Abstract
Metal–organic frameworks (MOFs) have attracted chemical and material research interests in recent years due to their incredibly prominent properties and unique structures. Although MOFs have many unique characteristics, including tunable pore structures and functionality, their application performances are hindered by their powdered crystalline state, intrinsic fragility, poor processability, and stability. Recent studies have shown that the incorporation of MOFs into hydrogel outperforms the MOFs in their crystalline state. This article presents the current development of stimulus-responsive MOF–hydrogel composites in terms of synthesis of MOF and preparation of MOF–hydrogel, characterization, and advancement of stimulus-responsive MOF–hydrogel composite in drug delivery and wound management. The article also discusses prospective study directions, delineating potential avenues for further exploration and innovation in this dynamic field.
Graphical abstract

1 Introduction
Metal–organic frameworks (MOFs) are a rapidly expanding family of crystalline porous materials that have attracted much research interest because of their tunable porosity and functionality, absorption ability, and chemical and thermal stability [1,2]. MOFs are formed by the linkages between metal nodes and organic ligands through coordination bonds [2]. The tunability of MOFs has shown great promise for a diverse range of applications such as gas adsorption, energy storage, catalysis, luminescent sensing, environmental remediation, and medical treatments [3–9]. In recent years, MOFs have been reported on their applications in drug delivery to encapsulate drug molecules within their porous structure and serve as a reservoir for the controlled release of drugs. The high porous structure makes it a prospect carrier for the controlled release of therapeutic agents. However, there are some limitations of the conventional MOFs in drug delivery and medical treatments such as fixed functionality and selectivity, rigid and poor adaptability characteristics toward changes in the bioenvironment [10], and processability [11]. Conventional MOFs have fixed functionality with fixed pore sizes and selectivity. Thus, it lacks precision in the controlled release of drugs at the localized sites. Conventional MOFs generally have a rigid structure and functionality and cannot undergo reversible structural transformations in response to external stimuli [12,13]. These limitations have hindered their responsiveness to specific biological conditions or stimuli-triggered behavior. Thus, it lacks precision in the controlled release of drugs at the localized sites.
To address these limitations, stimulus-responsive MOFs are the emerging functional materials that can be used in many frontier domains owing to their special structure and functionality with open activity sites and reversible physicochemical characteristics [14]. The stimulus-responsive MOFs can be synthesized by incorporating responsive elements such as switchable or stimulus-responsive linkers [15] to overcome the limitations of conventional MOFs and enable new applications in sensing [16], environmental remediation [17], and medical and biomedical applications [18,19]. The reversible physicochemical properties make the stimulus-responsive MOFs an ideal candidate to respond to physical stimuli (e.g., temperature and light), chemical changes, or biological changes (e.g., redox, pH, and adenosine triphosphate [ATP]) by modifying their physical or chemical properties [20,21]. In addition, both conventional and stimulus-responsive MOFs are also encountering stability and processability challenges as MOFs are often fragile and lack mechanical stability [14]. To overcome this limitation, embedding MOFs within a hydrogel matrix can provide mechanical support and enhance the stability of the MOFs. Hydrogel is a three-dimensional (3D) network of highly hydrated crosslinked hydrophilic polymers that can contain a wide range of structural forms and chemical compositions [22]. The 3D networks promote the dispersibility of MOFs in the hydrogel matrices and provide mechanical support to protect the MOFs from disintegration. Conversely, MOFs are also able to interact with the hydrogel components and tuning the hydrogel properties which could result in a synergistic effect of MOF–hydrogel composite [14]. The stable dispersion of MOFs in the hydrogel matrices makes it an attractive material in biological applications such as wound healing and drug delivery. This article provides a comprehensive overview of the emerging development of MOFs to stimulus-responsive MOFs and then the encapsulation of stimulus-responsive MOF by hydrogel to form stimulus-responsive MOF–hydrogel composites. This article also summarized the crucial characterization methods of the MOF and MOF–hydrogel and the advancement of stimulus-responsive MOF–hydrogel in medical applications to provide fundamental knowledge of stimulus-responsive MOF–hydrogel and its application in drug delivery and wound management.
2 MOFs in medical applications
The structure of MOFs greatly relies on the type of metal ions and organic ligands used in the synthesis reaction. The selection of metal centers (metal salt) and ligands (also known as linkers) will generate a particular MOF crystal structure with specific characteristics of pore size and adsorptive sites [23]. The formation of a 3D rigid framework is dependent on multidentate linkers such as carboxylates, imidazoles, and pyrazoles. This is due to their ability to aggregate the metal ions into M–O–C or M–N clusters, where the metal ions (M) are locked into the position in the network vertex by the carboxylate, imidazolate, or pyrazolate ligands [24]. Usually, low-toxicity metal ions such as Mg, Ca, Zr, Ti, Fe, Al, and Mn are used in MOF synthesis.
Back in 1999, the Hong Kong University of Science and Technology (HKUST) MOF was synthesized by Chui et al. [25] HKUST-1, also known as MOF-199, was produced using copper as the metal center. It contains Cu2+ ions that are coordinated by four carboxylate groups in a well-known paddle-wheel structure of copper acetate (Figure 1c). The 3D coordination polymer [Cu3(BTC)2(H2O)3] crystallizes with the formation of a highly porous cubic structure with a complicated 3D network of channels [26]. Matériaux de l′Institut Lavoisier (MIL) MOF was first synthesized and reported by Gérard Férey’s group back in 2002 [27]. These open frameworks of MOFs were derived using trivalent cations, such as iron(iii), vanadium(iii), and chromium(iii), which are then extended with the use of p-elements such as indium(iii), aluminum(iii), and gallium(iii) shown in Figure 1a and b. These classes of MOFs resemble zeolite topologies but differ in pore size, surface chemistry, and density [28]. As for zeolite imidazole framework (ZIF), it is an MOF sub-class that is composed of imidazolate linkers and tetrahedrally coordinated transition metal ions such as Zn, Co, Cu, and Fe [29]. Their porous crystalline structures adopt an analogous fashion to that of silicon and oxygen in zeolites (Figure 1d) [30]. Zirconium oxide-based MOFs such as UiO-66 (Figure 1e) with 12-coordinated Zr6(M3–O)4(M3–OH)4(CO2)12 clusters were originally reported by Cavka et al. in 2008 [31]. It was first synthesized at the University of Oslo and named after the university. These classes of MOFs have a face-centered-cubic topology with high chemical and thermal stability [31]. New zirconium MOFs were also synthesized from the Zr6O4(OH)4(CO2) n secondary building units (n = 6, 8, 10, or 12) and variously shaped carboxyl organic ligands to make extended porous frameworks such as the structures shown in Figure 1f and g [32].
![Figure 1
(a) Crystal structures of MIL-100 (Fe). Reuse under terms of CC-BY 4.0 license [33]. Copyright 2019 Molecules. (b) MIL-68. Adapted with permission from Volkringer et al. [34]. Copyright 2008 American Chemical Society. (c) HKUST-1. Adapted with permission from Sharma et al. [26]. Copyright 2021 American Chemical Society. (d) ZIF-8. Reuse under terms of CC-BY 4.0 licensed [35]. Copyright 2021 Applied Sciences. (e) UiO-66. Adapted with permission from Winarta et al. [36]. Copyright 2021 American Chemical Society. (f) UiO-68. Adapted with permission from Kutzscher et al. [37]. Copyright 2016 American Chemical Society. (g) MOF-808. Adapted with permission from Tan et al. [38]. Copyright 2019 American Chemical Society.](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_001.jpg)
(a) Crystal structures of MIL-100 (Fe). Reuse under terms of CC-BY 4.0 license [33]. Copyright 2019 Molecules. (b) MIL-68. Adapted with permission from Volkringer et al. [34]. Copyright 2008 American Chemical Society. (c) HKUST-1. Adapted with permission from Sharma et al. [26]. Copyright 2021 American Chemical Society. (d) ZIF-8. Reuse under terms of CC-BY 4.0 licensed [35]. Copyright 2021 Applied Sciences. (e) UiO-66. Adapted with permission from Winarta et al. [36]. Copyright 2021 American Chemical Society. (f) UiO-68. Adapted with permission from Kutzscher et al. [37]. Copyright 2016 American Chemical Society. (g) MOF-808. Adapted with permission from Tan et al. [38]. Copyright 2019 American Chemical Society.
The reported MOFs used in medical and biomedical applications are summarized in Table 1, according to the classes of MOFs. Table 1 provides a summary of a literature search conducted in the Web of Science database using the following keywords: “metal-organic framework,” “MOFs,” “hydrogel,” MOF-based hydrogel, and “medical applications.” The MOFs were used in drug delivery, biomarker, wound healing, and other non-specific biomedical applications. The metal ions used in MOF synthesis for medical applications are Eu, Ga, Fe, Au, Cu, Co, Zn, Zr, and Ni (Table 1). Fe, Cu, Co, and Zn are known to be low in toxicity and needed as essential trace elements for normal biological functioning, while Eu, Ga, Au, Zr, and Ni are usually chosen for MOF synthesis as a filler component to enhance the mechanical strength of the overall hydrogel composites [39]. The organic aromatic multicarboxylate ligands such as 2,3,4,5,6-benzenehexacarboxylic acid (TCI), trimesic acid (H3BTC), terephthalic acid (1,4-H2BDC), 1,4-dicarboxylic acid (1,4-NDC), and 1,4-benzenedicarboxylic acid (BDC) were used in the synthesis of MOFs mainly due to their rigid nature, topologies, and high thermal stability [40]. Other reported ligands that were being used in the synthesis of MOFs are heterocyclic compounds containing nitrogen donors such as 1,2-bis(4-pyridyl)ethylene, 2-methylimidazole, nicotinic acid, and amino-p,p′-terphenyldicarboxylic acid. Most of the particle sizes of the reported MOFs in Table 1 are in the range of 100–500 nm, while the larger particle size is up to 4 mm. The unique properties of MOFs, for instance, high porosity, pore size, and pore surface of MOFs, are determined by the properties of both metal ions and linkers. Also, the structural properties of the MOFs are affected by the solvent system, pH of the solvent, ratios of metal and ligand, and reaction temperature [41]. The reported MOFs in Table 1 were mainly synthesized by conventional techniques, i.e., solvothermal, hydrothermal, and room temperature in situ synthesis, where these reactions required water or solvents (ethanol, methanol, dimethylformamide [DMF], tetrahydrofuran, and dimethylacetamide [DMAC]) or a mixture of solvents as the reaction medium.
Summary of MOFs used in medical and biomedical applications
| Classes of MOF | MOF | Metal ions | Organic ligands | Synthesis method of MOF | Solvent used | Reaction temperature and duration | Particle size of MOF | Applications | Ref |
|---|---|---|---|---|---|---|---|---|---|
| MIL series | Eu@MIL-116(Ga) | Eu and Ga | TCI | Hydrothermal synthesis | Deionized water | Heated at 210°C for 20 h | N/A | Detection of antineoplastic mitoxantrone | [41] |
| MIL-100 | Fe | Trimesic acid (H3BTC) | One-pot synthesis | Deionized water | Room temperature stirred overnight | N/A | Controlled drug release | [42] | |
| HKUST series | Au@HKUST-1 | Au and Cu | Trimesic acid (H3BTC) | One-pot synthesis | Ethanol | Room temperature and stirred for 24 h | N/A | Detection of adenosine | [43] |
| HKUST-1 NP | Cu | Room temperature synthesis | Ethanol and water mixture | Room temperature and stir for 15 min | 285.6 ± 2.7 nm | Wound healing | [44] | ||
| HKUST-1 | Solvothermal synthesis | Ethanol and water mixture | Heated at 120°C for 12 h | 4 mm | Biomedical application | [45] | |||
| Cu-MOF and Zn-MOF | Cu, Co, and Zn | 1,2-Bis(4 pyridyl)ethylene | Hydrothermal synthesis | Distilled water | Heated at 80°C for 48 h | 100–500 μm | Antibacterial applications | [46] | |
| Co-MOF | Solvothermal synthesis | DMF and distilled water | Heated at 100°C for 72 h | ||||||
| ZIF series | ZIF-8 | Zn | 2-Methylimidazole | Solvothermal synthesis | Ethanol and water mixture | Heated at 120°C for 12 h | 4 mm | Biomedical application | [45] |
| ZIF-8 | One-pot synthesis | Deionized water | Room temperature stirred for 30 min | N/A | Controlled drug release | [41] | |||
| ZIF-8 | Solvothermal synthesis | Methanol | Heated at 80°C for 24 h | 691 ± 72 nm | Fabrication of customized tissues and organs | [47] | |||
| ZIF-8 | Room temperature synthesis | Methanol | Room temperature stirred for 24 h | 500 nm | Drug delivery materials | [48] | |||
| Zirconium series | UiO-66 | Zr | Terephthalic acid (1,4-H2BDC) | Solvothermal synthesis | DMF | Heated at 120°C for 24 h | 500 nm | Drug delivery materials | [48] |
| MOF-808 | Trimesic acid H3BTC | Solvothermal synthesis | DMF and formic acid mixture | Heated at 100°C for 7 days | N/A | Drug delivery and the removal of tumor contrast agent/dye | [49] | ||
| Others | Cu or Zn–vitamin frameworks | Cu and Zn | Nicotinic acid | Coaxial capillary microfluidic system with hierarchical injection channels | Calcium chloride solution | N/A | N/A | Tissue wound healing | [50] |
| Cu-TCPP(Co) MOF | Cu | 5,10,15,20-Tetrakis(4-carboxyphenyl)porphyrin-cobalt(ii) | Solvothermal synthesis | DMF | Heated at 120°C for 8 h | 10 nm | Chemiluminescence biosensor for thrombin detection | [51] | |
| EuNDC | Eu | 1,4-Dicarboxylic acid (1,4-NDC | Solvothermal synthesis | DMF | Heated at 115°C for 60 h | N/A | Fluorescence detection | [52] | |
| Dysprosium(iii) organic framework | Dy(NO3)3·6H2O | [1,3,5-Benzenetriyltris(carbonylimino)]tris-benzoic acid | Hydrothermal | NaOH (aq) | Heated at 160°C for 96 h | N/A | Detection of picric acid and treatment activity on human lung cancer cells | [53] | |
| 2D Ni–Fe MOF | Fe and Ni | Terephthalic acid (1,4-H2BDC) | Solvothermal synthesis | Deionized water and DMAC | Heated at 150°C for 3 h | N/A | Artificial cartilage materials | [54] | |
| Cd(ii) (MOF) | Cd(NO3)2·4H2O | Bis(2-carboxyethyl)isocyanurate | Solvothermal synthesis | DMF | Heated at 110°C for 3 h | N/A | Candidate drug for treating Hepatitis B | [55] | |
| PBNPs | Fe | Polyvinylpyrrolidone | Hydrothermal synthesis | Deionized water | Stirred at 60°C for 30 min | 100 nm | Wound healing | [18] |
As previously mentioned, conventional MOFs exhibit certain limitations regarding their fixed functionality and selectivity. Additionally, they demonstrate rigid and poor adaptability in response to changes in the bioenvironment, which in turn affects their processability. Upgradation of MOFs into stimulus-responsive MOFs would enhance the functionality, selectivity, and responsiveness of the MOFs toward complex biological environments.
2.1 Stimulus-responsive MOFs in medical–biomedical applications
Stimulus-responsive MOFs, also referred to as smart MOFs, exhibit detectable changes in their physical and/or chemical properties in response to one or more stimuli. These stimuli can either induce dynamic alterations in the overall structure or leave the structure unchanged and rigid. As a result of these alterations, MOFs can undergo reversible or irreversible changes. Reversible changes can be reversed by either halting the stimulus or applying an opposing stimulus, whereas irreversible changes are permanent and cannot be restored once the stimulus is stopped or countered [21]. The stimuli-responsive MOF can be classified into three categories mainly: physical, chemical, and multiple responsive stimuli. Physical responsive MOFs would change their physical and/or chemical and/or thermal properties in response to light and temperature. Conversely, chemical-responsive MOFs respond to analytes such as pH, biomolecules, and redox. As for multiple responsive stimuli, the MOFs can respond to multiple stimuli at the same time by altering their physical and/or chemical and/or thermal properties [56,57]. In the subsequent section, the emerging applications of different classes of stimulus-responsive MOFs are discussed.
2.1.1 Physical stimuli
2.1.1.1 Light-responsive MOFs
Light-mediated therapy is recognized for its non-invasive nature and precise control over time and space, making it a valuable therapeutic and diagnostic tool for specific areas both in vitro and in vivo using light at a particular wavelength. Diring et al. [58] introduced a novel method for releasing carbon monoxide (CO) using an MOF. They achieved this by embedding a manganese carbonyl complex, MnBr(bpydc)(CO)3 (bpydc = 5,5′-dicarboxylate-2,2′-bipyridine), into the walls of a durable Zr(iv)-based MOF via a post-synthetic immobilization process. CO release was triggered by visible light (wavelength of 460 nm and intensity of 300 W) and monitored using infrared and UV–Vis spectroscopy. Upon irradiation, the disappearance of the characteristic CO stretching vibration indicated changes in the Mn(i) coordination sphere, marking the release of CO (Figure 2). This light-responsive MOF is designed for precise targeting of tissues or organs, potentially revolutionizing tissue-specific therapy by providing localized CO release with anti-inflammatory benefits, while avoiding the toxicities associated with systemic CO administration in the bloodstream and lungs [58].
![Figure 2
Schematic representation of MnBr(bpy)(CO)3 loading onto UiO-67-bpy to synthesize CORF-1. Upon light irradiation, the loaded complex undergoes CO release. Reuse under terms of the CC-BY license [58]. Copyright 2016 Royal Society of Chemistry.](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_002.jpg)
Schematic representation of MnBr(bpy)(CO)3 loading onto UiO-67-bpy to synthesize CORF-1. Upon light irradiation, the loaded complex undergoes CO release. Reuse under terms of the CC-BY license [58]. Copyright 2016 Royal Society of Chemistry.
2.1.1.2 Temperature-responsive MOFs
Temperature-mediated therapy is used in oncology treatment due to its remote controllability and non-invasiveness. Jiang et al. [59] reported an isoreticular zirconium-cluster-based MOF denoted as ZJU-801 and applied it as a drug delivery system using heat to activate its drug release. The encapsulated diclofenac sodium (DS) drug (DS@ZJU-801) was released when the temperature was raised. The temperature rise caused the breaking of the π–π interactions between MOF and drug molecules (Figure 3). The release rate was reported to be ten times higher at 60°C than at 25°C due to the breakdown of π–π interactions between the MOF and drug molecules [59].

Schematic representation of π–π interactions between DS and MOF. As the temperature increased, the breaking down of π–π interactions released DS.
2.1.2 Chemical stimuli
2.1.2.1 pH-responsive MOFs
Compared to normal tissue, extracellular tumor tissues tend to exhibit a slightly lower pH because of their abnormal cell growth, whereby the cancerous cells consume nutrients and induce carbon dioxide generation, causing a decrease in the local pH of the tumor microenvironment. During the synthesis of MOFs, acidic or basic reaction conditions play a crucial role in the formation of MOFs in influencing the crystallization and coordination of the MOFs. Therefore, the coordination bonds of the MOFs can be modified to detect external pH [60].
Zhu et al. [61] reported on iron-based MIL-100 MOFs formed by a multifunctional core–shell with polypyrrole (PPy) nanoparticles (NPs) as cores and iron(iii) carboxylate MOFs (MIL-100) as the outer shell. MIL-100 was loaded with the anticancer drug, doxorubicin (DOX) by coating the drugs on its outer shell. Over 80% of DOX was released from MIL-100 at pH 5.0. The acidic environment caused gradual degradation of the MIL-100 structure, resulting in the weakening of the electrostatic interactions between DOX and MIL-100 shell (Figure 4) [61].
2.1.2.2 Redox-responsive MOFs
Tumor cells flourish in a reducing environment primarily controlled by the oxidation and reduction states of glutathione (GSH) and nicotinamide–adenine dinucleotide phosphate. This is because GSH controls the cellular reducing environment through fragmentation and disulfide bond formation, hence causing an increase in GSH concentration, which is used to regulate the tumor microenvironment. Due to these unique features, drug delivery systems have been designed to detect the reducing environment caused by the tumor cells to trigger drug release via the cleavage of the disulfide bond in GSH-sensitive materials [62]. Lei et al. [63] reported a redox-responsive MOF carrier using metal nodes such as aluminum, iron, or zirconium and 4,4′-dithiobisbenzoic acid (4,4′-DTBA) as the organic ligand where the disulfide bond was cleaved by GSH which is often overexpressed in tumor cells (Figure 5). MOF-Zr(DTBA) with zirconium as the metal ion was identified as a potential carrier for curcumin (CCM). Drug release of CCM-MOF-Zr(DTBA) and CCM@UiO-67 was demonstrated in vitro by exposing them to DL-dithiothreitol that mimics the tumor microenvironment. CCM-MOF-Zr(DTBA) shows a superior release behavior and enhanced cell death compared to CCM@UiO-67, indicating GSH responsiveness in the tumor microenvironment due to the cleavage of the disulfide bond, which accelerated the drug release from MOF-Zr(DTBA) [63]. In summary, the single-responsive MOFs for biomedical applications are driving targeted therapy. These MOFs, which are sensitive to light, temperature, and redox triggers, offer precise drug delivery. Light-responsive MOFs enable localized therapy, bypassing systemic toxicities. Temperature-triggered systems like ZJU-801 facilitate controlled drug release, which is beneficial in oncology. Redox-responsive MOFs like MOF-Zr(DTBA) exploit biomolecules in tumors for triggered drug release. These advancements are demonstrating a shift toward targeted therapy in order to minimize damage to healthy cells.
![Figure 5
Redox-responsive degradation of curcumin-loaded MOF (CCM@MOF-M(DTBA)) in tumor cells. Adapted with permission from Lei et al. [63]. Copyright 2018 American Chemical Society.](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_005.jpg)
Redox-responsive degradation of curcumin-loaded MOF (CCM@MOF-M(DTBA)) in tumor cells. Adapted with permission from Lei et al. [63]. Copyright 2018 American Chemical Society.
2.1.3 Dual- and multi-responsive MOFs
To enhance the development of cancer therapies, dual- and multi-responsive stimuli MOFs have emerged to give dual or multiple triggers rather than a single stimulus response due to the complex human body environment. Hence, by harnessing the synergistic effects of different stimuli, such as pH, temperature, light, and redox conditions, dual- or multi-responsive MOFs can provide enhanced control and precision in drug release, leading to more effective and targeted cancer treatments.
Nagata et al. reported dual pH and temperature-responsive MOFs for drug release via surface modification of MOFs using a copolymer of acrylic acid (AA) and N-isopropyl acrylamide (NIPAm) in the post-synthetic process [64]. The polymer exhibited both pH and temperature responsiveness, which allowed the release of procainamide from the MOF in an on–off manner. At pH 6.8 or temperature lower than 25°C, the polymer exhibited a fast release of procainamide. However, at pH 4 or higher temperature (>40°C), procainamide molecule release was suppressed, thereby demonstrating the MOFs were pH and temperature sensitive. This dual stimuli MOF was proposed to be used in controlled drug delivery by responding to the changes in pH and temperature during the therapy.
Jiang et al. [65] reported a dual ATP/pH-responsive MOF (ZIF-90), synthesized by a fast assembly process. ZIF-90 was loaded with DOX and conjugated to neuropeptide Y Y1 subtype receptor, as a new target site for breast cancer treatment via surface modification. The modified ZIF-90 reduced the premature DOX release and effectively triggered faster DOX release in the tumor cells in response to low pH conditions and high ATP in the tumor cells. By combining targeted delivery of DOX and dual-responsive DOX release, the study reported an 80% survival rate in MDA-MB-231 tumor-bearing mice after 40 days of treatment with minimal side effects on the liver and renal functions. The modified ZIF-90 was proposed to be used as a targeted triple-negative breast cancer treatment [65].
Tan et al. also reported on a multi-Ca2+/pH and thermo triple-responsive mechanized Zr-based MOFs for drug delivery in bone disease [66]. The loading of the drug 5-fluorouracil (5-Fu) into the CP5-based mechanized UiO-66-NH2 nanocarrier system was carried out by introducing the drug into the nanopores of the scaffold at room temperature. This was followed by the introduction of CP5 rings via host–guest complexation, forming pseudorotaxanes as the movable elements of the mechanized Zr-MOFs, thus achieving drug encapsulation. The drug release from the mechanized MOFs can be triggered by changes in calcium ion concentration (Ca2+), pH, and temperature. These triggers caused changes in the MOF structure or environment that resulted in the release of the encapsulated drug. The addition of Ca2+ ions led to an increase in the encapsulation capacity. The release was triggered in an acidic environment where the neutralization of CP5 sodium salts weakened the host–guest interactions, leading to the unblocking of the nanopores. Elevating the temperature to 60°C weakened these interactions and promoted the gradual release of 5-Fu. At normal body temperature or room temperature, the premature release of the drug was not significant. This aided the targeted delivery of the drug to the appropriate site, such as a tumor subjected to hyperthermia treatment. Overall, these reported stimuli-responsive MOFs contributed to a controlled and targeted drug delivery system capable of responding to specific physiological conditions or treatment protocols.
To summarize, the emergence of dual- and multi-responsive MOFs represents a significant step forward in cancer therapy, as compared to single-responsive MOF. The dual- and multi-responsive MOFs offer promising avenues due to their tunable properties in response to various stimuli such as pH, temperature, light, redox, and even combinations of stimuli-responsive MOFs such as pH/temperature-responsive and ATP/pH-responsive conditions. The dual- and multi-responsive MOFs present enhanced precision in targeted drug release for oncology therapy. The multi-responsiveness of MOFs holds potential for advancing combined therapies, such as chemoradiation therapy, complementing localized irradiation for a synergistic treatment effect.
2.2 Synthesis of stimulus-responsive MOFs
Regardless of conventional or stimulus-responsive MOFs, hydrothermal, solvothermal, and direct precipitation are the common and cost-effective synthesis methods for MOFs, as compared to non-conventional methods such as microwave-assisted method, electrochemical method, mechanochemical method, and sonochemical method. Typically, in the solvo/hydrothermal method, the metal ions and organic ligands are dissolved in the solvent and placed in an enclosed reaction vessel under high pressure and temperature conditions for the formation and self-assembly of MOF crystals. The commonly used organic solvents in the solvothermal method are acetonitrile [67], N,N-diethylformamide [68], DMF [68], ethanol, and methanol [69], whereas water is used in the hydrothermal method [70]. As for the direct precipitation method, it usually occurs at ambient temperature using DMF as the solvent.
Teflon-lined autoclaves are generally used for solvo/hydrothermal methods as a reaction chamber reacts under the desired conditions as they can withstand high temperature and pressure during the reaction. When compared to glass and quartz autoclaves, it exhibits a strong resistance to acid and alkaline media [71]. The advantage of this reaction method is that almost any metal ions and organic ligand material can be dissolved in the solvent and synthesized into MOFs by varying different reaction parameters, such as pressure, reaction time, reaction medium, temperature, pH, volume of autoclave, and the concentration of the reactants [71,72]. Figure 6 illustrates the synthesis of MOFs using solvothermal, hydrothermal, or direct precipitation methods. The synthesis of MOFs involves the process of crystallization, during which the nucleation and growth of crystals occur. The nucleation and growth of MOF crystals involve the self-arrangement of metal–oxygen clusters and organic ligands. Once the reaction comes to an end, the as-synthesized mixture will be cooled to room temperature. The products are then washed several times continuously with a deionized solution such as water to remove the remaining unreacted raw materials and impurities. Subsequently, anhydrous ethanol or other solvents are used to purify the product, and pure MOFs are obtained after vacuum drying. In recent years, the synthesis of bimetallic organic frameworks has been reported. Ngan Tran et al. reported on the synthesis of M/Fe-MOF (M-Co2+, Cu2+, and Mg2+) bimetallic MOFs by using solvothermal method [73]. The bimetallic MOFs were reported for their responsiveness toward visible light and responded by giving high photocatalytic degradation of RhB organic pigments. Tayyab et al. also reported on the synthesis of ternary In2S3/Nb2O5/Nb2C Schottky/S-scheme integrated heterojunction for efficient photocatalytic hydrogen production via one-pot in situ hydrothermal synthesis method [74].

Illustration of MOF formation in hydrothermal or solvothermal or direct precipitation synthesis.
2.2.1 Direct precipitation synthesis
Stimulus-responsive MOFs can also be synthesized via direct precipitation synthesis. The responsiveness can be imparted using a stimulus-responsive ligand [23]. In this method, MOFs are built using an organic ligand containing ionizable chemical groups that have the potential to undergo protonation and exhibit charge reversal when exposed to acidic conditions. Wang et al. [15] reported a synthesized ZIF-8 loaded with 5-FU (anticancer drug) that exhibited a pH-triggered controlled drug release property. ZIF-8 was synthesized by dissolving Zn(NO3)2·6H2O in deionized water and 2-methylimidazole in methanol. The zinc nitrate solution was added to the 2-methylimidazole solution while stirring at room temperature. The mixture was then mechanically stirred at 500 rpm for 30 min. ZIF-8 was collected by centrifugation, washed with methanol, and dried overnight in an oven at 80°C [15].
2.2.2 Surface modification
MOFs that are not inherently responsive to changes in the environment can be modified via surface modification by functionalizing the surface of MOFs with amine groups [72], responsive polymer materials such as chitosan [75], carboxymethylcellulose (CMC) [76], pectin [77], or other photo-responsive material [78] for biomedical applications.
Javanbakht et al. reported on the development of oral vehicles by coating the nano-encapsulating MOFs with a biopolymer. The bio-nanocomposite beads functioned as pH-responsive oral drug delivery systems for anti-inflammatory purposes. This method utilized a combination of different biopolymers, including gelatin, CMC, and κ-carrageenan (κ-Cr), along with various MOFs such as UiO-66 and Cu-MOF. When subjected to the conditions of the gastrointestinal tract, the biopolymer-coated MOFs exhibited responsiveness corresponding to the stomach’s acidic pH. Additionally, the biopolymer-coated MOFs contributed to the enhanced stability of the drug molecules, leading to improved therapeutic efficacy [78–82].
3 Stimulus-responsive MOF–hydrogel composites
While stimuli-responsive MOFs have garnered significant attention for their exceptional functionalization and stimulus responsiveness properties in medical and biomedical domains, it is important to acknowledge the prevailing challenges. The instability of MOFs over extended durations in diverse environmental conditions, coupled with issues related to processability, stands as noteworthy limitations that need to be addressed [23]. Hence, to address the challenges, stimulus-responsive MOFs were incorporated into the hydrogel to form stimulus-responsive MOF-hydrogel composites that can be shaped into various forms of materials to enhance their robustness, processability, and versatility for diverse ranges of products.
Table 2 summarizes the stimulus-responsive MOF–hydrogel composites that were used in medical applications based on the classes of hydrogel used. The stimulus-responsive MOF–hydrogel composites were found to be mostly used in drug delivery and wound treatment dressing applications. The reported single-responsive MOF–hydrogel composites are evident to demonstrate responsiveness to physical stimuli such as temperature, pH, and light, and chemical stimuli, namely ATP molecules, peroxide, reactive oxygen species (ROS), and glucose (Table 2). Conversely, the reported dual- and multi-responsive MOF–hydrogel composites are ATP and pH responsiveness, light and temperature responsiveness, light and pH responsiveness, and temperature and pH responsiveness. The stimulus-responsive MOF–hydrogel composites are mostly prepared by dispersing the stimuli-responsive MOFs in a polymer precursor solution, where the polymer components will surround the stimuli-responsive MOFs in which the chemical crosslinking occurs between the MOFs and polymer components via free radical polymerization [83,84]. The stimuli-responsive MOFs act as the dispersed phase and the hydrogel as the continuous phase. The responsiveness of the MOF–hydrogel composites can be contributed by the MOFs itself (ZIF-8, Cu@ZIF/GO x , Au@ZIF-8, LMNPs@MOFs, Ins@ZIF-8, and Pd(H)@ZIF-8), or via surface modification of the MOFs with responsive moieties, such as biopolymeric carboxymethylcellulose, pectin biopolymer, and polyethylene glycol-thioketal (PEG-TK), or by loading a responsive molecules (Prussian blue, Rose Bengal, glucose oxidase [GO x ], and sodium nitroprusside [SNP]) into the MOFs or polymeric hydrogel materials (methacryloyl chitosan, κ-Cr, alginate, poly-(polyethyleneglycol citrate-co-N-isopropylacrylamide) [PPCN], poly(lactic acid-co-glycolic acid)/poly(ethylene glycol) [PLGA-PEG-PLGA], poly(ethylene glycol)poly(ε-caprolactone-co-lactide), and polyacrylamide [PAM]).
Reported works on stimulus-responsive MOF–hydrogel in medical applications
| Classes of hydrogel | Hydrogel | Forms of MOF–hydrogel | Name of MOF | Synthesis method of MOF | Preparation method of MOF–hydrogel | Type of stimuli; stimulus responsiveness | Responsiveness moiety | Unique properties of MOF–hydrogel | Applications | Ref |
|---|---|---|---|---|---|---|---|---|---|---|
| Natural polymer-based hydrogel | Methacryloyl chitosan | Gel | Hydroxyapatite cored Mg-gallic acid MOF NPs | Hydrothermal | Ultraviolet (UV) irradiation (365 nm at 10 mW/cm2 for 60 min) grafting to form MOF-methacryloyl chitosan hydrogels | Single responsive (pH) | Methacryloyl chitosan hydrogel | Sustained release of Ca2+ at pH 7.4 provides a good microenvironment for new bone regeneration | Orthopedic implant | [85] |
| Pectin | Beads | Bio-MOF-11 | Hydrothermal | Grinding | Single responsive (pH) | Surface modification of bio-MOF-11 with pectin biopolymer | Long-term controlled drug release | Drug release | [79] | |
| κ-Cr | Beads | UiO-66 | Solvothermal | Direct mixing | Single responsive (pH) | κ-Cr | Sustained drug release | nontoxic oral delivery vehicle | [82] | |
| Cellulose | Gel | Cu@ZIF | Solvothermal | Direct mixing | Single responsive (peroxide sensitive) | GO x on the Cu ions doped graphited imidazolate framework (ZIF-8). Cu@ZIF/GO x s | Displayed significant antibacterial efficacy in the presence of normal and hyperglycemic levels of glucose in human blood | Hemostatic and antibacterial wound dressing | [86] | |
| Chitosan | Gel | PBNPs | Hydrothermal | Direct mixing | Single responsive (light) | PBNPs | Eliminates bacteria within 20 min through the combined effect of heat and an electropositive surface | Wound healing | [18] | |
| Crossed linked alginate/pectin hydrogel | Gel | ZIF-8 | Mixing at 30°C | Direct mixing | Single responsive (ROS) | SP-loaded ZIF-8 (SP@ZIF-8) NPs were fabricated and coated with PEG-TK to obtain a ROS-responsive trait | Showed high SP-loading efficiency and a novel release mechanism responsive to ROS stimuli | Wound dressing applications | [87] | |
| Sodium alginate carbohydrazide-modified methacrylated gelatin | Gel | Au@ZIF-8 | Self-assembly | Direct mixing | Single responsive (light) | Au@ZIF-8 | Accelerate wound closure | Wound treatment dressing | [88] | |
| PNIPAm/κ-carrageenan | Gel | ZIF-8 | In situ | In situ | Dual responsive (light and temperature) | Poly(N-isopropylacrylamide) (PNIPAm) and methacrylated κ-carrageenan (MA-κ-CA) | Controllable photothermal–chemical synergistic antibactericidal properties | Wound dressing | [89] | |
| Gelatin methacrylate (GelMA) | Hydrogel microsphere | ZIF-8 | Self-assembly | One-step microfluidic technology under UV | Single responsive (pH) | ZIF-8 | Sustained release of drug | Drug delivery | [19] | |
| CMC | Hydrogel beads | Cu-MOF | Direct mixing at 125°C for 24 h | Direct mixing | Single responsive (pH) | Surface modification of Cu-MOF with biopolymeric carboxy- methylcellulose (CMC) | Controlled drug delivery system | Drug delivery system for oral administration of Ibuprofen | [78] | |
| CMC | Bead | MOF-5 | Solvothermal | Direct mixing | Single responsive (pH) | Surface modification of MOF-5 with biopolymeric carboxy- methylcellulose (CMC) | Controlled release of 5-FU | Potential DDS for oral administration of 5-FU | [80] | |
| Alginate | Gel | Eutectic gallium indium (EGaIn) LMNPs and ZIF-8 NPs | Self-assembly | Direct mixing | Dual responsive (Light and temperature) | LMNPs@MOFs SupraParticles (light-responsive) Alginate hydrogel (temperature responsive) | Improve processability and functionality, control release of Zn2+ | Injectable dual functional antitumor and antibacterial biomedicine | [90] | |
| Gelatin–chitosan | Gel | Porphyrin Zr-MOF | Self-assembly | Direct mixing | Single responsive (NIR light) | SNP and Pt-modified porphyrin (PCN) MOF; in situ modification of gold particles and ZIF-8 | Promotes angiogenesis, collagen deposition, and suppressing inflammatory responses | Wound treatment | [91] | |
| Synthetic polymer-based hydrogel | Bicomponent hydrogels composed of melamine with salicylic acid, 3-hydroxybenzoic acid, and p-hydroxybenzoic acid | Gel | ZIF-8 | Solvothermal | In situ | Single responsive (pH) | Suppresses bacterial growth by causing cell deformation and rupture of the cell wall | Drug release | [92] | |
| Poloxamer 407 (P407) and poloxamer 188 (P188) serve as the primary hydrogel matrix, supplemented with HPMC (excipient) | Gel | ZIF-8 | One pot | Direct mixing | Multi-responsive (pH, temperature, and glucose-sensitive) | Ins@ZIF-8 (pH-responsive), Ins@ZIF-8/GO x -Gel (glucose-sensitive property); P407 and P188 (temperature responsive) | Improve the release time of the drug in vivo. Enhance the mechanical strength of the temperature-sensitive gel | Blood glucose regulation and diabetes treatment | [93] | |
| N-isopropylacrylamide (NIPAM) and a zwitterionic comonomer, [3 (methacryloylamino)propyl]dimethyl(3-sulfopropyl)ammonium hydroxide inner salt (MPDMSA) | Gel | NH2-MIL101 NPs | Microwave | Direct mixing | Single responsive (pH) | Poly(methacrylic acid-co-ethyl acrylate) (pH-sensitive); Eudragit L100-55 (transportation vehicle) | High exendin-4 loading content of more than 40% | Oral exendin-4 delivery | [94] | |
| PAM/starch hydrogel (PSH) | Gel | ZIF-8 NPs | In situ | In situ | Dual responsive (light and pH) | ZIF-8 (pH-responsive); Rose Bengal (photosensitizer drug) | Function as a smart photo-controllable topical formulation | Phototherapeutic delivery | [95] | |
| PLGA-PEG-PLGA | Gel | IRMOF-3 NPs | Solvothermal | Direct mixing | Dual responsive (Temperature and pH) | PLGA-PEG-PLGA (temperature-sensitive hydrogel) | Achieves steady delivery of dual drugs. High encapsulation capacity for Dox and Cel and enhanced chemotherapeutic efficiency for local tumor therapy | Drug delivery | [96] | |
| BP modified hyaluronic acid (HA-BP) | Gel | Zn-MOF | Direct mixing | Direct mixing | Dual responsive (ATP and pH) | Zn-MOF | Controlled release of drugs | Drug delivery | [97,98] | |
| PELA (poly(ethylene glycol)poly(ε-caprolactone-co-lactide)) | Gel | CuPP | Solvothermal | Direct mixing method (stirred for 1 h under vacuum (200 Pascal) at 120°C) | Dual responsive (Temperature and light) | PELA (poly(ethylene glycol)poly(ε-caprolactone-co-lactide)) | Unique physical property to completely seal the wound. Serve dual roles as a photothermal therapy (PTT) sensitizer and regulator of inflammation to promote normal wound healing | Wound treatment dressing | [99] | |
| PAM | Gel | Bimetallic MOF (Fe-Cu) | Hydrothermal | PAM gel and MOF(Fe-Cu)/GO x -PAM gel was freeze-dried. overnight | Single responsive (peroxide sensitive) | Bimetallic MOF(Fe-Cu) loaded with glucose oxidase | Excellent catalytic performance | Wound treatment dressing | [100] | |
| PAM | Hydrogel coating on MOF | NMOF (zirconium: aminotriphenyl dicarboxylic acid) | Direct mixing at 80°C for 5 days | Direct mixing | Single responsive (ATP) | ATP stimuli-responsive nucleic acid-based PAM hydrogel | Higher loading of DOX | Potential stimuli-responsive anticancer drug delivery system | [101] | |
| Acrylamide–poly(ethylene glycol)diacrylate (AP hydrogel) | Gel | ZIF-8 | Self-assembly | Direct mixing | Single responsive (pH) | Pd(H)@ZIF-8 | Pd(H)@ZIF-8@AP showed higher anti-inflammatory factors IL-10 | Helicobacter pylori treatment | [102] | |
| Pluronic 127 (P127) | Gel | Ni₃(HITP)₂ MOF | Ultrasonic-assisted direct mixing at 65°C | Direct mixing | Single responsive (temperature) | Pluronic 127 (P127) hydrogels | As a DDS for Ni₃(HITP)₂ MOF, | Wound therapy | [103] | |
| NiPAm/PEGMA | Microgel | MIL-101(Cr) | Hydrothermal | Direct mixing under Ar flow, from 25 to 70°C | Single responsive (temperature) | Poly(ethylene glycol)-graft-poly(N-isopropyl acrylamide) (PEG-g-PNIPAm) microgel layer | Improve the drug-delivery efficiency | Stimuli-responsive drug delivery system and biomimetic lubrication | [104] | |
| PLGA-PEG | Gel | MOF-818 | Solvothermal | Direct mixing | Single responsive (temperature) | PLGA-PEG hydrogel | Effectively adhered to would site as a wound dressing | Diabetic wound healing | [105] | |
| PPCN | Gel | HKUST-1 NPs | Direct mixing | Direct mixing | Single responsive (temperature) | Citrate-based PPCN hydrogel | Enhanced wound closure | Diabetic wound healing | [44] |
Most of the stimulus-responsive MOF–hydrogel composites are prepared via the direct mixing method and in situ synthesis method, while some used microfluidic, freeze-drying, and grinding methods (Table 2). The stimulus-responsive MOF–hydrogel composite is mainly shaped according to the shape of the MOFs since it is used to coat the outer layer of the MOFs or present as hydrogel sheet-form for wound dressing applications or drug delivery systems.
3.1 Preparation of stimulus-responsive MOF–hydrogel composites
3.1.1 Direct mixing method
Most of the reported stimulus-responsive MOF–hydrogel composites were prepared by the direct mixing method due to convenience and cost-effectiveness. In this approach, the preparation of hydrogels is altered by incorporating stimulus-responsive MOF particles into the liquid precursor or colloidal solution prior to gelation, as shown in Figure 7a. Following with gelation of the hydrogel matrices, the stimulus-responsive MOF particles are trapped within the hydrogel matrices to form stimulus-responsive MOF–hydrogel composites [106,107]. Javanbakht et al. [78] have successfully prepared a bio-nanocomposite carrier for oral delivery of the anticancer drug, 5-Fu, where 5-fluorouracil@MOF-5 NPs were coated with CMC via direct mixing method. The blend of CMC and 5-FU@MOF-5 was introduced into distilled water and stirred at room temperature. To crosslink the resulting mixture, it was gradually added dropwise into 100 mL of ferric chloride hexahydrate solution (0.01 M) [78].

Preparation of MOF–hydrogel composite via (a) direct mixing and (b) in situ synthesis.
3.1.2 In situ MOF synthesis
In situ synthesis is another commonly used preparation method for MOF–hydrogel composites whereby the hydrogel is first synthesized as the supporting material for the MOFs. The MOF crystals are grown inside the pores of the hydrogels by sequential adding in the metal precursor and organic linker to obtain the MOF–hydrogel, as shown in Figure 7b [106]. The in situ preparation method is employed to overcome the challenges associated with processing MOF particles into specifically shaped structures for practical industrial applications [108]. Zheng et al. [92] developed a porous ZIF-8 (HZIF-8) with pH responsiveness and antibacterial activity using a bicomponent hydrogel as a template through an in situ synthesis method. They transferred the transparent hot gel solution into 280 mL of preheated n-hexane, which contained 10 mL of Span 85 and stirred the mixture at 1,000 rpm for 1 h to emulsify it. The emulsion was then stirred at 0°C under mechanical stirring at 300 rpm for 1 h to induce gel formation. To the resulting gel, 20 mL of a 0.4 mol/L Hmim (1-hexyl-3-methylimidazolium) solution was added and stirred for 12 h. Afterward, an equal volume of ethanol was added to the sample to promote demulsification, and the resulting product was collected by centrifugation. The collected ZIF-8/gel was suspended in water at 70°C for 3 h, washed with methanol, and dried under vacuum at 80°C for 24 h [92].
3.2 Types of hydrogels used for stimulus-responsive MOF–hydrogel composites
The types of hydrogels used in the preparation of stimulus-responsive MOF–hydrogel composites can be classified into natural or synthetic polymers or a combination of both (Table 2). The selection of the hydrogels is mainly based on the compatibility of the MOFs with the hydrogel, the physical properties of the hydrogel such as swelling and water absorption, and the mechanical properties of the MOF–hydrogel composites, aiming to enhance the processability and scalability of the stimulus-responsive MOFs for the specific medical applications [109].
3.2.1 Natural-polymer-based hydrogel
Natural-polymer-based hydrogel is biocompatible, biodegradable, and has good cell adhesion properties. The natural polymers used in the preparation of stimulus-responsive MOF–hydrogel composites are classified into protein-based hydrogel and polysaccharide-based hydrogel. The protein-based hydrogel can be prepared using gelatin, while polysaccharide-based hydrogel can be prepared from alginate, cellulose, pectin, chitosan, and carrageenan [110].
3.2.1.1 Gelatin hydrogel
Gelatin (Figure 8a) can be derived from various natural sources and modified to produce hydrogels with excellent cytocompatibility and desirable biodegradable characteristics [111]. Gelatin-based hydrogels find applications in tissue engineering and drug delivery owing to their capacity to enhance cell adhesion and proliferation. Moreover, these hydrogels are suitable for wound dressing applications due to their impressive fluid absorbance properties [112]. For instance, Du et al. successfully synthesized a porphyrin MOF (PCN) encapsulated with photothermal-sensitive SNP modified with gold NPs to form a near-infrared (NIR) light-triggered stimulus-responsive MOF. The MOF was incorporated into a nontoxic gelatin–chitosan hydrogel via direct synthesis method [91].

General chemical structure of protein-based hydrogel used in stimulus-responsive MOF–hydrogel composites: (a) gelatin, (b) alginate, (c) pectin, (d) carrageenan, and (e) chitosan.
3.2.2 Alginate hydrogel
Alginate (Figure 8b) is a long-chain hydrophilic polysaccharide that is found in seaweeds, where it exists within the cell wall of the seaweed to provide strength and flexibility. It is a well-known natural polymer that can form hydrogel via ionotropic gelation under mild conditions [113]. Liu et al. [90] demonstrated a liquid metal micro/nanocomposite by combining eutectic gallium indium (EGaIn) LM NPs (NPs) and ZIF-8 NPs into the alginate hydrogel using an in situ synthesis approach. The resulting composites were used as light-activating nanofillers with functionalized light stimulus-responsive MOFs in the alginate-based hydrogel. The MOF–hydrogel was reported to exhibit physiological responsive sol–gel transformation, good injectability, controllable release of zinc ions, and stable photothermal performance for antitumor and antibacterial biomedicine [90].
3.2.3 Cellulose hydrogel
Cellulose is a hydrophobic polysaccharide found in natural fibers and plants such as linen and cotton. Specific bacteria such as Acetobacter xylinum can biologically synthesize cellulose [110]. Cellulose possesses unique properties such as biocompatibility, biodegradability, and insolubility in most solvents. Cellulose hydrogel has been a popular candidate for wound treatment and drug delivery applications due to its nontoxic nature [114]. Zhang et al. [86] reported a Cu@ZIF-8 MOFs biomimetic nanoreactor by assembling glucose oxidase (GO x ) onto the Cu ions doped graphitized imidazolate framework (ZIF-8), imparting glucose responsiveness to the overall Cu@ZIF-8/GO x . The resulting Cu@ZIF/GO x complex was encapsulated within a biocompatible hybrid hydrogel composed of bacterial cellulose (BC) and guar gum. The sustained release of GO x catalyzed the decomposition of glucose in the blood and triggered the in situ MOF-mediated catalytic activity by generating ˙OH radicals for bacteria eradication. Additionally, Cu@ZIF/GO x encapsulated in the hydrogel displayed a high swelling ratio for water absorption, excellent biocompatibility, and rapid hemostatic properties [86].
3.2.4 Pectin hydrogel
Pectin (Figure 8c) is an anionic, water-soluble polysaccharide that can be found in the cell walls of most plants. It can be extracted from different types of fruits through chemical or enzymatic methods. Due to its unique properties, such as biocompatibility, biodegradability, bioactivity, stabilizing, and gelling abilities, pectin has piqued the interest of researchers to further explore its medical and biomedical applications in wound healing, tissue engineering, drug delivery, and gene delivery [115,116]. Nabipour and Hu [79] synthesized a bio-MOF encapsulated with curcumin. The curcumin@bioMOF-11 was then functionalized by pH-sensitive pectin biopolymer to form a pH-responsive-MOF hydrogel and used as an anticancer drug carrier for controlled delivery [79].
3.2.5 Carrageenan
Carrageenan (Figure 8d) is extracted from a type of red algae called Irish Moss known as Chondrus crispus. It is a natural, linear, sulfated polysaccharide that exhibits unique gelling properties, with a strong negative charge, water absorption, and multiple functional groups, making it an ideal candidate for controlled and sustained drug delivery applications as well as wound healing and tissue engineering [117]. Feng et al. [89] synthesized a light and temperature-responsive hydrogel comprising poly(N-isopropylacrylamide) (PNIPAm) and methacrylated-κ-carrageenan (MA-κ-CA), incorporating PPy-polydopamine NPs (PPy-PDA NPs) and ZIF-8. The ZIF-8-hydrogel generated localized heat and gradually released Zn2+ to ensure safe and effective synergetic photothermal–chemical bactericidal capability upon application of NIR light [89].
3.2.6 Chitosan hydrogel
Chitosan (Figure 8e) is derived from chitin after undergoing the deacetylation process. It can be found as a component in the skeletons of invertebrates. It is a semicrystalline, cationic polysaccharide and is considered one of the most abundant natural biopolymers after cellulose [110]. Chitosan has great biocompatibility making it an ideal candidate in biomedical applications such as wound healing material. Chitosan hydrogel has been reported to possess mild antibacterial properties. Therefore, to improve its antibacterial property, the chitosan can be chemically modified to accurately design into a hydrogel structure [118].
Han et al. [18] prepared a photosensitive MOF–hydrogel composite composed of Prussian blue NPs (PBNPs) and chitosan using the direct mixing method. Chitosan was chemically modified with quaternary ammonium and a carbon–carbon double bond (C═C) to impart the hydrogel with an electropositive surface and robust structure. Within the hydrogel, PBNPs served as crosslinking points, absorbing polymers and reinforcing the structure, thereby enhancing the overall mechanical properties of the NIR light-responsive MOF-based hydrogel composite. Under NIR light irradiation, the stimulus-responsive MOF-based hydrogel composite exhibited antibacterial properties through the synergistic effect of heat and the electropositive surface of the chemically modified chitosan hydrogel. The presence of PBNPs facilitated NIR light absorption and light production through charge transfer between Fe3+ and Fe2+ within the PBNP structure [18].
3.3 Synthetic polymer-based hydrogel
Synthetic hydrogels have been widely used because they can be engineered to yield the desired mechanical and chemical properties for diverse ranges of products. Their mechanical strength can be modified into a slow degradation rate to provide good durability [119]. The commonly used synthetic hydrogel material for the preparation of stimulus-responsive MOF-based hydrogel composites include PAM, and block polymers such as PPCN, poly(lactic acid-co-glycolic acid) (PLGA)/poly(ethylene glycol) (PEG) (PLGA-PEG-PLGA), and N-isopropylacrylamide/ poly(ethylene glycol) dimethacrylate (NiPAm/PEGMA) (Table 2).
3.3.1 PAM hydrogel
PAM (Figure 9a) is a synthetic polymer of the acrylamide monomer. It is a colorless hydrogel material that is nontoxic, stable, non-immunogenic, and non-resorbable with hydrophilic, cohesive, viscoelastic, and biocompatible properties [120]. Tian et al. introduced a PAM hydrogel wound dressing incorporating a bimetallic MOF (Fe-Cu-MOF) loaded with GO x , termed MOF(Fe-Cu)/GO x -PAM gel. This PAM hydrogel offers an effective cascade-catalyzed system for accelerating wound healing, leveraging synergistic antibacterial and anti-inflammatory effects provided by the Fe-Cu-MOF. The authors noted that the catalytic activity of the bimetallic MOF(Fe-Cu) is approximately five times greater than that of the monometallic MOF(Fe), attributed to charge transfer from Cu(i) to Fe(iii). This accelerates the regeneration of Fe(ii) and enhances the subsequent production of hydroxyl radicals (˙OH) from the dissociation of H2O2. Within this hydrogel cascade-catalyzed system, an abundant supply of gluconic acid and H2O2 is continually generated by the breakdown of glucose through the enzyme GO x . This gluconic acid significantly enhances the peroxidase performance of MOF(Fe-Cu), enabling efficient decomposition of H2O2 to achieve effective antibacterial properties [100].

General chemical structure of synthetic polymers used in stimulus-responsive MOF–hydrogel composites: (a) PAM, (b) poly(polyethylene glycol-co-N-isopropylacrylamide) (PPCN), and (c) poly(d,l-lactide-coglycolide)-poly(ethy-leneglycol)-poly(d,l-lactide-coglycolide) (PLGA-PEG-PLGA).
3.3.2 PPCN hydrogel
PPCN (Figure 9b) is a biodegradable, temperature-responsive gel with intrinsic antioxidant properties. Temperature-responsive polymers are sensitive toward the balance of hydrophilic and hydrophobic groups as well as charge interactions, leading to a phase transition whenever there are temperature changes. PPCN can be synthesized through copolymerization by sequential polycondensation and radical polymerization by reacting citric acid, poly(ethylene glycol) (PEG), and poly-N-isopropylacrylamide (PNIPAAm) [121]. Xiao et al. [44] embedded copper-MOF NPs (HKUST-1 NPs) within an antioxidant temperature-responsive citrate-based hydrogel (PPCN). The incorporation of the HKUST-1 NPs into the hydrogel accelerated the wound healing of the diabetic mice and reduced the copper ion toxicity. The PPCN prevented the HKUST-1 NPs from disintegration, which also enabled the sustained release of copper ions without causing copper ion toxicity [44].
3.3.3 Poly(d,l-lactide-coglycolide)-poly(ethy-leneglycol)-poly(d,l-lactide-coglycolide) (PLGA-PEG-PLGA)
PLGA-PEG-PLGA (Figure 9c) is a triblock copolymer that can act as temperature-responsive copolymers, also known as thermogel. The block copolymers are composed of hydrophobic PLGA segments and hydrophilic PEG segments. The copolymer solution can form a high-viscosity gel at body temperature due to the weakening of the hydrogen bond of the PEG segments. At low temperatures, PLGA-PEG-PLGA is a free-flowing solution due to the interaction of PEG hydrophilic segments with water molecules. PLGA-PEG-PLGA has been heavily studied by researchers in clinical applications due to its biodegradability and biocompatibility [122]. Tan et al. [96] have prepared a hybrid nanocomposite in which MOFs were modified with temperature-sensitive PLGA-PEG-PLGA hydrogel to devise an injectable implant. DOX and celecoxib (Cel) were co-loaded into the MOF–hydrogel system for localized oral cancer therapy (DOX/Cel/MOFs@Gel). This formulation exhibited a high drug loading capacity, stable and pH-responsive release of both drugs, and enhanced toxic effects against oral cancer cells [96].
3.3.4 N-isopropylacrylamide/poly(ethylene glycol) dimethacrylate (NiPAm/PEGMA)
Poly(N-isopropylacrylamide) (PNiPAm) polymer-based hydrogels are recognized for their temperature-responsive behavior, particularly in proximity to a lower critical solution temperature (LCST) [123]. Hence, it demonstrates excellent lubricating performance and temperature-sensitive tribological properties [124]. Furthermore, by integrating hydrophilic polymer chains like PEG into the PNiPAm matrix, the LCST of the PNIPAm microgels can be elevated [123]. PEG is used for surface modification to enhance colloidal stability as well as reduce cytotoxicity in drug delivery systems [125]. Wu et al. [104] reported a MIL-101(Cr)@PEG-g-PNIPAm hybrid by growing the poly(ethylene glycol)-graft-poly(N-isopropylacrylamide) (PEG-g-PNIPAm) microgel layer using NIPAm, PEGMA, and N,N′-Methylenebisacrylamine (MBA) as the hydrogel materials on the MIL-101(Cr) surface with 2,2′-azobis(2-methyl-propionamidine) dihydrochloride (V50) to initiate polymerization via a one-pot soap-free emulsion method. At the osteoporosis site (∼40°C), the temperature exceeds the LCST of the MIL-101(Cr)@PNIPAm, which compromises its lubricating performance and affects drug release. Additionally, the colloidal stability of MIL-101(Cr)@PNIPAm is influenced by the transition of PNIPAm chains from hydrophilic to hydrophobic above the LCST. Therefore, the newly synthesized MIL-101(Cr)@PEG-g-PNIPAm hybrid, with the addition of PEG, elevates the LCST and stability of the microgel by creating a robust hydration layer around the ethylene glycol groups through hydrogen-bonding interactions with adjacent water molecules [104].
4 Characterization of stimulus MOF-based hydrogels
The resulting MOF–hydrogel composite prepared should be characterized with diverse physicochemical characterization techniques to determine the crystal structure, and physicochemical and thermal properties of the MOF and MOF–hydrogel composite.
4.1 FTIR spectroscopy analysis
FTIR is used to determine the functional groups and chemical bonds in the MOFs and hydrogel. The FTIR characteristic peaks of the commonly used MOFs and hydrogel are summarized in Table 3. Shifting in characteristic peaks is normally observed when MOFs are incorporated into the hydrogel. Nabipour and Hu [79] reported using FTIR spectroscopy to identify the functional groups of a pectin/curcumin@bio-MOF-11 hydrogel used for colon drug delivery. They compared the FTIR spectra of the pure pectin and the pectin/curcumin@bio-MOF-11. The FTIR spectra revealed that in the region of 400–800 cm−1, the Ca–O band was intensified, indicating successful crosslinking between the pectin polymeric chains and Ca2+ ions within the pectin/curcumin@bio-MOF-11 [79]. Conversely, Reddy et al. [95] reported that during the synthesis of nZIF-8@PAM/starch hydrogel via in situ synthesis, two prominent peaks at around 841 and 1,095 cm−1, corresponding to Zn(OH) bending vibrations, were observed when zinc salt was directly added into the acrylamide hydrogel. This indicated the formation of Zn(OH)2 precursor used for in situ MOF growth in the gel [95].
Common FTIR characteristic peaks of the selected MOFs and hydrogel materials
| Subjects | Characteristic peaks | Ref |
|---|---|---|
| Bio-MOF-11 |
|
[79] |
| MOF-5 |
|
[80] |
| HKUST-1 |
|
[44] |
| ZIF-8 |
|
[92,93] |
| UiO-66 |
|
[82] |
| Pectin |
|
[87] |
| Carrageenan |
|
[82] |
| Alginate |
|
[108] |
| Cellulose |
|
[114] |
| Chitosan |
|
[118] |
| PPCN |
|
[119] |
| PAM |
|
[120] |
4.2 Scanning electron microscopy (SEM) for morphology determination
The SEM characterization is often used for the determination of MOF–hydrogel morphology. High-resolution two-dimensional (2D) image generated by SEM displays the shapes and spatial variations of the hydrogel and the MOF crystals, thus providing information about the external morphology, phase mixture, and dispersion [126].
Chen et al. [101] utilized SEM to study the morphology of the UiO-68 NP MOFs, which had been surface-modified with an ATP-responsive nucleic acid-based PAM hydrogel. Figure 14c shows the SEM images of UiO-68 NP MOF carriers capped by the hybridization of [1] nucleic acid with [2] promoter to form duplex locking units. The particles revealed a bipyramidal structure with particle size ranges of 280–350 nm. Figure 14d shows the SEM image of UiO-68 NMOFs coated by a DNA crosslinked hydrogel. The “sharp” edges of the UiO-68 NMOFs were depleted after the hydrogel was coated on the UiO-68 NP MOFs, and the smooth surfaces of the particles were observed. The particle size of the MOF increased slightly with diameters in the range of 400–450 nm, indicating that the hydrogel was successfully coated onto the UiO-68 NMOFs [101].
4.3 Transmission electron microscopy (TEM)
TEM has been extensively utilized to determine the particle size, grain size, and crystallographic data of MOFs. This technique is particularly valuable for characterizing MOFs modified by incorporating NPs as it offers insights into the dispersion and size of the NPs [126]. Zhu et al. [87] conducted micromorphology characterization using TEM of the ZIF-8 NPs (SP@ZIF-8-PEG-TK) with surface modification with PEG-TK (Figure 10a). The SP@ZIF-8-PEG-TK NPs exhibited responsive release under stimulation by ROS. The SP@ZIF-8-PEG-TK NPs were observed to be rhombic and uniformly dodecahedral, resembling the blank ZIF-8 NPs (Figure 10b). These similar morphological characteristics suggest that the encapsulation of ZIF-8 into NPs did not affect the morphology of the MOF [87].
![Figure 10
TEM images of (a) ZIF-8 and (b) SP@ZIF-8-PEG-TK after surface modification with PEG-TK hydrogel. Reuse under terms of the CC-BY license [87]. Copyright 2016 Dove Medical Press. (c) MIL-101(Cr) NPs and (d) MIL-101(Cr)@PEG-g-PNIPAm hybrid after surface modification with poly(ethylene glycol)-graft-PNIPAm. Reuse under terms of the CC-BY license [104]. Copyright 2023 Elsevier.](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_010.jpg)
TEM images of (a) ZIF-8 and (b) SP@ZIF-8-PEG-TK after surface modification with PEG-TK hydrogel. Reuse under terms of the CC-BY license [87]. Copyright 2016 Dove Medical Press. (c) MIL-101(Cr) NPs and (d) MIL-101(Cr)@PEG-g-PNIPAm hybrid after surface modification with poly(ethylene glycol)-graft-PNIPAm. Reuse under terms of the CC-BY license [104]. Copyright 2023 Elsevier.
Additionally, Wu et al. [104] also reported on the morphological characterization of MIL-101(Cr) NPs and surface-modified MIL-101(Cr) NPs with poly(ethylene glycol)-graft-PNIPAm hydrogel using TEM. Figure 10c shows the morphology of MIL-101(Cr) NPs to be round dodecahedral. Figure 10d shows the surface-modified MIL-101(Cr) NPs demonstrated a more rounded in shape morphological structure [104].
4.4 Powder X-ray diffraction (PXRD) for structural and crystallinity determination
X-ray diffraction is used to deduce if single crystals or polycrystals are dominant in an MOF sample by using diffraction data with a mathematical adjustment, such as the non-linear least squares method and Scherrer’s equation to obtain the crystalline size once the diffraction peaks are identified from the non-linear least squares method [127]. Structural determination can be done by comparing the diffractogram of the synthesized MOFs with a reported one in literature or a simulated pattern generated by single crystal X-ray or computational modeling [128]. Figure 11a shows the PXRD patterns of PAM/starch hydrogel (PSH), ZIF-8, and nZIF-8 @PAM/starch hydrogel (ZPSH), which were prepared by the in situ synthesis method.
![Figure 11
X-ray diffraction patterns of (a) PAM/starch hydrogel (PSH), (b) ZIF-8, (c) nZIF-8 @PAM/starch hydrogel (ZPSH). Adapted with permission from Reddy et al. [95]. Copyright 2023 American Chemical Society.](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_011.jpg)
X-ray diffraction patterns of (a) PAM/starch hydrogel (PSH), (b) ZIF-8, (c) nZIF-8 @PAM/starch hydrogel (ZPSH). Adapted with permission from Reddy et al. [95]. Copyright 2023 American Chemical Society.
The significance of crystallinity becomes more pronounced concerning the physical attributes, wherein the interaction between ZIF-8 nanocrystals and the hydrogel plays a crucial role in both substance release and stability within the hydrogel environment. The XRD patterns revealed that the well-defined crystalline structure of the ZIF-8 nanocrystals within the nZIF-8@PAM/starch hydrogel composite remained intact (Figure 11c). Specifically, the XRD pattern of the nZIF-8@PAM/starch hydrogel composite exhibited distinct ZIF-8 crystal peaks that perfectly matched the anticipated pattern for ZIF-8 crystals (Figure 11b). In contrast, the XRD pattern of the PAM/starch hydrogel (Figure 11a) displayed a broad and non-crystalline curve, indicating its amorphous nature. The amorphous characteristic was also noted in the XRD pattern of the nZIF-8 PAM/starch composite hydrogel, suggesting that the introduction of ZIF-8 nanocrystals did not disrupt the overall amorphous nature of the hydrogel [95]. This characterization technique is extensively used for understanding the crystallization of MOF before and after its incorporation into hydrogel matrices.
4.5 Thermogravimetric analysis (TGA)
TGA is used to measure the thermal stability of MOFs and MOF–hydrogel composites after the incorporation of MOFs into the hydrogel. The decomposition of MOFs will vary depending on the carrier gas that will be used for TGA. The carrier gas used is usually N2, air, or O2 [129]. The decomposition of MOF usually occurs in several distinct steps. MOF is known to start decomposition when desolvation occurs, usually occurring at 150°C. After that, the TGA thermogram usually shows a plateau region where the solvent-free evacuated MOF is stable, until the temperature where the MOF starts to degrade and the TGA thermogram will reflect a secondary mass loss event [129]. The addition of MOFs into hydrogel might affect the thermal stability of the MOF–hydrogel composite. For instance, Nabipour and Hu [79] reported on a pH-sensitive pectin biopolymer used to functionalize curcumin@bio-MOF-11 to form a bio-nanocomposite hydrogel bead for drug delivery. TGA of both bio-MOF-11 and curcumin@bio-MOF-11 showed the degradation and structural collapse of the organic linkers between 270 and 390°C, whereby bio-MOF-11 undergoes structural degradation at approximately 390°C. As for curcumin@bio-MOF-11, it exhibited three distinct weight losses at approximately 250, 310, and 430°C. The initial weight loss corresponded to the phenolic compounds present in curcumin, the second weight loss indicated the breakdown of the organic linkers, and the final weight loss signified the degradation of the porous solid material [79].
4.6 Water content and hydrogel swelling studies
The determination of the water content of the hydrogel is important for controlled drug release and delivery. This is because water content influences the overall hydrogel integrity, solubility, and diffusion kinetics of substances [130]. The water content is estimated by measuring its insoluble part in the dried sample after immersion in deionized water. The gel fraction is then measured using the following equation:
where W i is the initial weight of the dried sample and W d is the weight of the dried part sample after extraction with water.
where W d and W s are the weights of the dried sample and swollen sample, respectively.
The swelling rate is measured by using free-absorbency capacity measurements at consecutive time intervals [131]. Javanbakht et al. [80] studied the effect of the synthesized CMC-coated 5-fluorouracil@MOF-5 nanohybrid on its swelling properties. The CMC-coated 5-fluorouracil@MOF-5 nanohybrid was reported to be pH-responsive. At a pH of 1.2, the carboxylate anions presented in CMC exist in a protonated form and lead to the formation of hydrogen-bonding interactions among the carboxylic acid groups of CMC. These hydrogen bonds act as physical crosslinkers, causing the CMC polymer network to contract or shrink. As a result, the swelling values of CMC were reduced in an acidic environment. However, as the pH was increased to 6.8 and 7.4, the carboxylic groups in the CMC network deprotonated leading to an increase in electrostatic repulsion between the carboxylate anions, thus resulting in higher swelling of the CMC network. Therefore, this pH-responsive CMC is a good protective layer to prevent the early release of 5-Fu from MOF-5 [80].
4.7 Rheology
Hydrogel of various types of polymers could mimic the soft tissue properties that vary greatly within the human body. Therefore, it is important to characterize the rheology of the polymer hydrogels. Rheology measurement assesses the extent to which a hydrogel can respond to stress by either absorbing energy (storage modulus) or undergoing stress relaxation to dissipate energy (loss modulus) [132]. Reddy et al. [95] reported on the rheological analysis of the synthesized PAM/starch hydrogel (PSH) and nZIF-8@PAM-starch hydrogel (ZPSH) using a rheometer at 25°C. When ZPSH was subjected to shear strain, it was observed that the shear strain (γc) of the ZPSH was higher compared to PSH, suggesting that the integration of nZIF-8 into the polymer network had improved the mechanical strength of PSH [95].
4.8 Mechanical strength test of the MOF–hydrogel composite
Compressive strength and tensile strength are generally used to measure the mechanical properties of hydrogels. Compressive strength refers to the ability of a certain material or structural element to withstand loads that reduce its size when force is applied. Conversely, tensile strength is the amount of load or stress that a material can withstand until it stretches and breaks [133]. The compressive strength and tensile strength of the MOF-based hydrogel are generally tested by a universal testing machine. Wang et al. reported on a donut-like MOF composed of copper–nicotinic acid (CuNA) synthesized via a solvothermal method that is incorporated into gelatin methacrylate (GelMA) hydrogel [134]. The compression properties of GelMA and CuNA@GelMA composites were evaluated using a universal machine tester. The trend indicates that the elasticity of CuNA@GelMAs was significantly superior to that of pure GelMA. When applied to a compression of 90%, the CuNA@GelMAs with different percentages of CuNA (3, 5, and 10%) incorporated into GelMA recovered into their initial state after the compression was removed. However, pure GelMA fractured when 90% strain was applied indicating that the incorporation of CuNA enhanced the mechanical properties of GelMA. This is due to the ionic crosslinking effect between copper ions of the MOFs and GelMA molecules whereby this effect enhanced the overall structure of the composite hydrogels, leading to improved mechanical strength and elasticity.
5 Application of stimulus-responsive MOF–hydrogels in medical treatments
Stimulus-responsive MOF–hydrogels combine the unique properties of MOFs and hydrogels, creating versatile and adaptive platforms capable of responding to external stimuli such as temperature, pH, light, or specific chemical triggers, opening the door to a wide array of applications in the field of medicine, offering drug delivery, on-demand therapeutic release, and wound management. Most of the reported stimulus-responsive MOF–hydrogel composites derived from biological or biocompatible materials for medical treatments were meant for wound management and drug delivery systems owing to the responsiveness of the MOF–hydrogel composites and the physicochemical properties of the hydrogels (Table 2). The stimulus-responsive MOF–hydrogels in medical treatments are categorized into single-responsive, dual-responsive, and multi-responsive. These variations in responsiveness have been designed to address the demands of critical applications such as precise drug delivery and effective wound management.
5.1 Drug delivery system
A drug delivery system plays an essential role in managing both the rate and the targeted sites of drug release within the body. The concept of controlled drug release involves employing a specialized vehicle that encapsulates the drug. This encapsulation enables the systematic and quantitative release of the drug, maintaining a consistent rate of delivery. The goal of controlled drug release is to uphold a stable concentration of the drug within the bloodstream, thereby optimizing therapeutic efficacy. Stimulus-responsive MOF–hydrogels have been reported to be used as carriers for drug delivery due to their responsive behavior enabling precise control over drug release, ensuring targeted delivery to the desired site of action within the body. The reported MOF–hydrogels were mainly used in the delivery of bioactive natural polyphenolic compounds or oncology drugs (Table 2).
5.1.1 Single-responsive MOF–hydrogels
The reported single-responsive MOF–hydrogels were mostly pH-responsive MOF–hydrogels for controlled release of ibuprofen, fluorouracil, exendin-4, itaconate, and natural polyphenolic compounds, such as curcumin. Nabipour and Hu [79] reported an MOF, namely bioMOF-11, synthesized from cobalt acetate tetrahydrate and adenine. The bioMOF-11 was encapsulated with curcumin followed by functionalizing it using a pH-sensitive pectin biopolymer to protect the drug formulation against the harsh conditions of the gastrointestinal tract (GIT). The release of 100% curcumin from curcumin@bio-MOF-11 was observed within 90 min due to the decomposition of bio-MOF-11 at pH 1.2. When curcumin@bio-MOF-11 was prepared in pectin hydrogel composite form, the controlled release rate was found to be 93% over 410 min at different pHs of 1.2, 6.8, and 7.4. The observed improvement in release behavior was attributed to the pH sensitivity of pectin, as shown in Figure 12. Furthermore, pectin functions as a protective layer, aiding in regulating the release of curcumin [79].
![Figure 12
Curcumin release profile of cucurmin@bio-MOF-11 and pectin/cucurmin@bio-MOF-11 hydrogel. Reuse under terms of the CC-BY license [79]. Copyright 2022 Springer Link.](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_012.jpg)
Curcumin release profile of cucurmin@bio-MOF-11 and pectin/cucurmin@bio-MOF-11 hydrogel. Reuse under terms of the CC-BY license [79]. Copyright 2022 Springer Link.
Javanbakht et al. reported on κ-Cr and CMC used for surface modification on three different MOFs, namely, UiO-66, Cu-MOF, and MOF-5 [82]. It was reported that the drug release of carrageenan/tramadol@UiO-66 at pH 1.2 was the lowest because of the attraction between UiO-66 and tramadol at acidic pH, which hindered the drug release. However, as the pH increased to 7.4 (simulating the intestinal fluid), the drug release rate dramatically increased. Ion-exchange between the buffered solution
In another study, Wu et al. reported thermosensitive microgel nanocarriers of PEG-g-PNIPAm copolymer functionalized on nanoMOFs (MIL-101 (Cr), namely MIL-101(Cr)@PEG-g-PNIPAm) [104]. This thermosensitive MIL-101(Cr)@PEG-g-PNIPAm-hydrogel composite was loaded with DS, aiming for the treatment of osteoarthritis. It was found that an initial rapid drug release was observed in both DS-MIL-101(Cr) and DS-MIL-101(Cr)@PEG-g-PNIPAm within the first 4 h, followed by a gradual release over time (Figure 13). After 72 h, DS-MIL-101(Cr) released approximately 67.4% of the drug at 37°C, with a similar release profile at 32°C. In contrast, DS-MIL-101(Cr)@PEG-g-PNIPAm exhibited reduced drug release, with approximately 43.1% released at 32°C and 36.6% at 37°C. This reduction was attributed to drug loss during the polymerization and washing processes. As the temperature increased to 42°C, the cumulative drug release of the DS-MIL-101(Cr)@PEG-g-PNIPAm further decreased to 31.7%, likely due to the partial collapse of PEG-g-PNIPAm microgel layers, which restricted drug release from the MOFs. Therefore, improving the anti-inflammatory drug-delivery efficiency in a controlled manner. Overall, the investigation revealed the temperature-dependent drug release profile of these materials, demonstrating the potential of DS-MIL-101(Cr)@PEG-g-PNIPAm as a controlled drug delivery system influenced by temperature variations.
![Figure 13
(a) Release profile of DS-MIL-101(Cr) in PBS at 32, 37, and 42°C. (b) DS-MIL-101(Cr)@PEG-g-PNIPAm in PBS at 32, 37, and 42°C. Reuse under terms of the CC-BY license [104]. Copyright 2023 Elsevier.](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_013.jpg)
(a) Release profile of DS-MIL-101(Cr) in PBS at 32, 37, and 42°C. (b) DS-MIL-101(Cr)@PEG-g-PNIPAm in PBS at 32, 37, and 42°C. Reuse under terms of the CC-BY license [104]. Copyright 2023 Elsevier.
Apart from the physical stimulus responsiveness, Chen et al. [101] developed an ATP-responsive nucleic acid-based PAM hydrogel functionalized on DOX-loaded UiO-68 nano-MOF (Figure 14). In the presence of ATP, which is known to be overexpressed in cancer cells, the nucleic acid-based PAM hydrogel coating underwent degradation through the formation of the ATP–aptamer complex. This degradation process led to the release of the DOX. The resulting hydrogel-coated UiO-68 nanoMOFs also exhibited a high degree of drug loading with a loading capacity of DOX at 79.1 nmol mg−1 and low background leakage (∼28%) of the drug loads as compared to the non-hydrogel-coated nanoMOFs. The cytotoxicity of the DOX-loaded hydrogel-coated UiO-68 nanoMOFs toward cancer cells was substantially higher (∼40%) as compared to the cytotoxicity of the non-hydrogel-coated nanoMOFs toward cancer cells (∼25%), indicating the efficiency of the nucleic acid-based PAM hydrogel.
![Figure 14
(a) Illustration of the synthesis of ATP-responsive DNA@PAM hydrogel NMOFs loaded dye/drugs. (b) SEM image of the nucleic acid with promoter-functionalized UiO-68 NMOFs before the deposition of the hydrogel. (c) SEM image of the hydrogel-coated NMOFs. Reuse under terms of the CC-BY license [101]. Copyright 2017 John Wiley and Sons.](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_014.jpg)
(a) Illustration of the synthesis of ATP-responsive DNA@PAM hydrogel NMOFs loaded dye/drugs. (b) SEM image of the nucleic acid with promoter-functionalized UiO-68 NMOFs before the deposition of the hydrogel. (c) SEM image of the hydrogel-coated NMOFs. Reuse under terms of the CC-BY license [101]. Copyright 2017 John Wiley and Sons.
5.1.2 Dual- and multi-responsive MOF–hydrogels
Dual-responsive MOF–hydrogels consist of the combination of two stimuli to enhance the effectiveness of its applications. The reported dual stimulus-responsive MOF–hydrogels are thermo- and photo-responsive, thermo- and shape-memory-responsive, thermo- and enzyme-responsive, or shape-memory and water-responsive hydrogels (Table 2).
Zeng et al. reported on an injectable pH and ATP-responsive MOF-based hydrogel composite, which possessed structurally dynamic properties such as shear-thinning and self-healing ability [97]. This stimulus-responsive MOF-based hydrogel composite was proposed to be a local depot for the sustained release of DOX drugs at tumor sites. The DOX-loaded MOF-based hydrogel composite was prepared by using a bisphosphonate (BP)-functionalized hyaluronic acid (HA-BP) macromolecule, which was mixed with a suspension of DOX-embedded Zn-MOF (MOF@DOX). The hydrogel network formed through dynamic coordination bonds between the BP group of HA-BP and Zn2+ ions on the MOF surface. In the presence of protons, these coordination bonds between Zn2+ ions and the imidazole groups in the MOF were cleaved, leading to drug release as the MOF framework collapsed. Additionally, ATP exhibited stronger coordination with Zn2+ ions than the imidazole groups, allowing it to competitively bind with Zn2+ ions, further promoting the collapse of the MOF and accelerating DOX release (Figure 15). In comparison to the HA-BP·MOF@DOX delivery system, MOF@DOX demonstrated faster drug release kinetics. The HA-BP·MOF@DOX hydrogel exhibited multiple stimuli-responsive drug release performances. For instance, only 2.6% of DOX was released from the HA-BP·MOF@DOX gel in 97 h in PBS of pH 7.4. In contrast, within the same period, 5.7, 13.4, 10.2, and 26.9% of DOX was released from HA-BP·MOF@DOX in mediums with pH 6.0, pH 5.0, pH 7.4 + 2 mmol/L ATP, and pH 5.0 + 2 mmol/L ATP, respectively. These data suggested that the presence of protons in the medium can cleave the coordination bond between Zn2+ ions and the imidazole group in the MOF, leading to drug release after framework collapse. Additionally, ATP exhibited stronger coordination with Zn2+ ions than the imidazole group, thereby competitively binding with Zn2+ ions and expediting MOF collapse and DOX release.
![Figure 15
Mechanism of drug release after the cleavage of coordination bond between Zn2+ ions and the imidazole group in the MOF [97].](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_015.jpg)
Mechanism of drug release after the cleavage of coordination bond between Zn2+ ions and the imidazole group in the MOF [97].
Moreover, compared to the HA-BP·MOF@DOX delivery system, MOF@DOX showed faster drug release kinetics. In 97 h, 4.7, 66.3, 33.1, and 100% of DOX was released from MOF@DOX in mediums with pH 7.4, pH 5.0, pH 7.4 + 2 mmol/L ATP, and pH 5.0 + 2 mmol/L ATP, respectively. The presence of particle-hydrogel superstructures in HA-BP·MOF@DOX inherently slowed down the degradation of MOF@DOX due to changes in proton and ATP molecules of osmotic conditions. This characteristic would prolong DOX release in the tumor microenvironment, even under low pH conditions and high ATP levels.
Reddy et al. [95] reported on a dual-responsive MOF–hydrogel prepared by the in situ growth of a nano-MOF (nZiF-8) embedded into a biocompatible PAM/starch hydrogel (PSH) composite. Photosensitizer, Rose Bengal (RB) (4,5,6,7-tetrachloro-2′,4′,5′,7′-tetraiodofluorescein), was loaded into the nZIF-PSH, and the interaction involving surface adsorption was observed. RBZIF-8 and RB-loaded PSH exhibited a burst release profile, rapidly releasing a significant portion of RB at a pH of 5.5 (∼90% drug release). In contrast, RBZPSH displayed a sustained release pattern over 18 h. This sustained release from RBZPSH was attributed to the π–π interactions between the imidazole ring of the nZIF-8 structure and the phenyl rings of RB embedded within the IPN of PSH. Additionally, ionic interactions between the halogen atoms (chlorine and iodine) of RB and zinc atoms within the nZIF-8 also contributed to the sustained release of RB from ZPSH. RBZPSH demonstrated significantly greater photodynamic antifungal activity compared to its monometallic counterparts [95].
Liu et al. reported on MOF-based injectable in situ gel used for multi-responsive insulin delivery [93]. The Ins@ZIF-8/GO x -gel was prepared via a one-pot synthesis method. The Ins@ZIF-8/GO x -gel exhibited several responsive characteristics. It displayed thermosensitivity as an in situ gel. When introduced into the body, the Ins@ZIF-8/GO x -gel solution transitioned into a semi-solid state due to the poloxamers that form the hydrogel. This is because when the temperature rises above the LCST, the poloxamers become insoluble and undergo phase separation from the solution, thus forming a hydrogel. In this process, the hydrogen bonding between the poloxamers and solvent is weakened, thereby leading to partial dehydration and aggregation of polymeric chains. As the temperature increases, the gradual removal of water molecules will reveal the hydrophobic domains of the polymer macromolecules, thereby promoting the formation of hydrophobic interactions between polymer macromolecules, known as the hydrophobic effect.
As the temperature surpassed the gelation threshold over time, the gel degraded due to interaction with body fluids, gradually releasing both GO x and Ins@ZIF-8. Additionally, the ZIF-8 component of the composite possessed pH sensitivity. In hyperglycemic conditions, GO x and glucose interaction led to the production of gluconic acid, resulting in a localized decrease in pH. While in this acidic environment (pH 5.0), the structure of ZIF-8 collapsed, facilitating the release of insulin. Ins@ZIF-8/GO x -gel was proposed to be an injectable hydrogel that is responsive to temperature and pH changes, which enabled the controlled and targeted release of insulin, making it a versatile candidate for diabetes treatment and blood glucose regulation.
5.2 Wound healing
Chronic non-healing wounds are one of the most significant impediments due to bacterial infection and have always remained a huge fundamental healthcare concern. Nonpharmacological and nonbiological strategies such as phototherapy are used in wound management because the regulatory cost and development time are significantly reduced if a medical therapy product is classified as a medical device. However, the effectiveness of phototherapy in wound treatment can be significantly reduced by pre-existing hypoxic microenvironments and biofilms. Incorporating stimulus-responsive MOF–hydrogels into wound care strategies holds great promise for enhancing the efficacy of treatments, reducing complications, and expediting the healing process due to controlled drug release, moisture retention, and antibacterial activity provided by MOFs that incorporate antibacterial ions or molecules, such as zinc ions (Zn2+), which are toxic to bacterial cells. The reported MOF–hydrogel composites used in wound management can be categorized into single-responsive and dual-responsive MOF–hydrogel, in which the reported single-responsive MOF–hydrogel composites are responded toward chemical or physical stimuli (Table 2).
5.2.1 Single-responsive MOF–hydrogels toward physical stimuli
The reported single-responsive MOF–hydrogels for wound healing were mainly aiming for photothermal therapies, where the MOF–hydrogels were proven to have their light or thermal responsiveness. The light responsiveness of the reported MOF–hydrogels was contributed by light-responsive chemicals, such as SNP, Prussian blue, or EGaIn being loaded or integrated into MOF. The hydrogel played the role of carrier matrix, some with injectable properties to aid in the delivery of the active ingredients and coverage of the wound. Some studies reported that the MOFs possessed antibacterial properties contributing to the wound healing properties of the MOF–hydrogel.
Du et al. developed a photothermal-sensitive SNP loaded into a Pt-modified porphyrin MOF (PCN) with gold particles modified in situ to form an NIR light-triggered phototherapeutic nano platform whereby it converted light into heat to kill bacteria [91]. The system was then incorporated into a hydrogel composed of gelatin and chitosan to form a multifunctional injectable hydrogel, known as PSPG hydrogel. The PSPG hydrogel exhibited remarkable catalase-like behavior, promoting the continuous decomposition of endogenous H2O2 into O2 provided by the platinum metal, which is a catalyst for in situ oxygen production due to its long-lasting catalytic and antioxidant properties. This enhanced the photodynamic therapy effect, specifically under hypoxic conditions. When subjected to NIR irradiation, the PSPG hydrogel generated hyperthermia and simultaneously produced ROS from the Pt-modified porphyrin MOF (PCN) while triggering the release of nitric oxide. These combined effects contributed to the removal of biofilms and disruption of the cell membranes of methicillin-resistant Staphylococcus aureus (MRSA) and Escherichia coli (E. coli) bacteria. The study evaluated the photothermal effect of PSPG hydrogel under 808 nm NIR laser irradiation, showcasing a rise in temperature with increasing concentrations of PCN. Notably, the 250 μg/mL hydrogel achieved a temperature of 65.3°C within 10 min, sufficient for bacterial eradication. The photothermal conversion ability of PSPG hydrogels was observed to increase with higher laser powers, demonstrating excellent photostability even after ten cycles of irradiation. The efficiency (η) of the PSPG hydrogel, calculated at 89.21%, surpassed that of PCN (85.73%), possibly due to enhanced material stability conferred by hydrogel formation. Furthermore, quantification of NO release from PSPG hydrogels revealed a positive correlation with hydrogel concentration and irradiation time, with significantly higher NO release observed under NIR irradiation compared to control conditions.
In another work, Han et al. [18] employed quaternary ammonium and double-bond modified chitosan (QCH) and an MOF NP, i.e., Prussian blue MOF NPs (PBNPs) produced by solvothermal synthesis using K4[Fe(CN)6] and FeCl3. The PBNPs were embedded with the modified chitosan hydrogel to form a photo-responsive hydrogel via free radical polymerization and named QCH/PB. The PBNPs were capable of absorbing 808 nm NIR light and generated light through the charge transfer between Fe3+ and Fe2+ ions within the structure, leading to an increase in the temperature of the QCH/PB hydrogel. Additionally, the modified chitosan hydrogel formed via quaternary ammonium and a carbon–carbon double bond (C═C) in the QCH/PB hydrogel exhibited strong electrostatic absorption, allowing them to capture bacteria and disrupted the surface potential of the bacterial membrane, inhibiting the normal metabolism of the bacteria and consequently suppressing bacterial respiration leading to an antibacterial ratio of up to 99.97% against Staphylococcus aureus and 99.93% against E. coli. As a result, the combination of photothermal effects from the Prussian blue MOF NPs and the inhibition of bacterial respiration provided by the modified chitosan hydrogel led to highly effective and rapid bacteria eradication [18]. These two studies highlighted the incorporation of photo-responsive MOFs into hydrogels to enhance bacteria eradication by generating hyperthermia upon NIR-light irradiation.
Liu et al. [90] developed a liquid metal micro/nanocomposite through the integration of EGaIn liquid metal NPs (LMNPs) and ZIF-8 MOFs within an alginate hydrogel. The resulting composites LMNPs@MOFs SupraParticles (LMSPs) were used as light-activating nanofillers with a functionalized light stimulus-responsive MOFs within the alginate-based hydrogel. The LMSPs/Alg hydrogel was prepared by injecting an LMSPs/Alg solution into phosphate-buffered saline containing CaCl2 to mimic gelation in the physiological environment, resulting in the formation of LMSPs/Alg-Ca2+ hydrogel upon injection. The LMNPs within the LMSPs enabled precise photothermal therapy (PTT) in the NIR window, providing localized heating for deep tissue penetration. The in vivo formation of the Alg-Ca2+ hydrogel at the injection site entrapped the therapeutic LMSPs, which generated localized heat around the nidus, thereby enhancing the efficacy of the PTT. Additionally, ZIF-8 NPs exhibited strong antibacterial effects owing to the toxicity of Zn2+ toward bacterial cells. Hence, the MOF shell within LMSPs underwent degradation, releasing bioactive ions (Zn2+) to facilitate a combination of chemo/PTT. Moreover, the hydrogels effectively maintained moisture in proximity to the heated wounds, which played a crucial role in expediting the healing of skin wounds through PTT [90].
A temperature-responsive MOF hydrogel has been gaining attention in the field of wound management due to its capability of undergoing temperature-driven phase transitions. Xiao et al. [44] synthesized a copper-MOF NP (HKUST-1 NPs) and embedded it within a thermoresponsive PPCN hydrogel and is named H-HKUST-1. The HKUST-1 and PPCN composite possessed a synergistic effect, where HKUST-1 did not interfere with the PPCN gelation by shielding the positive charge of the stored copper ions, and PPCN protected HKUST-1 from degradation. Hence, copper was released from H-HKUST-1 in a controlled manner, reducing copper ion toxicity. It was hypothesized that the copper ions of the HKUST-1 undergoing coordination with the carboxyl groups present in H3BTC molecules during the synthesis process. This coordination process effectively decreased the interactions between these carboxyl groups and those within PPCN, thus having limited influence on the gelation process. As a result of these interactions and coordination, H-HKUST-1 formed a stable hydrogel with only minor effects on the original LCST of PPCN. A comparison was made by treating diabetic mice with wound incisions between PPCN and H-HKUST-1 hydrogel. Figure 16 illustrates the quantitative analysis of digital images of wounds over time, demonstrating that wounds treated with H-HKUST-1 healed significantly faster than those treated with PBS, PPCN, and HKUST-1 NPs, particularly on days 7, 14, 21, and 29. Specifically, diabetic mice treated with H-HKUST-1 hydrogel exhibited accelerated healing compared to those treated with PBS, blank PPCN, and HKUST-1 NPs. The authors reported that the H-HKUST-1 hydrogels showed a higher wound closure rate due to the sustained release of noncytotoxic amounts of copper ions. This release promoted angiogenesis, collagen deposition, and re-epithelialization during wound healing, enhancing the overall healing process [44].
![Figure 16
The quantitative analysis of wound closure rates of wounds treated with PBS, PPCN, and HKUST-1 NPs and H-HKUST-1. Reuse under terms of the CC-BY license [106] Copyright 2016 John Wiley and Sons.](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_016.jpg)
The quantitative analysis of wound closure rates of wounds treated with PBS, PPCN, and HKUST-1 NPs and H-HKUST-1. Reuse under terms of the CC-BY license [106] Copyright 2016 John Wiley and Sons.
Wang et al. [103] reported Ni3(HITP)2 MOF nanorods synthesized via ultrasonic-assisted direct mixing (65°C) method using nickel acetic acid tetrahydrate, sodium acetate, and 2,3,6,7,10,11-hexaaminotriphenylene hexahydrochloride. The Ni3(HITP)2 MOF nanorods with antioxidant and anti-inflammatory effects, being encapsulated within a temperature-sensitive, nontoxic, and phase-changing synthetic hydrogel, Pluronic F127, was applied as an injectable and sprayable medical dressing. Pluronic F127 displayed distinct transitions between solution and gel states in an aqueous solution. In this work, the F127 hydrogel was formulated with a phase transition temperature of 21°C, considering the skin’s surface temperature is above 21°C. When the temperature exceeded this threshold, the aqueous solution transformed into a gel-like state to cover the wound evenly for a better wound-healing effect. Conversely, the solution reverted to its original state when the temperature dropped below 21°C. The Ni3(HITP)2/F127 hydrogel exhibited superoxide dismutase-like activity and anti-inflammatory properties, as well as the ability to promote angiogenesis and fibroblast migration due to the incorporation of Ni3(HITP)2 MOFs. This is because Ni3(HITP)2 MOF nanorods possessed high conductivity that can promote good free radical scavenging effect and reductive amino moiety in the molecular structure as well as abundant Ni–N4 active sites for electrical conductivity. When skin defects occur, there is a phenomenon in which negative charges present around the skin interact with positive charges within the wound, resulting in the creation of an endogenous electric field. This electric field is believed to play a significant role in directing and guiding the migration of cells toward the wound during the process of wound healing. As a result, the increased conductivity provided by Ni3(HITP)2 served to facilitate the transmission of electrical signals within biological tissues, thereby contributing to the promotion of wound healing [103].
5.2.2 Single-responsive MOF–hydrogels toward chemical stimuli
MOF–hydrogels have been reported to respond to chemical stimulation, such as pH changes, ions, or molecular signals, whereby they will undergo reversible structural alterations. These responsive behaviors enable controlled drug release, selective adsorption, or triggered chemical reactions [21]. The adaptability of MOF–hydrogels to changing chemical environments makes them a compelling avenue for wound treatment. The role of hydrogel in the chemical-responsive MOF–hydrogel composites is mainly serving as a carrier matrix for wound coverage and delivery of the active ingredients.
Zhang et al. [86] reported on a Cu-doped MOF nanocatalyst (GG-Cu@ZIF) loaded with GO x and encapsulated within a BC reinforced hydrogel for antibacterial and hemostatic therapies. The BC reinforced hydrogel serves as a matrix for encapsulating and immobilizing the Cu-doped MOF nanocatalyst (Cu@ZIF) loaded with GO x . This encapsulation is crucial for maintaining the stability and controlled release of the nanocatalyst within the wound site. The BC/GG-Cu@ZIF/GO x hydrogel displayed glucose decomposition capabilities and exhibited peroxidase-mimicking (POD). The Cu-doped MOF nanocatalyst with multienzyme mimetic activity loaded with GO x catalyzed in a physiological glucose environment into H2O2 and glucose acid through a self-activated cascaded reaction. The glucose acid further triggered the POD-mimicking activity, which could overcome the physiological pH limit and demand of exogenous H2O2 for POD-mimicking activity, producing exogenous •OH for bacteria inactivation (Figure 17). This BC/GG-Cu@ZIF/GO x hydrogel demonstrated high water absorption capacity, biocompatibility, and rapid hemostatic capability, both in vitro and in vivo. Experimental data revealed that the hydrogel exhibited a water absorption capacity of ∼3710% in a mass ratio of 2:1 of BC:GG. Additionally, in vitro evaluations demonstrated that the materials were cytocompatible and hemocompatible due to high cell viability in NIH-3T3 cells and low hemolysis ratios under 0.4%, confirming the material’s suitability for biomedical applications. Furthermore, hemostatic assays conducted in vivo showed rapid clotting times of within 1 min using a mouse liver hemorrhage model and a mouse-tail amputation model, underscoring the hydrogel’s efficacy in controlling bleeding. The absorption behavior of BC/GG-Cu@ZIF/GO x hydrogel contributed to its swift hemostasis by absorbing blood, which led to platelets aggregation and subsequently activated the blood coagulation cascades for hemostasis. Therefore, encapsulating Cu@ZIF/GO x within the hydrogel provided a synergistic effect of antibacterial properties and hemostasis for wound treatment [86].
![Figure 17
Catalytic cascaded reaction for bacteria inactivation and hemostasis of the BC/ Cu-doped MOF nanocatalyst/glucose oxidase (BC/GG-Cu@ZIF/GO
x
) hydrogel [86].](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_017.jpg)
Catalytic cascaded reaction for bacteria inactivation and hemostasis of the BC/ Cu-doped MOF nanocatalyst/glucose oxidase (BC/GG-Cu@ZIF/GO x ) hydrogel [86].
ROS acts as secondary messengers to immunocytes and non-lymphoid cells in the wound repair process. ROS regulates angiogenesis (formation of blood vessels) and contributes to the blood perfusion at the wound site. Furthermore, ROS serves as a crucial defender within the host’s immune system, particularly through the action of phagocytes. These specialized cells initiate an intense ROS burst directed at pathogens within wounds, effectively dismantling and destroying them. However, amidst this targeted assault, an overflow of ROS leaks into the surrounding environment, amplifying its bacteriostatic effects. This excess ROS acts as a formidable deterrent, further inhibiting bacterial growth and proliferation in the vicinity [135].
Zhu et al. reported on substance P (SP)-loaded ZIF-8 (SP@ZIF-8) NPs, which were then coated with PEG-TK to make them responsive to ROS [87]. SP is a neuropeptide composed of ten amino acids produced by both immune and neuronal cells. It is primarily located at the junction between the adventitia and media in the muscular layer, in the connective tissues surrounding blood vessels, and in the epidermis and dermis [136]. The SP@ZIF-8 NPs were incorporated into a hydrogel made of sodium alginate and pectin, which was then crosslinked with calcium chloride to create SP@ZIF-8-PEG-TK@CA dressings. The healing capabilities of SP@CA and SP@ZIF-8-PEG-TK@CA were assessed in vivo using an infected cutaneous wound mouse model. Treatment with SP@ZIF-8-PEG-TK@CA significantly reduced the wound area, showing a 12.6% reduction after 3 days compared to a 5.9% reduction in the SP@CA group. By day 7, SP@ZIF-8-PEG-TK@CA had reduced the wound area by 62.5%, compared to 42.5% in the SP@CA group. After 15 days, nearly complete wound closure (almost 100%) was achieved with SP@ZIF-8-PEG-TK@CA, significantly higher than in the SP@CA group.
Immunofluorescence staining revealed an abundance of CD32-positive cells at the injured sites treated with SP@ZIF-8-PEG-TK@CA, indicating an inflammatory response and accelerated wound healing. Furthermore, higher numbers of CD206-positive cells, indicative of M2 macrophage polarization associated with tissue repair, were observed at the late stage in mice treated with SP@ZIF-8-PEG-TK@CA compared to SP@CA dressings. These results suggested that SP@ZIF-8-PEG-TK@CA in hydrogel form promoted wound healing by stimulating M2 macrophage polarization [87].
5.2.3 Dual-responsive MOF–hydrogels
Dual-responsive MOF–hydrogel composites offer several advantages compared to single-responsive MOF–hydrogel composites due to their ability to respond to two different stimuli in the complex bioenvironment. For instance, Zhao et al. designed a dual light and temperature-responsive MCA-NI-AA/NP/ZIF8 hydrogel composed of PNIPAm and methacrylated κ-carrageenan (MA-κ-CA) incorporated with PPy-polydopamine NPs (PPy-PDA NPs) and ZIF-8 MOF, which offered both shape adaptability and synergistic bacterial elimination for wound healing application. The incorporation of PPy-PDA NPs into the MCA-NI-AA/ZIF8 hydrogel makes it a photothermal agent that converts photon energy into heat energy upon NIR-light irradiation, while the in situ formation of ZIF-8 in the hydrogel allowed the controlled release of Zn2+ ions over time. Through prolonged Zn2+ release, and when exposed to NIR irradiation, it generated localized heating and accelerated the release of Zn2+, thereby enhancing its therapeutic potential for treating highly infected wounds [137].
Nie et al. reported on light and temperature-sensitive polyMOF hydrogel incorporated with curcumin into a copper-tetrakis(4-carboxyphenyl)porphyrin MOF (Cu-TCPP MOF) and a poly(ethylene glycol)-poly(ε-caprolactone-co-lactide) hydrogel (known as CuPP-PELA polyMOF hydrogel) via in situ ring-opening polymerization [99]. Copper porphyrin (CuPP) served as a photothermal agent in the Cu-TPP MOF, where it was formed by the coordination of Cu2+ ions with porphyrin, a well-known photosensitive precursor that exhibits NIR absorption characteristics due to the d–d energy band transition of Cu2+ ions. When exposed to visible light, Cu-TCPP experienced a substantial local temperature increase, enabling the escape of water molecules from the voids/pores without causing damage to the crystal structure or disrupting proton conduction. Conversely, in the absence of light, water molecules re-entered the voids/pores, leading to reversible changes [138]. When CuPP-PELA polyMOF hydrogel was irradiated by NIR irradiation, the temperature of the CuPP-PELA polyMOF hydrogel increased due to CuPP. Curcumin was then released at a higher dose or released gradually post-irradiation at a lower dose, which functioned as both a sensitizer for PTT and a regulator for inflammation. As for the hydrogel, owing to its temperature-sensitive sol–gel transition properties, the hydrogel responded to body temperature and underwent a transition from a solution to a hydrogel state within the body, allowing it to adhere to the wound and form a complete seal. The study demonstrated the temperature of Cur/CuPP-PELA at the wound surface reached approximately 48.6°C within 5 min of NIR laser radiation. This significant increase in temperature suggested that the high efficacy of Cur/CuPP-PELA in combating pathogenic bacteria and infected surrounding tissue during PTT.
6 Conclusion
In conclusion, this article has delved into the diverse categories of stimulus-responsive MOF–hydrogel composites, their characterization methods, and their promising medical applications, specifically in wound healing and drug delivery. The integration of stimulus-responsive MOFs within hydrogels has garnered significant interest, primarily owing to their innate “smart” properties, enabling responses to various physical and chemical cues in the bioenvironment. These composite structures boast remarkable flexibility, superior structural integrity, and enhanced processability when compared to standalone MOFs. Leveraging the versatility of MOFs within hydrogel matrices unlocks boundless potential in biomedical realms, including but not limited to drug delivery, wound healing, and bio-diagnostics. The MOF–hydrogel composite materials offer a dynamic platform poised to revolutionize therapeutic and diagnostic strategies in the medical and biomedical fields.
6.1 Future prospective
While most studies have demonstrated the stimuli-responsive behavior of MOF–hydrogel composites, the underlying mechanisms governing these responses are not fully understood. Further research is needed to elucidate the specific interactions between the MOF and hydrogel components and how they contribute to or improve responsive behavior. Many studies have focused on the preparation of MOF–hydrogel composites with different combinations of MOFs and hydrogels; however, a critical aspect lies in understanding the compatibility of various hydrogel types with MOFs is essential. This comprehension is a necessity for judiciously selecting hydrogels that best complement stimulus-responsive MOFs. Comprehensive exploration involves systematic characterization, optimization, and conducting biodegradation and stability studies on these composite structures. Also, there is a lack of studies to address the challenges of the stimulus-responsive MOF–hydrogel, such as long-term stability and inflammatory response of the MOF–hydrogel composite in the development of stimulus-responsive MOF–hydrogel.
Additionally, multifunctional composites MOF–hydrogel composites have the potential to exhibit multiple stimuli-responsive behaviors simultaneously. At present, most studies are still focusing on a single stimulus responsiveness, such as pH or temperature. Investigating the design and synthesis of multifunctional composites that can respond to multiple stimuli in a controlled and predictable manner is an important research direction. In view of the enhanced physical, chemical, and biological properties of stimulus-responsive MOF–hydrogel composites, the applications of stimulus-responsive MOF–hydrogel composites should be expanded into advanced medical and biomedical applications such as bio-diagnostics, combined therapy, and precision medicine.
Acknowledgements
The authors would like to acknowledge the Malaysian Palm Oil Board (MPOB) for financial support.
-
Funding information: Authors state no funding involved.
-
Author contributions: Jia En, Toh – Investigation, methodology, visualization, writing– original draft and editing. Choy Sin, Lee – Conceptualization, methodology, visualization, supervision, investigation, writing-review and editing. Mallikarjuna Rao, Pichika – Supervision, methodology, writing – review and editing. Wei Huei, Lim- Supervision, methodology, writing – review. Bing Wei, Chua- Investigation, methodology, writing – original draft.
-
Conflict of interest: Authors state no conflict of interest.
-
Ethical approval: The conducted research is not related to either human or animal use.
-
Data availability statement: The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.
References
[1] Zhao X, Wang Y, Li D, Bu X, Feng P. Metal–organic frameworks for separation. Adv Mater. 2018 Sep;30(37):e1705189.Search in Google Scholar
[2] Cui J, Ren S, Sun B, Jia S. Optimization protocols and improved strategies for metal-organic frameworks for immobilizing enzymes: Current development and future challenges. Coord Chem Rev. 2018 Sep;370:22–41.Search in Google Scholar
[3] Wang C, Liu D, Lin W. Metal–organic frameworks as a tunable platform for designing functional molecular materials. J Am Chem Soc. 2013 Sep;135(36):13222–34.Search in Google Scholar
[4] Baumann AE, Burns DA, Liu B, Thoi VS. Metal-organic framework functionalization and design strategies for advanced electrochemical energy storage devices. Commun Chem. 2019 Jul;2(1):86.Search in Google Scholar
[5] Wasson MC, Buru CT, Chen Z, Islamoglu T, Farha OK. Metal–organic frameworks: A tunable platform to access single-site heterogeneous catalysts. Appl Catal A Gen. 2019 Sep;586:117214.Search in Google Scholar
[6] Miller SE, Teplensky MH, Moghadam PZ, Fairen-Jimenez D. Metal-organic frameworks as biosensors for luminescence-based detection and imaging. Interface Focus. 2016 Aug;6(4):20160027.Search in Google Scholar
[7] Sun Y, Zheng L, Yang Y, Qian X, Fu T, Li X, et al. Metal–organic framework nanocarriers for drug delivery in biomedical applications. Nanomicro Lett. 2020 Dec;12(1):103.Search in Google Scholar
[8] Khan M, Akmal Z, Tayyab M, Mansoor S, Zeb A, Ye Z, et al. MOFs materials as photocatalysts for CO2 reduction: Progress, challenges and perspectives. Carbon Capture Sci Technol. 2024 Jun;11:100191.Search in Google Scholar
[9] Shah SSA, Nazir MA, Khan K, Hussain I, Tayyab M, Alarfaji SS, et al. Solar energy storage to chemical: Photocatalytic CO2 reduction over pristine metal-organic frameworks with mechanistic studies. J Energy Storage. 2024 Jan;75:109725.Search in Google Scholar
[10] Chen Y, Li S, Pei X, Zhou J, Feng X, Zhang S, et al. A solvent‐free hot‐pressing method for preparing metal–organic‐framework coatings. Angew Chem. 2016 Mar;128(10):3480–4.Search in Google Scholar
[11] Neves P, Gomes AC, Amarante TR, Paz FAA, Pillinger M, Gonçalves IS, et al. Incorporation of a dioxomolybdenum(VI) complex in a ZrIV-based Metal–Organic Framework and its application in catalytic olefin epoxidation. Microporous Mesoporous Mater. 2015 Jan;202:106–14.Search in Google Scholar
[12] Krause S, Bon V, Stoeck U, Senkovska I, Többens DM, Wallacher D, et al. A Stimuli‐responsive zirconium metal–organic framework based on supermolecular design. Angew Chem Int Ed. 2017 Aug;56(36):10676–80.Search in Google Scholar
[13] Kanj AB, Müller K, Heinke L. Stimuli‐responsive metal‐organic frameworks with photoswitchable azobenzene side groups. Macromol Rapid Commun. 2018 Jan;39(1):e1700239.Search in Google Scholar
[14] Ding M, Cai X, Jiang HL. Improving MOF stability: approaches and applications. Chem Sci. 2019;10(44):10209–30.Search in Google Scholar
[15] Sun CY, Qin C, Wang XL, Yang GS, Shao KZ, Lan YQ, et al. Zeolitic imidazolate framework-8 as efficient pH-sensitive drug delivery vehicle. Dalton Trans. 2012;41(23):6906.Search in Google Scholar
[16] Bej S, Mandal S, Mondal A, Pal TK, Banerjee P. Solvothermal synthesis of high-performance d10-MOFs with hydrogel membranes @ “turn-On” monitoring of formaldehyde in solution and vapor phase. ACS Appl Mater Interfaces. 2021 Jun;13(21):25153–63.Search in Google Scholar
[17] Qiu J, Zhang X, Feng Y, Zhang X, Wang H, Yao J. Modified metal-organic frameworks as photocatalysts. Appl Catal B. 2018 Sep;231:317–42.Search in Google Scholar
[18] Han D, Li Y, Liu X, Li B, Han Y, Zheng Y, et al. Rapid bacteria trapping and killing of metal-organic frameworks strengthened photo-responsive hydrogel for rapid tissue repair of bacterial infected wounds. Chem Eng J. 2020 Sep;396:125194.Search in Google Scholar
[19] Yu H, Ren P, Pan X, Zhang X, Ma J, Chen J, et al. Intracellular delivery of itaconate by metal–organic framework-anchored hydrogel microspheres for osteoarthritis therapy. Pharmaceutics. 2023 Feb;15(3):724.Search in Google Scholar
[20] Li C, Wang K, Li J, Zhang Q. Recent progress in stimulus-responsive two-dimensional metal–organic frameworks. ACS Mater Lett. 2020 Jul;2(7):779–97.Search in Google Scholar
[21] Guan Q, Fang Y, Wu X, Ou R, Zhang X, Xie H, et al. Stimuli responsive metal organic framework materials towards advanced smart application. Mater Today. 2023 Apr;64:138–64.Search in Google Scholar
[22] Bahram M, Mohseni N, Moghtader M. An introduction to hydrogels and some recent applications. In: Emerging concepts in analysis and applications of hydrogels. London: InTech; 2016.Search in Google Scholar
[23] Furukawa H, Cordova KE, O’Keeffe M, Yaghi OM. The chemistry and applications of metal-organic frameworks. Science (1979). 2013 Aug;341(6149):e1230444.Search in Google Scholar
[24] Chowdhury MA. Metal‐organic‐frameworks as contrast agents in magnetic resonance imaging. ChemBioEng Rev. 2017 Aug;4(4):225–39.Search in Google Scholar
[25] Chui SSY, Lo SMF, Charmant JPH, Orpen AG, Williams ID. A chemically functionalizable nanoporous material [Cu3(TMA)2(H2O)3]n. Sci. 1999;283(5405):1148–50.Search in Google Scholar
[26] Sharma D, Rasaily S, Pradhan S, Baruah K, Tamang S, Pariyar A. HKUST-1 metal organic framework as an efficient dual-function catalyst: aziridination and one-pot ring-opening transformation for formation of β-Aryl sulfonamides with C–C, C–N, C–S, and C–O bonds. Inorg Chem. 2021 Jun;60(11):7794–802.Search in Google Scholar
[27] Millange F, Serre C, Férey G. Synthesis, structure determination and properties of MIL-53as and MIL-53ht: the first Criii hybrid inorganic–organic microporous solids: Criii(OH)·{O2C–C6H4–CO2}·{HO2C–C6H4–CO2H}x. Chem Commun. 2002 Apr;8:822–3.Search in Google Scholar
[28] Horcajada P, Surblé S, Serre C, Hong DY, Seo YK, Chang JS, et al. Synthesis and catalytic properties of MIL-100(Fe), an iron(III) carboxylate with large pores. Chem Commun. 2007;27:2820–2.Search in Google Scholar
[29] Hayashi H, Côté AP, Furukawa H, O’Keeffe M, Yaghi OM. Zeolite A imidazolate frameworks. Nat Mater. 2007 Jul;6(7):501–6.Search in Google Scholar
[30] Tan JC, Bennett TD, Cheetham AK. Chemical structure, network topology, and porosity effects on the mechanical properties of Zeolitic imidazolate frameworks. Proc Natl Acad Sci. 2010 Jun;107(22):9938–43.Search in Google Scholar
[31] Cavka JH, Jakobsen S, Olsbye U, Guillou N, Lamberti C, Bordiga S, et al. A new zirconium inorganic building brick forming metal organic frameworks with exceptional stability. J Am Chem Soc. 2008 Oct;130(42):13850–1.Search in Google Scholar
[32] Furukawa H, Gándara F, Zhang YB, Jiang J, Queen WL, Hudson MR, et al. Water adsorption in porous metal–organic frameworks and related materials. J Am Chem Soc. 2014 Mar;136(11):4369–81.Search in Google Scholar
[33] Chen G, Leng X, Luo J, You L, Qu C, Dong X, et al. In vitro toxicity study of a porous iron(III) metal‒organic framework. Molecules. 2019 Mar;24(7):1211.Search in Google Scholar
[34] Volkringer C, Meddouri M, Loiseau T, Guillou N, Marrot J, Férey G, et al. The kagomé topology of the gallium and indium metal-organic framework types with a mIL-68 structure: synthesis, XRD, solid-state NMR characterizations, and hydrogen adsorption. Inorg Chem. 2008 Dec;47(24):11892–901.Search in Google Scholar
[35] Schernikau M, Sablowski J, Gonzalez Martinez IG, Unz S, Kaskel S, Mikhailova D. Preparation and application of ZIF-8 thin layers. Appl Sci. 2021 Apr;11(9):4041.Search in Google Scholar
[36] Winarta J, Shan B, Mcintyre SM, Ye L, Wang C, Liu J, et al. A decade of UiO-66 research: a historic review of dynamic structure, synthesis mechanisms, and characterization techniques of an archetypal metal–organic framework. Cryst Growth Des. 2020 Feb;20(2):1347–62.Search in Google Scholar
[37] Kutzscher C, Nickerl G, Senkovska I, Bon V, Kaskel S. Proline functionalized UiO-67 and UiO-68 type metal–organic frameworks showing reversed diastereoselectivity in aldol addition reactions. Chem Mater. 2016 Apr;28(8):2573–80.Search in Google Scholar
[38] Tan K, Jensen S, Feng L, Wang H, Yuan S, Ferreri M, et al. Reactivity of atomic layer deposition precursors with OH/H 2 O-containing metal organic framework materials. Chem Mater. 2019 Apr;31(7):2286–95.Search in Google Scholar
[39] Furukawa H, Ko N, Go YB, Aratani N, Choi SB, Choi E, et al. Ultrahigh porosity in metal-organic frameworks. Science (1979). 2010 Jul;329(5990):424–8.Search in Google Scholar
[40] Hashemi B, Zohrabi P, Raza N, Kim KH. Metal-organic frameworks as advanced sorbents for the extraction and determination of pollutants from environmental, biological, and food media. TrAC Trends Anal Chem. 2017 Dec;97:65–82.Search in Google Scholar
[41] Lian X, Zhang Y, Wang J, Yan B. Antineoplastic mitoxantrone monitor: A sandwiched mixed matrix membrane (MMM) based on a luminescent MOF–hydrogel hybrid. Inorg Chem. 2020 Jul;59(14):10304–10.Search in Google Scholar
[42] Sultan S, Abdelhamid HN, Zou X, Mathew AP. CelloMOF: nanocellulose enabled 3D printing of metal–organic frameworks. Adv Funct Mater. 2019 Jan;29(2):e1805372.Search in Google Scholar
[43] Lin Y, Wang X, Sun Y, Dai Y, Sun W, Zhu X, et al. A chemiluminescent biosensor for ultrasensitive detection of adenosine based on target-responsive DNA hydrogel with Au@HKUST-1 encapsulation. Sens Actuators B Chem. 2019 Jun;289:56–64.Search in Google Scholar
[44] Xiao J, Chen S, Yi J, Zhang HF, Ameer GA. A cooperative copper metal–organic framework‐hydrogel system improves wound healing in diabetes. Adv Funct Mater. 2017 Jan;27(1):e1604872.Search in Google Scholar
[45] Garai A, Shepherd W, Huo J, Bradshaw D. Biomineral-inspired growth of metal–organic frameworks in gelatin hydrogel matrices. J Mater Chem B. 2013;1(30):3678.Search in Google Scholar
[46] Gwon K, Han I, Lee S, Kim Y, Lee DN. Novel metal–organic framework-based photocrosslinked hydrogel system for efficient antibacterial applications. ACS Appl Mater Interfaces. 2020 May;12(18):20234–42.Search in Google Scholar
[47] Hsieh CT, Ariga K, Shrestha LK, Hsu S. Development of MOF reinforcement for structural stability and toughness enhancement of biodegradable bioinks. Biomacromolecules. 2021 Mar;22(3):1053–64.Search in Google Scholar
[48] Liu H, Peng H, Xin Y, Zhang J. Metal–organic frameworks: a universal strategy towards super-elastic hydrogels. Polym Chem. 2019;10(18):2263–72.Search in Google Scholar
[49] Klein SE, Sosa JD, Castonguay AC, Flores WI, Zarzar LD, Liu Y. Green synthesis of Zr-based metal–organic framework hydrogel composites and their enhanced adsorptive properties. Inorg Chem Front. 2020;7(24):4813–21.Search in Google Scholar
[50] Yu Y, Chen G, Guo J, Liu Y, Ren J, Kong T, et al. Vitamin metal–organic framework-laden microfibers from microfluidics for wound healing. Mater Horiz. 2018;5(6):1137–42.Search in Google Scholar
[51] Lin Y, Sun Y, Dai Y, Sun W, Zhu X, Liu H, et al. A “signal-on” chemiluminescence biosensor for thrombin detection based on DNA functionalized magnetic sodium alginate hydrogel and metalloporphyrinic metal-organic framework nanosheets. Talanta. 2020 Jan;207:120300.Search in Google Scholar
[52] Zhao L, Gan J, Xia T, Jiang L, Zhang J, Cui Y, et al. A luminescent metal–organic framework integrated hydrogel optical fibre as a photoluminescence sensing platform for fluorescence detection. J Mater Chem C Mater. 2019;7(4):897–904.Search in Google Scholar
[53] Fan SJ, Sun R, Yan YB, Sun HB, Pang SN, Xu SD. A Dy(iii)–organic framework as a fluorescent probe for highly selective detection of picric acid and treatment activity on human lung cancer cells. Open Chem. 2020 Aug;18(1):1105–16.Search in Google Scholar
[54] Gao DY, Liu Z, Cheng ZL. 2D Ni-Fe MOF nanosheets reinforced poly(vinyl alcohol) hydrogels with enhanced mechanical and tribological performance. Colloids Surf A Physicochem Eng Asp. 2021 Feb;610:125934.Search in Google Scholar
[55] Huang MM, Xu LN. Construction of a new luminescent Cd(II) compound for the detection of Fe 3+ and treatment of Hepatitis B. Open Chem. 2021 Sep;19(1):953–60.Search in Google Scholar
[56] Wei M, Wan Y, Zhang X. Metal-organic framework-based stimuli-responsive polymers. J Compos Sci. 2021 Apr;5(4):101.Search in Google Scholar
[57] Karimzadeh S, Javanbakht S, Baradaran B, Shahbazi MA, Hashemzaei M, Mokhtarzadeh A, et al. Synthesis and therapeutic potential of stimuli-responsive metal-organic frameworks. Chem Eng J. 2021 Mar;408:127233.Search in Google Scholar
[58] Diring S, Carné-Sánchez A, Zhang J, Ikemura S, Kim C, Inaba H, et al. Light responsive metal–organic frameworks as controllable CO-releasing cell culture substrates. Chem Sci. 2017;8(3):2381–6.Search in Google Scholar
[59] Jiang K, Zhang L, Hu Q, Zhang Q, Lin W, Cui Y, et al. Thermal stimuli‐triggered drug release from a biocompatible porous metal–organic framework. Chem – A Eur J. 2017 Jul;23(42):10215–21.Search in Google Scholar
[60] Jia Q, Li Z, Guo C, Huang X, Song Y, Zhou N, et al. A γ-cyclodextrin-based metal-organic framework embedded with graphene quantum dots and modified with PEGMA: Via SI-ATRP for anticancer drug delivery and therapy. Nanoscale. 2019;11(43):20956–67.Search in Google Scholar
[61] Zhu YD, Chen SP, Zhao H, Yang Y, Chen XQ, Sun J, et al. PPy@MIL-100 nanoparticles as a pH- and Near-IR-irradiation-responsive drug carrier for simultaneous photothermal therapy and chemotherapy of cancer cells. ACS Appl Mater Interfaces. 2016 Dec;8(50):34209–17.Search in Google Scholar
[62] Qin Y, Fan J, Yang W, Shen B, Yang Y, Zhou Q, et al. Endogenous Cys-assisted GSH@AgNCs-rGO nanoprobe for real-time monitoring of dynamic change in GSH Levels regulated by natural drug. Anal Chem. 2020 Jan;92(2):1988–96.Search in Google Scholar
[63] Lei B, Wang M, Jiang Z, Qi W, Su R, He Z. Constructing redox-responsive metal–organic framework nanocarriers for anticancer drug delivery. ACS Appl Mater Interfaces. 2018 May;10(19):16698–706.Search in Google Scholar
[64] Nagata S, Kokado K, Sada K. Metal–organic framework tethering pH- and thermo-responsive polymer for ON–OFF controlled release of guest molecules. CrystEngComm. 2020;22(6):1106–11.Search in Google Scholar
[65] Jiang Z, Wang Y, Sun L, Yuan B, Tian Y, Xiang L, et al. Dual ATP and pH responsive ZIF-90 nanosystem with favorable biocompatibility and facile post-modification improves therapeutic outcomes of triple negative breast cancer in vivo. Biomaterials. 2019 Mar;197:41–50.Search in Google Scholar
[66] Tan LL, Song N, Zhang SXA, Li H, Wang B, Yang YW. Ca2+, pH and thermo triple-responsive mechanized Zr-based MOFs for on-command drug release in bone diseases. J Mater Chem B. 2016;4(1):135–40.Search in Google Scholar
[67] Leubner S, Stäglich R, Franke J, Jacobsen J, Gosch J, Siegel R, et al. Solvent impact on the properties of benchmark metal–organic frameworks: acetonitrile‐based synthesis of CAU‐10, Ce‐UiO‐66, and Al‐MIL‐53. Chem – Eur J. 2020 Mar;26(17):3877–83.Search in Google Scholar
[68] McKinstry C, Cathcart RJ, Cussen EJ, Fletcher AJ, Patwardhan SV, Sefcik J. Scalable continuous solvothermal synthesis of metal organic framework (MOF-5) crystals. Chem Eng J. 2016 Feb;285:718–25.Search in Google Scholar
[69] Zhang B, Luo Y, Kanyuck K, Saenz N, Reed K, Zavalij P, et al. Facile and template-free solvothermal synthesis of mesoporous/macroporous metal–organic framework nanosheets. RSC Adv. 2018;8(58):33059–64.Search in Google Scholar
[70] Byrappa K, Yoshimura M. Hydrothermal technology – principles and applications. In: Handbook of hydrothermal technology. Amsterdam: Elsevier; 2001. p. 1–52.Search in Google Scholar
[71] Rao BG, Mukherjee D, Reddy BM. Novel approaches for preparation of nanoparticles. Nanostructures for novel therapy: Synthesis, characterization and applications. Amsterdam: Elsevier; 2017 Jan. p. 11–36.Search in Google Scholar
[72] Han Y, Yang H, Guo X. Synthesis methods and crystallization of MOFs. In: Synthesis methods and crystallization. London: IntechOpen; 2020.Search in Google Scholar
[73] Ngan Tran TK, Ho HL, Nguyen HV, Tran BT, Nguyen TT, Thi Bui PQ, et al. Photocatalytic degradation of rhodamine B in aqueous phase by bimetallic metal-organic framework M/Fe-MOF (M = Co, Cu, and Mg). Open Chem. 2022 Jan;20(1):52–60.Search in Google Scholar
[74] Tayyab M, Liu Y, Liu Z, Pan L, Xu Z, Yue W, et al. One-pot in-situ hydrothermal synthesis of ternary In2S3/Nb2O5/Nb2C Schottky/S-scheme integrated heterojunction for efficient photocatalytic hydrogen production. J Colloid Interface Sci. 2022 Dec;628:500–12.Search in Google Scholar
[75] Tranchemontagne DJ, Hunt JR, Yaghi OM. Room temperature synthesis of metal-organic frameworks: MOF-5, MOF-74, MOF-177, MOF-199, and IRMOF-0. Tetrahedron. 2008 Sep;64(36):8553–7.Search in Google Scholar
[76] Abánades Lázaro I, Haddad S, Sacca S, Orellana-Tavra C, Fairen-Jimenez D, Forgan RS. Selective surface PEGylation of UiO-66 nanoparticles for enhanced stability, cell uptake, and pH-responsive drug delivery. Chem. 2017 Apr;2(4):561–78.Search in Google Scholar
[77] Abazari R, Mahjoub AR, Ataei F, Morsali A, Carpenter-Warren CL, Mehdizadeh K, et al. Chitosan immobilization on Bio-MOF nanostructures: a biocompatible PH-responsive nanocarrier for doxorubicin release on MCF-7 cell lines of human breast cancer. Inorg Chem. 2018 Nov;57(21):13364–79.Search in Google Scholar
[78] Javanbakht S, Pooresmaeil M, Hashemi H, Namazi H. Carboxymethylcellulose capsulated Cu-based metal-organic framework-drug nanohybrid as a pH-sensitive nanocomposite for ibuprofen oral delivery. Int J Biol Macromol. 2018 Nov;119:588–96.Search in Google Scholar
[79] Nabipour H, Hu Y. Development of fully bio-based pectin/curcumin@bio-MOF-11 for colon specific drug delivery. Chem Pap. 2022 May;76(5):2969–79.Search in Google Scholar
[80] Javanbakht S, Hemmati A, Namazi H, Heydari A. Carboxymethylcellulose-coated 5-fluorouracil@MOF-5 nano-hybrid as a bio-nanocomposite carrier for the anticancer oral delivery. Int J Biol Macromol. 2020 Jul;155:876–82.Search in Google Scholar
[81] Nezhad-Mokhtari P, Arsalani N, Javanbakht S, Shaabani A. Development of gelatin microsphere encapsulated Cu-based metal-organic framework nanohybrid for the methotrexate delivery. J Drug Delivery Sci Technol. 2019 Apr;50:174–80.Search in Google Scholar
[82] Javanbakht S, Shadi M, Mohammadian R, Shaabani A, Amini MM, Pooresmaeil M, et al. Facile preparation of pH-responsive k-Carrageenan/tramadol loaded UiO-66 bio-nanocomposite hydrogel beads as a nontoxic oral delivery vehicle. J Drug Deliv Sci Technol. 2019 Dec;54:101311.Search in Google Scholar
[83] Lin X, Guo L, Shaghaleh H, Hamoud YA, Xu X, Liu H. A TEMPO-oxidized cellulose nanofibers/MOFs hydrogel with temperature and pH responsiveness for fertilizers slow-release. Int J Biol Macromol. 2021 Nov;191:483–91.Search in Google Scholar
[84] Wu L, Chen G, Li Z. Layered rare‐earth hydroxide/polyacrylamide nanocomposite hydrogels with highly tunable photoluminescence. Small. 2017 Jun;13(23):e1604070.Search in Google Scholar
[85] Dong W, Zhao S, Wang Y, Zhou X, Jiang J, Dang J, et al. Stimuli-responsive metal–organic framework hydrogels endow long carbon fiber reinforced polyetheretherketone with enhanced anti-inflammatory and angiogenesis and osteogenesis. Mater Des. 2023 Jan;225:111485.Search in Google Scholar
[86] Zhang S, Ding F, Liu Y, Ren X. Glucose-responsive biomimetic nanoreactor in bacterial cellulose hydrogel for antibacterial and hemostatic therapies. Carbohydr Polym. 2022 Sep;292:119615.Search in Google Scholar
[87] Zhu Y, Yao Z, Liu Y, Zhang W, Geng L, Ni T. Incorporation of ROS-responsive substance P-loaded zeolite imidazolate framework-8 nanoparticles into a Ca2+-cross-linked alginate/ pectin hydrogel for wound dressing applications. Int J Nanomed. 2020 Jan;Volume 15:333–46.Search in Google Scholar
[88] Deng Z, Li M, Hu Y, He Y, Tao B, Yuan Z, et al. Injectable biomimetic hydrogels encapsulating Gold/metal–organic frameworks nanocomposites for enhanced antibacterial and wound healing activity under visible light actuation. Chem Eng J. 2021 Sep;420:129668.Search in Google Scholar
[89] Feng L, Chen Q, Cheng H, Yu Q, Zhao W, Zhao C. Dually‐thermoresponsive hydrogel with shape adaptability and synergetic bacterial elimination in the full course of wound healing. Adv Healthc Mater. 2022 Sep;11(18):e2201049.Search in Google Scholar
[90] Li J, Fu Z, Liu Y. Encapsulation of liquid metal nanoparticles inside metal–organic frameworks for hydrogel-integrated dual functional biotherapy. Chem Eng J. 2023 Feb;457:141302.Search in Google Scholar
[91] Du T, Xiao Z, Zhang G, Wei L, Cao J, Zhang Z, et al. An injectable multifunctional hydrogel for eradication of bacterial biofilms and wound healing. Acta Biomater. 2023 Apr;161:112–33.Search in Google Scholar
[92] Zheng H, Chen Y, Yang J, Hao P, Ren L, Zhou H. Bicomponent hydrogels assisted templating synthesis of hierarchically porous ZIF-8 for efficient antibacterial applications. J Mol Struct. 2023 Apr;1277:134824.Search in Google Scholar
[93] Liu Y, Ma N, Gao N, Ling G, Zhang P. Metal-organic framework-based injectable in situ gel for multi-responsive insulin delivery. J Drug Delivery Sci Technol. 2022 Sep;75:103604.Search in Google Scholar
[94] Zhou Y, Chen Z, Zhao D, Li D, He C, Chen X. A pH‐triggered self‐unpacking capsule containing zwitterionic hydrogel‐coated MOF nanoparticles for efficient oral exendin‐4 delivery. Adv Mater. 2021 Aug;33(32):e2102044.Search in Google Scholar
[95] Reddy YN, De A, Paul S, Pujari Ak, Bhaumik J. In situ nanoarchitectonics of a mof hydrogel: a self-adhesive and pH-Responsive smart platform for phototherapeutic delivery. Biomacromolecules. 2023 Apr;24(4):1717–30.Search in Google Scholar
[96] Tan G, Zhong Y, Yang L, Jiang Y, Liu J, Ren F. A multifunctional MOF-based nanohybrid as injectable implant platform for drug synergistic oral cancer therapy. Chem Eng J. 2020 Jun;390:124446.Search in Google Scholar
[97] Zeng Y, Zhang C, Du D, Li Y, Sun L, Han Y, et al. Metal-organic framework-based hydrogel with structurally dynamic properties as a stimuli-responsive localized drug delivery system for cancer therapy. Acta Biomater. 2022 Jun;145:43–51.Search in Google Scholar
[98] Du J, Fan L, Fu J. Hydrogels loaded with atenolol drug metal–organic framework showing biological activity. Sci Eng Compos Mater. 2023 Jul;30(1):e20220216.Search in Google Scholar
[99] Nie W, Huang Y, Wang Y, Kengla C, Scott Copus J, Sun J, et al. Temperature sensitive polyMOF hydrogel formed by in situ open-ring polymerization for infected chronic wound treatment. Chem Eng J. 2022 Oct;446:136948.Search in Google Scholar
[100] Tian M, Zhou L, Fan C, Wang L, Lin X, Wen Y, et al. Bimetal-organic framework/GOx-based hydrogel dressings with antibacterial and inflammatory modulation for wound healing. Acta Biomater. 2023 Mar;158:252–65.Search in Google Scholar
[101] Chen W, Liao W, Sohn YS, Fadeev M, Cecconello A, Nechushtai R, et al. Stimuli‐responsive nucleic acid‐based polyacrylamide hydrogel‐coated metal–organic framework nanoparticles for controlled drug release. Adv Funct Mater. 2018 Feb;28(8):e1705137.Search in Google Scholar
[102] Zhang W, Zhou Y, Fan Y, Cao R, Xu Y, Weng Z, et al. Metal–organic-framework-based hydrogen-release platform for multieffective helicobacter pylori targeting therapy and intestinal flora protective capabilities. Adv Mater. 2022 Jan;34(2):e2105738.Search in Google Scholar
[103] Zhang W, Dai X, Jin X, Huang M, Shan J, Chen X, et al. Promotion of wound healing by a thermosensitive and sprayable hydrogel with nanozyme activity and anti-inflammatory properties. Smart Mater Med. 2023;4:134–45.Search in Google Scholar
[104] Wu W, Liu J, Lin X, He Z, Zhang H, Ji L, et al. Dual-functional MOFs-based hybrid microgel advances aqueous lubrication and anti-inflammation. J Colloid Interface Sci. 2023 Aug;644:200–10.Search in Google Scholar
[105] Chao D, Dong Q, Yu Z, Qi D, Li M, Xu L, et al. Specific nanodrug for diabetic chronic wounds based on antioxidase-mimicking MOF-818 nanozymes. J Am Chem Soc. 2022 Dec;144(51):23438–47.Search in Google Scholar
[106] Cheng P, Wang C, Kaneti YV, Eguchi M, Lin J, Yamauchi Y, et al. Practical MOF Nanoarchitectonics: New Strategies for enhancing the processability of MOFs for practical applications. Langmuir. 2020 Apr;36(16):4231–49.Search in Google Scholar
[107] Chen F, Wang YM, Guo W, Yin XB. Color-tunable lanthanide metal–organic framework gels. Chem Sci. 2019;10(6):1644–50.Search in Google Scholar
[108] Zhu H, Zhang Q, Zhu S. Alginate hydrogel: a shapeable and versatile platform for in situ preparation of metal-organic framework-polymer composites. ACS Appl Mater Interfaces. 2016 Jul;8(27):17395–401.Search in Google Scholar
[109] Ahmed EM. Hydrogel: Preparation, characterization, and applications: A review. J Adv Res. 2015 Mar;6(2):105–21.Search in Google Scholar
[110] Varghese SA, Rangappa SM, Siengchin S, Parameswaranpillai J. Natural polymers and the hydrogels prepared from them. In: Hydrogels based on natural polymers. Amsterdam: Elsevier; 2020. p. 17–47.Search in Google Scholar
[111] Suvarnapathaki S, Nguyen MA, Wu X, Nukavarapu SP, Camci-Unal G. Synthesis and characterization of photocrosslinkable hydrogels from bovine skin gelatin. RSC Adv. 2019;9(23):13016–25.Search in Google Scholar
[112] Jaipan P, Nguyen A, Narayan RJ. Gelatin-based hydrogels for biomedical applications. MRS Commun. 2017 Sep;7(3):416–26.Search in Google Scholar
[113] Leong JY, Lam WH, Ho KW, Voo WP, Lee MFX, Lim HP, et al. Advances in fabricating spherical alginate hydrogels with controlled particle designs by ionotropic gelation as encapsulation systems. Particuology. 2016 Feb;24:44–60.Search in Google Scholar
[114] Kundu R, Mahada P, Chhirang B, Das B. Cellulose hydrogels: Green and sustainable soft biomaterials. Curr Res Green Sustain Chem. 2022;5:100252.Search in Google Scholar
[115] Munarin F, Tanzi MC, Petrini P. Advances in biomedical applications of pectin gels. Int J Biol Macromol. 2012 Nov;51(4):681–9.Search in Google Scholar
[116] Groult S, Buwalda S, Budtova T. Pectin hydrogels, aerogels, cryogels and xerogels: Influence of drying on structural and release properties. Eur Polym J. 2021 Apr;149:110386.Search in Google Scholar
[117] Joy R, Vigneshkumar PN, John F, George J. Hydrogels based on carrageenan. In: Plant and algal hydrogels for drug delivery and regenerative medicine. Amsterdam: Elsevier; 2021. p. 293–325.Search in Google Scholar
[118] Mohandas A, Deepthi S, Biswas R, Jayakumar R. Chitosan based metallic nanocomposite scaffolds as antimicrobial wound dressings. Bioact Mater. 2018 Sep;3(3):267–77.Search in Google Scholar
[119] Madduma‐Bandarage USK, Madihally SV. Synthetic hydrogels: Synthesis, novel trends, and applications. J Appl Polym Sci. 2021 May;138(19):e50376.Search in Google Scholar
[120] Anderson FA. Safety assessment of polyacrylamide and acrylamide residues in cosmetics. Int J Toxicol. 2005 Mar;24(2_suppl):21–50.Search in Google Scholar
[121] Alexander A, Ajazuddin, Khan J, Saraf S, Saraf S. Polyethylene glycol (PEG)–Poly(N-isopropylacrylamide) (PNIPAAm) based thermosensitive injectable hydrogels for biomedical applications. Eur J Pharm Biopharm. 2014 Nov;88(3):575–85.Search in Google Scholar
[122] Makadia HK, Siegel SJ. Poly Lactic-co-Glycolic Acid (PLGA) as biodegradable controlled drug delivery carrier. Polymer (Basel). 2011 Aug;3(3):1377–97.Search in Google Scholar
[123] Tzounis L, Doña M, Lopez-Romero JM, Fery A, Contreras-Caceres R. Temperature-controlled catalysis by core–shell–satellite AuAg@pNIPAM@Ag hybrid microgels: a highly efficient catalytic thermoresponsive nanoreactor. ACS Appl Mater Interfaces. 2019 Aug;11(32):29360–72.Search in Google Scholar
[124] Xu Y, Bao Y, Liu Z, Zheng Q, Dong Y, Song R, et al. Temperature-sensitive tribological performance of titanium alloy lubricated with PNIPAM microgels. Appl Surf Sci. 2022 Jan;572:151392.Search in Google Scholar
[125] Yu Q, Tian Y, Li M, Jiang Y, Sun H, Zhang G, et al. Poly(ethylene glycol)-mediated mineralization of metal–organic frameworks. Chem Commun. 2020;56(75):11078–81.Search in Google Scholar
[126] Titus D, James Jebaseelan Samuel E, Roopan SM. Nanoparticle characterization techniques. In: Green synthesis, characterization and applications of nanoparticles. Amsterdam: Elsevier; 2019. p. 303–19.Search in Google Scholar
[127] Nasiri S, Rabiei M, Palevicius A, Janusas G, Vilkauskas A, Nutalapati V, et al. Modified Scherrer equation to calculate crystal size by XRD with high accuracy, examples Fe2O3, TiO2 and V2O5. Nano Trends. 2023 Sep;3:100015.Search in Google Scholar
[128] Wilmer CE, Leaf M, Lee CY, Farha OK, Hauser BG, Hupp JT, et al. Large-scale screening of hypothetical metal–organic frameworks. Nat Chem. 2012 Feb;4(2):83–9.Search in Google Scholar
[129] Healy C, Patil KM, Wilson BH, Hermanspahn L, Harvey-Reid NC, Howard BI, et al. The thermal stability of metal-organic frameworks. Coord Chem Rev. 2020 Sep;419:213388.Search in Google Scholar
[130] Aoki H, Al-Assaf S, Katayama T, Phillips GO. Characterization and properties of Acacia senegal (L.) Willd. var. senegal with enhanced properties (Acacia (sen) SUPER GUMTM): Part 2 – Mechanism of the maturation process. Food Hydrocoll. 2007 May;21(3):329–37.Search in Google Scholar
[131] Zhang K, Feng W, Jin C. Protocol efficiently measuring the swelling rate of hydrogels. MethodsX. 2020;7:100779.Search in Google Scholar
[132] Chen MH, Wang LL, Chung JJ, Kim YH, Atluri P, Burdick JA. Methods to assess shear-thinning hydrogels for application as injectable biomaterials. ACS Biomater Sci Eng. 2017 Dec;3(12):3146–60.Search in Google Scholar
[133] Broughton W. Testing the mechanical, thermal and chemical properties of adhesives for marine environments. In: Adhesives in marine engineering. Amsterdam: Elsevier; 2012. p. 99–154.Search in Google Scholar
[134] Wang TL, Zhou ZF, Liu JF, Hou XD, Zhou Z, Dai YL, et al. Donut-like MOFs of copper/nicotinic acid and composite hydrogels with superior bioactivity for rh-bFGF delivering and skin wound healing. J Nanobiotechnol. 2021 Dec;19(1):275.Search in Google Scholar
[135] Dunnill C, Patton T, Brennan J, Barrett J, Dryden M, Cooke J, et al. Reactive oxygen species (ROS) and wound healing: the functional role of ROS and emerging ROS‐modulating technologies for augmentation of the healing process. Int Wound J. 2017 Feb;14(1):89–96.Search in Google Scholar
[136] Suvas S. Role of substance P neuropeptide in inflammation, wound healing, and tissue homeostasis. J Immunol. 2017 Sep;199(5):1543–52.Search in Google Scholar
[137] Feng L, Chen Q, Cheng H, Yu Q, Zhao W, Zhao C. Dually-thermoresponsive hydrogel with shape adaptability and synergetic bacterial elimination in the full course of wound healing. Adv Healthc Mater. 2022 Sep;11(18):e2201049.Search in Google Scholar
[138] Li B, Wang X, Chen L, Zhou Y, Dang W, Chang J, et al. Ultrathin Cu-TCPP MOF nanosheets: a new theragnostic nanoplatform with magnetic resonance/near-infrared thermal imaging for synergistic phototherapy of cancers. Theranostics. 2018;8(15):4086–96.Search in Google Scholar
© 2024 the author(s), published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Regular Articles
- Porous silicon nanostructures: Synthesis, characterization, and their antifungal activity
- Biochar from de-oiled Chlorella vulgaris and its adsorption on antibiotics
- Phytochemicals profiling, in vitro and in vivo antidiabetic activity, and in silico studies on Ajuga iva (L.) Schreb.: A comprehensive approach
- Synthesis, characterization, in silico and in vitro studies of novel glycoconjugates as potential antibacterial, antifungal, and antileishmanial agents
- Sonochemical synthesis of gold nanoparticles mediated by potato starch: Its performance in the treatment of esophageal cancer
- Computational study of ADME-Tox prediction of selected phytochemicals from Punica granatum peels
- Phytochemical analysis, in vitro antioxidant and antifungal activities of extracts and essential oil derived from Artemisia herba-alba Asso
- Two triazole-based coordination polymers: Synthesis and crystal structure characterization
- Phytochemical and physicochemical studies of different apple varieties grown in Morocco
- Synthesis of multi-template molecularly imprinted polymers (MT-MIPs) for isolating ethyl para-methoxycinnamate and ethyl cinnamate from Kaempferia galanga L., extract with methacrylic acid as functional monomer
- Nutraceutical potential of Mesembryanthemum forsskaolii Hochst. ex Bioss.: Insights into its nutritional composition, phytochemical contents, and antioxidant activity
- Evaluation of influence of Butea monosperma floral extract on inflammatory biomarkers
- Cannabis sativa L. essential oil: Chemical composition, anti-oxidant, anti-microbial properties, and acute toxicity: In vitro, in vivo, and in silico study
- The effect of gamma radiation on 5-hydroxymethylfurfural conversion in water and dimethyl sulfoxide
- Hollow mushroom nanomaterials for potentiometric sensing of Pb2+ ions in water via the intercalation of iodide ions into the polypyrrole matrix
- Determination of essential oil and chemical composition of St. John’s Wort
- Computational design and in vitro assay of lantadene-based novel inhibitors of NS3 protease of dengue virus
- Anti-parasitic activity and computational studies on a novel labdane diterpene from the roots of Vachellia nilotica
- Microbial dynamics and dehydrogenase activity in tomato (Lycopersicon esculentum Mill.) rhizospheres: Impacts on growth and soil health across different soil types
- Correlation between in vitro anti-urease activity and in silico molecular modeling approach of novel imidazopyridine–oxadiazole hybrids derivatives
- Spatial mapping of indoor air quality in a light metro system using the geographic information system method
- Iron indices and hemogram in renal anemia and the improvement with Tribulus terrestris green-formulated silver nanoparticles applied on rat model
- Integrated track of nano-informatics coupling with the enrichment concept in developing a novel nanoparticle targeting ERK protein in Naegleria fowleri
- Cytotoxic and phytochemical screening of Solanum lycopersicum–Daucus carota hydro-ethanolic extract and in silico evaluation of its lycopene content as anticancer agent
- Protective activities of silver nanoparticles containing Panax japonicus on apoptotic, inflammatory, and oxidative alterations in isoproterenol-induced cardiotoxicity
- pH-based colorimetric detection of monofunctional aldehydes in liquid and gas phases
- Investigating the effect of resveratrol on apoptosis and regulation of gene expression of Caco-2 cells: Unravelling potential implications for colorectal cancer treatment
- Metformin inhibits knee osteoarthritis induced by type 2 diabetes mellitus in rats: S100A8/9 and S100A12 as players and therapeutic targets
- Effect of silver nanoparticles formulated by Silybum marianum on menopausal urinary incontinence in ovariectomized rats
- Synthesis of new analogs of N-substituted(benzoylamino)-1,2,3,6-tetrahydropyridines
- Response of yield and quality of Japonica rice to different gradients of moisture deficit at grain-filling stage in cold regions
- Preparation of an inclusion complex of nickel-based β-cyclodextrin: Characterization and accelerating the osteoarthritis articular cartilage repair
- Empagliflozin-loaded nanomicelles responsive to reactive oxygen species for renal ischemia/reperfusion injury protection
- Preparation and pharmacodynamic evaluation of sodium aescinate solid lipid nanoparticles
- Assessment of potentially toxic elements and health risks of agricultural soil in Southwest Riyadh, Saudi Arabia
- Theoretical investigation of hydrogen-rich fuel production through ammonia decomposition
- Biosynthesis and screening of cobalt nanoparticles using citrus species for antimicrobial activity
- Investigating the interplay of genetic variations, MCP-1 polymorphism, and docking with phytochemical inhibitors for combatting dengue virus pathogenicity through in silico analysis
- Ultrasound induced biosynthesis of silver nanoparticles embedded into chitosan polymers: Investigation of its anti-cutaneous squamous cell carcinoma effects
- Copper oxide nanoparticles-mediated Heliotropium bacciferum leaf extract: Antifungal activity and molecular docking assays against strawberry pathogens
- Sprouted wheat flour for improving physical, chemical, rheological, microbial load, and quality properties of fino bread
- Comparative toxicity assessment of fisetin-aided artificial intelligence-assisted drug design targeting epibulbar dermoid through phytochemicals
- Acute toxicity and anti-inflammatory activity of bis-thiourea derivatives
- Anti-diabetic activity-guided isolation of α-amylase and α-glucosidase inhibitory terpenes from Capsella bursa-pastoris Linn.
- GC–MS analysis of Lactobacillus plantarum YW11 metabolites and its computational analysis on familial pulmonary fibrosis hub genes
- Green formulation of copper nanoparticles by Pistacia khinjuk leaf aqueous extract: Introducing a novel chemotherapeutic drug for the treatment of prostate cancer
- Improved photocatalytic properties of WO3 nanoparticles for Malachite green dye degradation under visible light irradiation: An effect of La doping
- One-pot synthesis of a network of Mn2O3–MnO2–poly(m-methylaniline) composite nanorods on a polypyrrole film presents a promising and efficient optoelectronic and solar cell device
- Groundwater quality and health risk assessment of nitrate and fluoride in Al Qaseem area, Saudi Arabia
- A comparative study of the antifungal efficacy and phytochemical composition of date palm leaflet extracts
- Processing of alcohol pomelo beverage (Citrus grandis (L.) Osbeck) using saccharomyces yeast: Optimization, physicochemical quality, and sensory characteristics
- Specialized compounds of four Cameroonian spices: Isolation, characterization, and in silico evaluation as prospective SARS-CoV-2 inhibitors
- Identification of a novel drug target in Porphyromonas gingivalis by a computational genome analysis approach
- Physico-chemical properties and durability of a fly-ash-based geopolymer
- FMS-like tyrosine kinase 3 inhibitory potentials of some phytochemicals from anti-leukemic plants using computational chemical methodologies
- Wild Thymus zygis L. ssp. gracilis and Eucalyptus camaldulensis Dehnh.: Chemical composition, antioxidant and antibacterial activities of essential oils
- 3D-QSAR, molecular docking, ADMET, simulation dynamic, and retrosynthesis studies on new styrylquinolines derivatives against breast cancer
- Deciphering the influenza neuraminidase inhibitory potential of naturally occurring biflavonoids: An in silico approach
- Determination of heavy elements in agricultural regions, Saudi Arabia
- Synthesis and characterization of antioxidant-enriched Moringa oil-based edible oleogel
- Ameliorative effects of thistle and thyme honeys on cyclophosphamide-induced toxicity in mice
- Study of phytochemical compound and antipyretic activity of Chenopodium ambrosioides L. fractions
- Investigating the adsorption mechanism of zinc chloride-modified porous carbon for sulfadiazine removal from water
- Performance repair of building materials using alumina and silica composite nanomaterials with electrodynamic properties
- Effects of nanoparticles on the activity and resistance genes of anaerobic digestion enzymes in livestock and poultry manure containing the antibiotic tetracycline
- Effect of copper nanoparticles green-synthesized using Ocimum basilicum against Pseudomonas aeruginosa in mice lung infection model
- Cardioprotective effects of nanoparticles green formulated by Spinacia oleracea extract on isoproterenol-induced myocardial infarction in mice by the determination of PPAR-γ/NF-κB pathway
- Anti-OTC antibody-conjugated fluorescent magnetic/silica and fluorescent hybrid silica nanoparticles for oxytetracycline detection
- Curcumin conjugated zinc nanoparticles for the treatment of myocardial infarction
- Identification and in silico screening of natural phloroglucinols as potential PI3Kα inhibitors: A computational approach for drug discovery
- Exploring the phytochemical profile and antioxidant evaluation: Molecular docking and ADMET analysis of main compounds from three Solanum species in Saudi Arabia
- Unveiling the molecular composition and biological properties of essential oil derived from the leaves of wild Mentha aquatica L.: A comprehensive in vitro and in silico exploration
- Analysis of bioactive compounds present in Boerhavia elegans seeds by GC-MS
- Homology modeling and molecular docking study of corticotrophin-releasing hormone: An approach to treat stress-related diseases
- LncRNA MIR17HG alleviates heart failure via targeting MIR17HG/miR-153-3p/SIRT1 axis in in vitro model
- Development and validation of a stability indicating UPLC-DAD method coupled with MS-TQD for ramipril and thymoquinone in bioactive SNEDDS with in silico toxicity analysis of ramipril degradation products
- Biosynthesis of Ag/Cu nanocomposite mediated by Curcuma longa: Evaluation of its antibacterial properties against oral pathogens
- Development of AMBER-compliant transferable force field parameters for polytetrafluoroethylene
- Treatment of gestational diabetes by Acroptilon repens leaf aqueous extract green-formulated iron nanoparticles in rats
- Development and characterization of new ecological adsorbents based on cardoon wastes: Application to brilliant green adsorption
- A fast, sensitive, greener, and stability-indicating HPLC method for the standardization and quantitative determination of chlorhexidine acetate in commercial products
- Assessment of Se, As, Cd, Cr, Hg, and Pb content status in Ankang tea plantations of China
- Effect of transition metal chloride (ZnCl2) on low-temperature pyrolysis of high ash bituminous coal
- Evaluating polyphenol and ascorbic acid contents, tannin removal ability, and physical properties during hydrolysis and convective hot-air drying of cashew apple powder
- Development and characterization of functional low-fat frozen dairy dessert enhanced with dried lemongrass powder
- Scrutinizing the effect of additive and synergistic antibiotics against carbapenem-resistant Pseudomonas aeruginosa
- Preparation, characterization, and determination of the therapeutic effects of copper nanoparticles green-formulated by Pistacia atlantica in diabetes-induced cardiac dysfunction in rat
- Antioxidant and antidiabetic potentials of methoxy-substituted Schiff bases using in vitro, in vivo, and molecular simulation approaches
- Anti-melanoma cancer activity and chemical profile of the essential oil of Seseli yunnanense Franch
- Molecular docking analysis of subtilisin-like alkaline serine protease (SLASP) and laccase with natural biopolymers
- Overcoming methicillin resistance by methicillin-resistant Staphylococcus aureus: Computational evaluation of napthyridine and oxadiazoles compounds for potential dual inhibition of PBP-2a and FemA proteins
- Exploring novel antitubercular agents: Innovative design of 2,3-diaryl-quinoxalines targeting DprE1 for effective tuberculosis treatment
- Drimia maritima flowers as a source of biologically potent components: Optimization of bioactive compound extractions, isolation, UPLC–ESI–MS/MS, and pharmacological properties
- Estimating molecular properties, drug-likeness, cardiotoxic risk, liability profile, and molecular docking study to characterize binding process of key phyto-compounds against serotonin 5-HT2A receptor
- Fabrication of β-cyclodextrin-based microgels for enhancing solubility of Terbinafine: An in-vitro and in-vivo toxicological evaluation
- Phyto-mediated synthesis of ZnO nanoparticles and their sunlight-driven photocatalytic degradation of cationic and anionic dyes
- Monosodium glutamate induces hypothalamic–pituitary–adrenal axis hyperactivation, glucocorticoid receptors down-regulation, and systemic inflammatory response in young male rats: Impact on miR-155 and miR-218
- Quality control analyses of selected honey samples from Serbia based on their mineral and flavonoid profiles, and the invertase activity
- Eco-friendly synthesis of silver nanoparticles using Phyllanthus niruri leaf extract: Assessment of antimicrobial activity, effectiveness on tropical neglected mosquito vector control, and biocompatibility using a fibroblast cell line model
- Green synthesis of silver nanoparticles containing Cichorium intybus to treat the sepsis-induced DNA damage in the liver of Wistar albino rats
- Quality changes of durian pulp (Durio ziberhinus Murr.) in cold storage
- Study on recrystallization process of nitroguanidine by directly adding cold water to control temperature
- Determination of heavy metals and health risk assessment in drinking water in Bukayriyah City, Saudi Arabia
- Larvicidal properties of essential oils of three Artemisia species against the chemically insecticide-resistant Nile fever vector Culex pipiens (L.) (Diptera: Culicidae): In vitro and in silico studies
- Design, synthesis, characterization, and theoretical calculations, along with in silico and in vitro antimicrobial proprieties of new isoxazole-amide conjugates
- The impact of drying and extraction methods on total lipid, fatty acid profile, and cytotoxicity of Tenebrio molitor larvae
- A zinc oxide–tin oxide–nerolidol hybrid nanomaterial: Efficacy against esophageal squamous cell carcinoma
- Research on technological process for production of muskmelon juice (Cucumis melo L.)
- Physicochemical components, antioxidant activity, and predictive models for quality of soursop tea (Annona muricata L.) during heat pump drying
- Characterization and application of Fe1−xCoxFe2O4 nanoparticles in Direct Red 79 adsorption
- Torilis arvensis ethanolic extract: Phytochemical analysis, antifungal efficacy, and cytotoxicity properties
- Magnetite–poly-1H pyrrole dendritic nanocomposite seeded on poly-1H pyrrole: A promising photocathode for green hydrogen generation from sanitation water without using external sacrificing agent
- HPLC and GC–MS analyses of phytochemical compounds in Haloxylon salicornicum extract: Antibacterial and antifungal activity assessment of phytopathogens
- Efficient and stable to coking catalysts of ethanol steam reforming comprised of Ni + Ru loaded on MgAl2O4 + LnFe0.7Ni0.3O3 (Ln = La, Pr) nanocomposites prepared via cost-effective procedure with Pluronic P123 copolymer
- Nitrogen and boron co-doped carbon dots probe for selectively detecting Hg2+ in water samples and the detection mechanism
- Heavy metals in road dust from typical old industrial areas of Wuhan: Seasonal distribution and bioaccessibility-based health risk assessment
- Phytochemical profiling and bioactivity evaluation of CBD- and THC-enriched Cannabis sativa extracts: In vitro and in silico investigation of antioxidant and anti-inflammatory effects
- Investigating dye adsorption: The role of surface-modified montmorillonite nanoclay in kinetics, isotherms, and thermodynamics
- Antimicrobial activity, induction of ROS generation in HepG2 liver cancer cells, and chemical composition of Pterospermum heterophyllum
- Study on the performance of nanoparticle-modified PVDF membrane in delaying membrane aging
- Impact of cholesterol in encapsulated vitamin E acetate within cocoliposomes
- Review Articles
- Structural aspects of Pt(η3-X1N1X2)(PL) (X1,2 = O, C, or Se) and Pt(η3-N1N2X1)(PL) (X1 = C, S, or Se) derivatives
- Biosurfactants in biocorrosion and corrosion mitigation of metals: An overview
- Stimulus-responsive MOF–hydrogel composites: Classification, preparation, characterization, and their advancement in medical treatments
- Electrochemical dissolution of titanium under alternating current polarization to obtain its dioxide
- Special Issue on Recent Trends in Green Chemistry
- Phytochemical screening and antioxidant activity of Vitex agnus-castus L.
- Phytochemical study, antioxidant activity, and dermoprotective activity of Chenopodium ambrosioides (L.)
- Exploitation of mangliculous marine fungi, Amarenographium solium, for the green synthesis of silver nanoparticles and their activity against multiple drug-resistant bacteria
- Study of the phytotoxicity of margines on Pistia stratiotes L.
- Special Issue on Advanced Nanomaterials for Energy, Environmental and Biological Applications - Part III
- Impact of biogenic zinc oxide nanoparticles on growth, development, and antioxidant system of high protein content crop (Lablab purpureus L.) sweet
- Green synthesis, characterization, and application of iron and molybdenum nanoparticles and their composites for enhancing the growth of Solanum lycopersicum
- Green synthesis of silver nanoparticles from Olea europaea L. extracted polysaccharides, characterization, and its assessment as an antimicrobial agent against multiple pathogenic microbes
- Photocatalytic treatment of organic dyes using metal oxides and nanocomposites: A quantitative study
- Antifungal, antioxidant, and photocatalytic activities of greenly synthesized iron oxide nanoparticles
- Special Issue on Phytochemical and Pharmacological Scrutinization of Medicinal Plants
- Hepatoprotective effects of safranal on acetaminophen-induced hepatotoxicity in rats
- Chemical composition and biological properties of Thymus capitatus plants from Algerian high plains: A comparative and analytical study
- Chemical composition and bioactivities of the methanol root extracts of Saussurea costus
- In vivo protective effects of vitamin C against cyto-genotoxicity induced by Dysphania ambrosioides aqueous extract
- Insights about the deleterious impact of a carbamate pesticide on some metabolic immune and antioxidant functions and a focus on the protective ability of a Saharan shrub and its anti-edematous property
- A comprehensive review uncovering the anticancerous potential of genkwanin (plant-derived compound) in several human carcinomas
- A study to investigate the anticancer potential of carvacrol via targeting Notch signaling in breast cancer
- Assessment of anti-diabetic properties of Ziziphus oenopolia (L.) wild edible fruit extract: In vitro and in silico investigations through molecular docking analysis
- Optimization of polyphenol extraction, phenolic profile by LC-ESI-MS/MS, antioxidant, anti-enzymatic, and cytotoxic activities of Physalis acutifolia
- Phytochemical screening, antioxidant properties, and photo-protective activities of Salvia balansae de Noé ex Coss
- Antihyperglycemic, antiglycation, anti-hypercholesteremic, and toxicity evaluation with gas chromatography mass spectrometry profiling for Aloe armatissima leaves
- Phyto-fabrication and characterization of gold nanoparticles by using Timur (Zanthoxylum armatum DC) and their effect on wound healing
- Does Erodium trifolium (Cav.) Guitt exhibit medicinal properties? Response elements from phytochemical profiling, enzyme-inhibiting, and antioxidant and antimicrobial activities
- Integrative in silico evaluation of the antiviral potential of terpenoids and its metal complexes derived from Homalomena aromatica based on main protease of SARS-CoV-2
- 6-Methoxyflavone improves anxiety, depression, and memory by increasing monoamines in mice brain: HPLC analysis and in silico studies
- Simultaneous extraction and quantification of hydrophilic and lipophilic antioxidants in Solanum lycopersicum L. varieties marketed in Saudi Arabia
- Biological evaluation of CH3OH and C2H5OH of Berberis vulgaris for in vivo antileishmanial potential against Leishmania tropica in murine models
Articles in the same Issue
- Regular Articles
- Porous silicon nanostructures: Synthesis, characterization, and their antifungal activity
- Biochar from de-oiled Chlorella vulgaris and its adsorption on antibiotics
- Phytochemicals profiling, in vitro and in vivo antidiabetic activity, and in silico studies on Ajuga iva (L.) Schreb.: A comprehensive approach
- Synthesis, characterization, in silico and in vitro studies of novel glycoconjugates as potential antibacterial, antifungal, and antileishmanial agents
- Sonochemical synthesis of gold nanoparticles mediated by potato starch: Its performance in the treatment of esophageal cancer
- Computational study of ADME-Tox prediction of selected phytochemicals from Punica granatum peels
- Phytochemical analysis, in vitro antioxidant and antifungal activities of extracts and essential oil derived from Artemisia herba-alba Asso
- Two triazole-based coordination polymers: Synthesis and crystal structure characterization
- Phytochemical and physicochemical studies of different apple varieties grown in Morocco
- Synthesis of multi-template molecularly imprinted polymers (MT-MIPs) for isolating ethyl para-methoxycinnamate and ethyl cinnamate from Kaempferia galanga L., extract with methacrylic acid as functional monomer
- Nutraceutical potential of Mesembryanthemum forsskaolii Hochst. ex Bioss.: Insights into its nutritional composition, phytochemical contents, and antioxidant activity
- Evaluation of influence of Butea monosperma floral extract on inflammatory biomarkers
- Cannabis sativa L. essential oil: Chemical composition, anti-oxidant, anti-microbial properties, and acute toxicity: In vitro, in vivo, and in silico study
- The effect of gamma radiation on 5-hydroxymethylfurfural conversion in water and dimethyl sulfoxide
- Hollow mushroom nanomaterials for potentiometric sensing of Pb2+ ions in water via the intercalation of iodide ions into the polypyrrole matrix
- Determination of essential oil and chemical composition of St. John’s Wort
- Computational design and in vitro assay of lantadene-based novel inhibitors of NS3 protease of dengue virus
- Anti-parasitic activity and computational studies on a novel labdane diterpene from the roots of Vachellia nilotica
- Microbial dynamics and dehydrogenase activity in tomato (Lycopersicon esculentum Mill.) rhizospheres: Impacts on growth and soil health across different soil types
- Correlation between in vitro anti-urease activity and in silico molecular modeling approach of novel imidazopyridine–oxadiazole hybrids derivatives
- Spatial mapping of indoor air quality in a light metro system using the geographic information system method
- Iron indices and hemogram in renal anemia and the improvement with Tribulus terrestris green-formulated silver nanoparticles applied on rat model
- Integrated track of nano-informatics coupling with the enrichment concept in developing a novel nanoparticle targeting ERK protein in Naegleria fowleri
- Cytotoxic and phytochemical screening of Solanum lycopersicum–Daucus carota hydro-ethanolic extract and in silico evaluation of its lycopene content as anticancer agent
- Protective activities of silver nanoparticles containing Panax japonicus on apoptotic, inflammatory, and oxidative alterations in isoproterenol-induced cardiotoxicity
- pH-based colorimetric detection of monofunctional aldehydes in liquid and gas phases
- Investigating the effect of resveratrol on apoptosis and regulation of gene expression of Caco-2 cells: Unravelling potential implications for colorectal cancer treatment
- Metformin inhibits knee osteoarthritis induced by type 2 diabetes mellitus in rats: S100A8/9 and S100A12 as players and therapeutic targets
- Effect of silver nanoparticles formulated by Silybum marianum on menopausal urinary incontinence in ovariectomized rats
- Synthesis of new analogs of N-substituted(benzoylamino)-1,2,3,6-tetrahydropyridines
- Response of yield and quality of Japonica rice to different gradients of moisture deficit at grain-filling stage in cold regions
- Preparation of an inclusion complex of nickel-based β-cyclodextrin: Characterization and accelerating the osteoarthritis articular cartilage repair
- Empagliflozin-loaded nanomicelles responsive to reactive oxygen species for renal ischemia/reperfusion injury protection
- Preparation and pharmacodynamic evaluation of sodium aescinate solid lipid nanoparticles
- Assessment of potentially toxic elements and health risks of agricultural soil in Southwest Riyadh, Saudi Arabia
- Theoretical investigation of hydrogen-rich fuel production through ammonia decomposition
- Biosynthesis and screening of cobalt nanoparticles using citrus species for antimicrobial activity
- Investigating the interplay of genetic variations, MCP-1 polymorphism, and docking with phytochemical inhibitors for combatting dengue virus pathogenicity through in silico analysis
- Ultrasound induced biosynthesis of silver nanoparticles embedded into chitosan polymers: Investigation of its anti-cutaneous squamous cell carcinoma effects
- Copper oxide nanoparticles-mediated Heliotropium bacciferum leaf extract: Antifungal activity and molecular docking assays against strawberry pathogens
- Sprouted wheat flour for improving physical, chemical, rheological, microbial load, and quality properties of fino bread
- Comparative toxicity assessment of fisetin-aided artificial intelligence-assisted drug design targeting epibulbar dermoid through phytochemicals
- Acute toxicity and anti-inflammatory activity of bis-thiourea derivatives
- Anti-diabetic activity-guided isolation of α-amylase and α-glucosidase inhibitory terpenes from Capsella bursa-pastoris Linn.
- GC–MS analysis of Lactobacillus plantarum YW11 metabolites and its computational analysis on familial pulmonary fibrosis hub genes
- Green formulation of copper nanoparticles by Pistacia khinjuk leaf aqueous extract: Introducing a novel chemotherapeutic drug for the treatment of prostate cancer
- Improved photocatalytic properties of WO3 nanoparticles for Malachite green dye degradation under visible light irradiation: An effect of La doping
- One-pot synthesis of a network of Mn2O3–MnO2–poly(m-methylaniline) composite nanorods on a polypyrrole film presents a promising and efficient optoelectronic and solar cell device
- Groundwater quality and health risk assessment of nitrate and fluoride in Al Qaseem area, Saudi Arabia
- A comparative study of the antifungal efficacy and phytochemical composition of date palm leaflet extracts
- Processing of alcohol pomelo beverage (Citrus grandis (L.) Osbeck) using saccharomyces yeast: Optimization, physicochemical quality, and sensory characteristics
- Specialized compounds of four Cameroonian spices: Isolation, characterization, and in silico evaluation as prospective SARS-CoV-2 inhibitors
- Identification of a novel drug target in Porphyromonas gingivalis by a computational genome analysis approach
- Physico-chemical properties and durability of a fly-ash-based geopolymer
- FMS-like tyrosine kinase 3 inhibitory potentials of some phytochemicals from anti-leukemic plants using computational chemical methodologies
- Wild Thymus zygis L. ssp. gracilis and Eucalyptus camaldulensis Dehnh.: Chemical composition, antioxidant and antibacterial activities of essential oils
- 3D-QSAR, molecular docking, ADMET, simulation dynamic, and retrosynthesis studies on new styrylquinolines derivatives against breast cancer
- Deciphering the influenza neuraminidase inhibitory potential of naturally occurring biflavonoids: An in silico approach
- Determination of heavy elements in agricultural regions, Saudi Arabia
- Synthesis and characterization of antioxidant-enriched Moringa oil-based edible oleogel
- Ameliorative effects of thistle and thyme honeys on cyclophosphamide-induced toxicity in mice
- Study of phytochemical compound and antipyretic activity of Chenopodium ambrosioides L. fractions
- Investigating the adsorption mechanism of zinc chloride-modified porous carbon for sulfadiazine removal from water
- Performance repair of building materials using alumina and silica composite nanomaterials with electrodynamic properties
- Effects of nanoparticles on the activity and resistance genes of anaerobic digestion enzymes in livestock and poultry manure containing the antibiotic tetracycline
- Effect of copper nanoparticles green-synthesized using Ocimum basilicum against Pseudomonas aeruginosa in mice lung infection model
- Cardioprotective effects of nanoparticles green formulated by Spinacia oleracea extract on isoproterenol-induced myocardial infarction in mice by the determination of PPAR-γ/NF-κB pathway
- Anti-OTC antibody-conjugated fluorescent magnetic/silica and fluorescent hybrid silica nanoparticles for oxytetracycline detection
- Curcumin conjugated zinc nanoparticles for the treatment of myocardial infarction
- Identification and in silico screening of natural phloroglucinols as potential PI3Kα inhibitors: A computational approach for drug discovery
- Exploring the phytochemical profile and antioxidant evaluation: Molecular docking and ADMET analysis of main compounds from three Solanum species in Saudi Arabia
- Unveiling the molecular composition and biological properties of essential oil derived from the leaves of wild Mentha aquatica L.: A comprehensive in vitro and in silico exploration
- Analysis of bioactive compounds present in Boerhavia elegans seeds by GC-MS
- Homology modeling and molecular docking study of corticotrophin-releasing hormone: An approach to treat stress-related diseases
- LncRNA MIR17HG alleviates heart failure via targeting MIR17HG/miR-153-3p/SIRT1 axis in in vitro model
- Development and validation of a stability indicating UPLC-DAD method coupled with MS-TQD for ramipril and thymoquinone in bioactive SNEDDS with in silico toxicity analysis of ramipril degradation products
- Biosynthesis of Ag/Cu nanocomposite mediated by Curcuma longa: Evaluation of its antibacterial properties against oral pathogens
- Development of AMBER-compliant transferable force field parameters for polytetrafluoroethylene
- Treatment of gestational diabetes by Acroptilon repens leaf aqueous extract green-formulated iron nanoparticles in rats
- Development and characterization of new ecological adsorbents based on cardoon wastes: Application to brilliant green adsorption
- A fast, sensitive, greener, and stability-indicating HPLC method for the standardization and quantitative determination of chlorhexidine acetate in commercial products
- Assessment of Se, As, Cd, Cr, Hg, and Pb content status in Ankang tea plantations of China
- Effect of transition metal chloride (ZnCl2) on low-temperature pyrolysis of high ash bituminous coal
- Evaluating polyphenol and ascorbic acid contents, tannin removal ability, and physical properties during hydrolysis and convective hot-air drying of cashew apple powder
- Development and characterization of functional low-fat frozen dairy dessert enhanced with dried lemongrass powder
- Scrutinizing the effect of additive and synergistic antibiotics against carbapenem-resistant Pseudomonas aeruginosa
- Preparation, characterization, and determination of the therapeutic effects of copper nanoparticles green-formulated by Pistacia atlantica in diabetes-induced cardiac dysfunction in rat
- Antioxidant and antidiabetic potentials of methoxy-substituted Schiff bases using in vitro, in vivo, and molecular simulation approaches
- Anti-melanoma cancer activity and chemical profile of the essential oil of Seseli yunnanense Franch
- Molecular docking analysis of subtilisin-like alkaline serine protease (SLASP) and laccase with natural biopolymers
- Overcoming methicillin resistance by methicillin-resistant Staphylococcus aureus: Computational evaluation of napthyridine and oxadiazoles compounds for potential dual inhibition of PBP-2a and FemA proteins
- Exploring novel antitubercular agents: Innovative design of 2,3-diaryl-quinoxalines targeting DprE1 for effective tuberculosis treatment
- Drimia maritima flowers as a source of biologically potent components: Optimization of bioactive compound extractions, isolation, UPLC–ESI–MS/MS, and pharmacological properties
- Estimating molecular properties, drug-likeness, cardiotoxic risk, liability profile, and molecular docking study to characterize binding process of key phyto-compounds against serotonin 5-HT2A receptor
- Fabrication of β-cyclodextrin-based microgels for enhancing solubility of Terbinafine: An in-vitro and in-vivo toxicological evaluation
- Phyto-mediated synthesis of ZnO nanoparticles and their sunlight-driven photocatalytic degradation of cationic and anionic dyes
- Monosodium glutamate induces hypothalamic–pituitary–adrenal axis hyperactivation, glucocorticoid receptors down-regulation, and systemic inflammatory response in young male rats: Impact on miR-155 and miR-218
- Quality control analyses of selected honey samples from Serbia based on their mineral and flavonoid profiles, and the invertase activity
- Eco-friendly synthesis of silver nanoparticles using Phyllanthus niruri leaf extract: Assessment of antimicrobial activity, effectiveness on tropical neglected mosquito vector control, and biocompatibility using a fibroblast cell line model
- Green synthesis of silver nanoparticles containing Cichorium intybus to treat the sepsis-induced DNA damage in the liver of Wistar albino rats
- Quality changes of durian pulp (Durio ziberhinus Murr.) in cold storage
- Study on recrystallization process of nitroguanidine by directly adding cold water to control temperature
- Determination of heavy metals and health risk assessment in drinking water in Bukayriyah City, Saudi Arabia
- Larvicidal properties of essential oils of three Artemisia species against the chemically insecticide-resistant Nile fever vector Culex pipiens (L.) (Diptera: Culicidae): In vitro and in silico studies
- Design, synthesis, characterization, and theoretical calculations, along with in silico and in vitro antimicrobial proprieties of new isoxazole-amide conjugates
- The impact of drying and extraction methods on total lipid, fatty acid profile, and cytotoxicity of Tenebrio molitor larvae
- A zinc oxide–tin oxide–nerolidol hybrid nanomaterial: Efficacy against esophageal squamous cell carcinoma
- Research on technological process for production of muskmelon juice (Cucumis melo L.)
- Physicochemical components, antioxidant activity, and predictive models for quality of soursop tea (Annona muricata L.) during heat pump drying
- Characterization and application of Fe1−xCoxFe2O4 nanoparticles in Direct Red 79 adsorption
- Torilis arvensis ethanolic extract: Phytochemical analysis, antifungal efficacy, and cytotoxicity properties
- Magnetite–poly-1H pyrrole dendritic nanocomposite seeded on poly-1H pyrrole: A promising photocathode for green hydrogen generation from sanitation water without using external sacrificing agent
- HPLC and GC–MS analyses of phytochemical compounds in Haloxylon salicornicum extract: Antibacterial and antifungal activity assessment of phytopathogens
- Efficient and stable to coking catalysts of ethanol steam reforming comprised of Ni + Ru loaded on MgAl2O4 + LnFe0.7Ni0.3O3 (Ln = La, Pr) nanocomposites prepared via cost-effective procedure with Pluronic P123 copolymer
- Nitrogen and boron co-doped carbon dots probe for selectively detecting Hg2+ in water samples and the detection mechanism
- Heavy metals in road dust from typical old industrial areas of Wuhan: Seasonal distribution and bioaccessibility-based health risk assessment
- Phytochemical profiling and bioactivity evaluation of CBD- and THC-enriched Cannabis sativa extracts: In vitro and in silico investigation of antioxidant and anti-inflammatory effects
- Investigating dye adsorption: The role of surface-modified montmorillonite nanoclay in kinetics, isotherms, and thermodynamics
- Antimicrobial activity, induction of ROS generation in HepG2 liver cancer cells, and chemical composition of Pterospermum heterophyllum
- Study on the performance of nanoparticle-modified PVDF membrane in delaying membrane aging
- Impact of cholesterol in encapsulated vitamin E acetate within cocoliposomes
- Review Articles
- Structural aspects of Pt(η3-X1N1X2)(PL) (X1,2 = O, C, or Se) and Pt(η3-N1N2X1)(PL) (X1 = C, S, or Se) derivatives
- Biosurfactants in biocorrosion and corrosion mitigation of metals: An overview
- Stimulus-responsive MOF–hydrogel composites: Classification, preparation, characterization, and their advancement in medical treatments
- Electrochemical dissolution of titanium under alternating current polarization to obtain its dioxide
- Special Issue on Recent Trends in Green Chemistry
- Phytochemical screening and antioxidant activity of Vitex agnus-castus L.
- Phytochemical study, antioxidant activity, and dermoprotective activity of Chenopodium ambrosioides (L.)
- Exploitation of mangliculous marine fungi, Amarenographium solium, for the green synthesis of silver nanoparticles and their activity against multiple drug-resistant bacteria
- Study of the phytotoxicity of margines on Pistia stratiotes L.
- Special Issue on Advanced Nanomaterials for Energy, Environmental and Biological Applications - Part III
- Impact of biogenic zinc oxide nanoparticles on growth, development, and antioxidant system of high protein content crop (Lablab purpureus L.) sweet
- Green synthesis, characterization, and application of iron and molybdenum nanoparticles and their composites for enhancing the growth of Solanum lycopersicum
- Green synthesis of silver nanoparticles from Olea europaea L. extracted polysaccharides, characterization, and its assessment as an antimicrobial agent against multiple pathogenic microbes
- Photocatalytic treatment of organic dyes using metal oxides and nanocomposites: A quantitative study
- Antifungal, antioxidant, and photocatalytic activities of greenly synthesized iron oxide nanoparticles
- Special Issue on Phytochemical and Pharmacological Scrutinization of Medicinal Plants
- Hepatoprotective effects of safranal on acetaminophen-induced hepatotoxicity in rats
- Chemical composition and biological properties of Thymus capitatus plants from Algerian high plains: A comparative and analytical study
- Chemical composition and bioactivities of the methanol root extracts of Saussurea costus
- In vivo protective effects of vitamin C against cyto-genotoxicity induced by Dysphania ambrosioides aqueous extract
- Insights about the deleterious impact of a carbamate pesticide on some metabolic immune and antioxidant functions and a focus on the protective ability of a Saharan shrub and its anti-edematous property
- A comprehensive review uncovering the anticancerous potential of genkwanin (plant-derived compound) in several human carcinomas
- A study to investigate the anticancer potential of carvacrol via targeting Notch signaling in breast cancer
- Assessment of anti-diabetic properties of Ziziphus oenopolia (L.) wild edible fruit extract: In vitro and in silico investigations through molecular docking analysis
- Optimization of polyphenol extraction, phenolic profile by LC-ESI-MS/MS, antioxidant, anti-enzymatic, and cytotoxic activities of Physalis acutifolia
- Phytochemical screening, antioxidant properties, and photo-protective activities of Salvia balansae de Noé ex Coss
- Antihyperglycemic, antiglycation, anti-hypercholesteremic, and toxicity evaluation with gas chromatography mass spectrometry profiling for Aloe armatissima leaves
- Phyto-fabrication and characterization of gold nanoparticles by using Timur (Zanthoxylum armatum DC) and their effect on wound healing
- Does Erodium trifolium (Cav.) Guitt exhibit medicinal properties? Response elements from phytochemical profiling, enzyme-inhibiting, and antioxidant and antimicrobial activities
- Integrative in silico evaluation of the antiviral potential of terpenoids and its metal complexes derived from Homalomena aromatica based on main protease of SARS-CoV-2
- 6-Methoxyflavone improves anxiety, depression, and memory by increasing monoamines in mice brain: HPLC analysis and in silico studies
- Simultaneous extraction and quantification of hydrophilic and lipophilic antioxidants in Solanum lycopersicum L. varieties marketed in Saudi Arabia
- Biological evaluation of CH3OH and C2H5OH of Berberis vulgaris for in vivo antileishmanial potential against Leishmania tropica in murine models
![Figure 4
Schematic representation of multifunctional NPs composed of a PPy core and a mesoporous MIL-100 shell [61] Adapted with permission from Zhu et al. [61]. Copyright 2016 American Chemical Society.](/document/doi/10.1515/chem-2024-0061/asset/graphic/j_chem-2024-0061_fig_004.jpg)