Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter July 28, 2017

Nanomaterials for cancer therapies

  • Qing Zhou

    Qing Zhou received his BS degree in Biopharmaceuticals from Soochow University in 2012. Since then, he has been pursuing his PhD degree under the supervision of Prof. Hong Wu in the Department of Pharmaceutical Analysis at the School of Pharmacy at the Fourth Military Medical University. His research interest is focused on the pH-sensitive drug delivery systems for cancer therapy.

    , Li Zhang

    Li Zhang received her BS degree in Stomatology from Soochow University in 2012 and her Master’s Degree in Stomatology from the Fourth Military Medical University in 2015. Since then, she is studying as a PhD candidate under the guidance of Prof. Yimin Zhao in the Department of Prosthetic Dentistry at the Fourth Military Medical University. Her current research directions include biofunctionalization of the titanium implant and nanotube for osteogenesis.

    and Hong Wu

    Hong Wu received her BS degree from Northwest University (China) in 1992 and her PhD degree from Xi’an Jiaotong University (China) in 2005. Dr. Wu is the Director of the Department of Pharmaceutical Analysis at the Fourth Military Medical University since 2013. She is now a professor working in the field of nanobiotechnology and nanomedicine to develop various functional nano drug delivery systems for cancer therapy, particular on polymer-based nano-drug carriers.

    EMAIL logo
From the journal Nanotechnology Reviews

Abstract

Cancer is one of the most deadly diseases in the world. In recent years, nanotechnology, as a unique technology, has been comprehensively applied in the therapy of cancer through diagnosis, imaging and theranostics. Additionally, with the emergence of advanced biomaterials which are capable of being applied in biomedical, research in cancer nanotechnology has made significant progress. Particularly, nanomaterials with dimensions below several hundred nanometers are intensively studied among these advanced biomaterials. In past decades, a number of organic and inorganic nanomaterials have emerged as novel tools for cancer diagnostics and therapeutics due to their unique characteristics, like their solubilization effect, drug protection, passive/active tumor targeting, controlled release of drugs which result in enhanced anticancer efficacy while reducing the side effects. In this review, we first provide a brief description of the key properties of nanomaterials, such as nanoparticle (NP) size, surface properties and tumor targeting. The major goal of this review is to summarize the achievements that have been made in the development of the application of nanomaterials for cancer therapies, along with a short description of their general characteristics and preparation of various kinds of nanoparticles.

1 Introduction

In the past few centuries, cancer has been one of the most serious threats to human health. Cancer patients have a sort survival expectation and poor life quality [1]. Although significant progress has been made in medical technology for cancer therapies, the mortality from cancers is still higher than expected and cancer treatment requires further research. Current cancer therapies mainly include surgery, chemotherapy, and radiotherapy. While surgery on many occasions is unable to completely remove all cancer cells in the human body, both chemotherapy and radiotherapy have severe toxic side effects on normal cells and poor specificities for cancer cells. For instance, doxorubicin (DOX), one of the most common chemotherapeutic agents, can induce apoptosis of rapidly dividing cells, but it can also result in apoptosis of numerous normal cells dividing rapidly.

In recent years, there have been significant achievements in the application of nanotechnology, especially in photonics, material science, supramolecular assemblies and drug delivery. In particular, the medical application of nanotechnology, which was called nanomedicine, promoted the development of various kinds of nanoparticles (NPs), such as polymeric micelles, carbon nanotubes, liposomes, etc. [2], [3]. Great effort has been spent on the engineering of nanoparticulate carriers, which function as efficient diagnostic or therapeutic tools for cancer [4], [5], [6]. NPs can accumulate at the site of tumors through enhanced permeability and retention (EPR), decreasing the side effects of drugs and increasing treatment efficiency. Various organic and inorganic nanomaterials have emerged as novel tools for cancer diagnosis and therapy due to their unique characteristics.

In this review, we detail the recent achievements in the development of the application of nanomaterials for cancer therapies with an emphasis on synthetic processes, therapeutic agent delivery and tumor imaging, including inorganic nanomaterials, like carbon nanotubes, silica NPs, gold NPs, magnetic NPs and quantum dots, and emerging organic nanomaterials, such as polymeric micelles, liposomes, dendrimers, etc. (Figure 1 and Table 1).

Figure 1: Representation of nanomaterials, including organic and inorganic nanoparticles, applied for cancer therapy.
Figure 1:

Representation of nanomaterials, including organic and inorganic nanoparticles, applied for cancer therapy.

Table 1:

Examples of nanomaterials for cancer therapies and imaging.

MaterialTypeSizeMain componentMain applications
OrganicPolymeric micelles20–200 nmPolymerTherapeutic and imaging agent carrier
Polymeric NPs10–1000 nmPolymerTherapeutic and imaging agent carrier
Liposomes10–1000 nmLipidTherapeutic and imaging agent carrier
Dendrimers1–15 nmPoly (amidoamine)Therapeutic and imaging agent carrier
Polymer-drug conjugates5–50 nmPolymerDrug carrier
InorganicSilica NPs20–100 nmSilicaTherapeutic and imaging agent carrier
Carbon nanotubes0.4–2 nm

1.4–100 nma
CarbonTherapeutic and imaging agent carrier; photodythermal therapy
Nanographene20–300 nmCarbonDrug carrier; photodythermal therapy
Gold NPs1–100 nmGoldRadiotherapy; imaging; drug carrier
Magnetic NPs10–50 nmIron oxideDrug carrier; magnetic hyperthermia; MRI
Quantum dots2–100 nmMetal compoundFluorescence imaging; photodynamic therapy; drug carrier
  1. aThe diameter of carbon nanotubes varies from 0.4 to 2 nm for SWNTs and from 1.4 to 100 nm for MWNTs.

2 Key properties of nanomaterials

2.1 Size

One of the advantages of nanomaterials is that their size is tunable. The size of NPs used in a drug delivery system should be large enough to prevent their rapid leakage into blood capillaries but small enough to escape capture by fixed macrophages that are lodged in the reticuloendothelial system (RES) [7]. Conventionally, systemically administered NPs should have diameters ranging from 10 to 200 nm, larger than 10 nm in diameter to avoid first-pass elimination through the kidneys while smaller than 200 nm to avoid sequestration by sinusoids in the spleen and fenestra of the liver, benefiting biodistribution and clearance/accumulation behavior [8]. Additionally, as the normal endothelium has an average effective pore size of 5 nm, particles smaller than that will rapidly extravasate across the endothelium [9]. The size of NPs has been shown to influence circulation half-life and tumor accumulation.

2.2 Surface characteristics

In addition to their size, the surface characteristics of NPs also play an important role in determining the life span and fate of nanomaterials during circulation. Surface characteristics can increase a NP’s stability and prolong its circulation in the blood, which then increases passive accumulation in tumors via the EPR effect. Moreover, surface characteristics can dramatically influence the electrostatic and hydrophobic interactions between NPs (aggregation) and clearance by opsonization [10]. Nanomaterials should ideally have a hydrophilic surface to escape macrophage capture. PEGylation is one of the most preferred methods to significantly reduce the binding of plasma proteins, interaction with opsonins and clearance by the RES system, and it refers to the process of covering the NPs with polyethylene glycol (PEG) or its derivatives through grafting, entrapping, adsorbing, or covalently binding to the NP surface [11], [12].

2.3 Passive and active targeting of a tumor

NPs could passively target tumor sites through the EPR effect. Unlike the blood vessels found in most normal tissues, angiogenic tumor vessels usually have larger pore sizes between adjacent vascular endothelial cells, resulting in preferential tumor accumulation of NPs and then release a large amount of the loaded anticancer agents specifically into the extracellular tumor medium or tumor cells, thereby allowing for effective anticancer therapy with minimum side effects [13]. The EPR effect, applicable for almost all rapidly growing solid tumors, is now becoming a guiding principle in tumor-targeting nanomedicine design. The past decades have witnessed that EPR plays an important role in the delivery of nanomaterials to tumors. However, the EPR effect depends on a number of factors such as the surface characteristics of NPs, immunogenicity, tumor characteristics, which thus may result in many challenges in the optimization of passive targeting [14]. Therefore, a full understanding of the degree of these factors in the EPR effect would be worthwhile.

Passive targeting could facilitate the efficient localization of NPs in the tumor interstitium but cannot further promote cellular uptake by cancer cells, however, which can be achieved by actively targeting NPs functionalized with ligands such as antibodies, peptides, nucleic acid aptamers and small molecules (Table 2), to receptors or other surface membrane proteins overexpressed on target cells [15]. This strategy exploits the differences between malignant cells and normal cells. Specific interactions between the ligands on the surface of NPs and receptors expressed on the tumor cells facilitate NP internalization by triggering receptor-mediated endocytosis. Tumor-targeted nanomaterials are being employed for early tumor diagnosis and therapy, which could reduce or eliminate the delivery of potentially toxic agents to healthy tissues, resulting in increased diagnostic or therapeutic effectiveness and decreased potential side effects [16]. Furthermore, active targeting has shown great potential in overcoming a variety of obstacles such as bypassing the blood-brain barrier [17], [18] and multi-drug resistance (MDR) in tumors [19], [20].

Table 2:

Targeting moieties for active tumor targeting.

Targeting moietiesTargetsType of cancer
FAFA receptorMCF-7, MGC803 and KB cells; ovarian cancer
TfTf receptorA549, PANC-1, BT-549, and MDA-MB-435 cells
HACD44 receptorSCC7, MDAMB-231, HCT116, and MCF-7 cells
RGD peptideIntegrin αvβ3U87MG cells; RCC
NGR motif peptideCD13MCF-7 cells; RCC
AptamerPSMALNCaP cells
NucleolinC6 glioma cells
  1. FA, Folic acid; Tf, transferrin; HA, hyaluronic acid; RCC, renal cell carcinoma; PSMA, prostate-specific membrane antigen.

3 Organic nanomaterials for cancer therapies

Because of the excellent properties, such as biological compatibility and degradability, natural or synthetic polymer-formed organic-based nanomaterials have been extensively used in the field of cancer therapies. They can roughly be divided into five types, i.e. polymeric micelles, polymeric NPs, liposomes, dendrimers and polymer-drug conjugates.

3.1 Polymeric micelles

3.1.1 General characteristics and synthetic aspects

Polymeric micelles (PMs), prepared from certain amphiphilic co-polymers, are spherical, nano-sized colloidal particles with a hydrophilic shell and a hydrophobic core [21], [22], [23], [24] (Figure 2). PEG is the most used in constructing the hydrophilic shell, which can help to stabilize the carriers and protect them from degradation by reducing unspecific interactions in vivo [25], [26]. To construct the hydrophobic core, several natural or synthetic polymers have been frequently taken into consideration, including polysaccharides, poly(ε-caprolactone) (PCL), poly(lactide) (PLA) and poly(lactic-co-glycolic acid) (PLGA). The hydrophobicity of the hydrophobic core provides a perfect medium to entrap hydrophobic drugs, helping to solve their poor water solubility. Except for the ability of solubilization of hydrophobic drugs, the small size with narrow distribution also makes polymeric micelles ideal nano-drug delivery systems, as it could avoid rapid renal excretion, helping to realize a long circulation time [27], [28].

Figure 2: Schematic of optimized polymeric micelle for anticancer therapy, bearing targeting ligands, contrast agents or imaging moieties and therapeutic drugs.
Figure 2:

Schematic of optimized polymeric micelle for anticancer therapy, bearing targeting ligands, contrast agents or imaging moieties and therapeutic drugs.

In the preparation of polymeric micelles, we need to compose a desirable amphiphilic block copolymer, and then a conversion of micelle at a critical micelle concentration (CMC) through several techniques including the dialysis method, the oil-in-water (o/w) emulsion method, the solvent evaporation method, the nano-precipitation method and the solid dispersion method [29], [30]. In the dialysis method, polymer and hydrophobic drugs are first dissolved in organic solvent, then the addition of small amounts of water and stirring is followed to form a water-miscible organic solvent. Finally, a dialysis bag is used for dialysis against excess water for several hours to remove the organic solvent [31]. As to the solid dispersion method, similar to the dialysis method, polymer and hydrophobic drugs are dissolved in organic solvent first. Then, a solid polymer matrix could be obtained by evaporating the organic solvent under reduced pressure [32]. In addition, except for physical entrapment, drug loading can also be completed by chemical conjugation.

3.1.2 Polymeric micelles for drug delivery

As they could load and deliver drugs to the desired function site, improving the pharmacokinetic of the loaded drug and decrease non-specific toxicity, polymeric micelles have been extensively explored as drug carriers in the past decades. In a recent study, in order to increase the drug-loading content, Liang et al. [33] used different p-conjugated moieties, including 7-carboxymethoxy coumarin, cinnamic acid and chrysin, to modify the terminal hydroxyl groups of PCL segments in mPEG-PCL micelles, as DOX could be loaded on the micelles through the evocation of the p-p stacking interaction. The results showed that the modification could improve the mPEG-PCL’s ability of crystallization; therefore, the drug-loading content increased significantly from 12.9% to 25.5%. As cRGD peptide could inhibit cancer metastasis, Makino et al. [34] prepared cRGD-conjugated polymeric micelles to impede the development of lymph node metastasis. Besides, an anticancer drug (1,2-diaminocylohexane)platinum(II) was entrapped into the micelles to further inhibit the metastasis progress. The results showed that the micelles could effectively inhibit lymph node metastasis by inhibition of the spreading of cancer cells.

Because the natural pH gradient exists in the tumor microenvironment and intracellular endo/lysosome, pH-sensitive degradable micelles are recently emerging as a promising platform for antitumor drug delivery [35], [36]. Recently, intracellularly acid-switchable micelles were prepared by Wang et al. [37] for accomplishing combinational therapy against drug-resistant tumors. The micelles were composed of a pH-sensitive diblock polymer, a photosensitizer and DOX. The results revealed that the activated micelles can induce notable reactive oxygen species (ROS) generation under near-infrared (NIR) laser irradiation for photodynamic therapies (PDT) and promoting tumor penetration of the chemotherapeutics. Furthermore, the micelles can cause significant temperature elevation via NIR laser illumination for PA imaging and photothermal therapies (PTT).

In our previous work, the pH-sensitive mixed micelles were prepared for cytosolic delivery of DOX [24]. The mixed micelles were composed of two block polymers, DSPE-PEG and pH-responsive poly(histidine)-PEG. The results showed that the micelles possessed a pH-dependent drug release property. Furthermore, while modified with nucleosome-specific monoclonal antibody, the micelles showed greatly enhanced endocytosis efficiency and antitumor efficacy. Similarly, we prepared other pH-sensitive micelles to protect and introduce the penetrating peptide (TAT) [38]. The micelles were composed of two block polymers, one was DOX-TAT and the other one was LHRH-PEG-PHIS-DOX, in which LHRH was an active targeting moiety, and PHIS was a pH-sensitive bond. Once they arrive at the tumor microenvironment, the micelles could respond to tumor extracellular pHe and dissociate and release DOX-TAT, which could overcome multidrug resistance and transport directly into cancer cells. The in vitro and in vivo results demonstrated that this kind of micelles showed the highest anticancer efficacy compared with the control group.

3.2 Polymeric nanoparticles

3.2.1 General characteristics and synthetic aspects

Polymeric NPs, one of the hotspot issues in the field of nanomedicine, present superior pharmacokinetic properties, including drug load and drug stability, compared to polymeric micelles. Composed of polymer matrix, NPs provide the media for the drugs absorption, dissolution, entrapment and encapsulation [39], [40], [41]. They could improve the therapeutic effect of anticancer treatment through passive targeting via the EPR effect.

Several methods are used for the synthesis of polymeric NPs, including emulsification and solvent evaporation/extraction [42], nanoprecipitation (solvent-displacement) [43], [44], supercritical antisolvent method [45], and salting out [46], [47]. Emulsification and solvent evaporation/extraction was the most used method to synthesize polymeric NPs. In this method, the polymer organic solution (oil phase) is first added into stabilizer-contained water to form the oil-in-water emulsion. Then, the organic solvent could be evaporated either under reduced pressure or by continuous stirring at room temperature. Evaporation of the organic solvent is followed by precipitation of the polymer as nanospheres of a few hundred nanometers in diameter. In 2014, a new method named electrospray nanoprecipitation was reported by Luo et al. [48] to prepare polymeric NPs, which combines agitated solvent displacement with electrospray. Compared with the agitated solvent displacement or the electrospray method, this combined technique could obtain polymeric NPs with a dramatically lower diameter range.

3.2.2 Polymeric nanoparticles for drug and gene delivery

Because of the superb characteristics, like biodegradability, biocompatibility and prolonged circulation, the polymeric NP-based delivery system for anticancer drugs has received extensive attention. In addition, polymeric NPs could deliver a great number of drugs for cancer therapies, including anticancer drugs, genes and proteins. Moreover, polymeric NPs may significantly reduce the adverse effects due to an alteration of the body distribution.

PEGylated PLGA NPs were employed by Danhier et al. [49] as a vehicle for the delivery of paclitaxel (PTX). Compared to Taxol, the polymeric NPs showed a threefold higher cytotoxicity on HeLa cells, resulting in remarkable inhibition of tumor growth. Polymeric NPs can be also used as a gene vector for extending brain tumor survival. Gene therapies have most often been performed using viral carriers. However, viruses pose significant safety risks due to their inherent toxicity, immunogenicity and tumorigenicity. In a recent study, polymeric NPs were introduced to deliver genes to extend in vivo brain tumor survival [50]. Two rat glioma cell lines, 9L and F98, were used to evaluate the efficacy of such NP formulations, and it was discovered that these treated animals displayed an enhanced survival rate compared to commercial reagent Lipofectamine 2000.

Hypoxia is a condition found in various intractable diseases including cancer. Thus, a hypoxia-responsive polymeric NP (HR-NPs) was prepared by Thambi et al. [51] for the tumor-targeting delivery of DOX. The HR-NPs were synthesized by carboxymethyl dextran conjugated with 2-nitroimidazole derivative, while DOX was entrapped into the NPs. Compared to normoxic cells, the DOX-loaded HR-NPs revealed higher cytotoxicity against hypoxic cells. As confirmed by microscopic observation, under hypoxic conditions, HR-NPs could promote cellular uptake of DOX, resulting in great antitumor efficacy in vivo.

3.2.3 Polymeric nanoparticles for imaging

Polymeric NPs loaded with imaging agents, namely, gadolinium complexes and magnetic NPs have been extensively explored to image cancer by magnetic resonance imaging (MRI). Typically, imaging agents were encapsulated into the core of the polymeric NPs. For instance, Li et al. [52] developed a novel folate-conjugated PEGylated PLGA-based NP for liver cancer therapeutics; sorafenib and magnetic NP were co-encapsulated into the NP. The results revealed that the polymeric NP possessed good imaging property. Furthermore, this NP formulation displayed great inhibition of tumor cell growth.

Recently, in order to realize combined diagnostic and therapeutic functions, Li et al. [53] prepared a theranostic NP, which was synthesized by the integration of a gadolinium-based MRI agent and an active gemcitabine metabolite through supramolecular self-assembly synthesis. The anticancer drug gemcitabine-5′-monophosphate was entrapped in such a NP for coordination with Gd(III). Thetheranostic NPs showed a strong T1 contrast signal for in vivo MRI of tumors and increased retention time. Moreover, the results showed that such a NP formulation could greatly inhibit the growth of MDA-MB-231 tumor in vivo.

3.3 Liposomes

3.3.1 General characteristics and synthetic aspects

Liposomes are self-assembling NPs with closed membrane structures. They are formed by dispersion of phospholipids featured with hydrophobic anionic/cationic long-chain tails and hydrophilic heads (Figure 3). Their specific structures enable water-soluble drugs to be entrapped in their aqueous core, while lipophilic drugs in the lipid bilayer(s) [54], [55]. In addition, liposomes can effectively load various bioactive molecules, including enzymes and nucleic acids [56], [57]. They have been proven to be beneficial for therapeutic compound stabilization, cellular and tissue uptake of therapeutic compounds and bio-distribution of compounds to target sites in vivo [58], [59], [60].

Figure 3: Schematic of liposomes used in cancer therapy. Liposomes are composed of lipids with PEG to increase their circulation time. They can be functionalized with therapeutic agents, targeting ligands, and imaging moieties or contrast agents.
Figure 3:

Schematic of liposomes used in cancer therapy. Liposomes are composed of lipids with PEG to increase their circulation time. They can be functionalized with therapeutic agents, targeting ligands, and imaging moieties or contrast agents.

Liposomes can be created from cholesterol and natural nontoxic phospholipids. Typically, liposomes can be synthesized in a solution containing the desired drug by hydration of the phospholipid mixture. After vortexing, unilamellar liposomes are extruded under high pressure, which can be further purified by column chromatography or ultracentrifugation. Jaafar-Maalej et al. [61] reported a new method to prepare liposome via the ethanol-injection technique and a membrane contactor module, which was specifically designed for colloidal system preparation. This method promoted pharmaceutical agents to be entrapped into liposomes.

3.3.2 Liposomes for drug delivery

In the past decades, liposomes have been extensively explored as an anticancer drug delivery system. They could be used to stabilize therapeutic agents, promote targeted site accumulation and endocytosis of these agents in vivo. At present, several liposomal anticancer drugs have been successfully applied into the clinic or clinical test stages [62], [63], [64]. For instance, Doxil, a PEGylated liposomal DOX, has been approved by the US Food and Drug Administration (FDA) for cancer therapy, as it could improve the plasma pharmacokinetics and tissue distribution.

In a recent study, H7K(R2)2-modified pH-responsive liposomes were prepared by Zhao et al. [65] for the tumor-targeting delivery of DOX. H7K(R2)2 referred to a targeting ligand but can respond to the mild acidic pH in the glioma microenvironment, acting as a cell-penetrating peptide (CPP). The in vitro results confirmed this kind of pH-triggered specific-targeting effect, while in vivo results showed a liposomal formulation could greatly improve the antiangiogenic activity.

In another similar work, Chiang et al. [66] introduced pH-responsive polymer-liposomes for the extracellular matrix (ECM)-targeting delivery of anticancer drugs. The results indicated that such ECM-targeting liposomes showed rapid drug-releasing profiles in acidic conditions. Because the ECM-targeting liposomes accumulated preferentially in the tumor site, the ECM-targeting liposomes showed exceptional anticancer activity in vivo and lower hepatic and renal toxicity.

3.3.3 Liposomes for imaging

Except for drug delivery, liposomes can also be functionalized with imaging contrast agents, providing combined diagnostic and therapeutic functions. Recently, a theranostic liposomal drug delivery system was prepared by Ren et al. [67] to realize a real-time image of bio-distribution by MRI and accomplish chemotherapy through the carried anticancer drug. Compared to commercial MRI contrast agent Omniscan®, this liposome showed a 36-fold higher T1 relaxation rate; moreover, its circulation time could reach 300 min in vivo. Besides, both hydrophobic and hydrophilic chemotherapeutic drugs could be loaded in such a liposome, resulting in synergetic therapy without noticeable toxicity.

Park et al. [68] prepared hyaluronic acid derivative-coated liposomes for tumor-targeting delivery of DOX and MRI contrast agent (Magnevist). Under HA and CD44 receptor interaction, the cellular uptake of DOX greatly increased compared to plain liposome. Furthermore, such a liposome appeared to be a promising tumor-targeting MRI probe for cancer diagnosis, which was confirmed by the in vivo contrast-enhancing effects.

3.4 Dendrimers

3.4.1 General characteristics and synthetic aspects

Dendrimers, as perfect monodisperse macromolecules, are characterized with a highly branched 3D architecture [5], [69], [70], [71]. They could load drugs and gene molecules through simple electrostatic interactions, encapsulations and covalent conjugations. Dendrimers possess empty internal cavities and an extremely higher density of surface functional group (-NH2 or -COOH), which makes them become attractive carriers for anticancer therapeutics. Moreover, due to the relatively small size, dendritic carriers (1–15 nm), can be cleared from the blood through the kidneys, which can decrease their toxicity in vivo.

Dendrimers can be prepared by the divergent and convergent synthesis approaches. In the divergent approach, the synthesis begins with the preparation of the core of the dendrimer, then the arms are attached to the core by adding building blocks in a step-by-step manner [72]. Most commercially available dendrimers are produced by the divergent approach. The convergent approach involves preassembly of the complete wedge-shaped branching units, which are coupled to the central core moiety in the final step [73]. This strategy can provide the best structural control by necessitating the synthesis of branches of various requisite sizes [74].

3.4.2 Dendrimers for drug and gene delivery

In recent years, dendrimers have been extensively introduced to act as drug carriers, and a variety of drugs can be entrapped or covalently conjugated to the dendrimers. Among all the dendrimers, polyamidoamine (PAMAM) was the most widely investigated dendrimer as its surface contains a great number of amine groups, which could be used to conjugate various functional moieties. Liu et al. [75] combined liposomes and dendrimers (PAMAM, G4.0) into hybrid nano-systems for the encapsulation of PTX to treat ovarian cancer. Compared with free PTX, the anticancer efficiency of PTX was greatly enhanced (37-fold higher) as it loaded in such a nano-formulation. Moreover, the anticancer effect could be improved by the combined function of PTX and the dendrimer.

A folate-PEGylated PAMAM dendrimer was prepared by Cheng et al. [76] for the delivery of DOX and enhanced tumor selectivity. DOX was covalently attached to PAMAM through an acid-labile linker. The results revealed that such a dendrimer formulation showed what was evidently pH-dependent drug release pattern. In addition, the cellular uptake of DOX was greatly improved, which was confirmed by fluorescence microscopy and flow cytometry. In 2014, Liu et al. [77] for the first time, prepared the G5 PAMAM dendrimer-based targeting nano-delivery system for the delivery of Hsp27 dsiRNA. Besides, a dual targeting peptide was used to decorate the dendrimer for simultaneously promoted cancer cell targeting. The results showed that this kind of nano-system could promote cancer cell targeting, resulting in enhanced gene silencing and anticancer efficacy.

3.4.3 Dendrimers for imaging

The unique morphology of dendrimers makes them a promising candidate for diagnostic applications. For example, in order to realize in vitro and in vivo computed tomography (CT) imaging of cancer cells, Wang et al. [78] introduced an acetylated dendrimer to entrap gold NPs. After treatment with such dendrimers, SPC-A1 cells can be detected by micro-CT images under X-ray. Moreover, the in vivo results also confirmed that the synthesized dendrimers’ administration could be imaged, whether intratumor or intraperitoneum.

In 2013, multifunctional dendrimer-based gold NPs (AuNPs), as a dual-modality contrast agent, were prepared by Li et al. [79] for in vitro and in vivo CT/MRI of breast cancer cells. In this study, gadolinium chelate (DOTA-NHS) and PEG monomethyl ether-modified G5 PAMAM dendrimers were used as templates to synthesize AuNPs, then the remaining dendrimer terminal amine groups were chelated and acetylated by Gd(III). The results showed that MCF-7 cells and the xenograft tumor model after treatment with such a contrast agent could be effectively imaged by CT or MR imaging. Recently, radionuclide (131)I-labeled dendrimers were prepared by Zhu et al. [80] for targeted single-photon emission computed tomography (SPECT) imaging and radiotherapy of tumors. In this study, G5-NH2 were first modified with 3-(4′-hydroxyphenyl)propionic acid-OSu (HPAO) and PEGylated folic acid (FA), and then the remaining dendrimer terminal amine groups were acetylated by radioactive iodine-131 [(131)I]. The results showed that such a dendrimer platform possessed good stability and high radiochemical purity for in vivo targeting SPECT imaging and radiotherapy on the FA receptor-overexpressing xenografted tumor model.

3.5 Polymer-drug conjugates

3.5.1 General characteristics

Polymer-drug conjugates are synthesized through conjugating drugs to water-soluble polymers, which contain a water-soluble polymer, an agent and/or a biodegradable linkage [81], [82], [83]. They could increase the tumor-targeting ability via the EPR effect and enable endocytic capture at the cellular level, which could promote lysosomotropic drug delivery. In addition, in spite of the development of multi-drug resistance, distinct cell uptake mechanisms make polymer-drug conjugates less sensitive to efflux pumps. To date, drug conjugates are the most successful nanomedicine therapeutics that are applied in clinical cancer treatment [82], [84], [85], [86]. Although many other devices achieved great progress, polymer-drug conjugates are still favored due to their distinct advantages.

3.5.2 Polymer-drug conjugates for drug delivery

Polymer-drug conjugates are drug delivery systems. Drug(s) can covalently attach to the functional groups of the polymer directly or through a spacer. They have been an excellent platform for drug delivery. Tu et al. [87] prepared a stimuli-responsive polymer-drug conjugate for improving cancer-targeting ability and anticancer efficacy. In this study, this polymer-drug conjugate consisted of a PEG, a matrix metalloproteinase 2 (MMP2)-sensitive peptide linker, a TAT and DOX. The polymer-drug conjugate possessed multifunctional properties, including an MMP2-mediated tumor-targeting ability and P-gp inhibition ability, resulting in enhanced cell endocytosis and followed by improved antitumor efficacy.

In our previous research, we prepared a pH-response charge convertible polymer-drug conjugate for the delivery of antitumor drugs [88]. In this polymer-drug conjugate, poly(β-l-malic acid) (PMLA) was used as a water-soluble polymer and was modified by polyethylenimine (PEI), a fragment antibody (HAb18 F(ab′)2), charge convertible group (2,3-dimethylmaleic anhydride) and DOX, which were covalently attached to the polymer. The in vitro results showed a significantly enhanced endocytosis, which was caused by receptor-mediated endocytosis and strong positive charge, resulting in increasing antitumor efficacy.

Owing to the complex molecular basis, single therapeutic agents could not effectively sustain cancer therapies. The majority of cancers are being treated by a combination of drugs [89]. Therefore, polymer-drug conjugate-based combination therapy would be an exciting therapy, which could achieve simultaneous delivery, providing synergetic drug effects [81]. In order to achieve combination therapies, DOX and mitomycin C (Mit C)-conjugated HPMA copolymer-based polymer-drug conjugate was prepared by Kostková et al. [90]. In this polymer-drug conjugate, DOX and Mit C, two drugs that have different action mechanisms, were conjugated to the polymer through a pH-sensitive hydrazone bond. Both in vitro and in vivo results showed a significant synergetic antitumor activity on EL-4 cells. In another similar work, a polymer-drug conjugate co-delivery of oxaliplatin and demethylcantharidin (DMC) was synthesized by Wang et al. [91]. The results revealed that this polymer-drug conjugate, which could be activated in two modes against cancer cells, resulting in higher cytotoxicity against SKOV-3 cells compared to free drugs used alone, which was mainly caused by the synergistic effect between oxaliplatin and DMC.

4 Inorganic nanoparticles for cancer therapies

Besides organic NPs used in the field of cancer therapies, various inorganic materials with interesting structures as well as unique chemical and physical properties have also been explored as NPs for cancer therapies.

4.1 Silica nanoparticles

4.1.1 General characteristics and synthetic aspects

As a major component of sand, silica is known for its compatibility in biological systems. A large variety of silica-based nanostructures have been synthesized in the past few decades and used in different applications [92]. The particle size, shape, porosity and surface chemistry of silica NPs can be successfully controlled during the synthesis process. It has been found that silica-based NPs with a special nanostructure can be used to encapsulate various antitumor agents for cancer therapies [93], [94], [95], [96], [97], [98].

Two major types of silica-based NPs have been successfully synthesized and studied. One is solid silica NPs (SiNPs) and the other is mesoporous silica NPs (MSNs). In recent years, functionalized SiNPs have been widely used as drug delivery carriers and optical imaging contrast agents. When used as optical contrast agents, SiNPs show remarkable properties, including biocompatibility, photophysical stability and favorable colloidal properties. Besides, the particle surface of silica can be modified with a series of functional groups such as aptamers and antibodies [99], [100]. MSNs also holds many special properties such as stable and rigid frameworks, high surface areas, large pore volumes, tunable pore sizes and have shown great promise as drug delivery vehicles during the past several decades [101], [102], [103], [104].

There are two main methods that are widely used to synthesize SiNPs, including the Stöber method and the reverse microemulsion method [105], [106]. The Stöber method mainly involves two processes, the controlled hydrolysis and condensation of a silica precursor. Tetraethoxysilane (TEOS) in ethanol and water was used in the Stöber method, and ammonia was used as a catalyst. The size of the particles can be controlled by adjusting the synthesis conditions. For the reverse microemulsion method, the synthesis process involves the ammonia-catalyzed polymerization of TEOS in water-in-oil or a reverse-phase microemulsion. The micelles of the microemulsion act as “nanoreactors”, where the particle growth is carried out. The final size of SiNPs is controlled by the water-to-organic solvent ratio. MSNs are usually prepared by the modified Stöber method, which relies on the cooperative self-assembly of supramolecular surfactant assemblies, acting as structure-directing agents and oligomeric silica species. The synthesis process can be conducted in either acidic or basic conditions.

4.1.2 Silica nanoparticles for drug and gene delivery

MSNs have been extensively used for drug and gene delivery due to their unique mesopores and nanochannels, which can render a high payload of the drug and easy stimuli-controllable release [101], [102]. For example, in order to overcome drug resistance in breast cancer, a multifunctional MSN carrier was prepared by Meng et al. [107] to co-deliver DOX and siRNA, which could target the P-glycoprotein (Pgp) drug exporter. This kind of delivery system showed a synergetic inhibition of tumor growth in vivo, providing a promising platform for multiple drug delivery and overcoming DOX resistance. In a recent work, Ngamcherdtrakul et al. [108] reported a PEI-PEG-coated and anti-HER2 monoclonal antibody (trastuzumab)-coupled MSNs for the targeting delivery of siRNA to breast cancer. The results showed that the synthetic antibody-NPs could induce cell apoptosis in HER2-positive breast cancer but not in HER2-negative (HER2) cells in vitro. More importantly, this synthetic antibody-NPs showed no noticeable systemic toxicity.

In our previous research, a SiO2@AuNP sequential drug delivery system was introduced to deliver three kinds of different drugs, including bio-drugs (monoclonal antibody), siRNA and a combination of chemotherapeutics with a synergetic effect [109]. Such a multifunctional NP system could target the colon cancer cell (colo-205) under mAb guidance, followed by endocytosis, endosome escape and sequential drug release. The siRNA released by the glutathione exchange was used to inhibit the Bcl-2 protein, while the sequential release of hydroxycamptothecine (HCPT) and DOX played a role in synergetic antitumor efficacy. The in vitro and in vivo results showed significantly increasing antitumor efficacy without noticeable systemic toxicity, providing a promising platform for multiple drug delivery and controllable sequential drug release.

4.1.3 Silica nanoparticles for imaging

In recent years, SiNPs loading fluorescent dye agents have achieved extensive interest in cancer imaging due to their brightness and photostability. Recently, a multifunctional theranostic magnetic MSN was prepared by Chen et al. [110] to achieve magnetic-enhanced tumor-targeting MRI. In order to realize reduction-triggered intracellular drug release, β-cyclodextrin (β-CD), acting as the gatekeeper, was anchored on the surface of MSN. The results showed that such MSN can realize active targeting of cancer cells and enhanced MRI in vivo under an external magnetic field, resulting in significantly increased antitumor efficacy without obvious systemic toxicity.

Iron oxide NPs seem to be promising diagnostic or therapeutic agents for cancer treatment. However, their in vivo application was hindered by the severely biological aggregation. Therefore, Hurley et al. [111] synthesized mesoporous silica-coated iron oxide NPs to prevent the aggregation. The results suggested that magnetic NPs coated with MSNs could avoid lower heating and imaging performance, which will be caused by aggregation. In another work, Chen et al. [112] developed a type of functionalized MSNs, coated with fluorescence-conjugated cyclodextrin, for pH-trigged drug delivery and imaging. In vitro results showed that drug release behavior can be controlled by the pH-responsive gatekeeper (cyclodextrin), which shall open the pores when facing acid environment. Moreover, such DOX-loaded functional MSNs showed greatly improved antitumor activity on both HepG2 and HeLa cells.

4.2 Carbon Nanomaterials

In recent years, carbon-based nanomaterials, namely carbon nanotubes (CNTs), fullerenes and graphene, have been developed as nanocarriers for drug delivery and as contrast probes for biomedical imaging. The carbon nanomaterials possess superb biocompatibility, sufficient surface-to-volume ratio, good thermal conductivity and rigid structural properties for post-chemical modification. They could deliver antitumor drugs and gene into cancer cells efficiently, in which the therapeutic molecules shall be safely carried, and then display enhanced antitumor efficiency [113], [114].

4.3 Carbon nanotubes

4.3.1 General characteristics and synthetic aspects

CNTs, which were first reported by Iijima in 1991, are made up of thin sheets of benzene ring carbons and rolled up into the shape of a seamless tubular structure [115]. Two categories of carbon nanotubes have been discovered, one is the single-walled carbon nanotubes (SWNTs) and the other is the multi-walled carbon nanotubes (MWNTs) [116]. Because of their unique properties in electrics, thermotics, optics, mechanics and biology, CNTs have largely been used in cancer therapies. In recent years, almost everywhere, in relation to cancer treatment, CNTs can be applied, including drug and gene delivery, photodynamic therapies and thermal therapies.

Many methods to synthesize CNTs have been reported over the past several years. The main synthesis techniques for CNT construction are of three types, including arc discharge, laser ablation and chemical vapor deposition [117], [118], [119]. All methods involve using a source of carbon and energy to form CNTs. Arc discharge is the most reported method for the synthesis of CNTs as it is very easy. In this method, carbon electrodes act as a source of carbon; a potential difference of about 20 V acts as a source of energy. In the laser ablation method, the carbon electrodes are used as a carbon source while laser pulses provide the energy source. Chemical vapor deposition is the most widely used method for the large-scale production of CNTs. In this method, hydrocarbons like CH4, acetylene, or carbon monoxide are utilized as a carbon source, and high temperature was used to provide sufficient energy for the decomposition of the hydrocarbons to form CNTs.

4.3.2 Carbon nanotubes for drug delivery

Because of tunable surface modifications and excellent physicochemical properties, CNTs have been extensively explored as drug carriers, and a variety of related research has been reported in the past several years [120], [121]. Liu et al. [122] reported SWNT-based drug delivery for the inhibition of tumor growth in vivo. PTX was conjugated to the branched PEG chains on SWNTs through a cleavable ester bond. Compared to clinical Taxol, such a SWNT-based formulation showed higher efficacy in inhibiting tumor growth, which was caused by prolonged blood circulation and much higher PTX accumulation at the tumor site through the EPR effect.

Recently, Al Faraj et al. [123] developed CD44 antibody-conjugated SWCNTs for targeting delivery of combining drugs to breast cancer and cancer stem cells (CSCs). The SWCNTs were modified with PEG for improved pharmacokinetics; combining PTX and salinomycin drugs as the conjugated carriers through a hydrazine bond, providing pH-dependent drug release. Compared with treatment using individual drug-conjugated nanocarriers or free drugs, this combined treatment showed considerably enhanced therapeutic efficacy against both breast cancer and CSCs, which was confirmed by bioluminescence and MRI. Zhang et al. [124] used modified SWCNTs for tumor-targeting delivery and controlled release of DOX. This system displayed superb stability under physiological conditions, while at acidic pH and conditions such as tumor microenvironment and intracellular endosomes and lysosomes, the DOX were quickly released and entered the cell nucleus, resulting in cell apoptosis. The results demonstrated that this nanoscale drug system possesses higher selectivity and effectiveness than that of the free drug.

4.3.3 Carbon nanotubes for imaging

As one of the darkest materials, CNTs could exhibit excellent absorbance in the NIR region; thus, they could be utilized as imaging contrast agents, making them achieve extensive attention in the field of cancer imaging. Recently, Al Faraj et al. [125] used CD44 antibody-conjugated SWCNTs as novel multimodality nanoprobes to realize specific targeting and imaging of breast CSC. MRI, single-photon emission CT, and NIR fluorescence imaging noninvasive imaging moieties were used to monitor the biodistribution and preferential homing of such SWCNTs toward the tumor site. The results showed that anti-CD44 SWCNT specifically target CD44-overexpressed cancer cells, and an enhanced colocalization was observed.

Similarly, in order to realize the targeted photoacoustic imaging of gastric cancer, Wang et al. [126] developed RGD-conjugated silica-coated gold nanorods, which were attached to the surface of MWNTs. The results revealed that these kinds of probes possessed excellent water solubility and low non-specific toxicity; moreover, they can target gastric cancer cells and obtain strong in vivo photoacoustic imaging.

Wu et al. [127] developed prostate stem cell antigen (PSCA) antibody-conjugated MWNTs [CNT-PEI(FITC)-mAb] for drug delivery and targeted ultrasound imaging. The results showed that such a MWNT-based formulation could realize specific targeting of PSCA-overexpressed cancer cells. In addition, CNT-PEI(FITC)-mAb showed high promise to be used as a targeted ultrasound contrast agent, which was confirmed in vitro and in vivo ultrasound imaging. Furthermore, the in vivo anticancer efficacy testing demonstrated that CNT-PEI(FITC)-mAb could realize targeted delivery of anticancer drugs to the tumor and inhibit the growth of tumor cells.

4.3.4 Carbon nanotubes for photothermal therapies

In the past several years, a variety of researches have demonstrated that CNTs possessed great potential to be applied for photothermal ablation of cancer cells due to their strong optical absorption in the NIR region [128], [129], [130], [131], [132], [133], [134]. Recently, manganese oxide (MnO)-coated CNTs were prepared by Wang et al. [129] to act as dual-modality lymph imaging agents for photothermal treatment of lung cancer metastasis. MnO-coated CNTs further covered by PEG was explored as a lymphatic theranostic agent to realize diagnosis and treat metastatic lymph nodes. The results showed that the desired regional lymph nodes were clearly imaged by T1-weighted MR of MnO and dark dye imaging of MWNTs. Furthermore, under the guidance of dual-modality imaging, metastatic lymph nodes could be killed by NIR irradiation as well.

In order to improve and optimize the antitumor efficacy, Zhang et al. [133] prepared Ru(II) complex-functionalized SWCNTs for PTT and PDT. Ru(II) complexes were covalently loaded on the SWCNTs by π-π interactions, and they could be efficiently released under the photothermal effect. Under two-photon laser irradiation, Ru(II) complexes can produce singlet oxygen species; thus, it could be used as a two-photon PDT agent. Compared to single therapy, such Ru(II) complex-functionalized SWCNTs possessed greater anticancer efficacy due to the combination of PTT and two-photon PDT, which was also confirmed by in vivo experiments. In 2011, Zhou et al. [134] reported a novel SWNT-based drug delivery system, which could realize mitochondria targeting for cancer PTT. This kind of SWNTs can efficiently convert 980-nm laser energy into heat and selectively destroy the targeting mitochondria, resulting in mitochondrial depolarization, caspase 3 activation, and cytochrome c release. The results revealed that the SWNTs+laser process could afford excellent efficacy in inhibiting tumor growth in the breast cancer model; moreover, in some cases, they could induce complete tumor regression.

4.4 Nanographene

4.4.1 General characteristics and synthetic aspects

Graphene, with a two-dimensional (2D) honeycomb structure composed of an atom-thick monolayer of carbon atoms, was first separated from graphite in 2004, and it is believed to be the basic structure (unit) for the composition of other graphitic materials [135], [136], [137]. Recently, graphene and its derivatives such as graphene oxide (GO) and reduced graphene (rGO) have attracted tremendous interest in biomedicine owing to their special physical, chemical and optical properties [138], [139], [140], [141], [142], [143]. During the past 10 years, their application in disease treatment based on various mechanisms has been widely explored, and extensive studies have been performed to develop graphene-based NPs for cancer therapies.

GO is an oxygenated derivative of graphene, which has the structure of a single-atom thick layer of graphene sheets and ranges from tens to several hundred nanometers according to the oxidizing conditions. It can be prepared using the Brodie, Staudenmaier, Hummers, and improved Hummers’ methods [144], [145], [146]. With large amounts of epoxide, carboxylic acid and hydroxyl groups on its surface, GO has good dispersibility and biocompatibility. Meanwhile, GO can be bio-functionalized by the introduction of various functional groups to synthesize GO derivatives with different properties and functions. To obtain rGO, GO is treated by chemical-reducing, thermal-reducing, or UV-reducing processes [147].

4.4.2 Nanographene for drug delivery

Nanographene has properties suitable for the delivery of various molecules. With its two-dimensional structures, nanographene provides relatively high surface areas and capacity for non-covalent π-π stacking and hydrophobic interactions with various drug molecules. In particular, GO could provide a great number of residual carboxylic acid, epoxide groups and hydroxide on its surfaces. Thus, it can also load various agents via covalent conjugation, hydrogen bonding and electrostatic interaction [148], [149], [150].

Recently, Zheng et al. [151] prepared poly-l-lysine-functionalized rGO and conjugated with anti-HER2 antibody for efficiently targeting the delivery of DOX to the nucleus of cancer cells. The results showed that cellular uptake of this kind of nanocarriers into MCF7/HER2 cells was significantly higher than the nanocarriers without anti-HER2 antibody, resulting in improved antitumor efficacy. Tian et al. [152] developed a GO-based nanocarrier, which can achieve targeted delivery of anticancer drugs by conjugation of the folate and self-monitoring both in vitro and in vivo as a labeled fluorescein peptide. The results showed that the drug-loaded nanocarrier can specifically carry the drugs into the folate receptor high-expressed cancer cells. In addition, drug-induced cancer cell apoptosis could be visualized by confocal fluorescence imaging. Living imaging in tumor-bearing mice further demonstrated the dual function mentioned above.

In 2013, Miao et al. [153] synthesized new rGO nanosheets, coated by cholesteryl hyaluronic acid (CHA), for the delivery of anticancer drugs. Compared to rGO, CHA-coated rGO nanosheets could display enhanced colloidal stability and increased safety in vivo. In vivo experiments revealed that CHA-coated nanosheets showed higher tumor accumulation than nanosheets without CHA coating, therefore, resulting in significantly improved antitumor efficacy.

4.4.3 Nanographene for photothermal therapies

Apart from its nature of being an excellent drug carrier, due to high NIR absorbance, nanographene can also serve as a photothermal agent that could convert the absorbed light into heat, inducing hyperthermia to the cancer cells [154]. For example, a dual anticancer drug-loaded GO was prepared by Tran et al. [155] to achieve a dual-in-dual synergetic treatment against cancer cells. Poloxamer 188 was used to stabilize GO and generate heat to induce cancer cell apoptosis under NIR laser irradiation. Compared with chemotherapy or photothermal therapy used separately, the combined therapy displayed a synergetic effect, resulting in higher antitumor efficacy.

In order to improve tumor targeting and photothermal treatment, a dual-ligand (folate and cRGD) targeting nGO was developed by Jang et al. [156]. The results revealed that cellular uptake of dual-ligand targeting nGO was much higher than single targeting FA-nGO or cRGD-nGO. cRGD-FA-nGO also showed a greatly higher tumor accumulation, resulting in complete ablation of tumor cells by photothermal therapy. In 2015, Kim et al. [143] prepared novel rGO-based NPs, conjugated with PEG-g-poly(dimethylaminoethyl methacrylate) and hyaluronic acid, for targeting PTT and pH-sensitive bioimaging. The rGO-based NPs could respond to the tumor environment to generate photothermal heat, resulting in the suppression of tumor growth. Moreover, the malignant tumor could be completely healed, which was demonstrated by histopathological studies.

4.5 Gold nanoparticles

4.5.1 General characteristics and synthetic aspects

AuNPs are a kind of colloidal or clustered particles with diameters in the range of a few to several hundreds of nanometers. AuNPs are made up of an Au core and/or a surface coating. The size and shape of AuNPs can be easily controlled through changing the synthesis condition (Figure 4). AuNPs have received particular research attention, as they exhibit special properties in labeling, delivery, heating and sensing. Multi-disciplinary research performed over the past decade has demonstrated well the potential of AuNP applied in cancer therapies [157], [158], [159].

Figure 4: The synthetic versatility of AuNPs. Various shapes and sizes of AuNPs controlled through changing the synthetic condition.
Figure 4:

The synthetic versatility of AuNPs. Various shapes and sizes of AuNPs controlled through changing the synthetic condition.

A number of chemical and physical methods were used in the synthesis of AuNPs. Usually, AuNPs are prepared by reducing metal salt precursors in either organic or water solvents [160]. In addition, another widely used method is seed-mediated growth of AuNPs, which was grown from small seed particles of gold [161], [162]. Because of the high affinity between the thiol group and gold, thiol-modified ligands were the most widely used to bind to the surface of AuNPs by the Au-thiol bond.

4.5.2 Gold nanoparticles for cancer radiotherapies

AuNPs have emerged as novel radiosensitizers owing to their high X-ray absorption, synthetic versatility and unique chemical, electronic and optical properties. In the past several years, multi-disciplinary research has demonstrated the potential of AuNP-based radiosensitizers and identified possible mechanisms underlying the observed radiation enhancement effects of AuNPs.

Recently, a combination therapy was reported by introducing core/shell NPs containing AuNPs and DOX [163]. This nanomedicine could realize both radiotherapies and chemotherapies by EPR effect-mediated passive targeting to cancer cells. AuNPs acted as a radiosensitizer, and they were incorporated into vesicles, which were prepared by liposome containing DOX. As shown in the results, the combination therapies achieved the highest antitumor efficacy when compared with the divided therapies. In another work, [164] glucose-capped AuNPs (Glu-AuNPs) were introduced to improve tumor cell targeting and radio-sensitization. Compared with X-rays used alone, this nanoplatform displayed a 1.5- to 2.0-fold enhancement in growth inhibition under 2 Gy of ortho-voltage irradiation.

CT contrast and radiosensitization usually increase with particle sizes of AuNPs, but there is a huge challenge to improve both by adjusting sizes under the requirements of in vivo application. Therefore, Dou et al. [165] found that AuNPs within 3–50 nm showed great size-dependent enhancements on CT imaging and radiotherapy (RT) [165]. They also demonstrated that about 13-nm-sized AuNPs could simultaneously possess excellent CT contrast ability and significant radioactive disruption. In vivo studies further indicated that this optimally sized AuNP improves real-time CT imaging and radiotherapeutic inhibition of tumors in living mice by effective accumulation at tumors with prolonged in vivo circulation times compared to clinically used small-molecule agents.

4.5.3 Gold nanoparticles for imaging

As detection methods are simple and convenient, AuNPs have been extensively explored for imaging of tumor cells. Compared with most organic dyes, AuNPs exhibit much higher absorption and scattering intensity. Therefore, AuNPs appear to be superb contrast agents for imaging [166].

Recently, Zhou et al. [167] prepared a cisplatin prodrug-conjugated gold nanocluster for fluorescence imaging and targeted therapies of the breast cancer. The fabrication was composed of fluorescence gold nanoclusters (GNC) conjugated with a cisplatin prodrug and folic acid (FA-GNC-Pt). Fluorescence imaging in vivo using the 4T1 tumor-bearing nude mouse model showed that FA-GNC-Pt NPs selectively accumulated in the orthotopic 4T1 tumor and generated a strong fluorescence signal due to the tumor targeting effect of FA. Moreover, they demonstrated that FA-GNC-Pt NPs could significantly inhibit the growth and lung metastasis of the orthotopically implanted 4T1 breast tumors. In another study, an aptamer-AuNP bioconjugate was prepared by Kim et al. [168] for both CT imaging and therapies of prostate cancer. This bioconjugate could realize targeted molecular CT imaging through aptamer-receptor specific binding, achieving an over fourfold greater CT intensity compared with a non-targeted cell.

4.5.4 Gold nanoparticles for drug and gene delivery

Because of special properties, namely, biocompatibility, size diversity and efficient translocation, AuNPs have also been extensively used as drug or gene carriers [169]. In 2014, Kumar et al. [170] synthesized glutathione-stabilized AuNPs for the delivery of platinum (IV) to prostate cancer cells by construction. The results revealed that the nanocarriers, themselves, showed no toxicity; however, once functionalized with a targeting peptide and loaded with a drug, they showed very high cytotoxicity against tumor cells.

Targeting of G-protein-coupled receptors (GPCRs) like somatostatin-14 (SST-14) could have a potential interest in delivering anticancer agents to tumor cells. Therefore, Abdellatif et al. [171] developed AuNPs coated with somatostatin as a promising delivery system for targeting somatostatin receptors. A cellular uptake study on HCC-1806 cell lines showed that the number of AuNPs-SST per cell was significantly higher compared to citrate-AuNPs when quantified using inductively coupled plasma spectroscopy. Moreover, the binding of AuNPs-SST to cells could be suppressed by the addition of an antagonist, indicating that the binding of AuNPs-SST to cells is due to receptor-specific binding. In order to achieve synergistic efficacy, Ren et al. [172] prepared NIR-radiation-responsive hollow AuNPs to sequentially co-deliver an microRNA inhibitor (miR-21i) and DOX. Compared to free DOX, this kind of nanoplatform showed a fourfold higher tumor accumulation, resulting in significantly improved antitumor efficacy.

4.6 Magnetic nanoparticles

4.6.1 General characteristics and synthetic aspects

Because of the unique physical properties such as magnetic targeting and MRI, magnetic NPs (MNPs) are an important class of nanomaterials, which could be utilized in the field of cancer therapies, including imaging of cancer cells, drug delivery and hyperthermia treatment [173], [174], [175], [176] (Figure 5).

Figure 5: Example of magnetic nanoparticle for cancer therapy. Nanocarrier, diagnosis and hyperthermia therapy.
Figure 5:

Example of magnetic nanoparticle for cancer therapy. Nanocarrier, diagnosis and hyperthermia therapy.

MNPs could be synthesized into various different compositions and phases, including iron oxides, pure metals, spinel-type ferromagnets and alloys. In recent years, a number of researches have been reported to synthesize MNPs, including several common methods, like thermal decomposition, co-precipitation, micelle synthesis and hydrothermal synthesis. MNPs prepared by these methods are of high quality [177], [178].

Co-precipitation provides a simple way to prepare iron oxides. During the hydrothermal and high-temperature reactions, iron salts in aqueous solutions and microemulsions were used. In high-temperature organic solvents, the thermal decomposition of iron pentacarbonyl and iron acetylacetonate is carried out. Surfactants can be used to synthesize monodisperse MNPs with a smaller size. For the microemulsion method, MNPs are prepared within micelles, in which the iron salt precursors are encapsulated and could control the nucleation and growth of the MNPs. For hydrothermal synthesis, a broad range of MNP nanostructure can be synthesized using these methods.

4.6.2 Magnetic nanoparticles for imaging and diagnostic applications

The most widely used application of MNPs is as MRI agents that can detect even the earliest stage disease, monitor tumor response to drug therapies and track cell migration [179], [180], [181], [182]. In a recent study, Bai et al. [183] developed a novel approach using a polymeric nanocapsule to engineer a triple-modal fluorescence/MR/SPECT imaging and magnetic targeting system, encapsulating hydrophobic superparamagnetic iron oxide NPs. When a static magnetic field was applied to the tumor for 1 h, up to ~3- and ~2.2-fold increase in tumor uptake at 1 and 24 h was achieved. Furthermore, the results indicated that the proposed system offered sufficient sensitivity to monitor its tumor uptake and magnetic targeting in vivo.

During the last decades, the development of early diagnostic methods for various tumors enabled an improvement in the treatment of cancer patients. This achievement, however, has not been achieved for pancreatic ductal adenocarcinoma (PDAC). Rosenberger et al. [184] developed targeted diagnostic magnetic NPs for the medical imaging of pancreatic cancer, which was prepared using recombinant human serum albumin (rHSA) and incorporated iron oxide (maghemite, γ-Fe2O3) NPs. Galectin-1 has been chosen as a target receptor as this protein is upregulated in pancreatic cancer and its precursor lesions but not in healthy pancreatic tissue nor in pancreatitis. Improved targeting and imaging properties were shown in mice using single-photon emission computed tomography-computer tomography (SPECT-CT), a handheld γ camera and MRI.

4.6.3 Magnetic nanoparticles for drug delivery

MNPs have been proven to be a very promising drug carrier, as they can be synthetized into various sizes and can be modified with different functional groups in order to load a number of molecules [185]. Usually, MNPs were coated with surfactants or polymers, such as PEG or dextran, to stabilize them and increase their biocompatibility. More importantly, MNPs can realize active targeting of cancer cells under an external magnetic field.

In a recent work, Farjadian et al. [186] prepared hydroxyl-modified magnetite NPs to deliver anticancer drug methotrexate (MTX). The modification resulted in the formation of highly hydrophilic NPs, which dispersed easily in water. The in vitro cell assays in the presence of free MTX and conjugated form showed an excellent anticancer effect of NPs with regard to soluble drugs. In a similar work, Hałupka-Bryl et al. [187] designed a novel MNP coated with PEG-b-poly(4-vinylbenzylphosphonate) to load DOX and realized targeting delivery. This formation displayed excellent relaxation properties under an external magnetic field. Thus, these kinds of MNPs seems to be good carriers to deliver magnetic anticancer drugs.

4.6.4 Magnetic nanoparticles for hyperthermia therapies

Under alternating magnetic fields, MNPs could be heated, and the heat produced can be directly employed for hyperthermia or indirectly used to induce drug release for killing cancer cells. Therefore, MNPs have been extensively explored as mediators of heat for hyperthermia therapies [188], [189], [190], [191], [192]. More importantly, magnetic hyperthermia appears to be a harmless approach to kill cancer cells, as it is based on the heat that is generated by the magnetic materials, themselves.

Hervault et al. [193] reported new magnetic nanocomposites (MNCs), which were composed of an iron oxide core and a thermo- and pH-responsive polymer shell. These novel MNCs could act as both a drug carrier and a hyperthermic agent. At a relatively low NP concentration, the MNCs were proven to preform as nano-heaters for hyperthermia therapies. In a previous work, Sadhukha et al. [194] prepared EGFR-targeted, inhalable SPIO NPs to realize targeting hyperthermia therapies in lung cancer. Because of the effect of EGFR, the tumor retention of SPIO NPs was greatly enhanced. Therefore, using these kinds of targeted SPIO NPs could contribute to significant inhibition of in vivo lung cancer growth by magnetic hyperthermia treatment.

Recently, Espinosa et al. [195] reported that iron oxide NPs had the dual capacity to act as both magnetic and photothermal agents. In cancer cells, the laser excitation restores the optimal efficiency of magnetic hyperthermia, resulting in a remarkable heating efficiency in the DUAL mode (up to 15-fold amplification). In solid tumors in vivo, single-mode treatments (magnetic or laser hyperthermia) reduced tumor growth, while dual treatment resulted in complete tumor regression, caused by heat-induced tumoral cell apoptosis and massive denaturation of the collagen fibers, and a long-lasting thermal efficiency over repeated treatments.

4.7 Quantum dots

4.7.1 General characteristics and synthetic aspects

Within the particle size of 2–100 nm, quantum dots (QDs) possess excellent tunable optical and passive-targeting (EPR effect) properties [196], [197], [198]. Size, composition and shape of QDs can be strictly regulated by the reaction temperature and choice of surfactants and precursors to give a number of QDs with a distinctive set of optical properties. Compared with traditional organic dyes, most of the QDs are semiconductor nanocrystals that could show better optical properties [199], [200]. They are featured with broad absorption and narrow emission spectra, and their emission maxima can be tuned between 450 and 850 nm through modulating their particle size. Recent developments in QD technology have already made a remarkable impact on cancer therapies.

The synthesis of QDs was first precipitated from a solution that contains metal ions (Cd, Ag, Pb, Hg, Zn, In) by adding into a hydroxide of S, Se, or Te [201]. In the preparation process of QDs, the size, shape, homogeneity and the surface structure need to be controlled, which are imperative for their unique properties mentioned above. Traditionally, trioctylphosphine oxide (TOPO) and trioctylphosphine (TOP) are used as coordinating solvents to stabilize the particles and protect from agglomeration.

4.7.2 Quantum dots for imaging

In the past decade, QDs have been explored as an excellent probe in cancer research due to the optical advantages [202], [203], [204], [205], [206]. Compared to conventional dyes, they are brighter, which means a very small number of QDs are sufficient to produce a signal, which is more photostable, affording for the acquisition of images after long periods of time and have a broader excitation spectrum. Furthermore, QDs could also be made to emit in the near-infrared window of the spectrum, where autofluorescence is considerably reduced and where no good dyes exist.

Rakovich et al. [207] prepared antibody-QD conjugate for detecting HER2 biomarker in breast and lung cancer cells. QD-based nanoprobes were applied for immunolabeling of breast and lung cancer cell lines via conjugating with single-domain anti-HER2 antibodies. The antibody-QD conjugates exhibited a superior staining efficiency in a panel of lung cancer cell lines with differential HER2 expression. In another study, Ye et al. [208] synthesized ZnO QDs for imaging and treating mouse tumors while showing no obvious toxicity on the body itself. ZnO QDs were synthesized and covered with polymer shells for the coordination with Gd3+ ions and adsorption of DOX to form ZnO-Gd-DOX QDs. This kind of QDs exhibited strong red fluorescence and possessed a high longitudinal relaxivity, superior to many other Gd3+-based nanoplatforms. Thus, except for effective drug delivery, fluorescence labeling and magnetic resonance imaging could be acquired simultaneously in the tumor-bearing mice as well. Kong et al. [209] designed a kind of PbS QDs, which encapsulated ribonuclease-A (RNase-A) for ultrasensitive fluorescence in vivo imaging in the second NIR region of the spectrum. Confirmed by the quantum yield (Φf) measurement, this PbS-based QDs were proven to be one of the brightest emitters in the second NIR region. The high Φf (~17.3%) and peak emission at ~1300 nm contributed to deep optical penetration to muscle tissues and superb imaging contrast at an ultra-low dose.

4.7.3 Quantum dots for photodynamic therapies

In previous research, QDs have proven to be successfully applied in PDT, and they could act as either photo-sensitizers, themselves, or an activator for another photo-sensitizer by providing energy [210]. They can match any PDT photosensitizer via changing their size and composition and used as an energy supplier.

Morosini et al. [211] reported the preparation of CdTe(S)-based hydrophilic QDs used as photosensitizers in the PDT of cancer. This kind of QDs were conjugated with folic acid for active targeting, and they can emit in the NIR region. The results showed significant difference in photodynamic efficiency between KB (overexpression of FR-α) and HT-29 cells (lacking FR-α) when the folic acid-conjugated QDs were evaluated. Besides, an enhanced photocytotoxicity effect was achieved under optimal conditions, demonstrating that this kind of QDs could be applied in targeted PDT. In 2015, Martynenko et al. [212] reported an effective reagent for photodynamic therapies, which was based on chlorin e6-ZnSe/ZnS QDs. Compared with chlorin e6 molecules, as shown by the PDT test, the effect of the complexes on the Erlich acsite carcinoma cell culture showed a two-fold enhancement of the cancer cell photodynamic destruction.

4.7.4 Quantum dots for drug and gene delivery

Because of the appropriate particle size, QDs could also be designed for simultaneous drug delivery, particularly for antitumor drug or gene delivery [213], [214], [215], [216]. For example, Li et al. [217] designed multifunctional QD-based complexes for the co-delivery of DOX and siRNA into multidrug resistance tumor cells and realized real-time tracking. The complexes were composed of two kinds of QDs, CdSe and ZnSe, which were modified with β-CD in order to couple L-Arg or L-His. The complexes then were used to simultaneously deliver DOX and siRNA, which could target the MDR1 gene, thus, realizing the reverse of the multidrug resistance of HeLa cells. Compared with free DOX, after treatment with the complexes, the number of apoptotic HeLa cells was greatly increased. Besides, this kind of system could be tracked by laser confocal microscopy due to the QDs’ intrinsic fluorescence. In another study, Zhu et al. [215] synthesized a system based on QDs for the delivery of HIF-1α siRNA into tumor cells. In this novel system, CdTe QDs were used as the core and covered by a functional shell, which was composed of 2-deoxyglucose (DG)-PEG. Besides, the compound of lipoic acid, lysine and 9-poly-d-arginin were connected with 2-deoxyglucose (DG)-PEG through a hydrazone bond. Results obtained in both in vitro and in vivo experiments demonstrated that this kind of delivery system could highly improve the delivery efficacy of the siRNA into tumor cells.

5 Conclusions

Nanomedicine, as a promising tool for cancer therapy, has received tremendous attention and yielded a great number of medical benefits. Recently, various kinds of nanomaterials, which could exhibit better cellular uptake, tumor site specificity, and prolonged circulation time after surface modification, have been investigated for imaging, diagnosis and treatment of cancers. Additionally, on account of their high surface area/volume ratio, nanomaterials can load numerous therapeutic drugs, which are delivered to tumor tissues via the EPR effect after penetrating into the leaky tumor vasculatures. Owning to these advantages, nanomaterial-based therapeutics have exhibited comparable or even superior anticancer efficacy to commercial formations, while displaying reduced side effects, providing new strategies to fight against carcinomas.

In this review, a variety of nanomaterials were presented including polymeric NPs, liposomes, dendrimers, polymeric micelles, polymer-drug conjugates, silica NPs, carbon nanotubes, nanographene, magnetic NPs, AuNPs and quantum dots. The investigation of these nanomaterials through improving their structure and cellular targeting ability has provided a more efficacious therapeutic delivery. Most of these works have shown great potential to create life-altering nanomedicine therapies for clinical cancer therapy. There are already some novel nanomaterial-based formations approved for cancer therapies available commercially, such as Abraxane, Doxil, and Embosphere, etc.

Despite excellent advantages, the successful translation of nanomaterial-based therapeutics into clinical practice still has many concerns and challenges including biocompatibility, pharmacokinetics and in vivo targeting efficacy. The most important concern would be nanomaterial toxicity and potential health risks, as little is known about how they behave in humans. Of all the nanomaterials discussed in this review, liposomes are the most developed, are clinically approved and currently possess the greatest number of clinical trials with some formulations already available in the marketplace, while many other nanomaterial-based formulations, especially inorganic ones, have not received such approval and success. This is probably due to the biocompatibility of the other nanomaterials that have not been investigated for the same duration in comparison with liposomes. A variety of nanomaterials are not biodegradable and may be retained inside the body for long periods of time after administration. The development of biocompatible and biodegradable nanomaterials for cancer therapy could, thus, have a much higher clinical value.

Although nanomaterials still have a number of technical limitations on clinical translation, they provide us with a direction for cancer therapy. There is good reason to believe that, in the near future, clinically featured nanomaterials will be developed, which will be followed with well-designed clinical trials to demonstrate their practical usage in clinical settings.

About the authors

Qing Zhou

Qing Zhou received his BS degree in Biopharmaceuticals from Soochow University in 2012. Since then, he has been pursuing his PhD degree under the supervision of Prof. Hong Wu in the Department of Pharmaceutical Analysis at the School of Pharmacy at the Fourth Military Medical University. His research interest is focused on the pH-sensitive drug delivery systems for cancer therapy.

Li Zhang

Li Zhang received her BS degree in Stomatology from Soochow University in 2012 and her Master’s Degree in Stomatology from the Fourth Military Medical University in 2015. Since then, she is studying as a PhD candidate under the guidance of Prof. Yimin Zhao in the Department of Prosthetic Dentistry at the Fourth Military Medical University. Her current research directions include biofunctionalization of the titanium implant and nanotube for osteogenesis.

Hong Wu

Hong Wu received her BS degree from Northwest University (China) in 1992 and her PhD degree from Xi’an Jiaotong University (China) in 2005. Dr. Wu is the Director of the Department of Pharmaceutical Analysis at the Fourth Military Medical University since 2013. She is now a professor working in the field of nanobiotechnology and nanomedicine to develop various functional nano drug delivery systems for cancer therapy, particular on polymer-based nano-drug carriers.

  1. Funding: National Natural Science Foundation of China, (Grant/Award Number: ‘No. 81271687’, ‘No. 81571786’).

References

[1] Siegel RL, Miller KD, Jemal A. Cancer statistics, 2016. CA Cancer J. Clin. 2016, 66, 7–30.10.3322/caac.21332Search in Google Scholar PubMed

[2] Tiwari M. Nano cancer therapies strategies. J. Cancer Res. Ther. 2012, 8, 19–22.10.4103/0973-1482.95168Search in Google Scholar PubMed

[3] Misra R, Acharya S, Sahoo SK. Cancer nanotechnology: application of nanotechnology in cancer therapies. Drug Discov. Today 2010, 15, 842–850.10.1016/j.drudis.2010.08.006Search in Google Scholar PubMed

[4] Ferrari M. Cancer nanotechnology: opportunities and challenges. Nat. Rev. Cancer 2005, 5, 161–171.10.1038/nrc1566Search in Google Scholar PubMed

[5] Yang H. Targeted nanosystems: advances in targeted dendrimers for cancer therapies. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 309–316.10.1016/j.nano.2015.11.012Search in Google Scholar PubMed PubMed Central

[6] Ma DD, Yang WX. Engineered nanoparticles induce cell apoptosis: potential for cancer therapies. Oncotarget 2016, 7, 40882–40903.10.18632/oncotarget.8553Search in Google Scholar PubMed PubMed Central

[7] Bawarski WE, Chidlowsky E, Bharali DJ, Mousa SA. Emerging nanopharmaceuticals. Nanomed. Nanotechnol. Biol. Med. 2008, 4, 273–282.10.1016/j.nano.2008.06.002Search in Google Scholar PubMed

[8] Davis ME, Shin DM. Nanoparticle therapeutics: an emerging treatment modality for cancer. Nat. Rev. Drug Discov. 2008, 7, 771–782.10.1142/9789814287005_0025Search in Google Scholar

[9] Rao J. Shedding light on tumors using nanoparticles. ACS Nano 2008, 2, 1984–1986.10.1021/nn800669nSearch in Google Scholar PubMed

[10] Torchilin VP. Targeted pharmaceutical nanocarriers for cancer therapy and imaging. AAPS J. 2007, 9, E128–E147.10.1208/aapsj0902015Search in Google Scholar PubMed PubMed Central

[11] Zalipsky S. Chemistry of polyethylene glycol conjugates with biologically active molecules. Adv. Drug Deliv. Rev. 1995, 16, 157–182.10.1016/0169-409X(95)00023-ZSearch in Google Scholar

[12] 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. 2010, 49, 6288–6308.10.1002/anie.200902672Search in Google Scholar

[13] Maeda H, Wu J, Sawa T, Matsumura Y, Hori K. Tumor vascular permeability and the EPR effect in macromolecular therapeutics: a review. J. Control. Release 2000, 65, 271–284.10.1016/S0168-3659(99)00248-5Search in Google Scholar

[14] Hall JB, Dobrovolskaia MA, Patri AK, McNeil SE. Characterization of nanoparticles for therapeutics. Nanomedicine 2007, 2, 789–803.10.2217/17435889.2.6.789Search in Google Scholar PubMed

[15] Wagner E. Programmed drug delivery: nanosystems for tumor targeting. Exp. Opin. Biol. Ther. 2007, 7, 587–593.10.1517/14712598.7.5.587Search in Google Scholar PubMed

[16] Phillips MA, Gran ML, Peppas NA. Targeted nanodelivery of drugs and diagnostics. Nano Today 2010, 5, 143–159.10.1016/j.nantod.2010.03.003Search in Google Scholar PubMed PubMed Central

[17] Veiseh O, Sun C, Fang C, Bhattarai N, Gunn J, Kievit F, Du K, Pullar B, Lee D, Ellenbogen RG, Olson J, Zhang M. Specific targeting of brain tumors with an optical/magnetic resonance imaging nanoprobe across the blood-brain barrier. Cancer Res. 2009, 69, 6200–6207.10.1158/0008-5472.CAN-09-1157Search in Google Scholar PubMed PubMed Central

[18] Juillerat-Jeanneret L. The targeted delivery of cancer drugs across the blood-brain barrier: chemical modifications of drugs or drug-nanoparticles? Drug Discov. Today 2008, 13, 1099–1106.10.1016/j.drudis.2008.09.005Search in Google Scholar PubMed

[19] Jabr-Milane LS, van Vlerken LE, Yadav S, Amiji MM. Multi-functional nanocarriers to overcome tumor drug resistance. Cancer Treat. Rev. 2008, 34, 592–602.10.1016/j.ctrv.2008.04.003Search in Google Scholar PubMed PubMed Central

[20] Saw PE, Park J, Jon S, Farokhzad OC. A drug-delivery strategy for overcoming drug resistance in breast cancer through targeting of oncofetal fibronectin. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 713–722.10.1016/j.nano.2016.10.005Search in Google Scholar PubMed

[21] Torchilin VP. Micellar nanocarriers: pharmaceutical perspectives. Pharm. Res. 2006, 24, 1–16.10.1007/s11095-006-9132-0Search in Google Scholar

[22] Zhang J, Ma PX. Polymeric core-shell assemblies mediated by host–guest interactions: versatile nanocarriers for drug delivery. Angew. Chem. Int. Ed. 2009, 48, 964–968.10.1002/anie.200804135Search in Google Scholar

[23] Nguyen DN, Green JJ, Chan JM, Langer R, Anderson DG. Polymeric materials for gene delivery and DNA vaccination. Adv. Mater. 2009, 21, 847–867.10.1002/adma.200801478Search in Google Scholar

[24] Wu H, Zhu L, Torchilin VP. pH-sensitive poly(histidine)-PEG/DSPE-PEG co-polymer micelles for cytosolic drug delivery. Biomaterials 2013, 34, 1213–1222.10.1016/j.biomaterials.2012.08.072Search in Google Scholar

[25] Kataoka K, Harada A, Nagasaki Y. Block copolymer micelles for drug delivery: design, characterization and biological significance. Adv. Drug Deliv. Rev. 2001, 47, 113–131.10.1016/S0169-409X(00)00124-1Search in Google Scholar

[26] Zhou Q, Hou Y, Zhang L, Wang J, Qiao Y, Guo S, Fan L, Yang T, Zhu L, Wu H. Dual-pH sensitive charge-reversal nanocomplex for tumor-targeted drug delivery with enhanced anticancer activity. Theranostics 2017, 7, 1806–1819.10.7150/thno.18607Search in Google Scholar

[27] Fonseca AC, Serra AC, Coelho JF. Bioabsorbable polymers in cancer therapies: latest developments. EPMA J. 2015, 6, 22.10.1186/s13167-015-0045-zSearch in Google Scholar

[28] Biswas S, Kumari P, Lakhani PM, Ghosh B. Recent advances in polymeric micelles for anti-cancer drug delivery. Eur. J. Pharm. Sci. 2016, 83, 184–202.10.1016/j.ejps.2015.12.031Search in Google Scholar

[29] Jones MC, Leroux JC. Polymeric micelles – a new generation of colloidal drug carriers. Eur. J. Pharm. Biopharm. 1999, 48, 101–111.10.1016/S0939-6411(99)00039-9Search in Google Scholar

[30] Kwon GS, Okano T. Polymeric micelles as new drug carriers. Adv. Drug Deliv. Rev. 1996, 21, 107–116.10.1016/S0169-409X(96)00401-2Search in Google Scholar

[31] Van Butsele K, Sibret P, Fustin CA, Gohy JF, Passirani C, Benoit J-P, Jerome R, Jerome C. Synthesis and pH-dependent micellization of diblock copolymer mixtures. J. Colloid Interface Sci. 2009, 329, 235–243.10.1016/j.jcis.2008.09.080Search in Google Scholar PubMed

[32] Yang X, Li ZJ, Wang N, Li L, Song L, He T, Sun L, Wang Z, Wu Q, Luo N, Yi C, Gong C. Curcumin-encapsulated polymeric micelles suppress the development of colon cancer in vitro and in vivo. Scientific Reports 2015, 5, 10322.10.1038/srep10322Search in Google Scholar PubMed PubMed Central

[33] Liang Y, Deng X, Zhang L, Peng X, Gao W, Cao J, Gu Z, He B. Terminal modification of polymeric micelles with π-conjugated moieties for efficient anticancer drug delivery. Biomaterials 2015, 71, 1–10.10.1016/j.biomaterials.2015.08.032Search in Google Scholar PubMed

[34] Makino J, Cabral H, Miura Y, Matsumoto Y, Wang M, Kinoh H, Mochida Y, Nishiyama N, Kataoka K. cRGD-installed polymeric micelles loading platinum anticancer drugs enable cooperative treatment against lymph node metastasis. J. Control. Release 2015, 220, Part B, 783–791.10.1016/j.jconrel.2015.10.017Search in Google Scholar PubMed

[35] Helmlinger G, Sckell A, Dellian M, Forbes NS, Jain RK. Acid production in glycolysis-impaired tumors provides new insights into tumor metabolism. Clin. Cancer Res. 2002, 8, 1284–1291.Search in Google Scholar

[36] Vander Heiden MG, Cantley LC, Thompson CB. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033.10.1126/science.1160809Search in Google Scholar PubMed PubMed Central

[37] Wang T, Wang D, Yu H, Wang M, Liu J, Feng B, Zhou F, Yin Q, Zhang Z, Huang Y, Li Y. Intracellularly acid-switchable multifunctional micelles for combinational photo/chemotherapies of the drug-resistant tumor. ACS Nano 2016, 10, 3496–3508.10.1021/acsnano.5b07706Search in Google Scholar PubMed

[38] Yang TH, Li F, Zhang HT, Fan L, Qiao Y, Tan G, Zhang H, Wu H. Multifunctional pH-sensitive micelles for tumor-specific uptake and cellular delivery. Polym. Chem. UK 2015, 6, 1373–1382.10.1039/C4PY01403KSearch in Google Scholar

[39] Masood F. Polymeric nanoparticles for targeted drug delivery system for cancer therapies. Mater. Sci. Eng. C 2016, 60, 569–578.10.1016/j.msec.2015.11.067Search in Google Scholar PubMed

[40] Devulapally R, Paulmurugan R. Polymer nanoparticles for drug and small silencing RNA delivery to treat cancers of different phenotypes. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2014, 6, 40–60.10.1002/wnan.1242Search in Google Scholar PubMed PubMed Central

[41] Tosi G, Bortot B, Ruozi B, Dolcetta D, Vandelli MA, Forni F, Severini GM. Potential use of polymeric nanoparticles for drug delivery across the blood-brain barrier. Curr. Med. Chem. 2013, 20, 2212–2225.10.2174/0929867311320170006Search in Google Scholar PubMed

[42] Patil YB, Toti US, Khdair A, Ma L, Panyam J. Single-step surface functionalization of polymeric nanoparticles for targeted drug delivery. Biomaterials 2009, 30, 859–866.10.1016/j.biomaterials.2008.09.056Search in Google Scholar

[43] Govender T, Stolnik S, Garnett MC, Illum L, Davis SS. PLGA nanoparticles prepared by nanoprecipitation: drug loading and release studies of a water soluble drug. J. Control. Release 1999, 57, 171–185.10.1016/S0168-3659(98)00116-3Search in Google Scholar

[44] Wang G, Chen Y, Wang P, Wang Y, Hong H, Li Y, Qian J, Yuan Y, Yu B, Liu C. Preferential tumor accumulation and desirable interstitial penetration of poly(lactic-co-glycolic acid) nanoparticles with dual coating of chitosan oligosaccharide and polyethylene glycol-poly(d,l-lactic acid). Acta Biomater. 2016, 29, 248–260.10.1016/j.actbio.2015.10.017Search in Google Scholar PubMed

[45] Tam YT, To KKW, Chow AHL. Fabrication of doxorubicin nanoparticles by controlled antisolvent precipitation for enhanced intracellular delivery. Colloids Surf. B Biointerfaces 2016, 139, 249–258.10.1016/j.colsurfb.2015.12.026Search in Google Scholar PubMed

[46] Owen SC, Patel N, Logie J, Pan G, Persson H, Moffat J, Sidhu SS, Shoichet MS. Targeting HER2 + breast cancer cells: lysosomal accumulation of anti-HER2 antibodies is influenced by antibody binding site and conjugation to polymeric nanoparticles. J. Control. Release 2013, 172, 395–404.10.1016/j.jconrel.2013.07.011Search in Google Scholar PubMed

[47] Liu Y, Chen Z, Liu C, Yu D, Lu Z, Zhang N. Gadolinium-loaded polymeric nanoparticles modified with Anti-VEGF as multifunctional MRI contrast agents for the diagnosis of liver cancer. Biomaterials 2011, 32, 5167–5176.10.1016/j.biomaterials.2011.03.077Search in Google Scholar PubMed

[48] Luo CJ, Okubo T, Nangrejo M, Edirisinghe M. Preparation of polymeric nanoparticles by novel electrospray nanoprecipitation. Polym. Int. 2015, 64, 183–187.10.1002/pi.4822Search in Google Scholar

[49] Danhier F, Lecouturier N, Vroman B, Jérôme C, Marchand-Brynaert J, Feron O, Préat V. Paclitaxel-loaded PEGylated PLGA-based nanoparticles: in vitro and in vivo evaluation. J. Control. Release 2009, 133, 11–17.10.1016/j.jconrel.2008.09.086Search in Google Scholar PubMed

[50] Mangraviti A, Tzeng SY, Kozielski KL, Wang Y, Jin Y, Gullotti D, Pedone M, Buaron N, Liu A, Wilson DR, Hansen SK, Rodriguez FJ, Gao GD, DiMeco F, Brem H, Olivi A, Tyler B, Green JJ. Polymeric nanoparticles for nonviral gene therapies extend brain tumor survival in vivo. ACS Nano 2015, 9, 1236–1249.10.1021/nn504905qSearch in Google Scholar PubMed PubMed Central

[51] Thambi T, Deepagan VG, Yoon HY, Han HS, Kim SH, Son S, Jo DG, Ahn CH, Suh YD, Kim K, Kwon IC, Lee DS, Park JH. Hypoxia-responsive polymeric nanoparticles for tumor-targeted drug delivery. Biomaterials 2014, 35, 1735–1743.10.1016/j.biomaterials.2013.11.022Search in Google Scholar PubMed

[52] Li YJ, Dong M, Kong FM, Zhou JP. Folate-decorated anticancer drug and magnetic nanoparticles encapsulated polymeric carrier for liver cancer therapeutics. Int. J. Pharm. 2015, 489, 83–90.10.1016/j.ijpharm.2015.04.028Search in Google Scholar PubMed

[53] Li L, Tong R, Li M, Kohane DS. Self-assembled gemcitabine–gadolinium nanoparticles for magnetic resonance imaging and cancer therapies. Acta Biomater. 2016, 33, 34–39.10.1016/j.actbio.2016.01.039Search in Google Scholar PubMed PubMed Central

[54] Huang S-L. Liposomes in ultrasonic drug and gene delivery. Adv. Drug Deliv. Rev. 2008, 60, 1167–1176.10.1016/j.addr.2008.03.003Search in Google Scholar PubMed

[55] Cai D, Gao W, He B, Dai W, Zhang H, Wang X, Wang J, Zhang X, Zhang Q. Hydrophobic penetrating peptide PFVYLI-modified stealth liposomes for doxorubicin delivery in breast cancer therapies. Biomaterials 2014, 35, 2283–2294.10.1016/j.biomaterials.2013.11.088Search in Google Scholar PubMed

[56] Pakunlu RI, Wang Y, Saad M, Khandare JJ, Starovoytov V, Minko T. In vitro and in vivo intracellular liposomal delivery of antisense oligonucleotides and anticancer drug. J. Control. Release 2006, 114, 153–16210.1016/j.jconrel.2006.06.010Search in Google Scholar PubMed

[57] Zhang J, Li X, Huang L. Non-viral nanocarriers for siRNA delivery in breast cancer. J. Control. Release 2014, 190, 440–450.10.1016/j.jconrel.2014.05.037Search in Google Scholar PubMed PubMed Central

[58] Peer D, Karp JM, Hong S, Farokhzad OC, Margalit R, Langer R. Nanocarriers as an emerging platform for cancer therapies. Nat. Nano. 2007, 2, 751–760.10.1201/9780429399039-2Search in Google Scholar

[59] Takara K, Hatakeyama H, Kibria G, Ohga N, Hida K, Harashima H. Size-controlled, dual-ligand modified liposomes that target the tumor vasculature show promise for use in drug-resistant cancer therapies. J. Control. Release 2012, 162, 225–232.10.1016/j.jconrel.2012.06.019Search in Google Scholar PubMed

[60] Rengan AK, Jagtap M, De A, Banerjee R, Srivastava R. Multifunctional gold coated thermo-sensitive liposomes for multimodal imaging and photo-thermal therapies of breast cancer cells. Nanoscale 2014, 6, 916–923.10.1039/C3NR04448CSearch in Google Scholar PubMed

[61] Jaafar-Maalej C, Charcosset C, Fessi H. A new method for liposome preparation using a membrane contactor. J. Liposome Res. 2011, 21, 213–220.10.3109/08982104.2010.517537Search in Google Scholar PubMed

[62] Marcato PD, Duran N. New aspects of nanopharmaceutical delivery systems. J. Nanosci. Nanotechnol. 2008, 8, 2216–2229.10.1166/jnn.2008.274Search in Google Scholar PubMed

[63] Slingerland M, Guchelaar HJ, Gelderblom H. Liposomal drug formulations in cancer therapies: 15 years along the road. Drug Discov. Today 2012, 17, 160–166.10.1016/j.drudis.2011.09.015Search in Google Scholar PubMed

[64] Chang HI, Yeh MK. Clinical development of liposome-based drugs: formulation, characterization, and therapeutic efficacy. Int. J. Nanomed. 2012, 7, 49–60.Search in Google Scholar

[65] Zhao Y, Ren W, Zhong T, Zhang S, Huang D, Guo Y, Yao X, Wang C, Zhang WQ, Zhang X, Zhang Q. Tumor-specific pH-responsive peptide-modified pH-sensitive liposomes containing doxorubicin for enhancing glioma targeting and anti-tumor activity. J. Control. Release 2016, 222, 56–66.10.1016/j.jconrel.2015.12.006Search in Google Scholar PubMed

[66] Chiang Y-T, Lo C-L. pH-Responsive polymer-liposomes for intracellular drug delivery and tumor extracellular matrix switched-on targeted cancer therapies. Biomaterials 2014, 35, 5414–5424.10.1016/j.biomaterials.2014.03.046Search in Google Scholar PubMed

[67] Ren L, Chen S, Li H, Zhang Z1, Zhong J1, Liu M1, Zhou X. MRI-guided liposomes for targeted tandem chemotherapies and therapeutic response prediction. Acta Biomater. 2016, 35, 260–268.10.1016/j.actbio.2016.02.011Search in Google Scholar PubMed

[68] German SV, Navolokin NA, Kuznetsova NR, Zuev VV, Inozemtseva OA, Anis’kov AA, Volkova EK, Bucharskaya AB, Maslyakova GN, Fakhrullin RF, Terentyuk GS, Vodovozova EL, Gorin DA. Liposomes loaded with hydrophilic magnetite nanoparticles: preparation and application as contrast agents for magnetic resonance imaging. Colloids Surf. B Biointerfaces 2015, 135, 109–115.10.1016/j.colsurfb.2015.07.042Search in Google Scholar PubMed

[69] Bharali DJ, Khalil M, Gurbuz M, Simone TM, Mousa SA. Nanoparticles and cancer therapies: a concise review with emphasis on dendrimers. Int. J. Nanomed. 2009, 4, 1–7.10.2147/IJN.S4241Search in Google Scholar

[70] Somani S, Dufes C. Applications of dendrimers for brain delivery and cancer therapies. Nanomedicine 2014, 9, 2403–2414.10.2217/nnm.14.130Search in Google Scholar PubMed

[71] Cai X, Hu J, Xiao J, Cheng Y. Dendrimer and cancer: a patent review (2006-present). Exp. Opin. Ther. Pat. 2013, 23, 515–529.10.1517/13543776.2013.761207Search in Google Scholar PubMed

[72] Newkome GR, Shreiner CD. Poly(amidoamine), polypropylenimine, and related dendrimers and dendrons possessing different 1→2 branching motifs: an overview of the divergent procedures. Polymer 2008, 49, 1–173.10.1016/j.polymer.2007.10.021Search in Google Scholar

[73] Khoee S, Hemati K. Synthesis of magnetite/polyamino-ester dendrimer based on PCL/PEG amphiphilic copolymers via convergent approach for targeted diagnosis and therapies. Polymer 2013, 54, 5574–5585.10.1016/j.polymer.2013.07.074Search in Google Scholar

[74] Grayson SM, Frechet JM. Convergent dendrons and dendrimers: from synthesis to applications. Chem. Rev. 2001, 101, 3819–3868.10.1021/cr990116hSearch in Google Scholar PubMed

[75] Liu Y, Ng Y, Toh MR, Chiu GNC. Lipid-dendrimer hybrid nanosystem as a novel delivery system for paclitaxel to treat ovarian cancer. J. Control. Release 2015, 220, Part A, 438–446.10.1016/j.jconrel.2015.11.004Search in Google Scholar PubMed

[76] Cheng L, Hu Q, Cheng L, Hu W, Xu M, Zhu Y, Zhang L, Chen D. Construction and evaluation of PAMAM-DOX conjugates with superior tumor recognition and intracellular acid-triggered drug release properties. Colloids Surf. B Biointerfaces 2015, 136, 37–45.10.1016/j.colsurfb.2015.04.003Search in Google Scholar PubMed

[77] Liu X, Liu C, Chen C, Bentobji M, Cheillan FA, Piana JT, Qu F, Rocchi P, Peng L. Targeted delivery of Dicer-substrate siRNAs using a dual targeting peptide decorated dendrimer delivery system. Nanomed. Nanotechnol. Biol. Med. 2014, 10, 1627–1636.10.1016/j.nano.2014.05.008Search in Google Scholar PubMed

[78] Wang H, Zheng L, Peng C, Guo R, Shen M, Shi X, Zhang G. Computed tomography imaging of cancer cells using acetylated dendrimer-entrapped gold nanoparticles. Biomaterials 2011, 32, 2979–2988.10.1016/j.biomaterials.2011.01.001Search in Google Scholar PubMed

[79] Li KA, Wen SH, Larson AC, Shen M, Zhang Z, Chen Q, Shi X, Zhang G. Multifunctional dendrimer-based nanoparticles for in vivo MR/CT dual-modal molecular imaging of breast cancer. Int. J. Nanomed. 2013, 8, 2589–2600.10.2147/IJN.S46177Search in Google Scholar PubMed PubMed Central

[80] Zhu J, Zhao L, Cheng Y, Xiong Z, Tang Y, Shen M, Zhao J, Shi X. Radionuclide 131I-labeled multifunctional dendrimers for targeted SPECT imaging and radiotherapies of tumors. Nanoscale 2015, 7, 18169–18178.10.1039/C5NR05585GSearch in Google Scholar

[81] Greco F, Vicent MJ. Combination therapies: opportunities and challenges for polymer-drug conjugates as anticancer nanomedicines. Adv. Drug Deliv. Rev. 2009, 61, 1203–1213.10.1016/j.addr.2009.05.006Search in Google Scholar PubMed

[82] Duncan R. Polymer conjugates as anticancer nanomedicines. Nat. Rev. Cancer 2006, 6, 688–701.10.1038/nrc1958Search in Google Scholar PubMed

[83] Duncan R, Vicent MJ, Greco F, Nicholson RI. Polymer-drug conjugates: towards a novel approach for the treatment of endocrine-related cancer. Endocr. Relat. Cancer 2005, 12 Suppl 1, S189–199.10.1677/erc.1.01045Search in Google Scholar PubMed

[84] Bonomi P. Paclitaxel poliglumex (PPX, CT-2103): macromolecular medicine for advanced non-small-cell lung cancer. Exp. Rev. Anticancer Ther. 2007, 7, 415–422.10.1586/14737140.7.4.415Search in Google Scholar PubMed

[85] Zhang L, Gu FX, Chan JM, Wang AZ, Langer RS, Farokhzad OC. Nanoparticles in medicine: therapeutic applications and developments. Clin. Pharmacol. Ther. 2008, 83, 761–769.10.1038/sj.clpt.6100400Search in Google Scholar PubMed

[86] Li C, Wallace S. Polymer-drug conjugates: recent development in clinical oncology. Adv. Drug Deliv. Rev. 2008, 60, 886–898.10.1016/j.addr.2007.11.009Search in Google Scholar PubMed PubMed Central

[87] Tu Y, Zhu L. Enhancing cancer targeting and anticancer activity by a stimulus-sensitive multifunctional polymer-drug conjugate. J. Control. Release 2015, 212, 94–102.10.1016/j.jconrel.2015.06.024Search in Google Scholar PubMed

[88] Zhou Q, Yang T, Qiao Y, Guo S, Zhu L, Wu H. Preparation of poly(beta-L-malic acid)-based charge-conversional nanoconjugates for tumor-specific uptake and cellular delivery. Int. J. Nanomed. 2015, 10, 1941–1952.10.2147/IJN.S78547Search in Google Scholar PubMed PubMed Central

[89] Broxterman HJ, Georgopapadakou NH. Anticancer therapeutics: “addictive” targets, multi-targeted drugs, new drug combinations. Drug Resist. Updat. 2005, 8, 183–197.10.1016/j.drup.2005.07.002Search in Google Scholar PubMed

[90] Kostková H, Etrych T, Říhová B, Kostka L, Starovoytová L, Kovář M, Ulbrich K. HPMA copolymer conjugates of DOX and mitomycin C for combination therapies: physicochemical characterization, cytotoxic effects, combination index analysis, and anti-tumor efficacy. Macromol. Biosci. 2013, 13, 1648–1660.10.1002/mabi.201300288Search in Google Scholar PubMed

[91] Wang E, Xiong H, Zhou D, Xie Z, Huang Y, Jing X, Sun X. Co-delivery of oxaliplatin and demethylcantharidin via a polymer–drug conjugate. Macromol. Biosci. 2014, 14, 588–596.10.1002/mabi.201300402Search in Google Scholar PubMed

[92] Vivero-Escoto JL, Huxford-Phillips RC, Lin W. Silica-based nanoprobes for biomedical imaging and theranostic applications. Chem. Soc. Rev. 2012, 41, 2673–2685.10.1039/c2cs15229kSearch in Google Scholar PubMed PubMed Central

[93] Liu X, Situ A, Kang Y, Villabroza KR, Liao Y, Chang CH, Donahue T, Nel AE, Meng H. Irinotecan delivery by lipid-coated mesoporous silica nanoparticles shows improved efficacy and safety over liposomes for pancreatic cancer. ACS Nano 2016, 10, 2702–2715.10.1021/acsnano.5b07781Search in Google Scholar PubMed PubMed Central

[94] Couleaud P, Morosini V, Frochot C, Richeter S, Raehm L, Durand J-O. Silica-based nanoparticles for photodynamic therapies applications. Nanoscale 2010, 2, 1083–1095.10.1039/c0nr00096eSearch in Google Scholar PubMed

[95] Prasad R, Aiyer S, Chauhan DS, Srivastava R, Selvaraj K. Bioresponsive carbon nano-gated multifunctional mesoporous silica for cancer theranostics. Nanoscale 2016, 8, 4537–4546.10.1039/C5NR06756ASearch in Google Scholar

[96] Yang S, Chen D, Li N, Xu Q, Li H, Gu F, Xie J, Lu J. Hollow mesoporous silica nanocarriers with multifunctional capping agents for in vivo cancer imaging and therapies. Small 2016, 12, 360–370.10.1002/smll.201503121Search in Google Scholar PubMed

[97] Paris JL, Cabañas MV, Manzano M, Vallet-Regí M. Polymer-grafted mesoporous silica nanoparticles as ultrasound-responsive drug carriers. ACS Nano 2015, 9, 11023–11033.10.1021/acsnano.5b04378Search in Google Scholar PubMed

[98] Hao X, Hu X, Zhang C, Chen S, Li Z, Yang X, Liu H, Jia G, Liu D, Ge K, Liang XJ, Zhang J. Hybrid mesoporous silica-based drug carrier nanostructures with improved degradability by hydroxyapatite. ACS Nano 2015, 9, 9614–9625.10.1021/nn507485jSearch in Google Scholar PubMed

[99] Wang L, Zhao W, Tan W. Bioconjugated silica nanoparticles: development and applications. Nano Res. 2008, 1, 99–115.10.1007/s12274-008-8018-3Search in Google Scholar

[100] Li X, Zhao W, Liu X, Chen K, Zhu S, Shi P, Chen Y, Shi J. Mesoporous manganese silicate coated silica nanoparticles as multi-stimuli-responsive T1-MRI contrast agents and drug delivery carriers. Acta Biomater. 2016, 30, 378–387.10.1016/j.actbio.2015.11.036Search in Google Scholar PubMed

[101] Mekaru H, Lu J, Tamanoi F. Development of mesoporous silica-based nanoparticles with controlled release capability for cancer therapies. Adv. Drug Deliv. Rev. 2015, 95, 40–49.10.1016/j.addr.2015.09.009Search in Google Scholar PubMed PubMed Central

[102] Hakeem A, Zahid F, Duan R, Asif M, Zhang T, Zhang Z, Cheng Y, Lou X, Xia F. Cellulose conjugated FITC-labelled mesoporous silica nanoparticles: intracellular accumulation and stimuli responsive doxorubicin release. Nanoscale 2016, 8, 5089–5097.10.1039/C5NR08753HSearch in Google Scholar

[103] Liu J, Luo Z, Zhang J, Luo T, Zhou J, Zhao X, Cai K. Hollow mesoporous silica nanoparticles facilitated drug delivery via cascade pH stimuli in tumor microenvironment for tumor therapies. Biomaterials 2016, 83, 51–65.10.1016/j.biomaterials.2016.01.008Search in Google Scholar

[104] Zhang M, Wang T, Zhang L, Li L, Wang C. Near-infrared light and pH-responsive polypyrrole@polyacrylic acid/fluorescent mesoporous silica nanoparticles for imaging and chemo-photothermal cancer therapies. Chemistry 2015, 21, 16162–16171.10.1002/chem.201502177Search in Google Scholar

[105] Stöber W, Fink A, Bohn E. Controlled growth of monodisperse silica spheres in the micron size range. J. Colloid Interface Sci. 1968, 26, 62–69.10.1016/0021-9797(68)90272-5Search in Google Scholar

[106] Arriagada FJ, Osseo-Asare K. Phase and dispersion stability effects in the synthesis of silica nanoparticles in a non-ionic reverse microemulsion. Colloids Surf. 1992, 69, 105–115.10.1016/0166-6622(92)80221-MSearch in Google Scholar

[107] Meng H, Mai WX, Zhang H, Xue M, Xia T, Lin S, Wang X, Zhao Y, Ji Z, Zink JI, Nel AE. Codelivery of an optimal drug/siRNA combination using mesoporous silica nanoparticles to overcome drug resistance in breast cancer in vitro and in vivo. ACS Nano 2013, 7, 994–1005.10.1021/nn3044066Search in Google Scholar PubMed PubMed Central

[108] Ngamcherdtrakul W, Morry J, Gu S, Castro DJ, Goodyear SM, Sangvanich T, Reda MM, Lee R, Mihelic SA, Beckman BL, Hu Z, Gray JW, Yantasee W. Cationic polymer modified mesoporous silica nanoparticles for targeted SiRNA delivery to HER2+ breast cancer. Adv. Funct. Mater. 2015, 25, 2646–2659.10.1002/adfm.201404629Search in Google Scholar PubMed PubMed Central

[109] Fan L, Zhang Y, Wang F, Yang Q, Tan J, Renata G, Wu H, Song C, Jin B. Multifunctional all-in-one drug delivery systems for tumor targeting and sequential release of three different anti-tumor drugs. Biomaterials 2016, 76, 399–407.10.1016/j.biomaterials.2015.10.069Search in Google Scholar PubMed

[110] Chen WH, Luo GF, Lei Q, Cao FY, Fan JX, Qiu WX, Jia HZ, Hong S, Fang F, Zeng X, Zhuo RX, Zhang XZ . Rational design of multifunctional magnetic mesoporous silica nanoparticle for tumor-targeted magnetic resonance imaging and precise therapies. Biomaterials 2016, 76, 87–101.10.1016/j.biomaterials.2015.10.053Search in Google Scholar PubMed

[111] Hurley KR, Ring HL, Etheridge M, Zhang J, Gao Z, Shao Q, Klein ND, Szlag VM, Chung C, Reineke TM, Garwood M, Bischof JC, Haynes CL. Predictable heating and positive MRI contrast from a mesoporous silica-coated iron oxide nanoparticle. Mol. Pharm. 2016, 13, 2172–2183.10.1021/acs.molpharmaceut.5b00866Search in Google Scholar PubMed

[112] Chen X, Yao X, Wang C, Chen L, Chen X. Mesoporous silica nanoparticles capped with fluorescence-conjugated cyclodextrin for pH-activated controlled drug delivery and imaging. Microporous Mesoporous Mater. 2015, 217, 46–53.10.1016/j.micromeso.2015.06.012Search in Google Scholar

[113] Chakrabarti M, Kiseleva R, Vertegel A, Ray SK. Carbon Nanomaterials for drug delivery and cancer therapies. J. Nanosci. Nanotechnol. 2015, 15, 5501–5511.10.1166/jnn.2015.10614Search in Google Scholar PubMed

[114] Sao R, Vaish R, Sinha N. Multifunctional drug delivery systems using inorganic nanomaterials: a review. J. Nanosci. Nanotechnol. 2015, 15, 1960–1972.10.1166/jnn.2015.9761Search in Google Scholar PubMed

[115] Iijima S. Helical microtubules of graphitic carbon. Nature 1991, 354, 56–58.10.1038/354056a0Search in Google Scholar

[116] Liang F, Chen B. A review on biomedical applications of single-walled carbon nanotubes. Curr. Med. Chem. 2010, 17, 10–24.10.2174/092986710789957742Search in Google Scholar PubMed

[117] Park S, Vosguerichian M, Bao Z. A review of fabrication and applications of carbon nanotube film-based flexible electronics. Nanoscale 2013, 5, 1727–1752.10.1039/c3nr33560gSearch in Google Scholar PubMed

[118] Gougis M, Tabet-Aoul A, Ma D, Mohamedi M. Laser synthesis and tailor-design of nanosized gold onto carbon nanotubes for non-enzymatic electrochemical glucose sensor. Sens. Actuat. B Chem. 2014, 193, 363–369.10.1016/j.snb.2013.12.008Search in Google Scholar

[119] Brownson DAC, Banks CE. The electrochemistry of CVD graphene: progress and prospects. Phys. Chem. Chem. Phys. 2012, 14, 8264–8281.10.1039/c2cp40225dSearch in Google Scholar PubMed

[120] Mehra NK, Palakurthi S. Interactions between carbon nanotubes and bioactives: a drug delivery perspective. Drug Discov. Today 2016, 21, 585–597.10.1016/j.drudis.2015.11.011Search in Google Scholar PubMed

[121] Wong BS, Yoong SL, Jagusiak A, Panczyk T, Ho HK, Ang WH, Pastorin G. Carbon nanotubes for delivery of small molecule drugs. Adv. Drug Deliv. Rev. 2013, 65, 1964–2015.10.1016/j.addr.2013.08.005Search in Google Scholar PubMed

[122] Liu Z, Chen K, Davis C, Sherlock S, Cao Q, Chen X, Dai H. Drug delivery with carbon nanotubes for in vivo cancer treatment. Cancer Res. 2008, 68, 6652–6660.10.1158/0008-5472.CAN-08-1468Search in Google Scholar PubMed PubMed Central

[123] Al Faraj A, Shaik AS, Ratemi E, Halwani R. Combination of drug-conjugated SWCNT nanocarriers for efficient therapy of cancer stem cells in a breast cancer animal model. J. Control. Release 2016, 225, 240–251.10.1016/j.jconrel.2016.01.053Search in Google Scholar PubMed

[124] Zhang X, Meng L, Lu Q, Fei Z, Dyson PJ. Targeted delivery and controlled release of doxorubicin to cancer cells using modified single wall carbon nanotubes. Biomaterials 2009, 30, 6041–6047.10.1016/j.biomaterials.2009.07.025Search in Google Scholar PubMed

[125] Al Faraj A, Shaik AS, Al Sayed B, Halwani R, Al Jammaz I. Specific targeting and noninvasive imaging of breast cancer stem cells using single-walled carbon nanotubes as novel multimodality nanoprobes. Nanomedicine 2016, 11, 31–46.10.2217/nnm.15.182Search in Google Scholar PubMed

[126] Wang C, Bao C, Liang S, Fu H, Wang K, Deng M, Liao Q, Cui D. RGD-conjugated silica-coated gold nanorods on the surface of carbon nanotubes for targeted photoacoustic imaging of gastric cancer. Nanoscale Res. Lett. 2014, 9, 264.10.1186/1556-276X-9-264Search in Google Scholar PubMed PubMed Central

[127] Wu H, Shi H, Zhang H, Wang X, Yang Y, Yu C, Hao C, Du J, Hu H, Yang S. Prostate stem cell antigen antibody-conjugated multiwalled carbon nanotubes for targeted ultrasound imaging and drug delivery. Biomaterials 2014, 35, 5369–5380.10.1016/j.biomaterials.2014.03.038Search in Google Scholar PubMed

[128] Liu X, Tao H, Yang K, Zhang S, Lee S-T, Liu Z. Optimization of surface chemistry on single-walled carbon nanotubes for in vivo photothermal ablation of tumors. Biomaterials 2011, 32, 144–151.10.1016/j.biomaterials.2010.08.096Search in Google Scholar PubMed

[129] Wang S, Zhang Q, Yang P, Yu X, Huang LY, Shen S, Cai S. Manganese oxide-coated carbon nanotubes as dual-modality lymph mapping agents for photothermal therapies of tumor metastasis. ACS Appl. Mater. Interfaces 2016, 8, 3736–3743.10.1021/acsami.5b08087Search in Google Scholar PubMed

[130] Zhou F, Xing D, Ou Z, Wu B, Resasco DE, Chen WR. Cancer photothermal therapies in the near-infrared region by using single-walled carbon nanotubes. BIOMEDO 2009, 14, 021009.10.1117/1.3078803Search in Google Scholar PubMed PubMed Central

[131] Luís FFN, John JK, Brent DVR, Rajagopal R, Daniel ER, Roger GH. Targeting single-walled carbon nanotubes for the treatment of breast cancer using photothermal therapies. Nanotechnology 2013, 24, 375104.10.1088/0957-4484/24/37/375104Search in Google Scholar PubMed

[132] Huang N, Wang H, Zhao J, Lui H, Korbelik M, Zeng H. Single-wall carbon nanotubes assisted photothermal cancer therapies: animal study with a murine model of squamous cell carcinoma. Lasers Surg. Med. 2010, 42, 798–808.10.1002/lsm.20968Search in Google Scholar PubMed

[133] Zhang P, Chao H, Huang H, Huang J, Chen H, Wang J, Qiu K, Zhao D, Ji L . Noncovalent ruthenium(II) complexes-single-walled carbon nanotube composites for bimodal photothermal and photodynamic therapy with near-infrared irradiation. ACS Appl. Mater. Interfaces 2015, 7, 23278–23290.10.1021/acsami.5b07510Search in Google Scholar PubMed

[134] Zhou F, Wu S, Wu B, Chen WR, Xing D. Mitochondria-targeting single-walled carbon nanotubes for cancer photothermal therapies. Small 2011, 7, 2727–2735.10.1002/smll.201100669Search in Google Scholar PubMed

[135] Geim AK, Novoselov KS. The rise of graphene. Nat. Mater. 2007, 6, 183–191.10.1142/9789814287005_0002Search in Google Scholar

[136] Kopelevich Y, Esquinazi P. Graphene physics in graphite. Adv. Mater. 2007, 19, 4559–4563.10.1002/adma.200702051Search in Google Scholar

[137] Rao CNR, Sood AK, Subrahmanyam KS, Govindaraj A. Graphene: the new two-dimensional nanomaterial. Angew. Chem. Int. Ed. 2009, 48, 7752–7777.10.1002/anie.200901678Search in Google Scholar PubMed

[138] Luo C, Deng Z, Li L, Clayton F, Chen AL, Wei R, Miles R, Stephens DM, Glenn M, Wang X, Jensen PE, Chen X. Association of rituximab with graphene oxide confers direct cytotoxicity for CD20-positive lymphoma cells. Oncotarget 2016, 7, 12806–12822.10.18632/oncotarget.7230Search in Google Scholar PubMed PubMed Central

[139] Chen YW, Liu TY, Chang P-H, Hsu P-H, Liu H-L, Lin H-C, Chen S-Y. A theranostic nrGO@MSN-ION nanocarrier developed to enhance the combination effect of sonodynamic therapies and ultrasound hyperthermia for treating tumor. Nanoscale 2016, 8, 12648–12657.10.1039/C5NR07782FSearch in Google Scholar

[140] Wei Y, Zhou F, Zhang D, Chen Q, Xing D. A graphene oxide based smart drug delivery system for tumor mitochondria-targeting photodynamic therapies. Nanoscale 2016, 8, 3530–3538.10.1039/C5NR07785KSearch in Google Scholar PubMed

[141] Gao S, Zhang L, Wang G, Yang K, Chen M, Tian R, Ma Q, Zhu L. Hybrid graphene/Au activatable theranostic agent for multimodalities imaging guided enhanced photothermal therapies. Biomaterials 2016, 79, 36–45.10.1016/j.biomaterials.2015.11.041Search in Google Scholar PubMed

[142] Gurunathan S, Han JW, Park JH, Kim ES, Choi Y-J, Kwon D-N, Kim J-H. Reduced graphene oxide-silver nanoparticle nanocomposite: a potential anticancer nanotherapies. Int. J. Nanomed. 2015, 10, 6257–6276.10.2147/IJN.S92449Search in Google Scholar PubMed PubMed Central

[143] Kim SH, Lee JE, Sharker SM, Jeong JH, In I, Park SY. In vitro and in vivo tumor targeted photothermal cancer therapies using functionalized graphene nanoparticles. Biomacromolecules 2015, 16, 3519–3529.10.1021/acs.biomac.5b00944Search in Google Scholar PubMed

[144] Zhu Y, Murali S, Cai W, Li X, Suk JW, Potts JR, Ruoff RS. Graphene and graphene oxide: synthesis, properties, and applications. Adv. Mater. 2010, 22, 3906–3924.10.1002/adma.201001068Search in Google Scholar PubMed

[145] Huang X, Yin Z, Wu S, Qi X, He Q, Zhang Q, Yan Q, Boey F, Zhang H. Graphene-based materials: synthesis, characterization, properties, and applications. Small 2011, 7, 1876–1902.10.1002/smll.201002009Search in Google Scholar PubMed

[146] Novoselov KS, Geim AK, Morozov SV, Jiang D, Zhang Y, Dubonos SV, Grigorieva IV, Firsov AA. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669.10.1126/science.1102896Search in Google Scholar PubMed

[147] Park S, An J, Jung I, Piner RD, An SJ, Li X, Velamakanni A, Ruoff RS. Colloidal suspensions of highly reduced graphene oxide in a wide variety of organic solvents. Nano Lett. 2009, 9, 1593–1597.10.1021/nl803798ySearch in Google Scholar PubMed

[148] You P, Yang Y, Wang M, Huang X, Huang X. Graphene oxide-based nanocarriers for cancer imaging and drug delivery. Curr. Pharm. Des. 2015, 21, 3215–3222.10.2174/1381612821666150531170832Search in Google Scholar PubMed

[149] Bikhof Torbati M, Ebrahimian M, Yousefi M, Shaabanzadeh M. GO-PEG as a drug nanocarrier and its antiproliferative effect on human cervical cancer cell line. Artif. Cells Nanomed. Biotechnol. 2016, 1–6.10.3109/21691401.2016.1161641Search in Google Scholar PubMed

[150] Zhao X, Yang L, Li X, Jia X, Liu L, Zeng J, Guo J, Liu P. Functionalized graphene oxide nanoparticles for cancer cell-specific delivery of antitumor drug. Bioconjug. Chem. 2015, 26, 128–136.10.1021/bc5005137Search in Google Scholar PubMed

[151] Zheng XT, Ma XQ, Li CM. Highly efficient nuclear delivery of anti-cancer drugs using a bio-functionalized reduced graphene oxide. J. Colloid Interface Sci. 2016, 467, 35–42.10.1016/j.jcis.2015.12.052Search in Google Scholar PubMed

[152] Tian J, Luo Y, Huang L, Feng Y, Ju H, Yu B-Y. Pegylated folate and peptide-decorated graphene oxide nanovehicle for in vivo targeted delivery of anticancer drugs and therapeutic self-monitoring. Biosens. Bioelectron. 2016, 80, 519–524.10.1016/j.bios.2016.02.018Search in Google Scholar PubMed

[153] Miao W, Shim G, Kang CM, Lee S, Choe YS, Choi HG, Oh YK. Cholesteryl hyaluronic acid-coated, reduced graphene oxide nanosheets for anti-cancer drug delivery. Biomaterials. 2013, 34, 9638–9647.10.1016/j.biomaterials.2013.08.058Search in Google Scholar PubMed

[154] Yang K, Hu L, Ma X, Ye S, Cheng L, Shi X, Li C, Li Y, Liu Z. Multimodal imaging guided photothermal therapies using functionalized graphene nanosheets anchored with magnetic nanoparticles. Adv. Mater. 2012, 24, 1868–1872.10.1002/adma.201104964Search in Google Scholar PubMed

[155] Tran TH, Nguyen HT, Pham TT, Choi JY, Choi HG, Yong CS, Kim JO. Development of a graphene oxide nanocarrier for dual-drug chemo-phototherapies to overcome drug resistance in cancer. ACS Appl. Mater. Interfaces 2015, 7, 28647–28655.10.1021/acsami.5b10426Search in Google Scholar PubMed

[156] Jang C, Lee JH, Sahu A, Tae G. The synergistic effect of folate and RGD dual ligand of nanographene oxide on tumor targeting and photothermal therapies in vivo. Nanoscale 2015, 7, 18584–18594.10.1039/C5NR05067GSearch in Google Scholar PubMed

[157] Singh M, Harris-Birtill DCC, Markar SR, Hanna GB, Elson DS. Application of gold nanoparticles for gastrointestinal cancer theranostics: a systematic review. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 2083–2098.10.1016/j.nano.2015.05.010Search in Google Scholar PubMed

[158] Popescu RC, Grumezescu AM. Metal based frameworks for drug delivery systems. Curr. Top. Med. Chem. 2015, 15, 1532–1542.10.2174/1568026615666150414145323Search in Google Scholar PubMed

[159] Kodiha M, Wang YM, Hutter E, Maysinger D, Stochaj U. Off to the organelles - killing cancer cells with targeted gold nanoparticles. Theranostics 2015, 5, 357–370.10.7150/thno.10657Search in Google Scholar PubMed PubMed Central

[160] Majdalawieh A, Kanan MC, El-Kadri O, Kanan SM. Recent advances in gold and silver nanoparticles: synthesis and applications. J. Nanosci. Nanotechnol. 2014, 14, 4757–4780.10.1166/jnn.2014.9526Search in Google Scholar PubMed

[161] Millstone JE, Wei W, Jones MR, Yoo H, Mirkin CA. Iodide ions control seed-mediated growth of anisotropic gold nanoparticles. Nano Lett. 2008, 8, 2526–2529.10.1021/nl8016253Search in Google Scholar PubMed PubMed Central

[162] Sanchez A, Diez P, Villalonga R, Martínez-Ruiz P, Eguílaz M, Fernández I, Pingarrón JM. Seed-mediated growth of jack-shaped gold nanoparticles from cyclodextrin-coated gold nanospheres. Dalton Trans. 2013, 42, 14309–14314.10.1039/c3dt51368hSearch in Google Scholar PubMed

[163] Kim K, Oh KS, Park DY, Lee JY, Lee BS, Kim IS, Kim K, Kwon IC, Sang YK, Yuk SH. Doxorubicin/gold-loaded core/shell nanoparticles for combination therapies to treat cancer through the enhanced tumor targeting. J. Control. Release 2016, 228, 141–149.10.1016/j.jconrel.2016.03.009Search in Google Scholar PubMed

[164] Roa W, Zhang X, Guo L, Shaw A, Hu X, Xiong Y, Gulavita S, Patel S, Sun X, Chen J, Moore R, Xing JZ. Gold nanoparticle sensitize radiotherapies of prostate cancer cells by regulation of the cell cycle. Nanotechnology 2009, 20, 375101.10.1088/0957-4484/20/37/375101Search in Google Scholar PubMed

[165] Dou Y, Guo Y, Li X, Li X, Wang S, Wang L, Lv G, Zhang X, Wang H, Gong X, Chang J. Size-tuning ionization to optimize gold nanoparticles for simultaneous enhanced CT imaging and radiotherapies. ACS Nano 2016, 10, 2536–2548.10.1021/acsnano.5b07473Search in Google Scholar PubMed

[166] Link S, El-Sayed MA. Spectral properties and relaxation dynamics of surface plasmon electronic oscillations in gold and silver nanodots and nanorods. J. Phys. Chem. B 1999, 103, 8410–8426.10.1021/jp9917648Search in Google Scholar

[167] Zhou F, Feng B, Yu H, Wang D, Wang T, Liu J, Meng Q, Wang S, Zhang P, Zhang Z, Li Y. Cisplatin prodrug-conjugated gold nanocluster for fluorescence imaging and targeted therapies of the breast cancer. Theranostics 2016, 6, 679–687.10.7150/thno.14556Search in Google Scholar PubMed PubMed Central

[168] Kim D, Jeong YY, Jon S. A drug-loaded aptamer−gold nanoparticle bioconjugate for combined CT imaging and therapies of prostate cancer. ACS Nano 2010, 4, 3689–3696.10.1021/nn901877hSearch in Google Scholar PubMed

[169] Sperling RA, Rivera Gil P, Zhang F, Zanella M, Parak WJ. Biological applications of gold nanoparticles. Chem. Soc. Rev. 2008, 37, 1896–1908.10.1039/b712170aSearch in Google Scholar PubMed

[170] Kumar A, Huo S, Zhang X, Liu J, Tan A, Li S, Jin S, Xue X, Zhao Y, Ji T, Han L, Liu H, Zhang X, Zhang J, Zou G, Wang T, Tang S, Liang XJ. Neuropilin-1-targeted gold nanoparticles enhance therapeutic efficacy of platinum(IV) drug for prostate cancer treatment. ACS Nano 2014, 8, 4205–4220.10.1021/nn500152uSearch in Google Scholar PubMed

[171] Abdellatif AAH, Zayed G, El-Bakry A, Zaky A, Saleem IY, Tawfeek HM. Novel gold nanoparticles coated with somatostatin as a potential delivery system for targeting somatostatin receptors. Drug Dev. Ind. Pharm. 2016, 1–37.10.3109/03639045.2016.1173052Search in Google Scholar PubMed

[172] Ren Y, Wang R, Gao L, Li K, Zhou X, Guo H, Liu C, Han D, Tian J, Ye Q, Hu YT, Sun D, Yuan X, Zhang N. Sequential co-delivery of miR-21 inhibitor followed by burst release doxorubicin using NIR-responsive hollow gold nanoparticle to enhance anticancer efficacy. J. Control. Release 2016, 228, 74–86.10.1016/j.jconrel.2016.03.008Search in Google Scholar PubMed

[173] Sun C, Lee JSH, Zhang M. Magnetic nanoparticles in MR imaging and drug delivery. Adv. Drug Deliv. Rev. 2008, 60, 1252–1265.10.1016/j.addr.2008.03.018Search in Google Scholar PubMed PubMed Central

[174] Hajba L, Guttman A. The use of magnetic nanoparticles in cancer theranostics: toward handheld diagnostic devices. Biotechnol. Adv. 2016, 34, 354–361.10.1016/j.biotechadv.2016.02.001Search in Google Scholar PubMed

[175] Gobbo OL, Sjaastad K, Radomski MW, Volkov Y, Prina-Mello A. Magnetic nanoparticles in cancer theranostics. Theranostics 2015, 5, 1249–1263.10.7150/thno.11544Search in Google Scholar PubMed PubMed Central

[176] Singh A, Sahoo SK. Magnetic nanoparticles: a novel platform for cancer theranostics. Drug Discov. Today 2014, 19, 474–481.10.1016/j.drudis.2013.10.005Search in Google Scholar PubMed

[177] Laurent S, Forge D, Port M, Roch A, Robic C, Vander Elst L, Muller RN. Magnetic iron oxide nanoparticles: synthesis, stabilization, vectorization, physicochemical characterizations, and biological applications. Chem. Rev. 2008, 108, 2064–2110.10.1021/cr068445eSearch in Google Scholar PubMed

[178] Lu A-H, Salabas EL, Schüth F. Magnetic nanoparticles: synthesis, protection, functionalization, and application. Angew. Chem. Int. Ed. 2007, 46, 1222–1244.10.1002/anie.200602866Search in Google Scholar PubMed

[179] Lee JH, Huh YM, Jun YW, Seo JW, Jang JT, Song HT, Kim S, Cho EJ, Yoon HG, Suh JS, Cheon J. Artificially engineered magnetic nanoparticles for ultra-sensitive molecular imaging. Nat. Med. 2007, 13, 95–99.10.1038/nm1467Search in Google Scholar PubMed

[180] Xie J, Liu G, Eden HS, Ai H, Chen X. Surface-engineered magnetic nanoparticle platforms for cancer imaging and therapies. Acc. Chem. Res. 2011, 44, 883–892.10.1021/ar200044bSearch in Google Scholar PubMed PubMed Central

[181] Hayashi K, Nakamura M, Sakamoto W, Yogo T, Miki H, Ozaki S, Abe M, Matsumoto T, Ishimura K. Superparamagnetic nanoparticle clusters for cancer theranostics combining magnetic resonance imaging and hyperthermia treatment. Theranostics 2013, 3, 366–376.10.7150/thno.5860Search in Google Scholar PubMed PubMed Central

[182] Song H-T, Choi J-S, Huh Y-M, Kim S, Jun YW, Suh JS, Cheon J. Surface modulation of magnetic nanocrystals in the development of highly efficient magnetic resonance probes for intracellular labeling. J. Am. Chem. Soc. 2005, 127, 9992–9993.10.1021/ja051833ySearch in Google Scholar PubMed

[183] Bai J, Wang JT, Rubio N, Protti A, Heidari H, Elgogary R, Southern P, Al-Jamal WT, Sosabowski J, Shah AM, Bals S, Pankhurst QA, Al-Jamal KT. Triple-modal imaging of magnetically-targeted nanocapsules in solid tumours in vivo. Theranostics 2016, 6, 342–356.10.7150/thno.11918Search in Google Scholar PubMed PubMed Central

[184] Rosenberger I, Strauss A, Dobiasch S, Weis C, Szanyi S, Gil-Iceta L, Alonso E, González Esparza M, Gómez-Vallejo V, Szczupak B, Plaza-García S, Mirzaei S, Israel LL, Bianchessi S, Scanziani E, Lellouche JP, Knoll P, Werner J, Felix K, Grenacher L, Reese T, Kreuter J, Jiménez-González M. Targeted diagnostic magnetic nanoparticles for medical imaging of pancreatic cancer. J. Control. Release 2015, 214, 76–84.10.1016/j.jconrel.2015.07.017Search in Google Scholar PubMed

[185] Tietze R, Zaloga J, Unterweger H, Lyer S, Friedrich RP, Janko C, Pöttler M, Dürr S, Alexiou C. Magnetic nanoparticle-based drug delivery for cancer therapies. Biochem. Biophys. Res. Commun. 2015, 468, 463–470.10.1016/j.bbrc.2015.08.022Search in Google Scholar PubMed

[186] Farjadian F, Ghasemi S, Mohammadi-Samani S. Hydroxyl-modified magnetite nanoparticles as novel carrier for delivery of methotrexate. Int. J. Pharm. 2016, 504, 110–116.10.1016/j.ijpharm.2016.03.022Search in Google Scholar PubMed

[187] Hałupka-Bryl M, Bednarowicz M, Dobosz B, Krzyminiewski R, Zalewski T, Wereszczyńska B, Nowaczyk G, Jarek M, Nagasaki Y. Doxorubicin loaded PEG-b-poly(4-vinylbenzylphosphonate) coated magnetic iron oxide nanoparticles for targeted drug delivery. J. Magnet. Magn. Mater. 2015, 384, 320–327.10.1016/j.jmmm.2015.02.078Search in Google Scholar

[188] Zhang W, Guo Z, Huang D, Liu Z, Guo X, Zhong H. Synergistic effect of chemo-photothermal therapies using PEGylated graphene oxide. Biomaterials 2011, 32, 8555–8561.10.1016/j.biomaterials.2011.07.071Search in Google Scholar PubMed

[189] You J, Zhang R, Zhang G, Zhong M, Liu Y, Van Pelt CS, Liang D, Wei W, Sood AK, Li C. Photothermal-chemotherapies with doxorubicin-loaded hollow gold nanospheres: a platform for near-infrared light-trigged drug release. J. Control. Release 2012, 158, 319–328.10.1016/j.jconrel.2011.10.028Search in Google Scholar PubMed PubMed Central

[190] Vivero-Escoto JL, Slowing II, Wu CW, Lin VS. Photoinduced intracellular controlled release drug delivery in human cells by gold-capped mesoporous silica nanosphere. J. Am. Chem. Soc. 2009, 131, 3462–3463.10.1021/ja900025fSearch in Google Scholar PubMed

[191] Lee SM, Park H, Choi JW, Park YN, Yun CO, Yoo KH. Multifunctional nanoparticles for targeted chemophotothermal treatment of cancer cells. Angew. Chem. 2011, 50, 7581–7586.10.1002/anie.201101783Search in Google Scholar PubMed

[192] Huang HC, Barua S, Kay DB, Rege K. Simultaneous enhancement of photothermal stability and gene delivery efficacy of gold nanorods using polyelectrolytes. ACS Nano 2010, 4, 1769–1770.10.1021/nn1001112Search in Google Scholar

[193] Hervault A, Dunn AE, Lim M, Boyer C, Mott D, Maenosono S, Thanh NT. Doxorubicin loaded dual pH- and thermo-responsive magnetic nanocarrier for combined magnetic hyperthermia and targeted controlled drug delivery applications. Nanoscale 2016, 8, 12152–12161.10.1039/C5NR07773GSearch in Google Scholar PubMed

[194] Sadhukha T, Wiedmann TS, Panyam J. Inhalable magnetic nanoparticles for targeted hyperthermia in lung cancer therapies. Biomaterials 2013, 34, 5163–5171.10.1016/j.biomaterials.2013.03.061Search in Google Scholar PubMed PubMed Central

[195] Espinosa A, Di Corato R, Kolosnjaj-Tabi J, Flaud P, Pellegrino T, Wilhelm C. Duality of iron oxide nanoparticles in cancer therapies: amplification of heating efficiency by magnetic hyperthermia and photothermal bimodal treatment. ACS Nano 2016, 10, 2436–2446.10.1021/acsnano.5b07249Search in Google Scholar PubMed

[196] Jin Y, Gao X. Plasmonic fluorescent quantum dots. Nat. Nano 2009, 4, 571–576.10.1038/nnano.2009.193Search in Google Scholar PubMed PubMed Central

[197] Chen W. Nanoparticle fluorescence based technology for biological applications. J. Nanosci. Nanotechnol. 2008, 8, 1019–1051.10.1166/jnn.2008.301Search in Google Scholar PubMed

[198] Samia AC, Dayal S, Burda C. Quantum dot-based energy transfer: perspectives and potential for applications in photodynamic therapies. Photochem. Photobiol. 2006, 82, 617–625.10.1562/2005-05-11-IR-525Search in Google Scholar PubMed

[199] Chen C, Peng J, Sun SR, Peng CW, Li Y, Pang DW. Tapping the potential of quantum dots for personalized oncology: current status and future perspectives. Nanomedicine 2012, 7, 411–428.10.2217/nnm.12.9Search in Google Scholar PubMed

[200] Chan WC, Nie S. Quantum dot bioconjugates for ultrasensitive nonisotopic detection. Science 1998, 281, 2016–2018.10.1126/science.281.5385.2016Search in Google Scholar PubMed

[201] He R, Gu H. Synthesis and characterization of monodispersed CdSe nanocrystals at lower temperature. Colloids Surf. A: Physicochem. Eng. Asp. 2006, 272, 111–116.10.1016/j.colsurfa.2005.07.017Search in Google Scholar

[202] Bwatanglang IB, Mohammad F, Yusof NA, Abdullah J, Hussein MZ, Alitheen NB, Abu N. Folic acid targeted Mn:ZnS quantum dots for theranostic applications of cancer cell imaging and therapies. Int. J. Nanomed. 2016, 11, 413–428.10.2147/IJN.S90198Search in Google Scholar PubMed PubMed Central

[203] Michalska M, Florczak A, Dams-Kozlowska H, Gapinski J, Jurga S, Schneider R. Peptide-functionalized ZCIS QDs as fluorescent nanoprobe for targeted HER2-positive breast cancer cells imaging. Acta Biomater. 2016, 35, 293–304.10.1016/j.actbio.2016.02.002Search in Google Scholar PubMed

[204] Wang Y, Wu B, Yang C, Liu M, Sum TC, Yong KT. Synthesis and characterization of Mn:ZnSe/ZnS/ZnMnS sandwiched QDs for multimodal imaging and theranostic applications. Small 2016, 12, 534–546.10.1002/smll.201503352Search in Google Scholar PubMed

[205] Qin MY, Yang XQ, Wang K, Zhang XS, Song JT, Yao MH, Yan DM, Liu B, Zhao YD. In vivo cancer targeting and fluorescence-CT dual-mode imaging with nanoprobes based on silver sulfide quantum dots and iodinated oil. Nanoscale 2015, 7, 19484–19492.10.1039/C5NR05620ASearch in Google Scholar PubMed

[206] Li C, Liang S, Zhang C, Liu Y, Yang M, Zhang J, Zhi X, Pan F, Cui D. Allogenic dendritic cell and tumor cell fused vaccine for targeted imaging and enhanced immunotherapeutic efficacy of gastric cancer. Biomaterials 2015, 54, 177–187.10.1016/j.biomaterials.2015.03.024Search in Google Scholar PubMed

[207] Rakovich TY, Mahfoud OK, Mohamed BM, Prina-Mello A, Crosbie-Staunton K, Van Den Broeck T, De Kimpe L, Sukhanova A, Baty D, Rakovich A, Maier SA, Alves F, Nauwelaers F, Nabiev I, Chames P, Volkov Y. Highly sensitive single domain antibody-quantum dot conjugates for detection of HER2 biomarker in lung and breast cancer cells. ACS Nano 2014, 8, 5682–5695.10.1021/nn500212hSearch in Google Scholar PubMed

[208] Ye DX, Ma YY, Zhao W, Cao HM, Kong JL, Xiong HM, Möhwald H.. ZnO-based nanoplatforms for labeling and treatment of mouse tumors without detectable toxic side effects. ACS Nano 2016, 10, 4294–300.10.1021/acsnano.5b07846Search in Google Scholar PubMed

[209] Kong Y, Chen J, Zhou D, Fang H, Heath G, Wo Y, Wang W, Li Y, Guo Y, Evans SD, Chen S. Highly fluorescent ribonuclease-A-encapsulated lead sulfide quantum dots for ultrasensitive fluorescence in vivo imaging in the second near-infrared window. Chem. Mater. 2016, 28, 3041–3050.10.1021/acs.chemmater.6b00208Search in Google Scholar PubMed PubMed Central

[210] Hahn MA, Keng PC, Krauss TD. Flow cytometric analysis to detect pathogens in bacterial cell mixtures using semiconductor quantum dots. Anal. Chem. 2008, 80, 864–872.10.1021/ac7018365Search in Google Scholar PubMed

[211] Morosini V, Bastogne T, Frochot C, Schneider R, François A, Guillemin F, Barberi-Heyob M. Quantum dot-folic acid conjugates as potential photosensitizers in photodynamic therapies of cancer. Photochem. Photobiol. Sci. 2011, 10, 842–851.10.1039/c0pp00380hSearch in Google Scholar PubMed

[212] Martynenko IV, Kuznetsova VA, Orlova AO, Kanaev PA, Maslov VG, Loudon A, Zaharov V, Parfenov P, Gun’ko YK, Baranov AV, Fedorov AV. Chlorin e6-ZnSe/ZnS quantum dots based system as reagent for photodynamic therapies. Nanotechnology 2015, 26, 055102.10.1088/0957-4484/26/5/055102Search in Google Scholar PubMed

[213] Habiba K, Encarnacion-Rosado J, Garcia-Pabon K, Villalobos-Santos JC, Makarov VI, Avalos JA, Weiner BR, Morell G. Improving cytotoxicity against cancer cells by chemo-photodynamic combined modalities using silver-graphene quantum dots nanocomposites. Int. J. Nanomed. 2016, 11, 107–19.10.2147/IJN.S95440Search in Google Scholar PubMed PubMed Central

[214] Qiu JC, Zhang RB, Li JH, Sang Y, Tang W, Rivera Gil P, Liu H. Fluorescent graphene quantum dots as traceable, pH-sensitive drug delivery systems. Int. J. Nanomed. 2015, 10, 6709–6724.10.2147/IJN.S91864Search in Google Scholar PubMed PubMed Central

[215] Zhu H, Zhang S, Ling Y, Meng G, Yang Y, Zhang W. pH-responsive hybrid quantum dots for targeting hypoxic tumor siRNA delivery. J. Control. Release 2015, 220, Part A, 529–544.10.1016/j.jconrel.2015.11.017Search in Google Scholar PubMed

[216] Dong H, Dai W, Ju H, Lu H, Wang S, Xu L, Zhou SF, Zhang Y, Zhang X. Multifunctional poly(l-lactide)-polyethylene glycol-grafted graphene quantum dots for intracellular microRNA imaging and combined specific-gene-targeting agents delivery for improved therapeutics. ACS Appl. Mater. Interfaces 2015, 7, 11015–11023.10.1021/acsami.5b02803Search in Google Scholar PubMed

[217] Li J-M, Wang Y-Y, Zhao M-X, Tan CP, Li YQ, Le XY, Ji LN, Mao ZW. Multifunctional QD-based co-delivery of siRNA and doxorubicin to HeLa cells for reversal of multidrug resistance and real-time tracking. Biomaterials 2012, 33, 2780–2790.10.1016/j.biomaterials.2011.12.035Search in Google Scholar PubMed

Received: 2016-11-23
Accepted: 2017-6-1
Published Online: 2017-7-28
Published in Print: 2017-10-26

©2017 Walter de Gruyter GmbH, Berlin/Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/ntrev-2016-0102/html
Scroll to top button