Journal Search Engine

View PDF Download PDF Export Citation Korean Bibliography PMC Previewer
ISSN : 1738-1894(Print)
ISSN : 2288-5471(Online)
Journal of Nuclear Fuel Cycle and Waste Technology Vol.21 No.1 pp.55-64
DOI : https://doi.org/10.7733/jnfcwt.2023.011

Phase Behavior of the Ternary NaCl-PuCl3-Pu Molten Salt

Toni Karlsson*, Cynthia Adkins, Ruchi Gakhar, James Newman, Steven Monk, Stephen Warmann
Idaho National Laboratory, 1955 N Fremont Ave, Idaho Falls, ID 83415, United States
* Corresponding Author. Toni Karlsson, Idaho National Laboratory, E-mail: Toni.Karlsson@inl.gov, Tel: +1-208-533-8230

September 6, 2022 ; January 9, 2023 ; March 7, 2023

Abstract


There is a gap in our understanding of the behavior of fused and molten fuel salts containing unavoidable contamination, such as those due to fabrication, handling, or storage. Therefore, this work used calorimetry to investigate the change in liquidus temperature of PuCl3, having an unknown purity and that had been in storage for several decades. Further research was performed by additions of NaCl, making several compositions within the binary system, and summarizing the resulting changes, if any, to the phase diagram. The melting temperature of the PuCl3 was determined to be 746.5°C, approximately 20°C lower than literature reported values, most likely due to an excess of Pu metal in the PuCl3 either due to the presence of metallic plutonium remaining from incomplete chlorination or due to the solubility of Pu in PuCl3. From the melting temperature, it was determined that the PuCl3 contained between 5.9 to 6.2mol% Pu metal. Analysis of the NaCl-PuCl3 samples showed that using the Pu rich PuCl3 resulted in significant changes to the NaCl-PuCl3 phase diagram. Most notably an unreported phase transition occurring at approximately 406°C and a new eutectic composition of 52.7mol% NaCl–38.7mol% PuCl3–2.5mol% Pu which melted at 449.3°C. Additionally, an increase in the liquidus temperatures was seen for NaCl rich compositions while lower liquidus temperatures were seen for PuCl3 rich compositions. It can therefore be concluded that changes will occur in the NaCl-PuCl3 binary system when using PuCl3 with excess Pu metal. However, melting temperature analysis can provide valuable insight into the composition of the PuCl3 and therefore the NaCl-PuCl3 system.



초록


    1. Background

    Access to accurate thermophysical property data is vital the in nuclear safety analysis, reactor design, licensing, and operation of advanced molten salt reactors (MSRs). MSRs are considered a potential game-changing technology for next-generation nuclear reactors that could be cost effective, safe, and more sustainable than traditional commercial nuclear power plants. A significant knowledge gap exists in the data for the fundamental properties relevant to fuels and coolants for MSRs which needs to be addressed in order to expedite the technical readiness level of different MSR design concepts. To this end, a proposed fuel salt, NaCl-PuCl3, and the binary system was investigated in this work.

    The authors are aware of only one other experimental study on the NaCl-PuCl3 system reported by Bjorklund et al. in 1959 in which the melting temperature of PuCl3 was determined to be 767°C, and the NaCl-PuCl3 eutectic composition contained 36mol% PuCl3 with a melting temperature of 453°C [1]. PuCl3 was reported to be blue green in color with a density of 5.708 g∙cm‒3 [2].

    There are several documented synthesis paths for producing PuCl3, with differing precursors. PuCl3 can be synthesized from the reaction of PuO2 or plutonium (III) oxalate with PCl5, SCl2, HCl-H2, COCl2, and CCl4. However, these are either not scalable reactions or do not yield a pure PuCl3 product. An aqueous synthesis route using plutonium (III) oxalate as the starting material is shown in Equations 12 [3].

    P u 2 ( C 2 O 4 ) 3 10 H 2 O + 3 C 6 C l 6 2 P u C l 3 + 3 C 3 C l 4 O + 3 C O 2 + 3 C O + 10 H 2 O
    (1)

    P u 2 ( C 2 O 4 ) 3 10 H 2 O + 6 H C l 2 P u C l 3 + 3 C O 2 + 3 C O + 13 H 2 O
    (2)

    The difficulties associated with obtaining a pure solution of Pu(III) and removing all water of crystallization complicates an aqueous preparation of PuCl3; however, despite these difficulties a 98% pure, PuCl3 has been obtained [3]. The synthesis and purification of PuCl3 is typically achieved through high temperature pyrochemical processing where the PuO2, feedstock is converted to Pu metal by direct reduction in CaCl2 through the multicycle direct oxide reduction (MDCOR) process [4]. Note the Pu metal produced by the MDCOR process is not always of high purity due to inherent impurities in the CaCl2 and the buildup of 241Am by the radioactive decay of 241Pu (t1/2 = 14.29 years). To remove 241Am, chlorine gas is sparged through the molten plutonium converting the Pu to PuCl3 which then reacts with 241Am to produce AmCl3 [5]. However, if no Am is present, this route can be used to produce PuCl3. Care must be taken during this synthesis path because incomplete chlorination will result in a PuCl3 product containing small quantities of Pu metal.

    Johnson et al. [6] studied the presence of Pu metal in PuCl3 and its effects on the melting temperature where it was determined that a eutectic transition exists at a composition of 7mol% Pu in PuCl3 and melts at 740°C. The effect on the NaCl-PuCl3 binary system when using PuCl3 containing Pu metal remain undetermined.

    2. Experimental

    For a detailed analysis of the binary NaCl-PuCl3 phase diagram, four samples were prepared using high-purity NaCl (99.99, anhydrous), and PuCl3. The PuCl3 had been in sealed storage at the at the Idaho National Laboratory as received from Los Alamos National Laboratory in 2010. The NaCl was conditioned by heating to 400°C under vacuum at 3.6 Pa for 4 hours in a distillation furnace to remove possible trace moisture impurities. The NaCl did not experience mass loss after the conditioning in the distillation furnace, indicating the material was of high purity. The NaCl and PuCl3 salts were handled inside argon-atmosphere gloveboxes where the moisture and oxygen levels were kept below 0.1 ppm and 5 ppm, respectively.

    No conditioning or purification was performed on the PuCl3 prior to use; the material was assumed to be highpurity PuCl3 but was not verified upon receipt in 2010. The nuclear material transaction report (NMTR) provided the plutonium isotopic breakdown of the PuCl3 sample as shown in Table 1, and visually, the PuCl3 appeared to be dark green-blue to grey in color. Using information from the NMTR, the PuCl3 was calculated to be 94.05wt% PuCl3 with 5.95wt% Pu-metal present. Since there were no means of PuCl3 purification at the time, the NaCl-PuCl3 phase diagram was investigated using the as received PuCl3 salt.

    Table 1

    Isotopic composition, in weight percent, of plutonium from the PuCl3 salt in 2010

    JNFCWT-21-1-55_T1.gif

    Sample preparation consisted of weighing the NaCl and PuCl3 in required proportions, then mixing the two components in a tapered, glassy carbon crucible, see Figs. 1(a) and Fig. 1(b). In total, six samples were used in this study. The compositions are provided in Table 2; two of the samples were the unary NaCl and PuCl3 salts. Four samples (S2– S5) were heated in a furnace to 850°C and cooled to an ambient temperature before being removed. All four salt mixtures were easily removed from the glassy carbon crucibles, and crucible-salt interaction was not observed. The salt buttons, see Fig. 1(c), were size reduced Fig. 1(d) using an agate mortar and pestle which was cleaned between samples. Each salt was placed in a corresponding airtight, leak-checked metal container before being transferred for analysis.

    JNFCWT-21-1-55_F1.gif
    Fig. 1

    Pictures of the salt blending process. (a) glassy carbon crucibles containing PuCl3 and NaCl, (b) PuCl3 and NaCl mixture prior to heating, (c) “button” of NaCl/PuCl3 mixture after heating, (d) pulverized eutectic button, used for thermal analysis.

    Table 2

    Compositions of the samples used to study the NaCl-PuCl3 phase diagram

    JNFCWT-21-1-55_T2.gif

    The blended samples were transferred to the Fresh Fuels Glovebox at MFC for analysis using the simultaneous thermal analyzer (STA) model STA449F3A-0171-M containing a type-S, differential scanning calorimeter-thermogravimetric (DSC/TG) sample carrier. Thermal analysis using the STA is capable of collecting DSC and TG data simultaneously for each sample, when properly calibrated. The STA was operated inside an inert argon glove box using ultra-high-purity argon protective gas and purge gas with a flow rate of 50 and 20 ml∙min–1, respectively. Non-reactive, unsealed glassy carbon crucibles with glassy carbon lids containing a small hole were used for all standards and samples. Samples were run using a temperature and sensitivity calibration, accurate in the range of 20 to 950°C, generated using five high purity calibration standards with heating rates of 10 and 2°C∙min−1. The melting or polymorphic transition temperature and peak area of each standard were calculated by averaging the onset and peak area for three heating cycles. The accuracy of the temperature calibration curve was verified using a silver standard where the experimental melting temperature was 962.0°C, within ± 1°C of the reported silver melting temperature of 961.78°C [7].

    During data collection and sample preparation, the glovebox atmosphere was maintained at less than 5 ppm O2 and 0.1 ppm moisture. The reported polymorphic, eutectic, and melting-point temperatures were determined by averaging the onset temperature (first peak) and peak temperature (subsequent peaks) from three heating cycles. The averaged transition temperature for each sample was then plotted against the heat rate, and the linear regression was used to remove the effects of thermal lag on the reported temperature value [8].

    Initially, a melting-point and stability analysis was performed on the PuCl3 to gain insight on purity and investigate the effects of Pu metal or possibly radiolysis after being in storage for several decades. Subsequent thermal analysis on the five remaining samples was performed to determine the effects of PuCl3 concentration on the melt/ transition temperature and provide insights into the stable temperature range for measuring future thermophysical properties.

    3. Results & Discussion

    3.1 Evaluation of PuCl3

    Fig. 2(a) shows the programmed temperature profile (black, dashed) for the STA along with the TG (blue) curve for the PuCl3 sample with a heating rate of 10°C∙min−1. In total, the sample experienced four heating and cooling cycles ranging from 650 to 820°C for the first heating cycle then 600 to 820°C for the remaining three cycles. After each cooling segment, the sample was held isothermally for 20 minutes before proceeding with the subsequent heating segment. At the start of the experiment, the relative sample mass was 100%, corresponding to a sample mass of 35.0 mg. During the four heating and cooling cycles, oscillations in the mass curve can be seen due to buoyancy of the sample carrier since the argon gas expands in the furnace causing this type of sample carrier to “float”. However, upon completing the experiment and returning to ambient temperature, the resulting mass gain was 0.27%, or 0.0945 mg, for the PuCl3 sample. No difference in the mass of the sample and crucible pre- and post- experiment, on a four-place analytical balance, was observed.

    JNFCWT-21-1-55_F2.gif
    Fig. 2

    STA results for PuCl3 with a heating rate of 10°C∙min−1. (a) temperature profile and TG curve, (b) first heating DSC curves: segment 1 (grey) and segment 2 (black), (c) DSC curves for heating cycles 2–4 with the first derivative of heating cycle 2 (grey dashed line), (d) cooling curves, cycles 2–4.

    When evaluating the temperature dependent changes in samples, typically the first heating curves are discarded. However, in the case of PuCl3, inspection of the first heating curve is warranted in order to observe inflections at lower temperatures and gain insight on the purity of the sample. The first heating segment occurring from 20 to 650°C in Fig. 2(b) shows no inflections or energy changes specifically below 200°C, indicating the sample had not been contaminated with moisture.

    DSC heating segments used to calculate the melting temperature of the PuCl3 are shown in Fig. 2(c). From these segments, it can be concluded that the sample did not interact with the glassy carbon crucible; otherwise, a noticeable variation (i.e., shift) in energy as a function of temperature would be observed for each heating cycle. For the PuCl3 used in this work, the singular broad peak occurring from 715 to 760°C is deemed the melting peak. To determine the onset temperature, the first derivative of the DSC curve is taken and shown for heating in Fig. 2(c). The average onset of the three heating cycles is 714°C, while the average peak-melting temperature was determined to be 749°C. Analysis of the 2°C∙min−1 heating curves yields an average melting onset temperature of 734°C and a peak-melting temperature of approximately 747°C.

    The narrow crystallization peak for each cooling cycle is shown in Fig. 2(d). Inspection of the cooling curves indicates that crystallization or solidification occurs over an approximate 12°C peak width, but temperature specific information on PuCl3 solidification cannot be extracted from the peaks due to supercooling of the salt. The shape and width of the peak on cooling indicate a pure sample with a liquid-to-solid transition and no evidence of a solid solution or solid transition. Cooling curves suggest a pure sample while the broad peak and lower experimental melting temperature extracted from the heating curve raise concerns over the purity of the PuCl3 and presence of Pu metal.

    Integrating peak areas for the heating (Fig. 2(c)) and cooling (Fig. 2(d)) curves of PuCl3 yield the enthalpy of fusion ( Δ H f o ) and crystallization ( Δ H c o ), respectively. The average experimentally determined values are Δ H f o = 102.2 ± 10 J∙g‒1 (35.8 kJ∙mol‒1) and Δ H c o = ‒101.4 ± 10 J∙g‒1 (35.5 kJ∙mol‒1). Calculated literature reported enthalpy of fusion values includes 55.0 ± 5.0 kJ∙mol‒1 [9] and 63.6 kJ∙mol‒1 [10]. The authors are unaware of any experimentally determined values for the enthalpy of fusion or crystallization for PuCl3, in correlation with the observation of others [11].

    There are several reported values for the melting temperature of PuCl3; the most widely used melting temperature value is 767 ± 2°C [1]. However, others have reported a PuCl3 melting temperature of approximately 760 ± 5°C [3] and 733 ± 10°C [12]. Another study using purified PuCl3 determined the material melted completely in the range of 760 to 765°C, and when cooled, crystallization began at 760°C [13]. The PuCl3 salt used in this study has a melting temperature of 746.5°C approximately 20°C lower than reported by Bjorklund et al. [1]. There are several possibilities, each discussed in detail, for the lower melting temperature observed in this work including: (1) moisture/oxygen contamination, (2) self-radiolysis, and (3) presence of Pu metal.

    • (1) Since the PuCl3 was stored in quality inspected, airtight, and leak-checked containers, oxygen or moisture contamination seemed unlikely. Moisture present in the PuCl3 salt upon heating would lead to the irreversible formation of PuOCl. According to the PuCl3-PuOCl phase diagram, the eutectic composition is 7mol% PuCl3 (93mol% PuOCl) with a melting temperature of 747°C [14, 15]. Natural formation of the eutectic composition is unlikely and prolonged storage woud have resulted in the majority of the PuCl3 sample being converted to PuOCl resulting in an approximate 7% mass loss in the bulk-stored sample. Intricate mass tracking and material monitoring occurs on plutonium compounds and a mass loss of this magnitude would have been documented; no documented mass loss was observed making the formation of PuOCl an unlikely scenario. In addition, slight deviations from the eutectic PuCl3-PuOCl composition would result in an increased melting temperature, which was not shown for the PuCl3 used in this study. This also suggests that the formation or presence of PuOCl was extremely unlikely.

    • (2) Radiolysis involves the breaking of chemical bonds or the direct displacement of atoms from a crystal lattice [16]. Most alpha particles produced by radioactive decay of plutonium in solids are self-absorbed [17]. 239Pu and 240Pu, the most abundant Pu isotopes in the PuCl3 sample, have half-lives of 24,000 and 6,561 years, respectively. The decay of 239Pu does not result in the release of β-particles nor high-energy γ-rays. However, low-energy γ-rays are released from the excited state 235U daughter product following alpha emission from the 239Pu parent [18]. In addition, two atomic (orbital) electrons from 239Pu are lost after the emission of the α-particle. The massenergy balance resulting from the decay of 239Pu is 222,676 MeV [18], which is mainly absorbed by the sample. Very little information on decomposition or self-radiolysis of PuCl3 exists in the literature, and the exact effects are unknown. Sufficient energy is released during 239Pu decay to thermally activate the defects in crystal structures and cause an in growth of decay products such as helium and uranium [19]. As with moisture/oxygen contamination, radiolysis would lead to a loss in sample mass. Thus, radiolysis is an unlikely but not implausible explanation for the lower PuCl3 melting temperature; however, no mass loss was recorded, and the Pu isotopes have long half-lives.

    • (3) The most probable cause for the lower observed melting temperature of the PuCl3 sample is the presence of Pu metal either from incomplete synthesis or Pu solubility in PuCl3. A calculated Pu- Cl phase diagram indicates a Pu-Cl eutectic composition of 26.8at% Pu and 73.2at% Cl with a melting temperature of 740°C [20]. The PuCl3 in this work has a melting temperature of 746.5°C; referencing the melting temperature to the calculated Pu-Cl phase diagram, two compositions of PuCl3 26.24at% Pu (73.76at% Cl) and 26.7at% Pu (73.3at% Cl) are possible.

    The simplified, Pu-PuCl3 phase diagram from Johnson et al. [6], reprinted with permission from Elsevier in Fig. 3, aids in determining the concentration of Pu present in the PuCl3. The mol% of Pu and PuCl3 for both possible compositions is shown in Table 3 along with the melting temperature at the composition determined from the Pu- PuCl3 phase diagram. The melting temperature of PuCl3 used in this study suggests there is 6.2mol% of Pu metal in the PuCl3 sample, which very closely corresponds to 5.95wt% Pu provided in the NMTR shipping information. The 6.2mol% Pu in PuCl3 has a melting temperature of 745.6°C while the 5.95wt% Pu in PuCl3 has a melting temperature of 746.5°C, matching the melting temperature of the material investigated here. The 7.87mol% Pu composition is deemed unfeasible due to the higher melting point and lack of any high-temperature peaks in the DSC curves shown in Fig. 2(c).

    JNFCWT-21-1-55_F3.gif
    Fig. 3

    Composition of the PuCl3 used in this work (JNFCWT-21-1-55I1.gif) plotted over the focused Pu-PuCl3 system. Reprinted from [6] with permission from Elsevier.

    Table 3

    Calculated values for the two possible compositions of the PuCl3 salt used in this study determined using the Pu-Cl phase diagram and the corresponding melting temperature determined in this work

    JNFCWT-21-1-55_T3.gif

    3.2 Binary Phase Diagram Study

    To further investigate the effects of PuCl3 containing Pu metal on the NaCl-PuCl3 phase diagram, six compositions on the NaCl-PuCl3 phase diagram were analyzed on the STA, including samples of PuCl3 and NaCl. A summary of the DSC curves generated for each sample with a heating rate of 10 and 2°C∙min−1 is shown in Fig. 4. For both the NaCl (Fig. 4(a)) and PuCl3 (Fig. 4(f)), salt samples with a single melting peak at both heating rates were observed, which is typical for a unary salt. The sample with 83.1mol% NaCl showed only two peaks; the first at 447°C is indicative of the eutectic transition while the second peak at 741°C correlates to the melting/liquidus temperature for that sample.

    JNFCWT-21-1-55_F4.gif
    Fig. 4

    DSC heating curves of the six samples used in this NaCl-PuCl3 phase diagram study in which Pre-Eut. = pre-eutectic peak, Eut. = eutectic peak, and MP = melting peak.

    The samples in Figs. 4(c)4(e) show three peaks, with the final peak being relatively more subtle. The first peak .abeled as the pre-eutectic peak is most likely a solid-to-solid phase transition due to the many crystal structures of Pu metal as well as the lack of reported liquid phase formation below 400°C. Analysis of the data generated from the STA yields the averaged values in Table 4 for the pre-eutectic, eutectic, and liquidus temperature, obtained by taking the average transition temperature for three heating cycles at each heating rate used in this study.

    Table 4

    Results from DSC curves generated in the STA

    JNFCWT-21-1-55_T4.gif

    The data from Table 4 were plotted against the experimental data generated by Bjorklund et al. [1] and are shown in Fig. 5. Bjorklund et al. purified the samples by bubbling HCl(g), ensuring full chlorination of the PuCl3. It has been reported that the incomplete conversion of Pu metal, hydride, or oxide to PuCl3 occurs when using a variety of synthesis paths [3] as seen in this work’s PuCl3. The resulting NaCl-PuCl3 phase diagram is of high importance for MSRs, pyroprocessing, fuel storage, and safeguarding material because the presence of unreacted metallic Pu can drastically affect the melting temperature and change the eutectic composition of the NaCl-PuCl3 binary which can in turn alter operating parameters such as temperature and Pu loading in the salt.

    JNFCWT-21-1-55_F5.gif
    Fig. 5

    NaCl-PuCl3 system developed in this work (black) with data and curve fit lines overlayed with the NaCl-PuCl3 binary data and trendlines (red) from Bjorklund et al.

    4. Conclusion

    The melting temperature of the PuCl3 used in this work, Pu-Cl phase diagram, and shipping information all aided in the conclusion that the PuCl3 in this study has approximatly 5.9–6.2mol% Pu present. This work highlights the importance of an in-depth understanding and characterization of the starting material, especially when used in nuclear applications utilizing the NaCl-PuCl3 system. The effects of Pu-metal due to insufficient chlorination or solubility of Pu metal in PuCl3, must be well understood, as even small concentrations will affect the melting temperature and eutectic composition. There are several important points to extract from the NaCl-PuCl3 phase diagram study using the PuCl3 containing Pu metal.

    • The presence of 5.9–6.2mol% Pu in the NaCl-PuCl3 system shifts the eutectic composition from 64 to 57mol% NaCl with a melting temperature of 449.3°C.

    • The eutectic composition of 36mol% PuCl3 determined by Bjorklund et al. melts at 453°C; however, using PuCl3 with 6mol% Pu metal, melting temperature at that eutectic composition increased by approximately 95°C, to 548°C.

    • The presence of Pu metal reduces the liquidus temperature on the PuCl3-rich side of the phase diagram and increases the liquidus temperature on the NaClrich side.

    • The production of PuCl3, especially if used to fabricate fuel salts for MSRs must ensure full conversion of Pu metal to PuCl3. Otherwise, undesired deviations from the NaCl-PuCl3 phase diagram will occur.

    • Melting temperature analysis of a PuCl3 sample is quick, nondestructive, and precise method for determining the level of chlorination while quantifying the amount of metal present in a PuCl3 sample with an accuracy of less than 0.5mol%.

    Acknowledgements

    This work was supported through the Idaho National Laboratory’s (INL) Laboratory Directed Research and Development Program under Department of Energy’s Idaho Operations Office Contract DE-AC07-05ID14517 and The Molten Salt Reactor Campaign, work package number AT- 22IN070502 “Thermochemical and Thermophysical Property Database Development–INL”. The authors would like to acknowledge and thank the staff in the Fuel Manufacturing Facility at INL for coordinating this research.

    Figures

    Tables

    References

    1. C.W. Bjorklund, J.G. Reavis, J.A. Leary, and K.A. Walsh, “Phase Equilibria in the Binary Systems PuCl3- NaCl and PuCl3-LiCl”, J. Phys. Chem., 63(10), 1774- 1777 (1959).
    2. J.H. Burns, J.R. Peterson, and J.N. Stevenson, “Crystallographic Studies of Some Transuranic Trihalides: 239PuCl3, 244CmBr3, 249BkBr3 and 249CfBr3”, J. Inorg. Nucl. Chem., 37(3), 743-749 (1975).
    3. T. S. Bakes, C.S. Garner, T.B. Johns, G.H. Noulton, and B. Weinstock. The Preperation of Plutonium Trichloride, Los Alamos National Laboratory Report, LA-112 (1944).
    4. T. Paget, J. Mcneese, K. Fife, M. Jackson, and R. Watson, “Chapter 6: Molten Salt Chemistry of Plutonium”, in: Plutonium Handbook, 201-286, D.L. Clark, D.A. Geeson, and R.J. Hanrahan, Jr., eds., American Nuclear Society, Illinois (2019).
    5. D.L. Clark, G. Jarvinen, C. Kowalczyk, J. Rubin, and M.A. Stroud, Actinide Research Quarterly, 3rd Quarter: Plutonium Processing, Los Alamos National Laboratory, New Mexico (2008).
    6. K.W.R. Johnson and J.A. Leary, “The Pu-PuCl3 System”, J. Inorg. Nucl. Chem., 26(1), 103-105 (1964).
    7. ASTM International. March 20 2018. “ASTM E967- 18: Standard Test Method for Temperature Calibration of Differential Scanning Calorimeters and Differential Thermal Analyzers.” ASTM International Homepage. Accessed Aug. 10 2022. Available from: https://www.astm.org/e0967-18.html.
    8. ASTM International. August 21 2018. “ASTM E3142- 18a: Standard Test Method for Thermal Lag of Thermal Analysis Apparatus.” ASTM International Hompage. Accessed Aug. 10 2022. Available from: https://www.astm.org/e3142-18a.html.
    9. R.J. Lemire, J. Fuger, H. Nitsche, P. Potter, M.H. Rand, J. Rydberg, K. Spahiu, J.C. Sullivan, W.J. Ullman, P. Vitorge, and H. Wanner, Chemical Thermodynamics, Volume 4: Chemical Thermodynamics of Neptunium and Plutonium, OECD Nuclear Energy Agency, ed., Elesvier, Amsterdam (2001).
    10. J.M. Cleveland, Volume 1: Plutonium Handbook, Compounds of Plutonium, O.J. Wick, ed., The American Nuclear Society, La Grange Park (1980).
    11. M. Gaune‐Escard, “Thermodynamics of Lanthanide Halides and Application to High‐Temperature Processes”, Scand. J. Metall., 34(6), 369-375 (2005).
    12. T.E. Phipps, G.W. Sears, R.L. Seifert, and O.C. Simpson, “The Vapor Pressure of Plutonium Halides”, J. Chem. Phys., 18(5), 713-723 (1950).
    13. H.P. Robinson, “Determination of the Melting Points of Plutonium (III) Chloride and Plutonium (III) Bromide”, in: The Transuranium Elements Research Papers Part II, G.T. Seaborg, J.J. Katz, and W.M. Manning, eds., 952-956, McGraw-Hill, New York (1949).
    14. M.A. Williamson and P.D. Kleinschmidt, “Thermodynamics and Sublimation Chemistry of Plutonium- Oxygen-Chlorine”, Proc. of the 15th Midwest High Temperature & Solid State Chemistry Conference, LA-UR-93-2129, June 14-16, 1993, Ithaca.
    15. J.G. Reavis and J.A. Leary, “Partial Phase Diagram of the PuCl3-PuOCl System”, J. Inorg. Nucl. Chem., 28(5), 1205-1208 (1966).
    16. D. Féron, “2-Overview of Nuclear Materials and Nuclear Corrosion Science and Engineering”, in: Nuclear Corrosion Science and Engineering, 31-56, Woodhead Publishing, Sawston (2012).
    17. J.M. Haschke, R.J. Martinez, R.E. Pruner II, B. Martinez, and T.H. Allen. Interaction of Plutonium With Diverse Materials in Moist Air and Nitrogen-Argon Atmospheres at Room Temperature, Los Alamos National Laboratory Technical Report, LA-13739-MS (2001).
    18. J.P. Adams and M.L. Carboneau. National Low-Level Waste Management Program Radionuclide Report Series Volume 2, Niobium-94, U.S. Department of Energy Report, DOE/LLW-127 (1995).
    19. B.A. Maiorov, J.B. Betts, F.F. Balakirev, and A. Migliori, “Temperature Dependent and Real Time Studies of Elastic Moduli of δ 239Pu and Alloys”, Proc. of the Plutonium Futures-The Science, LA-UR-18-25818, September 9-14, 2018, San Diego.
    20. H. Okamoto, “Cl-Pu (Chlorine-Plutonium)”, J. Phase Equilibria, 20(5), 534 (1999).

    Editorial Office
    Contact Information

    - Tel: +82-42-861-5851, 866-4157
    - Fax: +82-42-861-5852
    - E-mail: krs@krs.or.kr

    SCImago Journal & Country Rank