You are on page 1of 27

CNWRA 2005-003

MICROSTRUCTURAL ANALYSES AND MECHANICAL


PROPERTIES OF ALLOY 22

Prepared for

U.S. Nuclear Regulatory Commission


Contract NRC–02–02–012

Center for Nuclear Waste Regulatory Analyses


San Antonio, Texas
cnwra-2005-003

MICROSTRUCTURAL ANALYSES AND MECHANICAL


PROPERTIES OF ALLOY 22

Prepared for

U.S. Nuclear Regulatory Commission


Contract NRC–02–02–012

Prepared by

D.S. Dunn
Yi-Ming Pan
K.T. Chiang

Center for Nuclear Waste Regulatory Analyses


San Antonio, Texas

March 2005
ABSTRACT
Fabrication of the Alloy 22 waste package outer container will require multiple processes,
including welding and solution annealing. Effects of these thermally influenced fabrication
processes on the microstructure, impact strength, ductility, and fracture toughness were
evaluated using mill-annealed and welded Alloy 22. In the mill-annealed condition, Alloy 22
has high ductility and high fracture toughness. The precipitation of topologically close-packed
phases is known to reduce the ductility and impact strength of Alloy 22. The mechanical
properties of the mill-annealed material were influenced by the amount of topologically
close-packed phases at the grain boundaries. A decrease in the impact strength was observed
for the mill-annealed material after aging for 10 hours at 870 °C [1,598 °F]. The ductility and
fracture toughness of the mill-annealed material was significantly reduced after thermal aging
for 100 hours at 870 °C [1,598 °F]. Both gas-tungsten arc welded and gas-metal arc welded
Alloy 22 contain topologically close-packed phases in the as-welded and thermally aged
conditions. The welded material had higher yield strengths and lower impact strengths
compared with the mill-annealed material. Thermal aging of welded material for 1 hour at
870 °C [1,598 °F] did not result in a significant decrease in impact strength or fracture
toughness. A decrease in impact strength was observed after thermal aging for greater than
1 hour. Solution annealed welds had a larger fraction of topologically close-packed phase
precipitates, however, both the impact strength and fracture toughness were similar to the
as-welded material. From the results of this investigation, high fracture toughness and impact
strength are expected for Alloy 22 in the as-welded and welded + solution annealed conditions.
The ductility, fracture toughness, and impact strength of Alloy 22 are unlikely to be significantly
affected by the fabrication processes necessary to construct the waste package outer container.

iv
CONTENTS
Section Page

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix
ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-1

1.1 Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-2


1.2 Scope and Organization of the Report . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-2
1.3 Relevant DOE and NRC Agreements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1-3

2 MATERIALS AND METHODS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-1

2.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-1


2.2 Microstructural Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-2
2.3 Charpy V-Notch Impact Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-2
2.4 Fracture Toughness Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-3

3 METALLURGICAL ANALYSIS AND MECHANICAL PROPERTIES . . . . . . . . . . . . . 3-1

3.1 Metallurgical Analyses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-1


3.2 Charpy V-Notch Impact Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-4
3.3 Tensile Properties and Fracture Toughness . . . . . . . . . . . . . . . . . . . . . . . . . . 3-5

4 ASSESSMENT OF FABRICATION PROCESSES . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-1

5 CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5-1

6 REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-1

vi
FIGURES
Figure Page

3-1 Photographs Showing the Fusion Zones of the Gas-Tungsten Arc Welded (GTAW)
and Gas Metal Arc Welded (GMAW) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-2
3-2 Micrographs Showing Amount of Precipitates in Gas-Tungsten Arc Welded
(Left) and Gas-Metal Arc Welded (Right) Alloy 22 Welds in the As-Welded . . . . . . . 3-3
3-3 Charpy V-Notch Impact Test Results for Alloy 22 . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-5

viii
TABLES
Table Page

1-1 DOE and NRC Agreements Related to This Report . . . . . . . . . . . . . . . . . . . . . . . . . . 1-3

2-1 Composition of Alloy 22 Heats and ERNiCrMo-10 Filler Metal . . . . . . . . . . . . . . . . . . 2-1

3-1 Volume Fraction of Precipitates Measured on Thermally Aged Alloy 22 . . . . . . . . . . 3-1


3-2 Volume Fraction of Precipitates Measured on Welded and Welded + Heat(GTAW)
and Gas-Metal Arc Welded (GMAW) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-2
3-3 Room Temperature Tensile Properties of Alloy 22 Aged . . . . . . . . . . . . . . . . . . . . . . 3-6
3-4 Room Temperature Fracture Toughness of Alloy 22 Thermally Aged . . . . . . . . . . . . 3-7
3-5 Room Temperature Tensile Properties of Welded and Heat-Treated Alloy 22 . . . . . 3-8
3-6 Fracture Toughness Values of Welded and Welded + Heat Treated Alloy 22 . . . . . . 3-9

x
ACKNOWLEDGMENTS
This report was prepared to document work performed by the Center for Nuclear Waste Regulatory
Analyses (CNWRA) for the U.S. Nuclear Regulatory Commission (NRC) under Contract
No. NRC–02–02–012. The activities reported here were performed on behalf of the NRC Office
of Nuclear Material Safety and Safeguards, Division of High-Level Waste Repository Safety. The
report is an independent product of CNWRA and does not necessarily reflect the views or
regulatory position of NRC.

The authors gratefully acknowledge G. Cragnolino for the technical review, B. Sagar for
programmatic review, and C. Cudd and B. Long for the editorial reviews. Appreciation is due
A. Ramos for assistance in preparing this report.

QUALITY OF DATA, ANALYSES, AND CODE DEVELOPMENT

QUALITY OF DATA: Sources of data are referenced in each chapter. CNWRA-generated


laboratory data contained in this report meet quality assurance requirements described in the
CNWRA Quality Assurance Manual. Data from other sources, however, are freely used. The
respective sources of non-CNWRA data should be consulted for determining levels of quality
assurance. Experimental data have been recorded in CNWRA scientific notebook numbers 498,
503, 579, 636, and 641.

ANALYSES AND CODES: No scientific or engineering software was used in the generation of this
report.

xii
1 INTRODUCTION
Performance of the engineered barrier system during preclosure operations and after waste
emplacement is important to protect the public from any undue long-term risk, as recognized by the
U.S. Department of Energy (DOE) in its repository safety strategy for the potential repository at
Yucca Mountain (CRWMS M&O, 2000a). As an independent regulatory agency, the U.S. Nuclear
Regulatory Commission (NRC) has published licensing criteria for the disposal of high-level wastes
in the potential repository. According to 10 CFR 63.112, a preclosure safety analysis must
accompany a license application for construction authorization of a geologic repository. A
preclosure safety analysis is required to demonstrate the safety of the proposed design and
operations in the geologic repository operations area with regard to the overall preclosure
performance objectives through a systematic examination of the site; the design; and the potential
hazards, initiating events, and their resulting event sequences and potential radiological exposures
to workers and the public. This analysis should include a general description of the structures,
systems, components, equipment, and process activities at the geologic repository operations area.
In addition, 10 CFR 63.21(c)(3) requires the safety analysis report filed with the license application
must include a description and discussion of the designs of the various components of the geologic
repository operations area and the engineered barrier system including (i) the dimensions, material
properties, specifications, and analytical and design methods used along with any applicable codes
and standards; (ii) the design criteria used and their relationships to the preclosure performance
objectives specified in 10 CFR 63.111(b), 63.113(b), and 63.113(c); and (iii) the design bases and
their relation to the design criteria. In the postclosure period, 10 CFR 63.311 requires the
engineered barrier system to be designed so that, working in combination with natural barriers,
radiological exposures to the reasonably maximally exposed individual and release of radionuclides
into the accessible environment are limited.

For the preclosure period, the waste package is identified by DOE as the structure, system, or
component that prevents release of radionuclides in the event of waste package drops, objects
striking the waste package, waste package collisions, and fire or thermal hazards (CRWMS M&O,
2000a). In addition, the performance of the waste package, one of the two main components of the
engineered barrier subsystem, is noted by DOE among the principal factors for the postclosure
safety case (CRWMS M&O, 2000a). The reference waste package design in the DOE site
recommendation (CRWMS M&O, 2000b) consists of an outer container made of Alloy 22, a highly
corrosion-resistant nickel-chromium-molybdenum alloy (Ni–22Cr–13Mo–3W–4Fe), and an inner
container made of Type 316 nuclear grade stainless steel (low C-high N–Fe–18Cr–12Ni–2.5Mo).
Prior to closure, an inverted U-shaped drip shield, fabricated from Titanium Grade 7 (Ti–0.15Pd),
and Titanium Grade 24 (Ti–6Al–4V–0.05Pd) will be extended over the length of the emplacement
drifts to enclose the top and sides of the waste packages. The drip shield is designed to provide
additional protection to the waste package from mechanical loads as a consequence of rockfall, in
addition to diverting seepage water entering the drifts.

Components of the engineered barrier system must be designed to accommodate mechanical loads
as a consequence of loading, transfer, and emplacement operations. The waste package also may
be subjected to impacts as a result of drops. Engineered barrier system components, including the
waste package also may be mechanically loaded as a result of seismic activity, rockfall, and drift
degradation. The response of the waste package and the drip shield to mechanical loading will be
dependent on the design and mechanical properties of the engineered barrier system components.
The American Society of Mechanical Engineers Boiler and Pressure Vessel Code (American
Society of Mechanical Engineers, 2001a) provides requirements for the design, fabrication, and

1-1
inspection of nuclear components to assure component integrity for range of expected operating
conditions. Although the applicability and use of the American Society of Mechanical Engineers
Boiler and Pressure Vessel Code for the construction of the waste packages has not been
established, DOE indicated the waste packages will conform to the requirement of the American
Society of Mechanical Engineers Boiler and Pressure Vessel Code where practicable (Bechtel
SAIC Company LLC, 2001). Fabrication processes, including cold work during machining and
forming, welding, and postweld stress mitigation procedures such as solution annealing, laser
peening, and low plasticity burnishing, may alter the mechanical properties of the engineered
barrier system components. This report focuses on the effects of thermally influenced fabrication
processes on the mechanical properties of Alloy 22.

1.1 Objective
In support of the NRC prelicensing activities on topics important to the preclosure safety analysis
and postclosure performance of the potential repository, the Center for Nuclear Waste Regulatory
Analyses is conducting an independent technical assessment of the mechanical properties of the
waste package materials. This report provides results of microstructural analysis and
measurements, impact strength, ductility, and fracture toughness for Alloy 22 as a function of
metallurgical conditions including mill-annealed, thermally aged, as-welded, welded + solution
annealed, and welded + thermally aged.

1.2 Scope and Organization of the Report


The effects of fabrication processes on the mechanical disruption of the Alloy 22 outer container
have been addressed in NRC (2001, 2002). Thermally influenced fabrication processes may alter
the mechanical properties, the range of passive film stability, and the localized corrosion resistance
of the Alloy 22 outer container, which could lead to early through-wall penetration of the waste
package. Several limitations and deficiencies have been identified by NRC (2002) in the DOE
approach and in the technical bases provided for evaluation of the effects of fabrication processes
on the performance of the Alloy 22 waste package outer container.

The proposed DOE waste package design and the approach to evaluate the effects of fabrication
processes on the mechanical properties of Alloy 22 were previously reviewed by Dunn, et al. (2004)
The DOE approach is limited to the assessment of yield strength, ductility, and impact toughness
based on data obtained with welded and thermally aged specimens. Studies about the effects of
fabrication processes on the fracture toughness of Alloy 22 have not been performed. In addition,
the DOE assessment is limited to a single welding method and has not considered the complete
range of fabrication steps necessary to construct and seal the Alloy 22 waste package outer
container.

This report is organized into five chapters including an introduction as Chapter 1. The materials
and experimental methods used in this study are discussed in Chapter 2. Results and discussion
of the microstructural analysis, Charpy V-notch impact tests, tensile tests, and fracture toughness
tests are provided in Chapter 3. An assessment of the effects of fabrication processes on the waste
package mechanical properties is included in Chapter 4 where the actual fracture toughness
measurements are compared with previous estimates (Dunn, et al., 2004). Conclusions of the
study are included as Chapter 5.

1-2
1.3 Relevant DOE and NRC Agreements
As noted, the effects of fabrication processes on the microstructure and mechanical properties of
the waste package are considered in Subissue 2 of NRC (2001) and are incorporated in the
Degradation of Engineering Barriers and Mechanical Disruption of Engineered Barriers Integrated
Subissues and the Design of Structures, Systems, and Components Important to Safety and Safety
Controls (NRC, 2002). Through the process of prelicensing consultation for issue resolution
between DOE and NRC, these subissues are considered closed-pending according to the DOE and
NRC agreements. Agreements pertaining to mechanical properties and mechanical failure of
container materials for both preclosure activities and postclosure performance are listed in Table
1-1. According to the agreements for resolving all deficiencies and limitations identified in this
report, DOE agreed to provide the additional information requested prior to license application.

Table 1-1. DOE and NRC Agreements Related to This Report


Agreement Agreement Statement
PRE.7.03* Demonstrate that the allowed microstructural and compositional variations of
Alloy 22 base metal and the allowed compositional variations in the weld filler
metals used in the fabrication of the waste packages do not result in
unacceptable waste package mechanical properties. DOE will provide
justification that the American Society of Mechanical Engineers code case for
Alloy 22 results in acceptable waste package mechanical properties
considering allowed microstructural and compositional variations of Alloy 22
base metal and the allowed compositional variations in the weld filler metals
used in the fabrication of the waste packages. DOE agrees to provide the
information in fiscal year 2003 and document the information in the
waste package design
PRE.07.04* Demonstrate that the nondestructive evaluation methods used to inspect the
Alloy 22 and 316 nuclear grade plate material and closure welds are sufficient
and are capable of detecting all defects that may alter waste package
mechanical properties. DOE will provide justification that the nondestructive
evaluation methods used to inspect the Alloy 22 and 316 nuclear grade plate
material and welds are sufficient and are capable of detecting defects that
may adversely affect waste package preclosure structural performance. DOE
agrees to provide the information in fiscal year 2003 and document the
information in the Waste Package Operations Fabrication Process Report.
PRE.07.05* Provide justification that the mechanical properties of the disposal container
fabrication and waste package closure welds are adequately represented
considering the (1) range of welding methods used to construct the disposal
containers, (2) post weld annealing and stress mitigation processes,
and (3) post weld repairs. DOE agrees to provide the information in
fiscal year 2003 and document the information in the Waste Package
Operations Fabrication Process Report.

1-3
Table 1-1. DOE and NRC Agreements Related to This Report (continued)
Agreement Agreement Statement
CLST.2.08† Provide documentation of the path forward items in the “Subissue 2: Effects
of Phase Instability of Materials and Initial Defects on the Mechanical Failure
and Lifetime of the Containers” presentation, slide 16 [future rockfall
evaluations will address (1) effects of potential embrittlement of waste
package closure material after stress annealing due to aging, (2) effects of
drip shield wall thinning due to corrosion, (3) effects of hydrogen
embrittlement on titanium drip shield, and (4) effects of multiple rock blocks
falling on waste package and drip shield; future seismic evaluations will
address the effects of static loads from fallen rock on drip shield during
seismic events] DOE stated that the rockfall calculations addressing potential
embrittlement of the waste package closure weld and rock falls of multiple
rock blocks will be included in the next revision of the Analysis Model Report
ANL–UDC–MD–000001, Design Analysis for UCF Waste Packages, to be
completed prior to license application. Rock fall calculations addressing drip
shield wall thinning due to corrosion, hydrogen embrittlement of titanium, and
rock falls of multiple rock blocks will be included in the next revision of the
Analysis Model Report ANL–XCS–ME–000001, Design Analysis for the
Ex-Container Components, to be completed prior to license application.
Seismic calculations addressing the load of fallen rock on the drip shield will
be included in the next revision of the Analysis Model
Report ANL–XCS–ME–000001, Design Analysis for the Ex-Container
Components, to be completed prior to license application.
*Reamer, C.W. “U.S. Nuclear Regulatory Commission/U.S. Department of Energy Technical Exchange and
Management on Total System Performance Assessment and Integration (August 6–10, 2001).” Letter (August 23)
to S. Brocoum, DOE. Washington, DC: NRC. 2001.
†Schlueter, J.R. “U.S. Nuclear Regulatory Commission/U.S. Department of Energy Technical Exchange and
Management on Container Life and Source Term (September 12–13, 2000).” Letter (October 4) to S. Brocoum,
DOE. Washington, DC: NRC. 2000.

1-4
2 MATERIALS AND METHODS
Fabrication of the waste package outer container will require multiple fabrication processes
including cold rolling, machining, and welding. Postweld solution annealing will be used to mitigate
residual stresses arising from longitudinal and circumferential welding used in the manufacturing
of the Alloy 22 outer disposal container. Low plasticity burnishing and laser peening are possible
processes to mitigate residual stresses in the closure welds. Fabrication processes such as
welding and postweld heat treatments have been shown to adversely affect the localized corrosion
susceptibility of Alloy 22 in hot chloride solutions (Dunn, et al., 2003). The DOE evaluation of the
effects of fabrication processes on mechanical properties is limited to a single welding method
(Summers, et al., 2000, 2002). Studies performed by Summers, et al. (2000, 2002) did not evaluate
postweld solution annealing and was restricted to tensile and impact tests. Materials tested in this
investigation include a complete range of metallurgical conditions comprising mill-annealed,
thermally aged, as-welded (gas-tungsten arc welded and gas-metal arc welded), welded + thermally
aged, and welded + solution annealed.

2.1 Materials
Alloy 22 was analyzed and tested in a variety of metallurgical conditions expected to bound the
range of conditions for the waste package outer container. Chemical compositions of the materials
used in this study are provided in Table 2-1. Alloy 22 was tested in mill-annealed and thermally
aged conditions. Welded Alloy 22 was tested in as-welded, welded + solution annealed, and
welded + thermally aged conditions. Prior to use, confirmatory chemical analyses of the plate and
filler metals were performed. Plate materials were confirmed to meet the chemical composition
specified in ASTM B 575 for UNS N06022 alloy designation (ASTM International, 2001). Filler
materials were confirmed to meet the chemical composition specified in SFA 5.14 for ERNiCrMo-10
filler metal (American Society of Mechanical Engineers, 2001b).

Table 2-1. Composition of Alloy 22 Heats and ERNiCrMo-10 Filler Metal


Material Ni Cr Mo W Fe Co Si Mn V P S Cu C
Alloy 22 Heat
2277-3-3266 Bal 21.40 13.30 2.81 3.75 1.19 0.03 0.23 0.14 0.008 0.004 N/A 0.005
12.7-mm [0.5-in]
Alloy 22 Heat
2277-1-3164 Bal 21.15 13.47 3.26 3.93 1.27 0.02 0.23 0.11 0.007 0.003 N/A 0.003
25.4-mm [1-in]
ERNiCrMo-10
filler heat
WN813 Bal 22.24 13.7 3.13 2.37 0.41 0.02 0.34 0.01 0.003 0.001 0.01 0.003
2.38-mm
[0.0938-in]
ERNiCrMo-10
filler heat
XX1977BG11 Bal 20.25 14.13 2.99 2.56 0.07 0.06 0.20 0.04 0.008 0.001 0.09 0.005
1.14-mm
[0.045-in]

2-1
Mill-annealed and thermally aged specimens were prepared by cutting the 12.7-mm [0.5-in]-thick
plate to appropriate dimensions. Welded specimens were prepared by machining 25.4-mm [1-in]-
thick Alloy 22 plates to dimensions of 7.6 × 61 cm [3 × 24 in]. The weld joint was a single U-groove
design with an included angle of 42 degrees. Welding was performed by Roben Manufacturing
Company Incorporated (Lakewood, New Jersey). Two welding methods were used, gas-tungsten
arc welding (Welding Procedure Specification 43-7-0) and a combination of gas-tungsten arc
welding of the root pass and filler passes with gas-metal arc welding (Welding Procedure
Specifications 43-7-0 and 43-3-0). Both welding methods were qualified in accordance with the
criteria identified in the American Society of Mechanical Engineers Boiler and Pressure Vessel
Code (American Society of Mechanical Engineers, 2003). Gas-tungsten arc welding was performed
using matching filler metal and a 75-percent Argon 25-percent Helium shielding gas. Multiple
passes were required to complete the weld. The maximum interpass temperature was limited to
121 °C [250 °F]. Welding was conducted using current ranging from 75 to 180 amps, an electrode
voltage of 9 to 24 volts and a travel speed of 25 to 33 cm/s [10 to 13 in/s]. Gas-metal arc welding
was also performed using matching filler metal, a 75-percent Argon 25 percent Helium shielding
gas, and a maximum interpass temperature of 121 °C [250 °F]. Welding was conducted using
current ranging from 140 to 160 amps, an electrode voltage of 9 to 24 volts and a travel speed of
25 to 33 cm/s [10 to 13 in/s].

All welded materials used in tests of mechanical properties were inspected using radiographic
testing performed by IHI Southwest Technologies (San Antonio, Texas). The radiographic testing
was performed in accordance with procedure SWR-NN-RT1 using a Sperry X-ray machine
operated at 290 kV. Film from all welds was interpreted by an American Society for Nondestructive
Testing (ASNT) level II technician. All welds met the acceptance criteria of ASME NC-5320
(American Society of Mechanical Engineers 2001a).

Thermal aging was conducted at a temperature of 870 °C [1,598 °F] in a laboratory oven for times
up to 100 hours. After reaching the desired aging time, the material was removed from the oven,
water quenched, and stamped for identification. Solution annealing was performed in the same
oven at a temperature of 1,121 °C [2,050 °F] for 20 minutes. Subsequently, specimens were
removed from the oven, water quenched, and stamped for identification.

2.2 Microstructural Analysis


Microstructural analysis was performed at Southwest Research Institute® (San Antonio, Texas)
using an optical microscope connected to an image analyzer. Metallurgical samples were prepared
using standard polishing techniques and electrochemical etching in an oxalic acid solution. A short
etching time of a few seconds was used so that removal of precipitates as a result of over etching
could be minimized. The microstructure of these welded samples was analyzed using optical
microscopy. The volume percent of the precipitates was measured from optical micrographs using
imaging analysis software along the centerline of the weld fusion zone. An areal analysis method
was used to determine the amount of precipitation.

2.3 Charpy V-Notch Impact Tests


Charpy V-notch impact tests were performed in accordance with ASTM E23 (ASTM International,
2002a) by An-Tech Laboratories (Houston, Texas). Tests were conducted on a machine with an
impact energy range to 325 Joules [240 ft•lb]. All tests were conducted at room temperature. Mill-
annealed and thermally aged specimens were cut from 12.7-mm [0.5-in]-thick Alloy 22. Thermal

2-2
aging was conducted at 870 °C [1,598 °F] for times up to 100 hours. Welded material was
sectioned and heat treated to evaluate specimens in the as-welded, welded + thermally aged, and
welded + solution annealed conditions.

2.4 Fracture Toughness Tests


Tensile tests and fracture toughness measurements were performed by Westmoreland Laboratories
(Youngstown, Pennsylvania). Tensile specimens were machined and tested per ASTM E8 (ASTM
International, 2002b). Fracture toughness (JIc) specimens were machined and tested per ASTM
E1820–99a (ASTM International, 2000). Compact tension specimens were cut from the aged
materials, notched, and fatigue precracked to a final crack length to width ratio (a/W) of
approximately 0.5. Implications of these dimensions on the validity of the fracture toughness are
discussed with the test results. The specimens were side grooved to a depth equal to 20 percent
of the normal thickness (10 percent per side) as recommended by ASTM E1820–99a (ASTM
International, 2000). The fracture toughness specimens were tested at room temperature using a
servo-hydraulic test stand and an automated computer controlled machine. Test data were
analyzed using yield strength values from the same material.

Aging treatment on fracture toughness of Alloy 22 was evaluated using a 152 × 93 × 12.7-mm [6 ×
3.7 × 0.5-in]-thick mill-annealed plate. Aging treatments were conducted at 870 °C [1,598 °F] for
30 minutes, and 1, 10, and 100 hours followed by water quench. For comparison, mill-annealed
Alloy 22 was tested as a baseline. Two fracture toughness specimens and one tensile specimen
were machined and tested for each condition.

Welded Alloy 22 plates were machined into test specimens to evaluate mechanical properties and
fracture toughness of as-welded and postweld heat-treated materials. Both gas-tungsten arc
welded and gas-metal arc welded materials were evaluated in the as-welded, welded + thermally
aged, and welded + solution annealed conditions.

2-3
3 METALLURGICAL ANALYSIS AND MECHANICAL PROPERTIES
This study was conducted to compare the microstructure and mechanical properties of Alloy 22 as
a function of metallurgical condition. Results of the metallurgical analyses and mechanical
properties of Alloy 22 in the mill-annealed, thermally aged, as-welded (gas-tungsten arc welded and
gas-metal arc welded), welded + thermally aged, and welded + solution annealed conditions are
presented and discussed in the following sections.

3.1 Metallurgical Analyses


Table 3-1 provides results of volume fraction of precipitates measured on mill-annealed Alloy 22
after thermal aging at 870 °C [1,598 °F] for 30 minutes and 1, 10, and 100 hours. Each datapoint
in Table 3-1 represents the mean and standard deviation values of 20 measurements. The volume
fraction of precipitates was observed to increase with increasing thermal aging time.

Table 3-1. Volume Fraction of Precipitates Measured on Thermally Aged Alloy 22


Amount of Precipitates
Heat Treatment Condition (Volume Percent)
Mill-annealed {870 °C [1,598 °F]/30 minutes} 1.6 ± 0.4
Mill-annealed {870 °C [1,598 °F]/1 hour} 2.3 ± 0.5
Mill-annealed {870 °C [1,598 °F]/10 hours} 6.3 ± 0.6
Mill-annealed {870 °C [1,598 °F]/100 hours} 10.7 ± 1.2

Figure 3-1 shows the appearance of the fusion zones of the gas-tungsten arc welded and gas-metal
arc welded Alloy 22 welds. Table 3-2 presents the amount of precipitates in the welded samples
in the as-welded condition and after solution annealing at 1,125 °C [2,057 °F] for 20 minutes and
thermal aging at 870 °C [1,598 °F] for 30 minutes and 1 and 10 hours. The mean and standard
deviation values reported in Table 3-2 represent 20 measurements of the volume percent of
precipitates conducted for each welded sample. The relatively large standard deviation values
indicate an inhomogeneous distribution of the precipitates in the welded samples. It is apparent
in Table 3-2 that, in all conditions, the gas-tungsten arc welded sample has a slightly larger volume
fraction of precipitates relative to the gas-metal arc welding. Both solution annealing and thermal
aging treatments enhanced precipitation of the secondary phases. The behavior is far more
pronounced when aging at 870 °C [1,598 °F] was employed. Figure 3-2 displays representative
microstructures of the as-welded, solution annealed, and thermally aged Alloy 22 welds. As seen
in Figure 3-2, aging at 870 °C [1,598 °F] for 1 hour results in a significant increase in the size and
amount of the precipitates when compared with the as-welded and solution annealed samples.

Nickel-chromium-molybdenum alloys, such as Alloy 22, undergo phase transformations when


exposed to temperatures of approximately 600 °C [1,112 °F] and above that result in the formation
of secondary topologically close-packed phases (Gozlan, et al., 1991). Microstructural
characterization of the Alloy 22 welds indicates the high-temperature stability of

3-1
Figure 3-1. Photographs Showing the Fusion Zones of the Gas-Tungsten Arc Welded
(GTAW) and Gas-Metal Arc Welded (GMAW)

Table 3-2. Volume Fraction of Precipitates Measured on Welded and Welded + Heat
Treated Alloy 22
Amount of Precipitates
Heat Treatment Condition (Volume Percent)
Gas-Tungsten Arc Welded (as-welded) 0.97 ± 0.29
Gas-Tungsten Arc Welded {1,125 °C [2,057 °F]/20 minutes} 2.9 ± 0.7
Gas-Tungsten Arc Welded {870 °C [1,598 °F]/30 minutes} 3.2 ± 1.2
Gas-Tungsten Arc Welded {870 °C [1,598 °F]/1 hour} 6.8 ± 1.9
Gas-Tungsten Arc Welded {870 °C [1,598 °F]/10 hours} 18.6 ± 4.1
Gas-Metal Arc Welded (as-welded) 0.09 ± 0.05
Gas-Metal Arc Welded {1,125 °C [2,057 °F]/20 minutes} 1.4 ± 0.4
Gas-Metal Arc Welded {870 °C [1,598 °F]/30 minutes} 4.5 ± 1.6
Gas-Metal Arc Welded {870 °C [1,598 °F]/1 hour} 6.4 ± 2.0

the topologically close-packed phases in welded material as a result of the solidification-induced


molybdenum segregation. Dunn, et al. (2003) evaluated the effect of alloying element segregation
on the upper stability temperature of the P-phase in the Alloy 22 weld. The equilibrium
thermodynamic calculation predicted a solvus temperature for the P-phase of 1,271 °C [2,320 °F]
in the interdendritic regions based on the local compositions measured in the weld, suggesting the
P-phase is stable up to that temperature. As a result, further heat

3-2
(a) As-Welded (d) As-Welded

(b) 1,125 °C/20 Minutes (e) 1,125 °C/20 Minutes

(c) 870 °C/1 Hour (f) 870 °C/1 Hour


Figure 3-2. Micrographs Showing Amount of Precipitates in Gas-Tungsten Arc Welded (Left)
and Gas-Metal Arc Welded (Right) Alloy 22 Welds in the As-Welded Condition and After Heat
Treatments at 1,125 °C [2,057 °F] for 20 Minutes and 870 °C [1,598 °F] for 1 Hour

3-3
treatments of the Alloy 22 welds below the solvus temperature for the topologically close packed
phases would enhance precipitation of the secondary phases. This enhancement is evident in the
volume fractions of precipitates in both the solution annealed and thermally aged Alloy 22 welds,
which are much higher than the values for the as-welded materials. Comparison of experimentally
measured volume fractions of the precipitates revealed different degrees of precipitation
enhancement of the topologically close packed phases in the Alloy 22 welds after solution
annealing at 1,125 °C [2,057 °F] for 20 minutes and thermal aging at 870 °C [1,598 °F] for 1 hour.
Precipitation of the topologically close packed phases is fast at the aging temperature of 870 °C
[1,598 °F] as a consequence of the overall rate of nucleation and growth during the precipitation
process. This heat-treatment temperature is close to the nose of the time-temperature-precipitation
diagram of Alloy 22 (Heubner, et al., 1989), suggesting the fastest transformation rate to occur is
at this temperature.

3.2 Charpy V-Notch Impact Tests


Figure 3-3 shows results of Charpy V-notch impact tests conducted with mill-annealed, thermally
aged, and welded Alloy 22. The Charpy V-notch impact strength was measured at room
temperature. Face-centered cubic alloys such as 300 series stainless steels and nickel-chromium-
molybdenum alloys are not susceptible to brittle fracture at low temperatures, and the ductility of
Alloy 22 is not a strong function of temperature in the range 20–300 °C [68–572 °F]. For the mill-
annealed alloy, a high impact strength of 300 J [222 ft•lb] was measured. High impact strengths
also were observed after thermal aging at 870 °C [1,598 °F] for periods of 5 and 30 minutes.
Specimens thermally aged 30 minutes or less did not break during testing. Although the results for
specimens that do not break do not meet the criteria for valid Charpy V-notch impact tests, the
values are reported in Figure 3-3 and indicate high ductility and impact strength for the mill-
annealed material and the mill-annealed alloy that was thermally aged at 870 °C [1,598 °F] up to
30 minutes. A decreased impact strength was observed after 1 hour of aging. Thereafter, the
impact strength decreased with increasing thermal aging time. After 10 hours, the impact strength
was less than 100 J [74 ft•lb] and after 100 hours, the impact strength was less than 10 J [7 ft•lb].

Lower values of impact strength were observed with both gas-tungsten arc welded and gas-metal
arc welded Alloy 22. In the as-welded condition, the impact strength of both gas-tungsten arc
welded and gas-metal arc welded Alloy 22 was in the range 125–170 J [96–123 ft•lb]. Thermal
aging at 870 °C [1,598 °F] for 30 minutes resulted in a slight decrease in impact strength for both
welding methods. After 1 hour of aging, the impact strength was 56–74 J [41–55 ft•lb] and after 10
hours, the impact strength was 22–28 J [16–21 ft•lb]. Solution annealed {1,125 °C [2,057 °F] for
20 minutes/water quenched} welds had impact strengths similar to the as-welded specimens.

Results for the mill-annealed material are similar to the results reported by Summers, et al (2002).
In the current study, a decrease in impact strength was observed after thermal aging for 1 hour at
870 °C [1,598 °F]. Summers, et al. (2002) observed a decrease in impact strength after 10 hours
at 760 °C [1,400 °F]. Accelerated precipitation of brittle topologically close-packed phases in the
grain boundaries at higher temperatures is expected owing to the large activation energy for the
diffusion of alloying elements in the austenitic matrix. Dunn, et al. (2003) previously reported aging
times of a few minutes at 870 °C [1,598 °F] increased the localized corrosion susceptibility of Alloy
22 in hot chloride solutions. Intergranular attack in crevices was attributed to the presence of
discontinuous molybdenum-rich precipitates in the

3-4
Alloy 22 Charpy Impact Tests
Mill annealed and Thermally Aged
Gas-Tungsten Arc Welded and Thermally Aged
Gas-Tungsten Arc Welded and Solution Annealed
Gas-Metal Arc W elded and Thermally Aged
Gas-Metal Arc W elded and Solution Annealed
Charpy V-notch Impact Strength, J

350

300

250

200

150

100
No Aging

50

0.1 1 10 100
Agin g Time a t 87 0 oC [1,59 8 oF], h ours

Figure 3-3. Charpy V-Notch Impact Test Results for Alloy 22

grain boundaries. The fraction of grain boundary coverage increases with thermal aging time. The
impact strength of Alloy 22 is not affected adversely until the grain boundaries are
completely covered with topologically close-packed phase precipitates (CRWMS M&O, 2000c;
Summers, et al., 2002).

No effect of the welding method was observed in the impact strength tests. Impact strengths values
were similar for the gas-tungsten arc welded and the gas-metal arc welded Alloy 22 specimens.
The volume fraction of precipitates was greater for the gas-tungsten arc welded material in the as-
welded condition. After thermal aging for 1 hour, the volume fraction of precipitates was identical
for both the gas-tungsten arc welded and the gas-metal arc welded alloy. Solution annealing
slightly increased the volume fraction of precipitates but did not result in a significant decrease in
impact strength.

3.3 Tensile Properties and Fracture Toughness


Tensile tests were conducted for mill-annealed and mill-annealed + thermally aged Alloy 22 at 870
°C [1,598 °F] for 30 minutes and 1, 10, and 100 hours. Table 3-3 shows the results of tensile
properties including ultimate tensile strength, yield strength, percent elongation, reduction in area,
and modulus.

3-5
Table 3-3. Room Temperature Tensile Properties of Alloy 22 Aged at 870 °C [1,598 °F]
Ultimate
Tensile Yield Reduction
Specimen Strength Strength Elongation in Area Modulus
Condition MPa [ksi] MPa [ksi] (%) (%) MPa [ksi]
Mill annealed 787 347 71 79 1.95 × 105
[114.2] [50.3] [2.83 × 104]
Aged 870 °C 784 358 68 68 1.92 × 105
[1,598 °F] 30 minutes [113.7] [51.9] [12.79 × 104]
Aged 870 °C 783 356 66 64 1.65 × 105
[1,598 °F]1 hour [113.6] [51.7] [2.39 × 104]
Aged 870 °C 789 397 57 43 1.74 × 105
[1,598 °F] 10 hours [114.5] [53.2] [2.53 × 104]
Aged 870 °C 681 360 17 16 1.82 × 105
[1,598 °F] 100 hours [98.7] [52.2] [2.64 × 104]

The ultimate tensile strength and yield strength remain essentially constant for aging times to
10 hours. Aging for 100 hours decreased the ultimate tensile strength from 787 MPa [114.2 ksi]
to 681 MPa [98.7 ksi]. The ductility of the material, however, decreases gradually with increasing
aging times. For example, the percentage elongation of the alloy in the mill-annealed condition was
71 percent. After aging 1 hour at 870 °C [1,598 °F], the elongation was reduced to 66 percent, and
after 10 hours, the elongation was reduced to 57 percent. A prolonged aging for 100 hours at 870
°C [1,598 °F] reduced the elongation value to 17 percent.

Fracture toughness of Alloy 22 was determined using J-integral tests which consisted of both elastic
and plastic components (Anderson, 1995). Table 3-4 provides the room temperature values of
fracture toughness for the mill-annealed and mill-annealed + thermally aged Alloy 22. In Table 3-4
the parameters of J (in KJ/m2) and K (in MPa•m½) were used for determination of plane-strain
fracture toughness. The fracture toughness value KJIc decreased as aging time increased. Two
specimens aged at 870°C [1,598 °F] for 100 hours yielded valid JIc values. These values are
representative of the actual fracture toughness of the material after aging for 100 hours at 870 °C
[1,598 °F]. For these specimens, KJIc can be calculated from the Eq. (3-1)

JIcE
K JIc = (3-1)
1− ν 2
where E is Young’s modulus and ν is Poisson’s ratio.

The remaining specimens did not yield valid JIc values. These specimens had fracture toughness
values higher than could be measured using the specimen size tested. A larger specimen could
not be employed because the initial plate thickness was limited to 12.7 mm [0.5 in]. Although these
values are invalid because of size limitations, the results could be ranked by magnitude, keeping
in mind the values may not represent the absolute fracture toughness of the material tested. It is
apparent from the values listed in Table 3-4 the fracture toughness of Alloy 22 is high and only
diminished after extensive thermal aging.

3-6
Table 3-4. Room Temperature Fracture Toughness of Alloy 22 Thermally Aged at
870 °C [1,598 °F]
Specimen
Condition J kJ/m2 [ft•lb/in2]* KJ, MPa•m1/2 [ksi•in1/2]*
Mill-annealed 1,633 [781] 643 [585]
1,520 [726] 621 [565]
Thermally aged at 1,176 [562] 517 [470]
870 °C [1,598 °F]
30 minutes 1,286 [615] 540 [491]

Thermally aged at 1,054 [504] 488 [445]


870 °C [1,598 °F] 1 hour
1,196 [572] 521 [474]
Thermally aged at 366 [175] 288 [262]
870 °C [1,598 °F] 10 hours
362 [173] 287 [261]
Thermally aged at 40 [19]† 95 [87]‡
870 °C [1,598 °F] 100 hours
43 [21]† 99 [90]‡
*ASTM International. “Standard Test Method for Measurement of Fracture Toughness.” E1820–99a.
Annual Book of Standards. Vol. 3.01. West Conshohocken, Pennsylvania: ASTM International. 2000.
†valid JIc
‡valid KJIc

According to ASTM 1820 (ASTM International, 2000) the criteria for valid specimen thickness is
defined in Eq. (3-2).

Jq
B > 25 (3-2)
1 / 2( YS + UTS)
where B is the specimen thickness, Jq is the measured value of fracture toughness, YS is the yield
strength and UTS is the ultimate tensile strength. For ductile materials with relatively low yield and
ultimate tensile strengths, the required specimen thickness can be greater than 5 cm [2 in] thick.
For example, in order to be a valid JIc measurement, the mill annealed Alloy 22 material would need
to be 7.24 cm [2.85 in] thick. Solving Eq. (3-2) for Jq and using typical values of the yield and
ultimate tensile strength for the welded materials shows that for a 12.7 mm [0.5 in] thick Alloy 22
material, the maximum value for valid JIc measurements is 305 kJ/m2 [146 ft lb/in2]. Because this
value of fracture toughness is sufficiently high to assure failure of Alloy 22 will not occur as a result
of fracture (i.e., failure mode will be ductile), additional tests with thicker materials were considered
unnecessary.

Table 3-5 provides room temperature tensile data for the welded specimens. In general, the yield
strength was greater for the welded material than for the mill-annealed material. Thermal aging at
870 °C [1,598 °F] for 1 hour did not alter the yield strength of the gas-tungsten arc

3-7
Table 3-5. Room Temperature Tensile Properties of Welded and Heat-Treated Alloy 22
Tensile Yield Reduction
Specimen Strength Strength Elongation in Area Modulus
Condition MPa [ksi] MPa [ksi] (%) (%) MPa [ksi]
GTAW* as-welded 794 430 60 78 1.92 × 105
[115.2] [62.4] [2.78 × 104]
GTAW aged 870 °C 786 426 55 65 2.06 × 105
[1,598 °F] 1 hour [114.0] [61.8] [2.99 × 104]
GTAW aged 870 °C 616 441 10 18 1.94 × 105
[1,598 °F] 10 hours [89.4] [64.0] [2.81 × 104]
GTAW aged 870 °C 781 440 14 22 1.86 × 105
[1,598 °F] 100 hours [113.3] [63.9] [2.70 × 104]
GTAW solution 761 397 49 51 1.48 × 105
annealed 1,125 °C [110.3] [57.6] [2.14 × 104]
[2,057 °F]
20 minutes
GMAW† as-welded 780 545 46 58 1.90 × 105
[113.2] [79.1] [2.75 × 104]

GMAW aged 870 °C 776 364 37 28 1.91 × 105


[1,598 °F] 1 hour [112.6] [52.8] [2.77 × 104]
GMAW aged 870 °C 816 443 36 41 1.83 × 105
[1,598 °F] 10 hours [118.4] [64.2] [2.66 × 104]
GMAW aged 870 °C 758 448 13 21 2.26 × 105
[1,598 °F] 100 hours [110.0] [65.0] [3.28 × 104]
GMAW solution 587 386 19 29 1.84 × 105
annealed 1,125 °C [85.1] [56.1] [2.67 × 104]
[2,057 °F]
20 minutes
*GTAW = gas-tungsten arc welded
†GMAW = gas-metal arc welded

welded material, whereas the same thermal aging conditions decreased the yield strength of the
gas-metal arc welded alloy. Solution annealing at 1,125 °C [2,057 °F] did not affect the ductility
of the gas-tungsten arc welded alloy. In contrast, solution annealing significantly decreased
ductility of the gas-metal arc welded alloy. Note that the ductility of the gas-metal arc welded +
thermal aged specimen {870 °C [1,598 °F] for 1 hour} is also relatively low as compared to its
counterparts in gas-tungsten arc welded + thermal aging condition. Increasing the aging time
resulted in a decrease in the ductility of the gas-tungsten arc welded material. Results for the
gas-metal arc welded material had greater variability, however, thermal aging for extended
times resulted in a significant reduction in ductility.

3-8
Table 3-6 shows the fracture toughness values after welding and postweld heat treatments.
One set of welded specimens was subjected to aging treatment at 870 °C [1,598 °F] for 1 hour.
Another set of welded specimens was subjected to solution annealing treatment at 1,125 °C
[2,057 °F] for 20 minutes. In the as-welded conditions, the gas-tungsten arc welded and the
gas-metal arc welded specimens both showed the highest fracture toughness values of KJIc.
Aging at 870 °C [1,598 °F] for 1 hour or solution annealing treatment at 1,125 °C [2,057 °F] for
20 minutes decreased the KJIc values. Although none of the specimens produced valid fracture
toughness values because of specimen size limitations, it is apparent the fracture toughness
values of Alloy 22 are high in the as-welded, welded + thermally aged, and welded + solution
annealed conditions, indicating resistance to fracture.

Table 3-6. Fracture Toughness Values of Welded and Welded + Heat Treated Alloy 22
Specimen Condition J kJ/m2 [ft•lb/in2]* KJ MPa•m1/2 [ksi•in1/2]*
GTAW† as-welded 840 [401] 443 [403]
657 [314] 392 [357]
GTAW Aged 870 °C 227 [109] 231 [210]
[1,598 °F] 1 hour
425 [203] 316 [288]
GTAW Aged 870 °C 95 [45] 151 [137]
[1,598 °F] 10 hours
101 [48] 156 [142]
GTAW Aged 870 °C 47 [23] 106 [97]
[1,598 °F] 100 hours
42 [20] 100 [91]
GTAW solution annealed 1,043 [499] 505 [460]
1,125 °C [2,057 °F]
20 minutes 712 [340] 417 [380]
GMAW‡ as-welded 893 [427] 460 [418]
687 [328] 287 [367]
GMAW Aged 870 °C 299 [143] 264 [240]
[1,598 °F] 1 hour
437 [209] 319 [290]
GMAW Aged 870 °C 108 [52] 164 [149]
[1,598 °F] 10 hours
114 [55] 169 [154]
GMAW Aged 870 °C 57 [27] 119 [109]
[1,598 °F] 100 hours
46 [22] 107 [97]
GMAW solution annealed 664 [317] 394 [359]
1,125 °C [2,057 °F]
20 minutes 767 [367] 424 [386]
*ASTM International. “Standard Test Method for Measurement of Fracture Toughness.” E1820–99a:
Annual Book of Standards. Vol. 3.01. West Conshohocken, Pennsylvania: ASTM International. 2000.
†GTAW = gas-tungsten arc welded
‡GMAW = gas-metal arc welded

3-9
4 ASSESSMENT OF FABRICATION PROCESSES
The potential waste package design includes an Alloy 22 outer container with a Type 316
nuclear grade stainless steel inner container. Although the effects of fabrication processes on
the mechanical properties of Type 316 stainless steels were evaluated by Dunn, et al. (2004),
limited data were available on the effects of fabrication processes on the mechanical properties
of Alloy 22. Dunn, et al. (2004) estimated the fracture toughness of as-welded and welded +
thermally aged Alloy 22 using the Charpy V-notch data reported by Summers, et al. (2000,
2002) and the relationship of the Charpy V-notch impact strength to fracture toughness
developed for ferritic pressure vessel steels with yield strengths of more than 690 MPa [100 ksi]
(Barsom and Rolfe, 1970; Rolfe and Novak, 1970).

2
 K Ic   CVN 
  = 5.0  − 0.05 (4-1)
 σ YS   σ YS 

Although Iwadate, et al. (1977) used the relationship to estimate the KIc from Charpy V-notch
impact energy data for 21/4 Cr-1 Mo pressure vessel steels with yield strengths in the range
405–620 MPa [59–90 ksi], validity of the relationship for ductile nickel-chromium-molybdenum
alloys was not established. Nevertheless, estimated values of fracture toughness using
Eq. (4-1) were used to construct a failure assessment diagram (Dunn, et al., 2004). For
mill-annealed Alloy 22, the fracture toughness was estimated to be 304 MPa•m1/2 [276 ksi•in1/2].
In the as-welded condition, the fracture toughness was estimated using Eq. (4-1) to be
261 MPa•m1/2 [237 ksi•in1/2] and 63 MPa•m1/2 [57 ksi•in1/2] in the welded + thermally aged
condition {870 °C [1,598 °F] for 1 hour} based on an extrapolation of data for Alloy 22 thermally
aged at 593 to 760 °C [1,100 to 1,400 °F]. The extrapolated values of Charpy impact strength
in Dunn, et al. (2004) are significantly lower than the measured values included in this report.
For example, Dunn, et al. (2004) estimated the Charpy impact strength for welded Alloy 22
thermally aged at 870 °C [1,598 °F] for 1 hour to be 30 J [22 ft lb]. Measured values for the
Charpy impact strength for welded Alloy 22 thermally aged at 870 °C [1,598 °F] for 1 hour
shown in Figure 3-3 of this report are in the range of 75 to 100 J [56 to 74 ft lb], significantly
greater then the estimated values.

Although the measurements of fracture toughness of as-welded and welded + solution annealed
Alloy 22 are not valid according to the test standard criteria (ASTM International, 2000), data
provided in Table 3-6 clearly indicate the actual fracture toughness of welded + thermally aged
material is significantly greater than previously estimated by Dunn, et al. (2004). In addition, the
fracture toughness of the welded + solution annealed material was not reduced compared with
the as-welded condition. Data for the fracture toughness of the welded + solution annealed
material are consistent with the Charpy V-notch impact energy and the observed changes in the
volume fraction of precipitates in the weld microstructure.

Considering the data presented in this report, a significant reduction in the fracture toughness of
the waste packages as a result of welding and postweld heat treatment, including an improper
postweld heat treatment, seems unlikely. Although large variations and an overall reduction in
ductility were observed in the welded, welded + thermally aged, and welded + solution annealed
conditions, the fracture toughness of Alloy 22 is sufficiently high in all conditions tested to
assure that failure of the waste package outer containers by fracture will not occur. The failure
mode of Alloy 22 waste package outer containers including fabrication and closure welds is
expected to remain in the ductile collapse regime.

4-1
5 CONCLUSIONS
Microstructure, impact strength ductility, and fracture toughness of Alloy 22 were evaluated as a
function of metallurgical condition. Changes in the microstructure as a result of thermal aging at
temperatures that promote the precipitation of topologically close-packed phases can alter the
mechanical properties. Although short-term thermal aging is known to increase the localized
corrosion susceptibility of Alloy 22 in hot chloride solutions, no significant alteration of
mechanical properties was observed as a result of thermal aging at 870 °C [1,598 °F] for times
to 1 hour. Longer aging times were observed to decrease impact strength and fracture
toughness. The effect on these parameters, and on the ductility was more pronounced at an
aging time of 100 hours.

Both gas-tungsten arc welded and gas-metal arc welded Alloy 22 had higher yield strengths
and lower impact strengths compared with the mill-annealed material. Thermal aging for a
period of 1 hour at 870 °C [1,598 °F] did not result in a significant decrease in impact strength
or fracture toughness. Although solution annealed welds had larger fraction of topologically
close-packed phase precipitates, the impact strength and fracture toughness were similar to the
as-welded material.

Fabrication of the Alloy 22 container will require multiple processes including welding and
solution annealing. Results of this study suggest the thermal aging of Alloy 22 is unlikely to
affect adversely the mechanical properties of the waste package outer container. The high
fracture toughness, impact strength, and ductility of Alloy 22 are sufficient to ensure the
mechanical properties of the fabricated waste packages are acceptable.

5-1
6 REFERENCES
Anderson, T.L. “Fracture Mechanics: Fundamentals and Applications.” 2nd Edition. CRC
Press. Boca Raton, Florida. 1995.

American Society of Mechanical Engineers. “Welding and Brazing Qualifications.” 2003


Addenda Section IX. ASME Boiler and Pressure Vessel Code. New York City, New York:
American Society of Mechanical Engineers. 2003.

––––––. “Rules for Construction of Nuclear Power Plant Components, Division 1,


Subsection NC, Class 2 Components.” Section III. ASME Boiler and Pressure Vessel Code.
New York City, New York: American Society of Mechanical Engineers. 2001a.

––––––. “Materials, Part C—Specifications for Welding Rods, Electrodes, and Filler Metals.”
Section II. ASME Boiler and Pressure Vessel Code. New York City, New York: American
Society of Mechanical Engineers. 2001b.

ASTM International. “Standard Test Methods for Notched Bar Impact Testing of Metallic
Material.” E–23–02. Annual Book of Standards. Vol. 3.01. West Conshohocken,
Pennsylvania: ASTM International. 2002a.

––––––. “Standard Test Methods for Tension Testing of Metallic Materials.” E–8-01:
Annual Book of Standards. Vol. 3.01. West Conshohocken, Pennsylvania:
ASTM International. 2002b.

––––––. “Standard Specification for Low-Carbon Nickel-Molybdenum-Chromium, Low-Carbon


Nickel-Chromium-Molybdenum, Low-Carbon Nickel-Chromium-Molybdenum-Copper,
Low-Carbon Nickel-Chromium-Molybdenum-Tantalum, and Low-Carbon Nickel-Chromium-
Molybdenum-Tungsten Alloy Plate, Sheet, and Strip.” ASTM B–575–99a. Annual Book of
Standards. Volume 3.02: Wear and Erosion—Metal Corrosion. West Conshohocken,
Pennsylvania: ASTM International. 2001.

––––––. “Standard Test Method for Measurement of Fracture Toughness.” E–1820–99a:


Annual Book of Standards. Vol. 3.01. West Conshohocken, Pennsylvania:
ASTM International. 2000.

Barsom, J.M. and S.T. Rolfe. “Correlations Between KIc and Charpy V-Notch Test Results
in the Transition-Temperature Range.” Impact Testing of Metals. ASTM STP 466.
West Conshohocken, Pennsylvania: ASTM International. pp. 281–302. 1970.

Bechtel SAIC Company LLC. “Waste Package Operations Fabrication Process Report.”
TDR–EBS–ND–000003. Rev. 02. Las Vegas, Nevada: Bechtel SAIC Company LLC. 2001.

CRWMS M&O. “Repository Safety Strategy: Plan to Prepare the Safety Case to Support
Yucca Mountain Site Recommendation and Licensing Considerations.” TDR–WIS–RL–000001.
Rev. 04 ICN 01. Las Vegas, Nevada: CRWMS M&O. 2000a.

––––––. “Total System Performance Assessment for the Site Recommendation.”


TDR–WIS–PA–000001. Rev. 00 ICN 01. Las Vegas, Nevada: CRWMS M&O. 2000b.

6-1
––––––. “Aging and Phase Stability of Waste Package Outer Barrier.” ANL–EBS–MD–000002.
Rev. 00. Las Vegas, Nevada: CRWMS M&O. 2000c.

Dunn, D.S., Y.M. Pan, D. Daruwalla, and A. Csontos. “The Effects of Fabrication Processes on
the Mechanical Properties of Waste Packages—Progress Report.” San Antonio, Texas:
CNWRA. 2004.

Dunn, D.S., D. Daruwalla, and Y.-M. Pan. “Effect of Fabrication Processes on Material
Stability—Characterization and Corrosion.” CNWRA 2004-01. San Antonio, Texas:
CNWRA. 2003.

Gozlan, E., M. Bamberger, S.F. Dirnfeld, B. Prinz, and J. Klodt. “Topologically Close-Packed
Precipitations and Phase Diagrams of Ni-Mo-Cr and Ni-Mo-Fe and of Ni-Mo-Fe with Constant
Additions of Chromium.” Materials Science and Engineering. Vol. A141. pp. 85–95. 1991.

Heubner, U.L., E. Alepeter, M.B. Rockel, and E. Wallis. “Electrochemical Behavior and Its
Relation to Composition and Sensitization of Ni-Cr-Mo Alloys in ASTM G28 Solution.”
Corrosion. Vol. 45, No. 3. pp. 249–259. 1989.

Iwadate, T., T. Karaushi, and J. Watanabe. “Prediction of Fracture Toughness KIc of 21/4Cr-1Mo
Pressure Vessel Steels from Charpy V-Notch Test Results.” Flaw Growth and Fracture.
ASTM STP 631. Philadelphia, Pennsylvania: American Society for Testing and Materials.
pp. 493–506. 1977.

NRC. NUREG–1762, “Integrated Issue Resolution Status Report.” Rev. 0. Washington, DC:
NRC. August 2002.

––––––. “Issue Resolution Status Report, Key Technical Issue: Container Life and Source
Term.” Rev. 3B. Washington, DC: NRC. 2001.

Rolfe, S.T. and S.R. Novak. “Slow-Bend KIc Testing of Medium-Strength,


High-Toughness Steels.” Review of Developments in Plane Strain Fracture-Toughness
Testing. ASTM STP 463. West Conshohocken, Pennsylvania: ASTM International.
pp. 124–159. 1970.

Summers, T.S.E., R.B. Rebak, T.A. Palmer, and P. Crook. “Influence of Thermal Aging on
the Mechanical and Corrosion Properties of GTAW Welds on Alloy N06022.” Scientific
Basis for Nuclear Waste Management XXV. Symposium Proceedings 713. B.P. McGrail
and G.A. Cragnolino, eds. Warrendale, Pennsylvania: Materials Research Society.
pp. 45–52. 2002.

Summers, T.S.E., R.B. Rebak, and R.R. Seeley. “Influence of Thermal Aging on the
Mechanical and Corrosion Properties of C-22 Alloy Welds.” UCRL–JC–137727.
Livermore, California: Lawrence Livermore National Laboratory. 2000.

6-2

You might also like