6,585
Views
90
CrossRef citations to date
0
Altmetric
Review Article

Nanoparticle delivery of anticancer drugs overcomes multidrug resistance in breast cancer

, , , , &
Pages 3350-3357 | Received 17 Mar 2016, Accepted 12 Apr 2016, Published online: 29 May 2016

Abstract

Breast cancer is a serious threat to women's health, because multidrug resistance (MDR) has hampered treatment and prognosis. Nanodelivery of anticancer agents is a new technology to be exploited in the treatment of patients, because it bypasses multispecific drug efflux transporters such as P-glycoprotein (ABCB1), multidrug resistance protein-1 (MRP1, ABCC1) and breast cancer resistance protein (BCRP, ABCG2). Drugs can be delivered to tumor tissue by passive and active tumor targeting strategies, which may reduce or reverse drug resistance. This review will mainly focus on MDR-associated proteins, as well as various nanoparticle formulations developed to overcome MDR in breast cancer.

Introduction

Breast cancer is one of the most common cancers in women worldwide, and the incidence has been increasing over the past 20 years in most countries (Cecchini et al., Citation2015; Scalia-Wilbur et al., Citation2016). Chemotherapy is a mainstay in the treatment of breast cancer. However, tumor chemotherapy not only acts on tumor cells, but also exerts toxicity on non-cancer cells, and often results in rapid development of acquired drug resistance in tumor cells, particularly by upregulation of multidrug resistance (MDR) transporters. MDR is a phenomenon of cross resistance to multiple drugs with different structures, targets and functional mechanisms (Perez-Tomas, Citation2006; Ullah, Citation2008). At present, it is believed that MDR is the most important self-protection mechanism of tumor cells and also the most important reason for the failure of chemotherapy. In recent years, nanotechnologies have made considerable contributions to control drug release and to overcome MDR. Nanodrug delivery systems refer to technologies that deliver drugs by loading them into nanoliposomes, polymer micelles, inorganic metal nanoparticles or other nanomaterials, with a particle size between 1 nm and 1 μm. This review will mainly focus on recent progress of nanoparticle-mediated anticancer drug delivery in breast cancer and highlight recent progress in the field.

MDR in breast cancer

MDR mediated by drug efflux

The most frequent cause of tumor cells resistance to multiple drugs is overexpression of active drug efflux transporters, which intercept drugs at the level of cellular membranes. These are either located at the plasma membrane or alternatively at the nuclear membrane. This results in inadequate drug concentrations at cellular target structures and reduction of cancer cell toxicity. The most extensively studied MDR transporters include P-glycoprotein (P-gp/ABCB1), multidrug resistance protein-1 (MRP1/ABCC1), breast cancer resistance protein (BCRP, ABCG2) and lung resistance protein (LRP). P-gp, MRP1 and ABCG2 are members of the human family of ATP-binding cassette (ABC) transporters.

P-glycoprotein (P-gp)

P-gp, encoded by the MDR1 gene, is a 170 kDa plasma membrane protein which is produced from a 140 kDa precursor protein in the Golgi apparatus (Gottesman et al., Citation2002; Saraswathy & Gong, Citation2013). P-gp is composed of two pseudo-symmetrical halves, each consisting of an N-terminal transmembrane spanning domain (TMD) and a carboxy-terminal nucleotide binding domain (NBD). In the full length protein, these halves are arranged in tandem. Each of the TMDs has six transmembrane domains. Two pseudosymmetrical ATP-binding sites are formed at the interface of the NBDs (Chin et al., Citation1989; Gottesman et al., Citation2002). P-gp acquires energy for active transport of anticancer drugs through ATP binding and hydrolysis (Chang, Citation2003). It is able to intercept cytostatic drugs before they enter the cytoplasm and transport them back to the cell exterior against a concentration gradient. This leads to the phenotype of multidrug resistance, because P-gp has a multispecific substrate profile (Saraswathy & Gong, Citation2013). A large number of drugs used in cancer chemotherapy are substrates of P-gp, including anthracyclines, topoisomerase inhibitors, vinca alkaloids, camptothecin, and taxanes (Sun et al., Citation2004; Zhou, Citation2008). Studies have indicated that even a moderate increase in MDR1 expression (as low as 5-fold) were sufficient to cause doxorubicin resistance (Pajic et al., Citation2009).

Multidrug resistance protein

MRP, discovered in the process of studying the small cell lung cancer drug-resistant line H69AR, is another member of the human ABC family of plasma membrane associated transporters causing MDR (Cole et al., Citation1992). The MRP family consists of 13 proteins, most of which are associated with the transport of endogenous substances and xenobiotics. ABCC7/cystic fibrosis transmembrane conductance regulator (CFTR) is a chloride channel and ABCC8/sufonylurea receptor 1 (SUR1), and ABCC9/SUR2 are potassium channel regulators. Similar to P-gp MRP1 has been extensively studied over the past 25 years because of its prominent role in conferring multidrug resistance. MRP1 is a 190 kDa protein, consisting of 17 transmembrane domains with a P-gp-like core, but an additional N-terminal membrane spanning domain (TMD0) is present (Kruh & Belinsky, Citation2003; Leonessa & Clarke, Citation2003; Sodani et al., Citation2012). MRP1 is a basolateral transporter, the activity of which results in the movement of compounds into tissues that lie beneath the basement membrane. It is capable of transporting glutathione (GSH) and co-transporting drugs and-GSH and GSH-conjugated compounds (Evers et al., Citation1996). Efflux pumps involved in cellular export have been referred to as GSX pumps in the case of GSH conjugates (Ishikawa, Citation1992). Thence, the mechanism of MRP transport is different from that of P-gp (Kruh & Belinsky, Citation2003). Because MRP1 is ubiquitously expressed in human tissues, it is potentially present in most tumors and thus plays a role in drug resistance. MRP1 is able to confer resistance to anthracyclines, vinca alkaloids, epipodo-phyllotoxins, camptothecins and methotrexate, but not to taxanes, which are an important component of the P-gp profile. Besides, MRP1 is related to accelerated relapse in breast cancer, and a negative correlation between MRP1 expression and response to treatment has been found (Kruh & Belinsky, Citation2003; Sodani et al., Citation2012).

Breast cancer resistance protein

BCRP, encoded by the ABCG2 gene, was originally identified in an anticancer drug-resistant human breast cancer cell line under simultaneous treatment with mitoxantrone and inhibitors of P-gp. It was thus initially named mitoxantrone resistance protein (MXR) (Doyle et al., Citation1998; Ross et al., Citation1999; Chen et al., Citation2015). The protein, which belongs to the ABCG subfamily of ABC transporters, is an inverted half-transporter composed of an N-terminal nucleotide-binding site and a carboxyterminal transmembrane domain, which contains six membrane spanning helices (Gottesman et al., Citation2002; Gradhand & Kim, Citation2008). In its functional form, it is homodimeric and widely expressed in human tissues (Noguchi et al., Citation2009). Similar to ABCB1 and ABCC1 it plays an important role in intercellular drug absorption, distribution, metabolism, and excretion, as well as toxicity. The function as an active drug efflux transporter for anticancer drugs initially identified in breast cancer and later extended to other tumor types was eponymous for BCRP (Natarajan et al., Citation2012). In breast cancer cells, BCRP is not only expressed at the cell membrane, but also in intracellular vesicles that segregate and retain drugs, leading to drug compartmentalization. This is another reason for increased drug resistance of cells (Ifergan et al., Citation2005; Goler-Baron & Assaraf, Citation2011). Resistance of BCRP extends to the class of anthracyclines, but also to antifolatessuch as methotrexate (Doyle et al., Citation1998; Saraswathy & Gong, Citation2013).

Lung resistance protein

LRP, another resistance-related protein was initially identified in the non-small cell lung cancer cell line SW-1573/2R120, and is encoded by the LRP gene (Scheper et al., Citation1993). The gene for LRP was confirmed to regionally localize to chromosomal segment 16p11.2, a location approximately 27 cM proximal to MRP (16p13.1) (Slovak et al., Citation1995). Although both the MRP and LRP genes map to the short arm of chromosome 16, they are rarely co-amplified and are thus not normally located within the same amplicon. Primary sequence analysis of LRPcDNA revealed that its amino acid sequence was 87.7% homologous with the brown rat vault protein (major vault protein, MVP). LRP is not a member of the ABC transporter superfamily. It is found in the cytoplasm and at the nuclear membrane (Scheffer et al., Citation1995; Slovak et al., Citation1995). Vaults are cellular organelles broadly distributed and highly conserved among diverse eukaryotic cells, suggesting that they play a role in fundamental cellular processes. Vaults localize to nuclear pore complexes and may be the central plug of the nuclear pore complexes. Vault structure and localization support a transport function for this particle which could involve a variety of substrates. Vaults accordingly may play a role in drug resistance by regulating the nucleo-cytoplasmic transport of drugs (Izquierdo et al., Citation1996a). Histopathology demonstrated an increased amount of MVP/LRP protein in cancer, which results in lower anticancer drug levels in the nucleus (Szaflarski et al., Citation2011). Studies by Wood et al. (Wood & Streckfus, Citation2015) showed that LRP concentration in saliva of breast cancer patients in stageIwas significantly higher than that of healthy women. In a panel of 61 cancer cell lines which had not previously been undergone laboratory drug selection, LRP was a superior predictor for in vitro resistance to MDR-related drugs as compared to P-gp and MRP, and LRP's predictive value extended to MDR unrelated drugs, such as platinum compounds (Izquierdo et al., Citation1996b). Similar to the ABC MDRs, LRP also causes broad resistance to a partially overlapping panel of compounds including anthracyclines, alkaloids, epipodophyllotoxin, but in addition leads to resistance to cisplatin, melphalan and several atypical MDR drugs (Izquierdo et al., Citation1996a).

Other mechanisms of MDR

In addition to enhanced drug efflux mediated by MDR transporters, decreased drug concentration in the tumor is influenced by the tumor microenvironment, which is characterized by hypoxia and low pH (acidic cellular environment). Acidic pH outside cells limits uptake of weak base drugs such as doxorubicin, adriamycin and mitoxantrone (Chen et al., Citation2016). HIF-1α is a hypoxia-activated transcription factor that upregulates target genes by binding to the Hypoxia Response Element (HRE). MDR1 is a HIF-1 target gene and by that HIF-1-meditated P-gp expression results in drug resistance (Huang et al., Citation2010; Videira et al., Citation2014; Kapse-Mistry et al., Citation2014). Besides, by increasing water solubility of cytotoxic drugs, Glutathione S-transferase isoenzymes (GSTs) reduce toxicity, promote excretion and decrease the effective drug concentration (Huang et al., Citation2010). Altered DNA damage repair can also cause tumor MDR. TopoisomeraseII (TopoII) is an ATP-dependent enzyme that catalyzes the topological passing of two double-stranded DNA segments, and creates an intermediate DNA-enzyme complex referred to as the “cleavable complex” (Miller et al., Citation1981; Tewey et al., Citation1984). Reduced levels of TopoII catalytic activity decreases drug-induced TopoII-mediated cleavage of DNA, which leads to reduction of DNA breaks and lower cytotoxicity in MDR cells (Deffie et al., Citation1989; Ogiso et al., Citation2002). Alterations in apoptotic or antiapoptotic pathway such as loss of p53 (a transcription factor) (Oshika et al., Citation1998; Wang et al., Citation2011; Saha et al., Citation2012; Wang et al., Citation2013), abnormal expression of Bcl-2 family (Youle & Strasser, Citation2008; Saraswathy & Gong, Citation2013), increased activity of phosphatidylinositol-3-kinase (PI3K)/protein kinase B (Akt) (PI3k/Akt) (Jin et al., Citation2003; Zhang & Liu, Citation2011) and aberrant action of nuclear factor kappa beta (NF-κβ) (Bentires-Alj et al., Citation2003) also play a role in the development of MDR in breast cancer cells.

Nanoparticle encapsulated anticancer drugs overcome MDR in breast cancer cells

Anticancer drugs encapsulated in nanoparticles can actively or passively target tumor cells and thus improve the therapeutic effect at the target site. This can reduce systemic toxicity of chemotherapy drugs and bypass certain forms of multidrug resistance (Palakurthi et al., Citation2012). Blood vessels in most solid tumors possess the following unique characteristics: excessive angiogenesis and high vascular density, leaky vascular architecture, strengthened vascular permeability and impaired lymphatic drainage (Shing et al., Citation1985; Maeda & Matsumura, Citation1989; Skinner et al., Citation1990). The leaky blood vessels contain gaps in the basement membrane. The abnormal lymphatic drainage system of tumors fails to remove macromolecules and lipids from the interstitial space, trapping and retaining these over a longtime period. This phenomenon has been characterized as the tumor-selective enhanced permeability and retention (EPR) effect (Maeda & Matsumura, Citation1989; Maeda et al., Citation2000; Maeda, Citation2001). In passive targeting, nanoparticles pass through gaps in the leaky blood vessels and get trapped by the abnormal draining lymphatic system, generating the EPR effect. Passive targeting can also occur when positively-charged nanoparticles electrostatically interact with the negatively-charged sialic acid and phospholipids present on the surface of tumor associated endothelial cells (Patra & Turner, Citation2014; Cerqueira et al., Citation2015). Nanoparticles modified by active biomolecules such as nucleic acids, peptides, sugars and antibodies can actively bind to cancer cells. Ideally nanoparticles with high affinity selectively combine with molecules to target such as sugars, proteins, folate, transferrin, aptamers or lipids. These are overexpressed on the surface of cancer cells and nanoparticle delivery thus minimizes damage to non-cancer cells (Bertrand et al., Citation2014). Active targeting is applied to specifically recognize cells through the interactions of active biomolecules, enhance drug endocytosis by the cell, decrease cytotoxicity on non-cancer cells, increase the concentration of drug, and exploit the EPR effect (Patra & Turner, Citation2014; Cerqueira et al., Citation2015).

Nanoparticles that are taken up by the cell via endocytosis often bypass and evade the ABC-transporters responsible for efflux of cytotoxic drugs once released into the cytoplasm (Song et al., Citation2010; Huang et al., Citation2011; Cerqueira et al., Citation2015). There are four main mechanisms of endocytosis: clathrin-mediated endocytosis, caveolae-mediated endocytosis, macropinocytosis, and other endocytosis. Clathrin-mediated endocytosis is the most widely studied mechanism involved in receptor-mediated uptake of nanoparticles, apart from receptor-independent endocytosis. The relevant receptors include the transferrin, low density lipoprotein (LDL) receptor and the epidermal growth factor receptors (EGFR) including the human epidermal growth factor receptor 2 (HER2). Nanoparticles entering into cells via clathrin-mediated endocytosis end up in lysosomes and are degraded to release drugs by the acidic environment of lysosomes. Caveolae-mediated endocytosis generates cytosolic caveolar vesicles following nanoparticles binding to the cell membrane. Ligands associated with caveolae-mediated endocytosis include folic acid, albumin, and cholesterol. Macropinocytosis, another nonselective endocytosis pathway, relies on actin-driven membrane protusions, which subsequently fuzes with and separates from the plasma membrane to generate macropinosomes (Hillaireau & Couvreur, Citation2009; Saraswathy & Gong, Citation2013). Besides, nanoparticles containing co-encapsulated P-gp inhibitors and anti-cancer drugs can be used to overcome P-gp meditated MDR (Livney & Assaraf, Citation2013; Kibria et al., Citation2014; Cerqueira et al., Citation2015). Wong et al. (Wong et al., Citation2006) combined doxorubicin and elacridar, a P-gp inhibitor, in polymer-lipid hybrid nanoparticles. The in vitro result showed that simultaneous delivery of those two drugs enhanced the treatment of multidrug-resistant breast cancer ().

Table 1. Main mechanisms of nanoparticles reversing tumor MDR.

Liposomes

Liposomes have an aequeous core surrounded by a phospholipid bilayer. Therefore, they are able to carry both lipophilic and hydrophilic drugs (Qin et al., Citation2014). Liposomes, which are mainly composed of natural phospholipids, are thought to be pharmacologically inactive with minimal toxicity and therefore have good biocompatibility (Sercombe et al., Citation2015). It was shown that anionic membrane lipids (cardiolipin and phosphatidylserine) inhibited P-gp by direct interaction. This enhanced cellular absorption and cellular toxicity compared to free drugs (Kapse-Mistry et al., Citation2014). Phospholipids have been shown to be substrates of P-gp in many cell lines and P-gp was considered to be a phospholipid translocase with broad specificity (van Helvoort et al., Citation1996; Bosch et al., Citation1997). Lo (Citation2000) proposed that modulation of P-gp by phospholipids could be brought about by substrate substitution or membrane fluidization, thereby increasing epirubicin cytotoxicity. Encapsulation of drugs in liposomes prevents rapid inactivation, degradation and dilution in the circulation (Hua & Wu, Citation2013). Kang et al. (Citation2009) compared retention of Rhodamine 123, a fluorescent probe, often used as a paradigmatic P-gp and BCRP substrate, loaded into liposomes with different compositions in two kinds of breast cancer cells: wild type MCF-7/WT cells and P-gp overexpressing MCF-7/P-gp cells. The results showed that liposome encapsulation increased the retention of Rhodamine 123 in MCF-7/P-gp cell, but not in MCF-7/WT cells. Liposomes tend to accumulate preferentially in tumors because of the EPR effect. However, the delivery system is prone to rapid elimination by the reticuloendothelial system (RES) (Kang et al., Citation2009). To enhance liposome stability and improve the circulation times in the blood, PEGylated liposomes (long-circulating liposomes) were introduced. PEGylated liposome formulations of paclitaxel established by Yang et al. (Citation2007) showed increased biological half-life of paclitaxel (17.8 (±2.35)h versus 5.05 (±1.52)h, reduced uptake in the RES and increased uptake in tumor tissues in rats compared to conventional liposomal formulations. Liposomes, that are not specifically targeted, can become targeted liposomes by linking selective ligands to their surface, which bears great potential for site-specific delivery of drugs. Examples of such targeted liposomes are: immunoglobuline modified liposomes (immune liposomes (ILs)), folate- or sugar-modified liposomes. Modified with selective ligands, liposome encapsulated drugs undergo targeted endocytosis and thus are not subject to active drug efflux by MDR transporters (Kapse-Mistry et al., Citation2014). Alternatively, higher temperature and lower pH in tumor tissue can be exploited by manufacturing temperature and pH-sensitive liposomes, which release drugs in target tissue in a temperature and pH-dependent manner (Qin et al., Citation2014; Vaidya et al., Citation2016).

Polymer micelle nanoparticles

Polymeric micelle nanoparticles with a core-shell structure are spontaneously formed by amphiphilic copolymer. The shell is hydrophilic and the kernel is a hydrophobic drug reservoir composed by a hydrophobic block copolymer. The hydrophilic shell inhibits protein adsorption, which decreases the polymeric micelles involved in foreign body reaction while improving drug solubility (Kataoka et al., Citation2001). The pluronic block copolymers are one of the polymeric micelle nanoparticles widely used in various drug delivery systems. Pluronics are triblock copolymers of poly(ethylene oxide) (PEO) and poly(propylene oxide) (PPO), arranged in PEO-PPO-PEO structure. They have a core containing PPO blocks, which serves to incorporate hydrophobic drugs (Alakhova & Kabanov, Citation2014). Pluronic block copolymers are thought to associate with energy metabolism in MDR cancer cells. It is rapidly taken up by the cells via a caveolae-mediated endocytosis pathway and co-localizes with mitochondria after 15 min. Moreover, it suppresses complex I and complex IV of the mitochondrial respiratory chain, reduces oxygen consumption and leads to ATP depletion in MDR cells, while non-MDR cells require significantly higher doses of Pluronic to achieve similar depletion (Alakhova et al., Citation2010; Sahay et al., Citation2010). Studies showed that the Pluronic P85 (P85) resulted in a substantial decrease in ATP levels selectively in MDR cells, thereby inhibiting P-gp and sensitizing MDR cells (Batrakova et al., Citation2001). Using isolated Pgp-, MRP1-, and MRP2-overexpressing membranes, Batrakova et al. (Citation2004) evaluated the effects of P85 on the kinetic parameters Vmax, Km and first-order rate constant (Vmax/Km) of ATP hydrolysis by these ATPases. The effect of P85 on P-gp, MRP1 and MRP2 ATPase activity was as follows: MRP1 < MRP2 ≪ P-gp. Likewise, studies by Kabanov et al. (Citation2003) indicated that the effect of P85 on cells with respect to ATP depletion correlated with the levels of expression of the multidrug transporter, primarily that of P-gp.

Micelles carrying drugs can selectively accumulate and release their cargo in tumors through the EPR effect (passive targeting), or when modified by specific ligands (active targeting). Thus, the micelles are ideal carriers for active and passive targeting approaches with the aim to enhance efficacy and reduce toxicity. More recent researches (Lee et al., Citation2005; Mohajer et al., Citation2007) showed that micelles with pH sensitivity or surface modifications, have improved performance in overcoming MDR. Studies on folic acid modified pH-sensitive doxorubicin micelles (mixture of poly-histidine-polyethylene glycol-folic acid and poly-l-lactic acid-polyethylene glycol-folic acid) revealed that drug was released from the endosome into the cytoplasm and nucleus due to the formation of nanomicelles from early endosomes. The latter was triggered by lower pH in this compartment (about 6). The micelle drug delivery system overcame P-gp efflux due to folate-receptor mediated endocytosis. This increased intracellular drug accumulation, prolonged residence time of intracellular drug, and enhanced cytotoxicity in resistant MCF-7/DOX cells (Lee et al., Citation2005; Mohajer et al., Citation2007; Kim et al., Citation2008). Studies by Wang et al. (Citation2007) on four different paclitaxel (PTX) containing micelles prepared with Pluronic-P105, a mixture of P105 and L101, P105 modified by folic acid (FOL-P105) and a mixture of P105 and L101 modified by folic acid (FOL-PL) revealed that all micellar formulations had an improved anticancer effect on human breast cancer resistant MCF-7/DOX cells as compared to PTX in solution. FOL-P105 and FOL-PL micelles were more toxic than micelles without folic acid, indicating combined active targeting of ligands with the direct effect of Pluronic on multidrug resistant cells was superior in enhancing cytotoxicity (Kabanov et al., Citation2002; Batrakova et al., Citation2003,Citation2004; Sharma et al., Citation2008). Further research (Werle & Hoffer, Citation2006) revealed that Chitosan-(4-thiol Butanimidamide)/GSH-mixed micelles not only increased the absorption of Rhodamine123 and Saquinavir, but also inhibited P-gp-dependent ATPase, leading to decreased drug efflux.

Dendrimers

Dendrimers are polymers with spherical shape, dendritic skeleton and a number of functional groups on its surface. In this drug targeting delivery system drugs may either be encapsulated within the internal cavity of the polymer or coupled with the surface functional groups, which resulted in improved drug-stability and decreased cytotoxicity (Qin et al., Citation2014). It was considered that dendrimer phthalocyanine was internalized into cells via endocytosis. Anionic dendrimer phthalocyanine (DPc) electrostatically interacted with poly(ethylene glycol)-b-poly(l-lysine) block copolymer (PEG-PLL) to form polyion complex(PIC) micelles. A doxorubicin containing DPc-encapsulated PIC micelle (DPc/m) was shown to internalize into cells via endocytosis, and accumulate in the endosome/lysosome compartments after irradiation. Finally, the drug would accumulate in the nucleus in drug-resistant MCF-7 cells (Lu et al., Citation2011). A G3 PAMAM-NH(2) dendrimer-chlorambucil conjugate exerted a more potent antiproliferative effect and stronger inhibition of [(3)H]thymidine incorporation into DNA in both MDA-MB-231 and MCF-7 breast cancer cells than chlorambucil alone (Bielawski et al., Citation2011). Targeted dendrimer nanoparticles are often prepared with surface-modified ligands and the targeting moieties such as sugar, folate, antibody, peptide or epidermal growth factor, which are conjugated to the dendrimers cause preferential accumulation of drug in the target tissue or cells than untargeted controls or free drug, thereby producing a stronger reduction of the tumor size (Hong et al., Citation2007). It has been shown that in analogous manner biotin (a cofactor for carboxylation reactions taken up in rapidly proliferating tissues such as cancer cells) was conjugated to the dendrimer to improve the uptake of anticancer drugs in tumor cells (Yang et al., Citation2009).

Mesoporous silica nanoparticles

Mesoporous silica nanoparticles (MSNs) can hold large amounts of drugs due to high pore volume and surface area. Drugs are loaded into the mesoporous pore to improve drug-stability and have targeted and sustained action. MSNs contribute to improving the solubility of hydrophobic drugs due to the water-soluble silanol groups on their surface, and can be modified for targeted delivery on the basis of different functionalities. Conjugation of targeting moieties with specific recognition on tumor cells include folic acid, monoclonal antibodies, glycoproteins, peptides, nucleic acids, which provide a means for effectively and accurately delivering drugs to target tissues (Trewyn et al., Citation2007; Mamaeva et al., Citation2013). It was reported that the main pathway of cellular uptake of MSNs was macropinocytosis, which is a form of endocytosis that accompanies cell surface ruffling and that this uptake was distinct from micropinocytosis including clathrin-coated vesicle endocytosis. Doxorubicin-loaded MSNs (DMNs) induced higher accumulation of doxorubicin in drug-resistant tumors than free doxorubicin. MSNs themselves caused down-regulation of P-gp expression. This could be the main reason for the enhanced anti-cancer activity of DMNs against P-gp over-expressing MCF-7/ADR cells. When DMNs were endocytosed into cells, doxorubicin would not be available for efflux by P-gp (Shen et al., Citation2011). Huang et al. (Citation2011) used the endosomal pH-sensitive MSN (MSN-Hydrazone-Dox) for controlled release of doxorubicin in cells while evading its P-gp-mediated efflux. Both in vitro and in vivo studies showed that MSNs could serve as an efficient nanodelivery system entering cell via endocytosis and thus bypassing P-gp mediated efflux. Further studies (Slowing et al., Citation2006) showed that the MSN drug delivery system in combination with modifications by folic acid enhanced cellular uptake of drugs via receptor-mediated endocytosis. Moreover, Slowing et al. (Citation2006) investigated the possible mechanism of endocytosis of the MSNs by introducing different surface functionalities. The authors prepared a fluorescein-functionalized MSN (FITC-MSN), and the functional groups 3-amino-propyl (AP), N-(2-aminoethyl)-3-aminopropyl (AEAP), N-folate-3-aminopropyl (FAP), guanidinopropyl (GP), and 3-[N-(2-guanidinoethyl)guanidino]propyl (GEGP) were grafted onto the external surface of FITC-MSN. The results indicated that the FITC- and FAP-MSNs were endocytosed via a clathrin-pitted mechanism mediated by folic acid receptors, while the endocytosis of AP- and GP-MSNs were endocytosed via acaveolae-mediated mechanism. The uptake mechanism for GEGP remained unclear. These studies indicated surface functionalities to play a role in the uptake of MSNs. The cytotoxicity of the MSN drug delivery system with different pore size on MCF-7/ADR cells showed that the drug delivery system with a pore size of 12.6 nm was more powerful than that of 3.2 nm. This confirmed that the anti-tumor activity in breast cancer cells is dependent on pore size, whereby apoptosis increased with increasing nanoparticle pore size (Gao et al., Citation2011; Jia et al., Citation2013).

In addition to MSNs, inorganic nanoparticles used in tumor treatment include quantum dots, carbon nanotubes, silica nanoparticles, gold nanoparticles, iron oxide magnetic nanoparticles and ceramic nanoparticles, which provide additional opportunities for the treatment and diagnosis of tumors.

Conclusions and outlook

MDR in tumor cells is still one of the major impediments in the clinical treatment of breast cancers. Overall thorough research is needed for further understanding the mechanism of MDR as a prerequisite for reversing various MDR forms in breast cancer, avoid MDR-related pathways, and enhance the efficacy of anti-tumor drugs. Nanodrugs hold potential in mediating sustained drug release, selectively targeting tumor cells, bypassing drug efflux, improving drug therapeutic indices, and reversing an MDR phenotype in tumor cells. Compared with free drugs, drugs encapsulated by nanomaterials can enter the target cells through different mechanisms to avoid the drug transporters P-gp, MRP and BCRP and to increase the intracellular accumulation of drugs and by that reverse MDR. The mechanisms leading to MDR are complex, and application of nanocarriers combined with targeting technologies to overcome or reverse MDR has been recognized as an important and promising field of research in recent years. It also holds great potential for tumor diagnostic procedures.

Although nanomaterials, with excellent physiological properties have been applied more and more widely in cancer treatment and diagnosis, studying on the potential toxicity of nanomaterials on cells and organism has not yet caught up with the rapid development and wide use of this technology. Using high-throughput methods such as metabolomics to evaluate the potential toxicity of nanoparticles holds excellent prospect for research and clinical applications in the future.

Declaration of interest

The authors report no conflicts of interest in this work.

We acknowledge the financial support from the National Natural Science Foundation of China (81273707, 81173215), the Ministry of Education in the New Century Excellent Talents (NECT-12-0677), the Natural Science Foundation of Guangdong (S2013010012880), the Science and Technology Program of Guangzhou (2014J4500005), the Science Program of the Department of Education of Guangdong (2013KJCX0021, 2015KGJHZ012), and the Science and Technology Program of Guangdong (2015A050502027).

References

  • Alakhova DY, Kabanov AV. (2014). Pluronics and MDR reversal: an update. Mol Pharm 11:2566–78.
  • Alakhova DY, Rapoport NY, Batrakova EV, et al. (2010). Differential metabolic responses to pluronic in MDR and non-MDR cells: a novel pathway for chemosensitization of drug resistant cancers. J Control Release 142:89–100.
  • Batrakova EV, Li S, Alakhov VY, et al. (2003). Sensitization of cells overexpressing multidrug-resistant proteins by pluronic P85. Pharm Res 20:1581–90.
  • Batrakova EV, Li S, Elmquist WF, et al. (2001). Mechanism of sensitization of MDR cancer cells by pluronic block copolymers: selective energy depletion. Br J Cancer 85:1987–97.
  • Batrakova EV, Li S, Li Y, et al. (2004). Effect of pluronic P85 on ATPase activity of drug efflux transporters. Pharm Res 21:2226–33.
  • Bentires-Alj M, Barbu V, Fillet M, et al. (2003). NF-kappaB transcription factor induces drug resistance through MDR1 expression in cancer cells. Oncogene 22:90–7.
  • Bertrand N, Wu J, Xu X, et al. (2014). Cancer nanotechnology: the impact of passive and active targeting in the era of modern cancer biology. Adv Drug Deliv Rev 66:2–25.
  • Bielawski K, Bielawska A, Muszynska A, et al. (2011). Cytotoxic activity of G3 PAMAM-NH(2) dendrimer-chlorambucil conjugate in human breast cancer cells. Environ Toxicol Pharmacol 32:364–72.
  • Bosch I, Dunussi-Joannopoulos K, Wu RL, et al. (1997). Phosphatidylcholine and phosphatidylethanolamine behave as substrates of the human MDR1 P-glycoprotein. Biochemistry 36:5685–94.
  • Cecchini MJ, Yu E, Potvin K, et al. (2015). Concurrent or sequential hormonal and radiation therapy in breast cancer: a literature review. Cureus 7:e364.
  • Chen D, Lian S, Sun J, et al. (2016). Design of novel multifunctional targeting nano-carrier drug delivery system based on CD44 receptor and tumor microenvironment pH condition. Drug Deliv 23:808–13.
  • Cerqueira BB, Lasham A, Shelling AN, Al-Kassas R. (2015). Nanoparticle therapeutics: technologies and methods for overcoming cancer. Eur J Pharm Biopharm 97:140–51.
  • Chang G. (2003). Multidrug resistance ABC transporters. FEBS Lett 555:102–5.
  • Chen Z, Shi T, Zhang L, et al. (2015). Mammalian drug efflux transporters of the ATP binding cassette (ABC) family in multidrug resistance: a review of the past decade. Cancer Lett 370:153–64
  • Chin JE, Soffir R, Noonan KE, et al. (1989). Structure and expression of the human MDR (P-glycoprotein) gene family. Mol Cell Biol 9:3808–20.
  • Cole SP, Bhardwaj G, Gerlach JH, et al. (1992). Overexpression of a transporter gene in a multidrug-resistant human lung cancer cell line. Science 258:1650–4.
  • Deffie AM, Batra JK, Goldenberg GJ. (1989). Direct correlation between DNA topoisomerase II activity and cytotoxicity in adriamycin-sensitive and -resistant P388 leukemia cell lines. Cancer Res 49:58–62.
  • Doyle LA, Yang W, Abruzzo LV, et al. (1998). A multidrug resistance transporter from human MCF-7 breast cancer cells. Proc Natl Acad Sci USA 95:15665–70.
  • Evers R, Zaman GJ, Van Deemter L, et al. (1996). Basolateral localization and export activity of the human multidrug resistance-associated protein in polarized pig kidney cells. J Clin Invest 97:1211–18.
  • Gao Y, Chen Y, Ji X, et al. (2011). Controlled intracellular release of doxorubicin in multidrug-resistant cancer cells by tuning the shell-pore sizes of mesoporous silica nanoparticles. ACS Nano 5:9788–98.
  • Goler-Baron V, Assaraf YG. (2011). Structure and function of ABCG2-rich extracellular vesicles mediating multidrug resistance. PLoS One 6:e16007
  • Gottesman MM, Fojo T, Bates SE. (2002). Multidrug resistance in cancer: role of ATP-dependent transporters. Nat Rev Cancer 2:48–58.
  • Gradhand U, Kim RB. (2008). Pharmacogenomics of MRP transporters (ABCC1-5) and BCRP (ABCG2). Drug Metab Rev 40:317–54.
  • Hillaireau H, Couvreur P. (2009). Nanocarriers' entry into the cell: relevance to drug delivery. Cell Mol Life Sci 66:2873–96.
  • Hong S, Leroueil PR, Majoros IJ, et al. (2007). The binding avidity of a nanoparticle-based multivalent targeted drug delivery platform. Chem Biol 14:107–15.
  • Hua S, Wu SY. (2013). The use of lipid-based nanocarriers for targeted pain therapies. Front Pharmacol 4:143
  • Huang IP, Sun SP, Cheng SH, et al. (2011). Enhanced chemotherapy of cancer using pH-sensitive mesoporous silica nanoparticles to antagonize P-glycoprotein-mediated drug resistance. Mol Cancer Ther 10:761–9.
  • Huang J, He YM, Lin LW, et al. (2010). Advance of multidrug resistance mechanism of breast cancer reversal agents. BME Clin Med 14:556–63.
  • Ifergan I, Scheffer GL, Assaraf YG. (2005). Novel extracellular vesicles mediate an ABCG2-dependent anticancer drug sequestration and resistance. Cancer Res 65:10952–8.
  • Ishikawa T. (1992). The ATP-dependent glutathione S-conjugate export pump. Trends Biochem Sci 17:463–8.
  • Izquierdo MA, Scheffer GL, Flens MJ, et al. (1996a). Broad distribution of the multidrug resistance-related vault lung resistance protein in normal human tissues and tumors. Am J Pathol 148:877–87.
  • Izquierdo MA, Scheffer GL, Flens MJ, et al. (1996b). Relationship of LRP-human major vault protein to in vitro and clinical resistance to anticancer drugs. Cytotechnology 19:191–7.
  • Jia L, Shen J, Li Z, et al. (2013). In vitro and in vivo evaluation of paclitaxel-loaded mesoporous silica nanoparticles with three pore sizes. Int J Pharm 445:12–9.
  • Jin W, Wu L, Liang K, et al. (2003). Roles of the PI-3K and MEK pathways in ras-mediated chemoresistance in breast cancer cells. Br J Cancer 89:185–91.
  • Kabanov AV, Batrakova EV, Alakhov VY. (2002). Pluronic block copolymers for overcoming drug resistance in cancer. Adv Drug Deliv Rev 54:759–79.
  • Kabanov AV, Batrakova EV, Alakhov VY. (2003). An essential relationship between ATP depletion and chemosensitizing activity of pluronic block copolymers. J Control Release 91:75–83.
  • Kang DI, Kang HK, Gwak HS, et al. (2009). Liposome composition is important for retention of liposomal rhodamine in P-glycoprotein-overexpressing cancer cells. Drug Deliv 16:261–7.
  • Kapse-Mistry S, Govender T, Srivastava R, Yergeri M. (2014). Nanodrug delivery in reversing multidrug resistance in cancer cells. Front Pharmacol 5:159
  • Kataoka K, Harada A, Nagasaki Y. (2001). Block copolymer micelles for drug delivery: design, characterization and biological significance. Adv Drug Deliv Rev 47:113–31.
  • Kibria G, Hatakeyama H, Harashima H. (2014). Cancer multidrug resistance: mechanisms involved and strategies for circumvention using a drug delivery system. Arch Pharm Res 37:4–15.
  • Kim D, Lee ES, Park K, et al. (2008). Doxorubicin loaded pH-sensitive micelle: antitumoral efficacy against ovarian A2780/DOXR tumor. Pharm Res 25:2074–82.
  • Kruh GD, Belinsky MG. (2003). The MRP family of drug efflux pumps. Oncogene 22:7537–52.
  • Lee ES, Na K, Bae YH. (2005). Doxorubicin loaded pH-sensitive polymeric micelles for reversal of resistant MCF-7 tumor. J Control Release 103:405–18.
  • Leonessa F, Clarke R. (2003). ATP binding cassette transporters and drug resistance in breast cancer. Endocr Relat Cancer 10:43–73.
  • Livney YD, Assaraf YG. (2013). Rationally designed nanovehicles to overcome cancer chemoresistance. Adv Drug Deliv Rev 65:1716–30.
  • Lo YL. (2000). Phospholipids as multidrug resistance modulators of the transport of epirubicin in human intestinal epithelial Caco-2 cell layers and everted gut sacs of rats. Biochem Pharmacol 60:1381–90.
  • Lu HL, Syu WJ, Nishiyama N, et al. (2011). Dendrimer phthalocyanine-encapsulated polymeric micelle-mediated photochemical internalization extends the efficacy of photodynamic therapy and overcomes drug-resistance in vivo. J Control Release 155:458–64.
  • Maeda H. (2001). The enhanced permeability and retention (EPR) effect in tumor vasculature: the key role of tumor-selective macromolecular drug targeting. Adv Enzyme Regul 41:189–207.
  • Maeda H, Matsumura Y. (1989). Tumoritropic and lymphotropic principles of macromolecular drugs. Crit Rev Ther Drug Carrier Syst 6:193–210.
  • Maeda H, Wu J, Sawa T, et al. (2000). Tumor vascular permeability and the EPR effect in macromolecular therapeutics: a review. J Control Release 65:271–84.
  • Mamaeva V, Sahlgren C, Linden M. (2013). Mesoporous silica nanoparticles in medicine-recent advances. Adv Drug Deliv Rev 65:689–702.
  • Miller KG, Liu LF, Englund PT. (1981). A homogeneous type II DNA topoisomerase from HeLa cell nuclei. J Biol Chem 256:9334–9.
  • Mohajer G, Lee ES, Bae YH. (2007). Enhanced intercellular retention activity of novel pH-sensitive polymeric micelles in wild and multidrug resistant MCF-7 cells. Pharm Res 24:1618–27.
  • Natarajan K, Xie Y, Baer MR, Ross DD. (2012). Role of breast cancer resistance protein (BCRP/ABCG2) in cancer drug resistance. Biochem Pharmacol 83:1084–103.
  • Noguchi K, Katayama K, Mitsuhashi J, Sugimoto Y. (2009). Functions of the breast cancer resistance protein (BCRP/ABCG2) in chemotherapy. Adv Drug Deliv Rev 61:26–33.
  • Ogiso Y, Tomida A, Tsuruo T. (2002). Nuclear localization of proteasomes participates in stress-inducible resistance of solid tumor cells to topoisomerase II-directed drugs. Cancer Res 62:5008–12.
  • Oshika Y, Nakamura M, Tokunaga T, et al. (1998). Multidrug resistance-associated protein and mutant p53 protein expression in non-small cell lung cancer. Mod Pathol 11:1059–63.
  • Pajic M, Iyer JK, Kersbergen A, et al. (2009). Moderate increase in Mdr1a/1b expression causes in vivo resistance to doxorubicin in a mouse model for hereditary breast cancer. Cancer Res 69:6396–404.
  • Palakurthi S, Yellepeddi VK, Vangara KK. (2012). Recent trends in cancer drug resistance reversal strategies using nanoparticles. Expert Opin Drug Deliv 9:287–301.
  • Patra HK, Turner AP. (2014). The potential legacy of cancer nanotechnology: cellular selection. Trends Biotechnol 32:21–31.
  • Perez-Tomas R. (2006). Multidrug resistance: retrospect and prospects in anti-cancer drug treatment. Curr Med Chem 13:1859–76.
  • Qin Y, Li M, Qiu Q, et al. (2014). The application of nanoparticles drug system in cancer multidrug resistance. Prog Mod Biomed 14:5599–64.
  • Ross DD, Yang W, Abruzzo LV, et al. (1999). Atypical multidrug resistance: breast cancer resistance protein messenger RNA expression in mitoxantrone-selected cell lines. J Natl Cancer Inst 91:429–33.
  • Saha S, Adhikary A, Bhattacharyya P, et al. (2012). Death by design: where curcumin sensitizes drug-resistant tumours. Anticancer Res 32:2567–84.
  • Sahay G, Gautam V, Luxenhofer R, Kabanov AV. (2010). The utilization of pathogen-like cellular trafficking by single chain block copolymer. Biomaterials 31:1757–64.
  • Saraswathy M, Gong S. (2013). Different strategies to overcome multidrug resistance in cancer. Biotechnol Adv 31:1397–407.
  • Scalia-Wilbur J, Colins BL, Penson RT, Dizon DS. (2016). Breast cancer risk assessment: moving beyond BRCA 1 and 2. Semin Radiat Oncol 26:3–8.
  • Scheffer GL, Wijngaard PL, Flens MJ, et al. (1995). The drug resistance-related protein LRP is the human major vault protein. Nat Med 1:578–82.
  • Scheper RJ, Broxterman HJ, Scheffer GL, et al. (1993). Overexpression of a M(r) 110,000 vesicular protein in non-P-glycoprotein-mediated multidrug resistance. Cancer Res 53:1475–9.
  • Sercombe L, Veerati T, Moheimani F, et al. (2015). Advances and challenges of liposome assisted drug delivery. Front Pharmacol 6:286.
  • Sharma AK, Zhang L, Li S, et al. (2008). Prevention of MDR development in leukemia cells by micelle-forming polymeric surfactant. J Control Release 131:220–7.
  • Shen J, He Q, Gao Y, et al. (2011). Mesoporous silica nanoparticles loading doxorubicin reverse multidrug resistance: performance and mechanism. Nanoscale 3:4314–22.
  • Shing Y, Folkman J, Haudenschild C, et al. (1985). Angiogenesis is stimulated by a tumor-derived endothelial cell growth factor. J Cell Biochem 29:275–87.
  • Skinner SA, Tutton PJ, O'Brien PE. (1990). Microvascular architecture of experimental colon tumors in the rat. Cancer Res 50:2411–17.
  • Slovak ML, Ho JP, Cole SP, et al. (1995). The LRP gene encoding a major vault protein associated with drug resistance maps proximal to MRP on chromosome 16: evidence that chromosome breakage plays a key role in MRP or LRP gene amplification. Cancer Res 55:4214–19.
  • Slowing I, Trewyn BG, Lin VERSUS (2006). Effect of surface functionalization of MCM-41-type mesoporous silica nanoparticles on the endocytosis by human cancer cells. J Am Chem Soc 128:14792–3.
  • Sodani K, Patel A, Kathawala RJ, Chen ZS. (2012). Multidrug resistance associated proteins in multidrug resistance. Chin J Cancer 31:58–72.
  • Song XR, Zheng Y, He G, et al. (2010). Development of PLGA nanoparticles simultaneously loaded with vincristine and verapamil for treatment of hepatocellular carcinoma. J Pharm Sci 99:4874–9.
  • Sun J, He ZG, Cheng G, et al. (2004). Multidrug resistance P-glycoprotein: crucial significance in drug disposition and interaction. Med Sci Monit 10:Ra5–14.
  • Szaflarski W, Nowicki M, Zabel M. (2011). The structure of cellular vaults, their role in the normal cell and in the multidrug resistance of cancer. Postepy Biochem 57:266–73.
  • Tewey KM, Chen GL, Nelson EM, Liu LF. (1984). Intercalative antitumor drugs interfere with the breakage-reunion reaction of mammalian DNA topoisomerase II. J Biol Chem 259:9182–7.
  • Trewyn BG, Giri S, Slowing II, Lin VS. (2007). Mesoporous silica nanoparticle based controlled release, drug delivery, and biosensor systems. Chem Commun (Camb) 38:3236–45.
  • Ullah MF. (2008). Cancer multidrug resistance (MDR): a major impediment to effective chemotherapy. Asian Pac J Cancer Prev 9:1–6.
  • Vaidya B, Nayak MK, Dash D, et al. (2016). Development and characterization of highly selective target-sensitive liposomes for the delivery of streptokinase: in vitro/in vivo studies. Drug Deliv 23:801–7.
  • Van Helvoort A, Smith AJ, Sprong H, et al. (1996). MDR1 P-glycoprotein is a lipid translocase of broad specificity, while MDR3 P-glycoprotein specifically translocates phosphatidylcholine. Cell 87:507–17.
  • Videira M, Reis RL, Brito MA. (2014). Deconstructing breast cancer cell biology and the mechanisms of multidrug resistance. Biochim Biophys Acta 1846:312–25.
  • Wang QP, Wang Y, Wang XD, et al. (2013). Survivin up-regulates the expression of breast cancer resistance protein (BCRP) through attenuating the suppression of p53 on NF-κB expression in MCF-7/5-FU cells. Int J Biochem Cell Biol 45:2036–44.
  • Wang Y, Suh YA, Fuller MY, et al. (2011). Restoring expression of wild-type p53 suppresses tumor growth but does not cause tumor regression in mice with a p53 missense mutation. J Clin Invest 121:893–904.
  • Wang Y, Yu L, Han L, et al. (2007). Difunctional pluronic copolymer micelles for paclitaxel delivery: synergistic effect of folate-mediated targeting and pluronic-mediated overcoming multidrug resistance in tumor cell lines. Int J Pharm 337:63–73.
  • Werle M, Hoffer M. (2006). Glutathione and thiolated chitosan inhibit multidrug resistance P-glycoprotein activity in excised small intestine. J Control Release 111:41–6.
  • Wong HL, Bendayan R, Rauth AM, Wu XY. (2006). Simultaneous delivery of doxorubicin and GG918 (Elacridar) by new polymer-lipid hybrid nanoparticles (PLN) for enhanced treatment of multidrug-resistant breast cancer. J Control Release 116:275–84.
  • Wood N, Streckfus CF. (2015). The expression of lung resistance protein in saliva: a novel prognostic indicator protein for carcinoma of the breast. Cancer Invest 33:510–15.
  • Yang T, Cui FD, Choi MK, et al. (2007). Enhanced solubility and stability of PEGylated liposomal paclitaxel: in vitro and in vivo evaluation. Int J Pharm 338:317–26.
  • Yang W, Cheng Y, Xu T, et al. (2009). Targeting cancer cells with biotin-dendrimer conjugates. Eur J Med Chem 44:862–8.
  • Youle RJ, Strasser A. (2008). The BCL-2 protein family: opposing activities that mediate cell death. Nat Rev Mol Cell Biol 9:47–59.
  • Zhang Y, Liu Y. (2011). Progress in PI3K/Akt signaling pathway and cancer multidrug resistance. Chin J Clinicians (Electronic Edition) 5:446–9.
  • Zhou SF. (2008). Structure, function and regulation of P-glycoprotein and its clinical relevance in drug disposition. Xenobiotica 38:802–32.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.