Skip to main content
Medical Review logoLink to Medical Review
. 2024 Feb 20;4(1):5–30. doi: 10.1515/mr-2023-0058

Nanotechnology-based delivery systems to overcome drug resistance in cancer

Harsh Patel 1, Jiaxin Li 1,2, Letao Bo 1, Riddhi Mehta 3, Charles R Ashby Jr 1, Shanzhi Wang 1, Wei Cai 2,, Zhe-Sheng Chen 1,
PMCID: PMC10954245  PMID: 38515777

Abstract

Cancer nanomedicine is defined as the application of nanotechnology and nanomaterials for the formulation of cancer therapeutics that can overcome the impediments and restrictions of traditional chemotherapeutics. Multidrug resistance (MDR) in cancer cells can be defined as a decrease or abrogation in the efficacy of anticancer drugs that have different molecular structures and mechanisms of action and is one of the primary causes of therapeutic failure. There have been successes in the development of cancer nanomedicine to overcome MDR; however, relatively few of these formulations have been approved by the United States Food and Drug Administration for the treatment of cancer. This is primarily due to the paucity of knowledge about nanotechnology and the fundamental biology of cancer cells. Here, we discuss the advances, types of nanomedicines, and the challenges regarding the translation of in vitro to in vivo results and their relevance to effective therapies.

Keywords: nanotechnology, multidrug resistance, drug delivery systems, nanoformulations, enhanced permeability and retention effect

Introduction

“There’s plenty of room at the bottom”, a statement by Dr. Richard P. Feynman in 1959, revealed a new range of possibilities in almost all the realms of science [1]. Prof. Norio Taniguchi, in 1974, coined the term, “nanotechnology”, which is now defined as an area of technology that involves dimensions less than 100 nm, with a primary emphasis on the manipulation of individual atoms and molecules [2]. Nanomedicine involves the combination of nanotechnology with pharmaceutical and medical sciences [34]. The major aims of nanomedicine are the development of drugs with higher efficacies, lower toxicities, and the efficacy to overcome multidrug resistance (MDR) [5]. Nanoparticles (NPs) have high surface-to-volume ratios compared to bulk materials and have varying physical, chemical, and biological properties [5], [6], [7], [8], [9], [10], [11]. By definition, nanomedicines are drugs or biologics that incorporate NPs (usually <100 nm) to a greater magnitude than their bulk counterparts in vivo [12]. Nanomedicines are primarily used to treat cancerous tumors because these NP-drug conjugates act by passive targeting, characterized by their significant accumulation in the tumor [13]. NPs have been reported to have increased permeability through blood vessels and lower lymphatic drainage. This property is known as the enhanced permeability and retention (EPR) effect, and it ultimately causes significant accumulation of the drug in the tumor microenvironment (TME) [51112].

MDR, which causes a decrease or abolition of the efficacy of anticancer drugs that differ in their structure and mechanism of action in cancer cells, plays a major role in treatment failure, thereby increasing the risk of relapse and mortality [14]. Thus, novel drugs and techniques must be developed to surmount MDR in cancer. The delivery of anticancer drugs can be achieved by various routes of administration. The choice of the route of administration by a clinician depends on various factors, including but not limited to, the drug, type of cancer, efficacy, location of the tumor in the body, and toxicity of the drug, among others [15], [16], [17], [18]. Furthermore, certain anticancer drugs can be administered directly into the tumor [19]. Over the last 3 decades, there have been numerous innovations and developments in cancer drug delivery systems [162021]. Furthermore, the majority of these developments have some specific applications in drug delivery, such as: (1) anti-angiogenic drugs, which decrease angiogenesis in tumors; (2) biological therapies, such as gene therapy and viral oncolysis; (3) combinations of novel therapies with conventional chemotherapy and radiation therapy; (4) development of immune therapy, e.g., cancer vaccines, cytokine modulators, monoclonal antibodies and (5) nanobiotechnology for cancer therapy (targeted drug delivery systems) [16]. Chemotherapy remains the most commonly used systemic treatment to inhibit the growth and proliferation of cancer cells, progression of the disease, and metastasis [2223].

Despite the significant increase in research in nanobiotechnology and cancer nanomedicine, there is still a large disparity between scientific advancements and the use of these applications in the treatment of cancer patients. Figure 1 provides examples of advances in nanotechnology, with an emphasis on cancer nanomedicine. Nanomedicine has provided the concept of theranostics, defined as the simultaneous diagnosis and treatment of a disease [2425]. Furthermore, researchers have developed targeted drug delivery, which produces a lower frequency of adverse effects, greater efficacy, decreased immunogenicity, and a decrease in surgical intervention [2126]. Nanoformulations have a number of advantages compared to the non-nanoformulated parent drug, such as carrying a large amount and multiple drugs, maintaining therapeutic drug concentrations for a longer time, and permeating into cells by endocytosis, which bypasses certain resistance mechanisms [4]. However, numerous obstacles make it difficult to develop nanoformulations that can be used in humans, including biological barriers, safety profiling, and scaling-up processes [27].

Figure 1:

Figure 1:

Timeline of the history of the developments in nanotechnology in the field of medicine. The figure was adapted and revised from Salvioni et al. [14], Li et al. [17], and Shi et al. [28].

In this review, we will discuss the types of nanoformulations and the challenges in drug delivery and clinical translation. We will categorize and discuss the nanoformulations as metallic NPs, NPs from natural products, albumin NPs, polymeric micelles, liposomes, and polyphenolic compounds.

Challenges in drug delivery and the clinical translation of nanomedicine (Figure 2)

Figure 2:

Figure 2:

Rationale for developing cancer nanomedicines.

Despite the availability of non-invasive drug delivery platforms, most cancer nanomedicines use the intravenous route of administration for the transport of drug to the tumors [2930]. The infrastructure of NP delivery to solid tumors is based on the EPR effect [51112]; however, there are significant differences in the EPR effect between patients and the types of tumors. For example, tumor microenvironment (TME) properties, including extravasation, diffusivity, hemodynamic regulation, heterogeneity, etc., can have adverse effects, such as uncertainty of the concentration of NPs in the tumor vasculature [13]. Furthermore, lymphatic drainage is non-uniform throughout the tumor mass and there is a greater mechanical stress and a higher degree of functional loss in the bulkier regions of tumor, compared to the margins [31]. This factor must be considered in combination with other factors, such as extravasation and diffusivity, which affect the equilibrium of the NPs that accumulate inside the tumor mass [32]. Furthermore, tumor-associated factors affecting drug delivery depends on the type of material, its compatibility with tissue components, and the architecture of the tumor itself [16]. It has been reported NP accumulation increases in the interstitium of the tumor in the presence of a high number of phagocytic and dendritic cells [33]. Tumor-associated macrophages (TAMs) have been shown to significantly accumulate and retain nanoformulations and they can be gradually released in close proximity to tumor cells [34], [35], [36], [37].

In order to increase drug delivery to a tumor, it is critical to determine the interaction between the NPs and surface proteins. A corona-like structure is formed when a NP enters an organic environment, such as interstitial fluid, the extracellular matrix or blood, and the surface of the NP is rapidly enveloped by lipoproteins or cellular receptors [538], [39], [40], [41]. This produces a series of changes in the structure, size, stability, and surface properties of the NPs [39]. In addition to the changes in the physical properties of the NPs, the adsorption will produce a biological identity for the NP, which will determine its cellular uptake, intracellular trafficking, toxicity, and pharmacokinetic (PK) profile [3942], [43], [44], [45]. All of these changes will increase the efficacy of anticancer drug. Nanoparticle formulations of doxorubicin (DOX) in humans activate the complement system, producing a hypersensitivity reaction (non-immunoglobulin E (IgE) mediated) [46]. NP-protein interactions are dependent on the physicochemical properties of the NP, the source and concentration of the protein and the exposure time [47]. Studies have been conducted to predict the interaction of NPs with cells. Bigdeli et al. [48] identified a range of protein corona targets that could potentially improve the design of liposomes. Hajipour et al. [49] concluded that the type of the disease plays a critical role in determining the architecture of the corona. Walkey et al. [50] used qualitative structure-activity relationship (QSAR) to determine the interactions of NPs with biological components. A group of hyaluronan-binding proteins were identified in this study as mediators of interactions between the NPs and cells. The goal of this study was to establish a platform for building an extensive database of protein corona signatures and its biological responses for various types of NPs. Unfortunately, the majority of the studies were focused on NP-protein interactions in vitro. Moreover, in vivo corona formation and its correlation with pharmacokinetic parameters, biodistribution, and efficacy, have not been fully delineated. However, Sakulkhu et al. [51] investigated the in vivo protein corona interaction by using the distinct magnetic properties of supermagnetic iron oxide nanoparticles (SPIONs). They extracted NPs from rat sera after their interaction with the physiological system of rat and found differences between the types of corona formation in vitro and in vivo, such as the size of the corona, the family of proteins, surface charge and its biodistribution [51]. Similar future studies will hopefully benefit and strengthen the advancements in cancer nanomedicine.

It is well known that circulation affects the half-life of the drug. Drug molecules that need to reach poorly perfused tissues require a longer half-life than drugs that are distributed to tissues with a larger blood flow [52]. The opsonization of NPs, which causes their phagocytosis by the mononuclear phagocytic system (MPS), results from non-specific interactions between NP and certain serum proteins [52]. Therefore, the NP composition needs to be considered to decrease or avoid recognition by the immune system. For example, to prolong the time that NPs remain in circulation, polyethylene glycol (PEG) grafting is commonly used because PEGylated carriers delay absorption by the liver and the spleen [5354]. Alternatively, nanoformulations can be modified by adding molecules that will prevent the immune system from recognizing them as foreign antigens, such as the addition of the cluster of differentiation 47 (CD47) peptide (which decreases recognition by the MPS) [55] or by camouflaging the NP surface with a membrane derived from either red blood cells (RBCs), white blood cells (WBCs) or platelets [5657]. Factors, including abnormal and uneven tumor vasculature and perivascular TME, decrease extravasation of NPs from the systemic circulation to tumors [58]. This complication can be addressed by loading the nanoformulated drug into mesenchymal cells, macrophages and monocytes or by attaching NP to the membrane of these cells [58], [59], [60], [61].

It is important to note that the size and binding affinity of the nanocarrier affects the penetration of a drug into a tumor [62]. It has been reported that higher affinity antibodies that bind to the target antigens on cancer cells, are less likely to efficiently penetrate into cancer cells, compared to lower affinity antibodies, due to internalization [62]. NPs are usually greater in size than antibodies and consequently, they have a greater probability of being trapped in the extracellular matrix of the cancer cells [63]. It has been hypothesized that the modulation of the size of the NPs may be a solution to this problem. Smaller NPs can diffuse throughout the tumor tissue but particles less than 5 nm in diameter are readily removed by renal filtration [64], [65], [66]. An in vivo study reported that 15 × 54 nm nanorods more rapidly penetrated into orthotopic mammary tumors, compared to nanospheres that were 35 nm in diameter [67]. Tasciotti et al. [68] developed a multistage drug delivery system, using mesoporous silica particles 3.5 µm in size as an outer shell and a 20–30 nm NP-drug conjugate as the inner core. This formulation was incubated with human umbilical vein endothelial cells and the inner core NP were released and internalized (in distinct vesicles) [68]. Another research group developed a multistage delivery system, consisting of an outer layer of 100 nm NPs enclosing a 10 nm drug conjugated quantum dot (QD) [65]. These multistage quantum dot gelatin nanoparticles (QDGelNPs) were composed of a gelatin core with amino-PEG QDs conjugated to the surface, using 1-ethyl-3-[3-dimethylaminopropyl] carbodiimide hydrochloride/sulfo-N-hydroxysulfosuccinimide (EDC/sulfo-NHS) coupling chemistry. The 100 nm shell was degraded enzymatically and the 10 nm core NP was released deep into the dense collagen matrix of tumor. This formulation produced an effective delivery of drug to solid tumors.

Because most of the NPs are formulated to reach intracellular targets, their effective cellular uptake and internalization are required to produce efficacy. This can be done by placing targeted ligands on the NPs that will interact with specific receptors on the cancer cells, increasing the probability of their internalization into the cancer cells [69]. For example, the United States Food and Drug Administration (FDA)-approved drug, Abraxane®, utilizes nanoparticle albumin bound (nab) technology, and drug delivery was due to GP-60 receptor-mediated transcytosis through microvessel endothelial cells in the angiogenic tumor vasculature [70]. The uptake and internalization of Abraxane® in metastatic breast cancer cells was significantly greater than that of taxol or taxotere. This approach is critical for drug delivery via the intestinal mucosa and blood-brain barrier [71], [72], [73]. In addition to nab technology, it should be noted that gold nanoparticles (AuNPs) have the potential to be further modified to further increase their uptake and internalization by cancer cells [7475].

NP endosomal release is very important for the cytosolic delivery of the drug payload, particularly drugs that are small-interfering ribonucleic acid (siRNA)-based therapeutics, as these drugs must penetrate into cells to interact with the tumor mRNA to produce their therapeutic efficacy [7677]. A considerable number of siRNA-based therapeutics only have an endosomal release of 1 %–2 %, although formulations of this type have been successful in certain clinical trials [76]. The controlled release of a drug in the circulation is important in producing a constant level of drug in the blood, as it helps maintain the dosage level and dosing frequency, which can improve patient compliance. Furthermore, the controlled release of the drug decreases non-target accumulation, thus decreasing the probability of toxic effects [20]. For the controlled release of NPs, the Cmax should occur during the the infusion period; however, the concentration of the released drug will initially be lower than the Cmax. Subsequently, the Cmax of the free drug administered intravenously will be significantly greater than the Cmax of the free drug administered via NP delivery [78]. This lower release of the drug will significantly decrease the probability of toxicity. Furthermore, the EPR produces a differential accumulation of nanoformulations in the tumor to a greater magnitude than free or conventional drugs [79]. The same conventional drug, when released from the nanoformulation, typically accumulates to a greater extent in the tumor over a prolonged period of time [12]. Following internalization, NPs either release the loaded drug outside the cell or it is directed by the intracellular trafficking pathways and the drug is released inside the target cell [80]. Endosomal avoidance is very important for the cytosolic delivery of biomacromolecules, such as siRNA. The delivery of siRNA can be further increased by polymer-based NPs, cationic lipid, and lipid-like materials [778081]. The majority of RNA-interference (RNAi) therapeutics is formulated in liposomes and the targeted delivery of NPs delivery can further increase the cellular uptake of siRNA molecules [82]. However, only a small fraction of the siRNA is released from cellular endosomes [76] and approximately 70 % of the internalized siRNA may undergo exocytosis by Niemann-Pick type C1 protein [83]. Thus, alternative approaches will be required to produce NP formulations that avoid sequestration by endosomes. In addition to cytosol delivery, targeting mitochondria, the Golgi apparatus, nucleus, and the endoplasmic reticulum, have been utilized [84]. However, further studies must be conducted to characterize the delivery of NPs through the membranes of organelles [84], [85], [86], [87], [88], [89].

Conventional formulation techniques that are used to prepare NPs usually produce a greater magnitude of particle size heterogeneity in the mixture, compared to nanoformulations [6790]. Microfluidic technologies have more rapid self-assembly characteristics, with a significantly higher homogeneity and reproducibility, along with manageable physical and chemical properties [91], [92], [93], [94], [95], compared to sol-gel processes, molecular condensation or chemical reduction. Current approaches allow a greater control over the chemical composition, shape, size, drug loading, and surface properties [96]. However, the translation of this level of control over a larger scale of production will require the fulfillment of good manufacturing practice (GMP). The increase in complexity of the nanoformulations will also require advancements in chemistry, manufacturing, controls, and GMP challenges [28]. Advances in the scaling-up process may help in expediting the translation of these formulations from laboratory (grams) to commercial (kilograms) amounts [9798]. Although preclinical data suggests that NPs can increase drug delivery, a significant percentage of those studies were not translated into clinical trials. Furthermore, a study has shown that out of 94 % of successful Phase I trials, only 14 % proceeded to Phase III trials with favorable therapeutic results [99].

The in vitro assessment of NP formulations is essential to determine the biocompatibility with the targets and candidates before progressing to in vivo studies. Despite generating a large amount of information, in vitro studies lack the complexities of biological tissues and may not have the complex interactions of NPs with the physiological barriers [100]. Consequently, further developments are needed in the field of biomimetic devices [100], [101], [102]. However, animal models continue to provide critical information regarding the pharmacokinetics, efficacy, biodistribution, and toxicity, of the drugs being tested. It is imperative to note that PK scaling varies widely across species for different nanoformulations [103], [104], [105], [106]. Furthermore, human cancers are not always recapitulated in vivo and a single model cannot possibly contain all the biological characteristics of the disease. For example, EPR is more erratic in cancer patients than in animal models [13].

The TME play a significant role in cancer progression and metastasis and thus, it has become a target for cancer treatment [107], [108], [109]. Certain in vivo studies in mice models have shown that blocking angiogenesis causes tumor suppression and a decrease in metastasis [110111]. The targeting of TAMs and fibroblasts can be utilized for cancer treatment. Cellax (docetaxel-conjugate nanoparticles)-treated mice had a significant decrease in α-smooth muscle actin (α-SMA)-expressing fibroblasts, which, in turn, significantly decreased the tumor extracellular matrix (ECM) and interstitial fluid pressure (IFP), increased the vascular permeability and decreased the metastasis of breast cancer [112]. Thus, there is a direct relationship between ECM and IFP and drug penetration and subsequently, the metastatic potential of cancer cells.

Various classes of nanomedicines

Metallic NPs

Nanotechnology is used in medical and pharmaceutical applications, e.g., the delivery of biological materials, vaccines, and drugs [113]. NPs, particularly metallic NPs, have unique electrical, optical, magnetic, catalytic, and favorable biological characteristics [114]. Metallic NPs are a type of inorganic nanomaterials that are composed of titanium, silver, gold, ruthenium, zinc, selenium, iron, copper, gadolinium or hafnium, have been used for cancer treatment [115]. The characteristics and advantages of using metallic nanomaterials will be discussed in this section.

Because of its applications in many areas, metallic NPs, such as titanium dioxide (TiO2NPs), silver (AgNPs), gold (AuNPs), and ruthenium (ruthenium nanoparticles, RuNPs), represent useful and versatile molecules [116117]. TiO2NPs have a high capacity for the absorption of short-wavelength light, which has been widely utilized in various cosmetics and sunscreens [118]. Due to the innate antioxidant and antimicrobial efficacies of silver (which is slowly released from AgNPs as Ag+ ions), AgNPs have been used as biocidal compounds, treatment of multidrug resistant microbes, preservatives for food packing, and decontamination of water [119120]. Indeed, AgNPs have been used to inhibit microorganism proliferation, as they have broad antibacterial [114], antifungal, and antiviral efficacy [121]. Finally, there are in vitro data indicating that AgNPs induced the death of breast and colorectal cancer cells [122].

AuNPs have received increasing interest because of their stability, biocompatibility, tunable surface, facile synthesis, optical properties, and ease of chemical modifications, as AuNPs can be attached to hundreds of different molecules [116123]. AuNPs have been shown to be useful as diagnostic and therapeutic molecules [124]. Furthermore, AuNPs can be synthesized as solid spheres, nanospheres, nanorods, nanocages, nanoclusters, and nanostars, among others, thereby making AuNPs well-suited for various biomedical purposes (Figure 3) [125]. Consequently, AuNPs have the potential to be used for the diagnosis, treatment, and prevention of certain types of cancer [126]. AuNPs can be used for the early diagnosis of a thrombus by detecting biomarkers for direct imaging [116]. Recently, clinical trials have been conducted with AuNPs. In 2021, the clinical trial (NCT04907422) proposed a novel diagnostic and prognostic approach for the prompt and timely discovery of cancer stem cells in salivary gland tumors, using AuNPs conjugated to CD24. CD24 is used as a conjugate as it regulates the epithelial-mesenchymal transition in cancer cells [127]. The validation in this trial was done by real-time quantitative polymerase chain reaction (RT-qPCR). AuNPs have been reported to increase the efficiency of proton therapy, as the insertion of AuNPs increases the absorbed radiation dose in tumors [127]. A novel nucleotide-carrying AuNP formulation, known as NU-0129, is undergoing early Phase I clinical trials for patients with relapsing glioblastoma or gliosarcoma [128]. DOX tethered with AuNPs were synthesized for drug release into the acidic microenvironment of the cells, resulting in a reversal of adenosine triphosphate (ATP)-binding cassette (ABC) transporter activity and higher nuclear localization in MCF-7/ADR cells [129]. This represents an excellent example of mitigating MDR in cancer using nanomedicine. Luo et al. developed a prostate cancer (PCa)-targeted gold nanocluster radiosensitizer, coupled with monomethyl auristatin E (a potent cytotoxin), for radiotherapy and chemotherapy. This approach produced a greater nanocluster uptake by prostate-specific membrane antigen (PSMA)-positive cancer cells and it produced an increased efficacy of radiotherapy in vitro and in vivo [130]. RuNPs have been investigated as antibacterial molecules, catalysts, and pharmaceuticals, due to their biotoxicity, high surface-to-volume ratio, large optical properties, and high photothermal conversion rate [131]. Similar to AuNPs, RuNPs also have good biocompatibility [132133]. RuNPs target cancer by DNA damage-induced apoptosis, topoisomerase-II inhibition, inhibition of the mitogen-activated protein kinase (MAPK) signaling pathway, activation of the p53-dependent caspase-3 mediated signaling, upregulation of adenomatous polyposis coli (APC) and p53 genes, inhibition of proteasome activity, p53-independent activity, inhibition of the hypoxia-inducible factor-1 (HIF-1) pathway, anti-metastasis activity and the induction of dysfunction in lysosomal activity [134]. RuNPs are being developed as a nanoparticle drug delivery system for the photothermal treatment of cancer, as they have efficient near-infrared (NIR) efficacy in colorectal HIEC-6, Caco-2, SW480, HCT116, and CT26 WT cancer cells [132133]. Li et al. [135] reported that RuZ (ruthenium [II] complex), at a maximum concentration of 30 mg/mL, self-assembles in the cell culture media and had high penetration in SW620/Ad300, KB-C2, KB-ATO, H460/MX20, BEL-7404/CP20, BIU-87/DDP, and MDA-MB-231/Adr MDR cancer cell lines, with half maximal inhibitory concentration (IC50) values of 1.75, 2.64, 1.58, 0.70, 1.34, 1.26 and 1.96 μmol/L, respectively [135]. Lakshmi et al. [136] synthesized ruthenium (II)-curcumin liposome NPs and it produced cytotoxicity and nuclear damage in HeLa cells [IC50 value=99 μg/mL].

Figure 3:

Figure 3:

Various classes of cancer nanomedicines.

Gold nanostars are synthesized from HAuCl4 as precursors and hydroxylamine as a reductant above pH 11 [137]. Gold nanorods and nanospheres are some of the simplest nano-preparations in metallic NPs, with sizes of 60 nm and 20–40 nm, respectively [138]. Gold nanocages and nanoclusters are approximately 7–20 nm in size and they have positive and negative charges [139140]. Silver nanocrystals range in size from 1 to 100 nm and are used for diagnosis, drug delivery, treatment, and personal health care [121]. Liposomal ruthenium (II)-curcumin nanoformulation is a hybrid preparation that was developed to overcome the hydrophobicity of curcumin [136].

There are several physical and chemical methods that can be used to synthesize and stabilize metallic NPs [120141]. Currently, environmentally friendly methods (known as green chemistry) of nanoparticle synthesis have received increasing interest [131142], [143], [144], [145], [146], [147], [148]. In physical processes, metallic NPs are produced using evaporation/condensation techniques [149]. Conventionally, certain NPs materials, such as silver (Ag), gold (Au) and lead (II) sulfide (PbS), were synthesized using a tube furnace at atmospheric pressure [150]. Chemical reduction, using organic and inorganic reducing compounds, such as sodium citrate, ascorbate, and the Tollens reagent, is the most widely used chemical method to synthesize metallic NPs [151]. Since conventional methods of NP production involves using chemical compounds that produce environmental toxicity, green syntheses have been widely studied, which used eco-friendly and biocompatible procedures [151152]. The green-synthesis of TiO2NPs can be accomplished using extracts of clove (Syzygium aromaticum), black pepper (Piper nigrum), and coriander (Coriandrum sativum) [142]. AuNPs and AgNPs were first synthesized using a leaf extract from Perilla frutescens [141]. Bark [143153] and plant water extracts [146] may be a good source of reducing compounds for the synthesis of metallic NPs. Biological green synthesis has been reported using fungi, bacteria, and other microorganisms [144]. The biological AuNPs (conjugated with proteins, lipids, DNA or antibodies), and AgNPs (Trichoderma spp.-AgNP-T or Sclerotinia sp.-AgNP-S), which had antifungal efficacy in pathogenic fungi, were synthesized using the fungus, Trichoderma longibrachiatum [145148].

Metallic NPs have been of significant interest to researchers due to their extensive medical applications, such as targeted delivery, enhancement of anticancer efficacy, and increased contrast for imaging tumors [150154155]. Therapeutically, metallic NPs are considered a useful treatment due to their multifunctional physical, optical, and magnetic properties, which results in increased penetration and detection in the body [156]. Numerous studies have reported that metallic NPs have efficacy in liver, breast, colon, prostate, and human leukemic monocyte cancer cells [157], [158], [159], [160]. Metallic NPs are usually coupled with a targeting compound or molecule and are loaded with chemotherapeutic drug to increase therapeutic efficacy [161]. AuNPs have been utilized to a greater extent than other NPs because they are relatively easy to synthesize and have a tolerable safety profile [162]. AuNPs capped with bovine serum albumin (BSA) are effective carriers of methotrexate (MTX, an anticancer drug) to MCF-7 breast cancer cells, thereby increasing the drug efficacy, compared to conventional formulations [163]. Similarly, Majoumouo et al. reported that AuNPs had efficacy in MCF-7 and Caco-2 cancer cells [147]. NPs can penetrate the blood-brain barrier (BBB) due to their diminutive size, and surface properties, such as high interactivity and plasmon resonance [164], [165], [166]. AuNPs were developed as an alternative delivery platform to the central nervous system with modified ligands, using targeting moieties, such as transferrin (Tf) [167]. As nanocarriers, AuNPs provide a platform to attach biomolecules, such as oligonucleotides, proteins and peptides, and exosomes [154]. Zhang et al. [168] generated exosomes combined with AuNPs, which was used as the vehicle for the chemotherapeutic drug, DOX, for the treatment of melanoma. They used C57BL/6 mice and used the B16F10 tumor-bearing model to determine the in vivo biodistribution of EVdox@AuNP at a dose of 2.5 mg/kg via intravenous infusion. The results indicated that the AuNP formulation accumulated to a greater extent in the tumors, compared to unconjugated AuNPs. Following drug accumulation in the tumors, a laser beam was used to irradiate the tumors for 5 min. There was a slight increase in temperature (to 38.1 °C) in areas without NP accumulation, however, the areas that accumulated AuNPs had a rapid increase in temperature (to 46.1 °C) in just 1 min. These results suggest that the EVdox@AuNPs photothermal transformation, internalization, and retention in the tumor region [168]. Chen et al. [169] used DOX to prepare a thermo-responsive drug release formulation to determine its efficacy in KB-3-1 cancer cells. Their results indicated that the novel fragmented polymer nanotubes had favorable biocompatibility and were cytotoxic in KB-3-1 cancer cells (IC50=1.4 μmol/L). These specific types of polymer nanotubes have the potential to be used as efficient delivery systems for anti-cancer drugs because of their thermo-responsive gating system [169].

Metallic NPs have been evaluated as diagnostic and therapeutic molecules. Currently, however, only a few technologies based on metallic NPs have been approved by the FDA for diagnostic and therapeutic use [15]. For example, the nanodrug, Aurimune (CYT-6091), was produced by linking AuNPs to recombinant human tumor necrosis factor-α (rhTNF-α) and PEG [170]. The results of a Phase I clinical trial indicated that CYT-6091 was significantly more efficacious than rhTNF-α in patients with advanced solid tumors, compared to conventionally formulated treatment. Patents based on NPs, can be used to develop more precisely targeted therapeutics in support of the respective inventions. Chauhan et al. disclosed their invention (Patent No. US10456363B2), which is composed of modified cyclodextrin-coated magnetite NPs as a targeted nanocarrier for hydrophobic drugs [171]. The patent (WO 2013/176468 Al) reported the development of a liver-targeted drug delivery systems, using AuNPs [172]. Another patent (WO 2014/047318) for drug delivery, invented by Kaittanis et al. [15] is based on iron oxide NPs [173]. In breast cancer patients, superparamagnetic iron oxide nanoparticles (SPION) accumulate in the sentinel lymph nodes (SLN). In a Phase I clinical trial (NCT05359783), SPION was used as neoadjuvant therapy and a biopsy was done by minimal invasion to the tissue, resulting in a de-escalated, less complex surgery [174]. Bort et al. showed that the activation and guiding of irradiation by X-ray (AGuIX® NP) had improved accumulation and retention in tumors via an EPR effect [175]. A novel nanoradiosensitizer for the radiotherapy of cancers, NBTXR3, is a hafnium-based theranostic tool [176]. Additional research with metallic NPs should provide important data about the properties of metallic NPs that can be studied to provide efficacious cancer treatments (Table 1).

Table 1:

Various novel nanoparticles in clinical trials.

NCT identifiers and phase Topics Drugs Formulations References
NCT04789486 (Phase I/II) Nano-SMART: nanoparticles with MR guided SBRT in centrally located lung tumors and pancreatic cancer AGuIX Metallic NPs [175]
NCT04881032 (Phase I/II) AGuIX nanoparticles with radiotherapy plus concomitant temozolomide in the treatment of newly diagnosed glioblastoma (NANO-GBM) AGuIX Metallic NPs
NCT04899908 (Phase II) Stereotactic brain-directed radiation with or without aguix gadolinium-based nanoparticles in brain metastases AGuIX Metallic NPs [175, 177]
NCT05039632 (Phase I/II) Phase I/II randomized study of NBTXR3 activated by abscopal or RadScopal radiation in combination with immunotherapy (anti-PD-1/L-1) for patients with advanced solid malignancies Hafnium oxide-containing nanoparticles NBTXR3 Metallic NPs
NCT04505267 (Phase I) NBTXR3 and radiation therapy for the treatment of inoperable recurrent non-small cell lung cancer Hafnium oxide-containing nanoparticles NBTXR3 Metallic NPs [176]
NCT05359783 (Phase I/II) Sentinel node localization and staging with low dose superparamagnetic iron oxide (MAGSNOW) Superparamagnetic iron oxide Metallic NPs [174]
NCT03020017 (Phase I) NU-0129 in treating patients with recurrent glioblastoma or gliosarcoma undergoing surgery NU-019 SNA consists of nucleic acids arranged on the surface of a small spherical gold nanoparticle Metallic NPs [128]
NCT02837094 (Phase I) Enhanced epidermal antigen specific immunotherapy trial-1 (EE-ASI-1) C19-A3 GNP Metallic NPs
NCT05359783 (Phase II) Sentinel node localization and staging with low dose superparamagnetic iron oxide Superparamagnetic iron oxide Metallic NPs [174]
NCT05456022 (Phase II) Therapeutic efficacy of quercetin vs. its encapsulated nanoparticle on tongue squamous cell carcinoma cell line Quercetin-encapsulated PLGA-PEG nanoparticles (Nano-QUT) Polymeric micelles [178]
NCT04640480 (Phase I) Dose-finding study to evaluate the safety, tolerability, and pharmacokinetics of SNB-101 (SN-38) in patients with tumors SNB-101 Polymeric micelles [179]
NCT05893888 (Phase I/II) Safety and efficacy study of PRV211 in subjects with oral squamous cell carcinoma PRV211 (intraoperative cisplatin system) Polymeric micelles
NCT04751786 (Phase I) Dose escalation study of immunomodulatory nanoparticles (PRECIOUS-01) PRECIOUS-01 Polymeric micelles [180]
NCT05340725 (Phase II/III) Rectal dexmedetomidine niosomes for postoperative analgesia in pediatric cancer patients (DEX-NANO). Dexmedetomidine niosomes Liposomes [181]
NCT05700955 (Phase I) Neoadjuvant chemoimmunotherapy in recurrent glioblastoma Pembrolizumab and temozolomide Liposomes
NCT04675996 (Phase I) First-in-human study of INT-1B3 in patients with advanced solid tumors INT-1B3 Liposomes [182]
NCT05969041 (Phase I) Study of MT-302 in adults with advanced or metastatic epithelial tumors (MYE symphony) MT-302 Liposomes
NCT05497453 (Phase I/II) A Phase 1/2 study to evaluate OTX-2002 in patients with hepatocellular carcinoma and other solid tumor types known for association with the MYC oncogene (MYCHELANGELO I) OTX-2002 Liposomes
NCT05001282 (Phase I/II) A study to evaluate ELU001 in patients with solid tumors that overexpress folate receptor alpha (FRα) ELU001 CDC
NCT02439580 (Phase I) Effect of Annona muricata leaves on colorectal cancer patients and colorectal cancer cells Annona muricata extract, placebo Polyphenols

NPs, nanoparticles; NCT, national clinical trial; CDC, C’Dot drug conjugate.

Nanoformulations of natural products

Natural products and their numerous derivatives are valuable sources of compounds with anticancer efficacy, including extracts from animals, plants, and microorganisms, among others [183184]. Due to their effects on multiple cellular signaling pathways and their acceptable safety profile, more than half of all clinically used chemotherapeutic drugs are natural products [185186]. For anticancer therapy, various nanoformulations have been designed to: (1) decrease or abrogate MDR and/or the adverse and toxic effects produced by anticancer therapeutics and (2) improve targeted drug delivery and bioavailability [187]. In this section, we will discuss the anticancer efficacy of natural product nanoformulations.

Certain nanoformulations (i.e., dendrimers, polymers, liposomes, nano-emulsions, and micelles) can decrease particle sizes and have been used to increase bioavailability and minimize the adverse effects produced by the ligand(s) they deliver to the tumor site(s) [188]. The nanoformulations generally have a size range from 10 to 100 nm and can be classified as nanocarriers, no-carrier-added nanosuspensions and polymer-drug conjugates [189]. During the formulation of nano-drugs, investigators need to consider whether the drug is dissolved, entrapped, encapsulated or adhered to the drug carrier [28]. The formulation should target the tumor site and release the drug during the delivery process [4]. Various methods used to synthesize nanoformulations are selected based on the different nanoformulations (i.e., dendrimers, polymers, liposomes, nano-emulsions, and micelles) [190]. It is vital to indicate that certain natural products alone have off-target toxicity, low solubility, and low cell permeability, which can limit their anti-cancer efficacy [191]. Therefore, a drug delivery system, based on nanoformulation methods, represents an alternative strategy to significantly increase the chemopreventive and chemotherapeutic efficacy of natural products [183]. Nanopreparation techniques can be used to successfully deliver the natural products to the target site [157]. In this section, we discuss relevant nanoformulations of natural products that have the potential to be used as anticancer drugs.

(1E,6E)-1,7-bis (4-hydroxy-3-methoxyphenyl)-1,6-heptadiene-3,5-dione, also known as curcumin (CUR), a natural bioactive compound, which has antioxidant and anticancer efficacy, belongs to the curcuminoid subgroup of polyphenols (Figure 3) [192193]. Due to its poor bioavailability, certain nanotechnology strategies are utilized to improve the pharmacokinetic properties of CUR, thereby increasing the probability of significant therapeutic efficacy. One of these strategies is to deliver CUR to the target site, using various stimuli, such as light, pH, magnetic field, solubility, and chemical modifications [194]. A nanosystem was developed by encapsulating CUR into pH-responsive poly d,l-lactic-co-glycolic acid (PLGA) microspheres, which delivered hydrophobic drugs to the triple negative breast cancer cell lines, MCF-7, MDA-MB-231 and MDA-MB-468, that overexpress folate receptors [195]. Lai et al. [196] developed a pH-reactive hyaluronic acid-based NPs for targeted CUR delivery to increase its anticancer efficacy biosafety profile. Numerous researchers used photoresponsive nanoformulations to increase the stability and solubility of CUR. Alvi et al. [197] entrapped Au-liposome NPs as an adjuvant for photothermal therapy (PTT). NIR light exposure (780 nm for 5 min) in B16 melanoma cells activated the release of CUR nanocrystals, which coalesced to form CUR microcrystals for the continuous release of the active compound [197]. Studies have been conducted to increase the solubility, bioavailability, and stability of CUR [198199]. A self-assembling acylated ovalbumin nanogel for CUR was synthesized that had a relatively uniform size distribution and higher stability under high ionic strength and different pH [198]. Due to the poor water solubility of CUR (11 μg/L) [200], researchers have synthesized nanoformulations of CUR to surmount this problem. Mangalathillam et al. [201] developed curcumin-loaded chitin nanogels using biocompatible and biodegradable chitin. The formulation was cytotoxic to A375 melanoma cells, at concentrations from 0.1 to 1.0 mg/mL. Gou et al. [202] developed curcumin encapsulated with monomethoxy poly(ethylene glycol)-poly(ε-caprolactone) (MPEG-PCL) micelles, using a nano-precipitation method. This formulation decreased the viability (IC50=5.78 μg/mL) of C-26 colon cancer cells. CUR or Cur/MPEG-PCL micelles (curcumin, 100 mg/kg) were intravenously injected in rats and blood collection was performed at different intervals of time. For Cur/MPEG-PCL micelles, the following PK data was obtained: tmax=5 min, t1/2=34.2 min, area under the curve (AUC)(0→t)=47,642.1 μg/L/min, AUC(0→∞)=47,864.6 μg/L/min, and Cmax=430.5 μg/mL. These values were significantly higher than those for free curcumin and indicated that the encapsulation increased the t1/2, AUC(0→t), AUC(0→∞), and Cmax in vivo. Furthermore, a transgenic zebrafish model was used to ascertain the antiangiogenic efficacy of Cur/MPEG-PCL micelles. Zebrafish were exposed to free CUR (5 μg/mL) or Cur/MPEG-PCL micelles (5 μg/mL) for 72 h. The results indicated that both free CUR and Cur/MPEG-PCL micelles inhibited angiogenesis by different magnitudes (significantly greater in micellar CUR), resulting in the abnormal formation or absence of intersegmental vessels in the Zebrafish.

The compound (3R,5aS,6R,8aS,9R,12S,12aR)-octahydro-3,6,9-trimethyl-3,12-epoxy-12H-pyrano[4,3-j]-1,2-benzodioxepin-10(3H)-one, also known as artemisinin (ART), is present in the plant Artemisia annua [203]. In vitro and in vivo studies have reported that ART has antitumor, antifungal, anti-malarial, and anti-ulcer efficacy [204]. However, ART has a lower solubility in water, a shorter half-life and an increased first-passage metabolism, which limits its therapeutic efficacy [205]. The synthesis of ART-containing NPs improve the pharmacokinetic profile of ART and increase the in vivo efficacy of ART. Zhang et al. [206] synthesized hollow mesoporous manganese trioxide NPs for the in vivo delivery of ART. They synthesized a drug delivery system, TKD@RBCm-Mn2O3-ART, that specifically releases and produces a uniform dispersion of chemotherapeutic drugs in tumors. ART was loaded into Mn2O3 to facilitate the co-delivery of Mn2+ with ART. In vitro experiments indicated that the cumulative drug release reached a maximum of 97.5 % in the presence of glutathione. TKD@RBCm-Mn2O3-ART released Mn2+ and ART simultaneously, which produced high levels of reactive oxygen species (ROS) and DNA damage. In vivo results indicated that TKD@RBCm-Mn2O3/ART penetrated deep into solid tumors and produced a definitive diagnosis and treatment of breast cancer. Yu et al. [207] modified ART-loaded liposomes with activatable cell-penetrating peptides. This formulation was composed of Tf-coated, dihydroartemisinin (DHA), l-buthionine-sulfoximine (BSO), and CellROX-loaded liposomal nanoparticles (Tf-DBC NPs). ART was integrated into acidic pH-responsive liposomes that significantly decreased the proliferation of HepG2 cancer cells (IC50=2.5 μmol/L) and avoided oxidative stress to normal cells, as indicated by a fluorescence response of CellROX in HepG2 and normal cells. Tf-coated liposomal NPs encapsulating Cyanine7 (Tf-Cy7 NPs) were prepared to provide data about the circulation profile of the nanoformulation in tumor-bearing female BALB/c nude mice (that were subcutaneously injected with HepG2 cancer cells to create a HepG2 tumor model). After injecting 0.9 mg/kg via intravenous infusion, the circulation half-life (t1/2) was 4.81 h, which indicated the presence of Tf-Cy7 NPs in the bloodstream, thus indicating that the formulation was stable in vivo. In vivo and ex vivo analysis showed that fluorescence occurred only in tumorous areas after the intravenous injection of Tf-DBC NPs, indicating that the normal cells were not affected by the oxidative stress induced by the nanoformulation. Halevas et al. [208] designed a dendritic-linear-dendritic hybrid copolymer to encapsulate ART, using linear PEG chains and hyperbranched 2,2-bis(hydroxymethyl)propionic acid (bis-MPA) and this preparation was hydrophilic. The IC50 value of Art-loaded bis-MPA PEG6k-OH, pseudo-generation 4 (G4-PEG6k-OH) hyperbranched dendritic scaffolds (AHDS) in MCF-7 after 72 h of incubation was 30.5 μmol/L. The results indicated that the formulation maintained the anticancer efficacy of ART but was not toxic in normal 3T3 fibroblasts cells.

A significant interest in the development of nano-drugs has resulted in more inventions related to preparation methods and applications [209]. Nano-curcumin and its derivatives, such as isovanillin curcumin, ferulic acid curcumin, 3,5-di-tert-butyl-4-hydroxybenzaldehyde curcumin, syringaldehyde curcumin, and 4-methoxy-1-naphthaldehyde curcumin, have been the most common compounds evaluated for pharmacological efficacy in cancer cells. Xiao et al. [210] (patent no. WO2020088702A1) invented a method of making nanocrystals that increased the bioavailability of certain curcuminoids and camptothecin. The patent (WO2020249383A1) reported the invention of an active substance delivery system for curcumin to be used for the treatment of peritoneal metastases in numerous cancers [211]. Although the nanoparticle formulations of natural products have garnered increased attention, challenges, and problems, such as nanotoxicity, must be determined.

Albumin NPs

Albumin is the most abundant protein in the plasma that is involved in many physiological processes, such as maintenance of osmotic pressure, transport of hormones, ligands, drugs, and neutralization of free radicals, among others [212]. It is a multifunctional protein that is highly stable, biodegradable, biocompatible and non-immunogenic [213]. Albumin has numerous high-affinity hydrophobic binding sites that can bind a number of ligands and drugs [214]. Thus, based on these aforementioned properties, albumin is commonly used as a nanocarrier [213]. Albumin NPs can be formulated in combination with other nanocarriers, such as polymeric micelles, liposomes, metallic NPs, silica, and nanosheets, among others, to improve cellular uptake efficiency and minimize the immunogenicity of the parent drug [215], [216], [217], [218], [219], [220]. Albumin has a number of properties that make it a highly suited nanocarrier. Albumin has an isoelectric point of 4.25 at physiological pH [221] and thus, it has dense negatively charged areas that will induce a strong repulsive force towards negatively charged proteins, such as hemopexin, haptoglobin and transferrin, which increases its stability and half-life in the circulation [222], [223], [224]. Multiple functional groups in albumin, such as amino and carboxyl groups, facilitate the binding with ligands by forming covalent or non-covalent bonds [225]. The most commonly used albumin for nanoformulations is derived from BSA or human serum albumin (HSA), although ovalbumin can also be used [226]. HSA is a protein (molecular weight [MW]=133 kDa, 5.2 × 5.3 × 11.7 nm in dimensions) that consists of three homologous α-helical domains that are composed of ten anti-parallel helices that form a heart-shaped, asymmetric structure [227]. BSA NPs offer biocompatibility, non-toxicity, biodegradability, non-immunogenicity and a higher drug-binding capacity [228]. BSA (MW=69 kDa) is an acidic, globular, single polypeptide chain protein that is found abundantly in some mammals, with a half-life of 19 days [228]. Anti-cancer drug binding with albumin can be done via covalent and/or non-covalent conjugation [229]. Since albumin NPs are very versatile molecules, this conjugation can be manipulated as required.

The majority of albumin nanoformulations are prepared by desolvation, which involves: (1) the precipitation of NPs by depleting the solvation layer, using a dehydrating agent, such as ethanol; (2) densification and stabilization by chemical crosslinking; (3) additional purification to remove redundant solvents and crosslinking molecules [230]. Desolvation is a simple and inexpensive process that can be completed in approximately 3 h [231]. In addition, thermal gelation, emulsification (used for hydrophobic drugs) and electrostatic interactions (used for positively charged drugs), are used when desolvation is not achievable [226]. For other proteins, nanoformulations, such as gelatin NPs, collagen NPs, milk proteins, casein NPs, silk fibroin NPs, elastin NPs, and various other plant-based protein NPs have been used. Methods such as, pH variation [232], nano-spray drying [233], sonochemical method [18], phase separation [234], rapid laminar jet [235], milling [236], and polymer chain collapse [237], are the most commonly used preparatory methods [18233238239]. However, there is no perfect approach and each method has its advantages and disadvantages. Abraxane® (FDA approved) is one of the well-known nanoformulations that are based on albumin NPs [70240]. Other examples include DOX-loaded in BSA-dextran-chitosan NPs prepared by thermal gelation [241], poorly soluble 10-hydroxycamptothecin-loaded in BSA NPs via emulsification [242], and paclitaxel-loaded BSA NPs, using desolvation [243].

Albumin NPs target the cancer cells via active or passive targeting [228]. Active targeting includes binding of the NPs to the surface receptors entering the cell via internalization. In contrast, there is an EPR effect in passive targeting. Cancer cells usually express a high level of certain glycoproteins, such as Gp18, Gp30, and Gp60, secreted protein acidic rich in cysteine (SPARC) receptors, which are important binding sites for albumin NPs [228]. The accumulation of albumin-based NPs in solid tumors can be affected by the angiogenesis, leaky vasculature, and defective lymphatic drainage [213]. PEGylation increases the circulation half-life by >50-fold, decreases immunogenicity, and increases the accumulation of BSA nanoformulation of 5-fluorouracil in tumors due to the EPR effect [244]. HSA NPs represent a viable approach for targeted drug delivery by increasing drug bioavailability and distribution, decreasing drug toxicity and immunogenicity [70240]. Despite having several advantages, albumin NPs have certain drawbacks. For example, manipulation of particle size, shape, distribution, and stability poses major challenges. Furthermore, the factors affecting the toxicity of albumin NPs, such as the composition and surface properties must be precisely controlled. There is also batch-to-batch quality variation, which can impede the scaling-up process. A compelling approach to overcome this challenge is recombinant protein technology, where mono-dispersity and predefined characteristics of polymers along with the predictable arrangement of crosslinking groups, binding moieties at specific sites or their programmable degradation rates, makes them suitable and convenient for drug delivery and tissue engineering systems [226]. Moreover, there are limited publications about the immunogenicity of proteins, although there are no published reports about the antigenicity of albumin NPs, even after intravenous (IV) injections [245], [246], [247], [248].

Polymeric micelles

Polymeric micelles are composed of various amphiphilic polymers that preferentially self-assemble in an aqueous medium [249]. Each unit of polymeric micelles comprises a hydrophilic and a hydrophobic segment (Figure 3). These amphiphilic polymers are synthesized using various polymeric blocks, which can be customized, as needed, depending on the size, capacity, hydrophobicity, micellization capacity, and stability in blood [249]. Similar to the albumin NPs, the nanosize of the micelles facilitates penetration through the leaky vasculature and retention due to defective lymphatic drainage in tumors [250]. The outer layer of micellar encapsulation is typically hydrophilic and therefore, it evades detection by the reticulo-endothelial system [251]. Polymeric micelles can be modulated or adjusted by the addition of ligands on the surface that are involved in active targeting [250]. Various building blocks of polymers, such as PEG coupled with either phosphatidylethanolamine (PE), polypropylene oxide (PPO) triblock polymer (PEG-PPO-PEG), PEG-amino acids or PEG-carbonates, can be used individually or in combination. PEG-PE was one of the first lipid-based amphiphilic polymers to be used to integrate in the liposomal lipid bilayer to sterically stabilize liposomes [252]. In the next step of PEG-PE preparation by self-assembly, drugs are added and this process is primarily based on the physicochemical characteristics of the drug and the method of preparation [253]. PEG-PE self assembles at a low micellar concentration, allowing for the encapsulation of hydrophobic drugs [254255]. PEG-PE, coupled with vitamin-E micelles, has been used to formulate camptothecin, DOX and paclitaxel, with sizes ranging from 15 to 100 nm [256]. PEG-PPO-PEG triblock polymers, also known as pluronics or poloxamers, have core shell structures of pluronics, 10–100 nm in diameter and the preparation and formation process is temperature- and concentration-dependent [249]. These polymers, such as Pluronic® P85 carrying DOX and Pluronic® P123/F127 block copolymers carrying rhodamine-123, rhodamine-G, DOX and paclitaxel, have been reported to inhibit the ABC transporters, P-glycoprotein (P-gp), multidrug resistance-associated protein-1 (MRP1) and breast cancer resistance protein (BCRP), thereby effectively decreasing MDR [257], [258], [259]. Anti-cancer drugs, either individually or in combination with adjuvants or other anti-cancer drugs, can be loaded onto pluronics [260], [261], [262], [263], [264]. Du et al. [265] used α-tocopherol polyethylene glycol 2000 succinate (TPGS2k) to modify PEG NPs and the NPs were co-loaded with simvastatin and DOX [265]. The resulting formulation significantly decreased lipid raft formation (>80 %) in SW620/AD300 (colon cancer cells resistant to DOX) cells. In the presence of 20 μg/mL of free simvastatin (SV) or DOX, SW620 cell viability following incubation with SV@TPGS2k-PLGA NPs, DOX@TPGS2k-PLGA NPs and SV/DOX@TPGS2k-PLGA NPs was 28.1 , 25.9 and 13.2 %, respectively and the viability of SW620/AD300 cells was 27.2 , 24.1, and 11.3 %, respectively. The preparation of PEG-amino acids has only been recently reported in the scientific literature. Polypeptide chains can be linked to PEG and amino acids, based on the type of formulation desired, are used. For example, the amino acids, l-lysine, l-glutamate, and l-aspartate are used for hydrophilic preparations [266]. DOX [267268] and cisplatin [269] can be loaded to produce PEG-amino acids.

The preparation of polymeric micelles depends on the following factors and conditions: temperature, pH, micellar concentration, ionic charge, surface characteristics, interaction with organic counterparts, and kinetic stability [250]. Some of the methods used for the preparation of polymeric micelles include dissolution, dialysis, emulsion and solvent-casting, and film hydration [270], [271], [272], [273]. The characterization of micelles is clinically important as it helps to define and predict their interactions with other molecules [274]. The thorough characterization of micelles involves the determination of critical micellar concentration (CMC), the drug payload, and its release [274], [275], [276]. Generally, this characterization is accomplished using dynamic light scattering (DLS) [277], atomic force microscopy (AFM) [274278], cryo-transmission electric microscopy (Cryo-TEM) [279], X-ray scattering [280], [281], [282], and Foerster resonance energy transfer (FRET) [275276283284]. A study involving DOX formulated with a dextran-based polymeric nanosystem reported a higher toxicity in various cancer cells in vitro, compared to the parent drug [285]. SNB-101 is a novel micellar nanoformulation for the anticancer drug, SN-38, that has an approved investigational new drug (IND) application for the treatment of various cancers and it is in Phase I clinical trials for tumors (NCT04640480) [179]. A drug named PRECIOUS-01 is undergoing evaluation in Phase I clinical trials for advanced solid tumors. This is a PLGA-based nanocarrier that has promising therapeutic potential [180].

It is important to note that polymeric micelles can prematurely release drugs prior to reaching the target site [286]. Drug release is an essential step in drug delivery, as an optimum drug concentration must reach the tumor site(s) to achieve an intracellular level that is efficacious [287]. Membrane dialysis or ultracentrifugation can be used to determine the amount of a drug that is released from polymeric micelles [288], [289], [290], [291].

Liposomes

Dr. Alec Bangham discovered liposomes in 1964 [292], which are now some of the most commonly used nanosystems. Basically, liposomes are spherical systems that have two outer hydrophilic layers with a lipophilic layer in between [293]. Liposomes that have more than one concentric lipid bilayer are known as multilamellar vesicles (MLVs) [293]. Liposomes are typically classified into small unilamellar vesicles (SUV), large unilamellar vesicles (LUV), MLVs, and multi-vesicular vesicles (MVV) [294]. The best route of administration for liposomal drug delivery is the parenteral route, as these formulations will circumvent: (1) first pass metabolism; (2) adverse gastrointestinal (GI) effects; (3) poor GI tract absorption [294]. Furthermore, parenteral administration results in a higher bioavailability and increased efficacy, compared to oral or rectal administration [295]. Liposomes are biodegradable, highly biocompatible, and are prepared via self-assembly [296]. One of the most extensively used polymers is PEG. PEGylated liposomes are considered to be a “stealth formulation”, as these molecules are sterically stabilized liposomes that have a decreased propensity to aggregate, which decreases their interaction with opsonins, thereby decreasing their removal from the circulation by phagocytosis [297]. Doxil® was the first liposome-based DOX formulation and FDA approved liposome-based formulations include DaunoXome® (AIDS-related Kaposi’s sarcoma), Myocet® (metastatic breast cancer), DepoCyt® (neoplastic meningitis), Marqibo® (acute lymphoblastic leukemia), Onivyde™ (metastatic adenocarcinoma of pancreas), and Mepact® (metastatic osteosarcoma) [196]. Lipid-saporin combination of NPs can serve as a novel therapeutic regimen for ABCB1/ABCG2-positive drug-resistant cancers. Patel et al. developed palmitoyl carnitine-anchored nano-liposomes for targeted delivery of gemcitabine elaidate for the treatment of pancreatic cancer [298]. This combination therapy of gemcitabine elaidate with palmitoyl dl-carnitine chloride inhibits protein kinase-C inhibitor in a nano-liposomal carrier that had increased cellular uptake, producing a decrease in angiogenesis and increasing the anticancer potential in both 2D and 3D models of pancreatic tumors in vitro. Dinakar et al. exploited the abundance of folate receptors in breast cancer to develop a folate and poly-I-Lysine conjugate and coated it on liposomes for targeted drug delivery [299]. Their result indicates a decrease in the IC50 values and the induction of death in breast cancer cells via folate receptor-mediated internalization, followed by rapid drug release. Furthermore, they determined that luteolin formation inhibits cell migration and proliferation by regulating the expression of vascular endothelial growth factor (VEGF) and the induction of apoptosis by caspase-3 upregulation. Zhan et al. reported that liposomal dexmedetomidine could provide controlled, adjustable and on-demand local anesthesia in vivo [181]. INT-1B3 is a novel synthetic liposomal miR-193a-3p (tumor suppressor microRNA) mimicking NP [182]. INT-1B3 decreased target gene expression, resulting in decreased cell proliferation, increased cell cycle arrest, initiation of apoptosis, inhibition of migration, DNA damage and cell senescence in vivo [182]. In addition, research in the area of liposomal mRNA vaccine delivery (Figure 3) has gained an increased interest in recent years. These NPs have the unique advantages of controlled release, improved biosafety and high biocompatibility [80]. They also serve as immunotherapy adjuvants, in addition to delivering mRNA vaccines. This results in killing the cancer cells by reviving the anti-tumor immune response, maintaining the response and improving the threshold, producing prolonged survival and a better quality of life in patients with advanced tumors [300].

The laboratory manufacturing of liposomes primarily involves the dissolution of phospholipids in certain organic solvents, ethanol, isopropyl ether, diethyl ether, chloroform or methanol, either individually or in various combinations [296]. The fundamental molecular units of a liposome preparation are phospholipids and cholesterol [296]. At the CMC, typically in the nanomolar range, liposomes undergo self-assembly when exposed to an aqueous environment [301302]. In a method known as lipid film hydration, an organic solvent, such as trichloromethane, is used to dissolve the phospholipids [303304]. The lipophilic drug is subsequently added to the solvent to form a one-phase solution and the removal of the organic solvent is done under a vacuum, forming thin sheets of lipid with uniform distribution of the drug along the film [303304]. This procedure produces hydration of the sheets in the presence of an aqueous buffer that is above the glass transition phase (a reversible transition phase that occurs when an amorphous material is cooled or heated over a certain range of temperatures, which causes the material to become hard and brittle upon cooling and softer upon heating) of the lipid [295305]. If the drug is hydrophilic, it will dissolve in the aqueous buffer, leading to a particle size greater than 500 nm, with an entrapment efficiency of 10–30 %. In contrast, for lipophilic drugs, an optimum vesicle size is obtained, with an entrapment efficiency >90 % [306307].

Another technique, known as solvent dispersion, involves phospholipids being dissolved in a water-miscible organic solvent, usually ethanol [308], [309], [310]. If the lipophilic drug is not miscible in ethanol, other organic solvents, such as diethyl ether or ether–methanol mixture, may be used [296]. Instead of an aqueous buffer being added to the organic solution in the previous method, the mixture (organic solvent and drug) is added to the aqueous buffer, the solvent becomes diluted and there is a spontaneous formation of MLVs, usually of size above 500 nm [294]. This technique is best suited for lipophilic drugs.

Reverse phase evaporation, yet another technique used to formulate liposomes and it is the most frequently used technique for forming liposomes with hydrophilic drugs [311]. This technique can entrap a large amount of an aqueous core during the formation, yielding high entrapment efficiency [312313]. In a water/oil emulsion, the hydrophilic drug is dissolved in water and the phospholipid is dissolved in a water-immiscible solvent, usually diethyl ether or isopropyl ether [314]. The removal of the organic solvent using a vacuum leads to the gradual formation of a gel [296]. A liposomal dispersion is produced when the organic solvent is further evaporated [315]. The passive drug-loading efficiency for liposomes prepared via reverse phase evaporation is 30–50 % and it can be increased to >90 %, using active drug-loading techniques [311316317]. The stability of the final preparation depends on the physical and chemical environment and the storage conditions. The size of the carrier, the drug payload and chemical stability, can be significantly altered by improper storage conditions [295]. Importantly, the majority of marketed liposomal formulations need to be stored at 2–8 °C [295].

Scaling-up technologies for liposomal preparation are very limited. One of the most commonly used techniques is the rapid injection of ethanol and phospholipids into an aqueous medium, followed by extrusion of solvent, using a polycarbonate membrane. This approach ensures size reproducibility and the quality of the liposomes [308]. A typical commercial preparation involves the following sequential steps: preparation of buffer, filtration, preparation of a phospholipid solution, second filtration, lipid hydration, extrusion, diafiltration, dilution, sterile filtration, and filling [318], [319], [320], [321].

Polyphenols

An estimated 10,000 polyphenolic substances have a structure that includes multiple hydroxyl groups and aromatic rings [322323]. These compounds are present in high concentrations in various fruits and vegetables as secondary plant metabolites [323], and the most frequently occurring of these are the flavonoids (about 60 %) and phenolic acids (about 33 %) [324]. Polyphenols are primarily obtained from (1) fruits: plums, apricots, oranges, apples, tomatoes, cherries, peaches, berries, and other tropical fruits; (2) vegetables: broccoli, spinach, onions, carrots, olives, beans, capers, artichokes and cauliflowers; (3) herbs and spices: celery, rosemary, cloves, turmeric, parsley, thyme, mint, sage, dill weed, ginger, and curry and; (4) other sources: red wine, black or green tea, cocoa, coffee, fruit juices, beer, seeds, grains, and nuts [191322325], [326], [327], [328], [329], [330], [331]. Polyphenols have been reported to be efficacious in dyslipidemia, GI diseases, cardiovascular diseases, inflammatory disorders, neurological disorders and various cancers, due to their antioxidative and immunomodulatory effects [332], [333], [334]. Epigenetic modifications (primarily a result of DNA methyltransferases or histone deacetylases) also play an important role in the efficacy of dietary polyphenols. For example, the consumption of quercetin-rich food is significantly correlated with decreasing the risk of lung and gastric cancer [335], [336], [337]. In vitro studies indicate that resveratrol is efficacious in inhibiting the proliferation of stomach, breast, colon, prostate, lung, pancreatic, and thyroid cancer cell lines [338]. Green tea extract contains a high concentration catechins, such as epigallocatechin-3-gallate, epicatechin-3-gallate, and epigallocatechin, which inhibit cancer cell growth, metastasis, and angiogenesis in vitro [339], [340], [341]. An in vitro study has shown that epigallocatechin-3-gallate modulates the expression of regulatory proteins involved in cell cycle, activates caspases-3, caspase-8, and caspase-9 and suppresses the activation of nuclear factor-kappa B (NF-κB), resulting in the induction of apoptosis and the inhibition of cancer cell growth [342]. Cruciferous vegetables in the Brassica genus (e.g., cabbage, broccoli, kale) contain sulfur-rich compounds, known as glucosinates, that can decrease the risk of lung cancer and colon cancer in humans [343344].

Dietary polyphenols are primarily present in glycosylated forms, with sugar residues linked to the aromatic ring or the hydroxyl group [326345], which are more efficiently absorbed by the GI tract via passive diffusion [345]. Gallic acid, catechins, flavanones, and quercetin glucosides have been reported to have the highest GI absorption, compared to galloylated tea catechins, proanthocyanidins, and anthocyanins, which are among the least GI-absorbed polyphenols [346]. The extraction process used to obtain polyphenols from natural sources can be done using fresh, frozen, or dried samples and before the process begins, the material is exposed to milling, grinding, drying, and homogenization [347]. The freeze-drying process produces a higher yield of phenolic compounds, compared to air drying [348]. Liquid-liquid and solid-liquid extraction are widely used methods because of their applicability, efficiency, and ease of use [347349]. Solvents, such as ethanol, ethyl acetate, acetone, and diethyl ether, are generally used in various ratios with water. The extraction of nonpolar compounds, such as oils, waxes, chlorophyll and sterols and the extraction from less polar solvents, such as chloroform, benzene, dichloromethane and hexane, may be possible [347]. Solvents, such as methanol, yield a higher level of low molecular weight polyphenols, whereas acetone yields a higher level of high molecular weight flavanols [350], [351], [352], [353]. Modern techniques, like supercritical fluid technology, have been widely used for the separation of polyphenols from their natural sources [354]. This technique prevents the oxidation and degradation of the polyphenols [354355]. The main advantage of supercritical fluids is that they dynamically change the solvent characteristics and yield the desired levels of the extracts from raw materials [354]. Typically, supercritical CO2 is used as the solvent because of its high penetration into cellular materials or tissue explants and its solvent power [356]. Propane, argon, and sulfur hexafluoride (SF6) are supercritical fluids that have been used in supercritical fluid extraction methods [356]. Certain end products of polyphenol extraction include curcuminoids, stilbenes, tannins, lignans, phenolic acids, and flavonoids, i.e., flavanones, flavanols, anthocyanidins, catechins, isoflavones, and chalcones [347].

Discussion

Cancer is one of the leading causes of death in the world and the efficacy of anticancer drugs can be significantly decreased or abolished by MDR that occurs via a number of mechanisms. Furthermore, physiological factors, such as higher IFP, hypoxia, leaky tumor vasculature, low pH, and irregular drug penetration due to inaccessible location of cells, also decrease the efficacy of anti-cancer drugs (Figure 4) [28]. Unfortunately, increasing the dose of an anticancer drug, as well as drug combinations, significantly increases the probability of adverse effects and toxicity, thereby causing a decrease or absence of patient compliance, which increases cancer cell proliferation and a subsequent relapse. Studies over the past two decades suggest that certain nanomedicine formulations can overcome some of the mechanisms that produce MDR in cancer cells. There are certain advantages in using nanomedicine, compared to conventional preparations [357]. Nanoformulations effectively permeate and overcome biological barriers, such as transporters in cell membranes throughout the body. Some studies have reported that nano-preparations bypass ABC transporters and effectively kill the cancer cells. One study reported an increase in the nuclear localization of DOX in a few sets of sensitive- and drug-resistant cells when the drug was introduced with nanospheres, compared to parent drug [358]. It is important to state that cancer nanomedicine is not a “magic bullet” for the treatment of cancer. Certain factors, such as the interaction between NPs and surface proteins, i.e., protein corona formation, while maintaining the stability and half-life of the drug, must be taken into consideration. The size and binding affinity of NPs must be evaluated for the internalization of the therapeutic. Furthermore, the magnitude of cellular uptake is critical in determining the efficacy of nanoformulations. A drug may not enter the cell even after successful delivery near the target cell due to inadequate cellular uptake. Furthermore, NP endosomal release is crucial for successful drug delivery, especially for siRNA-based therapeutics. Although numerous researchers have studied the nanodelivery and co-delivery of NPs for the reversal of MDR, there have been ambiguous results in clinical trials [359], [360], [361]. This could have been due to a number of factors, such as a decrease in the effective release of a drug from the core [249]. Also, immune response reactivation is only seen in a small subset of patients. Finally, the absence of biomarkers to assess the efficacy of cancer nanotherapeutics is one of the main obstacles to attaining tailored immunotherapy.

Figure 4:

Figure 4:

An overview of factors affecting multidrug resistance in cancer.

As discussed in the main text, it is apparent that the field of cancer nanomedicine has great potential. Cancer nanomedicine caters unique advantages such as controlled and sustained release, better biosafety, higher biocompatibility, and lower systemic toxicity. Furthermore, certain types of NPs like liposomes and polymeric micelles serve as adjuvants in addition to delivering therapeutic payloads like mRNA vaccines. Nanoformulations can efficiently accumulate at the tumor site(s) via active or passive targeting or by the EPR effect. NPs have an increased circulation time and increased half-life. The interaction between NPs and biological components may improve drug delivery. Furthermore, nanomedicine can co-deliver a multi-drug or a multi-stage payload, which represents a promising approach to overcome MDR. Nonetheless, it will be critical to ascertain the toxicity profile of these novel therapeutics to determine if these compounds can be used to treat cancer in humans. Moreover, certain challenges, such as varied efficacy of industrial production, various side effects like pruritus, rashes, nausea, diarrhea and thyroid disorders, must be further evaluated. Consequently, in vivo studies must be conducted to determine the safety and efficacy of anticancer nanomedicines. Finally, researchers in academia and industry must develop scaling-up technologies without the loss of reproducibility. The gap in clinical translation of cancer nanomedicine can be further decreased by the development of humanized animal models, as there are significant structural and biochemical differences between tumors in rodents and humans. These humanized animal models must contain human-derived factors such as immune cells. EPR is vital for drug delivery; however, the uncertainty of the EPR effect in different patients leads to dubious results in clinical trials. This can possibly be solved by determining the level of the EPR effect in the patients. Although most NPs achieve active targeting, the mechanism behind this phenomenon remains unclear. However, it is firmly believed that the efficacy of cancer nanomedicine can be significantly improved by combining the advancements in cancer biology and nanotechnology.

The uniqueness of each case of cancer is a huge hurdle in the clinical translation of cancer drug research. In the past 50 years, a plethora of knowledge has been amassed, producing notable progress in the field of cancer nanomedicine. The era of personalized and customized nanomedicine has inevitably arrived that must help in overcoming unique but fundamental mechanisms of MDR.

Acknowledgments

The authors are thankful to the Teaching Assistantship received from the College of Pharmacy and Health Sciences, St. John’s University, New York.

Footnotes

Research ethics: Not applicable.

Informed consent: Not applicable.

Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

Competing interests: Authors state no conflict of interest.

Research funding: None declared.

Data availability: Not applicable.

Contributor Information

Wei Cai, Email: 20120941161@bucm.edu.cn.

Zhe-Sheng Chen, Email: chenz@stjohns.edu.

References

  • 1.Feynman RP. There’s plenty of room at the bottom. Resonance. 2011;16:890–905. doi: 10.1007/s12045-011-0109-x. [DOI] [Google Scholar]
  • 2.Nasrollahzadeh M, Sajadi SM, Sajjadi M, Issaabadi Z. Chapter 1 – an introduction to nanotechnology. In: Nasrollahzadeh M, Sajadi SM, Sajjadi M, Issaabadi Z, Atarod M, editors. Interface science and technology. Amsterdam, Netherlands: Elsevier; 2019. pp. 1–27. [Google Scholar]
  • 3.Yao C, Lu J. Introduction to nanomedicine. In: Webster TJ, editor. Nanomedicine: nanotechnology, biology, and medicine. Cambridge, England: Woodhead Publishing; 2012. pp. 3–19. [Google Scholar]
  • 4.Astruc D. Introduction to nanomedicine. Molecules. 2016;21:4. doi: 10.3390/molecules21010004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Nel AE, Madler L, Velegol D, Xia T, Hoek EM, Somasundaran P, et al. Understanding biophysicochemical interactions at the nano-bio interface. Nat Mater. 2009;8:543–57. doi: 10.1038/nmat2442. [DOI] [PubMed] [Google Scholar]
  • 6.Albanese A, Tang PS, Chan WCW. The effect of nanoparticle size, shape, and surface chemistry on biological systems. Annu Rev Biomed Eng. 2012;14:1–16. doi: 10.1146/annurev-bioeng-071811-150124. [DOI] [PubMed] [Google Scholar]
  • 7.Cimalla P, Werner T, Gaertner M, Mueller C, Walther J, Wittig D, et al. European conferences on biomedical optics. Proc SPIE 8802. Optical coherence tomography and coherence techniques VI. Munich, Germany: 2013. Magnetomotive imaging of iron oxide nanoparticles as cellular contrast agents for optical coherence tomography. [Google Scholar]
  • 8.Dreaden EC, Alkilany AM, Huang X, Murphy CJ, El-Sayed MA. The golden age: gold nanoparticles for biomedicine. Chem Soc Rev. 2012;41:2740–79. doi: 10.1039/c1cs15237h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Elsabahy M, Wooley KL. Design of polymeric nanoparticles for biomedical delivery applications. Chem Soc Rev. 2012;41:2545–61. doi: 10.1039/c2cs15327k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Kamaly N, Xiao Z, Valencia PM, Radovic-Moreno AF, Farokhzad OC. Targeted polymeric therapeutic nanoparticles: design, development and clinical translation. Chem Soc Rev. 2012;41:2971–3010. doi: 10.1039/c2cs15344k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Tang F, Li L, Chen D. Mesoporous silica nanoparticles: synthesis, biocompatibility and drug delivery. Adv Mater. 2012;24:1504–34. doi: 10.1002/adma.201104763. [DOI] [PubMed] [Google Scholar]
  • 12.Greish K. Enhanced permeability and retention (EPR) effect for anticancer nanomedicine drug targeting. Methods Mol Biol. 2010;624:25–37. doi: 10.1007/978-1-60761-609-2_3. [DOI] [PubMed] [Google Scholar]
  • 13.Bertrand N, Wu J, Xu X, Kamaly N, Farokhzad OC. Cancer nanotechnology: the impact of passive and active targeting in the era of modern cancer biology. Adv Drug Delivery Rev. 2014;66:2–25. doi: 10.1016/j.addr.2013.11.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Salvioni L, Rizzuto MA, Bertolini JA, Pandolfi L, Colombo M, Prosperi D. Thirty years of cancer nanomedicine: success, frustration, and hope. Cancers. 2019;11:1855. doi: 10.3390/cancers11121855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Ali Hazis NU, Aneja N, Rajabalaya R, David SR. Systematic patent review of nanoparticles in drug delivery and cancer therapy in the last decade. Recent Adv Drug Delivery Formulation. 2021;15:59–74. doi: 10.2174/1872211314666210521105534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Jain KK. An overview of drug delivery systems. Method Mol Biol. 2020;2059:1–54. doi: 10.1007/978-1-4939-9798-5_1. [DOI] [PubMed] [Google Scholar]
  • 17.Li Z, Tan S, Li S, Shen Q, Wang K. Cancer drug delivery in the nano era: an overview and perspectives (review) Oncol Rep. 2017;38:611–24. doi: 10.3892/or.2017.5718. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Patra JK, Das G, Fraceto LF, Campos EVR, Rodriguez-Torres MDP, Acosta-Torres LS, et al. Nano based drug delivery systems: recent developments and future prospects. J Nanobiotechnol. 2018;16:71. doi: 10.1186/s12951-018-0392-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Kim HY, Martin JH, McLachlan AJ, Boddy AV. Precision dosing of targeted anticancer drugs – challenges in the real world. Transl Cancer Res. 2017;6:S1500–11. doi: 10.21037/tcr.2017.10.30. [DOI] [Google Scholar]
  • 20.Adepu S, Ramakrishna S. Controlled drug delivery systems: current status and future directions. Molecules. 2021;26:5905. doi: 10.3390/molecules26195905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Laffleur F, Keckeis V. Advances in drug delivery systems: work in progress still needed? Int J Pharm. 2020;590:119912. doi: 10.1016/j.ijpharm.2020.119912. [DOI] [PubMed] [Google Scholar]
  • 22.Schirrmacher V. From chemotherapy to biological therapy: a review of novel concepts to reduce the side effects of systemic cancer treatment (review) Int J Oncol. 2019;54:407–19. doi: 10.3892/ijo.2018.4661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Patel H, Wu ZX, Chen Y, Bo L, Chen ZS. Drug resistance: from bacteria to cancer. Mol Biomed. 2021;2:27. doi: 10.1186/s43556-021-00041-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Guan X, Qin T, Qi T. Precision medicine in lung cancer theranostics: paving the way from traditional technology to advance era. Cancer Control. 2022;29 doi: 10.1177/10732748221077351. [DOI] [Google Scholar]
  • 25.Keek SA, Leijenaar RT, Jochems A, Woodruff HC. A review on radiomics and the future of theranostics for patient selection in precision medicine. Br J Radiol. 2018;91:20170926. doi: 10.1259/bjr.20170926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Ghasemiyeh P, Mohammadi-Samani S. Potential of nanoparticles as permeation enhancers and targeted delivery options for skin: advantages and disadvantages. Drug Des Dev Ther. 2020;14:3271–89. doi: 10.2147/dddt.s264648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Wu LP, Wang D, Li Z. Grand challenges in nanomedicine. Mater Sci Eng C. 2020;106:110302. doi: 10.1016/j.msec.2019.110302. [DOI] [PubMed] [Google Scholar]
  • 28.Shi J, Kantoff PW, Wooster R, Farokhzad OC. Cancer nanomedicine: progress, challenges and opportunities. Nat Rev Cancer. 2017;17:20–37. doi: 10.1038/nrc.2016.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Yu M, Wu J, Shi J, Farokhzad OC. Nanotechnology for protein delivery: overview and perspectives. J Controlled Release. 2016;240:24–37. doi: 10.1016/j.jconrel.2015.10.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Liechty WB, Caldorera-Moore M, Phillips MA, Schoener C, Peppas NA. Advanced molecular design of biopolymers for transmucosal and intracellular delivery of chemotherapeutic agents and biological therapeutics. J Controlled Release. 2011;155:119–27. doi: 10.1016/j.jconrel.2011.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Padera TP, Stoll BR, Tooredman JB, Capen D, di Tomaso E, Jain RK. Pathology: cancer cells compress intratumour vessels. Nature. 2004;427:695. doi: 10.1038/427695a. [DOI] [PubMed] [Google Scholar]
  • 32.Stylianopoulos T, Soteriou K, Fukumura D, Jain RK. Cationic nanoparticles have superior transvascular flux into solid tumors: insights from a mathematical model. Ann Biomed Eng. 2013;41:68–77. doi: 10.1007/s10439-012-0630-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Zamboni WC, Eiseman JL, Strychor S, Rice PM, Joseph E, Zamboni BA, et al. Tumor disposition of pegylated liposomal CKD-602 and the reticuloendothelial system in preclinical tumor models. J Liposome Res. 2011;21:70–80. doi: 10.3109/08982101003754385. [DOI] [PubMed] [Google Scholar]
  • 34.Daldrup-Link HE, Golovko D, Ruffell B, Denardo DG, Castaneda R, Ansari C, et al. MRI of tumor-associated macrophages with clinically applicable iron oxide nanoparticles. Clin Cancer Res. 2011;17:5695–704. doi: 10.1158/1078-0432.ccr-10-3420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Miller MA, Zheng YR, Gadde S, Pfirschke C, Zope H, Engblom C, et al. Tumour-associated macrophages act as a slow-release reservoir of nano-therapeutic Pt(IV) pro-drug. Nat Commun. 2015;6:8692. doi: 10.1038/ncomms9692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Miller MA, Gadde S, Pfirschke C, Engblom C, Sprachman MM, Kohler RH, et al. Predicting therapeutic nanomedicine efficacy using a companion magnetic resonance imaging nanoparticle. Sci Transl Med. 2015;7:314ra183. doi: 10.1126/scitranslmed.aac6522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Kalra AV, Spernyak J, Kim J, Sengooba A, Klinz S, Paz N, et al. Abstract 2065: magnetic resonance imaging with an iron oxide nanoparticle demonstrates the preclinical feasibility of predicting intratumoral uptake and activity of MM-398, a nanoliposomal irinotecan (nal-IRI) Cancer Res. 2014;74:2065. doi: 10.1158/1538-7445.am2014-2065. [DOI] [Google Scholar]
  • 38.Salvador-Morales C, Zhang L, Langer R, Farokhzad OC. Immunocompatibility properties of lipid-polymer hybrid nanoparticles with heterogeneous surface functional groups. Biomaterials. 2009;30:2231–40. doi: 10.1016/j.biomaterials.2009.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Mahmoudi M, Lynch I, Ejtehadi MR, Monopoli MP, Bombelli FB, Laurent S. Protein-nanoparticle interactions: opportunities and challenges. Chem Rev. 2011;111:5610–37. doi: 10.1021/cr100440g. [DOI] [PubMed] [Google Scholar]
  • 40.Cedervall T, Lynch I, Lindman S, Berggard T, Thulin E, Nilsson H, et al. Understanding the nanoparticle-protein corona using methods to quantify exchange rates and affinities of proteins for nanoparticles. Proc Natl Acad Sci USA. 2007;104:2050–5. doi: 10.1073/pnas.0608582104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Monopoli MP, Aberg C, Salvati A, Dawson KA. Biomolecular coronas provide the biological identity of nanosized materials. Nat Nanotechnol. 2012;7:779–86. doi: 10.1038/nnano.2012.207. [DOI] [PubMed] [Google Scholar]
  • 42.Ogawara KI, Furumoto K, Nagayama S, Minato K, Higaki K, Kai T, et al. Pre-coating with serum albumin reduces receptor-mediated hepatic disposition of polystyrene nanosphere: implications for rational design of nanoparticles. J Controlled Release. 2004;100:451–5. doi: 10.1016/j.jconrel.2004.07.028. [DOI] [PubMed] [Google Scholar]
  • 43.Walkey CD, Olsen JB, Guo H, Emili A, Chan WCW. Nanoparticle size and surface chemistry determine serum protein adsorption and macrophage uptake. J Am Chem Soc. 2012;134:2139–47. doi: 10.1021/ja2084338. [DOI] [PubMed] [Google Scholar]
  • 44.Salvati A, Pitek AS, Monopoli MP, Prapainop K, Bombelli FB, Hristov DR, et al. Transferrin-functionalized nanoparticles lose their targeting capabilities when a biomolecule corona adsorbs on the surface. Nat Nanotechnol. 2013;8:137–43. doi: 10.1038/nnano.2012.237. [DOI] [PubMed] [Google Scholar]
  • 45.Ritz S, Schöttler S, Kotman N, Baier G, Musyanovych A, Kuharev J, et al. Protein corona of nanoparticles: distinct proteins regulate the cellular uptake. Biomacromolecules. 2015;16:1311–21. doi: 10.1021/acs.biomac.5b00108. [DOI] [PubMed] [Google Scholar]
  • 46.Chanan-Khan A, Szebeni J, Savay S, Liebes L, Rafique NM, Alving CR, et al. Complement activation following first exposure to pegylated liposomal doxorubicin (Doxil): possible role in hypersensitivity reactions. Ann Oncol. 2003;14:1430–7. doi: 10.1093/annonc/mdg374. [DOI] [PubMed] [Google Scholar]
  • 47.Walkey CD, Chan WCW. Understanding and controlling the interaction of nanomaterials with proteins in a physiological environment. Chem Soc Rev. 2012;41:2780–99. doi: 10.1039/c1cs15233e. [DOI] [PubMed] [Google Scholar]
  • 48.Bigdeli A, Palchetti S, Pozzi D, Hormozi-Nezhad MR, Baldelli Bombelli F, Caracciolo G, et al. Exploring cellular interactions of liposomes using protein corona fingerprints and physicochemical properties. ACS Nano. 2016;10:3723–37. doi: 10.1021/acsnano.6b00261. [DOI] [PubMed] [Google Scholar]
  • 49.Hajipour MJ, Laurent S, Aghaie A, Rezaee F, Mahmoudi M. Personalized protein coronas: a “key” factor at the nanobiointerface. Biomater Sci. 2014;2:1210–21. doi: 10.1039/c4bm00131a. [DOI] [PubMed] [Google Scholar]
  • 50.Walkey CD, Olsen JB, Song F, Liu R, Guo H, Olsen DW, et al. Protein corona fingerprinting predicts the cellular interaction of gold and silver nanoparticles. ACS Nano. 2014;8:2439–55. doi: 10.1021/nn406018q. [DOI] [PubMed] [Google Scholar]
  • 51.Sakulkhu U, Maurizi L, Mahmoudi M, Motazacker M, Vries M, Gramoun A, et al. Ex situ evaluation of the composition of protein corona of intravenously injected superparamagnetic nanoparticles in rats. Nanoscale. 2014;6:11439–50. doi: 10.1039/c4nr02793k. [DOI] [PubMed] [Google Scholar]
  • 52.Owens DE, Peppas NA. Opsonization, biodistribution, and pharmacokinetics of polymeric nanoparticles. Int J Pharm. 2006;307:93–102. doi: 10.1016/j.ijpharm.2005.10.010. [DOI] [PubMed] [Google Scholar]
  • 53.Knop K, Hoogenboom R, Fischer D, Schubert US. Poly(ethylene glycol) in drug delivery: pros and cons as well as potential alternatives. Angew Chem Int Ed Engl. 2010;49:6288–308. doi: 10.1002/anie.200902672. [DOI] [PubMed] [Google Scholar]
  • 54.Gref R, Minamitake Y, Peracchia MT, Trubetskoy V, Torchilin V, Langer R. Biodegradable long-circulating polymeric nanospheres. Science. 1994;263:1600–3. doi: 10.1126/science.8128245. [DOI] [PubMed] [Google Scholar]
  • 55.Rodriguez PL, Harada T, Christian DA, Pantano DA, Tsai RK, Discher DE. Minimal “Self” peptides that inhibit phagocytic clearance and enhance delivery of nanoparticles. Science. 2013;339:971–5. doi: 10.1126/science.1229568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Hu CMJ, Fang RH, Wang KC, Luk BT, Thamphiwatana S, Dehaini D, et al. Nanoparticle biointerfacing by platelet membrane cloaking. Nature. 2015;526:118–21. doi: 10.1038/nature15373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Parodi A, Quattrocchi N, van de Ven AL, Chiappini C, Evangelopoulos M, Martinez JO, et al. Synthetic nanoparticles functionalized with biomimetic leukocyte membranes possess cell-like functions. Nat Nanotechnol. 2013;8:61–8. doi: 10.1038/nnano.2012.212. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Choi MR, Stanton-Maxey KJ, Stanley JK, Levin CS, Bardhan R, Akin D, et al. A cellular Trojan Horse for delivery of therapeutic nanoparticles into tumors. Nano Lett. 2007;7:3759–65. doi: 10.1021/nl072209h. [DOI] [PubMed] [Google Scholar]
  • 59.Alizadeh D, Zhang L, Hwang J, Schluep T, Badie B. Tumor-associated macrophages are predominant carriers of cyclodextrin-based nanoparticles into gliomas. Nanomedicine. 2010;6:382–90. doi: 10.1016/j.nano.2009.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Roger M, Clavreul A, Venier-Julienne MC, Passirani C, Sindji L, Schiller P, et al. Mesenchymal stem cells as cellular vehicles for delivery of nanoparticles to brain tumors. Biomaterials. 2010;31:8393–401. doi: 10.1016/j.biomaterials.2010.07.048. [DOI] [PubMed] [Google Scholar]
  • 61.Cheng H, Kastrup CJ, Ramanathan R, Siegwart DJ, Ma M, Bogatyrev SR, et al. Nanoparticulate cellular patches for cell-mediated tumoritropic delivery. ACS Nano. 2010;4:625–31. doi: 10.1021/nn901319y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Rudnick SI, Lou J, Shaller CC, Tang Y, Klein-Szanto AJP, Weiner LM, et al. Influence of affinity and antigen internalization on the uptake and penetration of Anti-HER2 antibodies in solid tumors. Cancer Res. 2011;71:2250–9. doi: 10.1158/0008-5472.can-10-2277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Jain RK, Stylianopoulos T. Delivering nanomedicine to solid tumors. Nat Rev Clin Oncol. 2010;7:653–64. doi: 10.1038/nrclinonc.2010.139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Tong R, Chiang HH, Kohane DS. Photoswitchable nanoparticles for in vivo cancer chemotherapy. Proc Natl Acad Sci USA. 2013;110:19048–53. doi: 10.1073/pnas.1315336110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Wong C, Stylianopoulos T, Cui J, Martin J, Chauhan VP, Jiang W, et al. Multistage nanoparticle delivery system for deep penetration into tumor tissue. Proc Natl Acad Sci USA. 2011;108:2426–31. doi: 10.1073/pnas.1018382108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Chauhan VP, Stylianopoulos T, Martin JD, Popović Z, Chen O, Kamoun WS, et al. Normalization of tumour blood vessels improves the delivery of nanomedicines in a size-dependent manner. Nat Nanotechnol. 2012;7:383–8. doi: 10.1038/nnano.2012.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Chauhan VP, Popović Z, Chen O, Cui J, Fukumura D, Bawendi MG, et al. Fluorescent nanorods and nanospheres for real-time in vivo probing of nanoparticle shape-dependent tumor penetration. Angew Chem Int Ed Engl. 2011;50:11417–20. doi: 10.1002/anie.201104449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Tasciotti E, Liu X, Bhavane R, Plant K, Leonard AD, Price BK, et al. Mesoporous silicon particles as a multistage delivery system for imaging and therapeutic applications. Nat Nanotechnol. 2008;3:151–7. doi: 10.1038/nnano.2008.34. [DOI] [PubMed] [Google Scholar]
  • 69.Farokhzad OC, Langer R. Impact of nanotechnology on drug delivery. ACS Nano. 2009;3:16–20. doi: 10.1021/nn900002m. [DOI] [PubMed] [Google Scholar]
  • 70.Desai N. Nanoparticle albumin bound (nab) technology: targeting tumors through the endothelial gp60 receptor and SPARC. Nanomed Nanotechnol Biol Med. 2007;3:339. doi: 10.1016/j.nano.2007.10.021. [DOI] [Google Scholar]
  • 71.Zhu X, Wu J, Shan W, Tao W, Zhao L, Lim JM, et al. Polymeric nanoparticles amenable to simultaneous installation of exterior targeting and interior therapeutic proteins. Angew Chem Int Ed Engl. 2016;55:3309–12. doi: 10.1002/anie.201509183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Pridgen EM, Alexis F, Kuo TT, Levy-Nissenbaum E, Karnik R, Blumberg RS, et al. Transepithelial transport of Fc-targeted nanoparticles by the neonatal Fc receptor for oral delivery. Sci Transl Med. 2013;5:213ra167. doi: 10.1126/scitranslmed.3007049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Cheng Y, Morshed RA, Auffinger B, Tobias AL, Lesniak MS. Multifunctional nanoparticles for brain tumor imaging and therapy. Adv Drug Delivery Rev. 2014;66:42–57. doi: 10.1016/j.addr.2013.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Gratton SEA, Ropp PA, Pohlhaus PD, Luft JC, Madden VJ, Napier ME, et al. The effect of particle design on cellular internalization pathways. Proc Natl Acad Sci USA. 2008;105:11613–8. doi: 10.1073/pnas.0801763105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Chithrani BD, Ghazani AA, Chan WCW. Determining the size and shape dependence of gold nanoparticle uptake into mammalian cells. Nano Lett. 2006;6:662–8. doi: 10.1021/nl052396o. [DOI] [PubMed] [Google Scholar]
  • 76.Gilleron J, Querbes W, Zeigerer A, Borodovsky A, Marsico G, Schubert U, et al. Image-based analysis of lipid nanoparticle-mediated siRNA delivery, intracellular trafficking and endosomal escape. Nat Biotechnol. 2013;31:638–46. doi: 10.1038/nbt.2612. [DOI] [PubMed] [Google Scholar]
  • 77.Kanasty R, Dorkin JR, Vegas A, Anderson D. Delivery materials for siRNA therapeutics. Nat Mater. 2013;12:967–77. doi: 10.1038/nmat3765. [DOI] [PubMed] [Google Scholar]
  • 78.Kalra AV, Kim J, Klinz SG, Paz N, Cain J, Drummond DC, et al. Preclinical activity of nanoliposomal irinotecan is governed by tumor deposition and intratumor prodrug conversion. Cancer Res. 2014;74:7003–13. doi: 10.1158/0008-5472.can-14-0572. [DOI] [PubMed] [Google Scholar]
  • 79.Iyer AK, Khaled G, Fang J, Maeda H. Exploiting the enhanced permeability and retention effect for tumor targeting. Drug Discovery Today. 2006;11:812–8. doi: 10.1016/j.drudis.2006.07.005. [DOI] [PubMed] [Google Scholar]
  • 80.Ozpolat B, Sood AK, Lopez-Berestein G. Liposomal siRNA nanocarriers for cancer therapy. Adv Drug Delivery Rev. 2014;66:110–6. doi: 10.1016/j.addr.2013.12.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Zuckerman JE, Davis ME. Clinical experiences with systemically administered siRNA-based therapeutics in cancer. Nat Rev Drug Discovery. 2015;14:843–56. doi: 10.1038/nrd4685. [DOI] [PubMed] [Google Scholar]
  • 82.Chen Y, Zhu X, Zhang X, Liu B, Huang L. Nanoparticles modified with tumor-targeting scFv deliver siRNA and miRNA for cancer therapy. Mol Ther. 2010;18:1650–6. doi: 10.1038/mt.2010.136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Sahay G, Querbes W, Alabi C, Eltoukhy A, Sarkar S, Zurenko C, et al. Efficiency of siRNA delivery by lipid nanoparticles is limited by endocytic recycling. Nat Biotechnol. 2013;31:653–8. doi: 10.1038/nbt.2614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Boddapati SV, D’Souza GGM, Erdogan S, Torchilin VP, Weissig V. Organelle-targeted nanocarriers: specific delivery of liposomal ceramide to mitochondria enhances its cytotoxicity in vitro and in vivo. Nano Lett. 2008;8:2559–63. doi: 10.1021/nl801908y. [DOI] [PubMed] [Google Scholar]
  • 85.Yameen B, Choi WI, Vilos C, Swami A, Shi J, Farokhzad OC. Insight into nanoparticle cellular uptake and intracellular targeting. J Controlled Release. 2014;190:485–99. doi: 10.1016/j.jconrel.2014.06.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Boddapati SV, Tongcharoensirikul P, Hanson RN, D’Souza GGM, Torchilin VP, Weissig V. Mitochondriotropic liposomes. J Liposome Res. 2005;15:49–58. doi: 10.1081/lpr-64958. [DOI] [PubMed] [Google Scholar]
  • 87.Marrache S, Dhar S. Engineering of blended nanoparticle platform for delivery of mitochondria-acting therapeutics. Proc Natl Acad Sci USA. 2012;109:16288–93. doi: 10.1073/pnas.1210096109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Cheng FY, Wang SPH, Su CH, Tsai TL, Wu PC, Shieh DB, et al. Stabilizer-free poly(lactide-co-glycolide) nanoparticles for multimodal biomedical probes. Biomaterials. 2008;29:2104–12. doi: 10.1016/j.biomaterials.2008.01.010. [DOI] [PubMed] [Google Scholar]
  • 89.Xu ZP, Niebert M, Porazik K, Walker TL, Cooper HM, Middelberg APJ, et al. Subcellular compartment targeting of layered double hydroxide nanoparticles. J Controlled Release. 2008;130:86–94. doi: 10.1016/j.jconrel.2008.05.021. [DOI] [PubMed] [Google Scholar]
  • 90.Park JH, von Maltzahn G, Zhang L, Schwartz MP, Ruoslahti E, Bhatia SN, et al. Magnetic iron oxide nanoworms for tumor targeting and imaging. Adv Mater. 2008;20:1630–5. doi: 10.1002/adma.200800004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Karnik R, Gu F, Basto P, Cannizzaro C, Dean L, Kyei-Manu W, et al. Microfluidic platform for controlled synthesis of polymeric nanoparticles. Nano Lett. 2008;8:2906–12. doi: 10.1021/nl801736q. [DOI] [PubMed] [Google Scholar]
  • 92.Kim Y, Lee Chung B, Ma M, Mulder WJM, Fayad ZA, Farokhzad OC, et al. Mass production and size control of lipid-polymer hybrid nanoparticles through controlled microvortices. Nano Lett. 2012;12:3587–91. doi: 10.1021/nl301253v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Rhee M, Valencia PM, Rodriguez MI, Langer R, Farokhzad OC, Karnik R. Synthesis of size-tunable polymeric nanoparticles enabled by 3D hydrodynamic flow focusing in single-layer microchannels. Adv Mater. 2011;23:H79–83. doi: 10.1002/adma.201004333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Chen D, Love KT, Chen Y, Eltoukhy AA, Kastrup C, Sahay G, et al. Rapid discovery of potent siRNA-containing lipid nanoparticles enabled by controlled microfluidic formulation. J Am Chem Soc. 2012;134:6948–51. doi: 10.1021/ja301621z. [DOI] [PubMed] [Google Scholar]
  • 95.Valencia PM, Basto PA, Zhang L, Rhee M, Langer R, Farokhzad OC, et al. Single-step assembly of homogenous lipid-polymeric and lipid-quantum dot nanoparticles enabled by microfluidic rapid mixing. ACS Nano. 2010;4:1671–9. doi: 10.1021/nn901433u. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Rolland JP, Maynor BW, Euliss LE, Exner AE, Denison GM, DeSimone JM. Direct fabrication and harvesting of monodisperse, shape-specific nanobiomaterials. J Am Chem Soc. 2005;127:10096–100. doi: 10.1021/ja051977c. [DOI] [PubMed] [Google Scholar]
  • 97.Bangham AD, Horne RW. Negative staining of phospholipids and their structural modification by surface-active agents as observed in the electron microscope. J Mol Biol. 1964;8:660–68. doi: 10.1016/s0022-2836(64)80115-7. [DOI] [PubMed] [Google Scholar]
  • 98.Xu J, Wong DHC, Byrne JD, Chen K, Bowerman C, DeSimone JM. Future of the particle replication in nonwetting templates (PRINT) technology. Angew Chem Int Ed Engl. 2013;52:6580–9. doi: 10.1002/anie.201209145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.He H, Liu L, Morin EE, Liu M, Schwendeman A. Survey of clinical translation of cancer nanomedicines-lessons learned from successes and failures. Acc Chem Res. 2019;52:2445–61. doi: 10.1021/acs.accounts.9b00228. [DOI] [PubMed] [Google Scholar]
  • 100.Huh D, Matthews BD, Mammoto A, Montoya-Zavala M, Hsin HY, Ingber DE. Reconstituting organ-level lung functions on a chip. Science. 2010;328:1662–8. doi: 10.1126/science.1188302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Albanese A, Lam AK, Sykes EA, Rocheleau JV, Chan WC. Tumour-on-a-chip provides an optical window into nanoparticle tissue transport. Nat Commun. 2013;4:2718. doi: 10.1038/ncomms3718. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Toh YC, Lim TC, Tai D, Xiao G, van Noort D, Yu H. A microfluidic 3D hepatocyte chip for drug toxicity testing. Lab Chip. 2009;9:2026–35. doi: 10.1039/b900912d. [DOI] [PubMed] [Google Scholar]
  • 103.Hrkach J, Von Hoff D, Ali MM, Andrianova E, Auer J, Campbell T, et al. Preclinical development and clinical translation of a PSMA-targeted docetaxel nanoparticle with a differentiated pharmacological profile. Sci Transl Med. 2012;4:128ra39. doi: 10.1126/scitranslmed.3003651. [DOI] [PubMed] [Google Scholar]
  • 104.Schultheis B, Strumberg D, Santel A, Vank C, Gebhardt F, Keil O, et al. First-in-human phase I study of the liposomal RNA interference therapeutic Atu027 in patients with advanced solid tumors. J Clin Oncol. 2014;32:4141–8. doi: 10.1200/jco.2013.55.0376. [DOI] [PubMed] [Google Scholar]
  • 105.Zuckerman JE, Gritli I, Tolcher A, Heidel JD, Lim D, Morgan R, et al. Correlating animal and human phase Ia/Ib clinical data with CALAA-01, a targeted, polymer-based nanoparticle containing siRNA. Proc Natl Acad Sci USA. 2014;111:11449–54. doi: 10.1073/pnas.1411393111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Eliasof S, Lazarus D, Peters CG, Case RI, Cole RO, Hwang J, et al. Correlating preclinical animal studies and human clinical trials of a multifunctional, polymeric nanoparticle. Proc Natl Acad Sci USA. 2013;110:15127–32. doi: 10.1073/pnas.1309566110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Meads MB, Gatenby RA, Dalton WS. Environment-mediated drug resistance: a major contributor to minimal residual disease. Nat Rev Cancer. 2009;9:665–74. doi: 10.1038/nrc2714. [DOI] [PubMed] [Google Scholar]
  • 108.Quail DF, Joyce JA. Microenvironmental regulation of tumor progression and metastasis. Nat Med. 2013;19:1423–37. doi: 10.1038/nm.3394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Joyce JA. Therapeutic targeting of the tumor microenvironment. Cancer Cell. 2005;7:513–20. doi: 10.1016/j.ccr.2005.05.024. [DOI] [PubMed] [Google Scholar]
  • 110.Hood JD, Bednarski M, Frausto R, Guccione S, Reisfeld RA, Xiang R, et al. Tumor regression by targeted gene delivery to the neovasculature. Science. 2002;296:2404–7. doi: 10.1126/science.1070200. [DOI] [PubMed] [Google Scholar]
  • 111.Murphy EA, Majeti BK, Barnes LA, Makale M, Weis SM, Lutu-Fuga K, et al. Nanoparticle-mediated drug delivery to tumor vasculature suppresses metastasis. Proc Natl Acad Sci USA. 2008;105:9343–8. doi: 10.1073/pnas.0803728105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Murakami M, Ernsting MJ, Undzys E, Holwell N, Foltz WD, Li SD. Docetaxel conjugate nanoparticles that target α-smooth muscle actin-expressing stromal cells suppress breast cancer metastasis. Cancer Res. 2013;73:4862–71. doi: 10.1158/0008-5472.can-13-0062. [DOI] [PubMed] [Google Scholar]
  • 113.Kreuter J. Nanoparticles – a historical perspective. Int J Pharm. 2007;331:1–10. doi: 10.1016/j.ijpharm.2006.10.021. [DOI] [PubMed] [Google Scholar]
  • 114.Vimbela GV, Ngo SM, Fraze C, Yang L, Stout DA. Antibacterial properties and toxicity from metallic nanomaterials. Int J Nanomed. 2017;12:3941–65. doi: 10.2147/ijn.s134526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Mody VV, Siwale R, Singh A, Mody HR. Introduction to metallic nanoparticles. J Pharm BioAllied Sci. 2010;2:282–9. doi: 10.4103/0975-7406.72127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Connor DM, Broome AM. Gold nanoparticles for the delivery of cancer therapeutics. Adv Cancer Res. 2018;139:163–84. doi: 10.1016/bs.acr.2018.05.001. [DOI] [PubMed] [Google Scholar]
  • 117.Zeng L, Gupta P, Chen Y, Wang E, Ji L, Chao H, et al. The development of anticancer ruthenium(II) complexes: from single molecule compounds to nanomaterials. Chem Soc Rev. 2017;46:5771–804. doi: 10.1039/c7cs00195a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Barnard AS. One-to-one comparison of sunscreen efficacy, aesthetics and potential nanotoxicity. Nat Nanotechnol. 2010;5:271–4. doi: 10.1038/nnano.2010.25. [DOI] [PubMed] [Google Scholar]
  • 119.Bhardwaj K, Dhanjal DS, Sharma A, Nepovimova E, Kalia A, Thakur S, et al. Conifer-derived metallic nanoparticles: green synthesis and biological applications. Int J Mol Sci. 2020;21:9028. doi: 10.3390/ijms21239028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Wei L, Lu J, Xu H, Patel A, Chen ZS, Chen G. Silver nanoparticles: synthesis, properties, and therapeutic applications. Drug Discovery Today. 2015;20:595–601. doi: 10.1016/j.drudis.2014.11.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Ge L, Li Q, Wang M, Ouyang J, Li X, Xing MMQ. Nanosilver particles in medical applications: synthesis, performance, and toxicity. Int J Nanomed. 2014;9:2399–407. doi: 10.2147/ijn.s55015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Khan MS, Alomari A, Tabrez S, Hassan I, Wahab R, Bhat SA, et al. Anticancer potential of biogenic silver nanoparticles: a mechanistic study. Pharmaceutics. 2021;13:707. doi: 10.3390/pharmaceutics13050707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Nicol JR, Dixon D, Coulter JA. Gold nanoparticle surface functionalization: a necessary requirement in the development of novel nanotherapeutics. Nanomedicine. 2015;10:1315–26. doi: 10.2217/nnm.14.219. [DOI] [PubMed] [Google Scholar]
  • 124.Lv J, Zhang L, Du W, Ling G, Zhang P. Functional gold nanoparticles for diagnosis, treatment and prevention of thrombus. J Controlled Release. 2022;345:572–85. doi: 10.1016/j.jconrel.2022.03.044. [DOI] [PubMed] [Google Scholar]
  • 125.Karathanasis E, Ghaghada KB. Crossing the barrier: treatment of brain tumors using nanochain particles. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2016;8:678–95. doi: 10.1002/wnan.1387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Liu XY, Wang JQ, Ashby CR, Zeng L, Fan YF, Chen ZS. Gold nanoparticles: synthesis, physiochemical properties and therapeutic applications in cancer. Drug Discovery Today. 2021;26:1284–92. doi: 10.1016/j.drudis.2021.01.030. [DOI] [PubMed] [Google Scholar]
  • 127.Torrisi L. Gold nanoparticles enhancing protontherapy efficiency. Recent Pat Nanotechnol. 2015;9:51–60. doi: 10.2174/1872210509999141222224121. [DOI] [PubMed] [Google Scholar]
  • 128.Kumthekar P, Rademaker A, Ko C, Dixit K, Schwartz MA, Sonabend AM, et al. A phase 0 first-in-human study using Nu-0129: a gold base spherical nucleic acid (SNA) nanoconjugate targeting BCL2L12 in recurrent glioblastoma patients. J Clin Oncol. 2019;37:3012. doi: 10.1200/jco.2019.37.15_suppl.3012. [DOI] [Google Scholar]
  • 129.Wang F, Wang YC, Dou S, Xiong MH, Sun TM, Wang J. Doxorubicin-tethered responsive gold nanoparticles facilitate intracellular drug delivery for overcoming multidrug resistance in cancer cells. ACS Nano. 2011;5:3679–92. doi: 10.1021/nn200007z. [DOI] [PubMed] [Google Scholar]
  • 130.Luo D, Wang X, Walker E, Springer S, Ramamurthy G, Burda C, et al. Targeted chemoradiotherapy of prostate cancer using gold nanoclusters with protease activatable monomethyl auristatin E. ACS Appl Mater Interfaces. 2022;14:14916–27. doi: 10.1021/acsami.1c23780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Kasinathan K, Marimuthu K, Murugesan B, Sathaiah M, Subramanian P, Sivakumar P, et al. Fabrication of eco-friendly chitosan functionalized few-layered WS2 nanocomposite implanted with ruthenium nanoparticles for in vitro antibacterial and anticancer activity: synthesis, characterization, and pharmaceutical applications. Int J Biol Macromol. 2021;190:520–32. doi: 10.1016/j.ijbiomac.2021.08.153. [DOI] [PubMed] [Google Scholar]
  • 132.Wang B, Wu S, Lin Z, Jiang Y, Chen Y, Chen ZS, et al. A personalized and long-acting local therapeutic platform combining photothermal therapy and chemotherapy for the treatment of multidrug-resistant colon tumor. Int J Nanomed. 2018;13:8411–27. doi: 10.2147/ijn.s184728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Xu M, Wen Y, Liu Y, Tan X, Chen X, Zhu X, et al. Hollow mesoporous ruthenium nanoparticles conjugated bispecific antibody for targeted anti-colorectal cancer response of combination therapy. Nanoscale. 2019;11:9661–78. doi: 10.1039/c9nr01904a. [DOI] [PubMed] [Google Scholar]
  • 134.Mahmud KM, Niloy MS, Shakil MS, Islam MA. Ruthenium complexes: an alternative to platinum drugs in colorectal cancer treatment. Pharmaceutics. 2021;13:1295. doi: 10.3390/pharmaceutics13081295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Li J, Zeng L, Wang Z, Chen H, Fang S, Wang J, et al. Cycloruthenated self-assembly with metabolic inhibition to efficiently overcome multidrug resistance in cancers. Adv Mater. 2022;34:e2100245. doi: 10.1002/adma.202100245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Lakshmi BA, Reddy AS, Sangubotla R, Hong JW, Kim S. Ruthenium(II)-curcumin liposome nanoparticles: synthesis, characterization, and their effects against cervical cancer. Colloids Surf B. 2021;204:111773. doi: 10.1016/j.colsurfb.2021.111773. [DOI] [PubMed] [Google Scholar]
  • 137.Minati L, Benetti F, Chiappini A, Speranza G. One-step synthesis of star-shaped gold nanoparticles. Colloids Surf A. 2014;441:623–8. doi: 10.1016/j.colsurfa.2013.10.025. [DOI] [Google Scholar]
  • 138.Bharadwaj KK, Rabha B, Pati S, Sarkar T, Choudhury BK, Barman A, et al. Green synthesis of gold nanoparticles using plant extracts as beneficial prospect for cancer theranostics. Molecules. 2021;26:6389. doi: 10.3390/molecules26216389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Paino IMM, Marangoni VS, de Oliveira RD, Antunes LMG, Zucolotto V. Cyto and genotoxicity of gold nanoparticles in human hepatocellular carcinoma and peripheral blood mononuclear cells. Toxicol Lett. 2012;215:119–25. doi: 10.1016/j.toxlet.2012.09.025. [DOI] [PubMed] [Google Scholar]
  • 140.Skrabalak SE, Chen J, Sun Y, Lu X, Au L, Cobley CM, et al. Gold nanocages: synthesis, properties, and applications. Acc Chem Res. 2008;41:1587–95. doi: 10.1021/ar800018v. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Basavegowda N, Lee YR. Synthesis of gold and silver nanoparticles using leaf extract of Perilla frutescens – a biogenic approach. J Nanosci Nanotechnol. 2014;14:4377–82. doi: 10.1166/jnn.2014.8646. [DOI] [PubMed] [Google Scholar]
  • 142.Bhullar S, Goyal N, Gupta S. Rapid green-synthesis of TiO2 nanoparticles for therapeutic applications. RSC Adv. 2021;11:30343–52. doi: 10.1039/d1ra05588g. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Burlacu E, Tanase C, Coman NA, Berta L. A review of bark-extract-mediated green synthesis of metallic nanoparticles and their applications. Molecules. 2019;24:4354. doi: 10.3390/molecules24234354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Guilger-Casagrande M, de Lima R. Synthesis of silver nanoparticles mediated by fungi: a review. Front Bioeng Biotechnol. 2019;7:287. doi: 10.3389/fbioe.2019.00287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Guilger-Casagrande M, Germano-Costa T, Pasquoto-Stigliani T, Fraceto LF, Lima R. Biosynthesis of silver nanoparticles employing Trichoderma harzianum with enzymatic stimulation for the control of Sclerotinia sclerotiorum . Sci Rep. 2019;9:14351. doi: 10.1038/s41598-019-50871-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Hassanisaadi M, Bonjar GHS, Rahdar A, Pandey S, Hosseinipour A, Abdolshahi R. Environmentally safe biosynthesis of gold nanoparticles using plant water extracts. Nanomaterials. 2021;11:2033. doi: 10.3390/nano11082033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Majoumouo MS, Sharma JR, Sibuyi NRS, Tincho MB, Boyom FF, Meyer M. Synthesis of biogenic gold nanoparticles from Terminalia mantaly extracts and the evaluation of their in vitro cytotoxic effects in cancer cells. Molecules. 2020;25:4469. doi: 10.3390/molecules25194469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Ponmurugan P. Biosynthesis of silver and gold nanoparticles using Trichoderma atroviride for the biological control of Phomopsis canker disease in tea plants. IET Nanobiotechnol. 2017;11:261–7. doi: 10.1049/iet-nbt.2016.0029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Lee SH, Jun BH. Silver nanoparticles: synthesis and application for nanomedicine. Int J Mol Sci. 2019;20:865. doi: 10.3390/ijms20040865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Abbasi E, Milani M, Fekri Aval S, Kouhi M, Akbarzadeh A, Tayefi Nasrabadi H, et al. Silver nanoparticles: synthesis methods, bio-applications and properties. Crit Rev Microbiol. 2016;42:173–80. doi: 10.3109/1040841x.2014.912200. [DOI] [PubMed] [Google Scholar]
  • 151.Li X, Xu H, Chen ZS, Chen G. Biosynthesis of nanoparticles by microorganisms and their applications. J Nanomater. 2011;2011:270974. doi: 10.1155/2011/270974. [DOI] [Google Scholar]
  • 152.Borase HP, Salunke BK, Salunkhe RB, Patil CD, Hallsworth JE, Kim BS, et al. Plant extract: a promising biomatrix for ecofriendly, controlled synthesis of silver nanoparticles. Appl Biochem Biotechnol. 2014;173:1–29. doi: 10.1007/s12010-014-0831-4. [DOI] [PubMed] [Google Scholar]
  • 153.Tanase C, Berta L, Coman NA, Roșca I, Man A, Toma F, et al. Investigation of in vitro antioxidant and antibacterial potential of silver nanoparticles obtained by biosynthesis using beech bark extract. Antioxidants. 2019;8:459. doi: 10.3390/antiox8100459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Ielo I, Rando G, Giacobello F, Sfameni S, Castellano A, Galletta M, et al. Synthesis, chemical-physical characterization, and biomedical applications of functional gold nanoparticles: a review. Molecules. 2021;26:5823. doi: 10.3390/molecules26195823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Si NT, Nhung NTA, Bui TQ, Nguyen MT, Nhat PV. Gold nanoclusters as prospective carriers and detectors of pramipexole. RSC Adv. 2021;11:16619–32. doi: 10.1039/d1ra02172a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Chenthamara D, Subramaniam S, Ramakrishnan SG, Krishnaswamy S, Essa MM, Lin FH, et al. Therapeutic efficacy of nanoparticles and routes of administration. Biomater Res. 2019;23:20. doi: 10.1186/s40824-019-0166-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Alphandéry E. Natural metallic nanoparticles for application in nano-oncology. Int J Mol Sci. 2020;21:E4412. doi: 10.3390/ijms21124412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Kim D, Jeong YY, Jon S. A drug-loaded aptamer-gold nanoparticle bioconjugate for combined CT imaging and therapy of prostate cancer. ACS Nano. 2010;4:3689–96. doi: 10.1021/nn901877h. [DOI] [PubMed] [Google Scholar]
  • 159.Nagaraj B, Musthafa SA, Muhammad S, Munuswamy-Ramanujam G, Chung WJ, Alodaini HA, et al. Anti-microbial and anti-cancer activity of gold nanoparticles phytofabricated using clerodin enriched clerodendrum ethanolic leaf extract. J King Saud Univ Sci. 2022;34:101989. doi: 10.1016/j.jksus.2022.101989. [DOI] [Google Scholar]
  • 160.Núñez C, Estévez SV, del Pilar Chantada M. Inorganic nanoparticles in diagnosis and treatment of breast cancer. JBIC J Biol Inorg Chem. 2018;23:331–45. doi: 10.1007/s00775-018-1542-z. [DOI] [PubMed] [Google Scholar]
  • 161.Dilnawaz F, Singh A, Mohanty C, Sahoo SK. Dual drug loaded superparamagnetic iron oxide nanoparticles for targeted cancer therapy. Biomaterials. 2010;31:3694–706. doi: 10.1016/j.biomaterials.2010.01.057. [DOI] [PubMed] [Google Scholar]
  • 162.Sztandera K, Gorzkiewicz M, Klajnert-Maculewicz B. Gold nanoparticles in cancer treatment. Mol Pharm. 2019;16:1–23. doi: 10.1021/acs.molpharmaceut.8b00810. [DOI] [PubMed] [Google Scholar]
  • 163.Murawala P, Tirmale A, Shiras A, Prasad BL. In situ synthesized BSA capped gold nanoparticles: effective carrier of anticancer drug methotrexate to MCF-7 breast cancer cells. Mater Sci Eng C. 2014;34:158–67. doi: 10.1016/j.msec.2013.09.004. [DOI] [PubMed] [Google Scholar]
  • 164.Schleh C, Semmler-Behnke M, Lipka J, Wenk A, Hirn S, Schäffler M, et al. Size and surface charge of gold nanoparticles determine absorption across intestinal barriers and accumulation in secondary target organs after oral administration. Nanotoxicology. 2012;6:36–46. doi: 10.3109/17435390.2011.552811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Hillyer JF, Albrecht RM. Gastrointestinal persorption and tissue distribution of differently sized colloidal gold nanoparticles. J Pharmaceut Sci. 2001;90:1927–36. doi: 10.1002/jps.1143. [DOI] [PubMed] [Google Scholar]
  • 166.Zhou Y, Peng Z, Seven ES, Leblanc RM. Crossing the blood-brain barrier with nanoparticles. J Controlled Release. 2018;270:290–303. doi: 10.1016/j.jconrel.2017.12.015. [DOI] [PubMed] [Google Scholar]
  • 167.Cai C, Wang M, Wang X, Wang B, Chen H, Chen J, et al. Transferrin adsorbed on PEGylated gold nanoparticles and its relevance to targeting specificity. J Nanosci Nanotechnol. 2018;18:5306–13. doi: 10.1166/jnn.2018.15435. [DOI] [PubMed] [Google Scholar]
  • 168.Zhang D, Qin X, Wu T, Qiao Q, Song Q, Zhang Z. Extracellular vesicles based self-grown gold nanopopcorn for combinatorial chemo-photothermal therapy. Biomaterials. 2019;197:220–8. doi: 10.1016/j.biomaterials.2019.01.024. [DOI] [PubMed] [Google Scholar]
  • 169.Chen G, Chen R, Zou C, Yang D, Chen ZS. Fragmented polymer nanotubes from sonication-induced scission with a thermo-responsive gating system for anti-cancer drug delivery. J Mater Chem B. 2014;2:1327–34. doi: 10.1039/c3tb21512a. [DOI] [PubMed] [Google Scholar]
  • 170.Libutti SK, Paciotti GF, Byrnes AA, Alexander HR, Gannon WE, Walker M, et al. Phase I and pharmacokinetic studies of CYT-6091, a novel PEGylated colloidal gold-rhTNF nanomedicine. Clin Cancer Res. 2010;16:6139–49. doi: 10.1158/1078-0432.ccr-10-0978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Joy P, Naduvilidam JK. Modified cyclodextrin coated magnetite nanoparticles for targeted delivery of hydrophobic drugs. . 2020. https://patentscope2.wipo.int/search/en/detail.jsf?docId=US293704527&_cid=JP2-LRK2V3-23560-1 Available from:
  • 172.Hahn SK, Li MY, Yang J, Jung HS. Liver targeted drug delivery systems using metal nanoparticles and preparing method thereof. . 2013. https://patentscope2.wipo.int/search/en/detail.jsf?docId=WO2013176468 Available from:
  • 173.Charalambos K, Grimm J. Delivery of therapeutic compounds with iron oxide nanoparticles patent. . 2014. https://patentscope2.wipo.int/search/en/detail.jsf?docId=WO2014047318&_cid=JP2-LRK2YA-27894-1 Available from:
  • 174.Jazrawi A, Pantiora E, Abdsaleh S, Bacovia DV, Eriksson S, Leonhardt H, et al. Magnetic-guided axillary ultrasound (MagUS) sentinel lymph node biopsy and mapping in patients with early breast cancer. A phase 2, single-arm prospective clinical trial. Cancers. 2021;13:4285. doi: 10.3390/cancers13174285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Bort G, Lux F, Dufort S, Crémillieux Y, Verry C, Tillement O. EPR-mediated tumor targeting using ultrasmall-hybrid nanoparticles: from animal to human with theranostic AGuIX nanoparticles. Theranostics. 2020;10:1319–31. doi: 10.7150/thno.37543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Ding S, Chen L, Liao J, Huo Q, Wang Q, Tian G, et al. Harnessing hafnium-based nanomaterials for cancer diagnosis and therapy. Small. 2023;19:e2300341. doi: 10.1002/smll.202300341. [DOI] [PubMed] [Google Scholar]
  • 177.Simonet S, Rodriguez-Lafrasse C, Beal D, Gerbaud S, Malesys C, Tillement O, et al. Gadolinium-based nanoparticles can overcome the radioresistance of head and neck squamous cell carcinoma through the induction of autophagy. J Biomed Nanotechnol. 2020;16:111–24. doi: 10.1166/jbn.2020.2871. [DOI] [PubMed] [Google Scholar]
  • 178.Zang X, Cheng M, Zhang X. Chen X. Quercetin nanoformulations: a promising strategy for tumor therapy. Food Funct. 2021;12:6664–81. doi: 10.1039/d1fo00851j. [DOI] [PubMed] [Google Scholar]
  • 179.Nanoparticle by SN bioscience announces phase 1 study results of SNB-101 (SN-38 nanoparticle anti-cancer drug) at ESMO. . 2023. prnewswire.com Available from:
  • 180.Creemers JHA, Pawlitzky I, Grosios K, Gileadi U, Middleton MR, Gerritsen WR, et al. Assessing the safety, tolerability and efficacy of PLGA-based immunomodulatory nanoparticles in patients with advanced NY-ESO-1-positive cancers: a first-in-human phase I open-label dose-escalation study protocol. BMJ Open. 2021;11:e050725. doi: 10.1136/bmjopen-2021-050725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Zhan C, Wang W, McAlvin JB, Guo S, Timko BP, Santamaria C, et al. Phototriggered local anesthesia. Nano Lett. 2016;16:177–81. doi: 10.1021/acs.nanolett.5b03440. [DOI] [PubMed] [Google Scholar]
  • 182.Telford BJ, Yahyanejad S, de Gunst T, den Boer HC, Vos RM, Stegink M, et al. Multi-modal effects of 1B3, a novel synthetic miR-193a-3p mimic, support strong potential for therapeutic intervention in oncology. Oncotarget. 2021;12:422–39. doi: 10.18632/oncotarget.27894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Kundu M, Majumder R, Das CK, Mandal M. Natural products based nanoformulations for cancer treatment: current evolution in Indian research. Biomed Mater. 2021;16 doi: 10.1088/1748-605x/abe8f2. Article Id: 044101. [DOI] [PubMed] [Google Scholar]
  • 184.Meerson A, Khatib S, Mahajna J. Natural products targeting cancer stem cells for augmenting cancer therapeutics. Int J Mol Sci. 2021;22:13044. doi: 10.3390/ijms222313044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Newman DJ, Cragg GM. Natural products as sources of new drugs over the nearly four decades from 01/1981 to 09/2019. J Nat Prod. 2020;83:770–803. doi: 10.1021/acs.jnatprod.9b01285. [DOI] [PubMed] [Google Scholar]
  • 186.Wang YJ, Huang Y, Anreddy N, Zhang GN, Zhang YK, Xie M, et al. Tea nanoparticle, a safe and biocompatible nanocarrier, greatly potentiates the anticancer activity of doxorubicin. Oncotarget. 2016;7:5877–91. doi: 10.18632/oncotarget.6711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Shim G, Kim MG, Kim D, Park JY, Oh YK. Nanoformulation-based sequential combination cancer therapy. Adv Drug Delivery Rev. 2017;115:57–81. doi: 10.1016/j.addr.2017.04.003. [DOI] [PubMed] [Google Scholar]
  • 188.Xing Y, Lu P, Xue Z, Liang C, Zhang B, Kebebe D, et al. Nano-strategies for improving the bioavailability of inhaled pharmaceutical formulations. Mini Rev Med Chem. 2020;20:1258–71. doi: 10.2174/1389557520666200509235945. [DOI] [PubMed] [Google Scholar]
  • 189.Anu Mary Ealia S, Saravanakumar MP. A review on the classification, characterisation, synthesis of nanoparticles and their application. IOP Conf Ser Mater Sci Eng. 2017;263:032019. doi: 10.1088/1757-899x/263/3/032019. [DOI] [Google Scholar]
  • 190.Jeevanandam J, Chan YS, Danquah MK. Nano-formulations of drugs: recent developments, impact and challenges. Biochimie. 2016;128/129:99. doi: 10.1016/j.biochi.2016.07.008. [DOI] [PubMed] [Google Scholar]
  • 191.Amin AR, Kucuk O, Khuri FR, Shin DM. Perspectives for cancer prevention with natural compounds. J Clin Oncol. 2009;27:2712–25. doi: 10.1200/jco.2008.20.6235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Hu Q, Luo Y. Chitosan-based nanocarriers for encapsulation and delivery of curcumin: a review. Int J Biol Macromol. 2021;179:125–35. doi: 10.1016/j.ijbiomac.2021.02.216. [DOI] [PubMed] [Google Scholar]
  • 193.Wong KE, Ngai SC, Chan KG, Lee LH, Goh BH, Chuah LH. Curcumin nanoformulations for colorectal cancer: a review. Front Pharmacol. 2019;10:152. doi: 10.3389/fphar.2019.00152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Ma Z, Wang N, He H, Tang X. Pharmaceutical strategies of improving oral systemic bioavailability of curcumin for clinical application. J Controlled Release. 2019;316:359–80. doi: 10.1016/j.jconrel.2019.10.053. [DOI] [PubMed] [Google Scholar]
  • 195.Pal K, Laha D, Parida PK, Roy S, Bardhan S, Dutta A, et al. An in vivo study for targeted delivery of curcumin in human triple negative breast carcinoma cells using biocompatible PLGA microspheres conjugated with folic acid. J Nanosci Nanotechnol. 2019;19:3720–33. doi: 10.1166/jnn.2019.16292. [DOI] [PubMed] [Google Scholar]
  • 196.Lai H, Ding X, Ye J, Deng J, Cui S. pH-responsive hyaluronic acid-based nanoparticles for targeted curcumin delivery and enhanced cancer therapy. Colloids Surf B. 2021;198:111455. doi: 10.1016/j.colsurfb.2020.111455. [DOI] [PubMed] [Google Scholar]
  • 197.Alvi SB, Appidi T, Deepak BP, Rajalakshmi PS, Minhas G, Singh SP, et al. The “nano to micro” transition of hydrophobic curcumin crystals leading to in situ adjuvant depots for Au-liposome nanoparticle mediated enhanced photothermal therapy. Biomater Sci. 2019;7:3866–75. doi: 10.1039/c9bm00932a. [DOI] [PubMed] [Google Scholar]
  • 198.Hu G, Batool Z, Cai Z, Liu Y, Ma M, Sheng L, et al. Production of self-assembling acylated ovalbumin nanogels as stable delivery vehicles for curcumin. Food Chem. 2021;355:129635. doi: 10.1016/j.foodchem.2021.129635. [DOI] [PubMed] [Google Scholar]
  • 199.Wang Z, Zhang RX, Zhang C, Dai C, Ju X, He R. Fabrication of stable and self-assembling rapeseed protein nanogel for hydrophobic curcumin delivery. J Agric Food Chem. 2019;67:887–94. doi: 10.1021/acs.jafc.8b05572. [DOI] [PubMed] [Google Scholar]
  • 200.Zhang L, Zhu W, Yang C, Guo H, Yu A, Ji J, et al. A novel folate-modified self-microemulsifying drug delivery system of curcumin for colon targeting. Int J Nanomed. 2012;7:151–62. doi: 10.2147/ijn.s27639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Mangalathillam S, Rejinold NS, Nair A, Lakshmanan VK, Nair SV, Jayakumar R. Curcumin loaded chitin nanogels for skin cancer treatment via the transdermal route. Nanoscale. 2012;4:239–50. doi: 10.1039/c1nr11271f. [DOI] [PubMed] [Google Scholar]
  • 202.Gou M, Men K, Shi H, Xiang M, Zhang J, Song J, et al. Curcumin-loaded biodegradable polymeric micelles for colon cancer therapy in vitro and in vivo. Nanoscale. 2011;3:1558–67. doi: 10.1039/c0nr00758g. [DOI] [PubMed] [Google Scholar]
  • 203.Tu Y. The discovery of artemisinin (qinghaosu) and gifts from Chinese medicine. Nat Med. 2011;17:1217–20. doi: 10.1038/nm.2471. [DOI] [PubMed] [Google Scholar]
  • 204.Charlie-Silva I, Fraceto LF, de Melo NFS. Progress in nano-drug delivery of artemisinin and its derivatives: towards to use in immunomodulatory approaches. Artif Cells Nanomed Biotechnol. 2018;46:S611–20. doi: 10.1080/21691401.2018.1505739. [DOI] [PubMed] [Google Scholar]
  • 205.Zhou X, Suo F, Haslinger K, Quax WJ. Artemisinin-type drugs in tumor cell death: mechanisms, combination treatment with biologics and nanoparticle delivery. Pharmaceutics. 2022;14:395. doi: 10.3390/pharmaceutics14020395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Zhang H, Li M, Zhu X, Zhang Z, Huang H, Hou L. Artemisinin co-delivery system based on manganese oxide for precise diagnosis and treatment of breast cancer. Nanotechnology. 2021;32 doi: 10.1088/1361-6528/abfc6f. Article Id: 325101. [DOI] [PubMed] [Google Scholar]
  • 207.Yu XA, Lu M, Luo Y, Hu Y, Zhang Y, Xu Z, et al. A cancer-specific activatable theranostic nanodrug for enhanced therapeutic efficacy via amplification of oxidative stress. Theranostics. 2020;10:371–83. doi: 10.7150/thno.39412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Halevas E, Mavroidi B, Kokotidou C, Mitraki A, Pelecanou M, Sagnou M. Advanced bis-MPA hyperbranched dendritic nanocarriers of artemisinin with anticancer potential. J Nanopart Res. 2021;23:135. doi: 10.1007/s11051-021-05250-0. [DOI] [Google Scholar]
  • 209.Nabil G, Bhise K, Sau S, Atef M, El-Banna HA, Iyer AK. Nano-engineered delivery systems for cancer imaging and therapy: recent advances, future direction and patent evaluation. Drug Discovery Today. 2019;24:462–91. doi: 10.1016/j.drudis.2018.08.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Xiao HJ, Vladimir S. Methods of making nanocrystals with enhanced biological availability and formulation for such nanocrystals preparation for use in anticancer. . https://patentscope2.wipo.int/search/en/detail.jsf?docId=WO2020088702 therapy; 2020. Available from:
  • 211.Reymond MA, Schneider M, Linder A, Sahoo R. Active substance delivery system. . 2020. https://patentscope2.wipo.int/search/en/detail.jsf?docId=WO2020249383&_cid=JP2-LRK3BR-41864-1 Available from:
  • 212.Miller A, Jedrzejczak WW. Albumin – biological functions and clinical significance. Postepy Hig Med Dosw. 2001;55:17–36. [PubMed] [Google Scholar]
  • 213.Zeeshan F, Madheswaran T, Panneerselvam J, Taliyan R, Kesharwani P. Human serum albumin as multifunctional nanocarrier for cancer therapy. J Pharmaceut Sci. 2021;110:3111–7. doi: 10.1016/j.xphs.2021.05.001. [DOI] [PubMed] [Google Scholar]
  • 214.Wartlick H, Michaelis K, Balthasar S, Strebhardt K, Kreuter J, Langer K. Highly specific HER2-mediated cellular uptake of antibody-modified nanoparticles in tumour cells. J Drug Target. 2004;12:461–71. doi: 10.1080/10611860400010697. [DOI] [PubMed] [Google Scholar]
  • 215.Wu H, Zhu L, Torchilin VP. pH-sensitive poly(histidine)-PEG/DSPE-PEG co-polymer micelles for cytosolic drug delivery. Biomaterials. 2013;34:1213–22. doi: 10.1016/j.biomaterials.2012.08.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Cheng Y, Zou T, Dai M, He XY, Peng N, Wu K, et al. Doxorubicin loaded tumor-triggered targeting ammonium bicarbonate liposomes for tumor-specific drug delivery. Colloids Surf B. 2019;178:263–8. doi: 10.1016/j.colsurfb.2019.03.002. [DOI] [PubMed] [Google Scholar]
  • 217.Dai M, Wu C, Fang HM, Li L, Yan JB, Zeng DL, et al. Thermo-responsive magnetic liposomes for hyperthermia-triggered local drug delivery. J Microencapsulation. 2017;34:408–15. doi: 10.1080/02652048.2017.1339738. [DOI] [PubMed] [Google Scholar]
  • 218.Luo J, Cheng Y, He XY, Liu Y, Peng N, Gong ZW, et al. Self-assembled CpG oligodeoxynucleotides conjugated hollow gold nanospheres to enhance cancer-associated immunostimulation. Colloids Surf B. 2019;175:248–55. doi: 10.1016/j.colsurfb.2018.12.001. [DOI] [PubMed] [Google Scholar]
  • 219.Luo J, Cheng Y, Gong ZW, Wu K, Zhou Y, Chen HX, et al. Self-assembled peptide functionalized gold nanopolyhedrons with excellent chiral optical properties. Langmuir. 2020;36:600–8. doi: 10.1021/acs.langmuir.9b03366. [DOI] [PubMed] [Google Scholar]
  • 220.Xia T, Lei C, Xu C, Peng N, Li Y, Yang XY, et al. Preparation and in vitro antitumor study of two-dimensional muscovite nanosheets. Langmuir. 2020;36:14268–75. doi: 10.1021/acs.langmuir.0c02393. [DOI] [PubMed] [Google Scholar]
  • 221.Rügheimer L, Hansell P, Wolgast M. Determination of the charge of the plasma proteins and consequent Donnan equilibrium across the capillary barriers in the rat microvasculature. Acta Physiol. 2008;194:335–9. doi: 10.1111/j.1748-1716.2008.01893.x. [DOI] [PubMed] [Google Scholar]
  • 222.Hobbs SK, Monsky WL, Yuan F, Roberts WG, Griffith L, Torchilin VP, et al. Regulation of transport pathways in tumor vessels: role of tumor type and microenvironment. Proc Natl Acad Sci USA. 1998;95:4607–12. doi: 10.1073/pnas.95.8.4607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Shen Z, Li Y, Kohama K, Oneill B, Bi J. Improved drug targeting of cancer cells by utilizing actively targetable folic acid-conjugated albumin nanospheres. Pharmacol Res. 2011;63:51–8. doi: 10.1016/j.phrs.2010.10.012. [DOI] [PubMed] [Google Scholar]
  • 224.Chacón M, Molpeceres J, Berges L, Guzmán M, Aberturas MR. Stability and freeze-drying of cyclosporine loaded poly(D,L lactide-glycolide) carriers. Eur J Pharm Sci. 1999;8:99–107. doi: 10.1016/s0928-0987(98)00066-9. [DOI] [PubMed] [Google Scholar]
  • 225.Quinlan GJ, Martin GS, Evans TW. Albumin: biochemical properties and therapeutic potential. Hepatology. 2005;41:1211–9. doi: 10.1002/hep.20720. [DOI] [PubMed] [Google Scholar]
  • 226.Elzoghby AO, Samy WM, Elgindy NA. Albumin-based nanoparticles as potential controlled release drug delivery systems. J Controlled Release. 2012;157:168–82. doi: 10.1016/j.jconrel.2011.07.031. [DOI] [PubMed] [Google Scholar]
  • 227.Mishra V, Heath RJ. Structural and biochemical features of human serum albumin essential for eukaryotic cell culture. Int J Mol Sci. 2021;22:8411. doi: 10.3390/ijms22168411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Solanki R, Rostamabadi H, Patel S, Jafari SM. Anticancer nano-delivery systems based on bovine serum albumin nanoparticles: a critical review. Int J Biol Macromol. 2021;193:528–40. doi: 10.1016/j.ijbiomac.2021.10.040. [DOI] [PubMed] [Google Scholar]
  • 229.Wang J, Zhang B. Bovine serum albumin as a versatile platform for cancer imaging and therapy. Curr Med Chem. 2018;25:2938–53. doi: 10.2174/0929867324666170314143335. [DOI] [PubMed] [Google Scholar]
  • 230.Lei C, Liu XR, Chen QB, Li Y, Zhou JL, Zhou LY, et al. Hyaluronic acid and albumin based nanoparticles for drug delivery. J Controlled Release. 2021;331:416–33. doi: 10.1016/j.jconrel.2021.01.033. [DOI] [PubMed] [Google Scholar]
  • 231.Jahanban-Esfahlan A, Dastmalchi S, Davaran S. A simple improved desolvation method for the rapid preparation of albumin nanoparticles. Int J Biol Macromol. 2016;91:703–9. doi: 10.1016/j.ijbiomac.2016.05.032. [DOI] [PubMed] [Google Scholar]
  • 232.Lammel AS, Hu X, Park SH, Kaplan DL, Scheibel TR. Controlling silk fibroin particle features for drug delivery. Biomaterials. 2010;31:4583–91. doi: 10.1016/j.biomaterials.2010.02.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Miele E, Spinelli GP, Miele E, Tomao F, Tomao S. Albumin-bound formulation of paclitaxel (Abraxane ABI-007) in the treatment of breast cancer. Int J Nanomed. 2009;4:99–105. doi: 10.2147/ijn.s3061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Tehrani SF, Bharadwaj P, Leblond Chain J, Roullin VG. Purification processes of polymeric nanoparticles: how to improve their clinical translation? J Controlled Release. 2023;360:591–612. doi: 10.1016/j.jconrel.2023.06.038. [DOI] [PubMed] [Google Scholar]
  • 235.DeFrates K, Markiewicz T, Gallo P, Rack A, Weyhmiller A, Jarmusik B, et al. Protein polymer-based nanoparticles: fabrication and medical applications. Int J Mol Sci. 2018;19:1717. doi: 10.3390/ijms19061717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.El-Eskandarany MS, Al-Hazza A, Al-Hajji LA, Ali N, Al-Duweesh AA, Banyan M, et al. Mechanical milling: a superior nanotechnological tool for fabrication of nanocrystalline and nanocomposite materials. Nanomaterials. 2021;11:2484. doi: 10.3390/nano11102484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Prasher A, Loynd CM, Tuten BT, Frank PG, Chao D, Berda EB. Efficient fabrication of polymer nanoparticles via sonogashira cross-linking of linear polymers in dilute solution. J Polym Sci Part A Polym Chem. 2016;54:209–17. doi: 10.1002/pola.27942. [DOI] [Google Scholar]
  • 238.Komlosh A, Weinstein V, Loupe P, Hasson T, Timan B, Konya A, et al. Physicochemical and biological examination of two glatiramer acetate products. Biomedicines. 2019;7:49. doi: 10.3390/biomedicines7030049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Kim MT, Chen Y, Marhoul J, Jacobson F. Statistical modeling of the drug load distribution on trastuzumab emtansine (Kadcyla), a lysine-linked antibody drug conjugate. Bioconjugate Chem. 2014;25:1223–32. doi: 10.1021/bc5000109. [DOI] [PubMed] [Google Scholar]
  • 240.Cortes J, Saura C. Nanoparticle albumin-bound (nab™)-paclitaxel: improving efficacy and tolerability by targeted drug delivery in metastatic breast cancer. EJC Suppl. 2010;8:1–10. doi: 10.1016/s1359-6349(10)70002-1. [DOI] [Google Scholar]
  • 241.Qi J, Yao P, He F, Yu C, Huang C. Nanoparticles with dextran/chitosan shell and BSA/chitosan core – doxorubicin loading and delivery. Int J Pharm. 2010;393:176–84. doi: 10.1016/j.ijpharm.2010.03.063. [DOI] [PubMed] [Google Scholar]
  • 242.Yang L, Cui F, Cun D, Tao A, Shi K, Lin W. Preparation, characterization and biodistribution of the lactone form of 10-hydroxycamptothecin (HCPT)-loaded bovine serum albumin (BSA) nanoparticles. Int J Pharm. 2007;340:163–72. doi: 10.1016/j.ijpharm.2007.03.028. [DOI] [PubMed] [Google Scholar]
  • 243.Zhao D, Zhao X, Zu Y, Li J, Zhang Y, Jiang R, et al. Preparation, characterization, and in vitro targeted delivery of folate-decorated paclitaxel-loaded bovine serum albumin nanoparticles. Int J Nanomed. 2010;5:669–77. doi: 10.2147/ijn.s12918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Kouchakzadeh H, Shojaosadati SA, Maghsoudi A, Vasheghani Farahani E. Optimization of PEGylation conditions for BSA nanoparticles using response surface methodology. AAPS PharmSciTech. 2010;11:1206–11. doi: 10.1208/s12249-010-9487-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Ulbrich K, Hekmatara T, Herbert E, Kreuter J. Transferrin- and transferrin-receptor-antibody-modified nanoparticles enable drug delivery across the blood-brain barrier (BBB) Eur J Pharm Biopharm. 2009;71:251–6. doi: 10.1016/j.ejpb.2008.08.021. [DOI] [PubMed] [Google Scholar]
  • 246.Li FQ, Su H, Wang J, Liu JY, Zhu QG, Fei YB, et al. Preparation and characterization of sodium ferulate entrapped bovine serum albumin nanoparticles for liver targeting. Int J Pharm. 2008;349:274–82. doi: 10.1016/j.ijpharm.2007.08.001. [DOI] [PubMed] [Google Scholar]
  • 247.Zensi A, Begley D, Pontikis C, Legros C, Mihoreanu L, Wagner S, et al. Albumin nanoparticles targeted with Apo E enter the CNS by transcytosis and are delivered to neurones. J Controlled Release. 2009;137:78–86. doi: 10.1016/j.jconrel.2009.03.002. [DOI] [PubMed] [Google Scholar]
  • 248.Mishra V, Mahor S, Rawat A, Gupta PN, Dubey P, Khatri K, et al. Targeted brain delivery of AZT via transferrin anchored pegylated albumin nanoparticles. J Drug Target. 2006;14:45–53. doi: 10.1080/10611860600612953. [DOI] [PubMed] [Google Scholar]
  • 249.Ghosh B, Biswas S. Polymeric micelles in cancer therapy: state of the art. J Controlled Release. 2021;332:127–47. doi: 10.1016/j.jconrel.2021.02.016. [DOI] [PubMed] [Google Scholar]
  • 250.Ghezzi M, Pescina S, Padula C, Santi P, Del Favero E, Cantù L, et al. Polymeric micelles in drug delivery: an insight of the techniques for their characterization and assessment in biorelevant conditions. J Controlled Release. 2021;332:312–36. doi: 10.1016/j.jconrel.2021.02.031. [DOI] [PubMed] [Google Scholar]
  • 251.Sawant RR, Torchilin VP. Multifunctionality of lipid-core micelles for drug delivery and tumour targeting. Mol Membr Biol. 2010;27:232–46. doi: 10.3109/09687688.2010.516276. [DOI] [PubMed] [Google Scholar]
  • 252.Needham D, Stoicheva N, Zhelev DV. Exchange of monooleoylphosphatidylcholine as monomer and micelle with membranes containing poly(ethylene glycol)-lipid. Biophys J. 1997;73:2615–29. doi: 10.1016/s0006-3495(97)78291-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Torchilin VP, Levchenko TS, Lukyanov AN, Khaw BA, Klibanov AL, Rammohan R, et al. P-nitrophenylcarbonyl-PEG-PE-liposomes: fast and simple attachment of specific ligands, including monoclonal antibodies, to distal ends of PEG chains via p-nitrophenylcarbonyl groups. Biochim Biophys Acta. 2001;1511:397–411. doi: 10.1016/s0005-2728(01)00165-7. [DOI] [PubMed] [Google Scholar]
  • 254.Wang J, Mongayt D, Torchilin VP. Polymeric micelles for delivery of poorly soluble drugs: preparation and anticancer activity in vitro of paclitaxel incorporated into mixed micelles based on poly(ethylene glycol)-lipid conjugate and positively charged lipids. J Drug Target. 2005;13:73–80. doi: 10.1080/10611860400011935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Dabholkar RD, Sawant RM, Mongayt DA, Devarajan PV, Torchilin VP. Polyethylene glycol–phosphatidylethanolamine conjugate (PEG–PE)-based mixed micelles: some properties, loading with paclitaxel, and modulation of P-glycoprotein-mediated efflux. Int J Pharm. 2006;315:148–57. doi: 10.1016/j.ijpharm.2006.02.018. [DOI] [PubMed] [Google Scholar]
  • 256.Sawant RR, Sawant RM, Torchilin VP. Mixed PEG-PE/vitamin E tumor-targeted immunomicelles as carriers for poorly soluble anti-cancer drugs: improved drug solubilization and enhanced in vitro cytotoxicity. Eur J Pharm Biopharm. 2008;70:51–7. doi: 10.1016/j.ejpb.2008.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Hong W, Chen D, Zhang X, Zeng J, Hu H, Zhao X, et al. Reversing multidrug resistance by intracellular delivery of Pluronic® P85 unimers. Biomaterials. 2013;34:9602–14. doi: 10.1016/j.biomaterials.2013.08.032. [DOI] [PubMed] [Google Scholar]
  • 258.Wei Z, Yuan S, Hao J, Fang X. Mechanism of inhibition of P-glycoprotein mediated efflux by Pluronic P123/F127 block copolymers: relationship between copolymer concentration and inhibitory activity. Eur J Pharm Biopharm. 2013;83:266–74. doi: 10.1016/j.ejpb.2012.09.014. [DOI] [PubMed] [Google Scholar]
  • 259.Gupta P, Jani KA, Yang DH, Sadoqi M, Squillante E, Chen ZS. Revisiting the role of nanoparticles as modulators of drug resistance and metabolism in cancer. Expert Opin Drug Metab Toxicol. 2016;12:281–9. doi: 10.1517/17425255.2016.1145655. [DOI] [PubMed] [Google Scholar]
  • 260.Chen Y, Zhang W, Huang Y, Gao F, Sha X, Fang X. Pluronic-based functional polymeric mixed micelles for co-delivery of doxorubicin and paclitaxel to multidrug resistant tumor. Int J Pharm. 2015;488:44–58. doi: 10.1016/j.ijpharm.2015.04.048. [DOI] [PubMed] [Google Scholar]
  • 261.Zhang W, Shi Y, Chen Y, Ye J, Sha X, Fang X. Multifunctional Pluronic P123/F127 mixed polymeric micelles loaded with paclitaxel for the treatment of multidrug resistant tumors. Biomaterials. 2011;32:2894–906. doi: 10.1016/j.biomaterials.2010.12.039. [DOI] [PubMed] [Google Scholar]
  • 262.Sahu A, Kasoju N, Goswami P, Bora U. Encapsulation of curcumin in Pluronic block copolymer micelles for drug delivery applications. J Biomater Appl. 2011;25:619–39. doi: 10.1177/0885328209357110. [DOI] [PubMed] [Google Scholar]
  • 263.Cai Y, Sun Z, Fang X, Fang X, Xiao F, Wang Y, et al. Synthesis, characterization and anti-cancer activity of Pluronic F68-curcumin conjugate micelles. Drug Delivery. 2016;23:2587–95. doi: 10.3109/10717544.2015.1037970. [DOI] [PubMed] [Google Scholar]
  • 264.Carlson LJ, Cote B, Alani AW, Rao DA. Polymeric micellar co-delivery of resveratrol and curcumin to mitigate in vitro doxorubicin-induced cardiotoxicity. J Pharmaceut Sci. 2014;103:2315–22. doi: 10.1002/jps.24042. [DOI] [PubMed] [Google Scholar]
  • 265.Du B, Zhu W, Yu L, Wang Y, Zheng M, Huang J, et al. TPGS2k-PLGA composite nanoparticles by depleting lipid rafts in colon cancer cells for overcoming drug resistance. Nanomed Nanotechnol Biol Med. 2021;35:102307. doi: 10.1016/j.nano.2020.102307. [DOI] [PubMed] [Google Scholar]
  • 266.Salmanpour M, Tamaddon A, Yousefi G, Mohammadi-Samani S. “Grafting-from” synthesis and characterization of poly (2-ethyl-2-oxazoline)-b-poly (benzyl L-glutamate) micellar nanoparticles for potential biomedical applications. Bioimpacts. 2017;7:155–66. doi: 10.15171/bi.2017.19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Dai S, Feng Y, Li S, Chen Y, Liu M, Zhang C, et al. Stereo complexation assisted assembly of poly(γ-glutamic acid)-graft-polylactide nano-micelles and their efficacy as anticancer drug carrier. Anti Cancer Agents Med Chem. 2018;18:302–11. doi: 10.2174/1871520617666170911170104. [DOI] [PubMed] [Google Scholar]
  • 268.Chen Y, Zhang L, Liu Y, Tan S, Qu R, Wu Z, et al. Preparation of PGA-PAE-micelles for enhanced antitumor efficacy of cisplatin. ACS Appl Mater Interfaces. 2018;10:25006–16. doi: 10.1021/acsami.8b04259. [DOI] [PubMed] [Google Scholar]
  • 269.Ahmad Z, Tang Z, Shah A, Lv S, Zhang D, Zhang Y, et al. Cisplatin loaded methoxy poly (ethylene glycol)-block-poly (L-glutamic acid-co-L-phenylalanine) nanoparticles against human breast cancer cell. Macromol Biosci. 2014;14:1337–45. doi: 10.1002/mabi.201400109. [DOI] [PubMed] [Google Scholar]
  • 270.Thipparaboina R, Chavan RB, Kumar D, Modugula S, Shastri NR. Micellar carriers for the delivery of multiple therapeutic agents. Colloids Surf B. 2015;135:291–308. doi: 10.1016/j.colsurfb.2015.07.046. [DOI] [PubMed] [Google Scholar]
  • 271.Aliabadi HM, Lavasanifar A. Polymeric micelles for drug delivery. Expert Opin Drug Delivery. 2006;3:139–62. doi: 10.1517/17425247.3.1.139. [DOI] [PubMed] [Google Scholar]
  • 272.Makhmalzade BS, Chavoshy F. Polymeric micelles as cutaneous drug delivery system in normal skin and dermatological disorders. J Adv Pharm Technol Res. 2018;9:2–8. doi: 10.4103/japtr.japtr_314_17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Gaucher G, Dufresne MH, Sant VP, Kang N, Maysinger D, Leroux JC. Block copolymer micelles: preparation, characterization and application in drug delivery. J Controlled Release. 2005;109:169–88. doi: 10.1016/j.jconrel.2005.09.034. [DOI] [PubMed] [Google Scholar]
  • 274.Nasrollahzadeh M, Atarod M, Sajjadi M, Sajadi SM, Issaabadi Z. Interface science and technology. Amsterdam: Elsevier; 2019. Plant-mediated green synthesis of nanostructures: mechanisms, characterization, and applications; pp. 199–322. [Google Scholar]
  • 275.Valerii MC, Benaglia M, Caggiano C, Papi A, Strillacci A, Lazzarini G, et al. Drug delivery by polymeric micelles: an in vitro and in vivo study to deliver lipophilic substances to colonocytes and selectively target inflamed colon. Nanomedicine. 2013;9:675–85. doi: 10.1016/j.nano.2012.11.007. [DOI] [PubMed] [Google Scholar]
  • 276.Wu DQ, Lu B, Chang C, Chen CS, Wang T, Zhang YY, et al. Galactosylated fluorescent labeled micelles as a liver targeting drug carrier. Biomaterials. 2009;30:1363–71. doi: 10.1016/j.biomaterials.2008.11.027. [DOI] [PubMed] [Google Scholar]
  • 277.Bhattacharjee S. DLS and Zeta potential – what they are and what they are not? J Controlled Release. 2016;235:337–51. doi: 10.1016/j.jconrel.2016.06.017. [DOI] [PubMed] [Google Scholar]
  • 278.Jalili N, Laxminarayana K. A review of atomic force microscopy imaging systems: application to molecular metrology and biological sciences. Mechatronics. 2004;14:907–45. doi: 10.1016/j.mechatronics.2004.04.005. [DOI] [Google Scholar]
  • 279.Milne JLS, Borgnia MJ, Bartesaghi A, Tran EEH, Earl LA, Schauder DM, et al. Cryo-electron microscopy – a primer for the non-microscopist. FEBS J. 2013;280:28–45. doi: 10.1111/febs.12078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 280.di Cola E, Grillo I, Ristori S. Small angle X-ray and neutron scattering: powerful tools for studying the structure of drug-loaded liposomes. Pharmaceutics. 2016;8:10. doi: 10.3390/pharmaceutics8020010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.Schilt Y, Berman T, Wei X, Barenholz Y, Raviv U. Using solution X-ray scattering to determine the high-resolution structure and morphology of PEGylated liposomal doxorubicin nanodrugs. Biochim Biophys Acta. 2016;1860:108–19. doi: 10.1016/j.bbagen.2015.09.012. [DOI] [PubMed] [Google Scholar]
  • 282.Dong YD, Boyd BJ. Applications of X-ray scattering in pharmaceutical science. Int J Pharm. 2011;417:101–11. doi: 10.1016/j.ijpharm.2011.01.022. [DOI] [PubMed] [Google Scholar]
  • 283.Dai X, Yue Z, Eccleston ME, Swartling J, Slater NKH, Kaminski CF. Fluorescence intensity and lifetime imaging of free and micellar-encapsulated doxorubicin in living cells. Nanomed Nanotechnol Biol Med. 2008;4:49–56. doi: 10.1016/j.nano.2007.12.002. [DOI] [PubMed] [Google Scholar]
  • 284.Ma PA, Liu S, Huang Y, Chen X, Zhang L, Jing X. Lactose mediated liver-targeting effect observed by ex vivo imaging technology. Biomaterials. 2010;31:2646–54. doi: 10.1016/j.biomaterials.2009.12.019. [DOI] [PubMed] [Google Scholar]
  • 285.Susa M, Iyer AK, Ryu K, Hornicek FJ, Mankin H, Amiji MM, et al. Doxorubicin loaded polymeric nanoparticulate delivery system to overcome drug resistance in osteosarcoma. BMC Cancer. 2009;9:399. doi: 10.1186/1471-2407-9-399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Miller T, Breyer S, van Colen G, Mier W, Haberkorn U, Geissler S, et al. Premature drug release of polymeric micelles and its effects on tumor targeting. Int J Pharm. 2013;445:117–24. doi: 10.1016/j.ijpharm.2013.01.059. [DOI] [PubMed] [Google Scholar]
  • 287.Yin Q, Shen J, Zhang Z, Yu H, Li Y. Reversal of multidrug resistance by stimuli-responsive drug delivery systems for therapy of tumor. Adv Drug Delivery Rev. 2013;65:1699–715. doi: 10.1016/j.addr.2013.04.011. [DOI] [PubMed] [Google Scholar]
  • 288.Imran M, Shah MR, Shafiullah . Amphiphilic block copolymers–based micelles for drug delivery. In: Grumezescu AM, editor. Design and development of new nanocarriers. New York, USA: William Andrew Publishing; 2018. pp. 365–400. [Google Scholar]
  • 289.de Andrade DF, Zuglianello C, Pohlmann AR, Guterres SS, Beck RCR. Assessing the in vitro drug release from lipid-core nanocapsules: a new strategy combining dialysis Sac and a continuous-flow system. AAPS PharmSciTech. 2015;16:1409–17. doi: 10.1208/s12249-015-0330-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.D’Souza S. A review of in vitro drug release test methods for nano-sized dosage forms. Adv Pharm. 2014;2014:304757. doi: 10.1155/2014/304757. [DOI] [Google Scholar]
  • 291.Gupta V, Trivedi P. In vitro and in vivo characterization of pharmaceutical topical nanocarriers containing anticancer drugs for skin cancer treatment. In: Grumezescu AM, editor. Lipid nanocarriers for drug targeting. Amsterdam, Netherlands: Elsevier; 2018. pp. 563–627. [Google Scholar]
  • 292.Daraee H, Etemadi A, Kouhi M, Alimirzalu S, Akbarzadeh A. Application of liposomes in medicine and drug delivery. Artif Cells Nanomed Biotechnol. 2016;44:381–91. doi: 10.3109/21691401.2014.953633. [DOI] [PubMed] [Google Scholar]
  • 293.Immordino ML, Dosio F, Cattel L. Stealth liposomes: review of the basic science, rationale, and clinical applications, existing and potential. Int J Nanomed. 2006;1:297–315. [PMC free article] [PubMed] [Google Scholar]
  • 294.Guimarães D, Cavaco-Paulo A, Nogueira E. Design of liposomes as drug delivery system for therapeutic applications. Int J Pharm. 2021;601:120571. doi: 10.1016/j.ijpharm.2021.120571. [DOI] [PubMed] [Google Scholar]
  • 295.Shah S, Dhawan V, Holm R, Nagarsenker MS, Perrie Y. Liposomes: advancements and innovation in the manufacturing process. Adv Drug Delivery Rev. 2020;154/155:102–22. doi: 10.1016/j.addr.2020.07.002. [DOI] [PubMed] [Google Scholar]
  • 296.Akbarzadeh A, Rezaei-Sadabady R, Davaran S, Joo SW, Zarghami N, Hanifehpour Y, et al. Liposome: classification, preparation, and applications. Nanoscale Res Lett. 2013;8:102. doi: 10.1186/1556-276x-8-102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 297.Lee Y, Thompson DH. Stimuli-responsive liposomes for drug delivery. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2017;9:e1450. doi: 10.1002/wnan.1450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 298.Patel A, Saraswat A, Patel H, Chen ZS, Patel K. Palmitoyl carnitine-anchored nanoliposomes for neovasculature-specific delivery of gemcitabine elaidate to treat pancreatic cancer. Cancers. 2022;15:182. doi: 10.3390/cancers15010182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 299.Dinakar YH, Karole A, Parvez S, Jain V, Mudavath SL. Folate receptor targeted NIR cleavable liposomal delivery system augment penetration and therapeutic efficacy in breast cancer. Biochim Biophys Acta Gen Subj. 2023;1867:130396. doi: 10.1016/j.bbagen.2023.130396. [DOI] [PubMed] [Google Scholar]
  • 300.Ren X, Zhang L, Zhang Y, Li Z, Siemers N, Zhang Z. Insights gained from single-cell analysis of immune cells in the tumor microenvironment. Annu Rev Immunol. 2021;39:583–609. doi: 10.1146/annurev-immunol-110519-071134. [DOI] [PubMed] [Google Scholar]
  • 301.Assil KK, Weinreb RN. Multivesicular liposomes. Sustained release of the antimetabolite cytarabine in the eye. Arch Ophthalmol. 1987;105:400–3. doi: 10.1001/archopht.1987.01060030120040. [DOI] [PubMed] [Google Scholar]
  • 302.Salim M, Minamikawa H, Sugimura A, Hashim R. Amphiphilic designer nano-carriers for controlled release: from drug delivery to diagnostics. Med Chem Commun. 2014;5:1602–18. doi: 10.1039/c4md00085d. [DOI] [Google Scholar]
  • 303.Zhang H. Thin-film hydration followed by extrusion method for liposome preparation. Methods Mol Biol. 2017;1522:17–22. doi: 10.1007/978-1-4939-6591-5_2. [DOI] [PubMed] [Google Scholar]
  • 304.Torchilin V, Weisslg V, editors. Liposomes: a practical approach. Oxford, England: Oxford University Press; 2003. [Google Scholar]
  • 305.Ghosh Dastidar D, Chakrabarti G. Thermoresponsive drug delivery systems, characterization and application. In: Mohapatra SS, Ranjan S, Dasgupta N, Mishra RK, Thomas S, editors. Applications of targeted nano drugs and delivery systems. Amsterdam, Netherlands: Elsevier; 2019. pp. 133–55. [Google Scholar]
  • 306.Mayer LD, Bally MB, Hope MJ, Cullis PR. Techniques for encapsulating bioactive agents into liposomes. Chem Phys Lipids. 1986;40:333–45. doi: 10.1016/0009-3084(86)90077-0. [DOI] [PubMed] [Google Scholar]
  • 307.Gruner SM, Lenk RP, Janoff AS, Ostro MJ. Novel multilayered lipid vesicles: comparison of physical characteristics of multilamellar liposomes and stable plurilamellar vesicles. Biochemistry. 1985;24:2833–42. doi: 10.1021/bi00333a004. [DOI] [PubMed] [Google Scholar]
  • 308.Charcosset C, Juban A, Valour JP, Urbaniak S, Fessi H. Preparation of liposomes at large scale using the ethanol injection method: effect of scale-up and injection devices. Chem Eng Res Des. 2015;94:508–15. doi: 10.1016/j.cherd.2014.09.008. [DOI] [Google Scholar]
  • 309.Pons M, Foradada M, Estelrich J. Liposomes obtained by the ethanol injection method. Int J Pharm. 1993;95:51–6. doi: 10.1016/0378-5173(93)90389-w. [DOI] [Google Scholar]
  • 310.Jaafar-Maalej C, Diab R, Andrieu V, Elaissari A, Fessi H. Ethanol injection method for hydrophilic and lipophilic drug-loaded liposome preparation. J Liposome Res. 2010;20:228–43. doi: 10.3109/08982100903347923. [DOI] [PubMed] [Google Scholar]
  • 311.Szoka F, Papahadjopoulos D. Procedure for preparation of liposomes with large internal aqueous space and high capture by reverse-phase evaporation. Proc Natl Acad Sci USA. 1978;75:4194–8. doi: 10.1073/pnas.75.9.4194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312.Brandl M. Liposomes as drug carriers: a technological approach. Biotechnol Annu Rev. 2001;7:59–85. doi: 10.1016/s1387-2656(01)07033-8. [DOI] [PubMed] [Google Scholar]
  • 313.Kataria S, Sandhu P, Bilandi A, Akanksha M, Kapoor B. Stealth liposomes: a review. Int J Res Ayurveda Pharm. 2011;2:1534–8. [Google Scholar]
  • 314.Pattni BS, Chupin VV, Torchilin VP. New developments in liposomal drug delivery. Chem Rev. 2015;115:10938–66. doi: 10.1021/acs.chemrev.5b00046. [DOI] [PubMed] [Google Scholar]
  • 315.Maherani B, Arab-Tehrany E, Mozafari MR, Gaiani C, Linder M. Liposomes: a review of manufacturing techniques and targeting strategies. Curr Nanosci. 2011;7:436–52. doi: 10.2174/157341311795542453. [DOI] [Google Scholar]
  • 316.Pidgeon C, McNeely S, Schmidt T, Johnson JE. Multilayered vesicles prepared by reverse-phase evaporation: liposome structure and optimum solute entrapment. Biochemistry. 1987;26:17–29. doi: 10.1021/bi00375a004. [DOI] [PubMed] [Google Scholar]
  • 317.Chen G, Li D, Jin Y, Zhang W, Teng L, Bunt C, et al. Deformable liposomes by reverse-phase evaporation method for an enhanced skin delivery of (+)-catechin. Drug Dev Ind Pharm. 2014;40:260–5. doi: 10.3109/03639045.2012.756512. [DOI] [PubMed] [Google Scholar]
  • 318.Anderson M, Omri A. The effect of different lipid components on the in vitro stability and release kinetics of liposome formulations. Drug Deliv. 2004;11:33–9. doi: 10.1080/10717540490265243. [DOI] [PubMed] [Google Scholar]
  • 319. ICHH Guideline. . https://database.ich.org/sites/default/files/ICH_Q3C-R8_Guideline_Step4_2021_0422_1.pdf Impurities: guideline for residual solvents Q3C(R8); 2021. Available from:
  • 320.Jensen GM. The care and feeding of a commercial liposomal product: liposomal amphotericin B (AmBisome®) J Liposome Res. 2017;27:173–9. doi: 10.1080/08982104.2017.1380664. [DOI] [PubMed] [Google Scholar]
  • 321.Rivnay B, Wakim J, Avery K, Petrochenko P, Myung JH, Kozak D, et al. Critical process parameters in manufacturing of liposomal formulations of amphotericin B. Int J Pharm. 2019;565:447–57. doi: 10.1016/j.ijpharm.2019.04.052. [DOI] [PubMed] [Google Scholar]
  • 322.Li AN, Li S, Zhang YJ, Xu XR, Chen YM, Li HB. Resources and biological activities of natural polyphenols. Nutrients. 2014;6:6020–47. doi: 10.3390/nu6126020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 323.González-Vallinas M, González-Castejón M, Rodríguez-Casado A, Ramírez de Molina A. Dietary phytochemicals in cancer prevention and therapy: a complementary approach with promising perspectives. Nutr Rev. 2013;71:585–99. doi: 10.1111/nure.12051. [DOI] [PubMed] [Google Scholar]
  • 324.Araújo JR, Gonçalves P, Martel F. Chemopreventive effect of dietary polyphenols in colorectal cancer cell lines. Nutr Res. 2011;31:77–87. doi: 10.1016/j.nutres.2011.01.006. [DOI] [PubMed] [Google Scholar]
  • 325.Fresco P, Borges F, Marques MPM, Diniz C. The anticancer properties of dietary polyphenols and its relation with apoptosis. Curr Pharm Des. 2010;16:114–34. doi: 10.2174/138161210789941856. [DOI] [PubMed] [Google Scholar]
  • 326.Mocanu MM, Nagy P, Szöllősi J. Chemoprevention of breast cancer by dietary polyphenols. Molecules. 2015;20:22578–620. doi: 10.3390/molecules201219864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.Weichselbaum E, Buttriss JL. Polyphenols in the diet. Nutr Bull. 2010;35:157–64. doi: 10.1111/j.1467-3010.2010.01821.x. [DOI] [Google Scholar]
  • 328.Bravo L. Polyphenols: chemistry, dietary sources, metabolism, and nutritional significance. Nutr Rev. 1998;56:317–33. doi: 10.1111/j.1753-4887.1998.tb01670.x. [DOI] [PubMed] [Google Scholar]
  • 329.Pérez-Jiménez J, Neveu V, Vos F, Scalbert A. Identification of the 100 richest dietary sources of polyphenols: an application of the Phenol-Explorer database. Eur J Clin Nutr. 2010;64:S112–20. doi: 10.1038/ejcn.2010.221. [DOI] [PubMed] [Google Scholar]
  • 330.Duthie GG, Duthie SJ, Kyle JA. Plant polyphenols in cancer and heart disease: implications as nutritional antioxidants. Nutr Res Rev. 2000;13:79–106. doi: 10.1079/095442200108729016. [DOI] [PubMed] [Google Scholar]
  • 331.Huang Y, Wang YJ, Wang Y, Yi S, Fan Z, Sun L, et al. Exploring naturally occurring ivy nanoparticles as an alternative biomaterial. Acta Biomater. 2015;25:268–83. doi: 10.1016/j.actbio.2015.07.035. [DOI] [PubMed] [Google Scholar]
  • 332.Biasutto L, Mattarei A, Sassi N, Azzolini M, Romio M, Paradisi C, et al. Improving the efficacy of plant polyphenols. Anti Cancer Agents Med Chem. 2014;14:1332–42. doi: 10.2174/1871520614666140627150054. [DOI] [PubMed] [Google Scholar]
  • 333.Feldman F, Koudoufio M, Desjardins Y, Spahis S, Delvin E, Levy E. Efficacy of polyphenols in the management of dyslipidemia: a focus on clinical studies. Nutrients. 2021;13:672. doi: 10.3390/nu13020672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 334.Dryden GW, Song M, McClain C. Polyphenols and gastrointestinal diseases. Curr Opin Gastroenterol. 2006;22:165–70. doi: 10.1097/01.mog.0000208463.69266.8c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Gibellini L, Pinti M, Nasi M, Montagna JP, de Biasi S, Roat E, et al. Quercetin and cancer chemoprevention. Evid Based Complementary Altern Med. 2011;2011:591356. doi: 10.1093/ecam/neq053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Lam TK, Rotunno M, Lubin JH, Wacholder S, Consonni D, Pesatori AC, et al. Dietary quercetin, quercetin-gene interaction, metabolic gene expression in lung tissue and lung cancer risk. Carcinogenesis. 2010;31:634–42. doi: 10.1093/carcin/bgp334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 337.Ekström AM, Serafini M, Nyrén O, Wolk A, Bosetti C, Bellocco R. Dietary quercetin intake and risk of gastric cancer: results from a population-based study in Sweden. Ann Oncol. 2011;22:438–43. doi: 10.1093/annonc/mdq390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 338.Udenigwe CC, Ramprasath VR, Aluko RE, Jones PJ. Potential of resveratrol in anticancer and anti-inflammatory therapy. Nutr Rev. 2008;66:445–54. doi: 10.1111/j.1753-4887.2008.00076.x. [DOI] [PubMed] [Google Scholar]
  • 339.Taniguchi S, Fujiki H, Kobayashi H, Go H, Miyado K, Sadano H, et al. Effect of (‒)-epigallocatechin gallate, the main constituent of green tea, on lung metastasis with mouse B16 melanoma cell lines. Cancer Lett. 1992;65:51–4. doi: 10.1016/0304-3835(92)90212-e. [DOI] [PubMed] [Google Scholar]
  • 340.Valcic S, Timmermann BN, Alberts DS, Wächter GA, Krutzsch M, Wymer J, et al. Inhibitory effect of six green tea catechins and caffeine on the growth of four selected human tumor cell lines. Anti Cancer Drugs. 1996;7:461–8. doi: 10.1097/00001813-199606000-00011. [DOI] [PubMed] [Google Scholar]
  • 341.Yang GY, Liao J, Kim K, Yurkow EJ, Yang CS. Inhibition of growth and induction of apoptosis in human cancer cell lines by tea polyphenols. Carcinogenesis. 1998;19:611–6. doi: 10.1093/carcin/19.4.611. [DOI] [PubMed] [Google Scholar]
  • 342.Gupta S, Hastak K, Afaq F, Ahmad N, Mukhtar H. Essential role of caspases in epigallocatechin-3-gallate-mediated inhibition of nuclear factor kappa B and induction of apoptosis. Oncogene. 2004;23:2507–22. doi: 10.1038/sj.onc.1207353. [DOI] [PubMed] [Google Scholar]
  • 343.Drewnowski A, Gomez-Carneros C. Bitter taste, phytonutrients, and the consumer: a review. Am J Clin Nutr. 2000;72:1424–35. doi: 10.1093/ajcn/72.6.1424. [DOI] [PubMed] [Google Scholar]
  • 344.Higdon JV, Delage B, Williams DE, Dashwood RH. Cruciferous vegetables and human cancer risk: epidemiologic evidence and mechanistic basis. Pharmacol Res. 2007;55:224–36. doi: 10.1016/j.phrs.2007.01.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 345.Manach C, Scalbert A, Morand C, Rémésy C, Jiménez L. Polyphenols: food sources and bioavailability. Am J Clin Nutr. 2004;79:727–47. doi: 10.1093/ajcn/79.5.727. [DOI] [PubMed] [Google Scholar]
  • 346.Manach C, Williamson G, Morand C, Scalbert A, Rémésy C. Bioavailability and bioefficacy of polyphenols in humans. I. Review of 97 bioavailability studies. Am J Clin Nutr. 2005;81:230S–42S. doi: 10.1093/ajcn/81.1.230s. [DOI] [PubMed] [Google Scholar]
  • 347.Stalikas CD. Extraction, separation, and detection methods for phenolic acids and flavonoids. J Sep Sci. 2007;30:3268–95. doi: 10.1002/jssc.200700261. [DOI] [PubMed] [Google Scholar]
  • 348.Chiou A, Karathanos VT, Mylona A, Salta FN, Preventi F, Andrikopoulos NK. Currants (Vitis vinifera L.) content of simple phenolics and antioxidant activity. Food Chem. 2007;102:516–22. doi: 10.1016/j.foodchem.2006.06.009. [DOI] [Google Scholar]
  • 349.Qiu Y, Liu Q, Beta T. Antioxidant properties of commercial wild rice and analysis of soluble and insoluble phenolic acids. Food Chem. 2010;121:140–7. doi: 10.1016/j.foodchem.2009.12.021. [DOI] [Google Scholar]
  • 350.Metivier RP, Francis FJ, Clydesdale FM. Solvent extraction of anthocyanins from wine pomace. J Food Sci. 1980;45:1099–100. doi: 10.1111/j.1365-2621.1980.tb07534.x. [DOI] [Google Scholar]
  • 351.Prior RL, Lazarus SA, Cao G, Muccitelli H, Hammerstone JF. Identification of procyanidins and anthocyanins in blueberries and cranberries (Vaccinium spp.) using high-performance liquid chromatography/mass spectrometry. J Agric Food Chem. 2001;49:1270–6. doi: 10.1021/jf001211q. [DOI] [PubMed] [Google Scholar]
  • 352.Guyot S, Marnet N, Drilleau J. Thiolysis-HPLC characterization of apple procyanidins covering a large range of polymerization states. J Agric Food Chem. 2001;49:14–20. doi: 10.1021/jf000814z. [DOI] [PubMed] [Google Scholar]
  • 353.Labarbe B, Cheynier V, Brossaud F, Souquet JM, Moutounet M. Quantitative fractionation of grape proanthocyanidins according to their degree of polymerization. J Agric Food Chem. 1999;47:2719–23. doi: 10.1021/jf990029q. [DOI] [PubMed] [Google Scholar]
  • 354.Meisen A. Supercritical fluid extraction: principles and practice, Mark A. McHugh and val J. Krukonis, 1986, page 527, Butterworth Publishers, Stoneham, MA, USA. Price: unknown. Can J Chem Eng. 1988;66:527. doi: 10.1002/cjce.5450660332. [DOI] [Google Scholar]
  • 355.King MB, Bott TR, editors. Extraction of natural products using near-critical solvents. Amsterdam, Netherlands: Springer Dordrecht; 1993. [Google Scholar]
  • 356.Dai J, Mumper RJ. Plant phenolics: extraction, analysis and their antioxidant and anticancer properties. Molecules. 2010;15:7313–52. doi: 10.3390/molecules15107313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 357.Li H, Gou R, Liao J, Wang Y, Qu R, Tang Q, et al. Recent advances in nano-targeting drug delivery systems for rheumatoid arthritis treatment. Acta Mater Med. 2023;2:23–41. doi: 10.15212/amm-2022-0039. [DOI] [Google Scholar]
  • 358.Cuvier C, Roblot-Treupel L, Millot JM, Lizard G, Chevillard S, Manfait M, et al. Doxorubicin-loaded nanospheres bypass tumor cell multidrug resistance. Biochem Pharmacol. 1992;44:509–17. doi: 10.1016/0006-2952(92)90443-m. [DOI] [PubMed] [Google Scholar]
  • 359.Pusztai L, Wagner P, Ibrahim N, Rivera E, Theriault R, Booser D, et al. Phase II study of tariquidar, a selective P-glycoprotein inhibitor, in patients with chemotherapy-resistant, advanced breast carcinoma. Cancer. 2005;104:682–91. doi: 10.1002/cncr.21227. [DOI] [PubMed] [Google Scholar]
  • 360.Fracasso PM, Brady MF, Moore DH, Walker JL, Rose PG, Letvak L, et al. Phase II study of paclitaxel and valspodar (PSC 833) in refractory ovarian carcinoma: a gynecologic oncology group study. J Clin Oncol. 2001;19:2975–82. doi: 10.1200/jco.2001.19.12.2975. [DOI] [PubMed] [Google Scholar]
  • 361.Lhommé C, Joly F, Walker JL, Lissoni AA, Nicoletto MO, Manikhas GM, et al. Phase III study of valspodar (PSC 833) combined with paclitaxel and carboplatin compared with paclitaxel and carboplatin alone in patients with stage IV or suboptimally debulked stage III epithelial ovarian cancer or primary peritoneal cancer. J Clin Oncol. 2008;26:2674–82. doi: 10.1200/jco.2007.14.9807. [DOI] [PubMed] [Google Scholar]

Articles from Medical Review are provided here courtesy of De Gruyter

RESOURCES