JBA-06756; No of Pages 17 Biotechnology Advances xxx (2013) xxx–xxx

Contents lists available at ScienceDirect

Biotechnology Advances journal homepage: www.elsevier.com/locate/biotechadv

Research review paper

Nanotechnology-based intelligent drug design for cancer metastasis treatment Yu Gao a,1, Jingjing Xie a,1, Haijun Chen a,b, Songen Gu a, Rongli Zhao a, Jingwei Shao a, Lee Jia a,⁎ a b

Cancer Metastasis Alert and Prevention Institute, College of Chemistry and Chemical Engineering, Fuzhou University, Fuzhou 350002, China Department of Pharmaceutical Engineering, College of Chemistry and Chemical Engineering, Fuzhou University, Fujian 350108, China

a r t i c l e

i n f o

Available online xxxx Keywords: Nanotechnology Nanomedicine Targeted drug delivery system Cancer metastasis therapy Nanoparticle platform

a b s t r a c t Traditional chemotherapy used today at clinics is mainly inherited from the thinking and designs made four decades ago when the Cancer War was declared. The potency of those chemotherapy drugs on in-vitro cancer cells is clearly demonstrated at even nanomolar levels. However, due to their non-specific effects in the body on normal tissues, these drugs cause toxicity, deteriorate patient's life quality, weaken the host immunosurveillance system, and result in an irreversible damage to human's own recovery power. Owing to their unique physical and biological properties, nanotechnology-based chemotherapies seem to have an ability to specifically and safely reach tumor foci with enhanced efficacy and low toxicity. Herein, we comprehensively examine the current nanotechnology-based pharmaceutical platforms and strategies for intelligent design of new nanomedicines based on targeted drug delivery system (TDDS) for cancer metastasis treatment, analyze the pros and cons of nanomedicines versus traditional chemotherapy, and evaluate the importance that nanomaterials can bring in to significantly improve cancer metastasis treatment. © 2013 Elsevier Inc. All rights reserved.

Contents 1. 2.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Current established nanoparticle platforms as drug delivery systems for cancer therapy 2.1. Lipid-based nanoparticle platforms . . . . . . . . . . . . . . . . . . . . 2.2. Polymer-based nanoparticle platforms . . . . . . . . . . . . . . . . . . . 2.3. Protein-based nanoparticle platforms . . . . . . . . . . . . . . . . . . . 2.4. Inorganic nanoparticle platforms . . . . . . . . . . . . . . . . . . . . . 3. Strategies in designing intelligent nanomedicine for enhanced cancer treatment . . . 3.1. Active targeting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Combination drug delivery approaches . . . . . . . . . . . . . . . . . . 3.3. Environment-response controlled release strategies . . . . . . . . . . . . . 3.4. Multi-stage delivery nanovectors . . . . . . . . . . . . . . . . . . . . . 3.5. Cancer nanotheranostics . . . . . . . . . . . . . . . . . . . . . . . . . 4. Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . Declaration of interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction Cancer remains a leading cause of death worldwide (Ferlay et al., 2010). Although years of intense biomedical research and billions of ⁎ Corresponding author. Tel./fax: +86 591 8357 6921. E-mail address: [email protected] (L. Jia). 1 These authors contributed equally to this work.

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

dollars in spending have increased our understanding of the underlying mechanisms of tumorigenesis and biology of cancer, cancer mortality surprisingly reached to the highest point as the top killer in the US population younger than 85 years old (Jemal et al., 2010). Among them, cancer metastasis attributes to approximately 90% of cancer-related deaths (Veiseh et al., 2011). Although immunotherapy, thermal therapy, phototherapy (Jia and Jia, 2012; Shao et al., 2013) and gene therapy are available as cancer treatment modalities, surgery, radiation, and/or

0734-9750/$ – see front matter © 2013 Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

2

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

chemotherapy continue to be the therapeutic options for most cancers over decades, each with its own limitations. Surgery and radiation therapy could be effective for the primary tumor, however, they may not be a good treatment choice for metastases. Chemotherapy with cytotoxic agents is commonly used for the whole-body treatment of recurrent disease. But the conventional anticancer drugs generally result in serious side effects in clinic (Sinha et al., 2006; Stortecky and Suter, 2010; Tsuruo et al., 2003). The side effects are associated with the formulation due to poor water solubility of the drug, non-specific distribution, severe toxicity to normal cells, inadequate drug concentrations at tumors or cancerous cells, and the development of multidrug resistance. Therefore, researchers are continuously seeking for improved anti-cancer therapies that can selectively target tumor cells with minimal side effects on normal tissues (Wang et al., 2008). Nanotechnology is the understanding of materials in the nano (10−9) size range, and involves imaging, measuring, modeling, and manipulating materials within that framework. Since its advent, nanotechnology has revolutionized a wide range of medical products, generic tools and biotechnology equipment. Nanomedicine focuses on application of nanotechnology in medicine for diagnosis, prevention, detection, and treatment of the disease. In particular, it has been used to design and development of targeted drug delivery system (TDDS) which could safely deliver therapeutic drugs to injury sites or specific cells. For formulations intended for i.v. administration, effective TDDSs could retain therapeutic drug in the vehicle, evade the reticuloendothelial system (RES) uptake, target to intended sites of injury, and release drug at the intended sites with required drug concentration (Mills and Needham, 1999). In the field of oncology, TDDS offers many potential benefits such as (1) avoiding the side effects of the clinical formulation for improving solubility, (2) protecting the entrapped therapeutic drug from degradation, (3) modifying pharmacokinetic and tissue distribution profile to increase drug distribution in tumor, (4) reducing toxicity to normal cells, and (5) increasing cellular uptake and internalization in cancer cells. In the past 20 years, many nanomedicines have been in preclinical development and some of them have been approved for use in clinic including for cancer therapy (Davis et al., 2008; Jain and Stylianopoulos, 2010; Peer et al., 2007). Besides used as drug delivery systems (DDSs) for cancer therapy, nanoparticles loaded with imaging agents were also found useful in imaging techniques applied for tumor diagnosis. Here we will focus on TDDS designed for i.v. administration and for delivering anticancer drugs including chemotherapeutic drugs and therapeutic genes. In this review, we first outline the different types of nanoparticle platforms currently being established for cancer treatment. We then present various strategies that have been employed in designing new effective TDDSs. 2. Current established nanoparticle platforms as drug delivery systems for cancer therapy There are diverse types of nanocarriers that have been synthesized for drug delivery including dendrimers, liposomes, solid lipid nanoparticles, polymersomes, polymer-drug conjugates, polymeric nanoparticles, peptide nanoparticles, micelles, nanoemulsions, nanospheres, nanoshells, carbon nanotubes, and gold nanoparticles, etc. (Fig. 1). In all these types, drugs can be entrapped inside, dissolved in the matrix, covalently linked to the backbone, or absorbed on the surface. From the aspect of the property, these nanocarriers could be divided into organic, inorganic, and organic/inorganic hybrid nanoparticles. From the perspective of formulation type, they could be divided into liposomes, micelles, emulsions, nanoparticles, etc (Jia, 2005). Ljubimova and Holler also proposed the term ‘nanopolymer’ meaning a single polymer molecule in the nanoscale range, to distinguish with ‘nano-polymer composites’ such as micelles and other self-assembled or aggregated forms in the point of whether they could dissociate in solutions (Ljubimova and Holler, 2012). Here, we will categorize these current established

nanoparticle platforms based on the difference in composition including lipid-based nanomedicine, polymer-based nanomedicine, peptidebased nanomedicine and inorganic nanomedicine for treating cancer. Some examples of nanomedicines that are approved for commercial use or still in clinical trials are listed in Table 1. 2.1. Lipid-based nanoparticle platforms Lipid-based nanoparticles have attracted great attention as DDS due to their attractive biological properties such as good biocompatibility, biodegradability, low immunogenicity, and the ability to deliver hydrophilic and hydrophobic drugs. Liposomes are the most widely used and studied examples (Jia et al., 2002), with bilayer membrane structures composed of phospholipids for stabilizing drugs, directing their cargo toward specific sites, and for overcoming barriers to cellular uptake. Their aqueous reservoir and the hydrophobic membrane allow them to encapsulate either hydrophilic or hydrophobic agents. The important milestone that led to the development of clinically suitable liposome formulations could be the inclusion of PEGylated lipids in the liposomes to protect liposomes from destruction by the RES, thus to increase circulation time and increase drug accumulation in the tumors. It is worthy to mention that Doxil®/caelyx, a PEGylated liposome formulation of the anticancer drug doxorubicin (DOX), was the first formulation approved for application in the clinic (Barenholz, 2012). With the aim to sitespecific delivery of cancer drugs to the cancerous tissues, the surface of liposomes can be modified with ligands or antibodies targeting those receptors overexpressed on cancer cell membranes (Gabizon et al., 2006). For tumor site-specific triggering drug release, liposomes were designed with responsive to changes in light (Leung and Romanowski, 2012), temperature (Park et al., 2013), acid (Mamasheva et al., 2011) or enzymes (Andresen et al., 2005). Though the work on modification of liposomes has achieved great progress, the application of liposomes in the clinic still poses several challenges including rapid clearance from the bloodstream, instability of the carrier, high production cost, and fast oxidation of some phospholipids. Solid lipid nanoparticles (SLN) is an alternative to liposomes, the matrix of which comprises of solid lipids. They exhibit major advantages such as less cytotoxicity than polymeric counterparts; stable formulations, excellent reproducibility, avoidance degradation of incorporated, controlled drug release, and potential application in intravenous, oral, dermal or topical routes (Uner and Yener, 2007). However, some limitations still exist such as undesired particle growth by agglomeration or coagulation, ineffective drug loading capacity, rapid drug expulsion during storage due to lipid crystallization and high water contents of the dispersions. Thus, modified SLN, so-called nanostructured lipid carriers (NLC) were developed to overcome these limitations and combine the advantages associated with SLN. In contrast to SLN which are made from solid lipids core containing triglycerides, glyceride mixtures, or waxes, NLC were composed of liquid lipid and solid lipid (preferably in a ratio of 30:70 up to 0.1:99.9) to form a nanosized solid particle matrix. The imperfect crystal or amorphous structure assures them to have enhanced drug loading and less drug expulsion during storage (Iqbal et al., 2012). Till now, SLN and NLC as colloidal drug carriers have been successfully multi-functionalized to transport drugs to the targeted cancer cells and achieve efficient drug release in a controlled manner, which confirm their promising application in cancer therapy. 2.2. Polymer-based nanoparticle platforms Polymer-based nanoparticle platforms show enormous potential for treating disease or repairing damaged tissues especially for cancer treatment, which relies on their remarkable properties including small size, excellent biocompatibility and biodegradability, prolonged circulation time in the bloodstream, enhanced drug loading capacity, and easy chemical modification or surface functionalization. The last two characters are the utmost important criteria for their clinical use. Generally

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

3

Fig. 1. Schematic illustration of representative nanoparticle platforms that have been synthesized for drug delivery for cancer therapy.

speaking, polymer-based nanomedicine can be categorized into three groups based on drug-incorporation mechanisms including polymerdrug conjugates by covalent conjugation, polymeric micelles by hydrophobic interactions, and polyplexes or polymersomes by encapsulation. Most of the polymers are approved by FDA as the commonly explored carriers for targeted drug delivery. Polymer-drug conjugates using water-soluble polymers as carriers have produced expected results, including a water-soluble polymeric carrier (natural or synthetic), a biodegradable linkage and an anticancer agent. Because polymer-drug conjugates can passively target to tumor cells by enhanced permeability and retention (EPR) effect (Matsumura and Maeda, 1986), many polymer-drug conjugates were under clinical evaluation. The most attractive representative of synthetic polymerdrug conjugates is poly (L-glutamic acid)-paclitaxel (CT-2103, Xyotax®), which has advanced to Phase III clinical trials (Bonomi, 2007). Polymerdrug conjugates have exhibited several superiorities such as enhanced therapeutic efficiency, fewer side effects, flexible drug administration and even increased patient compliance. However, many challenges still exist in the development of new generation of polymer-drug conjugates, including the design of novel polymers that have modulated degradation characteristics, polymerization methods allowing for controlling the weight distribution, and conjugation techniques available for sitespecific attachment. Amphiphilic block copolymers can self-assemble into different kinds of mesoscopic structures (micelles and vesicles), which is just up to the control about the volume ratio of hydrophilic to hydrophobic blocks (Antonietti and Förster, 2003; Zupancich et al., 2006). Polymersomes are self-assembled polymer vesicles formed by amphiphilic copolymers containing hydrophilic and hydrophobic segments, which are different from liposomes formed by amphiphilic phospholipids. The hydrophilic interior structure is suitable for encapsulating with water-soluble agents such as DNAs or proteins while the hydrophobic exterior bilayer membrane can be simultaneously entrapped with poorly water-soluble drugs. Compared with liposomes, polymersome exhibited more

prominent features such as higher loading capabilities, greater stabilities, and longer circulation time. The improvement of storage abilities is attributed to their own large hydrophic core and surface functionality through chemical synthesis and modification (Ghoroghchian et al., 2005, 2006). Polymeric micelles are self-assembling monolayers formed spontaneously under certain conditions including the concentrations of amphiphilic surfactants, pH, temperatures and ionic strength with a hydrophobic core and hydrophilic shell in the nanometer range. The properties of polymeric micelles such as small size, hydrophilic shell avoiding the uptake by the mononuclear phagocyte system (MPS), and the high molecular weight evading renal excretion made them effective passive targeting systems. Ligands such as small organic molecules, DNA/RNA aptamers, peptides, carbohydrates and monoclonal antibodies could be attached to the surface of micelles, not only increasing the accumulation at tumor sites but also increasing the cellular uptake in cancer cells via receptor-mediated endocytosis (Farokhzad et al., 2006; Sethuraman and Bae, 2007; Torchilin et al., 2003; Yoo and Park, 2004). Dendrimers are kinds of nanomaterials with super biological characteristics: small size (1–15 nm), high water solubility, regularly and highly branched three-dimensional architecture, nearly perfect monodispersibility in nature, and high payload. All these facilitate their applications in cancer or disease prevention and therapy. Polyamidoamine (PAMAM) dendrimer was one of the most studied dendrimers. It possesses multiple amine surface groups, and the number of the groups could be precisely controlled. Therefore, the multivalent conjugation could be achieved by attachment of targeting ligands, therapeutics agents, drugs, imaging contrast agents, genes or even chemical sensors to their terminal functional groups. M.H. Li et al. (2012) prepared the G5 PAMAM dendrimer-based multivalent methotrexates as dual acting nanoconjugates for cancer cell targeting. The study demonstrated that re-engineering dendrimer conjugates not only target KB cancer cells, but also inhibited dihydrofolate reductase.

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

4

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

Table 1 Examples of current established nanoparticle platforms approved for commercial use or undergoing clinical investigation for cancer therapy. Product

Nanoparticle platform

Drug

Current stage of development

Type of cancer

Doxil/caelyx

PEGylated liposomes

Doxorubicin

Approved by FDA

Myocet

Non-PEGylated liposomes

Doxorubicin

Approved in Europe and Canada

DaunoXome Abraxane ALN-VSP CRLX101 CALAA-01

Liposomes Albumin-bound nanoparticles Lipid nanoparticles Cyclodextrin nanoparticles Cyclodextrin-containing linear polymer, decorated with PEG and transferrin Polymer micelle PSMA-targeted polymeric nanoparticles Pegylated colloidal gold nanoparticles Heat-activated liposome Liposomes

Daunorubicin Paclitaxel siRNA Camptothecin siRNA

Approved in the USA Approved by FDA Phase I Phase II Phase I

Refractory Kaposi's sarcoma, recurrent breast cancer, ovarian cancer Combinational therapy of metastatic breast cancer, ovarian cancer, Kaposi's sarcoma Advanced HIV-associated Kaposi's sarcoma Various cancers Liver cancer Various cancers Solid melanoma tumors

Doxorubicin Docetaxel TNFi Doxorubicin Irinotecan and floxuridine

Phase I Phase I Phase II Phase III Phase II

Various cancers Advanced solid tumor cancers Solid tumors Hepatocellular carcinoma Advanced solid tumors

NK-911 BIND-014 Aurimune ThermoDox CPX-1

Thomas and his co-workers used antibody-conjugated dendrimers to bind antigen-expressing cells. The conjugates specifically bound to the antigen-expressing cells in a dose-and time-dependent manner with affinity similar to that of the free antibody (Thomas et al., 2008). 2.3. Protein-based nanoparticle platforms Protein-based nanomedicine platforms as one of the representatives have been paid serious attention owing to their biocompatibility, biodegradability as well as low toxicity. Protein-based nanomedicine platforms are usually consisted of naturally protein subunits of the same protein or the combination of natural or synthetic protein, and different types of drug molecules. There are a variety of proteins used and characterized for DDSs such as the plant-derived viral capsids (Liepold et al., 2007; Suci et al., 2007), the small Heat shock protein (sHsp) cage (Flenniken et al., 2005, 2006), albumin (Kratz, 2008; W. Lu et al., 2007), soy and whey protein (Chen et al., 2008; Gunasekaran et al., 2006), casein (Latha et al., 2000), collagen (Metzmacher et al., 2007) and the ferritin/apoferritin protein cage (Wu et al., 2008a, 2008b). The protein cage with hierarchical architectures derived from viruses has its various advantages on the cage's uniform nanometer size for drug loading and for avoidance of macromolecular aggregation, multifunctional groups on the surface available for conjugation with drugs, and superior biological characteristics beneficial for pharmacokinetics study. Albumin as a versatile protein carrier for improving drug targeting and pharmacokinetic properties is playing a vital role in the development of protein-based nanoparticles. It demonstrates prominent features of stability in a broad range of pH (4–9) and temperature (4 °C–60 °C), preferential uptake by tumor, and non-toxicity. Methotrexate-albumin conjugate, albumin-binding prodrug of DOX and albumin PTX nanoparticle (Abraxane) have been designed and now are under clinical trials (Miele et al., 2009). 2.4. Inorganic nanoparticle platforms Organic nanoparticles such as liposomes, dendrimers, polymeric micelles have made great advances in cancer diagnosis and therapy (Khemtong et al., 2009; Ljubimova et al., 2008). In contrast, inorganic nanoparticles such as gold nanoparticles (AuNPs), carbon nanotubes (CNTs), silica nanotubes, quantum dots (QDs), and super-paramagnetic iron oxide nanoparticles (SPIOs) have also been extensively developed and studied for biomedical applications due to their intrinsic unparallel physical and biological properties such as optical, electrochemical, magnetic characteristics. The biomedical applications of CNTs have been gradually proposed and recognized through preliminary studies in vitro and in vivo and even clinical tests, which is ascribed to their prominent physical and chemical properties. In general, CNTs can be classified to two categories:

single-walled carbon nanotubes (SWNTs, 0.4–2.0 nm in diameters, 20–1000 nm in lengths) and multi-walled carbon nanotubes (MWNTs, 1.4–100 nm in diameters, ≥1 μm in lengths). MWNTs provide potential platforms for large biomolecules delivery such as plasmids into cells, which is mainly due to the multiple layers of grapheme and larger diameters (Gao et al., 2006; Liu et al., 2005). SWNTs exhibit more attractive optical properties suitable for biological imaging (Cherukuri et al., 2004; O'Connell et al., 2002; Welsher et al., 2008). Functionalized SWNTs by covalent binding, adsorption, and electrostatic interaction can serve as novel drug delivery carriers in cancer therapy owing to their biocompatibility, little toxicity and enhanced water solubility (Feazell et al., 2007; Liu et al., 2007). Liu et al. (2008) prepared the SWNT-PTX conjugate by coupling PTX to the branched PEG chains on SWNTs and studied its antitumor effects in a xenograft murine 4T1 breast cancer model. They showed that SWNT delivery of PTX into xenograft tumors could have 10-fold higher tumor suppression efficacy than the clinical drug formulation Taxol. Distinction from other nanomaterials, mesoporous silica nanoparticles (MSNs) showed unique properties such as tunable particle size from 50 to 300 nm convenient for cell endocytosis; stable and rigid framework resistant to degradations induced by pH, heat, and mechanical stress; uniform and tunable pore size adjusted between 2 and 6 nm for the loading of different drug molecules; high surface area and large pore volume allowing for high drug loading; internal and external functional surfaces available for selective modification. MSNs could be functionalized through co-condensation, grafting, and imprint coating methods (Burleigh et al., 2001; Chen et al., 2006). The different surface functionalization of MSNs has great effects on the cellular uptake mechanism and the internalization efficiency of MSNs as well as the ability to escape the endolysosomes (Slowing et al., 2006). MSNs were reported having better biocompatibility compared with other silica-based materials. The viability of mammalian cells wasn't affected by the internalization of MSNs at concentrations below 100 μg/ml (Slowing et al., 2008). Similar results were found that injecting MSNs to the animals didn't pose any toxic side effects for 42 days (Kortesuo et al., 2000). Therefore, MSNs were widely employed as promising intracellular controlled release drug delivery carriers in cancer treatment (Slowing et al., 2007; Trewyn et al., 2007; Vallet-Regi et al., 2007). J. Lu et al. (2007) incorporated the hydrophobic anticancer drug camptothecin (CPT) into the pores of the prepared fluorescent mesoporous silica nanoparticles (FMSNs) and successfully achieved the controlled drug release to human cancer cells and induced tumor cell death. Magnetic nanoparticles (MNPs) have their own unique physical and biological features including controllable size distribution ranging from nanometers to micrometers, high magnetic flux density with the intrinsic penetrability for drug targeting, the ability to convert magnetic to heat, non-toxicity, biocompatibility, and injectability (Ito et al., 2005; Pankhurst et al., 2003). The magnetic nanoparticle-based DDSs could

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

be constructed by loading drug onto the particle coat via physical means such as electrostatic interaction instead of covalent conjugation (Kievit et al., 2011; Medarova et al., 2007). The physical, hydrodynamic, and physiological parameters have great effects on the drug delivery efficiencies of magnetic nanoparticles. Among the MNPs, SPIOs with the diameter of 5–100 nm, which show high magnetization in an external magnetic field, have demonstrated attractive possibilities in biomedical application. They could serve as good “nanotheranostics” for both targeted drug delivery and magnetic resonance imaging of tumor cells (Lee et al., 2013; Mouli et al., 2013; Yang et al., 2008; Zou et al., 2010). Usually, SPIOs are loaded with small-molecule-based therapeutics into polymer-based matrices (Quan et al., 2011). Great interest has been paid to gold nanoparticles (AuNPs) in recent years for their attractive properties including the strong and attractive optical properties in the near-infrared (NIR) region from 700 to 900 nm (Jain, 2009; Xia et al., 2011), easy modification with functional groups through formation of stable gold-thiolate bonds (Au\S) by reacting with disulfide (S\S) or thiol (\SH) groups (Huang et al., 2013), controllable particle size, shape and geometry (Kim et al., 2009), and diversely multi-functionalization with desired targeting ligands, specific antibodies or drugs. The routine applications of AuNPs in cancer therapy were photothermal therapy and radiation therapy, respectively, for their strong absorption cross-sections and X-ray emission characteristics (Sperling et al., 2008). AuNPs were also used as nanocarriers for drug delivery. Several strategies have been used to improve AuNPs accumulation in tumor cells specifically and efficient intracellular drug release, including the conjugation of AuNPs with appropriate surface ligands (membrane-translocating peptides) or specific antibodies (Huang et al., 2008), the coupling drugs of AuNPs through non-covalent (available for drug release) or covalent binding (requiring for second release), the external triggering methods such as glutathione (Hong et al., 2006), light or photothermal-mediated release (Agasti et al., 2009; Bikram et al., 2007), and the surface modification with amphiphilic reagents (PEG). Though advances have been made in the research field of AuNPs as TDDS for cancer therapy, more challenges are still confronted. The suitable types of AuNPs used as drug delivery (Cai et al., 2008; Chithrani et al., 2006), the delivery efficiency, the accuracy of targeting as well as the toxicity (Pan et al., 2009) were under re-evaluation and optimization prior to clinical application. 3. Strategies in designing intelligent nanomedicine for enhanced cancer treatment 3.1. Active targeting Active targeting utilizes targeting moieties to peripherally conjugate to nanoparticle systems for specifically targeting to tumor tissue, specific cancer cells, or even cellular organelles (Fig. 2). The most common used active targeting strategy involves the attachment of the targeting ligands such as folic acid, antibodies, aptamers, or proteins to the nanoparticles which recognizes receptors over-expressed on cancer cells (Ruoslahti et al., 2010). For example, in FR positive KB cells, uptake of folate-targeted liposomal arsenic PEGylated liposomes inserted with a small amount of DSPE-PEG3350-folate (0.3 mol%) was three to six times higher than that of nontargeted liposomal arsenic, leading to a 28-fold increase in cytotoxicity (H. Chen et al., 2009). Various antibodies that target receptors over-expressed on the surface of cancer cells such as vascular endothelial growth factor (VEGF), human epidermal receptor-2 (HER-2), tumor necrosis factor-α (TNFα), and epidermal growth factor receptor (EGFR) have been attached to nanoparticulate materials to achieve selective cancer cell targeting (Fay and Scott, 2011). When developing the antibody-conjugated nanoparticles, the affinity and the configuration of the antibodies, as well as the method to attach to the nanoparticles, are all key design factors that should be taken into consideration (Cheng and Allen, 2008; Firer and Gellerman, 2012; Rizk et al., 2009; Rudnick et al., 2011). The

5

characteristics of the antibodies will have great effects on the circulation time, cellular uptake, tolerability, and efficacy of the nanoparticulate systems. In a recent study, two gold nanoparticles with the surface partiallycovered and surface fully-covered by EGFR antibody cetuximab were designed to determine the cellular uptake mechanism of cetuximabconjugated nanoparticles. The endocytosis mechanism could be switched from a Cdc42-dependent pinocytosis/phagocytosis to original Dyn-2-dependent caveolar pathway when the nanoparticles were fully coated with cetuximab (Bhattacharyya et al., 2012). Compared with antibodies, antibody fragments such as Fab' and scFv are more widely used for active targeting because they have a smaller size easy to conjugate into nanoparticles. More importantly, they could keep their targeting specificity while reducing nonspecific antigen binding from Fc (Ansell et al., 2000; Chapman, 2002; Rothdiener et al., 2010; Sapra et al., 2004). Conjugation of the antibody on nanoparticles can be carried out by coupling the carboxylic acid groups or primary amine groups of the amino acid in the antibody to the primary amine groups or the carboxylic acid groups on the surface of the nanoparticles (Chapman, 2002). The interaction between biotin and avidin or streptavidin has also been exploited for designing antibody-conjugated nanoparticles. Wartlick et al. covalently modified the nanoparticles based on gelatin and HSA with the biotin-binding protein NeutrAvidin followed by the biotinylated herceptin conjugation to the surface of the nanoparticles. These nanoparticles could effectively internalize into HER-2overexpressing cells via receptor-mediated endocytosis observed by confocal laser scanning microscopy (Wartlick et al., 2004). Aptamers are kinds of oligonucleotides that are capable of binding to a large number of targets with high affinity and specificity. Since the advent of aptamer technology (Ellington and Szostak, 1990; Tuerk and Gold, 1990), aptamers have represented an interesting class of modern pharmaceuticals for therapy and diagnostics. Known as “chemical antibodies”, aptamers show many similar properties to traditional antibodies, however, they demonstrated a number of advantages over antibodies such as low immunogenic potential, easier to synthesize and modify, structural flexibility, higher affinity and specificity, and good stability (Majumder et al., 2009). Recently, aptamers have been conjugated to many types of molecules such as siRNAs, miRNA, proteins, and nanoparticles to improve their targeting efficiency, stability, and biodistribution profile (Kanwar et al., 2011). The chimeric aptamernanoparticle conjugates can bind to the target cells by the interaction of aptamer–receptor interaction, and finally enter into the target cells, resulting in release of the entrapped drugs (Estevez et al., 2010). In a study, branched PEI-PEG polyplexes modified with an anti-prostatespecific membrane antigens (PSMA) aptamer were used to co-deliver DOX and Bcl-xL-specific shRNA. The polyplexes could induce a synergistic and selective cancer cell death in PSMA-overexpressing prostate cancer cells (E. Kim et al., 2010). With the development of different kinds of biomolecular ligands, more and more new molecular-targeted nanoparticles are designed to target different receptors. Considering that one kind of ligand is commonly specific to only one or a limited few target receptors, and conjugation of large targeting molecules to nanoparticles can change their behaviors in vivo, pre-targeting strategy has been utilized to avoid development of multiple nanoparticle formulations with different targeting ligands and accommodate different targeting ligands without alternating pharmacokinetic profile of the nanoparticles. Pre-targeting is a multi-step process that first has a targeting ligand localized within a tumor by virtue of its anti-tumor binding site, followed by treatment with nanoparticles that recognize the targeting ligand conjugate on the cell surface. Using a cell targeting recombinant fusion protein (FP) composed of a single-chain antibody (scFv) and streptavidin (SA) to specifically pre-label the targeted cells, followed by application of a biotinylated nanoparticle that binds to the SA of the FP on the target cells, the nanoparticle system with two FPs, anti-CD20 and anti-TAG-72 CC49, could specifically target two model cancer cell lines, i.e., Ramos and Jurkat, respectively (Gunn et al., 2011). Besides using the

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

6

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

Fig. 2. Schematic illustration of active targeting strategies that have been used for design intelligent drug delivery systems for cancer therapy. The active targeting strategy involves the attachment of the targeting ligands such as folic acid, antibodies, aptamers, or proteins to the nanoparticles for specifically targeting to tumor tissue, tumor vasculature, specific cancer cells, or even cellular organelles (nucleus, cytoplasm, mitochondria). Pre-targeting is a multi-step process that first has a targeting ligand localize within a tumor by virtue of its antitumor binding site, followed by treatment with nanoparticles that recognize the targeting ligand conjugate on the cell surface. Dual-targeting strategy and sequential-targeting strategy have been applied to develop nanoparticles that can both penetrate the BBB and target the glioma cells.

interaction between biotin and avidin or streptavidin, bioorthogonal chemistry such as tetrazine/trans-cyclooctene cycloaddition reaction has also been employed to nanoparticles recognition of pre-labeled cells (Devaraj et al., 2009; Haun et al., 2010; Rossin et al., 2010). Dual-targeting strategy has been applied to develop nanoparticles for treatment of brain tumor. The treatment of brain tumor entails efficient delivery of therapeutic agents to specific regions of the brain after penetrating the blood–brain barrier (BBB). The BBB is a physiological barrier that selectively allows the entry of certain molecules. Many strategies have been employed to across the BBB such as the disruption BBB integrity by osmotic means or by ultrasound means, the use of endogenous carrier-mediated transporters, receptor-mediated transcytosis, and blocking of active efflux transporters. Nanotechnology is considered to be one of the most promising methods to deliver drugs across the BBB by attaching BBB-penetrating ligands to the surface of nanoparticles (Yang, 2010). To targeting infiltrated glioma cells or the glioblastoma multiform after crossing the BBB, nanoparticles were modified with dual ligands such as angiopep, transferrin, wheat germ agglutinin, which recognize the receptors over-expressed on both BBB and glioma cells for transporting the drug across the BBB and then targeting brain glioma (Du et al., 2009; He et al., 2011; Xin et al., 2011; Y. Li et al., 2012; Ying et al., 2010). Recently, a sequential-targeting strategy has been applied to develop nanoparticles that can come across the BBB and recognize the glioma cells subsequently. The exterior of the micelle was conjugated with transferrin to enhance the cellular uptake and BBB-penetrating through receptor mediated endocytosis. Then the loaded drug cyclo-[Arg–Gly–Asp–d–Phe–Lys] (c[RGDfK])-PTX conjugate (RP) was released from micelle subsequently to target integrin

over-expressed glioma cells (Zhang et al., 2012). This sequentialtargeting nanoparticulate system could not only protect the ligands from degradation during transportation across the BBB (Knisely et al., 2008), but also overcome the non-specific recognition of the receptors that are highly expressed throughout the brain (Bu et al., 1994). Many research groups have studied the mechanism of active targeting in solid tumors with ligand-modified nanoparticles. The improvement of cellular uptake of ligand-modified nanoparticles could be achieved through receptor-mediated endocytosis by tumor cells over-expressing corresponding receptors on the surface. However, whether ligand-modified nanoparticles could increase drug accumulation at the tumor site is largely dependent on the ligands. In a subcutaneous KB-3-1 xenograft model, the administration of the nanoparticles formed by ternary conjugate heparin-folic acid-PTX and additional PTX (HFT-T) enhanced the specific delivery of PTX into tumor tissues (Wang et al., 2009). Using transferrin-containing gold nanoparticles as study model, Choi et al. found that the content of targeting ligands significantly influences the number of nanoparticles localized within the cancer cells (Choi et al., 2010). Some contrary results were also found in some nanoparticulate systems such as antibody targeting of longcirculating lipidic nanoparticles and chlorotoxin labeled magnetic nanovectors that the targeting ligand only enhanced cellular uptake of nanoparticles, but did not affect the accumulation of nanoparticles at the tumor site (Kievit et al., 2010; Kirpotin et al., 2006). Besides the cancer cells, tumor vasculature is also a potential target for drug delivery (Jain and Stylianopoulos, 2010). In vivo phase display is a very useful tool to identify numerous peptides targeting the tumor vasculature (Li and Cho, 2012; Trepel et al., 2008). Several peptides

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

such as arginine–glycine–aspartic acid (RGD), asparagine–glycine– arginine (NGR) (Arap et al., 1998), TCP-1 (Li et al., 2010), iRGD (Sugahara et al., 2009; Sugahara et al., 2010), F3 (Porkka et al., 2002), and LyP-1 (Laakkonen et al., 2002) have been identified to specifically target tumor blood vessels or simultaneously recognize the tumor vasculature and cancer cells. Nanoparticles modified with peptides that targeting the tumor vasculature could enhance efficacy of the chemotherapeutic agents against human cancer xenografts (Chang et al., 2009), suppress tumor growth (Hood et al., 2002), and metastasis in mice (Murphy et al., 2008). Active targeting not only can direct the nanoparticles to specific cells of the tumor, but also been applied to deliver cargos to specific cellular organelles. As oncogenes and tumor suppressor genes play a crucial role in the processes of cancer development, gene therapy provides a tool to cure the disease at its source. The DNA-based medicines require carriers efficiently and safely carry the plasmid DNA into the nucleus of the desired cells. The cationic non-viral vectors which could interact with negatively charged DNA through electrostatic interactions to form polyplexes or lipoplexes, have been considered to be the most promising gene delivery systems compared to naked DNA or viral vectors (Al-Dosari and Gao, 2009). To successfully deliver genes into the nucleus, the cationic nanoparticles should circumvent a series of barriers including survival in the bloodstream, extravasation into the tissue, binding and internalization into the target cells, escaping from endosome, subcellular trafficking, and finally entry into nucleus (Khalil et al., 2006; Wiethoff and Middaugh, 2003). Numerous efforts have been made to develop effective and safe gene delivery systems and a large amount of strategies has been employed to improve systemic delivery and intracellular trafficking of cationic nanoparticles including attachment of ligand for cell-specific targeting and receptor-mediated endocytosis, use of protein transduction domain (PTD) such as TAT peptides to mediate cellular transduction, incorporation of pH-sensitive endosomolytic peptides, fusogenic peptides or membrane-destabilizing compounds to facilitate endosomal escape, employment of nuclear localization sequence to improve the uptake of plasmid DNA into the nucleus (Morille et al., 2008). Some multifunctional nanoparticles that could circumvent several biological barriers have been designed. For example, a multifunctional nano device consisting of poly(folate-poly(ethylene glycol)cyanoacrylate-co-hexadecylcyanoacrylate), dioleoyl phosphatidylethanolamine (DOPE), and DNA condensed by protamine sulfate (PS) was developed for nucleus delivery of DNA. The folic acid on the surface of the nano device could increase its active targeting ability to cancer cells. The PEG chain within the polymer could decrease its macrophages recognition and extend its half-life in blood circulation. DOPE could facilitate endosomal escape and PS could be served for nuclear transfer (Gao et al., 2007). Unlike the pDNA delivery to the cell nucleus, siRNA delivery involves fewer barriers because the target of delivery of siRNA is in the cytoplasm (Nguyen and Szoka, 2012). It is worthy to mention that a complex nanoparticle formulation CALAA01 (Calando Pharmaceuticals, Inc.), which consists of cyclodextrinbased polymer, transferrin targeting ligand, a hydrophilic PEG chain, and siRNA targeting ribonucleotide reductase M2, a critical biomolecule in DNA synthesis, has been recently shown to effectively deliver siRNA to humans (Davis et al., 2010). Mitochondria are playing an important role in regulating cell metabolism and cell death, and are involved in diverse physiological activities in the course of cancer development and progression. As an alternative subcellular target, some novel mitochondrial TDDSs have been developed (Yamada and Harashima, 2008). Torchilin's group developed mitochondrial-targeted liposomal drug-delivery system by incorporation a mitochondriotropic dye rhodamine-123 (Rh123)-PEG-DOPE into the liposomal lipid bilayer (Biswas et al., 2011) or by modification of a mitochondria-targeting triphenylphosphonium cation to liposome surfaces (Boddapati et al., 2008). The mitochondria-targeting liposomes could efficiently deliver the model drug to mitochondria to enhance its activity.

7

3.2. Combination drug delivery approaches To achieve better treatment efficacy, multimodality treatment or combination treatment is commonly used to treat cancer. Compared with single-modality treatment, multimodality treatment can do an excellent job with additive or even synergistic efficacy. On the one hand, using drugs acting through different molecular targets could delay or block the cancer adaptation processes from different aspects. On the other hand, the drugs with the same molecular target could function synergistically for higher therapeutic efficacy. The combination of chemotherapy, biologic therapy, endocrine therapy and/or thermotherapy has been investigated for their synergistic effects recently (Goodwin et al., 2012; Mehta et al., 2012). Although combination therapy has synergistic efficacy, it could also lead to increased toxicity in some cases. Some anticancer agents are mixed together for administration but they are eliminated independently, which could cause the additive adverse effects. The combination of taxanes with anthracyclines in first-line chemotherapy for metastatic breast carcinoma produces a significant benefit in activity, but with a significant cost in hematologic toxicity (Bria et al., 2005). In phase III trial, combination of PTX and bevacizumab significantly prolonged progression-free survival as compared with PTX alone, however, grade 3 or 4 hypertension, proteinuria, headache, and cerebrovascular ischemia were more frequent in patients receiving PTX plus bevacizumab (Miller et al., 2007). In addition, combination of anticancer agents which have different routes of administration would decrease patient satisfaction and compliance. For example, when employing the combination therapy of lapatinib and trastuzumab to treat ErbB2-positive breast cancer, patients needed to receive doses of lapatinib administered once daily (continuous) in combination with trastuzumab weekly (Storniolo et al., 2008). Nanomedicine can provide a fantastic platform for multimodality treatment. Nanoparticulate DDSs demonstrate many advantages over conventional formulations by physically blending multiple drugs together including: (1) improved solubility and bioavailability, (2) escape of elimination by macrophages and prolonged drug circulation half-life, (3) increased tumor site accumulation by passive or active targeting, (4) efficient internalization with the mechanism of endocytosis, (5) controlled pharmacokinetics of each drug, resulting in enhanced drug efficacy and reduced side effects. Many nanoparticulate platforms have been developed as co-delivery systems that can deliver a combination of small molecule drugs or a combination of small molecule drugs and macromolecular therapeutics. Liposomes are commonly used as co-delivery carriers with the ability to load both hydrophilic and hydrophobic drugs. In a study by Wong et al., co-encapsulation of vincristine and quercetin into a liposome formulation exhibited significant antitumor activity than free vincristine/ quercetin combinations (Wong and Chiu, 2011). Stealth liposomes (Ong et al., 2011) and targeted liposomes (Wu et al., 2007) were also developed as co-delivery carriers for cancer therapy. Some combination DDSs based on liposomes are currently in clinical trial. Liposomes encapsulated with irinotecan and floxuridine (CPX-1) (Batist et al., 2009) and liposomes containing cytarabine and daunorubicin (CPX351) (Feldman et al., 2011) have been entered into phase I trials. To ensure the product quality of complex liposome formulations, Zucker et al. showed the methods to characterize the critical features of LipoViTo where vincristine (VCR) and topotecan (TPT) were encapsulated in the same nanoliposome. The characterization methods are useful for a rational clinical development of liposomal formulations alike (Zucker et al., 2012). Cationic lipoplexes were initially proposed to co-delivery of anticancer drug and gene. The lipoplexes developed by L. Huang's team could co-deliver DNA and inflammatory suppressors into one immune cell (Liu et al., 2004). Co-delivery of antitumor agent docetaxel and DNA which suppress surviving protein, was achieved by a folate-modified multifunctional lipoplexes as a therapeutic approach for human

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

8

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

hepatocellular carcinoma (Xu et al., 2010). Cationic liposomes were also applied to co-deliver siRNA and an anticancer drug (Saad et al., 2008; Shim et al., 2011). In the preparation of lipoplexes, a condenser is always needed to condense pDNA to form the core of the cationic liposomes. But the cationic liposome:siRNA complexes are always formed by mixing siRNA and cationic liposomes together. Polymer-based nanoparticles have been extensively studied as codelivery carriers as the drug can be entrapped inside or covalently linked to the polymer matrix. Polyalkylcyanoacrylate nanoparticles were used to entrapped DOX and chemo-sensitizing compound cyclosporin A to achieve the synergistic effect in multidrug resistance (MDR) cancer cells (Soma et al., 2000) Amphiphilic block copolymers such as PEG-PLGA and PEG-PLA could self-assemble into micelles to co-deliver two chemotherapeutic agents (H. Wang et al., 2011; Shin et al., 2009). Micelles formed by cationic copolymers such as amphiphilic copolymer poly{(Nmethyldietheneamine sebacate)-co-[(cholesteryl oxocarbonylamido ethyl) methyl bis(ethylene) ammonium bromide] sebacate} (P(MDSco-CES)) (Wang et al., 2006) and triblock copolymers poly(N,Ndimethylamino-2-ethylmethacrylate)-polycaprolactone-poly(N,Ndimethylamino-2-ethyl methacrylate) (PDMAEMA-PCL-PDMAEMA) (Zhu et al., 2010) have been used to co-deliver genes and drugs to the same cells, in which hydrophobic anticancer drug PTX was entrapped in the core of the micelles, and genes were complexed onto their surface. PEI-graft-poly(ε-caprolactone) copolymer was also reported to codeliver DOX and a reporter gene (Qiu and Bae, 2007). A core-shell nanoparticle formed from folate coated PEGlated lipid shell and PLGA core was also developed for targeted co-delivery of drug and gene (Wang et al., 2010). Recently, a nanoparticle system formed by blend of poly(lactide)-D-α-tocopheryl polyethylene glycol succinate and carboxyl group-terminated TPGS (TPGS-COOH) copolymer was developed for triple modality treatment of cancer. The nanoparticles were formulated with docetaxel, herceptin and iron oxides for the chemo, bio, and thermo therapies (Mi et al., 2012). Besides entrapped inside the polymer-based nanoparticle systems, two or more drugs could be conjugated to one kind of polymer carrier for combination delivery. The aromatase inhibitor aminoglutethimide and DOX were simultaneously conjugated to HPMA copolymer. The results showed that the conjugate carrying two drugs was more potent than the combination of two polymer conjugates carrying only one drug (Vicent et al., 2005). The HPMAbased polymer was also used to simultaneously deliver DOX and antiinflammatory drug dexamethasone (DEX) (Krakovicova et al., 2009). In another study, DOX and gemcitabine were co-conjugated to the same HPMA polymer, which increased the efficacy of the combination of gemcitabine and DOX without increasing its toxicity (Lammers et al., 2009). Dendrimers have also been used for successfully synchronous delivery of two therapeutic agents by either complexation of two drugs or by conjugation of two drugs in the same polymer, as well as by complexation of one drug and simultaneous conjugation of another drug (Clementi et al., 2011; Kaneshiro and Lu, 2009; Kim et al., 2011; Lee et al., 2011). Also, nanoparticles formed by blending of selfassembled amphilic copolymer in the presence of a chemotherapeutic agent and polymer-prodrug conjugate could be used to achieve combination drug delivery (Kolishetti et al., 2010). Some inorganic nanoparticulate systems have been successfully employed for co-delivery. Successfully synchronous delivery of chemotherapeutic drug and siRNA was achieved by mesoporous silica nanoparticles coating with the cationic polymer. MSNs modified with PAMAM dendrimers were utilized to simultaneously deliver DOX and a Bcl-2 siRNA into MDR cancer cells (A. M. Chen et al., 2009). PEI functionalized MSNs were also used to synchronously deliver DOX and P-gp siRNA to reverse drug resistance in MDR cancer cells (Meng et al., 2010). The in vivo efficacy of the use of this dual drug/siRNA nanocarrier was also tested in a xenograft to overcome DOX resistance (Meng et al., 2013). Positively charged ammonium-functionalized MSNs could also immobilize negatively charged single strand DNA (ssDNA) for drug/ssDNA co-delivery (X. Ma et al., 2012). In a recent

work, hydrophilic-hydrophobic anticancer drug pairs, such as doxorubicin–paclitaxel and doxorubicin–rapamycin, could be loaded into magnetic mesoporous silica nanoparticles for simultaneous delivery of hydrophilic and hydrophobic drugs for combination treatment (Q. Liu et al., 2012). Magnetic nanoparticles embedded in polylactide-coglycolide matrixes were also designed as drug delivery and imaging vector for loading both hydrophilic and hydrophobic drugs (A. Singh et al., 2011). Layer-by-layer assembled charge-reversal functional gold nanoparticles were employed to co-deliver siRNA and plasmid DNA into cancer cells (Guo et al., 2010). Some hybrid systems such as lipid-polymer hybrid systems, lipidinorganic silica hybrid systems were developed for combination drug delivery. A polymer-caged nanobin (PCN) formed by liposomal core encapsulated with DOX and a polymer shell conjugated with cisplatin prodrug was designed to co-deliver of DOX and cisplatin. The PCN could exert synergistic cytotoxic effects of each drug against cancer cells at reduced doses (Lee et al., 2010). In another study, the protocells were designed with nanoporous silica cores enveloped by a lipid bilayer, which was further functionalized with poly(ethylene glycol),targeting peptides, and pH-responsive peptides, to deliver combinations of diverse drug cargos such as quantum dots, small molecules and oligonucleotides (Ashley et al., 2011). 3.3. Environment-response controlled release strategies Development of stimuli-responsive nanoparticles is a particularly appealing approach for the goal of increasing the specificity of drug delivery in vivo. Environmentally-responsive nanoparticles have the ability to produce physicochemical changes that regulate drug release at the target site when exposed to external stimuli. The differences between tumor environment and normal tissue such as pH value, protease expression, and the change of external conditions such as the local application of heat, ultrasound, light, magnetic field, or electric field could be served as stimuli (Fig. 3). The environmentally-responsive nanoparticles could improve tumor accumulation, tumor penetration of cancer therapeutics, and increase the intracellular localization of anticancer therapeutics, and thus further enhance the efficacy of antitumor therapeutics (MacEwan et al., 2010). Among these environmentally-responsive nanoparticles, pHresponsive nanoparticulate DDSs have been widely studied in the field of cancer therapy. A change in pH can be used in two ways to trigger drug release. First, the extracellular pH values of most human tumors (pHe) are found to be at an average value of ~ 7.0, which is a distinguishing phenotype of solid tumor to normal tissue (Vaupel, 2004). A second is used after cellular uptake when nanoparticles reached the endosomal and lysosomal compartments with low pH of about 5–6 (Murphy et al., 1984). Nanoparticles could be formulated with pH-responsive polymers that could change their physical and chemical properties in response to different pH values for pH-dependent drug release. One strategy is to take advantage of the changes in polymer protonation states to make pHdependent hydrophobic-to-hydrophilic transitions to affect polymer swelling or solubility, which will then acquire pH-responsive drug release. Expansile nanoparticles formulated by acrylate-based hydrophobic polymers modified with pH-labile protecting groups were stable at neutral pH, but in a mildly acidic pH, the protecting group was cleaved to reveal hydroxyls, which corresponded to a hydrophobic-to-hydrophilic transformation, resulting in swelling of the polymeric structure to create a hydrogel and subsequent drug release. The expanded nanoparticles can act as an intracellular or intratumoral depot for the chemotherapeutic agent PTX, affording higher intracellular drug concentrations (Zubris et al., 2012). PTX loaded expanded nanoparticles showed superior in vivo efficacy in murine tumor models including non-small cell lung cancer (Griset et al., 2009) and mesothelioma (Colson et al., 2011). PEGpoly(β-amino ester) polymers have also been used to design for pH-responsive nanoparticles targeting the low pH present in the

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

9

Fig. 3. Schematic illustration of environmental response drug delivery systems for cancer therapy. The low extracellular pH or up-regulated protease expression at the tumor environment, and the local application of heat, electric field, magnetic field, ultrasound or light could be utilized to trigger drug release from nanoparticulate drug delivery systems.

extratumoral and cellular microenvironment. The polymers are designed to incorporate some amines that are unprotonated at pH 7.4 to make the polymer insoluble in water, but are protonated at pH 6.4–6.8 to increase polymer solubility and induce a sharp micellization–demicellization transition for drug release (Shenoy et al., 2005; X.L. Wu et al., 2010). The micellization–demicellization transition was also found in DDSs formulated by combination of poly(L-histidine)-bPEG and PLLA-b-PEG (Lee et al., 2003). The low pH tumor microenvironment could also serve as stimulus to trigger a change in surface charge of nanoparticle which could facilitate uptake of nanoparticles by tumor cells. Poon et al. designed the trilayer nanoparticles which consist of iminobiotin modified PLL, the linker protein neutravidin, and biotin end-functionalized PEG layer. The PEG layer in the nanoparticles could selectively deshield when localized in low pH tumor microenvironment to expose positive charged nanoparticles for improving cellular uptake by tumor cells and decreasing non-specific cellular uptake by normal cells (Poon et al., 2011). Another work based on a charge-reversal nanogel triggered by the pHe was reported by Du et al. At physiological pH values, the nanogel was negatively charged with high positively charged drug loading capacity and was relatively inert to tumor cells. When exposed to the tumor microenvironment, the nanogel was turned to be positively charged to strengthen nanogel-cell interaction and enhance cellular uptake by cancer cells (Du et al., 2010). An alternate pH-triggered strategy involves the use of acid-labile linkers upon cleavage at a specific pH. Commonly used acid-labile crosslinkers include ester, hydrazone, carboxy dimethylmaleic anhydride,

orthoester, imine, vinylether, phosphoramidate, and so on (Gao et al., 2010). The prodrug strategy often conjugates drug molecules to macromolecular chains via pH-labile linkers. In response to the acidic extracellular or intracellular environment, these macromolecular carriers are able to release the drug to exert its efficacy (Ulbrich and Subr, 2004). In an example of this approach, cisplatin was conjugated to poly(ethylene glycol)-b-poly(L-lactide) using hydrazone cross-linkers to allow release of the drug after the nanoparticles were endocytosed by the target cells (Aryal et al., 2010). Likewise, a hydrazone linker was used to link the doxorubicinyl group and the PEG chain to the surface of the gold nanoparticles, which released DOX in acidic organelles after endocytosis (F. Wang et al., 2011). Acid-labile linkers were also used to synthesize acid-labile polymers for pH triggered rapid degradation to release the entrapped drugs. One example is the use of nanoparticles composed of a copolymer (poly-β-aminoester ketal-2) for efficient gene delivery. The nanoparticles can respond to endosomal pH and undergo a hydrophobic-hydrophilic switch and a rapid degradation “in series”, resulting in increased cytoplasmic delivery (Morachis et al., 2012). As many chemotherapeutics and some macromolecules such as DNA, exert their action in the nucleus, the pH-labile linkers have been used to develop charge-reversal polymers for the nuclear-targeted drug delivery. Modification of the positive amines in the cationic polymers such as PLL (Zhou et al., 2009) and PAMAM dendrimers (Shen et al., 2010) with negatively charged groups with acid-labile amides so that the polymer was negatively charged at physiological condition, but was hydrolyzed at low pH to restore its positive charge, can obtain efficient nuclear-targeted gene delivery. For favoring gene expression, pH labile linker was also employed to construct a

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

10

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

smart nanoassembly which has the ability for self-dePEGylation. The nanoassembly containing lipid envelope based on PEG-vinyl etherDOPE can self-remove PEG and recover the fusogenic ability of DOPE at low pH, resulting a higher transfection efficiency and much lower cytotoxicity than that of commercial Lipofectamine 2000 in several cancer cell lines (Xu et al., 2008). Over-expressed cancer-associated enzymes are also utilized to trigger drug release. Many kinds of nanomaterials such as liposomes, polymer-based nanoparticles, mesoporous silica nanoparticles and gold nanoparticles have been designed with high specificity for the enzyme stimulus (Andresen et al., 2010; Ghadiali and Stevens, 2008; Ulijn, 2006). The general strategy to design enzyme-responsive nanoparticle systems is to employ biological motifs that can be degraded by enzymes. By this approach, the self-assembled polymers could be synthesized with enzyme-responsive motifs to the encapsulated drug or the conjugated drug using enzyme-responsive linkers. When these nanomaterials encounter the enzyme, they could degrade to release the encapsulated drug or conjugated drug. In other cases, the nanoparticles can be designed to generate a change in the physical properties upon enzymatic stimulation, which could facilitate cellular uptake or intracellular delivery of nanoparticles. Matrix metalloproteinases (MMPs) especially MMP-2 and MMP-9, have been regarded to be up-regulated in the tumor microenvironment (Roy et al., 2009). Banerjee et al. had integrated an MMP-cleavable lipopeptide into the liposome formulation to prepare MMP-sensitive liposomes which can rapidly release their contents by MMP-9 (Banerjee et al., 2009). In another work, PEG-mesoporous silica nanoparticles (MSNPs) to respond to MMP for controlled drug delivery were engineered. The proteases present at the tumor site could trigger DOX release from the MSNPs, resulting in significant cellular apoptosis (N. Singh et al., 2011). Basel et al. incorporated peptides with a sequence that can be recognized by cancerassociated proteases into a polymer-stabilized liposome formulation. The liposomes were much stable and resistant to osmotic swelling, but released their payloads quickly in response to the protease presented at the tumor site (Basel et al., 2011). As abnormally high concentrations of phospholipase A2 (PLA2) have been found in the evading zone of tumors (Yamashita et al., 1993), several liposome systems consisting prodrugs of antitumor ether lipids (proAELs) were investigated for PLA2-triggered degradation, resulting in the release of antitumor ether lipids (AELs). As the AELs possess the ability to enhance transmembrane drug diffusion, encapsulation of the conventional chemotherapeutic drugs in liposomes containing proAELs can enhance the intracellular distribution of drug in cancer cells (Andresen et al., 2004). Sugar-based nanocarrier has also been designed to release anti-cancer drugs selectively to tumors. Bernardos et al. used saccharide derivatives which could respond to the lyzosomal amylase to modify the pore of the MSNPs for specific release of drugs by enzymatic stimulus. After the internalization of the nanoparticle, the saccharide molecular gate opened in the presence of the lyzosomal amylase, and the cytotoxic agent consequently released to attain a decreased cell viability (Bernardos et al., 2010). Differences in the reducing potential between the extra- and intracellular environment provide another stimulus for drug delivery. The disulfide linkage has been extensively employed to design redox sensitive nanocarriers for drug delivery due to their stability in the typically oxidizing extracellular environment and their lability in the elevated reducing intracellular environment (high glutathione concentration) to efficiently release entrapped cargos (Saito et al., 2003). Gao et al. synthesized a linear cationic click polymer containing disulfide bonds via the “click chemistry”. The polymer could facilitate efficient gene delivery through releasing DNA efficiently by the cleavage of disulfide bonds under the reduction condition (Gao et al., 2011). The disulfide bond was also used to link poly(epsilon–caprolactone) and poly(ethyl ethylene phosphate) to synthesize diblock copolymer. The micelles formed by the diblock copolymer could load drug in its inner core in aqueous solution, while release drug rapidly under glutathione

stimulus, leading to enhanced drug toxicity to tumor cells (Tang et al., 2009). The disulfide linkage was also applied to design micelles with the capability to self-remove PEG chain under the intracellular reducing environment. The successful detachment of PEG in endosome is beneficial for endosomal escape of micelles and enhanced gene transfection efficiency (Takae et al., 2008). An extrinsic stimuli can also be utilized to enhance drug distribution at the tumor site or improve intracellular drug accumulation by selectively release of its payload at the tumor tissue or cancer cells. Thermosensitive polymers that exhibit a volume phase transition in response to temperature have been developed for triggering drug release and local accumulation by application of heat. Ultrasound and electromagnetic (EM) fields have been employed as external stimuli to produce heat (O'Neill and Rapoport, 2011). Polymers based on N-isopropylacrylamide (NIPAm), N, N-diethylacrylamide and N-vinylcaprolactam monomers have a lower critical solution temperature (LCST) and polymers based on a combination of acrylamide and acrylic acid monomers showed a upper critical solution temperature (UCST) (Schmaljohann, 2006). Thermoresponsive chitosan-g-poly (N-vinylcaprolactam) polymers loaded with curcumin can specifically kill cancer cells at above their LCST (Rejinold et al., 2011). Hoare et al. produced thermosensitive nanoparticles by wrapping superparamagnetic iron oxide nanoparticles with a film formed by PNIPAm-based nanogels and ethyl cellulose. The entrapped drug molecules could transport across the film through heating the superparamagnetic nanoparticles to dissolve of the PNIPAm (Hoare et al., 2011). Inducing a local hyperthermia effect by the magnetic field at the level of the polymersome membrane could achieve triggered drug release from polymersomes encapsulate DOX together with superparamagnetic iron oxide nanoparticles (USPIO; γ-Fe2O3) (Oliveira et al., 2013). Similarly, ultrasound can be employed to trigger the release of DOX and other hydrophobic drugs from polymeric micelles at definite time and space (Husseini and Pitt, 2008, 2009). Some smart nanoparticles are designed to respond to a mixture of two or more stimuli. A nanoparticle system composed of a conducting polymer (polypyrrole) and a temperature-sensitive hydrogel matrix (PLGA-PEG-PLGA) were developed with response to temperature and electric field dual-stimulus for programmed drug delivery (Ge et al., 2012). Superparamagnetic maghemite (γ-Fe2O3) nanoparticles modified with poly (2-(dimethylamino)ethyl methacrylate) (pDMAEMA) which exhibit a pH- and temperature-dependent reversible agglomeration showed remarkable gene delivery efficiency in CHO-K1 cells (Majewski et al., 2012). A novel PEG and cRGD peptide modified poly(2-(pyridin-2-yldisulfanyl)ethyl acrylate) nanoparticle loaded with DOX (RPDSG/DOX) was designed to be both pH-responsive and redox sensitive. The RPDSG/DOX nanoparticle is stable in physiological condition while releasing DOX fast with the trigger of acidic pH and redox potential (Bahadur et al., 2012). Micelles formed by block copolymer which consisted of an acid-sensitive tetrahydropyran-protected 2hydroxyethyl methacrylate (HEMA) hydrophobic chain and a temperature-sensitive poly(N-isopropylacrylamide) (PNIPAM) hydrophilic chain can respond to triple stimuli including temperature, pH and redox potential (Klaikherd et al., 2009). 3.4. Multi-stage delivery nanovectors Between the point of intravenous administration and the tumor tissue, systemically administered nanoparticles should go through a three-step process: blood delivery to the blood vessels of the tumor, transported across the vessel wall into the interstitium, and migrated through the interstitium to reach cancer cells in tumor (Jain, 1999). To circumvent the multiplicity of biological barriers that the nanoparticles encounter after administration, and maximize drug localization and release in cancer cells, multistage nanovectors have been developed recently. The rationale for the multistage approach is based on the arrangement of different tasks to a single nanoparticle which could complete the tasks at different stages.

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

This concept was first applied to design a “nanocell” which overcomes the barriers unique to solid tumors. The nanocell was composed of a nuclear polymer-based nanoparticle containing a chemotherapy drug and a pegylated-lipid envelope, which entrapped an anti-angiogenesis agent. The anti-angiogenesis agent was first released from the outer envelope causing a vascular shutdown, and the chemotherapy drug was then released from the inner nanoparticle, which is trapped inside the tumor. The multistage release profile within a tumor results in improved therapeutic effects with reduced toxicity (Sengupta et al., 2005). Based on mathematical modeling, mesoporous silicon nanoparticles could be designed to carry the payload trafficking efficiently and controlling the release rate of the burden. Recently, Ferrari's group developed a multistage silicon nanocarrier system which was composed of mesoporous silican particles (also known as the first stage) and the entrapped nanoparticles (the second stage) loaded with anticancer therapeutics (the third stage). The first stage mesoporous silican particles could protect and ferry the inner nanoparticles until they recognize and dock at the tumor vasculature. Then the second stage nanoparticles released from the MSP with the biodegradation of porous multistage particles under physiological conditions. The released nanoparticles were able to extravasate through fenestrations of vessels and enter the tumor parenchyma, thus concentrating diagnostic and therapeutic agents within the target microenvironment (Tasciotti et al., 2008). This kind of multistage nanovectors was also applied to gene delivery. Liposomes consisted of dioleoyl phosphatidylcholine containing siRNA targeted against the EphA2 oncoprotein (the second-stage carriers) were loaded into mesoporous biodegradable silicon particles (the first-stage carriers). The mesoporous silicon particles allowed for the loading and release of second-stage nanocarriers in a sustained manner. Compared with the one-stage neutral nanoliposomes that require twice weekly injections to achieve continuous gene silencing, the multistage delivery methods could achieve sustained EphA2 gene silencing which could last for at least 3 weeks after a single i.v. administration (Tanaka et al., 2010). Similarly, superparamagnetic CaCO3 mesocrystals were used to encapsulate DOX, Au–DNA, and Fe3O4@silica nanoparticles for the co-delivery of drug and gene via a multistage method for treatment of cancer. The stage-one nanoparticles-CaCO3 system protected the encapsulated payloads from degradation and phagocytosis during navigation in the blood. After they docked to the vascular walls, the stage-one nanoparticles degraded to gradually release the stage-two nanoparticles and drugs (Zhao et al., 2010). A multistage system with sizeshrinking property was designed for deep tumor tissue penetration. The 100-nm multistage nanoparticles are composed of a gelatin core with surface covered with 10-nm QDs. After they exposed to the tumor microenvironment, the 100-nm nanoparticles “shrink” to 10nm nanoparticles due to the hydrolysis of gelatin by the MMPs. This shrinkable multistage system can combine the advantage of large 100nm nanoparticles that are suitable for the EPR effect, and small 10-nm nanoparticles that are suitable for diffusion in the collagen matrix of the interstitial space and penetration into the tumor parenchyma (Stylianopoulos et al., 2012; Wong et al., 2011). 3.5. Cancer nanotheranostics Cancer nanotheranostics is the use of nanotechnology for the combined therapeutics and diagnostics for cancer (Sumer and Gao, 2008). Besides for therapeutic purposes, potential applications of nanotheranostics range from the visualizing the blood circulation, biodistribution of drugs in real time, noninvasively assessing drug accumulation and drug release at the target site, facilitating triggered drug release and monitoring drug distribution, to predicting drug responses and evaluating drug efficacy longitudinally (Lammers et al., 2010). Integration of imaging capability into the design of nanoparticles made it possible to evaluate the fate of nanoparticles in real time, which will allow physicians to adjust the type and dosing of drugs for more personalized treatment regimens (Diou et al., 2012). The currently accessible imaging techniques include

11

MRI, single photon emission computed tomography (SPECT), positron emission tomography (PET), computer tomography (CT), and ultrasonography (US) (Mura and Couvreur, 2012). Nanotheranostics can be obtained by either attaching different imaging moieties (i.e. NIR probes, radionuclides) to the available nanocarriers or taking advantage of the intrinsic properties of some nanoparticle materials such as SPIOs for MRI and QDs for fluorescence imaging (Brigger et al., 2002). Till now, a variety of nanotheranostics that combine anti-cancer therapeutics with aforementioned imaging modalities has been developed using liposomes, micelles, polymers, gold-based nanomaterials, magnetic nanomaterials, carbon nanomaterials, and silica-based nanomaterials. From a recent summary of selected papers published between 2009 and 2012 on nanotheranostics, optical and MRI are the preferable modalities performed for imaging functionality, through use of NIR emission and magnetic agents, respectively (Wang et al., 2012). Magnetic nanoparticles have been used as “nanotheranostics” for both targeted drug delivery and tumor imaging due to their magnetic property as nanostructured contrast probes for MRI. Among the magnetic nanoparticles, SPIOs are the most commonly used nanomaterials. A number of polymers, including dextran, dendrimer, polyaniline, and polyvinylpyrrolidone, have been utilized to coat magnetic nanoparticles. Pluronic polymer F127 and β-cyclodextrin (β-CD) were coated onto the iron oxide core nanoparticles by a multi-layer approach for encapsulation of the anti-cancer drugs and for sustained drug release, respectively. The optimized water-dispersible SPIOs formulation showed improved MRI characteristics and improved therapeutic effects (Yallapu et al., 2011). By tuning the properties of the coating polymers, the triggered drug release could be realized in magnetic theranostic nanoparticles. A poly (beta-amino ester) (PBAE) copolymer was used to entrap SPIO and DOX for sensitive detection and effective treatment of cancer by pH sensitive controlled drug release (Fang et al., 2012). The magnetic nanoparticle formulation loaded with other imaging moieties could allow for multimodal imaging. Foy et al. loaded near-infrared dyes into a magnetic nanoparticle (MNP) formulation stabilized by an amphiphilic block copolymer to provide for both tumor MRI and optical imaging (Foy et al., 2010). To achieve targeted delivery, a variety of ligands has been attached to the outside layer of magnetic theranostic nanoparticles such as folate (Santra et al., 2009), cRGD, (Nasongkla et al., 2006) and antibody (Zou et al., 2010). Yang et al. designed an FR-targeted multifunctional polymer vesicle nanocarrier system loaded with SPIO and DOX to increase specific cellular uptake by FR positive cancer cells (Yang et al., 2010). Similarly, a multifunctional antibodyand fluorescence-labeled HuCC49ΔCH2-SPIO “nanotheranostics” was developed for combined targeted anticancer drug delivery and multimodel imaging of cancer cells (Zou et al., 2010). Cyclo(Arg–Gly– Asp–d–Phe–Cys) (c(RGDfC)) peptides were also employed to modify SPIO nanocarriers for targeted drug delivery and dual PET/MRI imaging (Yang et al., 2011). Magnetic theranostic nanoparticles can also be used for gene delivery. An MRI visible gene delivery system developed with a core of SPIO nanocrystals and a shell of biodegradable stearic acidmodified low molecular weight PEI (Stearic-LWPEI) via self-assembly showed synergistic advantages in the effective transfection of mcDNA and non-invasive MRI of gene delivery (Wan et al., 2013). Gold-based nanoparticles have been investigated as nanotheranostics due to their unique optical characteristics known as surface plasmon resonance and photothermal characteristics, which enable them not only to be used for imaging applications but also to induce photothermal effects for therapeutic purposes (Dykman and Khlebtsov, 2012; Saha et al., 2012). The optical and thermal properties of gold materials can be tuned by changing their morphology and surface properties, as spherical gold nanoparticles (AuNP), nanorod (AuNR), nanoshell, and nanocage exhibit different optical and thermal properties. Different kinds of gold-based nanoparticles such as high-photoluminescence-yield gold nanocubes (X. Wu et al., 2010), gold nanorod-in-shell nanostructures (Hu et al., 2009), silica-modified gold nanorods (Huang et al., 2011), and matrix metalloproteinase sensitive gold nanorod (Yi et al., 2010)

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

12

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

have been developed for combined imaging and photothermal therapy of cancer. Due to the well-established strategies for surface modification (ie, gold–thiol bonding) of gold-based nanoparticles, targeting ligands and chemotherapeutic agents could be conjugated to the surface of gold-based nanoparticles to achieve targeted drug delivery and diagnosing. Heo et al. designed a gold nanoparticle modified with PEG, biotin, PTX and rhodamine B linked β-CD which could specifically interact with cancer cells by biotin and effectively release the entrapped PTX under the intracellular glutathione condition (Heo et al., 2012). Another gold nanoparticle modified with a PSMA RNA aptamer was developed for simultaneous CT imaging and prostate cancer therapy. The PSMA aptamer conjugation could significantly improve CT intensity and DOX efficacy of gold nanoparticles for targeted LNCaP cells than that of non-targeted PC3 cells (D. Kim et al., 2010). Silica-based nanoparticles especially MSNs have been qualified as a new type of excellent theranostic nano-platform, due to their unique features, such as high surface area, large pore volume, tunable porosity, and facile functionalization, and many kinds of imaging and therapeutic agents have been encapsulated into MSNs to achieve theranostic purposes (Ambrogio et al., 2011; Lee et al., 2011). By decoration with magnetite nanocrystals, carbon or Si nanocrystals, MSNs could be used for synchronous drug delivery and imaging. He et al. encapsulated carbon and Si nanocrystals into the framework of mesoporous silica nanoparticles (CS-MSNs) to constructed a nanoparticle system with high payload of insoluble drugs and unique NIR-to-Vis luminescence imaging feature (He et al., 2012). Hyeon's group developed a dye-doped mesoporous silica shell containing a single Fe3O4 nanocrystal core as drug delivery system and dual MRI/fluorescence imaging agents (Kim et al., 2008). In a study by Cheng et al., mesoporous silica nanoparticle loading both contrast agent and drug was functionalized with biomolecular ligands cRGDyK peptides for targeted drug delivery to the over-expressed αvβ3 integrins of cancer cells and tumor imaging (Cheng et al., 2010). In addition, MSNs could be combined with gold composites to achieve multimodal imaging and multimodality treatment. Nanoparticles consisted of AuNRs-capped magnetic core and mesoporous silica shell designed by Ma et al could achieve synchronous chemotherapy, photo-thermotherapy, in vivo MR-, infrared thermal and optical imaging into one single system (M. Ma et al., 2012). In addition to the conventional nanomaterials AuNPs, SPIO, MSNs which have the unique properties, nanotheranostics using lipid- and polymer-based formulations have been widely studied in cancer research. They can be loaded with a variety of contrast agents such as SPIOs and gadolinium-based compounds for MRI applications. They can also be loaded or conjugated with radionuclide agents such as 64Cu, 99mTc, and 111In for radionuclide imaging, or loaded with fluorescent molecules or QDs for fluorescence imaging (Luk et al., 2012). Some of the lipid- and polymer-based nanotheranostics can be designed with environment sensitivity by incorporating thermosensitive polymer chains (Kono et al., 2011) or pH-sensitive linkages within polymer backbone (T. Liu et al., 2012). Single walled carbon nanotube (SWNT) is one potential candidate as a theranostic agent since it generates significant amounts of heat upon excitation with near-infrared light for the photothermal destruction of tumors (Kam et al., 2005; Moon et al., 2009). In addition, due to its strong optical absorbance, it has been used as contrast agents for photoacoustic imaging (De la Zerda et al., 2008). Mashal et al. reported that at frequency of 3 GHz, SWNTs could be used for both microwave-induced thermoacoustic imaging and hyperthermia treatment (Mashal et al., 2010). Other nanomaterials such as upper-conversion nanoparticles (UCNPs) (Cheng et al., 2011), QDs (Mitra et al., 2012), and silver nanoparticles (W. Wu et al., 2010) were also employed for theranostic applications. Some nanotheranostics can be designed by mixing drug loaded nanoparticulate systems with perfluoropentane (PFP) nano/microbubbles to facilitate ultrasound-mediated cellular uptake and ultrasonic tumor imaging (Rapoport et al., 2007).

4. Conclusions and outlook Since its advent in the field of cancer, nanotechnology has shown immense potential for cancer detection, prevention, and treatment. Nanoparticles have revolutionalized drug delivery, allowing for therapeutic agents to selectively targeting tumor tissue and cancer cells, while minimizing toxicity to normal cells. Liposome and protein based nanomedicine formulation are already approved for commercial use and many new formulations are in the phase II and phase III stages of evaluation for cancer therapy. In addition to drug delivery, some kinds of nanoparticles are promising materials for physical therapy methods such as hyperthermia to provide a non-invasive approach to treat cancer. Furthermore, some delicately designed nanoparticles can combine imaging and therapy of cancer together with the ability to monitor treatment process in real-time. Till now, increasingly sophisticated nanoparticles have been developed to combat cancer on the molecular scale through careful engineering of nanomedicines. Looking into the future, some considerations or directions require a concerted effort for success for a pharmaceutical scientist in developing nanoparticulate cancer therapeutics. The first direction is the optimal design of nanoparticles to be disease-specific. Because one tumor may be different from another, the primary tumor is different from its metastasis, and even the same tumor can change from one day to the next, the nanoparticles should be delicately designed on a case-by-case basis for personalized treatment. This is indeed a formidable task to choose specific nanocarrier or combinations which could lead to improved therapeutic outcomes considering the highly heterogeneous and continuously evolving nature of the tumor microenvironment. The second direction is the design of nanoparticles from the practical pharmaceutical point of view. The colloidal properties and the stability of nanoparticles, the time and cost on preparation of a formulation, whether the materials used in the formulation could receive approval from regulatory authorities, whether the preparation methods could be transformed into industrial processes etc. should be all taken into consideration when beginning to design the nanoparticles. The third direction is to address toxicology concerns of nanoparticles. Although many kinds of nanoparticles were employed as TDDS and showed great potential for cancer treatment, the safety issue of those nanoparticles is scarcely addressed. The fourth direction is the introduction of new interdisciplinary sciences such as computer science, analytical science, advanced instrumental techniques to develop suitable screening methodologies for determining optimal characteristics of nanocarriers, and to study the mechanism of action and the fate of nanoparticles in vivo, thus to design more intelligent TDDS. The last direction is to emerge new concepts from interdisciplinary collaborations to design nanoparticles with multiple functions, more than those functions mentioned above. Declaration of interest The authors do not have any conflicts of interest to declare. Acknowledgment This work was supported in part by the Fuzhou University Start-up Fund No. 033084, Chinese National Science Foundation Grant (Nos. 81201709, 81102388 and 81273548), the National Science Foundation for Fostering Talents in Basic Research of China (No. J1103303), and the Science and Technology Development Foundation of Fuzhou University (2013-XQ-8 and 2013-XQ-9). Reference Agasti SS, Chompoosor A, You CC, Ghosh P, Kim CK, Rotello VM. Photoregulated release of caged anticancer drugs from gold nanoparticles. J Am Chem Soc 2009;131:5728–9. Al-Dosari MS, Gao X. Nonviral gene delivery: principle, limitations, and recent progress. AAPS J 2009;11:671–81.

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx Ambrogio MW, Thomas CR, Zhao YL, Zink JI, Stoddart JF. Mechanized silica nanoparticles: a new frontier in theranostic nanomedicine. Acc Chem Res 2011;44:903–13. Andresen TL, Davidsen J, Begtrup M, Mouritsen OG, Jorgensen K. Enzymatic release of antitumor ether lipids by specific phospholipase A2 activation of liposome-forming prodrugs. J Med Chem 2004;47:1694–703. Andresen TL, Jensen SS, Jorgensen K. Advanced strategies in liposomal cancer therapy: problems and prospects of active and tumor specific drug release. Prog Lipid Res 2005;44:68–97. Andresen TL, Thompson DH, Kaasgaard T. Enzyme-triggered nanomedicine: drug release strategies in cancer therapy. Mol Membr Biol 2010;27:353–63. Ansell SM, Harasym TO, Tardi PG, Buchkowsky SS, Bally MB, Cullis PR. Antibody conjugation methods for active targeting of liposomes. Methods Mol Med 2000;25: 51–68. Antonietti M, Förster S. Vesicles and liposomes: a self‐assembly principle beyond lipids. Adv Mater 2003;15:1323–33. Arap W, Pasqualini R, Ruoslahti E. Cancer treatment by targeted drug delivery to tumor vasculature in a mouse model. Science 1998;279:377–80. Aryal S, Hu CM, Zhang L. Polymer–cisplatin conjugate nanoparticles for acid-responsive drug delivery. ACS Nano 2010;4:251–8. Ashley CE, Carnes EC, Phillips GK, Padilla D, Durfee PN, Brown PA, et al. The targeted delivery of multicomponent cargos to cancer cells by nanoporous particle-supported lipid bilayers. Nat Mater 2011;10:389–97. Bahadur R, Thapa B, Xu P. pH and redox dual responsive nanoparticle for nuclear targeted drug delivery. Mol Pharm 2012;9:2719–29. Banerjee J, Hanson AJ, Gadam B, Elegbede AI, Tobwala S, Ganguly B, et al. Release of liposomal contents by cell-secreted matrix metalloproteinase-9. Bioconjug Chem 2009;20: 1332–9. Barenholz Y. Doxil(R)—the first FDA-approved nano-drug: lessons learned. J Control Release 2012;160:117–34. Basel MT, Shrestha TB, Troyer DL, Bossmann SH. Protease-sensitive, polymer-caged liposomes: a method for making highly targeted liposomes using triggered release. ACS Nano 2011;5:2162–75. Batist G, Gelmon KA, Chi KN, Miller Jr WH, Chia SK, Mayer LD, et al. Safety, pharmacokinetics, and efficacy of CPX-1 liposome injection in patients with advanced solid tumors. Clin Cancer Res 2009;15:692–700. Bernardos A, Mondragon L, Aznar E, Marcos MD, Martinez-Manez R, Sancenon F, et al. Enzyme-responsive intracellular controlled release using nanometric silica mesoporous supports capped with “saccharides”. ACS Nano 2010;4:6353–68. Bhattacharyya S, Singh RD, Pagano R, Robertson JD, Bhattacharya R, Mukherjee P. Switching the targeting pathways of a therapeutic antibody by nanodesign. Angew Chem Int Ed Engl 2012;51:1563–7. Bikram M, Gobin AM, Whitmire RE, West JL. Temperature-sensitive hydrogels with SiO2–Au nanoshells for controlled drug delivery. J Control Release 2007;123: 219–27. Biswas S, Dodwadkar NS, Sawant RR, Koshkaryev A, Torchilin VP. Surface modification of liposomes with rhodamine-123-conjugated polymer results in enhanced mitochondrial targeting. J Drug Target 2011;19:552–61. Boddapati SV, D'Souza GG, Erdogan S, Torchilin VP, Weissig V. Organelle-targeted nanocarriers: specific delivery of liposomal ceramide to mitochondria enhances its cytotoxicity in vitro and in vivo. Nano Lett 2008;8:2559–63. Bonomi P. Paclitaxel poliglumex (PPX, CT-2103): macromolecular medicine for advanced non-small-cell lung cancer. Expert Rev Anticancer Ther 2007;7:415–22. Bria E, Giannarelli D, Felici A, Peters WP, Nistico C, Vanni B, et al. Taxanes with anthracyclines as first-line chemotherapy for metastatic breast carcinoma. Cancer 2005;103:672–9. Brigger I, Dubernet C, Couvreur P. Nanoparticles in cancer therapy and diagnosis. Adv Drug Deliv Rev 2002;54:631–51. Bu G, Maksymovitch EA, Nerbonne JM, Schwartz AL. Expression and function of the low density lipoprotein receptor-related protein (LRP) in mammalian central neurons. J Biol Chem 1994;269:18521–8. Burleigh M, Dai S, Hagaman E, Barnes C, Xue Z. Stepwise assembly of surface imprint sites on MCM-41 for selective metal ion separations. ACS Symposium Series: ACS Publications; 2001. p. 146–58. Cai W, Gao T, Hong H, Sun J. Applications of gold nanoparticles in cancer nanotechnology. Nanotechnol Sci Appl 2008;1:17–32. Chang DK, Chiu CY, Kuo SY, Lin WC, Lo A, Wang YP, et al. Antiangiogenic targeting liposomes increase therapeutic efficacy for solid tumors. J Biol Chem 2009;284: 12905–16. Chapman AP. PEGylated antibodies and antibody fragments for improved therapy: a review. Adv Drug Deliv Rev 2002;54:531–45. Chen H-T, Huh S, Lin VS-Y. A fine-tuning the functionalization of mesoporous silica. Prep Catal Sci Eng 2006;45. Chen L, Remondetto G, Rouabhia M, Subirade M. Kinetics of the breakdown of cross-linked soy protein films for drug delivery. Biomaterials 2008;29:3750–6. Chen AM, Zhang M, Wei D, Stueber D, Taratula O, Minko T, et al. Co-delivery of doxorubicin and Bcl-2 siRNA by mesoporous silica nanoparticles enhances the efficacy of chemotherapy in multidrug-resistant cancer cells. Small 2009a;5:2673–7. Chen H, Ahn R, Van den Bossche J, Thompson DH, O'Halloran TV. Folate-mediated intracellular drug delivery increases the anticancer efficacy of nanoparticulate formulation of arsenic trioxide. Mol Cancer Ther 2009b;8:1955–63. Cheng WW, Allen TM. Targeted delivery of anti-CD19 liposomal doxorubicin in B-cell lymphoma: a comparison of whole monoclonal antibody, Fab' fragments and single chain Fv. J Control Release 2008;126:50–8. Cheng S-H, Lee C-H, Chen M-C, Souris JS, Tseng F-G, Yang C-S, et al. Tri-functionalization of mesoporous silica nanoparticles for comprehensive cancer theranostics—the trio of imaging, targeting and therapy. J Mater Chem 2010;20:6149–57.

13

Cheng L, Yang K, Li Y, Chen J, Wang C, Shao M, et al. Facile preparation of multifunctional upconversion nanoprobes for multimodal imaging and dual-targeted photothermal therapy. Angew Chem Int Ed Engl 2011;50:7385–90. Cherukuri P, Bachilo SM, Litovsky SH, Weisman RB. Near-infrared fluorescence microscopy of single-walled carbon nanotubes in phagocytic cells. J Am Chem Soc 2004;126: 15638–9. Chithrani BD, Ghazani AA, Chan WC. Determining the size and shape dependence of gold nanoparticle uptake into mammalian cells. Nano Lett 2006;6:662–8. Choi CH, Alabi CA, Webster P, Davis ME. Mechanism of active targeting in solid tumors with transferrin-containing gold nanoparticles. Proc Natl Acad Sci U S A 2010;107:1235–40. Clementi C, Miller K, Mero A, Satchi-Fainaro R, Pasut G. Dendritic poly(ethylene glycol) bearing paclitaxel and alendronate for targeting bone neoplasms. Mol Pharm 2011;8:1063–72. Colson YL, Liu R, Southard EB, Schulz MD, Wade JE, Griset AP, et al. The performance of expansile nanoparticles in a murine model of peritoneal carcinomatosis. Biomaterials 2011;32:832–40. Davis ME, Chen ZG, Shin DM. Nanoparticle therapeutics: an emerging treatment modality for cancer. Nat Rev Drug Discov 2008;7:771–82. Davis ME, Zuckerman JE, Choi CH, Seligson D, Tolcher A, Alabi CA, et al. Evidence of RNAi in humans from systemically administered siRNA via targeted nanoparticles. Nature 2010;464:1067–70. De la Zerda A, Zavaleta C, Keren S, Vaithilingam S, Bodapati S, Liu Z, et al. Carbon nanotubes as photoacoustic molecular imaging agents in living mice. Nat Nanotechnol 2008;3:557–62. Devaraj NK, Upadhyay R, Haun JB, Hilderbrand SA, Weissleder R. Fast and sensitive pretargeted labeling of cancer cells through a tetrazine/trans‐cyclooctene cycloaddition. Angew Chem Int Ed 2009;48:7013–6. Diou O, Tsapis N, Fattal E. Targeted nanotheranostics for personalized cancer therapy. Expert Opin Drug Deliv 2012;9:1475–87. Du J, Lu WL, Ying X, Liu Y, Du P, Tian W, et al. Dual-targeting topotecan liposomes modified with tamoxifen and wheat germ agglutinin significantly improve drug transport across the blood–brain barrier and survival of brain tumor-bearing animals. Mol Pharm 2009;6:905–17. Du JZ, Sun TM, Song WJ, Wu J, Wang J. A tumor-acidity-activated charge-conversional nanogel as an intelligent vehicle for promoted tumoral-cell uptake and drug delivery. Angew Chem Int Ed Engl 2010;49:3621–6. Dykman L, Khlebtsov N. Gold nanoparticles in biomedical applications: recent advances and perspectives. Chem Soc Rev 2012;41:2256–82. Ellington AD, Szostak JW. In vitro selection of RNA molecules that bind specific ligands. Nature 1990;346:818–22. Estevez MC, Huang YF, Kang H, O'Donoghue MB, Bamrungsap S, Yan J, et al. Nanoparticle-aptamer conjugates for cancer cell targeting and detection. Methods Mol Biol 2010;624:235–48. Fang C, Kievit FM, Veiseh O, Stephen ZR, Wang T, Lee D, et al. Fabrication of magnetic nanoparticles with controllable drug loading and release through a simple assembly approach. J Control Release 2012;162:233–41. Farokhzad OC, Cheng J, Teply BA, Sherifi I, Jon S, Kantoff PW, et al. Targeted nanoparticle-aptamer bioconjugates for cancer chemotherapy in vivo. Proc Natl Acad Sci U S A 2006;103:6315–20. Fay F, Scott CJ. Antibody-targeted nanoparticles for cancer therapy. Immunotherapy 2011;3:381–94. Feazell RP, Nakayama-Ratchford N, Dai H, Lippard SJ. Soluble single-walled carbon nanotubes as longboat delivery systems for platinum(IV) anticancer drug design. J Am Chem Soc 2007;129:8438–9. Feldman EJ, Lancet JE, Kolitz JE, Ritchie EK, Roboz GJ, List AF, et al. First-in-man study of CPX-351: a liposomal carrier containing cytarabine and daunorubicin in a fixed 5:1 molar ratio for the treatment of relapsed and refractory acute myeloid leukemia. J Clin Oncol 2011;29:979–85. Ferlay J, Shin HR, Bray F, Forman D, Mathers C, Parkin DM. Estimates of worldwide burden of cancer in 2008: GLOBOCAN 2008. Int J Cancer 2010;127:2893–917. Firer MA, Gellerman G. Targeted drug delivery for cancer therapy: the other side of antibodies. J Hematol Oncol 2012;5:70. Flenniken ML, Liepold LO, Crowley BE, Willits DA, Young MJ, Douglas T. Selective attachment and release of a chemotherapeutic agent from the interior of a protein cage architecture. Chem Commun (Camb) 2005:447–9. Flenniken ML, Willits DA, Harmsen AL, Liepold LO, Harmsen AG, Young MJ, et al. Melanoma and lymphocyte cell-specific targeting incorporated into a heat shock protein cage architecture. Chem Biol 2006;13:161–70. Foy SP, Manthe RL, Foy ST, Dimitrijevic S, Krishnamurthy N, Labhasetwar V. Optical imaging and magnetic field targeting of magnetic nanoparticles in tumors. ACS Nano 2010;4:5217–24. Gabizon AA, Shmeeda H, Zalipsky S. Pros and cons of the liposome platform in cancer drug targeting. J Liposome Res 2006;16:175–83. Gao L, Nie L, Wang T, Qin Y, Guo Z, Yang D, et al. Carbon nanotube delivery of the GFP gene into mammalian cells. Chembiochem 2006;7:239–42. Gao Y, Gu W, Chen L, Xu Z, Li Y. A multifunctional nano device as non-viral vector for gene delivery: in vitro characteristics and transfection. J Control Release 2007;118:381–8. Gao W, Chan JM, Farokhzad OC. pH-Responsive nanoparticles for drug delivery. Mol Pharm 2010;7:1913–20. Gao Y, Chen L, Zhang Z, Chen Y, Li Y. Reversal of multidrug resistance by reductionsensitive linear cationic click polymer/iMDR1-pDNA complex nanoparticles. Biomaterials 2011;32:1738–47. Ge J, Neofytou E, Cahill III TJ, Beygui RE, Zare RN. Drug release from electricfield-responsive nanoparticles. ACS Nano 2012;6:227–33. Ghadiali JE, Stevens MM. Enzyme‐responsive nanoparticle systems. Adv Mater 2008;20: 4359–63.

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

14

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

Ghoroghchian PP, Frail PR, Susumu K, Blessington D, Brannan AK, Bates FS, et al. Near-infrared-emissive polymersomes: self-assembled soft matter for in vivo optical imaging. Proc Natl Acad Sci U S A 2005;102:2922–7. Ghoroghchian PP, Li G, Levine DH, Davis KP, Bates FS, Hammer DA, et al. Bioresorbable vesicles formed through spontaneous self-assembly of amphiphilic poly(ethylene oxide)-block-polycaprolactone. Macromolecules 2006;39:1673–5. Goodwin PJ, Thompson AM, Stambolic V. Diabetes, metformin, and breast cancer: lilac time? J Clin Oncol 2012;30:2812–4. Griset AP, Walpole J, Liu R, Gaffey A, Colson YL, Grinstaff MW. Expansile nanoparticles: synthesis, characterization, and in vivo efficacy of an acid-responsive polymeric drug delivery system. J Am Chem Soc 2009;131:2469–71. Gunasekaran S, Xiao L, Ould Eleya M. Whey protein concentrate hydrogels as bioactive carriers. J Appl Polym Sci 2006;99:2470–6. Gunn J, Park SI, Veiseh O, Press OW, Zhang M. A pretargeted nanoparticle system for tumor cell labeling. Mol Biosyst 2011;7:742–8. Guo S, Huang Y, Jiang Q, Sun Y, Deng L, Liang Z, et al. Enhanced gene delivery and siRNA silencing by gold nanoparticles coated with charge-reversal polyelectrolyte. ACS Nano 2010;4:5505–11. Haun JB, Devaraj NK, Hilderbrand SA, Lee H, Weissleder R. Bioorthogonal chemistry amplifies nanoparticle binding and enhances the sensitivity of cell detection. Nat Nanotechnol 2010;5:660–5. He H, Li Y, Jia XR, Du J, Ying X, Lu WL, et al. PEGylated Poly(amidoamine) dendrimer-based dual-targeting carrier for treating brain tumors. Biomaterials 2011;32:478–87. He Q, Ma M, Wei C, Shi J. Mesoporous carbon@silicon–silica nanotheranostics for synchronous delivery of insoluble drugs and luminescence imaging. Biomaterials 2012;33:4392–402. Heo DN, Yang DH, Moon HJ, Lee JB, Bae MS, Lee SC, et al. Gold nanoparticles surface-functionalized with paclitaxel drug and biotin receptor as theranostic agents for cancer therapy. Biomaterials 2012;33:856–66. Hoare T, Timko BP, Santamaria J, Goya GF, Irusta S, Lau S, et al. Magnetically triggered nanocomposite membranes: a versatile platform for triggered drug release. Nano Lett 2011;11:1395–400. Hong R, Han G, Fernandez JM, Kim BJ, Forbes NS, Rotello VM. Glutathione-mediated delivery and release using monolayer protected nanoparticle carriers. J Am Chem Soc 2006;128:1078–9. Hood JD, Bednarski M, Frausto R, Guccione S, Reisfeld RA, Xiang R, et al. Tumor regression by targeted gene delivery to the neovasculature. Science 2002;296:2404–7. Hu KW, Liu TM, Chung KY, Huang KS, Hsieh CT, Sun CK, et al. Efficient near-IR hyperthermia and intense nonlinear optical imaging contrast on the gold nanorod-in-shell nanostructures. J Am Chem Soc 2009;131:14186–7. Huang X, Jain PK, El-Sayed IH, El-Sayed MA. Plasmonic photothermal therapy (PPTT) using gold nanoparticles. Lasers Med Sci 2008;23:217–28. Huang P, Bao L, Zhang C, Lin J, Luo T, Yang D, et al. Folic acid-conjugated silica-modified gold nanorods for X-ray/CT imaging-guided dual-mode radiation and photothermal therapy. Biomaterials 2011;32:9796–809. Huang Q, Bao C, Lin Y, Chen J, Liu Z, Zhu L. Disulfide-phenylazide: a reductively cleavable photoreactive linker for facile modification of nanoparticle surfaces. J Mater Chem B 2013;1:1125–32. Husseini GA, Pitt WG. The use of ultrasound and micelles in cancer treatment. J Nanosci Nanotechnol 2008;8:2205–15. Husseini GA, Pitt WG. Ultrasonic-activated micellar drug delivery for cancer treatment. J Pharm Sci 2009;98:795–811. Iqbal MA, Md S, Sahni JK, Baboota S, Dang S, Ali J. Nanostructured lipid carriers system: recent advances in drug delivery. J Drug Target 2012;20:813–30. Ito A, Shinkai M, Honda H, Kobayashi T. Medical application of functionalized magnetic nanoparticles. J Biosci Bioeng 2005;100:1–11. Jain RK. Transport of molecules, particles, and cells in solid tumors. Annu Rev Biomed Eng 1999;1:241–63. Jain KK. The role of nanobiotechnology in drug discovery. Adv Exp Med Biol 2009;655: 37–43. Jain RK, Stylianopoulos T. Delivering nanomedicine to solid tumors. Nat Rev Clin Oncol 2010;7:653–64. Jemal A, Siegel R, Xu J, Ward E. Cancer statistics, 2010. CA Cancer J Clin 2010;60: 277–300. Jia L. Nanoparticle formulation increases oral bioavailability of poorly soluble drugs: approaches experimental evidences and theory. Curr Nanosci 2005;1:237. Jia X, Jia L. Nanoparticles improve biological functions of phthalocyanine photosensitizers used for photodynamic therapy. Curr Drug Metab 2012;13:1119–22. Jia L, Garza M, Wong H, Reimer D, Redelmeier T, Camden JB, et al. Pharmacokinetic comparison of intravenous carbendazim and remote loaded carbendazim liposomes in nude mice. J Pharm Biomed Anal 2002;28:65–72. Kam NW, O'Connell M, Wisdom JA, Dai H. Carbon nanotubes as multifunctional biological transporters and near-infrared agents for selective cancer cell destruction. Proc Natl Acad Sci U S A 2005;102:11600–5. Kaneshiro TL, Lu ZR. Targeted intracellular codelivery of chemotherapeutics and nucleic acid with a well-defined dendrimer-based nanoglobular carrier. Biomaterials 2009;30:5660–6. Kanwar JR, Roy K, Kanwar RK. Chimeric aptamers in cancer cell-targeted drug delivery. Crit Rev Biochem Mol Biol 2011;46:459–77. Khalil IA, Kogure K, Akita H, Harashima H. Uptake pathways and subsequent intracellular trafficking in nonviral gene delivery. Pharmacol Rev 2006;58:32–45. Khemtong C, Kessinger CW, Gao J. Polymeric nanomedicine for cancer MR imaging and drug delivery. Chem Commun 2009:3497–510. Kievit FM, Veiseh O, Fang C, Bhattarai N, Lee D, Ellenbogen RG, et al. Chlorotoxin labeled magnetic nanovectors for targeted gene delivery to glioma. ACS Nano 2010;4:4587–94.

Kievit FM, Wang FY, Fang C, Mok H, Wang K, Silber JR, et al. Doxorubicin loaded iron oxide nanoparticles overcome multidrug resistance in cancer in vitro. J Control Release 2011;152:76–83. Kim J, Kim HS, Lee N, Kim T, Kim H, Yu T, et al. Multifunctional uniform nanoparticles composed of a magnetite nanocrystal core and a mesoporous silica shell for magnetic resonance and fluorescence imaging and for drug delivery. Angew Chem Int Ed Engl 2008;47:8438–41. Kim CK, Ghosh P, Rotello VM. Multimodal drug delivery using gold nanoparticles. Nanoscale 2009;1:61–7. Kim D, Jeong YY, Jon S. A drug-loaded aptamer-gold nanoparticle bioconjugate for combined CT imaging and therapy of prostate cancer. ACS Nano 2010a;4:3689–96. Kim E, Jung Y, Choi H, Yang J, Suh JS, Huh YM, et al. Prostate cancer cell death produced by the co-delivery of Bcl-xL shRNA and doxorubicin using an aptamer-conjugated polyplex. Biomaterials 2010b;31:4592–9. Kim C, Shah BP, Subramaniam P, Lee KB. Synergistic induction of apoptosis in brain cancer cells by targeted codelivery of siRNA and anticancer drugs. Mol Pharm 2011;8: 1955–61. Kirpotin DB, Drummond DC, Shao Y, Shalaby MR, Hong K, Nielsen UB, et al. Antibody targeting of long-circulating lipidic nanoparticles does not increase tumor localization but does increase internalization in animal models. Cancer Res 2006;66: 6732–40. Klaikherd A, Nagamani C, Thayumanavan S. Multi-stimuli sensitive amphiphilic block copolymer assemblies. J Am Chem Soc 2009;131:4830–8. Knisely JM, Lee J, Bu G. Measurement of receptor endocytosis and recycling. Methods Mol Biol 2008;457:319–32. Kolishetti N, Dhar S, Valencia PM, Lin LQ, Karnik R, Lippard SJ, et al. Engineering of self-assembled nanoparticle platform for precisely controlled combination drug therapy. Proc Natl Acad Sci U S A 2010;107:17939–44. Kono K, Nakashima S, Kokuryo D, Aoki I, Shimomoto H, Aoshima S, et al. Multi-functional liposomes having temperature-triggered release and magnetic resonance imaging for tumor-specific chemotherapy. Biomaterials 2011;32:1387–95. Kortesuo P, Ahola M, Karlsson S, Kangasniemi I, Yli-Urpo A, Kiesvaara J. Silica xerogel as an implantable carrier for controlled drug delivery—evaluation of drug distribution and tissue effects after implantation. Biomaterials 2000;21:193–8. Krakovicova H, Etrych T, Ulbrich K. HPMA-based polymer conjugates with drug combination. Eur J Pharm Sci 2009;37:405–12. Kratz F. Albumin as a drug carrier: design of prodrugs, drug conjugates and nanoparticles. J Control Release 2008;132:171–83. Laakkonen P, Porkka K, Hoffman JA, Ruoslahti E. A tumor-homing peptide with a targeting specificity related to lymphatic vessels. Nat Med 2002;8:751–5. Lammers T, Subr V, Ulbrich K, Peschke P, Huber PE, Hennink WE, et al. Simultaneous delivery of doxorubicin and gemcitabine to tumors in vivo using prototypic polymeric drug carriers. Biomaterials 2009;30:3466–75. Lammers T, Kiessling F, Hennink WE, Storm G. Nanotheranostics and image-guided drug delivery: current concepts and future directions. Mol Pharm 2010;7:1899–912. Latha MS, Lal AV, Kumary TV, Sreekumar R, Jayakrishnan A. Progesterone release from glutaraldehyde cross-linked casein microspheres: in vitro studies and in vivo response in rabbits. Contraception 2000;61:329–34. Lee ES, Na K, Bae YH. Polymeric micelle for tumor pH and folate-mediated targeting. J Control Release 2003;91:103–13. Lee SM, O'Halloran TV, Nguyen ST. Polymer-caged nanobins for synergistic cisplatin– doxorubicin combination chemotherapy. J Am Chem Soc 2010;132:17130–8. Lee IH, An S, Yu MK, Kwon HK, Im SH, Jon S. Targeted chemoimmunotherapy using drug-loaded aptamer–dendrimer bioconjugates. J Control Release 2011;155:435–41. Lee GY, Qian WP, Wang L, Wang YA, Staley CA, Satpathy M, et al. Theranostic nanoparticles with controlled release of gemcitabine for targeted therapy and MRI of pancreatic cancer. ACS Nano 2013;7:2078–89. Leung SJ, Romanowski M. Light-activated content release from liposomes. Theranostics 2012;2:1020–36. Li ZJ, Cho CH. Peptides as targeting probes against tumor vasculature for diagnosis and drug delivery. J Transl Med 2012;10:S1. Li ZJ, Wu WK, Ng SS, Yu L, Li HT, Wong CC, et al. A novel peptide specifically targeting the vasculature of orthotopic colorectal cancer for imaging detection and drug delivery. J Control Release 2010;148:292–302. Li MH, Choi SK, Thomas TP, Desai A, Lee KH, Kotlyar A, et al. Dendrimer-based multivalent methotrexates as dual acting nanoconjugates for cancer cell targeting. Eur J Med Chem 2012a;47:560–72. Li Y, He H, Jia X, Lu WL, Lou J, Wei Y. A dual-targeting nanocarrier based on poly(amidoamine) dendrimers conjugated with transferrin and tamoxifen for treating brain gliomas. Biomaterials 2012b;33:3899–908. Liepold L, Anderson S, Willits D, Oltrogge L, Frank JA, Douglas T, et al. Viral capsids as MRI contrast agents. Magn Reson Med 2007;58:871–9. Liu F, Shollenberger LM, Huang L. Non-immunostimulatory nonviral vectors. FASEB J 2004;18:1779–81. Liu Y, Wu DC, Zhang WD, Jiang X, He CB, Chung TS, et al. Polyethylenimine‐grafted multiwalled carbon nanotubes for secure noncovalent immobilization and efficient delivery of DNA. Angew Chem 2005;117:4860–3. Liu Z, Sun X, Nakayama-Ratchford N, Dai H. Supramolecular chemistry on water-soluble carbon nanotubes for drug loading and delivery. ACS Nano 2007;1:50–6. Liu Z, Chen K, Davis C, Sherlock S, Cao Q, Chen X, et al. Drug delivery with carbon nanotubes for in vivo cancer treatment. Cancer Res 2008;68:6652–60. Liu Q, Zhang J, Sun W, Xie QR, Xia W, Gu H. Delivering hydrophilic and hydrophobic chemotherapeutics simultaneously by magnetic mesoporous silica nanoparticles to inhibit cancer cells. Int J Nanomedicine 2012a;7:999–1013. Liu T, Li X, Qian Y, Hu X, Liu S. Multifunctional pH-disintegrable micellar nanoparticles of asymmetrically functionalized beta-cyclodextrin-based star copolymer covalently

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx conjugated with doxorubicin and DOTA-Gd moieties. Biomaterials 2012b;33: 2521–31. Ljubimova JY, Holler E. Biocompatible nanopolymers: the next generation of breast cancer treatment? Nanomedicine (Lond) 2012;7:1467–70. Ljubimova JY, Fujita M, Ljubimov AV, Torchilin VP, Black KL, Holler E. Poly(malic acid) nanoconjugates containing various antibodies and oligonucleotides for multitargeting drug delivery. Nanomedicine (Lond) 2008;3:247–65. Lu J, Liong M, Zink JI, Tamanoi F. Mesoporous silica nanoparticles as a delivery system for hydrophobic anticancer drugs. Small 2007a;3:1341–6. Lu W, Wan J, She Z, Jiang X. Brain delivery property and accelerated blood clearance of cationic albumin conjugated pegylated nanoparticle. J Control Release 2007b;118: 38–53. Luk BT, Fang RH, Zhang L. Lipid- and polymer-based nanostructures for cancer theranostics. Theranostics 2012;2:1117–26. Ma M, Chen H, Chen Y, Wang X, Chen F, Cui X, et al. Au capped magnetic core/mesoporous silica shell nanoparticles for combined photothermo-/chemo-therapy and multimodal imaging. Biomaterials 2012a;33:989–98. Ma X, Nguyen KT, Borah P, Ang CY, Zhao Y. Functional silica nanoparticles for redox-triggered drug/ssDNA co-delivery. Adv Health Mater 2012b;1:690–7. MacEwan SR, Callahan DJ, Chilkoti A. Stimulus-responsive macromolecules and nanoparticles for cancer drug delivery. Nanomedicine (Lond) 2010;5:793–806. Majewski AP, Schallon A, Jerome V, Freitag R, Muller AH, Schmalz H. Dual-responsive magnetic core-shell nanoparticles for nonviral gene delivery and cell separation. Biomacromolecules 2012;13:857–66. Majumder P, Gomes KN, Ulrich H. Aptamers: from bench side research towards patented molecules with therapeutic applications. Expert Opin Ther Pat 2009;19:1603–13. Mamasheva E, O'Donnell C, Bandekar A, Sofou S. Heterogeneous liposome membranes with pH-triggered permeability enhance the in vitro antitumor activity of folate-receptor targeted liposomal doxorubicin. Mol Pharm 2011;8:2224–32. Mashal A, Sitharaman B, Li X, Avti PK, Sahakian AV, Booske JH, et al. Toward carbon-nanotube-based theranostic agents for microwave detection and treatment of breast cancer: enhanced dielectric and heating response of tissue-mimicking materials. IEEE Trans Biomed Eng 2010;57:1831–4. Matsumura Y, Maeda H. A new concept for macromolecular therapeutics in cancer chemotherapy: mechanism of tumoritropic accumulation of proteins and the antitumor agent smancs. Cancer Res 1986;46:6387–92. Medarova Z, Pham W, Farrar C, Petkova V, Moore A. In vivo imaging of siRNA delivery and silencing in tumors. Nat Med 2007;13:372–7. Mehta RS, Barlow WE, Albain KS, Vandenberg TA, Dakhil SR, Tirumali NR, et al. Combination anastrozole and fulvestrant in metastatic breast cancer. N Engl J Med 2012;367: 435–44. Meng H, Liong M, Xia T, Li Z, Ji Z, Zink JI, et al. Engineered design of mesoporous silica nanoparticles to deliver doxorubicin and P-glycoprotein siRNA to overcome drug resistance in a cancer cell line. ACS Nano 2010;4:4539–50. Meng H, Mai WX, Zhang H, Xue M, Xia T, Lin S, et al. 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. Metzmacher I, Radu F, Bause M, Knabner P, Friess W. A model describing the effect of enzymatic degradation on drug release from collagen minirods. Eur J Pharm Biopharm 2007;67:349–60. Mi Y, Liu X, Zhao J, Ding J, Feng SS. Multimodality treatment of cancer with herceptin conjugated, thermomagnetic iron oxides and docetaxel loaded nanoparticles of biodegradable polymers. Biomaterials 2012;33:7519–29. Miele E, Spinelli GP, Tomao F, Tomao S. Albumin-bound formulation of paclitaxel (abraxane ABI-007) in the treatment of breast cancer. Int J Nanomedicine 2009;4: 99–105. Miller K, Wang M, Gralow J, Dickler M, Cobleigh M, Perez EA, et al. Paclitaxel plus bevacizumab versus paclitaxel alone for metastatic breast cancer. N Engl J Med 2007;357:2666–76. Mills JK, Needham D. Targeted drug delivery. Expert Opin Ther Pat 1999;9:1499–513. Mitra RN, Doshi M, Zhang X, Tyus JC, Bengtsson N, Fletcher S, et al. An activatable multimodal/multifunctional nanoprobe for direct imaging of intracellular drug delivery. Biomaterials 2012;33:1500–8. Moon HK, Lee SH, Choi HC. In vivo near-infrared mediated tumor destruction by photothermal effect of carbon nanotubes. ACS Nano 2009;3:3707–13. Morachis JM, Mahmoud EA, Sankaranarayanan J, Almutairi A. Triggered rapid degradation of nanoparticles for gene delivery. J Drug Deliv 2012;2012. Morille M, Passirani C, Vonarbourg A, Clavreul A, Benoit JP. Progress in developing cationic vectors for non-viral systemic gene therapy against cancer. Biomaterials 2008;29: 3477–96. Mouli SK, Tyler P, McDevitt JL, Eifler AC, Guo Y, Nicolai J, et al. Image-guided local delivery strategies enhance therapeutic nanoparticle uptake in solid tumors. ACS Nano 2013;7:7724–33. Mura S, Couvreur P. Nanotheranostics for personalized medicine. Adv Drug Deliv Rev 2012;64:1394–416. Murphy RF, Powers S, Cantor CR. Endosome pH measured in single cells by dual fluorescence flow cytometry: rapid acidification of insulin to pH 6. J Cell Biol 1984;98: 1757–62. Murphy EA, Majeti BK, Barnes LA, Makale M, Weis SM, Lutu-Fuga K, et al. Nanoparticle-mediated drug delivery to tumor vasculature suppresses metastasis. Proc Natl Acad Sci U S A 2008;105:9343–8. Nasongkla N, Bey E, Ren J, Ai H, Khemtong C, Guthi JS, et al. Multifunctional polymeric micelles as cancer-targeted, MRI-ultrasensitive drug delivery systems. Nano Lett 2006;6:2427–30. Nguyen J, Szoka FC. Nucleic acid delivery: the missing pieces of the puzzle? Acc Chem Res 2012;45:1153–62.

15

O'Connell MJ, Bachilo SM, Huffman CB, Moore VC, Strano MS, Haroz EH, et al. Band gap fluorescence from individual single-walled carbon nanotubes. Science 2002;297: 593–6. Oliveira H, Perez-Andres E, Thevenot J, Sandre O, Berra E, Lecommandoux S. Magnetic field triggered drug release from polymersomes for cancer therapeutics. J Control Release 2013;169:165–70. O'Neill BE, Rapoport N. Phase-shift, stimuli-responsive drug carriers for targeted delivery. Ther Deliv 2011;2:1165–87. Ong JC, Sun F, Chan E. Development of stealth liposome coencapsulating doxorubicin and fluoxetine. J Liposome Res 2011;21:261–71. Pan Y, Leifert A, Ruau D, Neuss S, Bornemann J, Schmid G, et al. Gold nanoparticles of diameter 1.4 nm trigger necrosis by oxidative stress and mitochondrial damage. Small 2009;5:2067–76. Pankhurst QA, Connolly J, Jones S, Dobson J. Applications of magnetic nanoparticles in biomedicine. J Phys D Appl Phys 2003;36:R167. Park SM, Kim MS, Park SJ, Park ES, Choi KS, Kim YS, et al. Novel temperature-triggered liposome with high stability: formulation, in vitro evaluation, and in vivo study combined with high-intensity focused ultrasound (HIFU). J Control Release 2013;170:373–9. Peer D, Karp JM, Hong S, Farokhzad OC, Margalit R, Langer R. Nanocarriers as an emerging platform for cancer therapy. Nat Nanotechnol 2007;2:751–60. Poon Z, Chang D, Zhao X, Hammond PT. Layer-by-layer nanoparticles with a pH-sheddable layer for in vivo targeting of tumor hypoxia. ACS Nano 2011;5: 4284–92. Porkka K, Laakkonen P, Hoffman JA, Bernasconi M, Ruoslahti E. A fragment of the HMGN2 protein homes to the nuclei of tumor cells and tumor endothelial cells in vivo. Proc Natl Acad Sci U S A 2002;99:7444–9. Qiu LY, Bae YH. Self-assembled polyethylenimine-graft-poly(epsilon-caprolactone) micelles as potential dual carriers of genes and anticancer drugs. Biomaterials 2007;28:4132–42. Quan Q, Xie J, Gao H, Yang M, Zhang F, Liu G, et al. HSA coated iron oxide nanoparticles as drug delivery vehicles for cancer therapy. Mol Pharm 2011;8:1669–76. Rapoport N, Gao Z, Kennedy A. Multifunctional nanoparticles for combining ultrasonic tumor imaging and targeted chemotherapy. J Natl Cancer Inst 2007;99:1095–106. Rejinold NS, Muthunarayanan M, Divyarani VV, Sreerekha PR, Chennazhi KP, Nair SV, et al. Curcumin-loaded biocompatible thermoresponsive polymeric nanoparticles for cancer drug delivery. J Colloid Interface Sci 2011;360:39–51. Rizk SS, Luchniak A, Uysal S, Brawley CM, Rock RS, Kossiakoff AA. An engineered substance P variant for receptor-mediated delivery of synthetic antibodies into tumor cells. Proc Natl Acad Sci U S A 2009;106:11011–5. Rossin R, Verkerk PR, van den Bosch SM, Vulders RC, Verel I, Lub J, et al. In vivo chemistry for pretargeted tumor imaging in live mice. Angew Chem Int Ed Engl 2010;49: 3375–8. Rothdiener M, Beuttler J, Messerschmidt SK, Kontermann RE. Antibody targeting of nanoparticles to tumor-specific receptors: immunoliposomes. Methods Mol Biol 2010;624:295–308. Roy R, Yang J, Moses MA. Matrix metalloproteinases as novel biomarkers and potential therapeutic targets in human cancer. J Clin Oncol 2009;27:5287–97. Rudnick SI, Lou J, Shaller CC, Tang Y, Klein-Szanto AJ, Weiner LM, et al. Influence of affinity and antigen internalization on the uptake and penetration of Anti-HER2 antibodies in solid tumors. Cancer Res 2011;71:2250–9. Ruoslahti E, Bhatia SN, Sailor MJ. Targeting of drugs and nanoparticles to tumors. J Cell Biol 2010;188:759–68. Saad M, Garbuzenko OB, Minko T. Co-delivery of siRNA and an anticancer drug for treatment of multidrug-resistant cancer. Nanomedicine (Lond) 2008;3:761–76. Saha K, Agasti SS, Kim C, Li X, Rotello VM. Gold nanoparticles in chemical and biological sensing. Chem Rev 2012;112:2739–79. Saito G, Swanson JA, Lee KD. Drug delivery strategy utilizing conjugation via reversible disulfide linkages: role and site of cellular reducing activities. Adv Drug Deliv Rev 2003;55:199–215. Santra S, Kaittanis C, Grimm J, Perez JM. Drug/dye-loaded, multifunctional iron oxide nanoparticles for combined targeted cancer therapy and dual optical/magnetic resonance imaging. Small 2009;5:1862–8. Sapra P, Moase EH, Ma J, Allen TM. Improved therapeutic responses in a xenograft model of human B lymphoma (Namalwa) for liposomal vincristine versus liposomal doxorubicin targeted via anti-CD19 IgG2a or Fab' fragments. Clin Cancer Res 2004;10:1100–11. Schmaljohann D. Thermo- and pH-responsive polymers in drug delivery. Adv Drug Deliv Rev 2006;58:1655–70. Sengupta S, Eavarone D, Capila I, Zhao G, Watson N, Kiziltepe T, et al. Temporal targeting of tumour cells and neovasculature with a nanoscale delivery system. Nature 2005;436:568–72. Sethuraman VA, Bae YH. TAT peptide-based micelle system for potential active targeting of anti-cancer agents to acidic solid tumors. J Control Release 2007;118:216–24. Shao J, Dai Y, Zhao W, Xie J, Xue J, Ye J, et al. Intracellular distribution and mechanisms of actions of photosensitizer Zinc(II)-phthalocyanine solubilized in cremophor EL against human hepatocellular carcinoma HepG2 cells. Cancer Lett 2013;330:49–56. Shen Y, Zhou Z, Sui M, Tang J, Xu P, Van Kirk EA, et al. Charge-reversal polyamidoamine dendrimer for cascade nuclear drug delivery. Nanomedicine (Lond) 2010;5:1205–17. Shenoy D, Little S, Langer R, Amiji M. Poly(ethylene oxide)-modified poly(beta-amino ester) nanoparticles as a pH-sensitive system for tumor-targeted delivery of hydrophobic drugs. 1. In vitro evaluations. Mol Pharm 2005;2:357–66. Shim G, Han SE, Yu YH, Lee S, Lee HY, Kim K, et al. Trilysinoyl oleylamide-based cationic liposomes for systemic co-delivery of siRNA and an anticancer drug. J Control Release 2011;155:60–6. Shin HC, Alani AW, Rao DA, Rockich NC, Kwon GS. Multi-drug loaded polymeric micelles for simultaneous delivery of poorly soluble anticancer drugs. J Control Release 2009;140:294–300.

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

16

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx

Singh A, Dilnawaz F, Mewar S, Sharma U, Jagannathan NR, Sahoo SK. Composite polymeric magnetic nanoparticles for co-delivery of hydrophobic and hydrophilic anticancer drugs and MRI imaging for cancer therapy. ACS Appl Mater Interfaces 2011a;3: 842–56. Singh N, Karambelkar A, Gu L, Lin K, Miller JS, Chen CS, et al. Bioresponsive mesoporous silica nanoparticles for triggered drug release. J Am Chem Soc 2011b;133:19582–5. Sinha R, Kim GJ, Nie S, Shin DM. Nanotechnology in cancer therapeutics: bioconjugated nanoparticles for drug delivery. Mol Cancer Ther 2006;5:1909–17. Slowing I, Trewyn BG, Lin VS. Effect of surface functionalization of MCM-41-type mesoporous silica nanoparticles on the endocytosis by human cancer cells. J Am Chem Soc 2006;128:14792–3. Slowing II, Trewyn BG, Lin VS. Mesoporous silica nanoparticles for intracellular delivery of membrane-impermeable proteins. J Am Chem Soc 2007;129:8845–9. Slowing II, Vivero-Escoto JL, Wu CW, Lin VS. Mesoporous silica nanoparticles as controlled release drug delivery and gene transfection carriers. Adv Drug Deliv Rev 2008;60: 1278–88. Soma CE, Dubernet C, Bentolila D, Benita S, Couvreur P. Reversion of multidrug resistance by co-encapsulation of doxorubicin and cyclosporin A in polyalkylcyanoacrylate nanoparticles. Biomaterials 2000;21:1–7. Sperling RA, Rivera Gil P, Zhang F, Zanella M, Parak WJ. Biological applications of gold nanoparticles. Chem Soc Rev 2008;37:1896–908. Storniolo AM, Pegram MD, Overmoyer B, Silverman P, Peacock NW, Jones SF, et al. Phase I dose escalation and pharmacokinetic study of lapatinib in combination with trastuzumab in patients with advanced ErbB2-positive breast cancer. J Clin Oncol 2008;26:3317–23. Stortecky S, Suter TM. Insights into cardiovascular side-effects of modern anticancer therapeutics. Curr Opin Oncol 2010;22:312–7. Stylianopoulos T, Wong C, Bawendi MG, Jain RK, Fukumura D. Multistage nanoparticles for improved delivery into tumor tissue. Methods Enzymol 2012;508:109–30. Suci PA, Berglund DL, Liepold L, Brumfield S, Pitts B, Davison W, et al. High-density targeting of a viral multifunctional nanoplatform to a pathogenic, biofilm-forming bacterium. Chem Biol 2007;14:387–98. Sugahara KN, Teesalu T, Karmali PP, Kotamraju VR, Agemy L, Girard OM, et al. Tissue-penetrating delivery of compounds and nanoparticles into tumors. Cancer Cell 2009;16:510–20. Sugahara KN, Teesalu T, Karmali PP, Kotamraju VR, Agemy L, Greenwald DR, et al. Coadministration of a tumor-penetrating peptide enhances the efficacy of cancer drugs. Science 2010;328:1031–5. Sumer B, Gao J. Theranostic nanomedicine for cancer. Nanomedicine 2008;3:137–40. Takae S, Miyata K, Oba M, Ishii T, Nishiyama N, Itaka K, et al. PEG-detachable polyplex micelles based on disulfide-linked block catiomers as bioresponsive nonviral gene vectors. J Am Chem Soc 2008;130:6001–9. Tanaka T, Mangala LS, Vivas-Mejia PE, Nieves-Alicea R, Mann AP, Mora E, et al. Sustained small interfering RNA delivery by mesoporous silicon particles. Cancer Res 2010;70: 3687–96. Tang LY, Wang YC, Li Y, Du JZ, Wang J. Shell-detachable micelles based on disulfide-linked block copolymer as potential carrier for intracellular drug delivery. Bioconjug Chem 2009;20:1095–9. Tasciotti E, Liu X, Bhavane R, Plant K, Leonard AD, Price BK, et al. Mesoporous silicon particles as a multistage delivery system for imaging and therapeutic applications. Nat Nanotechnol 2008;3:151–7. Thomas TP, Shukla R, Kotlyar A, Liang B, Ye JY, Norris TB, et al. Dendrimer-epidermal growth factor conjugate displays superagonist activity. Biomacromolecules 2008;9: 603–9. Torchilin VP, Lukyanov AN, Gao Z, Papahadjopoulos-Sternberg B. Immunomicelles: targeted pharmaceutical carriers for poorly soluble drugs. Proc Natl Acad Sci U S A 2003;100:6039–44. Trepel M, Pasqualini R, Arap W. Screening phage-display peptide libraries for vascular targeted peptides. . Chapter 4Methods Enzymol 2008;445:83–106. Trewyn BG, Giri S, Slowing II, Lin VS. Mesoporous silica nanoparticle based controlled release, drug delivery, and biosensor systems. Chem Commun (Camb) 2007: 3236–45. Tsuruo T, Naito M, Tomida A, Fujita N, Mashima T, Sakamoto H, et al. Molecular targeting therapy of cancer: drug resistance, apoptosis and survival signal. Cancer Sci 2003;94: 15–21. Tuerk C, Gold L. Systematic evolution of ligands by exponential enrichment: RNA ligands to bacteriophage T4 DNA polymerase. Science 1990;249:505–10. Ulbrich K, Subr V. Polymeric anticancer drugs with pH-controlled activation. Adv Drug Deliv Rev 2004;56:1023–50. Ulijn RV. Enzyme-responsive materials: a new class of smart biomaterials. J Mater Chem 2006;16:2217–25. Uner M, Yener G. Importance of solid lipid nanoparticles (SLN) in various administration routes and future perspectives. Int J Nanomedicine 2007;2:289–300. Vallet-Regi M, Balas F, Arcos D. Mesoporous materials for drug delivery. Angew Chem Int Ed Engl 2007;46:7548–58. Vaupel P. Tumor microenvironmental physiology and its implications for radiation oncology. Seminars in radiation oncology. Elsevier; 2004198–206. Veiseh O, Kievit FM, Ellenbogen RG, Zhang M. Cancer cell invasion: treatment and monitoring opportunities in nanomedicine. Adv Drug Deliv Rev 2011;63: 582–96. Vicent MJ, Greco F, Nicholson RI, Paul A, Griffiths PC, Duncan R. Polymer therapeutics designed for a combination therapy of hormone-dependent cancer. Angew Chem Int Ed Engl 2005;44:4061–6. Wan Q, Xie L, Gao L, Wang Z, Nan X, Lei H, et al. Self-assembled magnetic theranostic nanoparticles for highly sensitive MRI of minicircle DNA delivery. Nanoscale 2013;5:744–52.

Wang Y, Gao S, Ye WH, Yoon HS, Yang YY. Co-delivery of drugs and DNA from cationic core-shell nanoparticles self-assembled from a biodegradable copolymer. Nat Mater 2006;5:791–6. Wang X, Yang L, Chen ZG, Shin DM. Application of nanotechnology in cancer therapy and imaging. CA Cancer J Clin 2008;58:97–110. Wang X, Li J, Wang Y, Cho KJ, Kim G, Gjyrezi A, et al. HFT-T, a targeting nanoparticle, enhances specific delivery of paclitaxel to folate receptor-positive tumors. ACS Nano 2009;3:3165–74. Wang H, Zhao P, Su W, Wang S, Liao Z, Niu R, et al. PLGA/polymeric liposome for targeted drug and gene co-delivery. Biomaterials 2010;31:8741–8. Wang F, Wang YC, Dou S, Xiong MH, Sun TM, Wang J. Doxorubicin-tethered responsive gold nanoparticles facilitate intracellular drug delivery for overcoming multidrug resistance in cancer cells. ACS Nano 2011a;5:3679–92. Wang H, Zhao Y, Wu Y, Hu YL, Nan K, Nie G, et al. Enhanced anti-tumor efficacy by co-delivery of doxorubicin and paclitaxel with amphiphilic methoxy PEG-PLGA copolymer nanoparticles. Biomaterials 2011b;32:8281–90. Wang LS, Chuang MC, Ho JA. Nanotheranostics—a review of recent publications. Int J Nanomedicine 2012;7:4679–95. Wartlick H, Michaelis K, Balthasar S, Strebhardt K, Kreuter J, Langer K. Highly specific HER2-mediated cellular uptake of antibody-modified nanoparticles in tumour cells. J Drug Target 2004;12:461–71. Welsher K, Liu Z, Daranciang D, Dai H. Selective probing and imaging of cells with single walled carbon nanotubes as near-infrared fluorescent molecules. Nano Lett 2008;8: 586–90. Wiethoff CM, Middaugh CR. Barriers to nonviral gene delivery. J Pharm Sci 2003;92: 203–17. Wong MY, Chiu GN. Liposome formulation of co-encapsulated vincristine and quercetin enhanced antitumor activity in a trastuzumab-insensitive breast tumor xenograft model. Nanomedicine 2011;7:834–40. Wong C, Stylianopoulos T, Cui J, Martin J, Chauhan VP, Jiang W, et al. Multistage nanoparticle delivery system for deep penetration into tumor tissue. Proc Natl Acad Sci U S A 2011;108:2426–31. Wu J, Lu Y, Lee A, Pan X, Yang X, Zhao X, et al. Reversal of multidrug resistance by transferrin-conjugated liposomes co-encapsulating doxorubicin and verapamil. J Pharm Pharm Sci 2007;10:350–7. Wu H, Engelhard MH, Wang J, Fisher DR, Lin Y. Synthesis of lutetium phosphate– apoferritin core–shell nanoparticles for potential applications in radioimmunoimaging and radioimmunotherapy of cancers. J Mater Chem 2008a;18:1779–83. Wu H, Wang J, Wang Z, Fisher DR, Lin Y. Apoferritin-templated yttrium phosphate nanoparticle conjugates for radioimmunotherapy of cancers. J Nanosci Nanotechnol 2008b;8:2316–22. Wu W, Shen J, Banerjee P, Zhou S. Core-shell hybrid nanogels for integration of optical temperature-sensing, targeted tumor cell imaging, and combined chemophotothermal treatment. Biomaterials 2010a;31:7555–66. Wu X, Ming T, Wang X, Wang P, Wang J, Chen J. High-photoluminescence-yield gold nanocubes: for cell imaging and photothermal therapy. ACS Nano 2010b;4: 113–20. Wu XL, Kim JH, Koo H, Bae SM, Shin H, Kim MS, et al. Tumor-targeting peptide conjugated pH-responsive micelles as a potential drug carrier for cancer therapy. Bioconjug Chem 2010c;21:208–13. Xia Y, Li W, Cobley CM, Chen J, Xia X, Zhang Q, et al. Gold nanocages: from synthesis to theranostic applications. Acc Chem Res 2011;44:914–24. Xin H, Jiang X, Gu J, Sha X, Chen L, Law K, et al. Angiopep-conjugated poly(ethylene glycol)-co-poly(epsilon-caprolactone) nanoparticles as dual-targeting drug delivery system for brain glioma. Biomaterials 2011;32:4293–305. Xu Z, Gu W, Chen L, Gao Y, Zhang Z, Li Y. A smart nanoassembly consisting of acid-labile vinyl ether PEG-DOPE and protamine for gene delivery: preparation and in vitro transfection. Biomacromolecules 2008;9:3119–26. Xu Z, Zhang Z, Chen Y, Chen L, Lin L, Li Y. The characteristics and performance of a multifunctional nanoassembly system for the co-delivery of docetaxel and iSur-pDNA in a mouse hepatocellular carcinoma model. Biomaterials 2010;31:916–22. Yallapu MM, Othman SF, Curtis ET, Gupta BK, Jaggi M, Chauhan SC. Multi-functional magnetic nanoparticles for magnetic resonance imaging and cancer therapy. Biomaterials 2011;32:1890–905. Yamada Y, Harashima H. Mitochondrial drug delivery systems for macromolecule and their therapeutic application to mitochondrial diseases. Adv Drug Deliv Rev 2008;60:1439–62. Yamashita S, Yamashita J, Sakamoto K, Inada K, Nakashima Y, Murata K, et al. Increased expression of membrane-associated phospholipase A2 shows malignant potential of human breast cancer cells. Cancer 1993;71:3058–64. Yang H. Nanoparticle-mediated brain-specific drug delivery, imaging, and diagnosis. Pharm Res 2010;27:1759–71. Yang L, Cao Z, Sajja HK, Mao H, Wang L, Geng H, et al. Development of receptor targeted magnetic iron oxide nanoparticles for efficient drug delivery and tumor imaging. J Biomed Nanotechnol 2008;4:439–49. Yang X, Grailer JJ, Rowland IJ, Javadi A, Hurley SA, Matson VZ, et al. Multifunctional stable and pH-responsive polymer vesicles formed by heterofunctional triblock copolymer for targeted anticancer drug delivery and ultrasensitive MR imaging. ACS Nano 2010;4:6805–17. Yang X, Hong H, Grailer JJ, Rowland IJ, Javadi A, Hurley SA, et al. cRGD-functionalized, DOX-conjugated, and (6)(4)Cu-labeled superparamagnetic iron oxide nanoparticles for targeted anticancer drug delivery and PET/MR imaging. Biomaterials 2011;32: 4151–60. Yi DK, Sun IC, Ryu JH, Koo H, Park CW, Youn IC, et al. Matrix metalloproteinase sensitive gold nanorod for simultaneous bioimaging and photothermal therapy of cancer. Bioconjug Chem 2010;21:2173–7.

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

Y. Gao et al. / Biotechnology Advances xxx (2013) xxx–xxx Ying X, Wen H, Lu WL, Du J, Guo J, Tian W, et al. Dual-targeting daunorubicin liposomes improve the therapeutic efficacy of brain glioma in animals. J Control Release 2010;141:183–92. Yoo HS, Park TG. Folate receptor targeted biodegradable polymeric doxorubicin micelles. J Control Release 2004;96:273–83. Zhang P, Hu L, Yin Q, Feng L, Li Y. Transferrin-modified c[RGDfK]-paclitaxel loaded hybrid micelle for sequential blood–brain barrier penetration and glioma targeting therapy. Mol Pharm 2012;9:1590–8. Zhao Y, Lu Y, Hu Y, Li JP, Dong L, Lin LN, et al. Synthesis of superparamagnetic CaCO3 mesocrystals for multistage delivery in cancer therapy. Small 2010;6:2436–42. Zhou Z, Shen Y, Tang J, Fan M, Van Kirk EA, Murdoch WJ, et al. Charge‐reversal drug conjugate for targeted cancer cell nuclear drug delivery. Adv Funct Mater 2009;19: 3580–9.

17

Zhu C, Jung S, Luo S, Meng F, Zhu X, Park TG, et al. Co-delivery of siRNA and paclitaxel into cancer cells by biodegradable cationic micelles based on PDMAEMA-PCL-PDMAEMA triblock copolymers. Biomaterials 2010;31:2408–16. Zou P, Yu Y, Wang YA, Zhong Y, Welton A, Galban C, et al. Superparamagnetic iron oxide nanotheranostics for targeted cancer cell imaging and pH-dependent intracellular drug release. Mol Pharm 2010;7:1974–84. Zubris KA, Colson YL, Grinstaff MW. Hydrogels as intracellular depots for drug delivery. Mol Pharm 2012;9:196–200. Zucker D, Andriyanov AV, Steiner A, Raviv U, Barenholz Y. Characterization of PEGylated nanoliposomes co-remotely loaded with topotecan and vincristine: relating structure and pharmacokinetics to therapeutic efficacy. J Control Release 2012;160:281–9. Zupancich JA, Bates FS, Hillmyer MA. Aqueous dispersions of poly (ethylene oxide)b-poly (gamma–methyl–epsilon–caprolactone) block copolymers. Macromolecules 2006;39:4286–8.

Please cite this article as: Gao Y, et al, Nanotechnology-based intelligent drug design for cancer metastasis treatment, Biotechnol Adv (2013), http://dx.doi.org/10.1016/j.biotechadv.2013.10.013

Nanotechnology-based intelligent drug design for cancer metastasis treatment.

Traditional chemotherapy used today at clinics is mainly inherited from the thinking and designs made four decades ago when the Cancer War was declare...
2MB Sizes 0 Downloads 0 Views