Next Article in Journal
Engineered Hybrid Vesicles and Cellular Internalization in Mammary Cancer Cells
Next Article in Special Issue
Carbosilane Dendritic Amphiphiles from Cholesterol or Vitamin E for Micelle Formation
Previous Article in Journal
Polymeric Nanoparticles Enable mRNA Transfection and Its Translation in Intervertebral Disc and Human Joint Cells, Except for M1 Macrophages
Previous Article in Special Issue
Dendritic Glycerol-Cholesterol Amphiphiles as Drug Delivery Systems: A Comparison between Monomeric and Polymeric Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

In Vivo Applications of Dendrimers: A Step toward the Future of Nanoparticle-Mediated Therapeutics

by
Krzysztof Sztandera
1,2,
José Luis Rodríguez-García
3 and
Valentín Ceña
1,2,*
1
Unidad Asociada Neurodeath, Instituto de Nanociencia Molecular, Universidad de Castilla-La Mancha, 02006 Albacete, Spain
2
Centro de Investigación Biomédica en Red en Enfermedades Neurodegenerativas, Instituto de Salud Carlos III, 28029 Madrid, Spain
3
Department of Internal Medicine, General University Hospital of Albacete, 02006 Albacete, Spain
*
Author to whom correspondence should be addressed.
Pharmaceutics 2024, 16(4), 439; https://doi.org/10.3390/pharmaceutics16040439
Submission received: 22 February 2024 / Revised: 17 March 2024 / Accepted: 20 March 2024 / Published: 22 March 2024

Abstract

:
Over the last few years, the development of nanotechnology has allowed for the synthesis of many different nanostructures with controlled sizes, shapes, and chemical properties, with dendrimers being the best-characterized of them. In this review, we present a succinct view of the structure and the synthetic procedures used for dendrimer synthesis, as well as the cellular uptake mechanisms used by these nanoparticles to gain access to the cell. In addition, the manuscript reviews the reported in vivo applications of dendrimers as drug carriers for drugs used in the treatment of cancer, neurodegenerative diseases, infections, and ocular diseases. The dendrimer-based formulations that have reached different phases of clinical trials, including safety and pharmacokinetic studies, or as delivery agents for therapeutic compounds are also presented. The continuous development of nanotechnology which makes it possible to produce increasingly sophisticated and complex dendrimers indicates that this fascinating family of nanoparticles has a wide potential in the pharmaceutical industry, especially for applications in drug delivery systems, and that the number of dendrimer-based compounds entering clinical trials will markedly increase during the coming years.

1. Introduction

Over the last few years, the development of nanotechnology has significantly increased. Nanotechnology allows researchers to synthesize many different nanostructures with controlled sizes, shapes, and properties. Such nanostructures include, among others, metal nanoparticles [1], polymeric micelles [2], polymersomes [3], dendrimers [4,5], and liposomes [6], with dendrimers being the best-characterized of these to date [7]. The name “dendrimer” comes from the combination of the Greek words “dendros” meaning “tree or branching”, and “meros” meaning “part of”. The high number of chemically modifiable terminal groups in their surface, as well as the void spaces present in their interior makes dendrimers excellent delivery agents for different therapeutic agents, including genetic material and drugs. Initial attempts to synthesize these branched polymers were first made in the 1940s. However, it was not until 1985 that the team of Donald Tomalia first described and synthesized poly(amidoamine) or PAMAM dendrimers [8]. To date, more than 100 families of dendrimers differing in their initiator cores, branching units and terminal moieties have been synthesized and evaluated [9]. Some of them, such as PAMAM [10], poly-L-lysine (PLL), bis-methyl propionic acid [11], and poly(propyleneimine) (PPI) dendrimers [12], have been commercialized while others like phosphorus dendrimers [13], Janus dendrimers [14], peptide dendrimers [15], rotaxane dendrimers [16], and carbosilane dendrimers [17] among others [18,19,20,21,22] have not. Amphiphillic Janus dendrimers and glycodendrimers are water-soluble mimics of biological membranes, including cell membrane glycans that self-assemble into unilamellar or onion-like dendrimersomes with predictable dimensions [23]. This dendrimer family has shown efficacy for the in vivo targeted delivery of mRNA to different organs based on changes in their chemical structure [24].

2. Dendrimer Structure

These interesting nanoparticles are three-dimensional sphere-shaped and monodispersed polymers with symmetrical branches [25]. They consist of three components: the central core, branches, and terminal groups (Figure 1). Branches are attached to the central core and end in terminal moieties that play a role as a dendrimer scaffold. As the number of branches increases, the dendrimer generation grows, and after generation 4 (G4), these nanoparticles adopt a globular structure [10] resembling the natural vesicles found in cells.
The synthesis of dendrimers can be strictly controlled, which provides high-purity homogeneous, monodispersed products with a well-defined structure, size, density, and uniform molecular weight [8,26,27]. This distinguishes them from classical linear polymers, that usually generate polydisperse products with different molecular weights.
There are two main ways to synthesize dendrimers: a divergent and a convergent approach (Figure 2). In the first method, monomers are added to the central core to engender the desired dendrimer generation. In contrast, during convergent synthesis, large segments (whole branches) are synthesized and then joined together to generate the dendrimer [26]. The use of click chemistry in the synthesis of dendrimers is also becoming increasingly popular. This system allows for the quick modification of substances without interfering with their structure [28]. Indeed, several reports about click chemistry in dendrimer synthesis with the use of thiol–yne, azide–alkyne, and Diels–Alder reactions have already been published [29,30,31].
Dendrimer synthesis requires a repeated series of reactions that increase dendrimer size and generation. Of note, there is a threshold generation level whereafter the attachment of further elements is impossible because the terminal groups form a closed membrane, with this phenomenon being referred to as the starburst effect [31]. Therefore, the synthesis of G10 is considered the highest possible level [32]. Furthermore, appropriate monomer selection is crucial in the development of dendrimers because the size, solubility, biocompatibility, multivalency, and ability to interact with biological systems depend on these initial units [33,34,35]. Moreover, manipulation of the dendrimer surface can prolong their blood circulation time or change their biodistribution in vivo.
The terminal groups of dendrimers can also be functionalized with biologically active molecules to provide them with wider application possibilities [36,37], thereby allowing them to encapsulate compounds (Figure 3). These cavities are usually hydrophobic and can generate hydrogen bonds to the transported compounds [38,39,40], which protects drugs from biodegradation before reaching the desired site of action [38]. However, because this premise is based on hydrophobic interactions, only hydrophobic compounds are used in this way. In addition to the use of internal cavities for storage, a wide range of terminal groups can also be conjugated covalently on the surface of dendrimers through specific linkers, including those sensitive to changes in pH or temperature, enzymatic or redox reactions, light irradiation, or hypoxia [41]. Non-covalent interactions between the terminal groups of dendrimers and different compounds do not interfere with the structure of these compounds but could be unstable in vivo. Thus, covalent bonding is the most stable connection between dendrimers and drugs but requires chemical modification of the compound by attaching linkers which could, in turn, change their properties [42] (Figure 3). The perspectives for the use of dendrimer-based systems as nanomedicines [43], biological adhesives [44], imaging agents [45], and in gene therapy [46] are all very promising. Dendrimers can be also used as vectors for the delivery of genes [47], nucleic acids [48], and drugs [49]. Moreover, research is currently underway to use dendrimers as adjuvants in vaccines [50] based on evidence that nanoparticles protect drugs from biodegradation, prolong their half-life in blood, improve their solubility, and provide targeted delivery [51].

3. Cellular Dendrimer Uptake

The main use of dendrimers to date is in the transport and delivery of different compounds to cell interiors. Thus, intracellular uptake is one of their most studied properties. Two main mechanisms are responsible for the intracellular transport of dendrimers: clathrin-mediated endocytosis (CME) and caveolin-mediated endocytosis (CVME). However, under some specific conditions, macropinocytosis can also be engaged in this process (Figure 4). During CME, clathrin is recruited to specialized regions of the cell membrane where it coats endocytic vesicles measuring 70–150 nm [52]. After internalization by the cell, the vesicles lose the clathrin coat and form early endosomes which later evolve into late endosomes and fuse with lysosomes [53]. This internalization pathway is common for receptor–ligand complexes such as low-density lipoprotein (LDL) particles [54] or epithelial growth factor [55], virus entry [56], and the uptake of neurotransmitters [57]. In turn, CMVE starts by generating 60–80 nm invaginations of the cell membrane in which proteins participating in endocytosis (e.g., caveolin-1) bind to lipid rafts. These vesicles either fuse with other caveolin vesicles to form caveosomes or with early endosomes, which can be transported to the Golgi apparatus or the smooth endoplasmic reticulum [58]. This pathway is used for, among other substances, the internalization of cholesterol [59], TGF-β receptor signaling [60] or toxins [61]. Macropinocytosis starts by creating vacuoles, termed macropinosomes, of variable sizes. After these enter cells, their pH decreases and they fuse with late endosomes or lysosomes [58]. This pathway is mainly used for the uptake of proteins [62], viruses [63], and antigens [64].
The type and efficacy of the cellular uptake of dendrimers is strongly dependent on the cell type, generation route, and surface charge of the nanoparticles in question. For instance, G4 PAMAM dendrimers are better internalized than lower generations [58]. Moreover, epithelial cells have a negative charge on their surface because of the presence of phosphate groups. Combined with a positive charge on the surface of cationic dendrimers, this creates electrostatic interactions that increase the efficiency of endocytosis [65]. Thus, it is extremely important to choose a dendrimer with an appropriate surface charge for every cell type. For example, PAMAM dendrimers with -NH2 moieties in MCF-7 cells are internalized by CME and micropinocytosis. However, in the case of A549 cells, they are internalized by both the CME and CVME pathways [66]. Furthermore, the chemical modification of dendrimers is also relevant: for instance, naked G4.5 PAMAM dendrimers are usually internalized through the CME, whereas after PEGylation (the modification of biological molecules by covalent conjugation with polyethylene glycol [PEG]), the main endocytic pathway changes to CVME [67].
The ideal nanocarriers should protect cargo compounds from biodegradation and release them into the targeted area. Moreover, they must also be biocompatible and not cause side effects [67]. Because of their shape, structure, and nanometric size, dendrimers can also penetrate into tumoral environments and be retained in the interstitium of tumors, a property termed the enhanced permeability and retention effect [51]. In addition, the possibility of almost unlimited dendrimer modifications by, for example, conjugation with antibodies, carbohydrates, folic acid (FA), or PEG, or by attaching active components through encapsulation, conjugation, or noncovalent interactions, makes these nanoparticles a promising alternative to regular cancer treatments [51]. In fact, nanosystems comprising dendrimers and drugs that result in improved cellular uptake and increased retention time compared to free drug delivery systems have been widely described in the academic literature [68,69,70].

4. In Vivo Use of Dendrimers

The cytotoxic activity, hemolytic properties, cellular uptake, or ability to transport and release different cargo compounds of many dendrimers have already been studied in vitro [9,71]. However, studies in isolated groups of cells do not accurately model the difficulties nanoparticles must overcome to reach their targets in vivo. These include biological barriers such as the blood–brain barrier (BBB), short plasma circulation times, hepatic or renal clearance, and limited distribution to sites of drug action [72]. Hence, more complex in vivo models of disease are now shedding new light on the usability of dendrimers in living systems. Studies performed at this stage aim to evaluate acute toxicity, dose dependence, and long-term toxicity in animal models, with parameters including pharmacodynamics and pharmacokinetics also being important.

4.1. Dendrimers as Carriers for Active Compounds

There is intense interest in dendrimers as delivery vehicles, especially for anti-cancer, anti-neurodegenerative, and anti-inflammation drugs [73,74,75,76], with the most important role of dendrimers in these systems being the efficient transport of drugs to the site of action. The almost infinite possibilities for dendrimer functionalization mean that compounds such as antibodies [77,78], folic acid [79], amino acids [80,81], sugar groups [82,83], or other specific ligands can be attached to their surface so that these nanosystems can target specific cells and tissues (Figure 5) [84]. For instance, these dendrimers could be designed to target cells with specific receptors or other substances secreted specifically by a given tumor type. Indeed, such modifications might also enable these molecules to cross the BBB [85]. Moreover, the use of pH-sensitive linkers can be used to release the drugs they carry only in a specific environment [86]. Thus, all the above-mentioned strategies could be combined to help improve drug efficiency and decrease side effects.
To date, most in vivo studies have focused on PAMAM dendrimers, although a few other dendrimer types have also been evaluated. For example, Chen et al. used phosphorus dendrimers to prepare micelles in order to encapsulate doxorubicin (DOX). They showed that this nanosystem can decrease the size of breast cancer tumor xenografts in mice by upregulating Bax, PTEN, and p53 [87]. Another group also used DOX combined with peptide dendrimers to target pancreatic ductal adenocarcinoma xenografts in mice, resulting in increased accumulation and internalization of the drug into these tumors. Additionally, these authors found that the efficacy of DOX was greater when it was co-administered with the dendrimer [88]. Furthermore, they reported that in a zebrafish model, the release of DOX in response to γ-radiation was increased when delivered by PAMAM dendrimers modified with L-cysteine because the latter acts as a radiosensitizer that allows for the release of the drug and inhibits cancer growth [89].
The use of PAMAM dendrimers to deliver other anti-cancer agents has also been extensively studied. For example, Bhadra et al. exploited naked and PEGylated PAMAM dendrimers to transport 5-fluorouracil and verified that this nanosystem was stable and biocompatible in albino rats by measuring the serum levels of the drug. Thus, this nanosystem proved to be safe for animals; and moreover, PEGylation reduced drug leakage and hemolysis, thereby indicating that PAMAM dendrimers are suitable for the prolonged delivery of 5-fluorouracil [90]. Additionally, Gupta et al. employed PAMAM dendrimers to transport berberine, a drug with potential anti-cancer activity. They prepared two formulations: in the first, berberine was conjugated with the dendrimers, while in the second, the drug was encapsulated inside the PAMAM structure. They then went on to evaluate the biocompatibility and safety of these formulations in albino rats, observing that the conjugated form was safer than the encapsulated one in this animal model. Moreover, the half-life of berberine when delivered in the conjugated form was also significantly increased when compared to free delivery of the drug [91].
In another study, PPI dendrimers anchored with polysorbate 80 were used to deliver docetaxel in a brain tumor mouse model. These authors found that this nanosystem reduced the tumor volume and moreover, the median survival rate was almost twice as high as in the case of mice treated with PPI-DOX (without polysorbate 80), and more than double when compared to the free delivery of DOX [92]. Finally, when G5 L-lysine dendrimers were used as a vehicle for SN-38 (an active irinotecan metabolite) in a murine colorectal cancer xenograft model, sustained blood levels of SN-38 were detected, which led to significant tumor regression. Moreover, this conjugate exhibited reduced gastrointestinal toxicity compared to free delivery of the drug. They also observed that a regimen of 4 mg/kg SN-38 in 4 doses at weekly intervals extended the survival of the mice to 70 days after delivery of the final dose [93]. More in vivo studies on the use of dendrimers as carriers of anticancer drugs are described in Table 1.
The effect of dendrimer-based preparation has been also studied on animal models of ocular, metabolic, or inflammation diseases. Simvastatin was entrapped into amino-terminated, hydroxyl-terminated, and pegylated PAMAM dendrimers to compare its cholesterol-reducing effects with the free drug in male albino rats. Since entrapment increased simvastatin solubility, it was described that the drug residence time when used in this dendrimer formulation was 3–5 times longer than when the free drug was used. Additionally, drug absorption and elimination rates also decreased significantly, indicating the controlled release of simvastatin from these dendrimer formulations [101]. In a similar vein, phosphorus dendrimers were used to transport azabisphosphonate to target monocytes where they produced anti-inflammatory effects in a model of rheumatoid arthritis in both IL-1ra(−/−) mice (genetically modified mice with silenced gene encoding interleukin-1 receptor antagonist) and mice undergoing K/BxN serum transfer, an animal model of arthritis where the serum from arthritic transgenic K/BxN mice is transferred to naive mice and manifestations of arthritis occur a few days later. This nanosystem suppressed disease by reducing the levels of inflammatory cytokines. Moreover, inhibition of the colony-stimulating factor receptor promoted the maintenance of normal synovial membranes and cartilage, and prevented bone erosion and anti-osteoclastic activity, both in mouse and human cells [102]. In another study, PAMAM dendrimers were used as carriers for pilocarpine nitrate and tropicamide, resulting in prolonged drug residence time for the ophthalmic route when compared to free delivery of the drug [103]. Neutral high-generation phosphorus dendrimers bearing 48 (G3) or 96 (G4) bisphosphonate groups on their surface showed no toxicity and good solubility, as well as chemical stability in aqueous solutions. Furthermore, the anti-inflammatory activity of these neutral phosphorus dendrimers was high in a mouse model of sub-chronic inflammation [76]. Finally, in other works, amino-bis(methylene phosphonate)-capped phosphorus dendrimers prevented the development of experimental autoimmune encephalomyelitis and inhibited the progression of the established disease [104] (Table 2).

4.2. Dendrimers in Infectious Diseases

The antiviral activity of dendrimers is based on mimicking the anionic cell surface, and so dendrimers developed taking this approach also have a negative surface charge. Consequently, viruses bind to dendrimers not to cells, leading to a decrease in the viral infection rate. For example, PLL dendrimers with naphthyl residues were able to inhibit herpes simplex virus infection [112]; and moreover, PAMAM dendrimers modified with naphthyl sulfonate exhibited anti-HIV activity. In both cases, dendrimers inhibited virus entry into cells and replication. Furthermore, other PAMAM dendrimers combined with sialic acid have been studied as inhibitors of influenza A virus infection [113].
Excessive use of antibiotics has been one of the most important factors leading to the emergence of new antibiotic-resistant pathogen variants, meaning that research into new types of anti-bacterial agents remains of vital importance. In this context, dendrimers with a positive surface charge have shown antibacterial activity by binding to the anionic surface of bacterial membranes and damaging it, thereby causing lysis. For instance, PPI dendrimers with alkyl ammonium moieties have been studied to counteract anti-Gram-positive and anti-Gram-negative bacteria [114]. Furthermore, PLL dendrimers with mannosyl surface groups inhibited the adhesion of Escherichia coli to blood cells, making them good candidates as antibacterial drugs [115].
To date, only a few in vivo studies have been carried out to examine the activity of dendrimers against viruses and bacteria. In this sense, lysine dendrimers formulated with SPL7013 were used in mouse and guinea pig models to evaluate anti-herpes simplex virus activity, where they provided protection against infection at 10 mg/mL doses. However, further studies on guinea pigs showed that the optimal dose ranges from 30 to 50 mg/mL [116]. In another study, Landers et al. exploited a G4 PAMAM dendrimer carrying sialic acid to treat infection with three influenza variants in a mouse model. They showed that this nanosystem prevented infection by the H3N2 subtype but was unable to prevent pneumonitis caused by the other two virus types [113]. Elsewhere, researchers synthesized poly(phosphorhydrazone) dendrimers carrying mannose units with oligommanoside caps differing in size, number, and length. They showed that G3 dendrimers with 48 trimannoside caps reduced the neutrophil influx by targeting the DC-SIGN murine homolog, SIGN-related 1. Thus, this nanosystem showed great promise for the treatment of lung inflammation caused by Mycobacterium tuberculosis infection [117]. Even more interestingly, a G2 polyanionic carbosilane dendrimer, G2-S16, with a silica core and 16 sulfonate end-groups exerted anti-HIV-1 activity at an early stage of viral replication, blocking the viral protein gp120/CD4 interaction. Furthermore, topical vaginal administration of a 3% G2-S16 gel prevented HIV-1JR-CSF transmission by 84% in humanized (h)-BLT mice with no presence of HIV-1 RNA in vaginal lesions, thereby taking us one step forward the development of G2-S16-based vaginal microbicides to prevent vaginal HIV-1 transmission in humans [118].

4.3. Dendrimers in Gene Delivery

Gene therapy involves two basic actions: enhancing the expression of or silencing a specific gene, and depending on the application in question, different nucleic acids acting upon the cytosol and/or nucleus are used [119]. For silencing a specific gene, antisense oligonucleotides, or small-interfering RNA (siRNA) are generally used, reaching some of the later the clinical setting [120]. Viral vectors were initially employed for the delivery of genetic material [121,122]. However, because of their immunogenicity, carcinogenicity, and difficulties in large-scale production, their biocompatibility and efficacy has been challenged. Considering all the above, a limited number of approved and commercially used gene therapies are currently available [123]. Thus, the search for the ideal synthetic carrier for genetic material delivery is still underway. The main goals for any novel synthetic nanocarrier would be to efficiently deliver genetic material to patient cells without causing toxicity [124]. Of note, current in vitro studies indicate that dendrimers could be more efficient carriers for genetic materials than biological vectors [125]. Several in vivo studies have been conducted in the central nervous system (Table 3). For instance, PEGylated PAMAM dendrimers carrying angiopep-2, a ligand for lipoprotein receptor-related protein 1 that would facilitate BBB crossing, were used to target glioma cells and deliver tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) as the therapeutic agent. This nanosystem exhibited good BBB penetration, and a favorable biodistribution and pharmacodynamic profile [126].
Elsewhere, cationic phosphorus dendrimers were used to deliver plasmid DNA encoding enhanced green fluorescent protein (GFP). Of note, this nanosystem did not show systemic toxicity after intratumoral injection of the formulation, suggesting that dendriplexes such as these could be good candidates for gene therapy [127]. Another study showed that compared to PAMAM dendrimers, polyethyleneimine (PEI) dendrimers transferred DNA to the airways in BALB/c mouse models more efficiently [128]. Similarly, Mekuria et al. used PAMAM dendrimers to create nanoclusters able to transfer DNA encoding GFP and p53. These authors proved that these dendriplexes were non-toxic to animals and also showed a high gene transfer rate [129]. In other work, poly(ether imine) (PETIM) dendrimers were also used to transport genetic material to enhance the immunogenicity and efficacy of a plasmid-based rabies vaccine in Swiss albino mice. The nanoformulations evaluated produced an earlier onset of a high-titer protective antibody response to rabies virus, thereby suggesting that PETIM dendriplexes were effective carriers for gene-based vaccines [130]. Moreover, Sheikh et al. utilized a polylysine-modified PEI (PEI-PLL) dendrimer to transport the vascular endothelial growth factor (VEGF) gene in a rat model of Parkinson disease where it prevented apoptosis and microglial activation [131].
Table 3. In vivo studies of dendrimers as carriers of genetic material. Transported active drugs are indicated in an italic font.
Table 3. In vivo studies of dendrimers as carriers of genetic material. Transported active drugs are indicated in an italic font.
DendrimerModification/
Active Drug
OutcomeRef.
Phosphorus dendrimerpyrrolidine ammonium
  • lack of systemic toxicity
[127]
Janus
dendrimers
mRNA
  • higher transfection efficiency than the positive control
  • organ specificity
[132]
siRNA
  • ability to deliver Hsp27 siRNA to a castration-resistant prostate cancer mode
  • gene silencing and potent anticancer activity
[133]
mRNA
  • protonated ionizable amines play role in c hanging delivery from the lung to the spleen and/or liver
  • replacing the interconnecting ester with the amide changed the delivery back to the lung
[134]
mRNA
  • targeting the spleen, liver, and lymph nodes
  • simplest synthetic vectors and the first system delivering equally to multiple organs
[135]
PAMAMpolyamidoamine
  • increased transfection in the lung
  • more flexible structure after complexes activation
[136]
fractured polyamidoamine
  • lower gene delivery efficiency in comparison with PEI dendrimers
[128]
lipids
  • increased GFP expression
  • improved serum biochemistry and hematological profile
  • no tissue necrosis
[137]
alkyl-carboxylate chain, PEG, and cholesteryl chloroformate
  • non-toxic, safe vector
  • inhibition of tumor growth
  • efficient gene delivery in TRAIL therapy
[138]
4,4′-dithiodibutryic acid (DA)
  • lack of systemic toxicity
  • efficient delivery of pDNA-p53
  • efficient cell cycle arrest at the G1 phase with upregulated p53 and p21 mRNA and protein expressions
[130]
triethanolamine core
  • efficient delivery of siRNA
  • gene silencing of Hsp27
  • significant anticancer activity in the prostate cancer
[139]
hydroxyl-terminated
  • protecting of payload from degradation
  • effective delivery of siRNA to the cells
  • knockdown of GFP expression
[140]
plasmid DNA
  • increased immunogenicity in comparison with the plasmid vector
[141]
Carbosilane dendrimer siRNA
FITC
  • detection of dendriplexes inside the brain
  • efficient transport of siRNA into the brain
[142]
Phospholipid peptide
dendrimers
-
  • more efficient intracellular uptake and endosome release
  • better siRNA releasing ability
  • more potent gene silencing and anticancer effects
[143]
Poly(ether imine) plasmid vaccine
  • providing 100% protection against virus infection
[130]
PPI plasmid DNA
  • induction of specific immunoglobulins and Th1 response
[144]
Modified dendrimer nanoparticle RNA
  • efficient immunization
[145]
PEI-PLL VEGF gene
  • prevented apoptosis and microglial activation in Parkinson’s disease
  • beneficial effects of PEI-PLL-mediated VEGF gene delivery in the dopaminergic system
[131]

4.4. Dendrimers as Diagnostic Agents

The number of patient cases requiring diagnostic imaging is steadily increasing. Therefore, it will also be essential to develop specific and efficient methods to speed up diagnosis and enable timelier treatments to begin. The imaging agents currently used in magnetic resonance imaging, (MRI), computed tomography (CT), and scintigraphy are of low specificity, have a short half-life, and some toxicity [146], and so their combination with dendrimers is being explored to help overcome these limitations (Table 4). Indeed, dendrimers can increase the efficacy of imaging by, for example, increasing accumulation of the imaging agent in tumor lesions [51]. Moreover, by functionalizing the surface of nanoparticles, these nanosystems can be leveraged to target the specific proteins expressed in certain cell types [78]. Finally, in addition to improving the properties of imaging agents, dendrimers can act as drugs per se to support treatment [147]. Thus, all these approaches can be used to deliver imaging agents to target cells with more specificity, while also limiting the dose needed and, consequently, the side effects of the currently used imaging agents. For example, in one study, PAMAM dendrimers that target epidermal growth factor receptor-2 were used as a contrast agent for CT and MRI of HER-2-positive breast cancer. These authors showed that the use of this nanosystem significantly enhanced MRI signal intensity by about 20% and doubled the CT resolution and contrast in a mouse model, thereby suggesting that it can efficiently target and image HER-2-positive tumors [78].
PAMAM dendrimer-based gold nanoparticles have also been evaluated as a dual-modality contrast agent for MRI and CT imaging of breast cancer to efficiently provide imaging of a xenograft tumor model [148]. Moreover, a more complex nanosystem comprising PAMAM dendrimer-entrapped gold nanoparticles loaded with gadolinium chelator/Gd(III) complexes for targeted dual-mode MRI and CT imaging of small tumors and carrying a RGD peptide produced good imaging results in tumors overexpressing αvβ3 integrin [149]. Similarly, the use of amphiphilic Janus dendrimer-based dendrimersomes to transport an MRI contrast agent and prednisolone resulted in anti-tumor activity in an animal model of melanoma [150]. Finally, a multifunctional dendrimer entrapping gold nanoparticles and gadolinium, and also carrying FA to target xenografts was generated using the keratin-forming HeLa tumor cell line. When tested, this nanosystem showed good potential as an imaging agent, suggesting that such systems could be used to design imaging agents for the diagnosis of different cancer types [151].
Table 4. In vivo studies of dendrimers in diagnostics. Transported active drugs are indicated in an italic font.
Table 4. In vivo studies of dendrimers in diagnostics. Transported active drugs are indicated in an italic font.
DendrimerModification/
Active Drug
OutcomeRef.
PAMAMencapsulated gold nanoparticles, chelated gadolinium, and anti-human HER-2 antibody
  • enhancing of MRI signal intensity by approx. 20%
  • improving CT resolution two times
[78]
thiolated cyclopeptide-based gold nanoparticle
entrapped gold nanoparticles loaded with gadolinium chelator/Gd(III) complexes
  • usability as a dual-mode nanoprobe for targeted CT/MR imaging of different types of αvβ3 integrin-overexpressing cancer
[149]
poly(amidoamine)/gold nanoparticles
gadolinium chelate
  • biocompatibility
  • efficient cellular uptake
  • usability in dual-mode MR/CT imaging of the xenograft tumor model after intravenous injection of the particles
[148]
FA-modified
entrapped gold nanoparticles loaded with gadolinium
  • high intensity of radiation suppression
  • improved MRI contrast
  • usability in dual mode nanoprobes for targeted CT/MR imaging xenograft tumor model in vivo via the FA receptor-mediated active targeting pathway
[149]
gadolinium-loaded dendrimer-entrapped gold nanoparticles
  • extended blood circulation time
  • total clearance within 24 h
  • usability in dual-mode CT/MR imaging of the heart, liver, kidney, and bladder
[152]
surface-PEGylated Gd-PAMAM dendrimers
  • higher relaxivities
  • stability in the blood
  • decreased plasma clearance
  • usability as a carrier for diagnostic or theranostic agents
[153]
N-succinimidyl (S)-acetyl(thiotetraethylene glycol with
Gadolinium
  • possessed a circulation half-life of >1.6 h
  • significant contrast enhancement in the abdominal aorta and kidneys for as long as 4 h
  • reduced circulation time as a result of thiol–disulfide exchange, and the degradation products were rapidly excreted via renal filtration
[154]
Arg-Gly-Asp (RGD)-modified conjugated with Fe3O4 NPs
  • high affinity for C6 cells that overexpress αvβ3 receptors in mice
  • excellent potential for use as contrast agents for targeted T2 MR imaging of specific tumors.
[155]
Janus dendrimersGdDOTAGA(C18)2 and prednisolone phosphate
  • contrast enhancement in the tumor area
  • good efficiency and displayed anti-tumor activity
  • usability in theranostic applications
[150]
GdDOTAGA(C18)2
  • improved relaxivity
  • long-term in vivo retention of the nanoprobe
[156]
poly(propylene imine) dendrimersdensely organized maltose shell (MAL DS)
tetraazacyclododecane tetraacetic acid (DOTA) ligands
  • fast metabolization and total clearance in 48 h
  • efficient contrast agent for MR imaging aorta, renal artery, kidney, and bladder
[157]
peptide dendrimersgadolinium
  • higher signal intensity enhancement
  • much higher Gd(III) concentration in blood
  • usability as an MRI probe
[158]
triazine dendrimergadolinium
  • high in vivo r1 relaxivity, desirable pharmacokinetics, and well-defined structure
[159]
folic acid- and gadolinium-labeled dendrimerFA/graphene oxide
gadolinium
  • long blood circulation time
  • excellent magnetic resonance angiography (MRA) images with high-resolution vascular structures
[160]
polyester dendrimerzwitterionized
  • minimal long-term Gd3+ retention in all organs and tissues
  • degraded into small fragments
  • significant capability of enhancing the MRI of metastases in the liver
[161]

5. Future Perspectives

The continuous development of nanotechnology makes it possible to produce increasingly sophisticated and complex dendrimers that have a wide spectrum of applications in medicine. Moreover, chemical modifications of their surfaces can further expand their capabilities as nanocarriers. Additionally, some dendrimers can act as drugs in and of themselves, as indicated in Section 4.1, thereby also contributing to treatment effectivity. Of note, further work will be essential to thoroughly understand the biological properties and actions of dendrimers on living organisms. To date, several in vitro studies in which dendrimers have proven to be excellent carriers have already been published. Unfortunately, not all the results from this initial research could be confirmed in in vivo preclinical phase drug development trials focusing on selecting optimal drugs and dosages, or in pharmacokinetic and pharmacodynamic studies. Thus, to date, there is scarce information about the chemical stability and long-term toxicity of these nanosystems, and standardized methods to evaluate nanoparticles are also scarce.
Therefore, so far, only a few dendrimer-based formulations have reached clinical trials, including safety and pharmacokinetic studies for hydroxyl-polyamidoamine dendrimer-N-actylcysteine conjugates (NCT03500627), hydroxyl-terminated PAMAM (NCT05105607), or as delivery agents for therapeutic compounds to treat COVID disease (NCT05208996), OP-101, a hydroxyl-polyamidoamine dendrimer-N-acetylcysteine conjugate [162], Phase I/II poly-L-lysine dendrimers for advanced non-Hodgkin lymphoma alone (NCT04214093) or in combination with voriconazole (NCT05205161), and poly-L-Lysine conjugated to a complex of 188Rhenium–ligand (nitro-imidazole-methyl-1,2,3-triazol-methyl-di-[2-pycolyl] amine) for inoperable liver tumor resistant to other therapies (NCT03255343). Although some compounds have successfully moved from in vivo studies to clinical trials, to date, only one dendrimer-based formulation has passed clinical trials to reach the market: VivaGel® (SPL7013), a G4 polylysine dendrimer that was approved for the treatment of bacterial vaginosis and protection against HIV [10]. However, the large amount of research data already published on dendrimers indicates that this fascinating family of nanoparticles has wide-ranging potential in the pharmaceutical industry, especially for applications in drug delivery systems. Indeed, the use of cutting-edge technologies has made it possible to develop advanced systems for delivering active agents into living systems. However, further research on dendrimers will still be needed to allow for medicine to reach previously unknown levels of sophistication through nanosystems, thereby making it possible to manage diseases that currently lack effective treatments.

Author Contributions

Conceptualization, K.S. and V.C.; Methodology, K.S. and J.L.R.-G.; Validation, J.L.R.-G. and V.C.; Formal Analysis, V.C.; Resources, V.C.; Writing—Original Draft Preparation, K.S.; Writing—Review and Editing, V.C.; Supervision, V.C.; Project Administration, V.C.; Funding Acquisition, V.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by MCINN and JCCM, with funding from the European Union NextGenerationEU (PRTR-C17.I1); project PID2020-120134RB-I00 from MINECO/AEI/FEDER/UE (MCIN/AEI/10.13039/501100011033), and University of Castilla-La Mancha project 2022-GRIN-34370 to V.C. K.S. is thankful for the support from the Foundation for Polish Science (FNP).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were generated for this manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sztandera, K.; Gorzkiewicz, M.; Klajnert-Maculewicz, B. Gold Nanoparticles in Cancer Treatment. Mol. Pharm. 2019, 16, 1–23. [Google Scholar] [CrossRef]
  2. Aliabadi, H.M.; Lavasanifar, A. Polymeric micelles for drug delivery. Expert Opin. Drug Deliv. 2006, 3, 139–162. [Google Scholar] [CrossRef]
  3. Sztandera, K.; Gorzkiewicz, M.; Wang, X.; Boye, S.; Appelhans, D.; Klajnert-Maculewicz, B. pH-stable polymersome as nanocarrier for post-loaded rose bengal in photodynamic therapy. Colloids Surf. B Biointerfaces 2022, 217, 112662. [Google Scholar] [CrossRef]
  4. Mignani, S.; Shi, X.; Ceña, V.; Rodrigues, J.; Tomas, H.; Majoral, J.P. Engineered non-invasive functionalized dendrimer/dendron-entrapped/complexed gold nanoparticles as a novel class of theranostic (radio)pharmaceuticals in cancer therapy. J. Control. Release. 2024, 332, 346–366. [Google Scholar] [CrossRef]
  5. Percec, V.; Sahoo, D. From Frank-Kasper, Quasicrystals, and Biological Membrane Mimics to Reprogramming In Vivo the Living Factory to Target the Delivery of mRNA with One-Component Amphiphilic Janus Dendrimers. Biomacromolecules 2024, 25, 1353–1370. [Google Scholar] [CrossRef]
  6. Bozzuto, G.; Molinari, A. Liposomes as nanomedical devices. Int. J. Nanomed. 2015, 10, 975–999. [Google Scholar] [CrossRef]
  7. Svenson, S.; Tomalia, D.A. Dendrimers in biomedical applications--reflections on the field. Adv. Drug Deliv. Rev. 2005, 57, 2106–2129. [Google Scholar] [CrossRef]
  8. Tomalia, D.A.; Baker, H.; Dewald, J.; Hall, M.; Kallos, G.; Martin, S.; Roeck, J.; Ryder, J.; Smith, P. A new class of polymers: Starburst-dendritic macromolecules. Polym. J. 1985, 17, 117–132. [Google Scholar] [CrossRef]
  9. Wang, J.; Li, B.; Qiu, L.; Qiao, X.; Yang, H. Dendrimer-based drug delivery systems: History, challenges, and latest developments. J. Biol. Eng. 2022, 16, 18. [Google Scholar] [CrossRef]
  10. Fox, L.J.; Richardson, R.M.; Briscoe, W.H. PAMAM dendrimer—Cell membrane interactions. Adv. Colloid. Interface Sci. 2018, 257, 1–18. [Google Scholar] [CrossRef]
  11. Stenstrom, P.; Manzanares, D.; Zhang, Y.; Cena, V.; Malkoch, M. Evaluation of Amino-Functional Polyester Dendrimers Based on Bis-MPA as Nonviral Vectors for siRNA Delivery. Molecules 2018, 23, 2028. [Google Scholar] [CrossRef]
  12. Singh, V.; Sahebkar, A.; Kesharwani, P. Poly (propylene imine) dendrimer as an emerging polymeric nanocarrier for anticancer drug and gene delivery. Eur. Polym. J. 2021, 158, 110683. [Google Scholar] [CrossRef]
  13. Mignani, S.; Shi, X.; Cena, V.; Shcharbin, D.; Bryszewska, M.; Majoral, J.P. In vivo therapeutic applications of phosphorus dendrimers: State of the art. Drug Discov. Today 2021, 26, 677–689. [Google Scholar] [CrossRef]
  14. Lu, J.; Atochina-Vasserman, E.N.; Maurya, D.S.; Shalihin, M.I.; Zhang, D.; Chenna, S.S.; Adamson, J.; Liu, M.; Ur, H.; Shah, R.; et al. Screening Libraries to Discover Molecular Design Principles for the Targeted Delivery of mRNA with One-Component Ionizable Amphiphilic Janus Dendrimers Derived from Plant Phenolic Acids. Pharmaceutics 2023, 15, 1572. [Google Scholar] [CrossRef]
  15. Sadler, K.; Tam, J.P. Peptide dendrimers: Applications and synthesis. Rev. Mol. Biotechnol. 2002, 90, 195–229. [Google Scholar] [CrossRef]
  16. Kwan, C.S.; Leung, K.C.F. Development and advancement of rotaxane dendrimers as switchable macromolecular machines. Mater. Chem. Front. 2020, 4, 2825–2844. [Google Scholar] [CrossRef]
  17. Sánchez-Nieves, J.; Ortega, P.; Muñoz-Fernández, M.Á.; Gómez, R.; De La Mata, F.J. Synthesis of carbosilane dendrons and dendrimers derived from 1,3,5-trihydroxybenzene. Tetrahedron 2010, 66, 9203–9213. [Google Scholar] [CrossRef]
  18. Xu, X.; Yuan, H.; Chang, J.; He, B.; Gu, Z. Cooperative Hierarchical Self-Assembly of Peptide Dendrimers and Linear Polypeptides into Nanoarchitectures Mimicking Viral Capsids. Angew. Chem. 2012, 124, 3184–3187. [Google Scholar] [CrossRef]
  19. Türp, D.; Nguyen, T.T.T.; Baumgarten, M.; Müllen, K. Uniquely versatile: Nano-site defined materials based on polyphenylene dendrimers. New J. Chem. 2012, 36, 282–298. [Google Scholar] [CrossRef]
  20. Feliu, N.; Walter, M.V.; Montañez, M.I.; Kunzmann, A.; Hult, A.; Nyström, A.; Malkoch, M.; Fadeel, B. Stability and biocompatibility of a library of polyester dendrimers in comparison to polyamidoamine dendrimers. Biomaterials 2012, 33, 1970–1981. [Google Scholar] [CrossRef]
  21. Simanek, E.E. Two Decades of Triazine Dendrimers. Molecules 2021, 26, 4774. [Google Scholar] [CrossRef]
  22. Turnbull, W.B.; Stoddart, J.F. Design and synthesis of glycodendrimers. J. Biotechnol. 2002, 90, 231–255. [Google Scholar] [CrossRef]
  23. Zhang, S.; Sun, H.J.; Hughes, A.D.; Moussodia, R.O.; Bertin, A.; Chen, Y.; Pochan, D.J.; Heiney, P.A.; Klein, M.L.; Percec, V. Self-assembly of amphiphilic Janus dendrimers into uniform onion-like dendrimersomes with predictable size and number of bilayers. Proc. Natl. Acad. Sci. USA 2014, 111, 9058–9063. [Google Scholar] [CrossRef]
  24. Sahoo, D.; Atochina-Vasserman, E.N.; Maurya, D.S.; Arshad, M.; Chenna, S.S.; Ona, N.; Vasserman, J.A.; Ni, H.; Weissman, D.; Percec, V. The Constitutional Isomerism of One-Component Ionizable Amphiphilic Janus Dendrimers Orchestrates the Total and Targeted Activities of mRNA Delivery. J. Am. Chem. Soc. 2024, 146, 3627–3634. [Google Scholar] [CrossRef]
  25. Pérez-Ferreiro, M.; Abelairas, A.M.; Criado, A.; Gómez, I.J.; Mosquera, J. Dendrimers: Exploring Their Wide Structural Variety and Applications. Polymers 2023, 15, 4369. [Google Scholar] [CrossRef]
  26. Tomalia, D.A. Birth of a new macromolecular architecture: Dendrimers as quantized building blocks for nanoscale synthetic polymer chemistry. Prog. Polym. Sci. 2005, 30, 294–324. [Google Scholar] [CrossRef]
  27. Bosman, A.W.; Janssen, H.M.; Meijer, E.W. About Dendrimers: Structure, Physical Properties, and Applications. Chem. Rev. 1999, 99, 1665–1688. [Google Scholar] [CrossRef]
  28. Franc, G.; Kakkar, A.K. Diels-alder “click” chemistry in designing dendritic macromolecules. Chem.—A Eur. J. 2009, 15, 5630–5639. [Google Scholar] [CrossRef]
  29. Shen, Y.; Ma, Y.; Li, Z. Facile synthesis of dendrimers combining aza-Michael addition with thiol-yne click chemistry. J. Polym. Sci. Part A Polym. Chem. 2013, 51, 708–715. [Google Scholar] [CrossRef]
  30. Ghirardello, M.; Öberg, K.; Staderini, S.; Renaudet, O.; Berthet, N.; Dumy, P.; Hed, Y.; Marra, A.; Malkoch, M.; Dondoni, A. Thiol-ene and thiol-yne-based synthesis of glycodendrimers as nanomolar inhibitors of wheat germ agglutinin. J. Polym. Sci. Part A Polym. Chem. 2014, 52, 2422–2433. [Google Scholar] [CrossRef]
  31. Hizal, G.; Tunca, U.; Sanyal, A. Discrete macromolecular constructs via the Diels-Alder “click” reaction. J. Polym. Sci. Part A Polym. Chem. 2011, 49, 4103–4120. [Google Scholar] [CrossRef]
  32. Bober, Z.; Bartusik-Aebisher, D.; Aebisher, D. Application of Dendrimers in Anticancer Diagnostics and Therapy. Molecules 2022, 27, 3237. [Google Scholar] [CrossRef]
  33. Mintzer, M.A.; Grinstaff, M.W. Biomedical applications of dendrimers: A tutorial. Chem. Soc. Rev. 2011, 40, 173–190. [Google Scholar] [CrossRef]
  34. Biricova, V.; Laznickova, A. Dendrimers: Analytical characterization and applications. Bioorganic Chem. 2009, 37, 185–192. [Google Scholar] [CrossRef] [PubMed]
  35. Fréchet, J.M.J. Functional polymers and dendrimers: Reactivity, molecular architecture, and interfacial energy. Science 1994, 263, 1710–1715. [Google Scholar] [CrossRef] [PubMed]
  36. Trinchi, A.; Muster, T.H. A review of surface functionalized amine terminated dendrimers for application in biological and molecular sensing. Supramol. Chem. 2007, 19, 431–445. [Google Scholar] [CrossRef]
  37. Fischer, M.; Vogtle, F. Dendrimers: From design to application—A progress report. Angew. Chem.—Int. Ed. 1999, 38, 884–905. [Google Scholar] [CrossRef]
  38. Gillies, E.R.; Fréchet, J.M.J. Dendrimers and dendritic polymers in drug delivery. Drug Discov. Today 2005, 10, 35–43. [Google Scholar] [CrossRef] [PubMed]
  39. D’Emanuele, A.; Attwood, D. Dendrimer-drug interactions. Adv. Drug Deliv. Rev. 2005, 57, 2147–2162. [Google Scholar] [CrossRef] [PubMed]
  40. Kojima, C.; Kono, K.; Maruyama, K.; Takagishi, T. Synthesis of polyamidoamine dendrimers having poly(ethylene glycol) grafts and their ability to encapsulate anticancer drugs. Bioconjugate Chem. 2000, 11, 910–917. [Google Scholar] [CrossRef] [PubMed]
  41. Leriche, G.; Chisholm, L.; Wagner, A. Cleavable linkers in chemical biology. Bioorganic Med. Chem. 2012, 20, 571–582. [Google Scholar] [CrossRef]
  42. Sztandera, K.; Marcinkowska, M.; Gorzkiewicz, M.; Janaszewska, A.; Laurent, R.; Zabłocka, M.; Mignani, S.; Majoral, J.P.; Klajnert-Maculewicz, B. In search of a phosphorus dendrimer-based carrier of rose bengal: Tyramine linker limits fluorescent and phototoxic properties of a photosensitizer. Int. J. Mol. Sci. 2020, 21, 4456. [Google Scholar] [CrossRef]
  43. Moreira, D.A.; Santos, S.D.; Leiro, V.; Pego, A.P. Dendrimers and Derivatives as Multifunctional Nanotherapeutics for Alzheimer’s Disease. Pharmaceutics 2023, 15, 1054. [Google Scholar] [CrossRef] [PubMed]
  44. Kongcharoen, H.; Coester, B.; Yu, F.; Aziz, I.; Poh, W.C.; Tan, M.W.M.; Tonanon, P.; Ciou, J.H.; Chan, B.; Webster, R.D.; et al. Magnetically Directed Co-nanoinitiators for Cross-Linking Adhesives and Enhancing Mechanical Properties. ACS Appl. Mater. Interfaces 2021, 13, 57851–57863. [Google Scholar] [CrossRef]
  45. Zhou, X.; Xu, X.; Hu, Q.; Wu, Y.; Yu, F.; He, C.; Qian, Y.; Han, Y.; Tang, J.; Hu, H. Novel manganese and polyester dendrimer-based theranostic nanoparticles for MRI and breast cancer therapy. J. Mater. Chem. B 2022, 11, 648–656. [Google Scholar] [CrossRef] [PubMed]
  46. Pedziwiatr-Werbicka, E.; Serchenya, T.; Shcharbin, D.; Terekhova, M.; Prokhira, E.; Dzmitruk, V.; Shyrochyna, I.; Sviridov, O.; Peña-González, C.E.; Gómez, R.; et al. Dendronization of gold nanoparticles decreases their effect on human alpha-1-microglobulin. Int. J. Biol. Macromol. 2018, 108, 936–941. [Google Scholar] [CrossRef] [PubMed]
  47. Garoff, H.; Simons, K. Location of the spike glycoproteins in the Semliki Forest virus membrane. Proc. Natl. Acad. Sci. USA 1974, 71, 3988–3992. [Google Scholar] [CrossRef]
  48. Zhang, P.; Li, Z.; Cao, W.; Tang, J.; Xia, Y.; Peng, L.; Ma, J. A PD-L1 Antibody-Conjugated PAMAM Dendrimer Nanosystem for Simultaneously Inhibiting Glycolysis and Promoting Immune Response in Fighting Breast Cancer. Adv. Mater. 2023, 35, e2305215. [Google Scholar] [CrossRef]
  49. Gorzkiewicz, M.; Marcinkowska, M.; Studzian, M.; Karwaciak, I.; Pulaski, L.; Klajnert-Maculewicz, B. Mesalazine–PAMAM Nanoparticles for Transporter-Independent Intracellular Drug Delivery: Cellular Uptake and Anti-Inflammatory Activity. Int. J. Nanomed. 2023, 18, 2109–2126. [Google Scholar] [CrossRef]
  50. Chowdhury, S.; Toth, I.; Stephenson, R.J. Dendrimers in vaccine delivery: Recent progress and advances. Biomaterials 2022, 280, 121303. [Google Scholar] [CrossRef]
  51. Bertrand, N.; Wu, J.; Xu, X.; Kamaly, N.; Farokhzad, O.C. Cancer nanotechnology: The impact of passive and active targeting in the era of modern cancer biology. Adv. Drug Deliv. Rev. 2014, 66, 2–25. [Google Scholar] [CrossRef] [PubMed]
  52. McMahon, H.T.; Boucrot, E. Molecular mechanism and physiological functions of clathrin-mediated endocytosis. Nat. Rev. Mol. Cell Biol. 2011, 12, 517–533. [Google Scholar] [CrossRef] [PubMed]
  53. Manzanares, D.; Cena, V. Endocytosis: The Nanoparticle and Submicron Nanocompounds Gateway into the Cell. Pharmaceutics 2020, 12, 371. [Google Scholar] [CrossRef] [PubMed]
  54. Goldstein, J.L.; Brown, M.S. The LDL receptor. Arter. Thromb. Vasc. Biol. 2009, 29, 431–438. [Google Scholar] [CrossRef]
  55. Sigismund, S.; Confalonieri, S.; Ciliberto, A.; Polo, S.; Scita, G.; Di Fiore, P.P. Endocytosis and signaling: Cell logistics shape the eukaryotic cell plan. Physiol. Rev. 2012, 92, 273–366. [Google Scholar] [CrossRef] [PubMed]
  56. Lakadamyali, M.; Rust, M.J.; Zhuang, X. Ligands for clathrin-mediated endocytosis are differentially sorted into distinct populations of early endosomes. Cell 2006, 124, 997–1009. [Google Scholar] [CrossRef] [PubMed]
  57. Saheki, Y.; De Camilli, P. Synaptic vesicle endocytosis. Cold Spring Harb. Perspect. Biol. 2012, 4, a005645. [Google Scholar] [CrossRef] [PubMed]
  58. Pelkmans, L.; Bürli, T.; Zerial, M.; Helenius, A. Caveolin-stabilized membrane domains as multifunctional transport and sorting devices in endocytic membrane traffic. Cell 2004, 118, 767–780. [Google Scholar] [CrossRef]
  59. Mayor, S.; Parton, R.G.; Donaldson, J.G. Clathrin-independent pathways of endocytosis. Cold Spring Harb. Perspect. Biol. 2014, 6, a016758. [Google Scholar] [CrossRef]
  60. Di Guglielmo, G.M.; Le Roy, C.; Goodfellow, A.F.; Wrana, J.L. Distinct endocytic pathways regulate TGF-beta receptor signalling and turnover. Nat. Cell Biol. 2003, 5, 410–421. [Google Scholar] [CrossRef]
  61. Rossetto, O.; Pirazzini, M.; Montecucco, C. Botulinum neurotoxins: Genetic, structural and mechanistic insights. Nat. Rev. Microbiol. 2014, 12, 535–549. [Google Scholar] [CrossRef]
  62. Commisso, C.; Davidson, S.M.; Soydaner-Azeloglu, R.G.; Parker, S.J.; Kamphorst, J.J.; Hackett, S.; Grabocka, E.; Nofal, M.; Drebin, J.A.; Thompson, C.B.; et al. Macropinocytosis of protein is an amino acid supply route in Ras-transformed cells. Nature 2013, 497, 633–637. [Google Scholar] [CrossRef]
  63. Mercer, J.; Helenius, A. Gulping rather than sipping: Macropinocytosis as a way of virus entry. Curr. Opin. Microbiol. 2012, 15, 490–499. [Google Scholar] [CrossRef] [PubMed]
  64. Norbury, C.C.; Chambers, B.J.; Prescott, A.R.; Ljunggren, H.G.; Watts, C. Constitutive macropinocytosis allows TAP-dependent major histocompatibility complex class I presentation of exogenous soluble antigen by bone marrow-derived dendritic cells. Eur. J. Immunol. 1997, 27, 280–288. [Google Scholar] [CrossRef]
  65. Albertazzi, L.; Serresi, M.; Albanese, A.; Beltram, F. Dendrimer internalization and intracellular trafficking in living cells. Mol. Pharm. 2010, 7, 680–688. [Google Scholar] [CrossRef]
  66. Perumal, O.P.; Inapagolla, R.; Kannan, S.; Kannan, R.M. The effect of surface functionality on cellular trafficking of dendrimers. Biomaterials 2008, 29, 3469–3476. [Google Scholar] [CrossRef] [PubMed]
  67. Bastien, E.; Schneider, R.; Hackbarth, S.; Dumas, D.; Jasniewski, J.; Röder, B.; Bezdetnaya, L.; Lassalle, H.P. PAMAM G4.5-chlorin e6 dendrimeric nanoparticles for enhanced photodynamic effects. Photochem. Photobiol. Sci. 2015, 14, 2203–2212. [Google Scholar] [CrossRef]
  68. Tabasi, H.; Babaei, M.; Abnous, K.; Taghdisi, S.M.; Saljooghi, A.S.; Ramezani, M.; Alibolandi, M. Metal–polymer-coordinated complexes as potential nanovehicles for drug delivery. J. Nanostruct. Chem. 2021, 11, 501–526. [Google Scholar] [CrossRef]
  69. Oddone, N.; Lecot, N.; Fernández, M.; Rodriguez-Haralambides, A.; Cabral, P.; Cerecetto, H.; Benech, J.C. In vitro and in vivo uptake studies of PAMAM G4.5 dendrimers in breast cancer. J. Nanobiotechnol. 2016, 14, 45. [Google Scholar] [CrossRef]
  70. Konda, S.D.; Aref, M.; Wang, S.; Brechbiel, M.; Wiener, E.C. Specific targeting of folate-dendrimer MRI contrast agents to the high affinity folate receptor expressed in ovarian tumor xenografts. Magn. Reson. Mater. Phys. Biol. Med. 2001, 12, 104–113. [Google Scholar] [CrossRef]
  71. Kharwade, R.; Badole, P.; Mahajan, N.; More, S. Toxicity and Surface Modification of Dendrimers: A Critical Review. Curr. Drug Deliv. 2022, 19, 451–465. [Google Scholar] [CrossRef] [PubMed]
  72. Shcharbin, D.; Janaszewska, A.; Klajnert-Maculewicz, B.; Ziemba, B.; Dzmitruk, V.; Halets, I.; Loznikova, S.; Shcharbina, N.; Milowska, K.; Ionov, M.; et al. How to study dendrimers and dendriplexes III. Biodistribution, pharmacokinetics and toxicity in vivo. J. Control. Release 2014, 181, 40–52. [Google Scholar] [CrossRef] [PubMed]
  73. Mahdavijalal, M.; Ahmad Panahi, H.; Moniri, E. Synthesis of PAMAM dendrimers anchored to WS2 nano-sheets for controlled delivery of docetaxel: Design, characterization and in vitro drug release. J. Drug Deliv. Sci. Technol. 2023, 79, 104066. [Google Scholar] [CrossRef]
  74. Sastri, K.T.; Gupta, N.V.; Sharadha, M.; Chakraborty, S.; Kumar, H.; Chand, P.; Balamuralidhara, V.; Gowda, D.V. Nanocarrier facilitated drug delivery to the brain through intranasal route: A promising approach to transcend bio-obstacles and alleviate neurodegenerative conditions. J. Drug Deliv. Sci. Technol. 2022, 75, 103656. [Google Scholar] [CrossRef]
  75. Benseny-Cases, N.; Álvarez-Marimon, E.; Aso, E.; Carmona, M.; Klementieva, O.; Appelhans, D.; Ferrer, I.; Cladera, J. In situ identification and G4-PPI-His-Mal-dendrimer-induced reduction of early-stage amyloid aggregates in Alzheimer’s disease transgenic mice using synchrotron-based infrared imaging. Sci. Rep. 2021, 11, 18368. [Google Scholar] [CrossRef]
  76. Posadas, I.; Romero-Castillo, L.; El, B.N.; Manzanares, D.; Mignani, S.; Majoral, J.P.; Cena, V. Neutral high-generation phosphorus dendrimers inhibit macrophage-mediated inflammatory response in vitro and in vivo. Proc. Natl. Acad. Sci. USA 2017, 114, E7660–E7669. [Google Scholar] [CrossRef]
  77. Oliveira, I.M.; Goncalves, C.; Oliveira, E.P.; Simon-Vazquez, R.; da Silva Morais, A.; Gonzalez-Fernandez, A.; Reis, R.L.; Oliveira, J.M. PAMAM dendrimers functionalised with an anti-TNF alpha antibody and chondroitin sulphate for treatment of rheumatoid arthritis. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 121, 111845. [Google Scholar] [CrossRef]
  78. Chen, J.S.; Chen, J.; Bhattacharjee, S.; Cao, Z.; Wang, H.; Swanson, S.D.; Zong, H.; Baker, J.R.; Wang, S.H. Functionalized nanoparticles with targeted antibody to enhance imaging of breast cancer in vivo. J. Nanobiotechnol. 2020, 18, 135. [Google Scholar] [CrossRef]
  79. Soltany, P.; Miralinaghi, M.; Pajoum Shariati, F. Folic acid conjugated poly (Amidoamine) dendrimer grafted magnetic chitosan as a smart drug delivery platform for doxorubicin: In-vitro drug release and cytotoxicity studies. Int. J. Biol. Macromol. 2024, 257, 127564. [Google Scholar] [CrossRef]
  80. Tan, G.; Li, J.; Liu, D.; Pan, H.; Zhu, R.; Yang, Y.; Pan, W. Amino acids functionalized dendrimers with nucleus accumulation for efficient gene delivery. Int. J. Pharm. 2021, 602, 120641. [Google Scholar] [CrossRef]
  81. Schito, A.M.; Schito, G.C.; Alfei, S. Synthesis and Antibacterial Activity of Cationic Amino Acid-Conjugated Dendrimers Loaded with a Mixture of Two Triterpenoid Acids. Polymers 2021, 13, 521. [Google Scholar] [CrossRef] [PubMed]
  82. Jatczak-Pawlik, I.; Gorzkiewicz, M.; Studzian, M.; Zinke, R.; Appelhans, D.; Klajnert-Maculewicz, B.; Pulaski, L. Nanoparticles for Directed Immunomodulation: Mannose-Functionalized Glycodendrimers Induce Interleukin-8 in Myeloid Cell Lines. Biomacromolecules 2021, 22, 3396–3407. [Google Scholar] [CrossRef] [PubMed]
  83. Sharma, R.; Liaw, K.; Sharma, A.; Jimenez, A.; Chang, M.; Salazar, S.; Amlani, I.; Kannan, S.; Kannan, R.M. Glycosylation of PAMAM dendrimers significantly improves tumor macrophage targeting and specificity in glioblastoma. J. Control. Release 2021, 337, 179–192. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, X.; Zhang, M.; Li, Y.; Cong, H.; Yu, B.; Shen, Y. Research Status of Dendrimer Micelles in Tumor Therapy for Drug Delivery. Small 2023, 19, e2304006. [Google Scholar] [CrossRef] [PubMed]
  85. de la Torre, C.; Jativa, P.; Posadas, I.; Manzanares, D.; Blanco, J.L.J.; Mellet, C.O.; Fernandez, J.M.G.; Cena, V. A beta-Cyclodextrin-Based Nanoparticle with Very High Transfection Efficiency Unveils siRNA-Activated TLR3 Responses in Human Prostate Cancer Cells. Pharmaceutics 2022, 14, 2424. [Google Scholar] [CrossRef]
  86. Caminade, A.M.; Turrin, C.O. Dendrimers for drug delivery. J. Mater. Chem. B 2014, 2, 4055–4066. [Google Scholar] [CrossRef]
  87. Chen, L.; Cao, L.; Zhan, M.; Li, J.; Wang, D.; Laurent, R.; Mignani, S.; Caminade, A.M.; Majoral, J.P.; Shi, X. Engineered Stable Bioactive per Se Amphiphilic Phosphorus Dendron Nanomicelles as a Highly Efficient Drug Delivery System to Take Down Breast Cancer in Vivo. Biomacromolecules 2022, 23, 2827–2837. [Google Scholar] [CrossRef]
  88. Huang, S.; Huang, X.; Yan, H. Peptide dendrimers as potentiators of conventional chemotherapy in the treatment of pancreatic cancer in a mouse model. Eur. J. Pharm. Biopharm. 2022, 170, 121–132. [Google Scholar] [CrossRef]
  89. Wu, S.Y.; Chou, H.Y.; Yuh, C.H.; Mekuria, S.L.; Kao, Y.C.; Tsai, H.C. Radiation-Sensitive Dendrimer-Based Drug Delivery System. Adv. Sci. 2018, 5, 1700339. [Google Scholar] [CrossRef]
  90. Bhadra, D.; Bhadra, S.; Jain, S.; Jain, N.K. A PEGylated dendritic nanoparticulate carrier of fluorouracil. Int. J. Pharm. 2003, 257, 111–124. [Google Scholar] [CrossRef]
  91. Gupta, L.; Sharma, A.K.; Gothwal, A.; Khan, M.S.; Khinchi, M.P.; Qayum, A.; Singh, S.K.; Gupta, U. Dendrimer encapsulated and conjugated delivery of berberine: A novel approach mitigating toxicity and improving in vivo pharmacokinetics. Int. J. Pharm. 2017, 528, 88–99. [Google Scholar] [CrossRef]
  92. Gajbhiye, V.; Jain, N.K. The treatment of Glioblastoma Xenografts by surfactant conjugated dendritic nanoconjugates. Biomaterials 2011, 32, 6213–6225. [Google Scholar] [CrossRef]
  93. England, R.M.; Hare, J.I.; Barnes, J.; Wilson, J.; Smith, A.; Strittmatter, N.; Kemmitt, P.D.; Waring, M.J.; Barry, S.T.; Alexander, C.; et al. Tumour regression and improved gastrointestinal tolerability from controlled release of SN-38 from novel polyoxazoline-modified dendrimers. J. Control. Release 2017, 247, 73–85. [Google Scholar] [CrossRef] [PubMed]
  94. Kirkpatrick, G.J.; Plumb, J.A.; Sutcliffe, O.B.; Flint, D.J.; Wheate, N.J. Evaluation of anionic half generation 3.5-6.5 poly(amidoamine) dendrimers as delivery vehicles for the active component of the anticancer drug cisplatin. J. Inorg. Biochem. 2011, 105, 1115–1122. [Google Scholar] [CrossRef]
  95. Luo, J.; Lin, X.X.; Li, L.L.; Tan, J.Q.; Li, P. β-Cyclodextrin and Oligoarginine Peptide-Based Dendrimer-Entrapped Gold Nanoparticles for Improving Drug Delivery to the Inner Ear. Front. Bioeng. Biotechnol. 2022, 10, 844177. [Google Scholar] [CrossRef]
  96. Xu, X.; Li, J.; Han, S.; Tao, C.; Fang, L.; Sun, Y.; Zhu, J.; Liang, Z.; Li, F. A novel doxorubicin loaded folic acid conjugated PAMAM modified with borneol, a nature dual-functional product of reducing PAMAM toxicity and boosting BBB penetration. Eur. J. Pharm. Sci. 2016, 88, 178–190. [Google Scholar] [CrossRef] [PubMed]
  97. Pishavar, E.; Ramezani, M.; Hashemi, M. Co-delivery of doxorubicin and TRAIL plasmid by modified PAMAM dendrimer in colon cancer cells, in vitro and in vivo evaluation. Drug Dev. Ind. Pharm. 2019, 45, 1931–1939. [Google Scholar] [CrossRef]
  98. Alibolandi, M.; Hoseini, F.; Mohammadi, M.; Ramezani, P.; Einafshar, E.; Taghdisi, S.M.; Ramezani, M.; Abnous, K. Curcumin-entrapped MUC-1 aptamer targeted dendrimer-gold hybrid nanostructure as a theranostic system for colon adenocarcinoma. Int. J. Pharm. 2018, 549, 67–75. [Google Scholar] [CrossRef]
  99. Li, N.; Guo, C.; Duan, Z.; Yu, L.; Luo, K.; Lu, J.; Gu, Z. A stimuli-responsive Janus peptide dendron-drug conjugate as a safe and nanoscale drug delivery vehicle for breast cancer therapy. J. Mater. Chem. B 2016, 4, 3760–3769. [Google Scholar] [CrossRef]
  100. Tardi, P.; Johnstone, S.; Harasym, N.; Xie, S.; Harasym, T.; Zisman, N.; Harvie, P.; Bermudes, D.; Mayer, L. In vivo maintenance of synergistic cytarabine:daunorubicin ratios greatly enhances therapeutic efficacy. Leuk. Res. 2009, 33, 129–139. [Google Scholar] [CrossRef]
  101. Kulhari, H.; Kulhari, D.P.; Prajapati, S.K.; Chauhan, A.S. Pharmacokinetic and pharmacodynamic studies of poly(amidoamine) dendrimer based simvastatin oral formulations for the treatment of hypercholesterolemia. Mol. Pharm. 2013, 10, 2528–2533. [Google Scholar] [CrossRef]
  102. Hayder, M.; Poupot, M.; Baron, M.; Nigon, D.; Turrin, C.O.; Caminade, A.M.; Majoral, J.P.; Eisenberg, R.A.; Fournié, J.J.; Cantagrel, A.; et al. A phosphorus-based dendrimer targets inflammation and osteoclastogenesis in experimental arthritis. Sci. Transl. Med. 2011, 3, 81ra35. [Google Scholar] [CrossRef]
  103. Vandamme, T.F.; Brobeck, L. Poly(amidoamine) dendrimers as ophthalmic vehicles for ocular delivery of pilocarpine nitrate and tropicamide. J. Control. Release 2005, 102, 23–38. [Google Scholar] [CrossRef]
  104. Hayder, M.; Varilh, M.; Turrin, C.O.; Saoudi, A.; Caminade, A.M.; Poupot, R.; Liblau, R.S. Phosphorus-Based Dendrimer ABP Treats Neuroinflammation by Promoting IL-10-Producing CD4+ T Cells. Biomacromolecules 2015, 16, 3425–3433. [Google Scholar] [CrossRef]
  105. Fruchon, S.; Mouriot, S.; Thiollier, T.; Grandin, C.; Caminade, A.M.; Turrin, C.O.; Contamin, H.; Poupot, R. Repeated intravenous injections in non-human primates demonstrate preclinical safety of an anti-inflammatory phosphorus-based dendrimer. Nanotoxicology 2015, 9, 433–441. [Google Scholar] [CrossRef]
  106. Spataro, G.; Malecaze, F.; Turrin, C.O.; Soler, V.; Duhayon, C.; Elena, P.P.; Majoral, J.P.; Caminade, A.M. Designing dendrimers for ocular drug delivery. Eur. J. Med. Chem. 2010, 45, 326–334. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, L.; Shi, C.; Wang, X.; Guo, D.; Duncan, T.M.; Luo, J. Zwitterionic Janus Dendrimer with distinct functional disparity for enhanced protein delivery. Biomaterials 2019, 215, 119233. [Google Scholar] [CrossRef]
  108. Akhtar, S.; Chandrasekhar, B.; Yousif, M.H.M.; Renno, W.; Benter, I.F.; El-Hashim, A.Z. Chronic administration of nano-sized PAMAM dendrimers in vivo inhibits EGFR-ERK1/2-ROCK signaling pathway and attenuates diabetes-induced vascular remodeling and dysfunction. Nanomedicine: Nanotechnol. Biol. Med. 2019, 18, 78–89. [Google Scholar] [CrossRef]
  109. Labieniec-Watala, M.; Przygodzki, T.; Sebekova, K.; Watala, C. Can metabolic impairments in experimental diabetes be cured with poly(amido)amine (PAMAM) G4 dendrimers?—In the search for minimizing of the adverse effects of PAMAM administration. Int. J. Pharm. 2014, 464, 152–167. [Google Scholar] [CrossRef] [PubMed]
  110. Karolczak, K.; Rozalska, S.; Wieczorek, M.; Labieniec-Watala, M.; Watala, C. Poly(amido)amine dendrimers generation 4.0 (PAMAM G4) reduce blood hyperglycaemia and restore impaired blood-brain barrier permeability in streptozotocin diabetes in rats. Int. J. Pharm. 2012, 436, 508–518. [Google Scholar] [CrossRef] [PubMed]
  111. Jiang, Y.; Zhao, W.; Xu, S.; Wei, J.; Lasaosa, F.L.; He, Y.; Mao, H.; Bolea Bailo, R.M.; Kong, D.; Gu, Z. Bioinspired design of mannose-decorated globular lysine dendrimers promotes diabetic wound healing by orchestrating appropriate macrophage polarization. Biomaterials 2022, 280, 121323. [Google Scholar] [CrossRef]
  112. Bourne, N.; Stanberry, L.R.; Kern, E.R.; Holan, G.; Matthews, B.; Bernstein, D.I. Dendrimers, a new class of candidate topical microbicides with activity against herpes simplex virus infection. Antimicrob. Agents Chemother. 2000, 44, 2471–2474. [Google Scholar] [CrossRef]
  113. Landers, J.J.; Cao, Z.; Lee, I.; Piehler, L.T.; Myc, P.P.; Myc, A.; Hamouda, T.; Galecki, A.T.; Baker, J.R. Prevention of influenza pneumonitis by sialic acid-conjugated dendritic polymers. J. Infect. Dis. 2002, 186, 1222–1230. [Google Scholar] [CrossRef]
  114. Chen, C.Z.; Cooper, S.L. Interactions between dendrimer biocides and bacterial membranes. Biomaterials 2002, 23, 3359–3368. [Google Scholar] [CrossRef] [PubMed]
  115. Nagahori, N.; Lee, R.T.; Nishimura, S.I.; Pagé, D.; Roy, R.; Lee, Y.C. Inhibition of adhesion of type 1 fimbriated Escherichia coli to highly mannosylated ligands. ChemBioChem 2002, 3, 836–844. [Google Scholar] [CrossRef] [PubMed]
  116. Bernstein, D.I.; Stanberry, L.R.; Sacks, S.; Ayisi, N.K.; Gong, Y.H.; Ireland, J.; Mumper, R.J.; Holan, G.; Matthews, B.; McCarthy, T.; et al. Evaluations of Unformulated and Formulated Dendrimer-Based Microbicide Candidates in Mouse and Guinea Pig Models of Genital Herpes. Antimicrob. Agents Chemother. 2003, 47, 3784–3788. [Google Scholar] [CrossRef] [PubMed]
  117. Blattes, E.; Vercellone, A.; Eutamène, H.; Turrin, C.O.; Théodorou, V.; Majoral, J.P.; Caminade, A.M.; Prandi, J.; Nigou, J.; Puzo, G. Mannodendrimers prevent acute lung inflammation by inhibiting neutrophil recruitment. Proc. Natl. Acad. Sci. USA 2013, 110, 8795–8800. [Google Scholar] [CrossRef] [PubMed]
  118. Sepulveda-Crespo, D.; Serramia, M.J.; Tager, A.M.; Vrbanac, V.; Gomez, R.; De La Mata, F.J.; Jimenez, J.L.; Munoz-Fernandez, M.A. Prevention vaginally of HIV-1 transmission in humanized BLT mice and mode of antiviral action of polyanionic carbosilane dendrimer G2-S16. Nanomedicine 2015, 11, 1299–1308. [Google Scholar] [CrossRef]
  119. Crooke, S.T.; Witztum, J.L.; Bennett, C.F.; Baker, B.F. RNA-Targeted Therapeutics. Cell Metab. 2018, 27, 714–739. [Google Scholar] [CrossRef]
  120. Perez-Carrion, M.D.; Posadas, I.; Cena, V. Nanoparticles and siRNA: A new era in therapeutics? Pharmacol. Res. 2024, 201, 107102. [Google Scholar] [CrossRef]
  121. Ribeil, J.-A.; Hacein-Bey-Abina, S.; Payen, E.; Magnani, A.; Semeraro, M.; Magrin, E.; Caccavelli, L.; Neven, B.; Bourget, P.; El Nemer, W.; et al. Gene Therapy in a Patient with Sickle Cell Disease. N. Engl. J. Med. 2017, 376, 848–855. [Google Scholar] [CrossRef]
  122. Rosenberg, J.B.; Kaminsky, S.M.; Aubourg, P.; Crystal, R.G.; Sondhi, D. Gene therapy for metachromatic leukodystrophy. J. Neurosci. Res. 2016, 94, 1169–1179. [Google Scholar] [CrossRef]
  123. Shahryari, A.; Saghaeian Jazi, M.; Mohammadi, S.; Razavi Nikoo, H.; Nazari, Z.; Hosseini, E.S.; Burtscher, I.; Mowla, S.J.; Lickert, H. Development and Clinical Translation of Approved Gene Therapy Products for Genetic Disorders. Front. Genet. 2019, 10, 868. [Google Scholar] [CrossRef]
  124. Sung, Y.K.; Kim, S.W. Recent advances in the development of gene delivery systems. Biomater. Res. 2019, 23, 8. [Google Scholar] [CrossRef]
  125. Xu, Q.; Wang, C.H.; Pack, D.W. Polymeric carriers for gene delivery: Chitosan and poly(amidoamine) dendrimers. Curr. Pharm. Des. 2010, 16, 2350–2368. [Google Scholar] [CrossRef]
  126. Huang, S.; Li, J.; Han, L.; Liu, S.; Ma, H.; Huang, R.; Jiang, C. Dual targeting effect of Angiopep-2-modified, DNA-loaded nanoparticles for glioma. Biomaterials 2011, 32, 6832–6838. [Google Scholar] [CrossRef]
  127. Chen, L.; Li, J.; Fan, Y.; Qiu, J.; Cao, L.; Laurent, R.; Mignani, S.; Caminade, A.M.; Majoral, J.P.; Shi, X. Revisiting Cationic Phosphorus Dendrimers as a Nonviral Vector for Optimized Gene Delivery Toward Cancer Therapy Applications. Biomacromolecules 2020, 21, 2502–2511. [Google Scholar] [CrossRef]
  128. Rudolph, C.; Lausier, J.; Naundorf, S.; Müller, R.H.; Rosenecker, J. In vivo gene delivery to the lung using polyethylenimine and fractured polyamidoamine dendrimers. J. Gene Med. 2000, 2, 269–278. [Google Scholar] [CrossRef]
  129. Mekuria, S.L.; Li, J.; Song, C.; Gao, Y.; Ouyang, Z.; Shen, M.; Shi, X. Facile Formation of PAMAM Dendrimer Nanoclusters for Enhanced Gene Delivery and Cancer Gene Therapy. ACS Appl. Bio Mater. 2021, 4, 7168–7175. [Google Scholar] [CrossRef]
  130. Ullas, P.T.; Madhusudana, S.N.; Desai, A.; Sagar, B.K.C.; Jayamurugan, G.; Rajesh, Y.B.R.D.; Jayaraman, N. Enhancement of immunogenicity and efficacy of a plasmid DNA rabies vaccine by nanoformulation with a fourth-generation amine-terminated poly(ether imine) dendrimer. Int. J. Nanomed. 2014, 9, 627–634. [Google Scholar] [CrossRef]
  131. Sheikh, M.A.; Malik, Y.S.; Xing, Z.; Guo, Z.; Tian, H.; Zhu, X.; Chen, X. Polylysine-modified polyethylenimine (PEI-PLL) mediated VEGF gene delivery protects dopaminergic neurons in cell culture and in rat models of Parkinson’s Disease (PD). Acta Biomater. 2017, 54, 58–68. [Google Scholar] [CrossRef]
  132. Zhang, D.; Atochina-Vasserman, E.N.; Maurya, D.S.; Huang, N.; Xiao, Q.; Ona, N.; Liu, M.; Shahnawaz, H.; Ni, H.; Kim, K.; et al. One-Component Multifunctional Sequence-Defined Ionizable Amphiphilic Janus Dendrimer Delivery Systems for mRNA. J. Am. Chem. Soc. 2021, 143, 12315–12327. [Google Scholar] [CrossRef]
  133. Yu, T.; Liu, X.; Bolcato-Bellemin, A.L.; Wang, Y.; Liu, C.; Erbacher, P.; Qu, F.; Rocchi, P.; Behr, J.P.; Peng, L. An Amphiphilic Dendrimer for Effective Delivery of Small Interfering RNA and Gene Silencing In Vitro and In Vivo. Angew. Chem. 2012, 124, 8606–8612. [Google Scholar] [CrossRef]
  134. Zhang, D.; Atochina-Vasserman, E.N.; Maurya, D.S.; Liu, M.; Xiao, Q.; Lu, J.; Lauri, G.; Ona, N.; Reagan, E.K.; Ni, H.; et al. Targeted Delivery of mRNA with One-Component Ionizable Amphiphilic Janus Dendrimers. J. Am. Chem. Soc. 2021, 143, 17975–17982. [Google Scholar] [CrossRef] [PubMed]
  135. Lu, J.; Atochina-Vasserman, E.N.; Maurya, D.S.; Sahoo, D.; Ona, N.; Reagan, E.K.; Ni, H.; Weissman, D.; Percec, V. Targeted and Equally Distributed Delivery of mRNA to Organs with Pentaerythritol-Based One-Component Ionizable Amphiphilic Janus Dendrimers. J. Am. Chem. Soc. 2023, 145, 18760–18766. [Google Scholar] [CrossRef] [PubMed]
  136. Navarro, G.; Tros de Ilarduya, C. Activated and non-activated PAMAM dendrimers for gene delivery in vitro and in vivo. Nanomed. Nanotechnol. Biol. Med. 2009, 5, 287–297. [Google Scholar] [CrossRef]
  137. Tariq, I.; Ali, M.Y.; Sohail, M.F.; Amin, M.U.; Ali, S.; Bukhari, N.I.; Raza, A.; Pinnapireddy, S.R.; Schäfer, J.; Bakowsky, U. Lipodendriplexes mediated enhanced gene delivery: A cellular to pre-clinical investigation. Sci. Rep. 2020, 10, 21446. [Google Scholar] [CrossRef] [PubMed]
  138. Pishavar, E.; Attaranzadeh, A.; Alibolandi, M.; Ramezani, M.; Hashemi, M. Modified PAMAM vehicles for effective TRAIL gene delivery to colon adenocarcinoma: In vitro and in vivo evaluation. Artif. Cells Nanomed. Biotechnol. 2018, 46, S503–S513. [Google Scholar] [CrossRef]
  139. Liu, X.; Liu, C.; Laurini, E.; Posocco, P.; Pricl, S.; Qu, F.; Rocchi, P.; Peng, L. Efficient delivery of sticky siRNA and potent gene silencing in a prostate cancer model using a generation 5 triethanolamine-core PAMAM dendrimer. Mol. Pharm. 2012, 9, 470–481. [Google Scholar] [CrossRef]
  140. Liyanage, W.; Wu, T.; Kannan, S.; Kannan, R.M. Dendrimer-siRNA Conjugates for Targeted Intracellular Delivery in Glioblastoma Animal Models. ACS Appl. Mater. Interfaces 2022, 14, 46290–46303. [Google Scholar] [CrossRef]
  141. Karpenko, L.I.; Apartsin, E.K.; Dudko, S.G.; Starostina, E.V.; Kaplina, O.N.; Antonets, D.V.; Volosnikova, E.A.; Zaitsev, B.N.; Bakulina, A.Y.; Venyaminova, A.G.; et al. Cationic polymers for the delivery of the ebola dna vaccine encoding artificial t-cell immunogen. Vaccines 2020, 8, 718. [Google Scholar] [CrossRef] [PubMed]
  142. Serramía, M.J.; Álvarez, S.; Fuentes-Paniagua, E.; Clemente, M.I.; Sánchez-Nieves, J.; Gómez, R.; De La Mata, J.; Muñoz-Fernández, M.Á. In vivo delivery of siRNA to the brain by carbosilane dendrimer. J. Control. Release 2015, 200, 60–70. [Google Scholar] [CrossRef]
  143. Dong, Y.; Chen, Y.; Zhu, D.; Shi, K.; Ma, C.; Zhang, W.; Rocchi, P.; Jiang, L.; Liu, X. Self-assembly of amphiphilic phospholipid peptide dendrimer-based nanovectors for effective delivery of siRNA therapeutics in prostate cancer therapy. J. Control. Release 2020, 322, 416–425. [Google Scholar] [CrossRef] [PubMed]
  144. Dutta, T.; Garg, M.; Jain, N.K. Poly(propyleneimine) dendrimer and dendrosome mediated genetic immunization against hepatitis B. Vaccine 2008, 26, 3389–3394. [Google Scholar] [CrossRef] [PubMed]
  145. Chahal, J.S.; Fang, T.; Woodham, A.W.; Khan, O.F.; Ling, J.; Anderson, D.G.; Ploegh, H.L. An RNA nanoparticle Vaccine against Zika virus elicits antibody and CD8+ T cell responses in a mouse model. Sci. Rep. 2017, 7, 252. [Google Scholar] [CrossRef]
  146. Longmire, M.; Choyke, P.; Kobayashi, H. Dendrimer-Based Contrast Agents for Molecular Imaging. Curr. Top. Med. Chem. 2008, 8, 1180–1186. [Google Scholar] [CrossRef] [PubMed]
  147. Posadas, I.; Romero-Castillo, L.; Ronca, R.A.; Karpus, A.; Mignani, S.; Majoral, J.P.; Munoz-Fernandez, M.; Cena, V. Engineered Neutral Phosphorous Dendrimers Protect Mouse Cortical Neurons and Brain Organoids from Excitotoxic Death. Int. J. Mol. Sci. 2022, 23, 4391. [Google Scholar] [CrossRef] [PubMed]
  148. Li, K.; Wen, S.; Larson, A.C.; Shen, M.; Zhang, Z.; Chen, Q.; Shi, X.; Zhang, G. Multifunctional dendrimer-based nanoparticles for in vivo MR/CT dual-modal molecular imaging of breast cancer. Int. J. Nanomed. 2013, 8, 2589–2600. [Google Scholar] [CrossRef]
  149. Chen, Q.; Wang, H.; Liu, H.; Wen, S.; Peng, C.; Shen, M.; Zhang, G.; Shi, X. Multifunctional Dendrimer-Entrapped Gold Nanoparticles Modified with RGD Peptide for Targeted Computed Tomography/Magnetic Resonance Dual-Modal Imaging of Tumors. Anal. Chem. 2015, 87, 3949–3956. [Google Scholar] [CrossRef]
  150. Filippi, M.; Catanzaro, V.; Patrucco, D.; Botta, M.; Tei, L.; Terreno, E. First in vivo MRI study on theranostic dendrimersomes. J. Control. Release 2017, 248, 45–52. [Google Scholar] [CrossRef]
  151. Chen, Q.; Li, K.; Wen, S.; Liu, H.; Peng, C.; Cai, H.; Shen, M.; Zhang, G.; Shi, X. Targeted CT/MR dual mode imaging of tumors using multifunctional dendrimer-entrapped gold nanoparticles. Biomaterials 2013, 34, 5200–5209. [Google Scholar] [CrossRef]
  152. Wen, S.; Li, K.; Cai, H.; Chen, Q.; Shen, M.; Huang, Y.; Peng, C.; Hou, W.; Zhu, M.; Zhang, G.; et al. Multifunctional dendrimer-entrapped gold nanoparticles for dual mode CT/MR imaging applications. Biomaterials 2013, 34, 1570–1580. [Google Scholar] [CrossRef]
  153. Kojima, C.; Turkbey, B.; Ogawa, M.; Bernardo, M.; Regino, C.A.S.; Bryant, L.H.; Choyke, P.L.; Kono, K.; Kobayashi, H. Dendrimer-based MRI contrast agents: The effects of PEGylation on relaxivity and pharmacokinetics. Nanomed. Nanotechnol. Biol. Med. 2011, 7, 1001–1008. [Google Scholar] [CrossRef]
  154. Huang, C.H.; Nwe, K.; Al Zaki, A.; Brechbiel, M.W.; Tsourkas, A. Biodegradable polydisulfide dendrimer nanoclusters as MRI contrast agents. ACS Nano 2012, 6, 9416–9424. [Google Scholar] [CrossRef]
  155. Yang, J.; Luo, Y.; Xu, Y.; Li, J.; Zhang, Z.; Wang, H.; Shen, M.; Shi, X.; Zhang, G. Conjugation of iron oxide nanoparticles with RGD-modified dendrimers for targeted tumor MR imaging. ACS Appl. Mater. Interfaces 2015, 7, 5420–5428. [Google Scholar] [CrossRef]
  156. Filippi, M.; Remotti, D.; Botta, M.; Terreno, E.; Tei, L. GdDOTAGA(C18)2: An efficient amphiphilic Gd(III) chelate for the preparation of self-assembled high relaxivity MRI nanoprobes. Chem. Commun. 2015, 51, 17455–17458. [Google Scholar] [CrossRef]
  157. Xiong, Z.; Wang, Y.; Zhu, J.; He, Y.; Qu, J.; Effenberg, C.; Xia, J.; Appelhans, D.; Shi, X. Gd-Chelated poly(propylene imine) dendrimers with densely organized maltose shells for enhanced MR imaging applications. Biomater. Sci. 2016, 4, 1622–1629. [Google Scholar] [CrossRef]
  158. Luo, K.; Liu, G.; She, W.; Wang, Q.; Wang, G.; He, B.; Ai, H.; Gong, Q.; Song, B.; Gu, Z. Gadolinium-labeled peptide dendrimers with controlled structures as potential magnetic resonance imaging contrast agents. Biomaterials 2011, 32, 7951–7960. [Google Scholar] [CrossRef]
  159. Lim, J.; Turkbey, B.; Bernardo, M.; Bryant, L.H.; Garzoni, M.; Pavan, G.M.; Nakajima, T.; Choyke, P.L.; Simanek, E.E.; Kobayashi, H. Gadolinium MRI contrast agents based on triazine dendrimers: Relaxivity and in vivo pharmacokinetics. Bioconjug. Chem. 2012, 23, 2291–2299. [Google Scholar] [CrossRef]
  160. Zhang, G.; Du, R.; Qian, J.; Zheng, X.; Tian, X.; Cai, D.; He, J.; Wu, Y.; Huang, W.; Wang, Y.; et al. A tailored nanosheet decorated with a metallized dendrimer for angiography and magnetic resonance imaging-guided combined chemotherapy. Nanoscale 2018, 10, 488–498. [Google Scholar] [CrossRef]
  161. Zhou, X.; Ye, M.; Han, Y.; Tang, J.; Qian, Y.; Hu, H.; Shen, Y. Enhancing MRI of liver metastases with a zwitterionized biodegradable dendritic contrast agent. Biomater. Sci. 2017, 5, 1588–1595. [Google Scholar] [CrossRef] [PubMed]
  162. Gusdon, A.M.; Faraday, N.; Aita, J.S.; Kumar, S.; Mehta, I.; Choi, H.A.; Cleland, J.L.; Robinson, K.; McCullough, L.D.; Ng, D.K.; et al. Dendrimer nanotherapy for severe COVID-19 attenuates inflammation and neurological injury markers and improves outcomes in a phase2a clinical trial. Sci. Transl. Med. 2022, 14, eabo2652. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Scheme of the dendrimer structure. For a detailed explanation, see the text.
Figure 1. Scheme of the dendrimer structure. For a detailed explanation, see the text.
Pharmaceutics 16 00439 g001
Figure 2. Scheme of the divergent and convergent methods for dendrimer synthesis. Symbols are the same as in Figure 1. For a detailed explanation, see the text.
Figure 2. Scheme of the divergent and convergent methods for dendrimer synthesis. Symbols are the same as in Figure 1. For a detailed explanation, see the text.
Pharmaceutics 16 00439 g002
Figure 3. Main strategies for transporting active compounds by dendrimers. For a detailed explanation, see the text.
Figure 3. Main strategies for transporting active compounds by dendrimers. For a detailed explanation, see the text.
Pharmaceutics 16 00439 g003
Figure 4. Main pathways for the intracellular uptake of dendrimers. For a detailed explanation, see the text.
Figure 4. Main pathways for the intracellular uptake of dendrimers. For a detailed explanation, see the text.
Pharmaceutics 16 00439 g004
Figure 5. The targeting methods of dendrimers. For a detailed explanation, see the text.
Figure 5. The targeting methods of dendrimers. For a detailed explanation, see the text.
Pharmaceutics 16 00439 g005
Table 1. In vivo studies of dendrimers as carriers for anti-cancer drugs. Transported active drugs are indicated in an italic font.
Table 1. In vivo studies of dendrimers as carriers for anti-cancer drugs. Transported active drugs are indicated in an italic font.
DendrimerModification/
Active Drug
OutcomeRef.
PAMAMcisplatin
  • selective accumulation in the tumor
  • decreased toxicity in comparison with free cisplatin
[94]
PEG and
5 fluorouracil
  • lack of significant hematological disturbances
[90]
β-cyclodextrin and oligoarginine peptide
entrapped gold nanoparticles and
dexamethasone
  • improved hearing in C57/BL6 mice
  • more effective tympanic injection than posterior ear injection, muscle injection, and tail vein injection
[95]
L-cysteine and doxorubicin
  • L-cysteine acts as a radiosensitizer
  • inhibition of cancer growth
[89]
berberine
  • longer half-life and AUC
[91]
folic acid and borneol
  • tumor growth inhibition
  • median survival time prolonged compared to free doxorubicin
[96]
cholesteryl chloroformate and alkyl-PEG
  • significantly decreased tumor growth rate
[97]
dendrimer–gold hybrid structure
  • effective anti-tumor agent
  • accurate CT imaging
  • usability in combined detection of therapy of different adenocarcinomas via an active, MUC-1-mediated targeting pathway
[98]
PPIPolysorbate 80
  • anti-cancer activity in brain tumor
  • longer survival time of the case of DTX–P80-PPI than DTX–PPI
  • higher targeting efficiency and biodistribution of ligand-conjugated dendrimer into the brain
[92]
phosphorus dendrimersdoxorubicin
(DOX)
  • good intrinsic anticancer activity
  • collective action with DOX to take down breast cancer via the upregulation of Bax, PTEN, and p53 proteins for enhanced cell apoptosis
[87]
Janus
dendrimers
doxorubicin (DOX)
  • minimal systemic toxicity
  • half-life approx. 16 h
  • 9 times higher tumor uptake in comparison with free DOX
  • complete tumor regression
  • 100% survival of the mice
  • antitumor effect of dendrimer–DOX similar to liposomal DOX
[99]
cytarabine and daunorubicin
  • superior therapeutic activity compared to free drug cocktails
  • prolonged maintenance of synergistic drug ratios in the bone marrow
[100]
peptide
dendrimers
doxorubicin
  • increased doxorubicin concentration in the tumor
  • enhanced anticancer efficacy of doxorubicin and gemcitabine
[88]
polylysine dendrimers polyoxazoline
  • prolonged circulation of the dendrimer
  • survival period of the mice extended beyond 70 days following the final dose
[93]
rotaxane dendrimerchlorambucil
  • accumulation in the reticuloendothelial system
  • enriching spleen and liver
[16]
Table 2. In vivo studies of dendrimers as carriers for drugs in non-cancer diseases. Transported active drugs are indicated in an italic font.
Table 2. In vivo studies of dendrimers as carriers for drugs in non-cancer diseases. Transported active drugs are indicated in an italic font.
DendrimerModification/
Active Drug
OutcomeRef.
Phosphorus
dendrimer
azabisphosphonate (ABP)
  • efficient anti-inflammatory activity
[76]
azabisphosphonate (ABP)
  • anti-inflammatory activity
  • selective targeting to monocytes
  • anti-osteoclastic activity
[102]
amino-bis(methylene phosphonate)
  • preventing of autoimmune encephalomyelitis development
  • inhibition of established disease progression
  • redirection of pathogenic myelin-specific CD4+ T cell response toward IL-10 production
[104]
amino-bis(methylene phosphonate)
  • inhibition of the onset and development of experimental arthritis
  • lack of adverse response
  • lack of lesion or non-physiological occurrence
[105]
carteolol
  • no eye irritation
  • 2.5 times larger quantity of mix of carteolol and dendrimer in comparison with free carteolol
[106]
Janus
dendrimers
carboxybetain and
α-lactalbumin
  • prolonged pharmacokinetic profiles of payloads in comparison with the PEGylated nanocarriers
[107]
PAMAMamino-terminated
  • improved diabetes-induced vascular remodeling and dysfunction
  • inhibited EGFR-ERK1/2-ROCK signaling
[108]
amino-terminated
  • intraperitoneal and subcutaneous administration are the most effective in suppressing the long-term markers of hyperglycemia
  • the lowest incidence of side effects was observed after subcutaneous administration
  • subcutaneous injection is the best way to compromise moderate PAMAM toxicity and effective reduction in the markers of long-term severe hyperglycemia
[109]
  • reduced levels of blood glucose glycated hemoglobin or protein oxidation, cholesterol, and triglycerides
  • higher terminal blood glucose in PAMAM-treated animals
amino-terminated
  • reduction of blood glucose concentration and hallmarks of late diabetic complications
  • reduced diabetes-induced permeabilization of BBB
  • enrichment of the biochemical hallmarks of severe hyperglycemia.
[110]
simvastatin
  • reduction in the increase in cholesterol level approx. 2 times and in triglyceride and low-density lipoprotein levels
  • better pharmacokinetic performance in comparison with free drug
  • three-to-five times higher residence time in comparison with free drug
  • controlled release of simvastatin decreasing absorption and elimination rates
[101]
Lysine
dendrimers
mannose
  • accelerated diabetes wound repair after topical administration
  • increased closure rate, collagen deposition, and angiogenesis, production of TGF-β1
[111]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sztandera, K.; Rodríguez-García, J.L.; Ceña, V. In Vivo Applications of Dendrimers: A Step toward the Future of Nanoparticle-Mediated Therapeutics. Pharmaceutics 2024, 16, 439. https://doi.org/10.3390/pharmaceutics16040439

AMA Style

Sztandera K, Rodríguez-García JL, Ceña V. In Vivo Applications of Dendrimers: A Step toward the Future of Nanoparticle-Mediated Therapeutics. Pharmaceutics. 2024; 16(4):439. https://doi.org/10.3390/pharmaceutics16040439

Chicago/Turabian Style

Sztandera, Krzysztof, José Luis Rodríguez-García, and Valentín Ceña. 2024. "In Vivo Applications of Dendrimers: A Step toward the Future of Nanoparticle-Mediated Therapeutics" Pharmaceutics 16, no. 4: 439. https://doi.org/10.3390/pharmaceutics16040439

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