Next Article in Journal
Extreme Rainfall and Flood Risk Prediction over the East Coast of South Africa
Previous Article in Journal
A Preliminary Approximation to Microbiological Beach Sand Quality along the Coast of the Department of Atlántico (Caribbean Sea of Colombia): Influence of the Magdalena River
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Removal of Molybdenum from Contaminated Groundwater Using Carbide-Derived Carbon

Qatar Environment and Energy Research Institute, Hamad Bin Khalifa University, Qatar Foundation, Doha P.O. Box 34110, Qatar
*
Authors to whom correspondence should be addressed.
Water 2023, 15(1), 49; https://doi.org/10.3390/w15010049
Submission received: 9 November 2022 / Revised: 8 December 2022 / Accepted: 14 December 2022 / Published: 23 December 2022
(This article belongs to the Section Wastewater Treatment and Reuse)

Abstract

:
In the present work, the removal of Mo from aqueous solutions and real groundwater by using the novel high-surface-area adsorbent carbide-derived carbon (CDC) was performed. The adsorbent was characterized using X-ray photoelectron spectroscopy (XPS), X-ray powder diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), energy-dispersive X-ray spectroscopy (EDS), Brunauer, Emmett, and Teller (BET) surface area analysis, and Fourier-transform infrared spectroscopy (FTIR). The effect of the operational parameters (contact time, CDC loading, Mo concentration, and pH) on the adsorptive performance of the sorbent in the batch adsorption mode was studied. The experimental work revealed that the adsorption of Mo onto CDC is a very fast process and provides 99% Mo removal in less than 30 min. The adsorption process was pH-dependent, achieving the maximum adsorptive removal at a pH range of 3–5. The highest adsorption capacity corresponded to 16.24 mg/g at a Mo concentration of 10 ppm, adsorbent loading of 0.6 g/L, and pH 3. Four models were used to analyze the adsorption isotherms of Mo onto CDC, which were Freundlich, Langmuir, Temkin, and Sips. The obtained adsorption results were also processed using four adsorption kinetic models: intra-particle diffusion, Elovich, second-order, and pseudo-first-order. The adsorption of Mo onto CDC was found to fit the Freundlich isotherm model, as confirmed by the highest R2 values (0.9118) and lowest SSE (0.4777), indicating the heterogeneous multilayer adsorption of Mo onto CDC. Likewise, the experimental adsorption data were found to be more consistent with the pseudo-second-order model. The main adsorption mechanisms contributing to Mo adsorption were found to be electrostatic interactions and ligand–ligand exchange, in addition to surface complexation or ion exchange between Mo ions and oxygen-containing groups on the CDC’s surface. Moreover, the removal efficiency under acidic conditions (pH: 3) was found to be stable and high (>99%), regardless of the Mo concentration (0.5–10 ppm) due to the characteristic PZC corresponding to CDC (pH 9.9). A performance test of the CDC using both real groundwater and GW spiked with 570 µg/L Mo showed an almost complete removal of Mo from GW. The regeneration tests confirmed that adsorbed Mo can be recovered from CDC by pH adjustment and the regenerated CDC can be reused.

1. Introduction

Molybdenum is an important microelement for human beings but might cause serious health effects when people are exposed to elevated levels of Mo in food and drinking water. According to the WHO, some of the potential adverse effects associated with the consumption of or exposure to high levels of Mo are tremors, fatigue, weakness, and joint pain [1]. The World Health Organization (WHO)’s recommended value for Mo in drinking water is 70 µg/L [2].
The reported Mo levels in groundwater samples collected from some parts of the world were found to exceed the WHO recommended value [3]. For instance, the Mo level in groundwater samples collected from La Pampa province in Argentina was as high as 89.2 μg/L [4]. Likewise, the survey study conducted by the United States national water-quality-assessment program on 3063 groundwater samples reported a maximum Mo level of 4700 μg/L. Similarly, the analysis of the Mo level in groundwater in Maryland (USA) reported a maximum Mo level of 143 μg/L [5]. In Qatar, the field sampling and analysis of 205 GW samples have shown elevated levels of molybdenum, with a mean of 26.9 µg/L and a maximum level of 103 µg/L [6]. The hydrogeochemical analysis of the samples found that the Mo level in Qatar reached 293 µg/L, 40 groundwater samples exceeded the US Environmental Protection Agency guidelines for short-term use for irrigation water, and 18 exceeded the USEPA guidelines for long-term use for irrigation water [7]. Industry is the main anthropogenic source of groundwater contamination in Qatar with molybdenum. Molybdenum is an important element that is used in manufacturing steel alloys, as it is added to increase their tensile strength and toughness, in addition to their corrosion resistance [8]. Moreover, the gigantic scale of the oil and gas industry in Qatar is also responsible for contaminating groundwater, as it uses Mo as a catalyst for the desulfurization process [6,9]. Mo is also used in many other industries, such as fertilizers, pigments, corrosion inhibitors, and lubricating oils [10,11,12].
Due to the shallow GW aquifers in Qatar (max 150 m deep), the elevated levels of Mo in GW can be attributed to anthropogenic sources, such as industry and agriculture [6]. The overuse of fertilizers (rich in Mo) in agriculture was found to be responsible for GW contamination with Mo. The other major anthropogenic source of GW contamination with Mo is the oil and gas industry. Mo is widely used as a catalyst to speed up the removal of sulfur from oil and gas [9]. The gigantic-scale production facilities for natural gas in Qatar, which make it the world’s largest producer of natural gas, might be one of the causes of the elevation in the Mo level in the GW of Qatar.
One of the most widely adopted techniques for groundwater treatment is adsorption due to the high simplicity of operation, removal efficiency, and cost-effectiveness. However, the common sorbents used for water treatment, such as activated carbon (AC), are not highly efficient for Mo removal from water. For instance, Xiong [13] compared the removal efficiency of Mo from aqueous solutions using AC and crosslinked persimmon tannin (CPT) adsorption gel. The experimental testing of both adsorbents showed that AC could not be used to remove Mo efficiently. Likewise, the performance testing of AC to remove Mo ions from water by the incorporation of oxygen-containing groups, such as Al2O3, onto AC showed that neither AC nor AC–Al2O3 were able to notably remove Mo from water [14]. The inability of AC and AC-Al2O3 to remove Mo despite the introduction of oxygen-containing groups was attributed to the fact that AC has a low point of zero charge (PZC) (2.73), which makes AC negatively charged at all pH values above 2.73. Even when AC was modified with oxygen-containing groups, AC–Al2O3 failed to adsorb Mo, as the PC of the AC–Al2O3 was 6.62, which was lower than the operating conditions of the reported adsorption experiments. Therefore, the search for novel sorbents with improved properties and high sorption capacities toward Mo is still an ongoing field of research.
Carbide-derived carbon (CDC) is a new carbon-based material. Owing to its favorable properties, such as its high surface area, electrical conductivity, and tunable nano-porous structure, CDC has caught the interest of material scientists around the world in numerous applications, such as supercapacitors, carbon capture, gas storage, etc. CDC was reported to have a higher surface area when compared with other carbon-based materials, such as CNTs and AC [15]. Previously, CDC was used for hydrogen capture [16,17], methane storage [18], electrolyte adsorption [19], and phosphate adsorption. The adsorptive removal of ibuprofen and phosphate from water was reported and showed that CDC achieved fast and efficient adsorptive removal of these pollutants from a model aqueous solution and treated sewage effluent [20,21].
The aim of this work is to, for the first time, investigate the adsorption of Mo from water by using CDC under various conditions, such as different pH levels, temperatures, CDC dosages, agitation speeds, and Mo feed concentrations.

2. Materials and Methods

2.1. Materials

The Mo solutions were prepared from a 1000 mg/l Mo standard solution (traceable from NIST), which was purchased from Merck (Darmstadt, Germany). Distilled water was used for the preparation of the Mo solutions.
The CDC (type: nanopore U) used in the current study was obtained from Carbon-Ukraine Company (Kiev, Ukraine). The CDC was synthesized from titanium carbide (TiC) powder, which was used as a precursor. The TiC powder was first heated to 800 °C in a stream of flowing argon gas—this step was followed by the treatment of the TiC with chlorine and hydrogen gases at 800 °C for 6.5 h and at 600 °C for 2 h, respectively.

2.2. Methods

The CDC microstructure and surface morphology were inspected by a field-emission scanning electron microscope (FE-SEM, QUANTA FEG 650, Thermo Fisher Scientific, Billerica, MA, USA) connected to an energy-dispersive spectrometer (EDS, Bruker Xflash 6l60, Germany), which was used for the elemental analysis of the CDC before and after Mo adsorption. The TEM images were acquired by a high-resolution transmission electron microscope (HR-TEM) (FEI Talos 200X, Thermo Fisher Scientific, Billerica, MA, USA) functioning at 200 kV. For TEM analysis, the CDC powder was sonicated in isopropyl alcohol for 5 min; then, 20 μL of the homogenized solution was placed over a 300-mesh lacey/copper carbon grid and dried at room temperature before imaging. Fourier-transform infrared spectroscopy (FT-IR, type; Thermo Fisher Scientific Nicolet iS10, USA) was utilized to analyze the developed chemical groups on the CDC’s surface, the spectra range was varied between 600 and 4000 cm−1. A Brunauer–Emmett–Teller (BET) surface area analyzer (Micromeritics, ASAP 2020, Norcross, GA, USA) was utilized to determine the surface area of the CDC. Barrett, Joyner, and Halenda (BJH) was utilized for the pore-size-distribution evaluation by the BET equipment. The BET was degassed at a temperature of 250 °C for 4 h, then N2 gas at 77 K was used as an adsorbate for the BET and BJH analyses. The zeta potential measurements with CDC particles were conducted using a Zetasizer (type; Nano ZS 90, Malvern Instruments Ltd., UK) functioning with an internal laser of 4.0 MW; the zeta potential values were evaluated at room temperature (25 °C) within a pH range of 3–10.

2.3. Mo Batch Adsorption Experiments

The adsorption testing was accomplished via batch adsorption tests. In all adsorption experiments, the CDC was added to 100 mL of the Mo solution in 250 mL conical flasks and shaken under the required conditions by using a Grant OLS Aqua Pro temperature-controlled shaker (Model OLS26, UK). While adsorption was taking place, samples for analysis were collected over time, filtered through a 0.22 micron PTFE membrane filter, and their Mo content was analyzed.
The determination of the level of Mo in water samples was conducted using inductively coupled plasma–optical emission spectrometry (5800 ICP OES, Agilent Technologies, Santa Clara, CA, USA) in accordance with an ISO-certified procedure (ISO 11885:2007). The calibration curve of the ICP-OES was created by using Mo solutions at different concentrations prepared from 1000 ppm Mo standard stock calibration solution (Agilent Technologies, USA).
The Mo adsorption capacity qe (mg of Mo/g of CDC) and removal efficiency from water were estimated according to Equations (1) and (2);
q e = ( C 0 C e ) × V m
R e m o v a l   E f f i c i e n c y   ( % ) = ( C 0 C e ) C 0 × 100
where C0 and Ce are the initial and final Mo concentrations (mg/L), respectively, V is the volume of the Mo solution (L), and m is the mass of the CDC (g). Each adsorption test was examined three times and the mean values of these trials were illustrated.
The effect of CDC dosage was tested using a CDC dosage ranging between 0.2 and 1.2 g/L. The effect of the pH level on the adsorption of Mo was investigated within a pH range of 3–10 by adjusting the pH of the Mo solution by adding aqueous HCl and NaOH solutions. The agitation speed range studied in the present work was 100–200 rpm. The effect of temperature in addition to the adsorption thermodynamics were tested over a temperature range of 25–45 °C (298.15–318.25 K). Moreover, the adsorption kinetics experiments were conducted by evaluating the equilibrium concentration of Mo after time intervals ranging between 1 min and 24 h.
The efficiency of CDC in removing Mo from multicomponent aqueous solutions containing heavy metals (As, Be, Cd, Cr, Mo, Ni, Pb, Se, and Ti), as well as at different ionic strengths of the aqueous solutions (0, 0.001, and 0.01 M NaCl) was investigated.
Finally, performance testing of the CDC using real GW was conducted by spiking GW with 250 and 500 µg/L molybdenum. The adsorption experiments and determination of the Mo level were conducted following the same procedure described above.

2.4. Adsorption Kinetics

The kinetics of Mo adsorption onto CDC were assessed by identifying the adsorption uptake as well as the rate-limiting step. The obtained adsorption results were processed as per the following adsorption kinetic models: intra-particle diffusion, Elovich, second-order, and pseudo-first-order [21]. Table 1 lists the mathematical expressions in addition to the linearized correlations of the used kinetic models.
The half-life (t1/2) or time required to adsorb half of the adsorbate in the solution was calculated using Equation (3):
t 1 / 2 = 1 K 2 q e

2.5. Adsorption Isotherms

The adsorption isotherm experiments were carried out at various Mo concentrations over a concentration range of 1–100 mg/L. The experiments were conducted for 6 h at pH 6, ambient temperature of 25 °C, and agitation speed of 150 RPM.
Four models were used to analyze the adsorption isotherms of Mo onto CDC: Freundlich, Langmuir, Temkin, and Sips [22,23,24] (Table 2).
The Langmuir model is used for the monolayer adsorption of an adsorbate onto an adsorbent’s surface, and assumes that all of the available adsorption sites are energetically homogeneous and the sorption activation energy for any adsorbate molecule is the same [25].
The separation factor or the Langmuir model’s dimensionless constant can be determined from Equation (4):
R L = 1 1 + K L C 0
where RL represents the adsorption favorability; if RL > 1, the adsorption is unfavorable; if RL = 0, it is irreversible; if RL = 1, it is linear; and for favorable adsorption, RL is in the range 0 to 1 [10].
The Freundlich model w used to describe heterogeneous, reversible, multi-layered, and non-ideal adsorption. The heterogeneity of the Freundlich model indicates that the adsorption energy varies according to the adsorption site within the adsorbent [26]. The Sips model, on the other hand, is reported to fit heterogeneous adsorption systems [27].
In the present work, a Microsoft Excel solver was used to fit the above-mentioned isotherm models by reducing the residual sum of squares of errors (RSS). RSS is used to determine the amount of discrepancy or variance between an experimentally measured adsorption capacity and the value predicted by isotherm models (through linear regression). Equation (5) was used to determine RSS.
R S S = i = 1 n ( q i ( E x p e r i m e n t a l ) q i ( M o d e l ) ) 2  
The total sum of squares (TSS) is similar to RSS (Equation (5)); however, in TSS, qi (model) is replaced with qi (mean). The standard deviation (SD) is used to measure the amount of dispersion associated with a particular group of data points. Equation (6) was used to estimate the value of SD, while Equation (7) was used to determine the value of the coefficient of determination (or R2).
S D = i = 1 n [ q i ( E x p e r i m e n t a l ) q i ( M o d e l ) ] 2 n 1
R 2 = 1 R S S T S S

2.6. Adsorption Thermodynamics

The evaluation of the adsorption thermodynamics can be achieved by calculating the thermodynamic parameters at various temperatures, as the adsorption process is temperature-dependent [28]. The importance of determining the thermodynamic parameters is to provide insight into the spontaneity of the adsorption process. In the present work, the adsorption thermodynamics parameters were evaluated over a temperature range of 25–45 °C. The Gibbs free energy (∆G0) at any temperature was estimated using Nernst’s equation [29]:
Δ G 0 = R T ln k e q  
where R is the universal gas constant (≈8.314 J. K−1. Mol−1), T is temperature (in Kelvin), and Keq is the adsorption equilibrium constant, which can be estimated from Equation (9).
K e q = q e c e  
Moreover, the Gibbs free energy (∆G0) can be estimated from the Gibbs–Helmholtz equation, which enables the calculation of the Gibbs free energy, given the enthalpy and entropy of the system.
Δ G 0 = Δ H 0 T . Δ S 0  
Substituting Equations (9) and (10) into the Gibbs–Helmholtz equation produces the following equation:
ln k e q = Δ S 0 R Δ H 0 R T
The simplified version of the Gibbs–Helmholtz equation can be used to calculate the changes in entropy (∆S0) and enthalpy (∆H0) by plotting 1 T on the x-axis and lnkeq on the y-axis. The enthalpy change can be obtained from the slope of the plot (or Δ H 0 R ) and the entropy change can be calculated from the y-intercept (or Δ S 0 R ).
The activation energy of the adsorption of Mo onto CDC was calculated from Equation (12):
E a = R T   ln ( k 0 k a e )
where k0 stands for the rate constant at temperature T and kae is the pseudo-second-order model constant.

2.7. Regeneration of CDC Adsorbent

The regeneration of the CDC adsorbent was conducted by placing Mo-saturated CDC samples in aqueous solutions at pH 11 and shaking the suspensions for 2 h. The desorption of the Mo ions from the CDC was evaluated by analyzing the Mo content in water samples following the same procedure described in Section 2.3 and by the XPS spectra of the CDC samples before and after the regeneration procedure.

3. Results and Discussion

The schematic presentation of the experimental work carried out in this study is shown in Figure 1. In the first stage, CDC material will be extensively characterized by using XPS, XRD, SEM, TEM, EDS, BET, and FTIR techniques, followed by the evaluation of the effects of operational parameters, such as the contact time, CDC loading, Mo concentration, and feed pH, on the adsorption process. Next, the adsorption isotherms and kinetics of Mo adsorption on CDC will be studied, followed by employing CDC for the treatment of real groundwater contaminated with Mo. Finally, the evaluation of the possibility of the regeneration of the used CDC by pH adjustment will be performed.

3.1. CDC Characterization

The SEM images were acquired in order to evaluate the size and shape of the CDC. Figure 2 presents the SEM images of the CDC particles at low and high magnifications. It can be observed that the CDC particles had irregular shapes with a wide range of sizes ranging between 200 nm and 2 µm. The CDC particles formed aggregates of irregular shapes. The TEM image is presented in Figure 3 and illustrates that a highly porous carbon material has been developed.
The pore size distribution of the CDC particles was studied by the BJH method, and the results are displayed in Figure 4a. As seen, the CDC contained mainly micropores with a significant peak at 2 nm. It can also be seen that most of the pores were between 1 and 4 nm. The mean pore size according to the BJH adsorption/desorption results was 2.8–3.1 nm. The estimated BET surface area of CDC in the present work was 1,120 m2/g. The BET surface area, along with the pore size, of the CDC were found to be in good agreement with other work reported in the literature, such as Xing [30], who reported a BET surface area and average pore size of pristine CDC of 1216 m2/g and 2.13 nm, respectively. The N2 adsorption–desorption curve of CDC is illustrated in Figure 4b. According to the IUPAC classification, N2 adsorption by CDC follows the type Ia sorption isotherm suggesting an ultra–highly microporous material. Similar curves were obtained for the CDC produced at 800 °C using TiC as a precursor.
Figure 5 depicts the XRD pattern of CDC, which was obtained by a Bruker D8 Advance spectrometer with a scan area of 3–90 2q degs. As seen, the absence of a strong peak indicates the amorphous structure of the CDC.
The zeta potential of the CDC is depicted in Figure 6. The zeta potential vs. pH plot shows that, with the exception of highly alkaline conditions (around pH 10), the zeta potential values of CDC were totally positively charged, with a decreasing overall positive charge along with pH. The point of zero charge (PZC) or point at which the net surface charge of CDC was zero occurred at 9.9. The surface charge of the CDC at the studied pH range in addition to the PZC was found to agree with results reported in the literature [20,21,31,32].
The elemental composition of the CDC before and after the adsorption experiments was evaluated by XPS and EDS analyses. Using the XPS technique, the main constituents of CDC before adsorption were carbon (87.3%), oxygen (6.8%), nitrogen (4.7%), chlorine (0.43%), and silicon (0.36%). No Mo was detected in the CDC sample before adsorption. On the other hand, the Mo content in the CDC sample after adsorption was 1.87%. The confirmation of the presence of Mo in samples using XPS was reported in the literature within the same range of binding energies [33,34,35,36,37]. EDS analysis of the CDC was also conducted prior to and after the adsorption process. Figure 7 shows that carbon is the main constituent of the CDC, besides oxygen and chloride. In addition to these elements, Mo was detected in the CDC samples collected after the adsorption process, as depicted in Figure 6. The SEM elemental mapping (Figure 7) also confirmed Mo adsorption on the CDC.
The corresponding functional groups on the CDC were determined by FTIR analysis. In the FTIR spectrum illustrated in Figure 8, a broad band was detected in the region of 3300–3500 cm−1, which belonged to the O–H stretching vibration, pointing out to the presence of adsorbed water molecules. Two more peaks positioned between 3000 and 2800 cm−1 belonged to the stretching vibration of C–H bonds, while the peaks at 1241 and 1574 cm−1 were linked to C–O–C and C=C stretching bonds, respectively [20,21,38,39,40]. Likewise, the presence of Mo in the FTIR spectra of the CDC samples after adsorption can be observed in Figure 8. The peaks at 597 and 898 cm−1 were reported in the literature to correspond to the stretching vibrations of Mo–O and Mo=O, respectively [41,42]. The Mo peaks in the FTIR spectra confirm the adsorption of Mo onto CDC.

3.2. Effect of Adsorption Time, CDC Dose, and Shaking Speed

The adsorption experiments with CDC were conducted in order to study the effect of contact time on the removal efficiency. To that end, the adsorption time was varied from 0.5 to 24 h while maintaining a feed Mo concentration of 1 ppm and CDC loading of 0.6 g/L. Figure 9 shows the Mo removal efficiency of CDC vs. adsorption time. As seen, the removal efficiency of CDC was found to exceed 99%, even after 0.5 h of adsorption. This can be attributed to the high surface area of CDC and fast adsorption kinetics.
The next adsorption experiments carried out were focused on the effect of the loading rate of CDC on the removal efficiency. Figure 10 shows the removal efficiency of CDC at loading rates ranging between 0.2 and 1.2 g/L. As seen, the removal efficiency of CDC was found to be higher than 99%, even with the loading rate being as low as 0.2 g/L.
Thereafter, the effect of the Mo concentration on the removal efficiency and adsorption capacity was studied. Figure 11 depicts the removal efficiency and adsorption capacity at different Mo concentrations in the feed solutions (1, 5, and 10 ppm). As seen, the removal of Mo from water and the adsorption capacity clearly decreased with increasing Mo concentration due to the saturation of the adsorption sites available on the CDC’s surface. In the present work, the reported removal efficiency and adsorption capacity were found to outperform those of other adsorbents cited in the literature for Mo removal from water (Table 3). For instance, Verbinnen [43] developed a zeolite-supported magnetite adsorbent for the removal of Mo from water. Despite having a lower Mo concentration (0.9 ppm) and higher adsorbent loadings (1–20 g/L), the reported adsorption capacities (18–23 mg/g) after a 24 h residence time were lower than those in the present work. Tu [44] prepared a novel biosorbent made from carbonized pomelo peel modified with nanoscale zero-valent iron and cetyl-trimethyl ammonium bromide for the removal of Mo from water. Despite using a higher loading of the biosorbent (1 g/L), the reported adsorption capacity (37–48 mg/g) was lower than the values observed in the present study. Lian [45] modified activated drinking water residuals by heat treatment at 600 °C followed by HCl treatment in order to increase their adsorption capacity for Mo in water. The produced residuals under the optimum conditions were reported to increase the adsorption capacity from 18.44 mg/g to 39.52 mg/g. Tu [44] synthesized crystalline ZnFe2O4 nanoparticles to remove Mo from water. Although the Mo concentrations reported (2.8–2.87 ppm) were lower than the ones used in the present work and the adsorbent loading was higher (2 g/L), the removal efficiency was as low as 2.5% and the adsorption capacity was 1.4 mg/g.

3.3. Adsorption Isotherms

In the present work, Langmuir, Freundlich, Temkin, and Sips isotherm models were applied to the experimental data. The isotherm models were compared by determining the correlation coefficient (R2) and the sum of the square error (SSE) values by fitting the experimental data to non-linear forms of the isotherm models. Figure 12 shows the experimental adsorption results fitted into Langmuir, Freundlich, Temkin, and Sips isotherm models. The isotherm parameters corresponding to the applied isotherm models are provided in Table 4. As seen from the statistical coefficient values (Table 4), the adsorption of Mo onto CDC can be better represented by the Freundlich isotherm model, as confirmed by the highest R2 values (0.9118) and lowest SSE (0.4777), indicating that heterogeneous multilayer adsorption of Mo occurred on the CDC. The Langmuir model also showed good compliance with the experimental data, with R2 of 0.831, indicating that the homogeneous monolayer adsorption clearly also took place (i.e., the adsorptive removal of Mo by CDC was a combination of a physical monolayer adsorption and a chemical adsorption process). Similar findings were reported for the adsorption of phosphate and surfactants on CDC [20,21,32].
The favorability of the adsorption process can be anticipated by calculating the value of the separation factor (RL) using Equation (4). In the present work, the calculated separation factor was within the range of 0.00003–0.0014, which was between 0 and 1, indicating a favorable adsorption process [48].

3.4. Adsorption Kinetics

The effect of contact time on the adsorptive removal of Mo by CDC is shown in figure. As seen in Figure 13, almost complete removal of Mo was achieved after 30 min of adsorption, indicating a very rapid adsorption process. The experimental data of Mo removal were evaluated by four different kinetic models: the pseudo-first-order, pseudo-second-order, Elovich, and intra-particle diffusion models. Table 5 lists the estimated correlation parameters in addition to the coefficient of determination (R2) for each of the studied models. It was found that the experimental adsorption data were more consistent with the pseudo-second-order model, as indicated by the highest correlation coefficient (R2) of 0.9998 for this model. On the contrary, the intra-particle diffusion and pseudo-first-order models exhibited relatively low correlation coefficients of about 0.6268 and 0.4269, respectively, which indicated the unsuitability of these models to fit the studied adsorption system. It seems that the Elovich model (R2 > 0.9) was in better agreement with the experimental data when compared with the intra-particle diffusion and pseudo-first-order models. The number of available adsorption sites in CDC can be estimated from the 1 β value, which was 0.024 in the present work. By referring to the Elovich model in Table 5, the value of the y-intercept (or 1 β ln ( α β ) ) can be used to estimate the adsorption capacity (qe) at the start of the adsorption experiment, which was found to be 0.881 mg/g. This value was found to agree with the experimental adsorption capacity value at t = 1, which corresponded to 0.878 mg/g. The half-life (t1/2), or time required to adsorb half of the adsorbate in the solution, was calculated using Equation (3). The half-life required for the CDC to adsorb half of the Mo in water was estimated to be 1.7 min. The fast kinetics of Mo adsorption by CDC might be beneficial for Mo removal from water, as it would reduce the required residence/contact time and eventually lower the operation cost.

3.5. Effects of Feed Temperature and Adsorption Thermodynamics

The effect of temperature on the adsorption of Mo onto CDC was investigated by evaluating the adsorption capacity at different feed solution temperatures (25, 35, and 45 °C) and at pH values of 6 and 9. As shown in Figure 14, increasing the temperature was observed to increase the Mo adsorption capacity of CDC. For instance, at pH 6, the adsorption capacity increased from 5.51 to 6.04 mg/g by elevating the feed solution temperature from 25 to 45 °C. The increase in the adsorption capacity of CDC with temperature can be attributed to the higher mass transfer rate of Mo ions, which resulted in faster and more efficient binding with CDC. The increase in the adsorption capacity of CDC with feed temperature in the present work was found to be in good agreement with other work reported in the literature [49,50,51].
The thermodynamic parameters of adsorption, such as the Gibbs free energy (∆G0), entropy (∆S0), and enthalpy (∆H0) were evaluated by using the Nernst Equation (8) in Section 2.6. Table 6 shows the thermodynamic parameters for the adsorption of Mo onto CDC at pH 6 and 9 and at a temperature range of 25–45 °C (298.15–318.25 K). At pH 6, the values of the Gibbs free energy changes were found to be −2.759, −3.243, and −3.727 kJ/mol at 25, 35, and 45 °C, respectively. The negative value associated with the Gibbs free energy at the studied temperatures indicated the spontaneity of the adsorption, whereas the increase in the amount of Gibbs free energy with temperature indicated that the favorability of the adsorption process increased with temperature, which was found to agree with other work in the literature [52]. The range of the Gibbs free energy observed in the present work was between 0 and −20 kJ/mol, which was reported in the literature to account for physisorption, which is a form of adsorption in which the interactions between Mo and CDC are mainly controlled by Van der Waals forces [53]. Conversely, Gibbs free energy values below −40 kJ/mol were reported to account for chemisorption (or chemical adsorption), which is an adsorption process taking place between Mo and particular active sites at the CDC’s surface [54].
On the other hand, the Gibbs free energy change associated with adsorption from basic media (pH 9) was found to demonstrate positive values (3.43–3.55 kJ/mol), indicating the unfavourability of the adsorption process.
Moreover, the entropy change associated with the studied adsorption process was found to correspond to 48.44 and −6.4 J/mol.K at pH 6 and 9, respectively. In general, a positive entropy change in adsorption signifies an increase in the degree of disorder/randomness in the system, as well as the higher affinity of the sorbate towards the sorbent [55,56,57]. The overall signs corresponding to the entropy change at pH 6 and 9 can be used as further evidence to support the experimental data, which showed a highly favorable adsorption process at pH 6 and unfavorable adsorption process at pH 9.
Finally, the estimated enthalpy change was positive and corresponded to 11.685 kJ/mol. The positive enthalpy value here supported the experimental finding that the adsorption process of Mo onto CDC was endothermic due to the existence of more energy in the system after adsorption compared with the system before adsorption [58]. On the other hand, exothermic reactions have been reported to have a negative reaction enthalpy difference due to the existence of less energy in the products compared with reactants as a result of the release of more energy than that absorbed [59].

3.6. Effect of pH

Due to the wide range of oxidation states of Mo that exist in natural waters, the formation of various Mo species in water is highly pH-dependent (Figure 15).
Equations (13)–(19) show some of the main reactions involving Mo species that take place in natural water bodies and affect the overall charge of Mo ions [61].
Mo + 2H2O → MoO2 + 4H+
MoO2 + H2O → MoO(OH)2
MoO2 + 2H2O → HmoO4 + 3H+
Mo + 4H+ → H2MoO2 + H2O
Mo + 2OH → Mo(OH)2+
Mo(OH)2+ +OH → Mo(OH)3
Mo(OH)3 +5OH → MoO42− + 4H2O
Figure 16 shows the removal efficiency vs. feed Mo concentration at three pH levels (3, 6, and 9). As seen, the adsorption of Mo by CDC was found to be pH-dependent. The removal efficiency under acidic conditions (pH: 3) was found to be stable and high (>99%), regardless of the Mo concentration (0.5–10 ppm). This can be attributed to the fact that, under acidic conditions, the surface of CDC was protonated by the available H+ ions in acidic water, which resulted in the formation of the positive charge on the sorbent’s surface that attracted negatively charged Mo species via attractive coulombic forces. This can be confirmed by the zeta potential curve of CDC displayed in Figure 6, which shows positive zeta potential values of the sorbent in the acidic feed solution. Figure 6 shows that, with increasing pH, the zeta potential values of CDC were found to decline until reaching the PZC at 9.9. Such zeta potential behavior of CDC correlated with the reduction in the Mo removal efficiency of CDC at pH 6 and 9. The decrease in the overall positive charge of CDC under neutral and basic pH values was caused by the continuous deprotonation of the sorbent’s surface by the addition of OH ions, which reduced the attractive coulombic forces between CDC and negatively charged Mo ions; hence, the adsorption process became less favorable.
This behavior was found to agree with other work published in the literature due to the variation in the speciation of Mo in water at different pH levels [58,60,61]. For instance, Al-Gaashani [14] tested the adsorption performance of AC modified with aluminum oxide in removing Mo ions from water. The adsorption tests, which were conducted under acidic, neutral, and alkaline conditions, demonstrated that the highest removal occurred under acidic conditions (specifically at pH 2) and that the removal efficiency was found to decrease along with pH until a very low removal efficiency was reached at pH 8. Al-Gaashani [14] correlated the decrease in the removal efficiency with pH due to the presence of negatively charged Mo ions, such as Mo7O246−, Mo7O23OH5−, Mo8O265−, H3Mo6O214−, H3Mo8O284−, Mo7O22(OH)24−, and Mo7O21(OH)33−. Similarly, Goldberg [62] reported the maximum adsorption capacity of some natural adsorbents, such as clay and soils, for Mo ions to exist at low pH levels. It was observed that the adsorption capacity decreased suddenly at pH levels above 5, reaching zero at pH 8.
Interestingly, as seen in Figure 16, at pH 9, the adsorption of Mo was still taking place, despite the repulsive interactions between Mo ions and the CDC’s surface. Clearly, other adsorption mechanisms, such as complexation or the ion exchange of Mo ions with oxygen-containing groups on the CDC’s surface, might also contribute to Mo adsorption [45]. The presence of oxygen-containing groups on the CDC’s surface was confirmed by XPS analysis in this study, as well as in previous work. For example, Xing [30] reported an average oxygen content in the CDC sample of 8.7%. Figure 17 shows the possible mechanisms of the adsorptive removal of Mo by the CDC.
  • Influence of co-existing ions on the adsorption of Mo by CDC
Water treatment in real processes or industry usually involves the treatment of complex multicomponent solutions composed of different ions that might simultaneously adsorb to the available sorption sites on a sorbent’s surface. For instance, some cations, anions, and heavy metal ions available in groundwater might adversely affect the adsorption efficiency and, consequently, the adsorption capacity of the sorbent [63,64,65]. Hence, the effect of competing ions on the Mo adsorption capacity of CDC was studied. Mixed solutions of As, Be, Cd, Cr, Mo, Ni, Pb, Se, and Ti ions with different concentrations (2 and 5 ppm) and pH values (2 and 4) were used during these adsorption experiments. Figure 18 shows the adsorption capacity (mg/g) of CDC toward the studied pollutants after adsorption for 1 h. As expected, the presence of the competing pollutant ions was observed to reduce the Mo adsorption capacity of CDC from 163 mg/g (in the single Mo solution) to 128 mg/g (in the mixed multicomponent solution) under the same experimental conditions. This can be attributed to the competition of the different metal species for the available adsorption spots inside the CDC.
It might be concluded that the competition between Mo and other species greatly depends on their ionic radii. As seen in Figure 19, the adsorption of Mo was more affected in the presence of As than Pb due to the fact that the atomic radius of As (1.1 Å) was closer to that of Molybdate (1.4 Å) than that of Pb (1.8 Å) [66]. The removal of Mo by CDC in the presence of the co-existing metals was found to outperform some of the work published in the literature. For instance, Verbinnen [43] conducted the simultaneous removal of Mo (0.8 mg/L), Sb (1.6 mg/L), and Se (0.208 mg/L) from industrial wastewater by using magnetite. It was found that higher adsorbent loading (20 g/L) and a longer residence time (24 h) were required to treat the wastewater, which was reflected in the drastic reduction in the adsorption capacity of magnetite down to 0.04 mg/g due to the competition of co-existing metals for the adsorption sites.
  • Effect of ionic strength of feed solution
The effect of the salinity of the aqueous solution on the adsorption of Mo was also studied by carrying out adsorption experiments at different NaCl concentrations in the feed solution. Figure 19 shows the Mo ion removal efficiency of CDC from 0, 0.001, and 0.01 M NaCl solutions. It was found that the higher the salinity level, the lower the removal efficiency of CDC. These findings can be explained by the reduction in electrostatic interactions between the Mo ions and adsorption sites on the CDC’s surface due to the shielding effect of the NaCl electrolyte. These findings correlate well with previous data on the impact of the ionic strength of solutions on the adsorption capacity of adsorbents reported in the literature [67,68].

3.7. Testing CDC in Groundwater

A performance test of CDC using real groundwater collected from the northern part of Qatar was conducted in the present work. The concentrations of the main elements in the GW samples were as follows: Ba (109 ppb), Ca (446 ppb), Cu (6.55), Mg (153 ppm), Mo (70 ppb), and Na (688 ppm). Figure 20 shows the removal efficiency of CDC from the unspiked GW sample along with samples spiked with 250 and 500 µg/L at pH 3. As seen, CDC was found to completely remove Mo from the unspiked GW sample (70 ppb). Moreover, the removal efficiencies for the CDC from GW samples spiked with 250 and 500 µg/L were 98.6% and 97.2%, respectively. The adsorption capacity for the removal of CDC from real groundwater had a range of 0.1–0.9 mg/g. These values were found to be higher than the adsorption capacities reported in the literature for the removal of Mo from synthetic water using zeolite-supported magnetite [43], acrylamide triazine polymer [46], and activated carbon [14].

3.8. Regeneration of CDC Adsorbent

The regeneration of the CDC adsorbent was conducted by placing the exhausted sorbent in an aqueous solution at pH 11 and shaking the solution with the sorbent for 2 h. The analysis of the aqueous solution after regeneration by using the ICP-OES method showed that more than 95% of the Mo adsorbed on CDC was released into the water. In addition, Figure 21 shows the XPS spectra of the exhausted CDC sample before and after regeneration. As seen, the peaks related to molybdate ions were lacking in the XPS spectra of the CDC sample after regeneration (Figure 21b). These findings indicate that the exhausted CDC adsorbent could be efficiently regenerated and Mo ions could be recovered from CDC by adjusting the pH of the aqueous solution.
This study shows the potential of using CDC as an efficient sorbent for molybdenum removal from wastewater in the batch mode at the bench scale level; however, future studies should focus on (1) column adsorption studies and (2) the technoeconomic feasibility of employing CDC for the treatment of wastewater. The adsorbents used in industry were reported in the literature to employ the continuous mode due to its outstanding features, such as increased productivity, decreased cost, increased quality, etc. The aim of the present work was to test the proof of concept by investigating the effectiveness of CDC in removing Mo from contaminated GW, which was found to be successful. The focus of future work must be on the continuous mode using adsorption columns; hence, the potential of scaling up to a continuous mode must be carried out using an up-flow fixed bed column and the breakthrough parameters (such as the breakthrough time, exhaust time, breakthrough volume, exhaust volume, adsorption capacity, mass transfer zone height, and moving rate) must be evaluated. Furthermore, technoeconomic feasibility studies must also be conducted in order to compare the benefits of using CDC as an adsorbent for wastewater treatment against the incurred cost. This will enable us to assess whether using this material on a large scale will be economically justified and the returns will overcome the expenditure.

4. Conclusions

For the first time, the adsorption of Mo from aqueous solutions using CDC was studied in a batch adsorption mode under different operating conditions, such as different adsorption times, CDC dosages, agitation speeds, Mo feed concentrations, pH values, and temperatures of the feed solutions. CDC samples before and after Mo adsorption were extensively characterized following SEM, TEM, XRD, XPS, BET, and FTIR methods.
The adsorption of Mo onto CDC was found to be better fit by the Freundlich isotherm model, as confirmed by the highest R2 values (0.9118) and lowest SSE (0.4777), while the experimental adsorption data were found to be more consistent with the pseudo-second-order kinetics adsorption model. Notably, the adsorption kinetics were very fast, and more than 99% removal of Mo from water was reached within a short adsorption time (less than 30 min). The main adsorption mechanisms contributing to Mo adsorption were suggested to be electrostatic interactions, ligand–ligand exchange, and surface complexation or ion exchange between Mo ions and oxygen-containing groups on the CDC surface. It should be mentioned that CDC almost completely removed Mo at a concentration of 1 ppm from water at an adsorbent loading of 0.6 g/L.
It was shown that Mo adsorption on CDC was pH-dependent, achieving the maximum adsorptive removal within a pH range of 3–5. The highest adsorption capacity corresponded to 16.24 mg/g at a Mo concentration of 10 ppm, adsorbent loading of 0.6 g/L, and pH 3. Moreover, the removal efficiency at acidic conditions (pH: 3) was found to be stable and high (>99%), regardless of the Mo concentrations (0.5–10 ppm).
The regenerability of the CDC was evaluated, and it was shown that about 95% of the Mo adsorbed could be recovered from CDC by regeneration at pH 11. Finally, the testing of the CDC using real groundwater showed that CDC could be used as an efficient adsorbent for the removal of Mo from groundwater; however, future work on the continuous fixed-bed-adsorption removal of Mo and techno-economic feasibility analysis are required to prove the efficiency of CDC for water treatment on an industrial scale.

Author Contributions

Conceptualization, V.K.; Methodology, S.S.; Investigation, Y.M. and S.S.; Writing—original draft, Y.M.; Writing—review & editing, J.L. and V.K.; Supervision, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barceloux, D.G.; Barceloux, D. Molybdenum. J. Toxicol. Clin. Toxicol. 1999, 37, 231–237. [Google Scholar] [CrossRef] [PubMed]
  2. Smedley, P.L.; Cooper, D.M.; Lapworth, D.J. Molybdenum distributions and variability in drinking water from England and Wales. Environ. Monit. Assess. 2014, 186, 6403–6416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Smedley, P.L.; Cooper, D.M.; Ander, E.L.; Milne, C.J.; Lapworth, D.J. Occurrence of molybdenum in British surface water and groundwater: Distributions, controls and implications for water supply. Appl. Geochem. 2014, 40, 144–154. [Google Scholar] [CrossRef]
  4. Smedley, P.L.; Nicolli, H.B. Molybdenum distributions and controls in groundwater from the Pampean aquifer of La Pampa Province, Argentina. In Molybdenum and its Compounds: Applications, Electrochemical Properties and Geological Implications; Nova Publisher: New York, NY, USA, 2014. [Google Scholar]
  5. Willis, S.S.; Johannesson, K.H. Controls on the geochemistry of rare earth elements in sediments and groundwaters of the Aquia aquifer, Maryland, USA. Chem. Geol. 2011, 285, 32–49. [Google Scholar] [CrossRef]
  6. Kuiper, N.; Rowell, C.; Shomar, B. High levels of molybdenum in Qatar’s groundwater and potential impacts. J. Geochem. Explor. 2015, 150, 16–24. [Google Scholar] [CrossRef]
  7. Ahmad, A.Y.; Al-Ghouti, M.A.; Khraisheh, M.; Zouari, N. Hydrogeochemical characterization and quality evaluation of groundwater suitability for domestic and agricultural uses in the state of Qatar. Groundw. Sustain. Dev. 2020, 11, 100467. [Google Scholar] [CrossRef]
  8. Olsson, J.; Wallen, B. Experience with a high molybdenum stainless steel in saline environments. Desalination 1983, 44, 241–254. [Google Scholar] [CrossRef]
  9. Zhang, M.; Liu, J.; Yang, J.; Chen, X.; Wang, M.; Li, H.; Zhu, W.; Li, H. Molybdenum-containing dendritic mesoporous silica spheres for fast oxidative desulfurization in fuel. Inorg. Chem. Front. 2019, 6, 451–458. [Google Scholar] [CrossRef]
  10. Rodriguez, A.A.; Tylczak, J.H.; Gao, M.C.; Jablonski, P.D.; Detrois, M.; Ziomek-Moroz, M.; Hawk, J.A. Effect of molybdenum on the corrosion behavior of high-entropy alloys CoCrFeNi2 and CoCrFeNi2Mo0. 25 under sodium chloride aqueous conditions. Adv. Mater. Sci. Eng. 2018, 2018, 908. [Google Scholar] [CrossRef] [Green Version]
  11. Farr, J. Molybdenum disulphide in lubrication. A review. Wear 1975, 35, 1–22. [Google Scholar] [CrossRef]
  12. Löbel, M.; Lindner, T.; Grimm, M.; Rymer, L.-M.; Lampke, T. Influence of Aluminum and Molybdenum on the Microstructure and Corrosion Behavior of Thermally Sprayed High-Entropy Alloy Coatings. J. Therm. Spray Technol. 2021, 31, 1366–1374. [Google Scholar] [CrossRef]
  13. Xiong, Y.; Chen, C.; Gu, X.; Biswas, B.K.; Shan, W.; Lou, Z.; Fang, D.; Zang, S. Investigation on the removal of Mo (VI) from Mo–Re containing wastewater by chemically modified persimmon residua. Bioresour. Technol. 2011, 102, 6857–6862. [Google Scholar] [CrossRef] [PubMed]
  14. Al-Gaashani, R.; Almasri, D.; Shomar, B.; Kochkodan, V. Preparation and properties of novel activated carbon doped with aluminum oxide and silver for water treatment. J. Alloys Compd. 2021, 858, 158372. [Google Scholar] [CrossRef]
  15. Rose, M.; Kockrick, E.; Senkovska, I.; Kaskel, S. High surface area carbide-derived carbon fibers produced by electrospinning of polycarbosilane precursors. Carbon 2010, 48, 403–407. [Google Scholar] [CrossRef]
  16. Naheed, L.; Koppel, M.; Paalo, M.; Alsabawi, K.; Lamb, K.E.; Mac, A.; Gray, E.; Jänes, A.; Webb, C.J. Hydrogen adsorption properties of carbide-derived carbons at ambient temperature and high pressure. Int. J. Hydrog. Energy 2021, 46, 15761–15772. [Google Scholar] [CrossRef]
  17. Fisher, J.E.; Gogotsi, Y.; Yildirim, T. Carbide-Derived Carbons with Tunable Porosity Optimized for Hydrogen Storage; University of Pennsylvania: Philadelphia, PA, USA, 2010. [Google Scholar]
  18. Yeon, S.-H.; Osswald, S.; Gogotsi, Y.; Singer, J.P.; Simmons, J.M.; Fischer, J.E.; Lillo-Ródenas, M.A.; Linares-Solano, Á. Enhanced methane storage of chemically and physically activated carbide-derived carbon. J. Power Sources 2009, 191, 560–567. [Google Scholar] [CrossRef]
  19. Trognko, L.; Lecante, P.; Ratel-Ramond, N.; Rozier, P.; Daffos, B.; Taberna, P.-L.; Simon, P. TiC-carbide derived carbon electrolyte adsorption study by ways of X-ray scattering analysis. Mater. Renew. Sustain. Energy 2015, 4, 17. [Google Scholar] [CrossRef] [Green Version]
  20. Almanassra, I.W.; Kochkodan, V.; Subeh, M.; Mckay, G.; Atieh, M.; Al-Ansari, T. Phosphate removal from synthetic and treated sewage effluent by carbide derive carbon. J. Water Process Eng. 2020, 36, 101323. [Google Scholar] [CrossRef]
  21. Almanassra, I.W.; Kochkodan, V.; Ponnusamy, G.; Mckay, G.; Ali Atieh, M.; Al-Ansari, T. Carbide Derived Carbon (CDC) as novel adsorbent for ibuprofen removal from synthetic water and treated sewage effluent. J. Environ. Health Sci. Eng. 2020, 18, 1375–1390. [Google Scholar] [CrossRef]
  22. Lim, L.; Priyantha, N.; Cheng, H.; Zaidi, N. Parkia speciosa (Petai) pod as a potential low-cost adsorbent for the removal of toxic crystal violet dye. Sci. Bruneiana 2018, 15, 99–106. [Google Scholar] [CrossRef]
  23. Lim, L.; Priyantha, N.; Tennakoon, D.T.B.; Hei Ing, C.; Dahri, M.K.; Suklueng, M. Breadnut peel as a highly effective low-cost biosorbent for methylene blue: Equilibrium, thermodynamic and kinetic studies. Arab. J. Chem. 2014, 175, 18. [Google Scholar] [CrossRef] [Green Version]
  24. Shikuku, V.; Chrispin, K.; Kengara, F. Errors in Parameters Estimation Using Linearized Adsorption Isotherms: Sulfadimethoxine Adsorption onto Kaolinite Clay. Chem. Sci. Int. J. 2018, 23, 1–6. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, C.; Boithias, L.; Ning, Z.; Han, Y.; Sauvage, S.; Sánchez-Pérez, J.-M.; Kuramochi, K.; Hatano, R. Comparison of Langmuir and Freundlich adsorption equations within the SWAT-K model for assessing potassium environmental losses at basin scale. Agric. Water Nonpoint Source Pollut. Manag. Watershed Scale Part II Overseen Dr Brent Cloth. 2017, 180, 205–211. [Google Scholar] [CrossRef]
  26. Alshameri, A.; He, H.; Zhu, J.; Xi, Y.; Zhu, R.; Ma, L.; Tao, Q. Adsorption of ammonium by different natural clay minerals: Characterization, kinetics and adsorption isotherms. Small Size–Big Sci. 3rd Asian Clay Conf. 2018, 159, 83–93. [Google Scholar] [CrossRef]
  27. Foo, K.Y.; Hameed, B.H. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 2010, 156, 2–10. [Google Scholar] [CrossRef]
  28. Ebelegi, A.; Nimibofa, A.; Donbebe, W. Interpretation of Adsorption Thermodynamics and Kinetics. Open J. Phys. Chem. 2020, 10, 166–182. [Google Scholar] [CrossRef]
  29. Nuengmatcha, P.; Mahachai, R.; Chanthai, S. Adsorption Capacity of The As-Synthetic Graphene Oxide for The Removal of Alizarin Red S Dye from Aqueous Solution. Orient. J. Chem. 2016, 32, 1399–1410. [Google Scholar] [CrossRef] [Green Version]
  30. Xing, W.; Liu, C.; Zhou, Z.; Zhou, J.; Wang, G.; Zhuo, S.; Xue, Q.; Song, L.; Yan, Z. Oxygen-containing functional group-facilitated CO 2 capture by carbide-derived carbons. Nanoscale Res. Lett. 2014, 9, 1–8. [Google Scholar] [CrossRef] [Green Version]
  31. Álvarez-Torrellas, S.; Munoz, M.; Gläsel, J.; de Pedro, Z.M.; Domínguez, C.M.; García, J.; Etzold, B.J.M.; Casas, J.A. Highly efficient removal of pharmaceuticals from water by well-defined carbide-derived carbons. Chem. Eng. J. 2018, 347, 595–606. [Google Scholar] [CrossRef]
  32. Almanassra, I.W.; Okonkwo, E.C.; Alhassan, O.; Atieh, M.A.; Kochkodan, V.; Al-Ansari, T. Stability and thermophysical properties test of carbide-derived carbon thermal fluid; a comparison between functionalized and emulsified suspensions. Powder Technol. 2020, 377, 415–428. [Google Scholar] [CrossRef]
  33. Tai, G.; Zeng, T.; Yu, J.; Zhou, J.; You, Y.; Wang, X.; Wu, H.; Sun, X.; Hu, T.; Guo, W. Fast and Large-Area Growth of Uniform MoS2 Monolayers on Molybdenum Foils. Nanoscale 2015, 8, 7226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Fominski, V.; Demin, M.; Nevolin, V.; Fominski, D.; Romanov, R.; Gritskevich, M.; Smirnov, N. Reactive Pulsed Laser Deposition of Clustered-Type MoSx (x~2, 3, and 4) Films and Their Solid Lubricant Properties at Low Temperature. Nanomaterials 2020, 10, 653. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Thomas, T.; Jayababu, N.; Shruthi, J.; Mathew, A.; Cerdán-Pasarán, A.; Hernández-Magallanes, J.; Sanal, K.; Reshmi, R. Room temperature ammonia sensing of α-MoO3 nanorods grown on glass substrates. Thin Solid Films 2021, 722, 138575. [Google Scholar] [CrossRef]
  36. Khademi, A.; Azimirad, R.; Zavarian, A.; Moshfegh, A. Growth and Field Emission Study of Molybdenum Oxide Nanostars. J. Phys. Chem. C-J Phys Chem. C 2009, 113, 6237. [Google Scholar] [CrossRef]
  37. Fu, D.; Fabre, B.; Loget, G.; Meriadec, C.; Ababou-Girard, S.; Cadot, E.; Leclerc-Laronze, N.; Marrot, J.; Ponfilly, Q. Polyoxothiometalate-Derivatized Silicon Photocathodes for Sunlight-Driven Hydrogen Evolution Reaction. ACS Omega 2018, 3, 13837–13849. [Google Scholar] [CrossRef]
  38. Valencia, C.; Valencia, C.; Zuluaga, F.; Zapata, M.V.; Mina, J.; Tovar, C.G. Synthesis and Application of Scaffolds of Chitosan-Graphene Oxide by the Freeze-Drying Method for Tissue Regeneration. Molecules 2018, 23, 2651. [Google Scholar] [CrossRef] [Green Version]
  39. Song, J.; Yuan, Q.; Liu, X.; Wang, D.; Fu, F.; Yang, W. Combination of Nitrogen Plasma Modification and Waterborne Polyurethane Treatment of Carbon Fiber Paper Used for Electric Heating of Wood Floors. BioResources 2015, 10, 5820–5829. [Google Scholar] [CrossRef] [Green Version]
  40. Maheswari, N.; Muralidharan, G. Controlled Synthesis of Nanostructured Molybdenum Oxide Electrodes for High Performance Supercapacitor Devices. Appl. Surf. Sci. 2017, 416, 94. [Google Scholar] [CrossRef]
  41. Myachina, M.; Gavrilova, N.; Nazarov, V. Formation of Molybdenum Blue Nanoparticles in the Organic Reducing Area. Molecules 2021, 26, 4438. [Google Scholar] [CrossRef]
  42. Verbinnen, B.; Block, C.; Hannes, D.; Lievens, P.; Vaclavikova, M.; Stefusova, K.; Gallios, G.; Vandecasteele, C. Removal of Molybdate Anions from Water by Adsorption on Zeolite-Supported Magnetite. Water Environ. Res. Res. Publ. Water Environ. Fed. 2012, 84, 753–760. [Google Scholar] [CrossRef]
  43. Tu, Y.-J.; Chan, T.-S.; Tu, H.-W.; Wang, S.-L.; You, C.-F.; Chang, C.-K. Rapid and efficient removal/recovery of molybdenum onto ZnFe2O4 nanoparticles. Chemosphere 2016, 148, 452–458. [Google Scholar] [CrossRef] [PubMed]
  44. Lian, J.; Zhou, F.; Chen, B.; Yang, M.; Wang, S.; Liu, Z.; Niu, S. Enhanced adsorption of molybdenum(VI) onto drinking water treatment residues modified by thermal treatment and acid activation. J. Clean. Prod. 2020, 244, 118719. [Google Scholar] [CrossRef]
  45. Romana, A. Arsenic and Molybdenum Ions Removal by Fe3O4 Embedded Acrylamide Triazine ybrid Polymer. Rev. Roum. Chim. 2017, 62, 139–148. [Google Scholar]
  46. Chen, Y.-C.; Lu, C. Kinetics, thermodynamics and regeneration of molybdenum adsorption in aqueous solutions with NaOCl-oxidized multiwalled carbon nanotubes. J. Ind. Eng. Chem. 2014, 20, 2521–2527. [Google Scholar] [CrossRef]
  47. Gönder, Z.B.; Vergili, I.; Kaya, Y.; Barlas, H. Adsorption of cationic and anionic surfactants onto organic polymer resin Lewatit VPOC 1064 MD PH. Environ. Geochem. Health 2010, 32, 267–273. [Google Scholar] [CrossRef]
  48. Namasivayam, C.; Kavitha, D. Removal of Congo Red from water by adsorption onto activated carbon prepared from coir pith, an agricultural solid waste. Dye Pigment. 2002, 54, 47–58. [Google Scholar] [CrossRef]
  49. Umpuch, C.; Sakaew, S. Removal of methyl orange from aqueous solutions by adsorption using chitosan intercalated montmorillonite. Songklanakarin J. Sci. Technol. 2013, 35, 451–459. [Google Scholar]
  50. Al-Degs, Y.S.; El-Barghouthi, M.I.; El-Sheikh, A.H.; Walker, G.M. Effect of solution pH, ionic strength, and temperature on adsorption behavior of reactive dyes on activated carbon. Dye. Pigment. 2008, 77, 16–23. [Google Scholar] [CrossRef]
  51. Almeida, C.A.P.; Debacher, N.A.; Downs, A.J.; Cottet, L.; Mello, C.A.D. Removal of methylene blue from colored effluents by adsorption on montmorillonite clay. J. Colloid Interface Sci. 2009, 332, 46–53. [Google Scholar] [CrossRef]
  52. Yu, Y.; Zhuang, Y.-Y.; Wang, Z.-H.; Qiu, M.-Q. Adsorption of water-soluble dyes onto modified resin. Chemosphere 2004, 54, 425–430. [Google Scholar] [CrossRef]
  53. Jaycock, M.J.; Parfitt, G.D. Chemistry of Interfaces; Ellis Horwood, Ltd.: Chichester, UK, 1981. [Google Scholar]
  54. Alkan, M.; Demirbaş, Ö.; Çelikçapa, S.; Doğan, M. Sorption of acid red 57 from aqueous solution onto sepiolite. J. Hazard. Mater. 2004, 116, 135–145. [Google Scholar] [CrossRef] [PubMed]
  55. Tahir, S.S.; Rauf, N. Removal of a cationic dye from aqueous solutions by adsorption onto bentonite clay. Chemosphere 2006, 63, 1842–1848. [Google Scholar] [CrossRef] [PubMed]
  56. Rytwo, G.; Huterer-Harari, R.; Dultz, S.; Gonen, Y. Adsorption of fast green and erythrosin-B to montmorillonite modified with crystal violet. J. Therm. Anal. Calorim. 2006, 84, 225–231. [Google Scholar] [CrossRef]
  57. Karaca, S.; Gürses, A.; Açıkyıldız, M.; Ejder, M. Adsorption of cationic dye from aqueous solutions by activated carbon. Microporous Mesoporous Mater. 2008, 115, 376–382. [Google Scholar] [CrossRef]
  58. Almanassra, I.W.; Kochkodan, V.; Mckay, G.; Atieh, M.A.; Al-Ansari, T. Review of phosphate removal from water by carbonaceous sorbents. J. Environ. Manag. 2021, 287, 112245. [Google Scholar] [CrossRef] [PubMed]
  59. Weidner, E.; Ciesielczyk, F. Removal of Hazardous Oxyanions from the Environment Using Metal-Oxide-Based Materials. Materials 2019, 12, E927. [Google Scholar] [CrossRef] [Green Version]
  60. Lee, C.-W. Molybdenum, Molybdenum Oxides, and their Electrochemistry. ChemSusChem 2012, 5, 1146–1161. [Google Scholar] [CrossRef]
  61. Goldberg, S.; Forster, H.; Godfrey, C. Molybdenum adsorption on oxides, clay minerals, and soils. Soil Sci. Soc. Am. J. 1996, 60, 425–432. [Google Scholar] [CrossRef] [Green Version]
  62. Xiao, B.; Thomas, K. Competitive Adsorption of Aqueous Metal Ions on an Oxidized Nanoporous Activated Carbon. Langmuir ACS J. Surf. Colloids 2004, 20, 4566–4578. [Google Scholar] [CrossRef]
  63. Parsons, D.; Salis, A. The Impact of the Competitive Adsorption of Ions at Surface Sites on Surface Free Energies and Surface Forces. J. Chem. Phys. 2015, 142, 134707. [Google Scholar] [CrossRef] [Green Version]
  64. Cano, T.; Offringa, N.D.; Willson, R.C. Competitive ion-exchange adsorption of proteins: Competitive isotherms with controlled competitor concentration. J. Chromatogr. A 2005, 1079, 116–126. [Google Scholar] [CrossRef] [PubMed]
  65. Greenwood, N.N.; Earnshaw, A. Chemistry of the Elements; Elsevier: Amsterdam, The Netherlands, 2012. [Google Scholar]
  66. Newcombe, G.; Drikas, M. Adsorption of NOM onto activated carbon: Electrostatic and non-electrostatic effects. Carbon 1997, 35, 1239–1250. [Google Scholar] [CrossRef]
  67. Alberghina, G.; Bianchini, R.; Fichera, M.; Fisichella, S. Dimerization of Cibacron Blue F3GA and other dyes: Influence of salts and temperature. Dyes Pigments 2000, 46, 129–137. [Google Scholar] [CrossRef]
  68. German-Heins, J.; Flury, M. Sorption of Brilliant Blue FCF in soils as affected by pH and ionic strength. Geoderma 2000, 97, 87–101. [Google Scholar] [CrossRef]
Figure 1. Schematic of the experimental work performed in this study.
Figure 1. Schematic of the experimental work performed in this study.
Water 15 00049 g001
Figure 2. SEM images of CDC particles at different magnifications.
Figure 2. SEM images of CDC particles at different magnifications.
Water 15 00049 g002aWater 15 00049 g002b
Figure 3. TEM images of CDC particles.
Figure 3. TEM images of CDC particles.
Water 15 00049 g003
Figure 4. (a) BJH pore size distribution curve of CDC; (b) N2 adsorption–desorption isotherm of CDC.
Figure 4. (a) BJH pore size distribution curve of CDC; (b) N2 adsorption–desorption isotherm of CDC.
Water 15 00049 g004
Figure 5. CDC zeta potential values vs. feed pH.
Figure 5. CDC zeta potential values vs. feed pH.
Water 15 00049 g005
Figure 6. Intensity of EDS spectra (counts per second) vs. energy (keV) for CDC samples before (red line) and after Mo adsorption (blue line).
Figure 6. Intensity of EDS spectra (counts per second) vs. energy (keV) for CDC samples before (red line) and after Mo adsorption (blue line).
Water 15 00049 g006
Figure 7. SEM elemental mapping of CDC before and after the adsorption of Mo.
Figure 7. SEM elemental mapping of CDC before and after the adsorption of Mo.
Water 15 00049 g007aWater 15 00049 g007b
Figure 8. FTIR spectra of CDC samples before (blue line) and after (red line) Mo adsorption.
Figure 8. FTIR spectra of CDC samples before (blue line) and after (red line) Mo adsorption.
Water 15 00049 g008
Figure 9. Removal efficiency of CDC vs. adsorption time: Mo concentration: 1 ppm, CDC loading: 0.6 g/L, pH: 5.2, agitation speed: 150 RPM.
Figure 9. Removal efficiency of CDC vs. adsorption time: Mo concentration: 1 ppm, CDC loading: 0.6 g/L, pH: 5.2, agitation speed: 150 RPM.
Water 15 00049 g009
Figure 10. Mo removal efficiency of CDC vs. CDC loading (g/L): Mo concentration: 1 ppm; pH: 5.2; agitation speed: 150 RPM; adsorption time: 1 h.
Figure 10. Mo removal efficiency of CDC vs. CDC loading (g/L): Mo concentration: 1 ppm; pH: 5.2; agitation speed: 150 RPM; adsorption time: 1 h.
Water 15 00049 g010
Figure 11. Removal efficiency and adsorption capacity of CDC at various Mo concentrations (ppm); CDC loading: 0.6 g/L; pH: 5.2; agitation speed: 150 RPM; adsorption time: 1 h.
Figure 11. Removal efficiency and adsorption capacity of CDC at various Mo concentrations (ppm); CDC loading: 0.6 g/L; pH: 5.2; agitation speed: 150 RPM; adsorption time: 1 h.
Water 15 00049 g011
Figure 12. Adsorption isotherm modeling of Mo removal by the CDC. pH 5.2, CDC dosage of 0.6 g/L, and adsorption time of 1 h.
Figure 12. Adsorption isotherm modeling of Mo removal by the CDC. pH 5.2, CDC dosage of 0.6 g/L, and adsorption time of 1 h.
Water 15 00049 g012
Figure 13. Nominal equilibrium concentration of Mo vs. time for the adsorption of Mo with CDC; CDC loading: 0.6 g/L; pH: 6; agitation speed: 150 RPM.
Figure 13. Nominal equilibrium concentration of Mo vs. time for the adsorption of Mo with CDC; CDC loading: 0.6 g/L; pH: 6; agitation speed: 150 RPM.
Water 15 00049 g013
Figure 14. Adsorption capacity of CDC vs. feed solution temperature at pH 6 and 9; CDC loading: 0.6 g/L, agitation speed: 150 RPM.
Figure 14. Adsorption capacity of CDC vs. feed solution temperature at pH 6 and 9; CDC loading: 0.6 g/L, agitation speed: 150 RPM.
Water 15 00049 g014
Figure 15. Speciation of Mo ions in water [60].
Figure 15. Speciation of Mo ions in water [60].
Water 15 00049 g015
Figure 16. Removal efficiency (a) and adsorption capacity (b) of CDC vs. Mo concentration (ppm) at different pH levels (3, 6, and 9); CDC loading: 0.6 g/L; agitation speed: 150 RPM.
Figure 16. Removal efficiency (a) and adsorption capacity (b) of CDC vs. Mo concentration (ppm) at different pH levels (3, 6, and 9); CDC loading: 0.6 g/L; agitation speed: 150 RPM.
Water 15 00049 g016aWater 15 00049 g016b
Figure 17. Schematic representation of the possible mechanisms of Mo adsorption by the CDC.
Figure 17. Schematic representation of the possible mechanisms of Mo adsorption by the CDC.
Water 15 00049 g017
Figure 18. Adsorption capacity of CDC toward various metals in a multicomponent aqueous solution (the heavy metal concentrations were 2 and 5 ppm at pH 2 and 4; CDC loading: 0.6 g/L; agitation speed: 150 RPM.
Figure 18. Adsorption capacity of CDC toward various metals in a multicomponent aqueous solution (the heavy metal concentrations were 2 and 5 ppm at pH 2 and 4; CDC loading: 0.6 g/L; agitation speed: 150 RPM.
Water 15 00049 g018
Figure 19. Mo removal efficiency and adsorption capacity of CDC at various NaCl concentrations in the feed solution.
Figure 19. Mo removal efficiency and adsorption capacity of CDC at various NaCl concentrations in the feed solution.
Water 15 00049 g019
Figure 20. Removal of Mo with CDC from different GW types; CDC loading: 0.6 g/L; pH: 3; agitation speed: 150 RPM.
Figure 20. Removal of Mo with CDC from different GW types; CDC loading: 0.6 g/L; pH: 3; agitation speed: 150 RPM.
Water 15 00049 g020
Figure 21. XPS (a,b) of CDC samples before and after Mo adsorption.
Figure 21. XPS (a,b) of CDC samples before and after Mo adsorption.
Water 15 00049 g021
Table 1. Adsorption kinetic models used in the study.
Table 1. Adsorption kinetic models used in the study.
ModelMathematical ExpressionLinearized FormXY
Pseudo-first-order d q t d t = K 1 ( q e q t ) ln ( q e q t ) = ln ( q e )   K1t ln ( q e q t ) vs. t
Pseudo-second-order d q t d t = K 2 ( q e q t ) 2 t q t = 1 K 2 q e 2 + ( 1 q e ) t t q t vs. t
Elovich d q t d t = α   e x p ( β q t ) q t = 1 β ln ( α β ) + ( 1 β ) ln ( t ) qt vs. ln t
Intra-particle diffusion q t = K I P D t + c q t = K I P D t + c qt vs. t
where qt is the adsorption capacity (in mg/g) at any given time (t, min) and qe is the equilibrium adsorption capacity (in mg/g). K1 and K2 represent the rate constants of the first- and second-order models in min−1 and g/mg/min, respectively. α and β, are constants related to the Elovich and intra-particle diffusion models. A and β are the initial rate of adsorption (in g/mg) and the desorption (mg/g/min) constant, respectively, whereas KIPD and C stand for the intraparticle intercept constant (mg/g/min−0.5) and the intraparticle intercept (mg/g), respectively.
Table 2. The adsorption isotherm models and adsorption parameters.
Table 2. The adsorption isotherm models and adsorption parameters.
Isotherm ModelNon-Linear ExpressionParameters
Langmuir q e = q m K L C e 1 + K L C e qm and KL stand for the maximum adsorption capacity of the monolayer and Langmuir constant
Freundlich q e = K F C e 1 n n is the heterogeneity factor and KF is the Freundlich constant (mg/g)/(dm3/mg)n
Temkin q e = R T ln ( K t C e ) b Kt is the equilibrium binding constant (in L/mg) and b is a parameter representing the variation in the adsorption energy values (in kJ/mol)
Sips q e = q m a x K s C e 1 n 1 + K s C e 1 n Ks and n are Sips’s constant and the model exponent, respectively
Table 3. Main findings reported in the literature for the removal of some heavy metals from water by adsorption onto various adsorbents, in addition to the present work.
Table 3. Main findings reported in the literature for the removal of some heavy metals from water by adsorption onto various adsorbents, in addition to the present work.
AdsorbentAdsorbent Loading (g/L)Contaminant Conc.Main FindingsRef
Crystalline ZnFe2O4 nanoparticles2Mo: 2.8 mg/LRemoval efficiency (%): 2.5–97;
qe: 1.4 mg/g;
residence time: 1 h.
[44]
Zeolite-supported magnetite20Mo: 0.9 mg/LRemoval efficiency (%): 99;
qe: 18–23 mg/g;
residence time: 24 h.
[43]
Acrylamide triazine polymer embedded with Fe3O41Mo: 2.5 mg/LRemoval efficiency (%): 99;
qe: 0.1 mg/g;
residence time: 2 h.
[46]
MWCNTs oxidized with NaOCl50Mo: 10 mg/LRemoval efficiency (%): 99;
qe: 6–12 mg/g;
residence time: >2 h.
[47]
Activated carbon modified with Al2O32.5Mo: 1 mg/LRemoval efficiency (%): 0;
qe (mg/g): 0;
residence time: 24 h.
[14]
CDC0.6Mo:1–50 mg/LRemoval efficiency: >99%;
qe (mg/g): 16.24;
residence time: 0.5 h
Present Work
Table 4. Isotherm parameters, correlation factors, and SSE of the tested models on Mo removal by the CDC.
Table 4. Isotherm parameters, correlation factors, and SSE of the tested models on Mo removal by the CDC.
ModelParametersR2SSE
LangmuirXm3.885 mg/g0.8310.9065
h 712.72 L/mg
RL0.00003–0.0014
Freundlichn14.1110.91180.4777
Kf0.323
Temkinb9644.11 J/mol0.85060.811
At438,470.18 L/mg
Sipsn2.0960.3712.096
Ks7.42 L/g
Table 5. Kinetic model parameters and correlation factors of the tested models on Mo adsorption by the CDC.
Table 5. Kinetic model parameters and correlation factors of the tested models on Mo adsorption by the CDC.
Qe (Experimental)4.18 mg/g
Pseudo-first-orderqe (model)2.98 mg/g
k10.0008 1 m i n
R20.4269
Pseudo-second-orderqe (model)1.672 mg/g
k20.3507 g m g m i n
R20.9998
Elovich α 6.28 m g g m i n
β 40.25 g/mg
R20.905
Intra-particle diffusionkIP0.0498 m g g m i n 0.5
c1.0076 mg/g
R20.6268
Table 6. Gibbs free energy, entropy, and enthalpy of the adsorption process of Mo onto CDC at various temperatures and pH 6 and 9; CDC loading: 0.6 g/L; agitation speed: 150 RPM.
Table 6. Gibbs free energy, entropy, and enthalpy of the adsorption process of Mo onto CDC at various temperatures and pH 6 and 9; CDC loading: 0.6 g/L; agitation speed: 150 RPM.
pHTemperature (°C) Δ S 0 (J/mol. K) Δ H 0 (kJ/mol) Δ G 0 (kJ/mol)
62548.4411.685−2.759
35−3.243
45−3.727
925−6.41.5223.43
353.49
453.55
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Manawi, Y.; Simson, S.; Lawler, J.; Kochkodan, V. Removal of Molybdenum from Contaminated Groundwater Using Carbide-Derived Carbon. Water 2023, 15, 49. https://doi.org/10.3390/w15010049

AMA Style

Manawi Y, Simson S, Lawler J, Kochkodan V. Removal of Molybdenum from Contaminated Groundwater Using Carbide-Derived Carbon. Water. 2023; 15(1):49. https://doi.org/10.3390/w15010049

Chicago/Turabian Style

Manawi, Yehia, Simjo Simson, Jenny Lawler, and Viktor Kochkodan. 2023. "Removal of Molybdenum from Contaminated Groundwater Using Carbide-Derived Carbon" Water 15, no. 1: 49. https://doi.org/10.3390/w15010049

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