Next Article in Journal
High-Efficiency Crystalline Silicon-Based Solar Cells Using Textured TiO2 Layer and Plasmonic Nanoparticles
Previous Article in Journal
Structure and Magnetic Properties of ErFexMn12−x (7.0 ≤ x ≤ 9.0, Δx = 0.2)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sorption Profile of Low Specific Activity 99Mo on Nanoceria-Based Sorbents for the Development of 99mTc Generators: Kinetics, Equilibrium, and Thermodynamic Studies

by
Mohamed F. Nawar
1,2,*,
Alaa F. El-Daoushy
2,
Metwally Madkour
3 and
Andreas Türler
1
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
Chemistry Department, Faculty of Science, Kuwait University, Safat 13060, Kuwait
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(9), 1587; https://doi.org/10.3390/nano12091587
Submission received: 7 April 2022 / Revised: 28 April 2022 / Accepted: 5 May 2022 / Published: 7 May 2022

Abstract

:
99Mo/99mTc generators play a significant role in supplying 99mTc for diagnostic interventions in nuclear medicine. However, the applicability of using low specific activity (LSA) 99Mo asks for sorbents with high sorption capacity. Herein, this study aims to evaluate the sorption behavior of LSA 99Mo towards several CeO2 nano-sorbents developed in our laboratory. These nanomaterials were prepared by wet chemical precipitation (CP) and hydrothermal (HT) approaches. Then, they were characterized using XRD, BET, FE-SEM, and zeta potential measurements. Additionally, we evaluated the sorption profile of carrier-added (CA) 99Mo onto each material under different experimental parameters. These parameters include pH, initial concentration of molybdate solution, contact time, and temperature. Furthermore, the maximum sorption capacities were evaluated. The results reveal that out of the synthesized CeO2 nanoparticles (NPs) materials, the sorption capacity of HT-1 and CP-2 reach 192 ± 10 and 184 ± 12 mg Mo·g–1, respectively. For both materials, the sorption kinetics and isotherm data agree with the Elovich and Freundlich models, respectively. Moreover, the diffusion study demonstrates that the sorption processes can be described by pore diffusion (for HT-synthesis route 1) and film diffusion (for CP-synthesis route 2). Furthermore, the thermodynamic parameters indicate that the Mo sorption onto both materials is a spontaneous and endothermic process. Consequently, it appears that HT-1 and CP-2 have favorable sorption profiles and high sorption capacities for CA-99Mo. Therefore, they are potential candidates for producing a 99Mo/99mTc radionuclide generator by using LSA 99Mo.

1. Introduction

There is an increasing interest in using 99mTc (T1/2 = 6.01 h) for diagnostic purposes in nuclear medicine and radiotracing applications in the industry [1,2,3,4]. This increased interest has heightened the need to produce 99Mo/99mTc generators on a large scale. The production of these generators includes different technologies, for instance, sublimation, electrochemical, solvent extraction, supported liquid membrane (SLM), and column chromatographic approaches [5,6,7,8,9]. Among these generators, the portable column chromatographic type is considered the primary source to supply ready-to-use 99mTc onsite [7,10]. The idea of these generators is to retain the parent 99Mo on a sorbent for decay so that the daughter 99mTc can then be easily separated in high purity by using an isotonic saline solution at desired time intervals [11,12].
The 99mTc parent, 99Mo, can be produced in nuclear reactors either by fission of 235U or by direct neutron irradiation of natural Mo or enriched 98Mo targets [10,13]. On the one hand, fission produces more than 95% of the 99mTc generators on the market because it provides 99Mo with high specific activity (>104 Ci/g Mo). On the other hand, this route has some disadvantages. The main ones are the prohibitive production costs, the generation of large amounts of radioactive waste, and the increasing proliferation risks [14,15,16]. To overcome these difficulties, many methods have been suggested for using neutron-activated 99Mo of low specific activity (LSA) (~5–10 Ci/g Mo) [15,17,18,19,20]. However, the low sorption capacity of commercial materials is the most critical limitation, which reduces their applicability. It is worth mentioning that the retention capacity of alumina (Al2O3), the material used, is limited to only 2–20 mg.g–1 for molybdate (99MoO42−) [8].
Recently, nanotechnology has shown that it has the potential to develop the properties of different materials. These newly developed materials provide attractive solutions in multidisciplinary sectors, such as biology, chemistry, and medical sciences [21]. Nanostructured materials possess a unique surface morphology, which derives from their small particle size structure, high surface area, and the large number of active sites. The successful synthesis of such nanostructured sorbents has been reported by using several approaches. Out of these techniques, wet-chemical precipitation and hydrothermal methods bring potential advantages. Wet chemical precipitation is a simple, cost-effective method that can be conducted at ambient temperature [22,23,24]. The hydrothermal route offers high product purity with uniform composition, less particle agglomeration, controlled surface morphology, and narrow particle size distribution [25,26]. Consequently, this unique surface morphology significantly improves sorption efficiency and high adsorbate selectivity [27]. In this context, the use of nanomaterial-based sorbents appears to be a promising solution to overcome the low sorption capacity of conventional sorbents for LSA 99Mo. These new sorbents might make it possible to produce 99Mo/99mTc generators by using LSA 99Mo.
In this paper, our main objective is to evaluate the sorption profile of LSA 99Mo on several CeO2 nanosorbents developed in our laboratory. To achieve this goal, we report the preparation of CeO2 NPs using wet chemical precipitation and hydrothermal methods. Moreover, we investigated the sorption efficiency of the prepared materials for Mo under different experimental conditions such as pH, the initial concentration of the molybdate solution, contact time, and temperature. In addition, the sorption kinetics, equilibrium isotherm, and thermodynamics were evaluated.

2. Materials and Methods

2.1. Materials

All chemicals were of analytical grade purity and were used without further purification. Milli-Q water was used for the preparation of solutions and washings. Cerium nitrate hexahydrate Ce(NO3)3·6H2O, sodium hydroxide, ethanol absolute, and nitric acid were purchased from Merck, Darmstadt, Germany.
99Mo radiotracer solution was obtained by eluting 57.1 GBq fission 99Mo alumina-based 99Mo/99mTc generator (Pertector, National Centre for Nuclear Research, Otwock, Poland) with 5 mL 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 Mirion Technologies, Inc., Florham Park, NJ, USA). The 99Mo eluate solution was passed through 0.45 micro-Millipore filter to retain any alumina particles. Then, the 99Mo solution was diluted with HNO3 solution to the desired pH value.

2.2. Instrumentation

Radiometric identifications and measurements were carried out by using a multichannel analyzer (MCA), Inspector 2000 model, Canberra Series, Mirion Technologies, Inc., Meriden, CT, USA, coupled with a high-purity germanium detector (HPGe). Samples of constant geometry were counted at a low dead time (<5%). The radioactivity levels were determined by quantifying the 740 keV photo peaks corresponding to 99Mo. A pH-meter with a microprocessor (Mettler Toledo, Seven Compact S210 model, Greifensee, Switzerland) was used to measure the pH values. A thermostated shaking water bath (Julabo GmbH, Seelbach, Germany) was used for all batch equilibrium studies. The X-ray powder diffraction (XRD) patterns were performed via a D8 ADVANCE diffractometer (Bruker AXS Inc, Madison, WI, USA) with a Cu-Kα radiation source (λ = 0.1542 nm) operating at 40 kV, 40 mA, and a scanning range of 2θ = 10–80° at room temperature. The Brunauer–Emmett–Teller (BET) surface area of the synthesized materials was measured using nitrogen sorption isotherms at −195 °C on a model Gemini VII, ASAP 2020 (Micromeritics Instrument, Norcross, GA, USA). Before the analysis, the sample was degassed for 12 h at 110 °C. Zeta potential (ζ) measurements were studied using a Zeta-sizer (Nano ZS, Malvern, UK) for isoelectric point (pHIEP) measurements. A JSM-7100F (JEOL) was used to conduct field emission scanning electron microscopy (FE-SEM) measurements.

2.3. Synthesis of Cerium Oxides Nanoparticles

2.3.1. Wet Chemical Precipitation Method

CeO2 NPs were prepared by a wet chemical method via the precipitation of cerium nitrate in an alkaline medium. In a typical procedure, 2 g of Ce(NO3)3·6H2O were dissolved in a mixture of Milli-Q water and ethanol absolute with a ratio of (1:4 v/v). Then, this solution was added drop-wise to a freshly prepared 0.4 M sodium hydroxide solution (pH~13) under continuous stirring, and the pH value was controlled at about 11. The mixture was stirred until a yellow suspension of CeO2 NPs was formed after approximately 30 min. After that, the formed precipitate was washed several times with de-ionized water and ethanol absolute so that the soluble salts and fine impurities could be removed. Subsequently, the washed precipitate was centrifuged at 4000 rpm for 15 min and dried in a drying oven at about 50–70 °C for 12 h. The resulting sample is denoted as CP (Chemical Precipitation).

2.3.2. Hydrothermal Modification Method

Hydrothermally modified CeO2 NPs were synthesized according to the same procedure as described in Section 2.3.1 with some modifications using a mixture of absolute ethanol and Milli-Q water (4:1 v/v). First, the formed NPs were transferred into a Teflon-lined stainless steel autoclave. Then, the autoclave was sealed and maintained at 150 °C for 12 h. Subsequently, the formed precipitate was washed several times with de-ionized water and ethanol absolute, centrifuged, and dried at about 70 °C for 12 h. The dried material is called HT (Hydrothermal Treatment).
Furthermore, the synthesized CeO2-CP and CeO2-HT NPs were heated at different temperatures: 200 °C and 400 °C with a heating rate of 10 °C/min for 2 h (Figure 1).

2.4. Batch Distribution Studies

2.4.1. Effect of Solution pH

The influence of medium pH on carrier-added 99Mo uptake on each sorbent was evaluated at a wide range of pH values. For this purpose, we added 30 mL of molybdate solution (50 mg·L−1) to 300 mg of each sorbent. The uptake behavior was investigated at different pH values, which ranged from 1 to 12. The initial pH of the 99Mo solution was adjusted by using either HNO3 or NaOH. The mixtures were shaken in a thermostatic shaker at 25 ± 1 °C for 24 h. After that, the liquid phases were separated, centrifuged, and then 1 mL aliquot was withdrawn for radiometric measurements. In these analyses, we measured the total solutions radioactivity before and after equilibration with an HPGe counter by using an appropriate gamma-ray peak (740 keV). Additionally, both initial and equilibrium pH values were measured. Finally, the uptake percent of each material was calculated as a function of solution pH by using the following equation:
Uptake % = A o   A e A o × 100
where Ao and Ae are the initial and equilibrium of 99Mo radioactivity expressed in counts/min.

2.4.2. Sorption Kinetics

To estimate the sorption rate of carrier-added 99Mo on the prepared CeO2 NPs, we determined the uptake percent of 99MoO42− ions (~pH 3) at different time intervals. Typically, 300 mg of each sorbent were suspended in 30 mL of 99MoO42− solution (50 mg·L−1 in 0.001 M HNO3) and shaken in thermostated shaker bath at 25 ± 1 °C for different time intervals (5 min–50 h). Then, 1 mL of the supernatant was separated and measured. The sorption of Mo on each material was followed with time until the equilibrium was established. Finally, we calculated the Mo capacity ( Q t ) in mg/g at each time (t) by using the following equation:
Q t = A o   A t A o × C o × V m
where Ao (count/min) is the initial activity, and A t is the 99Mo activity at a time (t), C o (mg·L−1) is the initial 99MoO42– ion concentration,   V (L) is the total volume, and m (g) is the sorbent weight.
Furthermore, for a comprehensive understanding of the reaction kinetics and diffusion mechanisms, we used the obtained results and applied different sorption kinetic and diffusion models.

2.4.3. Equilibrium Sorption Isotherms

Sorption isotherm studies were conducted by varying the initial 99MoO42– ion concentration from 50 to 5000 mg·L−1 while keeping the sorbent dose constant. Herein, we mixed 30 mL of 99Mo solution (pH ~ 3) with 300 mg of each material. The other parameters, such as equilibrium time and temperature, were maintained at 24 h and 25 ± 1 °C, respectively. Subsequently, the aqueous phases were pipetted, centrifuged, and measured. The equilibrium adsorption capacity (Qe) in mg·g−1 was calculated using Equation (2). The obtained results were evaluated using sorption isotherm models.
In order to evaluate sorption isotherm models, three error distribution functions were used. These functions are Chi-square ( X 2), average percentage error (APE), and root mean square error (RMSE). These functions can be calculated according to the following equations:
X 2 = ( Q e ( exp )   Q e ( cal ) ) 2   Q e ( cal )  
APE = ( 1 N ( | Q e ( exp )   Q e ( cal ) | / Q e ( exp ) ) / N ) × 100    
RMSE = ( 1 / N 2 ) × 1 N ( Q e ( exp )   Q e ( cal ) ) 2  

2.4.4. Temperature Effects

The change in molybdate uptake was studied under different reaction temperatures. The study was conducted between 298 and 343 K, and the other reaction conditions were kept constant. Moreover, to clarify the thermodynamic nature of the sorption process, we used the obtained data to investigate different thermodynamic parameters. These parameters are the standard enthalpy change (ΔH°), the standard entropy change (ΔS°), and the Gibbs free energy change (ΔG°).

3. Results and Discussion

In order to evaluate the sorption profile of the molybdate on the synthesized CeO2 NPs, different parameters were studied using the batch contact method. These parameters include solution pH, contact time, initial ion concentration (mi), and temperature. The impact of each parameter on the Mo sorption was studied separately. Meanwhile, the other factors were kept constant.

3.1. Structural Characterization of the Prepared CeO2 NPs

The XRD patterns of CeO2 nanoparticles prepared via wet chemical precipitation and hydrothermal methods are shown in Figure 2. The peaks are indexed using JCPDS card no: 34-0394. All samples of CeO2 nanoparticles have face centered cubic structure with the lattice parameters a = b = c = 5.411 Å and α = β = γ = 90°. The diffraction peaks were found at 28.56°, 33.08°, 47.47°, 56.36°, 59.08°, 69.40°, and 76.70° [28]. The absence of impurity indicates that both methods synthesize pure CeO2. The average crystallite size (D) of CeO2 nanoparticles is calculated by using the Debye–Scherrer equation:
D = K λ β cos θ
where λ is the wavelength of the Cu-Kα radiation, K is a constant (=0.89), θ is the diffraction angle, and β is the full width at half maximum (FWHM). The average crystallite size values for CeO2 nanoparticles were evaluated and listed in Table 1. The results revealed that the thermal treatment affects the crystallite size of the samples prepared by the wet chemical precipitation method. For these samples, elevating the temperature increases the crystallite size. However, in the case of the hydrothermal method, the thermal treatment does not influence the crystallite size.
We further examined the textural properties of the CeO2 nanoparticles using a nitrogen adsorption isotherm (Figure 3). All the samples exhibited Type IV isotherms with well-defined H2-type hysteresis loops, which indicates the mesoporous nature of the materials [29]. The BET surface area, pore volume, and pore size of all samples are tabulated in Table 1. The highest BET surface area of the mesoporous CeO2 is 187.2 m2·g−1. The surface area of the wet chemical precipitation method was increased upon increasing the temperature up to 200 °C and reached 175.3 m2·g−1, and further thermal treatment resulted in its reduction to 97.4 m2·g−1. Compared to CeO2 NPs prepared with the wet chemical precipitation method, the hydrothermally modified samples exhibit a higher surface area due to improved surface modification.
The zeta potential (ζ) measurements listed in Table 1 give insight into the surface charge. It revealed that the pHIEP of CeO2 NPs is at about seven. Therefore, the surface of CeO2 NPs bears a positive charge below pH ~ 7 and a negative charge above this pH value.
Additionally, as the morphology of the nanoparticles is a key characteristic factor for their performance, we investigated the effect of the preparation method with different temperatures on the CeO2 NPs morphology. Figure 4 presents typical FE-SEM images of CeO2 NPs prepared at different temperatures. As shown in the SEM images, the hydrothermally prepared CeO2 NPs exhibited spherical morphology with almost uniform characteristics. There is no obvious change in the morphology upon changing the temperature from 50 °C to 400 °C. The CeO2 NPs prepared by the wet chemical precipitation method showed a rough morphology compared with the hydrothermally modified analogous especially at a low temperature (50 °C). Upon increasing the temperature, the rough shape was regulated and semi-spherical NPs appeared as shown for CP-2 and CP-3.

3.2. Effect of Solution pH

Batch experiments were conducted to verify the usefulness of the synthesized CeO2 NPs for carrier-added 99Mo sorption from aqueous solutions. The pH of the Mo solution is the key factor that controls the sorption process, as it has a profound impact on determining the ionic state of the functional groups that exist on the sorbent surface. Furthermore, it affects the dissociation, complexation, and/or ionization of the adsorbed ions [30]. In order to predict the optimal experimental conditions for carrier-added 99Mo uptake, the sorption behavior of each material was evaluated at a wide range of pH values (1–12). The results are displayed in Figure 5a. The figure shows that the highest uptake values occur in an acidic medium (pH 2–4). After that, the uptake percent starts to decrease. This behavior can be explained based on two factors: the surface charge of the solid phase and the distribution of the molybdate species in the solution.
Since the external solution pH governs the sorbents’ surface acid-base properties, we attempted to clarify the isoelectric point (ζ) of each material. Table 1 shows that the pHIEP is reached at ~pH 7. Consequently, the CeO2 surface carries a positive charge at low pH (pH < 7). The change in the surface charge of the solid phase is because the active sites on the surface are hydrated and covered by amphoteric hydroxyl groups. Hence, based on the pH of the medium, these hydroxyl groups can undergo different reactions with either H+ or OH to develop positive or negative charges on the surface. Herein, at pH < 7, they are protonated, and the CeO2 surface develops a positive charge as follows:
Surface OH Surface +   H solution +     Surface OH 2 +
Furthermore, the data in Figure 5a can be interpreted by considering the speciation diagram of molybdenum shown in Figure 5b. The data were calculated with the help of the PHREEQC software (version 3). Molybdenum exists in the solution as molybdate (MoO42−). It polymerizes in an acidic medium to polymolybdate (Mo7O246−) with higher molybdenum content per unit charge [31]. Therefore, an electrostatic attraction between negatively charged molybdate anions and positively charged surfaces occurs.

3.3. Effect of Contact Time

The progress of Mo sorption onto the synthesized CeO2 was studied at different time intervals, and the results are represented in Figure S1. The data illustrate that for all sorbents, the uptake percent (U%) of Mo increases with increasing the contact time until a sort of equilibrium state is reached, after which negligible or no further uptake takes place. In more detail, the sorption process took place in two distinct steps. The first one involves a fast uptake rate, which may be attributed to the high availability of active sites on the sorbent surface at the beginning of the sorption reaction. The second one exhibits a relatively slow sorption rate until the equilibrium is established. That is because the number of active sites is rapidly diminished, and the ions have to compete for the sorption sites [32]. Accordingly, under our experimental conditions, the optimum contact times for CP-1, CP-2, CP-3, HT-1, HT-2, and HT-3 are 6 h, 1 h, 30 h, 15 min, 1 h, and 10 h, respectively.
Furthermore, the results revealed that compared to materials prepared via the wet chemical method, the hydrothermally modified samples introduce a faster reaction rate. However, heated materials at 400 °C (CP-3 and HT-3) exhibit a relatively slow sorption rate. This behavior may be explained because the thermal treatment of nanomaterials decreases the number of active surface sites, which reduces the sorbent reactivity [32].
Based on the earlier experimental and surface morphological evaluations, the hydrothermal modification improved the sorption profile of HT-1 considerably. Moreover, except for CP-2, the thermal treatment does not enhance the uptake capability of the heated materials. On the contrary, it adversely diminishes the sorption rate for the ones heated at 400 °C. Therefore, CP-2 and HT-1 were selected for subsequent evaluation studies.

3.3.1. Kinetic Studies

To design an effective sorption process, determining kinetic parameters is crucial [33]. Kinetic modeling provides helpful information about the rate of the sorption process. Therefore, we analyzed the sorption kinetic data of carrier-added 99Mo onto CP-2 and HT-1 by using three different kinetic models, namely, Lagergren-first-order (Equation (8)) [34], pseudo-second-order (Equation (9)) [35], and Elovich (Equation (10)) [36,37]. These models are presented in a non-linear regression form by using the equations as follows:
Q t =   Q e ( 1 exp K 1 t   )  
Q t = K 2 Q e 2 t 1 + K 2 Q e t  
Q t = 1 β ln ( 1 + α β t )
where Q t   and Q e (mg·g−1) are the amount of adsorbed molybdenum at a time (t) and equilibrium, respectively, k 1 (min−1) is the Lagergren-rate constant,   k 2 (g/mg·min) is the pseudo-second-order rate constant, α (mg/g·min) is the Elovich rate constant, and β (mg·g−1) is the desorption constant representing the degree of surface coverage.
Figure 6 displays non-linear fits of sorption kinetic data of Mo onto CP-2 and HT-1 according to the three aforementioned kinetic models by plotting calculated (qt) values versus contact time (t). The calculated kinetic parameters are presented in Table 2. Good non-linear plots are shown in Figure 6 for all kinetic models under our experimental conditions. Furthermore, by comparing the correlation coefficient values (R2) of the three applied models (Table 2), it can be seen that the three applied models have close R2 values. Nonetheless, we can notice that the Elovich model exhibits higher values (R2 = 0.999) than Lagergren-first-order model (R2 = 0.989 and 0.990) and the pseudo-second-order model (R2 = 0.991 and 0.990) for CP-2 and HT-1, respectively. Accordingly, we suggest that the Elovich model is the most appropriate to explain our sorption kinetics data. This finding may indicate that the prepared CeO2 matrices possess heterogeneous surfaces, and the uptake of carrier-added 99Mo develops multi-adsorption layers [36,37].

3.3.2. Diffusion Studies

Generally, the uptake of adsorbate ions on the sorbent material occurs through [38]: (i) the movement of adsorbate ions from the bulk solution to the sorbent surface (bulk diffusion), (ii) the diffusion of adsorbate ions through the liquid film covering the sorbent surface (film diffusion), (iii) internal diffusion of adsorbate ions within the interior pores of the sorbent material (pore diffusion), and (iv) the binding of adsorbate ions on available active sites. In order to investigate the rate-limiting step, three diffusion models are applied. These models are film diffusion (McKay, Equation (11)) [39], intra-particle diffusion (Weber and Morris, Equation (12)) [40], and pore diffusion model (Bangham, Equation (13)) [40]. The models are expressed by the following equations:
  ln ( 1 Q t Q e ) =   k f t  
Q t = K ip t 1 2 + C  
  log ( log ( C i C i   Q t w   ) ) = log ( K b W 2.303   V   ) +   σ   log   t  
where Q t   and Q e (mg·g−1) are the amount of adsorbed molybdenum at a time (t) and equilibrium, respectively. Ci (mg·L−1) is the initial Mo concentration, W (g) is the sorbent mass, and V (L) is the solution volume. k f (min−1), k ip (mg/g·min1/2), and σ are film diffusion, intra-particle diffusion, and pore diffusion rate constants, respectively.
The rate constants can be determined from the slope of the linear plots of ln ( 1 q t q e ) versus t (Equation (11)), q t versus t (Equation (12)), and log ( log ( C i C i   Q t w   ) ) versus t (Equation (13)). The results are shown in Figure 7. The calculated diffusion parameters and the correlation coefficient values (R2) are tabulated in Table 3. In order to estimate the best mechanism that describes our diffusion data, we compare the correlation coefficient values (R2) of each model. For CP-2, the application of the McKay model resulted in a higher value (R2 = 0.992) than both Weber and Morris model (R2 = 0.977) and the Bangham model (R2 = 0.967). These findings indicate that the sorption of Mo on CP-2 might occur through an external film diffusion mechanism. For HT-1, the correlation coefficient values of the three models were: McKay model (R2 = 0.968), Weber and Morris model (R2 = 0.946), and Bangham model (R2 = 0.980). Therefore, it can be concluded that the pore diffusion model might control Mo sorption on HT-1.

3.4. Equilibrium Isotherm Studies

Sorption isotherms describe how the adsorbed ions are dispersed between liquid and solid phases when the adsorption process reaches an equilibrium state at a constant temperature [41,42]. Consequently, a proper interpretation of equilibrium isotherms is necessary to improve the sorption mechanism and design an effective sorption system [31,43,44]. Figure 8 illustrates the influence of initial molybdate concentration on the equilibrium sorption capacity (Qe) and the uptake percent of Mo onto CP-2 and HT-1 at constant temperature (25 ± 1 °C). The figure shows that on increasing the initial Mo concentration, the uptake percent decreases, and on the contrary, the sorption capacity (Qe) increases. These findings are attributed to the increased mass driving force at higher molybdate concentrations [45,46,47].
Based on the sorbents’ behavior and the mode of interaction, the equilibrium isotherm data shown in Figure 8 were modeled. We applied three widely-used two-parameters, and isotherm models. These models are Freundlich [48], Langmuir [49], and Temkin [50]. The non-linear regressions of these models are represented by Equations (14)–(16), respectively:
Q e = K F C e 1 n
Q e = Q m K l C e 1 + ( K l C e )
Q e = RT b t ln ( A t C e )
where Qe (mg·g−1) is the equilibrium sorption capacity, Ce (mg·L−1) is the equilibrium concentration of molybdate ions, Qm (mg·g−1) is the maximum capacity. KF (mg1-n·Ln/g), KL (L·mg−1), and   A t   (L·g−1) are the Freundlich, Langmuir, and Temkin constants, respectively. The parameters n and B reflect the heterogeneity of adsorption sites and the heat of adsorption, respectively. R   is the universal gas constant (0.008314 kJ·mol−1·K−1), and T (K) is the absolute temperature.
Figure 9 introduces the non-linear fits of isotherm data of Mo onto CP-2 and HT-1 for the three isotherms models mentioned above by plotting calculated (Qe) values versus (Ce). The calculated equilibrium isotherm parameters and their corresponding correlation coefficient (R2) are given in Table 4. According to the correlation coefficient (R2) values, the Freundlich isotherm best describes the equilibrium data. It has a higher correlation coefficient (R2 = 0.986, 0.989) than Langmuir (R2 = 0.958, 0.983) and Temkin (R2 = 0.794, 0.756) for CP-2 and HT-1, respectively. This result is confirmed by the lower values of the error functions (X2), (APE), and (RMSE) for the Freundlich isotherm compared to the Langmuir and Temkin models (Table 4). Furthermore, the sorption intensity (n) values (for the Freundlich isotherm) are 3.35 (for CP-2) and 3.98 (for HT-1), which indicates a favorable sorption process. This result shows that CP-2 and HT-1 have heterogeneous surfaces with a wide distribution of non-equivalent sorption sites.
Furthermore, the maximum sorption capacity for Mo was evaluated under static conditions. We conducted repeated equilibrations of carrier-added 99Mo with CP-2 and HT-1 at 25 ± 1 °C until no further uptake occurred. The maximum sorption capacity was determined according to the following equation:
Q max = Uptake % 100 ×   C o × V m
where Q max (mg·g−1) is the maximum (saturation) sorption capacity, C o (mg·L−1) is the initial concentration,   V   (L) is the volume of the aqueous phase, and m   (g) is the matrix weight.
The results show that the sorption capacities of CP-2 and HT-1 reach 184 ± 12 and 192 ± 10 mg Mo·g−1, respectively. These values might be attributed to the high surface area of CP-2 and HT-1. In addition, the sorption behavior of each sorbent allows the formation of multi-adsorption layers on their surfaces, which is confirmed by kinetic and equilibrium isotherm studies. It is pertinent to point out that the sorption capacities of CeO2 nanosorbents we developed are much higher than the sorption capacity of conventional alumina (2–20 mg Mo·g−1) [8]. Therefore, the new materials have the potential to be implemented to produce 99Mo/99mTc generators by using LSA 99Mo.

3.5. Thermodynamic Studies

The effect of temperature on molybdenum uptake onto CP-2 and HT-1 was studied in the temperature range (298–343 K). With increasing the reaction temperature, a slight increase in carrier-added 99Mo uptake is observed on HT-1. Meanwhile, the sorption capacity of CP-2 increases noticeably. These findings indicate that the sorption process is endothermic in nature.
In order to understand the thermodynamic behavior of the sorption process, we investigated the values of the enthalpy change (ΔH°), the entropy change (ΔS°), and the free energy change (ΔG°) in our experimental temperature window. These parameters were calculated by using Vant’ Hoff equations [51], as follows:
lnK d = ( Δ H ° RT ) + ( Δ S ° R )
Δ G ° = Δ H ° T Δ S °  
Δ G ° = RTlnK d
where Kd (mg·mL−1) is the distribution coefficient, R is the universal gas constant (0.008314 kJ·mol−1·K−1), and T (K) is the absolute temperature.
Figure 10 shows linear plots of lnKd versus the reciprocal absolute temperature (1/T). The values of ∆H° and ∆S° can be determined from the slope and intercept, respectively. The calculated values of ΔH°, ΔS°, and ∆G° are listed in Table 5. The positive values of ΔH° indicate that the sorption processes of Mo onto CP-2 and HT-1 are endothermic. In addition, the positive values of ΔS° point to the increase in randomness at the solid–liquid interface. Moreover, the negative values of ∆G° confirm the spontaneous nature of the sorption processes under our experimental conditions [52].

4. Summary and Conclusions

Our main goal was to evaluate the sorption profile of LSA 99Mo on several CeO2 nanosorbents developed in our laboratory. First, these materials were synthesized by using simple wet-chemical precipitation (CP) and hydrothermal (HT) methods. Then, they were heated at different temperatures: 200 °C and 400 °C. Eventually, an assessment of the carrier-added 99Mo uptake behavior was carried out under different experimental conditions. We found that the hydrothermal modification improves the morphological and sorption profile of HT-1 considerably. Except for CP-2, the thermal treatment did not enhance the sorption capability of the heated materials. In contrast, it diminished the Mo sorption for the ones heated at 400 °C. In addition, sorption kinetics data of carrier-added 99Mo Mo onto CP-2 and HT-1 were most favorably described by the Elovich model. The diffusion data showed that film and pore diffusion models controlled the sorption process for CP-2 and HT-1, respectively. Moreover, out of the investigated isotherm models, the Freundlich model was the most appropriate to describe the equilibrium isotherms of carrier-added 99Mo on both sorbents. Furthermore, we studied the influence of reaction temperature on Mo uptake, and the results confirmed the spontaneous and endothermic nature of the sorption processes. Finally, we evaluated the maximum sorption capacity of CP-2 and HT-1 towards carrier-added 99Mo under static sorption conditions. Their capacity reached 184 ± 12 and 192 ± 10 mg Mo.g–1, respectively. From the outcome of our investigation, it is possible to conclude that HT-1 and CP-2 have favorable sorption profiles and appreciable uptake capacities for carrier-added 99Mo. Consequently, they have the potential for producing a 99mTc radionuclide generator by using LSA 99Mo. Clearly, the next stage of our research will investigate the feasibility of using these two sorbent materials to develop a useful 99Mo/99mTc generator.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12091587/s1, Figure S1: Effect of contact time on Mo uptake on synthesized CeO2 NPs.

Author Contributions

M.F.N.; Suggested the research idea, conducted the experiments, analyzed the data, drafted the original manuscript, and acquisition of the financial support for the project leading to this publication. A.F.E.-D. and M.M.; carried out the characterization measurements, co-analyzed the data, and reviewed the original draft. A.T.; supervised the project and reviewed the original draft. 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. In addition, Kuwait University facilities No. GS 01/01, GS 01/05, GE 01/07, and GS 03/01 are highly acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ruan, Q.; Gan, Q.; Zhang, X.; Fang, S.; Zhang, J. Preparation and Bioevaluation of Novel 99mTc-Labeled Complexes with a 2-Nitroimidazole HYNIC Derivative for Imaging Tumor Hypoxia. Pharmaceuticals 2021, 14, 158. [Google Scholar] [CrossRef] [PubMed]
  2. Gumiela, M. Cyclotron production of 99mTc: Comparison of known separation technologies for isolation of 99mTc from molybdenum targets. Nucl. Med. Biol. 2018, 58, 33–41. [Google Scholar] [CrossRef]
  3. Martini, P.; Boschi, A.; Cicoria, G.; Zagni, F.; Corazza, A.; Uccelli, L.; Pasquali, M.; Pupillo, G.; Marengo, M.; Loriggiola, M.; et al. In-house cyclotron production of high-purity Tc-99m and Tc-99m radiopharmaceuticals. Appl. Radiat. Isot. 2018, 139, 325–331. [Google Scholar] [CrossRef] [PubMed]
  4. Urbano, N.; Scimeca, M.; Tancredi, V.; Bonanno, E.; Schillaci, O. 99mTc-sestamibi breast imaging: Current status, new ideas and future perspectives. Semin. Cancer Biol. 2020, in press. [Google Scholar] [CrossRef] [PubMed]
  5. Chakravarty, R.; Dash, A.; Venkatesh, M. A novel electrochemical technique for the production of clinical grade 99mTc using (n,γ)99Mo. Nucl. Med. Biol. 2010, 37, 21–28. [Google Scholar] [CrossRef] [PubMed]
  6. Chang, S.H. Types of bulk liquid membrane and its membrane resistance in heavy metal removal and recovery from wastewater. Desalin. Water Treat. 2016, 57, 19785–19793. [Google Scholar] [CrossRef]
  7. I.A.E.A. Radiotracer Generators for Industrial Applications; International Atomic Energy Agency: Vienna, Austria, 2013. [Google Scholar]
  8. Molinski, V.J. A review of 99mTc generator technology. Int. J. Appl. Rad. Isot. 1982, 33, 811–819. [Google Scholar] [CrossRef]
  9. I.A.E.A. Cyclotron Based Production of Technetium-99m; International Atomic Energy Agency: Vienna, Austria, 2017. [Google Scholar]
  10. 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]
  11. Knapp, F., Jr.; Baum, R. Radionuclide generators-a new renaissance in the development of technologies to provide diagnostic and therapeutic radioisotopes for clinical applications. Curr. Radiopharm. 2012, 5, 175–177. [Google Scholar] [CrossRef]
  12. Osso, J.A., Jr.; Catanoso, M.F.; Barrio, G.; Brambilla, T.P.; Teodoro, R.; Dias, C.R.B.R.; Suzuki, K.N. Technetium-99m: New Production and Processing Strategies to Provide Adequate Levels for SPECT Imaging. Curr. Radiopharm. 2012, 5, 178–186. [Google Scholar] [CrossRef]
  13. N.A.P. Molybdenum-99 for Medical Imaging; The National Academies Press: Washington, DC, USA, 2016. [Google Scholar] [CrossRef] [Green Version]
  14. Qaim, S.M. The present and future of medical radionuclide production. Radiochim. Acta 2012, 100, 635–651. [Google Scholar] [CrossRef]
  15. I.A.E.A. Non-HEU Production Technologies for Molybdenum-99 and Technetium-99m; International Atomic Energy Agency: Vienna, Austria, 2013. [Google Scholar]
  16. I.A.E.A. Feasibility of Producing Molybdenum-99 on a Small Scale Using Fission of Low Enriched Uranium or Neutron Activation of Natural Molybdenum; International Atomic Energy Agency: Vienna, Austria, 2015. [Google Scholar]
  17. Evans, J.V.; Moore, P.W.; Shying, M.E.; Sodeau, J.M. Zirconium molybdate gel as a generator for technetium-99m—I. The concept and its evaluation. Int. J. Rad. Appl. Instrum. A 1987, 38, 19–23. [Google Scholar] [CrossRef]
  18. Maoliang, L. Production of gel-type Tc-99m generator for nuclear medicine. In Proceedings of the 12th KAIF/KNS Annual Conference, Seoul, Korea, 3–4 April 1997. [Google Scholar]
  19. Mostafa, M.; Saber, H.M.; El-Sadek, A.A.; Nassar, M.Y. Preparation and performance of 99Mo/99mTc chromatographic column generator based on zirconium molybdosilicate. Radiochim. Acta 2016, 104, 257–265. [Google Scholar] [CrossRef]
  20. Sekimoto, S.; Tatenuma, K.; Suzuki, Y.; Tsuguchi, A.; Tanaka, A.; Tadokoro, T.; Kani, Y.; Morikawa, Y.; Yamamoto, A.; Ohtsuki, T. Separation and purification of 99mTc from 99Mo produced by electron linear accelerator. J. Radioanal. Nucl. Chem. 2017, 311, 1361–1366. [Google Scholar] [CrossRef]
  21. Sebastian, V.; Arruebo, M.; Santamaria, J. Reaction Engineering Strategies for the Production of Inorganic Nanomaterials. Small 2014, 10, 835–853. [Google Scholar] [CrossRef] [PubMed]
  22. Jadhav, A. Wet Chemical Methods for Nanop article Synthesis. In Chemical Methods for Processing Nanomaterials, 1st ed.; Singh, V.N., Ed.; CRC Press: Boca Raton, FL, USA, 2021; pp. 49–58. [Google Scholar]
  23. Baig, N.; Kammakakam, I.; Falath, W. Nanomaterials: A review of synthesis methods, properties, recent progress, and challenges. Mater. Adv. 2021, 2, 1821–1871. [Google Scholar] [CrossRef]
  24. Priya, S.D.; Nesaraj, A.S.; Selvakumar, A.I. Facile wet-chemical synthesis and evaluation of physico-chemical characteristics of novel nanocrystalline NdCoO3-based perovskite oxide as cathode for LT-SOFC applications. Bull. Mater. Sci. 2021, 44, 115. [Google Scholar] [CrossRef]
  25. Gan, Y.X.; Jayatissa, A.H.; Yu, Z.; Chen, X.; Li, M. Hydrothermal Synthesis of Nanomaterials. J. Nanomater. 2020, 2020, 8917013. [Google Scholar] [CrossRef]
  26. Kigozi, M.; Ezealigo, B.N.; Onwualu, A.P.; Dzade, N.Y. Hydrothermal Synthesis of Metal Oxide Composite Cathode Materials for High Energy Application. In Chemically Deposited Nanocrystalline Metal Oxide Thin Films: Synthesis, Characterizations, and Applications, 1st ed.; Ezema, F.I., Lokhande, C.D., Rajan, J., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 489–508. [Google Scholar] [CrossRef]
  27. 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]
  28. Madkour, M.; Allam, O.G.; Abdel Nazeer, A.; Amin, M.O.; Al-Hetlani, E. CeO2-based nanoheterostructures with p–n and n–n heterojunction arrangements for enhancing the solar-driven photodegradation of rhodamine 6G dye. J. Mater. Sci. Mater. Electron. 2019, 30, 10857–10866. [Google Scholar] [CrossRef]
  29. Zhu, H.; Chen, Y.; Wang, Z.; Liu, W.; Wanga, L. Catalytic oxidation of CO over mesoporous copper-doped ceria catalysts via a facile CTAB-assisted synthesis. RSC Adv. 2018, 8, 14888–14897. [Google Scholar] [CrossRef] [Green Version]
  30. Ivanets, A.; Kitikova, N.; Shashkova, I.; Radkevich, A.; Stepanchuk, T.; Maslova, M.; Mudruk, N. One-Stage Adsorption Treatment of Liquid Radioactive Wastes with Complex Radionuclide Composition. Water Air Soil Pollut. 2020, 231, 144. [Google Scholar] [CrossRef]
  31. Metwally, S.S.; Attallah, M.F. Impact of surface modification of chabazite on the sorption of iodine and molybdenum radioisotopes from liquid phase. J. Mol. Liq. 2019, 290, 111237. [Google Scholar] [CrossRef]
  32. Hashem, A.; Sanousy, M.A.; Mohamed, L.A.; Okoye, P.U.; Hameed, B.H. Natural and Low-Cost P. turgidum for Efficient Adsorption of Hg(II) Ions from Contaminated Solution: Isotherms and Kinetics Studies. J. Polym. Environ. 2021, 29, 304–312. [Google Scholar] [CrossRef]
  33. Maamoun, I.; Eljamal, R.; Falyouna, O.; Bensaida, K.; Sugihara, Y.; Eljamal, O. Insights into kinetics, isotherms and thermodynamics of phosphorus sorption onto nanoscale zero-valent iron. J. Mol. Liq. 2021, 328, 115402. [Google Scholar] [CrossRef]
  34. Lagergren, S.K. About the theory of so-called adsorption of soluble substances. K. Sven. Vetensk. Handl. 1898, 24, 1–39. [Google Scholar]
  35. Jasper, E.E.; Ajibola, V.O.; Onwuka, J.C. Nonlinear regression analysis of the sorption of crystal violet and methylene blue from aqueous solutions onto an agro-waste derived activated carbon. Appl. Water Sci. 2020, 10, 132. [Google Scholar] [CrossRef]
  36. Hashem, A.; Badawy, S.M.; Farag, S.; Mohamed, L.A.; Fletcher, A.J.; Taha, G.M. Non-linear adsorption characteristics of modified pine wood sawdust optimised for adsorption of Cd(II) from aqueous systems. J. Environ. Chem. Eng. 2020, 8, 103966. [Google Scholar] [CrossRef]
  37. Chatterjee, R.; Majumder, C. Modelling of adsorption process in industrial wastewater treatment—A review. J. Indian Chem. Soc. 2019, 96, 499–506. [Google Scholar] [CrossRef]
  38. Pholosi, A.; Naidoo, E.B.; Ofomaja, A.E. Intraparticle diffusion of Cr(VI) through biomass and magnetite coated biomass: A comparative kinetic and diffusion study. S. Afr. J. Chem. Eng. 2020, 32, 39–55. [Google Scholar] [CrossRef]
  39. McKay, G.; Poots, V.J.P. Kinetics and diffusion processes in colour removal from effluent using wood as an adsorbent. J. Chem. Technol. Biotechnol. 1980, 30, 279–292. [Google Scholar] [CrossRef]
  40. Weber, W.J.; Morris, J.C. Kinetics of Adsorption on Carbon from Solution. J. Sanit. Eng. Div. 1963, 89, 31–59. [Google Scholar] [CrossRef]
  41. Hassan, H.S.; Elmaghraby, E.K. Retention behavior of cesium radioisotope on poly (acrylamido-sulfonic acid) synthesized by chain polymerization. Appl. Radiat. Isot. 2019, 146, 40–47. [Google Scholar] [CrossRef] [PubMed]
  42. Khandaker, S.; Toyohara, Y.; Saha, G.C.; Awual, M.R.; Kuba, T. Development of synthetic zeolites from bio-slag for cesium adsorption: Kinetic, isotherm and thermodynamic studies. J. Water Process. Eng. 2020, 33, 101055. [Google Scholar] [CrossRef]
  43. Benmessaoud, A.; Nibou, D.; Mekatel, E.H.; Amokrane, S. A Comparative Study of the Linear and Non-Linear Methods for Determination of the Optimum Equilibrium Isotherm for Adsorption of Pb2+ Ions onto Algerian Treated Clay. Iran. J. Chem. Chem. Eng. 2020, 39, 153–171. [Google Scholar] [CrossRef]
  44. Mahmoud, M.E.; Saad, E.A.; El-Khatib, A.M.; Soliman, M.A.; Allam, E.A. Adsorptive removal of radioactive isotopes of cobalt and zinc from water and radioactive wastewater using TiO2/Ag2O nanoadsorbents. Prog. Nucl. Energy 2018, 106, 51–63. [Google Scholar] [CrossRef]
  45. Karimzadeh, L.; Lippold, H.; Stockmann, M.; Fischer, C. Effect of DTPA on europium sorption onto quartz—Batch sorption experiments and surface complexation modeling. Chemosphere 2020, 239, 124771. [Google Scholar] [CrossRef]
  46. Luo, W.; Chen, J.; Lin, H.; Ye, X. Biomass base membrane with phenol hydroxy-amino group for ultraselective adsorption of radioactive Co(II) in trace concentration. Sep. Purif. Technol. 2021, 272, 118878. [Google Scholar] [CrossRef]
  47. Awual, M.R.; Khraisheh, M.; Alharthi, N.H.; Luqman, M.; Islam, A.; Karim, M.R.; Rahman, M.M.; Khaleque, M.A. Efficient detection and adsorption of cadmium(II) ions using innovative nano-composite materials. Chem. Eng. J. 2018, 343, 118–127. [Google Scholar] [CrossRef]
  48. Zsigmondy, R. Über amikroskopische Goldkeime. I. Z. Für Phys. Chem. 1906, 56U, 65–76. [Google Scholar] [CrossRef] [Green Version]
  49. 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]
  50. Temkin, M.; Pyzhev, V. Recent modifications to Langmuir isotherms. Acta Physiochim. URSS 1940, 12, 217–225. [Google Scholar]
  51. Chen, T.; Wang, Q.; Lyu, J.; Bai, P.; Guo, X. Boron removal and reclamation by magnetic magnetite (Fe3O4) nanoparticle: An adsorption and isotopic separation study. Sep. Purif. Technol. 2020, 231, 115930. [Google Scholar] [CrossRef]
  52. Marcu, C.; Balla, A.; Balázs, J.Z.S.; Lar, C. Adsorption Isotherms and Thermodynamics for Chromium(VI) Using an Anion Exchange Resin. Anal. Lett. 2021, 54, 1783–1793. [Google Scholar] [CrossRef]
Figure 1. The synthesis protocol of CeO2 NPs using wet chemical precipitation and hydrothermal modification approaches.
Figure 1. The synthesis protocol of CeO2 NPs using wet chemical precipitation and hydrothermal modification approaches.
Nanomaterials 12 01587 g001
Figure 2. XRD patterns of CeO2 NPs at different temperatures prepared via (a) wet chemical precipitation method; (b) hydrothermal modification.
Figure 2. XRD patterns of CeO2 NPs at different temperatures prepared via (a) wet chemical precipitation method; (b) hydrothermal modification.
Nanomaterials 12 01587 g002
Figure 3. N2 adsorption–desorption isotherms of CeO2 NPs at different temperatures prepared via (a) wet chemical precipitation method; (b) hydrothermal modification.
Figure 3. N2 adsorption–desorption isotherms of CeO2 NPs at different temperatures prepared via (a) wet chemical precipitation method; (b) hydrothermal modification.
Nanomaterials 12 01587 g003
Figure 4. FE-SEM images of CeO2 NPs at different temperatures prepared via hydrothermal modification (HT) and wet chemical precipitation method (CP).
Figure 4. FE-SEM images of CeO2 NPs at different temperatures prepared via hydrothermal modification (HT) and wet chemical precipitation method (CP).
Nanomaterials 12 01587 g004
Figure 5. Effect of solution pH on (a) the Mo uptake on synthesized CeO2 NPs (Co = 50 mg·L−1, V/m = 100 mL·g−1, and temperature = 25 ± 1 °C); (b) speciation of molybdenum.
Figure 5. Effect of solution pH on (a) the Mo uptake on synthesized CeO2 NPs (Co = 50 mg·L−1, V/m = 100 mL·g−1, and temperature = 25 ± 1 °C); (b) speciation of molybdenum.
Nanomaterials 12 01587 g005
Figure 6. Non-linear fitting of kinetics models (Lagergren-first-order, pseudo-second-order, and Elovich) for the sorption of 99Mo on (a) CP-2; (b) HT-1.
Figure 6. Non-linear fitting of kinetics models (Lagergren-first-order, pseudo-second-order, and Elovich) for the sorption of 99Mo on (a) CP-2; (b) HT-1.
Nanomaterials 12 01587 g006
Figure 7. Diffusion plots of (a) McKay; (b) Weber and Morris; (c) Bangham for the sorption of carrier-added 99Mo on CP-2 and HT-1.
Figure 7. Diffusion plots of (a) McKay; (b) Weber and Morris; (c) Bangham for the sorption of carrier-added 99Mo on CP-2 and HT-1.
Nanomaterials 12 01587 g007aNanomaterials 12 01587 g007b
Figure 8. Effect of initial molybdate concentration on the uptake percent and equilibrium sorption capacity (Qe) of carrier-added 99Mo on CP-2 and HT-1 (pH = 3, V/m = 100 mL·g−1, t = 24 h, and temperature = 25 ± 1 °C).
Figure 8. Effect of initial molybdate concentration on the uptake percent and equilibrium sorption capacity (Qe) of carrier-added 99Mo on CP-2 and HT-1 (pH = 3, V/m = 100 mL·g−1, t = 24 h, and temperature = 25 ± 1 °C).
Nanomaterials 12 01587 g008
Figure 9. Equilibrium isotherms (Langmuir, Freundlich, and Temkin) of carrier-added 99Mo on (a) CP-2; (b) HT-1.
Figure 9. Equilibrium isotherms (Langmuir, Freundlich, and Temkin) of carrier-added 99Mo on (a) CP-2; (b) HT-1.
Nanomaterials 12 01587 g009
Figure 10. Van’t Hoff plot for the sorption of carrier-added 99Mo on CP-2 and HT-1 (Co = 1000 mg·L−1, pH = 3, V/m = 100 mL·g−1, and t = 24 h).
Figure 10. Van’t Hoff plot for the sorption of carrier-added 99Mo on CP-2 and HT-1 (Co = 1000 mg·L−1, pH = 3, V/m = 100 mL·g−1, and t = 24 h).
Nanomaterials 12 01587 g010
Table 1. BET surface area (SBET), average pore volume and diameter, crystallite size, and isoelectric point (pHIEP) for CeO2 NPs prepared at different conditions.
Table 1. BET surface area (SBET), average pore volume and diameter, crystallite size, and isoelectric point (pHIEP) for CeO2 NPs prepared at different conditions.
SampleSBET,
(m2·g1)
Pore Volume, (cm3·g1)Pore Size,
(nm)
Crystallite Size, (nm)pHIEP
CP-1150.20.0852.262.506.91
CP-2175.30.0572.384.626.92
CP-397.40.0312.335.666.94
HT-1187.20.1282.753.546.99
HT-2179.80.1222.713.617.00
HT-3114.70.0782.713.507.02
Table 2. Kinetic parameters for the sorption of carrier-added 99Mo on CP-2 and HT-1.
Table 2. Kinetic parameters for the sorption of carrier-added 99Mo on CP-2 and HT-1.
Kinetic ModelParameterCP-2HT-1
Lageregen-first-orderqe,1 (mg·g−1) 4.686 4.881
K1 (min−1) 0.171 0.170
R20.9900.989
Pseudo-second-orderqe,2 (mg·g−1)4.751 4.930
K2 (g/mg·min)0.1280.127
R20.9910.990
Elovichα (mg/g·min) 4.724 × 10 16 3.877 × 1016
β (g·mg−1)9.874 9.48458
R20.999 0.999
Table 3. Diffusion parameters for the sorption of carrier-added 99Mo on CP-2 and HT-1.
Table 3. Diffusion parameters for the sorption of carrier-added 99Mo on CP-2 and HT-1.
Diffusion ModelParameterCP-2HT-1
McKay
(Film diffusion)
  k f   (min−1)0.0130.004
intercept−2.218−2.425
R2 0.992 0.968
Weber and Morris
(Intra-particle diffusion)
k ip   (mg/g·min1/2)0.0360.019
Intercept (C) *4.2244.531
R20.9770.946
Bangham
(Pore diffusion)
σ **0.0340.024
intercept−2.462−2.431
R20.9670.980
* The intercept is influenced by the boundary layer thickness. ** σ value should be <1.
Table 4. Isotherm parameters and error functions calculations for the sorption of carrier-added 99Mo on CP-2 and HT-1.
Table 4. Isotherm parameters and error functions calculations for the sorption of carrier-added 99Mo on CP-2 and HT-1.
Isotherm ModelParameterCP-2HT-1
LangmuirQmax (mg·g1) 96.083 109.114
KL (L·mg1)0.0050.012
R2 0.958 0.983
X241.989138.843
APE28.7536.032
RMSE7.19011.185
FreundlichKF (mg1-n·Ln·g1) 8.668 15.078
n 3.350 3.981
R2 0.986 0.989
X23.8884.678
APE11.44215.304
RMSE4.0605.524
Temkinbt0.2600.266
At (L·g1) *1.535 8.199
R2 0.794 0.756
X232.40770.355
APE81.781137.723
RMSE15.91728.175
* At is the equilibrium-binding constant that corresponds to the maximum binding energy.
Table 5. Thermodynamic parameters for the sorption of carrier-added 99Mo on CP-2 and HT-1.
Table 5. Thermodynamic parameters for the sorption of carrier-added 99Mo on CP-2 and HT-1.
CeO2 NPsTemperature
(K)
ΔG°
(kJ·mol−1)
ΔH°
(kJ·mol−1)
ΔS°
(kJ·mol−1·K−1)
R2
CP-2298−12.33823.6200.1210.970
313−14.149
323−15.355
333−16.562
HT-1298−14.6767.9590.0760.867
313−15.816
323−16.575
333−17.335
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.; 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. https://doi.org/10.3390/nano12091587

AMA Style

Nawar MF, El-Daoushy AF, 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(9):1587. https://doi.org/10.3390/nano12091587

Chicago/Turabian Style

Nawar, Mohamed F., Alaa F. El-Daoushy, Metwally Madkour, and Andreas Türler. 2022. "Sorption Profile of Low Specific Activity 99Mo on Nanoceria-Based Sorbents for the Development of 99mTc Generators: Kinetics, Equilibrium, and Thermodynamic Studies" Nanomaterials 12, no. 9: 1587. https://doi.org/10.3390/nano12091587

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