Next Article in Journal
Modeling and Strength Calculations of Parts Made Using 3D Printing Technology and Mounted in a Custom-Made Lower Limb Exoskeleton
Previous Article in Journal
Investigation of the Potential of Repurposing Medium-Density Fiberboard Waste as an Adsorbent for Heavy Metal Ion Removal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Physico-Chemical and Biological Features of Fluorine-Substituted Hydroxyapatite Suspensions

by
Carmen Steluta Ciobanu
1,*,
Daniela Predoi
1,*,
Simona Liliana Iconaru
1,
Mihai Valentin Predoi
2,
Krzysztof Rokosz
3,
Steinar Raaen
4,
Catalin Constantin Negrila
1,
Nicolas Buton
5,
Liliana Ghegoiu
1 and
Monica Luminita Badea
6
1
National Institute of Materials Physics, Atomistilor Street, No. 405A, 077125 Magurele, Romania
2
Department of Mechanics, University Politehnica of Bucharest, BN 002, 313 Splaiul Independentei, Sector 6, 060042 Bucharest, Romania
3
Faculty of Electronics and Computer Science, Koszalin University of Technology, Sniadeckich 2, PL 75-453 Koszalin, Poland
4
Department of Physics, Norwegian University of Science and Technology (NTNU), Realfagbygget E3-124 Høgskoleringen 5, NO 7491 Trondheim, Norway
5
HORIBA Jobin Yvon S.A.S., 6–18, Rue du Canal, 91165 Longjumeau CEDEX, France
6
Faculty of Horticulture, University of Agronomic Sciences and Veterinary Medicine, 59 Marasti Blvd., 011464 Bucharest, Romania
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(14), 3404; https://doi.org/10.3390/ma17143404
Submission received: 20 June 2024 / Revised: 28 June 2024 / Accepted: 8 July 2024 / Published: 10 July 2024

Abstract

:
Infections related to orthopedic/stomatology surgery are widely recognized as a significant health concern. Therefore, the development of new materials with superior biological properties and good stability could represent a valuable alternative to the classical treatments. In this paper, the fluorine-substituted hydroxyapatite (FHAp) suspension, with the chemical formula Ca10(PO4)6(OH)2−2xF2x (where x = 0.05), was prepared using a modified coprecipitation technique. Stability studies were conducted by zeta potential and ultrasound measurements for the first time. The X-ray diffraction (XRD) patterns of FHAp powders displayed a hexagonal structure akin to that of pure hydroxyapatite (HAp). The XPS general spectrum revealed peaks corresponding to the constituent elements of fluorine-substituted hydroxyapatite such as calcium, phosphorus, oxygen, and fluorine. The purity of the obtained FHAp samples was confirmed by energy-dispersive X-ray spectroscopy (EDS) studies. The FHAp morphology was evaluated by scanning electron microscopy (SEM) measurements. Fourier-transform infrared spectroscopy (FTIR) studies were performed in order to study the vibrational properties of the FHAp samples. The FHAp suspensions were tested for antibacterial activity against reference strains such as Staphylococcus aureus 25923 ATCC, Escherichia coli ATCC 25922, and Candida albicans ATCC 10231. Additionally, the biocompatibility of the FHAp suspensions was assessed using human fetal osteoblastic cells (hFOB 1.19 cell line). The results of our biological tests suggest that FHAp suspensions are promising candidates for the future development of new biocompatible and antimicrobial agents for use in the biomedical field.

1. Introduction

Nowadays, hydroxyapatite (HAp) has attracted the attention of the scientific community, especially as a crucial substitute/coating material for biomedical applications (e.g., orthopedics and dentistry applications, etc.) [1,2]. The use of hydroxyapatite in various biomedical applications is attributed to its excellent bioactivity, biocompatibility, and osteoconductivity, as well as its chemical and biological similarity to the mineral component of bone tissue [1,2]. According to previous studies, the enhancement of the HAp features can be achieved through the incorporation of various dopants in their structure [3,4]. HAp possesses a structure that permits the replacement of ions at the Ca2+ and OH sites with various cations and anions [5,6]. Methods such as wet precipitation or the solid-state method have been employed to develop new HAp-based biomaterials doped with Ag+, Cu2+, Mg2+, Zn2+, and F, as documented in prior research [7,8,9,10,11,12]. Fluoride (F), is a trace element found in natural bone, blood, and dental enamel, and it is known to promote and enhance new bone growth and formation [13]. On the other hand, fluoride plays a key role in the prevention of dental caries by strengthening the resilience of tooth surface minerals against acidic dissolution under the oral cavity in low pH conditions [14,15]. Various synthesis methods have been employed to develop fluorine-substituted hydroxyapatite (with various F ion concentration), each yielding to different morphologies and degrees of crystallinity, as follows: wet chemical methods, hydrothermal methods, the multiple emulsion technique, sol–gel, wet precipitation, microwave synthesis, the dry solid-state method, the electrode deposition technique, the modified wet chemical process and mechanochemical techniques [15]. Among these synthesis methods, the one based on the wet chemical precipitation technique has gained popularity due to its simplicity, cost-effectiveness, and ease of application in industrial manufacturing [16].
Staphylococci cause around 80% of orthopedic implant-associated infections, and Staphylococcus aureus (S. aureus) is responsible for one-third of them [17]. Also, Candida species can infect orthopedic implants [17]. Furthermore, both S. aureus and Candida albicans have the ability to form biofilms, which is a key step in the development of implant infections [17]. On the other hand, Escherichia coli (E. coli) is a leading cause of Gram-negative bacterial infections of orthopedic implants [18]. It is well known that infections associated with orthopedic implants pose a significant health issue.
A potential solution to this problem may be repented by the incorporation of antimicrobial substances into hydroxyapatite, which may prevent the growth and development of microbes on the implant surface [17,18]. The study conducted by Nasker, P. and coworkers [19] was focused on the development of hydroxyapatite nano-powders with various fluorine concentrations using hydrothermal processing and by analyzing their structure and effects on cell viability (using mouse osteoblast cell line (MC3T3-E1)) and bacterial growth (against Gram-negative (Escherichia coli) and Gram-positive (Staphylococcus aureus) microbial strains). Their findings underlined that the studied powders were non-toxic and that those with 50% or higher fluorine substitution impeded the proliferation of osteoblast cells [19]. Additionally, fluorine-substituted HAp exhibited moderate antibacterial properties [19].
In the research paper reported by Stanić, V. and colleagues [20], the results of fluorine-substituted hydroxyapatite obtained by the neutralization method were presented [20]. The results of antimicrobial tests showed that the efficiency of the fluorapatite materials against the Streptococcus mutans microbial strain depends on the fluoride concentration [20]. More than that, Tredwin, C.J. et al. [21] reported the fabrication of hydroxyapatite and fluor-hydroxyapatite (with various fluoride ion concentrations) the sol–gel method, along with in vitro biological assay outcomes on the human osteosarcoma (HOS) cell line [21]. Their results highlighted that a high fluoride concentration within the HAp structure leads to enhanced biocompatibility [21]. Another study conducted by S. Shanmugam and coworkers [22] revealed the superior antibacterial efficiency of fluorapatite against Gram-negative (E. coli), Gram-positive (S. aureus) and fungus (C. albicans) strains [22].
In this context, the novelty of this study presents the development of a fluorine-substituted hydroxyapatite (FHAp, Ca10(PO4)6(OH)2−2xF2x, x = 0.05) suspension by an adapted coprecipitation method and its complex characterization from both a physico-chemical and biological point of view. In this manuscript are reported, for the first time, the results of ultrasound studies conducted on FHAp suspensions that provide valuable information about their stability. Other techniques such as X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), dynamic light scattering (DLS) and zeta potential (ZP) were employed in the FHAp sample characterization. The in vitro biological properties of FHAp suspensions were also evaluated.

2. Materials and Methods

2.1. Materials

Calcium nitrate (Ca(NO3)2·4H2O), diammonium hydrogen phosphate ((NH4)2HPO4), ammonium fluoride (NH4F), an ammonium hydroxide (NH4OH) 30% solution and double-distilled water were used in this study. The calcium nitrate, diammonium hydrogen phosphate and ammonium hydroxide were purchased from Sigma-Aldrich, St. Louis, MO, USA. The ammonium fluoride was purchased from Merck Romania S.R.L., Bucharest, Romania.

Synthesis of Fluorine Substituted Hydroxyapatite (FHAp)

The suspension of FHAp (Ca10(PO4)6(OH)2−2xF2x, x = 0.05) was obtained by an adapted co-precipitation method [11,12]. To obtain the suspension, a 0.5 M calcium nitrate solution was prepared. The aqueous solutions of diammonium hydrogen phosphate and ammonium fluoride were added drop by drop to the calcium nitrate solution under continuous stirring at 100 °C. The Ca/P ratio was 1.67. The pH during the synthesis was maintained at 10. The pH was kept constant by adding ammonium hydroxide (NH4OH) 30% solution. At the end of the drip, the resulting solution was stirred for 12 h at 100 °C. At the end, the formed precipitate was washed several times by centrifugation. After the last wash, the resulting precipitate was redispersed in 100 mL of double-distilled water under continuous stirring for 12 h. The resulting suspension was analyzed from a physico-chemical and biological point of view.

2.2. Characterization Techniques

2.2.1. Ultrasound, Dynamic Light Scattering (DLS) and Zeta (ζ) Potential Studies

The stability of the FHAp suspension was evaluated by non-destructive ultrasound (US) studies. For these experiments, double-distilled water was used as reference. The protocol and instrument used for the US studies were described in our previous study [23]. The experimental setup used for the ultrasound measurements of the FHAp suspension is presented in Figure 1 [23]. Therefore, the spectral amplitude as a function of frequency for the FHAp suspension relative to the reference fluid (double-distilled water), in addition to attenuation as a function of frequency, were employed to assess the stability of the FHAp suspension.
The dynamic light scattering (DLS) and Zeta (ζ) potential studies were performed with the aid of a SZ-100 Nanoparticle Analyzer (Horiba-SAS France, Longjumeau, France) at 25 ± 1 °C [23]. For this purpose, the FHAp suspension was diluted in water 10 times prior to the measurements. The reported values represent the mean value of 3 determinations.

2.2.2. X-ray Diffraction

The X-ray diffraction (XRD) patterns were registered using a Bruker D8 Advance diffractometer with CuKα radiation (λ = 1.5418 Å) (Bruker, Karlsruhe, Germany) equipped with a LynxEye™ 1D high-efficiency one-dimensional linear detector. Data regarding the FHAp powders were acquired in the 2θ range of 20–70° with a step size of 0.02° and a time of 5 s per step.

2.2.3. Fourier Transform Infrared Spectroscopy

The Fourier-Transform Infrared spectroscopy (FTIR) spectra for the obtained sample were acquired at ambient temperature using a Perkin Elmer Spectrum BX II spectrometer (Waltham, MA, USA). The FTIR spectrometer was operated in ATR mode. The experimental data were recorded in a wavelength range between 3800 and 450 cm−1.

2.2.4. X-ray Photoelectron Spectroscopy

The X-ray photoelectron spectroscopy (XPS) investigations were conducted using a SES 2002 instrument (Scienta Omicron, Taunusstein, Germany). The XPS measurements utilized a monochromatic Al K (alpha) X-ray source with an energy of 1486.6 eV. The scan analyses and the experiments respected protocols established in the previous studies [24,25]. The CasaXPS 2.3.14 software was used [26] for data analysis. All binding energy (BE) values presented in this study were charge-corrected to C1s at 284.8 eV.

2.2.5. Scanning Electron Microscopy

Scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDS) were used to study the morphology and chemical composition of the FHAp suspension. For this purpose, a Hitachi S4500 microscope (Hitachi, Tokyo, Japan) equipped with an energy-dispersive X-ray spectroscopy system for detection and chemical analysis was employed.

2.3. In Vitro Biological Assays

2.3.1. Antimicrobial Activity

The antibacterial activity of the FHAp suspensions was tested against the following reference strains: Staphylococcus aureus 25923 ATCC, Escherichia coli ATCC 25922, and Candida albicans ATCC 10231. The antimicrobial properties of the FHAp suspensions were determined using microbial suspensions of 1.5 × 108 CFU/mL corresponding to a 0.5 McFarland density obtained from 15 to 18 h bacterial cultures developed on solid media [25,27]. A control was used using the non-inoculated culture media (sample blank). The absorbances were measured at 620 nm, and the microbial viability was represented as a percentage compared to the value obtained for the control sample.
Microbial   cell   viability =   A   s a m p l e A   s a m p l e   b l a n k A   s t r a i n   c o n t r o l A   b l a n k   ×   100

2.3.2. Hemolysis

The hemocompatibility of the FHAp suspensions was evaluated by studying the hemolytic activity. The hemolysis assays were performed using sheep red blood cells (RBCs) using a modified procedure previously described by Das et al. [28]. For this purpose, a 5% RBC suspension was loaded into a tube and incubated at 37 °C for 30 min. Both water and a NaCl solution were added to the RBC suspensions to obtain the positive and negative controls, respectively. The FHAp suspensions at different concentrations were mixed with the RBC suspensions, and all the groups were incubated at 37 °C for 4 h. The hemolysis% of the FHAp suspensions was determined with the aid of the following equation:
Hemolysis   ( % ) = O D   s a m p l e O D   n e g a t i v e   c o n t r o l O D   p o s i t i v e   c o n t r o l O D   n e g a t i v e   c o n t r o l   ×   100
The hemolysis % assays were performed in triplicate and the results were depicted as mean ± SD.
Human fetal osteoblastic cells (hFOB 1.19 cell line) were grown in Ham’s F12: Dulbecco Modified Eagle’s Medium (1:1) without phenol red, with 2.5 mM of L-glutamine and 0.3 mg/mL of G418, and with 10% fetal bovine serum (Gibco, Waltham, MA, USA) at 37 °C in a humidified atmosphere with 5% CO2. The cells were seeded at a cell density of 3 × 104 cells/ cm2 on the tissue culture plastic surface, which served as a control, or on the top of the tested samples, which were previously sterilized under UV light. After 24 h of incubation in standard conditions, biocompatibility tests were performed.

2.3.3. MTT Assay

The biocompatibility of the FHAp suspensions was studied with the aid of Human fetal osteoblastic cells (hFOB 1.19 cell line). For this purpose, the cells were grown in Dulbecco Modified Eagle’s Medium enhanced with L-glutamine and with 10% fetal bovine serum (Gibco, USA) at 37 °C in a humidified atmosphere with 5% CO2. The cells were seeded at a cell density of 3 × 104 cells/cm2 on the tissue culture plastic surface, which served as a control. The cellular viability was determined using the 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT; Sigma-Aldrich, USA) assay. The cellular viability was evaluated at three different time intervals (24, 48 and 72 h). After each incubation period, the medium was removed and the cells were incubated with 1 mg/mL of MTT for 4 h at 37 °C. The absorbance was measured at 595 nm using a microplate reader (Flex Station 3, Molecular Devices, San Jose, CA, USA) and the cell viability was quantified from the absorbance values.

3. Results

The size of the FHAp particles in suspension (hydrodynamic diameter, DH), as well as the stability of these suspensions, were analyzed by classical techniques such as dynamic light scattering (DLS) and zeta potential (ZP). For the FHAp suspensions, stability studies using ultrasonic measurements were also performed to confirm the stability of these suspensions. The Dynamic Light Scattering (DLS) technique and zeta potential are widely used in biochemistry, biotechnology and pharmaceutical development [29,30]. The DLS technique offers the ability to quickly and easily measure the distribution and size of nanoparticles. The disadvantage of this technique is that the analyzed particle suspensions must be diluted. However, by using the DLS technique, we can obtain information about the quality of the sample before using other expensive and time-consuming bio-physical methods. Since errors may appear in the analysis process as a result of the dilution of the suspension obtained following the synthesis process, in this study we used ultrasonic measurements. Ultrasonic measurements were performed on the concentrated suspensions obtained after dilutions. When the DLS data were weighted by numbers, the hydrodynamic diameter (DH) of the FHAp particles in suspension was 65.3 nm (Figure 2a). When the DLS data were weighted by volume, the DH of the FHAp particles in suspension was 78.69 nm (Figure 2b). The measured zeta potential (ZP) value of the FHAp suspension was −27.58 mV, which indicates moderate stability. The measured value for the zeta potential shows that the FHAp suspension does not have good stability.
In order to obtain more accurate information on the stability of the FHAp suspension, ultrasound measurements were performed on undiluted suspensions. First, 100 mL of the FHAp suspension was poured into a special transparent cubical container. Two coaxial ultrasonic transducers were immersed in the suspension and were distanced by 16 mm, facing each other. The transducers’ axis was at 29 mm from the flat bottom of the container box. The FHAp suspension was continuously stirred for 15 min at 750 rot/min in order to obtain a good homogeneity of the solid particles. Immediately after stopping the magnetic stirring machine, the acquisition of the 2000 ultrasonic signals began, recorded every 5 s from the oscilloscope. Each recorded signal was an average of 32 signals on the oscilloscope, reducing the experimental noise.
A superposition of the recorded signals is shown in Figure 3. From right to left, all 2000 signals covering 10,000 s of process evolution are plotted as water flow. The rapid initial evolution of amplitudes was followed by a progressive and slow variation in the overall amplitudes.
This complex evolution is detailed in Figure 4. An initial small decrease is followed by a nonlinear increase in the measured signal amplitudes. The temporal change in properties manifests as a variation in the frequency spectra of all recorded FHAp sample signals. These spectra are shown in Figure 5 and, for comparison, the spectrum of the reference liquid is also plotted (double distilled water, in dotted blue line).
The evolution of signals in time is represented by the wide distribution of the 2000 spectral curves, which differ significantly from the spectrum of the reference liquid, which has a peak at 26.2 MHz. The spectra evolve from ones with lower peak amplitudes of 0.021 at 22 MHz to amplitudes close to those of the reference liquid, at 26 MHz. This important evolution during the experiment indicates a continuous variation in the suspension concentration, which diminishes in time. The ultrasonic signals are also attenuated by the suspension. The time-averaged attenuation plot is shown in Figure 6. Compared against the standard attenuation in the reference liquid (red dotted line), the attenuation is much larger for the FHAp sample in the higher frequency ranges, reaching 103 nepper/m at 35 MHz. In the frequency range of 15–23.5 MHz, the attenuation is lower than the attenuation in the reference liquid.
Another FHAp suspension characteristic is the spectral stability, representing the amplitude of the frequency component of each spectrum, as a function of time (Figure 7).
Some remarkable features can be associated with this sample. During the first 100 s, all spectral amplitudes decrease, a fact that can be attributed to the settlement of the larger particles in suspension. Then, the amplitudes begin to rise in a more or less pronounced manner. For example, the 35 MHz spectral component is no longer following the curve at 32 MHz, but keeping a relatively constant amplitude up to 3000 s, followed by a pronounced increase. A specific feature is the tendency to concentrate the amplitudes after 9000 s of evolution, followed by a tendency to spread again, a phenomenon associated with the passage of the sedimentation plane in front of the transducer’s axis. The relative spectral amplitudes larger than the unit (value for the reference liquid) indicate that the suspension has not sedimented even after 10,000 s.
The FHAp suspension is not very stable; the stability parameter S = d A ¯ A d t = 0.00016   s 1 , in which A is the signal amplitude, with the bar above indicating averaging. This value confirms the continuous slow variation in properties during the test. The results obtained from ultrasound measurements regarding the stability of FHAp suspensions were in agreement with those obtained from zeta potential measurements. The information regarding the stability of suspensions is very important for biological materials because stability directly influences their efficiency [31,32].
The phase composition, lattice parameters, and crystallite size of the FHAp powder obtained by an adapted coprecipitation method were determined by comparing the XRD pattern (Figure 8a) with the standard JCPDS cards for hydroxyapatite (HAp, 9-0432). The X-ray diffraction pattern of the FHAp powder revealed the peaks (002), (210), (211), (300), (202), (310), (222), (213), and (004), showing the formation of hexagonal pure HAp (space group P63/m). The lattice constants were determined through least-squares refinements using the well-defined positions of the most intense reflections, namely (002), (211), (300), (202) and (310). Based on the information provided, it appears that the diffraction maxima moved slightly to the right. This fact was very well observed in the case of the (002) plane when the 2θ value changed from 25.879° (according to JCPDS #09-0432) to 25.942° (Figure 8b). The diffraction spectrum did not exhibit peaks corresponding to other phases.
This behavior could be due to the incorporation of fluorine in the hydroxyapatite structure, as shown in previous reports in the literature [33,34,35]. Moreover, S. Kannan et al. [34] argued that the replacement of OH- ions placed in disordered positions with negatively charged F ions leads to a change in the network parameters. The calculated lattice constants of the FHAp sample were a = b = 0.9358 nm and c = 0.6861 nm. The calculated crystallite size of FHAp was 19.55 nm. The X-ray diffraction (XRD) patterns of the FHAp powders exhibited a hexagonal structure similar to pure hydroxyapatite (in agreement with the JCPDS cards for hydroxyapatite, 9-0432). This resemblance suggests that FHAp retains the fundamental crystallographic arrangement characteristic of HAp. The obtained result is in full agreement with previous studies in which it was shown that when analyzing apatite powders with substituted elements using X-ray diffraction, the resulting diffraction patterns can resemble those of hexagonal hydroxyapatite (HAp), even if there are significant differences in the element concentrations [34,36,37,38,39].
In Figure 9a,b, both the FTIR (absorbance) and 2nd derivative spectra of the FHAp sample are presented. The results of the FTIR studies reveal the presence of the HAp in the analyzed sample without highlighting the formation of new phases following the substitution with F ions. Thus, in the FTIR spectra (Figure 9a), the maxima that belong to the PO43− groups are observed at approximately 470 cm−1 (O–P–O symmetric bending, ν2), 568 cm−1 and 604 cm−1 (O–P–O antisymmetric bending, ν4), 963 cm−1 (P–O symmetric stretching, ν1), and around 1034 cm−1/1096 cm−1 (P–O antisymmetric stretching, ν3) [25,40,41,42,43]. Additionally, the maxima that could be attributed to carbonate groups (ν3 antisymmetric, CO32−) are evident at around 1455 and 1427 cm−1 [25,40,41,42,43]. Furthermore, the maxima observed at 1642 cm−1 could be attributed to the H–O–H bands of the water lattice [44]. The maxima observed in the 3200–3600 cm−1 spectral domain are usually associated with the antisymmetric and symmetric vibrations of water and the OH group within the HAp lattice [45].
The second derivative (SD) studies allowed the differentiation of peaks within the absorption FTIR spectrum that are normally too tightly spaced to be separately identified [46]. We performed these SD studies in the next spectral domains: 500 cm−1–700 cm−1 and 900 cm−1–1150 cm−1. It could be observed that there was no split in the weak maxima centered at 964 cm−1 (that belongs to the ν1 vibration of phosphate groups in the HAp structure) [46]. In the ν4(PO43−) region, sive sub-bands centered at 539 cm−1, 655 cm−1, 589 cm−1, 577 cm−1 and 604 cm−1 were detected [46]. No other intense sub-bands were detected in this spectral domain. The second derivative spectra obtained for the 900 cm−1–1150 cm−1 spectral domain was dominated by the sub-band centered at 1035 cm−1. In the same spectral domain, the sub-bands centered at 1008 cm−1 and 1096 cm−1 were observed. All these sub-bands could be attributed to the ν3 (PO43−) from the hydroxyapatite lattice [46]. The lack of additional significant sub-bands suggests that the studied FHAp samples are pure. These FTIR and SD findings are in good agreement with the ones previously reported by Leung, Y. et al. [46].
The XPS general spectrum of FHAp was collected and is shown in Figure 10a. The XPS general spectrum of FHAp shows peaks corresponding to the constituent elements: calcium (Ca), phosphorus (P), oxygen (O), and fluorine (F). The F1s × ray photoelectron spectroscopy of the samples and its fitting results are also examined in Figure 10b. The narrow XPS scan of F1s revealed two peaks located at 684.2 eV and 686.6 eV, respectively. In agreement with previous studies [47,48], the peak located at 684.2 eV is the imprint of F in the FHAp structure and certifies the fact that fluoride ions were successfully incorporated into the HAp structure. The peak observed at 686.6 eV is due to the reaction between Ca2+ and F ions, in agreement with previous studies [49,50].
Information about the FHAp samples’ morphology was obtained by SEM. For these studies, a drop of FHAp suspension was placed on the carbon tape and dried prior to SEM examination. The results of the scanning electron microscopy studies and EDS studies are presented in Figure 11. The SEM micrographs of FHAp reveal that the particles were obtained at a nanometric scale and had a pronounced tendency to form agglomerates.
Information about the chemical composition of the FHAp suspension was obtained via EDS studies and their results are shown in Figure 11b. The main line observed in the EDS spectra belongs to the calcium (Ca K), phosphate (P K), oxygen (O K) and fluorine (F K). All these chemical elements are the main constituents of the FHAp structure. No other noticeable lines could be observed in Figure 11b, which suggests the FHAp’s purity. These results are consistent with the ones obtained in XPS and XRD studies.
The biological properties of the FHAp suspensions were evaluated through both hemocompatibility and biocompatibility studies. The hemolytic properties of the materials suggested their ability to induce lysis or rupture in red blood cells (erythrocytes), therefore indicating the biocompatibility of the tested materials. These studies are crucial because they reveal important information about their safety and biocompatibility regarding their use in contact with blood or biological tissues. Materials that exhibit high hemolytic activity can cause damage to red blood cells, release hemoglobin, and cause adverse physiological responses like inflammation, thrombosis, and organ damage, making them unsuitable for biomedical applications. On the other hand, materials that possess a low hemolytic index are considered suitable for use in biomedical applications because they pose a minimal risk of adverse reactions and have better compatibility with biological systems. The hemocompatibility of the FHAp suspensions was tested according to their hemolytic activity. The results are depicted in Figure 12.
The data obtained through the hemocompatibility assays showed that none of the tested concentrations of FHAp suspensions caused hemolysis. Furthermore, the values obtained were well within the acceptable hemocompatibility limits for a biomaterial. The results highlighted that the FHAp suspensions exhibited a hemolytic activity lower than 1%, with the hemolysis index increasing with concentration. The results of the hemocompatibility assays demonstrated that the FHAp suspensions exhibited a low hemolytic index, making them suitable for further cytotoxicity assessments to confirm their efficacy and safety in biomedical applications [51].
The cytotoxic response of the FHAp suspensions was investigated with the aid of MTT cell viability and LDH release analyses. The cytotoxic response of the FHAp suspensions towards hFOB 1.19 cells was evaluated at three different time intervals (24, 48 and 72 h). The results of the LDH release and MTT cell viability assays are depicted graphically in Figure 13a,b. The results highlighted that the FHAp suspensions did not exhibit any adverse effects on hFOB 1.19 cells at the tested time intervals. The MTT assay results indicated that the cellular viability after 24 h of incubation with the FHAp suspensions was higher than 96% and reached a value of 99% after 72 h of incubation. The results are in good agreement, confirming the biocompatibility of hydroxyapatite and fluoride-doped hydroxyapatite on different cell lines [13,19,52,53,54,55,56,57,58,59,60,61,62,63,64]. These findings were also confirmed by the results obtained from the extracellular LDH release analysis (Figure 13a). The LDH assay measures the lactate dehydrogenase released into the environment from the cytoplasm due to cell membrane damage. The results obtained from the LDH-specific activity assay confirmed the MTT results and highlighted that the FHAp suspensions did not exhibit any cytotoxic activity against the hFOB 1.19 cells. Both the MTT and LDH results indicated that the FHAp suspension exhibited good biocompatibility towards hFOB 1.19 cells, and all the data suggested that the FHAp suspensions could be successfully used in biomedical applications.
In recent years, the emergence of microbial strains with resistance to conventional antibiotics has become a major global health issue. Due to the limited availability of effective antimicrobial agents and therapies for clinical use, there is a growing need for the development of novel antimicrobial agents. Consequently, the development of new materials with enhanced biological properties and efficacy against drug-resistant microbial strains is of significant global interest. Ongoing research and innovation in this field are crucial to addressing the challenges posed by drug-resistant microorganisms to the global health system. In this study, the antimicrobial activity of FHAp suspensions was evaluated. The antimicrobial activity of the FHAp suspensions was tested against some of the most common microbial strains responsible for the appearance of infections, namely Staphylococcus aureus 25923 ATCC, Escherichia coli 25922 ATCC, and Candida albicans 10231 ATCC strains. The antimicrobial assays were performed at three different time intervals and the experiments were performed in triplicate. The results were depicted graphically as mean ± SD.
The results of the quantitative antimicrobial assays presented in Figure 14, Figure 15 and Figure 16 show that the FHAp suspensions inhibited the growth of all the microbial strains tested for all the tested intervals. Moreover, the results emphasize that the FHAp suspensions reduced the microbial cell viability by more than 50% for all the tested microbial strains, even in the first stages of development (first 24 h). On the other hand, the results highlight that the antimicrobial activity of the FHAp suspensions was correlated with the suspension concentrations, as well as with the incubation time and the type of microbial strain that they were tested against.
The results showed that the tested microbial strain was more susceptible to higher concentrations of FHAp suspensions and also that the reduction in microbial cell viability decreased with the increase in the exposure time. Therefore, the most prominent reduction in the microbial cell viability was achieved after 72 h of exposure to the FHAp suspensions in the case of all the tested microbial strains. More than that, the results showed that the microbial strains most susceptible to the FHAp suspensions was the E. coli bacterial strain. As is depicted in Figure 14, Figure 15 and Figure 16, all the tested microbial strains were inhibited by the FHAp suspensions at even lower concentrations of 0.09 mg/mL. The data also suggest that the lowest antimicrobial activity was against the fungal strain C. albicans. These results are in good agreement with other studies that have suggested that C. albicans appears more resistant to silver fluorapatite than the bacterial species tested [65]. This behavior is attributed to the fact that Candida albicans cells are typically surrounded by an exopolymeric substance that protects them from adverse environmental conditions. The results of the antibacterial assays are in good agreement with previous research on the antibacterial properties of fluoride and fluoride composites [13,19,20,53,54,55,56,65,66,67,68,69,70,71,72,73,74]. In their work, Kus-Liskiewicz et al. [65] showed that there is a time-dependent bactericidal effect on the gram-negative E. coli strain when it is exposed to pure fluorapatite (FAP). Similar results were obtained by Bala et al. [66] in their studies emphasizing the strong antibacterial activity that is exhibited by calcium fluoride nanoparticles against Escherichia coli, Pseudomonas aeruginosa, Bacillus badius, and Staphylococcus aureus. The antimicrobial activity of the FHAp was attributed to the presence of F ions in the HAp lattice. Even though the exact mechanisms responsible for the antimicrobial activity of nanomaterials are still unclear, there are several proposed mechanisms that could be responsible for the antimicrobial activity of fluoride ions. The fluoride ions released from FHA can affect the microbial metabolism in numerous ways. For example, they can inhibit the glycolytic enzyme enolase. More than that, the fluoride complexes can mimic phosphate, forming complexes with ADP at enzyme reaction centers, leading to the inhibition of proton-translocating F-ATPases. ATPase is crucial for maintaining the intracellular pH by pumping out protons, and its inhibition disrupts the microbial metabolism and aciduric capabilities. Beyond that, fluoride ions could inhibit acid-producing bacteria such as Fusobacterium nucleatum (F. n) and Treponema species, which are periodontal pathogens, thus helping to maintain the acid–base balance necessary for proper osteogenesis.
A schematic representation of the proposed antimicrobial mechanism of FHAp is depicted in Figure 17.
Fluoride ions are released as bacterial metabolism begins and the pH decreases, which inhibits bacterial growth, stabilizes the microenvironment, and resists the inflammatory process. Another mechanism by which F ions may interact with bacterial cells is through interference with key metabolic enzymes. However, when developing new fluoride composites, it is essential to consider that the fluoride concentrations required for antimicrobial effects often exceed those needed to reduce apatite solubility [19,20,54,55,56,65,67,68,69,70,71,72,73,74].
The antimicrobial activity of FHAp suspensions could be attributed to a series of mechanisms. First, it can be attributed to the release of fluoride ions that have the ability to disrupt the bacterial metabolism and impair enzyme function, thus compromising the membrane’s integrity and inducing oxidative stress. All of these combined effects lead to the death of bacterial cells, making FHAp suspensions suitable for use in antimicrobial applications.
The results of our studies are in alignment with previously reported studies regarding the antimicrobial activity of fluoride-based materials. Furthermore, our data suggest that FHAp suspensions could be successfully used for the future development of antimicrobial agents.

4. Conclusions

In this study, we report the development of a FHAp suspension through an adapted coprecipitation method. The degree of stability of the FHAp suspension was evaluated by zeta potential and ultrasound measurements. Both measurements showed good stability. X-ray diffraction analysis of the FHAp powders resulted in diffraction patterns that resembled pure HAp with a hexagonal structure. The XPS analysis confirms the presence of Ca, P, O, and F in FHAp, and the specific peaks provide valuable insights into the chemical composition and incorporation of fluoride ions. The nanometric dimension of the particles and their tendency to form an agglomerate were revealed by the SEM results. The presence of hydroxyapatite and the lack of supplementary phases in the analyzed samples were underlined by the FTIR results. The results of the in vitro biological assay suggest that FHAp suspensions could be employed for the future development of new antimicrobial agents.

Author Contributions

Conceptualization, C.S.C., M.V.P., S.L.I. and D.P.; methodology, C.S.C., D.P., S.L.I. and M.V.P.; software, C.S.C., D.P., S.L.I. and M.V.P.; validation, C.S.C., D.P., S.L.I., M.V.P., K.R., S.R., N.B., C.C.N., L.G. and M.L.B.; formal analysis, C.S.C., D.P., S.L.I., M.V.P., K.R., S.R., N.B., C.C.N., L.G. and M.L.B.; investigation, C.S.C., D.P., S.L.I., M.V.P., K.R., S.R., N.B., C.C.N., L.G. and M.L.B.; resources, C.S.C., D.P., S.L.I. and M.V.P.; data curation, C.S.C., D.P., S.L.I., M.V.P., K.R. and S.R.; writing—original draft preparation, C.S.C., D.P., S.L.I., M.V.P., K.R., S.R., N.B., C.C.N., L.G. and M.L.B.; writing—review and editing, C.S.C., D.P., S.L.I., M.V.P., K.R., S.R., N.B., C.C.N., L.G. and M.L.B.; visualization, C.S.C., D.P., S.L.I., M.V.P., K.R., S.R., N.B., C.C.N., L.G. and M.L.B.; supervision, M.V.P., S.L.I., C.S.C. and D.P.; project administration, C.S.C. and D.P.; funding acquisition, D.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Core Program of the National Institute of Materials Physics, granted by the Romanian Ministry of Research, Innovation and Digitalization through the Project PC1-PN23080101.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

Author Nicolas Buton was employed by the company HORIBA Jobin Yvon S.A.S. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Hench, L.L. Bioceramics: From concept to clinic. J. Am. Ceram. Soc. 1991, 74, 1487–1510. [Google Scholar] [CrossRef]
  2. Rajabi-Zamani, A.H.; Behnamghader, A.; Kazemzadeh, A. Synthesis of nanocrystalline carbonated hydroxyapatite powder via nonalkoxide sol–gel method. Mater. Sci. Eng. C 2008, 28, 1326–1329. [Google Scholar] [CrossRef]
  3. Vu, A.A.; Robertson, S.F.; Ke, D.; Bandyopadhyay, A.; Bose, S. Mechanical and biological properties of ZnO, SiO2, and Ag2O doped plasma sprayed hydroxyapatite coating for orthopaedic and dental applications. Acta Biomater. 2019, 92, 325–335. [Google Scholar] [CrossRef] [PubMed]
  4. Roy, M.; Fielding, G.A.; Beyenal, H.; Bandyopadhyay, A.; Bose, S. Mechanical, in vitro antimicrobial, and biological properties of plasma-sprayed silver-doped hydroxyapatite coating. ACS Appl. Mater. Interfaces 2012, 4, 1341–1349. [Google Scholar] [CrossRef] [PubMed]
  5. White, T.; Ferraris, C.; Kim, J.; Madhavi, S. Apatite–An adaptive framework structure. Rev. Mineral. Geochem. 2005, 57, 307–401. [Google Scholar] [CrossRef]
  6. Sarath Chandra, V.; Elayaraja, K.; Thanigai Arul, K.; Ferraris, S.; Spriano, S.; Ferraris, M.; Asokan, K.; Narayana Kalkura, S. Synthesis of magnetic hydroxyapatite by hydrothermal–microwave technique: Dielectric, protein adsorption, blood compatibility and drug release studies. Ceram. Int. 2015, 41, 13153–13163. [Google Scholar] [CrossRef]
  7. Banerjee, D.; Bose, S. Effects of aloe vera gel extract in doped hydroxyapatite-coated titanium implants on in vivo and in vitro biological properties. ACS Appl. Bio Mater 2019, 2, 3194–3202. [Google Scholar] [CrossRef] [PubMed]
  8. Basu, S.; Basu, B. Unravelling doped biphasic calcium phosphate: Synthesis to application. ACS Appl. Bio Mater. 2019, 2, 5263–5297. [Google Scholar] [CrossRef] [PubMed]
  9. Baikie, T.; Ng, G.M.H.; Madhavi, S.; Pramana, S.S.; Blake, K.; Elcombe, M.; White, T.J. The crystal chemistry of the alkaline-earth apatites A10(PO4)6CuxOy(H)z(A = Ca, Sr and Ba). Dalt. Trans 2009, 2009, 6722–6726. [Google Scholar] [CrossRef]
  10. Predoi, S.A.; Ciobanu, S.C.; Chifiriuc, C.M.; Iconaru, S.L.; Predoi, D.; Negrila, C.C.; Marinas, I.C.; Raaen, S.; Rokosz, K.; Motelica-Heino, M. Sodium bicarbonate-hydroxyapatite used for removal of lead ions from aqueous solution. Ceram. Int. 2024, 50, 1742–1755. [Google Scholar] [CrossRef]
  11. Ciobanu, C.S.; Massuyeau, F.; Constantin, L.V.; Predoi, D. Structural and physical properties of antibacterial Ag-doped nano-hydroxyapatite synthesized at 100 C. Nanoscale Res. Lett. 2011, 6, 1–8. [Google Scholar] [CrossRef]
  12. Predoi, D.; Iconaru, S.L.; Ciobanu, S.C.; Predoi, S.-A.; Buton, N.; Megier, C.; Beuran, M. Development of Iron-Doped Hydroxyapatite Coatings. Coatings 2021, 11, 186. [Google Scholar] [CrossRef]
  13. Wang, L.; Wang, M.; Li, M.; Shen, Z.; Wang, Y.; Shao, Y.; Zhu, Y. Trace fluorine substituted calcium deficient hydroxyapatite with excellent osteoblastic activity and antibacterial ability. Cryst. Eng. Comm. 2018, 20, 5744–5753. [Google Scholar] [CrossRef]
  14. Bhattacharjee, A.; Bandyopadhyay, A.; Bose, S. Plasma sprayed fluoride and zinc doped hydroxyapatite coated titanium for load-bearing implants. Surf. Coat. Technol. 2022, 440, 128464. [Google Scholar] [CrossRef]
  15. Feroz, S.; Khan, A.S. Fluoride-substituted hydroxyapatite. In Handbook of Ionic Substituted Hydroxyapatites; Woodhead Publishing: Sawston, UK, 2020; pp. 175–196. [Google Scholar]
  16. Gentile, P.; Wilcock, C.J.; Miller, C.A.; Moorehead, R.; Hatton, P.V. Process Optimisation to Control the Physico-Chemical Characteristics of Biomimetic Nanoscale Hydroxyapatites Prepared Using Wet Chemical Precipitation. Materials 2015, 8, 2297–2310. [Google Scholar] [CrossRef]
  17. Tan, G.; Xu, J.; Chirume, W.M.; Zhang, J.; Zhang, H.; Hu, X. Antibacterial and Anti-Inflammatory Coating Materials for Orthopedic Implants: A Review. Coatings 2021, 11, 1401. [Google Scholar] [CrossRef]
  18. Crémet, L.; Broquet, A.; Brulin, B.; Jacqueline, C.; Dauvergne, S.; Brion, R.; Asehnoune, K.; Corvec, S.; Heymann, D.; Caroff, N. Pathogenic potential of Escherichia coli clinical strains from orthopedic implant infections towards human osteoblastic cells. Pathog. Dis. 2015, 73, ftv065. [Google Scholar] [CrossRef]
  19. Nasker, P.; Mukherjee, M.; Kant, S.; Tripathy, S.; Sinha, A.; Das, M. Fluorine substituted nano hydroxyapatite: Synthesis, bio-activity and antibacterial response study. Ceram. Int. 2018, 44, 22008–22013. [Google Scholar] [CrossRef]
  20. Stanić, V.; Dimitrijević, S.; Antonović, D.G.; Jokić, B.M.; Zec, S.P.; Tanasković, S.T.; Raičević, S. Synthesis of fluorine substituted hydroxyapatite nanopowders and application of the central composite design for determination of its antimicrobial effects. Appl. Surf. Sci. 2014, 290, 346–352. [Google Scholar] [CrossRef]
  21. Tredwin, C.J.; Young, A.M.; Abou Neel, E.A.; Georgiou, G.; Knowles, J.C. Hydroxyapatite, fluor-hydroxyapatite and fluorapatite produced via the sol–gel method: Dissolution behaviour and biological properties after crystallisation. J. Mater. Sci. Mater. Med. 2014, 25, 47–53. [Google Scholar] [CrossRef]
  22. Shanmugam, S.; Gopal, B. Copper substituted hydroxyapatite and fluorapatite: Synthesis, characterization and antimicrobial properties. Ceram. Int. 2014, 40, 15655–15662. [Google Scholar] [CrossRef]
  23. Predoi, D.; Iconaru, S.L.; Predoi, M.V.; Motelica-Heino, M.; Guegan, R.; Buton, N. Evaluation of Antibacterial Activity of Zinc-Doped Hydroxyapatite Colloids and Dispersion Stability Using Ultrasounds. Nanomaterials 2019, 9, 515. [Google Scholar] [CrossRef] [PubMed]
  24. Iconaru, S.L.; Predoi, M.V.; Chapon, P.; Gaiaschi, S.; Rokosz, K.; Raaen, S.; Motelica-Heino, M.; Predoi, D. Investigation of Spin Coating Cerium-Doped Hydroxyapatite Thin Films with Antifungal Properties. Coatings 2021, 11, 464. [Google Scholar] [CrossRef]
  25. Iconaru, S.L.; Groza, A.; Gaiaschi, S.; Rokosz, K.; Raaen, S.; Ciobanu, S.C.; Chapon, P.; Predoi, D. Antimicrobial Properties of Samarium Doped Hydroxyapatite Suspensions and Coatings. Coatings 2020, 10, 1124. [Google Scholar] [CrossRef]
  26. Casa Software Ltd. CasaXPS: Processing Software for XPS, AES, SIMS and More. 2009. Available online: www.casaxps.com (accessed on 10 March 2024).
  27. Marinas, I.C.; Ignat, L.; Maurușa, I.E.; Gaboreanu, M.D.; Adina, C.; Popa, M.; Chifiriuc, M.C.; Angheloiu, M.; Georgescu, M.; Iacobescu, A.; et al. Insights into the physico-chemical and biological characterization of sodium lignosulfonate-silver nanosystems designed for wound management. Heliyon 2024, 10, e26047. [Google Scholar] [CrossRef] [PubMed]
  28. Das, D.; Nath, B.C.; Phukon, P.; Kalita, A.; Dolui, S.K. Synthesis of ZnO nanoparticles and evaluation of antioxidant and cytotoxic activity. Colloids Surf. B Biointerfaces 2013, 111, 556–560. [Google Scholar] [CrossRef] [PubMed]
  29. Powers, K.W.; Brown, S.C.; Krishna, V.B.; Wasdo, S.C.; Moudgil, B.M.; Roberts, S.M. Research strategies for safety evaluation of nanomaterials. Part VI. Characterization of nanoscale particles for toxicological evaluation. Toxicol. Sci. 2006, 90, 296–303. [Google Scholar] [CrossRef] [PubMed]
  30. Murdock, R.C.; Braydich-Stolle, L.; Schrand, A.M.; Schlager, J.J.; Hussain, S.M. Characterization of nanomaterial dispersion in solution prior to in vitro exposure using dynamic light scattering technique. Toxicol. Sci. 2008, 101, 239–253. [Google Scholar] [CrossRef]
  31. Sharma, S.; Shukla, P.; Misra, A.; Mishra, P.R. Interfacial and colloidal properties of emulsified systems: Pharmaceutical and biological perspective. In Colloid and Interface Science in Pharmaceutical Research and Development; Elsevier: Amsterdam, The Netherlands, 2014; pp. 149–172. [Google Scholar] [CrossRef]
  32. Moriarty, T.F.; Poulsson, A.H.C.; Rochford, E.T.J.; Richards, R.G. 4.8 Bacterial Adhesion and Biomaterial Surfaces; Ducheyne, P., Ed.; Comprehensive Biomaterials II; Elsevier: Amsterdam, The Netherlands, 2017; pp. 101–129. [Google Scholar] [CrossRef]
  33. Rodríguez-Lorenzo, L.M.; Hart, J.N.; Gross, K.A. Influence of fluorine in the synthesis of apatites. Synthesis of solid solutions of hydroxy-fluorapatite. Biomaterials 2003, 24, 3777–3785. [Google Scholar] [CrossRef] [PubMed]
  34. Kannan, S.; Rebelo, A.; Ferreira, J.M.F. Novel synthesis and structural characterization of fluorine and chlorine co-substituted hydroxyapatites. J. Inorg. Biochem. 2006, 100, 1692–1697. [Google Scholar] [CrossRef]
  35. Rodríguez-Lorenzo, L.M.; Hart, J.N.; Gross, K.A. Structural and chemical analysis of well-crystallized hydroxyfluorapatites. J. Phys. Chem. B 2003, 107, 8316–8320. [Google Scholar] [CrossRef]
  36. Gibson, I.R.; Rehman, I.; Best, S.M.; Bonfield, W. Characterization of the transformation from calcium deficient apatite to b-tricalcium phosphate. J. Mater. Sci. Mater. Med. 2000, 12, 799–805. [Google Scholar] [CrossRef] [PubMed]
  37. Kannan, S.; Lemos, I.A.F.; Rocha, J.H.G.; Ferreira, J.M.F. Synthesis and characterization of magnesium substituted biphasic mixtures of controlled hydroxyapatite/b-tricalcium phosphate ratios. J. Solid State Chem. 2005, 178, 3190–3196. [Google Scholar] [CrossRef]
  38. Gibson, I.R.; Bonfield, W. Preparation and characterization of magnesium/carbonate co-substituted hydroxyapatites. J. Mater. Sci. Mater. Med. 2002, 13, 685–693. [Google Scholar] [CrossRef] [PubMed]
  39. Ptáček, P. Apatites and their Synthetic Analogues Synthesis, Structure, Properties and Applications; BoD–Books on Demand: Norderstedt, Germany, 2016. [Google Scholar] [CrossRef]
  40. Predoi, D.; Ciobanu, S.C.; Iconaru, S.L.; Predoi, M.V. Influence of the Biological Medium on the Properties of Magnesium Doped Hydroxyapatite Composite Coatings. Coatings 2023, 13, 409. [Google Scholar] [CrossRef]
  41. Predoi, S.A.; Ciobanu, S.C.; Chifiriuc, M.C.; Motelica-Heino, M.; Predoi, D.; Iconaru, S.L. Hydroxyapatite Nanopowders for Effective Removal of Strontium Ions from Aqueous Solutions. Materials 2023, 16, 229. [Google Scholar] [CrossRef] [PubMed]
  42. Kannan, S.; Rocha, J.H.G.; Agathopoulos, S.; Ferreira, J.M.F. Fluorine-substituted hydroxyapatite scaffolds hydrothermally grown from aragonitic cuttlefish bones. Acta Biomater. 2007, 3, 243–249. [Google Scholar] [CrossRef] [PubMed]
  43. Kannan, S.; Ventura, J.M.; Ferreira, J.M.F. In situ formation and characterization of flourine-substituted biphasic calcium phosphate ceramics of varied F-HAP/β-TCP ratios. Chem. Mater. 2005, 17, 3065–3068. [Google Scholar] [CrossRef]
  44. Arends, J.; Christoffersen, J.; Christoffersen, M.R.; Eckert, H.; Fowler, B.O.; Heughebaert, J.C.; Nancollas, G.H.; Yesinowski, J.P.; Zawacki, S.J. A calcium hydroxyapatite precipitated from an aqueous solution; an international multimethod analysis. J. Cryst. Growth 1987, 84, 512–532. [Google Scholar] [CrossRef]
  45. Constantin, L.V.; Iconaru, S.; Ciobanu, C.S. Europium doped hydroxyapatite for applications in environmental field. Rom. Rep. Phys. 2012, 64, 788–794. [Google Scholar]
  46. Leung, Y.; Walters, M.A.; LeGeros, R.Z. Second derivative infrared spectra of hydroxyapatite. Spectrochim. Acta Part A Mol. Spectrosc. 1990, 46, 1453–1459. [Google Scholar] [CrossRef]
  47. Stranick, M.A.; Root, M.J. Influence of strontium on monofluorophosphate uptake by hydroxyapatite XPS characterization of the hydroxyapatite surface. Colloids Surf. 1991, 55, 137–147. [Google Scholar] [CrossRef]
  48. Wang, Y.; Zhang, S.; Zeng, X.; Ma, L.L.; Weng, W.; Yan, W.; Qian, M. Osteoblastic cell response on fluoridated hydroxyapatite coatings. Acta Biomater. 2007, 3, 191–197. [Google Scholar] [CrossRef]
  49. Cheng, K.; Han, G.; Weng, W.; Qu, H.; Du, P.; Shen, G.; Yang, J.; Ferreira, J.M.F. Sol–gel derived fluoridated hydroxyapatite films. Mater. Res. Bull. 2003, 38, 89–97. [Google Scholar] [CrossRef]
  50. Cheng, K.; Zhang, S.; Weng, W. The F content in sol–gel derived FHA coatings: An XPS study. Surf. Coat. Technol. 2005, 198, 237–241. [Google Scholar] [CrossRef]
  51. Radha, G.; Balakumar, S.; Venkatesan, B.; Vellaichamy, E. Evaluation of hemocompatibility and in vitro immersion on microwaveassisted hydroxyapatite–alumina nanocomposites. Mater. Sci. Eng. C 2015, 50, 143–150. [Google Scholar] [CrossRef]
  52. Qu, H.; Wei, M. The effect of fluoride contents in fluoridated hydroxyapatite on osteoblast behavior. Acta Biomater. 2006, 2, 113–119. [Google Scholar] [CrossRef]
  53. Yin, X.; Bai, Y.; Zhou, S.-j.; Ma, W.; Bai, X.; Chen, W. solubility, mechanical and biological properties of fluoridated hydroxyapatite/calcium silicate gradient coatings for orthopedic and dental applications. J. Therm. Spray Technol. 2020, 29, 471–488. [Google Scholar] [CrossRef]
  54. Sun, J.; Wu, T.; Fan, Q.; Hu, Q.; Shi, B. Comparative study of hydroxyapatite, fluor-hydroxyapatite and Si-substituted hydroxyapatite nanoparticles on osteogenic, osteoclastic and antibacterial ability. RSC Adv. 2019, 9, 16106–16118. [Google Scholar] [CrossRef]
  55. Gholipourmalekabadi, M.; Sameni, M.; Hashemi, A.; Zamani, F.; Rostami, A.; Mozafari, M. Silver- and fluoride-containing mesoporous bioactive glasses versus commonly used antibiotics: Activity against multidrug-resistant bacterial strains isolated from patients with burns. Burns 2016, 42, 131–140. [Google Scholar] [CrossRef]
  56. Cheng, K.; Weng, W.; Wang, H.; Zhang, S. In vitro behavior of osteoblast-like cells on fluoridated hydroxyapatite coatings. Biomaterials 2005, 26, 6288–6295. [Google Scholar] [CrossRef]
  57. Cai, Y.; Zhang, S.; Zeng, X.; Wang, Y.; Qian, M.; Weng, W. Improvement of bioactivity with magnesium and fluorine ions incorporated hydroxyapatite coatings via sol–gel deposition on Ti6Al4V alloys. Thin Solid Film. 2009, 517, 5347–5351. [Google Scholar] [CrossRef]
  58. Lau, K.H.W.; Baylink, D.J. Molecular Mechanism of Action of Fluorideon Bone Cells. J. Bone Miner. Res. 1998, 13, 1660–1667. [Google Scholar] [CrossRef]
  59. Farley, J.R.; Wergedal, J.E.; Baylink, D.J. Fluoride directly stimulates proliferation and alkaline phosphatase activity of bone forming cells. Science 1983, 222, 330–332. [Google Scholar] [CrossRef]
  60. Baylink, D.J. Serum fluoride levels. In Primers on the Metabolic Bone Diseases and Disorders of Mineral Metabolism, 2nd ed.; Favus, M.J., Ed.; Raven Press: New York, NY, USA, 1993; pp. 262–263. [Google Scholar]
  61. Hall, B.K. Sodium fluoride as an initiator of osteogenesi from embryonic mesenchyme in vitro. Bone 1987, 8, 111–116. [Google Scholar] [CrossRef]
  62. Wergedal, J.E.; Lau, K.-H.W.; Baylink, D.J. Fluoride and bovine bone extract influence cell proliferation and phosphatase activities in human bone cell cultures. Clin. Orthop. Relat. Res. 1988, 233, 274–282. [Google Scholar] [CrossRef]
  63. Modrowski, D.; Miravet, L.; Feuga, M.; Bannie, F.; Marie, P.J. Effect of fluoride on bone and bone cells in ovariectomized rats. J Bone Miner. Res 1992, 7, 961–969. [Google Scholar] [CrossRef]
  64. Bellows, C.G.; Aubin, J.E.; Heersche, J.N.M. Differential effects of fluoride during initiation and progression of mineralization of osteoid nodules formed in vitro. J. Bone Miner. Res. 1993, 8, 1357–1363. [Google Scholar] [CrossRef]
  65. Kus-Liśkiewicz, M.; Rzeszutko, J.; Bobitski, Y.; Barylyak, A.; Nechyporenko, G.; Zinchenko, V.; Zebrowski, J. Alternative Approach for Fighting Bacteria and Fungi: Use of Modified Fluorapatite. J. Biomed. Nanotechnol. 2019, 15, 848–855. [Google Scholar] [CrossRef]
  66. Bala, W.A.; Benitha, V.S.; Jeyasubramanian, K.; Hikku, G.S.; Sankar, P.; Kumar, S.V. Investigation of anti-bacterial activity and cytotoxicity of calcium fluoride nanoparticles. J. Fluor. Chem. 2017, 193, 38–44. [Google Scholar] [CrossRef]
  67. Nurhaerani; Arita, K.; Shinonaga, Y.; Nishino, M. Plasma-based fluorine ion implantation into dental materials for inhibition of bacterial adhesion. Dent. Mater. J. 2006, 25, 684. [Google Scholar]
  68. Yoshinari, M.; Oda, Y.; Kato, T.; Okuda, K. Influence of surface modifications to titanium on antibacterial activity in vitro. Biomaterials 2001, 22, 2043. [Google Scholar] [CrossRef]
  69. Hayacibara, M.F.; Rosa, O.P.D.S.; Koo, H.; Torres, S.A.; Costa, B.; Cury, J.A. Effects of fluoride and aluminum from ionomeric materials on S. mutans biofilm. J. Dent. Res. 2003, 82, 267–271. [Google Scholar] [CrossRef]
  70. Marquis, R.E. Antimicrobial actions of fluoride for oral bacteria. Can. J. Microbiol. 1995, 41, 955–964. [Google Scholar] [CrossRef]
  71. Marquis, R.E.; Clock, S.A.; Mota-Meira, M. Fluoride and organic weak acids as modulators of microbial physiology. FEMS Microbiol. Rev. 2003, 26, 493–510. [Google Scholar] [CrossRef]
  72. Sturr, M.G.; Marquis, R.E. Inhibition of proton-translocating ATPases of Streptococcus mutans and Lactobacillus casei by fluoride and aluminum. Arch. Microbiol. 1990, 155, 22–27. [Google Scholar] [CrossRef]
  73. Barbier, O.; Arreola-Mendoza, L.; Del Razo, L.M. Molecular mechanisms of fluoride toxicity. Chem.-Biol. Interact. 2010, 188, 319–333. [Google Scholar] [CrossRef]
  74. Agalakova, N.I.; Gusev, G.P. Molecular mechanisms of cytotoxicity and apoptosis induced by inorganic fluoride. ISRN Cell Biol. 2012, 2012, 403835. [Google Scholar] [CrossRef]
Figure 1. Experimental setup used for ultrasound measurements of FHAp suspension: schematics (a) and image (b) [23].
Figure 1. Experimental setup used for ultrasound measurements of FHAp suspension: schematics (a) and image (b) [23].
Materials 17 03404 g001
Figure 2. Dynamic light scattering (DLS) measurements of the size dispersion of the FHAp, showing data weighted by the number (a) and volume (b) of particles, respectively.
Figure 2. Dynamic light scattering (DLS) measurements of the size dispersion of the FHAp, showing data weighted by the number (a) and volume (b) of particles, respectively.
Materials 17 03404 g002
Figure 3. Time evolution of the recorded signals, from left to right over 10,000 s.
Figure 3. Time evolution of the recorded signals, from left to right over 10,000 s.
Materials 17 03404 g003
Figure 4. Recorded signal amplitudes during the experiment.
Figure 4. Recorded signal amplitudes during the experiment.
Materials 17 03404 g004
Figure 5. Spectral amplitudes of all 2000 signals.
Figure 5. Spectral amplitudes of all 2000 signals.
Materials 17 03404 g005
Figure 6. Time-averaged attenuation for the investigated frequency range.
Figure 6. Time-averaged attenuation for the investigated frequency range.
Materials 17 03404 g006
Figure 7. Relative spectral amplitudes vs. time.
Figure 7. Relative spectral amplitudes vs. time.
Materials 17 03404 g007
Figure 8. The peaks associated with the JCPDS # 09–0432 card (blue lines) and X-ray diffraction pattern of FHAp powder (a); relative shift in (002) in FHAp (b).
Figure 8. The peaks associated with the JCPDS # 09–0432 card (blue lines) and X-ray diffraction pattern of FHAp powder (a); relative shift in (002) in FHAp (b).
Materials 17 03404 g008
Figure 9. FTIR (a) and second derivative (b) spectra obtained for FHAp.
Figure 9. FTIR (a) and second derivative (b) spectra obtained for FHAp.
Materials 17 03404 g009
Figure 10. XPS general spectrum (a) and high-resolution spectra of F1s (b).
Figure 10. XPS general spectrum (a) and high-resolution spectra of F1s (b).
Materials 17 03404 g010
Figure 11. SEM micrographs (a) and EDS spectra (b) of FHAp (Ca10(PO4)6(OH)2−2xF2x, x = 0.05).
Figure 11. SEM micrographs (a) and EDS spectra (b) of FHAp (Ca10(PO4)6(OH)2−2xF2x, x = 0.05).
Materials 17 03404 g011
Figure 12. Percentage hemolysis of sheep red blood cells (RBCs) exposed to different concentrations of FHAp suspensions.
Figure 12. Percentage hemolysis of sheep red blood cells (RBCs) exposed to different concentrations of FHAp suspensions.
Materials 17 03404 g012
Figure 13. Biocompatibility: (a) cell viability (%) determined through MTT assay and (b) LDH release (%) of hFOB 1.19 cells exposed to FHAp suspensions at different time intervals; (p > 0.05) in all comparisons.
Figure 13. Biocompatibility: (a) cell viability (%) determined through MTT assay and (b) LDH release (%) of hFOB 1.19 cells exposed to FHAp suspensions at different time intervals; (p > 0.05) in all comparisons.
Materials 17 03404 g013
Figure 14. Antimicrobial activity of FHAp suspensions on S. aureus at different time intervals.
Figure 14. Antimicrobial activity of FHAp suspensions on S. aureus at different time intervals.
Materials 17 03404 g014
Figure 15. Antimicrobial activity of FHAp suspensions on E. coli at different time intervals.
Figure 15. Antimicrobial activity of FHAp suspensions on E. coli at different time intervals.
Materials 17 03404 g015
Figure 16. Antimicrobial activity of FHAp suspensions on C. albicans at different time intervals.
Figure 16. Antimicrobial activity of FHAp suspensions on C. albicans at different time intervals.
Materials 17 03404 g016
Figure 17. Schematic representation of the antimicrobial mechanism of FHAp suspensions.
Figure 17. Schematic representation of the antimicrobial mechanism of FHAp suspensions.
Materials 17 03404 g017
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ciobanu, C.S.; Predoi, D.; Iconaru, S.L.; Predoi, M.V.; Rokosz, K.; Raaen, S.; Negrila, C.C.; Buton, N.; Ghegoiu, L.; Badea, M.L. Physico-Chemical and Biological Features of Fluorine-Substituted Hydroxyapatite Suspensions. Materials 2024, 17, 3404. https://doi.org/10.3390/ma17143404

AMA Style

Ciobanu CS, Predoi D, Iconaru SL, Predoi MV, Rokosz K, Raaen S, Negrila CC, Buton N, Ghegoiu L, Badea ML. Physico-Chemical and Biological Features of Fluorine-Substituted Hydroxyapatite Suspensions. Materials. 2024; 17(14):3404. https://doi.org/10.3390/ma17143404

Chicago/Turabian Style

Ciobanu, Carmen Steluta, Daniela Predoi, Simona Liliana Iconaru, Mihai Valentin Predoi, Krzysztof Rokosz, Steinar Raaen, Catalin Constantin Negrila, Nicolas Buton, Liliana Ghegoiu, and Monica Luminita Badea. 2024. "Physico-Chemical and Biological Features of Fluorine-Substituted Hydroxyapatite Suspensions" Materials 17, no. 14: 3404. https://doi.org/10.3390/ma17143404

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