Cesium Adsorption/Desorption Behavior of Clay Minerals Considering Actual Pollution Conditions in Fukushima | Scientific Reports Scientific Reports

2021-11-22 07:51:44 By : Ms. Leah Li

Thank you for visiting Nature. The browser version you are using has limited support for CSS. For the best experience, we recommend that you use a newer version of the browser (or turn off the compatibility mode in Internet Explorer). At the same time, to ensure continued support, we will display sites without styles and JavaScript.

Scientific Reports Volume 6, Article Number: 21543 (2016) Cite this article

Taking into account the actual pollution conditions in Fukushima, 137Cs radioisotope and imaging plate (IP) autoradiography were used to perform cesium adsorption/desorption experiments on various clay minerals. Drop a 50 μl solution containing 0.185 ~ 1.85 Bq 137Cs (10-11 ~ 10-9 molL-1 of 137Cs) on the substrate where various mineral particles are arranged. It was discovered that part of the vermiculized biotite was called "weathered biotite" (WB) in this study. The adsorption capacity of 137Cs from Fukushima far exceeds that of other clay minerals (fresh biotite, illite, montmorillonite, high Lingshi, halloysite, diaspore and imogolite) are on the same substrate. When WB is not present on the substrate, the amount of 137Cs adsorbed to other clay minerals increases significantly, which means that the selective adsorption of WB leads to the consumption of radioactive cesium in the solution and the reduction of the adsorption of coexisting minerals. The adsorption of Cs to WB lasted about one day, and the adsorption of iron montmorillonite was completed within one hour. The 137Cs adsorbed in WB is difficult to be leached by hydrochloric acid at a pH of 1, especially in samples with a long adsorption time. The presence/absence of WB absorbed radioactive cesium is a key factor affecting the dynamics and fate of Fukushima radioactive cesium.

Since the Fukushima Daiichi Nuclear Power Plant (FDNPP) accident in March 2011, a detailed understanding of the chemical state and dynamics of radioactive cesium that causes high air dose rates in polluted areas has been an important issue. Research conducted before the accident showed that mica minerals were responsible for the retention of radioactive cesium in the soil1,2. Several groups suggested that illite, a dioctahedral interlayer defect mica containing aluminum as the main octahedral cation, is the main adsorption mineral of radioactive cesium in the soil. They studied its Cs adsorption properties 4,5,6, 7,8,9. Some of them 5, 7, 9 indicate that the structure of illite contains a variety of adsorption sites with different affinities and densities. From similar Cs adsorption experiments 10, 11, 12, and 13, such results for actual soil or sediment are also reported. In these studies, the location with the highest affinity (but low density) is considered to be on the frayed edge of the mica mineral, and is called the "frayed edge location" (FES)1,14. On the other hand, there are several work reports that hydroxyl-inserted vermiculite (HIV) can also effectively adsorb and fix Cs ions15,16,17,18. The origin of HIV high Cs affinity is attributed to the wedge-shaped interlayer transition zone from mica to HIV, which provides suitable adsorption sites for Cs and Rb, similar to FES in mica.

On the other hand, Mukai et al.19 reported that part of the vermiculized biotite was called "weathered biotite (WB)" in their research. It originated from the weathering of granite in Fukushima and was contaminated by collected radioactive cesium. The main substance in the soil. From Fukushima. Komarneni et al.20 reported the superior adsorption properties of altered phlogopite (such as vermiculite). However, a recent study investigating the Cs adsorption capacity of various clay minerals shows that vermiculite is not superior to other mica clay minerals such as illite and montmorillonite. Therefore, we hope to find out whether Fukushima's WB is better than other Fukushima clay minerals in adsorbing Cs through further experiments. First, a Cs adsorption experiment was performed using a solution with a cesium concentration as low as sub-ppm (~10-5 molL-1). However, the Cs adsorption characteristics of minerals depend on the concentration of cesium in the solution and/or the ratio of the amount of minerals in the solution to the amount of Cs ions. The actual concentration of radioactive cesium in raindrops that cause radioactive contamination is considered to be very low. For example, the rainfall in Iitate Village, a heavily polluted area in Fukushima, in the weeks after the accident was approximately 10 mm according to the records of the Japan Meteorological Agency23. On the other hand, the amount of 137Cs deposited on Iitate-village per unit area in the literature is ~106 Bq/m2. According to these values, the concentration in raindrops should be on the order of 10 ppt (10-10 molL-1). In order to discuss the cesium contamination event in Fukushima soil, a similar low concentration of Cs should be used for adsorption experiments. However, this is close to or below the detection limit of the most sensitive analytical instruments. If radioactive cesium itself is used as a cesium source, and the adsorption/desorption process is estimated by measuring the radiation of the radioisotope, this problem can be solved, just like the previous work 6,13,22 did. In addition, we used quantitative autoradiography with imaging plates (IP) instead of counting gamma rays to measure the radiation in individual mineral particles, thereby assessing the amount of 137C adsorbed in minerals. By using IP autoradiography, we can study the kinetic aspects of the reaction between the solution and various minerals (details will be described later). Compared with the reaction between a solution and a single mineral, this method is believed to better reproduce the reaction between raindrops and the soil composed of various minerals actually occurring in Fukushima. In this study, the results of this unique adsorption/desorption experiment on clay minerals that may be present in Fukushima soil are shown. They clearly demonstrate the importance of WB as a radioactive cesium adsorbent for Fukushima.

Eight mineral species, each with four or five particles with a particle size of about 50 μm, are arranged in a 4 mm x 7 mm square area on the acrylic substrate with Kapton double-sided tape (Figure S1). Then, a 50 μL solution of 137Cs containing 3.7, 37, and 370 Bq/mL (0.185, 1.85, and 18.5 Bq in the solution) was dropped to cover all the particles on the substrate. The readout image of IP is shown in Figure 1. Under any condition of different 137Cs concentration and immersion time, the amount of 137Cs adsorbed by WB collected from Fukushima is much higher than other minerals. In the case of low 137Cs concentration and/or short reaction time, only 137Cs was detected at the WB imaging position, indicating that WB is the only substance that adsorbs 137Cs. At a concentration of 18.5 Bq/50 μL, the amount of 137Cs adsorbed by WB is about two orders of magnitude higher than that of other clay minerals after one day of reaction. These results clearly show that in the weathered granite soil, the adsorption capacity of WB for cesium far exceeds that of other clay minerals, and its cesium concentration is equivalent to that of Fukushima wet deposits. This result is consistent with our previous results that WB 19 is often found in fine radioactive particles collected in Fukushima contaminated soil. Our recent survey25 shows that the radioactivity of the site radioactive soil particles is usually about 10-3 ~ 10-2 Bq, which roughly corresponds to the results in Figure 1, with a concentration range of 0.185 Bq/50 μL to 1.85 Bq/50 μL. At such a low concentration level, it is still uncertain where the actual adsorption site of cesium is in the WB structure. Considering the extremely low concentration of radioactive cesium, the adsorption sites are likely to be those with the highest affinity in minerals. As suggested by previous work14,18, these sites may be the transitional zone in FES or HIV, where non-hydrated cesium ions with an ionic radius larger than that of potassium ions exist stably. However, further complicated experiments are needed to prove this hypothesis. On the other hand, the Cs adsorption capacity of illite, which was considered to be the main Cs adsorbent in contaminated soil in previous work 4,5,6,7,8,9, is much lower than that of WB, although it is obvious that minerals The type itself does not control the adsorption performance.

The IP readout image matrix covering the substrate with various mineral particles (five particles per species) adsorbed radioactive cesium from the solution.

The radioactive input and reaction time of the solution are located at the top and left, respectively. The number in the upper right corner of each image is the estimated percentage of radioactivity (or 137Cs) absorbed from the IP signal to the entire mineral particle. The arrangement of mineral particles is shown in Figure S1. The abbreviations in the lower right corner indicate FB: fresh biotite, WB: weathered biotite, K: kaolinite, H: halloysite. IL: illite, M: montmorillonite, A: diaspore, IM: Imogenite.

In the experiment shown in Figure 1, we used montmorillonite as a representative of the montmorillonite group. This group includes various mineral species with different compositions. As shown in Figure 1, no smectite particles appeared in the IP read image at all, indicating that the adsorption capacity is very weak. Therefore, we prepared several smectites and compared their adsorption capacity using the same procedure (Figure 2a). In the experiment, 137Cs adsorbed by montmorillonite was the least, while the iron montmorillonite SWa-1 adsorbed the highest amount of cesium. Soapstone and chlorite are at an intermediate level. In our previous work, this iron montmorillonite was also found in the rice fields of Fukushima’s restaurant village as a Cs adsorbent next to WB. It was also found that SWa-1 adsorbed more 137Cs than other clay minerals except WB (Figure 2b). However, when SWa-1 and WB coexist on the same substrate, its adsorption capacity for SWa-1 can also be ignored (Figure 2c).

Use (a) several types of montmorillonite, (b) iron montmorillonite (SWa-1) and several mineral types other than montmorillonite, (c) WB and SWa-1 to cover the IP of the substrate Read out the image. For each mineral species, four particles are arranged on the substrate. The radioactivity input into the solution is 1.85 Bq, and the reaction time is 24 hours. The meaning of the figure on the upper right of each image and the abbreviation of the mineral are the same as in Figure 1.

From the estimates of the total amount of radioactive cesium absorbed in minerals from the IP images, it is obvious that in these experiments, the radioactive cesium in the solution was not depleted by WB adsorption (Figures 1 and 2). Therefore, it is suspected that the reason for the low amount of Cs adsorbed in other minerals coexisting with WB is that these minerals have low Cs affinity at low cesium concentrations. However, Figure 2 shows that SWa-1 without WB on the same substrate adsorbs much more cesium than with WB, which means that radioactive cesium is consumed by adsorption to WB. One explanation for this result is that most of the cesium is actually adsorbed by WB and cesium is depleted in the solution, but a considerable amount of cesium adsorbed on WB is not tightly adsorbed; therefore, wash the solution before IP autoradiography When, remove it with the solution.

In addition to the amount of cesium adsorbed, the time required for WB and SWa-1 to complete adsorption is also different. Figure 3 shows the IP readout images of WB and SWa-1 after immersion for 1 hour, 1 day, and 1 week, and shows the percentage of 137Cs adsorbed in minerals. With the passage of time, the amount of cesium adsorbed by WB gradually increased, but it was almost completely adsorbed to SWa-1 after 1 hour, and no further adsorption occurred, which indicates that there are differences in adsorption sites and/or mechanisms between the two minerals . In Figure 3a, the amount of cesium adsorbed on WB after one week seems to be greater than after one day, indicating that the adsorption lasted more than one day. However, it should be noted that the mineral particles in Figure 3 are not the same for each run. It is possible that the third WB particle running from the top within a week has extremely high adsorption capacity, and the adsorption capacity of WB within one day and one week may be equivalent.

The four mineral particles of the IP readout image matrix (top) WB and (bottom) SWa-1 covering the substrate reacted with the 1.85 Bq 137Cs solution for different times.

Please note that only one mineral species is placed on the substrate, and the mineral particles are different for each run.

Desorption experiments were carried out using WB and SWa-1 Cs adsorption particles and various electrolyte solutions for elution. We chose four electrolytes; 1 mol L-1 ammonium acetate (CH3COONH4), 1 mol L-1 cesium chloride (CsCl), 1 mol L-1 magnesium nitrate (Mg(NO3)2) and 0.1 mol L- 1 Hydrochloric acid (HCl)). 1 mol L-1 of ammonium acetate is a typical solution for extracting ion-exchange cations from soil samples7,27. Cesium chloride was used to check the isotope exchangeability of the adsorption site. According to reports, magnesium nitrate can effectively remove cesium adsorbed in vermiculite. Finally, choose 0.1 molL-1 (pH 1) hydrochloric acid to detect whether the cesium adsorbed in the mineral is desorbed by the gastric juice of the organism. Desorption was performed by immersing mineral particles with adsorbed radioactive cesium in these 50 μL of each solution for one day. The results are shown in Figure 4. The radioactive cesium in WB is difficult to reduce in ammonium acetate and cesium chloride solutions, but it is reduced to about half in magnesium nitrate and hydrochloric acid. In contrast, half of the adsorbed cesium in SWa-1 was eluted by ammonium acetate and cesium chloride, and almost completely by magnesium nitrate and hydrochloric acid. Finally, the desorption experiment was performed with WB particles. As shown in Figure 3, 137Cs was adsorbed into it. The reaction time was 1 hour, 1 day, and 1 week, and "natural" radioactive WB particles collected from Fukushima were used (Figure 5) The percentage of cesium desorbed in the particles that have been adsorbed for one hour is greater than that of the particles that have been adsorbed for one week, but the amount of 137Cs adsorbed is much higher in the particles that have reacted for one week. Although the actual mechanism is unclear, there may be an "aging" effect to increase the affinity of adsorption over time. On the other hand, the actual WB particles that absorbed radioactive cesium in Fukushima released almost half of the radioactive cesium through the hydrochloric acid solution. Since these particles have been in the field for more than two years, it is expected that the adsorbed radioactive cesium is more immobilized than the laboratory prepared for one week of processing, but in fact there is no significant difference between the field samples and those samples. In the laboratory. The reason is not clear, but it may be due to the different weathering stages between the two WB samples.

IP reads out the image matrix, showing the characteristics of desorbing 137Cs from (top) WB and (bottom) SWa-1 using various electrolyte solutions.

Initially, 137Cs was adsorbed onto mineral particles from 50 μL of solution at a concentration of 1.85 Bq for one day. Elution was performed by immersing the substrate in an electrolyte solution (50 μL) for one day. The number at the bottom of each image indicates the percentage of 137Cs remaining in the mineral particles after immersion.

The IP readout image showed that the HCl solution (pH 1) desorbed 137 Cs from the WB particles, and it adsorbed 137 Cs from the solution of different reaction cycles (18.5 Bq) and the "natural" radioactive WB particles collected from Fukushima.

Note that the image before elution is the same as the image at the top of Figure 3, except for "natural" particles. The total radioactivity before elution is shown at the top, and the percentage of 137Cs remaining in the mineral particles after elution is shown at the bottom.

In summary, the experimental results in this study clearly show that if the substance is present in the soil, WB can become the main adsorbent of radioactive cesium, and the kinetics (retention/diffusion) of radioactive cesium in the soil strongly depends on the presence or absence of WB or Do not. A recent work29 reported that in Fukushima, the transfer factor of radioactive cesium from soil to rice is completely different between vermiculite-containing and vermiculite-free soil. Our results clearly support the view that this phenomenon actually occurred in Fukushima.

WB and its original (fresh biotite: FB) were collected from an outcrop in Ono City, Fukushima, Japan, where an older type of Abukuma granite30 granodiorite was exposed. The area is located approximately 40 kilometers southwest of FDNPP. The width and thickness of the biotite crystals are only a few millimeters. Through electron probe microanalysis, the composition of FB was determined to be (K0.88Na0.01Ca0.01)(Fe1.33Mg1.08Mn0.02Al0.24Ti0.16)(Si2.83Al1.17)O10(OH)2. There are various WB crystals in different weathering stages, and their changes mainly depend on their position in the outcrop. In this study, we chose the one that seemed most weathered. Although there are significant differences in the chemical composition between individual crystals and within the crystals (Figure S2), the average composition is roughly expressed as (K0.16,Ca0.07)(Fe0.71Mg0.64Al0.25Ti0.22) (Si2. 67Al1 .33) O10(OH)2, showing the characteristics of two octahedrons, iron oxidation and potassium leaching through the hydration or "vermiculization" of the intermediate layer. The XRD spectra of WB and FB are shown in Figure S3.

For several clay minerals, clay specimens distributed by the Japan Clay Science Society (JCSS) were used. They are montmorillonite (JCSS-3101), soapstone (JCSS-3501) and illite (JCSS-5101)31. The iron montmorillonite and nontronite used are SWa-1 and NG-1, respectively, provided by the source clay minerals of the Clay Mineral Society. The kaolinite specimen was produced in a small country in Yamagata Prefecture, Japan. X-ray diffraction shows that the sample is completely single mineral and has an orderly accumulation. Halloysite comes from Eureka 33, Nevada, USA. The diaspore specimens are from Kitakami 34, Iwate Prefecture, Japan, and the Imogenites are from Kurayoshi City, Tottori Prefecture, Japan 35.

The biotite crystals and consolidated clay minerals were crushed and sieved to 25-53 μm in an agate mortar. The fine powdered minerals are compacted into granules with a hydraulic press, and then crushed and sieved to obtain similar particle sizes. The typical appearance of these mineral particles is shown in Figure S4.

Dilute the carrier-free 137Cs solution (nominal 3.7 MBq mL-1, Eckert & Ziegler Isotope Products, California, USA) to 370 Bq mL-1 (8.4 × 10-10 molL-1), 37 Bq mL-1 (8.4 × 10−11 molL−1) and 3.7 Bq mL−1 (8.4 × 10−12 molL−1) and reverse osmosis water. Since the 137Cs solution is a "reactor produced" isotope product, it may contain an amount of 133Cs equivalent to the 137Cs in the solution.

Select mineral particles of similar size under a stereo microscope and arrange them on the acrylic substrate using Kapton double-sided tape, using vacuum tweezers connected to a micro-manipulator (Quick Pro, Micro Support Co., Ltd.) (Figure S1) . Use a micropipette to drop 50 μL of 137C solution containing 0.185, 1.85, and 18.5 Bq onto the minerals on the substrate. Then seal the substrate in a styrene box to avoid drying, and then gently rinse with about 10 mL of running water. For desorption experiments, Cs adsorbed minerals react with four electrolyte solutions; CH3COONH4 (1 molL-1), CsCl (1 molL-1), Mg(NO3)2 (1 molL-1) and HCl (0.1 molL-1) . The pH values ​​of these solutions were 6.97, 4.26, 5.06, and 1.06, respectively. Drop each 50 μL of the solution onto the substrate and rinse gently after 24 hours.

The adsorption of 137C in minerals was estimated by IP autoradiography. The substrate containing minerals was covered with a 20 μm thick water-soluble flat sheet and exposed to IP (BAS-MS, Fujifilm) in the dark for 24 hours. Use an IP reader (FLA-7000, Fuji Film) to scan the generated IP image. Use Gatan Digital Micrograph 3.10.0 to color the read image to be linearly proportional to the IP signal. The amount of 137C in each mineral particle was estimated based on the calibration curve between the IP signal integrated at the point formed by the radioactive particle and the radioactivity of the particle measured by a germanium radiation spectrometer (Figure S5). On the other hand, a germanium radiation spectrometer was used to measure the total amount of 137Cs on the substrate and compare it with the total amount of 137Cs estimated from a single particle on the same substrate based on the IP signal. The difference between the two measurements is 13%, which shows the accuracy of using IP to estimate the amount of 137C.

How to cite this article: Mukai, H. etc. The cesium adsorption/desorption behavior of clay minerals considering the actual pollution conditions of Fukushima. science. Rep. 6, 21543; doi: 10.1038/srep21543 (2016).

Sawhney, BL Selective adsorption and immobilization of cations by clay minerals-a review. Clay clay miner. 20, 93-& (1972).

Francis, CW & Brinkley, FS cs-137 Preferential adsorption of mica minerals in contaminated freshwater sediments. Nature 260, 511–513 (1976).

CAS Article ADS Google Scholar 

Rieder, M. etc. The nomenclature of mica. can. mineral. 36, 905–912 (1998).

Brouwer, E., Baeyens, B., Maes, A. & Cremers, A. Cesium and rubidium ion balance in illite clay. J. Physics. Chemistry 87, 1213–1219 (1983).

Comans, RNJ, Haller, M. & Depreter, P. Adsorption of cesium on illite-non-equilibrium behavior and reversibility. Geochemistry. Cosmic chemistry. Journal 55, 433–440 (1991).

CAS Article ADS Google Scholar 

Poinssot, C., Baeyens, B. & Bradbury, MH Experimental and modeling study of cesium adsorption on illite. Geochemistry. Cosmic chemistry. Journal 63, 3217–3227 (1999).

CAS Article ADS Google Scholar 

de Koning, A. & Comans, RNJ The reversibility of the adsorption of radioactive cesium on illite. Geochemistry. Cosmic chemistry. Journal 68, 2815–2823 (2004).

CAS Article ADS Google Scholar 

Fuller, AJ etc. Incorporation and retention of cesium in the illite interlayer. Application Clay Science. 108, 128–134 (2015).

Brouwer, E., Baeyens, B., Maes, A. & Cremers, A. Cesium and rubidium ion balance in illite clay. J. Physics. Chemistry 87, 1213–1219 (1983).

Evans, DW, Alberts, JJ & Clark, RA Reversible ion exchange immobilization of cesium-137 caused migration from reservoir sediments. Geochemistry. Cosmic chemistry. Journal 47, 1041–1049 (1983).

CAS Article ADS Google Scholar 

Cremers, A., Elsen, A., Depreter, P. & Maes, A. Quantitative analysis of the retention of radioactive cesium in soil. Nature 335, 247–249 (1988).

CAS Article ADS Google Scholar 

Bradbury, MH & Baeyens, B. A generalized adsorption model for the concentration-dependent absorption of cesium by argillaceous rocks. J Contam. Hydraulic. 42, 141–163 (2000).

Fuller, AJ etc. The ionic strength and pH-dependent multi-site adsorption of Cs on mica aquifer sediments. Application Geochemistry. 40, 32–42 (2014).

Jackson, ML Weathering of primary and secondary minerals in soil. Translation internationally. soil. science. 4, 281–292 (1968).

Elprince, AM, Rich, CI & Martens, DC temperature and aluminum hydroxy interlayer affect the adsorption of trace amounts of radioactive cesium on sediments near water-cooled nuclear reactors. Water resources. Reservoir 13, 375–380 (1977).

CAS Article ADS Google Scholar 

Nakao, A., Funakawa, S. & Kosaki, T. Hydroxy-Al polymer prevents wear edge sites of illite minerals in acid soils: A study of different weathering stages in southwestern Japan. EUR. J. Soil Science. 60, 127–138 (2009).

Goto, M. etc. The interaction of radioactive and stable cesium with hydroxy intercalated vermiculite particles in the soil of the Savannah River site in South Carolina, USA. Clay clay miner. 62, 161–173 (2014).

CAS Article ADS Google Scholar 

Zaunbrecher, LK, Cygan, RT & Elliott, WC Adsorption molecular model of cesium and rubidium on weathered mica minerals. J. Physics. Chemistry A 119, 5691–5700 (2015).

Mukai, H. et al. The form of radioactive soil particles in the Fukushima contaminated area was determined by IP autoradiography and microanalysis. environment. science. technology. 48, 13053–13059 (2014).

CAS Article ADS Google Scholar 

Komarneni, S. & Roy, ​​R. Cesium selective ion sieve made by topological leaching of Phlogopite. Science 239, 1286–1288 (1988).

CAS Article ADS Google Scholar 

National Institute of Materials Science (NIMS). Adsorbent database for the purification of radioactive materials after the Fukushima Daiichi Nuclear Power Plant accident. Available at: http://reads.nims.go.jp/index_en.html (accessed on December 28, 2015).

Ohnuki, T. & Kozai, N. Adsorption behavior of radioactive cesium by non-mica minerals. J. Nuclear. science. technology. 50, 369–375 (2013).

Japan Meteorological Agency. Past weather data archive (Japanese). Available at: http://www.data.jma.go.jp/obd/stats/etrn/index.php (accessed on December 28, 2015).

Japan Atomic Energy Agency (JAEA). The extended site of the radiation dose distribution map, etc. Available at: http://ramap.jmc.or.jp/map/eng/ (accessed on December 28, 2015).

Satoko, M. etc. The mineralogical characterization of radioactive particles in Fukushima soil using synchrotron radiation m-XRD. J. Miner. gasoline. Science, under review.

Fujii, E. etc. The adsorption of cesium on Fukushima paddy soil. Clay science. 19, 17-22 (2015).

Tessier, A., Campbell, PGC & Bisson, M. Sequential extraction procedure of micro-particle trace metal forms. anus. Chemistry 51, 844–851 (1979).

Morimoto, K. et al. Desorption of Cs ions embedded in vermiculite clay by cation exchange with Mg2 ions. Chemical Wright. 41, 1715–1717 (2012).

Kato, N. etc. Potash fertilizer and other materials as countermeasures to reduce the radioactive cesium content in rice: the results of emergency experiments in response to the Fukushima Daiichi nuclear power plant accident in 2011. Soil science. Plant Nutrition 61, 179–190 (2015).

Kamei, A., Takagi, T. and Kubo, K. Geology and petrology of the Abukuma granite in the Hiyama area of ​​Fukushima Prefecture in northeastern Japan. bull. Jell. Survive. Jpn 54, 395–409 (2003). (Japanese with English abstract)

Miyawaki, R. et al. Some reference data of JCSS clay specimens. Nendo Kagaku (Journal of the Japanese Society of Clay Science) 48, 158–198 (2010). (Japanese with English abstract)

Manso, A. et al. The redox mechanism of iron in dioctahedral montmorillonite: I. Refer to the crystal chemistry of nontronite for oxidation. Yes. mineral. 85, 133–152 (2000).

CAS Article ADS Google Scholar 

Kogure, T., Mori, K., Kimura, Y. & Takai, Y. Use a new TEM with a computer-assisted minimum dose system to unravel the stacked structure in the tubular halloysite. Yes. mineral. 96, 1776–1780 (2011).

CAS Article ADS Google Scholar 

Henmi, T. The effect of the ratio of SiO2-Al2O3 on the thermal reaction of hydroalane. Clay clay miner. 28, 92-96 (1980).

CAS Article ADS Google Scholar 

Yoshinaga, M. and Yamaguchi, N. Imagolite appears as a gel film in pumice and slag beds in the western and central parts of Honshu and Hokkaido. Soil science. Plant nutrition. 16, 215–223 (1970).

We thank M. Pentrak and JW Stucki for donating SWa-1 specimens, N. Matsue for donating water glass specimens, and S. Wada for donating Immo tassel specimens. Part of this work was supported by scientific research grants (15H04222, 15H02149, and 24340133) from the Japan Society for the Promotion of Science (JSPS). This work was also supported by the Fukushima Environmental Restoration Contract signed with JAEA, which was titled "Study on the Cs Adsorption and Desorption Process on Clay Minerals".

Faculty of Science, University of Tokyo, Bunkyo, 113-0033, Tokyo, Japan

Hiroki Mukai, Kiriko, Ryosuke Kikuchi, Toshihiro Kogure

Department of Agricultural and Life Sciences, University of Tokyo, Bunkyo, 113-8657, Tokyo, Japan

Atsushi Hirose, Keitaro Tanoi & Tomoko M. Nakanishi

Quantum Beam Science Center of Japan Atomic Energy Agency and Fukushima Environmental Safety Center, 1-1-1 Kouto, Sayo-gun, Sayo-gun, 679-5148, Hyogo, Japan

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

TK designed research, supervised experiments and wrote manuscripts; HM conducted most of the adsorption/desorption experiments and summarized the results. AH prepared a radioactive cesium solution and assisted HM in his experiments. SM is mainly dedicated to making calibration curves between IP signal and radioactivity. RK collected WB specimens in Fukushima. KT, TN and TY have provided many important ideas and suggestions for the study of TK.

The author declares that there are no competing economic interests.

This work has been licensed under the Creative Commons Attribution 4.0 International License Agreement. The images or other third-party materials in this article are included in the Creative Commons license of the article, unless otherwise stated in the credit line; if the material is not included under the Creative Commons license, the user will need permission from the license holder to copy The material. To view a copy of this license, please visit http://creativecommons.org/licenses/by/4.0/

Mukai, H., Hirose, A., Motai, S. etc. The cesium adsorption/desorption behavior of clay minerals considering the actual pollution conditions of Fukushima. Scientific Report 6, 21543 (2016). https://doi.org/10.1038/srep21543

DOI: https://doi.org/10.1038/srep21543

Anyone you share the following link with can read this content:

Sorry, there is currently no shareable link in this article.

Provided by Springer Nature SharedIt content sharing program

Journal of Soils and Sediments (2021)

Journal of Radioanalysis and Nuclear Chemistry (2021)

Environmental Science and Pollution Research (2021)

By submitting a comment, you agree to abide by our terms and community guidelines. If you find content that is abusive or does not comply with our terms or guidelines, please mark it as inappropriate.

Scientific Report (Sci Rep) ISSN 2045-2322 (online)