You are on page 1of 359

RPP-RPT-50941, Rev.

TABLE OF CONTENTS 1.0 2.0 EXECUTIVE SUMMARY .................................................................................................1 BACKGROUND INFORMATION ....................................................................................3 2.1 2.2 2.3 2.4 3.0 3.1 PROBLEM STATEMENT ......................................................................................3 PREVIOUS WORK 2010 WTP REPORT ...........................................................4 IMPACT OF 2010 WTP REPORT PISA ISSUED ..............................................5 PLUTONIUM OXIDE RESOLUTION PLANS .....................................................6 REPROCESSING FACILITIES ..............................................................................9 3.1.1 Bismuth Phosphate Process .........................................................................9 3.1.2 REDOX (Reduction Oxidation) Process ...................................................10 3.1.3 PUREX (Plutonium-Uranium Extraction) Process ....................................10 PLUTONIUM CONVERSION .............................................................................11 3.2.1 Plutonium Isolation 231-Z Isolation Building ...........................................11 3.2.2 Plutonium Finishing Plant..........................................................................11 3.2.3 PUREX N-Cell...........................................................................................12 3.2.4 Other Materials Processed at REDOX and PUREX ..................................12 WASTE TREATMENT, STORAGE, AND DISPOSAL FACILITIES ...............13 3.3.1 Discharges to Ground ................................................................................13 3.3.2 Single-Shell Tanks .....................................................................................13 3.3.3 Double-Shell Tanks ...................................................................................13 HISTORICAL PFP PROCESSING DIAGRAMS ................................................15 OVERVIEW OF PLUTONIUM PROCESSING HISTORY ................................25 231-Z ISOLATION FACILITY PLUTONIUM OPERATION ............................27 PFP PRODUCTION LINES ..................................................................................29 4.4.1 RG Line ......................................................................................................32 4.4.2 Remote Mechanical (RM) A Line .............................................................35 4.4.3 Remote Mechanical (RM) B Line..............................................................36 4.4.4 Remote Mechanical (RM) C Line..............................................................36 SCRAP RECOVERY (RECUPLEX, PRF, MT, 232-Z) .......................................39 4.5.1 RECUPLEX ...............................................................................................39 4.5.2 Plutonium Reclamation Facility (PRF) ......................................................44 4.5.3 Ion-Exchange Recovery of Plutonium .......................................................52 Official Use Only ii

GENERAL HISTORICAL DESCRIPTION OF HANFORD SITE ...................................9

3.2

3.3

4.0

PLUTONIUM FINISHING PLANT (Z PLANT) .............................................................14 4.1 4.2 4.3 4.4

4.5

RPP-RPT-50941, Rev. 0

4.6 4.7

PLUTONIUM FINISHING PLANT LABS ..........................................................54 PFP SUPPORT FACILITIES ................................................................................56 4.7.1 242-Z Waste Treatment Facility ................................................................56 4.7.2 241-Z Sump Tanks .....................................................................................64 4.7.3 241-Z-361 Settling Tank ............................................................................72 4.7.4 Trenches and Crib ......................................................................................75 4.7.5 Conclusion of Plutonium Losses from PFP to Cribs .................................83 SCRAP MATERIAL PROCESSED OUTSIDE OF PFP......................................85 PLUTONIUM PARTICLE SIZE SUMMARY .....................................................85 DIFFICULTIES WITH PLUTONIUM SOLIDS ..................................................90 4.10.1 Solids in 242-Z and PRF ............................................................................90 4.10.2 Use of Filtration and Centrifuges to Limit Loss from PFP Aqueous Waste Stream .............................................................................................93 4.10.3 Conclusions Regarding Centrifuge Operation .........................................107 4.10.4 Plutonium Tetrafluoride Carry-Over and Impact in Tank Farms ............110 OVERVIEW OF PLUTONIUM ACCOUNTABILITY METHODS .................111 PFP WASTE SOLUTION TRANSFERS TO TANK FARMS ..........................117 4.12.1 PFP Aqueous Liquid Discharges after May 1973....................................117 4.12.2 Acid Waste Discharge to 242-T May 1973December 1980 ..................117 4.12.3 Neutralized Waste Discharge to 244-TX December 1981November 2004..........................................................................................................120 4.12.4 Waste Transfer by Truck (216-Z-8, 241-Z-361, TK D9) ........................122 4.12.5 Waste from Solution Stabilization using Magnesium Hydroxide or Oxalic Acid (2003-2004) .........................................................................128 ESTIMATE OF TOTAL PLUTONIUM DISCHARGE TO TANK FARMS FROM PFP...........................................................................................................129 4.13.1 Accountable Losses 1973-2004 ...............................................................129 4.13.2 PRF MUFs and Impact on Waste ............................................................133 CONCLUSION OF PFP LOSSES INCLUDING ESTIMATE OF LARGE PARTICLE FRACTION .....................................................................................138 PROCESSES THAT DID NOT PROCESS PUO2 ..............................................145 5.1.1 Bismuth Phosphate, PUREX, REDOX, Fission Product Recovery, Uranium Recovery ...................................................................................145 5.1.2 Critical Mass Laboratory .........................................................................145 5.1.3 Review of 222-S Laboratory waste collected in the 219-S Waste Handling Facility .....................................................................................151 Official Use Only iii

4.8 4.9 4.10

4.11 4.12

4.13

4.14 5.0

FUEL PROCESSING FACILITIES (T/B PLANT AND REDOX AND PUREX) ........145 5.1

RPP-RPT-50941, Rev. 0

5.1.4 5.1.5 5.2 5.3

231-Z Filtrate Transfers to T-Plant ..........................................................152 Liquid Waste Receipts at the 204-S and 204-AR Facilities ....................154

PROCESSES THAT HAD PUO2 BUT DID NOT DISCHARGE IT TO TANK FARMS ....................................................................................................155 PROCESSES THAT MAY HAVE DISCHARGED PUO2 TO TANK FARMS ................................................................................................................160 5.3.1 REDOX and PUREX Processing of PFP Nitrate Solutions ....................160 5.3.2 Miscellaneous Fuel Campaigns ...............................................................170 CONCLUSION TO NON-PFP PROCESSES .....................................................181 INTRODUCTION TO TANK FARMS CHARACTERIZATION .....................183 CHARACTERIZATION OF WASTE IN TANKS SY-102 AND TX-118 ........184 6.2.1 Tank SY-102 History and Characterization .............................................184 6.2.2 Tank TX-118 History and Characterization Data ....................................188 PLANNED LABORATORY WORK .................................................................189 RECOMMENDATIONS .....................................................................................190

5.4 6.0 6.1 6.2

TANK FARM CHARACTERIZATION AND CHEMISTRY .......................................182

6.3 6.4 7.0 8.0 9.0 10.0

REVIEW PROCESS........................................................................................................191 CONCLUSION ................................................................................................................195 REFERENCES ................................................................................................................199 APPENDICES .................................................................................................................229

TABLE OF APPENDIXES APPENDIX A SCOPE OF THE TWO TEAMS ..................................................................... A-1 APPENDIX B LIST OF ASSUMPTIONS TO VALIDATE AND COMMENT ....................B-1 APPENDIX C 5 REPORTS......................................................................................................C-1 APPENDIX D PLUTONIUM AND ABSORBER CONCENTRATIONS IN FAST AND LOW SETTLING FRACTION ............................................................................. D-1 APPENDIX E INDEPENDENT REVIEW TEAM CHARTER .............................................. E-1 APPENDIX F PLUTONIUM INDEPENDENT REVIEW GROUP REPORT ....................... F-1 APPENDIX G TEAM BIOGRAPHIES .................................................................................. G-1 APPENDIX H REVIEW OF SOLIDS IN PRF EFFLLUENT ............................................... H-1 APPENDIX I TANK TABLES ................................................................................................. I-1 APPENDIX J PU OXIDE-METAL DISCHARGE TO TANK FARM ....................................J-1 Official Use Only iv

RPP-RPT-50941, Rev. 0

LIST OF TABLES

ES-1. 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.

Summary PuO2 and Metal Results ....................................................................................... 2 Single-Shell Tanks at Hanford ........................................................................................... 14 Double-Shell Tanks at Hanford .......................................................................................... 14 Summary of Process Timeframes....................................................................................... 24 Historical Task Numbers and Task Descriptions ............................................................... 30 Overview of PFP Plutonium Processing Activities, 1949 to 2004..................................... 31 Plutonium Scrap Streams to RECUPLEX for Plutonium Recovery (HW-35030) ............ 40 PFP Laboratory Aqueous Waste ........................................................................................ 54 Composition of Hanford CAW Solution ............................................................................ 57 CAW Composition Continuous Waste Treatment Flowsheet ............................................ 58 242-Z Tank Room Assay ................................................................................................... 61 Sources and Volumes of Discharges to 241-Z Facility in 1970 ......................................... 67 Estimated Contaminants in Low-Salt Wastes and Nominal Composition in 1969 ............ 67 241-Z Tank Values Before and After Final Cleaning ........................................................ 69 Inventory Estimates for Sludge Remaining in Tank 241-Z-361 ........................................ 74 RECUPLEX Input to Z-9 Crib ........................................................................................... 82 Summary PSD Data Used in Later Analysis ...................................................................... 90 Examples of Build up of Plutonium Containing Solids in 242-Z ...................................... 92 Examples of the Causes of Presence of Solids in the PRF Solvent Extraction System ..... 92 Flow Sheet Data for Prototype Solids Handling System ................................................... 98 E1W centrifuge Sludge Characterization June 1973 ....................................................... 99 Results on Centrifuged PRF CAW from 1984 and 1987 ................................................. 106 Truck Transfers 241-Z 361 to 241-TX-101 Based on Shift Logbooks Entries................ 125 Total Reported Pu discharged from PFP to Tank Farms Based on Accountability Data (in Kilograms) .......................................................................................................... 131 Splits of Total Pu Discharged as Liquid Waste................................................................ 132 Average CAW Pu Concentration during the 1980s ......................................................... 133 Plutonium Scrap and Waste Loss and MUF Data from PRF ........................................... 134 PRF Plutonium Losses and ID Data from the 1980s........................................................ 136 PRF Operating Hours, Output and Estimated Material Unaccounted for ........................ 137 Total Assumed Pu Discharged from PFP to Tank Farms with Major Process Line Activity ............................................................................................................................. 137 Total Assumed Pu Discharged from PFP Directly to Tank Farms with Percentages Assigned to Pu Waste Types ............................................................................................ 140 Kgs in Waste Stream by Form and Particle Size.............................................................. 142 Official Use Only v

RPP-RPT-50941, Rev. 0

32. 33. 34. 35. 36. 37. 38. 39. 40.

Data Used on Large Particle Determination..................................................................... 143 Pu Recovered at the Processing Plants (MT) ................................................................... 154 Plutonium Transfers from Z Plant to REDOX and PUREX ............................................ 162 Transfers of Z Plant Dissolver Solutions to REDOX and PUREX .................................. 165 Miscellaneous Fuels Dissolved at REDOX and PUREX Plants ...................................... 174 Summary of PuO2 Particle Discharges from REDOX and PUREX Plants ..................... 180 Likely Plutonium Oxide Discharges to the Tank Farms .................................................. 181 Best-Basis Inventory Values for Pu isotopes in Tanks SY-102, TX-105, TX-109, and TX-118....................................................................................................................... 183 Comparison of April 2000 Surface Grab Sample with Underlying PFP Waste and with SY-101 Sludge. ........................................................................................................ 185

LIST OF FIGURES

1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21.

Hanford Process Timeline .................................................................................................. 12 Plutonium Finishing Plant Liquid Waste Discharge Routing: 19441949 Timeframe ..... 16 Plutonium Finishing Plant Liquid Waste Discharge Routing: 19491955 Timeframe ..... 17 Plutonium Finishing Plant Liquid Waste Discharge Routing: 19551962 Timeframe ..... 18 Plutonium Finishing Plant Liquid Waste Discharge Routing: 19621964 Timeframe ..... 19 Plutonium Finishing Plant Liquid Waste Discharge Routing: 19641973 Timeframe ..... 20 Plutonium Finishing Plant Liquid Waste Discharge Routing: 19731982 Timeframe ..... 21 Plutonium Finishing Plant Liquid Waste Discharge Routing: 19831989 Timeframe ..... 22 Plutonium Finishing Plant Liquid Waste Discharge Routing: 19902004 Timeframe ....... 23 Historical Flowsheet for 231-Z .......................................................................................... 29 RG Line Flowsheet ............................................................................................................. 33 Early PFP Operation ........................................................................................................... 33 RMC Line Process Pictorial ............................................................................................... 38 RMC Line Flow Diagram Nitrate to Tetrafluoride (PFD-Z-190-00002 B-0) ................. 38 RMC Line Flow Diagram Metal Reduction (PFD-Z-190-00002 B-0) ........................... 39 Simplified Flow Diagram of RECUPLEX Recovery and Coupling Operations (HW-35030) ....................................................................................................................... 41 Simplified RECUPLEX Solvent Extraction Flowsheet (HW-35030) ............................... 43 Scrap Plutonium Recovery Sources for PRF Facility ........................................................ 45 PRF Plutonium Scrap Processing Pictorial ........................................................................ 46 Plutonium Reclamation Flowsheet (PFD-Z-180-00001 Rev. C-1) .................................... 52 Batch Waste Treatment Process ......................................................................................... 63 Official Use Only vi

RPP-RPT-50941, Rev. 0

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.

Continuous 242-Z Waste Treatment Flowsheet ................................................................. 64 Schematic drawing of a 241-Z Tank and Tank Array ........................................................ 66 Tank D4 Showing Initial Deposits and Final Deposits ...................................................... 70 May 2005 Inspection of TK D-5 Interior Floor South Side and Agitator ....................... 71 Access Cut through TK-D5 Tank Wall for Interior Cleaning ............................................ 71 Removing Deposit from TK-D8 during Decontamination and Decommissioning ............ 72 361-Z Settling Tank System ............................................................................................... 73 Overview of Z Plant Complex Liquid Waste Disposal Sites ............................................. 77 Description of the 216-Z-9 Trench and Dimensions .......................................................... 80 Plutonium Concentrations on Surface of 216-Z-9 Crib ..................................................... 81 Particle Size Distribution for Hanford Scrap Pu Oxide ..................................................... 87 Particle Size Distribution for RFETS Chloride Scrap Oxide Sent to Hanford .................. 87 Particle Size Distribution for RFETS Pu Oxide from Peroxide Calcination ..................... 88 Particle Size Distribution for Pu Oxide from Magnesium Hydroxide Precipitation and Calcination ................................................................................................................... 88 Particle Size Distribution for Pu Oxide from Burned Metal .............................................. 89 Particle Size Distributions for Hanford Oxalate Source PuO2 ........................................... 89 Wesfalia Centrifuge Sketch ................................................................................................ 94 Photo of Similar Centrifuge ............................................................................................... 94 CAW Centrifuge in 6th Floor............................................................................................. 95 Western States Centrifuges ................................................................................................ 98 Dissolver Pan Filter in MT-5 for Dissolver Solutions ..................................................... 103 Mid 1980s MT Dissolver System ..................................................................................... 105 Leaching test on MT Dissolver Centrifuge Sludge .......................................................... 109 PRF MUF Trend Analysis 19731975 (ARH-CD-411) .................................................. 135 Timeline for Hot Semiworks ............................................................................................ 149 Simplified Liquid Waste Diagram ................................................................................... 160 PFP Filtrate Recycle to REDOX Plant ............................................................................. 169 Filtrate Recycle to PUREX Plant ..................................................................................... 170 REDOX Process Flow Schematic Dissolution of Specialty Fuels ................................ 177 PUREX Process Flow Schematic Dissolution of Specialty Fuels ................................ 178

Official Use Only vii

RPP-RPT-50941, Rev. 0

ACRONYMS & ABBREVIATIONS

AFAN AL Al(NO3)3 ANN ANN ARHCO ARIES BBI BiPO4 BNW B-PID CCl4 CML CPS CSER CSS CSSG D&D DBBP DBP DL DNFSB DOR DSA DST ER ERDA HF HLSW HN HNO3/HF ID IDMS

ammonium fluoride and ammonium nitrate Analytical Laboratory Aluminum nitrate nonahydrate aluminum nitrate nonahydrate ammonium nitrate Atlantic Richfield Hanford Company Advanced Recovery and Integration Extraction System Best Basis Inventory bismuth-phosphate Battelle Northwest book-physical inventory difference carbon tetrachloride Critical Mass Laboratory criticality prevention specification criticality safety evaluation report Criticality Safety Subcommittee Criticality Safety Support Group decontamination and decommissioning dibutylbutyl phosphonate dibutyl phosphate Development Laboratory Defense Nuclear Facilities Safety Board Daily Operating Reports Documented Safety Analysis double-shell tank electroreduction Energy Research and Development Administration hydrogen fluoride high and low salt waste hydrolyamine nitrate nitric/hydrofluoric inventory difference Integrated Document Management System Official Use Only viii

RPP-RPT-50941, Rev. 0

IDMS La LaF3 LANL LLW MBA MgO MIBK MIS MOX MSMPR MT MT MTU MUF NCSE NDA Np-237 NPH PFP PISA PN PNL PNNL PPSL PRC PRF PRT PRTR PSD PSHS Pu PuF4 PuO2 PUREX

Isotopic dilution mass spectrometry lanthanum lanthanum fluoride Los Alamos National laboratory low-level waste material balance magnesium oxide methyl isobutyl ketone Material Identification and Surveillance mixed oxide mixed suspension / mixed product removal metric tons Miscellaneous Treatment metric tons uranium material unaccounted for nuclear criticality safety evaluation nondestructive assay neptunium-237 normal parrafin hydrocarbon Plutonium Finishing Plant potential inadequacy in the safety analysis partial neutralization Pacific Northwest Laboratory Pacific Northwest National Laboratory Plutonium Process Support Laboratory Plant Review Committee Plutonium Reclamation Facility pre-reduction tank Plutonium Recycle Test Reactor particle size distribution Prototype Solids Handling System plutonium Plutonium Tetrafluoride plutonium oxide Plutonium Uranium Extraction

Official Use Only ix

RPP-RPT-50941, Rev. 0

RB REDOX RFETS RG RGL RLW RMA RMB RMC S&C SC SEM SEV SRS SS&C SST SX TBP TRU TRUEX TWINS U U-233 VRT WRPS WTP XRD

Reception and Blending Reduction-Oxidation Rocky Flats Environmental Technology Site Rubber Glove Rubber Glove Line radioactive liquid waste Remote Mechanical A Remote Mechanical B Remote Mechanical C sand and crucible Slag and Crucible Scanning Electron Microscopy spherical equivalent volume Savannah River Site sand, slag, and crucible single-shell tank solvent extraction system tributyl phosphate transuranic transuranic extraction Tank Waste Information Network System uranium uranium-233 vacuum receiver tank Washington River Protection Solutions Waste Treatment and Immobilization Plant X-ray Diffraction

Official Use Only x

RPP-RPT-50941, Rev. 0

1.0

EXECUTIVE SUMMARY

Based upon the results of the M-3 mixing studies for the Waste Treatment and Immobilization Plant (WTP) pretreatment facility, a concern emerged over the settling of greater than 10 micron plutonium-bearing particles in WTP process vessels. Specifically, members of the WTP technical staff believed that the Plutonium Finishing Plant (PFP) may have transferred large, dense plutonium oxide particles to the tank farms which could then be sent to WTP for processing. Therefore, WTP management commissioned a three-person team (WTP team) of former Hanford employees with extensive PFP and tank waste experience to investigate if large, dense plutonium-bearing particles could be present in the tank farms. In February 2011, the draft report prepared by this team was given to Washington River Protection Solutions (WRPS) management for their review. The draft report indicated that there was a possibility that large, dense plutonium-bearing particles could be present in the waste tanks. WRPS concluded that the presence of such particles could exceed the analyses of the criticality safety evaluation report (CSER) for the waste tanks. In response to these findings, and because the original team did not have broad access to process documentation and no access to classified documentation, WRPS management commissioned a second team (WRPS team) to investigate the issue and determine the extent of the problem using all available sources of information. The scope of the WRPS teams review was to: Determine all source facilities that contributed plutonium oxide (PuO2) or plutonium oxalate to the tank farm Determine all the tanks that received PuO2 or oxalate Determine how much plutonium (Pu) was sent to each tank and identify its chemical and physical form Determine the particle size and density of the plutonium disposed to the tank farms, specifically looking for particles greater than 10 microns.

As part of the review, questions about the disposal of plutonium fluoride compounds (e.g., PuF4) to the Hanford waste tanks were investigated. Accordingly, the WRPS team reviewed the chemistry of Pu oxalate and Pu fluoride compounds to determine if large dense particles would persist in the alkaline tank waste environment. The investigation determined that oxalate and other crystalline forms, when not calcined and after neutralization, were converted to light, fine particles of low-solubility plutonium hydrous oxide, PuO2x H2O, and would not be an issue. In addition, the precipitated plutonium would be bonded to other particles or poisons, making a co-precipitated agglomerate. In this manner, they would be no different than the other plutonium hydrous oxide discarded as reprocessing operations losses in the tank farm. Plutonium fluoride compounds likewise were found to form finely divided low-solubility plutonium hydrous oxide. When reviewing all the processes that sent Pu to the tank farms, three facilities; PFP, Plutonium Uranium Extraction (PUREX) plant, and Reduction-Oxidation (REDOX) were found that sent Official Use Only 1

RPP-RPT-50941, Rev. 0

large dense particulate plutonium-bearing material. The materials were primarily in the form of calcined plutonium dioxide, PuO2. This team estimated that approximately 100kg of Pu was sent to the tank farms from plutonium oxide processing in different facilities. Of this quantity, approximately 30 kg was disposed as large (>10 micron) and dense calcined PuO2 and plutonium metal fines, with the balance discarded in the form of plutonium nitrate, plutonium hydroxide, or plutonium oxalate and compounds that would readily decompose in alkaline waste solution to form fine plutonium hydrous oxide. The PFP facility contributed a majority of the calcined PuO2 with approximately 23 kg plutonium sent that was greater than 10 microns. Of that 23 kg, the WRPS team estimates that PFP discarded as much as 2.5 kg of plutonium as fine metal particles from incomplete metal burning subsequent to oxide dissolution. The PUREX and REDOX facilities sent approximately 7 kgs of similar calcined plutonium oxide material. Eight tanks received an appreciable quantity (>750 g) of oxide or metal from the three facilities. They are TX-105, -109 and -118, 244-TX, SY-102, C-102, AN-101, and S-108. In addition there were eight more tanks that received minimal amounts (<400 g) of oxide or metal and they are A-105, BX-101, S-107, -111,SX-114, B-101, TX-101, and C-104. Below in Table ES-1 is a summary of the results. For a more detailed listing tabulated by see Appendix J and for information tabulated by year and facility see Appendix I. Table ES-1. Summary PuO2 and Metal Results
Tank SST 241-S-111 SST 241-SX-114 SST 241-C-104 SST 241-BX-101 SST 241-S-107 SST 241-B-101 SST 241-A-105 SST 241-TX-101 SST 241-C-102 SST 241-S-108 SST 241-TX-105 SST 241-TX-118 SST 241-TX-109 DST 241-AN-101 DST 241-SY-102 DCRT 244-TX Pu Oxide (g) 30 30 40 150 260 320 400 400 770 1000 2320 2710 4050 1600 10685 845 Pu Metal fines(g) 0 0 0 0 0 0 0 0 0 0 54 9 90 0 1965 155

Official Use Only 2

RPP-RPT-50941, Rev. 0

The size of the disposed Pu metal and dioxide particles were determined based upon the method used to process the plutonium and historical sources. The PuO2 derived from oxalate processing has a particle size of 1 to 40 microns and a particle density of 811 g/cc. The PuO2 derived from metal burning has a particle size from 40 to 100 microns and a density of 8-11 g/cc. The Pu metal fines from rework of material and scrap recovery (largely as unreacted metal) have an estimated particle size of 40 to 100 microns and a density of approximately 19 g/cc. In determining the 30 kgs of material sent to the tank farm, several margins of safety were used and are discussed later in this report. If the margins of safety were reduced, the amount of material sent to the tank farm could be lowered to approximately 18 kgs. However, this investigation has no basis to remove the margins at this time. The plutonium loss inventory estimates made by this team are consistent with the 65-130 kg of Pu (midpoint 98 kg) sent to the tank farms estimated by the WTP team. The present investigation has refined the estimates, provided information on the plutonium chemical and physical properties, particularly the density and particle size, and delineated the quantities and waste tanks that received the plutonium. The results of this team were reviewed by an independent review team of 4 experts in nuclear operations management and oversight, scientific investigation, nuclear and criticality safety, radiological characterization, and plutonium chemistry and processing. They spent approximately one month reviewing the report, receiving presentations and interviewing team personnel. The team concluded that the investigation and data analysis activities were complete, thorough and comprehensive with the conclusions being reasonable, conservative, and useful in full recognition of the uncertainties inherent with the historical waste data quality. In summary, the WPRS team estimates that approximately 30 kgs of plutonium present in the tank farms was delivered as >10-micron particulate Pu oxide and Pu metal. This inventory is located in 16 tanks; eight with minimal quantities, and eight with appreciable quantities that could challenge the CSER. Particle sizes range from 10100 microns and densities range from 811 g/cc with approximately 2.5 kg present as 19 g/cc metal.

2.0 2.1

BACKGROUND INFORMATION PROBLEM STATEMENT

The current CSER assumes that Pu in tank waste is in finely divided forms intermixed with neutron absorbers. A recent report issued by the WTP has identified the possibility that large PuO2 particles could be in the waste, having originated from PFP. The PuO2, being large and dense, could potentially segregate from the lighter or smaller neutron absorbers during mixing operations, thus invalidating the assumptions in the CSER. This phenomenon could also cause issues with the use of pulse jet mixers at WTP. There are concerns that the pulse jet mixers may not effectively mix the dense PuO2 particles above 10 microns effective spherical diameter.

Official Use Only 3

RPP-RPT-50941, Rev. 0

The purpose of this investigation is to determine the presence of PuO2 in the waste tanks. If the PuO2 exists, estimate the quantity, determine the particle size and density, and determine the current location of the PuO2 in the tank farm. The scope of this study includes both PFP and other processing facilities that could be sources of PuO2. Additionally, the potential for the creation of PuO2 in the tank farm itself (by particle growth, or decomposition of plutonium oxalate or fluoride) will also be investigated.

2.2

PREVIOUS WORK 2010 WTP REPORT

An investigation was initiated in December of 2009 by WTP Criticality Safety to review historical information on plutonium solids that were entrained in PFP aqueous waste and were subsequently transferred to the tank farms. The purpose of this study was to have Hanford Site subject matter experts review available data and records concerning discards of plutonium liquid wastes from the PFP, particularly as related to the concerns surfaced by the DNFSB about plutonium quantities, plutonium particle sizes, and plutonium particle density. The study was completed the following year and a draft report, Historical Overview of Solids in PFP Aqueous Waste Transferred to Tank Farms: Quantity of Plutonium, Particle Size Distribution, and Particle Density, was issued in June 2010 (24590-CM-HC4-W000-00176-T02-01-00002 Rev 00B) and revised in 2011 (24590-CM-HC4-W000-00176-T02-01-00001). The key findings of this review of available historical information are summarized below: Over its lifetime, PFP realized a persistent problem with plutonium solids being entrained in liquid solutions, pipelines, tanks, and waste discards. The source for most of the plutonium solids entrained in the plutonium solutions and waste streams were from the recovery and recycle of out-of-specification plutonium materials. Pu solids from scrap recovery may be significantly different from product oxide. Plutonium oxide particulates entrained in PFP waste solutions were discarded to cribs and trenches and varied in size up to 100 m in diameter. Plutonium accountability methods on liquid waste discards were biased low, presumably because the methods frequently did not adequately account for the plutonium solids that were entrained in the liquid waste. WPRS estimates that the amount of plutonium transferred from PFP to tank farms may be underestimated by a factor of approximately 2. Efforts to minimize plutonium solids in PFP solutions were implemented over time, utilizing various combinations of filters and centrifuges. These strategies helped to reduce the solids entrainment, but they did not resolve material unaccounted for (MUF) issues nor eliminate significant plutonium solids accumulation in various tanks. Actual data, on the plutonium solids entrained in plutonium waste solutions transferred to tank farms, is unavailable because these wastes were uncharacterized for plutonium solids. Official Use Only 4

RPP-RPT-50941, Rev. 0

Particle densities for plutonium oxides prepared by various methods have been reported to range from approximately 311 gm/cc. A most likely estimate of 65 kgs of plutonium was sent from PFP to the tank farms. The WTP team believes that the lowest estimate would be 30 kgs and a worst case estimate of 130 kgs. Values outside this range would be difficult to support.

The findings of the preliminary investigation indicated that plutonium processing activities in PFP generated plutonium materials during both production and scrap recovery operations whose particle size distributions included a significant fraction that was greater than 10 microns in diameter (spherical equivalent volume [SEV]). Strategies were implemented to minimize entrainment of these solid materials in the waste streams utilizing filters and centrifuges during the timeframe that waste was transferred to tank farms. Data was identified to indicate that significant amounts of these solids still remained in the waste solutions. Therefore, it is highly probable that plutonium solids with particle sizes greater than 10 microns were entrained in the waste solutions transferred to tank farms.

2.3

IMPACT OF 2010 WTP REPORT PISA ISSUED

After the WTP report was issued, WRPS evaluated its impact on the Hanford tank farms. Even though the report was a draft, WRPS management, in consultation with DOE, decided to consider the report as new information. On February 14, 2011, the Plant Review Committee (PRC) convened to review the results of the preliminary evaluation and determined that the new information in the WTP document on plutonium particles in tank farms tanks from PFP aqueous waste constitutes a Potential Inadequacy in the Safety Analysis (PISA) to the Documented Safety Analysis (DSA). As a result, the PISA was categorized as a Group 3B, Documented Safety Analysis Inadequacies, (2) SC-3 event. This review led to issuance of occurrence report EM-RP-CORPS-TANK FARM-2011-0004, Aqueous Waste Particle Size Transferred from Plutonium Finishing Plant May be Larger than Previously Evaluated. WRPS determined that a PISA exists for the Tank Farm Documented Safety Analysis because a recent WTP document casts doubt that the PuO2 particles in the aqueous waste transferred from PFP to the Tank Farms are of a size less than 10 microns equivalent spherical particles. If the PuO2 particles are large, dense particles, then they potentially could preferentially settle and concentrate in a diluted liquid environment such as exists during retrieval operations. The information in the WTP document is related to whether presence of large PuO particles might result in differential settling that would result in exceeding the 2.6 grams/l limit identified in CPS-T-149-00012, Criticality Prevention Specification. Under current management direction, the WTP report is to be treated as identifying a potential to exceed the criticality prevention specifications. A Red Arrow entry was made in the Central Shift Managers logbook. This red arrow restriction prohibits mixer pump operations and retrieval of sludge from SY-102, TX-101, TX-118 and 244-TX until a peer review and associated technical evaluation of the results of a

Official Use Only 5

RPP-RPT-50941, Rev. 0

recently issued WTP report, Potential Criticality in Hanford Tanks Resulting from Retrieval of Tank Waste, are completed. After the PRC review and the issuing of the occurrence report, the Criticality Safety Subcommittee (CSS) convened to review the issue. They spent approximately three weeks reviewing documents and interviewing people to determine how the issue should be investigated. Draft memo CCN:228133, Plan/Schedule for WTP/Tank Farms Issue Identification/Resolution, was issued and WRPS began to assemble a team to investigate the issue based upon the CSS recommendations. In mid-March members were identified and a plan of action was developed (external letter WRPS-1101477, Contract Number DE-AC27-08RV14800 Washington River Protection Solutions LLC Plutonium Oxide Resolution Plan) to resolve the issue.

2.4

PLUTONIUM OXIDE RESOLUTION PLANS

Guidance for Follow-On Investigation The authors of the WTP 2010 study indicated they did not have sufficient time, support, resources, nor access to historical Hanford documents that were publicly unavailable. They knew from previous work in PFP that some of those unavailable documents might provide significant additional information and further clarification in many areas. Based upon this information, the authors of the 2010 WTP report were asked to develop a list of assumptions and open issues that, if confirmed or nullified, would either validate, clarify, or possibly better define their findings. This listing is provided in Appendix B, List of Assumptions to Validate and Comment. The items on this list are categorized into: Statements of Fact, Assumptions, and Areas Needing Further Investigation. The WRPS team used this listing to help focus the follow-on investigation and ensure the investigation was thorough and complete. Criticality Safety Support Group/Washington River Protection Solutions Teams Since 2008, the Criticality Safety Support Group (CSSG) has been reviewing the development of the WTP nuclear criticality safety evaluation (NCSE) because WTP is a new facility design. In mid 2009, the DNFSB began questioning the possibility of Pu solids accumulation based upon the WTP mixing studies. In December 2009, the CSSG began reviewing the tank farm criticality status due to impending preparations for waste mobilization. In addition, the CSSG was requested to review the WTP mixing study for criticality implications. Based upon these reviews, in January 2010, WTP commissioned a study to determine the source and amount of Pu solids that could be present in tank waste. The draft report was given to WRPS in early February 2011. An independent review team made up of members of the CSSG was asked to evaluate the report and determine the impact. That review resulted in a draft

Official Use Only 6

RPP-RPT-50941, Rev. 0

plan/schedule being developed in late February (draft memo CCN:228133 [Appendix A, pages 47]) After the draft plan was issued, WRPS commissioned a team to conduct the investigation outlined by the independent review team. A team was assembled including the original authors of the WTP report, Pacific Northwest National Laboratory (PNNL), WTP, WRPS, and contract personnel Team members as follows: Walt Isom Ted Venetz Jacob Reynolds Cal Delegard Susan Jones David Place David Lini Robert Watrous David Bowers Richard Hoyt Tom Jones Joseph Teal Team Lead SRS Plutonium and Tank Farm processing PFP Team lead and PFP process engineer Purex/Redox Team lead and Tank Farm Waste Characterization Staff Scientist, PFP Lab and PNNL Chemist and Development Lab PFP and PNNL Purex and Tank Farm process engineer Consultant PFP processing and chemistry Consultant Purex and Redox process engineer Consultant Tank Farm and Evaporator process engineer Consultant Pu Precipitation expert and PFP processing Consultant Tank Farm and PFP liquid discharge Consultant Tank Farm and PFP processing

From the original draft plan, a separate resolution plan was developed to determine the extent of the problem (external letter WRPS-1101477 [Appendix A, pages 13]). The plan defined the scope of the review to answer four main questions: Determine all the source facilities that contributed PuO2 or Pu oxalate to the tank farms Determine all the tanks that received the PuO2 or oxalate How much was sent to each tank and its form Determine particle size and density of the Pu.

In addition, the plan would have the WRPS team investigate the disposal of plutonium fluoride compounds (e.g., PuF4) to the Hanford waste tanks. As such, the team would review the chemistry of Pu oxalate and Pu fluoride compounds to determine if they formed large dense particles when made alkaline before being sent to the tank farms. The plan set up two WRPS teams to investigate the issues. The first WRPS team (PFP) was to review the PFP facility to determine what was sent to other facilities for processing and to the tank farms as waste. The team would use members familiar with PFP operations with access to confidential documents to answer the four questions and to close open items in the 2010 report. The second WRPS team (PUREX/REDOX) was to review all other site facilities that may have received PFP material for processing or sent PuO2 or Pu oxalates to the tank farm. The investigation started by reviewing the PUREX, REDOX, Hot Semi Works, and the Critical Mass Laboratory (CML) facilities. The goal for this team was to confirm the transfers from the PFP

Official Use Only 7

RPP-RPT-50941, Rev. 0

facility to the other facilities, in addition to reviewing other processing campaigns undertaken by those facilities that could have sent PuO2 and Pu Oxalate to the tank farm. The second WRPS team used PNNL personnel to assist investigating the fate of Pu oxalate and Pu fluoride compounds in the alkaline tank waste discussed above. The SY-102 tank sample crystalline structure was the starting point to determine what it is, how they may have been formed, and its impact to this study. The teams were not assigned to investigate the Pu nitrate solutions that, when neutralized, hydrolyze to form light, fine particles. The teams were directed to use all sources of information available such as, but not limited to, classified documents, logbooks, weekly and monthly reports, reports in the Integrated Document Management System (IDMS), and interviewing current and past workers to establish how the facilities operated. The final product for this effort is a technical report that answers the questions above. This report is an overarching document that builds on the existing WTP work completed in 2010. The plan is not to revise the existing document, but to review and identify open issues/assumptions by understanding how the document was prepared. With access to documents and other resources, including classified documents unavailable to the WTP team, the WRPS team could resolve or close many of the open issues based upon new-found information. With the new information, the original WTP authors will decide if revision is needed to the WTP original document. If, after this review, documentation is unavailable to close an open issue, the WRPS team will develop a best-basis estimate with a defensible technical justification to close any remaining open issues. This report is constructed based on a chronological process history. Like the investigation, it began with the start of facility processing and continues until the facilities were shut down. The history of how the facility was operated is essential to understanding how material could have been released to the tank farms. The PFP section gives a description of the process and then reviews how the process was modified over time and how those changes affected the waste discharges. It discusses the interaction between the PFP facility and the other processing facilities and how the wastes were discharged to the tank farm. The Fuel Processing and Tank Characterization section follows the same format, reviewing all the facilities at Hanford and their capability to send Pu dioxide to the tank farm. This section contains the review of the chemistry of Pu oxalate and Pu fluoride compounds to determine if they made large dense particles when made alkaline before and after being sent to the tank farms. The purpose of the WRPS team was only to document the specifics concerning Pu oxide and its various forms that exist in the tank farms. Future studies will determine the impact and necessary corrective actions regarding WRPS waste retrieval and WTP mixing efforts.

Official Use Only 8

RPP-RPT-50941, Rev. 0

3.0

GENERAL HISTORICAL DESCRIPTION OF HANFORD SITE

The Hanford Site in Washington State was a primary source of United States plutonium production. Between 1944 and 1989, Hanford produced 60 percent of the Pu. The Hanford Site received uranium (U) billets from other DOE facilities for fuel fabrication, irradiated the fuel in eight production reactors and one co-generation reactor, and reprocessed the fuel to extract Pu and U. The Pu was converted into metal and either shipped to other sites or manufactured into weapons components. The U was shipped offsite for recycling into metal. The waste from reprocessing and manufacturing was discharged to ground and underground storage tanks. Fuel fabrication was carried out in Hanfords 300 Area from 1943 to 1986. The nuclear fuel fabricated at the Hanford Site was solid natural (later low enriched) U metal fuel elements, clad in aluminum-silicon (later Zircaloy-2)a. Fuel was then irradiated in nine reactors clustered along a 13mile stretch of the Hanford shoreline of the Columbia River, known as the 100 Areas. Reactor operations themselves did not generate tank wastes. Decontamination wastes and ion exchange resin wastes from the reactor operations were discharged to the waste tanks.

3.1

REPROCESSING FACILITIES

Reprocessing nuclear fuel evolved with time at the Hanford Site. The initial process was a co-precipitation process which was followed by two solvent extraction processes. All of these processes were enhanced to improve recovery of product and to increase throughput. 3.1.1 Bismuth Phosphate Process

Irradiated fuel elements were transported to the Sites 200 Areas for reprocessing. The earliest radiochemical processing operations at Hanford began at the 221-T Building, also known as T Plant, in December 1944. The T Plant was supported by the 224T Concentration Building, the 222T Process Control Laboratory and other associated support structures. The next reprocessing facility built at Hanford was known as the B Plant, which began processing irradiated U in April 1945. The T and B Plants used the bismuth-phosphate (BiPO4) process, a batch, co-precipitation process. The bismuth phosphate process recovered plutonium from irradiated fuel through a series of co-precipitation processes. The initial operations in the canyon building recovered plutonium from the fission products and uranium by the co-precipitation of plutonium phosphate and bismuth phosphate. After two additional precipitation steps to decontaminate the Pu, the product from the bismuth phosphate 221 Building was purified and concentrated in 224-T or 224-B. In the Concentration Building (224 Building), the product was purified, product volume

Zircaloy-2 is an alloy composed chiefly of zirconium, blended with small amounts of tin, iron, chromium and nickel. It is not a trademarked product name.

Official Use Only 9

RPP-RPT-50941, Rev. 0

was reduced, and the carrier was switched to lanthanum fluoride. The final purification was completed in the 231-Z facility where a plutonium nitrate paste was produced. The BiPO4 process was a batch operation that did not recover uranium, generated large waste volumes, and did not have a production rate that was considered adequate; which led to the development of two solvent extraction processes for reprocessing the fuel. The REDOX process, a continuous solvent extraction process, was developed first (Section 3.1.2). The B Plant was shut down in 1952, as soon as the REDOX Plant became operational. The T Plant was shut down when the PUREX plant started up in 1956 (Section 3.1.3). 3.1.2 REDOX (Reduction Oxidation) Process

The REDOX process was housed in the 202-S Building and began operating in January 1952. The REDOX process recovered both Pu and U using continuous, countercurrent solvent extraction process. The aluminum cladding on the irradiated fuel slugs was removed by a dissolution process. Then the exposed uranium fuel was dissolved in boiling nitric acid. The resulting solution was processed through a series of packed solvent extraction columns. The organic phase used was methyl isobutyl ketone (MIBK), commonly known as hexone. Through a series of oxidation-reduction operations the Pu and U were separated from each other and also from the fission products. The uranium and plutonium nitrate streams were further purified and the final plutonium product stream was transferred to the 231-Z facility for further purification. Once the REDOX plutonium product met the feed specifications for the 234-5 Building, the plutonium product from REDOX was transferred directly to the 234-5Z Building. REDOX was designed to process an average of 1 to 2.5 metric tons (MT) of natural irradiated U per day. A series of upgrades increased the production rate to 11 to 12 metric tons uranium (MTU) per day by 1958. Part of the Phase II Capacity Increase project constructed the 233-S Plutonium Concentration Building. The REDOX plant began processing slightly enriched irradiated uranium fuel assemblies, known as E-metal in 1958. During 1961-62, new anion exchange Pu concentration equipment was installed in the 233-S Building. REDOX ceased operations in 1967. 3.1.3 PUREX (Plutonium-Uranium Extraction) Process

The PUREX process was developed due to the need for even greater production capacity. PUREX extraction process was similar to REDOX process, but used tributyl phosphate (TBP) in normal parrafin hydrocarbon (NPH) as extractant and solvent, respectively. The PUREX plant was designed to process 200 MT of irradiated, aluminum-clad U per month. The plant (also known as 202-A Plant) soon surpassed its design capacity. By October 1957, PUREX demonstrated a sustained production rate of 20 MTU per day. By 1958, PUREX was processing 79% of Hanfords total Pu output. A series of reliability improvements at the PUREX plant during 1959-61 brought enlarged and reconfigured equipment. In late 1965, PUREX systems were again modified to allow the plant to sustain normal operations at four times capacity factor, or 33 MTU/day. Official Use Only 10

RPP-RPT-50941, Rev. 0

Beginning in 1963, the PUREX plant was modified to allow it to process various fuel types, including fuel from N Reactor. N Reactor fuel elements were zirconium clad solid uranium fuel tube-in-tube arrangement. The outer tube contained 0.95 enriched uranium and some of the inner tubes were enriched to 1.25 percent 235U. In 1966, annular dissolvers and a special Zirflex process were emplaced in PUREX to accommodate N Reactor fuel. The process dissolved the fuel coating (Zircaloy-2) with a mixture of ammonium fluoride and ammonium nitrate (AFAN). The uranium metal was then dissolved in nitric acid. Once PUREX began separating N Reactor fuel, the PUREX processing rate fell to about 2,000 MTU/year for N Reactor fuel, from 5,000 to 7,000 for aluminum clad fuel. In 1972, the PUREX plant began a shutdown that lasted for 11 years. In 1983, the PUREX plant started reprocessing N Reactor fuel for approximately five years. It closed for about six weeks in early 1988, and again for a year beginning in December 1988. After a short stabilization run during late 1989March 1990, the plant again closed, and was placed on standby status in October 1990. The government issued a final closure order in December, 1992, bringing the end of plutonium production and processing at the Hanford Site.

3.2

PLUTONIUM CONVERSION

Hanford developed a number of processes to convert Pu nitrate to either metal or oxides. These processes allowed for either further processing of the Pu or aided in the conversion to product forms such as metal, oxide or fabricated weapons components. 3.2.1 Plutonium Isolation 231-Z Isolation Building

The Pu nitrate product from the 224-B and T Buildings containing excessive impurities, was sent to the 231-Z Building in the 200 West Area. In 231-Z, the Pu-bearing solution was purified by a two-cycle precipitation process with hydrogen peroxide, sulfates, ammonium nitrate, and sulfuric acid. Initially the final product was concentrated Pu nitrate paste that was shipped to the Los Alamos Site. The plutonium product sent to the 234-5 Building was initially concentrated plutonium nitrate, then plutonium peroxide or plutonium oxalate. 3.2.2 Plutonium Finishing Plant

When the 234-5Z facility came online in 1949, it became the major receiver of the plutonium product from 231-Z, REDOX and PUREX. The construction of the 234-5Z facility represented a major expansion of Hanfords role in nuclear weapons production. This facility was designed to convert plutonium nitrate into metal and then into weapons components. Many Hanford documents refer to the collection of Z facilities supporting plutonium finishing activities as Z Plant. As plutonium processing capabilities expanded in areas of scrap recovery and waste treatment over the years, the expanded plutonium processing operation was referred to as the Plutonium Finishing Plant.

Official Use Only 11

RPP-RPT-50941, Rev. 0

3.2.3

PUREX N-Cell

During the PUREX shutdown period starting in the 1970s, the M-Cell and N-Cell within the PUREX plant were modified to convert Pu nitrate to Pu oxide. Preparation tanks were added to a partitioned section of M-Cell, and a precipitation, calcination, blending and oxide product can loadout process was installed in N-Cell. Additional changes were made to the loadout configuration of Q Cell. 3.2.4 Other Materials Processed at REDOX and PUREX

The PUREX and REDOX plants also processed small quantities of other irradiated materials. In 1962, J and Q Cells in PUREX were modified to separate neptunium-237 (Np-237). During 1965 and 1966, PUREX processed 664 tons of powdered thorium oxide fuel targets that had been irradiated to produce uranium-233 (U-233). Another campaign in 1970 processed 820 kilograms (kgs) of pelletized thorium oxide targets. At REDOX, mixed oxide (MOX) fuel from the Plutonium Recycle Test Reactor (PRTR), as well as pelletized U fuels from the Shippingport (Pennsylvania) commercial reactor were processed during 1963-1967. PUREX also reprocessed some PRTR fuel in 19691972. Figure 1. Hanford Process Timeline

Official Use Only 12

RPP-RPT-50941, Rev. 0

3.3

WASTE TREATMENT, STORAGE, AND DISPOSAL FACILITIES

Waste from the various reprocessing and Pu processes was categorized by the amount of the radioactivity associated with the stream. Waste streams containing the least amount of radioactivity were discharged to surface ponds and trenches. Waste streams containing moderate levels of activity were discharged to sub-surface cribs and trenches. Waste streams with the highest level radioactivity were discharged to underground tanks. 3.3.1 Discharges to Ground

The largest of the waste discharged to the ground from the process facilities (including PFP) consisted of cooling water and steam condensates that were, by design, not contaminated with radioactivity. These cooling water streams were discharged to surface ponds for evaporation and infiltration. However, over time all of the large 200-Area ponds became contaminated from various process upsets, equipment failures and errors in waste transfer operations. More contaminated waste streams from process operations were discharged to cribs and trenches (WHC-MR-0227, Tank Wastes Discharged Directly to the Soil at the Hanford Site). Before 1973, all PFP wastes had been ground-disposed. In 1973, PFP waste (excluding cooling water) began to be routed to the 241-TX tank farm. With the withdrawal of the SSTs from service, the PFP waste was later discharged to 241-SY-102. During the 1970s, additional wastes were routed to Hanford tanks, including some 300 Area laboratory wastes, decontamination wastes from N Reactor, and ion exchange resin wastes from the N fuel storage basin. As work progressed at Hanford, fewer and fewer streams were designated for ground disposal without treatment. This practice was terminated in the mid 1980s with the construction of effluent treatment facilities. 3.3.2 Single-Shell Tanks

The installation and operation of single-shell tank (SST) farms included the design and construction of individual farms, the associated chemical processing operations that generated tank waste, evolutionary features of SST farms, and various secondary waste handling operations. Table 1 provides a list of the SST farms, their capacities, and time periods of initial periods of operation. The chemical processing work isolated long-life fission products, minimized waste volumes, and eliminated free liquids. Streams with lesser radionuclide content were discharged from the tanks system to the soil column. 3.3.3 Double-Shell Tanks

After 241-AX farm, construction and use of double-shell tanks (DSTs) began. Hanfords DSTs were all constructed of carbon steel, with stronger grades of steel being used over time. Each DST had an annulus between its two walls, and was then surrounded by reinforced concrete. Some tanks were equipped to accommodate self-boiling wastes and all had leak detection capability. The majority of the waste in the DST system came from stabilization operations from the SST system. The tanks in 241-AZ tank farm also received waste from PUREX clean out operations. Official Use Only 13

RPP-RPT-50941, Rev. 0

Table 1. Single-Shell Tanks at Hanford


200 East Area B&C Tank Farms: Built 1943-44 twelve 530,000-gal tanks (each) Plus four 50,000-gal tanks (each) BX Tank Farm: Built 1946-47 twelve 530,000-gal tanks BY Tank Farm: Built 1948-49 twelve 758,000-gal tanks A Tank Farm: Built 1953-56 six 1-million-gal (Mgal) tanks AX Tank Farm: Built 1963-65 four 1-Mgal tanks 200 West Area T&U Tank Farms: Built 1943-45 twelve 530,000-gal tanks (each) Plus four 50,000-gal tanks (each) TX Tank Farm: Built 1947-48 eighteen 758,000-gal tanks TY Tank Farm: Built 1951-52 six 758,000-gal tanks S Tank Farm: Built 1950-51 twelve 758,000-gal tanks SX Tank Farm: Built 1953-55 fifteen 1-Mgal tanks

Table 2. Double-Shell Tanks at Hanford


200 East Area AY Tank Farm: Built 1968-70 two 1M-gal tanks AZ Tank Farm: Built 1970-74 two 1M-gal tanks AW Tank Farm: Built 1976-80 six 1.16M-gal tanks AN Tank Farm: Built 1977-80 seven 1.16M-gal tanks AP Tank Farm: Built 1982-86 eight 1.16-Mgal tanks 200 West Area SY Tank Farm: Built 1974-76 three 1M-gal tanks

4.0

PLUTONIUM FINISHING PLANT (Z PLANT)

This section examines the operation of the Hanford plutonium facilities over their history of operation. It is broken down into a review of the processing lines, process changes, support facilities, solids issues, and an estimate of how much plutonium may have been released to the tank farm. Process flow diagrams, with emphasis on the waste routings and how they changed over time are presented for all operating timeframes. A discussion of historical production is provided along with descriptive detail about the conversion processes (nitrate to metal or oxide) and the scrap recovery processes; which will assist in understanding how plutonium oxide was released to the tank farms. Additionally, a discussion of laboratories and other support facilities

Official Use Only 14

RPP-RPT-50941, Rev. 0

is also provided, including waste treatment facilities, tanks, and cribs that received or collected PFP waste solutions. Particle size data on plutonium in waste solutions is very limited, primarily related to soil studies on crib sites that received waste. Therefore, particle size data on PFP product and scrap oxides is presented to show the range of values expected. The difficulties experienced with solids accumulations in PFP and the waste streams, which was a life-long problem at PFP, is discussed along with methods taken to control and prevent accumulations and conclusions in regard to their effectiveness. A discussion of accountability methods used on waste streams is presented with emphasis on the problems of unmeasured particulate. The presentation of the historical data and historical problems with solids accumulation and discharge is used to focus on the magnitude of the problem that was expected to continue after May 1973 when PFP waste solutions were routed to Tank Farms. The changes in waste solution receipt and treatment are presented along with an estimate of total plutonium discharge to tank farms from PFP. The plutonium estimate includes a value for unmeasured particulate, based on a review of material un-accounted for analysis, after centrifuges were employed on the PRF waste stream. Finally the fate of each form of particulate plutonium discharges is explored and an estimate of the amount of large particulate plutonium, greater than 10 microns, is presented along with conclusions in regard to the discharge of plutonium to tank farms from the PFP.

4.1

HISTORICAL PFP PROCESSING DIAGRAMS

Figures 2 through 9 depict the flow of plutonium-containing liquid waste discharges from PFP and 231-Z over their operating lifetimes and how the processes and waste routings changed over time. The individual figures depict the individual plant operations that generated the liquid waste and the routes of the liquid wastes through the plant to the 241-Z Sump tanks to the disposal site. The diagrams were used to trace the past operations and investigate the way materials were handled, how PFP operated and how waste streams flowed during various periods of operation. A summary of the process changes shown on these diagrams is detailed in Table 3.

Official Use Only 15

RPP-RPT-50941, Rev. 0

Figure 2. Plutonium Finishing Plant Liquid Waste Discharge Routing: 19441949 Timeframe

Official Use Only 16

RPP-RPT-50941, Rev. 0

Figure 3. Plutonium Finishing Plant Liquid Waste Discharge Routing: 19491955 Timeframe

Official Use Only 17

RPP-RPT-50941, Rev. 0

Figure 4. Plutonium Finishing Plant Liquid Waste Discharge Routing: 19551962 Timeframe

Official Use Only 18

RPP-RPT-50941, Rev. 0

Figure 5. Plutonium Finishing Plant Liquid Waste Discharge Routing: 19621964 Timeframe

Official Use Only 19

RPP-RPT-50941, Rev. 0

Figure 6. Plutonium Finishing Plant Liquid Waste Discharge Routing: 19641973 Timeframe

Official Use Only 20

RPP-RPT-50941, Rev. 0

Figure 7. Plutonium Finishing Plant Liquid Waste Discharge Routing: 19731982 Timeframe

Official Use Only 21

RPP-RPT-50941, Rev. 0

Figure 8. Plutonium Finishing Plant Liquid Waste Discharge Routing: 19831989 Timeframe

Official Use Only 22

RPP-RPT-50941, Rev. 0

Figure 9. Plutonium Finishing Plant Liquid Waste Discharge Routing: 19902004 Timeframe

Official Use Only 23

RPP-RPT-50941, Rev. 0

Table 3. Summary of Process Timeframes


Time frame 19441949 Description 231-Z operations only, filtrate from peroxide precipitation recycled to 224-T 231-Z (peroxide precipitation) 234-5 oxalate precipitation, calcination and hydro-fluorination in Rubber Gloveline and RMA (1952). Filtrates recycled to 224-T and REDOX. RECUPLEX (March 1955) 233-S (April 1955) PUREX (Jan 1956) RMA Continuous Task I-II (1957) RMC Line (1960) Limited in-house scrap recovery. RECUPLEX down Filtrate recycled to REDOX/PUREX Oxide/skulls dissolved and sent to PUREX/REDOX Filtrate sent to Ion Exchange (Oct 1962) RMC in metal production RMA in oxide production Plutonium Reclamation Facility -236-Z (1964) Waste Treatment Facility 242-Z (1965) Concerns Solids in filtrate carry though process. 224-T discharges to 200 series T farm tanks Solids in filtrate carry though process. Oxide and PuF4 entrained in furnace off-gas, (Tech manual indicates 0.02 gm per 320 gm charge). Mitigation Filtrate treated with permanganate and heating Noted Off Normals

19491955

Filtrate treated with permanganate and heating.

19551962

Oxide particulate in RECUPLEX CAW (unmeasured). Solids in filtrate carry though process. Oxide and PuF4 entrained in furnace offgas. Oxide particulate in dissolver solutions sent to separations facilities. Solids carry over in filtrate. Oxide and PuF4 entrained in furnace offgas.

100 micron filters on RECUPLEX waste discharge. Sintered SS Blowback filters on calciner off-gas, carbon filters on FLUR off-gas stream Permanganate added to filtrate. Cake wash discontinued to save volume.

Greater than 3 times the accountability value was discharged to Z-9. Oxide particle found in later soils characterization work

19621964

Solids noted in filtrate and some dissolver solutions, sent back to PFP for filtration.

19641973

Oxide particulate in dissolver solutions. Solids carry over in filtrate. Oxide and PuF4 entrained in furnace off-gas.

CAW crud filter (1967) CAW centrifuge (1967) Prototype Solids Handling System Project (1973) Analytical methods found to significantly under report Pu in liquid waste (1967) Centrifuge Operation improved and specification for use improved. TK-D5 monitored for buildup using NDA

Pu buildup in 242-Z tanks led to continuous process. MUFs out of control for extended periods, blamed on solids, new centrifuges recommended. Blowback filters found corroded. One left out and oxide proliferated through PRF. Anecdotal evidence of 241-Z to 242-T transfer line noted as full of solids. Evidence of continued unmeasured losses from PRF after centrifuges was put in operation.

19731982

Waste Discharged as acid to Tank Farms -242-T RMC Shutdown, only RMA operated as oxide line. 242-Z/PRF shutdown in 1976. Cleanout PRF/RMA stabilization run in 1978 1979

Oxide particulate in dissolver solutions. Solids carry over in filtrate. Oxide entrained in furnace off-gas. Transfers of sludge from Z-8 and SN from 241-Z-361 to tank farms

Official Use Only 24

RPP-RPT-50941, Rev. 0

Table 3. Summary of Process Timeframes


Time frame 19821989 Description RMC and PRF restarted. Waste neutralized in 241-Z TK-D5 and discharged to 244-TX Concerns Oxide particulate in dissolver solutions. Solids carry over in filtrate. Oxide and PuF4 entrained in furnace off-gas. Mitigation Centrifuge operation required for operation. Ferric nitrate additions made to waste to dilute Pu solids content. Noted Off Normals Solids still causing high waste loss in PRF. RMC carbon filters checked weekly for powder plugging. Filtrate Evaporator failure results in 2500 gm loss to Tk-D4 1100 grams noted in TK-D7

19902004

Major Process lines shutdown, PRF training run aborted, many tanks flushed, Solution stabilization via MgO and oxalate precipitation.

Filtrate from precipitation discharged to tank farms without treatment. Discharge limit raised to 10X historical value (1200 gms/batch)

Absorbers credited in TK D5

4.2

OVERVIEW OF PLUTONIUM PROCESSING HISTORY

The Hanford Site was selected for the construction and operation of nuclear reactors and separation plants during the early 1940s for the production of plutonium as part of the Manhattan Project during World War II. Plutonium production, finishing, and various metal component manufacturing at Hanford were dynamic processes that continually evolved to meet the ever-changing requirements for special nuclear materials to support various defense and non-defense needs. The 231-Z Plutonium Isolation Facility, the first Z building, became operational in 1944/1945 and provided the final purification and isolation steps for plutonium from the bismuth phosphate process operating at the B and T Plants from 1944 through 1956. After 1956, this facility became the focal point for plutonium metallurgy research and development activities. The initial operation in 231-Z Isolation Facility was the conversion of plutonium incorporated within a lanthanum fluoride carrier into plutonium nitrate paste or glass for shipment offsite. A detailed description of the 231-Z separations chemistry can be found in the Hanford Technical Manual (HW-10475) and a short synopsis of 231-Z plutonium activities is provided in section 4.3 During the mid 1940s, the Hanford operating contractor was requested to initiate the design for a facility that had the capabilities to convert plutonium nitrate solutions to oxalate, oxide, and finally metal, and the capabilities to fabricate the metal into weapon components. The design of PFP was initiated in 1947, construction was initiated in 1948, and hot operations were initiated in July1949 (HNF-EP-0924, History and Stabilization of the Plutonium Finishing Plant (PFP) Complex, Hanford Site). The 234-5Z facility became the major receiver of the plutonium product from 231-Z, REDOX, and PUREX. The construction of the 234-5Z facility represented a major expansion of Hanfords role in nuclear weapons production. This facility was designed to convert plutonium

Official Use Only 25

RPP-RPT-50941, Rev. 0

nitrate into metal and then into weapons components. Many Hanford documents refer to the collection of Z facilities supporting plutonium finishing activities as Z Plant. As plutonium processing capabilities expanded in areas of scrap recovery and waste treatment over the years, the expanded plutonium processing operation was referred to as the Plutonium Finishing Plant. The PFP Complex was designed to be a self sufficient stand-alone plant for the conversion of plutonium nitrate into oxide and metal products, with all necessary scrap recovery, analytical, and developmental capabilities. The functional capabilities included such things as: Chemical processing for conversion of plutonium nitrate to oxide and metal Metallurgical processing for fabrication of weapon components Scrap processing systems for recovery of plutonium Impure nitrate purification processing Waste treatment processing Radiochemical/chemical analytical laboratories Development laboratory Secured plutonium storage (vaults and vault type rooms) Maintenance and utility support systems.

The PFP processed more than 95% of the plutonium produced at Hanford and a more thorough discussion of its process systems and processing history is provided in Sections 4.3 to 4.6. In the spring of 1962, PFP was fully engaged in conversion of plutonium nitrate solutions into metal buttons and on into formed weapons components. Scrap recovery to support this work was conducted in the RECUPLEX facility. The work was occurring at the height of the Cold War and all available site reactors and separations plants were operating. The RECUPLEX criticality incident in April 1962 (HW-74723) interrupted the production cycle because PFP scrap forms could not be processed on an as-generated basis. The most immediate problem involved dealing with filtrate from the oxalate precipitation step. This liquid stream contained an average of about two percent of the total plutonium in the feedstock to the plant. From about two weeks following the criticality event, substantial efforts were made to resume processing plutonium nitrate to metal. The temporary resolution for filtrate involved expedited construction of several hundred batch cans of nominal 10-liter capacity each, and transport of button line filtrate in those cans, by truck to the REDOX and PUREX canyon fuel processing facilities for rework in their processes. That mode of accommodating the filtrate continued until an ion-exchange capability was installed in room 232 of the 234-5Z Building. By the end of November 1962, the ion exchange equipment had picked up the filtrate processing load. Scrap solutions continued to be sent to PUREX and REDOX for recovery until PRF was started in 1964. Several additional plutonium facilities were also established at the Hanford Site. Interest in the use of plutonium for commercial nuclear reactors grew in the 1950s, which led to the construction of the Plutonium Fabrication Pilot Plant (308/308-A Bldg) in 1960 for plutonium reactor fuel development activities. This facility, which was located in the 300 Area, included Official Use Only 26

RPP-RPT-50941, Rev. 0

numerous plutonium metallurgical capabilities to support plutonium metal and alloy reactor fuel development studies, and plutonium oxide fuel fabrication capabilities to support plutonium oxide fuel development studies. This facility operated from 1960 until 1992. Prior to restart of PUREX in the early 1980s, it was decided to expand PUREX processing capabilities. This processing expansion included the ability to convert the plutonium nitrate to plutonium oxide in a new processing system to be installed in N-Cell of the PUREX canyon. This plutonium processing system became known as the PUREX Oxide Line, Project B-175, and the N-Cell Line (see section 5.4).

4.3

231-Z ISOLATION FACILITY PLUTONIUM OPERATION

The 231-Z facility operated from January 1945 until January 1957 to purify Pu (NO3)4 solutions that were produced at the T and B Plants (plutonium bismuth phosphate process) and the REDOX facility (methyl isobutyl ketone solvent extraction process). The 231-Z operations were fairly straightforward. The generalized plutonium peroxide flow sheet used at the 231-Z Facility is depicted in Figure 10 [HW-29200, Plutonium Purification and Fabrication Technical Manual, HW-26365, Brief Summary of Separations Processes, HW-10475]. The feed stream to 231-Z was plutonium in a lathanium fluoride carrier from the 224-T or 224-B facility (operational December 1944March 1956) and B Plant (operational April 1945August 1952), which contained an excessive quantity of lanthanum (La). The plutonium nitrate from REDOX (operational September 1951April 1967) that did not meet product specification for 234-5 Building was also processed at the 231-Z Facility. Once the REDOX plutonium nitrate product started meeting 234-5Z feed specification, the REDOX material went directly to the 234-5 Building. For the material from T and B plants, two peroxide precipitations were used to achieve a final product nitrate (at least 95% pure). The feed solution to 231-Z was first filtered and then treated with ammonium sulfite to reduce any hexavalent plutonium that might be present, and ammonium sulfate and hydrogen peroxide were added to precipitate plutonium peroxide. The precipitate was washed, filtered, and dissolved in nitric acid, with the corresponding decomposition of the peroxides which yielded a plutonium nitrate solution. A second precipitation was performed to get a product that would meet final product specifications. If the product was to be shipped offsite, the final plutonium nitrate solution was concentrated in a still to a thick paste. If the product was to be sent to the 234-5 Building for conversion to plutonium oxide, a final oxalate precipitation was performed. The oxalate product slurry was filtered to collect the precipitate and the filtrate was collected in a vacuum receiver (CT-1). When the RG Line Task I process was operational (1949early 1950s) plutonium nitrate was shipped to the 234-5 Building. The filtrate from 231-Z processing that was recycled to the 224-T Concentration Facility was treated with 40% NaNO2 to react with excess peroxide and finally with 4% KMnO4 until a permanent coloration appeared. All the various filtrate streams from first, second, etc., precipitations and washes were combined together and sent back for plutonium recovery. Typically, 6.38 kg of CT-1 solution contained on the order of 0.12 kg Pu, and 2.13 kg La. Official Use Only 27

RPP-RPT-50941, Rev. 0

The 234-5 Building production was started up in 1949. Initially 231-Z supplied a plutonium paste to the RG Line. When the RMA Line replaced the RG Line, the feed was in the form of plutonium III oxalate and was later changed to a Pu(IV) oxalate. [HW-10475; HW-29200, Plutonium Purification and Fabrication Technical Manual] The La(NO)3-Pu(NO)4 solution (F-10-P) sent to the Isolation Building (231-Z) from the T and B Plants contained about four times as much La as Pu and also included smaller amounts of other impurities (K, Ca, Bi, Fe, Cr, etc.). The plutonium nitrate product solution was transferred via a PR can to the Isolation Building (231-Z) at a nominal concentration of 20-40 g Pu/L nitric acid solution. The REDOX process started-up in January 1952. The plutonium nitrate product (III-BP) was sent to the 231-Z Facility for further purification. The REDOX product contained about 10-20 g/L Pu. Initially at least, the REDOX plutonium was 80-90% Pu(VI) and the valence was adjusted to Pu(IV) before making plutonium (IV) oxalate precipitate. The REDOX process recovered both plutonium and uranium from a solvent extraction process. The uranium and plutonium nitrate streams were further purified and the final plutonium product stream was transferred to the 231-Z facility for further purification. Once the REDOX plutonium product met the feed specifications for the 234-5 Building, the plutonium product from REDOX was transferred directly to the 234-5Z Building. Some plutonium scrap solution from the 234-5 Building operations was transferred for recovery and recycle via a 231-Z purification process [HW-19991, Final Report Production Test 234-2 Recycle of Skull Solution to the Isolation Process, HW-29200]. The skull of residual plutonium metal removed from the pouring crucible was dissolved in nitric acid in laboratory equipment to produce a relatively pure plutonium nitrate solution. The product solution was sent to the 231-Z operation where the solution was adjusted to 231-Z first cycle flow sheet concentrations and processed normally through 231-Z. Available records are incomplete, but by 1952 a plutonium oxalate process (Task I) capability was operational at the 234-5 Building. The 231-Z Facility was supplying plutonium oxalate cake for Task II from the T Plant bismuth-phosphate recovery. The REDOX plutonium nitrate product would have met the 234-5 plutonium nitrate feed specification by 1953 (assumption) and have been transferred directly to the Remote Mechanical A (RMA) Task I (batch oxalate precipitation) operation.

Official Use Only 28

RPP-RPT-50941, Rev. 0

Figure 10. Historical Flowsheet for 231-Z

4.4

PFP PRODUCTION LINES

A wide variety of plutonium processing and handling capabilities were established in the PFP facility during its more than 50 year life to handle a significant array of plutonium material types and to handle plutonium having a wide spectrum of isotope distributions. A significant overview of PFP history and its process systems has been documented by Gerber (HNF-EP-0924), and a detailed overview of its operating history, production quantities, and scrap generation and recovery activities has been documented by Hoyt and Teal [HNF-22064, Plutonium Finishing Plant Operations Overview (19492004)]. Initial operations at the PFP were focused at the conversion of pure weapon grade plutonium nitrate solutions into plutonium metal and subsequent fabrication into weapon components. Operations that were directly related to the production lines were designated with Task numbers, see Table 4. Historical information in operations reports usually refers to work activities by these Task numbers, as is done in some of the tables presented in this report.

Official Use Only 29

RPP-RPT-50941, Rev. 0

Table 4. Historical Task Numbers and Task Descriptions


Task Number I II III IV V VI VII VIII Task Descriptions Oxalate Precipitation (includes infiltration and calcinations) (also referred to as Purification or Wet Chemistry) Hydrofluorination (Dry Chemistry) Reduction (to metal) Ingoting and Casting Machining (Shaping) Cleaning Coating Final Inspection

There were three separate generations of plutonium conversion lines installed and operated within PFP: Rubber Glove (RG) Line, Remote Mechanical A (RMA) Line, and Remote Mechanical C (RMC) Line. Another line, the Remote Mechanical B (RMB) Line, was also installed after the RMA Line. However, this line became obsolete and was removed before ever initiating hot operations. The initial plutonium metal production and weapon component fabrication line, the RG Line, ran from 1949 until ~1953. The RG Line was a hands-on batch operated system; and after a few years of difficult operation it was replaced with the remote mechanical RMA Line. The RMA Line became operational in 1952 and proved so successful that the RG Line was phased out. As plutonium production requirements accelerated, a second metal production and weapons component fabrication line was installed to supplement the RMA Line capabilities. This new process, the RMC Line, became operational in 1960. The RMA Line was eventually converted to an oxide line in 1968 when the hydrofluorinator was converted to a second stage calciner. Subsequently this line became to be known as the RMA Oxide Line and it remained operational through 1979. The RMC Metal Line operated through 1973 and then remained shutdown until its final campaigns during 19851989 time period. Plutonium production activities significantly diminished in the 1970s and 1980s, after which attention focused on the stabilization and safe storage of remaining plutonium inventories. Thermal stabilization of plutonium solids began in the early 1990s. A Thermal Stabilization and Packaging system was subsequently installed in 234-5Z in the mid 1990s and another was installed in the 2736-ZB Facility in ~2000. Solutions were processed through either batch Mg(OH)2 precipitation and calcination or batch plutonium(IV) oxalate precipitation and calcination in 234-5Z during 2003. All materials in inventory requiring stabilization and packaging to the new storage and stabilization standard were so treated by April 2004, at which time efforts on legacy holdup removal, facility cleanout and demolition became the central mission focus.

Official Use Only 30

RPP-RPT-50941, Rev. 0

A summary of the more significant PFP missions and activities is provided in Table 5, along with their timeframes of operation. These activities, while not all-encompassing, provide a realistic depiction of the work performed during PFPs more than fifty-year lifetime. Additional non-production operations such as analytical work, development work, and plutonium storage supported all of these activities. Many of the non-defense missions involved small amounts of plutonium, various plutonium compounds, and plutonium having a wide variety of Pu-240 isotopic content. Table 5. Overview of PFP Plutonium Processing Activities, 1949 to 2004
Overview of PFP Plutonium Processing (Time Period 1949-2004) Processing Activity Processing Involving Hanford Generated Pu Weapons Grade Material Production RG Line Weapons Grade Material Production RMA Line Weapons Grade Material Production RMC Line High Pu-240 Oxide and/or Metal RMA Line High Pu-240 Oxide and/or Metal RMC Line High Pu-240 Component Fabrication Research Special Oxide Production for EURATOM (Hanford activities) PRTR 20% Pu-240 Nitrate Conversion to Nitrate, Oxide, Metal Nitrate Blending for SEFOR Fuel (13&20% w WG to 8%) Shippingport 24% Pu-240 Nitrate Conversion to Nitrate, Oxide, Metal BNW Criticality Safety Lab FFTF Cores 1&2 Oxide in RMA Line High Pu-240 Oxide for SRS Cf Deep Burn Experiment & Other Blend SRS & Hanford Oxide in RMC Line, FFTF Cores 3&4 SRS 3% Metal Burn To Oxide, Dissolve, Blend w FG, to WG Metal Subtotal First Generation Scrap Recovery & Stabilization Processing Pu Recycled From RECUPLEX SX Pu Recycled From PRF SX Pre-DNFSB 94-1 Stabilization Activities DNFSB 94-1 Stabilization & Packaging Activities (234-5Z) DNFSB 94-1 Stabilization & Packaging Activities (2376ZB) Solution Stabilization (Mg hydroxide & oxalate batch precipitations) <5 < 10 <1 <2 <2 < 0.5 55 62 64 79; 84 88 90 94 95 04 01 04 99 02 <2 > 25 < 25 < 1.5 >3 < 0.5 > 0.5 < 0.5 <1 < 0.5 < 0.5 <2 < 0.5 <1 < 0.5 > 60 49 52 52 64 60 73; 85 89 58; 64; 68 76 63 73 62 64 59 71 63 66; 68 70; 72 65 70 66 67 68; 70 68 76 71 77 78 81 84 Estimated Total (Mt) Approximate Time Frame (Years)

Official Use Only 31

RPP-RPT-50941, Rev. 0

Table 5. Overview of PFP Plutonium Processing Activities, 1949 to 2004


Overview of PFP Plutonium Processing (Time Period 1949-2004) Processing Activity Subtotal Processing Involving Off-Site Receipts UK Receipts & Metal Conversion To Nitrate, Oxide & Metal NFS West Valley Nitrate Receipt & Conversion ZPPR Ingoting In RMC Line ZPPR Scrap Returns Processing (assume to oxide) CSMO Scrap Receipts/Processing RFP Ash Stabilization RFP Oxide Processing RFP Oxide Processing For Chloride Removal Misc. WG Scrap Processing Blend SRS & Hanford Oxide in RMC Line, FFTF Cores 3&4 SRS 3% Metal Burn To Oxide, Dissolve, Blend w FG, to WG Metal Subtotal Total Pu Processed Summary For Years 1973 2004 Total Pu Processed (including recycle) < 15 <5 <2 <5 < 0.5 < 0.5 < 0.5 < 0.5 < 0.5 <5 <1 <1 < 20 < 100 66 70; 76 66 71 67 69 69 72 72 77 85 86 85 86 03 60 73 77 78 81 84 Estimated Total (Mt) < 20 Approximate Time Frame (Years)

4.4.1

RG Line

The Rubber Glove (RG) Line was the first process line installed in the PFP. This line began running hot feed on July 5, 1949. Operations were manual and required operators to work at both open face hoods and gloveboxes. The process consisted of: (1) batch plutonium (IV) oxalate precipitation and filtration, (2) batch combination of calcination/hydrofluorination, (3) batch plutonium tetrafluoride reduction to metal, (4) casting and shaping operations, (5) weapon component machining and cleaning operations, (6) component coating, and (7) inspection operations. Material transfer between process areas was also accomplished manually. The RG Line was beset with numerous operating difficulties involving inefficient conversions, frequent accidents involving spread of plutonium contamination into unintended locations, and contamination of personnel. A simplified flowsheet is shown in Figure 11, and a diagram of the primary process operations for the RG Line is shown in Figure 12 for plutonium nitrate receipt from the fuel processing facilities through metal production, component fabrication, and component assembly for storage and shipment.

Official Use Only 32

RPP-RPT-50941, Rev. 0

Figure 11. RG Line Flowsheet

Figure 12. Early PFP Operation

Official Use Only 33

RPP-RPT-50941, Rev. 0

The RG Line operated from 1949 through 1953, was shut down in 1953, and then placed in stand-by from 1953 to 1957 (see Figure 3 for liquid waste discharge routing from RG line operations). Final removal of the RG Line occurred in early 1957. The RG Line was initially built to operate on a Pu III oxalate precipitation process, Task I. A small amount of hydrogen peroxide was added and the solution agitated for valence adjustment to the +4 state. Then the precipitator was heated to ~35C and 1M oxalic acid was added over a period of about one hour. Then the slurry was transferred to a platinum-lined filter boat and the filtrate drawn off. The oxalate cake was washed with a dilute oxalic acid-nitric acid mix. The filter boat and plutonium oxalate cake was then transferred to the hydrofluorination furnace, which was pre-heated to something less than 100C to prevent spattering of the cake. Once the oxalate cake had been dried the furnace was slowly heated to 300C while air was drawn through the Pu oxalate cake. After the Pu oxalate had been completely converted to Pu oxide, the furnace temperature was increased while a mixture of hydrogen fluoride and oxygen was passed through to convert the plutonium oxide to plutonium tetrafluoride. The reaction was complete after the furnace reached a temperature of 600C. The time to complete this calcination & hydrofluorination process was 5 8 hours. Upon cooling, the plutonium tetrafluoride was transferred to a reduction-mixing area where the plutonium tetrafluoride was mixed with metallic calcium which serves as a reducer and iodine which served as a booster. Some granules of gallium oxide were added to some batches to assure a desired plutonium metal alloy. The mixture was then placed into a magnesium oxide crucible, which was then placed into a reduction furnace, and the furnace was heated to ~750C. The reaction between calcium and iodine reacted initially to form calcium iodine at a lower temperature than the calcium-plutonium tetrafluoride reacted, generating heat to quickly boost the temperature up to initiate the calcium-tetrafluoride reaction. The second reaction was also exothermic, raising the temperature further. The plutonium was reduced to metal with the formation of calcium fluoride; and the temperature was high enough that the plutonium metal, calcium iodide, and calcium fluoride were all molten. Molten plutonium pooled at the bottom of the reduction vessel with the salt flowing to the top. Upon cooling, the crucible was broken and the plutonium metal button was separated from the crucible and salt. The crucible fragments and salt were saved for processing and residual plutonium recovery and the button was cleaned with a pickling solution, weighed, sampled, and placed into interim storage. The plutonium metal buttons were then transferred to the fabrication area of the RG Line. This is where the plutonium metal was cast into rough weapons component. After machining, the components were cleaned, coated to prevent oxidation and reduce contamination, inspected, and fitted for assembly. The parts were then packaged and placed into storage while awaiting shipment off-site. Residues from the casting and machining, and any reject parts, were recycled to appropriate steps in the RG Line or wet recovery areas.

Official Use Only 34

RPP-RPT-50941, Rev. 0

4.4.2

Remote Mechanical (RM) A Line

With the need for increased production and to address Lessons Learned on the RG Line, a new process line was designed, constructed, and initiated hot operations in March 1952. The RMA Line utilized the same chemical and metallurgical processes used in the RG Line. However, the RMA Line was automated to minimize the direct operator interface in order to minimize occupational exposure. See Figures 3, 4, 5, 6 and 7 for liquid waste discharge routing from RMA line operations) The RMA Line began to process the plutonium (IV) oxalate that was prepared in 231-Z. By March 1953, the RMA Line began handling all of the plutonium work. In 1955 a batch calcination step, separate from the hydrofluorination process, was added. Then in 1958 significant process enhancements were added to increase production capability. These included replacing the batch precipitation, calcination, and hydrofluorination steps with continuous plutonium (IV) oxalate precipitation, continuous plutonium oxalate filtration, continuous plutonium oxalate calcination to oxide, and continuous oxide hydrofluorination to plutonium tetrafluoride. The RMA Line was taken completely out of service in 1964 and then in December of 1965 the AEC terminated fabrication of weapon components at Hanford, with all fabrication activities being transferred to the Rocky Flats Plant in Colorado. The RMA Line was down during 1965-1967. In Sept. 1967, a glovebox in the RMA Line was reactivated in order to grind plutonium-aluminum scraps into turnings suitable for plutonium recovery. In 1968, Tasks I-Ill of the RMA Line were cleaned out and reactivated to participate in expanded programs for commercial (i.e., nondefense) nuclear development involving high-exposure plutonium oxide and experiments such as oxides for the Hanford Fast Flux Test Facility. To support this work, the continuous hydrofluorinator was converted to a second stage continuous calciner. Subsequently the RMA Line never produced metal again and became known as the Oxide Line. Gloveboxes associated with weapon component fabrication were placed in stand-by, and subsequently most were removed during 19741976. As a result of the explosion in the americium recovery process in the 242-Z facility, the RMA Line was shut down and put in stand-by in 1976, and had its last campaign in 1979 to convert some miscellaneous high Pu-240 solutions to oxide. Cleanup of the line, with metal plates placed over many of the glove ports, was completed by years end. In 1984, it was decided to keep the RMA Line in standby status. It was never operated again. A description of the RMA Line process after it was converted to the RMA Oxide Line is as follows. The plutonium nitrate solution was loaded into measurement tanks and, after mixing and sampling, was vacuum-transferred into storage tanks. The solution was then vacuum-transferred on a batch basis from to the PRT in glovebox H-7A, where the acid concentration was adjusted and hydrogen peroxide was added to adjust the valence to plutoniumIV. From the PRT, the batch was gravity transferred to the THT so that another batch could be transferred to the PRT. The solution in the THT was pumped continuously to the precipitation reactor in hood H-9A, where oxalic acid solution was added to precipitate plutonium oxalate on a Official Use Only 35

RPP-RPT-50941, Rev. 0

continuous basis. The slurry formed in the reactor overflowed into the vacuum drum filter pan, where it was picked up by vacuum on the rotating drum filter and discharged into the calciner. The filtrate passed through the drum to the VRT by vacuum and was pumped to H-7A, where the plutonium oxalate and plutonium nitrate were neutralized (killed) with potassium permanganate. The filtrate was pumped from hood H-7A to Reclamation (PRF), where the plutonium was recovered. The oxalate precipitate was carried through the two-stage calciner where it was converted into plutonium dioxide. The first stage calciner (screw type) was operated at 375400C. The second stage calciner (vibrating type, which was previously a hydrofluorinator) was operated at 525-550C. The plutonium dioxide was discharged at the end of the second stage calciner into tare weighed containers. The storage containers were weighed, labeled, sampled, sealed and stored for shipment or blending with other powder. 4.4.3 Remote Mechanical (RM) B Line

A second remote mechanical line, RMB Line, was designed at the time that the RMA Line was designed. However, it was never commissioned to operate and became obsolete relative to the continual upgrades that occurred on the RMA Line. It was finally removed from the facility in 19571959 never having operated with plutonium. 4.4.4 Remote Mechanical (RM) C Line

Another remote mechanical line, the RMC Line, was constructed and installed in the PFP to increase production capabilities. The RMC Line featured some radiation shielding and was generally operable remotely, although maintenance and cleanup of residues and spilled materials still resulted in exposure problems. The RMC Line consisted of a glovebox line similar to that of the RMA Line with all of its improvements. The RMC Line began routine processing with hot plutonium nitrate feed in October 1960. By November, the new line was operating so well that it was performing all of the Task I-II work and 80 percent of the button firings. The RMA Line operated intermittently after 1960 and the RMA Task III section (reduction furnaces) was abandoned in favor of the RMC furnaces for firing charges. By 1964, the RMC Line was being used for the bulk of the metal processing and the throughput was increased substantially. In 1964 and 1965, the RMC Line was successfully converting in excess of four metric tons of plutonium per year into specification metal. The RMC Line was the only operating plutonium line in PFP during 1965 to 1968. As with the RMA Line, the weapon component fabrication capabilities were put in standby in 1966 when component fabrication activities were transferred to the Rocky Flats Plat. Most of the component fabrication gloveboxes were subsequently removed in 19741976 along with those of the RMA Line. From19661968 the RMC Line processed non-defense plutonium work. The non-defense work include blending and casting ingots for the ZPPR facility and producing a special lot of oxide material high in Pu-240. In September 1968, in support of the expanded work load, Task I-II portions of the RMA Line were cleaned out and reactivated. From that time until the A-Lines final closure in 1979, the RMA Line was known as the oxide line and the RMC Line was Official Use Only 36

RPP-RPT-50941, Rev. 0

known as the metal line, although the RMC Line would produce both oxide and metal. The line continued to operate on a mixed schedule for the next several years producing weaponsgrade plutonium and fuel-grade plutonium and oxides to support both Defense and Non-Defense missions through 1973, at which time it was put in standby. This line was reactivated to produce weapon grade metal in the mid 1980s and its last campaign ended in June 1989. An RMC Line process pictorial is shown in Figure 13 and a flow diagram for converting the plutonium nitrate to plutonium tetrafluoride is shown in Figure 14. The RMC Line process flow diagram for reducing the plutonium tetrafluoride to metal is shown in Figure 15. The plutonium conversion line processes generated various plutonium-containing scrap materials. These include filtrate, scrap powders, reduction residues (slag and crucible), pickling solutions, off specification metal and metal samples. The filtrate from the plutonium oxalate precipitation process passed through the drum filter cloth and into the vacuum receiver tank (VRT) by vacuum, and was then pumped to hood HC-7, where the excess oxalic acid was treated with potassium permanganate. The filtrate was then pumped to the Plutonium Reclamation Facility for concentration and plutonium recovery. Plutonium oxide powder scrap accumulated in the calciner off gas filter and on the floor of the calincer glovebox. The plutonium oxide was converted to PuF4 in the vibrating hydrofluorinator. During most years, the hydrofluorinator off-gas was scrubbed with a water-jet and transferred the solution to waste tanks, neutralized and pumped to a settling tank which drained to soil-waste disposal crib. During the last campaigns during the 1980s, the hydrofluorinator off-gas was scrubbed with either aluminum nitrate nonahydrate (ANN) or KOH. Spent scrubber solution was tanked, neutralized, and then transferred to Tank Farms for disposal. The plutonium tetrafluoride reduction to plutonium metal generated the reduction sand, slag, and crucible fragments that dropped into a crusher and were ground for size reduction for future recovery. Plutonium oxide, fluoride and metal fines could be present in this material. The plutonium metal button was pickled (cleaned) and the pickling solution was sent to the Plutonium Reclamation Facility for plutonium recovery. Off specification metal, metal samples, and reject buttons were burned to an oxide and recycled with scrap powders. During the early years when weapon components were being fabricated, the metal buttons were transferred to the fabrication area for casting, pressing, shaping, machining, coating, inspection, and assembly operations. This would have generated plutonium scrap as metal, skuls, metal turnings, and fabrication oil.

Official Use Only 37

RPP-RPT-50941, Rev. 0

Figure 13. RMC Line Process Pictorial

Figure 14. RMC Line Flow Diagram Nitrate to Tetrafluoride (PFD-Z-190-00002 B-0)

Official Use Only 38

RPP-RPT-50941, Rev. 0

Figure 15. RMC Line Flow Diagram Metal Reduction (PFD-Z-190-00002 B-0)

4.5

SCRAP RECOVERY (RECUPLEX, PRF, MT, 232-Z)

Scrap recovery operations during the early years were not performed in a dedicated process. Scrap was recycled in a few isolated gloveboxes or by directly incorporating clean scrap back into specific unit operations of the RG or RMA Lines. Miscellaneous batch dissolvers for relatively easy to dissolve scrap, such as oxide, oxalate, skulls, metal, etc., were established and the impure plutonium nitrate solutions were recycled back to other facilities. Out of the need to establish controlled, systematic, integrated, and coupled processes to recover and purify plutonium from the myriad feed and scrap streams within the PFP, first the RECUPLEX and then the Plutonium Reclamation Facility were constructed and operated. 4.5.1 RECUPLEX

The RECUPLEX scrap recovery and purification process was established and initiated hot operations in 1955. (HW-22596, RECUPLEX Feasibility Report) This scrap recovery system, which was a rather complex scrap purification facility, included numerous dissolvers for the dissolution of solids and a solvent extraction process for plutonium nitrate purification of dissolver product, filtrate solutions, and for final purification of feed from the separation plants. The RECUPLEX facility included about one hundred process vessels that were enclosed in three large hoods or glove-boxes located in room 221 of the 234-5Z Building. Several of the process vessels in RECUPLEX were of large diameter that was not geometrically safe for criticality prevention. The RB Process hood was utilized for reception and blending of the various feed

Official Use Only 39

RPP-RPT-50941, Rev. 0

streams, treatment of the Task I supernates, and neutralization of waste solutions. The SE Process hood contained the solvent extraction system and the product concentration equipment. The recovery and purification functions of RECUPLEX were achieved by means of a solvent extraction process which was basically similar to the plutonium purification cycles of the REDOX and PUREX plants except that RECUPLEX used TBP in a carbon tetrachloride was deliuent. There were 11 scrap streams generated during plutonium metal conversion and fabrication operations that were identified for plutonium recovery in RECUPLEX, shown in Table 6, and they were integrated into RECUPLEX processing as shown in Figure 16. Table 6. Plutonium Scrap Streams to RECUPLEX for Plutonium Recovery (HW-35030)
Plutonium Content (Percent of Plutonium Feed to Task I) 2. 2. 1. 0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5 <0.5

Type Waste Slag and Crucible Skulls Supernates Scrap Powder Sample Remnants Inactive Turnings Stripping Solution Pickling Solution Boat Cleaning Solution Solid Laboratory Wastes Liquid Laboratory Wastes

Source Tasks III & IV Task IV Task I Task II Task III Task V Task VI Task III Task I Laboratories Laboratories

Physical State Solid Solid Liquid Solid Solid Solid Liquid Liquid Liquid Solid Liquid

Feed pretreatment and dissolution operations for the bulk of the scrap materials were accomplished in RECUPLEX. Some of the smaller streams, however, such as casting skulls and laboratory wastes, were dissolved and pretreated in other facilities within the building. Dissolution of the casting skulls, inactive turnings, and metal sample remnants was accomplished by a boiling nitric acid-hydrofluoric acid mixture in the skull recovery hood. The dissolver solution was cooled, filtered, and diluted. After treatment with hydrogen peroxide to stabilize the plutonium in the more extractable (IV) valence state, the solution was transferred by vacuum to the Reception and Blending (RB) Process hood.

Official Use Only 40

RPP-RPT-50941, Rev. 0

Figure 16. Simplified Flow Diagram of RECUPLEX Recovery and Coupling Operations (HW-35030)

A boat cleaning station was employed to flush the Task I filter boats with a hot nitric acidaluminum nitrate (ANN) mixture to dissolve entrained solid plutonium fluoride using a large excess of ANN to minimize corrosion. The solution was cooled, transferred by vacuum to Task I and from there to the RB Process hood. Laboratory solid and liquid wastes were subjected to a series of metathesis operations to remove plutonium-complexing agents and corrosive materials. They were periodically transferred by RC can to Task I for load-in to the RB Process hood. The button pickling and coating stripping solutions were transferred in suitable containers to Task I for direct load-in to the Slag and Crucible (SC) and RB Process hoods respectively. The SC Process hood was where the solid reduction slag and crucibles, pouring and casting crucibles, unreacted reduction powders, and fluoride-powder sweepings were dissolved in the hood by the controlled-fluoride process. (Button pickling solutions and wet chemistry supernates were also treated simultaneously with the solid materials.) Concentrated nitric acid was sufficient to dissolve the bulk of the solid materials. However, the plutonium metal and oxides require the presence of hydrofluoric acid in addition to the nitric acid for complete dissolution within an operable time cycle. The dissolver solution was refluxed at a temperature just below the boiling point for a period of eight to ten hours to precipitate dissolved silica and coagulate it to an easily filterable form. After cooling to approximately 80C, the solution was filtered to remove the silica and transferred by vacuum to the RB Process hood. The silica was slurried from the filter and transferred by steam jet to the 216-Z-8 settling tank and overflow crib (french drain) located east of the 234-5 Building. Official Use Only 41

RPP-RPT-50941, Rev. 0

The supernates from Task I operations contained oxalate ions. Provisions were included in Task I for destroying the oxalate ion with potassium permanganate and hydrogen peroxide. The supernates may also be treated in the SC hood dissolvers by processing them with the slag-andcrucible material, as previously mentioned. This latter procedure reduced the capacities of the dissolvers and posed a radiation exposure problem for operating personnel. Prior to their purification in the solvent extraction system, the various recovery-coupling feed streams were blended and adjusted with respect to their salting strength, acidity, and plutonium valence state in the RB Process hood. The bulk of the solutions in the dilute feed make-up were filtrate solutions from Task I and the product from the reduction slag and crucible dissolvers. The blended feed solutions were filtered prior to solvent extraction processing. The solvent extraction system utilized 15% TBP in carbon tetrachloride solvent in three columns: a simple extraction column, a compound extraction-scrubbing column, and a stripping column. More efficient pulse columns with perforated plate cartridges were employed rather than the typical packed columns because of limited height restrictions. The pulsing motion of the column liquids was provided by the expansion and contraction of a Teflon bellows located in a six-inch diameter pipe directly connected to each bottom end section of the column. The organic phase was nonflammable and heavier than water. The most outstanding characteristic of the RECUPLEX solvent extraction system was the use of a reflux-scrub stream to obtain high plutonium concentrations in the product solutions (HW-35030, RECUPLEX Operating Manual). The limitations of dilute nitric acid as a plutonium stripping agent restricted the plutonium concentration in the product solution to approximately four times that of the feed in a non-reflux system. By returning 5095 per cent of the aqueous effluent from the stripping column to the extraction-scrub column as a scrubbing agent, the concentration ratio of the product and feed streams was increased to approximately 100, the maximum concentration being limited to 7090 grams per liter due to plutonium saturation of the organic phase in the scrub section. The aqueous wastes from the solvent extraction columns were large in volume, contained high concentrations of salts, and were low in plutonium content. They were accumulated in the RB Process hood and periodically neutralized with caustic solution to a point just short of precipitating solids. The neutralized solutions were then transferred by gravity to a new type of crib (216-Z-9) located east of the 234-5 Building. The silica waste slurry from the SC Process hood was transferred by steam jet from vessels D-10 and D-11 to an underground storage-overflow tank (216-Z-8), where the solids (silica, filter aid, and any undissolved plutonium) settled out and from which the supernates overflowed to a small crib. Liquid wastes from all auxiliary operations in the process facilities were discharged to the D-6 waste disposal system of the 234-5 Building. These wastes included caustic-scrubber solutions, jacket, coil, and condenser effluents, and excess chemical drainage from the addition vessels. The organic extractant in the solvent extraction system normally flowed through a closed cycle without any continuous treatment. However, gradual depletion of TBP necessitated periodic

Official Use Only 42

RPP-RPT-50941, Rev. 0

addition of new solvent. Moreover, TBP hydrolysis products (such as dibutyl phosphate [DBP]) which strongly complexed with plutonium in the organic phase, tended to accumulate within the system, requiring periodic batch treatments for their removal. The solvent treatment included three successive batch aqueous washes: (1) a dilute oxalic acid solution which removed complexed plutonium, (2) a dilute sodium carbonate solution which removed any organic hydrolysis products, and (3) a water wash which removed residual soda ash. The clean solvent was returned to the columns and the unwashed solvent was similarly separated and treated. A simplified RECUPLEX solvent extraction flowsheet is given in Figure 17. The RECUPLEX process was originally a Research and Development semi-works operation. In 1959 this facility was converted from a semi-works R&D plant to a dedicated process facility. This process eventually became cumbersome and difficult to operate as equipment problems developed and visibility into the gloveboxes became extremely poor. As the facility was operated, there occurred a steady buildup of fission products from separations plant feed. Near the end of life of RECUPLEX, the hood floors contained a gelatinous soup of partially-dissolved dry-box gloves, miscellaneous plastics, corroded tools, failed equipment, and considerable plutonium. RECUPLEX operated until 1962 when a criticality accident occurred during some non-routine operations (HW-74723, Final Report of Accidental Nuclear Excursion, RECUPLEX Operation, 234-5 Facility). All operations were terminated, the facility cleaned out, and the equipment and gloveboxes were removed in 1967. Figure 17. Simplified RECUPLEX Solvent Extraction Flowsheet (HW-35030)

Official Use Only 43

RPP-RPT-50941, Rev. 0

4.5.2

Plutonium Reclamation Facility (PRF)

A decision to dismantle and retire the RECUPLEX facility had been made and the design for an improved scrap recovery process system was already in progress at the time of the 1962 RECUPLEX criticality. This new process, the Plutonium Reclamation Facility (PRF) became operational in 1964. Purification of plutonium scrap materials in the Plutonium Reclamation Facility was effected using a continuous liquid-liquid solvent extraction process that was conducted in a single perforated-plate pulse column 4 inches in diameter and 57 feet tall. The extraction medium consisted of tri-butyl phosphate in carbon tetrachloride as the carrier fluid. The PRF systems, process, and equipment were based largely upon the predecessor scrap recovery facility, RECUPLEX. The scrap treatment operations were called the Miscellaneous Treatment (MT) section. Capabilities originally established for scrap treatment included solid scrap dissolution and leaching, electrolytic metal dissolution, metal oxidation, etc. An illustration showing the many different scrap materials that were identified as PRF feeds for plutonium recovery is shown in Figure 18. The PRF operated from 19641976, at which time it was shutdown and placed in standby. The facility was reactivated in early 1978 and ran through February 1979 processing miscellaneous solutions and scrap materials. The facility remained in hot standby for the rest of 1979 and until 1984. The PRF was reactivated again, with the exception of the sand, slag, and crucible dissolvers, to support scrap treatment and filtrate concentration during the final campaigns of the RMC Line, 19851989. The last operation of the solvent extraction system was in 1987, except for some cold training runs that were performed in 1994. Filtrate from the RMC Line was concentrated in the PRF filtrate concentrator and loaded out and stored in 10 liter batch cans after the solvent extraction system was shut down, and this continued until the RMC Line completed its last run in 1989. Details of the primary process flow-sheets used in the PRF are extensively described in the following documents: (PFD-Z-180-00001, PRF Solvent Extraction; PFD-Z-180-00004, Slag and Crucible Dissolver Flowsheet (Standard Feed) 236-Z Building; RFO-F-2, Alternate Tributylphosphate Solvent Extraction Flowsheet for Recovery of Plutonium from Scrap; RFO-F-10, Plutonium-Uranium Coextraction-Partitioning Flowsheet; ARH-CD-704, Tributyl Phosphate Solvent Extraction Flowsheet for Recovery of Plutonium from Scrap). A more detailed description of the Miscellaneous Treatment area and the solvent extraction system is provided in the following sections.

Official Use Only 44

RPP-RPT-50941, Rev. 0

Figure 18. Scrap Plutonium Recovery Sources for PRF Facility

Official Use Only 45

RPP-RPT-50941, Rev. 0

Figure 19. PRF Plutonium Scrap Processing Pictorial

4.5.2.1 Miscellaneous Treatment. The Miscellaneous Treatment (MT) Section of the PRF was comprised of what in the early RECUPLEX days was call pretreatment, which were those process operations necessary to convert plutonium scrap into a suitable liquid nitrate form that could be subsequently purified as feed to the production operations. These operations included size reduction operations, metal burning stations, hot plates for thermal stabilization, and various types of dissolvers and associated filters and centrifuges. These operations are described below in conjunction with descriptions of their associated scrap feed materials.

Official Use Only 46

RPP-RPT-50941, Rev. 0

Miscellaneous Treatment Plutonium-bearing solids that could not be processed in the sand and crucible (S&C) dissolvers, were processed for plutonium recovery in the Miscellaneous Treatment (MT) Section. These scraps included pure or impure plutonium-uranium mixed oxides, fluorides, oxalates, silicates, contaminated carbon, or organic sludges or residues, metal plutonium, casting skulls (burned to oxide), incinerator ash, polystyrene cubes, plutonium oxide and metal that did not meet specifications, plutonium-bearing solids and miscellaneous scrap received from various PFP operations and from other plutonium scrap generators. Operations usually included the stabilization and conversion to oxide on hot plates, dissolution, leaching, product solution filtering and centrifuging. Plutonium metal was dissolved in the electrolytic dissolvers or burned to an oxide and the oxide subsequently dissolved in pot dissolvers. Plutonium oxideimpregnated polystyrene cubes were burned and distilled to produce a plutonium oxide product. Plutonium metal chips and bulk were burned to produce a plutonium oxide product. The resulting residual solids and plutonium-bearing residual solids from other processes were leached with a near boiling nitric-hydrofluoric acid mixture in heated pots. Aluminum nitrate nonahydrate [Al(NO3)3 9H2O] (ANN) was added later to complex fluoride. The plutonium solutions from dissolution and leaching were ultimately processed in the PRF solvent extraction process. An overview of process operations conducted in MT is as follows: Scrap Feed Receipt Plutonium scrap feed for dissolution was inspected for removal of extraneous material, and batch sized. The screened scrap was divided into batches containing up to 800 g of plutonium and transferred for dissolution. Plutonium Burning The plutonium metal for burning was weighed and placed on an empty burning plan. The metal was initially ignited by a propane torch; however in later years this was replaced by an electric igniter for safety reasons. The torch was extinguished after satisfactory ignition. The material was stirred occasionally to prevent the plutonium from being smothered by oxide powder. When the burning was completed, the cooled material was screened and packaged and transferred for dissolution. Acid Dissolution Dissolution of plutonium oxide powders and plutonium-bearing compounds was achieved using boiling nitric acid (6M, 12M, 16M), hydrofluoric acid (0.35M), aluminum nitrate nonahydrate (ANN) (1.8M). Sodium fluoride was also employed as another source of additional fluoride ion. Scrap containing less than 20% organic and up to 800 g of plutonium were placed in batch dissolvers containing approximately 3 L of 12 M HNO3 + 0.35M HF. For difficult to dissolve material, the fluoride concentration was increased by the addition of sodium fluoride to the dissolver solution. The plutonium-bearing scrap and dissolver solution were heated to boiling for one hour and then briefly cooled to allow settling of undissolved solids. The dissolver solution generally contained about 1% solids. Alternate dissolver recipes were used for different types of scrap. For plutonium-fluoride containing solids, 3 L of 1.5 M Al(NO3)3 2 .0 M HNO3 were added to the dissolver. An alternate oxide dissolver solution was 3 L of 12 M HNO3 and 50 g of NaF. An alternate fluoride Official Use Only 47

RPP-RPT-50941, Rev. 0

dissolver solution was 2.5 L of 12 M HNO3 and 0.5 L of 50% Al(NO3)3. Three liters of 12 M HNO3 was recommended for incinerator ash. Calcium fluoride was preferred when scrap contained silicates (e.g., incinerator ash). Excess foaming was experienced periodically from dissolution of certain scrap materials. Fifteen milliliters of 500 ppm antifoam in 1 M HNO3 was added as needed. The antifoaming agent was a proprietary commercially available material. Recovered undissolved solids, i.e., dissolver heels, with plutonium content greater than 10 g were recycled to the batch dissolvers. Solids with plutonium content less than economically recoverable were stabilized for transuranic (TRU) solid waste disposal. Material to be stored for future recovery was stabilized. Material to be disposed of as TRU waste was cemented. The scrap was cemented in polyjars and discarded as solid TRU waste. Electrolytic Dissolution The process was designed to dissolve plutonium metal in nitric acid by application of an electric current. Plutonium metal was placed between two electrodes of the dissolver and concentrated nitric acid was pumped through the system. A potential of 15-24 volts was applied to the electrodes causing the plutonium metal to dissolve on the anodic side yielding plutonium nitrate. Heat generated by the electrodes was removed by circulating the acid through a primary heat exchanger. Gases produced by the electrolysis are removed and diluted by a vacuum system. Polycube Processing Polycubes were used in criticality studies in the 1960s. The plutonium recovery process involved pyrolysis and distillation of the polystyrene and burning of the residual carbon. The cube coatings were split to allow the escape of distilling gases and the coatings were burned rather than removed. The still temperature was maintained at 450C and alarmed at 490C. The off-gas from the still was drawn through a scrubber containing carbon tetrachloride. The cubes were broken in half and placed in the boats. The boats were charged to the still on a 45 minute cycle. The boats were then transferred to the furnace potion of glove box 4. Each distilled cube was crushed and evenly distributed into two boats. The boats were charged to the burning furnace which was maintained at 800C and alarmed at 950C. Oxide recovered from polycubes would be considered high-fired and difficult to dissolve. 4.5.2.2 Continuous Slag and Crucible Dissolvers. One of the primary scrap streams from a nitrate to metal conversion process was the slag and crucible stream from the reduction of plutonium tetrafluoride to metal, as previously mentioned. These reduction residues were processed in continuous slag and crucible dissolvers. These dissolvers were located in the PRF solvent extraction process cell and were always operated in conjunction with the solvent extraction system. Plutonium metal reguli were created at the Plutonium Finishing Plant using the bomb reduction process. Ingredients in the furnace charge included Plutonium tetra-fluoride powder and calcium metal, with a small quantity of elemental iodine as an initiator or accelerant. After firing

Official Use Only 48

RPP-RPT-50941, Rev. 0

in an induction furnace, the pressure vessel contained a two-kilogram metal button and a calcium-fluoride slag lump. An operator retrieved the button and sent it to an acid pickling bath to clean the surface of the button. The materials remaining behind: MgO sand for nesting the MgO crucible, the crucible fragments, and the calcium-fluoride slag lump, were fed into a hammer mill for size reduction. The fragments were then placed into two metal cans. A typical reduction charge resulted in about 20 grams of plutonium remaining in the scrap fragments. The following descriptions of the slag and crucible dissolver operation were taken from the hazard evaluation, SD-CP-SAR-013 Rev 1. In the dissolver, the can dropped to a stop below the top of the acid column where three separate roughly sequential dissolution reactions take place. The first two reactions take place relatively fast compared to the third. Initially the can reacted with heated nitric acid which breached the can and then the finely divided metals within reacted vigorously with the acid solution. The MgO sand and larger chips of the broken MgO crucible would tend to settle and dissolve slowly while undergoing agitation from the steam and nitrogen sparges. The initial heat for reaction was provided by the heated acid. After the first can of a process run was processed, the acid heaters were turned off since the heat-of-reaction and the steam sparges were sufficient to maintain the proper operating temperature. The product overflowed continuously to the solids settling tank through a centrifuge and then to a makeup tank as feed for the solvent extraction system. Product overflow from the dissolvers was centrifuged and then transferred to the PRF solvent extraction system. Undissolved solids were flushed out and filtered for recovery or disposal as waste, depending upon the plutonium content. Plutonium was also recovered from many different types of scrap in these dissolvers, including incinerator ash, plutonium-aluminum alloy turnings, plutonium-Be sources, etc. Sodium fluoride and mercuric nitrate were added to the acid dissolution media for some of these materials. Dissolution of ash in these dissolvers proved difficult with undissolved heels and carbon causing poor visibility, interfering with the aqueousorganic interface during solvent extraction operations, and solids buildup. 4.5.2.3 Filtrate Treatment and Concentration. The filtrates from the RMA and RMC vacuum drum filter operation were treated with potassium permanganate to destroy the residual oxalate in the filtrate. Other recoverable solutions from the process, including condensate from the calciner off-gas, were also gathered in the glovebox and transferred to PRF along with the filtrate solutions. The filtrate solution along with other dilute recoverable solutions resulting from spills that were collected and condensate from the product concentrator were routinely concentrated in the filtrate evaporator. The feed solution to the evaporator was chemically adjusted to the required nitrate concentration and an antifoam agent was added to avoid frothing. The feed was pumped into the evaporator which was steam heated. The condensate was normally routed to the D-4 waste tank in the 241-Z Waste Facility for disposal. The concentrate from the evaporator was routed back to the CA column for plutonium recovery. A more detailed process description may be found in Process Flow Sheet PFD-Z-180-00001.

Official Use Only 49

RPP-RPT-50941, Rev. 0

The filtrate from the button line precipitation process was concentrated in the filtrate evaporator to reduce the volume of solution before going to feed makeup for solvent extraction. Rework from waste treatment, floor solutions, and other solutions needing concentration was also run through the evaporator. Vapor from the evaporator entered a shell and tube heat exchanger while the filtrate flowed into two concentrate receiver tanks. When a full tank of concentrate was accumulated, it was sampled for accountability and then pumped to feed makeup for solvent extraction. The condensate was routed to the Tank D-4 waste system. 4.5.2.4 Solvent Extraction. Purification of plutonium scrap materials in the Plutonium Reclamation Facility was effected using a continuous liquid-liquid solvent extraction process that was conducted in a single perforated-plate pulse column four inches in diameter and fiftyseven feet tall. The extraction medium consisted of tri-butyl phosphate in carbon tetrachloride as the carrier fluid. The PRF systems, process, and equipment were based largely upon the RECUPLEX scrap recovery facility. The Plutonium Reclamation Facility housed a continuous chemical processing system. Purification of the plutonium was accomplished by counter-currently contacting the aqueous phase with a heavy tributyl phosphate (TBP) in carbon tetrachloride organic phase. It was typically operated on a round-the-clock basis on either a five or seven days-per-week schedule because of: (1) the large quantities of scrap materials that were generated at the PFP, and (2) the sensitivity of process performance to relatively small changes in feed types, flow-rates, and chemistry. Relatively small aberrations in process parameters resulted in long term perturbations of process stability. The operating regimen consisted of first achieving stabilized operation of the solvent refining system with acceptably-low waste losses, and then gradual introduction of plutonium- laden feed-stocks including solutions from a variety of dissolution vessels, concentrated filtrate from a button line, and various rework liquors. Purified plutonium nitrate solution was concentrated and periodically returned to a 234-5Z Building button line as a feedstock blend. Solvent Extraction Process Operation Impure dilute aqueous plutonium solutions from the RMA and RMC lines and from solid scrap recovery operations in the MT section of the 236-Z facility were piped to the feed tanks. Adjustments to the feed chemical composition and acidity were made to satisfy the specific feed characteristics for the extraction process. Acceptable feed was transferred from the feed tank for delivery into the solvent extraction system (SX) (Figure 20). The feed solution was transferred by pumping at a controlled rate to the CA column where the feed was contacted with an acidic approximately 20% tributyl phosphate in carbon tetrachloride extractant solution. The column was pulsed to assist in mixing the carbon tetrachloride solution with the acidic aqueous feed solution. The plutonium from the feed solution was extracted into the carbon tetrachloride solution and the aqueous stream, depleted of plutonium, was routed to waste tanks. In these tanks the residual plutonium concentration was determined and, if found acceptable for disposal, the spent solution was transferred to liquid waste disposal.

Official Use Only 50

RPP-RPT-50941, Rev. 0

The plutonium bearing carbon tetrachloride stream was next pumped to the CC column where it was contacted with dilute aqueous nitric acid/hydrolyamine nitrate (HN) mixture, which stripped the plutonium from the carbon tetrachloride solution into the aqueous nitric acid solution. The depleted carbon tetrachloride solution was pumped through the CO, CU, CX, and OA columns where the solvent solution was washed to strip residual plutonium, remove the tributyl phosphate decomposition products, recovered and chemically adjusted for reuse. Aqueous waste streams from these columns were either recycled or sent to liquid waste disposal. The aqueous nitric acid product stream, now rich in plutonium, was pumped from the CC column to the feed tank for the product evaporator. In the product evaporator the aqueous plutonium was steam stripped of residual carbon tetrachloride and was concentrated by evaporation to the desired plutonium concentration, usually about 200- 300 grams plutonium per liter. Normally several passes through the concentrator were required to reach the desired concentration. The concentrated plutonium product was stored in product receiver tanks. The acidic distillate was condensed and routed back into the SX process to recover any residual entrained plutonium. The concentrated plutonium product after sampling and testing was pumped to the RMA Line or RMC Line feed tanks. In the normal Pu (IV) flowsheet, uranium transferred to the organic in the CA co1umn and then tended to build up in the CC column. This type of operation was continued until uranium concentration in the product exceeded a specified level. The composition of the CCX was then changed to what was called the uranium depletion makeup. When this makeup was used the uranium stayed with the organic until reaching the CU column. In the CU column the uranium was stripped to the CUU stream, which was then discarded with other aqueous waste streams to tank farms. Generally, the concentration of plutonium in the CCR stream must be kept lower during uranium depletion than during normal uranium buildup operation. Plutonium solids entrained in the solutions handled in the PRF solvent extraction system caused operations problems and facility outages, as in other systems within the plant. This is discussed in greater detail in later portions of this section. A more detailed description of the extraction process can be found in the Process Flow Sheet (PFD-Z-180-00001), RHO-MA-246, Plutonium Reclamation Facility Engineering Training Manual, SD-CP-TI-122, Plutonium Reclamation Facility Engineering Technical Reference, and the Training Manual on Solvent Extraction.

Official Use Only 51

RPP-RPT-50941, Rev. 0

Figure 20. Plutonium Reclamation Flowsheet (PFD-Z-180-00001 Rev. C-1)

4.5.3

Ion-Exchange Recovery of Plutonium

Project CGC-978 provided continuous anion exchange facilities at Z Plant for purifying and concentrating plutonium from the RMC Button Line aqueous wastes, including Task I filtrate, Task III button pickling solution, and glove box sump solutions. No formal scope design was issued because of project urgency. The urgency was due to the loss of RECUPLEX from the accidental criticality. Until ion exchange was operational, button line filtrates were recycled back to the separations plants. The Project encompasses the following: (1) feed preparation and rate control equipment, including new transfer pumps and instrumentation in the HC-7 and H-7A glove boxes of the RMC and RMA Button Lines, respectively; (2) chemical preparation equipment, and (3) the ion exchange equipment, including three batch-loaded Pyrex glass columns, two parallel-operated waste receivers, a chemical pump tank for washing and eluting the columns, a product receiver, and associated piping, pumps, valves and instrumentation. The ion-exchange system was housed in the new HA-46 glove box and tank cell located in the old Official Use Only 52

RPP-RPT-50941, Rev. 0

Task I area of the RMA Button Line. The HA-46 vessels were of favorable geometry, except for the overflow tank. The overflow tank was filled with borated glass Raschig rings. The glovebox with an external sump overflow catch tank contained all equipment except the waste receivers and the chemical pump tank, which were located in the cell physically isolated from the operating room by the glove box and a carbon steel barrier wall. The waste stream from the anion exchange columns was drained by gravity to the underground waste sump pit tank D-6 for crib disposal, and was expected to contain less than 0.001 gram plutonium per liter. The product stream from the columns was transferred by vacuum to the HC-6 glove box storage and blending tanks for Task I processing, and averaged ~30 grams plutonium per liter. The RMC Button Line Task I filtrate was received continuously in the existing filtrate storage tank in the HC-7 glove box. It was then transferred by pump to the existing filtrate kill tank filtrate catch tank U-tube, which overflowed to the filtrate storage tank in the H-7A glove box of the RMA Button Line. During the HC-7 to H-7A transfer the filtrate was adjusted with a nitric acid-aluminum nitrate addition. The chemically-adjusted feed was passed through a steamjacketed heater to attain the optimum temperature for ion exchange processing. The heated feed then passed in series downward through two of the three ion exchange columns into one of the two parallel-operated waste tanks, where it was agitated by pump recirculation, sampled for plutonium assay, and drained by gravity to the waste sump tank D-6. When the first of the series-operated columns was loaded with plutonium, as denoted by resin color change, valving was changed to send the feed downward through the remaining two columns in series. The loaded column was then washed with heated high molarity nitric acid wash solution, which was routed to waste. Plutonium was removed from the resin by eluting with heated low molarity nitric acid. The product solution was routed to the product tank for plutonium assay and vacuum transfer to the HC-6 glove box storage and blending tanks while awaiting Task I processing. Miscellaneous aqueous wastes, including the HA-46 glove box sump and overflow catch tank materials were collected batch-wise in the pre-reduction tank (PRT) of the H-7A glovebox, chemically adjusted, and fed slowly into the suction side of the Task I filtrate pump feeding the ion exchange column (HW-75134). The ion exchange facility operated during the time period 19621964. Early operation of the ion-exchange facility produced prohibitively high waste losses, which were traced to significant quantities of Pu(VI) in the filtrate. Since Pu(VI) does not form an anion complex, as does Pu(IV), this species did not load on an anion resin. As a result, ferrous sulfamate, which reduces the Pu(VI) to Pu(III) and Pu(IV), and sodium nitrite, which oxidizes any Pu(III) to Pu(IV), were added to the process. Another second change from early operations was a procedural change, calling for recycle of the dilute product obtained at the beginning and end of the elution step, which resulted in product solution containing 5060 g Pu/liter being produced, rather than the 30 g Pu/liter shown originally on early flowsheets.

Official Use Only 53

RPP-RPT-50941, Rev. 0

4.6

PLUTONIUM FINISHING PLANT LABS

There were three laboratory facilities within the 234-5Z Building: The Analytical Laboratory (some documents talk of the Engineering Laboratory), The Standards Laboratory, and the Development Laboratory (Plutonium Finishing Support LaboratoryPPSL) [Section 5.4.14, 1995 SAR]. The aqueous plutonium containing wastes streams from the laboratories were classified as low salt waste. Jones (HNF-1989, Tank 241-Z-361 Process and Characterization History) indicated that the approximately 100 g of plutonium/year were discarded in the approximately 1,740 Kgallons of aqueous waste (Table 7). During periods of laboratory waste only, actual accountability values are used rather than the nominal 100-gram per year estimate. Table 7. PFP Laboratory Aqueous Waste
Stream Contaminated lab wastes Source Lab sink drains Thousands of Gallons/yr 1,740 Pu grams/yr 100 Chemical Content Miscellaneous lab chemicals

Analytical Laboratory The Analytical Laboratory was located in rooms 131 through 157-A of the 234-5Z Building. The Laboratory performed analytical and physical test analyses on plutonium samples supporting process control, product quality verification, and nuclear material accountability. The samples came from the plant operations within PFP, the research laboratory within PFP, the PUREX plant, and special test samples from other locations. The types of measurements performed on samples received included fissile material content, impurities content, isotopic content, visual checks for color, phase (aqueous or organic) and solids content for liquids, specific gravity (or density), sinerability, surface area and loss-onignition of oxide powders, and material weights. Hoods in the Analytical Laboratory were categorized as dry hoods (those handling powder or metal) and solution hoods (those handling solutions only). Solutions were normally received in 30 mL (nominal volume) sample bottles and solid material was normally received in 300 g lots. After solution samples were analyzed, they were collected until sufficient volume was available for concentration in the waste recovery facility located in Room 152. Organic solutions were separated from aqueous solutions. The product solution concentrate was sent to the PRF. Wet chemistry hoods utilized slurp bottles or vacuum receiver vessels that simultaneously accumulated sample solutions no longer needed, and concentrated them, via a constant air flow to the 26-inch vacuum system. These solutions were processed through the Laboratory Recycle Solution Concentrator in Room 152. Official Use Only 54

RPP-RPT-50941, Rev. 0

Chemicals incompatible with the laboratory recycle solution concentrator, the plutonium liquidliquid extraction process, or which contain precious metals were segregated for special recovery rather than being recycled with other slurp solutions. Low level laboratory waste was permitted with prior approval to be routed to the PFP waste tanks in the 241-Z Building, if they were incompatible with recycle material. Room 143 was used for handling liquid waste not compatible with the waste recovery system and disposal of as low-level waste (LLW) to the D-4 tank in 241-Z. Hood 5 was used to dispose of LLW via a vacuum-transfer system to the D-4 tank for transfer to waste disposal. The hood also contained a waste aspirator to transfer the waste from the 4-L bottles to the sink drain. Room 152, Waste Laboratory, was used to reprocess liquid laboratory plutonium wastes. Glovebox 552 contained three storage tanks, a distillate tank, vacuum trap, steam concentrator, caustic scrubber, water condenser, and a load-out tank. Overheads from the concentrator were condensed, collected in the distillate tank, and sampled and analyzed for plutonium content prior to transfer to the 241-Z D-4 tank for disposal (or recycled if necessary). The concentrator bottoms were transferred to the load-out tank. When the load-out tank had accumulated 8 L of product, it was transferred to batch tanks in Room 227. As a backup, vacuum transfer could be made to a PR can. Radiochemical Standards Laboratory The Radiochemical Standards Laboratory was located in Rooms 221-, 221-D and 221-E. The Standards Laboratory was responsible for the preparation, verification, packaging, and the distribution of actinide-containing destructive and non-destructive standards needed in the laboratories. Development Laboratory (Plutonium Process Support Laboratory [PPSL]) The PPSL was located in Rooms 179-B through 191, Room 202, and Room 235-C. The PPSL supported numerous programs across the Hanford Site and on occasion, other U.S. Department of Energy Sites. The PPSL supported processing facilities flowsheet verification and development studies, troubleshooting, new technology implementation, and process demonstration. The above activities were performed on both bench-and pilot-plant scales. The PPSL also analyzed non-routine samples for process support and developed new methods which were transferred to other organizations when the methodology was well established and the need became routine. The majority of the work was performed in gloveboxes, with some open-faced hoods. The PPS, for the most part, used non-fixed or semi-fixed equipment which allowed changing the equipment, depending on the particular study being conducted. Material was received from the customer. After the particular study requested by the customer was performed, the material was removed. Waste (liquid or solid) if low enough was discarded or returned with the original material if reprocessing was required the material was transferred to the PRF. Low level wastes and cooling water was disposed of via drain lines that connected to the radioactive D-4 drain line. Official Use Only 55

RPP-RPT-50941, Rev. 0

4.7

PFP SUPPORT FACILITIES

A number of liquid effluent streams were generated at the PFP. This section will review liquid aqueous and organic effluents were routinely discharged as waste to a variety of disposal sites over the lifetime of the PFP. The radioactive liquid waste (RLW) streams were generated by the process operations to convert plutonium nitrate to plutonium fluoride/oxide/metal, activities to recover plutonium and americium from scrap and waste, plutonium purification activities, and by activities conducted in the Analytical Laboratory (AL) and Development Laboratory (DL). The Plutonium Reclamation Facility aqueous plutonium-containing waste stream from the solvent extraction process (CAW) flowed to the 242-Z Waste Treatment Facility for the recovery of the residual plutonium in the CAW waste stream. The aqueous waste stream from the 242-Z plutonium extraction process was collected in waste receiver tanks and discarded to the appropriate crib. The other plutonium containing liquid waste streams generated at the PFP streams were collected in the 241-Z Waste Retention Facility for temporary storage and treatment before disposal. Neutralization of the acid streams was accomplished in the 241-sump tanks. These neutralized waste streams were discharged via the 241-Z Settling tank to a variety of cribs and trenches. After May 1973 all the plutonium containing waste streams were collected in the 241-Z Waste Retention Facility for temporary storage before transfer to tank farm facilities. 4.7.1 242-Z Waste Treatment Facility

The function of the 242-Z Facility was to treat waste streams from the 236-Z Plutonium Reclamation Facility (PRF) and the 241-Z Sump for further plutonium and americium recovery. In 1964 PRF was activated to recover plutonium from scrap. In the early years of plutonium scrap processing operations, the aqueous raffinate (CAW) stream from the tributyl phosphate (TBP) plutonium process was routed to trenches (cribs). The Waste Treatment Facility (242-Z Building) was constructed to recover the low levels of plutonium in the CAW waste stream and the Americium Recovery Facility was added in 1964 to recover americium-241 The initial process utilized a static bed solvent extraction column (W-2) to recover plutonium and batch extraction/separation (Tanks W-12, W-13, and W-14) to recover Am241 using dibutylbutyl phosphonate (DBBP) in carbon tetrachloride (CCl4) from the CAW waste stream from PRF. The PRF stream was routed through a static bed of dibutyl butyl phosphonate (DBBP) in carbon tetrachloride (CCl4). The recovered plutonium stream was rerouted to PRF. The 241Americium was recovered in the W-14 cation exchange concentration column (later W-14, and W-14A). Originally it was intended to purify and recover the americium-241 in the 242-Z Facility. This proved unworkable and the 241Am eluted from the column was loaded out and shipped to the Pacific Northwest Laboratory (PNL). The static bed solvent extraction process for recovering plutonium was used until 1970 (Figure 21). So many problems were encountered with the static bed and batch separation Official Use Only 56

RPP-RPT-50941, Rev. 0

process that the batch process was replaced with a continuous solvent extraction process which was installed in the 236-Z Process cell (Figure 22). A minor and less well defined operation was the recovery of plutonium form 241-Z Building waste streams utilizing the W-1 ion exchange column. The feed stream to the 242-Z Waste Treatment Facility was the aqueous raffinate (CAW) stream from the PRF operations. Before the 242-Z facility became operational this waste stream was discarded to waste via the cribs. The CAW waste stream contained about 2-10 mg Pu/L. However this value was known to vary widely depending on the characteristics of the feed streams to the PRF solvent extraction unit and the operation of the extraction unit. This plutonium content of the waste stream was based on laboratory analysis that only reported soluble plutonium in the sample. It was generally accepted that at least 100 g Pu/month (assay value) was lost via the CAW waste stream. During time of high throughput, CAW waste losses could have been as high as 1 kg/month. The waste stream was known to include unextractable plutonium in the form of plutonium containing organic complexes and undissolved plutonium containing solids. A nominal composition of Hanford CAW is provided in the following table. Table 8. Composition of Hanford CAW Solution
Component NO3 H
+ -1

Concentration (M) 5.0 2.2 0.8 0.5 0.3 (normal values, 0.1) 0.009 0.222 0.001 .0007 .006 0.0003 2-10 mg/L 2-10 mg/L

Al+3 Na FFe
+3 +

Si Ca
+2

Cr+3 Mg
+2

Ni+2 Pu
241

Am

Based on the 1970 Continuous Waste Treatment Flowsheet the composition of the CAW stream had the composition shown in the following table.

Official Use Only 57

RPP-RPT-50941, Rev. 0

Table 9. CAW Composition Continuous Waste Treatment Flowsheet


Component H+ AlF
+2

Concentration (M) 2.0 .03 0.2 1.9

Al+3 Na
+

Fe+3, Ca+2, Mg+2, etc. NO3Pu Am SpG flow 5.4 7.2 mg/L 0.22 mg/L 1.2 200 L/hr

Performance in the early years of facility operations was generally good. Six hundred and seventeen grams of plutonium were recovered from 82,690 L of CAW in June 1965. In October 1965, 270 g of plutonium was recovered from 81,172 L of CAW, and 56 g plutonium was recovered from 56,787 L of CAW in December. From 1966 through 1968, rates remained approximately the same, averaging, on a monthly basis, about 65,000L of feed and yielding about 200 g of plutonium. Rates dropped in early 1969 when processing of plutonium-bearing ash from the 232-Z Incinerator. Trouble was encountered in the W-2 solvent extraction column with high plutonium holdups. In the summer of 1966, monitoring indicated a 1,500 g plutonium hold-up in the extraction column. Indications were that the polyethylene Raschig rings were expanding and becoming absorbed with Pu-DBBP-CCl4. This phenomenon continued through 19681969 with plutonium hold-up levels in Tanks W3 and W-4 hovering between 200800 g of plutonium, with some monitoring indicating the presence of 1,500 to 2,400 g of plutonium (HNF-EP-0924) The static bed ion exchange operation consisted of a cation exchange column, W-1, a solvent exchange column, W-2, two solvent extraction waste receiver tanks W-3 and W-4, a product receiver tank, W-5, a sump tank, W-6, and associated feed tanks and instrumentation. The original system was designed to process the CAW waste stream at the rate of 125 L/hour. It was expected that the column would be stripped of accumulated plutonium every one to two weeks (HW-67010, Design Scope of the Waste Treatment Facility, Z-Plant, Project CGC-912). The CAW waste flowed by gravity upward through a static bed of 35% DDDP-CCl4 in the Raschig (-inch borated polyethylene rings) ring filled W-2 column, where the plutonium was extracted into the organic phase. A flow diagram of the waste treatment and americium recovery batch operation is depicted in Figure 21. When the W-2 column plutonium reached a predetermined value (approximately 230 g) the plutonium was washed with dilute nitric acid and stripped with sodium carbonate solution and transferred to the W-5 product receiver tank. The Official Use Only 58

RPP-RPT-50941, Rev. 0

W-5 plutonium nitrate product was pumped to the Plutonium Reclamation Facility (PRF). The aqueous phase was collected in the W-3 and W-4 waste receiver tanks for discard to the appropriate crib. Degraded solvent, after sampling and accountability was routed to the cribs. While the flowsheet showed 98% of the extractable plutonium removal from the CAW waste, in actual practice the extent of plutonium extraction from the CAW was found to be approximately 40 to 95% of the plutonium sent to the Waste Treatment Facility. The inextractable plutonium (thought to be complexed plutonium-fluo-nitrate hydroxide solids) accounted for the actual loss of plutonium from the process. (letter, Waste Treatment Processing without Americium Ion Exchange). The organic phase from the W-2 column, containing the americium-241and residual plutonium was routed to Tank W-13, the americium-241 strip tank for recovery of Am241. The acidity was adjusted and the aqueous acid stream was passed through the ion exchange column W-14. The ion-exchange resin consisted of Dowex-50W-X8 cation resin. The feed stream to the ion exchange column was nominally 37.3 mg/L americium-241 and 0.4 mg/L plutonium. After the column was loaded with americium-241, it was eluted with concentrated nitric acid and the recovered americium-241 loaded into PR cans for transfer to PNL. The product stream from the column was nominally 3.0 g/L Am and 0.1 g/L plutonium. The raffinate aqueous stream from the column was then discarded to Tank D-4 of the 241-Z facility. Column W-1 was not part of the main plutonium recovery process. It was originally installed to treat any waste in 241-Z Facility that needed to be reworked before it could be routed to the cribs. The W-1 column was intended to reprocess neutralized batches of waste from the 241-Z sump pits. Operation of the column was poor from the start because of poor neutralization control and pluggage of the column from solids in the sump system. The entire operation was abandoned in 1963 and nuclear materials assigned a hold-up of 177 g to the resin which was believed to still be in the column in 1977 (letter, Removal of W-1 Column Resin). During its years of operation, the W-2 solvent extraction column experienced significant plutonium holdup (Gerber, 1997). Gamma scans of the column indicated about 1,500 g of plutonium holdup. (letter, Determination of the Pu Build-Up in Waste Treatment Facility Tanks, W-2, W-3, and W-4). In 1969, Tanks W-2, W-3, and W-4 were neutron monitored and gamma counted. Neutron monitoring results indicated there was a maximum of 2,400, 1,000, and 1,250 g of plutonium in W-2, W-3, and W-4, respectively. The probable hold-up was 200, 400, and 400 grams in W-2, W-3, and W-4, respectively. Gamma counting resulted in a maximum of 575 and 640 g of plutonium in W-3 and W-4. The W-2 Column was neutron monitored routinely to determine the amount of plutonium absorbed on the Raschig rings (Evaluation of Process Test PRF-69-2, Improved Control of Solids Build-Up in the Waste Treatment Facility). Between June 1969 and April 16, 1970, the plutonium accumulation varied from 911 g plutonium on August 28, 1969 to 1,675 g on April 16, 1970. The plutonium accumulation was removed in April 1971. Usage of W-2 as a packed column had ended. (Process Test PRF-71-2, Leaching of W-2 Rings). In 1971, the polyethylene Raschig rings from the W-2 column were removed. The rings contain a total of 2,750 g of plutonium by gamma count. [ARH-LD-213 B, Atlantic Richfield Hanford Company Monthly ReportJanuary 1976).

Official Use Only 59

RPP-RPT-50941, Rev. 0

During routine washing of the organic waste treatment (W-12) tank it was determined that the W-12 Tank contained 650-700 g of plutonium. A continuous plutonium-americium solvent extraction package was installed to replace the static bed ion exchange system. The continuous waste treatment solvent extraction columns were located in the process cell of the 236-Z Building with auxiliary tankage in the cell of the 242-Z Building The waste treatment process utilized three continuous countercurrent, solvent extraction columns (WE-1, WS-1, and WS-2) to separate the CAW into three streams. These were a lowlevel aqueous waste stream (E1W) [flowsheet value 3.5X10-2 gPu/L] that was discarded to the Z-18 crib, a plutonium-rich stream that was recycled back via the filtrate concentrator (236-Z) or tank 15 of the slag and crucible dissolver process to the plutonium recovery process, and an americium nitrate stream that went to the americium recovery ion exchange process. Figure 22 shows the flowsheet for the continuous waste treatment process (Process Test PRF-70-8, Operation of Continous Waste Treatment (HCE-637) for Start-up.) The CAW from the plutonium recovery solvent extraction process was adjusted in pH with sodium hydroxide and became the waste treatment aqueous feed (E1F). The E1F was contacted with a solvent, 30% DBBP, 70% CCl4, in the WE-1 column. The majority of the plutonium and americium contained in the E1F was extracted into the solvent. The aqueous waste (E1W) from the WE-1 extraction column was sampled and discarded, if sufficiently low in plutonium and 241 Am, to Crib Z-18 The extract leaving the WE-1 column (E1P) flowed to the WS-1 column, where americium was selectively stripped. The americium-rich aqueous stream (S1P) from the WS-1 column flowed to the WS-2 column where plutonium was stripped from the organic by dilute nitric-hydrofluoric acid solution. The plutonium-rich aqueous stream (S2P) from the WS-2 column was recycled to the plutonium recovery process via the filtrate evaporator or tank 15. The organic effluent from the WS-2 column was then reused by the process. The americium-rich aqueous stream from the WS-1 column (SIP) was passed through the ion exchange column W-14 or W-14A and the raffinate discarded to Tank D-4 of the 241-Z Waste Retention Facility. On August 30, 1976, ion exchange Column W-14A in the 242-Z Waste Treatment Facility exploded. The column was used to recover americium-241 (241Am) utilizing Dowex 50, X-8 cation exchange resin. Several of the tanks in the tank room contained resin loaded plutonium, and aqueous and organic solutions containing gram quantities of plutonium and 241 Am. Following the accident, the 242-Z facility closed permanently. Beginning in early 1977, activities were undertaken to determine the amount of plutonium and americium-241 that were in the various tanks and columns in the 242-Z Building and develop plans to disposition the nuclear material.

Official Use Only 60

RPP-RPT-50941, Rev. 0

On January 16, 1976, the Tanks in the 242-Z Tank Room were assayed (letter, Plutonium in 242-Z Tanks). The results of the assay are presented below: Table 10. 242-Z Tank Room Assay
Tank W-2 W-3 W-4 W-5 W-6 W-12 W-13 Plutonium (grams) 70 10 10 40 5 100 35125

Until the plutonium was finally dispositioned monitoring the 242-Z Facility tanks was a semicontinuous activity. The activities are summarized below in bullet fashion. W-1 column about 180 g plutonium (letter, W-1 Column). Transfer 1,440 L of organic from W-12 to W-13. Tank 13 expected to contain about 225 g of plutonium (letter, Safety Analysis Waste Liquid in W-2, W-3, W-4, W-5, and W-6 Tanks in 242-Z Building). The resin from Column 14 (the second of the two columns used for americium-241 recovery) was removed (letter, Removal of W-14 Coupons,). An estimated 250L of organic liquid leaked to the tank room floor from Tank W-13. Chemical analysis of a sample of the liquid indicated that the plutonium concentration was 0.026 g/L. The total inventory in the room was estimated to be at less than 200 g plutonium (Occurrence Report 3 78-74, Leakage of Organic Liquid in 242-Z Tank Room.) The resin was removed from the W-1 column. The disposal of liquids in Tanks W-3, W-4, and W-5 was completed by April 26, 1979 by transfer to the 241-Z sump. The volumes of liquid from each tank and Pu/Am241content were: Tank W-3 790 L, 5.5 g Pu and 4.7 g Am Tank W-4 150 L, 8.4 g Pu and 6.3 g Am Tank W-5 110 L, 4.0 g Pu and 5.0 g Am (letter, Americium Exchange Column Accident).

Official Use Only 61

RPP-RPT-50941, Rev. 0

In February 1990, all the organic solutions within the 242-Z processing system had been transferred to Tank W-4. They comprised 80 liters of mostly DBBP. The plutonium content was estimated at 16 grams. Also all aqueous solutions have been discarded to the 241-Z. (letter, Status of 242-Z Tank Solutions) June 1982, that the contents of Tank W-4 in 242-Z were discarded to Tank D-5 (241-Z) for final disposal. (letter, Weekly Report for Week Ending June 4, 1982)

One of the last documents that was found discussing the status of 242-Z reported that the plutonium in 242-Z was originally present in the form of residues or sludges associated with wet processes. The estimated holdup was 330 g of plutonium, but Nuclear Material Accountability records showed no plutonium in 242-Z. Therefore the holdup was dispersed. Two-thirds of the holdup was estimated to be in 9 tanks (20-30 g per tank) and one third on floors and hoods. It was also believed that the hold-up estimate was a factor of 8-10 high based on the fact that all the solution was transferred to the PRF and the columns were drained and the rashing rings disposed (letter, WHC-MR-0265, PFP SAR Chapter 9 Questions on 242-Z Inventory and Ion Exchange Resin) In regard to the final cleanout of the 242-Z facility, only the plutonium in the aqueous solution was discarded as waste to the tank farms. This amounted to less than 100 g of total plutonium. Based on the operations conducted in the in the 242-Z facility, significant amounts of insoluble plutonium contained in the waste solution was deposited in the process equipment and tanks (as discussed elsewhere). How much of the plutonium solids actually passed through the facility and was subsequently transferred to tank farms is unknown. Only a few grams of plutoniumcontaining solids would have been discharged to the underground waste storage tanks during facility cleanout. Once 242-Z was shutdown and PRF continued to operate, Pu losses to liquid waste would have been higher. Soluble losses would have gone up and insoluable solids that periodically settled out in 242-Z process vessels would not have been trapped and would have been carried out with the CAW waste stream.

Official Use Only 62

RPP-RPT-50941, Rev. 0

Figure 21. Batch Waste Treatment Process

Official Use Only 63

RPP-RPT-50941, Rev. 0

Figure 22. Continuous 242-Z Waste Treatment Flowsheet

4.7.2

241-Z Sump Tanks

An aqueous liquid waste disposal system was constructed as part of the 234-5 building construction in the late 1940s. This waste disposal system included the 241-Z sump tanks, 361-Z settling tank, and a series of engineered soil waste-disposal sites (216-Z-1, Z-2, Z-1A, and later Z-3 and Z-12). Initially the aqueous liquid waste-disposal system supported the Rubber Glove Line (RGL) operations and supporting laboratory operations and, as they came online, supported the Remote Mechanical A (RMA) Line and Remote Mechanical C (RMC) Line operations. The waste disposal facility was designed to handle dilute acidic plutoniumOfficial Use Only 64

RPP-RPT-50941, Rev. 0

contaminated aqueous waste streams. As other PFP operations came online additional waste streams capable of being processed through this facility were included. Note that during periods of discharge to cribs, the 241-Z sump tanks could not have been used to dispose of high-salt wastes coming from the solvent extraction systems because of some unique aspects of aqueous aluminum chemistry. The 241-Z Waste disposal facility was composed of five interconnecting reinforced concrete cells that each contained a 4300-gal stainless steel tank. A schematic drawing of a tank and the array of tanks are shown in Figure 23. From 1949 until May of 1973, the low-salt aqueous waste streams coming from various points within PFP were collected in these sump tanks in the 241-Z facility, blended, neutralized to a pH ~ 10 and then transferred to the 241-Z-361 settling tank. Sources of the low-salt waste streams in 1969 are shown in Table 11, along with estimated volumes and amounts of plutonium lost in that stream. Table 12 provides estimates of the composition and quantities of chemicals lost in the low-salt waste stream during 1969. The lowsalt discharges after 1973 were expected to be similar, with the exception of the 232-Z incinerator stream, which was shutdown, and 234-5Z Process operations stream, where the water jet was replaced with the ANN/KOH HF scrubber system. In the RMA and RMC lines, during metal production the plutonium oxide was converted to the tetrafluoride using a continuous flow process. A six-fold excess of hydrogen fluoride was used in this process. Excess hydrogen fluoride gas from this conversion was discharged through the water aspirator used to draw a low vacuum on the hydrofluorinator system. Prior to 1973, this excess hydrogen fluoride gas was captured in the water aspirator and discharged to the 241-Z sump tanks, and carried some plutonium fluoride solids. The dilute acidic waste stream from the aspirator was neutralized and transferred to the 241-Z-361 settling tank. After its installation in 1973 (ARH-2597, Safety Analysis-Plutonium Effluent Clean-up), the aluminum nitrate-based hydrogen fluoride scrubber system reduced the volume of this waste stream going to the 241-Z sump tanks from 1.9-million gal/yr down to 60,000 gal/yr (D&D-30349, Study of Liquid Effluents and Hazardous Constituents Generated and Discharged by the Plutonium Finishing Plant). Laboratory waste steams going to the 241-Z facility included liquid wastes from both the analytical and development laboratories. These waste streams would likely have carried small quantities of common laboratory chemicals and radionuclide contamination. Compositions of liquid wastes going to 241-Z from hood floor drains are ill-defined. The scrubber and wash solutions from the off-gas scrubber of the Z Plant incinerator are of interest because these solutions could have contained plutonium oxide particles.

Official Use Only 65

RPP-RPT-50941, Rev. 0

Figure 23. Schematic drawing of a 241-Z Tank and Tank Array

(Taken from Kasper 1982, RHO-ST-44)

Official Use Only 66

RPP-RPT-50941, Rev. 0

Table 11. Sources and Volumes of Discharges to 241-Z Facility in 1970

(Taken from Kasper 1982, RHO-ST-44. The reference Knights et al., 1970 is: Knights, LM, DL Merrick, and GW Upington, Waste Management Program Plutonium Finishing Plant, ARH-1740, Atlantic Richfield Hanford Company, Richland, WA)

Table 12. Estimated Contaminants in Low-Salt Wastes and Nominal Composition in 1969

(Taken from Kasper 1982, RHO-ST-44. The reference Knights et al., 1970 is: Knights, LM, DL Merrick, and GW Upington, Waste Management Program Plutonium Finishing Plant, ARH-1740, Atlantic Richfield Hanford Company, Richland, WA)

Official Use Only 67

RPP-RPT-50941, Rev. 0

4.7.2.1 Solids Buildup and Neutron Monitoring of 241-Z Tanks. The presence of solids buildup in the 241-Z tanks has been a concern in the past. (letter RE 5.13.1, Neutron Monitoring of 241-Z Tanks). Solids were believed to be in the bottom of D-7 Tank in 241-Z. On May 16 and 17 in 1962, the D-4, D-5, D-6, and D-7 tanks in 241-Z were monitored with a neutron probe. Positive results were obtained on the bottom of the D-7 tank via the monitor well and riser, but not in the D-4, D-5, D-6, or D-8 tanks. Roughly 260 to 660 g of plutonium was estimated in the solids on the bottom of D-7 Tank. Acid flushes on October 29 and November 2, 1962 both with cold and hot with nitric acid only removed about 10 g of plutonium. A follow-on neutron counting of the same locations indicated that the plutonium content had doubled since the May 1962 assay. Solids were subsequently removed in November 13, 1962. Analysis of two types of solid from the tank indicated that the dark solids and suspended solids showed a negligible amount of plutonium. Tests showed incomplete dissolution of the fast settling solids in 10M nitric0.5 M HF and indicated about100 or more fold quantity of plutonium in the heavy, light colored solids. On December 6, 1962, the tank was entered and approximately 3 gallons of sand and gravel were recovered. The plutonium content of the solids was about 18 g. Preliminary analytical results indicated about 35 g of plutonium, inconsistent with the neutron monitoring results. It was concluded that the D-7 tank had been emptied of significant quantities of solid. The installation of pump agitators in the D-7 and D-8 tanks was recommended to minimize the accumulation of solids in the future. Neutron monitoring was eventually discontinued in favor of non-destructive assay. In 1976, NDA of TK D-5 was performed in the presence of two feet of liquid, which made results difficult to interpret (internal letter, NDA Measurement of 241-Z, TK D-5). The scientist reported; 20 gms of Pu in floating solids 200 gms of Pu in solution 20 kg of Pu in solids.

The recommendation was made to empty the tank prior to any more NDA efforts. Jetting the tank empty prior to NDA would become common practice. In the early 1980s when PFP wastes were neutralized at 241-Z in TK D-5 prior to transfer to Tank Farms the process specification was changed to include routine NDA of TK D-5 (internal letter 65240-80-161, Technical Basis Process Specification Limits 10.4.f, 10.4.h &10.4.i). Neutralization would cause solids to precipitate and these could remain in the vessel. NDA would be required for every 400 grams throughput to ensure buildup was not occurring. Calculations showed that solutions should not remain if the solid to liquid ratio was at least 1 to 4. In 1985, during RMC line operation, a significant release of plutonium occurred to the TK D-4 tank [RHO-RE-EV-85, 1985 Remote Mechanical C (RMC) Line Campaign Report] During the RMC line operation in the 1980s, only the filtrate evaporator in PRF would be operated and concentrated filtrate would be stored in the PRF tanks for future processing.

Official Use Only 68

RPP-RPT-50941, Rev. 0

Filtrate evaporator overheads and steam condensate would be discharged to TK D-4. On September 23, 1985, a criticality prevention specification (CPS) violation was declared when it was discovered that a failure of the tube bundle caused a release of filtrate directly to TK D-4 and that about 2500 grams of Pu had accumulated in a tank with a normal limit of 100 grams and criticality limit of 400 grams. A recovery plan was put into place on September 24, 1985, that called for dilution and transfer of the accumulation to tank farms in small 100 grams batches until the accumulation was removed. All 241-Z tanks involved, TK D-4, D-5, D-7, and D-8 were non-destructively assayed after completion of the last transfer to ensure no significant heel remained (Recovery Plan 9/253/1985 241-Z CPS Violation). The last waste transfers from TK D5 in 2004 included significant Pu content. At one point over 1000 grams were estimated to have collected in TK D8, resulting in a CPS violation. Potassium permanganate was added to TK-D8 to reduce the accumulated heel. After transfer to TK-D5, large ferric nitrate additions were made resulting in high solids content. The next to last waste transfer, Batch 286 was analyzed to have over 23% solids (HNF-27735, Operating, Analytical and Engineering Data for 241-Z Tank D5 Batch 286). The D-4, D-7, and D-8 tanks were transferred to D-5 using steam jet. The normal D5 turbine pump and jet could not be used the transfer to tank farms due to the high solids content. Dual air-operated diaphragm pumps were installed and used for the final transfers and flush water. An air operated pump was specified because of the high solids content, (HNF-FMP-03-18568, 241-Z Tank D5 Transfer System Modification). The D-5 tank was flushed twice with water during the last transfer, (HNF-30206, 241-Z D-5 Cell RCRA Closure). When the 241-Z Facility was closed in 2005-2006, some minor solids buildup in the tanks was noted. It is estimated, using NDA, that 348 grams of Pu (criticality value) were remaining in the facility prior to cleanout. The interior of the tanks were photographed prior to cleaning. Access holes were cut into the sides of the tanks for physical removal of the deposits and cleaning the tank interior. Plutonium-bearing material that was removed was disposed as solid waste. A photograph of TK-D4, is shown in Figure 24 showing initial deposits and final deposits. TK-D-5 was inspected and found to contain a small amount of solids and some staining on the walls. A photograph taken during this remote inspection is shown in Figure 25. A photograph of the tank interior that was taken through the access hole in shown in Figure 26. Cleaning of Tank D-8 is shown in Figure 27. Table 13 shows the before and after plutonium criticality values determined by NDA for the 241-Z tanks. (Safeguards values were less). Table 13. 241-Z Tank Values Before and After Final Cleaning
Tank Before Cleaning (after last waste transfer After final Cleaning Waste removed
(2) (1)

D-4 137 gms 11 gms ~116 gms

D-5 7 gms 20 gms ~10 gms

D-6 9 gms 6 gms ~6

D-7 72 gms 25 gms ~61

D-8 123 gms 26 gms ~82 gms

(1) internal letter M2100-05-174, Summary of NDA Result for 241-Z B.D. Keele to G.A. Johnston, November 8, 2004 (2) HNF-33999, 241-Z As-Left Characterization, R.W. Bloom and S. D. Elingston, June 2007

Official Use Only 69

RPP-RPT-50941, Rev. 0

Although solids build-up was a periodic occurrence in the 241-Z facility, the mechanical agitation provided by the paddle agitators was sufficient to keep most solids suspended. Routine NDA was adequate to ensure no buildup was occurring during operations known to generate solids, but the NDA had to be done with a near- empty tank to avoid erroneous results. Airoperated pumps were required for the last waste transfers with high Pu content and high solids content. No significant build-up of solids was present when the tanks were closed. Figure 24. Tank D4 Showing Initial Deposits and Final Deposits

Official Use Only 70

RPP-RPT-50941, Rev. 0

Figure 25. May 2005 Inspection of TK D-5 Interior Floor South Side and Agitator

Figure 26. Access Cut through TK-D5 Tank Wall for Interior Cleaning

Official Use Only 71

RPP-RPT-50941, Rev. 0

Figure 27. Removing Deposit from TK-D8 during Decontamination and Decommissioning

4.7.3

241-Z-361 Settling Tank

Prior to routing PFP low salt waste solution to tanks farms in 1973, these solutions went to cribs via the 241-Z Facility and the 241-Z-361 Settling Tank. The 241-Z-361 Settling Tank (361-Z tank) is a rectangular structure, constructed of reinforced concrete with a 3/8-in. carbon steel liner (Figure 28, taken from RHO-ST-44, 216-Z-12 Transuranic Crib Characterization: Operational History and Distribution of Plutonium and Americium) provides the tank plan view and a side view. The interior of the tank is 13 by 26 ft with the depth varying from 17 ft at the inlet to 18 ft at the outlet. The tank had a working volume of approximately 4,900 ft3 or 139,000 L and currently contains approximate 75,000 L of sludge. This settling tank was associated with the disposal of PFP low-salt waste streams. These waste streams were treated in the 241-Z sump tanks before they were transferred to the 241-Z-361 settling tank. Liquids from the 241-Z-361 settling tank overflowed to engineered soil waste-disposal sites. An extensive amount of information is available on the historical operations of the 241-Z-361 Settling Tank (HNF-1989; HNF-2012, Engineering Study of the Critically Issues Associated with Hanford Tank 241-Z-361; D&D-30359, Study of Liquid Effluents and Hazardous Constituents Generated and Discharged by the Plutonium Finishing Plant; RHO-ST-44, Transuranic Crib Characterization: Operational History and Distribution of Plutonium and Americium).

Official Use Only 72

RPP-RPT-50941, Rev. 0

Figure 28. 361-Z Settling Tank System

(Taken from Kasper 1982, RHO-ST-44)

In the past, there were considerable concerns about the quantities of plutonium in the 241-Z-361 settling tank and the related potential criticality issues. One study estimated that a little over a kg of plutonium/year was being lost in the low-salt waste stream going through the 241-Z sump tanks to the 361-Z settling tank. Efforts to resolve potential criticality issues in the 361-Z tank are documented in Lipke (1997) and Jones (1998). Based on data collected to support resolution

Official Use Only 73

RPP-RPT-50941, Rev. 0

of potential criticality issues, Lipke concluded that the 241-Z-361 settling tank contained approximately 30 kg of plutonium in the 75 m3 of sludge remaining in this tank. This settling tank operated from 1949 until 1973 (beginning in May 1973, these PFP aqueous wastes were sent to the tank farms). The most recent characterization effort for the sludge currently in the 361-Z tank was reported by Hampton and Miller (HNF-8735, 241-Z-361 Tank Characterization Report). Two cores were taken from the 361-Z tank and characterized to address regulatory issues driven by RCRA. Later, analytical data included in HNF-8735 was used to develop an inventory estimate for the sludge currently in this tank (SGW-35955, Inventory Estimates for Sludge Currently in Tank 241-Z-361). These inventory estimates are shown in the following table.

Table 14. Inventory Estimates for Sludge Remaining in Tank 241-Z-361

(from Jones 2008)

Previous estimates conclude that there is approximately 30 kg of plutonium in the there 241-Z-361 tank. Based on studies of the waste streams being discharged through this tank it is likely the plutonium particles would be small and associated with the other metal oxides found in this tank. However, no solids characterization data were found that would provide quantitative information of plutonium particle sizes in the sludge in this tank. It is likely that the majority of plutonium currently in the 361-Z tank originated from the hydrofluorinator systems in the RMA and RMC lines. There was a five to six fold excess in the amount of hydrogen fluoride gas that passed through the high-temperature hydrofluorinator system. The hydrogen fluoride gas then passed through two successive carbon filters (with a 69 micron pore size [ARH-2597]) then went to the water aspirator. The dilute hydrofluoric solution produced by the water aspirator, likely containing some plutonium fluoride particles, went to the 241-Z sump tanks for neutralization to a pH around 10 and then transferred to the 361-Z settling tank. Hydrolysis of the plutonium fluoride in the 241-Z sump tanks and 361-Z settling tank would likely have been slow because of the insolubility of plutonium tetrafluoride solids around a pH 10. Since this neutralized waste stream passed through a settling tank with an approximate 3 to 6-day residence time (based on data from Table 14 and previously reported estimate in SGW-35955), most of the plutonium solids would be expected to have settled out in the 241-Z-361 tank. The close agreement Official Use Only 74

RPP-RPT-50941, Rev. 0

between the quantity of plutonium lost through the 241-Z-sump tanks, based on accountability measurements, and the amount currently in the 361-Z settling tank and the reasonably small quantities of plutonium found in characterization of the Z-12 crib (RHO-ST-44) supports the assumption that the 361-Z settling tank functioned reasonably well in removing solids from the low-salt waste stream. Liquids overflowing from the 241-Z-361 settling tank gravity-drained to subsurface waste disposal sites. Because the 241-Z sump tanks/241-Z-361 settling tank system provided an efficient method for processing liquids into a pH range where the plutonium and americium would be expected to quickly sorbed by the soil, it is likely that as many of the PFP liquid waste streams as possible were processed through this facility. It is also clear that waste streams containing high concentrations of aluminum ions (i.e., high-salt waste streams) could not have been processed through the 241-Z sump tanks and 241-Z-361 settling tank because of the limited capability to neutralize high-aluminum waste streams in that facility. When complete neutralization of solutions containing high concentrations of aluminum ions is required, then a reverse strike technique, such as that used in the REDOX Process, is required. [After 1980, when high salt (CAW) wastes were being neutralized in 241-Z sump tanks this waste stream was neutralized using a reverse strike method.] 4.7.4 Trenches and Crib

An understanding of the characteristics of plutonium solids found in PFP process wastes that were discharged to the soil column would be helpful in defining the nature of plutonium solids that may have been transferred to the tank farms after May 1973. With the decision to transfer PFP aqueous liquids to the tank farms, a number of changes were made in PFP waste disposal practices. These changes could have significantly modified the nature of plutonium solids going out in these waste streams. This issue is discussed in Section 4.0. As shown in Figure 29, there are about two-dozen liquid waste disposal sites associated with the Z Plant Complex and 231-Z Facility operations. However, only five Z Plant Complex waste sites received liquid waste discharges from the 241-Z-361 settling tank, one waste site received all RECUPLEX high salt waste discharges, and two waste sites received essentially all of the PRF high-salt waste discharges. Extensive details on Z Plant waste site construction and operations are available in RHO-LD-114, Existing Data on the 216-Z Liquid Waste Sites. PFP Low-Salt Discharge Sites From June 1949 to June 1952, liquids drained from the 241-Z-361 settling tank to the interconnected 216-Z-1 and 216-Z-2 cribs, then from July 1952 until March 1959 these liquids drained to 216-Z-3 crib. These 3 cribs were designed to overflow to the 216-Z-1A tile field. From March 1959 until May 1973 the liquids from the 361-Z settling tank drained to the 216-Z-12 crib. Although the three cribs, 216-Z-1, -2, and -3 were plumbed to allow liquids to overflow to the 216-Z-1A tile field, it is unclear if any plutonium contamination from the 216-Z-1, -2, and -3 cribs overflow reached the Z-1A tile field. PFP High-Salt Discharge Sites The RECUPLEX high-salt waste stream was discharged to the 216-Z-9 covered trench (Z-9 crib). The 216-Z-9 crib also received the spent organic solvents from the RECUPLEX solvent extraction operations. There was clean water flushes associated Official Use Only 75

RPP-RPT-50941, Rev. 0

with waste discharges to the 216-Z-9 crib. There was a high-silica waste stream associated with the reprocessing of the magnesium oxide/calcium fluoride solids originating in the plutonium reduction process. This high-silica waste stream was discharged to the 216-Z-8 tank and French drain and appears to have been far more benign than the high-salt waste going to the 216-Z-9 trench. From 1964 until 1969, PRF high-salt wastes were discharged to three locations in the 216-Z-1A tile field; from 1969 through May 1973, these wastes were discharged to the 216-Z-18 crib. For two very short periods of time in 1966 and 1967, the PRF high-salt waste was discharged to the 216-Z-1 and 216-Z-2 cribs while piping was being installed to move the point of discharge of wastes in the 216-Z-1A tile field. Spent Organic Solvents As in the case of the RECUPLEX high salt waste discharges to the 216-Z-9 crib, spent organic solvents from PRF were also discharged to the waste sites that received the PRF high-salt aqueous waste stream. Reprocessed lard oil/carbon tetrachloride, called fab oil, was discharged to the cribs that received the high-salt and spent organic solvents from the solvent extraction systems. An unpublished whitepaper contains extensive detail on the history of processing and disposal of lard oil at PFP (letter, Clean-Up and Disposal of Machining Effluents). Large particulate plutonium was removed from the lard oil by filtration. The Fab oil was filtered prior to storage with a 50 micron and 5 micron size filter (HW-76399, Process Design Basis for Fabrication Oil Storage and Sampling Facility 234-5 Building). The Fab oil has been reported to contain 0.1 2 gl Pu as fines or as a solution or colloidal suspension (HW-75275, Flowsheet and Equipment for Plutonium Recovery from Fabrication Oil). The plutonium-bearing particles have been reported to range in size from 0.00115 microns, and the particles that were >0.5 microns contained as high as 36% Pu (HW-74350, Recovery of Plutonium from Fabrication Oil).

Official Use Only 76

RPP-RPT-50941, Rev. 0

Figure 29. Overview of Z Plant Complex Liquid Waste Disposal Sites

(Taken from Kasper 1982, RHO-ST-44)

Official Use Only 77

RPP-RPT-50941, Rev. 0

There have been a number of field and laboratory characterization efforts at waste sites that received Z Plant high-risk waste streams. These included studies at 216-Z-9, 216-Z-1A and 216-Z-12 to evaluate the extent of plutonium movement and characterization. Recently, two additional boreholes were drilled near 216-Z-9 to collect additional information on plutonium movement in the soil column. A description of the 216-Z-9 crib and dimensions are shown in Figure 30 (HNF-31792, Characterization Information for the 216-Z-9 Crib at the Plutonium Finishing Plant). In the late 1960s and early 1970s there was considerable concern about the amount of plutonium that had accumulated in the top layer of soil in the 216-Z-9 covered trench. Field characterization results suggested the neutron flux at the surface was approaching levels that could result in a criticality accident. This led to extensive characterization activities at 216-Z-9. Results from these studies can be found in Smith 1973 (ARH-2915, Nuclear Reactivity Evaluations of 216-Z-9 Enclosed Trench). This document reported that the plutonium concentrations were highest in the first few inches of soil (see Figure 31) and generally dropped by two orders of magnitude at 1 ft below the base of the trench (ARH-2915, Nuclear Reactivity Evaluations of 216-Z-9 Enclosed Trench). Based on very limited data, the plutonium concentration remained relatively constant between 1 ft and 8 ft below the trench floor. Based on the characterization data reported in ARH-2915, it was estimated that the first foot of soil in 216-Z-9 contained 38 kg of plutonium (the range was 26 to 69 kg). Analytical data from soils removed from 216-Z-9 showed that 58 kg of plutonium had been removed. It has been suggested that the total plutonium discharges to 216-Z-9 were in the order of 100 kg (HNF-31792, RHO-ST-21, Report on Plutonium Mining Activities at 216-Z-9 Enclosed Trench). Based on accountability records, 27.4 kg of plutonium was discharged to 216-Z-9 (Owens 1981). A number of studies have evaluated the nature of plutonium solids found in two high-salt waste disposal sites, 216-Z-9 crib and 216-Z-1A tile field. These include Ames (BNWL-1812, Characterization of Actinide Bearing Soils: Top Sixty Centimeters of 216-Z-9 Enclosed Trench), Price and Ames (ARH-SA-232, Characterization of Actinide-Bearing Sediments Underlying Liquid Waste Disposal Facilities at Hanford), Ames (BNWL-2117, Proceedings of an ActinideSediment Reactions Working Meeting at Seattle, Washington on February 10-ll, 1976), Price et al (RHO-ST-17, Distribution of Plutonium and Americium Beneath the 216-Z-1A Crib: A Status Report), and Felmy (2011, private communication). Ames (BNWL-1812, BNWL-2117) reported results from solids characterization of soil core samples collected from the high-salt receiver site, 216-Z-9 trench. Price discussed solids characterization of soil samples from 216-Z-9 and the 216-Z-1A tile field (ARH-SA-232, RHO-ST-17). Felmy reported preliminary information from an ongoing investigation of soil samples evaluated by Ames in earlier studies. Each of the studies has reported the existence of large plutonium oxide particles in soil samples from near the point of release in the liquid waste disposal sites. Ames reported plutonium oxide particles up to 100 microns in diameter that appear to be crystalline plutonium oxide. Price (ARH-SA-232) reported that the analysis of seven of the eight plutonium particles examined indicated the presence of PuO2. Microprobe analysis of the same particles showed that they contained 70 to 95 wt% PuO2. Felmy reported that analysis of a less than 40-micron

Official Use Only 78

RPP-RPT-50941, Rev. 0

fraction of soil samples from 216-Z-9 crib found numerous PuO2 particles, one a near-spherical 30-micron particle and the second a 30-by-50 micron barrel shaped particle. The first Ames paper (BNWL-1812) stated the plutonium oxide particles in the first 15 centimeters of soil were generally 10 microns or less but the other two publications (ARH-SA-232, BNWL-2117) state that the plutonium oxide particles generally fell in the 2 to 25 micron range. The sludge overlaying the soil contained plutonium oxide particles that were 10, 25, 50, and 100 m, and had plutonium concentration that varied from 62 to 84% (BNWL-1812, letter Characterization of 216-Z-9 Trench Soils X-ray Diffraction Identification of Plutonium Particles). The 216-Z-9 and 216-Z-1A soils studies report found two types of plutonium solids in these soil samples. The first was crystalline plutonium oxide particles, generally falling into the 2 to 25 micron diameter range, found in the first few centimeters of the soil column. The second type of plutonium solid found deeper in the soil column was non-crystalline polymeric material associated with altered soil particles. They suggest that the more mobile plutonium was associated with plutonium that was in solution as Pu(IV) in the acidic high-salt waste stream being discharged. They postulate that as this waste stream was neutralized by interaction of acidic waste with the soil as mobile plutonium was removed from solution. The significant point is that evaluation of the high-salt waste disposal sites provides evidence for the routine discharge of both plutonium oxide particles with size up to 100 m in diameter and soluble plutonium species in the high-salt/spent organic solvent waste stream. The extensive analyses of the plutonium plume in the soil column below the 216-Z-9 crib provide a rough estimate of the fraction of plutonium oxide (mainly found near-surface) to total plutonium discharged to the waste site.

Official Use Only 79

RPP-RPT-50941, Rev. 0

Figure 30. Description of the 216-Z-9 Trench and Dimensions

(From Teal 2007)

Official Use Only 80

RPP-RPT-50941, Rev. 0

Figure 31. Plutonium Concentrations on Surface of 216-Z-9 Crib

(from ARH-2915)
Price et al (1979) reported on detailed field and laboratory studies of plutonium and americium in the 216-Z-1A tile field (that had received PRF high-salt waste from 1965 until 1969). The highest concentrations of plutonium and americium were found immediately below the base of the 216-Z-1A tile field (consistent with observations from 216-Z-9), lower levels of these actinide contaminants had moved down considerably below the bottom of the 216-Z-1A tile field. Plutonium oxide particles much greater than 10 microns were identified in soil samples near the base of the crib, similar to those that had been found in the 216-Z-9 covered trench. The same mobility patterns found at 216-Z-9 and 216-Z-1A would be expected at 216-Z-18 since all three waste sites received very similar waste types.

Official Use Only 81

RPP-RPT-50941, Rev. 0

All of the Z Plant waste sites that received the high-salt waste also received the spent organic solvents including organic degradation products and lard oil/carbon tetrachloride mixtures from the solvent. The 216-Z-9 trench is projected to have received some 60 tons of the carbon tetrachloride/lard oil mixture, as shown in Table 15. Table 15. RECUPLEX Input to Z-9 Crib

(Taken from Teal (2007) and Letter, L.E. Bruns to A.E. Smith, RECUPLEX Inputs to Z-9 Crib, April 10, 1973.)

Detailed characterization of low-salt waste disposal sites is limited to the 216-Z-12 crib. In 1969, Crawley (ARH-1278, Plutonium-Americium Soil Penetration at 234-5 Building Crib Sites) reported that the plutonium and americium in the 216-Z-12 crib were concentrated in the soil near the level that liquid waste entered the crib with little evidence of downward movement (ref). This plutonium sorption profile was expected in cribs receiving the neutralized low-salt waste stream coming from the 241-Z-361 settling tank. In 1981 Kasper (RHO-ST-44) reported a detailed field and laboratory characterization of plutonium and americium contamination in the characterization soil column below the 216-Z-12 crib. Although there does not seem to have been a quantitative analysis done on data from the multiple well locations, the quantities of actinides found do not seem to indicate that large quantities of plutonium were discharged to the soil column in the lowsalt waste stream (certainly not the 25.1 kg of plutonium stated to be in 216-Z-12 crib by Kasper). The plutonium sorption behavior observed at 216-Z-12 is consistent with what would Official Use Only 82

RPP-RPT-50941, Rev. 0

be expected for a slightly basic waste stream being discharged to the Hanford soil column. Felmy (2011, private communication) reported that STEM analysis of a Z-12 crib soil sample found no plutonium oxide particles. A number of gamma-ray and neutron logging studies have been reported for drywells associated with the 216-Z-12 crib. These data were reviewed in a recent document [SGW-42004, Summary Report for Selected Boreholes within the 216-Z-1A Tile Field, 216-Z-9 Trench, and 216-Z-12 Crib]. These studies indicate that actinide contamination is found near the top of the soil column with little evidence for downward movement. A high neutron flux and gamma-rays associated with fluoride alpha-N reactions were found in the region of high actinide activity indicating the likely presence of plutonium fluoride solids, or at least plutonium in close contact with fluoride ions. 4.7.5 Conclusion of Plutonium Losses from PFP to Cribs

Characterization of soil samples from two PFP high-salt waste disposal sites (Z-9 and Z-1A) indicated that plutonium discharged involved plutonium oxide particles that appeared to have been filtered out of the liquid waste stream in the first few inches of the soil column. At both sites the plutonium concentrations drop by a couple orders of magnitude in the first foot of soil. As discussed in Sections 4.11 and 4.12, plutonium solids in the liquid waste stream were not accurately measured nor accounted for in the plutonium accountability records. The available characterization data from 216-Z-9 crib is useful in understanding the magnitude of unmonitored plutonium losses in the high-salt waste stream and to some extent the particle size of the insoluble plutonium. It has been estimated that approximately 4.5 metric tons of plutonium were processed through RECUPLEX during its seven years of operations. An estimate is that approximately 100 kg of plutonium was discharged to the 216-Z-9 crib (HNF-31792). These numbers would suggest a 3.3% loss of the total amount of plutonium recycles through RECUPLEX going to the liquid waste streams to 216-Z-9 crib. Plutonium oxide solids are one source of the unmonitored plutonium in the high-salt waste streams. Results from the series of Ames and Price documents, and unpublished results for Andy Felmy, indicated that it is likely that plutonium oxide particles with diameters significantly greater than 10 microns make up a significant mass fraction of this unmonitored plutonium losses in the high-salt waste stream going to the soil column. One source of plutonium oxide particles is from the recycling of plutonium metal. In the first step, plutonium metal was burned to form the oxide (HW-77531, Characteristics of Burning Plutonium). This oxide was then dissolved in nitric/hydrofluoric acid and processed through the solvent extraction systems for recovery of plutonium as the nitrate. The dissolution of plutonium oxide in acids is sensitive to the thermal history of the oxide. (see discussion in Appendix C, Report 3). Other sources of oxide particles were the slag and crucible wastes and incinerator ash. Plutonium oxide particles were entrained in the acidic feed to the solvent extraction systems. The particles would have exited the solvent extraction column in the aqueous (CAW) waste stream or as silica-degraded organophosphate containing plutonium solids in the plutonium-rich Official Use Only 83

RPP-RPT-50941, Rev. 0

organic stream existing at the bottom of the CA column. The solids tended to accumulate at the interface between the organic and aqueous phases at the top of the CA column. Materials that accumulated at this interface were referred to as interface crud. Periodically, as needed, during column operations, this interface crud layer was discharged to the aqueous (CAW) waste stream (see Appendix C, Report 4). Another source of plutonium reaching the soil disposal sites was in the spent organic solvents. The degraded organics led to poor plutonium recoveries in the solvent extraction systems. In the early years, organic solvents were replaced as needed and the spent organic solvents were discharged to the soil column. Dibutyl-phosphate (DBP) was formed in the decomposition of tributyl-phosphate (TBP). DBP forms a much stronger complex with plutonium(IV) than does TBP. The Pu(IV)-DPB complex remains in the organic phase during the plutonium stripping step in the solvent extraction systems. The Pu(IV)-DBP complex was a likely source of much of the mobile plutonium observed in 216-Z-9 and Z-1A cribs. There were significant changes in organic waste disposal practices once aqueous wastes were being sent to the tank farms. Smetana (ARH-CD-323, Z Plant Liquid Waste Disposal Through the 241-Z Vault) reports that about 2% of the organic solvent used at PFP was lost in the aqueous waste stream going to the tank farms, mostly as entrained droplets. Spent organic solvents were loaded out and stored in retrievable storage in PFP burial grounds. Enhanced solvent clean-up steps were added in 1973 to extend the operating life of the organic solvents (ARH-323). Such steps likely increased the amount of plutonium going out in the aqueous waste stream but such losses likely would not have involved large plutonium particle sizes. An additional source of mobile plutonium would likely have been from complexed Pu(IV) in the feed stream going to the CA column. If plutonium (IV) were complexed with an anion such as fluoride, chloride, sulfate, or phosphate then it would not have been extracted into the organic phase in the CA column. Such complexed plutonium species would likely have gone out in the CAW aqueous waste stream and then to cribs. The plutonium complexes would not have survived long in the soil column environment. The presence of insoluble plutonium containing particles was a constant issue over the lifetime of PFP. During the operating history of PRF, efforts to removed insoluble unaccountable plutonium were intensified. The more rigorous use of centrifuge technology to reduce losses of plutonium solids in the high-salt waste stream is an example of changes implemented at PFP. In summary, the Pu discharge to the cribs prior to 1973 reveals the following: 1. Based on historical data from characterization of soil samples taken from PFP high-salt waste disposal sites used from 1955 until 1969, plutonium oxide particles were discharged to the soil column in the CAW/spent organic solvent waste stream. 2. The plutonium oxide particles found in high-salt waste disposal sites were reported to be in the range of 2-25 microns, with individual particles found to be up to 100 microns.

Official Use Only 84

RPP-RPT-50941, Rev. 0

3. The awareness in the late 1960s that non measured plutonium was being lost in the highsalt waste stream and that these losses were associated with plutonium oxide solids, led to the development and implementation of new techniques for reducing plutonium losses in the high-salt waste stream. 4. It has been estimated that approximately 4.5 metric tons of plutonium was processed through RECUPLEX during its 7 years of operations. An upper bound estimate is that approximately 100 kg of plutonium was discharged to 216-Z-9 crib. These numbers would suggest a 3.3% loss of the total amount of plutonium recycles through RECUPLEX going to the liquid waste streams to 216-Z-9 crib. 4.8 SCRAP MATERIAL PROCESSED OUTSIDE OF PFP

As discussed in Section 4 Plutonium Finishing Plant (Z Plant), the PFP Complex, including 231-Z Facility, was designed to be a self sufficient stand-alone plant for the conversion of plutonium nitrate into oxide and metal products, with all necessary scrap recovery, analytical, and developmental capabilities. The bulk of the plutonium solids discharged to the underground waste storage tank came from the PFP plutonium entrained in the liquid waste discharged to 242-T after 1973. This is discussed in Sections 4.11 and 4.13. However the PFP did transfer liquid plutonium containing scrap streams to the T Plant, REDOX, and PUREX for plutonium recovery. Both PUREX and REDOX processed irradiated fuel containing plutonium dioxide. Those liquid waste streams from those operations would have contributed solid plutonium particles to the underground waste tanks. A discussion of the operations conducted at those facilities is in Section 5.0, Fuel Processing Facilities (T/B Plant and REDOX and PUREX).

4.9

PLUTONIUM PARTICLE SIZE SUMMARY

A comprehensive review of PFP product and scrap particle size data was presented by Hoyt in 2005, (NMS-15232, Overview of Plutonium Oxide Particle Size Distributions and Respirable Fractions). This work included data produced by LANL that was later found to be suspect. Particle size data presented in the original WTP report included some of this LANL data but it was accompanied with a disclaimer that follow work had shown it to be in error. LANL has begun re-analysis of the some of the original materials. When the WTP report was revised after peer review, it included as much re-analysis of the LANL material as was currently available. The WRPS review team reviewed all current particles size data compiled by the original WTP report authors (24590-CM-HC4-W000-00176-T02-01-00001) and did not locate any new, more comprehensive information. The WRPS team saw no reason to question the WTP report validity or completeness; thus the particle data from that report is presented in this section. A significant amount of plutonium scrap characterization work, including particle size distribution (PSD) measurements, was performed on scrap materials at all of the DOE Sites and the DOE National Laboratories. Excess nuclear material from all DOE sites, including the National Laboratories, were stabilized and packaged into new 3013 long-term storage containers, and placed into storage at the Savannah River Site (SRS) (DOE-STD-3013-2004, Stabilization, Official Use Only 85

RPP-RPT-50941, Rev. 0

Packaging, and Storage of Plutonium-Bearing Materials). Most of these stabilization and repackaging activities were accomplished in the 1990s and through ~2005. To support this stabilization and storing mission, a Material Identification and Surveillance (MIS) project was established to characterize the material that was being placed into storage and to assess the performance of the new storage container over time. Initial work was accomplished at the Los Alamos National laboratory (LANL) on representative samples of all the different types of PuO2 materials from each site that were placed into the new storage containers. These oxide samples were characterized both physically and chemically. In 2010 LANL began to re-analyze some of the oxide PSDs with a new particle size analyzer. Some of these new PSD measurements have included both Hanford materials and plutonium oxides from the Rocky Flats Environmental Technology Site (RFETS) that were sent to Hanford for plutonium recovery. These PSDs are shown below in Figures 32, 33, 34 and 35. The data presented in Figure 32 indicate that the particles for scrap oxide are significantly larger than those reported for product Pu oxide (Figure 37), with 60-75wt% of the oxide being contained in particles larger than 10 m. Figures 33 and 34 indicate that 28-75 wt% of the RFETS Pu oxide from chloride contaminated and peroxide scrap was >10 m in size. The PSDs shown in Figure 35 indicate that Pu oxide from batch magnesium hydroxide precipitation and calcination at Hanford could range from 11-97 wt% >10 m. This wide range is not unexpected for this process since it was run only as a batch stabilization process, without any rigorous process control or specific product oxide requirements. In addition, new PSDs have been measured by LANL on burned metal oxide from their Advanced Recovery and Integration Extraction System (ARIES) process. These PSDs are also shown below in Figure 37 along with some previously reported burned metal oxide data.1 This data illustrates what has been reported in the literature; that the particle size of Pu oxide from burned metal depends upon the oxidation temperature. Room temperature oxidation results in very fine particles, whereas higher oxidation temperatures result in very large particles. The oxidation is exothermic and any metal oxidation performed in conditions where the oxidation atmosphere is not controlled is near impossible to perform at a specific temperature. Consequently, metal oxidation usually results in plutonium oxide powder having a very wide range in particle size, and it is frequently bi-modal. In addition, burned metal oxide usually contains small particles of unreacted metal in sizes up to BB size. This is the result of some of the oxide forming an adhering protective layer, and the oxide that does spall off, forming a bed of powder that inhibits oxygen diffusion and slows the reaction.

Venetz-Particle size data from ongoing analysis of oxide at Los Alamos as part of the DOE Materials identification and Surveillance (MIS) Working Group supporting DOE-STD-3013, Stabilization, Packaging, and Storage of Pu Bearing Materials.

Official Use Only 86

RPP-RPT-50941, Rev. 0

Figure 32. Particle Size Distribution for Hanford Scrap Pu Oxide LANL MIS
100 90 80

Volume Percent Less Than

70 60 50 40 30 20 10 0 0.1 1 10 100 1000

Note : T his is new unpublished 2010 MIS data from LANL. Hanford Scrap PuO2, Misc. Glovebox Scrap, Restab at 950 C Hanford PRF Scrap, Floor Sw eepings, Restab at 950 C

Particle Size (Microns)

Figure 33. Particle Size Distribution for RFETS Chloride Scrap Oxide Sent to Hanford
100 90

Volume Percent Less Than

80 70 60 50 40 30 20 10 0 0.1 1 10 100
Note : T he MIS data is new unpublished 2011 MIS data from LANL.
LANL MIS Data RFETS ER Sc rape Out, LotA1+A4 LANL MIS Data RFETS ER Sc rape Out, Lot# A3 LANL MIS Data RFETS ER Sc rape Out, Lot# A6 LANL MIS Data RFETS Pyro Sc rap Oxide

Particle Size (Microns)

Official Use Only 87

RPP-RPT-50941, Rev. 0

Figure 34. Particle Size Distribution for RFETS Pu Oxide from Peroxide Calcination
100 90

Volume Weight Percent Less Than

80 70 60 50 40 30 20 10 0 0.1 1 10 100
LANL MIS data on RFETS Pu Oxide from Peroxide Prec ip/Cal

Note : T his is new unpublished 2011 MIS data from LANL.

Particle Size (Microns)

Figure 35. Particle Size Distribution for Pu Oxide from Magnesium Hydroxide Precipitation and Calcination

Official Use Only 88

RPP-RPT-50941, Rev. 0

Figure 36. Particle Size Distribution for Pu Oxide from Burned Metal
1.0 0.9
Pu Oxide From
Room Temperature

LANL MIS ARIES Metal Oxidation Unpublished Data Referenced In BNWL-205

Com bined Data Reported In LA-12315-MS

Weight Fraction Less Than

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0 0.1

Metal Oxidation,
Then Re-Calcined At

100 and 900 C, RFP-927

Note : T he MIS data is new unpublished 2011 MIS data from LANL.
2 Gram Turnings 12 Gram Cylinders LA- 12315- MS RFP- 927: 100 C RFP- 927: 900 C LANL ARIES Metal Oxidation LANL ARIES, Short Can Oxide Mishima, BNWL- 205 Bloomster, BNWL- 205

Data In HW-77531 (Data Referenced From AERO/CONF 8, 1960) BNWL-205

10

100

1000

Particle Size (Microns)

Figure 37. Particle Size Distributions for Hanford Oxalate Source PuO2

(As Measured in the LANL MIS Program Reported in Historical Data, and for Some Recent Hanford Measurements)

From the data summarized and presented in the WTP report (24590-CM-HC4-W000-00176T02-01-00002), the following PSD data extracted, see Table 16. This data is used later for the estimate of large particulate plutonium discharged to tank farms, see section 4.14.

Official Use Only 89

RPP-RPT-50941, Rev. 0

Table 16. Summary PSD Data Used in Later Analysis


Summary Particle Size Data: Volume percent greater than Plutonium Material Pure Pu Oxide Product Pu Oxide from high temperature metal oxidization %>10 m 4589% 64100% %>20 m 540% 28100% %>30 m 08% 10100% %>40 m 05% 5100%

Practical Upper Limit on Particle Sizes Discharged to Waste The PSD data presented includes large particles, greater that 100 micron, but the subsequent analysis used a 100 micron upper limit for particles that were lost in the liquid waste stream. The basis for the 100 micron upper limit for particles discharged in the waste stream was picked based on the consensus judgment of several plutonium processing subject matter experts. This was supported by several observations: 40 micron is the practical upper limit for oxide from oxalate precipitation. 100 micron was the largest oxide particle seen in Z-9 as reported by Ames. Oxide was usually screened prior to dissolution. 100 micron seemed like the practical upper limit of particles that could be carried out of the dissolver pots and through the solvent extraction column and tanks. Settling calculation studies for the MT dissolver pots that were found in the PFP history files show 10 micron particles would settle in about 30 minutes and that 128 micron particles would settle in 8 minutes. In 1978, settling times in MT were given as 10 minutes in training materials. Later, after the settling calculation was done in 1984, this was changed to 30 minutes.

4.10

DIFFICULTIES WITH PLUTONIUM SOLIDS

4.10.1 Solids in 242-Z and PRF The presence of plutonium in solid particles in the PRF solvent extraction columns was not an uncommon occurrence. The plutonium-containing solids were not extracted in the plutonium recovery operation. Additionally, the plutonium solids were not measured by the analytical procedures used to measure the accountable plutonium in the aqueous waste solutions discarded to waste. The loss of solid plutonium containing particles in the CAW stream was readily apparent by the build-up of plutonium in the columns in the 242-Z Waste Treatment Facility. From 1965 to 1970, plutonium was recovered in 242-Z in Raschig ring-filled columns containing DBBP in CCl4. The extractable plutonium formed a Pu-DBBP complex that was soluble in CCl4. Some portion of the solid plutonium-containing particles was trapped/filtered on the Raschig rings.

Official Use Only 90

RPP-RPT-50941, Rev. 0

This plutonium buildup in the W-2 and W-3 columns was a continual problem. This gradual build-up of non recoverable plutonium was noted for several years and the plutonium would build up to kilogram levels. Periodically, the plutonium-containing solids were removed from the columns and the plutonium recovered. Table 17 provides examples of solid build-up that were reported. While the recovered plutonium reduced the measured inventory difference, it did nothing for the under accounting for the amount of plutonium going to the waste. For the period of time that the 242-Z operated, the quantity of unmeasured solid plutonium discarded to waste was reduced (as noted earlier). In 1970, the static solvent extraction process was replaced by a continuous solvent extraction system. The solvent extraction columns were placed in the PRF Canyon in the PRF building, but the receiver tanks were in 242-Z. This did not reduce the problem of the presence of solids in the aqueous waste stream, but it did reduce the buildup of the plutonium accumulation in the 242-Z tanks. Plutonium-containing solid particles still settled out in the 242-Z tanks (effectively a wide spot in the waste line that allowed large plutonium-containing particles to settle out and not ultimately end up in the discarded aqueous plutonium-containing waste. A laboratory study on the solids removed from Raschig rings in W-12 Tank contained greater than 90% of the trapped plutonium (ARH-195- RD-BK-1, Monthly Activities Reports Separations Chemistry Lab CY 1971 Book 1). Two contacts of boiling HNO3-HF recovered 97% of the plutonium. The amount of solids on the rings suggested that a great deal of solids were obtained from CAW or caustic-adjusted CAW. The high surface area of the W-2 packed column loaded with CCl4 appeared to create a trap for CAW solids. The plutonium was not absorbed onto the rings but filtered out as solids (ARH-423-RD, Plutonium Process Engineering Monthly Reports, September 1967). The source of the solids in the PRF and 242-Z has been described in monthly and engineering reports from the mid 1960s. Table 17 and Table 18 provide the select details noted in some monthly reports. The sources of and composition of the plutonium containing solids included: Pu-DBP precipitate in the CC column. This is attributed to low acid strength and a buildup of dibutyl phosphate (DBP) in the organic. Precipitation of plutonium fluorides and carbonates Dark green floating solids were identified as a silica-organophosphate containing plutonium. High plutonium loss in CAW would sometimes correspond to ash dissolver runs. Incinerator ash did not dissolve completely and solids passed through the in-line filters.

Because of the large amount of unaccounted for material, the introduction of centrifuges and filters in the PRF solvent extraction operation was initiated to recover plutonium containing solids, including the interfacial crud from the solvent streams. This is discussed in detail in the next section.

Official Use Only 91

RPP-RPT-50941, Rev. 0

Table 17. Examples of Build up of Plutonium Containing Solids in 242-Z


Date September 1967 April 1968 March 1969 April 1969 September 1971 Notes There is a gradual build up of plutonium in the W-2 column. The plutonium cannot be removed by vigorous flushing Build-up of solids in W-3 and W-4 tanks in 242-Z approximately 100 to 150 g of plutonium respectively About 700 g of plutonium solids are held-up in the W-3 and W-4 columns as of February Forty six quarts of plutonium-bearing solids were removed from the W-3 and W-4 columns containing about 800 g of plutonium. Gamma monitoring of the W-2 column indicates that there is a gradual build up of plutonium. It cannot be removed by vigorous flushing. It is assumed to be associated with the organic absorbed into the plastic rings.

Table 18. Examples of the Causes of Presence of Solids in the PRF Solvent Extraction System
Date April 1968 Comment During testing of the Pu-U partition flowsheet, there was formation of a Pu-DBP precipitate in the CC column. This is attributed to low acid strength and a build-up of DBP in the organic. Plutonium solids lodged in the CX column in the packing, caused by an acid imbalance that caused precipitation of plutonium fluorides and carbonates. Golf-ball sized solids, believed to PuF3 precipitated in filter and plugged outlet. Laboratory extraction efforts indicated that the plutonium being lost was not extractable and that it was, for the most part, not particulate. High plutonium loss in CAW due to ash run Build up of organic degradation, dibutyl and monobutyl phosphate affected process efficiency in the PRF SX. Led to high Pu loadings in organic and resulted in precipitation of PuF4. PRF tested a high Pu processing rate using only dissolved PuO2, the CAW aqueous had unusually high Pu concentration. Appeared to be insufficient stages in the CA column. Continuous dissolver flowsheet; 91% Pu dissolved in 9 M HNO3 with 0.16 to 0.47 M NaF. Increasing fluoride to 0.6M resulted in precipitation of plutonium fluoride. Decreasing fluoride to 0.07 to 0.2 resulted in only 82% dissolution. Dark green floating solids in CCP were identified as a silica-organophosphate containing plutonium.

May 1968 May 1968 September 1969 November 1969 October 1969

May 1971 July 1971

Official Use Only 92

RPP-RPT-50941, Rev. 0

4.10.2 Use of Filtration and Centrifuges to Limit Loss from PFP Aqueous Waste Stream This section details the history of filters and centrifuges in PFP, so that an analysis of their effectiveness can be determined with respect to solids losses. During the period 1956 to 1966, the inventory difference at the 234-5 Building was estimated to be about 600 kg Pu, (RHO-CD-194, A Study of the 234-5 Building Inventory Difference for the Years 1956 through 1966). The inventory difference (ID), also referred to as the material unaccounted for (MUF) and earlier as the book-physical inventory difference (B-PID) is the algebraic difference between the book and physical inventory quantities. Positive numbers indicate that the physical inventory is less than that reported in the records. Namely material has been lost. The resulting investigation, suggested about 100 kg of the Pu ID could be attributed to understatement of liquid discharge transfers. This was attributed to the liquid waste not being adequately sampled or analyzed. It was concluded based on studies and analyses from samples of cribs that received 234-5 Building plutonium containing liquid wastes that there could be three times more plutonium lost to liquid waste than the measured discards and accounted-for discards indicated. It was generally recognized there were solid plutonium particles in the waste solutions. In 1967 it was realized that significant plutonium was being lost as unmeasured solids in the PRF liquid waste stream. Analytical methods at the time did not account for any solids that might be present (ISO-659 RD, Z Plant Report Task I-II & Recovery Operations) and ARH-423-RD-DEL). The new method (E6-B AEA) reported 3 to 10 times more plutonium the previous method (PuA-20A) for duplicate sample analysis. Changes were made in the method of waste sample analysis. Process changes were also made to capture solids in the extraction column aqueous waste stream (CAW) by use of an interface crud filter and ultimately to use of centrifuges to remove the particulates. In July 1967, a CAW crud filter with 25 micron rating was installed to capture accumulated interface crud. In October crud filter use was stated as being intermittent and mentioned that 4.8 grams of Pu was captured in 3 days. The filter rating was increased to 40 micron and discontinued shortly afterwards because of flow problems. In September the CAW centrifuge was received and cold testing began. In ARH-108, Safety Analysis Report and Hazards Review CAW Centrifuge it is stated, Change in methods of analyzing waste samples indicated 3 to 10 times the amount of Pu as reported with the previous analysis. This is attributed to solids not removed by the extraction process. Removal by filtration has not proven satisfactory. Lab tests (no reference is made) show solids have average size of 3 micron and maximum size of 12 micron can be removed by centrifugation. Under normal conditions CAW solids are less than 2 gms/cc. The centrifuge was a Westfalia Centrifuge, or Centrico Centrifuge, Model SA OOH 205. This was a sophisticated unit capable of continuous de-sludging. Specifications in the PFP historical files stated; Capacity 80 gpm 8000 rpm

Official Use Only 93

RPP-RPT-50941, Rev. 0

Sediment specific gravity 23 500 bowl volume. Rated up to 5500 Gs. Solids removal rate is assumed to be constant.

A sketch and pictures of Wesfailia centrifuge picture are given in Figures 39, 40, and 41 and the installation drawings are H-2-26271, Centrifuge Installation Pulse Column Hood 6th floor, dated 10/6/1967. Figure 38. Wesfalia Centrifuge Sketch

Figure 39. Photo of Similar Centrifuge

Official Use Only 94

RPP-RPT-50941, Rev. 0

Figure 40. CAW Centrifuge in 6th Floor

During the first 2 days on operation in October 1967 it was reported 13 gms of Pu bearing solids were removed and observation made of a large volume of Pu bearing solids. In February 1968, the monthly reports state the centrifuge was removing approximately 0.6 g of Pu/hr. They reported that there were about 60 g of air-dried solids per liter of composited sludge. Three to five percent of the dried solids were plutonium (ARH-423-RD-DEL, Monthly Reports 1968 Plutonium Process Engineering, LM Knights). The early Westfalia centrifuge operation was plagued with mechanical and operability issues. Frequently the unit would fail to reseat after de-sludging and continuously discharged to the floor. Frequent dismantlement and cleanings were done and large accumulations of solids were removed from the internals. Operability through 1968 was probably less than 50%. After vendor assistance and modifications, operability was improved to 8090% in 1969. A detailed review of biweekly and monthly reports is provided in Appendix H of this document. Discussion of replacement units are mentioned in the reports. An April 1, 1969 letter, L.M. Knights to W.P. Ingalls, Solution Clarification Hood the Plutonium Reclamation Facility, discussed requirements for 1972 budget item for feed and/or waste stream clarification. Justifications stated as: Eliminate build-up of Pu bearing solids in processing equipment Eliminate unmeasured plutonium losses, and Permit better control of liquid waste losses.

The need date was stated as after installation of continuous waste treatment in 242-Z but before waste management facilities are modified (i.e., the cessation of crib discharge for HLW and

Official Use Only 95

RPP-RPT-50941, Rev. 0

LSW). The stated expectations are of little Pu in the solids (1 gram/liter), that they are refractory in nature and difficult to dissolve. Design parameters are stated as: 10 liters of solids per week Centrifuge should be capable of 1000 to 1500 Gs Minimum filter porosity of 5 microns 40 micron filter suitable after they are coated with solids Solids both gelatinous and crystalline

In 1971 development work on improved dissolution and solids liquid clarification was reported (ARH-2077, Continuous Dissolver Development Program Progress Report. This included tests on an 8 diameter bowl centrifuge (Barrett Company brand) and dissolver overflow runs on ash, slag and crucible and other feeds. Better than 95% of the solids were removed by the centrifuge. Sock filters were less effective, having flow problems and metal frit filters and hydroclones were dismissed as ineffective. In 1972, it became clear that significant solids were still being lost to waste and also present thought-out the PRF. Justification was made to the AEC for 6 additional centrifuges to clarify feed, waste, and product streams (Letter ATA-5G, R.P. Corlew to O.J. Elgert, Prototype Solids Handling System for Z Plant Process). Justification was cited as Almost all dissolved and /or leached plutonium scrap or rework solutions contain undissolved solids and alleviation of processing problems as plugged lines, pump abrasion, SX column instability, impure product, and difficulties in accountability. This was also cited in a safeguards survey done by AEC in which they comment that the Material Un-accounted For (MUF) were out-of-control for several reporting periods (7 of 12 months) and it was attributed to unrecognized loss from the presence of solids in a supposedly liquid system (SSSM-1797, Survey No. 41, Facilities HVA and HVB). Additional process design information can be found in ARH-2078, Prototype Solids Handling Systems for the Plutonium Finishing Plant. The original CAW centrifuge helped reduce waste loss but problems continued with plugged lines, column stability and product purity. Little performance information is presented but the report indicates past efforts to use screens and filters (100 micron size rating) were not successful. Testing is recommended as no data is available on operational characteristics. The recommendation is made for a belt driven unit made by the Barrett Company, with design emphasis placed on critically safe geometry. ARH-2078, Addendum, Prototype Solids Handling System for Plutonium Finishing P.C. Ely and J. E. Hammelman, March 14, 1972. Configuration changes were made with a switch to direct drive centrifuges made by the Western States Company. The centrifuges selected were Western States Model superfuges selected primarily because they were geometrically safe in regard to criticality. The Western States vendor file gives Superfuge model Vendor drawing 032-1020-00-4, 8 diameter bowl, 3450 rpm, Hp. The Official Use Only 96

RPP-RPT-50941, Rev. 0

separation efficiency for these units, rated at 1000Gs is reduced over that of the Westfalia unit, 5500 Gs, but their simplicity in design and operations was preferred. This project added 6 centrifuges throughout the PRF designated as: A for Continuous Slag & Crucible dissolver overflow B1 and B2 for PRF Extraction Column Aqueous waste (CAW) C for Pot-type scrap oxide dissolvers in Miscellaneous Treatment (MT) D for Product solutions E for EIW (242-Z waste to TK-D5)

The relevant drawings are for the Prototype Solids Handling System (PSHS) are: H-2-26840, General Layout and Master Drawing List H-2-26841, sheet 1, PFD PSHS Dissolver Centrifuge H-2-26841, sheet 2, PFD PHHS CAW Centrifuge H-2-26841, sheet 3, PFD PSHS MT Centrifuge H-2-26841, sheet 4, PFD PSHS CCP and EIW centrifuges H-2-26842, sheet 1, EFD PSHS Dissolver Centrifuge H-2-26842, sheet 2, EFD PHHS CAW Centrifuge H-2-26842, sheet 3, EFD PSHS MT Centrifuge H-2-26842, sheet 4, EFD PSHS CCP and EIW centrifuges

A picture and vendor sketch of the Western States centrifuge unit is shown below in Figure 41.

Official Use Only 97

RPP-RPT-50941, Rev. 0

Figure 41. Western States Centrifuges

B1 and B2 Centrifuge in MT-6

Western States Superfuge Drawing

Flowsheet data used in design of the PSHS is presented in Table 19. Little information on particle size is presented. The presence of insoluble and undissolved PuO2 in PRF streams was recognized during the design of the PSHS. Table 19. Flow Sheet Data for Prototype Solids Handling System
Centrifuge Service Average Particle Size Specific Gravity Totals A SS&C Dissolver >0.5 1.3 1.4 17, 480 gms moist cake containing 8,400 gms ash, 256 gms PuO2 (22 hrs operations of 2 dissolvers) 14% of charged PuO2 not leachable 1.3 1.4 1.8 1.3 1.4 1.3 1.4 B1/B2 CAW C MT Dissolver D Product E E1W (242-Z, Waste Treatment)

1104 gms wet 6.24 kg moist cake 522 gms wet 960 gms wet solids solids containing with 3.12 kg PuO2 solids containing 48 gms Pu per 78 gms Pu 24 hrs 2.2 gms insoluble Pu/hr 1% of un-dissolved PuO2 in centrifuge overflow

Other notes

Official Use Only 98

RPP-RPT-50941, Rev. 0

An August 16, 1972 Work Authorization CP-1439, Atlantic Richfield Hanford Company (ARHCO) to JA Jones gave direction to install the B-1, B-2, C, D, and E centrifuges and support equipment. The A centrifuge was to be installed later. The centrifuges began operation in May 1973, the same month PFP high and low salt waste was routed to Tank Farms. Operability issues were apparent from the start, especially with bearing problems and outlet flow restrictions. A review of the initial operability of the PSHS centrifuges is found in letters in the PFP engineering files and a review of monthly reports from the period. The review of monthly reports is included in the Appendix H. Table 20. E1W centrifuge Sludge Characterization June 1973

Official Use Only 99

RPP-RPT-50941, Rev. 0

During the second month (June 1973) it was reported that 940 gms and 704 gms of Pu were captured by these centrifuges. Solids from EIW centrifuge were characterized and results are reported in Table 20. There were solids present in the waste effluent from 242-Z. It is unclear whether the CAW centrifuge, which should have clarified the feed to the 242-Z process, was operable at the time (ARHCO letter, P.C. Ely to M.J. Klem, Centrifuge Solids, June 19, 1973). The letter stated the particle density was obtained by measuring the volume of water displaced and the particle size is calculated from the settling rate. Large, slow settling particles were present. Problems with early PSHS centrifuge operation are evident in letters from the PFP engineering files. June 22, 1973 Letter, Status of Centrifuge System R.D. Fox to D.L. Merrick et.al. Centrifuge A, line modification not initiated Centrifuge B-1 and B-2, Discharge line plugs, centrifuges overflow Centrifuge C, High pressure drop in discharge line causes leakage Centrifuge D, Operational Centrifuge E, excessive equipment failure reduces usage July 16, 1973 Letter Prototype Centrifuge Shaft Bearings, M.J. Klem to R.E. Felt. Bearings fail rapidly, life extended to 5 weeks with use of purge air. Other changes recommended. July 30, 1973 letter, R.E. Felt to C.M. Peabody, Centrifuge Operation makes recommendations made on various PSHS centrifuges: A centrifuge Never operated and unit badly corroded. Move unit downstream of solids collections tank 53. C Centrifuge Unit breaks down frequently, requiring use of filter for transfers. Recommend using one of the B centrifuge as a spare. (B centrifuges apparently not used for CAW) D Centrifuge Discrepancies about operations, Recommendation to recirculate solution through unit prior to sampling. E centrifuge Used on solids from EIW (waste treatment), Observations that DW-5 disengager downstream fills with solids. Solids small enough to escape CAW centrifuge will surely escape the E centrifuge. October 17, 1973 ARCO letter D.A. Turner to D.T. Crawley, Results of Flow restriction test. Describe results of test designed to reduce C centrifuge overflow problems. Dec 10, 1973 letter, D.A. Turner to W.H. Koontz, Design request Hood 6 Centrifuge Discharge Piping. Recommends changes to B and C centrifuges, cites performance as unsatisfactory.

Official Use Only 100

RPP-RPT-50941, Rev. 0

Specific data on the remainder of centrifuge operation from 1974 until PRF shutdown in 1976 could not be located. A discussion of solids in waste effluent was located in 1975 (ARH-LD-206, ARCO Monthly ReportJune 1975), which indicate the subject solids were not removed by the centrifuge. Floating solids in the plutonium recovery extraction column feed resulted in high waste losses from both the plutonium recovery and the waste treatment extraction columns. The floating solids which were not removed by the in-line centrifuge were attributed to the processing of a small amount of ash in the miscellaneous treatment dissolvers. Mention of floating plutonium occurs frequently in the history of PFP recovery operations and it is not always tied to ash processing. It is more often attributed to plutonium complexed with organic or particulate bound in an agglomeration of degraded organic. A detailed analysis of some of these floating solids was done in 1988 by PFP research chemists. (WHC Letter 12221-PSL88-058, Analysis of Suspended Solids in Solutions from Tanks BT-1, 2-3, G. S. Barney, 4/13/1988). He identified large particles, rods 10 by 100 microns and rounded particles up to 20 microns. They contained measureable but little Pu (estimated as ~2/gms/liter of solution) and required aggressive methods for dissolution. It was concluded they contained degraded solvent, silicate and plutonium. In the late 1970s PRF was planned for a restart to do a cleanout run and correspondence was located describing centrifuge requirements. Nearly continuous CAW centrifuge operation would become a requirement to solvent extraction operation and efforts were made to improve operability of both waste centrifuge units. After efforts to improve operability of the Westfalia unit were exhausted, it appears that parallel Western States units (B1 and B2) were put into service on the CAW stream. The following is a chronological history of these efforts: April 20 1977 Memo, DTC to DGH CAW Centrifuge and CAW Monitor. The B-1 and B-2 centrifuges were salvaged to get EIW, CCP, S&C and Hd 6 feed centrifuges operating, since the 6th floor (Westfalia) unit was operating satisfactorily. EIW and S&C centrifuges not operating and could be backup for CAW, but hydraulic problems in Hd 6 have not been fixed from Dec 1973 request. RHO Letter 60413-78-101, D.W. Reberger to D.G. Harlow, Revision to Process Specification Numbers 10 and 8, June 12, 1978. This letter made CAW centrifuge operation a requirement to run solvent extraction. The centrifuge could not be bypassed for greater than hour in a given 8 hour shift.

Official Use Only 101

RPP-RPT-50941, Rev. 0

Letter 60413-78-115, CAW Centrifuge, D.T. Crawley to R.J. Thomas, June 28, 1978. This letter provided information for replacement centrifuge, good information is not available. Cites references for 50 gms/hr removal rate and average size of 3 with maximum of size of 12 . Current CAW centrifuge rated at 5000 Gs, and replacement rated at 1000 Gs. Crawley also stated the two Western States units were installed on the CAW stream, but never used. He does state that the EIW centrifuge was used routinely. June 28, 1978 letter, K.H. Oma to R.J. Thomas, CAW Centrifuge Installation in Hood 6 of Miscellaneous Treatment. This letter recommended the configuration of routing CAW from the Column through the two centrifuges in MT-6 and into a small CAW pump tank. This would become the configuration used throughout the rest of PRF operation, into the 1980s. July 10, 1978 RHO letter, D.W. Reberger to P.A. Miskimin, CAW Centrifuge Problems. This letter documented the results of two days meeting with Westfalia vendor to improve operability and maintenance on centrifuge. Slugging of interface crud was identified as a major problem. The unit is described as too sophisticated. August 4, 1978 letter, D.W. Reberger to G.C. Oberg, Backup CAW Centrifuge Installation. Mechanical problems with the Westfalia centrifuge have resulted in extended shutdowns of solvent extraction. Requested analysis to use previously installed B-1 and B-2 centrifuges as backup. Comments were made that these units were installed and never operated. August 10, 1978 letter, D.W. Reberger to C.L. Owen, CAW Centrifuge Flushing. Changes are recommended to desludging cycle to improve operation and bowl re-seating.

PRF was idle from the end of the cleanout run in 1979 until it was restarted in 1984. During PRF operation in the 1980s, centrifuges were required on the CAW stream (B1 and B2 units) and the MT dissolver stream (C centrifuge). In addition, prior to the centrifuge, the MT solutions were first allowed to settle and filtered using a pan filter, shown in Figure 42. The Westfalia unit was abandoned and no longer used. All centrifuges used in the 1980s were the Western States superfuge units. Changes were made to improve the change out of the gearbox which was the major cause of downtime. Centrifuge failure was attributed to PRF and MT system downtime as documented in Campaign Reports.

Official Use Only 102

RPP-RPT-50941, Rev. 0

Figure 42. Dissolver Pan Filter in MT-5 for Dissolver Solutions

(25 m poly propylene cloth placed inside pan shown)

Solids in the waste stream and MT dissolver stream continued to be problem. These caused high waste losses and contributed to accountability issues. Significant changes to dissolution were not made until 1987. No report was available for 1984 PRF operations, but a review of the reports for 1985, 1986, and 1987 include the following statements: 1985 Operations High waste losses due to Pu-Bearing fines passing through centrifuge and into the CAW. Ineffective solids removal in Miscellaneous Treatment. The pan Filter and C centrifuge are unable to remove fines. Lab studies are recommended to improve dissolution. 45 charges were lost in MT due to C centrifuge failure, many maintenance problems noted, use of settling and pan filters helped. 44 hours of SX down time attributed to B1/B2 centrifuge failure. Approximately 69 hours downtime attributed to high solids content in 241ZTK-D5. High solids seen in SX feed from MT dissolver solution. Total alpha and extraction methods for analysis of CAW disagree by over 100% in some cases. After filtering the CAW through millipore filters the discrepancy between methods is normal.

Official Use Only 103

RPP-RPT-50941, Rev. 0

1986 Operations Although the B and C centrifuges did not cause major downtime, they failed numerous times during the campaign. Most failures were due to gearbox failure (evidence the unit continue to run submerged with restricted outlet flows). 16 hours of downtime in SX attributed to centrifuge failure. Centrifuge overloaded with solids, excess solids plugging storage tanks. C centrifuge failure caused of 22.5% of downtime in MT, rodding lines plugged with solids attributed 9%. Flush water used to dilute solids in Tank Farms Receiver Tank. 21,090 gallons of flush water used in 1986, compared to 28,700 gallons in 1985. In 1986, 3 batches of waste exceeded specification for solids content, versus 22 batches in 1985. Recommendations made for improved dissolution and solids separation in the future.

1987 Operations 2% of MT downtime attributed to C centrifuge failure, no downtime in SX attributed to B1/B2 centrifuge. Changes in MT dissolution have greater reduced recycle and solids content. Recycle rate in 1987 was 13% versus 70% in prior campaigns. CC column (Pu stripping) experienced unstable interface attributed to solids and crud formation. Crud drop released 15 liters of slurry, dark brown and high in solids.

A study in 1986 examined plutonium oxide scrap dissolution and PRF and other DOE sites in which recommendations are made for improved performance. (SD-CP-ES-096, Recommendations for Improved Dissolution of Plutonium Oxide Scrap at the Plutonium Reclamation Facility.) The mid 1980s oxide dissolution system is shown in Figure 43. It was estimated the 40% of the solids in the centrifuge overflow are <0.2 micron. It was recognized that the most significant problem at PRF was the solid-liquid separation and that similar problems were seen at other sites.

Official Use Only 104

RPP-RPT-50941, Rev. 0

Figure 43. Mid 1980s MT Dissolver System

Laboratory experiments dealing with the transuranic extraction (TRUEX) process were conducted during the 1980s on centrifuged PRF CAW solutions. Small amount of Pu-bearing solids were present in these samples and had to be removed by ultra filtration. Summary results from these reports are presented in Table 21 and discussed below. SD-CP-DTR-006, Plutonium Recovery and TRU removal from PRF CAW, C.H. Delegard, October 1984. The report analyzed centrifuged PRF CAW from 1984 operations for residual TRU content. The original solution had a reported <1% solids but was cloudy and showed solids when left undisturbed. Analysis found that 15% of the activity (mostly Pu) was removal by ultra filtration, and therefore particulate. SD-CP-TRP-016, TRUEX Solvent Extraction and Solution Clarification Results for FY1986 PFP Waste, Naser et.al. This document continued testing of PFP effluent samples. The solids in the PRF CAW sample from 1984 were analyzed along with RMC scrubber solution. Only solids were found in the PRF samples. Analysis showed the solids were mostly Mistron, a mineral additive used in the PRF process. About 0.1 wt% of the sample was Pu oxide. A rough particle size distribution is reported with 22 volume % of the particles greater than 9 micron. Official Use Only 105

RPP-RPT-50941, Rev. 0

SD-WM-DTR-020, TRUEX Process Demonstration Tests with Plutonium Finishing Plant Waste, A. J. Naser, August 1988. This document reported on testing of two samples of centrifuged PRF CAW from the 1987 operation. Both samples had low solids content, 0.11 and 0.028% by volume. Mean particle size was 17.6 and 16.4 micron, with a wide distribution. The first sample contained rods (20 to 100 micron) and platelets (5 to 40 micron) with smaller, angular particle (0.5 to 2 micron). The second sample has rods (10 to 50 micron) and small angular particles (<1 to 10 micron). These are reported to be oxalate crystals, silicates and Mistron. The nature of the plutonium in the samples is not discussed, but significant TRU content is removed by ultra filtration. Table 21. Results on Centrifuged PRF CAW from 1984 and 1987

Official Use Only 106

RPP-RPT-50941, Rev. 0

Centrifuge Sludge Recovery The sludges recovered from the centrifuges were assayed and usually returned to dissolution for recovery if greater than 4 gms/liter Pu. Lean sludges were discarded as solid waste. Studies on recovery were examined to provide insight on the amount of high fired and refractory oxide that might be present. This is complicated by the fact that some unrecovered Pu in sludge may be present as soluble Pu complexed with organic degradation products and as particulate trapped in an organic matrix. Centrifuge sludges from dissolution were simply recycled and sludges from CAW centrifuge were sometimes fired or burned to remove and destroy organic prior to recycle. In March 1969, burning and leaching tests were done on CAW centrifuge sludge (ARH-1048 RD, Z Plant Report Task I-II Recovery Operations). Pu values were not reported but 15% of the material was lost by burning at 700C for 1 hour and 31% was lost by leaching in B-Acid (concentrated nitric-HF) for hour. Battelle Northwest Laboratory experiments with caustic-peroxide fusion for Pu recovery from centrifuge sludge were conducted in 1975 (BNWL-B-419, Plutonium Recovery from Incinerator Ash and Centrifuge Sludge by Peroxide Fusion). When the sludge was fired to 800C for one hour to destroy organics, 23% weight loss was seen. Pu recovery using the fusion process ranged from 63 to 95%. The leaching of EIW centrifuge sludge reported previously in Table 20 showed that between 20 and 61% of the Pu was leached using aggressive methods. In the 1980s, the PFP development lab performed dissolution studies on Rocky Flats oxides, dissolver heels and centrifuge sludge from dissolution (internal letters 65454-85-023, Recommendations for Dissolution of Rocky Flats Oxide and Resulting Centrifuge Sludge, 65454-85-032, Chemical Analysis of Rocky Flats Oxide Sludge and Heel). These were performed in part in response to address excessive waste due to Pu-bearing fines materials passing through the centrifuge and into the CAW. This testing showed about 30% of the sludge dissolves in B acid with NaF to boost fluoride, and the Pu content in the sludge was about 20 wt%. The sludge and sludge heel was shown by X-ray diffraction to contain PuO2. Other tests shown in Figure 44 for centrifuge sludge leaching studies tests showed 25 to 50% of the Pu was leached using aggressive leachants, B acid with additional fluoride as NaF or HF. 4.10.3 Conclusions Regarding Centrifuge Operation In 1967, PFP realized that significant plutonium was being lost as unmeasured particulate in the PRF CAW liquid waste stream. Initial efforts to capture this particulate were centered around use of a Westfalia centrifuge on the CAW stream. About 0.5 to 1 gram/Pu per hour was being captured. Operability of the centrifuge was poor for the first several years but improved marginally in the late 1970s. Operability was a challenge due to the sophisticated design of the Westfalia centrifuge. Official Use Only 107

RPP-RPT-50941, Rev. 0

In 1971, MUF problems associated with unmeasured and uncaptured particulate were still occurring, so an extensive installation of new centrifuges was initiated on multiple PRF streams, waste, dissolvers, and product. These new centrifuges (Western States) were less sophisticated but not capable of the same separation efficiency as the prior unit. Initial operation of these units was also plagued by maintenance problems. The CAW centrifuge was essentially not operated until 1978, when they replaced the troublesome Westfaila centrifuge. CAW centrifuge operation was not requirement for solvent extraction operation until 1978. It was required throughout the PRF campaigns in the 1980s. The Western States unit ran with restricted outlets that caused the bowl to be submerged and some fraction of the solution to be bypassed. This may have been as high as 20%. The design basis for all centrifuges assumed a 12 micron maximum particle size and 3 micron average. Limited analysis is available on centrifuge sludge but larger particles are routinely found in the centrifuge sludge, up to 40 micron in some analyses. Pu solids were detected downstream of CAW centrifuges as solids buildup in W3 and W4 in 242-Z. There are periodic reports of floating plutonium solids that are not removed by centrifugation. This represents plutonium potentially in particulate form (captured in organic crud) that escaped the centrifuge. The last periods of operation of PRF in the 1980s still saw the presence of Pu fines that contributed to high waste losses and discrepancies in waste analysis. Particles > 10 micron were found in centrifuged PRF CAW solutions in the 1980s. These were typically lean in Pu, about 1-2%. Burning studies of centrifuge sludge indicated a large fraction of organic material in the range of 1525%. Leaching studies on sludge with aggressive leach solutions showed a large amount of plutonium was not leached, typically less than half, and this could be indicative of high-fired refractory oxide material. Overall, it can be concluded that some progress in plutonium solids capture was realized when the centrifuges were operated. The operation was limited and intermittent at first, and became more widespread and reliable later. Nonetheless, centrifuge design and operation was not perfect, operability problems persisted and particulate plutonium discharge continued.

Official Use Only 108

RPP-RPT-50941, Rev. 0

Particulate losses, primarily unmeasured, that occurred during early periods of waste discharge to cribs without centrifugation were on the order of 3.5 times the measured, accountable loss. With centrifuge operation, this unmeasured particulate loss is assumed to be reduced to about equal the measured and accountable loss. This is discussed in greater detail in section 4.13 when estimate of total plutonium discharged to tank farms is presented. Figure 44. Leaching test on MT Dissolver Centrifuge Sludge

Official Use Only 109

RPP-RPT-50941, Rev. 0

4.10.4 Plutonium Tetrafluoride Carry-Over and Impact in Tank Farms Plutonium Tetrafluoride (PuF4) would have gotten into the liquid waste streams from the Plutonium Finishing Plant (PFP) by two routes: entrained fluoride particulate in the metal line fluorinator off-gas system and entrained fluoride particulate in PRF waste, which could come from the PuF4 precipitation in dissolver solution or from precipitation in the columns. The fluorinator-off gas scrubber was historically a water-jet used to provide vacuum and an aqueous scrub of the off gas from the plutonium fluorinator to remove HF in the gas stream. This aqueous waste stream was estimated to contribute about 100 grams plutonium/year to the waste. The off-gas from the hydrofluorinator was filtered through carbon filters to reduce particulate entrainment. The filters had an average pore rating of 69 microns and minimum particle retention size of 25 microns. Plant personnel indicated that about 2 grams of PuF4 per week escaped the carbon filters and that the filters retained about 2500 gms PuF4 per week (ARH-2597, Safety Analysis Plutonium Effluent Cleanup, Z Plant). This aqueous waste stream was routed to 241-Z and neutralized with soda ash or caustic before routing to the 361-Z Tank and the cribs. PuF4 is soluble in concentrated base solutions. In 241-Z tanks the waste was adjusted to pH 2 and treated with ammonium nitrate (ANN) to convert any solid PuF4 to the soluble Pu(IV) ion. In the absence of aluminum ion the hydrolysis reaction is slow. There is evidence that PuF4 could have survived and reached the cribs. When the 241-Z waste solutions were transferred to the waste tanks after 1973, the off gas scrubber was converted to a re-circulating system using aqueous ANN and later an aqueous KOH. The PuF4 should have all been dissolved during scrubbing of the solution before transfer to the 241-Z Building. In the PRF, scrap PuF4 that was routinely recovered from the glovebox floors and the glovebox filters was recycled as scrap to the dissolvers. If fluoride was known to be present in these powders, ANN was added to the dissolver to facilitate dissolution. In the presence of aluminum ion the dissolution of PuF4 should be complete. However there has been mention in the Z Plant weekly reports of undissolved PuF4 solids entrained in the PRF solvent extraction column solutions, including mention of golf-ball sized solids plugging filters (ARH-294 RD DEL, Z Plant Report Task I-II & Recovery Operations, May 20, 1968. Pu fluoride was also known to precipitate in the dissolver pots when other scrap forms were dissolved, due to high plutonium concentration and presence of fluoride added to aid in dissolution. ANN was added to the dissolver solutions to complex the free fluoride prior to introduction into solvent extraction. Pu fluoride would also continuously precipitate in the CO column, used for stripping residual plutonium from the organic, as dilute nitric/HF was used as the extractant. The column would be shutdown weekly, filled with ANN and air-sparged for the weekend to re-dissolve the precipitated fluoride. Although methods were employed to minimize the formation of precipitated PuF4 and re-dissolve precipitated materials, it is likely that some PuF4 was also entrained in the waste stream from PRF. During 19731976, subsequent processing through the 242-Z Waste Official Use Only 110

RPP-RPT-50941, Rev. 0

Treatment Facility should have further reduced any PuF4 solids in the final waste stream. The reaction rates would have been slow, but any PuF4 should have been hydrolyzed in the acid waste stream. After neutralization in TK D-5 was began, any PuF4 solids would have been subject to alkaline hydrolysis as described in this document (Appendix C, Report 2, Metathesis of Plutonium(IV) Oxalate and Plutonium(III)/(IV) Fluorides in Alkaline Media to Form PuO2xH2O).

4.11

OVERVIEW OF PLUTONIUM ACCOUNTABILITY METHODS

Accountability Measurement Methods in Liquid Waste Streams This section examines the history of laboratory analysis for Pu in waste streams and details the uncertainties. Plutonium Analysis History The PUREX and PFP laboratories no longer exist and the REDOX Laboratory shifted away from a plutonium separations process control facility decades ago. To a large extent, records of analytical methods capabilities have vanished with the laboratories. A partial history of methods used for plutonium accountability has been assembled from the records that could be located but these assembled records likely are incomplete. This review examines plutonium analysis accountability in liquid waste streams with emphasis on the sources of under accounting for plutonium in solids. Throughout the Hanford plutonium processing history and waste accumulation, examples of plutonium in unmeasured (or incompletely measured) solids have been identified and methods to better characterize the waste have been identified. The quantity of plutonium in fuel was determined by calculation of the type of fuel and the operating conditions in the nuclear reactor where the plutonium was born. It is noted that plutonium was a very small mass fraction of the feed to the dissolvers for enriched and natural uranium production reactor fuels. In contrast, the mixed oxide fuels (e.g., PRTR) would have contained vastly more plutonium with respect to uranium. The first step in plutonium separations was to remove cladding from fuel elements. The cladding removal waste (CRW) was the first liquid waste stream and a potential source of loss of plutonium in solids. An extensive investigation into CRW solids as a source of MUF was conducted in 1967. After the cladding was removed, the fuel, consisting primarily of uranium, fission products and plutonium, was dissolved. The dissolver solution had very high beta-gamma activity due to the fission products. The uranium was determined by using the solution density and a conversion factor. Plutonium concentration was based on the ratio factor from the production reactors. From a mass balance standpoint, the largest fraction of material in the dissolver was uranium while the fission products were the primary contributors to the radioactivity of the dissolver solution. Periodically this factor was adjusted to better align the calculated values with the observed concentration from either special analytical investigation or to better balance the inputs Official Use Only 111

RPP-RPT-50941, Rev. 0

with the plant outputs. In the separations process, the fission products were removed and went to tank waste or another facility for additional separations. The plutonium and recovered uranium from the solvent extraction processes were relatively pure materials. Free of interferences, plutonium and uranium could be measured by very accurate and precise accountability methods. Specific examples of plutonium analytical methods are provided in the following paragraphs. In 1949, HW-14744, Tentative Analytical Methods and Estimated Samples for Process Control of Redox Production Plant No. 1, plutonium is measured by various lanthanum fluoride carrier precipitation methods followed by alpha counting using an alpha proportional counter. The proportionality factor is a function of two variables, the specific activity (disintegrations/unit time/unit weight) of the alpha emitter, and the counting efficiency or geometry of the counting instrument. As described in HW-59434 J, Chemical Processing Department Research and Engineering Operation Monthly Report February, 1959, plutonium was measured by direct X-ray absorption and uranium by density. These were considered routine and nonspecific methods for process control. The more precise independent and specific methods were to measure separated uranium and plutonium by controlled potential coulometry. The uranium density method was biased high by 0.6% compared to the test method but plutonium in the final product agreed with 0.15%. However, REDOX dissolver solution containing coating waste for processing I&E and cold slugs had a bias that was greater by a factor of two to three. The more specific method of measuring uranium involved first performing a TBP extraction and then measuring the X-ray absorption. It was mentioned that similar studies were being conducted on PUREX solutions. Because the amount of plutonium was determined by ratio to uranium, improvements in the uranium measurement had a corresponding improvement in the plutonium value assigned to dissolver solutions. By this time, the alpha scintillation method for plutonium had been improved and was considered ready to be tested on process samples. PUREX and REDOX monthly input values were examined in RL-CAO-165, Historical Review Plutonium Ratio Receipt Measurement July 1960 through December 1964 according to the following assay: Pu (ratio determined) = U (reactor stated) AT/U (dissolver solution assays) x Pu/AT (product solution assays) where AT indicates alpha total values for plutonium. The reactor-stated values were based on uranium core weights adjusted for reactor fuel burnout. Chemical Processing Division dissolver and product solution values were corrected for analytical bias and the AT assays on dissolver solutions were corrected for counts attributed to uranium, americium, and curium. The Pu/AT assays were needed to convert the AT counts in the dissolver to gram quantities of plutonium. It was stated that prior to 1958, dissolver solutions were measure by standard volume-sample assay methods with reliability of 6% on a monthly basis and 2% on a yearly basis. In 1959, the ratio method was instituted so that the dissolver volume was no longer needed. Analytical assays

Official Use Only 112

RPP-RPT-50941, Rev. 0

were taken on monthly composites at PUREX and on individual dissolver solution batches at REDOX. PUREX had monthly uncertainties of 2.7% while REDOX had uncertainty of 1.7%. According to HW-72283 DEL, Chemical Processing Nuclear Materials Control, in January 1962, PUREX reported that a new plutonium standard was put into use at all three control laboratories. Statement was made that loadout measurements at PUREX and REDOX continue to show improvements. The standard was used for both plutonium X-ray and alpha determinations. The D-5 density measurement for uranium and plutonium by ratio to uranium was used because it was available and convenient. In ARH-1904 DEL, 200 Areas Operation Monthly ReportApril 1971, investigation into the measurement of plutonium in the plutonium reclamation process. Poor analytical precision was reported. Analytical Techniques As shown above, early analyses of plutonium in solution streams relied on separation and alpha counting. Such techniques require complete dissolution of the plutonium-bearing materials so that the chemical separations could occur. Gamma spectrometry methods instituted in the 1980s helped overcome some of the problems attendant to under-reporting of plutonium in waste and other process streams. The following paragraphs describe the alpha and gamma counting approaches. Alpha Counting For samples containing a pure radionuclide, a constant value for the specific activity may be assumed. Plutonium, however, is normally produced as a mixture of alpha-emitting isotopes with widely varying half-lives, and as a result the specific activity may vary from batch to batch or from one source of plutonium to another. Consequently, when total alpha measurements were used for plutonium assays, an experimental specific activity factor (the F factor, grams plutonium/alpha count/minute) was used. A given F factor, which may apply for a single batch or for all plutonium produced over a period of time, is determined by directly relating the plutonium chemical analysis of a sample to its alpha counting rate. It should be noted that total alpha counting instruments are blind in that they cannot distinguish between alpha particles from different sources. Therefore, contamination of the plutonium by other alpha emitters such as americium or curium would contribute to the plutonium signal. The isotopic composition of the plutonium would also be altered by its exposure history in the reactor and thus would also affect the specific activity of the plutonium. Consequently, use of alpha counting results in radioassay is valid only for solutions of known alpha composition achieved by scrupulous purification of the plutonium and knowledge of the plutonium isotopic distribution. The total alpha activity of aqueous solution was determined by transferring the sample onto a 20 mm stainless steel or platinum counting disc, drying, flaming, and measuring the alpha activity with a standard alpha counting instrument. A similar process was used to determine the Official Use Only 113

RPP-RPT-50941, Rev. 0

total alpha activity of organic solutions. In this case, the organic solution was evaporated on a stainless steel disc which is heated around the edges in such a manner that a temperature gradient exists between the edge and the center of the disc. This allowed the solvent to evaporate more rapidly from the outer edge and therefore retained the liquid on the flat counting disc. The heater was usually maintained 40-60 degrees C above the boiling point of the solvent. The method was applicable to the evaporation of organic solutions such as hexone (methyl isobutyl ketone), TBP (tributyl phosphate), and TTA (Thenoyltrifluoracetone). The total alpha activity of solid-liquid slurries presents challenges in that absorption (and consequent under-reporting) could occur because of the high absorption of the alpha particles by the residual solids. The absorption by the solids was minimized by uniformly spreading the solid-liquid slurry over the maximum disc surface area. The slurry was then dried, flamed, and counted in the normal manner. Issues attendant to the determination of plutonium in solidsbearing streams were addressed as described in the following studies. At PRF in 1964, based on CAW stream samples and final waste batches, it was concluded that about 80 percent of the plutonium was held up in the DBBP column. Studies thus were undertaken to improve analytical and sampling methods to better account for the plutonium holdup. Analyses of the column feed taken from the slag and crucible dissolver showed pink solids which were assumed to be PuF4 (HW-83876, Chemical Processing Department Monthly Report for August 1964). In 1975, plutonium values in Miscellaneous Treatment (MT) centrifuge sludge samples measured by wet chemistry were about 60% of the nondestructive assay (NDA) results. This shortfall was found to be due to incomplete dissolution of plutonium which occurred in the destructive method (Letter, D. E. Turner to J. V. Panesko, Miscellaneous Treatment Centrifuge Sludge Sample (Including Summary of Results, 9/17/1975)]. The Chemical Technology Laboratory attempted to leach the plutonium from the sample and was unable to completely dissolve the plutonium. Based on Isotopic Source Assay Nondestructive Assay System and Calorimetry, coupled with the repeated leaching total material recovered by the CTL, the higher value was established to be the accurate quantity of plutonium in the sludge. The findings from this study emphasized that the chemistry methods that were designed to measure soluble plutonium under-reported plutonium in some samples because of the presence of the refractory solids. On the other hand, at the PFP, use of AT for plutonium measurements was subject to fewer interferences from other sources of alpha emitting materials (e.g., americium and curium) because those would have been removed at the separations facilities. However, americium-241 present in undissolved residues from older scrap material would have been a possible exception, especially for the fuel or reactor grades of plutonium. In waste, over-reporting of plutonium values would result from not accounting for the quantity of americium. As was demonstrated in several reports, there were categories of material such as the Rocky Flats ash and resulting centrifuge sludge from MT-4 dissolver solution, that were resistant to ordinary and even aggressive dissolution methods. The dissolution methods developed for dissolving these difficult scrap materials give evidence that plutonium assays could have been seriously under-reported when less aggressive or even equivalent methods were used in routine assays.

Official Use Only 114

RPP-RPT-50941, Rev. 0

Thus, it was observed that the ash and centrifuge sludge scrap materials did not dissolve in the MT-4 dissolver. When the laboratories conducted acid leaching studies, even after five contacts with heated, 15.4 M HNO3/0.1 M HF, some plutonium-bearing solids remained. Given that the leaching procedure conducted specifically to try to dissolve the solids was able to dissolve only about 60% of the material in the first contact, the analytical method most certainly would have also resulted in low bias of the plutonium value in wastes containing these types of solids. The laboratory method PuA-20A employed heating with strong acid to bring plutonium into solution from the sample, followed by plutonium valence adjustment, and TTA extraction of the plutonium from the aqueous phase. The organic solution was transferred to a stainless steel planchet and the plutonium measured by alpha spectrometry. The plutonium by method PuA-20A was specific to plutonium based on extraction chemistry but required that the plutonium in the sample be completely dissolved. The less specific method, E6-B, which simply dried an aliquot of sample directly onto a planchet for counting, would include the undissolved solids, attenuated to an unknown extent by the alpha particle absorption, as plutonium suspended in the sample by mixing and transferred to the planchet. However, in the E6-B method, americium in the sample would not be separated and thus would apply a high bias to the plutonium analytical result. Realization of the impact of undissolved solids was acknowledged in ISO-659, April 12, 1967, the Z-Plant Report from 2400 3-26-67 to 2400 4-9-67. In this report, the statement is made that the plutonium in MT dissolvers was not completely measured because plutonium in the solids was under-reported. A special study was conducted to determine the cause of MUF problems. The amount of plutonium discharged to cribs as solids was estimated to be up to 2.5-times the amount reported based on laboratory measurements. Gamma Solution Counter A partial resolution of the under-reporting of plutonium-bearing solids discharged to the waste and plutonium in other solids-bearing streams was made by the use of the gamma solution counter. The gamma solution counter relies on the gamma ray emission from anything present in the solution or solution-solid sample. The gamma counter is capable of not only determining the plutonium isotopic distribution, but could also determine the americium-241. In letter #65454-84-155 Pu/Am Concentrations, Pu Isotopic Composition and Thermal Behavior of Hood 7-C Floor Sludge, C. H. Delegard to R. L. Crocker, it is stated that plutonium isotopic composition was determined via two methodsmass spectrometry and a new gamma spectrometry technique implemented by Dennis Fazzari of AL-QA. The results of both methods were in good agreement. The Pu and Am concentrations in the sludge filtrate each were determined by two independent methods. Americium-241 concentration was determined by the gamma spectrometric method described above and also by gamma counting at the 222-S counting room. The results agreed well. This is the first reference to use of the isotopic solution counter for plutonium measurements at PFP and the observation was made that the gamma approach is more straightforward and probably yielded a better result.

Official Use Only 115

RPP-RPT-50941, Rev. 0

It must be noted that even with the gamma counter, the taking of representative solids-bearing samples still poses a risk of under-reporting plutonium concentration because of the settling of solids within the sample-collection vessel. The 1987 report RHO-SS-MA-30, Computerized Model for Calculating the Variance of Inventory Difference documents the use of a computer model which calculates the variance of an inventory difference (ID) for a processing area, material balance (MBA), or an entire plant. This variance is used in calculating warning and alarm limits. Note that the Z-Plant Analytical methods contained two methods for solution counter with one specifying CAW. The specific methods could not be located in the electronic files available, but it can be surmised that additional treatment for the CAW was included in this method and was likely a result of the investigations into MUF losses from not counting all of the solids in CAW. Analytical Technique Summary The following list shows the methods employed as evaluated for plutonium accountability measurements at PUREX and Z-Plant but does not indicate the sample points for each. The PUREX methods include accounting for uranium as well as plutonium, while Z-Plant measurements account only for plutonium.
Purex Analytical Isotopic dilution mass spectrometry (IDMS) Amperometric Titration (Chem assay) Alpha Count TK. F 15-16 Alpha Count / Extraction TK. F-18 Laser Fluorometric Davis Gray (Potentiometric Endpoint) Alpha Count TK. U3U4 Pu III Color Spectrophotometric Mass Spectromnetry Solution Counter TK. D-5 Solids Alpha Count TK G8 R8 TK. C-1 Davies Gray NDA Portable Gamma Segmented Gamma Calorimeter Neutron well Counter NDA Portable Gamma Segmented Gamma Calorimeter Neutron well Counter NAI Counter Canyon NDA Hood NDA Transfer Line NDA Filter Box NDA Misc. NDA NAI of 26 Vacuum Analytical Amperometric Titration (Chem Assay) Alpha Counting Mass Spectrometry Solution Counter Solution Counter (CAW) Z-Plant

Official Use Only 116

RPP-RPT-50941, Rev. 0

In summary, throughout the Hanford plutonium processing history, examples of plutonium in unmeasured (or incompletely measured) solids have been identified and analytical methods changed to better characterize the waste. In 1967, it was recognized that the analysis methods in use (PuA-20A) did not did not account for any solids that might be present. The new replacement method (E6-B AEA) reported 3 to 10 times more plutonium the previous method. This under-measurement was a significant contributor to the PFP material unaccounted for (MUF) due to large unreported discharges to cribs which occurred before the E6-B method was instituted. Even with the new methods, the problem of unaccounted plutonium in solids was not totally resolved, as problems with solids measurement in liquids waste continued to be an issue. In 1985, it was noted in the PRF campaign report that the total alpha and extraction methods for analysis of CAW disagreed by over 100% in some cases. Only after filtering the CAW did the discrepancy between methods become normal, further substantiating the influence of solids in plutonium analysis discrepancies.

4.12 PFP WASTE SOLUTION TRANSFERS TO TANK FARMS 4.12.1 PFP Aqueous Liquid Discharges after May 1973 In May 1973, discharge of high-risk PFP liquid wastes to the soil column ended. Beginning on May 15, 1973, the PFP high-salt and low-salt waste streams were comingled in the 241-Z sump tanks and transferred to the 242-T evaporator for neutralization and disposal in the Hanford underground storage tanks. At this point the 361-Z settling tank was taken out of service, as were the 216-Z-18 and 216-Z-12 cribs. In 1973, there were significant changes in the way organic solvents were treated in PRF and 242-Z so as to minimize the quantities of spent organic solvents requiring storage and at some point, disposal (ARH-CD-323, Z Plant Liquid Waste Disposal through the 241-Z Vault). The routine disposal of PFP aqueous liquid waste to the tank farms continued from 1973 until 2004. The disposal practices can be divided into two periods. From 1973 until the 242-T Evaporator ceased operations in 1980, PFP liquid aqueous wastes were blended in with much larger volumes of tank supernatant being processed through the evaporator. The comingled PFP aqueous wastes ended up as T-2 Saltcake and would reside in T-farm complex. The 242-T Facility operated until its shutdown as an evaporator in April of 1976, but continued to receive and dispose of PFP acidic aqueous wastes that were neutralized with 241-TX-118 supernatant until 1980. Plutonium solids transferred from PFP would have remained in tank 241-TX-118; however, Tank TX-118 supernatants were transferred to other tanks during this time period. Beginning in 1981, PFP aqueous wastes were neutralized in 241-Z sump tanks and transferred to the 244-TX tank and then to the double-shell tank, 241-SY-102 until PFP was decommissioned. 4.12.2 Acid Waste Discharge to 242-T May 1973December 1980 The 242-T Evaporator Facility was built in the early 1950s to reclaim non-boiling waste storage capacity in the existing underground waste storage tanks. The unit operated until 1955 and then was shut down. The facility was next modified in 1965 from batch operation to continuous operation and returned to service. During the period of time it operated, the 242-T facility and Official Use Only 117

RPP-RPT-50941, Rev. 0

the 241-TX tank farm underwent many modifications to improve evaporator operations and increase the capabilities of the system to concentrate and store salt waste in the 241-TX tank farm including the receipt and disposal of acid waste from PFP. The 242-T Facility operated until its shutdown as an evaporator in April of 1976 but, continued to receive and dispose of PFP waste until the end of 1980. (RL-SEP-396, 242-T Evaporator Facility Information Manual, RHO-CD-1410, 242-T Evaporator Facility Shutdown/Standby Plan, ARH-CD-178, Operational Safety Analysis 242-T Waste Evaporator) The 242-T evaporator was modified in March 1972 by AEC Project HCP-669 to provide for the disposal of the PFP high and low salt waste (HLSW) via the 242-T Evaporator Facility to the underground single shell tank system. This project provided for modification to the 241-Z Facility for collection of the acidic waste in tank D-5, connection of the tank D-5 transfer pump and steam jet to each of two new 2 inch encased stainless steel transfer lines (HSW 202 and 203 respectively) connected to the new tank R-1 located in the new 242-TA concrete vault by the 242-T Evaporator Facility. It also provided for modifications of the Evaporator Facility to enable the transfer of the HLSW from tank R1 into the existing Evaporator tank B-1 for blending and neutralization of the HLSW waste with Tank Farm waste and its subsequent storage in the 241-TX single shell tank farm (HNF-EP-0924, History and Stabilization of the Plutonium Finishing Plant (PFP) Complex, ARH-2335, Design Criteria Disposal of Z Plant Wastes, H-2-27334, 5, & 6). Project HCP-669 was completed and went operational on May 15, 1973. The best description of the 242-T Evaporator and how it was operated is contained in the Operational Safety Analysis 242-T Waste Evaporator (ARH-CD-178, Operational Safety Analysis 242-T Waste Evaporator, also H-2-37256, Engineering Flow Diagram Waste Evaporator, sheets1-3). To briefly summarize the process, waste was continuously transferred from the 750,000 gallon 241-TX-118 Evaporator Feed Tank at 20 to 40 gallons per minute via a pump and 2 inch transfer line into the 11,000 gallon near atmospheric evaporator vessel, located inside of the 242-T Building, where it was heated by immersed steam coils. The aqueous condensate was collected in condensate tanks, sampled and sent to a crib. The concentrated slurry from the evaporator was continuously drained by gravity to the 750,000 gallon primary 241-TX bottoms receiver tank. The cooled waste was transferred through a series of tanks until it was returned to 241-TX-118 Feed Tank. Fresh dilute feed was periodically introduced into the system and waste that couldnt be concentrated any further was removed from the system, usually to the 241-U Tank Farm where it eventually became feed for the 242-S Evaporator. The acidic HLSW from the PFP tank D-5 was transferred in nominal 2,500 gallon batches to tank R1 in the 242-TA Receiver Vault. Tank R1 was equipped with a mechanical agitator, transfer pump and two neutron monitors in wells installed inside the 4,200 gallon tank. After waste was received in tank R1 it was transferred to the 4,200 gallon tank B1 inside the 242-T Evaporator Building feed cell. Tank B1 was equipped with a mechanical agitator, and a pair of neutron monitors attached to the outside of the tank. There was also a single neutron monitor installed under the evaporator vessel. Waste was transferred from tank B1 via a jet eductor installed in the feed line from the 241-TX-118 Feed Tank. After the waste went through the jet and as it was mixed with the feed from 241-TX-118, it was sent to the evaporator if it was operating or was recirculated back to 241-TX-118 if the evaporator was down. There were at

Official Use Only 118

RPP-RPT-50941, Rev. 0

least two different occasions when the neutron monitors showed high readings and had to be temporarily set to higher limits. One occasion was in August 1973 when R1 alarmed from increased plutonium in the HLSW batches and the other in September 1973 when R1 alarmed from high americium content in the HLSW (Tank Farms Engineering Monthly Reports Sept 1973, 10071905598 and October 1973, 1007190599). Searches were made to locate copies of procedures and data sheets used for the operation of the Evaporator, including the 241-TX Farm, but were unsuccessful. The few data sheets that were found were not adequate to characterize the operation of the facility. A review of log books, Waste Status Summary Reports, Monthly Reports, Tank Farm Engineering Technicians Work Sheets, drawings, and tank history documents including WHC-MR-0132, A History of the 2000 Area Tank Farms, covering the period of time from May 1973 through December 1980 noted: Project HCP-669, which modified the 242-T Evaporator for receipt of acidic HLSW from the PFP tank D-5 and neutralizing and disposing of it to the 241-TX tank farm, became operational on May 15, 1973. 241-TX-118 was the dedicated feed tank for the evaporator and for blending HLSW waste from May 15, 1973 to April 15, 1976 and for HLSW blending only from April 15, 1976 to December 31, 1980. There were evaporator operation stand downs from June 14, 1973 until July 28, 1973 (241-T-106 tank leak) and from November 17, 1973 until March 25, 1974 (241-S-107 tank transfer leak). During the stand downs, HLSW processing to 241-TX-118 was allowed after brief pauses. 241-TX-105 was the primary bottoms tank for receiving slurry from May 15 through November 17, 1973. (WHC-SD-WM-TP-293, Tank 241-TX-105 Tank Characterization Plan, Tank Farms Monthly Report November 1973). 241-TX-109 was the primary bottoms tank for receiving slurry from March 25, 1974 through April 15, 1976 when the evaporator portion of the process was permanently shut down as the result of a plugged evaporator dropout line on April 4, 1976. The decision was made to discontinue the evaporator process but continue with HLSW neutralization using 241-TX-118. (ARH-LD-216 B, Atlantic Richfield Hanford Company Monthly Report April 1976,). The intermediate bottoms tanks between the primary receiver and the feed tank were 241-TX-102, TX-104, TX-106, TX-107, TX-110, TX-111, and TX-112. 241-TX-112 was unique in that it had a water cooled shell and tube heater exchanger for cooling the waste slurry before return to the feed tank. It operated from before May 1973 and its use was discontinued on May 17, 1973 so it was not used with the PFP waste cycle. The last HLSW transfer from PFP tank D-5 was received and blended to 241-TX-118 on 11-8-80. From this time forward the HLSW waste was neutralized and sent to the

Official Use Only 119

RPP-RPT-50941, Rev. 0

244-TX Double Contained Receiver Tank facility when it went operational in December 1981. Single Shell Tank (SST) use for liquid waste storage was discontinued by January 1, 1981. It was required that all SSTs have less than 50 kgallons of free liquid. 241-TX-118 was one of the last tanks in the 241-TX tank farms to be active and achieve the 50 kgallon goal. The free liquid in the tank was transferred to 241-SY-102 starting on November 12 and ending on November 16, 1980.

From the perspective of tank farm operations, HLSW data sheets and the evaporator data sheets provide volumes and quantities of PFP plutonium received and whether it was blended to the 241-TX-118 feed tank or the evaporator and thus the first bottoms tank. All sampling of Pu for material balances was performed at PFP. There was no independent sampling performed at the tank farms. The assumption was made that heavy non soluble large particles of plutonium oxide would rapidly settle in the first 241-TX tank that the neutralized waste was received in. Therefore most of the large dense plutonium would be expected in 241-TX-105, 241-TX-109 and 241-TX-118. Soluble and very slow settling small particles can be expected to be found in lesser amounts in the intermediate bottoms tanks that were between the first bottoms receiver tank and the 241-TX-118 feed tank. An effort was made to locate all of the data sheets for the PFP waste transfers and for operation of 242-T. While some were available, most were not found. The collection of data sheets was compiled (RPP-17017, 242-T and 242-TA Vault Information from Engineering Notebooks and Archived Data) which also has a collection of tables from an engineering log for D5 transfers to tank R1 to tank B1 to 241-TX-118 for the period 1-7-76 to 6-5-80. Waste volume information is available in the Waste Status Reports but not plutonium quantity information. These records are in Records Holding Area Box 89705 for 1974 through 1978. In summary, tank farm records could not be adequately reconstructed to provide a meaningful estimation of the quantity of plutonium or the volume of liquid waste transferred from PFP to the underground storage tanks. The discussions in 4.13, Estimate of total Plutonium Discharge to Tank Farms from PFP, based on PFP records, provide a much clearer understanding of the quantities of plutonium transferred to the tank farms; and was used in this investigation. 4.12.3 Neutralized Waste Discharge to 244-TX December 1981November 2004 In response to a DNFSB commitment to discontinue use of single shell tanks (SST) by January 1, 1981, the 244-TX Double Contained Receiver Tank (244-TX DCRT) was constructed by Project B-180, Salt Well Receiver Vessels, between the 241-TX and 241-TY Tank Farms. The new facility was for receipt and temporary storage of interstitial liquid from the salt well pumping systems installed in the 241-T, TX and TY tank farms and for neutralized high low salt waste (HLSW) from Z-Plant. When the DCRT was full, the waste was transferred to the 241-SY Tank Farm. The facility consisted of a nominal 25,000 gallon operating capacity, 31,000 gallon maximum, 12 foot diameter by 35 foot long horizontal carbon steel catch tank contained within a partially Official Use Only 120

RPP-RPT-50941, Rev. 0

steel lined concrete vault. The vault and tank were equipped with active ventilation, leak detection, liquid level and sludge level measurement, neutron monitoring, agitation recirculation system, water dilution capable transfer out system, and a number of interlock and control systems The 244-TX DCRT was equipped with an agitation system to prevent the buildup of solids from saltwell liquors and plutonium from PFP. The agitation system used a pump to circulate waste through a header system inside the tank that ran the length along the top of the tank and had a series of branch lines that ended in nozzles that discharged along the bottom of the tank. Initially the tank was equipped with a single agitator pump but since the transfer out pump was interlocked with the agitator pump operation, a second pump was added to increase availability. Only one pump was operated at a time. There were also three neutron monitoring probes, one at each end and one in the middle, located underneath the tank. The HLSW waste previously transferred from the PFP Tank D-5 via the 2 inch stainless steel encased HSW- 202 and 203 transfer lines to the 242-TA R1 Tank was rerouted to the 244-TX DCRT. The HSW lines were cut at the edge of the 241-TX tank farm and rerouted to the DCRT via encased carbon steel lines. Since the new transfer lines, the DCRT, and most of the rest of the double shell tank (DST) transfer and storage systems were constructed using carbon steel and storage and neutralization for acid waste were not provided by Project B-180, the acidic waste from the D-5 tank required neutralization at PFP prior to transfer to 244-TX. The 244-TX DCRT facility went operational and received its first waste transfer from the PFP D-5 Tank on December 1, 1981 and salt well pumping liquid waste from 241-TX Farm was initiated on December 3, 1981. The waste received was then transferred from 244-TX to Tank 241-SY-102 on December 4, 1981. Waste receipts from salt wells was completed by April 2000 and the final transfer from the PFP D-5 Tank was received in 244-TX on November 8, 2004. Although waste from the 244-TX DCRT could be physically transferred to any of the 241-SY tanks, it was only sent to tank 241-SY-102, except for the final transfer prior to facility deactivation which was completed on July 10, 2005 which went to tank 241-SY-101. The BBI FY06 Q1 update assumed that the final transfer to tank 241-SY-101 consisted only of supernatant because of a lack of effective agitation in the 244-TX DCRT and there was no change to the 244-TX sludge inventory. The transfer route from 244-TX DCRT to the 241-SY Tank Farm contains a high point at the 241-TX 152 Diversion Box and a low point at the 244-S Pump Pit. The 244-S low point could be drained after a transfer was completed and the line flushed but this was not the normal practice. The 244-S Facility transferred waste to 241-SY-102 from the 241-S and SX tank farm salt wells and other miscellaneous facilities. During the operating life of the 244-TX DCRT, transfer line failures occurred in the V-402 and V-398 lines. The leakage from the failed lines was collected in the transfer line secondary containment, which drained to 244-TX and the 241-U-301B catch tank respectively. Leak detection interlocks which shut the transfers down, limited the volume of waste that could have leaked. Only negligible amounts of waste from PFP would have been lost to the secondary

Official Use Only 121

RPP-RPT-50941, Rev. 0

containment. The secondary containment was flushed and the transfer lines were repaired and returned to service. An effort was made to locate all of the PFP waste transfer data sheets for 244-TX DCRT and these were not readily found. The more recent transfers have documentation in Tank Waste Information Network System (TWINS) and in the electronic data kept by Operations. Therefore, the estimates made for PFP of what was sent would be a more accurate representation of the plutonium sent to and received by the 244-TX DCRT. (see 4.13, Estimate of Total Plutonium Discharge to Tank Farms from PFP). The 244-TX DCRT was removed from active service on June 30, 2005 and isolation completed by June 23, 2006. 4.12.4 Waste Transfer by Truck (216-Z-8, 241-Z-361, TK D9) Three sets of truck transfers were made from PFP tanks to tank farm facilities (216-Z-8, 241-Z-361, TK-D9). They are described in the following sections. 4.12.4.1 216-Z-8 Pump out September- October 1974. The solid and liquid plutonium containing wastes from 216-Z-8 were to be recovered and transferred to the 241-TX-109 single shell tank in the 241-TX Tank Farm via tank truck (Letter, 1974, Removal of Waste from Tank 216-Z-8, D. A. Turner to D. G. Harlow, Letter, 1974, Z8 to 109-TX Transfers, R. M. Smithers to P. E. Alley, AEC-4783, 1974, Waste Tank Survey Contract AT (45-1) 2130). The 216-Z-8 Tank was a buried forty foot long by eight foot diameter horizontal tank with a nominal 56,000 liters (15,000 gallon) capacity. The tank was estimated to contain 28,968 liters of liquid (7,653 gallons) and 1,879 liters (496 gallons) of sludge. Samples from the tank provided worst case plutonium concentrations of .006 g/l aqueous and .755 g/l sludge. The total plutonium content of the tank was estimated to be 1,592 grams (174 grams in the liquid and 1,418 grams in the sludge) (Letter, 1974 Characterization of 216-Z-8 Tank and 361-Z Settling Tank, C. M. Peabody and D. A. Turner to D. G. Harlow). The best description of the 216-Z-8 pump out and transfer to 241-TX-109 is contained in an Arco memo, Attachment Final Report: 216-Z-8 Transfer by JVP (JV Panesko) (IDMS ref 293-003108). The following is a summary of the report. There are two inlets located at one end of the tank and a single outlet located 40 feet away at the other end of the tank. There are also four risers along the top of the tank. A 12 inch opening at 20 feet from the inlet which was buried but uncovered and used for the turbine pump installed on 10-18-74. A buried 24 inch manhole at 26 feet from the inlet and two 4 inch risers which extended above grade at 30 feet (used for samples and flushes) and 37.7 feet from the inlet (also see ref h-2-1). The report states that liquid was pumped from a 4 inch riser using a submersible electric pump. Four thousand gallons of flush solution was added and pumping was terminated on 9-10-74 when the pump plugged. The turbine pump was then installed and another one thousand gallon flush added and pumping was completed on 10-19-74. This report stated that 28,840 Liters (7,600 gallons) of liquid and 18,700 Liters (5000 gallons) of flush solution, for a total of 47,000 (12,600 gallons) were pumped from 216-Z-8 and that 46,200 Liters (12,350 gallons) of that total were received into Tank 241-TX-109. There were 18 cm (7 inches) of sludge remaining in the tank for a Official Use Only 122

RPP-RPT-50941, Rev. 0

volume of about 1870 Liters (500 gallons). A sample of the sludge contained 0.02 g/l of plutonium and a pH of 6. Although the report doesnt estimate the remaining plutonium, Burton reported after completion of the 216-Z-8 pumping, the AEC concluded that 1.2 kilograms of plutonium remained in the tank (AEC-4783, 1974, Waste Tank Survey Contract AT(45-1)-2130 IDMS ref D196844287). WHC-SD-DD-TI-057, 1991, Summary of Radioactive Underground Tanks Managed by Hanford Restoration Operations, provided a summary of the history of 216-Z-8 based on most of the same reference letters discussed above and estimated from the Final Report (IDMS ref 293-003108) result that only 38 grams of plutonium remained in the tank (and by difference 1,554 grams went to 241-TX-109). Reference RHO-RE-EV-46 P, The 216-Z-8 French Drain Characterization Study, contained information on the source and history of the waste which went into the 216-Z-8 tank and reported additional information on the pumping and photographs that were taken afterwards. The 216-Z-8 Tank and French drain operated from July 1955 to April 1962. The waste sent to 216-Z-8 was derived from slag crucible scrap recovery and consisted of neutralized filter washes that were composed of aluminum silicate precipitate from the crucible dissolution and pure diatomaceous earth powder used to aid in filtering. The waste transfers were small volumes consisting of 40 to 50 liters of waste, sampled for accountability, then jetted out to the 216-Z-8 tank and followed by a 200-liter water flush. The pipelines from the PFP building wall to the 216-Z-8 tank were 1-1/2 inch diameter stainless steel lines that were about 99 meters (325 feet) in length and have a total drop of about 2.4 meter (8 feet) in elevation (ref H-2-1). The lines entered at the top of the end of the tank and were immediately terminated in 90 degree horizontally orientated elbows pointed out towards the walls, so there wasnt much velocity flowing to the discharge located 12 meters (40 feet) away at the far end of the tank providing settling time before overflow to the French drain. It was reported that the tank didnt overflow to the French drain until October 1957 and that a total of 9,590 liters with a total of 48.2 grams of plutonium was overflowed (which would equate to about .005 g/l). As discussed in Ref Burton June 10, 1974 the tank was missing about 26,500 Liters (7,000 gallons) of liquid from sometime before it was removed from service until the time it was pumped. As a part of the studies this document reports on, neutron activation analysis by copper foils was performed in the 216-Z-8 tank and concluded that a remaining inventory of 1.4 kilograms of plutonium is plausible. And finally the document concludes at the end that, the worst case estimate, made in 1974 and supported by measurements made during the current study, of 1.5 kg in the tank is plausible. In summary, from review of Daily Operating Reports (DOR), Technicians notes, log books, and the discussion above, the first two transfers of waste from the 216-Z-8 tank received in tank 241-TX-109 on 9-5-74 (3822 gal) and 9-10-74 (2338 gal) were undiluted liquid waste. The transfer on 9-10-74 was a little short of the expected volume and can be attributed to pump failure. The transfer on September 18, 1974 (1650 gal) was made after flushing had started so the volume could have been in the truck from before flushing started or was after the 4000 gallons had been added but there was another pump failure noted in the DOR. On October 18, 1974 the turbine pump was installed and additional 1000 gallon flush added and the tank pumped on 10-19-74 and emptied to the TX-109 on 10-23-74 (3025 gal). The total volume

Official Use Only 123

RPP-RPT-50941, Rev. 0

reported to have been pumped from 216-Z-8 from the Final Report (293-003108) was 12,600 gallons. The Tank Farm Technicians notes indicate that a total of 10,451 gallons were received plus another 809 gallons of flush water used during unloading. Reference 293-003108 says Tank Farms received 12,350 gallons. The cause of the differences in received volumes reported is not known. The volume received in 241-TX-109 is determined by measuring the change in the tank liquid level. The tank was equipped with an automated gage that read to .05 inches. One inch of change in the 75 foot diameter tank is equal to 2,750 gallons. Although the 241-TX-109 level gage was very accurate, the tank was known to have crust material floating on top of the liquid which could easily lead to inaccurate material balances. To estimate the quantity of plutonium transferred from 216-Z-8 to 241-TX-109 the following data and assumptions were used: Volume transferred = 47,691 liters (12,600 gallons) from Panesko (293-003108) Volume of Sludge = 1,879 liters (496 gallons) from Peabody, Turner to Harlow ltr w/att Aqueous Concentration = .006 g/l from Peabody, Turner to Harlow ltr w/att Sludge Concentration = .755 g/l from Peabody, Turner to Harlow ltr w/att Assumed Volume of Sludge mobilized and suspended in transferred volume = 10%

From the above there were 286 grams of Pu in the liquid (47,691 l x .006 g/l) and 142 grams of suspended Pu solids (1879 lx .1 x 0.755 g/l) were transferred to 241-TX-109 for a total of 428 grams. The amount of solids transferred may not seem conservative, so the following analysis is provided as an explanation. Given the .02 g/l Pu sample results after pumping, this sample would have been taken from the same location that was disturbed by the addition of 4000 or more gallons of water after most of the free liquid had been removed from the tank so a low value compared to that previously obtained at that location would not be unexpected. It is also likely that the actual distribution of Pu solids is not homogenous given the 40 foot length of the tank, the small volumes transferred and the low velocity of the entering waste. It was observed that the sludge was deeper at the receiving end of the tank. It is likely that the heavier solids were deposited near the inlet but, much of the solids were extremely fine, easily suspended and slow settling so were more likely to be evenly distributed along the length of the tank so the sample point was likely representative of this type material. As there was no agitation system to mix the sludge with the flush water it is likely that the only material that was suspended and remained in suspension was the extremely fine material. The heavier material would not have been suspended or would have resettled prior to final pumping. The turbine pump would have only influenced the solids in its immediate vicinity of a 3 foot diameter area as the solids were not particularly fluid. Additionally as reported in Burton (D196844287) and RHO-RE-EV-46P there was little evidence that any significant volume of solids were removed by the pumping operations. There is no known independent sampling for plutonium from the tank truck, as the criticality specification was written to allow all of Z-8 to be transferred without the need for sampling as it was considered bounding (ARH-145 82.20, Criticality Prevention Specifications-Removal of Silica Gel waste from Tank 216-Z-8 to Tank 241-TX-109 Via Tank Truck and ARH-CD-132, Criticality Specification for241-TX-109). Official Use Only 124

RPP-RPT-50941, Rev. 0

4.12.4.2 241-Z-361 Pump out April-May 1975. The 241-Z-361 tank is a buried carbon steel lined concrete vault with internal dimensions of 23 feet long by 13 feet wide and a flat sloping bottom 17 feet deep at the inlet and 18 feet at the outlet with a nominal operating capacity of 36,660 gallons at overflow. The 241-Z-361 tank was used for solids settling prior to overflow to the cribs. The tank was put into service in 1949 and operated continuously until May 15, 1973 when wastes began to be sent to the Tank Farms (see section 4.7.3, 241-Z Settling Tank, for further detail). Details on the history of 241-Z-361 and on many of the follow-on investigations into the sludge characterization and criticality safety can be found in the references (WHC-SD-DD-TI-057, HNF-EP-0924, HNF-2012, HNF-1989, HNF-1692, HNF-8735, SGW-35955, SGW-39385, H-2-16024, H-2-16460). For the 241-Z-361 tank, the decision was made to remove only liquids from the tank and transfer it to the 241-TX-101 single shell tank in the 241-TX Tank Farm via use of a tank truck in 1975. A review of the Shift log books resulted in the following transfer volume estimates; Table 22. Truck Transfers 241-Z 361 to 241-TX-101 Based on Shift Logbooks Entries
Date April 28, 1975 April 29, 1975 April 30, 1975 May 1, 2005 May 2, 1975 May 6, 1975 May 7, 1975 May 8, 1975 Total Volume (Liters) 13500 12,500 12,500 12,500 12,500 12,500 10,900 7,800 94,700 Volume (Gallons) 3,575 3,300 3,300 33000 3,300 3,300 2887 2,062 25,024

As corroboration, the Tank Farm Engineering technicians waste status reports total for the eight transfers in April and May 1975 was 25,990 gallons. The most useful information available is from historical records, (1974 memo, Lundgren to Gaylord, 361-Z Settling Tank Analysis) and (1975 Letter, Crawley to Parazin, Request for DesignLiquid Evaporation from 241-Z-361 Tank). Corresponding PFP reports substantiate the Tank Farm data. A total tank volume calculated from measurements of 139,541 liters (36,866 gal) which would be at the overflow level to the crib, estimated that 84,509 liters (22,327 gal) was liquid and 55,032 liters (14539 gal) were sludge. The plan was to leave the solids and only pump the liquid. It was stated by Crawley that approximately 80,000 liters

Official Use Only 125

RPP-RPT-50941, Rev. 0

(21,136 gal) was pumped from 241-Z-361 and sent to the tank farms by tank truck leaving about 800 liters of liquid on the sludge. To estimate the quantity of plutonium transferred from 241-Z-361 to 241-TX-101 the following data and assumptions were used: Volume transferred = 98,410 liters (26,000 gallons) rounded value from Tank Farm Engineering technicians waste status reports which is largest volume reported Liquid Concentration = 0.0004 g/l Pu from the average of the two supernatant samples [(.0002 + .0006)/2] (reported in memo Lundgren to Gaylord) Sludge concentration = 0.91 g/l Pu from sample results (reported in memo Lundgren to Gaylord) Assumed 0.5% Pu solids entrainment in the liquid as stated goal was to leave the solids.

From the above there were 39 grams of Pu in the liquid (98,410 l x .0004 g/l) and 448 grams of entrained Pu solids (98,410 l X 0.91 g/l X .005) transferred to 241-TX-109 for a total of 487 grams. The Lundgren memo to Gaylord was used as the basis of the criticality specification (ARH-CD-296, Criticality Specification for 241-TX-101) for the transfer and represented the most current sample results available at the time of the transfer. The value of 0.91 g/l Pu is for wet sludge and is near the highest individual values reported in later detailed studies of 241-Z-361 to include HNF-2012, HNF-1989, HNF-8735 and SGW-35955 and remains conservative. There is no known independent sampling for plutonium from the tank truck, as the criticality specification was written to allow all of the 241-Z-361 liquid to be transferred without the need for sampling as it was considered bounding (ARH-CD-296). 4.12.4.3 Truck Transfer of PRF Uranium Solutions from Tank D9. (October 1972 March 1973). Mixed uranium-plutonium scrap was processed by the PFP from October 1972 to March 1973. Rather than dispose of the separated uranium to cribs, the uranium product streams were transferred by tanker truck to the PUREX Plant for neutralization and disposal to tank farms. The hazards evaluation prepared for this processing, ARH-2221, Hazards Evaluation Plutonium-Uranium Partitioning 236-Z Building Project HCE-651, states the uranium enrichment would be controlled below a limit of 1% 235U and that the expected plutonium concentration would be 100 ppm plutonium in the uranium solution. Below is a synopsis of the transfers made, based on limited documentation for the actual transfers. The first load of uranium waste was transferred to the 212-A Facility at the PUREX Plant on October 10, 1972 (ARH-2416 RD DEL, Chemical Processing Division Monthly Report Summaries Jan. 1972 December 1973). This transfer was reported to contain 130 grams plutonium and 55 kilograms of uranium in 3,500 gallons of solution. The second and third batches of PRF uranium partition waste were received at PUREX in November 1972 and transferred to tank 241-C-104 via tank F18 [Internal letter from C. M. Walker to R. L. Walser

Official Use Only 126

RPP-RPT-50941, Rev. 0

Monthly Report November 1972, (Walker, C. M. 1972-11-10)]. The total 241-C-104 fissile material of 42.09 kg reported after these transfers (Walker, C. M. 1972-11-10) is only 0.19 kg higher than the starting tank 241-C-104 inventory reported in an internal letter from C. M. Walker to C. J. Francis, Monthly Report October 1972 (Walker, C. M. 1972-10-14). By difference, the plutonium contained in the second and third transfers totaled 60 grams. No data could be located for transfers 4 and 5. The sixth and seventh trailer-loads were delivered to PUREX in February 1973 and routed to underground storage after neutralization (ARH-2416 RD DEL). The Richland Operations Office monthly report for March 1973 (PWM-550 #3 DEL, Nuclear Materials Management Branch Monthly Report for March 1973) indicates that processing of the mixed scrap had been discontinued at PRF in March and that an additional trailer-load of uranyl nitrate solution was transferred to PUREX for disposal into tank farms. Based on available transfer information, a total of eight batches of uranium solution contaminated with plutonium were loaded in at the PUREX 212-A Loadout Facility, transferred to tank F18 for neutralization and finally transferred to tank 241-C-104. The total quantity of plutonium contained in these transfers was likely less than 600 grams, based upon 190 g in three transfers of eight total transfers. There would have been no significant potential for plutonium oxide particles in the transfers. Examination of the engineering flow diagram for the processing of mixed oxide scrap at PRF (H-2-26721, Engineering Flow Diagram Pu U Partitioning) and the process flowsheet document (PFD-Z-180-00002, Plutonium-Uranium Coextraction-Partitioning Flowsheet) show that any plutonium oxide particles in feed would have to traverse four solvent extraction columns to reach the uranium waste stream. Plutonium oxide particles in the feed would have to carry over to the heavy-organic CAP stream, then not be stripped to the aqueous CCP stream, not be stripped to the aqueous COP plutonium product stream and finally stripped to the aqueous CUU uranium product stream. This sequence of transfer for plutonium oxide particles is very unlikely. The process flowsheet (PFD-Z-180-00002) does indicate a potential for precipitating plutonium fluoride in the CO column if its solubility limit is exceeded due to high plutonium and/or low nitric acid concentrations. The product uranium solution is filtered twice prior transfer to the 241-Z-D9 uranium holding tank as shown on drawing H-2-26722, Engineering Flow Diagram Pu U Partitioning & Uranium Loadout Facility. The specifications for these filters were not identified in the drawing and could not be determined. The process flowsheet document (PFD-Z-180-00002) indicates that the uranium waste stream (CUU) was also sampled and assayed to determine if the stream needed to be reworked before disposal. The potential for transfer of plutonium fluoride solids to tank 241-Z-9, and hence to PUREX and tank farms, would be unlikely. In summary, the transfer of uranium solution from PFP to tank farms via PUREX should not have included significant plutonium particles other than the normal plutonium hydroxide generated when acidic solutions containing soluble plutonium are neutralized with caustic.

Official Use Only 127

RPP-RPT-50941, Rev. 0

4.12.5 Waste from Solution Stabilization using Magnesium Hydroxide or Oxalic Acid (2003-2004) When production was terminated at the PFP, plutonium-bearing solutions in storage required processing to convert the plutonium to a solid form for stabilization. When the DNFSB Recommendation 94-1 solutions stabilization campaign was performed in the 1999 to 2004 time frame, the goal was simply to convert the solutions to a separated solid plutonium product suitable for thermal stabilization and a liquid waste stream. Solution stabilization using magnesium hydroxide and oxalate precipitation processes operated from 1999 to 2002 [CCP-AK-RL-101, Rev. 6, 2011, Central Characterization Project Acceptable Knowledge Summary Report for Hanford Plutonium Finishing Plant (PFP) ContactHandled Transuranic Debris Waste Stream: RLMPFPDD]. The solutions processing equipment located in Room 230C of the 234-5 Z Building used first magnesium hydroxide and then oxalic acid as the precipitating agents to facilitate removal of the plutonium from the solutions for stabilization and packaging. The filtrate and flush water were discharged to tank D-8. Oxalic acid was added to the plutonium nitrate solutions and plutonium was removed from solution by precipitation as the plutonium oxalate solid. Operating conditions for both the oxalate and magnesium hydroxide precipitation processes favored the formation of smaller solids particles by rapidly adding the precipitating agent to solutions at room temperature. In most operations, the fine, poorly crystalline particles would be considered undesirable, but for the purposes of limiting large, dense particle discharge to the filtrates sent to the waste, the operating conditions are advantageous. Solution containing precipitated solids was filtered through a nominal 100-m fiber filter, and the filtrate was collected in holding tanks. The large pore size for the filters would have been rapidly diminished by the solids cake built up in the filter pan, effectively reducing the size of particles that could pass through to the filtrate. Filtrate was tested for plutonium concentration, and when acceptable plutonium concentration levels were determined, the filtrate was sent directly to the 241-Z waste facility with no effort made to kill the excess oxalic acid. The filtrate was collected with other waste in 241-Z D-8 tank, where the excess oxalic acid could form additional plutonium oxalate solids. The fine plutonium oxalate solids in the filtrate also could age to form larger particulates. This leads up to the possibility that very large particles, up to 100 microns, could have been discharged or formed in 241-Z facility and also have the opportunity grow while waiting transport for final disposition. Two characteristics were unusual about these process streams compared to other PFP operationsthe stream from the magnesium hydroxide process was caustic and the stream from the oxalate process contained high concentrations of oxalic acid. Material sent to D-8 or D-7 was typically acidic and not neutralized until treated in TK D-5 for transfer to tank farms. For the oxalate process, the excess oxalic acid was not killed but rather was sent directly to waste. Both streams had the potential for precipitating other waste containing plutonium if other plutonium-bearing waste streams were discharged during the solution stabilization campaign.

Official Use Only 128

RPP-RPT-50941, Rev. 0

For the plutonium oxalate and hydrous oxide particles, additional particle growth post filtration is expected because the precipitation was carried out at ambient temperatures where equilibrium particle formation is slow. The analysis given in Report 1 in Appendix C shows, however, that growth likely would be slow for the alkaline magnesium hydroxide stream. However, plutonium oxalate particle growth could be substantial in the acidic oxalic acid process. The estimated total quantity of plutonium discharged during this period was estimated to be less than 2 kg. Of that quantity, if any formed large particles, it certainly would not be the entire quantity due to these process streams. Prior to discharge to tank farms, the oxalate filtrate was treated with sodium hydroxide to neutralize the acid waste. As shown in Report 2 of Appendix C, plutonium oxalate is metathesized in strongly alkaline solution. The density and structure of the resulting hydrolysis product is not known, but it does have the potential to generate large agglomerates of dense but small plutonium hydrous oxide particles. As noted in Report 2, these agglomerates likely are only weakly bound and are unlikely to survive even minimal slurrying and pumping. Solutions processed through the magnesium hydroxide or oxalate precipitation processes also had the potential to contain plutonium oxide particles. These would have been subjected to the same filtration as the particles produced by hydroxide or oxalate and are not expected to contribute to PuO2 particles in the tank farms. 4.13 ESTIMATE OF TOTAL PLUTONIUM DISCHARGE TO TANK FARMS FROM PFP.

This section provides an estimate of the total plutonium discharged for waste transferred directly from PFP. The estimate includes an accountable fraction, primarily soluble plutonium and a fraction estimated as particulate unaccounted for based on analysis of PRF MaterialUnaccounted-For (MUF) Analysis. The estimates of material transferred to the tank farm from truck transfers that originated at PFP were estimated separately as discussed in section 4.12.3 are accounted for in the Non-PFP Processes in section 5.4. 4.13.1 Accountable Losses 1973-2004 An estimate of the total plutonium discharged from PFP to the tank farms, based on accountability data, is provided in Table 23. It was compiled by comparing the sources listed, which covered several time frames. When conflicting data was found, the highest value from any source was used. These values report only soluble losses and an additional provision was made later for under-accounted material. A brief discussion of the sources and their limitations are as follows: WHC-EP-0793, Estimation of Plutonium in Hanford Site Waste Tanks based on Historical Records, September 1994. This document estimated Pu in all tanks farms and estimated discharges by year to Tank farms from PFP that started in 1973 through 1994. The data was based on source data and transfer data sheets. No data could be located for 1978 to 1981, even though it was known the PRF operated for a cleanout run in 19781979. The authors indicated the total might be higher. Official Use Only 129

RPP-RPT-50941, Rev. 0

ARH-1194, Tank Farm Data, May 1973 to March 1978, J. Murphy. This was an older document of working papers generated by safeguards. A table of transfers from PRO (Plutonium Recovery Operations) to 242-T is given with transfers starting in May 1973. The data discontinues after March 1978 and the reference given for continuance could not be located in classified or unclassified records. RHO-CD-492, Material Status Information Summary, March 9, 1979, J.W. Jordon. This document reconciles the PRF inventory from the last inventory in April 1976, just prior to the shutdown from 242-Z column explosion, to the time of the next inventory, completed after cleanout of the facility in February 1979. The value listed as liquids waste discard should account for all the plutonium lost during the PRF cleanout run in 19781979, listed as missing in WHC-EP-0793. The Pu discards would have been higher than normal, given the degraded nature of the organic and the large amounts of floor solutions and sludges that were processed during the cleanout. The PRE/LANMAS (LAN Material Accountability System) values are taken from the recent reporting of the accountability database by safeguards personnel. The values in the current system start in 1983 and continue through the time of the last transfer in 2004. The periods of peak discharges in 19841988 are included. In Table 23, the years with 0 grams Pu reported transferred are likely to represent years with no waste transfers from PFP. Several periods of no waste transfers are known to have occurred. The major process lines at PFP (PRF, RMC) were not operating and waste discharge to 241-Z would have been limited to small amount of laboratory solutions. For example for one period from May 1990 to November 1991, only one transfer of ~2,500 gallons to tank farms occurred. (15000-91-ECV-143, Waste Minimization ReportNovember). Other years with zero discharge reported, specifically 1980 and 1981, may have had small amounts of Pu, but were not reported in any of the references used in Table 23. For example, 224-T logbooks report that for 7 of 12 transfers in 1980, a total of 120 grams were transferred from PFP. In 1981, after the outage to connect 244-TX, one transfer in December was made to test the transfer line. The transfer was primarily flush water and the amount of Pu was likely negligible. Also, no process activity was occurring during 19801981 and minor under-reporting during these periods of no process activity is judged insignificant. Table 23 shows that the periods of highest discharge are during periods of PRF operation; 1973 to 1976 (PRF was shutdown after the explosion in 242-Z). 19781979 (cleanout run) Restart campaigns in 19841987.

Official Use Only 130

RPP-RPT-50941, Rev. 0

Discards during these time periods account for about 90% of the total. Discards during early periods PRF operation (19731976) would be expected to be lower in total plutonium than later periods (19781987), since the 242-Z process was operating and it removed about of the soluble Pu from the PRF CAW stream. A closer examination of the overall reported PRF aqueous waste can be performed by a review of the 1987 PRF campaign report (WHC-88-0071). The plutonium by waste stream is reported in Table 24. The CAW is the majority contributor to volume and total Pu content. The average CAW by year is shown in Table 25. PRF Operation was improved in 1987, in part from improved dissolution of Pu oxide, but waste values still typically exceeded flow sheet expectations.

Table 23. Total Reported Pu discharged from PFP to Tank Farms Based on Accountability Data (in Kilograms)
Year 1973 1974 1975 1976 1977 1978 1979 1980 1981 1982 1983 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 0.047 0.517 0.517 0.109 0.109 6.343 5.443 5.556 2.373 1.768 0.306 0.151 0.045 9.659 6.878 3.406 3.212 0.649 0.156 0.103 WHC-EP-0793 1.974 3.045 2.066 0.692 0.229 ARH-1194 1.432 2.693 2.059 0.684 0.436 0.027 7.941 RHO-CD-492 PRE/LANMAS Maximum 1.974 3.045 2.066 0.692 0.436 0.027 7.941 0 0 0.109 0.109 9.659 6.878 5.556 3.212 1.768 0.306 0.151 0 0.047 0.517 0

Official Use Only 131

RPP-RPT-50941, Rev. 0

Table 23. Total Reported Pu discharged from PFP to Tank Farms Based on Accountability Data (in Kilograms)
Year 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 Totals 30.728 7.331 7.941 0.646 0.274 25.814 0.03 0.239 WHC-EP-0793 ARH-1194 RHO-CD-492 PRE/LANMAS Maximum 0 0 0 0 0 0.03 0.239 0 0.646 0.274 45.682

WHC-EP-0793, Estimation of Plutonium in Hanford Site Waste Tanks based on Historical records, September 1994 ARH-1194, Tank farm data, May 1973 to March 1978, J. Murphy RHO-CD-492, Material Status Information Summary, March 9, 1979, J.W. Jordon Other point sources not listed but likely included in totals above; ZUTS Trend analysis for FY 1985 = 9.1 Kg, 1985 PRF Campaign report = 3.8 kg, 1986 PRF Campaign Report = 3.54 kg. 1987 PRF Campaign Report =2.252 kg The ANSI scrap code(F00) for non-conforming solutions for Safeguards results came from PRE/LANMAS, Safeguards concurs with results from 1983 thru 2004 data from PRE/LANMAS.

Table 24. Splits of Total Pu Discharged as Liquid Waste


Stream CAW CXP LSW FY 84 73.2% 16.8% 10.0% FY85 87.3% 0.9% 11.8% FY86 87% 3% 10.0% FY87 70.9% 11.9% 17.2%

CAW = CA column waste, the principal aqueous waste stream. CXP = CX column product, organic wash waste containing soluble Pu complexed with degradation products LSW Low salts waste, evaporator overheads and concentrator steam condensate, lab waste.

Official Use Only 132

RPP-RPT-50941, Rev. 0

Table 25. Average CAW Pu Concentration during the 1980s


Year FY84 FY85 FY86 FY87 Average CAW Concentration 0.021 g/l 0.017 g/l 0.032 g/l 0.0094 g/l % of CAW batches under flowsheet NA 20% 10% 50%

4.13.2 PRF MUFs and Impact on Waste As previously stated, a major realization that significant material unaccounted for (MUF) was occurring and was attributed to the presence of unmeasured solids came in 1967 with the change in accountability methods (ISO-659 RD, ARH-423-RD). Even after centrifugation of the PRF CAW, initiated in late 1967, MUF in PRF continued to be a problem. At year end in 1970, an AEC survey concluded MUF values were outside of uncertainty limits for 7 of 12 reporting periods (SSSM-1797, Survey No. 41, Facilities HVA and HVB). The totals reported for 1970 for PRF were: Normal Operating Loss (includes solid and liquids waste) = 7,059 gms Material Unaccounted For = 21,146 gms As the problems were attributed to the presence of solids and unrecognized waste loss, additional centrifuges were installed by the Prototype Solids Handling Project (ARH-2078) and became partially operational the same month that PFP aqueous waste transfers were routed to tank farms. Effectiveness of the actions can be gauged based on review of MUF analysis in the early 1970s presented in Table 26. In June of 1975, after several years of solids removal using centrifuges, the trending of PRF MUFs showed a continuing under-accounting trend, see Figure 45, (ARH-CD-411-A-DEL, Nuclear Materials ReviewPlutonium Finishing Plant Sections Third and Fourth Quarters, Fiscal year 1975). This report showed a continuing 2.9% underaccounting trend (reported as % of throughput). One of the actions coming out of this report included determination and effectiveness of the solids removal system, including centrifuge and filters. It was requested that Engineering evaluate the extension of this system to the other PRF aqueous waste streams, CX column carbonate waste (CXP) and CU column water wash (CUU). No evidence of this evaluation could be located.

Official Use Only 133

RPP-RPT-50941, Rev. 0

Table 26. Plutonium Scrap and Waste Loss and MUF Data from PRF
Material Unaccounted for (MUF) Scrap and Waste Loss kg kg 3.2 3.1 6.1 2.6 22.4 8.1 1.5 (15.3) 20.4 (21.6) 5.7 2.9

Time Period 1/70-6/70 1/71-6/71 1/72-6/72 6/72-11/72 1/74-12/74 12/746/75

Report/Comments ARH-1824-DEL, measurement biases due to solids in process continued ARH-2194-DEL, Un-dissolved solids in PRF caused high bias, overstating PRF product ARH-2584-DEL, One sources of bias given as difficulties caused by un-dissolved solids ARH-2728-DEL, Overstated PRF Product and understated filtrate returns ARH-3024-B-DEL, un-dissolved solids listed as one source of bias ARH-CD-411-A-DEL, Effectiveness of solids system removal is questioned.

ARH-1824-DEL, Nuclear Materials Review Plutonium Finishing Section, Third and Fourth Quarters Fiscal Year 1970 ARH-2194-DEL, Nuclear Materials Review Plutonium Finishing Section, Third and Fourth Quarters Fiscal Year 1971 ARH-2584-DEL, Nuclear Materials Review Plutonium Finishing Section, Third and Fourth Quarters Fiscal Year 1972 ARH-2728 DEL, Nuclear Materials Review Plutonium Finishing Section, First and Second Quarters Fiscal Year 1973 ARH-3024-B-DEL, Nuclear Materials Review Plutonium Finishing Section, Calendar Year 1974 ARH-CD-411-A-DEL, Nuclear Materials Review Plutonium Finishing Section, Third and Fourth Quarters Fiscal Year 1972

Changes in inventory procedures began to routinely include the PRF canyon and glovebox holdup values as determined by non-destructive assay (NDA). This may have had the effect of masking any remaining MUF issues or inventory differences that might have occurred, since such large error was associated with facility NDA. Inventory difference values for a period in the 1980s that included some PRF operating periods is present in Table 27. Values are similar to those seen in the 1970s. Using the bias factor for material under-accounted for in the above analysis and the production totals for the PRF campaigns of the 1980s the amount of material under-accounted for in these campaigns is estimated and presented in Table 28. A total of 23 kgs of Pu was estimated for the PRF campaigns of the 1980s.

Official Use Only 134

RPP-RPT-50941, Rev. 0

Not all Pu material under accounted for can be attributed to liquid waste. Many sources for this exist. A study on inventory differences at PFP over years of peak production was conducted in the 1970s (RHO-CD-194, A Study of the 234-5 Building Inventory Difference for the Years 1956 through 1966. In this study it was estimated that the 662 kgs Pu inventory difference at PFP during this time might be assigned to the following sources: 1. 2. 3. 4. 5. Equipment Burial Vacuum system Deposition Liquids waste Scrap receipts Routine Burial 200 kg 7 kg 100 kg 280 kg 75 kg

Figure 45. PRF MUF Trend Analysis 19731975 (ARH-CD-411)

A later Energy Research and Development Administration (ERDA) study on MUF focused on the total site plutonium inventory difference and concluded that from 1946 through 1976 that the total Pu inventory difference may be 1,286 kilograms (1978 ERDA letter, Plutonium Inventory Difference at Hanford). They stated that the contributions to the MUF from under accounted Pu in liquid waste, supportable by analysis, was 100kg of this total with a possible additional amount of 50kg more. By contrast, the contributions attributed to under-accounted Pu in the solid waste stream were stated as over three times this amount. Official Use Only 135

RPP-RPT-50941, Rev. 0

Therefore it is reasonable to assume that of the ~3% of throughput bias seen in the MUF trend analysis, that about one-third or 1% was under-accounted Pu in liquid waste, 1% was under accounted in the solid waste, and 1% was solids discharged with the liquid waste. The remainder of the MUF (2%) is attributed to the other sources, including Solid Waste, Equipment Holdup, Error in Receipts and other minor sources. Since accountable PRF waste loss is also about 1% of throughput, the overall total amount of Pu is liquid waste discharged to tank farms from PFP are about double the Table 23 value of 45 kg for a total of about 90 kg. The assumption that total loss in aqueous waste (soluble and insoluble) is about double the reported loss (soluble) is supported by the observations in the 1985 PRF campaign report that the total alpha and extraction methods for analysis of CAW disagreed by over 100% in some cases and that after filtering the CAW through millipore filters the discrepancy between methods was normal (RHO-RE-EV-75, 1985 PRF Weapons Grade Campaign Report). The bulk of the under accounted Pu in the waste stream sent to the tank farm is assumed to be particulate Pu; primarily refractory Pu oxide with some Pu fluoride and potentially undissolved Pu metal fines from dissolution of oxide from burned metal and other refractory forms. In Table 29 the assumed total loss values are listed by year, based on the accountability maximum loss values from Table 23. The greatest discharges occurred when process lines were operating. Table 27. PRF Plutonium Losses and ID Data from the 1980s
Normal Operating Loss (NOL) (grams) 576 Inventory Difference (ID) (grams) (3891)

Date 12/30/1982 to 11/8/1983 11/9/1983 to 11/3/1984 12/1/1984 to 6/11/1985

Source RHO-SS-VS-14, no process operations and gain due to facility cleanout and facility NDA RHO-SS-SR-128, 1984 PRF Campaign RHO-SS-SR-128 Rev1, 1985 PRF Campaign

9560 3826

5856 9348

RHO-SS-VS-14, Resolution of Material Balance Area 102 Plutonium Reclamation Facility, Unusual Inventory Difference RHO-SS-SR-128, Plutonium Reclamation Facility Material Balance and LEID Summary November 8, 1983-November 30, 1984 RHO-SS-SR-128, Rev 1, Plutonium Reclamation Facility Material Balance and LEID Summary December 1, 1984-June 11, 1985

Official Use Only 136

RPP-RPT-50941, Rev. 0

Table 28. PRF Operating Hours, Output and Estimated Material Unaccounted for
Year 1984 1985 1986 1987 Operating hours 2669.25 1537.25 867.75 1071.5 Kgs Loadout out 128 249 202 189 Estimated material under accounted for (3% bias adjustment) 3.8 kg 7.5 kg 6.1 kg 5.7 kg

Table 29. Total Assumed Pu Discharged from PFP to Tank Farms with Major Process Line Activity
Accountability Maximum (from Table 1) kgs 1.974 3.045 2.066 0.692 0.436 0.027 7.941 0 0 0.109 0.109 9.659 6.878 5.556 3.212 1.768 0.306 0.151 0 0.047 PRF Dissolution/ Solvent Extraction X X X X 242-Z Waste Treatment X X X X RMC RMA Oxide Metal Production Production X X X X

Year 1973 1974 1975 1976 1977 1978 1979 1980 1981 1982 1983 1984 1985 1986 1987 1988 1989 1990 1991 1992

Total Assumed Value kgs 3.9 6.1 4.1 1.4 0.9 0.1 15.9 0 0 0.2 0.2 19.3 13.8 11.1 6.4 3.5 0.6 0.3 0 0.1

X X X X X X X X

Official Use Only 137

RPP-RPT-50941, Rev. 0

Table 29. Total Assumed Pu Discharged from PFP to Tank Farms with Major Process Line Activity
Accountability Maximum (from Table 1) kgs 0.517 0 0 0 0 0 0 0.03 0.239 0 0.646 0.274 45.682 PRF Dissolution/ Solvent Extraction 242-Z Waste Treatment RMC RMA Oxide Metal Production Production

Year 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 Totals

Total Assumed Value kgs 1 0 0 0 0 0 0 0.1 0.5 0 1.3 0.5 91.4

4.14

CONCLUSION OF PFP LOSSES INCLUDING ESTIMATE OF LARGE PARTICLE FRACTION

An estimate of plutonium in liquid waste discharged by year was previously reported in section 4.13. This estimate included a soluble portion (measured Pu) determined from accountability records and an insoluble and largely unmeasured portion. The portion of material unmeasured was estimated from review of PRF MUF trend analysis and plutonium analysis methods. Various forms of plutonium would exist in the waste stream, including; Interfacial Crud which is Pu complexed with organic degradation products (MBP, DBP). The crud also acts as trap or carrier for solid particles, some potentially large in size. The estimate for crud includes only the soluble and complexed component. The insoluble particulate component is estimated separately. (A) Refer to columns in Tables 30 and 31 for the letter designators. PuO2 formed from metal burning, or other thermal oxidation methods like slag and crucible or ash. (B) PuO2 from precipitation and calcination of oxalate. (C)

Official Use Only 138

RPP-RPT-50941, Rev. 0

PuF4 from entrainment in the hydroflurinator off gas, or precipitation in scrap dissolution, or precipitation in the solvent extraction columns. (D) Pu hydroxide from caustic precipitation. (E) Pu oxalate discharge directly to waste. (F) Pu metal fines present in slag and crucible, oxide from metal burning and oxide received from offsite for recovery. (G) Pu nitrate the primarily soluble form as un-extractable loss. (H)

The amount of the forms would vary with time based on what processes operated and what type of feeds were being dissolved and processed in PRF. Members of the team who are Process, PFP, and Pu chemistry experts apportioned the percentage of total plutonium discharged from each of the various sources based on the plant operating history and the feeds processed. The results are shown in Table 30. Using these estimated percentage and the total assumed discard values, the kilograms of each plutonium waste form were determined and are presented in Table 31.

Official Use Only 139

RPP-RPT-50941, Rev. 0

Table 30. Total Assumed Pu Discharged from PFP Directly to Tank Farms with Percentages Assigned to Pu Waste Types
Total Assumed Value kgs 3.9 6.1 4.1 1.4 0.9 0.1 15.9 0 0 0.2 0.2 19.3 13.8 11.1 6.4 3.5 0.6 0.3 0 0.1 1 0 0 0 40 40 0/100 0/100 10 10 Scrap Oxides, solutions RF oxide, Scrap oxides Burned Metal, Filtrate, Scrap Oxide, RF oxide Filtrate, Scrap Oxide, RF oxide 15 10 10 10 Sludges, floor solution Sludges, floor solution 25 25 Crud (Pu-DBP Pu-MBP) (A) 2 2 2 2 Total PuO2 (B) + (C) 45 45 45 45 50 12.5 12.5 50 50 50 50 20 25 20 25 10 10 50 PuO2 split Metal/precip (B)/(C) 50/50 50/50 50/50 50/50 0/100 50/50 50/50 0/100 0/100 0/100 0/100 0/100 50/50 80/20 50/50 0/100 0/100 0/100 15 10 10 10 40 40 5 10 5 12.5 12.5 Pu F4 (also Na2PuF6, CaPuF6) (D) 2 2 2 2 Pu hydroxide (E) Pu Oxalate (F) Pu Metal fines (G) 1 1 1 1 Pu Nitrate (H) 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 50 PRF Training Run Only limited lab and dry operation RMC last operated Feed had more burned metal oxide PRF Last operated on hot feed Cadmium, Iron additions started in D5 No discharges, until Dec 1981 No transfers Neutralized in TK-D5 244TX/SY102 PRF Cleanout run PRF Cleanout, last RMA oxide run PFP/242-Z and RMA operation ceases

Year 1973 1974 1975 1976 1977 1978 1979 1980 1981 1982 1983 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996

PRF Feeds Metal, Scrap Oxide, SS&C Ash, Filtrate

Comments The fate of metal fines and of crud discussed in App. C

TF Rcpt Acid waste sent to 242-T TX105 from May 1973 to March 1974 TX 109 from late March 1974 to April 1976 TX 118 throughout until 1980

Official Use Only 140

RPP-RPT-50941, Rev. 0

Table 30. Total Assumed Pu Discharged from PFP Directly to Tank Farms with Percentages Assigned to Pu Waste Types
Total Assumed Value kgs 0 0 0 0.1 0.5 0 1.3 0.5 91.4 50 50 50 50 Crud (Pu-DBP Pu-MBP) (A) Total PuO2 (B) + (C) PuO2 split Metal/precip (B)/(C) Pu F4 (also Na2PuF6, CaPuF6) (D) Pu hydroxide (E) Pu Oxalate (F) Pu Metal fines (G) Pu Nitrate (H) 50 50 50 50 50 50 50 50 Batch Oxalate Precipitation Large solids content Mg(OH)2 precipitation

Year 1997 1998 1999 2000 2001 2002 2003 2004 Totals

PRF Feeds

Comments

TF Rcpt

Crud values are soluble Pu complexed with organic degradation products, It can include particulate Pu, PuO2, Pu metal fines, PuF42.5 H20 counted in their columns. After burning ~50% of crud was dissolved in B acid

Official Use Only 141

RPP-RPT-50941, Rev. 0

Table 31. Kgs in Waste Stream by Form and Particle Size


Crud (A) 0.08 0.12 0.08 0.03 0 0.03 3.98 0 0 0 0 2.90 1.38 1.11 0.64 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 10.34 From From metal Precipitation burning (B) (C) 0.88 1.37 0.92 0.32 0 0.01 0.99 0 0 0 0 0 1.73 1.78 0.8 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 8.79 1.00 0.00 8.79 0.88 1.37 0.92 0.32 0.45 0.01 0.99 0 0 0.1 0.1 3.86 1.73 0.44 0.8 0.35 0.06 0.15 0 0.04 0.4 0 0 0 0 0 0 0 0 0 0 0 12.97 0.90 11.67 0.00 0.00 0.00 Pu F4 (D) 0.08 0.12 0.08 0.03 0 0.01 1.99 0 0 0 0 2.90 1.38 1.11 0.64 1.4 0.24 0 0 0.01 0.1 0 0 0 0 0 0 0 0 0 0 0 10.09 Pu hydroxide (E) 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0.05 0.25 0 0 0 0.30 Pu Oxalate (F) 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0.65 0.25 0.90 Metal Fine (G)s 0.04 0.06 0.04 0.01 0 0 0 0 0 0 0 0 0.69 1.11 0.32 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 2.28 1.00 2.28 0.00 22.73 Pu Nitrate (H) 1.95 3.05 2.05 0.7 0.45 0.05 7.95 0 0 0.1 0.1 9.65 6.9 5.55 3.2 1.75 0.3 0.15 0 0.05 0.5 0 0 0 0 0 0 0.05 0.25 0 0.65 0.25 45.65

Year 1973 1974 1975 1976 1977 1978 1979 1980 1981 1982 1983 1984 1985 1986 1987 1988 1989 1990 1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 Totals fraction of particles >10 um Total Large particles (kgs)

Total 3.9 6.1 4.1 1.4 0.9 0.1 15.9 0 0 0.2 0.2 19.3 13.8 11.1 6.4 3.5 0.6 0.3 0 0.1 1 0 0 0 0 0 0 0.1 0.5 0 1.3 0.5 91.30

Official Use Only 142

RPP-RPT-50941, Rev. 0

Of the forms of plutonium, some will not persist and can be removed from the quantities of concern. These include; Crud, which will breakdown and allow the soluble Pu component to from a hydroxide. The insoluble particles trapped in the crud binder will be released when it degrades and are accounted for separately. (see Appendix C, Report 4, Interfacial Crud Disposition in Alkaline Tank Waste) Pu fluoride and Pu oxalate particles which will undergo alkaline hydrolysis as described in section (see Appendix C, Report 2, Metathesis of Plutonium(IV) Oxalate and Plutonium(III)/(IV) Fluorides in Alkaline Media to Form PuO2xH2O) Pu hydroxide, which will form small, amorphous particles less than 10 micron in size. Pu nitrate, which was neutralized to a hydroxide in either tank farms or 241-Z and coprecipitated as fine particulate less than 10 micron.

Pu oxide made from precipitate and from metal burning are two forms that will persist. The Pu metal fines are not thermodynamically preferred and might not persist, but a small fraction is known to survive oxidization and dissolution in nitric/HF acid so it is assumed on conservative basis that some fine particulate, perhaps protected in a ceramic oxide coat which could exist in the Tank Farms, (see Appendix C, Report 3, Plutonium(IV) Oxide and Unburned Plutonium Metal from Operations to Burn Plutonium Metal (BURNT METAL) In letter report 24590-CM-HC4-W000-00176-T02-01-00002, Historical Overview of Solids in PFP Aqueous Waste Transferred to Tank Farms: Quantity of Plutonium, Particle Size Distribution, and Particle Density, a comprehensive data review was conducted on particle size and density on plutonium prepared and processed by various methods. Extracts of this particle size data was presented in section 4.9. The WRPS team reviewed the basis and found no new information in this investigation to question the data and relevant portions used in the estimate of large particles are presented in Table 32. Table 32. Data Used on Large Particle Determination
Form Oxide from Metal Burning Oxide from Oxalate Precipitation Pu Metal Fines Particle Size >10um 55100% 4589% 100% Particle Size >40 um 13100% 05% 100% Particle Density (gms/cc) 4.411.1 3.57 19

Using conservative fractions for particle size greater than 10 micron from the data compiled in the above report, the total kilograms of the three persistent forms with potential for particle size greater than 10 micron are estimated and shown in at the bottom of Table 31. The majority of large particles are PuO2 from oxide precipitation and metal burning. A small amount of Pu metal fines are present as well.

Official Use Only 143

RPP-RPT-50941, Rev. 0

From this analysis, the following conclusions can be shown: About 91 kilograms of Pu were discharged to tank farms from PFP from 1973-2004. About of the total was measured and accounted for and discharged as nitrate, converted to a hydroxide, which is fine particulate less than 10 micron. The remaining 45 kgs was not measured and not accounted for, primarily as particulate forms of calcined plutonium. Some metal fines could persist in tank waste as particles greater than 10 micron. The form and amount of the large particles are estimated at approximately 20.5 kg of Pu oxide and 2 kg as Pu metal fines. No specific error determinations were made, however, the reported values are believed to be conservative estimates and the margins of safety are detailed below.

Margin of Safety The conservatism in the estimates are detailed to provide the users of this document with an understanding of how the estimates were made and establish a lower limit for the amount of material sent to the tank farms: The largest value from any data source was used for accountable discard values. A 2X multiplier was used for unaccounted for material. Even though the number was higher during earlier RECUPLEX operations, the 2 X values reflects improved accountability measurements and efforts to capture solids. The largest fraction of MUF (one-third) that is believed to be attributed to liquid waste was used. DOE documents indicate this number could be more on the order of one sixth. The largest values of particle size distributions were used. 100% greater than 10 micron for oxide from burned metal 90% greater than 10 microns for calcined oxide Pu metal fines are assumed to persist in waste.

In order to establish a lower limit from the margins of safety an analysis for each tank was performed and recorded in Appendix I. The table used the lowest values of the PSD greater than 10 micron and assumed that the metal fines corrode to a fine oxide after 20-30 years in the waste tanks. This resulted in the total amount of plutonium oxide being sent to the tank farms to be reduced to approximately 18kgs.

Official Use Only 144

RPP-RPT-50941, Rev. 0

5.0

FUEL PROCESSING FACILITIES (T/B PLANT AND REDOX AND PUREX)

Earlier sections of this report covered the history of Hanford and the potential discharge of PuO2 from the PFP to the Tank farms. This section covers all other facilities that potentially could have discharged PuO2 to the tank farms. The team evaluated a number of facilities that contributed plutonium to the tank farm. These facilities were evaluated to see if there was a source of large and dense plutonium particles (PuO2). This section is organized into three major sub-sections: (1) Facilities that did not process PuO2, so could not have contributed PuO2 to the tank farm. (2) Facilities that processed plutonium but had no credible pathway for that PuO2 to reach the tank farm. (3) Facilities that processed PuO2 that may have discharged PuO2 to the tank farm.

5.1 5.1.1

PROCESSES THAT DID NOT PROCESS PUO2 Bismuth Phosphate, PUREX, REDOX, Fission Product Recovery, Uranium Recovery

The three major plutonium separation processes used at Hanford are the Bismuth Phosphate (B Plant and T Plant), REDOX, and PUREX processes. These processes dissolved metallic fuel from Hanford reactors to extract plutonium. The plutonium in the fuel was dispersed in a matrix that easily dissolved rather than as an insoluble particulate. Thus, the plutonium in this fuel would completely dissolve. There were no high temperatures or other conditions in these facilities that could have converted plutonium to PuO2 except for PUREX N Cell operation after 1984 (see section 5.2). Thus, it can be concluded that although these facilities processed large quantities of reactor fuel and generated the majority of plutonium containing liquid waste, they did not contribute PuO2 to the tank farm under normal operating conditions. In addition to the normal operation of PUREX and REDOX, both REDOX and PUREX had special campaigns that potentially could have contributed PuO2 to the tank farms. These special campaigns are discussed in Section 5.3. Some of the high-level waste discharged from Hanford was retrieved for radionuclide removal. Two such campaigns were the uranium recovery from some high-level waste in U plant and the fission products (Sr-90 and Cs-137) recovery from some high level wastes in B Plant. These processes did not add any additional plutonium to the waste, so they were also excluded as a source of PuO2 to the tank farm. 5.1.2 Critical Mass Laboratory

In 1949, a facility for plutonium criticality studies was built near the100-F Area as part of the P 11 project and called the Critical Mass Laboratory (CML). The P-11 Facility consisted of the 120 Experimental Building and 123 Control Building (HW-24514 DEL, Critical Mass Studies of Plutonium Solutions). The P-11 Facility operated until November 1951 when a criticality occurred in the 120 Building that resulted in extensive plutonium contamination inside the building. Decontamination of the building was almost complete when, on December 4, 1951, spontaneous combustion of cleaning materials caused a fire that completely burned the inside of Official Use Only 145

RPP-RPT-50941, Rev. 0

the building. Some soil contamination was caused by water used to fight the fire carrying plutonium onto the soil. The contamination was stabilized. In 1974, the 120 Building was removed and the associated crib and underground piping were decontaminated and demolished (DOE/RL-88-30 Rev. 19). The area was released from radiological controls. No source of plutonium particle waste transfers to tank farms was identified from this facility. In 1959, General Electric built the 209-E Critical Mass Laboratory adjacent to the Hot Semi-Works facility in 200 East Area. The 209-E CML operated from 1960 to 1986. It was decontaminated and decommissioned between 1986 and 1988, and was placed into an unmanned standby condition at the beginning of 1989 (WHC-EP-0342, 209-E Laboratory Reflector Water Stream-Specific Report). Most of the waste from this facility was in the form of low-level radioactive waste and cooling water that went to cribs. During the operation of the 209-E critical mass laboratory there were no direct transfers to tank farms identified. Wastes with recoverable quantities of plutonium and uranium were collected in waste storage tanks, while low level contaminated liquid waste was transferred to cribs. Sampling of the waste hold-up tank contents was performed prior to discharge (HW-66266, Hazards Summary Report for the Hanford Plutonium Critical Mass Laboratory). After sampling and verifying that the waste met discharge requirements, the waste from this tank was discharged by pump to an underground crib or dry well and absorbed into the soil (HW-57603, Preliminary Hazards Study of the Hanford Plutonium Critical Mass Laboratory). From the mixing hold tanks was a waste storage tank that was intended for the accumulation of used experimental solutions and for containing recoverable concentrations of plutonium. Plutonium nitrate solutions that were not directly reused within the CML were shipped to PFP for recovery (1980 Letter R80-1035, J. H. Roecker to Mr. O. J. Elgert, Request for Approval to Transport PR and 10-L Containers (Contract DE-AC06-77RL0l030); 1984 Letter, D. D. Scott to H. E. Ransom, Concurrence to Ship Plutonium Nitrate from 209-E, Critical Mass Laboratory). The purpose of the waste hold up tank was to prevent the inadvertent transfer of recoverable plutonium to the waste stream. This further supports the observation that at the CML, plutonium was considered an asset be recovered after use in experiments. Plutonium potentially contaminated waste was transferred to a tank that had a capacity to sample transfer to and transfer from the holding tank waste from the tank after sampling to meet release limits was transferred to the cribs (HW-68857, Hanford Plutonium CML Nuclear Safety, Operational Limitations, Instructions, and Emergency Procedures). Although the preparation for shut-down included removal of fissionable material from the numerous storage and processing tanks in the CML, none of the material generated from the tank work entered the tank farms during shutdown. Solutions containing plutonium and uranium for recovery were transferred to 224-T and later to Z-Plant where they remained until the solution stabilization campaign. Criticality research was also conducted with solid special nuclear material and fuels. The Plutonium Finishing Plant (PFP) received hundreds of polystyrene compacts (polycubes) containing PuO2 from the Critical Mass Laboratory. A facility for recovering the PuO2 was built in Miscellaneous Treatment (MT) Glovebox 4 in the Plutonium Reclamation Facility (PRF). Not all of the polycubes were processed. PuO2 from this pyrolysis was either reused as the oxide for additional experiments in the CML or dissolved through the MT scrap recovery dissolvers. The

Official Use Only 146

RPP-RPT-50941, Rev. 0

miscellaneous treatment and PRF recovered plutonium polycubes by pyrolysis of the polystyrene contained in the cubes. From this source, any of plutonium oxide solids would be generated with other MT dissolver solutions and be present in PRF liquid waste. These would be included with the Pu oxide estimates from PRF in section 4.13 and 4.14. During decontamination and decommissioning (D&D) activities, all contaminated process tanks were flushed and the resulting liquids were sent to waste tank storage or absorbed and disposed as solid waste (CCP-AK-RL-110, Central Characterization Project Acceptable Knowledge Summary Report for Hanford Building 209E Contact-Handled Transuranic Debris Waste from Decontamination and Decommissioning, Waste Stream RLCCP209ED1; WMP-32037, Acceptable Knowledge Summary Report for 209-E Critical Mass Laboratory Mixed Transuranic Debris). While there may have been potential for solids in this waste stream, the total plutonium remaining in the facility prior to D&D was an estimated 15 grams. Solid waste measurements reported 29 grams removed. It is unlikely that any plutonium beyond contamination levels was sent from 209-E CML to tank farms. 5.1.2.1 Hot Semiworks (201-C). The activities performed in the Hot-Semiworks (201-C) did not contribute to any large plutonium oxide particles, in particular from the Plutonium Finishing Plant, that might be present in the in the high level underground waste storage tanks. There is no data that indicates that the work performed at the Hot Semiworks with irradiated fuel or irradiated waste solutions sent plutonium oxide particles to the waste tanks. The 201-C (Hot Semiworks) went operational in October 1949 [PNNL-13524, Historical Time Line and Information about the Hanford Site). The Hot Semiworks was built to: (1) develop optimum conditions for the economic operations of the Reduction-Oxidation (REDOX) and tributyl phosphate (TBP) plants, (2) procure engineering design data which would allow the specification of process equipment required for new processes such as Plutonium Uranium Extraction (PUREX), (3) provide facilities for the study of future processes and engineering problems on a semiworks scale employing radioactive process solutions, and (4) provide facilities for immediate troubleshooting for urgent separation plant problems (HW-31767, Hot Semiworks REDOX Studies). The work performed at the hot Semiworks can be divided into two time periods; the work from 1952 through 1956, and the work from 1960 through 1967. In the first phase the Hot Semiworks performed pilot level studies to support the work conducted at the reprocessing facilities using irradiated fuel from the Hanford reactors. At a pilot fuel reprocessing plant, reactor fuel was brought into the Semiworks and dissolved. The plutonium was recovered and shipped to another facility. The high level waste was transferred to waste storage tanks. The REDOX process was tested processing irradiated uranium fuel elements to recover plutonium from 19521953. The facility was modified in the mid 1950s to serve as a pilot plant for demonstrating the PUREX flowsheet and the Waste Metal recovery flowsheet. (HW-15825, Modification of the Hot Semiworks to Accommodate the TBP Waste Metal Recovery Flowsheet; HW-22771, Modification of the Hot Semiworks to Accommodate the PUREX Process Flowsheet; HW-22772, Modification of the Hot Semiworks to Accommodate Waste Metal Recovery Flowsheet using 30% TBP in CCl4,). A time line of the operations conducted in the Hot Semiworks is shown in

Official Use Only 147

RPP-RPT-50941, Rev. 0

Figure 46 (WHC-SD-DD-TI-039, Tank 241-CX-71 Preliminary Waste Characterization). The Hot Semiworks was placed in standby mode in July 1957. A thorough study was conducted to trace the flow path of the radioactive wastes transferred from the Hot Semiworks through the Hanford Site waste tanks (RPP-15408, Origin of Wastes in C-200 Series Single-Shell Tanks; RPP-RPT-23177, Origin of Waste in Tank 241-AW-105). There were three tanks associated with the operations at the Hot Semiworks; 241-CX-70, 241-CX-71, and 241-CX-72. The radioactive waste from the REDOX process research runs were concentrated and transferred to tank, TK-70 (also designated as 241-CX-70) located at the Hot Semiworks 241-CX tank farm facilities. The high radioactive wastes from the demonstration of the PUREX flowsheet process research runs were concentrated and transferred initially to Tanks 241-C-201 through 241-C-204. The liquid inventory from Tank CX70 was pumped out of the tank in 1979, leaving approximately 39,000L (10,300 gal) of soft sludge in the tank (WHC-SD-DD-TI-057, Summary of Radioactive Underground Tanks Managed by Hanford Restoration Operations). The supernatant was pumped to Tank 241-C-104. Analysis of the supernatant in 1974 indicated a plutonium concentration of 3.5E-04 g/L (9.24E-05 g/gal). In 1988, the sludge was sluiced out of 241-CX-70 and pumped via the 244-CR vault to Tank 241-AW-105 in 200 East Area Tank Farms. A total of approximately 39,000L (10,000 gal) of waste was removed using 530,000L (140,000 gal) of water. The remaining sludge and water remaining in CX-70 were drummed and transferred to the Hanford Site Central Waste Complex in 1992 (RPP-RPT-23177). Available historical documentation (Cummings 1989 memo to Harlow/Teal, Disposition and Isolation of Tanks 270-E-1, 270-W, 241-CX-70, 241-CX-71, and 241-CX-72] indicated that Tank 241-CX-71 was used during the REDOX and PUREX operations and the decontamination flushes following the completion of the PUREX operations. After the decontamination flushes the tank was no longer used after June 1957. The tank received approximately 33.3 million L (8.8 million gallons) of waste. The waste contained on the average 8.7E-04 g/L (0.0033 g/gal) of uranium and 2.4E-08 g/L (9.3E-08 g/gal) plutonium. Tank-CX-72 was used as an experimental tank to study the characteristics of self-concentrating wastes from the PUREX process. The waste in the tank was thought to contain 150 to 200 g of 239 Pu. In the fall of 1986 the tank was decommissioned and filled with grout. Neutron and gamma measurements indicated about 10 feet of contaminated sludge in the tank below the grout. Several cribs: 216-C-1, 216-C-3, 216-C-4, 216-C-5, 216-C6, and 216-C-10 were also associated with activities conducted at the Hot Semiworks. Process condensates from the evaporation of the radioactive waste were transferred to Crib 216-C-6. Organic solvent waste from the PUREX research process runs were transferred to Crib 216-C-4. Crib-216-C-1 was used during the REDOX operations, the PUREX operations and the decontamination flushes following completion of the PUREX operations. Crib 216-C-1 was in service from January 1953 through

Official Use Only 148

RPP-RPT-50941, Rev. 0

June 1957. During that time, waste flows to the crib reportedly totaled 2.3 E+7 L (6.16 E+6 gallons). Figure 46. Timeline for Hot Semiworks

Tanks 241-C-201 through 241-C-204 were used from May 1955 through October 1956 to receive and store neutralized radioactive waste originating from research and development activities conducted at the 201-C Hot Semiworks (RPP-15408, Origin of Wastes in 200 Series Single- Shell Tanks). Tank 241-C-201 received Hot Semiworks radioactive waste in May 1951 and was reported as being filled with 206,300 L (54,500 gallons) of waste as of November 1955. Tank 241-C-202 received waste from the Hot Semiworks beginning in November 1955 and was reported filled with 206,300 L (54,500 gallons) of liquid waste in May 1956. Tanks 241-C-203 and 241-C-204 received liquid waste from December 1955 to November 1954 and each tank was reported to contain 130,600 L (34,500 gallons) of waste. Tanks 241-C-201 through 241-C-204 did not receive any additional waste after November 1956. Approximately 7,200 L (1,900 gallons) of supernatant was transferred from Tank 241-C-203 to Tank 241-C-109 in January through March 1970. The supernatant contained in tanks 241-C-201 through 241-C-204 was pumped to Tank 241-C-104 in April through June 1970. With the exception of Tank 241-C-204, these tanks contained only a heel of supernatant and sludge. The liquid in Tank 241-C-204 was

Official Use Only 149

RPP-RPT-50941, Rev. 0

transferred to single shell tanks 241-C-104 and 241-C-109 in 1970. Residual liquids were subsequently transferred from these tanks into single-shell tank 241-C-106 in October 1980. In 1960 the Hot Semiworks was selected for a strontium recovery program. Modifications were made to the facility from 19601967, and processing for strontium was initiated in early May 1961 and completed in October 1961 (HW-72666, Hot Semiworks Stronium-909 Recovery Program; RL-SA-40, Status of Fission Product Recovery at Hanford). The program involved the separation of strontium and rare earths from PUREX acid waste by lead carrier sulfate precipitation and subsequent separation of rare earths from strontium by oxalate precipitation. Strontium was further purified by di (2-ethylhexyl) phosphoric acid (D2EHPA) solvent extraction, followed by preferential stripping of the strontium from the solvent with 1M citric acid. Cerium-144 was also separated from the rare earths by D2EHPA acid solvent extraction using silver catalyzed persulfate oxidation of Ce(III) to Ce(IV). Promethium-147 was subsequently separated from the cerium free rare earths by chromatographic ion exchange. Strontium-90 and the rare earths, which include cerium-144 and promethium-147, were removed from the PUREX acid waste in the PUREX plant. The PUREX acid waste stream from the waste concentrator was centrifuged to remove solids, primarily silica and sulfates. The solids were leached with nitric acid to recover residual fission products. The leachate was combined with the solids-free supernatant solution. A lead-carrier sulfate precipitation process was used to recover strontium, cerium, and the rare earths. The cerium and rare earths were precipitated as the sodium sulfate-rare earth sulfate double salt. The precipitate was metathesized with a mixture of sodium hydroxide and sodium carbonate to remove lead sulfate and to convert strontium, cerium, and the rare earths to acid-soluble carbonates. The precipitate was dissolved in nitric acid. The strontium was separated from the rare earths in B Plant equipment. The strontium fraction was transferred to the Strontium Semiworks. The purification was performed in two pulse columns using D2EHPA extractant in a TBP diluent. Cerium and promethium were separated from lead and the rare earths with the D2EHPA system used for strontium. Cerium was separated from promethium and the rare-earth by batch solvent extraction using D2EHPA after oxidizing the cerium to Ce(IV) by silver catalyzed sodium persulfate. Ce(IV) was extracted into the organic phase, effectively separating it from the other rare earths and promethium. The Ce(IV) was reduced to Ce(III) and recovered in nitric acid. Promethium was recovered from the cerium-free rare earth fraction by chromatographic ion exchange. During 1966, REDOX processed Shippingport Reactor (pressurized water reactor [PWR]) fuels. An americium-curium campaign successfully conducted in December recovered approximately 1,200 g of raffinate from the PWR waste. The solution was shipped to the Strontium Semiworks for purification. A counter current solvent extraction process was used to separate americium, curium and the rare earths from the impurities. In that step, americium, curium, and the rare earths were extracted as solvated salts of D2EHPA. Thermal concentration, followed by sugar

Official Use Only 150

RPP-RPT-50941, Rev. 0

denitration yielded a product suitable as feed for the follow-on ion-exchange which was conducted at Battelle Northwest (BNW) Laboratories. (RHO-CD-505 RD, Synopsis of REDOX Plant Operations; BNWL-SA-1492, Simultaneous Recovery and Purification of Pm, Am, and Cm by the Use of Alternating DTPA and NTA Cation Exchange Flowsheets). Tanks 241-C-107, 241-C-108, 241-C-109, 241-C-111, and 241-C-112 received radioactive waste solutions from the Strontium Semiworks from 1961 through 1967. Tank 241-CX-70 did not receive any waste from the Strontium Semiworks (PPP-PRT-23177). The spent organic was discharged to a specific organic retention crib at the end of the program. Elsewhere documents stated that the organic following the Strontium Semiworks shutdown in 1967, the organic solvent was transferred to the REDOX for incorporation in the REDOX process. This latter seems somewhat suspect, since the REDOX Plant stopped processing fuel in January 1967, and was shut down in April, with final layaway completed by July. Based on the review of the operations that were conducted in the hot Semiworks from 1949 (Hot-Semiworks went operational Oct 1949; elsewhere documents stated that the Hot Semiworks was constructed 19511952 and went operational in November 1952) to November 1967 (Strontium Semiworks deactivated October-November 1967) there is no data that indicates that plutonium particles were discarded to the Hanford Tank Farms. Irradiated fuel was dissolved when the Hot Semiworks was conducting pilot plant demonstration of the REDOX and PUREX flowsheets. The radioactive liquid wastes that were generated in the pilot plant operations would have been of the same composition as those radioactive liquid wastes generated at REDOX and PUREX. In the case of the REDOX, waste was discarded to CX-70. The bulk of the sludge containing the preponderance of the plutonium was ultimately sent to tank 241-AW-105. The remaining sludge heel was drummed and transferred to the Hanford Site Central Waste Complex. The radioactive waste from the PUREX pilot plant studies was sent to the tanks 241-C-201 through 241-C-204. The radioactive liquid waste was later transferred to tanks 241-C-104, 241-C-109, and 241-C-106. The operations conducted at the then-called Strontium Semiworks to recover strontium-90, cerium-147, and americium-curium consisted of waste streams recovered from REDOX and PUREX. These waste streams and the processes used at the process plants to recover the raffinate streams that were transferred to the Strontium Semiworks, would have reduced the concentration of any plutonium to extremely low levels. This fact practically precludes the possibility of the existence of large plutonium oxide particles in the strontium radioactive waste stream. 5.1.3 Review of 222-S Laboratory waste collected in the 219-S Waste Handling Facility

The 222-S laboratory has transferred a small amount of waste to the Tank Farm. The report WRPS-1104536, Evaluation of Plutonium Content of 222-S Laboratory Waste Collected in the 219-S Waste Handling Facility, evaluated the amount of plutonium sent to tank farm from the laboratory. This report determined that all of the 222-S laboratory waste was sent to the REDOX facility prior to 1999, where it was an inconsequential (tens of grams per year) in comparison to the amount of plutonium already in the REDOX facility. After 1999, a maximum of 0.6 grams of Pu per transfer was transferred to the tank farm resulting in a yearly discharge of only about 10-20 grams. The report notes that the plutonium was in a form that was easily suspended;

Official Use Only 151

RPP-RPT-50941, Rev. 0

therefore it was light and fine. Thus, the amount of plutonium sent to the tank farm from the 222-S laboratory is inconsequential, and is in a form that is of no concern. 5.1.4 231-Z Filtrate Transfers to T-Plant

The aqueous acid filtrate from the plutonium peroxide/oxalate stream from 231-Z Facility (generally referred to as the byproduct stream) was recycled for recovery of plutonium. The available information indicates that virtually all the recycle material was processed at the 224T Concentration Building. The 231-Z Facility operated from January 1945 through January 1957. Plutonium-containing filtrate recycled to Concentration Facility was treated with 40% NaNO2 and finally with 4% KMnO4 until a permanent coloration appeared to react with excess hydrogen peroxide or oxalic acid and dissolve any plutonium peroxide/oxalate solids. All the various filtrate streams from first, second, etc., precipitations and washes were combined together and sent back for plutonium recovery. At the Concentration Building the plutonium-containing solution was introduced into the purification step along with the T Plant canyon plutonium nitrate product stream from the bismuth phosphate plutonium recovery operation. A summary of the steps in the Concentration Building for the recovery of a plutonium nitrate solution suitable for purification in the 231-Z Building are as follows: Bismuth Phosphate Cross-Over 1. Pu (IV) to Pu(VI) Oxidation The product solution was received from the canyon building. The Pu(IV) was oxidized to Pu (VI) using sodium dichromate at 45oC in the presence of NaBiO3. H3PO4 was added to the solution, which precipitated Zr and Nb fission products as phosphates, but left the Pu (VI) in solution. 2. Lanthanum Fluoride LaF3 Precipitation The solution from the Pu(VI) phosphate step was treated with hydrofluoric acid and La(NH4)2(NO3)5 and KMnO4 to precipitate La and rare earth fission products as the fluorides while leaving the Pu(VI) in solution. Oxalic acid was added to reduce Pu(VI) to Pu(IV) and hydrofluoric acid was added along with La(NH4)2(NO3)5 which precipitated plutonium (IV) fluoride and Lanthanum fluoride. 3. Plutonium Metathesis The PuF4 and LaF3 solids were treated with 50 wt% solution of KOH to precipitate the corresponding hydroxides. After washing solids and removal from the centrifuge the plutonium and lanthanum hydroxides were dissolved using 60 wt% nitric acid. The Official Use Only 152

RPP-RPT-50941, Rev. 0

plutonium nitrate product solution was transferred via a PR can to the Isolation Building (231-Z) at a nominal concentration of 20-40 g Pu/L nitric acid solution. Any plutonium from 231-Z filtrate cycle to the Concentrator Building would be indistinguishable from the normal plutonium operating losses from T Plant. 5.1.4.1 Plutonium in Recycle Solutions from 231-Z. The 231-Z facility handled all plutonium nitrate product that was recovered from the bismuth phosphate operations at the T and B Plants. It most likely also received the majority of the plutonium nitrate product from the REDOX Plant during 1952. The REDOX plutonium nitrate product would have met the 234-5 plutonium nitrate feed specification by 1953 (assumption) and have been transferred directly to the RMA Task I (batch operation) operation. As depicted in Table 33, T Plant and B Plants between 1945 and 1956 recovered approximately three metric tons of plutonium. This material would have been purified in the 231-Z Facility. In 1952 REDOX recovered about 0.46 metric tons of plutonium which could have been purified in 231-Z Building. A conservative is that the 231-Z facility handled 3,500 kg of plutonium during its operational lifetime. The plutonium peroxide flowsheet showed that the killed combined supernatant and wash (CT-1) contained 0.3 g plutonium/L in 63L of solution or 18.9 g of plutonium. The Feed F-10-P solution contained 375 g of plutonium. This represents 5% plutonium lost to the filtrate stream that is recycled to the T Plant Concentration building. If the 231-Z Facility handled 3,500 kg, then nominally, 175 kg of plutonium would have been transferred to the Concentration Building at T Plant. The bismuth phosphate flowsheet showed that in the Lanthanum fluoride Pu(VI) to Pu(IV) reduction and precipitation step the plutonium precipitation waste solution contained 0.7 g of Pu/L. The lanthanum fluoride by-product LaF3 precipitation product contained 346 g of Pu/batch. In the plutonium metathesis step, another 0.1 g of plutonium/L was lost in the plutonium metathesis 224T solution for a 345 g plutonium/batch content in the plutonium product metathesis slurry. In total, for a 346 g plutonium/batch feed to the 224 Concentration Building, 0.8 g/L of plutonium was lost in the tank waste. This translates into a 0.23% plutonium loss. Based on 175 kg of plutonium recycled to the T Concentration Building from the 231-Z facility 0.4 kg of 231-Z plutonium could have been transferred to tank farms indistinguishable from the normal 224T Concentration Building plutonium waste losses. The filtrate recycle streams at 231-Z were treated with oxidizing agents and heat to destroy excess hydrogen peroxide, oxalic acid, and entrained solid plutonium peroxide or plutonium oxalate. Plutonium peroxide is easily redissolved with HNO3 (above 3 molar). Plutonium peroxide is stable for several hours in dilute nitric acid (1M). After 24, hours the peroxide shows evidence of dissolution probably as a result of decomposition of the peroxide linkage in the molecule. In water, the peroxide is stable for 48 hours or longer in the absence of excess peroxide. In more concentrated acid, the peroxide slowly dissolves at room temperature.

Official Use Only 153

RPP-RPT-50941, Rev. 0

These processes and the dissolution of the plutonium hydroxide in 60% nitric acid during the plutonium metathesis step would suggest that negligible 231-Z recycled plutonium was in the form of plutonium oxide and/or large particles. The metathesis of plutonium oxalate is discussed in Appendix C Methathesis of Plutonium (IV) Oxalate and Plutonium (IV) fluorides in alkaline media to Form Pu)2.xH2O. Table 33. Pu Recovered at the Processing Plants (MT)
T-Plant CY Year 1945 1946 1947 1948 1949 1950 1951 1952 1953 1954 1955 1956 Total Calc Dissolving (tons) 334.00 291.00 306.00 388.00 393.00 424.00 560.00 295.00 246.00 1,011.00 1,266.00 50.00 Product (MT) 0.08 0.07 0.08 0.10 0.10 0.11 0.18 0.16 0.10 0.21 0.26 0.02 1.47 1.53 Total Pu produced 1945 1956 B-Plant Dissolving (tons) 302 573 345 398 356 460 508 118 3,060.00 Product (MT) 0.08 0.14 0.09 0.10 0.09 0.12 0.13 0.03 0.77 T&B Total 0.16 0.22 0.16 0.20 0.19 0.22 0.30 0.19 0.87 0.21 0.26 0.02 3.00 5.97 T&B Plants WHC-EP0793 0.06 0.21 0.14 0.14 0.18 0.28 0.39 0.23 0.12 0.46 0.47 0.02 2.70 5.48 2.98 2.78 822.13 1,430.60 1,579.30 2,425.58 0.41 0.74 0.86 0.97 0.46 0.70 0.72 0.90 0.1 Jan-Feb REDOX Dissolving (tons) Product (MT) WCH-EP0793

5.1.5

Liquid Waste Receipts at the 204-S and 204-AR Facilities

The 204-S and 204-AR facilities received radioactive and mixed chemical/radioactive liquid wastes from the Hanford 100, 200, 300 and 400 Areas and transferred these wastes to the tank farms. Equipment for unloading railroad tank cars and tanks was installed at the 204-S facility in 1957 (RHO-CD-1129, 204-S Facility Shutdown Plan). Wastes unloaded at the 204-S facility were transferred to underground tanks, primarily tank 241-S-107 and 241-U-107. The 204-S unloading equipment was not enclosed in a building and was shut down in 1981 after waste transfers to the single-shell tanks were discontinued and the new 204-AR facility was completed. The 204-AR facility transferred only to double-shell tank system and had the additional capability of adjusting the hydroxide and nitrite ion concentrations of the wastes. Although railroad tank cars and tanker trucks were fully enclosed in the 204-AR building, the initial hookup to the transfer piping, vent lines, weight factor instrumentation and sluice water was performed manually; therefore the solutions transferred were limited to liquid wastes with relatively low radioactivity. Wastes received at these facilities were generated at several Hanford facilities. The primary waste sources are:

Official Use Only 154

RPP-RPT-50941, Rev. 0

1. 100 Area reactor liquid wastes Reactor decontamination waste containing phosphate and complexing agents Ion exchange regeneration waste containing sulfate Sand filter backwash waste. 2. Liquid waste from 300 Area primarily laboratory wastes 3. Liquid waste from 400 Area from equipment decontamination and laboratory wastes. 4. T Plant equipment decontamination wastes 5. 222-S laboratory wastes. The 100 Area and 300 Area wastes constituted most of the waste volumes received. Typical radionuclide concentrations are given in RHO-CD-757, 1981, Safety Analysis Report 204-AR Waste Unloading Facility for the 100 and 300 Area wastes. The principal radionuclide hazard in 100 Area reactor wastes are gamma emitters: cobalt-60 at 0.15 Ci/mL was the primary concern. The typical total alpha activity of 300 Area wastes was given as 1.12E-02 Ci/mL for liquids and 3.3E-02 Ci/mL for entrained solids. These activities would equivalent to approximately 1.8E-04 g Pu/L and 5.3E-04 g Pu/L respectively if all of the alpha activity was assigned to 239Pu. The wastes received at the 204-S and 204-AR had low plutonium concentrations. The initial specification for the 204-AR facility was < 15 grams fissile content per rail car or tanker truck (FDM-T-290-00001, 204-AR-Rail Car Unloading Facility Facility Description Manual). In later years criticality prevention was switched to a concentrations basis; 0.01 gram fissile material per gallon (OSD-T-151-00008, Operating Specifications for 204-AR Waste Unloading Facility). Based on the sources for the transfers to tank farms, these wastes would not be a significant source of plutonium particles, including any large, dense particles.

5.2

PROCESSES THAT HAD PUO2 BUT DID NOT DISCHARGE IT TO TANK FARMS

The PUREX N Cell Plutonium Oxide Facility (Projects 175 and 175a) was constructed as part of PUREX restart effort and was placed in service in 1984. The facility operated only through 1991. Four liquid waste streams were routed to the PUREX solvent extraction system for recovery of plutonium losses (RHO-MA-116, PUREX Technical ManualAddendum I Plutonium Oxide Production and Rework Facility). No N Cell liquid wastes were routed directly to tank farms. A simplified diagram of the PUREX waste routings is provided in Figure 47. N Cell processing is described in detail in RHO-MA-116. A process flow diagram for N Cell is provided in drawing H-2-65483. The N Cell liquid waste streams routed to the PUREX solvent extraction system for plutonium recovery were:

Official Use Only 155

RPP-RPT-50941, Rev. 0

Tanks N15 and N16 which primarily contained filtrate from the plutonium oxalate vacuum drum filter which had been heat treated to destroy the oxalate ion (oxalate-killed). Tanks N21 and N22 which contained process condensates from the filtrate evaporator and the first stage calciner. Tank N30 which contained entrained liquid from the process vacuum system and sump liquids vacuum transferred from the wet gloveboxes. Tank N53 which contains plutonium nitrate solutions generated by dissolving of plutonium oxide or plutonium oxalate solids for recycle to the PUREX solvent extraction process.

These waste streams are discussed individually in the following sections. Tanks N15 and N16 Oxalate-Killed Filtrate The primary source for these waste collection tanks was the filtrate from the vacuum drum filter. Liquids drawn through the woven polyethylene filter cloth were routed to the TK-N11 evaporator feed tank. The porosity of the filter cloth was 120 microns (RHO-MA-116), so the filtrate would have contained small plutonium oxalate particles well as soluble plutonium oxalate and oxalic acid. The filtrate was routed to the TK-N13 filtrate concentrator for concentration to > 9 molar nitric acid. The high nitric acid concentration and boiling temperature destroyed the oxalate ion and dissolved plutonium oxalate particles (SD-CP-ES-043, The Destruction of Oxalic Acid by Nitric Acid at the Boiling Point) to form soluble plutonium nitrate. The evaporator overflow was routed to one of two holding tanks (TK-N15 or TK-N16) where the temperature was maintained near boiling during and after filling of the 90 liter batches (RHO-MA-116). This assured complete destruction of the oxalate so that all plutonium will be present as soluble plutonium nitrate. The batch was sampled after digestion and analyzed for plutonium concentration, nitric acid, oxalate ion and visual observation for solid particles (< 1% is the typical avisual detection limit) [FSS-P-080-00002, PUREX Sample Schedule]. All of these analyses were required for critical mass control. Batches which contained a high oxalate concentration or particles were recycled to TK-N11 for reprocessing to complete destruction of the oxalate ion. After confirmation of these parameters, the solution was pumped to PUREX solvent extraction feed makeup tank (TK-E6) for normal solvent extraction recovery of the plutonium. Solutions from tanks N21, N22 (process condensates) or N30 (wet glovebox sumps) having oxalate ion concentrations of 0.1 molar (FSS-P-080-00002) were also processed through the TK-N13 filtrate concentrator for oxalate destruction and returned to the PUREX solvent extraction system for plutonium recovery through tanks N15 and N16. The N15/N16 pump outlet for batch transfers to TK-E6 was filtered, but only to 100 mesh (~150 microns) per RHO-MA-116. The potential for returning plutonium oxalate particles to the PUREX solvent extraction system was low due the extensive processing to destroy the oxalate ion and verification by sample analysis. Official Use Only 156

RPP-RPT-50941, Rev. 0

Tanks N21 and N22 Process Condensates The N21 and N22 tanks collected process condensates from the TK-N13 filtrate concentrator and the CA-N9 calciner. Normally, these condensates contained extremely low levels of plutonium and no particulates. The calciner off-gas sintered-metal filters had a mean pore size of 5 microns and would have effectively removed 100% of the particles > 1 micron (RHO-MA-116). The primary off-standard condition which could have introduced plutonium particles into these tanks was failure of the sintered metal filters in the calciner off-gas system. N Cell operators and supervisors polled did not recall that this failure occurring during operation of N Cell. Each batch of process condensate was sampled in tank N21 or N22 and analyzed for plutonium, oxalate ion and a visual observation. Process condensates batches having oxalate concentrations of 0.1 molar oxalate ion were transferred to tank N11for oxalate ion destruction. The process condensates were normally transferred to TK-F10 in the solvent extraction back-cycle waste system. Tank F10 was the feed tank for the E-H4 concentrator which operates at atmospheric boiling and a nitric acid concentration of ~ 8 molar (RHO-RE-FL-2, PUREX Flowsheet Reprocessing N Reactor 6% Fuel). Trace amounts of small oxalate particles would have likely dissolved during concentration in the E-H4 concentrator. The E-H4 evaporator concentrate was recycled to the HA column, with the aqueous waste routed to the HLW waste concentrator, sugar denitration, waste neutralization and eventually to the aging waste tanks in tank farms. The process condensates transferred from tanks N21 and N22 during normal operations presented a very low risk of transferring > 10 micron plutonium oxide or plutonium oxalate particles to the PUREX solvent extraction system and hence to tanks farms. Tank N30 Glovebox Sump Waste The N30 vacuum tank served a collection tank for entrained liquid in the vacuum header and for liquids in the wet glovebox sumps. The wet gloveboxes included gloveboxes N1A/N2A, N1B, N2B, N2C and N3. Liquids spilled in the N6 glovebox were handled internally in that glovebox and were not routed to tank N30. With the exception of the N3 glovebox, the liquids collected in the sumps would have contained plutonium nitrate solutions with little potential for either plutonium oxalate or plutonium oxide particles. The N3 glovebox contained the plutonium oxalate precipitator/vacuum drum filter and could have contained plutonium oxalate particles. Solid wastes from the N3 glovebox were packaged in polyethylene jars and transferred to the N6 Plutonium Rework glovebox for recovery of the plutonium. Residual plutonium oxalate particles would be transferred to tank N30 during wash down of the N3 glovebox with 1.2 molar or 12 molar nitric acid. Wash downs of the N3 glove box were not a flowsheet operation, but were not uncommon. Tank N30 was equipped with a 25 micron filter in the pump outlet (RHO-MA-116). The contents of tank N30 would be recirculated through this filter prior to sampling of a batch. The samples were analyzed for plutonium and oxalate ion concentrations and visually inspected prior to releasing the solution for transfer to TK-E6 via TK-L11 (FSS-P-080-00002). Some intermediate size oxalate particles (< 25 micron) could have been included in these transfers if Official Use Only 157

RPP-RPT-50941, Rev. 0

the batch included wash downs of the N3 glovebox were included in the batch. Batches having oxalate ion concentrations of 0.1 molar (FSS-P-080-00002) would have been processed through the TK-N13 filtrate concentrator for oxalate destruction and returned to the PUREX solvent extraction system through tanks N15 and N16. The tank N30 transfers to the PUREX solvent extraction system could have contained plutonium oxalate particles of up to 25 microns when wash downs of the N3 glovebox occurred. The overflow capacity of tank N30 was 24 liters. Assuming that the oxalate ion concentration was at the Criticality Prevention Specification limit of 0.1 molar, a 20 L transfer from tank N30 and 2,000 gallon minimum volume in TK-E6 required for rework transfers (RHO-MA-116), the resulting oxalate ion concentration in TK-E6 would have been 2.6 E-04 molar. A portion of any plutonium oxalate recycled through tank N30 would have been dissolved in the nitric acid used for wash down of the N3 glovebox or by disappropriation of the plutonium oxalate to uranium oxalate and plutonium nitrate in the feed makeup tank E6 and the H1 feed tank to the solvent extraction system. Undissolved solids would likely have collected in high-level waste concentrator feed tank F7 which was operated without agitation during processing to prevent the introduction of an organic phase into the E-F6 concentrator. At shut down of the solvent extraction system, any accumulated organic would be removed from tank F7 and the residual liquids would have been processed through the waste treatment system, neutralized and disposed to the aging wastes tanks. Tank N53 Rework Glovebox Product The N6 glovebox contained all equipment used for rework of plutonium oxide or plutonium oxalate solid wastes. A detailed process flow diagram for the N6 glovebox is provided in drawing H-2-75643, Process Flow Diagram. This glovebox was isolated from other N Cell gloveboxes with the exception of a secondary criticality drain overflow from the N7 glovebox installed in 1987 (H-2-75645, Engineering Flow Diagram N-6 Glove Box & Pipe Chase); there was no process transfer piping to or from other N Cell gloveboxes or equipment. Solid wastes containing plutonium were packaged in product cans or polyethylene jars and bagged into the glovebox for processing. Plutonium oxide or plutonium oxalate was dissolved in one of two dissolvers and underwent two stages of filtration prior to transfer to tank N53. The first stage of filtration was a pan filter utilizing a Teflon media with 5 micron retention (RHO-MA-116). The filtered solutions must then pass through a 1-micron retention disc filter during transfer to tank N53 (RHO-MA-116). The original filters were replaced with ones having a larger solids volume capacity during early operation of the N6 rework glovebox. Drawing H-2-75657, Rev. 3 shows the final configuration where 3-micron FilteriteTM canister filters are installed in both in the N52-1 position (replacing the pan filter) and in the N52/53-1 position (replacing the parallel disc filters). Upon collection of a batch of plutonium nitrate solution in tank N53, recirculation and sampling of the batch was performed. The sampling schedule (FSS-P-080-00002) specified analyses for plutonium, nitric acid and oxalate ion concentrations and visual inspection. Solutions not meeting specifications would have been re-filtered or reprocessed through the dissolvers.

Official Use Only 158

RPP-RPT-50941, Rev. 0

Plutonium nitrate solution meeting the criticality specifications was then recycled the tank E6 solvent extraction feed makeup tank via TK-L11. Given the double-filtration and sample verification, there was very limited possibility that plutonium particles exceeding 10 microns were recycled to PUREX solvent extraction system via tank N53 and hence to tank farms via the high-level waste stream. The only method for recycling of plutonium oxide from the N Cell dry gloveboxes or plutonium oxide product containers is dissolution in the N6 glovebox. Actions at Plant Shutdown PUREX standby and deactivation activities are described in HNF-SP-1147, PUREX/UO3 Facilities Deactivation Lessons Learned History. PUREX was officially placed in standby in October 1990 until an options study and revised Environmental Impact Statement were completed. N Cell and the PR Room underwent a major cleanout campaign in 1991. Plutonium oxide powder was removed from N Cell gloveboxes, reworked through the N6 rework glovebox and recycled to tank E6 by blending with recycled uranium nitrate solution. It was expected that the PUREX solvent extraction system would be restated to recover these product materials. Operations of N Cell were still subject to criticality prevention specifications while PUREX was in standby mode. Note that dissolution in the N6 glovebox was the only method for recycling of plutonium oxide from the N Cell dry gloveboxes or plutonium oxide product containers. A final shutdown order for the PUREX Plant was received on December 22, 1992 (HNF-SP-1147). The shutdown order precluded restart of any equipment and the uranium and plutonium solutions stored in tanks D5 and E6 (approximately 9kg Pu and 5 metric tons uranium) were eventually discarded to tank 241-AW-105 in 1995. No further transfers were made between N Cell and the PUREX solvent extractions system after receipt of the final shutdown order. N Cell scrap materials containing recoverable plutonium were packaged and shipped to PFP. Residual plutonium liquids were also transported to PFP. All N Cell equipment was eventually removed from the gloveboxes and disposed of as transuranic waste as part of the facility deactivation. In summary, there was limited potential for transferring plutonium oxide particles of > 10 micron diameter to the PUREX solvent extraction system and hence to the waste tanks (tanks 241-AZ-101 and 241-AZ-102). All plutonium oxide was removed from the process in product containers, as packaged solid waste or packaged in product cans for recycle to the N6 rework glovebox. The liquid plutonium nitrate product from the N6 glovebox was double filtered and the product sampled and analyzed prior to transfer to the PUREX solvent extraction system for plutonium recovery. There was a more credible potential for transferring plutonium oxalate particles of up to 25 microns to the solvent extraction system via the collection of N3 glovebox wash downs in tank N30. Any plutonium oxalate particles reaching the aging waste storage tanks (tanks 241-AZ-101 and 241-AZ-102) would be expected to decompose into light fine particulate [see Metathesis of Plutonium(IV) Oxalate and Plutonium(III)/(IV) Fluorides in Alkaline Media to Form PuO2xH2O report in Appendix C of this document]. Official Use Only 159

RPP-RPT-50941, Rev. 0

Figure 47. Simplified Liquid Waste Diagram

5.3 5.3.1

PROCESSES THAT MAY HAVE DISCHARGED PUO2 TO TANK FARMS REDOX and PUREX Processing of PFP Nitrate Solutions

Except for the time before the RECUPLEX Facility came online, the 234-5 Building (Z Plant) was essentially self-sufficient when it came to reprocessing plutonium residues and solutions from the different Task operations. The 234-5 Building received aqueous plutonium solutions from REDOX and PUREX and did not tend to ship plutonium containing solutions to REDOX and PUREX. The significant exception was between 1962 and 1964, after the RECUPLEX accident. Detailed records on the quantity and form of plutonium that was shipped to REDOX and PUREX from the 234-5 Building over the operating lifetimes of the facilities were not located. What was located were the Chemical Process Statistics Books I & II (handwritten) which provided nuclear material information on the plant operations including transfer and receipt of plutonium between 234-5, and PUREX, and between 234-5 and REDOX from 1957 through 1972, (HW63089, Chemical Processing Statistics Book I; HW-63090, Chemical Processing Statistics Book II). A notation indicated that the data came from various Accountability Material Official Use Only 160

RPP-RPT-50941, Rev. 0

Balance and Net Production Reports. The data did not specify the composition, physical state, properties, only that it was plutonium. Table 34 shows the quantity of plutonium transferred between Z Plant and REDOX and PUREX. The only weekly or monthly reports and other correspondence that described the quantity of plutonium transferred from 234-5 to PUREX and REDOX was for the period between 1962 and 1964, and 1972 during the period at 234-5 that RECUPLEX was down and the startup of PRF and then when some unirradiated PRTR fuel scrap that had been stored at 234-5 Building was processed at PUREX. Columns two to four of Table 34 show the kilograms of plutonium sent from PUREX and REDOX to 234-5 Facility based on McDonalds report from the shipping and receiving plants. The next two columns show the amount of plutonium received by PUREX and REDOX based on those plants records. The last six columns are the quantities of plutonium shipped to PUREX and REDOX based on PFP data. The records are for plutonium shipped from the Product Recovery process and the Plutonium Processing operation. There is a reasonable correlation between what was shipped and received at the various Plants. Note that PFP is processing about 4,000+ kg of plutonium per year. No information was found that would indicate that the plutonium shipped between 234-5 and PUREX and REDOX was anything other than plutonium nitrate solution. Based upon the lack of documentation as compared to the 19621964 campaigns, these returns are expected to be plutonium nitrate solutions that were high in fission products incompatible with the PFP process. The plutonium shipped from PUREX and REDOX was plutonium nitrate solution to be used as feed for Task I, oxalate precipitation and those plants did not produce any plutonium product other than plutonium nitrate solution. The PFP plutonium returns also are almost certainly plutonium nitrate solutions with the exception of the unirradiated plutonium scrap shipped to PUREX for dissolution and plutonium recovery in 1972 time frame. A 1961 report on the processing of plutonium scrap at Hanford (HW-70361, The Processing of Pu Scrap at Hanford), stated The material most adaptable for recycling to PUREX and REDOX Plants are relatively pure plutonium solutions derived from the dissolution of skulls, reject buttons, and similar materials, and filtrate from the plutonium oxalate precipitation process. Other recycle streams (this included plutonium solid scrap streams) are incompatible with the separation processes or are considered too variable in nature to risk of process upsets which might result in poor decontamination or in plutonium being discarded to waste. Dissolved metal and skull solutions could be made compatible with the extraction processes at REDOX and PUREX, and with the ion exchange tail end process at PUREX. One monthly report (HW-27288, Chemical Process Department Monthly Report) mentioned tests to demonstrate that the filtrate could be handled in REDOX and that it was compatible with the nominal REDOX IAF plutonium stream. Some 1965 Chemical Processing Department Monthly Reports (RL-SEP-706, RL-SEP-654, RL-SEP-837) stated: Occasional (bi-monthly) purging of uranium from the Reclamation Facility system by routing either concentrated supernatants or plutonium nitrate product solution to PUREX or REDOX for partitioning.

Official Use Only 161

RPP-RPT-50941, Rev. 0

Uranium contamination in the plutonium nitrate was a factor of 8 above specification. The uranium was subsequently removed at Z Plant and returned to REDOX for reprocessing. Laboratory tests indicate the feasibility of reprocessing Z Plant supernate from the oxalate strike (RCR) in PUREX without adverse effect on waste losses or column performance. Such recycle is desirable to remove uranium from the Z-Plant process system.

The plutonium associated with plutonium processing removals appears to be plutonium nitrate filtrate from Task I. This is the same designation that PFP used to record the PFP receipts from the processing plants. These transfers increase during the 19621964 period. The large plutonium transfer in 1964 is discussed later in this section. As stated previously in this report the filtrate solutions were treated and heated to destroy excess oxalic acid and dissolve any solids. The waste stream from filtrate processing is not considered a source of large plutonium particulates to the tank farms. The plutonium transferred to PUREX and REDOX from PFP product recovery would seem to be plutonium nitrate from the dissolution of skulls and reject buttons, but this cant be confirmed. It is a relatively minor stream in comparison to the plutonium processing removal stream and should have the same characteristics of plutonium nitrate from the dissolution of plutonium oxide which is discussed later in this section Table 34. Plutonium Transfers from Z Plant to REDOX and PUREX
PFP Receipts PUREX Plutonium Processing Pu(kg) PFP Receipts REDOX Plutonium Processing Pu(Kg) PFP Product Recovery 234-5 to 234-5 to Removals PUREX REDOX to PUREX Pu (Kg) 36.48 44.78 523 430 685 509 733 1136 963 1040 131 67.52 2.83 39.16 106.79 42.67 227.03 76.65 57.06 25.58 0 0 0.53 0.17 27.79 14.2 0 2.56 0.74 69.05 73.74 21.1 35.24 3.74 Pu (kg) 4.33 17.27 39.72 2.07 15.41 0.4 0 0 16.74 27.82 0 0 115.89 69.5 226.54 19.85 40.48 29.59 PFP Plutonium Processing Removals Total 234-5 to PUREX to PUREX Pu(kg) Pu(KG) 4.33 34.01 67.54 2.07 15.41 116.29 69.5 226.54 19.85 40.48 29.59 PFP PFP Product Plutonium Recovery Processing Total Removals Removals to 234-5 to REDOX REDOX to REDOX Pu (kg) 0.11 2.46 0 0 0 0 0 8.48 0 0 0 69.03 82.4 21.36 35.46 3.74 0 Pu(kg) Pu (Kg) 0.11 10.94 0 0 0 69.03 82.4 21.36 35.46 3.74 0

PUREX to 234-5 Pu(Kg) 776 2,535 3,005 3,687 3,494 3,559 3,256 3,743 3,233 2,564 3,411 3,207 2,091 643 1,978

REDOX to234-5 Pu(Kg) 471 675

2806 3553 3489 3552 3466 3646 4366 2332 3301 3208 2091 643 1977

529 437 708 564 733 1136 983 1001

Official Use Only 162

RPP-RPT-50941, Rev. 0

In April 1962, the criticality incident at Z Plant shut down the RECUPLEX Facility. Until the Plutonium Reclamation Facility (PRF) came online in March of 1964, the Z Plant capability to recover plutonium from line generated residues was severely restricted. This was a particularly acute problem with the plutonium- containing filtrate stream from the plutonium oxalate precipitation process. As an interim fix to the filtrate handling problem, the RMC Button Line Filtrate Handling Facility was installed in October 1962 to recover plutonium from the oxalate precipitation filtrate stream. For the period from April 1962 into 1964 literally thousands of liters of filtrate and dissolver solution containing about four hundred kilograms of recoverable plutonium were transferred to the REDOX and PUREX Plants. Five hundred specially fabricated batch cans, similar to PR cans, were fabricated for this effort. The containers were referred to as SN (supernatant) cans or emergency PR cans. The plutonium in the solutions was predominately soluble plutonium (IV) nitrate. Before the plutonium containing oxalate filtrate left the 234-5 Building the stream was treated with KMnO4 to destroy excess oxalic acid and dissolve any plutonium oxalate particles, and NaNO2 to dissolve any MnO2. Reports describing the shipment of plutonium containing solutions to the PUREX process included the following assumptions/conditions, (HW-73579, Special Plutonium Recycle at PUREX; HW-80168, Campaign Rework of Z-Plant Material at PUREX); The supernatants contain relatively small quantities of plutonium (on the order of 0.5 g/L). Solutions will be shipped as treated (with KMnO4 and NaNO2) nitric acid solutions in PR cans. Plutonium solids with a particle size greater than 40 microns shall not be introduced into the PUREX Plant. Rework solution should be transferred through a less than 40 micron filter to prevent plugging of PR Room lines and valves. The solutions will be introduced into PUREX at the normal rework position, the HAF Make-up Tank E-6

Z Plant Supernatant solutions were being received by the processing plants by May 1962, (HW-71990, Process Summary of PUREX Plant Operations January 1962 through December 1962). As of May 19, 1962, 300 cans of Z Plant supernatant were processed and about 6,700 g of plutonium was processed (elsewhere in the report it indicated that the 300 SN cans contained about 11.5 kg of plutonium). In May 1962 a new plutonium recycle unloading and monitoring system was installed in the 233-S Plutonium Concentration Building which provided facilities for receiving Z Plant Task I filtrate, (HW-73884, Chemical Processing Department Monthly Report May 1962). The solutions would be unloaded to Tank L-22 before being transferred to the REDOX process via the recycle tank. In July 1962, REDOX reported that the ion exchange unit was operated to process approximately 650 cans of Z Plant filtrate and 75 cans were processed via the H-4 metal solution oxidation tank, (HW-74505, Chemical Processing Department Monthly Report July 1962). Also in July PUREX processed 1,030 cans of Z Plant filtrate. In August 1962, an additional 505 recycle containers were processed at REDOX, (HW-74804, Chemical Processing Department Monthly Report August 1962). In

Official Use Only 163

RPP-RPT-50941, Rev. 0

October 1962, REDOX processed another 480 cans of Task I supernatant from Z Plant, (HW-75470, Chemical Processing Department Monthly Report October 1962). By December 1962, the 234-5 Building RMC Button Line Filtrate Handling Facility handled all the filtrate generated by the button line plus miscellaneous waste solutions (HW-76054, Chemical Processing Department Monthly Report December 1962). This appeared to be the end of the major thrust of the Z Plant filtrate processing at the REDOX and PUREX plants. Based on the available data, at least 3,040 cans of supernatant were processed through PUREX and REDOX. This value seems low as explained below. A calculation made based on probable plutonium processing capability during this time calculated a somewhat larger number of filtrate cans were shipped. The quantity of plutonium transferred was estimated. [Private communication with Joseph Teal] Site production during fiscal years 1962 and 1963 was approximately 4.1 metric tons per year. For this estimate, seven months of production involving shipment of filtrate to the primary plants was assumed: 7 months/12 months/year x 4,100 kilograms/year = 2,392 kilograms of plutonium production; Approximate stoichiometric filtrate quantity generated was ~ 30 liters per kilogram of plutonium x 25% excess = 2,392 kg x 30 liters/kg x 1.25 = 89,700 liters of filtrate transported.

Although there were rather wide fluctuations in filtrate concentration due to equipment and chemical reaction problems, 2.3 grams of plutonium/liter of filtrate was considered a reasonable approximation for the plutonium concentration in the filtrate. This value was derived by averaging twenty-five weekly observations recorded in HW-72224 RD, during the time interval of interest. Therefore, 89,700 liters x 2.3 grams plutonium per liter = ~ 103 kilograms to REDOX and PUREX during the seven month interval. And 89,700 liters loaded at nine liters per batch can = 9,967 cans transferred over 210 days = 47 cans per day. The cans were trucked fifteen per load in a closed van truck at about one load per shift. This value is in line with the value in Table 34 for 1962, but the values for 19621964 also include plutonium nitrate solution from the dissolution of plutonium oxide, which is discussed below. As stated earlier in this section and elsewhere in this document, the oxalate filtrate plutonium stream was treated to dissolve any plutonium solids and destroy any residue oxalic acid before it was processed at RECUPLEX, PRF, 224 T, PUREX, or REDOX. This waste stream is not considered to have contributed large particulate PuO2 to the waste tanks. In addition to recovering plutonium from Z plant filtrate operation, REDOX and PUREX also processed plutonium nitrate solution from the dissolution of plutonium oxide in laboratory-type dissolvers located in Rooms 149 and 161, and Hood 42 in Z Plant.

Official Use Only 164

RPP-RPT-50941, Rev. 0

A December 1963 report provided details on the solutions that were candidate for PUREX processing, (HW-80168, Campaign Rework of Z-Plant Material at PUREX). Two Z-Plant process streams (MR and EPT) containing about 200Kg of plutonium were to be shipped to PUREX (McDonald showed that 227 kg of plutonium were transferred from Z Plant to PUREX in 1964). The EPT solution was the product of the RMC Button Line Filtrate Handling Facility which was installed in October 1962. The solution was expected to contain about 30-60 g Pu/L in a 0.7 to 1.3 M nitric acid medium. The EPT solutions could also contain a trace of resin fines. The MR rework solution was the product of plutonium scrap dissolution. The MR solution would vary from 40 to 150 g of plutonium/L in 10-15 M HNO3, 0.1-0.3 M F- and 0.1 to 0.3 M Al+3 media. The rework solution could contain an appreciable quantity of post-precipitated aluminum nitrate crystals. The solutions were normally not expected to contain organic phases, plutonium solids, or other materials deleterious to the PUREX process, although some insoluble plutonium and other substances had been noted. A review of Z Plant Weekly Reports for 1963 and 1964 provided data on the plutonium nitrate dissolver solutions. Beginning in July 1963, the weekly reports differentiated between the dissolution of unstabilized (alpha phase) plutonium and stabilized (delta phase) plutonium dissolution and the fact that delta phase plutonium was sent to REDOX and PUREX. Then in September, it was reported that only the dissolvers in Hood42 were to be used for delta phase plutonium dissolution. The alpha phase plutonium nitrate was suitable for recycle with button line feed. (HW-76162 RD, Z Plant Weekly Report Task I-II and Incinerator from January 1, 1963 to December 31, 1963; HW-80356 RD, Z Plant Weekly Report Task I-II and Recovery Operation from January 1, 1964 to December 31, 1964). Table 35 provides data on the amount of plutonium dissolved, and when identified, which processing plant received the solution. Chemical Processing Monthly reports for the 19621964 time frame reported similar statistics. The July 1962 report (HW-74505) reported that the plutonium was being recycled through the 233-S Ion Exchange unit and via recycle to H-4. The September 1963 Monthly Report (HW-79097) reported that REDOX processed 49 Kg of plutonium in stabilized plutonium recycle solution through the plutonium decontamination cycle. Based on the Chemical Processing Monthly Reports, about 74 Kg of plutonium was recycled through REDOX during 1963. Table 35. Transfers of Z Plant Dissolver Solutions to REDOX and PUREX
Date (Weekly Report) Plutonium (Kg) Rooms 149 & 161 March 31, 1963 April 7, 1963 April 15, 1963 April 29, 1963 May 6, 1963 May 20, 1963 May 27, 1963 11 44 41 46.8 38 19 16 8.7 3.7 10.6 10 Hood-42 Plant

Official Use Only 165

RPP-RPT-50941, Rev. 0

Table 35. Transfers of Z Plant Dissolver Solutions to REDOX and PUREX


Date (Weekly Report) Plutonium (Kg) Rooms 149 & 161 June 3, 1963 June 9, 1963 June 17, 1963 June 24, 1963 July 1, 1963 July 8, 1963 July 15, 1963 July 22, 1963 July 29, 1963 September 9, 1963 September 16, 1963 September 23, 1963 September 30, 1963 October 7, 1963 October 14, 1963 October 21, 1963 October 28,1963 November 4, 1963 November 11,1963 November 18, 1963 December 10, 1963 December 16, 1963 December 2, 1963 December 29, 1963 January 4, 1964 January 13, 1964 January 20, 1964 January 27, 1964 February 10, 1964 February 24, 1964 March 9, 1964 March 29, 1964 45.7 36.8 37.8 22.2 38.9 30.3 23 31.5 12.2 38.6 16.4 17.5 34 53.9 10 & 21 38.4 20.5 51.9 51.4 55.2 1 21.5 17.8 3.6 PUREX 42.5 kg to PUREX 56.2 kg to PUREX 10.8 2.8 16.5 17 10.4 7.0 23.5 10.9 & 5.5 20.1 15.2 & 2.2 14 0.3 21.5 14 21.3 30.6 27.6 Accepted by both plants Hood-42 only for REDOX 8.9 8.7 0.5 0.98 for REDOX REDOX (glass dissolver Room 149) Hood-42 Plant

Official Use Only 166

RPP-RPT-50941, Rev. 0

Table 35. Transfers of Z Plant Dissolver Solutions to REDOX and PUREX


Date (Weekly Report) Plutonium (Kg) Rooms 149 & 161 April 12, 1964 April 27, 1964 May 11, 1964 May 25, 1964 June 8, 1964 Approximate Total Plutonium Dissolver Solution Processed 37.4 3.5 18.8 49.7 28.1 PRF coming on line All dissolver solution to PRF To Button line Feed 130.9kg, To PUREX and REDOX 310.48 kg Hood-42 Plant

Route of PFP Recycle Solutions through the REDOX Plant Figure 48 illustrates the route of PFP filtrate solutions or dissolver solutions as they were reworked through the REDOX plant in the late fifties and early sixties. As shown by Figure 48, recycle solution (potentially containing Pu oxalate or oxide solids) was transported by truck to the REDOX load-in station in the 233-S facility. Recycle solution was vacuum-transferred from RC or PR cans to tank L-22. From L-22 it was either directly routed (without filtration) to solvent extraction purification or reworked via the moving bed ion exchange column (L-18). On Figure 48 the routing of recycle solution and any contained Pu solids is shown as a bold process stream line. In the option to rework solutions via the H-4 tank and solvent extraction, the low concentration of soluble plutonium in the filtrate was extracted into Hexone solvent in the HA column (along with virgin plutonium from current Hanford reactor fuel dissolution). After being stripped from the hexone in the HC column, both plutonium sources were processed through two additional cycles of solvent extraction and a final purification in the L-18 Moving Bed Ion Exchange contactor. Plutonium solids which may have been present in recycle solutions from PFP would likely have been carried in the HA column raffinate stream, through the D-12 waste concentrator and on to disposal in the S or SX tank farm. In all, these solids (if present) would have had to pass through nine tanks, one packed solvent extraction column, and one evaporator/concentrator before finding their way to the tank farm. Although intuition would argue that a fraction of these heavy particles should have been retained by settling in the plant vessels, no such separation factor has been assumed in estimating PuO2 particle additions to underground receiver tanks (see table REDOX/PUREX/Other in the summary). In the option to rework recycle solutions via the ion exchange system, any solids would have had to find their way through two additional concentrators, the L-18 contactor (where solids would likely be retained on the resin bed), and the 1-S solvent extraction column and not likely be discharged to the tank farm.

Official Use Only 167

RPP-RPT-50941, Rev. 0

Tanks Receiving PFP/REDOX Recycle Solutions During the period, July through September, 1963, when PFP dissolver solutions were being recycled to the REDOX plant (see Table 35), the REDOX high-level waste was being routed to underground tank S-108 (WHC-SD-WM-TI-614). Table 35 also contains data to suggest that a total of 310.5 kg Pu (from PFP dissolver solutions) was recycled to either the PUREX and REDOX plants, and of this recycle quantity, roughly 33% was routed to the REDOX plant while 67% was routed to the PUREX plant. Thus, it is estimated that 102 kg Pu in PFP dissolver solution from oxide from burned metal may have been recycled to the REDOX plant. Coupling this REDOX recycle value (102 kg Pu) with the assumed one percent PuO2 that could have remained un-dissolved, leads to the prediction that approximately 1 kg of Pu may have been transferred to tank S-108 in the form of PuO2 particles. Since the REDOX process applied no filtration step to the PFP dissolver recycle stream, the size of these PuO2 particles could have been as large as 100 microns, with essentially 100% of this quantity being greater than 10 microns in size. This is based on PSD data showing that oxide from burnt metal typically had micron size of 40-100 micron. Route of PFP Recycle Solutions through the PUREX Plant The reworking of PFP recycle solutions through the PUREX plant is depicted in Figure 49. RC or PR cans of solution (potentially containing Pu oxalate or oxide particles) were transported by truck to the PUREX plant where they were loaded into the process at either of two load-in stations: 1) a hood located in the sample gallery, or 2) the normal PR hood containing tank L-11. From these two load-in stations recycle solution was routed to the process feed make-up tank, E-6. In the E-6 tank soluble plutonium in the recycle solution was combined with normal feed from the dissolution of Hanford metal fuels and fed via the H-1 tank to the HA solvent extraction pulse column. There, plutonium was extracted into the Tri-Butyl Phosphate/normal paraffin hydrocarbon solvent stream, separated from uranium in the 1B column, further purified in the second plutonium extraction cycle and finally processed through the N-Cell moving bed ion exchange system. However, any solid Pu oxalate or oxide particles potentially present in the HA column feed stream would have reported to the HAW raffinate and then travelled through the PUREX highlevel waste treatment cell before being routed to an underground tank. From load-in to discharge to an underground tank, any particles would have had to pass through seven tanks (one, an unagitated settling tank), the HA column extraction section and the F-6 evaporator. Tanks Receiving PFP/PUREX Recycle Solutions During the first quarter 1964 when PFP dissolver solutions were being recycled to the PUREX plant (see Table 35), the PUREX high-level waste (1WW) was being routed to underground tank A-105 (WHC-SD-WM-TI-615).

Official Use Only 168

RPP-RPT-50941, Rev. 0

As discussed above, data given in Table 35 suggests that of the total dissolver solution recycled to the REDOX + PUREX plants, 67% (or 208 kg Pu) was routed to the PUREX plant. Coupling this PUREX recycle quantity (208 kg Pu) with the assumed one percent PuO2 that could have remained un-dissolved, leads to the prediction that roughly 2 kg Pu may have entered the PUREX plant in the form of PuO2 particles. During this operating period recycle streams were passed through a 40 micron filter. Oxide from metal burning typically had a particle size of 40 to 100 micron and it is estimated that less than 20% made it past the filter. Information on the filter media was unavailable to indicate a capture percentage, so 20% was assumed. Thus, during this operating period, the quantity of PuO2 particles that may have been transferred to the PUREX high-level waste receiver tank (A-105) is conservatively estimated at 20% of 2kg equaling 400 g Pu. It is also estimated that 100% of these PuO2 particles were over 10 microns in size, based on PSD data for oxide metal oxidation that typically had a micron size of 40-100 micron after dissolution. Figure 48. PFP Filtrate Recycle to REDOX Plant
May 1962 through Sept. 1963 - Running Flow Sheet #7 (RL-Sep-243)
Pu Finishing Plant
Button Line Filtrate KMnO4 NaNO2

233-S Facility
XAW Rework Thru Ion Exch. XAF TK L-22 Rework Thru Solvent Exch. L-12 Feed Concentrator L-18 Moving Bed Ion Exch. L-13 Product Concentrator XAP Pu Nitrate To PFP

Oxalate Kill Tank


Head Tank

RC Can Load-In Hood

Filter

Dissolved U, Pu

TK H4 Oxidizer HAF H A

Pu Intermediate Product (2AP)

RC Can Load-Out

Hanford U Metal Fuel

U Product

TK-F8

1 S

HL Waste To Tank Farm

Back Cycle Salt (2DW)

D-14 Concentrator

D-12 Concentrator

Multi Purpose Dissolver

REDOX Solvent Extr. System (Simplified)


TKH10 Cladding Waste To Tank Farm

Official Use Only 169

RPP-RPT-50941, Rev. 0

Figure 49. Filtrate Recycle to PUREX Plant


Pu Finishing Plant
Button Line Filtrate KMnO4 NaNO2

May 1962 through Jan. 1964


N-Cell

PUREX Plant
Pu Nitrate To PFP

RC Load-In Hood (Sample Gallery)

Moving Bed Ion Exch.

Pu Intermediate Product (2AP) Oxalate Kill Tank


Head Tank

H A

PR Load-In Hood
Filter TK E6 TK H-1

C o l u m n

3WB Back Cycle Waste

E-H4 Concentrator NaOH

U Product

RC Can Load-Out

TK L-11
TK F7
Dissolved U, Pu

E-F6 Waste Concentrator

F-26

F-15 Sugar Denitration

F-16 Waste Neutralizer

Hanford U Metal Fuel

HL Waste to Aging Tanks

TKS D-3, D-4, D-5

Simplified PUREX Solvent Extraction System


Cladding Waste To Tank Farm

POT Dissolver

TKS E-3 E-5

5.3.2

Miscellaneous Fuel Campaigns

Beginning in 1957 the REDOX and PUREX plants began processing small campaigns of miscellaneous or specialty fuels from the Hanford Plutonium Recycle Test Reactor (PRTR), and plutonium oxide scrap materials from the Plutonium Finishing Plant as well as from fuel fabrication operations for other research reactors. Much of the specialty fuel was of a design that mixed uranium oxide with high density, PuO2 particles. In contrast to the plutonium contained in regular Hanford Uranium metal fuel rods, the PuO2 particles from these specialty fuels could potentially segregate during the fuel dissolution process (i.e., the PuO2 could remain partially undissolved) and be routed by entrainment to the tank farms via either the cladding waste stream or the high-level raffinate stream. The following sections discuss the quantities of PuO2 that were processed, the particle sizes present, the tendency of fuel core oxide materials to disintegrate to separable particles that could be entrained in solution transfers out of the dissolver, and finally the estimate of PuO2 quantities that may have been entrained in these transfers to specific underground tanks. Pu Oxide Particle Size and Quantities in Fuel and Scrap PRTR mixed oxide fuels that were processed in the REDOX and PUREX plants during the sixties and early seventies certainly were a source of PuO2 particles of up to 46 microns in size (see Fuel Fabrication Techniques, later in this section). Thus, these particles were at risk of being lost to underground waste tanks.

Official Use Only 170

RPP-RPT-50941, Rev. 0

Table 36 lists, in chronological order, 13 specialty fuel dissolution campaigns that were run between 1957 and July, 1972. Of the 13 various specialty dissolution campaigns run in REDOX and PUREX, the largest quantity of PuO2 came from Plutonium Recycle Test Reactor (PRTR) fuels. Note that some of the PRTR fuels were also of a Pu/Al alloy form which should not have led to the presence of PuO2 particles in the waste. As listed in Table 36, the first campaign of mixed oxide fuel (0.48% PuO2 UO2) was run in the REDOX Plant in June, 1965. PRTR mixed oxide fuels (0.48% and 1.0% PuO2 UO2) were next run in the PUREX Plant in early 1969, followed by a mixture of oxide fuels (0.48%, 1.0%, 1.5%, 2.0% and 4.0% PuO2- UO2) that were run in February, 1970. Three additional campaigns of miscellaneous, unirradiated mixed oxide scrap and PRTR materials were then processed through PUREX in June, 1970, February, 1972, and June, 1972. In the February 1972 campaign to dissolve mixed oxide scrap, the quantity of Plutonium processed (2.5 kg Pu as reported in Table 5.3-1) was derived from Table 36 reference #22 (PPD-494-2, Processing Statistics February 1972) after noting that the planned campaign of seven dissolver charges (to have contained 3.5 kg Pu) was terminated after only five charges were run. Thus, the quantity of Pu processed in this five-charge campaign was estimated as five-sevenths of the total 3.5 kg Pu, or 2.5 kg Pu. This same reference forecasts that the remaining two dissolver charges were to have been processed separately after campaign # 13 in July or August. However, Table 36 reference # 21 (ARH-2416 RD, Chemical Processing Division Monthly Report Summaries Jan. 1972 Dec. 1973) gives no indication that two charges of mixed oxide scrap were processed during the July to August 1972 period, and after August PUREX was placed on standby status until the mid-eighties. Presumably, the two changes worth of scrap were included as part of specialty fuel campaign # 12 -- mixed oxide scrap run in June 1972. Irradiated zirconium-clad special uranium-niobium and uranium-niobium-zirconium alloy fuels were processed in PUREX in July and August 1972 as the final campaign before extended PUREX shut-down. Despite extensive successful laboratory flowsheet testing performed with non-irradiated fuel of both types, the process losses for the irradiated fuels were unexpectedly high with as much as (6 kg) of the Pu lost to cladding removal waste and to solids residues from dissolution of the alloy fuel itself (ARH-2665). According to the lab characterization of process samples, about 50% of the niobium from the fuel alloys precipitated in the dissolver as a niobium pentoxide, Nb2O5, residue. Tests showed that the Pu was co-precipitated and incorporated into the solid phase Nb2O5 matrix and was not released to form acid-soluble plutonium hydrous oxide by the alkaline metathesis flowsheet step using 6 M KOH. Significant uranium also was incorporated into the Nb2O5 and was not released by metathesis. Because the host Nb2O5 dissolves in neither acid (dissolver solution) nor base (tank waste), and the Nb2O5 solids did not release the plutonium and uranium to acid-soluble forms by alkaline metathesis, the incorporated plutonium and uranium will likely continue to persist in the tank waste co-precipitated in the refractory Nb2O5 particles. The Nb2O5 particle size is unknown but could be greater than 10 m. However, the density of amorphous Nb2O5, the most likely form arising under the process conditions, is 4.36 g/cm3 (Holtzberg et al. 1957). This density is much lower than that of PuO2 (~11 g/cm3) and comparable to that of other tank waste sludge solid phases (e.g., goethite or

Official Use Only 171

RPP-RPT-50941, Rev. 0 FeOOH, 4.13 g/cm3; hematite or Fe2O3, 5.04 g/cm3) so that separation by hydrodynamic settling is not an issue. The plutonium dilution with vastly greater quantities of niobium and uranium also affords another margin of criticality safety. According to process records, the niobium-bearing solids residues in cladding removal waste were discharged to Tank C-104. It has been assumed that no plutonium oxide was contributed from losses during the processing of the special alloy fuel and that assumed 0 values are can be entered in Table 36. In total, approximately 70 kg of Pu passed through the REDOX and PUREX dissolvers in the form of PuO2 particles that could have been as large as 46 microns at the beginning of the dissolution step. Conservatism in Pu Quantity Estimates Table 36 gives estimates of the kilograms of total Pu initially present in the pre-irradiated fuel or unirradiated scrap where these data were available in plant processing records. For these mixed oxide fuels that were irradiated (fuel campaign #5), exposure records giving megawatt days per fuel assembly have not been found and burn-out of initial PuO2 cannot be estimated. Thus, for the sake of conservatism and estimating simplicity, the plutonium quantities given for preirradiated fuel are assumed for irradiated fuel at the time of dissolution (i.e., for the PuO2 at risk value). In other cases (Fuel campaigns #7 and #8), PRTR reactor records provided results of burnout code calculations predicting the quantity of total Pu in each irradiated fuel assembly and the sum of these values has been entered in Table 36 as an approximation of PuO2 at risk. Again a degree of conservatism is introduced since total Pu in irradiated fuel represents a combination of burned-out initial PuO2 and Pu bred from the initial fuel U-238. Since bred Pu is not at risk of separating from its parent U-238, the actual quantity of separable (at risk) PuO2 remaining in irradiated fuel is somewhat less by an unknown amount than the total Pu. Of course, for fuel campaigns #7 and #8 fuel exposure data is available and it should be possible to rerun a computer code simulating burn-out in PRTR fuels. Unfortunately, time does not allow such calculations to be completed for this report. Our best estimate as to the conservatism in Table 36 numbers is provided by a simplified burnout calculation by Smith and Kutcher using the THERMOS code (Battelle Northwest Letter, R. I. Smith to R. A. Watrous, Plutonium Content of Irradiated PRTR Fuel Elements). These results indicated that for typical fuel exposure levels the greatest percentage burn-out in initial plutonium occurred in the 0.48% fuel. These results also indicated that the total Pu quantity after irradiation was roughly the same as in unirradiated fuel again, due to the combined effects of initial PuO2 burnout and bred Pu growth. For the 0.48% fuel with an exposure of 5900 MWd/MTU, however, burn-out was estimated to reduce initial PuO2 by about a factor of three (32% remaining). Likewise, for 1.0% fuel after 4600 MWd/MTU exposure, 46% of initial PuO2 was predicted to remain.

Official Use Only 172

RPP-RPT-50941, Rev. 0

The THERMOS code was not run to predict burn-out effects in 2.0% and 4.0% PRTR fuel, and so the degree of conservatism in quantity values for these fuels cannot be estimated, other than to observe that for the same exposure level percentage burn-out values will be less than for the 1.0% fuel. In all cases describing irradiated fuel, the values in Table 36 conservatively overestimate the quantity of PuO2 at risk due to the fact that the burnout of initial Pu during irradiation is not accounted for. Fuel Fabrication Techniques A variety of fuel fabrication methods were tested with the PRTR mixed oxide fuels, as described in references, (HW-79290, Specifications for Swage Compacted, Mixed Oxide (UO2 PuO2), Fuel Elements for the PRTR (Mark I-M); HW-79291, Specifications for Vibrationally Compacted Mixed Oxide (UO2 PuO2), Fuel Elements for the PRTR (Mark I-L). Two or more versions of the basic PRTR fuel element were fabricated and irradiated. Mark I-L elements were vibrationally compacted while Mark I-M elements were swage compacted. Each of these two compaction techniques were applied to either (1) physical mixtures of ceramic grade UO2 and PuO2 or (2) sized particles of PuO2 dispersed in UO2 as formed by a high energy rate, pneumatic-impaction technique (Nupac). A third fabrication technique, pelletized mixed oxide, was also tested on four elements. In all of these fabrication versions, the particle size of the initial PuO2 was specified as less than 46 microns (~325 mesh, Tyler sieve). Further details can be found regarding the PuO2 particle size profile for this ~325 mesh source material. However, for simplicity, 46 microns is conservatively assumed as the particle size for all PRTR plutonium oxide in this study. Dissolution Testing Laboratory testing of PRTR fuel specimens was done to determine the effect of Fuel Fabrication methods and fuel irradiation on dissolution rates of UO2 and PuO2, (BNWL-204 PT2, Aqueous Decladding and Dissolution of Plutonium Recycle Test Reactor Fuels, Part 2: PuO2-UO2 Fuels). These tests also provided qualitative information on the degree of fuel core disintegration during the decladding step. Test results indicated that as long as dissolvent acids contain 0.1M free fluoride ion, dissolution of PuO2 particles would be sufficiently rapid (regardless of Fuel Fabrication method or the degree of fuel irradiation) to essentially completely dissolve each fuel charge. Thus, the accumulation of residual (heel) PuO2 particles in the plant dissolver should have been minimal.

Official Use Only 173

RPP-RPT-50941, Rev. 0

Table 36. Miscellaneous Fuels Dissolved at REDOX and PUREX Plants


Units: Kg of Pu Plant Redox Redox Redox Redox Redox Redox Purex Purex Purex Purex Purex Purex Purex
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

Cladding Waste

High-Level Waste References (1) (2),(3), (4) (5), (6) (7), (8) (9), (10), (11) (12) (13), (14), (15) (16), (17) (18), (19) (20), (21), (22) (21), (23) (21), (24), (25) (21)

Fuel Source Chalk Riv. Fuel PRTR PRTR PRTR PRTR Shippingport PRTR PRTR & EBWR MOX Scrap MOX Scrap PRTR & misc. PRTR special alloy

Diss'n Date 19571958 2/19/1963 Oct-63 6/17/1965 11/23/1966 ca. Feb-69 Feb-70 Jun-70 Feb-72 3/29/1972 6/8/1972 7/19/1972

Fuel Core Type UO2 Pu/Al Alloy UO2 Pu/Al Alloy UO2PuO2 MOX UO2 Blanket UO2PuO2 MOX UO2PuO2 MOX unirr'd MOX unirr'd MOX Pu/Al & U/Al unirr'd MOX Scrap special alloy

Initial Pu 0 8.3 0 11.2 6.6 0 20.3 18.01 0.43 2.5 53.72 22.5

PuO2 at risk 0 0 0 0 6.6 0 20.3 18.01 0.43 2.5 0 22.5

Receiver Receiver Tank (26,27) Tank (26, 27) S-107 C-102 C-104 C-104 C-104 C-104 C-104 SX-111 or SX-114 BX-101 B-101 B-101 B-101 B-101 B-101

0 as oxide assume 0

RHO-CD-505 RD, 7/6/1978, C. E. Jenkins and C. B. Foster, Synopsis of REDOX Plant Operations. HW-76848,3/21/1963, General Electric, Chemical Processing Department Monthly Report for February, 1963. HW-77138, 4/22/1963, General Electric, Chemical Processing Department Monthly Report for March, 1963. HW-76472, 2/5/1963, G. L. Hanson, Proposed Chemical Flowsheets for the Decladding and Dissolution of PRTR Plutonium-Aluminum Fuel. HW-79480, 11/21/1963, General Electric Co., Chemical Processing Department Monthly Report for October, 1963. HW 81950, 3/6/1964, Wallace W. Schulz and Gordon L. Hanson, Decladding and Dissolution of Irradiated PRTR Zircaloy-Clad UO2 Fuel. HW-81620, 4/21/1964, General Electric Co., Chemical Processing Department Monthly Report for March 1964. HW-82089, 5/22/1964, General Electric Co., Chemical Processing Department Monthly Report for April, 1964. RL-SEP-618, 7/21/1965, General Electric Co., Chemical Processing Department Monthly Report for June, 1965. RL-SEP-654, 8/23/1965, General Electric Co., Chemical Processing Department Monthly Report for July, 1965. RL-SEP-397, 4/8/1965, G. L. Hanson, Chemical Flowsheets for the Decladding and Dissolution of PRTR 0.48 Percent PuO2-UO2 Fuel. ISO-642, 1/20/1967, Isochem Inc., Chemical Processing Division Monthly Report for December, 1966. ARH-9552, Jan. 15, 1969, R. A. Watrous, Criticality Analysis of PRTR Fuel Processing in the PUREX Plant -- Part 2: Dissolution and Solvent Extraction Processing of Fuels Containing Up to 1% PuO2 in UO2 ARH923 REV1, March 11, 1969, W. E. Matheison and G. A. Nicholson, Proposed Chemical FlowsheetProcessing of PRTR 0.5 Percent and 1 Percent PuO2UO2 Fuels in the PUREX Plant W. E. Matheison and G. J. Raab, 8/5/1969, Draft #4 undocumented, PRTR Fuel Processing at the PUREX Plant, 0.5% and 1.0% PuO2 in UO2 Fuel ARH-1478, December 15, 1969, R. A. Watrous, Proposed Chemical Flowsheet Processing of PRTR Fuels in the PUREX Plant February, 1970 Letter, R. A. Watrous to Distribution, R. A. Watrous, February 27, 1970, Guidelines for PRTR Plutonium Accountability. ARH-1695, May 8, 1970, R. A. Watrous, Criticality Safety Analysis of Mixed Oxide Scrap Processing in the PUREX Plant. CPD-350 #6, 7/13/1970, US AEC-RLO, Processing Statistics June 1970. ARH-2317, 12/28/71, W. E. Matheison and G. A. Nicholson, Supplemental Criticality Prevention Specification -- Mixed Oxide Scrap Processing in the PUREX Plant, Atlantic Richfield Hanford Co., Richland Washington. ARH-2416 RD, 1/1/1972, H. W. Murray, Chemical Processing Division Monthly Report Summaries Jan. 1972Dec. 1973 PPD 494 #2, 3/15/1972, US AEC, Richland Operations Office, Processing Statistics February 1972. ARH-2389, March 3, 1972, W. E. Matheison and G. A. Nicholson, Chemical Flowsheet Plutonium-Aluminum Processing in the PUREX Plant. PPD-493 #6, US AEC RLO, Monthly Status and Progress Report June 1972. ARH-2480, May 1, 1972, W. E. Matheison, PUREX Chemical Flowsheet Mixed Oxide Processing in the PUREX Plant. WHC-SD-WM-TI-614. Rev. 1, Jan 29, 1996, S. F. Agnew. et al., Waste Status and Transaction Record Summary for the Southwest Quadrant of the Hanford 200 Area. WHC-SD-WM-TI-615. Rev. 1, Jan 29, 1996, S. F. Agnew. et al., Waste Status and Transaction Record Summary for the Northeast Quadrant of the Hanford 200 Area.

Official Use Only 174

RPP-RPT-50941, Rev. 0

Testing to determine the effects of fuel irradiation and fabrication method on the tendency of oxide fuels to disintegrate to a powder (mud) during the decladding process was too limited to allow reliable predictions. Schulz tested unirradiated fuel specimens and found that most core materials disintegrated to a mud when the cladding was removed. In contrast, core materials that had been irradiated to 5000 MWD/t tended to remain as intact chunks of fuel. However, the actual irradiated PRTR fuels had accumulated irradiation exposures that varied from 2000 to over 6000 MWD/ton, (Draft #4, PRTR Fuel Processing at the PUREX Plant, 0.5% and 1.0% PuO2 in UO2 Fuel, 8/5/1969, W. E. Matheison and G. J. Raab). Without additional testing of specimens irradiated to intermediate exposure levels, it is difficult to predict the disintegration behavior of the individual fuels. Note, that the degree to which declad oxide materials might disintegrate and settle to the dissolver vessel floor would likely increase the entrainment of core solids to underground tanks during the several jet out transfers associated with decladding and metathesis steps. Dissolution Process Flowsheets The dissolution processes for specialty fuels tended to involve many extra steps, depending on how the fuel elements were charged to the dissolver vessel, the type of cladding, the need for a metathesis step, and finally the number of core dissolution cuts. Typical dissolution steps were: 1. Dissolution of the disposable aluminum shipping containers in sodium hydroxide sodium nitrate solution. The resulting solution of sodium aluminate was then jet transferred to tank farms. Note that for some PRTR campaigns, the fuel elements were charged in stainless steel baskets which eliminated the need for a canister dissolution step. 2. Mixed oxide scrap materials were sometimes charged in mild steel cans, requiring an initial can dissolution step using nitric acid. 3. Dissolution of the zircaloy cladding and end-fittings in ammonium fluoride ammonium nitrate solution. The resulting solution of ammonium hexafluorozirconate was then jet transferred to tank farms. 4. Metathesis of uranium fluoride compounds (byproducts of Zr cladding dissolution) by digestion in sodium or potassium hydroxide solution. This step was sometimes employed to allow better control of the fluoride ion concentration during core dissolution. Resulting solutions of sodium fluoride were jet transferred to tank farms. 5. Dissolution of the PuO2 UO2 fuel core material in nitric acid ammonium fluoride solution. The resulting U/Pu dissolver product solution was then jet transferred to solvent extraction feed make-up tanks. Each of these steps required a solids settling period before solutions were jet transferred out of the dissolver vessel, plus one or more low volume rinses. Each jet out of solution or rinse offered an opportunity for oxide particles to be entrained to the tank farm, depending on: a) the

Official Use Only 175

RPP-RPT-50941, Rev. 0

degree to which oxide core material had disintegrated during cladding dissolution and metathesis steps, b) particle settling times prior to each jet out transfer, and c) the clearance of the jet out leg above the dissolver vessel floor. Dissolution of the oxide core was controlled by the initial nitric acid and fluoride ion concentrations plus time and temperature, all based on laboratory test experience. Completeness of PuO2 dissolution on individual dissolver charges was often confirmed by sampling the dissolver product solution. For a specialty fuel processing campaign, the over-all material balance for uranium or plutonium typically showed that losses were relatively small. Particle Pathways Figures 50 and 51 show the possible pathways of normally dissolved fuel and incompletely dissolved oxide particles through the REDOX and PUREX processes. In both separations plants there were two main pathways by which particles could have been routed to underground waste tanks: (1) the cladding waste route, and (2) the first extraction cycle raffinate (high level waste) route. The cladding waste route had the potential for entraining large PuO2 particles typical of the preirradiated fuel, mitigated by the degree to which oxide core material would remain intact during the Zirflex and metathesis steps. On the second and subsequent charges of an oxide fuel campaign, the cladding waste route also had the potential of entraining the smaller, residual heel particles left behind from the previous core dissolution steps. The first extraction cycle raffinate (HAW) route had the potential for entraining small residual particles that remained undissolved after the core dissolution step. Note that particles following the raffinate route had to be entrained through multiple vessels and extraction columns before reaching the underground tanks. For example, in the REDOX process, particles would have had to pass through ten tanks, two packed extraction columns and two waste concentrators ample opportunity to be partially retained within the plant. Particles entrained via the REDOX cladding waste route; however, only had to pass through one tank before reaching an underground waste tank. It also should be noted that the cladding waste route had more opportunity to entrain particles (four to seven jet transfers per dissolver charge cycle) than did the raffinate route (one jet transfer). In summary, expect to find more particles entrained via the cladding waste route, and these particles could be of varying particle sizes.

Official Use Only 176

RPP-RPT-50941, Rev. 0

Figure 50. REDOX Process Flow Schematic Dissolution of Specialty Fuels

Official Use Only 177

RPP-RPT-50941, Rev. 0

Figure 51. PUREX Process Flow Schematic Dissolution of Specialty Fuels

Actual PuO2 Loss Experience Although the preceding sections describe various ways in which PuO2 particles could have been routed from dissolver vessels to underground tanks, operating accountability records show that actual percentage losses, although greater than losses typical of the normal 100 Area reactor fuels, were still relatively small.

Official Use Only 178

RPP-RPT-50941, Rev. 0

For example in a detailed processing review of the first PRTR campaign run in the PUREX Plant the overall recovery for plutonium was 95.2 percent, indicating an apparent loss of 4.8%, (Draft #4, PRTR Fuel Processing at the PUREX Plant, 0.5% and 1.0% PuO2 in UO2 Fuel, 8/5/1969, W. E. Matheison and G. J. Raab). In the case of this one campaign, a later flush of the E Cell cladding waste vessels recovered 2.3% of the PRTR plutonium that was processed. These data imply that significant (about 5%) Pu solids were likely carried in the jet-out transfers from the dissolver to cladding waste or dissolver to high-level waste. Due to the decision to flush the cladding waste vessels during this first PUREX campaign, the overall loss of Pu was limited to about 2.5%. However, in other specialty fuel dissolution campaigns, the potential for a Pu loss of 5% must be considered possible since the flushes were not performed. Estimate of Specialty Fuel PuO2 that may have entered Underground Tanks. Table 36 conservatively estimates the mass of Pu considered to be at risk for the six fuel or scrap dissolution campaigns. The sum of these six values is estimated as 70.3 kg Pu. Coupling the Pu-at-risk quantity with the conservative estimate of Pu solids entrainment loss (5%) leads to the prediction that 3.5 kg of Pu may have been entrained to underground waste tanks as a result of all specialty fuel processing in the REDOX and PUREX plants. Prediction of Tanks Receiving PuO2 Particles from Specialty Fuels Table 36 lists those underground tanks that were receiving cladding waste or high-lever waste from the REDOX and PUREX plants at the times these plants were processing mixed oxide fuels. Note that while the total loss of Pu particles is being estimated as 5 percent, the loss via the cladding waste route is being calculated, below, as 4.2%, allowing for an estimated 0.8% loss (five times less) to occur via the high-level waste route. This assumption of a five times greater fractional loss occurring via the cladding waste route reflects the greater number of jet transfers made between the dissolver and the cladding waste tank (four to seven transfers per dissolver charge), compared to the single jet transfer of dissolver product solution per charge. (See discussion of Particle Pathways, above.) Coupling the fuel quantity and receiver tank data from Table 36 with the PuO2 loss percentages, above, leads to a prediction that the predominant loss of PuO2 particles from the REDOX plant (4.2% of 6.6 Kg Pu or 0.28 kg Pu) would likely have been routed via the cladding waste stream to tank S-107. A much smaller fraction of the initial 6.6 kg Pu (estimated to be 0.8% or 0.05 kg) was likely entrained to the high-level waste receiver tanks, SX-111 or SX-114. Likewise, for the PUREX Plant, the cladding waste from the February, 1969 PRTR fuel campaign is estimated to have carried 0.85 kg of Pu in the form of PuO2 particles to tank C-102 (4.2% of 20.3 kg Pu in the fuel), while the high-level waste stream may have carried 0.16 kg of Pu particles to tank BX-101. Note that at this point in time, the PUREX high-level waste was not being neutralized but was being routed as Current Acid Waste through the AR vault (tank AR-001) and on to waste separation processing in B-Plant where any PuO2 particles would have Official Use Only 179

RPP-RPT-50941, Rev. 0

been collected in the feed centrifuge and from there, been routed via the B-Plant low-level waste stream to tank BX-101. Also, in the years 1970 through 1972, PUREX plant cladding waste is similarly estimated to have carried 1.8 kg of Pu particles to receiver tank C-104 (4.2% of 43.4 kg of Pu contained in Fuel # 8, 9, 10, and 12; see Table 36), while the high-level waste stream carried 0.35 kg of Pu to tank B-101. In summary, the estimate of Pu oxide discharged to underground tanks from dissolution and processing of mixed oxide fuels and scrap in the REDOX and PUREX plants is based on the following: 13 Miscellaneous Fuel Campaigns were run in Dissolvers between 1957 and 1972. Six campaigns were MOX fuel which could have introduced Pu oxide particles to underground tanks. A total of 70 kg Pu (in the form of PuO2) was dissolved (no allowance for burn-out reduction was assumed in this estimate.) If the effect of initial PuO2 burnout could be estimated, the quantity of dissolved PuO2 may be as low as 40 to 50 kg Pu, (but this value cannot be known with any certainty). The initial PuO2 particle size was < 46 microns as per fuel fabrication specifications. The effect of fuel exposure on the disintegration of fuel core particles during decladding was assumed to remain constant in all dissolution campaigns. Fractional loss of PuO2 to underground tanks was assumed to be 5%, based on detailed accountability measurements during the first PUREX PRTR fuel campaign. Apportionment of this total loss between Cladding Waste receiver tanks and High-Level receiver tanks was based on the relative number of jet-out transfers to the two receiver tanks (roughly 5x more to cladding waste than to high-level waste assumed constant in all campaigns). The tendency of one discharge route or the other to hinder the entrainment of heavy particles (due to the number of plant vessels on the route) was beyond our ability to estimate and was, therefore, ignored. Table 37. Summary of PuO2 Particle Discharges from REDOX and PUREX Plants
Initial Fuel Plant REDOX PUREX PUREX Total: (Kg Pu) 6.6 20.3 43.44 70.3 S-107 C-102 C-104 Clad Waste Receiver Tk Kg Pu recvd 0.28 0.85 1.82 2.95 HL Waste Receiver Tk SX-111 or 114 BX-101 B-101 Kg Pu recvd 0.05 0.16 0.35 0.56

Official Use Only 180

RPP-RPT-50941, Rev. 0

5.4

CONCLUSION TO NON-PFP PROCESSES

This section of the report has determined that a number of Hanford facilities could not have contributed PuO2 to the Hanford tank farms. These are the B, U, and T Plants, the normal operations of PUREX and REDOX, the Hot Semi-Works, the Critical Mass Laboratory, and the 222-S laboratory. These facilities could not have contributed PuO2 because they did not process any PuO2. Additionally, the PUREX N-cell could not have discharged PuO2 to the tank farm, even though it did process PuO2, because there was no pathway for the PuO2 to get to tank farms. There were several non-PFP processes that may have contributed PuO2 to the tank farms. These are the re-work of PFP scrap solutions at REDOX and PUREX, as well as the processing of mixed oxide specialty fuels at these same two facilities. In addition several truck transfers discussed in section 4.13.3 are included in these results. Table 38 displays the location, quantities and particle properties of PuO2 that may have been discharged to the tank farms from these operations. In addition to the tanks in the table, tank AN-101 is also expected to contain plutonium oxide because C-104 has been retrieved into AN-101 in 2010. The particle sizes listed in Table 38 are based on the source of the PuO2. Table 38. Likely Plutonium Oxide Discharges to the Tank Farms
Source PFP Dissolver Solutions 196364 REDOX PFP Dissolver Solutions 196364 PUREX PRTR Specialty Fuels 1965 REDOX PRTR Specialty Fuels 1965 REDOX PRTR Specialty Fuels 1969 PUREX PRTR Specialty Fuels 1969 PUREX Specialty Fuels 1970-72 PUREX Chemical Form Quantity Lost Tank Farm Location S-108 Density 8-11 g/cc Percent > 10 micron 100% > 10 due to being oxide from burned metal 100% > 10 due to being oxide from burned metal Quantity > 10 micron 1kg Particle Size microns 40-100 Comments No filter used at REDOX 100% went to tank farm Only 20% of material made it past 40 micron filter to tank farm Sieved with 325 mesh use RMA PSD- 4 Distributed 50% to each tank

Nitrite with 1Kg based on 1% oxide loss of 100kg entrainment Nitrite with 2Kg based on 1% oxide loss of 200kg entrainment Oxide 280 g based on 4.2% lost of 6.6 kg 53 g based on 0.8% lost of 6.6 kg 850 g based on 4.2% of 20.3kg 160 g based on 0.8% of 20.3 1820 g based on 4.2% of 43.4kg

A-105

8-11 g/cc

2 X 0.2 = 0.4 Kg

40-100

S-107

8-11 g/cc 45-90% based upon PFP oxide PSD

280 X 0.9 = 252

10-40

Oxide

S-111 and 8-11 g/cc 45-90% based 53 X 0.9 = 48 SX-114 upon PFP oxide PSD C-102 8-11 g/cc 45-90% based upon PFP oxide PSD 8-11 g/cc 45-90% based upon PFP oxide PSD 850 X 0.9 = 768 160 X 0.9 = 144 1820 X 0.9 = 1640

10-40

Oxide

10-40

Oxide

BX-101

10-40

Oxide

C-104 8-11 g/cc 45-90% based tranfered upon PFP to AN-101 oxide PSD in 2010 B-101 8-11 g/cc 45-90% based upon PFP oxide PSD

10-40

98% is in AN-101 and 2% in C-104 1600 and 40 g respectively

Specialty Fuels 1970-72 PUREX

Oxide

350 g based on 0.8% of 43.4

350 X 0.9 = 315

10-40

Official Use Only 181

RPP-RPT-50941, Rev. 0

Table 38. Likely Plutonium Oxide Discharges to the Tank Farms


Source Tank Pumpouts /Truck Transfers 216-Z-8 Chemical Form Oxide/ soluble Quantity Lost 290 grams of Pu in liquid and 142 grams oxide from PFP based upon 10% entrainment Tank Farm Location TX-109 Density 142 g was 811 g/cc Percent > 10 micron Of the 142 g, 50% is oxide from precipitate and based upon PFP oxide PSD of 4590% Quantity > 10 micron 71 X 0.9 = 64 g Particle Size microns 50% is 1040 micron and 50% is 40-100 micron Comments 50/50 mixture of PFP burnt metal and oxide based upon slag and crucible source

Truck Transfer 241-Z-361

Oxide/ less than 40 grams soluble/PuF of Pu in liquid and 4 448 grams oxide from PFP assuming 0.5% entrainment

TX-101

448 g 45-90% based was upon PFP 8-11 g/cc oxide PSD

448 X 0.9 = 0.403 g

10-40 micron

A 0.5% entrainment used since Z-361 was trying not to remove solids

6.0

TANK FARM CHARACTERIZATION AND CHEMISTRY

This section discusses the chemistry of plutonium in the tank farm. Five papers were written by Calvin Delegard of Pacific Northwest National Laboratory on Pu-chemistry in the tank farms can be found in Appendix C of this report. These papers are denoted the Delegard Papers here after the author. The main body of the report in Section 6 covers plutonium inventory and characterization in the tank farm. The first Delegard paper is Ostwald Ripening and Its effect on PuO2 Particle Size in Hanford Tank waste, which discusses the unlikelihood of appreciable growth of PuO2 particles in the Tank Farms under tank farm conditions. The second Delegard paper is Metathesis of Plutonium(IV) Oxalate and Plutonium(III)/(IV)Fluoride in Alkaline Medium to Form PuO2.xH2O, which shows that plutonium(IV) Oxalate and plutonium(III) and (IV) Fluoride cannot survive in the tank farm, and thus are not of concern. The third Delegard paper is called Plutonium(IV) Oxide from Burnt Plutonium Metal, which shows that plutonium metal can survive in the tank farm even though it is not thermodynamically stable because of slow conversion kinetics. The 4th Delegard paper is called Coprecipitation from Nitric Acid Media of Plutonium(IV) with Uranium(VI) in Alkaline Solutions, which shows that plutonium can co-precipitate with uranium when acidic solutions containing both metals are neutralized with sodium hydroxide. This paper also explains the neutralization process used for treating soluble plutonium at all fuel processing facilities.

Official Use Only 182

RPP-RPT-50941, Rev. 0

The fifth Delegard paper, Interfacial Crud Disposition in Alkaline Tank Waste, discusses interfacial crud formation and decomposition. As discussed in Section 4, interfacial Crud is a possible mechanism for the translocation of large plutonium particles from PFP to the tanks.

6.1

INTRODUCTION TO TANK FARMS CHARACTERIZATION

This section discusses the characterization of the tanks that received waste that could have contained large, dense plutonium particles, such as PuO2. The two tanks that will be discussed are SY-102 and TX-118. The other two tanks receiving large quantities of PFP waste (TX-105 and TX-109) have not been sampled previously, so there is no sample data to discuss for these tanks. Tanks that received waste from non-PFP sources (see section 5) have too little large plutonium particles in them relative to their total plutonium inventories for a discussion of the plutonium inventories to be useful. As a background for this section, a Pu inventory for the PFP waste receiver tanks is reported here. The Best-Basis Inventory reports for SY-102, TX-105, TX-109, and TX-118 report the quantity of Pu-238, 239 and 240 (Disselkamp, 2009c;d;e; Place, 2011). Table 39 contains the current Best-Basis Inventory for these Pu isotopes, converted to Kg. Only the Pu associated with the PFP waste layers in SY-102 and TX-118 is reported in Table 39, but the total tank Pu is reported for the other two tanks because a PFP waste layer could not be distinctly identified in them. It should be noted that results of the present study may result in an update in the Best-Basis Inventory for tanks TX-105 and TX-109. The sample-based Pu-inventories for TX-118 and SY-102 are still regarded as the most accurate inventory information available for these two tanks. Table 39. Best-Basis Inventory Values for Pu isotopes in Tanks SY-102, TX-105, TX-109, and TX-118
Tank Name 241-SY-102 241-SY-102 241-SY-102 241-TX-105 241-TX-105 241-TX-105 241-TX-109 241-TX-109 241-TX-109 241-TX-118 241-TX-118 241-TX-118 Total (all four tanks) Pu Isotope 238Pu 239Pu 240Pu 238Pu 239Pu 240Pu 238Pu 239Pu 240Pu 238Pu 239Pu 240Pu All Three Isotopes Kg of Pu 7.89E-03 3.59E+01 2.46E+00 5.64E-04 4.56E+00 2.61E-01 6.55E-05 2.64E+00 7.84E-02 6.43E-03 5.17E+01 2.97E+00 100.58

Official Use Only 183

RPP-RPT-50941, Rev. 0

6.2 6.2.1

CHARACTERIZATION OF WASTE IN TANKS SY-102 AND TX-118 Tank SY-102 History and Characterization

6.2.1.1 Tank SY-102 History. This paragraph provides a history of tank SY-102 and is taken from Disselkamp (2009a).Tank 241-SY-102 went into service during the second quarter of 1977 when it began receiving supernatant from various SSTs (Agnew et al. 1997b) as the feed tank for the 242-S Evaporator until 1980. Since 1980, tank 241-SY-102 has served as the staging tank for cross-site waste transfers from the 200 West Area tanks and processes to the 200 East Area DSTs. It has received saltwell liquor from various SSTs, as well as waste from DST 241-SY-101 in 1999 and 2000. It has also received waste from various Hanford Site processes, including Transuranic (TRU) waste from the PFP, miscellaneous waste from the 222-S Laboratory, dilute liquid waste from the 300 and 400 Area laboratories, dilute phosphate waste from the 231-Z laboratories, and decontamination waste from T-Plant. For the purposes of this present study, the Z plant waste is the most important because it contributed the majority of the plutonium to the tank. Tank SY-102 has been core sampled in 1988, 1990, 1997, and 2000. A review of the data and associated tank stratigraphy for the 1988 and 1990 core samples is recorded in Birnbaum et al. 1993. They determined that a saltcake layer existed on the bottom of tank SY-102 prior to the addition of PFP sludge in 1981. Birnbaum et al. (1993) believed that some of the saltcake dissolved before or during the time period when PFP sludge was being added to the tank. The data from the 2000 core sample are reported in Bushaw (2002a;b) and summarized in the Tank Characterization Report for SY-102 (Disselkamp, 2009a). The analytical data for the 1997 Core sample is reported in Steen (1998). A layer of sludge has settled out over the surface of the PFP sludge, creating a surface layer that covers the PFP waste (Place, 2011). This layer is believed to be solids that were entrained during the dissolution of saltcake in SY-101 and S-102 when these tanks were retrieved into SY-102. The Best-Basis Inventory combines the lower sludge layer and the middle PFP plant sludge layer for convenience (Place, 2011). Evidence for a sludge layer overlying the PFP waste can be derived from the plutonium concentrations reported in the Means and Confidence Interval Reports inside the Tank Characterization Report (Disselkamp, 2009a). A grab sample was taken from the SY-102 sludge surface in April, 2000, immediately after a transfer of waste from SY-101 to SY-102. The sludge surface grab sample was taken in April 2000 and gave a Pu239+240 concentration of 0.0534 microcuries/gram. This contrasts the Pu239+240 concentration reported for a composite of SY-102 sludge taken earlier year 2000 (Prior to the SY-101 transfer), where the Pu239+240 concentration was 9.25 microcuries/gram, or nearly 200 times higher than the grab sample from the sludge surface. The concentration in the grab sample is consistent with the Pu-239+240 concentration in the sludge left behind in SY-102, which is reported to be 0.0679 microcuries/gram (Disselkamp, 2009b). This result supports the conclusion that a sludge layer overlies the Z plant waste in SY-102. Table 40, below, compares the April 2000 grab sample results for the surface layer in SY-102 with the 2007 core sample results from tank SY-101. This table also compares the data to the 2000 Core composite taken from SY-102, which is composed primarily of PFP waste. As can be seen by comparing these key components, the waste on the Official Use Only 184

RPP-RPT-50941, Rev. 0

surface resembles the sludge in SY-101 much more than the PFP sludge in the 2000 SY-102 core composite. This grab sample was taken prior to the retrieval of S-102 into SY-102. The S-102 transfer also covered the surface of the PFP sludge, providing an additional barrier to the PFP waste (Place, 2011). Therefore, the headspace over the sludge in SY-102 could still be used for liquid storage and retrieval without disturbing the PFP sludge. Table 40. Comparison of April 2000 Surface Grab Sample with Underlying PFP Waste and with SY-101 Sludge.
2000 Core Composite Sample (Mostly PFP Waste) 2,230 23,500 14,600 104 9.250 8.350 1,220 0.00684 2000 Grab Sample of SY-102 Surface 550 17,500 1,060 65 0.053 2.230 339 0.00658 2007 Core Sample of Residual Sludge in SY-101 497 22,400 1,440 67 0.068 2.850 426 0.00669

Analyte Ca Cr Fe Sr-90 Pu239+240 U235 U238 U235/U238 Ratio

Units ug/g ug/g ug/g uCi/g uCi/g ug/g ug/g undefined

Callaway and Cooke (2004) identified a particle that contained both plutonium and bismuth. The primary use of bismuth at Hanford was the bismuth phosphate-plutonium phosphate coprecipitation method to recover plutonium from irradiated fuel elements at the T and B plants in the 19441956 time-frames. BiONO3, H3PO4, HNO3, and NaNO2 (to assure Pu IV) were added to Pu(NO3)4 to co-precipitate Bi PO4 and Pu3(PO4)4. The solid was separated from the dissolved uranium, fission products, rare earths, etc. This plutonium-bearing bismuth phosphate waste stream was dissolved for plutonium recovery before begin discharged to the tank farm, so any Be-P-Pu particle found in the tank farms could not have originated from the Bismuth Phosphate process. The question arises then, where did the Bismuth-plutonium particles identified in report Callaway and Cooke (2004) come from? Both bismuth and plutonium are much too insoluble in the caustic waste for the Bi-Pu particle to have reasonably formed by dissolution of the solids and re-precipitation as a mixed phase (Rai and Others, 2010; Delegard and Gallagher, 1983). Both bismuth and plutonium, however, are much more soluble under acidic conditions. A likely location for plutonium and bismuth to precipitate, therefore, would be a location where an acidic stream containing dissolved bismuth and plutonium are neutralized so that they precipitate together. A reasonable candidate for such a stream is when the acidic waste from PFP was blended with evaporator feed in the 242-T evaporator feed vessel. Here, acidic Pu-bearing PFP waste is blended with caustic waste. Disselkamp et al. (2009a; c) shows that there were substantial quantities of bismuth in both Tank SY-102 and tank TX-118. This indicates that there was bismuth in the 242-T evaporator feeds available to react with the soluble plutonium in

Official Use Only 185

RPP-RPT-50941, Rev. 0

the PFP streams. This bismuth had to be small particle size or it would not have been transferred with the liquid waste to the evaporator. These small bismuth particles would likely have rapidly dissolved in the acidic feed stream, which would have made dissolved bismuth available to react with plutonium. Those particles would then have been deposited in tank SY-102. 6.2.1.2 Tank SY-102 Plutonium Concentration. The current Best-Basis Inventory (BBI) for Tank SY-102 (Place, 2011) applied the mean concentration from the 2000 core composite to the PFP sludge layer. This mean concentration of Pu239+240 was 9.25 uCi/g (Table 40) with an upper 95% confidence limit for the mean of 16.6 uCi/g (Disselkamp, 2009a). This value is somewhat higher than the mean (8.35 uCi/g) value for the 1997 core sampling event reported in the Means and Confidence Interval Tables of the Tank Characterization Report (Disselkamp, 2009a). Thus, using the 2000 core sample for plutonium in the BBI is conservative in this regard. The Pu239+240 reported for the 1990 core sampling event was 10.1 uCi/g, but this value was based on a single analyses (no replication), as reported in the 1995 Tank Characterization Report (Winters and DiCenso, 1995). The Appendix of Winters and DiCenso (1995) indicates that the Pu-239 concentration in the 1988 Core sample ranged from 0.13 to 9.73 uCi/g and the Pu240 concentration ranged from 0.039 to 3.40 uCi/g. The results indicate that the current BestBasis Inventory for Tank SY-102 is reasonable. 6.2.1.3 Tank SY-102 Particle Size and Settling. Onishi et al. (1996), performed particle size measurements on sludge from SY-102. They did not record which sample they used, but the study predates the 1997 core sample, so it must have been either from the 1988 or 1990 core sampling event. They identified amorphous silica in the sludge, which may indicate that the sample is contaminated with silica from glass dissolution during storage in glass jars. This contamination may or may not have affected the particle size measurements appreciably. Onishi et al. (1996) did note, however, that the particle size they reported did not agree well with Scheele and Peterson (1990) particle size data from the sludge layer underlying the PFP waste. Scheele and Peterson (1990) measured the particle size of the bottom segment of the 1988 Core sample. Based on the low Pu concentration (Pu239 = 0.13 uCi/g) and low calcium concentration (0.015 milliemol/gram) they reported for this segment, this segment appears to be nearly exclusively from the sludge underlying the PFP waste. Thus, the particle size data reported by Scheelle and Peterson (1990) are not representative of the PFP waste of interest for this study. Onishi et al. (1996) suspended the sludge in de-ionized water, 0.1 M NaNO3 or 1 M NaNO3 during particle size measurements. All of these suspension mediums are much more dilute than the native concentration of the SY-102 supernatant and interstitial liquid (Disselkamp et al., 2009a). Their results show that the particle size increased with the sodium concentration of the liquid. The mean volume-weighted particle size was 6 microns in 0.1 M NaNO3 but 7.24 microns in 1 M NaNO3. High salt concentrations are known to increase aggregation of particles. This increase in particle size observed by Onishi et al. (1996) with increase salt concentration is therefore consistent with theory. Given that the liquid in the tank is much higher than 1 M sodium concentration, the average volume weighted particle size in the tank is likely higher than seven microns.

Official Use Only 186

RPP-RPT-50941, Rev. 0

Wells and Others (2011) also investigated the particle size of PuO2 in tank SY-102. They did not perform any experimental investigation; they simply relied on the data provided by Callaway and Cooke (2004). Therefore, the present study is considered more reliable than Wells et al. (2011) in regards to the particle size and density of PuO2 in waste. Bratzel (1985) performed settling studies on solids collected from tank SY-102. They did not say how the samples were collected, but given the data, and the fact that most of the samples contained only liquids, these samples appeared to be grab samples rather than core samples. They performed the gravity settling experiment for 48 hours in a centrifuge cone. They then measured the total plutonium concentration in the top and bottom portions of the settled solids. They found that the Pu concentration was nearly equal in the top and settled solids layers during gravity settling. These values were Pu = 7.44 uCi/g in the upper layer and 7.86 uCi/g in the lower layer. These results indicate that Pu does not selectively settle out in this waste. Bratzel (1985) also performed a centrifugation experiment, where the centrifuge cones were centrifuged for 20 minutes. The plutonium concentration was again measured in the top and bottom layers of the centrifuge cone. In the centrifuge experiments, there was some segregation of the plutonium. The Pu concentration in the upper layer was 22.4 uCi/g and the concentration in the bottom layer was 3.93 uCi/g. These results indicate that Pu is primarily associated with lighter or smaller particles in the waste that do not settle rapidly under centrifugation. The ratio of Pu in the upper and lower centrifuged solids layer can be used as a worst case estimate of the fraction of plutonium associated with large or dense settling particles. This ratio is 3.93 Ci/(3.93 + 22.4) = 0.15, or 15%. Thus, a maximum of 15% of the plutonium in SY-102 is in the rapid settling fraction of the waste. Herting (2010) has also reviewed properties of PuO2 in SY-102. He primarily reviewed the Z-9 Crib samples (see Section 4.0 of this report) and the Callaway and Cooke (2004) data. Herting (2010) did not perform any additional experimental testing on SY-102. He noted, however, that Calloway and Cooke (2004) looked at 579 plutonium-bearing particles, and found that less than 50 had dimensions larger than 10 microns in length. Most of the particle size data on plutonium oxide discussed in section 4 is from measurements of the Pu-oxide product or Pu-oxide scrap generated in PFP. The Calloway and Cooke (2004) study was on the waste actually in Tank SY-102. Calloway and Cooke (2004) separated the SY-102 sample into a fast settling (large size and dense) and slow settling fractions through settling tests. They indicate that 65% of the plutonium mass in the fast settling fraction was greater than 10 microns in length because the large-size particles had more mass per particle. Nonetheless, they found that 95% of the plutonium was associated with the fines (slow settling) fraction of the waste (Page 87 of Calloway and Cooke, 2004) rather than the fast settling fraction. Therefore only (0.65 *5%) = 3.2% of the plutonium mass was associated with particles that had a dimension greater than 10 microns. Of those, they report that 78% of the Pu bearing particles were associated with only plutonium and oxygen and classified as plutonium oxide (Page 85 of Calloway and Cooke, 2004). Thus (0.78) *3.2% = 2.5%) only 2.5% of the plutonium in the overall sample was associated with true plutonium oxide with dimensions greater than 10 microns. Thus, using the particle size of the Pu-oxide product is conservative.

Official Use Only 187

RPP-RPT-50941, Rev. 0

Calloway and Cooke (2004) measured the plutonium concentration and a number of other key constituents in the slow and fast settling fractions of the SY-102 Waste. Appendix D (Plutonium and Absorber Concentrations in Fast and Low Settling Fraction) reproduces the concentrations of Pu-239, Fe, Cd, Cr, Mn and Ni (key absorbers) in the two size fractions (from Table 2.3-1 of Calloway and Cooke, 2004). Appendix D also has the ratio of Pu-239 to each of the key absorbers in both the rapidly settling and slow settling fractions. Of note is that Pu to absorber ratio is actually lower in the fast settling fraction than the small settling fraction. This data indicates that plutonium does not selectively settle out without neutron absorbers in an experiment that was specifically designed to separate as much large Pu particulate from absorbers as possible. Thus, concerns that the plutonium will selectively settle out without absorbers is not supported by this data. 6.2.2 Tank TX-118 History and Characterization Data

6.2.2.1 Tank TX-118 History. This section provides a history of tank TX-118 and is taken from Disselkamp (2009c). The 241-TX Tank Farm was constructed during 1947 and 1948 in the 200 West Area. The TX Tank Farm contains eighteen 100-series tanks. The 100-series tanks have a capacity of 2869 kL (758 kgal), a diameter of 22.9 m (75.0 ft), and an operating depth of 7.3 m (23.94 ft). Tank 241-TX-118 is the third tank in a three tank cascade with tanks 241-TX-116 and 241-TX-117. Tank 241-TX-118 went into service in 1951 when T plant bismuth phosphate process first decontamination cycle waste cascaded from tank 241-TX-117 and was received from tanks 241-T-104, 241-T-105, 241-T-108, and 241-T-109 (Agnew et al. 1997b). Between 1951 and 1955 tank 241-TX-118 was the feed tank for the 242-T Evaporator and received evaporator feed from tanks 241-T-105, 241-T-106, 241-T-107, 241-T-108, 241-T-109, 241-TX-109, 241-TX-110, 241-TX-111, 241-TX-112, 241-TX-113, 241-TX-114, 241-TX-115, 241-TX-116, 241-TX-117, 241-TY-101, 241-TY-104, 241-TY-105, 241-TY-106, 241-U-110, 241-U-111, and 241-U-112. Tank 241-TX-118 also received small additions of water on four occasions during the time period. Waste was sent from tank 241-TX-118 to various cribs and to tanks 241-T-109, 241-TX-103, 241-TX-113, 241-TX-116, 241-TX-117, and 241-TY-102. No activity in tank 241-TX-118 was recorded for 1956. Activity in the tank resumed in 1957, when decontamination waste was received from U Plant. Tank 241-TX-118 received decontamination waste from U Plant and T Plant between 1957 and 1965. During that time period waste was sent from tank 241-TX-118 to tanks 241-TY-103, 241-TY-104, 241-TX-108, 241-TX-114, 241-TX-116. Between 1965 and the first quarter of 1976, tank 241-TX-118 was an active feed tank for the 242-T evaporator, and waste was transferred between tank 241-TX-118 and the seventeen other TX-farm tanks. Waste was also transferred between tank 241-TX-118 and 241-BX-106, 241-S-107, 241-SX-103, 241-SX-105, 241-U-101, 241-U-102, 241-U-103, 241-U-104, 241-U-105, 241-U-108, 241-U-109, 241-U-110, and various T-Farm and TY-Farm tanks during that time period. Water from miscellaneous sources was added to the tank between 1971 and 1978. Between 1973 and 1976 Z Plant waste was also added to tank 241-TX-118. In 1977 a partial neutralization (PN) process was applied to tank 241-TX-118. A portion of the residual liquors in the tank was converted into solids when the caustic in the slurry supernate was converted to sodium nitrate by reaction with nitric acid. This process was implemented to reduce the volume of liquors requiring double-shell tank storage. Between 1977 and 1978, acid Official Use Only 188

RPP-RPT-50941, Rev. 0

neutralization was applied to the tank alternately with waste additions from various TX-farm tanks. 6.2.2.2 Tank TX-118 Stratigraphy and Characterization. Given the tank history in the previous section, Tank TX-118 would be expected to have a PFP-sludge layer overlying a saltcake. The Best-Basis Inventory, reported in Disselkamp 2009c, indeed reports this to be the case. The BBI reports the upper layer to be a NA layer because it is actually a blend of waste types, because of the partial neutralization campaigns that were adding waste to the tank at the same time as the PFP waste. The Means and Confidence Interval Tables in Disselkamp (2009c) divide the core sample results in the tank between an upper and lower saltcake. The PFP affected saltcake can be identified in the upper layer by the high gross alpha value of 48.1 microcuries per gram, compared to 1.37 microcuries per gram in the lower saltcake. The mean plutonium 239/240 concentration in the upper saltcake was 20.2 microcuries per gram whereas the mean Pu-238 concentration was 4.54 microcuries per gram (Disselkamp, 2009c). No particle size data was found for Tank TX-118. Nonetheless, a bulk particle size for the waste in tank TX-118 would not meaningfully describe the particle size of the Pu-bearing particles anyway because of the large amount of salt in the tank samples.

6.3

PLANNED LABORATORY WORK

As noted in other places in this report, there is still some uncertainty about the particle size and properties of Pu-Particles in Hanford waste. Consequently, this team has proposed some additional laboratory testing to increase knowledge about the nature of plutonium particles in the waste. A laboratory test plan has been completed (McCoskey et al., 2011), but this work has not been completed because of maintenance activities at the laboratory. This work is planned to possibly determine the particle size and chemical form of plutonium in tanks SY-102 and TX-118. Additionally, data on Pu particles from the Z-9 crib will be collected because there is a greater chance of success for this sample because of the much larger concentration of Pu than in the tank samples. Lastly, a control tank (AZ-102) will be analyzed; a tank that we believe received no large plutonium particles, but yet has large amounts of plutonium in a precipitated form. The criterion for selecting samples from the archive is documented in McCoskey et al. (2011). The strategy for these samples is to try to concentrate the large plutonium particles to increase the likelihood of isolation the Pu particles to obtain reliable particle size and chemical identification. The samples will be wet-sieved to obtain a fraction of waste concentrated in large particles, where the large-Pu particles are most likely to be. Subsequently, three separate digestion methods will be employed that will dissolve other major waste constituents but will not dissolve PuO2. The first of these is water, which will dissolve all of the soluble salts in the samples. The second method of concentrated sodium hydroxide, should remove aluminum from the sample. The last extraction is an acid dissolution that should remove iron and some leftover aluminum from the previous extraction. The amount of plutonium removed from these extractions will be quantified because we know that any Pu removed by these extractions cannot be in the form of PuO2. Thus, these extractions will help narrow down the upper bound Official Use Only 189

RPP-RPT-50941, Rev. 0

concentration of PuO2 the waste. Residues for the three extractions will be analyzed by X-ray Diffraction (XRD) and Scanning Electron Microscopy (SEM). XRD will identify crystalline Pu-compounds, such as PuO2, if they are in sufficient concentration. SEM will be used to take pictures of the Pu-particles so that the particle size can be determined. There is reasonable confidence that sufficient particles will be found in the SY-102 and TX-118 to be observed by SEM, as evidenced by the fact that Cooke and Callaway (2004) have previously found large Pu-particles in SY-102 with a similar technique. Hanford is less likely to successfully isolate sufficient PuO2 for conclusive identification by XRD, because of the poor detection limit for this technique. Nonetheless, XRD is the only method available to conclusively identify the chemical form of the Pu solids. Because of the small amount of plutonium oxide material in these tanks compared the total volume of waste, successful identification of the Pu-bearing phase by XRD is not assured.

6.4

RECOMMENDATIONS

This section considers laboratory work that may be of interest for future engineering evaluations of large plutonium particles. Note that the project uses the Data Quality Objective process to identify specific data needs, and this section does not constitute a Data Quality Objective. This section simply identifies areas where the current waste data is meager with regards to PuO2bearing waste. Of note is that two of the tanks that have received possible PuO2-bearing waste from PFP have never been sampled (TX-105 and TX-109). Sampling these two tanks to see if there is elevated plutonium concentrations would substantiate the process knowledge identified in the present report. Given that much of the process knowledge for the quantity of plutonium sent to those tanks is based on un-measured or un-accounted for plutonium, there are substantial uncertainties in these estimates. Sampling these tanks would improve the Best-Basis Inventory estimate for these tanks, which would help us more precisely define the amount of plutonium that came from PFP. Onishi et al. (1996) performed physical property testing on SY-102 waste. The waste they used however; was stored in glass jars for long periods of time, which likely altered the physical properties of the waste. The identification of amorphous silicon oxide indicates that there might have been some degradation of the glass container. Reaction of the waste with the glass container may have altered the physical properties of the waste. Therefore, analysis of the SY-102 (or other PFP waste) with fresh samples would add to the credibility of the results. Fresh SY-102 samples are currently unavailable.

Official Use Only 190

RPP-RPT-50941, Rev. 0

7.0

REVIEW PROCESS

The review process for this investigation was designed to provide a technical review of the work to insure it has the proper investigative rigor and a result that was defendable for future safety basis, criticality and design documentation. The process was comprised of several parts and is described in the following sections: Data Qualification Validation of the 2010 WTP report Data Analyses for margin of safety Independent review team.

Data Qualification It was expected that the team would identify directly relevant data and identify supporting data needed to prepare this document. Data from numerous sources such as logbooks, personal communications and letters, analytical data, classified data, shift notebooks, PFP engineering files, and other sources were expected to be in conflict or not fully understood. This existing data needed to be evaluated by several technical disciplines and subject matter discipline experts to determine if it was valid for the intended purpose in this report. The guidance provided in ASME NQA-1-2004, Nonmandatory Appendix 3.1, Guidance on Qualification of Existing Data was reviewed by the team lead and several members of the team. The guidance was considered as the selection and quality of the data collected was being reviewed by team members throughout the investigation process. Sections such as 402, 404, and 500 were evaluated for relevance and considered when evaluating the data for validity and for documenting the data in the final report. Numerous attributes from these sections were considered before the data was considered for use in this evaluation. These included: the technical adequacy of the equipment and procedures used to collect and analyze the data, the extent to which the data demonstrate the properties and ranges of interest (e.g., physical, chemical, geological, mechanical), prior uses of the data, prior peer or other professional reviews of the data and their results, extent and reliability of the documentation associated with the data, and the extent and quality of corroborating data or confirmatory test results. The data was qualified for use by several methods. These include data corroboration and peer review. Several examples include; samples of specific tank contents were identified for confirmatory testing in future efforts. The peer review method was used to independently evaluate data to determine if the methodology was acceptable or the data had been used in a similar range of applications. Use of the peer review method for this purpose included an evaluation of the data acquisition and development approach to determine the acceptability of the uncertainties associated with the data acquisition or development methodology, the adequacy and appropriateness of the interpretations derived from the data, and the extent to which the uncertainties affect the interpretations, conclusions, and overall validity of the data.

Official Use Only 191

RPP-RPT-50941, Rev. 0

This process was conducted in numerous team meetings and documented in the meeting minutes. There were also numerous peer to peer reviews in evaluating the data. All conclusions and comments by individual peer reviewers are included in this report. Data Analyses for Margin of Safety While preparing the estimate for the amount and size of the plutonium being sent to the tank farms, certain margins of safety were documented to insure the result was sufficient for future needs. Each area below describes where margin was integrated into the estimate and can assist the future users of this document in understanding of the margin of safety. In Appendix I, for each tank, the assumptions used in determining the amount of material sent to that tank are summarized. The range of material sent to each tank is based upon the teams best estimate and what it would be if the margin was removed. Material Unaccounted For The material losses for PFP were reviewed throughout the entire processing life of PFP. Three separate documented investigations and numerous informal ones were reviewed as part of this investigation. The consensus of those reviews is that the unaccounted losses for PFP during the time frame of interest (1973-2004) have been established at a value equal to the accountable losses. It was also assumed that all metal fines that came from incomplete burning are coated with a fine passive layer that would prevent them from dissolving in the dissolver and corroding tank caustic environment. Without that layer it would be expected after 20-30 years in the tank the metal would be expected to corrode to a small fine particle. Production Losses for PUREX and REDOX for Specialty Campaigns All production losses for PUREX and REDOX were assumed to be as large as or larger than the historical losses in the facility. These losses were assumed to be Pu oxide in addition to the soluble losses. All reactor calculations were used as the starting inventory. The calculations were not always conservative and some losses can be attributed to incomplete burnup or that the Pu was consumed in the reactors. These losses were not accounted for in the loss estimates. Particle Size Distributions When reviewing the particle size distribution table a range existed for the various sizes of particles lost. The largest value was used if it gave the most conservative value. For example for PFP metal oxide production, the percentage of 10 40 micron particles produced were 45-90% of product produced. The calculations assumed the 90% value vice the 45% value giving the largest amount lost.

Official Use Only 192

RPP-RPT-50941, Rev. 0

Particle Size Limit The PSD data presented specifically for oxide from metal burning includes a large particles, greater that 100 micron, but the subsequent analysis used a 100 micron upper limit for particles that were lost in the liquid waste stream. The basis for 100 micron upper limit for particles discharge in the waste stream was picked based on the consensus judgment of several plutonium processing subject matter experts. This was supported by several observations: 40 micron is the practical upper limit for oxide from oxalate precipitation. 100 micron was the largest oxide particle seen in Z-9 as reported by Ames. Oxide was usually screened prior to dissolution. Based on engineering judgment, a 100 micron particle would be the upper limit of particles that could be carried out of the dissolver pots and through the solvent extraction column and tanks. Settling calculation studies for the MT dissolver pots that were found in the PFP history files show 10 micron particles would settle in about 30 minutes and that 128 micron particles would settle in 8 minutes. In 1978 settling times in MT were given as 10 minutes in training materials. Later, after the settling calculation was done in 1984, this was changed to 30 minutes.

Filter Losses When accounting for filter bypass, vendor data was not available. The standard pass through for current filters was doubled to account for the uncertainties in filter capture and possible installation bypass issues. Holdup and Cleanouts It was assumed when calculating process losses that no holdup occurred in the facilities. When a clean out was done, the MCA valves were used to determine how much material was flushed from the facility to the tank farm. In essence, the flushes were counting material already accounted for in the normal production losses, giving a conservative value. Validation of the 2010 WTP Report As part of this investigation, the authors developed a list of assumptions, statement of facts and areas to investigate to facilitate this investigation. (Appendix A; also mentioned in Section 2.2 of this document) The team reviewed the information and agreed that the areas needing investigation were valid. The WRPS team used these areas to begin their investigation. At the end of the investigative stage, the team lead reviewed the list to insure all areas were reviewed and to some degree discussed in the report. The PFP investigation covered all assumptions and items for investigation except numbers 4 and 7. They are discussed in various sections in the write up on PFP.

Official Use Only 193

RPP-RPT-50941, Rev. 0

Items 4 and 7 were investigated by the PUREX/REDOX team and are discussed in the Tank Farm Characterization and Chemistry section. The two items were not fully completed and this report has recommendations in that same section for sample and analyses to be completed. The lab analysis was scheduled to be completed however, due to the lab being down for the later part of the investigation for maintenance the results are not available for this report. Once the data is received, it will be given to the criticality and safety basis groups to use in conjunction with this report to determine its impact. Independent Review Team To insure the results of this investigation would be able to be used for safety basis and criticality analysis along with being a design input an independent team review was established as part of the resolution plan. A charter was developed that provided the outline for the review process (Appendix E). This process was used during the report writing phase so that if issues were found they could be corrected prior to the report being issued. The team was comprised of four experts in nuclear operations management and oversight, scientific investigation, nuclear and criticality safety, radiological characterization and plutonium chemistry and processing. The team members were: Frank R. McCoy, III, Lead Dr. Fred Beranek Dr. Raymond G. Wymer Dr. Michael T Ryan, PhD C.H.P. The scope of the review was to review documents, receive presentations and interview presenters and other team members associated with the WRPS Plutonium Oxide Team in order to: A. Ascertain the technical merit, completeness of data analysis, and overall thoroughness of investigation activities conducted by the Project to quantify the extent to which plutonium oxide or metal with a particle size greater than 10 microns and particle density greater than 8 grams/cm3 may exist in the Hanford Tank Farms and its distribution within the Tank Farm. Ascertain the reasonableness of conclusions and recommendations documented in this report of the Projects activities. Provide input and recommendations to enhance the technical validity of the Projects conclusions and improve the Projects report. Provide appropriate text for input into the Projects report regarding the results of the IRGs review.

B.

C.

D.

Official Use Only 194

RPP-RPT-50941, Rev. 0

The Independent review team concluded the following from their review: The IRG concludes that the investigation and data analysis activities of the Plutonium Oxide Evaluation Project were complete, thorough, and comprehensive. As reflected in the Teams recommendations, and in full recognition of the uncertainties inherent with historical waste data quality, the IRG concludes there is still valuable technical merit in the results and conclusions of the Projects activities. The IRG concludes that the conclusions and recommendations presented by the Plutonium Oxide Evaluation Project, again in recognition of the uncertainties inherent with historical waste data quality, are reasonable, conservative and useful.

The IRG made several recommendations: That plutonium particle quantities and properties be formally portrayed on a tank-bytank, and where possible layer by layer, basis. Link the plutonium particle quantity and property data with their associated assumptions and analyzed ranges of estimated values. When developing plans to prepare tanks contents for feed to the WTP, efforts should be made to account for plutonium content and form in each tank and related considerations regarding criticality.

The IRG recommends that during tank retrieval and transfer operations, samples taken for other parameters, should be considered for analysis for plutonium particle size and density to provide more definitive characterization (Appendix F).

8.0

CONCLUSION

Based upon the results of the mixing studies for the Waste Treatment and Immobilization Plant (WTP) pretreatment facility, a concern emerged over the settling of greater than 10 micron plutonium-bearing particles in WTP process vessels. Specifically, the site technical staff believed that the plant may have transferred large, dense plutonium oxide particles to the tank farms which could then be sent to WTP for processing. Based upon this information, WTP management commissioned a three-person team of former Hanford employees with extensive PFP and tank waste experience to investigate if large, dense plutonium-bearing particles could be present in the tank farms. In February 2011, the draft report prepared by this team was given to Washington River Protection Solutions (WRPS) management for their review. The draft report indicated that there was a possibility that large, dense plutonium-bearing particles could be present in the waste tanks. WRPS concluded that the presence of such particles could exceed the analyses of the criticality safety evaluation report (CSER) for the waste tanks. In response to these findings, and because the original team did not have broad access to process documentation or access to Official Use Only 195

RPP-RPT-50941, Rev. 0

classified documentation, WRPS management commissioned a second team to investigate the issue and determine the extent of the problem using all available sources of information. The team was comprised of the original authors of the WTP report, PNNL, WTP, WRPS, and contract personnel. The scope of the teams review was to: Determine all source facilities that contributed plutonium oxide (PuO2) or plutonium oxalate to the tank farm Determine all the tanks that received PuO2 or oxalate Determine how much plutonium was sent to each tank and identify its chemical and physical form Determine the particle size and density of the plutonium disposed to the tank farms, specifically looking for particles greater than 10 microns.

As part of this review, questions about the disposal of plutonium fluoride compounds (e.g., PuF4) to the Hanford waste tanks were also to be investigated. Accordingly, the team reviewed the chemistry of Pu oxalate and Pu fluoride compounds to determine their fate when neutralized and sent to the tank farms. The investigation determined that the plutonium and other crystalline forms, when not calcined and after neutralization, were light, fine particles of low-solubility hydrous plutonium oxide, PuO2xH2O, and would not be an issue. Examples of the forms investigated are listed below. Pu nitrate, which was neutralized to a hydroxide in either tank farms or 241-Z and co-precipitated as light, fine particulate. Interfacial crud (typically emulsions), which will breakdown and allow the soluble Pu component to form a hydroxide. The insoluble Pu particles trapped in the interfacial crud binder will be released when it degrades. Pu fluoride and Pu oxalate particles which will undergo alkaline hydrolysis to Pu hydroxide, which is formed as small, amorphous particles less than 10 micron in size when nitrate is mixed with caustic solution.

The precipitated plutonium hydroxide could be bonded to other particles or poisons making a co-precipitated agglomerate. In this manner, they were no different than the other plutonium hydrous oxide discarded as processing operational losses to the tank farm. Plutonium fluoride compounds likewise were found to form finely divided low-solubility plutonium hydrous oxide. These forms were removed from the quantities of concern and therefore are not included in the results discussed. Upon completion of the investigation phase, each team whose members were Process and Pu chemistry experts developed a table that apportioned the fraction of total plutonium discharged Official Use Only 196

RPP-RPT-50941, Rev. 0

from each of the various sources based on the plant operating history and the feeds processed. Using the estimated fractions and the total assumed discard values, the amount of each plutonium waste form was determined. There is approximately 900 kgs of Pu listed in best basis inventory for the tank farms. This includes all forms of Pu, oxalate and calcined material. Of the total 900 kgs, approximately 100 kgs can be attributed to the PFP processing campaigns in PFP and the other processing facilities such as PUREX and REDOX. The nitrate, oxalate, fluoride, and hydroxide accounts for approximately 70 kgs of the total 100 kgs sent to the tank farm and are not considered to be of concern. Of the numerous facilities investigated for liquid and solids discharged to the tank farms, only three PFP, PUREX and REDOX sent calcined or refractory oxide material to the tank farms. These facilities accounted for the remaining 30 kgs of material. The PFP facility contributed a majority of the calcined material with approximately 20 kg of calcined oxide and 2.5 kg of metal fines greater than 10 microns. The PUREX and REDOX facilities processed some PFP material along with several special campaigns of oxide and mixed oxide. This resulted in remaining approximately 7 kgs of oxide greater than 10 microns being sent to the tank farms. The 30 kg value is the WRPS teams best estimate. The actual summation of all the material shown in this report being sent to the tank farm is slightly more than 28 kg. However, due to the nature of the information which this estimate was made, a conservative value of 30 kg was chosen. In determining the 30kg amount of material sent to the tank farm, several margins of safety were used. As discussed in section 7 some examples of those margins are: The largest values of any source for accountable values were used for PFP source material A 2X multiplier was applied for the fraction of material unaccounted for (MUF) associated with the liquid waste for stream from PFP during the period of interest (1973-2004) The most conservative values of particle size distributions were used for PFP source material Pu metal fines were assumed to persist in the Tank Farm for PFP source material No Pu was assumed to be fissioned in reactor operation with MOX fuels for REDOX/PUREX Source Material Assumed that all lost Pu was PuO2 for mixed oxide fuel processed REDOX/PUREX Source Material.

Official Use Only 197

RPP-RPT-50941, Rev. 0

If the margins of safety are reduced, the amount of material sent to the tank farm could be lowered to approximately 18 kgs. However, this investigation has no basis to remove the margins at this time. The material was sent to 16 tanks in the tank farm. Eight tanks received an appreciable quantity (>750 g) of oxide or metal from the three facilities. They are 244-TX, AN-101, C-102, S-108, SY-102, TX-105, TX-109, and TX-118. In addition there were eight more tanks that received minimal amounts (<400 g) of oxide or metal and they are A-105, B-101, BX-101, C-104, S-107, S-111, SX-114, and TX-101. The particle size distribution was determined based upon the methods used to process the plutonium and taken from historical sources. Pu oxide from calcination of oxalate has a particle size of 1 to 40 microns and a density of 8 11 g/cc. Pu oxide from metal oxidation has a particle size from 40 to 100 microns and a density of 8 11 g/cc. Pu metal fines from rework of material and scrap recovery has a particle size of 40 to 100 microns and a density of approximately 19 g/cc. The Pu metal fines are not thermodynamically stable in tank waste and may not have survived the extended storage time. A Summary table in Appendix J tabulates the results by year and processing facility. A second table in Appendix I lists the margin of safety and calculates the range of material of concern on a tank-by-tank basis. The conclusions of the WRPS team are consistent with the results of the previous work, where it was estimated that between 65 130 kg of all forms of Pu (midpoint 98) were sent to the tank farms; however the original report made no attempt at defining the amount or specific size of each form. The results of the WRPS team were reviewed by an independent team of 4 experts in nuclear operations management and over site, scientific investigation, nuclear and criticality safety, radiological characterization and plutonium chemistry and processing. The team spent approximately one month reviewing the report, receiving presentations and interviewing team personnel. The team concluded that the investigation and data analysis activities were complete, through and comprehensive. Also, the conclusions and recommendations presented were reasonable, conservative and useful in full recognition of the uncertainties inherent with the historical waste data quality. In summary there is the potential for approximately 30 kg of Pu Oxide and Pu metal fines greater than 10 microns in the tank farm. It is located in 16 tanks, 8 that have appreciable quantities that challenge the CSER. Particle sizes range from 10 100 microns and the density range is mainly in the 8 11 g/cc for the oxide with approximately 2.5 g of 19 g/cc Pu metal fines material.

Official Use Only 198

RPP-RPT-50941, Rev. 0

9.0

REFERENCES

12134-R5, 1980. Status of 242-Z Tank Solutions (Contract DE-AC06-77RL01030), (external letter J.E. Kinser to O.J. Elgert, February 25), Rockwell Hanford Operations, Richland, Washington. 12221-PSL88-005, January 13, 1988. Particle Size Analyses of Plutonium Fluoride Samples from the RMC Line in 234-5 Building, Internal Memo from Barney, G.S., 12221-PSL88-005, January 13, 1988, Westinghouse Hanford Company, Richland, Washington. 12221-PSL88-058, 1988. Analysis of Suspended Solids in Solution from Tanks BT-1, 2, 3, (internal letter from Plutonium Process Support Laboratories to D.M. Gerboth, April 13), Westinghouse Hanford Company, Richland, Washington. 12820-89-LHR-001, 1989. Waste Characterization of Plutonium Finishing Plant Waste to 241-Z, (internal letter from C.A. Barrington to L.H. Rogers, January 18) Westinghouse Hanford Company, Richland, Washington. 15000-91-ECV-143, Waste Minimization Report November from Plutonium Finishing Plant, Westinghouse House Hanford, Richland, Washington. 24590-CM-HC4-W000-00176-T02-01-00001, (R.C. Hoyt, J. Teal, T. Jones, September 2011) Historical Overview of Solids in PFP Aqueous Waste Transferred to Tank Farms: Quantity of Plutonium, Particle Size Distribution, and Particle Density, Bechtel, Inc., Richland, Washington 24590-CM-HC4-W000-00176-T02-01-00002 Rev 00B, (R.C. Hoyt, J. Teal, T. Jones, June 2010) Historical Overview of Solids in PFP Aqueous Waste Transferred to Tank Farms: Quantity of Plutonium, Particle Size Distribution, and Particle Density, Bechtel, Inc., Richland, Washington. 293-003105, July 17, 1974. Criticality Specification Letter 293-003105, 7-17-1974, ARH-145 attached. 60413-78-101, 1978. Revision to Process Specification Numbers 10 & 8, (internal letter from D.W. Reberger to D.G. Harlow, June 12), Rockwell International, Richland, Washington. 60413-78-115, CAW Centrifuge, D.T. Crawley to R.J. Thomas, June 28, 1978. Atlantic Richfield Hanford Company, Richland, Washington. 60413-78-136, Backup CAW Centrifuge Installation, August 4, 1978. D.W. Reberger to G.C. Oberg, Rockwell International, Richland, Washington. Official Use Only 199

RPP-RPT-50941, Rev. 0

60413-78-142, CAW Centrifuge Flushing, August 10, 1978. D.W. Reberger to C.L. Owen, Rockwell International, Richland, Washington. 65240-80-161, 1980. Technical Basis Process Specification Limits 10.4.f, 10.4.h, & 10.4.i, (internal letter from D.W. Reberger to File, October 2), Rockwell Hanford Operations, Richland, Washington. 65453-85-006, 1985. Tank 102-SY Plutonium Settling Study , Bratzel, D.R, 1985, Letter from D.R. Bratzel to L.A. Gale dated January 10, 1985, Rockwell International, Richland, Washington. 65454-84-155, 1984. Pu/Am Concentrations, Pu Isotopic Composition and Thermal Behavior of Hood 7-C Floor Sludge (internal letter from C.H. Delegard and D.G. Bouse to R.L. Crocker, December 4), Rockwell Hanford Operations, Richland, Washington. 65454-85-023, Recommendations for Dissolution of Rocky Flats Oxide and Resulting Centrifuge Sludge, February 8, 1985. E.T. Abramowski to C.H. Delegard/D.G. Bouse, Rockwell International, Richland, Washington. 65454-85-032, 1985. Chemical Analyses of Rocky Flats Oxide Sludge and Heel, (internal letter from C.H. Delegard to E.T. Abramowski, February 25), Rockwell International, Richland, Washington. AE-4783, 1974. Waste Tank Survey Contract AT(45-1)-2130, (internal letter from G. Burton, Jr. to O.J. Elgert, November 1), Atomic Energy Commission, Richland, Washington. AEC-4783, 1974. Waste Tank Survey Contract AT (45-1) 2130, (letter from G. Burton, Jr. to O.J. Elgert, July 5), Atlantic Richfield Hanford Company, Richland, Washington. AEC-4783, 1974. Waste Tank Survey Contract AT(45-1)-2130, (internal letter from G. Burton, Jr. to O.J. Elgert, September 12), Atomic Energy Commission, Richland, Washington. Alenchikova, I.F., L.L. Zaitseva, L.V. Lipis, V.V. Fomin, and N.T. Chebotarev, 1958. Preparation and Properties of Certain Double Fluorides of Tetrapositive Plutonium, Proceedings of the Second United Nations International Conference on the Peaceful Uses of Atomic Energy, Volume 28, Basic Chemistry in Nuclear Energy, United Nations, Geneva, Switzerland. ARH-923 REV1, March 11, 1969. W. E. Matheison and G. A. Nicholson, Proposed Chemical Flowsheet Processing of PRTR 0.5 Percent and 1 Percent PuO2UO2 Fuels in the Purex Plant, Atlantic Richfield Hanford Company, Richland, Washington.

Official Use Only 200

RPP-RPT-50941, Rev. 0

ARH-1048 RD, 1969. Z Plant Report Task I-II Recovery Operations, Atlantic Richfield Hanford Company, Richland, Washington. ARH-108, 1967. Safety Analysis Report and Hazards Review CAW Centrifuge, Atlantic Richfield Hanford Company, Richland, Washington. ARH-11 RD, Z Plant Report Task I-II Recovery Operations, September 5, 1967 through December 31, 1967. Atlantic Richfield Hanford Company, Richland, Washington. ARH-1194, July 7, 1969. Tank Farm Data, May 1973 to March 1978, J.G. Murphy, Atlantic Richfield Hanford Company, Richland, Washington. ARH-1278, 1969. Plutonium-Americium Soil Penetration at 234-5 Building Crib Sites, Atlantic Richfield Hanford Company, Richland, Washington. ARH-1478, December 15, 1969. R.A. Watrous, Proposed Chemical Flowsheet Processing of PRTR Fuels in the Purex Plant February, 1970, Atlantic Richfield Hanford Company, Richland, Washington. ARH-1695, May 8, 1970. R.A. Watrous, Criticality Safety Analysis of Mixed Oxide Scrap Processing in the Purex Plant, Atlantic Richfield Hanford Company, Richland, Washington. ARH-1824, 1970. Nuclear Materials Review Plutonium Finishing Section, Third and Fourth Quarters Fiscal Year 1970, Atlantic Richfield Hanford Company, Richland, Washington. ARH-1904 DEL, April 1971. 200 Areas Operation Monthly Report, Atlantic Richfield Hanford Company, Richland, Washington. File name: ARH-1950-RD, 1971. Monthly Activities Reports Separations Chemistry Lab CY 1971 Book 1, Atlantic Richfield Hanford Company, Richland, Washington. ARH-2077, 1971. Continuous Dissolver Development Program Progress Report March 1970 to March 1971, Atlantic Richfield Hanford Company, Richland, Washington. ARH-2078, 1971. Prototype Solids Handling Systems for Plutonium Finishing, Atlantic Richfield Hanford Company, Richland, Washington. ARH-2078 ADD, 1972. Addendum, Prototype Solids Handling Systems for Plutonium Finishing, Atlantic Richfield Hanford Company, Richland, Washington.

Official Use Only 201

RPP-RPT-50941, Rev. 0

ARH-2194-DEL, August 30, 1971. DEL, Nuclear Materials ReviewPlutonium Finishing SectionThird and Fourth Quarters Fiscal Year 1971, R.W. Doerr, Atlantic Richfield Hanford Company, Richland, Washington. ARH-2221, 1972. Hazards Evaluation Plutonium-Uranium Partitioning 236-Z Building Project HCE-651, Atlantic Richfield Hanford Company, Richland, Washington. ARH-2317, 12/28/71. W.E. Matheison and G.A. Nicholson, Supplemental Criticality Prevention Specification Mixed Oxide Scrap Processing in the Purex Plant, Atlantic Richfield Hanford Company, Richland Washington. ARH-2335, 1972. Design Criteria Disposal of Z Plant Wastes, Atlantic Richfield Hanford Company, Richland, Washington. ARH-2343, Panesko, J.V., 1972. Hydrofluoric Acid Scrubber Systems, Atlantic Richfield Hanford Company, Richland, Washington. Available at: http://www.osti.gov/bridge/servlets/purl/1016165-iGj45w/1016165.pdf. ARH-2389, March 3, 1972. W. E. Matheison and G.A. Nicholson, Chemical Flowsheet Plutonium-Aluminum Processing in the Purex Plant. Atlantic Richfield Hanford Company, Richland, Washington. ARH-2416 RD DEL, 1974. Chemical Processing Division Monthly Report Summaries January 1972 December 1973, Atlantic Richfield Hanford Company, Richland, Washington. ARH-2480, May 1, 1972. W. E. Matheison, Purex Chemical Flowsheet Mixed Oxide Processing in the Purex Plant. Atlantic Richfield Hanford Company, Richland, Washington. ARH-2584-DEL, August 30, 1972. Nuclear Materials ReviewPlutonium Finishing Section Third and Fourth Quarters Fiscal Year 1972, R.W. Doerr, Atlantic Richfield Hanford Company, Richland, Washington. ARH-2597, 1972. Safety Analysis Plutonium Effluent Clean-up, Z Plant, Atlantic Richfield Hanford Company, Richland, Washington. ARH-2728-DEL, February 1, 1973. Nuclear Materials Review-Plutonium Finishing Section First and Second Quarters Fiscal Year 1973, R.W. Doerr, Atlantic Richfield Hanford Company, Richland, Washington. ARH-2915, 1973. Nuclear Reactivity Evaluations of 216-Z-9 Enclosed Trench, Atlantic Richfield Hanford Company, Richland, Washington.

Official Use Only 202

RPP-RPT-50941, Rev. 0

ARH-3024-DEL, January 31, 1975. Nuclear Materials Review-Plutonium Finishing Section Calendar Year 1974, R.W. Doerr, Atlantic Richfield Hanford Company, Richland, Washington. ARH-423-RD-DEL, 1968. Monthly Reports 1968, Monthly Reports from Process Engineering, Process Engineering Section 1 of 2, Atlantic Richfield Hanford Company, Richland, Washington. ARH-423-RD-DEL, 1968. Monthly Reports 1968. Plutonium Process Engineering Section 2 of 2, L.M. Knights, Atlantic Richfield Hanford Company, Richland, Washington. ARH-959-2, 1969. Criticality Analysis of PRTR Fuel Processing in the Purex Plant Part 2: Dissolution and Solvent Extraction Processing of Fuels Containing up to 1% PuO2 in UO2, Atlantic Richfield Hanford Company, Richland, Washington. ARH-CD-132, July 26, 1974. Criticality Specification for 241-TX-109 J.C. Womack, Atlantic Richfield Hanford Company, Richland, Washington. ARH-CD-178, 1974. Operational Safety Analysis 242-T Waste Evaporator, Atlantic Richfield Hanford Company, Richland, Washington. ARH-294 RD-DEL, 1968. Z Plant Report Task I-II & Recovery Operations, January 1, 1968 through December 31, 1968, Engineers of Plutonium Process Engineering Atlantic Richfield Hanford Company, Richland, Washington. ARH-CD-296, 1975. Criticality Specification for 241-TX-101, Atlantic Richfield Hanford Company, Richland, Washington. ARH-CD-323, 1976. Z Plant Liquid Waste Disposal through the 241-Z Vault, Atlantic Richfield Hanford Company, Richland, Washington. ARH-CD-336A, 1975. Production and Waste Management Division Waste Status Summary January 1, 1975 through March 31, 1975, J.D. Anderson, Atlantic Richfield Hanford Company, Richland, Washington. ARH-CD-411 A DEL, 1975. Nuclear Materials Review Plutonium Finishing Section Third and Fourth Quarters, Fiscal Year 1975, Atlantic Richfield Hanford Company, Richland, Washington. ARH-CD-704, 1976. Tributyl Phosphate Solvent Extraction Flowsheet for Recovery of Plutonium from Scrap, Atlantic Richfield Hanford Company, Richland, Washington.

Official Use Only 203

RPP-RPT-50941, Rev. 0

ARH-F-109, Oma, K.H., 1977. Plutonium Metal Electrolytic Dissolution Flowsheet for the Plutonium Reclamation Facility, Atlantic Richfield Hanford Company, Richland, Washington. ARH-LD-205 B, 1975. Atlantic Richfield Hanford Company Monthly Report, Atlantic Richfield Hanford Company, Richland, Washington. ARH-LD-206 B, 1975. Atlantic Richfield Hanford Company Monthly Report, June 1975, Atlantic Richfield Hanford Company, Richland, Washington. ARH-LD-213 B, 1976. Atlantic Richfield Hanford Company Monthly Report January 1976, Atlantic Richfield Hanford Company, Richland Washington. ARH-SA-232, 1975. Characterization of Actinide-Bearing Sediments Underlying Liquid Waste Disposal Facilities at Hanford, Atlantic Richfield Hanford Company, Richland, Washington. ATA-5G, R.P. Corlew to O.J. Elgert, Prototype Solids Handling System for Z Plant Processes Contract AT (45- )-2130, U.S. Atomic Energy Commission, Richland Operations Office, Richland, Washington. BNWL-204 PT2, November, 1966. W W. Schulz, Aqueous Decladding and Dissolution of Plutonium Recycle Test Reactor Fuels, Part 2: PuO2-UO2 Fuels, Pacific Northwest Laboratory, Richland, Washington. BNWL-205, 1965. Plutonium Release Studies: I. Release from the Ignited Metal, Mishima, J, Pacific Northwest Laboratory, Richland, Washington. BNWL-357. 1966. Plutonium Release Studies: II. Release from Ignited Bulk Metallic Pieces, Mishima, J, Battelle Northwest, Richland, Washington. BNWL-B-296, 1973. Nature of Actinide Species Retained by Sediments at Hanford: Interim Progress Report, J.L. Swanson, Battelle Pacific Northwest Laboratory, Richland, Washington. BNWL-B-419, 1975. Special Distribution, Plutonium Recovery from Incinerator Ash and Centrifuge Sludge by Peroxide Fusion an Interim Report, Battelle Pacific Northwest Laboratories, Richland, Washington. BNWL-CC-649, 1966. Disposal Characteristics of Plutonium and Americium in a High Salt Acid Waste, Battelle-Northwest, Richland, Washington.

Official Use Only 204

RPP-RPT-50941, Rev. 0

BNWL-SA-1492, Simultaneous Recovery and Purification of Pm, Am, and Cm by the Use of Alternating DTPA and NTA Cation Exchange Flowsheets, E.J. Wheelwright, F.P. Roberts, and L.A. Bray, Pacific Northwest National Laboratory, Richland Washington. BNWL-1642, 1972. Development of an Electrolytic Dissolver for Plutonium Metal, Wheelwright, E.J., Battelle Pacific Northwest Laboratories, Richland, Washington. BNWL-1812, 1974. Characterization of Actinide Bearing Soils: Top Sixty Centimeters of 216-Z-9 Enclosed Trench, Rev. 0, Pacific Northwest Laboratories, Richland, Washington BNWL-2117, 1976. Proceedings of an Actinide-Sediment Reactions Working Meeting at Seattle, Washington on February 10-ll, 1976, section Actinide Occurrences in Sediments Following Ground Disposal of Acid Wastes at 216-Z-9, Battelle Pacific Northwest Laboratory, Richland, Washington. Bundschuh, T., R. Knopp, R. Mller, J.I. Kim, V. Neck, and Th. Fanghnel, 2000. Application of LIBD to the Determination of the Solubility Product of Thorium(IV)-Colloids, Radiochimica Acta 88:625-629. Burney, G.A. and F.W. Tober, 1965. Precipitation of Plutonium Trifluoride, I&EC Process Design and Development 4(1):28-32. CCN:228133, Plan/Schedule for WTP / Tank Farms Issue Identification/Resolution, Washington River Protection Solutions, LLC, Richland, Washington. CCP-AK-RL-101, 2011. Central Characterization Project Acceptable Knowledge Summary Report for Hanford Plutonium Finishing Plant (PFP) Contact-Handled Transuranic Debris Waste Stream: RLMPFPDD, Rev. 6, Washington TRU Solutions, LLC, Carlsbad, New Mexico. CCP-AK-RL-110, 2011. Central Characterization Project Acceptable Knowledge Summary Report for Hanford Building 209E Contact-Handled Transuranic Debris Waste from Decontamination and Decommissioning, Waste Stream RLCCP209ED1, Rev. 0, Carlsbad, New Mexico, Washington TRU Solutions, LLC. CH2M-0400872, 2004. Distribution of Plutonium-Rich Particles in Tank 241-SY-102 Sludge, (external letter from W.S. Callaway and G.A. Cooke to K.H. Abel, May 17), CH2M HILL Hanford Group, Inc., Richland, Washington. CP-15584, 2005. Documented Safety Analysis for 209-E Facility Critical Mass Laboratory, Rev. 1, Fluor Hanford, Inc., Richland, Washington.

Official Use Only 205

RPP-RPT-50941, Rev. 0

Clark, D.L., S.S. Hecker, G.D. Jarvinen, and M.P. Neu, 2006. Plutonium, Chapter 7, in The Chemistry of the Actinide and Transactinide Elements, 3rd edition, L.R. Morss, N.M. Edelstein, J. Fuger, and J.J. Katz, Editors, Springer, Dordrecht, The Netherlands. Clark, D.L., S.S. Hecker, G.D. Jarvinen, and M.P. Neu, 2006. Plutonium, Chapter 7 in The Chemistry of the Actinide and Transactinide Elements, 3rd edition, L.R. Morss, N.M. Edelstein, J. Fuger, and J.J. Katz, Editors, Springer, Dordrecht, The Netherlands. Clark, D.L., S.S. Hecker, G.D. Jarvinen, and M.P. Neu, 2006. Plutonium, Page 936, Chapter 7, in The Chemistry of the Actinide and Transactinide Elements, 3rd edition, L.R. Morss, N.M. Edelstein, J. Fuger, and J.J. Katz, Editors, Springer, Dordrecht, The Netherlands. Clark, S.B., and C. Delegard, 2002. Plutonium in Concentrated Solutions, Chapter 7 in Advances in Plutonium Chemistry 1967-2000, D.C. Hoffmann, Senior Editor, American Nuclear Society, La Grange Park, Illinois. Cooke, G.A., 2011. Personal communication from co-author of Callaway and Cooke (2008). CPD-350 #6-DEL, 7/13/1970, US AEC-RLO, Processing Statistics June 1970, U.S. Atomic Energy Commission Richland Hanford Operations, Richland, Washington. CPS-T-1-49-00012, 2011. Criticality Prevention Specification, Rev. B-16, Washington River Protection Solutions, Richland Washington. D&D-30349, 2006. Study of Liquid Effluents and CERCLA Hazardous Constituents Generated and Discharged by the Plutonium Finishing Plant, Rev. 0, Fluor Hanford, Inc., Richland, Washington. Dawson, J.K., R.M. Elliott, R. Hurst, and A.E. Truswell, 1954a. The Preparation and Some Properties of Plutonium Fluorides, Journal of the Chemical Society 558-564. Dawson, J.K., R.W.M. DEye, and A.E. Truswell, 1954b. The Hydrated Tetrafluorides of Uranium and Plutonium, Journal of the Chemical Society 3922-3929. DD-48234, 2011. Data Quality Objectives Summary Report for the 209E Facility, Rev 0, CH2M HILL Plateau Remediation Company, Richland, Washington. Deichman, E.N. and I.V. Tananaev, 1962. Plutonium Fluorides, Soviet Radiochemistry 4(1):56-62. Delegard, C.H. 1985. Solubility of PuO2xH2O in Alkaline Hanford High-Level Waste Solution. RHO-RE-SA-75 P, Rockwell Hanford Operations, Richland, Washington. Available at: www.osti.gov/bridge/servlets/purl/5402793-8aDMm2/5402793.pdf. Also found as Official Use Only 206

RPP-RPT-50941, Rev. 0

Solubility of PuO2xH2O in Alkaline Hanford High-Level Waste Solution, CH Delegard, 1987. Radiochimica Acta 41:11-21. DOE/RL-88-30 2010, Hanford Site Waste Management Units Report, Rev. 19, CH2M HILL Plateau Remediation Company, Richland, Washington. DOE/RL-96-82, 2004. Hanford Facility Dangerous Waste Closure Plan, 241-Z Treatment and Storage Tanks, Rev. 1, U.S. Department of Energy, Richland Operations Office, Richland, Washington. DOE-STD-3013-2004, 2004. Stabilization, Packaging, and Storage of Plutonium-Bearing Materials, U.S. Department of Energy, Washington, D.C. Ecology, EPA, and DOE, 1989. Hanford Federal Facility Agreement and Consent Order TriParty Agreement, 2 vols., as amended, State of Washington Department of Ecology, U.S. Environmental Protection Agency, and U.S. Department of Energy, Olympia, Washington Esch, R. A., 1998. Tank 241-TX-118, Core 236 Analytical Results for the Final Report, HNF-1651, Rev. 0, Waste Management of Hanford, Inc., Richland, Washington. Evaluation of Process Test PRF-69-2, Improved Control of Solids Build-Up in the Waste Treatment Facility, June 10, 1971. Atlantic Richfield Hanford Company, Richland, Washington. Fanghnel, Th. and V Neck. 2002. Aquatic Chemistry and Solubility Phenomena of Actinide Oxides/Hydroxides, Pure and Applied Chemistry 74(10):1895-1907. FDM-T-290-00001, 1986. 204-AR-Rail Car Unloading Facility Facility Description Manual, Rev/Mod A-0, Westinghouse Hanford Company, Richland, Washington. Fedoseev, A.M., N.N. Krot, N.A. Budantseva, A.A. Bessonov, M.V. Nikonov, M.S. Grigoriev, A.Yu Garnov, V.P. Perminov, and L.N. Astafurova. 1998. Interaction of Pu(IV,VI) Hydroxides/Oxides with Metal Hydroxides/Oxides in Alkaline Media. PNNL-11900, Pacific Northwest National Laboratory, Richland, Washington. Available at: http://www.osti.gov/bridge/servlets/purl/665966-ij8eO9/webviewable/665966.pdf. See also: Fedoseev, A.M., N.A. Budantseva, M.V. Nikonov, and M.S. Grigor'ev. 2000. Interaction of Pu(IV) Hydroxide with Hydroxides of Some d and f Elements in Alkaline Solutions: I. Systems Pu(IV)-Fe(III) and Pu(IV)-Ni(II). Radiochemistry 42(5):405-412. FSS-P-080-00002, 1985. PUREX Sample Schedule, Rev/Mod C-0, Rockwell Hanford Operations, Richland, Washington.

Official Use Only 207

RPP-RPT-50941, Rev. 0

Gelman, A.D., A.I. Moskvin, L.M. Zaitsev, and M.P. Mefodeva.,. 1962. Complex Compounds of Transuranium Elements, pages 121-123, Consultants Bureau, New York, New York. Gelman, AD. and L.P. Sokhina,. 1958. Oxalate Complex Compounds of Plutonium (IV), Journal of Inorganic Chemistry, USSR 3(5):49-56. Griffo, J.S., W.B. Brown, and F.D. Lonadier, 1964. Solubility of Plutonium Compounds, MLM-1191, Mound Laboratory, Miamisburg, Ohio. Guillaumont, R.T. Fanghnel, J. Fuger, I. Grenthe, V. Neck, D.A. Palmer, and M.H. Rand. 2003. Page 318 of Update on the Chemical Thermodynamics of Uranium, Neptunium, Plutonium, Americium and Technetium. Elsevier, Amsterdam, The Netherlands. Guillaumont, RT. Fanghnel, J. Fuger, I. Grenthe, V. Neck, D.A. Palmer, and M.H. Rand. 2003. Update on the Chemical Thermodynamics of Uranium, Neptunium, Plutonium, Americium and Technetium. Elsevier, Amsterdam, the Netherlands. H-2-26271, 1967. Centrifuge Installation Pulse Column Hood 6th Floor, Isochem Inc., Richland, Washington. April 1, 1969 letter, L.M. Knights to L.M. Ingalls. Solution Clarification Hood the Plutonium Reclamation Facility, Atlantic Richfield Hanford Company, Richland, Washington. H-2-26721, 1972. Engineering Flow Diagram Pu U Partitioning, Rev. 2, Hanford Engineering Services, Richland, Washington. H-2-26722, 1970. Engineering Flow Diagram Pu U Partitioning & Uranium Loadout Facility, Rev. 3, Hanford Engineering Services, Richland, Washington. H-2-65483, 1983. Process Flow Diagram, Rev. 1, Vitro Engineering Corporation, Richland, Washington. H-2-75643, 1983. Process Flow Diagram, Rev. 1, Kaiser Engineers Hanford Company, Richland, Washington. H-2-75645, 1997. Engineering Flow Diagram N-6 Glove Box & Pipe Chase, Rev. 10, Fluor Daniel Hanford, Inc., Richland Washington. H-2-75657, 1987. Piping N-6 Glove Box Section & Details, Rev. 3, Kaiser Engineers Hanford Company Operations, Richland, Washington. Haire, R.G., M.H. Lloyd, M.L. Beasley, and W.O. Mulligan, 1971. Aging of Hydrous Plutonium Oxide, Journal of Electron Microscopy 20(1):8-16.

Official Use Only 208

RPP-RPT-50941, Rev. 0

Harlow, D.G. and L.J. Olguin, 1977. Current Status and Safety Hazards in Miscellaneous Treatment, Internal letter to RB Gelman, March 8, 1977, Atlantic Richfield Hanford Company, Richland, Washington. Haschke, J.M. and J.L. Stakebake, 2006. Handling, Storage, and Disposition of Plutonium and Uranium, Chapter 29 in The Chemistry of the Actinide and Transactinide Elements, 3rd edition, L.R. Morss, N.M. Edelstein, J. Fuger, and J.J. Katz, Editors, Springer, Dordrecht, The Netherlands. Haschke, J.M. and T.H. Allen. 2002. Equilibrium and Thermodynamic Properties of the PuO2+x Solid Solution. Journal of Alloys and Compounds 336:124131. Haschke, J.M., T.H. Allen, and L.A. Morales, 2000. Surface and Corrosion Chemistry of Plutonium, Los Alamos Science, No. 26, NG Cooper, editor, Los Alamos National Laboratory, Los Alamos, New Mexico, Available at: http://www.fas.org/sgp/othergov/doe/lanl/pubs/00818031.pdf. Haschke, J.M., T.H. Allen, and L.A. Morales, 2001. Reactions of Plutonium Dioxide with Water and Hydrogen-Oxygen Mixtures: Mechanisms for Corrosion of Uranium and Plutonium, Journal of Alloys and Compounds 314:78-91. HEP-687-2-9, 1973. Characterization of 216-Z-9 Trench Soils X- Ray Diffraction Identification of Plutonium Particles, Battelle Pacific Northwest Laboratories, Richland, Washington. HNF-1681 Bell, K. E. 1999. Tank 241-TX-118, Final Report for Rotary-Mode Core Samples 259 and 260, Rev. 0, Fluor Daniel Hanford, Inc., Richland, WA HNF-1692, 2000. Tank 241-Z-361, Cores 263 and 264 Analytical Results for the Final Report, Rev. 0B, Fluor Hanford, Richland, Washington. HNF-1989, 1998. Tank 241-Z-361 Process and Characterization History, Rev. 1, B&W Hanford Company, Richland, Washington. HNF-2012, 1997. Engineering Study of the Criticality Issues Associated with Hanford Tank 241-Z-361, Rev. 0, Duke Engineering and Services, Richland, Washington. HNF-2012, 1997. Engineering Study of the Criticality Issues Associated with Hanford Tank 241-Z-361, Rev. 0, Duke Engineering and Services, Richland, Washington. HNF-22064, 2004. (DEL Revision 0), Plutonium Finishing Plant Operations Overview (1949-2004), Fluor Hanford, Richland, Washington.

Official Use Only 209

RPP-RPT-50941, Rev. 0

HNF-27735, 2006. Operating, Analytical and Engineering Data for 241-Z Tank D5 Batch 286, Rev. 0, Fluor Hanford, Inc., Richland, Washington. HNF-30206, 2006. 241-Z D-5 Cell RCRA Closure, Rev. 0, Fluor Hanford, Inc., Richland, Washington. HNF-31792, 1007. Characterization Information for the 216-Z-9 Crib at the Plutonium Finishing Plant, Rev. 0, Fluor Hanford, Richland, Washington. HNF-3337, 1998. Authorization Basis for the 209-E Building, Rev. 0, Fluor Daniel Hanford, Inc., Richland, Washington. HNF-33999, 2007. 241-Z As Left Characterization, Rev. 0, Fluor Hanford, Inc., Richland, Washington. HNF-7437, 2002. Tank 241-SY-102, Cores 284, 285, and 286 Analytical Results for the Final Report, Rev. 0A, Fluor Hanford, Inc., Richland, Washington. HNF-7758, 2001. E Critical Mass Laboratory, Emergency Preparedness Hazards Assessment, Rev. 0, Fluor Hanford, Inc., Richland, Washington. HNF-8735, 2001. 241-Z-361 Tank Characterization Report, Rev. 0, Fluor Hanford, Richland, Washington. HNF-EP-0924, 1997. History and Stabilization of the Plutonium Finishing Plant (PFP) Complex, Hanford Site, Rev. 0, Fluor Hanford Company, Richland Washington. HNF-EP-0924, 1997. History and Stabilization of the Plutonium Finishing Plant (PFP) Complex, Hanford Site, Fluor Hanford, Inc., Richland, Washington. HNF-FMP-03-18568, 241-Z Tank D5 Transfer System Modification, Fluor Hanford, Inc., Richland, Washington. HNF-SD-WM-DP-267, Steen, F.H., 1998. Tank 241-SY-102, Cores 211 and 213 Analytical Results for the Final Report, HNF-SD-WM-DP-267, Rev. 0, Waste Management Hanford, Inc., Richland, Washington. HNF-SP-1147, 1997. PUREX/UO3 Facilities Deactivation Lessons Learned History, Rev. 2, B&W Hanford Company, Richland, Washington. Hobbs, DT, 1999. Precipitation of Uranium and Plutonium from Alkaline Salt Solutions, Nuclear Technology 128:103-112.

Official Use Only 210

RPP-RPT-50941, Rev. 0

Holtzberg, F, A Reisman, M. Berry, and M. Berkenblit. 1957. Chemistry of the Group VB Pentoxides. VI. The Polymorphism of Nb2O5, Journal of the American Chemical Society 79(9):2039-2043. HW-10475, 1944. Hanford Technical Manual, Sections A, B, and C, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-14744, 1949. Tentative Analytical Methods and Estimated Sample Sizes for Process Control of REDOX Production Plant No. 1, General Electric Company, Hanford Works, Richland, Washington. HW-15825, 1950. Modification of the Hot Semiworks to Accommodate the TBP Waste Metal Recovery Flowsheet, Hanford Atomic Production Operation, Richland Washington. HW-19991, 1951. Production Test 234-2 Recycle of Skull Solution to Isolation Process, Hanford Atomic Products Operation, Richland Washington. HW-22596, 1951. RECUPLEX Feasibility Report, Hanford Atomic Products Operation, Richland, Washington. HW-22771, Modification of the Hot Semiworks to Accommodate the PUREX Process Flowsheet, ED Bragg, November, 1951. Hanford Atomic Production Operation, Richland Washington. HW-22772, 1951. Modification of the Hot Semiworks to Accommodate Waste Metal Recovery Flowsheet Using 30% TBP in CCl4, Hanford Atomic Production Operation, Richland Washington. HW-245141952, Critical Mass Studies of Plutonium Solutions, General Electric, Neucleonics Division, Richland, Washington. HW-26365, 1952. Brief Summary of Separations Processes, Hanford Atomic Products Operation, Richland Washington. HW-29200-DEL, 1954. Plutonium Purification and Fabrication Technical Manual, Hanford Atomic Products Operation, Richland Washington. HW-31767, 1954. Hot Semiworks REDOX Studies, Hanford Atomic Production Operation, Richland Washington. HW-35030, 1955. RECUPLEX Operating Manual, General Electric, Hanford Atomic Products Operation, Richland, Washington.

Official Use Only 211

RPP-RPT-50941, Rev. 0

HW-40910, Burger, L.L., 1955. The Chemistry of Tributyl Phosphate A Review, HW-40910, Hanford Atomic Products Operation, Richland, Washington. HW-42111, 1956. Counting Methods and Calculations Used by the Analytical LaboratoriesSeparations Section, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-44585, 1956. Crystallite Sizes of PuO2 Powder, Pallmer, P.G, Hanford Atomic Products Operation, Richland, Washington. HW-45128, 1956. Thermal Decomposition of Plutonium(IV)Oxalate and Hydrofluorination of Plutonium(IV) Oxalate and Oxide, Myers, M.N. Hanford Atomic Products Operation, Richland, Washington. HW-49597, 1957. Conversion Chemistry of Plutonium Nitrate, A, Harmon, K.M. and W.H. Reas, Hanford Atomic Products Operation, Richland, Washington. HW-57603, 1958. Preliminary Hazards Study of the Hanford Plutonium Critical Mass Laboratory, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-59434, 1959. Chemical Processing Department Research and Engineering Operation Monthly Report February, 1959, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-63089, 1959. Chemical Processing Statistics Book I, December 14, 1959, D McDonald, Hanford Atomic Products Operation, Richland, Washington. HW-63090, 1959. Chemical Processing Statistics Book II, December 14, 1959, D McDonald, Hanford Atomic Products Operation, Richland Washington. HW-64851, Bruns, L.E., 1960. Solvent Extraction Flowsheet for New Plutonium Reclamation Facility, Bruns, L.E., Hanford Atomic Products Operation, Richland, Washington. HW-65727, 1960. Removal of Plutonium from 234-5 Sump Wastes and from RECUPLEX Extraction Wastes, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-66266, 1960. Hazards Summary Report for the Hanford Plutonium Critical Mass Laboratory, General Electric, Hanford Atomic Products Operation, Richland, Washington.

Official Use Only 212

RPP-RPT-50941, Rev. 0

HW-67010, 1960. Design Scope of the Waste Treatment Facility, Z-Plant, Project CGC-912, Hanford Atomic Products Operation, Richland, Washington. HW-67240, 1960. Minutes of Critical Mass Laboratory Program Meeting, General Electric, Hanford Atomic Products Operations, Richland, Washington. HW-68857, 1961. Hanford Plutonium Critical Mass Laboratory Nuclear Safety, Operational Limitations, Instructions, and Emergency Procedures, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-72283 DEL, 1962. Chemical Processing Nuclear Materials Control, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-72666, 1963. Hot Semiworks Stronium-90 Recovery Program, Hanford Atomic Products Operation, Richland, Washington. HW-73579, 1962. Special Plutonium Recycle at PUREX, BF Judson and HC Rathvon, Hanford Atomic Products Operation, Richland, Washington. HW-74723, 1962. Final Report of Accidental Nuclear Excursion, RECUPLEX Operation, 234-5 Facility, U.S. Atomic Energy Commission, Richland, Washington. HW-76472, 1963. Proposed Chemical Flowsheets for the Decladding and Dissolution of PRTR Plutonium-Aluminum Fuel, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-76848, 1963. Chemical Processing Department Monthly Report for February, 1963, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-77083, 1963. Leaching of Plutonium from the 216-Z-9 Crib, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-77138, 1963. Chemical Processing Department Monthly Report for March, 1963, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-77531, 1964. Characteristics of Burning Plutonium, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-79290, 1963. Specifications for Swage Compacted, Mixed Oxide (UO2 PuO2), Fuel Elements for the PRTR (Mark I-M), General Electric, Hanford Atomic Products Operation, Richland, Washington.

Official Use Only 213

RPP-RPT-50941, Rev. 0

HW-79291, 1963. Specifications for Vibrationally Compacted, Mixed Oxide (UO2 PuO2), Fuel Elements for the PRTR (Mark I-L), General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-79480, 1963. Chemical Processing Department Monthly Report for October, 1963, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-80168, 1963. Campaign Rework of Z-Plant Material at PUREX, BF Judson and JB Kendall, Hanford Atomic Products Operation, Richland, Washington. HW-81620, 1964. Chemical Processing Department Monthly Report for March, 1964, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-81950, 1964. Decladding and Dissolution of Irradiated PRTR Zircaloy-Clad UO2 Fuel, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-82089, 1964. Chemical Processing Department Monthly Report for April 1964, General Electric, Hanford Atomic Products Operation, Richland, Washington. HW-83876, 1964. Chemical Processing Department Monthly Report for August, 1964, Hanford Atomic Products Operation, Richland, Washington. ISO-642, 1967. Chemical Processing Division Monthly Report for December, 1966, Isochem Inc., Richland, Washington. ISO-659 RD, 1967. Z Plant Report Task I-II Recovery Operations, January 1, 1967, through September 5, 1967, Isochem Inc., Richland, Washington. ISO-659, Z Plant Report Task I-II and Recovery Operations, January 1, 1967 through December 21, 1967, Isochem, Inc., Richland, Washington. ISO-756, 1967. Burning and Extinguishing Characteristics of Plutonium Metal Fires, Felt, RE, Isochem, Incorporated, Richland, Washington. JA Teal, RC Hoyt, April 1997. Clean-Up and Disposal of Machining Effluents Jenkins, I.L., F.H. Moore, and M.J. Waterman. 1963. X-Ray Powder Crystallographic Data on Plutonium Oxalates and Their Isomorphs, Chemistry and Industry (London) 1:35-36, As described by JM Cleveland, 1970, The Chemistry of Plutonium, page 407, Gordon and Breach Science Publishers, Inc.

Official Use Only 214

RPP-RPT-50941, Rev. 0

Jenkins, I.L., F.H. Moore, and M.J. Waterman. 1965a. X-ray Powder Crystallographic Data on Plutonium and Other Oxalates I: The Oxalates of Plutonium(III) and Plutonium(VI) and Their Isomorphs. Journal of Inorganic and Nuclear Chemistry 27(1):77-80. Jenkins, I.L., F.H. Moore, and M.J. Waterman. 1965b. X-Ray Powder Crystallographic Data on Plutonium and Other Oxalates II: Plutonium(IV) Oxalate Dihydrate, Uranium(IV) Oxalate Hexahydrate, Uranium(IV) Oxalate Dihydrate and Thorium Oxalate Dihydrate, Journal of Inorganic and Nuclear Chemistry 27(1):81-87. Keller, C. and M. Salzer, 1967. Ternre Fluoride des Typs MeIIMeIVF6 mit LaF3-Struktur, Journal of Inorganic and Nuclear Chemistry 29:2925-2934. Klem, M.J, 1972. Evaluation of Process Test PRF 67-5 Filtration of CA Column Interface Jettings, Internal memo, June 20, 1972, Atlantic Richfield Hanford Company, Richland, Washington. Kolman, D.G. and L.P. Colletti, 2008. Aqueous Corrosion Behavior of Plutonium Metal and PlutoniumGallium Alloys Exposed to Aqueous Nitrate and Chloride Solutions. Journal of the Electrochemical Society, 155 (12):C565-C570. Krot, N.N., V.P. Shilov, A.M. Fedoseev, A.B. Yusov, A.A. Bessonov, N.A. Budantseva, S.I. Nikitenko, G.M. Plavnik, T.P. Puraeva, M.S. Grigoriev, A. Yu Garnov, A.V. Gelis, V.P. Perminov, and L.N. Astafurova, and C.H. Delegard, 1998. Alkaline Treatment of Acidic Solution from Hanford K Basin Sludge Dissolution, PNNL-11944, Pacific Northwest National Laboratory, Richland, Washington. Available at: http://www.osti.gov/bridge/product.biblio.jsp?osti_id=2507. LA-12315-MS, Haschke, J.M., 1992. Evaluation of Source-Term Data for Plutonium Aerosolization, Los Alamos National Laboratory, Los Alamos, New Mexico. LA-13069-MS, Haschke, J.M., 1995. Reactions of Plutonium and Uranium with Water: Kinetics and Potential Hazards, 12315-MS, Los Alamos National Laboratory, Los Alamos, New Mexico. LAB-PLN-11-00009, 2011. Plutonium Particle Characterization of Hanford High-Level Waste Tanks, Washington River Protection Solutions LLC, Richland, Washington. LA-12676-MS, 1993. Chemical Composition of Hanford Tank SY-102, Los Alamos National Laboratory, Birnbaum, E, S. Agnew, G. Jarvinen, and S. Yarbro Los Alamos, New Mexico.

Official Use Only 215

RPP-RPT-50941, Rev. 0

LA-UR-02-5920, Kolman, DG, 2002. The Aqueous Corrosion Behavior of Plutonium Metal Exposed to Nitric Acid Solution, Los Alamos National Laboratory, Los Alamos, New Mexico. LA-UR-04-1453, Eller, P.G., R.W. Szempruch, and J.W. McClard, 2004. Summary of Plutonium and Oxide Storage Package Failures, Los Alamos National Laboratory, Los Alamos, New Mexico. LA-UR-74-356, Martell, C.J., 1974. The Effect of Particle Size on the Carrier-Distillation Analysis of PuO2, Los Alamos Scientific Laboratory, Los Alamos, New Mexico. LA-UR-97-311, 1997. Waste Status and Transaction Record Summary (WSTRS), Rev. 4, Los Alamos National Laboratory, Los Alamos, New Mexico. Lee, E., C. Ribeiro, E. Longo and E. Leite, 2005. Oriented Attachment: An Effective Mechanism in the Formation of Anisotropic Nanocrystals, Journal of Physical Chemistry B 109:20842-20846. Letter, Americium Exchange Column Accident, .February 13, 1979, J.E. Kinser to O.J. Elgert, February 13, 1979. Letter, CA Column Interface Crud Sampled on September 27, 1971, 1971a, Panesko, J.V., Internal letter from Panesko, J.V., to A.J. Low, October 1, 1971, Atlantic Richfield Hanford Company, Richland, Washington. Letter, CAW Centrifuge and CAW Monitor, April 20 1977, Memo, DTC to DGH (hand written) Atlantic Richfield Hanford Company, Richland, Washington. Letter, CAW Centrifuge Installation in Hood 6 of Miscellaneous Treatment, June 28, 1978 letter, K.H. Oma to R.J. Thomas, Atlantic Richfield Hanford Company, Richland, Washington. Letter, CAW Centrifuge Problems, July 10, 1978 RHO letter, D.W. Reberger to P.A. Miskimin, Rockwell International, Richland, Washington. Letter, CAW Centrifuge Solids from Burnt Oxide Run, internal letter from Panesko, J.V., to A.J. Low, August 26, 1971, Atlantic Richfield Hanford Company, Richland, Washington. Letter, Centrifuge Operation, July 30, 1973. R.E. Felt to C.M. Peabody, Atlantic Richfield Hanford Company, Richland, Washington. Letter, Centrifuge solids, June 19, 1973. P.C. Ely to M.J. Klem, Atlantic Richfield Hanford Company, Richland, Washington. Official Use Only 216

RPP-RPT-50941, Rev. 0

Letter, Characterization of 216-Z-8 Tank and 361-Z Settling Tank, 1974. (from C.M. Peabody and D.A. Turner to D.G. Harlow, May 31), Atlantic Richfield Hanford Company, Richland, Washington. Letter, Concurrence to Ship Plutonium Nitrate from 209-E, Critical Mass Laboratory, June 7, 1984. from D.D. Scott to H.E. Ransom, Pacific Northwest Laboratory, Richland, Washington. Letter, Design Criteria: Interim Disposal of PFP Wastes, Chiaramonte, GR, 1971. Draft, Atlantic Richfield Hanford Company, Richland, Washington. Letter, Design request Hood 6 Centrifuge Discharge Piping Dec 10, 1973. D.A. Turner to W.H. Koontz, Atlantic Richfield Hanford Company, Richland, Washington. Letter, Determination of Plutonium in Waste Sumps, 234-5 Building to J.J. Courtney from L.M. Knights, June 8, 1962. Letter, Determination of the Pu Build-Up in Waste Treatment Facility Tanks, W-2, W-3, and W-4, 1969 (internal letter from P.C. Ely to R.E. Olsen, March 10), Atlantic Richfield Hanford Company, Richland, Washington. Letter, Disposition and Isolation of Tanks 270-E-1, 270-W, 241-CX-70, 241-CX-71, and 241-CX-72, Cummings 1989 ref Memo, Harlow/Teal, July 1974. Letter, Disposition of 361-Z Tank Contents, (attachment from J. R. Irish to M. L. Short), July 24, 1974. Atlantic Richfield Hanford Company, Richland, Washington Letter, Guidelines for PRTR Plutonium Accountability, February 27, 1970. R.A. Watrous to Distribution, Atlantic Richfield Hanford Company, Richland, Washington. Letter, Miscellaneous Treatment Centrifuge Sludge Sample (Including Summary of Results, 9/17/1975, August 19, 1975, from D.E. Turner to J.V. Panesko, Atlantic Richfield Hanford Company, Richland, Washington. Letter, Monthly Report October 1972, 1972-10-14, (internal letter from C.M. Walker to C.J. Francis), Atlantic Richfield Hanford Company, Richland, Washington. Letter, Monthly Report November 1973, November 14, 1973, (internal letter from W.R. Christensen to R.L. Walser), Atlantic Richfield Hanford Company, Richland, Washington Letter, NDA Measurement of 241-Z, D-5 Tank, 1976, (internal letter from G.A. Westsik to C.L. Owen, October 26), Atlantic Richfield Hanford Company, Richland, Washington. Official Use Only 217

RPP-RPT-50941, Rev. 0

Letter, Neutron Monitoring of 241-Z Tanks, January 10, 1963, D.J. Crawley to J.J. Courtney, General Electric, Richland, Washington. Letter, Plutonium in 242-Z Tanks, 1976, (internal letter, J.A. Compton and C.H. Kindle to D.L. Ubelacker, January 20), Atlantic Richfield Hanford Company, Richland, Washington. Letter, Plutonium Inventory Difference at Hanford, February 9, 1978, Energy Research and Development Agency, Richland, Washington. Letter, Prototype Centrifuge Shaft Bearings, July 16, 1973, M.J. Klem to R.E. Felt. Atlantic Richfield Hanford Company, Richland, Washington. Letter, Removal of W-1 Column Resin, 1977, (internal letter from R.J. Loberg to Distribution, March 15), Atlantic Richfield Hanford Company, Richland, Washington. Letter, Removal of W-14 Coupons, 1977, (internal letter, N.A. Brube to D.G. Harlow, February 28), Atlantic Richfield Hanford Company, Richland, Washington. Letter, Removal of Waste from Tank 216-Z-8, 1974, from D. A. Turner to D. G. Harlow, July 1), Atlantic Richfield Hanford Company, Richland, Washington. Letter, Results of Flow Restriction Test, October 17, 1973, ARCO letter D.A. Turner to D.T. Crawley, Atlantic Richfield Hanford Company, Richland, Washington. Letter, Safety Analysis Waste Liquid in W-2, W-3, W-4, W-5, and W-6 Tanks in 242-Z Building, 1977 (internal letter, R.J, Loberg to D.G. Harlow, May 9), Atlantic Richfield Hanford Company, Richland, Washington. Letter, Status of Centrifuge System June 22, 1973, R.D. Fox to D.L. Merick et.al. Atlantic Richfield Hanford Company, Richland, Washington. Letter, Tentative Indications from the Burning of a Portion of One Dow Button and Dissolution of Oxide, Campbell, M.H. and J.V. Panesko, 1971, Internal letter to L.M. Knights, October 13, 1971, Atlantic Richfield Hanford Company, Richland, Washington. Letter, W-1 Column, 1977, (internal letter, M.R. Fox to R.B. Gelman, February 9), Atlantic Richfield Hanford Company, Richland, Washington. Letter, Waste Treatment Processing without Americium Ion Exchange, September 21, 1976, (internal letter from R.E. Felt to R.B. Gelman, Atlantic Richfield Hanford Company, Richland, Washington.

Official Use Only 218

RPP-RPT-50941, Rev. 0

Letter, Weekly Report for Week Ending June 4, 1982, June 2, 1982, W.F. Skiba to E.J. Kosiancic. Letter, Z8 to 109-TX Transfers, 1974, from R.M. Smithers to P.E. Alley, et.al., August 9), Atlantic Richfield Hanford Company, Richland, Washington. Letter, Monthly Report November 1972, 1972-11-10, (internal letter from C.M. Walker to R. L. Walser), Atlantic Richfield Hanford Company, Richland, Washington. Liu, Y. K. Kathan, W. Saad and R. Prudhomme, 2007. Ostwald Ripening of -Carotene Nanoparticles, Physical Review Letters 98(3):036102-1 036102-4. M2100-05-174, 2004. Summary of NDA Result for 241-Z (internal letter, B.D. Keele to G.A. Johnston, November 8), Fluor Hanford, Inc., Richland, Washington. Machuron-Mandard, X. and C. Madic, 1996. Plutonium Dioxide Particle Properties as a Function of Calcination Temperature, Journal of Alloys and Compounds 235:216-224. Mandleburg, C.J., K.E. Francis, and R. Smith, 1961. The Solubility of Plutonium Trifluoride, Plutonium Tetra-Fluoride, and Plutonium(IV) Oxalate in Nitric Acid Mixtures, Journal of the Chemical Society 2464-2468. Martz, J.C., J.M. Haschke, and JL Stakebake, 1994. A Mechanism for Plutonium Pyrophoricity, Journal of Nuclear Materials 210:130-142. NMS-15232, 2005. Overview of Plutonium Oxide Particle Size Distributions and Respirable Fractions, Rev. 0, Fluor Hanford, Richland, Washington. Occurrence Report EM-RP--WRPS-TANKFARM2011-0004, February 14, 2011. Aqueous Waste Particle Size Transferred from Plutonium Finishing Plant May be Larger than Previously Evaluated, Washington River Protection Solutions, LLC, Richland, Washington. ORNL/TM-9565, 1985. Alpha Radiolysis and Other Factors Affecting Hydrolysis of Tributyl Phosphate, Lloyd, MH and RL Fellows, Oak Ridge National Laboratory, Oak Ridge, Tennessee. OSD-T-151-00008, 1994. Operating Specifications for the 204-AR Waste Unloading Facility, Rev/Mod E-1, Westinghouse Hanford Corporation, Richland, Washington. PFD-Z-180-00001, 1983. PRF Solvent Extraction, Rockwell Hanford Operations, Richland, Washington.

Official Use Only 219

RPP-RPT-50941, Rev. 0

PFD-Z-180-00002, 1983. Plutonium-Uranium Coextraction - Partitioning Flowsheet, Rev/Mod B-0, Rockwell Hanford Operations, Richland, Washington. PFD-Z-180-00004, 1987. Slag and Crucible Dissolver Flowsheet (Standard Feed) 236-Z Building, Rev. B-0, Rockwell Hanford Operations, Richland, Washington. Plutonium Reclamation Z Plant, Training Manual on Solvent Extraction, January 1978, Rockwell Hanford Operations, Richland, Washington. PNL-D-372, 1984. A Vulnerability Analysis of the Critical Mass Laboratory at Battelle, Pacific Northwest Laboratory, A Second Look, Pacific Northwest Laboratory, Richland, Washington. PNNL-11352, 1996. Tank SY-102 Waste Retrieval Assessment:Rheological Measurements and Pump Jet Mixing Simulations, Onishi, Y, et al., Pacific Northwest National Laboratory, Richland, Washington. PNNL-13524, 2001. Historical Time Line and Information About the Hanford Site, Pacific Northwest National Laboratory, Richland, Washington. PPD - 494 #2, 1972. Processing Statistics - February 1972, U.S. Atomic Energy Commission, Richland Operations Office, Richland, Washington. PPD-493 #6, 1972. US AEC RLO, Monthly Status and Progress Report - June 1972, U.S. Atomic Energy Commission, Richland, Washington. Process Test PRF-70-8, Operation of Continous Waste Treatment (HCE-637) for Start-up, October 10, 1970, Atlantic Richfield Hanford Company, Richland, Washington Process Test PRF-71-2, Leaching of W-2 Rings, M.H.Curtis, June 1971, Atlantic Richfield Hanford Company, Richland, Washington. PWM-550-2-DEL,1973. Nuclear Materials Management Branch Monthly Report for February 1972, U.S. Atomic Energy Commission Richland Operations Office, Richland, Washington. PWM-550 -3-DEL, 1973. Nuclear Materials Management Branch Monthly Report for March 1973, U S. Atomic Energy Commission Richland Operations Office, Richland, Washington. PWM-550-5-DEL, Nuclear Materials Management Branch Monthly Report for May 1973, U.S. Atomic Energy Commission Richland Operations Office, Richland, Washington.

Official Use Only 220

RPP-RPT-50941, Rev. 0

PWM-550-6-DEL, Nuclear Materials Management Branch Monthly Report for June 1973, U.S. Atomic Energy Commission Richland Operations Office, Richland, Washington. R80-1035, 1980. Request for Approval to Transport PR and 10-L Containers (Contract DEAC06-77RL0l030), (internal letter from J. H. Roecker to Mr. O. J. Elgert, March 28), Rockwell Hanford Operations, Richland, Washington. Rai, D.M. Yui, H.T. Schaef, and A. Kitamura, 2010. Thermodynamic Model for Bi(PO4)(cr) and Bi(OH)3(am) Solubility in the Aqueous Na+-H+-H2PO4HPO42OH- -Cl- -H2O System, J. Solution Chemistry, 39, 99-1019. Recovery Plan, 1985. 241-Z CPS Violation, September 24, Rockwell Hanford Operations, Richland, Washington. RFO-F-10, 1980. Plutonium-Uranium Coextraction-Partitioning Flowsheet, Rockwell Hanford Operations, Richland, Washington. RFO-F-2, 1980. Alternate Tributylphosphate Solvent Extraction Flowsheet for Recovery of Plutonium from Scrap, Rev. 1, Rockwell Hanford Operations, Richland, Washington. RFP-491, 1966. The Oxidation and Ignition of Plutonium, Thompson, M.A., The Dow Chemical Company, Rocky Flats Division, Golden, Colorado. RFP-503, 1965. Properties of Plutonium Dioxide, Moseley, J.D. and R.O. Wing, The Dow Chemical Company, Rocky Flats Division, Golden, Colorado. RFP-927, 1967, Properties of Plutonium Oxide, Part II. Rocky Flats, Molen, G.F and R.D. White The Dow Chemical Company, Golden, Colorado. RHO-CD-1129, 1980. 204-S Facility Shutdown Plan, Rockwell Hanford Operations, Richland, Washington. RHO-CD-1410, 1981. 242-T Evaporator Facility Shutdown/Standby Plan, Rockwell Hanford Operations, Richland, Washington. RHO-CD-194, A Study of the 234-5 Building Inventory Difference for the Years 1956 Through 1966, J.D. Anderson, Rockwell Hanford Operations, Richland, Washington. RHO-CD-492, March 9, 1979. Material Status Information Summary, Rockwell Hanford Operations, Richland, Washington. RHO-CD-505 RD, 1978. Synopsis of REDOX Plant Operations, Rockwell Hanford Operations, Richland, Washington. Official Use Only 221

RPP-RPT-50941, Rev. 0

RHO-CD-505 RD, 1978. Synopsis of Redox Plant Operations, Rockwell Hanford Operations, Richland, Washington. RHO-CD-757, 1981. Safety Analysis Report, 204-AR Waste Unloading Facility, Rockwell Hanford Operations, Richland, Washington. RHO-LD-114, 1981. Existing Data on the 216-Z Liquid Waste Sites, Rockwell Hanford Company, Richland, Washington. RHO-MA-116, 1983. PUREX Technical Manual Addendum I Plutonium Oxide Production and Rework Facility, Rockwell Hanford Operations, Richland, Washington. RHO-MA-246, 1980. Plutonium Reclamation Facility Engineering Training Manual, Rockwell Hanford Operations, Richland, Washington. RHO-RE-EV-46 P, 1984. The 216-Z-8 French Drain Characterization Study, Rockwell Hanford Operations, Richland, Washington. RHO-RE-EV-75, 1985. FY 85 PRF Weapons Grade Campaign Report, Rockwell Hanford Operations, Richland, Washington. RHO-RE-EV-85, 1986. 1985 Remote Mechanical C (RMC) Line Campaign Report, Rockwell Hanford Operations, Richland, Washington. RHO-RE-EV-86, 1986. FY86 PRF Weapons Grade Campaign Report, Rockwell Hanford Operations, Richland, Washington. RHO-RE-FL-2, 1985. PUREX Flowsheet Reprocessing N Reactor 6% Fuel, Rev. 1, Rockwell Hanford Operations, Richland, Washington. RHO-RE-ST-3 P, 1983. Effects of Hanford High-Level Waste Components on the Solubility of Cobalt, Strontium, Neptunium, Plutonium, and Americium, Rockwell International, Richland, Washington. RHO-RE-ST-53, 1987. Solubility of PuO2xH2O in PUREX Plant Metathesis Solutions, Rockwell Hanford Operations, Richland, Washington. RHO-SD-CP-PTP-035, 1987. PUREX Headend Changes to Metathesis and CRW Waste Processing, Keith, L.M. and A.G. Westra, Rockwell Hanford Operations, Richland, Washington. RHO-SS-MA-30, 1987. Computerized Model for Calculating the Variance of Inventory Difference, Rev. 1, Rockwell Hanford Operations, Richland, Washington. Official Use Only 222

RPP-RPT-50941, Rev. 0

RHO-SS-SR-128, 1984. Plutonium Reclamation Facility Material Balance and LEID Summary November 8, 1983- November 30, 1984, G.P. Kodman, Rockwell Hanford Operations, Richland, Washington. RHO-SS-SR-128, Rev. 1, 1985. Plutonium Reclamation Facility Material Balance and LEID Summary December 1, 1984 June 11, 1985 J.V. Matkevich, R.C. Martinson, D.H. Nichols, Rockwell Hanford Operations, Richland, Washington. RHO-SS-VS-14, 1983. Resolution of Material Balance Area 102 Plutonium Reclamation Facility, Unusual Inventory Difference, D.E. Six, Rockwell Hanford Operations, Richland, Washington. RHO-ST-17, 1979. Distribution of Plutonium and Americium beneath the 216-Z-1A crib: A Status Report, Rockwell Hanford Operations, Richland, Washington. RHO-ST-21, 1978. Report on Plutonium Mining Activities at 216-Z-9 Enclosed Trench, Rockwell Hanford Operations, Richland, Washington. RHO-ST-44, 1982. 216-Z-12 Transuranic Crib Characterization: Operational History and Distribution of Plutonium and Americium, Rev. 0, Rockwell Hanford Operations, Richland, Washington. RL-CAO-165, 1965. Historical Review Plutonium Ratio Receipt Measurement July 1960 through December 1964, General Electric, Hanford Atomic Products Operation, Richland, Washington. RL-SA-40, Status of Fission Product Recovery at Hanford, S.J. Beard and P.W. Smith, August 1965, Hanford Atomic Products Operation, Richland, Washington. RL-SEP-396, 1965. 242-T Evaporator Facility Information Manual, Hanford Atomic Products Operation, Richland, Washington. RL-SEP-397, 1965. Chemical Flowsheets for the Decladding and Dissolution of PRTR 0.48 Percent PuO2-UO2 Fuel, General Electric, Hanford Atomic Products Operation, Richland, Washington. RL-SEP-618, 1965. Chemical Processing Department Monthly Report for June, 1965, General Electric, Hanford Atomic Products Operation, Richland, Washington. RL-SEP-654, 1965. Chemical Processing Department Monthly Report for July, 1965, General Electric, Hanford Atomic Products Operation, Richland, Washington.

Official Use Only 223

RPP-RPT-50941, Rev. 0

Rockwell Occurrence Report 78-74, Leakage of Organic Liquid in 242-Z Tank Room, July 20, 1978, Rockwell Hanford Operations, Richland, Washington. RPP-15408, Origin of Wastes in C-200 Series Single-Shell Tanks, Rev. 0, CH2M HILL Hanford Group, Inc., Richland, Washington. RPP-17017, 2003. 242-T and 242-TA Vault Information from Engineering Notebooks and Archived Data, Rev. 0, CH2M HILL Hanford Group, Inc., Richland, Washington. RPP-RPT-23177, Origin of Waste in Tank 241-AW-105, Rev. 0, CH2M HILL Hanford Group, Inc., Richland, Washington. RPP-RPT-4307, Disselkamp, R. A. 2009c, 2009 Auto-TCR for Tank 241-TX-118, RPP-RPT4307, Rev. 0, Washington River Protection Solutions, LLC, Richland, Washington. RPP-RPT-43071, 2009. 2009 Auto-TCR for Tank 241-SY-101, Rev. 0, Washington River Protection Solutions, LLC., Richland, Washington. RPP-RPT-43072, 2009. 2009 Auto-TCR for Tank 241-SY-102, Rev. 0, Washington River Protection Solutions, LLC., Richland, Washington. RPP-RPT-43179, 2009. 2009 Auto-TCR for Tank 241-TX-105, Rev. 0, Washington River Protection Solutions, LLC, Richland, Washington. RPP-RPT-43183, 2009. 2009 Auto-TCR for Tank 241-TX-109, Rev. 0, Washington River Protection Solutions, LLC, Richland, Washington. RPP-RPT-43192, 2009. Auto-TCR for Tank 241-TX-118, Rev. 0, Washington River Protection Solutions, LLC, Richland, Washington. RPP-RPT-44643, 2010. Derivation of Best-Basis Inventory for Tank 241-SY-102, Rev. 1, Washington River Protection Solutions, LLC, Richland, Washington. RPP-RPT-44643, 2011. Derivation of Best-Basis Inventory for Tank 241-SY-102, Place, D.E Washington River Protection Solutions, LLC., Richland, Washington. Rudisill, T, 2011. Personal e-mail communication PuF4.2.5H2O, 23 June 2011, to C Delegard. Pacific Northwest National Laboratories, Richland, Washington. Scheele, R.D. and M.E. Peterson, 1990. Results of the Characterization of Samples of Waste From Double-Shell Tank 102-SY, Correspondence N0. 9000455, Letter from M.E. Peterson to A.J. Diliberto, January, 1990, Battelle, Pacific Northwest National Laboratories, Richland, Washington. Official Use Only 224

RPP-RPT-50941, Rev. 0

Schindler, PW. 1967. Heterogeneous Equilibria Involving Oxides, Hydroxides, Carbonates, and Hydroxide Carbonates. Pages 196 to 221 of Equilibrium Concepts in Natural Water Systems, Advances in Chemistry Series 67, W Stumm, Symposium Chairman, American Chemical Society, Washington, DC, See also W Stumm and JJ Morgan, 1996, Aquatic Chemistry Chemical Equilibria and Rates in Natural Waters, 3rd edition, pages 413-414 and 806-809, John Wiley and Sons, New York, New York. SD-CP-DTR-006, 1984. Plutonium Recovery and TRU Removal from PRF CAW, Rev. 0, Rockwell Hanford Operations, Richland, Washington. SD-CP-ES-043, 1984. The Destruction of Oxalic Acid by Nitric Acid at the Boiling Point, Rev. 0, Rockwell Hanford Operations, Richland, Washington. SD-CP-ES-096, 1987. Recommendations for Improved Dissolution of Plutonium Oxide Scrap at the Plutonium Reclamation Facility, Rev. 0, Rockwell Hanford Operations, Richland, Washington. SD-CP-SAR-013, 1987. Slag and Crucible Dissolver Hazard Identification and Evaluation, Rev. 1, Westinghouse Hanford Company, Richland, Washington. SD-CP-TI-122, 1988. Plutonium Reclamation Facility Engineering Technical Reference, Rev. 0, Westinghouse Hanford Company reissue of RHO-MA-246, Richland, Washington. SD-CP-TRP-016, 1986. TRUEX Solvent Extraction and Solution Clarification Results for FY-1986 PFP Waste, Rockwell Hanford Operations, Richland, Washington. SD-WM-DTR-020, 1988. TRUEX Process Demonstration Tests with Plutonium Finishing Plant Waste, Waste Management Hanford, Inc., Richland, Washington. SGW-35955, 2008. Inventory Estimates for Sludge Currently in Tank 241-Z-361, Rev. 0, Fluor Hanford, Inc., Richland, Washington. SGW-39385, 2009. Z Plant Complex Waste Streams Discharged to the Soil Column (1949 to 1973), Rev. 0, CH2MHILL Plateau Remediation Company, Richland, Washington. SGW-42004, 2009. Summary Report for Selected Boreholes Within the 216-Z-1A Tile Field, 216-Z-9 Trench, Rev. 0, CH2MHILL Plateau Remediation Company, Richland, Washington. Shannon, D, 1976. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides, Acta Crystallographica Section A A32:751-767. Official Use Only 225

RPP-RPT-50941, Rev. 0 Shaw, DJ, 1980. Colloid and Surface Chemistry, 3rd edition, Butterworth & Co. Ltd., New York, New York. SRNL-L7100-2009-00046, 2009. Review of Plutonium Oxidation Literature, Savannah River National Laboratory, Aiken, South Carolina, Available at: http://sti.srs.gov/fulltext/SRNL-L7100-2009-00046R1.pdf. SSSM-1797, December 31, 1970. Survey No. 41, Report on the Safeguards and SS Materials Management Practices and Procedures at Atlantic Richfield Hanford Company Facility HVA and HVB Richland Washington, U.S. Atomic Energy Commission, Richland, Washington. Stakebake, J.L. 1981. Kinetics for the Reaction of Hydrogen with a Plutonium-1 Weight Percent Gallium Alloy Powder, Journal of the Electrochemical Society 128(11):23832388. Stakebake, J.L. and L.M. Steward. 1973. Water Vapor Adsorption on Plutonium Dioxide, Journal of Collo,d and Interface Science 42(2):328-333. Stewart, K, 1961. Vixen A Trials: Experiments to Study the Release of Particulate Materials During the Combustion of Plutonium, Uranium and Beryllium in a Petrol Fire. AWRE Report T15/60, Atomic Weapons Research Establishment, Aldermaston, United Kingdom. As cited by Haschke 1992. US Patent 2,931,702., Duffield, R.B., 1960, Metathesis of Plutonium Carrier Lanthanum Fluoride Precipitate with an Alkali, US Patent Office, Washington, DC. W.E. Matheison and G.J. Raab, 8/5/1969, Draft #4 - undocumented, PRTR Fuel Processing at the Purex Plant, 0.5% and 1.0% PuO2 in UO2 Fuel Wagman, D.D., W.H. Evans, V.B. Parker, R.H. Schumm, I. Halow, S.M. Bailey, K.L. Churney, and R.L. Nuttall. 1982. The NBS Tables of Chemical Thermodynamic Properties Selected Values for Inorganic and C1 and C2 Organic Substances in SI Units. Journal of Physical and Chemical Reference Data 11, Supplement 2. WHC-88-00071, 1988. FY 1987 PRF Weapons Grade Campaign Report, Westinghouse Hanford, Richland, Washington. WHC-EP-0342, Addendum 31, UC-630, 1990, 209-E Laboratory Reflector Water StreamSpecific Report, Westinghouse Hanford Company, Richland, Washington. WHC-EP-0793, 1994. Estimation of Plutonium in Hanford Site Waste Tanks Based on Historical Records, Rev. 0, Westinghouse Hanford Company, Richland, Washington. Official Use Only 226

RPP-RPT-50941, Rev. 0

WHC-MR-0132, 1990. History of 200 Area Tank Farms, J.D. Anderson, Westinghouse Hanford, Richland, Washington. WHC-MR-0265, 1990. PFP SAR Chapter 9 Questions on 242-Z Inventory and Ion Exchange Resins, R.M. Marusich to PFP Engineering, Westinghouse Hanford, Richland, Washington. WHC-SD-CP-TI-204, 1996. Stabilization of Polycubes Engineering Study, Rev. 0A, Westinghouse Hanford Company, Richland, Washington. WHC-SD-DD-ES-008, 1989. Recommendations for the Sampling and Decommissioning of Tank 241-CX-72,Westinghouse Hanford Company, Richland, Washington. WHC-SD-DD-TI-039, 1989. Tank 241-CX-71 Preliminary Waste Characterization, Rev. 0, Westinghouse Hanford Company, Richland, Washington. WHC-SD-DD-TI-057, 1991. Summary of Radioactive Underground Tanks Managed by Hanford Restoration Operations, Rev. 0, Westinghouse Hanford Company, Richland Washington. WHC-SD-EN- ES-019, 1992. Semi-Works Aggregate Area Management Study Technical Baseline Report, Rev. 0, Westinghouse Hanford Company, Richland, Washington. WHC-SD-WM-ER-366, 1995. Tank Characterization Report for Double-Shell Tank 241-SY-102, Winters, W.I. and A.T. DiCenso, Westinghouse Hanford Company, Richland, Washington. WHC-SD-WM-TI-614, 1995, Waste Status and Transaction Record Summary for the Southwest Quadrant of the Hanford 200 Area, Westinghouse Hanford Company, Richland, Washington. WHC-SD-WM-TI-615, Waste Status and Transaction Record Summary for the Northeast Quadrant of the Hanford 200 Area, Westinghouse Hanford Company, Richland, Washington. WMP-32037, 2007. Acceptable Knowledge Summary Report for 209-E Critical Mass Laboratory Mixed Transuranic Debris, Rev. 0, Fluor Hanford, Inc., Richland, Washington. WRPS-1101477, 2011. Plutonium Oxide Resolution Plan, (external letter from W.L. Isom to M.T. Coyle and R.L. Garrett, May 3), Washington River Protection Solutions, Richland, Washington.

Official Use Only 227

RPP-RPT-50941, Rev. 0

Yamamura, T, A Kitamura, A Fukui, S Nishikawa, T Yamamoto, and H Moriyama, 1998. Solubility of U(VI) in Highly Basic Solutions, Radiochimica Acta 83:139-146. Yusov, A.B., A. Yu Garnov, V.P. Shilov, I.G. Tananaev, M.S. Grigorev, and N.N. Krot. 2000. Plutonium(IV) Precipitation from Alkaline Solutions. I: Effect of Precipitation and Coagulation Conditions on Properties of Hydrated Plutonium Dioxide PuO2xH2O. Radiochemistry 42(2):151-156, See also Plutonium(IV) Precipitates Formed in Alkaline Media in the Presence of Various Anions, NN Krot, VP Shilov, AB Yusov, IG Tananaev, MS Grigoriev, AYu Garnov, VP Perminov, and LN Astafurova, 1998, PNNL-11901, Pacific Northwest National Laboratory, Richland, Washington. Available at: http://www.osti.gov/bridge/purl.cover.jsp?purl=/665911-NwS0BW/webviewable/. Yusov, A.B., A. Yu Garnov, V.P. Shilov, I.G. Tananaev, M.S. Grigorev, and N.N. Krot. 2000a. Plutonium(IV) Precipitation from Alkaline Solutions. I: Effect of Precipitation and Coagulation Conditions on Properties of Hydrated Plutonium Dioxide PuO2xH2O. Radiochemistry 42(2):151-156. Yusov, A.B., A. Yu Garnov, V.P. Shilov, I.G. Tananaev, M.S. Grigorev, and N.N. Krot. 2000b. Plutonium(IV) Precipitation from Alkaline Solutions. I: Effect of Anions on Composition and Properties of Hydrated Plutonium Dioxide PuO2xH2O. Radiochemistry 42(2):157-160. Yusov, A.B., N.A. Budantseva, A.V. Ananev, and A.M. Fedoseev. 2000. Coprecipitation of Aluminum with Hydroxides of Tetra-, Penta-, and Hexavalent Actinides. Radiochemistry 42(5):413-416. Zimmer, E and J Borchardt, 1986. Crud Formation in the PUREX and THOREX Processes, Nuclear Technology 75:332-337.

Official Use Only 228

RPP-RPT-50941, Rev. 0

10.0

APPENDICES

Appendices follow.

Official Use Only 229

RPP-RPT-50941, Rev. 0

APPENDIX A SCOPE OF THE TWO TEAMS

Official Use Only A-1

RPP-RPT-50941, Rev. 0

Official Use Only A-2

RPP-RPT-50941, Rev. 0

Official Use Only A-3

RPP-RPT-50941, Rev. 0

Official Use Only A-4

RPP-RPT-50941, Rev. 0

Official Use Only A-5

RPP-RPT-50941, Rev. 0

Official Use Only A-6

RPP-RPT-50941, Rev. 0

Official Use Only A-7

RPP-RPT-50941, Rev. 0

Official Use Only A-8

RPP-RPT-50941, Rev. 0

Official Use Only A-9

RPP-RPT-50941, Rev. 0

Official Use Only A-10

RPP-RPT-50941, Rev. 0

Official Use Only A-11

RPP-RPT-50941, Rev. 0

Official Use Only A-12

RPP-RPT-50941, Rev. 0

Official Use Only A-13

RPP-RPT-50941, Rev. 0

APPENDIX B LIST OF ASSUMPTIONS TO VALIDATE AND COMMENT

Official Use Only B-1

RPP-RPT-50941, Rev. 0

LIST OF ASSUMPTIONS TO VALIDATE AND COMMENT The following is a list of assumptions and beliefs that the authors of the 2010 Draft Report (24590-CM-HC4-W000-00176-T02-01-00002 Rev 00B) have identified as worth having validated or confirmed in order to provide additional clarification and technical support related to the 2010 Draft Report. Items on the list are separated into Statements of Fact, Assumptions, and Areas Needing Further Investigation. It is recommended that specific details of documentation that will describe acceptable evidence to support a successful validation be established to help facilitate the validation effort. Statements of Fact 1. Historical PFP aqueous waste accountability did not adequately account for plutonium solids entrained within the waste solution, resulting in Book plutonium accountability records that were significantly biased low and under estimating the amount of plutonium discharged in PFP waste solutions. PFP efforts to minimize plutonium solids within product and waste solutions utilizing filters and centrifuges were fraught with difficulties and frequent times of poor performance, resulting in plutonium solids continuing to be present in these streams. The plutonium solids that could have been entrained within waste streams at PFP are best represented by the following materials: plutonium (IV) oxalate, plutonium tetrafluoride, and either pure or impure plutonium oxide. The impure plutonium oxide would be from plutonium metal oxidation, scrap dissolver heels, and ash residues from miscellaneous plutonium scrap incineration. The time interval that PFP waste was transferred to Tank Farms, the pathway utilized, and the primary tanks (SY-102 and TX-118) that received PFP wastes after April 1973 are known. Fundamental chemistry used in metal lines and solvent extraction systems remained reasonably constant over the operating lifetime of PFP. That is, the processes used to convert Pu-nitrate into metal used the same basic steps in RGL, RMA, and RMC lines. The basic approach used in Pu scrap reprocessing was based on acid dissolution of Pu-oxide or other Pu solids. Solvent extraction chemistry remained reasonably constant from RECUPLEX to PRF. 242-Z provided Am-141 recovery and enhanced recovery of soluble Pu in the high-salt waste stream. The chemistries of the high-salt waste stream coming from the solvent extraction systems would have reflected scrap recovery operations ongoing at the time of waste generation. Scrap recovery during the time waste streams were going to the tank farms (after 4/73) may be significantly different from high-salt waste streams going to the soil column preApril 1973.

2.

3.

4.

5.

6.

Official Use Only B-2

RPP-RPT-50941, Rev. 0

Assumptions 1. The major source for Pu solids in the high-salt waste stream likely originated from the entrainment of dissolver heels solids in product solutions that were sent to solvent extraction for plutonium recovery and purification. These heels were the remnant after dissolution attempts of impure residues and oxides for plutonium recovery. Dissolution followed by solvent extraction was a routine method in scrap reprocessing over the lifetime of PFP operations. This included the dissolution of impure oxide that was routinely generated from the burning of impure plutonium metal. Thus, the chemical and physical characteristics of the Pu-oxide particles discharged in the CAW waste stream likely remained reasonably constant over time. Additional small amounts of plutonium would have been tied up in various organic based particles that also were entrained in the solvent extraction waste solutions. The amount of under-accounted for plutonium in waste streams that were transferred to Tank Farms is significant; and it is reasonable that the actual amount of plutonium transferred to Tank Farms falls within the range between the Book amounts of plutonium transferred, totaling ~30 Kgs, and a high estimate of 130 Kgs. Any total estimate of plutonium discard to Tank Farms from the PFP of less than 30 Kgs or over 130 Kgs is difficult to support, and the most likely amount is ~65 Kgs.

2.

Areas Needing Further Investigation 1. It is believed that additional useful information may be contained within either classified documents or documents that are now part of the Hanford-net that the authors did not have access to. These resources would have included historical Hanford documents, laboratory notebooks, letters, technical letter books, weeklies, monthlies, operation reports, etc. The information sought would have been related to such things as: a. b. c. d. e. f. The amount of plutonium discarded in waste streams (to cribs, trenches, and Tank Farms) Particle size analysis of sludge and solids in waste streams & discharges to cribs and trenches Laboratory results on the particle size of dissolver heels Operation and use of filters and centrifuges to minimize solids entrainment in aqueous streams Accountability techniques, procedures, and difficulties associated with waste solution discards Information related to material unaccounted for MUF associated with liquid waste discards from PFP, and related to plutonium solution transfers between PFP and other facilities (REDOX, PUREX, 231-Z, 224-T, etc).

Official Use Only B-3

RPP-RPT-50941, Rev. 0

g. h. i. j. k. l. 2.

Particle size distribution, particle density, and SEM analyses of miscellaneous plutonium materials encountered during PFP historical operations Waste solution treatment prior to discard General issues relating to plutonium solids buildup occurring in process lines and process tanks, 241-Z Sump tanks (TK-D4 TK-D8), 241-Z-361 Settling tank. Solids buildup within PFP waste solution transfer pathways, lines and tanks All relevant historical PFP flowsheets. Historical PFP plant physical configurations.

Reliability of the Pu values and transfer timeline of PFP waste transferred to Tank Farms as shown in the Roetman Table in WHC-EP-0793 Rev 0, Figure 4 and Table 4 of the 2010 Draft Report. Several references in the PFP Letter Book Index list historical letters that address particle size in dissolution tests and the PSD of dissolver heels. The PSD of these heels are of interest as they would be examples of typical solids that could be found in PFP sludge. These letters are: a. Internal Letter #65454-85-120, Characterization, Particle Size and Dissolutions Tests of Three Rocky Flats Oxides, C.H. Delegard and D. G. Bouse to K. S. Kalkofen, 8/30/1985. Internal Letter # 65930-85-A-090, MT-5 Dissolver Solution/Heel Analysis and Heel Particle Size Distribution, K. S. Kalkofen to C. H. Delegard, 7/15/1985.

3.

b. 4.

Evaluate documentation that plutonium tetrafluoride and plutonium oxalate solids occasionally ended up in PFP waste solutions, cribs, trenches, and may have been transferred to Tank Farms or any of the canyon facilities and subsequently to Tank Farms. Evaluate documentation that demonstrates the samples from 102-SY, which are analyzed and reported in CH2M-0400872, are adequate representations and characterization of the plutonium solids in the 102-SY tank and are better than just a shot-in-the-dark. This is very important since the results of these analyses are being used as part of the justification for a 10 micron bounding particle size for Pu oxide. The presence of Pu particles in high-salt liquid waste discharges likely remained reasonably constant over the operational lifetime of the Z Plant complex. That is, if Pu-oxide particles were lost to the soil column from 1954 until 1973, then it is likely these types of particles would continue to have been generated and that some of them could have been sent to the tank farms, depending upon the successfulness of any implemented solids reduction strategy.

5.

6.

Official Use Only B-4

RPP-RPT-50941, Rev. 0

7.

Careful review of Z Plant (and maybe tank farm) records in the 1968 to 1990 time period would clarify upgrades in waste treatment activities in Z Plant that were implemented to reduce Pu solids in waste streams going to the tank farms. These activities included the use of centrifuges and filtering systems to reduce the quantities of Pu solids in various solutions, including the liquid wastes going to the tank farm. Review of operating reports and documents may identify how effective these solids removal strategies were. Document that the amount of under-accounted for plutonium in waste streams that were transferred to Tank Farms is significant; and it is reasonable that the actual amount of plutonium transferred to Tank Farms falls within the range between the Book amounts of plutonium transferred, totaling ~30 Kgs, and a high estimate of 130 Kgs. Any total estimate of plutonium discard to Tank Farms from the PFP of less than 30 Kgs or over 130 Kgs is difficult to support, and the most likely amount is ~65 Kgs. Develop documentation that demonstrates the plutonium oxide particles in the samples analyzed in CH2M-0400872 actually/probably came from PFP, rather than having come from other facilities discharging Pu contaminated waste streams or may grown while being stored in the tanks.

8.

9.

Official Use Only B-5

RPP-RPT-50941, Rev. 0

APPENDIX C 5 REPORTS

Report 1 Ostwald Ripening and Its Effect on PuO2 Particle Size in Hanford Tank Waste Report 2 Metathesis of Plutonium(IV) Oxalate and Plutonium(III)/(IV) Fluorides in Alkaline Media to Form PuO2xH2O Report 3 Plutonium(IV) Oxide and Unburned Plutonium Metal from Operations to Burn Plutonium Metal (BURNT METAL) Report 4 Interfacial Crud Disposition in Alkaline Tank Waste Report 5 Coprecipitation from Nitric Acid Media of Plutonium(IV) with Uranium(VI) in Alkaline Solution

Official Use Only C-1

RPP-RPT-50941, Rev. 0

REPORT 1 OSTWALD RIPENING AND ITS EFFECT ON PUO2 PARTICLE SIZE IN HANFORD TANK WASTE

Ostwald Ripening and Its Effect on PuO2 Particle Size in Hanford Tank Waste Ostwald ripening is a process by which small particles dissolve, due to their enhanced solubility arising from their high surface area and energy, and larger particles grow to result in a coarsening of the solid particle or an overall increase in the mean particle size (Liu et al. 2007). Ostwald ripening is not expected to occur to a significant extent for PuO2 particles in alkaline Hanford tank waste. Instead, any micronscale crystalline plutonium oxide found in the tank waste almost certainly entered the tank in that form and, because of the low solubility of plutonium phases in alkaline media, was not formed by processes within the tank. Plutonium that entered the alkaline tank waste by precipitation through neutralization from acid solution is present as 2- to 3-nm (0.002 to 0.003-m) scale PuO2xH2O [plutonium(IV) hydrous oxide] crystallite particles and grows from that point at exceedingly slow rates. These conclusions are reached by both general considerations of the phenomenon of Ostwald ripening and specific observations of the behaviors of PuO2 and PuO2xH2O upon aging in alkaline solution. General Considerations A precipitate/solvent system at saturation appears to be static, but is in fact dynamic, with both dissolution and crystallization of the precipitate occurring simultaneously and at the same or nearly the same rate. Ostwald ripening occurs because the solubility of the smaller precipitate particles is greater than the solubility of the larger particles. The smaller particles have higher specific surface area than the larger particles and thus a greater proportion of the imperfect crystal structure that is present at the solidsolution interface. As a result of their greater solubility, the smaller particles tend to dissolve and their solutes deposit onto the surfaces of the larger and less soluble particles. The rate of crystal growth by Ostwald ripening is proportional to the solubility of the precipitate as described mathematically by Liu and colleagues (2007; and many other references). For precipitates with very low solubility, like PuO2, the rates of dissolution and crystallization are so slow that re-distribution of particle sizes due to Ostwald ripening may take place only on geologic time scales, not over the time scale of tank waste storage. An example cited in the literature (Lee et al. 2005) states that Ostwald ripening is not an important mechanism for tin oxide systems, as expected by the low solubility of this oxide. Another example states that In a sol of a highly insoluble substance, such as silver iodide hydrosol, this phenomenon [Ostwald ripening] will be of little consequence, since both large and small particles have extremely little tendency to dissolve (Shaw 1980). The potential growth of PuO2 particles by Ostwald ripening is further attenuated in Hanford tank waste sludges which contain abundant low-solubility metal hydroxide surfaces, such as those of ferric hydroxide, that have high affinity for plutonium, thus lowering plutoniums effective concentration, while the metal hydroxide particles themselves provide physical barriers to impede dissolved plutonium from diffusing from dissolving PuO2 particles to the growing PuO2 particles. The only evidence of potential re-crystallization of PuO2 occurring in tank waste is the observation (Cooke 2011) that some plutonium-rich particles in the tank 241-SY-102 samples (Callaway and Cooke 2008) have crystalline structures (including sharply defined edges) that suggest slow growth of crystals. However, an alternative explanation for the existence of these crystalline structures is that the particles entered the waste tank that way and have neither grown nor dissolved appreciably since that time.

Official Use Only C-2

RPP-RPT-50941, Rev. 0

Effect of Aging on PuO2xH2O Particle Size Information about the rate of Ostwald ripening of plutonium hydrous oxide, PuO2xH2O can be deduced from results of long-term studies of the plutonium concentrations observed in alkaline solution of variable NaOH concentration containing PuO2xH2O, created by making plutonium nitrate solutions alkaline in NaOH solution, and by knowledge of the effects of particle size on the solubilities of PuO2xH2O and fully crystalline PuO2. According to the work of Schindler (1967), later extended to understanding the solubility of thorium(IV) oxide/hydroxide according to particle size (Bundschuh et al. 2000; Fanghnel t and Kim 2002(2)), the solubility products of colloidal particles present at time t, K sp (colloid) , and fully
0 crystallized large crystals, K sp (l arg e crystal) , of the same chemical composition and structure are related

to the size, dt, of the colloidal particles at time t by Equation 1 (derivation shown in Appendix 1):
t 0 log K sp (colloid) log K sp (l arg e crystal) =

x d t (nm)

Equation 1

The value of x is a function of the molecular weight of the solid (PuO2 in the present case), PuO2 particle density, the radii of the constituent Pu4+ and O2- ions, and the surface area-to-volume ratio factor of 6 for the assumed spherical PuO2xH2O colloidal particles of diameter dt. As shown in Appendix 1, the value of x for PuO2 is 18.46 nm. Critically reviewed values of K 0 (l arg e crystal) for crystalline PuO2, PuO2 (cr), and K 0 (colloid) , for the sp sp amorphous hydrous oxide of PuO2 initially formed before significant aging has occurred, PuO2 (am,hydr.) or PuO2xH2O, are 10-64.04 and 10-58.33, respectively (Guillaumont et al. 2003). By input of values for x, 0 K sp (l arg e crystal) , and K 0 (colloid) to Equation 1, the birth size, d0, of the freshly formed PuO2xH2O sp particle may be determined as shown in Equations 2 and 3.
0 0 log K sp (colloid) log K sp (l arg e crystal) = 58.33 + 64.04 = 5.71 =

18.46 d 0 (nm)

Equation 2

d 0 (nm) =

18.46 = 3.2 nm 5.71

Equation 3

The d0 = 3.2-nm birth size of PuO2xH2O predicted by this analysis compares closely with the 3-nm size observed for PuO2xH2O solids precipitated with ammonium hydroxide and then air-dried at 70C as measured by XRD reflection line broadening (Westrum 1949). It is also close to the 2.5-nm particle size for PuO2xH2O precipitated by neutralizing Pu(IV) nitrate from alkaline solution after 5 hours of aging at room temperature and also for PuO2xH2O obtained by 2 hours of aging at 75C (Yusov et al. 2000). The particle size in the tests by Yusov and colleagues (2000) was measured by XRD line broadening from the 111 plane appearing at 28.55 2-. The research also found that the NaOH concentration (at 0.1, 1, and 10 M) had negligible effect on particle size in tests done with 2 hours of 75C aging. Preparation of the PuO2xH2O under 180-200C hydrothermal conditions produced larger 4.5-nm particles.

(2) Note that Th(IV) is chemically analogous to the Pu(IV) system.

Official Use Only C-3

RPP-RPT-50941, Rev. 0

Previously unpublished results of separate tests by Cal Delegard in the ~19831987 time frame include XRD patterns of PuO2xH2O, prepared by precipitating plutonium nitrate solutions in NaOH solution, at various aging times. The XRD line broadening of the PuO2xH2O produced by precipitating Pu(IV) nitrate in 5 M NaOH solution and separated after 4 days of room temperature aging for the 111 plane reflection at 28.55 2- indicates a particle size of ~2.1-nm compared with the 3.2-nm birth size predicted by the thermodynamic analysis shown in Equation 3 and the 2.5-nm PuO2xH2O particle size measured by Yusov and colleagues (2000). The particle size estimate was obtained using the Scherrer equation. (3) With ~14 months of aging under room temperature (~22C) 5 M NaOH, the PuO2xH2O particles prepared from Pu(IV) nitrate increased size marginally to ~2.6 nm as shown by the slight narrowing of the 28.55 2- peak. Similarly broad XRD diffraction patterns as found in 5 M NaOH after 14 months were observed for PuO2xH2O prepared from Pu(IV) nitrate in 2 M and 10 M NaOH and aged 14 months. After ~38 months of aging, the particle size of PuO2xH2O isolated from NaOH solution was ~3.0 nm. Greater particle size was observed for PuO2xH2O prepared by precipitating slowly reducing Pu(VI) (added as nitrate) in 1 M NaOH and aged about 9 months. In this case the particle size was about 11 nm as determined by XRD line broadening. The greater particle size shown for the Pu(VI)-origin material is attributed to the higher concentration of the dissolved plutonium, present largely as Pu(VI), compared with the Pu(IV) present in the other testing. The higher dissolved plutonium concentration present while the Pu(VI) was being radiolytically reduced to the much less soluble Pu(IV) allowed slower precipitation and greater growth to occur on the developing PuO2xH2O particles. The PuO2xH2O XRD patterns for particles prepared from Pu(IV) nitrate in 5 M NaOH after 4 days and 14 months, from Pu(IV) nitrate in NaOH at 38 months, and from Pu(VI) nitrate in 1 M NaOH after 9 months aging are shown in Appendix 2. Other studies showed that freshly formed PuO2xH2O particles (aged <1 hour) prepared by addition of 10 to 130 g/L Pu(IV) in 1 to 3 M nitric acid to excess ammonium hydroxide solution (pH ~12) were less than 2 nm as estimated by electron micrography (Haire et al. 1971). The precipitates were better defined and more discrete after washing to pH 6 to 8 followed by 4 hours of reflux at 95C to 100C. The particles had the XRD pattern of PuO2 while XRD line broadening showed the particle size to be less than 2.5 nm. Overall, the published laboratory studies, measurements from previously unpublished studies, and theoretical predictions indicate that the PuO2xH2O birth particles produced by making Pu(IV) nitrate solution alkaline are about 2 to 3 nm in diameter. Particle size is expected to increase with aging by Ostwald ripening. According to Equation 1, this aging t is reflected in the decreased solubility product, K sp (colloid) , of the growing particle as the aging time, t, increases. As shown in Appendix 1, the aging colloids influence is also reflected in the colloids t decreased solubility [i.e., plutonium concentration, K s, 4 (colloid ) ] according to Equation 4.

(3) d(nm) =

K 0.9 0.154 nm 0.9 0.154 nm = = cos (radians, 2) cos(0.249 radians) (radians, 2) 0.969

where K is a shape factor (typically around 0.9), is the wavelength of the copper K X-ray (0.154056 nm), is the full width, in radians, 2-, at the half maximum of (in this case) the 111 plane diffraction peak at 28.55 2-, and is the Bragg angle (28.55/2 or 0.249 radians) of the 111 plane diffraction peak.

Official Use Only C-4

RPP-RPT-50941, Rev. 0

t log K s , 4 (colloid) log K 0, 4 (l arg e crystal) = 5.71 + log (so lub ility) = s

18.46 d t (nm)

Equation 4

The relationship of the change in PuO2xH2O particle size as solubility decreases given in Equation 4 is illustrated in Figure 1. For example, a 100-fold decrease in plutonium solubility from that of a 3.2-nm birth particle indicates that the PuO2xH2O particle has grown to ~5.0 nm as shown in Equations 5 and 6.
t 0 log K s , 4 (colloid) log K s, 4 (l arg e crystal) = 5.71 + log (so lub ility) = 5.71 2.00 =

18.46 d t (nm)

Equation 5

d t (nm ) =

18.46 = 5.0 nm 5.71 2.00

Equation 6

1000

100

dt, nm
10
Birth particle, d0 = 3.2 nm

1 -6 -5 -4 -3 -2 -1 0

log( solubility)

Figure 1. PuO2xH2O Particle Size, dt, as a Function of Decrease in Solubility from That of the Birth Particle
The effect of aging time on the plutonium concentration above PuO2xH2O in NaOH solution at room temperature (~22C) has been studied (Delegard 1985). The results in Figure 2 show that the plutonium concentration increases with NaOH concentration but, for any given NaOH concentration, decreases in proportion to the square root of time (i.e., t, where t is time) beyond a few days of aging. The decreasing plutonium concentration is observed from the initial 2-day sampling time for 1 M NaOH and from 7 days onward for all concentrations less than 9 M NaOH. The decreases in Pu concentration continued through the end of testing at about 2 years for solutions ranging from 1 to 10 M NaOH. The decreasing Pu concentration evidently is caused by aging or Ostwald ripening of the initial ~3.2-nm PuO2xH2O birth particles as they grow.

Official Use Only C-5

RPP-RPT-50941, Rev. 0

Figure 2. Plutonium Concentrations above Solids Precipitated from Pu(IV) Nitrate in Room-Temperature NaOH Solutions as a Function of Time (Delegard 1985) Because of the t dependence, a ~10-fold plutonium concentration decrease is observed between the sampling performed at 7 days and the last sampling done about 100-times later at 768 or 877 days (at 3 and 4 M NaOH) for tests run at 1 M to 8 M NaOH. Conservatively, an additional 10-fold concentration decrease might occur with a further 100-fold increase of aging time. Thus, if aging time were extended another factor of 100 from 800 days to 80,000 days or ~220 years, the Pu concentration would decrease by an additional factor of 10. Therefore, as much as a 100-fold decrease in plutonium concentration is projected to occur after 220 years of ~22C (room temperature) aging from the time of tank discharge as 3.2-nm PuO2xH2O birth particles. However, this extended aging would attain a projected terminal particle size of only ~5.0 nm as seen in Equation 6. Effect of Temperature on PuO2xH2O Particle Size
The work by Yusov and colleagues (2000) showed that moderate increase in PuO2xH2O particle size occurred with increasing preparation temperature. Thus, 2.5-nm particles were found after 5 hours of room temperature aging and 2 hours of 75C aging while 4.5-nm particles were produced at 160-200C (aging time not given). The effect of temperature on the growth of PuO2xH2O particles in Hanford tank waste is of interest because the waste has underwent and continues to undergo heating by both radiolytic

Official Use Only C-6

RPP-RPT-50941, Rev. 0

and waste evaporation processes. However, aside from the tests of Yusov and colleagues (2000), no explicit tests of the effects of temperature and time on PuO2xH2O particle size have been performed. The data shown in Figure 2 provide some information on the rate of PuO2xH2O particle growth at room temperature (~22C) through the rate of plutonium concentration change. The concentration dependence on time can be interpreted to mean that the reaction is controlled by diffusion. In this case, the t dependence of plutonium concentration found above the aging PuO2xH2O particles implies that the crystallization reaction is controlled by the diffusion rate of plutonium through solution from the dissolving PuO2xH2O particles to the PuO2xH2O particles growing by Ostwald ripening. The diffusion rate of solutes, such as dissolved plutonium, through solution is known to increase with increasing temperature (Adamson 1979). Further, the diffusion rate increase with temperature occurs because of increasing solution fluidity where fluidity is inverse viscosity. Therefore, the rate of the crystallization reaction should increase at higher temperature in proportion to the corresponding increase in the solution fluidity. The dependence of the solution fluidity on temperature follows an Arrhenius relationship in which the logarithm of the fluidity increases in proportion to the inverse absolute temperature (Adamson 1979). The activation energies calculated for aqueous solution fluidity are low compared with activation energies typically found for chemical reactions meaning that the temperature dependence of fluidity and, by extension, solute diffusion also is low. For pure water, the activation energy for fluidity is 15.8 kJ/mole. At this activation energy, the fluidity of water increases about a factor of 2.5 by raising the temperature from 22C (room temperature) to 70C and by a factor of 3.4 when raised to 90C. In contrast, chemical reaction rates typically increase about 2-fold for every 10C temperature increase. Therefore, a reaction temperature increase from 22C to 90C would increase ordinary chemical reaction rates by a factor of about 100. Sodium hydroxide solutions have higher viscosities than water and also have higher fluidity activation energies. In addition, both viscosity and fluidity activation energy increase with increasing NaOH concentration. These properties of NaOH solution are described in Appendix 3. It is also noted that NaOH is the dominant contributor to the viscosity of tank waste solution with sodium aluminate having a lesser contribution the other constituent salts (e.g., sodium nitrate and nitrite, NaNO3 and NaNO2) having little influence on viscosity. As seen in Appendix 3, the fluidity activation energy for 3 M NaOH is ~17 kJ/mole and the fluidity increases about a factor of 2.7 between 22C and 70C, the temperature at the high end of typical waste vacuum evaporation operations. However, boiling temperatures have been experienced for some tanks due to radiolytic heating. The fluidity of 3 M NaOH increases about a factor of 3.7 between 22C and 90C. For 8 M NaOH (i.e., at the high end of tank waste salt concentration), the activation energy for fluidity is ~25 kJ/mole and the fluidity increases by a factor of 4.1 between 22C and 70C and by a factor of 6.7 between 22C and 90C. As a first approximation, the higher waste storage temperatures increase NaOH solution fluidity, and thus on the rate of PuO2xH2O particle growth, by about a factor of 5. It was seen in the previous section that the 100-fold decrease in plutonium concentration projected to occur after 220 years of ~22C (room temperature) aging would grow the 3.2-nm PuO2xH2O birth particles to ~5.0 nm. Therefore, the approximately 5-fold more rapid aging that would occur at actual tank waste temperatures and fluidities would decrease the time to attain the ~5.0 nm particle size about 5-fold to about 220/5 = 44 years.

Official Use Only C-7

RPP-RPT-50941, Rev. 0

Based on this analysis, the higher tank waste storage temperatures would elicit little additional progress toward transforming the PuO2xH2O solids to micron-scale particles by Ostwald ripening. Instead, the PuO2xH2O particles would grow by Ostwald ripening to no greater than tens of nanometers in size even after decades of hydrothermal aging. References Adamson, AW. 1979. Pages 280-282 and 616 of A Textbook of Physical Chemistry, 2nd edition, Academic Press, New York, New York. Bundschuh, T, R Knopp, R Mller, JI Kim, V Neck, and Th Fanghnel. 2000. Application of LIBD to the Determination of the Solubility Product of Thorium(IV)-Colloids. Radiochimica Acta 88:625-629. Callaway, WS and GA Cooke. 2008. Distribution of Plutonium-Rich Particles in Tank 241-SY-102 Sludge. CH2M-0400872, May 6, 2008, CH2M Hill Plateau Remediation Company, Richland, Washington. Clark, DL, SS Hecker, GD Jarvinen, and MP Neu. 2006. Plutonium. Page 936, Chapter 7, in The Chemistry of the Actinide and Transactinide Elements, 3rd edition, LR Morss, NM Edelstein, J Fuger, and JJ Katz, Editors, Springer, Dordrecht, The Netherlands. Clark, SB and C Delegard. 2002. Plutonium in Concentrated Solutions, Chapter 7 in Advances in Plutonium Chemistry 1967-2000, DC Hoffmann, Senior Editor, American Nuclear Society, La Grange Park, Illinois. Cooke, GA. 2011. Personal communication from co-author of Callaway and Cooke (2008). Delegard, CH. 1985. Solubility of PuO2xH2O in Alkaline Hanford High-Level Waste Solution. RHORE-SA-75 P, Rockwell Hanford Operations, Richland, Washington. Available at: www.osti.gov/bridge/servlets/purl/5402793-8aDMm2/5402793.pdf. Also found as Solubility of PuO2.xH2O in Alkaline Hanford High-Level Waste Solution, CH Delegard, 1987, Radiochimica Acta 41:11-21. Dow. 2011. Viscosity Table For Pure (Salt Free) Caustic Soda Solution. Form No. 102-00394-1203. The Dow Chemical Company. Available at: http://www.dow.com/webapps/lit/litorder.asp?filepath=causticsoda/pdfs/noreg/102-00394.pdf&pdf=true. For supplementary viscosity data, see https://tecci.bayer.de/io-tra-pro/emea/de/docId2857611/Caustic_soda_solution.pdf. Fanghnel, Th. and V Neck. 2002. Aquatic Chemistry and Solubility Phenomena of Actinide Oxides/Hydroxides. Pure and Applied Chemistry 74(10):18951907. Guillaumont, R, T Fanghnel, J Fuger, I Grenthe, V Neck, DA Palmer, and MH Rand. 2003. Page 318 of Update on the Chemical Thermodynamics of Uranium, Neptunium, Plutonium, Americium and Technetium. Elsevier, Amsterdam, The Netherlands. Haire, RG, MH Lloyd, ML Beasley, and WO Mulligan. 1971. Aging of Hydrous Plutonium Oxide. Journal of Electron Microscopy 20(1):8-16.

Official Use Only C-8

RPP-RPT-50941, Rev. 0

Lee, E, C Ribeiro, E Longo and E Leite. 2005. Oriented Attachment: An Effective Mechanism in the Formation of Anisotropic Nanocrystals. Journal of Physical Chemistry B 109:20842-20846. Liu, Y, K Kathan, W Saad and R Prudhomme. 2007. Ostwald Ripening of -Carotene Nanoparticles. Physical Review Letters 98(3):036102-1 036102-4. Schindler, PW. 1967. Heterogeneous Equilibria Involving Oxides, Hydroxides, Carbonates, and Hydroxide Carbonates. Pages 196 to 221 of Equilibrium Concepts in Natural Water Systems, Advances in Chemistry Series 67, W Stumm, Symposium Chairman, American Chemical Society, Washington, DC. See also W Stumm and JJ Morgan, 1996, Aquatic Chemistry Chemical Equilibria and Rates in Natural Waters, 3rd edition, pages 413-414 and 806-809, John Wiley and Sons, New York, New York. Shannon, D. 1976. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallographica Section A A32:751-767. Shaw, DJ. 1980. Colloid and Surface Chemistry, 3rd edition, Butterworth & Co. Ltd., New York, New York. Yusov, AB, AYu Garnov, VP Shilov, IG Tananaev, MS Grigorev, and NN Krot. 2000. Plutonium(IV) Precipitation from Alkaline Solutions. I: Effect of Precipitation and Coagulation Conditions on Properties of Hydrated Plutonium Dioxide PuO2xH2O. Radiochemistry 42(2):151-156. See also Plutonium(IV) Precipitates Formed in Alkaline Media in the Presence of Various Anions, NN Krot, VP Shilov, AB Yusov, IG Tananaev, MS Grigoriev, AYu Garnov, VP Perminov, and LN Astafurova, 1998, PNNL-11901, Pacific Northwest National Laboratory, Richland, Washington. Available at: http://www.osti.gov/bridge/purl.cover.jsp?purl=/665911-NwS0BW/webviewable/. Westrum, EF, Jr. 1949. The Preparation and Properties of Plutonium Oxides, paper 6.57, pages 936944 of The Transuranium Elements Research Papers, GT Seaborg, JJ Katz, and WM Manning, editors, National Nuclear Energy Series IV-14B, McGraw-Hill Book Company, Inc., New York, New York.

Official Use Only C-9

RPP-RPT-50941, Rev. 0

Attachment 1 Derivation of Equations for the Effect of PuO2 Particle Size on PuO2 Solubility Product and Solubility

Derivation of Equation 1 and the Value of x in Equation 1


According to the derivation given by Schindler (1967), the Gibbs free energy change, G, observed for the equilibrium between a massive solid, AB (S0), with molar surface area, S, tending to zero and particle size, d, tending to infinity and a finely divided solid AB (S), of the same composition: AB (S0) AB (S) Reaction 1A

is related to the solubility products of the two respective AB phases by , the average Gibbs free energy of the solid-liquid interface, and the molar surface area:
G = RT ln K 0 (S) sp
0 K sp (S 0)

2 S 3

Equation 1A

where: R = gas constant T = absolute temperature. The molar surface area for particles of size d is calculated as shown in Equation 2A:
S= M M 6 = d d

Equation 2A

where M = formula weight of the solid /d = 6/d = surface area-to-volume ratio of a sphere or cube of diameter or edge length d = solid density. Schindler (1967) derived Equation 3A to estimate the average Gibbs free energy of the solid-liquid interface:

=
where

0 3RT ln K sp (S 0)

2 N A 4ri2

Equation 3A

NA = Avogadros number ri = ionic radii of the constituent ions.

Official Use Only C-10

RPP-RPT-50941, Rev. 0

Values pertinent to PuO2 can be entered into Equations 2A and 3A as shown in Equations 4A and 5A, respectively, to determine S and . The PuO2 data and other numerical values are: M = 271 g/mole PuO2 = 11.46 g/cm3 (Clark et al. 2006) R = 8.314 J/moledegree T = 298 0 log K sp = -64.04 (Guillaumont 2003) NA = 6.0231023/mole rPu 4+ = 0.096 nm, 8-coordinate (Shannon 1976)

rO 2 = 0.138 nm, 4-coordinate (Shannon 1976)


271 g 6 141.9 cm 3 1.419 10 4 m 3 M mole PuO 2 S= = = = 11.46 g mole PuO 2 d mole PuO 2 d d d 3 cm
=
0 3RT ln K sp (S 0)

Equation 4A

2 N A 4ri2 3

0 3RT ln K sp (S 0) 2 2 2 N A 4 (rPu 4 + + rO 2 )

8.314 J 298 (64.04 ) mol deg = 2 6.02 10 23 4 [(0.96 10 10 m) 2 + (1.38 10 10 m) 2 ] 1.113 J = m2

Equation 5A

The values for S and from Equations 4A and 5A then may be entered into Equation 1A to obtain, with algebraic rearrangement, Equation 6A:
2 S t 0 log K sp (S) log K sp (S 0) = 3 2.303 RT 2 1.113 J 1.419 10 4 m 3 3 mole PuO 2 d t 10 9 nm m2 = 8.314 J m 2.303 298 mole deg 18.46 = d t ( nm)

Equation 6A

The result of Equation 6A shows that the value of x used in Equation 1 is 18.46. Equation 6A is used to determine the effect of particle size, dt, on the solubility product observed for the finely divided PuO2xH2O particles or, conversely, the influence of the observed solubility product on the particle size.

Official Use Only C-11

RPP-RPT-50941, Rev. 0

Derivation of Equation 4
The solubility product for PuO2 expresses the equilibrium Reaction 2A: PuO2 (am, hydr. or cr.) + 2 H2O (l) Pu4+ (aq) + 4 OH- (aq). Reaction 2A

The solubility equilibrium of the charge-neutral plutonium hydroxide dissolved species from the solid phase PuO2 (am, hydr. or cr.) is shown in Reaction 3A: PuO2 (am, hydr. or cr.) + 2 H2O (l) Pu(OH)4 (aq). Reaction 3A

0 The solubility of the Pu(OH)4 (aq) solution species, K s, 4 , is related to the PuO2 (am, hydr.) solubility product

and the cumulative formation constant of the Pu(OH)4 (aq) solution species, 0 ,1 , by Equation 7A, which 4 can be rearranged to Equation 8A. The analogous Equation 9A can be written for large crystal PuO2 (cr).
0 log K s , 4 (colloid) = log [Pu(OH) 4 ( aq) ] = log (K 0 colloid) + log 0,1 sp 4

Equation 7A Equation 8A Equation 9A

0 log (K sp colloid) = log K 0, 4 (colloid) log 0,1 s 4

0 0 log (K sp l arg e crystal) = log K s , 4 (l arg e crystal) log 0,1 4

0 0 Substitution of log (K sp colloid) and log (K sp l arg e crystal) from Equations 8A and 9A into Equation 2

yields Equation 10A in which the log 0 ,1 values in Equations 8A and 9A are identical and whose 4 difference is zero:
0 0 log K s, 4 (colloid) log K s , 4 (l arg e crystal) = 5.71 =

18.46 d 0 (nm)

Equation 10A

The decrease in solubility of Pu (i.e., Pu solution concentration) at any time after the highest solubility established by the ~3.2-nm size birth particle thus is reflected in the increase in size of the particle, dt, at that later time, t. This is seen in Equation 11A, which is a recasting of Equation 10A to include the influence of the solubility change, log (solubility), for the growing colloid:
0 t log K s, 4 (colloid) log K s, 4 (l arg e crystal) = 5.71 + log (so lub ility) =

18.46 d t (nm)

Equation 11A

Official Use Only C-12

RPP-RPT-50941, Rev. 0

It is noted that the formation of the neutral complex Pu(OH)4 aq becomes the first step in subsequent dissolution equilibria needed to form the anionic hydroxide-complexed Pu(IV) and Pu(V) species postulated to form in strongly alkaline solution [e.g., Pu(OH)5- or PuO2(OH)43-, respectively; Clark and Delegard 2002]. The influences of the subsequent hydroxide complexation and oxidation steps in the plutonium solute species equilibria are fixed for a given hydroxide chemical activity (NaOH concentration) and REDOX potential (established in Hanford tank waste by influences such as nitrate and nitrite concentration and radiolysis). Therefore, the change in the solubility equilibria established for Pu(OH)4 (aq) and any subsequent dissolved species over an aging/ripening PuO2 (am, hydr.) solid phase is reflected only in the change in the solid phases solubility due to aging.

Official Use Only C-13

RPP-RPT-50941, Rev. 0

Attachment 2 XRD Patterns of PuO2xH2O Precipitated from Pu(IV) and Pu(VI) Nitrate in NaOH Solution

Figure A.2.1. XRD Pattern for PuO2xH2O Precipitated from Pu(IV) Nitrate in 5 M NaOH after 4 Days of Aging (full width at the half maximum of the peak at 28.55 2- is 0.069 radians)

Official Use Only C-14

RPP-RPT-50941, Rev. 0

Figure A.2.2. XRD Pattern for PuO2xH2O Precipitated from Pu(IV) Nitrate in 5 M NaOH after 14 Months of Aging (full width at the half maximum of the peak at 28.55 2- is 0.055 radians)

Official Use Only C-15

RPP-RPT-50941, Rev. 0

Figure A.2.3. XRD Pattern for PuO2xH2O Precipitated from Pu(IV) Nitrate in NaOH Solution after 38 Months of Aging (full width at the half maximum of the peak at 28.55 2- is 0.047 radians)

Official Use Only C-16

RPP-RPT-50941, Rev. 0

Figure A.2.4. XRD Pattern for PuO2xH2O Precipitated from Pu(VI) Nitrate in 1 M NaOH after 9 Months of Aging (full width at the half maximum of the peak at 28.55 2- is 0.0132 radians)

Official Use Only C-17

RPP-RPT-50941, Rev. 0

Attachment 3 Sodium Hydroxide Solution Viscosities and Fluidities as Functions of Concentration and Temperature
The dependence of sodium hydroxide (NaOH) solution viscosities on temperature and concentration were derived from data plots compiled and published by Dow Chemical Company (Dow 2011). Values from the Dow plots were cross-checked and found to be consistent with values from other published sources. The solution fluidity, which is the inverse of the solution viscosity, is known to follow an Arrhenius dependence on temperature in which the natural logarithm of the fluidity is linearly dependent on the inverse absolute temperature with a slope of -Ea/R where Ea is the activation energy of the fluidity and R is the gas constant, 8.3145 J/moleK (Adamson 1979). The plots of the temperature dependencies of water and NaOH solution fluidities given in Figure A.3.1 follow the Arrhenius equation with high fidelity over NaOH concentrations up to 40 wt% (~14.3 M) and 10 to 90C (and beyond). The Ea values for solution fluidity, taken from the slopes given in Figure A.3.1, are seen to increase with NaOH concentration (Figure A.3.2), almost doubling between pure water and 40 wt% NaOH. This means that the fluidities of the NaOH-rich Hanford tank waste solutions become increasingly temperature-dependent as the NaOH concentration increases.
T, C

1.4 1.0

90

80

70

60

50

40

30

20

10

0 0%
y = -1903.0x + 6.4508 R2 = 0.9878 y = -2046.3x + 6.3062 R2 = 0.9920 y = -2197.9x + 6.4753 R2 = 0.9914 y = -2496.2x + 7.0335 R2 = 0.9939 y = -2981.8x + 8.1247 R2 = 0.9935 y = -3322.9x + 8.7003 R2 = 0.9916 y = -3382.5x + 8.5963

ln Fluidity (1/Viscosity, centipoise-1)

0.6 0.2 -0.2 -0.6 -1.0 -1.4 -1.8 -2.2 -2.6 -3.0 -3.4 0.0027

10%

15%

20%

25%

30%

35% R2 = 0.9948
y = -3649.1x + 9.0759

40% R2 = 0.9983

0.0029

0.0031

0.0033

0.0035

0.0037

1/T, K-1

Figure A.3.1. Arrhenius Plot of Water and NaOH Solution Fluidities (NaOH concentrations in weight%; straight-line fits of the data are shown to the right)

Official Use Only C-18

RPP-RPT-50941, Rev. 0

31 30 29 28 27 26 25 24 23 22 21 20 19 18 17 16 15 0 2 4 6 8 10 12 14

Ea, kJ/mol

[NaOH], M

Figure A.3.2. Activation Energy, Ea, for Fluidity as a Function of NaOH Concentration
The multiplicative factor fluidity increases as functions of NaOH concentration and temperature shown in Figure A.3.3 were derived from the fluidity data to determine how much the PuO2xH2O crystallization rate in a given NaOH solution would increase with increasing aging temperature. For example, it is seen that the fluidity of 6 M NaOH increases about a factor of 5 when the temperature increases from 22C to 90C. Therefore, the time observed for PuO2xH2O to grow to a given size in the room-temperature (22C) aging tests described in Figure 2 and the related text would decrease by a factor of 5 if the aging had occurred at 90C instead of room temperature.
11
90 C

10

Fluidity Factor Increase above 22 C

9 8 7 6 5
60 C 70 C 80 C

4 3 2 1 0 2 4 6 8 10 12 14
50 C 40 C 30 C

[NaOH], M

Figure A.3.3. Factor Increase in Fluidity above Room Temperature (22C) for NaOH Solution

Official Use Only C-19

RPP-RPT-50941, Rev. 0

REPORT 2 METATHESIS OF PLUTONIUM(IV) OXALATE AND PLUTONIUM(III)/(IV) FLUORIDES IN ALKALINE MEDIA TO FORM PUO2XH2O Metathesis of Plutonium(IV) Oxalate and Plutonium(III)/(IV) Fluorides in Alkaline Media to Form PuO2xH2O
Introduction
Plutonium oxalate, Pu(C2O4)26H2O, and plutonium tetrafluoride, PuF4, were solid phase chemical intermediates formed in the preparation of plutonium metal from plutonium nitrate solutions in the Plutonium Finishing Plant, PFP. Plutonium oxalate also was an intermediate in the preparation of plutonium dioxide, PuO2, at both the PFP and at N Cell operations in PUREX. Hydrated plutonium tetrafluoride, PuF42.5H2O, and the double salts NaPuF5 and Na2PuF6, all solid phases with low aqueous solubility, likely were formed in fluoride-assisted dissolution of PuO2 values in scrap processing and in other points in Hanford processes. The hydrate PuF42.5H2O also could form in stripping of Pu(IV) from the organic phase in the PRF solvent extraction operations within the PFP in which fluoride was added to enhance this stripping while the trivalent fluoride salt PuF3xH2O could form in reductive stripping. Inadvertent discharge of Pu(C2O4)26H2O, PuF3xH2O, PuF4, PuF42.5H2O, NaPuF5, and Na2PuF6 solids to the Hanford waste tanks could have occurred due to process operation upsets and as trace material losses in scrubber solutions and process wastes from these facilities. The chemical behaviors of Pu(C2O4)26H2O, PuF3xH2O, PuF4, PuF42.5H2O, and NaPuF5, and Na2PuF6 with aging and after contacting alkaline tank wastes are of interest in understanding their likely current disposition in the tank farms. Besides the plutonium oxalate and fluoride compounds, plutonium also was introduced to the tank farms as PuO2 largely suspected to have been added by way of incomplete dissolution of Pu-bearing scrap in the PFP and ensuing loss of the undissolved solids in raffinates from solvent extraction processes in that plant. Losses of PuO2 to tank farms also could have arisen during reprocessing of non-irradiated and irradiated plutonium-uranium oxide mixed oxide [(U,Pu)O2; MOX) fuel in the REDOX and PUREX plants. Finally, plutonium addition to the tank farms is expected to have occurred in the form of PuO2xH2O [Pu(IV) hydrous oxide, sometimes called plutonium hydroxide] produced by the action of hydroxide ion on dissolved Pu(IV) [and even Pu(III) and Pu(VI)] nitrate solutions. The plutonium added to the tank farm wastes from these vectors largely were accompanied by dissolved ferric nitrate, Fe(NO3)3, which, when made alkaline, would co-precipitate as ferric hydroxide, Fe(OH)3, with the PuO2xH2O formed from plutonium nitrate acid solution neutralization and accompany tramp PuO2 particles. The Fe(NO3)3 was present from process sources [e.g., stainless steel corrosion and oxidation of ferrous sulfamate, Fe(NH2SO3)2 reductant] and also as material intentionally added to act as a neutron absorber for criticality safety purposes. The PuO2xH2O formed by precipitation of Pu(IV) nitrate solution in nitric acid (HNO3) by mixing with the precipitating alkaline agent solution either in direct (alkali-to-acid) or reverse (acid-to-alkali) strike has been examined and found by XRD to be very finely crystalline PuO2 containing associated water described as PuO2xH2O (Yusov et al. 2000a and others). Based on XRD analyses, the PuO2xH2O crystallite size for solids prepared by precipitation in room temperature and 75C alkaline solution is about 0.0025 m and about 0.0045 m when prepared hydrothermally at 160 to 200C. The value of x in PuO2xH2O is variable, depending on how well the PuO2xH2O is dried. At ~82% relative humidity

Official Use Only C-20

RPP-RPT-50941, Rev. 0

(controlled in a humidor by 25% sulfuric acid), x is about 2.6. At ~35% relative humidity (50% sulfuric acid control), x is 2.3 and x is 1.5 when dried at 15% relative humidity (saturated potassium hydroxide, KOH). Thermal analyses show that the association of the H2O with the PuO2 is indiscrete meaning that the water is not bound within distinct hydrate compounds but rather is evolved continuously when heated from 50C to 250C with a clear endotherm at 110C proportional to x. Assuming that the water association with the underlying finely crystalline PuO2 is physical only and that x, the number of H2O molecules per PuO2 is 2.3, the particle density of PuO2xH2O (x=2.3) is 5.14 /cm3.(4) For spherical underlying PuO2 particles with 2.5-nm diameter, the surface water layer of PuO2xH2O where x is 2.3 at 35% relative humidity is about 1.6 H2O molecules thick. This thickness is similar to that found for PuO2 prepared from metal oxidation in air and calcined in oxygen at 800C in which ~2 molecules of water thickness was observed at room temperature under ~30-50% relative humidity (Haschke et al. 2001; Stakebake and Steward 1973). However, the particle density of PuO2xH2O in aqueous suspension would revert to that of the core PuO2 because the surface and bulk water would be indistinguishable. The technical chemical literature was examined to determine the outcomes of interactions of Pu(IV) oxalate, Pu(C2O4)26H2O, Pu(III) fluoride, and plutonium(IV) fluoride as PuF4, PuF42.5H2O, NaPuF5, and Na2PuF6 with alkaline solution. The following analysis and discussion was prepared to determine if these compounds are unstable with respect to the formation of PuO2 or PuO2xH2O either by radiolysis or by alkaline hydrolysis. Information on the initial particle size of Pu(C2O4)26H2O, PuF3xH2O, PuF4, PuF42.5H2O, NaPuF5, and Na2PuF6 and the expected particle sizes of their products after contact with alkaline solution also is discussed.

Plutonium Oxalate
Plutonium(IV) oxalate was prepared in PFP operations by co-addition of ~1 M oxalic acid solution and ~0.2 M Pu(IV) in ~1-3 M nitric acid (HNO3) to a stirred and warmed precipitation vessel. The solution mixture was digested at ~60C to encourage crystal growth and the slurry decanted for filtration in a continuous mixed suspension / mixed product removal (MSMPR) crystallizer. The product Pu(C2O4)26H2O particle size was 5 to 7 m as determined by X-ray analysis (Myers 1956). Processing in 2001-2002 to recover plutonium from stored nitrate solutions during the de-inventory of the PFP also used Pu(IV) oxalate precipitation. In this case, the precipitation occurred in an air-sparged batch reactor by adding solid oxalic acid crystals to the feed Pu(IV)-bearing solution in HNO3. The precipitation occurred at lab temperature and digestion time was about 1 hour before filtration. However, under these conditions, solids formation attributed to Pu(IV) oxalate post-precipitation was observed in the process filtrates. The process filtrates including precipitated solids were discarded to waste and ultimately to tank farms. The particle size distribution of the Pu(IV) oxalate solids in this stream is not known. No information on the particle density of Pu(IV) oxalate hexahydrate, or even of the thorium(IV) or uranium(IV) chemical analogues, was found in the technical literature. The densities of Pu(IV) oxalate dihydrate and the thorium analogue are 3.085 and 3.391, respectively, while that of plutonium(III) oxalate decahydrate is 3.115 g/cm3 (Jenkins et al. 1965a and b). Based on these values, Pu(IV) oxalate hexahydrate density is estimated to be between 3 and 4 g/cm3.

(4) Density PuO 2 2H 2 O =

271 g PuO 2 + 2 18 g H 2 O 5.16 g = . 3 3 cm 3 cm cm 271 g PuO 2 + 2 18 g H 2 O 11.54 g 1 .0 g

Official Use Only C-21

RPP-RPT-50941, Rev. 0

The stabilities of plutonium(IV) oxalate hexahydrate to radiolysis and to alkaline hydrolysis are examined in this section.

Radiolysis
Radiolytic decomposition of Pu(C2O4)26H2O is observed with the ultimate formation of PuOCO32H2O (Gelman et al. 1962; Gelman and Sokhina 1958). The half-life of the material to decomposition by its own alpha radiation is 64 days (Jenkins et al. 1963). A similar rate was observed by Myers (1956). The intermediate presence of trivalent plutonium, Pu(III), as Pu2(C2O4)3nH2O is observed during radiolytic decomposition (Gelman and Sokhina 1958). Based on these observations, Pu(C2O4)26H2O particles are not stable to chemical decomposition with respect to autogenous alpha radiolysis.

Alkaline Hydrolysis
Plutonium(IV) oxalate is known to hydrolyze in strong alkaline solution to form plutonium hydroxide (or plutonium hydrous oxide, PuO2xH2O). The balanced chemical reaction is: Pu(C2O4)26H2O + 4 OH- PuO2xH2O + 2 C2O42- + (8-x) H2O The most direct account found in the technical literature of the reaction of Pu(IV) oxalate with hydroxide ion, OH-, describes metathesizing Pu(C2O4)26H2O in 4 M KOH solution. This was done to release the oxalate ion, C2O42-, for analysis after first filtering out the precipitated plutonium hydroxide (Myers 1956). Other laboratory tests show that PuO2xH2O forms exclusively when dissolved acidic Pu(IV) nitrate solution is added to 60C alkaline solutions, at 0.2 and 1 M NaOH, containing 0.1 M oxalate (Yusov et al. 2000b). The precipitated PuO2xH2O contained no detectible oxalate (<5 mol% with respect to plutonium) and has the same hygroscopic (water absorption) properties as the PuO2xH2O prepared by precipitation in the absence of oxalate. Together, the experiments of Yusov and colleagues (2000b) and those of Myers (1956) indicate that Pu(IV) oxalate is unstable with respect to the formation of PuO2xH2O in alkaline solutions at least above 0.2 M NaOH. The stability of Pu(IV) oxalate at lower alkalinity can be inferred by study of the behavior of sodium plutonium oxalate double salts (Gelman and Sokhina 1958). In these tests, the double salt of formula Na4[Pu(C2O4)4]5H2O was prepared by mixing Pu(C2O4)26H2O solids in water, adding sodium oxalate, Na2C2O4, solids and heating at ~80C until both solids dissolved by formation of higher Pu(IV) oxalate complexes. The sodium-plutonium oxalate double salt then was precipitated from the aqueous solution by adding ethyl alcohol. The Na4[Pu(C2O4)4]5H2O precipitate, when dissolved in water, was found to produce a solution of pH 4.5 to 4.7. However, raising the pH to about 7.5 to 8 caused destruction of the double salt and precipitation of plutonium hydroxide. Therefore, even with stoichiometrically excess oxalate compared to formation of Pu(C2O4)26H2O, as would be the case for Na4[Pu(C2O4)4]5H2O dissolution, PuO2xH2O forms preferentially under very mildly alkaline pH ~8 conditions. Analysis of the thermodynamic stability of Pu(IV) oxalate to form PuO2xH2O by hydrolysis in alkaline solution was not possible because Gibbs free energy of formation values for Pu(C2O4)26H2O were not found in the technical literature. Despite this paucity of thermodynamic knowledge, abundant chemical observation shows that Pu(C2O4)26H2O is unstable, to even mildly alkaline hydrolysis conditions, to formation of PuO2xH2O.

Official Use Only C-22

RPP-RPT-50941, Rev. 0

Plutonium(III) and Plutonium(IV) Fluoride Salts


The process sources and properties of PuF3xH2O and the Pu(IV) fluorides PuF4, PuF42.5H2O, NaPuF5, and Na2PuF6 and their stabilities to alkaline hydrolysis are examined based on findings reported in the technical literature.

Plutonium(III) and (IV) Fluoride Sources and Properties


The process sources and properties of the Pu(IV) fluorides PuF3xH2O, PuF4, PuF42.5H2O, NaPuF5, and Na2PuF6 are examined. PuF3xH2O A report of the potential presence of PuF3 (likely PuF3xH2O with x ~0.75 or 0.4; Clark et al. 2006) was reported to be the cause of decreased CAX flow in May 1968 (pages 53-54, Engineers of Plutonium Process Engineering 1968). The PuF3xH2O, described as golf ball size and readily and completely soluble in warm aluminum nitrate solution, were collected downstream of the process feed tank filter. The particle size of PuF3xH2O that could have formed in the process is not known but, given the low solubility of PuF3 (Griffo et al. 1964), the individual particles likely were small. Studies designed to optimize the precipitation of PuF3 by mixing of Pu(III) in nitric acid with hydrofluoric acid solution showed that 10-15 m agglomerates could be made with moderate agitation but that with greater agitation, the particle size was ~2-3 m (Burney and Tober 1965). Subsequent lab tests at PFP under the same chemical process conditions that produced the golf ball size agglomerates also produced black solids at the filter. The PuF3xH2O is blue-violet in color, and thus may have appeared black when observed in plant processing, and has a particle density of the anhydrous PuF3 is 9.32 g/cm3 (Clark et al. 2006). PuF4 The particle size of PuF4 formed by hydrofluorination using hydrogen fluoride (HF) gas under conditions used at the PFP is 2 to 10 m (Myers 1956). Measurement of PuF4 particle size produced by the Remote Mechanical C (RMC) Line under routine 1980s operation in the PFP showed the weight-average particle size to be ~22 m with about 87 wt% of the material being greater than 10 m and about 0.5 wt% less than 2 m (Barney 1988). The density of PuF4 is 7.04 g/cm3 (Clark et al. 2006). It is noted that the solid particulate PuF4 (or starting PuO2) most likely to report to the downstream system used to scrub excess corrosive and toxic HF from the offgas would be that lofted by flow of the impinging HF gas and should be strongly skewed to the finer particles. PuF42.5 H2O Plutonium(IV) fluoride, in the form PuF42.5 H2O, arises by precipitation of Pu(IV) from acid solution containing fluoride (Dawson et al. 1954a). Conversion of PuF4 to PuF42.5 H2O is known to occur in HF solution (Dawson et al. 1954b) and it is likely that this conversion also transpires in water. Precipitation of pink PuF42.5 H2O was often observed in dissolution of scrap plutonium oxides because fluoride ion was added to HNO3 to complex Pu4+ as dissolved PuF3+ to improve plutonium dissolution from the source PuO2. At PFP, the fluoride often was added as hydrofluoric acid (dissolved HF). The fluoride ion concentration used in the dissolution processing generally ranged up to ~0.5 M while the HNO3 concentration ranged from about 6 to 12 M. During scrap dissolution, the fluoride was consumed

Official Use Only C-23

RPP-RPT-50941, Rev. 0
by the PuF3+ complex and also by reaction with silica (SiO2; to form SiF4) and by complexation with aluminum ion (Al3+) or other metal ions present as impurities in the scrap. Because of these reactions, additional fluoride would be required for the plutonium scrap dissolution to continue. However, if the fluoride concentration became too high, Pu(IV) fluoride precipitated as PuF42.5 H2O. Precipitates of what was probably PuF42.5 H2O, but was reported as pink PuF4, also have been observed in the organic wash column (CO) of the Plutonium Reclamation Facility (Knights 1971). The precipitation of the pink compound was found to be enhanced due to the presence of DBP, a hydrolytic and radiolytic decomposition product of the TBP extractant used at PFP. The solubility of PuF42.5H2O in the 2.5 M HNO3 / 0.25 M HF process stream is reported to be in excess of 8 grams plutonium per liter. However, the pink precipitate was observed even at concentrations less than 3 grams Pu/liter. According to the process report, the pink PuF4 precipitate coagulated and was heavy enough to exist with the organic effluent (which contained carbon tetrachloride diluent and thus was the lower phase in the extraction process). Solids reported to contain PuF4, but likely PuF42.5 H2O, and other miscellaneous solids were noted to be plugging a filter between Tank 39 and the CX column in May 1968 (page 50 of Engineers of Plutonium Process Engineering 1968). The solids were attributed to prior process operations. No information was found in the process or technical literature on PuF42.5 H2O particle size. However, the conditions under which the PuF42.5 H2O formed during scrap dissolution, such as extended times at high temperature followed by slow cooling and relatively high concentrations of fluoride and Pu(IV), would favor crystal growth while the low solubility would limit crystal size. The particle size is likely to be 1-10 m based on related formation condition and properties of PuF3xH2O (Burney and Tober 1965). The density of PuF42.5 H2O is 4.89 g/cm3 (Clark et al. 2006). NaPuF5 and Na2PuF6 The precipitation of Pu(IV) fluoride in the form of NaPuF5 or Na2PuF6 is likely if NaF were used, as it often was at PFP, as the fluoride source in dissolving plutonium-bearing scrap. Sodium also was present in many plutonium-bearing scrap materials processed at PFP, most obviously in electroreduction (ER) scrap containing NaCl/KCl (eutectic sodium/potassium chloride salt) from the Rocky Flats Site. Lab studies show that NaPuF5 forms by mixing PuF42.5H2O solids with NaF in water (Deichman and Tananaev 1962). With NaF concentrations above about 0.25 M, the salt Na2PuF6 is observed. The starting PuF42.5H2O, which has solubility of ~0.00013 M in water, is described as large acicular (needleshaped) crystals but the particle size is not given. The product NaPuF5 (solubility decreasing from ~0.00013 to ~0.000001 M as NaF concentration increased to 0.25 M) is described as fine gray-green needles. The density of NaPuF5 is 6.03 g/cm3 (Clark et al. 2006). The Na2PuF6, with ~0.000001 M solubility in ~0.25 to 0.5 M NaF, is described as large brown needles, but no quantitative particle size information is provided (Deichman and Tananaev 1962), and has a density of 5.84 g/cm3 (Clark et al. 2006). It is reported separately that when an excess of NaF is added to a solution of Pu(IV) in nitric acid, NaPuF5 is obtained as a dense green precipitate. On standing, the green solid becomes pink, converting to the Na2PuF6 salt (Alenchikova et al. 1958). Tenuous XRD evidence of a sodium-plutonium fluoride double salt in the heel of a Rocky Flats oxide laboratory dissolution test at the PFP also has been reported (Delegard 1985).

Official Use Only C-24

RPP-RPT-50941, Rev. 0

Other Pu(IV) Fluoride Salts Savannah River Site operations used potassium fluoride (KF) in their digestions of plutonium oxide values and have identified the analogous KPuF5 salt as well as KPu2F9 by X-ray diffractometry (Rudisill 2011). However, the particle size of the potassium double salts was not measured. Based on these observations, potassium-Pu(IV) fluoride double salts also might have formed in PFP scrap recovery operations for scraps, such as ER salts, which contained potassium. The precipitation of Pu(IV) fluoride in the form of CaPuF6 or other calcium-Pu(IV) fluoride double salt also might be credible if calcium were present in the scrap. For example, calcium fluoride, CaF2, is present in sand, slag, and crucible (SS&C) scrap. Calcium fluoride also was used as a source of fluoride in dissolution of plutonium values in scrap recovery operations at the PFP. The density of CaPuF6 is 6.65 g/cm3 (Keller and Salzer 1967). Plant-scale preparation of this compound is reported to occur by mixing approximately two volumes of ~50 g Pu(IV)/liter solution in 4.5 M HNO3 with one volume of 5.6 M HF to make CaPuF6 under a solution containing only 0.3 g Pu/liter (Harmon and Reas 1957). No information on the existence of the analogous MgPuF6 compound was found in the technical literature but its existence is likely and could have formed in processing SS&C scrap which contained magnesium oxide (MgO) sand and crucible materials.

Alkaline Hydrolysis of Plutonium(IV) Fluoride Compounds


Direct experimental and process plant observations of the interaction of plutonium tetrafluoride with alkaline solution have been recorded in the technical literature. These observations and supporting thermodynamic analyses indicate that Pu(IV) fluoride compounds are unstable to formation of PuO2xH2O in sodium-rich alkaline media. The earliest information on the stability of PuF4 in alkaline solution was developed in the Manhattan Project for plutonium separations from irradiated uranium fuel. In the Concentration step in the Bismuth Phosphate Process to recover plutonium, metathesis (double decomposition) of lanthanum fluoride (LaF3) and coprecipitated trace PuF4 was performed in which the LaF3/PuF4 solids were converted to their respective hydroxides La(OH)3/PuO2xH2O in hydroxide solution (Anonymous 1944; Duffield 1960): LaF3/PuF4 (s) + 3 KOH (aq) La(OH)3/PuO2xH2O (s) + 3 KF (aq) It is likely that the neat plutonium(IV) fluoride would be PuF42.5H2O when precipitated under these conditions. The LaF3-PuF4 precipitate, produced by treating mixed La(III) and Pu(IV) nitrate solution with 0.2 M to 0.5 M HF by either direct or reverse strike, took place using 15% KOH (~4 M KOH). To assure completeness in this plutonium-rich stream, the KOH metathesis was run at 80C for 90 minutes. It was found that at KOH concentrations lower than 15% or at temperatures lower than 75C the metathesis conversion was less rapid. The behavior of carrier-free plutonium and the performance in NaOH are similar to the behaviors observed with carrier and with KOH: It is to be understood that the process may be carried out with equally good results with a plutonium fluoride uncontaminated by carrier when the plutonium is in sufficient concentration and it will form precipitates without a carrier. As to the reagents used in the metathesis steps of my process, the hydroxide or carbonate of sodium may be used in place of the potassium basic salts (Duffield 1960).

Official Use Only C-25

RPP-RPT-50941, Rev. 0

The metathesized solids resemble the original LaF3/PuF4 mixture in appearance. The LaF3 characteristics are described as follows: Lanthanum fluoride as normally precipitated is a finely divided amorphous, hydrated, flocculent mass, which even under the electron microscope shows no definite crystalline structure, but rather a small ill-defined mass less than 0.01 micron in size (page 706 of Anonymous 1944). These observations imply that the metathesized solids likewise should be very finely particulate. Plutonium fluoride decomposition in KOH solution also occurred in the step immediately following the chemical removal of Zircaloy cladding from irradiated N Reactor uranium metal fuel in the Zirflex process at the PUREX plant at Hanford. The Pu(IV) fluoride was present with much greater quantities of U(IV) fluoride and undissolved irradiated uranium metal fuel in the cladding dissolver heels. Both compounds likely were present as the respective actinide fluoride hydrate salts of the form AnF42.5H2O. The metathesis reaction in the KOH solution: is used to convert uranium and plutonium fluorides to uranium and plutonium oxides to permit solvent extraction recovery and to limit corrosion of the dissolvers. The fluoride is removed from the dissolvers in the spent metathesis solution as soluble potassium fluoride. (Keith and Westra 1987) Laboratory testing undertaken to support the flowsheet modifications described by Keith and Westra (1987) showed that solid plutonium trifluoride, PuF3xH2O, oxidizes instantly to Pu(IV), as indicated by change from its original lavender color to form a green solid precipitate when added to solutions of 25 different compositions composed of mixed KOH (1 to 3 M) and KF (0 to 2 M). Plutonium(III) nitrate, initially blue in color, instantly precipitated to a lavender solid that, with shaking, turned green when added to the same KOH/KF solutions. These observations gave qualitative demonstration of the preferential formation of green PuO2xH2O over the pink PuF4 or PuF42.5H2O during oxidation of the Pu(III) even though abundant fluoride was present in many of these experiments (Delegard 1987). Plutonium solution concentrations for these tests tended to the same concentrations observed for Pu(IV) addition to KOH solutions of the same chemical activity, further supporting the thesis that Pu(IV) hydrous oxide, PuO2xH2O, is the solubility-controlling solid phase. For most of the PFP history, the scrubber used concentrated KOH solution. For other periods, the scrubber solution was monobasic aluminum nitrate [Monoban, AlOH(NO3)2] or aluminum nitrate [Al(NO3)3] (Panesko 1972). Small PuF4 particles were lofted from flowing HF gas and captured in KOH, AlOH(NO3)2, or Al(NO3)3 scrubber solutions in hydrofluorination operations in the PFP. The PuF4 particles metathesize upon contact with KOH solution to form Pu(IV) hydrous oxide: PuF4 (s) + 4 KOH (aq.) + (x-2) H2O (aq.) PuO2xH2O (s) + 4 KF (aq.) Contact of the PuF4 particles with the AlOH(NO3)2 or Al(NO3)3 solution decompose the PuF4 compound by abstracting fluoride in a series of steps to culminate in the following net reaction to dissolve the plutonium as Pu(NO3)4 or as soluble plutonium-fluoride complexes in excess Al(NO3)3: 3 PuF4 (s) + 4 Al(NO3)3 (aq.) 3 Pu(NO3)4 (aq.) + 4 AlF3 (s)

Official Use Only C-26

RPP-RPT-50941, Rev. 0

During much of PFP operating history, these scrubber solutions ultimately were discharged to the underground waste storage tanks and thus entered NaOH solution where the Pu(NO3)4 or Pu(IV) fluoride complexes would have hydrolyzed. As outlined in Table 1, thermodynamic analysis shows that the PuF4 hydrolysis reaction: PuF4 (s) + 4 NaOH (aq., 1 m) PuO2 (s) + 4 NaF (aq., 1 m) + 2 H2O (aq.) is favored (i.e., the free energy of the reaction, Grxn, is negative) in 1 molal (~1 M) sodium hydroxide, NaOH, solution to form PuO2 solid and 1 molal (~1 M) sodium fluoride, NaF, solution.

Table 1. Free Energy of the PuF4 Hydrolysis Reactions in 1 Molal NaOH and in Water
Grxn, kJ/mol, in Reactants PuF4 (s) NaOH (aq, 1 m) H2O (aq.) Products PuO2 (s) or PuO2xH2O (s)* NaF (aq., 1 m) HF (aq. 1 m) H2O (aq.) Grxn Grxn = Gf0 (products) Gf0 (reactants) = -998.113 -965.520 -540.680 -296.820 -237.129 Product PuO2 PuO2xH2O -474.258 In 1 m NaOH -201.750 -169.157 In Water 45.606 78.199 -998.113 -965.520 -2162.720 -1187.280 -998.113 -965.520 Gf0, kJ/mol 1 m NaOH 1756.741 1676.600 474.258 Water 1756.741

-1756.741 -419.150 -237.129

Thermodynamic data for PuF4 and PuO2 from Guillaumont et al. 2003. Thermodynamic data for non-Pu phases from Wagman et al. 1982. * The log10Ksp for PuO2 = -64.04 and the log10Ksp for PuO2xH2O = -58.33 where Ksp is the solubility product of the respective PuO2 and PuO2xH2O phases (Guillaumont et al. 2003). Grxn. (PuO2PuO2xH2O) = -2.303 RT log K = -2.303 RT (-64.04+58.33) = 32.593 kJ/mol. Gf (PuO2xH2O) = Gf (PuO2) + 32.593 = -998.113 + 32.593 = -965.520 kJ/mol.

However, as also shown in Table 1, the reaction of PuF4 in neutral water: PuF4 (s) + 2 H2O (aq.) PuO2 (s) + 4 HF (aq., 1 m)

Official Use Only C-27

RPP-RPT-50941, Rev. 0

to form PuO2 solid and 1 molal hydrogen fluoride, HF, aqueous solution is not favored (i.e., Grxn is negative). Thermodynamic data on the standard Gibbs free energy of formation (Gf0) of PuF42.5H2O, NaPuF5, and Na2PuF6 were not found. However, none of these compounds form even in the presence of excess fluoride and sodium in alkaline solution. Instead, PuO2xH2O is found from metathesis of PuF42.5H2O in alkaline solution abundant in both sodium and fluoride. The experimental observations, PUREX plant experience, and the thermodynamic analyses show that PuF4 is unstable to hydrolysis to PuO2 or PuO2xH2O in alkaline solution. The compounds PuF42.5H2O, NaPuF5, and Na2PuF6 are demonstrably unstable to formation of PuO2xH2O in alkaline solution. Thermodynamic and experimental information on PuF4 stability in mildly alkaline solution (e.g., pH 810) was not found. Discussion of tests to measure the solubility of PuF4 in water and in HNO3 solution does not include information on the stability of PuF4 in water but does indicate that PuF4 has low water solubility (0.00025 M Pu while the solubility of PuF3, likely as PuF3xH2O, in excess fluoride and 0.05 M HNO3 is ~0.0002 M Pu; Mandleburg et al. 1961). Anhydrous PuF4 converts to PuF42.5H2O in aqueous HF solution. The solubility of PuF42.5H2O in water (0.00012 M Pu; Deichman and Tananaev 1962) is similar to that of PuF4 suggesting both likely are controlled by the same PuF42.5H2O solid phase. The thermodynamic analysis also shows that PuF4 is stable to hydrolysis by water alone. The thermodynamic analyses indicate that chemically sufficient hydroxide ion must be present to satisfy the reaction stoichiometry of one hydroxide per fluoride.

Particle Size Impacts


Whenever a very insoluble solid quickly forms due to reaction such as metathesis, the products particle size tends to be very small. On this basis, the PuO2xH2O formed by Pu(IV) fluoride or Pu(IV) oxalate metathesis also should be small [noting that Pu(III) fluoride is rapidly oxidized to Pu(IV)]. If the tank waste system is relatively quiescent when the reaction takes place, the insoluble product could loosely adhere to the surface of the reactant particle and thus form an agglomerated PuO2xH2O product. In such a case, the surface shell formed by the precipitating PuO2xH2O product might seal the underlying particle undergoing metathesis and protect or at least impede the underlying material from further metathesis. However, given the decade or more time that the Pu(IV) fluoride and Pu(IV) oxalate particles have been present in the waste tanks, the survival of such armored particles is unlikely. Nevertheless, if left undisturbed, weak PuO2xH2O agglomerates of dimension similar to the original Pu(IV) fluoride and Pu(IV) oxalate particles may remain. Given the anticipated small size of the PuO2xH2O formed by Pu(IV) fluoride compound or Pu(IV) oxalate metathesis and the expected weak inter-particle adherence of the product PuO2xH2O, the agglomerated plutonium-bearing particles are unlikely to survive even minimal slurrying and pumping. Instead, the particles are likely to be significantly size-reduced and blended with the accompanying waste solids.

Conclusions
Process and laboratory chemical evidence shows that Pu(IV) oxalate, Pu(IV) fluoride, and sodiumplutonium fluoride double salts are unstable to decomposition by metathesis to PuO2xH2O in alkaline solution and that Pu(III) fluoride is rapidly oxidized to Pu(IV). Plutonium(IV) oxalate also is unstable to hydrolysis in neutral pH solution. Thermodynamic calculations show that PuF4 is unstable to alkaline hydrolysis but stable to hydrolysis in neutral solution. Lack of Gibbs free energy of formation data on Pu(IV) oxalate, PuF42.5H2O, and the sodium-plutonium fluoride double salts preclude calculation of the

Official Use Only C-28

RPP-RPT-50941, Rev. 0

thermodynamic stabilities of these compounds to hydrolysis to form PuO2xH2O in neutral or alkaline solution. The particle size of Pu(IV) oxalate prepared under routine PFP production conditions was ~5-7 m. Later de-inventory processing produced limited quantities of post-precipitated Pu(IV) oxalate [i.e., Pu(IV) oxalate precipitating downstream of solid-liquid separation at pan filters] that ultimately was discharged to tank farms and had unknown particle size. The particle size of PuF4 lost as aerosols to HF scrubbers in hydrofluorination processing of PuO2 at the PFP also is unknown but, because the PuF4 left as an aerosol, likely was micron to sub-micron scale. The particle sizes of PuF42.5H2O and the sodium-plutonium fluoride double salts arising from plutonium scrap dissolution operations and the PuF42.5H2O from organic solvent wash operations also are unknown but are expected to be ~1-10 m. The particle densities of Pu(C2O4)26H2O (~3-4 g/cm3), PuF3xH2O (9.32 g/cm3), PuF4 (7.04 g/cm3), PuF42.5H2O (4.89 g/cm3), NaPuF5 (6.03 g/cm3), and Na2PuF6 (5.84 g/cm3) are lower than that of PuO2 (11.46 g/cm3; Clark et al. 2006). Overall, the constituent particle sizes of the PuO2xH2O hydrolysis products arising from the metathesis reactions of Pu(IV) oxalate and the Pu(III) and Pu(IV) fluoride compounds, and the mechanical resilience of agglomerates produced by these reactions in quiescent alkaline solutions, are unknown. However, the PuO2xH2O metathesis product has extremely low solubility with the outcome being that the particles from the oxalate and fluoride compound metatheses are likely to be small (sub-micron). In addition, the product PuO2xH2O agglomerates are likely to be weak and have minimal resistance to shear caused by slurrying and pumping. No such PuO2xH2O agglomerates have yet been observed in actual tank waste samples.

References
Alenchikova, IF, LL Zaitseva, LV Lipis, VV Fomin, and NT Chebotarev. 1958. Preparation and Properties of Certain Double Fluorides of Tetrapositive Plutonium. Proceedings of the Second United Nations International Conference on the Peaceful Uses of Atomic Energy, Volume 28, Basic Chemistry in Nuclear Energy, United Nations, Geneva, Switzerland. Anonymous. 1944. Hanford Engineering Works Technical Manual, Sections A, B and C. Pages 706-716 of Section C, HW-10475 ABC, EI du Pont de Nemours, Richland, Washington. Barney, GS. 1988. Particle Size Analyses of Plutonium Fluoride Samples from the RMC Line in 234-5 Building. Internal Memo 12221-PSL88-005, January 13, 1988, Westinghouse Hanford Company, Richland, Washington. Burney, GA and FW Tober. 1965. Precipitation of Plutonium Trifluoride. I&EC Process Design and Development 4(1):28-32. Clark, DL, SS Hecker, GD Jarvinen, and MP Neu. 2006. Plutonium. Chapter 7, in The Chemistry of the Actinide and Transactinide Elements, 3rd edition, LR Morss, NM Edelstein, J Fuger, and JJ Katz, Editors, Springer, Dordrecht, The Netherlands. Dawson, JK, RM Elliott, R Hurst, and AE Truswell. 1954a. The Preparation and Some Properties of Plutonium Fluorides. Journal of the Chemical Society 558-564. Dawson, JK, RWM DEye, and AE Truswell. 1954b. The Hydrated Tetrafluorides of Uranium and Plutonium. Journal of the Chemical Society 3922-3929.

Official Use Only C-29

RPP-RPT-50941, Rev. 0

Deichman, EN and IV Tananaev. 1962. Plutonium Fluorides. Soviet Radiochemistry 4(1):56-62. Delegard, CH. 1985. Chemical Analysis of Rocky Flats Oxide Sludge and Heel. Internal letter to ET Abramowski, 65454-85-032, 25 February 1985, Rockwell Hanford Operations, Richland, Washington. Delegard, CH. 1987. Solubility of PuO2xH2O in PUREX Plant Metathesis Solutions. RHO-RE-ST-53 P, Rockwell Hanford Operations, Richland, Washington. Duffield, RB. 1960. Metathesis of Plutonium Carrier Lanthanum Fluoride Precipitate with an Alkali. US Patent 2,931,702. US Patent Office, Washington, DC. Engineers of Plutonium Process Engineering. 1968. Z Plant Report Task I-II & Recovery Operations, January 1, 1968 through December 31, 1968, ARH-294 RD (with deletions), Atlantic Richfield Hanford Company, Richland, Washington. Gelman, AD and LP Sokhina. 1958. Oxalate Complex Compounds of Plutonium (IV). Journal of Inorganic Chemistry, USSR 3(5):49-56. Gelman, AD, AI Moskvin, LM Zaitsev, and MP Mefodeva. 1962. Complex Compounds of Transuranium Elements, pages 121-123, Consultants Bureau, New York, New York. Griffo, JS, WB Brown, and FD Lonadier. 1964. Solubility of Plutonium Compounds. MLM-1191, Mound Laboratory, Miamisburg, Ohio. Guillaumont, R, T Fanghnel, J Fuger, I Grenthe, V Neck, DA Palmer, and MH Rand. 2003. Update on the Chemical Thermodynamics of Uranium, Neptunium, Plutonium, Americium and Technetium. Elsevier, Amsterdam, the Netherlands. Harmon, KM and WH Reas. 1957. Conversion Chemistry of Plutonium Nitrate. HW-49597 A, Hanford Atomic Products Operation, Richland, Washington. Haschke, JM, TH Allen, and LA Morales. 2001. Reactions of Plutonium Dioxide with Water and Hydrogen-Oxygen Mixtures: Mechanisms for Corrosion of Uranium and Plutonium. Journal of Alloys and Compounds 314:78-91. Jenkins, IL, FH Moore, and MJ Waterman. 1963. X-Ray Powder Crystallographic Data on Plutonium Oxalates and Their Isomorphs. Chemistry and Industry (London) 1:35-36. As described by JM Cleveland, 1970, The Chemistry of Plutonium, page 407, Gordon and Breach Science Publishers, Inc. Jenkins, IL, FH Moore, and MJ Waterman. 1965a. X-ray Powder Crystallographic Data on Plutonium and Other Oxalates I: The Oxalates of Plutonium(III) and Plutonium(VI) and Their Isomorphs. Journal of Inorganic and Nuclear Chemistry 27(1):77-80. Jenkins, IL, FH Moore, and MJ Waterman. 1965b. X-Ray Powder Crystallographic Data on Plutonium and Other Oxalates II: Plutonium(IV) Oxalate Dihydrate, Uranium(IV) Oxalate Hexahydrate, Uranium(IV) Oxalate Dihydrate and Thorium Oxalate Dihydrate. Journal of Inorganic and Nuclear Chemistry 27(1):81-87.

Official Use Only C-30

RPP-RPT-50941, Rev. 0

Keith, LM and AG Westra. 1987. PUREX Headend Changes to Metathesis and CRW Waste Processing. RHO-SD-CP-PTP-035, Rev. 0, Rockwell Hanford Operations, Richland, Washington. Keller, C and M Salzer. 1967. Ternre Fluoride des Typs MeIIMeIVF6 mit LaF3-Struktur. Journal of Inorganic and Nuclear Chemistry 29:2925-2934. Knights, LM. 1970. Monthly Report Plutonium Process Engineering Section. December 1970, ARH-423-RD-DEL, Atlantic Richfield Hanford Company, Richland, Washington. Mandleburg, CJ, KE Francis, and R. Smith. 1961. The Solubility of Plutonium Trifluoride, Plutonium Tetra-Fluoride, and Plutonium(IV) Oxalate in Nitric Acid Mixtures. Journal of the Chemical Society 2464-2468. Myers, MN. 1956. Thermal Decomposition of Plutonium(IV)Oxalate and Hydrofluorination of Plutonium(IV) Oxalate and Oxide. HW-45128, Hanford Atomic Products Operation, Richland, Washington. Panesko, JV. 1972. Hydrofluoric Acid Scrubber Systems. ARH-2343. Atlantic Richfield Hanford Company, Richland, Washington. Available at: http://www.osti.gov/bridge/servlets/purl/1016165iGj45w/1016165.pdf. Rudisill, T. 2011. Personal e-mail communication PuF4.2.5H2O, 23 June 2011, to C Delegard. Stakebake, JL and LM Steward. 1973. Water Vapor Adsorption on Plutonium Dioxide. Journal of Colloid and Interface Science 42(2):328-333. Wagman, DD, WH Evans, VB Parker, RH Schumm, I Halow, SM Bailey, KL Churney, and RL Nuttall. 1982. The NBS Tables of Chemical Thermodynamic Properties Selected Values for Inorganic and C1 and C2 Organic Substances in SI Units. Journal of Physical and Chemical Reference Data 11, Supplement 2. Yusov, AB, AYu Garnov, VP Shilov, IG Tananaev, MS Grigorev, and NN Krot. 2000a. Plutonium(IV) Precipitation from Alkaline Solutions. I: Effect of Precipitation and Coagulation Conditions on Properties of Hydrated Plutonium Dioxide PuO2xH2O. Radiochemistry 42(2):151-156. Yusov, AB, AYu Garnov, VP Shilov, IG Tananaev, MS Grigorev, and NN Krot. 2000b. Plutonium(IV) Precipitation from Alkaline Solutions. I: Effect of Anions on Composition and Properties of Hydrated Plutonium Dioxide PuO2xH2O. Radiochemistry 42(2):157-160.

Official Use Only C-31

RPP-RPT-50941, Rev. 0

REPORT 3 PLUTONIUM(IV) OXIDE AND UNBURNED PLUTONIUM METAL FROM OPERATIONS TO BURN PLUTONIUM METAL (BURNT METAL)

Plutonium(IV) Oxide and Unburned Plutonium Metal from Operations to Burn Plutonium Metal (BURNT METAL) Introduction
Alpha phase unalloyed plutonium (Pu) metal was prepared at the Hanford Site Plutonium Finishing Plant (PFP) in a series of chemical process steps by conversion of Pu(IV) nitrate through Pu(IV) oxalate, Pu(IV) oxide, and Pu(IV) fluoride intermediates. The pure product alpha () phase Pu metal took the physical form of a button with size and shape near that of a hockey puck. Plutonium metal was then assayed and material found not meeting product specifications for purity was recycled through PFP unit operations. Plutonium metal also was alloyed with gallium to stabilize the softer and more machinable delta () phase for use in weapons. Fabrication of weapons parts at the PFP produced machining swarf (turnings and cuttings). This swarf also was recycled through PFP operations. Over most of the PFP process history, the first step in the recycle of plutonium metal was burning the metal to form plutonium dioxide, PuO2. Ignition of the plutonium metal was performed by either flame (gas torch) or by Calrod-type electrical resistance loops. Once ignited, the Pu metal continued burning until completion or near completion. The materials occasionally were stirred during burning in plant operations to enhance air access to the metal and improve the completeness of the oxidation step. Oxide from metal burning in Hood 1 of Miscellaneous Treatment (MT-1) would be screened and charged directly into the dissolver pots in MT-2A and MT-5. However, if the oxide was to be packaged for storage or shipment, it generally would be sieved and thoroughly oxidized by heating in a pot or muffle furnace. This secondary oxidation was required to ensure complete oxidization of the metal. Observations of gains-on-ignition or GOIs often were made for oxide from metal burning when analyzed for moisture content by gravimetric loss-on-ignition. The weight gains were caused by oxidation of unburned or unreacted metal by atmospheric oxygen during heating and would offset or mask any actual moisture content. Incomplete oxidation of metal also was suspected in incidents where food pack containers holding burnt metal collapsed in storage from residual unburned metal consuming the packaging atmosphere and creating a vacuum sufficient to lead to can collapse (see, for example, cases PANEL-5 and PANEL-8 in Table 3 of Eller et al. 2004). Over the period 1973 to 1976, off-specification plutonium metal also was digested for subsequent processing by an electrolytic dissolver. The dissolution occurred in a strong nitric/hydrofluoric (HNO3/HF) acid medium that became depleted in nitrate by the accompanying reduction of nitrate to oxides of nitrogen, NOx during electrolysis (Wheelwright 1972; Oma 1977; Harlow and Olguin 1977). In practice, however, metal dissolution in the electrolytic dissolver was incomplete with very fine and readily suspendible solids observed particularly if HNO3/HF concentrations were not maintained above 10 M and 0.05 M, respectively. The two electrolytic Pu metal dissolvers operated in MT-3.

Characteristics of Burning Plutonium Metal in Air


An overview of plutonium metal burning has recently been prepared (Korinko 2009). As shown there and elsewhere, the product of burning plutonium metal in air is PuO2. However, Pu2O3 is an intermediate that forms on the surface of the burning metal (Haschke et al. 2000). The ignition temperature, metal

Official Use Only C-32

RPP-RPT-50941, Rev. 0

oxidation rate, and the PuO2 particle size distribution from burning Pu metal are strong functions of the atmosphere and the oxidation temperature (Haschke et al. 2000). In one set of tests conducted at Rocky Flats, Pu metal (in the form of flat sheet 1.270.510.1 or 0.2 cm) ignited in air at 35% relative humidity between 310C and 505C and reached a maximum temperature of 780C (Thompson 1966). The burning temperature thus exceeded the Pu metal melting point of 640C (Clark et al. 2006). The Pu metal ignition temperature increases in a step fashion with increase in the size of the solid. For Pu metal thickness less than about 0.2 mm, which includes powder and most machining chips, ignition in air occurs at ~150-200C; at greater thicknesses, ignition occurs between ~450-520C (Figure 29.6 of Haschke and Stakebake 2006). With massive Pu metal pieces, such as buttons, the oxidation is self-heating (autothermic) and the reaction rate in air is about 0.14 g Pu cm2 min1 at 500C to ~1000C, irrespective of alloying or humidity, corresponding to a linear penetration rate of about 5 mm Pu metal thickness per hour (Haschke and Allen 2002). The invariance of the burning rate over this broad temperature span has been attributed to imposition of an oxygen gas-depleted zone around the burning metal with non-reactive nitrogen gas providing a diffusion impediment to inflowing atmospheric oxygen. Because of the combined opposing effects of high reaction exothermicity and oxygen gas diffusion limits, significant but steady reaction temperatures can be attained such that bulk burning Pu metal often glows to a bright orange color (see, for example, Figure 3 of Haschke et al. 2000) and are observed to be 930C to 1000C within bulk pieces ranging from 570 to 1770 grams (Mishima 1966). With smaller Pu metal pieces, the reaction is not autothermic meaning that external heat must be supplied to sustain burning. Thus, smaller pieces ignited by a flame may be extinguished if, for example, the piece is held in a forceps which can conduct heat away. Characteristics of burning plutonium metal; including the burning of 1-3 kg Pu metal buttons and smaller 0.2-1 kg Pu metal pieces (both -phase), ~0.7 kg -phase metal pieces, casting skulls, and turnings; are described in detail by Felt (1967). The burning of Pu metal buttons was described as follows: Metal ignition usually required 60 to 70 sec of torch contact to establish a satisfactory burn. Spread of the burn throughout the metal took 12 to 15 min, and reached a peak temperature of about 825C. Following the peak temperature, a decrease of 200C occurred in the next 30 min, leveling to 600C; with a gradual decrease of 20C per hour until completely oxidized. The burning metal was a bright cherry-red color with a gradually-increasing surface oxide coating. There was no flaming or flashing from the burning mass. After reaching the peak temperature, the metal was 150C above its melting point. Containment and extinguishment became a problem of handling molten plutonium metal encased by an oxide-molten metal shell. Any disturbance of the shell would cause the metal to spew. Because autothermic conditions diminish and smothering by product PuO2 and oxygen depletion or nitrogen blanketing increases at later stages, incomplete combustion frequently is observed in burning large metal pieces. Anecdotal PFP experience has been related in which storage of oxide from burnt metal in food pack cans under air atmosphere has resulted in inward collapse of the cans (see also Table 3 of Eller et al. 2004). The inward collapse has been attributed to the presence of unburned metal which reacted with both oxygen to form PuO2 and nitrogen to form PuN, plutonium nitride. Note that it is only by consumption of nitrogen that sufficient differential pressure; i.e., >0.2 atmospheres; is generated to collapse a food pack can. Therefore, complete oxidation of Pu metal buttons often required supplemental furnace heating in air (to make so-called twice-burned oxide). Alternatively, subsequent scrap recovery

Official Use Only C-33

RPP-RPT-50941, Rev. 0

processing of burnt Pu metal buttons had to deal with variable but generally small amounts of unburned metal.

Plutonium Oxide Particle Size Found by Burning Plutonium Metal in Air


The high temperatures attained by burning large Pu metal pieces such as buttons causes sintering of the product PuO2 to occur and leads to increased PuO2 particle agglomerate size when compared with oxidation of Pu metal in air at lower temperatures. Because of the high sintering temperature and the nature of the Pu metal burning process, in which spallation of oxide layers from the massive metal occurs, the PuO2 particle agglomerate size distribution tends to much greater size than observed for PuO2 prepared from other starting materials. For example, calcination of 5 to 7 m diameter Pu(IV) oxalate produces 2 to 5 m PuO2 particles when calcined at 300 to 620C (Myers 1956). The PuO2 particles prepared by calcining plutonium salts themselves are agglomerates of much smaller crystallites that become larger with increase in calcination temperature. Crystallite sizes as functions of source material and calcination temperature as determined by X-ray diffraction line broadening techniques (i.e., by the Scherrer equation) are shown in Table 1 (Moseley and Wing 1965; Pallmer 1956). It is seen that the Pu(IV) oxalate particles calcined at 300 to 620C are agglomerates with diameters equivalent to about 800 to 1000 crystallite diameters. Later studies confirm that increasing Pu(IV) oxalate calcination temperature from 450 to 1050C increases PuO2 crystallite size and decreases the specific particle surface area (e.g., surface area per gram). However, the agglomerate particle size distribution remains relatively unchanged with increasing calcination temperature and is centered at about 5 m (Machuron-Mandard and Madic 1996).

Table 1. PuO2 Crystallite Size as Functions of Plutonium Compound and Calcination Temperature
Calcination Temp., C 240 400 600 800 1000 Crystallite Size,(a) nm, for PuO2 Prepared from Nitrate 11.5 13.0 11.8 100 100 Peroxide 3.8 5.2 27.7 48.0 100 Oxalate
(b)

Size, nm Calcination Hydroxide Temp., C Metal(c) 5.6 6.1 10.0 29.1 55.0 100 300 500 700 900 105 122 167 269 442

4.1 5.7 13.7 27.5 100

(a) Moseley and Wing 1956. (b) Crystallite sizes for PuO2 prepared by calcination of Pu(IV) oxalate at 270, 300, 480, 510, 550, and 600C were 3, 11, 11, 17, 25, and 28 nm, respectively (Pallmer 1956). (c) Metal air-oxidized at room temperature and then calcined at 100, 300, 500, 700, and 900C (Molen and White 1967).

As shown in Table 1, PuO2 crystallite sizes produced by room temperature air-oxidation of Pu metal and then calcined at 100 to 900C also increase with increasing calcination temperature (Molen and White 1967). However, the crystallites are much larger than the PuO2 crystallites produced by calcination of the plutonium nitrate, peroxide, oxalate, and hydroxide compounds. Like the PuO2 agglomerates produced by calcining the plutonium compounds, the particle size of the PuO2 agglomerates produced by calcining PuO2 obtained by room-temperature oxidation of metal did not change appreciably with subsequent

Official Use Only C-34

RPP-RPT-50941, Rev. 0

increased calcination temperature. Over 90% of the size distribution was in the range of 1 to 5 m with the largest particles being about 16 to 45 m. The larger sizes were observed at 700 and 900C. Particle size distribution data from three different studies of high temperature bulk Pu metal oxidation in air (Stewart 1961; Mishima 1966; Stakebake 1981) have been spliced into a single size distribution as shown in Table 2 and Figure 1 (Haschke 1992). The resulting mass-based spliced size distribution is seen to be bimodal with maxima centered at about 30 m and 600 m.

Table 2. Size Distribution for PuO2 Prepared by Burning Pu Metal above 500C
Size, m 0.25 0.35 0.5 1.5 3 5 8 15 30 52 77 150 300 600 1000 1700 2500 3000 Mass Fraction 3.80E-09 9.10E-09 6.70E-08 4.20E-07 4.50E-05 6.40E-05 4.20E-04 1.90E-03 1.70E-02 8.40E-03 1.10E-02 9.20E-02 1.55E-01 2.63E-01 2.23E-01 2.18E-01 4.90E-03 6.80E-03 Cumulative Mass Fraction Particle Fraction* 3.80E-09 1.29E-08 7.99E-08 5.00E-07 4.55E-05 1.09E-04 5.29E-04 2.43E-03 1.94E-02 2.78E-02 3.88E-02 1.31E-01 2.86E-01 5.49E-01 7.72E-01 9.90E-01 9.95E-01 1.00E+00 1.24E-01 1.10E-01 6.90E-02 1.10E-01 2.04E-01 6.30E-02 1.02E-01 1.24E-01 8.00E-02 7.30E-03 3.80E-03 3.20E-03 7.10E-04 1.50E-04 2.80E-05 5.50E-06 3.80E-08 2.30E-08 Cumulative Particle Fraction 1.24E-01 2.34E-01 3.03E-01 4.13E-01 6.17E-01 6.80E-01 7.82E-01 9.06E-01 9.86E-01 9.93E-01 9.97E-01 1.00E+00 1.00E+00 1.00E+00 1.00E+00 1.00E+00 1.00E+00 1.00E+00

* Particle fractions calculated assuming spherical particle shape.

Official Use Only C-35

RPP-RPT-50941, Rev. 0

0.30

0.25

0.20

Mass Fraction

0.15

0.10

0.05

0.00
0.25 0.35 0.5 1.5 3 5 8 15 30 52 77 150 300 600 1000 1700 2500 3000

Size, m

Figure 1. Plutonium Oxide Mass Size Distribution for Plutonium Metal Burned above 500C
The PuO2 particle size data presented in Table 2 is depicted in log (size) normal (cumulative mass fraction) form in Figure 2. The slope break into two segments demonstrates that the particle size distribution is bimodal. The particle size data provided in Figure 1 are supported by an Arrhenius correlation of the logarithm of plutonium oxide scale thickness as a function of inverse temperature based on observations of oxidation of Pu metal in air at 90, 257, and 400C (Martz et al. 1994). According to the Arrhenius correlation equation,(5) the plutonium oxide thickness should be about 120 m at 500C and 280 m at 1000C. These extrapolated scale thickness data are consistent with the particle sizes shown in Table 2 and Figure 1. The particle size distribution of PuO2 derived by burning metal and which had then been ground for an unspecified time in a mortar was 20 wt% between 75 and 150 m, 46 wt% between 45 and 75 m, and 34 wt% below 45 m (Martell 1974). Further grinding was successful in decreasing all PuO2 below 45 m. In yet another study, the burning of a portion of a gallium-bearing Dow metal button (apparently from Rocky Flats) at 600 to 650C produced coarse oxide such that 29 wt% did not pass a 600-m screen (Campbell and Panesko 1971). According to this report, Most of this coarse oxide consisted of small layers curled at the edges; with a few bigger, brittle layers which were retrievable with tweezers. These accounts thus confirm the coarseness of the PuO2 produced by self-burnt metal.

(5) PuO2 thickness, m = 1043 e(-1673/T) where T is in K.

Official Use Only C-36

RPP-RPT-50941, Rev. 0

10000

1000

Size, m

100

10

0.01

0.1

1 2

10

50

90 95 98 99

99.9 99.99

Cumulative Mass%

Figure 2. Cumulative Mass Size Distribution for Plutonium Oxide from Metal Burned above 500C
Efforts to find information in the technical literature on the size of the constituent PuO2 crystallites arising from high temperature burning of metal were not successful. However, in analogy with 900 to 1000C calcination of other plutonium-bearing materials to form PuO2, it is expected that the crystallite sizes are about a factor of 10- to 50-times smaller than the agglomerated particles arising from burning plutonium metal in air. The PuO2 particles generated by burning bulk Pu metal pieces in air thus are uncommonly large compared with the PuO2 particles prepared by calcination of plutonium compounds or the PuO2 found from room temperature air-oxidation of Pu metal. According to Figure 2, 50 mass% of the PuO2 produced by burning Pu metal in air is greater than 400 m diameter and 99.9 mass% is greater than 10 m diameter.

Corrosion of Plutonium Metal Residues from Burnt Plutonium Metal in Tank Waste Solution
Besides containing large agglomerates, the product from burning scrap Pu metal in air at PFP likely also contained small amounts of unburned Pu metal because of diminishing quantity of burning material and self-extinguishing by smothering. The plutonium oxide and the associated unburned metal then was processed for recovery at the PFP by dissolution in B acid, a mixture of ~6-12 M nitric acid (HNO3) containing ~0.5 M fluoride. Both PuO2 and the Pu metal dissolve in this reagent, but dissolution rates for the large particle size and high-fired oxide (because it was formed at temperatures around 800-1000C) would be slow. Dissolution of Pu metal in this reagent also would be slow as the metal forms a protective oxide layer in HNO3 that also must yield to fluoride. Furthermore, the metal likely would have relatively low surface area relative to its size (i.e., the metal particles would be millimeter- rather than m-scale). Discharge of these plutonium-rich solids from PFP processes to the tank farms would be difficult owing to their rapid settling caused by their large particle size and high density. However, these particles also represent the materials of most concern to criticality safety in tank farms. The PuO2 and Pu metal discharged to Hanford tank farms are not expected to be altered by interaction with the alkaline tank farm waste solutions. As noted elsewhere in discussions of Ostwald ripening and

Official Use Only C-37

RPP-RPT-50941, Rev. 0

particle growth of PuO2xH2O, PuO2 is thermodynamically stable in alkaline solution. Therefore, the large sintered PuO2 particles formed by burning Pu metal would remain as large PuO2 particles in tank waste. Studies of Pu metal corrosion rate aqueous solution are few. An overview of Pu metal corrosion rates in aqueous solution provided in Table 3 shows it to be stable in alkaline solution (Kolman 2002, as abstracted from prior studies). Similarly, a study of -phase Pu-Ga alloy metal corrosion in saturated alkaline lithium hydroxide (~ 4.8 M LiOH; pH ~13.7) showed no measureable corrosion occurred as could be determined by weight change or hydrogen gas evolution (Haschke 1995).

Table 3. Summary of Aqueous Corrosion Rates for Plutonium Metal (Kolman 2002)
Rapid Attack HNO3/0.005 M HF HCl HBr 72% HClO4 85% H3PO4 Conc. CCl3COOH Artificial sea water Slow Attack H 2O HF Dilute H2SO4 Dilute CH3COOH Conc. CF3COOH Dilute CCl3COOH Tapwater No Attack HNO3 Conc. H2SO4 Alkaline solution Glacial CH3COOH

More recent studies and compilation of prior work shows that Pu metal corrosion rates increase with increasing salt concentration (see Figure 1 of Haschke 1995). Although Pu metal corrosion rates increase with increasing sodium chloride (NaCl) concentration and decreased pH, sodium nitrate (NaNO3), the solute generally of highest concentration in Hanford tank waste, has much less impact (Kolman and Colletti 2008). No studies of the corrosion of plutonium metal in Hanford tank waste were found. Based on the published literature, which show that Pu metal is stable to corrosion in alkaline solution and is only marginally affected by NaNO3, it is likely that Pu metal corrodes at very slow rates in tank waste solution. However, Hanford tank waste is complex and contains many dissolved components. The impacts of REDOX-active materials present in the tank waste, such as nitrite, NO2-, or organics, (e.g., glycolate, HOCCO2-) have not been examined and may be significant. Any plutonium metal that has corroded is certain to be PuO2xH2O and would likely have characteristically small particle size owing to its extremely low solubility. Because no studies of plutonium metal corrosion in Hanford tank waste have been performed and other studies of Pu metal corrosion in alkaline solution showed the metal to be stable, no plutonium metal corrosion product could have been made to be studied.

Conclusions
The products of plutonium metal burning operations in the PFP, PuO2 and unburned Pu metal, are expected to be the largest and most dense Pu-bearing materials potentially discharged to the Hanford tank waste system. Over 99 wt% of the plutonium oxide particles produced by burning scrap Pu metal buttons (and possibly other metal forms such as turnings and skulls) are expected to be greater than 10 m. Residues of unburned plutonium metal are anticipated to be present in the products from Pu metal button burning unless supplemental furnace oxidation had been used. The residual Pu metal and the large and

Official Use Only C-38

RPP-RPT-50941, Rev. 0

high-fired PuO2 particles produced in burning Pu metal scrap would be difficult to dissolve in PFP operations and, with difficulty (owing to their high sedimentation rates), could have been discharged to the Hanford waste tanks. Once there, little change in the chemical or physical form of the high-fired PuO2 would be expected as this material is stable in the extant alkaline solution conditions. Plutonium metal corrosion rates in alkaline solution are low. Although no confirmatory lab tests have been done, it is likely that any Pu metal lost to the Hanford tank waste would have corroded in the ~20 years or greater time since discharge. In such case, the product would be PuO2xH2O having the extremely small particle size characteristic of this material when formed in aqueous solution.

References
Campbell, MH and JV Panesko. 1971. Tentative Indications from the Burning of a Portion of One Dow Button and Dissolution of Oxide. Internal letter to LM Knights, October 13, 1971, Atlantic Richfield Hanford Company, Richland, Washington. Clark, DL, SS Hecker, GD Jarvinen, and MP Neu. 2006. Plutonium. Chapter 7 in The Chemistry of the Actinide and Transactinide Elements, 3rd edition, LR Morss, NM Edelstein, J Fuger, and JJ Katz, Editors, Springer, Dordrecht, The Netherlands. Eller, PG, RW Szempruch, and JW McClard. 2004. Summary of Plutonium and Oxide Storage Package Failures. LA-UR-04-1453, Los Alamos National Laboratory, Los Alamos, New Mexico. Felt, RE. 1967. Burning and Extinguishing Characteristics of Plutonium Metal Fires. ISO-756, Isochem, Incorporated, Richland, Washington. Harlow, DG and LJ Olguin. 1977. Current Status and Safety Hazards in Miscellaneous Treatment. Internal letter to RB Gelman, March 8, 1977, Atlantic Richfield Hanford Company, Richland, Washington. Haschke, JM. 1992. Evaluation of Source-Term Data for Plutonium Aerosolization. LA-12315-MS, Los Alamos National Laboratory, Los Alamos, New Mexico. Haschke, JM. 1995. Reactions of Plutonium and Uranium with Water: Kinetics and Potential Hazards. LA-13069-MS, Los Alamos National Laboratory, Los Alamos, New Mexico. Haschke, JM, TH Allen, and LA Morales. 2000. Surface and Corrosion Chemistry of Plutonium. Los Alamos Science, No. 26, NG Cooper, editor, Los Alamos National Laboratory, Los Alamos, New Mexico. Available at: http://www.fas.org/sgp/othergov/doe/lanl/pubs/00818031.pdf. Haschke, JM and TH Allen. 2002. Equilibrium and Thermodynamic Properties of the PuO2+x Solid Solution. Journal of Alloys and Compounds 336:124131. Haschke, JM and JL Stakebake. 2006. Handling, Storage, and Disposition of Plutonium and Uranium. Chapter 29 in The Chemistry of the Actinide and Transactinide Elements, 3rd edition, LR Morss, NM Edelstein, J Fuger, and JJ Katz, Editors, Springer, Dordrecht, The Netherlands. Kolman, DG. 2002. The Aqueous Corrosion Behavior of Plutonium Metal Exposed to Nitric Acid Solution. LA-UR-02-5920, Los Alamos National Laboratory, Los Alamos, New Mexico.

Official Use Only C-39

RPP-RPT-50941, Rev. 0

Kolman, DG and LP Colletti. 2008. Aqueous Corrosion Behavior of Plutonium Metal and Plutonium Gallium Alloys Exposed to Aqueous Nitrate and Chloride Solutions. Journal of the Electrochemical Society, 155 (12):C565-C570. Korinko, P. 2009. Review of Plutonium Oxidation Literature. SRNL-L7100-2009-00046 Rev. 1, Savannah River National Laboratory, Aiken, South Carolina. Available at: http://sti.srs.gov/fulltext/SRNL-L7100-2009-00046R1.pdf. Machuron-Mandard, X and C Madic. 1996. Plutonium Dioxide Particle Properties as a Function of Calcination Temperature. Journal of Alloys and Compounds 235:216-224. Martell, CJ. 1974. The Effect of Particle Size on the Carrier-Distillation Analysis of PuO2. LA-UR-74-356, Los Alamos Scientific Laboratory, Los Alamos, New Mexico. Martz, JC, JM Haschke, and JL Stakebake. 1994. A Mechanism for Plutonium Pyrophoricity. Journal of Nuclear Materials 210:130-142. Mishima, J. 1966. Plutonium Release Studies II. Release from Ignited Bulk Metallic Pieces. BNWL-357, Battelle Northwest, Richland, Washington. Molen, GF and RD White. 1967. Properties of Plutonium Oxide, Part II. RFP-927, The Dow Chemical Company, Golden Colorado. Note that Part I is the report by Moseley and Wing (1965). Moseley, JD and RO Wing. 1965. Properties of Plutonium Dioxide. RFP-503, The Dow Chemical Company, Golden, Colorado. Myers, MN. 1956. Thermal Decomposition of Plutonium(IV)Oxalate and Hydrofluorination of Plutonium(IV) Oxalate and Oxide. HW-45128, Hanford Atomic Products Operation, Richland, Washington. Oma, KH. 1977. Plutonium Metal Electrolytic Dissolution Flowsheet for the Plutonium Reclamation Facility. ARH-F-109, Atlantic Richfield Hanford Company, Richland, Washington. Pallmer, PG Jr. 1956. Crystallite Sizes of PuO2 Powder. HW-44585, Hanford Atomic Products Operation, Richland, Washington. Stakebake, JL. 1981. Kinetics for the Reaction of Hydrogen with a Plutonium-1 Weight Percent Gallium Alloy Powder. Journal of the Electrochemical Society 128(11):2383-2388. Stewart, K. 1961. Vixen A Trials: Experiments to Study the Release of Particulate Materials During the Combustion of Plutonium, Uranium and Beryllium in a Petrol Fire. AWRE Report T15/60, Atomic Weapons Research Establishment, Aldermaston, United Kingdom. As cited by Haschke 1992. Thompson, MA. 1966. The Oxidation and Ignition of Plutonium. RFP-491, The Dow Chemical Company, Rocky Flats Division, Golden, Colorado. Wheelwright, EJ. 1972. Development of an Electrolytic Dissolver for Plutonium Metal. BNWL-1642, Battelle Pacific Northwest Laboratories, Richland, Washington.

Official Use Only C-40

RPP-RPT-50941, Rev. 0

REPORT 4 INTERFACIAL CRUD DISPOSITION IN ALKALINE TANK WASTE

Interfacial Crud Disposition in Alkaline Tank Waste


Interfacial crud was generated by the interaction of plutonium and other metal ions with solvent degradation products in the Plutonium Reclamation Facility and by the presence of finely particulate solids that sought the organic-aqueous interface in solvent extraction processing. These crud materials, both of which could contain plutonium, acted as vectors for plutonium losses from the Plutonium Finishing Plant.

Introduction
Plutonium recovery occurred in the PFP Plutonium Reclamation Facility (PRF) by solvent extraction in 20% [TBP; OP(O(CH2)3CH3)3] dissolved in carbon tetrachloride (CCl4) diluent. An early flowsheet outlines the key features of this process (Bruns 1960). The plutonium, originating from dissolved impure scrap and present in the tetravalent oxidation state, Pu(IV), was extracted from aqueous solution containing ~2.5 M nitric acid (HNO3) and other dissolved metal nitrate salts (principally magnesium, calcium, iron, and aluminum as well as trace americium-241, 241 Am, in-grown from radioactive decay of plutonium-241, 241Pu). The extraction by the TBP-CCl4 organic solution, which is denser than the aqueous phase, occurred in a pulse column in the PRF glovebox canyon. Scrub solution, nominally 1 M HNO3, was used to remove impurities from the Pu-loaded TBP by another pulse column operation. The purified Pu then was stripped from the TBP-CCl4 into the aqueous phase by contact in another pulse column with solution containing a chemical reductant, usually hydroxylamine, HONH2. The hydroxylamine served to chemically reduce the Pu(IV) dissolved in the TBP-CCl4 to trivalent plutonium, Pu(III), which is not extracted in TBP-CCl4. The organic then was recycled to repeat the process while the plutonium aqueous solution was processed further to plutonium metal and oxide product forms. Both chemical and radiolytic decomposition of TBP occur because of the respective effects of acid hydrolysis in HNO3 (Burger 1955; Lloyd and Fellows 1985) and alpha radiation damage (Lloyd and Fellows 1985). The decompositions in both cases involve the stepwise loss of butoxy [-O(CH2)3CH3] groups as butanol, H3C(CH2)2CH2OH, to make, in order, dibutyl phosphate [DBP; -O2P(O(CH2)3CH3)2], monobutyl phosphate [MBP; 2-O3P(O(CH2)3CH3)], and phosphate, PO43-. When the solvent is properly maintained by alkaline clean-up steps in the solvent extraction process, decomposition beyond DBP normally is not given an opportunity to occur. However, complexes form by the interaction of DBP with polyvalent metal ions, including Pu4+ and Pu3+, and, in the PRF, the Pu(III,IV)-DBP complexes have high affinity for the TBP-CCl4 phase that resists stripping. Precipitation of thorium(IV), chemically analogous to Pu(IV), as the compound Th(DBP)4 occurs and appears at the interface between the organic and aqueous phases. This has led to the speculation that Pu(DBP)4 also can precipitate (Zimmer and Borchardt 1986). Finely divided solids such as silica gels and clays present in solvent extraction systems also seek the organic-aqueous interface and can create and stabilize emulsions that impede and ultimately thwart the rapid organic-aqueous disengagement needed for successful solvent extraction operations. The solids appearing at the interface also can include undissolved solid plutonium dioxide, PuO2. These solid materials and the less soluble metal-organic complexes appearing at the organic-aqueous interface are known collectively as crud.

Official Use Only C-41

RPP-RPT-50941, Rev. 0

In the PFP, dissolution of Pu-bearing scrap almost invariably produced silica gels from the acid decomposition of tramp silicate minerals present in the scrap (e.g., ash, crucible impurities) or present in infiltrated dust particles. In practice, any time solution feeds with high solids loading were processed, problems with crud formation would arise. Therefore, crud was an early and continuing issue in proper solvent extraction operations in the PRF. To help address problems with crud accumulating randomly in the organic-aqueous mixtures, the PRF operation flowsheet called for addition of Mistron, a magnesium talc mineral, to the solvent extraction system. The Mistron, added at about 100 ppm to the feed solution (Bruns 1960), served to coagulate the crud materials at the interface and help in efficiently sweeping them from the interface by occasional decantation. As described by Klem (1972): CA Column interface solids (crud) are a complex mixture formed by mistron (MgSiO3) and other impurities in the feed or by combination of metallic impurities with solvent decomposition product to form insoluble compounds. The crud fills the columns glass disengaging section and must be removed to prevent it overflowing with the high salt aqueous waste stream (CAW) to the Waste Treatment Facility. Crud removal is achieved by jetting the solids together with up to 10 liters of aqueous and organic to the Z-18 crib. The crud would appear at the organic-aqueous interface or, occasionally, at the top of the aqueous layer in the CA column which provided the primary plutonium solvent extraction contact. When the interface crud layer became too thick in the CA column (i.e., measured in feet, not inches), the crud was pushed to overflow to the CA column waste (CAW) centrifuge by raising the interface level. The supernatant solution from the centrifuge would be discharged to the CAW while the solids were collected for later processing (burning and leaching) or for crib Z 18 discharge (Klem 1972). However, overloading of the centrifuge would cause loss of crud with the supernatant solution for discharge with the CAW that, after 1973, ultimately would go to tank farms.

Plutonium Losses through Crud Discards


Crud solids would contain not only the silica gels and the Mistron but also some undissolved plutonium, most likely present as PuO2, as well as Pu contained in poorly soluble organic (DBP) complexes. The decanted crud sometimes was collected for burning and plutonium recovery but crud also was lost, in the 1970s, to the Z-18 crib with the CAW discharges and later, entrained with the CAW, inadvertently to direct discard to the tank farms. With decreasing drive to recover plutonium from relatively lean and hard-to-process scraps like crud, simple discard of collected crud by cementation occurred by the late 1980s even as inadvertent loss to the tank farms by entrainment with the CAW continued. The crud contained highly variable amounts of plutonium. In one account (Panesko 1971a), a 500-mL sample of crud was filtered, washed with CCl4, and dried to produce 6.7 grams of fine, light, gray solids containing about 2.5 grams of plutonium (i.e., 37 wt% Pu in the solids or 5 grams Pu/liter of initial wet crud). Other crud samples reported in the same letter contained solids that were 26 wt% and 0.8 wt% plutonium (the original crud volumes were not reported). In another account (Panesko 1971b), 120 grams of dried centrifuged solids collected from the CAW of a burnt oxide run were described as light, gray, homogeneous (flour-like) solids. The 120 g of dried solids arose from ~300 mL of settled crud sludge. The dried solids contained 13 to 16 wt% Pu (i.e., ~60 grams of Pu/liter of initial wet crud) with the major other elements being silicon and phosphorus. As noted in the letter report, the presence of silicon and phosphorus suggested that an organic emulsion was the source of the centrifuged crud.

Official Use Only C-42

RPP-RPT-50941, Rev. 0

Plutonium losses also occurred by its retention in Pu-DBP compounds that would partition to the crud or be removed from the organic phase by the organic clean-up performed by treatment with sodium carbonate, Na2CO3, solution. With the carbonate treatment, the Pu-DBP complex would be decomposed to form water-soluble sodium dibutyl phosphate, NaDBP, and either dissolved aqueous Pu(IV) carbonate complexes, if the carbonate concentration were sufficiently high, or precipitated Pu(IV) hydrous oxide, PuO2xH2O, as an extremely finely divided solid (nanometer-scale) in the wash if the carbonate concentration were depleted.

Fate of Plutonium-Crud in the Tank Waste


As described, plutonium in crud took two forms as a Pu(III)- or Pu(IV)-organic complexes, generally with DBP, and as undissolved scrap solids, primarily as PuO2. These Pu-bearing streams would have been discharged by entrainment in the CAW. Also present in the organic phase would be Pu-DBP complexes that were scrubbed from the organic by carbonate wash in the COO column waste stream and sent to waste. Both CAW and COO waste would have been treated in D-5 before discharge to tank farms and thus would have been blended with ferric nitrate as a neutron absorber in D-5 and with excess sodium hydroxide, either in D-5 or in waste blending operations further downstream. The Pu-DBP complexes, whether present in the crud or scrubbed from the organic, would hydrolyze and, in the case of Pu(III), oxidize in the highly alkaline tank waste to form finely divided PuO2xH2O. Because ferric nitrate was also present, coprecipitation of the Pu(IV) with the ferric hydroxide, Fe(OH)3, would occur when the stream was made alkaline. However, any solid PuO2 or other solid that had already resisted acid digestion at PFP prior to solvent extraction treatment in the PRF would remain in that form when discharged to tank farms and, given the low solubility of PuO2 in alkaline media, would remain unchanged.

References
Bruns, LE. 1960. Solvent Extraction Flowsheet for New Plutonium Reclamation Facility. HW-64851, Hanford Atomic Products Operation, Richland, Washington. Burger, LL. 1955. The Chemistry of Tributyl Phosphate A Review. HW-40910, Hanford Atomic Products Operation, Richland, Washington. Klem, MJ. 1972. Evaluation of Process Test PRF 67-5 Filtration of CA Column Interface Jettings. Internal note, June 20, 1972, Atlantic Richfield Hanford Company, Richland, Washington. Lloyd, MH and RL Fellows. 1985. Alpha Radiolysis and Other Factors Affecting Hydrolysis of Tributyl Phosphate. ORNL/TM-9565, Oak Ridge National Laboratory, Oak Ridge, Tennessee. Panesko, JV. 1971a. CA Column Interface Crud Sampled on September 27, 1971. Internal letter to AJ Low, October 1, 1971, Atlantic Richfield Hanford Company, Richland, Washington. Panesko, JV. 1971b. CAW Centrifuge Solids from Burnt Oxide Run. Internal letter to AJ Low, August 26, 1971, Atlantic Richfield Hanford Company, Richland, Washington. Zimmer, E and J Borchardt. 1986. Crud Formation in the PUREX and THOREX Processes. Nuclear Technology 75:332-337.

Official Use Only C-43

RPP-RPT-50941, Rev. 0

REPORT 5 COPRECIPITATION FROM NITRIC ACID MEDIA OF PLUTONIUM(IV) WITH URANIUM(VI) IN ALKALINE SOLUTION

Coprecipitation from Nitric Acid Media of Plutonium(IV) with Uranium(VI) in Alkaline Solution Introduction
The interaction of mixed nitric acid solutions of tetravalent plutonium, Pu(IV), and hexavalent uranium, U(VI), in their coprecipitation in alkaline solution is of interest in understanding the disposition of certain plutonium streams relegated to Hanford tank waste. Both of the individual Pu(IV) and U(VI) components produce low solubility solid phases when made alkaline. Acidic Pu(IV) nitrate solution precipitates in alkaline media to form plutonium(IV) hydrous oxide, PuO2xH2O, while acidic U(VI) nitrate solution precipitates in alkaline media to form clarkeite, Na2U2O73H2O (alternatively, Na[UO2(O)OH]H2O). Uranium(VI) typically is present in a vastly greater concentration than the Pu(IV) such that Pu(IV), in coprecipitation with the U(VI), would be incorporated into the clarkeite phase. The solubilities of Pu(IV) and U(VI) in strongly alkaline solution are compared in Figure 1. The solubility curves show that both solid phases are amphoteric, becoming more soluble as the NaOH concentration increases, and that U(VI) in 1 M ionic strength (adjusted using sodium perchlorate) is more soluble than Pu(IV) in a mixture of 1 M sodium nitrate (NaNO3) with 1 M sodium nitrite (NaNO2). For example, the uranium molar concentration above clarkeite is about 200-times greater than the molar concentration of plutonium over Pu(IV) hydrous oxide in 1 M NaOH.

Coprecipitation of Plutonium with Uranium in Alkaline Solution


Coprecipitation interactions of Pu(IV) with multicharged metals aluminum, iron, nickel, cobalt, chromium, zirconium, lanthanum, and uranium [Al(III), Fe(III), Ni(II), Co(III), Cr(III), Zr(IV), La(III), and U(VI), present as UO22+ in acid solution], all found in acidic wastes from plutonium processing, have been studied at plutonium concentrations comparable to those of the metals by reacting HCl solutions of the plutonium-metal salt pairs with NaOH solution at various terminal NaOH concentrations (Fedoseev et al. 1998). Infrared spectroscopy (IR) and X-ray diffractometry (XRD) of the solids and study of the differential dissolution rates of the plutonium/metal oxide, hydroxide, or salt precipitates in HCl solution showed that Pu(IV) interacts strongly with U(VI) as well as with Fe(III), Co(III), Cr(III), Zr(IV), and La(III). The IR spectra of the mixed Pu(IV) precipitates with these metal ions differed significantly from those of the individual components. In addition, the XRD patterns for the mixed precipitates showed no PuO2xH2O reflections. Because the plutonium dissolution rates in HCl were more rapid for the mixed compounds than for similarly prepared pure PuO2xH2O solid phase, intimate interaction of the plutonium with the associated metal in the formation of the coprecipitate is inferred. However, under the same preparation conditions, no analogous IR, XRD, or differential dissolution evidence of interaction was found between Pu(IV) and Ni(II) and little interaction of Pu(IV) with Al(III) hydroxides was found.

Official Use Only C-44

RPP-RPT-50941, Rev. 0

0.0001
Na2U2O7.3H2O PuO2.xH2O

0.00001

0.000001

[An], M
0.0000001 0.00000001 0.000000001 0.01

0.1

10

[OH ], M
Figure 1. Solubility of Clarkeite in I = 1.0 M (NaClO4) after One Year (Yamamura et al. 1998) and Pu(IV) Hydrous Oxide in 1 M NaNO3 / 1 M NaNO2 after 56 Days (Delegard 1985) as Functions of NaOH Concentration
These data show that intimate interaction of plutonium(IV) with uranium(VI), present as Na2U2O73H2O, occurs in the coprecipitation of Pu(IV) with U(VI) in alkaline solution even under relatively high Pu(IV) concentration conditions where precipitation of a uranium-free plutonium-bearing phase (e.g., PuO2xH2O) might be favored. Although no tests of Pu(VI) coprecipitation with U(VI) were performed, it is fully expected that complete and intimate combination of these two isomorphous components also should occur. In further work reported in the same study (Fedoseev et al. 1998), four preparations of tracer amounts of Pu(IV) in 1 M nitric acid (HNO3) solution containing U(VI), with or without added Fe(III) and Al(III), were precipitated in excess NaOH solution. The test solution compositions are shown in Table 1. Dissolution of plutonium from the washed mixed alkaline precipitates containing U(VI) but no Fe(III) were virtually complete in 0.25 and 0.5 M NaHCO3 solutions, whether or not Al(III) was present. The precipitates in these cases also dissolved almost completely, as would be expected given the high U(VI) solubility in carbonate solution. However, with Fe(III) present in the precipitate, less than 10% of the tracer plutonium dissolved. Even heating of the Fe(III)-bearing solids in the bicarbonate solution failed to leach the plutonium. These findings indicate the preferential coprecipitation of plutonium with Fe(III) over Al(III) or U(VI) compounds and the resistance of Fe(III) precipitates to dissolution by bicarbonate.

Official Use Only C-45

RPP-RPT-50941, Rev. 0

Table 1. Resistance of Pu to Carbonate Leaching from U, Fe, and Al Co-Precipitates


Metal:Pu Mole Ratio U(VI) 1000 1000 1000 1000 Fe(III) 0 0 390 390 Al(III) 0 400 0 400 Pu Dissolved in 0.25 or 0.5 M NaHCO3 95% 95% 5%-8% 5%-8%

Three different simulated acidic PUREX and Savannah River Site H-Canyon modified (HM) PUREX high-activity and low-activity waste (HAW and LAW) solutions of compositions given in Table 2 and containing varying (sometimes zero) concentrations of U(VI), Fe(III), Al(III), Mn(II), and Ni(II) nitrates with tracer concentrations of plutonium were treated to achieve 0.6 M NaOH (Hobbs 1999). Both the initial and final concentrations of Pu and U after the alkaline treatment are shown in Table 2. It is seen that the uranium concentrations observed in the 0.6 M NaOH [Hobbs 1999; (1.0-2.7)10-5 M] are about a factor of 2 greater than observed by Yamamura and colleagues (1998) after one year of aging in NaOH solution (~810-6 M; see Figure 1). These findings suggest that sodium diuranate, Na2U2O73H2O, controlled the uranium concentration. The plutonium concentrations are much more variable across the diverse solution compositions. In the blank test containing only plutonium in 0.6 M NaOH, the average Pu solution concentration was 8.410-7 M (610-7 M) after contact times of 1, 7, 14, 28, 42, and 56 days. This value was about a factor of 20 higher than the ~410-8 M concentration predicted by the solubility of PuO2xH2O shown in Figure 1 (Delegard 1985).

Table 2. Composition of PUREX-Type Solutions for Plutonium Coprecipitation Testing


Acid Feed Concentration, M Solution PUREX-HAW PUREX-LAW HM-HAW HM-LAW U-Blank Pu-Blank Pu/U-Blank Pu 7.110
-6

Concentration in 0.6 M NaOH, M Mn Ni 5.710 0 0 0 0 0 0


-2

U 4.210
-2

Fe 1.110
-1

Al 4.210 0 1.2100 0 0 0 0
-1

Pu 2.810
-9

U 1.710-5 2.010-5 1.010-5 2.910-6 1.910-5 Not detd. 2.710-5

0 2.510-2 0 0 0 0 0

9.010-6 8.510-5 0 0 7.110-6 7.110-6

3.810-3 1.210-5 6.210-6 4.210-2 0 4.210-2

5.310-1 1.810-2 1.510-1 0 0 0

4.610-9 1.510-6 Not detd. Not detd. 8.410-7 3.510-9

It is seen that plutonium coprecipitated with Fe(III) and U(VI) in the tests of simulated PUREX plant wastes made 0.6 M in NaOH. With Fe(III) or U(VI) present in the original solutions, the plutonium concentrations ranged from (2-4)10-9 M, about 100-times lower than the observed solubility of pure PuO2xH2O prepared in side-by-side tests. However, little plutonium removal occurred from simulated HM-HAW solution containing 1.2 M aluminum but low (1.210-5 M) uranium and low (1.810-2 M) iron concentrations. Thus, plutonium effectively coprecipitated with iron and uranium solid phases but not

Official Use Only C-46

RPP-RPT-50941, Rev. 0

with aluminum solid phases in agreement with the findings of Fedoseev and colleagues (1998). Partial crystallization of the precipitates also occurred with aging. After two months, the precipitates were analyzed by XRD. Although most of the solids remained amorphous, crystalline goethite (-FeOOH), gibbsite [-Al(OH)3], bayerite [-Al(OH)3], and sodium diuranate (described as Na2U2O7) were identified. Significantly, aging did not affect the plutonium solution concentration meaning that the Pu remained associated with the U and Fe carrier solids. Tests to determine the effects of coprecipitation conditions were run with a mixed simulant solution of composition 1 M HNO3, 0.096 M UO2(NO3)2, 0.246 M Fe(NO3)3, 0.033 M Al(NO3)3, 0.001 M Na2SiO3, and 1.910-4 M Pu(NO3)4 (Krot et al. 1998). The Pu(IV) coprecipitated from this solution effectively, decreasing to (5-50)10-9 M Pu in centrifuged supernates over a wide range of precipitation conditions (0.01 to 1 M excess NaOH, direct/reverse strike mixing, 40C and 80C mixing and digestion, varied U/Fe/Al ratios). The uranium concentrations found at 0.01 M and 1 M NaOH were consistent with the concentrations observed by Yamamura and colleagues (1998). As in Hobbs (1999) studies, the solid phases formed were largely amorphous. However, solids invariably included sodium diuranate (Na2U2O7), as was observed by Hobbs (1999) and Yamamura and colleagues (1998), occasionally goethite (-FeOOH), and, in one case, bayerite [-Al(OH)3)]. The solids crystallite particle size distributions centered at 2 to 3 nm with the Fe(III) hydroxide crystallites slightly smaller than those of sodium diuranate. Parametric studies also were performed on alkaline coprecipitation of 1.910-4 M Pu(IV) from HNO3 solutions containing U(VI), Fe(III), and Al(III) in varying concentrations summing to 37.5 g (metal)/L concentration (Krot et al. 1998). The sodium diuranate precipitate dissolved extensively when contacted with simulated high carbonate waste solution, 1 M Na2CO3 in 5 M NaNO3, and a fraction of the plutonium leached into solution. Precipitating aluminum with uranium did not diminish plutonium leaching. However, plutonium leaching by carbonate-bearing solution was significantly lower for solids having Fe(III) than for solids containing U(VI) only or U(VI) and Al(III) but no Fe(III). Plutonium leaching by a simulated alkaline waste solution containing a high concentration of organic complexing agent (0.1 M EDTA in 1 M NaOH and 5 M NaNO3) was low from U/Fe/Al solids of any studied composition. Therefore, the association of plutonium with the U, Fe, and Al macro-components increased in the order Al << U < Fe.

Conclusions
Experiments performed in several laboratories and in different test regimes confirm that Pu(IV) coprecipitates effectively with U(VI) in alkaline solution. Characterization of precipitates from roughly equimolar Pu and U indicate that interaction of Pu with U is intimate and that PuO2xH2O does not form as a separate phase. The coprecipitation of micro-quantities of Pu with macro-quantities of U produces sodium diuranate or clarkeite, Na2U2O73H2O. However, the host sodium diuranate phase is susceptible to dissolution in carbonate or bicarbonate solution. With the dissolution of the host sodium diuranate, dissolution of the contained plutonium also occurs. The tests also show that in the presence of both Fe(III) and U(VI), plutonium preferentially is carried from the alkaline solution with the precipitating ferric hydroxide. When coprecipitated with Fe(III) hydroxide phases, the plutonium is much less susceptible to leaching by carbonate solutions than when coprecipitated with sodium diuranate. Coprecipitation of plutonium with aluminum is negligible.

Official Use Only C-47

RPP-RPT-50941, Rev. 0

References
Delegard, CH. 1985. Solubility of PuO2xH2O in Alkaline Hanford High-Level Waste Solution. RHO-RE-SA-75 P, Rockwell Hanford Operations, Richland, Washington. Available at: www.osti.gov/bridge/servlets/purl/5402793-8aDMm2/5402793.pdf. Also found as Solubility of PuO2xH2O in Alkaline Hanford High-Level Waste Solution, CH Delegard, 1987, Radiochimica Acta 41:11-21. In addition, in Figure 1, one datum at 1 M NaOH and all data below 1 M NaOH are from CH Delegard unpublished sources. Fedoseev, AM, NN Krot, NA Budantseva, AA Bessonov, MV Nikonov, MS Grigoriev, AYu Garnov, VP Perminov, and LN Astafurova. 1998. Interaction of Pu(IV,VI) Hydroxides/Oxides with Metal Hydroxides/Oxides in Alkaline Media. PNNL-11900, Pacific Northwest National Laboratory, Richland, Washington. Available at: http://www.osti.gov/bridge/servlets/purl/665966ij8eO9/webviewable/665966.pdf. See also: Fedoseev, AM, NA Budantseva, MV Nikonov, and MS Grigorev. 2000. Interaction of Pu(IV) Hydroxide with Hydroxides of Some d and f Elements in Alkaline Solutions: I. Systems Pu(IV)Fe(III) and Pu(IV)-Ni(II). Radiochemistry 42(5):405-412. Yusov, AB, NA Budantseva, AV Ananev, and AM Fedoseev. 2000. Coprecipitation of Aluminum with Hydroxides of Tetra-, Penta-, and Hexavalent Actinides. Radiochemistry 42(5):413-416. Hobbs, DT. 1999. Precipitation of Uranium and Plutonium from Alkaline Salt Solutions. Nuclear Technology 128:103-112. Krot, NN, VP Shilov, AM Fedoseev, AB Yusov, AA Bessonov, NA Budantseva, SI Nikitenko, GM Plavnik, TP Puraeva, MS Grigoriev, AYu Garnov, AV Gelis, VP Perminov, and LN Astafurova, and CH Delegard. 1998. Alkaline Treatment of Acidic Solution from Hanford K Basin Sludge Dissolution. PNNL-11944, Pacific Northwest National Laboratory, Richland, Washington. Available at: http://www.osti.gov/bridge/product.biblio.jsp?osti_id=2507. Yamamura, T, A Kitamura, A Fukui, S Nishikawa, T Yamamoto, and H Moriyama. 1998. Solubility of U(VI) in Highly Basic Solutions. Radiochimica Acta 83:139-146.

Official Use Only C-48

RPP-RPT-50941, Rev. 0

APPENDIX D PLUTONIUM AND ABSORBER CONCENTRATIONS IN FAST AND LOW SETTLING FRACTION

Official Use Only D-1

RPP-RPT-50941, Rev. 0

Plutonium and Absorber Concentrations in Fast and Low Settling Fraction


Anlayte Fe (g/g) Cd (g/g) Cr (g/g) Mn (g/g) Ni (g/g) Pu-239 (g/g) Pu:Fe Ratio Pu:Cd Ratio Pu:Cr Ratio Pu:Mn Ratio Pu:Ni Ratio Slow Settling Fraction 41000 662 31400 11,200 913 456 0.011 0.689 0.015 0.041 0.499 Rapid Settling Fraction 38500 514 23700 7900 699 282 0.007 0.549 0.012 0.036 0.403

Official Use Only D-2

RPP-RPT-50941, Rev. 0

APPENDIX E INDEPENDENT REVIEW TEAM CHARTER

Official Use Only E-1

RPP-RPT-50941, Rev. 0

1.0

PURPOSE

The Independent Review Group (IRG) for the Plutonium Oxide Evaluation Project is formed as an initiative to provide oversight and review of the results associated with the plutonium mixing/settling suspected issues and subsequent recommendations. The IRG should be involved throughout the last phase of the activities to help review and validate analytical data provided to the project. This data will form the basis for conclusions and recommendations that will be documented in a final report. Additionally, the IRG will provide technical review of all documents generated as to validity and content. The IRG will provide needed comments and suggestions as required in order to ensure that all conclusions and recommendations generated are consistent and defensible. Every effort shall be made to ensure the same IRG members stay on the project.

2.0

OBJECTIVES

Participate in group meetings to evaluate validity of analytical data and provide technical recommendations and conclusions based on the data. Review all draft documents and provide recommendations to enhance the technical validity of all conclusions. Solicit feedback from peers and the project team regarding the conclusions. Provide input to the data review process and the document preparation. The IRG will have the responsibility to provide review of the available data, question the team on the data as to its merit and completeness, and provide suggestions on best usage and conclusions based on available data. The team will also provide appropriate text for incorporation into the draft report. They will provide input and suggestions for comment resolution as needed. The team will also provide an independent review of the draft report. They will review for technical merit, validity of recommendations and conclusions, completeness of data analysis, and overall thoroughness of the report to answer the initial questions.

3.0

MEMBERSHIP

Membership on the IRG is provided via either a subcontract or a ROS with the URS Corporation. Adequate funding will be provided to ensure adequate time and resources are provided for IRG success. The IRG will have a Team Lead as designated by the Chief Engineer. The IRG is comprised of the following personnel: Dr. Raymond G. Wymer Frank R. McCoy, III, Team Lead Michael T. Ryan, Ph. D., C.H.P. Dr. Fred Beranek

Official Use Only E-2

RPP-RPT-50941, Rev. 0

4.0

RULES OF PRACTICE

The protocol for review of data and documents shall be: IRG Team Lead shall review the IRG Roles and Responsibilities with IRG members prior to kick off agenda. Project Manager performs kick off agenda and expectations Project Manager introduces the project team IRG members shall be allowed to participate freely and ask questions necessary to understand the issues and the data IRG members shall discuss the data and provide any recommendations or suggestions directly to the project Manager and the IRG Team Lead Project Managers shall incorporate the recommendations as appropriate. IRG members will assist in preparation of senior management presentation The IRG shall review the guidance provided in ASME NQA-1-2004, Nonmandatory Appendix 3.1, Guidance on Qualification of Existing Data. Sections such as 402, 404, and 500 will be evaluated for relevance and potentially provide the foundation for evaluating the data for validity and for documenting the data in the final report.

5.0

DELIVERABLES AND SCHEDULE

The IRG team will provide the following deliverables: Review draft sections of the report starting August 11-12. Continue review of draft sections as they are completed. Provide assistance in compilation of final report during last two weeks of August. Provide final review and comment on report August 26. Review final report for understanding of conclusions. August 11-22. Assist team in preparation of senior management presentation. August 22-26. Provide written letter to WRPS/WTP senior management that they have reviewed the final report and presentation and understand and agree with the conclusions and recommendations. August 29. Participate in senior management presentation week of August 29.

Official Use Only E-3

RPP-RPT-50941, Rev. 0

APPENDIX F PLUTONIUM INDEPENDENT REVIEW GROUP REPORT

Official Use Only F-1

RPP-RPT-50941, Rev. 0

Official Use Only F-2

RPP-RPT-50941, Rev. 0

Official Use Only F-3

RPP-RPT-50941, Rev. 0

Official Use Only F-4

RPP-RPT-50941, Rev. 0

Official Use Only F-5

RPP-RPT-50941, Rev. 0

Official Use Only F-6

RPP-RPT-50941, Rev. 0

Official Use Only F-7

RPP-RPT-50941, Rev. 0

Official Use Only F-8

RPP-RPT-50941, Rev. 0

Official Use Only F-9

RPP-RPT-50941, Rev. 0

Official Use Only F-10

RPP-RPT-50941, Rev. 0

Official Use Only F-11

RPP-RPT-50941, Rev. 0

Official Use Only F-12

RPP-RPT-50941, Rev. 0

Official Use Only F-13

RPP-RPT-50941, Rev. 0

Official Use Only F-14

RPP-RPT-50941, Rev. 0

Official Use Only F-15

RPP-RPT-50941, Rev. 0

Official Use Only F-16

RPP-RPT-50941, Rev. 0

Official Use Only F-17

RPP-RPT-50941, Rev. 0

APPENDIX G TEAM BIOGRAPHIES

Official Use Only G-1

RPP-RPT-50941, Rev. 0

David Bowers has more than 32 years of operations and engineering experience in the Hanford Tank Farms and related facilities. Relevant to this report he was an operations shift manager from 1974 to 1978 with direct supervisory responsibility for the 242-S and 242-T Evaporators and related supporting operations to include transfer of waste between tanks and facilities. From 1979 to 1983 he was the lead process engineer for operation of 242-S, 242-T, and West Area tank farms waste transfer operations and for the startup and operation of the 244-TX DCRT and the TX farm salt well system. The next 12 years were spent as a staff engineer to operations management resolving problems and providing project support followed by another 12 years in process engineering as a cognizant and project engineer for tank farm projects. David graduated from Oregon State University in 1974 with a B.S. G.S. interdisciplinary degree combining the graduate program in radiation science with the undergraduate nuclear engineering and technology program. Cal Delegard has 31 years experience in applied/process chemistry and nuclear materials at Hanford with experience in environmental chemistry of radioactive elements, plutonium process, waste, and storage chemistry, uranium, iron, copper, and zirconium corrosion, and technical liaison for projects to study transuranium element and technetium chemistry in alkaline media at the Institute of Physical Chemistry of the Russian Academy of Sciences. Cal was twice appointed to the International Atomic Energy Agency in Vienna where for eight years he designed and led nuclear material safeguards inspections at reactors, fuel conversion and fabrication plants, and reprocessing plants. He also helped develop methods for destructive and nondestructive nuclear materials verification. His publications include many reports for Hanford contractors and Pacific Northwest, Los Alamos, and Brookhaven National Laboratories; Journal of Nuclear Materials Management; Radiochimica Acta; Nuclear and Chemical Waste Management; Journal of Alloys and Compounds; Radiochemistry; numerous symposium proceedings; and contributions to book chapters. Cal received his B.S. in chemistry from Washington State University in 1970. Dr. Daniel L. Herting has a Ph.D. in Inorganic Chemistry from Oregon State University and has been a process chemist at Hanford for 32 years. During that time he has provided chemical and physical characterization of tank farm wastes, support to 242-S and 242-A evaporator operations, development of the TRUEX solvent extraction process and related pretreatment activities, support to various PUREX process development and troubleshooting activities, development of various saltcake and sludge retrieval methods, and has served for the past 2 years as Chief Chemist for the Waste Treatment and Immobilization Plant (WTP). Richard Hoyt has PhD and MS Degrees in Nuclear Engineering and a BS Degree in Chemical Engineering from Iowa State University. He has over 30 years experience in the nuclear field and has been involved with plutonium processing activities at Hanford for 28 years. He acted as technical lead in numerous plutonium processing facility pre-conceptual design studies and as Hanford Site technical representative in various DOE plutonium technical exchange working groups. He has participated in and chaired plutonium processing reviews at numerous DOE sites. During his tenure at Hanford he has directed plutonium precipitation development activities, established himself as a knowledgeable expert in the control of plutonium oxalate precipitation, plutonium oxide particle size distributions, and particle structures. He also has

Official Use Only G-2

RPP-RPT-50941, Rev. 0

documented substantial overviews of historical PFP process operations involving plutonium production and scrap recovery. He currently is active in providing plutonium processing consulting services. Walter Isom has a Bachelor of Science Degree in Mechanical Engineering and an MBA. He has over 30 years experience at SRS in Operations, Maintenance, Design Authority, Project and System engineering. During his tenure at SRS he has supported numerous facilities and processes including Savannah River National Lab, Plutonium/Uranium separations and metal processing, Naval Fuel fabrication, Salt Waste processing, Sludge Waste processing and Waste Retrieval. Tom Jones holds a PhD in chemistry and has worked at the Hanford site for over 30-years on many of the major Hanford programs. He was involved with the characterization of deep basalt groundwater systems in the BWIP Project, characterization of Hanford underground waste tanks core samples, supported development of the Best-Basis Inventory (BBI) inventory estimates for Hanford underground tanks, developed inventory estimates for Hanford single-shell tank leaks, and developed a comprehensive description of liquid discharges from Hanfords plutonium finishing operations from 1949 through 1973. David Lini has a Ph.D in Chemistry and an MBA. He has over 30 years nuclear related experience at Hanford in plutonium processing at PUREX, PFP, and the Plutonium Technology Laboratory, Program Management, and Strategic Studies. During his tenure he has supported numerous projects including BWIP, SIS, Uranium sales and disposition, Nuclear Materials Management, and the disposition of excess unirradiated plutonium. Jacob Reynolds has Bachelors and Masters Degrees in chemistry. Jacob has 11 years experience modeling and evaluating Hanford waste characterization data for a variety of engineering applications. This experience includes authoring more than 50 reports, calculations, or evaluations on waste process engineering or process chemistry. He has published more than 20 papers in journals or conference proceedings. His current responsibilities involve managing chemical data and waste inventories for the high-level waste tanks stored at the Hanford site. Terry Sams has over thirty three years of experience in the DOE complex. Mr. Sams is currently the Manager for Process Engineering Analysis for Washington River Protection Services. His prior assignment at Hanford tank farms was to build a team to develop and implement retrieval technologies, closure technologies, and develop a regulatory strategy for tank closures. He built the team and was instrumental in the startup of the next tank retrievals, including C-106 and the C-200 tanks. He leveraged technical ingenuity to lead research and development efforts seeking creative technological solutions to radioactive waste challenges throughout his career in the nuclear field. Prior to moving to Hanford, Mr. Sams was located at Oak Ridge where he led a team of engineers, cost estimators, and schedulers in the development of the technolgy processess for the U-233 disposition project. He designed numerous flow sheets and diagrams for the dissolution and separation of U-233 and Th-228. Mr. Sams also spent twenty years working for Lockheed Martin Energy Systems in Oak Ridge as a Senior Project Manager or as a Project Engineer. He managed large teams of engineers, scientists, and

Official Use Only G-3

RPP-RPT-50941, Rev. 0

craftsmen involved in complex remediation projects. He performed cost estimation for life cycle analysis, including material handling, transportation, treatment, and disposal for numerous radioactive and hazardous waste streams. As a DOE technical consultant, Mr. Sams evaluated new and innovative treatment technologies. He performed plant design and equipment selection for numerous waste treatment plants across the DOE complex. During the first ten years of this tenure, Mr. Sams was performing research in the development of waste treatment technologies, bio-remediation technologies, and direct coal liquefaction processes. Mr. Sams is a Chemical Engineer from the University of Tennessee where he earned both a BS and an MS degree. He is also a certified Project Manager Professional. He has over 70 publications in the areas of technology development, tank closure and retrieval, waste treatment, energy research, and others. Joseph Teal has been employed at Hanford for 52 years, including 18 years at the Plutonium Finishing Plant and several years at Tank Farms. Ted Venetz has a BS in Chemical Engineering from Montana State University and Masters Degree in Engineering and Technology Management from Washington State University. He has over 30 years experience at Hanford primarily in plutonium processing, stabilization, storage and transportation. Ted is a member of the DOE Materials Identification and Surveillance Working Group (MIS-WG) and has been a technical contributor to the DOE-STD-3013 since 1996. He has provided technical assistance to the DOE on various initiatives related to improving integration, nuclear materials management and consolidation, and plutonium storage, surveillance and disposition issues. Ted is currently a principal engineer with Washington River Protection Solutions and responsible for the Single Shell Tank Integrity Project. Bob Watrous comes to the Second Review Team on Plutonium Particulates Transferred to Tank Farms via E2 Consulting Engineers and brings with him a 43-year background in Hanford fuel reprocessing and waste management technology. He has experience in flowsheet design and R & D for both the REDOX and PUREX fuel separation processes, with specialties in criticality control for fuel dissolution operations. He helped bring computer process control online in the weapons fabrication portion of the 234-5Z plant. For two years Bob served on the Combined Operational Planning team (AECOP), operated by ORNL for the AEC. Since then he has worked to clean up his waste footprint by devoting many years in early phases of high-level waste management process development and waste vitrification technology.

Official Use Only G-4

RPP-RPT-50941, Rev. 0

APPENDIX H REVIEW OF SOLIDS IN PRF EFFLLUENT

Official Use Only H-1

RPP-RPT-50941, Rev. 0

Table A. Review of Solids in PRF Effluent


A review of solids in PRF effluent, the need for centrifuge and centrifuge operability for the Westfalia and Western States Centrifuges follows, based on review of biweekly and monthly reports: Month 4/12/1967 4/24/1967 Comment Special Study on PRF MUF started, 20% of Pu from MT dissolver is not being measured. Interface crud, CAW and CCW to be studied with both methods. Study for causes of PRF MUF continues. Actual Pu in CCW (to cribs) is 2.5 times that reported. Test on waste focused on polymer. Particle size analysis in CAW shows 0.5 to 12 micron particles with average of 3.5. Centrifuge located for installation in CAW stream. MUF study should focus on 1) Can solids in CCW be removed by centrifugation, 2) what is the composition of solids, 3) positive methods of analysis of these solutions. Centrifuge tests on CCW reported, specific gravity of solids = 3.1. Insufficient solids available for X ray diffraction. Development work on solids analysis continues. A.H. Case reports new methods of analysis of waste solutions to detect unmeasured losses. Tests planned to check his method. Five CCW samples analyzed by the normal method, samples centrifuged and only one shows solids. A.H. Case has not reported values from new method. Analysis of CCW samples by A.H. Case shows Pu content was 10X higher than that reported with old method. Meeting was held to address uncertainties in analysis and verification program described. Confirmed that existing method will not measure all Pu in CWW stream. Based on Process Control Lab work, a new method will be proposed The method for CCW is now the PuA-E6-B methods versus the old PuA-20-A. Pu loss to the cribs will be factor of 3 more than previously reported. CCW analysis now running ~ 0.013 gm/liter using new method, compared to old method values of 0.001 to 0.003 gm/liter. Centrifuge to be used to remove solids from CAW has arrived and is being tested on cold solutions (Mistron and water). CAW Centrifuge began process testing October 26. CA column interface filter installed and performing satisfactory for 2 days, 25 micron filter being used. Pu in first filter change out 4.8 gms. 40 micron filter to be tested to improve filtration rates. CAW centrifuge being installed and testing expected to begin CA column interface filter not used, vacuum cannot be obtained CAW centrifuge (Centrico Model SAOOH-2005) started on 10/24/1967. Unit automatically de-sludges into a collection tank. Some solids settle in 1-2 hours, other gave no indication of settling. Pu in initial samples is 6, 3, 2, and 3 grams. Ref 1 1

5/15/1967

5/29/1967

6/12/1967 6/26/1967 7/10/1967

1 1 1

7/24/1967 8/7/1967 8/21/1967

1 1 1

October 1967 10/2/1967 10/16/1967

2 4 4

10/30/1967

Official Use Only H-2

RPP-RPT-50941, Rev. 0

Table A. Review of Solids in PRF Effluent


A review of solids in PRF effluent, the need for centrifuge and centrifuge operability for the Westfalia and Western States Centrifuges follows, based on review of biweekly and monthly reports: Month 11/13/1967 Comment CAW losses are 3 times normal, undissolved fine solids are probable cause. The CAW centrifuge has been idle most of the time, due to operational and mechanical problems. Solids not completely removed during de-sludging and bowl will not re-seat. Fine grey-black solids dispersed through internal and corrosion noted. CAW centrifuge operating erratically, manufacture consulted, gaskets changed, bowl not re-seating consistently. Lab test on solids shows about 3 wt% Pu in dried solids. 75% of Pu can be leached in 57% HNO3 in 2 hours. CAW centrifuge unexpected started operating properly, believed from running with no feed. Unit operated 99 hours and removed 40 gms of Pu (0.4 gm/hr) Would not reseat on 12/5 and centrifuge dismantled after 125 hours, solids plugging found throughout internals including clarified solution outlet. Unit cleaned and returned to operation for the rest of week, removing about 0.6 gms/week. CAW centrifuge down since 12/15 with periodic operation from 12/13. Operated about 135 hours. Long discussion of re-seating and operational problems, with margin note IT DONT WORK CAW centrifuge operated 84 hours the first week and 31 hours the second until the piston failed to re-seat again. Routine maintenance and all the old tricks failed to reseat the piston. On 1/28/1968 a factory installed shim was installed and piston re-seated on startup. Unit on line the rest of the week, about 100 hours Large discussion on centrifuge. In operation 400 hrs, preliminary problems fixed. Collecting 0.6 gms of Pu per hr, sludge averages 3-5 wt% Pu. 40 to 60% of the plutonium can be leached by refluxing 2 hrs in 15.8M HNO3/0.6 M HF or 16 M HNO3. CAW centrifuge operated satisfactorily (110 hours) without re-seating problems. Unit shutdown due to leaks on gage. After shim installation unit has operated 210 hours without re-seating problems, previous record was 135 hours. Removal rate is 0.6 gms/hr CAW centrifuge operated about 200 hours without re-seating or gasket problems. Removal rate is 0.67 gms/hr Operated 460 hours, shutdown for cleaning and inspection. Solids clinging to internals. Back in service and only ran 4 days, corrosion and seal problems On 2/27 unit shutdown for inspection. Unit had operated 460 hours without problems. Solids plugging internal through-out, 2 quarts of solids removed, containing 11 gms Pu. Unit cleaned, re-gasketed and returned to use on 2/29, operation excellent until 3/4 when piston would not re-seat. Multiple mechanical problems noted, peeling coating, worn, nut, and bad seal. Ref 4

11/27/1967

12/11/1967

1/2/1968

1/15/1968

1/29/1968 February 1968

5 2

2/13/1968

2/26/1968 March 1968 ~3/14/1968

5 2 5

Official Use Only H-3

RPP-RPT-50941, Rev. 0

Table A. Review of Solids in PRF Effluent


A review of solids in PRF effluent, the need for centrifuge and centrifuge operability for the Westfalia and Western States Centrifuges follows, based on review of biweekly and monthly reports: Month 5/20/1968 Comment CAW flow rate dropping and SX shutdown. Tk-39-1 filter found to contain only a screen. Tk-39-2 filter intact but discharge plugged with golf-ball size ball solids. Lab tests show solids believed to be PuF3, as they dissolved readily and complete in warm ANN. Laboratory extraction tests on CXF showed 15 times more Pu in product than in feed. After filtration with 40 micron filter (readily apparent black solids), the concentration factors were equal. First mention of CAW centrifuge since 3/14. Unit likely not operated since Pu/U partitioning was being tested, and ash not being processed. Maintenance and additional suppliers information allowed for 125 hours of operation this period. Periodic re-seating problems noted and additional recommendations are made. Centrifuge operated 200 hours this period, one day piston would not re-seat. There were of 320 total centrifuge operating hours in June. Centrifuge operated 100 hours and was not good. Bowl dropped out twice causing CAW to spew out, gage fell off, and excessive vibration noted. Bearings are to be replaced. CAW operated 2 days and was valved out due to bearing noise. Centrifuge cleaned out and bearing replaced. Unit operated 100 hours or about 50% of the time, due to problems re-seating bowl and lack of necessity because ash was not being processed. Centrifuge operated 41% of column time Centrifuge operated 22.5% of column time Centrifuge operated 41% of column time Centrifuge operated 44% of column time. CAW overflow line from TK 32A found plugged with fine solids. Centrifuge operated 30% of column time Centrifuge operated 38% of column time Reports on re-burned ash show smaller particles in solution. Centrifuge operated about 64% (100 hours) of column time. Removal rates were 0.52 gms Pu/hr. A newly installed CA column crud filter is being tested. Centrifuge operated 100% of column time. Removal rate was 0.43 gms/hr, lower than previous periods even though more ash was processed. CAW overflow line to TK 32A plugged with solids again. Centrifuge operated 62% of column time. Ref 5

6/17/68

7/1/1968 7/15/1968

5 5

7/29/1968 8/12/1968

5 5

8/26/1968 9/12/1968 9/23/1968 10/7/1968 10/21/1968 11/4/1968 11/18/1968 12/8/1968 12/16/1968

5 5 5 5 5 5 5 5 5

12/30/1968

Official Use Only H-4

RPP-RPT-50941, Rev. 0

Table A. Review of Solids in PRF Effluent


A review of solids in PRF effluent, the need for centrifuge and centrifuge operability for the Westfalia and Western States Centrifuges follows, based on review of biweekly and monthly reports: Month 1/13/1969 Comment Centrifuge operated 30% of column time, removing about 0.22 gms/hr. Bowl reseating problems continue. New technique for sludge processing by soaking into rags, shredding and burning in the incinerator. Solids flushed from both columns. New CA column crud filter started on 1/23/1969. After a few filtration pressure drop increases and solids flow stopped. About one pint of grey-black solids removed. Centrifuge operated 0% of column time CA column crud filter was leaking and is used infrequently due to long filtration times. CA column interface crud removal system operating (25 micron sock filter), collecting 1-2 pints per week of solids with 10-15 grams Pu. 40 micron filter being tested as backup to centrifuge for removal of CAW solids. Pu solids detected in Waste Treatment facility tanks W-3 and W-4. Up to 1200 grams estimated. Will not dissolve in nitric-HF or HCl. Detected by neutron and gamma monitoring. CAW centrifuge was not operated. Data from burning and leaching sludge is presented. CAW centrifuge operate 95% of column time and removed 0.21 gms/hr CA crud filter continues to function acceptably. Waste Treatment by-passed due to Pu buildup in Tanks W-2, W-3 and W-4. 46 quarts of solids, containing 800 grams of Pu removed from W-3 and W-4, attributed to difficult to dissolve solids. Additional neutron monitoring to be installed. Discussion of Barrett clarifuge testing, 95% of solids removed by 40 micron filter. CA crud filter continues to function acceptably, collecting 5-10 gms of Pu/pint of solids. CAW centrifuge operated 70% of column time with no reports of collected sludge. The piston would not re-seat and maintenance reported large amounts of flaking and peeling on internal components. Hole being cut in W-4 in Waste Treatment The interface crud filter continues to operate The CAW centrifuge operated 100% of column time, (198 hours), collecting 0.07 gms Pu/hr. 44 quarts of solids removed from W-3 and W-4, estimated at 800 gms Pu The interface crud filter continues to operate, 33 gms Pu was removed in two batches of solids. The CAW centrifuge operated 96% of column time, (183 hours), collecting 0.19 gms Pu/hr Ref 6

2/3/1969

2/17/1969 March 1969

6 2

3/3/1969 3/17/1969

6 6

April 1969

4/1/1969

4/14/1969

4/28/1969

Official Use Only H-5

RPP-RPT-50941, Rev. 0

Table A. Review of Solids in PRF Effluent


A review of solids in PRF effluent, the need for centrifuge and centrifuge operability for the Westfalia and Western States Centrifuges follows, based on review of biweekly and monthly reports: Month 5/19/1969 Comment The interface crud filter continues to operate, 72 gms of Pu was collected in residue. The CAW centrifuge operated 99% of column time, (325hours), collecting 0.077 gms Pu/hr. On 5/6 the centrifuge was shutdown for cleaning and maintenance, only a handful of solids was found and the unit had operated 477 hours without major problems. Some improvement is attributed to longer processing ash. Extensive discussion of cleanout and processing of W-3 and W-4 Corrects prior CAW sludge removal rate to 0.26 gms Pu/hr. CA column crud filter slurry pump failed and was replaced CAW centrifuge operated 92% of column time (169 hours) and average removal was 0.44 gms Pu/hr. Stripping of W-2 in waste treatment continues, so far 1673.6 gms of Pu have been stripped from W-2. The CAW centrifuge operated 92% of column time, (200 hours), collecting 0.16 gms Pu/hr. On 6/10 it was shutdown because the piston failed to seat after de-sludging. This is the first malfunction after about 1100 hours of operation. About 3 pints of solids build-up with 71 gms Pu was removed. This developed after 22 days of operation. Ash Processing resumed. The CAW centrifuge operated 100% of column time, (196 hours), collecting 0.2 gms Pu/hr, noted as normal even with ash processing resumed. The CAW centrifuge operated 99% of column time, (192 hours), collecting 0.42 gms Pu/hr. Operation was excellent except for periodic and unexpected bowl droppings. The CAW centrifuge operated 97% of column time, (210 hours), collecting 0.44 gms Pu/hr. On 7/22 the centrifuge was opened and inspected, very little solids (<0.5 liter) were found. CAW centrifuge operated 216 hours or 54.4% of the time. Accumulation rate was 0.32 gms Pu/hr. Bearing worn and centrifuge down 9/12 to 9/24. CAW centrifuge operated 377 hours or 82% of the time. Accumulation Rate 0.32 gms Pu/hr. New gamma counting procedure adopted for Pu assay of sludge. CA column waste stream Pu content increased to 96 mg/l during ash processing and was reduced to 16 mg/l after additional centrifugation. Centrifuge operated 435 hours or 92% of the time and total Pu recovery was 343 grams Centrifuge operated 457 hours or 99% of the time and total Pu recovery rate was 1.1 grams/hr. Ref 6

6/2/1969

2/16/1969

6/30/1969

7/14/1969

8/4/1969

Sept 1969 October 1969

2 2

November 1969

December 1969

Official Use Only H-6

RPP-RPT-50941, Rev. 0

Table A. Review of Solids in PRF Effluent


A review of solids in PRF effluent, the need for centrifuge and centrifuge operability for the Westfalia and Western States Centrifuges follows, based on review of biweekly and monthly reports: Month June 1970 November 1970 April 1971 May 1971 August 1971 Comment Problems with ash dissolving showed large heels and Pu content in CAW centrifuge sludge Recalibration of product receiver tanks showed large bias due to solids buildup Centrifuge can remove 90% of the heavy and light phase solids from ash dissolution, filters blind constantly Shortage across button line attributed to solids in PRF product A backflush filter was evaluated for solids removal from the 08 dissolver overflow. A 240 mesh and 5-10 micron filter were tested. Filters plug rapidly and require excessive backflushing Testing of prototype centrifuge for feed solution clarification continued Filters not effective in removing fines from oxide dissolution. Centrifuge planned for future installation Under-accounting in oxide dissolution of 20% reduced to 3% by reducing nitric/HF to reduce PuF4 formation, improved settle/decant, use of filter on settled solution and reduced charge size. Discharges from PRF are excessive and contributors were the CA column interface crud and high salt waste. Recommendations are made for reactivation of the dissolver solution filters, removal of the crud using the centrifuge, use of finer filters in waste treatment. Losses were reduced from 1082 gms in May to 503 gms in June by use of the MT dissolver filters and removal of the interface crud using the centrifuge. The PSHS installation is being planned, with the unit on slag and crucible dissolver installed and test are underway to produce a thicker sludge Studies on synthetic ash show over the 1 micron particles can be removed by the centrifuge It is desirable to increase the RPM of the Western States Centrifuge to improve separation of particles. The Vendor states this is not possible with the existing design. Testing of a automatic sludge removal dip leg was unsuccessful Installation of centrifuges in MT-6 for removal of trace solids was completed by testing indicates alterations are necessary. PFP high and low salt waste now discharged to Tank Farms The centrifuge performance for the PSHS system has been satisfactory with the exception of frequent bearing failure. Air purged bearing are being installed. All centrifuge installed in PHSH project are operable except the B1 and B2 unit installed on the PRF CAW stream. Ref 3 3 3 3 3

September 1971 October 1971 November 1971

3 3 2

May 1972

June 1972 September 1972 October 1972 November 1972

3 3 3 3

February 1973 February 1973 May 15, 1973 May 1973 May 1973

3 7 3, 8 3 6

Official Use Only H-7

RPP-RPT-50941, Rev. 0

Table A. Review of Solids in PRF Effluent


A review of solids in PRF effluent, the need for centrifuge and centrifuge operability for the Westfalia and Western States Centrifuges follows, based on review of biweekly and monthly reports: Month June 1973 Comment The five new centrifuges installed in PRF have recovered 940 grams in May and 704 grams in June. Solids from the final centrifuge (E Centrifuge on the EIW waste treatment stream) are characterized. Mechanical and Hydraulic problems continue with PHSH centrifuge, sludge from final waste stream collected and characterized to gage economics of future recovery efforts Centrifuge bearing life has doubled to 5 weeks, which is still unacceptable and additional changes are planned The Western States representative has visited the site and indicates the centrifuge outlet flow lines are too restricted and they are operating with the bowl submerged. Centrifuge bearing life is now at 11 weeks by better provision of purge air and reduction of out flow restrictions on 3 units. Other 2 unit still be studied. E centrifuge failed at 11 weeks, trying new bearings and grease. Plastic lids are failing and being replaced with metal. Request for redesign of the MT-6 centrifuge (B1, B2, and C units) to reduce the outlet flow restrictions. Ref 3

June 1973

July 1973 August 1973

3 3

September 1973 October 1973 December 1973

3 3 3

ISO-659 RD, Z Plant Report Task I-II Recovery Operations, 1967 ARH-423-RD-DEL, PPE Monthlies, 1968 ARH-423-RD-DEL, PPE Monthlies 19701972 ARH-11 RD, Z Plant Report Task I-II Recovery Operations, September 5, 1967 through December 31, 1967 ARH-294 RD, Z Plant Report Task I-II Recovery Operations, January 1, 1968 through December 31, 1968 ARH-1048 RD Z Plant Report Task I-II Recovery Operations, 1969 PWM-550-2-DEL, Nuclear Materials Management Branch Monthly Report for February 1972 (title wrong, really 1973) PWM-550-5-DEL, Nuclear Materials Management Branch Monthly Report for May 1973 PWM-550-6-DEL, Nuclear Materials Management Branch Monthly Report for June 1973

Official Use Only H-8

RPP-RPT-50941, Rev. 0

APPENDIX I TANK TABLES

Official Use Only I-1

RPP-RPT-50941, Rev. 0

Tank A-105

Best Estimate of Particulate Plutonium: 400 g


Sources
Type Pu oxide from metal burning Amount (gms) 400 Size 40-100 microns Density 8-11 g/cc

Assumptions and Margins Quantity based upon 1% entrainment similar to PFP since came as nitrate solution for solvent extraction only A 40 micron filter was used in PUREX to filter the material assume 20% not trapped (2000 g*0.2=400 g) most particles near 40 micron since source was 40-100 microns No material hold up in PUREX process tanks No particles were left in shipping container after vacuum removal

Estimate without Margin Typical filters are 95-99% efficient, could reduces losses to as low as 20 g (2000*0.01 =20)

Range: 20-400 g

Official Use Only I-2

RPP-RPT-50941, Rev. 0

Tank S-108 Best Estimate of Particulate Plutonium: 1kg Sources


Type Pu oxide from metal burning Amount (gms) 1000 Size 40-100 microns Density 8-11 g/cc

Assumptions and Margins Quantity based upon 1% entrainment similar to PFP since came as nitrate solution for solvent extraction only No filter was used in REDOX to filter the material assume 100% went to tank farm No material hold up in REDOX process tanks No particles were left in shipping container after vacuum removal

Estimate without Margin NA Range: NA

Official Use Only I-3

RPP-RPT-50941, Rev. 0

Tank C-104
Best Estimate of Particulate Plutonium: 40 g

Sources
Type Pu oxide from precipitate Amount (gms) 1640 Size 10-40 microns Density 8-11 g/cc

Assumptions and Margins Based upon 4.8% total losses, rounded to 5% for all campaigns Cladding waste losses were 4.2% and High Level Waste losses were 0.8% Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% A total of 1640 gms oxide particulate was discharged to Tank C-104 during PUREX operation but C-104 was transferred to AN-101 in 2010 with 98% of Ci moved. Therefore assumed ~2% of the Pu particulate remained in tank C-104 (1640*0.02=33g) No material hold up in PUREX process tanks No particles were left in shipping container after vacuum removal Effects of burnout are neglected

Estimate without Margin Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 18 g (40*0.45 =18)

Range: 18-40 g

Official Use Only I-4

RPP-RPT-50941, Rev. 0

Tank AN-101 Cladding waste Best Estimate of Particulate Plutonium: 1600 g Sources
Type Pu oxide from precipitate Amount (gms) 1640 Size 10-40 microns Density 8-11 g/cc

Assumptions and Margins Based upon 4.8% total losses, rounded to 5% for all campaigns Cladding waste losses were 4.2% and High Level Waste losses were 0.8% Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% A total of 1640 gms oxide particulate was discharged to Tank C-104 during PUREX operation but C-104 was transferred to AN-101 in 2010 with 98% of Ci moved. Therefore assumed 98% of the particulate transferred to tank AN-101 (1640*0.98=1607) No material hold up in PUREX process tanks No particles were left in shipping container after vacuum removal Effects of burnout are neglected

Estimate without Margin Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 720 g (1600*0.45 =720)

Range: 720-1600

Official Use Only I-5

RPP-RPT-50941, Rev. 0

Tank B-101 High Level waste Best Estimate of Particulate Plutonium: 320 g Sources
Type Pu oxide from precipitate Amount (gms) 320 Size 10-40 microns Density 8-11 g/cc

Assumptions and Margins Based upon 4.8% total losses, rounded to 5% for all campaigns Cladding waste losses were 4.2% and High Level Waste losses were 0.8% Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% No material hold up in PUREX process tanks No particles were left in shipping container after vacuum removal Effects of burnout are neglected

Estimate without Margin Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 144 g (320*0.45 =144)

Range: 144-320 g

Official Use Only I-6

RPP-RPT-50941, Rev. 0

Tank BX-101 Best Estimate of Particulate Plutonium: 150 g Sources


Type Pu oxide from precipitate Amount (gms) 150 Size 10-40 microns Density 8-11 g/cc

Assumptions and Margins Based upon 4.8% total losses, rounded to 5% for all campaigns Flush of system recovered 2.3% of 4.8% for 1 campaign so could be as low as 2.5% Cladding waste losses were 4.2% and High Level Waste losses were 0.8% Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% No material hold up in PUREX process tanks No particles were left in shipping container after vacuum removal Effects of burn-out are neglected

Estimate without Margin Reducing the losses 50% reduces the amount lost to 75 g (160*0.9*0.5 =72) Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 36 g (160*0.5*0.45 =36)

Range: 36-150 g

Official Use Only I-7

RPP-RPT-50941, Rev. 0

Tank C-102 Best Estimate of Particulate Plutonium: 770 g Sources


Type Pu oxide from precipitate Amount (gms) 770 Size 10-40 microns Density 8-11 g/cc

Assumptions and Margins Based upon 4.8% total losses, rounded to 5% for all campaigns Flush of system recovered 2.3% of 4.8% for 1 campaign so could be as low as 2.5% loses Cladding waste losses were 4.2% and High Level Waste losses were 0.8% Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% No material hold up in PUREX process tanks No particles were left in shipping container after vacuum removal Effects of burnout are neglected

Estimate without Margin Reducing the losses 50% reduces the amount lost to 383 g (850*0.9*0.5 = 383) Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to g (850*0.5*0.45 = 191)

Range: 191-770

Official Use Only I-8

RPP-RPT-50941, Rev. 0

Tank S-107 Best Estimate of Particulate Plutonium: 260 g Sources


Type Pu oxide from precipitate Amount (gms) 260 Size 10-40 microns Density 8-11 g/cc

Assumptions and Margins Based upon 4.8% total losses, rounded to 5% for all campaigns Cladding waste losses were 4.2% and High Level Waste losses were 0.8% Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% No material hold up in REDOX process tanks No particles were left in shipping container after vacuum removal Effects of burnout are neglected

Estimate without Margin Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 126 g (280*0.45 =126)

Range: 126-260 g

Official Use Only I-9

RPP-RPT-50941, Rev. 0

Tank TX-101 Best Estimate of Particulate Plutonium: 400 g Sources


Type Pu oxide from precipitate Amount (gms) 400 Size 10-40 microns Density 8-11 g/cc

Assumptions and Margins PFP losses were based upon 0.5% entrainment since goal was not to entrain solids Distribution of the waste stream is based upon process operations and feeds being processed apportioned by consensus of subject matter experts Particles greater than 100 micron did not leave facility Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% TX 101 received 448 g from the pump-out of Z-361 Catch tank, assumed to be oxide from calcined precipitate (448*0.9=403 g)

Estimate without Margin Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 202 g (448*0.45 = 202)

Range: 200-400 g

Official Use Only I-10

RPP-RPT-50941, Rev. 0

Tank S-111 High Level waste Best Estimate of Particulate Plutonium: 30 g Sources
Type Pu oxide from precipitate Amount (gms) 30 Size 10-40 microns Density 8-11 g/cc

Assumptions and Margins Based upon 4.8% total losses, rounded to 5% for all campaigns Cladding waste losses were 4.2% and High Level Waste losses were 0.8% Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% Split losses 50/50 between 2 tanks S-111 and SX-114 No material hold up in REDOX process tanks No particles were left in shipping container after vacuum removal Effects of burnout are neglected

Estimate without Margin Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 12 g (53*0.5*0.45 =12)

Range: 12-30

Official Use Only I-11

RPP-RPT-50941, Rev. 0

Tank SX-114 High Level Waste Best Estimate of Particulate Plutonium: 30 g Sources
Type Pu oxide from precipitate Amount (gms) 30 Size 10-40 microns Density 8-11 g/cc

Assumptions and Margins Based upon 4.8% total losses, rounded to 5% for all campaigns Cladding waste losses were 4.2% and High Level Waste losses were 0.8% Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% No material hold up in REDOX process tanks No particles were left in shipping container after vacuum removal Effects of burnout are neglected

Estimate without Margin Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 12 g (53*0.5*0.45 =12)

Range: 12-30

Official Use Only I-12

RPP-RPT-50941, Rev. 0

Tank TX-105 Best Estimate of particulate Plutonium: 2.37kg


Type Pu oxide from precipitation Pu oxide from metal burning Pu metal fines Amount (gms) 1100 1220 54 Size 10-40 microns 40-100 microns 40-100 microns Density 8-11 g/cc 8-11 g/cc 19 g/cc

Assumptions and Margins Unmeasured losses were assumed to be 100% of measured loses Distribution of the waste stream is based upon process operations and feeds being processed apportioned by consensus of subject matter experts Particles greater than 100 micron did not leave facility Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% All burnt metal was greater than 10 micron- estimates used 100% (section 4.10) Pu metal fines did not dissolve in tank farm and 100% greater than 10 micron TX 105 received discharges reported for all of 1973 and of 1974

Estimate without Margin TX118 was used during down times of the evaporator as the drop tank. Will assume 50% of material (proportional) could have gone to TX118 vice TX-105 (equals 1.19 kg) Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 550 g Pu metal fines dissolving in waste over last 20 years would reduce quantity to zero

Range: 0.86 to 2.37 kg

Official Use Only I-13

RPP-RPT-50941, Rev. 0

Tank TX-109 Best Estimate of particulate Plutonium: 4.14kg


Type Pu oxide from precipitation Pu oxide from metal burning Pu metal fines Amount (gms) 1925 2125 90 Size 10-40 microns 40-100 microns 40-100 microns Density 8-11 g/cc 8-11 g/cc 19 g/cc

Assumptions and Margins Unmeasured losses were assumed to be 100% of measured loses Distribution of the waste stream is based upon process operations and feeds being processed apportioned by consensus of subject matter experts Particles greater than 100 micron did not leave facility Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% All burnt metal was greater than 10 micron- estimates used 100% (section 4.10) Pu metal fines did not dissolve in tank farm and 100% greater than 10 micron TX 109 received discharges reported for 3/4 of 1974, all of 1975, and 1/3 of 1976 TX 109 received 130 gms from the pump-out of Z-8 Catch tank, assumed to be 50% oxide from metal burning and 50% oxide from calcined precipitate

Estimate without Margin TX118 was used during down times of the evaporator as the drop tank. Will assume 50% of material (proportional) could have gone to TX118 vice TX-109 (equals 2.07 kg) Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 963 g Pu metal fines dissolving in waste over last 20 years would reduce quantity to zero

Range: 1.55 to 4.14 kg

Official Use Only I-14

RPP-RPT-50941, Rev. 0

Tank TX-118 Best Estimate of particulate Plutonium: 2.72kg


Type Pu oxide from precipitation Pu oxide from metal burning Pu metal fines Amount (gms) 1500 1210 9 Size 10-40 microns 40-100 microns 40-100 microns Density 8-11 g/cc 8-11 g/cc 19 g/cc

Assumptions and Margins Unmeasured losses were assumed to be 100% of measured loses Distribution of the waste stream is based upon process operations and feeds being processed apportioned by consensus of subject matter experts Particles greater than 100 micron did not leave facility Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% All burnt metal was greater than 10 micron- estimates used 100% (section 4.10) Pu metal fines did not dissolve in tank farm and 100% greater than 10 micron TX 118 received discharges reported for 2/3 of 1976 and all of 1977, 1978, 1979

Estimate without Margin (upper and lower) TX118 was used during down times of the evaporator as the drop tank. Will assume 50% of material (proportional) could have gone to TX118 vice TX-105 and TX-109 adding 3.26 kg Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 750 g Pu metal fines dissolving in waste over last 20 years would reduce quantity to zero

Range: 1.96 to 5.98 kg

Official Use Only I-15

RPP-RPT-50941, Rev. 0

Tank SY-102 Best Estimate of particulate Plutonium: 12.65kg


Type Pu oxide from precipitation Pu oxide from metal burning Pu metal fines Amount (gms) 6700 3985 1965 Size 10-40 microns 40-100 microns 40-100 microns Density 8-11 g/cc 8-11 g/cc 19 g/cc

Assumptions and Margins Unmeasured losses were assumed to be 100% of measured loses Distribution of the waste stream is based upon process operations and feeds being processed apportioned by consensus of subject matter experts Particles greater than 100 micron did not leave facility Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% All burnt metal was greater than 10 micron- estimates used 100% (section 4.10) Pu metal fines did not dissolve in tank farm and 100% greater than 10 micron SY 102 received discharges reported for all of 1981-2004 Apportioned 1000 grams to tank 244-TX based upon BBI inventory of 108 Ci. BBI inventory would be 1370 grams if the ratio of Pu239/240 was 90/10. Assigned a proportional amount of each plutonium type remained.

Estimate without Margin Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 3350 g Pu metal fines dissolving in waste over last 20 years would reduce quantity to zero

Range: 7.34 to 12.65 kg

Official Use Only I-16

RPP-RPT-50941, Rev. 0

Tank 244-TX Horizontal tank before SY-102 Best Estimate of particulate Plutonium: 1000 g
Type Pu oxide from precipitation Pu oxide from metal burning Pu metal fines Amount (gms) 530 315 155 Size 10-40 microns 40-100 microns 40-100 microns Density 8-11 g/cc 8-11 g/cc 19 g/cc

Assumptions and Margins Unmeasured losses were assumed to be 100% of measured loses Distribution of the waste stream is based upon process operations and feeds being processed apportioned by consensus of subject matter experts Particles greater than 100 micron did not leave facility Calcined precipitant PSD range was 45-89% greater than 10 micron- estimates used 90% All burnt metal was greater than 10 micron- estimates used 100% (section 4.10) Pu metal fines did not dissolve in tank farm and 100% greater than 10 micron SY 102 received discharges reported for all of 1981-2004 Apportioned 1000 grams to tank 244-TX based upon BBI inventory of 108 Ci. BBI inventory would be 1370 grams if the ratio of Pu239/240 was 90/10. Assigned a proportional amount of each plutonium type remained

Estimate without Margin Using 45% PSD >10 microns for oxide from calcined precipitate, the amount would be reduced to 265 g Pu metal fines dissolving in waste over last 20 years would reduce quantity to zero

Range: 580 to 1000 g

Official Use Only I-17

RPP-RPT-50941, Rev. 0

APPENDIX J PU OXIDE-METAL DISCHARGE TO TANK FARM

Official Use Only J-1

RPP-RPT-50941, Rev. 0

Pu Oxide-Metal Discharge to Tank Farm


Year 1963/64 1963/64 1969 1969 1969 1969 1969 1970-72 1970-72 Facility PUREX REDOX PUREX PUREX REDOX REDOX REDOX PUREX PUREX Type PuO metal PuO metal Pu Oxide Pu Oxide Pu Oxide Pu Oxide Pu Oxide Pu Oxide Pu Oxide Size (microns) 40-100 40-100 10-40 10-40 10-40 10-40 10-40 10-40 10-40 Density (g/cc) 8-11 8-11 8-11 8-11 8-11 8-11 8-11 8-11 8-11 Amount (Kg) Receiving Tank 0.4 1 0.77 0.15 0.26 0.03 0.03 0.04 1.6 A-105 S-108 C-102 BX-101 S-107 S-111 SX-114 C-104 AN-101 Comments Based upon 1% losses and filtered to 40 micron with 20% passing filter Based on 1% losses and no filter Based on total 5% losses/4.2% in cladding and 0.8% HLW Based upon 0.8% losses via HLW Based on total 5% losses/4.2% in cladding and 0.8% HLW Based upon 0.8% losses via HLW / 50% in S-111 and 50% in SX-114 Based upon 0.8% losses via HLW Amount not transferred to AN-101 based upon volume Based on 5% losses/4.2% in cladding Transferred to AN-101 in 2010 Based upon 0.8% losses via HLW Based on 0.5% entrainment/goal was not to remove solids Based on 10% entrainment/goal was to remove solids Based on 10% entrainment/goal was to remove solids TX-105 primary bottoms receiver

1970-72 1975 1974 1974 1973 1973 1973 1974 1974 1974 1975 1975 1975 1976 1976 1976 1977 1978

PUREX 241-Z361 216-Z-8 216-Z-8 PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP

Pu Oxide Pu Oxide PuO metal Pu Oxide PuO precp PuO metal Metal fines PuO precp PuO metal Metal fines PuO precp PuO metal Metal fines PuO precp PuO metal Metal fines PuO precp PuO precp

10-40 10-40 40-100 10-40 10-40 40-100 40-100 10-40 40-100 40-100 10-40 40-100 40-100 10-40 40-100 40-100 10-40 10-40

8-11 8-11 8-11 8-11 8-11 8-11 19 8-11 8-11 19 8-11 8-11 19 8-11 8-11 19 8-11 8-11

0.32 0.4 0.07 0.07 0.79 0.88 0.04 1.24 1.37 0.06 0.83 0.92 0.04 0.28 0.32 0.01 0.4 0.01

B-101 TX-101 TX-109 TX-109 TX-105 TX-105 TX-105 TX-109 TX-109 TX-109 TX-109 TX-109 TX-109 TX-118 TX-118 TX-118 TX-118 TX-118

TX-109 made primary bottoms receiver 4/74

Evaporator shut down in 4/76. All waste to TX-118

Official Use Only J-2

RPP-RPT-50941, Rev. 0

Pu Oxide-Metal Discharge to Tank Farm


Year 1978 1979 1979 1982 Facility PFP PFP PFP PFP Type PuO metal PuO precp PuO metal PuO precp Size (microns) 40-100 10-40 40-100 10-40 Density (g/cc) 8-11 8-11 8-11 8-11 Amount (Kg) Receiving Tank 0.01 0.9 1 0.09 TX-118 TX-118 TX-118 244-TX/SY-102 244-TX BBI inventory is 108 Ci based upon a 90/10 split of 239 /240 equals 1370 g will assign 1000 g to 244-TX 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 244-TX/SY-102 No processing in 1991 244-TX/SY-102 244-TX/SY-102 1979 estimates are inclusive of some prior discharges Comments

1983 1984 1985 1985 1985 1986 1986 1986 1987 1987 1987 1988 1989 1990 1992 1993

PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP PFP

PuO precp PuO precp PuO precp PuO metal Metal fines PuO precp PuO metal Metal fines PuO precp PuO metal Metal fines PuO precp PuO precp PuO precp PuO precp PuO precp

10-40 10-40 10-40 40-100 40-100 10-40 40-100 40-100 10-40 40-100 40-100 10-40 10-40 10-40 10-40 10-40

8-11 8-11 8-11 8-11 19 8-11 8-11 19 8-11 8-11 19 8-11 8-11 8-11 8-11 8-11

0.09 3.47 1.55 1.7 0.7 0.4 1.8 1.1 0.72 0.8 0.32 0.32 0.05 0.14 0.04 0.36

Official Use Only J-3

You might also like