive Peptide Hydrogel for Glucose-Responsive Insulin Delivery. Acta Biomater. 2017, 51, 294–303. [Google Scholar] [CrossRef] [PubMed]
  117. Volpatti, L.R.; Matranga, M.A.; Cortinas, A.B.; Delcassian, D.; Daniel, K.B.; Langer, R.; Anderson, D.G. Glucose-Responsive Nanoparticles for Rapid and Extended Self-Regulated Insulin Delivery. ACS Nano 2019, 14, 488–497. [Google Scholar] [CrossRef] [PubMed]
  118. Wang, J.; Wang, Z.; Yu, J.; Kahkoska, A.R.; Buse, J.B.; Gu, Z. Glucose-responsive Insulin and Delivery Systems: Innovation and Translation. Adv. Mater. 2020, 32, 1902004. [Google Scholar] [CrossRef] [PubMed]
  119. Tanna, S.; Sahota, T.S.; Sawicka, K.; Taylor, M.J. The Effect of Degree of Acrylic Derivatisation on Dextran and Concanavalin A Glucose-Responsive Materials for Closed-Loop Insulin Delivery. Biomaterials 2006, 27, 4498–4507. [Google Scholar] [CrossRef] [PubMed]
  120. Matsumoto, A.; Ikeda, S.; Harada, A.; Kataoka, K. Glucose-Responsive Polymer Bearing a Novel Phenylborate Derivative as a Glucose-Sensing Moiety Operating at Physiological PH Conditions. Biomacromolecules 2003, 4, 1410–1416. [Google Scholar] [CrossRef] [PubMed]
  121. Liang, Y.; Li, M.; Yang, Y.; Qiao, L.; Xu, H.; Guo, B. PH/Glucose Dual Responsive Metformin Release Hydrogel Dressings with Adhesion and Self-Healing via Dual-Dynamic Bonding for Athletic Diabetic Foot Wound Healing. ACS Nano 2022, 16, 3194–3207. [Google Scholar] [CrossRef] [PubMed]
  122. Luo, J.; Cao, S.; Chen, X.; Liu, S.; Tan, H.; Wu, W.; Li, J. Super Long-Term Glycemic Control in Diabetic Rats by Glucose-Sensitive LbL Films Constructed of Supramolecular Insulin Assembly. Biomaterials 2012, 33, 8733–8742. [Google Scholar] [CrossRef]
  123. Gu, Z.; Aimetti, A.A.; Wang, Q.; Dang, T.T.; Zhang, Y.; Veish, O.; Cheng, H.; Langer, R.S.; Anderson, D.G. Injectable Nano-Network for Glucose-Mediated Insulin Delivery. ACS Nano 2013, 7, 4194–4201. [Google Scholar] [CrossRef] [PubMed]
  124. de Almeida, M.S.; Susnik, E.; Drasler, B.; Taladriz-Blanco, P.; Petri-Fink, A.; Rothen-Rutishauser, B. Understanding Nanoparticle Endocytosis to Improve Targeting Strategies in Nanomedicine. Chem. Soc. Rev. 2021, 50, 5397–5434. [Google Scholar] [CrossRef]
  125. Rahim, M.A.; Jan, N.; Khan, S.; Shah, H.; Madni, A.; Khan, A.; Jabar, A.; Khan, S.; Elhissi, A.; Hussain, Z. Recent Advancements in Stimuli Responsive Drug Delivery Platforms for Active and Passive Cancer Targeting. Cancers 2021, 13, 670. [Google Scholar] [CrossRef]
  126. Wang, N.; Cheng, X.; Li, N.; Wang, H.; Chen, H. Nanocarriers and Their Loading Strategies. Adv. Heal. Mater. 2019, 8, 1801002. [Google Scholar] [CrossRef] [PubMed]
  127. Ding, C.; Li, Z. A Review of Drug Release Mechanisms from Nanocarrier Systems. Mater. Sci. Eng. C 2017, 76, 1440–1453. [Google Scholar] [CrossRef]
  128. Huang, J.; Zhang, X.; Fu, K.; Wei, G.; Su, Z. Stimulus-Responsive Nanomaterials under Physical Regulation for Biomedical Applications. J. Mater. Chem. B 2021, 9, 9642–9657. [Google Scholar] [CrossRef] [PubMed]
  129. Ahsan, A.; Farooq, M.A.; Parveen, A. Thermosensitive Chitosan-Based Injectable Hydrogel as an Efficient Anticancer Drug Carrier. ACS Omega 2020, 5, 20450–20460. [Google Scholar] [CrossRef]
  130. Yang, F.; Shi, K.; Hao, Y.; Jia, Y.; Liu, Q.; Chen, Y.; Pan, M.; Yuan, L.; Yu, Y.; Qian, Z. Cyclophosphamide Loaded Thermo-Responsive Hydrogel System Synergize with a Hydrogel Cancer Vaccine to Amplify Cancer Immunotherapy in a Prime-Boost Manner. Bioact. Mater. 2021, 6, 3036–3048. [Google Scholar] [CrossRef] [PubMed]
  131. Garrett, M.C.; O’Shea, T.M.; Wollenberg, A.L.; Bernstein, A.M.; Hung, D.; Staarman, B.; Soto, H.; Deming, T.J.; Sofroniew, M.V.; Kornblum, H.I. Injectable Diblock Copolypeptide Hydrogel Provides Platform to Deliver Effective Concentrations of Paclitaxel to an Intracranial Xenograft Model of Glioblastoma. PLoS ONE 2020, 15, e0219632. [Google Scholar] [CrossRef]
  132. Qian, Z.; Zhao, N.; Xu, S.; Yuan, W. In Situ Injectable Thermoresponsive Nanocomposite Hydrogel Based on Hydroxypropyl Chitosan for Precise Synergistic Calcium-Overload, Photodynamic and Photothermal Tumor Therapy. Carbohydr. Polym. 2024, 324, 121487. [Google Scholar] [CrossRef]
  133. He, J.; Zou, H.; Zhou, J.; Deng, C. Thermoresponsive MXene-Based Hydrogel for Controlled Anticancer Drug Release. J. Drug Deliv. Sci. Technol. 2024, 91, 105207. [Google Scholar] [CrossRef]
  134. Lu, Y.-J.; Lan, Y.-H.; Chuang, C.-C.; Lu, W.-T.; Chan, L.-Y.; Hsu, P.-W.; Chen, J.-P. Injectable Thermo-Sensitive Chitosan Hydrogel Containing CPT-11-Loaded EGFR-Targeted Graphene Oxide and SLP2 ShRNA for Localized Drug/Gene Delivery in Glioblastoma Therapy. Int. J. Mol. Sci. 2020, 21, 7111. [Google Scholar] [CrossRef]
  135. Darge, H.F.; Andrgie, A.T.; Hanurry, E.Y.; Birhan, Y.S.; Mekonnen, T.W.; Chou, H.-Y.; Hsu, W.-H.; Lai, J.-Y.; Lin, S.-Y.; Tsai, H.-C. Localized Controlled Release of Bevacizumab and Doxorubicin by Thermo-Sensitive Hydrogel for Normalization of Tumor Vasculature and to Enhance the Efficacy of Chemotherapy. Int. J. Pharm. 2019, 572, 118799. [Google Scholar] [CrossRef]
  136. Liu, M.; Song, X.; Wen, Y.; Zhu, J.-L.; Li, J. Injectable Thermoresponsive Hydrogel Formed by Alginate-g-Poly (N-Isopropylacrylamide) That Releases Doxorubicin-Encapsulated Micelles as a Smart Drug Delivery System. ACS Appl. Mater. Interfaces 2017, 9, 35673–35682. [Google Scholar] [CrossRef]
  137. Chen, X.; Liu, Z.; Parker, S.G.; Zhang, X.; Gooding, J.J.; Ru, Y.; Liu, Y.; Zhou, Y. Light-Induced Hydrogel Based on Tumor-Targeting Mesoporous Silica Nanoparticles as a Theranostic Platform for Sustained Cancer Treatment. ACS Appl. Mater. Interfaces 2016, 8, 15857–15863. [Google Scholar] [CrossRef] [PubMed]
  138. Chen, B.; Wang, G.; Huang, C.; Sun, Y.; Zhang, J.; Chai, Z.; Guo, S.-S.; Zhao, X.-Z.; Yuan, Y.; Liu, W. A Light-Induced Hydrogel Responsive Platform to Capture and Selectively Isolate Single Circulating Tumor Cells. Nanoscale 2022, 14, 3504–3512. [Google Scholar] [CrossRef] [PubMed]
  139. Xing, C.; Chen, S.; Liang, X.; Liu, Q.; Qu, M.; Zou, Q.; Li, J.; Tan, H.; Liu, L.; Fan, D. Two-Dimensional MXene (Ti3C2)-Integrated Cellulose Hydrogels: Toward Smart Three-Dimensional Network Nanoplatforms Exhibiting Light-Induced Swelling and Bimodal Photothermal/Chemotherapy Anticancer Activity. ACS Appl. Mater. Interfaces 2018, 10, 27631–27643. [Google Scholar] [CrossRef] [PubMed]
  140. Hou, M.; Yang, R.; Zhang, L.; Zhang, L.; Liu, G.; Xu, Z.; Kang, Y.; Xue, P. Injectable and Natural Humic Acid/Agarose Hybrid Hydrogel for Localized Light-Driven Photothermal Ablation and Chemotherapy of Cancer. ACS Biomater. Sci. Eng. 2018, 4, 4266–4277. [Google Scholar] [CrossRef]
  141. Wu, Y.; Chen, F.; Huang, N.; Li, J.; Wu, C.; Tan, B.; Liu, Y.; Li, L.; Yang, C.; Shao, D. Near-Infrared Light-Responsive Hybrid Hydrogels for the Synergistic Chemo-Photothermal Therapy of Oral Cancer. Nanoscale 2021, 13, 17168–17182. [Google Scholar] [CrossRef]
  142. Chen, X.; Wang, M.; Yang, X.; Wang, Y.; Yu, L.; Sun, J.; Ding, J. Injectable Hydrogels for the Sustained Delivery of a HER2-Targeted Antibody for Preventing Local Relapse of HER2+ Breast Cancer after Breast-Conserving Surgery. Theranostics 2019, 9, 6080. [Google Scholar] [CrossRef]
  143. Jeong, B.; Kim, S.W.; Bae, Y.H. Thermosensitive Sol–Gel Reversible Hydrogels. Adv. Drug Deliv. Rev. 2012, 64, 154–162. [Google Scholar] [CrossRef]
  144. Bellotti, E.; Schilling, A.L.; Little, S.R.; Decuzzi, P. Injectable Thermoresponsive Hydrogels as Drug Delivery System for the Treatment of Central Nervous System Disorders: A Review. J. Control. Release 2021, 329, 16–35. [Google Scholar] [CrossRef]
  145. Cook, M.T.; Haddow, P.; Kirton, S.B.; McAuley, W.J. Polymers Exhibiting Lower Critical Solution Temperatures as a Route to Thermoreversible Gelators for Healthcare. Adv. Funct. Mater. 2021, 31, 2008123. [Google Scholar] [CrossRef]
  146. Lanzalaco, S.; Armelin, E. Poly (N-Isopropylacrylamide) and Copolymers: A Review on Recent Progresses in Biomedical Applications. Gels 2017, 3, 36. [Google Scholar] [CrossRef]
  147. Xiao, Y.; Gu, Y.; Qin, L.; Chen, L.; Chen, X.; Cui, W.; Li, F.; Xiang, N.; He, X. Injectable Thermosensitive Hydrogel-Based Drug Delivery System for Local Cancer Therapy. Colloids Surf. B Biointerfaces 2021, 200, 111581. [Google Scholar] [CrossRef] [PubMed]
  148. Dasgupta, A.; Mondal, J.H.; Das, D. Peptide Hydrogels. RSC Adv. 2013, 3, 9117–9149. [Google Scholar] [CrossRef]
  149. Nagahama, K.; Takahashi, A.; Ohya, Y. Biodegradable Polymers Exhibiting Temperature-Responsive Sol–Gel Transition as Injectable Biomedical Materials. React. Funct. Polym. 2013, 73, 979–985. [Google Scholar] [CrossRef]
  150. Liow, S.S.; Dou, Q.; Kai, D.; Karim, A.A.; Zhang, K.; Xu, F.; Loh, X.J. Thermogels: In Situ Gelling Biomaterial. ACS Biomater. Sci. Eng. 2016, 2, 295–316. [Google Scholar] [CrossRef] [PubMed]
  151. Di Martino, M.; Sessa, L.; Diana, R.; Piotto, S.; Concilio, S. Recent Progress in Photoresponsive Biomaterials. Molecules 2023, 28, 3712. [Google Scholar] [CrossRef]
  152. Li, L.; Scheiger, J.M.; Levkin, P.A. Design and Applications of Photoresponsive Hydrogels. Adv. Mater. 2019, 31, 1807333. [Google Scholar] [CrossRef] [PubMed]
  153. Peng, K.; Zheng, L.; Zhou, T.; Zhang, C.; Li, H. Light Manipulation for Fabrication of Hydrogels and Their Biological Applications. Acta Biomater. 2022, 137, 20–43. [Google Scholar] [CrossRef] [PubMed]
  154. Guimarães, C.F.; Ahmed, R.; Marques, A.P.; Reis, R.L.; Demirci, U. Engineering Hydrogel-Based Biomedical Photonics: Design, Fabrication, and Applications. Adv. Mater. 2021, 33, 2006582. [Google Scholar] [CrossRef]
  155. Tong, R.; Hemmati, H.D.; Langer, R.; Kohane, D.S. Photoswitchable Nanoparticles for Triggered Tissue Penetration and Drug Delivery. J. Am. Chem. Soc. 2012, 134, 8848–8855. [Google Scholar] [CrossRef]
  156. Jiang, Z.; Li, H.; You, Y.; Wu, X.; Shao, S.; Gu, Q. Controlled Protein Delivery from Photosensitive Nanoparticles. J. Biomed. Mater. Res. A 2015, 103, 65–70. [Google Scholar] [CrossRef] [PubMed]
  157. Pianowski, Z.L.; Karcher, J.; Schneider, K. Photoresponsive Self-Healing Supramolecular Hydrogels for Light-Induced Release of DNA and Doxorubicin. Chem. Commun. 2016, 52, 3143–3146. [Google Scholar] [CrossRef] [PubMed]
  158. Jiang, J.; Qi, B.; Lepage, M.; Zhao, Y. Polymer Micelles Stabilization on Demand through Reversible Photo-Cross-Linking. Macromolecules 2007, 40, 790–792. [Google Scholar] [CrossRef]
  159. Xing, Y.; Zeng, B.; Yang, W. Light Responsive Hydrogels for Controlled Drug Delivery. Front. Bioeng. Biotechnol. 2022, 10, 1075670. [Google Scholar] [CrossRef] [PubMed]
  160. Pourjavadi, A.; Heydarpour, R.; Tehrani, Z.M. Multi-Stimuli-Responsive Hydrogels and Their Medical Applications. New J. Chem. 2021, 45, 15705–15717. [Google Scholar] [CrossRef]
  161. 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]
  162. El-Husseiny, H.M.; Mady, E.A.; Hamabe, L.; Abugomaa, A.; Shimada, K.; Yoshida, T.; Tanaka, T.; Yokoi, A.; Elbadawy, M.; Tanaka, R. Smart/Stimuli-Responsive Hydrogels: Cutting-Edge Platforms for Tissue Engineering and Other Biomedical Applications. Mater. Today Bio 2022, 13, 100186. [Google Scholar] [CrossRef] [PubMed]
  163. Thang, N.H.; Chien, T.B.; Cuong, D.X. Polymer-Based Hydrogels Applied in Drug Delivery: An Overview. Gels 2023, 9, 523. [Google Scholar] [CrossRef] [PubMed]
  164. Güngör, A.; Özdemir, T.; Genç, R. Investigation of Use in 5-FU Release: Synthesis of Temperature and PH Responsive P (NVCL-Co-VIm)/PVP Hydrogels. Polym. Bull. 2024, 81, 2091–2109. [Google Scholar] [CrossRef]
  165. Fathi, M.; Alami-Milani, M.; Geranmayeh, M.H.; Barar, J.; Erfan-Niya, H.; Omidi, Y. Dual Thermo-and PH-Sensitive Injectable Hydrogels of Chitosan/(Poly(N-Isopropylacrylamide-Co-Itaconic Acid)) for Doxorubicin Delivery in Breast Cancer. Int. J. Biol. Macromol. 2019, 128, 957–964. [Google Scholar] [CrossRef]
  166. Zhao, C.; Zhuang, X.; He, P.; Xiao, C.; He, C.; Sun, J.; Chen, X.; Jing, X. Synthesis of Biodegradable Thermo-and PH-Responsive Hydrogels for Controlled Drug Release. Polymer 2009, 50, 4308–4316. [Google Scholar] [CrossRef]
  167. Wu, Y.; Wang, Y.; Long, L.; Hu, C.; Kong, Q.; Wang, Y. A Spatiotemporal Release Platform Based on PH/ROS Stimuli-Responsive Hydrogel in Wound Repairing. J. Control. Release 2022, 341, 147–165. [Google Scholar] [CrossRef] [PubMed]
  168. Jung, J.M.; Kim, S.H.; Phan, V.H.G.; Thambi, T.; Lee, D.S. Therapeutic Effects of Boronate Ester Cross-Linked Injectable Hydrogels for the Treatment of Hepatocellular Carcinoma. Biomater. Sci. 2021, 9, 7275–7286. [Google Scholar] [CrossRef]
  169. Ma, Z.; Ma, R.; Wang, X.; Gao, J.; Zheng, Y.; Sun, Z. Enzyme and PH Responsive 5-Flurouracil (5-FU) Loaded Hydrogels Based on Olsalazine Derivatives for Colon-Specific Drug Delivery. Eur. Polym. J. 2019, 118, 64–70. [Google Scholar] [CrossRef]
  170. Palmese, L.L.; Fan, M.; Scott, R.A.; Tan, H.; Kiick, K.L. Multi-Stimuli-Responsive, Liposome-Crosslinked Poly(Ethylene Glycol) Hydrogels for Drug Delivery. J. Biomater. Sci. Polym. Ed. 2020, 32, 635–656. [Google Scholar] [CrossRef]
  171. Li, X.; Su, X. Multifunctional Smart Hydrogels: Potential in Tissue Engineering and Cancer Therapy. J. Mater. Chem. B 2018, 6, 4714–4730. [Google Scholar] [CrossRef] [PubMed]
  172. Zhang, Y.; Li, J.; Pu, K. Recent Advances in Dual-and Multi-Responsive Nanomedicines for Precision Cancer Therapy. Biomaterials 2022, 291, 121906. [Google Scholar] [CrossRef] [PubMed]
  173. Jo, Y.-J.; Gulfam, M.; Jo, S.-H.; Gal, Y.-S.; Oh, C.-W.; Park, S.-H.; Lim, K.T. Multi-Stimuli Responsive Hydrogels Derived from Hyaluronic Acid for Cancer Therapy Application. Carbohydr. Polym. 2022, 286, 119303. [Google Scholar] [CrossRef] [PubMed]
  174. Gou, S.; Xie, D.; Ma, Y.; Huang, Y.; Dai, F.; Wang, C.; Xiao, B. Injectable, Thixotropic, and Multiresponsive Silk Fibroin Hydrogel for Localized and Synergistic Tumor Therapy. ACS Biomater. Sci. Eng. 2019, 6, 1052–1063. [Google Scholar] [CrossRef]
  175. Buwalda, S.J.; Vermonden, T.; Hennink, W.E. Hydrogels for Therapeutic Delivery: Current Developments and Future Directions. Biomacromolecules 2017, 18, 316–330. [Google Scholar] [CrossRef]
  176. Zhang, X.; Achazi, K.; Steinhilber, D.; Kratz, F.; Dernedde, J.; Haag, R. A Facile Approach for Dual-Responsive Prodrug Nanogels Based on Dendritic Polyglycerols with Minimal Leaching. J. Control. Release 2014, 174, 209–216. [Google Scholar] [CrossRef] [PubMed]
  177. Chen, H.; Xu, J.; Sun, J.; Jiang, Y.; Zheng, W.; Hu, W.; Qian, H. Recent Advances on Thermosensitive Hydrogels-Mediated Precision Therapy. Asian J. Pharm. Sci. 2024, 19, 100911. [Google Scholar] [CrossRef]
  178. Cook, A.B.; Decuzzi, P. Harnessing Endogenous Stimuli for Responsive Materials in Theranostics. ACS Nano 2021, 15, 2068–2098. [Google Scholar] [CrossRef] [PubMed]
  179. Yin, Z.-C.; Wang, Y.-L.; Wang, K. A PH-Responsive Composite Hydrogel Beads Based on Agar and Alginate for Oral Drug Delivery. J. Drug Deliv. Sci. Technol. 2018, 43, 12–18. [Google Scholar] [CrossRef]
  180. Tan, J.; Luo, Y.; Guo, Y.; Zhou, Y.; Liao, X.; Li, D.; Lai, X.; Liu, Y. Development of Alginate-Based Hydrogels: Crosslinking Strategies and Biomedical Applications. Int. J. Biol. Macromol. 2023, 239, 124275. [Google Scholar] [CrossRef] [PubMed]
  181. Salehi, S.; Naghib, S.M.; Garshasbi, H.R.; Ghorbanzadeh, S.; Zhang, W. Smart Stimuli-Responsive Injectable Gels and Hydrogels for Drug Delivery and Tissue Engineering Applications: A Review. Front. Bioeng. Biotechnol. 2023, 11, 1104126. [Google Scholar] [CrossRef] [PubMed]
  182. Esther, M.; Solanki, R.; Dhanka, M.; Thareja, P.; Bhatia, D. Self-Healing, Injectable Chitosan-Based Hydrogels: Structure, Properties and Biological Applications. Mater. Adv. 2024; advance article. [Google Scholar] [CrossRef]
  183. Gu, D.; O’Connor, A.J.; GH Qiao, G.; Ladewig, K. Hydrogels with Smart Systems for Delivery of Hydrophobic Drugs. Expert. Opin. Drug Deliv. 2017, 14, 879–895. [Google Scholar] [CrossRef]
  184. Ghasemiyeh, P.; Mohammadi-Samani, S. Hydrogels as Drug Delivery Systems; Pros and Cons. Trends Pharm. Sci. 2019, 5, 7–24. [Google Scholar]
  185. Larrañeta, E.; Stewart, S.; Ervine, M.; Al-Kasasbeh, R.; Donnelly, R.F. Hydrogels for Hydrophobic Drug Delivery. Classification, Synthesis and Applications. J. Funct. Biomater. 2018, 9, 13. [Google Scholar] [CrossRef]
  186. Li, J.; Mooney, D.J. Designing Hydrogels for Controlled Drug Delivery. Nat. Rev. Mater. 2016, 1, 16071. [Google Scholar] [CrossRef]
  187. Popescu, M.-T.; Mourtas, S.; Pampalakis, G.; Antimisiaris, S.G.; Tsitsilianis, C. PH-Responsive Hydrogel/Liposome Soft Nanocomposites for Tuning Drug Release. Biomacromolecules 2011, 12, 3023–3030. [Google Scholar] [CrossRef] [PubMed]
  188. Tao, H.; Parthiban, A. Stimuli-Responsive Copolymers and Their Applications. In Synthesis and Applications of Copolymers; John Wiley & Sons: Hoboken, NJ, USA, 2014; pp. 274–306. [Google Scholar]
  189. Wu, J.; Xue, W.; Yun, Z.; Liu, Q.; Sun, X. Biomedical Applications of Stimuli-Responsive “Smart” Interpenetrating Polymer Network Hydrogels. Mater. Today Bio 2024, 25, 100998. [Google Scholar] [CrossRef] [PubMed]
  190. Anju, S.; Prajitha, N.; Sukanya, V.S.; Mohanan, P.V. Complicity of Degradable Polymers in Health-Care Applications. Mater. Today Chem. 2020, 16, 100236. [Google Scholar] [CrossRef]
Figure 1. Stimulus-responsive hydrogels: synthesis materials, route for administration, and applications.
Figure 1. Stimulus-responsive hydrogels: synthesis materials, route for administration, and applications.
Gels 10 00440 g001
Figure 2. Schematic representation of tumor microenvironment (TME) and conditions prevailing in it, which are utilized as stimuli for targeted drug delivery to tumors.
Figure 2. Schematic representation of tumor microenvironment (TME) and conditions prevailing in it, which are utilized as stimuli for targeted drug delivery to tumors.
Gels 10 00440 g002
Figure 3. Classification and common synthesis methods for preparation of hydrogels.
Figure 3. Classification and common synthesis methods for preparation of hydrogels.
Gels 10 00440 g003
Figure 4. Some important developments in hydrogels as drug delivery systems.
Figure 4. Some important developments in hydrogels as drug delivery systems.
Gels 10 00440 g004
Figure 5. Illustration of hydrogel preparation (Aa). Illustration of the state of gel precursors and the hydrogel (Ab). Demonstration of the pH-induced phase transitions of the hydrogel: (Ba) gel state, (Bb) sol state when concentrated HCl was added, (Bc) recovery of the gel state when NaOH solution was added, (Bd) dissolution of the regenerated gel when concentrated HCl was included again, and (Be) the gel after five cycles of regeneration. Studies conducted in vivo following administration of free DOX and Hydrogel-DOX; time-dependent volume change curves of tumor-bearing animals (Ca). Tumor-bearing mice’s body weight change curves over time (Cb). Typical pictures of tumors removed from mice (Cc). Figures are reprinted from [51] with permission (Copyright © 2015, American Chemical Society).
Figure 5. Illustration of hydrogel preparation (Aa). Illustration of the state of gel precursors and the hydrogel (Ab). Demonstration of the pH-induced phase transitions of the hydrogel: (Ba) gel state, (Bb) sol state when concentrated HCl was added, (Bc) recovery of the gel state when NaOH solution was added, (Bd) dissolution of the regenerated gel when concentrated HCl was included again, and (Be) the gel after five cycles of regeneration. Studies conducted in vivo following administration of free DOX and Hydrogel-DOX; time-dependent volume change curves of tumor-bearing animals (Ca). Tumor-bearing mice’s body weight change curves over time (Cb). Typical pictures of tumors removed from mice (Cc). Figures are reprinted from [51] with permission (Copyright © 2015, American Chemical Society).
Gels 10 00440 g005
Figure 6. In vivo antitumor efficacy of Herceptin-loaded hydrogel. Illustration depicting the development of a HER2+ breast tumor relapse nude mice model (A). Tumor volume changes (B) and ex vivo tumor weight (C) on day 28 after treatment. Digital images of tumors (D) and body weights of mice following the various treatments (E). (* p < 0.05, ** p < 0.01 and *** p < 0.001). Figures are reprinted from [142] with permission (Ivyspring International Publisher).
Figure 6. In vivo antitumor efficacy of Herceptin-loaded hydrogel. Illustration depicting the development of a HER2+ breast tumor relapse nude mice model (A). Tumor volume changes (B) and ex vivo tumor weight (C) on day 28 after treatment. Digital images of tumors (D) and body weights of mice following the various treatments (E). (* p < 0.05, ** p < 0.01 and *** p < 0.001). Figures are reprinted from [142] with permission (Ivyspring International Publisher).
Gels 10 00440 g006
Figure 7. Schematic diagram of the mechanism of action of the prepared SH- and DOX-loaded light-responsive hydrogel (A) and in vivo anticancer activity (B). Thermographic imaging (Ba), matching temperature variation of mice under laser irradiation (Bb), change in mouse body weight over time (Bc), images of the excised tumors on the 14th day after different treatments (Bd), change in normalized tumor volume over time (Be), average weight of excised tumors on the 14th day after different treatments (Bf), histopathological analysis of excised tumors and H&E staining (Bg) and TUNEL staining on the 14th day after various treatments (Bh) (scale bars: 200 μm). Figures are reprinted from [140] with permission (Copyright © 2018 American Chemical Society).
Figure 7. Schematic diagram of the mechanism of action of the prepared SH- and DOX-loaded light-responsive hydrogel (A) and in vivo anticancer activity (B). Thermographic imaging (Ba), matching temperature variation of mice under laser irradiation (Bb), change in mouse body weight over time (Bc), images of the excised tumors on the 14th day after different treatments (Bd), change in normalized tumor volume over time (Be), average weight of excised tumors on the 14th day after different treatments (Bf), histopathological analysis of excised tumors and H&E staining (Bg) and TUNEL staining on the 14th day after various treatments (Bh) (scale bars: 200 μm). Figures are reprinted from [140] with permission (Copyright © 2018 American Chemical Society).
Gels 10 00440 g007
Figure 8. Schematic representation of multi-responsive silk fibroin-based hydrogel synthesis, action of mechanism, and in vivo antitumor efficacy. Figures reprinted from [174] with permission (Copyright © 2020, American Chemical Society).
Figure 8. Schematic representation of multi-responsive silk fibroin-based hydrogel synthesis, action of mechanism, and in vivo antitumor efficacy. Figures reprinted from [174] with permission (Copyright © 2020, American Chemical Society).
Gels 10 00440 g008
Figure 9. Uptake and release of DOX from nanogel (A). Synthesis mechanism of DOX nanogel (B). Cellular (intracellular/subcellular) uptake of DOX nanogel within HeLa cells (Ca,Cb). Figures reprinted from [176] with permission (Copyright © 2013 Elsevier B.V.).
Figure 9. Uptake and release of DOX from nanogel (A). Synthesis mechanism of DOX nanogel (B). Cellular (intracellular/subcellular) uptake of DOX nanogel within HeLa cells (Ca,Cb). Figures reprinted from [176] with permission (Copyright © 2013 Elsevier B.V.).
Gels 10 00440 g009
Table 1. List of chemical stimuli-responsive hydrogels.
Table 1. List of chemical stimuli-responsive hydrogels.
StimulusPolymer(s)Drug/Dye/LigandCross-Linking AgentPreparation MethodCancer/Model/RouteFindingsReference
pHCS and PEOAmoxicillin, MetronidazoleGlyoxalCross-linkingPeptic ulcersIn the acidic environment of the stomach fluid, prepared hydrogels may be helpful for the localized administration of antibiotics.[52]
PAA: PEOSAM, NAM, CHC, PDNTDICCross-linkingGI tractThe pH-dependent swelling of IPN granules in the matrix significantly determines drug release across all studied types.[57]
Gelatin: PEORiboflavinGlyoxalCross-linkingFor oral deliveryGelatin and gelatin-PEO hydrogels swell based on pH and high--eight PEO.[60]
PEG and L-LactideDOX and TETMA–PLLA–PEG–PLLA–MACross-linkingGI tractpH-mediated drug release observed (slow release of DOX in acidic buffers as well as fast release of TET).[75]
PEG-6000 and MAA-MBARadical polymerization reactionMale albino rabbitsThe nanogels were well-tolerated with no toxic effects in animals.[63]
Gelatin and Pluronic F127 CURFCHOSchiff base cross-linkingDiabetes mellitusDisplayed antibacterial and antioxidative activity and biocompatibility, facilitated wound closure, and enhanced tissue regeneration.[76]
Lap®/CS/PVACURLap®CURCross-linkingBreast cancer cells (MDA-MB 231) and bacteria S. aureus, E. coli, and H. pyloriHad good blood compatibility, excellent antioxidant properties, and antibacterial activity.[69]
PEI-Co-MAAMesalazineMBAFree radical polymerizationColorectal diseasesHydrophilic drugs can be delivered to colon sites via hydrogels.[70]
Black seed extract and β-CD, MAAPerindopril ErbumineMBAFree radical polymerization-At alkaline pHs, hydrogels demonstrated more swelling and in vitro drug release compared to acidic pHs, with no adverse effects observed in animals.[71]
CMTKG/PVP/PAMDSMBAFree radical polymerization-A higher drug release was observed at physiological pH (pH 7.4) compared to acidic pH (pH 1.2) from hydrogel.[77]
PU−PEI and PU−CACiprofloxacin,
Bromophenol blue and Pyronin Y
-Aminolysis and Stiglich esterification mechanism, physical cross-linkingChronic woundsPU–PEI films exhibited significantly higher antibacterial activity than PU–CA films, and they discharged more cargo at an acidic pH than PU–CA films did at an alkaline pH.[74]
CMA and CSDS-Ionic complexationDermal drug delivery, HaCaT cellsThe viability of HaCaT cells was nearly 100% in the presence of hydrogels and DS, indicating the potential of CMA/CS PECs for pH-responsive dermal drug delivery.[50]
DF-PEG, PAHyDOX-Chemical cross-linkingHuman fibrosarcomaThe DOX-loaded hydrogel exhibited enhanced efficacy, achieving approximately 80% tumor inhibition by day 20, suggesting its potential as a highly effective treatment for human fibrosarcoma.[51]
RedoxDPA-DSDMA, PEGPDS, BSA--L929 mouse fibroblastsWhen hydrogels were treated with thiol-containing reducing agents, they broke down quickly, facilitating the release of the encapsulated payload (such as BSA) more quickly.[33]
PEG-SH and Fe-EDTADextranDVSOne-pot cross-linkingNIH/3T3 mouse fibroblastsThese gels offer a potentially useful platform for separating the behavior of degradation in response to reduction stimuli from the initial mechanical properties.[78]
ResilinRGDDTSSP-NIH/3T3 fibroblastsDemonstrated the degradation and cytocompatibility of DTSSP-cross-linked RZ10-RGD, showcasing their potential for biomedical applications.[79]
Abbreviations—CS: chitosan; PEO: poly (ethylene oxide); SAM: salicylamide; NAM: nicotinamide; CHC: conidine·HCl; PDN: prednisolone; PAA: poly(acrylic acid); TDIC: tolylene-2;4-diisocyanate; PLLA-PEG-PLLA: poly(L-lactide)-co-polyethyleneglycol-co-poly(L-lactide); MA: methacrylic anhydride; DOX: doxorubicin; TET: tetracycline; GI tract: gastrointestinal tract; MAA: methacrylic acid; MBA: N-methylene bis(acrylamide); CUR: curcumin; FCHO: benzaldehyde-modified Pluronic F127 polymer; Lap®/CS/PVA: laponite RD/chitosan/polyvinyl alcohol; PEI: polyethyleneimine; β-CD: β-cyclodextrin; CMTKG/ PVP/PAM: carboxymethyl tamarind kernel gum/polyvinylpyrrolidone/polyacrylamide; DS: diclofenac sodium; PU−PEI: polyurethane–polyethylenimine; PU−CA: polyurethane–carboxylic-acid-modified films; CMA: carboxymethylagarose; DF-PEG: dibenzaldehyde-terminated PEG; PAHy: polyaspartylhydrazide; PEG: poly(ethylene glycol); PDS: pyridyl disulfide; DPA: 2-(Diisopropylamino) ethyl methacrylate, DSDMA: Bis(2-methacryloyloxyethyl) Disulfide, BSA: bovine serum albumin; DVS: divinyl sulfone; Fe-EDTA: ferric ethylenediaminetetraacetic acid; DTSSP: 3,3′-dithiobis(sulfosuccinimidyl propionate).
Table 2. List of biological stimuli-responsive hydrogels.
Table 2. List of biological stimuli-responsive hydrogels.
StimulusPolymer(s)Drug/Dye/LigandCross-Linking AgentPreparation MethodCancer/Model/RouteFindingsReference
Enzyme (HRP) Alginate and tyraminePhenolHRP/H2O2Enzymatic cross-linking-Findings demonstrated the viability of a unique alginate synthesized with phenols as an alternate to typical unmodified alginates.[104]
Enzyme (GOx)N-hydroxyimide–heparin DOXEDC/NHS, GoxRadical polymerization reactionHeLa, HepG2, and NIH-3T3 cellsDrug was released from hydrogel in an enzyme-responsive manner.[112]
Enzyme (MMP)TmTMZ, BG--C6 glioma cells, BALB/c nude male mice, orthotropic glioma modelHydrogels reduced MGMT expression in vivo, rendering TMZ-resistant glioma cells more responsive to TMZ treatment. Additionally, post-surgery, these hydrogels significantly enhanced TMZ efficacy in glioma growth inhibition and reduced the recurrence of TMZ-resistant gliomas.[109]
Enzyme (β-gal)PLGA−
PEG−
PLGA
5-FU,
LAPONITE
2-ethyl-hexanoate as a catalystBulk ring-opening co-polymerizationMCF-7, female nude mice (ICR-nu/nu), PC-3 cellsProdrug 5-FU−β-gal and nanocomposite gels were injected locally once, and the combination had long-lasting anticancer activity in vivo with no side effects.[113]
Enzyme (MMP)dPGDOXPeptideNano-precipitationHeLa cells, MCTSs, primary fibroblastsThe digested multistage pNGs demonstrated enhanced diffusive transport through a dense gel matrix. pNGs facilitated the infiltration of functional chemotherapeutic medication into deeper tissue regions in tumor-like MCTSs.[111]
Enzyme (MMP)PEGDOX, MIONPs, RGDS-Surface-initiated photopolymerizationHeLa cells, mouse fibroblastTargeted nanocarriers internalized and efficiently carried and released DOX into the nucleus of HeLa cells within 2 h.[110]
Enzyme (dextranases)Dextran-HDI or DDI-SD rats, human colonic fermentation modelDextran hydrogels were degraded in vitro by a model dextranase, as well as in vivo in rats and a human colonic fermentation model.[114]
GlucosePBA and glucose PBAAIBN as an initiatorRadical polymerizationInsulin deliveryMechanism was not studied but could be used for insulin delivery[115]
GlucoseIA-0 peptideGox, Catalase, Insulin-Solid-phase methodSTZ-induced diabetic miceIn vitro and in vivo studies demonstrated that the developed hydrogels can regulate blood glucose levels.[116]
Abbreviations—HRP: horseradish peroxidase; Gox: glucose oxidase; DOX: doxorubicin; EDC: 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide; NHS: N-hydroxysuccinimide; TMZ: temozolomide; MMP: matrix metalloproteinase; BG: O6-benzylguanine; Tm: triglycerol monostearate; MGMT: O6-methylguanine-DNA methyltransferase; β-gal: β-galactosidase; 5-FU: 5-fluorouracil; dPG: dendritic polyglycerol; PLGA−PEG−PLGA: poly(dl-lactide-co-glycolide)-b-poly(ethylene glycol)-b-poly(dl-lactide-co-glycolide); MCTS: multicellular tumor spheroid; MIONPs: magnetic iron oxide nanoparticles; HDI: 1,6-hexamethylenediisocyanate; DDI: 1,12-dodecamethylenediisocyanate; SD rats: Sprague-Dawley rats; PBA: phenylboronic acid; AIBN: azobis (isobutyronitrile).
Table 3. List of physical stimuli-responsive hydrogels.
Table 3. List of physical stimuli-responsive hydrogels.
StimulusPolymer(s)Drug/Dye/LigandCross-Linking AgentPreparation MethodCancer/Model/RouteFindingsReference
TemperatureChitosanDSF-Physical cross-linkingSMMC-7721 cellsHigh biocompatibility hydrogels that quickly gelled at body temperature and showed dose-dependently greater cytotoxicity compared to the free DSF solution may be given at room temperature.[129]
PLA-PEG-PLACTX and CpG-ODN--CT26 cells, CT26-bearing miceThe outcomes demonstrated that this combined approach decreases CTX toxicity while generating a cytotoxic T cell response that efficiently suppresses tumor growth, extends survival, and significantly increases the tumor cure rate.[130]
L-lysine and L-alanine-based diblock copolypeptidePTX--Glioblastoma (HK308 cells)Hydrogel loaded with paclitaxel caused less cellular inflammation, tissue damage, and reactive astrocytes than either the hydrogel or cremaphor-taxol (the usual taxol carrier). In vivo studies suggested local tumor control and improved survival.[131]
HPCS and F127-CHOICG and BSA, CaO2 NPs, Bi2S3-Schiff-base linkingL929 cells, 4 T1 cells, BALB/c nude miceICG@CaO2-BSA nanoparticles’ CaO2 broke down in the TME to produce Ca2+ and H2O2. In addition, ICG produced ROS when exposed to NIR radiation. Furthermore, when Bi2S3 nanorods and ICG were exposed to near-infrared radiation, they produced a photothermal effect that raised the temperature of tumor tissues, which helped to precisely destroy tumor cells.[132]
MXene nanosheetsDOX, FeCl2 solution, gellan gum-Physical cross linkingA549 and L-929 cellsMXene@GG demonstrated superior photothermal properties and precise drug release control. Additionally, cell studies confirmed MXene@GG’s high biocompatibility and the sustained anticancer efficacy of DOX.[133]
CS-g-PNIPAMGO-CET/
CPT11 and shRNA
NIPAM and MAAFree radical polymerizationU87 cells (glioblastoma), 3T3 fibroblasts, BALB/c nude miceIn vitro studies suggested cell apoptosis, reduced SLP2 protein expression, and inhibited cell migration. In vivo studies confirmed 40% tumor size compared with the untreated control group after 12 days.[134]
PDLLA-PEG-PDLLABVZ and DOXStannous octoate as catalystRing-opening polymerizationHaCaT and HeLa cells, HeLa xenograft nudeIn vitro studies showed negligible cytotoxicity on HeLa and HaCaT cells. In vivo studies suggested that hydrogel co-loaded with BVZ and DOX effectively suppressed tumors for 36 days after a single intratumoral injection, with no harm to vital organs.[135]
Alginate-grafted PNIPAMDOXEDC, NHS, MES bufferATRPAT3B-1 cellsDOX was gradually released from hydrogel, and had enhanced cellular uptake, good biocompatibility, and increased efficacy in inducing cancer cell death. [136]
LightAzobenzene and α-CD functionalized HAMSNs-AuNBs, DOXNIR radiationIn situ self-assemblyHaCaT and SCC cells, MCSUpregulation of hyaluronidase (HAase) near the tumor tissue caused hydrogel HA degradation and the release of the drug from hydrogel, which could be taken up by tumor cells and the drug is delivered to cell nuclei.[137]
CSMAGnp substrate, LAP405 nm laser-MCF-7, HepG2, and HeLa cells, healthy and cancer patients’ bloodStudy indicated that the isolation platform had acceptable biocompatibility and had isolated the selected cells successfully. This light-responsive hydrogel has potential for use in clinical applications.[138]
Ti3C2 MXene/celluloseDOXECHChemical cross-linkingHepA1-6, SMMC-7721, HepG2, U-118MG and U-251 MG cells, BALB/c or C57BL/6 miceThe results showed the promise of the nanoplatform for use in cancer therapy by demonstrating that the combination of PTT and adjuvant chemotherapy delivered via this nanoplatform destroyed tumors instantly and prevented tumor relapse. Notably, DOX was released from the hydrogel and had excellent photothermal action.[139]
Humic acid/agaroseSH and DOX-Physical cross-linking4T1 cells, 4T1 tumor-bearing BALB/c miceIn vivo studies suggested improved antineoplastic efficacy of hydrogel drugs in tumoral tissues compared to the local distribution of free drugs.[140]
MCMSNs, DOX--3T3 mouse fibroblasts and Cal27 human OSCC, female BALB/c miceChemotherapy and phototherapy together produced a less toxic, long-lasting synergistic antitumor impact both in vitro and in vivo.[141]
Abbreviations—DSF: disulfiram; CTX: cyclophosphamide; CpG-ODN: cytosine-phosphate-guanine oligonucleotide; PLA-PEG-PLA–poly(D, L-lactide)-poly(ethylene glycol)-poly(D, L-lactide); ICG: indocyanine green; BSA: bovine serum albumin; CaO2 NPs: calcium peroxide nanoparticles; Bi2S3: bismuth sulfide nanorods; HPCS: hydroxypropyl chitosan; F127-CHO: aldehyde-modified Pluronic F127; TME: tumor microenvironment; H2O2: hydrogen peroxide; ROS: reactive oxygen species; DOX: doxorubicin; GO-CET/CPT11: CS-g-PNIPAM: chitosan-g-poly(N-isopropylacrylamide); shRNA: stomatin-like protein 2 (SLP2) short hairpin RNA (shRNA); NIPAM: poly(N-isopropylacrylamide); MAA: mercaptoacetic acid; GO-CET/CPT11: irinotecan (CPT-11) to cetuximab (CET)-conjugated graphene oxide; PDLLA-PEG-PDLLA: poly (d, l-lactide)–poly (ethylene glycol)–poly (d; l-lactide); BVZ: bevacizumab; ATRP: atom transfer radical polymerization (ATRP); NHS: N-hydroxysuccinimide; MES: 2-(N-morpholino)-ethanesulfonic acid; MSNs-AuNBs: EDC: 1-ethyl-3-(dimethylaminopropyl)carbodiimide, gold nanobipyramid (AuNB)-conjugated mesoporous silica nanoparticles; azobenzene and α-CD functionalized HA: azobenzene and α-cyclodextrin-functionalized hyaluronic acid; SCC: human squamous carcinoma cells; MSC: multicellular spheroids; CSMA: chondroitin sulfate methacryloyl; Gnp substrate: gelatin nanoparticle-modified flat glass; LAP: lithium phenyl-2,4,6-trimethylbenzoylphosphinate; ECH: epichlorohydrin; PTT: photothermal performance; OSCC: oral squamous cell carcinoma; SH: sodium humate; MC: methylcellulose; MSNs: mesoporous silica nanoparticles.
Table 4. List of multi-stimuli-responsive hydrogels.
Table 4. List of multi-stimuli-responsive hydrogels.
StimuliPolymer(s)Drug/Dye/LigandCross-Linking AgentPreparation MethodCancer/Disease Model/RouteFindingsReference
Temperature and pHPNVCL, Vim, PVP5-FUMBAFree radical polymerizationNeoplastic cellsHydrogels of P(NVCL-co-VIm)/PVP across various pH and temperature conditions offer promise for targeted drug delivery applications.[164]
Temperature and pHPNIAAm-co-IA and CSDOXGPFree radical polymerizationBreast cancer (MCF-7 cells)Lower concentrations in an acidic environment (37 °C) demonstrated faster DOX release than a neutral pH and 40 °C. The hydrogels were cytocompatible and had negligible or no cytotoxicity according to the cytotoxicity analysis.[165]
Temperature and pHPGA and PNHLysozymeCarbodiimideRadical polymerization, ring-opening polymerization-The hydrogel’s potential as a smart drug carrier was highlighted by the quicker rates of lysozyme release at pH 7.4 along with a decreased cross-linking density and PNH content. At pH 4.0, the release of lysozyme was slowed due to protonation of the PGA portion.[166]
pH and glucosePEGS-PBA-BA and CS-DA-LAGrGO@
PDA and metformin
-Double dynamic bond between Schiff bases and phenylboronate estersType II diabetic foot woundsWith their increased adhesion, stimuli-responsive metformin release, and self-healing properties, the PC hydrogels were shown to be effective in helping chronic athletic diabetic foot wounds to recover.[121]
pH and ROS POD, CEDS and MFGroups from POD and CESchiff base linkages and boronic ester bondsChronic diabetic woundResults both in vitro and in vivo studies showed anti-infection, anti-oxidation, and anti-inflammatory effects at first, which were followed by enhanced angiogenesis and faster wound healing.[167]
Temperature and pH-PCLADOX-pH-GA, BA-Covalent cross-linkingHepatocellular carcinomaThe in vivo investigation demonstrated the efficacious inhibition of tumor growth by the DOX-releasing hydrogel depot. These results demonstrated the pH-responsive hydrogel’s intriguing potential for localized anticancer therapy.[168]
pH and enzymeHEMA and MAA5-FUOLZ-ACRadical copolymerizationHCT116 colon cells, rat colonic fluidLocal 5-FU release occurred at a colon location, and high 5-FU concentrations overcame cancer therapy resistance by promoting necroptosis in colon cancer cells.[169]
Temperature and enzymePEG, MMP peptideDOX, TSLsMichael-type reaction responsible for cross linkingThiol–maleimide reaction, chemical cross-linkingAoAF and NIH3T3 cellsInvestigations into in situ drug delivery and degradation demonstrated that the TSL-gel reacts to local environmental factors such as temperature and enzymatic stimulation.[170]
Abbreviations—5-FU: 5-fluorouracil; PNVCL: poly(N-vinylcaprolactam); Vim: 1-vinylimidazole; PVP: polyvinylpyrrolidone; PNIAAm-co-IA: poly (N-isopropylacrylamide-co-itaconic acid); MBA: N, N′-methylene bisacrylamide; DOX: doxorubicin; CS: chitosan; PNH: poly(N-isopropylacrylamide-co-2-hydroxyethyl methacrylate); GP: glycerophosphate; PGA: poly(l-glutamic acid); CS-DA-LAG: dihydrocaffeic acid and l-arginine-cografted chitosan; PEGS-PBA-BA: polyethylene glycol-co-poly(glycerol sebacic acid); POD: phenylboronic acid-grafted oxidized dextran; CE: caffeic acid-grafted ε-polylysine; rGO@PDA: polydopamine-coated reduced graphene oxide; DS: diclofenac sodium; MF: mangiferin; DOX-pH-GA: glucuronic acid-bearing doxorubicin; BA: boronic acid; PCLA: poly(ε-caprolactone-co-lactide)-b-poly(ethylene glycol)-b-poly(ε-caprolactone-co-lactide; HEMA: hydroxyethyl methacrylate; MAA: methacrylic acid; OLZ-AC: acryloyl chloride-modified olsalazine; TSLs: liposomes; AoAFs: human aortic adventitial fibroblasts.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Solanki, R.; Bhatia, D. Stimulus-Responsive Hydrogels for Targeted Cancer Therapy. Gels 2024, 10, 440. https://doi.org/10.3390/gels10070440

AMA Style

Solanki R, Bhatia D. Stimulus-Responsive Hydrogels for Targeted Cancer Therapy. Gels. 2024; 10(7):440. https://doi.org/10.3390/gels10070440

Chicago/Turabian Style

Solanki, Raghu, and Dhiraj Bhatia. 2024. "Stimulus-Responsive Hydrogels for Targeted Cancer Therapy" Gels 10, no. 7: 440. https://doi.org/10.3390/gels10070440

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