Next Article in Journal
Recent Progress of Nanodiamond Film in Controllable Fabrication and Field Emission Properties
Next Article in Special Issue
Preparation and Antibacterial Properties of a Composite Fiber Membrane Material Loaded with Cationic Antibacterial Agent by Electrospinning
Previous Article in Journal
Atomistic Investigation on the Blocking Phenomenon of Crack Propagation in Cu Substrate Reinforced by CNT
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Iodine Removal by Porous Biochar-Confined Nano-Cu2O/Cu0: Rapid and Selective Adsorption of Iodide and Iodate Ions

Hebei Key Laboratory of Heavy Metal Deep-Remediation in Water and Resource Reuse, School of Environmental and Chemical Engineering, Yanshan University, Qinhuangdao 066004, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(3), 576; https://doi.org/10.3390/nano13030576
Submission received: 10 January 2023 / Revised: 28 January 2023 / Accepted: 29 January 2023 / Published: 31 January 2023

Abstract

:
Iodine is a nuclide of crucial concern in radioactive waste management. Nanomaterials selectively adsorb iodine from water; however, the efficient application of nanomaterials in engineering still needs to be developed for radioactive wastewater deiodination. Artemia egg shells possess large surface groups and connecting pores, providing a new biomaterial to remove contaminants. Based on the Artemia egg shell-derived biochar (AES biochar) and in situ precipitation and reduction of cuprous, we synthesized a novel nanocomposite, namely porous biochar-confined nano-Cu2O/Cu0 (C-Cu). The characterization of C-Cu confirmed that the nano-Cu2O/Cu0 was dispersed in the pores of AES biochar, serving in the efficient and selective adsorption of iodide and iodate ions from water. The iodide ion removal by C-Cu when equilibrated for 40 min exhibited high removal efficiency over the wide pH range of 4 to 10. Remarkable selectivity towards both iodide and iodate ions of C-Cu was permitted against competing anions (Cl/NO3/SO42−) at high concentrations. The applicability of C-Cu was demonstrated by a packed column test with treated effluents of 1279 BV. The rapid and selective removal of iodide and iodate ions from water is attributed to nanoparticles confined on the AES biochar and pore-facilitated mass transfer. Combining the advantages of the porous biochar and nano-Cu2O/Cu0, the use of C-Cu offers a promising method of iodine removal from water in engineering applications.

1. Introduction

Nuclear power produces about 10% of the total global electricity and about 5% of the total in China, and it will continue to be a viable alternative to fossil fuels in the future [1]. While nuclear power contributes to electricity supply and reduces CO2 emissions, it also poses significant challenges in terms of radioactive waste treatment and disposal. The radionuclides produced by nuclear reaction processes such as in uranium fission mainly contain 129/131I, 127Xe, 134/137Cs, 90Sr, 99Tc, and 79Se [1]. Among these, iodine is particularly problematic due to its volatility. Before treatments involving high temperatures such as vitrification, the liquidus iodine needs to be captured efficiently to avoid the release of gaseous wastes [2]. Additionally, the emergency accident in Fukushima, Japan, released radioactive iodine into the water and soil [3,4]. Radioactive iodine poses a significant threat to living creatures, including humans, and is considered one of the most dangerous radionuclides. As one of the halogen elements, iodine is highly soluble and mobile in terms of its chemical properties. It can quickly enter the human body through the food chain and can cause damage due to its chemical toxicity and radioactivity. Therefore, it is necessary to develop efficient methods to remove radioactive iodine.
Among the different halide removal treatments, adsorption still attracts significant interest because of its operational simplicity and flexibility [5,6]. Activated carbon removes gaseous iodine via physical adsorption, and the ion exchange resin enriches iodine in the form of liquidus ions [7,8]. In recent years, metals such as copper, silver, and bismuth have had a special selective adsorption capacity for iodine [9,10]. Among them, Cu(I)- and Cu(I)-based composite materials are the most promising owing to their high removal efficiency for aqueous iodide anions and their economic cost-effectiveness. Cu(I)- and Cu(I)-based composites remove iodide ions from the water via different mechanisms. In acidic solutions, the cuprous cation Cu+ and iodide anion I produce insoluble CuI precipitates, while in neutral and basic solutions, the iodide anions exchange with the hydroxyl group of Cu-OH and form an inner sphere–surface complex [10]. Moreover, the transient Cu(I) species can be generated by combining Cu(II)-containing compounds and monomeric copper Cu0 [11]. However, several issues inhibit the direct use of Cu(I) oxide or compounds for iodine removal from water. First, in neutral and acidic water, Cu+ is extremely unstable under oxidizing conditions and undergoes complex chemical reactions that change the valence state and affect the removal of iodine. Secondly, the oxidation states of iodine, mainly for iodate anions (IO3) in natural water, mean it is difficult for it to adsorb on solids based on Cu(I) [12]. Thirdly, the solvability of Cu+ may be problematic in water because copper can be very toxic when ingested in excess and is also highly toxic to aquatic life [13]. These issues are the constraints that affect the selective removal of iodine from water using copper-based materials.
In addition to the concern regarding the selectivity in the iodine removal process, the fast kinetic properties are also the focus of iodine removal materials. This is due to the differences in the hazardous effects of the different iodine isotopes on the environment and people. There are 24 radioactive isotopes of iodine with different adverse effects on the loss of control in the environment. The most typical radioactive iodine is 129I, which has an extremely long half-life of 1.57 × 107 years [14]. Its environmental risk is persistent and receives much attention in radioactive iodine removal. In contrast, the half-life of the radioactive isotope 131I is only 8.02 days, so in some cases it is not considered a long-term hazard [15]. However, the high radioactivity of 131I at 4.6 × 1015 Bq/g makes it a particularly acute risk to human health [2,16]. Thus, it is crucial to remove radioactive iodine quickly via emergency treatments; unfortunately, this has not been fully emphasized.
Porous materials are traditionally used in adsorption. Conventional activated carbon molecules are widely used for iodine adsorption because of their high surface area, large porosity, good thermal and chemical stability, and potential for facile modification. Additionally, compared to the developing synthetic adsorbents such as porous organic polymers (POPs) and metal–organic frameworks (MOFs), activated carbons are cheaper and simpler to prepare with higher productivity rates [3]. Therefore, the use of porous carbon materials is industrially feasible for iodine enrichment at the present stage. However, the use of low-cost porous structures to provide efficient and selective binding sites for the rapid adsorption of both iodide and iodate ions in water remains a challenge. Sorbent materials should have a unique morphology, the absence of any blocked bulk volume, low energy and overall costs, and high chemical durability. Natural biosubstrates have drawn wide attention for their high porosity, large surface areas, functional groups, and low costs. As an emerging adsorbent and nanoparticle host, Artemia egg shell-based composites have proven effective in the enrichment of phosphate, fluoride, and lead, as well as in the degradation of organic contaminants as catalysts [17,18,19,20,21,22,23,24,25,26]. The hierarchical porous macropore–mesopore structure facilitates nanoparticle dispersion and contaminant mass transfer. Thus, the biochar derived from Artemia egg shells may stably disperse Cu(I)-containing nanoparticles, offering the possibility of rapid iodine removal from water bodies. To the best of our knowledge, such research has not been reported.
Herein, we synthesized a novel porous nanocomposite noted as C-Cu via the in situ precipitation and reduction of Cu2O and Cu0 onto a biomatrix, namely Artemia egg shell-derived biochar (AES biochar). The iodine removal performance by the C-Cu was investigated through a series of adsorption experiments with iodide and iodate ions. The engineering applicability of C-Cu was evaluated using a fixed-bed column test. The advantageous properties of the porous biochar and nano-Cu2O/Cu0 are promising for real applications involving iodine removal by nanoparticles.

2. Materials and Methods

2.1. Materials and Chemicals

The Cu(NO3)2⋅3H2O was purchased from Aladdin Reagent Co., Shanghai, China. The KI, NaI, NaIO3, Na2SO4, NaCl, and NaNO3 were purchased from Kemio Chemical Reagent Co., Tianjin, China. The polyethylene imine (99% pure, MW 1800) was purchased from Ron Reagent Co., Shanghai, China. The commercial Cu2O was purchased from Sinopharm Chemical Reagent Co., Beijing, China. All chemicals were used without any further purification. The Artemia egg shells were provided by the Fisheries Research Institute of Qinhuangdao, Qinhuangdao, China.

2.2. Synthesis of Artemia Egg Shell Biochar-Hosted Nano-Cu2O/Cu0

As the raw material of the biochar, Artemia egg shells were washed with deionized water to remove salt and impurities. The washed Artemia egg shells were calcinated under vacuum at 550 °C for 3 h with a nitrogen flow rate of 60 mL/min, and the resulting product was Artemia egg shell biochar, noted as AES biochar. The Cu+ precursor solution was prepared with 1.06 g of polyethyleneimine (PEI) and 1.52 g of Cu(NO3)2⋅3H2O dissolved in 100 ml of deionized water. The Cu+ precursor solution and 1.0 g of AES biochar were vigorously mixed by stirring of 12 h. The mixture was then transferred to a 200 mL tetrafluoroethylene hydrothermal reactor and heated hermetically at 220 °C for 2 h. The solid was then filtered and dried at 60 °C to obtain Artemia egg shell biochar-hosted nano-Cu2O/Cu0, denoted as C-Cu. The copper concentration in the filtrate was determined via atomic absorption spectrometry (AAS, AA6800, Shimadzu, Kyoto, Japan) to calculate the copper content of the C-Cu.

2.3. Characterization

The morphology and structure of the adsorbents were recorded using a scanning electron microscope (SEM, S-3400N II, Hitachi, Tokyo, Japan) and transmission electron microscope (TEM, JEM-2010FX, JEOL, Tokyo, Japan). The specific surface area (SSA) and pore size distribution of the C-Cu and AES biochar were determined via N2 adsorption/desorption with a surface area and pore size analyzer (ASAP 2420, Micromeritics, Norcross, GA, USA). The specific surface areas were calculated from the Brunauer– Emmett–Teller (BET) equation. The total pore volumes were estimated using the adsorbed N2 amount at the relative pressure P/P0 of 0.99, and the micropore volumes were calculated from the t-plot method. The pore size distribution was obtained based on the Barrett–Joyner–Halenda (BJH) method. X-ray diffraction (XRD) data for the Cu2O, AEC biochar, and Cu-C were collected using an X-ray diffractometer (D/max2550PC, Rigaku, Tokyo, Japan) at a scanning speed of 5° 2θ/min over a range of 10° to 100°. The zeta potentials of the C-Cu, AES biochar, and Cu2O were determined at different pH values using a Zetasizer (Nano ZS-90) from Malvern Instruments, Malvern, UK.

2.4. Iodine Adsorption

In the iodine adsorption experiments, stable isotopes of 127I was used instead of radioactive iodine 129/131I. Solutions containing I or IO3 were prepared with sodium iodide or sodium iodate, respectively. Batch experiments were used for static adsorption, and the C-Cu was placed into solutions with 10 mg/L of I or IO3 at pH = 7.0 ± 0.2. The adsorption system was maintained at 25 °C using a thermostatic shaker with an agitation speed of 180 rpm. The dosage of C-Cu during the kinetic initiation process was 1.0 g and the solution volume was 1000 mL. As the adsorption process continued to various time points, supernatant samples were taken and analyzed using a spectrophotometer for the concentrations of I or IO3 [27]. In the isotherm experiments, 50 mL I or IO3 solutions with different initial concentrations of 1, 5, 10, 20, 50, 100, and 200 mg/L were contacted with 0.025 g of C-Cu for 24 h to allow adsorption equilibrium to be achieved. The pH of the solutions was adjusted to different values ranging between 2 and 12 to investigate the effect of the solution pH on the iodide and iodate ion adsorption onto the C-Cu. Na2SO4, NaCl, and NaNO3 were used in the selective adsorption experiments, and the competing anions of SO42−, Cl, and NO3 had molar ratios of 1, 5, 10, 20, 50, and 100 mol/mol. Cu2O and AEC biochar were used as the reference materials for the static adsorption experiments.
The column adsorption experiment was conducted to evaluate the working conditions of the adsorbents in the fixed-bed reactor to investigate the removal efficiency in engineering applications. A column with a 10 mm inner diameter was packed with 1.5 g of C-Cu to form a 25-mm-high fixed bed. The influent-simulated radioactive wastewater was percolated into the column at a constant flow rate using a peristaltic pump containing I = 2 mg/L, Cl = 150 mg/L, SO42− = 50 mg/L, NO3 = 50 mg/L, F = 50 mg/L, ReO4 = 10 mg/L, and HCO3 = 50 mg/L. An automatic collector collected the effluent samples for the iodide ion concentration measurements.

3. Results and Discussion

3.1. Characterization

The microstructure and chemical components of the Artemia egg shell biochar-hosted nano-Cu2O/Cu0 (C-Cu) is shown in Figure 1.
The C-Cu sample possessed a rough surface (Figure 1a) and the length and width were around 2–3 μm (Figure 1b). Fine particles can be observed in the TEM images in Figure 1b,c. In the XRD pattern in Figure 1d, the AES biochar is mainly amorphous carbon, which is also shown in the C-Cu sample. The characteristic lines of Cu2O (JCPDS No. 65-3288) exposed well-crystalized Cu2O in the commercial cuprous oxide and C-Cu, while Cu0 (JCPDS No. 04-0836) was only detected in the C-Cu. The particle sizes of the Cu2O and Cu0 in the C-Cu calculated from Scherrer’s formula were 36 nm and 35 nm, respectively. The lattices in the HR-TEM images clearly present Cu2O (111) and Cu0 (111) planes in Figure 1e,f. As listed in Table 1, the Brunauer–Emmett–Teller (BET) specific surface area and pore volume of C-Cu decreased after the dispersion of Cu species. However, the micropore volume and the average Barrett–Joyner–Halenda (BJH) pore size were comparable to the AES biochar. The AES biochar showed a micro–meso–macro pore distribution (Figure 1h inset) with the most probable pore size of 4.8 nm, in accordance with the desorption hysteresis typical for mesoporous materials (Figure 1h). However, the mesopore and macropore volumes of the C-Cu were about 50% of those of the AES biochar, indicating that the dispersed nanoparticles of Cu2O and Cu0 blocked some of the mesopores and macropores of the C-Cu. Thus, the hierarchical porous structure of the AES (Figure 1g) remained in the C-Cu, with the pores confined to the nanoparticle dispersion. The characterization results demonstrated that nanoparticles of Cu2O and Cu0 were loaded successfully on the AES biochar using the abovementioned synthesis method. The content of Cu in the C-Cu was 117.8 mg/g.

3.2. Iodine Removal Efficiency by C-Cu

The iodine removal efficiency was evaluated using the removal kinetics and isotherm adsorption of the C-Cu, and the experimental data and fitting results are illustrated in Figure 2 and Figure 3, respectively. Pseudo-first order and pseudo-second order models are shown in Equations (1) and (2), respectively [28].
Pseudo-first order model:
Q t = Q e ( 1 e k 1 t )
Pseudo-second order model:
Q t = Q e 2 k 2 t 1 + Q e k 2 t
Here, Qt (mg/g) is the adsorbed amount of iodide ions or iodate ions at a certain time t (min); Qe (mg/g) is the adsorption capacity of the iodide ions or iodate ions at equilibrium; k1 (1/min) and k2 (g⸱mg−1⸱min−1) are the kinetic rate constants of the pseudo-first order and pseudo-second order models, respectively.
It is clear from the kinetic data of the removal of iodide ions by C-Cu in Figure 2a that the rate of iodide ion adsorption was very rapid, reaching 50% of the maximum adsorption capacity at 10 min and the adsorption equilibrium at 40 min. In Figure 2b, the iodate ion adsorption capacity reached half of its maximal capacity at 20 min, and then the adsorption rate slowed and equilibrated at 60 min. The adsorption of iodide and iodate ions onto C-Cu was fitted using pseudo-first order and pseudo-second order models, and the results more consistent with the pseudo-second order kinetic model (Table 2). This kinetic fitting results explained the external and internal diffusion and adsorption of iodide and iodate anions onto the active sites of the porous C-Cu [28].
The Langmuir and Freundlich models shown in Equations (3) and (4) are used to fit the isotherm adsorption data, respectively [29].
Langmuir isotherm model:
Q e = Q m K L C e 1 + K L C e
Freundlich isotherm model:
Q e = K F C e 1 n
Here, Qe (mg/g) and Ce (mg/L) are the adsorbed amount and concentration of iodide ions or iodate ions at equilibrium; Qm (mg/g) is the maximum adsorption capacity of the Langmuir model; KL (L/mg) is the ratio of the adsorption and desorption rates; KF (L1/n⸱mg(1−1/n)⸱g−1) and n are the Freundlich model constants.
In Figure 3, the maximum adsorption capacity values of the iodide ions decreased with the increase in temperature, indicating that low temperatures were more favorable for C-Cu to adsorb iodide ions. At 20 °C, the maximum adsorption capacity reached 86.8 mg/g. Regarding the adsorption of iodate ions onto the C-Cu, the temperature presented opposite effects on the capacity compared with the iodide ions; the maximum adsorption capacity values increased with the increases in temperature to 29.8 mg/g, 37.4 mg/g, and 43.1 mg/g at 20 °C, 40 ℃, and 60 °C, respectively. Table 3 shows the regression coefficients of the iodine adsorption data onto C-Cu fitted by the Langmuir and Freundlich isothermal models, where the data are more consistent with the Freundlich model.
The removal mechanism of iodine by C-Cu was extrapolated using XRD patterns of the C-Cu before and after iodine adsorption, with Cu2O used as for comparison. The Cu2O and Cu0 crystal diffraction peaks were present in the C-Cu. After the adsorption of iodide ions, the C-Cu showed new diffraction peaks at 2θ = 25.6°, 50.0°, and 67.5°, corresponding to the (111), (311), and (311) crystal planes of γ-CuI (JCPDS No. 06-0246), respectively, which proved the generation of CuI. The stronger intensity of the diffraction peak of γ-CuI in the XRD pattern of C-Cu-I compared to Cu2O after I absorption further indicated the stronger I-binding ability of C-Cu.
Compared with the iodine removal materials listed in Table 4, the as-prepared Artemia egg shell biochar-hosted nano-Cu2O/Cu0 showed effective iodine removal performance in terms of the adsorption kinetics and maximum actual adsorption capacity (Qa) [30,31,32,33]. The hierarchical porous structure of the Artemia egg shell biochar facilitates the faster mass transfer of iodide and iodate anions onto the C-Cu surface, leading to equilibrium being achieved much quicker at 40 min. Moreover, the hierarchical porous structure promotes nano-Cu2O and Cu0 dispersion on the AES biochar surface, which provides abundant active sites for iodide or iodate ion sorption. Thus, high maximal adsorption capacities are achieved for iodide and iodate ions.

3.3. Effect of Solution pH on Iodine Removal

The iodine species vary at different solution pH levels, mainly occurring as iodide anions and iodate anions [13]. The effects of the solution pH on iodide and iodate anions are illustrated in Figure 4. In the experiment, the AES biochar removed 2.6% to 10.5% of the iodine as both iodide and iodate anions in the pH range of 2.1 to 12.0, while the C-Cu achieved good iodine removal at neutral pH. Over a wide pH range of 4.0 to 9.1, C-Cu maintained iodide ion removal rates higher than 81.5%, with maximum removal rates of around 96.4% at pH = 6.1 and 7.2. With increasing pH levels of 10.0, 10.9, and 12.0, the rate of iodide ion removal by C-Cu rapidly declined, reaching only 10.9% at 12.0. This pattern was consistent with the iodide ion removal by Cu2O, but due to the wider pH range and higher iodine uptake rates, the C-Cu was more efficient for iodine removal. The removal rate was related to the nanoparticle stability at different solution pH levels, as shown in Table 5. In acid solutions, the nanoparticles on the C-Cu leached by up to 26.8% at pH 3.1, which decreased the iodide and iodate ion adsorption and removal rates at acidic pH values. On the other hand, the nanoparticles remained stable in near-neutral to alkaline solutions, with a leaching rate of 2.0% at pH 7.2. Moreover, the zeta potentials of the AES biochar, Cu2O, and C-Cu, as illustrated in Figure 4c, can be used to explain the iodide ion removal efficiency. All sorbents showed decreased seta potentials with increasing pH values, but the AES biochar showed negative zeta potentials across the majority of the experimental pH range, which increased the repulsion towards the iodide anions on the biochar surface. In contrast, both Cu2O and C-Cu showed positive zeta potentials in the acidic to near-neutral pH range, with zero charge potentials of 5.10 and 6.75 for Cu2O and C-Cu, respectively. Positive potentials promoted high iodide ion removal efficiency by C-Cu by enhancing the electrostatic attraction to the iodide anions. The lower removal rates of iodide ions by Cu2O at pH values of 2.0 to 4.0 in Figure 4a might be attributed to the instability of Cu2O in acid solutions, which was mitigated by the pore confinement of Cu2O on C-Cu and led to an iodide ion removal rate of around 44.1% at pH = 2.0. Although the zeta potential partly indicated how the pH of the solution affected the removal of iodide ions, the electrostatic effect was not the only reason for the iodide ion removal. Even at lower pH values, Cu2O showed low iodate anion removal efficiency, despite showing positive zeta potentials. In comparison, C-Cu demonstrated removal rates of 43.7% for iodate ions in an iodate solution (Figure 4b), and 22.3% for iodate ions and 93.7% for iodide ions in an iodide and iodate solution, both at pH = 7.0 ~ 7.1. In addition to the electrostatic effect, a more complicated mechanism of iodine removal by C-Cu was indicated by the different affinities of the iodide and iodate ions on Cu2O and C-Cu.

3.4. High Selectivity of Iodide and Iodate Ions against Competitive Anions

As shown in Figure 5, C-Cu demonstrated remarkable selectivity towards both iodide and iodate ions against competing anions (Cl/NO3/SO42−) at high concentrations. Neither the competing anion species nor the concentration significantly weakened the adsorption of iodide ions onto C-Cu in the experimental. The distribution coefficient Kd increased by 18-fold when compared to the AES biochar, indicating that C-Cu has an extraordinarily high affinity for iodide ions that is not only brought about by electrostatic effects. The adsorption selectivity of C-Cu for iodide ions increased by around 20% over that of Cu2O, which should also be mentioned because it suggested that the affinity sites were used to their greater potential. The hierarchical porous structure of the AES biochar promoted the in situ generation of nano-Cu2O and Cu0 and uniform nanoparticle dispersion on the porous surface, which was attributed to the presence of high-affinity sites for iodide ions. The high selectivity of C-Cu towards iodine was also exhibited by the iodate ion adsorption. Even while the adsorption of C-Cu towards the iodate ions was somewhat reduced by the three competing anions with high concentrations, the Kd value of the iodate ions against 100-fold Cl (mol/mol) for the C-Cu was still 8 times higher than that of the AES biochar. Therefore, C-Cu offered superior iodine removal performance for natural water sources where large amounts of anions were present.

3.5. Continuous Adsorption Performance

To further investigate the applicability of C-Cu in an engineering continuous flow treatment system, a fixed-bed column adsorption experiment was conducted, with commercial Cu2O powder used as a comparison, and the experimental data are shown in Figure 6. The concentration of iodide ions in the influent was 2.0 mg/L, and the concentration in the effluent of 941 BV remained stable at about 0.78 mg/L, which provided an iodide ion removal rate of about 61%. After this, the I concentration in the effluent started to rise and a breakthrough occurred when the effluent reached 1279 BV. The fixed-bed column packed with C-Cu processed more than four times as much water as the Cu2O column did, while also having a lower iodide ion concentration in the effluent, which was ascribed to the hierarchical porous structure of the C-Cu and pore-confined nanoparticles. The large water treatment volumes and high iodine removal performance rates under engineering conditions demonstrated the application potential of C-Cu for the deep treatment of iodine-containing wastewater.

4. Conclusions

A novel adsorbent was synthesized via the in situ precipitation and reduction of Cu2O and Cu0 on a porous biomatrix of Artemia egg shell biochar. SEM, TEM, XRD, and BET analyses confirmed that approximately 35 nm of nano-Cu2O/Cu0 dispersed in the pores of the AES biochar. The C-Cu induced a rapid adsorption equilibrium for iodide and iodate ions. The adsorption isotherm experiments showed that the maximum actual adsorption capacities of the iodide and iodate ions by C-Cu reached 86.8 and 43.1 mg/g at pH = 7.0 ± 0.2, respectively. Moreover, for the iodide and iodate ions, the C-Cu achieved Kd values about 18 and 8 times greater than for the AES biochar against Cl at a 100-fold molar ratio. These excellent adsorption performance results were attributed to mechanisms including electrostatic interactions and precipitation. The continuous adsorption experiments demonstrated the potential of C-Cu for engineering applications. Overall, the rapid and selective adsorption of iodide and iodate ions onto C-Cu makes this nanocomposite attractive for efficiently removing radioactive iodine from water.

Author Contributions

Conceptualization, Q.S.; formal analysis, M.W. and X.Z.; investigation, Z.L. and Y.N.; writing—original draft preparation, J.L.; writing—review and editing, Q.S. and S.W.; funding acquisition, Q.S. and S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Hebei Province (No. B2021203036, E2022203011) and the Key Project of the Hebei Education Department (No. ZD2021103).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jin, K.; Lee, B.; Park, J. Metal-organic frameworks as a versatile platform for radionuclide management. Coord. Chem. Rev. 2021, 427, 213473. [Google Scholar] [CrossRef]
  2. Pénélope, R.; Campayo, L.; Fournier, M.; Gossard, A.; Grandjean, A. Solid sorbents for gaseous iodine capture and their conversion into stable waste forms. J. Nucl. Mater. 2022, 563, 153635. [Google Scholar] [CrossRef]
  3. Xie, W.; Cui, D.; Zhang, S.-R.; Xu, Y.-H.; Jiang, D.-L. Iodine capture in porous organic polymers and metal–organic frameworks materials. Mater. Horiz. 2019, 6, 1571–1595. [Google Scholar] [CrossRef]
  4. Lehto, J.; Koivula, R.; Leinonen, H.; Tusa, E.; Harjula, R. Removal of Radionuclides from Fukushima Daiichi Waste Effluents. Sep. Purif. Rev. 2019, 48, 122–142. [Google Scholar] [CrossRef]
  5. Zhang, X.; Gu, P.; Liu, Y. Decontamination of radioactive wastewater: State of the art and challenges forward. Chemosphere 2019, 215, 543–553. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, Q.; Bolisetty, S.; Cao, Y.; Handschin, S.; Adamcik, J.; Peng, Q.; Mezzenga, R. Selective and Efficient Removal of Fluoride from Water: In Situ Engineered Amyloid Fibril/ZrO2 Hybrid Membranes. Angew. Chem. Int. Ed. Engl. 2019, 58, 6012–6016. [Google Scholar] [CrossRef] [PubMed]
  7. Decamp, C.; Happel, S. Utilization of a mixed-bed column for the removal of iodine from radioactive process waste solutions. J. Radioanal. Nucl. Chem. 2013, 298, 763–767. [Google Scholar] [CrossRef]
  8. Sinha, P.K.; Lal, K.B.; Ahmed, J. Removal of radioiodine from liquid effluents. Waste Manag. 1997, 17, 33–37. [Google Scholar] [CrossRef]
  9. Asmussen, R.M.; Matyas, J.; Qafoku, N.P.; Kruger, A.A. Silver-functionalized silica aerogels and their application in the removal of iodine from aqueous environments. J. Hazard. Mater. 2019, 379, 119364. [Google Scholar] [CrossRef]
  10. Lefèvre, G.; Bessière, J.; Ehrhardt, J.-J.; Walcarius, A. Immobilization of iodide on copper(I) sulfide minerals. J. Environ. Radioact. 2003, 70, 73–83. [Google Scholar] [CrossRef] [PubMed]
  11. LefÈvre, G.; Alnot, M.; Ehrhardt, J.J.; BessiÈre, J. Uptake of Iodide by a Mixture of Metallic Copper and Cupric Compounds. Environ. Sci. Technol. 1999, 33, 1732–1737. [Google Scholar] [CrossRef]
  12. Zhang, S.; Xu, C.; Creeley, D.; Ho, Y.-F.; Li, H.-P.; Grandbois, R.; Schwehr, K.A.; Kaplan, D.I.; Yeager, C.M.; Wellman, D.; et al. Iodine-129 and Iodine-127 Speciation in Groundwater at the Hanford Site, US: Iodate Incorporation into Calcite. Environ. Sci. Technol. 2013, 47, 9635–9642. [Google Scholar] [CrossRef] [PubMed]
  13. Moore, R.C.; Pearce, C.I.; Morad, J.W.; Chatterjee, S.; Levitskaia, T.G.; Asmussen, R.M.; Lawter, A.R.; Neeway, J.J.; Qafoku, N.P.; Rigali, M.J.; et al. Iodine immobilization by materials through sorption and redox-driven processes: A literature review. Sci. Total Environ. 2020, 716, 132820. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, L.; Liu, J.; Zhang, W.; Zhou, J.; Luo, D.; Li, Z. Uranium (U) source, speciation, uptake, toxicity and bioremediation strategies in soil-plant system: A review. J. Hazard. Mater. 2021, 413, 125319. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, Z.; Zhang, L.; Zhang, K.; Lu, Y.; Chen, J.; Wang, S.; Hu, B.; Wang, X. Application of carbon dots and their composite materials for the detection and removal of radioactive ions: A review. Chemosphere 2022, 287, 132313. [Google Scholar] [CrossRef]
  16. Zhang, X.; Maddock, J.; Nenoff, T.M.; Denecke, M.A.; Yang, S.; Schröder, M. Adsorption of iodine in metal-organic framework materials. Chem. Soc. Rev. 2022, 51, 3243–3262. [Google Scholar] [CrossRef] [PubMed]
  17. Qian, C.; Yin, J.; Zhao, J.; Li, X.; Wang, S.; Bai, Z.; Jiao, T. Facile preparation and highly efficient photodegradation performances of self-assembled Artemia eggshell-ZnO nanocomposites for wastewater treatment. Colloids Surf. Physicochem. Eng. Asp. 2021, 610, 125752. [Google Scholar] [CrossRef]
  18. Song, Y.; Song, X.; Sun, Q.; Wang, S.; Jiao, T.; Peng, Q.; Zhang, Q. Efficient and sustainable phosphate removal from water by small-sized Al(OH)(3) nanocrystals confined in discarded Artemia Cyst-shell: Ultrahigh sorption capacity and rapid sequestration. Sci. Total Environ. 2022, 803, 150087. [Google Scholar] [CrossRef]
  19. Wang, B.; Xia, J.; Mei, L.; Wang, L.; Zhang, Q. Highly Efficient and Rapid Lead(II) Scavenging by the Natural Artemia Cyst Shell with Unique Three-Dimensional Porous Structure and Strong Sorption Affinity. ACS Sustain. Chem. Eng. 2018, 6, 1343–1351. [Google Scholar] [CrossRef]
  20. Wang, S.; Ao, J.; Lv, F.; Zhang, Q.; Jiao, T. The enhanced antibacterial performance by the unique Artemia egg shell-supported nano-Ag composites. J. Taiwan Inst. Chem. Eng. 2016, 61, 336–341. [Google Scholar] [CrossRef]
  21. Wang, S.; Lv, F.; Jiao, T.; Ao, J.; Zhang, X.; Jin, F. A Novel Porous Carrier Found in Nature for Nanocomposite Materials Preparation: A Case Study of Artemia Egg Shell-Supported TiO2 for Formaldehyde Removal. J. Nanomater. 2015, 2015, 963012. [Google Scholar] [CrossRef] [Green Version]
  22. Wang, S.; Lv, X.; Fu, M.; Wang, Z.; Zhang, D.; Sun, Q. Risk assessment of Artemia egg shell-Mg-P composites as a slow-release phosphorus fertilizer during its formation and application in typical heavy metals contaminated environment. J. Environ. Manag. 2022, 329, 117092. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, S.; Ma, M.; Zhang, Q.; Sun, G.; Jiao, T.; Okazaki, R.K. Efficient Phosphate Sequestration in Waters by the Unique Hierarchical 3D Artemia Egg Shell Supported Nano-Mg(OH)(2) Composite and Sequenced Potential Application in Slow Release Fertilizer. ACS Sustain. Chem. Eng. 2015, 3, 2496–2503. [Google Scholar] [CrossRef]
  24. Wang, S.-F.; Sun, S.-C.; Okazaki, R.K. Comparative study on thermotolerance of Artemia resting eggs from Qinghai-Xizang Plateau, China. Aquaculture 2010, 307, 141–149. [Google Scholar] [CrossRef]
  25. Zhao, J.; Yin, J.; Zhong, J.; Jiao, T.; Bai, Z.; Wang, S.; Zhang, L.; Peng, Q. Facile preparation of a self-assembled Artemia cyst shell-TiO2-MoS2 porous composite structure with highly efficient catalytic reduction of nitro compounds for wastewater treatment. Nanotechnology 2020, 31, 085603. [Google Scholar] [CrossRef] [PubMed]
  26. Zhao, X.; Song, Y.; Zhao, Z.; Gao, W.; Peng, Q.; Zhang, Q. Efficient fluoride removal from water by amino Acid-enriched Artemia Cyst motivated Sub-10 nm La(OH)(3) confined inside superporous skeleton. Sep. Purif. Technol. 2022, 283, 120205. [Google Scholar] [CrossRef]
  27. Zhao, Q.; Zhu, L.; Lin, G.; Chen, G.; Liu, B.; Zhang, L.; Duan, T.; Lei, J. Controllable Synthesis of Porous Cu-BTC@polymer Composite Beads for Iodine Capture. ACS Appl. Mater. Interfaces 2019, 11, 42635–42645. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, J.; Guo, X. Adsorption kinetic models: Physical meanings, applications, and solving methods. J. Hazard. Mater. 2020, 390, 122156. [Google Scholar] [CrossRef]
  29. Wang, J.; Guo, X. Adsorption isotherm models: Classification, physical meaning, application and solving method. Chemosphere 2020, 258, 127279. [Google Scholar] [CrossRef]
  30. Mao, P.; Qi, B.; Liu, Y.; Zhao, L.; Jiao, Y.; Zhang, Y.; Jiang, Z.; Li, Q.; Wang, J.; Chen, S.; et al. AgII doped MIL-101 and its adsorption of iodine with high speed in solution. J. Solid State Chem. 2016, 237, 274–283. [Google Scholar] [CrossRef]
  31. Mao, P.; Liu, Y.; Jiao, Y.; Chen, S.; Yang, Y. Enhanced uptake of iodide on Ag@Cu2O nanoparticles. Chemosphere 2016, 164, 396–403. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, X.; Gu, P.; Li, X.; Zhang, G. Efficient adsorption of radioactive iodide ion from simulated wastewater by nano Cu2O/Cu modified activated carbon. Chem. Eng. J. 2017, 322, 129–139. [Google Scholar] [CrossRef]
  33. Chen, J.; Wang, J.; Gao, Q.; Zhang, X.; Liu, Y.; Wang, P.; Jiao, Y.; Zhang, Z.; Yang, Y. Enhanced removal of I- on hierarchically structured layered double hydroxides by in suit growth of Cu/Cu2O. J. Environ. Sci. 2020, 88, 338–348. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) SEM image and TEM images of C-Cu at (b) 0.5 μm and (c) 100 nm. (d) XRD patterns of AES biochar, Cu2O, and C-Cu. HR-TEM images of C-Cu with (e) Cu2O and (f) Cu0 planes. (g) SEM image of AES biochar. (h) N2 sorption isotherms of C-Cu and AES biochar. Inset: Pore size distribution.
Figure 1. (a) SEM image and TEM images of C-Cu at (b) 0.5 μm and (c) 100 nm. (d) XRD patterns of AES biochar, Cu2O, and C-Cu. HR-TEM images of C-Cu with (e) Cu2O and (f) Cu0 planes. (g) SEM image of AES biochar. (h) N2 sorption isotherms of C-Cu and AES biochar. Inset: Pore size distribution.
Nanomaterials 13 00576 g001
Figure 2. Data on the adsorption of (a) iodide ions and (b) iodate ions onto C-Cu fitted by kinetic models: C0 = 10 mg/L, adsorbents 1.0 g/L, 25 °C, pH = 7.0 ± 0.2. (c) XRD patterns of Cu2O, Cu2O-I, and C-Cu-I.
Figure 2. Data on the adsorption of (a) iodide ions and (b) iodate ions onto C-Cu fitted by kinetic models: C0 = 10 mg/L, adsorbents 1.0 g/L, 25 °C, pH = 7.0 ± 0.2. (c) XRD patterns of Cu2O, Cu2O-I, and C-Cu-I.
Nanomaterials 13 00576 g002
Figure 3. Isotherm adsorption of iodide ions onto C-Cu at (a) 20 °C, (b) 40 °C, and (c) 60 °C and of iodate ions at (d) 20 °C, (e) 40 °C, and (f) 60 °C. C0 = 10 mg/L, C-Cu dose 0.5 g/L, pH = 7.0 ± 0.2.
Figure 3. Isotherm adsorption of iodide ions onto C-Cu at (a) 20 °C, (b) 40 °C, and (c) 60 °C and of iodate ions at (d) 20 °C, (e) 40 °C, and (f) 60 °C. C0 = 10 mg/L, C-Cu dose 0.5 g/L, pH = 7.0 ± 0.2.
Nanomaterials 13 00576 g003
Figure 4. Effect of the solution pH on (a) iodide anion and (b) iodate anion removal rates by Cu2O, AES biochar, and C-Cu. (c) Effect of the solution pH on iodide ion and iodate ion adsorption onto C-Cu. C0 = 10 mg/L, adsorbent dosage 0.5 g/L, 25 °C. (d) Zeta potentials of Cu2O, AES biochar, and C-Cu.
Figure 4. Effect of the solution pH on (a) iodide anion and (b) iodate anion removal rates by Cu2O, AES biochar, and C-Cu. (c) Effect of the solution pH on iodide ion and iodate ion adsorption onto C-Cu. C0 = 10 mg/L, adsorbent dosage 0.5 g/L, 25 °C. (d) Zeta potentials of Cu2O, AES biochar, and C-Cu.
Nanomaterials 13 00576 g004
Figure 5. Effects of competitive ions on the removal of iodide and iodate anions: (ac) I adsorption against SO42− Cl, and NO3; (d) Kd value of I against Cl; (eg) IO3 adsorption against SO42−, Cl, and NO3; (h) Kd value of IO3 against Cl; adsorbents 0.5 g/L, 25 °C, pH = 7 ± 0.2.
Figure 5. Effects of competitive ions on the removal of iodide and iodate anions: (ac) I adsorption against SO42− Cl, and NO3; (d) Kd value of I against Cl; (eg) IO3 adsorption against SO42−, Cl, and NO3; (h) Kd value of IO3 against Cl; adsorbents 0.5 g/L, 25 °C, pH = 7 ± 0.2.
Nanomaterials 13 00576 g005
Figure 6. Column adsorption of iodide ions onto Cu2O and C-Cu. C0 = 2.0 mg/L, adsorbents 1.5 g, 25 °C, pH = 7.0 ± 0.2, Cl = 150 mg/L, SO42− = 50 mg/L, NO3 = 50 mg/L, F = 50 mg/L, ReO4 = 10 mg/L, HCO3 = 50 mg/L, 20 BV/h.
Figure 6. Column adsorption of iodide ions onto Cu2O and C-Cu. C0 = 2.0 mg/L, adsorbents 1.5 g, 25 °C, pH = 7.0 ± 0.2, Cl = 150 mg/L, SO42− = 50 mg/L, NO3 = 50 mg/L, F = 50 mg/L, ReO4 = 10 mg/L, HCO3 = 50 mg/L, 20 BV/h.
Nanomaterials 13 00576 g006
Table 1. Specific surface area, pore volume, and pore size values of AES biochar and C-Cu.
Table 1. Specific surface area, pore volume, and pore size values of AES biochar and C-Cu.
AdsorbentsSpecific Surface Area
(m2/g)
Pore Volume
(cm3/g)
Micropore Volume
(cm3/g)
BJH Average Pore Size
(nm)
AES biochar130.0450.0047.8
C-Cu80.0230.0037.3
Table 2. Adsorption kinetic parameters of iodide and iodate ions on C-Cu.
Table 2. Adsorption kinetic parameters of iodide and iodate ions on C-Cu.
Pseudo-First OrderPseudo-Second Order
Qe (mg/g)k1 (1/min)R2Qe (mg/g)k2 (g⸱mg−1⸱min−1)R2
I9.80.06300.95410.60.00970.987
IO37.50.04670.9939.10.00570.994
Table 3. Adsorption isotherm parameters of iodide and iodate ions on C-Cu.
Table 3. Adsorption isotherm parameters of iodide and iodate ions on C-Cu.
IodineT (K)Langmuir ModelFreundlich Model
Qm (mg/g)KL (L/mg)R2KF
(L1/n⸱mg(1−1/n)⸱g−1)
nR2
I29376.30.5500.93223.81−0.2690.944
31373.10.1110.94915.96−0.3160.974
33354.50.0200.9074.38−0.4450.921
IO329329.80.0440.9854.14−0.3720.968
31337.40.0420.9774.75−0.3900.972
33343.10.0440.9385.64−0.3890.994
Table 4. Iodide ion removal performance parameters of different adsorbents.
Table 4. Iodide ion removal performance parameters of different adsorbents.
AdsorbentsAdsorption ConditionsEquilibrium Time
(Min)
Qa (mg/g)Reference
IIO3
Ag@MIL-101C0 = 12.7–508 mg/L, pH = 7.0, dose 1.0 g/L1802.1[30]
Ag@Cu2OC0 = 2.5~38 mg/L, pH = 3~10, dose 1.0 g/L18025.4[31]
Cu2O-CC0 = 1~20 mg/L, pH = 7, dose 1.0 g/L12041.2[32]
Cu/Cu2O-LDHC0 = 0~240 mg/L, pH = 6.5, dose 1.0 g/L265137.8[33]
C-CuC0 = 1~200 mg/L, pH = 4~10, dose 0.5 g/L40 (I), 60 (IO3)86.843.1This work
Table 5. Cu leaching rates (%) in solutions with different pH values.
Table 5. Cu leaching rates (%) in solutions with different pH values.
Solution pHCu leaching Rate (%)Solution pHCu leaching Rate (%)
3.026.88.01.1
4.011.99.01.0
5.07.210.00.8
6.04.611.00.1
7.02.012.00.1
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

Li, J.; Wang, M.; Zhao, X.; Li, Z.; Niu, Y.; Wang, S.; Sun, Q. Efficient Iodine Removal by Porous Biochar-Confined Nano-Cu2O/Cu0: Rapid and Selective Adsorption of Iodide and Iodate Ions. Nanomaterials 2023, 13, 576. https://doi.org/10.3390/nano13030576

AMA Style

Li J, Wang M, Zhao X, Li Z, Niu Y, Wang S, Sun Q. Efficient Iodine Removal by Porous Biochar-Confined Nano-Cu2O/Cu0: Rapid and Selective Adsorption of Iodide and Iodate Ions. Nanomaterials. 2023; 13(3):576. https://doi.org/10.3390/nano13030576

Chicago/Turabian Style

Li, Jiaqi, Mengzhou Wang, Xu Zhao, Zitong Li, Yihui Niu, Sufeng Wang, and Qina Sun. 2023. "Efficient Iodine Removal by Porous Biochar-Confined Nano-Cu2O/Cu0: Rapid and Selective Adsorption of Iodide and Iodate Ions" Nanomaterials 13, no. 3: 576. https://doi.org/10.3390/nano13030576

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