Next Article in Journal
Assessment of the Influence of the Selected Range of Visible Light Radiation on the Durability of the Gel with Ascorbic Acid and Its Derivative
Next Article in Special Issue
Cross-Linked Hyaluronate and Corticosteroid Combination Ameliorate the Rat Experimental Tendinopathy through Anti-Senescent and -Apoptotic Effects
Previous Article in Journal
Genomic Insights of First ermB-Positive ST338-SCCmecVT/CC59 Taiwan Clone of Community-Associated Methicillin-Resistant Staphylococcus aureus in Poland
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

NIR-Triggered Generation of Reactive Oxygen Species and Photodynamic Therapy Based on Mesoporous Silica-Coated LiYF4 Upconverting Nanoparticles

1
Department of Chemical and Materials Engineering, National Kaohsiung University of Science and Technology, Kaohsiung 807, Taiwan
2
Department of Chemical and Materials Engineering, National University of Kaohsiung, Kaohsiung 811, Taiwan
3
Graduate Institute of Natural Products, Kaohsiung Medical University, Kaohsiung 807, Taiwan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(15), 8757; https://doi.org/10.3390/ijms23158757
Submission received: 19 July 2022 / Revised: 2 August 2022 / Accepted: 3 August 2022 / Published: 6 August 2022
(This article belongs to the Special Issue Materials for Bioengineering and Biotechnology)

Abstract

:
To date, the increase in reactive oxygen species (ROS) production for effectual photodynamic therapy (PDT) treatment still remains challenging. In this study, a facile and effective approach is utilized to coat mesoporous silica (mSiO2) shell on the ligand-free upconversion nanoparticles (UCNPs) based on the LiYF4 host material. Two kinds of mesoporous silica-coated UCNPs (UCNP@mSiO2) that display green emission (doped with Ho3+) and red emission (doped with Er3+), respectively, were successfully synthesized and well characterized. Three photosensitizers (PSs), merocyanine 540 (MC 540), rose bengal (RB), and chlorin e6 (Ce6), with the function of absorption of green or red emission, were selected and loaded into the mSiO2 shell of both UCNP@mSiO2 nanomaterials. A comprehensive study for the three UCNP@mSiO2/PS donor/acceptor pairs was performed to investigate the efficacy of fluorescence resonance energy transfer (FRET), ROS generation, and in vitro PDT using a MCF-7 cell line. ROS generation detection showed that as compared to the oleate-capped and ligand-free UCNP/PS pairs, the UCNP@mSiO2/PS nanocarrier system demonstrated more pronounced ROS generation due to the UCNP@mSiO2 nanoparticles in close vicinity to PS molecules and a higher loading capacity of the photosensitizer. As a result, the three LiYF4 UCNP@mSiO2/PS nanoplatforms displayed more prominent therapeutic efficacies in PDT by using in vitro cytotoxicity tests.

1. Introduction

In recent decades, the development of nanotechnology resulted in the design of various nanomaterials for biological and therapeutic applications such as gene therapy, tissue engineering, drug delivery, molecular imaging, etc. [1,2,3,4]. In comparison with conventional therapies, nanomaterials are more efficient for cancer therapy because they can be designed appropriately to target affected tumor sites and discriminatorily distribute their loads with fewer side effects. Among the various nanomaterials, mesoporous silica nanoparticles (MSNs) have emerged as favorable inorganic material platforms for biomedical applications since 2001 [5,6,7]. Due to their large surface areas, tunable sizes, high loading capacity, thermal and photostability, facile modification and considerable biocompatibility, mesoporous silica (mSiO2) functionalized nanocomposites have been widely used in many biomedical applications such as drug delivery, photodynamic therapy (PDT), etc. [8,9,10,11,12,13,14,15]. Compared to other kind of nanoparticles, the unique properties of these inorganic nanomaterials allowed for building efficient complex nanostructures. Consequently, mesoporous silica nanoparticles such as MCM-41, MCM-48, SBA-15, and core-shell and hollow MSNs have been used in drug delivery due to their distinctive properties [16,17,18,19].
Meanwhile, there has been an awareness of the medical potential of PDT for more than 25 years. PDT pertaining to the photochemical reactions of photosensitizers (PSs) is a treatment approach utilizing light irradiation at proper wavelengths. In a PDT treatment, the photosensitizers are stimulated and produce reactive oxygen species (ROS) to give rise to oxidative stress and cellular damages [20]. Compared to the conventional cancer therapy, PDT has several advantages including precise tumor targeting and negligible systemic toxicity; it is less invasive than surgery, and is accessible for repeated therapies [21,22,23]. Among them, the utmost advantage of PDT can be attributed to its selective therapeutics for tumor cells by using appropriate light under specific control. However, most PS molecules need to be motivated by ultraviolet or visible light that has poor tissue-penetration depth, and thus restricts the phototherapy of large or internal tumors. This drawback can be avoided by adsorbing the photosensitizers on the surface or outer shell of upconversion nanoparticles (UCNPs). Upconversion nanoparticles can absorb two or more photons and give off the light at shorter wavelength than the excitation wavelength. Consequently, the utilization of UCNPs as nanocarriers facilitates the development of near-infrared (NIR) light-triggered PDT. This treatment is in relation to upconverted fluorescence emissions of UCNPs and fluorescence resonance energy transfer (FRET) between UCNPs and the photosensitizers. Several approaches have been employed to load the PS molecules onto the surface of UCNPs, which include coating with a mesoporous silica layer and encapsulating in a polymer shell [24,25,26,27,28]. To secure more efficient remedial effects, it is suggested to exploit intense and appropriate upconversion (UC) emission at certain wavelengths and achieve a higher loading capacity of the PS molecules. To intensify the selectivity with respect to tumor cells and the efficiency of PDT, fabricating UCNP-PS nanocarrier systems with high therapeutic efficacy by using more efficient and powerful synthetic strategies remains challenging.
In recent years, lanthanide-doped upconversion nanoparticles (UCNPs) have gained tremendous focus for various applications such as biological imaging, therapeutics, photovoltaics, photonics, etc. [29,30,31,32,33,34]. Due to the bounteous and unique energy levels, lanthanide-doped UCNPs can display numerous distinctive characteristics including low autofluorescence, narrow bandwidths, long luminescence lifetime, less photobleaching, large anti-Stokes shift, superior photostability, and deep penetration depth for tissue [33,35,36,37,38,39,40,41,42]. The intrinsic photophysical and photochemical properties of UCNPs render them particularly applicable for phototherapy. Over the past decade, combining UCNPs and photosensitizers as theranostic platforms have been widely applied in PDT via in vitro and in vivo tests [43,44]. To endow their promising versatility for practical applications, a wide range of surface modification methods has been adopted for the design of nanocarriers with distinctive physico-chemical, toxicological, and pharmacological properties [31,45,46]. In particular, studies on light-controlled drug delivery systems based on MSNs have been executed extensively because light under specific control can provide well-regulated drug release both spatially and temporally, and thus present great potentials for further biomedical applications. Taking the advantages of UCNPs in phototherapy, the functionalization of the UCNP surface for various biomedical applications is intriguing for researchers in the clinical theranostic field.
In a UCNP—PS nanocarrier system, UCNPs act as donors, while the PS molecules function as acceptors to generate ROS. Through the FRET process between UCNPs and photosensitizers, ROS can be generated and used in PDT [47,48,49]. Since the efficacy of PDT treatment depends on the efficiency of the corresponding ROS generation, the loading capacity of PS molecules on the surface of UCNPs and selection of the photosensitizer for effective FRET between UCNP—PS (donor-acceptor) system is of crucial importance. In this study, we report a more effective approach for PDT treatment based on the surface modification of LiYF4 UCNPs by using MSNs as compared to the PDT efficacy via the oleate-capped and ligand-free type UCNPs in our previous study [50]. To the best of our knowledge, it is hard to find relevant literature with regard to ROS generation and PDT treatment through the donor-acceptor FRET pair of the LiYF4 type UCNP@mSiO2 nanomaterials and the organic dye-based photosensitizers. It is suggested that the comparison of the PDT efficacy of the three kinds of UCNPs with different surface morphologies could be applied in all UCNP—PS nanocarrier systems. Both green emission UCNPs and red emission UCNPs composed of LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2 and LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2 core/shell nanoparticles were synthesized, respectively. In order to establish an effective upconversion resonance energy transfer process, two photosensitizers, merocyanine 540 (MC 540) and rose bengal (RB), acted as the energy acceptor of green emission, and the other photosensitizer, chlorin e6 (Ce6), utilized as the energy acceptor of red emission were selected for this kind of donor—acceptor system. The in vitro cytotoxicity tests were also performed to evaluate the efficacy of the MSN-based surface modification for UCNPs in PDT.

2. Materials and Methods

2.1. Synthesis of Mesoporous Silica-Coated LiYF4:Yb3+/Ho3+@LiYF4:Yb3+ and LiYF4:Yb3+/Er3+@ LiYF4:Yb3+ UCNPs

Two kinds of LiYF4 UCNPs capped with oleic acid and doped with Yb3+/Ho3+ (green emission) and Yb3+/Er3+ (red emission) ions were synthesized, respectively, in accordance with the previously reported method [51,52]. Subsequently, the removal of both oleate ligand from the core/shell type UCNPs composed of LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2 and LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2 were carried out via the strategy proposed by the Capobianco group [53]. Upon the preparation of oleate-free UCNPs, the mesoporous silica was then coated on the surface of UCNPs (UCNP@mSiO2) (Scheme 1). The experiment was achieved through a NaOH-based etching method. In a typical experiment, 0.25 g of hexadecyl trimethyl ammonium chloride (CTAC) and 0.025 g of ligand-free UCNPs were dispersed in a 50 mL aqueous solution under ultrasonic vibrating and then the temperature was raised to 75 °C. After one hour, 0.15 mL of tetraethyl orthosilicate (TEOS) was added dropwise and then 0.12 mL of NaOH aqueous solution was added into the flask. The resulting mixture was stirred for another 1 h. The reaction was added with an aliquot of ethanol to obtain the crude UCNP@mSiO2 (CTAC) via centrifugation. The crude product was redispersed in 2 wt% NaCl aqueous solution to remove CTAC and then recovered by centrifugation. This step was repeated three times to acquire the final product. The UCNP@mSiO2 nanomaterials with the function of green emission and red emission were prepared for this work. The final products were dispersed and stored in deionized water and ready for use.

2.2. Adsorption of Photosensitizers into the Hollow Pores of UCNP@mSiO2

In this study, 6 mg of photosensitizers merocyanine 540 (MC 540), rose bengal (RB), and chlorin e6 (Ce6) were individually loaded into the pores of the mesoporous silica by soaking 10 mg of their corresponding UCNP@mSiO2 nanoparticles in 10 mL of solvent for 24 h at room temperature. Deionized (DI) water was used as the solvent of MC 540 and RB, while ethanol was used as the solvent of Ce6. The nanoparticles were then centrifuged and washed with the solvent (DI water or ethanol) three times to remove non-adsorbed molecules of the photosensitizers. The loading capacity of three photosensitizers was obtained from the calibration curve of their UV-vis absorption spectra at the corresponding peaks. The loading capacity of each photosensitizer was calculated as follow: loading capacity (%, w/w) = (weight amount of photosensitizer incorporated into nanoparticles)/(weight amount of nanoparticles) × 100%.

2.3. NIR-Induced Reactive Oxygen Species (ROS) Generation

The ROS production upon NIR irradiation was determined by using a 9,10-anthracenediyl-bis(methylene) dimalonic acid (ABDA) probe in the aqueous solution. Two mesoporous silica- coated UCNPs loaded with three different photosensitizers can be all designated as UCNP@mSiO2/PS or individually denoted as Ho3+-doped UCNP@mSiO2/MC 540, Ho3+-doped UCNP@mSiO2/RB, and Er3+-doped UCNP@mSiO2/Ce6. The above three nanohybrids were dispersed in DI water with a concentration of 1 mg/mL, while the ABDA solution was prepared in a concentration of 3 mg/mL. Then 2.5 mL of each nanohybrid aqueous solution was mixed with 0.5 mL of ABDA solution. The solutions were placed in quartz cuvettes and irradiated for 60 min using 3 W/cm2 980 nm laser. The absorption intensities for the three peaks at 359 nm, 378 nm and 399 nm were recorded every 10 min interval by UV-vis spectrophotometer. ROS generation was investigated by detecting the absorptions of ABDA at the three peaks.

2.4. Cytotoxicity Tests of Cancer Cells

Cytotoxicity tests were measured using CCK-8 kits. The MCF-7 cells were seeded in 96-well plates. After cultivation for 24 h, 100 μL of UCNP@mSiO2/PS nanomaterials was added into the culture medium at different concentrations, with five parallel wells for each concentration (0, 50, 100, 250, and 500 μg/mL). After irradiation with 980 nm NIR for 5 min, the cells were then incubated for another 24 h in an incubator and added with 10 μL of CCK-8 and 90 μL of medium assay per well. After 3 h of incubation, the cell viability was obtained by an ELISA reader.

2.5. Characterization

X-ray photoelectron spectroscopy (XPS) analysis was performed by a JEOL JAMP-9500F Auger electron spectrometer. Energy-dispersive X-ray spectroscopy (EDS) was conducted in a FEI Genesis XM 4i energy dispersive X-ray analysis system. Wide-angle X-ray diffraction (WAXD) analysis were performed with a Bruker D8 ADVANCE diffractometer, using Cu Kα radiation with a step size of 0.05° and a scanning speed of 4°/min. Transmission electron microscopy (TEM) data were collected with a JEOL JEM1230 transmission electron microscope. Ultraviolet-visible (UV-vis) spectroscopic analysis was implemented by a Perkin Elmer Lambda 35 UV-vis spectrophotometer. Photoluminescence (PL) spectra were recorded on a Hitachi F-7000 fluorescence spectrophotometer (Japan). Dynamic light scattering (DLS) measurements were performed using a Brookhaven Instruments 90PLUS instrument. The emission spectra of the upconversion nanoparticles were recorded upon irradiation with 980 nm NIR laser using a SDL980-LM-5000T (Shanghai Dream Lasers Technology Co., Ltd., Shanghai, China) laser diode at 3 W/cm2.

3. Results and Discussion

3.1. Synthesis of Mesoporous Silica-Coated UCNPs

A general strategy to achieve the NIR-triggered ROS generation in UCNPs is to modify the surface of UCNPs with the function of adsorption of photosensitizers. Since the as-synthesized UCNPs were oleate-capped and hydrophobic, the surface modification via the coating of mesoporous silica (mSiO2) were utilized to render them hydrophilic for the related biological applications. An overcoating of mesoporous silica on the UCNP surface has several advantages, including low cytotoxicity, high chemical stability, etc. Above all, the large surface area and pore volume of mesoporous silica secure easier adsorption as well as a higher loading capacity of various therapeutic materials. For the preparation of mesoporous silica-coated UCNPs, many studies employed the reverse micro-emulsion method to directly coat mesoporous SiO2 on the hydrophobic UCNPs. However, this method needs to add a surfactant in the reaction mixture. In addition, the existence of organic ligands (such as oleic acid) on the UCNP surface may hinder the process of FRET and the homogeneity of mesoporous silica layer. More importantly, the agglomeration phenomenon of UCNPs caused by the organic ligands during the mSiO2 coating process resulted in the lower loading capacity (%, w/w) for every gram of mSiO2-coated UCNP particles, as we found in some previous studies [43,48]. Consequently, the preparation of mesoporous silica-coated UCNPs in this study was carried out through a modified approach unlike most reported methods. The synthetic route of UCNP@mSiO2 was achieved by coating the silica layer on the surface of oleate-free UCNPs with the addition of CTAC template and NaOH etchant to form the mesoporous silica shell. In this work, two kinds of mesoporous silica-coated UCNPs with the emission in the green (doped with Yb3+/Ho3+) and red (doped with Yb3+/Er3+) regions were prepared. The green emission UCNP is adopted as the donor for the photosensitizers MC 540 and rose bengal because both PSs have strong absorption in the green region, while the red emission UCNP is selected for the photosensitizer Ce6 due to its strong absorption band in the red region. Firstly, both oleate-free type UCNPs, green emission of LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2 nanoparticles and red emission of LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2 nanoparticles were synthesized according to the published article [50]. The coating of silica on the surface of UCNPs was then carried out by adding TEOS and using CTAC as the template to form the silica-coated UCNPs (UCNP@SiO2). Afterwards, the mesoporous silica coating was achieved by the NaOH etching and ion exchange with NaCl to obtain the mesoporous silica-coated UCNPs (UCNP@mSiO2). After coating with the mesoporous silica, it is apparent that both mSiO2-coated UCNPs displayed excellent dispersibility in aqueous solutions.

3.2. Structural Characterization of the m-SiO2 Coated UCNPs

X-ray diffraction (XRD) analysis was used to clarify the phase structures of both Ho3+-doped and Er3+-doped oleate-capped core/shell UCNPs and mesoporous silica-coated UCNPs. As regards both mSiO2-coated UCNPs, the XRD patterns shown in Figure 1a,b are similar to those of oleate-capped UCNPs, which are indexed as the standard data (JCPDS No. 17-0874) of the tetragonal LiYF4 crystal. It is obvious that the coating of a thin layer mesoporous silica did not affect the crystal structure of the synthesized UCNPs. It should be noted that the WAXD diffractograms of both ligand-free UCNPs are omitted here due to the fact that they have the identical structures with those of oleate-capped UCNPs [51].
The surface morphology of both mSiO2-coated UCNPs is observed with TEM and shown in Figure 2 and Figure 3. As evident from Figure 2 and Figure 3, a thin layer of mesoporous silica is present on the surface of LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2 and LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2 UCNPs. Compared to the ligand-free UCNPs, a homogeneous coating of mSiO2 layer on the UCNP surface appears both in the long axis and short axis of the tetragonal crystals.
As seen in Figure 2 and Figure 3, with respect to both green emission and red emission UCNPs, it is evident that the particle sizes increased after the mesoporous silica layer was coated. For easier comparison, a representative particle of each oleate-free and mesoporous silica-coated UCNP sample was chosen and the size was labeled, as shown in Figure 2 and Figure 3. The long axis and short axis for the ligand-free Ho3+-doped UCNPs and the Er3+-doped UCNPs are 159.64 nm and 69.13 nm, and 152.19 nm and 71.62 nm, respectively. The calculated aspect ratios (major axis/minor axis) are approximately 2.31 and 2.12. After overcoating the mSiO2 layer, it is apparent that both ligand-free UCNPs display the increases in size in both long axis and short axis. The long axis and short axis for the Ho3+-doped and Er3+-doped nanoparticles are 185.24 nm and 108.40 nm and 177.11 nm and 102.78 nm, respectively.
Specific surface area measurements of both mSiO2-coated UCNPs were performed by the Brunauer-Emmett-Teller (BET) method through the adsorption of nitrogen gas. According to the IUPAC classification scheme for mesoporous materials, both N2 adsorption/desorption isotherms of Ho3+-doped and Er3+-doped UCNP@mSiO2 (Figure S1a,b) indicate that the isotherms of UCNP@mSiO2 nanomaterials exhibit a type-IV with a hysteresis loop. The surface areas for both UCNP@mSiO2 are determined to be 139.1 and 232.8 m2/g, respectively, using the BET method. The average pore radii (Figure S1c) are determined to be 13.7 and 13.0 nm, correspondingly, suggesting that both as-synthesized UCNP@mSiO2 nanocomposites are porous structures and can be used for loading photosensitizers or drugs.
Dynamic light-scattering (DLS) analyses were carried out to further confirm the change in the particle size after the surface of UCNPs was coated with mSiO2. As shown in Figure S2, for the green emission LiYF4:Yb3+0.25/Ho3+0.01@ LiYF4:Yb3+0.2 UCNPs, the average particle size changed from 145.6 nm (ligand-free) to 183.4 nm after covering with mSiO2. Similarly, as shown in Figure S3, the average particle size of the red emission LiYF4:Yb3+0.25/Er3+0.01@ LiYF4:Yb3+0.2 UCNPs increased from 141.2 nm (ligand-free) to 174.3 nm (mSiO2-coated). The results of DLS data testified the successful coating of mesoporous silica on the surface of oleate-free UCNPs. Figure S4 depicts that the zeta potentials for Ho3+-doped and Er3+-doped UCNP@mSiO2 are around −19.05 and −23.29 mV, respectively.
The successful coating of an mSiO2 layer on the ligand-free UCNPs was further verified by the XPS and EDS data, as shown in Figure 4. The XPS spectra in Figure 4a show the Ho 3d band for the Ho3+-doped ligand-free UCNPs and mSiO2-coated UCNPs. Similarly, the existence of Er 3d band in Figure 4b confirms the structure of both Er3+-doped ligand-free and mSiO2-coated UCNPs. As compared to both ligand-free UCNPs, the presence of intense peaks corresponding to Si 2p and O 1s indicates the successful coating of mSiO2 on the UCNP surface. Nevertheless, both Si 2p and O 1s peaks are also perceptible in both ligand-free type UCNPs because the XPS measurements were performed on the glass substrates. The EDS spectra exhibit the existence of the elements of Y, Yb, F, Si, and O in the Ho3+-doped UCNP@mSiO2 and the Er3+-doped UCNP@mSiO2 (Figure 4c,d). The appearance of Si and O elements confirms the overcoating of mSiO2 layer on both UCNPs. The invisibility of Ho and Er elements in each kind of UCNP is due to the tiny doping amount (1 mol% of Ho or Er) compared to the host material of LiYF4 (74 mol% of Y). In contrast, the presence of Cu element can be attributed to the analyzed materials that were deposited on the copper substrates for examination.

3.3. Optical Properties

Since the PL intensity is a crucial factor for the efficiency of the FRET process between the UCNP donor and the PS acceptor, the PL spectra of both UCNPs before and after coating of mesoporous silica layer are shown in Figure 5a,b. As reported in our previous article, the PL spectra are almost the same for both UCNPs before and after removing the oleate ligand. Therefore, only the PL spectra of oleate-capped and mSiO2-coated UCNPs are compared in this figure. For the oleate-capped Ho3+-doped UCNPs, the PL spectrum presents the characteristic emission peaks of 540 nm and 650 nm (Figure 5a). With regard to the oleate-capped Er3+-doped nanoparticles, the PL spectrum exhibits three emission peaks around 525, 550, and 650 nm (Figure 5b). It is evident that overcoating the mesoporous silica layer on the surface of UCNPs has no prominent influence on the PL intensity. Conversely, it is beneficial for UCNPs to generate more ROS because the mSiO2 layer can adsorb a greater amount of photosensitizer.

3.4. Loading Capacities of Photosensitizers and Fluorescence Resonance Energy Transfer

Taking advantage of the green emission and red emission function of UCNP@mSiO2 nanocomposites, a therapeutic nanoplatform for photodynamic therapy (PDT) is constructed by incorporating photosensitizers into the UCNP@mSiO2 nanohybrids. Specifically, we have chosen MC 540 and RB as the photosensitizers of Ho3+-doped UCNP@mSiO2 in PDT treatment, as they have strong UV/Vis absorptions around 400–600 nm that overlap well with the dominant green emission (540 nm) of Ho3+-doped UCNP materials (Figure S5a,b). However, Ce 6 has been adopted as the photosensitizer of Er3+-doped UCNP@mSiO2 because it has a strong absorption overlapping with the red emission (650 nm) of the Er3+-doped UCNPs, as shown in Figure S5c. Due to the efficient overlap between the emissions of UCNP@mSiO2 nanoparticles and the absorptions of photosensitizers, the effective FRET is expected to occur from UCNP@mSiO2 nanomaterials to photosensitizers.
For the NIR-triggered PDT, the therapeutic strategy is associated with the photochemical reactions of photosensitizers to generate cytotoxic ROS to destroy nearby cancer cells. Accordingly, in order to generate a large amount of ROS for PDT, it is crucial that the effective FRET takes place between UCNPs and photosensitizers upon NIR irradiation. Directly coating the mSiO2 layer for loading more PS molecules on the UCNP surface is considered to facilitate more ROS generation compared to other surface modification strategy.
To perform the effective FRET process between UCNP@mSiO2 nanocomposites and photosensitizers, the loading capacities (%, w/w) of both mesoporous silica-coated UCNPs with their corresponding photosensitizers were determined by the UV-vis absorption spectra of standard photosensitizer solutions at different concentrations. In order to confirm that coating mSiO2 layer on the surface of UCNPs is a more efficient modality to produce ROS for PDT compared to other synthesized UCNPs with similar compositions, the loading capacities of oleate-capped and ligand-free UCNPs with their correlative photosensitizers were also listed in Table 1. As seen in Table 1, for the three kinds of UCNP—PS pair, denoted as Ho3+-doped UCNPs/MC 540, Ho3+-doped UCNPs/RB, and Er3+-doped UCNPs/Ce6, respectively, the adsorbed amounts of photosensitizers for mSiO2-coated UCNPs are the largest among the three different types of UCNPs. It is noteworthy that the adsorbed amount of MC 540 reaches about 30.6% (w/w) for every gram of Ho3+-doped UCNP@mSiO2. This result demonstrates that increasing the surface area of UCNPs by means of coating the UCNP surface with the mesoporous silica is a promising strategy to raise the adsorbed amounts of photosensitizers on the surface of UCNPs.
The efficacy of FRET between UCNPs and photosensitizers could be evaluated from the quenching efficiency of PL spectra of UCNPs before and after adsorption of the photosensitizers. Based on the peak intensity of the green emission and red emission before (Fmax) and after (Ft) adsorption of the photosensitizer molecules on the surface of UCNPs, the quenching efficiency (%) can be obtained using the equation of (Fmax − Ft)/Fmax. The PL spectra due to FRET between both Ho3+-doped UCNP@mSiO2 and Er3+-doped UCNP@mSiO2 with their corresponding photosensitizers are shown in Figure 6. For comparison, the PL spectra of the ligand-free type UCNPs with their corresponding photosensitizers are also displayed in this figure. As noted above, the PL intensities of both ligand-free UCNPs are almost the same as those of oleate-capped UCNPs. It is evident that effective FRET processes between UCNP@mSiO2 nanocomposites and their corresponding PSs can be observed from Figure 6.
The calculated quenching efficiencies of the three different types of UCNP—PS pairs are illustrated in Figure 7. As seen in Figure 7, it is apparent that among the three different UCNP—PS pairs, the quenching efficiency resulting from the FRET of Ho3+-doped UCNPs to MC 540 is the greatest due to the larger spectral overlap region as shown in Figure S5. Compared to the oleate-capped UCNPs, both ligand-free and mSiO2-coated UCNPs have greater quenching efficiencies, which can be ascribed to the close adjacency between the FRET pairs of the UCNP donors and the PS acceptors and the larger adsorbed amount of PS molecules. The quenching efficiencies for Ho3+-doped UCNP@mSiO2 with MC 540 or rose bengal, and Er3+-doped UCNP@mSiO2 with Ce6 are ca. 76.2%, 64.6%, and 44.1%, respectively. The results shown in Figure 7 demonstrate that the three mSiO2-coated UCNP/PS pairs have comparable quenching efficiencies with their corresponding ligand-free UCNP/PS pairs. In spite of this, the results which will be shown below will confirm that mSiO2-coated UCNP/PS pairs have more pronounced effect for ROS production in PDT as compared to the ligand-free type UCNP/PS pairs. It can be attributed to that mSiO2-coated UCNPs have greater loading capacities of photosensitizers in comparison with ligand-free UCNPs.

3.5. Evaluation of ROS Generation of UCNP@mSiO2/PS Nanomaterials

The generation of reactive oxygen species (ROS) from the three UCNP@mSiO2/PS nanomaterials were evaluated using 9,10-anthracenediyl-bis(methylene) dimalonic acid (ABDA) as the chemical probe to detect the singlet oxygen 1O2. The quenching reaction between ABDA and singlet 1O2 resulted in a decrease in UV-vis absorption intensity of ABDA. Consequently, the ROS production can be well monitored by measuring the decrease in the absorption signals of the characteristic peaks at 359 nm, 378 nm and 399 nm of ABDA under NIR irradiation. The absorption changes of the three characteristic peaks for Ho3+-doped UCNP@mSiO2/MC 540, Ho3+-doped UCNP@mSiO2/RB, and Er3+-doped UCNP@mSiO2/Ce6 donor-acceptor nanoplaforms under different irradiation times of 980 nm NIR are shown in Figure 8 and Figure S6. As seen in this figure, the absorbances of the three characteristic peaks for the three donor-acceptor pairs display a significant downward trend, indicating the occurrence of effective ROS generation. It can be seen that the At/Amax (%) curves of three characteristic peaks are almost the same. The values of At/Amax (%) for the three donor-acceptor systems upon 980 nm laser irradiation for 60 min decrease by about 27%, 26%, and 22%, as is evident from Figure 8 and Figure S6 and Table 2. It is pronounced that the Ho3+-doped UCNP@mSiO2/MC 540 donor-acceptor pair exhibited the largest ROS generation among the three donor-acceptor pairs owing to the larger spectral overlap area and the greater loading capacity of MC 540 for efficient FRET. Most importantly, as illustrated in Figures S7 and S8 and Table 2, it is apparent that after overcoating the UCNP surface with the mesoporous silica layer on the UCNP surface, the generation amounts of ROS for the three donor-acceptor pairs are significantly increased compared to their corresponding oleate-capped and ligand-free type donor-acceptor pairs. Consequently, it is suggested that surface modification of UCNPs via coating an mSiO2 layer may be a facile and efficient method for ROS generation in PDT.

3.6. In Vitro Photodynamic Therapy of UCNP@mSiO2/PS Nanomaterials with MCF-7 Cancer Cells

The cytotoxicity test of the three UCNP@mSiO2/PS nanomaterials was performed on a human breast-cancer cell line MCF-7. Figure 9a,b illustrates the viability of MCF-7 cells incubated with UCNP@mSiO2/PS at different concentrations without and under laser light irradiation, respectively. As shown in Figure 9a, it is evident that the cell viability is still over 75% when up to 500 μg/mL of UCNP@mSiO2/PS nanocomposites were incubated with the cells without NIR exposure. Conversely, if the three UCNP@mSiO2/PS nanocomposites were incubated with the cells at different concentrations and irradiated with NIR laser light, the cell viabilities for the three nanocomposites significantly drop to ca. 39% at the concentration of 500 μg/mL (Figure 9b), suggesting the occurrence of effectual ROS generation for PDT. Figure 10 is images of MCF-7 breast cells incubated with various UCNPs and PSs without and with NIR irradiation, respectively. The number of cells under NIR irradiation is less than that of particle-treated controls (unirradiated samples), which agreed with the results presented in Figure 9.

4. Conclusions

In this work, two mSiO2-coated LiYF4 UCNPs with the corresponding emission in green and red regions were prepared and used as the donors for the study of ROS generation and in vitro PDT treatment. Three PSs were employed as the acceptors to fabricate the PDT theranostic nanoplatform via the donor/acceptor FRET process. In comparison with both oleate-capped and ligand-free UCNPs, the mSiO2-coated UCNPs achieved a high loading capacity of 30.6% (w/w) for MC 540. The results of quenching efficiencies between both mSiO2-coated UCNPs and their corresponding photosensitizers indicated that the FRET efficacies were close to those of ligand-free type UCNP/PS pairs and much greater than those of oleate-capped UCNP/PS pairs due to the close proximity of the donor-acceptor pair after ligand removal. Furthermore, the quenching efficiency was more prominent for Ho3+-doped UCNPs as compared to Er3+-doped UCNPs because of the larger spectral overlap for effective FRET process between the UCNPs and the PS molecules. With regard to the three UCNP@mSiO2/PS nanocarriers, due to the effect of close adjacency between donors and acceptors as well as the higher loading capacities of PS molecules, the amounts of ROS generation were more prominent as compared to the ligand-free and the oleate-capped UCNP/PS pairs. In vitro cytotoxicity tests indicated that the cell viabilities for the three UCNP@mSiO2/PS nanocomposites dropped to ca. 39% at the concentration of 500 μg/mL, suggesting that the LiYF4 UCNP@mSiO2/PS donor-acceptor nanoplatform could be utilized as an effectual approach in photodynamic therapy.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms23158757/s1.

Author Contributions

T.-H.H.: Conceptualization, Supervision, Writing—original draft. C.-H.Y.: Conceptualization, Investigation, Formal analysis. Z.-E.J.: Software, Investigation, Data curation, Methodology. H.-Y.L.: Resources, Data curation, Supervision. Y.-F.C.: Data curation, Investigation. T.-L.W.: Funding acquisition, Resources, Writing—original draft, Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

We gratefully acknowledge the support of the Ministry of Science and Technology, Taiwan, under Grant no. MOST 109-2221-E-390-024.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article and its Supplementary Materials.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Medina-Cruz, D.; Mostafavi, E.; Vernet-Crua, A. Green nanotechnology-based drug delivery systems for osteogenic disorders. Expert Opin. Drug Deliv. 2020, 17, 341–356. [Google Scholar] [CrossRef] [PubMed]
  2. Mostafavi, E.; Medina-Cruz, D.; Kalantari, K.; Taymoori, A.; Soltantabar, P.; Webster, T.J. Electroconductive nanobiomaterials for tissue engineering and regenerative medicine. Bioelectricity 2020, 2, 120–149. [Google Scholar] [CrossRef] [PubMed]
  3. Zare, H.; Ahmadi, S.; Ghasemi, A. Carbon nanotubes: Smart drug/gene delivery carriers. Int. J. Nanomed. 2021, 16, 1681–1706. [Google Scholar] [CrossRef] [PubMed]
  4. Mostafavi, E.; Zarepour, A.; Barabadi, H.; Zarrabi, A.; Truong, L.B.; Medina-Cruz, D. Antineoplastic activity of biogenic silver and gold nanoparticles to combat leukemia: Beginning a new era in cancer theragnostic. Biotechnol. Rep. 2022, 34, e00714. [Google Scholar] [CrossRef] [PubMed]
  5. Vallet-Regi, M.; Rámila, A.; del Real, R.P.; Pérez-Pariente, J. A new property of MCM-41: Drug delivery system. Chem. Mater. 2001, 13, 308–311. [Google Scholar] [CrossRef]
  6. Qiang, C.; Luo, Z.S.; Pang, W.Q.; Fan, Y.W.; Chen, X.H.; Cui, F.Z. Dilute solution routes to various controllable morphologies of MCM-41 Silica with a basic medium. Chem. Mater. 2001, 13, 258–263. [Google Scholar]
  7. Fowler, C.E.; Khushalani, D.; Lebeau, B.; Mann, S. Nanoscale materials with mesostructured interiors. Adv. Mater. 2001, 13, 649–652. [Google Scholar] [CrossRef]
  8. Liu, J.; Bu, W.; Pan, L.; Shi, J. NIR-triggered anticancer drug delivery by upconverting nanoparticles with integrated azobenzene-modified mesoporous silica. Angew Chem. Int. Ed. 2013, 52, 4375–4379. [Google Scholar] [CrossRef]
  9. Lee, J.E.; Lee, N.; Kim, T.; Kim, J.; Hyeon, T. Multifunctional mesoporous silica nanocomposite nanoparticles for theranostic applications. Acc. Chem. Res. 2011, 44, 893–902. [Google Scholar] [CrossRef]
  10. Zhao, W.R.; Gu, J.L.; Zhang, L.X.; Chen, H.R.; Shi, J.L. Fabrication of uniform magnetic nanocomposite spheres with a magnetic core/mesoporous silica shell structure. J. Am. Chem. Soc. 2005, 127, 8916–8917. [Google Scholar] [CrossRef]
  11. Lee, J.E.; Lee, N.; Kim, H.; Kim, J.; Choi, S.H.; Kim, J.H.; Kim, T.; Song, I.C.; Park, S.P.; Moon, W.K.; et al. Uniform mesoporous dye-doped silica nanoparticles decorated with multiple magnetite nanocrystals for simultaneous enhanced magnetic resonance imaging, fluorescence imaging, and drug delivery. J. Am. Chem. Soc. 2010, 132, 552–557. [Google Scholar] [CrossRef] [PubMed]
  12. He, Q.; Shi, J. Mesoporous silica nanoparticle based nano drug delivery systems: Synthesis, controlled drug release and delivery, pharmacokinetics and biocompatibility. J. Mater. Chem. 2011, 21, 5845–5855. [Google Scholar] [CrossRef]
  13. Tu, H.L.; Lin, Y.S.; Hung, Y.; Lo, L.W.; Chen, Y.F.; Mou, C.Y. In vitro studies of functionalized mesoporous silica nanoparticles for photodynamic therapy. Adv. Mater. 2009, 21, 172–174. [Google Scholar] [CrossRef]
  14. Cheng, S.H.; Lee, C.H.; Yang, C.S.; Tseng, F.G.; Mou, C.Y.; Lo, L.W. Mesoporous silica nanoparticles functionalized with an oxygen-sensing probe for cell photodynamic therapy: Potential cancer theranostics. J. Mater. Chem. 2009, 19, 1252–1257. [Google Scholar] [CrossRef]
  15. Hong, S.H.; Choi, Y. Mesoporous silica-based nanoplatforms for the delivery of photodynamic therapy agents. J. Pharm. Investig. 2018, 48, 3–17. [Google Scholar] [CrossRef] [Green Version]
  16. Porrang, S.; Davaran, S.; Rahemi, N.; Allahyari, S.; Mostafavi, E. How advancing are mesoporous silica nanoparticles? A comprehensive review of the literature. Int. J. Nanomed. 2022, 17, 1803–1827. [Google Scholar] [CrossRef]
  17. Narayan, R.; Nayak, U.Y.; Raichur, A.M.; Garg, S. Mesoporous silica nanoparticles: A comprehensive review on synthesis and recent advances. Pharmaceutics 2018, 10, 118. [Google Scholar] [CrossRef] [Green Version]
  18. Yang, P.; Gai, S.; Lin, J. Functionalized mesoporous silica materials for controlled drug delivery. Chem. Soc. Rev. 2012, 41, 3679–3698. [Google Scholar] [CrossRef]
  19. Kim, M.K.; Ki, D.H.; Na, Y.G.; Lee, H.S.; Baek, J.S.; Lee, J.Y.; Lee, H.K.; Cho, C.W. Optimization of Mesoporous Silica Nanoparticles through Statistical Design of Experiment and the Application for the Anticancer Drug. Pharmaceutics 2021, 13, 184. [Google Scholar] [CrossRef]
  20. Castano, A.P.; Mroz, P.; Hamblin, M.R. Photodynamic therapy and anti-tumour immunity. Nat. Rev. Cancer. 2006, 6, 535–545. [Google Scholar] [CrossRef] [Green Version]
  21. Svensson, J.; Johansson, A.; Grafe, S.; Gitter, B.; Trebst, T.; Bendsoe, N.; Andersson-Engels, S.; Svanberg, K. Tumor selectivity at short times following systemic administration of a liposomal temoporfin formulation in a murine tumor model. Photochem. Photobiol. 2007, 83, 1211–1219. [Google Scholar] [CrossRef] [PubMed]
  22. Sanabria, L.M.; Rodriguez, M.E.; Cogno, I.S.; Vittar, N.B.R.; Pansa, M.F.; Lamberti, M.J.; Rivarola, V.A. Direct and indirect photodynamic therapy effects on the cellular and molecular components of the tumor microenvironment. Biochim. Biophys. Acta 2013, 1835, 36–45. [Google Scholar] [CrossRef] [PubMed]
  23. Yang, Y.; Hu, Y.; Wang, H. Targeting antitumor immune response for enhancing the efficacy of photodynamic therapy of cancer: Recent advances and future perspectives. Oxid. Med. Cell Longev. 2016, 2016, 5274084. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Yang, S.; Li, N.; Liu, Z.; Sha, W.; Chen, D.; Xu, Q.; Lu, J. Amphiphilic copolymer coated upconversion nanoparticles for near-infrared light-triggered dual anticancer treatment. Nanoscale 2014, 6, 14903–14910. [Google Scholar] [CrossRef] [PubMed]
  25. Zhao, L.; Peng, J.; Huang, Q.; Li, C.; Chen, M.; Sun, Y.; Lin, Q.; Zhu, L.; Li, F. Near-infrared photoregulated drug release in living tumor tissue via yolk-shell upconversion nanocages. Adv. Funct. Mater. 2014, 24, 363–371. [Google Scholar] [CrossRef]
  26. Qian, H.S.; Guo, H.C.; Ho, P.C.L.; Mahendran, R.; Zhang, Y. Mesoporous-silica-coated up-conversion fluorescent nanoparticles for photodynamic therapy. Small 2009, 5, 2285–2290. [Google Scholar] [CrossRef] [PubMed]
  27. Wu, X.J.; Xu, D. Formation of yolk/SiO2 shell structures using surfactant mixtures as template. J. Am. Chem. Soc. 2009, 131, 2774–2775. [Google Scholar] [CrossRef]
  28. Budijono, S.J.; Shan, J.; Yao, N.; Miura, Y.; Hoye, T.; Austin, R.H.; Ju, Y.; Prud’homme, R.K. Synthesis of stable block-copolymer-protected NaYF4:Yb3+, Er3+ up-converting phosphor nanoparticles. Chem. Mater. 2010, 22, 311–318. [Google Scholar] [CrossRef]
  29. Chen, G.; Qiu, H.; Prasad, P.N.; Chen, X. Upconversion nanoparticles: Design, nanochemistry, and applications in theranostics. Chem. Rev. 2014, 114, 5161–5214. [Google Scholar] [CrossRef]
  30. DaCosta, M.; Doughan, S.; Han, Y.; Krull, U. Lathanide upconversion nanoparticles and applications in bioassays and bioimaging: A review. Anal. Chim. Acta 2014, 832, 1–33. [Google Scholar] [CrossRef]
  31. Wang, F.; Banerjee, D.; Liu, Y.; Chen, X.; Liu, X. Upconversion nanoparticles in biological labeling, imaging, and therapy. Analyst 2010, 135, 1839–1854. [Google Scholar] [CrossRef] [PubMed]
  32. Wu, X.; Chen, G.; Shen, J.; Li, Z.; Zhang, Y.; Han, G. Upconversion nanoparticles: A versatile solution to multiscale biological imaging. Bioconjugate Chem. 2015, 26, 166–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Chen, C.; Li, C.; Shi, Z. Current advances in lanthanide-doped upconversion nanostructures for detection and bioapplication. Adv. Sci. 2016, 3, 1600029. [Google Scholar] [CrossRef] [Green Version]
  34. Goldschmidt, J.C.; Fischer, S. Upconversion for photovoltaics-a review of materials, devices and concepts for performance enhancement. Adv. Opt. Mater. 2015, 3, 510–535. [Google Scholar] [CrossRef]
  35. Sun, L.; Wei, R.; Feng, J.; Zhang, H. Tailored lanthanide-doped upconversion nanoparticles and their promising bioapplication prospects. Coord. Chem. Rev. 2018, 364, 10–32. [Google Scholar] [CrossRef]
  36. Fan, Y.; Liu, L.; Zhang, F. Exploiting lanthanide-doped upconversion nanoparticles with core/shell structures. Nano Today 2019, 25, 68–84. [Google Scholar] [CrossRef]
  37. Bagheri, A.; Arandiyan, H.; Boyer, C.; Lim, M. Lanthanide-doped upconversion nanoparticles: Emerging intelligent light-activated drug delivery systems. Adv. Sci. 2016, 3, 1500437. [Google Scholar] [CrossRef] [Green Version]
  38. Zhang, Q.; Yang, F.; Xu, Z.; Chaker, M.; Ma, D. Are lanthanide-doped upconversion materials good candidates for photocatalysis? Nanoscale Horiz. 2019, 4, 579–591. [Google Scholar] [CrossRef]
  39. Kang, D.; Jeon, E.; Kim, S.; Lee, J. Lanthanide-doped upconversion nanomaterials: Recent advances and applications. BioChip J. 2020, 14, 124–135. [Google Scholar] [CrossRef] [Green Version]
  40. Lee, G.; Park, Y.I. Lanthanide-doped upconversion nanocarriers for drug and gene delivery. Nanomaterials 2018, 8, 511. [Google Scholar] [CrossRef] [Green Version]
  41. Feng, W.; Sun, L.D.; Zhang, Y.W.; Yan, C.H. Synthesis and assembly of rare earth nanostructures directed by the principle of coordination chemistry in solution-based process. Coord. Chem. Rev. 2010, 254, 1038–1053. [Google Scholar] [CrossRef]
  42. Chatterjee, D.K.; Gnanasammandhan, M.K.; Zhang, Y. Small upconverting fluorescent nanoparticles for biomedical applications. Small 2010, 6, 2781–2795. [Google Scholar] [CrossRef]
  43. Idris, N.M.; Gnanasammandhan, M.K.; Zhang, J.; Ho, P.C.; Mahendran, R.; Zhang, Y. In vivo photodynamic therapy using upconversion nanoparticles as remote-controlled nanotransducers. Nat. Med. 2012, 18, 1580–1585. [Google Scholar] [CrossRef]
  44. Idris, N.M.; Jayakumar, M.K.G.; Bansal, A.; Zhang, Y. Upconversion nanoparticles as versatile light nanotransducers for photoactivation applications. Chem. Soc. Rev. 2015, 44, 1449–1478. [Google Scholar] [CrossRef]
  45. Biju, V. Chemical modifications and bioconjugate reactions of nanomaterials for sensing, imaging, drug delivery and therapy. Chem. Soc. Rev. 2014, 43, 744–764. [Google Scholar] [CrossRef] [PubMed]
  46. Sedlmeier, A.; Gorris, H.H. Surface modification and characterization of photon-upconverting nanoparticles for bioanalytical applications. Chem. Soc. Rev. 2015, 44, 1526–1560. [Google Scholar] [CrossRef] [Green Version]
  47. Liu, J.; Yang, F.; Feng, M.; Wang, Y.; Peng, X.; Lv, R. Surface plasmonic enhanced imaging-guided photothermal/photodynamic therapy based on lanthanide-metal nanocomposites under single 808 nm laser. ACS Biomater. Sci. Eng. 2019, 5, 5051–5059. [Google Scholar] [CrossRef] [PubMed]
  48. Tang, M.; Zhu, X.; Zhang, Y.; Zhang, Z.; Zhang, Z.; Mei, Q.; Zhang, J.; Wu, M.; Liu, J.; Zhang, Y. Near-infrared excited orthogonal emissive upconversion nanoparticles for imaging-guided on-demand therapy. ACS Nano 2019, 13, 10405–10418. [Google Scholar] [CrossRef]
  49. Chen, Y.; Ren, J.; Tian, D.; Li, Y.; Jiang, H.; Zhu, J. Polymer-upconverting nanoparticle hybrid micelles for enhanced synergistic chemo-photodynamic therapy: Effects of emission-absorption spectral match. Biomacromolecules 2019, 20, 4044–4052. [Google Scholar] [CrossRef]
  50. Ho, T.H.; Hong, S.Y.; Yang, C.H.; Chen, Y.F.; Lin, H.Y.; Wang, T.L. Preparation of green emission and red emission ligand-free upconverting nanoparticles for investigation of the generation of reactive oxygen species applied to photodynamic therapy. J. Alloys Compd. 2021, 893, 162323. [Google Scholar] [CrossRef]
  51. Wang, F.; Han, Y.; Lim, C.S.; Lu, Y.H.; Wang, J.; Xu, J.; Chen, H.; Zhang, C.; Hong, M.; Liu, X. Simultaneous phase and size control of upconversion nanocrystals through lanthanide doping. Nature 2010, 463, 1061–1065. [Google Scholar] [CrossRef] [PubMed]
  52. Zhao, X.Y.; Tan, M.C. Ce dopant effects on NaYF4 particle morphology and optical properties. J. Mater. Chem. 2015, C3, 10207–10214. [Google Scholar] [CrossRef]
  53. Bogdan, N.; Vetrone, F.; Ozin, G.A.; Capobianco, J.A. Synthesis of ligand-free colloidally stable water dispersible brightly luminescent lanthanide–doped upconverting nanoparticles. Nano Lett. 2011, 11, 835–840. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthetic route for the preparation of mSiO2-coated UCNP.
Scheme 1. Synthetic route for the preparation of mSiO2-coated UCNP.
Ijms 23 08757 sch001
Figure 1. Wide-angle X-ray diffractograms of the oleate-capped and mSiO2-coated core/shell nanoparticles. (a) LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2, (b) LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2.
Figure 1. Wide-angle X-ray diffractograms of the oleate-capped and mSiO2-coated core/shell nanoparticles. (a) LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2, (b) LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2.
Ijms 23 08757 g001
Figure 2. TEM images of LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2 UCNPs. (a) Ligand-free nanoparticles; (b) mSiO2-coated nanoparticles.
Figure 2. TEM images of LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2 UCNPs. (a) Ligand-free nanoparticles; (b) mSiO2-coated nanoparticles.
Ijms 23 08757 g002
Figure 3. TEM images of LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2 UCNPs. (a) Ligand-free nanoparticles; (b) mSiO2-coated nanoparticles.
Figure 3. TEM images of LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2 UCNPs. (a) Ligand-free nanoparticles; (b) mSiO2-coated nanoparticles.
Ijms 23 08757 g003
Figure 4. XPS spectra of (a) ligand-free and mSiO2-coated LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2 nanoparticles and (b) ligand-free and mSiO2-coated LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2 nanoparticles. EDS spectra of (c) mSiO2-coated LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2 nanoparticles and (d) mSiO2-coated LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2 nanoparticles.
Figure 4. XPS spectra of (a) ligand-free and mSiO2-coated LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2 nanoparticles and (b) ligand-free and mSiO2-coated LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2 nanoparticles. EDS spectra of (c) mSiO2-coated LiYF4:Yb3+0.25,Ho3+0.01@LiYF4:Yb3+0.2 nanoparticles and (d) mSiO2-coated LiYF4:Yb3+0.25,Er3+0.01@LiYF4:Yb3+0.2 nanoparticles.
Ijms 23 08757 g004
Figure 5. PL spectra of (a) oleate-capped and mSiO2-coated LiYF4: Yb3+0.25/Ho3+0.01 @ LiYF4:Yb3+0.2 and (b) oleate-capped and mSiO2-coated LiYF4: Yb3+0.25/Er3+0.01 @ LiYF4:Yb3+0.2 UCNPs under 980 nm NIR irradiation.
Figure 5. PL spectra of (a) oleate-capped and mSiO2-coated LiYF4: Yb3+0.25/Ho3+0.01 @ LiYF4:Yb3+0.2 and (b) oleate-capped and mSiO2-coated LiYF4: Yb3+0.25/Er3+0.01 @ LiYF4:Yb3+0.2 UCNPs under 980 nm NIR irradiation.
Ijms 23 08757 g005
Figure 6. PL spectra of nanoparticles under 980 nm NIR excitation before and after adsorption of PS molecules. (a) Ligand-free, and; (b) mSiO2-coated LiYF4: Yb3+0.25/Ho3+0.01 @ LiYF4:Yb3+0.2 with MC 540; (c) ligand-free, and; (d) mSiO2-coated LiYF4: Yb3+0.25/Ho3+0.01 @ LiYF4:Yb3+0.2 with rose bengal; (e) ligand-free, and; (f) mSiO2-coated LiYF4: Yb3+0.25/Er3+0.01 @ LiYF4:Yb3+0.2 with Ce6.
Figure 6. PL spectra of nanoparticles under 980 nm NIR excitation before and after adsorption of PS molecules. (a) Ligand-free, and; (b) mSiO2-coated LiYF4: Yb3+0.25/Ho3+0.01 @ LiYF4:Yb3+0.2 with MC 540; (c) ligand-free, and; (d) mSiO2-coated LiYF4: Yb3+0.25/Ho3+0.01 @ LiYF4:Yb3+0.2 with rose bengal; (e) ligand-free, and; (f) mSiO2-coated LiYF4: Yb3+0.25/Er3+0.01 @ LiYF4:Yb3+0.2 with Ce6.
Ijms 23 08757 g006
Figure 7. Comparison of quenching efficiency of Ho3+-doped UCNP@mSiO2/MC 540 pair, Ho3+-doped UCNP@mSiO2/RB pair and Er3+-doped UCNP@mSiO2/Ce6 pair with their corresponding oleate-capped and ligand-free type adducts after the occurrence of FRET.
Figure 7. Comparison of quenching efficiency of Ho3+-doped UCNP@mSiO2/MC 540 pair, Ho3+-doped UCNP@mSiO2/RB pair and Er3+-doped UCNP@mSiO2/Ce6 pair with their corresponding oleate-capped and ligand-free type adducts after the occurrence of FRET.
Ijms 23 08757 g007
Figure 8. Under NIR exposure, the UV-vis absorbance change of Ho3+-doped UCNP@mSiO2/MC 540 donor-acceptor system (a) as a function of irradiation time and (b) as a function of irradiation time and absorption wavelength.
Figure 8. Under NIR exposure, the UV-vis absorbance change of Ho3+-doped UCNP@mSiO2/MC 540 donor-acceptor system (a) as a function of irradiation time and (b) as a function of irradiation time and absorption wavelength.
Ijms 23 08757 g008
Figure 9. (a) Viability of MCF-7 cells incubated with UCNP@mSiO2/PS at different concentrations for 24 h without NIR exposure. (b) Viability of MCF-7 cells incubated with UCNP@mSiO2/PS at different concentrations for 24 h and irradiation with 980 nm for 5 min.
Figure 9. (a) Viability of MCF-7 cells incubated with UCNP@mSiO2/PS at different concentrations for 24 h without NIR exposure. (b) Viability of MCF-7 cells incubated with UCNP@mSiO2/PS at different concentrations for 24 h and irradiation with 980 nm for 5 min.
Ijms 23 08757 g009
Figure 10. Optical and DAPI-stained images (from top to bottom) of Ho3+-UCNP@mSiO2/MC 540, Ho3+-UCNP@mSiO2/RB and Er3+-UCNP@mSiO2/Ce6 without (left) and with (right) NIR irradiation. All scale bars are in 100 μm.
Figure 10. Optical and DAPI-stained images (from top to bottom) of Ho3+-UCNP@mSiO2/MC 540, Ho3+-UCNP@mSiO2/RB and Er3+-UCNP@mSiO2/Ce6 without (left) and with (right) NIR irradiation. All scale bars are in 100 μm.
Ijms 23 08757 g010
Table 1. Loading capacities for the oleate-capped, ligand-free, and mSiO2-coated UCNPs with their corresponding photosensitizers.
Table 1. Loading capacities for the oleate-capped, ligand-free, and mSiO2-coated UCNPs with their corresponding photosensitizers.
UCNP Samples/PhotosensitizersLoading Capacity (%, w/w)
Oleate-Capped UCNPsLigand-Free UCNPsmSiO2-Coated UCNPs
Ho3+-doped UCNPs/MC 54014.624.830.6
Ho3+-doped UCNPs/RB13.722.628.1
Er3+-doped UCNPs/Ce611.918.825.6
Table 2. ROS generation for oleate-capped UCNPs, ligand-free UCNPs, and mSiO2-coated UCNPs with their corresponding photosensitizers.
Table 2. ROS generation for oleate-capped UCNPs, ligand-free UCNPs, and mSiO2-coated UCNPs with their corresponding photosensitizers.
UCNP Samples/PhotosensitizersROS Generation from the Absorbance Change (%)
Oleate-Capped
UCNPs
Ligand-Free
UCNPs
mSiO2-Coated
UCNPs
Ho3+-doped UCNPs/MC 54012.421.027.1
Ho3+-doped UCNPs/RB9.816.126.2
Er3+-doped UCNPs/Ce66.915.322.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ho, T.-H.; Yang, C.-H.; Jiang, Z.-E.; Lin, H.-Y.; Chen, Y.-F.; Wang, T.-L. NIR-Triggered Generation of Reactive Oxygen Species and Photodynamic Therapy Based on Mesoporous Silica-Coated LiYF4 Upconverting Nanoparticles. Int. J. Mol. Sci. 2022, 23, 8757. https://doi.org/10.3390/ijms23158757

AMA Style

Ho T-H, Yang C-H, Jiang Z-E, Lin H-Y, Chen Y-F, Wang T-L. NIR-Triggered Generation of Reactive Oxygen Species and Photodynamic Therapy Based on Mesoporous Silica-Coated LiYF4 Upconverting Nanoparticles. International Journal of Molecular Sciences. 2022; 23(15):8757. https://doi.org/10.3390/ijms23158757

Chicago/Turabian Style

Ho, Tsung-Han, Chien-Hsin Yang, Zheng-En Jiang, Hung-Yin Lin, Yih-Fung Chen, and Tzong-Liu Wang. 2022. "NIR-Triggered Generation of Reactive Oxygen Species and Photodynamic Therapy Based on Mesoporous Silica-Coated LiYF4 Upconverting Nanoparticles" International Journal of Molecular Sciences 23, no. 15: 8757. https://doi.org/10.3390/ijms23158757

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop