Open Access Article
Edgar Mendes
ab,
Pietro Apràc,
Ana Belchiorad,
Federico Picollo
ce,
Marta M. Alves
fg,
Rodica Mihaela Dinica
h,
Maria João Mourai,
Sofia Sturarice,
Teresa Pinheiro
*bd and
Maria Paula Cabral Campello
*ad
aCentro de Ciências e Tecnologias Nucleares, Instituto Superior Técnico, Universidade de Lisboa, Campus Tecnológico e Nuclear, Estrada Nacional 10, Km 139.7, 2695-066 Bobadela, Portugal. E-mail: edgar.mendes@tecnico.ulisboa.pt; anabelchior@ctn.tecnico.ulisboa.pt; pcampelo@ctn.tecnico.ulisboa.pt
biBB-Institute of Bioengineering and Biosciences, Instituto Superior Técnico, University of Lisbon, 1049-001 Lisbon, Portugal. E-mail: teresa.pinheiro@tecnico.ulisboa.pt
cNational Institute of Nuclear Physics, Section of Turin, Via P. Giuria 1, 10125 Turin, Italy. E-mail: pietro.apra@to.infn.it; federico.picollo@unito.it; sofia.sturari@unito.it
dDepartamento de Engenharia e Ciências Nucleares (DECN), Instituto Superior Técnico, Universidade de Lisboa, Estrada Nacional 10, 2695-066 Bobadela, Portugal. E-mail: anabelchior@ctn.tecnico.ulisboa.pt; pcampelo@ctn.tecnico.ulisboa.pt
ePhysics Department, and Nanomaterials for Industry and Sustainability (NIS) Inter-Departmental Centre University of Turin, Via P. Giuria 1, 10125 Turin, Italy. E-mail: federico.picollo@unito.it; sofia.sturari@unito.it
fCentro de Química Estrutural (CQE), Institute of Molecular Sciences (IMS), Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001, Lisboa, Portugal. E-mail: martamalves@tecnico.ulisboa.pt
gDepartamento de Engenharia Química (DEQ), Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisboa, Portugal
hFaculty of Sciences and Environment, Department of Chemistry Physical and Environment, “Dunărea de Jos” University of Galati, 111 Domnească Street, 800201 Galati, Romania. E-mail: rodica.dinicas@ugal.ro
iInstituto Nacional de Investigação Agrária e Veterinária, I.P. - Laboratório Químico Agrícola Rebelo da Silva Tapada da Ajuda, P-1301-596 Lisboa, Portugal. E-mail: mjoao.mouras@iniav.pt
First published on 7th January 2026
Nanodiamonds (ND) possess unique properties, including high biocompatibility, tunable surface chemistry, and stable photoluminescence, that make them highly attractive for biomedical applications. In this study, we synthesized gold-coated nanodiamonds (NDAu) using a green chemistry route based on Nymphaea alba root extract as a natural reducing agent. The hybrids were produced from two types of ND with median diameters of 50 nm and 230 nm, which were subjected to different thermal treatments prior to the gold coating to modulate their surface properties. The functionalized particles were comprehensively characterized using a combination of spectroscopic techniques (UV-Vis spectroscopy, ATR-FTIR spectroscopy, Raman spectroscopy, PIXE), Powder X-ray Diffraction (PXRD), electron microscopy (SEM and TEM), and zeta potential. These techniques evidenced the impact of the thermal treatments on the NDs, reported the influence of the plant extracts on the final nanoparticles, as well as confirmed and quantified the presence of metallic gold in this material. Moreover, we carried out biological evaluation on A549 lung cell line to assess their cytotoxicity, cellular uptake, and impact on cell survival. Our results confirmed the efficacy of the gold-coating method, elucidating the modifications in particles structural, physical and chemical properties due to functionalization, and the interaction with cells. These nanoparticles could then be used for various biomedical applications, such as drug delivery or as potential radiosensitizers.
New conceptsIn this work, we present a novel and sustainable approach to synthesize hybrid gold-nanodiamond (NDAu) nanostructures using green chemistry principles. The final nanoparticles apart from retaining properties of nanodiamonds (such as photoluminescence), gained plasmonic and catalytic properties that enhance imaging, sensing, and therapeutic applications from the gold-coating. By employing Nymphaea alba root extracts as eco-friendly reducing and stabilizing agents, we successfully functionalized thermally treated nanodiamonds of two distinct sizes (50 nm and 230 nm) with gold, circumventing the use of toxic reagents or harsh conditions. To the best of our knowledge, this is the first report on the green synthesis of hybrid gold-coated nanodiamonds, which results in biocompatible, hybrid nanostructures with potential biomedical applications. Moreover, through a radiosensitization process, these gold-conjugated nanodiamonds could provide notable benefits in augmenting the effectiveness of radiation therapy, as well as having great contributions in imaging and photothermal therapy. All these finding led us to conclude that our work represents a significant advancement beyond previous studies that focused on nanodiamonds modified via conventional chemical routes. Our method not only introduces a sustainable alternative for NDAu synthesis but also opens new possibilities for tuning surface properties through natural phytochemicals. |
Proton therapy utilizes proton beams with a distinctive depth-dose profile that concentrates the radiation dose at the final microns of their path, forming a Bragg peak. This enables highly precise targeting of the radiation dose, minimizing harm to healthy tissues, unlike X-rays, which deliver their maximum dose to superficial tissues. However, proton therapy has a Relative Biologic Effectiveness (RBE) of 1.1, making it only 10% more effective than conventional radiation therapy.4 Additionally, it requires costly equipment and specialized expertise. Despite these limitations, proton therapy is experiencing global growth, with over 100 operational facilities and over 50 under development.5 Consequently, investigating new radiosensitizers—agents that enhance the biological impact of radiation—is vital to improving the efficacy of cancer therapy in general and proton therapy in particular.
Nanodiamonds (ND) are particularly attractive for biomedical applications due to their biocompatibility, ease of functionalization, and ability to penetrate biological barriers, offering potential solutions to the limitations of conventional approaches in cancer theranostics.6–8 Their unique properties, such as the presence of Nitrogen Vacancy (NV) centers – characterized by a nitrogen atom substituting a carbon atom and a nearby lattice vacancy – further enhance their applicability in biomedical sciences. When excited with radiation in the range of 465 to 565 nm, the NV center emits light in the red spectrum, making NDs highly photostable with low photobleaching compared to common fluorescent dyes. This photoluminescent property, combined with the ability to induce additional NV centers, through high-energy particle irradiation followed by annealing at 600–800 °C, positions NDs as a valuable tool for bioimaging.7–18
NDs have garnered significant attention in research not only for their biomedical applications but also for their potential as radiosensitizers. Furthermore, nanodiamonds have been investigated for their ability to enhance radiosensitivity during proton and photon irradiation.19–21 Additionally, surface modifications enable the conjugation of therapeutic agents, such as drugs or nucleic acids, and imaging probes, such as fluorophores or radiolabels, turning nanodiamonds into multifunctional platforms for targeted drug delivery and real-time disease monitoring.6–8,22–27 A particularly promising modification involves the conjugation of gold onto the surface of nanodiamonds. Gold nanoparticles are well known for having exceptional surface and electronic properties, which make them perfect for a variety of biomedical applications.
The resulting composite—nanodiamond-gold hybrids—represents a cutting-edge development in nanotechnology, offering a synergistic combination of the unique properties of nanodiamonds and gold nanoparticles. NDs provide exceptional photostability, biocompatibility, and versatile surface chemistry, while gold nanoparticles contribute plasmonic and catalytic properties that enhance imaging, sensing, and therapeutic applications. This integration creates multifunctional systems with enhanced performance in bioimaging techniques, such as surface-enhanced Raman spectroscopy, photoacoustic imaging and multimodal imaging, leveraging the plasmonic properties of gold.28–31 These hybrid nanoplatforms also demonstrate significant potential in biosensing and diagnostics, including immunosensors for the detection of cancer biomarkers like HER2, which exhibit high sensitivity and specificity in human serum samples.32 Hybrid architectures, such as silica-encapsulated nanodiamonds coated with a gold shell, are effective in selective thermal ablation of cancer cells using near-infrared laser irradiation.33 Additionally, their catalytic properties and DNA-based coupling mechanisms enable intracellular biosensing and therapeutic delivery, establishing nanodiamond-gold hybrids as versatile tools for advancing precision medicine.34
Through a radiosensitization process, gold-conjugated nanodiamonds provide notable benefits in augmenting the effectiveness of radiation therapy, in addition to their contributions to imaging and photothermal therapy. Because of their high atomic number (Z), which increases the local deposition of radiation energy, gold nanoparticles are well-established radiosensitizers. Gold can promote the generation of secondary electrons when applied to the surface of nanodiamonds, which increases DNA damage in cancer cells when exposed to ionizing radiation.20,35–39 This effect is particularly advantageous for treating resistant tumors, where conventional radiation doses may be insufficient to achieve therapeutic outcomes. By leveraging gold's radiosensitizing properties, gold-functionalized nanodiamonds can potentially lower the required radiation dose, minimizing damage to surrounding healthy tissues while maintaining therapeutic efficacy. In proton therapy, AuNPs trigger mechanisms like the Coulomb nanoradiator effect and Auger cascades. Proton interactions with gold atoms strip electrons, generating reactive oxygen species (ROS) and X-rays, while secondary electron ejections via the Auger effect further increase tumor damage.40–43 The physical, optical, and biological properties of nanoparticles strongly depend on their size, shape, surface functionalization, and the dielectric properties of their environment. These characteristics are intrinsically linked to the experimental conditions employed during their synthesis.44
Herein, we developed hybrid gold-nanodiamond (NDAu) nanostructures using two types of commercial NDs with median sizes of 50 nm and 230 nm. The NDs were annealed at 800 °C for 2 hours in a nitrogen atmosphere, yielding annealed nanodiamonds (Ann NDs 50 and Ann NDs 230). These were subsequently oxidized in air at 500 °C for 12 hours, producing oxidized nanodiamonds (Ox NDs 50 and Ox NDs 230).20 The gold coating on the NDs was synthesized using green chemistry methods with Nymphaea alba root extracts as reducing agents, in line with sustainable practices. Nymphaea alba, also known as the European white-water lily, is rich in phytochemicals whose composition varies by geographical location and plant part. These extracts are known for their antioxidant, anti-inflammatory, and hepatoprotective properties, as well as their traditional use in treating diabetes, urinary tract infections, and other ailments. The key bioactive effects are largely attributed to polyphenols such as ellagic acid and flavonoids like quercetin and kaempferol.45,46
Green chemistry emphasizes reducing the environmental and health impacts of nanoparticle production by avoiding toxic solvents and pollutants. It leverages biological molecules such as polyphenols, citric acid, flavonoids, and reductases found in plants, algae, or microorganisms.45–49 These methods are versatile, cost-effective, and can incorporate biological waste from agricultural or food industries, lowering the carbon footprint of nanoparticle synthesis. Furthermore, green nanoparticles exhibit improved biocompatibility, as harmful chemical contaminants are avoided. Organic compounds from plant extracts integrated into the nanoparticle surfaces can enhance cellular uptake and confer additional beneficial properties, such as antimicrobial or anticancer activity.50
While green chemistry has been widely used to produce metal or metal oxide nanoparticles, to our knowledge, this is the first report of hybrid gold-nanodiamonds synthesized using this approach. Comparable methods, such as the environmentally friendly synthesis of silver nanoparticles on nanodiamonds using tea polyphenols, have demonstrated the potential of such strategies.50 Hybrid metal-ND for multipurpose biological applications, such as imaging or for hyperthermia, have also been synthesized using conventional chemical methods.39
In this study, the synthesized hybrid gold-nanodiamond underwent comprehensive physicochemical characterization. Their biological activity was preliminarily evaluated through cytotoxicity assays, cell survivability and colony formation assessments, and cellular uptake studies. These evaluations were conducted using the A549 lung cancer cell line, since it is a critical cancer located near vital structures, complicating treatment due to the risk of damaging healthy tissues. Moreover, these cancers often exhibit resistance to conventional therapies, underscoring the need for innovative and targeted treatment approaches.51,52 These biological assays were intended as a first assessment of cytocompatibility, supporting future in-depth biomedical investigations.
The A549 lung carcinoma epithelial cell was purchased from the American Type Culture Collection. The cell lines were cultured in Dulbecco's Modified Eagle Medium supplemented with 10% fetal bovine serum (FBS), supplemented with 1% of antibiotic–antimycotic solution. The cells were kept at 37 °C and 5% CO2 in a humidified atmosphere (Heraeus, Germany). All products for cellular studies, including the cellular media and supplements, were acquired from Gibco (Thermo Fisher Scientific), unless otherwise stated.
Table 1 lists the acronyms for each gold-coated ND and the specific amounts of reagents and precursors used at each synthesis step, and an outline of the different synthesis stages is also provided in Scheme 1, which includes the starting nanodiamonds (A), thermal pretreatment (B), surface oxidation (C), and gold reduction with Nymphaea alba root extract (D).
| Sample | Precursor | Volume of HAuCl4 (mL) |
|---|---|---|
| Ann NDAu 50 | 28 mg Ann ND 50 | 2 + 8 |
| Ox NDAu 50 | 27 mg Ox ND 50 | 2 + 8 |
| Ann NDAu 230 | 18 mg Ann ND 230 | 1.5 + 4.5 |
| Ox NDAu 230 | 17 mg Ox ND 230 | 1.5 + 4.5 |
:
3 nitric to hydrochloric acids molar ratios) of NDAu samples. Raman spectra were acquired with a Horiba LabRAM HR800 Evolution using a 532 nm solid-state laser at 20 mW output power. SEM analysis was conducted using a Phenom ProX G6 or Hitachi S-2400, operating with an electron beam at 15 or 20 kV, respectively. TEM imaging was performed on a JEOL 1200EX and the size distribution and mean diameter of the annealed and oxidized NDAu samples were assessed.
The effectiveness of Nymphaea alba root extracts in promoting the green-synthesis of gold coated nanodiamonds was demonstrated through a comprehensive set of complementary techniques. UV-vis, infrared, and Raman spectroscopies together with powder X-ray diffraction (PXRD), proton induced X-ray emission (PIXE), scanning electron microscopy (SEM), transmission electron microscopy (TEM), and measurements of DLS and zeta potential, consistently revealed changes in surface chemistry and morphology. Altogether, these results confirmed the efficiency in situ formation of the gold coating process on various nanodiamonds types, validating the extracts’ dual role as both reducing and stabilizing agents in the synthesis process.
It is anticipated that NDs' poorly absorbing particles will have UV-vis spectra that are strongly influenced by λ4, resulting in a curve that is more akin to exponential decay, with absorbance rising quickly as wavelength drops, especially in the UV region.58 The UV-Vis spectra of Ann NDAu 50 and Ox NDAu 50 exhibit a distinct absorption band at 545 nm, indicative of gold nanoparticle formation via localized surface plasmon resonance (LSPR).51 In the spectra of Ann NDAu 230 and Ox NDAu 230, this band is red-shifted to approximately 550 nm, consistent with the formation of larger nanoparticles and/or the presence of aggregates. Additionally, all NDAu samples exhibit absorption bands below 300 nm, which are attributed to polyphenolic and flavonoid compounds originating from the Nymphaea alba root extracts. This spectral feature supports the involvement of these phytochemicals in both the reduction of gold ions and the stabilization of the resulting nanoparticles.46 These findings are corroborated by the FTIR and UV-Vis spectra shown in Fig. 1 for the pristine and gold-coated Ann ND 50 samples. Corresponding spectra for the remaining samples are provided in Fig. S1 and S2.
![]() | ||
| Fig. 1 (A) UV-Vis and (B) FTIR spectra of 50 nm Ann ND, Ann NDAu, plant extracts and commercial AuNPs. | ||
The FTIR spectra of the pristine nanodiamond samples exhibit characteristic stretching and bending vibrations associated with annealed or oxidized NDs. Key absorption bands include a broad peak at 3420 cm−1 (O–H stretch of water), 1618 cm−1 and 1630 cm−1 (O–H bending), 1384 cm−1 (C–N stretching), and 1090 cm−1 and 1110 cm−1 (C–O stretching).59,60 The C–N band at 1384 cm−1 is less prominent in the Ox NDs, compared to the annealed counterpart, indicating effective nitrogen removal. The Ann ND shows smaller bands at 2920 cm−1 and 2849 cm−1 (CH2 stretching), while the Ox ND has unique bands at 1774 cm−1 (C
O stretching from ketones) and 1263 cm−1 (C–O bending from esters). In the NDAu spectra new peaks emerge at 1700 cm−1 (C
C stretching of carboxylic acids) and 1445 cm−1 (C–O–H bending flavonoids), which align with the presence of polyphenols and flavonoids, such as ellagic acid and quercetin, reported as key (C–O of flavonoids), consistent with the presence of polyphenols such as ellagic acid and quercetin, previously identified in N. alba root extracts.45 Additional peaks at 1582 cm−1 and 1508 cm−1 (aromatic C
C stretching) and 1366 cm−1 (O–H bending of phenolic groups) further support the successful contribution of these phytochemical components present in the extracts not only as a reducing, but also as a source of functional groups that enhance the stability and bioactivity of the hybrid nanostructures. The presence of esters and polyols, evidenced by the bands at 1111 cm−1 and 1200 cm−1, respectively, reinforce the involvement of plant-derived molecules in surface functionalization, potentially improving the biomedical applicability of the resulting nanodiamonds.45,46,59,60 Although metal NPs should not have any characteristic FTIR bands, an ATR-FTIR spectrum was recorded for commercial AuNPs (Aldrich Chemistry, 5 nm diameter) stabilized in citrate buffer in order to verify that neither of the bands in the NDAu samples arise from the gold in the surface. As expected, the only bands present in the AuNPs spectrum correspond to the O–H stretching of water and a C
O symmetrical stretching from the citrate molecule.61
Fig. 2 shows the Raman spectra of the 50 nm pristine nanodiamonds and their gold-coated counterparts. As shown in Fig. 2A and B, the spectra of the pristine NDs display features characteristic of this type of material: the diamond peak (red) at around 1320 cm−1, which corresponds to sp3 bonded carbon; the D band (green) at 1330–1340 cm−1, indicating defects or disordered sp2 carbon; and the G band (light blue) at around 1590 cm−1, related to in-plane sp2 carbon (graphite-like carbon). The diamond peak is usually located at 1332 cm−1, but it can be shifted to lower Raman shifts in the case of nanodiamonds.62 A fourth band, the D” band (dark blue), is associated to defects probably located on the nanodiamond surface.59–61 In the case of the gold-coated nanodiamonds, peak fitting was not performed due to a marked loss in Raman signal. This attenuation arises from the gold coating, which does not produce a Raman signal and reduces the overall signal intensity through light absorption, scattering and refraction. Notably, the diamond peak was significantly diminished or absent, confirming the presence of a gold layer coating on the nanodiamonds.
The Raman spectra for the 230 NDs and corresponding NDAus are shown in Fig. 3. Interestingly, unlike other NDAu samples, the Ann NDAu 230 sample retains a noticeable diamond peak (Fig. 3C), likely less influenced by the gold shell, which alters Raman signals. This suggests that the Raman signal may still be partially transmitted through a thinner or less uniform gold shell. Supporting this hypothesis, the gold content determined by PIXE analysis (Table 4) shows that this sample contains a comparatively lower amount of gold on the surface. These results point to an inverse correlation between the intensity of the diamond band and the surface gold content, with the Raman signal detection in NDAu systems supporting altered thickness and uniformity of the green synthesised gold layer.
Furthermore, Fig. 3B illustrates the effects of the oxidation process on nanodiamonds following annealing. The D and G bands are considerably less intense than the diamond peak after oxidation treatment, indicating effective removal of surface sp2 carbon species, including graphitic domains and amorphous carbon. This confirms that the oxidation process successfully eliminates non-diamond carbon and residual organic contaminants from the ND surface, enhancing the purity of the sp3 carbon core.
Table 2 presents the ratios derived from the deconvoluted Raman spectra of the 50 nm and 230 nm nanodiamonds, providing insights into their carbon structure. These ratios, calculated from the maxima of each deconvoluted peak, reflect the contributions of different carbon species. Specifically, the ratios of the diamond peak to the G band and the diamond peak to the D band indicate the sp3/sp2 ratio, while the D-to-G ratio relates to the structural ordering of sp2 carbon.63 The data clearly show that the oxides of NDs exhibit a higher diamond Raman signal than sp2-phase carbon. This confirms the efficacy of the oxidation process in selectively etching the graphitic shells. The D-to-G ratio, which decreases as sp2 carbon becomes more ordered, shows minimal variation through the samples. All ND batches underwent an annealing process to promote the conversion of amorphous carbon phases into graphitic ones, to standardise the surface terminations, and to facilitate the oxidation process. This results in comparable ratios for all the analysed NDs, with only a minimal decrease after oxidation at 500 °C.63,64 Notably, the changes in the diamond peak ratios are more pronounced in the 230 nm NDs than in the 50 nm NDs. While the 50 nm samples exhibit approximately a 50% increase in the diamond peak to G band ratio after oxidation (from 1.50 to 2.52), the 230 nm samples demonstrate a more significant increase of around 160% (from 2.79 to 7.63), indicating a higher relative quantity of sp3-bonded carbon compared to sp2-bonded carbon in the larger samples. The greater size of the 230 nm NDs likely contributes to their higher diamond peak ratios in the annealed samples (Ann NDs), as the innermost regions of these larger particles were less affected by the annealing conditions, preserving more of the diamond core. In the Ox NDs, the D-to-G ratio decreases slightly as a result of the oxidation process. This decrease reflects a reduction in disordered sp2 carbon, which is attributed to superficial etching that selectively enhances the sp3-dominant diamond signal.
To assess the crystallinity and grain size of the nanodiamonds and their gold-coated counterparts, powder X-ray diffraction (PXRD) was performed on these samples within the 2θ range of 10° to 100° (Fig. 4 and Fig. S3). The PXRD patterns of the nanodiamonds (NDs) exhibit characteristic diffraction peaks, which can be indexed to the diamond crystal planes (111), (220), and (311), at approximately 2θ = 44°, 2θ = 75°, and 2θ = 91°, respectively (Table 3). These values are consistent with the Crystallography Open Database (COD 9012304) and confirm the pure crystalline nature of annealed and oxidized NDs, with no other crystalline phases detected.65,66
| Sample | Phase | Lattice parameter (Å)* | 2θ (°) | Crystallite size (nm) |
|---|---|---|---|---|
| Ann ND 50 | Diamond | 3.56 | 44.04 | 23.67 ± 0.06 |
| Ox ND 50 | Diamond | 3.56 | 44.04 | 24.92 ± 0.11 |
| Ann ND 230 | Diamond | 3.56 | 44.05 | 33.85 ± 0.09 |
| Ox ND 230 | Diamond | 3.56 | 44.04 | 36.81 ± 0.11 |
| Ann NDAu 50 | Metallic Au | 4.07 | 38.43 | 7.31 ± 0.02 |
| Ox NDAu 50 | Metallic Au | 4.07 | 38.43 | 7.76 ± 0.02 |
| Ann NDAu 230 | Metallic Au | 4.07 | 38.34 | 12.41 ± 0.02 |
| Ox NDAu 230 | Metallic Au | 4.07 | 38.32 | 11.49 ± 0.02 |
For the NDAus, the PXRD patterns reveal eight diffraction peaks. Three of these correspond to the crystal planes of diamond, while the remaining five are indexed to the (111), (200), (220), (311), and (222) lattice planes, characteristic of the face-centered cubic (FCC) structure of metallic gold, in agreement with the Crystallography Open Database (COD 9008463). The patterns confirm the purity of the samples, with no crystalline impurities detected.
Using Scherrer's equation, the crystallite sizes were estimated based on the highest-intensity diffraction peaks. For the gold coating, the crystallite sizes were 7.31 nm and 7.76 nm for the Ann NDAu 50 and Ox NDAu 50 samples, respectively. In the NDAu 230 samples, larger gold crystallite sizes were observed: 12.41 nm for the annealed and 11.49 nm for the oxidized variants, respectively. These values are in good agreement with those reported in the literature for gold nanoparticles synthesized via green chemistry approaches.66 The diamond crystallite sizes in the Ann ND 50 and OX ND 50 were 23.67 nm and 24.92 nm, respectively, while the values increased to 33.85 nm and 36.81 nm for the Ann ND 230 and Ox ND 230 samples, respectively. These differences may reflect the influence of particle size and surface treatments on the diffraction profile. It is important to note that the crystallite size reflects the dimensions of coherent diffracting domains and not necessarily the full physical size of the particles. Deviations between crystallite and particle sizes may arise from the presence of amorphous regions, surface defects, or polycrystalline structures, particularly in the case of coated or functionalized nanoparticles. Additional details regarding the diffraction peaks and crystallite size calculations for additional lattice planes are provided in Tables S2 and S3.
PIXE analysis revealed variations in the composition of the modified nanodiamonds between annealed and oxidized samples. Table 4 presents the mass percentages of gold in the NDAu samples, as determined by PIXE, along with the relative carbon content of both pristine and gold-coated nanodiamonds, determined through dry combustion. These carbon measurements provide additional insight into the extent of surface modification and gold deposition. In general, a decrease in the carbon content after gold coating was observed, which is consistent with the incorporation of heavier elements on the nanodiamond surface.
Trace amounts of other elements, such as calcium, iron, and potassium, were also detected (Table S4), but their contributions were minimal and did not significantly affect the elemental balance.
The PIXE technique, based on the emission of characteristic X-rays following the relaxation of electron clouds excited by energetic proton beams from a particle accelerator, allows for precise multi-elemental analysis. The low error values for all gold measurements (<0.2%) underscore the high analytical precision of the method.
Regarding the NDAu 50 samples, the gold content was slightly higher in the oxidized material (39.50%) compared to the annealed one (37.34%), although the values were relatively close. This minor discrepancy may be attributed to the oxidized nanodiamonds presenting a higher density of oxygen-containing functional groups, which likely enhanced their interaction with the gold-containing reaction bath, promoting more effective adsorption and deposition of gold ions.
In contrast, more pronounced differences were observed for the NDAu 230 samples. The oxidized material exhibited a significantly higher gold content (41.06%) compared to its annealed counterpart (19.62%), more than doubling the amount of deposited gold. The carbon content measurements also reflect this trend, with a more substantial reduction in carbon content after oxidation and coating, consistent with a more extensive surface modification. These results highlight the critical role of oxidative treatment in preparing the surface of larger nanodiamonds for efficient gold deposition. The markedly lower gold content in the annealed sample may be associated with incomplete surface coverage, reduced gold–nanodiamond interaction due to residual surface contaminants, or insufficient surface activation. Conversely, oxidation likely removed such impurities and introduced polar functional groups, increasing the reactivity of the surface.
Another factor that may have contributed to the improved gold deposition on oxidized samples is their enhanced hydrophilicity. Compared to annealed nanodiamonds, which tend to be more hydrophobic, the oxidized surfaces exhibit a greater affinity for aqueous environments. This increased wettability could facilitate more intimate contact between the nanodiamond surface and the plant-derived aqueous reaction medium, thereby improving the efficiency and uniformity of gold ion adsorption and subsequent reduction. Taken together, these observations suggest that the chemical nature of the nanodiamond surface alone does not fully account for the differences in gold uptake. Physical parameters, such as particle size and surface-to-volume ratio, also play a critical role. These trends underscore the influence of particle size and surface-to-volume ratio on gold deposition. Smaller particles (50 nm), with their higher surface-to-volume ratios, appear to support more uniform gold coverage under both annealing and oxidation conditions. Larger particles (230 nm), however, seem to require oxidative treatment to expose sufficient reactive surface for comparable gold deposition. These results highlight the importance of surface chemistry and treatment methods in tailoring gold content and coating quality in nanodiamond modifications, particularly for applications that require precise control over coating properties, such as drug delivery, radiosensitization, or catalysis.
To further investigate these trends, SEM and TEM analyses were performed to characterize the morphology and size distribution of green-synthesized hybrid gold-nanodiamonds, focusing on their shape and size distribution. Fig. 5 shows representative SEM (A–D) and TEM (E–H) images of the NDAu samples. The images indicate that NDAu tend to form small aggregates, which are more evident in the annealed samples. The corresponding particle size distributions of the NDAu samples are reported in Fig. 6.
![]() | ||
| Fig. 5 SEM (left) and TEM (right) images of NDAu annealed (Ann) and oxidized (Ox). 50 nm Ann NDAu (A) and (E); 50 nm Ox NDAu (B) and (F) 230 nm Ann NDAu (C) and (G); 230 nm Ox NDAu (D) and (H). | ||
The direct in situ synthesis of gold nanoparticles in a nanodiamond dispersion typically produces spherical or quasi-spherical particles. SEM and TEM images indicate that numerous small gold nanoparticles are distributed across the ND surfaces. However, the extent of AuNPs coverage varied, and not all diamond surfaces were fully coated. The gold nanoparticles exhibited a broad size distribution, suggesting uncontrolled growth after nucleation on the ND surface. This lack of size regulation can be attributed to the absence of capping agents and the intrinsic non-uniformity of the ND surface, which likely influenced nucleation and growth dynamics.67 In future studies, it would be beneficial to explore variations in both the concentration and type of plant extracts used in the synthesis of AuNPs. Adjusting these parameters can significantly influence the size, shape, and functional properties of the resulting nanoparticles.
For instance, a study by Teimuri-Mofrad et al. (2017) discusses the impact of different plant extracts on the green synthesis of AuNPs, highlighting how variations in extract composition can affect nanoparticle characteristics.68 By systematically varying the extract type and concentration, we can optimize the synthesis process to tailor AuNPs for specific applications, such as drug delivery, catalysis, therapy or imaging.44,46,68–70
Energy dispersive X-ray spectroscopy (EDX) analysis confirmed that the NDAu were composed solely of carbon and gold, with no detectable impurities or artifacts. Red crosses in the Fig. S4–S7 indicate the locations where EDS spectra were acquired (see Fig. S4–S7 and Table S5). The ζ-potential of nanoparticles reflects the electric potential at the interface between the particle surface and the surrounding liquid medium. This potential arises from the distribution of ions around the nanoparticle, particularly within the electric double layer, and serves as a key indicator of colloidal stability. As shown in Table 5, gold coating resulted in a significant decrease in the ζ-potential of all annealed NDAu samples, shifting toward more negative values. The ζ-potential distribution of the nanodiamonds is presented in the (Table S6), providing a visualization of these variations. This trend was particularly evident for Ann NDAu 230, where the ζ-potential decreased by almost 35 mV (from −3.84 mV to −39.3 mV), compared to a decrease of around 30 mV for Ann NDAu 50 (from −26.1 mV to −56.9 mV). This change correlated with better dispersibility in suspension, suggesting that surface modification enhanced the stability of NDAu. Indeed, it is well-known that nanoparticles with high absolute ζ-potential values tend to repel each other more strongly, reducing aggregation and improving stability, while values close to 0 mV (±5 mV) indicate weaker repulsion and potential instability.71 In contrast, the ζ-potential of Ox NDAu 50 remained relatively constant but increased for the Ox NDAu 230 batch. These variations suggest that while surface modification influences ζ-potential, factors such as particle size and surface-to-volume ratio also play a role, with smaller NDAu samples consistently exhibiting more negative ζ-potential values. The observed changes in the ζ-potential of NDAu samples can be significantly influenced by the presence of polyphenolic compounds and carboxyl groups introduced during the synthesis process. These functional groups, commonly found in plant extracts used for nanoparticle synthesis, play a crucial role in determining the surface charge and stability of nanoparticles. Polyphenols, abundant in many plant extracts, contain phenolic hydroxyl (–OH) and carboxyl (–COOH) groups. At physiological pH, these groups can deprotonate, imparting a negative charge on the nanoparticle surface. This increase in negative surface charge enhances the electrostatic repulsion between particles, thereby improving colloidal stability. For instance, a study on chitosan-functionalized hydroxyapatite nanoparticles demonstrated that polyphenols provide a negative ζ-potential due to the deprotonation of phenolic and carboxylic groups at pH 7.4, leading to increased stability of the colloidal system.72,73 The surface charge of nanoparticles, as indicated by ζ-potential, is a critical factor in their interaction with cellular membranes. Cell membranes typically possess a net negative charge due to the presence of phospholipid bilayers. Nanoparticles with highly negative ζ-potentials may experience electrostatic repulsion, potentially reducing cellular uptake. Conversely, nanoparticles with positive ζ-potentials can interact more readily with cell membranes, facilitating uptake but also posing a risk of cytotoxicity due to membrane disruption. A study investigating polymeric nanoparticles found that those with positive surface charges exhibited stronger interactions with cells and higher cytotoxicity compared to negatively charged nanoparticles.74–76
| Sample | Zeta potential, ζ (mV) |
|---|---|
| Ann ND 50 | −26.1 ± 0.6 |
| Ox ND 50 | −54.8 ± 1.8 |
| Ann ND 230 | −3.8 ± 0.2 |
| Ox ND 230 | −67.7 ± 1.1 |
| Ann NDAu 50 | −56.9 ± 1.0 |
| Ox NDAu 50 | −54.9 ± 0.2 |
| Ann NDAu 230 | −39.3 ± 0.6 |
| Ox NDAu 230 | −40.8 ± 0.8 |
Different extracts contain varying amounts and types of polyphenols and other functional groups, which can modulate the surface chemistry and charge of the nanoparticles, thereby impacting key aspects of NDAu stability and its interactions with cells. Given the influence of size, ζ-potential, and other physicochemical characteristics of nanoparticles on their stability and cellular interactions, understanding these surface properties is essential for evaluating NDAu's performance in biomedical applications. The colloidal stability of the NDAu samples was assessed by UV-Vis spectroscopy 0, 3, 6 and 24 hours after suspension in cell culture medium (Fig. S8). The most notable result is the red-shift of the of the characteristic gold plasmon peak, indicative of an increase in the apparent particle size.77 This shift is attributed to the formation of a protein corona, resulting from the adsorption of biomolecules from the medium onto the NDAu surface. Notably, the red-shift is already observable at 0 h, underscoring the rapid onset of protein adsorption. Despite these changes, the gold plasmon peak remains detectable at all time points, suggesting that the particles maintain their overall integrity over 24 h. Accordingly, we investigate the biological evaluation of NDAu, focusing on cell viability by cytotoxicity assays, cell survival, clonogenic assay, and NDAu cellular uptake. These experiments offer valuable insights into the biological behavior of NDAu and their potential applications in radiosensitization and drug delivery.
The cytotoxicity of the synthesized NDAus was, also, evaluated in A549 cells incubated with oxidized and annealed NDAu 50 and NDAu 230 at concentrations ranging from 1 to 100 µg mL−1 MTT assays (Fig. S9 and S10). The results obtained using both the MTT and ATP luminescence assays were similar. After 24 hours of incubation, both Ox NDAu 50 and Ox NDAu 230 showed low toxicity, with cell viability relative to the controls remaining above 80% even at a concentration of 100 µg mL−1 (Fig. S9 and S10). Comparable results were obtained for the annealed NDAu samples (Fig. S9). However, after 48 hours of incubation, the viability of cells treated with Ann NDAu 230 had decreased significantly from 90% to below 70% (Fig. S10). This decrease in cellular viability may be attributed to the larger particle size, increased tendency to aggregate and lower dispersibility, which could lead to NDAu 230 deposition and accumulation on the surface of cells and subsequent interference with cell function.
The results of the clonogenic assay performed on A549 cells are presented in Fig. 8. The cells were incubated with Ox NDAu 50 for 24 hours, in the absence of irradiation, and following proton irradiation. The data indicate that incubation with 10 and 20 µg mL−1 of NDAus significantly reduced the survival fraction (SF). However, control samples incubated with NDAus maintained SF values above 60%, corresponding to a modifying factor ratio (SF without NDAus/SF with NDAus) below 1.6. Upon proton beam irradiation, the SF decreased markedly in both the control group and the NDAu-treated cells (p < 0.001, relative to non-irradiated counterparts). Despite the relatively high radiation dose (17 Gy), which would be expected to attenuate the radiosensitizing effect, a significant difference persisted between the SF of irradiated controls and that of cells treated with 20 µg mL−1 of Ox 50 nm NDAus, with a modifying factor ratio of 2.2. Future studies will investigate lower radiation doses to better evaluate the radiosensitizing potential of these nanoplatforms.
Cellular uptake of NDAu was evaluated using gold (Au) as a marker to localize the nanoplatform. Due to their superior biological activity, only the Ann NDAu 50 and Ox NDAu 50 samples were included in the uptake analysis. A549 cells were incubated with 20 µg mL−1 of Ann and Ox NDAu 50 and the results are presented in Fig. 9. Notably, Ox NDAu exhibited a more uniform Au distribution across the cell and a smoother content gradient compared to the Ann NDAu. This observation is consistent with the superior dispersibility of Ox NDAu 50 in aqueous environments. In contrast, scattered deposits of Au were observed throughout the cells treated with Ann NDAu.
Quantitative analysis was performed using PIXE with matrix composition and thickness parameters derived from EBS. Au concentrations were measured at hotspots and coldspots, regions of high and low Au signal intensity respectively, illustrated in the distribution maps of Fig. 9B and C as points 1 and 2. The quantitative analysis (Fig. 9G) of multiple points across six individual cells per condition revealed distinct and significant differences in Au distribution. Cells treated with Ann NDAu 50 showed a difference in Au concentration of approximately thirteen-fold between hot and cold spots, whereas cells treated with Ox NDAu 50 showed only a three-fold difference. Specifically, in Ann NDAu 50 treated cells, the Au concentration (on a dry weight basis) was 7.0 ± 1.8 mg g−1 in coldspots and 84 ± 22 mg g−1 in hotspots (p < 0.001). For Ox NDAu 50, the corresponding values were 28 ± 4 mg g−1 and 81 ± 9 mg g−1 (p < 0.001). The Au levels in the hotspots did not differ significantly between the Ann and Ox NDAu 50 samples, indicating that these regions likely correspond to agglomerates of NDAu particles. These agglomerates appear larger in cells treated with Ann NDAu 50. However, a significantly higher Au content was found in coldspots of Ox NDAu 50 treated cells than in Ann NDAu 50 (p < 0.01). Three-dimensional (3D) Au distribution maps, corrected for beam energy loss at different depths obtained from EBS spectral analysis78 (Fig. 9D and E), shows a high number of events extending across the full cellular thickness (approximately 900–1000 nm, a typical thickness for freeze-dried cells as measurements were performed in vacuum). This confirms the internalization of Au in Ox NDAu 50 treated A549 cells, and therefore more efficient nanoparticle uptake. In the Ann NDAu 50 treated cells the Au deposits mostly remain in the upper surface layers of the cell.
SEM images of A549 cells also incubated with 20 µg mL−1 Ox NDAu 50 (Fig. 9F), further support NDAu internalization showing distinct features of internalized aggregates compared to those on the cell surface. Taken together, the Au distribution profiles suggest enhanced dispersion of Ox NDAu, likely due to differences in surface modifications that improve dispersion and may facilitate cellular uptake, as evidenced by the more homogeneous Au distribution throughout the cellular volume compared to Ann NDAu 50.
Biological assays performed on A549 lung carcinoma cell line indicated that the NDAus, especially those based on oxidized NDs, were biocompatible at therapeutic concentrations and were taken up efficiently by cells. Higher cellular uptake of Ox NDAu indicates a more biologically active surface modification that can be useful for targeted therapeutic delivery or ND based radiosensitizers. Furthermore, the differences seen in the gold distribution patterns in Ann and Ox samples highlight the significant role that surface chemistry plays in dictating biological interactions. Furthermore, the efficient uptake of Ox NDAu by A549 cells resulted in enhanced proton-induced cell killing, highlighting the importance of further elucidating the mechanisms of radiosensitisation.
In summary, our findings suggest that the green-synthesized NDAu are nanoplatforms for various biomedical uses, such as bioimaging, targeted drug delivery, and radiosensitization. Their prospective application as radiosensitizers in proton therapy provides intriguing prospects for improving outcomes in cancer therapy by augmenting tumor-specific radiation damage while reducing detrimental effects on proximal healthy tissue. Future research efforts will be geared towards assessing the radiosensitizing potential of these NDAu platforms under a proton beam as well as further exploring surface modifications that could be performed on these platforms.
| This journal is © The Royal Society of Chemistry 2026 |