+ All Categories
Home > Documents > amin2015.pdf

amin2015.pdf

Date post: 18-Dec-2015
Category:
Upload: narayanan158
View: 213 times
Download: 0 times
Share this document with a friend
Popular Tags:
49
Surface Modification and Local Orientations of Surface Molecules in Nanoth- erapeutics Md. Lutful Amin, Jae Yeon Joo, Dong Kee Yi, Seong Soo A. An PII: S0168-3659(15)00235-7 DOI: doi: 10.1016/j.jconrel.2015.04.017 Reference: COREL 7638 To appear in: Journal of Controlled Release Received date: 25 December 2014 Revised date: 9 April 2015 Accepted date: 12 April 2015 Please cite this article as: Md. Lutful Amin, Jae Yeon Joo, Dong Kee Yi, Seong Soo A. An, Surface Modification and Local Orientations of Surface Molecules in Nanotherapeu- tics, Journal of Controlled Release (2015), doi: 10.1016/j.jconrel.2015.04.017 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Transcript
  • Surface Modification and Local Orientations of Surface Molecules in Nanoth-erapeutics

    Md. Lutful Amin, Jae Yeon Joo, Dong Kee Yi, Seong Soo A. An

    PII: S0168-3659(15)00235-7DOI: doi: 10.1016/j.jconrel.2015.04.017Reference: COREL 7638

    To appear in: Journal of Controlled Release

    Received date: 25 December 2014Revised date: 9 April 2015Accepted date: 12 April 2015

    Please cite this article as: Md. Lutful Amin, Jae Yeon Joo, Dong Kee Yi, Seong Soo A.An, Surface Modication and Local Orientations of Surface Molecules in Nanotherapeu-tics, Journal of Controlled Release (2015), doi: 10.1016/j.jconrel.2015.04.017

    This is a PDF le of an unedited manuscript that has been accepted for publication.As a service to our customers we are providing this early version of the manuscript.The manuscript will undergo copyediting, typesetting, and review of the resulting proofbefore it is published in its nal form. Please note that during the production processerrors may be discovered which could aect the content, and all legal disclaimers thatapply to the journal pertain.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    1

    Surface Modification and Local Orientations of Surface

    Molecules in Nanotherapeutics

    (REVISED VERSION II)

    Ms. Ref. No.: JCR-D-14-01819

    Md. Lutful Aminad, Jae Yeon Jooa, Dong Kee Yibc*, Seong Soo A. Ana**

    aDepartment of BioNano Technology, Gachon University, Gyeonggi-do, Republic of Korea.

    bDepartment of Chemistry, Myongji University, Yongin, Gyeonggi-do, Republic of Korea.

    cDepartment of Energy and Biotechnology, Myongji University

    dDepartment of Pharmacy, Stamford University Bangladesh

    * Corresponding author at: Department of Chemistry, Myongji University, 116 Myongji-ro,

    Cheoin-gu, Yongin, Gyeonggi-do 449 728, Republic of Korea, Tel: +82 31 330 6178, Email:

    [email protected]; [email protected]

    ** Corresponding author at: Department of BioNano Technology, Gachon University, 1342

    Seongnamdaero, Sujeong-gu, Seongnam, Gyeonggi-do 461 701; Republic of Korea, Tel: +82

    31 750 8981, Email: [email protected]; [email protected]

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    2

    ABSTRACT

    Nanotechnology has emerged as a powerful tool for various therapeutic applications, solving

    many difficulties in both diagnosis and treatment. However, many obstacles in complex

    biological systems have impeded the successful application of therapeutic nanoparticles, and

    fine-tuning nanoparticle properties has become extremely important in developing highly

    effective nanomedicines. To this end, particles have been engineered in various ways, with a

    special emphasis on surface modifications. The nanoparticle surface contacts the biological

    environment, and is a crucial determinant of the response. Thus, surface coating, surface charge,

    conjugated molecules, shape, and topography have enormous impacts on the total behavior of

    nanoparticles, including their biodistribution, stability, target localization, cellular interaction,

    uptake, drug release, and toxicity. Hence, engineering of the particle surface would provide

    wider dimensions of control for the specific and precise functions in the development of smart

    nanomedicines. Moreover, local orientation of nanoparticles in vivo and orientations of surface

    molecules are critical for their efficacy. Herein, we analyze surface functionalities, focusing on

    their mechanisms and respective advantages, and summarize results of surface engineering and

    renovating applications of nanoparticles.

    Keywords: targeting strategy, surface engineering, surface orientation, surface topography, drug

    delivery, nanotoxicity.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    3

    1. INTRODUCTION

    The emergence of nanotechnology in medical science has improved drug delivery systems in

    terms of solubility, membrane permeability, drug targeting, and controlled release [1,2]. Over the

    last few decades, development of therapeutic nanocarriers has become an important area of

    research, as they can improve the delivery of active molecules to the target site and overcome

    biological barriers [1,3]. In complex biological systems, nanomedicines have faced obstacles

    such as rapid clearance, lack of proper interaction, nonspecific targeting, inability to enter

    targeted cells and the core of tissues, and uncontrolled drug release [4,5]. Several investigations

    have found that nanoparticle surfaces greatly impact their performance [6,7]. Because

    nanoparticle surfaces come in contact with biological systems, surface characteristics determine

    how the nanoparticles will behave in vivo and the overall effectiveness of the particles, including

    stability, cellular interaction, uptake, drug delivery, and toxicity [4,6,8,9]. Therefore, engineering

    of the particle surface provides another dimension of control for specifically tailoring the

    functions and developing smart nanoparticles.

    Nanoparticle surfaces have been engineered in various ways for specific functions depending on

    the intended therapeutic use. Surface modifications such as attachment of bioactive and

    penetrating molecules, use of stimulus-responsive materials, surface coating, and modification of

    surface charge, texture, and shape allow nanoparticles to serve as smart devices that interact

    selectively with targeted cells, enter the core of tissues, and release drugs at a controlled rate,

    escaping from body clearance [4,10]. Surface modification with hydrophilic biocompatible

    polymers increases the circulation time and delays the removal of nanoparticles by the

    mononuclear phagocyte system (MPS), preventing rapid clearance prior to reaching the targeted

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    4

    tissue and delivering drugs [11,12]. The surface charge of nanoparticles has been modified

    according to the charge of the target cells or tissues to increase cellular interaction and

    internalization [13]. Conjugating active molecules on the surface of nanoparticles converts them

    into guided vehicles that exhibit cell-specific targeted interactions and penetrate to the core of

    the target tissue [1,14,15]. Attachment of specific penetrating peptides to nanoparticles is also

    effective for enhancing cellular uptake [16]. Nanoparticle surfaces have been decorated with

    stimulus-responsive materials to release drugs in response to various stimuli [17]. Importantly,

    the orientation and alignment of surface molecules affect their functions and the overall

    efficiency of nanomedicines [15,16]. In addition, modification of surface roughness and

    nanoparticle shapes significantly affect cellular interactions and drug release [18], and

    nanoparticle toxicity is related to surface properties, as variation in surface characteristics has

    significantly lowered the level of toxicity [19,20]. Combining these properties has allowed

    development of intelligent nanoparticles to meet delivery and therapeutic objectives [21].

    Surface engineering thus has become a major tool to address the particular challenges of

    nanotherapeutics. Studies have shown how surface engineered nanoparticles can efficiently

    overcome the limitations of other delivery systems [22]. Nowadays, proper surface

    functionalization has become a prerequisite for effective application of nanoparticles [23]. In this

    review, we focus on different surface characteristics of nanoparticles and discuss conventional

    and newer functionalities that have been discovered to increase the efficiency of nanoparticles.

    We further review the effect of surface molecule and nanoparticle orientations on function.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    5

    2. SURFACE MODIFICATION FOR ENHANCED STABILITY OF

    NANOPARTICLES

    Surface properties play the most important role for nanoparticle stability in blood circulation.

    Conventional nanoparticles with unmodified surfaces are unstable and rapidly cleared by the

    MPS before reaching the target tissue. Nanoparticles are recognized by the immune system

    through opsonization, a process by which a particle becomes covered with opsonin proteins,

    making it more visible to the MPS (liver, spleen, lungs, and bone marrow), and thereby

    increasing clearance [24]. Different functional groups on nanoparticle surfaces are the primary

    determinants of stability [24,25]. Usually, the degree of hydrophobicity determines the extent of

    opsonin binding: a high level of hydrophilicity is associated with lower opsonin binding and

    higher stability of nanoparticles in blood circulation, whereas hydrophobic surfaces induce

    nanoparticle aggregation through hydrophobic interactions and minimize surface energy, which

    can trigger opsonization and rapid elimination [26].

    The stability of nanoparticles in blood has been increased by surface modification. The surface

    has been coated with biocompatible hydrophilic polymers/surfactants or copolymers with

    hydrophilic properties. Hydrophilic polymers on the surface of the nanoparticles repel other

    molecules by steric effects. Thus, nanoparticles are not covered with opsonin proteins and are

    protected from rapid clearance [27]. Widely used surface-coating materials for prolonging the

    circulation time include polyethylene glycol (PEG), polyethylene oxide (PEO),

    polyvinylpyrrolidone (PVP), polyacrylic acid, dextran, poloxamer, poloxamine, and polysorbate

    (Tween-80). PEG is the most widely used material and has efficiently enhanced the stability of

    nanoparticles in many studies [26,27,28].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    6

    In addition to the density of the polymer on the nanoparticle surface, polymer conformation

    plays a significant role in determining stability [27]. PEG molecules with brush-like

    configurations on the surface of nanoparticles reduce phagocytosis and complement activation,

    whereas mushroom-like structures of PEG are potent complement activators and favor

    phagocytosis [11,29]. Polymer conformation is described using the Flory radius, F (Eq. 1), which

    is determined by the number of monomers per polymer chain, n, and the length of one monomer,

    , in Angstroms ( = 3.5 for PEG). Hence, F increases with increasing molecular weight of

    PEG. Polymers that are less densely grafted to the nanoparticle surface form mushroom-like

    structures wherein the mean distance between each grafting site is larger than the polymer size

    (Flory radius); thus, the individual polymer chains remain separated without interacting. When

    the mean distance is decreased and F is increased, each polymer chain overlaps, resulting in

    higher grafting density and polymer interaction. It forms a brush like conformation PEG,

    extending from the NP surface [30]. Brush-like PEG chains on the surface establish a

    hydrophilic lining that prevents attachment of molecules to the nanoparticles [31]. PEG densities

    below 9% on the surface of nanoparticles form a mushroom-like structure, and those above 9%

    form a more rigid, brush-like morphology [32]. Larger nanoparticles can be coated with smaller

    lengths of PEG (3,40010,000 monomers) because increases in hydrodynamic radius can shorten

    the half-life [33].

    Equation 1:

    F = n3/5 -------------------------------------------------------------------------------------------------- (1)

    Superparamagnetic iron oxide nanoparticles (45 nm) were coated with 2030 dextran chains in

    a brush-like conformation, which reduced the rapid clearance of the nanoparticles from the

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    7

    bloodstream with a prolonged half-life (t1/2) of 34 h [34]. Figure 1 illustrates the effect of PEG

    density on serum exposure and macrophage uptake. Neutral functional groups on nanoparticle

    surfaces provide excellent protection from opsonization, whereas charged functional groups are

    responsible for active nanoparticle interaction with target cells [35,36]. A zwitterionic copolymer

    showed excellent shielding of nanoparticles in blood circulation and maximized their interactions

    with target tissue. Yuan et al. developed a zwitterionic polymer-based nanoparticle for enhanced

    drug delivery to tumors. The nanoparticles were neutrally charged at physiological conditions

    and thus showed a prolonged circulation time (t1/2 24.05 h in a three-compartment

    pharmacokinetic model). After leaking into the tumor, the nanoparticles became positively

    charged in the acidic extracellular environment, and were efficiently taken up by tumor cells

    [35]. In our previous study, we modified silica nanoparticles with a PEG-conjugated

    polyethyleneimine copolymer, and studied the biodistribution profile by ICP-MS and

    fluorescence imaging. The observed biodistribution indicated that the particles were retained for

    a long time, and efficiently entered A549 lung cancer cells [37].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    8

    Figure 1. Influence of PEG density on serum protein adsorption to gold nanoparticles and their

    subsequent uptake by macrophages. (Adapted with permission from C.D. Walkey, J.B. Olsen, H.

    Guo, A. Emili, W.C. Chan. Nanoparticle size and surface chemistry determine serum protein

    adsorption and macrophage uptake. J. Am. Chem. Soc. 134 (2012) 2139-2147. Copyright (2012)

    American Chemical Society.)

    Although hydrophilic materials on the surface of nanoparticles prevent them from being

    opsonized, hydrophobic characteristics are often required for purposes such as controlled release

    and increasing membrane permeability and cellular uptake [38]. Amphiphilic copolymers with

    both hydrophilic and hydrophobic residues have shown promising results addressing these

    issues. The hydrophobic part interacts with the hydrophobic core, whereas the hydrophilic part

    remains outside to protect the nanoparticles from opsonization. Subtle Change in the degree of

    surface hydrophilicity and hydrophobicity has also shown substantial effect on circulation and

    uptake. [38,39]. Block copolymers have been popular for designing nanomaterials, incorporating

    desired properties of several polymers [40]. Novel amphiphilic triblock copolymers consisting of

    PEG as the hydrophilic segment and poly(-caprolactone) as the hydrophobic block were used

    successfully to encapsulate the poorly water-soluble anticancer drug 4 -demethyl-

    epipodophyllotoxin via hydrophobic interactions, resulting in controlled drug release [41].

    3. SURFACE CHARGE AND CELLULAR INTERACTIONS

    Because cell membranes are charged, nanoparticle surface charge plays a major role in active

    cell interaction and cellular uptake. Neutral and negatively charged nanoparticles show lower

    levels of internalization than positively charged particles because of the negative charge of the

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    9

    cell membrane [35]. Several cationic polymers such as poly(L-lysine) (PLL), poly(ethylenimine)

    (PEI), chitosan, and diethylaminoethyl-dextran (DEAE-DEX) have been used successfully for

    cellular uptake. PEI bears the highest density of charged groups on its surface and greatly

    increases membrane permeability, as reported in several studies [42-46].

    Cationic particles bind to negatively charged groups on the cell surface (e.g., sialic acid) and

    translocate across the plasma membrane [43]. The amount of positive charge is correlated with

    cellular internalization. Clathrin-dependent endocytosis occurs at lower charge density, whereas

    strongly charged surface coatings bypass typical uptake mechanisms in certain cells [13]. When

    the endocytic pathway is inhibited, compensatory endocytosis causes higher internalization of

    cationic particles [47]. Among cationic polymers that do not possess fusogenic activity, some

    can disrupt endosomal membranes through their buffering capacity [48]. Moreover, plasma

    proteins can bind to the cationic surface of nanoparticles and increase cellular uptake [49].

    Depolarization of the plasma membrane and formation of holes in lipid bilayers by cationic

    nanoparticles is common regardless of the shape, chemical composition, deformability, charge

    density, or size of the particles [50]. Positively charged nanoparticles are also effective in

    mucosal drug delivery. Nanoparticles coated with chitosan or hyaluronic acid showed increased

    in vivo association to gastrointestinal tract mucosae, nasal mucosae, and the cornea [48].

    Efficient uptake of cross-linked chitosan nanoparticles (e.g., particles having a zeta potential of

    +15 to +30 mV) by Caco-2 cells has been reported in several studies [51]. Positively charged

    nanoparticles have also shown increased penetration through the skin [52,53].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    10

    Despite these findings showing greater interactions, a positively charged surface is not a

    prerequisite for efficient endocytosis. Neutral or negatively charged nanoparticles also show

    efficient cellular uptake, especially when conjugated with targeting ligands. There is also

    evidence of internalization of negatively charged particles without targeting ligand. Nonspecific

    binding and clustering of the particles on the few cationic sites on the plasma membrane are

    believed to be responsible for their internalization [47,54,55]. Oligonucleotide-modified gold

    nanoparticles were examined for cellular uptake and were readily taken up by mouse endothelial

    cells even with negative surface charge [56]. Further study revealed the adsorption of serum

    proteins on the surface of the nanoparticles through electrostatic and hydrophobic interactions,

    which allowed them to interface with the cell membrane [56,57]. The possible role of

    nanoparticle surface charge on entry through the apical plasma membrane (facing the external

    environment) was investigated. Both cationic and anionic nanoparticles were targeted mainly to

    the clathrin endocytic pathway. A few nanoparticles (both positive and negative) were thought to

    be internalized through a macropinocytosis-dependent pathway. Many nanoparticles

    transcytosed and accumulated at the basolateral membrane, and some anionic nanoparticles

    transited through the degradative lysosomal pathway. It was therefore suggested that cationic

    nanoparticles may be promising carriers for transcytosing drugs [58].

    Cationic nanoparticles have been found to cross the blood-brain barrier (BBB) to a greater

    extent, resulting in increased brain penetration [59]. However, highly cationic and anionic

    nanoparticles were found to cause charge mediated disruption of the BBB. Neutral and slightly

    negative NPs were more effective in delivering therapeutics in brain, maintaining the integrity of

    BBB [60].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    11

    4. CELL TARGETING LIGAND ON THE SURFACE OF

    NANOPARTICLE

    Targeting molecules are the major components in active tissue targeting. Specific targeting

    moieties can be conjugated to the surface of a core particle that recognizes the unique surface

    properties of the targeted cells. Molecular interactions between nanoparticle and cell surfaces are

    based on ligand-receptor binding theory [61]. Targeting agent-linked nanoparticles are

    internalized through the pathway utilized by the unconjugated ligand, and the higher

    concentrations of targeting agents on nanoparticle surfaces cause stronger cell interactions

    compared to those observed with the ligand alone [40]. Figure 2 depicts a comparison of cellular

    uptake among nanoparticles conjugated to a targeting moiety, PEGylated nanoparticles, and

    unconjugated nanoparticles. It shows a high cellular uptake of folate-conjugated nanoparticles

    and greater internalization of PEGylated particles than that observed with unmodified

    nanoparticles because of increased circulation time. Various molecules such as folic acid, biotin,

    transferrin, peptides, enzyme substrates, mono- or oligosaccharides (which specifically interact

    with the membrane protein lectin), and antibodies have been successfully used as the targeting

    moiety [62]. Table 1 summarizes some investigations of ligand-conjugated nanocarriers and their

    extent of cellular uptake. In most of these studies, nanoparticles were successfully internalized

    despite bearing negative surface charges. The uptake of particles was associated with the

    targeting moiety rather than with the effect of surface charge. Successful uptake was also shown

    for positively charged particles. These findings suggest that the surface charge of nanoparticles is

    not significant when targeting moieties are conjugated.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    12

    Figure 2. Confocal microscopic images of KB cells treated with (A) PLGA nanoparticles, (B)

    functional PLGA nanoparticles (PLL-PEG coated), and (C) folate-conjugated nanoparticles.

    (Adapted with permission from S.H. Kim, J.H. Jeong, K.W. Chun, T.G. Park. Target-specific

    cellular uptake of PLGA nanoparticles coated with poly(l-lysine)-poly(ethylene glycol)-folate

    conjugate. Langmuir 21 (2005) 88528857. Copyright (2012) American Chemical Society.)

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    13

    Table 1.

    Different targeting moieties on nanoparticle surfaces and extent of their internalization.

    Targeting

    Moiety

    Nanoparti

    cle

    Size

    (nm)

    Surface

    Charge

    (mV)

    Targeted

    Tissue

    Result Conjugation

    Method

    Referen

    ces

    Folate

    PLGA

    Nanoparti

    cles

    108.9 -18.4 KB 6.7-fold

    higher uptake

    Covalent

    Binding

    (amide

    linkage)

    [63]

    Iron

    Oxide

    Nanoparti

    cles

    60-80 -12.5 HeLa >3-fold

    higher uptake

    Covalent

    Binding

    (amide

    linkage)

    [64]

    Transferri

    n

    PLGA

    Nanoparti

    cles

    220 -8 PC3 3-fold higher

    uptake

    Linker

    Chemistry

    [65]

    Biotin Gold

    Nanoparti

    cles

    20-40 - HeLa,

    A549,

    MG63

    >2-fold

    higher uptake

    Covalent

    Binding

    (amide

    linkage)

    [66]

    Chitosan

    Nanoparti

    cles

    165.3 +16.5 MCF-7

    cells

    >2-fold

    higher uptake

    Covalent

    Binding

    (amide

    linkage)

    [67]

    Monoclo

    nal

    Antibody

    PLGA

    Nanoparti

    cles

    124.2 +12.0 MDA-

    MB-231

    cells

    >2-fold

    higher uptake

    Covalent

    Binding

    (amide

    [68]

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    14

    linkage)

    RGD

    Peptide

    Albumin

    Nanoparti

    cles

    130 -30.77 BxPC-3 >2-fold

    higher uptake

    (after 24 h)

    Linker

    Chemistry

    [69]

    Galactose Gold

    Nanoparti

    cles

    50 -30 Hepatocy

    te

    16-fold

    higher uptake

    Au-S Binding [70]

    Targeting moieties are attached to nanoparticles with longer chains so that they extend outside of

    the dense polymer brush, thereby avoiding steric hindrance and enabling binding to the targeted

    receptors [61,71]. However, increased ligand concentrations on nanoparticle surfaces can

    enhance rapid clearance due to adsorption of opsonin. Gu et al. used aptamers as a targeting

    ligand for prostate cancer-specific antigens (PSMA) and found greater accumulation of

    nanoparticles in the liver and the spleen with higher surface ligand density. The nanoparticles

    also showed lower tumor localization compared to the particles functionalized with lower

    densities of aptamers [72]. These findings were supported by Kirpotin et al., who used an HER-2

    antibody and observed higher clearance rates with dense ligand loading. High targeting moiety

    surface density makes the stealth polymer layer ineffective, resulting in serum protein binding

    [73], which can diminish the ability of the ligands to interact with the target receptors [74].

    Therefore, the number, density, and orientation of targeting ligands on the surface of

    nanoparticles are important for effective targeting.

    The active site of the ligand must be available in the correct three-dimensional arrangement to

    bind to the receptor [75]. Particular consideration is required to ensure that serum proteins do not

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    15

    cover the targeting moiety. As previously discussed, the addition of targeting moieties may

    compromise the stealth feature of nanoparticles and can accelerate their clearance [76]. Salvati et

    al. developed transferrin-modified PEGylated silica nanoparticles (86 nm) and observed that the

    accumulation and uptake of transferrin-modified particles was lower than that of the unmodified

    particles. The number of transferrin molecules per nanoparticle was high (~250), and a protein

    corona formed on the nanoparticles [74]. McNeeley et al. observed a similar effect for

    doxorubicin-loaded, folate-modified PEGylated liposomal formulations: the incorporation of

    0.15% (of total lipid formulation) folate conjugates resulted in a 41.8% smaller area under the

    curve (AUC), whereas formulations containing 0.2% folate conjugates demonstrated a 61.9%

    reduction in AUC compared to that observed with non-targeted liposomes [77]. These results

    imply that low ligand concentration can increase half-life and improve biodistribution of

    nanoparticles.

    Lee et al. developed folate-modified tetrahedron-shaped oligonucleotide nanoparticles (28.6 nm)

    for siRNA delivery and showed that at least three folate molecules per nanoparticle are required

    for optimal delivery to cancer cells. Increasing the number of folate molecules did not increase

    the cellular uptake. The nanoparticles showed a longer circulation time in vivo (t1/2 24.2 min).

    They found that gene silencing occurred only when the ligands were in the appropriate spatial

    orientation. The density and location of folate on the tetrahedron were varied, and greater gene

    silencing was observed when the local density of three folic acids on the tetrahedron was

    maximized (encompassing a face of the tetrahedron). However, when the density of the three

    folic acid molecules was decreased (greater distance from each other), no gene silencing was

    observed. They speculated that higher local ligand density might influence the intracellular

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    16

    trafficking pathway of nanoparticles through the cells and affect gene silencing [78]. Elias et al.

    developed ligand conjugated magnetic nanoparticle and showed that an intermediate ligand

    density can provide greater cellular association compared with higher and lower ligand densities.

    Even though multivalency increases binding avidity, over-crowding can prevent ligands from

    obtaining the correct orientation. Furthermore, higher ligand density can saturate receptor

    binding [79]. The receptor density reaches an entropic limit at critically large ligand density,

    resulting in insufficient adhesion to overcome membrane deformation cost [80].

    These investigations indicate that the number and orientation of the targeting ligand should be

    considered for proper biodistribution and cellular uptake. Optimizing the number of ligands and

    their orientations is critical for nanotherapeutics and is not equivalent to increasing the number of

    ligands per nanoparticle.

    4.1 METHODS CURRENTLY BEING DEVELOPED FOR TARGETING

    Methods for increasing target specificity are currently being developed, and some show

    increased efficiency over conventional targeting strategies. These methods are based on surface

    engineering and involve two steps. In the first step, signaling molecules or surface labels are

    delivered to the target cells; in the second step, therapeutic nanoparticles with complimentary

    surface functionality target those molecules. This has been described by two approaches called

    communication control system and chemical targeting system; both effectively amplify the

    location of a tumor and subsequently target the amplified tumor with therapeutic particles [81].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    17

    Von Maltzahn et al. reported the communication control system where tumor-targeted tissue

    factor (tTF-RGD) was delivered as a signaling module to activate the coagulation cascade in

    tumors. Coagulation transglutaminase factor XIII (FXIII)-binding peptides were conjugated on

    the surface of doxorubicin loaded liposomes which acted as the receiving module to target the

    coagulation. This autonomous communication system improved the tumor targeting efficiency

    by up to 40-fold relative to particles without a communication system [82].

    The chemical targeting system involves in vivo metabolic labeling with artificial chemical

    groups. Unnatural azide-functionalized glycans are delivered to cells and become incorporated

    by glycosylation. These azide functional groups are nonreactive and compatible with endogenous

    biological molecules. Cells effectively "tag" glycoproteins with the azide group by using an

    intrinsic glycan synthesis process and express azide-modified sialic acids on their surface. This

    azide group then can be specifically targeted by an alkyne-activated compound, using copper-

    free "click" chemistry. Various cancer cell lines including A549, U87MG, MCF-7, and KB cells

    have been shown to express azide groups homogeneously on the cytoplasmic membrane [83,84].

    Lee et al. reported tumor targeting using a chemical targeting system (Figure 3). In the first step,

    an unnatural glycan, tetraacetylated N-azidoacetyl-D-mannosamine (Ac4ManNAz), was used to

    introduce azide groups on the cell surface. Ac4ManNAz was delivered to cells by using chitosan

    nanoparticles, which accumulated in tumor cells by enhanced permeability and retention effect.

    Cells expressed azide-containing glycoproteins in a dose-dependent manner. In the second step,

    Ce6-loaded bicyclo[6.1.0]-nonyne-modified nanoparticles were delivered to A549 human lung

    cancer cells for effective photodynamic therapy. The nanoparticles showed significantly higher

    tumor accumulation than did unmodified particles [84].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    18

    Figure 3. Tumor targeting strategy via metabolic glycoengineering and click chemistry.

    (Adapted with permission from S. Lee, H. Koo, J.H. Na, S.J. Han, H.S. Min, S.J. Lee, S.H. Kim,

    S.H. Yun, S.Y. Jeong, I.C. Kwon, K. Choi, K. Kim. Chemical tumor-targeting of nanoparticles

    based on metabolic glycoengineering and click chemistry. ACS Nano 8 (2014) 20482063.

    Copyright (2012) American Chemical Society.)

    5. PENETRATING AGENTS

    Cell-penetrating peptides (CPPs) are short peptide sequences that can assist nanoparticle entry

    into the cell and have delivered drugs 200 times larger than their established size limit for

    bioavailability. Surface modification with CPPs enhances the cell permeability of nanoparticle-

    based drug delivery systems [85,86]. Unlike targeting ligands, CPPs can assist nanoparticle

    internalization to various types of cells (not specific for a particular cell type) and are especially

    useful for internalizing nanoparticles by passive targeting [87,88]. Trans-activating

    transcriptional activator (TAT) has been widely tested and has shown potential for enhancing

    cellular uptake. Other CPPs include penetratin, transportan, VP-22, MPG, Pep-1, MAP, SAP,

    PPTG1, hCT (932), RGD, and SynB [89].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    19

    TAT moieties were attached to the surface of long-circulating PEGylated liposomes and PEG-

    phosphatidylethanolamine (PEG-PE)-based micelles by using TATp-short PEG-PE derivatives

    and showed high level of internalization [90]. Superparamagnetic iron oxide particles,

    conjugated to the TAT peptide sowed approximately 100-fold higher cell penetration than non-

    modified particles [91]. Zhao et al. developed paclitaxel-loaded polymeric liposomes with

    surface-conjugated TAT peptide and folic acid (size: 6070 nm; zeta potential: 28.69 mV). A

    cellular internalization study showed that TAT peptide-modified particles, and folate-TAT

    peptide-modified particles had 2.63-fold and 3.38-fold higher uptakes, respectively, than

    unmodified particles in A549 cells, and 6.67-fold and 24.27-fold higher uptakes, respectively,

    compared to unmodified particles in KB cells [88].

    Todorova et al. studied the effects of TAT peptide concentration and surface arrangement on

    membrane permeation capacity using 3-nm gold nanoparticles. They found that maximal

    penetration occurred with 5.4% TAT (Figure 4). A high TAT concentration could reduce the

    stability of nanoparticles in solution and result in lower cell penetration ability. Dense

    arrangement of TAT peptides resulted in like-charge repulsion, which changed the solution

    conformations to disfavor membrane interactions [92]. Therefore, it is necessary to control and

    maintain appropriate surface coverage of CPPs to produce stable and cell-permeating

    nanoparticles.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    20

    Figure 4. TEM images of nanoparticles in the cytosol showing the effect of CALNNTAT

    concentrations on the internalization of gold nanoparticles by HeLa cells. (a) 1.8%; (b) 3.6%; (c)

    5.4%; and (d) 9.0%. (Adapted with permission from N. Todorova, C. Chiappini, M. Mager, B.

    Simona, I.I. Patel, M.M. Stevens, I. Yarovsky. Surface presentation of functional peptides in

    solution determines cell internalization efficiency of tat conjugated nanoparticles. Nano Lett 14

    (2014) 5229-5237. Copyright (2012) American Chemical Society.)

    Petersen et al. developed penetratin-conjugated gold nanoparticles and demonstrated up to 100%

    uptake within 2 h in a preliminary biological study [93]. Most CPPs are positively charged, and

    zeta potential is thought to play a major role in cellular uptake; however, CPP-conjugated

    nanoparticles with negative zeta potentials also show effective cellular uptake. PEG-poly(lactic

    acid) nanoparticles were functionalized by penetratin for efficient biodistribution and brain drug

    delivery. The particles (100 nm) with a zeta potential of -4.42 mV showed enhanced cellular

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    21

    uptake in MDCK-MDR cells via both lipid raft-mediated endocytosis and direct translocation

    processes. They also exhibited significantly enhanced brain uptake and reduced interactions with

    non-targeted tissues in in vivo pharmacokinetic and biodistribution studies [94].

    Although the exact mechanism by which CPPs enter cells remains unclear, two general

    mechanisms, direct penetration and endocytosis, have been described for their internalization.

    Direct penetration occurs through various mechanisms including inverted micelle formation,

    pore formation, the carpet-like model, and the membrane thinning model. The first step in all of

    these pathways is interaction between the positively charged CPP and negatively charged

    elements of the cell membrane (e.g., heparan sulfate and the phospholipid bilayer), which results

    in stable or transient destabilization of the membrane. In contrast, endocytic pathways include

    phagocytosis for large particles and pinocytosis for smaller particles. In pinocytosis, vesicles are

    formed after the inward folding of the outer surface of the cell membrane. Clathrin or caveolin

    pits are involved in receptor-mediated endocytosis [95].

    6. EFFECT OF SHAPE

    In addition to other surface properties, the shape of nanoparticles was found to be useful for their

    interaction with cells, cellular uptake, biodistribution, and subsequent fate [96]. For drug

    delivery, zero-order release was achieved with a hemi-spherical particle that degraded at its face

    [97]. Transferrin-coated rod-shaped nanoparticles were developed with much lower

    internalization efficiency than transferrin-coated spherical nanoparticles. The uptake of the rod-

    shaped nanoparticles was speculated to be more dependent on their width than length [98].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    22

    Internalization kinetics and efficacy can be affected by local surface curvature, which determines

    the interactions between cell and particle surfaces. Spherical gold particles with diameters of 14

    or 75 nm showed 375500% greater uptake in HeLa cells than 74 14 nm rod-shaped particles.

    Chan et al. hypothesized that the differences in particle curvature were responsible for the

    differences in cellular uptake, which was affected by particle contact area with the cell

    membrane [18]. Florez et al. evaluated particle shape and demonstrated that ellipsoidal

    nanoparticles had higher cell surface binding but lower uptake compared to spherical particles.

    The negative correlation between aspect ratio and uptake rate was related to the lower uptake of

    non-spherical nanoparticles, which was further explained by the larger radius of curvature for

    non-spherical particles adsorbed on the cells [99]. A similar effect was observed for polystyrene

    nanoparticles upon change from a three-dimensional spherical particle to a two-dimensional disk,

    which enhanced cell surface binding but reduced cellular uptake. The surface charges of the

    spherical and disk-shaped particles were 13.6 and 16.8 mV, respectively, which suggests the

    role of shape in cellular uptake irrespective of surface charge [100,101].

    The effect of nanoparticle shape and orientation on the pattern of cell penetration was examined.

    Rod-shaped nanoparticles were most rapidly internalized among all shapes when they were

    longitudinally perpendicular to the cell membrane. When the rod was oriented tangentially to the

    cell membrane, the internalization rate significantly decreased. This was explained by greater

    difficulty enclosing the particles in their tangential orientation. In contrast, the penetration of

    spherical nanoparticles was independent of contact angle because of their symmetric shape

    (Figure 5) [102].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    23

    Figure 5. Effect of shape and nanoparticle orientation on cellular uptake.

    Xu et al. reported the effects of the shape of layered double hydroxide (LDH) particles (rods with

    width 3060 nm and length 100200 nm and hexagonal sheets with width 50150 nm and

    thickness 1020 nm) on cellular internalization. Both shapes were readily taken up by CHO-K1,

    NIH 3T3, and HEK 293T cells. The rod-like particles specifically targeted the nucleus, whereas

    the sheet-like particles were retained in the cytoplasm. The uptake of the differently shaped

    particles was primarily associated with strong positive charge. Both types possessed high zeta

    potential (around +40 mV), which interacted with the negatively charged cell membrane

    irrespective of shape. Further investigation revealed that the LDH particles were internalized

    through clathrin-mediated endocytosis [103]. Gratton et al. demonstrated uptake of PRINT-

    fabricated nanoparticles of different sizes and aspect ratios in HeLa cells and found that

    cylindrical nanoparticles had the highest percentage of cellular internalization over time. Cubic

    particles showed similar internalization profiles to those of cylindrical particles. However,

    internalization depended on positive surface charge. When the charge was reversed, uptake of

    particles decreased dramatically [104].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    24

    In blood circulation, nanoparticle motion encounters different forces. Hemodynamic forces are

    fluid resistant forces that oppose the motion of nanoparticles in blood and are proportional to the

    shape of the nanoparticle in a fluid with a moderate to high Reynolds number. An investigation

    of fluid resistant forces demonstrated that the force increased with the radius of spherical

    particles. Additionally, non-spherical particles experience tumbling and rolling because of

    unequal moments resulting from these forces [105,106]. Self-assembled filamentous micelles

    with very high aspect ratios (2260 nm diameter, 28 m length) were found circulate ten times

    longer (~ 1 week) than spherical particles [107]. Polystyrene particles of different geometric

    shapes but otherwise identical properties (dimensions, surface area, volume, and chemistry) were

    fabricated, and their engulfment by macrophages was monitored. The local particle shape at the

    point of initial contact determined their phagocytic fate. Macrophages that attached to elliptical

    disks along the major axis internalized them very quickly (

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    25

    1.6% of the injected disk-shaped particles remained in the blood at 1 and 30 min, respectively

    [110]. Although more investigation is needed to fully determine the role of particle shape in

    clearance, these investigations imply that non-spherical particles can be retained longer in blood

    circulation and suggest a relationship between particle movement in blood and clearance. As the

    spherical particles maintain stable orientations during movement, the complex tumbling and

    rolling motions of non-spherical particles (with high aspect ratios) might interfere with the

    macrophage attachment and phagocytosis.

    From these experiments, we conclude that, in blood circulation, non-spherical particles with high

    aspect ratios are phagocytosed less than spherical particles are. However, spherical particles

    show better performance in entering target cells than do the non-spherical particles.

    7. SURFACE TEXTURE OF NANOCARRIERS

    The surface texture of nanoparticles can significantly affect their cell attachment, interaction,

    uptake, and drug delivery [111,112]. Although smoothness is important for controlled drug

    release, roughness also has beneficial effects. Surface roughness promotes cell binding through

    nonspecific binding forces that enhance cellular uptake. Direct penetration into the cell

    membrane may occur through a combination of nonspecific attractive forces (e.g., roughness and

    charge) on the surface of spiked particles without the need for ligand-receptor interactions [111].

    Niu et al. developed silica nanoparticles with a viral capsid-like rough structure on the surface.

    These particles were very effective in penetrating the membrane, disrupting the endosome, and

    delivering siRNA (Figure 6) [113]. Nanoscale surface roughness greatly minimizes repulsive

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    26

    interactions (e.g., electrostatic and hydrophilic interactions) between cells and nanoparticles,

    thereby promoting adhesion and possibly facilitating engulfment by cells [111].

    Figure 6. Silica nanoparticles with capsid like rough surface and their cellular uptake. (Adapted

    with permission from Y. Niu, M. Yu, S.B. Hartono, J. Yang, H. Xu, H. Zhang, J. Zhang, J. Zou,

    A. Dexter, W. Gu, C. Yu. Nanoparticles mimicking viral surface topography for enhanced

    cellular delivery. Adv. Mater. 25 (2013) 6233-6237. Copyright John Wiley and Sons)

    Particles can be fabricated with different surface textures such as ridges on the sides, with a

    variation in surface areas and surface roughness. These surface properties can be used to alter the

    drug release pattern of nanoparticles or to affect their movement in the bloodstream [114]. The

    performance of aerosols was enhanced by increasing the surface roughness in several studies.

    Particles tend to become cohesive, partially amorphous, and physically unstable in aerosols,

    making powder deagglomeration and precise delivery to the lungs difficult because of their

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    27

    cohesion forces. A small variation in the surface roughness was important for reducing the

    cohesion and performance variability of powdered BSA [115]. Rough surfaces were also useful

    for mucoadhesion of nanoparticles for mucoadhesive drug delivery. Ageros et al. developed

    cyclodextrinpoly(anhydride) nanoparticles for bioadhesion within the gut, where high

    bioadhesive capacity was observed for nanoparticles with a rough surface morphology [116]. In

    contrast, surface smoothness is needed to minimize the variation in drug release and can be an

    important factor in the biodistribution of nanoparticles [112,113]. Proteins and biomolecules tend

    to bind more to the rough surfaces of nanoparticles than to smoother surfaces [113]; therefore,

    nanoparticles with smoother surfaces are more stable in biological systems [24].

    8. STIMULI RESPONSIVE DELIVERY

    Drug release can be triggered by internal and external stimuli. The carrier system can be

    designed with materials to control the release of drug in the targeted area [117]. Materials or

    chemical groups sensitive to various stimuli are conjugated to the surface of nanoparticles and,

    upon exposure to the stimulus, undergo a conformational change, cleavage, or other chemical

    changes that result in drug release [18].

    8.1 SURFACE SWITCHING

    Recently, surface switching has been investigated for stimuli responsive delivery. In this

    technique, surface molecules or chemical groups can toggle activity upon exposure to different

    stimuli. This method includes changing wettability (hydrophilicity-hydrophobicity), surface

    charge, or shape in response to temperature, pH, electric field, light, and ionic strength

    [118,119].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    28

    Yoo et al. reported low-molecular-weight PLGA nanoparticles (160310 nm) that were able to

    switch their shape from elliptical disk to spherical in a stimulus-responsive manner. Elliptical

    disk-shaped particles show prolonged circulation and increased tumor accumulation but exhibit

    slower cellular uptake compared to spherical particles. The disk-shaped particles could become

    spherical in the targeted area for efficient internalization. The shape-switching characteristics

    were achieved by a subtle balance between polymer viscosity and interfacial tension. The PLGA

    molecules possess a carboxylic acid end group, and the interfacial tension of PLGA particles was

    controlled through ionization of these carboxylic acid groups. At physiological pH, the end

    carboxyl groups are charged, causing low hydrophobicity and low interfacial tension. When the

    pH was lowered, carboxylate groups were protonated, thus increasing hydrophobicity and

    inducing the shape switch with an exponential response, where interfacial tension provided the

    main driving force for shape change [120].

    Rotello et al. recently reported surface charge-based DNA uncaging from the surface of

    photolabile gold nanoparticles. The surface was designed with a photocleavable o-nitrobenzyl

    ester linker and a quaternary ammonium salt to switch the surface from cationic to anionic upon

    irradiation. DNA was complexed with the nanoparticles. After irradiation with near-UV light, the

    nitrobenzyl linkage was cleaved to produce an anionic carboxylate group, resulting in DNA

    release. In vitro studies were conducted using a T7 RNA polymerase assay, which revealed

    restored DNA transcription upon UV irradiation [121].

    Lahann et al. developed a smart surface that changes its wettability in response to an electrical

    potential. Gold nanoparticles with a self-assembled 16-mercaptohexadecanoic acid (MHA)

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    29

    monolayer were fabricated to possess a hydrophobic chain capped by a hydrophilic carboxylate

    group. Upon applying the electrical potential, the negatively charged carboxylate groups were

    attracted towards the positive gold surface, resulting in a conformational change of the MHA

    hydrophobic chains. Because of this chain bending, the aliphatic MHA chains were changed

    from an all-trans conformation to a mixture of trans and gauche conformations. This mixed

    conformation increased intermolecular van der Waals contacts and exposed hydrophobic chains

    to the surface, converting the hydrophilic surface to a moderately hydrophobic state [122].

    8.2 ENZYME RESPONSIVE SURFACE DESIGN

    Enzymes have also been targeted for responsive drug delivery. In some diseases, abnormal

    production of enzymes differentiates diseased cells from normal healthy cells. Nanomaterials

    have been designed with appropriate substrates, mostly peptides, attached to the surface of the

    nanoparticle, that are cleaved by the respective enzyme, causing release of the drug [119]. For

    example, matrix metalloproteinase (MMP) is overexpressed by many cancer cells and is needed

    for their metabolism. Specific peptide sequences attached to the surface of nanoparticles are

    cleaved between Leu and Gly residues by MMP [119,123]. We previously reported the

    development of MMP-sensitive gold nanorods using a peptide sequence Gly-Pro-Leu-Gly-Val-

    Arg-Gly-Cys conjugated to a fluorescent probe (Cy5.5) for simultaneous imaging and treatment

    of cancer by hyperthermia. The enzyme-sensitive nanorods were studied in HeLa cells, and the

    Cy5.5 activated by MMP enzymes in the cell culture media was successfully visualized [123].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    30

    9. EFFECT OF SURFACE PROPERTIES ON NANOPARTICLE

    TOXICITY

    The toxicity of nanomedicines has become a major issue impeding their effective application.

    Several toxic effects of nanoparticles such as endocrine disruption and neurotoxicity have been

    reported [124,125]. The abundance of macrophages and accumulation of nanoparticles cause

    organs such as the liver and spleen to be highly affected by toxicity. Organs with high blood flow

    (e.g., kidney and lung) are also prone to nanoparticle-induced toxicity. Inside cells, nanoparticles

    may interact with cellular components, alter cell function, damage DNA, arrest the cell cycle,

    cause mutagenesis, create reactive oxygen species (ROS), and induce apoptosis. As nanoparticles

    have a large surface area, various transition metals, chemicals, and proteins can be attached to

    their surface and produce ROS [126-128]. The main molecular mechanism underlying

    nanotoxicity is free radical formation and induction of oxidative stress by ROS, which can be

    formed by several mechanisms including the phagocytic cell response to nanoparticles, presence

    of transition metals, and environmental factors [126].

    Different nanoparticle surface properties such as charge and functional groups can affect toxicity

    [127]. Surface protection plays a critical role in preventing toxicity by reducing deposition of

    transition metals, chemicals, and proteins and the generation of ROS [129]. Gold nanoparticles

    without surface modifications have shown in vivo toxicity in many studies. However, surface-

    modified gold nanoparticles exhibit lower toxicity over long exposure [130,131].

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    31

    9.1 ROLE OF DIFFERENT SURFACE MOLECULES IN NANOTOXICITY

    Various reports indicate that nanoparticle toxicity is highly dependent on the type of particle,

    surface material and their molecular weights, and particle concentrations [132]. Drug carriers can

    be inorganic functional materials that impart unique physicochemical properties such as

    photothermal activity (gold nanorod), magnetic moment (iron oxide nanoparticles), porous

    structure (mesoporous silica nanoparticles), and fluorescence (quantum dots [QDs]). However,

    inorganic materials usually show toxicity [133]. Similarly, organic materials such as different

    synthetic polymers also show toxicity in vivo. Cationic polymers appear to be especially toxic to

    different organs. Interestingly, lower toxicity was observed for a lower molecular weight cationic

    polymer (PEI) [134]. Low-molecular-weight dextran on nanoparticle surfaces also showed lower

    toxicity than did high-molecular-weight dextran [135]. Furthermore, when toxic polymers are

    linked to hydrophilic polymers such as PEG, the combined amphiphilic polymers can reduce

    toxicity [136].

    Hoshino et al. developed carboxylic acid, polyalcohol, and amine-modified QDs and evaluated

    their cytotoxicity. The cytotoxicity of QDs was related to the surface-conjugated molecules but

    not the nanocrystalline particle itself. Among all the particles, hydroxyl-modified QD was the

    least toxic [137]. Our group developed silica-coated gold nanorods, minimized the size of the

    silica layer to sub-nanometer thickness, and performed cytotoxicity studies. In our investigation,

    a very thin layer of silica effectively reduced the toxicity of gold nanorods with a half-maximal

    inhibitory concentration >400 g/mL. The optothermal efficiency of bare rods was maintained

    even after coating with silica [138]. These results suggest that hydrophilic nanoparticle surfaces

    are important for preventing toxicity due to the protection of surface from binding of different

    molecules.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    32

    Cetyltrimethylammonium bromide (CTAB) is a cationic surfactant commonly used to stabilize

    gold nanorods during synthesis. CTAB is highly toxic, causing mitochondrial damage and

    generating ROS. Displacement of CTAB with safe materials such as PEG and citric acid or

    conjugation with biocompatible hydrophilic materials can significantly reduce gold-induced

    toxicity [129,139]. Wang et al. reported that CTAB -containing nanorods showed no toxicity in

    vivo when conjugated to poly(styrenesulfonate) [129]. Boca et al. examined the toxicity of

    chitosan-capped gold nanoparticles in Chinese hamster ovary cells and found that more than

    85% of the cells were viable even after a long period of exposure [130].

    9.2 SURFACE CHARGE AND GEOMETRY IN NANOTOXICITY

    Positively charged nanoparticles were reported to be toxic in many studies because of their

    ability to destabilize the negatively charged cell membrane through electrostatic interactions

    [127]. Positively charged particles can further damage the outer membrane of mitochondria,

    which is slightly negatively charged and has an electrochemical gradient very similar to that of

    the plasma membrane, which is very sensitive to charge. If positively charged nanoparticles

    accumulate outside the mitochondria, the mitochondrial membrane potential is imbalanced,

    resulting in membrane damage and release of pro-apoptotic proteins into the cytosol [140].

    Goodman et al. fabricated cationic and anionic nanoparticles functionalized with quaternary

    ammonium and carboxylate groups, respectively. Their results demonstrated that cationic

    nanoparticles were moderately toxic, whereas anionic nanoparticles were non-toxic to cells

    [141]. The toxicity of poly(amidoamine) PAMAM dendrimers was evaluated in mice. Positively

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    33

    charged particles showed toxicity, whereas negatively charged particles from the same materials

    were non-toxic [127,142].

    Yu et al. evaluated the toxicity of silica nanoparticles in animal model. Nonporous silica

    nanoparticles (120 nm) exhibited low systemic toxicity and the highest maximum tolerated dose

    (MTD) (450 mg/kg) compared to mesoporous silica nanoparticles (MTDs: 3065 mg/kg).

    However, toxicity decreased when mesoporous silica nanoparticles were modified with primary

    amine groups (MTDs: 100150 mg/kg) [143]. It suggests the protection of large porous surface

    for reducing toxicity. Although the surface was modified with amines, toxicity depends on

    charge density. Yu et al. showed increased cellular association of silica nanoparticles with higher

    zeta potentials (+17 mV versus +32 mV), which they explained by a surface charge threshold

    (> +30 mV) above which the amine functionalities facilitated nanoparticle-cell interaction

    through electrostatic interactions [144]. Adsorption of serum proteins on the nanoparticle surface

    can also decrease the overall charge and toxicity [145].

    Different particle shapes and aspect ratios did not significantly affect particle toxicity. However,

    gold nanorods were shown to be more toxic than their spherical counterparts [145]. Further

    studies revealed that the toxicity was associated with surface CTAB molecules and was

    attenuated when CTAB molecules were replaced by other molecules as discussed above.

    10. PERSPECTIVE

    Although the effects of different surface characteristics have been discovered and various

    methods of surface modification have evolved, control of conjugation at specific sites with

    ordered individual components will require optimization to achieve reproducible and effective

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    34

    surface modifications [146]. Particle aggregation must be carefully considered as it can diminish

    the success of particle engineering [147]. Our current observation suggests that nanomedicine

    will be based on smart surface functionalities and will rely on development of drug carriers as a

    platform for prolonged blood circulation, interaction with specific sites, and accurate drug

    delivery in addition to addressing toxicity concerns during and after drug delivery [146, 148].

    Surface engineering will supplement existing techniques and will be valuable for newer

    applications. Since understanding of the interactions between nanoparticle surface features and

    the biological milieu is increasing, future particle design will enable precise behavioral control of

    nanomedicines, which may be applied as real-time monitoring tools for biodistribution and drug

    delivery [146, 147]. Functionalized moieties on the nanohybrid structure will reform adjuvant

    cancer therapy with selective and precise targeting and destruction of cancer cells, which will

    improve long-term survival.

    Smart surfaces of nanoparticles may be utilized for ultrasensitive biorecognition through the

    interaction of biomarkers and the nanoparticle surface, which will accelerate treatment of

    complicated disorders [146]. As present targeting strategies involve cellular internalization only,

    there is a high probability that surface chemistry can be designed to direct and localize

    nanoparticles to a particular area [149,150]. In addition to intracellular delivery, surface

    functionalization will play a key role for delivery of therapeutics to various sites. As several

    studies have already shown that surface modification of nanoparticles can enhance their cell

    attachment, future formulations for drug delivery to the extracellular matrix, which strongly

    affects cell growth and death, will be based on engineered surfaces. Hence, design of smart

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    35

    surfaces will be the most critical ground to improve cell responses of nanomedicines for

    particular applications.

    Conclusion

    Herein, we analyzed how various functionalities on nanoparticle surfaces affect their desired

    behavior. Several studies have successfully engineered nanoparticle surfaces to overcome the

    limitations of conventional nanoparticles where surface modification was deemed the most

    efficient technique to increase the efficacy of nanomedicines. This review demonstrated how

    surface modification and local orientations of surface molecules are important for the efficiency

    and utility of nanotherapeutics. Based on our current understanding of nanoparticle surfaces, we

    predict that surface functionalization will change and revolutionize the therapeutic applications

    of nanoparticles and treatment patterns of different diseases in the near future.

    Acknowledgment

    This work was supported by the grant of National Research Foundation of Korea (NRF), funded

    by Korean government (MEST) (2012R1A2A2A03046819), and (2013R1A1A2005329).

    REFERENCES

    [1] A.H. Faraji, P. Wipf. Nanoparticles in cellular drug delivery. Bioorg. Med. Chem. 17 (2009)

    2950-2962.

    [2] S.R. Mudshinge, A.B. Deore, S. Patil, C.M. Bhalgat. Nanoparticles: emerging carriers for drug

    delivery. Saudi Pharm. J. 19 (2011) 129141.

    [3] T.C. Yih, M. Al-Fandi. Engineered nanoparticles as precise drug delivery systems. J. Cell.

    Biochem. 97 (2006) 1184-1190.

    [4] R. Singh, J.W. Lillard Jr. Nanoparticle-based targeted drug delivery. Exp. Mol. Pathol. 86 (2009)

    215-223.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    36

    [5] A.A. Manzoor, L.H. Lindner, C.D. Landon, J.Y. Park, A.J. Simnick, M.R. Dreher, S. Das, G.

    Hanna, W. Park, A. Chilkoti, G.A. Koning, T.L. ten Hagen, D. Needham, M.W. Dewhirst.

    Overcoming limitations in nanoparticle drug delivery: triggered, intravascular release to improve

    drug penetration into tumors. Cancer Res. 72 (2012) 5566-5575.

    [6] J.V. Georgieva, D. Kalicharan, P.O. Couraud, I.A. Romero, B. Weksler, D. Hoekstra, I.S.

    Zuhorn. Surface characteristics of nanoparticles determine their intracellular fate in and

    processing by human blood-brain barrier endothelial cells in vitro. Mol. Ther. 19 (2011) 318-325.

    [7] G.W. Bumgarner, R. Shashidharamurthy, S. Nagarajan, M.J. Souza, P. Selvaraj. Surface

    engineering of microparticles by novel protein transfer for targeted antigen/drug delivery. J.

    Control. Release. 137 (2009) 90-97.

    [8] M. Mahmoudi, A. Simchi, M. Imani. Recent advances in surface engineering of

    superparamagnetic iron oxide nanoparticles for biomedical applications. J. Iran Chem. Soc. 7

    (2010) S1-S27.

    [9] N. Kasinathan, H.V. Jagani, A.T. Alex, S.M. Volety, J.V. Rao. Strategies for drug delivery to the

    central nervous system by systemic route. Drug Deliv. (2014) DOI:

    10.3109/10717544.2013.878858.

    [10] L. Jabr-Milane, L. van Vlerken, H. Devalapally, D. Shenoy, S. Komareddy, M. Bhavsar,

    M. Amiji. Multi-functional nanocarriers for targeted delivery of drugs and genes. J. Control.

    Release. 130 (2008) 121-128.

    [11] S.D. Li, L. Huang. Stealth nanoparticles: high density but sheddable PEG is a key for

    tumor targeting. J. Control. Release 145 (2010) 178-181.

    [12] D.E. Owens 3rd, N.A. Peppas. Opsonization, biodistribution, and pharmacokinetics of

    polymeric nanoparticles. Int. J. Pharm. 307 (2006) 93-102.

    [13] T.H. Chung, S.H. Wu, M. Yao, C.W. Lu, Y.S. Lin, Y. Hung, C.Y. Mou, Y.C. Chen, D.M.

    Huang. The effect of surface charge on the uptake and biological function of mesoporous silica

    nanoparticles in 3T3-L1 cells and human mesenchymal stem cells. Biomaterials 28 (2007) 2959-

    2966.

    [14] T. Sen, I.J. Bruce. Surface engineering of nanoparticles in suspension for particle based

    bio-sensing. Scientific Reports 2 (2012) 564. DOI: 10.1038/srep00564

    [15] R. Gref, P. Couvreur, G. Barratt, E. Mysiakine. Surface-engineered nanoparticles for

    multiple ligand coupling. Biomaterials 24 (2003) 4529-4537.

    [16] C.J. Cheng, W.M. Saltzman. Enhanced siRNA delivery into cells by exploiting the

    synergy between targeting ligands and cell-penetrating peptides. Biomaterials 32 (2011) 6194-

    6203.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    37

    [17] R. Cheng, F. Meng, C. Deng, H.A. Klok, Z. Zhong. Dual and multi-stimuli responsive

    polymeric nanoparticles for programmed site-specific drug delivery. Biomaterials 34 (2013)

    3647-3657.

    [18] M. Caldorera-Moore, N. Guimard, L. Shi, K. Roy. Designer nanoparticles: incorporating

    size, shape and triggered release into nanoscale drug carriers. Expert Opin. Drug Deliv. 7 (2010)

    479-495.

    [19] S. Mura, H. Hillaireau, J. Nicolas, B. Le Droumaguet, C. Gueutin, S. Zanna, N. Tsapis, E.

    Fattal. Influence of surface charge on the potential toxicity of PLGA nanoparticles towards Calu-

    3 cells. Int. J. Nanomedicine 6 (2011) 2591-2605.

    [20] J.P. Ryman-Rasmussen, J.E. Riviere, N.A. Monteiro-Riviere. Surface coatings determine

    cytotoxicity and irritation potential of quantum dot nanoparticles in epidermal keratinocytes. J.

    Invest. Dermatol. 127 (2007) 143-153.

    [21] S. Yu, G. Wu, X. Gu, J. Wang, Y. Wang, H. Gao, J. Ma. Magnetic and pH-sensitive

    nanoparticles for antitumor drug delivery. Colloids Surf. B Biointerfaces 103 (2013) 15-22.

    [22] K. Ulbrich, T. Hekmatara, E. Herbert, J. Kreuter. Transferrin- and transferrin-receptor-

    antibody-modified nanoparticles enable drug delivery across the blood-brain barrier (BBB). Eur.

    J. Pharm. Biopharm. 71 (2009) 251-256.

    [23] R. Mout, D.F. Moyano, S. Rana, V.M. Rotello. Surface functionalization of nanoparticles

    for nanomedicine. Chem. Soc. Rev. 41 (2012) 2539-2544.

    [24] Y. Sheng, C. Liu, Y. Yuan, X. Tao, F. Yang, X. Shan, H. Zhou, F. Xu. Long-circulating

    polymeric nanoparticles bearing a combinatorial coating of PEG and water-soluble chitosan.

    Biomaterials 30 (2009) 2340-2348.

    [25] H. Wang, P. Zhao, X. Liang, X. Gong, T. Song, R. Niu, J. Chang. Folate-PEG coated

    cationic modified chitosan--cholesterol liposomes for tumor-targeted drug delivery. Biomaterials

    31 (2010) 4129-4138.

    [26] M.L. Hans, A.M. Lowman. Biodegradable nanoparticles for drug delivery and targeting.

    Curr. Opin. Sol. St. Mater. Sci. 6 (2002) 319327.

    [27] M. Elsabahy, K.L. Wooley. Design of polymeric nanoparticles for biomedical delivery

    applications. Chem. Soc. Rev. 41 (2012) 2545-2561.

    [28] S. Runa, A. Hill, V.L. Cochran, C.K. Payne. PEGylated nanoparticles: protein corona and

    secondary structure. Physical Chemistry of Interfaces and Nanomaterials XIII 9165 (2014)

    91651F1-91651F-8.

    [29] J.V. Jokerst, T. Lobovkina, R.N. Zare, S.S. Gambhir. Nanoparticle PEGylation for

    imaging and therapy. Nanomedicine (Lond) 6 (2011) 715-728.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    38

    [30] P.L. Hansen, J.A. Cohen, R. Podgornik, V.A. Parsegian. Osmotic properties of

    poly(ethylene glycols): quantitative features of brush and bulk scaling laws. Biophys. J. 84 (2003)

    350-355.

    [31] J. Yang, A. Vitale, R. Bongiovanni, J. Nie. Synthesis and characterization of siloxane

    photopolymers used for microfluidic devices. DOI: 10.1039/C4NJ01773K

    [32] S. Biswas, V.P. Torchilin. Nanopreparations for organelle-specific delivery in cancer.

    Advanced Drug Delivery Reviews 66C (2013) 26-41.

    [33] J.V. Jokerst, T. Lobovkina, R.N. Zare, S.S. Gambhir. Nanoparticle PEGylation for

    imaging and therapy. Nanomedicine (Lond) 6 (2011) 715-728.

    [34] R. Weissleder, A. Bogdanov, E.A. Neuwelt, M. Papisov. Long-circulating iron oxides for

    MR imaging. Advanced Drug Delivery Reviews 16 (1995) 321334.

    [35] Y.Y. Yuan, C.Q. Mao, X.J. Du, J.Z. Du, F. Wang, J. Wang. Surface charge switchable

    nanoparticles based on zwitterionic polymer for enhanced drug delivery to tumor. Adv. Mater. 24

    (2012) 5476-5480.

    [36] J.E. Rosen, F.X. Gu. Surface functionalization of silica nanoparticles with cysteine: a

    low-fouling zwitterionic surface. Langmuir 27 (2011) 10507-10513.

    [37] H. Lee, D.Y. Yi. A study of Gd loaded silica nanoparticles for both optical and magnetic

    resonance imaging of cells. Materials Letters 100 (2013) 98-101.

    [38] S.K. Sahoo, J. Panyam, S. Prabha, V. Labhasetwar. Residual polyvinyl alcohol associated

    with poly (D,L-lactide-co-glycolide) nanoparticles affects their physical properties and cellular

    uptake. J. Control. Release 82 (2002) 105-114.

    [39] L. Zhang, F.X. Gu, J.M. Chan, A.Z. Wang, R.S. Langer, O.C. Farokhzad. Nanoparticles

    in medicine: therapeutic applications and developments. Clin. Pharmacol. Ther. 83 (2008) 761-

    769.

    [40] S. Chen, X.Z. Zhang, S.X. Cheng, R.X. Zhuo, Z.W. Gu. Functionalized amphiphilic

    hyperbranched polymers for targeted drug delivery. Biomacromolecules 9 (2008) 2578-2585.

    [41] Y. Zhang, R.X. Zhuo. Synthesis and in vitro drug release behavior of amphiphilic

    triblock copolymer nanoparticles based on poly (ethylene glycol) and polycaprolactone.

    Biomaterials 26 (2005) 6736-6742.

    [42] Y. Zhang, M. Yang, N.G. Portney, D. Cui, G. Budak, E. Ozbay, M. Ozkan, C.S. Ozkan.

    Zeta potential: a surface electrical characteristic to probe the interaction of nanoparticles with

    normal and cancer human breast epithelial cells. Biomed. Microdevices 10 (2008) 321-328.

    [43] A. Verma, F. Stellacci. Effect of surface properties on nanoparticlecell interactions.

    Small 6 (2010) 1221.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    39

    [44] J. Ramos, J. Forcada, R. Hidalgo-Alvarez. Cationic polymer nanoparticles and nanogels:

    from synthesis to biotechnological applications. Chem. Rev. 114 (2014) 367428.

    [45] D. Hhn, K. Kantner, C. Geidel, S. Brandholt, I. De Cock, S.J. Soenen, P. Rivera Gil,

    J.M. Montenegro, K. Braeckmans, K. Mllen, G.U. Nienhaus, M. Klapper, W.J. Parak. Polymer-

    coated nanoparticles interacting with proteins and cells: focusing on the sign of the net charge.

    ACS Nano 7 (2013) 3253-3263.

    [46] M. Bivas-Benita, S. Romeijn, H.E. Junginger, G. Borchard. PLGA-PEI nanoparticles for

    gene delivery to pulmonary epithelium. Eur. J. Pharm. Biopharm. 58 (2004) 1-6.

    [47] X. Shi, T.P. Thomas, L.A. Myc, A. Kotlyar, J.R. Baker, Jr. Synthesis, characterization,

    and intracellular uptake of carboxyl-terminated poly(amidoamine) dendrimer-stabilized iron

    oxide nanoparticles. Phys. Chem. Chem. Phys. 9 (2007) 5712-5720.

    [48] H. Hillaireau, P. Couvreur. Nanocarriers' entry into the cell: relevance to drug delivery.

    Cell. Mol. Life Sci. 66 (2009) 2873-2896.

    [49] C.C. Fleischer, C.K. Payne. Nanoparticle surface charge mediates the cellular receptors

    used by protein-nanoparticle complexes. The Journal of Physical Chemistry B 116 (2012) 8901-

    8907.

    [50] P.R. Leroueil, S.A. Berry, K. Duthie, G. Han, V.M. Rotello, D.Q. McNerny, J.R. Baker

    Jr., B.G. Orr, M.M. Holl. Wide varieties of cationic nanoparticles induce defects in supported

    lipid bilayers. Nano Lett. 8 (2008) 420-424.

    [51] D. Guarnieri, A. Guaccio, S. Fusco, P.A. Netti. Effect of serum proteins on polystyrene

    nanoparticle uptake and intracellular trafficking in endothelial cells. J. Nanopart. Res.13 (2011)

    42954309.

    [52] X. Wu, K. Landfester, A. Musyanovych, R.H. Guy. Disposition of charged nanoparticles

    after their topical application to the skin. Skin Pharmacol. Physiol. 23(2010) 117-123.

    [53] S. Shanmugam, C.K. Song, S. Nagayya-Sriraman, R. Baskaran, C.S. Yong, H.G. Choi,

    D.D. Kim, J.S. Woo, B.K. Yoo. Physicochemical characterization and skin permeation of

    liposome formulations containing clindamycin phosphate. Arch. Pharm. Res. 32 (2009) 1067

    1075.

    [54] S. Patil, A. Sandberg, E. Heckert, W. Self, S. Seal. Protein adsorption and cellular uptake

    of cerium oxide nanoparticles as a function of zeta potential. Biomaterials 28 (2007) 4600-4607.

    [55] L.J. Mortensen, G. Oberdrster, A.P. Pentland, L.A. Delouise. In vivo skin penetration of

    quantum dot nanoparticles in the murine model: the effect of UVR. Nano Lett. 8 (2008) 2779-

    2787.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    40

    [56] N.L. Rosi, D.A. Giljohann, C.S. Thaxton, A.K. Lytton-Jean, M.S. Han, C.A. Mirkin.

    Oligonucleotide-modified gold nanoparticles for intracellular gene regulation. Science 312 (2006)

    1027-1030.

    [57] J.A. Khan, B. Pillai, T.K. Das, Y. Singh, S. Maiti. molecular effects of uptake of gold

    nanoparticles in HeLa cells. Chembiochem. 8 (2007) 1237-1240.

    [58] O. Harush-Frenkel, N. Debotton, S. Benita, Y. Altschuler. Targeting of nanoparticles to

    the clathrin-mediated endocytic pathway. Biochem. Biophys. Res. Commun. 353 (2007) 26-32.

    [59] Y. Jallouli, A. Paillard, J. Chang, E. Sevin, D. Betbeder. Influence of surface charge and

    inner composition of porous nanoparticles to cross blood-brain barrier in vitro. Int. J. Pharm. 344

    (2007) 103-109.

    [60] P.R. Lockman, J.M. Koziara, R.J. Mumper, D.D. Allen. Nanoparticle surface charges

    alter blood-brain barrier integrity and permeability. J. Drug. Target. 12 (2004) 635-641.

    [61] Y. Zhong, F. Meng, C. Deng, Z. Zhong. Ligand-directed active tumor-targeting

    polymeric nanoparticles for cancer chemotherapy. Biomacromolecules 15 (2014) 1955-1969.

    [62] J.D. Byrne, T. Betancourt, L. Brannon-Peppas. Active targeting schemes for nanoparticle

    systems in cancer therapeutics. Adv. Drug. Deliv. Rev. 60 (2008) 1615-1626.

    [63] S.H. Kim, J.H. Jeong, K.W. Chun, T.G. Park. Target-specific cellular uptake of PLGA

    nanoparticles coated with poly(L-lysine)-poly(ethylene glycol)-folate conjugate. Langmuir 21

    (2005) 8852-8857.

    [64] F. Sonvico, S. Mornet, S. Vasseur, C. Dubernet, D. Jaillard, J. Degrouard, J. Hoebeke, E.

    Duguet, P. Colombo, P. Couvreur. Folate-conjugated iron oxide nanoparticles for solid tumor

    targeting as potential specific magnetic hyperthermia mediators: synthesis, physicochemical

    characterization, and in vitro experiments. Bioconjug. Chem. 16 (2005) 1181-8.

    [65] S.K. Sahoo, W. Ma, V. Labhasetwar. Efficacy of transferrin-conjugated paclitaxel-loaded

    nanoparticles in a murine model of prostate cancer. Int. J. Cancer 112 (2004) 335-340.

    [66] D.N. Heo, D.H. Yang, H.J. Moon, J.B. Lee, M.S. Bae, S.C. Lee, W.J. Lee, I.C. Sun, I.K.

    Kwon. Gold nanoparticles surface-functionalized with paclitaxel drug and biotin receptor as

    theranostic agents for cancer therapy. Biomaterials 33 (2012) 856-866.

    [67] X. Tian X, H. Yin, S. Zhang, Y. Luo, K. Xu, P. Ma, C. Sui, F. Meng, Y. Liu, Y. Jiang, J.

    Fang. Bufalin loaded biotinylated chitosan nanoparticles: an efficient drug delivery system for

    targeted chemotherapy against breast carcinoma. Eur. J. Pharm. Biopharm. 87 (2014) 445-453.

    [68] H. Chen, J. Gao, Y. Lu, G. Kou, H. Zhang, L. Fan, Z. Sun, Y. Guo, Y. Zhong.

    Preparation and characterization of PE38KDEL-loaded anti-HER2 nanoparticles for targeted

    cancer therapy. J. Control. Release 128 (2008) 209-216.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    41

    [69] S. Ji, J. Xu, B. Zhang, W. Yao, W. Xu, W. Wu, Y. Xu, H. Wang, Q. Ni, H. Hou, X. Yu.

    RGD-conjugated albumin nanoparticles as a novel delivery vehicle in pancreatic cancer therapy.

    Cancer Biol. Ther. 13 (2012) 206-215.

    [70] J.M. Bergen, H.A. von Recum, T.T. Goodman, A.P. Massey, S.H. Pun. Gold

    nanoparticles as a versatile platform for optimizing physicochemical parameters for targeted drug

    delivery. Macromol. Biosci. 6 (2006) 506-516.

    [71] M. Wang, M. Thanou. Targeting nanoparticles to cancer. Pharmacol. Res. 62 (2010) 90-

    99.

    [72] F. Gu, L. Zhang, B.A. Teply, N. Mann, A. Wang, A.F. Radovic-Moreno, R. Langer, O.C.

    Farokhzad. Precise engineering of targeted nanoparticles by using self-assembled biointegrated

    block copolymers. PNAS. 105 (2008) 2586-2591.

    [73] D.B. Kirpotin, D.C. Drummond, Y. Shao, M.R. Shalaby, K. Hong, U.B. Nielsen, J.D.

    Marks, C.C. Benz, J.W. Park. Antibody targeting of long-circulating lipidic nanoparticles does

    not increase tumor localization but does increase internalization in animal models. Cancer Res. 66

    (2006) 6732-6740.

    [74] A. Salvati, A.S. Pitek, M.P. Monopoli, K. Prapainop, F.B. Bombelli, D.R. Hristov,

    P.M. Kelly, C. berg, E. Mahon, K.A. Dawson. Transferrin-functionalized

    nanoparticles lose their targeting capabilities when a biomolecule corona adsorbs on the surface.

    Nature Nanotechnology 8 (2013) 137143.

    [75] A. Albanese, P.S. Tang, W.C. Chan. The effect of nanoparticle size, shape, and surface

    chemistry on biological systems. Annual review of biomedical engineering 14 (2012) 1-16.

    [76] E. Mahon, A. Salvati, F.B. Bombelli, I. Lynch, K.A. Dawson. Designing the

    nanoparticlebiomolecule interface for targeting and therapeutic delivery. Journal of Controlled

    Release 161 (2012) 164-174.

    [77] K.M. McNeeley, A. Annapragada, R.V. Bellamkonda. Decreased circulation time offsets

    increased efficacy of PEGylated nanocarriers targeting folate receptors of glioma.

    Nanotechnology 18 (2007) 385101.

    [78] H. Lee, A.K. Lytton-Jean, Y. Chen, K.T. Love, A.I. Park, E.D. Karagiannis, A. Sehgal,

    W. Querbes, C.S. Zurenko, M. Jayaraman, C.G. Peng, K. Charisse, A. Borodovsky, M.

    Manoharan, J.S. Donahoe, J. Truelove, M. Nahrendorf, R. Langer, D.G. Anderson. Molecularly

    self-assembled nucleic acid nanoparticles for targeted in vivo siRNA delivery. Nat. Nanotechnol.

    7 (2012) 389-393.

    [79] D.R. Elias, A. Poloukhtine, V. Popik, A. Tsourkas. Effect of ligand density, receptor

    density, and nanoparticle size on cell targeting. Nanomedicine 9 (2013) 194-201.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    42

    [80] H. Yuan, J. Li, G. Bao, S. Zhang. Variable nanoparticle-cell adhesion strength regulates

    cellular uptake. Phys. Rev. Lett. 105 (2010) 138101.

    [81] J.Y. Yhee, S. Lee, K. Kim. Advances in targeting strategies for nanoparticles in cancer

    imaging and therapy. Nanoscale 6 (2014) 13383-13390.

    [82] G. von Maltzahn, J.H. Park, K.Y. Lin, N. Singh, C. Schwppe, R. Mesters, W.E. Berdel,

    E. Ruoslahti, M.J. Sailor, S.N. Bhatia. Nanoparticles that communicate in vivo to amplify tumour

    targeting. Nat. Mater. 10 (2011) 545-552.

    [83] P.V. Chang, D.H. Dube, E.M. Sletten, C.R. Bertozzi. A strategy for the selective imaging

    of glycans using caged metabolic precursors. J. Am. Chem. Soc. 132 (2010) 9516-9518.

    [84] S. Lee, H. Koo, J.H. Na, S.J. Han, H.S. Min, S.J. Lee, S.H. Kim, S.H. Yun, S.Y. Jeong,

    I.C. Kwon, K. Choi, K. Kim. Chemical tumor-targeting of nanoparticles based on metabolic

    glycoengineering and click chemistry. ACS Nano 8 (2014) 20482063.

    [85] K.M. Stewart, K.L. Horton, S.O. Kelley. Cell-penetrating peptides as delivery vehicles

    for biology and medicine. Org. Biomol. Chem. 6 (2008) 2242-2255.

    [86] W.L. Munyendo, H. Lv, H. Benza-Ingoula, L.D. Baraza, J. Zhou. Cell penetrating

    peptides in the delivery of biopharmaceuticals. Biomolecules 2 (2012) 187-202.

    [87] A.A. Kale, V.P. Torchilin. Enhanced transfection of tumor cells in vivo using Smart

    pH-sensitive TAT-modified pegylated liposomes. J. Drug. Target. 15 (2007) 538-545.

    [88] P. Zhao, H. Wang, M. Yu, S. Cao, F. Zhang, J. Chang, R. Niu. Paclitaxel-loaded, folic-

    acid-targeted and TAT-peptide-conjugated polymeric liposomes: in vitro and in vivo evaluation.

    Pharm Res 27 (2010) 1914-1926.

    [89] F. Heitz, M.C. Morris, G. Divita. Twenty years of cell-penetrating peptides: from

    molecular mechanisms to therapeutics. Br. J. Pharmacol. 157 (2009) 195-206.

    [90] R.M. Sawant, J.P. Hurley, S. Salmaso, A. Kale, E. Tolcheva, T.S. Levchenko, V.P.

    Torchilin. "SMART" drug delivery systems: double-targeted pH-responsive pharmaceutical

    nanocarriers. Bioconjug. Chem. 17 (2006) 943-949.

    [91] S. Trabulo, A.L. Cardoso, M. Mano, M.C.P.D. Lima. Cell-penetrating peptides

    mechanisms of cellular uptake and generation of delivery systems. Pharmaceuticals (Basel) 3

    (2010) 961993.

    [92] N. Todorova, C. Chiappini, M. Mager, B. Simona, I.I. Patel, M.M. Stevens, I. Yarovsky.

    Surface presentation of functional peptides in solution determines cell internalization efficiency

    of tat conjugated nanoparticles. Nano Lett. 14 (2014) 5229-5237.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    43

    [93] S. Petersen, A. Barchanski, U. Taylor, S. Klein, D. Rath, S. Barcikowski. Penetratin-

    conjugated gold nanoparticles design of cell-penetrating nanomarkers by femtosecond laser

    ablation. J. Phys. Chem. C 115 (2011) 51525159.

    [94] H. Xia, X. Gao, G. Gu, Z. Liu, Q. Hu, Y. Tu, Q. Song, L. Yao, Z. Pang, X. Jiang, J.

    Chen, H. Chen. Penetratin-functionalized PEG-PLA nanoparticles for brain drug delivery. Int. J.

    Pharm. 436 (2012) 840-580.

    [95] F. Madani, S. Lindberg, U. Langel, S. Futaki, A. Grslund. Mechanisms of cellular

    uptake of cell-penetrating peptides. J. Biophys. (2011) 414729.

    [96] J.A. Champion, Y.K. Katare, S. Mitragotri. Particle shape: a new design parameter for

    micro- and nanoscale drug delivery carriers. J. Control. Release 121 (2007) 3-9.

    [97] D.S. Hsieh, W.D. Rhine, R. Langer. Zero-order controlled-release polymer matrices for

    micro- and macromolecules. J. Pharm. Sci. 72 (1983) 17-22.

    [98] B.D. Chithrani, W.C. Chan. Elucidating the mechanism of cellular uptake and removal of

    protein-coated gold nanoparticles of different sizes and shapes. Nano Lett. 7 (2007) 1542-1550.

    [99] L. Florez, C. Herrmann, J.M. Cramer, C.P. Hauser, K. Koynov, K. Landfester, D. Crespy,

    V. Mailnder. How shape influences uptake: interactions of anisotropic polymer nanoparticles

    and human mesenchymal stem cells. Small 8 (2012) 2222-2230.

    [100] Y. Zhang, S. Tekobo, Y. Tu, Q. Zhou, X. Jin, S.A. Dergunov, E. Pinkhassik, B. Yan.

    Permission to enter cell by shape: nanodisk vs nanosphere. ACS Appl. Mater. Interfaces 4 (2012)

    4099-4105.

    [101] S. Dasgupta, T. Auth, G. Gompper. Shape and orientation matter for the cellular uptake

    of nonspherical particles. Nano Lett. 14 (2014) 687-693.

    [102] R. Toy, P.M. Peiris, K.B. Ghaghada, E. Karathanasis. Shaping cancer nanomedicine: the

    effect of particle shape on the in vivo journey of nanoparticles. Nanomedicine (Lond) 9 (2014)

    121-134.

    [103] Z.P. Xu, M. Niebert, K. Porazik, T.L. Walker, H.M. Cooper, A.P. Middelberg, P.P. Gray,

    P.F. Bartlett, G.Q. Lu. Subcellular compartment targeting of layered double hydroxide

    nanoparticles. J. Control. Release 130 (2008) 8694.

    [104] S.E. Gratton, P.A. Ropp, P.D. Pohlhaus, J.C. Luft, V.J. Madden, M.E. Napier, J.M.

    DeSimone. The effect of particle design on cellular internalization pathways. Proc. Natl. Acad.

    Sci. U S A. 105 (2008) 11613-11618.

    [105] P. Decuzzi, R. Pasqualini, W. Arap, M. Ferrari. Intravascular delivery of particulate

    systems: does geometry really matter?. Pharmaceutical research 26 (2009) 235-243.

  • ACCE

    PTED

    MAN

    USCR

    IPT

    ACCEPTED MANUSCRIPT

    44

    [106] A.J. Goldman, R.G. Cox, H. Brenner. Slow viscous motion of a sphere parallel to a plane

    wall--II Couette flow. Chemical Engineering Science 22 (1967) 653660.

    [107] Y. Geng, P. Dalhaimer, S. Cai, R. Tsai, M. Tewari, T. Minko, D.E. Discher. Shape

    effects of filaments versus spherical particles in flow and drug delivery. Nat. Nano 2 (2007) 249

    255.

    [108] J.A. Champion, S. Mitragotri. Role of target geometry in phagocytosis. Proc. Natl. Acad.

    Sci. U S A. 103 (2006) 4930-4934.

    [109] X. Huang, L. Li, T. Liu, N. Hao, H. Liu, D. Chen, F. Tang. The shape effect of

    mesoporous silica nanoparticles on biodistribution, clearance, and biocompatibility in vivo. ACS

    Nano 5 (2011) 5390-5399.

    [110] S. Muro, C. Garnacho, J.A. Champion, J. Leferovich, C. Gajewski, E.H. Schuchman, S.

    Mitragotri, V.R. Muzykantov. Control of endothelial targeting and intracellular delivery of

    therapeutic enzymes by modulating the size and shape of ICAM-1-targeted carriers. Mol. Ther.

    16 (2008) 1450-1458.

    [111] A.E. Nel, L. Mdler, D. Velegol, T. Xia, E.M.V. Hoek, P. Somasundaran, F. Klaessig, V.

    Castranova, M. Thompson. Understanding biophysicochemical interactions at the nanobio

    interface. Nature Materials 8 (2009) 543 557.

    [112] Z. Zhang, S.S. Feng. Nanoparticles of pol


Recommended