Nanomaterial synthesis
BTNPs were synthesized by hydrothermal synthesis. Briefly, titanium hydroxide precursors were washed with CO2-free, deionized water. Then, the gels were suspended together with Ba(OH)2· 8 H2O in a 1,4-butanediol/water mixture (1:2). The resulting suspension was placed in a 700 mL Teflon container within a stainless-steel pressure vessel. The reaction vessel was then heated at a rate of 5 °C/min to 220 °C and kept for 48 h. The resulting powders were washed with pH-adjusted (pH = 10) CO2-free deionized water to remove the unreacted barium present in the solution and to prevent the incongruent dissolution of barium ions from the BaTiO3 particle surface. GO nanoflakes were synthesized as previously reported1. BTNPs and GO were autoclaved through vapor steam (30 min at 121 °C) to ensure their sterilization, according to ISO standard 17665-1:2006.
Nanomaterial characterization
The size of BTNPs was estimated by the BET method (Brunauer–Emmett–Teller theory) using a Belsorp-mini II (Belsorp) based on the curves of adsorption/desorption of N2 by the nanoparticle powder. The specific surface area (SSA) was calculated, and the particle diameter was determined as follows:
where ρ= 6.02 g/cm3 is the density of BaTiO3.
The size and morphology of the BTNPs were also analyzed through Transmission Electron Microscopy (TEM). A drop of autoclaved BTNPs in water suspension (100 µg/mL) was deposited onto a 300-mesh carbon-coated copper grid (TedPella). TEM analysis was carried out using a Libra 120 Plus microscope (Carl Zeiss, Oberkochen, Germany) operating at an accelerating voltage of 120 keV, equipped with an in-column omega filter for energy-filtered imaging, and with a bottom-mounted 12 bit 2k × 2k CCD camera (TRS).
The piezoelectric properties of autoclaved BTNPs were investigated through piezoelectric force microscopy (PFM), performed using an Icon Bruker AFM system (Dimension Icon, Bruker Co., USA), in the Peak Force PFM modality. A silicon probe (DDESP-V2, Bruker, Billerica, MA, USA), with a measured spring constant of 132.5 N/m, a resonant frequency of 486 kHz, and a deflection sensitivity of 57.4 nm/V was used. The amplitude of the piezoelectric signal and the hysteresis (sample bias from -10 to 10 V), were acquired in the vertical direction via lock-in detection by applying to the tip an alternating current voltage (Vac) of 2 V at 300 kHz, outside the tip resonance frequency. Five independent samples were analyzed with a scan frequency of 0.25 Hz, and the average value of the d33 piezoelectric coefficient was calculated as follows:
where A is the amplitude signal (pm). A reference sample made of polyvinyl fluoride (PVDF) in the form of a thin film (Goodfellow, thickness: 28 µm, d33: ~ -20 pC/N) was also analyzed to properly calibrate the PFM amplitude signal.
Atomic force microscopy (AFM) measurements were carried out on GO nanoflakes, deposited on Si/SiO2 wafers, using a Bio FastScan scanning probe microscope (Bruker, Dimension Icon & FastScan Bio, Karlsruhe, Germany). All images were obtained using PeakForce Quantitative Nanomechanical Mapping mode with a Fast Scan C (Bruker) silicon probe (spring constant: 0.45 N/m). The images were captured in the retrace direction with a scan rate of 1.5 Hz. The resolution of the images was 512 samples/line.
Nanomaterial coating
A PGA (degree of esterification < 80%, Carbosynth, Staad, St. Gallen, Switzerland) solution was prepared at a concentration of 2.5 mg/mL in deionized water and then filtrated (filter size: 0.22 μm) at room temperature (RT). The autoclaved BTNPs were added in a ratio of 1:1 w/w to the polymeric solutions. Then, a sonication process with an ultrasound probe (power: 25 W, time: 30 min, frequency: 20 kHz, Bandelin SonoPuls HD4050, Berlin, Germany) allowed enhancing the interaction between the polymer and the BTNPs, favoring nanomaterial dispersion in aqueous media.
The coating of GO nanoflakes with PDA was performed as follows: autoclaved GO (5 mg/mL) was suspended in an aqueous solution made of dopamine hydrochloride (Sigma-Aldrich) at a concentration of 5 mg/mL in deionized water, previously filtered (filter size: 0.22 μm, material: PES) and adjusted in terms of pH by drop addition of 1 M NaOH solution (Sigma-Aldrich) to achieve a value of 8.5. The solution was sonicated with an ultrasound probe (power: 25 W, time: 300 s, frequency: 20 kHz). Finally, the mixture was stirred vigorously for 24 h at room temperature in the dark.
Dynamic light scattering (DLS) and zeta potential measurements were performed using a Zetasizer NanoZS90 (Malvern Instruments Ltd., Worcestershire, UK), analyzing the average size and polydispersity index (PDI) immediately after sonication and 3 and 7 days from the nanomaterial preparation. The samples were dispersed in deionized water and cell culture medium (Chondrocyte Growth Medium without phenol red, Cell Inc. Application), setting the concentration for all sample types to 100 μg/mL. Six independent samples were analyzed for each sample type.
X-ray photoelectron spectroscopy (XPS) analysis was carried out to verify the coating presence on the BTNPs and GO nanoflakes. XPS was performed using a Nexsa spectrometer (Thermo Scientific, Sunnyvale, USA) equipped with a monochromatic, micro-focused, low-power Al Ka X-ray source (photon energy: 1486.6 eV). High-resolution spectra were acquired at a pass energy of 50 eV. The source power was 72 W. The measurements were carried out under ultra-high-vacuum conditions, at a base pressure of 5 × 10−10 torr (not higher than 3 × 10−9 torr). The spectra obtained were analyzed and deconvoluted using the Vision software (Kratos). Overlapping signals were analyzed after deconvolution into Gaussian/Lorentzian-shaped components.
Assessment of nanomaterial cytotoxicity on human chondrocytes
The nanomaterial cytotoxicity was preliminarily evaluated on human articular chondrocytes (Cell Applications Inc., Boston, MA, USA), by carrying out Live/Dead assay, DNA quantification, metabolic activity analysis and LDH release quantification. The detailed protocols used for these tests are described in Supplementary information, section S8.1.
Nanocomposite hydrogel preparation
VitroGel RGD® was purchased from Well Bioscience (North Brunswick, NJ, USA) and prepared following the manufacturer’s protocol. Briefly, the VG-RGD solution was mixed at RT with the Dilution Solution Type 1® (The Well Bioscience, North Brunswick, NJ, USA) at the ratio 1:2 up to obtain a uniform mixture. Then, Dulbecco Modified Eagle Medium (DMEM) (Life Technologies, Bleiswijk, The Netherlands) with a suspension of ASCs to reach the final cell density into the hydrogel of 2 × 106 cells/mL was added at the ratio of 4:1 (pre-crosslinked solution: DMEM with cells) at RT and mixed. Hydrogels doped with GO nanoflakes and BTNPs (referred as nanocomposite) were obtained following the same procedure but adding the nanomaterials at concentrations of 25 µg/mL and 50 µg/mL, respectively, and mixing at RT until obtaining a uniform solution. Finally, 300 μL of both cell-laden non-doped or nanocomposite mixtures were gently transferred into a cell crown (Scaffdex, Finland), and inserted into a 24-wells plate. After 20 min of stabilization at RT, further 150 µL of DMEM were placed over the hydrogel for 1 h to allow saturation of the ionic crosslinking. Finally, 1.5 mL of DMEM were added to each well and the samples incubated at 37 °C and 5% CO2.
TEM imaging of the nanocomposite hydrogel
For ultrastructural evaluation, the hydrogels were fixed with 2.5% glutaraldehyde in 0.1 M cacodylate buffer (pH = 7.4) for 1 h at RT and for 3 h at 4°C. Afterward, samples were postfixed with 1% osmium tetroxide in 0.1 M cacodylate buffer for 2 h at 4°C, dehydrated in an ethanol series, infiltrated with propylene oxide and embedded in Epon resin. Cross-sections of each hydrogel were cut to allow internal analysis. Ultrathin sections (80 nm-thick) were stained with uranyl acetate and lead citrate (15 min each) and observed with a Jeol Jem 1011 transmission electron microscope (Jeol Jem, USA), operated at 100 kV. Images were captured using an Olympus digital camera and iTEM software. Unstained ultrathin sections were observed with a Zeiss Libra 120 plus TEM operating at 120 keV, and equipped with a Bruker XFlash 6T-60 SDD detector for Energy Dispersive X-ray spectroscopy (EDX).
Chemical-physical characterization of the nanocomposite hydrogel
The chemical features of the sample were characterized through Fourier Transform Infrared Spectroscopy (FT-IR) and by Nuclear Magnetic Resonance (NMR) spectra recording. The physical features were assessed by rheometry (which also allowed estimating the shear stresses acting on the cells during injection), uniaxial compression tests, tribological measurements, and degradation tests in different media. The detailed protocols are described in Supplementary information, section S8.2.
Assessment of material injectability
Injectability tests were performed by compressing a syringe piston loaded with the hydrogel solution using a traction/compression machine (model 2444, Instron, Norwood, MA, USA). The syringe (6 mL) was equipped with different needle sizes (18G, 20G, 22G and 24G, length: 3.8 cm), and pushed using a speed of 2.5 mm/s, compatible with EN ISO 7886‑1:2018. regulating the use of syringes. The force needed to allow material injection was recorded by the load cell of the instrument.
Assessment of nanocomposite hydrogel stability on the cartilage tissue
The stability of the hydrogels onto the cartilage tissue was assessed upon injecting the material solutions while varying the angle that the injector tip formed with respect to the cartilage tissue. The tissue was harvested from an adult bovine knee. A drop of ~20 µL was poured onto the cartilage, and a photo was taken after 2 s to assess its stability. Five trials were performed for each angle.
To evaluate the adhesion strength of the hydrogels to the cartilage tissue, a custom set-up was used, as reported in Trucco et al.2 Cartilage samples from the knee of an adult bovine were cut using a surgical instrument for bone/cartilage biopsies (Longueur) with an inner diameter of 6.4 mm, and fit to the setup. Then, 400 µL of hydrogel were delivered onto the cartilage and left crosslinking. After hydrogel crosslinking, the hydrogel-hosting part was hooked to the load cell of the traction test machine, and the test was performed in traction modality (speed: 1 mm/min) until reaching the mechanical failure of the interface. Force curves as a function of the displacement were obtained from each test, and the adhesion strength (in kPa) was determined by dividing the force by the contact area between the hydrogel and the cartilage tissue. From each adhesion strength curve, the maximum adhesion strength value (in kPa) was obtained.
Controlled ultrasound stimulation
Two US systems, one for 38 kHz low-frequency stimulation and the other one for high-frequencies (1 MHz and 5 MHz) stimulation were used in this work (Fig. 2a,b). The detailed protocols are described in Supplementary information, section S8.3.
FEM simulations of the BTNP – US wave interaction
FEM analyses were carried out using COMSOL Multiphysics (V6.0), run on a MacBook Pro M1 Max ARM64 processor, with 64 GB RAM. The COMSOL “MEMS” and “Acoustics” modules were chosen to include the relevant physics of the acoustic pressure wave and the piezoelectric and dielectric response of the BTNP. The detailed methods are described in Supplementary information, section S8.4.
In vitro culture of human ASCs
ASCs were purchased from Lonza (Pharma&Biotech, Switzerland) (N=6) and were expanded by seeding 7,500 cells/cm2 in T150 culture flasks and culturing them in α-MEM containing 5% isogrowth (IsoCellsGROWTH, Euroclone, Pero, IT) and 1% penicillin/streptomycin (Life Technologies) at 37°C in a 5% CO2 incubator. Before encapsulation in the hydrogel, ASCs were phenotypically characterized for the CD markers CD31, CD34, CD45, CD73, CD90, CD105, CD166 as previously reported3 and were analyzed for differentiation capability by using specific osteogenic and chondrogenic media as previously described4,5 to check that they satisfied the minimal criteria for defining multipotent mesenchymal stem cells6.
ASCs encapsulated in the bare or nanocomposite hydrogel were cultured with chondrogenic medium (high‐glucose DMEM supplemented with 50 mg/mL ITS + premix, 10−7 M dexamethasone, 50 μg/mL ascorbate–2phosphate, 1‐mM sodium pyruvate, and 100 U/mL – 100 μg/mL penicillin–streptomycin, Sigma Aldrich) containing chondrogenic factors TGF‐β3 (10 ng/mL) and BMP6 (10 ng/mL), both from Miltenyi Biotech, Auburn, CA, USA (Extended data Fig. 5a) or in inflammatory conditions (+ IL1β, 10 ng/mL) (R&D Systems, Inc., Minneapolis, MN, USA) (Extended data Fig. 8c). Cell culture medium was changed three times a week.
ASC-laden hydrogels treated with or without US (+ and -US) following specific experimental designs (Extended data Fig. 5a, 6c, 7a, 7e, 8c) were cultured for 2, 3, 10 and 28 days and evaluated for cell viability, cytotoxicity, and metabolic activity, gene expression, released factors, protein analysis and immunohistochemistry.
Viability of ASCs in the nanocomposite hydrogel
The viability of ASCs encapsulated in the nanocomposite hydrogel was evaluated by Live/Dead kit (Life Technologies). Samples were washed in D-PBS (Aurogene Srl, Rome, IT) and incubated with Live/Dead solution for 35 min at 37°C. Then, hydrogels were washed again with D-PBS and imaged, to discriminate live cells (in green) and the nuclei of dead cells (in red) with a fluorescence microscope (Nikon Instruments Europe BW). Quantitative analysis of stained slides was performed on five microscopic fields (×200 magnification) for each section. The analysis was performed using a Red/Green/Blue (RGB) tool within the Software NIS-Elements, at an Eclipse 90i microscope (Nikon Instruments Europe BV). The total number of green cells stained and the total number of positive-stained red cells were acquired. Data were expressed as a percentage of viable cells.
Cytotoxicity was assessed with a LDH assay kit (Roche, Mannheim, Germany). The supernatant was collected after 2 and 10 days and tested for the absorbance values at 490 nm by a microplate reader TECAN Infinite® 200 PRO (Tecan Italia S.r.l., Cernusco Sul Naviglio, Italy).
Cell metabolic activity was analyzed by Alamar Blue test. Briefly, the samples were incubated with 10% Alamar Blue (Life Technologies), and after 5 h, the absorbance was read at 570 and 600 nm using an automated spectrophotometric plate reader TECAN Infinite® 200 PRO (Tecan). The results were expressed as percentages of AlamarBlue reduction, as indicated by the manufacturer's data sheet (BioRad Laboratories).
For evaluating cell distribution within the hydrogels, the samples were fixed in 10% formaldehyde in D-PBS for 40 min, washed in D-PBS, dehydrated in ethanol, and embedded in paraffin. Thin sections (5 µm) were cut and stained for hematoxylin-eosin (Bioptica, Milan, Italy), then the slides were analyzed through a light microscope (Nikon Instruments Europe BW).
RNA isolation and quantitative PCR
Total RNA was extracted by treating all samples with 1 mL of Eurogold RnaPure (EuroClone S.p.a.). The samples were then immediately snap frozen in liquid nitrogen (-196 °C) and stored in a freezer at −80 °C. RNA extraction was performed by homogenizing samples and following the Eurogold manufacturer's instructions. The samples were then treated with DNase I (DNA-free Kit) and the RNA was quantified using a Nanodrop® spectrophotometer (EuroClone S.p.a). Reverse transcription was performed using Super Script® Vilo™ cDNA synthesis Kit (Life Technologies), according to the manufacturer's protocol.
qRT-PCR was performed by using TBGreen® Premix ExTaq™ (Takara Bio Inc. Shiga 525-0058, Japan) with LightCycler®2.0 (Roche Molecular Biochemicals). The gene markers quantified were: aggrecan (ACAN), collagen type 1 alpha 1 chain (COL1A1), collagen type 2 alpha 1 chain (COL2A1), collagen type 10 alpha 1 chain (COL10A1), the proliferation marker Ki-67 (MKI67), matrix metalloproteinase 13 (MMP13), SRY-Box Transcription Factor 9 (SOX9) and Tissue inhibitor of metalloproteinase 1 (TIMP1) (see Supplementary information, section S8.5). The efficiency of all primers was confirmed as high (>90%) and comparable. For each target gene, Crossing Point (CP) values were calculated and normalized to the CP of the housekeeping reference gene Glyceraldehyde-3-Phosphate Dehydrogenase (GAPDH) according to the formula 2-ΔCt and expressed as a percentage of the reference gene.
Immunohistochemistry and cytokine release measurements
On day 10 or 28, both hydrogel and nanocomposite treated with or without US were fixed in 10% formaldehyde in D-PBS for 40 min, washed in PBS, dehydrated in ethanol, and embedded in paraffin. Immunohistochemistry techniques were used to evaluate collagen type 2, proteoglycan and collagen type 1 expression. Serial sections of 5 µm were incubated for 60 min at RT with monoclonal mouse anti-human collagen type 2 (10 µg/mL), anti-human proteoglycan (5 µg/mL), anti-human collagen type 1 (5 µg/mL), all from Chemicon International, Temecula, CA, USA and polyclonal rabbit anti-human beta galactosidase (1 µg/mL) (from Proteintech Group, Rosemont, Illinois, USA), rinsed, and then sequentially incubated at RT for 20 min with multilinker biotinylated secondary antibody and alkaline phosphatase-conjugated streptavidin (Biocare Medical, Walnut CreeK, CA, USA). The colorimetric reactions were developed using fast red (Biocare Medical) counterstained with hematoxylin and mounted with glycerol jelly. The sections were evaluated with a bright field microscope (Nikon Instruments Europe BW). Negative and isotype-matched control sections were performed.
Apoptotic cells were detected using the In Situ Cell Death Detection Kit, AP (Merck, Darmstadt, Germany), following the manufacturer's instructions. Semiquantitative analyses on the stained slides were performed by acquiring 20 microscopic fields (×200 magnification) for each section. The analysis was performed using RGB with the software NIS-Elements and using an Eclipse 90i microscope (Nikon Instruments Europe BV). Briefly, we acquired the total number of blue-stained nuclei and the total number of positive-stained red cells and data were expressed as a percentage of positive cells.
The analysis of IL6 and IL8 release in the supernatant was purused using multiplex bead-based sandwich immunoassay kits (BioRad Laboratories Inc., Segrate, Italy) following the manufacturer's instructions.
Proteomic analysis, liquid chromatography-Tandem mass spectrometry (LC-MS/MS) and bioinformatic analysis
The total proteins were extracted and analyzed to assess the differential protein expression between the samples (nanocomposite with and without the application of US). The detailed protocols for sample treatment, data collection and analyses are reported in Supplementary information, section S8.6.
In vitro genotoxicity tests and in vivo biocompatibility tests
In vitro genotoxicity tests were performed by following ISO 10993-3:2015 (Biological evaluation of medical devices - Part 3: Tests for genotoxicity, carcinogenicity and reproductive toxicity) by applying the Ames and micronuclei tests. The Ames bacterial reverse mutation assay (Ames MPF™ Penta II kit, Xenometrix AG, Switzerland), was performed on four Salmonella typhimurium strains and one Escherichia coli strain, evaluating revertant colonies after a 90 min exposure to the nanocomposite hydrogel and a 48 h culture period. The cell micronuclei assay was performed on human lymphoblastoid TK6 cell line (ATCC, lot 59429029), for 3 and 24 h exposure periods, after which the relative population doubling (RPD) and the micronuclei frequencies were assessed.
All in vivo procedures were conducted strictly following the Italian Law on animals used for scientific purposes (Decree n. 26/2014): the project was authorized by the Italian Ministry of Health (n. 777/2021- PR) on the 3rd November 2021. Skin irritation tests were carried out following ISO 10993-23 (2021) on New Zealand SPF white male rabbits. Nanocomposite hydrogel, negative control and a positive known sensitizer were topically applied on the shaved dorsum region. After 4 h exposure, the treated sites were scored for erythema and oedema at 1, 24, 48 and 72 h. The Primary Irritation Index (PII) (minimum 0- maximum 8) was calculated according to the ISO 10993-23 standard. Acute systemic toxicity tests were carried out following ISO 10993-11 (2018) by single dose intramuscular nanocomposite injections on Sprague Dawley male rats. Clinical observations, signs of illness, pain, injury at the main apparatuses and systems, any behavioral alteration, and weight, water and food intake measurements were registered at baseline and at 24, 48, 72 h after treatment. Delayed type hypersensitivity tests were carried out following ISO 10993-10 (2010) on Dunkin Hartley guinea pigs, scoring erythema and oedema by Magnusson and Kligman grading scale after 24 h and 48 h7. The detailed protocols are reported in Supplementary information, section S8.7.
Statistical analyses
All data were analyzed using GraphPad Prism version 9.0.0 for Windows (GraphPad Software, San Diego, California USA, www.graphpad.com). D’Agostino-Pearson Normality test was performed on all data; data showing a normal distribution were analzyed using parametric tests, while data . data showing a non-normal distribution were analzyed using non-parametric tests.
Experimental data concerning DNA, LDH, metabolic analyses, DLS measurements, rheological indexes (K and n), estimated shear stress to the cells (using different needles), degradation rate, injection force, adhesion strength and COF were analyzed by applying a non-parametric Kruskal–Wallis test and Dunn’s multiple comparison test to analyze significant differences between groups.
Data concerning compressive modulus, swelling ratio and sol-gel fraction were analyzed by applying a non-parametric Mann-Whitney U-test to compare non-doped and doped hydrogels.
Experimental data derived from in vitro tests on ASCs were analyzed by applying a Mann-Whitney test or Wilcoxon test or Kruskal–Wallis one-way ANOVA and Dunn’s multiple comparison tests or Friedman and Dunn’s multiple comparison tests to analyze significant differences between groups.
Data from in vitro genotoxicity and in vivo biocompatibility tests were analyzed by applying a Shapiro-Wilk test and a Student’s t test when comparison versus CTR- was needed; otherwise, a two-way ANOVA followed by Sidak’s multiple comparison test was conducted.
For all tests, the significance threshold was set at p<0.05.
Sample size, randomization and blinding
For in vitro tests, the sample size was chosen based on previous laboratory experience considering a minimum of at least two independent experiments and a triplicate of independent samples. For genotoxicity and in vivo tests, the sample size was established based on the OECD guidelines and UNI EN ISO 10993 standard, which define the minimum number of samples/animals per group and test and guarantee the statistical validity of the results. No method of randomization was followed and no animals were excluded from this study. For in vitro tests, investigators were not blinded to sample allocation during the experiments and assessment of results. For in vivo tests, caregivers and the veterinary doctor were not blinded, whereas outcome assessors were blinded to the subject’s allocation.
Fig. 1. In c, the image is representative of five independent experiments; in f, g, h and i, n = 5 per group.
Fig. 3. In b, the images are representative of ten independent experiments; in c, the images are representative of six independent experiments, n=6 per group; in d, n=4 per group; in e, n=16 per group; in f, the images are representative of three independent experiments, n=12 per group; in g, the images are representative of five independent experiments, n=20 per group; in h, six independent experiments, n=18 per group; in i, six independent experiments, n=12 per group.
Fig. 4. In a, the image is representative of two independent experiments; in b, the image is representative of two independent experiments; in c, two independent experiments, n=10 per group; in d, two independent experiments, n=5 per group; in e, two independent experiments, n=5 per group; in f, two independent experiments, n=7 per group, the images are representative of two independent experiments; in g, duplicate cultures were performed for testing negative and positive controls and three different concentrations of the test substance. Micronuclei were scored on 2,000 cells, equally divided among the two replicates, for each tested condition; in h, healthy young adult New Zealand SPF white male rabbits were used for testing the nanocomposite hydrogel and negative control (n=3) and for positive control (n=1); in i, healthy Sprague Dawley male rats were used for testing the nanocomposite hydrogel and negative control (n=5 per group); in j, healthy Dunkin Hartley male and female guinea pigs were used for testing negative control (n=5), positive controls (n=10) and nanocomposite hydrogel (n=10).
Extended data Fig. 1. In a, the image is representative of five independent experiments; in b, three independent experiments; in c, five independent experiments per group (particle and background); in e, images are representative of two independent experiments; in f, n=6 per group; in g, n=2, images are representative of five images per sample analyzed; in h, n=3 per group; in i and j, n=4 per group.
Extended data Fig. 2. In b, images are representative of two independent experiments; in c, n=6 per group; in e, n=2, images are representative of five images per sample analyzed; in f, n=3 per group; in g and h, n=4 per group.
Extended data Fig. 3. In b, FT-IR graphs are representative of two independent experiments; in c, d, e, f, g, and h, n = 5 per group; in j, n = 5 per each angle tested; in k and j, n = 5 per group.
Extended data Fig. 5. In b, the images are representative of ten independent experiments; in c, the images are representative of six independent experiments, n=6 per group; in d, n=4 per group; in e, n=16 per group; in f, the images are representative of three independent experiments, n=12 per group; in g, the images are representative of five independent experiments, n=20 per group; in h, 6 independent experiments, n=18 per group; in i, six independent experiments, n=12 per group.
Extended data Fig. 6. In a, the image is representative of three independent experiments; in b, n=3 per group; in d, images are representative of two independent experiments; in e, the image is representative of two independent experiments.
Extended data Fig. 7. In b, the image is representative of three independent experiments; in c, d, e, f and g, two independent experiments, n=4 per group.
Extended data Fig. 8. In a and b, data refer to two independent experiments.
Extended data Fig. 9. In a, the Ames experiment was performed on five different bacterial strains, n=3 for each material and each concentration tested; in c, Micronuclei experiments were run in duplicates.
Data availability
The datasets generated and analyzed during the current study are available from the corresponding author on reasonable request. Mass spectrometry and proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE8 partner repository with the dataset identifier PXD038147 and 10.6019/PXD038147. Username: [email protected]. Password: PipAjUfZ.
Code availability
The code of the COMSOL routine used for FEM simulations of nanoparticles-US waves interactions is available in Supplementary Data File 1.