Hexavalent Actinide Extraction Using N,N-Dialkyl Amides - Industrial

May 15, 2017 - The possibility of accounting for aqueous phase actinide–nitrate or organic phase extractant–nitric acid complexation to provide a ...
0 downloads 10 Views 464KB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Hexavalent Actinide Extraction Using N,N-Dialkyl Amides Kevin McCann, Bruce J. Mincher, Nicholas C. Schmitt, and Jenifer C. Braley Ind. Eng. Chem. Res., Just Accepted Manuscript • Publication Date (Web): 15 May 2017 Downloaded from http://pubs.acs.org on May 17, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Industrial & Engineering Chemistry Research is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

Hexavalent Actinide Extraction Using N,N-Dialkyl Amides Kevin McCann,1 Bruce J. Mincher,2 Nicholas C. Schmitt,2 and Jenifer C. Braley1 1

2

Colorado School of Mines, Golden, CO, USA 80401

Idaho National Laboratory, Idaho Falls, ID USA 83415

ABSTRACT Recently, efforts towards closing the nuclear fuel cycle have considered oxidizing americium to a hexavalent, linear-dioxo cation for co-recovery with hexavalent U, Np, and Pu from molar nitric acid using solvating extraction ligands such as tributyl phosphate or diamyl amylphosphonate. This work assesses solvent extraction recovery of sodium bismuthate oxidized americium by N,N-di-(2-ethylhexyl)butyramide (DEHBA), N,N-di-(2-ethylhexyl)isobutyramide (DEHiBA), and N,N-dihexyloctanamide (DHOA). Extraction efficiency between the monoamides was found to increase in the order of DEHiBA < DHOA < DEHBA. For all monoamides, oxidized americium extraction was less than 50% from 4 M HNO3 and below. Extraction efficiency above 50% was obtained using concentrations of 5 M HNO3 or higher. The DEHBA extractant provided the highest distribution value of 5.4 at 7 M HNO3. Distribution values were found to be stable for up to 45 seconds aqueous/organic phase mixing times and indicated decreased reduction of hexavalent americium relative to separations completed with

ACS Paragon Plus Environment

1

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 17

organophosphorus extractants. The simultaneous co-extraction of U, Np, Pu, and Am was demonstrated using DEHiBA, and was found to decrease with increasing atomic number (DU > DNp > DPu > DAm). Interestingly, a break in recovery was observed where lighter actinides, U and Np, were comparably recovered relative to the heavier actinides, Pu and Am, in this study. This observation seems to be related to differences in the extracted metal complex for light actinides compared to heavy actinides. The metal:extractant dependence suggests the presence of 1:1 and 1:2 complexes for U and Np, while data collected for Pu and Am suggests 1:2 the presence of complexes.

INTRODUCTION Concepts for closing the nuclear fuel cycle typically include separating select actinides (U – Am) from other fuel components to allow for their fissioning in light water reactors (e.g., U, Pu, and possibly Np) or in fast spectrum nuclear reactors (all actinides). In such an advanced fuel cycle, radioactive fission products and activation products would be deposited into geological storage for several hundred years to allow for their decay to stable isotopes. Partitioning and transmutation of actinides substantially reduces the long term radiotoxicity and the volume of high level waste relative to a once-through fuel cycle. Advanced fuel cycles considered to date typically involve two or three post-irradiation separations. The first separation would recover U, Pu and perhaps Np from Am, Cm and fission products. Subsequent steps seek to separate Am from the lanthanide fission products. Closing the nuclear fuel cycle would be simpler if the actinides U through Am could be recovered as a group from fission products and Cm. Such an approach would offer improved proliferation resistance because Pu would not be separated from the other transuranic elements

ACS Paragon Plus Environment

2

Page 3 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

(except for Cm). In light of the significant redox instability of U3+, Np3+ and Am4+ and the weak extraction of pentavalent actinides, AnO2+, the search for a common oxidation state for U-Am leads to the hexavalent state (AnO22+). The linear dioxo cation geometry is unique to hexavalent and pentavalent actinides; this characteristic geometry impacts both redox chemistry and patterns of stability of actinide complexes. Because this geometry is not seen elsewhere in the mix of fission and activation products found in dissolved used fuel, this feature of actinide chemistry can be exploited to separate these hexavalent actinides from all other species using liquid-liquid extraction technology. These motivations have led to a renewed interest in hexavalent americium extraction for nuclear fuel cycle applications. Americium is difficult to oxidize, though once achieved, americium has comparable solvent extraction chemistry to other actinyl ions. Strong chemical oxidizing agents such as sodium bismuthate1 or copper periodate,2,3 as well electrochemical oxidation using novel, surface functionalized electrodes4 have been used to prepare Am(VI) in nitric acid solution.

Only sodium bismuthate and copper periodate have demonstrated

americium oxidation followed by its recovery as a hexavalent cation in solvent extraction processes.3,5–7 Both the recovery of hexavalent actinides (U - Am) in the presence of either copper periodate or sodium bismuthate produces decreasing extraction efficiency with increasing atomic mass.5

The most viable extractant for these separations has been diamyl amyl

phosphonate (DAAP), but DAAP will recover both tetravalent and hexavalent elements and contains phosphorus that can complicate low-level waste management after DAAP’s use in an extraction. An extractant with selectivity for just hexavalent actinides and that adheres to the CHON principal (reagents containing carbon, hydrogen, oxygen and nitrogen) would be preferential.

ACS Paragon Plus Environment

3

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 17

Monoamides have been proposed as extractants for nuclear fuel cycle applications, with advantages over phosphorus containing reagents that include greater selectivity for uranium over fission products and possibly higher radiolytic stability.8 The selectivity of a given monoamide for both hexavalent and tetravalent actinide recovery, versus just hexavalent actinide recovery, depends on the alkyl chain substitution at the carbonyl carbon. The n-alkane substituted amides have overall higher extraction efficiencies while branch-chain alkanes are used for selection of U(VI) over Th(IV), Pu(IV) and Zr(IV).9,10 For example, N,N-di-(2-ethylhexyl)butyramide (DEHBA) extracts U(VI) and Pu(IV), while N,N-di-(2-ethylhexyl)isobutyramide (DEHiBA) selectively extracts U(VI), although with lower DU than for the unbranched amide.11 Thus, for applications involving U/Pu co-extraction, DEHBA would be preferred and for applications involving selective extraction of U, DEHiBA would be preferred.

The structures of these

compounds are shown in Figure 1.

Figure 1 Monoamide structures discussed in this study: DEHBA (left), DEHiBA (middle) and DHOA (right). Since the monoamide ligands can recover hexavalent uranium, they may be able to recover other hexavalent actinides – including americium. This could make them ideal candidates for group actinide recovery of hexavalent actinides. This report considers the solvent extractionbased recovery of bismuthate-oxidized americium using n-alkane substituted DEHBA, the branched alkane DEHiBA, and also N,N-dihexyloctanamide (DHOA), a monoamide that has

ACS Paragon Plus Environment

4

Page 5 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

received extensive previous investigation as a TBP replacement compound.12 The utility of the monoamides in the simultaneous co-extraction of U, Np, Pu and Am is also demonstrated.

EXPERIMENTAL METHODS The monoamides DEHBA and DHOA were synthesized via nucleophilic attack of acyl chloride by secondary amines containing the desired alkyl chains.8 A 1.2:1 molar ratio of butyryl chloride or 1: 1 octanoyl chloride, diluted in chloroform, was added dropwise to a mixture of the secondary amine, bis(2-ethylhexyl)amine or di-n-hexylamine, diluted in chloroform and triethylamine. The dropwise addition was done in a nitrogen atmosphere at 5°C under vigorous stirring. Once mixed, the solution was refluxed for 2 hours at 60° C. The solution was then filtered and washed with water, 10 wt % sodium carbonate, and 1 M HCl to remove unreacted triethylamine as well as residual alkyl chlorides. The solution was dried with sodium sulfate overnight. To purify the product, chloroform was removed by rotary evaporation followed by vacuum distillation. Characterization by H-NMR showed the product to be greater than 99% pure. DEHiBA was purchased from Technocomm (Edinburgh, UK). The product was also characterized by NMR upon receipt and found to be greater than 99% pure. Solvent extraction experiments were completed using equal organic:aqueous phase volumes. The aqueous phase of appropriate acidity was traced with 30 µM uranium, 2 µM neptunium, 20 µM plutonium, and/or 10 µM americium, as appropriate. The radiotracers used were natural uranium,

237

Np,

239

Pu, and

243

Am, from stocks on-hand at INL. Cerium-139 radiotracer was

purchased from Eckert and Ziegler Isotope Products (Valencia CA USA) and used at an activity of 1000 Bq mL-1. A two hour pre-equilibration of the organic phase with an acidic bismuthatecontaining aqueous phase was completed prior to the extraction contact. The aqueous phase was also treated for two hours with 60 mg mL-1 sodium bismuthate powder to oxidize the actinides to

ACS Paragon Plus Environment

5

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 17

their hexavalent actinyl valence state. After these pre-treatments, the extraction was completed with a 10 sec contact at 23 ± 2 oC of an organic phase consisting of 1 M monoamide in dodecane. Short contact times are necessary to prevent Am(VI) reduction by reaction with the ligands themselves. The phases were separated with 1 min centrifugation, and then aliquots were removed for analysis. Caution! The isotopes 237Np, 239Pu, 243Am, 139Ce, and natural uranium are radioactive. All radioactive materials were handled in radioactively controlled facilities that are equipped with personal safety equipment. Distribution ratios were measured as the concentration or activity of the actinide metal in the organic phase, over the concentration or activity in the aqueous phase. Post extraction americium characterization for both phases was completed by γ-ray counting, using the 243Am 74.7 keV line and a high purity germanium detector, while

139

Ce was measured at 165.9 keV. Uranium, Np,

and Pu were determined in both phases using a Thermo X series 2 ICP-MS with a Teflon sample introduction system and platinum cones. During analysis of organic solutions, 0.2 L/min of 20 % oxygen in argon was added to the spray chamber to aid in combustion of the organic material and to reduce build up on the cones. Organic solutions were emulsified with Triton TX100 into the normal 1 % nitric acid solution used to dilute samples prior to analysis. The aqueous phases were also treated with the surfactant to ensure consistency of the sample matrix.

RESULTS AND DISCUSSION General considerations regarding Am(VI) recovery The distribution data presented below is anticipated to be largely descriptive of Am(VI). The partitioning of Am(VI) is attributed because of the limited distribution of Am(III) (vide infra), classically low distribution of pentavalent actinides,13 and thermodynamic difficulty associated with producing Am(IV).14 Previous work has provided solvent extraction evidence of Am(VI)

ACS Paragon Plus Environment

6

Page 7 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

over a wide range of nitric acid concentrations, in which bismuthate-oxidized americium was coextracted with the other hexavalent actinides by TBP 5 and DAAP.6 Spectroscopic evidence of Am(VI) production using sodium bismuthate has been shown for 0.1 M HNO3,5 and higher acid concentrations in our own unpublished work. However, oxidation yields are seldom quantitative, and contact with the aqueous phase reduces americium, mandating the short contact times used. Therefore, the distribution ratios of americium are presented with the caveat that mixed oxidation states may be present, particularly at low acid concentration. Monoamide extractants have been demonstrated to recover metal ions through solvating mechanisms, where the charge neutralization necessary for efficient metal partitioning to a neutral organic phase is afforded by co-extraction of counter anions, such as nitrate, equation 1, ⇌  +  + MNO  S

(1)

For the systems considered here, metal efficient metal recovery should happen under high acid concentrations (3-7 M) where significant amounts of nitrate are available to encourage metal partitioning. As will be discussed below, low americium partitioning at low acid concentrations can also be a result of incomplete oxidation of americium to the hexavalent state.

Am(VI) extraction by DEHBA, DEHiBA and DHOA The solvent extraction distribution ratios for bismuthate-oxidized americium from nitric acid solution using 1 M solutions of DEHBA, DEHiBA or DHOA in dodecane are shown in Figure 2. The extraction of Am(III) using DEHBA is also presented and shows no significant extraction of trivalent americium at any acid concentration. The average DAmIII was 0.002. The extraction of oxidized americium (nominally Am(VI)) was evident for all three ligands. Increasing DAm occurs with nitric acid concentration and possibly reaches a maximum at 7 M HNO3.

ACS Paragon Plus Environment

7

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 17

This americium extraction patterns are consistent with those previously reported for bismuthate-oxidized americium extraction by DAAP, where extraction attributable to Am(VI) increased with acid concentration.6 These extraction patterns are attributed to a combination of inefficient americium extraction at low acid concentrations, courtesy the low solubility of the oxidant under these conditions,15 and the necessity of forming the neutral ligand complex for recovery by the organic phase. Figure 2 also shows that DEHBA offers the highest DAm of nearly 5.5 at 7 M HNO3, versus approximately 2.1 for the other monoamides. For the comparison between DEHBA and DEHiBA, these results are consistent with those previously measured for the extraction of U(VI), where higher extraction efficiency was reported for the n-alkane substituted monoamide.11 The longer-chain DHOA behaved more similarly to the branch-chain DEHiBA with regard to americium extraction efficiency. The highest DAm of 4.5–5.5 occurred for DEHBA extraction of oxidized americium from 6–7 M HNO3, and is consistent with DAAP extraction where DAm of 3–4 are typically reported at this high acidity. Thus, the single-stage extraction efficiency for bismuthate-oxidized americium is slightly higher for DEHBA (85 %) than for DAAP extraction (80 %), using the same ligand concentration.6,7

ACS Paragon Plus Environment

8

Page 9 of 17

10 1 0.1

DAm

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

0.01 0.001 0.0001 1

2

3

4

5

6

7

8

[HNO3], (M) Figure 2 The solvent extraction distribution ratios for nominally Am(VI) over a range of nitric acid concentrations using 1 M DEHBA (squares), DEHiBA (triangles), and DHOA (circles). The extraction of non-oxidized Am(III) using DEHBA is shown for comparison (filled diamonds). The error bars shown represent ± 7% uncertainty based on repetitive extractions (N = 5 for DEHiBA and N = 2 for DEHBA). Other data shown are for a single extraction. Effect of contact time Previous work with TBP5 and DAAP6 extractants have shown that short contact times were necessary to ensure adequate extraction of americium. This is presumably due to reduction of the Am(VI) by the organic phase, though more rigorous studies regarding the species responsible for reduction or the reduction mechanism have not been outlined. The effect of contact time on Am recovery by DEHBA was considered for extraction from 6.5 M HNO3. Contact times varied from 8–45 seconds. Figure 3 shows extraction values with a 68% confidence interval (one sigma, 10% error) and 95% confidence interval (two sigma, 20%). For less than 30 second contact times no change occurs within the one sigma values, and decreases slightly at 45 seconds. When two sigma values are included, there is no loss in extraction efficiency from 8 to 45 second contact times. By comparison, the DAm decreased by one-third with increased DAAP contact time from 15 to 30 sec for extraction from 6.5 M HNO3.6

ACS Paragon Plus Environment

9

Industrial & Engineering Chemistry Research

10

DAm

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 17

1 0

10

20

30 40 Contact Time (s)

50

60

70

Figure 3 The effect of contact time on distribution ratios for the extraction of bismuthate oxidized americium from 6.5 M HNO3 at different contact times with DEHBA (closed circles) and DAAP data from reference #6 (open circles). Black error bars encompass values within one sigma (10%) and red error bars account for values within two sigma (20%). Ce (IV) extraction using DEHBA and DEHiBA Among the lanthanides, only cerium is oxidized to Ce(IV) by sodium bismuthate, and has been shown to be extracted by DAAP.6,7 The potential extraction of Ce(IV) by monoamides DEHBA and DEHiBA was also investigated here. Figure 4 shows the distribution ratios for the extraction of tetravalent cerium were high for DEHBA, but the extraction of cerium was inefficient for the branched-chain DEHiBA. This result is consistent with those reported by others for Zr(IV), Th(IV), and Pu(IV).9–11 Due to the high cerium extraction efficiency using DEHBA, a separation of Am(VI) from Ce(IV) in nitric acid concentrations of 6–7 M would require DEHiBA as a ligand, and a separation factor between Am and Ce of ~ 2 is obtained using 1 M DEHiBA. Efficient extraction of both americium and cerium using DEHBA might be preferable, followed by selective stripping of the americium with dilute hydrogen peroxide, as was previously demonstrated using DAAP.6

ACS Paragon Plus Environment

10

Page 11 of 17

100

10

DCe(IV)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

1

0.1 0

1

2

3

4

5

6

7

8

[HNO3], (M) Figure 4 The solvent extraction distribution ratios for Ce(IV) extraction using 1 M DEHiBA (open diamonds) with ± 7 % error bars for duplicate extractions; and a single set of 1 M DEHBA extractions (closed circles). Co-extraction of the actinyl ions using DEHBA and DEHiBA Any oxidant sufficient to oxidize Am(III) to Am(VI), with its standard redox couple of 1.72 V,16 should also oxidize neptunium and plutonium to the hexavalent state. Uranium already exists as U(VI) under acidic aqueous conditions. Thus, treatment of an actinide containing acidic solution with sodium bismuthate is expected to prepare a solution containing U(VI), Np(VI), Pu(VI) and Am(VI), all as their hexavalent AnO22+ cations, to enable a group hexavalent actinide extraction. Here acidic solutions containing all four actinides in tracer amounts were oxidized with sodium bismuthate and subjected to 1 M DEHiBA extraction. The results in Figure 5 indicate that americium was extracted with the same efficiency as shown in Figure 2, when in the presence of these additional actinides.

ACS Paragon Plus Environment

11

Industrial & Engineering Chemistry Research

10

D

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 17

1

0.1 3

4

5

6

7

8

[HNO3], M Figure 5 The solvent extraction distribution ratios for nominally U(VI) (squares), Np(VI) (diamonds), Pu(VI) (triangles) and Am(VI) (circles) over a range of nitric acid concentrations using 1 M DEHiBA/dodecane extraction. All data are the means of duplicate co-extractions with error bars shown of ± 7%. Previously reported co-extraction experiments using TBP as the extractant produced data with a regular decrease in extraction efficiency across the actinide series, with americium being the least well recovered.5 The same pattern, DU > DNp > DPu > DAm, is shown for DEHiBA extraction in Figure 5; however, the actinides appear in two groups with DU and DNp being similar, and DPu and DAm being similar. To further investigate this result, the extraction stoichiometry was assessed for all four metals by varying the DEHBA (0.1-0.5 M) concentration in the presence of 6.5 M HNO3, as shown in Figure 6. Although DEHiBA and DEHBA have slightly different structures, the same pattern is observed for the straight chain monoamide, with DU > DNp > DPu > DAm and again forming the two groups. The extractant dependence, possibly reflecting ligand stoichiometry, increased across the series with a range of 1.4–1.9. However, these also fell into two groups with a mean value of 1.5 ± 0.2 for U/Np, and 1.9 ± 0.1 for Pu/Am. This suggests a change in stoichiometry from mixed 1:1 and 1:2 metal:ligand complexes for U/Np to predominantly 1:2 complexes for Pu/Am. Recent

ACS Paragon Plus Environment

12

Page 13 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

studies have suggested high nitric acid concentration increases the amount of protonated monoamide ligands (HL) in the organic phase, bringing about an outer-sphere coordination environment.17,18 Similarly, elevated nitric acid conditions give rise to anionic trinitrate uranyl species, [UO2(NO3)3]-.19–21 The mixed 1:1 and 1:2 stoichiometry could be due a protonated monoamide complexing with anionic uranyl (UO2(NO3)3HL) and neptunyl species while the 1:2 ratio is synonymous with two ligands extracting the neutral hexavalent actinide dinitrate species (UO2(NO3)2L2). The results are consistent with observations made by Condamines who observed a change in U(VI):DEHBA metal:ligand stoichiometry from 2.38 at 1 M HNO3 to 1.53 at 10 M HNO3, and proposed an anionic UO2(NO3)3HL species at high nitric acid concentrations.21 For comparison to these DEHBA results, the solvation number was also measured for Am(VI) using the branch chain DEHiBA and a value of 1.9 ± 0.2 was obtained (data not shown), consistent with the results in Figure 6. Spectroscopic assessment of extracted uranium complex recovered from high acid conditions could assist in confirming the presence of protonated monoamides recovering anionic uranyl and neptunyl. It is acknowledged that accurate determination of a metal-extractant ratio using slope relationships relies on low ligand and metal concentrations, such that ligand aggregation effects and/or substantial changes in free ligand concentration do not perturb the stoichiometry under investigation. Low acid concentrations are also preferred. Here the acid concentration was kept at 6.5 M to dissolve sufficient oxidizing agent, and higher than desirable ligand concentrations of 0.1–0.5 M were necessary to ensure measurable DAm. While the extractant dependence reported here may not be indicative of the actual stoichiometry, they should be adequate to reveal comparative trends in extractant dependence as discussed above.

ACS Paragon Plus Environment

13

Industrial & Engineering Chemistry Research

The possibility of accounting for aqueous phase actinide-nitrate or organic phase extractantnitric acid complexation to provide a corrected distribution ratio was not deemed possible at this time or particularly informative. Information available in the literature regarding hexavalent transuranic-nitrate speciation in high nitric acid conditions is too limited to complete corrections for the transuranic actinides in this study, particularly americium. Condamines has previously assessed the constants for the formation of the 1:1, 1:2 and 2:1 nitric acid species, but since the organic phase extractant:nitric acid speciation would be constant, different trends in extractant dependences would not be provided by these corrections. 0 -1 1.42 ± 0.06

-2

log D

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 17

-3 1.66 ± 0.06

-4

1.85 ± 0.08

-5 1.9 ± 0.1

-6 -2.5

-2

-1.5

-1

-0.5

log [DEHBA], (M) Figure 6 Solvation number study for U(VI) (squares), Np(VI) (diamonds), Pu(VI) (triangles) and Am(VI) (circles) using a range of DEHBA concentrations from 0.10–0.50 M in dodecane with 6.5 M HNO3.

CONCLUSIONS The monoamide DEHBA is a promising ligand for use in hexavalent americium extraction. As expected, DEHBA provides increased extraction efficiency for Am(VI) over DEHiBA and DHOA. Interestingly, the recovery of hexavalent americium using DEHBA is somewhat greater than achieved using DAAP or TBP, and appears to be less reducing toward Am(VI) than those

ACS Paragon Plus Environment

14

Page 15 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

ligands. Although DEHBA provided the highest Am(VI) distribution ratios, Ce(IV) was extracted with even greater efficiency by that ligand. Cerium (IV) extraction by DEHiBA had lower distribution ratios than Am(VI) above 6.5 M HNO3. In co-extraction of hexavalent uranium, neptunium, plutonium and americium, the actinides fall into two groups with behavior of the two light and two heavy actinides being similar. This is true both for distribution ratios and extractant dependences, as illustrated by extractions with DEHiBA and DEHBA. The magnitude of the solvation number increases across these actinyl ions from 1.4–1.9, indicating a higher ligand to metal ratio in the extracted complex for the higher z-numbers. The reason for the shift in solvation number is not completely obvious currently, but could be related to a protonated monoamide extracting anionic trinitrato uranyl and neptunyl species. Future work should further examine the possibility of trinitrato actinyl recovery, characterize americium oxidation states in the organic phase using spectrophotometry and study actinide recovery in the presence of used nuclear fuel simulants. Corresponding Author: E-mail: [email protected] Acknowledgement: Work at Colorado School of Mines was funded by the US Nuclear Energy University Program under contract DENE0008289. Work at Idaho National Laboratory was sponsored by the U.S. D.O.E. Office of Nuclear Energy Fuel Cycle R&D program, Sigma Team for Advanced Actinide Recycle program under Idaho Operations contract DE-AC07-05ID14517.

REFERENCES (1)

Runde, W. H.; Mincher, B. J. Higher Oxidation States of Americium: Preparation, Characterization and Use for Separations. Chem. Rev. 2011, 111 (9), 5723–5741.

(2)

Sinkov, S. I.; Lumetta, G. J. Americium(III) Oxidation by copper(III) Periodate in Nitric Acid Solution as Compared with the Action of Bi(V) Compounds of Sodium, Lithium, and Potassium. Radiochim. Acta 2015, 103 (8), 541–552.

ACS Paragon Plus Environment

15

Industrial & Engineering Chemistry Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 17

(3)

McCann, K.; Brigham, D. M.; Morrison, S.; Braley, J. C. Hexavalent Americium Recovery Using Copper(III) Periodate. Inorg. Chem. 2016, 55 (22), 11971–11978.

(4)

Dares, C. J.; Lapides, A. M.; Mincher, B. J.; Meyer, T. J. Electrochemical Oxidation of 243Am(III) in Nitric Acid by a Terpyridyl-Derivatized Electrode. Science (80-. ). 2015, 350 (6261), 652–655.

(5)

Mincher, B. J.; Martin, L. R.; Schmitt, N. C. Tributylphosphate Extraction Behavior of Bismuthate-Oxidized Americium. Inorg. Chem. 2008, 47 (15), 6984–6989.

(6)

Mincher, B. J.; Martin, L. R.; Schmitt, N. C. Diamylamylphosphonate Solven Extraction of Am(VI) From Nuclear Fuel Raffinate Simulant Solution. Solvent Extr. Ion Exch. 2012, 30, 445–456.

(7)

Mincher, B. J.; Schmitt, N. C.; Tillotson, R. D.; Elias, G.; White, B. M.; Law, J. D. Characterizing DAAP as an Am Ligand for Nuclear Fuel Cycle Applications. Solvent Extr. Ion Exch. 2014, 32, 153–166.

(8)

Gasparini, G. M.; Grossi, G. Long Chain Disubstituted Aliphatic Amides as Extracting Agents in Industrial Applications of Solvent Extraction. Solvent Extr. Ion Exch. 1986, 4 (6), 1233–1271.

(9)

Siddall, T. H. Effects of Structure of N,N-Disubstituted Amides on Their Extraction of Actinide and Zirconium Nitrates and of Nitric Acid. J. Phys. Chem. 1960, 1863–1866.

(10)

Pathak, P. N.; Kumbhare, L. B.; Manchanda, V. K. Structural Effects in N,N-Dialkyl Amides on Their Extraction Behavior toward Uranium and Thorium. Solvent Extr. Ion Exch. 2001, 19 (1), 105–126.

(11)

Prabhu, D. R.; Mahajan, G. R.; Nair, G. M. Di(2-Ethyl Hexyl) Butyramide and di(2-Ethyl Hexyl)isobutyramide as Extractants for uranium(VI) and plutonium(IV). J. Radioanal. Nucl. Chem. 1997, 224 (1–2), 113–117.

(12)

Kulkarni, P. G.; Gupta, K. K.; Gurba, P. B.; Janardan, P.; Changrani, R. D.; Dey, P. K.; Manchanda, V. K. Solvent Extraction Behaviour of Some Fission Product Elements with DHOA and TBP under Simulated PUREX Process Conditions. Radiochim. Acta 2006, 94, 325–329.

(13)

Nash, K. L.; Braley, J. C. Challenges for Actinide Separations in Advanced Nuclear Fuel Cycles. In Nuclear Energy and the Environment; ACS Symposium Series; American Chemical Society, 2010; Vol. 1046, pp 19–38.

(14)

Asprey, L. B.; Penneman, R. A. First Observation of Aqueous Tetravalent Americium. J. Am. Chem. Soc. 1961, 83, 2200.

(15)

Hara, M.; Suzuki, S. Oxidation of Americium(III) with Sodium Bismuthate. J. Radioanal. Chem. 1977, 36, 95–104.

(16)

Bratsch, S. G. Standard Electrode Potentials and Temperature Coefficients in Water at 298.15 K. J. Phys. Chem. 1989, 18, 1–21.

ACS Paragon Plus Environment

16

Page 17 of 17

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Industrial & Engineering Chemistry Research

(17)

Acher, E.; Hacene Cherkaski, Y.; Dumas, T.; Tamain, C.; Guillaumont, D.; Boubals, N.; Javierre, G.; Hennig, C.; Solari, P. L.; Charbonnel, M.-C. Structures of Plutonium(IV) and Uranium(VI) with N , N -Dialkyl Amides from Crystallography, X-Ray Absorption Spectra, and Theoretical Calculations. Inorg. Chem. 2016, 55 (11), 5558–5569.

(18)

Acher, E.; Dumas, T.; Tamain, C.; Boubals, N.; Solari, P. L.; Guillaumont, D. Inner to Outer-Sphere Coordination of plutonium(IV) with N,N-Dialkyl Amide: Influence of Nitric Acid. Dalt. Trans. 2017, 23–26.

(19)

Smith, N. A.; Czerwinski, K. R. Speciation of the Uranyl Nitrate System via Spectrophotometric Titrations. J. Radioanal. Nucl. Chem. 2013, 298 (3), 1777–1783.

(20)

Ikeda-Ohno, A.; Hennig, C.; Tsushima, S.; Scheinost, A. C.; Bernhard, G.; Yaita, T. Speciation and Structural Study of U(IV) and -(VI) in Perchloric and Nitric Acid Solutions. Inorg. Chem. 2009, 48 (15), 7201–7210.

(21)

Condamines, N.; Musikas, C. The Extraction by N,N-Dialkylamides II. Extraction of Actinide Cations. Solvent Extr. Ion Exch. 1992, 10 (1), 69–100.

Table of Contents Figure

ACS Paragon Plus Environment

17