Hybrid mesoporous silica nanoparticles · 2.3. Synthesis of RAFT Agent Functionalized MSNs...

10
Hybrid mesoporous silica nanoparticles Elizabete Coutinho e-mail: [email protected] CQFM Centro de Química-Física Molecular, Instituto Superior Técnico, 1049-001 Lisboa, Portugal. ABSTRACT: A decrease in the side effects of a drug can be obtained if it is delivered in an effective and timely manner to the needed location. Therefore, systems are needed to improve the transport of the drug and the location of the target cells, resulting in an increase in drug efficiency and decreased toxicity to healthy tissues. Nanoparticles have gained much importance in medicine due to the fact that they have high area to volume ratio and, after appropriate functionalization, the ability to recognize target sites. In this work, hybrid spherical mesoporous silica nanoparticles (MSNs) were synthesized with a core-shell design. The mesoporous silica core has a well-defined morphology with pores that can incorporate molecules. The MSNs are characterized by ordered pores with diameters arround 2-2.5 nm and volumes above 1 mL / g. RAFT polymerization (Reversible Addition- Fragmentation chain Transfer), associated with the grafting-from method, were used to grow a biocompatible polymer shell with temperature conformation dependence. This conformational change from extended to collapsed can be used to control the release of molecules incorporated in the pores, providing a convenient platform for controlled release of drugs. The MSNs have diameters between 140 and 220 nm. We prove the successful release of molecules from the MSNs pores through thermal stimulation of the polymeric shell. These nanoparticles open new pathways in the development of controlled release systems. KEYWORDS: mesoporous silica nanoparticles, core-shell nanoparticles, controlled release, functionalization, RAFT. 1. INTRODUCTION The hydrophobic nature of most chemotherapeutic agents makes them poorly soluble in water and therefore limits the administration in higher doses. Thus, systems that improve the transport of the drug and its delivery to the target cells are required, resulting in enhanced drug efficacy and decreased toxicity to healthy tissues.These ideas continue to be a priority in cancer therapy. Mesoporous silica nanoparticles (MSNs) have a mesoporous structure with well-defined pore diameters between 2 and 4 nm, and volume exceeding 1 mL/g pore. 1 In this structure, the mesopores are aligned and appear in form of honeycombs making it very interesting and important for many applications. The pores act as individual reservoirs without links between them. These internal structures can be controlled by the initial reagents or the surfactant used. The mesoporous silica nanoparticles have diameters of 40 to hundreds of nanometers and are a biocompatible three-dimensional

Transcript of Hybrid mesoporous silica nanoparticles · 2.3. Synthesis of RAFT Agent Functionalized MSNs...

  • Hybrid mesoporous silica nanoparticles

    Elizabete Coutinho

    e-mail: [email protected]

    CQFM – Centro de Química-Física Molecular, Instituto Superior Técnico, 1049-001 Lisboa, Portugal.

    ABSTRACT: A decrease in the side effects of a drug can be obtained if it is delivered in

    an effective and timely manner to the needed location. Therefore, systems are needed to

    improve the transport of the drug and the location of the target cells, resulting in an increase in

    drug efficiency and decreased toxicity to healthy tissues. Nanoparticles have gained much

    importance in medicine due to the fact that they have high area to volume ratio and, after

    appropriate functionalization, the ability to recognize target sites.

    In this work, hybrid spherical mesoporous silica nanoparticles (MSNs) were synthesized

    with a core-shell design. The mesoporous silica core has a well-defined morphology with pores

    that can incorporate molecules. The MSNs are characterized by ordered pores with diameters

    arround 2-2.5 nm and volumes above 1 mL / g. RAFT polymerization (Reversible Addition-

    Fragmentation chain Transfer), associated with the grafting-from method, were used to grow a

    biocompatible polymer shell with temperature conformation dependence. This conformational

    change from extended to collapsed can be used to control the release of molecules

    incorporated in the pores, providing a convenient platform for controlled release of drugs. The

    MSNs have diameters between 140 and 220 nm. We prove the successful release of molecules

    from the MSNs pores through thermal stimulation of the polymeric shell. These nanoparticles

    open new pathways in the development of controlled release systems.

    KEYWORDS: mesoporous silica nanoparticles, core-shell nanoparticles, controlled

    release, functionalization, RAFT.

    1. INTRODUCTION

    The hydrophobic nature of most

    chemotherapeutic agents makes them poorly

    soluble in water and therefore limits the

    administration in higher doses. Thus, systems

    that improve the transport of the drug and its

    delivery to the target cells are required,

    resulting in enhanced drug efficacy and

    decreased toxicity to healthy tissues.These

    ideas continue to be a priority in cancer

    therapy. Mesoporous silica nanoparticles

    (MSNs) have a mesoporous structure with

    well-defined pore diameters between 2 and 4

    nm, and volume exceeding 1 mL/g pore.1 In

    this structure, the mesopores are aligned and

    appear in form of honeycombs making it very

    interesting and important for many

    applications. The pores act as individual

    reservoirs without links between them. These

    internal structures can be controlled by the

    initial reagents or the surfactant used. The

    mesoporous silica nanoparticles have

    diameters of 40 to hundreds of nanometers

    and are a biocompatible three-dimensional

  • network with silanol groups (≡Si-OH) on the

    surface.2

    The main objective of this work is to

    develop hybrid MSNs with a thermoresponsive

    polymer shell to obtain a smart release control

    mechanism, able to accommodate large drug

    loads and deliver their cargo on demand to a

    desired location.3,4,5

    The hybrid MSNs were

    prepared by incorporating a perylenediimide

    (PDI) derivative in the silica network (MSN-PDI)

    for bioimaging. The surface of MSNs were

    funcionalized with (3-aminopropyl)

    triethoxysilane (APTES). The RAFT

    polymerization method associated with the

    grafting from method was used to growth the

    polymeric shell.6,7

    A fluorescent probe,

    Sulforhodamine B (SRB), was incorporated into

    the hybrid MSNs in order to study the controlled

    release using fluorescence techniques.

    2. MATERIALS AND METHODS

    2.1. Materials

    Absolute ethanol (EtOH, 99,9% Scharlau),

    hydrochloric acid (HCl, 37% Panreac),

    ammonium hydroxide solution (NH4OH, 25%

    Fluka), tetraethyl orthosilicate (TEOS, 98%

    Aldrich), N-cetyltrimethylammonium bromide

    (CTAB, 99% Sigma), (3-aminopropyl)

    triethoxysilane (APTES, 98% Sigma-Aldrich), N-

    (3-dimethylaminopropyl) carbodiimide (EDC,

    98% Sigma-Aldrich) were used without further

    purification. The RAFT agent, 3-

    (benzylsulfanylthiocarbonylsulfanyl) propionic

    acid was used for RAFT polimerization, 2-(2-

    methoxyethoxy)ethyl methacrylate (MEO2MA,

    95% Sigma-Aldrich, Mn=188,22 g mol-1

    ) and

    poly(ethylene glycol) methyl ether methacrylate

    (OEGMA, 98% Sigma-Aldrich, Mn=475 g mol-1

    ),

    2,2-azobis (2-methylpropionitrile) (AIBN, 99%

    Sigma-Aldrich) were used without further

    purification. Commercial toluene and

    dichloromethane were distilled over calcium

    hydride prior to use. The Perylenediimide

    derivative was synthetized according to the

    literature.8

    2.2. Synthesis of the MSNs

    The perylenediimide-functionalized mesoporous

    silica nanoparticles (MSN-PDI) were prepared

    by a sol-gel modified process.9 In a

    polypropylene flask CTAB (0.113 g) was

    dissolved in 58.7 mL of 0.5 M NH4OH at 50ºC.

    2.5 mL of a TEOS solution (0.2 M in ethanol),

    0.5 mL of a PDI solution (6 mg PDI and 15 mL

    EtOH) and 9.2 mL of absolute ethanol were

    added to the surfactant solution dropwise while

    stirring. The stirring was continued for 5 h. TEOS

    solution (2.5 mL, 0.2 M in ethanol) and PDI

    solution (0.5 mL) were added in the mixture and

    kept stirring for 1 h. The mixture was then aged

    at 50º C for 24 h. The nanoparticles were

    recovered by centrifugation at 19118 g for 20

    min each cycle. The MSNs were washed 2 times

    with ethanol/water and 3 times with ethanol. The

    MSNs were dried at 50ºC. The amino

    modification of the silica surface was performed

    by dispersing the obtained nanoparticles in a

    solution of APTES in dry toluene. The ratio of

    the reaction mixture used was 0.2 g MSNs: 10

    mL toluene: 0.468 mL APTES. The reaction

    occurs at 125ºC under reflux and argon

    atmosphere for 24 h. Finally, the nanoparticles

    were centrifuged for three cycles (19118 g, 20

    min), redispersed in absolute ethanol and finally

    dried overnight at 50º C. To remove the

    surfactant, the nanoparticles were placed in a

    polypropylene flask with an acidic ethanol

    solution (0.5 M HCl, 20 mL of this solution for

    each 500 mg of particles) and stirred for 2 h at

    40º C. The nanoparticles were centrifuged three

  • times and redispersed in ethanol (19118 g, 10

    min each cycle) and subsequently dried

    overnight at 50° C.

    2.3. Synthesis of RAFT Agent Functionalized

    MSNs (MSN-RAFT)

    To immobilize the RAFT agent on the

    nanoparticles surface, the EDC and dry

    dichloromethane were added to the MSN-PDI-

    APTES and RAFT agent. The mixture was

    stirred under an argon atmosphere at room

    temperature for 24 hours. The nanoparticles

    functionalized with RAFT agent (MSN-PDI-

    RAFT) were recovered by centrifugation in 3

    cycles (19118 g, 20 min each cycle), and

    redispersed in absolute ethanol.

    2.4. Synthesis of MSN-Poly

    The polymer shell was synthesized by

    introducing MSN-PDI-RAFT, OEGMA, MEO2MA

    and ethanol in the schlenk tube. The mixture

    was placed in an ultrasonic bath for 30 minutes

    and stirred for 45 minutes at room temperature

    under argon atmosphere. A solution of AIBN (~8

    mg, 10 mL ethanol) was stirred for 45 min at

    room temperature under argon atmosphere.

    After, the schlenk tube was placed in an oil bath

    at 70° C and 30 µl of the AIBN solution was

    added to the mixture. The reaction mixture was

    maintained under argon atmosphere for 24

    hours. At the end of the reaction, was

    centrifuged in 3 cycles (8497 g, 10 min each

    cycle) and redispersed in absolute ethanol.

    2.5. Loading and Release of Sulforhodamine

    B

    For loading hybrid nanoparticles with

    sulforhodamine B (MSN-POLY-SRB) a solution

    of SRB (pH 7, 4.47 × 10-3

    M) was prepared. First,

    3 mg of MSN-POLY were added to 3 mL of the

    SRB solution and the mixture was stirred

    overnight at 20° C. The temperature was

    changed to 50° C and mantained during 3 hours.

    1 mL of this dispersion was centrifuged in 3

    cycles (8497 g, 10 min each cycle, 40° C), and

    redispersed in phosphate buffer to remove the

    SRB which was not loaded in the MSNs.

    Supernatants were stored to measure the SRB

    which wasn´t incorporated in MSN-POLY. The

    amount of SRB loaded into MSN-POLY was

    calculated from the difference between the

    concentration of SRB solution used in the loading

    and the concentration the separated medium

    after centrifugations. For the release experiment,

    200 µL of the dispersion with MSN-POLY and

    SRB were placed in a dialysis tube and 3.5 mL of

    phosphate buffer were introduced in the cell with

    a magnetic stirrer. A kinetic study was performed

    by changing the temperature between 20° C and

    50° C, each 20 minutes, during for 4 hours.

    3. Methods

    TEM images were obtained in a Hitachi

    transmission electron microscope (model H-

    8100 with a LaB6 filament) with an accelerator

    voltage of 200 kV. One drop of particles

    dispersion in ethanol was placed on a carbon

    grid and dried in air before observation. Images

    were processed with the Fiji software. Zeta

    potential (ξ) and hydrodynamic diameter (Dh)

    measurements of the particles were obtained

    using a Zetasizer Nano ZS, model ZEN3600.

    The absorption spectra were recorded on a

    Jasco V-660 spectrophotometer with a Peltier

    temperature control. Fluorescence emission and

    excitation spectra were recorded on a Horiba-

    Jobin Yvon Fluorolog 3 spectrofluorimeter and

    Right Angle geometry was used. All spectra

    were recorded using 1 cm x 1 cm quartz cells.

  • 4. RESULTS AND DISCUSSION

    The fluorescent MSNs were obtained through a

    sol–gel modified procedure, using CTAB as

    template and ethanol as solvent. The excitation

    and emission spectra of MSN-PDI dispersed in

    ethanol are similar to the spectra of free PDI in

    ethanol (Figure 1) meaning that the

    incorporation of the PDI in the MSNs does not

    affect the fluorescence properties of the dye.

    Figure 1 - Normalized excitation (dashed lines, λemi =

    560 nm) and emission (solid lines, λexc = 500 nm) for

    PDI (green) and MSN-PDI 7a (blue).

    The surface of the MSNs was modified with

    APTES to incorporate primary amino groups on

    the external surface.10

    Then, the template was

    removed using an acidic ethanol solution

    allowing the pores to be completely available.

    The synthesis of the polymer shell starts with the

    anchoring of the RAFT agent on the external

    surface of the MSN-PDI particles. The amino

    groups on the external surface reacted with the

    RAFT agent to obtain MSN-PDI-RAFT. These

    nanoparticles were used on the RAFT synthesis

    of the PEG-acrylate monomers, to obtain core-

    shell silica-polymer nanoparticles (MSN-POLY)

    (Figure 2).

    Figure 2 - Different steps involved in the preparation of

    hybrid MSNs coated with a thermoresponsive

    biocompatible PEG-acrylate copolymer, incorporating a

    PDI dye in the silica network.

    The diameters of the mesoporous silica

    nanoparticles were measured using TEM (DTEM)

    and DLS (Dh). Table 1 summarizes the average

    diameters obtained for the different synthesis

    with both techniques. As seen in Table 1, the

    hydrodynamic diameter is higher than the

    diameter obtained by TEM, which shows the

    tendency of nanoparticles to aggregate in

    dispersion. However, the Dh is usually higher

    than DTEM due to the hydration layer around the

    nanoparticle in the particles dispersion.

    0

    0,2

    0,4

    0,6

    0,8

    1

    1,2

    400 500 600 700

    Inte

    nsi

    ty (

    a.u

    )

    Wavelength(nm)

  • Table 1 - Average diameters and standard deviations

    obtained by DLS and TEM for the core of MSNs at 25º

    C.

    Sample Dh (nm) DTEM (nm)

    MSN-PDI 2a 210±7 150±32

    MSN-PDI 3a 220±5 160±35

    MSN-PDI 3b 180±14 170±27

    MSN-PDI 5b 170±2 160±34

    MSN-PDI 7a 160±2 140±38

    MSN-PDI 7b 170±10 160±44

    From the TEM images of MSN-PDI (Figure 3), it

    is possible to observe the spheroid shape and

    the mesopores network along the nanoparticles

    (Figure 3 C). The RAFT polymerization was

    conducted to synthesize a copolymer formed

    from a mixture of two PEG-acrylates. The

    thermossensitive behavior of the polymer shell

    was tested by analyzing the Dh of the hybrid

    nanoparticles at 25 and 40º C, for different

    samples (Table 2).

    Table 2- Average hydrodynamic diameters of different

    samples of hybrid nanoparticles dispersed in water at

    25 and 40° C.

    The average Dh for sample MSN-POLY 3b was

    290 ± 5 nm at 25ºC, decreasing to 180 ± 10 nm

    at 40ºC, providing the temperature-switchable

    collapsed/expanded conformation behavior. The

    average hydrodynamic diameter for sample

    MSN-PDI 3b at 25ºC (Table 1) was 180 ± 14

    nm. This value is equal to that obtained for the

    MSN-POLY at 40ºC, which indicates that at this

    temperature the chains of polymer shell are

    collapsed onto the silica surface.

    Figure 3 - Transmission electron microscope images of

    MSN-PDI, A and B showing the particle morphology

    and C showing the structure mesoporous.

    The Dh of MSN-POLY 7a and MSN-POLY 7a

    at 40ºC are lower than Dh of MSN-PDI,

    because the silica nanoparticles (MSN-PDI)

    are more likely to aggregate than the silica

    nanoparticles with polymer shell (MSN-POLY)

    which are more stable and so are better

    Sample Dh (nm) 25ºC Dh (nm) 40ºC

    MSN-POLY 2a 270±21 180±26

    MSN-POLY 3a 210±11 150±7

    MSN-POLY 3b 290±5 180±10

    MSN-POLY 5b 190±2 140±1

    MSN-POLY 7a 310±16 180±12

    MSN-POLY 7b 190±5 150±3

    A

    B

    C

  • dispersed. The TEM images for MSN-POLY 7b

    (Figure 3 A) clearly show the polymer shell

    around the silica nanoparticles. We also

    measured the Zeta potencial in the different

    steps of the synthesis of the nanoparticles

    (Figure 4). The zeta potential of the silica

    nanoparticles with polymer (MSN-POLY)

    increased compared with the RAFT agent

    nanoparticles (MSN-PDI-RAFT), suggesting

    the surface modification of the nanoparticles

    with the polymer.11

    Figure 4- Average of Zeta potential of the different steps

    of the synthesis of MSNs. The nanoparticles were

    dispersed in water and measured at 25° C.

    Figure 5 – Dinamic light scattering (DLS) of MSN-POLI

    2a (A) and and MSN-POLI 7a (B) with LCST at 36,3ºC

    and 35,6ºC, respectively.

    The LCST of the polymer shell in MSN-POLY

    was characterized by DLS, by measuring de Dh

    from 20ºC to 50ºC (Figure 5). The polymer shell

    present a LCST arround 36°C.12

    To determine

    the hysteresis we perform rapid cooling and

    heating cycles (Figure 6). At temperatures below

    the transition temperature (20°C), the diameter

    was 218 nm, and above this temperature (50°C)

    was 116 nm. When MSN-POLY was cooling, the

    diameter was 114 nm at 50° C and 208 nm at

    20°C, which indicates that the polymer chains

    expanded at low temperature, but collapse onto

    the silica surface at high temperature.13

    Figure 6 - Transmission electron microscope image of

    MSN-POLY 7b (A) showing a core-shell nanoparticle

    and representation of the hydrodynamic diameters at

    different temperature (B). The figure present a heating

    cycle (20 - 50ºC, black) and followed by a cooling cycle

    (50 - 20ºC, gray).

    100

    150

    200

    250

    20 30 40 50

    Dia

    me

    ter

    (n

    m)

    Temperature ( °C)

    A

    B

    A

    B

  • 3.2. Release of sulforhodamine B

    The MSNs coated with a polymeric shell were

    tested as controlled release containers for

    sulforhodamine B (SRB) molecules. The

    emission (λexc = 520 nm) and excitation (λemi =

    620 nm) spectra were recorded at 20° C and 50°

    C (Figure 6), because this temperature will be

    used in the release experiment. In figure 6, we

    can observe that the fluorescence intensity at 50

    ° C is lower than at 20° C. This happens due to

    the increase in the non-radiactive components

    with temperature leading to a decrease in the

    quantum yield. A ratio of fluorescence

    intensities, RF = IF (50ºC) / IF (20ºC), was used to

    correct the fluorescence intensities in the

    release experiments. The SRB was loaded into

    MSN-POLY and the study of release were

    performed by switching the temperature

    between 50 and 20° C, every 20 minutes. The

    emission of free SRB and SRB loaded MSN-

    POLY was followed for 4 hours. The kinetics

    started at 50°C because at this temperature the

    polymer shell was collapsed and the MSNs were

    loaded. The fluorescence intensity obtained at

    50ºC for the SRB and MSN-SRB were

    normalized with the ratio RF (Figure 7).

    Figure 7 - Excitation (dashed lines, λemi = 620 nm) and

    emission (solid lines, λexc = 520 nm) spectra of SRB at

    20° C (red) and 50° C (blue).

    Figure 8 - Fluorescence intensity obtained for SRB

    (blue) and MSN-SRB (green), measured continuously

    during 4 hours at λemi = 585 nm and λexc = 565 nm. The

    fluorescence intensities measured at 50° C for SRB

    (red) and MSN-SRB (purple) were normalized with the

    RF ratio.

    The release profiles of MSN-POLY can be

    observed by applying heating - cooling cycles

    to the nanoparticles. The values at each

    plateau (20ºC and 50ºC) were fitted with a 1st

    order equation (Figure 9). The slope of the

    different temperature for SRB and MSN-SRB

    are compiled in Figure 10. The slopes of MSN-

    SRB at 20°C are identical over time because

    at this temperature the polymer chains are

    hydrated and expanded allowing SRB

    molecules to flow from the core to the shell but

    forcing them to remain there due to the

    electrostatic interations with the polymer

    chains. The slopes for MSN-SRB at 50ºC are

    higher than for free SRB at 50ºC, because by

    increasing the temperature to 50ºC, the

    extended polymer chains of MSN-SRB shrink

    rapidly, releasing the SRB molecules that were

    in the polymer shell and thus increasing the

    fluorescent intensity in the cell bottom

    compartment. The slopes decrease over time,

    because the amount of SRB released

    decreases as the particle load is depleted.

  • Figure 9- Fluorescence intensities obtained at 20 ° C for

    SRB (red) and MSN-SRB (blue) and normalized

    fluorescence intensities obtained for the SRB at 50º C

    (green) SRB and MSN-SRB at 50° C (purple). The

    fluorescence intensities were measured to λemi= 585 nm

    and λexc = 565 nm.

    Figure 10 - Representation of the slopes obtained

    from the adjustments made to the variations of the

    fluorescence intensity of SRB (gray) and MSN-SRB

    (blue) at 20° C, and SRB (yellow) and MSN-SRB

    (orange) at 50° C.

    The representation of the slopes versus time

    (Figure 10) shows the free SRB slopes at 20°C

    are similar over time (due to the slow diffusion of

    free SRB molecules in solution by the

    membrane). The free SRB slopes at 50°C not

    significantly changed (although, the slopes were

    higher than slopes at 20°C, due to the increased

    porosity of the membrane with the temperature).

    Figure 11 - Schematic representation of the behavior

    of MSN-POLY loaded with sulforhodamine B. The

    polymer at 50º C (a) is collapsed closing the pores.

    Decreasing the temperature to 20ºC (b), the SRB

    molecules flow from the pores to the hydrated

    expanded polymer shell, remaining there duo to

    electrostatic interations. Upon increasing the

    temperature to 50ºC (c), the polymer shell collapsed

    squeezing the SRB molecules.

    The proposed mechanism of release is

    presented in Figure 11. Initially, at 50ºC, the

    MSNs were fully loaded and the polymer chains

    are collapsed closing the pores entrance (Figure

    11 a). By decreasing the temperature to 20°C,

    SRB molecules flow to the polymer shell and

    remain there (Figure 11 b). Increased the

    temperature to 50°C, the polymer chains

    collapsed and SRB is released (Figure 11 c). By

    decreasing the temperature to 20°C again, a

    new cycle starts (Figure 11 d), with the system

    behaving as a nano-pump for the drug model.

    CONCLUSIONS

    A novel drug delivery carrier based on a

    mesoporous silica core and a stimui-responsive

    polymer shell has been successfully prepared.

    The MSNs have incorporated in their structure a

    0

    500

    1000

    1500

    2000

    2500

    3000

    0-2

    0

    23

    -42

    47

    -62

    68

    -80

    90

    -10

    0

    11

    0-1

    20

    13

    0-1

    42

    15

    0-1

    60

    17

    0-1

    80

    19

    0-2

    00

    Slo

    pe

    Time (min)

  • fluorescent PDI derivative for bioimaging, and

    after template removal the pores are completely

    available to accommodate molecules. The

    characterization by TEM shows that the

    nanoparticles are mesoporous, monodispersed

    and they have a spheroid shape. DLS

    measurements showed the termoresponsive

    behavior of the polymer shell. The release of a

    model molecule (SRB) from the MSN-POLY

    could be controlled by adjusting the

    temperature. This hybrid nanocontainer have

    potential applications as a nano-pump drug

    delivery system controlled by temperature.

    REFERENCES

    1 A. Rodrigues, T. Ribeiro, F. Fernandes, J. P.

    Farinha, C. Baleizão, “Intrinsically Fluorescent

    Silica Nanocontainers: a promising theranostic

    platform”, Microsc. Microanal., 2013, 19, 1216-

    1221, DOI: 10.1017/S1431927613001517.

    2 B. Trewyn, I Slowing, U. Giri, H. Chen, V. Lin,

    "Synthesis and Functionalization of a

    Mesoporous Silica Nanoparticle Based on the

    Sol–Gel Process and Applications in Controlled

    Release", Acc. Chem. Res.2007,40,846–853,

    DOI:10.1021/ar600032u.

    3 A. Babu, A. K. Templeton, A. Munshi, R.

    Ramesh, “Nanoparticle-Based Drug Delivery for

    Therapy of Lung Cancer:Progress and

    Challenges” Journal of Nanomaterials, 2013,

    11, DOI: 10.1155/2013/863951.

    4 Kurkina T., Balasubramanian K., “Towards in

    vitro molecular diagnostics using

    nanostructures”, Cellular and Molecular Life

    Science. 2012, 69, 373-388, DOI:

    10.1007/s00018-011-0855-7.

    5 Sanvicens N., Marco MP., “Multifunctional

    nanoparticles--properties and prospects for their

    use in human medicine”, Trends in

    Biotechnology, 2008, 26, 425-433, DOI:

    10.1016/j.tibtech.2008.04.005.

    6 A. Favier, M. Charreyre, ”Experimental

    Requirements for an Efficient Control of Free-

    Radical Polymerizations via the Reversible

    Addition-Fragmentation Chain Transfer (RAFT)

    Process”, Macromolecular Rapid

    Communications, 2006, 27, 653–692; DOI:

    10.1002/marc.200500839.

    7 M. Beija, J. Marty, M. Destarac,” RAFT/MADIX

    polymers for the preparation of

    polymer/inorganic nanohybrids”, Progress in

    Polymer Science, 2011, 36,845–886, DOI:

    10.1016/j.progpolymsci.2011.01.002.

    8 T. Ribeiro, C. Baleizão, and J. P. Farinha;”

    Synthesis and Characterization of

    Perylenediimide Labeled Core-Shell

    HybridSilica-Polymer Nanoparticles”, J. Phys.

    Chem. C 2009, 113, 18082–18090, doi:

    10.1021/jp906748r.

    9 Y. Lin, C. Tsai, H. Huang, C. Kuo, Y. Hung, D.

    Huang, Y. Chen, C. Mou, “Well-Ordered

    Mesoporous Silica Nanoparticles as Cell

    Markers”, Chemistry of Materials, 2005,17,4570-

    4573, DOI: 10.1021/cm051014c.

    10 I. Slowing, J. L. Vivero-Escoto, B. G. Trewyn,

    V. S. Lin.” Mesoporous silica nanoparticles:

    Structural design and applications”, Journal of

    Material Chemistry, 2010, 20, 7924–7937, DOI:

    10.1039/c0jm00554a.

  • 11 Y. Zhang, M. Yang, N. G. Portney, “Zeta

    potential: a surface electrical characteristic to

    probe the interaction of nanoparticles with

    normal and cancer human breast epithelial

    cells”, Biomedical Microdevices, 2008, 10, 321–

    328, DOI 10.1007/s10544-007-9139-2.

    12 J.F. Lutz, A. Hoth, "Preparation of Ideal PEG

    Analogues with a Tunable Thermosensitivity by

    Controlled Radical Copolymerization of 2-(2-

    Methoxyethoxy)ethyl Methacrylate and

    Oligo(ethylene glycol) Methacrylate,"

    Macromolecules, 2006,vol. 39, pp. 893-896,

    DOI: 10.1021/ma0517042.

    13 D. Roy, J. Cambre, B. Sumerlin, "Future

    perspectives and recent advances in stimuli-

    responsive materials", Progress in Polymer

    Science, 2010, 35, 278–301, DOI:

    10.1016/j.progpolymsci.2009.10.002.