Next Article in Journal
Metabolic Alteration Analysis of Steroid Hormones in Niemann–Pick Disease Type C Model Cell Using Liquid Chromatography/Tandem Mass Spectrometry
Previous Article in Journal
Microgravity and Space Medicine 2.0
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Applications of Metal-Organic Frameworks as Drug Delivery Systems

1
Coriolan Dragulescu Institute of Chemistry, 24 Mihai Viteazul Blv., 300223 Timisoara, Romania
2
Department of Biology-Chemistry, Faculty of Chemistry, Biology, Geography, West University Timisoara, 16 Pestalozzi Street, 300115 Timisoara, Romania
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(8), 4458; https://doi.org/10.3390/ijms23084458
Submission received: 21 March 2022 / Revised: 6 April 2022 / Accepted: 15 April 2022 / Published: 18 April 2022
(This article belongs to the Section Molecular Pharmacology)

Abstract

:
In the last decade, metal organic frameworks (MOFs) have shown great prospective as new drug delivery systems (DDSs) due to their unique properties: these materials exhibit fascinating architectures, surfaces, composition, and a rich chemistry of these compounds. The DSSs allow the release of the active pharmaceutical ingredient to accomplish a desired therapeutic response. Over the past few decades, there has been exponential growth of many new classes of coordination polymers, and MOFs have gained popularity over other identified systems due to their higher biocompatibility and versatile loading capabilities. This review presents and assesses the most recent research, findings, and challenges associated with the use of MOFs as DDSs. Among the most commonly used MOFs for investigated-purpose MOFs, coordination polymers and metal complexes based on synthetic and natural polymers, are well known. Specific attention is given to the stimuli- and multistimuli-responsive MOFs-based DDSs. Of great interest in the COVID-19 pandemic is the use of MOFs for combination therapy and multimodal systems.

1. Introduction

It has been reported that DDS contains two main components: a carrier and a drug. The optimal properties of both components are precisely checked by careful selection of these constituents [1].
Conventional DDS are tablets, capsules, granules, and ointments; syrups for oral administration; and suppositories or solutions for intravenous administration. Owing to several disadvantages and various limitations, such as pure absorption for target sites, repeated dosage several times per day, high dose needed, fluctuations in plasma drug level, difficult to monitor, poor bioavailability issues, critical toxicities, side effects, and premature excretion from the body, conventional DSSs are unable to reach sustained release [2]. In this respect, new DDSs are required and being investigated.
Currently, numerous DDSs have been created to reduce side effects and increase therapeutic efficacy. Thus, inorganic materials such as carbon nanotubes [3,4], graphene [5], magnetic nanoparticles (NPs) of iron oxide [6,7], silicon-based materials and their oxides [8,9], and gold nanoparticles [10,11,12] were used.
Metal–organic frameworks (MOFs) are an interesting and versatile class of coordination polymers, assembled from metal ion/clusters and organic linkers. MOFs have diverse properties, such as high porosity and surface area, large pore volume for encapsulation of various substances/gazes, and special chemical and thermal stability [13,14,15,16]. They have been used in various applications, such as gas storage [17], photochemistry [18], catalysis [19,20,21], flame retardants [22,23,24], corrosion protection [25,26], separation processes [27], adsorption properties [28,29,30,31,32,33,34], diagnostic [35], antimicrobial properties [36,37,38], and delivery of a large variety of active drug molecules, cosmetics, and biological gases [39,40,41]. Owing to these characteristics, MOFs have become the most promising materials for use in biomedicine [42,43].
The characteristics, such as nontoxic effects of MOFs, directed and stimulus-based delivery systems, multiple drug loaded properties, and continuous release, have enriched the use of MOFs in drug delivery, biocompatibility, and biodegradability in the last decade. One of its most significant properties, which is continuously explored by scientists, is MOFs capability of interacting with biological systems based on various stimuli encountered inside our system, which includes precise and sustained drug release capacity and improved solubility of amorphous and poorly soluble drugs in variable conditions such as pH, temperature, light, magnetic, pressure, glucose level, multiple stimuli-responsive systems [44,45].
MOFs were originally synthesized by a method called solvothermal synthesis, which has been the most popular technique for MOFs synthesis. After two decades from the first MOFs reference, new improvements in synthesis approaches were accomplished. New methods, which involved the use of mechanochemical, electrochemical, sonochemical, and microwave-assisted technique, have been reported in literature with both advantages and disadvantages, as presented in Figure 1 [46,47,48,49].
Because of the difference in several orders of magnitude between the size of nano-MOFs and the size of cancer cells, MOFs show the ability to association a porous structure with extraordinary drug loading over various kinds of interactions. Therefore, MOFs are perfectly suited as in vivo DDSs. In the past decades, the number of reviews and research papers dedicated to MOFs, Figure 2A, and MOFs and DDSs, Figure 2B, has experienced a steady increase, responding to an emergent request for such smart materials.
A wide variety of factors influence the drug delivery of various MOFs. Besides these factors, we also need to consider the physiochemical properties of MOF together with drug molecule properties, which permit fitting drugs inside the carrier MOF for effective delivery to the desired position. Compared to other molecules that act as drug delivery systems, the mechanism of MOFs drug delivery permits a slow and manageable liberation of drug molecules [50].
Some model examples of drugs delivered by MOFs are presented in the following review, such as: indocianine green (ICG), doxorubicin hydrochloride (DOX), 5-fluorouracil (5-Fu), caffeine, cidofovir, folic acid (FA), calcein (cal) curcumin (Cur), and paracetamol, together with the synthetic classical and new strategies, biocompatibility, and versatility of loading capabilities, Figure 3.
Additionally, the stimuli and multistimuli responsive variety and type of cancer cell used were considered. Stimuli are used to control how drug delivery systems and drug release act and function. Two types of stimuli are explored: endogenous stimuli features of the pathological sites: pH or temperature of local sites, redox status, and hypoxia, and exogenous stimuli: ultrasound irradiation, temperature, light, and magnetic fields.

2. Modification of MOFs—Cargo Loading Strategies

Wang and coworkers represent three cargo loading strategies: encapsulation, direct assembly, and post-synthesis strategy, to load MOFs with larger amounts of drugs, exemplified in Figure 4. The classification of strategies was made taking into account the location of the drug and the effect of cargo and host–guest interactions within the MOF frameworks.
  • Encapsulation strategy: Cargo is situated inside of the MOFs pores or channels by noncovalent bonding interaction. This type of strategy does not alter the MOF framework structures.
  • Direct assembly strategy: The interaction between MOFs and cargo is coordination bonds. The cargo participate in the synthetic reaction as ligands to partly contribute to the construction of MOFs.
  • Post-synthesis strategy: The cargo molecules are located in MOFs surfaces. These molecules serve as linkers for pre-synthesized MOFs. The chemical interactions implied by this strategy are coordination bonds and covalent bonds between metal nodes/organic linkers and used cargo. This type of strategy does not alter the MOF framework structures. The second possibility as a post-synthesis strategy is adsorption on MOF surfaces. In general, the main forces in adsorption are weak interactions, such as π–π and Van der Waals interactions, and hydrogen bonding.

2.1. Encapsulation Strategy

An extensive series of hydrophilic, hydrophobic, and amphiphilic drug molecules were encapsulated in the MOFs. The library of investigated MOF included in this review—MIL-53(Fe), MIL-53-NH2 (Fe),MOF-74 (Fe),MOF-74 (Zn), MIL-101 (Cr), MIL-100 (Fe), MIL-89 (Fe), MIL-101-NH2 (Fe), MIL-100 (Fe), UiO-66(Zr), MIL-127 (Fe), MIL-100 (Fe), Mi-UiO-68, MIL-88@ZIF-8, ZIF-8, UiO-66-NH2, MIL-88A(Fe)—use various metal–organic framework templates, such as HKUST, UiO, ZIF, MIL, and MOF-78. The MOFs investigated have various stabilities in biological environments.
For biomedical and pharmaceutical applications, the functionalization of MOFs represents a new area of research for binding therapeutic molecules on MOFs surfaces. However, the desired modification must not only reduce the interaction and increase the stability of MOFs with the biologic medium, but it must also simplify the drugs’ passage through the physiological barriers in order to reach the targeted delivery center. Numerous ways were investigated to modify the surface of MOFs materials by optimizing the modification conditions and applying diverse biomolecules, polymers, and ligands, among other molecules [52,53,54]. Recall the excellent physical and chemical properties of MOFs.
Various parameters, such as pH value, an optimal buffer, nanoparticle size, and surface adaptation, are very important, in addition to analytical approaches and methodologies appropriate to control the MOF stability. The chemical stability of an MOF is unstated as the capacity to keep the long-ranged ordered structure when the material is exposed to a specific environment. Normally, there are two central ways of measuring material degradation: decomposition to the original building blocks (in this way, the components can supplementarily react with various compounds existing in the environment) and amorphization. The chemical stability of MOFs is explained on the basis of pH, time, and temperature.
A large number of researchers combined their efforts to explain MOF stability in acids and bases. In an acidic environment, the MOF degradation is principally caused by competing protons and metal ions trying to coordinate with the organic ligand. In a basic environment, MOF decomposition involves the replacement of organic ligands by hydroxide ions. Consequently, MOFs based on high-valent metal ions and carboxylate ligands are estimated to be rather stable in acids and less resistant to bases. Inversely, the MOFs based on soft divalent metal ions and nitrogen-based ligands are expected to be more stable in basic solutions and further unstable in acids.
The chemical stability of MOFs and other materials used as drug delivery systems as a function of time is a very important factor to be investigated. Optimal stability is influenced by the application and route of drug administration. Depending on applications, the time range can vary from several days to a few hours, which might be sufficient. As a target, 24 h should be implemented as the initial treatment time, with the possibility of extended times depending on the initial results and needs. For applications in nanomedicine, the preferred temperature is 37 °C in studies involving the chemical stability of MOFs [55].
Demel and colleagues explored the stability of UiO-66 MOF in four different buffers: N-ethylmorpholine (NEM), phosphate buffer (PB), 4-(2-hydroxyethyl) piperazine-1-ethane sulfonic acid (HEPES), and 2-amino-2-(hydroxymethyl)-1,3-propanediol (TRIS). The work demonstrated that the fate of zirconium is a major concern in UiO-66 stability. TRIS and PB buffers, depending on concentration, can act as ligands to coordinate with zirconium ions. The kinetics of terephthalate release are accelerated by increasing the temperature between 25 and 37 degrees Celsius [56].
Particle size is an important property of a nanocarrier. Depending on the intended application and the route of administration, different MOF particle sizes are preferred [57]. Particles ranging from 1 to 100 nm exhibit distinctive physicochemical properties. This size allows them to be useful as drug carriers, therapeutics, and profound tissue contrast agents [58]. For instance, nanoparticles smaller than 200 nm are believed to be capable of passive targeting in cancer therapy due to their enhanced permeability and retention effect [59].
To conclude, in gastrointestinal treatments, the mesh pore arrangement of the intestinal mucosa wall (50–1800 nm) works as a sieve that defines which particles are absorbed [60]. The size of MOFs and various NPs used as DDS influenced movement half-life and tumor growth.
Researchers have validated Lewis acid catalytic activity stemming from open metal sites in MOFs. By carefully choosing MOF materials that possesses open metal sites and functionalizing an organic ligand with an organic basic group, in order to obtain materials that present both basic and acidic functional sites [61].
The surface modification of MOFs aids in improving their water stability and drug-loading enrichment, thus altering the degradation pattern and controlling the release of drug molecules.
Indocyanine green (ICG), a water-soluble photosensitizer dye permitted by the US Food and Drug Administration (US FDA) for clinical usage, is being broadly investigated in phototherapy, together with iron oxide nanoparticles—the only nanoparticles approved for clinical practice [62]. As a complex molecule composed of two hydrophobic polycyclic parts holding two additional sulfonate anions at each terminal site (Figure 3a), ICG is capable of interacting with several metal ions, such as the common coordination metal ions Zn2+, Mn2+, Fe3+, and Pt2+, and additional trace metals that take part in maintaining homeostasis [63,64]. In the near-infrared (NIR) region, ICG displays light absorption (λabs, max  =  780 nm) and fluorescence emission (λem, max  =  800 nm) applied to biologically transparent systems [65]. Yang and coworkers investigated one-step encapsulation of ICG in ZIF-8 nanoparticles (NPs). The encapsulation was evidenced using near infrared because ICG@ZIF-8 possesses an absorption band at 783 nm and possesses high photothermal conversion efficiency. In vivo and in vitro studies were performed which reveal that ICG@ZIF-8 after loading DOX forms ICG@ZIF-8-DOX NPs, which display the chemical and photothermal synergistic treatment for tumors [66]. In some cases, ICG effectiveness is limited by photobleaching and its tissue selectivity is low.
Doxorubicin hydrochloride (DOX) is the first line chemotherapeutic drug for various types of cancer such as lymphoblastic leukemia, ovarian cancer, and breast cancer [67,68], Figure 3b.
Yao and coworkers synthesized a water-stable MIL-101(Fe)-C4H4 via a microwave-assisted technique starting from iron (III) chloride hexahydrate and naphthalene-1,4-dicarboxylic acid. The nano-MOFs operated as sponges when saturated in DOX alkaline aqueous solution with an experimental loading capacity of up to 24.5 wt%, with and an upholding loading efficiency of about 98% [69].
5-Fluorouracil (5-Fu), a pyrimidine analogue (Figure 3c), has been an important anticancer agent for managing an extensive diversity of tumors because it can be combined with DNA and RNA, leading to cancer cell death. 5-Fu, a small-molecular-weight drug, was investigated and used in biomedical applications as a model anticancer drug [70,71].
Caffeine, Figure 3d, is a liporeductor and an amphiphilic drug, with an effective one molecule volume of 165.5 Å3. Caffeine molecules were often investigated as a model active component; many preparations of caffeine were contained within numerous diverse matrices. The combination with MOFs makes these cosmetic-containing composite MOF-based patches encouraging nominees for new devices applied on the skin and used in cosmetic applications [72]. Liedana and coworkers investigated one-pot encapsulation of caffeine drugs in ZIF-8 a zeolitic imidazolate MOF [73]. This one-pot strategy reveals a larger quantity of caffeine encapsulated when compared with the encapsulation in two steps, which was presented by Gaudin and co-workers [74] using MIL-88B_2OH as a drug delivery system. Furthermore, by comparing the two MOFs as caffeine drug delivery systems, it was observed that ZIF-8 provides a slower delivery, most likely due to the strong interaction formed between caffeine and ZIF-8.
With an excellent affinity for the folate receptors, folic acid (FA, Figure 3f) is used in combination with various MOFs as drug carriers to reduce the toxicity of anticancer drugs to normal cells and increase the anticancer effects on cancer cells [75,76]. In most cases, the folic acid receptors are overstretched on the external surface of cancer cells. Therefore, FA is broadly used as a model targeting drug for proficiency in delivering drugs into tumor cells [77]. Owing to its high stability and good compatibility in physiological environments, FA acts as a perfect targeting molecule for delivery systems.
Calceine (cal), a hydrophylic fluorescent molecule, nontoxic to the cell, which can be traced by confocal microscopy, was encapsulated by Orellana-Tavra and coworkers in UiO-66 MOF as DDS, forming cal@UiO-66 MOF. UiO-66 has a cubic structure with two types of cavities (ca. 11 and 8 Å and uses Zr—a low toxicity metal). The material was ball-milled, forming an amorphous material named cal@amUiO-66 MOF. The ball-milling amorphization process allows the entrapment of the calceine inside the porosity. This process significantly increases drug release times from 2 days (cal@UiO-66 MOF) up to 30 days (cal@amUiO-66) in a drug delivery system (DDS) [78]. Calcein-loaded MOFs triggered higher cargo internalization related to the free molecule of calcein; enhanced cargo internalization, minimizing undesired side effects. Cal@MIL-101(Cr) outperformed Cal@UiO-66 at concentrations higher than 10 mg m L−1, while the calcein transport efficacy was appropriate at 1 mg mL−1 with an enriched uptake related to free calcein, which was higher after 24 h [79].
Over the past decades, curcumin, the active constituent of the Curcuma longa plant, has received great consideration as an anti-inflammatory agent, antioxidant, and anticancer representative [80]. Metal–organic frameworks (MOFs) are employed as possible drug-delivery systems for controlling the pharmacokinetic rate of drug delivery. Combining two MOF-74 (a solid with a pore size of 12 nm, designed by combining 2,5-dihydroxyterephthalic acid with divalent metal ions Mg2+ and Zn2+ (Mg-MOF-74 and Zn-MOF-74) with different solubility of the metal nodes, Tomeh and coworkers investigated the pharmacokinetic eliberation rate of curcumin as a drug. In the study, the ratio of Mg-MOF-74 and Zn-MOF-74 was modified to the following limits: 80:20, 60:40, 40:60, and 20:80 wt% Mg:Zn to control the pharmacokinetic concern at a rate of 30 wt% curcumin. The drug delivery trials were performed in saline phosphate buffered by varying the time from 0 to 24 h, and keeping the temperature at 37.4 °C. The revealed amount of curcumin was found to be higher when the concentration of Mg-MOF-74 was higher. The drug concentration heightened both the raw delivery and the pharmacokinetic rate of drug release.
Gao and coworkers [81] fabricated a nanoplatform with extended circulating properties. UiO-66, a zirconium-based MOF, was used as a transporter for O2 storage. UiO-66 was linked with indocyanine green (ICG) as a first step by coordination reaction, followed by surface modification by covering with red blood cell (RBC) membranes, Figure 5.
C. Chu and co-workers [82] investigated zinc(II) coordination-based nanoformulations (zinc(II)-dipicolylamine) adept at loading indocyanine green (ICG) and therapeutic genes. For various therapy strategies, such as fluorescence behavior combined with photoacoustic-imaging-guided photo/gene therapy and zinc(II)-dipicolylamine-based photosensitizer metal–organic nanostructures, were synthesized to accomplish various therapy strategies. Considerable work has been performed in this direction and valuable research outcomes have been evidenced. Nevertheless, challenges still remain for clinical interpretation, and because of this some important factors are still unsatisfied, such as biocompatibility, limited efficacy, and complex procedure of synthesis.
Fang and colleagues [83] published a useful study of a versatile nanoplatform composed of core-shell ZnS@ZIF-8 nanoparticles combined with ICG and tirapazamine (TPZ), denoted as ZnS@ZIF-8/ICG/TPZ (ZSZIT), designed and prepared to enable H2S-sensitized chemo- or photodynamic synergistic therapy. The degradation of core–shell structure and H2S release properties are important features of nanoparticles for the antitumor purposes. The study exhibit a considerable anticancer effect, in vitro and in vivo, produced by the joint effects of intracellular reactive oxygen species, H2S, and activated TPZ empowered by ZnS@ZIF-8/ICG/TPZ nanoparticles.
You and coworkers studied the effect of a Pt-based oxygen nanogenerator UiO-66-like MOF [84], which can continuously use endogenous H2O2 inside the tumor microenvironment of the solid tumor by the contained Pt nanozyme, enriching the oxygen concentration in situ, which considerably improved tumor hypoxia and supplementary heightened the effect of photodynamic therapy. The stable, porous Zr-BDC—a sustainable substitute to the conventional UiO-66(Zr) MOF—enables the loading of large volumes of Pt nanozymes.
Fan and coworkers made a multimodal imaging-led synergistic cancer photoimmunotherapy by using MIL101-NH2 dual-dressed with photoacoustic and fluorescent indicator donors ICG and immune adjuvants (CpG ODN- cytosine-phosphate-guanine sequence-oligo nucleotides) as the core transporter. The strategy behind this material is to use it to tune cold tumors to hot. When light activates the photothermal therapy, the immune system is triggered to show a long-acting antitumor result. The dual-dressed MIL101-NH2 brings a favorable method used for the prognosis of cancer and its treatment [85].
Wu and coworkers constructed a photosensitizer, ICG, and a chemotherapy-induced drug, DOX, which were encapsulated step by step inside the nanopores of the MIL-88 core and ZIF-8 shell to exhibit a synergistic photothermal, photodynamic, and chemotherapy nanoplatform. Both coreshell MOFs, MIL-88, and ZIF-8 have two functional regions for codrug delivery. DOX relief could be triggered by a low pH value and the speed of this effect is increased by near-infrared light irradiation [86].
Su and coworkers [87] encapsulated curcumin and ICG into ZIF-8 pores in a one-pot synthesis method and synthesized the compound entitled ICG&Cur@ZIF-8, which exhibits pH-controlled drug delivery and good photothermal performance. The codelivery of two drugs at the same time opens the way to new opportunities for biomedical application. The percentage of drug loading capacity was calculated using the experimental UV–VIS spectra, at 9.6 wt% for Cur and 12.3 wt% for ICG, respectively.

2.2. Direct Assembly Strategy

Magnetic nanoparticles of MOF act as a drug delivery system by directing the cure treatment location without disturbing the other cells of the body. In this respect, Gautam and coworkers synthesized a porous and flexible Cu-based (MOF), Cu-benzene tricarboxylic acid, well known as HKUST-1. The crystalline growth of Cu-BTC has been optimized using a hydrothermal technique. A selection of nontoxic solvents was used in the synthesis [88]. The absorption of paracetamol medication on HKUST-1 at 10 h and 48 h is notable.
Yang and coworkers [89] fabricated a nano-MOF by using a direct assembly strategy. In order to test the compound as photothermal therapy, a MOF composed of Mn2+ and IR825 near-infrared (NIR) dye was synthesized. The prepared compound was named Mn-IR825. The first step was to coat it with a polydopamine (PDA) shell. The last step was to functionalize it with polyethylene glycol (PEG), thus becoming Mn-IR825@PDA−PEG, a compound with a great photostability. The DDS offers strong contrast in T1-weighted MRI, IR825 with strong NIR optical absorbance. The tests were performed on mice and the Mn-IR825@PDA−PEG showed rapid renal excretion, thus decreasing the long-term toxicity effect. During laser irradiation at 808 nm, tumors in mice with Mn-IR825@PDA−PEG injection were totally removed without reappearance within 60 days. Therefore, Mn-IR825@PDA−PEG is a good candidate for applications in cancer theranostics.
Lu and coworkers [90] produced the first chlorin-based metal−organic framework nanomaterial named DBC-UiO, which was used to kill two colon cancer in cell mouse models by applying DBC-UiO-induced photodynamic therapy (PDT).
Rabiee and coworkers [91] used two-dimensional MOF-5 inserted MXene nanostructures in the codelivery of DOX and to increase its bioavailability; their interaction with the pCRISPR was studied. The MXene/MOF-5 was coated with alginate and chitosan. Cytotoxicity tests were conducted on different cell lines such as HeLa, PC12, HEK-293, and HepG2. The results emphasize the outstanding cell viability at very low and high concentrations. The chitosan-coated MOFs presented higher relative cell viability, more than 60% in all cell lines investigated. The alginate-coated MOFs placed second in cell viability, with approximately 50%. Additionally, a significant drug delivery of 35.7% was accomplished, due to the interactions between the MXene/mMOF-5 and the doxorubicin drug.
Sun and coworkers investigated one porphyrinic MOF nanoplatform (H-PMOF) and achieved enhanced photodynamic therapy effectiveness related to nonhollow MOF nanoparticles. Additionally, H-PMOF was investigated as a drug delivery system to coload DOX and indocyanine green (ICG) with an excellent drug-loading ability of 635%. Additionally, cancer cell tissue cover-up of the (DOX and ICG)@H-PMOF composite nanoparticles gave a biomimetic nanoplatform, namely (DOX and ICG)@H-PMOF@mem, with a remaining tumor-targeting and immune-escaping capacity. The following multistimuli: near-infrared laser-triggering and pH-controlled were used for DOX release from (DOX and ICG)@H-PM OF@mem. MOF-based multifunctional nanomedicine development and design are being considered for the combination of cancer healing and specific theranostics [92].

2.3. Post-Synthesis Strategy

Alves and coworkers [93] applied a post-synthesis strategy to reach a tumor-targeting drug-loaded MOF using “click” chemistry. As a starting MOF, N3-bio-MOF-100 was used and folic acid molecules were anchored on its surface. The anticancer model drug used in the investigations was curcumin, measuring the drug-loading encapsulation effectiveness at 24.02 percent after soaking the MOF for 1 day in a curcumin solution and 25.64% after 3 days of soaking. The synthesized MOF DDS was successfully characterized by 1H-NMR, FTIR, and LC–MS. The N3-bio-MOF-100 post-synthesized was investigated as a stimuli-responsive drug release. The studies were performed on the breast cancer 4T1 cell line. Curcumin has been shown to increase in release in acidic microenvironments.
Yang and coworkers used a post-synthetic modification of nanoscale zinc MOFs (nanoMOFs) functionalized with folate to build tumor cell-targeted material, named FA-IRMOF-3 by conjugating the -NH2 group of IRMOF-3 with the -COOH groups of folic acid. The FA-IRMOF-3 was stocked with 5-Fu by impregnating it with 5.0 mg of drug (5-Fu) for 1, 3, 5, or 7 days. During drug loading tests, nanoMOFs functionalized with folate rapidly adsorb up to 24 wt% of 5-fluorouracil while retaining the FA-IRMOF-3’s crystalline structure. In folate receptor-positive KB and HeLa cells, 5-Fu and 5-Fu-FA-IRMOF-3 cytotoxicity were investigated. The results showed that 5-Fu exhibited higher cytotoxicity than 5-Fu-IRMOF-3, and the cytotoxicity of 5-Fu-FA-IRMOF-3 was stronger compared with free 5-Fu [94].
The literature assessment of MOF-based nanomaterials for DDS is presented in Table 1.

3. Stimuli Responsive for MOFs as DDSs

Cai and coworkers [110] explained the basic mechanisms of stimuli-responsive DDSs described in Figure 6. After injection, the stimuli-responsive drug delivery system (MOFs or other types of nanomaterials) gathers at the tumor site and is activated in two ways: one enhanced by permeability and preservation effect, and the other one enhanced by receptor–ligand affinity. When stimulated by specific triggers, the DDSs undergo a transformation, which induces drug delivery at the planned site in a controlled manner. Paralleled with healthy cells, the tumor cells microenvironment has different characteristics: higher temperature or lower pH, which helps to develop internal stimuli-responsive DDS. The external (light, magnetic field, temperature, ultrasound irradiation) stimuli-responsive systems envisage the MOFs gathered at the tumor site and are activated by triggers placed outside the body. The external stimuli-responsive systems offer improved continuous drug release capacity and have established a higher potential for biomedical application.
One way to unlock MOFs and relieve their loads is through the following method: the loaded MOFs with chemical functionalities attached or absorbed on their surface are secured, acting as entrances for the activated release of the loads [111,112,113,114]. Among the MOF nanocarriers generated by outside stimuli, due to the acidic tumor microenvironment, the pH-responsive MOFs are the most broadly investigated [115,116,117]. MOFs based on iron [118], zinc [119,120,121], zirconium [122,123], gadolinium [124], hafnium [125], and europium [126] were investigated for pH responsive stimuli in tumor treatment.
Post-synthetic modification in MOFs, particularly cation exchange, has been a highly explored research area due to its potential applications in various domains, providing novel functional materials. Hamisu and coworkers [127] thoroughly described how the hard–soft acid–base (HSAB) principle permits an overabundance of experimental clarifications that manage cation exchange at the secondary building units (SBUs) of MOFs.
The strength or weakness of MOFs can be affected by numerous factors, such as: interaction of metal ions and organic linkers that describes the strength of the M–L bonds, coordination geometry, pore surface properties, and working environments [128,129]. The HSAB principle is appropriate to define a moderately stable MOF system, in which a soft acid binds more strongly to a soft base, and a hard acid has a preference for a hard base. MOFs prepared from borderline divalent cations and azolate linkers, such as the ZIF-based MOFs ZIF-8 [130], ZIF-71 [131], and ZIF-90 [132], are found to be the most stable for a wide range of pH and a large variety of solvents.
In this respect, stable MOFs, based on hard–hard interactions, imply high valent cations and carboxylate linkers, or those based on soft–soft interaction, the case of low valent cations and azolate linkers, which are more challenging to break and interchange due to strong M–L bonds. In this case, external energy is necessary to overcome the bond dissociation energy and to facilitate the cation exchange.
To achieve improved therapeutic efficacy, supplementary efforts should be dedicated to the development of MOF-based, multiple stimuli-responsive drug delivery systems.

4. Conclusions and Future Perspectives

In conclusion, we studied the recent significant progress in nanometal–organic framework. The library of MOFs investigated in this review included MIL-53(Fe), MIL-53-NH2 (Fe),MOF-74 (Fe),MOF-74 (Zn), MIL-101 (Cr), MIL-100 (Fe), MIL-89 (Fe), MIL-101-NH2 (Fe), MIL-100 (Fe), UiO-66(Zr), MIL-127 (Fe), MIL-100 (Fe), Mi-UiO-68, MIL-88@ZIF-8, ZIF-8, UiO-66-NH2, and MIL-88A(Fe), which use various metal–organic framework templates, such as HKUST, UiO, ZIF, MIL, and MOF-78 as drug delivery systems, and categorized the cargo delivery strategies into three types: encapsulation, direct assembly as a classical approach, and post-synthesis strategy as a new approach. The MOFs investigated have different stabilities in biological surroundings. The surface and available anchorages for a large number of cargos—indocianine green (ICG), doxorubicin hydrochloride (DOX), 5-fluorouracil (5-Fu), caffeine, cidofovir, acid folic (FA), calcein, curcumin and paracetamol—are well established. In comparison to commonly used cargo carriers such as carbon nanotubes, graphenes, magnetic nanoparticles, biomolecules, polymers, inorganic silica, and gold nanoparticles, the study of MOF and its nanoscale development are being investigated using both experimental and computational approaches based on cargo loading and release within the pores of MOF at the atomic level. Among the various classes of NPs compounds used as DDSs, MOFs are considered an encouraging class of nanocarriers for DDSs, exhibiting a well-defined arrangement, extraordinary surface area and porosity, versatile pore size, and easy chemical functionalization. Comparing the two types of stimuli, the external responsive stimuli have greater flexibility. Each stimulus-responsive modality has its own limitations.
However, by surveying the current investigation, there are still numerous tasks that need to be accomplished. Taking into account the biomedical application, it is necessary to carefully investigate the toxicity and solubility of the selected building blocks and metal nodes for synthesized MOFs, as well as the biocompatibility issues. In addition to toxicity, efforts should be made to improve the size and morphology of MOF nanoparticles, to certify the prolonged and retarded blood circulation, and to precisely control the drug release and target the tumor after drug administration. Notwithstanding these challenges, the bioapplication of MOF nanoparticles has significantly improved since the first report on MOF for drug loading in 2006, and MOFs as DDS are expected to achieve extensive advances in nanomedicine.

Author Contributions

Conceptualization, A.V. and B.M.; methodology, A.V. and B.M.; software, A.V. and B.M..; validation, A.V. and B.M.; investigation, A.V. and B.M.; resources, A.V. and B.M..; data curation, A.V. and B.M.; writing—original draft preparation, A.V. and B.M.; writing—review and editing, A.V. and B.M.; visualization, A.V. and B.M.; supervision, A.V. and B.M.; project administration, A.V. and B.M.; funding acquisition, A.V. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the “Coriolan Dragulescu” Institute of Chemistry, Program 2, Project 2.2, and by a grant from the Romanian National Authority for Scientific Research, CNCS—UEFISCDI, project number PN-III-P4-PCE2021-0089.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

5-Fu5-Fluorouracil
CalCalcein
CpG ODNCytosine-phosphate-guanine sequence–oligo nucleotides
CurCurcumin
DDSsDrug delivery systems
DNADeoxyribonucleic acid
HKUSTHong Kong University of Science and Technology
HEPES4-(2-hydroxyethyl) piperazine-1-ethane sulfonic acid
ICGIndocianine green
MOFsMetal Organic Frameworks
US FDAUS Food and Drug Administration
NIRNear-infrared
MILMaterials of Institute Lavoisier
NEMN-ethylmorpholine
DOXDoxorubicin hydrochloride
FAFolic acid
HSABHard–Soft Acid–Base
H-PMOFPorphyrinic MOF nanoplatform
PDTPhotodynamic therapy
PDAPolydopamine
PEGPolyethylene glycol
PB Phosphate buffer
RBCRed blood cell
RNARibonucleic acid
TRIS2-Amino-2-(hydroxymethyl)-1:3-propanediol
UiOUniversitetet I Oslo
ZIFZeolitic imidazole framework

References

  1. Dzhardimalieva, G.I.; Rabinskiy, L.N.; Kydralieva, K.A.; Uflyand, I.E. Recent advances in metallopolymer-based drug delivery systems. RSC Adv. 2019, 9, 37009–37051. [Google Scholar] [CrossRef] [Green Version]
  2. Adepu, S.; Ramakrishna, S. Controlled Drug Delivery Systems: Current Status and Future Directions. Molecules 2021, 26, 5905. [Google Scholar] [CrossRef]
  3. Zhang, W.; Zhang, Z.; Zhang, Y. The application of carbon nanotubes in target drug delivery systems for cancer therapies. Nanoscale Res. Lett. 2011, 6, 555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Rahamathulla, M.; Bhosale, R.R.; Osmani, R.A.M.; Mahima, C.K.; Johnson, P.A.; Hani, U.; Ghazwani, M.; Begum, M.Y.; Alshehri, S.; Ghoneim, M.M.; et al. Carbon nanotubes: Current perspectives on diverse applications in targeted drug delivery and therapies. Materials 2021, 14, 6707. [Google Scholar] [CrossRef] [PubMed]
  5. Jampilek, J.; Kralova, K. Advances in drug delivery nanosystems using graphene-based materials and carbon nanotubes. Materials 2021, 14, 1059. [Google Scholar] [CrossRef]
  6. Aslam, H.; Shukrullah, S.; Naz, M.Y.; Fatima, H.; Hussain, H.; Ullah, S.; Assiri, M.A. Current and future perspectives of multifunctional magnetic nanoparticles based controlled drug delivery systems. J. Drug Deliv. Sci. Technol. 2021, 67, 102946. [Google Scholar] [CrossRef]
  7. Kalubowilage, M.; Janik, K.; Bossmann, S.H. Magnetic Nanomaterials for Magnetically-Aided Drug Delivery and Hyperthermia. Appl. Sci. 2019, 9, 2927. [Google Scholar] [CrossRef] [Green Version]
  8. Twana, M.; Keng, W.; Wing, M.L.; Vitaliy, V.K. Silica nanoparticles in transmucosal drug delivery. Pharmaceutics 2020, 12, 751. [Google Scholar]
  9. Vallet-Regí, M.; Colilla, M.; Izquierdo-Barba, I.; Manzano, M. Mesoporous silica nanoparticles for drug delivery: Current insights. Molecules 2018, 23, 47. [Google Scholar] [CrossRef] [Green Version]
  10. Zhu, F.; Tan, G.; Zhong, Y.; Jiang, Y.; Cai, L.; You, Z.; Liu, S.; Ren, F. Smart nanoplatform for sequential drug release and enhanced chemo-thermal effect of dual drug loaded gold nanorod vesicles for cancer therapy. J. Nanobiotechnol. 2019, 17, 44. [Google Scholar] [CrossRef]
  11. Zhang, H.; Gao, Z.; Li, X.; Li, L.; Ye, S.; Tang, B. Multiple-mRNA-controlled and heat-driven drug release from gold nanocages in targeted chemo-photothermal therapy for tumors. Chem. Sci. 2021, 12, 12429–12436. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, W.; Li, X.; Tang, K.; Song, Z.; Luo, X.A. AuNP-capped cage fluorescent biosensor based on controlled-release and cyclic enzymatic amplification for ultrasensitive detection of ATP. J. Mater. Chem. B 2020, 8, 5945. [Google Scholar] [CrossRef] [PubMed]
  13. Maranescu, B.; Visa, A.; Mracec, M.; Ilia, G.; Maranescu, V.; Simon, Z.; Mracec, M. Lamellar Co2+ vinylphosphonate metal organic framework.PM3 semi-empirical analysis of structural properties. Rev. Roum. Chim. 2011, 56, 473–482. [Google Scholar]
  14. Visa, A.; Mracec, M.; Maranescu, B.; Maranescu, V.; Ilia, G.; Popa, A.; Mracec, M. Structure simulation into a lamellar supramolecular network and calculation of the metal ions/ligands ratio. Chem. Cent. J. 2012, 6, 91. [Google Scholar] [CrossRef] [Green Version]
  15. Visa, A.; Maranescu, B.; Bucur, A.; Iliescu, S.; Demadis, K. Synthesis and characterization of a novel phosphonate metal organic framework starting from copper salts. Phosphorus Sulfur Silicon Relat. Elem. 2014, 189, 630–639. [Google Scholar] [CrossRef]
  16. Maranescu, B.; Visa, A.; Ilia, G.; Simon, Z.; Demadis, K.; Colodrero, R.M.P.; Cabeza, A.; Vallcorba, O.; Rius, J.; Choquesillo-Lazarte, D. Synthesis and characterization of styryl phosphonic acid and its use as new ligand for phosphonate metal organic framework. J. Coord. Chem. 2014, 67, 1562–1572. [Google Scholar] [CrossRef]
  17. Li, H.; Li, L.; Lin, R.B.; Zhou, W.; Zhang, Z.; Xiang, S.; Chen, B. Porous metal-organic frameworks for gas storage and separation: Status and challenges. Energy Chem. 2019, 1, 100006. [Google Scholar] [CrossRef]
  18. Dolgopolova, E.A.; Rice, A.M.; Martina, C.R.; Shustova, N.B. Photochemistry and photophysics of MOFs: Steps towards MOF-based sensing enhancements. Chem. Soc. Rev. 2018, 47, 4710–4728. [Google Scholar] [CrossRef]
  19. Bazaga-García, M.; Vílchez-Cózar, Á.; Maranescu, B.; Olivera-Pastor, P.; Marganovici, M.; Ilia, G.; Cabeza Díaz, A.; Visa, A.; Colodrero, R.M.P. Synthesis and electrochemical properties of metal(II)-carboxyethylphenylphosphinates. Dalton. Trans. 2021, 50, 6539–6548. [Google Scholar] [CrossRef]
  20. Song, J.; Hua, M.; Huang, X.; Visa, A.; Wu, T.; Hou, M.; Zhang, Z.; Han, B. Highly Efficient Meerwein-Ponndorf-Verley Reductions Over a Robust Zirconium-Organoboronic Acid Hybrid. Green Chem. 2021, 23, 1259–1265. [Google Scholar] [CrossRef]
  21. Iliescu, S.; Ilia, G.; Pascariu, A.; Popa, A.; Plesu, N. Novel synthesis of phosphorus containing polymers under inverse phase transfer catalysis. Polymer 2006, 47, 6509–6512. [Google Scholar] [CrossRef]
  22. Nabipour, H.; Wang, X.; Song, L.; Hu, Y. Metal-organic frameworks for flame retardant polymers application: A critical review. Compos. Part A Appl. Sci. 2020, 139, 106113. [Google Scholar] [CrossRef]
  23. Drehe, M.; Simulescu, V.; Ilia, G. Progress in the development of flame retardants. Rev. Chem. Eng. 2006, 24, 263–302. [Google Scholar]
  24. Macarie, L.; Simulescu, V.; Ilia, G. Ultrasonic irradiation used in synthesis of aminophosphonates. Monatsh Chem. 2019, 150, 163–171. [Google Scholar] [CrossRef]
  25. Maranescu, B.; Lupa, L.; Tara Lunga Mihali, M.; Plesu, N.; Maranescu, V.; Visa, A. The corrosion inhibitor behavior of iron in saline solution by the action of magnesium carboxyphosphonate. Pure Appl. Chem. 2018, 90, 1713–1722. [Google Scholar] [CrossRef]
  26. Maranescu, B.; Plesu, N.; Visa, A. Phosphonic acid vs phosphonate metal organic framework influence on mild steel corrosion protection. Appl. Surface Sci. 2019, 497, 143734. [Google Scholar] [CrossRef]
  27. Li, H.; Wang, K.; Sun, Y.; Lollar, C.T.; Li, J.; Zhou, H.-C. Recent advances in gas storage and separation using metal–organic frameworks. Mater. Today 2018, 21, 108–121. [Google Scholar] [CrossRef]
  28. Lupa, L.; Maranescu, B.; Visa, A. Equilibrium and kinetic studies of chromium ions adsorption on Co(II) based phosphonate metal organic frameworks. Sep. Sci. Technol. 2018, 53, 1017–1026. [Google Scholar] [CrossRef]
  29. Nistor, M.A.; Muntean, S.G.; Maranescu, B.; Visa, A. Phosphonate metal organic frameworks used as dyes removal materials from wastewaters. Appl. Organomet. Chem. 2020, 34, e5939. [Google Scholar] [CrossRef]
  30. Maranescu, B.; Lupa, L.; Visa, A. Synthesis, characterization and rare earth elements adsorption properties of phosphonate metal organic frameworks. Appl. Surface Sci. 2019, 481, 83–91. [Google Scholar] [CrossRef]
  31. Baby, R.; Hussein, M.Z.; Abdullah, A.H.; Zainal, Z. Nanomaterials for the treatment of heavy metal contaminated water. Polymers 2022, 14, 583. [Google Scholar] [CrossRef] [PubMed]
  32. Visa, A.; Maranescu, B.; Lupa, L.; Crisan, L.; Borota, A. New efficient adsorbent materials in the removal process of Cd(II) from aqueous solutions. Nanomaterials 2020, 10, 899. [Google Scholar] [CrossRef] [PubMed]
  33. Maranescu, B.; Lupa, L.; Visa, A. Synthesis, characterizations and Pb(II) sorption properties of cobalt phosphonate materials. Pure Appl. Chem. 2016, 88, 979–992. [Google Scholar] [CrossRef] [Green Version]
  34. Maranescu, B.; Popa, A.; Lupa, L.; Maranescu, V.; Visa, A. Use of chitosan complex with aminophosphonic groups and cobalt for the removal of Sr2+ ions. Sep. Sci. Technol. 2018, 53, 1058–1064. [Google Scholar]
  35. Al Sharabati, M.; Sabouni, R.; Ghaleb, A.; Husseini, G.A. Biomedical applications of metal–organic frameworks for disease diagnosis and drug delivery: A review. Nanomaterials 2022, 12, 277. [Google Scholar] [CrossRef]
  36. Popa, A.; Ilia, G.; Iliescu, S.; Dehelean, G.; Pascariu, A.; Bora, A.; Pacureanu, L. Mixed quaternary ammonium and phosphonium salts bound to macromolecular supports for removal bacteria from water. Mol. Cryst. Liq. Cryst. 2004, 418, 195–203. [Google Scholar] [CrossRef]
  37. Pettinari, C.; Pettinari, R.; Di Nicola, C.; Tombesi, A.; Scuri, S.; Marchetti, F. Antimicrobial MOFs. Coord. Chem. Rev. 2021, 446, 214121. [Google Scholar] [CrossRef]
  38. Popa, A.; Ilia, G.; Pascariu, A.; Iliescu, S.; Plesu, N. Grafted styrene-divinylbenzene copolymers containing benzaldehyde and their Wittig reactions with various phosphonium salts. Chin. J. Polym. Sci. 2005, 23, 651–656. [Google Scholar] [CrossRef]
  39. Quijia, C.R.; Lima, R.; Silva, C.; Alves, C.R.; Frem, R.; Chorilli, M. Application of MIL-100(Fe) in drug delivery and biomedicine. J. Drug. Deliv. Sci. Technol. 2021, 61, 102217. [Google Scholar] [CrossRef]
  40. Vassaki, M.; Kotoula, C.; Turhanen, P.; Choquesillo-Lazarte, D.; Demadis, K.D. Calcium and strontium coordination polymers as controlled delivery systems of the anti-osteoporosis drug risedronate and the augmenting effect of solubilizers. Appl. Sci. 2021, 11, 11383. [Google Scholar] [CrossRef]
  41. Vaughn, J.; Wu, H.; Efremovska, B.; Olson, D.H.; Mattai, J.; Ortiz, C.; Puchalski, A.; Li, J.; Pan, L. Encapsulated recyclable porous materials: An effective moisture-triggered fragrance release system. Chem. Comm. 2013, 49, 5724–5726. [Google Scholar] [CrossRef] [PubMed]
  42. Ma, X.; Lepoitevin, M.; Serre, C. Metal–organic frameworks towards bio-medical applications. Mater. Chem. Front. 2021, 5, 5573–5594. [Google Scholar] [CrossRef]
  43. Sun, Y.; Zheng, L.; Yang, Y.; Fu, F.; Li, X.; Yang, Z.; Yan, H.; Cui, C.; Tan, W.N. Metal–organic framework nanocarriers for drug delivery in biomedical applications. Nano-Micro Lett. 2020, 12, 103. [Google Scholar] [CrossRef] [PubMed]
  44. Mallakpour, S.; Nikkhoo, E.; Hussain, C.M. Application of MOF materials as drug delivery systems for cancer therapy and dermal treatment. Coord. Chem. Rev. 2022, 451, 214262. [Google Scholar] [CrossRef]
  45. Jeyaseelan, C.; Jain, P.; Soin, D.; Gupta, G. Metal organic frameworks: An effective application in drug delivery systems. Inorg. Nano-Met. Chem. 2021. [Google Scholar] [CrossRef]
  46. Martí-Rujas, J. Structural elucidation of microcrystalline MOFs from powder X-ray diffraction. Dalton Trans. 2020, 49, 13897. [Google Scholar] [CrossRef]
  47. Torresi, S.; Famulari, A.; Martí-Rujas, J. Kinetically Controlled Fast Crystallization of M12L8 Poly-[n]-catenanes Using the 2,4,6-Tris(4-pyridyl)benzene Ligand and ZnCl2 in an Aromatic Environment. J. Am. Chem. Soc. 2020, 142, 9537–9543. [Google Scholar] [CrossRef]
  48. Guo, F.; Wang, H.-C.; Famulari, A.; Lu, H.-D.; Martí-Rujas, J. Dynamic behaviour in nonporous hybrid metal–organic materials via mechanochemical and gas–solid reactions. CrystEngComm 2018, 20, 6721–6726. [Google Scholar] [CrossRef]
  49. Silva, P.; Vilela, S.M.F.; João, P.C.; Tome, J.P.C.; Almeida Paz, F.A. Multifunctional metal–organic frameworks: From academia to industrial applications. Chem. Soc. Rev. 2015, 44, 6774–6803. [Google Scholar] [CrossRef] [Green Version]
  50. Sun, C.-Y.; Qin, C.; Wang, X.-L.; Suet, Z.M. Metal–organic frameworks as potential drug delivery systems. Expert Opin. Drug Deliv. 2013, 10, 89–101. [Google Scholar] [CrossRef]
  51. Wang, L.; Zheng, M.; Xie, Z. Nanoscale metal–organic frameworks for drug delivery: A conventional platform with new promise. J. Mater. Chem. B 2018, 6, 707–717. [Google Scholar] [CrossRef] [PubMed]
  52. Petric, M.; Crisan, L.; Crisan, M.; Micle, A.; Maranescu, B.; Ilia, G. Synthesis and QSRR Study for a Series of Phosphoramidic Acid Derivatives. Heteroatom Chem. 2013, 24, 138–145. [Google Scholar] [CrossRef]
  53. Gheonea, R.; Mak, C.; Crasmareanu, E.C.; Simulescu, V.; Plesu, N.; Ilia, G. Surface modification of SnO2 with phosphonic acids. J. Chem. 2017, 2017, 2105938. [Google Scholar] [CrossRef] [Green Version]
  54. Mohammed, L.; Gomaa, H.G.; Ragab, D.; Zhu, J. Magnetic nanoparticles for environmental and biomedical applications: A review. Particuology 2017, 30, 1–14. [Google Scholar] [CrossRef]
  55. Bunzen, H. Chemical stability of metal-organic frameworks for applications in drug delivery. ChemNanoMat 2021, 7, 998–1007. [Google Scholar] [CrossRef]
  56. Bůžek, D.; Adamec, S.; Lang, K.; Demel, J. Metal–organic frameworks vs. buffers: Case study of UiO-66 stability. Inorg. Chem. Front. 2021, 8, 720–734. [Google Scholar] [CrossRef]
  57. Velásquez-Hernández, M.J.; Linares-Moreau, M.; Astria, E.; Carraro, F.; Alyami, M.Z.; Khashab, N.M.; Sumby, C.J.; Doonan, C.J.; Falcaro, P. Towards applications of bioentities@MOFs in biomedicine. Coord. Chem. Rev. 2021, 429, 213651. [Google Scholar] [CrossRef]
  58. Raliya, R.; Chadha, T.S.; Hadad, K.; Biswas, P. Perspective on nanoparticle technology for biomedical use. Curr. Pharm. Des. 2016, 22, 2481–2490. [Google Scholar] [CrossRef] [Green Version]
  59. Zhou, Q.; Zhang, L.; Wu, H. Nanomaterials for cancer therapies. Nanotechnol. Rev. 2017, 6, 473–496. [Google Scholar] [CrossRef]
  60. Lai, S.K.; Wang, Y.-Y.; Hida, K.; Cone, R.; Hanes, J. Nanoparticles reveal that human cervicovaginal mucus is riddled with pores larger than viruses. Proc. Natl. Acad. Sci. USA 2010, 107, 598–603. [Google Scholar] [CrossRef] [Green Version]
  61. Li, B.; Chrzanowski, M.; Zhang, Y.; Ma, S. Applications of metal-organic frameworks featuring multi-functional sites. Coord. Chem. Rev. 2016, 307, 106–129. [Google Scholar] [CrossRef] [Green Version]
  62. Hannah, A.; Wilson, K.; Homan, K.; Emelianov, S. Indocyanine green-loaded photoacoustic nanodroplets: Dual contrast nanoconstructs for enhanced photoacoustic and ultrasound imaging. ACS Nano 2014, 8, 250–259. [Google Scholar] [CrossRef] [PubMed]
  63. Dai, Q.; Ren, E.; Xu, D.; Zeng, Y.; Chen, C.; Liu, G. Indocyanine green-based nanodrugs: A portfolio strategy for precision medicine. Prog. Nat. Sci. 2020, 30, 577–588. [Google Scholar] [CrossRef]
  64. Jung, H.S.; Verwilst, P.; Sharma, A.; Shin, J.; Sessler, J.L.; Kim, J.S. Organic molecule-based photothermal agents: An expanding photothermal therapy universe. Chem. Soc. Rev. 2018, 47, 2280–2297. [Google Scholar] [CrossRef] [PubMed]
  65. Bhattacharjee, A.; Mihir, G.S.; Purkait, K. Fe3O4 promoted metal organic framework MIL-100(Fe) for the controlled release of doxorubicin hydrochloride. Micropor. Mesopor. Mat. 2018, 259, 203–210. [Google Scholar] [CrossRef]
  66. Yang, C.; Xu, J.; Yang, D.; Wang, X.; Liu, B.; He, N.; Wang, Z. ICG@ZIF-8: One-step encapsulation of Indocyanine Green in ZIF-8 and use as a therapeutic nanoplatform. Chin. Chem. Lett. 2018, 29, 1421–1424. [Google Scholar] [CrossRef]
  67. Bhattacharjee, A.; Purkaita, M.K.; Gumma, S. Loading and release of doxorubicin hydrochloride from iron(III) trimesate MOF and zinc oxide nanoparticle composites. Dalton Trans. 2020, 49, 8755–8763. [Google Scholar] [CrossRef]
  68. Yao, L.; Tang, Y.; Cao, W.; Yuanjing Cui, Y.; Qian, G. Highly efficient encapsulation of doxorubicin hydrochloride in metal–organic frameworks for synergistic chemotherapy and chemodynamic therapy. ACS Biomater. Sci. Eng. 2021, 7, 4999–5006. [Google Scholar] [CrossRef]
  69. Mojardin, L.; Botet, J.; Quintales, L.; Moreno, S.; Salas, M. New Insights into the RNA based mechanism of action of the anticancer drug 5′-fluorouracil in eukaryotic cells. PLoS ONE 2013, 8, e78172. [Google Scholar] [CrossRef] [Green Version]
  70. Stefaniak, K.R.; Epley, C.C.; Novak, J.J.; McAndrew, M.L.; Cornell, H.D.; Zhu, J.; McDaniel, D.K.; Davis, J.L.; Allen, I.C.; Morris, A.J.; et al. Photo-triggered release of 5-fluorouracil from a MOF drug delivery vehicle. Chem. Commun. 2018, 54, 7617–7620. [Google Scholar] [CrossRef]
  71. Sun, X.Y.; Zhang, H.J.; Zhao, X.Y.; Sun, Q.; Yuan-Yuan Wanga, Y.-Y.; Gao, E.Y. Dual functions of pH-sensitive cation Zr-MOF for 5-Fu: Large drug-loading capacity and high-sensitivity fluorescence detection. Dalton Trans. 2021, 50, 10524–10532. [Google Scholar] [CrossRef] [PubMed]
  72. Márquez, A.G.; Hidalgo, T.; Lana, H.; Cunha, D.; Blanco-Prieto, M.J.; Álvarez-Lorenzo, C.; Boissière, C.; Sánchez, C.; Serrea, C.; Horcajada, P. Biocompatible polymer–metal–organic framework composite patches for cutaneous administration of cosmetic molecules. J. Mater. Chem. B 2016, 4, 7031–7040. [Google Scholar] [CrossRef] [PubMed]
  73. Liédana, N.; Galve, A.; Rubio, C.; Téllez, C.; Coronas, J. CAF@ZIF-8: One-step encapsulation of caffeine in MOF. ACS Appl. Mater. Interfaces 2012, 4, 5016–5021. [Google Scholar] [CrossRef] [PubMed]
  74. Gaudin, C.; Cunha, D.; Ivanoff, E.; Horcajada, P.; Chevé, G.; Yasri, A.; Loget, O.; Serre, C.; Maurin, G. A Quantitative Structure Activity Relationship Approach to Probe the Influence of the Functionalization on the Drug Encapsulation of Porous Metal-organic Frameworks. Microporous Mesoporous Mater. 2012, 157, 124–130. [Google Scholar] [CrossRef]
  75. Xu, W.; Lou, Y.; Chen, W.; Kang, Y. Folic acid decorated metal-organic frameworks loaded with doxorubicin for tumor-targeted chemotherapy of osteosarcoma. Biomed Technol. 2020, 65, 229–236. [Google Scholar] [CrossRef]
  76. Prabaharan, M.; Grailer, J.J.; Pilla, S.; Steeber, D.A.; Gong, S.Q. Folate-conjugated Amphiphilic Hyperbranched Block Copolymers based on Boltorn®H40, Poly(l-lactide) and Poly(ethylene glycol) for Tumor-targeted Drug Delivery. Biomaterials 2009, 30, 3009–3019. [Google Scholar] [CrossRef]
  77. Mansoori, G.A.; Brandenburg, K.S.; Shakeri-Zadeh, A. A Comparative study of two folate-conjugated gold nanoparticles for cancer nanotechnology applications. Cancers 2010, 2, 1911–1928. [Google Scholar] [CrossRef] [Green Version]
  78. Orellana-Tavra, C.; Baxter, E.F.; Tian, T.; Bennett, T.D.; Slater, N.K.H.; Cheetham, A.K.; Fairen-Jimenez, D. Amorphous metal–organic frameworks for drug delivery. Chem. Commun. 2015, 51, 13878–13881. [Google Scholar] [CrossRef] [Green Version]
  79. Yalamandala, B.N.; Shen, W.T.; Min, S.-H.; Chiang, W.-H.; Chang, S.J.; Hu, S.-H. Advances in functional metal-organic frameworks based on-demand drug delivery systems for tumor therapeutics. Adv. NanoBiomed Res. 2021, 1, 2100014. [Google Scholar] [CrossRef]
  80. Tomeh, M.A.; Hadianamrei, R.; Zhao, X. A review of curcumin and its derivatives as anticancer agents. Int. J. Mol. Sci. 2019, 20, 1033. [Google Scholar] [CrossRef] [Green Version]
  81. Gao, S.; Zheng, P.; Li, Z.; Feng, X.; Yan, W.; Chen, S.; Guo, W.; Liu, D.; Yang, X.; Wang, S.; et al. Biomimetic O2- evolving metal-organic framework nanoplatform for highly efficient photodynamic therapy against hypoxic tumor. Biomaterials 2018, 178, 83–94. [Google Scholar] [CrossRef] [PubMed]
  82. Chu, C.C.; Ren, E.; Zhang, Y.M.; Yu, J.W.; Lin, H.R.; Pang, X.; Zhang, Y.; Liu, H.; Qin, Z.N.; Cheng, Y.; et al. Zinc(II)-dipicolylamine coordination nanotheranostics: Toward synergistic nanomedicine by combined photo/gene therapy. Angew. Chem. Int. Ed. 2019, 58, 269–272. [Google Scholar] [CrossRef] [PubMed]
  83. Fang, C.; Cen, D.; Wang, Y.F.; Wu, Y.J.; Cai, X.J.; Li, X.; Han, G.R. ZnS@ZIF-8 core-shell nanoparticles incorporated with ICG and TPZ to enable H2S-amplified synergistic therapy. Theranostics 2020, 10, 7671–7682. [Google Scholar] [CrossRef]
  84. You, Q.; Zhang, K.Y.; Liu, J.Y.; Liu, C.L.; Wang, H.Y.; Wang, M.T.; Ye, S.Y.; Gao, H.Q.; Lv, L.T.; Wang, C.; et al. Persistent regulation of tumor hypoxia microenvironment via a bioinspired Pt-based oxygen nanogenerator for multimodal imaging-guided synergistic phototherapy. Adv. Sci. 2020, 7, 1903341. [Google Scholar] [CrossRef] [PubMed]
  85. Fan, Z.; Liu, H.; Xue, Y.; Lin, J.; Fu, Y.; Xia, Z.; Pan, D.; Zhang, J.; Qiao, K.; Zhang, Z.; et al. Reversing cold tumors to hot: An immunoadjuvant-functionalized metal-organic framework for multimodal imaging-guided synergistic photoimmunotherapy. Bioact. Mater. 2021, 6, 312–325. [Google Scholar] [CrossRef] [PubMed]
  86. Wu, B.; Fu, J.; Zhou, Y.; Luo, S.; Zhao, Y.; Quan, G.; Pan, X.; Wu, C. Tailored core-shell dual metal-organic frameworks as a versatile nanomotor for effective synergistic antitumor therapy. Acta Pharm. Sin. B 2020, 10, 2198–2211. [Google Scholar] [CrossRef]
  87. Su, Y.; Ni, W.; Zhang, H.; Jiang, K.; Wen, F. One-pot synthesis of ICG&Cur@ZIF-8 nanocomposites with pH-controlled drug delivery and good photothermal performance. Z. Anorg. Allg. Chem. 2021. [Google Scholar] [CrossRef]
  88. Gautam, S.; Singhal, J.; Lee, H.K.; Chae, K.H. Drug delivery of paracetamol by metal-organic frameworks (HKUST-1): Improvised synthesis and investigations. Mater. Today Chem. 2022, 23, 100647. [Google Scholar] [CrossRef]
  89. Yang, Y.; Liu, J.; Liang, C.; Feng, L.; Fu, T.; Dong, Z.; Chao, Y.; Li, Y.; Lu, G.; Chen, M.; et al. Nanoscale metal-organic particles with rapid clearance for magnetic resonance imaging-guided photothermal therapy. ACS Nano 2016, 23, 2774–2781. [Google Scholar] [CrossRef]
  90. Lu, K.; He, C.; Lin, W. A Chlorin-Based Nanoscale metal–organic framework for photodynamic therapy of colon cancers. J. Am. Chem. Soc. 2015, 137, 7600–7603. [Google Scholar] [CrossRef] [Green Version]
  91. Rabiee, N.; Bagherzadeh, M.; Jouyandeh, M.; Zarrintaj, P.; Saeb, M.R.; Mozafari, M.; Shokouhimehr, M.; Varma, R.S. Natural Polymers Decorated MOF-MXene nanocarriers for Co-delivery of doxorubicin/pCRISPR. ACS Appl. Bio Mater. 2021, 4, 5106–5121. [Google Scholar] [CrossRef] [PubMed]
  92. Sun, X.; He, G.; Xiong, C.; Wang, C.; Lian, X.; Hu, L.; Li, Z.; Dalgarno, S.J.; Yang, Y.-W.; Tian, J. One-pot fabrication of hollow porphyrinic MOF nanoparticles with ultrahigh drug loading toward controlled delivery and synergistic cancer therapy. ACS Appl. Mater. Interfaces 2021, 13, 3679–3693. [Google Scholar] [CrossRef] [PubMed]
  93. Alves, R.C.; Schulte, Z.M.; Luiz, M.T.; da Silva, P.B.; Frem, R.C.G.; Rosi, N.L.; Chorilli, M. Breast Cancer targeting of a drug delivery system through postsynthetic modification of curcumin@N3-bio-MOF-100 via Click Chemistry. Inorg. Chem. 2021, 60, 11739–11744. [Google Scholar] [CrossRef] [PubMed]
  94. Yang, B.; Shen, M.; Liu, J.; Ren, F. Post-Synthetic Modification nanoscale metal-organic frameworks for targeted drug delivery in cancer cells. Pharm. Res. 2017, 34, 2440–2450. [Google Scholar] [CrossRef] [PubMed]
  95. Horcajada, P.; Chalati, T.; Serre, C.; Gillet, B.; Sebrie, C.; Baati, T.; Eubank, J.F.; Heurtaux, D.; Clayette, P.; Kreuz, C.; et al. Porous metal-organic-framework nanoscale carriers as a potential platform for drug delivery and imaging. Nat. Mater. 2010, 9, 172–178. [Google Scholar] [CrossRef]
  96. Gao, X.; Zhai, M.; Guan, W.; Liu, J.; Liu, Z.; Damirin, A. Controllable Synthesis of a Smart Multifunctional Nanoscale Metal–Organic Framework for Magnetic Resonance/Optical Imaging and Targeted Drug Delivery. ACS Appl. Mater. Interfaces 2017, 9, 3455–3462. [Google Scholar] [CrossRef] [PubMed]
  97. Hu, Q.; Yu, J.; Liu, M.; Liu, A.; Dou, Z.; Yang, Y. A low cytotoxic cationic metal-organic framework carrier for controllable drug release. J. Med. Chem. 2014, 57, 5679–5685. [Google Scholar] [CrossRef]
  98. Lawson, S.; Newport, K.; Schueddig, K.; Rownaghi, A.A.; Rezaei, F. Optimizing ibuprofen concentration for rapid pharmacokinetics on biocompatible zinc-based MOF-74 and UTSA-74. Mater. Sci. Eng. C 2020, 117, 111336. [Google Scholar] [CrossRef]
  99. Taylor-Pashow, K.M.L.; Della Rocca, J.; Xie, Z.; Tran, S.; Lin, L. Postsynthetic Modifications of iron-carboxylate nanoscale metal-organic frameworks for imaging and drug delivery. J. Am. Chem. Soc. 2009, 131, 14261–142632. [Google Scholar] [CrossRef] [Green Version]
  100. Han, L.; Qi, H.; Zhang, D.; Ye, G.; Zhou, W.; Hou, C.; Xud, W.; Sun, Y. A facile and green synthesis of MIL-100(Fe) with high-yield and its catalytic performance. New J. Chem. 2017, 41, 13504–13509. [Google Scholar] [CrossRef]
  101. Zhang, Y.; Wang, L.; Liu, L.; Lin, L.; Liu, F.; Xie, Z.; Tian, H.; Chen, X. Engineering metal–organic frameworks for photoacoustic imaging-guided chemo-/photothermal combinational tumor therapy. ACS Appl. Mater. Interfaces 2018, 10, 41035–41045. [Google Scholar] [CrossRef] [PubMed]
  102. Xue, T.; Xu, C.; Wang, Y.; Wang, Y.; Tian, H.; Zhang, Y. Doxorubicin-loaded nanoscale metal–organic framework for tumor-targeting combined chemotherapy and chemodynamic therapy. Biomater. Sci. 2019, 7, 4615–4623. [Google Scholar] [CrossRef] [PubMed]
  103. Devautour-Vinot, S.; Martineau, C.; Diaby, S.; Ben-Yahia, M.; Miller, S.; Serre, C.; Horcajada, P.; Cunha, D.; Taulelle, F.; Maurin, G. Caffeine confinement into a series of functionalized porous zirconium MOFs: A joint experimental/modeling exploration. J. Phys. Chem. C 2013, 117, 11694–11704. [Google Scholar] [CrossRef]
  104. Chen, X.; Zhuang, Y.; Rampal, N.; Hewitt, R.; Divitini, G.; O’Keefe, C.A.; Liu, X.; Whitaker, D.J.; Wills, J.W.; Jugdaohsingh, R.; et al. Formulation of metal–organic framework-based drug carriers by controlled coordination of methoxy PEG phosphate: Boosting colloidal stability and redispersibility. J. Am. Chem. Soc. 2021, 143, 13557–13572. [Google Scholar] [CrossRef]
  105. Cunha, D.; Mouna, B.Y.; Hall, S.; Miller, R.S.; Chevreau, H.; Elkaïm, E.; Maurin, G.; Horcajada, P.; Serre, C. Rationale of drug encapsulation and release from biocompatible porous metal–organic frameworks. Chem. Mater. 2013, 25, 2767–2776. [Google Scholar] [CrossRef]
  106. Ye, X.; Liu, D. Metal–Organic Framework UiO-68 and Its Derivatives with sufficiently good properties and performance show promising prospects in potential industrial applications. Cryst. Growth Des. 2021, 21, 4780–4804. [Google Scholar] [CrossRef]
  107. Li, S.; Wang, K.; Shi, Y.; Cui, Y.; Chen, B.; He, B.; Dai, W.; Zhang, H.; Wang, X.; Zhong, C.; et al. Novel biological functions of ZIF-NP as a delivery vehicle: High pulmonary accumulation, favorable biocompatibility, and improved therapeutic outcome. Adv. Funct. Mater. 2016, 26, 2715. [Google Scholar] [CrossRef]
  108. Tan, L.-L.; Song, N.; Zhang, S.X.-A.; Li, H.; Wang, B.; Yang, Y.-W. Ca2+, pH and thermo triple-responsive mechanized Zr-based MOFs for on-command drug release in bone diseases. J. Mater. Chem. B 2016, 4, 135. [Google Scholar] [CrossRef]
  109. He, S.; Wu, L.; Li, X.; Sun, H.; Xiong, T.; Liu, J.; Chengxi Huang, C.; Xu, H.; Sun, H.; Chen, W.; et al. Metal-organic frameworks for advanced drug delivery. Acta Pharm. Sin. B 2021, 11, 2362–2395. [Google Scholar] [CrossRef]
  110. Cai, W.; Wang, J.; Chu, C.; Chen, W.; Wu, C.; Liu, G. Metal–organic framework-based stimuli-responsive systems for drug delivery. Adv. Sci. 2018, 6, 1801526. [Google Scholar] [CrossRef] [Green Version]
  111. Luo, Z.; Jiang, L.; Yang, S.; Li, Z.; Soh, W.M.W.; Zheng, L.; Loh, X.J.; Wu, Y.L. Light-induced redox-responsive smart drug delivery system by using selenium-containing polymer@MOF Shell/Core nanocomposite. Adv. Healthc. Mater. 2019, 8, 1900406. [Google Scholar] [CrossRef] [PubMed]
  112. Diring, S.; Carne-Sanchez, A.; Zhang, J.; Ikemura, S.; Kim, C.; Inaba, H.; Kitagawa, S.; Furukawa, S. Light responsive metal–organic frameworks as controllable CO-releasing cell culture substrates. Chem. Sci. 2017, 8, 2381–2386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Jin, Z.; Zhao, P.; Zhang, J.; Yang, T.; Zhou, G.; Zhang, D.; Wang, T.; He, Q. Intelligent metal carbonyl metal–organic framework nanocomplex for fluorescent traceable H2O2-triggered CO delivery. Chem. Eur. J. 2018, 24, 11667–11674. [Google Scholar] [CrossRef] [PubMed]
  114. Pobłocki, K.; Drzeżdżon, J.; Kostrzewa, T.; Jacewicz, D. Coordination Complexes as a New Generation Photosensitizer for Photodynamic Anticancer Therapy. Int. J. Mol. Sci. 2021, 22, 8052. [Google Scholar] [CrossRef]
  115. Cabrera-García, A.; Checa-Chavarria, E.; Rivero-Buceta, E.; Moreno, V.; Fernández, E.; Botella, P. Amino modified metal-organic frameworks as pH-responsive nanoplatforms for safe delivery of camptothecin. J. Colloid Interface Sci. 2019, 541, 163–174. [Google Scholar] [CrossRef]
  116. Tan, L.L.; Li, H.; Qiu, Y.C.; Chen, D.X.; Wang, X.; Pan, R.Y.; Wang, Y.; Zhang, S.X.; Wang, B.; Yang, Y.W. Stimuli-responsive metal–organic frameworks gated by pillar[5]arene supramolecular switches. Chem. Sci. 2015, 6, 1640–1644. [Google Scholar] [CrossRef] [Green Version]
  117. Karimzadeh, S.; Javanbakht, S.; Baradaran, B.; Shahbazi, M.A.; Hashemzaei, M.; Mokhtarzadeh, A.; Santos, H.A. Synthesis and therapeutic potential of stimuli-responsive metal-organic frameworks. Chem. Eng. J. 2021, 408, 127233. [Google Scholar] [CrossRef]
  118. Zhu, Y.D.; Chen, S.P.; Zhao, H.; Yang, Y.; Chen, X.Q.; Sun, J.; Fan, H.S.; Zhang, X.D. 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, 8, 34209–34217. [Google Scholar] [CrossRef]
  119. Zou, Z.; Li, S.Q.; He, D.G.; He, X.X.; Wang, K.M.; Li, L.L.; Yang, X.; Li, H.F. A versatile stimulus-responsive metal–organic framework for size/morphology tunable hollow mesoporous silica and pH-triggered drug delivery. J. Mater. Chem. B 2017, 5, 2126–2132. [Google Scholar] [CrossRef]
  120. Tian, Z.F.; Yao, X.X.; Zhu, Y.F. Simple synthesis of multifunctional zeolitic imidazolate frameworks-8/graphene oxide nanocrystals with controlled drug release and photothermal effect. Microporous Mesoporous Mater. 2017, 237, 160–167. [Google Scholar] [CrossRef]
  121. Tang, L.; Shi, J.F.; Wang, X.L.; Zhang, S.H.; Wu, H.; Sun, H.F.; Jiang, Z.Y. Coordination polymer nanocapsules prepared using metal-organic framework templates for pH-responsive drug delivery. Nanotechnology 2017, 28, 275601. [Google Scholar] [CrossRef] [PubMed]
  122. Zhu, X.; Gu, J.; Wang, Y.; Li, B.; Li, Y.; Zhao, W.; Shi, J. Inherent anchorages in UiO-66 nanoparticles for efficient capture of alendronate and its mediated release. Chem. Commun. 2014, 50, 8779–8782. [Google Scholar] [CrossRef] [PubMed]
  123. Yang, Y.; Hu, Q.; Zhang, Q.; Jiang, K.; Lin, W.; Yang, Y.; Cui, Y.; Qian, G. A large capacity cationic metal–organic framework nanocarrier for physiological pH responsive drug delivery. Mol. Pharm. 2016, 13, 2782–2786. [Google Scholar] [CrossRef]
  124. Kundu, T.; Mitra, S.; Patra, P.; Goswami, A.; Diaz, D.; Banerjee, R. Mechanical Downsizing of a Gadolinium(III)-based Metal–Organic Framework for Anticancer Drug Delivery. Chem.-Eur. J. 2014, 20, 10514–10518. [Google Scholar] [CrossRef] [PubMed]
  125. Liu, J.J.; Wang, H.R.; Yi, X.; Chao, Y.; Geng, Y.H.; Xu, L.G.; Yang, K.; Liu, Z. pH-sensitive dissociable nanoscale coordination polymers with drug loading for synergistically enhanced chemoradiotherapy. Adv. Funct. Mater. 2017, 27, 1703832. [Google Scholar] [CrossRef]
  126. Duan, F.; Feng, X.C.; Yang, X.J.; Sun, W.T.; Jin, Y.; Liu, H.F.; Ge, K.; Li, Z.H.; Zhang, J.C. A simple and powerful co-delivery system based on pH-responsive metal-organic frameworks for enhanced cancer immunotherapy. Biomaterials 2017, 122, 23–33. [Google Scholar] [CrossRef] [PubMed]
  127. Hamisu, A.M.; Ariffin, A.; Wibowo, A.C. Cation Exchange in Metal-Organic Frameworks (MOFs): The Hard-Soft Acid-Base (HSAB) Principle Appraisal. Inorganica Chim. Acta 2020, 511, 119801. [Google Scholar] [CrossRef]
  128. Vardali, S.C.; Manousi, N.; Barczak, M.; Giannakoudakis, D.A. Novel Approaches Utilizing Metal-Organic Framework Composites for the Extraction of Organic Compounds and Metal Traces from Fish and Seafood. Molecules 2020, 25, 513. [Google Scholar] [CrossRef]
  129. Wang, C.; Liu, X.; Demir, N.K.; Chen, J.P.; Li, K. Applications of water stable metal–organic frameworks. Chem. Soc. Rev. 2016, 45, 5107–5134. [Google Scholar] [CrossRef]
  130. Park, K.S.; Ni, Z.; Côté, A.P.; Choi, J.Y.; Huang, R.; Uribe-Romo, F.J.; Chae, H.K.; O’Keeffe, M.; Yaghi, O.M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. USA 2006, 103, 10186–10191. [Google Scholar] [CrossRef] [Green Version]
  131. Fei, H.; Cahill, J.F.; Prather, K.A.; Cohen, S.M. Tandem postsynthetic metal ion and ligand exchange in zeolitic imidazolate frameworks. Inorg. Chem. 2016, 52, 4011–4016. [Google Scholar] [CrossRef] [PubMed]
  132. Huang, A.; Dou, W.; Caro, J.R. Steam-Stable Zeolitic Imidazolate Framework ZIF-90 Membrane with Hydrogen Selectivity through Covalent Functionalization. J. Am. Chem. Soc. 2010, 132, 15562–15564. [Google Scholar] [CrossRef] [PubMed]
Figure 1. MOFs synthesis—alternative reaction conditions. Features shown in blue are advantages and black are disadvantages for all described methods.
Figure 1. MOFs synthesis—alternative reaction conditions. Features shown in blue are advantages and black are disadvantages for all described methods.
Ijms 23 04458 g001
Figure 2. Number of publications from Web of Science: (A) “Metal Organic Frameworks” and (B) “Metal Organic Frameworks and Drug Delivery Systems”, from 1996 and 2009, through 21 March 2022.
Figure 2. Number of publications from Web of Science: (A) “Metal Organic Frameworks” and (B) “Metal Organic Frameworks and Drug Delivery Systems”, from 1996 and 2009, through 21 March 2022.
Ijms 23 04458 g002
Figure 3. Some examples of drugs delivered by MOFs as DDS: (a)—indocianine green, (b)—doxorubicin hydrochloride, (c)—5-fluorouracil, (d)—caffeine, (e)—cidofovir, (f)—acid folic, (g)—calcein, (h)—curcumin.
Figure 3. Some examples of drugs delivered by MOFs as DDS: (a)—indocianine green, (b)—doxorubicin hydrochloride, (c)—5-fluorouracil, (d)—caffeine, (e)—cidofovir, (f)—acid folic, (g)—calcein, (h)—curcumin.
Ijms 23 04458 g003
Figure 4. Encapsulation, direct assembly, and post-synthesis of cargo-loading strategies for MOFs. Reproduced with permission from [51]; Copyright (Wang, 2018), RSC J. Mater. Chem. B.
Figure 4. Encapsulation, direct assembly, and post-synthesis of cargo-loading strategies for MOFs. Reproduced with permission from [51]; Copyright (Wang, 2018), RSC J. Mater. Chem. B.
Ijms 23 04458 g004
Figure 5. Synthesis of O2@UiO-66@ICG@RBC (A). Schematic mechanism of NIR-triggered O2 enhanced and discharging PDT (B). Reproduced with permission from [81]; Copyright (Gao, 2018), Elsevier Biomaterials.
Figure 5. Synthesis of O2@UiO-66@ICG@RBC (A). Schematic mechanism of NIR-triggered O2 enhanced and discharging PDT (B). Reproduced with permission from [81]; Copyright (Gao, 2018), Elsevier Biomaterials.
Ijms 23 04458 g005
Figure 6. Schematic illustration of metal–organic frameworks (MOFs)-based stimuli-responsive system for drug delivery. Reproduced from reference [110]; Copyright (Cai, 2018) Wiley Online Library.
Figure 6. Schematic illustration of metal–organic frameworks (MOFs)-based stimuli-responsive system for drug delivery. Reproduced from reference [110]; Copyright (Cai, 2018) Wiley Online Library.
Ijms 23 04458 g006
Table 1. A literature assessment of MOF-based nanomaterials for drug delivery systems.
Table 1. A literature assessment of MOF-based nanomaterials for drug delivery systems.
MOFsComponentsPore Size or Volume/Particle SizeTherapeutic AgentDrug Loading PercentageOutcomeRef.
OrganicInorganic
MIL-53(Fe)Terephtalic acidFeCl3·6H2O8.6 Å/ 350 nmIbuprofen
Caffeine
22
23.1
antitumoral and retroviral drugs against cancer and AIDS[95]
MIL-53-NH2 (Fe)2-Amino-terephtalic acidFeCl3·6H2O120 nm5-Fu28magnetic
resonance,
optical imaging,
targeted drug
delivery
[96]
MOF-74-Fe1.4-Dihydroxy terephtalic acidFeCl2·4H2O-/200–800 nm × some μmIbuprofen15.9low cytotoxicity,
efficient drug loading capacity,
controllable drug release
[97]
MOF-74-Zn2.5-Dihydroxyterephthalic acidZn(NO3)2 6H2O12.7 A/20 nmIbuprofen50fast kinetics
(k = 0.27 h−1/2),
high drug conc. in first 10 h.
[98]
MIL-101 (Cr)1,4-benzenedicarboxylates,trimeric chromium(III) octahedral clusters34 ÅIbuprofen -can adsorb 138 wt% ibuprofen,
release the total amount slowly in 6 days
[99]
MIL-100 (Fe)Trimesic acidFe (NO3)3·9H2O or FeCl3·6H2O29 A/200 nmIbuprofen
urcumin
33antitumoural and retroviral drugs,
photoacoustic imaging-guided chemo-photothermal combinational tumor therapy
[95,100,101]
MIL-89 (Fe)Muconic acidFeCl3·6H2O11A/50–100 nmCidofovir14antitumoral and retroviral drugs[95]
MIL-101-NH2 (Fe)Aminoterephtalic acidFeCl3·6H2O34 A/120 nmCidofovir41.9 [95]
MIL-100 (Fe)Trimesic acid Fe (NO3)3·9H2O or FeCl3·6H2O1.2 cm3/g
/102.8 nm
DOX
Caffeine
28
24.2
released DOX in a pH-dependent manner,
breast cancer treatment
[102]
UiO-66(Zr)1,4-Benzenedicarboxylic acidZrCl45–7 ACaffeine
DOX
22.4
15.1
caffeine molecules are preferentially located in the smaller cages,
DOX-containing PEGylated nanoMOFs exhibit notable redispersibility
[103,104]
MIL-127 Fe3,3′,5,5′-Azobenzene
tetracarboxylate
trimers of iron(III) octahedra4 ACaffeine15.9carriers for topical administration of caffeine[105]
MIL-100 (Fe)Trimesic acid{Fe3O} trimer200 nmDOX9.1full release of drug in 5 days[95]
Mi-UiO-68maleimide-attached H2L ligandZrCl425.6 ADOX
FA (acid folic)
4.84multifunctional cancer treatment system[106]
MIL-88@ZIF-8BDC-NH2, 2-Me-IMFeCl3·6H2O
Zn(NO3)2·6H2O
-/1.3 nmICG
DOX
3.58%
21.69%
MIL-88-ICG@ZIF-8-DOX
core-shell dual MOF for synergistic cancer, photothermal and photodynamic therapy
[86]
ZIF-8Me-IMZn(NO3)2·6H2O200 nm5-Fu21.2biological purposes: bio–nano interaction,
pulmonary accumulation,
antitumor therapy
[107]
UiO-66-NH2NH2-BDC{Zr6O8} cluster100 nm5-Fu3.1multistimuli responsive in bone diseases:
increasing Ca2+ concentration, decreasing pH, thermal therapy
[108]
MIL-88A (Fe)Fumaric acidFeCl3·6H2O6 A/150 nmCidofovir2.6 [109]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Maranescu, B.; Visa, A. Applications of Metal-Organic Frameworks as Drug Delivery Systems. Int. J. Mol. Sci. 2022, 23, 4458. https://doi.org/10.3390/ijms23084458

AMA Style

Maranescu B, Visa A. Applications of Metal-Organic Frameworks as Drug Delivery Systems. International Journal of Molecular Sciences. 2022; 23(8):4458. https://doi.org/10.3390/ijms23084458

Chicago/Turabian Style

Maranescu, Bianca, and Aurelia Visa. 2022. "Applications of Metal-Organic Frameworks as Drug Delivery Systems" International Journal of Molecular Sciences 23, no. 8: 4458. https://doi.org/10.3390/ijms23084458

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

Article Metrics

Back to TopTop