Next Article in Journal
First-Principles Calculations to Investigate the Effect of Van der Waals Interactions on the Crystal and Electronic Structures of Tin-Based 0D Hybrid Perovskites
Next Article in Special Issue
Fabrication of a Potential Electrodeposited Nanocomposite for Dental Applications
Previous Article in Journal
Structural and Luminescent Peculiarities of Spark Plasma Sintered Transparent MgAl2O4 Spinel Ceramics Doped with Cerium Ions
Previous Article in Special Issue
Preparation, Microstructural Characterization and Photocatalysis Tests of V5+-Doped TiO2/WO3 Nanocomposites Supported on Electrospun Membranes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evaluating the Sorption Affinity of Low Specific Activity 99Mo on Different Metal Oxide Nanoparticles

1
Department of Chemistry, Biochemistry, and Pharmaceutical Sciences, Faculty of Science, University of Bern, Freiestrasse 3, CH-3012 Bern, Switzerland
2
Radioactive Isotopes and Generators Department, Hot Laboratories Center, Egyptian Atomic Energy Authority, Cairo 13759, Egypt
3
Radiation Protection and Civil Defense Department, Nuclear Research Center, Egyptian Atomic Energy Authority, Cairo 13759, Egypt
4
Egypt Second Research Reactor, Nuclear Research Center, Egyptian Atomic Energy Authority, Cairo 13759, Egypt
*
Author to whom correspondence should be addressed.
Inorganics 2022, 10(10), 154; https://doi.org/10.3390/inorganics10100154
Submission received: 7 August 2022 / Revised: 16 September 2022 / Accepted: 23 September 2022 / Published: 26 September 2022
(This article belongs to the Special Issue New Advances into Nanostructured Oxides)

Abstract

:
99Mo/99mTc generators are mainly produced from 99Mo of high specific activity generated from the fission of 235U. Such a method raises proliferation concerns. Alternative methods suggested the use of low specific activity (LSA) 99Mo to produce 99mTc generators. However, its applicability is limited due to the low adsorptive capacity of conventional adsorbent materials. This study attempts to investigate the effectiveness of some commercial metal oxides nanoparticles as adsorbents for LSA 99Mo. In a batch equilibration system, we studied the influence of solution pH (from 1–8), contact time, initial Mo concentration (from 50–500 mg∙L−1), and temperature (from 298–333 K). Moreover, equilibrium isotherms and thermodynamic parameters (changes in free energy ΔG0, enthalpy change ΔH0, and entropy ΔS0) were evaluated. The results showed that the optimum pH of adsorption ranges between 2 and 4, and that the equilibrium was attained within the first two minutes. In addition, the adsorption data fit well with the Freundlich isotherm model. The thermodynamic parameters prove that the adsorption of molybdate ions is spontaneous. Furthermore, some investigated adsorbents showed maximum adsorption capacity ranging from 40 ± 2 to 73 ± 1 mg Mo∙g−1. Therefore, this work demonstrates that the materials used exhibit rapid adsorption reactions with LSA 99Mo and higher capacity than conventional alumina (2–20 mg Mo∙g−1).

1. Introduction

99Mo/99mTc radioisotope generators have a growing importance in nuclear medicine investigations. They are the primary source of supplying 99mTc radionuclide for diagnostic purposes [1,2,3]. 99mTc is considered the workhorse of all nuclear medicine applications [4,5]. It is involved in more than 80% of all in vivo diagnostic procedures because of its ideal nuclear characteristics, such as the short half-life of 6 h, absence of beta particles, and emission of a mono-energetic photon with low energy at 140 keV [3,6]. Therefore, this leads to less radiation exposure dose to the patients, and it produces a high-quality image for better diagnosis aspects. Furthermore, its unique labeling chemistry allows the use of a wide range of 99mTc-labelled compounds to visualize different body organs [7,8]. For instance, 99mTc-DTPA and 99mTc-MAG3 are used to monitor renal functions [9]. In addition, 99mTc-tetrofosmin, 99mTc-sestamibi, and 99mTc-teboroxime are utilized for the diagnosis of cardiac disease [10]. Moreover, 99mTc-lidofenin is applied for liver diagnostics [11]. Furthermore, 99mTc-medronate, 99mTc-propyleneamineoxime, and 99mTc-MDP (methylene diphosphonate) are involved in skeletal imaging, cerebral perfusion, and diagnosis of bone metastases, respectively [12,13,14,15].
Among the developed 99Mo/99mTc generators, the chromatographic column type is the most widely used system [3,16]. This system is based on adsorbing 99Mo on a column filled with a suitable material from which 99mTcO4 can be easily eluted while 99Mo remains adsorbed [3]. The differences between these generators include the column material and the origin of the parent, 99Mo. The main practical difficulties linked to the preparation of 99Mo/99mTc generators are the low sorption capacity of the bulk conventional inorganic sorbents usually used. These sorbents have low sorption capacity (2–20 mg Mo/g) due to the low availability of active sites and relatively limited surface area [3]. Consequently, such sorbents require a parent of high specific activity to prepare a useful generator of a proper radioactivity level. A high specific activity parent can be produced from the fission of 235U. Fission-produced 99Mo faces some critical difficulties. For example, sophisticated infrastructures and well-qualified personnel are needed to separate and purify 99Mo from the irradiated 235U target and other fission products. In addition, a considerable level of radioactive waste is generated during the manufacturing process, which increases the cost of production [17,18]. Alternatively, research studies focused on developing clinical-grade chromatographic 99Mo/99mTc generators based on 99Mo of low specific activity (LSA) [3,19,20]. However, this proposal demands using high-capacity sorbents to compensate for the LSA 99Mo and make it more reliable from the economic point of view [17,21].
The use of advanced nanomaterials has generated a growing interest in developing diagnostic 99mTc generators [3]. Nawar and Türler [3] highlighted several nanomaterial adsorbents that have been developed for 99Mo/99mTc generator application. This class of sorbents possesses appreciable adsorption capacity and unique performance [20]. In this regard, the utilization of advanced commercial metal-oxide nanoparticles is an exciting idea due to their improved properties. In contrast to traditional sorbents, these nano-adsorbents have large surface-to-volume ratios, enhanced porosity, improved surface reactivity, and significant radiation resistance and chemical stability [21,22]. Therefore, they show high adsorption efficiency and selectivity [23].
In this study, we intend to evaluate the sorption efficiency of some commercially available nano-metal oxides towards LSA 99Mo. To achieve this goal, we investigated the adsorption behavior of the selected materials for LSA 99Mo under different experimental conditions. These conditions include the pH, initial concentration of molybdate ions, contact time, and temperature. In addition, to better understand their sorption behavior, the sorption kinetics, equilibrium isotherms, and thermodynamic behavior were evaluated.

2. Results and Discussion

2.1. Effect of Solution pH

The solution pH has a profound impact on the efficiency of the adsorption process. The influence of pH can be clarified by understanding its role in varying the ionic state of the functional groups on the adsorbent surface. Moreover, it affects the ionization and/or the dissociation of the studied ions [24]. In this context, a batch equilibration experiment was conducted at a pH range from 1 to 8 to determine the optimum pH value that shows the maximum 99Mo retention on each adsorbent. Figure 1a depicts the distribution coefficients (Kd) of CA-99Mo at different pH values. The data presented in this figure show that higher Kd values are observed at pH values (2–4). Beyond this region, the Kd values decrease with increasing the solution pH, which agrees with previously published studies [25].
Since the adsorbents are metal oxides, they might have similar surface chemistry. Moreover, since the adsorption process depends mainly on the aqueous phase’s pH values and the adsorbent material’s surface characteristics, we investigated the isoelectric point (pHIEP) of each adsorbent (Table 1). The pHIEP measurements help to clarify the sorption mechanism. The sorbent surface carries a positive charge at pH < pHIEP, zero charge at pH~pHIEP, and is negatively charged at pH > pHIEP. Consequently, there is a change in the pHIEP of the sorbent with the pH of an aqueous solution. Nawar et al. [22] reported that this behavior might occur because amphoteric hydroxyl groups cover the adsorbent surface. Hence, based on the pH of the medium, these groups develop different reactions in different pH media, resulting in positive or negative charges appearing on the adsorbent surface. Herein, at pH < pHIEP, they are protonated, and the surface develops a positive charge as follows:
Adsorbent OH Surface +   H solution +     Adsorbent OH 2 +
The data presented in Figure 1a can be interpreted by considering the speciation diagram of molybdenum shown in Figure 1b [22]. The speciation data are generated using the PHREEQC software (version 3) to determine the predominant Mo species at different pHs for the following conditions: C0 = 50 mg∙L−1 at 298   ±   1 K and using the built-in database of stability constants [22]. At acidic medium, the molybdate anionic species exist and polymerize, increasing the molybdenum content per unit charge as follows:
7   M 99 oO 4 2 + 8   H +   M 99 o 7 O 24 6 + 4   H 2 O  
Consequently, this results in favorable interactions between negatively charged molybdenum polyanions and positively charged adsorbents surfaces [26]. At higher pH values, the speciation shifts to less negatively charged Mo species, and the density of hydroxyl groups (OH) increases in solution. These hydroxyl anions compete with less negatively charged molybdenum anions to retain the available active sites on adsorbents surfaces, explaining the low Kd distribution values at higher pH values [22,27].
Moreover, based on the isoelectric point (pHIEP) of each sorbent material (Table 1) and the measured final solution pH (Figure 1c), it can be observed that Kd values start to decrease when the final solution pH exceeds the sorbent’s pHIEP, which can be attributed to the expected change in the surface charge of the sorbent material. As previousely mentioned, at solution pH values above the pHIEP, the sorbent surface becomes predominately negatively charged. As a result, repulsion between the negatively charged sorbent surface and the negatively charged molybdenum polyanions takes place, leading to the observed decrease in Kd values [22,28].
It can also be observed that both silicon oxide and aluminosilicate nanoparticles possess small particle sizes (5–20 and 4.5–4.8 nm) and high surface area (590–690 and 900–1100 m2∙g−1), respectively (Table 1). However, both adsorbents show a weak affinity for Mo species. This behavior may be attributed to their poor stability with increasing pH values. At high pH values, the dissolution of silica occurs, resulting in the formation of monomeric ortho-silicic acid (H4SiO4), which can be explained due to the presence of more hydroxyl groups. These hydroxyl groups are chemisorbed on the adsorbent surface, which increases the number of coordination bonds around the silicon atom to more than four bonds. Consequently, it may lead to Si-O bond rupture, and the silicon atom dissolves as Si(OH)4 and ortho-silicic acid [29].
According to the obtained results, we selected six adsorbents that showed high distribution coefficient values towards CA-99Mo for the subsequent investigations. These adsorbents are CeO2-544841, ZrO2-544760, TiO2-637254, Al2TiO5-634143, CeO2/ZrO2-634174, and CeO2-700290.

2.2. Adsorption Isotherm

Equilibrium isotherms are essential in describing the adsorption mechanisms for the interaction of Mo(VI) ions with the surfaces of the investigated metal oxides NPs. These mechanisms describe the adsorption process successfully. Here, we investigated equilibrium data obtained for adsorption of CA-99Mo on CeO2-544841, ZrO2-544760, TiO2-637254, Al2TiO5-634143, CeO2/ZrO2-634174, and CeO2-700290 with various isotherm models to find out which one is the most suitable for describing the obtained adsorption equilibrium data.

2.2.1. Freundlich Isotherm

Many studies have utilized the Freundlich adsorption isotherm model proposed as a general power equation used to describe the adsorption of radionuclides in a large number of studies [30,31,32]. The Freundlich isotherm has the form shown as follows:
q e = K f C e 1 n f
where qe (mg∙g−1) is the concentration of CA-99Mo adsorbed and Ce (mg∙L−1) is the concentration of Mo remaining in the solution. Kf (mg1−nLn∙g−1) and nf (dimensionless) are constants unique to each combination of adsorbent and adsorbate.

2.2.2. Langmuir Isotherm

Langmuir (1918) developed an equation to describe the adsorption of gases on a solid surface that was subsequently adapted to describe the adsorption of solutes onto solids in aqueous solutions [31,33,34], as shown in Equation (4):
q e = n L K L C e 1 + K L C e
where qe (mg∙g−1) is the total concentration of solute adsorbed, KL (L∙mg−1) is an equilibrium constant, and nL (mg∙g−1) is the adsorption capacity.
Figure 2 presents the experimental adsorption equilibrium data obtained for Mo ions on the investigated metal oxide adsorbents as a plot of adsorption equilibrium capacity (qe) against initial concentration (C0). It is observed that there is an increase in the amount of Mo ions taken up with the increase in the initial metal ion concentration. This increase in the adsorbate uptake can be explained by the driving force for mass transfer [34].
The non-linear forms of both isotherm models were applied to the measured adsorption data (Ce versus qe), and the data were displayed in Figure 3. Adsorption parameters were optimized using the add-ins “Solver” function in Microsoft Excel. Table 2 gives the Freundlich parameters (Kf and nf), Langmuir parameters (KL and nL), and the goodness of fit of the model lines to the experimental data (R2). Based on the regression coefficient values reported in Table 2, it is observed that good to excellent correlations between the experimental results and the fitted data of the Freundlich isotherm model were obtained for all the investigated sorbents. In contrast, the Langmuir model failed to fit any equilibrium sorption isotherm of the CA-99Mo on all tested adsorbents; lower R2 values were obtained.
These findings suggest that CA-99Mo adsorption on metal oxide nanomaterials under investigation mainly occurred through multilayer adsorption at heterogeneous surfaces [31,35]. The Freundlich adsorption constant (nf) is usually used as a measure of adsorption intensity as follows; (i) nf < 1 indicates that adsorption takes place via a chemical process, (ii) nf = 1 shows linear adsorption, (iii) while nf > 1 indicates physisorption [35]. The nf values displayed in Table 2 were higher than 1, indicating that CA-99Mo adsorption on the materials used in this study was physisorption and favorable under the investigated conditions. Furthermore, the closer the 1/n value to 0 than unity (ranging from 0.10 to 0.25), the more heterogeneous the surface is, implying a broad distribution of adsorption sites on the adsorbent surface [32,33].

2.3. Thermodynamic Studies

We determined the amount of CA-99Mo adsorbed on the surface of the materials investigated in the current study as a function of temperature (T) using adsorption thermodynamic parameters. These parameters include the Gibbs free energy ΔG0 (kJ∙mol−1), the standard enthalpy change ΔH0 (kJ∙mol−1), and the standard entropy change ΔS0 (J∙mol−1∙K−1). They were investigated at different temperatures (298, 313, 323, and 333 K) using Equations (5) and (6) [34,36,37] and are tabulated in Table 3:
G 0 = RTlnK d
ln K d = S 0 R H 0 RT  
where R is the universal gas constant (8.314 J∙mol−1∙K−1), T is the absolute temperature (K), and Kd (mL∙g−1) is the distribution coefficient.
Figure 4 shows linear plots of ln Kd versus (1/T). The calculated ΔG0 values at each temperature for all nano-adsorbents are ΔG0 < 0, which implies that the Mo(VI) adsorption process on the surfaces of all adsorbents is spontaneous and the reaction is feasible. Likewise, ΔG0 values decrease with increasing temperature, indicating that the degree of spontaneity can be enhanced by increasing the temperature. Furthermore, the adsorption process is physisorption (−20 < ΔG0 < 0) [38]. The positive values of ΔS0S0 > 0) report random adsorption reactions of CA-99Mo at all adsorbents surfaces. The values of ΔH0 are positive (ΔH0 > 0) for both TiO2-637254 and CeO2/ZrO2-634174, implying that CA-99Mo adsorption at their surfaces is endothermic [39]. While for CeO2-544841, ZrO2-544760, Al2TiO5-634143, and CeO2-700290, the change in enthalpy (ΔH0) is negative (ΔH0 < 0), indicating that the adsorption of CA-99Mo at their surfaces is exothermic [38,40].

2.4. Determining the Maximum Sorption Capacity

In order to evaluate the maximum sorption capacity of each adsorbent, the equilibrations of CA-99Mo with each adsorbent were performed separately. Batch equilibrations were repeated until no further 99Mo(IV) uptake was observed, and the adsorbents became fully saturated with 99Mo. After each equilibration, 1 mL aliquot was decanted, centrifuged, and counted. Ultimately, the maximum sorption capacity q max for each material was calculated by applying the following equation:
q max = U % 100 × C o × V m mg · g 1
where U% is the uptake percent of CA-99Mo, C0(mg∙L−1) is the starting Mo(IV) concentration, V (L) is the liquid phase volume, and m   (g) is the adsorbent weight. Figure 5 shows CA-99Mo maximum sorption capacity on different studied metal oxides NPs. It can be concluded that the studied metal oxide NPs show better sorption capacity than conventional alumina currently used in 99Mo/99mTc generators. Nonetheless, the obtained capacities are insufficient for developing a clinical-grade 99mTc generator based on LSA 99Mo.

2.5. Effect of Contact Time

The effect of contact time on the uptake percent of CA-99Mo was monitored for an initial Mo(IV) concentration of 50 mg∙L−1 (pH~3), using an adsorbent dose of 200 mg. The reaction temperature was adjusted to 298 ± 1 K. The results are shown in Figure 6. The results show that the Mo uptake sharply increased at the beginning of the adsorption process and reached a constant value (a plateau value) in the first two minutes. This behavior indicates a rapid and almost instantaneous removal of CA-99Mo from the solution, and a dynamic equilibrium is established under the given experimental conditions. In order to design an effective adsorption process, determining the kinetic parameters is crucial. The kinetic data shown in Figure 6 revealed that the equilibrium for adsorption of Mo on metal oxide nano-adsorbents is already reached at the very beginning of the adsorption process. Consequently, using the current methodology, such data cannot be modeled with adsorption kinetic models.

3. Materials and Methods

3.1. Materials

All chemicals are of analytical grade purity (A. R. grade) and were used without further purification. Milli-Q water was used for the preparation of solutions and washings. Sodium hydroxide and nitric acid were purchased from Merck, Darmstadt, Germany. The metal oxide nanomaterials were purchased from different suppliers (Table 1).
99Mo radiotracer solution was obtained by eluting a 40 GBq fission 99Mo alumina-based 99Mo/99mTc generator (Pertector, manufactured by National Centre for Nuclear Research, POLATOM, Otwock, Poland) with 5 mL of 1 M NaOH solution after ~7 d from the calibration date. The total 99Mo radioactivity was measured with a Capintec Radioisotopes Calibrator (model CRC-55tR Capintec, Inc., Florham Park, NJ, USA). The 99Mo eluate solution was passed through a 0.45 micro-Millipore filter to retain alumina particles. Then, the 99Mo solution was treated with nitric acid to attain the desired pH value.

3.2. Batch Equilibrium Studies

A batch equilibration experiment was conducted to investigate the adsorption behavior of carrier-added (CA) 99Mo (Mo(IV) treated with 99Mo) on several commercial metal oxide nanoparticles (NPs) under different conditions. These conditions included the influence of pH, contact time, reaction temperature, and initial adsorbate concentration. In a series of clean glass bottles, we added 200 mg of each adsorbent to 20 mL of 99Mo(IV) solution of a given concentration and pH value. Subsequently, the mixtures were shaken in a thermostatic shaker water bath (Julabo GmbH, Seelbach, Germany) at 298 ± 1 K for 24 h. Eventually, the supernatant solution was collected, centrifuged, and 1 mL was separated for radiometric measurements. For all radiometric identifications and γ-spectrometry, we used a multichannel analyzer (MCA) of Inspector 2000 model, Canberra Series, Mirion Technologies, Inc., Meriden, CT, USA, coupled with a high-purity germanium coaxial detector (HPGe). All samples have fixed geometry and were counted at a low dead time (<2%). The measurements were done by using an appropriate gamma-ray peak of 740 keV.

3.2.1. Distribution Ratio (Kd)

The distribution coefficient (Kd) values of CA-99Mo were investigated at a wide range of pH (from 1–8). For adjusting the desired pH value of the solutions, few drops of 0.5 M nitric acid or 0.5 M sodium hydroxide were added. The pH values of the solutions were measured before and after reaching the equilibrium state. pH values were determined using a pH-meter with a microprocessor (Mettler Toledo, Seven Compact S210 model, Greifensee, Switzerland).

3.2.2. Adsorption Isotherm

In order to determine the sorption isotherms, we used different initial molybdate ion concentrations from 50 to 500 mg∙L−1 while keeping the adsorbent amount constant. Moreover, the solution pH, equilibrium time, and reaction temperature were kept at pH~3, 24 h, and 298 ± 1 K, respectively. In addition, the equilibrium adsorption capacity (qe) was calculated. Finally, we used the obtained results to determine the sorption isotherm model.

3.2.3. Thermodynamic Studies

The reaction temperature effect on the uptake of carrier-added 99Mo was studied at four different reaction temperatures (298, 313, 323, and 333 K). At each temperature, we added 20 mL of CA-99Mo solution (pH 3) in contact with 200 mg of the adsorbent material for 24 h. From the resulting data, we calculated different thermodynamic parameters, namely the standard enthalpy change (ΔH0), standard entropy change (ΔS0), and Gibbs free energy change (ΔG0).

3.2.4. Effect of Contact Time

In order to investigate the 99Mo adsorption rate on the studied metal oxides NPs, we monitored the progress of the uptake capacity of 99MoO42– ions (50 mg∙L−1 and pH~3) at different time slots. The adsorption of CA-99Mo was followed with time until the equilibrium was established. Finally, we calculated the 99Mo capacity ( q t ) in mg∙g−1 at each time (t).

3.3. Calculations

The adsorption data of CA-99Mo include uptake percent (U%), distribution coefficient (Kd), equilibrium capacity (qe), and equilibrium concentration (Ce). These data were calculated according to the following equations:
U % = ( A i A f ) A i × 100
q e = U % 100 × C 0 × V m mg · g 1  
C e = A i   A i × U % 100 mg · L 1
K d   = A i   A f A i × V m mL · g 1
where Ai and Af are the initial and final 99Mo radioactivity in counts/min. C 0 (mg∙L−1) is the initial concentration of CA-99Mo, V (L) and V (mL) represent the volume of liquid phases, and m (g) is the weight of the solid phase.

4. Summary and Conclusions

The main objective of this study was to evaluate the adsorption affinity of different commercial metal oxides NPs purchased from different suppliers towards LSA 99Mo. All experiments were conducted at static equilibrium conditions. We studied the distribution ratio of CA-99Mo in a pH range of 1 to 8. The optimum adsorption pH was found to be in the range of pH 2 to 4. In addition, the Freundlich isotherm model fitted the experimental data of the CA-99Mo on all adsorbent materials investigated in this study. Moreover, we determined the values of enthalpy change (ΔH0), entropy change (ΔS0), and free energy change (ΔG0) at the different reaction temperatures. Furthermore, the maximum adsorption capacities were evaluated, and the best adsorbents showed a capacity of 40 ± 2 to 73 ± 1 mg Mo∙g−1. Summing up the results, it can be concluded that the adsorption behavior of the materials investigated depends on the solution pH, contact time, initial metal ion concentration, and temperature. Furthermore, the investigated materials showed higher static sorption capacities than conventional alumina (2–20 mg Mo∙g−1). Nonetheless, they are not suitable to build a useful 99Mo/99mTc generator using LAS 99Mo for radiopharmaceutical applications. Since the available specific activity of LAS 99Mo is 2.5–5 Ci/g Mo, approximately 20–25 g of each material would be required to prepare a 99mTc generator of 37 GBq (1 Ci). Using such a massive amount of sorbent material per generator would deteriorate the elution performance and the radioactive concentration of the produced 99mTc.

Author Contributions

Conceptualization, M.F.N.; methodology, M.F.N. and A.F.E.-D.; software, M.F.N., A.F.E.-D. and A.A.; validation, M.F.N., A.F.E.-D., A.A., M.A.S. and A.T.; formal analysis, M.F.N., A.F.E.-D. and A.A.; investigation, M.F.N.; data curation, M.F.N., A.F.E.-D. and A.T.; writing—original draft preparation, M.F.N.; writing—review and editing, A.F.E.-D., A.A., M.A.S. and A.T.; visualization, M.F.N. and A.F.E.-D.; supervision, A.T.; project administration, M.F.N. and A.T.; funding acquisition, M.F.N. All authors have read and agreed to the published version of the manuscript.

Funding

The research was funded by the Swiss National Science Foundation (grant number CRSII5_180352). Mohamed F. Nawar gratefully acknowledges the funding support of the Swiss Government Excellence fellowships program (fellowship No: 2017.1028).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The authors would like to express their sincere thanks to Marcel Langensand, Managing Director, Medeo AG, CH-5040 Schöftland, Switzerland, for his valuable support in supplying 99Mo/99mTc generators to conduct this research study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cutler, C.S. Supply of Mo-99: Focus on U.S. supply. Trends in Radiopharmaceuticals (ISTR-2019). In Proceedings of the International Symposium, Vienna, Austria, 28 October–1 November 2019; International Atomic Energy Agency (IAEA): Vienna, Austria, 2020. [Google Scholar]
  2. Munir, M.; Sriyono; Abidin; Sarmini, E.; Saptiama, I.; Kadarisman; Marlina. Development of mesoporous γ-alumina from aluminium foil waste for 99Mo/99mTc generator. J. Radioanal. Nucl. Chem. Artic. 2020, 326, 87–96. [Google Scholar] [CrossRef]
  3. Nawar, M.F.; Türler, A. New strategies for a sustainable 99mTc supply to meet increasing medical demands: Promising solutions for current problems. Front. Chem. 2022, 10, 926258. [Google Scholar] [CrossRef] [PubMed]
  4. Capogni, M.; Pietropaolo, A.; Quintieri, L.; Angelone, M.; Boschi, A.; Capone, M.; Cherubini, N.; De Felice, P.; Dodaro, A.; Duatti, A.; et al. 14 MeV Neutrons for 99Mo/99mTc Production: Experiments, Simulations and Perspectives. Molecules 2018, 23, 1872. [Google Scholar] [CrossRef] [PubMed]
  5. Duatti, A. Review on 99mTc radiopharmaceuticals with emphasis on new advancements. Nucl. Med. Biol. 2021, 92, 202–216. [Google Scholar] [CrossRef] [PubMed]
  6. Boschi, A.; Uccelli, L.; Martini, P. A Picture of Modern Tc-99m Radiopharmaceuticals: Production, Chemistry, and Applications in Molecular Imaging. Appl. Sci. 2019, 9, 2526. [Google Scholar] [CrossRef]
  7. Mohan, A.-M.; Beindorff, N.; Brenner, W. Nuclear Medicine Imaging Procedures in Oncology. Metastasis 2021, 2294, 297–323. [Google Scholar] [CrossRef]
  8. Kniess, T.; Laube, M.; Wüst, F.; Pietzsch, J. Technetium-99m based small molecule radiopharmaceuticals and radiotracers targeting inflammation and infection. Dalton Trans. 2017, 46, 14435–14451. [Google Scholar] [CrossRef] [PubMed]
  9. Momin, M.; Abdullah, M.; Reza, M. Comparison of relative renal functions calculated with 99m Tc-DTPA and 99m Tc-DMSA for kidney patients of wide age ranges. Phys. Med. 2018, 45, 99–105. [Google Scholar] [CrossRef]
  10. Fang, W.; Liu, S. New 99mTc Radiotracers for Myocardial Perfusion Imaging by SPECT. Curr. Radiopharm. 2019, 12, 171–186. [Google Scholar] [CrossRef]
  11. Marlina, M.; Lestari, E.; Abidin, A.; Hambali, H.; Saptiama, I.; Febriana, S.; Kadarisman, K.; Awaludin, R.; Tanase, M.; Nishikata, K.; et al. Molybdenum-99 (99Mo) Adsorption Profile of Zirconia-Based Materials for 99Mo/99mTc Generator Application. At. Indones. 2020, 46, 91–97. [Google Scholar] [CrossRef]
  12. Papagiannopoulou, D. Technetium-99m radiochemistry for pharmaceutical applications. J. Label. Compd. Radiopharm. 2017, 60, 502–520. [Google Scholar] [CrossRef] [PubMed]
  13. Sharma, S.; Jain, S.; Baldi, A.; Singh, R.K.; Sharma, R.K. Intricacies in the approval of radiopharmaceuticals-regulatory per-spectives and the way forward. Curr. Sci. 2019, 116, 47–55. [Google Scholar] [CrossRef]
  14. Lassen, A.; Stokely, E.; Vorstrup, S.; Goldman, T.; Henriksen, J.H. Neuro-SPECT: On the development and function of brain emission tomography in the Copenhagen area. Clin. Physiol. Funct. Imaging 2021, 41, 10–24. [Google Scholar] [CrossRef]
  15. Chen, B.; Wei, P.; Macapinlac, H.A.; Lu, Y. Comparison of 18F-Fluciclovine PET/CT and 99mTc-MDP bone scan in detection of bone metastasis in prostate cancer. Nucl. Med. Commun. 2019, 40, 940–946. [Google Scholar] [CrossRef] [PubMed]
  16. Hasan, S.; Prelas, M.A. Molybdenum-99 production pathways and the sorbents for 99Mo/99mTc generator systems using (n, γ) 99Mo: A review. SN Appl. Sci. 2020, 2, 1782. [Google Scholar] [CrossRef]
  17. Marlina; Ridwan, M.; Abdullah, I.; Yulizar, Y. Recent progress and future challenge of high-capacity adsorbent for non-fission molybdenum-99 (99Mo) in application of 99Mo/99mTc generator. AIP Conf. Proc. 2021, 2346, 030003. [Google Scholar] [CrossRef]
  18. Munir, M.; Herlina; Sriyono; Sarmini, E.; Abidin; Lubis, H.; Marlina. Influence of GA Siwabessy Reactor Irradiation Period on The Molybdenum-99 (99Mo) Production by Neutron Activation of Natural Molybdenum to Produce Technetium-99m (99mTc). J. Phys. Conf. Ser. 2019, 1204, 012021. [Google Scholar] [CrossRef]
  19. Nawar, M.F.; El-Daoushy, A.F.; Ashry, A.; Türler, A. Developing a Chromatographic 99mTc Generator Based on Mesoporous Alumina for Industrial Radiotracer Applications: A Potential New Generation Sorbent for Using Low-Specific-Activity 99Mo. Molecules 2022, 27, 5667. [Google Scholar] [CrossRef]
  20. Nawar, M.F.; Türler, A. Development of New Generation of 99Mo/99mTc Radioisotope Generators to Meet the Continuing Clinical Demands. In Proceedings of the 2nd International Conference on Radioanalytical and Nuclear Chemistry (RANC 2019), Budapest, Hungary, 5–10 May 2019. [Google Scholar]
  21. Moreno-Gil, N.; Badillo-Almaraz, V.E.; Pérez-Hernández, R.; López-Reyes, C.; Issac-Olivé, K. Comparison of the sorption behavior of 99Mo by Ti-, Si-, Ti-Si-xerogels and commercial sorbents. J. Radioanal. Nucl. Chem. Artic. 2021, 328, 679–690. [Google Scholar] [CrossRef]
  22. Nawar, M.F.; El-Daoushy, A.F.; Madkour, M.; Türler, A. Sorption Profile of Low Specific Activity 99Mo on Nanoceria-Based Sorbents for the Development of 99mTc Generators: Kinetics, Equilibrium, and Thermodynamic Studies. Nanomaterials 2022, 12, 1587. [Google Scholar] [CrossRef]
  23. Sakr, T.M.; Nawar, M.F.; Fasih, T.; El-Bayoumy, S.; Abd El-Rehim, H.A. Nano-technology contributions towards the development of high performance radioisotope generators: The future promise to meet the continuing clinical demand. Appl. Radiat. Isot. 2017, 129, 67–75. [Google Scholar] [CrossRef] [PubMed]
  24. Monir, T.; El-Din, A.S.; El-Nadi, Y.; Ali, A. A novel ionic liquid-impregnated chitosan application for separation and purification of fission 99Mo from alkaline solution. Radiochim. Acta 2020, 108, 649–659. [Google Scholar] [CrossRef]
  25. Van Tran, T.; Nguyen, V.H.; Nong, L.X.; Nguyen, H.-T.T.; Nguyen, D.T.C.; Nguyen, H.T.T.; Nguyen, T.D. Hexagonal Fe-based MIL-88B nanocrystals with NH2 functional groups accelerating oxytetracycline capture via hydrogen bonding. Surf. Interfaces 2020, 20, 100605. [Google Scholar] [CrossRef]
  26. Dash, A.; Chakravarty, R.; Ram, R.; Pillai, K.; Yadav, Y.Y.; Wagh, D.; Verma, R.; Biswas, S.; Venkatesh, M. Development of a 99Mo/99mTc generator using alumina microspheres for industrial radiotracer applications. Appl. Radiat. Isot. 2012, 70, 51–58. [Google Scholar] [CrossRef] [PubMed]
  27. Moret, J.; Alkemade, J.; Upcraft, T.; Oehlke, E.; Wolterbeek, H.; van Ommen, J.; Denkova, A. The application of atomic layer deposition in the production of sorbents for 99Mo/99mTc generator. Appl. Radiat. Isot. 2020, 164, 109266. [Google Scholar] [CrossRef]
  28. Reedijk, J.; Poeppelmeier, K. Comprehensive Inorganic Chemistry II: From Elements to Applications, 2nd ed.; Elsevier Ltd.: Singapore, 2013; pp. 1–7196. [Google Scholar] [CrossRef]
  29. Bernauer, U.; Bodin, L.; Chaudhry, Q.; Coenraads, P.J.; Dusinska, M.; Gaffet, E.; Panteri, E.; Rousselle, C.; Stepnik, M.; Wijnhoven, S.; et al. SCCS Opinion on Solubility of Synthetic Amorphous Silica (SAS)-SCCS/1606/19; Publications Office of the European Union: Luxembourg, 2020; Available online: https://hal.archives-ouvertes.fr/hal-03115473 (accessed on 23 September 2022).
  30. Ashry, A.; Bailey, E.H.; Chenery, S.R.N.; Young, S.D. Kinetic study of time-dependent fixation of U(VI) on biochar. J. Hazard Mater. 2016, 320, 55–66. [Google Scholar] [CrossRef]
  31. Ashry, A. Adsorption and Time Dependent Fixation of Uranium (VI) in Synthetic and Natural Matrices. Ph.D. Thesis, University of Nottingham, Nottingham, UK, 2017. [Google Scholar]
  32. Mahmoud, M.R.; Soliman, M.A.; Allan, K.F. Removal of Thoron and Arsenazo III from radioactive liquid waste by sorption onto cetyltrimethylammonium-functionalized polyacrylonitrile. J. Radioanal. Nucl. Chem. Artic. 2014, 300, 1195–1207. [Google Scholar] [CrossRef]
  33. Mahmoud, M.R.; Sharaf El-deen, G.E.; Soliman, M.A. Surfactant-impregnated activated carbon for enhanced adsorptive re-moval of Ce(IV) radionuclides from aqueous solutions. Ann. Nucl. Energy 2014, 72, 134–144. [Google Scholar] [CrossRef]
  34. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef] [Green Version]
  35. Akbas, Y.A.; Yusan, S.; Sert, S.; Aytas, S. Sorption of Ce(III) on magnetic/olive pomace nanocomposite: Isotherm, kinetic and thermodynamic studies. Environ. Sci. Pollut. Res. 2021, 28, 56782–56794. [Google Scholar] [CrossRef]
  36. Alothman, Z.A.; Naushad, M.; Ali, R. Kinetic, equilibrium isotherm and thermodynamic studies of Cr(VI) adsorption onto low-cost adsorbent developed from peanut shell activated with phosphoric acid. Environ. Sci. Pollut. Res. 2013, 20, 3351–3365. [Google Scholar] [CrossRef] [PubMed]
  37. Zhang, J.; Deng, R.j.; Ren, B.Z.; Hou, B.; Hursthouse, A. Preparation of a novel Fe3O4/HCO composite adsorbent and the mechanism for the removal of antimony (III) from aqueous solution. Sci. Rep. 2019, 9, 13021. [Google Scholar] [CrossRef] [PubMed]
  38. Zheng, H.; Wang, Y.; Zheng, Y.; Zhang, H.; Liang, S.; Long, M. Equilibrium, kinetic and thermodynamic studies on the sorption of 4-hydroxyphenol on Cr-bentonite. Chem. Eng. J. 2008, 143, 117–123. [Google Scholar] [CrossRef]
  39. Naiya, T.K.; Bhattacharya, A.K.; Mandal, S.; Das, S.K. The sorption of lead(II) ions on rice husk ash. J. Hazard. Mater. 2009, 163, 1254–1264. [Google Scholar] [CrossRef] [PubMed]
  40. Hamdaoui, O.; Naffrechoux, E. Modeling of adsorption isotherms of phenol and chlorophenols onto granular activated carbon. Part I. Two-parameter models and equations allowing determination of thermodynamic parameters. J. Hazard Mater. 2007, 147, 381–394. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Effect of initial pH on (a) the distribution coefficients (Kd) of CA-99Mo on different metal oxides NPs (C0 = 50 mg∙L−1, V/m = 100 mL∙g−1, and temperature = 298 ± 1 K), (b) Speciation of molybdenum [22], and (c) variation of the final pH values.
Figure 1. Effect of initial pH on (a) the distribution coefficients (Kd) of CA-99Mo on different metal oxides NPs (C0 = 50 mg∙L−1, V/m = 100 mL∙g−1, and temperature = 298 ± 1 K), (b) Speciation of molybdenum [22], and (c) variation of the final pH values.
Inorganics 10 00154 g001
Figure 2. The influence of initial molybdate concentration on the equilibrium sorption capacity (qe) of CA-99Mo on different metal oxides NPs (pH = 3, V/m =100 mL∙g−1, t = 24 h, and temperature = 298 ± 1 K).
Figure 2. The influence of initial molybdate concentration on the equilibrium sorption capacity (qe) of CA-99Mo on different metal oxides NPs (pH = 3, V/m =100 mL∙g−1, t = 24 h, and temperature = 298 ± 1 K).
Inorganics 10 00154 g002
Figure 3. Adsorption isotherms: (a) Langmuir and (b) Freundlich of CA-99Mo on different metal oxides NPs.
Figure 3. Adsorption isotherms: (a) Langmuir and (b) Freundlich of CA-99Mo on different metal oxides NPs.
Inorganics 10 00154 g003
Figure 4. Van’t Hoff plot for the sorption of CA-99Mo on different metal oxides NPs.
Figure 4. Van’t Hoff plot for the sorption of CA-99Mo on different metal oxides NPs.
Inorganics 10 00154 g004
Figure 5. The maximum sorption capacity of different metal oxide NPs for CA-99Mo.
Figure 5. The maximum sorption capacity of different metal oxide NPs for CA-99Mo.
Inorganics 10 00154 g005
Figure 6. Effect of contact time on CA-99Mo uptake on different metal oxide NPs (C0 = 50 mg∙L−1, pH = 3, V/m = 100 mL∙g−1, and temperature = 298 ± 1 K).
Figure 6. Effect of contact time on CA-99Mo uptake on different metal oxide NPs (C0 = 50 mg∙L−1, pH = 3, V/m = 100 mL∙g−1, and temperature = 298 ± 1 K).
Inorganics 10 00154 g006
Table 1. Description of the analyzed commercial metal oxides NPs *.
Table 1. Description of the analyzed commercial metal oxides NPs *.
No.NameDescriptionParticle Size,
(nm)
Surface Area,
(m2∙g−1)
Isoelectric Point
(pHIEP)
CeO2-544841 Cerium oxide-SA-544841Molecular formula: CeO2
Molecular weight: 172.11
Density: 7.13 g∙mL−1 at 298 K
<25N.A5
ZrO2-544760 Zirconium oxide-SA-544760Molecular formula: ZrO2
Molecular weight: 123.22
Density: 5.89 g∙mL−1 at 298 K
<100≥256.1
TiO2-637254 Titanium oxide-SA-637254Molecular formula: TiO2
Molecular weight: 79.87
Density: 3.9 g∙mL−1 at 298 K
<2545–556.6
SnO2-549657 Tin oxide-SA-549657Molecular formula: SnO2
Molecular weight: 150.71
Density: 6.95 g∙mL−1 at 298 K
≤10020.13.8
SiO2-637246 Silicon oxide-SA-637246Molecular formula: SiO2
Molecular weight: 60.08
Density: 2.2–2.6 g∙mL−1 at 298 K
5–20590–6902.5
AlCeO3-637866 Cerium aluminium oxide-SA-637866Molecular formula: AlCeO3
Molecular weight: 215.1
≤80N.A4.8
Al2TiO5-634143 Aluminium titanium oxide-SA-634143Molecular formula: Al2TiO5
Molecular weight: 181.83
<25N.A6.4
Al2TiO5-14484 Aluminium titanium oxide-AA-14484Molecular formula: Al2TiO5
Molecular weight: 181.86
100 meshN.A6.5
CeO2/ZrO2-634174 Cerium zirconium oxide-SA-634174Molecular formula: (CeO2)·(ZrO2)
Molecular weight: 295.34
Density: 6.61 g∙mL−1 at 298 K
<50N.A6.7
SiO2/Al2O3-643653 Aluminosilicate-SA-643653Molecular formula: (SiO2)x(Al2O3)y
pore volume:
0.8–1.1 cm3∙g−1
mesostructured, pore size: 2–4 nm
4.5–4.8900–11006
CeO2-700290 Cerium oxide-SA-700290Molecular formula: CeO2
Molecular weight: 172.11
Density: 7.13 g∙mL−1 at 298 K
<50304.5
* The information was provided by the supplier. Only the isoelectric point data were determined experimentally. Abbreviations: AA: Alfa Aesar (Kandel, Germany); N.A: Not Available; SA: Sigma-Aldrich (Buchs, Switzerland).
Table 2. Isotherm parameters calculations for the adsorption of CA-99Mo on different metal oxides NPs.
Table 2. Isotherm parameters calculations for the adsorption of CA-99Mo on different metal oxides NPs.
Isotherm ModelParameterCeO2-544841ZrO2-544760TiO2-637254Al2TiO5-634143CeO2/ZrO2-634174CeO2-700290
LangmuirnL (mg∙g−1)26.70416.81436.98010.20723.47019.603
KL (L∙mg−1)0.4070.9930.3110.09073.5370.254
R20.9110.9570.9300.8360.8700.954
FreundlichKF (mg1−nLn∙g−1)10.5149.05813.0646.36411.5068.876
nf5.0108.2944.07910.3256.3466.644
R20.9820.9680.9890.8980.9550.966
Table 3. Thermodynamic parameters for the sorption of CA-99Mo on different metal oxides NPs.
Table 3. Thermodynamic parameters for the sorption of CA-99Mo on different metal oxides NPs.
AdsorbentTemperature
(K)
ΔG0
(kJ∙mol−1)
ΔH0
(kJ∙mol−1)
ΔS0
(J∙mol−1∙K−1)
CeO2-544841298−9.8 ± 2.9−5.1 ± 1.516.0 ± 4.7
313−10.1 ± 3.0
323−10.2 ± 3.0
333−10.4 ± 3.1
ZrO2-544760298−7.8 ± 1.4−1.4 ± 0.721.3 ± 2.2
313−8.1 ± 1.4
323−8.3 ± 1.4
333−8.5 ± 1.5
TiO2-637254298−10.9 ± 3.14.7 ± 1.552.0 ± 5.0
313−11.7 ± 3.1
323−12.2 ± 3.2
333−12.7 ± 3.2
Al2TiO5-634143298−7.2 ± 2.0−4.5 ± 19.1 ± 3.2
313−7.4 ± 2.0
323−7.5 ± 2.0
333−7.6 ± 2.0
CeO2/ZrO2-634174298−9.0 ± 1.41.3 ± 0.734.6 ± 2.2
313−9.5 ± 1.4
323−9.9 ± 1.4
333−10.2 ± 1.4
CeO2-700290298−8.2 ± 1.9−0.1 ± 0.927.2 ± 3.0
313−8.6 ± 1.9
323−8.9 ± 2.0
333−9.1 ± 2.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nawar, M.F.; El-Daoushy, A.F.; Ashry, A.; Soliman, M.A.; Türler, A. Evaluating the Sorption Affinity of Low Specific Activity 99Mo on Different Metal Oxide Nanoparticles. Inorganics 2022, 10, 154. https://doi.org/10.3390/inorganics10100154

AMA Style

Nawar MF, El-Daoushy AF, Ashry A, Soliman MA, Türler A. Evaluating the Sorption Affinity of Low Specific Activity 99Mo on Different Metal Oxide Nanoparticles. Inorganics. 2022; 10(10):154. https://doi.org/10.3390/inorganics10100154

Chicago/Turabian Style

Nawar, Mohamed F., Alaa F. El-Daoushy, Ahmed Ashry, Mohamed A. Soliman, and Andreas Türler. 2022. "Evaluating the Sorption Affinity of Low Specific Activity 99Mo on Different Metal Oxide Nanoparticles" Inorganics 10, no. 10: 154. https://doi.org/10.3390/inorganics10100154

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