Naphthalocyanine-Based Biodegradable Polymeric Nanoparticles for Image-Guided Combinatorial Phototherapy

advertisement
Naphthalocyanine-Based Biodegradable Polymeric Nanoparticles for
Image-Guided Combinatorial Phototherapy
Taratula, O., Doddapaneni, B. S., Schumann, C., Li, X., Bracha, S., Milovancev,
M., ... & Taratula, O. (2015). Naphthalocyanine-Based Biodegradable Polymeric
Nanoparticles for Image-Guided Combinatorial Phototherapy. Chemistry of
Materials. doi:10.1021/acs.chemmater.5b03128
10.1021/acs.chemmater.5b03128
American Chemical Society
Accepted Manuscript
http://cdss.library.oregonstate.edu/sa-termsofuse
Naphthalocyanine-Based Biodegradable Polymeric Nanoparticles for Image-Guided
Combinatorial Phototherapy
Olena Taratula,§ Bhuvana S. Doddapaneni,§ Canan Schumann,§ Xiaoning Li,§ Shay Bracha, #
Milan Milovancev, # Adam W. G. Alani, §,* Oleh Taratula§,*
§
Department of Pharmaceutical Sciences, College of Pharmacy, Oregon State University,
Portland, Oregon 97201, United States
#
Department of Clinical Sciences, College of Veterinary Medicine, Oregon State University,
Corvallis, OR 97331, United States
*
Address correspondence to Oleh.Taratula@oregonstate.edu, Adam.Alani@oregonstate.edu
Abstract (250 words)
Image-guided phototherapy is extensively considered as a promising therapy for cancer
treatment. To enhance translational potential of this modality, we developed a single agent-based
biocompatible nanoplatform that provides both real time near-infrared (NIR) fluorescence
imaging and combinatorial phototherapy with dual photothermal and photodynamic therapeutic
mechanisms. The developed theranostic nanoplatform consists of two building blocks: (1) silicon
naphthalocyanine (SiNc) as NIR fluorescence imaging and phototherapeutic agent and (2) a
copolymer, poly(ethylene glycol)-block-poly(ε-caprolactone) (PEG-PCL) as the biodegradable
SiNc carrier. Our simple, highly reproducible and robust approach results in preparation of
spherical, monodisperse SiNc-loaded PEG-PCL polymeric nanoparticles (SiNc-PNP) with a
hydrodynamic size of 37.66 ± 0.26 nm (polydispersity index = 0.06) and surface charge of -2.76
± 1.83 mV. The SiNc-loaded nanoparticles exhibit a strong NIR light absorption with an
1
extinction coefficient of 2.8 x 105 M-1cm-1 and efficiently convert the absorbed energy into
fluorescence emission (F = 11.8%), heat (T ~ 25 C) and reactive oxygen species. Moreover,
the SiNc-PNP are characterized by superior photostability under extensive photoirradiation and
structure integrity during storage at room temperature over a period of 30 days. Following
intravenous injection, the SiNc-PNP accumulated selectively in tumors and provided high lesionto-normal tissue contrast for sensitive fluorescence detection. Finally, Adriamycin-resistant
tumors treated with a single intravenous dose of SiNc-PNP (1.5 mg/kg) combined with 10 min of
a 785 nm light irradiation (1.3 W/cm2) were completely eradicated from the mice without cancer
recurrence or side effects. The reported characteristics make the developed SiNc-PNP a
promising platform for future clinical application.
KEYWORDS: naphthalocyanine, photothermal therapy, photodynamic therapy, near-infrared
imaging, photostability, theranostic, biodegradability
2
INTRODUCTION
Image-guided phototherapy is extensively considered as a promising, non-invasive therapeutic
modality for intraoperative cancer treatment.1-8 It combines dual photothermal and photodynamic
therapeutic mechanisms, which are different from chemo- and radiotherapies and, therefore, has
significant potential for treatment of resistant cancer cells.7-9 Furthermore, phototherapy results
in minimal side effects because it selectively kills cancer cells by non-toxic photoactive agents
upon activation by a specific wavelength of targeted light.3,
4, 6, 8
Light-activated photoactive
agents can concurrently generate both intratumoral heat and toxic reactive oxygen species
(ROS), leading to cancer cell destruction.3, 4, 6, 8 Moreover, combining phototherapy with realtime fluorescence imaging enables tracking of accumulation of photoactive agents within tumor
tissue and allows for precise focusing of the laser beam on the tumor targets, to both activate
therapeutic mechanisms and minimize exposure of healthy organs.3, 4
To translate this promising therapeutic modality to the clinic, multifunctional theranostic
nanoplatforms that provide both real-time fluorescence imaging and phototherapy are urgently
needed. Phototherapeutic and imaging agents capable of near-infrared (NIR) light absorption
(~700 -950 nm) and fluorescence have attracted significant attention as the main building blocks
for these nanoplatforms.1, 2, 4, 8, 10, 11 Because biological tissues and body fluids show minimal
light absorption and autofluorescence in the NIR spectral region, NIR phototherapeutic drugs
and fluorophores can provide for an efficient treatment and imaging of deep-seated tumors.12, 13
In recent years, the number of the required nanoplatforms has been developed by encapsulation
of separate NIR imaging and phototherapy agents into a single nanocarrier.2, 7, 10, 11, 14, 15 Despite
their efficiency being demonstrated in preclinical studies, the high complexity and complicated
synthetic procedure could limit further translational potential of these nanoplatforms.3
3
If NIR imaging and phototherapeutic properties could be merged into a single agent,
substantially improved cancer theranostic outcomes and a high translational potential could be
expected. Recently, nanoporphyrin with absorption and emission maximum in a 600 – 700 nm
spectral range have been employed to develop a smart “all-in-one” nanoplatform that
conveniently integrates fluorescence imaging and phototherapeutic functions.3 To further
improve the theranostic efficiency of the “all-in-one” nanoplatforms, there is a high interest in
application of photoactive agents capable to absorb and emit light in the 750 - 900 nm spectral
region. Application of these dyes offers several advantages for image-guided phototherapy
including enhanced tissue penetration of excitation and emission light, lower autofluorescence
from surrounding tissue and body fluids, and decreased skin photosensitivity as a potential side
effect of phototherapy.16-19 The FDA-approved NIR dye, indocyanine green (ICG) with
photophysical properties in a 750 - 850 nm spectral region, has been considered as a main
building block for the construction of theranostic nanoplatforms composed of a single
phototherapeutic agent.4,
20-23
However, ICG exhibits poor photostability and, therefore,
photodegradation of the dye can be misread as an absence of fluorescence signal during the
image-guided phototherapy.24 In addition, photo-degradation of the photoactive agent also
compromises ROS and heat generation abilities, diminishing its therapeutic potential.8, 13, 24, 25
Finally, the quantum yield of ICG is quite low (F = 2.7%), which significantly reduces its
efficiency as the fluorescence imaging agent.26
Recent reports indicated that, owing to its unique intrinsic properties including enhanced
absorption in the spectral range 750 - 800 nm, silicon naphthalocyanine (SiNc) can be efficiently
activated with NIR light to produce fluorescence, ROS and heat for imaging and combinatorial
treatment of tumors.8,
25, 27, 28
Furthermore, it was revealed that dendrimer-encapsulated SiNc
4
molecules with axial substituents exhibit superior photostability compared to ICG.8 Clinical
application of SiNc, however, is limited by low water solubility, aggregation and lack of cancer
specificity.8, 29 Particularly, an intrinsic tendency of SiNc molecules to aggregate results in selfquenching between the bound dyes, which diminishes their phototherapeutic efficacy and
imaging ability.30 To fully develop its clinical potential and overcome the above-mentioned
obstacles, there is a critical need to develop a highly efficient SiNc-based nanoplatform. Ideally,
this nanoplatform has to be (i) biocompatible and biodegradable, (ii) simple and reproducible in
preparation, and (iii) not complex, in that it contains a small number of components. In addition,
it has to increase the water solubility of the SiNc molecules to clinically relevant concentrations
and decrease their self-aggregation in aqueous solutions to preserve the fluorescence emission
and phototherapeutic properties in the physiological environment. Finally, it should provide
specific delivery of SiNc to the cancer tissue within a short period of time after systemic
administration, while being rapidly eliminated from healthy organs.
Herein, we report the development of a desirable multifunctional nanoplatform for image-guided
phototherapy, which contains only two components: (1) SiNc as NIR imaging and
phototherapeutic agent, and (2) a biodegradable-biocompatible block copolymer, PEG-PCL,31, 32
as the SiNc carrier (Figure 1). The developed procedure is simple, reproducible and results in
fabrication of stable, highly uniform nanoparticles with excellent imaging and phototherapeutic
properties. The obtained characteristics make the fabricated theranostic nanoplatform promising
material for future clinical application.
MATERIALS AND METHODS
Materials. SiNc (2,3-Naphthalocyaninato-bis(trihexylsiloxy)silane) was obtained from SigmaAldrich
Co.
(Milwaukee,
WI).
PEG-PCL
5
(methoxy
poly(ethylene
glycol)-b-poly(ε-
caprolactone), MW = 15000 Da, PDI=1.17) was obtained from Advanced Polymer Materials Inc.
(Montreal, Canada). All other reagents were purchased from Fisher Scientific, Inc. (Fairlawn,
NJ) and VWR International, LLC (Radnor, PA).
Preparation of SiNc-based Polymeric Nanoparticles. SiNc nanoparticles were constructed
using modified solvent evaporating method.33 Breifly, PEG-PCL (20 mg) and SiNc (0.6 mg)
were dissolved in 2 mL of THF, The solution was transferred into 20 mL scintillation vial and to
the constantly stirred solution 2 mL of saline was added in a 1-step process. THF was removed
under vacuum using a rotary evaporator and the final volume was adjusted to 2 mL using saline.
The evaporation cycle was divided into three segments, with the first segment lasting 7 min at
420 mbar, the second for 7 min at 320 mbar and the final segment of 6 min at 200 mbar. The
temperature of the water bath was maintained at 45 oC with a rotation of 100 rpm for the round
bottom flask. The final volume was adjusted to 2 mL with saline, the nanoparticles solution was
centrifuged (5 min, 10,000 rpm) and the supernatant was filtered using a 13 mm 0.2 µM nylon
filter.
Characterization. The hydrodynamic diameter, size distribution and zeta potential of
naphthalocyanine loaded polymeric nanoparticle (SiNc-PNP) was determined using dynamic
light scattering (DLS, Malvern ZetaSizer NanoSeries, Malvern, UK) at 25 C as previously
described.7, 8, 34-36 The morphology of the polymeric nanoparticles was observed using cryogenic
transmission electron microscopy cryoTEM. The 3 µl of nanoparticle suspensions (0.3 mg/ml)
were pipetted onto glow-discharged (30 sec, 30 mAmp) copper Quantifoil holey carbon support
grids (Ted Pella 658-300-CU) and vitrified on liquid ethane using a Mark IV Vitrobot (FEI,
Hillsboro, OR). The conditions utilized were 100% humidity, blot force 0 and blotting times
between 2-4 seconds. Low-dose conditions were used to acquire images on a FEI Krios-Titan
6
equipped with a Falcon II direct electron detector (FEI, Hillsboro, OR). Cryo-EM images were
collected with a defocus range of 2-4 µm.
The fluorescence of SiNc-PNP was recorded on UV-1800 Shimadzu spectrophotometer
(Carlsbad, CA) and absorbance spectra were taken on Cary Eclipse R3896 fluorescence Varian
spectrophotometer (Mulgrave Victoria, Australia).
To determine SiNc loading capacity, the SiNc-PNP solution was freeze-dried, and the obtained
pellets were weighted and re-dissolved in THF. SiNc encapsulation efficiency was evaluated as
the percentage of SiNc weight (quantified based on UV–Vis absorption of SiNc at 772 nm)
encapsulated into PNP over the initial weight of SiNc used for loading. The loading capacity (%)
and encapsulation efficiency (%) of the SiNc-PNP were calculated based on the equation
below:37
Loading Capacity (%) = mass of SiNc in SiNC-PNP / total mass of the SiNc-PNP × 100
Encapsulation efficiency (%) = mass of SiNc in SiNc-PNP / initial amount of SiNc × 100.
Fluorescence quantum yield for SiNc-PNP was calculated in aqueous solution (ex = 750 nm)
with indocyanine green (ICG) as reference (0.4)38 as follows:
SiNc = (AbsICG/AbsSiNc) x (AreaSiNc/AreaICG) x ICG x 100
where SiNc is a quantum yield of SiNc-PNP; AbsICG and AbsSiNc are absorbances of ICG and
SiNc-PNP, respectively; AreaICG and AreaSiNc are fluorescence emission areas of ICG and SiNcPNP, respectively; where ICG is a quantum yield of SiNc-PNP.
Water was employed as solvent for ICG and SiNc-PNP and, therefore, refractive index was not
used in the equation.
Brightness of SiNc-PNP relative to ICG was calculated by using the previously reported
equation: 39
7
Brightness = (SiNc x SiNc) x (ICG x ICG)
where SiNc and ICG are extinction coefficients of SiNc-PNP and ICG, respectively.
Stability of SiNc-PNP (SiNc encapsulation, size, surface charge) in aqueous media over one
month was assessed at room temperature protected from light.
Singlet Oxygen 1O2 Generation. As previously reported, 1O2 production was measured by
SOSG assay in a 96 well plate.7, 35. Corresponding wells were treated for 5 min at 785 nm and
1.3 W/cm2 while the dark controls were protected from light. Instantly, the samples were
evaluated
on
Varian
Cary
Eclipse
R3896
fluorescence
spectrophotometer
using
excitation/emission of 504/525 nm.
Photothermal Effect. A 785 nm laser diode (CW, ThorLabs, Newton, NJ, 1.3 W/cm2) was used.
For temperature increase evaluation, 300 µL of SiNc-PNP aqueous solutions (0.3 mg/mL) in the
500 L vials was exposed for 20 min to the NIR laser. The thermocouple was used to record
temperature at selected time points. As a control, water was tested under the same irradiation
conditions. Absorption of SiNc-PNP solutions in water at 0.3 mg/mL SiNc concentration was
evaluated prior to and after 30 min laser treatment (1.3 W/cm2, 785 nm) in order to evaluate
photostability.
Dark Cytotoxicity. To determine the dark cellular cytotoxicity of SiNc-PNP a modified cell
viability assay based on Calcein AM (Biotium, Inc., Hayward, CA) was employed. Human
ovarian epithelial carcinoma/Adriamycin-resistant derivative cell line (A2780/AD) were seeded
in two 96-well plates (1 x 104 cells/well) and the next day were incubated with SiNc-PNP (2.0,
5.0, 10.0, 25.0, 50.0, and 100 µg/mL) in RMPI 1640 (10% FBS) in the dark conditions. As a
control, A2780/AD cells only were used. After 36 hours, cell viability was assessed with Calcein
AM solution (10 µM in DPBS buffer) for 1 hour. Fluorescence was recorded on a multiwell
8
Synergy HT plate reader (BioTek Instruments, Winooski, VT) with a 485/528 nm
excitation/emission filters. The cell viability (%) was expressed as a percentage relative to the
control cells.
Cellular Internalization. A2780/AD cancer cells (2 x 105 cells/well) were plated in 6-well
plates followed by 36 hours incubation with SiNc-PNP (30 µg/mL). Cells were washed with
DPBS and imaged with EVOS® FL Cell Imaging System (Life Technologies, Grand Island,
NY).
Quantification of Intracellular SiNc Content. The SiNc content in cancer cells incubated with
SiNc-PNP was estimated via recording UV-Vis absorption. Cells were seeded in T-25 flask to
reach 80% confluency followed by 36 hours incubation with SiNc-PNP (30 g/mL). Next, 5 x
106 cells were lysed with 50 mM NaOH and diluted in THF to dissolve SiNc at 1:9 ratio (100 L
cell lysate/900 L THF). The absorbance was recorded at 772 nm and the amount of SiNc
uptaken by cells was calculated by using the standard curve.
ROS Measurements. DCFH-DA (2´,7´-dichlorodihydrofluorescein diacetate) was employed to
evaluate intracellular ROS levels.7,
35
Cancer cells seeded in 96-well plates (1×104 cells/well)
were incubated with SiNc-PNP (30 µg/mL) for 36 h. After incubation, cell were treated with 10
µM DCFH-DA for 30 min prior to 10 min light treatment (785 nm, 1.3 W/cm2). Separately, the
untreated cells and SiNc-PNP without light treatment were used as a control a dark control,
respectively. Fluorescence was recorded on a plate reader (λabs = 485 nm and λem = 528 nm).
Combinatorial Phototherapy on Cell Pellet. The T-25 cell culture flask with A2780/AD cells
of 80% confluency were incubated for 36 h with SiNc-PNP (50 g/mL) in RPMI 1640 (10 %
FBS). Next, 2 × 106 cells were centrifuged (5 min, 1000 rpm) to give the cell pellet. Samples (in
250 L of culture media) were photoirradiated (785 nm, 10 min, 1.3 W/cm2). Temperature was
9
recorded before and after 10 min light treatment using a fiber optic temperature probe inside of
the pellets. After the light experiment, the cells were seeded in a 96-well plate (1 × 104
cells/well). After 24 h, the cell viability was evaluated by Calcein AM assay along with the
controls: untreated cells, cells incubated with SiNc-NP and untreated cells exposed to the NIR
laser (785 nm, 1.3 W/cm2, 10 min).
In vivo NIR Fluorescence Imaging and Biodistribution. Approval for the animal studies was
given by the Oregon State University IACUC and all procedures were performed according to
IACUC policies. A subcutaneous A2780/AD tumor animal model was developed via injection of
A2780/AD cells (5 × 106) in RPMI 1640 into the flank of 5-week-old athymic nu/nu female mice
(Harlan Laboratories, Inc., USA). The tumor sizes were measured for 25 days and were
calculated as 0.5 × width2 × length. At tumor volume of ~ 40 mm3, the mice were injected
intravenously with 100 L SiNc-PNP in saline (0.3 mg/mL) at the dose of 1.5 mg/kg. Then, the
whole animal and harvested tumor, heart, liver, spleen, lung, and kidney were imaged using
Pearl Impulse Small Animal Imaging System (LI-COR, Lincoln, NE) with 800 nm channel at
0.5, 2, 24 and 96 h post-injection.
In vivo Phototherapy. With tumor volume of ~ 40 mm3, the mice were distributed to one of four
treatment groups (n=20, 5 mice/group) and injected with saline (control), saline and followed by
exposure to 10 min of NIR light (light control), SiNc-PNP (1.5 mg/mL) (SiNc-PNP dark) or
SiNc-PNP (1.5 mg/kg) followed by NIR (1.3 W/cm2, 785 nm) irradiation (SiNc-PNP light). At
this point, fluorescent images were obtained to determine the SiNc location. At 24 h, the mice in
the specific experimental groups were treated once with phototherapy for 10 min (785 nm, 1.3
W/cm2) under isoflurane anesthesia. Tumor size and body weight were documented for all mice
groups during 25 days after phototherapy. Mice were euthanized when tumor diameter was more
10
than 1200 mm3. Intratumoral temperature was measured in 10-s intervals using a fiber optic
temperature probe as previously described.8
Ex vivo Histology. Tissue samples of control (untreated and NIR light only treated) tumors,
tumor treated SiNc-PNP dark and SiNc-PNP light were dissected from mice 6 h postphotoirradiation. The formalin-fixed paraffin embedded tissues were sectioned and collected
onto slides as reported earlier.8 Hematoxylin and eosin (H&E, IHC WORLD, LLC Ellicott City,
MD), Hoechst 33342 (NucBlue® Live ReadyProbes® Reagent, Life Technologies, Grand
Island, NY) and caspase-3 (Cell Signaling Technology, Danvers, MA) staining were performed
following manufacturer’s protocol and observed by a Carl Zeiss Axio Imager Z1 fluorescent
microscope (Thornwood, NY).
Data Analysis. Data were analyzed and presented as mean values ± standard deviation (SD)
from three to eight independent measurements. The statistic difference was analyzed using t-test
where P value of 0.05 was considered significant.
RESULTS AND DISCUSSION
Synthesis of SiNc-PNP. To prepare the proposed nanoplatform, we used SiNc molecules
modified with two organic substituents, trihexylsilyloxide (R), through their conjugation to
hydroxyl groups at the silicon center of naphthalocyanine (Figure 1). The application of
substituted SiNc provides several important advantages for our nanoplatform. First, we
previously demonstrated that the appropriate hydrophobic substituents can facilitate
encapsulation of phthalocyanine and naphthalocyanine dyes into the hydrophobic interior of the
nanoparticles.7, 8, 35 Second, these substituents provide steric hindrance of these dyes and reduce
intermolecular aggregation inside of the nanoparticles, minimizing the self-quenching effect and
11
thus improving their fluorescence imaging and ROS generating properties.7, 8, 35 Third, several
reports have shown that SiNc with the substituents in axial positions exhibits superior
photostability in comparison to non-modified molecules. 8, 30
To overcome in vitro and in vivo challenges associated with low intrinsic solubility of the SiNc
(<1 ng/mL, evaluated by Advanced Chemistry Development ACD/Universal LogSw GALAS
Module, Percepta 14.0.0 (Build 1996)), amphiphilic copolymer, PEG-PCL, which consist of a
hydrophilic poly(ethylene glycol) (PEG) and a hydrophobic poly(-caprolactone) (PCL) blocks
(Figure 1), has been employed as a carrier. The PEG is nontoxic material and cleared by the US
Food and Drug Administration (FDA) for internal use in the human body.40 The PCL is
biocompatible and biodegradable polyester approved by the FDA for biomedical applications in
humans.31 The PEG-PCL copolymer was extensively tested for various biological applications
and all studies have reported low cytotoxicity and complete clearance from the body.32, 41, 42 As a
result, this co-polymer is characterized by controlled biodegradability, excellent biocompatibility
and self-assembly in aqueous solution to form polymeric nanoparticles (PNP) with a
hydrophobic core for loading non-water soluble drugs.32 In addition, this copolymer, while
forming nanoparticles, decreased drug uptake by the liver and kidney, and also prolonged drug
retention in the blood.43
A simple, highly reproducible and robust approach for transformation of hydrophobic SiNc into
water-soluble nanoplatforms has been developed through loading of SiNc into the hydrophobic
core of PEG-PCL-based polymeric nanoparticles (Figure 1). The encapsulation procedure is
based on solubilizing SiNc and PEG-PCL in THF followed by addition of saline to produce an
emulsion. After evaporation of the organic solvent under vacuum, stable SiNc-PNP are obtained
with drug loading capacity and encapsulation efficiency of 3% and 98%, respectively. According
12
to the quantitative UV–Vis absorption measurements the formulated nanoparticles were able to
solubilize 0.30 ± 0.01 mg/mL (3 x 105 ng/mL) of SiNc and retain the compound at the initial
concentrations for tested 30 days period. Thus, incorporation of SiNc into the polymeric
nanoparticles (PNP) increased its intrinsic aqueous solubility (<1 ng/mL) by 3 x 105 fold, thereby
achieving therapeutically relevant dosing concentrations for in vivo assessment. Our findings
were consistent with published results, where the incorporation of a hydrophobic compound into
amphiphilic diblock copolymers like PEG-PCL increases their aqueous solubility and stability.32,
44
According to the DLS analysis, the hydrodynamic size of the SiNc-PNP was 37.66 ± 0.26 nm
with highly monodisperse size distribution as indicated by polydispersity index (PDI) of 0.06 ±
0.02 (Figure 2A). PDI values of less than 0.1 specify unimodal distribution of nanoparticles.
Cryogenic transmission electron microscopy (cryo-TEM) images revealed a spherical
morphology for SiNc-PNP with an average size of 19.3 ± 1.5 nm (Figure 2B) and further
confirmed preparation of highly monodisperse nanoparticles by using the developed approach
(Figure S1). The observed discrepancy between SiNc-PNP sizes measured by TEM and DLS is
in good agreement with the previous studies and it is related to the fact that DLS yields a larger
average size by measuring the hydro-dynamic diameter of the particle including the solvation
layers, while cryoTEM detects their dehydration morphology.4, 34 Nevertheless, both techniques
demonstrated that the developed nanoplatforms have a required size to minimize their rapid
clearance by the kidneys (<10 nm), identification by macrophages (<100 nm), and to increase
blood circulation time and accumulation in cancer tumors via passive targeting (<200 nm).45-47
Moreover, the recent studies indicated that nanoparticles with a size ≤50 nm have an advantage
in terms of the deep penetration and retention in the tumor tissue.48
13
In addition to the size, surface modification and charge of nanoparticles are also important
factors affecting their clearance and biodistribution after systemic administration.47,
49, 50
The
employed amphiphilic copolymer, PEG-PCL, introduced a biocompatible, non-charged PEG
layer to the nanoparticles surface, which provides for almost neutral zeta potential of the
prepared SiNc-PNP (-2.76 ± 1.83 mV). The number of reports indicated that PEG layer and
neutral/slightly negative surface charge significantly reduce non-specific interaction with plasma
proteins and, therefore, minimize opsonization and rapid uptake of nanocarriers by the
reticuloendothelial system.47, 49, 50
Of note, the developed nanoparticles are characterized by a high stability during storage. Thus,
no noticeable change in nanoparticle size distribution, zeta potential or drug loading was detected
over the one month period at the room temperature, implying that SiNc-PNP formulation is
stable in aqueous environment (Figure 2A and Table 1). The detected long shelf life of the
nanoplatform further indicates its high translation potential. The stability of SiNc-PNP in the
solution system similar to physiological environment such as phosphate buffer solution (PBS)
was also evaluated. No significant changes in nanoparticles hydrodynamic diameter (37.01 ±
0.18 nm vs 36.58 ± 0.20 nm) and UV-Vis absorption spectra (Figure S2) were detected over 24 h
period at the room temperature.
Theranostic Properties of SiNc-PNP. The effective photoactive theranostic agents should
exhibit a strong NIR absorption to assure efficient conversion of the absorbed light into
fluorescence emission, ROS and heat. SiNc loaded in the polymeric nanoparticles revealed a
strong NIR absorption in water with extinction coefficient of 2.8 x 105 M-1cm-1 which is 2.8-fold
higher as compared to the FDA approved clinical imaging agent, ICG (1.0 x 105 M-1cm-1).39
Upon excitation with NIR light, the SiNc-PNP demonstrated intense fluorescence emission in the
14
760-810 nm wavelength range (Figure 2C). It is known that most NIR dyes including SiNc are
characterized by the small Stokes shift.27 In the current studies to avoid crosstalk between the
excitation light and resulting fluorescence signal, the SiNc-loaded nanoparticles were excited at
750 nm and the fluorescence was recorded in 760 – 850 nm spectral range. Fluorescence
quantum yield (F) of 11.8% was calculated for SiNc-PNP in water using ICG as reference
solution.38 The fluorescence quantum yield of free SiNc was also evaluated in THF and F of
30.0 % was calculated. Aggregation of SiNc molecules inside of the polymeric nanoparticles
could induce quenching effect. To minimize this effect, we employed SiNc molecules modified
with trihexylsilyloxide. These substituents provide steric hindrance of these dyes and reduce
intermolecular aggregation inside of the PEG-PCL nanoparticles, minimizing the self-quenching
effect. Therefore, the relatively high quantum yield of 11.8% for SiNc-PNP in water was
achieved. Of note, the nanoplatform-loaded SiNc has a higher quantum yield in aqueous
environment (11.8%) than NIR dyes such as ICG (2.7%) and IRDye800 (~4-10%), currently
employed in clinical trials.26, 51, 52 Furthermore, SiNc-PNP are approximately 10 times brighter
than ICG as determined according to the previously reported equation.39
Jin et al. recently developed highly fluorescent SiNc-doped polymer dots (Pdots) and the
quantum yield of the doped NIR dye was reported to be 11%.27 The high fluorescence intensity
of SiNc was achieved by excitation of Pdots with 457 nm light and following energy transfer
from Pdot matrix to NIR dye. However, the fluorescence of the SiNc-loaded Pdots excited with
NIR light at 763 nm was substantially dimmer (F = 7%) than the Pdots excited at 450 nm. It is
important to indicate that SiNc-loaded in our polymeric nanoparticles exhibit high fluorescence
emission (F = 11.8%) following excitation with NIR light at 750 nm, which is strongly required
for the deeper tissue penetration. In addition, SiNc-PNP in water revealed an intense NIR
15
fluorescence signal detected by Pearl Imaging System (Figure 2C, inset) using 800 nm imaging
channel, which eludes tissue autofluorescence. Thus, the PEG-PCL-based nanocarrier can
efficiently solubilize SiNc molecules in water, prevent them from aggregation, and, therefore,
preserves the fluorescent properties of the dye.
To test properties of SiNc-PNP as an agent for photodynamic therapy (PDT), production of
singlet oxygen (1O2) was evaluated by the SOSG assay.53 After treatment with a laser diode (785
nm, 5 min 1.3 W/cm2,), SiNc-PNP demonstrated a 20-fold increase in 1O2 generation when
compared to the water control (Figure 2D). Thus, SiNc within the developed PEG-PCL
polymeric nanoparticles is capable of an efficient 1O2 generation and therefore demonstrates
great potential as a PDT agent.
To assess heat generation properties of SiNc loaded in the polymeric nanoparticle, the SiNc-PNP
solution was irradiated with the 785 nm NIR light (1.3 W/cm2) and temperature variations were
recorded during 20 min. Figure 2E indicates that the nanoparticles demonstrate rapid heating
under photoirradiation, reaching 40 °C in 2.5 min and 50 °C in 5.5 min. The attained temperature
could be sustained for the required duration when the laser diode is on. We also demonstrated
that the desired temperature increase can be achieved by varying laser power density (Figure S4),
providing the possibility for controlling intratumoral temperature during treatment. In the same
experimental settings, water did not trigger a significant temperature increase (∆T = 1.1 °C),
suggesting that the combination of SiNc-PN and NIR light was responsible for the heat
generation (Figure 2E). Our results are in a good agreement with the previous reports. For
example, Sheng et al reported that temperature of PBS only increased by 1.9 °C upon exposure to
a 808 nm laser irradiation with a power density of 1.0 W/cm2.4 The similar result was reported
by Li et al after irradiation of PBS only with 808 nm laser diode (8.0 W/cm2).24
16
Recently, Yang et al reported a highly efficient PTT agent, the polypyrrole organic
nanoparticles.54 It was demonstrated that the temperature of aqueous solution containing
polypyrrole organic nanoparticles (0.25 mg/mL) increased from ~ 15 °C to ~ 35°C (~∆T = 20 °C)
under 808 nm laser irradiation at a power density of 0.5 W/cm2. The Figure S4 indicates that the
developed SiNc-PNP demonstrate slightly lower photothermal conversion properties, which
could be related to a fact that SiNc-PNP can simultaneously generate fluorescence, heat and
ROS. The temperature of aqueous solution of SiNc-PNP (0.30 mg/mL) increased from ~ 25 °C to
~ 40°C (~∆T = 15 °C) under 785 nm laser irradiation at the same power density (Figure S4).
Photostability. The NIR organic dyes, including ICG, undergo irreversible photobleaching upon
photoirradiation, which may significantly affect the efficiency of fluorescence imaging and
phototherapy.8, 13, 25, 55 Therefore, we evaluated the potential photobleaching of SiNc loaded into
the PEG-PCL nanoparticles under extensive photoirradiation. The 30 min exposure of the SiNcPNP solution to NIR light (785 nm, 1.3 W/cm2) did not result in a visible reduction of SiNc
fluorescence or absorption intensity (Figure 2F). In addition, SiNc-PNP were capable of
elevating and upholding the temperature of a water solution at 54-55 oC upon irradiation for the
extended period of time (Figure 2E). In contrast, previous reports revealed that exposure of an
ICG solution to the laser light of a similar power density resulted in a significant reduction in
absorbance, fluorescence and photothermal properties, indicating photodegradation of ICG.8, 25 It
was also reported that encapsulation of the NIR dyes into the micelles only reduces their
photobleaching under photoirradiation as compared to free dyes.1 Finally, Sheng et al. indicated
that free ICG decomposes under storage conditions and loading of this dye into nanoparticles
only delays its degradation process.4
17
SiNc encapsulated into the developed PNP demonstrated a high optically stable behavior under
both extensive photoirradiation and storage conditions (Figure 2F and Figure S3), and thus this
nanoplatform has a high potential for theranostic applications. The superior photostability of
SiNc-PNP could be related to the following two reasons. First of all, the employed SiNc with
axial substituents is characterized by minimal photobleaching compared to unsubstituted
naphthalocyanines.8,
30
It was reported that absorption of free substituted SiNc was only
decreased by 10% upon exposure to a 785 nm laser diode (1.3 W/cm2) for 30 min.8 Additionally,
the loading of SiNc into the hydrophobic core of the polymeric nanoparticles further protects the
dye molecules from photodegradation and decomposition.4, 8
In vitro Theranostic Efficiency of SiNc-PNP. The Adriamycin-resistant A2780/AD human
ovarian carcinoma cells were employed to evaluate theranostic properties of SiNc-PNP in vitro.
Negligible cytotoxicity, a critical parameter for safe and effective phototherapy and bioimaging,
was observed for SiNc-PNP (2 – 100 g/mL) under dark conditions (Figure 3A). Since
fluorescence imaging, PDT and photothermal therapy (PTT) efficiency strongly rely on the
amount of phototherapeutic drug delivered into cancer cells,56 we quantitatively measured the
cellular uptake of SiNc-PNP with absorption spectroscopy. According to these measurements,
the quantity of SiNc delivered into cells by the polymeric nanoparticles was found to be 9.1 pg
per cell. Cellular internalization of SiNc-PNP was also verified with fluorescence microscopy by
taking advantage of SiNc intrinsic fluorescence (Figure 3B) and thus, confirming its imaging
ability. As shown in Figure 3B, the cancer cells emit bright fluorescence after incubation with
the SiNc-PNP.
To assess the effect of combinatorial phototherapy mediated by the developed SiNc-PNP in
vitro, A2780/AD cancer cells (5 ×106 cells) were incubated with the non-toxic concentration of
18
SiNc-PNP (50 g/mL, Figure 3A) and exposed to NIR irradiation (785 nm, 1.3 W/cm2) for 10
min. Cell viability measurements revealed more than 95% cancer cell death when compared to
the appropriate controls (Figure 3C). In contrast, laser irradiation alone or SiNc-PNP under dark
conditions were not capable of substantial killing of cancer cells, indicating safety and selectivity
of the combinatorial phototherapy. The obtained results are supported by the earlier publications
which pointed out that PTT has a synergistic effect with PDT, and only combinatorial
phototherapy can substantially induce cancer cell death as compared with PDT or PTT alone.2-4,
8, 14
To validate that both therapeutic mechanisms contributed to the destruction of SiNc-PNPtransfected cancer cells under exposure to a NIR laser, we evaluated alterations in intracellular
ROS and temperature during the combinatorial therapy. Following photoirradiation the
temperature inside the cell pellets increased by 17 oC reaching a value of 50 oC. In contrast,
temperature increase in non-treated cell pellets exposed to NIR light was minimal (<1 0C). In
addition to the photothermal therapy, it was further confirmed that the PDT mechanism was also
activated during the treatment. As indicated in Figure 3D, intracellular ROS level increased 26fold following irradiation of SiNc-PNP-treated cancer cells for 10 min with the 785 nm laser
light.
In Vivo Evaluation of Fluorescence Imaging and Phototherapy. Successful in vitro
assessment of SiNc-PNP as a NIR theranostic agent provided us with expectations for effective
in vivo bioimaging and cancer phototherapy applications. To assess the imaging feasibility and
biodistribution behavior of the developed SiNc-PNP, the intrinsic fluorescence of nanoparticleencapsulated SiNc after intravenous injection into nude mice bearing subcutaneous A2780/AD
tumors was monitored in vivo and ex vivo in various tissues at different time points. Figure 4
19
demonstrates that only weak autofluorescence was detected in the liver before SiNc-PNP
administration. After 30 min post injection, the entire body of the mouse emitted bright NIR
fluorescence indicating the distribution of SiNc-NP (Figure 4A). In addition, ex vivo images
revealed strong fluorescence intensity in major organs such as the liver, heart, kidneys, and lungs
(Figure 4B). After 2 h, strong fluorescence was still observed in the entire mouse body; however
fluorescence intensity significantly decreased in the liver, heart, kidneys and lungs,
demonstrating rapid clearance of SiNc-PNP from these normal organs (Figure 4). After 24 h, the
SiNc-loaded nanoparticles accumulated predominantly in tumors and provided high lesion-tonormal tissue contrast for sensitive fluorescence detection (Figure 4). The developed SiNc-PNP
have the optimal characteristics including small hydrodynamic size, neutral charge and
PEGylated surface, which provide an efficient passive tumor targeting and penetration.57 Li et al.
reported significant accumulation of PEGylated theranostic nanoparticles with a diameter of 32
nm in ovarian cancer tumors 24 h after intravenous administration.3 The cancer cell targeting
ligands could potentially improve internalization of the nanoparticles by cancer cells. However,
it will also increase complexity, complicate synthetic procedure, raise manufacturing cost and
therefore limit translational potential of these nanoplatforms.
Despite a fact that SiNc is not biodegradable molecule, SiNc was cleared from healthy organs
within 96 h after intravenous injection. This is an important characteristic for the developed
nanoplatform, because the US FDA requires all injected contrast agents to be completely cleared
from the body in a reasonable time period.45,
58
The tumor targeting capability and rapid
clearance of SiNc-PNP highlights the possibility of developing this unique nanoplatform for
phototherapy treatment.
20
To investigate the in vivo anticancer efficacy of the combinatorial phototherapy, mice bearing
subcutaneous xenograft of Adriamycin-resistant ovarian cancer cells were intravenously injected
with SiNc-PNP at the dose of 1.5 mg/kg. Under NIR fluorescence imaging guidance, which
confirmed SiNc accumulation in the tumor at 24-h post-injection, the laser beam (785 nm, 1.3
W/cm2) was precisely focused on the tumor and maintained for 10 min. The growth rate of
tumors was monitored for up to 25 days every 2-3 days and the day of injection was considered
as day 0 (Figure 5A). The untreated tumors and tumors treated with NIR light only and SiNcPNP without light (dark) grew rapidly, indicating that the A2780/AD tumor growth was not
affected by NIR laser alone or non-activated SiNc-PNP (Figure 5A). When the SiNc-PNPtreated tumors were exposed to the continuous irradiation with NIR laser beam for 10 min, the
tumor tissue was completely eradicated on day 3-5 post-injection, and there was no tumor
recurrence detected during 25 days.
Acute toxicity due to treatment with SiNc-PNP was determined in female athymic mice. Postinjection, mice were monitored for 25 days for changes in behavior, weight loss > 15 % and
death. During the course of the study none of the mice treated exhibited weight lost, changes in
behavior or died. Weight changes between the treatment and control groups are presented in
Figure 5B. As seen in the figure, none of the mice in the treatment groups were significantly
different as compared to the control mice, indicating that SiNc-PNP treatment shows no adverse
effects.
As demonstrated in Figure 5C, after exposure to NIR laser irradiation (1.30 W/cm2), the
intratumor temperature increased up to ~ 47 C, which is sufficient to ablate malignant cells. In
contrast, SiNc-PNP-free tumors did not demonstrate significant change in intratumoral
temperature under the same laser irradiation (∆T = 1.3 °C). Finally, not even slight skin burn was
21
noticed after irradiating the tumors for 15 min, suggesting that 785 nm laser light at a power
density 1.3 W/cm2 is safe for phototherapy. It is important to mention that NIR lasers with a
similar power density are widely used to conduct combinatorial phototherapy on living body. For
example, Li et al. reported that a 690 nm laser diode with a power density of 1.25 W/cm2 was
employed to perform combinatorial phototherapy in vivo.3 In another study, Li et al. employed a
808 nm laser at a power density of 1.4 W/cm2 to generate phototherapy in mice and the
temperature changes in saline-treated tumors was only ~ 2 °C.24
Based on our in vitro studies (Figure 3D) and the observed strong anticancer effect in vivo
(Figure 5A), we suggest that PDT contributed to the therapeutic effect of PTT and the
photoirradiation of SiNc-treated tumors resulted in generation of intracellular ROS. However,
the in vivo real time measurements of intratumoral ROS during the combinatorial phototherapy is
challenging to experimentally support this point and future studies are needed to further elucidate
this potential mechanism of tumor ablation.
The observed therapeutic outcome of the combinatorial phototherapy is in good agreement with
a previous report, which indicated that simultaneous PDT/PTT therapy mediated by ICG-loaded
nanoparticles resulted in complete inhibition of tumor growth with no detected cancer
recurrence.4
To investigate morphological changes and apoptotic processes in tumor tissues during the
phototherapeutic treatment and for corresponding controls, histological studies were carried out
in tumors excised from the mice. Changes in morphology in tumor tissue were evaluated using
Hematoxylin and eosin (H&E) staining for tumor sections. H&E staining showed that the
phototherapy treated sample had more condensed nuclei (dark purple) suggesting an increase in
cellular apoptosis as well as marked increase in tissue damage when compared to control
22
samples (Figure 6A). The Hoechst stained images in the control and light only tissue samples as
well as in dark sample all demonstrated a minimal amount of cells positive for nuclear
condensation and membrane blebbing all of which are hallmarks of cells undergoing apoptosis.59
The phototherapy treated sample conversely showed a pronounced increase in the number of
cells positive for apoptosis as indicated by the highly fluorescent cellular nuclei (Figure 6B).
Tumor tissue staining with a caspase-3 antibody further verified that only combinatorial
phototherapy result in the substantial apoptosis (red fluorescence) (Figure 6C).
Cumulatively, the above experiments demonstrated that PEG-PCL polymeric nanoparticles can
safely and selectively deliver SiNc to tumors after systemic administration while preserving its
fluorescence imaging and phototherapeutic properties. Due to high photostability, SiNc-PNP
offers accurate detection of the tumor tissue during prolong imaging time, which also allows for
precise treatment with NIR light with minimal side effects toward healthy organs.
CONCLUSIONS
Herein we have reported the development and use of biodegradable-biocompatible SiNc-loaded
polymeric nanoparticles for image-guided combinatorial phototherapy, which consist of only two
building blocks. The developed preparation procedure is simple, reproducible, and provides
highly uniform nanoparticles with the required structural features for enhanced passive tumor
targeting after systemic administration. SiNc-PNP demonstrated excellent photostability and
efficient generation of NIR fluorescence, ROS and hyperthermia under NIR light illumination
needed for tumor imaging and cancer eradication via combined PDT and PTT mechanisms. After
efficient tumor accumulation of SiNc-NP following intravenous injection, ovarian cancer tumors
were delineated with NIR fluorescence and completely eradicated after irradiation with a single
23
dose of 785 nm NIR light. Thus, our strategy provides cancer imaging with a single-agent
theranostic nanoplatform and subsequent phototherapeutic treatment with great potential for
clinical translation. We also expect that the developed simple approach can be employed for a
variety of NIR phototherapeutic and imaging agents with limited translational potential due to
their hydrophobic nature.
Acknowledgement. The authors are grateful for funding support provided by the OSU College of
Pharmacy, OSU Venture Development Fund, OSU General Research Fund and the Medical
Research Foundation of Oregon Health and Science University. Electron microscopy was
performed at the Multiscale Microscopy Core (MMC) with technical support from the Oregon
Health & Science University (OHSU)-FEI Living Lab and the OHSU Center for Spatial Systems
Biomedicine (OCSSB).
Supporting Information. Cryo-TEM-based size distribution histograms of SiNc-PNP,
absorption spectra of SiNc-PNP aqueous solution before and after storing at room temperature,
and temperature profiles of SiNc-PNP aqueous solution exposed to the 785 nm NIR laser diode
with various power densities. This material is available free of charge via the Internet at
http://pubs.acs.org.
24
Table of Contents
25
Figures and Tables
=
SiNc-PNP
SiNc
+
PEG-PCL
leaky blood
vessel
Imaging
Phototherapy
Figure 1. Schematic illustration of a single-agent based theranostic nanoplatform (SiNc-PNP)
designed for image-guided combinatorial phototherapy. The nanoplatform is prepared by
encapsulation of a theranostic agent, SiNc, into the hydrophobic interior of PEG-PCL polymeric
nanoparticles. After intratumoral accumulation of SiNc-PNP, presumably facilitated by passive
targeting, the tumor tissue can be detected via NIR fluorescence imaging. When irradiated with
the NIR laser, SiNc-PNP transform the absorbed light energy into toxic ROS and heat to
eradicate tumor tissue.
26
Volume, %
20
Freshly prepared
30 days
A
1000
B
Fluorescence, a.u.
25
15
10
5
0
600
400
200
0
750
50 nm
10 20 30 40 50 60 70 80
C
800
Water
SiNc-PNP
200
0
Dark
80
Light
Water
SiNc-PNP
E
Absorbance, a.u.
400
90
**
D
Temperature,C
SOSG Fluorescence, a.u.
600
800
850
900
Wavelength, nm
Size, nm
70
60
50
40
30
20
0
5
10
15
Time, min
20
1.0
F
0.5
Before
After
0.0
600
700
800
Wavelength, nm
900
Figure 2. (A) Size distribution of SiNc-PNP measured by Dynamic Light Scattering before and
after storing for 30 days at room temperature. The inset shows the actual photo of vial with 0.3
mg/mL SiNc-PNP in aqueous solution. (B) Representative cryo-TEM image of SiNc-PNP. (C)
Fluorescence spectrum (ex = 750 nm) of SiNc-PNP water solution (SiNc = 0.05 mg/mL). The
inset represents fluorescence image of SiNc-PNP solution recorded with Pearl Imaging System at
800 nm. (D) Singlet oxygen generation by SiNc-PNP (0.02 mg/mL) without (dark) and in the
presence of NIR irradiation (light) (785 nm, 1.3 W/cm2, 5 min). **P < 0.005 when compared
with water as control. (E) Temperature profiles of water (black) and SiNc-PNP (0.3 mg/mL)
aqueous solution (red) exposed to the NIR light (785 nm, 1.3 W/cm2). (F) Absorption spectra
and fluorescence images (Insets) of SiNc-PNP aqueous solution before and after irradiation with
NIR light (30 min, 1.3 W/cm2).
27
Cell viability,%
120
100
E
B
A
cellular
internalization
SiNc-PNP
80
60
40
20
0
cell pellet
formation
2.0 5.0 10.0 25.0 50.0 100.0
Cell viability, %
150
120
C
Control
NIR light
SiNc-PNP dark
SiNc-PNP light
90
60
30
0
**
ROS increase, a.u.
SiNc concentration, g/mL
2000
1500
**
D
ROS
increase
Cells control
SiNc-PNP
1000
500
0
T ~ 50 C
Dark
Light
Figure 3. (A) Viability of A2780/AD ovarian carcinoma cells treated with SiNc-PNP under dark
conditions. (B) Representative fluorescence microscopy image demonstrating intracellular
accumulation of SiNc-PNP. (C) Combinatorial (PDT+PTT) therapeutic effect against A2780/AD
ovarian carcinoma cells. The cells were treated with 50 μg/mL of SiNc-PNP and exposed to a
laser light for 10 min (785 nm, 1.3 W/cm2). **P < 0.005 when compared to the controls. (D)
ROS generation in SiNc-PNP-treated A2780/AD cells following NIR light irradiation (785 nm,
1.3 W/cm2). **P < 0.005 when compared the controls. (E) In vitro evaluation of the
combinatorial photherapy mediated by the SiNc-PNP.
28
A
B
Liver
Kidneys
Spleen
Heart
Lungs
Tumor
Before
1.0
0.5 h
2h
24 h
0.0
96 h
Figure 4. The biodistribution of SiNc-PNP in nude mice. In vivo NIR fluorescence imaging of
the mice bearing A2780/AD tumor (A) and ex vivo imaging of SiNc-PNP in liver, kidneys,
spleen, heart, lungs, and tumor (B) after i.v. injected with SiNc-PNP (1.5 mg/kg) before and at
0.5, 2, 24 and 96 hours post-injection, respectively.
29
Tumor volume, mm3
1500
A
1200
900
600
300
0
0
Body weight, g
5
10
15
20
Time, days
25
B
25
20
Control
NIR light
SiNc-PNP dark
SiNc-PNP light
15
10
Intratumoral temperature, oC
Control
NIR light
SiNc-PNP dark
SiNc-PNP light
0
55
5
10
15
20
Time, days
NIR light
SiNc-PNP light
C
50
25
45
Laser “off”
40
35
30
Laser “on”
0
250
500
Time, s
750
1000
Figure 5. (A) Tumor growth profiles in mice with A2780/AD xenografts after the following
treatments (a) Control (black curve): no treatment, (b) NIR light (red curve): tumors irradiated
with a laser light (785 nm, 1.3 W/cm2) for 10 min, (c) SiNc-PNP dark (blue curve): mice
intravenously injected with SiNc-PNP (1.5 mg/kg) without light exposure and (d) SiNc-PNP
light (green curve): mice intravenously injected with SiNc-PNP (1.5 mg/kg) and tumors were
irradiated for 10 min with a laser light (785 nm, 1.3 W/cm2). (B) Changes in body weights of
treated mice compared to control groups. Means ± SD are shown. Arrows represent one injection
on day 0. (C). Intratumoral temperature profile during the phototherapy mediated by SiNc-PNP
30
(1.5 mg/kg) in combination with NIR light (red curve). Tumors irradiated with the NIR laser
light only were employed as the control (black curve).
Control
NIR light
SiNc-PNP dark
SiNc-PNP light
H&E
A
Hoechst
B
Caspase-3
C
Figure 6. Representative images of H&E (A), Hoechst (B), and caspase-3 (C) stained tumor
sections harvested from the mice after the following treatments: control (no treatment), NIR light
(tumor irradiated with a laser (785 nm, 1.3 W/cm2)), SiNc-PNP dark (mice treated with SiNcPNP without light exposure) and SiNc-PNP light (phototherapy).
Table 1. Stability of the SiNc-PNP
Time
Size, nm
PDI
Zeta potential, mV
SiNc loading, mg/mL
Day, 0
Day, 30
37.66 ± 0.26
37.16 ± 0.83
0.06 ± 0.02
0.08 ± 0.01
-2.76 ± 1.83
-2.84 ± 1.68
0.30 ± 0.01
0.29 ± 0.02
31
REFERENCES
(1).
Guo, M.; Mao, H.; Li, Y.; Zhu, A.; He, H.; Yang, H.; Wang, Y.; Tian, X.; Ge, C.; Peng,
Q.; Wang, X.; Yang, X.; Chen, X.; Liu, G.; Chen, H. Dual imaging-guided
photothermal/photodynamic therapy using micelles. Biomaterials 2014, 35, 4656-4666.
(2).
Jang, B.; Park, J. Y.; Tung, C. H.; Kim, I. H.; Choi, Y. Gold nanorod-photosensitizer
complex for near-infrared fluorescence imaging and photodynamic/photothermal therapy in vivo.
ACS Nano 2011, 5, 1086-1094.
(3).
Li, Y.; Lin, T. Y.; Luo, Y.; Liu, Q.; Xiao, W.; Guo, W.; Lac, D.; Zhang, H.; Feng, C.;
Wachsmann-Hogiu, S.; Walton, J. H.; Cherry, S. R.; Rowland, D. J.; Kukis, D.; Pan, C.; Lam, K.
S. A smart and versatile theranostic nanomedicine platform based on nanoporphyrin. Nat.
Commun. 2014, 5, 4712.
(4).
Sheng, Z.; Hu, D.; Zheng, M.; Zhao, P.; Liu, H.; Gao, D.; Gong, P.; Gao, G.; Zhang, P.;
Ma, Y.; Cai, L. Smart human serum albumin-indocyanine green nanoparticles generated by
programmed assembly for dual-modal imaging-guided cancer synergistic phototherapy. ACS
Nano 2014, 8, 12310-12322.
(5).
Chen, Q.; Liang, C.; Wang, X.; He, J.; Li, Y.; Liu, Z. An albumin-based theranostic
nano-agent for dual-modal imaging guided photothermal therapy to inhibit lymphatic metastasis
of cancer post surgery. Biomaterials 2014, 35, 9355-9362.
(6).
Chu, M.; Li, H.; Wu, Q.; Wo, F.; Shi, D. Pluronic-encapsulated natural chlorophyll
nanocomposites for in vivo cancer imaging and photothermal/photodynamic therapies.
Biomaterials 2014, 35, 8357-8373.
(7).
Taratula, O.; Patel, M.; Schumann, C.; Naleway, M. A.; Pang, A. J.; He, H.
Phthalocyanine-loaded graphene nanoplatform for imaging-guided combinatorial phototherapy.
Int. J. Nanomedicine 2015, 10, 2347-2362.
(8).
Taratula, O.; Schumann, C.; Duong, T.; Taylor, K. L. Dendrimer-encapsulated
naphthalocyanine as a single agent-based theranostic nanoplatform for near-infrared fluorescence
imaging and combinatorial anticancer phototherapy. Nanoscale 2015, 7, 3888-3902.
(9).
Gottesman, M. M. Mechanisms of cancer drug resistance. Ann. Rev. Med. 2002, 53, 615627.
(10). Wang, Y.; Wang, H.; Liu, D.; Song, S.; Wang, X.; Zhang, H. Graphene oxide covalently
grafted upconversion nanoparticles for combined NIR mediated imaging and
photothermal/photodynamic cancer therapy. Biomaterials 2013, 34, 7715-7724.
(11). Gollavelli, G.; Ling, Y. C. Magnetic and fluorescent graphene for dual modal imaging
and single light induced photothermal and photodynamic therapy of cancer cells. Biomaterials
2014, 35, 4499-4507.
(12). Frangioni, J. V. In vivo near-infrared fluorescence imaging. Curr. Opin. Chem. Biol.
2003, 7, 626-634.
(13). Gibbs, S. L. Near infrared fluorescence for image-guided surgery. Quant. Imaging Med.
Surg. 2012, 2, 177-187.
(14). Tian, B.; Wang, C.; Zhang, S.; Feng, L.; Liu, Z. Photothermally enhanced photodynamic
therapy delivered by nano-graphene oxide. ACS Nano 2011, 5, 7000-7009.
(15). Singh, A. K.; Hahn, M. A.; Gutwein, L. G.; Rule, M. C.; Knapik, J. A.; Moudgil, B. M.;
Grobmyer, S. R.; Brown, S. C. Multi-dye theranostic nanoparticle platform for bioimaging and
cancer therapy. Int. J. Nanomedicine 2012, 7, 2739-2750.
32
(16). Sevick-Muraca, E. M.; Rasmussen, J. C. Molecular imaging with optics: primer and case
for near-infrared fluorescence techniques in personalized medicine. J. Biomed. Opt. 2008, 13,
041303.
(17). Adams, K. E.; Ke, S.; Kwon, S.; Liang, F.; Fan, Z.; Lu, Y.; Hirschi, K.; Mawad, M. E.;
Barry, M. A.; Sevick-Muraca, E. M. Comparison of visible and near-infrared wavelengthexcitable fluorescent dyes for molecular imaging of cancer. J. Biomed. Opt. 2007, 12, 024017.
(18). Marshall, M. V.; Rasmussen, J. C.; Tan, I. C.; Aldrich, M. B.; Adams, K. E.; Wang, X.;
Fife, C. E.; Maus, E. A.; Smith, L. A.; Sevick-Muraca, E. M. Near-infrared fluorescence imaging
in humans with indocyanine green: a review and update. Open Surg. Oncol. J. 2010, 2, 12-25.
(19). Ethirajan, M.; Chen, Y.; Joshi, P.; Pandey, R. K. The role of porphyrin chemistry in
tumor imaging and photodynamic therapy. Chem. Soc. Rev. 2011, 40, 340-362.
(20). Bahmani, B.; Guerrero, Y.; Bacon, D.; Kundra, V.; Vullev, V. I.; Anvari, B.
Functionalized polymeric nanoparticles loaded with indocyanine green as theranostic materials
for targeted molecular near infrared fluorescence imaging and photothermal destruction of
ovarian cancer cells. Lasers Surg. Med. 2014, 46, 582-592.
(21). Zhao, P.; Zheng, M.; Yue, C.; Luo, Z.; Gong, P.; Gao, G.; Sheng, Z.; Zheng, C.; Cai, L.
Improving drug accumulation and photothermal efficacy in tumor depending on size of ICG
loaded lipid-polymer nanoparticles. Biomaterials 2014, 35, 6037-6046.
(22). Wang, Y. G.; Kim, H.; Mun, S.; Kim, D.; Choi, Y. Indocyanine green-loaded
perfluorocarbon nanoemulsions for bimodal (19)F-magnetic resonance/nearinfrared fluorescence
imaging and subsequent phototherapy. Quant. Imaging Med. Surg. 2013, 3, 132-140.
(23). Zheng, X.; Xing, D.; Zhou, F.; Wu, B.; Chen, W. R. Indocyanine green-containing
nanostructure as near infrared dual-functional targeting probes for optical imaging and
photothermal therapy. Mol. Pharm. 2011, 8, 447-456.
(24). Li, Y.; Wen, T.; Zhao, R.; Liu, X.; Ji, T.; Wang, H.; Shi, X.; Shi, J.; Wei, J.; Zhao, Y.;
Wu, X.; Nie, G. Localized electric field of plasmonic nanoplatform enhanced photodynamic
tumor therapy. ACS Nano 2014, 8, 11529-11542.
(25). Mathew, S.; Murakami, T.; Nakatsuji, H.; Okamoto, H.; Morone, N.; Heuser, J. E.;
Hashida, M.; Imahori, H. Exclusive photothermal heat generation by a gadolinium
bis(naphthalocyanine) complex and inclusion into modified high-density lipoprotein nanocarriers
for therapeutic applications. ACS Nano 2013, 7, 8908-8916.
(26). Russin, T. J.; Altinoglu, E. I.; Adair, J. H.; Eklund, P. C. Measuring the fluorescent
quantum efficiency of indocyanine green encapsulated in nanocomposite particulates. J. Phys.
Condens. Matter. 2010, 22, 334217.
(27). Jin, Y. H.; Ye, F. M.; Zeigler, M.; Wu, C. F.; Chiu, D. T. Near-infrared fluorescent dyedoped semiconducting polymer dots. ACS Nano 2011, 5, 1468-1475.
(28). Song, L. P.; Li, H.; Sunar, U.; Chen, J.; Corbin, I.; Yodh, A. G.; Zheng, G.
Naphthalocyanine-reconstituted LDL nanoparticles for in vivo cancer imaging and treatment. Int.
J. Nanomedicine 2007, 2, 767-774.
(29). Josefsen, L. B.; Boyle, R. W. Unique diagnostic and therapeutic roles of porphyrins and
phthalocyanines in photodynamic therapy, imaging and theranostics. Theranostics 2012, 2, 916966.
(30). Ali, H.; van Lier, J. E. Metal complexes as photo- and radiosensitizers. Chem. Rev. 1999,
99, 2379-2450.
33
(31). Letchford, K.; Burt, H. A review of the formation and classification of amphiphilic block
copolymer nanoparticulate structures: micelles, nanospheres, nanocapsules and polymersomes.
Eur. J. Pharm. Biopharm. 2007, 65, 259-269.
(32). Wei, X.; Gong, C.; Gou, M.; Fu, S.; Guo, Q.; Shi, S.; Luo, F.; Guo, G.; Qiu, L.; Qian, Z.
Biodegradable poly(epsilon-caprolactone)-poly(ethylene glycol) copolymers as drug delivery
system. Int. J. Pharm. 2009, 381, 1-18.
(33). Marsden, H. R.; Gabrielli, L.; Kros, A. Rapid preparation of polymersomes by a water
addition/solvent evaporation method. Polymer Chem. 2010, 1, 1512 - 1518.
(34). Taratula, O.; Dani, R. K.; Schumann, C.; Xu, H.; Wang, A.; Song, H.; Dhagat, P.;
Taratula, O. Multifunctional nanomedicine platform for concurrent delivery of chemotherapeutic
drugs and mild hyperthermia to ovarian cancer cells. Int. J. Pharm. 2013, 458, 169-180.
(35). Taratula, O.; Schumann, C.; Naleway, M. A.; Pang, A. J.; Chon, K. J.; Taratula, O. A
multifunctional theranostic platform based on phthalocyanine-loaded dendrimer for imageguided drug delivery and photodynamic therapy. Mol. Pharm. 2013, 10, 3946-3958.
(36). Dani, R. K.; Schumann, C.; Taratula, O. Temperature-tunable iron oxide nanoparticles
for remote-controlled drug release. AAPS PharmSciTech 2014, 15, 963-972.
(37). Ding, M.; Zhou, L.; Fu, X.; Tan, H.; Li, J.; Fu, Q. Biodegradable gemini multiblock
poly(ε-caprolactone urethane)s toward controllable micellization. . Soft Matter 2010, 6, 20872092.
(38). Philip, R.; Penzkofer, A.; Biiumler, W.; Szeimies, R. M.; Abels, C. Absorption and
fluorescence spectroscopic investigation of indocyanine green. J. Photocbem. Photobiol. A:
Chem. 1996, 96, 137-148.
(39). Davies-Venn, C. A.; Angermiller, B.; Wilganowski, N.; Ghosh, P.; Harvey, B. R.; Wu,
G.; Kwon, S.; Aldrich, M. B.; Sevick-Muraca, E. M. Albumin-binding domain conjugate for
near-infrared fluorescence lymphatic imaging. Mol. Imaging Biol. 2012, 14, 301-314.
(40). Zhou, S.; Deng, X.; Yang, H. Biodegradable poly(epsilon-caprolactone)-poly(ethylene
glycol) block copolymers: characterization and their use as drug carriers for a controlled delivery
system. Biomaterials 2003, 24, 3563-3570.
(41). Sun, T. M.; Du, J. Z.; Yan, L. F.; Mao, H. Q.; Wang, J. Self-assembled biodegradable
micellar nanoparticles of amphiphilic and cationic block copolymer for siRNA delivery.
Biomaterials 2008, 29, 4348-4355.
(42). Gong, C. Y.; Wu, Q. J.; Dong, P. W.; Shi, S.; Fu, S. Z.; Guo, G.; Hu, H. Z.; Zhao, X.;
Wei, Y. Q.; Qian, Z. Y. Acute toxicity evaluation of biodegradable in situ gel-forming controlled
drug delivery system based on thermosensitive PEG-PCL-PEG hydrogel. J. Biomed. Mater. Res.
B Appl. Biomater. 2009, 91, 26-36.
(43). Lin, W. J.; Chen, Y. C.; Lin, C. C.; Chen, C. F.; Chen, J. W. Characterization of
pegylated copolymeric micelles and in vivo pharmacokinetics and biodistribution studies. J.
Biomed. Mater. Res. B Appl. Biomater. 2006, 77, 188-194.
(44). Mudshinge, S. R.; Deore, A. B.; Patil, S.; Bhalgat, C. M. Nanoparticles: emerging
carriers for drug delivery. Saudi Pharm. J. 2011, 19, 129-141.
(45). Longmire, M.; Choyke, P. L.; Kobayashi, H. Clearance properties of nano-sized particles
and molecules as imaging agents: considerations and caveats. Nanomedicine 2008, 3, 703-717.
(46). Maruyama, K. Intracellular targeting delivery of liposomal drugs to solid tumors based
on EPR effects. Adv. Drug Deliv. Rev. 2011, 63, 161-169.
(47). Alexis, F.; Pridgen, E.; Molnar, L. K.; Farokhzad, O. C. Factors affecting the clearance
and biodistribution of polymeric nanoparticles. Mol. Pharm. 2008, 5, 505-515.
34
(48). Huo, S.; Ma, H.; Huang, K.; Liu, J.; Wei, T.; Jin, S.; Zhang, J.; He, S.; Liang, X. J.
Superior penetration and retention behavior of 50 nm gold nanoparticles in tumors. Cancer Res.
2013, 73, 319-330.
(49). Duan, X.; Li, Y. Physicochemical characteristics of nanoparticles affect circulation,
biodistribution, cellular internalization, and trafficking. Small 2013, 9, 1521-1532.
(50). Petros, R. A.; DeSimone, J. M. Strategies in the design of nanoparticles for therapeutic
applications. Nat. Rev. Drug Discov. 2010, 9, 615-627.
(51). Azhdarinia, A.; Wilganowski, N.; Robinson, H.; Ghosh, P.; Kwon, S.; Lazard, Z. W.;
Davis, A. R.; Olmsted-Davis, E.; Sevick-Muraca, E. M. Characterization of chemical,
radiochemical and optical properties of a dual-labeled MMP-9 targeting peptide. Bioorg. Med.
Chem. 2011, 19, 3769-3776.
(52). Ohnishi, S.; Lomnes, S. J.; Laurence, R. G.; Gogbashian, A.; Mariani, G.; Frangioni, J.
V. Organic alternatives to quantum dots for intraoperative near-infrared fluorescent sentinel
lymph node mapping. Mol. Imaging 2005, 4, 172-181.
(53). Kim, S.; Fujitsuka, M.; Majima, T. Photochemistry of Singlet Oxygen Sensor Green. J..
Phys. Chem. B 2013, 117, 13985-13992.
(54). Yang, K.; Xu, H.; Cheng, L.; Sun, C.; Wang, J.; Liu, Z. In vitro and in vivo near-infrared
photothermal therapy of cancer using polypyrrole organic nanoparticles. Adv. Mater. 2012, 24,
5586-5592.
(55). Gioux, S.; Choi, H. S.; Frangioni, J. V. Image-guided surgery using invisible nearinfrared light: fundamentals of clinical translation. Mol. Imaging 2010, 9, 237-255.
(56). Almeida, R. D.; Manadas, B. J.; Carvalho, A. P.; Duarte, C. B. Intracellular signaling
mechanisms in photodynamic therapy. Biochim.Biophys. Acta-Rev. Cancer 2004, 1704, 59-86.
(57). Alexis, F.; Pridgen, E.; Molnar, L. K.; Farokhzad, O. C. Factors affecting the clearance
and biodistribution of polymeric nanoparticles. Mol. Pharm. 2008, 5, 505-515.
(58). Choi, H. S.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J. P.; Itty Ipe, B.; Bawendi, M. G.;
Frangioni, J. V. Renal clearance of quantum dots. Nat. Biotechnol. 2007, 25, 1165-1170.
(59). Fazal, N.; Al-Ghoul, W. M. Thermal injury-plus-sepsis contributes to a substantial
deletion of intestinal mesenteric lymph node CD4 T cell via apoptosis. Int. J. Biol. Sci. 2007, 3,
393-401.
35
Download